Next Article in Journal
A Multicriteria Motion Planning Approach for Combining Smoothness and Speed in Collaborative Assembly Systems
Next Article in Special Issue
A Maturity Estimation of Bell Pepper (Capsicum annuum L.) by Artificial Vision System for Quality Control
Previous Article in Journal
Mineral Carbonation of CO2 in Mafic Plutonic Rocks, II—Laboratory Experiments on Early-Phase Supercritical CO2‒Brine‒Rock Interactions
Previous Article in Special Issue
Application of Sterilization Process for Inactivation of Bacillus Stearothermophilus in Biomedical Waste and Associated Greenhouse Gas Emissions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Developments in the Use of Lipase Transesterification for Biodiesel Production from Animal Fat Waste

1
Instituto de Tecnología Química (CSIC-UPV), Universitat Politècnica de València, Camino de Vera s/n, 46022 Valencia, Spain
2
Instituto de Agroquímica y Tecnología de Alimentos (CSIC), Avenue Agustín Escardino 7, Paterna, 46980 Valencia, Spain
*
Author to whom correspondence should be addressed.
Current address: CNRS, 25 rue des Martyrs, B.P. 166, 38042 Grenoble CEDEX 9, France.
Appl. Sci. 2020, 10(15), 5085; https://doi.org/10.3390/app10155085
Submission received: 30 June 2020 / Revised: 20 July 2020 / Accepted: 21 July 2020 / Published: 23 July 2020

Abstract

:
Biodiesel constitutes an attractive source of energy because it is renewable, biodegradable, and non-polluting. Up to 20% biodiesel can be blended with fossil diesel and is being produced and used in many countries. Animal fat waste represents nearly 6% of total feedstock used to produce biodiesel through alkaline catalysis transesterification after its pretreatment. Lipase transesterification has some advantages such as the need of mild conditions, absence of pretreatment, no soap formation, simple downstream purification process and generation of high quality biodiesel. A few companies are using liquid lipase formulations and, in some cases, immobilized lipases for industrial biodiesel production, but the efficiency of the process can be further improved. Recent developments on immobilization support materials such as nanoparticles and magnetic nanomaterials have demonstrated high efficiency and potential for industrial applications. This manuscript reviews the latest advances on lipase transesterification and key operational variables for an efficient biodiesel production from animal fat waste.

Graphical Abstract

1. Introduction

Animal byproducts generated in the European Union slaughterhouses represent nearly 17 million tons per year and, from them, 5 million tons inedible byproducts result from rendering and are mostly used for energy generation like biofuels and biodiesel [1,2,3]. After rendering byproducts, fat is obtained from beef tallow, mutton tallow, pork lard and chicken fat [4,5]. Such fat is majorly composed of triacylglycerols with fatty acids of 16 to 18 carbons. The most abundant saturated fatty acids are palmitic (16:0) and stearic (18:0) acids; the major monounsaturated fatty acid is oleic acid (18:1) and the most abundant polyunsaturated fatty acids are linoleic (18:2) and arachidonic (20:4) acids [6,7]. Animal fat waste is also obtained from the meat processing industry and from recycled waste from the cooking business [8,9] that are classified as yellow grease if the content of free fatty acids is lower than 15% by weight and brown grease when it is higher than 15% [10]. In 2019, more than 800 thousand tons of animal fats, equivalent to 6% of total feedstock, were used to produce biodiesel in the European Union [11,12], while 8.4% of total feedstock was used in the case of the US, consisting of mainly 74 tons of poultry fat, 132 tons of tallow and 243 thousand tons of white grease [13].
Biodiesel produced from animal fats is cheaper than when made from vegetable oils. An additional advantage is that fossil CO2 reduction is higher when using animal fat for biodiesel generation; nearly 80% CO2 reduction may be reached for animal fat in comparison to 30% reduction when using vegetable oil [14,15]. The bioenergy demand is continuously increasing and in 2050 it is expected to reach 30% of the fuel consumed in the world for road transport [15,16]. Research on biodiesel production is trying to maximize the yield and minimize the costs by using better catalysts that can be reused and improve the transesterification efficiency [17,18]. Furthermore, the feedstock used as raw material for biodiesel production represents up to 80% of the total cost [19] and it explains its variability in different geographic areas depending on the climate and agriculture [20].
Total biodiesel world production has been increasing progressively year by year, reaching nearly 45 million tons in 2019 [12]. The European Union has the largest biodiesel production through its 202 plants producing more than 14 million tons of biodiesel in 2019 [11,21]. More than 5.6 million tons of biodiesel were produced in the US in 2019 through its 91 plants [13,22]. Nearly 80% of new diesel vehicles are prepared for B20 use that consists of fossil diesel blended with 20% biodiesel [13].
Transesterification through alkaline catalysis is the preferred process at industrial biodiesel production plants [23]. However, raw materials like animal fat that contain moisture and free fatty acids are troublesome for alkaline transesterification due to soap formation. Acid catalysis does not have such troubles, but the reaction is much slower than alkaline catalysis, needs a larger size reactor and requires a higher alcohol to fat molar ratio [24]. Heterogeneous catalysts are not sensitive to the presence of free fatty acids and moisture, can catalyze esterification and transesterification simultaneously, and can be separated from the reaction media. However, such solid catalysts tend to form three phases resulting in a reduced reaction rate and high energy consumption [25]. The simultaneous esterification and transesterification also occur with supercritical technology where high temperature and pressure conditions (i.e., >250 °C and 10 MPa) increase the solubility and reduce the mass transfer limitation resulting in good efficiency but with high energy consumption [26,27,28]. Pseudo catalytic transesterification using biochar as the porous material for the pseudo-catalytic reaction at more than 300 °C has the same advantages as supercritical transesterification, but also has high energy consumption [29,30]. Therefore, animal fats may be processed for biodiesel production through enzymatically catalyzed transesterification even though some issues, like the cost of lipase and its poor stability, can be improved through immobilization [31]. Lipases have the advantage to generate biodiesel under mild reaction conditions through the conversion of free fatty acids and triacylglycerols in the presence of an acyl acceptor [32]. This manuscript reviews and discusses the latest advances in the use of free and immobilized lipases for an efficient transesterification of animal fat waste.

2. Mechanisms of Action of Lipases

Lipases, triacylglycerol ester hydrolases (EC 3.1.1.3), are serine hydrolases with an active site containing an amino-acid triad of serine, histidine and aspartate [32]. Lipases are obtained from a variety of sources such as animal and plant tissues and microorganisms. Lipases show a wide range of pH and temperature for activity and vary from strain to strain regarding specificity and hydrolysis rate [33] Lipases exhibit good stability in non-aqueous mediums and exhibit neutral pH range; such stability is increased when the enzyme is immobilized.
Lipases can catalyze esterification, inter-esterification, and trans-esterification reactions in non-aqueous environments. Lipases catalyze the hydrolysis of triacylglycerols at the aqueous-non aqueous interface but these enzymes can also catalyze the synthesis of esters from alcohols and long chain fatty acids in low moisture environment [33]. Lipases follow a two-step mechanism for the generation of fatty acid methyl esters in transesterification reactions, usually through the Ping-Pong Bi Bi mechanism [34].
Most triacylglycerol lipases are regiospecific because they can only hydrolyze primary ester bonds at the sn-1 and sn-3 positions, external positions within the triacylglycerol, and can generate either one free fatty acid and diacylglycerol, or two free fatty acids and 2-monoacylglycerol that remain unhydrolyzed. The full process from triacylglycerols into biodiesel and glycerol as end products is shown in Figure 1. Regiospecificity is characteristic of extracellular bacterial lipases from Bacillus sp. [35,36].
Monoacylglycerol lipases (EC 3.1.1.23) catalyze the hydrolysis at the specific sn-2 position of 2-monoacylglycerol into free fatty acid and glycerol. Such lipases may be present in the enzyme extract and masked when measuring activity with standard activity methods like those based on triolein hydrolysis measurement. Monoacylglycerol lipases have been the object of few studies [37], although they might be present in some microbial enzyme preparations [38]. Other lipases are nonspecific and can act on any of the ester bonds of the triacylglycerol and therefore break down the triacylglycerol to release free fatty acids and glycerol as the final products. This is the case of lipases from Staphylococcus aureus and hyicus [39], Geotrichum candidum, Corynebacterium acnes, Penicillium cyclopium [24] and Chromobacterium viscosu [40]. Another alternative for the hydrolysis of monoacylglycerols is the acyl migration in the glycerol backbone from the sn-2 position to sn-1 or sn-3 positions [41].
The specificity of lipases depends on the length of fatty acids, presence of double bonds, branched groups and, consequently, reaction rates may have important variations depending on the composition of triacylglycerols present in the fat waste. Lipases are especially active against medium to long chain fatty acids, which are those more usual in animal fat waste [17].

3. Sources of Lipases

Most lipases originated from microorganisms and are produced in fermenters under controlled conditions (see Table 1). Lipases are produced by a variety of gram-positive and gram-negative bacterial strains, especially from the genera of Pseudomonas [42,43], also by filamentous fungus that are commercially important such as those belonging to the genera of Rhizopus sp. [44], Aspergillus sp. [45], Penicillium sp. [46], Geotrichum sp. [33], Mucor sp. [47] and Thermomyces sp. [48]. Lipases produced from yeasts are also relevant such as those from Candida sp. [49,50].
Extracellular lipases are secreted into the production medium and recovered from the microorganism broth. Then, lipases are further separated and purified but downstream processing is costly. Intracellular lipases imply the use of whole cell microorganisms and this fact reduces the costs of enzyme extraction and purification but the efficiency and biodiesel yield is low when catalyzing an oily substrate due to mass transfer limitations for substrate penetration and product release [28,51]. Some whole cell biocatalysts used to produce biodiesel are filamentous fungi like Aspergillus and Rhizopus [49].

4. Free Lipase

Lipases constitute an attractive catalyst for transesterification in those wastes containing large amounts of moisture and free fatty acids, which is the case of animal fat and is what makes it troublesome for alkaline transesterification. Table 2 shows some examples of the use of free lipases for biodiesel production from animal fat waste. The use of lipases has relevant advantages over conventional alkali catalysts. The most relevant are the absence of soap formation in the reactor, insensitivity to water content and acidity value, moderate reaction conditions, broad substrate range, good purity of biodiesel after transesterification and absence of pollutants, especially when treating cooking oil waste containing large amounts of free fatty acids [75]. On the contrary, there are also important disadvantages such as the high enzyme costs, poor enzyme stability, and the enzyme deactivation by alcohol [76] and partly by the generated glycerol [66].
As said, lipases are sensitive to the alcohol, in most cases methanol, used for biodiesel production and this fact increases the operational costs. There are some alternatives to avoid enzyme damage by methanol: stepwise addition of methanol to reaction mixtures avoiding a high concentration [65], the use of co-solvents like hydrophilic tert-butanol that dissolve glycerol and methanol and therefore allow high transesterification yields and rates [77], also the addition of longer-chain alcohols [78], or methyl or ethyl acetate as acyl acceptors [79,80]. Another solution is the use of novel lipases that can support one-step addition of high methanol concentration [81]. In this sense, another solution is the use of tools like protein engineering, recombinant methods and metabolic engineering that are used to improve thermostability as well as stability in organic solvents [80]. Glycerol may be extracted with organic solvents although the enzyme activity may be affected [81].
Lipase transesterification requires an extended time of reaction, and has a slow conversion rate as shown in Table 2. The recovery of the enzyme is rather difficult and the enzyme stability is poorer at high temperature and pH [81,82,83,84]. All these troubles have hindered its adoption at an industrial scale and therefore, transesterification with alkaline catalysis is still preferred at biodiesel-producing industrial plants [23]. However, such troubles experienced with enzymes can be partly overcome through its immobilization on a solid material that acts as enzyme carrier and increases its stability and efficiency [81,85,86]. Immobilization also allows an easy downstream separation from the product and decreases cost [82]. However, recyclability is an issue because lipases tend to lose activity after continuous operation [86]. In any case, it was reported that the use of soluble lipases might be more competitive if the commercial enzymes would have a price 50 times lower than the immobilized lipase [25,87]

5. Immobilized Lipase

5.1. Types of Supports and Immobilization Procedures

The immobilization of lipases consists of the retention of the enzyme at the surface of the support material. In this way, immobilized lipases show an improved efficiency and reduced costs, with longer enzyme stability and better resistance to denaturation by alcohol. There are many available supports of organic, synthetic and inorganic nature for lipase immobilization. Such materials may vary in characteristics such as particle dimensions, shape, pore volume, hydrophobicity, and density and must be stable to physical, chemical, and microbial degradation [93]. Porous supports with controlled pore distribution are very interesting for lipase immobilization because they offer an extensive surface area and therefore, higher enzyme loading. However, caution must be observed if pores are too small because they could get blocked by the enzyme, reducing its efficiency.
There is a large variety of immobilization procedures (see Figure 2) such as adsorption, covalent binding, cross-linking, entrapment, or encapsulation that have been developed to enhance the catalytic activity, and its stability, and make possible the reutilization of the enzyme in relation to the soluble lipase [94]. Some methods like cross-linking enzyme aggregates are not considered in this review because even though they are inexpensive, highly efficient, and do not need support for immobilization, they have rather poor mechanical stability [95].
Immobilization by adsorption on materials such as water-absorbing polymer, hydrophobic macroporous polypropylene particles or silica gel is simple, but can result in an undesirable leakage making it necessary to assure the retention of the enzyme by additional ionic or covalent bonds. This can be an ion exchange resin or cross-linking with glutaraldehyde [68]; performance was improved by crosslinking with glutaraldehyde. The stability in acid pH was improved as well as thermostability at 45–50 °C. It retained 80% of relative biodiesel production after 5 consecutive batches [96]. Adsorption of lipase from Burkholderia cepacia was compared to covalent immobilization on epoxy acrylic resin. The adsorbed enzyme gave a higher conversion than the covalent one after a three-step addition of ethanol, 68% vs 47% [97]. The covalently immobilized enzyme showed lower affinity towards diglycerides and monoglycerides; this was attributed to blockage of the active groups by the covalent bonds to the support material, which resulted in enzyme rigidity [98]. Mesoporous materials are attractive because they have high surface area, larger pore volume, absence of toxicity, and good stability [99]. Examples of reported immobilized lipases used in recent studies are shown in Table 3.
There is a good affinity for immobilization of lipases on hydrophobic supports [104], giving a fast and good attachment by hydrophobic adsorption [105]. Transesterification of waste lard was tested with immobilized lipase B from Candida antarctica with the assistance of ultrasound for improving the dispersion and collision of the reagent molecules. Ultrasonic wave amplitude, ultrasonic cycles, and reaction parameters were optimized and a kinetic model was developed [69]. Pulsed ultrasound irradiation increased by about 3 times the synthesis rate of fatty acid ethyl ester by lipases immobilized on hydrophobic carriers like octadecyl-sepabeads [106].
Porous silica nanoflowers have center-radial pore structure that allows the load of lipase inside the structure to have good mass transfer for substrates and products [107]. Dichlorodimethylsilane was used to modify the silica nanoflowers for the adsorption of Candida antarctica lipase B and the biocatalytic pickering emulsion was constructed [94]. Pickering emulsion stabilized by hybrid nanoparticles [108], solid particles [109], or carbon nanotubes crosslinked with lipase [110], have been constructed and successfully used [111,112]. In this way, this emulsion facilitates biphasic reactions and simplifies the recovery of lipase that remains in its microenvironment [113].
Metal-organic frameworks provide advantages for immobilization: they can be easily separated, they offer an extended surface area that can be tuned, they have adequate pore size, have structural and functional diversity and good stability. Immobilization may be through adsorption, encapsulation, and coprecipitation [114]. Lipase is strongly adsorbed by metal-chelating affinity immobilization that is a simple technique with the advantage that support may be reused [54]. Desorption agents may cause the desorption of the enzyme that can be also achieved by changing the pH value [115]. The compound n-hexane could reduce the deactivation of AGMNP-CO2+-PFL from methanol. It was reported that biodiesel production from oil transesterification was higher with n-hexane than using tert-butanol [70,94].
Encapsulation immobilization entraps lipase by a co-precipitation method and crosslinking agents like glutaraldehyde are used to interconnect the enzymes. However, there is a high mass transfer resistance that reduces its efficiency.
The entrapment of cells having lipase activity appears to be a simple and attractive technique. Lipase immobilized in silicon granules or calcium alginate beads, with glucose supplementation for cells maintenance, achieved an increased number of cycles, 28 instead of 23, while keeping 90% activity [33]. Whole cell, recombinant methods and protein and metabolic engineering are promising options to increase lipase applications [82].

5.2. Magnetic Nanocarriers

Materials like magnetite (Fe3O4) are used as support for immobilization because they allow a rapid separation from the reaction medium when an external magnetic field is applied [116]. The development of magnetic nanoparticles (MNPs) as a support for enzymes immobilization has been recently reviewed [117]. Typical magnetic nanomaterials include iron oxide (Fe3O4 and γ-Fe2O3), alloy-based (CoPt3 and FePt), pure metal (Fe and Co), and spinel-type ferromagnet (MgFe2O4, MnFe2O4, and CoFe2O4) [117]. Examples of lipase immobilized in various types of magnetic nanoparticles are shown in Table 4.
Magnetic nanoparticles (MNPs) have good biocompatibility and non-toxicity but tend to aggregate and oxidize, so they need to be functionalized on the surface and use a cross-linking agent to bind the enzyme. One way is by using silica coating where a silica shell is formed on the surface by using amino-functional reagents like 3-aminopropyl triethoxysilane (APTES). Fe3O4 particles were encapsulated with mesoporous silicon and modified with APTES or 3-mercaptopropyl trimethoxysilane (MPTMS) followed by binding of the lipase with glutaraldehyde. APTES-modified Fe3O4 particles were reported to give better yield of biodiesel (90%) than MPTMS particles [75]. Burkholderria sp. lipase on Fe3O4 MNPs also achieved 90% conversion [123].
Another way to protect MNPs is with organic polymers, including synthetic polymers and biopolymers. The polymer can be either incorporated into the precursor solution to form Fe3O4 MNPs or externally to create the core shell [117]. The polymer surface provides numerous functional groups that facilitate the enzyme binding. Magnetic chitosan binds the lipase by covalent attachment [124].
Separation of nanobiocatalysts is difficult in an oily system [119] but the magnetic properties of MNPs can facilitate the separation of enzyme from reaction media. In this way, the reaction may be immediately finished as well as using the enzyme for further uses [117].

5.3. Coimmobilization

Some works have proposed to use coimmobilization of enzymes. The advantage is that the first enzyme releases the product that is transferred to the next coimmobilized enzyme with a short diffusional distance. This is especially relevant for lipases due to their specificity for triacylglycerols ester bonds. The mixture of 1,3-specific lipase and a non-specific lipase enhances the global activity because it removes the limiting acyl-migration step. Several coimmobilized systems have been studied for biodiesel production like Candida rugosa lipase and Rhyzopus oryzae lipase simultaneously on silica gel [74], Candida antarctica lipase B and Thermomyces lanuginosus lipase, on the surface of the Phichia pastoris cell [57], and Rhizomucor miehei and Candida antarctica lipases on epoxy-functionalized silica [67,100].
Lipases are coimmobilized on the same material surface in order to get better global activity and improved enzyme specificity and selectivity for hydrolysis of triacylglycerols as well as those generated diacylglycerols and monoacylglycerols that must be further hydrolyzed [96]. However, the most active enzyme may get some loss of activity through this procedure [125]. A different coimmobilization strategy was proposed by immobilizing several lipases layer-by-layer using abcoating with polyethylenimine [126]. Other authors have used coating with PEI/glutaraldehyde to form 5 enzyme layers of lipases A and B from Candida antarctica, lipases from Rhizomocur miehei, and Themomyces lanuginosus and phospholipase Lecitase Ultra [125]. Although it gives an innovative way for fats hydrolysis some problems might arise either from inhibition by coating agents used, high costs of different lipases used or steric hindrance for accessibility of triacylglycerols to the active site of lipases immobilized in the inner layers.

6. Industrial Applications of Lipase-Catalyzed Biodiesel

Even though transesterification through alkaline catalysis is the preferred process in the majority of industrial biodiesel production plants [23], a few lipase-based processes have already been implemented to plant-scale operation. The collaboration of Novozymes (Bagsvaerd, Denmark) with Piedmont Biofuels (Pittsboro, NC, USA) resulted in a patent application to produce fatty acid alkyl esters, by a lipolytic enzyme in a solution containing triacylglycerol, alcohol, water, and glycerol [93,127]. Viesel Fuel (Terrac Stuart, FL, USA) upgraded in 2013 its facility through an enzymatic process developed by Novozymes (Denmark) to use brown grease and waste cooking oil to produce up to 11 million gallons biodiesel per year using Eversa Transform® lipase from Novozymes, a soluble lipase produced by a genetically modified strain of Aspergillus oryzae [128], and an ion exchange resin system for removal of remaining free fatty acids during crude biodiesel refining [129,130]. Viesel Fuel, Novozymes and Tactical Fabrication also collaborated with Buster Biofuels to upgrade its facility in San Diego (CA, USA) to produce up to 5 million gallons per year [131]. Lvming and Environmental Protection Technology Co. Ltd. (Shanghai, China) used lipase of Candida sp. to produce 10,000 tons per year from waste frying oil [95]. A plant in Sumaré (Sao Paulo, Brazil) produces biodiesel from mixed beef tallow and soybean oil using Callera® Trans L lipase in a batch reactor [132]. These companies are using liquid lipase formulations but the efficiency of the process can be improved further by using recent developments in immobilized lipases. So, Hunan Rivers Bioengineering Co. Ltd. (Hunan, China) was reported to use Novozym 435® lipase in a stirred tank reactor to produce 20,000 tons of biodiesel per year. The enzyme is a lipase B from Candida antarctica immobilized on a resin consisting of macroporous support formed by poly(methyl methacrylate) crosslinked with divinylbenzene [133]. New technology protected with patents [134] has been provided by EnzymoCore, a leading global producer company founded in 2007 in Israel and with several active biodiesel plants around the world. This company has developed modified-immobilized enzymes, supported on solid organic resins, with high resistance to methanol and able to produce biodiesel from any type of oil or fat, even those cheap and with very large content of free fatty acids and polar lipids [135].

7. Conclusions

Animal fat waste, usually resulting from slaughterhouses, the meat processing industry, and cooking facilities, is being increasingly used for biodiesel production. Transesterification through alkaline catalysis is the preferred process at industrial biodiesel production plants. Transesterification with lipases has not been attractive for industry yet because of the higher operative costs in comparison to alkaline catalysis; transesterification with lipases has problems including poor enzyme stability, difficulties in reusability, and denaturation by alcohol although they are not affected by water and free fatty acids typically found in animal fats. However, a few companies could solve such troubles since they are running liquid lipase formulations for producing biodiesel from cooking oil waste at industrial scale. However, the efficiency of the process can be further improved. Recent developments in immobilized lipases and availability of different types of supports such as mesoporus materials, silica nanoflowers, pickering emulsion, and metal-organic frameworks demonstrate improved efficiency and reduced costs. Immobilization of the enzyme in such materials increases its stability and makes it more resistant to denaturation by alcohol. Magnetic nanomaterials constitute an even better support for enzyme immobilization because they can be recovered when an external magnetic field is applied. These nanoparticles are functionalized on the surface by coating with silica or organic polymers that enhance the efficiency of the process. The entrapment of whole cells with lipase activity, appears to be simple and efficient although more research is needed. Coimmobilization of lipases is an innovative process, but not so attractive for industrial application. It needs further research because of the need for different lipases that increases the costs and the efficiency affected by steric difficulties for enzymes to hydrolyze triacylglycerols.

Author Contributions

Conceptualization, F.T.; resources, F.T.-R. and L.M.; writing—original draft preparation, F.T.-R.; writing—review and editing, F.T.-R., L.M. and F.T.; supervision, F.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by European Marie Curie project, grant number 614281 (HIGHVALFOOD) and European Regional Development Fund.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; they had no role in the collection, analyses, or interpretation of data; and they had no role in the writing of the manuscript, or in the decision to publish the results.

References

  1. Rosson, E.; Sgarbossa, P.; Pedrielli, F.; Mozzon, M.; Bertani, R. Bioliquids from raw waste animal fats: An alternative renewable energy source. Biomass Convers. Biorefin. 2020, 1–16. [Google Scholar] [CrossRef]
  2. EFPRA. Rendering in Numbers. 2016. Available online: https://efpra.eu/wp-content/uploads/2016/11/Rendering-in-numbers-Infographic.pdf (accessed on 20 March 2020).
  3. Toldrá, F.; Mora, L.; Reig, M.; Mora, L. New insights into meat by-product utilization. Meat Sci. 2016, 120, 54–59. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Akhil, U.S.; Alagumalai, A. A Short Review on Valorization of Slaughterhouse Wastes for Biodiesel Production. ChemistrySelect 2019, 4, 13356–13362. [Google Scholar] [CrossRef]
  5. Barik, D.; Vijayaraghavan, R. Effects of waste chicken fat derived biodiesel on the performance and emission characteristics of a compression ignition engine. Int. J. Ambient Energy 2018, 41, 88–97. [Google Scholar] [CrossRef]
  6. Prates, J.; Alfaia, C.; Alves, S.; Bessa, R. Fatty acids. In Handbook of Analysis of Edible Animal by-Products; Nollet, L.M.L., Toldrá, F., Eds.; CRC Press: Boca Raton, FL, USA, 2011; pp. 137–159. [Google Scholar]
  7. Mora, L.; Toldrá-Reig, F.; Prates, J.A.M.; Toldrá, F. Cattle by-products. In Byproducts from Agriculture and Fisheries: Adding Value for Food, Feed, Pharma and Fuels; Simpson, B.K., Aryee, A.N., Toldrá, F., Eds.; Wiley: Chichester, UK, 2020; pp. 43–55. [Google Scholar]
  8. Baladincz, P.; Hancsók, J. Fuel from waste animal fats. Chem. Eng. J. 2015, 282, 152–160. [Google Scholar] [CrossRef]
  9. Mora, L.; Toldrá-Reig, F.; Reig, M.; Toldrá, F. Possible uses of processed slaughter by-products. In Sustainable Meat Production and Processing; Galanakis, C.M., Ed.; Academic Press/Elsevier: London, UK, 2019; pp. 145–160. [Google Scholar]
  10. Banković-Ilić, I.B.; Stojković, I.J.; Stamenković, O.S.; Veljkovic, V.B.; Hung, Y.-T. Waste animal fats as feed stocks for biodiesel production. Renew. Sustain. Energy Rev. 2014, 32, 238–254. [Google Scholar] [CrossRef] [Green Version]
  11. Ramos, M.; Dias, A.P.S.; Puna, J.F.; Gomes, J.; Bordado, J.C. Biodiesel Production Processes and Sustainable Raw Materials. Energies 2019, 12, 4408. [Google Scholar] [CrossRef] [Green Version]
  12. Flach, B.; Lieberz, S.; Bolla, S. EU Biofuels Annual 2019, Gain Report NL9022, USDA Foreign Agricultural Service. 2019. Available online: http://gain.fas.usda.gov/Pages/Default.aspx (accessed on 6 May 2020).
  13. US Energy Information Administration. Monthly Biodiesel Production Report; US Department of Energy: Washington, DC, USA, 2020. Available online: https://www.eia.gov/biofuels/biodiesel/production/biodiesel.pdf (accessed on 5 May 2020).
  14. IPPR. Time for Change: A New Vision for the British Economy—The Interim Report of the IPPR Commission on Economic Justice. 2017. Available online: http://www.ippr.org/cej-time-for-change (accessed on 25 March 2020).
  15. IRENA. Global Energy Transformation: A Roadmap to 2050. 2018. Available online: https://www.irena.org/-/media/Files/IRENA/Agency/Publication/2018/Apr/IRENA_Report_GET_2018.pdf (accessed on 14 April 2020).
  16. Cernat, A.; Pana, C.; Negurescu, N.; Lazaroiu, G.; Nutu, C.; Fuiorescu, D.; Toma, M.; Nicolici, A. Combustion of preheated raw animal fats-diesel fuel blends at diesel engine. J. Therm. Anal. Calorim. 2019, 140, 2369–2375. [Google Scholar] [CrossRef]
  17. Toldrá-Reig, F.; Mora, L.; Toldrá, F. Trends in biodiesel production from animal fat waste. Appl. Sci. 2020, 10, 3644. [Google Scholar]
  18. Lawan, I.; Garba, Z.N.; Zhou, W.; Zhang, M.; Yuan, Z. Synergies between the microwave reactor and CaO/zeolite catalyst in waste lard biodiesel production. Renew. Energy 2020, 145, 2550–2560. [Google Scholar] [CrossRef]
  19. Bušić, A.; Kundas, S.; Morzak, G.; Belskaya, H.; Marđetko, N.; Santek, M.I.; Komes, D.; Novak, S.; Šantek, B. Recent Trends in Biodiesel and Biogas Production. Food Technol. Biotechnol. 2018, 56, 152–173. [Google Scholar] [CrossRef] [PubMed]
  20. Mahlia, T.; Syazmi, Z.; Mofijur, M.; Abas, A.P.; Bilad, M.; Ong, H.C.; Silitonga, A. Patent landscape review on biodiesel production: Technology updates. Renew. Sustain. Energy Rev. 2020, 118, 109526. [Google Scholar] [CrossRef]
  21. Bockey, D. The significance and perspective of biodiesel production—A European and global view. OCL 2019, 26, 40. [Google Scholar] [CrossRef]
  22. Biodiesel. Available online: https://www.biodiesel.org/what-is-biodiesel/biodiesel-basics (accessed on 21 March 2020).
  23. Kristi, M.; Milbrandt, A.; Lewis, J.; Schwab, A. Bioenergy Industry Status 2017 Report; National Renewable Energy Laboratory: Golden, CO, USA, 2018. Available online: https://www.nrel.gov/docs/fy20osti/75776.pdf (accessed on 5 May 2020).
  24. Canakci, M.; Sanli, H. Biodiesel production from various feedstocks and their effects on the fuel properties. J. Ind. Microbiol. Biotechnol. 2008, 35, 431–441. [Google Scholar] [CrossRef] [PubMed]
  25. Wancura, J.H.C.; Tres, M.V.; Jahn, S.L.; de Oliveira, J.V. Lipases in liquid formulation for biodiesel production: Current status and challenges. Biotechnol. Appl. Biochem. 2019. [Google Scholar] [CrossRef] [PubMed]
  26. Shin, H.-Y.; Lee, S.-H.; Ryu, J.-H.; Bae, S.-Y. Biodiesel production from waste lard using supercritical methanol. J. Supercrit. Fluids 2012, 61, 134–138. [Google Scholar] [CrossRef]
  27. Marulanda, V.F.; Anitescu, G.; Tavlarides, L.L. Investigations on supercritical transesterification of chicken fat for biodiesel production from low-cost lipid feedstocks. J. Supercrit. Fluids 2010, 54, 53–60. [Google Scholar] [CrossRef]
  28. Thangaraj, B.; Solomon, P.-R.; Muniyandi, B.; Ranganathan, S.; Lin, L. Catalysis in biodiesel production—A review. Clean Energy 2019, 3, 2–23. [Google Scholar] [CrossRef] [Green Version]
  29. Vakros, J. Biochars and Their Use as Transesterification Catalysts for Biodiesel Production: A Short Review. Catalysts 2018, 8, 562. [Google Scholar] [CrossRef] [Green Version]
  30. Lee, J.; Jung, J.-M.; Oh, J.-I.; Ok, Y.S.; Lee, S.-R.; Kwon, E.E. Evaluating the effectiveness of various biochars as porous media for biodiesel synthesis via pseudo-catalytic transesterification. Bioresour. Technol. 2017, 231, 59–64. [Google Scholar] [CrossRef]
  31. Issariyakul, T.; Kulkarni, M.G.; Dalai, A.K.; Bakhshi, N.N. Production of biodiesel from waste fryer grease using mixed methanol/ethanol system. Fuel Process Technol. 2007, 88, 429–436. [Google Scholar] [CrossRef]
  32. Melani, N.B.; Tambourgi, E.B.; Silveira, E. Lipases: From Production to Applications. Sep. Purif. Rev. 2019, 49, 143–158. [Google Scholar] [CrossRef]
  33. Maldonado, R.R.; Lopes, D.B.; Aguiar-Oliveira, E.; Kamimura, E.S.; Macedo, G.A. A review on geotrichum lipases: Production, purification, immobilization and applications. Chem. Biochem. Eng. 2016, 30, 439–454. [Google Scholar] [CrossRef]
  34. Alzuhair, S.; Ling, F.W.; Limsong, J. Proposed kinetic mechanism of the production of biodiesel from palm oil using lipase. Process Biochem. 2007, 42, 951–960. [Google Scholar] [CrossRef]
  35. Sugihara, A.; Tani, T.; Tominaga, Y. Purification and characterization of a novel thermostable lipase from Bacillus sp. J. Biochem. 1991, 109, 211–216. [Google Scholar]
  36. Lanser, A.C.; Manthey, L.K.; Hou, C.T. Regioselectivity of new bacterial lipases determined by hydrolysis of triolein. Curr. Microbiol. 2002, 44, 336–340. [Google Scholar] [CrossRef]
  37. Tsurumura, T.; Tsuge, H. Substrate selectivity of bacterial monoacylglycerol lipase based on crystal structure. J. Struct. Funct. Genom. 2014, 15, 83–89. [Google Scholar] [CrossRef]
  38. Li, P.-Y.; Zhang, Y.-Q.; Zhang, Y.; Jiang, W.-X.; Wang, Y.-J.; Zhang, Y.-S.; Sun, Z.-Z.; Li, C.-Y.; Zhang, Y.-Z.; Shi, M.; et al. Study on a Novel Cold-Active and Halotolerant Monoacylglycerol Lipase Widespread in Marine Bacteria Reveals a New Group of Bacterial Monoacylglycerol Lipases Containing Unusual C(A/S)HSMG Catalytic Motifs. Front. Microbiol. 2020, 11, 9. [Google Scholar] [CrossRef]
  39. Davranov, K. Microbial lipases in biotechnology (review). Appl. Biochem. Microbiol. 1994, 30, 527–534. [Google Scholar]
  40. Jaeger, K.-E.; Ransac, S.; Dijkstra, B.W.; Colson, C.; Heuvel, M.; van Misset, O. Bacterial lipases. FEMS Microbiol. Rev. 1994, 15, 29–63. [Google Scholar] [CrossRef]
  41. Wei, L.; Li, R.W.; Qiang, L.; Wei, D.; Liu, D.H. Acyl migration and kinetics study of 1(3)-positional specific lipase of Rhizopus oryzae-catalyzed methanolysis of triglyceride for biodiesel production. Process Biochem. 2010, 45, 1888–1893. [Google Scholar]
  42. Sánchez, D.A.; Tonetto, G.M.; Ferreira, M.L. Burkholderia cepacia lipase: A versatile catalyst in synthesis reactions. Biotechnol. Bioeng. 2018, 115, 6–24. [Google Scholar] [CrossRef] [PubMed]
  43. Encinar, J.M.; González, J.F.; Sánchez, N.; Nogales-Delgado, S. Sunflower oil transesterification with methanol using immobilized lipase enzymes. Bioproc. Biosyst. Eng. 2019, 42, 157–166. [Google Scholar] [CrossRef] [PubMed]
  44. Riyadi, F.A.; Alam, M.Z.; Salleh, M.N.; Salleh, H.M. Optimization of thermostable organic solvent-tolerant lipase production by thermotolerant Rhizopus sp. Using solid-state fermentation of palm kernel cake. 3 Biotech 2017, 7, 300. [Google Scholar] [CrossRef]
  45. Oliveira, F.; Moreira, C.; Salgado, J.M.; Abrunhosa, L.; Venancio, A.; Belo, I. Olive pomace valorization by Aspergillus species: Lipase production using solid-state fermentation. J. Sci. Food Agric. 2016, 96, 3583–3589. [Google Scholar] [CrossRef] [Green Version]
  46. Pandey, N.; Dhakar, K.; Jain, R.; Pandey, A. Temperature dependent lipase production from cold and pH tolerant species of Penicillium. Mycosphere 2016, 7, 1533–1545. [Google Scholar] [CrossRef]
  47. Calabrò, V.; Ricca, E.; De Paola, M.G.; Curcio, S.; Iorio, G. Kinetics of enzymatic trans-esterification of glycerides for biodiesel production. Bioproc. Biosyst. Eng. 2010, 33, 701–710. [Google Scholar] [CrossRef] [PubMed]
  48. Dizge, N.; Aydiner, C.; Imer, D.Y.; Bayramoglu, M.; Tanriseven, A.; Keskinler, B. Biodiesel production from sunflower, soybean, and waste cooking oils by transesterification using lipase immobilized onto a novel microporous polymer. Bioresour. Technol. 2009, 100, 1983–1991. [Google Scholar] [CrossRef] [PubMed]
  49. Gog, A.; Roman, M.; Tos, M.; Paizs, C.; Dan Irimie, F. Biodiesel production using enzymatic transesterification e Current state and Perspectives. Renew. Energy 2012, 39, 10–16. [Google Scholar] [CrossRef]
  50. Shah, S.; Gupta, M.N. Lipase catalyzed preparation of biodiesel from Jatropha oil in a solvent free system. Process Biochem. 2007, 42, 409–414. [Google Scholar] [CrossRef]
  51. Cubides-Roman, D.C.; Perez, V.H.; de Castro, H.F.; Orrego, C.E.; Giraldo, O.H.; Silveira, E.G.; David, G.F. Ethyl esters (biodiesel) production by Pseudomonas fluorescens lipase immobilized on chitosan with magnetic properties in a bioreactor assisted by electromagnetic field. Fuel 2017, 196, 481–487. [Google Scholar] [CrossRef]
  52. Li, W.; Du, W.; Liu, D. Rhizopus oryzae IFO 4697 whole cell catalyzed methanolysis of crude and acidified rapeseed oils for biodiesel production in tert-butanol system. Process Biochem. 2007, 42, 1481–1485. [Google Scholar] [CrossRef]
  53. Kaieda, M.; Samukawa, T.; Kondo, A.; Fukuda, H. Effect of methanol and water contents on production of biodiesel fuel from plant oil catalyzed by various lipases in a solvent-free system. J. Biosci. Bioeng. 2001, 91, 12–15. [Google Scholar] [CrossRef]
  54. Li, K.; Wang, J.; He, Y.; Cui, G.; Abdulrazaq, M.A.; Yan, Y. Enhancing enzyme activity and enantioselectivity of Burkholderia cepacia lipase via immobilization on melamine glutaraldehyde dendrimer modified magnetic nanoparticles. Chem. Eng. J. 2018, 351, 258–268. [Google Scholar] [CrossRef]
  55. Kim, S.H.; Kim, S.; Park, S.; Kim, H.K. Biodiesel production using cross-linked Staphylococcus haemolyticus lipase immobilized on solid polymeric carriers. J. Mol. Catal. B Enzym. 2013, 85–86, 10–16. [Google Scholar] [CrossRef]
  56. Shah, S.; Sharma, S.; Gupta, M.N. Biodiesel preparation by lipase-catalyzed transesterification of Jatropha oil. Energy Fuels 2004, 18, 154–159. [Google Scholar] [CrossRef]
  57. Ji, Q.; Wang, B.; Tan, J.; Zhu, L.; Li, L. Immobilized multienzymatic systems for catalysis of cascade reactions. Process Biochem. 2016, 51, 1193–1203. [Google Scholar] [CrossRef]
  58. Handayani, R.; Wahyuningrum, D.; Zulfikar, M.A.; Nurbaiti, S.; Radiman, C.L.; Buchari, R. The synthesis of biodiesel catalyzed by Mucor miehei lipase immobilized onto aminated polyethersulfone membranes. Bioresour. Bioproc. 2016, 3, 22. [Google Scholar] [CrossRef] [Green Version]
  59. Costa Rodrigues, R.; Volpato, G.; Ayub, M.A.Z.; Wada, K. Lipase-catalysed ethanolysis of soybean oil in a solvent-free system using central composite design and response surface methodology. J. Chem. Technol. Biotechnol. 2008, 83, 849–854. [Google Scholar] [CrossRef]
  60. Ashjari, M.; Garmroodi, M.; Asl, F.A.; Emampour, M.; Yousefi, M.; Lish, M.P.; Habibi, Z.; Mohammadi, M. Application of multi-component reaction for covalent immobilization of two lipases on aldehyde-functionalized magnetic nanoparticles; production of biodiesel from waste cooking oil. Process Biochem. 2020, 90, 156–167. [Google Scholar] [CrossRef]
  61. Chen, G.; Ying, M.; Li, W. Enzymatic conversion of waste-cooking oils into alternative fuel biodiesel. Appl. Biochem. Biotechnol. 2006, 132, 911–921. [Google Scholar] [CrossRef]
  62. Arumugam, A.; Ponnusami, V. Production of biodiesel by enzymatic transesterification of waste sardine oil and evaluation of its engine performance. Heliyon 2017, 3, 00486. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Lv, L.; Dai, L.; Du, W.; Liu, D. Effect of water on lipase NS81006-catalyzed alcoholysis for biodiesel production. Process Biochem. 2017, 58, 239–244. [Google Scholar] [CrossRef]
  64. Touqeer, T.; Mumtaz, M.W.; Mukhtar, H.; Irfan, A.; Akram, S.; Shabbir, A.; Rashid, U.; Nehdi, I.A.; Choong, T.S.Y. Fe3O4-PDA-Lipase as Surface Functionalized NanoBiocatalyst for the Production of Biodiesel Using Waste Cooking Oil as Feedstock: Characterization and Process Optimization. Energies 2020, 13, 177. [Google Scholar] [CrossRef] [Green Version]
  65. Kaieda, M.; Samukawa, T.; Matsumoto, T.; Ban, K.; Kondo, A.; Shimada, Y.; Noda, H.; Nomoto, F.; Ohtsuka, K.; Izumoto, E.; et al. Biodiesel fuel production from plant oil catalyzed by Rhizopus oryzae lipase in a water-containing system without an organic solvent. J. Biosci. Bioeng. 1999, 88, 627–631. [Google Scholar] [CrossRef]
  66. Duarte, S.H.; Hernández, G.L.P.; Canet, A.; Benaiges, M.D.; Maugeria, F.; Valero, F. Enzymatic biodiesel synthesis from yeast oil using immobilized recombinant Rhizopus oryzae lipase. Bioresour. Technol. 2015, 183, 175–180. [Google Scholar] [CrossRef]
  67. Shahedi, M.; Yousefi, M.; Habibi, Z.; Mohammadi, M.; As’habi, M.A. Co-immobilization of Rhizomucor miehei lipase and Candida antarctica lipase B and optimization of biocatalytic biodiesel production from palm oil using response surface methodology. Renew. Energy 2019, 141, 847–857. [Google Scholar] [CrossRef]
  68. Matsuda, T.; Marukado, R.; Mukouyama, M.; Harada, T.; Nakamura, K. Asymmetric reduction of ketones by Geotrichum candidum: Immobilization and application to reactions using supercritical carbon dioxide. Tetrahedron Asymmetry 2008, 19, 2272–2275. [Google Scholar] [CrossRef]
  69. Adewale, P.; Dumont, J.-M.; Ngadi, M. Enzyme-catalyzed synthesis and kinetics of ultrasonic assisted methanolysis of waste lard for biodiesel production. Chem. Eng. J. 2016, 284, 158–165. [Google Scholar] [CrossRef]
  70. Wang, L.; Liu, X.; Jiang, Y.; Liu, P.; Zhou, L.; Ma, L.; He, Y.; Li, H.; Gao, J. Silica Nanoflowers-Stabilized Pickering Emulsion as a Robust Biocatalysis Platform for Enzymatic Production of Biodiesel. Catalysts 2019, 9, 1026. [Google Scholar] [CrossRef] [Green Version]
  71. Antonio, D.C.; Amancio, L.P.; Rosset, I.G. Biocatalytic Ethanolysis of Waste Chicken Fat for Biodiesel Production. Catal. Lett. 2018, 148, 3214–3222. [Google Scholar] [CrossRef]
  72. Lara, P.V.; Park, E.Y. Potential application of waste activated bleaching earth on the production of fatty acid alkyl esters using Candida cylindracea lipase in organic solvent system. Enzym. Microb. Technol. 2004, 34, 270–277. [Google Scholar] [CrossRef]
  73. Matinja, A.I.; Zain, N.A.M.; Suhaimi, M.S.; Alhassan, A.J. Optimization of biodiesel production from palm oil mill effluent using lipase immobilized in PVA alginate- sulfate beads. Renew. Energy 2019, 135, 1178–1185. [Google Scholar] [CrossRef]
  74. Lee, J.H.; Kim, S.B.; Yoo, H.Y.; Lee, J.H.; Han, S.O.; Park, C.; Kim, S.W. Co-immobilization of Candida rugosa and Rhyzopus oryzae lipases and biodiesel production. Korean J. Chem. Eng. 2013, 30, 1335–1338. [Google Scholar] [CrossRef]
  75. Christopher, L.P.; Kumar, H.; Zambare, V.P. Enzymatic biodiesel: Challenges and opportunities. Appl. Energy 2014, 119, 497–520. [Google Scholar] [CrossRef]
  76. Bandikari, R.; Qian, J.; Baskaran, R.; Liu, Z.; Wu, G. Bio-affinity mediated immobilization of lipase onto magnetic cellulose nanospheres for high yield biodiesel in one time addition of methanol. Bioresour. Technol. 2018, 249, 354–360. [Google Scholar] [CrossRef]
  77. López, E.N.; Medina, A.R.; Moreno, P.A.G.; Cerdán, L.E.; Valverde, L.M.; Grima, E.M. Biodiesel production from Nannochloropsis gaditana lipids through transesterification catalyzed by Rhizopus oryzae lipase. Bioresour. Technol. 2016, 203, 236–244. [Google Scholar] [CrossRef]
  78. Iso, M.; Chen, B.; Eguchi, M.; Kudo, T.; Shrestha, S. Production of biodiesel fuel from triglycerides and alcohol using immobilized lipase. J. Mol. Catal. B Enzym. 2001, 16, 53–58. [Google Scholar] [CrossRef]
  79. Goembira, F.; Saka, S. Optimization of biodiesel production by supercritical methyl acetate. Bioresour. Technol. 2013, 131, 47–52. [Google Scholar] [CrossRef] [Green Version]
  80. Razack, S.A.; Duraiarasan, S. Response surface methodology assisted biodiesel production from waste cooking oil using encapsulated mixed enzyme. Waste Manage. 2016, 47, 98–104. [Google Scholar] [CrossRef]
  81. Wang, X.; Qin, X.; Li, D.; Yang, B.; Wang, Y. One-step synthesis of high-yield biodiesel from waste cooking oils by a novel and highly methanol-tolerant immobilized lipase. Bioresour. Technol. 2017, 235, 18–24. [Google Scholar] [CrossRef] [PubMed]
  82. Hwang, H.T.; Qi, F.; Yuan, C.; Zhao, X.; Ramkrishna, D.; Liu, D.; Varma, A. Lipase-Catalyzed Process for Biodiesel Production: Protein Engineering and Lipase Production. Biotechnol. Bioeng. 2014, 111, 639. [Google Scholar] [CrossRef] [PubMed]
  83. Pollardo, A.A.; Lee, H.; Lee, D.; Kim, S.; Kim, J. Solvent effect on the enzymatic production of biodiesel from waste animal fat. J. Clean. Prod. 2018, 185, 382–388. [Google Scholar] [CrossRef]
  84. Kumar, D.; Das, T.; Giri, B.S.; Verma, B. Preparation and characterization of novel hybrid bio-support material immobilized from Pseudomonas cepacia lipase and its application to enhance biodiesel production. Renew. Energy 2020, 147, 11–24. [Google Scholar] [CrossRef]
  85. Al-Zuhair, S.; Hasan, M.; Ramachandran, K. Kinetics of the enzymatic hydrolysis of palm oil by lipase. Process Biochem. 2003, 38, 1155–1163. [Google Scholar] [CrossRef]
  86. Chesterfield, D.M.; Rogers, P.L.; Al-Zaini, E.O.; Adesina, A.A. Production of biodiesel via ethanolysis of waste cooking oil using immobilised lipase. Chem. Eng. J. 2012, 207, 701–710. [Google Scholar] [CrossRef]
  87. Remonatto, D.; Santin, C.M.T.; Oliveira, D.; Di Luccio, M.; Oliveira, J.V. FAME Production from Waste Oils Through Commercial Soluble Lipase Eversa®Catalysis. Ind. Biotechnol. 2016, 12, 254–262. [Google Scholar] [CrossRef]
  88. Wancura, J.H.C.; Rosset, D.V.; Tres, M.V.; Oliveira, J.V.; Mazutti, M.A.; Jahn, S.L. Production of biodiesel catalyzed by lipase from Thermomyces lanuginosus in its soluble form. Can. J. Chem. Eng. 2018, 96, 2361–2368. [Google Scholar] [CrossRef]
  89. Lee, K.T.; Foglia, T.A.; Chang, K.S. Production of alkyl ester as biodiesel from fractionated lard and restaurant grease. J. Am. Oil Chem. Soc. 2002, 79, 191–195. [Google Scholar] [CrossRef]
  90. Huang, Y.; Zheng, H.; Yan, Y. Optimization of Lipase-Catalyzed Transesterification of lard for Biodiesel Production Using Response Surface Methodology. Appl. Microbiol. Biotechnol. 2010, 160, 504–515. [Google Scholar] [CrossRef]
  91. Lu, J.; Nie, K.; Xie, F.; Wang, F.; Tan, T. Enzymatic synthesis of fatty acid methyl esters from lard with immobilized Candida sp. 99–125. Process Biochem. 2007, 42, 1367–1370. [Google Scholar] [CrossRef]
  92. Da Silva, J.R.P.; da Costa, F.P.; Lerin, L.A.; Ninow, J.L.; Oliveira, V.; de Oliveira, D. Application of Different Methodologies to Produce Fatty Acid Esters Using the Waste Chicken Fat Catalyzed by Free NS 40116 Lipase. Ind. Biotechnol. 2019, 5, 293–302. [Google Scholar] [CrossRef]
  93. Poppe, J.K.; Fernandez-Lafuente, R.; Rodrigues, R.C.; Záchia Ayub, M.A. Enzymatic reactors for biodiesel synthesis: Present status and future prospects. Biotechnol. Adv. 2015, 33, 511–525. [Google Scholar] [CrossRef] [PubMed]
  94. Wang, J.; Li, K.; He, Y.; Wang, Y.; Han, X.; Yan, Y. Enhanced performance of lipase immobilized onto Co2+-chelated magnetic nanoparticles and its application in biodiesel production. Fuel 2019, 255, 115794. [Google Scholar] [CrossRef]
  95. Cui, J.D.; Jia, S.R. Optimization protocols and improved strategies of cross-linked enzyme aggregates technology: Current development and future challenges. Crit. Rev. Biotechnol. 2015, 35, 15–28. [Google Scholar] [CrossRef]
  96. Yan, J.; Yan, Y.; Liu, S.; Hu, J.; Wang, G. Preparation of cross-linked lipase-coated micro-crystals for biodiesel production from waste cooking oil. Bioresour. Technol. 2011, 102, 4755–4758. [Google Scholar] [CrossRef]
  97. Lopresto, C.G.; De Paola, M.G.; Albo, L.; Policicchio, M.F.; Chakraborty, S.; Calabro, V. Comparative analysis of immobilized biocatalyst: Study of process variables in trans-esterification reaction. 3 Biotech 2019, 9, 443. [Google Scholar] [CrossRef]
  98. Mateo, C.; Palomo, J.M.; Fernandez-Lorente, G.; Guisan, J.M.; Fernandez-Lafuente, R. Improvement of enzyme activity, stability and selectivity via immobilization techniques. Enzym. Microb. Technol. 2007, 40, 1451–1463. [Google Scholar] [CrossRef]
  99. Zhou, Z.; Hartmann, M. Progress in enzyme immobilization in ordered mesoporous materials and related applications. Chem. Soc. Rev. 2013, 42, 3894–3912. [Google Scholar] [CrossRef]
  100. Babaki, M.; Yousefi, M.; Habibi, Z.; Mohammadi, M. Process optimization for biodiesel production from waste cooking oil using multi-enzyme systems through response surface methodology. Renew. Energy 2017, 105, 465–472. [Google Scholar] [CrossRef] [Green Version]
  101. Da Rós, P.C.M.; Silva, G.A.M.; Mendes, A.A.; Santos, J.C.; de Castro, H.F. Evaluation of the catalytic properties of Burkholderia cepacia lipase immobilized on non-commercial matrices to be used in biodiesel synthesis from different feedstocks. Bioresour. Technol. 2010, 101, 5508–5516. [Google Scholar] [CrossRef] [PubMed]
  102. Angulo, B.; Fraile, J.M.; Gil, L.; Herrerías, C.I. Comparison of Chemical and Enzymatic Methods for the Transesterification of Waste Fish Oil Fatty Ethyl Esters with Different Alcohols. ACS Omega 2020, 5, 1479−1487. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Marin-Suarez, M.; Mendez-Mateos, D.; Guadix, A.; Guadix, E.M. Reuse of immobilized lipases in the transesterification of waste fish oil for the production of biodiesel. Renew. Energy 2019, 140, 1–8. [Google Scholar] [CrossRef]
  104. Rodrigues, R.C.; Virgen-Ortíz, J.J.; dos Santos, J.C.; Berenguer-Murcia, Á.; Alcantara, A.R.; Barbosa, O.; Ortiz, C.; Fernandez-Lafuente, R. Immobilization of lipases on hydrophobic supports: Immobilization mechanism, advantages, problems, and solutions. Biotechnol. Adv. 2019, 37, 746–770. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Manoel, E.A.; dos Santos, J.C.; Freire, D.M.; Rueda, N. Fernandez-Lafuente, R. Immobilization of lipases on hydrophobic supports involves the open form of the enzyme. Enzym. Microb. Technol. 2015, 71, 53–57. [Google Scholar] [CrossRef] [PubMed]
  106. Quilles, J.C., Jr.; Ferrarezi, A.L.; Borges, J.P.; Rossi, J.S.; Bocchini, D.A.; Gomes, E.; da Silva, R.; Boscolo, M. Ultrasound affects the selectivity and activity of immobilized lipases applied to fatty acid ethyl ester synthesis. Acta Sci-Technol. 2020, 42, 46582. [Google Scholar] [CrossRef] [Green Version]
  107. Zhou, W.J.; Fang, L.; Fan, Z.; Albela, B.; Bonneviot, L.; De Campo, F.; Pera Titus, M.; Clacens, J.M. Tunable catalysts for solvent-free biphasic systems: Pickering interfacial catalysts over amphiphilic silica nanoparticles. J. Am. Chem. Soc. 2014, 136, 4869–4872. [Google Scholar] [CrossRef]
  108. Jiang, Y.; Liu, X.; Chen, Y.; Zhou, L.; He, Y.; Ma, L.; Gao, J. Pickering emulsion stabilized by lipase-containing periodic mesoporous organosilica particles: A robust biocatalyst system for biodiesel production. Bioresour. Technol. 2014, 153, 278–283. [Google Scholar] [CrossRef]
  109. Zhang, W.; Fu, L.; Yang, H. Micrometer-Scale Mixing with Pickering Emulsions: Biphasic Reactions without Stirring. ChemSusChem 2014, 7, 391–396. [Google Scholar] [CrossRef]
  110. Wang, L.; Liu, X.; Jiang, Y.; Zhou, L.; Ma, L.; He, Y.; Gao, J. Biocatalytic Pickering Emulsions Stabilized by Lipase-Immobilized Carbon Nanotubes for Biodiesel Production. Catalysts 2018, 8, 587. [Google Scholar] [CrossRef] [Green Version]
  111. Shi, J.; Wang, X.; Zhang, S.; Tang, L.; Jiang, Z. Enzyme-conjugated ZIF-8 particles as efficient and stable Pickering interfacial biocatalysts for biphasic biocatalysis. J. Mater. Chem. B 2016, 4, 2654–2661. [Google Scholar] [CrossRef] [PubMed]
  112. Chen, Z.; Zhao, C.; Ju, E.; Ji, H.; Ren, J.; Binks, B.P.; Qu, X. Design of Surface-Active Artificial Enzyme Particles to Stabilize Pickering Emulsions for High-Performance Biphasic Biocatalysis. Adv. Mater. 2016, 28, 1682–1688. [Google Scholar] [CrossRef] [PubMed]
  113. Yang, B.; Leclercq, L.; Clacensb, J.M.; Rataj, V.N. Acidic/amphiphilic silica nanoparticles: New eco-friendly Pickering interfacial catalysis for biodiesel production. Green Chem. 2017, 19, 4552–4562. [Google Scholar] [CrossRef]
  114. Talin, A.A.; Centrone, A.; Ford, A.C.; Foster, M.E.; Stavila, V.; Haney, P.; Kinney, R.A.; Szalai, V.; El Gabaly, F.; Yoon, H.P.; et al. Tunable electrical conductivity in metal-organic framework thin-film devices. Science 2014, 343, 66–69. [Google Scholar] [CrossRef] [PubMed]
  115. Lin, J.; Liu, Y.; Chen, S.; Le, X.; Zhou, X.; Zhao, Z.; Ou, Y.; Yang, J. Reversible immobilization of laccase onto metal-ion-chelated magnetic microspheres for bisphenol A removal. Int. J. Biol. Macromol. 2016, 84, 189–199. [Google Scholar] [CrossRef]
  116. Netto, C.G.; Toma, H.E.; Andrade, L.H. Superparamagnetic nanoparticles as versatile carriers and supporting materials for enzymes. J. Mol. Catal. B Enzym. 2013, 85, 71–92. [Google Scholar] [CrossRef]
  117. Bilal, M.; Zhao, Y.; Rasheed, T.; Iqbal, H.M.N. Magnetic nanoparticles as versatile carriers for enzymes immobilization: A review. Int. J. Biol. Macromol. 2018, 120, 2530–2544. [Google Scholar] [CrossRef]
  118. Yang, H.; Zhang, W. Surfactant Imprinting Hyperactivated Immobilized Lipase as Efficient Biocatalyst for Biodiesel Production from Waste Cooking Oil. Catalysts 2019, 9, 914. [Google Scholar] [CrossRef] [Green Version]
  119. Ngo, T.P.N.; Li, A.; Tiew, K.-W.; Li, Z. Efficient transformation of grease to biodiesel using highly active and easily recyclable magnetic nanobiocatalyst aggregates. Bioresour. Technol. 2013, 145, 233–239. [Google Scholar] [CrossRef]
  120. Chiaradia, V.; Soares, N.S.; Valério, A.; de Oliveira, D.; Araújo, P.H.H.; Sayer, C. Immobilization of Candida antarctica Lipase B on Magnetic Poly(Urea-Urethane) Nanoparticles. Appl. Biochem. Biotechnol. 2016, 180, 558–575. [Google Scholar] [CrossRef]
  121. Miao, C.; Yang, L.; Wang, Z.; Luo, W.; Li, H.; Lv, P.; Yuan, Z. Lipase immobilization on amino-silane modified superparamagnetic Fe3O4 nanoparticles as biocatalyst for biodiesel production. Fuel 2018, 224, 774–782. [Google Scholar] [CrossRef]
  122. Mijone, P.D.; Vilas-Boas, R.N.; Bento, H.B.S.; Rodrigues, B.S.B.; de Castro, R.H.F. Coating and incorporation of iron oxides into a magnetic-polymer composite to be used as lipase support for ester syntheses. Renew. Energy 2020, 149, 1167–1173. [Google Scholar] [CrossRef]
  123. Tran, D.-T.; Chen, C.-L.; Chang, J.-S. Immobilization of Burkholderia sp. Lipase on a ferric silica nanocomposite for biodiesel production. J. Biotechnol. 2012, 158, 112–119. [Google Scholar] [CrossRef] [PubMed]
  124. Xie, W.; Wang, J. Immobilized lipase on magnetic chitosan microspheres for transesterification of soybean oil. Biomass. Bioenergy 2012, 36, 373–380. [Google Scholar] [CrossRef]
  125. Arana-Peña, S.; Rios, N.S.; Mendez-Sanchez, C.; Lokha, Y.; Carballares, D.; Gonçalves, L.R.B.; Fernandez-Lafuente, R. Coimmobilization of different lipases: Simple layer by layer enzyme spatial ordering. Int. J. Biol. Macromol. 2020, 145, 856–864. [Google Scholar] [CrossRef]
  126. Virgen-Ortíz, J.J.; dos Santos, J.C.S.; Berenguer-Murcia, Á.; Barbosa, O.; Rodrigues, R.C.; Fernandez-Lafuente, R. Polyethylenimine: A very useful ionic polymer in the design of immobilized enzyme biocatalysts. J. Mater. Chem. B 2017, 5, 7461–7490. [Google Scholar] [CrossRef] [Green Version]
  127. Nielsen, P.M. Production of Fatty Acid Alkyl Esters. World Patent WO/2012/098114, 26 July 2012. [Google Scholar]
  128. Fraga, F.C.; Valério, A.; de Oliveira, V.A.; Di Luccio, M.; de Oliveira, D. Effect of magnetic field on the Eversa®Transform 2.0 enzyme: Enzymatic activity and structural conformation. Int. J. Biol. Macromol. 2019, 122, 653–658. [Google Scholar] [CrossRef]
  129. Viesel. Available online: https://gstarbio.com/es/ (accessed on 25 June 2020).
  130. Chen, X.; Li, L.; Deng, L.; Pedersen, J.N.; Li, L.; Guo, Z.; Cong, F.; Xu, X. Biodiesel Production Using Lipases. In Lipid Modification by Enzymes and Engineered Microbes; Bornscheuer, U.T., Ed.; Academic Press/AOCS Press: London, UK, 2018; pp. 343–373. [Google Scholar]
  131. Kotrba, R. Realizing the Vision to Reclaim, Recycle, Refuel. Biodiesel Magazine. 2015. Available online: http://www.biodieselmagazine.com/articles/265452/realizing-the-vision-to-reclaim-recycle-refuel (accessed on 22 June 2020).
  132. Wancura, J.H.C.; Rosset, D.V.; Brondani, M.; Mazutti, M.A.; de Oliveira, J.V.; Tres, M.V.; Jahn, S.L. Soluble lipase-catalyzed synthesis of methyl esters using a blend of edible and nonedible raw materials. Bioproc. Biosyst. Eng. 2018, 41, 1185–1193. [Google Scholar] [CrossRef]
  133. Ortiz, C.; Ferreira, M.L.; Barbosa, O.; dos Santos, J.C.S.; Rodrigues, R.C.; Berenguer-Murcia, A.; Briand, L.E.; Fernandez-Lafuente, R. Novozym 435: The “perfect” lipase immobilized biocatalyst? Catal. Sci. Technol. 2019, 9, 2380. [Google Scholar] [CrossRef] [Green Version]
  134. Sobhi, B. Modified-Immobilized Enzymes of High Tolerance to Hydrophilic Substrates in Organic Media. US Patent US9,068,175 B2, 30 June 2015. [Google Scholar]
  135. Enzymocore. Available online: https://enzymocore.com/about-us/our-enzymatic-technology/ (accessed on 22 June 2020).
Figure 1. Transesterification of animal fat to biodiesel. TGL: triacylglycerol lipase; nsTGL: non specific triacylglycerol lipase; MGL: monoacylglycerol lipase.
Figure 1. Transesterification of animal fat to biodiesel. TGL: triacylglycerol lipase; nsTGL: non specific triacylglycerol lipase; MGL: monoacylglycerol lipase.
Applsci 10 05085 g001
Figure 2. Major types of biological catalysis for biodiesel generation.
Figure 2. Major types of biological catalysis for biodiesel generation.
Applsci 10 05085 g002
Table 1. Bacteria, yeasts and filamentous fungi producing lipase and sources of isolation.
Table 1. Bacteria, yeasts and filamentous fungi producing lipase and sources of isolation.
Lipase OriginReference
Pseudomonas fluorescens[52]
Burkholderia cepacia[42,50,53,54]
Staphylococcus haemolyticus[55]
Chromobacterium viscosum[56]
Phichia pastoris[57]
Mucor miehei[58]
Thermomyces lanuginosus[59,60]
Aspergillus oryzae[61]
Aspergillus niger[62,63]
Aspergillus terreus[64]
Rhizopus oryzae[41,65,66]
Rhizomucor miehei[60,67]
Geotrichum candidum[68]
Candida antarctica[66,69,70,71]
Candida cylindracea[72]
Candida rugosa[50,73,74]
Table 2. Biodiesel production with various free soluble lipases.
Table 2. Biodiesel production with various free soluble lipases.
Lipase SourceFeedstockConditions
(T, alcohol:oil, t)
Yield (%)References
A. nigerWaste cooking oil45 °C, 4.2:1, 30 h90 a[63]
T. lanoginosusBeef tallow35 °C, 4.5:1, 6 h84.6 a[88]
C. antarcticaLard30 °C, 1:1, 72 h74 a[89]
C. antarcticaLard50 °C, 5:1, 20 h97.2 b[90]
Candida spLard40 °C, 3:1, 30 h87.4 b[91]
T. lanoginosusChicken fat-89.04 b[92]
C. antarcticaChicken fat32 °C, 3:1, 24 h96 a[71]
a Biodiesel yield (wt/wt.%) was determined as the methyl esters amount produced by the lipases in the reaction process divided by the initial amount of esters; or b by the amount of oil.
Table 3. Biodiesel production with various immobilized lipases.
Table 3. Biodiesel production with various immobilized lipases.
Lipase SourceFeedstockImmobilizationYield a (%)No. CyclesReferences
A. nigerSardine oilActivated carbon94.5 a5[62]
T. lanoginosusLardSilica gel97.6 a20[90]
C. antarcticaWaste cooking oilSilica nanoflower pickering emulsion98.5 a15[70]
B. cepaciaCastor oilPolyvinylalcohol alginate75 b6[85]
R. miehei and
C. antarctica
Waste cooking oilEpoxy functionalized silica91.5 a14[100]
Streptomyces spWaste cooking oilXAD 1180 resin95.45 a4[79]
B. cepaciaBeef tallowPolysiloxane–polyvinyl alcohol (SiO2–PVA) hybrid composite89 a-[101]
C. antarcticaWaste fish oilAcrylic resin95 a4[102]
C. antarcticaWaste fish oilAcrylic resin75 a10[103]
a Biodiesel yield (wt/wt.%) was determined as the methyl esters amount produced by the lipases in the reaction process divided by the initial amount of esters, or b by the amount of oil.
Table 4. Biodiesel production with various immobilized lipases using magnetic nanoparticles as carriers.
Table 4. Biodiesel production with various immobilized lipases using magnetic nanoparticles as carriers.
Lipase SourceCarrierImmobilizationYield a (%)No. CyclesReferences
B. cepaciaSilica coated hydroxyapatite and glutaraldehydeEncapsulation984[118]
A. terreusIron oxide polydopamineCovalent bonding927[64]
T. lanuginosis & C. antarcticaCore-shell structured iron oxideCovalent bonding9911[119]
P. fluorescensCo2+ chelated, with 3-glycidoxypropyltrimethoxysylaneAdsorption9510[70]
R. miehei and T. lanuginosaSilica core shell iron oxide with tryethylamineCovalent bonding93.15[60]
C. antarcticaPoly(urea-urethane) encapsulated magnetiteEncapsulation958[120]
C. antarcticaMagnetic iron oxide with 1-Butyl-3-methylimidazolium tetrafluoroborate & 3-aminopropyltriethoxysilaneCovalent bonding89.45[121]
B. cepaciaPolysiloxane–polyvinyl alcohol hybrid magnetic-polymer composite Covalent bonding96.5-[122]
a Biodiesel yield (wt/wt.%) was determined as the methyl esters amount produced by the lipases in the reaction process divided by the initial amount of esters.

Share and Cite

MDPI and ACS Style

Toldrá-Reig, F.; Mora, L.; Toldrá, F. Developments in the Use of Lipase Transesterification for Biodiesel Production from Animal Fat Waste. Appl. Sci. 2020, 10, 5085. https://doi.org/10.3390/app10155085

AMA Style

Toldrá-Reig F, Mora L, Toldrá F. Developments in the Use of Lipase Transesterification for Biodiesel Production from Animal Fat Waste. Applied Sciences. 2020; 10(15):5085. https://doi.org/10.3390/app10155085

Chicago/Turabian Style

Toldrá-Reig, Fidel, Leticia Mora, and Fidel Toldrá. 2020. "Developments in the Use of Lipase Transesterification for Biodiesel Production from Animal Fat Waste" Applied Sciences 10, no. 15: 5085. https://doi.org/10.3390/app10155085

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