Next Article in Journal
Microwave-Assisted Hydrothermal Synthesis of Pure-Phase Sodalite (>99 wt.%) in Suspension: Methodology Design and Verification
Previous Article in Journal
Enhanced Degradation of Ethylene in Thermo-Photocatalytic Process Using TiO2/Nickel Foam
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Esters in the Food and Cosmetic Industries: An Overview of the Reactors Used in Their Biocatalytic Synthesis

by
Salvadora Ortega-Requena
,
Claudia Montiel
,
Fuensanta Máximo
,
María Gómez
,
María Dolores Murcia
and
Josefa Bastida
*
Department of Chemical Engineering, Faculty of Chemistry, Campus of Espinardo, University of Murcia, 30100 Murcia, Spain
*
Author to whom correspondence should be addressed.
Materials 2024, 17(1), 268; https://doi.org/10.3390/ma17010268
Submission received: 5 December 2023 / Revised: 29 December 2023 / Accepted: 3 January 2024 / Published: 4 January 2024
(This article belongs to the Section Catalytic Materials)

Abstract

:
Esters are versatile compounds with a wide range of applications in various industries due to their unique properties and pleasant aromas. Conventionally, the manufacture of these compounds has relied on the chemical route. Nevertheless, this technique employs high temperatures and inorganic catalysts, resulting in undesired additional steps to purify the final product by removing solvent residues, which decreases environmental sustainability and energy efficiency. In accordance with the principles of “Green Chemistry” and the search for more environmentally friendly methods, a new alternative, the enzymatic route, has been introduced. This technique uses low temperatures and does not require the use of solvents, resulting in more environmentally friendly final products. Despite the large number of studies published on the biocatalytic synthesis of esters, little attention has been paid to the reactors used for it. Therefore, it is convenient to gather the scattered information regarding the type of reactor employed in these synthesis reactions, considering the industrial field in which the process is carried out. A comparison between the performance of the different reactor configurations will allow us to draw the appropriate conclusions regarding their suitability for each specific industrial application. This review addresses, for the first time, the above aspects, which will undoubtedly help with the correct industrial implementation of these processes.

1. Introduction

In 1850, scientist Alexander William Williamson accidentally discovered the synthesis of ethers based on the reaction between an alcohol and an alkyl iodide in the presence of sulfuric acid [1]. Since then, esters have been recognized as one of the most crucial organic compounds for industrial applications, with uses in fields, such as food production, cosmetics, lubricants, pharmaceuticals, biodiesel additives, and various others [2]. The formation of an ester through the reaction of an alcohol and an organic acid has been a topic of great interest amongst scientists since the beginning, and it is widely regarded as the most effective method for studying the catalytic activity of acids because of its precision, ease of development, and reversibility [3].
Esters can be synthesized by means of esterification between acids and alcohols, as well as through transesterification, alcoholysis, or acidolysis reactions. Although the classical methods have been widely studied, the catalysts used have been refined and optimized over time to facilitate more efficient, productive, and eco-friendly procedures [4]. However, these traditional procedures often generate hazardous byproducts, have a considerable environmental impact, and require high energy consumption. To address these issues, enzymatic biocatalysis has been proposed as a revolutionary advancement in the biotechnology industry that demonstrates potential as a sustainable alternative to conventional processing methods for a wide range of everyday products. This method employs enzymes instead of traditional chemical catalysts to increase reaction rates [5]. Enzymes possess numerous properties that render them highly intriguing as they are the most effective catalysts in nature and operate under exceedingly mild conditions, including low pressure and temperature. Therefore, biocatalytic industrial processes for ester synthesis are classified under the “Green Chemistry” category since they comply with many of its “12 principles” [6].
Lipases (triacylglycerol hydrolases EC.3.1.1.3) are widely used enzymes in biocatalysis. Their biological function is to hydrolyze triglycerides and generate free fatty acids and glycerol, but they are known for their broad specificity. This is due to their ability to accept a variety of substrates besides glycerides, including amides. Therefore, lipases are employed in vitro to catalyze various reactions beyond their natural hydrolytic function. These reactions include esterification, acidolysis, interesterification, transesterification, aminolysis, and perhydrolysis, as well as a range of promiscuous reactions [7]. Lipases are highly stable, making them a suitable choice for use in diverse reaction media, including not only aqueous environments but also organic solvents, ionic liquids, supercritical fluids, and deep eutectic solvents (DESs) [8,9].
Lipases can be sourced from animals, plants, and microorganisms. Microbial lipases are the most valuable type when compared to those derived from plants or animals. This is due to the range of catalytic activities, the high production yield, the ease of genetic manipulation, the absence of seasonal fluctuations, the consistent supply, greater stability, and notably, the rapid growth rate of microorganisms in cost-effective culture media, such as byproducts from other industries [10]. Among the bacterial lipases that are commonly utilized in the industrial sector, Candida antarctica lipase B (CALB) is the most extensively used enzyme with the largest number of patents. Candida rugosa lipase (CRL), another significant yeast lipase, is a blend of different isoforms that is commercially available and documented as “generally recognized as safe” (GRAS) for use in the food industry [11]. Phospholipases from Fusarium oxysporum, Thermomyces lanuginosus, Aspergillus niger, and Trichoderma reesei are also employed in different industries [12].
Given the high cost of enzymes, the ability to reuse the biocatalyst is a desirable benefit that is made possible through immobilization, which also has the added benefit of improving enzyme stability. Enzyme immobilization is commonly practiced to modify and improve enzyme properties, including specificity, activity, and kinetic parameters. Moreover, the immobilization of enzymes results in improved separation from the reaction medium and promotes their reuse. Various methods for enzyme immobilization have been proposed, but ongoing research aims to find simpler and more cost-effective routes to obtain immobilized derivatives for industrial applications. Adsorption and covalent binding are both commonly used methods for immobilizing lipases onto a support material. Adsorption is a rapid and straightforward technique that typically results in minimal structural changes in the lipase, as the interactions between the enzyme and support are weak. In contrast, covalent binding induces strong interactions between the enzyme and support, which reduces the risk of enzyme desorption [13].
The availability of several commercial preparations on the market with exceptional properties, in terms of activity and stability, supports the use of immobilized lipases on an industrial scale. Undoubtedly, the most used immobilized lipase is Novozym® 435, a derivative of Candida antarctica lipase B (CALB), which has been on the market since 1992 and is marketed by Novozymes (Bagsvaerd, Denmark). The immobilizing support is Lewatit VP OC 1600, which is a macroporous acrylic polymer resin onto which CALB is adsorbed through interfacial activation [9]. Novozym® 435 does not aggregate, which would cause the loss of active sites in the enzyme, and has good stability over a wide pH range, especially in alkaline media. Furthermore, one of the characteristics of this enzyme is its ability to function in non-aqueous media (organic solvents or even solvent-free conditions) as it requires only a minimal aqueous layer to maintain its enzymatic activity. In addition, it has been reported in the literature that this thermophilic lipase can operate at temperatures above 100 °C and maintain activity even at 150 °C [14,15].
Lipozyme® RM IM, also produced by Novozymes, is a widely used lipase in the industry that comes from Rhizomucor miehei and offers higher conversion efficiency at lower temperatures compared to other biocatalysts. It possesses unique characteristics that allow it to cleave sn-1,3 bonds, has great stability, and has high activity even with a low water content, making it suitable for use in reactions involving organic solvents. Novozymes has also marketed Lipozyme® TL IM, which is a specific lipase derived from Thermomyces lanuginosus. It is highly effective for rearrangement reactions, such as interesterification, especially at positions 1 and 3 of triglycerides. This lipase also exhibits thermophilic properties, as it maintains suitable activity at high temperatures, up to 65 °C [5].
At present, Novozymes has restructured its product range and renamed its commercialized enzymes. Under the “Fine Chemicals” category, they now offer three immobilized lipases: Sustine® 110 IM (formerly known as Novozym® 435), Sustine® 120 IM, and Sustine® 130 IM (both lipases specific for 1,3-positions). In addition, four other immobilized lipases are available in the “Oils and Fats” section: Lipozyme® 435, Lipozyme® TL IM, Lipura® Flex, and Lipura® Select, the latter being specific for 1,3 positions. Unlike what was common on the Novozymes website, they currently do not specify the origin of their immobilized lipase preparations (www.novozymes.com/en/products, accessed on 2 January 2024).
In addition to Novozymes immobilized lipases, other commercial preparations, such as Chirazyme L-2 (from Candida antarctica) or Amano Lipase PS (from Burkholderia cepacia), are also used in ester synthesis.
Even though the first studies on ester synthesis with an immobilized lipase date back to the mid-20th century, there are very few studies that explicitly reference the reactor used in this process (geometry and configuration, operating conditions, etc.), with the majority focusing almost exclusively on batch reactors (BRs). There are even fewer studies found that describe the use of other types of reactors, such as packed-bed reactors (PBRs) or fluidized bed reactors (FBRs) [16]. Furthermore, there is a lack of research on the development of mathematical models for reaction kinetics and reactor design.
Therefore, in this present study, a systematic compilation of literature published in recent years (since 2000) is carried out, describing the reactor used in the synthesis of esters with immobilized lipases, considering the influence of the reactor configuration on the achieved conversion, as well as exploring the potential use of alternative reactors different from the conventional ones (such as membrane reactors, microreactors, etc.). Given the high number of papers found (>4900 in WOS based on the search terms “lipase” + “ester synthesis” between 2000 and 2023), in this review, only the articles that describe the enzymatic synthesis of esters used in the food and cosmetic industries have been considered. These two industries are the main ones involved in producing high-purity compounds that can be labeled as “natural”. Additionally, the processes align with the principles of “Green Chemistry” making them environmentally sustainable. Other important industrial sectors that also use esters in the formulation of their products are the biodiesel and biolubricant industries. In these cases, although the purity of the compounds used is not a primary factor when commercializing them, the growing interest in the development of sustainable processes has led to the publication, in recent years, of many studies. For this reason, this part of the study deserves to be dealt with in adequate depth in another review.

2. Reactors Used in the Biocatalytic Synthesis of Esters with Application in the Food Industry

The significant expansion of the food and beverages industry worldwide plays a pivotal role in propelling this market. Revenue in the food market for 2023 is estimated at US $9.36 trillion, with a projected annual growth rate of 6.74% (CAGR 2023–2028). The largest segment within this market is confectionery and snacks, accounting for a market volume of US $1.66 trillion in 2023, according to data sourced from Statista (https://www.statista.com/outlook/cmo/food/worldwide, accessed on 2 January 2024). The rising consumption of packaged food products and beverages across the globe, owing to evolving dietary patterns among the population, is expected to further boost the demand for additives used during food processing to enhance quality and nutritional content. The global food additives market had a valuation of US $98.40 billion in 2022 and is anticipated to experience a compound annual growth rate (CAGR) of 5.8% from 2023 to 2030, based on data from Grand View Research (https://www.grandviewresearch.com/industry-analysis/food-additives-market, accessed on 2 January 2024).
Esters occupy a prominent place among food additives, as they are used in a wide range of applications. Many references can be found describing the synthesis of sugar esters (used as emulsifiers, foaming agents, coating agents, or even stabilizers), aromatic esters, and even specific food additives. This industry, being regulated by strict quality standards, demands certain purities and the absence of byproducts from its products resulting from synthesis processes. For this reason, manufacturers of food additives are increasingly shifting from traditional chemical synthesis to alternative processes, with biocatalytic synthesis using immobilized enzymes being a standout method. The importance of developing new sustainable processes for the synthesis of esters used in the food industry is evident from the large number of articles found in the WOS database (>500 papers based on the search terms “ester synthesis” + “lipase” + ”food industry” between 2000 and 2023). As mentioned previously, only those papers that explicitly refer to the reactor used in enzymatic synthesis, studying various aspects, such as the influence of geometry or the configuration on the conversion achieved, have been considered. All the information collected in the database has been categorized according to the type of reactor used: tank (discontinuous and continuous), tubular (packed-bed and fluidized bed), and other types of reactors.
In Table 1 [17,18,19,20,21,22,23,24,25,26,27,28], articles describing the enzymatic synthesis of esters with an immobilized lipase in tank reactors are compiled. Since this type of reactor is the most widely used in the chemical industry, only studies reporting the use of a tank reactor with a volume greater than 50 mL have been compiled for the purpose of this review. Numerous studies conducted in small vessels (screw cap vials, Eppendorf vials, etc.) used for preliminary investigations into the development of new products/processes have not been considered. However, it is surprising that very few studies have looked specifically at how the characteristics of tank reactors affect the outcome of the synthesis process. Moreover, these reactors are commonly used in industry, and new studies focus more on introducing other types of reactors than on improving existing ones. Out of the 30 articles found on WOS for “ester synthesis” + “lipase” + “tank reactor” + “food” during the considered period, only 12 articles were selected for this review. The table shows that the types of esters produced enzymatically in tank reactors for this type of industry are not very diverse: sugar esters [17,18,19,25], flavors [24,26], or emulsifiers [20,21,22,23,28].
It is convenient to specify that sugar esters, which are surfactants obtained from a sugar and a fatty acid, are natural ingredients widely used in detergents, cosmetics, pharmaceuticals, and the food industry. This wide applicability means that the papers about these compounds included in this review have been classified according to their application in the food or cosmetics industry as indicated by the authors in the introduction to the papers, although all of these studies could be included in either of the two sections. As for the immobilized lipase utilized, researchers primarily choose to use commercial preparations [17,18,19,23,24,26,27], with Novozymes products being the most popular option. A few papers describe and optimize the immobilization process [20,21,22,28], and only in one study, the immobilized lipase consists of a non-growth state microorganism (Rhizopus microsporus) adsorbed onto a porous support [25].
Due to its ease of operation, the most used tank reactor is the batch reactor, although it operates in a non-steady state, which complicates the design equations. Only two references have been found that describe continuous operation in tank reactors [17,25]. In the first one, immobilized lipase particles are confined in a stainless-steel basket, and in the second, there is no explicit reference to the procedure used to retain the solid biocatalyst within the vessel. The fact that they operate in steady state does not seem to compensate for the operational difficulties that a continuous process has. Only in a few situations is fed-batch the operational option selected. It is described in the literature that short-chain acids can provoke the deactivation of the enzyme. To prevent this phenomenon, an excess of alcohol is often used, and, while it can positively affect enzymatic activity, it also complicates and increases the cost of separation and purification operations for the final product. Therefore, the use of a fed-batch reactor has been proposed, in which the acid is fed into the reactor in successive additions until the appropriate molar substrate ratio is achieved [26].
One of the strengths of using tank reactors is their operation in complete mixing, which ensures homogeneity in the reaction medium (concentrations and temperatures). For this reason, the stirring geometry and speed are crucial facts. It is quite common for publications to not explicitly reference the type of stirrer used [17,27], so it is assumed that a magnetic stirrer is employed, which is the most common in laboratory-scale processes [18,24,26]. However, the detrimental effect of this stirring and mixing procedure on the physical structure of the solid particles of immobilized lipase [9] leads researchers to use overhead stirrers when aiming to implement these processes on an industrial scale [19,20,21,22,23,25,28]. The significance of the stirring device is highlighted in two papers where the performances of different types of agitator blades are compared and their influence on process productivity is studied [19,23].
The removal of water generated in esterification is a vitally important aspect since, if not performed correctly, it can shift the reaction equilibrium towards hydrolysis. In most processes, molecular sieves are chosen for this purpose [17,18,25,26], although the option of conducting reactions in open-air reactors to allow for water evaporation or even using N2 bubbling and a vacuum has also been described [20,21,22,23,27,28]. Some of the synthesized products are high-molecular-weight esters that have high viscosity, so many reactions are carried out in organic solvents, such as acetone, 2-methyl-2-butanol, and n-hexane [17,18,19,25]. However, to simplify the final product separation and purification and comply with the principles of “Green Chemistry”, many researchers choose to perform ester synthesis in solvent-free reaction media [20,21,22,23,26,28]. Special mention of the use of supercritical CO2 should be made [18,24], which is a trend in recent years.
The results published in the reviewed papers are very promising, and conversions over 90% are achieved in most of them [19,21,22,23,28]. In addition, the authors point out that certain operating variables must be controlled to obtain good yields, such as the quantity of molecular sieves for water removal [17], using diverse methods for water removal [20], adjusting the enzyme concentration and substrates molar ratio [18], or implementing a different operating mode (fed-batch) [26]. In summary, it can be affirmed that the use of batch reactors is the primary choice when approaching the biocatalytic synthesis of esters, and in most cases, successful results are obtained.
As mentioned above, the tank reactors are not usually chosen for continuous operation; instead, it is more common to implement tubular reactors, either a packed-bed or fluidized bed, because they provide better results achieving higher productivities per unit of reactor volume. However, the main advantage of using tubular reactors is undoubtedly that, in the absence of mechanical agitators, the immobilized enzyme particles are not damaged and can therefore maintain their catalytic capacity for a longer period.
The use of tubular reactors for ester synthesis is rather infrequent in the chemical industry and in the food industry. However, the use of this type of reactor for the biocatalytic synthesis of food ingredients has been the subject of numerous studies published in recent years. Thus, in the last 23 years, 111 articles were found in WOS using the keywords “lipase” + “ester synthesis” + “packed bed reactor”. If the search is carried out with a change in the type of reactor to “fluidized bed reactor”, an additional 15 articles will be added to the list. Among these, only 16 articles, specifically related to the food industry, are included in Table 2 [29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44]. In this case, there was no restriction on the reactor volume as all the papers refer to preliminary studies, and the reactor dimensions are relatively small, usually in the order of a few centimeters or even millimeters. Only five papers describe the use of a fluidized bed reactor [33,37,38,39,40], with the packed-bed configuration being the most common. Not all processes are continuous, as there are specific cases where part of the reaction medium is recirculated [37] or even the entire reactants are recirculated through the bed using a storage tank as a reservoir, making the operation occur in batch cycles [38,40].
The characteristics of the above-cited processes are very similar to those described in the papers gathered in the previous table. The synthesized esters are mostly sucrose esters [29,30,31,32,34,38,40], although flavor compounds [35,36,39,41,42], modified fats, and structured lipids are also described [33,37,43,44]. Only four studies report the use of lipases that are in-lab immobilized by the authors [36,37,38,39], while the majority opted for the use of commercial immobilized lipases. Moreover, many processes are conducted in solvent-free reaction media [33,34,37,38,39,43,44], whereas those using organic solvents employ acetone [29,30,31,32,40], iso-octane [35], n-heptane [41], etc. Special mention should be made of those that propose the use of supercritical CO2, which is a trend in recent years, as previously mentioned [36,42].
The outcomes from utilizing tubular reactors vary, although most authors agree that both packed-bed and fluidized bed reactors offer promising potential for the future due to their ease of use, scalability, affordability, and effectiveness. The growing interest in implementing these reactors on an industrial scale is evidenced by studies comparing the performance of the commonly used tank reactor with other reactor configurations, including membrane, packedbed, and fluidized bed reactors. Table 3 [45,46,47,48,49] provides a compilation of these studies.
Several studies have highlighted the advantages of the packed-bed configuration in terms of conversion and productivity, making it a preferred option for synthesizing butyl butyrate [47] and isoamyl acetate [48]. On the other hand, some experts have noted that the fluidized bed reactor exhibits better performance in the production of amyl caprylate when compared to the batch reactor [46]. The use of a membrane reactor [45] and DES in batch and tubular reactors [49] also appears to be a viable alternative for producing esters for use in the food industry.
Table 4 compiles literature [50,51,52,53,54,55,56] on enzymatic ester synthesis using non-conventional reactors.
One notable category is microreactors [52,53,54,55]. The increasing use and advancement of microfluidic systems offers promising opportunities for researchers investigating catalytic processes. The benefits of miniaturized reactors are numerous. The compact size of these systems allows for portable applications and reduces the amount of reactant consumption and required samples. Additionally, improved heat and mass transfers enhance reaction control, while smaller reaction channels and chambers enable a more in-depth analysis. Moreover, higher degrees of automation present opportunities for both industrial production and research applications [57]. However, to date, most of the devices described have been utilized in healthcare and pharmaceutical applications with minimal research focused on synthesizing products for industries outside of these fields, such as esters.
Table 4 shows the characteristics of four processes used to synthesize different esters with food applications using microreactors, both tank [52,55] and packed-bed [53,54], with excellent results, proposing an interesting alternative to traditional reactors. On the other hand, the same table shows two other studies that investigate the operability of a tank reactor with a membrane [50,51] and a tank reactor with a rotating basket operating in successive batches [56]. In both cases, the esters synthesized belong to the sucrose esters group, obtaining conversions of 93% and 80%, respectively.

3. Reactors Used in the Biocatalytic Synthesis of Esters with Applications in the Cosmetic Industry

Cosmetics are commonly recognized as a category of items associated with personal grooming, particularly skincare. They have a history that likely dates to the beginnings of human civilization. The widespread use of cosmetic products in daily life gained momentum with the emergence of synthetic organic chemistry, which made it feasible for people to access desired ingredients and formulations with relative ease. Today, cosmetics hold a significant place, especially considering the recent “wellness” movement. Manufacturers must continually enhance their products to maintain a competitive edge in a market where consumers expect more choices and increasingly effective solutions. It is worth noting that most cosmetic products have a shelf life of less than five years, and manufacturers reformulate approximately 25% of their products each year [58].
In 2023, the Beauty and Personal Care market was estimated to generate a revenue of approximately US $625.70 billion. This market is anticipated to experience an annual growth rate of 3.32% from 2023 to 2028 according to data published by Statista (https://www.statista.com/outlook/cmo/beauty-personal-care/worldwide, accessed on 2 January 2024). The global cosmetic ingredients market, as of 2022, had a market size of about US $32 billion. It is expected to reach approximately US $55.44 billion by 2032, with a recorded compound annual growth rate (CAGR) of 5.7% during the forecast period from 2023 to 2032 (https://www.precedenceresearch.com/cosmetic-ingredients-market, accessed on 2 January 2024). As a result, cosmetic chemicals constitute a significant sector within the chemical industry primarily served by chemical companies, like BASF, Evonik Industries, Clariant, and Rhodia. Cosmetic products, overall, are predominantly promoted by international corporations, such as Procter & Gamble, L’Oreal, Unilever, Beiersdorf, and Colgate-Palmolive [58].
The worldwide beauty industry is typically segmented into five primary categories: skincare, hair care, color cosmetics (makeup), fragrances, and toiletries. The formulation of these cosmetic products is primarily influenced by factors, like their intended use, manufacturer preferences, and the target market. Esters, among all the classes of organic compounds employed in cosmetics, have diverse applications within the cosmetic sector. They serve as emollients in creams, act as surfactants in shampoos, function as antioxidants in anti-aging creams, contribute to fragrances in perfumes, and provide flavors in lip cosmetics, based on their distinct properties [59].
At present, the industrial production of cosmetic esters involves high-temperature synthesis with either an acid or a base catalyst, requiring temperatures as high as 150–240 °C. These elevated-temperature conditions result in the production of products of inferior quality (inappropriate for skin applications) that require additional treatments and expenses. Enzymatic processes offer a compelling solution to address these challenges, as they operate at lower temperatures (30–70 °C) and lower pressures, resulting in the creation of ultrapure, colorless, and odorless products. Esters produced through biocatalysis can be considered environmentally friendly, aligning with the growing consumer demand for “green” and “natural” products [59]. This is a primary driver behind the substantial number of publications on enzymatic cosmetic ingredient synthesis. Furthermore, the strong interest in implementing these processes on an industrial scale has motivated researchers to conduct applied research using reactors of a significant volume, along with the development of kinetic and mass transfer studies to facilitate the process scale-up.
These efforts have resulted in the commercialization of several cosmetic ingredients obtained via biocatalysis. Evonik Industries AG was the first in this field and currently offers five emulsifiers synthesized through biocatalytic processes: isoamyl cocoate (Tegosoft AC MB), cetyl ricinoleate (Tegosoft CR MB), decyl cocoate (Tegosoft DC MB), myristyl myristate (Tegosoft MM MB), and oleyl erucate (Tegosoft OER MB). The company website highlights that these esters have been “produced by an eco-efficient (enzymatic) process leading to a minimized environmental footprint” (https://personal-care.evonik.com, accessed on 2 January 2024). Afterward, the Eastman Company produced 2-ethylhexyl palmitate using GEM™ technology, which, according to the website (https://www.eastman.com, accessed on 2 January 2024), is “a bio-catalytic process that uses enzymes and closely controlled manufacturing conditions to eliminate high temperatures, strong acids, and unwanted by-products, consumes less energy compared with conventional manufacturing processes”. As far as we know, only these two companies have commercialized cosmetic ingredients using enzymatic processes.
In the WOS database, a search using the terms “ester synthesis” + “lipase” + “cosmetic” for the years 2000–2003 yielded 254 papers. As mentioned in the previous section, the tank reactor appears as the primary choice for the biocatalytic synthesis of cosmetic esters, although a search in WOS with the terms “ester synthesis” + “lipase” + “tank reactor” + “cosmetic” over the last 23 years yielded only five articles. To incorporate a greater number of studies, the search parameters were adjusted by removing the term “cosmetic”, resulting in 62 papers, of which 43 studies met the established criteria (tank reactors with a volume superior than 50 mL and cosmetic applications). Table 5 [60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102] compiles these papers. The esters synthesized fall into these categories: emollient esters [60,61,62,65,69,70,71,73,74,79,80,81,82,83,84,85,86,89,91,92,93,94,95,96,97,98,99,100,102], fragrant esters [63,76,77,78,87,90], sugar esters [64,67,75], and derivatives of active ingredients [66,68,72,88,101].
The rise in the use of biocatalytic synthesis to produce cosmetic esters in the industry has made commercial immobilized lipases the preferred choice for potential manufacturers and researchers. This guarantees a constant supply of enzymes. According to Table 5, among 43 papers reviewed, 36 report the use of commercial lipases from different companies: Novozymes [60,61,62,63,64,65,66,68,70,72,73,74,75,76,77,78,79,80,82,88,90,91,92,93,96,97,98,99,101], Fermenta Biotech Ltd. (Thane, India) [71,81,83,84,86,89,95], and Purolite (King of Prussia, PA, USA) [80]. In five articles, researchers describe the use of an in-lab immobilized lipase [61,69,85,94,100,102], and only two studies involve microbial cells with lipase activity [67,87]. It is evident that commercial immobilized enzymes come with a high cost. However, the immobilization processes also incur significant expenses, which can potentially surpass the cost of commercial enzymes. In one of the referenced papers, an economic analysis of the synthesis process for a mixture of wax esters similar to spermaceti (used in cosmetics for extremely cold climates) was conducted. The study demonstrated that the direct production costs of one gram of this product using in-lab immobilized lipase was comparable to that obtained using the commercial lipase Novozym® 435 [85], revealing that, after thorough optimization of the immobilization process, the option of using a lipase immobilized by the manufacturers themselves can be a valid alternative to commercial products, thus avoiding excessive dependence on certain production sectors.
In most of the papers compiled in Table 5, it can be observed that they chose to develop their investigations in large tank reactors, including 2 L [65] and 300 L [82], to speed up the industrial applicability of the investigations. Within them, studies are conducted to determine the most suitable agitator geometry and to address scale-up issues. For the same reason, most researchers choose batch reactors, which are easy to operate and provide very good results in terms of productivity. There are only two references in which continuous tank reactors were employed, and these make explicit mention of the procedure used to retain the solid particles of the immobilized enzyme inside the reactor: a stainless-steel basket [64] and a membrane [87]. Of particular importance are the papers describing the synthesis of cosmetic ingredients, mainly emollients, in batch reactors using ultrasound as an energy source [74,76,81,83,86,89,95]. On the other hand, the inhibitory effect of short-chain fatty acids (pKa ≤ 4.8) on lipase activity is once again shown in some of the biocatalytic processes, as previously noted in the production of esters for the food industry [26]. Thus, heptanoic [92] and caprylic acids [98] exert this inhibitory effect, which is avoided by using fed-batch reactors that maintain the acid concentration at an optimal level. The authors suggest using either successive acid additions [92] or continuous addition using a peristaltic pump [98]. In other cases, the fractional addition of alcohols, such as butanol and ethanol, is employed to prevent their potential inhibitory effect [79]. Finally, one study was found in which one of the substrates (maltose) was added in batches because of its partial solubility in the reaction medium [64].
As for the removal of water formed in the reactions, the procedures used are similar to those described in the previous section. These methods involve the use of molecular sieves [64,67,71,75,81], vacuum operation with dry N2 bubbling [66,69,70,72,73,80,85,91,93,97,100], and atmospheric evaporation [69,70,92,96,98,99]. Regarding the use of organic solvents as a reaction medium, solvent-free systems are becoming more prevalent [61,62,66,68,69,70,71,72,73,74,75,79,80,81,82,83,84,85,86,89,91,92,93,95,96,97,98,99,100,101,102]. This is probably due to the additional costs associated with the final separation and purification steps of the synthesized product, as well as the need to eliminate any trace of solvents from the final product, which could interfere with its use on the skin. Other alternative solvents, such as supercritical CO2 [60,63] and eutectic mixtures [101], have also been employed. In all cases, very high final conversions are reported, highlighting the feasibility of an enzymatic process for the synthesis of esters with cosmetic applications in tank reactors, even in cases where the obtained compounds have a high molecular weight, which could potentially complicate mass- and heat-transfer processes within the reactor.
It is evident that there is a high interest in implementing these processes on an industrial scale, which has led numerous researchers to expand their fundamental studies. This expansion involves not only using large-volume tanks but also developing kinetic models based on the mechanism of the studied reactions, which enables the design of the reactor for its future scale-up [62,65,71,74,77,78,81,84,88,90,94,95,97,102].
The high economic cost of such processes is perhaps the main drawback that opponents of biocatalytic synthesis cite as a reason not to pursue them as an alternative to traditional chemical routes. Therefore, economic studies have been conducted [85,93,98,99] demonstrating that the unit cost of these esters, when the immobilized lipase is appropriately reused, is comparable to that described in the literature for fine chemical compounds [103]. On the other hand, increasing environmental awareness is prompting manufacturers to implement sustainable processes that align with the principles of “Green Chemistry”.
However, the perception that biocatalysis is “environmentally friendly” and a “technologically robust approach from an industrial point of view” has not always been substantiated with convincing metrics. It is common to rely on somewhat empty claims regarding the ecological nature of a particular biocatalytic process [104]. For this reason, in recent years it has become popular to include the so-called “green metrics” in the development studies of biocatalytic processes, as can be seen in some of the most recent papers of those compiled in Table 5 [98,99,101]. On the other hand, only one paper [100] deals with the design and simulation of a plant for the production of an ester mixture, which, according to the authors, would produce 173.25 kg of product per working day with a purity of 99.55%. This is an avenue to be explored if these processes are to be successfully implemented on an industrial scale.
Among the 126 articles that WOS provides based on the searches “lipase” + “ester synthesis” + “packed bed reactor” and “lipase” + “ester synthesis “ + “fluidized bed reactor”, the manual screening has allowed for the selection of the 22 shown in Table 6 [105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120,121,122,123,124,125,126], which describe ester-synthesis processes with applications in cosmetics, using the packed-bed reactor in 18 of them and the fluidized bed reactor in only 4 of them, the latter corresponding to the most recent publications.
Esters synthesized in tubular reactors are used as emollients [106,108,109,113,114,115,117,120,121,122,126], fragrances [110,116,119,123], surfactants [105,107,112,118,124,125], or active ingredients [111]. The biocatalysts used are the same as those employed in the tank reactors. Noteworthy is the greater presence of in-lab immobilized lipases [110,114,118,119,120,121,125,126]. These data are important as they indicate that the authors are trying to introduce modifications to the traditional processes using the tank reactor, which is displayed not only in the attempt to avoid the use of commercial immobilized enzymes, but also in novel reactor configurations that include recirculation [107,114,119,120,124,126] or batch use with reuses [115,122]. A special configuration uses an additional column of molecular sieves to remove water from the reaction medium [110,114].
Undoubtedly, the most important factor to highlight in this group of papers is the high presence of studies that develop kinetic and mass transfer mathematical models, which are of great importance for scaling up [120,121,125,126]. In addition, there are also outstanding works in which the simulation of the industrial plant and an economic study are carried out based on data obtained on a pilot-plant scale [112,122].
During the bibliographical search, several papers have been found that are concerned with a comparative study of different types of reactors to select the most suitable one for the process under study. Table 7 shows these nine papers [127,128,129,130,131,132,133,134,135].
As can be observed, most of the papers point out that both packed and fluidized bed reactors give better results than tank reactors [127,128,129,130,132,134,135], highlighting the advantage of using tubular reactors, especially fluidized bed and bubble column reactors, since these configurations avoid the mechanical damage that stirrers can cause to the solid particles of immobilized lipase. It is very interesting to note a study that states that the best option for synthesizing kojic acid derivatives is the batch reactor, which is unusual for these type of studies [133]. In other reports, the results obtained in a conventional batch reactor are compared with those obtained in another reactor of the same geometry but equipped with ultrasound [131] or N2 bubbling [135]. Both give higher conversions than the batch reactor. Finally, it is also important to highlight the incorporation of microreactors in the processes for obtaining cosmetic esters, although the number of papers found is very small. Table 8 [136,137,138] shows the main characteristics of these reactors.
As can be seen, two articles describe the use of packed-bed microreactors [137,138] and the third one involves a membrane reactor with a single channel [136]. In all processes, good results are obtained showing the promising future of these type of reactors in the industry.

4. Conclusions

Esters are compounds of a diverse nature and structure, making them applicable in a wide variety of industrial segments. Among these, the food and cosmetic industries stand out. In recent years, these two productive sectors have become highly conscious of the use of high-purity compounds, not only to meet strict international regulations but also to satisfy the increasingly demanding preferences of consumers. In this context, biocatalysis emerges as an alternative that provides food and cosmetic product manufacturers with the indispensable esters to be used in their formulations. The mild operating conditions of enzymatic processes allow for the synthesis of high-purity compounds with almost the complete absence of undesired by-products. Additionally, biocatalytic synthesis aligns with many of the 12 principles of “Green Chemistry”, enabling products to be labeled as “natural”.
On the other hand, the significant presence of commercial immobilized lipases in the international market with high activity and stability has encouraged companies to incorporate biocatalytic processes into their production lines. However, to avoid dependence on the supply of immobilized enzymes, the lack of which could affect production, numerous studies in the literature explore innovative methods for lipase immobilization and their application in ester synthesis. The controversy surrounding the potentially excessive cost of preparing the biocatalyst compared to the high price of commercial immobilized enzymes also features prominently in the papers surveyed in this review.
It appears that manufacturers of food additive and cosmetic ingredients predominantly conduct their production using tank reactors, mostly in batch mode. For this reason, many studies aim to investigate the possibilities of applying other operation procedures (continuous) or even different reactor configurations. Numerous papers describe the successful performance of continuous reactors, both tank and packed-bed. Special mention should be made of attempts to incorporate fluidized bed reactors, which, in addition to providing satisfactory results, are particularly suitable for preventing the breakage of immobilized enzyme particles. Furthermore, in recent years, there has been emerging interest in the incorporation of new microreactors, although many studies will be necessary before their implementation on an industrial scale.
On the other hand, it is crucial to highlight the need for developing kinetic, mass-transfer, and reactor-design models, which are of decisive importance when designing and simulating a biocatalytic ester-synthesis plant. Additionally, there is a need to raise awareness among process engineers about the importance of conducting sustainability and economic studies, which are essential for the successful industrial-scale implementation of these production processes.
Finally, it is important to highlight that the introduction of artificial intelligence to the sustainable biocatalytic synthesis of esters holds significant promise. By leveraging artificial intelligence, various aspects of the manufacturing process can be optimized to improve efficiency, reduce the environmental impact, and enhance overall sustainability. AI can contribute to the design of eco-friendly processes, precise control of production parameters, and real-time monitoring, leading to more resource-efficient and environmentally friendly synthesis practices. Additionally, the integration of artificial intelligence can open new avenues for innovation and the development of novel solutions to address challenges in the sustainable production of esters for applications in various industries, such as food and cosmetics.

Author Contributions

Conceptualization, J.B. and F.M.; methodology, M.D.M. and M.G.; writing—original draft preparation, S.O.-R.; writing—review and editing, C.M.; funding acquisition, J.B. and F.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded with support from RTI 2018-094908-B-I00 MCIN/AEI/10.13039/501100011033/ and ERDF “A way of making Europe”.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

BRbatch reactor
CAGRcompound annual growth rate
CALBCandida antarctica lipase B
CRLCandida rugosa lipase
DESdeep eutectic solvent
FBRfluidized bed reactor
GRASgenerally recognized as safe
PBRpacked-bed reactor
WOSWeb of Science

References

  1. Wisniak, J. Alexander William Williamson. Educ. Quim. 2009, 7, 360–368. [Google Scholar] [CrossRef]
  2. Sirsam, R.; Hansora, D.; Usmani, G.A. A mini-review on solid acid catalysts for esterification reactions. J. Inst. Eng. India Ser. E 2016, 97, 167–181. [Google Scholar] [CrossRef]
  3. Jia, M.; Jiang, L.; Niu, F.; Zhang, Y.; Sun, X. A novel and highly efficient esterification process using triphenylphosphine oxide with oxalyl chloride. R. Soc. Open Sci. 2023, 5, 171988. [Google Scholar] [CrossRef] [PubMed]
  4. Montiel, M.C.; Maximo, F.; Serrano-Arnaldos, M.; Ortega-Requena, S.; Murcia, M.D.; Bastida, J. Biocatalytic solutions to cyclomethicones problem in cosmetics. Eng. Life Sci. 2019, 19, 370–388. [Google Scholar] [CrossRef] [PubMed]
  5. Chook, K.Y.; Aroua, M.K.; Gew, L.T. Enzyme biocatalysis for sustainability applications in reactors: A systematic review. Ind. Eng. Chem. Res. 2023, 62, 10800–10812. [Google Scholar] [CrossRef]
  6. Anastas, P.; Eghbali, N. Green chemistry: Principles and practice. Chem. Soc. Rev. 2010, 39, 301–312. [Google Scholar] [CrossRef] [PubMed]
  7. Bornscheuer, U.T. Lipase-catalyzed syntheses of monoacylglycerols. Enzym. Microb. Technol. 1995, 17, 578–586. [Google Scholar] [CrossRef]
  8. Klibanov, A.M. Improving enzymes by using them in organic solvents. Nature 2001, 409, 241–246. [Google Scholar] [CrossRef]
  9. Ortiz, C.; Lujan Ferreira, M.; Barbosa, O.; dos Santos, J.C.S.; Rodrigues, R.C.; Berenguer-Murcia, A.; Briand, L.E.; Fernandez-Lafuente, R. Novozym 435: The “perfect” lipase immobilized biocatalyst? Catal. Sci. Technol. 2019, 9, 2380. [Google Scholar] [CrossRef]
  10. Chandra, P.; Enespa; Singh, R.; Arora, P.K. Microbial lipases and their industrial applications: A comprehensive review. Microb. Cell Fact. 2020, 19, 169. [Google Scholar] [CrossRef]
  11. Moura, M.V.H.; da Silva, G.P.; de Oliveira Machado, A.C.; Torres, F.A.G.; Freire, D.M.G.; Almeida, R.V. Displaying lipase B from Candida antarctica in Pichia pastoris using the yeast surface display approach: Prospection of a new anchor and characterization of the whole cell biocatalyst. PLoS ONE 2015, 10, e0141454. [Google Scholar] [CrossRef] [PubMed]
  12. de Maria, L.; Vind, J.; Oxenbøll, K.M.; Svendsen, A.; Patkar, S. Phospholipases and their industrial applications. Appl. Microbiol. Biotechnol. 2007, 74, 290–300. [Google Scholar] [CrossRef] [PubMed]
  13. Remonatto, D.; Miotti, R.H.; Monti, R.; Bassan, J.C.; de Paula, A.V. Applications of immobilized lipases in enzymatic reactors: A review. Process Biochem. 2022, 114, 1–20. [Google Scholar] [CrossRef]
  14. Ragupathy, L.; Ziener, U.; Dyllick-Brenzinger, R.; von Vacano, B.; Landfester, K. Enzyme-catalyzed polymerizations at higher temperatures: Synthetic methods to produce polyamides and new poly(amide-co-ester)s. J. Mol. Catal. B Enzym. 2012, 76, 94–105. [Google Scholar] [CrossRef]
  15. Homann, M.J.; Vail, R.; Morgan, B.; Sabesan, V.; Levy, C.; Dodds, D.R.; Zaks, A. Enzymatic hydrolysis of a prochiral 3-substituted glutarate ester, an intermediate in the synthesis of an NK1/NK2 dual antagonist. Adv. Synth. Catal. 2001, 343, 744–749. [Google Scholar] [CrossRef]
  16. Lindeque, R.M.; Woodley, J. Reactor selection for effective continuous biocatalytic production of pharmaceuticals. Catalysts 2019, 9, 262. [Google Scholar] [CrossRef]
  17. Zhang, X.; Kobayashi, T.; Watanabe, Y.; Fuji, T.; Adachi, S.; Nakanishi, K.; Matsuno, R. Lipase-catalyzed synthesis of monolauroyl maltose through condensation of maltose and lauric acid. Food Sci. Technol. Res. 2003, 9, 110–113. [Google Scholar] [CrossRef]
  18. Sabeder, S.; Habulin, M.; Knez, Z. Comparison of the esterification of fructose and palmitic acid in organic solvent and in supercritical carbon dioxide. Ind. Eng. Chem. Res. 2005, 44, 9631–9635. [Google Scholar] [CrossRef]
  19. Keng, P.S.; Basri, M.; Ariff, A.B.; Abdul Rahman, M.B.; Abdul Rahman, R.N.Z.; Salleh, A.B. Scale-up synthesis of lipase-catalyzed palm esters in stirred-tank reactor. Bioresour. Technol. 2008, 99, 6097–6104. [Google Scholar] [CrossRef]
  20. Bodalo, A.; Bastida, J.; Maximo, M.F.; Montiel, M.C.; Murcia, M.D.; Ortega, S. Influence of the operating conditions on lipase-catalysed synthesis of ricinoleic acid estolides in solvent-free systems. Biochem. Eng. J. 2009, 44, 214–219. [Google Scholar] [CrossRef]
  21. Gomez, J.L.; Bastida, J.; Maximo, M.F.; Montiel, M.C.; Murcia, M.D.; Ortega, S. Solvent-free polyglycerol polyricinoleate synthesis mediated by lipase from Rhizopus arrhizus. Biochem. Eng. J. 2011, 54, 111–116. [Google Scholar] [CrossRef]
  22. Ortega-Requena, S.; Gómez, J.L.; Bastida, J.; Maximo, F.; Montiel, M.C.; Murcia, M.D. Study of different reaction schemes for the enzymatic synthesis of polyglycerol polyricinoleate. J. Sci. Food Agric. 2014, 94, 2308–2316. [Google Scholar] [CrossRef] [PubMed]
  23. Ortega-Requena, S.; Bodalo-Santoyo, A.; Bastida-Rodriguez, J.; Maximo-Martin, M.F.; Montiel-Morte, M.C.; Gomez-Gomez, M. Optimized enzymatic synthesis of the food additive polyglycerol polyricinoleate (PGPR) using Novozym® 435 in a solvent free system. Biochem. Eng. J. 2014, 84, 91–97. [Google Scholar] [CrossRef]
  24. dos Santos, P.; Zabot, G.L.; Meireles, M.A.A.; Mazutti, M.A.; Martinez, J. Synthesis of eugenyl acetate by enzymatic reactions in supercritical carbon dioxide. Biochem. Eng. J. 2016, 114, 1–9. [Google Scholar] [CrossRef]
  25. Martinez-Ruiz, A.; Tovar-Castro, L.; Garcia, H.S.; Saucedo-Castañeda, G.; Favela-Torres, E. Continuous ethyl oleate synthesis by lipases produced by solid-state fermentation by Rhizopus microsporus. Bioresour. Technol. 2018, 265, 52–58. [Google Scholar] [CrossRef]
  26. de Meneses, A.C.; Gomes Almeida Sa, A.; Lerin, L.A.; Corazza, M.L.; Hermes de Araujo, P.H.; Sayer, C.; de Oliveira, D. Benzyl butyrate esterification mediated by immobilized lipases: Evaluation of batch and fed-batch reactors to overcome lipase-acid deactivation. Process Biochem. 2019, 78, 50–57. [Google Scholar] [CrossRef]
  27. Kim, N.H.; Kim, H.; Choi, N.; Kim, Y.; Kim, B.H.; Kim, I.H. Production of stearidonic acid-rich triacylglycerol via a two-step enzymatic esterification. Food Chem. 2019, 270, 332–337. [Google Scholar] [CrossRef]
  28. Ortega-Requena, S.; Serrano-Arnaldos, M.; Montiel, M.C.; Maximo, F.; Bastida, J.; Murcia, M.D. Biocatalytic synthesis of polymeric esters used as emulsifiers. Chem. Biochem. Eng. Q. 2019, 33, 79–86. [Google Scholar] [CrossRef]
  29. Watanabe, Y.; Miyawaki, Y.; Adachi, S.; Nakanishi, K.; Matsuno, R. Continuous production of acyl mannoses by immobilized lipase using a packed bed reactor and their surfactant properties. Biochem. Eng. J. 2001, 8, 213–216. [Google Scholar] [CrossRef]
  30. Kuwabara, K.; Watanabe, Y.; Adachi, S.; Nakanishi, K.; Matsuno, R. Continuous production of acyl L-ascorbates using a packed bed reactor with immobilized lipase. J. Am. Oil Chem. Soc. 2003, 80, 895–898. [Google Scholar] [CrossRef]
  31. Piao, J.; Kobayashi, T.; Adachi, S.; Nakanishi, K.; Matsuno, R. Continuous synthesis of lauroyl or oleoyl erythritol by a packed bed reactor with an immobilized lipase. Process Biochem. 2004, 39, 681–686. [Google Scholar] [CrossRef]
  32. Piao, J.; Adachi, S. Enzymatic preparation of fatty acid esters of sugar alcohols by condensation in acetone using a packed bed reactor with immobilized Candida antarctica lipase. Biocatal. Biotransform. 2004, 22, 269–274. [Google Scholar] [CrossRef]
  33. Osorio, N.M.; Gusmao, J.H.; da Fonseca, M.M.; Ferreira-Dias, S. Lipase-catalysed interesterification of palm stearin with soybean oil in a continuous fluidised-bed reactor. Eur. J. Lipid Sci. Technol. 2005, 107, 455–463. [Google Scholar] [CrossRef]
  34. de Freitas, L.; dos Santos, J.C.; Zanin, G.M.; de Castro, H.F. Packed bed reactor running on Babassu oil and glycerol to produce monoglycerides by enzymatic route using immobilized Burkholderia cepacia lipase. Appl. Biochem. Biotechnol. 2010, 161, 372–381. [Google Scholar] [CrossRef] [PubMed]
  35. Rahman, N.K.; Kamaruddin, A.H.; Uzir, M.H. Enzymatic synthesis of farnesyl laurate in organic solvent: Initial water activity, kinetics mechanism, optimization of continuous operation using packed bed reactor and mass transfer studies. Bioprocess. Biosyst. Eng. 2011, 34, 687–699. [Google Scholar] [CrossRef] [PubMed]
  36. Escandell, J.; Wurm, D.J.; Belleville, M.P.; Sanchez, J.; Harasek, M.; Paolucci-Jeanjean, D. Enzymatic synthesis of butyl acetate in a packed bed reactor under liquid and supercritical conditions. Catal. Today 2015, 255, 3–9. [Google Scholar] [CrossRef]
  37. Paula, A.V.; Nunes, G.F.M.; Osorio, N.M.; Santos, J.C.; de Castro, H.F.; Ferreira-Dias, S. Continuous enzymatic interesterification of milkfat with soybean oil produces a highly spreadable product rich in polyunsaturated fatty acids. Eur. J. Lipid Sci. Technol. 2015, 117, 608–619. [Google Scholar] [CrossRef]
  38. Hidayat, C.; Fitria, K.; Supriyanto; Hastuti, P. Enzymatic synthesis of bio-surfactant fructose oleic ester using immobilized lipase on modified hydrophobic matrix in fluidized bed reactor. Agric. Agric. Sci. Procedia 2016, 9, 353–362. [Google Scholar] [CrossRef]
  39. Kirdi, R.; Akacha, N.B.; Messaoudi, Y.; Gargouri, M. Enhanced synthesis of isoamyl acetate using liquid-gas biphasic system by the transesterification reaction of isoamyl alcohol obtained from fusel oil. Biotechnol. Bioprocess. Eng. 2017, 22, 413–422. [Google Scholar] [CrossRef]
  40. Jia, C.; Wang, H.; Zhang, W.; Zhang, X.; Feng, B. Efficient enzyme-selective synthesis of monolauryl mannose in a circulating fluidized bed reactor. Process Biochem. 2018, 66, 28–32. [Google Scholar] [CrossRef]
  41. Salvi, H.M.; Kamble, M.P.; Yadav, G.D. Synthesis of geraniol esters in a continuous-flow packed bed reactor of immobilized lipase: Optimization of process parameters and kinetic modeling. Appl. Biochem. Biotechnol. 2018, 184, 630–643. [Google Scholar] [CrossRef] [PubMed]
  42. Dias, A.L.B.; Ubeyitogullari, A.; Hatami, T.; Martinez, J.; Ciftci, O.N. Continuous production of isoamyl acetate from fusel oil under supercritical CO2: A mass transfer approach. Chem. Eng. Res. Des. 2021, 176, 23–33. [Google Scholar] [CrossRef]
  43. Souza-Gonçalves, J.; Fialho, A.; Soares, C.M.F.; Ossrio, N.M.; Ferreira-Dias, S. Continuous production of dietetic structured lipids using crude acidic olive pomace oils. Molecules 2023, 28, 2637. [Google Scholar] [CrossRef] [PubMed]
  44. Ai, H.; Lee, Y.Y.; Xie, X.; Tan, C.P.; Lai, O.M.; Li, A.; Wang, Y.; Zhang, Z. Structured lipids produced from palm-olein oil by interesterification: A controllable lipase-catalyzed approach in a solvent-free system. Food Chem. 2023, 412, 135558. [Google Scholar] [CrossRef]
  45. Lozano, P.; Villora, G.; Gómez, D.; Gayo, A.B.; Sanchez-Conesa, J.A.; Rubio, M.; Iborra, J.L. Membrane reactor with immobilized Candida antarctica lipase B for ester synthesis in supercritical carbon dioxide. J. Supercrit. Fluids 2004, 29, 121–128. [Google Scholar] [CrossRef]
  46. Saponjic, S.; Knezevic-Jugovic, Z.D.; Bezbradica, D.I.; Zuza, M.G.; Saied, O.A.; Boskovic-Vragolovic, N.; Mijin, D.Z. Use of Candida rugosa lipase immobilized on sepabeads for the amyl caprylate synthesis: Batch and fluidized bed reactor study. Electron. J. Biotechnol. 2010, 13, 12–13. [Google Scholar] [CrossRef]
  47. Matte, C.R.; Bordinhao, C.; Poppe, J.K.; Rodrigues, R.C.; Hertz, P.F.; Ayub, M.A.Z. Synthesis of butyl butyrate in batch and continuous enzymatic reactors using Thermomyces lanuginosus lipase immobilized immobead 150. J. Mol. Catal. B Enzym. 2016, 127, 67–75. [Google Scholar] [CrossRef]
  48. dos Santos, P.; Meireles, M.A.A.; Martinez, J. Production of isoamyl acetate by enzymatic reactions in batch and packed bed reactors with supercritical CO2. J. Supercrit. Fluids 2017, 127, 71–80. [Google Scholar] [CrossRef]
  49. Delavault, A.; Opochenska, O.; Laneque, L.; Soergel, H.; Muhle-Goll, C.; Ochsenreither, K.; Syldatk, C. Lipase-catalyzed production of sorbitol laurate in a “2-in-1” deep eutectic system: Factors affecting the synthesis and scalability. Molecules 2021, 26, 2759. [Google Scholar] [CrossRef]
  50. Yan, Y.; Bornscheuer, U.T.; Stadler, G.; Lutz-Wahl, S.; Reuss, M.; Schmid, R.D. Production of sugar fatty acid esters by enzymatic esterification in a stirred-tank membrane reactor: Optimization of parameters by response surface methodology. J. Am. Oil Chem. Soc. 2001, 78, 147–153. [Google Scholar] [CrossRef]
  51. Yan, Y.; Bornscheuer, U.T.; Schmid, R.D. Efficient water removal in lipase-catalyzed esterifications using a low-boiling-point azeotrope. Biotechnol. Bioeng. 2002, 78, 31–34. [Google Scholar] [CrossRef]
  52. Antczak, T.; Patura, J.; Szczesna-Antczak, M.; Hiler, D.; Bielecki, S. Sugar ester synthesis by a mycelium-bound Mucor circinelloides lipase in a micro-reactor equipped with water activity sensor. J. Mol. Catal. B Enzym. 2004, 29, 155–161. [Google Scholar] [CrossRef]
  53. Woodcok, L.L.; Wiles, C.; Greenway, G.M.; Watts, P.; Wells, A.; Eyley, S. Enzymatic synthesis of a series of alkyl esters using Novozyme 435 in a packed bed, miniaturized, continuous flow reactor. Biocatal. Biotransform. 2008, 26, 501–507. [Google Scholar] [CrossRef]
  54. Bhavsar, K.V.; Yadav, G.D. n-Butyl levulinate synthesis using lipase catalysis: Comparison of batch reactor versus continuous flow packed bed tubular microreactor. J. Flow Chem. 2018, 8, 97–105. [Google Scholar] [CrossRef]
  55. Mathpati, A.C.; Kalghatgi, S.G.; Mathpati, C.S.; Bhanagea, B.M. Immobilized lipase catalyzed synthesis of n-amyl acetate: Parameter optimization, heterogeneous kinetics, continuous flow operation and reactor modeling. J. Chem. Technol. Biotechnol. 2018, 93, 2906–2916. [Google Scholar] [CrossRef]
  56. Holtheuer, J.; Tavernini, L.; Bernal, C.; Romero, O.; Ottone, C.; Wilson, L. Enzymatic synthesis of ascorbyl palmitate in a rotating bed reactor. Molecules 2023, 28, 644. [Google Scholar] [CrossRef] [PubMed]
  57. Enders, A.; Grunberger, A.; Bahnemann, J. Towards small scale: Overview and applications of microfluidics in biotechnology. Mol. Biotechnol. 2022. [Google Scholar] [CrossRef] [PubMed]
  58. Ansorge-Schumacher, M.B.; Thum, O. Immobilised lipases in the cosmetics industry. Chem. Soc. Rev. 2013, 42, 6475–6490. [Google Scholar] [CrossRef]
  59. Khan, N.R.; Rathod, V.K. Enzyme catalyzed synthesis of cosmetic esters and its intensification: A review. Process Biochem. 2015, 50, 1793–1806. [Google Scholar] [CrossRef]
  60. Laudani, C.G.; Habulin, M.; Kneza, Z.; Della Porta, G.; Reverchon, E. Lipase-catalyzed long chain fatty ester synthesis in dense carbon dioxide: Kinetics and thermodynamics. J. Supercrit. Fluids 2007, 41, 92–101. [Google Scholar] [CrossRef]
  61. Petersson, A.E.V.; Adlercreutz, P.; Mattiasson, B. A water activity control system for enzymatic reactions in organic media. Biotechnol. Bioeng. 2007, 97, 235–241. [Google Scholar] [CrossRef] [PubMed]
  62. Trubiano, G.; Borio, D.; Errazu, A. Influence of the operating conditions and the external mass transfer limitations on the synthesis of fatty acid esters using a Candida antarctica lipase. Enzym. Microb. Technol. 2007, 40, 716–722. [Google Scholar] [CrossRef]
  63. Habulin, M.; Sabeder, S.; Acebes Sampedro, M.; Knez, Z. Enzymatic synthesis of citronellol laurate in organic media and in supercritical carbon dioxide. Biochem. Eng. J. 2008, 42, 6–12. [Google Scholar] [CrossRef]
  64. Liu, Q.; Jia, C.; Kim, J.M.; Jiang, P.; Zhang, X.; Feng, B.; Xu, S. Lipase-catalyzed selective synthesis of monolauroyl maltose using continuous stirred tank reactor. Biotechnol. Lett. 2008, 30, 497–502. [Google Scholar] [CrossRef] [PubMed]
  65. Radzi, S.M.; Mohamad, R.; Basri, M.; Salleh, A.B.; Ariff, A.; Abdul Rahman, M.B.; Abdul Rahman, R.N.Z.R. Kinetics of enzymatic synthesis of liquid wax ester from oleic acid and oleyl alcohol. J. Oleo Sci. 2010, 59, 127–134. [Google Scholar] [CrossRef] [PubMed]
  66. El-Boulifi, B.; Ashari, S.E.; Serrano, M.; Aracil, J.; Martínez, M. Solvent-free lipase-catalyzed synthesis of a novel hydroxyl-fatty acid derivative of kojic acid. Enzym. Microb. Technol. 2014, 55, 128–132. [Google Scholar] [CrossRef] [PubMed]
  67. Guo, D.H.; Jin, Z.; Xu, Y.S.; Wang, P.; Lin, Y.; Han, S.Y.; Zheng, S.P. Scaling-up the synthesis of myristate glucose ester catalyzed by a CALB-displaying Pichia pastoris whole-cell biocatalyst. Enzym. Microb. Technol. 2015, 75–76, 30–36. [Google Scholar] [CrossRef]
  68. Jumbri, K.; Rozy, M.F.A.H.; Ashari, S.E.; Mohamad, R.; Basri, M.; Masoumi, H.E.F. Optimisation and characterisation of lipase- catalysed synthesis of a kojic monooleate ester in a solvent-free system by response surface methodology. PLoS ONE 2015, 10, e0144664. [Google Scholar] [CrossRef]
  69. Montiel, M.C.; Serrano, M.; Maximo, M.F.; Gomez, M.; Ortega-Requena, S.; Bastida, J. Synthesis of cetyl ricinoleate catalyzed by immobilized Lipozyme® CalB lipase in a solvent-free system. Catal. Today 2015, 255, 49–53. [Google Scholar] [CrossRef]
  70. Ortega, S.; Montiel, M.C.; Serrano, M.; Maximo, M.F.; Bastida, J. “100% natural” myristyl myristate. Study of the biocatalytic synthesis process. Afinidad 2015, 570, 155–159. [Google Scholar]
  71. Khan, N.; Jadhav, S.; Rathod, V.K. Enzymatic synthesis of n-butyl palmitate in a solvent-free system: RSM optimization and kinetic studies. Biocatal. Biotransform. 2016, 34, 99–109. [Google Scholar] [CrossRef]
  72. Mouad, A.M.; Taupin, D.; Lehr, L.; Yvergnaux, F.; Porto, A.L.M. Aminolysis of linoleic and salicylic acid derivatives with Candida antarctica lipase B: A solvent-free process to obtain amphiphilica amides for cosmetic application. J. Mol. Catal. B Enzym. 2016, 126, 64–68. [Google Scholar] [CrossRef]
  73. Serrano-Arnaldos, M.; Maximo-Martın, M.F.; Montiel-Morte, M.C.; Ortega-Requena, S.; Gomez-Gomez, E.; Bastida-Rodrıguez, J. Solvent-free enzymatic production of high quality cetyl esters. Bioprocess. Biosyst. Eng. 2016, 39, 641–649. [Google Scholar] [CrossRef] [PubMed]
  74. Tomke, P.D.; Rathod, V.K. Enzyme as biocatalyst for synthesis of octyl ethanoate using acoustic cavitation: Optimization and kinetic study. Biocatal. Agric. Biotechnol. 2016, 7, 145–153. [Google Scholar] [CrossRef]
  75. Ye, R.; Hayes, D.G.; Burton, R.; Liu, A.; Harte, F.M.; Wang, Y. Solvent-free lipase-catalyzed synthesis of technical-grade sugar esters and evaluation of their physicochemical and bioactive properties. Catalysts 2016, 6, 78. [Google Scholar] [CrossRef]
  76. Deshmukh, A.R.; Rathod, V.K. Intensification of enzyme catalysed synthesis of hexyl acetate using sonication. Green. Process Synth. 2017, 6, 55–62. [Google Scholar] [CrossRef]
  77. Hidalgo, A.M.; Sanchez, A.; Gomez, J.L.; Gomez, E.; Gomez, M.; Murcia, M.D. Kinetic study of the enzymatic synthesis of 2-phenylethyl acetate in discontinuous tank reactor. Ind. Eng. Chem. Res. 2018, 57, 11280–11287. [Google Scholar] [CrossRef]
  78. Murcia, M.D.; Gomez, M.; Gomez, E.; Gomez, J.L.; Hidalgo, A.M.; Sanchez, A.; Vergara, P. Kinetic modelling and kinetic parameters calculation in the lipase-catalysed synthesis of geranyl acetate. Chem. Eng. Res. Des. 2018, 138, 135–143. [Google Scholar] [CrossRef]
  79. Pereira, G.N.; Holz, J.P.; Giovannini, P.P.; Oliveira, J.V.; de Oliveira, D.; Lerin, L.A. Enzymatic esterification for the synthesis of butyl stearate and ethyl stearate. Biocatal. Agric. Biotechnol. 2018, 16, 373–377. [Google Scholar] [CrossRef]
  80. Serrano-Arnaldos, M.; Bastida, J.; Maximo, F.; Ortega-Requena, S.; Montiel, C. One-step solvent-free production of a spermaceti analogue using commercial immobilized lipases. ChemistrySelect 2018, 3, 748–752. [Google Scholar] [CrossRef]
  81. Khan, N.R.; Gawas, S.D.; Rathod, V.K. Enzyme-catalysed production of n-butyl palmitate using ultrasound assisted esterification of palmitic acid in a solvent-free system. Bioprocess. Biosys. Eng. 2018, 41, 1621–1634. [Google Scholar] [CrossRef]
  82. Basri, M.; Abdul Rahman, N.F.; Kassim, M.A.; Shahruzzaman, R.M.H.R.; Mokles, M.S.N. Lipase-catalyzed Production and Purification of Palm Esters Using Stirred Tank Reactors (STR). J. Oleo Sci. 2019, 68, 329–337. [Google Scholar] [CrossRef] [PubMed]
  83. Gawas, S.D.; Khan, N.; Rathod, V.K. Application of response surface methodology for lipase catalyzed synthesis of 2-ethylhexyl palmitate in a solvent free system using ultrasound. Braz. J. Chem. Eng. 2019, 36, 1007–1017. [Google Scholar] [CrossRef]
  84. Jaiswal, K.S.; Rathod, V.K. Enzymatic synthesis of cosmetic grade wax ester in solvent free system: Optimization, kinetic and thermodynamic studies. SN Appl. Sci. 2019, 1, 949. [Google Scholar] [CrossRef]
  85. Serrano-Arnaldos, M.; Ortega-Requena, S.; Montiel, M.C.; Maximo, F.; Bastida, J.; Murcia, M.D. Preliminary economic assessment: A valuable tool to establish biocatalytic process feasibility with an in-lab immobilized lipase. J. Chem. Technol. Biotechnol. 2019, 94, 409–417. [Google Scholar] [CrossRef]
  86. Khan, N.R.; Rathod, V.K. Enzymatic synthesis of cetyl oleate in a solvent free medium using microwave irradiation and physicochemical evaluation. Biocatal. Biotransform. 2020, 38, 114–122. [Google Scholar] [CrossRef]
  87. Perdomo, I.C.; Contente, M.L.; Pinto, A.; Romano, D.; Fernandes, P.; Molinari, F. Continuous preparation of flavour-active acetate esters by direct biocatalytic esterification. Flavour. Fragr. J. 2020, 35, 190–196. [Google Scholar] [CrossRef]
  88. Corovic, M.; Milivojevic, A.; Simovic, M.; Banjanac, K.; Pjanovic, R.; Bezbradica, D. Enzymatically derived oil-based L-ascorbyl esters: Synthesis, antioxidant properties and controlled release from cosmetic formulations. Sus. Chem. Phar. 2020, 15, 100231. [Google Scholar] [CrossRef]
  89. Gawas, S.D.; Rathod, V.K. Ultrasound assisted green synthesis of 2-ethylhexyl stearate: A cosmetic bio-lubricant. J. Oleo Sci. 2020, 69, 1043–1049. [Google Scholar] [CrossRef]
  90. Gomez, J.L.; Gomez, M.; Murcia, M.D.; Gomez, E.; Hidalgo, A.M.; Montiel, C.; Martinez, R. Biosynthesis of benzyl acetate: Optimization of experimental conditions, kinetic modelling and application of alternative methods for parameters determination. Bioresour. Technol. Rep. 2020, 11, 100519. [Google Scholar] [CrossRef]
  91. Murcia, M.D.; Serrano-Arnaldos, M.; Ortega-Requena, S.; Maximo, F.; Bastida, J.; Montiel, M.C. Optimization of a sustainable biocatalytic process for the synthesis of ethylhexyl fatty acids esters. Catal. Today 2020, 6346, 98–105. [Google Scholar] [CrossRef]
  92. Serrano-Arnaldos, M.; Garcia-Martinez, J.J.; Ortega-Requena, S.; Bastida, J.; Maximo, F.; Montiel, M.C. Reaction strategies for the enzymatic synthesis of neopentyl glycol diheptanoate. Enzym. Microb. Technol. 2020, 132, 109400. [Google Scholar] [CrossRef] [PubMed]
  93. Serrano-Arnaldos, M.; Montiel, M.C.; Ortega-Requena, S.; Maximo, F.; Bastida, J. Development and economic evaluation of an eco-friendly biocatalytic synthesis of emollient esters. Bioprocess. Biosys. Eng. 2020, 43, 495–505. [Google Scholar] [CrossRef] [PubMed]
  94. Silva, M.V.C.; Rosa, C.M.R.; Aguiar, L.G.; Oliveira, P.C.; de Castro, H.F.; Freitas, L. Synthesis of isopropyl palmitate by lipase immobilized on a magnetized polymer matrix. Chem. Eng. Technol. 2020, 43, 1741–1748. [Google Scholar] [CrossRef]
  95. Jaiswal, K.; Saraiya, S.; Rathod, V.K. Intensification of enzymatic synthesis of decyl oleate using ultrasound in solvent free system: Kinetic, thermodynamic and physicochemical study. J. Oleo Sci. 2021, 70, 559–570. [Google Scholar] [CrossRef]
  96. Montiel, M.C.; Asensi, M.; Gimeno-Martos, S.; Maximo, F.; Bastida, J. Sustainable biocatalytic procedure for obtaining new branched acid esters. Materials 2021, 14, 6847. [Google Scholar] [CrossRef]
  97. Serrano-Arnaldos, M.; Murcia, M.D.; Ortega-Requena, S.; Montiel, M.C.; Maximo, F.; Gomez, E.; Bastida, J. A simplified kinetic model to describe the solvent-free enzymatic synthesis of wax esters. J. Chem. Technol. Biotechnol. 2021, 96, 2325–2335. [Google Scholar] [CrossRef]
  98. Maximo, F.; Asensi, M.; Serrano-Arnaldos, M.; Ortega-Requena, S.; Montiel, C.; Bastida, J. Biocatalytic intensified process for the synthesis of neopentyl glycol dicaprylate/dicaprate. Sus. Chem. Phar. 2022, 30, 100882. [Google Scholar] [CrossRef]
  99. Montiel, C.; Gimeno-Martos, S.; Ortega-Requena, S.; Serrano-Arnaldos, M.; Maximo, F.; Bastida, J. Green production of a high-value branched-chain diester: Optimization based on operating conditions and economic and sustainability criteria. Appl. Sci. 2023, 13, 6177. [Google Scholar] [CrossRef]
  100. Montiel, M.C.; Serrano-Arnaldos, M.; Yagüe, C.; Ortega-Requena, S.; Maximo, F.; Bastida, J. Development of an industrial sustainable process for wax esters production: Enzyme immobilization, process optimization, and plant simulation. J. Chem. Technol. Biotechnol. 2023, 98, 2295–2304. [Google Scholar] [CrossRef]
  101. Nieto, S.; Bernal, J.M.; Villa, R.; Garcia-Verdugo, E.; Donaire, A.; Lozano, P. Sustainable setups for the biocatalytic production and scale-up of panthenyl monoacyl esters under solvent-free conditions. ACS Sustain. Chem. Eng. 2023, 11, 5737–5747. [Google Scholar] [CrossRef] [PubMed]
  102. Rangel, A.B.S.; Silva, M.V.C.; de Assis, G.P.; Rosa, C.M.R.; dos Santos, J.C.; de Freitas, L. Synthesis and characterization of magnetic poly(STY-EGDMA) particles for application as biocatalyst support in octyl oleate ester synthesis: Kinetic and thermodynamic parameters and mathematical modeling. Catal. Lett. 2023, 153, 3284–3296. [Google Scholar] [CrossRef]
  103. Tufvesson, P.; Lima-Ramos, J.; Nordblad, M.; Woodley, J.M. Guidelines and cost analysis for catalyst production in biocatalytic processes. Org. Process Res. Dev. 2011, 15, 266–274. [Google Scholar] [CrossRef]
  104. de Maria, P.D. Biocatalysis, sustainability, and industrial applications: Show me the metrics. Curr. Opin. Green. Sustain. Chem. 2021, 31, 100514. [Google Scholar] [CrossRef]
  105. Arcos, J.A.; Garcia, H.S.; Hill, C.G. Continuous enzymatic esterification of glycerol with (poly)unsaturated fatty acids in a packed bed reactor. Biotechnol. Bioeng. 2000, 68, 563–570. [Google Scholar] [CrossRef]
  106. Wehtje, E.; Costes, D.; Adlercreutz, P. Continuous lipase-catalyzed production of wax ester using silicone tubing. J. Am. Oil Chem. Soc. 1999, 76, 1489–1493. [Google Scholar] [CrossRef]
  107. Laszlo, J.A.; Compton, D.L.; Eller, F.J.; Taylor, S.L.; Isbell, T.A. Packed bed bioreactor synthesis of feruloylated monoacyl- and diacylglycerols: Clean production of a “green” sunscreen. Green Chem. 2003, 5, 382–386. [Google Scholar] [CrossRef]
  108. Chang, S.W.; Shaw, J.F.; Yang, C.K.; Shieh, C.J. Optimal continuous biosynthesis of hexyl laurate by a packed bed bioreactor. Process Biochem. 2007, 42, 1362–1366. [Google Scholar] [CrossRef]
  109. Ju, H.Y.; Yang, C.K.; Yen, Y.H.; Shieh, C.J. Continuous lipase-catalyzed synthesis of hexyl laurate in a packed bed reactor: Optimization of the reaction conditions in a solvent-free system. J. Chem. Technol. Biotechnol. 2009, 84, 29–33. [Google Scholar] [CrossRef]
  110. Serri, N.A.; Kamaruddin, A.H.; Len, K.Y.T. A continuous esterification of malonic acid with citronellol using packed bed reactor: Investigation of parameter and kinetics study. Food Bioprod. Process 2010, 88, 327–332. [Google Scholar] [CrossRef]
  111. Zhao, H.; Liu, J.; Lv, F.; Ye, R.; Bie, X.; Zhang, C.; Lu, Z. Enzymatic synthesis of lard-based ascorbyl esters in a packed bed reactor: Optimization by response surface methodology and evaluation of antioxidant properties. LWT Food Sci. Technol. 2014, 57, 393–399. [Google Scholar] [CrossRef]
  112. Compton, D.L.; Goodell, J.R.; Grallc, S.; Evans, K.O.; Cermak, S.C. Continuous, packed bed, enzymatic bioreactor production and stability of feruloyl soy glycerides. Ind. Crops Prod. 2015, 77, 787–794. [Google Scholar] [CrossRef]
  113. Ganguly, S.; Nandi, S. Process optimization of lipase catalyzed synthesis of diesters in a packed bed reactor. Biochem. Eng. J. 2015, 102, 2–5. [Google Scholar] [CrossRef]
  114. Shen, H.; Tao, Y.; Cui, C.; Zhang, Y.; Chen, B.; Tan, T. Synthesis of 2-ethyl hexanol fatty acid esters in a packed bed bioreactor using a lipase immobilized on a textile membrane. Biocatal. Biotransform. 2015, 33, 44–50. [Google Scholar] [CrossRef]
  115. Wan, F.L.; Teng, Y.L.; Wang, Y.; Li, A.J.; Zhang, N. Optimization of oligoglycerol fatty acid esters preparation catalyzed by Lipozyme 435. Grasas Aceites 2015, 66, e088. [Google Scholar] [CrossRef]
  116. Machado, J.R.; Pereira, G.N.; dos Santos de Oliveira, P.; Zenevicz, M.C.; Lerin, L.; Barreto de Oliveira, R.D.R.; Cavalcanti, S.C.H.; Ninow, J.L.; de Oliveira, D. Synthesis of eugenyl acetate by immobilized lipase in a packed bed reactor and evaluation of its larvicidal activity. Process Biochem. 2017, 58, 114–119. [Google Scholar] [CrossRef]
  117. Zenevicz, M.C.P.; Jacques, A.; Silva, M.J.A.; Furigo, A.; Oliveira, V.; de Oliveira, D. Study of a reactor model for enzymatic reactions in continuous mode coupled to an ultrasound bath for esters production. Bioprocess. Biosys. Eng. 2018, 41, 1589–1597. [Google Scholar] [CrossRef]
  118. Shahrin, N.A.; Yong, G.G.; Serri, N.A. Fructose stearate esterify in packed bed reactor using immobilized lipase. IOP Conf. Ser. Mat. Sci. Eng. 2020, 716, 012017. [Google Scholar] [CrossRef]
  119. Wang, L.; Chen, G.; Tang, J.; Ming, M.; Jia, C.; Feng, B. Continuous biosynthesis of geranyl butyrate in a circulating fluidized bed reactor. Food Biosci. 2019, 27, 60–65. [Google Scholar] [CrossRef]
  120. Silva, M.V.C.; Rangel, A.B.S.; Aguiar, L.G.; de Castro, H.F.; de Freitas, L. Continuous enzymatic synthesis of 2-ethylhexyl oleate in a fluidized bed reactor: Operating conditions, hydrodynamics, and mathematical modeling. Ind. Eng. Chem. Res. 2020, 59, 19522–19530. [Google Scholar] [CrossRef]
  121. Silva, M.V.C.; Souza, A.B.; de Castro, H.F.; Aguiar, L.G.; de Oliveira, P.C.; de Freitas, L. Synthesis of 2-ethylhexyl oleate catalyzed by Candida antarctica lipase immobilized on a magnetic polymer support in continuous flow. Bioprocess. Biosys. Eng. 2020, 43, 615–623. [Google Scholar] [CrossRef] [PubMed]
  122. 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]
  123. Huang, S.M.; Huang, H.Y.; Chen, Y.M.; Kuo, C.H.; Shieh, C.J. Continuous production of 2-phenylethyl acetate in a solvent-free system using a packed bed reactor with Novozym® 435. Catalysts 2020, 10, 714. [Google Scholar] [CrossRef]
  124. Hollenbach, R.; Muller, D.; Delavault, A.; Syldatk, C. Continuous flow glycolipid synthesis using a packed bed reactor. Catalysts 2022, 12, 551. [Google Scholar] [CrossRef]
  125. Vilas Boas, R.M.; Lima, R.; Silva, M.V.C.; de Freitas, L.; Aguiar, L.G.; de Castro, H.F. Continuous production of monoacylglycerol via glycerolysis of babassu oil by immobilized Burkholderia cepacia lipase in a packed bed reactor. Bioprocess. Biosys. Eng. 2021, 44, 2205–2215. [Google Scholar] [CrossRef] [PubMed]
  126. Silva, M.V.C.; Rangel, A.B.S.; Rosa, C.M.R.; de Assis, G.P.; Aguiar, L.G.; de Freitas, L. Development of a magnetically stabilized fluidized bed bioreactor for enzymatic synthesis of 2-ethylhexyl oleate. Bioprocess. Biosys. Eng. 2023, 46, 1665–1676. [Google Scholar] [CrossRef] [PubMed]
  127. Bousquet, M.P.; Willemot, R.M.; Monsan, P.; Boures, E. Enzymatic Synthesis of α-butylglucoside linoleate in a packed bed reactor for future pilot scale-up. Biotechnol. Prog. 2000, 16, 589–594. [Google Scholar] [CrossRef]
  128. Hilterhaus, L.; Thum, O.; Liese, A. Reactor concept for lipase-catalyzed solvent-free conversion of highly viscous reactants forming two-phase systems. Org. Process Res. Dev. 2008, 12, 618–625. [Google Scholar] [CrossRef]
  129. Couto, R.; Vidinha, P.; Peres, C.; Ribeiro, A.S.; Ferreira, O.; Oliveira, M.V.; Macedo, E.A.; Loureiro, J.M.; Barreiros, S. Geranyl acetate synthesis in a packed bed reactor catalyzed by novozym in supercritical carbon dioxide and in supercritical ethane. Ind. Eng. Chem. Res. 2011, 50, 1938–1946. [Google Scholar] [CrossRef]
  130. Damnjanovic, J.J.; Zuza, M.G.; Savanovic, J.K.; Bezbradica, D.I.; Mijin, D.Z.; Boskovic-Vragolovic, N.; Knezevic-Jugovic, Z.D. Covalently immobilized lipase catalyzing high-yielding optimized geranyl butyrate synthesis in a batch and fluidized bed reactor. J. Mol. Catal. B Enzym. 2012, 75, 50–59. [Google Scholar] [CrossRef]
  131. Khan, N.R.; Jadhav, S.V.; Rathod, V.K. Lipase catalysed synthesis of cetyl oleate using ultrasound: Optimisation and kinetic studies. Ultrason. Sonochem. 2015, 27, 522–529. [Google Scholar] [CrossRef] [PubMed]
  132. Corovic, M.; Milivojevic, A.; Carevic, M.; Banjanac, K.; Tanaskovic, S.J.; Bezbradica, D. Batch and semicontinuous production of l-ascorbyloleate catalyzed by CALB immobilized onto Purolite® MN102. Chem. Eng. Res. Des. 2017, 126, 161–171. [Google Scholar] [CrossRef]
  133. Lajis, A.F.B.; Hamid, M.; Ahmad, S.; Ariff, A.B. Comparative study of stirred and fluidized tank reactor for hydroxyl-kojic acid derivatives synthesis and their biological activities. Turk. J. Biochem. 2018, 43, 205–219. [Google Scholar] [CrossRef]
  134. Vilas Boas, R.N.; Ceron, A.A.; Bento, H.B.S.; de Castro, H.F. Application of an immobilized Rhizopus oryzae lipase to batch and continuous ester synthesis with a mixture of a lauric acid and fusel oil. Biomass Bioenergy 2018, 119, 61–68. [Google Scholar] [CrossRef]
  135. Satyawali, Y.; Cauwenberghs, L.; Dejonghe, W. Lipase-catalyzed solvent-free synthesis of polyglycerol 10 (PG-10) esters. Chem. Biochem. Eng. Q. 2019, 33, 501–509. [Google Scholar] [CrossRef]
  136. Carvalho, C.M.L.; Aires-Barros, M.R.; Cabral, J.M.S. A continuous membrane bioreactor for ester synthesis in organic media: I. Operational characterization and stability. Biotechnol. Bioeng. 2001, 72, 127–135. [Google Scholar] [CrossRef]
  137. Cvjetkoa, M.; Vorkapic-Furac, J.; Znidarsic-Plaz, P. Isoamyl acetate synthesis in imidazolium-based ionic liquids using packed bed enzyme microreactor. Process Biochem. 2012, 47, 1344–1350. [Google Scholar] [CrossRef]
  138. Giovannini, P.P.; Catani, M.; Massi, A.; Sacchetti, G.; Tacchini, M.; de Oliveira, D.; Lerin, L.A. Continuous production of eugenol esters using enzymatic packed bed microreactors and an evaluation of the products as antifungal agents. Flavour. Fragr. J. 2019, 34, 201–210. [Google Scholar] [CrossRef]
Table 1. Biocatalytic synthesis of esters with applications in the food industry using tank reactors.
Table 1. Biocatalytic synthesis of esters with applications in the food industry using tank reactors.
EsterBiocatalystCharacteristicsReference
Monolauroyl maltoseChirazyme® L-2 C2
immobilized Candida antarctica lipase B
Batch and continuous stirred tank reactors
Volume: 300 mL
Immobilized lipase packed into a stainless-steel basket
Solvent: acetone
Water removal: molecular sieves
Conversion: 60% after 90 h
[17]
Fructose
palmitate
Novozym® 435
immobilized Candida antarctica lipase B
Batch reactor
Volume: 100 mL
Solvent: 2-methyl-2-butanol, supercritical CO2
Water removal: molecular sieves
Conversion: 78% after 72 h
[18]
Oleyl palm
ester
Lipozyme® RM IM
immobilized
Rhizomucor miehei lipase
Batch reactor with different impellers
Volume: 2 L and scale up to 50 L
Solvent: n-hexane
Conversion: 97.2% after 5 h
[19]
Ricinoleic acid
estolides
Candida rugosa lipase in-lab
immobilized in Lewatit
MonoPlusMP64
Batch reactor
Volume: 100 mL
Solvent: solvent-free
Water removal: atmospheric evaporation and vacuum (comparison)
Conversion: 68% after 24 h
[20]
Polyglycerol polyricinoleateCandida rugosa, Rhizopus arrhizus, and Rhizopus oryzae
lipases in-lab immobilized in Lewatit MonoPlusMP64
Batch reactor (two steps)
Volume: 100 mL
Solvent: solvent-free
Water removal: vacuum
Conversion: 91.5% after 125 h, 98% after 320 h
[21,22]
Polyglycerol polyricinoleateNovozym® 435
immobilized Candida antarctica lipase B
Batch reactor with different impellers
Volume: 100 mL
Solvent: solvent-free
Water removal: vacuum and dry N2 bubbling
Conversion: 99.3% after 55 h
[23]
Eugenyl
acetate
Lipozyme® 435 and
Novozym® 435
immobilized Candida antarctica lipase B
Batch reactor
Volume: 100 mL
Solvent: supercritical CO2
Kinetic model
Conversion: 45% after 6 h
[24]
Ethyl
oleate
Dry biocatalyst of supported Rhizopus microsporus with
lipase activity
Continuous stirred tank reactor
Volume: 700 mL
Solvent: n-hexane
Water removal: molecular sieves
Conversion: 90% after 14 h
[25]
Benzyl
butyrate
Novozym® 435
immobilized Candida antarctica lipase B
Lipozyme® TL-IM
immobilized Thermomyces lanuginosus lipase
Lipozyme® RM IM
immobilized Rhizomucor miehei lipase NS 88011
Non-commercial immobilized Candida antarctica lipase B
Batch and fed-batch reactors
Volume: 500 mL
Solvent: solvent-free
Water removal: molecular sieves
Conversion: 80% after 12 h
[26]
Stearidonic acid-rich
triacylglycerol
Novozym® 435
immobilized Candida antarctica lipase B
Lipozyme® TL-IM
immobilized Thermomyces
lanuginosus lipase
Batch reactor (two steps)
Volume: 50 mL
Water removal: vacuum
Conversion: 86.4% after 12 h
[27]
Polyglycerol polyricinoleateCandida antarctica lipase B
in-lab immobilized in
Lewatit MonoPlusMP64
Batch reactor
Volume: 100 mL
Solvent: solvent-free
Water removal: vacuum and dry N2 bubbling
Conversion: 98% after 159 h
[28]
Table 2. Biocatalytic synthesis of esters with applications in the food industry using tubular reactors.
Table 2. Biocatalytic synthesis of esters with applications in the food industry using tubular reactors.
EsterBiocatalystCharacteristicsReference
Acyl
mannoses
Chirazyme® L-2 C2
immobilized Candida antarctica lipase B
Packed-bed reactor (continuous)
10 mm i.d. × 50 mm
Residence time: 12 min
Solvent: acetonitrile, acetone, 2-methyl-2-propanol, 2-methyl-2-butanol
Conversion: 40% for 16 days
[29]
Acyl L-
ascorbates
Chirazyme® L-2 C2
immobilized Candida antarctica lipase B
Packed-bed reactor (continuous)
4.6 mm i.d. × 150 mm
Residence time: 5 min
Solvent: acetone
Productivity: 1.6–1.9 kg/L for 11 days
[30]
Lauroyl and oleoyl
erythritol
Chirazyme® L-2 C2
immobilized Candida antarctica lipase B
Packed-bed reactor (continuous)
10 mm i.d. × 50 mm
Residence time: 4.5 min
Solvent: acetone
Productivity: 1.25–1.6 kg/L for 14 days
[31]
Fatty acid esters of sugar
alcohols
Chirazyme® L-2 C2
immobilized Candida antarctica lipase B
Packed-bed reactor (continuous)
20 mm i.d. × 50 mm
Residence time: 15 min
Solvent: acetone
Productivity: 1.3–2 kg/L for 2 days
[32]
Esters of palm stearin with soybean oilNovozym® 435
immobilized Candida antarctica lipase B
Fluidized bed reactor (continuous)
2 cm i.d. × 20 cm
Residence time: 19 min
Solvent: solvent-free
Conversion: 10–45% for 21 days
[33]
Monoglycerides of
Babassu oil
Lipase PS—Batch number: 01022TD
in-lab immobilized
Burkholderia cepacia lipase
Packed-bed reactor (continuous)
1.5 cm i.d. × 5.5 cm
Residence time: 356 min
Solvent: solvent-free
Conversion: 25–33% for 22 days
[34]
Farnesyl laurateLipozyme® RM IM
immobilized Rhizomucor miehei lipase
Packed-bed reactor (continuous)
1.2 cm i.d × 9.24 cm
Residence time: 22 min
Solvent: iso-octane
Kinetic and mass transfer model
Conversion: 98.07% for 3 h
[35]
Butyl
acetate
Candida antarctica lipase B
in-lab immobilized in porous
γ-alumina pellets
Packed-bed reactor (continuous)
12 g biocatalyst
Flow rate: 0.5–10 mL/min
Solvent: n-hexane, supercritical CO2 (comparison)
Productivities: 119 µmol/min × g pellets and 501 µmol/min × g pellets
[36]
Esters of milkfat with soybean oilRhizopus oryzae lipase in-lab
immobilized in polysiloxane–
polyvinyl alcohol particles
Novozym® 435
immobilized Candida antarctica lipase B
Fluidized bed reactor (recirculating and continuous)
20 mm i.d. × 200 mm
Residence time: 12 min and 6 min
Solvent: solvent-free
Conversion: 52% and 27% for 190 h
[37]
Fructose oleic esterCandida rugosa lipase in-lab
immobilized in modified
Amberlite IRA-96
Fluidized bed reactor (batch recirculating)
10 mm i.d. × 160 mm
Residence time: 42.78–213.91 min
Solvent: solvent-free
Conversion: 197.06% (mixture of mono- di- and tri-
esters) 15 cycles
[38]
Isoamyl
acetate
Aspergillus oryzae lipase in-lab immobilized in calcium
alginate beads
Gas-liquid fluidized bed reactor (continuous)
0.8 mm i.d. × 143 mm
Continuous ethanol removal: N2 flow
Solvent: solvent-free
Conversion: 89.55% for 60 min
[39]
Monolauroyl maltoseNovozym® 435
immobilized Candida antarctica lipase B
Fluidized bed reactor (recirculating)
10 mm i.d. × 300 mm
Flow rate: 1 mL/min
Solvent: acetone
Conversion: 30% for 5 days
[40]
Geraniol
esters
Novozym® 435
immobilized Candida antarctica lipase B
Lipozyme® TL-IM
immobilized Thermomyces
lanuginosus lipase
Lipozyme® RM IM
immobilized Rhizomucor miehei lipase
Packed-bed reactor (continuous)
3.0 mm i.d. × 100 mm
Residence time: 5–25 min
Solvent: n-heptane
Kinetic model
Conversion: 87% for 25 h
[41]
Isoamyl
acetate
Novozym® 435
immobilized Candida antarctica lipase B
Packed-bed reactor (continuous)
8 mm i.d. × 200 mm
Residence time: 36.5 min
Solvent: supercritical CO2
Mathematical model
Conversion: 95.5%
[42]
Structured
lipids from
olive oil
Lipozyme® TL-IM
immobilized Thermomyces
lanuginosus lipase
Lipozyme® RM IM
immobilized Rhizomucor miehei lipase
Packed-bed reactor (continuous)
2 cm i.d. × 20 cm
Residence time: 10.9 and 20 min
Solvent: solvent-free
Conversion: 70% for 70 h
[43]
Structured
lipids from palm-olein
Lipozyme® TL-IM
immobilized Thermomyces
lanuginosus lipase
Packed-bed reactor (continuous)
30 mm i.d. × 48 cm
Solvent: solvent-free
Lipid composition
[44]
Table 3. Biocatalytic synthesis of esters with applications in the food industry. Comparison of different reactors.
Table 3. Biocatalytic synthesis of esters with applications in the food industry. Comparison of different reactors.
EsterBiocatalystCharacteristicsReference
Butyl
butyrate
Novozym® 435
immobilized Candida antarctica lipase B
Batch reactor and membrane reactor (recirculating)
Volume BR: 400 mL
Volume MR: 175 mL
Solvent: supercritical CO2
Selectivity: ≥99% after 3 h BR
≥99% after 7 cycles of 6 h MR
Better results with MR
[45]
Amyl
caprylate
Candida rugosa lipase in-lab
immobilized on
Sepabeads EC-EP
Batch reactor and fluidized bed reactor (recirculating)
Volume BR: 10 mL
Volume FBR: 80 mL, 10 mm i.d. × 136 mm, residence time: 3.53–0.75 min
Solvent: isooctane
Water removal: molecular sieves
Conversion: ≥99% after 24 h BR
90.2% for 70 h FBR
Better results with FBR
[46]
Butyl
butyrate
Thermomyces lanuginosus lipase (TLL) in-lab immobilized on Immobead 150Batch reactor, packed-bed reactor, packed-bed reactor with glass beads and fluidized bed reactor (continuous)
Volume BR: 10 mL
PBR: 10 mm i.d. × 65 mm
Solvent: n-hexane
Conversion: 21%, 85% and 60%
Better results with PBR with glass beads
[47]
Isoamyl
acetate
Novozym® 435
immobilized Candida antarctica lipase B
Batch reactor and packed-bed reactor (continuous)
Volume BR: 100 mL
PBR: 3.37 mm i.d. × 0.33 m, residence time: 5.6–11 s
Solvent: supercritical CO2
Mathematical model (mass transfer, kinetic and reactor design)
Better results with PBR
[48]
Sorbitol
laurate
10 immobilized lipases
6 lipases in solution
Batch reactor and “tube-reactor” (discontinuous in orbital shaker)
Volume BR: 500 mL
Conversion: 28% after 48 h BR
50% after 48 h TR
[49]
Table 4. Biocatalytic synthesis of esters with applications in the food industry using other reactor configurations.
Table 4. Biocatalytic synthesis of esters with applications in the food industry using other reactor configurations.
EsterBiocatalystCharacteristicsReferences
Sugar fatty
acid esters
Chirazyme® L-2
immobilized Candida antarctica lipase B
Continuous stirred membrane tank reactor
Membrane area: 23 cm2.
Bottom part for the reaction 58 mm i.d. × 2 mm
Solvent: ethyl methyl ketone, n-hexane
Water removal: azeotrope and membrane evaporation
Conversion: 93% after 48 h
[50,51]
Sugar estersMycelium-bound
Mucor circinelloides lipase
Batch microreactor with water activity sensor
Volume: 37 mL
Solvent: di-n-pentyl and petroleum ethers
Water activity influence
Conversion: 72% after 20 min
[52]
Alkyl estersNovozym® 435
Immobilized Candida antarctica lipase B
Packed-bed miniaturized reactor (continuous)
1.65 mm i.d. × 30 mm
100 mg lipase, flow rate of 1 µL/min
Solvent: n-hexane
Conversion: 92% for 2 h
[53]
n-Butyl
levulinate
Lipozyme® RM IM
immobilized Rhizomucor miehei lipase
Lipozyme® TL-IM
immobilized Thermomyces
lanuginosus lipase
Novozym® 435
immobilized Candida antarctica lipase B
Packed-bed microreactor (continuous)
3 mm i.d. × 100 mm
Residence time: 1–5 min
Solvent: tert-butyl methyl ether, 1,4 dioxane, acetonitrile, toluene
Conversion: 85% for 25 h (6 runs)
[54]
n-Amyl
acetate
Burkholderia cepacia lipase
in-lab immobilized on a
biodegradable polymer
Coated film microreactor (batch and continuous)
Volume: 18 mL
Solvent: n-hexane
Mathematical model (dispersion model)
Productivities: 8.16 and 6.54 mmol/g h
[55]
Ascorbyl
palmitate
Lipozyme® 435
immobilized Candida antarctica lipase B
Lipozyme® TL-IM
immobilized Thermomyces
lanuginosus lipase
Lipozyme® RM IM
immobilized Rhizomucor miehei lipase
Lipozyme® Novo 40086
immobilized Rhizomucor miehei lipase
Amano lipase PS
immobilized Burkholderia
cepacia lipase
Batch reactor: rotating basket, sequential batches
Volume: 500 mL
Solvent: 2-methyl-2-butanol
Water removal: molecular sieves
Conversion: 80% each batch (4 batches)
[56]
Table 5. Biocatalytic synthesis of esters with applications in the cosmetic industry using tank reactors.
Table 5. Biocatalytic synthesis of esters with applications in the cosmetic industry using tank reactors.
EsterBiocatalystCharacteristicsReference
n-Octyl
oleate
Lipozyme® RM IM
immobilized Rhizomucor miehei lipase
Batch reactor
Volume: 102 mL
Solvent: supercritical CO2
Conversion: 88% after 5 h
[60]
Cetyl
palmitate
Novozym® 435
immobilized Candida antarctica lipase B
Candida rugosa lipase in-lab
immobilized in MP 1000
Batch reactor
Volume: 0.6 L
Solvent: solvent-free
Water activity measurement and control
Conversion: 73% after 192 h
[61]
Ethyl
oleate
Novozym® 435
immobilized Candida antarctica lipase B
Batch reactor
Volume: 250 mL
Solvent: solvent-free
Kinetic model
Conversion: 90% after 5.5 h
[62]
Citronellol laurateNovozym® 435
immobilized Candida antarctica lipase B
SP 382 immobilized Candida antarctica lipase B
Lipozyme® RM IM
immobilized Rhizomucor miehei lipase
Batch reactor
Volume: 100 mL
Solvent: n-heptane, supercritical CO2
Conversion: 74% after 5 h
[63]
Monolauryl
maltose
Novozym® 435
immobilized Candida antarctica lipase B
Continuous stirred tank reactor
Lipase in a stainless-steel basket
Volume: 250 mL
Solvent: acetone
Water removal: molecular sieves (addition)
Successive maltose addition due to insolubility
Recycling lauric acid and solvent
Productivity: 9.2 g/d L reactor for 10 days
[64]
Oleyl
oleate
Novozym® 435
immobilized Candida antarctica lipase B
Batch reactor (different agitators)
Volume: 2 L
Solvent: hexane
Kinetic model
Conversion: >90% after 1 h (Rushton turbine)
[65]
Kojic acid
ricinoleate
Novozym® 435
immobilized Candida antarctica lipase B
Lipozyme® RM IM
immobilized Rhizomucor miehei lipase
Lipozyme® TL-IM
immobilized Thermomyces
lanuginosus lipase
Batch reactor
Volume: 500 mL
Solvent: solvent-free
Water removal: vacuum
Conversion: 87.4% after 6 h
[66]
Fatty acid
glucose ester
Candida antarctica lipase B
displaying-Pichia pastoris strain GS115/CALB-GCW21-42
Batch reactor
Volume: 5 mL, 2 L and 5 L
Solvent: different organic solvents
Water removal: molecular sieves
Conversion: 90% after 96 h
[67]
Kojic acid
monooleate
Novozym® 435
immobilized Candida antarctica lipase B
Batch reactor
Volume: 125 mL
Solvent: solvent-free
Water removal: atmospheric evaporation
Conversion: 44.46% after 5 h
[68]
Cetyl
ricinoleate
Candida antarctica lipase B
in-lab immobilized in
Lewatit MonoPlusMP64
Batch reactor
Volume: 50 mL and 100 mL
Solvent: solvent-free
Water removal: atmospheric evaporation and
vacuum with dry N2 bubbling
Conversion: 98% after 3 h
[69]
Myristyl
myristate
Novozym® 435
immobilized Candida antarctica lipase B
Batch reactor
Volume: 50 mL and 100 mL
Solvent: solvent-free
Water removal: atmospheric evaporation and
vacuum with dry N2 bubbling
Conversion: 99% after 2 h
[70]
n-Butyl
palmitate
Fermase CALB 10000
immobilized Candida antarctica lipase B
Batch reactor
Volume: 250 mL
Solvent: solvent-free
Water removal: molecular sieves
Kinetic model
Conversion: 91.25% after 4 h
[71]
Amphiphilic
amides
Novozym® 435
immobilized Candida antarctica lipase B
Batch reactor
Volume: 50 mL
Solvent: solvent-free
Ethanol removal: vacuum
Conversion: 99% after 20 h
[72]
Cetyl fatty
acid esters
Novozym® 435
immobilized Candida antarctica lipase B
Batch reactor
Volume: 100 mL
Solvent: solvent-free
Water removal: vacuum with dry N2 bubbling
Conversion: 98.5% after 1.5 h
[73]
Octyl
ethanoate
Novozym® 435
immobilized Candida antarctica lipase B
Batch reactor with ultrasound
Volume: 50 mL
Solvent: solvent-free
Kinetic model
Conversion: 97.31% after 20 min
[74]
Oleic acid
sugar esters
Novozym® 435
immobilized Candida antarctica lipase B
Batch reactor
Volume: 100 mL
Solvent: solvent-free
Water removal: molecular sieves
Conversion: 96.6% after 6 days
[75]
Hexyl
acetate
Lipozyme® RM IM
immobilized Rhizomucor miehei lipase
Batch reactor with ultrasound
Volume: 50 mL
Solvent: hexane
Conversion: 85% after 4 h
[76]
2-Phenylethyl acetateNovozym® 435
immobilized Candida antarctica lipase B
Batch reactor
Volume: 50 mL
Solvent: hexane
Kinetic model
Conversion: 95.42% after 2 h
[77]
Geranyl
acetate
Novozym® 435
immobilized Candida antarctica lipase B
Batch reactor
Volume: 50 mL
Solvent: hexane
Kinetic model
Conversion: 98.4% after 160 min
[78]
Butyl stearate ethyl stearateNovozym® 435
immobilized Candida antarctica lipase B
Batch and fedbatch reactors
Volume: 250 mL
Solvent: solvent-free
Conversion: 92% after 24 h
[79]
Spermaceti
analogue
Lipozyme® RM IM
immobilized Rhizomucor miehei lipase
CalB immo Plus
immobilized Candida antarctica lipase B
Batch reactor
Volume: 100 mL
Solvent: solvent-free
Water removal: vacuum with dry N2 bubbling
Conversion: 98% after 2 h
[80]
n-Butyl
palmitate
Fermase CALB 10000
immobilized Candida antarctica lipase B
Batch reactor with ultrasound
Volume: 100 mL
Solvent: solvent-free
Water removal: molecular sieves
Kinetic model
Conversion: 96.6% after 50 min
[81]
Palm oil
esters
Lipozyme® RM IM
immobilized Rhizomucor miehei lipase
Lipozyme® TL-IM
immobilized Thermomyces
lanuginosus lipase
Batch reactor (scale-up)
Volume: 2 L, 15 L, and 300 L
Solvent: n-hexane and solvent-free
Conversion: >90% after 3 h
[82]
2-Ethylhexyl palmitateFermase CALB 10000
immobilized Candida antarctica lipase B
Batch reactor with ultrasound
Volume: 50 mL
Solvent: solvent-free
Conversion: 96.56% after 2 h
[83]
Cetyl
caprate
Fermase CALB 10000
immobilized Candida antarctica lipase B
Batch reactor
Volume: 50 mL
Solvent: solvent-free
Kinetic model
Conversion: 95% after 80 min
[84]
Spermaceti
analogue
Candida antarctica lipase B
in-lab immobilized in
different supports
Batch reactor
Volume: 100 mL
Solvent: solvent-free
Economic study
Water removal: vacuum with dry N2 bubbling
Conversion: >90% after 1 h
[85]
Cetyl
oleate
Fermase CALB 10000
immobilized Candida antarctica lipase B
Batch reactor with ultrasound
Volume: 50 mL
Solvent: solvent-free
Conversion: 97.5% after 20 min
[86]
Isoamyl
and cinnamyl acetate
Lyophilized mycelium of
Aspergillus oryzae
Continuous stirred tank membrane reactor
Volume: 200 mL
Residence time: 500 min
Solvent: n-heptane
Conversion: 98% for 10 days
[87]
Fatty acid ascorbyl
esters
Novozym® 435
immobilized Candida antarctica lipase B
Batch reactor
Volume: 100 mL
Solvent: organic solvent
Kinetic model
Conversion: high conversion depending on the fatty acid
[88]
2-Ethylhexyl stearateFermase CALB 10000
immobilized Candida antarctica lipase B
Batch reactor with ultrasound
Volume: 50 mL
Solvent: solvent-free
Conversion: 95.87% after 3 h
[89]
Benzyl
acetate
Novozym® 435
immobilized Candida antarctica lipase B
Novozym® 40086
immobilized Rhizomucor miehei lipase
Lipozyme® TL-IM
immobilized Thermomyces
lanuginosus lipase
Batch reactor
Volume: 50 mL
Solvent: n-hexane and n-heptane
Kinetic model
Conversion: >90% after 2.5 h
[90]
2-Ethylhexyl
palmitate and stearate
Novozym® 435
immobilized Candida antarctica lipase B
Novozym® 40086
immobilized Rhizomucor miehei lipase
Batch reactor
Volume: 100 mL
Solvent: solvent-free
Water removal: vacuum with dry N2 bubbling
Conversion: 98% after 45 min
[91]
Neopentyl
glycol
diheptanoate
Novozym® 435
immobilized Candida antarctica lipase B
Batch and fed-batch reactors
Volume: 50 mL
Solvent: solvent-free
Water removal: atmospheric evaporation
Conversion: 95% after 6 h
[92]
Spermaceti
analogue
Novozym® 435
immobilized Candida antarctica lipase B
Lipozyme® TL-IM
immobilized Thermomyces
lanuginosus lipase
Batch reactor
Volume: 100 mL
Solvent: solvent-free
Economic study
Water removal: vacuum with dry N2 bubbling
Conversion: >97.5% after 2 h
[93]
Isopropyl
palmitate
Penicillium camemberti lipase
in-lab immobilized on
magnetized poly(styrene-
codivinylbenzene)
Batch reactor
Volume: 280 mL
Solvent: heptane
Kinetic model
Conversion: 85.65% after 12 h
[94]
Decyl
oleate
Fermase CALB 10000
immobilized Candida antarctica lipase B
Batch reactor with ultrasounds
Volume: 50 mL
Solvent: solvent-free
Kinetic model
Conversion: 97.14% after 25 min
[95]
2-Ethylhexyl
2-methylhexanoate
Novozym® 435
immobilized Candida antarctica lipase B
Batch reactor
Volume: 50 mL
Solvent: solvent-free
Economic study and green metrics
Water removal: atmospheric evaporation
Conversion: 99.74% after 5 h
[96]
Spermaceti
analogue
Novozym® 435
immobilized Candida antarctica lipase B
Lipozyme® RM IM
immobilized Rhizomucor miehei lipase
Lipozyme® TL-IM
immobilized Thermomyces
lanuginosus lipase
CalB immo Plus
immobilized Candida antarctica lipase B
Batch reactor
Volume: 100 mL
Solvent: solvent-free
Kinetic model
Water removal: vacuum with dry N2 bubbling
Conversion: >90% after 1 h
[97]
Neopentyl
glycol
dicaprylate/
dicaprate
Lipozyme® 435
immobilized Candida antarctica lipase B
Batch and fed-batch reactors
Volume: 50 mL
Solvent: solvent-free
Economic study and green metrics
Water removal: atmospheric evaporation
Conversion: 92.5% after 6 h
[98]
Neopentyl
glycol
dilaurate
Novozym® 435
immobilized Candida antarctica lipase B
Novozym® 40086
immobilized Rhizomucor miehei lipase
Batch reactor
Volume: 50 mL
Solvent: solvent-free
Economic study and green metrics
Water removal: atmospheric evaporation
Conversion: >90% after 6 h
[99]
Spermaceti
analogue
Candida antarctica lipase B
in-lab immobilized in
Purolite® Lifetech™
ECR8285
Batch reactor
Volume: 100 mL
Solvent: solvent-free
Water removal: vacuum with dry N2 bubbling
Conversion: 97% after 1 h
Production plant simulation using aspenONE suite v10
[100]
Panthenyl monoacyl
ester
Novozym® 435
immobilized Candida antarctica lipase B
Batch reactor
Volume: 500 mL
Solvent: solvent-free (eutectic mixture)
Green metrics
Conversion: 87−95% after 6 h
[10]
Octyl
oleate
Candida antarctica lipase B
in-lab immobilized on
magnetic poly(STY-EGDMA) particles
Batch reactor
Volume: 100 mL
Solvent: solvent-free
Kinetic model
Conversion: 57% after 24 h
[102]
Table 6. Biocatalytic synthesis of esters with applications in the cosmetic industry using tubular reactors.
Table 6. Biocatalytic synthesis of esters with applications in the cosmetic industry using tubular reactors.
EsterBiocatalystCharacteristicsReference
Mono-, di-,
and triacyglycerols
from (poly)unsaturated fatty acids
Chirazyme® L-9
immobilized Mucor miehei
lipase
Packed-bed reactor (continuous)
0.32 cm i.d. × 20 cm
0.47 cm i.d. × 8.9 cm
0.63 cm i.d. × 5 cm
0.79 cm i.d. × 3.2
Residence time: 15 min
Solvent: hexane, 2-propanol, ethyl acetate, formic acid
Conversion: 80–90% for 12 days
[105]
Cetyl
palmitate
SP 435 immobilized
Candida antarctica lipase B
Packed-bed reactor (continuous)
Silicone and PVC tube: 3 mm i.d.
Flow rate: 0.005 g/min
Productivity: 7.2 g/day
Conversion: 99.1% for 7 days
[106]
Feruloylated monoacyl-
and
diacyl
glycerols
Novozym® 435
immobilized Candida antarctica lipase B
Packed-bed reactor (recirculating)
2.5 cm i.d. × 30 cm
Flow rate: 2 mL/min
Solvent: solvent-free
Water removal: molecular sieves
Conversion: 60% after 140 h
[107]
Hexyl
laurate
Lipozyme® IM-77
immobilized Rhizomucor miehei lipase
Packed-bed reactor (continuous)
0.25 cm i.d. × 25 cm
Residence time: 0.43 min
Solvent: n-hexane
Conversion: 97%
[108]
Hexyl
laurate
Lipozyme® IM-77
immobilized Rhizomucor miehei lipase
Packed-bed reactor (continuous)
0.25 cm i.d. × 25 cm
Flow rate: 0.55 mL/min
Solvent: solvent-free
Production rate: 87.44 μmol/min
[109]
Citronellyl
malonate
Candida rugosa lipase
in-lab immobilized on
Amberlite MB-1
Packed-bed reactor (continuous)
1.2 cm i.d. × 24 cm
Flow rate: 1 mL/min
Solvent: iso-octane
Kinetic model
Water removal: molecular sieves
Conversion: 90% (steady state after 180 min)
[110]
Lard-based ascorbyl
esters
Novozym® 435
immobilized Candida antarctica lipase B
Packed-bed reactor (continuous)
2 cm i.d. × 10 cm or 25 cm
Flow rate: 0.07 mL/min
Solvent: tert-amyl alcohol
Water removal: molecular sieves
Conversion: 50.50%
[111]
Feruloyl soy glyceridesNovozym® 435
immobilized Candida antarctica lipase B
Packed-bed reactor (continuous, pilot scale)
Four 304-stainless steel columns 9.8 cm i.d. × 132 cm
Flow rate: 2.5 mL/min
Solvent: solvent-free
Conversion: 65% for 4.5 months
[112]
Dibehenyl adipate
Dibehenyl
sebacate
Lipozyme® RM IM
immobilized Rhizomucor miehei lipase
Lipozyme® TL-IM
immobilized Thermomyces
lanuginosus lipase
SP 435 immobilized
Candida antarctica lipase B
NS40013 immobilized
Candida antarctica lipase B
Packed-bed reactor (continuous)
1 in i.d. × 12 in
Flow rate: 3 mL/min
Solvent: isooctane
Water removal: vacuum
Conversion: 89% and 91% for 5 h; 20 reuses
[113]
2-Ethylhexyl palmitateCandida sp. 99–125 lipase
in-lab immobilized
on a fabric membrane
Packed-bed reactor (recirculating)
40, 60, 90 mm i.d. × 630, 280, 124 mm
Residence time: 160 s
Solvent: solvent-free
Study of H/D influence on conversion
Water removal: molecular sieves
Conversion: 95% for 300 h (30 batches)
[114]
Polyglycerol fatty acid
esters
Lipozyme® 435
immobilized Candida antarctica lipase B
Bubble column reactor (batches)
Volume: 2 L
Solvent: solvent-free
Water removal: vacuum and N2 bubbling
Conversion: 95.82% for 4.25 h (10 batches)
[115]
Eugenyl
acetate
Lipozyme® TL-IM
immobilized Thermomyces
lanuginosus lipase
Packed-bed reactor (continuous)
15 mm i.d. × 55 mm
Residence time: 55, 7, and 4 min
Solvent: solvent-free
Conversion: 93.1%
[116]
Soybean-free fatty acids
ethyl esters
Novozym® 435
immobilized Candida antarctica lipase B
Lipozyme® TL-IM
immobilized Thermomyces
lanuginosus lipase
Packed-bed reactor with ultrasound (continuous)
14.5 mm i.d. × 171 mm
Flow rate: 2.5 mL/min
Conversion: 95% at 6 min residence time
[117]
Fructose
stearate
Rhizomucor miehei lipase
in-lab immobilized into
chicken eggshells
Packed-bed reactor (continuous)
10 mm i.d.× 90 mm
Flow rate: 0.074 mL/min
Solvent: ethanol
Product concentration: 7.252 × 10−1 mol/L
[118]
Geranyl
butyrate
Candida rugosa lipase
in-lab immobilized on
different resins
Fluidized bed reactor (recirculating)
1.4 cm i.d. × 17 cm
Flow rate: 0.07 mL/min
Residence time: 4.7 h
Solvent: n-heptane
Water removal: molecular sieves
Conversion: 77% for 12 h
[119]
2-Ethylhexyl
oleate
Candida antarctica lipase
in-lab immobilized on
STY-DVB-M particles
Fluidized bed reactor (continuous with recirculation)
15 mm i.d. × 202 mm
Residence time: 6, 12, and 18 h
Solvent: solvent-free
Mathematical model (kinetic and mass transfer)
Conversion: 48.24% for 8 days
[120]
2-Ethylhexyl
oleate
Candida antarctica lipase
in-lab immobilized on
STY-DVB-M particles
Packed-bed reactor (continuous)
11 mm i.d. × 166 mm
Residence time: 3, 6, and 12 h
Solvent: solvent-free
Kinetic model
Conversion: 60% for 16 days
[121]
2-Ethylhexyl
oleate
Novozym® 435
immobilized Candida antarctica lipase B
Packed-bed reactor (semicontinuous)
12 mm i.d. × 300 mm
Flow rate: 1.5 mL/min
Solvent: solvent-free
Water removal: molecular sieves
Economic study and process plant simulation
Conversion: >95% for 12 cycles × 720 h each
[122]
2-Phenylethyl acetateNovozym® 435
immobilized Candida antarctica lipase B
Packed-bed reactor (continuous)
0.46 cm i.d. × 25 cm
Flow rate: 1, 3, and 5 mL/min
Solvent: solvent-free
Conversion: 100% for 10 min (lower flow rate)
[123]
Glucose
mono
decanoate
Novozym® 435
immobilized Candida antarctica lipase B
Packed-bed reactor (continuous recycling glucose)
XK16 column from Cytiva
Flow rate: 0.5 mL/min
Residence time: 13 min
Productivity: 1228 µmol/L h
[124]
Monoacyl
glycerols of
Babassu oil
Burkholderia cepacia lipase
in-lab immobilized on
SiO2–PVA particles
Packed-bed reactor (continuous)
15 mm i.d. × 55 mm
Residence time: 9.8 h
Mathematical model (mass transfer)
Productivity: 52.3 mg/g h
[125]
2-Ethylhexyl
oleate
Candida antarctica lipase
in-lab immobilized on
STY-DVB-M particles
Fluidized bed reactor magnetically stabilized (continuous recycling substrate)
15 mm i.d. × 202 mm
Flow rate: 0.044 mL/min
Solvent: solvent-free
Residence time: 12 h
Kinetic model
Conversion: 55.63% for 16 days
[126]
Table 7. Biocatalytic synthesis of esters with applications in the cosmetic industry. Comparison of different reactors.
Table 7. Biocatalytic synthesis of esters with applications in the cosmetic industry. Comparison of different reactors.
EsterBiocatalystCharacteristicsReference
α-Butylglucoside linoleateChirazyme® L-2 C2
immobilized Candida antarctica
lipase B
Chirazyme® L-9
immobilized Mucor miehei
lipase
Batch reactor and packed-bed reactor (recirculating with mixing tank)
BR: rotary evaporator (Büchi, R-114)
PBR: 20 cm i.d. × 150 cm, flow rate: 4.5 mL/min
Solvent: decane
Water removal: vacuum
Conversion: >90% for >5 cycles × 70 h each
Better results with PBR and Chirazyme L-9
[127]
Myristyl myristateNovozym® 435
immobilized Candida antarctica lipase B
Batch reactor, packed-bed reactor, and bubble column reactor
Solvent: solvent-free
Water removal: vacuum
Mathematical model (kinetic and mass transfer)
Conversion: 99.6% after 5.5 h bubble column reactor
after 17 h packed-bed reactor
after 24 h batch reactor
Better results with bubble column reactor
[128]
Geranyl
acetate
Novozym® 435
immobilized Candida antarctica lipase B
Batch reactor, packed-bed reactor, and PBR series
PBR: 3.2 mm i.d. × different lengths
Solvent: supercritical CO2 and supercritical ethane
Mathematical model (kinetic and reactor design)
Batter results with two reactors in series
[129]
Geranyl
butyrate
Candida rugosa lipase
in-lab immobilized on
Sepabeads® EC-EP,
Sepabeads® EC-HA and
Purolite® A-109
Batch reactor and fluidized bed reactor (recirculating)
Volume BR: 100 mL
PBR: 10 mm i.d. × 136 mm
Solvent: isooctane
Mathematical model (hydrodynamic)
Conversion: >99.9% after 48 h BR
78.9% after 10 h FBR
Better results with FBR
[130]
Cetyl
oleate
Fermase CALB 10000
immobilized Candida antarctica lipase B
Batch reactor and batch reactor with ultrasound
Volume BR: 250 mL
Volume BR ultrasound: 50 mL
Solvent: solvent-free
Water removal: molecular sieves
Kinetic model
Conversion: 95.96% after 2 h BR
95.96% after 30 min BR with ultrasound
Better results with BR (stirred) with ultrasound
[131]
Ascorbyl
oleate
Candida antarctica lipase
in-lab immobilized on
Purolite® MN102
Batch reactor and fluidized bed reactor (recirculating)
Volume BR: 5 mL
FBR: 9 mm i.d. × 136 mm
Solvent: tert-butanol
Water removal: molecular sieves
Mathematical model (kinetic and hydrodynamic)
Better results with FBR (no damage of particles)
[132]
Kojic acid
derivatives
with fatty
acids
Novozym® 435
immobilized Candida antarctica lipase B
Lipozyme® RM IM
immobilized Rhizomucor miehei lipase
Lipozyme® TL-IM
immobilized Thermomyces
lanuginosus lipase
Batch reactor and fluidized tank reactor
Volume BR: 100 mL
FBR: The same BR sparged with air
Solvent: solvent-free
Better results with BR
[133]
Isoamyl
laurate
Five microbial lipases
In-lab immobilized on
epoxy-polysiloxane-hydroxyethylcellulose and
styrene-divinylbenzene
Batch reactor and packed-bed reactor (continuous)
Volume BR: 20 mL
PBR: 15 mm i.d. × 55 mm; flow rate: 1.8 mL/h; residence time: 3.12 h
Solvent: solvent-free
Conversion: 81.26% after 24 h BR
0.8 mol/L h PBR for 168 h
Better results with PBR
[134]
Polyglycerol-10 laurate
Polyglycerol-10 caprylate
Novozym® 435
immobilized Candida antarctica lipase B
Batch reactor (with mechanical agitation and bubbling) and fluidized bed reactor (batch operation)
Volume BR: Duran bottle (unknown volume)
FBR: 20 mm i.d. × 35 cm
Solvent: solvent-free
Water removal: dry N2 bubbling
Conversion: ≈100% after 20 h and 22 h
Better results with N2 bubbling
[135]
Table 8. Biocatalytic synthesis of esters with applications in the cosmetic industry using other reactor configurations.
Table 8. Biocatalytic synthesis of esters with applications in the cosmetic industry using other reactor configurations.
EsterBiocatalystCharacteristicsReference
Hexyl
acetate
Fusarium solani pisi
cutinase cloned and
expressed
in Escherichia coli
Membrane reactor stainless-steel monochannel
ultrafiltration module (continuous)
Volume: 100 mL
Membrane area: 38 cm2
Homogenization achieved through partial recirculation
Flow rate: 0.1 mL/min
Solvent: iso-octane (reversed micelles)
Mathematical model (reactor design)
Good performance of the MR
[136]
Isoamyl
acetate
Novozym® 435
immobilized Candida antarctica lipase B
Packed-bed microreactor (microchannel, continuous)
1 cm width × 450 µm height × 75 mm length
Solvent: ionic liquid
Conversion: 92% in 15 min (multiple runs for 2 weeks)
[137]
Eugenyl
esters
Novozym® 435
immobilized Candida antarctica lipase B
Lipozyme® RM IM
immobilized Rhizomucor miehei lipase
Packed-bed microreactor (continuous)
0.5 cm i.d. × 5 cm
Solvent: solvent-free
Conversion: 82% N435 and 90% RM IM for 26 h (acetate)
[138]
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

Ortega-Requena, S.; Montiel, C.; Máximo, F.; Gómez, M.; Murcia, M.D.; Bastida, J. Esters in the Food and Cosmetic Industries: An Overview of the Reactors Used in Their Biocatalytic Synthesis. Materials 2024, 17, 268. https://doi.org/10.3390/ma17010268

AMA Style

Ortega-Requena S, Montiel C, Máximo F, Gómez M, Murcia MD, Bastida J. Esters in the Food and Cosmetic Industries: An Overview of the Reactors Used in Their Biocatalytic Synthesis. Materials. 2024; 17(1):268. https://doi.org/10.3390/ma17010268

Chicago/Turabian Style

Ortega-Requena, Salvadora, Claudia Montiel, Fuensanta Máximo, María Gómez, María Dolores Murcia, and Josefa Bastida. 2024. "Esters in the Food and Cosmetic Industries: An Overview of the Reactors Used in Their Biocatalytic Synthesis" Materials 17, no. 1: 268. https://doi.org/10.3390/ma17010268

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