Next Article in Journal
Study of Calcination Temperature Influence on Physicochemical Properties and Photodegradation Performance of Cu2O/WO3/TiO2
Previous Article in Journal
Construction of A NiS/g-C3N4 Co-Catalyst-Based S-Scheme Heterojunction and Its Performance in Photocatalytic CO2 Reduction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advanced Biocatalytic Processes for the Conversion of Renewable Feedstocks into High-Value Oleochemicals

by
João H. C. Wancura
1,
Eliane Pereira Cipolatti
2,
Evelin Andrade Manoel
3,4,
Febri Odel Nitbani
5,
Angie Vanessa Caicedo-Paz
6,
Cassamo Ussemane Mussagy
6,
Tamer M. M. Abdellatief
7,
Ahmad Mustafa
8 and
Luigi di Bitonto
9,*
1
Laboratory of Biomass and Biofuels (L2B), Federal University of Santa Maria, Santa Maria 97105-900, RS, Brazil
2
Chemical Engineering Department, Institute of Technology, Universidade Federal Rural do Rio de Janeiro (UFRRJ), Seropédica 23890-000, RJ, Brazil
3
Pharmaceutical Biotechnology Program, Faculty of Pharmacy, Federal University of Rio de Janeiro, Rio de Janeiro 21941-853, RJ, Brazil
4
Biochemistry Department, Chemistry Institute, Federal University of Rio de Janeiro, Rio de Janeiro 21941-901, RJ, Brazil
5
Department of Chemistry, Faculty of Science and Engineering, University of Nusa Cendana, Jl. Adisucipto, Penfui, Kupang 85001, Nusa Tenggara Timur, Indonesia
6
Escuela de Agronomía, Facultad de Ciencias Agronómicas y de los Alimentos, Pontificia Universidad Católica de Valparaíso, Quillota 2260000, Chile
7
Chemical Engineering Department, Faculty of Engineering, Minia University, EL-Minya 61519, Egypt
8
Center of Excellence, October University for Modern Sciences and Arts (MSA), Giza 12451, Egypt
9
Water Research Institute (IRSA), National Research Council (CNR), Via F. de Blasio 5, 70132 Bari, Italy
*
Author to whom correspondence should be addressed.
Catalysts 2025, 15(6), 600; https://doi.org/10.3390/catal15060600
Submission received: 9 April 2025 / Revised: 14 June 2025 / Accepted: 15 June 2025 / Published: 17 June 2025
(This article belongs to the Special Issue Sustainable Enzymatic Processes for Fine Chemicals and Biodiesel)

Abstract

Oleochemicals, which are obtained from vegetable and animal fats and oils, have become indispensable in the food, cosmetics, pharmaceutical and biofuel industries. Traditionally, they are synthesized using chemical catalysts, a process that is often associated with high energy requirements and a considerable environmental impact. Biocatalysis, using enzymes such as lipases, has emerged as a transformative alternative that offers high specificity, environmental friendliness and cost-efficiency. This review comprehensively examines the current state of biocatalysis for oleochemical synthesis, highlighting key reactions such as esterification and transesterification and their integration into industrial processes. A bibliometric analysis uncovers global trends and collaborations, while case studies illustrate cost efficiency and scalability. The article outlines recommendations and future research directions to advance biocatalytic processes. This review is intended to be an important resource for researchers and industries transitioning to sustainable oleochemical production.

Graphical Abstract

1. Introduction

Oleochemicals are biodegradable compounds consisting of a diverse group of aliphatic chemicals derived from plant or animal sources, usually derived from vegetable oils or fats. These compounds are versatile raw materials as they can be used in a variety of industries for the production of a wide range of products such as detergents, lubricants, emulsifiers, polymers, cosmetics, food additives, pharmaceuticals, herbicides, fuels and others [1,2]. The industrial processing of these raw materials leads to chemical reactions that produce various compounds which, depending on the technology used, pose significant hazards and usually require appropriate management to prevent damage to water and soil ecosystems or air pollution [3,4].
As an alternative, the scientific endeavours of recent decades suggest the use of enzymes as biocatalysts for oleochemical production processes. Their use in various processes is associated with several advantages over chemical catalysts, mainly because each enzyme is highly specific with respect to the substrate it catalyses, which eliminates the possibility of unwanted parallel reactions, facilitates process control and produces products of high purity [5]. A classic example is the production of biodiesel from oily sources, where conventional alkaline transesterification requires high-purity feedstocks (moisture < 200 ppm and free fatty acids < 0.5%wt) to prevent saponification of the triglycerides [6,7].
Replacing the chemical catalyst with an enzyme allows the use of feedstocks with a high free fatty acid contents, as lipases can catalyse both the esterification of the free fatty acids and the transesterification of their triglycerides to methyl esters of fatty acids, the main component of biodiesel [8,9]. This approach can significantly reduce process costs. In addition, the optimal reaction conditions commonly used in enzymatic synthesis are mild, which significantly reduces the energy requirements of the process [10,11].
Large-scale enzymatic reactions are increasingly seen as a sustainable alternative to conventional chemical processes. Although biocatalysts were more expensive than chemical catalysts in the past [12,13], recent advances have significantly improved their cost/benefit ratio. The favourable cost/benefit ratio of enzymatic processes makes biocatalysis a promising candidate for scaling up from the laboratory to the industrial level. In addition to their well-known environmental benefits, enzymatic processes are also becoming economically friendly and offer a sustainable solution that maintains profitability.
The optimization of these technologies is expanding their application possibilities so that they can compete with conventional methods and often outperform them in terms of efficiency and environmental impact. In addition, a clear understanding of the economic benefits of biocatalysis is driving innovation that reduces costs and improves process performance. Over the last decade, significant progress has been made in reducing the cost of enzymes and expanding their industrial use through successful pilot and commercial applications [14,15].
This review aims to fill a critical gap in the literature by comprehensively evaluating the cost efficiency of biocatalysis in oleochemical synthesis. The cost components of enzymatic processes, including enzyme production and operating costs, are evaluated in comparison to conventional chemical catalysts. Successful applications of biocatalysis, such as the production of biodiesel and the utilization of by-products such as glycerol, are highlighted through an analysis of case studies. This review also examines the trade-offs between economic feasibility and sustainability and shows how biocatalysis can strike a balance between these factors. Finally, future research priorities and technological innovations are outlined to overcome the current economic barriers and make biocatalysis a practical and scalable solution for the oleochemical industry. By addressing these aspects, this review aims to guide researchers and industry representatives towards the utilization of biocatalysis as a cost-effective and sustainable alternative for oleochemical synthesis.

2. Biocatalysis in Oleochemistry: Current Status and Progress

2.1. Oleochemical Synthesis and Chemical Modification of Fatty Acids: An Overview

Throughout the history of oleochemistry, various methods have been essential for the synthesis and modification of fatty acids and oils and have developed in parallel with advances in chemical engineering, biotechnology and materials science. According to Behr and Seidensticker [16], oleochemicals can be categorized into basic types, including fatty acids, fatty acid methyl esters or other alkyl groups, fatty alcohols and fatty amines. Industrial oil products also include vegetable and seed oils, which are primarily obtained from animal or vegetable sources by processes such as extraction, refining, conversion and stabilization. Firstly, the oils are extracted or crushed by mechanical pressing and/or solvent extraction. This is followed by the refining process, in which pigments, metals, phospholipids, flavourings and other undesirable substances are removed. At this stage, the oil can be modified by hydrogenation, winterization or crystallization. After refining, the fatty acids are stabilized in order to achieve certain properties, such as greater stability and safety for the final products [17,18].
In industry, fats and oils are often modified by chemical or biocatalytic reactions. Common chemical strategies include hydrolysis in alkaline media and high-pressure steam. Alkaline hydrolysis usually produces fatty acid soaps, which require further treatment, e.g., acidification, to obtain pure free fatty acids (FFAs). In this reaction, fats and oils are hydrolysed with sodium or potassium hydroxide. The search for new methods to produce fatty acid derivatives began in the late 19th century. Twitchell and Colgate-Emery, for example, used an acidic naphthalene stearosulphonic acid catalyst for industrial fatty acid modification [19]. Although this approach was an improvement over previous methods, it was associated with problems such as high energy requirements and corrosiveness, resulting in lower quality products. To overcome these problems, the process was modified and converted from a batch to a continuous process.
The chemical processes in oleochemistry are primarily hydrolysis reactions, which are influenced by factors such as temperature and pressure. In addition, other reagents can be used in the industry instead of water, leading to the formation of various fatty acid derivatives, e.g., methanol in methanolysis or amines in aminolysis. Other reactions such as hydrogenation, esterification, transesterification, oxidation and epoxidation are examples of reactions that can be carried out using chemical methods [20].
Figure 1 gives an overview of common reactions that are used in chemical processes involving fatty acids as a model for oleochemical products.
Biocatalysis for the modification of fatty acids has been researched since the early 1900s, with enzymatic hydrolysis proving to be a promising application for the chemical industry. The advantages of using enzymes, both in soluble (non-immobilized) and immobilized form, as well as whole cells and engineered modified microorganisms are widely recognized. Biocatalysis can work through different mechanisms, such as the chemical modification of fatty acids and oils from renewable resources or the conversion of fatty acids into alkyl esters that can be applied in several industrial contexts [21].
Research in the field of oleochemical biocatalysis focuses on the development of sustainable, cost-effective methods for the synthesis of oleochemicals with multiple applications, taking into account the challenges related to enzyme stability, substrate compatibility and process scalability [5]. These methods offer notable environmental advantages as they produce bioproducts with low toxicity and high biodegradability and operate under milder conditions [22], reducing the formation of by-products compared to conventional chemical processes and offering flexibility in reactor design. Despite these advantages, challenges such as high energy consumption, the cost of the enzymes and problems related to enzyme inactivation and stability remain. However, the ability to reusing biocatalysts can mitigate these concerns, making the overall process economically viable and environmentally sustainable. Recently, enzymatic reactions and whole-cell biotransformations have gained importance due to their high selectivity, mild operating conditions and sustainability. Enzymes such as lipases [22,23,24,25], phospholipases [26,27], esterases, yeast P450-monooxygenases [28,29,30,31] and oxidases are used to catalyse various transformations in oleochemistry, including transesterification, hydrolysis and oxidation reactions. As a result, enzymes are increasingly used in the synthesis of many oleochemical products, as objectively illustrated in Figure 1.
Enzymes such as lipases can also be combined with other chemical reactions. For example, enzymes do not directly catalyse hydrogenation reactions, which involve the addition of hydrogen atoms to unsaturated compounds, typically carbon–carbon double bonds, to produce saturated compounds. However, enzymes can indirectly influence hydrogenation processes through their role in metabolic pathways. For instance, certain enzymes involved in lipid metabolism can catalyse the synthesis or degradation of unsaturated fatty acids, which are substrates for industrial hydrogenation reactions. This relationship allows the production of modified fats tailored to specific properties.
Lipases (EC 3.1.1.3, triacylglycerol hydrolases) are most frequently used enzymes in biocatalytic processes for the conversion of oleochemical compounds [32]. This preference is mainly due to the fact that fats and oils serve as natural substrates that do not require cofactors and that lipases are chemo- and regioselective and exhibit high activity and stability [33,34,35,36]. Numerous lipases have been documented to be used in various oleochemical applications, including those from Rhizomucor miehei, Thermomyces lanuginosus, Burkholderia cepacia, Candida rugosa and Rhizopus oryzae [37,38].
In oleochemistry, lipases are frequently used for the synthesis of fatty acid methyl esters (e.g., biodiesel) and the modification of lipids [39]. A typical example is the use of ricinoleic acid as a substrate for lipase catalysis. This compound, which is also known as cis-12-hydroxy-9-octadecenoic acid or 12-hydroxyoleic acid, is contained in large quantities in castor oil. This unsaturated fatty acid can be used as a starting material for the production of various compounds with specific industrial applications, such as lubricants, polyesters and other oleochemical products [40,41,42,43,44,45].
Further applications can be found in the production of chemical products for the cosmetics industry, in particular through esterification reactions catalysed by biocatalysts. In addition, these processes are used in the food industry to convert conventional oils into structured lipids with functional properties that benefit human metabolism; this is performed by modifying the fatty acids at specific positions of the triglycerides through interesterification. Various emollient esters such as myristyl myristate, decyl cocoate, acetyl ricinoleate and isocetyl palmitate can be obtained from natural sources such as coconut oil or kernel oil [46].
Other enzymes such as thioesterases, phospholipases and lipoxygenases have also been extensively studied in biocatalysis and oleochemistry [47]. Lipoxygenases, for example, play a crucial role in the production of hydroperoxides, important intermediates in the synthesis of oleochemicals. However, despite their potential, the industrial application of these enzymes remains difficult due to problems such as low stability, limited catalytic activity and the lack of efficient overexpression systems capable of producing sufficient enzyme quantities [48]. Conversely, phospholipases can be used to upgrade waste cooking oils to oleochemical feedstocks [49].
Upgrading processes such as degumming, neutralization, bleaching and deodorization are essential steps in the production of high-quality vegetable oils for various oleochemical applications. These steps improve the quality, stability and nutritional properties of vegetable oils and thus support their use in oleochemistry.

2.2. Recent Innovations and Chemical Mechanisms of Biocatalysis in Oleochemistry

In recent years, innovative strategies have been developed to make oleochemical processes more profitable and scalable, including the increasing use of microorganisms as bioreagents for the production of intermediates or high-value products. This is made possible by genetic manipulation to increase the yield and scalability and the use of low-cost matrices [50,51]. Metabolic engineering serves as a key tool in the search for energy-rich oleochemicals, liquid fuels for transport and high-value chemicals, especially from renewable resources, by optimizing genetic and regulatory processes to increase cellular production while minimizing energy input [52]. The most common strategies include blocking competing metabolic pathways, redirecting metabolism to oil synthesis pathways and the overexpression of specific enzymes, which are often accompanied by manipulation of tolerant cells to intermediates to reduce stress. The integration of biocatalysis into oleochemical processes aims to develop more sustainable and efficient methods, such as the production of biopolymers, enzymatic biodiesel and the valorisation of glycerol to high-value compounds such as 1,3-propanediol, acrolein and epichlorohydrin using green solvents and ionic liquids to improve efficiency and environmental sustainability [15].
On the chemical side, enzymatic oleochemistry involves key reactions such as hydrolysis, acidolysis, esterification, transesterification, enzymatic oxidation and enzymatic epoxidation, which are summarized in Equations (1)–(6).
Triglyceride + 3 H 2 O   L i p a s e   Glycerol + 3 Free   fatty   acids
Ester   # 1 + Fatty   acid   # 2   L i p a s e   Ester   # 2 + Fatty   acids   # 1
Fatty   acid + Alcohol   L i p a s e   Ester + H 2 O
Triglyceride + Alcohol   L i p a s e   3 Acyl   esters + Glycerol
Unsaturated   fatty   acid + O 2   O x i d a s e   Hydroperoxide
Unsaturated   fatty   acid + H 2 O 2   P e r o x i d a s e   Epoxidized   fatty   acid + H 2 O
Looking at the reaction mechanism of these reactions, enzymatic catalysis begins with the recognition and activation of the substrate by the biocatalyst, a critical step to ensure the specificity and efficiency of the process. For lipases, this involves the formation of an enzyme–substrate complex facilitated by interfacial activation, in which hydrophobic interfaces lead to a conformational change that exposes the active site of the enzyme. Once the substrate is positioned at the catalytic site, it is oriented to favour the cleavage or formation of bonds, such as ester or amide bonds.
Lipases operate via a conserved catalytic triad of serine, histidine and an acid residue (aspartic acid or glutamic acid). In this mechanism, serine acts as a nucleophile and attacks the electrophilic carbonyl carbon of the substrate, while histidine removes a proton from serine, which increases its reactivity. The acid residue stabilizes the histidine through electrostatic interactions and thus promotes proton transfer.
This process leads to the formation of a covalent “acyl-enzyme” intermediate, a key step in hydrolysis, esterification and transesterification processes. Enzymes also have complementary domains that confer high selectivity for acyl chain length and nucleophilic groups (such as alcohol, amine and water) and enable reactions with high regio- and enantioselectivity. After the formation of the acyl-enzyme intermediate, the enzyme transfers the acyl group to a second nucleophile, e.g., an alcohol, an amine or water, which completes the reaction. This transfer step is essential for the conversion of the substrate into the desired oleochemical products and is strictly regulated by the microenvironment of the active site, which stabilizes the transition state through hydrogen bonding, steric effects and specific hydrophobic interactions that increase both the selectivity and efficiency of catalysis. Following product release, the enzyme is regenerated in its active form, allowing several reaction cycles that increase productivity and reduce enzyme consumption. Lipases also exhibit regiospecificity, recognizing and selectively reacting with functional groups at particular positions, and enantioselectivity, distinguishing between enantiomers of chiral compounds.
This specificity results from the highly organized architecture of their active sites, which impose spatial and electronic constraints to precisely control the positioning and orientation of functional groups during modification. Such selectivity enables the production of optically pure compounds with a high enantiomeric excess compared to conventional chemical methods, which is essential for applications in the fields of pharmaceuticals, fragrances, surfactants and polymers. This section provides an overview of various classes of enzymes involved in fatty acid modification, highlighting their structures, mechanisms and industrial applications.

2.2.1. Hydratases

Fatty acid hydratases, classified under enzyme EC 4.2.1.53, are specialised catalysts that enable the stereo- and regioselective addition of water molecules to carbon–carbon double bonds in fatty acids. This enzymatic activity leads to the formation of chiral hydroxy fatty acids, which are of great biological and industrial importance. These enzymes contain flavin adenine dinucleotide (FAD) as a prosthetic group and mainly target the cis-9 and cis-12 double bonds in monounsaturated fatty acids and polyunsaturated fatty acids (PUFAs) with chain lengths of C14 to C22.
The most common reaction mediated by fatty acid hydratases is the addition of water to the cis-9 double bond in oleic acid to form (S)-10-hydroxystearic acid (denoted as (S)-10-HSA, Figure 2) [53,54].
These hydroxy derivatives are valuable starting materials for the synthesis of γ-lactones, compounds that are frequently used as flavouring agents in the food industry [55,56]. Microbial hydration of oleic acid was first observed in 1962, when Wallen et al. [57] documented that Pseudomonas species converted it into 10-hydroxystearic acid. Since then, numerous other bacteria and fungi from different genera have catalysed similar reactions, expanding our understanding of microbial fatty acid hydration.
The first cloning and detailed characterization of a fatty acid hydratase was performed by Bevers et al. in 2009 [58]. This was followed by research by Volkov and colleagues, who successfully identified and reclassified enzymes from Streptococcus pyogenes, Bifidobacterium breve and Lactobacillus species as specific (cis)-9-hydratases [59,60,61]. The structural elucidation of a (cis)-9-hydratase from Lactobacillus acidophilus was an important milestone, as the enzyme was crystallized in both ligand-bound and unbound states (PDB codes: 4IA6 and 4IA5). This enzyme functions as a homodimer that has an extended substrate binding channel and an N-terminal lid domain in each protomer, which facilitate fatty acid recognition and catalysis. Further progress was made in 2015 with the structural resolution of the oleate hydratase from Elizabethkingia meningoseptica (PDB code: 4UIR), which was studied together with mutagenesis experiments to elucidate the reaction mechanism [62].
These studies revealed that although FAD is not directly involved in the addition of water, it plays a crucial role in maintaining the proper organization of the active site and stabilizing charge formation during the transition state. These findings led to attempts to recombinantly express other (cis)-9-hydratases, including those from Lysinibacillus fusiformis, Stenotrophomonas maltophilia and Macrococcus caseolyticus [63,64,65].
Interestingly, many of these hydratases show high specificity for (cis)-9 double bonds, while their activity is generally limited for (cis)-12 double bonds, such as those in linoleic acid. However, there are some exceptions; for example, a (cis)-12 hydratase from Lactobacillus acidophilus has been identified that predominantly catalyses the formation of the 13-hydroxy derivative of linoleic acid [66,67]. In addition, Hirata and colleagues demonstrated that certain (cis)-12-hydratases from Lactobacillus species can also act on longer-chain PUFAs such as arachidonic acid (C20) and docosahexaenoic acid (DHA, C22) and convert them into the corresponding 15- and 14-hydroxy fatty acids [68]. These enzymes are remarkably efficient. For example, recombinant (cis)-9-hydratases expressed in Escherichia coli have been shown to tolerate high concentrations of oleic acid up to 1 M, allowing for scalable biotransformation [69]. In practical applications, whole-cell biotransformations with E. coli harbouring (cis)-9-hydratases from Stenotrophomonas maltophilia have shown remarkable efficiency, achieving a yield of 91% relative to the starting oleic acid and a final concentration of 46 g/L 10-hydroxystearic acid in the culture medium [65].
In summary, fatty acid hydratases are versatile and potent enzymes with significant potential for industrial biocatalysis that enable the regio- and stereoselective modification of fatty acids to produce valuable hydroxy derivatives. Ongoing structural and functional studies deepen our understanding and open new opportunities for the development of these enzymes for tailored applications in the fields of flavours, pharmaceuticals and bio-based materials.

2.2.2. Hydroxylases

Hydroxylases play a crucial role in lipid biochemistry, especially in the specific modification of fatty acids. Among them, the 12-hydroxylases (enzyme class EC 1.14.13.26) are the best-researched enzymes of this group. These enzymes catalyse the hydroxylation of oleic acid at the 12th carbon position and convert it into ricinoleic acid, an important precursor for the production of various oleochemical derivatives, including lubricants, polyesters and other valuable bio-based products (Figure 3).
The presence of 12-hydroxylases has been demonstrated in plants such as Ricinus communis, commonly known as castor bean, which is known for its high content of ricinoleic acid. Structurally, these enzymes have a di-iron centre that favours their catalytic activity. They utilize molecular oxygen and NADPH to carry out the hydroxylation process, which is highly regio- and stereoselective and ensures precise modification of the fatty acid substrate. Interestingly, 12-hydroxylases are evolutionarily related to 12-desaturases, enzymes that typically introduce a double bond into fatty acids, particularly in the conversion of oleic acid to linoleic acid. This close relationship suggests a common evolutionary origin and a similar structural framework. In fact, scientists have shown that by exchanging only four amino acids in a 12-desaturase for the corresponding amino acids in a fatty acid hydroxylase, desaturase can be converted into an efficient 12-hydroxylase. This remarkable switch highlights the subtle structural differences that determine the specificity and activity of the enzyme [44,70]. Overall, hydroxylases like 12-hydroxylases are important tools in biocatalysis and metabolic engineering, opening up avenues for the production of customized fatty acids for industrial applications. Their ability to catalyse highly specific reactions underlines their importance both for natural lipid metabolism and for biotechnological innovations.

2.2.3. Hydroperoxidases

In addition to hydroxylation processes, hydroperoxidation at specific activated sites in fatty acids, in particular at the α-carbon (α-C-H bonds) and at allyl positions, was also investigated in detail. In particular, α-dioxygenases (α-DOx) are a key class of enzymes involved in these transformations (Figure 4) [71].
The main products of α-DOx activity are α-hydroperoxyacids, which are spontaneously decarboxylated to aldehydes. Given the high reactivity and versatility of aldehydes as starting materials, they serve as valuable intermediates for a variety of chemical syntheses, suggesting that interest in α-DOx enzymes is likely to increase in future research. Despite their potential, the application of α-DOx in synthetic chemistry is still relatively limited, and there are only a handful of studies investigating their utility [72,73]. In contrast, lipoxygenases (LOXs) are better-researched enzymes. They catalyse the hydroperoxidation of the cis,cis-1,4-pentadiene system found in polyunsaturated fatty acids. These enzymes usually facilitate the subsequent cleavage of the hydroperoxide intermediate by hydroperoxide lyases, which leads to the formation of volatile compounds.
This enzymatic system of combining lipoxygenase and hydroperoxide lyase is well established in the flavour and fragrance industries to produce C6 aldehyde odours known as “green leaf volatiles” [74]. Recently, interest in the utilization of these enzymatic transformations to produce bifunctional polymer precursors from unsaturated fatty acids has increased, expanding their potential applications in materials science [75,76]. Another interesting class are the diol synthases. These enzymes catalyse allylic hydroperoxidation reactions similar to lipoxygenases: the generated hydroperoxide intermediates are directly used as oxygen donors in intramolecular hydroxylation reactions. Diol synthases exhibit regioselectivity, e.g., 1,2-, 1,3- or 1,4-positions relative to the hydroperoxidation site [77,78,79] (see Figure 5). Despite their promising catalytic capabilities, these enzymes have not yet attracted much attention in the biocatalysis community. In summary, the unique potential of α-dioxygenases and diol synthases needs to be further explored, even though lipoxygenases and their associated metabolic pathways are better established and widely utilized. Their ability to selectively modify fatty acids though hydroperoxidation opens up avenues for innovative applications in synthesis, materials science and flavour chemistry and promises exciting developments in the future.

2.2.4. Lipoxygenases

Lipoxygenases (LOXs) are a family of oxidative enzymes characterized by their non-haem iron or manganese cofactors. The formation of conjugated hydroperoxy derivatives known as hydroperoxydienoic acids is the primary function of these enzymes when they catalyse the dioxygenation of polyunsaturated fatty acids containing a cis-1,4-pentadiene structure. Their primary function involves catalysing the dioxygenation of polyunsaturated fatty acids (PUFAs) that contain a cis-1,4-pentadiene structure, leading to the formation of a conjugated hydroperoxy derivative known as 15-Hydroxyeicosatetraenoic acid (denoted as (S)-15-HETE, Figure 6) [80].
The ability of LOXs to produce hydroperoxides has opened the door to numerous industrial applications. In the food sector, for instance, LOXs are used to bleach coloured components by oxidizing carotenoids, resulting in a clearer appearance for products like flour. LOX-mediated oxidation also helps with bleaching processes in the paper and textile industries. In addition, LOXs are also used to modify lipids from various raw materials and to produce valuable oleochemicals and aromatic compounds, such as (Z)-3-hexenal, which contributes to fresh, green fragrances.
Despite their promising potential, the widespread industrial utilization of LOXs faces significant challenges. The main obstacles include the low stability of the enzymes under operating conditions, the relatively low catalytic activity and the lack of efficient expression systems that can produce large amounts of active enzymes. These limitations hinder the scalability and economic viability of LOX-based processes, as noted by Heshof et al. (2016) [48]. One notable commercial application of LOXs is in the bleaching of wheat flour. Here, soybean flour enriched with LOX enzymes is used to oxidize carotenoid pigments, thereby enhancing the flour’s whiteness [81].
To improve their functional properties, scientists have turned to protein engineering techniques to modify LOXs. In recent years, efforts have focused on modifying their regio- and stereospecificity, understanding structure–function relationships [82], optimizing metal ion selectivity and improving enzyme stability [83]. Such modifications have identified critical amino acids in the active site of the enzyme, providing insight into how these enzymes can be tailored for specific tasks. A remarkable aspect of LOX biochemistry is its stereoselectivity, which appears to depend on a single conserved amino acid residue in the active site. In LOXs that produce (S)-stereoisomers, this residue is typically an alanine, whereas in (R)-stereoisomers, it is a glycine. Mutational studies have shown that replacing glycine with an alanine can partially change the stereospecificity of the enzyme, effectively converting an (S)-LOX into an (R)-LOX, or vice versa [84,85].
Laboratory-scale studies on OX-mediated hydroperoxide production have yielded promising results. For example, 13-hydroperoxy-(9Z,11E)-octadecadienoic acid was successfully synthetized from linoleic acid using LOXs derived from Gaeumannomyces graminis and Pseudomonas aeruginosa. Both biotransformations showed high yields (88% and 75%) with respect to the starting linoleic acid and good selectivity (74% and 61%). It is noteworthy that the reaction with Pseudomonas aeruginosa LOX required a much lower substrate concentration (10 g/L) and longer incubation times (48 hours) compared to the G. graminis enzyme [84,85].
Scale-up experiments with G. graminis LOX at industrially relevant linoleic acid concentrations (100–300 g/L) showed a decrease in yield; however, a volumetric productivity of about 3.6 g per litre per hour was still achieved [86]. The combination of LOXs with hydroperoxide lyases (HPLs) represents an interesting route to produce green fragrances, which are highly valued in the fragrance and flavour industries. In initial experiments, crude homogenates from soybean (as LOX source) and guava (as HPL source) were used to produce C6 compounds like (Z)-3-hexenal and (Z)-3-hexenol, which contribute significantly to the characteristic green aroma. HPLs are classified as CYP74 enzymes and carry out the homolytic isomerization of fatty acid hydroperoxides to unstable hemi-acetals, which are then broken down into C6-aldehydes and C12-oxo acids (see Figure 7).
The C6-aldehyde can be reduced to its corresponding alcohol, although isomerization to (Z)-2-aldehyde and (Z)-2-alcohol has also been observed. The functional expression of the guava HPL gene in Escherichia coli has led to significant progress in this field. Subsequently, directed evolution techniques, including DNA shuffling and random mutagenesis, were applied to improve the performance of the enzyme. The result of these efforts was a variant with fifteen times higher productivity, primarily due to improved expression and increased thermostability of the enzyme [87]. More recently, an enzyme cascade process was developed in which the engineered 13-HPL enzyme, used as a crude cell lysate, was combined with a commercial ketoreductase. Using this approach, (Z)-3-hexenol was successfully produced with an isomeric purity of 99% and a high concentration of approximately 8 g/L, demonstrating the potential for the scalability of the process [88]. In summary, LOXs are versatile enzymes with significant industrial potential, especially when combined with other biocatalysts. Ongoing research into enzyme engineering, process optimization and scale-up strategies continues to push the boundaries of their applications from food processing to fragrance manufacturing despite existing challenges related to stability and expression efficiency.

2.2.5. Decarboxylases

The oxidative decarboxylation of fatty acids to produce terminal alkenes has emerged as a fascinating area of enzymatic bioconversion that primarily involves a series of haem-dependent monooxygenases. This conversion was initially documented with a P450 monooxygenase derived from Jeotgalicoccus species, known as OleTJe (Figure 8) [89].
Since then, similar activity has been observed as a secondary or side reaction in several P450 enzymes, including P450Bsβ [90] and a range of additional P450 monooxygenases [91,92,93]. In addition to haem-based systems, non-haem iron-containing decarboxylases, like UndB from Pseudomonas species, have also been identified that can catalyse decarboxylation processes [94,95,96,97]. An innovative approach to synthetize terminal alkenes is to combine OleT with an alditol oxidase enzyme, as shown in Figure 9 [98,99]. This strategy aims to create a comprehensive biocatalytic pathway that valorises natural triglycerides and converts them directly into valuable alkenes. Such holistic approaches are promising for sustainable chemical production utilizing biomass components.
More recently, a new class of enzymes, the so-called fatty acid photodecarboxylases (FAPs), has attracted considerable attention. Sorigué et al. [100,101] characterized FAPs, which represent a breakthrough in light-driven decarboxylation. The enzyme from Chlorella variabilis (CvFAP) has become the focus of research, especially considering its potential for biofuel production. Its unique mechanism is based on the photoactivation of a flavin prosthetic group, which enables the direct conversion of fatty acids into alkanes when exposed to light. This process offers several advantages over the conventional biodiesel process based on fatty acid methyl esters (FAMEs), including the reaction being simple and irreversible and the fact that products with higher caloric content are produced.
However, despite the promising prospects of FAPs in the production of next-generation biofuels, their practical application faces significant obstacles. Chief among these is the poor photostability of the enzyme, which limits the operational longevity and, consequently, the economic viability of large-scale processes [102]. A deeper understanding of the mechanisms underlying FAP inactivation and advances in enzyme engineering to improve stability are required to achieve sustained enzyme activity over extended periods of time. To overcome these challenges, several innovative strategies have been pursued to integrate FAPs into synthetic pathways aimed at producing higher-value compounds. Researchers have engineered CvFAP variants to improve selectivity towards different substrates, especially focusing on light-driven kinetic resolution reactions. Li and colleagues [103] have successfully modified CvFAP to achieve better discrimination between α-substituted carboxylic acids and to distinguish between cis- and trans-unsaturated fatty acids.
Such modifications open up exciting opportunities to tailor enzyme activity to specific bioconversion targets and extend the potential applications of FAPs beyond biofuel production to speciality chemicals and fine intermediates. Further research on their stability, specificity and mechanism will be crucial to fully realise their industrial potential.

2.2.6. P450 Monooxygenases

Cytochrome P450 enzymes, commonly known as CYPs, constitute a large family characterized by their prosthetic haem group. These enzymes are capable of catalysing specific oxidation reactions, both regio- and stereoselectively, across a broad spectrum of substrates. A characteristic feature of CYPs is their dependence on a redox partner system that facilitates electron transfer from NAD(P)H to the haem centre of the enzyme, enabling catalytic activity. These enzymes are essential for the metabolism of xenobiotics, foreign compounds such as drugs and environmental chemicals; they are also involved in the biosynthesis of several natural products, including steroids, fatty acids and secondary metabolites [104,105]. Despite their remarkable versatility and potential for biotechnological applications, the industrial use of CYPs remains limited.
Challenges such as those involving enzyme stability and the need for complex electron transfer systems hinder their widespread use. In addition, many CYPs are either membrane-bound or membrane-associated, which complicates their extraction, purification and utilization as isolated enzymes. Their dependence on NADPH in stoichiometric amounts further complicates their practical application.
A typical reaction mediated by P450 monooxygenases is the hydroxylation of fatty acids, whereby ω-hydroxy fatty acids are formed directly from free fatty acids (Figure 10). These hydroxylated products can subsequently be oxidized to dicarboxylic acids, as has been shown in various whole-cell biotransformations with Candida species [106]. Such compounds are multifunctional and have significant industrial value, as they serve as precursors for polymer production and fragrances [107]. Among the different CYPs, the enzyme CYP102A1 from Bacillus megaterium, also known as P450-BM3, stands out as the most practical and well-studied example.
This enzyme is a natural fusion protein that integrates both the haem domain and the reductase domain into a single polypeptide chain, giving it exceptional catalytic efficiency. P450-BM3 has the highest turnover frequency among the P450 monooxygenases and is specialised in the subterminal hydroxylation of long-chain fatty acids (from C12 to C22) with high regio- and stereoselectivity [108].
During the development of the protein, extensive efforts were made to modify P450-BM3 to improve its activity, selectivity, stability and substrate range. These modifications include altering its regio-, stereo- and chemoselectivity and reducing its dependence on cofactors. Studies have investigated how specific amino acid substitutions affect the performance of the enzyme, with particular attention paid to residue F87, which has been identified as a key factor for substrate specificity [109,110].
Recent advances include iterative cycles of targeted or random mutagenesis to generate P450-BM3 variants capable of producing terminal hydroxy fatty acids, which are valuable intermediates in chemical synthesis. Some of the most successful variants contained up to 11 amino acid substitutions, three of which, F87, I263 and A328, were close to the substrate binding pocket and had a significant impact on the activity of the enzyme [111].
In addition to P450-BM3, several other self-sufficient CYPs from Bacillus and Streptomyces species have been characterized and used for biotransformations. For example, P450 monooxygenases from Sorangium cellulosum and Fusarium oxysporum were overexpressed in microbial hosts such as Saccharomyces cerevisiae and Escherichia coli and enabled the biosynthesis of medium-chain ω-hydroxy acids, highlighting the versatility and potential of these enzymes for sustainable biocatalysis [112,113].

3. Environmental Benefits of Biocatalysts in Oleochemistry Compared to Chemical Catalysts

In the synthesis of oleochemical products, catalysts play an essential and multifaceted role, significantly influencing the overall efficiency, selectivity and environmental footprint of the process. Traditionally, both homogeneous and heterogeneous chemical catalysts are used to facilitate these reactions.
Homogeneous catalysts, which are soluble in the reaction medium, offer notable advantages such as fast reaction rates, simple operating procedures and easy mixing, resulting in efficient process kinetics. Common examples of homogeneous catalysts are inorganic bases such as sodium hydroxide (NaOH), potassium hydroxide (KOH), calcium hydroxide (Ca(OH)2), potassium carbonate (K2CO3) and sodium carbonate (Na2CO3), as well as inorganic acids such as hydrochloric acid (HCl), sulphuric acid (H2SO4), phosphoric acid (H3PO4) and p-toluenesulphonic acid [114,115,116,117].
Alkaline homogeneous catalysts are particularly valued for their high reaction speed, low reagent consumption and corrosion resistance, which can increase process efficiency. Acid catalysts, on the other hand, often allow reactions under milder temperature and pressure conditions, which can reduce energy requirements and potentially improve product selectivity. Despite these advantages, homogeneous catalysts have notable limitations. They are difficult to recover and dispose of, resulting in additional purification steps to recover the final products and environmental problems due to the release of toxic waste that needs to be disposed of [118,119]. Some catalysts are corrosive, which can damage equipment, and side reactions can produce unwanted by-products.
Heterogeneous alkaline catalysts such as calcium oxide (CaO) or oxides of alkali and alkaline earth metals require the use of refined oils, as they are sensitive to the presence of free fatty acids, which can lead to the formation of soaps, resulting in their deactivation [120]. The use of heterogeneous acid catalysts has been extensively studied, as they can be easily separated at the end of the process. Examples of effective catalysts are strong acid exchange resins (Amberlyst 15, 20BD, 35, Lewatit K2621, K2620, Tulsion T-42, Dowex M 31) [121,122,123], magnetically separable iron oxide on a sulphonated graphene oxide [124], zeolites [125] and metal oxides [126,127,128,129]. Although heterogeneous catalysts are easier to recover, they can be deactivated by fouling or poisoning, especially when processing raw materials with high FFA contents. In addition, these reactions often require high temperatures and long periods of time, which increases energy consumption and costs.
Lipases have been shown to accelerate and catalyse several chemical reactions that use lipid substrates [130]. Each lipase has an active site consisting of different amino acids so that it can recognize several lipid molecules. For example, the immobilized lipase from Thermomyces lanuginosus (Lipozym TL IM) has an active site containing amino acids such as lysine, aspartic acid and glutamic acid [131] and can catalyse a variety of reactions, including the esterification of fatty acids, hydrolysis, alcoholysis, transesterification and the acidolysis of oils and fats. Lipozym TL IM exhibits regioselectivity by catalysing the ethanolysis reaction of trilaurin and trimyristin, leading to the formation of β-monoglycerides, in particular 2-monolaurin [132] and 2-monomyristin [133].
Similarly, the α-monoglyceride compounds, α-glycerol monolaurate (1-monolaurin) and α-glycerol monocaprate (1-monocapryn), were prepared by transesterification of methyl laurate or methyl caprate with 1,2-acetonide glycerol using Lipozym TL IM (Figure 11). These compounds were then deprotected with Amberlyst-15 [134,135].
Lipids from food waste can be converted into biodiesel using methanol (molar ratio 1:5) and Novozym® 435 as a catalyst. The conversion rate of lipids to biodiesel was 90% at a reaction time of 24 h and 40 °C [136]. Other food waste can be enzymatically hydrolysed to produce hydrosilate, which is subsequently cultured into microalgae biomass. The transesterification of microalgae lipids with methanol and lipase can also be used to produce biodiesel [137].
Lipase enzymes offer advantages in terms of green chemistry and sustainability: they are safe and easy to handle, non-toxic, ecologically harmless, energy efficient, renewable and biodegradable [138]. As a result, the amount of waste is reduced to an absolute minimum. Certain types of lipases eliminate the need for toxic solvents and can even occur in aqueous solution [139].
Lipases have found significant strategic and potential application in lipid processes, especially in biogenic and eutectic solvents [140]. It is more productive and environmentally friendly to use lipases in solvent-free systems for lipid conversion processes [141]. These parameters suggest that reaction processes including lipase are more environmentally friendly, less harmful, safer and guarantee long-term viability.
However, it is clear that most oleochemical synthesis processes involving lipases take place in organic solvents, as the lipid substrates are more easily dissolved in them. The solubility of the reagent affects the stability and activity of the enzyme, which in turn controls the reaction rates.
In addition, immobilized lipase is a heterogeneous catalyst that offers several advantages both in terms of industrial production and environmental protection. Easy separation from the target product, reusability and decomposition in the environment upon loss of activity contribute to its favourable properties. The use of immobilized lipase significantly reduces the environmental impact of the catalyst.
Finally, oleochemicals are traditionally synthesized using chemical catalysts. Although these methods are effective, they are often energy-intensive, harmful for the environment and generate large amounts of waste. In contrast, enzymatic catalysis, particularly using lipases, has proven to be a sustainable alternative that is compatible with the principle of green chemistry. This section addresses the environmental benefits of enzymatic processes, with a focus on reducing greenhouse gas emissions, minimizing toxic by-products, energy efficiency and water conservation.

3.1. Reduction of Carbon Dioxide Emissions

One of the most important environmental benefits of enzymatic catalysis is its ability to significantly reduce CO2 emissions. Conventional chemical catalysis often requires high temperatures (160–220 °C), which leads to significant energy consumption and consequently increased greenhouse gas emissions. In contrast, enzymatic catalysis operates efficiently under milder conditions (35–45 °C), which significantly reduces energy consumption and the associated emissions.
Mustafa and Niikura [142] pointed out that the enzymatic production of isopropyl palmitate leads to a reduction in CO2 emissions by more than 50% compared to chemical processes. This reduction results from lower operating temperatures and the absence of CO2-generating side reactions.
Similarly, Pasha et al. [143] analysed the life cycle emissions of enzymatic versus chemical biodiesel production and found that enzymatic processes can reduce CO2 emissions from soybean oil and used cooking oil by up to 78.9% and 63.1%, respectively. These results emphasize the potential of biocatalysis to reduce the environmental impact of industrial oleochemical production.

3.2. Minimization of Toxic By-Products

Chemical catalysis often involves the use of hazardous reagents such as strong acids or alkaline compounds, which can lead to the formation of significant quantities of toxic by-products. These substances pose a serious environmental risk, including soil and water contamination, and require extensive post-reaction treatment, further increasing the environmental impact of the process. In contrast, lipase-catalysed reactions are inherently cleaner, highly specific and generate minimal waste.
Hosney and Mustafa [144] demonstrated that the enzymatic production of 2-ethylhexyl esters generated significantly less hazardous waste compared to chemical processes. This fact not only simplifies waste disposal but also reduces the ecological footprint of the process. In addition, enzymatic reactions generally produce more environmentally friendly by-products that are biodegradable or reusable. In the production of biodiesel, for example, the by-product glycerin, obtained by enzymatic processes, is of higher purity and can be more easily converted into value-added products such as 1,3-propanediol or acrolein.

3.3. Improved Biodegradability and Compatibility with Green Solvents

Chemical catalysis usually uses volatile organic compounds as reactants, which contributes to air pollution and pose health risks. However, enzymatic catalysis is highly compatible with environmentally friendly solvents such as ionic liquids, supercritical CO2 and deep eutectic solvents.
Mustafa and Niikura [142] investigated the use of immobilized Candida antarctica lipase in green solvent systems for esterification reactions. The study showed high reaction efficiency and minimal environmental impact, emphasizing the synergy between biocatalysis and using tert-butanol as a solvent. In addition, the enzymatic products are often more biodegradable, which reduces the long-term environmental impact compared to their chemically synthesized counterparts.

4. Market and Economic Considerations

4.1. General Arguments

A direct comparison of the costs of lipase with those of conventional chemical catalysts should be avoided at all costs. Instead, the focus should be on assessing the overall economics of the process. Firstly, enzymatic processes often require less capital investment than chemical technologies due to their simpler processes and lower equipment requirements. Secondly, enzymatic processes usually consume less energy, which makes them economically attractive [145]. Furthermore, the environmental benefits of enzymatic processes should not be overlooked. Given the growing consumer awareness and interest in environmentally friendly products, the clean nature of enzymatically produced goods has become a valuable marketing advantage.
A comprehensive economic evaluation includes several factors that influence the cost of biocatalysts. These include the “lipase production cost” (by submerged or solid-state cultivation), the “lipase reaction cost” (considering the substrate and the final product) and the “type of support” used in the immobilized lipase process, which directly affects the immobilization method chosen.
Mustafa et al. [146] demonstrated that the manufacturing costs for an enzymatic process for the production of monostearin were 10% higher than those for the chemical counterpart. However, when considering the capital cost of the plant, the process showed a positive return on investment and a net present value, emphasizing its competitiveness with chemical processes. Economic evaluations should focus on the most influential factors, even if not all favourable conditions are present. For example, Andrade et al. [147] identified enzyme reusability—at least 300 cycles—as a critical threshold for economic feasibility.
Alternatively, efforts can focus on reducing the cost for biocatalyst immobilization, as Budžaki et al. [148] found that support materials can account for almost 90% of the total price of a biocatalyst. Optimizing this aspect can significantly increase the overall efficiency of the process. This review highlights the key recommendations to promote investment in enzymatic technologies for oleochemical and biodiesel production. Enzyme suppliers should not only provide biocatalysts but also offer comprehensive solutions to help customers sustain their business. In addition, engineering companies supplying production equipment should work closely with enzyme producers to provide robust and reliable enzymatic technologies for esterification or hydrolysis systems.
Companies introducing enzymatic processes need to ensure that their employees are well-trained to optimize operations and maximize profitability. Ultimately, the affordability and simplicity of enzymatic processes are the main reasons for their growing appeal to manufacturers.
Considering these approaches and the competitive overall costs frequently cited by researchers, the enzymatic process can be a technically and economically viable competitor to conventional chemical processes, which underpins the reasons why the use of these biocatalysts in the synthesis of oleochemicals has been intensively researched over the last two decades, as shown by the bibliometric data in Section 4.2.

4.2. Bibliometric Evaluation

Bibliometric analyses are an interesting way of recognizing trends, mapping knowledge, investigating collaborations and following the scientific development of a particular topic. Accordingly, a bibliometric study of scientific documents dealing with the application of biocatalysts in the field of oleochemistry was carried out on the scientific basis of the Scopus® platform.
In order to select the largest possible number of works, the following string was defined with Boolean operators: (biocatalys* OR bio-catalys* OR enzym*) AND (transesterification OR trans-esterification OR esterification OR hydroesterification OR hydro-esterification OR hydrolysis-esterification OR methanolysis OR ethanolysis OR ethoxylation OR neutralization OR amination OR amidation) AND (“vegetable oil*”) OR (fat OR “animal fat*” OR “animal oil*” OR tallow OR suet OR “poultry fat*” OR “abdominal fat*”) OR (triglyceride* OR triacylglicerol*) OR (“medium chain triglyceride*” OR “medium chain triacylglycerol*”) OR (diglyceride* OR diacylglycerol*) OR (monoglyceride* OR monoacylglycerol*) OR (“fatty acid*” OR lipid*) OR (“fatty alcohol*” OR “fatty acyl alcohol*” OR “aliphatic alcohol*” OR “alkyl alcohol*”) OR (biodiesel OR “fatty ester*” OR “fatty acid ester*” OR “fatty acid alkyl ester*” OR FAME* OR “methyl ester*” OR “fatty acid methyl ester*” OR “fatty acid ethyl ester*”) OR (glycerol OR glycerin* OR propanetriol* OR triacetin*) OR (“sugar ester*” OR “fatty acid sugar ester*” OR “sugar fatty acid ester*” OR “carbohydrate fatty ester*” OR “carbohydrate fatty acid ester*” OR “glycolipid*”) OR (“lubricant ester*” OR biolubricant*) OR (“wax ester*” OR “ester wax*” OR “acyl wax ester*” OR “cosmetic wax*”) OR (emulsifier* OR surfactant* OR “emulsion stabilizer*”) OR (“2-ethyl hexyl oleate” OR “2-ethyl hexyl stearate” OR “2-ethyl hexyl ester*”) OR (“isopropyl ester*” OR “isopropyl myristate” OR “isopropyl palmitate”) OR (“sorbitan ester*” OR “sorbitan stearate” OR “sorbitan oleate”) OR (“glyceryl monostearate”) OR (soap* OR detergent*).
The search yielded a total of 7160 documents published between January 2004 and March 2024. The data was then exported to VOSviewer® software (version 1.6.20) to enable a textual and geographical analysis of the topic.
Figure 12a,b show the network visualization (periodic and temporal, respectively) derived from the frequency of keywords used by the authors in the documents, considered and characterized by a representation of the co-occurrence of frequently used terms in the dataset.
To improve the understanding and the context of the bibliographic map, the minimum number of occurrences of each keyword was set at 15, i.e., each keyword had to occur at least 15 times within the observation period to be included in the analysis. A total of 223 keywords were considered and presented in 11 clusters, with the largest cluster comprising 159 keywords. The most important keywords were “biodiesel” (1258 occurrences), “lipase” (1133), “transesterification” (904), “esterification” (639), “immobilization” (263), “kinetics” (248), “biocatalysis” (186), “immobilized lipase” (164), “lipases” (155) and “response surface methodology” (145 occurrences).
In addition, a geographical mapping of the VOSViewer® global bibliographic linkage network was created based on the same string/dataset, as shown in Figure 13a,b. The maximum number of countries per document considered in the analysis was 25, with each nation having at least 10 documents, resulting in a total number of 60 countries in five clusters. China (1278 documents and 35,128 citations), the United States (1118 documents and 57,799 citations), India (674 documents and 24,839 citations), Brazil (535 documents and 14,947 citations), Spain (436 documents and 14,265 citations), Germany (363 documents and 13,509 citations), Japan (350 documents and 11,655 citations), Canada (311 documents and 13,338 citations), France (283 documents and 12,615 citations), Malaysia (279 documents and 11,214 citations) and South Korea (257 documents and 7814 citations) are the most important countries conducting research on this topic.
Spatial bibliographic analysis provides information about the performance of a study network, which may include different collaborators around the world (Figure 13). Clusters representing the frequency of the co-occurrence of articles provide a direct indication of the relationships between these groups. The degree of co-occurrence of each country is indicated by coloured groups, where countries with the same colour form a cluster.
The information presented in Figure 13a indicates a strong research alliance in the field of enzymatic synthesis of oleochemicals involving Eastern countries, especially countries known for their scientific development, such as China, India and South Korea. As you can see, there is also a solid partnership between North American countries (United States and Canada) and European countries, particularly Germany, the United Kingdom, Italy and Sweden. Brazil and Japan also stand out and are at the top of the other two influential groups related to the topic under study, with both countries still maintaining close relations with the groups led by China and the United States, respectively.
Scientific collaboration between the researchers from different countries is common in a globalized society, of which many examples can be found. Interestingly, Razzaghi et al [149], in collaboration with authors from Iran, Italy, Canada, India, China and Mexico, investigated the application of nano-biocatalysts in different industrial processes and evaluated the immobilization of enzymes in different nano-compounds with the aim of improving their stability in the reaction system for subsequent reuse. Similarly, Melo et al. [150] discuss, in an article co-authored by Brazilian and Polish scientists, the research progress, trends and updates related to magnetic biocatalysts. Pereira et al. [151], in a partnership between Brazilian and French researchers, comprehensively describe the importance of lipases in the synthesis of phytosterol esters.
Figure 14 shows the evolution of published works on the topics of “biocatalysis” and “oleochemical synthesis” (Scopus® data platform) using the same keywords used to create the bibliographic maps presented previously. Using this data, it is possible to establish a certain consistency in the works published since 2011. However, it is important to emphasize that the vast majority of these works include aspects related to the evaluation of the technical feasibility of the subject. This situation is confirmed by the inclusion of terms relating to economic assessment in the search. By adding the string (econom* OR “economic review*” OR “economic examination*” OR “economic discussion*” OR “economic assessment*” OR "economic stud*” OR “economic evaluation*” OR “economic appraisal*” OR “economic* consideration*” OR “operational cost*”) to the other keywords, only 339 documents were found in the period from 2004 to March 2024, that is less than 5% of the total number of papers. Therefore, this critical debate emphasizes that further efforts need to be made to enable the commercialization of the enzymatic process, focusing on a technoeconomic discussion.

5. Limitations, Challenges and Final Considerations

A major limitation in the production of oleochemicals is probably the availability of a wide range of natural plant oil substances, as most bio-based oleochemicals are derived from these natural renewable resources.
Knowledge regarding the morphological and chemical biodiversity of each region, combined with the conversion of these resources, is a competitive advantage for the development of new products and technologies with high added value, especially in the field of oleochemistry. As biodiversity is reflected in the possibilities of oleochemistry, Brazil seems to be a promising region due to its diversity of exotic plants rich in oils, such as “pinhão-bravo” (Jatropha mollissima Pohl Baill), “pinhão-manso” (Jatropha curcas), “pequi” (Caryocar brasiliense), “oiticica” (Licania rígida Benth), “baru” (Dipteryx alata) [152,153] and “macaúba” (Acrocomia aculeata) [154]. In various biotechnological fields, the use of microorganisms is an alternative approach to conserve natural resources and increase the production of organic products. The main obstacles to the expansion of biocatalytic biodiesel production using enzymes are primarily of an economic and infrastructural nature. Firstly, the high cost of enzymes, especially specialised lipases, has a significant impact on the economic viability of the overall process. The production and purification of these enzymes on an industrial scale is still expensive, which makes them less competitive than conventional chemical processes. Secondly, the established nature of chemical catalysis, based on mature infrastructure and proven cost efficiency, is a significant barrier to adoption. Industry players are often reluctant to switch from established chemical processes to enzymatic processes in the absence of clear, proven economic benefits on a large scale. To overcome these barriers, enzyme costs must be reduced, enzyme stability and reusability improved and the economic viability of enzymatic biodiesel production demonstrated in real industrial settings. It is clear that metabolic engineering optimizes the production of oleochemical derivatives. Microorganisms such as Escherichia coli, Saccharomyces cerevisiae, Yarrowia lipolytica, Rhodococcus opacus and Rhodosporidium toruloides can produce oleochemical products such as triacylglycerols (TAGs) and FFAs at amounts approaching 100 g L−1 with a theoretical yield of more than 90%, as discussed by Yan and Pfleger [3]. Governmental policies on enzymatic biodiesel and oleochemicals focus mainly on promoting the introduction of enzymatic processes for their production with respect to the use of chemical catalysts. The overall objective is to support the growth of the enzyme sector while achieving environmental and economic targets [155,156]. Accordingly, some key components of this policy should include the following:
i. Incentives and subsidies. To encourage the use of enzymatic processes, governments can provide financial support in the form of tax credits, grants or low-interest loans to producers. These measures can help to offset the higher initial costs associated with enzymatic production [157];
ii. Regulatory framework. Governments should establish standards and regulations to ensure the safety, quality and environmental sustainability of enzymatic biodiesel/oleochemicals. These frameworks often include guidelines for the use of the enzyme, product standards and sustainability practices. Compliance with these regulations is usually a prerequisite for obtaining government incentives or market entry;
iii. Renewable energy requirements. Many countries require that a certain percentage of transport fuels come from renewable sources, especially biodiesel. Producers of enzymatic biodiesel can contribute to meeting these requirements and may be eligible for additional incentives under renewable energy initiatives [158];
iv. Environmental initiatives. Enzymatic biodiesel production is being promoted as a more environmentally friendly alternative as it requires less energy and produces minimal chemical waste. Governments can prioritize the introduction of biodiesel in public transport and official vehicle fleets to reduce greenhouse gas emissions;
v. Education and public relations. Public education campaigns and outreach programs can be funded to raise awareness of the benefits of enzymatic oleochemicals. These efforts will target producers, farmers and the general public to promote broad acceptance of this technology;
vi. Trade and export opportunities. Governments can promote the international trade in enzymatic biodiesel/oleochemicals and related technologies to support economic growth and build global renewable energy partnerships;
vii. Research cooperation. Partnerships between academia, research institutions and industry stakeholders can be supported to drive innovation and accelerate the commercialization of enzymatic biodiesel technologies.
Overall, these policies aim to create a favourable environment for enzymatic oleochemistry and biodiesel production that contributes to greenhouse gas emission reduction, energy security and long-term sustainability. This can be achieved through a combination of financial incentives, regulatory measures, educational initiatives and international cooperation, all aimed at promoting the growth of the enzymatic oleochemical industry.

Author Contributions

Conceptualization, A.M. and L.d.B.; methodology, validation, investigation and writing—original draft preparation J.H.C.W., E.P.C., E.A.M., F.O.N., A.V.C.-P., C.U.M. and T.M.M.A.; writing—review and editing, A.M. and L.d.B.; supervision, A.M. and L.d.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Dataset available upon request from the authors.

Acknowledgments

J.H.C.Wancura is grateful for the scholarship of the Human Resources Program of the Brazilian Agency of Petroleum, Natural Gas, and Biofuels—PRH/ANP through the Human Resources Training Program for Petroleum and Biofuels Processing (PRH 52.1).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Marella, E.R.; Holkenbrink, C.; Siewers, V.; Borodina, I. Engineering Microbial Fatty Acid Metabolism for Biofuels and Biochemicals. Curr. Opin. Biotechnol. 2018, 50, 39–46. [Google Scholar] [CrossRef] [PubMed]
  2. Zhang, Y.; Nielsen, J.; Liu, Z. Yeast Based Biorefineries for Oleochemical Production. Curr. Opin. Biotechnol. 2021, 67, 26–34. [Google Scholar] [CrossRef]
  3. Yan, Q.; Pfleger, B.F. Revisiting Metabolic Engineering Strategies for Microbial Synthesis of Oleochemicals. Metab. Eng. 2020, 58, 35–46. [Google Scholar] [CrossRef] [PubMed]
  4. Su, J.; van Dyk, S.; Saddler, J. Repurposing Oil Refineries to “Stand-Alone Units” That Refine Lipids/Oleochemicals to Produce Low-Carbon Intensive, Drop-in Biofuels. J. Clean. Prod. 2022, 376, 134335. [Google Scholar] [CrossRef]
  5. Wancura, J.H.C.; Brondani, M.; dos Santos, M.S.N.; Oro, C.E.D.; Wancura, G.C.; Tres, M.V.; Oliveira, J.V. Demystifying the Enzymatic Biodiesel: How Lipases Are Contributing to Its Technological Advances. Renew. Energy 2023, 216, 119085. [Google Scholar] [CrossRef]
  6. Awogbemi, O.; Desai, D.A. Application of computational technologies for transesterification of waste cooking oil into biodiesel. Biomass Bioenergy 2025, 194, 107620. [Google Scholar] [CrossRef]
  7. Knothe, G.; Razon, L.F. Biodiesel Fuels. Prog. Energy Combust. Sci. 2017, 58, 36–59. [Google Scholar] [CrossRef]
  8. Wancura, J.H.C.; Tres, M.V.; Jahn, S.L.; Oliveira, J.V. Lipases in Liquid Formulation for Biodiesel Production: Current Status and Challenges. Biotechnol. Appl. Biochem. 2020, 67, 648–667. [Google Scholar] [CrossRef]
  9. Steinke, G.; Cavali, M.; Wancura, J.H.C.; Magro, J.D.; Priamo, W.L.; Mibielli, G.M.; Bender, J.P.; de Oliveira, J.V. Lipase and Phospholipase Combination for Biodiesel Production from Crude Soybean Oil. BioEnergy Res. 2022, 15, 1555–1567. [Google Scholar] [CrossRef]
  10. Santolin, L.; Fiametti, K.G.; da Silva Lobo, V.; Wancura, J.H.C.; Oliveira, J.V. Enzymatic Synthesis of Eugenyl Acetate from Essential Oil of Clove Using Lipases in Liquid Formulation as Biocatalyst. Appl. Biochem. Biotechnol. 2021, 193, 3512–3527. [Google Scholar] [CrossRef]
  11. Finco, G.F.; Fiametti, K.G.; Lobo, V.d.S.; da Silva, E.A.; Palú, F.; Wancura, J.H.C.; Rodrigues, M.L.F.; Valério, A.; de Oliveira, J.V. Kinetic Study of Liquid Lipase-catalyzed Glycerolysis of Olive Oil Using Lipozyme. J. Am. Oil Chem. Soc. 2022, 99, 559–568. [Google Scholar] [CrossRef]
  12. Hosney, H.; Al-Sakkari, E.G.; Mustafa, A.; Ashour, I.; Mustafa, I.; El-Shibiny, A. A Cleaner Enzymatic Approach for Producing Non-Phthalate Plasticiser to Replace Toxic-Based Phthalates. Clean Technol. Environ. Policy 2020, 22, 73–89. [Google Scholar] [CrossRef]
  13. Wancura, J.H.C.; Rosset, D.V.; Tres, M.V.; Oliveira, J.V.; Mazutti, M.A.; Jahn, S.L. Production of Biodiesel Catalyzed by Lipase from Thermomyces lanuginosus in Its Soluble Form. Can. J. Chem. Eng. 2018, 96, 2361–2368. [Google Scholar] [CrossRef]
  14. Wancura, J.H.C.; Fantinel, A.L.; Ugalde, G.A.; Donato, F.F.; de Oliveira, J.V.; Tres, M.V.; Jahn, S.L. Semi-Continuous Production of Biodiesel on Pilot Scale via Enzymatic Hydroesterification of Waste Material: Process and Economics Considerations. J. Clean. Prod. 2021, 285, 124838. [Google Scholar] [CrossRef]
  15. Mustafa, A.; Faisal, S.; Ahmed, I.A.; Munir, M.; Cipolatti, E.P.; Manoel, E.A.; Pastore, C.; di Bitonto, L.; Hanelt, D.; Nitbani, F.O.; et al. Has the Time Finally Come for Green Oleochemicals and Biodiesel Production Using Large-Scale Enzyme Technologies? Current Status and New Developments. Biotechnol. Adv. 2023, 69, 108275. [Google Scholar] [CrossRef] [PubMed]
  16. Behr, A.; Seidensticker, T. The Basics of Oleochemistry—Basic Oleochemicals. In Chemistry of Renewables; Springer: Berlin/Heidelberg, Germany, 2020; pp. 37–60. [Google Scholar] [CrossRef]
  17. Gharby, S. Refining Vegetable Oils: Chemical and Physical Refining. Sci. World J. 2022, 2022, 6627013. [Google Scholar] [CrossRef] [PubMed]
  18. O’Brien, R.D. FATS and OILS: Formulating and Processing for Applications, 2nd ed.; CRC Press: Boca Raton, FL, USA, 2004; ISBN 9780203483664. [Google Scholar]
  19. Richardson, A.S.; Conley, C. Process of Decomposing Fats or Oils into Fatty Acids and Glycerin. U.S. Patent 601603A, 29 March 1898. [Google Scholar]
  20. Lascaray, L. Industrial Fat Splitting. J. Am. Oil Chem. Soc. 1952, 29, 362–366. [Google Scholar] [CrossRef]
  21. Papadaki, A.; Cipolatti, E.P.; Aguieiras, E.C.G.; Cerqueira Pinto, M.C.; Kopsahelis, N.; Freire, D.M.G.; Mandala, I.; Koutinas, A.A. Development of Microbial Oil Wax-Based Oleogel with Potential Application in Food Formulations. Food Bioprocess Technol. 2019, 12, 899–909. [Google Scholar] [CrossRef]
  22. Schmid, U.; Bornscheuer, U.T.; Soumanou, M.M.; McNeill, G.P.; Schmid, R.D. Highly Selective Synthesis of 1,3-Oleoyl-2-Palmitoylglycerol by Lipase Catalysis. Biotechnol. Bioeng. 1999, 64, 678–684. [Google Scholar] [CrossRef] [PubMed]
  23. Halldorsson, A.; Haraldsson, G.G. Fatty Acid Selectivity of Microbial Lipase and Lipolytic Enzymes from Salmonid Fish Intestines toward Astaxanthin Diesters. J. Am. Oil Chem. Soc. 2004, 81, 347–353. [Google Scholar] [CrossRef]
  24. Röttig, A.; Wenning, L.; Bröker, D.; Steinbüchel, A. Fatty Acid Alkyl Esters: Perspectives for Production of Alternative Biofuels. Appl. Microbiol. Biotechnol. 2010, 85, 1713–1733. [Google Scholar] [CrossRef] [PubMed]
  25. Adamczak, M.; Bornscheuer, U.T.; Bednarski, W. The Application of Biotechnological Methods for the Synthesis of Biodiesel. Eur. J. Lipid Sci. Technol. 2009, 111, 800–813. [Google Scholar] [CrossRef]
  26. Jackisch, B.O.; Simmler-Hübenthal, H.; Zschau, W.; Bornscheuer, U.; Durban, M.; Riemer, C. Use of a Thermostable Phospholipase in the Degumming of an Oil or Fat, and a Method for Obtaining a Thermostable Phospholipase. European Patent Application No. EP1788080A, 23 May 2007. pp. 1–56. [Google Scholar]
  27. Sein, A.; Hitchman, T.; Dayton, L.G. Enzymes in vegetable oil degumming processes. In Industrial Enzyme Applications; Wiley-VHC: Weinheim, Germany, 2019; Chapter 4.2; pp. 323–350. [Google Scholar] [CrossRef]
  28. Daniel, H.; Otto, R.T.; Binder, M.; Syldatk, C. Production of Sophorolipids from Whey-Development of a Two-Stage Process with Cryptococcus curvatus ATCC 20509 and Candida bombicola ATCC 22214. Appl. Microbiol. Biotechnol. 1999, 51, 40–45. [Google Scholar] [CrossRef] [PubMed]
  29. Ratledge, C. Fatty Acid Biosynthesis in Microorganisms Being Used for Single Cell Oil Production. Biochimie 2004, 86, 807–815. [Google Scholar] [CrossRef]
  30. Sakuradani, E.; Ando, A.; Ogawa, J.; Shimizu, S. Improved Production of Various Polyunsaturated Fatty Acids through Filamentous Fungus Mortierella alpina Breeding. Appl. Microbiol. Biotechnol. 2009, 84, 1–10. [Google Scholar] [CrossRef] [PubMed]
  31. Felse, P.A.; Shah, V.; Chan, J.; Rao, K.J.; Gross, R.A. Sophorolipid Biosynthesis by Candida bombicola from Industrial Fatty Acid Residues. Enzym. Microb. Technol. 2007, 40, 316–323. [Google Scholar] [CrossRef]
  32. Ramani, K.; Boopathy, R.; Vidya, C.; Kennedy, L.J.; Velan, M.; Sekaran, G. Immobilisation of Pseudomonas gessardii Acidic Lipase Derived from Beef Tallow onto Mesoporous Activated Carbon and Its Application on Hydrolysis of Olive Oil. Process Biochem. 2010, 45, 986–992. [Google Scholar] [CrossRef]
  33. Bornscheuer, U.T. Enzymes in Lipid Modification: An Overview. In Lipid Modification by Enzymes and Engineered Microbes; Academic Press: Cambridge, MA, USA, 2018; pp. 1–9. [Google Scholar] [CrossRef]
  34. Jaeger, K.E.; Ransac, S.; Dijkstra, B.W.; Colson, C.; van Heuvel, M.; Misset, O. Bacterial Lipases. FEMS Microbiol. Rev. 1994, 15, 29–63. [Google Scholar] [CrossRef]
  35. Javed, S.; Azeem, F.; Hussain, S.; Rasul, I.; Siddique, M.H.; Riaz, M.; Afzal, M.; Kouser, A.; Nadeem, H. Bacterial Lipases: A Review on Purification and Characterization. Prog. Biophys. Mol. Biol. 2018, 132, 23–34. [Google Scholar] [CrossRef]
  36. Schmid, R.D.; Verger, R. Lipases: Interfacial Enzymes with Attractive Applications. Angew. Chem. Int. Ed. 1998, 37, 1608–1633. [Google Scholar] [CrossRef]
  37. Caicedo-Paz, A.V.; Mussagy, C.U.; Mesa, V.; Junior, R.H.M.; Valenzuela, R.; de Paula, A.V.; Martinez-Galan, J.P. Enzymatic synthesis of structured lipids from sacha inchi (Plukenetia volubilis) oil with capric acid via acidolysis reaction in stirred tank and packed bed mini reactors. Food Biosci. 2024, 58, 103769. [Google Scholar] [CrossRef]
  38. Caicedo-Paz, A.V.; Mediavilla, M.; Farías, C.; Valenzuela, R.; Mussagy, C.U.; Martinez-Galan, J.P. Sustainable one-pot solvent-free enzymatic synthesis of capric acid-rich structured lipids to enhance the nutritional value of grape seed oil. Process Biochem. 2024, 144, 160–167. [Google Scholar] [CrossRef]
  39. Gunstone, F.D.; Hamilton, R.J. Oleochemical Manufacture and Applications; Wiley-Blackwell: Hoboken, NJ, USA, 2001; Volume 325. [Google Scholar]
  40. Gamayurova, V.S.; Zinov’Eva, M.E.; Tran, H.T.T. Features of the Enzymatic Hydrolysis of Castor Oil. Catal. Ind. 2013, 5, 269–273. [Google Scholar] [CrossRef]
  41. Mensah, M.B.; Awudza, J.A.M.; O’Brien, P. Castor Oil: A Suitable Green Source of Capping Agent for Nanoparticle Syntheses and Facile Surface Functionalization. R. Soc. Open Sci. 2018, 5, 180824. [Google Scholar] [CrossRef] [PubMed]
  42. Nitbani, F.O.; Tjitda, P.J.P.; Wogo, H.E.; Detha, A.I.R. Preparation of Ricinoleic Acid from Castor Oil: A Review. J. Oleo Sci. 2022, 71, 781–793. [Google Scholar] [CrossRef]
  43. Rathod, V.K.; Pandit, A.B. Effect of Various Additives on Enzymatic Hydrolysis of Castor Oil. Biochem. Eng. J. 2009, 47, 93–99. [Google Scholar] [CrossRef]
  44. Broun, P.; Shanklin, J.; Whittle, E.; Somerville, C. Catalytic Plasticity of Fatty Acid Modification Enzymes Underlying Chemical Diversity of Plant Lipids. Science 1998, 282, 1315–1317. [Google Scholar] [CrossRef]
  45. Thakkar, K.; Kachhwaha, S.S.; Kodgire, P. Multi-Response Optimization of Transesterification Reaction for Biodiesel Production from Castor Oil Assisted by Hydrodynamic Cavitation. Fuel 2022, 308, 121907. [Google Scholar] [CrossRef]
  46. Metzger, J.O.; Bornscheuer, U. Lipids as Renewable Resources: Current State of Chemical and Biotechnological Conversion and Diversification. Appl. Microbiol. Biotechnol. 2006, 71, 13–22. [Google Scholar] [CrossRef]
  47. Goh, E.B.; Baidoo, E.E.K.; Keasling, J.D.; Beller, H.R. Engineering of Bacterial Methyl Ketone Synthesis for Biofuels. Appl. Environ. Microbiol. 2012, 78, 70–80. [Google Scholar] [CrossRef] [PubMed]
  48. Heshof, R.; de Graaff, L.H.; Villaverde, J.J.; Silvestre, A.J.D.; Haarmann, T.; Dalsgaard, T.K.; Buchert, J. Industrial Potential of Lipoxygenases. Crit. Rev. Biotechnol. 2016, 36, 665–674. [Google Scholar] [CrossRef]
  49. Rincón, L.A.; Ramírez, J.C.; Orjuela, A. Assessment of Degumming and Bleaching Processes for Used Cooking Oils Upgrading into Oleochemical Feedstocks. J. Environ. Chem. Eng. 2021, 9, 21–23. [Google Scholar] [CrossRef]
  50. Le Borgne, S.; Quintero, R. Biotechnological Processes for the Refining of Petroleum. Fuel Process. Technol. 2003, 81, 155–169. [Google Scholar] [CrossRef]
  51. Robles-Medina, A.; Gonzalez-Moreno, P.A.; Esteban-Cerdan, L.; Molina-Grima, E. Biocatalysis: Towards Ever Greener Biodiesel Production. Biotechnol. Adv. 2009, 27, 398–408. [Google Scholar] [CrossRef]
  52. Royce, L.A.; Yoon, J.M.; Chen, Y.; Rickenbach, E.; Shanks, J.V.; Jarboe, L.R. Evolution for Exogenous Octanoic Acid Tolerance Improves Carboxylic Acid Production and Membrane Integrity. Metab. Eng. 2015, 29, 180–188. [Google Scholar] [CrossRef] [PubMed]
  53. di Bitonto, L.; Todisco, S.; Gallo, V.; Pastore, C. Urban sewage scum and primary sludge as profitable sources of biodiesel and biolubricants of new generation. Bioresour. Technol. Rep. 2020, 9, 100382. [Google Scholar] [CrossRef]
  54. di Bitonto, L.; Angelini, A.; Pastore, C. Energy Recovery from Municipal Sewage Sludge: An Environmentally Friendly Source for the Production of Biochemicals. Appl. Sci. 2024, 14, 4974. [Google Scholar] [CrossRef]
  55. Jo, Y.S.; An, J.U.; Oh, D.K. γ-Dodecelactone production from safflower oil via 10-hydroxy-12(Z)-octadecenoic acid intermediate by whole cells of Candida boidinii and Stenotrophomonas nitritireducens. J. Agric. Food Chem. 2014, 62, 6736–6745. [Google Scholar] [CrossRef]
  56. An, J.U.; Joo, Y.C.; Oh, D.K. New biotransformation process for production of the fragrant compound γ-dodecalactone from 10-hydroxystearate by permeabilized Waltomyces lipofer cells. Appl. Environ. Microbiol. 2013, 79, 2636–2641. [Google Scholar] [CrossRef]
  57. Wallen, L.; Benedict, R.; Jackson, R. The microbiological production of 10-hydroxystearic acid from oleic acid. Arch. Biochem. Biophys. 1962, 99, 249–253. [Google Scholar] [CrossRef] [PubMed]
  58. Bevers, L.E.; Pinkse, M.W.; Verhaert, P.D.; Hagen, W.R. Oleate hydratase catalyzes the hydration of a nonactivated carbon-carbon bond. J. Bacteriol. 2009, 191, 5010–5012. [Google Scholar] [CrossRef] [PubMed]
  59. Rosberg-Cody, E.; Liavonchanka, A.; Gobel, C.; Ross, R.P.; O’Sullivan, O.; Fitzgerald, G.F.; Stanton, C. Myosin-cross-reactive antigen (MCRA) protein from Bifidobacterium breve is a FAD-dependent fatty acid hydratase which has a function in stress protection. BMC Biochem. 2011, 12, 9. [Google Scholar] [CrossRef]
  60. Yang, B.; Chen, H.; Song, Y.; Chen, Y.Q.; Zhang, H.; Chen, W. Myosin-cross-reactive antigens from four different lactic acid bacteria are fatty acid hydratases. Biotechnol. Lett. 2013, 35, 75–81. [Google Scholar] [CrossRef]
  61. Volkov, A.; Liavonchanka, A.; Kamneva, O.; Fiedler, T.; Goebel, C.; Kreikemeryer, B.; Feussner, I. Myosin cross-reactive antigen of Streptococcus pyogenes M49 encodes a fatty acid double bond hydratase that plays a role in oleic acid detoxification and bacterial virulence. J. Biol. Chem. 2010, 285, 10353–10361. [Google Scholar] [CrossRef]
  62. Engleder, M.; Pavkov-Keller, T.; Emmerstorfer, A.; Hromic, A.; Schrempf, S.; Steinkellner, G.; Wriessnegger, T.; Leitner, E.; Strohmeier, G.A.; Kaluzna, I.; et al. Structure-based mechanism of oleate hydratase from Elizabethkingia meningoseptica. ChemBioChem 2015, 16, 1730–1734. [Google Scholar] [CrossRef]
  63. Kang, W.R.; Seo, M.J.; Shin, K.C.; Park, J.B.; Oh, D.K. Comparison of biochemical properties of the original and newly identified oleate hydratases from Stenotrophomonas maltophilia. Appl. Environ. Microbiol. 2017, 83, e03351-16. [Google Scholar] [CrossRef]
  64. Kim, B.N.; Joo, Y.C.; Kim, Y.S.; Kim, K.R.; Oh, D.K. Production of 10-hydroxystearic acid from oleic acid and olive oil hydrolyzate by an oleate hydratase from Lysinibacillus fusiformis. Appl. Microbiol. Biotechnol. 2012, 95, 929–937. [Google Scholar] [CrossRef] [PubMed]
  65. Jeon, E.Y.; Lee, J.H.; Yang, K.M.; Joo, Y.C.; Oh, D.K.; Park, J.B. Bioprocess engineering to produce 10-hydroxystearic acid from oleic acid by recombinant Escherichia coli expressing the oleate hydratase gene of Stenotrophomonas maltophilia. Proc. Biochem. 2012, 47, 941–947. [Google Scholar] [CrossRef]
  66. Kim, K.R.; Oh, H.J.; Park, C.S.; Hong, S.H.; Park, J.Y.; Oh, D.K. Unveiling of novel regio-selective fatty acid double bond hydratases from Lactobacillus acidophilus involved in the selective oxyfunctionalization of mono- and di-hydroxy fatty acids. Biotechnol. Bioeng. 2015, 112, 2206–2213. [Google Scholar] [CrossRef]
  67. Song, J.W.; Lee, J.H.; Bornscheuer, U.T.; Park, J.B. Microbial synthesis of medium chain α,ω-dicarboxylic acids and ω-aminocarboxylic acids from renewable long chain fatty acids. Adv. Synth. Catal. 2014, 356, 1782–1788. [Google Scholar] [CrossRef]
  68. Hirata, A.; Kishino, S.; Park, S.B.; Takeuchi, M.; Kitamura, N.; Ogawa, J. A novel unsaturated fatty acid hydratase toward C16 to C22 fatty acids from Lactobacillus acidophilus. J. Lipid Res. 2015, 56, 1340–1350. [Google Scholar] [CrossRef] [PubMed]
  69. Takeuchi, M.; Kishino, S.; Park, S.B.; Hirata, A.; Kitamura, N.; Saika, A.; Ogawa, J. Efficient enzymatic production of hydroxy fatty acids by linoleic acid 9 hydratase from Lactobacillus plantarum AKU 1009a. J. Appl. Microbiol. 2016, 120, 1282–1288. [Google Scholar] [CrossRef] [PubMed]
  70. Broadwater, J.A.; Whittle, E.; Shanklin, J. Desaturation and hydroxylation. J. Biol. Chem. 2002, 227, 15613–15620. [Google Scholar] [CrossRef] [PubMed]
  71. Kim, I.J.; Bayer, T.; Terholsen, H.; Bornscheuer, U. α-Dioxygenases (α-DOXs): Promising Biocatalysts for the Environmentally Friendly Production of Aroma Compounds. ChemBioChem 2022, 23, 202100693. [Google Scholar] [CrossRef]
  72. Kaehne, F.; Buchhaupt, M.; Schrader, J. A Recombinant α-Dioxygenase from Rice to Produce Fatty Aldehydes Using E. coli. Appl. Microbiol. Biotechnol. 2011, 90, 989–995. [Google Scholar] [CrossRef]
  73. Maurer, S.; Schewe, H.; Schrader, J.; Buchhaupt, M. Investigation of Fatty Aldehyde and Alcohol Synthesis from Fatty Acids by αDox- or CAR-Expressing Escherichia coli. J. Biotechnol. 2019, 305, 11–17. [Google Scholar] [CrossRef]
  74. Otte, K.B.; Kittelberger, J.; Kirtz, M.; Nestl, B.M.; Hauer, B. Whole-Cell One-Pot Biosynthesis of Azelaic Acid. ChemCatChem 2014, 6, 1003–1009. [Google Scholar] [CrossRef]
  75. Coenen, A.; Ferrer, M.; Jaeger, K.E.; Schörken, U. Synthesis of 12-Aminododecenoic Acid by Coupling Transaminase to Oxylipin Pathway Enzymes. Appl. Microbiol. Biotechnol. 2023, 107, 2209–2221. [Google Scholar] [CrossRef]
  76. Qi, Y.K.; Pan, J.; Zhang, Z.J.; Xu, J.H. Whole-Cell One-Pot Biosynthesis of Dodecanedioic Acid from Renewable Linoleic Acid. Bioresour. Bioprocess. 2023, 11, 55. [Google Scholar] [CrossRef]
  77. Seo, M.J.; Shin, K.C.; An, J.U.; Kang, W.R.; Ko, Y.J.; Oh, D.K. Characterization of a Recombinant 7,8-Linoleate Diol Synthase from Glomerella cingulate. Appl. Microbiol. Biotechnol. 2016, 100, 3087–3099. [Google Scholar] [CrossRef] [PubMed]
  78. Shin, K.C.; Seo, M.J.; Kang, S.H.; Oh, D.K. Production of 8,11-Dihydroxy Fatty Acids from Oleic and Palmitoleic Acids by Escherichia Coli Cells Expressing Variant 6,8-Linoleate Diol Synthases from Penicillium oxalicum. Biotechnol. Prog. 2022, 38, e3267. [Google Scholar] [CrossRef]
  79. Seo, M.J.; Shin, K.C.; Oh, D.K. Production of 5,8-Dihydroxy-9,12(Z,Z)-Octadecadienoic Acid From Linoleic Acid by Whole Recombinant Escherichia coli Cells Expressing Diol Synthase From Aspergillus nidulans. Appl. Microbiol. Biotechnol. 2014, 98, 7447–7456. [Google Scholar] [CrossRef] [PubMed]
  80. Andreou, A.; Feussner, I. Lipoxygenases: Structure and reaction mechanism. Phytochemistry 2009, 70, 1504–1510. [Google Scholar] [CrossRef]
  81. de Roos, A.L.; van Dijk, A.A.; Folkertsma, B. Bleaching of Dairy Products. U.S. Patent 20060127533 A1, 15 May 2006. [Google Scholar]
  82. Palmieri-Thiers, C.; Alberti, J.C.; Canaan, S.; Brunini, V.; Gambotti, C.; Tomi, F.; Oliw, E.H.; Berti, L.; Maury, J. Identification of putative residues involved in the accessibility of the substrate-binding site of lipoxygenase by site-directed mutagenesis studies. Arch. Biochem. Biophys. 2011, 509, 82–89. [Google Scholar] [CrossRef] [PubMed]
  83. Guo, F.; Zhang, C.; Bie, X.; Zhao, H.; Diao, H.; Lu, F.; Lu, Z. Improving the thermostability and activity of lipoxygenase from Anabaena sp. PCC7120 by directed evolution and site-directed mutagenesis. J. Mol. Catal. B 2014, 107, 23–30. [Google Scholar] [CrossRef]
  84. Coffa, G.; Schneider, C.; Brash, A.R. A comprehensive model of positional and stereo control in lipoxygenases. Biochem. Biophys. Res. Commun. 2005, 338, 87–92. [Google Scholar] [CrossRef]
  85. Coffa, G.; Brash, A.R. A single active site residue directs oxygenation stereospecificity in lipoxygenases: Stereocontrol is linked to the position of oxygenation. Proc. Natl. Acad. Sci. USA 2004, 101, 15579–15584. [Google Scholar] [CrossRef]
  86. Villaverde, J.J.; van der Vlist, V.; Santos, S.A.O.; Haarmann, T.; Langfelder, K.; Pirttimaa, M.; Nyyssölä, A.; Jylhä, S.; Tamminen, T.; Kruus, K.; et al. Hydroperoxide production from linoleic acid by heterologous Gaeumannomyces graminis tritici lipoxygenase: Optimization and scaleup. Chem. Eng. J. 2013, 217, 82–90. [Google Scholar] [CrossRef]
  87. Bruhlmann, F.; Bosijokovic, B.; Ullmann, C.; Auffray, P.; Fourage, L.; Wahler, D. Directed evolution of a 13-hydroperoxide lyase (CYP74B) for improved process performance. J. Biotechnol. 2013, 163, 339–345. [Google Scholar] [CrossRef]
  88. Bruhlmann, F.; Bosijokovic, B. Efficient biochemical cascade for accessing green leaf alcohols. Org. Process Res. Dev. 2016, 20, 1974–1978. [Google Scholar] [CrossRef]
  89. Rude, M.A.; Baron, T.S.; Brubaker, S.; Alibhai, M.; Del Cardayre, S.B.; Schirmer, A. Terminal olefin (1-alkene) biosynthesis by a novel P450 fatty acid decarboxylase from Jeotgalicoccus species. Appl. Environ. Microbiol. 2011, 77, 1718–1727. [Google Scholar] [CrossRef] [PubMed]
  90. Chen, H.; Huang, M.; Yan, W.; Bai, W.J.; Wang, X. Enzymatic regio-and enantioselective C–H oxyfunctionalization of fatty acids. ACS Catal. 2021, 11, 10625–10630. [Google Scholar] [CrossRef]
  91. Jiang, Y.; Li, Z.; Wang, C.; Zhou, Y.J.; Xu, H.; Li, S. Biochemical characterization of three new α-olefin-producing P450 fatty acid decarboxylases with a halophilic property. Biotechnol. Biofuels 2019, 12, 79. [Google Scholar] [CrossRef]
  92. Lee, J.W.; Niraula, N.P.; Trinh, C.T. Harnessing a P450 fatty acid decarboxylase from Macrococcus caseolyticus for microbial biosynthesis of odd chain terminal alkenes. Metab. Eng. Commun. 2018, 7, e00076. [Google Scholar] [CrossRef]
  93. Xu, H.; Ning, L.; Yang, W.; Fang, B.; Wang, C.; Wang, Y.; Xu, J.; Collin, S.; Laurent, F.; Li, S. In vitro oxidative decarboxylation of free fatty acids to terminal alkenes by two new P450 peroxygenases. Biotechnol. Biofuels 2017, 10, 208. [Google Scholar] [CrossRef]
  94. Iqbal, T.; Murugan, S.; Das, D. A chimeric membrane enzyme and an engineered whole-cell biocatalyst for efficient 1-alkene production. Sci. Adv. 2024, 10, eadl2492. [Google Scholar] [CrossRef] [PubMed]
  95. Iqbal, T.; Murugan, S.; Rajendran, K.; Sidhu, J.S.; Das, D. Unraveling the conversion of fatty acids into terminal alkenes by an integral membrane enzyme, UndB. ACS Catal. 2023, 13, 15516–15525. [Google Scholar] [CrossRef]
  96. Manley, O.M.; Fan, R.; Guo, Y.; Makris, T.M. Oxidative decarboxylase UndA utilizes a dinuclear iron cofactor. J. Am. Chem. Soc. 2019, 141, 8684–8688. [Google Scholar] [CrossRef]
  97. Rui, Z.; Harris, N.C.; Zhu, X.; Huang, W.; Zhang, W. Discovery of a family of desaturase-like enzymes for 1-alkene biosynthesis. ACS Catal. 2015, 5, 7091–7094. [Google Scholar] [CrossRef]
  98. Jiang, Y.; Li, Z.; Zheng, S.; Xu, H.; Zhou, Y.J.; Gao, Z.; Meng, C.; Li, S. Establishing an enzyme cascade for one-pot production of α-olefins from low-cost triglycerides and oils without exogenous H2O2 addition. Biotechnol. Biofuels 2020, 13, 52. [Google Scholar] [CrossRef] [PubMed]
  99. Matthews, S.; Tee, K.L.; Rattray, N.J.; McLean, K.J.; Leys, D.; Parker, D.A.; Blankley, R.T.; Munro, A.W. Production of alkenes and novel secondary products by P450 Ole TJE using novel H2O2-generating fusion protein systems. FEBS Lett. 2017, 591, 737–750. [Google Scholar] [CrossRef] [PubMed]
  100. Sorigué, D.; Légeret, B.; Cuiné, S.; Blangy, S.; Moulin, S.; Billon, E.; Beisson, F. An algal photoenzyme converts fatty acids to hydrocarbons. Science 2017, 357, 903–907. [Google Scholar] [CrossRef]
  101. Sorigué, D.; Légeret, B.; Cuiné, S.; Morales, P.; Mirabella, B.; Guédeney, G.; Beisson, F. Microalgae synthesize hydrocarbons from long-chain fatty acids via a light-dependent pathway. Plant Physiol. 2016, 171, 2393–2405. [Google Scholar] [CrossRef]
  102. Ma, Y.; Wang, Y.; Wu, B.; Zhou, J.; Yang, S.; Zhang, F.; Luo, K.; Wang, Y.; Hollmann, F. Photobiocatalysis: More than just an interesting lab curiosity? Chem. Catal. 2024, 4, 101077. [Google Scholar] [CrossRef]
  103. Li, D.; Han, T.; Xue, J.; Xu, W.; Xu, J.; Wu, Q. Engineering fatty acid photodecarboxylase to enable highly selective decarboxylation of trans fatty acids. Angew. Chem. 2021, 133, 20863–20867. [Google Scholar] [CrossRef]
  104. Bernhardt, R.; Urlacher, V.B. Cytochrome P450s as promising catalysts for biotechnological application: Chances and limitations. Appl. Microbiol. Biotechnol. 2014, 98, 6185–6203. [Google Scholar] [CrossRef]
  105. Urlacher, V.B.; Girhard, M. Cytochrome P450 monooxygenases: An update on perspectives for synthetic application. Trends Biotechnol. 2012, 30, 26–36. [Google Scholar] [CrossRef]
  106. Van Bogaert, I.N.; Groeneboer, S.; Saerens, K.; Soetaert, W. The role of cytochrome P450 monooxygenases in microbial fatty acid metabolism. FEBS J. 2011, 278, 206–221. [Google Scholar] [CrossRef]
  107. Biermann, U.; Bornscheuer, U.; Meier, M.A.R.; Metzger, J.O.; Schafer, H.J. Oils and fats as renewable raw materials in chemistry. Angew. Chem. Int. Ed. 2011, 50, 3854–3871. [Google Scholar] [CrossRef]
  108. Munro, A.W.; Leys, D.G.; McLean, K.J.; Marshall, K.R.; Ost, T.; Daff, S.; Dutton, P.L. P450 BM3: The very model of a modern flavocytochrome. Trends Biochem. Sci. 2002, 27, 250–257. [Google Scholar] [CrossRef] [PubMed]
  109. Dietrich, J.A.; Yoshikuni, Y.; Fisher, K.J.; Woolard, F.X.; Ockey, D.; McPhee, D.J.; Keasling, J.D. A novel semi-biosynthetic route for artemisinin production using engineered substrate-promiscuous P450BM3. ACS Chem. Biol. 2009, 4, 261–267. [Google Scholar] [CrossRef] [PubMed]
  110. Seifert, A.; Vomund, S.; Grohmann, K.; Kriening, S.; Urlacher, V.B. Rational design of a minimal and highly enriched CYP102A1 mutant library with improved regio-, stereo- and chemoselectivity. ChemBioChem 2009, 10, 853–861. [Google Scholar] [CrossRef]
  111. Bruhlmann, F.; Fourage, L.; Ullmann, C.; Haefliger, O.P.; Jeckelmann, N. Engineering cytochrome P450 BM3 of Bacillus megaterium for terminal oxidation of palmitic acid. J. Biotechnol. 2014, 184, 17–26. [Google Scholar] [CrossRef] [PubMed]
  112. Khatri, Y.; Hannemann, F.; Girhard, M.; Kappl, R.; Hutter, M.; Jeckelmann, N.; Dubois, C.; Wahler, D. A natural heme-signature variant of CYP267A1 from Sorangium cellulosum So ce56 executes diverse ω-hydroxylation. FEBS J. 2015, 282, 74–88. [Google Scholar] [CrossRef]
  113. Durairaj, P.; Malla, S.; Nadarajan, S.P.; Lee, P.G.; Jung, E. Fungal cytochrome P450 monooxygenases of Fusarium oxysporum for the synthesis of ω-hydroxy fatty acids in engineered Saccharomyces cerevisiae. Microb. Cell Fact. 2015, 14, 45. [Google Scholar] [CrossRef]
  114. Keera, S.T.; El Sabagh, S.M.; Taman, A.R. Transesterification of vegetable oil to biodiesel fuel using alkaline catalyst. Fuel 2011, 90, 42–47. [Google Scholar] [CrossRef]
  115. Tubino, M.; Junior, J.G.R.; Bauerfeldt, G.F. Biodiesel synthesis: A study of the triglyceride methanolysis reaction with alkaline catalysts. Catal. Comm. 2016, 75, 6–12. [Google Scholar] [CrossRef]
  116. Canakci, M. The potential of restaurant waste lipids as biodiesel feedstocks. Bioresour. Technol. 2007, 98, 183–190. [Google Scholar] [CrossRef]
  117. Nurhayati, N.; Anita, S.; Amri, T.A.; Linggawati, A. Esterification of Crude Palm Oil Using H2SO4 and Transesterification Using CaO Catalyst Derived from Anadara granosa. Indones. J. Chem. 2017, 17, 309–315. [Google Scholar] [CrossRef]
  118. Verma, P.; Sharma, M.P. Review of process parameters for biodiesel production from different feedstocks. Renew. Sustain. Energy Rev. 2016, 62, 1063–1071. [Google Scholar] [CrossRef]
  119. Eze, V.C.; Phan, A.N.; Harvey, A.P. Intensified one-step biodiesel production from high water and free fatty acid waste cooking oils. Fuel 2018, 220, 567–574. [Google Scholar] [CrossRef]
  120. Tavizón-Pozos, J.A.; Chavez-Esquivel, G.; Suárez-Toriello, V.A.; Santolalla-Vargas, C.E.; Luévano-Rivas, O.A.; Valdés-Martínez, O.U.; Lopez, A.T.; Rodriguez, J.A. State of art of alkaline earth metal oxides catalysts used in the transesterification of oils for biodiesel production. Energies 2021, 14, 1031. [Google Scholar] [CrossRef]
  121. Abidin, S.Z.; Haigh, K.F.; Saha, B. Esterification of free fatty acids in used cooking oil using ion-exchange resins as catalysts: An efficient pretreatment method for biodiesel feedstock. Ind. Eng. Chem. Res. 2012, 51, 14653–14664. [Google Scholar] [CrossRef]
  122. Park, J.Y.; Kim, D.K.; Lee, J.S. Esterification of free fatty acids using water-tolerable Amberlyst as a heterogeneous catalyst. Bioresour. Technol. 2010, 101, S62–S65. [Google Scholar] [CrossRef]
  123. Özbay, N.; Oktar, N.; Tapan, N. Esterification of free fatty acids in waste cooking oils (WCO): Role of ion-exchange resins. Fuel 2008, 87, 1789–1798. [Google Scholar] [CrossRef]
  124. D’Souza, R.; Vats, T.; Chattree, A.; Siril, P.F. Graphene supported magnetically separable solid acid catalyst for the single step conversion of waste cooking oil to biodiesel. Renew. Energy 2018, 126, 1064–1073. [Google Scholar] [CrossRef]
  125. Sani, Y.M.; Daud, W.M.A.W.; Aziz, A.A. Activity of solid acid catalysts for biodiesel production: A critical review. Appl. Catal. A Gen. 2014, 470, 140–161. [Google Scholar] [CrossRef]
  126. Su, F.; Guo, Y. Advancements in solid acid catalysts for biodiesel production. Green Chem. 2014, 16, 2934–2957. [Google Scholar] [CrossRef]
  127. Wang, A.; Quan, W.; Zhang, H.; Li, H.; Yang, S. Heterogeneous ZnO-Containing Catalysts for Efficient Biodiesel Production. RSC Adv. 2021, 11, 20465–20478. [Google Scholar] [CrossRef]
  128. Nie, Z.; Jiang, P.; Zhang, P.; Liu, D.; Haryono, A.; Zhao, M.; Zhao, H. Zirconia-Coated Magnetic Fe2O3 Nanoparticles Supported 12-Tungstophosphoric Acid: A Novel Environmentally Friendly Catalyst for Biodiesel Production. J. Chin. Chem. Soc. 2021, 68, 837–848. [Google Scholar] [CrossRef]
  129. Okechukwu, O.D.; Joseph, E.; Nonso, U.C.; Kenechi, N.-O. Improving Heterogeneous Catalysis for Biodiesel Production Process. Clean. Chem. Eng. 2022, 3, 100038. [Google Scholar] [CrossRef]
  130. Park, J.Y.; Park, K.M. Lipase and Its Unique Selectivity: A Mini-Review. J. Chem. 2022, 1, 7609019. [Google Scholar] [CrossRef]
  131. Fernandez-Lafuente, R. Lipase from Thermomyces lanuginosus: Uses and Prospects as An Industrial Biocatalyst. J. Mol. Catal. B Enzym. 2010, 62, 197–212. [Google Scholar] [CrossRef]
  132. Nitbani, F.O.; Jumina, J.; Tjitda, P.J.P.; Wogo, H.E.; Detha, A.I.; Nurohmah, B.A. Synthesis of 2-Monolaurin from Pure Lauric Acid. In Proceedings of the AIP Conference Proceedings, Yoyakarta, Indonesia, 30 September 2020. [Google Scholar]
  133. Jumina; Nurmala, A.; Fitria, A.; Pranowo, D.; Sholikhah, E.N.; Kurniawan, Y.S.; Kuswandi, B. Monomyristin and Monopalmitin Derivatives: Synthesis and Evaluation as Potential Antibacterial and Antifungal Agents. Molecules 2018, 23, 3141. [Google Scholar] [CrossRef]
  134. Nitbani, F.; Angwarmasse, L.; Bessy, E.; Wogo, H.E.; Detha, A.I.; Tjitda, P.J. Improved Synthesis of α-Glycerol Monolaurate Using Lipozyme TL IM. J. Oleo Sci. 2022, 71, 1013–1020. [Google Scholar] [CrossRef] [PubMed]
  135. Nitbani, F.O.; Tjitda, P.J.P.; Detha, A.I.R.; Nitti, F. Modification of A-Glycerol Mono Caprate Synthesis from Capric Acid. Rasayan J. Chem. 2023, 16, 549–556. [Google Scholar] [CrossRef]
  136. Karmee, S.K. Moving towards the Application of Biocatalysis in Food Waste Biorefinery. Fermentation 2023, 9, 73. [Google Scholar] [CrossRef]
  137. Lee, S.Y.; Khoiroh, I.; Vo, D.V.N.; Senthil Kumar, P.; Show, P.L. Techniques of Lipid Extraction from Microalgae for Biofuel Production: A Review. Environ. Chem. Lett. 2020, 19, 231–251. [Google Scholar] [CrossRef]
  138. Badgujar, K.C.; Badgujar, V.C.; Bhanage, B.M. Lipase as a Green and Sustainable Material for Production of Levulinate Compounds: State of the Art. Mater. Sci. Energy Technol. 2022, 5, 232–242. [Google Scholar] [CrossRef]
  139. Kumar, A.; Dhar, K.; Kanwar, S.S.; Arora, P.K. Lipase Catalysis in Organic Solvents: Advantages and Applications. Biol. Proced. Online 2016, 18, 1–11. [Google Scholar] [CrossRef] [PubMed]
  140. Domínguez de María, P. Biocatalysis, Sustainability, and Industrial Applications: Show Me the Metrics. Curr. Opin. Green Sustain. Chem. 2021, 31, 100514. [Google Scholar] [CrossRef]
  141. Satyawali, Y.; Cauwenberghs, L.; Maesen, M.; Dejonghe, W. Lipase Catalyzed Solvent Free Synthesis of Monoacylglycerols in Various Reaction Systems and Coupling Reaction with Pervaporation for in situ Water Removal. Chem. Eng. Process. Intensif. 2021, 166, 108475. [Google Scholar] [CrossRef]
  142. 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]
  143. Pasha, M.K.; Rahim, M.; Dai, L.; Liu, D.; Du, W.; Guo, M. Comparative Study of a Two-Step Enzymatic Process and Conventional Chemical Methods for Biodiesel Production: Economic and Environmental Perspectives. Chem. Eng. J. 2024, 489, 151254. [Google Scholar] [CrossRef]
  144. Hosney, H.; Mustafa, A. Semi-Continuous Production of 2-Ethyl Hexyl Ester in a Packed Bed Reactor: Optimization and Economic Evaluation. J. Oleo Sci. 2020, 69, 31–41. [Google Scholar] [CrossRef]
  145. Holm, H.C.; Nielsen, P.M.; Longin, F. Chapter 15—Upscaling of Enzymatic Processes. In Lipid Modification by Enzymes and Engineered Microbes; AOCS Press: Urbana, IL, USA, 2018; pp. 343–373. ISBN 9780128131671. [Google Scholar]
  146. Mustafa, A.; Niikura, F.; Pastore, C.; Allam, H.A.; Hassan, O.B.; Mustafa, M.; Inayat, A.; Salah, S.A.; Salam, A.A.; Mohsen, R. Selective Synthesis of Alpha Monoglycerides by a Clean Method: Techno-Economic and Environmental Assessment. Sustain. Chem. Pharm. 2022, 27, 100690. [Google Scholar] [CrossRef]
  147. Andrade, T.A.; Martín, M.; Errico, M.; Christensen, K.V. Biodiesel Production Catalyzed by Liquid and Immobilized Enzymes: Optimization and Economic Analysis. Chem. Eng. Res. Des. 2019, 141, 1–14. [Google Scholar] [CrossRef]
  148. Budžaki, S.; Miljić, G.; Tišma, M.; Sundaram, S.; Hessel, V. Is There a Future for Enzymatic Biodiesel Industrial Production in Microreactors? Appl. Energy 2017, 201, 124–134. [Google Scholar] [CrossRef]
  149. Razzaghi, M.; Homaei, A.; Vianello, F.; Azad, T.; Sharma, T.; Nadda, A.K.; Stevanato, R.; Bilal, M.; Iqbal, H.M.N. Industrial Applications of Immobilized Nano-Biocatalysts. Bioprocess Biosyst. Eng. 2022, 45, 237–256. [Google Scholar] [CrossRef]
  150. Melo, R.L.F.; Sales, M.B.; de Castro Bizerra, V.; de Sousa Junior, P.G.; Cavalcante, A.L.G.; Freire, T.M.; Neto, F.S.; Bilal, M.; Jesionowski, T.; Soares, J.M.; et al. Recent Applications and Future Prospects of Magnetic Biocatalysts. Int. J. Biol. Macromol. 2023, 253, 126709. [Google Scholar] [CrossRef]
  151. Pereira, A.S.; de Souza, A.H.; Fraga, J.L.; Villeneuve, P.; Torres, A.G.; Amaral, P.F.F. Lipases as Effective Green Biocatalysts for Phytosterol Esters’ Production: A Review. Catalysts 2022, 12, 88. [Google Scholar] [CrossRef]
  152. Lisboa, M.C.; Wiltshire, F.M.S.; Fricks, A.T.; Dariva, C.; Carrière, F.; Lima, Á.S.; Soares, C.M.F. Oleochemistry Potential from Brazil Northeastern Exotic Plants. Biochimie 2020, 178, 96–104. [Google Scholar] [CrossRef] [PubMed]
  153. Monteiro, G.d.M.; Carvalho, E.E.N.; Lago, R.C.d.; Silva, L.G.M.d.; Souza, L.R.d.; Costa, C.A.R.d.; Boas, E.V.d.B.V. Compositional Analysis of Baru (Dipteryx alata Vogel) Pulp Highlighting Its Industrial Potential. Food Res. Int. 2025, 201, 115548. [Google Scholar] [CrossRef] [PubMed]
  154. Sorita, G.D.; Favaro, S.P.; Rodrigues, D.d.S.; Silva, W.P.d., Jr.; Leal, W.G.d.O.; Ambrosi, A.; Di Luccio, M. Aqueous Enzymatic Extraction of Macauba (Acrocomia aculeata) Pulp Oil: A Green and Sustainable Approach for High-Quality Oil Production. Food Res. Int. 2024, 182, 114160. [Google Scholar] [CrossRef] [PubMed]
  155. Borin, G.P.; Alves, R.F.; Júnior, A.D.N.F. Current Status of Biotechnological Processes in the Biofuel Industries. In Bioprocessing for Biomolecules Production; Wiley: Hoboken, NJ, USA, 2019; pp. 47–69. [Google Scholar]
  156. Gonçalves, T.d.S.; Oro, C.E.D.; Wancura, J.H.C.; dos Santos, M.S.N.; Junges, A.; Dallago, R.M.; Tres, M.V. Challenges for Energy Guidelines in Crop-Based Liquid Biofuels Development in Brazil. Next Sustain. 2023, 2, 100002. [Google Scholar] [CrossRef]
  157. Wassell, C.S.; Dittmer, T.P. Are Subsidies for Biodiesel Economically Efficient? Energy Policy 2006, 34, 3993–4001. [Google Scholar] [CrossRef]
  158. Morone, P.; Cottoni, L.; Giudice, F. Biofuels: Technology, Economics, and Policy Issues. In Handbook of Biofuels Production; Elsevier: Amsterdam, The Netherlands, 2023; pp. 55–92. [Google Scholar]
Figure 1. An overview of the common chemical reactions used to produce oleochemical products.
Figure 1. An overview of the common chemical reactions used to produce oleochemical products.
Catalysts 15 00600 g001
Figure 2. Synthesis pathway to produce (S)-10-hydroxystearic acid from oleic acid catalysed by hydratase.
Figure 2. Synthesis pathway to produce (S)-10-hydroxystearic acid from oleic acid catalysed by hydratase.
Catalysts 15 00600 g002
Figure 3. Synthesis of ricinoleic acid from oleic acid catalysed by hydroxylase.
Figure 3. Synthesis of ricinoleic acid from oleic acid catalysed by hydroxylase.
Catalysts 15 00600 g003
Figure 4. Synthesis of aldehydes from fatty acids catalysed by hydroperoxidase.
Figure 4. Synthesis of aldehydes from fatty acids catalysed by hydroperoxidase.
Catalysts 15 00600 g004
Figure 5. Mechanism of diol synthetase for the synthesis of 1,2-, 1,3-, 1,4- diols.
Figure 5. Mechanism of diol synthetase for the synthesis of 1,2-, 1,3-, 1,4- diols.
Catalysts 15 00600 g005
Figure 6. Synthesis of (S)-15-HETE from arachidonic acid by using lipoxygenase.
Figure 6. Synthesis of (S)-15-HETE from arachidonic acid by using lipoxygenase.
Catalysts 15 00600 g006
Figure 7. Combined use of 13-Lipoxygenase and 13-hydroperoxidase for the synthesis of (Z)-3-hexanal starting from linolenic acid.
Figure 7. Combined use of 13-Lipoxygenase and 13-hydroperoxidase for the synthesis of (Z)-3-hexanal starting from linolenic acid.
Catalysts 15 00600 g007
Figure 8. Oxidative decarboxylation of carboxylic acids using OleT.
Figure 8. Oxidative decarboxylation of carboxylic acids using OleT.
Catalysts 15 00600 g008
Figure 9. Use of triglycerides for the synthesis of terminal alkenes. The triglyceride is first hydrolysed to glycerol and fatty acids. The fatty acids are then converted to terminal alkenes by OleT with the partial use of H2O2, which is formed during the oxidation of glycerol to glyceric acid (catalysed by alditol oxidase, AldO).
Figure 9. Use of triglycerides for the synthesis of terminal alkenes. The triglyceride is first hydrolysed to glycerol and fatty acids. The fatty acids are then converted to terminal alkenes by OleT with the partial use of H2O2, which is formed during the oxidation of glycerol to glyceric acid (catalysed by alditol oxidase, AldO).
Catalysts 15 00600 g009
Figure 10. P450 monooxygenase-catalysed hydroxylation of fatty acids to produce hydroxy-(ω)-fatty acids.
Figure 10. P450 monooxygenase-catalysed hydroxylation of fatty acids to produce hydroxy-(ω)-fatty acids.
Catalysts 15 00600 g010
Figure 11. Chemical structure of monoglycerides.
Figure 11. Chemical structure of monoglycerides.
Catalysts 15 00600 g011
Figure 12. Periodic (a) and temporal (b) VOSviewer® representation of the co-occurrence of author keywords in scientific articles published between 2004 and March 2024 on the topic “biocatalysis” and “oleochemical synthesis”. The data is based on the scientific platform Scopus®.
Figure 12. Periodic (a) and temporal (b) VOSviewer® representation of the co-occurrence of author keywords in scientific articles published between 2004 and March 2024 on the topic “biocatalysis” and “oleochemical synthesis”. The data is based on the scientific platform Scopus®.
Catalysts 15 00600 g012aCatalysts 15 00600 g012b
Figure 13. Periodic (a) and temporal (b) VOSviewer® representation of the bibliographic linkage of nations with at least ten published documents between 2004 and March 2024 on the subject of “biocatalysis” and “oleochemical synthesis”. The data is based on the scientific platform Scopus®.
Figure 13. Periodic (a) and temporal (b) VOSviewer® representation of the bibliographic linkage of nations with at least ten published documents between 2004 and March 2024 on the subject of “biocatalysis” and “oleochemical synthesis”. The data is based on the scientific platform Scopus®.
Catalysts 15 00600 g013
Figure 14. Evolution of the scientific research on the enzymatic synthesis of oleochemicals. The green bars represent studies focusing on the technical aspects of the topic, while the blue bars correspond to papers addressing technical/economic considerations.
Figure 14. Evolution of the scientific research on the enzymatic synthesis of oleochemicals. The green bars represent studies focusing on the technical aspects of the topic, while the blue bars correspond to papers addressing technical/economic considerations.
Catalysts 15 00600 g014
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

Wancura, J.H.C.; Cipolatti, E.P.; Manoel, E.A.; Nitbani, F.O.; Caicedo-Paz, A.V.; Mussagy, C.U.; Abdellatief, T.M.M.; Mustafa, A.; di Bitonto, L. Advanced Biocatalytic Processes for the Conversion of Renewable Feedstocks into High-Value Oleochemicals. Catalysts 2025, 15, 600. https://doi.org/10.3390/catal15060600

AMA Style

Wancura JHC, Cipolatti EP, Manoel EA, Nitbani FO, Caicedo-Paz AV, Mussagy CU, Abdellatief TMM, Mustafa A, di Bitonto L. Advanced Biocatalytic Processes for the Conversion of Renewable Feedstocks into High-Value Oleochemicals. Catalysts. 2025; 15(6):600. https://doi.org/10.3390/catal15060600

Chicago/Turabian Style

Wancura, João H. C., Eliane Pereira Cipolatti, Evelin Andrade Manoel, Febri Odel Nitbani, Angie Vanessa Caicedo-Paz, Cassamo Ussemane Mussagy, Tamer M. M. Abdellatief, Ahmad Mustafa, and Luigi di Bitonto. 2025. "Advanced Biocatalytic Processes for the Conversion of Renewable Feedstocks into High-Value Oleochemicals" Catalysts 15, no. 6: 600. https://doi.org/10.3390/catal15060600

APA Style

Wancura, J. H. C., Cipolatti, E. P., Manoel, E. A., Nitbani, F. O., Caicedo-Paz, A. V., Mussagy, C. U., Abdellatief, T. M. M., Mustafa, A., & di Bitonto, L. (2025). Advanced Biocatalytic Processes for the Conversion of Renewable Feedstocks into High-Value Oleochemicals. Catalysts, 15(6), 600. https://doi.org/10.3390/catal15060600

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