Abstract
Due to diminishing fossil fuel supplies and rising energy needs, there has been an ever-increasing demand for renewable energy sources. The available renewable energy resources, such as solar, wind, hydropower, and biofuels, provide a new way of supplying the world's energy needs. Biofuels stand out among them because they are sustainable and have the potential to bring the idea of a global bioeconomy to life. As a result of their production of biofuels like biomethane, biohydrogen, and biodiesel, atmospheric CO2 is being fixed, eventually lowering the world's carbon footprint. Current developments in the production of bioenergy have concentrated on producing biodiesel among other biofuels. Biodiesel is being produced from a variety of feedstocks using a number of processes, including transesterification, micro-emulsion, direct mixing, and pyrolysis. The most popular method among these is transesterification, which makes use of a variety of catalysts. As a result of the development of nanotechnology, nanocatalysts with desirable properties, such as increased catalytic activity, increased surface area, and superior thermal stability, have been made and modified. In this review, various nanocatalyst types and manufacturing processes are examined in relation to transesterification. It explores how crucial nanocatalysts are in boosting biodiesel production, highlights potential barriers, and makes recommendations for their widespread use in the future.
Keywords: Biofuels, Biodiesel, Feedstock, Renewable energy, Transesterification, Nanocatalysts
Graphical abstract
Highlights
-
•
Role of nanomaterials in biodiesel production is explained.
-
•
Different types of nanocatlysts are deliniated.
-
•
Different feedstocks for biodiesel production have been discussed.
-
•
Reusability of nanomaterials and its economic effects are explained.
Abbreviations
- TAG
Triacylglyceride
- GHG
Greenhouse gas
- FFA
Free fatty acid
- FAME
Fatty-acid-methyl-ester
- CNT
Carbon nanotube
- MWNT
Multi-walled carbon nanotube
- s-MWCNT
Sulfonated multiwalled carbon nanotube
- BCL
Burkholderia cepacia lipase
- MNP
Magnetic nanoparticle
- ZSM-5
Zeolite Socony Mobil-5
- NBC
Nanobiocatalyst
1. Introduction
A vast majority of global energy demand is met by fossil fuels, which poses threats to the environment. Continuous globalization and industrialization have increased the use of fossil fuel products such as petroleum and coal, rendering their reserves in an endangered state [1]. The annual worldwide energy requirement is approximately 580 TJ, a whopping 80% of which is met by burning conventional fossil fuels. Increasing population and accessibility necessitate greater energy consumption, further accelerating the depletion of fossil fuel resources [2]. The shortage of energy, limited deposits, rapid increment in petroleum oil prices, and adverse effects of greenhouse gas emissions have necessitated a shift of focus to alternative energy sources. The pursuit of environmentally friendly technology is propelling research attempts to discover viable energy sources that are carbon-neutral, biodegradable, renewable, and non-toxic [3]. With technological advancements, implementing many cutting-edge concepts can drive perceptions and methodologies of curating renewable energy to new horizons and offer greener solutions to address rising demands while conserving the environment [4]. The existing renewable energy sources, such as, hydro, geothermal, solar, wind, and biofuels, contribute to meeting only about 18% of global energy demand, mostly due to the non-implementation of the technologies on a large scale [5]. As a substitute for traditional fuels, biofuel production has garnered much attention from scientific circles in recent decades. Biomass-derived energy is expected to generate half of the net energy demand in industrial nations by 2050 because of its accessibility and eco-friendly characteristics, such as low carbon footprint and generating minimal concentrations of sulfur. The combustion of fossil fuels releases around 27 billion tonnes of CO2 each year, with this figure expected to surge by 60% by 2030. As a result, employing biomass-derived fuels is critical to lowering the carbon footprint [6]. Fig. 1 shows the contribution of biomass-derived energy in meeting global energy demands. Biofuel production primarily focuses on biomass conversion to fluid fuels such as biogas, biohydrogen and biodiesel, which is accomplished mostly by photosynthesis. Among them, biodiesel as a possible fuel is rapidly developing, thus limiting and minimizing the utilization of carbon energy and the rest of nonrenewable resources [7].
Fig. 1.
Contribution of different renewable energy sources for meeting global energy demands.
Biodiesel is a sustainable energy source made composed of hydrocarbons having a long carbon chain and an ester group (-COOR) that is produced from many types of feedstocks, such as oils derived from plants and animals or other lipids termed triacylglycerides (TAGs) [8]. Biodiesel generated from non-food sources is a cost-efficient and reliable substitute for petroleum-derived fuels. It can be utilized directly or mixed in any ratio with conventional diesel. However, production costs, feedstock availability, and complexities in production methods are among the most important elements influencing commercial biodiesel production [9]. Biodiesel generation technologies such as pyrolysis, the supercritical fluid process, and transesterification are widely employed. Transesterification is the most extensively utilized of these processes [10]. Some recent works in the field of biodiesel production can be seen Table 1.
Table 1.
Different nanocatalysts for the synthesis of biodiesel using several feedstocks.
| Feedstock | Nanomaterial | Yield (%) | Reference |
|---|---|---|---|
| Mesua ferrea | Co doped ZnO | 98.03 | [11] |
| Waste cooking oil | MgO | 93.3 | [12] |
| Waste cooking oil | ZnFe2O4 | 98.6 | [13] |
| Mesua ferrea linin | Egg shell derived Cao | 94 | [14] |
| FFA | MoO3/B-ZSM-5 | 98 | [15] |
| Goat fat | MgO | 93.12 | [16] |
| Waste cooking oil | CaO over Zr- doped MCM-41 | 88.5 | [17] |
| Rapessed oil | Potassium impregnated | 96.1 | [18] |
Transesterification is a catalytic process that requires using methanol to transform triglycerides into fatty acid chains in the presence of a homogeneous/heterogeneous catalyst. Homogeneous catalysis is where the catalyst stays in the same state as reactants throughout the process. The catalyst is heterogeneous if it is in a different phase than the reactants [19]. Contrary to those that are catalyzed heterogeneously, homogeneously catalyzed processes are often quicker and call for less loading. However, solid catalysts are very expensive and have a smaller surface area, thereby depreciating the catalyst's total effectiveness [20]. The Development of nanotechnology has led nanomaterials to execute an important role in enhancing biodiesel output and grade while shortening reaction time. Using a nanocatalyst decreases catalyst weight, the reaction temperature, and the oil-to-alcohol ratio [21,22]. It has already been established that nanocatalysts significantly contribute to accelerating the velocity of reaction without getting damaged. It also decreases the activation energy of the reactants substantially.
Furthermore, the quantity of nanocatalysts used is crucial in delimiting biodiesel production output [23]. It was further reported that with no catalyst loading in the specified reaction prerequisites, net biodiesel production was zero, implying that the presence and number of nanocatalysts determine the efficacy of biodiesel synthesis [24]. The esterification process mediated by nanocatalysts has various distinctive benefits, such as a faster mixing rate with the reactants, shorter reaction periods, and quick and simple purification [25].
This review explores different types of nanomaterials used in mediating the transesterification process and emphasizes the importance of nanotechnology in biodiesel production. In addition, this article evaluates their applicability and related obstacles in the transesterification process, focusing on reusability, selectivity, catalytic activity and catalyst loading [[26], [27], [28]]. The prime motive of the review is to gain a thorough understanding of the application of nanotechnology in bioenergy generation and serve as a roadmap for future innovations for sustainable development.
2. Biodiesel
Biofuels are mostly sourced from renewable resources such as wood, bagasse, feedstock, edible and non-edible commodities, and cellulosic and lignocellulosic materials. They are generated using a biological fixation process, resulting in the generation of several kinds of biofuels like bioethanol, biohydrogen, biomethane, and biodiesel [29]. Biodiesel is a fatty acid alkyl ester synthesized by transesterifying triglycerides in presence of an alcohol. Depending on its composition, the generated biodiesel mostly comprises various fatty acids. It has a low sulfur content, is non-toxic, renewable, and biodegradable, and has fewer hazardous gas emissions. Biodiesel can be used in combination with conventional diesel to scale back on the number of pollutants and environmentally hazardous gases generated by petrol, diesel and other fuels [30].
2.1. Feedstock for biodiesel production
A vast range of possible sources for the feedstock, including as oilseeds, algae, and animal fats, have been studied for the manufacturing of biodiesel. The choice of raw materials for biodiesel synthesis is an important area of consideration since some of the feedstock materials are expensive and utilizing them will render the production unprofitable [31]. Feedstock is mostly constituted of oils and fats sourced from organic biomass of plants and animals and can be classified into three main generations.
-
1.
First generation feedstock comprising of edible oils derived from seeds of sunflower [32], canola, rapeseed, coconut, etc. [33].
-
2.
Second generation feedstock consisting of non-edible oils like jatropha oil [34], oil derived from sewage and kitchen wastes, discarded frying oil, etc. [35].
-
3.
Third generation feedstock which mainly include oil derived from microorganisms such as algae, yeast and other lipid-rich microbes [36].
The three generations of feedstocks come with their own benefits and drawbacks, a few of which have been listed in Table 2 and Fig. 2 [31].
Table 2.
Advantages and Disadvantages of three generations of feedstock.
| Feedstock generation | Source | Advantages | Disadvantages | Reference |
|---|---|---|---|---|
| First | Edible oil | 1.Environmentally safe | 1.Food vs Feed problem 2.Lack of agricultural lands for high yield of feedstock and biodiesel |
[[37], [38], [39], [40], [41]] |
| 2.Easy to handle | ||||
| 3.Sustainable | ||||
| Second | Non – edible oil | 1.Sustainable | Purification process is required which is often complex | [[42], [43], [44], [45], [46]] |
| 2.Easy to handle | ||||
| 3.Economically more viable | ||||
| 4.No Food vs Fuel duel | ||||
| 5.Promotes recycling | ||||
| Third | Yeast, algae, microbial biomass | 1.Sustainable | 1.Has not been applied in commercial set ups 2.Lack of statistical knowledge on scale-up |
[[47], [48], [49], [50], [51]] |
| 2.High yields of biodiesel | ||||
| 3.Easily available | ||||
| 4.Economically viable, if used to full potential | ||||
| 5.Doesn't lead to Food vs Feed competition |
Fig. 2.
The three generations of feedstock for Biodiesel production.
The fourth generation of biofuels' raw materials are genetically altered microalgae. These microbes have undergone genetic modification to boost photosynthesis, produce more biofuel, and create an artificial carbon sink. Several algal strains, including Chlamydomonas reinhardtii sp., Phaeodactylum tricornutum sp., and Thalassiosira pseudonana sp., have undergone genetic modification to speed up development and improve their capacity to flourish in environments with little nutrients. The reduction of GHGs, CO2 assimilation, and use of genetically modified microalgae as a wastewater treatment medium are all benefits for the environment. Although though the research is still in its early stages, the fourth generation of biofuel aims to have the least negative environmental impact when compared to prior generations. As a result, the majority of literature research concentrate on the effects of gene alteration on the environment.
3. Transesterification in biodiesel production
Biodiesel production from feedstock oils is achieved via three main routes: pyrolysis, emulsification and transesterification. All of the techniques have some degree of viability, as well as certain benefits and drawbacks that are detailed in Table 3 [52]. Since traditional methods like pyrolysis, blending, and emulsification technologies for synthesis are incompetent for large volumes of biodiesel output, transesterification or alcoholysis is the standard method for producing biodiesel.
Table 3.
Techniques used for Biodiesel production (Adapted from Karmakar and Halder [53]).
| Technique | Advantage | Disadvantage |
|---|---|---|
| Direct blending | No specific feedstock requirement and no need of additional preparations | Not suitable for long haul use |
| Higher viscosity could lead to other issues and blockage | ||
| Thermal cracking or pyrolysis | Potent process with almost no contaminating byproducts. | Expensive set-up and quite costly process |
| No requirement for additional detachment or filtration | Low biodiesel quality because of impurities | |
| Microemulsification | Straightforward cycle | Fragmented ignition and arrival of inordinate carbon stores |
| It is utilized for extremely sticky feedstock | ||
| Transesterification | Gentle response conditions | Requirement of large volumes of water |
| High conversion rate | Development of side residues | |
| Effortless handling | Treatment required for the unsaturated triglycerides | |
| Enzyme catalyzed transesterification | Possess the capacity to transesterify both triglycerides and FFAs, and demand a low oil to alcohol molar ratio. | The creation of waste water hinders the approaches' efficiency. |
| Are able to produce biodiesel with fewer processing steps, using less energy in the process. | A low rate of reaction, especially when immobilized for enhancement of their reusability |
Transesterification involves reacting triglycerides with alcohols (usually methanol) with the aid of a catalyst to speed up the reaction and create fatty acid esters (often, fatty-acid-methyl-esters, or FAME) [54]. Most oils derived from organic biomass are triglycerides and hence, the glycerol backbone needs to be removed to obtain fatty acid ester chains, having similar structural and combustion properties as conventional diesel. The glycerol backbone is removed using transesterification method, which is sped up using a catalyst. The reaction proceeds in stages, beginning with triacylglycerol and progressing through di- and monoacylglycerols to glycerol, generating methyl ester at each stage [55].
The most promising technique is the reversible transesterification reaction, as depicted in Fig. 3. FAMEs (biodiesel) and glycerol are produced by reacting fatty acids (oils) with alcohol in the presence of a suitable catalyst. A powerful base or acid, such as hydrochloric acid or sulfuric acid, or a strong base, such as sodium hydroxide, sodium methoxide, or potassium hydroxide, can be used as a catalyst [56]. Since alcohol is very weakly soluble in the oil phase and uncatalyzed reactions are exceedingly sluggish, the catalyst is typically used to initiate the reaction and serves as an alcohol solubilizer. The catalyst promotes the solubility of alcohol and, as a result, makes it possible for the reaction to advance at a tolerable rate [57]. One of the most crucial factors in the production of biodiesel is the catalyst's unique surface area. Because there are more active areas for reactions to occur on the catalyst's surfaces, a nanocatalyst with a better surface/volume ratio performs better. The transesterification reaction benefits from the high specific surface area of nanocatalysts. A popular nanocatalyst for the synthesis of biodiesel is calciumoxide because of its low cost and potent catalytic activity. CaO can therefore be used to make biodiesel either by itself or in combination with other substances. Other nanocatalysts used in the transesterification process include MgO, SrO, and ZnO. Once more, Fe3O4, CuFe2O4, activated carbon, and other substances can be utilized to improve the properties of these catalysts [[58], [59], [60], [61], [62], [63], [64], [65]].
Fig. 3.
Transesterification of triglycerides.
Transesterification involves utilization of catalysts for conversion of triglycerides to alkyl esters. Based on the phase and type of catalysts used, the process can be divided into.
1Homogeneous Catalysis
2Heterogeneous Catalysis
3.1. Homogeneous catalysis
In homogeneous catalysis, a chemical which is in the same phase as the reaction system is used to catalyze a set of reactions. It is the most favored catalyst to be used in the synthesis of biodiesel because it is more convenient and takes short time to execute the reaction [66]. Homogenous catalysis can involve the use of either acid or base catalysts, with base catalysts being preferred over the former. Catalysts most preferred include alkali metal hydroxides and alkoxides, which operate at low temperatures to provide maximum yields [67].
3.1.1. Base catalyst
Alkaline liquids such as alkali metal oxides including methoxides, sodium or potassium hydroxide, and carbonates are examples of homogeneous base catalysts. Base catalysts have a strong transesterification activity [68], however a notable downside is that oil with considerable FFA levels cannot be entirely converted into biodiesel and instead stays as soap in large volumes. The process may still be accelerated with an alkali catalyst up to 5% FFAs, but more catalyst is needed to make up for the catalyst that is lost to soap [69].
3.1.2. Acid catalyst
In case of acid catalysts, Brønsted acids such as sulfonic and sulfuric acids, hydrochloric acid, catalyze the esterification reaction [70]. Alkyl esters are produced in exceedingly high yields using these catalysts. However, because of the slower reactions compared to alkali-catalyzed reactions and the higher energy requirements, the procedure is economically burdensome [71]. Acid catalysis is capable of catalyzing both transesterification and esterification processes and is not influenced by the presence of FFA. The separation problems with homogeneous acid catalysis are identical to those with homogeneous base catalysis, notwithstanding these additional benefits.
3.2. Heterogeneous catalysis
In heterogeneous catalysis, the phases or states of catalysts differ from the reactants. These are the kinds of catalysts which frequently form active site during a reaction with their reactants [72]. It is remarkably effective and may be readily recycled as per the nature of catalyst. Based on the type utilized, heterogeneous catalysis is further classified into alkali-based, acid-based and enzyme-based catalyzed processes [73].
3.2.1. Alkali/base catalyst
The use of heterogeneous basic catalyst seeks to get around limitations like saponification, which prevents extraction of glycerol from methyl ester phase when a homogeneous base catalyst is used. Under moderate circumstances, these catalysts also exhibit remarkable catalytic activity [74]. Other benefits of these catalysts include their non-corrosiveness, environmental friendliness, and ease of disposal. Additionally, they may be made to have better activity, greater selectivity, and a longer catalyst lifespan and are simple to remove from the reaction environment. Recent research demonstrated the successful use of heterogeneous based-catalysts made of various oxides, zeolites, and clay [75,76]. Commercialization of such catalysts is a significant challenge due to a series of major drawbacks, which include low FFA levels, limitations in diffusion, large molar alcohol to oil proportion, high cost when in comparison to conventional reactions, and catalyst syphoning stills [77]. Heterogeneous base catalysis employs a diverse range of metal-based oxides, such as alkali, alkaline earth, and transition metal oxides. Boron group-derived and waste-derived catalysts are other types of catalysts used in this type of catalysis [78].
3.2.2. Acid catalyst
Heterogeneous acid catalysis is frequently performed using inorganic polymeric materials, which allow for simple separation from the reaction media, reusability, and the promotion of environmentally friendly technologies [79]. Compared to homogeneous acid catalysts, heterogeneous ones are less poisonous and corrosive and cause less environmental issues [80]. Numerous acid sites with varying degrees of Brønsted and Lewis acidity may be found in these catalysts. Application of these catalysts is constrained at some key points because nano porous inorganic oxides frequently exhibit low catalytic activity. These points include high solubility, low thermal stability, low surface area, obstacles in fabricating chemically antagonistic features, and low amounts of reactive sites [81].
3.2.3. Enzyme-based catalyst
Enzyme-based catalysts, also called biocatalysts, are derived from living creatures which facilitate chemical processes keeping their own chemical makeup stable [82]. Extracellular and intracellular lipases are the two main types of enzyme-based biocatalysts that are often utilized in the manufacture of biodiesel. The enzymes extracted from microbial broth and then purified are known as extracellular lipases. In contrast, intracellular lipase stays either inside cell or in cellular walls that produce new cells. Bottlenecks such as generation of unwanted byproducts, such as soaps and pigments, which obstruct purification from glycerol and di- and monoacylglycerols, can be minimized by using biocatalysts. The source of the enzyme and the process parameters affect how well the biocatalyzed transesterification process works. Contrary to chemical catalysts, a large spectrum of triglyceride sources can use biocatalysts, with FFA in the range of 0.5–80% [80]. Easy product removal, a process temperature ranging from 35 to 45 °C, absence of any byproducts, and the possibility to reuse catalysts are additional benefits of enzymatic biodiesel generation [83]. The complexity of separation and purification processes and their accompanying expense are limitation of utilizing extracellular enzymes as a catalyst. This can be reduced by utilizing microbial cells as whole-cell catalysts, resulting in satisfactory biodiesel production [84].
4. Nanomaterials in transesterification
Nanomaterials has been in spotlight in recent times due to various advantageous morphological and physico-chemical properties. Nanomaterials are particles which are approximately 1–100 nm (10−9 – 10−8 m) in diameter and are much smaller than particle size of bulk materials which are usually in ranges of microns (10−6 m). Due to smaller size, these particles have various characteristic properties such as a high surface area-volume ratio, enhanced reactivity and photo-catalytic activity and low cytotoxicity [85]. This makes their utilization more advantageous over conventional materials. Usage of nanomaterials in transesterfication has lately received a lot of emphasis for biodiesel generation due to excellent catalytic performance [86]. Compared to traditional catalysts, nanocatalysts have greater surface area, which improves their efficiency. Additionally, these catalysts have excellent saponification resistance, a high surface area-to-volume ratio, greater stability, and significant reusability [87]. Another key factor influencing the reaction conditions is designing nanocatalysts to be more stable at high temperatures, since this enhances the nanocatalyst's likelihood of being recyclable and reusable. As a result, many kinds of catalysts made of both organic and inorganic materials, such as ion exchange resin and enzymes, as well as oxides, and metal salts, are being produced. The creation of nanocatalysts made of polymers, zeolite, and carbon-based nanomaterials has recently attracted consideration from many researchers [88]. These catalysts primarily display appealing qualities of significant catalytic activity, shortening reaction time and lowering reaction temperature. Hence, production of biofuel will be significantly increased by the modification and manufacture of nanocatalysts. Numerous techniques, including chemical vapor deposition, impregnation, precipitation, and electrochemical deposition can be used to create the newly formed nanocatalyst. Some noteworthy nanocatalysts that have been used in transesterification producing biodiesel include metal-oxide nanocatalysts such as CaO, ZnO, zeolites, hydrotalcites and nanobiocatalysts [89]. Some types of nanocatalysts can be seen in Fig. 4.
Fig. 4.
Various types of nanocatalysts.
4.1. Metal-oxide nanocatalysts
Out of all types, metal-oxide based nanocatalysts are the most preferred and accounts for a pivotal role in maximization of biodiesel production. Oxides of several metals such as Fe [90], Mg, Zn, Ca, etc. have been used for making of nanoparticles to be used for catalysis in transesterification.
Kumar et al. reported four types of calcium oxide (CaO) nanoparticles, namely, doped CaO, waste CaO, pure CaO and loaded CaO, which are used in transesterification of different kinds of feedstock oils [91]. Amalgamation of CaO nanoparticles via sol-gel method was done by Bharti et al. which was then used for catalysis in transesterification of feedstock oil extracted from soyabeans [92]. After optimization and synthesis in optimal conditions, the process gave a 97.61% biodiesel yield. The synthesized catalyst had a higher surface area/volume ratio than conventional CaO catalysts and had a reusability of up to 6 cycles with minimal drop in conversion efficiency. High concentrations of fatty acids can be readily converted by the CaO nanocatalyst, however long-term usage is not recommended owing to leaching and active site blocking. Doping with metal and metal oxides has the potential to curb the limitation. Higher yields of biodiesel were obtained by synthesizing CaO nanoparticles from discarded chicken eggshells and then using wet impregnation method to dope them with Cu and Zn [93]. On comparison, Zn/Cao was the most efficient with reusability of up to 7 cycles and a net conversion rate of 80%. Efficiency of Zn-doped CaO nanoparticles has also been studied by Ref. [20], which had a biodiesel production efficiency of 96.74% and could be reused for up to 5 cycles. Additionally, KF/CaO nanocatalysts with size ranges between 30 and 100 nm were manufactured by impregnation technique where CaO was doped [94]. Their efficacy in the synthesis of biodiesel using Chinese tallow seed as feedstock was assessed. According to this study, 96.8% of biodiesel was produced, demonstrating the nanocatalyst's potential for the biodiesel sector. CaO, as a nanocatalyst, has been broadly explored and analyzed for catalyzing the transesterification process. However, this catalyst has its own bottlenecks. Leaching which is mostly brought on by the catalyst being hydrolyzed, has a detrimental effect on the conversion efficiency. This leads to saponification, lowering the catalyst's overall catalytic activity and specificity. Comparative study of different metal-oxide catalyst can be seen in Table 4.
Table 4.
Biodiesel production using different metal-oxide catalyst.
The nanocatalyst derived from MgO/MgAl2O4 was described by Rahmani Vahid & Haghighi [103]. It was produced primarily using the combustion technique, succeeded by impregnation. The ideal fuel ratio was found to be 1.5, with 60.6 m2/g specific surface area and having an average pore diameter size of 6.3 nm. Further testing of the improved catalyst for generation of biodiesel revealed an efficiency of conversion of 95.7% and a reusability of 3–5 cycles. At 60 °C and methanol to oil ratio of 12:1, the creation of Li doped MgO nanocatalyst has been reported to have a yield of 93.9% [104]. Due to the addition of Li, these catalysts provide more basic site creation, which enhance the net production efficiency. Furthermore, using CaO and MgO nanoparticles studied the synthesis of biodiesel using cooking oils as the feedstock [105]. When compared to MgO nanocatalysts, CaO nanoparticles significantly increased the biodiesel yield.
Production of fatty acid methyl ester (FAME) was achieved by transesterifying palm oil involving catalysis by TiO2–ZnO nanocatalyst. Ti replaces Zn in the lattice structure, creating deformities and thereby making the catalyst more stable and increasing its catalytic activity [106]. Sulfate-doped TiO2 catalyst was reported to be efficient in refining petroleum and in production of biodiesel with cooking oil used as feedstock. Hematite was used to provide SO4/Fe–Al–TiO2 catalyst its magnetic properties, while alumina served as a buffer layer and the catalyst, on transesterifying cooking oil, had a 96% conversion efficiency [107]. Doping of potassium taratarate on TiO2 nanocatalyst increased the basicity of the catalyst, leading to a net production of 98.9% on transesterification of linseed oil [10]. Alves et al. showed a comparison analysis of Fe/Cd along with Fe/Sn oxide nanoparticles having magnetic properties to transesterify soybean oil [108]. Fe/Sn oxide nanoparticles outperformed with an 84% yield of biodiesel, exhibiting the highest catalytic activity out of the duo. In addition to having remarkable resistance at high temperatures, these nanoparticles further promote effective reaction between free fatty acids and alcohol (specifically methanol), increasing output. The nanocatalysts and related reports discussed here provide sufficient evidence to derive the resolution that nanocatalysts curated using multiple metal oxides are among the most effective heterogeneous catalysts for producing biodiesel using a variety of feedstocks.
4.2. Carbon nanocatalysts
In accordance with various structures and sizes of the generated nanocomposite materials, nanocatalysts made from carbon materials such as carbon nanotubes, graphene as well as reduced graphene oxides have demonstrated superior physico-chemical characteristics [109].
Carbon nanotubes (CNTs) are reported to be an excellent catalysts, and beneficial applications in fuel cells and other electrocatalytical devices, due to various structural, mechanical, thermal, and biocompatibility properties. Generally, CNTs are made from graphite sheets wrapped into cylindrical shapes, measuring in nanometric dimentions, high surface area, and are exceptionally biocompatible in nature. Enzymes such as, lipases immobilized or conjugated with CNTs (oe functionalized multi-walled carbon nanotubes - MWNTs) were shown to have better activity and stability. The ester hydrolysis of MWNT-immobilized lipase and free lipase in an aqueous media was studied over time, and increased amounts of aldehydes, carboxylates, and peroxides were reported as the end products [110]. Sulphonated multiwalled CNT was used as a nanocatalyst in Guan et al. research to trans-esterify triglycerides [111]. For the esterification of fatty acid derived from palm, Shuit et al. reported employing sulfonated multiwalled carbon nanotubes (s-MWCNT) [112]. Loading of ZnO nanoparticles on microwalled carbon nanotubes (MWCNT) led to decrement of surface area and increment of active sites, enhancing rate of reaction and mass transfer [113].
For the synthesis of biodiesel, some researchers have further employed lipase enzymes which have been immobilized rather than carbon nanotubes because of their improved thermal strength and recyclable nature. Fan et al., described the fabrication of multiwalled CNTs that were loaded with Fe-oxide, coupled with polyamidoamine dendrimers, and then anchored over Burkholderia cepacia lipase (BCL) [114]. Due to the surface-linked dendrimers, the designed catalyst demonstrated increased catalytic activity, producing an overall yield of 92.8% and outstanding recyclability to an approximate of 20 subsequent cycles. Nanocatalysts made of biochar are very effective for catalyzing transesterification and is primarily made via pyrolysis and carbonization of biomass derived from various feedstock. Using the carbonization approach, more carbon-based nanocatalysts can easily generated from fly ash and sugar-based nanomaterial [115]. However, biochar based nanocatalysts have low thermal stability and have lengthened time of reaction. Graphene oxide also shown improved fatty acid conversion due to more surface area and unique functional groups. TiO2/reduced graphene oxide nanocomposites were created by Verma et al. via transesterification of used cooking oil. Under optimal conditions, the catalyst had a conversion efficiency of 98% and reusability of up to 3 cycles [116].
4.3. Magnetic nanocatalysts
Magnetic nanocatalysts may be produced from a variety of materials, such as metal oxides, iron, cobalt, nickel, and platinum alloys, which are used in a variety of ways [117]. Magnetic nanoparticles (MNPs) have several favorable characteristics: higher molecular weight and smaller size (thus, a good ratio of surface area to volume); quantum characteristics; and capacity to transport other chemicals (such as drugs). MNPs also have a variety of possible applications in biofuel industries, like the synthesis of sugars using lignocellulosic stuff by using immobilized enzymes [118]. MNPs can be utilized as enzyme immobilization support, or used to attach other catalytic nanoparticles, making them ideal nanoparticle for use in several applications. One maor advantage of using ‘MNPs-enzymes’ is the possibility of collecting the enzyme complex using magnetic field, and resuisng it more than once. By using a high-frequency magnetic field, these nanocatalysts may also be utilized for hydrogenation, photooxidation, and inductive heating. Magnetic nanoparticles also act as carriers for enzymes, which will be more manageable in a magnetic field range. Rai and colleagues utilized magnetic nanoparticles to immobilize the lipase enzyme, which they then used to make biodiesel using palm oil [110]. The biocatalyst was quite stable and catalytically active, and it was possible to detach immobilized enzyme by using a magnetic field for five consecutive batches, while retaining >80% activity.
4.4. Zeolite nanocatalysts
Zeolite materials are being used more often now than ever before. Strong acidic sites, a large surface area, specificity, and distinctive molecular sieving features are recognized for these compounds' potential catalytic activity [119]. Nanozeolites are hydrophobic in nature with large external surface areas, and their great solubility in both polar and non-polar solvents facilitates improved enzyme access to the substrate [120]. Synthetic-based materials are far more widely employed in commercial sectors than natural zeolite materials. ZSM-5, X, Y, and beta zeolites are commercially accessible synthetic zeolites that are primarily employed in the manufacturing of biodiesel [121].
Using NaX zeolite loaded with KOH, Xie et al. demonstrated the transesterification of soybean oil [122]. To gain the highest yield of production, this zeolite was synthesized and examined using various compositions, and the net biodiesel yield was 85.6%. Sunflower oil was trans-esterified using CaO nanoparticles and NaX zeolite under optimal circumstances, producing 93.5% biodiesel [123]. The utilization of NaY as well as KOH/NaY zeolites as catalysts in the transesterification of lipid extracted from Madhuca indica has been reported by another group [124]. These zeolites have a conversion rate of 98.9% and may be recycled and used repeatedly. CaO/NaY–Fe3O4 is a nanomagnetic-based zeolite catalyst that was reported to exhibit outstanding stability at high temperatures and easy separation method. With canola oil used as feedstock material, upon analysis of various metrics, the zeolites showed a conversion efficiency of 95.37% [125].
ZSM-based nanocatalysts are being synthesized and used for catalyzing transesterification. ZSM-5, also known as Zbio was modified to form mesoporous zeolites. Due to the availability of acid active sites, ZBio has more activation energy. Biodiesel production of 87% was attained using these zeolites and was greatly influenced by the higher energy of activation [126]. Thus, the above discussions may be used to draw the conclusion that functionalizing zeolites with nanocatalysts will open new possibilities for scaling-up biodiesel production by transesterification of numerous feedstock materials [127].
4.5. Nanobiocatalysts
Biocatalysts used in transesterification conventionally use intracellular and extracellular enzymes such as lipase to enhance biodiesel output. However, these catalysts are laden with bottlenecks such as complex methods of purification, low reusability, and relatively unstable nature of free enzymes. To combat such issues, immobilizing enzymes on solid substructures is considered a novel approach for better biodiesel yields. Nanoparticles can be effective support for enzyme immobilization due to increased Brownian motion and less stearic and conformational constraints. Nanobiocatalysts employed in transesterification constitute two main components: nanocatalyst part containing nanoparticles to be used as support; and biological part comprising of the immobilized enzymes [128]. Based on the nature of nanocatalyst, nanobiocatalysts can be sub-divided into inorganic nanobiocatalysts, magnetic nanobiocatalysts, carbon scaffold-based nanobiocatalysts and mesoporous nanobiocatalysts.
Due to larger surface area, existence of more active sites, and superparamagnetic characteristics, magnetic nanomaterials (MNP) are one of the most often employed materials as a nanocatalytic component of nanobiocatalysts (NBCs). Ferrite colloids such as maghemite (γ-Fe2O3) and magnetite (Fe2O3) are the most often employed magnetic nanomaterials [90,129]. To facilitate the attachment of enzymes for immobilization, these nanomaterials' surfaces are typically further modified with certain functional groups, such as amides, alcohols, etc. Using the solvothermal technique, Touqeer et al. developed Fe3O4 NPs which were covalently immobilized with lipase extracted from Aspergillus terreus AH-F2 [1]. The NBCs were then utilized to catalyze transesterification of wasted cooking oil to make biodiesel.
One of the most often used inorganic nanocarriers for the fabrication of the NBCs is composed of graphene oxide, carbon nanotubes, silica (SiO2), and alumina (Al2O3), among others. The availability of different functional moieties such as COO- groups, OH groups, and epoxy groups allows for effective lipase immobilization without the need of unusual cross-linking. Graphene oxide was used by Li et al. for immobilization of lipase and the catalyst had a hydrolytic effectiveness of 80%, validating graphene oxide as a useful nanomaterial for NBCs [130]. One of the most recent advanced classes to appear in academic manuscripts as a nanoscaffolds for immobilizing enzymes needed to produce NBCs is metal organic frameworks. Possessing characteristics such as surface porosity, optimum surface area, these have been proven to be amazingly effective in immobilizing a variety of enzymes for use in transesterification [131]. Some of the recently published work on using different types of nanomaterials for biodiesel synthesis, different types of feedstock, experimental conditions, and biodiesel yields are highlighted in Table 5.
Table 5.
Different types of nanomaterials used for biodiesel synthesis.
| No. | Type of feedstock | Type of nanomaterials | Experimental conditions | Biodiesel Yield (%) | References |
|---|---|---|---|---|---|
| 1 | Wild mustard seed oil | Lanthanum titanium dioxide (LaTiO3) nanoparticles | MeOH/oil ratio (4:1), catalyst (100 mg), reaction time (60 min), 80 °C | 92.21 | [132] |
| 2 | Canola oil | αFe2O3 1-x/ZnO x nanoparticles | MeOH/oil ratio (11.25:1), ultrasonic power of 278.46 W, irradiation power time 29.22 min, catalyst (47.24%), 65 °C | 94.21 | [133] |
| 3 | Waste cooking oil (WCO) | TiO2 nano-catalyst | MeOH/oil ratio (10:1), 0.01% TiO2 nano-catalyst coupled with 0.3% NaOH, reaction time (60 min), 60 °C | 95 | [134] |
| 4 | Mango seed (Mangifera indica) extracts | ZnO nanocatalyst | MeOH/oil ratio (7:2), 10 mL of ZnO nano catalyst, reaction time (60 min), 50 °C | 85 | [135] |
| 5 | Waste cooking oil (WCO) | Candida antarctica Lipase B immobilized on a magnetic hybrid sol-gel nanocomposite | MeOH/oil ratio (4:1), catalyst (1 g), reaction time (30 h), 40 °C | 96 | [136] |
| 6 | Citrus medicaseed oil | Green CuO nanoparticles | MeOH/oil ratio (8:1), catalyst 0.18%, reaction time (120 min), 85 °C | 93 | [137] |
| 7 | Canola oil | ZnO nanocatalyst | MeOH/oil ratio (11.19:1), ultrasound irradiation time (31.98 min), nanocatalyst (3.17%) | 90.16 | [138] |
| 8 | Waste palm cooking oil (WPCO) | Calcium oxide with a strontium ion additive (Sr–CaO) | MeOH/oil ratio (9:1), 5%w/w catalyst, reaction time (3 h), 80 °C | 99.33 | [139] |
| 9 | Waste cooking oil (WCO) | Heterogeneous nanocatalyst-based magnetic polyaniline (MPANI) decorated with cobalt oxide | MeOH/oil ratio (10:1), MPANI to Co3O4 ratio of 1:1, 0.2 g catalyst, reaction time (240 min), 90 °C | 93 | [140] |
| 10 | Sunflower oil | ZnO nanoparticle | MeOH/oil ratio (20:1), catalyst (4.7%), reaction time (3 h), 70 °C | 75 | [141] |
| 11 | Olive oil | Magnetite nanoparticles-immobilized lipase fromBacillus subtilis | MeOH/oil ratio (12:1), catalyst (1% w/v), reaction time (1 h), 37 °C | 45 | [142] |
| 12 | Oleaginous fungal lipid from Fusarium solani (FS12) | Bimetallic gold-silver core-shell nanoparticle (Au@Ag NPs) | MeOH/oil ratio (20:1), 3% Au@Ag NPs, reaction time (30 min), 70 °C | 91.28 ± 0.19 | [143] |
| 13 | Zanthoxylum armatumseed oil | Silver oxide nanoparticles | MeOH/oil ratio (7:1), catalyst 0.5%, reaction time (2 h), 90 °C | 95 | [144] |
| 14 | Waste cooking oil (WCO) | CaO nanocatalyst from Acalypha indica | MeOH/oil ratio (11.8:1), catalyst (2.4%), reaction time (70 min), 63.7 °C | 94.74 | [145] |
| 15 | Terminalia chebulaoil | CuO nanoparticles (from T. chebula leaves | MeOH/oil ratio (9:1), catalyst (3%), reaction time (60 min), 60 °C, stirring speed of (650 rpm) | 97.1 | [146] |
| 16 | Palm oil | Magnetic mesoporous solid acid catalysts (Fe3O4@SBA-15-NH2-HPW) | MeOH/oil ratio (20:1), catalyst 4%, reaction time (5 h), 150 °C | 91 | [147] |
| 17 | Lipids from Chlorella pyrenoidosa | Aspergillus niger (KP001169) lipase enzyme immobilized into Fe3O4nanoparticles | MeOH/oil ratio (4:1), immobilized lipase (300 mg), reaction time (6 h), 45 °C, stirring (400 rpm) | 90 | [148] |
| 18 | Waste cooking oil (WCO) | Lanthanum phosphate foam heterogeneous nanocatalyst (LaPO4 foam) | MeOH/oil ratio (5:1), LaPO4 (2.5%), reaction time (120 min), 90 °C | 91 | [149] |
| 19 | Non-edible seed oil ofChamaerops humilis | CoO nanoparticles | MeOH/oil ratio (6:1), catalyst (0.54%), reaction time (180 min), 60 °C | 92 | [150] |
| 20 | Microalgal oil (Chlorella vulgaris) | CaO nanoparticles | MeOH/oil ratio (9:0.6), catalyst (4%), reaction time (4 h), 80 °C | 67 | [151] |
| 21 | Waste seed oil ofCitrus aurantium | Zirconium oxide | MeOH/oil ratio (6:1), catalyst (0.5%), reaction time (120 min), 87.5 °C | 94 | [152] |
| 22 | Jatropha curcas seed oil (JCSO) | Lipase- polydopamine (PDA)-TiO2 nanoparticles | MeOH/oil ratio (6:1), NPs concentration (10%), reaction time (30 h), 37 °C, water percentage (0.5%) | 92 | [153] |
| 23 | Waste cooking oil (WCO) | Graphene oxide and bimetal zirconium/strontium oxide nanoparticles (GO@ZrO2–SrO) | MeOH/oil ratio (4:1), material ratio of (1:0.5 w/w) of GO: ZrO2–SrO, reaction time (90 min), 120 °C | 91 | [154] |
| 24 | Waste cooking oil (WCO) | Tungstophosphoric acid nanoparticles | MeOH/oil ratio (12:1), catalyst 3% (w/w of oil), reaction time (5 h), 65 °C | 91 | [155] |
| 25 | Waste cooking oil (WCO) | Magnetic nano-sized solid catalyst derived from bio-wasteCitrus sinensis peel ash (CSPA)@Fe3O4 | MeOH/oil ratio (6:1), catalyst (6%), reaction time (3 h), 65 °C | 98 | [156] |
| 26 | Pongamia pinnata raw oil | Fe3O4nanoparticles impregnated eggshell (CES–Fe3O4) | MeOH/oil ratio (12:1), catalyst (2%), reaction time (2 h), 65 °C | 98 | [157] |
| 27 | Blend of neem and rubber seed oils (NO-RSO) | Nanoparticles from elephant-ear tree (E. cyclocarpum) pod husk | MeOH/oil ratio (11.4:1), catalyst (2.96%), reaction time (5.88 min), microwave heating power (150 W) | 98.77 ± 0.16 | [158] |
| 28 | Microalgae oil | Fe3O4/ZnMg(Al)O solid base catalyst | MeOH/oil ratio (12:1), catalyst (3%), reaction time (3 h), 65 °C | 94 | [159] |
| 29 | Waste cooking oil (WCO) | Magnetic SO4/Fe–Al–TiO2 solid acid catalyst | MeOH/oil ratio (10:1), magnetic catalyst (3%), reaction time (2.5 h), 90 °C | 96 | [107] |
| 30 | Waste scum oil | TiO2 nanoparticles | MeOH/oil ratio (9:1), catalyst (0.9%), reaction time (90 min), 60 °C, stirring speed (300 rpm) | 93.5 | [160] |
| 31 | Rapeseed oil | Potassium impregnated Fe3O4–CeO2 nanocatalyst | MeOH/oil ratio (7:1), catalyst (4.5%), reaction time (120 min), 65 °C | 96.13 | [161] |
5. Reusability and recovery of nanoparticles
The most important benefits of utilizing heterogeneous nanocatalysts in the biodiesel manufacturing process are their reusability and recovery. The nanocatalyst is used in several cycles during these processes to create biodiesel, and it is recovered and reused at every step. Chemical methods are commonly used to recover nanocatalysts. Heterogeneous catalysts make it simple and rapid to recover both the desired end product and byproduct from the reaction mixture. With this form of catalyst, a washing step is not necessary. Several advantages, such as quick and simple separation from the reaction mixture and quicker mixing of the reactants and catalyst, were suggested for the esterification process utilizing nanocatalysts.
6. Techno-economic analysis
As a medium-term substitute for petroleum-based diesel fuel, biodiesel—which is defined as the monoalkyl esters of long chain fatty acids—has drawn significant attention from all over the world. Oils can be transesterified with short-chain alcohols like methanol to create biodiesel, or free fatty acids can be esterified to create biodiesel (FFA). Glycerol is the lone consequence of the transesterification reaction, which is occurring in the presence of an appropriate catalyst like an alkali or an acid. The key benefits of biodiesel are that it is generated locally from renewable resources and emits less carbon monoxide, particulates, and sulfur dioxide. Besides from being biodegradable, non-toxic, and safer to handle, it also produces 78% less carbon dioxide. The primary drawbacks are its high cost, increased nitrogen oxide emissions, and worries about how it would affect the longevity of diesel engines when utilized in their purest form.
7. Challenges and prospects
Excellent surface properties, small size, biocompatibility (for most types), ease of synthesise, proven applications in a wide variety of fields, are some of the highly useful traits of nanomaterials. However, thers is a flipside of using such nanonaterials – nanotoxicology, which is under the lens, with the possible effects on flora and fauna. The main attributes of nanomaterials, which makes them exrtremyly useful are also a possible casue of concern that can directly or indirectly trigger numerous cytotoxic impacts in humans. In addition, the majority of metal-oxide nanoparticles are now known to be extremely harmful to specific organs, if entered in the human body. The studies have highlighted hazardous effects of nanoparticles, using Wistar rodents medels, where it showed fanatical changes in the liver, extended platelet count, and delicate disturbance. It has likewise been discovered that titanium was bio-aggregated in the liver, lungs, and mind of uncovered rodents, causing neuro-behavioral changes in the creatures and, additionally, organ wounds. However, the negative effects of extensicve usage or release of nanoparticles in the environment is not yet completely studied, which is still a matter of concern. Therefor, there is a widespread concern about not just the advantages of these incredibly helpful particles yet in addition their harmful effects, especially when applied in large scale applications, such as, biofuel-biodiesel prodcution. These contradicting opinions and possibilities provides worldwide experts an opportunity to propose and study long term effects, and how to avert negative impacts of using nanomaterials, while benefiting the society.
In light of the review, future investigations can greatly outperform the current biodiesel production restrictions. Creating a strategy for the large-scale use of non-edible feedstock oils and investigating ways to boost the production of biomass derived from algae, which contains a significant amount of oil. Developing nanocatalysts with multifunctional active sites that can endure disintegration during the trans-esterification reaction while also minimizing the impact of leaching Nanomaterials and polymeric materials are used to functionalize carbonaceous and zeolite-based catalysts in order to increase their catalytic activity, specificity, and stability. Using structural modification based on the shape and size of nano-based materials to improve the kinetics of the reactions involved in the generation of biodiesel. developing statistical methods for efficient large-scale reaction condition optimization. Finally, technological Development is required to scale up the reaction process in order to reduce mass transfer resistance, high energy consumption, and by-product utilization. boosting the possibility of preserving the ideal reaction conditions and so raising the production of biodiesel.
8. Conclusions
Biodiesel production holds an enormous potential to cause a paradigm shift in global energy sources and can lead to a re-imagination of recycling of resources, products, and residues in light of sustainability. This can help to achieve eco-friendly routes of consumption and make the vision of a global circular bioeconomy a reality. However, scale-up of biodiesel production is laden with bottlenecks such as low output, dilemma of food vs feed and lack of enough statistical know-how to fully exploit the potential of biofuels. Transesterification is the popularly preferred route for production of biodiesel due to its economic viability and sustainability, coupled with minimal drawbacks and a high output. But, the biggest limitation is transesterification is a slow proceeding process and the conventional catalysts used to fasten the process often pose newer set of challenges.
Usage of nanoparticles as catalysts present a new advent in biodiesel production that can curb the existent bottlenecks. In this review, the various nanocatalysts and their role in enhancing biodiesel output via transesterification have been explored and discussed in detail. The unique characteristics of the nanocatalysts include a high surface area-to-volume ratio, varied shape, enhanced activity, and specificity of the catalyst. Numerous functional groups can be used to significantly alter nanocatalysts, increasing their overall acidic character, reducing leaching of active metals, and increasing number of active sites for the catalytic process. The quality and output of biodiesel can be enhanced by using nanocatalyst in the right ratios to increase various reaction parameters.
As per the review, future studies can focus on up-scaling of biodiesel production from waste-derived and biomass-derived feedstocks, using nanomaterials. Synthesis and functionalization of nanocatalysts having minimal leaching and disintegration and possessing various active sites is another area of consideration and research. Finally, technical Development is required to scale up the reaction process in order to reduce mass transfer resistance, high energy consumption, and by-product utilization. With further research, advanced technological methods need to be devised to fully understand and avail the effectiveness of biodiesel in meeting global energy requirements in a renewable and environmentally sound manner.
Ethical approval
Not applicable.
Author contribution statement
All authors listed have significantly contributed to the development and the writing of this article.
Data availability statement
Data included in article/supp. material/referenced in article.
Funding
Sanket Joshi would like to kindly acknowledge research fund from Sultan Qaboos University, under HMTF scheme (SR/DVC/OGRC/21/01), to perform such study.
Declaration of competing interest
The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.
Acknowledgements
Authors kindly acknowledge their respective Universities for providing kind support.
References
- 1.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 nano biocatalyst for the production of biodiesel using waste cooking oil as feedstock: characterization and process optimization. Energies. 2020;13:177. doi: 10.3390/en13010177. [DOI] [Google Scholar]
- 2.Martchamadol J., Kumar S. Thailand's energy security indicators. Renew. Sustain. Energy Rev. 2012;16:6103–6122. doi: 10.1016/j.rser.2012.06.021. [DOI] [Google Scholar]
- 3.Arbab M.I., Varman M., Masjuki H.H., Kalam M.A., Imtenan S., Sajjad H., Rizwanul Fattah I.M. Evaluation of combustion, performance, and emissions of optimum palm–coconut blend in turbocharged and non-turbocharged conditions of a diesel engine. Energy Convers. Manag. 2015;90:111–120. doi: 10.1016/j.enconman.2014.11.017. [DOI] [Google Scholar]
- 4.Sulaiman S., Zik N.A.F.A., Jamal P. Thermally produced nano catalyst for biodiesel production: a review. Journal of Advanced Research in Fluid Mechanics and Thermal Sciences. 2018;52:139–147. [Google Scholar]
- 5.Winoto V., Yoswathana N. Optimization of biodiesel production using nanomagnetic CaO-based catalysts with subcritical methanol transesterification of rubber seed oil. Energies. 2019;12:230. doi: 10.3390/en12020230. [DOI] [Google Scholar]
- 6.Demirbas A. Biofuels sources, biofuel policy, biofuel economy and global biofuel projections. Energy Convers. Manag. 2008;49:2106–2116. doi: 10.1016/j.enconman.2008.02.020. [DOI] [Google Scholar]
- 7.Prokopowicz A., Zaciera M., Sobczak A., Bielaczyc P., Woodburn J. The effects of neat biodiesel and biodiesel and HVO blends in diesel fuel on exhaust emissions from a light duty vehicle with a diesel engine. Environ. Sci. Technol. 2015;49:7473–7482. doi: 10.1021/acs.est.5b00648. [DOI] [PubMed] [Google Scholar]
- 8.Hoekman S.K., Robbins C. Review of the effects of biodiesel on NOx emissions. Fuel Process. Technol. 2012;96:237–249. doi: 10.1016/j.fuproc.2011.12.036. [DOI] [Google Scholar]
- 9.Perez V.H., Silveira Junior E.G., Cubides D.C., David G.F., Justo O.R., Castro M.P.P., Sthel M.S., de Castro H.F. In: Biofuels in Brazil: Fundamental Aspects, Recent Developments, and Future Perspectives. da Silva S.S., Chandel A.K., editors. Springer International Publishing; Cham: 2014. Trends in biodiesel production: present status and future directions; pp. 281–302. [DOI] [Google Scholar]
- 10.Ambat I., Srivastava V., Sillanpää M. Recent advancement in biodiesel production methodologies using various feedstock: a review. Renew. Sustain. Energy Rev. 2018;90:356–369. doi: 10.1016/j.rser.2018.03.069. [DOI] [Google Scholar]
- 11.Borah M.J., Devi A., Borah R., Deka D. Synthesis and application of Co doped ZnO as heterogeneous nanocatalyst for biodiesel production from non-edible oil. Renew. Energy. 2019;133:512–519. [Google Scholar]
- 12.Ashok A., Kennedy L.J., Vijaya J.J., Aruldoss U. Optimization of biodiesel production from waste cooking oil by magnesium oxide nanocatalyst synthesized using coprecipitation method. Clean Technol. Environ. Policy. 2018;20:1219–1231. doi: 10.1007/s10098-018-1547-x. [DOI] [Google Scholar]
- 13.Ashok A., Kennedy L.J. Magnetically separable zinc ferrite nanocatalyst for an effective biodiesel production from waste cooking oil. Catal. Lett. 2019;149:3525–3542. doi: 10.1007/s10562-019-02906-4. [DOI] [Google Scholar]
- 14.Boro J., Deka D., Konwar L.J. Transesterification of non edible feedstock with lithium incorporated egg shell derived CaO for biodiesel production. Fuel Process. Technol. 2014 doi: 10.1016/j.fuproc.2014.01.022. (accessed March 17, 2023) [DOI] [Google Scholar]
- 15.Mohebbi S., Rostamizadeh M., Kahforoushan D. Effect of molybdenum promoter on performance of high silica MoO3/B-ZSM-5 nanocatalyst in biodiesel production. Fuel. 2020;266 doi: 10.1016/j.fuel.2020.117063. [DOI] [Google Scholar]
- 16.Rasouli H., Esmaeili H. Characterization of MgO nanocatalyst to produce biodiesel from goat fat using transesterification process. 3 Biotech. 2019;9:429. doi: 10.1007/s13205-019-1963-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Dehghani S., Haghighi M. Sono-enhanced dispersion of CaO over Zr-Doped MCM-41 bifunctional nanocatalyst with various Si/Zr ratios for conversion of waste cooking oil to biodiesel. Renew. Energy. 2020;153:801–812. doi: 10.1016/j.renene.2020.02.023. [DOI] [Google Scholar]
- 18.Ambat I., Srivastava V., Haapaniemi E., Sillanpää M. Nano-magnetic potassium impregnated ceria as catalyst for the biodiesel production. Renew. Energy. 2019;139:1428–1436. [Google Scholar]
- 19.Ruhul A.M., Kalam M.A., Masjuki H.H., Fattah I.M.R., Reham S.S., Rashed M.M. State of the art of biodiesel production processes: a review of the heterogeneous catalyst. RSC Adv. 2015;5:101023–101044. doi: 10.1039/C5RA09862A. [DOI] [Google Scholar]
- 20.Borah M.J., Das A., Das V., Bhuyan N., Deka D. Transesterification of waste cooking oil for biodiesel production catalyzed by Zn substituted waste egg shell derived CaO nanocatalyst. Fuel. 2019;242:345–354. doi: 10.1016/j.fuel.2019.01.060. [DOI] [Google Scholar]
- 21.Ghosh N., Halder G. Current progress and perspective of heterogeneous nanocatalytic transesterification towards biodiesel production from edible and inedible feedstock: a review. Energy Convers. Manag. 2022;270 doi: 10.1016/j.enconman.2022.116292. [DOI] [Google Scholar]
- 22.Dhawane S.H., Kumar T., Halder G. Recent advancement and prospective of heterogeneous carbonaceous catalysts in chemical and enzymatic transformation of biodiesel. Energy Convers. Manag. 2018;167:176–202. doi: 10.1016/j.enconman.2018.04.073. [DOI] [Google Scholar]
- 23.Ibrahim M.L., Nik Abdul Khalil N.N.A., Islam A., Rashid U., Ibrahim S.F., Sinar Mashuri S.I., Taufiq-Yap Y.H. Preparation of Na2O supported CNTs nanocatalyst for efficient biodiesel production from waste-oil. Energy Convers. Manag. 2020;205 doi: 10.1016/j.enconman.2019.112445. [DOI] [Google Scholar]
- 24.Degfie T.A., Mamo T.T., Mekonnen Y.S. Optimized biodiesel production from waste cooking oil (WCO) using calcium oxide (CaO) nano-catalyst. Sci. Rep. 2019;9 doi: 10.1038/s41598-019-55403-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Ramsurn H., Gupta R.B. Nanotechnology in solar and biofuels. ACS Sustainable Chem. Eng. 2013;1:779–797. doi: 10.1021/sc400046y. [DOI] [Google Scholar]
- 26.Foo W.H., Koay S.S.N., Chia S.R., Chia W.Y., Tang D.Y.Y., Nomanbhay S., Chew K.W. Recent advances in the conversion of waste cooking oil into value-added products: a review. Fuel. 2022;324 doi: 10.1016/j.fuel.2022.124539. [DOI] [Google Scholar]
- 27.Aghbashlo M., Peng W., Tabatabaei M., Kalogirou S.A., Soltanian S., Hosseinzadeh-Bandbafha H., Mahian O., Lam S.S. Machine learning technology in biodiesel research: a review. Prog. Energy Combust. Sci. 2021;85 doi: 10.1016/j.pecs.2021.100904. [DOI] [Google Scholar]
- 28.Hosseinzadeh-Bandbafha H., Nizami A.-S., Kalogirou S.A., Gupta V.K., Park Y.-K., Fallahi A., Sulaiman A., Ranjbari M., Rahnama H., Aghbashlo M., Peng W., Tabatabaei M. Environmental life cycle assessment of biodiesel production from waste cooking oil: a systematic review. Renew. Sustain. Energy Rev. 2022;161 doi: 10.1016/j.rser.2022.112411. [DOI] [Google Scholar]
- 29.Khan O., Yadav A.K., Khan M.E., Parvez M. Characterization of bioethanol obtained from Eichhornia Crassipes plant; its emission and performance analysis on CI engine. Energy Sources, Part A Recovery, Util. Environ. Eff. 2021;43:1793–1803. doi: 10.1080/15567036.2019.1648600. [DOI] [Google Scholar]
- 30.Akia M., Yazdani F., Motaee E., Han D., Arandiyan H. A review on conversion of biomass to biofuel by nanocatalysts. Biofuel Research Journal. 2014;1:16–25. doi: 10.18331/BRJ2015.1.1.5. [DOI] [Google Scholar]
- 31.Najeeb J., Akram S., Mumtaz M.W., Danish M., Irfan A., Touqeer T., Rashid U., Ghani W.A.W.A.K., Choong T.S.Y. Nanobiocatalysts for biodiesel synthesis through transesterification—a review. Catalysts. 2021;11:171. doi: 10.3390/catal11020171. [DOI] [Google Scholar]
- 32.Nogales-Delgado S., Encinar J.M., González J.F. Safflower biodiesel: improvement of its oxidative stability by using BHA and TBHQ. Energies. 2019;12:1940. doi: 10.3390/en12101940. [DOI] [Google Scholar]
- 33.Ghani N.A.A., Channip A.-A., Chok Hwee Hwa P., Ja’afar F., Yasin H.M., Usman A. Physicochemical properties, antioxidant capacities, and metal contents of virgin coconut oil produced by wet and dry processes. Food Sci. Nutr. 2018;6:1298–1306. doi: 10.1002/fsn3.671. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Encinar J.M., Nogales S., González J.F. Biodiesel and biolubricant production from different vegetable oils through transesterification. Engineering Reports. 2020;2 doi: 10.1002/eng2.12190. [DOI] [Google Scholar]
- 35.Kassem Y., Çamur H., Alassi E. Biodiesel production from four residential waste frying oils: proposing blends for improving the physicochemical properties of methyl biodiesel. Energies. 2020;13:4111. doi: 10.3390/en13164111. [DOI] [Google Scholar]
- 36.Muhammad G., Alam M.A., Mofijur M., Jahirul M.I., Lv Y., Xiong W., Ong H.C., Xu J. Modern developmental aspects in the field of economical harvesting and biodiesel production from microalgae biomass. Renew. Sustain. Energy Rev. 2021;135 doi: 10.1016/j.rser.2020.110209. [DOI] [Google Scholar]
- 37.Sarwer A., Hussain M., Al-Muhtaseb A.H., Inayat A., Rafiq S., Khurram M.S., Ul-Haq N., Shah N.S., Alaud Din A., Ahmad I., Jamil F. Suitability of biofuels production on commercial scale from various feedstocks: a critical review. ChemBioEng Rev. 2022;9:423–441. doi: 10.1002/cben.202100049. [DOI] [Google Scholar]
- 38.Brandao M., Heijungs R., Cowie A.R. On quantifying sources of uncertainty in the carbon footprint of biofuels: crop/feedstock, LCA modelling approach, land-use change, and GHG metrics. Biofuel Research Journal. 2022;9:1608–1616. doi: 10.18331/BRJ2022.9.2.2. [DOI] [Google Scholar]
- 39.Gutschmann B., Huang B., Santolin L., Thiele I., Neubauer P., Riedel S.L. Native feedstock options for the polyhydroxyalkanoate industry in Europe: a review. Microbiol. Res. 2022;264 doi: 10.1016/j.micres.2022.127177. [DOI] [PubMed] [Google Scholar]
- 40.Thanigaivel S., Priya A.K., Dutta K., Rajendran S., Vasseghian Y. Engineering strategies and opportunities of next generation biofuel from microalgae: a perspective review on the potential bioenergy feedstock. Fuel. 2022;312 doi: 10.1016/j.fuel.2021.122827. [DOI] [Google Scholar]
- 41.Singh A., Prajapati P., Vyas S., Gaur V.K., Sindhu R., Binod P., Kumar V., Singhania R.R., Awasthi M.K., Zhang Z., Varjani S. A comprehensive review of feedstocks as sustainable substrates for next-generation biofuels. Bioenerg. Res. 2023;16:105–122. doi: 10.1007/s12155-022-10440-2. [DOI] [Google Scholar]
- 42.Narisetty V., Okibe M.C., Amulya K., Jokodola E.O., Coulon F., Tyagi V.K., Lens P.N.L., Parameswaran B., Kumar V. Technological advancements in valorization of second generation (2G) feedstocks for bio-based succinic acid production. Bioresour. Technol. 2022;360 doi: 10.1016/j.biortech.2022.127513. [DOI] [PubMed] [Google Scholar]
- 43.Goh B.H.H., Chong C.T., Ong H.C., Milano J., Shamsuddin A.H., Lee X.J., Ng J.-H. Strategies for fuel property enhancement for second-generation multi-feedstock biodiesel. Fuel. 2022;315 doi: 10.1016/j.fuel.2022.123178. [DOI] [Google Scholar]
- 44.Wicki L., Naglis-Liepa K., Filipiak T., Parzonko A., Wicka A. Is the production of agricultural biogas environmentally friendly? Does the structure of consumption of first- and second-generation raw materials in Latvia and Poland matter? Energies. 2022;15:5623. doi: 10.3390/en15155623. [DOI] [Google Scholar]
- 45.Kanan M., Habib M.S., Shahbaz A., Hussain A., Habib T., Raza H., Abusaq Z., Assaf R. A grey-fuzzy programming approach towards socio-economic optimization of second-generation biodiesel supply chains. Sustainability. 2022;14 doi: 10.3390/su141610169. [DOI] [Google Scholar]
- 46.Areniello M., Matassa S., Esposito G., Lens P.N.L. Biowaste upcycling into second-generation microbial protein through mixed-culture fermentation. Trends Biotechnol. 2023;41:197–213. doi: 10.1016/j.tibtech.2022.07.008. [DOI] [PubMed] [Google Scholar]
- 47.Sachin Powar R., Singh Yadav A., Siva Ramakrishna Ch, Patel S., Mohan M., Sakharwade S.G., Choubey M., Kumar Ansu A., Sharma A. Algae: a potential feedstock for third generation biofuel. Mater. Today: Proc. 2022;63 doi: 10.1016/j.matpr.2022.07.161. A27–A33. [DOI] [Google Scholar]
- 48.Goswami L., Kayalvizhi R., Dikshit P.K., Sherpa K.C., Roy S., Kushwaha A., Kim B.S., Banerjee R., Jacob S., Rajak R.C. A critical review on prospects of bio-refinery products from second and third generation biomasses. Chem. Eng. J. 2022;448 doi: 10.1016/j.cej.2022.137677. [DOI] [Google Scholar]
- 49.Maliha A., Abu-Hijleh B. Energy Syst; 2022. A Review on the Current Status and Post-pandemic Prospects of Third-Generation Biofuels. [DOI] [Google Scholar]
- 50.Ahmad A., Yadav A.K., Singh A. Biodiesel yield optimisation from a third-generation feedstock (microalgae spirulina) using a hybrid statistical approach. Int. J. Ambient Energy. 2023:1–12. doi: 10.1080/01430750.2023.2169758. [DOI] [Google Scholar]
- 51.Wong K.H., Tan I.S., Foo H.C.Y., Chin L.M., Cheah J.R.N., Sia J.K., Tong K.T.X., Lam M.K. Third-generation bioethanol and L-lactic acid production from red macroalgae cellulosic residue: prospects of Industry 5.0 algae. Energy Convers. Manag. 2022;253 doi: 10.1016/j.enconman.2021.115155. [DOI] [Google Scholar]
- 52.Zahan K.A., Kano M. Biodiesel production from palm oil, its by-products, and mill effluent: a review. Energies. 2018;11:2132. doi: 10.3390/en11082132. [DOI] [Google Scholar]
- 53.Karmakar B., Halder G. Progress and future of biodiesel synthesis: advancements in oil extraction and conversion technologies. Energy Convers. Manag. 2019;182:307–339. doi: 10.1016/j.enconman.2018.12.066. [DOI] [Google Scholar]
- 54.Mahlia T.M.I., Syazmi Z.A.H.S., Mofijur M., Abas A.E.P., Bilad M.R., Ong H.C., Silitonga A.S. Patent landscape review on biodiesel production: technology updates. Renew. Sustain. Energy Rev. 2020;118 doi: 10.1016/j.rser.2019.109526. [DOI] [Google Scholar]
- 55.Ong H.C., Milano J., Silitonga A.S., Hassan M.H., Shamsuddin A.H., Wang C.-T., Indra Mahlia T.M., Siswantoro J., Kusumo F., Sutrisno J. Biodiesel production from Calophyllum inophyllum-Ceiba pentandra oil mixture: optimization and characterization. J. Clean. Prod. 2019;219:183–198. doi: 10.1016/j.jclepro.2019.02.048. [DOI] [Google Scholar]
- 56.Nabi MdN., Hussam W.K., Afroz H.M.M., Rashid A.B., Islam J., Mukut A.N.M.M.I. Investigation of engine performance, combustion, and emissions using waste tire Oil-Diesel-Glycine max biodiesel blends in a diesel engine. Case Stud. Therm. Eng. 2022;39 doi: 10.1016/j.csite.2022.102435. [DOI] [Google Scholar]
- 57.Nabi MdN., Hussam W.K., Rashid A.B., Islam J., Islam S., Afroz H.M.M. Notable improvement of fuel properties of waste tire pyrolysis oil by blending a novel pumpkin seed oil–biodiesel. Energy Rep. 2022;8:112–119. doi: 10.1016/j.egyr.2022.10.246. [DOI] [Google Scholar]
- 58.Thitsartarn W., Kawi S. An active and stable CaO–CeO2 catalyst for transesterification of oil to biodiesel. Green Chem. 2011;13:3423–3430. doi: 10.1039/C1GC15596B. [DOI] [Google Scholar]
- 59.Soares Dias A.P., Bernardo J., Felizardo P., Correia M. Biodiesel production over thermal activated cerium modified Mg-Al hydrotalcites. Energy. 2013;41:344–353. doi: 10.1016/j.energy.2012.03.005. [DOI] [Google Scholar]
- 60.Sivaprakash G., Mohanrasu K., Ravindran B., Jin Chung W., Al Farraj D.A., Soliman Elshikh M., Al Khulaifi M.M., Alkufeidy R.M., Arun A. Integrated approach: Al2O3-CaO nanocatalytic biodiesel production and antibacterial potential silver nanoparticle synthesis from Pedalium murex extract. J. King Saud Univ. Sci. 2020;32:1503–1509. doi: 10.1016/j.jksus.2019.12.004. [DOI] [Google Scholar]
- 61.Pulluri G., Padal K., Jaikumar S. Experimental investigation on a compression ignition engine operating with Al2O3 nanoparticles and Schleichera oleosa biodiesel: combustion, vibration, and noise approach. Nanotechnology for Environmental Engineering. 2022 doi: 10.1007/s41204-022-00299-1. [DOI] [Google Scholar]
- 62.Hossain M.A., Rahman F., Mamun M., Naznin S., Rashid A. 2017. Comparative Analysis of Emission Characteristics and Noise Test of an I.C. Engine Using Different Biodiesel Blends. [DOI] [Google Scholar]
- 63.Cako E., Wang Z., Castro-Muñoz R., Rayaroth M.P., Boczkaj G. Cavitation based cleaner technologies for biodiesel production and processing of hydrocarbon streams: a perspective on key fundamentals, missing process data and economic feasibility – a review. Ultrason. Sonochem. 2022;88 doi: 10.1016/j.ultsonch.2022.106081. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Bin Rashid A. Utilization of nanotechnology and nanomaterials in biodiesel production and property enhancement. J. Nanomater. 2023;2023 doi: 10.1155/2023/7054045. [DOI] [Google Scholar]
- 65.Alaei S., Haghighi M., Toghiani J., Rahmani Vahid B. Magnetic and reusable MgO/MgFe2O4 nanocatalyst for biodiesel production from sunflower oil: influence of fuel ratio in combustion synthesis on catalytic properties and performance. Ind. Crop. Prod. 2018;117:322–332. doi: 10.1016/j.indcrop.2018.03.015. [DOI] [Google Scholar]
- 66.Liu S., McDonald T., Wang Y. Producing biodiesel from high free fatty acids waste cooking oil assisted by radio frequency heating. Fuel. 2010;89:2735–2740. doi: 10.1016/j.fuel.2010.03.011. [DOI] [Google Scholar]
- 67.Gardy J., Rehan M., Hassanpour A., Lai X., Nizami A.-S. Advances in nano-catalysts based biodiesel production from non-food feedstocks. J. Environ. Manag. 2019;249 doi: 10.1016/j.jenvman.2019.109316. [DOI] [PubMed] [Google Scholar]
- 68.Endalew A.K., Kiros Y., Zanzi R. Inorganic heterogeneous catalysts for biodiesel production from vegetable oils. Biomass Bioenergy. 2011;35:3787–3809. doi: 10.1016/j.biombioe.2011.06.011. [DOI] [Google Scholar]
- 69.Helwani Z., Othman M.R., Aziz N., Kim J., Fernando W.J.N. Solid heterogeneous catalysts for transesterification of triglycerides with methanol: a review. Appl. Catal., A. 2009;363:1–10. doi: 10.1016/j.apcata.2009.05.021. [DOI] [Google Scholar]
- 70.Schuchardt U., Sercheli R., Vargas R.M. Transesterification of vegetable oils: a review. J. Braz. Chem. Soc. 1998;9:199–210. doi: 10.1590/S0103-50531998000300002. [DOI] [Google Scholar]
- 71.Silitonga A.S., Shamsuddin A.H., Mahlia T.M.I., Milano J., Kusumo F., Siswantoro J., Dharma S., Sebayang A.H., Masjuki H.H., Ong H.C. Biodiesel synthesis from Ceiba pentandra oil by microwave irradiation-assisted transesterification: ELM modeling and optimization. Renew. Energy. 2020;146:1278–1291. doi: 10.1016/j.renene.2019.07.065. [DOI] [Google Scholar]
- 72.Melero J.A., Iglesias J., Morales G. Heterogeneous acid catalysts for biodiesel production: current status and future challenges. Green Chem. 2009;11:1285–1308. doi: 10.1039/B902086A. [DOI] [Google Scholar]
- 73.Marwaha A., Dhir A., Mahla S.K., Mohapatra S.K. An overview of solid base heterogeneous catalysts for biodiesel production. Catal. Rev. 2018;60:594–628. doi: 10.1080/01614940.2018.1494782. [DOI] [Google Scholar]
- 74.Calero J., Luna D., Sancho E.D., Luna C., Bautista F.M., Romero A.A., Posadillo A., Verdugo C. Development of a new biodiesel that integrates glycerol, by using CaO as heterogeneous catalyst, in the partial methanolysis of sunflower oil. Fuel. 2014;122:94–102. doi: 10.1016/j.fuel.2014.01.033. [DOI] [Google Scholar]
- 75.Knothe G., Steidley K.R. The effect of metals and metal oxides on biodiesel oxidative stability from promotion to inhibition. Fuel Process. Technol. 2018;177:75–80. doi: 10.1016/j.fuproc.2018.04.009. [DOI] [Google Scholar]
- 76.Kalavathy G., Baskar G. Synergism of clay with zinc oxide as nanocatalyst for production of biodiesel from marine Ulva lactuca. Bioresour. Technol. 2019;281:234–238. doi: 10.1016/j.biortech.2019.02.101. [DOI] [PubMed] [Google Scholar]
- 77.Quah R.V., Tan Y.H., Mubarak N.M., Khalid M., Abdullah E.C., Nolasco-Hipolito C. An overview of biodiesel production using recyclable biomass and non-biomass derived magnetic catalysts. J. Environ. Chem. Eng. 2019;7 doi: 10.1016/j.jece.2019.103219. [DOI] [Google Scholar]
- 78.Semwal S., Arora A.K., Badoni R.P., Tuli D.K. Biodiesel production using heterogeneous catalysts. Bioresour. Technol. 2011;102:2151–2161. doi: 10.1016/j.biortech.2010.10.080. [DOI] [PubMed] [Google Scholar]
- 79.Cao F., Chen Y., Zhai F., Li J., Wang J., Wang X., Wang S., Zhu W. Biodiesel production from high acid value waste frying oil catalyzed by superacid heteropolyacid. Biotechnol. Bioeng. 2008;101:93–100. doi: 10.1002/bit.21879. [DOI] [PubMed] [Google Scholar]
- 80.Aransiola E.F., Ojumu T.V., Oyekola O.O., Madzimbamuto T.F., Ikhu-Omoregbe D.I.O. A review of current technology for biodiesel production: state of the art. Biomass Bioenergy. 2014;61:276–297. doi: 10.1016/j.biombioe.2013.11.014. [DOI] [Google Scholar]
- 81.Hama S., Noda H., Kondo A. How lipase technology contributes to evolution of biodiesel production using multiple feedstocks. Curr. Opin. Biotechnol. 2018;50:57–64. doi: 10.1016/j.copbio.2017.11.001. [DOI] [PubMed] [Google Scholar]
- 82.Amini Z., Ilham Z., Ong H.C., Mazaheri H., Chen W.-H. State of the art and prospective of lipase-catalyzed transesterification reaction for biodiesel production. Energy Convers. Manag. 2017;141:339–353. doi: 10.1016/j.enconman.2016.09.049. [DOI] [Google Scholar]
- 83.Christopher L.P., Kumar Hemanathan, Zambare V.P. Enzymatic biodiesel: challenges and opportunities. Appl. Energy. 2014;119:497–520. doi: 10.1016/j.apenergy.2014.01.017. [DOI] [Google Scholar]
- 84.Mardhiah H.H., Ong H.C., Masjuki H.H., Lim S., Lee H.V. A review on latest developments and future prospects of heterogeneous catalyst in biodiesel production from non-edible oils. Renew. Sustain. Energy Rev. 2017;67:1225–1236. doi: 10.1016/j.rser.2016.09.036. [DOI] [Google Scholar]
- 85.Schwirn K., Tietjen L., Beer I. Why are nanomaterials different and how can they be appropriately regulated under REACH? Environ. Sci. Eur. 2014;26:4. doi: 10.1186/2190-4715-26-4. [DOI] [Google Scholar]
- 86.Qiu F., Li Y., Yang D., Li X., Sun P. Heterogeneous solid base nanocatalyst: preparation, characterization and application in biodiesel production. Bioresour. Technol. 2011;102:4150–4156. doi: 10.1016/j.biortech.2010.12.071. [DOI] [PubMed] [Google Scholar]
- 87.Rahmani Vahid B., Haghighi M., Alaei S., Toghiani J. Reusability enhancement of combustion synthesized MgO/MgAl2O4 nanocatalyst in biodiesel production by glow discharge plasma treatment. Energy Convers. Manag. 2017;143:23–32. doi: 10.1016/j.enconman.2017.03.075. [DOI] [Google Scholar]
- 88.Rasouli H., Esmaeili H. Characterization of MgO nanocatalyst to produce biodiesel from goat fat using transesterification process. 3 Biotech. 2019;9:429. doi: 10.1007/s13205-019-1963-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Zuliani A., Ivars F., Luque R. Advances in nanocatalyst design for biofuel production. ChemCatChem. 2018;10:1968–1981. doi: 10.1002/cctc.201701712. [DOI] [Google Scholar]
- 90.Banerjee S., Rout S., Banerjee S., Atta A., Das D. Fe2O3 nanocatalyst aided transesterification for biodiesel production from lipid-intact wet microalgal biomass: a biorefinery approach. Energy Convers. Manag. 2019;195:844–853. doi: 10.1016/j.enconman.2019.05.060. [DOI] [Google Scholar]
- 91.Kumar D., Sharma S., Srivastava N., Shukla S., Gaurav* K. Advancement in the utilization of nanocatalyst for transesterification of triglycerides. Journal of Nanoscience and Technology. 2018:374–379. doi: 10.30799/jnst.111.18040302. [DOI] [Google Scholar]
- 92.Bharti P., Singh B., Dey R.K. Process optimization of biodiesel production catalyzed by CaO nanocatalyst using response surface methodology. J Nanostruct Chem. 2019;9:269–280. doi: 10.1007/s40097-019-00317-w. [DOI] [Google Scholar]
- 93.Rahman W.U., Fatima A., Anwer A.H., Athar M., Khan M.Z., Khan N.A., Halder G. Biodiesel synthesis from eucalyptus oil by utilizing waste egg shell derived calcium based metal oxide catalyst. Process Saf. Environ. Protect. 2019;122:313–319. doi: 10.1016/j.psep.2018.12.015. [DOI] [Google Scholar]
- 94.Wen L., Wang Y., Lu D., Hu S., Han H. Preparation of KF/CaO nanocatalyst and its application in biodiesel production from Chinese tallow seed oil. Fuel. 2010;89:2267–2271. doi: 10.1016/j.fuel.2010.01.028. [DOI] [Google Scholar]
- 95.Savitri S. Development of heterogeneous alumina supported base catalyst for biodiesel production. Mater. Sci. 2012 https://www.academia.edu/2569672/Development_of_Heterogeneous_Alumina_Supported_Base_Catalyst_for_Biodiesel_Production (accessed March 17, 2023) [Google Scholar]
- 96.Fallah Kelarijani A., Gholipour Zanjani N., Kamran Pirzaman A. Ultrasonic assisted transesterification of rapeseed oil to biodiesel using nano magnetic catalysts. Waste Biomass Valor. 2020;11:2613–2621. doi: 10.1007/s12649-019-00593-1. [DOI] [Google Scholar]
- 97.Degfie T.A., Mamo T.T., Mekonnen Y.S. Optimized biodiesel production from waste cooking oil (WCO) using calcium oxide (CaO) nano-catalyst. Sci. Rep. 2019;9 doi: 10.1038/s41598-019-55403-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Thitsartarn W., Kawi S. An active and stable CaO–CeO2 catalyst for transesterification of oil to biodiesel. Green Chem. 2011;13:3423–3430. doi: 10.1039/C1GC15596B. [DOI] [Google Scholar]
- 99.Dahlquist E. Bioresource Technology; 2010. Biodiesel Production from Waste Cooking Oil Catalyzed by TiO 2–MgO Mixed Oxides.https://www.academia.edu/6938354/Biodiesel_production_from_waste_cooking_oil_catalyzed_by_TiO_2_MgO_mixed_oxides (accessed March 17, 2023) [DOI] [PubMed] [Google Scholar]
- 100.Alaei S., Haghighi M., Toghiani J., Rahmani Vahid B. Magnetic and reusable MgO/MgFe2O4 nanocatalyst for biodiesel production from sunflower oil: influence of fuel ratio in combustion synthesis on catalytic properties and performance. Ind. Crop. Prod. 2018;117:322–332. doi: 10.1016/j.indcrop.2018.03.015. [DOI] [Google Scholar]
- 101.Lee H.V., Taufiq-Yap Y.H., Hussein M.Z., Yunus R. Transesterification of jatropha oil with methanol over Mg–Zn mixed metal oxide catalysts. Energy. 2013;49:12–18. [Google Scholar]
- 102.Asghar M. Ghasemi. Transesterification of waste cooking oil to biodiesel using Ca and Zr mixed oxides as heterogeneous base catalysts. Fuel Process. Technol. 2012;97:45–51. doi: 10.1016/j.fuproc.2012.01.010. [DOI] [Google Scholar]
- 103.Rahmani Vahid B., Haghighi M. Urea-nitrate combustion synthesis of MgO/MgAl2O4 nanocatalyst used in biodiesel production from sunflower oil: influence of fuel ratio on catalytic properties and performance. Energy Convers. Manag. 2016;126:362–372. doi: 10.1016/j.enconman.2016.07.050. [DOI] [Google Scholar]
- 104.Wen Z., Yu X., Tu S.-T., Yan J., Dahlquist E. Synthesis of biodiesel from vegetable oil with methanol catalyzed by Li-doped magnesium oxide catalysts. Appl. Energy. 2010;87:743–748. doi: 10.1016/j.apenergy.2009.09.013. [DOI] [Google Scholar]
- 105.Tahvildari K., Anaraki Y.N., Fazaeli R., Mirpanji S., Delrish E. The study of CaO and MgO heterogenic nano-catalyst coupling on transesterification reaction efficacy in the production of biodiesel from recycled cooking oil. J Environ Health Sci Engineer. 2015;13:73. doi: 10.1186/s40201-015-0226-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106.Madhuvilakku R., Piraman S. Biodiesel synthesis by TiO2–ZnO mixed oxide nanocatalyst catalyzed palm oil transesterification process. Bioresour. Technol. 2013;150:55–59. doi: 10.1016/j.biortech.2013.09.087. [DOI] [PubMed] [Google Scholar]
- 107.Gardy J., Osatiashtiani A., Céspedes O., Hassanpour A., Lai X., Lee A.F., Wilson K., Rehan M. A magnetically separable SO4/Fe-Al-TiO2 solid acid catalyst for biodiesel production from waste cooking oil. Appl. Catal. B Environ. 2018;234:268–278. doi: 10.1016/j.apcatb.2018.04.046. [DOI] [Google Scholar]
- 108.Alves M.B., Medeiros F.C.M., Sousa M.H., Rubim J.C., Suarez P.A.Z. Cadmium and tin magnetic nanocatalysts useful for biodiesel production. J. Braz. Chem. Soc. 2014;25:2304–2313. doi: 10.5935/0103-5053.20140238. [DOI] [Google Scholar]
- 109.Nizami A.-S., Rehan M. Towards nanotechnology-based biofuel industry. Biofuel Research Journal. 2018;5:798–799. doi: 10.18331/BRJ2018.5.2.2. [DOI] [Google Scholar]
- 110.Rai M., dos Santos J.C., Soler M.F., Marcelino P.R.F., Brumano L.P., Ingle A.P., Gaikwad S., Gade A., da Silva S.S. Strategic role of nanotechnology for production of bioethanol and biodiesel. Nanotechnol. Rev. 2016;5:231–250. doi: 10.1515/ntrev-2015-0069. [DOI] [Google Scholar]
- 111.Guan Q., Li Y., Chen Y., Shi Y., Gu J., Li B., Miao R., Chen Q., Ning P. Sulfonated multi-walled carbon nanotubes for biodiesel production through triglycerides transesterification. RSC Adv. 2017;7:7250–7258. doi: 10.1039/C6RA28067F. [DOI] [Google Scholar]
- 112.Shuit S.H., Ng E.P., Tan S.H. A facile and acid-free approach towards the preparation of sulphonated multi-walled carbon nanotubes as a strong protonic acid catalyst for biodiesel production. J. Taiwan Inst. Chem. Eng. 2015;52:100–108. doi: 10.1016/j.jtice.2015.02.018. [DOI] [Google Scholar]
- 113.Asri N.P., Yuniati Y., Hindarso H., Suprapto, Yogaswara R.R. Biodiesel production from Kesambi (Schleichera oleosa) oil using multi-walled carbon nanotubes supported zinc oxide as a solid acid catalyst. IOP Conf. Ser. Earth Environ. Sci. 2020;456 doi: 10.1088/1755-1315/456/1/012003. [DOI] [Google Scholar]
- 114.Fan Y., Su F., Li K., Ke C., Yan Y. Carbon nanotube filled with magnetic iron oxide and modified with polyamidoamine dendrimers for immobilizing lipase toward application in biodiesel production. Sci. Rep. 2017;7 doi: 10.1038/srep45643. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.Abdullah S.H.Y.S., Hanapi N.H.M., Azid A., Umar R., Juahir H., Khatoon H., Endut A. A review of biomass-derived heterogeneous catalyst for a sustainable biodiesel production. Renew. Sustain. Energy Rev. 2017;70:1040–1051. doi: 10.1016/j.rser.2016.12.008. [DOI] [Google Scholar]
- 116.Verma M., Godbout S., Brar S.K., Solomatnikova O., Lemay S.P., Larouche J.P. Biofuels production from biomass by thermochemical conversion technologies. Int. J. Chem. Eng. 2012 doi: 10.1155/2012/542426. [DOI] [Google Scholar]
- 117.Laurent S., Forge D., Port M., Roch A., Robic C., Vander Elst L., Muller R.N. Magnetic iron oxide nanoparticles: synthesis, stabilization, vectorization, physicochemical characterizations, and biological applications. Chem. Rev. 2008;108:2064–2110. doi: 10.1021/cr068445e. [DOI] [PubMed] [Google Scholar]
- 118.Abraham R.E., Verma M.L., Barrow C.J., Puri M. Suitability of magnetic nanoparticle immobilised cellulases in enhancing enzymatic saccharification of pretreated hemp biomass. Biotechnol. Biofuels. 2014;7:90. doi: 10.1186/1754-6834-7-90. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119.Al-Ani A., Darton R.J., Sneddon S., Zholobenko V. Nanostructured zeolites: the introduction of intracrystalline mesoporosity in basic faujasite-type catalysts. ACS Appl. Nano Mater. 2018;1:310–318. doi: 10.1021/acsanm.7b00169. [DOI] [Google Scholar]
- 120.Zhang Z., Yuan X., Miao S., Li H., Shan W., Jia M., Zhang C. Effect of Fe additives on the catalytic performance of ion-exchanged CsX zeolites for side-chain alkylation of toluene with methanol. Catalysts. 2019;9:829. doi: 10.3390/catal9100829. [DOI] [Google Scholar]
- 121.Abukhadra M.R., Ibrahim S.M., Yakout S.M., El-Zaidy M.E., Abdeltawab A.A. Synthesis of Na+ trapped bentonite/zeolite-P composite as a novel catalyst for effective production of biodiesel from palm oil; Effect of ultrasonic irradiation and mechanism. Energy Convers. Manag. 2019;196:739–750. doi: 10.1016/j.enconman.2019.06.027. [DOI] [Google Scholar]
- 122.Xie W., Huang X., Li H. Soybean oil methyl esters preparation using NaX zeolites loaded with KOH as a heterogeneous catalyst. Bioresour. Technol. 2007;98:936–939. doi: 10.1016/j.biortech.2006.04.003. [DOI] [PubMed] [Google Scholar]
- 123.Wu H., Zhang J., Wei Q., Zheng J., Zhang J. Transesterification of soybean oil to biodiesel using zeolite supported CaO as strong base catalysts. Fuel Process. Technol. 2013;109:13–18. doi: 10.1016/j.fuproc.2012.09.032. [DOI] [Google Scholar]
- 124.Pratap S.R., Shamshuddin S.Z.M., Thimmaraju N., Shyamsundar M., Reena S.S. Kinetics of transesterification of Madhuca Indica oil over modified zeolites: biodiesel synthesis. Bangladesh J. Sci. Ind. Res. 2015;50:271–278. doi: 10.3329/bjsir.v50i4.25836. [DOI] [Google Scholar]
- 125.Firouzjaee M.H., Taghizadeh M. Optimization of process variables for biodiesel production using the nanomagnetic catalyst CaO/NaY-Fe3O4. Chem. Eng. Technol. 2017;40:1140–1148. doi: 10.1002/ceat.201600406. [DOI] [Google Scholar]
- 126.Vieira S.S., Magriotis Z.M., Santos N.A.V., Saczk A.A., Hori C.E., Arroyo P.A. Biodiesel production by free fatty acid esterification using lanthanum (La3+) and HZSM-5 based catalysts. Bioresour. Technol. 2013;133:248–255. doi: 10.1016/j.biortech.2013.01.107. [DOI] [PubMed] [Google Scholar]
- 127.Mohebbi S., Rostamizadeh M., Kahforoushan D. Effect of molybdenum promoter on performance of high silica MoO3/B-ZSM-5 nanocatalyst in biodiesel production. Fuel. 2020;266 doi: 10.1016/j.fuel.2020.117063. [DOI] [Google Scholar]
- 128.Avhad M.R., Marchetti J.M. A review on recent advancement in catalytic materials for biodiesel production. Renew. Sustain. Energy Rev. 2015;50:696–718. doi: 10.1016/j.rser.2015.05.038. [DOI] [Google Scholar]
- 129.Ferreira L.F.P., de Oliveira T.M., Toma S.H., Toyama M.M., Araki K., Avanzi L.H. Superparamagnetic iron oxide nanoparticles (SPIONs) conjugated with lipase Candida antarctica A for biodiesel synthesis. RSC Adv. 2020;10:38490–38496. doi: 10.1039/D0RA06215D. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130.Li Q., Fan F., Wang Y., Feng W., Ji P. Enzyme immobilization on carboxyl-functionalized graphene oxide for catalysis in organic solvent. Ind. Eng. Chem. Res. 2013;52:6343–6348. doi: 10.1021/ie400558u. [DOI] [Google Scholar]
- 131.Zhong L., Feng Y., Wang G., Wang Z., Bilal M., Lv H., Jia S., Cui J. Production and use of immobilized lipases in/on nanomaterials: a review from the waste to biodiesel production. Int. J. Biol. Macromol. 2020;152:207–222. doi: 10.1016/j.ijbiomac.2020.02.258. [DOI] [PubMed] [Google Scholar]
- 132.Rezania S., Mahdinia S., Oryani B., Cho J., Kwon E.E., Bozorgian A., Rashidi Nodeh H., Darajeh N., Mehranzamir K. Biodiesel production from wild mustard (Sinapis Arvensis) seed oil using a novel heterogeneous catalyst of LaTiO3 nanoparticles. Fuel. 2022;307 doi: 10.1016/j.fuel.2021.121759. [DOI] [Google Scholar]
- 133.Maleki B., Ashraf Talesh S.S. Optimization of the sono-biodiesel in the attendance of ZnO nanoparticles, process yield enhancement: box behnken design. Journal of Chemical and Petroleum Engineering. 2022;56:1–14. doi: 10.22059/jchpe.2021.330251.1361. [DOI] [Google Scholar]
- 134.Elkelawy M., Etaiw S.E.H., Bastawissi H.A.-E., Marie H., Radwan A.M., Dawood M.M., Panchal H. WCO biodiesel production by heterogeneous catalyst and using cadmium (II)-based supramolecular coordination polymer additives to improve diesel/biodiesel fueled engine performance and emissions. J. Therm. Anal. Calorim. 2022;147:6375–6391. doi: 10.1007/s10973-021-10920-1. [DOI] [Google Scholar]
- 135.Saka A., Enkosa E., Jule L.T., Nagaprasad N., Subramanian K., Ramaswamy K. Biomass Conv. Bioref.; 2022. Biofuel Production from Mango (Mangifera Indica) Seed Extracts through Zinc Oxide Nanoparticle. [DOI] [Google Scholar]
- 136.Parandi E., Safaripour M., Abdellattif M.H., Saidi M., Bozorgian A., Rashidi Nodeh H., Rezania S. Biodiesel production from waste cooking oil using a novel biocatalyst of lipase enzyme immobilized magnetic nanocomposite. Fuel. 2022;313 doi: 10.1016/j.fuel.2021.123057. [DOI] [Google Scholar]
- 137.Rozina S.R., Ahmad ChiaM., Sultana S., Zafar M., Asif S., Bokhari A., Nomanbhay S., Mubashir M., Khoo K.S., Show P.L. Green synthesis of biodiesel from Citrus medica seed oil using green nanoparticles of copper oxide. Fuel. 2022;323 doi: 10.1016/j.fuel.2022.124285. [DOI] [Google Scholar]
- 138.Maleki B., Ashraf Talesh S.S. Optimization of ZnO incorporation to αFe2O3 nanoparticles as an efficient catalyst for biodiesel production in a sonoreactor: application on the CI engine. Renew. Energy. 2022;182:43–59. doi: 10.1016/j.renene.2021.10.013. [DOI] [Google Scholar]
- 139.Prokaewa A., Smith S.M., Luengnaruemitchai A., Kandiah M., Boonyuen S. Biodiesel production from waste cooking oil using a new heterogeneous catalyst SrO doped CaO nanoparticles. Journal of Metals, Materials and Minerals. 2022;32:79–85. doi: 10.55713/jmmm.v32i1.1149. [DOI] [Google Scholar]
- 140.Bahadoran A., Ramakrishna S., Oryani B., Ahmed Al-Keridis L., Rashidi Nodeh H., Rezania S. Biodiesel production from waste cooking oil using heterogeneous nanocatalyst-based magnetic polyaniline decorated with cobalt oxide. Fuel. 2022;319 doi: 10.1016/j.fuel.2022.123858. [DOI] [Google Scholar]
- 141.Salim S.M., Izriq R., Almaky M.M., Al-Abbassi A.A. Synthesis and characterization of ZnO nanoparticles for the production of biodiesel by transesterification: kinetic and thermodynamic studies. Fuel. 2022;321 doi: 10.1016/j.fuel.2022.124135. [DOI] [Google Scholar]
- 142.Amruth Maroju P., Ganesan R., Ray Dutta J. Biofuel generation from food waste through immobilized enzymes on magnetic nanoparticles. Mater. Today: Proc. 2023;72:62–66. doi: 10.1016/j.matpr.2022.05.555. [DOI] [Google Scholar]
- 143.Al-Zaban M.I., AlHarbi M.A., Mahmoud M.A., Bahatheq A.M. Production of biodiesel from oleaginous fungal lipid using highly catalytic bimetallic gold-silver core-shell nanoparticle. J. Appl. Microbiol. 2022;132:381–389. doi: 10.1111/jam.15176. [DOI] [PubMed] [Google Scholar]
- 144.Rozina M., Ahmad S., Asif J.J., Klemeš M., Mubashir A., Bokhari S., Sultana A., Mukhtar M., Zafar A.A., Bazmi S., Ullah M.S., Khan A.K., Koyande M., Mofijur P.-L.Show. Conversion of the toxic and hazardous Zanthoxylum armatum seed oil into methyl ester using green and recyclable silver oxide nanoparticles. Fuel. 2022;310 doi: 10.1016/j.fuel.2021.122296. [DOI] [Google Scholar]
- 145.Cholapandian K., Gurunathan B., Rajendran N. Investigation of CaO nanocatalyst synthesized from Acalypha indica leaves and its application in biodiesel production using waste cooking oil. Fuel. 2022;312 doi: 10.1016/j.fuel.2021.122958. [DOI] [Google Scholar]
- 146.Yatish K.V., Prakash R.M., Ningaraju C., Sakar M., GeethaBalakrishna R., Lalithamba H.S. Terminalia chebula as a novel green source for the synthesis of copper oxide nanoparticles and as feedstock for biodiesel production and its application on diesel engine. Energy. 2021;215 doi: 10.1016/j.energy.2020.119165. [DOI] [Google Scholar]
- 147.Zhang P., Liu P., Fan M., Jiang P., Haryono A. High-performance magnetite nanoparticles catalyst for biodiesel production: immobilization of 12-tungstophosphoric acid on SBA-15 works effectively. Renew. Energy. 2021;175:244–252. doi: 10.1016/j.renene.2021.05.033. [DOI] [Google Scholar]
- 148.Shalini P., Deepanraj B., Vijayalakshmi S., Ranjitha J. Synthesis and characterisation of lipase immobilised magnetic nanoparticles and its role as a catalyst in biodiesel production. Mater. Today: Proc. 2021 doi: 10.1016/j.matpr.2021.07.027. [DOI] [Google Scholar]
- 149.Rezania S., Korrani Z.S., Gabris M.A., Cho J., Yadav K.K., Cabral-Pinto M.M.S., Alam J., Ahamed M., Nodeh H.R. Lanthanum phosphate foam as novel heterogeneous nanocatalyst for biodiesel production from waste cooking oil. Renew. Energy. 2021;176:228–236. doi: 10.1016/j.renene.2021.05.060. [DOI] [Google Scholar]
- 150.Rozina M., Ahmad M., Alruqi M.Zafar. Cleaner production of biodiesel from novel and non-edible seed oil of Chamaerops humilis using recyclable cobalt oxide nanoparticles: a contribution to resilient and sustainable world. J. Clean. Prod. 2022;369 doi: 10.1016/j.jclepro.2022.133378. [DOI] [Google Scholar]
- 151.Davoodbasha M., Pugazhendhi A., Kim J.-W., Lee S.-Y., Nooruddin T. Biodiesel production through transesterification of Chlorella vulgaris: synthesis and characterization of CaO nanocatalyst. Fuel. 2021;300 doi: 10.1016/j.fuel.2021.121018. [DOI] [Google Scholar]
- 152.Rozina M., Ahmad M.Zafar. Conversion of waste seed oil of Citrus aurantium into methyl ester via green and recyclable nanoparticles of zirconium oxide in the context of circular bioeconomy approach. Waste Manag. 2021;136:310–320. doi: 10.1016/j.wasman.2021.10.001. [DOI] [PubMed] [Google Scholar]
- 153.Zulfiqar A., Mumtaz M.W., Mukhtar H., Najeeb J., Irfan A., Akram S., Touqeer T., Nabi G. Lipase-PDA-TiO2 NPs: an emphatic nano-biocatalyst for optimized biodiesel production from Jatropha curcas oil. Renew. Energy. 2021;169:1026–1037. doi: 10.1016/j.renene.2020.12.135. [DOI] [Google Scholar]
- 154.Jume B.H., Gabris M.A., Rashidi Nodeh H., Rezania S., Cho J. Biodiesel production from waste cooking oil using a novel heterogeneous catalyst based on graphene oxide doped metal oxide nanoparticles. Renew. Energy. 2020;162:2182–2189. doi: 10.1016/j.renene.2020.10.046. [DOI] [Google Scholar]
- 155.Sahabdheen A.B., Arivarasu A. Synthesis and characterization of reusable heteropoly acid nanoparticles for one step biodiesel production from high acid value waste cooking oil – performance and emission studies. Mater. Today: Proc. 2020;22:383–392. doi: 10.1016/j.matpr.2019.07.249. [DOI] [Google Scholar]
- 156.Changmai B., Rano R., Vanlalveni C., Rokhum L. A novel Citrus sinensis peel ash coated magnetic nanoparticles as an easily recoverable solid catalyst for biodiesel production. Fuel. 2021;286 doi: 10.1016/j.fuel.2020.119447. [DOI] [Google Scholar]
- 157.Chingakham Ch, David A., Sajith V. Fe3O4 nanoparticles impregnated eggshell as a novel catalyst for enhanced biodiesel production. Chin. J. Chem. Eng. 2019;27:2835–2843. doi: 10.1016/j.cjche.2019.02.022. [DOI] [Google Scholar]
- 158.Falowo O.A., Oloko-Oba M.I., Betiku E. Biodiesel production intensification via microwave irradiation-assisted transesterification of oil blend using nanoparticles from elephant-ear tree pod husk as a base heterogeneous catalyst. Chemical Engineering and Processing - Process Intensification. 2019;140:157–170. doi: 10.1016/j.cep.2019.04.010. [DOI] [Google Scholar]
- 159.Chen Y., Liu T., He H., Liang H. Fe3O4/ZnMg(Al)O magnetic nanoparticles for efficient biodiesel production. Appl. Organomet. Chem. 2018;32 doi: 10.1002/aoc.4330. [DOI] [Google Scholar]
- 160.Muralidharan N.G., Ranjitha J. Optimization of biodiesel production from dairy waste scum using response surface methodology. Res. J. Pharm. Technol. 2019;12:342. doi: 10.5958/0974-360X.2019.00062.3. [DOI] [Google Scholar]
- 161.Ambat I., Srivastava V., Haapaniemi E., Sillanpää M. Nano-magnetic potassium impregnated ceria as catalyst for the biodiesel production. Renew. Energy. 2019;139:1428–1436. doi: 10.1016/j.renene.2019.03.042. [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Data Availability Statement
Data included in article/supp. material/referenced in article.





