Catalysis Research is an international peer-reviewed Open Access journal published quarterly online by LIDSEN Publishing Inc. This periodical is devoted to publishing high-quality papers that describe the most significant and cutting-edge research in all areas of catalysts and catalyzed reactions. Its aim is to provide timely, authoritative introductions to current thinking, developments and research in carefully selected topics.

Topics contain but are not limited to:

  • Photocatalysis
  • Electrocatalysis
  • Environmental catalysis
  • Biocatalysis, enzymes, enzyme catalysis
  • Catalysis for biomass conversion
  • Organocatalysis, catalysis in organic and polymer chemistry
  • Nanostructured Catalysts
  • Catalytic materials
  • Computational catalysis
  • Kinetics of catalytic reactions

It publishes a variety of article types: Original Research, Review, Communication, Opinion, Comment, Conference Report, Technical Note, Book Review, etc.

There is no restriction on paper length, provided that the text is concise and comprehensive. Authors should present their results in as much detail as possible, as reviewers are encouraged to emphasize scientific rigor and reproducibility.

Publication Speed (median values for papers published in 2023): Submission to First Decision: 4.1 weeks; Submission to Acceptance: 13.0 weeks; Acceptance to Publication: 9 days (1-2 days of FREE language polishing included)

Current Issue: 2024  Archive: 2023 2022 2021
Open Access Review

Lipases: Sources of Acquisition, Ways of Production, and Recent Applications

Myllena R. Santos 1,†, Daniela B. Hirata 2,†, Joelise A. F. Angelotti 2,‡,*

  1. Postgraduation in Biotechnology, Federal University of Alfenas, Gabriel Monteiro da Silva Street, 700. Center, Alfenas 37130-001, MG, Brazil

  2. Institute of Chemistry, Federal University of Alfenas, Gabriel Monteiro da Silva Street, 700. Center, Alfenas 37130-001, MG, Brazil

‡ Current Affiliation: Institute of Chemistry, Federal University of Alfenas

† These authors contributed equally to this work.

Correspondence: Joelise A. F. Angelotti

Academic Editors: Pedro Fernandes and Sandra Aparecida de Assis

Special Issue: Development of Enzymatic and Whole Cell Based Processes Towards the Production of Added Value Goods from Renewable Resources

Received: February 16, 2022 | Accepted: April 26, 2022 | Published: May 07, 2022

Catalysis Research 2022, Volume 2, Issue 2, doi:10.21926/cr.2202013

Recommended citation: Santos MR, Hirata DB, Angelotti JAF. Lipases: Sources of Acquisition, Ways of Production, and Recent Applications. Catalysis Research 2022;2(2):35; doi:10.21926/cr.2202013.

© 2022 by the authors. This is an open access article distributed under the conditions of the Creative Commons by Attribution License, which permits unrestricted use, distribution, and reproduction in any medium or format, provided the original work is correctly cited.

Abstract

Enzymes are extensively used in biotechnological processes in several areas of industry. They are sustainable and safe, and their specificity is another characteristic that improves the performance in the process. Among enzymes, lipase is relevant due to the ability to play different roles in the industry and the possibility of collecting them from microbial sources that are found in industrial residues. This can reduce the costs of enzyme production. In relation to that, lipase immobilization is an interesting process that allows the enzymes to be reused and improves enzyme robustness. Among them, the cross-linked enzyme aggregates (CLEAs) methodology is attractive due to its simplicity, low cost (given the absence of support), and greater interaction with the substrate. Thus, in this review, we discussed the potential of lipase. We reviewed the traditional and new sources of obtaining lipases, along with the ways of improving production, activity, and application in the industry.

Keywords

Biocatalyst; Industrial wastes; immobilization; lipase; filamentous fungi

1. Introduction

Enzymes are globular proteins that participate in chemical reactions, making the process faster. They reduce the activation energy, which is essential for the reaction to occur; therefore, enzymes are commonly applied in biotechnological procedures since these molecules have important characteristics, such as specificity and biodegradability, and they can operate in easily controllable environments [1,2].

In the industrial sector, enzymes are widely used. Besides being sustainable and safe, their application guarantees better performance, providing products with greater functionality and higher yields [3]. Enzymes are used in several fields because of their wide applicability and specificity [4]. The global market of enzymes has been expanding strongly, and in 2019 it was estimated to be $9.9 billion, with a compound annual growth rate of 7.1% from 2020 to 2027 [5].

Hydrolases are a class of enzymes widely applied in the industrial field, with amylases, proteases, and lipases being the most used enzymes [6]. Among the enzymes that are popular in the industrial market, microbial lipases are quite prominent since they have great biotechnological potential and numerous applications, thus attracting higher interest commercially [7].

Lipases are proteins that act in aqueous and non-aqueous systems, in which a phenomenon known as interfacial activation occurs. In the presence of water, lipases catalyze triglyceride hydrolysis to generate molecules of glycerol and fatty acids. However, in media with low water activity, lipases perform reverse reactions, such as esterification and transesterification [8,9]. The main reactions catalyzed by lipases can be observed in Figure 1, Figure 2 and Figure 3.

Click to view original image

Figure 1 Hydrolysis.

Click to view original image

Figure 2 Esterification.

Click to view original image

Figure 3 Transesterification.

During metabolic transformations, lipases are extensively used because of their characteristics, such as specificity, availability, no requirement for the use of cofactors, and the ability to operate in large substrate concentrations [10]. Furthermore, they may have regioselectivity and enantioselectivity [11]. Lipases work in open and closed conformations; this mechanism of action of the enzyme is controlled by a kind of lid that hides the catalytic site of the enzyme. A microhydration condition at the active site of the enzyme is necessary for it to move the lid and expose its active site for catalysis [12]. However, in some enzymes, such as the lipase B of Candida antarctica, the catalytic site is blocked only partially since its lid is small [13]. Lipase B from Candida antarctica adopts a larger opening in aqueous conditions compared to other lipases [14].

Given their robustness, lipases can also be reused, which reduces the cost. The enzymes used need to be recovered for reuse. Recovering enzymes is a very complex process when the enzymes are in their soluble form. To avoid this problem, enzymes might be used in their immobilized form [15].

In this review, we discussed the traditional and alternative sources of lipases, as well as their most recent applications in the pharmaceutical and food industry, among others. The bibliographic review contextualized the most relevant aspects of lipases, providing an overview of the studies that have been conducted, from their production to application. We performed a bibliometric analysis, where the ScienceDirect, PubMed Central, National Library of Medicine, Capes Periodicals, and Google Scholar databases were searched. In this study, we discussed the potential of the lipase enzyme, highlighting its most recent applications, which might serve as a reference for future studies.

2. Traditional Sources

Lipases can be found in several sources and are produced by animals, plants, and microorganisms, including bacteria, fungi, and yeasts. With an increase in the use of this enzyme, it is necessary to discover new sources to supply industrial needs [16]. Industries are particularly interested in lipases of microbial origin which have several favorable characteristics, including resistance to high temperatures, action in a wide pH range, stability in organic solvents, and higher productivity compared to the lipases from other sources [17].

2.1 Plant Lipases

Plant lipases can be found in several types of plants, being present in seeds, grains, fruits, and leaves. However, most lipases are extracted from seeds, especially because they have a greater potential to perform hydrolysis compared to those from the other parts. These sources of lipases have a low production cost and are highly stable in the interesterification processes [18].

The enzymes found in plants have advantages compared to microbial enzymes since their production requires simpler methodologies and does not demand the use of genetic engineering techniques. These factors reduce the cost of these enzymes by 20 times the cost of the enzymes obtained by fermentation [19]. Plant-derived enzymes are accepted more commonly for application in the food and pharmaceutical areas. However, some disadvantages are associated with the use of lipases, such as the low availability of plants with lipolytic activity in nature, seasonality, and a decrease in the lipolytic capacity, considering the techniques applied for purification [20 apud 18].

Tavares et al. [21] studied the production of lipases obtained from castor seeds and evaluated their hydrolytic potential in cambre oil. The authors observed that the seeds in the natural form showed emulsifying properties, which gives a higher advantage to the hydrolytic reactions. Kouteu et al. [22] investigated the lipase activity on crude extracts of seeds of Adansonia suarezensis, Adansonia grandidieri, Moringa drouhardii, Moringa oleifera, Jatropha mahafalensis, and Jatropha curcas, and found that all extracts showed ethanolysis and hydrolysis activities.

Lipase was also found in orange residues by Okino-delgado and Fleuri [23]. These enzymes showed good biochemical characteristics, such as action in the broad pH range of 6-9 and optimal temperature around 20 °C to 60 °C, implying high thermostability, and the lipase activity measured was 68.5 lipase U/g. Moreau et al. [24] used sorghum grains as the raw material to extract lipases. The lipolytic activity was estimated by the release of fatty acids. Initially, the levels of free fatty acids were 3.76 ±0.38%, and in the samples stored for 7 and 14 days at 4 °C, the levels were 4.63 ±0.08 and 5.12 ±1.08%, indicating that sorghum grains have lipolytic enzymes.

2.2 Lipases of Animal Origin

Animals are another source of lipases, and the main sources of these enzymes were, initially, the swine and human pancreas. However, their production for commercial purposes currently includes microorganisms as the main source [25], besides other animals, such as sheep, calf, and lamb [3]. To make the use of porcine lipases for industrial application feasible, these enzymes undergo recombinant processes, which allows for even greater yields. Their application in industrial processes is not something usual, given the difficulty in the steps of extraction and purification to obtain the final product [25]. Another reason that prevents the use of pancreatic lipases obtained from animals is that, mostly, these enzymes are impure. Some porcine lipases have amino acids that give a bitter taste, limiting their application in the food industry [26].

The application of animal lipases is more limited to the procedures that involve clinical diagnosis [18]. Despite the low use of these enzymes in the biotechnological processes, the use of porcine lipases is still observed in some activities, such as treating tannery effluents [27] and wastewater from the dairy industry [28].

Sea bass liver lipases were extracted, and subsequently, a crude lipase extract was prepared for use in the degreasing of fish skin, with a lipase activity of 7.24 U/g of tissue. The enzyme exhibited the highest hydrolytic activity for p-NPP at 50 °C and pH 8.0 [29].

2.3 Microbial Lipases

Lipases can be produced by microorganisms, and there is a particular interest of industries in this source of lipase, which is widely used in the field of biotechnology [30]. Microbial lipases are preferred because of several characteristics that make them more attractive when compared to plant and animal lipases. Besides the ease of genetically modifying microorganisms, this source provides high enzyme yields, allowing the production of highly stable lipases [31]. Their capacity to tolerate organic solvents is one of the factors that make them necessary to conduct biotechnological processes [32,33]. Microbial lipases are active in a broad range of pH and temperature, and they can also have regioselectivity, chemoselectivity, and enantioselectivity [34].

The reactions catalyzed by lipases represented approximately 20% of the biotransformations performed in 2019, which demonstrates the versatility of this class of enzymes for catalyzing hydrolysis and synthesis reactions [35,36,37,38].

Among the lipase-producing microorganisms are fungi, bacteria, and yeasts [39]. Bacterial lipases are resistant to the conditions used in the industry and can be used in several applications, being found intracellularly or extracellularly or attached to the membrane [40]. In the food industry, bacterial lipases are used less often than fungal lipases because of some features, such as lower specificity to the substrate and because they cannot tolerate high temperatures, unlike fungal lipases. Conversely, the production of lipases by bacteria shows higher yields [41]. To obtain bacterial lipases, the most commonly used method is submerged fermentation (SF) [42]. This type of fermentation is especially required for large-scale production since it permits monitoring of the physicochemical parameters necessary for the development of the microorganisms involved [43].

Among the several groups of lipase-producing microorganisms, fungi are the most attractive for industrial application since these sources are frequently extracellular, allowing a simpler extraction [44 apud 60]. Fungal lipases have specific characteristics required by the biotechnological sector, such as stability, specificity, and ease of production [18]. Lipase production by fungi occurs mainly by solid-state fermentation (SSF) processes, which provides an advantage to lipid production since SSF is an extremely economical method [45].

Many microorganisms are good lipase producers. Among them, fungi, especially of the genera Rhizopus, Aspergillus, Penicillium, Geotrichum, Mucor, and Rhizomucor, yeasts of the genera Candida, Yarrowia, Pichia, Rhodotorula, and Saccharomycopsis, and bacteria of the genera Bacillus, Pseudomonas, Burkholderia, and Staphylococcus [4] are notable. Table 1, Table 2, and Table 3 show the lipases obtained from the sources mentioned in Section 2.1.

Table 1 Plant sources for lipase production.

Table 2 Animal sources for lipase production.

Table 3 Microbial sources for lipase production.

 

2.3.1 Lipases from Aspergillus niger

The fungal species Aspergillus niger is used in biotechnology for its ability to produce enzymes at high concentrations, as well as for its capacity to produce pharmaceutical supplies that are beneficial to human and animal health [87]. Aspergillus niger is an efficient producer of extracellular enzymes, such as amylases, cellulase, and lipases [88]. Notably, A. niger lipases are recognized by the FDA (US Food and Drug Administration) as GRAS (generally recognized as safe) substances for use in formulations of products that might be in direct contact with the human body [89].

Several studies have reported on the use of A. niger as an efficient lipase producer. Suyanto, Soetarto, and Cahyanto [90] obtained high lipolytic activity (10.83 U.mL-1) with A. niger by solid-state fermentation using agro-industrial residues as the substrate. Submerged fermentation can also be used in lipase production by A. niger, as reported by Lima et al. [55], who evaluated the effect of the variables on lipase production by Aspergillus niger C through submerged fermentation. The kinetics of lipase production in the study showed that high yields can be obtained in a brief period of fermentation which is required in industrial processes, since saving time and energy is important during large-scale production.

One of the main applications of Aspergillus niger lipases is biodiesel production [91]. Aliyah et al. [92] used the immobilized enzyme in four cycles for synthesizing biodiesel. Other applications include their use for treating industrial effluents [93,94]. These lipases might be used in the food industry, considering that they have been used for synthesizing flavor esters [95,96].

Several studies have been published on lipase production by strains of Aspergillus niger. The lipase activity of this fungus depends on the substrate used, as shown in Table 4.

Table 4 The enzymatic activity of Aspergillus niger using different substrates.

2.4 Biochemical Characteristics

The biochemical properties of lipases vary according to the sources, i.e., microorganisms, animals, or plants [105]. Given the great variety of microorganisms with lipolytic capacity, the lipases produced by the different strains present distinct operating characteristics. Some factors can inhibit or stimulate the production of this enzyme. The presence of certain compounds, such as glucose, fructose, and glycerol, leads to the inhibition of lipase production. However, its production is promoted in situations where there are concentrations of free fatty acids, triglycerides, or complex glycans [106].

The cultivation of microbes for enzyme production is affected by several factors. The hydrogen potential (pH) of the fermentation process is an extremely important factor [107]. Some parameters, such as temperature, the composition of the medium, volume of inoculation, aeration, and agitation, can interfere in the process of enzyme production and acquisition [108]. Other conditions that affect lipase production are the nutrients used, such as the carbon and nitrogen sources [30]. In general, lipases act in a pH range between 6.0 and 8.0 and a temperature range of 30-40 °C; however, these values may vary according to the source of the enzyme [109].

Although there is a consensus that these enzymes may act in an optimum pH range of 4-9 and temperature range of 25-70 °C [108], some studies found enzymatic activity at an extremely acidic pH of 2.0 [104], and a highly basic pH of 11.0 [110]. Other studies showed the capacity of lipases to act at high-temperature ranges. For example, Liu, Li, and Yan [111] characterized the lipase of Pseudomonas fluorescens and found that the lipase acted at an optimum temperature of 70 °C and remained active at temperatures between 80-100 °C. Colla et al. [45] characterized lipases from Aspergillus flavus and Aspergillus niger by submerged and solid-state fermentation, respectively. The lipases produced by submerged fermentation had 80% stability at an acidic pH, whereas the lipases obtained by solid-state fermentation showed stability higher than 60% at an alkaline pH (Table 5).

Table 5 The biochemical characteristics of some microbial lipases.

The ocean is a favorable environment to obtain enzymes from, as it has a large microbial diversity that favors the acquisition of enzymes with desirable characteristics, such as stability at various pH and temperature ranges and under high pressure and salinity [124]. Lipases from marine microorganisms were recently investigated. Promising sources were found, considering that these microorganisms have activity under extreme conditions [43]. A study performed by Balduyck et al. [125] involved the analysis of two microalgae species, including Nannochloropsis oculata and Tisochrysis lutea, and the lipases of these marine microorganisms exhibited high activity at 20 °C. Some lipases are active in a wide range of temperatures and pH, as reported by Latip et al. [126], who investigated the properties of lipase of Pseudomonas from Antarctica. The lipase produced by the bacterium revealed thermostability at 10-70 °C, and it was also very stable in a broad range of pH, which ranged from 5.0 to 10.0 and showed optimum activity at pH 8.

2.5 Lipase Production: Type of Fermentation and Nutrients

Submerged fermentation (SF) and solid-state fermentation (SSF) are used in enzyme production, in which the substrates involved depend on the type of fermentation. In SF, the substrates are dissolved in a liquid medium, whereas in SSF, solid substrates are used [127]. Lipases have broad specificity to diverse substrates [40,128,129]. This capacity to act on different substrates is a key characteristic of lipases, and their enzymatic yield depends on the substrate used. Some of their substrates are triacylglycerides, fatty acid esters, lipids, and synthetic and natural oils [130].

Geoffry and Achur [131] used the effluent from a palm oil factory as the substrate and obtained high lipase activity (7.8 U/mL), thus revealing that the effluent of palm oil factories is a good alternative for the substrate. The lipase produced by Streptomyces sp. has good potential in the catalytic processes, although its production is slightly low. Thus, to optimize the conditions of fermentation, Lan, Qu, Yang, and Wang [132] used 4-nitrophenyloctanoate as the substrate and, together with the other optimized parameters, the level of lipase expression increased to 442 U/mL. Salihu, Bala, and Alam [133] showed that peanut pie is a promising substrate for lipase production by Aspergillus niger. To reduce the costs of enzyme production, Rodrigues et al. [61] used sesame (Sesamum indicum), sunflower (Helianthus annus), and linseed (Linum usitatissimum L.) seed meals as substrates since they have a high lipid value. For all substrates, enzyme production was confirmed; however, the maximum production (160 U) occurred when sunflower seeds were used.

For producing microbial lipases, regardless of the type of fermentation used, carbon and nitrogen sources must be used as nutrients. These nutritional factors directly influence enzyme productivity [30]. During lipase production, the medium must be rich in nutrients. For better enzyme yields, carbon sources, such as glucose, xylose, maltose, lactose, sucrose, glycerol, and other compounds can be added [134]. Numerous oils are described as carbon sources for lipid production; olive oil, palm oil, sunflower oil, and almond oil presented higher enzyme yields in a study by Colla et al. [42]. The synthesis of microbial lipases is controlled by various conditions; however, the carbon source used in the process is the determining factor for the amount of enzyme produced by the microorganism [135].

In several studies, different carbon sources were tested for lipase production. Jesus et al. [136] evaluated lipase production by Bacillus subtilis using manipueira as the carbon source, which is a good alternative source, considering that it can exhibit lipolytic activity in a medium in the absence of inducers. Castiglioni, Costa, and Alegre [137] found a positive effect on the activity of the lipase produced by Burkholderia cepacia when the concentrations of soy oil were increased, thus demonstrating that soy oil is a good alternative. Castro et al. [138], in their study, analyzed lipase production by the fungus Aspergillus westerdijkiae, and in their experiment, a comparison of enzyme production was performed in a medium supplemented with olive oil as the carbon source and another medium without the addition of olive oil. Their results demonstrated that the yield of lipase activity was around 40,000 U/g in the medium containing 1% olive oil. A reduction in the activity was found when the fungus was cultivated without olive oil, presenting a yield of only 12,000 U/g.

Using appropriate nitrogen sources is crucial to the fermentation process, and the addition of these nutrients directly influences cell growth, as well as enzyme production [139]. The nitrogen sources favor microbial growth, and among the most applied are corn steep liquor, soy flour, yeast extract, peptone, urea, nitrate, and ammonium salts [140]. Many studies used yeast extract as the nitrogen source [55,85].

Penha et al. [121], in their experiment, concluded that pure palm pie was an incomplete substrate for lipase production by A. niger 11T53A14; however, ammonium sulfate solution, as a nitrogen source, was capable of positively influencing lipase production. Salwoom et al. [141] analyzed the nutritional factors involved in lipase production by bacterial strains obtained from the soil of Signy Island. In that study, various carbon (sucrose, maltose, lactose, glucose, and arabinose) and nitrogen (casein, yeast extract, peptone, and tryptone) sources were tested; only the addition of glucose as a carbon source significantly increased enzymatic activity compared to the mean. Other carbon sources reduced lipase production by more than 50%, relative to the nitrogen sources used. A slight increase in lipase production occurred in the presence of peptone, and applying other nitrogen sources slightly decreased the enzymatic activity compared to the mean production.

To achieve a high enzymatic activity, several points must be considered. One of them is the presence of an ideal choice of inducers, a variable that must be investigated by researchers to increase lipase production [142]. In a study, Reinehr et al. [102] tested the production of lipases by Aspergillus in the presence of two inducers, i.e., soy oil and glycerol. Their results demonstrated that higher hydrolysis activities were found using 2% soy oil as an inducer, with activity values of 7.69 U/g (A. fumigatus), 1.03 U/g (A. niger), and 24.17 U/g (A. niger O-4). Patel et al. [143] optimized the production of lipase from Pseudomonas aeruginosa in their experiments, analyzing several oils as inducers. Among the vegetable oils tested, the best inducer for lipase production was olive oil. A reduction of 15.8% in the activity was found when the inducer was substituted by coconut oil, and the other oils used were classified in the descending order for enzyme production as castor oil > peanut oil > cottonseed oil > sunflower oil. Excessive lipid concentrations in the growth medium might result in cytotoxic effects on the microorganism [144]. Prabaningtyas et al. [99] described the production of a fungal lipase. In their study, enzyme production was evaluated at various inducer concentrations (2%, 4%, and 8% g/g of growth medium), and the highest lipid activity (163.33 U/g of growth medium) was obtained when 4% of the inducer was added to the fermentation process.

Another factor that might interfere in lipase production is the agitation of the growth medium [42]. To optimize the parameters for the production of lipase of G. thermoleovorans DA2, Fotouh, Bayoumi, and Hassan [145] adjusted the agitation rate at different rotations per minute, i.e., 40, 80, 120, and 150 rpm. The maximum lipase activity (1,021.91U/mL) was found at 120 rpm, and in the other rotations, the enzymatic activity was lower. Lipase production by Bacillus cereus HSS was analyzed by Hassan, Latif, and Ali [146]. The production was evaluated under static conditions and agitation at 120 rpm. They found that the maximum lipase activity was 285 U/mL under the condition of agitation and 225 U/mL in the stationary mode.

2.6 Lipases Produced by Microorganisms Isolated from Industrial Residues

Microorganisms that produce lipase can be found in several types of environments, especially in industrial residues [147]. These industrial effluents are characterized as regions that favor microbial growth and development, as these areas contain high concentrations of free fatty acids that can be assimilated as nutrients and substrates for their metabolism [57]. Peil et al. [148], in their study, isolated microorganisms that could produce lipases in different industrial effluents. The samples were collected in slaughterhouses and dairy industries, and 21 bacteria and seven fungi were isolated. Extracellular lipase production was observed in 71.43% of the bacteria and 57.14% of the fungi.

Lipase production by microorganisms isolated from residues of vegetable oil and petroleum refineries was investigated by Sperb et al. [149]. A total of 24 fungi were isolated and were subjected to qualitative tests for the presence of lipases; only three of them tested positive for lipase activity, including Aspergillus niger and two Rhizopus species. Padilla, Tabarez, and Orduz [57] collected samples from industrial effluents contaminated with vegetable oils and isolated microorganisms. In that study, the lipolytic enzymatic activity was evaluated under different conditions of temperature and pH. In total, 149 microorganisms were isolated, of which only 37 showed lipolytic activity. For enzyme production, olive oil was used as the substrate, and the optimum pH for the enzymatic activity was 8.0 at 37 °C.

To isolate lipase-producing yeasts, Salgado et al. [150] used wastewater from oil mills as a source of microorganisms. They isolated 32 yeasts, five of which were subjected to tests for determining extracellular lipase activity. Olive oil was used as a supplement in the synthetic liquid medium (YEP). Only JT5 showed a high lipase activity of 0.85 U/mL, and the identification of the strain revealed the yeast Magnusiomyces capitatus. The maximum enzyme production by the microorganism mentioned was achieved using the effluent as the substrate, supplemented with yeast extract and olive oil as carbon sources.

Industrial residues from palm oil (palm pie, palm bark, and fiber) were collected for isolating microorganisms with lipolytic activity by Kuncharoen et al. [151], and 16 yeasts were isolated. The strain Trichosporon insectorum 4E-1D was the best lipase producer in palm oil and soy oil as carbon sources (28.19 ±4.84 U/mL in palm oil and 22.63 ±0.18 U/mL in soy oil). An industrial effluent rich in organic matter is vinasse, which is a liquid residue from distilleries. It is formed by fermentation, mainly performed in the sugar and alcohol industry [152]. This industrial residue was used by Guan et al. [153], who isolated a new Lactobacillus species with lipolytic activity. Three strains were isolated and subjected to lipase activity using olive oil as the substrate. The lipase activities of the three strains varied from 5.88 to 17.79 U/mL. Padma and Vasudevan [154] collected samples from vegetable market waste tanks to isolate proteolytic and lipolytic microorganisms. The isolate that could produce lipase was a bacterium in the genus Bacillus; however, the study did not perform quantitative analysis for determining enzyme production.

Dairy wastewaters are conducive to microbial development, and they were the target of the study conducted by Ficarra et al. (2016) [155], who used these wastewaters as the source of microorganisms to isolate lipase producers. Ogunnusi and Olorunfemi [156] isolated lipolytic bacteria from slaughterhouse effluents and cow manure samples. In total, 12 isolates were obtained, seven of which were from slaughterhouse effluents, and the bacteria were tested for lipase production, with a broth supplemented with oil used as the production medium by submerged fermentation. The bacterium with the highest lipolytic activity was identified as Pseudomonas aeruginosa (6.0 × 10-3 mg/mL). The study demonstrated the potential these residues have as sources for enzyme production. Palm oil residues were used for isolating lipase-producing bacteria, and the microorganisms were subjected to fermentation. Hasan et al. [157] evaluated the effect of different carbon and nitrogen sources. Seven bacterial isolates were obtained, and one isolate of the genus Bacillus had a maximum yield of 0.168 µg/mL.min using palm oil as a carbon source, whereas, for the nitrogen source, the addition of tryptone to the medium showed the highest enzymatic activity (0.135 µg/mL.min). The optimum pH and temperature for fermentation were 7 and 37 °C, respectively.

Based on these studies, we infer that microbial growth and development occur in different types of environments with distinct characteristics. Among them, industrial residues have been extensively studied as they are a great source of microorganisms (Figure 4). The results, mostly positive, led many authors to search for new lipase-producing microorganisms.

Click to view original image

Figure 4 The segments and industries with wastewater rich in organic matter from which lipase-producing microorganisms can be isolated.

2.7 Enzyme Immobilization

The use of enzymes in the industrial field has been impaired by some negative factors associated with these biomolecules. The possibility of denaturation because of their low stability and the complexity in the recovery of the enzyme for consecutive uses make their application challenging in the free form [158]. To solve these adversities, enzyme immobilization techniques aim at facilitating the steps of enzyme separation and recovery since one disadvantage associated with the use of enzymes in the free form is their low stability related to parameters such as temperature and pH [159]. This process of immobilization is based on a bond between the protein and the surface of an insoluble matrix (solid support) [160].

The methods for enzyme immobilization include adsorption, covalent bonding, entrapment/encapsulation, and cross-linking, as demonstrated in Figure 5 [161]. Regarding the immobilization techniques, there is no single methodology that can be used for all enzymes; therefore, for determining the method, some points must be considered, such as the simplicity of the procedure, low cost, and choosing a procedure that provides a satisfactory catalytic activity, besides a technique that allows good operational stability [2].

Click to view original image

Figure 5 Methods of enzyme immobilization: adsorption, entrapment, covalent bonding, and cross-linking.

An immobilization technique that has attracted attention is known as CLEAs (Cross-Linked Enzyme Aggregates). It is an advantageous method since its methodology is relatively simple and requires a short time to be performed [162]. Proposed by Sheldon [163], CLEAs have interesting characteristics that are desired in industrial biotransformations. The method is very accessible, does not require the use of transporters, two or more enzymes can be co-immobilized, the cost is low, and it allows the retention of high enzymatic activities, as well as has good thermal stability, besides high tolerance to organic solvents. In the literature, there are reports on the use of CLEAs to immobilize various enzymes, such as lipases [164], cellulases [165], proteases [166], beta-galactosidase [167], amyloglucosidase [168], laccase [169], and acetylesterase [170].

CLEAs are based on two stages. Initially, the enzyme is precipitated by precipitating agents, such as inorganic salts, polymers, and organic solvents, among others. This is followed by the action of cross-linking agents, among which glutaraldehyde is used most commonly as a bifunctional reagent, or by multifunctional agents through reactions with the amino groups present on the surface of the enzyme [171]. Furthermore, this method combines purification and immobilization in a single step, which makes the whole procedure faster and simpler [172].

Dietary proteins might be used alternatively when the enzyme presents a low content of residues of the amino groups on its surface. Thus, to improve the process of cross-linking, the addition of substances that contain many amino groups, such as bovine serum albumin and proteins, is implemented. These substances are called co-feeders [173]. Mafra et al. [174] evaluated the concentration necessary for bovine serum albumin (BSA) (20, 40, and 60 mg) for catalase immobilization in a comparative analysis of enzyme immobilization with and without BSA. It was found that the use of BSA reduces the resistance to mass transfer, and the efficacy factor was 3.7 times greater when bovine serum albumin was used. Two protein feeders were studied by Goetze et al. [175], including feather meal and BSA, for preparing combi-CLEAs to clarify grape juice. Both feeders showed positive effects on the recovery of the activity of the enzyme pectinase; however, the use of feather meal showed higher activity recoveries and was cheaper than using BSA in the immobilization of pectinase.

Several studies have been published regarding lipase immobilization using CLEAs. Picó et al. [176] immobilized lipase B from Candida antarctica by the CLEAs method using ammonium sulfate and glutaraldehyde. They found that the reuse of the enzyme was very satisfactory, with biodiesel conversions possible for up to 10 catalytic cycles, retaining 100% of the initial activity. In another study, a lipase from A. niger was precipitated using ammonium sulfate, and cross-linking was shown for the first time with the addition of Sodium Tripolyphosphate (TPP). The use of TPP generated favorable results since it provided highly stable enzymes at broader temperature and pH ranges compared to the free enzymes [177]. An advantage of this technique is its ability to make the enzyme more pH stable. JiaoJiao et al. [178] showed that lipase from Candida rugosa could be immobilized by the CLEAs methodology and its behavior improved as the enzyme became more stable at a particular pH compared to the free enzyme.

CLEAs and combi-CLEAs are promising immobilization methods, especially because they do not require the use of previously purified enzymes. However, the optimization of the precipitation and cross-linking parameters and the interaction between the parameters is necessary. Their capacity to improve enzyme stability can benefit bioprocesses in industrial applications [179].

2.8 Industrial Applications

The use of microbial enzymes in the industry is an alternative that has been growing considerably, especially because of their use in the food sector, such as in meat processing, dairy industry, beverage production, bakery, and the pharmaceutical industry [180]. Among the industrial enzymes, lipases are prominent due to their versatility, given their exclusive properties [25]. They are widely used in the field of biotechnology, biodiesel production, drug formulation, food industry, and detergent industry [181]. Other sectors also benefit from the use of lipases, such as the production of cosmetics and the textile and leather industries [182] (Figure 6).

Click to view original image

Figure 6 The application of lipase in the industrial sectors. Source: Adapted from Gonçalves Filho, Silva, and Guidini [183].

2.8.1 Lipases in the Food Industry

The food industry uses lipase for various processes, including increasing food quality and yield, providing aromas to certain products, and removing oils and fats. It is used in dairy products, bakeries, breweries, and processing meat, among others [133]. In the dairy industry, lipase hydrolyzes the fat present in milk. Lipase also enhances the flavor of cheese and accelerates its ripening process [184]. Biolipolysis is a process in which lipases are incorporated to remove the fat present in the meat during their processing, to obtain lean meat, and is frequently used in fish meat. Furthermore, microbial lipases are used to refine the flavor of rice, and lipase can also be used in the production of omega-3 [185].

Flavoring substances, such as esters, can be incorporated into the food, cosmetics, and pharmaceutical areas. Natural sources, such as fruits, have these esters. However, their yield is low, and throughout the ripening process of the fruit, their composition is affected. Thus, the use of these enzymes is a feasible alternative for synthesizing these aromatic esters [186]. Lipases are powerful biocatalysts that act in the production of polyunsaturated fatty acids and the synthesis of esters and alcohols. These compounds generated by the action of lipase are used for producing healthy foods and even in the formulation of food supplements [136]. Souza, Ribeiro, and Coelho [187] used a lipase from Yarrowia lipolytica, which was obtained by submerged fermentation using soybean meal, to synthesize various flavoring esters with commercial value in the food industry. Free and immobilized lipases were used for catalyzing the enzymatic synthesis of isoamyl acetate and isopentyl acetate esters. Ester conversions were 98.4% and 73.7% for isoamyl acetate and isopentyl acetate, respectively, in a study by Bayramoglu et al. [188]. These results demonstrate the capacity of lipases as potential biocatalysts in the application of aromatic esters.

Another very important function of lipase is its use to expand the shelf life of some shelf products, such as bakery products, and to improve the flavor of juices, soups, sauces, cheeses, and baked foods. Some properties of noodles improved with the addition of lipase [189]. Cao et al. [190], in their study, found an enhancement in blue cheese by the application of lipase, which influenced the characteristic aroma of the cheese, besides contributing to the typical flavor of the cheese by synthesizing methyl ketones. A study by Schaffarczyk, Ostdal, and Koehler [191] evaluated the influence of the application of lipase in baking. The addition of this enzyme resulted in an increase in the volume of bread from 56% to 58%, depending on the type and concentration of the lipase added. Some lipases can be applied in several sectors, such as the lipase from Candida rugosa (CRL), which can help to enhance the aroma and flavor of foods and can be used in the industry of fats and oils, besides being used in other areas [192].

2.8.2 Lipases in Biodiesel Production

Renewable energy sources are now being investigated, especially to substitute the use of the non-renewable ones, which are being depleted rapidly. The production of biodiesel using enzymes received emphasis for being a highly feasible and eco-friendly process, where the degradable oils and fats with a high free fatty acid (FFA) content are used for biodiesel production. These compounds are converted by esterification and transesterification reactions mediated by biocatalysts such as lipases, thus, generating biodiesel as the final product [193].

The lipases from Burkholderia cepacia, Rhizomucor miehei, and Candida rugosa were described by Fan et al. [72] as promising catalysts for biodiesel production, and a comparison of biodiesel production by these three lipases was performed using soybean oil and residual vegetable oil. The results showed that the residual vegetable oil can be effectively converted into biodiesel by the lipase of R. miehei through the transesterification process. Luliano et al. [194] used the lipase of Candida rugosa to produce biodiesel, using a methodology that involved the conversion of grains used in breweries into biodiesel in the presence of methanol. Good results were obtained with a high yield (98%), and the process also had a low cost since the catalyst could be reused even after four cycles. According to Amini et al. [195], biodiesel can be efficiently produced using the oil of Ocimum basilicum seeds as a new source of biodiesel. Immobilized lipase (Novozym 435) and methanol were used for production. The immobilized lipase was reused for up to seven cycles.

Lipases can be used as efficient biocatalysts for producing biodiesel. However, some disadvantages are associated with this practice. For example, the biodiesel generated by this methodology has a low yield, and its use in machines might damage these devices. Additionally, in one of the last production steps, it is difficult to remove the unreacted methanol while producing biodiesel. Thus, additional techniques need to be used to remove the impurities and obtaining high-quality fuel [196].

However, several studies have reported promising results with the use of lipases as catalysts in biodiesel production. For example, in a study by Costa et al. [197], the researchers investigated biodiesel production through the reactions of hydrolysis and esterification. The lipases of Geotrichum candidum were used for hydrolyzing soybean oil, and the subsequent step in free fatty acid esterification was mediated by the immobilized lipase from Pseudomonas fluorescens (PEL). The hydroesterification mechanism used in the study produced high-quality biodiesel, and the reactions presented favorable viscosity and purity. The hydrolysis reactions presented a complete conversion of the soybean oil within a short period, and the free fatty acids did not require purification, which was advantageous since it reduced the cost. The synthesis of biodiesel by transesterification, catalyzed by a lipase of Aspergillus niger, was studied by Shalini et al. [198]. The immobilized enzyme presented a total biodiesel yield of 90%, and it could be used five times for synthesizing biodiesel with no loss in its activity. Thus, it is an efficient catalyst in biodiesel production. The performance of studies at a laboratory scale, pilot projects, as well as process optimization have advanced the field of catalysts. These studies help to improve enzyme activity and productivity [199].

The use of lipases as biocatalysts has several advantages; however, it is necessary to study the existing methodologies and search for new ways to broaden the use of lipases in the production of this biofuel [200].

2.8.3 Lipases in the Pharmaceutical Industry

In the pharmaceutical sector, the production of pure enantiomers is one of the most significant methodologies since a large percentage of the drugs in progress are chiral. Thus, this production of pure enantiomers is mediated by the action of biocatalysts, especially lipases [201]. This process can also occur by chemical methods; however, the use of enzymatic catalysts is preferable due to their efficiency and selectivity. Additionally, many enantiomers can be produced using these enzymes [202].

Lipases can produce enantiomers. For example, ibuprofen can be produced from certain compounds using a lipase of Rhodothermus marinus [203]; atenolol can be produced using a lipase of Candida rugosa [204]; milnacipran, a medicine used in fibromyalgia therapy, can be produced using lipases [205]. Other drugs, such as nebacetin, naproxen, ascorbic acid, derivatives of quinolones, ketoprofen, and chloramphenicol, are medicines produced using different types of lipases [206].

The extensive use of these enzymes in the pharmaceutical area derives from their characteristics of regioselectivity and stereospecificity, which are desirable in the biocatalysts used for synthesizing pharmaceutical products [207]. The process of biocatalysis might be visited in the future with a stronger emphasis on the production of medicines and pharmaceutical intermediates [208].

2.8.4 Lipases in the Detergent Industry

Lipases have numerous applications, among which their action in detergent production is emphasized [200]. Lipases break triglyceride molecules into fatty acids, thus generating less emulsion and making it easier to remove triglycerides [209]. In the detergent industry, the aqueous medium is alkaline, and thus, the lipases used in the formulation of detergents must be alkalophilic [210]. The use of enzymes in the formulation of detergents has some advantages, such as a reduction in the time necessary for washing and increasing the life span of the fabric [211].

Studies have shown that microbial lipases can efficiently remove fat. For example, García-Silvera et al. [111] successfully used a lipase of Serratia marcescens to remove triacylglycerol of olive oil from cotton fabric. Shrimp hepatopancreas lipases were used in the production of a detergent, and its effectiveness was compared to that of commercial detergents for clothes; positive results were found [212]. A lipase produced by Bacillus subtilis showed desirable properties for application in the production of detergents. It was highly resistant to surfactants, oxidizing agents, and commercial detergents and thus could be a strong candidate for use in the cleaning industry [213].

The use of alkaline lipase in the production of detergents can decrease the use of surfactants and other compounds that, from an ecological point of view, are not favorable [214]. Since commercial detergents contain dangerous chemical compounds, both for human and environmental health, lipases represent an alternative for the use of these substances, besides being preferable for their application at room temperature and maintaining fabric quality [40].

2.8.5 Lipases in the Treatment of Effluents

Many industrial residues are released in aquatic environments, which results in the death of aquatic organisms, as these residues contain many oily compounds, which impair the diffusion of oxygen in the water. To solve this environmental problem, enzymes are used to assist in reducing the content of oils and fats present in these residues [215].

Ktata et al. [216] performed the biological treatment of wastewaters using the lipase produced by Aeribacillus pallidus. The treatment of the effluents generated by the industry of canned tuna production was highly efficient (96.11%) when lipase was used for hydrolyzing wastewater at 50 °C for 1 h. The lipase of Bacillus sp. was evaluated regarding its capacity to hydrolyze fats in different industrial effluents (dairy, bakery, and poultry). The enzyme could hydrolyze more than 50% of the initial fat present in all these effluents, and thus, it was considered to be a good alternative in the pretreatment of these effluents [217]. A reduction of fat by three times was identified in the pretreatment of the effluent from a slaughterhouse when a lipase of Fusarium verticillioides was used. The enzymatic activity had an optimum temperature and pH of 45 °C and 5.5, respectively, with the enzyme being stable at a broad pH range (4.0 to 8.0). These properties are desirable for industrial applications [218].

This pretreatment step using lipases is highly feasible since enzymatic hydrolysis allows the subsequent step of the treatment by microbes to be simpler [219]. Studies and new technologies have shown that the enzymes in the pretreatment of industrial effluents reduce the costs involved in this process, besides being eco-friendly [220]. Lipases are often used in the industry; however, their use in the treatment of effluents is limited by their high cost. Thus, the use of these biocatalysts to eliminate the lipids present in the effluents needs further investigation, considering that obtaining high-value by-products in waters with high concentrations of lipids can become attractive and influence the economy [221].

2.8.6 Lipases in Cosmetics

Lipases have been used in the cosmetics industry to produce active compounds or in the catalysis of chemical products, such as esters, aromas, and active agents [222].

In the field of enzymocosmetics, lipases are especially efficient in the formulation of cleaning products and creams for cellulite treatment [223]. Lipases are also used in the preparation of slimming products, as they can break down fat deposits; moreover, hair and makeup products are also created by lipase action. Several products have been synthesized by the catalytic action of lipases, and many of them have been commercialized [222].

A concern of the cosmetic industry is the production of aromas, which are used in fragrances and other products, and this is an important aspect, especially for the consumers. Thus, the industry seeks biotechnological alternatives for the development of natural aromas [224]. Biotechnological processes have been investigated by the cosmetics industry, and the use of biotechnology has led to the formulation of new products, as well as the optimization of the existing methodologies. The use of enzymes has allowed this sector to provide better products to consumers [225].

2.8.7 Lipases in the Production of Biosensors

Lipases can be used to produce biosensors since their characteristics make them attractive for the manufacture of these materials. Several microbial and animal lipases are used for developing these devices [226]. The advantages of using enzymes for the manufacture of biosensors include high specificity and selectivity, the possibility of regeneration, high yields, and real-time diagnosis. Many biosensors based on enzymatic actions are used in the field of health, such as for domestic monitoring of blood glucose (glucometer) and in portable clinical analyzers [227].

The lipases of Candida antartica and Aspergillus oryzae were used by Hasanah et al. [228] for developing an optical biosensor to detect triglycerides (TGs). The immobilized lipases can hydrolyze triglycerides, releasing glycerol and free fatty acids. The released acid decreases the pH and the acidity of the medium, which is then detected by an optical pH sensor. Other lipase-based biosensors have been described in some studies. Zehani et al. [229] prepared a conductometric enzymatic biosensor using a lipase of Candida rugosa (CRL). This equipment could detect organophosphate pesticides (diazinon, methyl parathion, and methyl paraoxon) in an aqueous medium. In this device, lipase was responsible for catalyzing diazinon hydrolysis. The biosensor presented positive results since it was capable of measuring concentrations as low as 60 µg. Also, for the detection of pesticides, a biosensor was produced with lipases of Ceratobasidium sp. by Barboza et al. [230] for carbendazim determination, and in this sensor, the lipase hydrolyzed the substrate p-nitrophenyl acetate (pNPA).

The stability of enzymatic biosensors is required in these devices, and their specificity is an important parameter in avoiding false signals. The development of these biosensors is an example of the cooperation between the Biological Sciences and Engineering. Biosensors can be used to monitor health at home, as well as, in hospitals [231].

2.8.8 Lipase in the Textile Industry

Lipases are used in the textile industry, and one of their applications is based on the treatment of fabric. Some conventional treatments use substances that may result in the loss of fabric weight and a higher probability of causing an itch. The substitution of these compounds by lipases reduces the possibility of a reaction, as well as the loss of fabric weight. This enzymatic treatment makes certain fabrics, such as polyester, softer and efficiently modifies the properties of the fabric [232].

Lipases can remove natural oils and fats. Additionally, a high level of dyeing and cleaning performed by this enzyme imparts a high degree of hydrophilicity to cellulosic textiles [233]. Polyester fabrics were treated with lipases by El-Shemy, El-Hawary, and El-Sayed [234]. The enzymatic treatment provided better dyeing capacity, as well as higher wettability and recovery of moisture, and the action of lipase did not promote deterioration of the fabric under the reaction conditions. Taleb et al. [235], in their study, revealed that Bacillus aryabhattai can be used to efficiently produce a lipase with a strong ability to remove the lipid barrier from the surface of wool fibers. The enzyme proved to be efficient for bio-cleaning of wool up to five times. The use of the lipase was successful since its application did not deteriorate the fiber properties. Several studies have used lipases for cleaning, restoring, and conserving historic fabrics [236].

The advantage of using enzymes in the treatment of fabric over the conventional methodologies is that the properties of polyethylene terephthalate are maintained since enzymes are large proteins and do not enter the material. Moreover, lipases do not corrode the fabric, an effect that occurs when other substances are used, such as in alkaline treatments [237].

2.8.9 Lipases in Bioremediation

Lipases can be used for the bioremediation of lipid residues [132]. They can clean and decontaminate soil contaminated with greasy residues by converting these contaminants into non-toxic substances [238]. The lipase from Pseudomonas sp. was used by Sahoo, Sahu, and Subudhi [239] for the bioremediation of soil contaminated with hydrocarbons from automotive oils. The enzyme used mediated the degradation of this oil, causing 92.6% of the automotive oil to be degraded, thus demonstrating its efficacy in this process.

However, some disadvantages are also associated with the use of enzymes in bioremediation, such as the high cost of enzyme production, the possible inactivation of the enzyme by the conditions used in the process, and limited knowledge of the behavior of the enzymes when they interact with pollutants. Therefore, more studies need to be conducted on the use of enzymes in bioremediation to optimize the parameters and improve the technique [240].

Several industrial sectors use biocatalysis mediated by lipase, and the methodologies used have been studied and optimized. Lipases have been improved due to the association of biotechnology, genetic engineering, and protein engineering, which has allowed the production of several high-quality products. Although there are still some limitations, these inconveniences can be resolved as new technologies are developed [241].

3. Conclusion

Lipases are versatile enzymes used in several industrial sectors, with their activity and biochemical characteristics being greatly influenced by the production source, formulation of the culture medium, and the fermentation method. New microorganisms isolated from residual sources can be an alternative to minimize production costs, besides being the source of new lipases with interesting characteristics required by different industrial segments.

Acknowledgments

We would like to thank Bioprocess Laboratory and Federal University of Alfenas.

Author Contributions

Writing and original draft preparation, Myllena R. Santos; Writing and data curation Daniela B. Hirata; Writing-review and supervision Joelise A. F. Angelotti.

Funding

This study was funded by the Coordination for the Improvement of Higher Education Personnel - Brazil (CAPES) - Financing Code 001 and by the Federal Center for Technological Education through a program to encourage personnel improvement.

Competing Interests

The authors have declared that no competing interests exist.

References

  1. Malajovich MA. Biotecnologia. 2nd ed. Rio de Janeiro: Axcel Books do Brasil Editora; 2016.
  2. Resende RR, Soccol CR, França LR. Biotecnologia aplicada à agro & indústria: Fundamentos e Aplicações. 4th ed. São Paulo: Blucher; 2016. pp.1069. [CrossRef]
  3. Guerrand D. Lipases industrial applications: Focus on food and agroindustries. Oilseeds Fats Crops Lipids. 2017; 24: D403. [CrossRef]
  4. Liu X, Kokare C. Microbial enzymes of use in industry. In: Biotechnology of microbial enzymes. London: Academic Press; 2017. pp.267-298. [CrossRef]
  5. Grand view research. Enzymes market size, share & trends analysis report by type (industrial, specialty), by product (carbohydrase, proteases), by source (microorganisms, animals), by region, and segment forecasts, 2021-2028. San Francisco, CA: Grand view research; 2021; 978-1-68038-022-4.
  6. Homaei A. Enzyme immobilization and its application in the food industry. Adv Food Biotechnol. 2015; 9: 145-164. [CrossRef]
  7. Sharma AK, Sharma V, Saxena J. A review on applications of microbial lipases. Int J Biotech Trends Technol. 2016; 19: 1-5. [CrossRef]
  8. Gao Z, Chu J, Jiang T, Xu T, Wu B, He B. Lipase immobilization on functionalized mesoporous TiO2: Specific adsorption, hyperactivation and application in cinnamyl acetate synthesis. Process Biochem. 2018; 64: 152-159. [CrossRef]
  9. Aguieiras EC, Cavalcanti-Oliveira ED, Freire DM. Current status and new developments of biodiesel production using fungal lipases. Fuel. 2015; 159: 52-67. [CrossRef]
  10. Kazlauskas R. Hydrolysis and formation of carboxylic acid and alcohol derivatives. In: Organic synthesis using biocatalysis. Amsterdam: Academic Press; 2016. pp.127-148. [CrossRef]
  11. Casas-Godoy L, Gasteazoro F, Duquesne S, Bordes F, Marty A, Sandoval G. Lipases: An overview. In: Lipases and phospholipases. Methods in molecular biology. New York: Humana Press; 2018. pp.3-38. [CrossRef]
  12. Khan FI, Lan D, Durrani R, Huan W, Zhao Z, Wang Y. The lid domain in lipases: Structural and functional determinant of enzymatic properties. Front Bioeng Biotechnol. 2017; 5: 16. [CrossRef]
  13. Monteiro RR, Virgen-Ortiz JJ, Berenguer-Murcia A, da Rocha TN, dos Santos JC, Alcantara AR, et al. Biotechnological relevance of the lipase A from Candida antarctica. Catal Today. 2021; 362: 141-154. [CrossRef]
  14. Luan B, Zhou R. A novel self-activation mechanism of Candida antarctica lipase B. Phys Chem Chem Phys. 2017; 19: 15709-15714. [CrossRef]
  15. Liu DM, Chen J, Shi YP. Advances on methods and easy separated support materials for enzymes immobilization. Trends Analyt Chem. 2018; 102: 332-342. [CrossRef]
  16. Bancerz R. Przemysłowe zastosowania lipaz. Postepy Biochem. 2017; 63: 335-341.
  17. Geoffry K, Achur RN. Screening and production of lipase from fungal organisms. Biocatal Agric Biotechnol. 2018; 14: 241-253. [CrossRef]
  18. Sarmah N, Revathi D, Sheelu G, Yamuna Rani K, Sridhar S, Mehtab V, et al. Recent advances on sources and industrial applications of lipases. Biotechnol Prog. 2018; 34: 5-28. [CrossRef]
  19. Mounguengui RW, Brunschwig C, Baréa B, Villeneuve P, Blin J. Are plant lipases a promising alternative to catalyze transesterification for biodiesel production? Prog Energy Combust Sci. 2013; 39: 441-456. [CrossRef]
  20. Seth S, Chakravorty D, Dubey VK, Patra S. An insight into plant lipase research-challenges encountered. Protein Expr Purif. 2014; 95: 13-21. [CrossRef]
  21. Tavares F, Petry J, Sackser PR, Borba CE, Silva EA. Use of castor bean seeds as lipase source for hydrolysis of crambe oil. Ind Crops Prod. 2018; 124: 254-264. [CrossRef]
  22. Nanssou Kouteu PA, Baréa B, Barouh N, Blin J, Villeneuve P. Lipase activity of tropical oilseed plants for ethyl biodiesel synthesis and their typo-and regioselectivity. J Agric Food Chem. 2016; 64: 8838-8847. [CrossRef]
  23. Okino-Delgado CH, Fleuri LF. Obtaining lipases from byproducts of orange juice processing. Food Chem. 2014; 163: 103-107. [CrossRef]
  24. Moreau RA, Harron AF, Powell MJ, Hoyt JL. A comparison of the levels of oil, carotenoids, and lipolytic enzyme activities in modern lines and hybrids of grain sorghum. J Am Oil Chem Soc. 2016; 93: 569-573. [CrossRef]
  25. Borrelli GM, Trono D. Recombinant lipases and phospholipases and their use as biocatalysts for industrial applications. Int J Mol Sci. 2015; 16: 20774-20840. [CrossRef]
  26. Sharma S, Kanwar SS. Organic solvent tolerant lipases and applications. Sci World J. 2014; 2014: 625258. [CrossRef]
  27. de Oliveira Lima V, Prates KV, Fazolo A. Avaliação da eficiência de bactérias lipolíticas e lipase suína no pré-tratamento de efluente de curtume. Rev Gestao Sustentabilidade Ambiental. 2019; 8: 131-151. [CrossRef]
  28. Mendes AA, Pereira EB, Furigo Jr A, Castro HF. Anaerobic biodegradability of dairy wastewater pretreated with porcine pancreas lipase. Braz Arch Biol Technol. 2010; 53: 1279-1284. [CrossRef]
  29. Sae-Leaw T, Benjakul S. Lipase from liver of seabass (Lates calcarifer): Characteristics and the use for defatting of fish skin. Food Chem. 2018; 240: 9-15. [CrossRef]
  30. Bharathi D, Rajalakshmi G. Microbial lipases: An overview of screening, production and purification. Biocatal Agric Biotechnol. 2019; 22: 101368. [CrossRef]
  31. Gupta R, Kumari A, Syal P, Singh Y. Molecular and functional diversity of yeast and fungal lipases: Their role in biotechnology and cellular physiology. Prog Lipid Res. 2015; 57: 40-54. [CrossRef]
  32. Thapa S, Li H, OHair J, Bhatti S, Chen FC, Nasr KA, et al. Biochemical characteristics of microbial enzymes and their significance from industrial perspectives. Mol Biotechnol. 2019; 61: 579-601. [CrossRef]
  33. Salihu A, Alam MZ. Solvent tolerant lipases: A review. Process Biochem. 2015; 50: 86-96. [CrossRef]
  34. Wang W, Zhou W, Li J, Hao D, Su Z, Ma G. Comparison of covalent and physical immobilization of lipase in gigaporous polymeric microspheres. Bioprocess Biosyst Eng. 2015; 38: 2107-2115. [CrossRef]
  35. Hills G. Industrial use of lipases to produce fatty acid esters. Eur J Lipid Sci Technol. 2003; 105: 601-607. [CrossRef]
  36. Hasan F, Shah AA, Hameed A. Industrial applications of microbial lipases. Enzyme Microb Technol. 2006; 39: 235-251. [CrossRef]
  37. Dong H, Li J, Li Y, Hu L, Luo D. Improvement of catalytic activity and stability of lipase by immobilization on organobentonite. Chem Eng J. 2012; 181: 590-596. [CrossRef]
  38. Chandra P, Singh R, Arora PK. Microbial lipases and their industrial applications: A comprehensive review. Microb Cell Factories. 2020; 19: 169. [CrossRef]
  39. Mahale PK, Desai SV, Hombalimath VS, Achappa S. Isolation, screening and characterization of lipase producing strain from oil contaminated soil of Hubballi, Karnataka. Int J Basic Appl Biol. 2014; 2: 198-201.
  40. Javed S, Azeem F, Hussain S, Rasul I, Siddique MH, Riaz M, et al. Bacterial lipases: A review on purification and characterization. Prog Biophys Mol Biol. 2018; 132: 23-34. [CrossRef]
  41. Fatima S, Ajmal R, Badr G, Khan RH. Harmful effect of detergents on lipase. Cell Biochem Biophys. 2014; 70: 759-763. [CrossRef]
  42. Colla LM, Primaz AL, Benedetti S, Loss RA, Lima MD, Reinehr CO, et al. Surface response methodology for the optimization of lipase production under submerged fermentation by filamentous fungi. Braz J Microbiol. 2016; 47: 461-467. [CrossRef]
  43. Patnala HS, Kabilan U, Gopalakrishnan L, Rao RM, Kumar DS. Marine fungal and bacterial isolates for lipase production: A comparative study. Adv Food Nutr Res. 2016; 78: 71-94. [CrossRef]
  44. Ferraz JL, Souza LO, Silva TP, Franco M. Obtaining of microbial lipase: A brief review. Recen. 2018; 20: 30-53. [CrossRef]
  45. Colla LM, Ficanha AM, Rizzardi J, Bertolin TE, Reinehr CO, Costa JA. Production and characterization of lipases by two new isolates of Aspergillus through solid-state and submerged fermentation. BioMed Res Int. 2015; 2015: 725959. [CrossRef]
  46. Silva FA, Kopp W, Giordano RLC. Extração de lipase de mamona (Ricinus communis L.) em fase aquosa. Blucher Chem Eng Proc. 2015; 1: 1754-1761. [CrossRef]
  47. Avelar MH, Cassimiro DM, Santos KC, Domingues RC, de Castro HF, Mendes AA. Hydrolysis of vegetable oils catalyzed by lipase extract powder from dormant castor bean seeds. Ind Crops Prod. 2013; 44: 452-458. [CrossRef]
  48. Chen CC, Gao GJ, Kao AL, Tsai CT, Tsai ZC. Two novel lipases purified from rice bran displaying lipolytic and esterification activities. Int J Biol Macromol. 2019; 139: 298-306. [CrossRef]
  49. Amid M, Manap MY, Hussin M, Mustafa S. A novel aqueous two phase system composed of surfactant and xylitol for the purification of lipase from pumpkin (Cucurbita moschata) seeds and recycling of phase components. Molecules. 2015; 20: 11184-11201. [CrossRef]
  50. Huang WC, Chen CY, Wu SJ. Almond skin polyphenol extract inhibits inflammation and promotes lipolysis in differentiated 3T3-L1 adipocytes. J Med Food. 2017; 20: 103-109. [CrossRef]
  51. Bavandi H, Habibi Z, Yousefi M. Porcine pancreas lipase as a green catalyst for synthesis of bis-4-hydroxy coumarins. Bioorg Chem. 2020; 103: 104139. [CrossRef]
  52. Zarai Z, Ali MB, Fendri A, Louati H, Mejdoub H, Gargouri Y. Purification and biochemical properties of Hexaplex trunculus digestive lipase. Process Biochem. 2012; 47: 2434-2439. [CrossRef]
  53. Weidlich S, Hoffmann KH, Woodring J. Secretion of lipases in the digestive tract of the cricket Gryllus bimaculatus. Arch Insect Biochem Physiol. 2015; 90: 209-217. [CrossRef]
  54. Rueda-López S, Martínez-Montaño E, Viana MT. Biochemical characterization and comparison of pancreatic lipases from the Pacific bluefin tuna, Thunnus orientalis; totoaba, Totoaba macdonaldi; and striped bass, Morone saxatilis. J World Aquac Soc. 2017; 48: 156-165. [CrossRef]
  55. Lima LG, Gonçalves MM, Couri S, Melo VF, Sant’Ana GC, Costa AC. Lipase production by Aspergillus niger C by submerged fermentation. Braz Arch Biol Technol. 2019; 62: e19180113. [CrossRef]
  56. Cujilema-Quitio MC, León-Revelo G, Rizo Porro M, Taramona Ruiz L, Ramos-Sánchez LB. Producción de lipasas por fermentación sólida con Aspergillus niger: Influencia del PH. Cen Az. 2018; 45: 1-9.
  57. Pedroza-Padilla CJ, Romero-Tabarez MA, Orduz S. Actividad lipolítica de microorganismos aislados de aguas residuales contaminadas con grasas. Biotecnol Sect Agropecu Agroind. 2017; 15: 36-44. [CrossRef]
  58. Aguiar GP, Martins VG, Martins PC, Boschero RA, Prentice-Hernández C. Produção de lipase microbiana a partir de resíduos de corvina. Rev Eng Tecnol. 2018; 10: áginas-118.
  59. Piscitelli A, Tarallo V, Guarino L, Sannia G, Birolo L, Pezzella C. New lipases by mining of Pleurotus ostreatus genome. PloS One. 2017; 12: e0185377. [CrossRef]
  60. Rodrigues C, Cassini ST, Antunes PW, Keller RP, Gonçalves RF. Isolamento e seleção de fungos produtores de lipases com base na atividade lipásica e no potencial hidrolítico sobre óleo comestível de soja e escuma de caixa de gordura. Eng Sanit Ambientv. 2016; 21: 507-518. [CrossRef]
  61. Rodrigues ML, da Silva EA, Borba CE, Oliveira AC, Kruger C, Raimundo RW, et al. Produção de enzimas hidrolíticas pelo fungo endofítico Penicillium sp. isolado das folhas de Ricinus communis L. Rev Bras Energ Renováveis. 2015; 4: 129-145. [CrossRef]
  62. Romano IP, Santos VS, Louzada AC, Pereira Junior RC, Carmo EJ, Mota AJ, et al. Avaliação da biomassa de fungos amazônicos como fonte de lipases para biocatálise. Quím Nova. 2020; 43: 146-154. [CrossRef]
  63. Sethi BK, Nanda PK, Sahoo S. Characterization of biotechnologically relevant extracellular lipase produced by Aspergillus Terreus NCFT 4269.10. Braz J Microbiol. 2016; 47: 143-149. [CrossRef]
  64. Onofre SB, Abatti D, Refosco D, Tessaro AA, Onofre JA, Tessaro AB. Selection of filamentous fungi producing lipases from residual waters of slaughterhouses. Afr J Biotechnol. 2017; 16: 247-253. [CrossRef]
  65. Sales P, Fonseca T, Amaral F, Silva C. Production of lipase by Cunninghamella echinulata UCP 1308 in through submerged fermentation in different culture media. Engevista. 2017; 19: 1340-1351. [CrossRef]
  66. Maldonado RR, Lopes DB, Aguiar-Oliveira E, Kamimura ES, Macedo GA. A review on Geotrichum lipases: Production, purification, immobilization and applications. Chem Biochem Eng Q. 2016; 30: 439-454. [CrossRef]
  67. Louhasakul Y, Cheirsilp B, Prasertsan P. Valorization of palm oil mill effluent into lipid and cell-bound lipase by marine yeast Yarrowia lipolytica and their application in biodiesel production. Waste Biomass Valorization. 2016; 7: 417-426. [CrossRef]
  68. Soares MD, Facundes BC, Júnior AF, da Silva EM. Assessment of lipolytic activity of isolated microorganisms from the savannah of the Tocantins. Acta Sci Biol Sci. 2015; 37: 471-475. [CrossRef]
  69. Carvalho T, Finotelli PV, Bonomo RC, Franco M, Amaral PF. Evaluating aqueous two-phase systems for Yarrowia lipolytica extracellular lipase purification. Process Biochem. 2017; 53: 259-266. [CrossRef]
  70. Badgujar KC, Bhanage BM. Lipase immobilization on hyroxypropyl methyl cellulose support and its applications for chemo-selective synthesis of β-amino ester compounds. Process Biochem. 2016; 51: 1420-1433. [CrossRef]
  71. Martin LS, Ceron A, Oliveira PC, Zanin GM, de Castro HF. Different organic components on silica hybrid matrices modulate the lipase inhibition by the glycerol formed in continuous transesterification reactions. J Ind Eng Chem. 2018; 62: 462-470. [CrossRef]
  72. Fan Y, Su F, Li K, Ke C, Yan Y. Carbon nanotube filled with magnetic iron oxide and modified with polyamidoamine dendrimers for immobilizing lipase toward application in biodiesel production. Sci Rep. 2017; 7: 45643. [CrossRef]
  73. Yele VU, Desai K. A new thermostable and organic solvent-tolerant lipase from Staphylococcus warneri; optimization of media and production conditions using statistical methods. Appl Biochem Biotechnol. 2015; 175: 855-869. [CrossRef]
  74. Gricajeva A, Bendikienė V, Kalėdienė L. Lipase of Bacillus stratosphericus L1: Cloning, expression and characterization. Int J Biol Macromol. 2016; 92: 96-104. [CrossRef]
  75. Zare A, Bordbar AK, Jafarian F, Tangestaninejad S. Candida rugosa lipase immobilization on various chemically modified Chromium terephthalate MIL-101. J Mol Liq. 2018; 254: 137-144. [CrossRef]
  76. Barriuso J, Vaquero ME, Prieto A, Martínez MJ. Structural traits and catalytic versatility of the lipases from the Candida rugosa-like family: A review. Biotechnol Adv. 2016; 34: 874-885. [CrossRef]
  77. Zarinviarsagh M, Ebrahimipour G, Sadeghi H. Lipase and biosurfactant from Ochrobactrum intermedium strain MZV101 isolated by washing powder for detergent application. Lipids Health Dis. 2017; 16: 177. [CrossRef]
  78. Gutiérrez-Arnillas E, Rodríguez A, Sanromán MA, Deive FJ. New sources of halophilic lipases: Isolation of bacteria from Spanish and Turkish saltworks. Biochem Eng J. 2016; 109: 170-177. [CrossRef]
  79. Li Y, Liu TJ, Zhao MJ, Zhang H, Feng FQ. Screening, purification, and characterization of an extracellular lipase from Aureobasidium pullulans isolated from stuffed buns steamers. J Zhejiang Univ Sci B. 2019; 20: 332-342. [CrossRef]
  80. Rios NS, Pinheiro BB, Pinheiro MP, Bezerra RM, dos Santos JC, Gonçalves LR. Biotechnological potential of lipases from Pseudomonas: Sources, properties and applications. Process Biochem. 2018; 75: 99-120. [CrossRef]
  81. Haq A, Adeel S, Khan A, Khan MA, Rafiq M, Ishfaq M, et al. Screening of lipase-producing bacteria and optimization of lipase-mediated biodiesel production from jatropha curcas seed oil using whole cell approach. BioEnergy Res. 2020; 13: 1280-1296. [CrossRef]
  82. Martínez-Ruiz A, Tovar-Castro L, García HS, 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. [CrossRef]
  83. Khan MT, Kaushik AC, Malik SI, Khan AS, Wei DQ, Sajjad W, et al. Characterization and synthetic biology of lipase from Bacillus amyloliquefaciens strain. Arch Microbiol. 2020; 202: 1497-1506. [CrossRef]
  84. Cao SL, Huang YM, Li XH, Xu P, Wu H, Li N, et al. Preparation and characterization of immobilized lipase from Pseudomonas cepacia onto magnetic cellulose nanocrystals. Sci Rep. 2016; 6: 20420. [CrossRef]
  85. Aamri LE, Hafidi M, Scordino F, Krasowska A, Lebrihi A, Orlando MG, et al. Arthrographis curvata and Rhodosporidium babjevae as new potential fungal lipase producers for biotechnological applications. Braz Arch Biol Technol. 2020; 63: e20180444. [CrossRef]
  86. Martinez-Corona R, Banderas-Martínez FJ, Pérez-Castillo JN, Cortes-Penagos C, González-Hernández JC. Avocado oil as an inducer of the extracellular lipase activity of Kluyveromyces marxianus L-2029. Food Sci Technol. 2019; 40: 121-129. [CrossRef]
  87. Cairns TC, Nai C, Meyer V. How a fungus shapes biotechnology: 100 years of Aspergillus niger research. Fungal Biol Biotechnol. 2018; 5: 13. [CrossRef]
  88. Bellaouchi R, Abouloifa H, Rokni Y, Hasnaoui A, Ghabbour N, Hakkou A, et al. Characterization and optimization of extracellular enzymes production by Aspergillus niger strains isolated from date by-products. J Genet Eng Biotechnol. 2021; 19: 50. [CrossRef]
  89. U.S. Food and Drug Administration. Microorganisms & microbial-derived ingredients used in food (partial list) [Internet]. Silver Spring: U.S. Food and Drug Administration; 2018. Available from: https://www.fda.gov/food/generally-recognized-safe-gras/microorganisms-microbial-derived-ingredients-used-food-partial-list.
  90. Suyanto E, Soetarto ES, Cahyanto MN. Production and optimization of lipase by Aspergillus niger using coconut pulp waste in solid state fermentation. J Phys Conf Ser. 2019; 1374: 012005. [CrossRef]
  91. Guldhe A, Singh P, Kumari S, Rawat I, Permaul K, Bux F. Biodiesel synthesis from microalgae using immobilized Aspergillus niger whole cell lipase biocatalyst. Renew Energ. 2016; 85: 1002-1010. [CrossRef]
  92. Aliyah AN, Edelweiss ED, Sahlan M, Wijanarko A, Hermansyah H. Solid state fermentation using agroindustrial wastes to produce Aspergillus niger lipase as a biocatalyst immobilized by an adsorption-crosslinking method for biodiesel synthesis. Int J Technol. 2016; 7: 1393-1404. [CrossRef]
  93. Zulkifli NN, Rasit N. lipase production from solid state fermentation of copra waste associated fungus Aspergillus niger. UMT J Undergrad Res. 2020; 2: 33-40.
  94. Golunski SM, Mulinari J, Camargo AF, Venturin B, Baldissarelli DP, Marques CT, et al. Ultrasound effects on the activity of Aspergillus niger lipases in their application in dairy wastewater treatment. Environ Qual Manag. 2017; 27: 95-101. [CrossRef]
  95. ShanZi C, Lei C, KangMing T, MengDi L, FuPing L, ZhengXiang W. Aspergillus niger lipase CutA having flavor ester synthesis activity. Food Ferment Ind. 2019; 45: 1-5.
  96. Cong S, Tian K, Zhang X, Lu F, Singh S, Prior B, et al. Synthesis of flavor esters by a novel lipase from Aspergillus niger in a soybean-solvent system. 3 Biotech. 2019; 9: 244. [CrossRef]
  97. Putri DN, Khootama A, Perdani MS, Utami TS, Hermansyah H. Optimization of Aspergillus niger lipase production by solid state fermentation of agro-industrial waste. Energy Rep. 2020; 6: 331-335. [CrossRef]
  98. Costa TM, Hermann KL, Garcia-Roman M, Valle RD, Tavares LB. Lipase production by Aspergillus niger grown in different agro-industrial wastes by solid-state fermentation. Brazilian J Chem Eng. 2017; 34: 419-427. [CrossRef]
  99. Prabaningtyas RK, Putri DN, Utami TS, Hermansyah H. Production of immobilized extracellular lipase from Aspergillus niger by solid state fermentation method using palm kernel cake, soybean meal, and coir pith as the substrate. Energy Procedia. 2018; 153: 242-247. [CrossRef]
  100. Mukhtar H, Hanif M, Rehman A, Nawaz A, Haq IU. Studies on the lipase production by Aspergillus niger through solid state fermentation. Pak J Bot. 2015; 47: 351-354.
  101. Nema A, Patnala SH, Mandari V, Kota S, Devarai SK. Production and optimization of lipase using Aspergillus niger MTCC 872 by solid-state fermentation. Bull Natl Res Cent. 2019; 43: 82. [CrossRef]
  102. Reinehr CO, Bortoluzzi L, Morais VQ. Lipases production with hydrolytic activity by aspergillus using agroindustrial byproducts, soy oil and glycerol. Recen. 2016; 18: 97-115.
  103. Adio OQ, Kareem SO, Osho MB, Omemu AM. Production of lipases in solid-state fermentation by Aspergillus niger F7-02 With Agricultural Residues. J Microbiol Biotechnol Food Sci. 2021; 2021: 509-512. [CrossRef]
  104. dos Santos EA, Lima ÁS, Soares CM, de Aquino Santana LC. Lipase from Aspergillus niger obtained from mangaba residue fermentation: Biochemical characterization of free and immobilized enzymes on a sol-gel matrix. Acta Sci Technol. 2017; 39: 1-8. [CrossRef]
  105. Zhou X, Han Y, Lv Z, Tian X, Li H, Xie P, et al. Simultaneously achieve soluble expression and biomimetic immobilization of Candida antarctica lipase B by introducing polyamine tags. J Biotechnol. 2017; 249: 1-9. [CrossRef]
  106. Reis RL, Leão NS, Souza AF, Silva GK, Luna MA, Silva CA, et al. Evaluation of biotechnological potential of Asergillus parasiticus UCP 1281 in the wastewater biotreatment of dairy industry and lipids production. E-xacta. 2015; 8: 31-42. [CrossRef]
  107. Lall WS, Sirohi R, Prakash V. Otimização do processo para a produção de lipase por fermentação em estado sólido. World J Pharm Pharm Sci. 2014; 3: 703-712.
  108. Kishan G, Gopalakannan P, Muthukumaran C, Muthukumaresan KT, Kumar MD, Tamilarasan K. Statistical optimization of critical medium components for lipase production from Yarrowia lipolytica (MTCC 35). J Genet Eng Biotechnol. 2013; 11: 111-116. [CrossRef]
  109. Pascoal A, Estevinho LM, Martins IM, Choupina AB. Novel sources and functions of microbial lipases and their role in the infection mechanisms. Physiol Mol Plant Pathol. 2018; 104: 119-126. [CrossRef]
  110. Santos RC, de Araújo KB, Zubiolo C, Soares CM, Lima AS, de Aquino Santana LC. Microbial lipase obtained from the fermentation of pumpkin seeds: Immobilization potential of hydrophobic matrices. Acta Sci Techno. 2014; 36: 193-201. [CrossRef]
  111. Liu W, Li M, Yan Y. Heterologous expression and characterization of a new lipase from Pseudomonas fluorescens Pf0-1 and used for biodiesel production. Sci Rep. 2017; 7: 15711. [CrossRef]
  112. Alami NH, Nasihah L, Umar RL, Kuswytasari ND, Zulaika E, Shovitri M. Lipase production in lipolytic yeast from Wonorejo mangrove area. AIP Conf Proc. 2017; 1854: 020001. [CrossRef]
  113. Pereira AD, Fontes-Sant’Ana GC, Amaral PF. Mango agro-industrial wastes for lipase production from Yarrowia lipolytica and the potential of the fermented solid as a biocatalyst. Food Bioprod Process. 2019; 115: 68-77. [CrossRef]
  114. Saraswat R, Bhushan I, Gupta P, Kumar V, Verma V. Production and purification of an alkaline lipase from Bacillus sp. for enantioselective resolution of (±)-Ketoprofen butyl ester. 3 Biotech. 2018; 8: 491. [CrossRef]
  115. García-Silvera EE, Martínez-Morales F, Bertrand B, Morales-Guzmán D, Rosas-Galván NS, León-Rodríguez R, et al. Production and application of a thermostable lipase from Serratia marcescens in detergent formulation and biodiesel production. Biotechnol Appl Biochem. 2018; 65: 156-172. [CrossRef]
  116. Phukon LC, Chourasia R, Kumari M, Godan TK, Sahoo D, Parameswaran B, et al. Production and characterisation of lipase for application in detergent industry from a novel Pseudomonas helmanticensis HS6. Bioresour Technol. 2020; 309: 123352. [CrossRef]
  117. Satti SM, Abbasi AM, Marsh TL, Auras R, Hasan F, Badshah M, et al. Statistical optimization of lipase production from Sphingobacterium sp. strain S2 and evaluation of enzymatic depolymerization of poly (lactic acid) at mesophilic temperature. Polym Degrad Stab. 2019; 160: 1-13. [CrossRef]
  118. Elegado F, Legaspi CL, Paet JM, Querubin F, Tolentino JE, Vilela J, et al. Screening, identification and optimization of extracellular lipase production of yeast (Cryptococcus flavescens) isolated from a tree canopy fern in the Mount Makiling Forest Reserve, Philippines. AIP Conf Proc. 2019; 2155: 020029. [CrossRef]
  119. Penha ED, Viana LD, Gottschalk LM, Terzi SD, Souza EF, Freitas SC, et al. Aproveitamento de resíduos da agroindústria do óleo de dendê para a produção de lipase por Aspergillus Níger. Ciênc Rural. 2015; 46: 755-761. [CrossRef]
  120. Lima RT, Alves AM, de Paula AV, de Castro HF, Andrade GS. Mycelium-bound lipase from Penicillium citrinum as biocatalyst for the hydrolysis of vegetable oils. Biocatal Agric Biotechnol. 2019; 22: 101410. [CrossRef]
  121. Venkatesagowda B, Ponugupaty E, Barbosa AM, Dekker RF. Solid-state fermentation of coconut kernel-cake as substrate for the production of lipases by the coconut kernel-associated fungus Lasiodiplodia theobromae VBE-1. Ann Microbiol. 2015; 65: 129-142. [CrossRef]
  122. Hermansyah H, Andikoputro MI, Alatas A. Production of lipase enzyme from Rhizopus oryzae by solid state fermentation and submerged fermentation using wheat bran as substrate. AIP Conf Proc. 2019; 2085: 020013. [CrossRef]
  123. Neethu CS, Rahiman KM, Rosmine E, Saramma AV, Hatha AM. Utilization of agro-industrial wastes for the production of lipase from Stenotrophomonas maltophilia isolated from Arctic and optimization of physical parameters. Biocatal Agric Biotechnol. 2015; 4: 703-709. [CrossRef]
  124. Navvabi A, Razzaghi M, Fernandes P, Karami L, Homaei A. Novel lipases discovery specifically from marine organisms for industrial production and practical applications. Process Biochem. 2018; 70: 61-70. [CrossRef]
  125. Balduyck L, Veryser C, Goiris K, Bruneel C, Muylaert K, Foubert I. Optimization of a Nile Red method for rapid lipid determination in autotrophic, marine microalgae is species dependent. J Microbiol Methods. 2015; 118: 152-158. [CrossRef]
  126. Latip W, Abd Rahman RN, Leow AT, Shariff FM, Ali MS. Expression and characterization of thermotolerant lipase with broad pH profiles isolated from an Antarctic Pseudomonas sp strain AMS3. PeerJ. 2016; 4: e2420. [CrossRef]
  127. Farinas CS, Pirota R, Fonseca R, Bertucci Neto V. Desenvolvimentos em fermentação em estado sólido para produção de enzimas de interesse agroindustrial. Conceitos e aplicações da instrumentação para o avanço da agricultura. Brasília: Empraba; 2014. pp.211-241.
  128. Borowiecki P, Justyniak I, Ochal Z. Lipase-catalyzed kinetic resolution approach toward enantiomerically enriched 1-(β-hydroxypropyl) indoles. Tetrahedron Asymmetry. 2017; 28: 1717-1732. [CrossRef]
  129. Kumar A, Dhar K, Kanwar SS, Arora PK. Lipase catalysis in organic solvents: Advantages and applications. Biol Proced Online. 2016; 18: 2. [CrossRef]
  130. Negi S. Lipases: A promising tool for food industry. In: Green bio-processes. Singapore: Springer; 2019. pp.181-198. [CrossRef]
  131. Geoffry K, Achur RN. Optimization of novel halophilic lipase production by Fusarium solani strain NFCCL 4084 using palm oil mill effluent. J Genet Eng Biotechnol. 2018; 16: 327-334. [CrossRef]
  132. Lan D, Qu M, Yang B, Wang Y. Enhancing production of lipase MAS1 from marine Streptomyces sp. strain in Pichia pastoris by chaperones co-expression. Electron J Biotechnol. 2016; 22: 62-67. [CrossRef]
  133. Salihu A, Bala M, Alam MZ. Lipase production by Aspergillus niger using sheanut cake: An optimization study. J Taibah Univ Sci. 2016; 10: 850-859. [CrossRef]
  134. de Morais Junior WG, Kamimura ES, Ribeiro EJ, Pessela BC, Cardoso VL, de Resende MM. Optimization of the production and characterization of lipase from Candida rugosa and Geotrichum candidum in soybean molasses by submerged fermentation. Protein Expr Purif. 2016; 123: 26-34. [CrossRef]
  135. Dobrev G, Zhekova B, Dobreva V, Strinska H, Doykina P, Krastanov A. Lipase biosynthesis by Aspergillus carbonarius in a nutrient medium containing products and byproducts from the oleochemical industry. Biocatal Agric Biotechnol. 2015; 4: 77-82. [CrossRef]
  136. de Jesus MV, de Oliveira TS, Ferreira RD, de Lima AM, da Silva Rodrigues JR, Silva CF, et al. Produção de lipase utilizando manipueira como fonte alternativa de carbono. Scie Plena. 2016; 12. Doi: 10.14808/sci.plena.2016.054208. [CrossRef]
  137. Castiglioni GL, Costa JA, Alegre RM. Estudo da produção de lipase por Burkholderia cepacia. Eng Sanit Ambient. 2018; 23: 637-644. [CrossRef]
  138. de Castro FF, Pinheiro AB, Nassur CB, Barbosa-Tessmann IP. Mycelium-bound lipase from a locally isolated strain of Aspergillus westerdijkiae. Biocatal Agric Biotechnol. 2017; 10: 321-328. [CrossRef]
  139. Gottschalk LM, Paredes RD, Teixeira RS, Silva AS, Bon EP. Efficient production of lignocellulolytic enzymes xylanase, β-xylosidase, ferulic acid esterase and β-glucosidase by the mutant strain Aspergillus awamori 2B. 361 U2/1. Braz J Microbiol. 2013; 44: 569-576. [CrossRef]
  140. Singh R, Kumar M, Mittal A, Mehta PK. Microbial enzymes: Industrial progress in 21st century. 3 Biotech. 2016; 6: 174. [CrossRef]
  141. Salwoom L, Raja Abd Rahman RN, Salleh AB, Mohd Shariff F, Convey P, Pearce D, et al. Isolation, characterisation, and lipase production of a cold-adapted bacterial strain Pseudomonas sp. LSK25 isolated from Signy Island, Antarctica. Molecules. 2019; 24: 715. [CrossRef]
  142. Soleymani S, Alizadeh H, Mohammadian H, Rabbani E, Moazen F, Sadeghi HM, et al. Efficient media for high lipase production: One variable at a time approach. Avicenna J Med Biotechnol. 2017; 9: 82-86.
  143. Patel H, Ray S, Patel A, Patel K, Trivedi U. Enhanced lipase production from organic solvent tolerant Pseudomonas aeruginosa UKHL1 and its application in oily waste-water treatment. Biocatal Agric Biotechnol. 2020; 28: 101731. [CrossRef]
  144. Zarevúcka M. Olive oil as inductor of microbial lipase. In: Olive oil-constituents, quality, health properties and bioconversions. Rijeka: InTech Europe; 2012. pp.457-471. [CrossRef]
  145. Abol Fotouh DM, Bayoumi RA, Hassan MA. Production of thermoalkaliphilic lipase from Geobacillus thermoleovorans DA2 and application in leather industry. Enzyme Res. 2016; 2016: 9034364. [CrossRef]
  146. Hassan SW, Abd El Latif HH, Ali SM. Production of cold-active lipase by free and immobilized marine Bacillus cereus HSS: Application in wastewater treatment. Front Microbiol. 2018; 9: 2377. [CrossRef]
  147. Cortez DV, de Castro HF, Andrade GS. Potential catalytic of mycelium-bound lipase of filamentous fungi in biotransformation processes. Quím Nova. 2017; 40: 85-96.
  148. Peil GH, KuSS AV, Rave AF, VillArreAl JP, Hernandes YM, Nascente PS. Bioprospecting of lipolytic microorganisms obtained from industrial effluents. An Acad Bras Cienc. 2016; 88: 1769-1779. [CrossRef]
  149. Sperb JG, Costa TM, Vaz DA, Valle JA, Valle RD, Tavares LB. Análise qualitativa da produção de lipases e biossurfactantes por fungos isolados de resíduos oleosos. Engevista. 2015; 17: 385-397. [CrossRef]
  150. Salgado V, Fonseca C, Lopes da Silva T, Roseiro JC, Eusébio A. Isolation and identification of Magnusiomyces capitatus as a lipase-producing yeast from olive mill wastewater. Waste Biomass Valorization. 2020; 11: 3207-3221. [CrossRef]
  151. Kuncharoen N, Techo S, Savarajara A, Tanasupawat S. Identification and lipolytic activity of yeasts isolated from foods and wastes. Mycology. 2020; 11: 279-286. [CrossRef]
  152. Cazetta ML, Celligoi MA. Aproveitamento do melaço e vinhaça de cana-de-açúcar como substrato para produção de biomassa protéica e lipídica por leveduras e bactéria. Semin Ciênc Exatas Tecnol. 2005; 26: 105-112. [CrossRef]
  153. Guan C, Tao Z, Wang L, Zhao R, Chen X, Huang X, et al. Isolation of novel Lactobacillus with lipolytic activity from the vinasse and their preliminary potential using as probiotics. AMB Express. 2020; 10: 91. [CrossRef]
  154. Padma Priya R, Vasudevan N. Isolation and molecular characterization of proteolytic and lipolytic bacterial isolates from the market vegetable wastes. Int J Sci Invent Today. 2018; 7: 40-53.
  155. Ficarra FA, Santecchia I, Lagorio SH, Alarcón S, Magni C, Espariz M. Genome mining of lipolytic exoenzymes from Bacillus safensis S9 and Pseudomonas alcaliphila ED1 isolated from a dairy wastewater lagoon. Arch Microbiol. 2016; 198: 893-904. [CrossRef]
  156. Ogunnusi TA, Olorunfemi O. Isolation and identification of Proteolytic and Lipolytic Bacteria in cow dung and abattoir effluent from Ekiti general abattoir, Ekiti state, Nigeria. J Adv Microbiol. 2018; 11: 42508. [CrossRef]
  157. Hasan NA, Nawahwi MZ, Yahya N, Othman NA. Identification and optimization of lipase producing bacteria from palm oil contaminated waste. J Fundam Appl Sci. 2018; 10: 300-310.
  158. Sirisha VL, Jain A, Jain A. Enzyme immobilization: An overview on methods, support material, and applications of immobilized enzymes. Adv Food Nutr Res. 2016; 79: 179-211. [CrossRef]
  159. Sarno M, Iuliano M, Polichetti M, Ciambelli P. High activity and selectivity immobilized lipase on Fe3O4 nanoparticles for banana flavour synthesis. Process Biochem. 2017; 56: 98-108. [CrossRef]
  160. Eş I, Vieira JD, Amaral AC. Principles, techniques, and applications of biocatalyst immobilization for industrial application. Appl Microbiol Biotechnol. 2015; 99: 2065-2082. [CrossRef]
  161. Souza LT, Veríssimo LA, Pessela BC, Santoro RR, Resende RR, Mendes AA. Imobilização enzimática: princípios fundamentais e tipos de suporte. In: Biotecnologia aplicada à agro & indústria. São Paulo: Blucher; 2017. pp.529-568. [CrossRef]
  162. Cui J, Lin T, Feng Y, Tan Z, Jia S. Preparation of spherical cross-linked lipase aggregates with improved activity, stability and reusability characteristic in water-in-ionic liquid microemulsion. J Chem Technol Biotechnol. 2017; 92: 1785-1793. [CrossRef]
  163. Sheldon, Roger A. Cross-linked enzyme aggregates as industrial Biocatalysts. Org Process Res Dev. 2011; 15: 213-223. [CrossRef]
  164. Muley AB, Awasthi S, Bhalerao PP, Jadhav NL, Singhal RS. Preparation of cross-linked enzyme aggregates of lipase from Aspergillus niger: Process optimization, characterization, stability, and application for epoxidation of lemongrass oil. Bioprocess Biosyst Eng. 2021; 44: 1383-1404. [CrossRef]
  165. Podrepšek GH, Knez Ž, Leitgeb M. Activation of cellulase cross-linked enzyme aggregates (CLEAs) in scCO2. J Supercrit Fluids. 2019; 154: 104629. [CrossRef]
  166. Asgher M, Bashir F, Iqbal HM. Protease-based cross-linked enzyme aggregates with improved catalytic stability, silver removal, and dehairing potentials. Int J Biol Macromol. 2018; 118: 1247-1256. [CrossRef]
  167. Ulrich L, Mafra A, Tardioli P, Ribeiro M. Imobilização de beta-galactosidase utilizando a técnica de Cleas (cross-linked enzyme aggregates). Blucher Chem Eng Proc. 2017; 1: 1680-1685. [CrossRef]
  168. Amaral-Fonseca M, Kopp W, Giordano RD, Fernández-Lafuente R, Tardioli PW. Preparation of magnetic cross-linked amyloglucosidase aggregates: Solving some activity problems. Catalysts. 2018; 8: 496-517. [CrossRef]
  169. Hong J, Jung D, Park S, Oh Y, Oh KK, Lee SH. Immobilization of laccase via cross-linked enzyme aggregates prepared using genipin as a natural cross-linker. Int J Biol Macromol. 2021; 169: 541-550. [CrossRef]
  170. Doraiswamy N, Sarathi M, Pennathur G. Cross-linked esterase aggregates (CLEAs) using nanoparticles as immobilization matrix. Prep Biochem Biotechnol. 2019; 49: 270-278. [CrossRef]
  171. Bilal M, Iqbal HM, Guo S, Hu H, Wang W, Zhang X. State-of-the-art protein engineering approaches using biological macromolecules: A review from immobilization to implementation view point. Int J Biol Macromol. 2018; 108: 893-901. [CrossRef]
  172. Samoylova YV, Sorokina KN, Piligaev AV, Parmon VN. Preparation of stable cross-linked enzyme aggregates (CLEAs) of a Ureibacillus thermosphaericus esterase for application in malathion removal from wastewater. Catalysts. 2018; 8: 154-173. [CrossRef]
  173. Cui JD, Liu RL, Li LB. A facile technique to prepare cross-linked enzyme aggregates of bovine pancreatic lipase using bovine serum albumin as an additive. Korean J Chem Eng. 2016; 33: 610-615. [CrossRef]
  174. Mafra AC, Kopp W, Beltrame MB, Giordano RD, de Arruda Ribeiro MP, Tardioli PW. Diffusion effects of bovine serum albumin on cross-linked aggregates of catalase. J Mol Catal B Enzym. 2016; 133: 107-116. [CrossRef]
  175. Goetze D, Foletto EF, da Silva HB, Silveira VC, Dal Magro L, Rodrigues RC. Effect of feather meal as proteic feeder on combi-CLEAs preparation for grape juice clarification. Process Biochem. 2017; 62: 122-127. [CrossRef]
  176. Picó EA, López C, Cruz-Izquierdo Á, Munarriz M, Iruretagoyena FJ, Serra JL, et al. Easy reuse of magnetic cross-linked enzyme aggregates of lipase B from Candida antarctica to obtain biodiesel from Chlorella vulgaris lipids. J Biosci Bioeng. 2018; 126: 451-457. [CrossRef]
  177. Qian J, Zhao C, Ding J, Chen Y, Guo H. Preparation of nano-enzyme aggregates by crosslinking lipase with sodium tripolyphosphate. Process Biochem. 2020; 97: 19-26. [CrossRef]
  178. Jiaojiao X, Yan Y, Bin Z, Feng L. Improved catalytic performance of carrier-free immobilized lipase by advanced cross-linked enzyme aggregates technology. Bioprocess Biosyst Eng. 2022; 45: 147-158. [CrossRef]
  179. Xu MQ, Wang SS, Li LN, Gao J, Zhang YW. Combined cross-linked enzyme aggregates as biocatalysts. Catalysts. 201; 8: 460. [CrossRef]
  180. Markets and Markets. Microbial lipases market worth 590.2 million USD by 2023 [Internet]. Northbrook: Markets and Markets; 2018. Available from: https://www.marketsandmarkets.com/PressReleases/microbial-lipase.asp.
  181. Yadav AN, Verma P, Sachan SG, Saxena AK. Biodiversity and biotechnological applications of psychrotrophic microbes isolated from Indian Himalayan regions. EC Microbiol Eco. 2017; 1: 48-54.
  182. Avhad MR, Marchetti JM. Uses of enzymes for biodiesel production. In: Advanced bioprocessing for alternative fuels, biobased chemicals, and bioproducts. Sawston: Woodhead Publishing; 2019. pp.135-152. [CrossRef]
  183. Silva AG, Guidini CZ. Lipases: Sources, immobilization methods, and industrial applications. Appl Microbiol Biotechnol. 2019; 103: 7399-7423. [CrossRef]
  184. Memarpoor-Yazdi M, Karbalaei-Heidari HR, Doroodmand MM. Enantioselective hydrolysis of ibuprofen ethyl ester by a thermophilic immobilized lipase, ELT, from Rhodothermus marinus. Biochem Eng J. 2018; 130: 55-65. [CrossRef]
  185. Kumar K, Yadav AN, Kumar V, Vyas P, Dhaliwal HS. Food waste: A potential bioresource for extraction of nutraceuticals and bioactive compounds. Bioresour Bioprocess. 2017; 4: 18. [CrossRef]
  186. Su L, Hong R, Guo X, Wu J, Xia Y. Short-chain aliphatic ester synthesis using Thermobifida fusca cutinase. Food Chem. 2016; 206: 131-136. [CrossRef]
  187. de Souza CE, Ribeiro BD, Coelho MA. Characterization and application of Yarrowia lipolytica lipase obtained by solid-state fermentation in the synthesis of different esters used in the food industry. Appl Biochem Biotechnol. 2019; 189: 933-959. [CrossRef]
  188. Bayramoglu G, Celikbicak O, Kilic M, Arica MY. Immobilization of Candida rugosa lipase on magnetic chitosan beads and application in flavor esters synthesis. Food Chem. 2022; 366: 130699. [CrossRef]
  189. Ray A. Application of lipase in industry. Asian J Pharm Technol. 2012; 2: 33-37.
  190. Cao M, Fonseca LM, Schoenfuss TC, Rankin SA. Homogenization and lipase treatment of milk and resulting methyl ketone generation in blue cheese. J Agric Food Chem. 2014; 62: 5726-5733. [CrossRef]
  191. Schaffarczyk M, Østdal H, Koehler P. Lipases in wheat breadmaking: Analysis and functional effects of lipid reaction products. J Agric Food Chem. 2014; 62: 8229-8237. [CrossRef]
  192. Melis S, Pauly A, Gerits LR, Pareyt B, Delcour JA. Lipases as processing aids in the separation of wheat flour into gluten and starch: Impact on the lipid population, gluten agglomeration, and yield. J Agric Food Chem. 2017; 65: 1932-1940. [CrossRef]
  193. Sandoval G, Casas-Godoy L, Bonet-Ragel K, Rodrigues J, Ferreira-Dias S, Valero F. Enzyme-catalyzed production of biodiesel as alternative to chemical-catalyzed processes: Advantages and constraints. Curr Biochem Eng. 2017; 4: 109-141. [CrossRef]
  194. Iuliano M, Sarno M, De Pasquale S, Ponticorvo E. Candida rugosa lipase for the biodiesel production from renewable sources. Renew Energ. 2020; 162: 124-133. [CrossRef]
  195. Amini Z, Ong HC, Harrison MD, Kusumo F, Mazaheri H, Ilham Z. Biodiesel production by lipase-catalyzed transesterification of Ocimum basilicum L. (sweet basil) seed oil. Energy Convers Manag. 2017; 132: 82-90. [CrossRef]
  196. Pérez MM, Gonçalves EC, Vici AC, Salgado JC, de Moraes Polizeli MD. Fungal lipases: Versatile tools for white biotechnology. In: Recent advancement in white biotechnology through fungi. Cham: Springer; 2019. pp.361-404. [CrossRef]
  197. Costa MJ, Silva MR, Ferreira EE, Carvalho AK, Basso RC, Pereira EB, et al. Enzymatic biodiesel production by hydroesterification using waste cooking oil as feedstock. Chem Eng Process. 2020; 157: 108131. [CrossRef]
  198. Shalini P, Deepanraj B, Vijayalakshmi S, Ranjitha J. Synthesis and characterisation of lipase immobilised magnetic nanoparticles and its role as a catalyst in biodiesel production. Mater Today Proc. 2021. Doi: 10.1016/j.matpr.2021.07.027. [CrossRef]
  199. Pasha MK, Dai L, Liu D, Du W, Guo M. Biodiesel production with enzymatic technology: Progress and perspectives. Biofuel Bioprod Biorefin. 2021; 15: 1526-1548. [CrossRef]
  200. Sankaran R, Show PL, Chang JS. Biodiesel production using immobilized lipase: Feasibility and challenges. Biofuel Bioprod Biorefin. 2016; 10: 896-916. [CrossRef]
  201. Sawant G, Ghosh S, Banesh S, Bhaumik J, Banerjee UC. In silico approach towards lipase mediated chemoenzymatic synthesis of (S)-ranolazine, as an anti-anginal drug. RSC Adv. 2016; 6: 49150-49157. [CrossRef]
  202. Bhardwaj KK, Gupta R. Synthesis of chirally pure enantiomers by lipase. J Oleo Sci. 2017; 66: 1073-1084. [CrossRef]
  203. Memarpoor-Yazdi M, Karbalaei-Heidari HR, Khajeh K. Production of the renewable extremophile lipase: Valuable biocatalyst with potential usage in food industry. Food Bioprod Proces. 2017; 102: 153-166. [CrossRef]
  204. Sikora A, Chełminiak-Dudkiewicz D, Ziegler-Borowska M, Marszałł MP. Enantioseparation of (RS)-atenolol with the use of lipases immobilized onto new-synthesized magnetic nanoparticles. Tetrahedron Asymmetry. 2017; 28: 374-380. [CrossRef]
  205. Sanfilippo C, Nicolosi G, Patti A. Milnacipran as a challenging example of aminomethyl substrate for lipase-catalyzed kinetic resolution. J Mol Catal B Enzym. 2014; 104: 82-86. [CrossRef]
  206. Carvalho AC, Fonseca TD, Mattos MC, Oliveira MD, Lemos TL, Molinari F, et al. Recent advances in lipase-mediated preparation of pharmaceuticals and their intermediates. Int J Mol Sci. 2015; 16: 29682-29716. [CrossRef]
  207. Stauch B, Fisher SJ, Cianci M. Open and closed states of Candida antarctica lipase B: Protonation and the mechanism of interfacial activation1. J Lipid Res. 2015; 56: 2348-2358. [CrossRef]
  208. Sun H, Zhang H, Ang EL, Zhao H. Biocatalysis for the synthesis of pharmaceuticals and pharmaceutical intermediates. Bioorg Med Chem. 2018; 26: 1275-1284. [CrossRef]
  209. Rahman IA, Rahman RN, Salleh AB, Basri M. Formulation and evaluation of an automatic dishwashing detergent containing T1 lipase. J Surfactants Deterg. 2013; 16: 427-434. [CrossRef]
  210. Bhosale H, Shaheen U, Kadam T. Characterization of a hyperthermostable alkaline lipase from Bacillus sonorensis 4R. Enzyme Res. 2016; 2016: 4170684. [CrossRef]
  211. Agobo KU, Arazu VA, Uzo K, Igwe CN. Microbial lipases: A prospect for biotechnological industrial catalysis for green products: A review. Ferment Technol. 2017; 6: 2.
  212. Kuepethkaew S, Sangkharak K, Benjakul S, Klomklao S. Laundry detergent-stable lipase from Pacific white shrimp (Litopenaeus vannamei) hepatopancreas: Effect of extraction media and biochemical characterization. Int J food Prop. 2017; 20: 769-781. [CrossRef]
  213. Saraswat R, Verma V, Sistla S, Bhushan I. Evaluation of alkali and thermotolerant lipase from an indigenous isolated Bacillus strain for detergent formulation. Electron J Biotechnol. 2017; 30: 33-38. [CrossRef]
  214. Niyonzima FN, More SS. Microbial detergent compatible lipases. J Sci Ind Res. 2015; 74: 105-113.
  215. Hu J, Cai W, Wang C, Du X, Lin J, Cai J. Purification and characterization of alkaline lipase production by Pseudomonas aeruginosa HFE733 and application for biodegradation in food wastewater treatment. Biotechnol Biotechnol Equip. 2018; 32: 583-590. [CrossRef]
  216. Ktata A, Krayem N, Aloulou A, Bezzine S, Sayari A, Chamkha M, et al. Purification, biochemical and molecular study of lipase producing from a newly thermoalkaliphilic Aeribacillus pallidus for oily wastewater treatment. J Biochem. 2020; 167: 89-99. [CrossRef]
  217. Balaji L, Chittoor JT, Jayaraman G. Optimization of extracellular lipase production by halotolerant Bacillus sp. VITL8 using factorial design and applicability of enzyme in pretreatment of food industry effluents. Prep Biochem Biotechnol. 2020; 50: 708-716. [CrossRef]
  218. Facchini FD, Vici AC, Pereira MG, Jorge JA, de Moraes MD. Enhanced lipase production of Fusarium verticillioides by using response surface methodology and wastewater pretreatment application. J Biochem Technol. 2016; 6: 996-1002.
  219. Soares JL, Cammarota MC, Gutarra ML, Volschan I. Reduction of scum accumulation through the addition of low-cost enzymatic extract in the feeding of high-rate anaerobic reactor. Water Sci Technol. 2019; 80: 67-74. [CrossRef]
  220. Garrido JV, Silva CF. Prospecção tecnológica e científica da utilização de lipases microbianas para tratamento de efluentes. Cad Prospec. 2016; 9: 298-306. [CrossRef]
  221. Nimkande VD, Bafana A. A review on the utility of microbial lipases in wastewater treatment. J Water Process Eng. 2022; 46: 102591. [CrossRef]
  222. Ansorge-Schumacher MB, Thum O. Immobilised lipases in the cosmetics industry. Chem Soc Rev. 2013; 42: 6475-6490. [CrossRef]
  223. Beli CM, Mageste JM, Taketani NF. Bioprospecção de enzimas para cosmética: Seu impacto na biotecnologia. Rev Ensaios Pioneiros. 2019; 3: 10-24. [CrossRef]
  224. Baránková E, Dohnal V. Effect of additives on volatility of aroma compounds from dilute aqueous solutions. Fluid Phase Equilib. 2016; 407: 217-223. [CrossRef]
  225. Dias RF, Carvalho CD. Bioeconomia no Brasil e no mundo: Panorama atual e perspectivas. Rev Virtual Quím. 2017; 9: 410-430. [CrossRef]
  226. Priyanka P, Kinsella G, Henehan GT, Ryan BJ. Isolation, purification and characterization of a novel solvent stable lipase from Pseudomonas reinekei. Protein Expr Purif. 2019; 153: 121-130. [CrossRef]
  227. Nguyen HH, Lee SH, Lee UJ, Fermin CD, Kim M. Immobilized enzymes in biosensor applications. Materials. 2019; 12: 121. [CrossRef]
  228. Hasanah U, Sani ND, Heng LY, Idroes R, Safitri E. Construction of a hydrogel pectin-based triglyceride optical biosensor with immobilized lipase enzymes. Biosensors. 2019; 9: 135. [CrossRef]
  229. Zehani N, Kherrat R, Dzyadevych SV, Jaffrezic-Renault N. A microconductometric biosensor based on lipase extracted from Candida rugosa for direct and rapid detection of organophosphate pesticides. Int J Environ Anal Chem. 2015; 95: 466-479. [CrossRef]
  230. de Moura Barboza A, da Silva AB, da Silva EM, de Souza WP, Soares MA, de Vasconcelos LG, et al. A biosensor based on microbial lipase immobilized on lamellar zinc hydroxide-decorated gold nanoparticles for carbendazim determination. Anal Methods. 2019; 11: 5388-5397. [CrossRef]
  231. Sonawane A, Manickam P, Bhansali S. Stability of enzymatic biosensors for wearable applications. IEEE Rev Biomed Eng. 2017; 10: 174-186. [CrossRef]
  232. Kumar JA, Kumar MS. A study on improving dyeability of polyester fabric using lipase enzyme. Autex Res J. 2020; 20: 243-249. [CrossRef]
  233. Kalantzi S, Mamma D, Kalogeris E, Kekos D. Improved properties of cotton fabrics treated with lipase and its combination with pectinase. Fibres Text East Eur. 2010; 18: 86-92.
  234. El-Shemy NS, El-Hawary NS, El-Sayed H. Basic and reactive-dyeable polyester fabrics using lipase enzymes. J Chem Eng Process Technol. 2016; 7: 1. [CrossRef]
  235. Abou Taleb M, Gomaa SK, Wahba MI, Zaki RA, El-Fiky AF, El-Refai HA, et al. Bioscouring of wool fibres using immobilized thermophilic lipase. Int J Biol Macromol. 2022; 194: 800-810. [CrossRef]
  236. Darwish SS, Hassan ME, Ahmed HE, El Fadl MA, El Bhery MB. Evaluation of Effectiveness of Covalently immobilized α-amylase and lipase in cleaning of historical textiles. Biointerface Res Appl Chem. 2020; 11: 9952-9962. [CrossRef]
  237. Choudhury AK. Enzyme applications in textile chemical processing. In: Sustainable technologies for fashion and textiles. Sawston: Woodhead Publishing; 2020. pp.91-115. [CrossRef]
  238. Larik IA, Qazi MA, Phulpoto AH, Haleem A, Ahmed S, Kanhar NA. Stenotrophomonas maltophilia strain 5DMD: An efficient biosurfactant-producing bacterium for biodegradation of diesel oil and used engine oil. Int J Environ Sci Technol. 2019; 16: 259-268. [CrossRef]
  239. Sahoo RK, Sahu A, Subudhi E. Bioremediation of hydrocarbon using bacterial lipase from waste biomass. Iran J Sci Technol Trans A Sci. 2020; 44: 1287-1293. [CrossRef]
  240. Kumar A, Gudiukaite R, Gricajeva A, Sadauskas M, Malunavicius V, Kamyab H, et al. Microbial lipolytic enzymes-promising energy-efficient biocatalysts in bioremediation. Energy. 2020; 192: 116674. [CrossRef]
  241. Kirana S, Arshada Z, Nosheenb S, Kamala S, Gulzara T, Majeeda MS, et al. Microbial lipases: Production and applications: A review. J Biochem Biotechnol Biomater. 2016; 1: 7-20.
Newsletter
Download PDF Download Citation
0 0

TOP