Next Article in Journal
Solanum lycopersicum, a Model Plant for the Studies in Developmental Biology, Stress Biology and Food Science
Next Article in Special Issue
Prebiotic, Antipathogenic Bacteria and Hypocholesterolemia Properties of Fermented Rice Bran Extracts Derived from Black Rice and Germinated Brown Rice
Previous Article in Journal
Allicin Promoted Reducing Effect of Garlic Powder through Acrylamide Formation Stage
Previous Article in Special Issue
Recent Developments in Fermented Cereals on Nutritional Constituents and Potential Health Benefits
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Future of Structured Lipids: Enzymatic Synthesis and Their New Applications in Food Systems

1
JNU-UPM International Joint Laboratory on Plant Oil Processing and Safety, Department of Food Science and Engineering, Jinan University, 601 Huangpu Ave West, Guangzhou 510632, China
2
School of Science, Monash University Malaysia, Bandar Sunway 47500, Selangor, Malaysia
3
Guangdong Joint International Research Centre of Oilseed Biorefinery, Nutrition and Safety, Guangzhou 510632, China
*
Author to whom correspondence should be addressed.
Submission received: 15 July 2022 / Revised: 4 August 2022 / Accepted: 8 August 2022 / Published: 10 August 2022

Abstract

:
Structured lipids (SLs) refer to a new type of functional lipid obtained by modifying natural triacylglycerol (TAG) through the restructuring of fatty acids, thereby altering the composition, structure, and distribution of fatty acids attached to the glycerol backbones. Due to the unique functional characteristics of SLs (easy to absorb, low in calories, reduced serum TAG, etc.), there is increasing interest in the research and application of SLs. SLs were initially prepared using chemical methods. With the wide application of enzymes in industries and the advantages of enzymatic synthesis (mild reaction conditions, high catalytic efficiency, environmental friendliness, etc.), synthesis of SLs using lipase has aroused great interest. This review summarizes the reaction system of SL production and introduces the enzymatic synthesis and application of some of the latest SLs discussed/developed in recent years, including medium- to long-chain triacylglycerol (MLCT), diacylglycerol (DAG), EPA- and DHA-enriched TAG, human milk fat substitutes, and esterified propoxylated glycerol (EPG). Lastly, several new ways of applying SLs (powdered oil, DAG plastic fat, inert gas spray oil, and emulsion) in the future food industry are also highlighted.

1. Introduction

Lipids are among the vital nutrients needed for supporting normal physiological activities and aiding in the maintenance of human health by providing essential fatty acids. However, excessive fat consumption can lead to obesity and related diseases such as diabetes, high blood pressure, hyperlipidemia, hardening of the arteries, and heart disease. At present, obesity and cardiovascular diseases have become health threats worldwide. A variety of fat-substitute products, such as starch, cellulose, protein, pectin, and glucose, have been developed to control excessive fat consumption [1], although they have not been shown to be effective in obese individuals [2]. The lack of self-discipline on strict dietary regimes for dieters may cause weight rebounds [3]. This might be related to the low palatability of low-fat diets. As a result, developing novel oils and fats with weight loss function is of great significance in replacing traditional edible oils and fats.
Structured lipids (SLs) represent a form of oil that has unique functional properties, such as being easily absorbed, low in calories, and low in saturated fat. SLs are obtained by modifying natural triacylglycerol (TAG) through the restructuring of fatty acids, thereby altering the type, content, structure, and/or distribution of fatty acids attached to the glycerol backbone [4]. However, SL products are still in the laboratory preparation stage, and most present reaction systems are not suitable for food production, while the production cost is high. Hence, how to reduce the cost by changing the preparation method and selecting decent large-scale reactors to improve the industrialization and practical application of SLs is still a hot research topic. The SLs can be produced using chemical, enzymatic, and genetic modification methods, among which chemical methods are the primary preparation approach practiced in industries due to their low cost, ease of processing, mass production, etc. However, chemical methods have some disadvantages, such as low specificity, high energy consumption, high reaction temperature, and easy degradation of reactants and products, and the products may exhibit risk factors that are not conducive to consumer health. Therefore, researchers are actively exploring the preparation of SL using enzymatic methods with mild conditions, strong specificity, and low risk factors.
The uniqueness of SLs lies in their structural characteristics, which influence the stomach emptying rate and fat digestion, thus providing the desired physiological attributes that improve the nutritional properties [5]. SLs can reduce the accumulation of fat in the body and improve metabolic disorders such as fat malabsorption, thereby reducing the incidence of obesity and cardiovascular disease, in addition to being beneficial for fetal development. However, using SLs has some drawbacks such as poor oxidative stability and short product shelf-life. Adding synthetic antioxidants can improve the stability of SLs; however, this does not conform to the concept of healthy food. On another note, the oil’s quality and stability can be enhanced by establishing a new and stable food system for SLs that aligns well with its application and indirectly expands the diverse range of available fat and oil products. In the future, SLs could be applied in microcapsule oils, plastic fats, and food emulsions to better meet people’s health needs. Some novel application methods, such as in spray oil stored in an inert gas container, could also improve the oxidative stability.
Most of the previous reviews summarized the specific nutritional functions of some SLs, but there is less understanding of the future applications in food systems for recent SLs. In this regard, this article encompasses a review of the enzymatic production and application of recent SLs, including medium- to long-chain TAG (MLCT), diacylglycerol (DAG), eicosapentaenoic acid (EPA)- and docosahexaenoic acid (DHA)-enriched TAG, and human milk fat substitutes (HMFSs) (Figure 1). New applications of SLs in food systems in future are also discussed.

2. Reaction System for SL Production

The modification of fat and oil can be achieved through chemical, enzymatic, and genetic methods. The traditional chemical method for the preparation of SL has unique advantages, but it is not suitable to be used in food systems because of its disadvantages such as the production of dangerous factors at high temperature. Application of the genetic method is limited by the problems of market supervision and consumers’ skeptical attitude toward the products. At present, the enzymatic synthesis of SLs shows great advantages, e.g., mild reaction conditions, avoidance of the production of risk factors, and the generation of immobilized enzymes, which greatly reduce the cost of enzymatic production [4]. Table 1 shows the advantages and disadvantages of the three methods for preparing SLs.
Enzymatic synthesis takes place in a solvent-free system under relatively moderate reaction conditions. This process has the advantages of being nontoxic (no hazardous organic solvents) and easy to operate, with rapid and simple purification steps, using enzymes that can be reused for numerous cycles, which is conducive to the manufacturing of high-value products [6]. As a result, the solvent-free approach has broad development prospects for the industrial-scale enzymatic synthesis of SL.
For oils with a high viscosity coefficient, it is preferable for the reaction to be performed in an organic solvent solution. This is because the presence of organic solvents can significantly reduce the viscosity of the reaction system and increase the solubility of the oil, allowing the reaction to be carried out at a lower temperature and, thus, consuming less energy, as well as reducing the occurrence of acyl migration and hydrolysis reactions [7]. The lipase is suspended in an organic solvent and may be filtered and separated rapidly when the reaction has completed. Nevertheless, purification of the end product can be a challenging task. Most researchers used organic solvents during SL production in order to improve the substrate solubility; however, this is not suitable for food applications because of the solvent residues. However, the industrial production of edible-grade functional oils has begun to expand the use of safer, continuous, solvent-free systems.
The enzyme reactor is the core device for an enzyme-catalyzed reaction. It provides an appropriate setting for the reaction to occur under optimal enzyme-catalyzed reaction conditions, allowing the substrate to be converted into products to the greatest extent possible. Despite the fact that several studies have shown that enzymatic transesterification to synthesize SL has apparent benefits [17,18,19], production at an industrial scale is still out of favor. Successful industrial-scale production requires to be not only economically feasible, but also strongly supported by pilot plant technology. The two most popular reactors used in enzymatic transesterification are stirred tank reactor and packed-bed reactor (PBR) (Figure 2), with the former being more commonly studied. SLs with enriched DHA and ARA at sn2 were successfully produced by reacting oil substrates in a solvent-free stirred tank system [20]. Lee et al. [21] discovered that enzymatic transesterification using palm oil as the raw material produced 60% palm oil-based MLCT in a pilot-scale test. Irrespective of the scale of the reactor, stirred tanks have intrinsic benefits such as inexpensive equipment costs, a simple device, complete reaction, and controllable reaction conditions. However, the immobilized enzyme must be retrieved periodically throughout the reaction, making the conditions tedious. Furthermore, the stirring paddle has a high shear force, making it easier to shatter and break enzyme particles, which inhibits enzyme activity. PBR has been successfully used as a large-scale reactor in industrial-scale application studies. PBR is also the most extensively employed reactor for heterogeneous catalytic systems involving solid–liquid interfaces. PBR has shown its unrivalled advantages in synthesizing SL, particularly in industrial settings. The scale-up of PBR for enzymatic transesterification of edible waste oil is also feasible, as confirmed by Halim et al. [22]. PBR offers a simple design structure, high efficiency, ease of operation, better stability of the immobilized enzymes, mild reaction conditions, nontoxicity, and high recyclability [23].
Traditional enzyme reactors employ mechanical stirring or liquid flow for mass transfer, but the reaction efficiency may be affected by the reaction system’s high viscosity or the non-miscibility of reactants [24,25]. However, vigorous stirring to increase mechanical power substantially increases the risk of rupturing the enzyme carrier, thereby leading to enzyme inactivation. The concept of a bubble column reactor (Figure 2) was initially proposed by Hilterhaus et al. [26], in which an immobilized enzyme catalyzed the esterification of lauric acid and polyglycerol in a solvent-free system. The high-viscosity reaction system was gently and effectively agitated by bubbles in BCR using nitrogen. This stirring method has excellent heat and mass transfer efficiency, low destruction of enzyme support, high catalyst durability, and particular feasibility for heterogeneous systems and exothermic reactions [27]. Table 2 summarizes the reactors used for the enzymatic synthesis of SLs in recent years.

3. Enzymatic Preparation and Application of SLs

3.1. Medium—To Long-Chain Triacylglycerol (MLCT)

A new type of SL, MLCT, composed of fatty acids of medium- and long-chain lengths randomly positioned on the same glycerol backbone, was synthesized to improve the safety and efficacy of medium-chain TAG (MCT) fat emulsions and avoid the shortcomings of MCT/long-chain TAG (LCT) physically mixed fat emulsions. Due to its simultaneous supplies of medium-chain fatty acids (MCFAs; fatty acids with 6–12 carbon atoms in the carbon chain) and long-chain fatty acids (LCFAs; fatty acids with more than 12 carbon atoms in the carbon chain), MLCT maintains the advantages of MCT in terms of easy digestion and provision of a rapid source of energy, while the LCFAs contained in MLCT supplement essential fatty acids. Compared to the physical mixture, MLCT exhibits superior fatty-acid metabolism, allowing it to be cleared from the bloodstream more quickly while improving nitrogen balance without reducing liver function. It has been reported that MLCT can be used to control obesity, fat malabsorption, and other metabolic disorders, while also improving insulin resistance to regulate diabetes, as it is well regarded as a healthy cooking oil [21,37]. Nevertheless, MLCT is rarely found in nature in a high concentration, necessitating its synthesis.
MLCT is synthesized mainly through one-step or multistep enzyme-catalyzed reactions with lipase. Depending on the reaction substrate, one-step enzymatic catalysis includes acidolysis, interesterification, and esterification [21] (Figure 3). Multistep enzymatic catalysis combines two types of one-step reactions (esterification and acidolysis, esterification and acidolysis, etc.) to produce MLCT. Acidolysis, the reaction between triacylglycerol and free fatty acids, is the most widely used method for producing MLCT [37,38]. Usually, capric acid, caprylic acid, and lauric acid are utilized as the source of MCFAs, whereas corn oil, mango kernel fat, avocado oil, soybean oil, Echium seed oil, canola oil, corn oil, olive oil, and microbial oil are used to provide LCFAs. In esterification, the substrates are generally glycerol and fatty acids. Esterification provides a higher concentration of MLCT as compared to transesterification and acidolysis due to the higher purity of MCFAs and LCFAs. Nevertheless, esterification has some drawbacks in that it possesses no natural antioxidants and is costly due to the expensive fatty acids and glycerol. Interesterification is the reaction involving the exchange of two acyl groups between two ester molecules or TAGs. Most of the current studies prepared MLCTs by inter-esterifying oils rich in MCFAs (palm kernel oil, coconut oil, Cinnamomum camphora seed oil, etc.) with common vegetable oils such as rapeseed oil, soybean oil, palm oil, and Camellia oil. This process is preferred because it generates a small amount of byproducts such as free fatty acid, MAG, and DAG. Polyunsaturated fatty acids (PUFAs) are usually used as the substrate in a two-step reaction. Some studies were aimed at esterifying glycerol and PUFAs to generate tri-PUFAs, followed by acidolysis of the tri-PUFAs and MCFAs to produce MLM SLs [39]. Table 3 summarizes the enzymatic synthesis of MLCT in recent years.
Fat and oil are required to provide energy to the body and maintain human health. Moreover, food products are bestowed with unique flavors attributable to the features of different fatty acids present in oil. Hence, fats should not be entirely removed from foods in order to fulfill both the health and the technological aspects. Nevertheless, consuming foods with high-fat content can be detrimental to health. MLCT offers great potential as a replacement for traditional oils, especially in high-fat diets, to manage/prevent weight gain and body fat accumulation. Several current studies have demonstrated the potential of MLCT as an ingredient in edible oils, energy bars, margarine, shortening, and beverages. Recently, MLCT has been sought after as valuable functional oil for human milk fat substitute due to the close similarity of its triacylglycerol and fatty-acid composition with breast milk, particularly the MLL species. MLCT has become a research hotspot, and it is widely employed in the food, cosmetic, and pharmaceutical sectors as a functional oil [39,48].

3.2. Diacylglycerol (DAG)

DAG is a class of SL in which one of the fatty-acid chains on the glycerol backbone is replaced by a hydroxyl [49]. It can be found in trace amounts in natural oils and fats [50,51]. The two DAG isomers, namely, 1,3-DAG and 1,2(2,3)-DAG, are named according to the positions of hydroxyl groups at sn2 and sn3 (1), with ratios of about 7:3 and 6:4, respectively [52]. This ratio may also approach 1:1 due to the occurrence of acyl migration during deodorization and storage. Because of its symmetrical structure, 1,3-DAG is more thermodynamically stable in terms of physical and chemical characteristics. Generally, the melting point of 1,3-DAG is about 10 °C higher than that of TAG with the same fatty-acid composition, while the melting point of 1,2-DAG is similar to that of TAG with the same fatty-acid composition. Furthermore, 1,2-DAG tends to form α and β’ crystals, while 1,3-DAG is more symmetrical and tends to form stable β crystals [53]. Generally speaking, TAG products mainly focus on the type and content of fatty acids, while the development of DAG products pays attention to the nutritional properties of DAG arising from its unique structure. DAG is beneficial to health mainly because of two aspects: inhibiting the increase in postprandial serum TAG and suppressing the accumulation of body fat, mainly because of the difference in the metabolic modes compared to TAG [54]. Most 1,3-DAG is hydrolyzed into glycerol and free fatty acids, which are transported to the liver for energy by β-oxidation. Part of the intermediate product 1-MAG, which is not entirely hydrolyzed, is absorbed by small intestinal epithelial cells and resynthesized into new TAG. Nevertheless, this process is significantly less efficient than direct intake of TAG; hence, DAG plays a significant role in the prevention and treatment of obesity [55].
DAG is traditionally produced via chemical catalytic glycerolysis, which is conducted at high temperatures [56]. With the development of enzymes for food application in the industry, commercial lipases have been widely utilized to prepare DAG via partial hydrolysis, glycerolysis, esterification, and transesterification [52] (Figure 3). TAG may be over-hydrolyzed to MAG and fatty acids instead of DAG during hydrolysis, which limits the use of partial hydrolysis process for DAG production. As for glycerolysis, one of the substrates, i.e., glycerol, is easily adsorbed on the surface of the enzyme carrier, forming a hydrophilic layer that restricts the interaction between lipase and oil, limiting the reaction [57]. To ensure that the reaction proceeds forward, the esterification procedure must continually eliminate the water generated during the reaction. Moreover, glycerol and oil phases that are immiscible in glycerolysis and esterification reduce the mass transfer of the reaction process [58]. There are a variety of approaches to solve this challenge, including the introduction of organic solvents of different polarities to enhance the glycerolysis process by promoting the uniform distribution of two phases in the reaction system. Surfactants/ionic liquids, supersonic assistance, supercritical fluids, or suitable reactors can also increase the contact surface of reaction substrates with enzymes and further improve the yield [59,60,61,62]. However, these technologies are too costly and energy-intensive for industrial manufacturing, making them unsuitable for large-scale production. Table 4 summarizes the enzymatic synthesis of DAG in recent years.
Substituting DAG for normal edible oil can not only affect appetite but also inhibit weight gain due to DAG’s unique metabolic route in the human body [54,72]. As a result, DAG may be utilized to formulate foods with weight loss function that are nutritious and healthy. Bread baked with DAG-rich fat was reported to be fluffier, taste better, and be easily molded [73]. Lecithin, DAG, and other additives have been developed as promoters that may be dissolved in instant drinks to accelerate the dissolution of solids and provide products with a smooth mouthfeel. Dough prepared with DAG shortening has good oil-retaining properties, is easy to cut, and has good stability [74].
The presence of a hydroxyl group in the DAG molecule endows DAG with certain hydrophilicity. In terms of food applications, this feature may influence the retention of some flavor substances, hydrolysis during cooking, selection of food emulsifiers, and solubility of some functional components in foods. DAG can also be used in the pharmaceutical industry [75]. DAG has the function of inhibiting bile-acid secretion; it may be manufactured into tablets, powders, and capsules, as well as suppositories and injections, for the prevention and treatment of diarrhea. In addition, DAG can be mixed directly with medications, acting as a slow-release agent to regulate drug release and enhance the functionality of other active ingredients in the medication. DAG is also a synthetic raw material in the chemical industry, and it can also be utilized in skin care products to increase the skin’s water-holding capacity [76]. DAG containing special fatty acids also has a good repair effect on frostbitten skin.
Alkylglycerol (AKG) belongs to the ether lipid family. It is a hydroxyl compound of the glycerin molecule, with the sn1 or sn3 bit connected to the hydrocarbon group (alkyl or alkenyl) via an ether bond; therefore, it is also known as hydrocarbon ether glycerin, including alkyl ether glycerin and alkylene glycerin. Alkylglycerols are acylated into alkyl glycerides, whereby the hydroxyl group of the glycerol molecule sn1 is linked to the hydrocarbon group via an ether bond, while the sn2 and sn3 hydroxyl groups form DAG with two molecular fatty acids, which can be regarded as derivatives of DAG (Figure 1). Due to the different length and saturation of the alkyl group, there are many kinds of AKGs such as batyl alcohol and selachyl alcohol. Research has shown that AKG regulates the transformation of body fat in infants, thereby preventing obesity [77].

3.3. EPA and DHA-Enriched TAG

EPA and DHA are polyunsaturated fatty acids that are abundant in marine animals and plants; for example, fish oil contains high levels of EPA and DHA. The concentration of EPA and DHA in fish oil products on the market is around 18% and 12%, respectively. Several studies have shown that EPA and DHA are important for fetal development, including retinal, neuronal, and immune development. EPA and DHA may also be associated with cardiovascular function as anti-inflammation and anti-coagulant agents in managing coronary and peripheral artery disease. In addition, they have been shown to be effective in weight management and cognitive function improvement for those with mild Alzheimer’s disease [78,79]. Consumers are becoming more health-conscious, and there is a greater demand for EPA and DHA products. Hence, developing products with high concentrations of EPA and DHA is of top priority.
The enrichment of EPA and DHA in fats and oils can be achieved via physical, chemical, or enzymatic methods such as low-temperature solvent crystallization, supercritical fluid extraction, urea complexation, and silver nitrate complexation [80,81,82,83,84,85]. Via these techniques, EPA and DHA in fish oil are first hydrolyzed or trans-esterified into their free form or ethyl ester before enrichment. Because of its moderate reaction conditions, remarkable specificity, high catalytic efficiency, low energy consumption, and excellent product safety, the enzymatic approach has become a preferred method for enriching EPA and DHA in fish oil in recent years [86,87]. Table 5 summarizes the synthesis of EPA- and DHA-enriched TAG in recent years.
The primary component of natural fish oil is a TAG of mixed fatty acids. Via the positional or acyl selectivity of lipase during the lipid modification process, i.e., hydrolysis [86], transesterification, and esterification, fish oil with high concentrations of EPA and DHA can be obtained (Figure 3). The cis-carbon double-bond structure of EPA and DHA results in a highly coiled fatty-acid chain, leading to close proximity of the methyl group with the ester bond in the glycerol ester molecule, thus resulting in the active site of the lipase being difficult to be accessed by EPA, as well as an ester bond formed between DHA and glycerol. In contrast, the straight-chain structure of saturated and unsaturated fatty acids (low degree of unsaturation) does not suffer from steric hindrance or has low steric hindrance, making them more susceptible to hydrolysis. Under the activity of enzymes, the saturated and unsaturated fatty acids (low degree of unsaturation) on glycerol esters are hydrolyzed, thereby increasing the EPA and DHA contents. Transesterification refers to the lipase-catalyzed acyl exchange reaction between the glyceride and free fatty acids (EPA and DHA) or alcohol or EPA and DHA esters, with the intention of enhancing EPA and DHA. Alcoholysis, acidolysis, and transesterification are the three types of transesterification processes [100,101]. To accomplish the goal of EPA and DHA enrichment, the enzymatically catalyzed esterification process selectively catalyzes EPA and DHA into their free fatty-acid form, causing them to react with glycerol and further enrich the glycerol esters. This process fully hydrolyzes fish oil glyceride into free fatty acids, concentrates and enriches them to high concentrations of free EPA and DHA, or saponifies fish oil glyceride ethyl ester to obtain free fatty acids, before esterifying them with glycerol under lipase catalysis [102]. The lymphatic recovery of EPA and DHA given as TAG was faster and more efficient than corresponding ethyl esters and free acids, particularly shortly after administration. The limited related reports call for more studies exploring the enrichment of fish oil products with EPA and DHA in the form of TAG. Thus, economical and efficient enzymatic processes to prepare EPA- and DHA-enriched TAG should be investigated in the future.

3.4. Human Milk Fat Substitutes

Human milk fat is considered the best source of dietary fat for infants, providing 50% of the dietary energy demand for infants [103], in addition to being an important source of essential fatty acids, fat-soluble vitamins, and bioactive compounds [104]. Human milk fat TAG’s composition consists mainly of linoleic acid, palmitic acid, and oleic acid. According to the position distribution, over 70% of palmitic acid in human milk fat is found in the central position. UPU-type TAGs (unsaturated fatty acids at sn1,3 positions and palmitic acid at the sn2 position) are typical TAGs in HMF. Hence, human milk fat is abundant in 1-oleoyl-2-palmitoyl-3-oleoylglycerol (OPO) and 1-oleoyl-2-palmitoyl-3-linoleoylglycerol (OPL) [105]. The milk fat TAG structure and the distribution of fatty acids significantly impact the nutrition absorption in infants. Infants more readily absorb fats containing palmitic acid at the sn2 position under the age of 12 months with lower total soaps, palmitate soaps, and calcium [106]. The addition of OPO and OPL into infant formula can, thus, be more easily absorbed by the human body [107], suggesting a new research direction for human milk fat substitute formulation.
There are three main types of enzymatic reactions, i.e., acidolysis, transesterification, and multistep enzymatic catalysis, to synthesize OPO and OPL (Figure 3). The most direct method is the acidolysis of an oil rich in palmitic acid at the sn2 position with linoleic acid and/or oleic acid, catalyzed by widely used 1,3-specific lipases. Transesterification is a reaction between a TAG and TAG or an ester, usually occurring between oils rich in sn2 palmitoyl and sn1,3 oleoyl. Due to its simplicity, a one-step enzymatic method has been used in many studies, but it also has some disadvantages, such as the difficulty in converting intermediate DAG to HMFS and the complexity of purification due to the existence of byproducts [108,109]. To overcome these shortcomings, researchers have studied multistep enzymatic catalytic reactions (alcoholysis and esterification, acidolysis and acidolysis, transesterification and acidolysis, etc.). Firstly, the natural oil rich in sn2 palmitoyl is subjected to alcoholysis by 1,3-specific lipase, yielding sn2 palmitoyl MAG. Then, the MAG is esterified with linoleic and oleic acid to form the OPO/OPL [110]. Table 6 summarizes the synthesis of OPO/OPL in recent years.
The majority of studies focused on the OPO synthesis process. In contrast to OPO, OPL contains linoleic acid, which is an essential fatty acid that can be elongated and desaturated to form other omega-6 polyunsaturated fatty acids, such as γ-linolenic acid and arachidonic acid, which are vital for brain and retinal development during pregnancy [114]. Furthermore, LA is an indispensable structural component of certain skin ceramides, critical for maintaining the water barrier of the skin epidermis. Most vegetable oils do not contain OPL and the composition of TAG in animal fats differs from that of human milk fat; hence, it is critical to investigate OPL synthesis. Future research studies should focus on the development of OPO and OPL through enzymatic synthesis using plant-based oil resources as a healthy alternative to the currently used animal fats, which has a high cholesterol content. In addition, the enzymatic cost can be further reduced through choosing the economic lipase or using the reusable immobilized lipases.
Figure 3. Approaches for enzymatic modification to obtain SLs. (a) The main enzymatic synthesis methods of medium-long chain triacylglycerol (MLCT) [6,41,42]. (b) The main enzymatic synthesis methods of diacylglycerol (DAG) [53,63,65]. (c) The main enzymatic synthesis methods of EPA- and DHA-enriched TAG [86,88,90,92]. (d) The main enzymatic synthesis methods of human milk fat substitutes [108,109]. (R, M, L, C18C34O2, C16H32O2, C18H32O2 represent fatty acids, medium-chain fatty acids, long-chain fatty acids, oleic acid, palmitic acid, linoleic acid, respectively).
Figure 3. Approaches for enzymatic modification to obtain SLs. (a) The main enzymatic synthesis methods of medium-long chain triacylglycerol (MLCT) [6,41,42]. (b) The main enzymatic synthesis methods of diacylglycerol (DAG) [53,63,65]. (c) The main enzymatic synthesis methods of EPA- and DHA-enriched TAG [86,88,90,92]. (d) The main enzymatic synthesis methods of human milk fat substitutes [108,109]. (R, M, L, C18C34O2, C16H32O2, C18H32O2 represent fatty acids, medium-chain fatty acids, long-chain fatty acids, oleic acid, palmitic acid, linoleic acid, respectively).
Foods 11 02400 g003aFoods 11 02400 g003b

3.5. Esterified Propoxylated Glycerol (EPG)

Esterified propoxylated glycerol (EPG) is an alternative fat consisting of acetyl epoxides and polyols. It is first generated by conjugating glycerol and propylene oxide under alkaline conditions, followed by esterification with fatty acid. Its propoxy group is bound between the glycerol backbone and the fatty acid (Figure 4). Studies have shown that EPG is nontoxic to humans in the context of expected levels of intake and does not accumulate in humans after ingestion [115]; it has been certified by the US FDA as generally recognized as safe (GRAS) [116]. EPG has both the rheological properties of liquid oil and the texture properties of solid oil, along with good palatability. The state of the two phases in EPG depends on the average propoxylation degree (the number of propylene glycol multimers between the glycerol backbone and fatty acid) and the fatty-acid composition, whereby a higher average propoxylation degree results in a more liquid EPG [115].
The similarity of the structural characteristics of EPG and TAG can be used to predict their physicochemical properties. The intermediate metabolite 2-MAG produced after digestion of TAG is finally converted into TAG chylomicrons that increase the metabolic burden [117], while the propylene glycol group that exists between the glycerol backbone and acyl fatty acids in EPG is separated from the free fatty acids after digestion. The methyl group blocks lipase from accessing the ester bond [115], preventing the formation of 1-MAG, thereby avoiding the generation of TAG chylomicrons. For production applications, EPGs with a high average degree of propoxylation are less sensitive to lipases and, thus, may have more beneficial effects in human metabolism. Esterified propoxyglycerol, based on non-GMO vegetable oils (primarily canola), was invented to reduce fat; it is utilized in the preparation of candy, chocolate bars, and baked goods with reduced calories, as well as other applications, e.g., snack foods (chips, corn-flavored snacks, and chicken nuggets), vegetable protein products, beverages, coffee, tea, and dairy analogs. It can replace 50–85% of fat in most applications, reducing calories by up to 45%. At present, there are fewer reports on EPG synthesis and applications. In the future, EPG can be used in many fields as a new lipid substitute to traditional oils. Therefore, the enzymatic synthesis, characteristics, and application properties of EPG used in different food systems deserve to be studied by researchers in future work.

4. Future Applications of SLs in Food Systems

Most SLs are made up of unsaturated fatty acids that get easily oxidized. As a result, the oxidation issue has emerged as the most significant challenge for their application and storage. Meanwhile, many dietary supplements have a limited absorption rate when consumed directly. The absorption efficacy of certain fat-soluble functional elements by the human body via novel food systems is an important fundamental aspect for food scientists, which remains a research challenge. In recent years, SLs as functional food ingredients with biological activity and health-promoting functions have steadily attracted increasing attention. Henceforth, this section briefly introduces several potential applications of structural oils.

4.1. Powdered Oil

Microencapsulated oil, commonly known as powdered oil, is obtained by wrapping a core material with a water-soluble wall material. It is an oil-in-water (O/W) product made using vegetable oil, maize syrup, high-quality protein, stabilizer, emulsifier, and other auxiliary components using high-tech microcapsule technology developed from the concepts of bioengineering and food engineering. At present, the methods for preparing microencapsulated oil include freeze-drying, spray-drying, complex coacervation, and electrostatic spraying.
In general, the major difficulty in producing healthy oils for food applications is associated with the susceptibility of oils to oxidation. Light, heat, oxygen, etc. can lead to rapid oxidative degradation of oily food. At the same time, spoilage caused by the oxidation of unsaturated fatty acids affects their subsequent application in the food system; for example, the oxidation of fish oil affects the product quality, flavor, color, and shelf-life. The microencapsulation of oils can effectively solve the above problems through the preparation of stable powders with excellent performance. Due to the presence of the wall material layer, which isolates the oil from the external environment, the shelf-stability of the microencapsulated oil product is enhanced. Because the oil particles are embedded in the capsule wall, the product possesses a variety of unique features. Attributed to their superior solubility, fluidity, and small particle size, powdered oil products are widely used in the production of infant dairy products, formula milk powder, milk-containing drinks, cakes, cold food, drinks, pasta, candy, meat products, etc., thereby facilitating their digestion and absorption in the human digestive system [118,119,120]. At present, there are few reports on the microencapsulation of SLs, with more studies focused on the microencapsulation of traditional oils. In the future, SLs, which are easily oxidized and unstable, can be prepared into microcapsule oils to improve their availability. Table 7 summarizes the preparation of powdered oil in recent years.

4.2. DAG-Based Plastic Fat

Baking oil is a kind of specialty oil used in the food industry. Emulsification and kneading are used to create a water-in-oil (W/O) emulsion system from base oil. The oil level is typically about 80%, and the quality of the oil directly influences the color, flavor, and taste of baked goods. DAG made from typical unsaturated vegetable oils is easily oxidized; however, DAG made from high-melting-point vegetable oils such as palm oil have higher durability. DAG also has a greater melting point and broader plasticity than TAG with the same fatty-acid makeup, and it may be used to replace high-melting-point solid fat in specialty oils. DAG has been used in baking oil, mainly due to its amphiphilicity, lower interfacial tension between oil and water, and ease of emulsification. Although it is utilized as a food additive or a crystallization inhibitor [53], there are few accounts of DAG being employed as the primary stearin base material. The crisping function of baking oil is based on solid-phase oil. DAG’s high melting point provides a solid foundation for its polymeric skeleton in a W/O emulsion. The universality of DAG and ordinary edible oil makes it practicable to convert DAG into plastic fat products with beneficial functional properties to replace hydrogenated and saturated fat.
The majority of DAG-rich oil research focused on improving the process, while numerous studies were aimed at improving the process optimization and physiological functions; however, there is a scarcity of studies exploring the physical and chemical properties of DAG-enriched oils, as well as the impact of processing conditions on the DAG molecular structure [54,59,129,130]. There are limited studies looking at the influence of the molecular structure, physicochemical characteristics, crystallization behavior, and interaction mechanism of high-melting-point DAG on W/O emulsion systems. As a result, understanding the crystallization and emulsification mechanisms of DAG-enriched oils and fats is critical for developing a baking oil system with good stability and operability, excellent nutrition profile, and low degree of saturation, which is critical for the food industry’s transformation and development.

4.3. Inert Gas Spray Oil

Using traditional cooking oil enables us to evenly and quantitatively control the oil usage, whereas oil spray allows the edible oil to be evenly spread out, thereby aiding in controlling and reducing the intake of dietary oil, which is beneficial to health. The use of oil spray originated from a Western cooking habit; currently, the demand for edible spray oil in the world is gradually increasing. The design of edible spray oil allows the application of a controllable quantity of oil. As the packaging of edible spray oil is generally not too large with good sealing performance, it is convenient to carry around. Edible spray oil uses a binary packaging design, which is a multiprotection material packaging system formed by valves, multilayer vacuum bags, and aluminum cans, which are explosion-proof and fireproof, while preventing the leakage of oil (Figure 5). Because of the system’s opacity and good sealing performance, the product’s exposure to air and ultraviolet rays is reduced, hindering the oil’s deterioration. An inert gas pressure container can be proposed to improve the oxidative stability as an alternative to microencapsulation, which also could improve the quality and stability of unsaturated SLs. This will allow improving the atomization efficiency and enhancing the sealing effect of spray oil, in addition to preventing exposure to oxygen and prolonging the shelf-life.

4.4. Emulsion

An emulsion is made up of two incompatible phases (generally oil and water), wherein one of the spherical droplets is dispersed in the other phase. According to the difference between the continuous phase and dispersed phase, emulsions can be divided into O/W and W/O. At present, emulsion research mainly involves microemulsions, nanoemulsions, Pickering emulsions, high-internal-phase emulsions, etc. The application of unsaturated SLs and other functional bioactives in the food industry is limited due to their inherent properties; for example, carotenoids, polyphenols, unsaturated fatty acids, etc. are often poorly soluble in water, sensitive to oxygen, and easy to degrade under light and heat conditions. In recent years, utilizing emulsion systems to carry functional bioactives via SL delivery systems has emerged as an attractive approach with fascinating prospects to improve their biological efficacy (Figure 6).
It was found that microemulsions, nanoemulsions, and Pickering emulsions can be prepared to load DHA/EPA, regulate their release in the gastrointestinal tract, protect them from adverse environmental effects, and improve their solubility and bioavailability [131]. The preparation of these emulsions usually requires the use of artificial surfactants, but the use of synthetic ingredients may lead to environmental damage and adverse health effects on consumers; therefore, it is necessary to design and select green, healthy natural surfactants to prepare emulsions. Liu and Li [132,133] found that medium- to long-chain DAG can be used to produce Pickering emulsions and high-internal-phase emulsions, thereby improving the stability of the emulsion, which enables the use of DAG emulsions in the preparation of low-calorie margarine, spreadable sauce, and cosmetics. Guo et al. [134] recently used an MLCT-based emulsion system loaded with vitamin D to perform in vitro and in vivo digestion experiments, which demonstrated the improved bioavailability of vitamin D. Studies have demonstrated the possibility of preparing emulsions with diglycerol as an emulsifier and preparing emulsions loaded with easily oxidized structural oils. Currently, the application of SLs in emulsion systems is attracting increasing attention. In the future, researchers could explore the properties of SLs in emulsion systems for the transport of nutrients, so as to improve their intake, which is conducive to achieving a healthy diet.

5. Conclusions

Enzymatic synthesis shows great advantages in the preparation of SLs. The enzyme reactor is the core device of an enzyme-catalyzed reaction, and selecting the appropriate reactor is beneficial to the performance of the reaction. Further research should be carried out to achieve the large-scale application of PBR and CBR in industry. Novel oils such as MLCT, DAG, EPA- and DHA-enriched TAG, human milk fat substitutes, and EPG can be prepared through esterification, transesterification, acidolysis, and their combinations. In the future, it will be crucial to explore ways to extend the shelf-life of easily oxidized and unstable structural functional lipids by applying SLs to some food systems such as powdered oil, plastic fat, and emulsions, while also improving their functional characteristics and maximizing their nutritional properties. The application of high-melting-point DAG as a novel base stock has large potential value in the production of specialty fats with low saturated fat content, high stability, and decent operability. As compared to other oils with a higher melting point, the utilization of MLCT with a lower melting point, as well as DHA- and EPA-enriched TAG, represents a new research pathway for the development of novel carrier systems essential for the delivery of healthy functional elements in the future.

Author Contributions

J.Z.: Software, Investigation, Writing—Original draft preparation. Y.-Y.L.: Validation, Formal analysis, Writing—Review & Editing. Y.M.: Validation, Formal analysis, Visualization. Y.W.: Resources, Writing—Review & Editing. Z.Z.: Conceptualization, Methodology, Investigation, Writing—Original draft preparation, Supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China under grant 32001734, and the Department of Science and Technology of Guangdong Province under the grant 2020A1515111153, 2020B020226010 and 2016YT03H132.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

We thank the National Natural Science Foundation of China, and the Department of Science and Technology of Guangdong Province for financial support, and Xiaodong Xie, Xiaoxue Sun, and Hongzeng Ai for assistance in collecting referenced data.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wylie-Rosett, J. Fat substitutes and health: An advisory from the Nutrition Committee of the American Heart Association. Circulation. 2002, 105, 2800–2804. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Heshka, S.; Anderson, J.W.; Atkinson, R.L.; Greenway, F.L.; Hill, J.O.; Phinney, S.D.; Pi-Sunyer, F.X. Weight loss with self-help compared with a structured commercial program: A randomized trial. JAMA. 2003, 289, 1792–1798. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Takase, H. Metabolism of diacylglycerol in humans. Asia Pac. J. Clin. Nutr. 2007, 16 (Suppl. S1), 398–403. [Google Scholar] [PubMed]
  4. Guo, Y.; Cai, Z.; Xie, Y.; Ma, A.; Zhang, H.; Rao, P.; Wang, Q. Synthesis, physicochemical properties, and health aspects of structured lipids: A review. Compr. Rev. Food Sci. Food Saf. 2020, 19, 759–800. [Google Scholar] [CrossRef] [PubMed]
  5. Mu, H.; Porsgaard, T. The metabolism of structured triacylglycerols. Prog. Lipid Res. 2005, 44, 430–448. [Google Scholar] [CrossRef]
  6. Yang, K.; Bi, Y.; Sun, S.; Yang, G.; Ma, S.; Liu, W. Optimisation of N ovozym-435-catalysed esterification of fatty acid mixture for the preparation of medium-and long-chain triacylglycerols (MLCT) in solvent-free medium. Int. J. Food Sci. Technol. 2014, 49, 1001–1011. [Google Scholar] [CrossRef]
  7. Abd Razak, N.N.; Pérès, Y.; Gew, L.T.; Cognet, P.; Aroua, M.K. Effect of reaction medium mixture on the lipase catalyzed synthesis of diacylglycerol. Ind. Eng. Chem. Res. 2020, 59, 9869–9881. [Google Scholar] [CrossRef]
  8. Sitompul, J.P.; Gusdinar, T.; Anggadiredja, K.; Rahman, H. Synthesis of Structured Triacylglycerols Based on Canarium Oil for Food Application. J. Eng. Technol. Sci. 2018, 50, 87–98. [Google Scholar] [CrossRef] [Green Version]
  9. Huang, J.; Lu, Y.; Jin, J.; Song, Z.; Tang, J. Chemical transesterification of flaxseed oil and medium-chain triacylglycerols: MLCT yield, DAG content, physicochemical properties, minor compounds and oxidation stability. Int. J. Food Sci. Technol. 2021, 56, 5160–5167. [Google Scholar] [CrossRef]
  10. Guebara, S.A.B.; Ract, J.N.R.; Vitolo, M. Conversion of caprylic acid into mono-, di-and triacylglycerols using immobilized lipase. Arab. J. Sci. Eng. 2018, 43, 3631–3637. [Google Scholar] [CrossRef]
  11. Bornscheuer, U.T. Enzymes in lipid modification. Annu. Rev. Food Sci. Technol. 2018, 9, 85–103. [Google Scholar] [CrossRef] [PubMed]
  12. Utama, Q.D.; Sitanggang, A.B.; Adawiyah, D.R.; Hariyadi, P. Lipase-catalyzed interesterification for the synthesis of medium-long-medium (MLM) structured lipids–A review. Food Technol. Biotechnol. 2019, 57, 305. [Google Scholar] [CrossRef]
  13. Phuah, E.T.; Tang, T.K.; Lee, Y.Y.; Choong, T.S.Y.; Tan, C.P.; Lai, O.M. Review on the current state of diacylglycerol production using enzymatic approach. Food Bioprocess Technol. 2015, 8, 1169–1186. [Google Scholar] [CrossRef]
  14. Feltes, M.M.C.; de Oliveira, D.; Block, J.M.; Ninow, J.L. The production, benefits, and applications of monoacylglycerols and diacylglycerols of nutritional interest. Food Bioprocess Technol. 2013, 6, 17–35. [Google Scholar] [CrossRef]
  15. van Erp, H.; Bryant, F.M.; Martin-Moreno, J.; Michaelson, L.V.; Bhutada, G.; Eastmond, P.J. Engineering the stereoisomeric structure of seed oil to mimic human milk fat. Proc. Natl. Acad. Sci. USA 2019, 116, 20947–20952. [Google Scholar] [CrossRef] [Green Version]
  16. Zam, W. Structured lipids: Methods of production, commercial products and nutraceutical characteristics. Prog. Nutr. 2015, 17, 198–213. [Google Scholar]
  17. Wang, X.; Xiao, J.; Zou, W.; Han, Z.; Jin, Q.; Wang, X. Improved enzymatic synthesis route for highly purified diacid 1,3-diacylglycerols. Process Biochem. 2015, 50, 388–394. [Google Scholar] [CrossRef]
  18. Utama, Q.D.; Sitanggang, A.B.; Adawiyah, D.R.; Hariyadi, P. Lipase-catalyzed synthesis of medium-long-medium-type of structured lipids from refined bleached deodorized olein. Appl. Food Biotechnol. 2020, 7, 85–94. [Google Scholar]
  19. Tang, W.; Wang, X.; Huang, J.; Jin, Q.; Wang, X. A novel method for the synthesis of symmetrical triacylglycerols by enzymatic transesterification. Bioresour. Technol. 2015, 196, 559–565. [Google Scholar] [CrossRef]
  20. Álvarez, C.A.; Akoh, C.C. Preparation of infant formula fat analog containing capric acid and enriched with DHA and ARA at the sn-2 position. J. Am. Oil Chem. Soc. 2016, 93, 531–542. [Google Scholar] [CrossRef]
  21. Lee, Y.Y.; Tang, T.K.; Chan, E.S.; Phuah, E.T.; Lai, O.M.; Tan, C.P.; Tan, J.S. Medium chain triacylglycerol and medium-and long chain triacylglycerol: Metabolism, production, health impacts and its applications–a review. Crit. Rev. Food Sci. Nutr. 2022, 62, 4169–4185. [Google Scholar] [CrossRef] [PubMed]
  22. Halim, S.F.A.; Hassan, H.; Amri, N.; Bashah, N.A.A. Transesterification catalyzed by Lipozyme TLIM for biodiesel production from low cost feedstock. In AIP Conference Proceedings; AIP Publishing LLC: Melville, NY, USA, 2015; Volume 1660, p. 070089. [Google Scholar]
  23. Zhang, Z.; Zhang, S.; Lee, W.J.; Lai, O.M.; Tan, C.P.; Wang, Y. Production of Structured triacylglycerol via enzymatic interesterification of medium-chain triacylglycerol and soybean oil using a pilot-scale solvent-free packed bed reactor. J. Am. Oil Chem. Soc. 2020, 97, 271–280. [Google Scholar] [CrossRef]
  24. Wang, Y.; Zhao, M.; Ou, S.; Song, K.; Han, X. Preparation of diacylglycerol-enriched palm olein by phospholipase A1-catalyzed partial hydrolysis. Eur. J. Lipid Sci. Technol. 2009, 111, 652–662. [Google Scholar] [CrossRef]
  25. Baeza-Jiménez, R.; Miranda, K.; García, H.S.; Otero, C. Lipase-catalyzed glycerolysis of fish oil to obtain diacylglycerols. Grasas Y Aceites. 2013, 64, 3. [Google Scholar]
  26. Hilterhaus, L.; Thum, O.; Liese, A. Reactor concept for lipase-catalyzed solvent-free conversion of highly viscous reactants forming two-phase systems. Org. Process Res. Dev. 2008, 12, 618–625. [Google Scholar] [CrossRef]
  27. Chen, J.; Lee, W.J.; Qiu, C.; Wang, S.; Li, G.; Wang, Y. Immobilized Lipase in the Synthesis of High Purity Medium Chain Diacylglycerols Using a Bubble Column Reactor: Characterization and Application. Front. Bioeng. Biotechnol. 2020, 8, 466. [Google Scholar] [CrossRef]
  28. Kavadia, M.R.; Yadav, M.G.; Odaneth, A.A.; Lali, A.M. Synthesis of designer triacylglycerols by enzymatic acidolysis. Biotechnol. Rep. 2018, 18, e00246. [Google Scholar] [CrossRef]
  29. Diwan, B.; Gupta, P. Synthesis of MCFA and PUFA rich oils by enzymatic structuring of flax oil with single cell oils. LWT 2020, 133, 109928. [Google Scholar] [CrossRef]
  30. Moreira, D.K.T.; de Pádua Gandra, R.L.; Zuin, J.C.; Ract, J.N.R.; Ribeiro, A.P.B.; Macedo, J.A.; Macedo, G.A. Synthesis and characterization of structured lipid rich in behenic acid by enzymatic interesterification. Food Bioprod. Process. 2020, 122, 303–310. [Google Scholar] [CrossRef]
  31. Kowalska, M.; Woźniak, M.; Krzton-Maziopa, A.; Tavernier, S.; Pazdur, Ł.; Żbikowska, A. Development of the emulsions containing modified fats formed via enzymatic interesterification catalyzed by specific lipase with various amount of water. J. Dispers. Sci. Technol. 2019, 40, 192–205. [Google Scholar] [CrossRef]
  32. Bassan, N.; Rodrigues, R.H.; Monti, R.; Tecelão, C.; Ferreira-Dias, S.; Paula, A.V. Enzymatic modification of grapeseed (Vitis vinifera L.) oil aiming to obtain dietary triacylglycerols in a batch reactor. LWT 2019, 99, 600–606. [Google Scholar] [CrossRef] [Green Version]
  33. Ribeiro, M.D.; Ming, C.C.; Lopes, T.I.; Grimaldi, R.; Marsaioli, A.J.; Gonçalves, L.A. Synthesis of structured lipids containing behenic acid from fully hydrogenated Crambe abyssinica oil by enzymatic interesterification. J. Food Sci. Technol. 2017, 54, 1146–1157. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Liu, M.; Fu, J.; Teng, Y.; Zhang, Z.; Zhang, N.; Wang, Y. Fast production of diacylglycerol in a solvent free system via lipase catalyzed esterification using a bubble column reactor. J. Am. Oil Chem. Soc. 2016, 93, 637–648. [Google Scholar] [CrossRef]
  35. Zhang, Z.; Song, J.; Lee, W.J.; Xie, X.; Wang, Y. Characterization of enzymatically interesterified palm oil-based fats and its potential application as cocoa butter substitute. Food Chem. 2020, 318, 126518. [Google Scholar] [CrossRef] [PubMed]
  36. Li, G.; Chen, J.; Ma, X.; Zhang, Z.; Liu, N.; Wang, Y. Enzymatic preparation and facile purification of medium-chain, and medium-and long-chain fatty acid diacylglycerols. LWT 2018, 92, 227–233. [Google Scholar] [CrossRef] [Green Version]
  37. Korma, S.A.; Zou, X.; Ali, A.H.; Abed, S.M.; Jin, Q.; Wang, X. Preparation of structured lipids enriched with medium-and long-chain triacylglycerols by enzymatic interesterification for infant formula. Food Bioprod. Process. 2018, 107, 121–130. [Google Scholar] [CrossRef]
  38. Sousa, V.; Campos, V.; Nunes, P.; Pires-Cabral, P. Incorporation of capric acid in pumpkin seed oil by sn-1,3 regioselective lipase-catalyzed acidolysis. OCL. 2018, 25, A302. [Google Scholar] [CrossRef] [Green Version]
  39. Lee, Y.Y.; Tang, T.K.; Lai, O.M. Health benefits, enzymatic production, and application of medium-and long-chain triacylglycerol (MLCT) in food industries: A review. J. Food Sci. 2012, 77, R137–R144. [Google Scholar] [CrossRef]
  40. Koh, S.P.; Tan, C.P.; Lai, O.M.; Arifin, N.; Yusoff, M.S.A.; Long, K. Enzymatic synthesis of medium-and long-chain triacylglycerols (MLCT): Optimization of process parameters using response surface methodology. Food Bioprocess Technol. 2010, 3, 288–299. [Google Scholar] [CrossRef]
  41. Wang, Y.; Xia, L.; Xu, X.; Xie, L.; Duan, Z. Lipase-catalyzed acidolysis of canola oil with caprylic acid to produce medium-, long-and medium-chain-type structured lipids. Food Bioprod. Process. 2012, 90, 707–712. [Google Scholar] [CrossRef]
  42. Zhao, M.L.; Hu, J.N.; Zhu, X.M.; Li, H.Y.; Li, J.; Fan, Y.W.; Deng, Z.Y. Enzymatic synthesis of medium-and long-chain triacylglycerols–enriched structured lipid from C innamomum camphora seed oil and camellia oil by L ipozyme RM IM. Int. J. Food Sci. Technol. 2014, 49, 453–459. [Google Scholar] [CrossRef]
  43. Khodadadi, M.; Kermasha, S. Modeling lipase-catalyzed interesterification of flaxseed oil and tricaprylin for the synthesis of structured lipids. J. Mol. Catal. B Enzym. 2014, 102, 33–40. [Google Scholar] [CrossRef]
  44. Adhikari, P.; Shin, J.A.; Lee, J.H.; Kim, H.R.; Kim, I.H.; Hong, S.T.; Lee, K.T. Crystallization, physicochemical properties, and oxidative stability of the interesterified hard fat from rice bran oil, fully hydrogenated soybean oil, and coconut oil through lipase-catalyzed reaction. Food Bioprocess Technol. 2012, 5, 2474–2487. [Google Scholar] [CrossRef]
  45. Yang, H.; Mu, Y.; Chen, H.; Su, C.; Yang, T.; Xiu, Z. Sn-1,3-specific interesterification of soybean oil with medium-chain triacylglycerol catalyzed by Lipozyme TL IM. Chin. J. Chem. Eng. 2014, 22, 1016–1020. [Google Scholar] [CrossRef]
  46. Yuan, T.; Wei, W.; Wang, X.; Jin, Q. Biosynthesis of structured lipids enriched with medium and long-chain triacylglycerols for human milk fat substitute. LWT 2020, 128, 109255. [Google Scholar] [CrossRef]
  47. Zou, X.; Zhang, S.; Cheng, Y.; Huang, J.; He, X.; Jiang, X.; Zhang, H. Lipase-Catalyzed Interesterification of Schizochytrium sp. Oil and Medium-Chain Triacylglycerols for Preparation of DHA-Rich Medium and Long-Chain Structured Lipids. J. Am. Oil Chem. Soc. 2021, 98, 253–267. [Google Scholar] [CrossRef]
  48. Jennings, B.H.; Shewfelt, R.L.; Akoh, C.C. Food applications of a rice bran oil structured lipid in fried sweet potato chips and an energy bar. J. Food Qual. 2010, 33, 679–692. [Google Scholar] [CrossRef]
  49. Lee, W.J.; Zhang, Z.; Lai, O.M.; Tan, C.P.; Wang, Y. Diacylglycerol in food industry: Synthesis methods, functionalities, health benefits, potential risks and drawbacks. Trends Food Sci. Technol. 2020, 97, 114–125. [Google Scholar] [CrossRef]
  50. Cicero, N.; Albergamo, A.; Salvo, A.; Bua, G.D.; Bartolomeo, G.; Mangano, V.; Dugo, G. Chemical characterization of a variety of cold-pressed gourmet oils available on the Brazilian market. Food Res. Int. 2018, 109, 517–525. [Google Scholar] [CrossRef]
  51. Salvo, A.; Rotondo, A.; La Torre, G.L.; Cicero, N.; Dugo, G. Determination of 1,2/1,3-diglycerides in Sicilian extra-virgin olive oils by 1H-NMR over a one-year storage period. Nat. Prod. Res. 2017, 31, 822–828. [Google Scholar] [CrossRef]
  52. Arpi, N.; Lubis, Y.M.; Supardan, M.D.; Mustapha, W.A.W. Diacylglycerol-enriched oil production using chemical glycerolysis. Eur. J. Lipid Sci. Technol. 2016, 118, 1880–1890. [Google Scholar]
  53. Lo, S.K.; Tan, C.P.; Long, K.; Yusoff, M.; Affandi, S.; Lai, O.M. Diacylglycerol oil—Properties, processes and products: A review. Food Bioprocess Technol. 2008, 1, 223–233. [Google Scholar] [CrossRef]
  54. Devi, B.P.; Gangadhar, K.N.; Prasad, R.B.N.; Sugasini, D.; Rao, Y.P.C.; Lokesh, B.R. Nutritionally enriched 1,3-diacylglycerol-rich oil: Low calorie fat with hypolipidemic effects in rats. Food Chem. 2018, 248, 210–216. [Google Scholar] [CrossRef] [PubMed]
  55. Lee, Y.Y.; Tang, T.K.; Phuah, E.T.; Tan, C.P.; Wang, Y.; Li, Y.; Lai, O.M. Production, safety, health effects and applications of diacylglycerol functional oil in food systems: A review. Crit. Rev. Food Sci. Nutr. 2020, 60, 2509–2525. [Google Scholar] [CrossRef] [PubMed]
  56. Feltes, M.M.C.; Villeneuve, P.; Baréa, B.; Barouh, N.; De Oliveira, J.V.; De Oliveira, D.; Ninow, J.L. Enzymatic production of monoacylglycerols (MAG) and diacylglycerols (DAG) from fish oil in a solvent-free system. J. Am. Oil Chem. Soc. 2012, 89, 1057–1065. [Google Scholar] [CrossRef]
  57. Kristensen, J.B.; Xu, X.; Mu, H. Diacylglycerol synthesis by enzymatic glycerolysis: Screening of commercially available lipases. J. Am. Oil Chem. Soc. 2005, 82, 329–334. [Google Scholar] [CrossRef]
  58. Zhong, N.; Li, L.; Xu, X.; Cheong, L.Z.; Zhao, X.; Li, B. Production of diacylglycerols through low-temperature chemical glycerolysis. Food Chem. 2010, 122, 228–232. [Google Scholar] [CrossRef]
  59. Zhao, X.; Sun, Q.; Qin, Z.; Liu, Q.; Kong, B. Ultrasonic pretreatment promotes diacylglycerol production from lard by lipase-catalysed glycerolysis and its physicochemical properties. Ultrason. Sonochemistry. 2018, 48, 11–18. [Google Scholar] [CrossRef]
  60. Moquin, P.H.; Temelli, F. Production of monoolein from oleic acid and glycerol in supercritical carbon dioxide media: A kinetic approach. J. Supercrit. Fluids. 2008, 44, 40–47. [Google Scholar] [CrossRef]
  61. Feltes, M.M.C.; Vladimir Oliveira, J.; Treichel, H.; Block, J.M.; de Oliveira, D.; Ninow, J.L. Assessment of process parameters on the production of diglycerides rich in omega-3 fatty acids through the enzymatic glycerolysis of fish oil. Eur. Food Res. Technol. 2010, 231, 701–710. [Google Scholar] [CrossRef]
  62. Kahveci, D.; Guo, Z.; Özçelik, B.; Xu, X. Optimisation of enzymatic synthesis of diacylglycerols in binary medium systems containing ionic liquids. Food Chem. 2010, 119, 880–885. [Google Scholar] [CrossRef]
  63. Garba, U.; Singanusong, R.; Jiamyangyuen, S.; Thongsook, T. Preparing spray-dried cholesterol free salad dressing emulsified with enzymatically synthesized mixed mono-and diglycerides from rice bran oil and glycerol. J. Food Sci. Technol. 2021, 58, 968–977. [Google Scholar] [CrossRef]
  64. Valério, A.; Rovani, S.; Treichel, H.; de Oliveira, D.; Oliveira, J.V. Optimization of mono and diacylglycerols production from enzymatic glycerolysis in solvent-free systems. Bioprocess Biosyst. Eng. 2010, 33, 805–812. [Google Scholar] [CrossRef] [PubMed]
  65. Liu, N.; Wang, Y.; Zhao, Q.; Zhang, Q.; Zhao, M. Fast synthesis of 1,3-DAG by Lecitase® Ultra-catalyzed esterification in solvent-free system. Eur. J. Lipid Sci. Technol. 2011, 113, 973–979. [Google Scholar] [CrossRef] [Green Version]
  66. Zhong, N.; Gui, Z.; Xu, L.; Huang, J.; Hu, K.; Gao, Y.; Li, B. Solvent-free enzymatic synthesis of 1,3-Diacylglycerols by direct esterification of glycerol with saturated fatty acids. Lipids Health Dis. 2013, 12, 65. [Google Scholar] [CrossRef] [Green Version]
  67. von der Haar, D.; Stäbler, A.; Wichmann, R.; Schweiggert-Weisz, U. Enzyme-assisted process for DAG synthesis in edible oils. Food Chem. 2015, 176, 263–270. [Google Scholar] [CrossRef]
  68. Zhong, N.; Chen, W.; Liu, L.; Chen, H. Immobilization of Rhizomucor miehei lipase onto the organic functionalized SBA-15: Their enzymatic properties and glycerolysis efficiencies for diacylglycerols production. Food Chem. 2019, 271, 739–746. [Google Scholar] [CrossRef] [PubMed]
  69. Zhao, X.; Zhao, F.; Zhong, N. Production of diacylglycerols through glycerolysis with SBA-15 supported Thermomyces lanuginosus lipase as catalyst. J. Sci. Food Agric. 2020, 100, 1426–1435. [Google Scholar] [CrossRef] [PubMed]
  70. Liu, N.; Wang, Y.; Zhao, Q.; Cui, C.; Fu, M.; Zhao, M. Immobilisation of lecitase® ultra for production of diacylglycerols by glycerolysis of soybean oil. Food Chem. 2012, 134, 301–307. [Google Scholar] [CrossRef]
  71. Vázquez, L.; González, N.; Reglero, G.; Torres, C. Solvent-free lipase-catalyzed synthesis of diacylgycerols as low-calorie food ingredients. Front. Bioeng. Biotechnol. 2016, 4, 6. [Google Scholar] [CrossRef] [Green Version]
  72. Anikisetty, M.; Krishna, A.G.; Panneerselvam, V.; Kamatham, A.N. Diacylglycerol (DAG) rich rice bran and sunflower oils modulate lipid profile and cardiovascular risk factors in Wistar rats. J. Funct. Foods. 2018, 40, 117–127. [Google Scholar] [CrossRef]
  73. Cheong, L.Z.; Tan, C.P.; Long, K.; Idris, N.A.; Yusoff, M.S.A.; Lai, O.M. Baking performance of palm diacylglycerol bakery fats and sensory evaluation of baked products. Eur. J. Lipid Sci. Technol. 2011, 113, 253–261. [Google Scholar] [CrossRef]
  74. Doucet, J. Shortening System Products Therewith and Methods for Making the Same. U.S. Patent No. 5,908,655, 1 June 1999. [Google Scholar]
  75. Obitte, N.C.; Zorn, K.; Oroz-Guinea, I.; Bornscheuer, U.T.; Klein, S. Enzymatically modified Shea butter and Palm Kernel oil as potential lipid drug delivery matrices. Eur. J. Lipid Sci. Technol. 2018, 120, 1700332. [Google Scholar] [CrossRef] [Green Version]
  76. Macierzanka, A.; Szeląg, H.; Szumała, P.; Pawłowicz, R.; Mackie, A.R.; Ridout, M.J. Effect of crystalline emulsifier composition on structural transformations of water-in-oil emulsions: Emulsification and quiescent conditions. Colloids Surf. A Physicochem. Eng. Asp. 2009, 334, 40–52. [Google Scholar] [CrossRef]
  77. Yu, H.; Dilbaz, S.; Coßmann, J.; Hoang, A.C.; Diedrich, V.; Herwig, A.; Röszer, T. Breast milk alkylglycerols sustain beige adipocytes through adipose tissue macrophages. J. Clin. Investig. 2019, 129, 2485–2499. [Google Scholar] [CrossRef] [Green Version]
  78. Zhang, T.T.; Xu, J.; Wang, Y.M.; Xue, C.H. Health benefits of dietary marine DHA/EPA-enriched glycerophospholipids. Prog. Lipid Res. 2019, 75, 100997. [Google Scholar] [CrossRef]
  79. Fard, S.G.; Loh, S.P.; Turchini, G.M.; Wang, B.; Elliott, G.; Sinclair, A.J. Microencapsulated tuna oil results in higher absorption of DHA in toddlers. Nutrients. 2020, 12, 248. [Google Scholar] [CrossRef] [Green Version]
  80. Hong, S.I.; Ma, N.; No, D.S.; Choi, N.; Baik, J.Y.; Kim, C.T.; Kim, I.H. Enrichment of DHA from tuna oil in a packed bed reactor via lipase-catalyzed esterification. J. Am. Oil Chem. Soc. 2014, 91, 1877–1884. [Google Scholar] [CrossRef]
  81. Nguyen, T.T.; Zhang, W.; Barber, A.R.; Su, P.; He, S. Significant enrichment of polyunsaturated fatty acids (PUFAs) in the lipids extracted by supercritical CO2 from the livers of Australian rock lobsters (Jasus edwardsii). J. Agric. Food Chem. 2015, 63, 4621–4628. [Google Scholar] [CrossRef]
  82. Shanmugam, K.; Donaldson, A.A. Extraction of EPA/DHA from 18/12EE fish oil using AgNO3 (aq): Composition, yield, and effects of solvent addition on interfacial tension and flow pattern in mini-fluidic systems. Ind. Eng. Chem. Res. 2015, 54, 8295–8301. [Google Scholar] [CrossRef]
  83. Lin, W.; Wu, F.W.; Yue, L.; Du, Q.G.; Tian, L.; Wang, Z.X. Combination of urea complexation and molecular distillation to purify DHA and EPA from sardine oil ethyl esters. J. Am. Oil Chem. Soc. 2014, 91, 687–695. [Google Scholar] [CrossRef]
  84. Crexi, V.T.; Monte, M.L.; Monte, M.L.; Pinto, L.A. Polyunsaturated fatty acid concentrates of carp oil: Chemical hydrolysis and urea complexation. J. Am. Oil Chem. Soc. 2012, 89, 329–334. [Google Scholar] [CrossRef]
  85. Wang, W.; Li, T.; Ning, Z.; Wang, Y.; Yang, B.; Ma, Y.; Yang, X. A process for the synthesis of PUFA-enriched triacylglycerols from high-acid crude fish oil. J. Food Eng. 2012, 109, 366–371. [Google Scholar] [CrossRef]
  86. Akanbi, T.O.; Adcock, J.L.; Barrow, C.J. Selective concentration of EPA and DHA using Thermomyces lanuginosus lipase is due to fatty acid selectivity and not regioselectivity. Food Chem. 2013, 138, 615–620. [Google Scholar] [CrossRef]
  87. Akanbi, T.O.; Barrow, C.J. Lipase-catalysed incorporation of EPA into emu oil: Formation and characterisation of new structured lipids. J. Funct. Foods. 2015, 19, 801–809. [Google Scholar] [CrossRef]
  88. Zhang, Z.; Liu, F.; Ma, X.; Huang, H.; Wang, Y. Two-stage enzymatic preparation of eicosapentaenoic acid (EPA) and docosahexaenoic acid (DHA) enriched fish oil triacylglycerols. J. Agric. Food Chem. 2018, 66, 218–227. [Google Scholar] [CrossRef]
  89. Gao, K.; Chu, W.; Sun, J.; Mao, X. Identification of an alkaline lipase capable of better enrichment of EPA than DHA due to fatty acids selectivity and regioselectivity. Food Chem. 2020, 330, 127225. [Google Scholar] [CrossRef]
  90. Castejón, N.; Señoráns, F.J. Strategies for Enzymatic Synthesis of Omega-3 Structured Triacylglycerols from Camelina sativa Oil Enriched in EPA and DHA. Eur. J. Lipid Sci. Technol. 2019, 121, 1800412. [Google Scholar] [CrossRef]
  91. Li, D.; Wang, W.; Qin, X.; Li, X.; Yang, B.; Wang, Y. A novel process for the synthesis of highly pure n-3 polyunsaturated fatty acid (PUFA)-enriched triacylglycerols by combined transesterification and ethanolysis. J. Agric. Food Chem. 2016, 64, 6533–6538. [Google Scholar] [CrossRef]
  92. Xu, L.; Zhang, L.; Li, D.; Liu, P.; Tan, C.P.; Wang, W.; Wang, Y. Deep Eutectic Solvents Enable the Enhanced Production of n-3 PUFA-Enriched Triacylglycerols. Eur. J. Lipid Sci. Technol. 2017, 119, 1700300. [Google Scholar] [CrossRef]
  93. Lei, Q.; Ba, S.; Zhang, H.; Wei, Y.; Lee, J.Y.; Li, T. Enrichment of omega-3 fatty acids in cod liver oil via alternate solvent winterization and enzymatic interesterification. Food Chem. 2016, 199, 364–371. [Google Scholar] [CrossRef] [PubMed]
  94. Wang, J.; Wang, X.D.; Zhao, X.Y.; Liu, X.; Dong, T.; Wu, F.A. From microalgae oil to produce novel structured triacylglycerols enriched with unsaturated fatty acids. Bioresour. Technol. 2015, 184, 405–414. [Google Scholar] [CrossRef] [PubMed]
  95. He, Y.; Wang, X.; Zhang, Y.; Guo, Z.; Jiang, Y.; Chen, F. Enzymatic ethanolysis subjected to Schizochytrium biomass: Sequential processing for DHA enrichment and biodiesel production. Energy Convers. Manag. 2019, 184, 159–171. [Google Scholar] [CrossRef]
  96. Solaesa, Á.G.; Sanz, M.T.; Falkeborg, M.; Beltrán, S.; Guo, Z. Production and concentration of monoacylglycerols rich in omega-3 polyunsaturated fatty acids by enzymatic glycerolysis and molecular distillation. Food Chem. 2016, 190, 960–967. [Google Scholar] [CrossRef] [Green Version]
  97. Haq, M.; Park, S.K.; Kim, M.J.; Cho, Y.J.; Chun, B.S. Modifications of Atlantic salmon by-product oil for obtaining different ω-3 polyunsaturated fatty acids concentrates: An approach to comparative analysis. J. Food Drug Anal. 2018, 26, 545–556. [Google Scholar] [CrossRef] [Green Version]
  98. Naranjo, J.M.D.; Callejón, M.J.J.; Vásquez, M.P.; Rios, L.A.; Medina, A.R. Optimization of the enzymatic synthesis of structured triacylglycerols rich in docosahexaenoic acid at sn-2 position by acidolysis of Aurantiochytrium limacinum SR21 oil and caprylic acid using response surface methodology. J. Appl. Phycol. 2021, 33, 2031–2045. [Google Scholar] [CrossRef]
  99. Zou, X.; Ye, L.; He, X.; Wu, S.; Zhang, H.; Jin, Q. Preparation of DHA-rich medium-and long-chain triacylglycerols by lipase-catalyzed acidolysis of microbial oil from Schizochytrium sp. with medium-chain fatty acids. Appl. Biochem. Biotechnol. 2020, 191, 1294–1314. [Google Scholar] [CrossRef]
  100. Valverde, L.M.; Moreno, P.A.G.; Cerdán, L.E.; López, E.N.; López, B.C.; Medina, A.R. Concentration of docosahexaenoic and eicosapentaenoic acids by enzymatic alcoholysis with different acyl-acceptors. Biochem. Eng. J. 2014, 91, 163–173. [Google Scholar] [CrossRef]
  101. Hita, E.; Robles, A.; Camacho, B.; Ramírez, A.; Esteban, L.; Jiménez, M.J.; Molina, E. Production of structured triacylglycerols (STAG) rich in docosahexaenoic acid (DHA) in position 2 by acidolysis of tuna oil catalyzed by lipases. Process Biochem. 2007, 42, 415–422. [Google Scholar] [CrossRef]
  102. Yang, Z.; Jin, W.; Cheng, X.; Dong, Z.; Chang, M.; Wang, X. Enzymatic enrichment of n-3 polyunsaturated fatty acid glycerides by selective hydrolysis. Food Chem. 2021, 346, 128743. [Google Scholar] [CrossRef]
  103. Qin, X.L.; Wang, Y.M.; Wang, Y.H.; Huang, H.H.; Yang, B. Preparation and characterization of 1,3-dioleoyl-2-palmitoylglycerol. J. Agric. Food Chem. 2021, 59, 5714–5719. [Google Scholar] [CrossRef] [PubMed]
  104. Zhang, L.S.; Chu, M.Y.; Zong, M.H.; Yang, J.G.; Lou, W.Y. Efficient Production of 1, 3-Dioleoyl-2-Palmitoylglycerol through Rhodococcus opacus Fermentation. J. Am. Oil Chem. Soc. 2020, 97, 851–860. [Google Scholar] [CrossRef]
  105. Gao, L.; Cheng, X.; Yu, X.; Wang, X.; Jin, Q.; Wang, X. Lipase-mediated production of 1-oleoyl-2-palmitoyl-3-linoleoylglycerol by a two-step method. Food Biosci. 2020, 36, 100678. [Google Scholar] [CrossRef]
  106. Wang, X.; Jiang, C.; Xu, W.; Miu, Z.; Jin, Q.; Wang, X. Enzymatic synthesis of structured triacylglycerols rich in 1,3-dioleoyl-2-palmitoylglycerol and 1-oleoyl-2-palmitoyl-3-linoleoylglycerol in a solvent-free system. LWT 2020, 118, 108798. [Google Scholar] [CrossRef]
  107. Sun, C.; Wei, W.; Zou, X.; Huang, J.; Jin, Q.; Wang, X. Evaluation of triacylglycerol composition in commercial infant formulas on the Chinese market: A comparative study based on fat source and stage. Food Chem. 2018, 252, 154–162. [Google Scholar] [CrossRef]
  108. Wei, W.; Jin, Q.; Wang, X. Human milk fat substitutes: Past achievements and current trends. Prog. Lipid Res. 2019, 74, 69–86. [Google Scholar] [CrossRef]
  109. Hasibuan, H.A.; Sitanggang, A.B.; Andarwulan, N.; Hariyadi, P. Enzymatic Synthesis of Human Milk Fat Substitute-A Review on Technological Approaches. Food Technol. Biotechnol. 2021, 59, 475–495. [Google Scholar] [CrossRef]
  110. Liu, C.; Zhang, Y.; Zhang, X.; Nie, K.; Deng, L.; Wang, F. The two-step synthesis of 1,3-oleoyl-2-palmitoylglycerol by Candida sp. 99–125 lipase. J. Mol. Catal. B Enzym 2016, 133, S1–S5. [Google Scholar] [CrossRef]
  111. Cao, X.; Pan, Y.; Qiao, M.; Yuan, Y. Synthesis of human milk fat substitutes based on enzymatic preparation of low erucic acid acyl-donors from rapeseed oil. Food Chem. 2022, 387, 132907. [Google Scholar] [CrossRef]
  112. Sun, D.; Li, S.; Shang, J.; You, L.; Wang, M.; Sun, C.; Wang, X. Process optimization of simultaneous enzymatic production of 1,3-dioleoyl-2-palmitoylglycerol and 1-oleoyl-2-palmitoyl-3-linoleoylglycerol. J. Am. Oil Chem. Soc. 2021, 98, 1167–1176. [Google Scholar] [CrossRef]
  113. Gu, S.; Feng, F.; Li, Y.; Wei, W.; Lu, R. Aspergillus oryzae lipase-catalyzed synthesis of dioleoyl; palmitoyl-rich triacylglycerols in two reactors. J. Am. Oil Chem. Soc. 2016, 93, 1347–1354. [Google Scholar] [CrossRef]
  114. Saini, R.K.; Keum, Y.S. Omega-3 and omega-6 polyunsaturated fatty acids: Dietary sources, metabolism, and significance—A review. Life Sci. 2018, 203, 255–267. [Google Scholar] [CrossRef] [PubMed]
  115. Bechtel, D.H. Article series: Safety of esterified propoxylated glycerol (EPG), a nonabsorbable fat replacer. Regul. Toxicol. Pharmacol. 2014, 70, S91–S94. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. GRAS. Notice for Esterified Propoxylated Glycerol (EPG) for Use in Select Commercial Frying Applications; GRAS: Silver Spring, MD, USA, 2018.
  117. Yanai, H.; Tomono, Y.; Ito, K.; Furutani, N.; Yoshida, H.; Tada, N. Diacylglycerol oil for the metabolic syndrome. Nutr. J. 2007, 6, 43. [Google Scholar] [CrossRef] [Green Version]
  118. Busolo, M.A.; Torres-Giner, S.; Prieto, C.; Lagaron, J.M. Electrospraying assisted by pressurized gas as an innovative high-throughput process for the microencapsulation and stabilization of docosahexaenoic acid-enriched fish oil in zein prolamine. Innov. Food Sci. Emerg. Technol. 2019, 51, 12–19. [Google Scholar] [CrossRef]
  119. Böger, B.R.; Acre, L.B.; Viegas, M.C.; Kurozawa, L.E.; Benassi, M.T. Roasted coffee oil microencapsulation by spray drying and complex coacervation techniques: Characteristics of the particles and sensory effect. Innov. Food Sci. Emerg. Technol. 2021, 72, 102739. [Google Scholar] [CrossRef]
  120. Bakry, A.M.; Abbas, S.; Ali, B.; Majeed, H.; Abouelwafa, M.Y.; Mousa, A.; Liang, L. Microencapsulation of oils: A comprehensive review of benefits, techniques, and applications. Compr. Rev. Food Sci. Food Saf. 2016, 15, 143–182. [Google Scholar] [CrossRef]
  121. Avramenko, N.A.; Chang, C.; Low, N.H.; Nickerson, M.T. Encapsulation of flaxseed oil within native and modified lentil protein-based microcapsules. Food Res. Int. 2016, 81, 17–24. [Google Scholar] [CrossRef]
  122. Hermanto, R.F.; Khasanah, L.U.; Atmaka, W.; Manuhara, G.J.; Utami, R. Physical characteristics of cinnamon oil microcapsule. In IOP Conference Series: Materials Science and Engineering; IOP Publishing: Philadelphia, PA, USA, 2016; Volume 107, p. 012064. [Google Scholar]
  123. Yuan, Y.; Kong, Z.Y.; Sun, Y.E.; Zeng, Q.Z.; Yang, X.Q. Complex coacervation of soy protein with chitosan: Constructing antioxidant microcapsule for algal oil delivery. LWT 2017, 75, 171–179. [Google Scholar] [CrossRef]
  124. Huang, H.; Hao, S.; Li, L.; Yang, X.; Cen, J.; Lin, W.; Wei, Y. Influence of emulsion composition and spray-drying conditions on microencapsulation of tilapia oil. J. Food Sci. Technol. 2014, 51, 2148–2154. [Google Scholar] [CrossRef] [Green Version]
  125. Granados-Vallejo, M.; Espinosa-Andrews, H.; Guatemala-Morales, G.M.; Esquivel-Solis, H.; Arriola-Guevara, E. Oxidative stability of green coffee oil (Coffea arabica) microencapsulated by spray drying. Processes 2019, 7, 734. [Google Scholar] [CrossRef] [Green Version]
  126. Carneiro, H.C.; Tonon, R.V.; Grosso, C.R.; Hubinger, M.D. Encapsulation efficiency and oxidative stability of flaxseed oil microencapsulated by spray drying using different combinations of wall materials. J. Food Eng. 2013, 115, 443–451. [Google Scholar] [CrossRef] [Green Version]
  127. Aghbashlo, M.; Mobli, H.; Madadlou, A.; Rafiee, S. Influence of wall material and inlet drying air temperature on the microencapsulation of fish oil by spray drying. Food Bioprocess Technol. 2013, 6, 1561–1569. [Google Scholar] [CrossRef]
  128. Aghbashlo, M.; Mobli, H.; Madadlou, A.; Rafiee, S. Fish oil microencapsulation as influenced by spray dryer operational variables. Int. J. Food Sci. Technol. 2013, 48, 1707–1713. [Google Scholar] [CrossRef]
  129. Li, D.; Zhong, X.; Faiza, M.; Wang, W.; Lian, W.; Liu, N.; Wang, Y. Simultaneous preparation of edible quality medium and high purity diacylglycerol by a novel combined approach. LWT 2021, 150, 111949. [Google Scholar] [CrossRef]
  130. Diao, X.; Guan, H.; Kong, B.; Zhao, X. Preparation of diacylglycerol from lard by enzymatic glycerolysis and its compositional characteristics. Korean J. Food Sci. Anim. Resour. 2017, 37, 813. [Google Scholar]
  131. Zhou, Q.; Wei, Z. Food-grade systems for delivery of DHA and EPA: Opportunities, fabrication, characterization and future perspectives. Crit. Rev. Food Sci. Nutr. 2021. [Google Scholar] [CrossRef]
  132. Li, G.; Lee, W.J.; Tan, C.P.; Lai, O.M.; Wang, Y.; Qiu, C. Tailored rigidity of W/O Pickering emulsions using diacylglycerol-based surface-active solid lipid nanoparticles. Food Funct. 2021, 12, 11732–11746. [Google Scholar] [CrossRef]
  133. Liu, Y.; Lee, W.J.; Tan, C.P.; Lai, O.M.; Wang, Y.; Qiu, C. W/O high internal phase emulsion featuring by interfacial crystallization of diacylglycerol and different internal compositions. Food Chem. 2022, 372, 131305. [Google Scholar] [CrossRef]
  134. Guo, Y.; Xu, Y.; Zhang, T.; Wang, Y.; Liu, R.; Chang, M.; Wang, X. Medium and long-chain structured triacylglycerol enhances vitamin D bioavailability in an emulsion-based delivery system: Combination of in vitro and in vivo studies. Food Funct. 2022, 13, 1762–1773. [Google Scholar] [CrossRef]
Figure 1. The structure of five structured lipids. (R, M, L, C18C34O2, C16H32O2, and C18H32O2 represent fatty acids, medium-chain fatty acids, long-chain fatty acids, oleic acid, palmitic acid, and linoleic acid, respectively).
Figure 1. The structure of five structured lipids. (R, M, L, C18C34O2, C16H32O2, and C18H32O2 represent fatty acids, medium-chain fatty acids, long-chain fatty acids, oleic acid, palmitic acid, and linoleic acid, respectively).
Foods 11 02400 g001
Figure 2. The Schematic diagram of several common enzyme reactors.
Figure 2. The Schematic diagram of several common enzyme reactors.
Foods 11 02400 g002
Figure 4. The systhesis mechanism of esterified propoxylated glycerol (EPG).
Figure 4. The systhesis mechanism of esterified propoxylated glycerol (EPG).
Foods 11 02400 g004
Figure 5. The schematic diagram of several spray oils: (a) is traditionally used to generate spray oil by compressing oil, (b) is used to form spray oil by compressing inert gas, and (c) is used to generate spray oil by compressing propane and oil together.
Figure 5. The schematic diagram of several spray oils: (a) is traditionally used to generate spray oil by compressing oil, (b) is used to form spray oil by compressing inert gas, and (c) is used to generate spray oil by compressing propane and oil together.
Foods 11 02400 g005
Figure 6. Schematic diagram SLs-based emulsions. (a) Oil-in-water emulsion; (b) water-in-oil emulsion; (c) Pickering emulsion formed from DAG-based nanoparticles as surfactants; (d) oil-in-water emulsion loaded with oil-soluble substances.
Figure 6. Schematic diagram SLs-based emulsions. (a) Oil-in-water emulsion; (b) water-in-oil emulsion; (c) Pickering emulsion formed from DAG-based nanoparticles as surfactants; (d) oil-in-water emulsion loaded with oil-soluble substances.
Foods 11 02400 g006
Table 1. The advantages and disadvantages of the use of different methods in SLs production.
Table 1. The advantages and disadvantages of the use of different methods in SLs production.
Enzymatic MethodChemical MethodGenetic Method
Reaction conditionMild; Efficiency; GreenStrict to catalyst and substrates; Flexible; Stability; ChemicalsStrong pertinence; Complex process; Long cycle
Production cost (Energy consumption, catalyst reuse ability)High lipase cost; Continuous; Low energy consumptionLow catalyst cost; Batch; High energy requestHigh breeding cost
Production efficiencyHigh yield; SpecificityEasy to degrade; RandomHigh yield; Specificity
Production securityHigh security in solvent-free systemChemical and metal issuesPotential gene security issues
References [4,8,9,10,11,12] [4,13,14] [4,15,16]
Table 2. Current enzyme reactors for SLs production.
Table 2. Current enzyme reactors for SLs production.
EquipmentSubstratesTarget ProductsConditionsAdvantagesReferences
Orbital shakerFully hydrogenated soybean oil, caprylic acidMLCTPyLip enzyme acidolysis at 60 °C for 1 h, hexaneSubstrates mixed evenly; Controllable reaction conditions [28]
Rotary shakerFlax oil, single cell oils, caprylic acidMLCTImmobilized TLL lipase acidolysis at 50 °C for 24 h, solvent-free systemControllable reaction conditions; Low energy consumption [29]
StirrerExtra virgin olive oil, soybean oil, fully hydrogenatedcrambe oilBehenic acid enriched SLLipozyme TL IM interesterification
at 60 °C for 4 h, solvent-free system
Controllable reaction conditions; Low energy consumption [30]
ShakerMutton tallow, hemp oilDAGImmobead 150 lipase interesterification at 60 °C for 6 hSubstrate mixed evenly, Controllable reaction conditions [31]
Batch reactorGrapeseed oil, capric acid, caprylic acidMLCTLipozyme RM IM® lipase acidolysis at 45 °C for 24 h, solvent-free systemControllable reaction conditions [32]
Magnetic stirrerHigh oleic sun-flower oil, fully hydrogenated Crambe abyssinica oilBehenic acid enriched SL Lipozyme TL IM interesterification
at 70 °C for 3 h, solvent-free system
Low cost; Controllable reaction conditions [33]
Bubble column reactorCaprylic acid, capric acidHigh Purity Medium Chain DAGNovozym 435 esterification at 60 °C for 30 min, solvent-free systemMild; Efficiency; Durable catalyst [27]
Bubble column reactorPalm oil deodorizer distillate, oleic acid, glycerolDAGLipozyme 435 esterification
at 60 °C for 30 min, solvent-free system
Mild; Efficiency; Durable catalyst [34]
Packed bed reactorPalm olein, fully hydrogenated palm oil, palm kernel oil Cocoa butter substitutesLipozyme TL IM interesterification at 65 °C with feed flow rate of 70 mL/min, solvent-free systemMild; Efficiency; Durable catalyst [35]
Magnetic stirrerMAG, caprylic acidDAGNovozyme 435 esterification at 65 °C for 30 min, solvent-free systemLow cost; Controllable reaction conditions [36]
Table 3. Current enzymatic synthesis methods to produce MLCT.
Table 3. Current enzymatic synthesis methods to produce MLCT.
SubstratesType and System of ReactionOptimal ConditionsCharacteristicsReferences
Caprylic acid, capric acid, oleic acid, glycerolVacuum and solvent-free system, esterificationNovozym 435 esterification at 90 °C for 12.37 hMLCT content, Enzyme activity [6]
Arachidonic acid single cell oils, MCTsSolvent-free system, transesterificationLipozyme 435 transesterification at 90 °C for 3 h.Fatty acid and TAG composition, Melting and crystallization behavior [37]
Glycerol, capric acid, oleic acidVacuum and solvent-free system, esterificationLipozyme RM IM esterification at 70 °C for 14 hMLCT content [40]
Canola oil, caprylic acidSolvent-free system, acidolysisLipozyme RM IM acidolysis at 50–60 °C for 15 hFatty acid and TAG composition, Refractive index, Melting profile [41]
Cinnamomum camphora seed oil, camellia oilSolvent-free system, transesterificationLipozyme RM IM transesterification at 60 °C for 3 hFatty acid and TAG composition, Tocopherol analysis [42]
Flaxseed oil, tricaprylinOrganic solvent system, transesterificationLipozyme TL IM transesterification at 41.49–50.00 °C for 4.00–4.01 hBioconversion yield [43]
Hydrogenated soybean oil, rice bran oil, coconut oilSolvent-free system, transesterificationLipozyme TL IM transesterification at 65 °C for 24 hFatty acid and TAG composition, Polymorphism, Crystal Morphology, Analysis of Tocopherols and Phytosterols [44]
Soybean oil, MCTSolvent-free system, transesterificationLipozyme TL IM transesterification at 55 °C for 30–40 minFatty acid and TAG composition, DAG content, Acyl migration [45]
Catfish oil, basa catfish oilSolvent-free system, transesterificationNS 40086 lipase transesterification at 60 °C for 3 hFatty acid and TAG composition, [46]
Microbial oil, MCT containing 99% of caprylic acidSolvent-free system, transesterificationNS 40086 lipase transesterification at 60 °C for 8 hFatty acid and TAG composition [47]
Table 4. Current enzymatic synthesis methods to produce DAG.
Table 4. Current enzymatic synthesis methods to produce DAG.
SubstratesType and System of ReactionConditionsCharacteristicsReferences
Menhaden oilSolvent-free system, glycerolysisNovozym 435 glycerolysis at 70 °C for 24 hDAG content, Enzyme activity, Positional Analysis of TAG, Fatty acid composition [56]
Rice bran oilSolvent-free system, glycerolysisC.antarctica lipase glycerolysis at 70 °C for 24 hMAG and DAG content, Particle size [63]
Olive oilSolvent-free system, glycerolysisNovozym 435 glycerolysis at 70 °C for 4 hFatty acid composition, Enzyme activity [64]
Oleic acidSolvent-free system, esterificationLecitase® Ultra esterification at 40 °C for 1.5  hAcylglycerols compositions, Enzyme activity [65]
Lauric acidSolvent-free system, esterificationLipozyme RM IM esterification at 50 °C for 3 hDAG content, Reusability of lipase [66]
Rapeseed oil, MAG, oleic acidSolvent-free system, esterificationImmobilized lipase EC3.1.1.3 at 60 °C for 6 hAcid value, MAG and DAG content [67]
Soybean oilSolvent-free system, glycerolysisImmobilized RML glycerolysis at 60 °C for 24 hDAG content, Enzyme activity, TAG conversion [68]
Soybean oilSolvent-free system, glycerolysisImmobilized TLL glycerolysis at 60  °C for 12  h, MAG, DAG, and TAG content, Enzyme activity [69]
Soybean oilSolvent-free system, glycerolysisImmobilisation lecitase® ultra glycerolysis at 45 °C for 12 hTAG composition, DAG content [70]
Short- and medium-chain fatty acid ethyl estersSolvent-free system, transesterificationNovozym 435 transesterification at 65 °C for 24 h, Lipozyme RM IM transesterification at 65 °C for 32 h.MAG, DAG, and TAG content [71]
Table 5. Current enzymatic synthesis methods to produce EPA- and DHA-enriched products.
Table 5. Current enzymatic synthesis methods to produce EPA- and DHA-enriched products.
SubstratesType of ReactionConditionsReferences
Fish oilTwo-stage enzymatic refining process: hydrolysis, transesterification400SD hydrolysis at 35 °C for 10 h, sodium phosphate buffer; Novozym 435 transesterification at 60 °C for 6 h, solvent-free system [88]
Codfish oilHydrolysisOUC-Lipase 6 hydrolysis at 40 °C, for 36 h, Tris-HCl buffer [89]
Camelina sativa oil, omega-3 fatty acid ethyl estersTwo-step reaction: ethanolysis, transesterificationLipozyme TL IM ethanolysis for 1 h, ethanol; Novozym 435 transesterification at 35 °C for 4 h, solvent-free system [90]
Glycerol, DHA/EPA-rich ethyl esters Two-step enzymatic reaction: transesterification, ethanolysisNovozym 435 transesterification at 60 °C for 24 h, solvent-free system;
Immobilized SMG1-F278N ethanolysis at 30 °C for 96 h, n-hexane, ethanol
[91]
Glycerol, n-3 PUFAEsterificationNovozym 435 at 50 °C for 50 h, deep eutectic solvents [92]
Arctic cod liver oilAlternate winterization and enzymatic interesterificationAlternate winterization at −80 °C for 24 h, acetone; Lipozyme TL IM interesterification at 40 °C for 2.5 h, solvent-free system [93]
Microalgae oil and oleic acidAcidolysisLipase RM IM acidolysis at 65 °C for 6 h, solvent-free system [94]
Schizochytrium sp. biomassEthanolysisFour liquid formulated enzymes CALA, PLA, RM and TL ethanolysis at 35 °C for 96 h, ethanol [95]
Refined sardine oil, glycerol and tertpentanolGlycerolysisLipozyme 435 glycerolysis at 50 °C for 2 h, tert-Pentanol [96]
Salmon frame bone oilAlcoholysis, esterificationNovozym 435 alcoholysis at 37 °C for 3 h, ethanol; Lipozyme RM IM esterification at 55 °C for 48 h, solvent-free system [97]
Caprylic acid, A. limacinum SR21 oilAcidolysisLipozyme TL IM acidolysis at 37 °C for 30.4 h, hexane [98]
Microbial oil, medium-chain fatty acidsAcidolysisLipozyme RM IM acidolysis at 55 °C for 6 h, solvent-free system [99]
Table 6. Current enzymatic synthesis methods to produce human milk fat substitute SLs.
Table 6. Current enzymatic synthesis methods to produce human milk fat substitute SLs.
EquipmentSubstratesConditionsType of ReactionReferences
StirrerLeaf lard, camellia oil fatty acidsFractionation at 60 °C for 20 min followed by 34 °C for 10 h; Lipozyme RM IM acidolysis at 45 °C for 6 h, solvent-free systemFractionation and acidolysis [103]
Batch reactorPalm stearin, oleic acidNS40086 lipase acidolysis at 60 °C for 4 h, hexane, solvent-free systemAcidolysis [105]
Magnetic stirrerPalm stearin, oleic acid, linoleic acidNS40086 lipase acidolysis at 60 °C for 4 h, solvent-free systemAcidolysis [106]
StirrerRapeseed oil, tripalmitinCandida cylindracea lipase hydrolysis at 40 °C for 2 h, Tris-HCl buffer; Novozym 40086 acidolysis at 40 °C for 2 h, n-hexaneHydrolysis and acidolysis [111]
Magnetic stirrer Soy oil, palm kernel stearin, palm stearin, oleic acid, linoleic acidLipozyme TL IM interesterification at 60 °C for 5 h, solvent free system;
Lipozyme RM IM interesterification at 56 °C for 7 h, solvent free system
Interesterification [112]
Stirred tank reactor and continuous PBRPalm stearin, oleic acidStirred tank reactor: Immobilized AOL lipase at 65 °C for 1.5 h, solvent-free system
PBR: Immobilized AOL lipase at 62.09 °C for 3 h, solvent-free system
Acidolysis [113]
Table 7. Approaches for preparation of powdered oil products.
Table 7. Approaches for preparation of powdered oil products.
Core MaterialsWall MaterialsPreparation MethodsReferences
DHA-enriched fish oilZeinElectrospraying assisted by pressurized gas [118]
Roasted coffee oilStarch, gelatin, gum arabicSpray drying and complex coacervation [119]
Flaxseed oilLentil protein, maltodextrinFreeze-drying [121]
Cinnamon oilMaltodextrin, gum arabicSpray drying [122]
Algal oilSoy protein isolate, chitosanComplex coacervation [123]
Tilapia oilTrehalose, gelatin, sucrose, xanthanSpray drying [124]
Coffee oilMesquite gum, octenyl succinic anhydride modified starchesSpray drying [125]
Flaxseed oilMaltodextrin, gum Arabic, whey protein concentrate, modified starch Spray drying [126]
Fish oilSkim milk powder, whey protein concentrate, whey protein isolate, milk protein concentrate, sodium caseinateSpray drying [127]
Fish oilSkim milk powderSpray drying [128]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhou, J.; Lee, Y.-Y.; Mao, Y.; Wang, Y.; Zhang, Z. Future of Structured Lipids: Enzymatic Synthesis and Their New Applications in Food Systems. Foods 2022, 11, 2400. https://0-doi-org.brum.beds.ac.uk/10.3390/foods11162400

AMA Style

Zhou J, Lee Y-Y, Mao Y, Wang Y, Zhang Z. Future of Structured Lipids: Enzymatic Synthesis and Their New Applications in Food Systems. Foods. 2022; 11(16):2400. https://0-doi-org.brum.beds.ac.uk/10.3390/foods11162400

Chicago/Turabian Style

Zhou, Jun, Yee-Ying Lee, Yilin Mao, Yong Wang, and Zhen Zhang. 2022. "Future of Structured Lipids: Enzymatic Synthesis and Their New Applications in Food Systems" Foods 11, no. 16: 2400. https://0-doi-org.brum.beds.ac.uk/10.3390/foods11162400

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

Article Metrics

Back to TopTop