Next Article in Journal
Design Optimization of a Direct-Drive Electrically Excited Synchronous Generator for Tidal Wave Energy
Previous Article in Journal
Design and Implementation of Low-Cost Phasor Measurement Unit: PhasorsCatcher
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Biodiesel Is Dead: Long Life to Advanced Biofuels—A Comprehensive Critical Review

Departamento de Química Orgánica, Universidad de Cordoba, Campus de Rabanales, Ed. Marie Curie, 14014 Cordoba, Spain
*
Author to whom correspondence should be addressed.
Energies 2022, 15(9), 3173; https://doi.org/10.3390/en15093173
Submission received: 24 March 2022 / Revised: 13 April 2022 / Accepted: 25 April 2022 / Published: 26 April 2022

Abstract

:
Many countries are immersed in several strategies to reduce the carbon dioxide (CO2) emissions of internal combustion engines. One option is the substitution of these engines by electric and/or hydrogen engines. However, apart from the strategic and logistical difficulties associated with this change, the application of electric or hydrogen engines in heavy transport, e.g., trucks, shipping, and aircrafts, also presents technological difficulties in the short-medium term. In addition, the replacement of the current car fleet will take decades. This is why the use of biofuels is presented as the only viable alternative to diminishing CO2 emissions in the very near future. Nowadays, it is assumed that vegetable oils will be the main raw material for replacing fossil fuels in diesel engines. In this context, it has also been assumed that the reduction in the viscosity of straight vegetable oils (SVO) must be performed through a transesterification reaction with methanol in order to obtain the mixture of fatty acid methyl esters (FAMEs) that constitute biodiesel. Nevertheless, the complexity in the industrial production of this biofuel, mainly due to the costs of eliminating the glycerol produced, has caused a significant delay in the energy transition. For this reason, several advanced biofuels that avoid the glycerol production and exhibit similar properties to fossil diesel have been developed. In this way, “green diesels” have emerged as products of different processes, such as the cracking or pyrolysis of vegetable oil, as well as catalytic (hydro)cracking. In addition, some biodiesel-like biofuels, such as Gliperol (DMC-Biod) or Ecodiesel, as well as straight vegetable oils, in blends with plant-based sources with low viscosity have been described as renewable biofuels capable of performing in combustion ignition engines. After evaluating the research carried out in the last decades, it can be concluded that green diesel and biodiesel-like biofuels could constitute the main alternative to addressing the energy transition, although green diesel will be the principal option in aviation fuel.

1. Introduction

Nowadays, most of the countries worldwide are making an unprecedented effort to reduce anthropogenic greenhouse gas (GHG) emissions to carry out a decarbonization process, which significantly affects the energy sources applied. In this sense, the Treaty of Paris [1] and the European New Green Deal, as well as the REDII, aim to achieve a climate-neutral Europe by 2050 [2]. Therefore, several countries have also implemented their own energy and climate policy framework for 2030 and beyond, advancing in decarbonization and promoting innovation in order to achieve a viable new climate economy low in CO2 emissions [3].
Considering that the choice of green hydrogen as the main energy vector for the decarbonization of the planet seems definitive, biofuels should receive a secondary role in the current research and development priorities of transportation, including cars, trucks, ships, planes, etc. However, the transition from current energy sources to this new technology requires a period of several decades, in accordance with the planning carried out by the same countries involved in these international agreements [4].
Notwithstanding the possibility of building a transport fleet operating with new technologies and being neutral in CO2 emissions, it is mandatory to consider the temporary rate of the substitution of current vehicles working with internal combustion engines (ICEs) in order to avoid an economic chaos of unpredictable consequences. In this sense, the replacement of the enormous number of vehicles that operate with ICEs needs to be carried out in such a way that they can continue operating throughout their useful life with diesel fuels or, alternatively, with biofuels with similar properties. This fact does not constitute a trivial problem due to the very high number of vehicles currently in use and the fact that those vehicles that are being built now and in the next two or three decades must be added to the list [5]. Consequently, the reduction in emissions in this long transition period involves a reduction in fossil fuels and increase in other fuels that allow for their operation in ICEs, together with the incorporation of hydrogen-powered engines and other emerging technologies. In this way, a smooth transition to a scenario without fossil fuels could be foreseen [6].
In this sense, biofuels can be easily integrated into the logistics of the global transportation system. In fact, the goal pursued by EU is that biofuels constitute 30% of all fuels by 2030 [7]. Despite this goal being easy to achieve considering the technical issues, the substitution of fossil fuels with biofuels is considered unattainable in this deadline due to the impossibility of having enough agricultural land to carry out the necessary crops, since bioethanol and biodiesel (the most widely biofuels employed) require enormous agricultural resources to fulfill these purposes [8,9,10,11,12].
Therefore, it is also mandatory to introduce electric engines in the transport sector, since biofuels will not be able to completely replace all fossil fuels currently operating ICEs, considering that 20% of the global emissions of GHG are contributed by this sector [13,14,15].
In summary, to carry out the planned energy transition efficiently and sustainably, it is essential to have biofuels that are technically and economically feasible to not only be able to gradually replace the fossil fuels used by the current diesel engines but also to be used in a long indeterminate time horizon in trucks, ships, and especially airplanes, where the introduction of electric engines cannot yet be considered in a predictable time due to the technology immaturity [5,16,17,18]. Hence, regardless of the progress that will be obtained in the coming decades with respect to the introduction of electric engines in the transport sector, research on biofuels presents the maximum interest, not only to facilitate the necessary energy transition with the profitability of current conventional ICEs, but also for its application in more specific sectors, such as trucks, boats, and aircrafts, along a temporarily indefinite period.
On the other hand, the introduction of biofuels in the energy market dedicated to transport is in its early stages, since only 1.6% of biofuels are used currently, with respect to the total transport fuel worldwide spent [19], including the biodiesel production worldwide, which is around four billion liters. In this context, straight vegetable oils (SVO) and animal fats can be considered as the main option to achieve the substitution of fossil diesel fuel. In order to carry this out, SVO or animal fats are transformed into biodiesel by a transesterification reaction with methanol, producing a mixture of fatty acid methyl esters (FAMEs) [20]. This relatively simple process can reduce oil viscosity to the level of the values of conventional fossil diesel (4–5 mm2/s) and constitutes the only currently applied industrial method to convert vegetable oils into biodiesel [21].
However, the transesterification process reveals a serious drawback at the industrial scale, which is associated with the production of a relatively high amount of glycerol generated as a by-product (10% by weight of the total biodiesel produced). Apart from reducing the performance of the process, the glycerol must be eliminated since the high temperatures reached in engines favor the formation of glycerol polymers, as well as acrolein, which has a high toxicity. In fact, glycerol production is considered the most important barrier, and has, so far, prevented the consolidation of biodiesel as the biofuel that can replace fossil fuels [22].
To overcome the glycerol production problem, different alternatives have been developed in the last decades to transform vegetable oils into high-quality biofuels without the production of glycerol. Thus, several oxygenated biofuels that integrate glycerol as soluble derivatives have been described, e.g., Gliperol, DMC-Biod, or Ecodiesel [23,24]. These biofuels are known as “biodiesel-like biofuels”. In addition, biofuels obtained from triglycerides by different processes, such as cracking, pyrolysis, hydrodeoxygenation, and hydrotreating of vegetable oils, have also been described. These are high-quality renewable diesel fuels generally known as “green diesel” or “renewable diesel”, exhibiting a similar composition to fossil fuel [25]. Finally, the possibility of using various additives in mixtures with SVO is also being evaluated to reduce the kinematic viscosity of the mixtures to the levels required by ICEs. For this reason, biofuels made up of mixtures of SVO and renewable solvents have been described. Since these compounds generally have low octane numbers, these biofuels are used in blends with oils, obtaining the so-called LVLC (less viscous and lower in cetane) fuels [26,27].
Figure 1 summarizes the different methods of transforming animal fats and vegetable oils into biofuels, avoiding the generation of glycerol, exhibiting all of the many advantages with respect to conventional biodiesel. This review is an overall view of the current research in fuel development alternatives to biodiesel, intending also to evaluate the strengths and weaknesses of these alternative processes described in Figure 1 in order to consider them as fuels in different transport sectors, such as heavy road vehicles, aviation, and/or maritime transportation sectors. In addition, this review aims to claim to the scientific community that the research in these alternatives to biodiesel must be made mandatory due to their strong dependence on fossil energy sources, since oil is the main energy source supplying approximately 95% of the sector’s energy consumption [28].

2. Strengths and Weaknesses of Biodiesel as Renewable Biofuel in Current Diesel Engines

Biodiesel is defined as a mixture of long chain fatty acid methyl esters derived from renewable lipid sources, such as vegetable oil or animal fat, that can be used in compression ignition engines with little or no modifications. Until now, the use of a homogeneous alkaline transesterification chemical process with methanol has been initially chosen to address the biodiesel production [29,30,31,32]. In fact, biodiesel is, to date, the liquid biofuel produced at a greater quantity, due to the simplicity of its chemical process and its rheological properties, like fossil diesel [33,34,35]. In addition, it can be produced from different feedstocks, depending on the availability of the crop in the region. Among other advantages, biodiesel exhibits biodegradability, non-toxicity, renewability, a high cetane number, a high flash point, and its high oxygen content allows for its complete combustion in engines, reducing the amount of particulate matter, hydrocarbons, and gases, such as carbon monoxide (CO), CO2, and sulfur oxides (SOx). Furthermore, biodiesel has a very low sulfur content and very low aromatic components, as well as other pollutant emissions. Nevertheless, a slight increase in nitrogen oxide (NOx) emissions is usually described in comparison to diesel fuel [36,37,38]. Due to the high flashpoint that biodiesel exhibits, at around 150 °C, it is very safe for transportation and storage [39,40,41]. In addition, biodiesel perfectly fits into existing engines without any modification and it can be used in its pure form or blended with petroleum-based fuels without modification of existing engines or with only minor modifications [42,43,44,45,46,47]. Moreover, biodiesel exhibits better lubricant properties than fossil diesel, which allow for the extension of the engine life, and also allow for a reduction in carbon dioxide emissions by 78% in comparison with fossil diesel. In addition, the biodegradability of biodiesel is certainly high, ranging from 80.4% to 91.2% after 30 days, whereas the biodegradability of fossil diesel is only 24.5% [48]. Taking into consideration all of the advantages abovementioned, it is understandable that biodiesel has become a research hot spot during the last years, resulting in an increase in scientific publications and patents [49], as can be seen in Figure 2. Thus, for only microalgae biodiesel production, more than ten thousand patents have been published in the last 20 years [50,51,52]. Furthermore, in the last twenty years, almost forty-four thousand articles have been published, producing a growing increase year after year, demonstrating the growing interest in the problem.
Despite these efforts made to achieve better processes, better catalysts, and better sources of raw materials, it is currently concluded that the generation of glycerol represents a barrier that is difficult to overcome for the industrial production of biodiesel [53]. An alternative could be the reduction in the production cost of biodiesel. Nevertheless, the biodiesel industry strongly depends on the cost of the feedstock employed as a raw material. Despite the fact that some feedstock, such as non-edible oil and waste cooking oil, can be obtained at a good cost, they usually need a higher cost in their manufacture processing to produce standard-quality biodiesel [54,55,56]. The true magnitude of this problem has been proven in all of its consequences when the industrial-scale production of biodiesel has begun in the last three decades. The management of the huge amounts of wastes, where glycerol is the main component, is a problem with a very difficult solution [57,58], and there are still no industrial processes capable of integrating the enormous amount of glycerol. Furthermore, this glycerol obtained as a by-product also exhibits a very low quality, since it is in a mixture with other products, such as methanol, water, salts, and some amounts of monoglycerides [59].
Therefore, for being employed as a biofuel, the biodiesel obtained must be cleaned and separated from these by-products. The additional cleaning process is usually carried out by successive washing steps with water, so that a large consumption of water, energy, and time to obtain the glycerol elimination is required in order to obtain the limits established by the quality standard EN 14214 and the ASTM D6751, which are the European and the American ones, respectively [60,61,62,63,64]. These limits establish that the amount of glycerol should not exceed the 0.02% in the refined biodiesel in order to prevent its reaction with oxygen at high temperatures inside the engine, which would either produce acrolein or would polymerize generating deposits of carbonaceous compounds on the injector nozzles, pistons, and valves in the engines, consequently reducing the efficiency of the engine and its service life [65,66,67]. Therefore, it is clear that the industrial production of biodiesel requires a very complex design in order to avoid the presence of glycerol in the final biofuel [68], as is shown in Figure 3. In summary, the transesterification reaction is usually carried out in a batch reactor under constant stirring at 60 °C. Then, glycerol is separated together with the excess of methanol by decantation. Then, methanol is recovered by distillation. This crude biodiesel contains catalyst residues that must be neutralized and eliminated.
As aforementioned, biodiesel must be subjected to several washing steps with water, although the purification process also requires a drying process in an evaporator to remove held residual water [60]. Alternatively, the purification of biodiesel may also be obtained by ultrafiltration and dry washing, employing fumed silica sorbent, molecular distillation, organic resins, and biomass-based adsorbent or starch and cellulose as adsorbents of impurities [69,70,71,72,73,74]. This vast number of studies devoted to obtaining methodologies that are economically viable show that this step is one of the main factors that lead to an unprofitable biodiesel production [75].
Consequently, there is not a practical solution for the problem associated with the destabilizing glycerol price in the global market, since there are no industrial processes capable of adsorbing the increasing glycerol production [76,77]. To minimize this problem, multiple investigations are being carried out in order to valorize this crude glycerol [78,79,80]; see Figure 4.
Another element of vulnerability associated with the production of conventional biodiesel is related to the low atomic yield (or atomic efficiency) of the process. The atom yield is an important concept in green chemistry, and is far from the concept of chemical yield. In fact, a high-yielding process can still result in a substantial quantity of by-products, as is the case for biodiesel production. These green metrics are crucial for determining the sustainability and environmental impact of biodiesel production [81,82,83].
In summary, it is commonly accepted that the greatest contribution to determining the cost of biodiesel is determined by the price of feedstock, which occupies 70% of the biodiesel production cost [84,85,86]. Therefore, independently of looking at increasing the sources of triglycerides, available at appropriate prices for their transformation into biofuels, by optimizing the parameters influencing the production process of biodiesel, costs could be reduced by up to 30%. In addition, savings could be obtained by avoiding the management of residual glycerol, obtained together with conventional biodiesel, as well as the increase of at least 10% of the final product, if it is no glycerol is generated as a by-product. Thus, the search for different renewable biofuels integrating glycerol is still encouraged, while also avoiding several collective drawbacks, such as being energy-intensive, tedious in recovering glycerol, difficult in removing the acid or base catalyst from the product, the further treatment of alkaline wastewater, and the interference of free fatty acids and water in the reaction [87].

3. Green Diesel Fuels

Taking into account the complexity of the processes needed to perform the biodiesel purification, as well as the final management of the glycerol generated, the thermal conversion of fats and oils can constitute an alternative route for obtaining diesel-like biofuels, usually called green diesel, renewable diesel, or green fuels [25]. Green diesel is obtained directly from natural oils and fats via the UOP/Eni EcofiningTM process, providing a significant reduction in greenhouse gas emissions (GHGs) [88]. The production process implies a deoxygenation step and can be applied to different feedstocks, e.g., vegetable oils, animal fats, fatty acids, and waste cooking oils. Usually, the deoxygenation reaction occurs in the liquid phase following three pathways: decarboxylation, decarbonylation, and hydrodeoxygenation [89]. In this way, very similar hydrocarbons to those from crude oil are obtained when subjecting these compounds to different thermal treatments, such as cracking or pyrolysis, involving changes in the chemical structure of the triglycerides [90].
The thermal treatment can be also conducted in the presence of a catalyst, i.e., catalytic hydrocracking. This method requires an acid catalyst and a free-oxygen atmosphere. This method consumes less thermal energy and also produces a lower amount of coke that diminishes the catalyst deactivation by pore blockage and/or the catalytic poisoning. An important variation of the hydrocracking process consists of the hydroprocessing of triglycerides and petroleum gas oil simultaneously, by their co-processing in the same hydrotreating unit. In this section, we will focus on these processes.
Nowadays, the commercial plants for the production of renewable diesel have been installed all over the world. Currently, over 5.5 billion liters of renewable diesel is produced globally and is forecasted to grow up to 13 billion liters in 2024. Neste is the most important petroleum refining company, although some other oil and gas companies, such as ENI and Total, are also producing a significant amount of renewable diesel [91]. The current scenario of green diesel production worldwide is shown in Figure 5.

3.1. Pyrolysis or Cracking

In general, this technique involves a destructive distillation process that produces a change in the chemical structure of the compound through the irreversible breaking of chemical bonds, leading to smaller molecules. The thermal pyrolysis can occur either by the application of heat or in the presence of a catalyst (catalytic pyrolysis). The irreversible reaction is highly endothermic, requiring high heat transfer rates. These conditions can be reached by burning a fraction of the products to produce the thermal energy required for the reaction.
These processes are usually performed in an inert atmosphere, in a temperature range of 573–1573 K. Three different pyrolysis methods have been described, according to the reaction temperature, that influence the yield and reaction times. Conventional pyrolysis takes place in the temperature range of 550–900 K; the Fast pyrolysis in the range of 850–1250 K; and the Flash pyrolysis occurring within the temperature range of 1050–1300 K. Slow pyrolysis is also employed, although it is conducted under lower temperatures, taking a longer time to yield appropriate bio-oils [92].
The pyrolysis liquids obtained from different raw material containing triglycerides have different properties and their characteristics are strongly dependent on the reactor type used, the temperature employed, and the operational conditions. The green oil obtained through the pyrolysis of biomass can be used not only as biofuel but also as a raw material to synthesize other value-added chemicals. Furthermore, with this technique, low-quality oils and fats can be employed, contrary to what occurs in the transesterification process. For instance, an environmentally friendly renewable feedstock such as microalgae can be processed into an array of products via pyrolysis, yielding useful chemicals such as light olefins, alkanes, syngas, and biochar, as well as the bio-oils with less oxygen, more hydrocarbons, and higher gross heating values than the bio-oils derived from cellulosic biomass [93,94]. Even sewage sludge has been used to produce green diesel through a pyrolytic process [95].
Despite the cost of raw materials being around 80% of the total cost of bio-oil production, the pyrolysis method is much simpler and less expensive than other methods in producing green diesel, since it is carried out at atmospheric pressure and does not need hydrogen as a co-reagent, as abovementioned [25,96]. In summary, to produce green diesel through triglycerides, any of the three types of pyrolysis have lower costs than conventional transesterification. In addition, a wide variety of raw materials can be employed, including microalgae or waste materials, such as waste cooking oils [97,98,99,100].

3.2. Catalytic Cracking or Deoxygenation

Diesel-like hydrocarbons can also be obtained by the triglyceride catalytic cracking, comprising the deoxygenation (DO) or elimination of the oxygen atoms, obtaining hydrocarbon molecules with a lower molecular weight than the original molecules. Figure 6 shows a basic scheme of the triglyceride thermal cracking process in the presence of a catalyst. Three different fractions are generated during the process: the solid fraction is usually called coke, the liquid fraction is called bio-oil, and the gaseous stream is known as biogas. In this respect, the temperature and residence time are the key factors for this process.
Catalytic cracking occurs via a multiple-reaction process including steps such as dehydration, decarboxylation, and decarbonylation. This process produces a liquid fuel with great stability, fantastic combustion properties, and a suitable viscosity for being employed in current combustion engines. In addition, the catalytic process can be performed either in batch or continuous within a favorable temperature range of 300–450 °C [101]. The catalytic deoxygenation is carried out via hydrodeoxygenation, decarboxylation, and decarbonylation reaction pathways. Thus, oxygen is separated from the fatty acid structure of vegetable oils as H2O, CO2, and CO, respectively. The reaction pathways for fatty acid deoxygenation include both liquid and gas phase reactions. The liquid phase reactions consist of direct decarboxylation and decarbonylation, both occurring simultaneously. On the other hand, the deoxygenation of saturated fatty acids involving H2 occurs via indirect decarboxylation and direct hydrogenation for the production of n-alkanes. The deoxygenation of fatty acids via decarboxylation/decarbonylation in the liquid phase produces CO2, CO, H2, and H2O, which further undergo gas phase reactions, such as the methanation of CO2 and CO and water-gas-shift reaction, as is shown in Figure 7 [102]. Decarboxylation/decarbonylation results in an advantageous process for the production of diesel-like hydrocarbons in comparison to hydrodeoxygenation because neither H2 atmosphere nor any sulfide metal catalysts are employed.
Thus, it has been described that catalytic cracking allows for the efficient use of vegetable oils as biofuels through the deoxygenation of triglycerides. In addition, as can be seen in Figure 8, hydrodeoxygenation processes are obtained even in the absence of a hydrogen atmosphere [103]. These renewable hydrocarbon blends are chemically analogous to fossil petroleum-based fuels, having a good distribution (>32% bio-gasoline, >50% green diesel, and <11% heavy fraction). Moreover, they can be fractionated and used in different formulations depending on the types of desired fuels [104]. As the composition of the products may vary due to the solid catalyst used, many investigations evaluate the use of several catalysts. In addition, the coke formation limits the use of heterogeneous catalysts due to deactivation or poisoning, so this requires an additional regeneration process for its reuse, making the entire process very complex. On the other hand, liquid biofuels, depending on their energy density, have fundamental importance in the final energy consumption. In this way, most of the research is being conducted to maximize the amount of liquid bio-products.
In this respect, many different solid catalysts have been described, including oxides, zeolites, mesoporous materials, and their composites, as well as commercial FCC catalysts [104,105,106,107,108,109]. Interestingly, the deoxygenation paths can be regulated by altering the synthesized methods of the same catalysts [110].
Among the catalysts tested, metal-supported mesoporous materials with a small particle size have been considered as optimum catalysts for the production of diesel-like hydrocarbons via the deoxygenation of fatty acids [89,111,112,113,114,115]. Furthermore, this process exhibits a higher selectivity to diesel-like hydrocarbons than that to fatty acid ester. The deoxygenation process, catalyzed by noble-metal-supported catalysts and employing a small amount of H2, is recommended to obtain a higher yield of diesel-like hydrocarbons than that employing atmospheres rich in hydrogen, due to the formation of coke being favored at those conditions [111]. In the absence of hydrogen, the deoxygenation of triglycerides usually gives CO2, CO, and H2O, where, through the deoxygenation reaction paths, i.e., decarboxylation yielding CO2 and decarbonylation yielding CO, it results in a partial loss of the carbon amount contained in the triglyceride feedstock [116,117]. However, hydrodeoxygenation, yielding H2O, can convert most of the carbon content in the feedstock to hydrocarbons [118,119,120].
Regarding the production cost of green diesel, it is assumed that it is highly dependent on the synthesis procedure. The removal of oxygenate-bonded compounds via deoxygenation under a hydrogen-free atmosphere is more economic than hydrodeoxygenation and pyrolysis. The reason why hydrodeoxygenation is costly lies in the high consumption of hydrogen during the process. Regarding this part, pyrolysis has a lower cost, although the hydrocarbon product is mainly composed by light fractions. In addition, olefins could also be problematic, because they are associated with a lower stability due to the potential formation of gums or insoluble materials. Thus, to saturate the double bonds, the hydrorefining process or direct hydrocracking could be also an option.
In summary, despite the fact that the industrial application of pyrolysis and/or catalytic cracking still has some obstacles to overcome, the current refineries can be suitable for upgrading these green fuels production.

3.3. Catalytic Hydrocracking or Hydrodeoxygenation

In addition to the pyrolysis or cracking processes, hydrocracking or hydrodeoxygenation constitutes a very suitable methodology to convert fats and oils into fuels similar to fossil diesel [118]. One of the main strengths of this procedure is that it can be carried out in the currently existing oil refineries [121,122,123]. Some results indicate that green diesel production by catalytic hydroprocessing, located in a petroleum refinery, appears to be the most cost-effective option compared to conventional biodiesel [124].
In this procedure, hydrocarbons, mainly n-paraffins, are obtained from triglycerides. The reaction temperatures range from 300 to 450 °C and the hydrogen pressures are above 3 MPa. Furthermore, CO, CO2, and water are obtained as by-products. The hydroprocessing of triglycerides involves the hydrogenation of the double bonds of the side chains, the hydrogenation of the double bonds of the fatty acids, and the removal of oxygen in esters bonds [125]. The reactions involved in hydroprocessing can be classified into two groups: (a) hydrotreating and (b) hydrocracking. The hydrotreating of vegetable oils leads to C15–C18 hydrocarbons, which is so-called “green diesel”, “renewable diesel”, or “bio-hydrogenated diesel”.

3.3.1. Hydrotreating

As can be seen in Figure 9, the first step consists of the hydrogenation of double bonds of triglycerides. The removal of oxygen in the form of CO2 (decarboxylation) and H2O (dehydration) occurs in the next step. This suggests that the hydrodeoxygenation requires a large amount of hydrogen due to the additional hydrogenation of double bonds existing in the triglycerides, attaining the total hydrogenation, which yields hydrocarbons and water as the only reaction products, since all oxygen atoms are eliminated as water.
In the last decades, several heterogeneous catalysts capable of transforming vegetable oils into alkanes through a hydrotreating process have been described [127]. These hydrocarbons exhibit boiling points in the range of gasoline or diesel and, therefore, they can be used as fuels without any modification. These green fuels can be classified as naphtha, jet fuel, and diesel; see Figure 10 [90]. Green fuels are obtained from triglycerides by using the same process currently used in the hydrotreating of vacuum gas oil [128,129,130,131,132]. Thus, there is an increasing interest in developing the best catalytic systems, as well as the most favorable operating conditions with the most favorable techno-economic conditions, taking advantage of the facilities currently existing in refineries [133,134,135,136,137,138,139].

3.3.2. Hydrocracking

Regarding the hydrocracking or (hydro)decarboxylation process, all of the oxygen atoms in the triglyceride molecules are eliminated as carbon dioxide, so only hydrocarbons with odd carbon atom numbers can be obtained from the fatty acids. A hydro-prefix is employed to point out that hydrogen is involved in the reaction. Thus, it has been proposed that hydrogen is needed to break the fatty acid moiety loose from the triglyceride. Once the fatty acids are released, they undergo a subsequent decarboxylation step to yield hydrocarbon and CO2 [140]. Both reaction pathways are schematically depicted in Figure 11.
In contrast to that occurring in the hydrotreating process, where the reaction occurred on the metal acid sites, in the hydrocracking process, the reaction is carried out on acid sites of amorphous supports. Therefore, by modulating the balance between the acidic and hydrotreating centers in the catalyst, it is possible to achieve hydrocarbons with a boiling point in the range of jet fuel or gasoline. Moreover, the acidity of catalysts also increases the isomerization degree of the molecules, thus boosting the properties of the green liquid fuels, such as a lower pour point and a higher octane number.
Accordingly, various studies have shown that renewable liquid alkanes can be obtained by treating mixtures of vegetable oils and heavy vacuum gas oils (HVO) in hydrogen streams over conventional catalysts employed in the same hydrocracking units. These liquid hydrocarbons exhibit appropriate characteristics, such as a low acidity index, adequate density, viscosity, and a high cetane index. In addition, different by-products are obtained during the process, mainly consisting of hydrogen (H2), CO, CO2, oxygen (O2), nitrogen (N2), hydrogen sulfide (H2S), methane (CH4), and C2–C6 hydrocarbons. Despite quite a lot of by-products being obtained, the yield of liquid hydrocarbons is significantly high, at around 80% by weight [141,142,143,144,145,146,147]. In summary, the fossil fuel hydrotreatment process, initially developed to reduce the diesel sulfur amount to fulfill the specification of each country, has also been applied for obtaining a high-quality diesel biofuel from vegetable oils or animal fats [148,149,150,151,152,153].
Considering all of the advantages provided by the application of cracking treatments of vegetable oils, the research in this field has grown in recent decades, as can be seen in Figure 12. It is worth noting the growing relevance in the application of cracking techniques to produce biofuels for use in aviation [137,154,155,156,157,158,159,160,161,162]. Thus, the environmental impact and the dependence on fossil fuels in the aeronautical sector have promoted the demand for alternative and greener fuels. In this respect, while, in road transport, several biofuels or electricity can be used, in aviation, only high-quality paraffinic biofuels can be currently considered. Thus, biomass must be transformed into hydrocarbons that are fully compatible with the existing fossil fuel systems, so the implementation of renewable alternative green fuels is currently the main challenge for this sector. This current interest is shown in the high number of publications devoted to research on aviation green fuels production processes, as is also collected in Figure 12.
At present, short-term solutions, such as blending biofuels with jet fossil fuel, are gaining strength. With this in mind, the obtention of greener fuels that allow for a reduction in greenhouse gases and pollutant emissions without any significant changes in the existing fleets of the aviation companies is gaining strength. In fact, green diesel has the same chemical properties as fossil diesel, so it can be used in the current tanks, pipelines, trucks, and pumps without important infrastructure changes. Accordingly, during the last years, several companies supplying aviation fossil fuels have shown interest in the hydroprocessing of lipid feedstocks to produce renewable green liquid fuels. However, some technologies have already started to be commercialized for this purpose.
In this respect, UOP Honeywell Co. (Chickasaw, AL, USA), a refining technology company, is offering an alternative process to obtain green fuels from various vegetable oils and fats, consisting of converting the non-edible vegetable oils to green diesel [163]. Haldor Topsøe has also developed a proper hydrotreating technology, designated HydroFlex, to produce renewable fuels such as green diesel and jet fuel from non-edible raw grease material [164].
Similarly, many other hydroprocessing plants around the world are currently boosting jet biofuels production with several companies, such as SkyNRG Fly Green Fund in the Nordics, Project Solaris in South Africa, Initiative Towards sustainable Kerosene for Aviation (ITAKA) project in Europe, and Commercial Aviation Alternative Fuels Initiative (CAAFI) Petrobras, ConocoPhillips, Haldor Topsøe, or BP [165]. In general, they all have adopted a co-processing method, where the biofuel is directly mixed with petroleum feedstocks. In this way, after the hydrotreating of the mixture, a fuel product with a lower sulfur content is obtained, reducing the carbon footprint.
Likewise, various studies based on the technological preparation of biofuels have shown that the hydroprocessing of fatty acids and esters (HEFA) route has many advantages with regard to the production facilities and final properties of the biofuels, being one of the four alternative fuels approved in the ASTM D7566, with a maximum blending proportion allowed of 50% [166].
Despite the fact that the production costs of aviation biofuels are higher than conventional fossil fuels, limiting their use on a commercial scale, there is a great number of companies producing renewable fuels with a growing market, so it seems that biofuel for aircraft has a promising economic future [167].
For a better comparison between green diesel, biodiesel, and fossil diesel, some of the most important data have been collected in Table 1.

3.4. Environmental and Economic Impact of Green Diesel

Considering the green diesel emissions in internal combustion engines, the company Neste has reported the results of exhaust emission tests that have been performed in trucks, buses, and also in passenger cars, some with neat green diesel and some with a blend of 85% green diesel and 15% petroleum diesel. The results showed that those engines operating with green diesel reduced the emissions of CO, CO2, unburned hydrocarbons (HC), nitrogen oxides (NOx), and solid particulates [90,170]. CO emissions were, on average, 27, 38, and 45% lower than the EN 590 levels in the case of trucks and buses. The reduction in CO usually means higher CO2 due to a more complete combustion, but Neste claims that CO2 emissions were also reduced due to the higher H/C atomic ratio of the green diesel fuel. Therefore, most of the communications published to date manifest a reduction in greenhouse emissions using green diesel in comparison to those obtained with biodiesel or fossil diesel.
Regarding the economic issues, despite the fact that most of the papers are focused on the chemistry and engineering behind these (bio)fuels, companies are already seeing the economic benefits of switching to renewable fuels. According to a recent study [168], a green diesel fuel tested in a Class 8 truck reduced the lifecycle emission by 66%, saving 1217 t of carbon after one million miles. According to the authors, approximately USD 0.021/mi—USD 0.015/mi are saved from reduced exhaust replacement parts and downtime spent clearing diesel particulate filters (DPFs), USD 0.005/mi from a 75% oil cost reduction, and the remainder resulting from reducing the amount of diesel particulate filters (DPFs) required.
Despite these data seeming insignificant, saving USD 0.021/mi using renewable diesel, considering that there are around three million of these trucks just in US roads, each truck would have saved an average of USD 1317.77/year, or USD 5.15 B/year in savings. In terms of emissions reductions, if all Class 8 trucks had used this biofuel, it would have saved more than 297 million metric tons of carbon dioxide (CO2) per year according to the U.S. Energy Information Administration.

4. Biodiesel-like Biofuels

Another emerging alternative to obtain conventional biodiesel, with a higher atom efficiency than transesterification, consists of obtaining some derivatives of glycerol in the same transesterification process, i.e., the glycerol is integrated into the blend as a derivative. Thus, the atomic yield increases to 100%, and, above all, the separation and cleaning of residual glycerol dissolved in the FAME mixture are avoided. To achieve this, some acyl acceptors (ester molecules) instead of short chain alcohols must be employed. Hence, the corresponding glycerol ester is obtained together with the FAME (or FAEE). The reaction products are constituted by lipophilic compounds completely miscible with fossil fuels, attaining a new biofuel that is very similar to biodiesel, but avoiding the presence of free glycerol, as aforementioned [171,172].
It is also interesting to note that these processes that avoid the production of glycerol, using other acyl donor compounds instead of methanol, can be performed with the same catalysts as those employed in the conventional transesterification processes, such as several acid or alkali catalysts, under homogeneous or heterogeneous conditions, as well as with several free or immobilized lipases, or without a catalyst, under supercritical conditions.

4.1. Biodiesel-like Biofuels Integrating the Glycerol as Glycerol Triacetate

To improve the biodiesel manufacturing process, the Industrial Chemistry Research Institute patented a new type of biofuel by the interesterification of triglycerides with methyl acetate in the presence of a strong acid catalyst [173]. The reaction products consist of a mixture of FAMEs and glycerol triacetate (triacetin) and was called Gliperol (Figure 13) [173,174,175].
After that, several studies optimized the reaction conditions, i.e., oil/methyl acetate molar ratio (from 1:3 to 1:9), reaction temperatures (from 40 to 200 °C), etc. The results obtained have shown that Gliperol exhibits similar fuel properties to biodiesel, although the process itself improves not only the yield but also factors such as the efficiency and the economic feasibility. Thus, the production cost is around 30–35% lower than the biodiesel production cost, but the incorporation of glycerol in the mixture also reduces the ecological costs associated with the biodiesel purification and processing. In addition, the combustion of Gliperol improves the greenhouse gases emissions [176].
Likewise, different types of catalysts have been investigated, from homogeneous basic catalysts, such as potassium hydroxide, potassium methoxide, and polyethylene glycolate, etc. [177,178,179,180,181,182,183,184], to different heterogeneous catalysts [185,186,187,188,189,190,191,192,193,194]. In addition, the use of lipases as biocatalysts, in solvent-free systems [195,196,197,198,199,200,201,202], in ionic liquids [203,204,205,206,207], supercritical conditions [208,209,210,211,212,213,214,215,216,217,218,219], or ultrasound-assisted interesterification has also been studied [183,208,220,221,222,223,224,225,226,227].
Another possibility is the use of ethyl acetate as an acyl acceptor instead of methyl acetate. In this case, triacetin and a mixture of ethyl esters of fatty acids or FAEE are obtained. This biofuel has been considered as a biofuel-like Gliperol, although with a greener character, since ethyl acetate has a renewable character. However, despite this acyl acceptor exhibiting a similar behavior to methyl acetate in the interesterification process, it is still less studied [228,229,230,231,232].
According to various studies, the presence of triacetin improves the biodiesel behavior, since triacetin act as an anti-knocking additive when it is used along with biodiesel in diesel engines, improving the performance and reducing pollutant emissions [233,234,235,236,237,238,239,240,241,242,243,244,245,246]. Other studies have shown that a 10% by weight of triacetin in triacetin/biodiesel or triacetin/diesel blends exhibited the best results in the engine [247,248,249,250,251,252]. In addition, the triacetin also exhibits a positive effect on the cloud point and cold filter plugging point [253,254,255,256,257,258,259].
Regarding the cost analysis, the fact that Gliperol obtains the triacetin in the same process of FAMEs synthesis makes this product as viable from a technical and economic point of view, according to many studies that support the economic viability of the interesterification of triglycerides [260,261,262,263,264,265,266].
Therefore, it can be concluded that Gliperol exhibits clear techno-economic advantages compared to conventional biodiesel, being a suitable methodology to obtain a biofuel with a certain amount of a well-recognized additive, the triacetin, that improves the quality of biodiesel and practically meets the quality standards ASTM D6451 and EN 14214.

4.2. Biodiesel-like Biofuels Integrating the Glycerol as Glycerol Carbonate

Regarding the use of dimethyl carbonate (DMC) as an acyl acceptor, FAME blends with glycerol esters from lipids can be obtained, yielding alternative co-products in biodiesel solutions. This reagent is especially attractive since it is cheap, neither toxic for human health nor for the environment because it is neutral and odorless, non-corrosive, non-toxic, inexpensive, an effective non-flammable solvent, and less toxic than methanol [267,268,269]. In addition, the reagents to produce DMC, methanol, and carbon monoxides can be obtained from the synthesis gas [270,271]. These facts make DMC one of the most widely used reagents in fine chemistry. Among all of the options available for DMC, the production of glycerol carbonate has gained great attention, since glycerol carbonate is an added-value product that can be employed as an additive to fossil diesel but also allows for the valorization of glycerol [272,273,274,275,276,277,278,279]. However, the use of the residual glycerol obtained in the synthesis of biodiesel does not suppose an adequate solution to the main problem that this glycerol generates, which focuses on the need to intensely clean the biodiesel produced, which must contain a maximum of 0.04% of this impurity, according to standard N14014.
On the contrary, a biofuel produced with DMC and vegetable oil or fats as raw materials must be considered as an alternative to biodiesel because it is totally derived from renewable resources, improves the atomic yield, and avoids the formation of glycerol. Thus, through an interesterification process in which dimethyl carbonate operates as an acyl acceptor, a new biodiesel-like biofuel integrating glycerol carbonate is directly produced, whose abbreviation is DMC-BioD; see Figure 14 [280]. However, FAME and DMC are not the only products obtained in the interesterification of triglycerides with DMC, since cyclic esters of glycerol carbonate fatty acids molecules (FAGC), a small amount of glycerol dicarbonate (GDC), and glycerol carbonate (GC) are also obtained [281]. These mixtures, including glycerol derivative molecules, exhibit physical and rheological properties that allow for their use as a biofuel.
As can be seen in Figure 13, the difference between DMC-BioD and conventional biodiesel is the presence of some amounts of fatty acid glycerol carbonate monoesters (FAGCs), glycerol dicarbonate (GDC), and glycerol carbonate (GC), together with the corresponding FAMEs that constitute biodiesel [282,283,284,285,286]. Regarding the techno-economic analysis of processes for biodiesel coproduction with glycerol carbonate, they all indicate the profitability of the process because not only is the atomic yield improved (the formation of glycerol is avoided) but also the same basic catalysts described in the production of biodiesel, such as KOH, sodium methylate, sodium hydride, some amines, or different alkaline solids, can also be employed for this procedure [287,288,289,290,291,292,293,294]. Moreover, the DMC-BioD production has also been studied under supercritical conditions [295,296,297,298], employing lipases as a biocatalyst [299,300,301,302,303,304,305,306] and even ionic liquids [307,308,309].
In summary, the substitution of methanol with DMC in order to produce DMC-Biod instead of biodiesel could simplify the synthesis process and increase the atomic yield by up to 100. In addition, all of the reaction products can be used as biofuel, not being necessary for the separation of nonreacted DMC, because it is an effective additive for diesel engines due to its high oxygen content [310,311,312,313,314,315,316,317].
With respect to diethyl carbonate (DEC) as an acyl acceptor in the substitution of DMC, it must be said that it has been much less investigated, although more analogous results than those with DMC have been obtained, since ethyl esters (FAEEs) can be successfully employed as biofuels [298,318,319,320,321,322,323,324].

4.3. Biodiesel-like Biofuels Obtained by Incorporating Glycerol as Monoglycerides in the Selective Transesterification Process of Oils and Fats

Another strategy for the preparation of a new type of biodiesel-like biofuel that also integrates glycerol in its composition has been developed in the last decade. In this case, the reaction takes place through the 1,3-regiospecific transesterification of triglycerides, generating only two equivalent molecules of FAMEs (or FAEEs), maintaining the third fatty acid as monoglyceride (MG). This method was initially applied using lipases as a biocatalyst, given its 1,3-selective character, called Ecodiesel-100® [325]; see Figure 15. At the beginning, several research studies were described using pig pancreatic lipase (PPL), either in free form [326,327] or immobilized [328,329].
This reaction allowed for a reduction in the triglycerides’ viscosity and also displayed important advantages in comparison with the conventional biodiesel production. Analogously to what happened in the obtention of the abovementioned biodiesel-like biofuels, 100% of atomic efficiency is achieved, avoiding the formation of glycerol. In addition, it is worth mentioning that the operating conditions of the enzymatic process are much softer than those employed for biodiesel production, and acidic or alkaline impurities are not generated, reducing the environmental impact of the process [330]. Moreover, a specific feature of the enzymatic process is the use of ethanol, instead of methanol, so that a blend of fatty acid ethyl esters (FAEEs), together with MGs are obtained.
Ecodiesel, like biodiesel, has excellent advantages, such as renewability, a high octane number, high flash point, good lubricity, low viscosity, high biodegradability, being able to reduce important amounts of the environmental pollutant emissions, and being able to be used in internal combustion engines without any modification [331,332,333,334]. In fact, monoacylglycerides (MG) enhance the lubricity of biodiesel, as was demonstrated by recent studies [335,336,337,338]. In addition, all of the reagents used in the synthesis of Ecodiesel remain in the final blend that constitutes the biofuel. This represents great simplicity in the synthesis process, since there are no residues or final refining process of biofuel, as well as a total atomic yield of 100%, which provides a cost reduction that results in the technical and economic feasibility of the process.
The main handicap of this procedure is related to the economic difficulty of carrying out this process on an industrial scale, because of the high price of pig pancreatic lipases, even in immobilized form on an inorganic support. That is why, in recent years, a high number of low-cost lipases, both in free and immobilized form, have been tested, demonstrating the technical efficiency of the process [339,340,341,342,343,344,345,346,347,348,349], but without solving the economic difficulties associated with the high cost of the different lipases investigated.
To overcome this problem, some heterogeneous catalysts have also been investigated, e.g., KF [350] or CaO [351,352], as well as homogeneous alkaline catalysts such as sodium methoxide, operating at low enough temperatures, which contribute to a low cost of the procedure [353]. Nevertheless, in order to reach the same selectivity as that obtained with enzymatic catalysis, it is necessary to operate under the kinetic control of the chemical process, i.e., a less basic catalyst than the alkali metals must be employed and/or operate under softer reaction conditions, a lower temperature, and/or a lower concentration of ethanol. In addition, in these cases, methanol can also be used, obtaining the FAMEs blend, together with monoglycerides. According to the results, the weaker surface basic sites of heterogeneous catalysts are strong enough to perform the transesterification of primary alcohols (such as those in positions 1 and 3 of glycerol) but are not enough strong to obtain the methanolysis of secondary alcohols (such as those in positions 2 of glycerol), which is much more difficult to achieve [350,351,352].
Regarding the behavior of this biofuel in the engine, results obtained with 30% of Ecodiesel in diesel/Ecodiesel blends have shown that there are no differences with pure fossil diesel in terms of power generation, attaining a significant reduction in the emission of pollutants, at around 40% less, although a small increase in fuel consumption was obtained [353].
In summary, Ecodiesel can be obtained with similar catalysts to those conventionally employed in transesterification but operating at much softer conditions. In addition, there is no need to perform any additional purification operation, since glycerol is not produced.
Table 2 collects the technical and economic advantages that the three processes reviewed to obtain a comparison between biodiesel-like biofuels and biodiesel.

5. Straight Vegetable Oils (SVO) Blending with Less Viscous and Lower Cetane (LVLC) Biofuels

Pure vegetable oils (SVO) could constitute an ideal alternative to fossil diesel due to the similar properties that they exhibit. In addition, SVOs are renewable, non-toxic, biodegradable, and do not contain sulfur, which could make them very suited to replace fossil diesel. In addition, vegetable oils can be obtained from agricultural or industrial sources, avoiding the costs associated with the transesterification to obtain Biodiesel. The main drawback in the use of SVO is their high viscosity, which dramatically alters fuel spray characteristics, atomization quality, and volatility, leading to severe carbon deposits, fuel injector clogging, and rapid wearing of fuel pump components [26].
To solve this problem, the incorporation of SVO into blends with diesel in optimal proportions could be an option. This method can operate along the current strategy of reducing the level of CO2 emissions as much as possible. Thus, a recent study supports the possibility of using 10% of vegetable oils of different sources in blends with diesel [354].
However, to achieve higher percentages of fossil diesel substitution using SVO, another method consisting of blends of SVO with other lower viscosity biofuel, with the goal of reaching viscosity values similar to fossil diesel, has been proposed. In this way, an effective and inexpensive method to obtain dual biofuels able to be blended efficiently with diesel fossil is being investigated. Thus, blends of a light biofuel a with low viscosity, based on Melaleuca Cajuputi oil (MCO) and SVO, to obtain higher substitution values of diesel fossil, have recently been proposed [355,356,357].
Since low-viscosity biofuels also have a low cetane number, they are often referred to as low viscosity low cetane (LVLC) compounds. Thus, biofuels composed of alcohol such as methanol, ethanol, or butanol, and plant-based light biofuels such as eucalyptus and pine oil, have been chosen and classified as LVLC fuels [358]. The high-viscosity values of some of these oils, e.g., castor oil (226.2 cSt) and pine oil (1.3 cSt), are balanced with the low viscosity of alcohols, achieving a suitable viscosity value for blending in different proportions and attaining a final viscosity in the range of 2.0–4.5 cSt, which are the values imposed by the EN590 normative. However, these blends exhibit a lower cetane number, so a limit usually exists in which they can be blended in order to be correctly employed in an engine [27]. Analogously to pine and castor oil, many other compounds, such as eucalyptus oil, orange oil, or camphor oil, have been studied as LVLC fuels for use in double or triple blends with biodiesel, or even with fossil diesel [27,359,360,361,362,363,364,365,366]. These low viscous vegetable oils improved the performance of the biodiesel-fueled CI engine, described as operating as biofuels either by themselves or in blends with fossil diesel [367,368,369]. Moreover, some of these ternary blends (vegetable oil, alcohols, and fossil diesel) have demonstrate that they are able to reduce the emission pollutants [370,371,372,373,374,375,376,377,378].
In addition to vegetable oils and some other low-viscosity natural products, triple mixtures with molecules obtained by synthesis from renewable commodities are also being evaluated, so they can also be considered as renewable compounds. Thus, diethyl ether (DEE), which can be synthesized from bioethanol, has also been considered as a biofuel. DEE is a butanol isomer that has shown good properties for being blended with diesel, which are as follows: a high cetane number, reasonable energy density, high oxygen content, low autoignition temperature, broad flammability limits, and high miscibility with vegetable oils and diesel fossil. In addition, DEE has been reported as a low-emission renewable fuel and high-quality combustion improver when it is used in blends with diesel fossil and several vegetable oils [366,379,380,381,382,383,384,385,386,387,388]. The incorporation of DEE in the triple blend, diesel/vegetable oil/DEE, allows for the substitution of fossil diesel by up to 40% by volume, improving the engine power with fewer emissions, as well as improving the cold flow behavior of fuel.
As with DEE, the possibility of using acetone (ACE) in triple blends with fossil diesel and straight vegetable oils as alternatives has also been investigated. ACE is an oxygenated additive that also complies with the requirements for being blended with vegetable oils and fossil diesel, i.e., acetone exhibits a very low kinematic viscosity that balances the high viscosity of SVOs. Likewise, the oxygen content, the low autoignition temperature, and the very low cloud point and pour point values, make acetone a good candidate for being employed in triple blends. On the other hand, although acetone is currently produced from fossil resources, it can also be obtained from renewable resources, either from ethanol [25] or from cellulose through a typical acetone–butanol–ethanol (ABE) fermentation process. Regarding the results obtained in the engine with a triple blend of diesel/vegetable oil/acetone, in which, acetone is at around 16–18% by volume, a considerable reduction in emissions of air pollutants, as well as a good power engine, were attained. Nevertheless, the fuel consumption was slightly higher than with fossil diesel [389]. Acetone has also been described as a fuel additive in biodiesel–diesel blends [390].
Another organic compound object of study for these purposes is the ethyl acetate (EA). It has been tested in blends of diesel/sunflower oil/EA or diesel/castor oil/EA. The results obtained indicated that triple blends composed of up to 24% of EA in the case of sunflower oil and 36% of EA in the case of castor oil allow for the substitution of 60–80% of fossil diesel, providing engine power values that are very similar to conventional diesel [391]. In addition, the EA properties make it a fuel that is very safe for transportation [392,393,394,395,396].
Diethyl carbonate (DEC) has also been reported as an effective oxygenated additive, lowering soot and NOx emissions with an improvement in the engine performance [397,398,399,400,401,402]. The main advantage of DEC is that it can be synthesized from bioethanol and CO2, which could contribute to the reduction in the atmospheric CO2 to a large extent [403]. Very recently, DEC has been addressed as a solvent for vegetable oils for use in diesel engines in triple blends [404], as well as dimethyl carbonate (DMC) [405,406,407,408,409,410,411]. In addition, it has been tested in blends with biodiesel [412] and in biodiesel/diesel blends [413]. Results have shown that, in all cases, the use of DMC notably improves the engine performance and exhaust emissions from C.I. engines [413]. To move forward in the substitution of fossil fuels with others of renewable character, the strategy has even been applied in triple mixtures with gasoline. In this way, these triple blends allow for the substitution of up to 40% of fossil diesel with sunflower oil, and up to 25% with castor oil, with a significant reduction in the emission of pollutants also being obtained with these triple blends [414,415].
Therefore, we can conclude that, in just a decade, it has been possible to verify a suitable strategy to reduce the viscosity of vegetable oils to the values required by current CI engines by their blending with low-viscosity solvents (LVS). In this way, SVO/LVS/fossil diesel triple blends can be obtained with a suitable composition to comply with regulations of the EN 14,214 standard. In addition, to achieve a higher fossil diesel replacement, compounds derived from renewable sources represent the better option. On the other hand, the use of oxygen-rich compounds as viscosity reducer solvents allows for a better combustion process and reduced emissions. In this respect, up until this moment, some light vegetable oils (orange, camphor, eucalyptus, and pine oil) and higher alcohols (1-propanol, 2-propanol, 1-butanol, 2-butanol, and 1-pentanol), as well as several renewable oxygenated compounds, such as diethyl ether, acetone, ethyl acetate, diethyl carbonate, and dimethyl carbonate, have been described as viscosity reducers of SVOs. Overall, the exhaust emissions were significantly reduced with the use of these blends, resulting in a similar or slightly lower engine performance than that exhibited by conventional diesel. Moreover, the behavior of blends at low temperatures is usually improved by using these less viscous oxygenated compounds.

6. Summary and Concluding Remarks

Very recently, ambitious targets were established by the European New Green Deal and the REDII, aiming to achieve a climate-neutral Europe by 2050, with the transport sector being the most critical area to decarbonize. Given the urgency of the deadlines imposed by policy makers, it is time to determine which biofuels have the necessary maturity to be incorporated into this substitution process. In this respect, a wide range of biofuels are under development. Although some of them have been extensively commercialized, the so-called conventional biofuels (starch, sugar-based ethanol, biodiesel, etc.), some others, generally called non-conventional, are far from commercialization; in particular, biofuels produced from lignocellulosic residues (i.e., agricultural and forestry residues, e.g., straws, stoves, bagasse, woody biomass). However, all renewable biofuels are considered important for the long-run decarbonization of the transportation sector due to incompatibilities that other low-carbon fuel, such as electricity or hydrogen, exhibit when being applied to heavy-duty fleets and air transports.
In this respect, most of the research in the literature focuses on the production of biodiesel to supplement petroleum-based diesel. However, as has been discussed in the present review, biodiesel presents techno-economic difficulties that prevent its massive application on an industrial scale, mainly due to the production of glycerol as a by-product. Thus, it is understandable that transesterification is not adequate on its own to take advantage of vegetable oils as substitutes for fossil fuels in ID engines.
Therefore, in this review, only those methodologies that can be incorporated immediately into the process of substituting fossil fuels with biofuels obtained from vegetable oils are addressed. We do not consider the “drop-in” fuel production from lignocellulosic sources as viable, which will probably be an important procedure in the medium-long term, but cannot be applied immediately.
Independently of the current energy scenario, the complete replacement of petroleum-derived fuels with biofuels is practically impossible in the short-medium term, since the production of all of the raw materials needed is impossible, not only due to economic difficulties and the high prices of vegetable oils but also because of the shortage of agricultural land suitable for it. Thus, in this review, three alternatives to obtaining biofuels that do not produce glycerol have been evaluated and compared with the conventional biodiesel, as can be seen in Table 3.
In addition, these advanced methods (biodiesel-like biofuel, green diesel, and/or LVLC solvents blended with vegetable oils) could be applied without any problem from a technical point of view. It is a matter of discriminating which of them are more viable from a techno-economic point of view. Another important aspect is the need to produce adequate fuels for air transport. In this respect, hydrocracking would constitute the ideal solution, since it can simultaneously produce high-quality biojet and biogasoline.
Finally, it can be concluded that, according to the checked bibliography in this comprehensive review, any of the alternative methods proposed in Table 2 are able to compete advantageously with conventional biodiesel in order to achieve the gradual replacement of fossil fuels by some other fuels of renewable nature, operating in the car fleet currently in use. Thus, any of the selected advanced biofuels must always be within the acceptable limits prescribed by ASTM D 6751, currently in force for conventional fossil diesel fuel. However, publications from the academic world have not yet become aware of this fact. Thus, it can be verified in Figure 2 that the increase in the publications in the Web of Science database with the keywords “biodiesel” from year 2000 to 2021 grows continuously year after year, without noticing any change in this trend. This is despite the fact that, from 2015, there was a sharp increase in the number of publications dedicated to the study of so-called green diesel, as shown in Figure 12. In addition, we can verify the sudden eruption from this date, which was at the beginning of several commercial plants for the production of renewable diesel all over the world; see Figure 5. Therefore, it seems that the process of replacing fossil fuels continues, but that conventional biodiesel is no longer the chosen candidate for this process.

Author Contributions

D.L. designed the study; L.A.-D. performed the experiments and wrote the paper; R.E. reviewed and edited the paper; R.E., D.L. and F.M.B. supervised the study and revised the manuscript; C.L., F.J.L.-T., J.C. and A.A.R. made intellectual contributions to this study. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

The authors are thankful to MINECO-ENE2016-81013-R (AEI/FEDER, EU), MICIIN (Project ref. PID2019-104953RB-100), Consejería de Transformación Económica, Industria, Conocimiento y Universidades de la Junta de Andalucía (UCO-FEDER Project CATOLIVAL, ref. 1264113-R, 2018 call and Project ref. P18-RT-4822) and FEDER Funds for financial support. Technical assistance of staff at Central Service for Research Support (SCAI) and Institute of Nanochemistry (IUNAN) of the University of Córdoba is also thanked.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Obergassel, W.; Arens, C.; Hermwille, L.; Kreibich, N.; Mersmann, F.; Ott, H.E.; Wang-Helmreich, H. Phoenix from the ashes: An analysis of the Paris Agreement to the United Nations Framework Convention on Climate Change. Eur. J. Int. Law 2015, 21, 9–77. [Google Scholar]
  2. Chiaramonti, D.; Talluri, G.; Scarlat, N.; Prussi, M. The challenge of forecasting the role of biofuel in EU transport decarbonisation at 2050: A meta-analysis review of published scenarios. Renew. Sustain. Energy Rev. 2021, 139, 110715. [Google Scholar] [CrossRef]
  3. Oberthür, S. Where to go from Paris? The European Union in climate geopolitics. Glob. Aff. 2016, 2, 119–130. [Google Scholar] [CrossRef]
  4. Gota, S.; Huizenga, C.; Peet, K.; Medimorec, N.; Bakker, S. Decarbonising transport to achieve Paris Agreement targets. Energy Effic. 2019, 12, 363–386. [Google Scholar] [CrossRef]
  5. Kalghatgi, G. Is it really the end of internal combustion engines and petroleum in transport? Appl. Energy 2018, 225, 965–974. [Google Scholar] [CrossRef]
  6. Sisco, M.R.; Pianta, S.; Weber, E.U.; Bosetti, V. Global climate marches sharply raise attention to climate change: Analysis of climate search behavior in 46 countries. J. Environ. Psychol. 2021, 75, 101596. [Google Scholar] [CrossRef]
  7. Dafnomilis, I.; Hoefnagels, R.; Pratama, Y.W.; Schott, D.L.; Lodewijks, G.; Junginger, M. Review of solid and liquid biofuel demand and supply in Northwest Europe towards 2030—A comparison of national and regional projections. Renew. Sustain. Energy Rev. 2017, 78, 31–45. [Google Scholar] [CrossRef] [Green Version]
  8. Schreyer, F.; Luderer, G.; Rodrigues, R.; Pietzcker, R.C.; Baumstark, L.; Sugiyama, M.; Brecha, R.J.; Ueckerdt, F. Common but differentiated leadership: Strategies and challenges for carbon neutrality by 2050 across industrialized economies. Environ. Res. Lett. 2020, 15, 114016. [Google Scholar] [CrossRef]
  9. Gomiero, T. Large-scale biofuels production: A possible threat to soil conservation and environmental services. Appl. Soil Ecol. 2018, 123, 729–736. [Google Scholar] [CrossRef]
  10. Rocha, M.H.; Capaz, R.S.; Lora, E.E.S.; Nogueira, L.A.H.; Leme, M.M.V.; Renó, M.L.G.; del Olmo, O.A. Life cycle assessment (LCA) for biofuels in Brazilian conditions: A meta-analysis. Renew. Sustain. Energy Rev. 2014, 37, 435–459. [Google Scholar] [CrossRef]
  11. Bindraban, P.S.; Bulte, E.H.; Conijn, S.G. Can large-scale biofuels production be sustainable by 2020? Agric. Syst. 2009, 101, 197–199. [Google Scholar] [CrossRef]
  12. Demirbas, A. Biofuels securing the planet’s future energy needs. Energy Convers. Manag. 2009, 50, 2239–2249. [Google Scholar] [CrossRef]
  13. Abas, N.; Kalair, A.; Khan, N. Review of fossil fuels and future energy technologies. Futures 2015, 69, 31–49. [Google Scholar] [CrossRef]
  14. Fulton, L.M.; Lynd, L.R.; Körner, A.; Greene, N.; Tonachel, L.R. The need for biofuels as part of a low carbon energy future. Biofuels Bioprod. Biorefining 2015, 9, 476–483. [Google Scholar] [CrossRef]
  15. Mandley, S.; Daioglou, V.; Junginger, H.; van Vuuren, D.; Wicke, B. EU bioenergy development to 2050. Renew. Sustain. Energy Rev. 2020, 127, 109858. [Google Scholar] [CrossRef]
  16. Kurzawska, P.; Jasiński, R. Overview of Sustainable Aviation Fuels with Emission Characteristic and Particles Emission of the Turbine Engine Fueled ATJ Blends with Different Percentages of ATJ Fuel. Energies 2021, 14, 1858. [Google Scholar] [CrossRef]
  17. Riboldi, C.E. An optimal approach to the preliminary design of small hybrid-electric aircraft. Aerosp. Sci. Technol. 2018, 81, 14–31. [Google Scholar] [CrossRef]
  18. Witcover, J.; Williams, R.B. Comparison of “Advanced” biofuel cost estimates: Trends during rollout of low carbon fuel policies. Transp. Res. Part D Transp. Environ. 2020, 79, 102211. [Google Scholar] [CrossRef]
  19. Nogueira, L.A. Does biodiesel make sense? Energy 2011, 36, 3659–3666. [Google Scholar] [CrossRef]
  20. Ramos, M.; Dias, A.P.S.; Puna, J.F.; Gomes, J.; Bordado, J.C. Biodiesel production processes and sustainable raw materials. Energies 2019, 12, 4408. [Google Scholar] [CrossRef] [Green Version]
  21. Mohadesi, M.; Aghel, B.; Maleki, M.; Ansari, A. Production of biodiesel from waste cooking oil using a homogeneous catalyst: Study of semi-industrial pilot of microreactor. Renew. Energy 2019, 136, 677–682. [Google Scholar] [CrossRef]
  22. Estevez, R.; Aguado-Deblas, L.; Bautista, F.M.; Luna, D.; Luna, C.; Calero, J.; Posadillo, A.; Romero, A.A. Biodiesel at the Crossroads: A Critical Review. Catalysts 2019, 9, 1033. [Google Scholar] [CrossRef] [Green Version]
  23. Luque, R.; Herrero-Davila, L.; Campelo, J.M.; Clark, J.H.; Hidalgo, J.M.; Luna, D.; Marinas, J.M.; Romero, A.A. Biofuels: A technological perspective. Energy Environ. Sci. 2008, 1, 542–564. [Google Scholar] [CrossRef]
  24. Calero, J.; Luna, D.; Sancho, E.D.; Luna, C.; Bautista, F.M.; Romero, A.A.; Posadillo, A.; Berbel, J.; Verdugo-Escamilla, C. An overview on glycerol-free processes for the production of renewable liquid biofuels, applicable in diesel engines. Renew. Sustain. Energy Rev. 2015, 42, 1437–1452. [Google Scholar] [CrossRef]
  25. Abdulkareem-Alsultan, G.; Asikin-Mijan, N.; Lee, H.; Rashid, U.; Islam, A.; Taufiq-Yap, Y. A Review on Thermal Conversion of Plant Oil (Edible and Inedible) into Green Fuel Using Carbon-Based Nanocatalyst. Catalysts 2019, 9, 350. [Google Scholar] [CrossRef] [Green Version]
  26. Mat, S.C.; Idroas, M.Y.; Teoh, Y.; Hamid, M.F. An Investigation of Viscosities, Calorific Values and Densities of Binary Biofuel Blends. In MATEC Web of Conferences; EDP Sciences: Les Ulis, France, 2017; p. 4. [Google Scholar]
  27. Prakash, T.; Geo, V.E.; Martin, L.J.; Nagalingam, B. Evaluation of pine oil blending to improve the combustion of high viscous (castor oil) biofuel compared to castor oil biodiesel in a CI engine. Heat Mass Transf. 2018, 55, 1491–1501. [Google Scholar] [CrossRef]
  28. Ibarra-Gonzalez, P.; Rong, B.-G. A review of the current state of biofuels production from lignocellulosic biomass using thermochemical conversion routes. Chin. J. Chem. Eng. 2019, 27, 1523–1535. [Google Scholar] [CrossRef]
  29. Demirbas, A. Future energy sources. In Waste Energy for Life Cycle Assessment; Springer: Berlin, Germany, 2016; pp. 33–70. [Google Scholar]
  30. Mythili, R.; Venkatachalam, P.; Subramanian, P.; Uma, D. Production characterization and efficiency of biodiesel: A review. Int. J. Energy Res. 2014, 38, 1233–1259. [Google Scholar] [CrossRef]
  31. Siraj, S.; Kale, R.; Deshmukh, S. Effects of thermal, physical, and chemical properties of biodiesel and diesel blends. Am. J. Mech. Ind. Eng 2017, 2, 24–31. [Google Scholar] [CrossRef]
  32. Nigam, P.S.; Singh, A. Production of liquid biofuels from renewable resources. Prog. Energy Combust. Sci. 2011, 37, 52–68. [Google Scholar] [CrossRef]
  33. Mohiddin, M.N.B.; Tan, Y.H.; Seow, Y.X.; Kansedo, J.; Mubarak, N.; Abdullah, M.O.; San Chan, Y.; Khalid, M. Evaluation on feedstock, technologies, catalyst and reactor for sustainable biodiesel production: A review. J. Ind. Eng. Chem. 2021, 98, 60–81. [Google Scholar] [CrossRef]
  34. Raheem, I.; Mohiddin, M.N.B.; Tan, Y.H.; Kansedo, J.; Mubarak, N.M.; Abdullah, M.O.; Ibrahim, M.L. A review on influence of reactor technologies and kinetic studies for biodiesel application. J. Ind. Eng. Chem. 2020, 91, 54–68. [Google Scholar] [CrossRef]
  35. Aydın, S. Comprehensive analysis of combustion, performance and emissions of power generator diesel engine fueled with different source of biodiesel blends. Energy 2020, 205, 118074. [Google Scholar] [CrossRef]
  36. Thangaraj, B.; Solomon, P.R.; Muniyandi, B.; Ranganathan, S.; Lin, L. Catalysis in biodiesel production—A review. Clean Energy 2019, 3, 2–23. [Google Scholar] [CrossRef] [Green Version]
  37. Noor, C.M.; Noor, M.; Mamat, R. Biodiesel as alternative fuel for marine diesel engine applications: A review. Renew. Sustain. Energy Rev. 2018, 94, 127–142. [Google Scholar] [CrossRef]
  38. Sharma, A.K.; Sharma, P.K.; Chintala, V.; Khatri, N.; Patel, A. Environment-friendly biodiesel/diesel blends for improving the exhaust emission and engine performance to reduce the pollutants emitted from transportation fleets. Int. J. Environ. Res. Public Health 2020, 17, 3896. [Google Scholar] [CrossRef] [PubMed]
  39. Álvarez, A.; Lapuerta, M.n.; Agudelo, J.R. Prediction of flash-point temperature of alcohol/biodiesel/diesel fuel blends. Ind. Eng. Chem. Res. 2019, 58, 6860–6869. [Google Scholar] [CrossRef]
  40. Do Nascimento, D.C.; Carareto, N.D.D.; Neto, A.M.B.; Gerbaud, V.; da Costa, M.C. Flash point prediction with UNIFAC type models of ethylic biodiesel and binary/ternary mixtures of FAEEs. Fuel 2020, 281, 118717. [Google Scholar] [CrossRef]
  41. Hazrat, M.; Rasul, M.; Khan, M.; Mofijur, M.; Ahmed, S.; Ong, H.C.; Vo, D.-V.N.; Show, P.L. Techniques to improve the stability of biodiesel: A review. Environ. Chem. Lett. 2021, 19, 2209–2236. [Google Scholar] [CrossRef]
  42. Chandran, D. Compatibility of diesel engine materials with biodiesel fuel. Renew. Energy 2020, 147, 89–99. [Google Scholar] [CrossRef]
  43. Chidambaranathan, B.; Gopinath, S.; Aravindraj, R.; Devaraj, A.; Krishnan, S.G.; Jeevaananthan, J. The production of biodiesel from castor oil as a potential feedstock and its usage in compression ignition Engine: A comprehensive review. Mater. Today Proc. 2020, 33, 84–92. [Google Scholar] [CrossRef]
  44. Jayakumar, M.; Karmegam, N.; Gundupalli, M.P.; Gebeyehu, K.B.; Asfaw, B.T.; Chang, S.W.; Balasubramani, R.; Awasthi, M.K. Heterogeneous base catalysts: Synthesis and application for biodiesel production—A review. Bioresour. Technol. 2021, 331, 125054. [Google Scholar] [CrossRef]
  45. Alagumalai, A.; Mahian, O.; Hollmann, F.; Zhang, W. Environmentally benign solid catalysts for sustainable biodiesel production: A critical review. Sci. Total Environ. 2021, 768, 144856. [Google Scholar] [CrossRef]
  46. Vignesh, P.; Kumar, A.P.; Ganesh, N.S.; Jayaseelan, V.; Sudhakar, K. A review of conventional and renewable biodiesel production. Chin. J. Chem. Eng. 2020, 40, 1–17. [Google Scholar] [CrossRef]
  47. Chozhavendhan, S.; Singh, M.V.P.; Fransila, B.; Kumar, R.P.; Devi, G.K. A review on influencing parameters of biodiesel production and purification processes. Curr. Res. Green Sustain. Chem. 2020, 1, 1–6. [Google Scholar] [CrossRef]
  48. Syafiuddin, A.; Hao, C.J.; Yuniarto, A.; Hadibarata, T. The current scenario and challenges of biodiesel production in Asian countries: A review. Bioresour. Technol. Rep. 2020, 12, 100608. [Google Scholar] [CrossRef]
  49. Zhang, M.; Gao, Z.; Zheng, T.; Ma, Y.; Wang, Q.; Gao, M.; Sun, X. A bibliometric analysis of biodiesel research during 1991–2015. J. Mater. Cycles Waste Manag. 2018, 20, 10–18. [Google Scholar] [CrossRef]
  50. Li, D.; Du, W.; Fu, W.; Cao, X. A Quick Look Back at the Microalgal Biofuel Patents: Rise and Fall. Front. Bioeng. Biotechnol. 2020, 8, 1035. [Google Scholar] [CrossRef]
  51. Mahlia, T.; Syazmi, Z.; Mofijur, M.; Abas, A.P.; Bilad, M.; Ong, H.C.; Silitonga, A. Patent landscape review on biodiesel production: Technology updates. Renew. Sustain. Energy Rev. 2020, 118, 109526. [Google Scholar] [CrossRef]
  52. Rawat, J.; Gupta, P.K.; Pandit, S.; Priya, K.; Agarwal, D.; Pant, M.; Thakur, V.K.; Pande, V. Latest Expansions in Lipid Enhancement of Microalgae for Biodiesel Production: An Update. Energies 2022, 15, 1550. [Google Scholar] [CrossRef]
  53. Bazooyar, B.; Shariati, A.; Hashemabadi, S.H. Economy of a utility boiler power plant fueled with vegetable oil, biodiesel, petrodiesel and their prevalent blends. Sustain. Prod. Consum. 2015, 3, 1–7. [Google Scholar] [CrossRef]
  54. 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, 100904. [Google Scholar] [CrossRef]
  55. Gebremariam, S.; Marchetti, J. Economics of biodiesel production. Energy Convers. Manag. 2018, 168, 74–84. [Google Scholar] [CrossRef]
  56. Rochelle, D.; Najafi, H. A review of the effect of biodiesel on gas turbine emissions and performance. Renew. Sustain. Energy Rev. 2019, 105, 129–137. [Google Scholar] [CrossRef]
  57. Nda-Umar, U.; Ramli, I.; Taufiq-Yap, Y.; Muhamad, E. An Overview of Recent Research in the Conversion of Glycerol into Biofuels, Fuel Additives and other Bio-Based Chemicals. Catalysts 2019, 9, 15. [Google Scholar] [CrossRef] [Green Version]
  58. Quispe, C.A.; Coronado, C.J.; Carvalho, J.A., Jr. Glycerol: Production, consumption, prices, characterization and new trends in combustion. Renew. Sustain. Energy Rev. 2013, 27, 475–493. [Google Scholar] [CrossRef]
  59. Smirnov, A.A.; Selishcheva, S.A.; Yakovlev, V.A. Acetalization catalysts for synthesis of valuable oxygenated fuel additives from glycerol. Catalysts 2018, 8, 595. [Google Scholar] [CrossRef] [Green Version]
  60. Atadashi, I.; Aroua, M.K.; Aziz, A.A.; Sulaiman, N. The effects of water on biodiesel production and refining technologies: A review. Renew. Sustain. Energy Rev. 2012, 16, 3456–3470. [Google Scholar] [CrossRef]
  61. Gomes, M.G.; Santos, D.Q.; de Morais, L.C.; Pasquini, D. Purification of biodiesel by dry washing, employing starch and cellulose as natural adsorbents. Fuel 2015, 155, 1–6. [Google Scholar] [CrossRef]
  62. Squissato, A.L.; Fernandes, D.M.; Sousa, R.M.; Cunha, R.R.; Serqueira, D.S.; Richter, E.M.; Pasquini, D.; Munoz, R.A. Eucalyptus pulp as an adsorbent for biodiesel purification. Cellulose 2015, 22, 1263–1274. [Google Scholar] [CrossRef]
  63. Tajziehchi, K.; Sadrameli, S. Optimization for free glycerol, diglyceride, and triglyceride reduction in biodiesel using ultrafiltration polymeric membrane: Effect of process parameters. Process Saf. Environ. Prot. 2021, 148, 34–46. [Google Scholar] [CrossRef]
  64. Govindaraju, R.; Chen, S.-S.; Wang, L.-P.; Chang, H.-M.; Pasawan, M. Significance of Membrane Applications for High-Quality Biodiesel and Byproduct (Glycerol) in Biofuel Industries. Curr. Pollut. Rep. 2021, 7, 128–145. [Google Scholar] [CrossRef]
  65. Li, R.; Liang, N.; Ma, X.; Chen, B.; Huang, F. Free glycerol removal from biodiesel using anion exchange resin as a new type of adsorbent. Ind. Eng. Chem. Res. 2018, 57, 17226–17236. [Google Scholar] [CrossRef]
  66. Shahbaz, K.; Mjalli, F.; Hashim, M.; AlNashef, I. Using deep eutectic solvents based on methyl triphenyl phosphunium bromide for the removal of glycerol from palm-oil-based biodiesel. Energy Fuels 2011, 25, 2671–2678. [Google Scholar] [CrossRef]
  67. Reis, M.; Cardoso, V. Biodiesel production and purification using membrane technology. In Membrane Technologies for Biorefining; Elsevier: Amsterdam, The Netherlands, 2016; pp. 289–307. [Google Scholar]
  68. Karmakar, B.; Halder, G. Progress and future of biodiesel synthesis: Advancements in oil extraction and conversion technologies. Energy Convers. Manag. 2019, 182, 307–339. [Google Scholar] [CrossRef]
  69. Catarino, M.; Ferreira, E.; Dias, A.P.S.; Gomes, J. Dry washing biodiesel purification using fumed silica sorbent. Chem. Eng. J. 2020, 386, 123930. [Google Scholar] [CrossRef]
  70. Rodriguez, N.E.; Martinello, M.A. Molecular distillation applied to the purification of biodiesel from ethanol and soybean oil. Fuel 2021, 296, 120597. [Google Scholar] [CrossRef]
  71. Limmun, W.; Sansiribhan, S. Water-spray washing technique as a purification process in the production of biodiesel. In E3S Web of Conferences; EDP Sciences: Les Ulis, France, 2020; p. 3006. [Google Scholar]
  72. Sokač, T.; Gojun, M.; Tušek, A.J.; Šalić, A.; Zelić, B. Purification of biodiesel produced by lipase catalysed transesterification by ultrafiltration: Selection of membranes and analysis of membrane blocking mechanisms. Renew. Energy 2020, 159, 642–651. [Google Scholar] [CrossRef]
  73. Sandouqa, A.; Al-Shannag, M.; Al-Hamamre, Z. Biodiesel purification using biomass-based adsorbent manufactured from delignified olive cake residues. Renew. Energy 2020, 151, 103–117. [Google Scholar] [CrossRef]
  74. De Jesus, S.S.; Ferreira, G.F.; Maciel, M.R.W.; Maciel Filho, R. Biodiesel purification by column chromatography and liquid-liquid extraction using green solvents. Fuel 2019, 235, 1123–1130. [Google Scholar] [CrossRef]
  75. Suthar, K.; Dwivedi, A.; Joshipura, M. A review on separation and purification techniques for biodiesel production with special emphasis on Jatropha oil as a feedstock. Asia-Pac. J. Chem. Eng. 2019, 14, e2361. [Google Scholar] [CrossRef]
  76. Abomohra, A.E.-F.; Elsayed, M.; Esakkimuthu, S.; El-Sheekh, M.; Hanelt, D. Potential of fat, oil and grease (FOG) for biodiesel production: A critical review on the recent progress and future perspectives. Prog. Energy Combust. Sci. 2020, 81, 100868. [Google Scholar] [CrossRef]
  77. Anuar, M.R.; Abdullah, A.Z. Challenges in biodiesel industry with regards to feedstock, environmental, social and sustainability issues: A critical review. Renew. Sustain. Energy Rev. 2016, 58, 208–223. [Google Scholar] [CrossRef]
  78. Kaur, J.; Sarma, A.K.; Jha, M.K.; Gera, P. Valorisation of crude glycerol to value-added products: Perspectives of process technology, economics and environmental issues. Biotechnol. Rep. 2020, 27, e00487. [Google Scholar] [CrossRef]
  79. Kosamia, N.M.; Samavi, M.; Uprety, B.K.; Rakshit, S.K. Valorization of biodiesel byproduct crude glycerol for the production of bioenergy and biochemicals. Catalysts 2020, 10, 609. [Google Scholar] [CrossRef]
  80. Ripoll, M.; Betancor, L. Opportunities for the valorization of industrial glycerol via biotransformations. Curr. Opin. Green Sustain. Chem. 2021, 28, 100430. [Google Scholar] [CrossRef]
  81. Martinez-Guerra, E.; Gude, V.G. Assessment of sustainability indicators for biodiesel production. Appl. Sci. 2017, 7, 869. [Google Scholar] [CrossRef] [Green Version]
  82. Gude, V.G.; Martinez-Guerra, E. Green chemistry with process intensification for sustainable biodiesel production. Environ. Chem. Lett. 2018, 16, 327–341. [Google Scholar] [CrossRef]
  83. Shaheen, A.; Sultana, S.; Lu, H.; Ahmad, M.; Asma, M.; Mahmood, T. Assessing the potential of different nano-composite (MgO, Al2O3-CaO and TiO2) for efficient conversion of Silybum eburneum seed oil to liquid biodiesel. J. Mol. Liq. 2018, 249, 511–521. [Google Scholar] [CrossRef]
  84. Yan, Y.; Li, X.; Wang, G.; Gui, X.; Li, G.; Su, F.; Wang, X.; Liu, T. Biotechnological preparation of biodiesel and its high-valued derivatives: A review. Appl. Energy 2014, 113, 1614–1631. [Google Scholar] [CrossRef]
  85. Yaqoob, H.; Teoh, Y.H.; Sher, F.; Farooq, M.U.; Jamil, M.A.; Kausar, Z.; Sabah, N.U.; Shah, M.F.; Rehman, H.Z.U.; Rehman, A.U. Potential of waste cooking oil biodiesel as renewable fuel in combustion engines: A Review. Energies 2021, 14, 2565. [Google Scholar] [CrossRef]
  86. Kemp, W.H. Biodiesel: Basics and Beyond: A Comprehensive Guide to Production and Use for the Home and Farm; Aztext Press: Tamworth, Australia, 2006. [Google Scholar]
  87. Abd Manaf, I.S.; Embong, N.H.; Khazaai, S.N.M.; Rahim, M.H.A.; Yusoff, M.M.; Lee, K.T.; Maniam, G.P. A review for key challenges of the development of biodiesel industry. Energy Convers. Manag. 2019, 185, 508–517. [Google Scholar] [CrossRef]
  88. Kalnes, T.; Marker, T.; Shonnard, D.R. Green diesel: A second generation biofuel. Int. J. Chem. React. Eng. 2007, 5, A48. [Google Scholar] [CrossRef]
  89. Hongloi, N.; Prapainainar, P.; Prapainainar, C. Review of green diesel production from fatty acid deoxygenation over Ni-based catalysts. Mol. Catal. 2021, 111696. [Google Scholar] [CrossRef]
  90. Douvartzides, S.L.; Charisiou, N.D.; Papageridis, K.N.; Goula, M.A. Green diesel: Biomass feedstocks, production technologies, catalytic research, fuel properties and performance in compression ignition internal combustion engines. Energies 2019, 12, 809. [Google Scholar] [CrossRef] [Green Version]
  91. Zhang, B.; Wu, J.; Yang, C.; Qiu, Q.; Yan, Q.; Li, R.; Wang, B.; Wu, J.; Ding, Y. Recent developments in commercial processes for refining bio-feedstocks to renewable diesel. BioEnergy Res. 2018, 11, 689–702. [Google Scholar] [CrossRef]
  92. Panchasara, H.; Ashwath, N. Effects of Pyrolysis Bio-Oils on Fuel Atomisation—A Review. Energies 2021, 14, 794. [Google Scholar] [CrossRef]
  93. Lee, X.J.; Ong, H.C.; Gan, Y.Y.; Chen, W.-H.; Mahlia, T.M.I. State of art review on conventional and advanced pyrolysis of macroalgae and microalgae for biochar, bio-oil and bio-syngas production. Energy Convers. Manag. 2020, 210, 112707. [Google Scholar] [CrossRef]
  94. Yang, C.; Li, R.; Zhang, B.; Qiu, Q.; Wang, B.; Yang, H.; Ding, Y.; Wang, C. Pyrolysis of microalgae: A critical review. Fuel Processing Technol. 2019, 186, 53–72. [Google Scholar] [CrossRef]
  95. Haghighat, M.; Majidian, N.; Hallajisani, A. Production of bio-oil from sewage sludge: A review on the thermal and catalytic conversion by pyrolysis. Sustain. Energy Technol. Assess. 2020, 42, 100870. [Google Scholar] [CrossRef]
  96. Knothe, G. Biodiesel and renewable diesel: A comparison. Prog. Energy Combust. Sci. 2010, 36, 364–373. [Google Scholar] [CrossRef]
  97. Czajczyńska, D.; Anguilano, L.; Ghazal, H.; Krzyżyńska, R.; Reynolds, A.; Spencer, N.; Jouhara, H. Potential of pyrolysis processes in the waste management sector. Therm. Sci. Eng. Prog. 2017, 3, 171–197. [Google Scholar] [CrossRef]
  98. Cortez, L.; Franco, T.T.; Valença, G.; Rosillo-Calle, F. Perspective Use of Fast Pyrolysis Bio-Oil (FPBO) in Maritime Transport: The Case of Brazil. Energies 2021, 14, 4779. [Google Scholar] [CrossRef]
  99. He, Y.; Zhao, Y.; Chai, M.; Zhou, Z.; Sarker, M.; Li, C.; Liu, R.; Cai, J.; Liu, X. Comparative study of fast pyrolysis, hydropyrolysis and catalytic hydropyrolysis of poplar sawdust and rice husk in a modified Py-GC/MS microreactor system: Insights into product distribution, quantum description and reaction mechanism. Renew. Sustain. Energy Rev. 2020, 119, 109604. [Google Scholar] [CrossRef]
  100. Nascimento, L.; Ribeiro, A.; Ferreira, A.; Valério, N.; Pinheiro, V.; Araújo, J.; Vilarinho, C.; Carvalho, J. Turning Waste Cooking Oils into Biofuels—Valorization Technologies: A Review. Energies 2021, 15, 116. [Google Scholar] [CrossRef]
  101. Pattanaik, B.P.; Misra, R.D. Effect of reaction pathway and operating parameters on the deoxygenation of vegetable oils to produce diesel range hydrocarbon fuels: A review. Renew. Sustain. Energy Rev. 2017, 73, 545–557. [Google Scholar] [CrossRef]
  102. Hermida, L.; Abdullah, A.Z.; Mohamed, A.R. Deoxygenation of fatty acid to produce diesel-like hydrocarbons: A review of process conditions, reaction kinetics and mechanism. Renew. Sustain. Energy Rev. 2015, 42, 1223–1233. [Google Scholar] [CrossRef]
  103. Shimada, I.; Kato, S.; Hirazawa, N.; Nakamura, Y.; Ohta, H.; Suzuki, K.; Takatsuka, T. Deoxygenation of triglycerides by catalytic cracking with enhanced hydrogen transfer activity. Ind. Eng. Chem. Res. 2017, 56, 75–86. [Google Scholar] [CrossRef]
  104. Zhang, J.; Wu, Z.; Li, X.; Zhang, Y.; Bao, Z.; Bai, L.; Wang, F. Catalytic Cracking of Inedible Oils for the Production of Drop-In Biofuels over a SO42–/TiO2-ZrO2 Catalyst. Energy Fuels 2020, 34, 14204–14214. [Google Scholar] [CrossRef]
  105. Wang, H.; Lin, H.; Zheng, Y.; Ng, S.; Brown, H.; Xia, Y. Kaolin-based catalyst as a triglyceride FCC upgrading catalyst with high deoxygenation, mild cracking, and low dehydrogenation performances. Catal. Today 2019, 319, 164–171. [Google Scholar] [CrossRef]
  106. Khan, S.; Lup, A.N.K.; Qureshi, K.M.; Abnisa, F.; Daud, W.M.A.W.; Patah, M.F.A. A review on deoxygenation of triglycerides for jet fuel range hydrocarbons. J. Anal. Appl. Pyrolysis 2019, 140, 1–24. [Google Scholar] [CrossRef]
  107. Choo, M.-Y.; Oi, L.E.; Ling, T.C.; Ng, E.-P.; Lin, Y.-C.; Centi, G.; Juan, J.C. Deoxygenation of triolein to green diesel in the H2-free condition: Effect of transition metal oxide supported on zeolite Y. J. Anal. Appl. Pyrolysis 2020, 147, 104797. [Google Scholar] [CrossRef]
  108. Soni, V.K.; Dhara, S.; Krishnapriya, R.; Choudhary, G.; Sharma, P.R.; Sharma, R.K. Highly selective Co3O4/silica-alumina catalytic system for deoxygenation of triglyceride-based feedstock. Fuel 2020, 266, 117065. [Google Scholar] [CrossRef]
  109. Gamal, M.S.; Asikin-Mijan, N.; Khalit, W.N.A.W.; Arumugam, M.; Izham, S.M.; Taufiq-Yap, Y. Effective catalytic deoxygenation of palm fatty acid distillate for green diesel production under hydrogen-free atmosphere over bimetallic catalyst CoMo supported on activated carbon. Fuel Processing Technol. 2020, 208, 106519. [Google Scholar] [CrossRef]
  110. Ding, Z.; Zhao, T.; Zhu, Q.; Liao, S.; Ning, L.; Bi, Y.; Chen, H. Facile synthesis Cobalt catalysts for regulating the deoxygenation pathways in producing diesel-like hydrocarbon fuels. Biomass Bioenergy 2020, 143, 105879. [Google Scholar] [CrossRef]
  111. Baharudin, K.B.; Abdullah, N.; Taufiq-Yap, Y.H.; Derawi, D. Renewable diesel via solventless and hydrogen-free catalytic deoxygenation of palm fatty acid distillate. J. Clean. Prod. 2020, 274, 122850. [Google Scholar] [CrossRef]
  112. Gousi, M.; Kordouli, E.; Bourikas, K.; Simianakis, E.; Ladas, S.; Panagiotou, G.D.; Kordulis, C.; Lycourghiotis, A. Green diesel production over nickel-alumina nanostructured catalysts promoted by zinc. Catal. Today 2020, 355, 903–909. [Google Scholar] [CrossRef]
  113. Jiraroj, D.; Jirarattanapochai, O.; Anutrasakda, W.; Samec, J.S.; Tungasmita, D.N. Selective decarboxylation of biobased fatty acids using a Ni-FSM-16 catalyst. Appl. Catal. B Environ. 2021, 291, 120050. [Google Scholar] [CrossRef]
  114. Loe, R.; Lavoignat, Y.; Maier, M.; Abdallah, M.; Morgan, T.; Qian, D.; Pace, R.; Santillan-Jimenez, E.; Crocker, M. Continuous catalytic deoxygenation of waste free fatty acid-based feeds to fuel-like hydrocarbons over a supported Ni-Cu catalyst. Catalysts 2019, 9, 123. [Google Scholar] [CrossRef] [Green Version]
  115. Cheah, K.W.; Yusup, S.; Loy, A.C.M.; How, B.S.; Skoulou, V.; Taylor, M.J. Recent advances in the catalytic deoxygenation of plant oils and prototypical fatty acid models compounds: Catalysis, process, and kinetics. Mol. Catal. 2021, 111469. [Google Scholar] [CrossRef]
  116. Abdulkareem-Alsultan, G.; Asikin-Mijan, N.; Mustafa-Alsultan, G.; Lee, H.; Wilson, K.; Taufiq-Yap, Y.H. Efficient deoxygenation of waste cooking oil over Co3O4–La2O3-doped activated carbon for the production of diesel-like fuel. RSC Adv. 2020, 10, 4996–5009. [Google Scholar] [CrossRef] [Green Version]
  117. Kamaruzaman, M.F.; Taufiq-Yap, Y.H.; Derawi, D. Green diesel production from palm fatty acid distillate over SBA-15-supported nickel, cobalt, and nickel/cobalt catalysts. Biomass Bioenergy 2020, 134, 105476. [Google Scholar] [CrossRef]
  118. Yu, C.; Yu, S.; Li, L. Upgraded methyl oleate to diesel-like hydrocarbons through selective hydrodeoxygenation over Mo-based catalyst. Fuel 2022, 308, 122038. [Google Scholar] [CrossRef]
  119. Chen, J.; Zhu, Y.; Li, W.; Luo, F.; Li, S.; Li, X.; Huang, Y.; Zhang, A.; Xiao, Z.; Wang, D. Production of diesel-like hydrocarbons via hydrodeoxygenation of palmitic acid over Ni/TS-1 catalyst. Biomass Bioenergy 2021, 149, 106081. [Google Scholar] [CrossRef]
  120. Ameen, M.; Azizan, M.T.; Yusup, S.; Ramli, A.; Yasir, M. Catalytic hydrodeoxygenation of triglycerides: An approach to clean diesel fuel production. Renew. Sustain. Energy Rev. 2017, 80, 1072–1088. [Google Scholar] [CrossRef]
  121. Manara, P.; Bezergianni, S.; Pfisterer, U. Study on phase behavior and properties of binary blends of bio-oil/fossil-based refinery intermediates: A step toward bio-oil refinery integration. Energy Convers. Manag. 2018, 165, 304–315. [Google Scholar] [CrossRef]
  122. van Dyk, S.; Su, J.; Mcmillan, J.D.; Saddler, J. Potential synergies of drop-in biofuel production with further co-processing at oil refineries. Biofuels Bioprod. Biorefining 2019, 13, 760–775. [Google Scholar] [CrossRef] [Green Version]
  123. Bezergianni, S.; Dimitriadis, A.; Kikhtyanin, O.; Kubička, D. Refinery co-processing of renewable feeds. Prog. Energy Combust. Sci. 2018, 68, 29–64. [Google Scholar] [CrossRef]
  124. Glisic, S.B.; Pajnik, J.M.; Orlović, A.M. Process and techno-economic analysis of green diesel production from waste vegetable oil and the comparison with ester type biodiesel production. Appl. Energy 2016, 170, 176–185. [Google Scholar] [CrossRef]
  125. Ko, C.H.; Park, S.H.; Jeon, J.-K.; Suh, D.J.; Jeong, K.-E.; Park, Y.-K. Upgrading of biofuel by the catalytic deoxygenation of biomass. Korean J. Chem. Eng. 2012, 29, 1657–1665. [Google Scholar] [CrossRef]
  126. Veriansyah, B.; Han, J.Y.; Kim, S.K.; Hong, S.-A.; Kim, Y.J.; Lim, J.S.; Shu, Y.-W.; Oh, S.-G.; Kim, J. Production of renewable diesel by hydroprocessing of soybean oil: Effect of catalysts. Fuel 2012, 94, 578–585. [Google Scholar] [CrossRef]
  127. Hájek, M.; Vávra, A.; de Paz Carmona, H.; Kocík, J. The Catalysed Transformation of Vegetable Oils or Animal Fats to Biofuels and Bio-Lubricants: A Review. Catalysts 2021, 11, 1118. [Google Scholar] [CrossRef]
  128. El-Sawy, M.S.; Hanafi, S.A.; Ashour, F.; Aboul-Fotouh, T.M. Co-hydroprocessing and hydrocracking of alternative feed mixture (vacuum gas oil/waste lubricating oil/waste cooking oil) with the aim of producing high quality fuels. Fuel 2020, 269, 117437. [Google Scholar] [CrossRef]
  129. Carmona, H.D.P.; Akhmetzyanova, U.; Tišler, Z.; Vondrova, P. Hydrotreating atmospheric gasoil and co-processing with rapeseed oil using supported Ni-Mo and Co-Mo carbide catalysts. Fuel 2020, 268, 117363. [Google Scholar] [CrossRef]
  130. De Paz Carmona, H.; Svobodová, E.k.; Tišler, Z.k.; Akhmetzyanova, U.; Strejcová, K.i. Hydrotreating of Atmospheric Gas Oil and Co-Processing with Rapeseed Oil Using Sulfur-Free PMoCx/Al2O3 Catalysts. ACS Omega 2021, 6, 7680–7692. [Google Scholar] [CrossRef]
  131. Horáček, J.; Akhmetzyanova, U.; Skuhrovcova, L.; Tišler, Z.; de Paz Carmona, H. Alumina-supported MoNx, MoCx and MoPx catalysts for the hydrotreatment of rapeseed oil. Appl. Catal. B Environ. 2020, 263, 118328. [Google Scholar] [CrossRef]
  132. Plazas-González, M.; Guerrero-Fajardo, C.A.; Sodré, J.R. Modelling and simulation of hydrotreating of palm oil components to obtain green diesel. J. Clean. Prod. 2018, 184, 301–308. [Google Scholar] [CrossRef]
  133. Chu, P.L.; Vanderghem, C.; MacLean, H.L.; Saville, B.A. Financial analysis and risk assessment of hydroprocessed renewable jet fuel production from camelina, carinata and used cooking oil. Appl. Energy 2017, 198, 401–409. [Google Scholar] [CrossRef]
  134. Kruger, J.S.; Christensen, E.D.; Dong, T.; Van Wychen, S.; Fioroni, G.M.; Pienkos, P.T.; McCormick, R.L. Bleaching and hydroprocessing of algal biomass-derived lipids to produce renewable diesel fuel. Energy Fuels 2017, 31, 10946–10953. [Google Scholar] [CrossRef]
  135. Karatzos, S.; van Dyk, J.S.; McMillan, J.D.; Saddler, J. Drop-in biofuel production via conventional (lipid/fatty acid) and advanced (biomass) routes. Part I. Biofuels Bioprod. Biorefining 2017, 11, 344–362. [Google Scholar] [CrossRef]
  136. Sousa, F.P.; Silva, L.N.; de Rezende, D.B.; de Oliveira, L.C.A.; Pasa, V.M. Simultaneous deoxygenation, cracking and isomerization of palm kernel oil and palm olein over beta zeolite to produce biogasoline, green diesel and biojet-fuel. Fuel 2018, 223, 149–156. [Google Scholar] [CrossRef]
  137. Scaldaferri, C.A.; Pasa, V.M.D. Hydrogen-free process to convert lipids into bio-jet fuel and green diesel over niobium phosphate catalyst in one-step. Chem. Eng. J. 2019, 370, 98–109. [Google Scholar] [CrossRef]
  138. Romero-Izquierdo, A.G.; Gutiérrez-Antonio, C.; Gómez-Castro, F.I.; Hernández, S. Hydrotreating of triglyceride feedstock to produce renewable aviation fuel. Recent Innov. Chem. Eng. Former. Recent Pat. Chem. Eng. 2018, 11, 77–89. [Google Scholar] [CrossRef]
  139. Kordouli, E.; Sygellou, L.; Kordulis, C.; Bourikas, K.; Lycourghiotis, A. Probing the synergistic ratio of the NiMo/γ-Al2O3 reduced catalysts for the transformation of natural triglycerides into green diesel. Appl. Catal. B Environ. 2017, 209, 12–22. [Google Scholar] [CrossRef]
  140. Kubička, D.; Tukač, V. Hydrotreating of triglyceride-based feedstocks in refineries. In Advances in Chemical Engineering; Elsevier: Amsterdam, The Netherlands, 2013; Volume 42, pp. 141–194. [Google Scholar]
  141. Sági, D.; Baladincz, P.; Varga, Z.; Hancsók, J. Co-processing of FCC light cycle oil and waste animal fats with straight run gas oil fraction. J. Clean. Prod. 2016, 111, 34–41. [Google Scholar] [CrossRef]
  142. Boonyasuwat, S.; Tscheikuna, J. Co-processing of palm fatty acid distillate and light gas oil in pilot-scale hydrodesulfurization unit over commercial CoMo/Al2O3. Fuel 2017, 199, 115–124. [Google Scholar] [CrossRef]
  143. Carmona, H.D.P.; De La Torre Alfaro, O.; Alayón, A.B.; Vázquez, M.R.; Hernández, J.M. Co-processing of straight run gas oil with used cooking oil and animal fats. Fuel 2019, 254, 115583. [Google Scholar] [CrossRef]
  144. Herrador, J.M.H.; Psenicka, M.; Horacek, J.; Tisler, Z.; Vrablik, A.; Cerny, R.; Murat, M. Co-processing of Waste Cooking Oil and Light Cycle Oil with NiW/(Pseudoboehmite + SBA-15) Catalyst. Chem. Eng. Technol. 2019, 42, 512–517. [Google Scholar] [CrossRef]
  145. Tóth, O.; Holló, A.; Hancsók, J. Co-processing a waste fatty acid mixture and unrefined gas oil to produce renewable diesel fuel-blending components. Energy Convers. Manag. 2019, 185, 304–312. [Google Scholar] [CrossRef]
  146. Vlasova, E.; Porsin, A.; Aleksandrov, P.; Nuzhdin, A.; Bukhtiyarova, G. Co-processing of rapeseed oil—Straight run gas oil mixture: Comparative study of sulfide CoMo/Al2O3-SAPO-11 and NiMo/Al2O3-SAPO-11 catalysts. Catal. Today 2021, 378, 119–125. [Google Scholar] [CrossRef]
  147. Pelemo, J.; Inambao, F.L.; Onuh, E.I. Potential of Used Cooking Oil as Feedstock for Hydroprocessing into Hydrogenation Derived Renewable Diesel: A Review. IJERT 2020, 13, 500–519. [Google Scholar] [CrossRef]
  148. Hachemi, I.; Kumar, N.; Mäki-Arvela, P.; Roine, J.; Peurla, M.; Hemming, J.; Salonen, J.; Murzin, D.Y. Sulfur-free Ni catalyst for production of green diesel by hydrodeoxygenation. J. Catal. 2017, 347, 205–221. [Google Scholar] [CrossRef]
  149. Zhao, X.; Wei, L.; Cheng, S.; Julson, J. Review of heterogeneous catalysts for catalytically upgrading vegetable oils into hydrocarbon biofuels. Catalysts 2017, 7, 83. [Google Scholar] [CrossRef]
  150. Abdulkareem-Alsultan, G.; Asikin-Mijan, N.; Mansir, N.; Lee, H.V.; Zainal, Z.; Islam, A.; Taufiq-Yap, Y.H. Pyro-lytic de-oxygenation of waste cooking oil for green diesel production over Ag2O3-La2O3/AC nano-catalyst. J. Anal. Appl. Pyrolysis 2019, 137, 171–184. [Google Scholar] [CrossRef]
  151. Arun, N.; Sharma, R.V.; Dalai, A.K. Green diesel synthesis by hydrodeoxygenation of bio-based feedstocks: Strategies for catalyst design and development. Renew. Sustain. Energy Rev. 2015, 48, 240–255. [Google Scholar] [CrossRef]
  152. Asikin-Mijan, N.; Lee, H.V.; Juan, J.C.; Noorsaadah, A.; Ong, H.C.; Razali, S.; Taufiq-Yap, Y.H. Promoting deoxygenation of triglycerides via Co-Ca loaded SiO2-Al2O3 catalyst. Appl. Catal. A Gen. 2018, 552, 38–48. [Google Scholar] [CrossRef]
  153. Kordulis, C.; Bourikas, K.; Gousi, M.; Kordouli, E.; Lycourghiotis, A. Development of nickel based catalysts for the transformation of natural triglycerides and related compounds into green diesel: A critical review. Appl. Catal. B Environ. 2016, 181, 156–196. [Google Scholar] [CrossRef]
  154. Neuling, U.; Kaltschmitt, M. Techno-economic and environmental analysis of aviation biofuels. Fuel Process. Technol. 2018, 171, 54–69. [Google Scholar] [CrossRef]
  155. Lu, C. When will biofuels be economically feasible for commercial flights? Considering the difference between environmental benefits and fuel purchase costs. J. Clean. Prod. 2018, 181, 365–373. [Google Scholar] [CrossRef]
  156. Deane, J.; Pye, S. Europe’s ambition for biofuels in aviation-A strategic review of challenges and opportunities. Energy Strategy Rev. 2018, 20, 1–5. [Google Scholar] [CrossRef]
  157. Prussi, M.; O’connell, A.; Lonza, L. Analysis of current aviation biofuel technical production potential in EU28. Biomass Bioenergy 2019, 130, 105371. [Google Scholar] [CrossRef]
  158. Why, E.S.K.; Ong, H.C.; Lee, H.V.; Gan, Y.Y.; Chen, W.-H.; Chong, C.T. Renewable aviation fuel by advanced hydroprocessing of biomass: Challenges and perspective. Energy Convers. Manag. 2019, 199, 112015. [Google Scholar] [CrossRef]
  159. Lin, C.-H.; Chen, Y.-K.; Wang, W.-C. The production of bio-jet fuel from palm oil derived alkanes. Fuel 2020, 260, 116345. [Google Scholar] [CrossRef]
  160. Pacheco, G.a.; Silva, A.; Costa, M.r. Single-Droplet Combustion of Jet A-1, Hydroprocessed Vegetable Oil, and Their Blends in a Drop-Tube Furnace. Energy Fuels 2021, 35, 7232–7241. [Google Scholar] [CrossRef]
  161. Hari, T.K.; Yaakob, Z.; Binitha, N.N. Aviation biofuel from renewable resources: Routes, opportunities and challenges. Renew. Sustain. Energy Rev. 2015, 42, 1234–1244. [Google Scholar] [CrossRef]
  162. Cabrera, E.; de Sousa, J.M.M. Use of Sustainable Fuels in Aviation—A Review. Energies 2022, 15, 2440. [Google Scholar] [CrossRef]
  163. Sotelo-Boyás, R.; Trejo-Zárraga, F.; Hernández-Loyo, F.D.J. Hydroconversion of triglycerides into green liquid fuels. Hydrogenation 2012, 338, 338. [Google Scholar]
  164. Verdier, S.; Alkilde, O.F.; Chopra, R.; Gabrielsen, J.; Grubb, M. Hydroprocessing of Renewable Feedstocks-Challenges and Solutions; Haldor Topsoe A/S: Copenhagen, Denmark, 2019. [Google Scholar]
  165. Vasquez, M.C.; Silva, E.E.; Castillo, E.F. Hydrotreatment of vegetable oils: A review of the technologies and its developments for jet biofuel production. Biomass Bioenergy 2017, 105, 197–206. [Google Scholar] [CrossRef]
  166. Ng, K.S.; Farooq, D.; Yang, A. Global biorenewable development strategies for sustainable aviation fuel production. Renew. Sustain. Energy Rev. 2021, 150, 111502. [Google Scholar] [CrossRef]
  167. Alherbawi, M.; McKay, G.; Mackey, H.R.; Al-Ansari, T. Jatropha curcas for jet biofuel production: Current status and future prospects. Renew. Sustain. Energy Rev. 2021, 135, 110396. [Google Scholar] [CrossRef]
  168. Prakash, T. Renewable Diesel: The Fuel of the Future; Future Bridge: Newark, NJ, USA, 2021. [Google Scholar]
  169. O’brien, P. Business Trends: Battle of the Biofuels: Renewable Diesel vs. Biodiesel; Hydrocarbon Processing: Houston, TX, USA, 2021. [Google Scholar]
  170. Argüelles, A.; Amezcua-Allieri, M.A.; Ramírez, L.F. Life cycle assessment of green diesel production by hydrodeoxygenation of palm oil. Front. Energy Res. 2021, 9, 296. [Google Scholar] [CrossRef]
  171. Tan, K.T.; Ang, G.T. Recent trends and advances in glycerol-free biodiesel production. In Advanced Bioprocessing for Alternative Fuels, Biobased Chemicals, and Bioproducts; Elsevier: Amsterdam, The Netherlands, 2019; pp. 153–164. [Google Scholar]
  172. Sakdasri, W.; Komintarachat, C.; Sawangkeaw, R.; Ngamprasertsith, S. A Review of Supercritical Technologies for Lipid-Based Biofuels Production: The Glycerol-free Processes. Eng. J. 2021, 25, 1–14. [Google Scholar] [CrossRef]
  173. Uthoff, S.; Bröker, D.; Steinbüchel, A. Current state and perspectives of producing biodiesel-like compounds by biotechnology. Microb. Biotechnol. 2009, 2, 551–565. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Kijenski, J.; Lipkowski, A.; Walisiewicz-Niedbalska, W.; Gwardiak, H.; Rozyczki, K.; Pawlak, I. A Biofuel for Compression-Ignition Engines and A Method for Preparing the Biofuel. European Patent EP1580255, 2004. [Google Scholar]
  175. Ryms, M.; Lewandowski, W.M.; Januszewicz, K.; Klugmann-Radziemska, E.; Ciunel, K. Methods of liquid biofuel production-the biodiesel example. Proc. ECOpole 2013, 7, 67. [Google Scholar] [CrossRef]
  176. Żółtowski, A.; Grzelak, P.L. Emissions from engines fuelled with biofuels. J. KONES 2018, 25, 533–539. [Google Scholar]
  177. Kijeński, J. Biorefineries-from biofuels to the chemicalization of agricultural products. Pol. J. Chem. Technol. 2007, 9, 42–45. [Google Scholar] [CrossRef] [Green Version]
  178. Casas, A.; Ramos, M.J.S.; Pérez, A.N. Product separation after chemical interesterification of vegetable oils with methyl acetate. Part II: Liquid–liquid equilibrium. Ind. Eng. Chem. Res. 2012, 51, 10201–10206. [Google Scholar] [CrossRef]
  179. Casas, A.; Ramos, M.J.; Pérez, Á. Kinetics of chemical interesterification of sunflower oil with methyl acetate for biodiesel and triacetin production. Chem. Eng. J. 2011, 171, 1324–1332. [Google Scholar] [CrossRef]
  180. Casas, A.; Ramos, M.J.; Perez, A. New trends in biodiesel production: Chemical interesterification of sunflower oil with methyl acetate. Biomass Bioenergy 2011, 35, 1702–1709. [Google Scholar] [CrossRef]
  181. Casas, A.; Ramos, M.J.; Pérez, Á. Methanol-enhanced chemical interesterification of sunflower oil with methyl acetate. Fuel 2013, 106, 869–872. [Google Scholar] [CrossRef]
  182. Visioli, L.J.; Trentini, C.P.; de Castilhos, F.; da Silva, C. Esters production in continuous reactor from macauba pulp oil using methyl acetate in pressurized conditions. J. Supercrit. Fluids 2018, 140, 238–247. [Google Scholar] [CrossRef]
  183. Kampars, V.; Šustere, Z.; Kampare, R. Biofuel via interesterification of rapeseed oil with methyl acetate in presence of potassium t-butoxide/THF. Int. Multidiscip. Sci. GeoConference: SGEM 2018, 18, 163–170. [Google Scholar]
  184. Brondani, L.; Simões, S.; Celante, D.; Castilhos, F. Kinetic modeling of supercritical interesterification with heterogeneous catalyst to produce methyl esters considering degradation effects. Ind. Eng. Chem. Res. 2018, 58, 816–827. [Google Scholar] [CrossRef]
  185. Dhillon, S.S.; Tan, K.T. Optimization of biodiesel production via methyl acetate reaction from cerbera odollam. Adv. Energy Res. 2016, 4, 325. [Google Scholar] [CrossRef]
  186. Ribeiro, J.S.; Celante, D.; Brondani, L.N.; Trojahn, D.O.; da Silva, C.; de Castilhos, F. Synthesis of methyl esters and triacetin from macaw oil (Acrocomia aculeata) and methyl acetate over γ-alumina. Ind. Crops Prod. 2018, 124, 84–90. [Google Scholar] [CrossRef]
  187. Māliņš, K.; Kampars, V.; Kampare, R.; Sustere, Z.; Arenta, A. Production of biodiesel and triacetin by interesterification of rapeseed oil. In Key Engineering Materials; Trans Tech Publications Ltd.: Freienbach, Switzerland, 2018; pp. 129–133. [Google Scholar]
  188. Simões, S.; Ribeiro, J.; Celante, D.; Brondani, L.; Castilhos, F. Heterogeneous catalyst screening for fatty acid methyl esters production through interesterification reaction. Renew. Energy 2020, 146, 719–726. [Google Scholar] [CrossRef]
  189. Gupta, A.R.; Yadav, S.V.; Rathod, V.K. Enhancement in biodiesel production using waste cooking oil and calcium diglyceroxide as a heterogeneous catalyst in presence of ultrasound. Fuel 2015, 158, 800–806. [Google Scholar] [CrossRef]
  190. Galia, A.; Centineo, A.; Saracco, G.; Schiavo, B.; Scialdone, O. Interesterification of rapeseed oil catalyzed by tin octoate. Biomass Bioenergy 2014, 67, 193–200. [Google Scholar] [CrossRef]
  191. Man, I.-C.; Soriga, S.G.; Parvulescu, V. Effect of Ca and Sr in MgO (100) on the activation of methanol and methyl acetate. Catal. Today 2018, 306, 207–214. [Google Scholar] [CrossRef]
  192. Alves, M.A.; Pinheiro, N.S.; Brondani, L.N.; Celante, D.; Ketzer, F.; Castilhos, F. Assessment of niobium phosphate as heterogeneous catalyst in esterification with methyl acetate. J. Chem. Technol. Biotechnol. 2019, 94, 3172–3179. [Google Scholar] [CrossRef]
  193. Sustere, Z.; Murnieks, R.; Kampars, V. Chemical interesterification of rapeseed oil with methyl, ethyl, propyl and isopropyl acetates and fuel properties of obtained mixtures. Fuel Process. Technol. 2016, 149, 320–325. [Google Scholar] [CrossRef]
  194. Dhawan, M.S.; Barton, S.C.; Yadav, G.D. Interesterification of triglycerides with methyl acetate for the co-production biodiesel and triacetin using hydrotalcite as a heterogenous base catalyst. Catal. Today 2021, 375, 101–111. [Google Scholar] [CrossRef]
  195. Casas, A.; Ramos, M.J.; Pérez, Á. Adsorption equilibrium and kinetics of methyl acetate/methanol and methyl acetate/water mixtures on zeolite 5A. Chem. Eng. J. 2013, 220, 337–342. [Google Scholar] [CrossRef]
  196. Xu, Y.; Du, W.; Liu, D. Study on the kinetics of enzymatic interesterification of triglycerides for biodiesel production with methyl acetate as the acyl acceptor. J. Mol. Catal. B Enzym. 2005, 32, 241–245. [Google Scholar] [CrossRef]
  197. Orçaire, O.; Buisson, P.; Pierre, A.C. Application of silica aerogel encapsulated lipases in the synthesis of biodiesel by transesterification reactions. J. Mol. Catal. B Enzym. 2006, 42, 106–113. [Google Scholar] [CrossRef]
  198. Ognjanovic, N.; Bezbradica, D.; Knezevic-Jugovic, Z. Enzymatic conversion of sunflower oil to biodiesel in a solvent-free system: Process optimization and the immobilized system stability. Bioresour. Technol. 2009, 100, 5146–5154. [Google Scholar] [CrossRef]
  199. Talukder, M.M.R.; Das, P.; Fang, T.S.; Wu, J.C. Enhanced enzymatic transesterification of palm oil to biodiesel. Biochem. Eng. J. 2011, 55, 119–122. [Google Scholar] [CrossRef]
  200. Vasudevan, P.T.; Fu, B. Environmentally sustainable biofuels: Advances in biodiesel research. Waste Biomass Valorization 2010, 1, 47–63. [Google Scholar] [CrossRef]
  201. Marx, S. Glycerol-free biodiesel production through transesterification: A review. Fuel Processing Technol. 2016, 151, 139–147. [Google Scholar] [CrossRef]
  202. Norjannah, B.; Ong, H.C.; Masjuki, H.; Juan, J.; Chong, W. Enzymatic transesterification for biodiesel production: A comprehensive review. RSC Adv. 2016, 6, 60034–60055. [Google Scholar] [CrossRef]
  203. Alavijeh, R.S.; Tabandeh, F.; Tavakoli, O.; Karkhane, A.; Shariati, P. Enzymatic production of biodiesel from microalgal oil using ethyl acetate as an acyl acceptor. J. Oleo Sci. 2015, 64, 69–74. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Ruzich, N.I.; Bassi, A.S. Investigation of enzymatic biodiesel production using ionic liquid as a co-solvent. Can. J. Chem. Eng. 2010, 88, 277–282. [Google Scholar] [CrossRef]
  205. Ruzich, N.I.; Bassi, A.S. Proposed kinetic mechanism of biodiesel production through lipase catalysed interesterification with a methyl acetate acyl acceptor and ionic liquid (BMIM)(PF6) co-solvent. Can. J. Chem. Eng. 2011, 89, 166–170. [Google Scholar] [CrossRef]
  206. Muhammad, N.; Elsheikh, Y.A.; Mutalib, M.I.A.; Bazmi, A.A.; Khan, R.A.; Khan, H.; Rafiq, S.; Man, Z. An overview of the role of ionic liquids in biodiesel reactions. J. Ind. Eng. Chem. 2015, 21, 1–10. [Google Scholar] [CrossRef]
  207. Quintana-Gómez, L.; Ladero, M.; Calvo, L. Enzymatic production of biodiesel from alperujo oil in supercritical CO2. J. Supercrit. Fluids 2021, 171, 105184. [Google Scholar] [CrossRef]
  208. Kashyap, S.S.; Gogate, P.R.; Joshi, S.M. Ultrasound assisted synthesis of biodiesel from karanja oil by interesterification: Intensification studies and optimization using RSM. Ultrason. Sonochemistry 2019, 50, 36–45. [Google Scholar] [CrossRef]
  209. Andreani, L.; Rocha, J. Use of ionic liquids in biodiesel production: A review. Braz. J. Chem. Eng. 2012, 29, 1–13. [Google Scholar] [CrossRef] [Green Version]
  210. Campanelli, P.; Banchero, M.; Manna, L. Synthesis of biodiesel from edible, non-edible and waste cooking oils via supercritical methyl acetate transesterification. Fuel 2010, 89, 3675–3682. [Google Scholar] [CrossRef]
  211. Niza, N.M.; Tan, K.T.; Lee, K.T.; Ahmad, Z. Biodiesel production by non-catalytic supercritical methyl acetate: Thermal stability study. Appl. Energy 2013, 101, 198–202. [Google Scholar] [CrossRef]
  212. Goembira, F.; Matsuura, K.; Saka, S. Biodiesel production from rapeseed oil by various supercritical carboxylate esters. Fuel 2012, 97, 373–378. [Google Scholar] [CrossRef] [Green Version]
  213. Bernal, J.M.; Lozano, P.; García-Verdugo, E.; Burguete, M.I.; Sánchez-Gómez, G.; López-López, G.; Pucheault, M.; Vaultier, M.; Luis, S.V. Supercritical synthesis of biodiesel. Molecules 2012, 17, 8696–8719. [Google Scholar] [CrossRef] [Green Version]
  214. Lourinho, G.; Brito, P. Advanced biodiesel production technologies: Novel developments. Rev. Environ. Sci. Bio/Technol. 2015, 14, 287–316. [Google Scholar] [CrossRef]
  215. Komintarachat, C.; Sawangkeaw, R.; Ngamprasertsith, S. Continuous production of palm biofuel under supercritical ethyl acetate. Energy Convers. Manag. 2015, 93, 332–338. [Google Scholar] [CrossRef]
  216. Goembira, F.; Saka, S. Advanced supercritical Methyl acetate method for biodiesel production from Pongamia pinnata oil. Renew. Energy 2015, 83, 1245–1249. [Google Scholar] [CrossRef]
  217. Farobie, O.; Matsumura, Y. State of the art of biodiesel production under supercritical conditions. Prog. Energy Combust. Sci. 2017, 63, 173–203. [Google Scholar] [CrossRef]
  218. Patil, P.D.; Reddy, H.; Muppaneni, T.; Deng, S. Biodiesel fuel production from algal lipids using supercritical methyl acetate (glycerin-free) technology. Fuel 2017, 195, 201–207. [Google Scholar] [CrossRef] [Green Version]
  219. Lamba, N.; Gupta, K.; Modak, J.M.; Madras, G. Biodiesel synthesis from Calophyllum inophyllum oil with different supercritical fluids. Bioresour. Technol. 2017, 241, 767–774. [Google Scholar] [CrossRef]
  220. Mahfud, M.; Ansori, A. Box-Behnken Design for Optimization on Biodiesel Production from Palm Oil and Methyl Acetate using Ultrasound Assisted Interesterification Method. Period. Polytech. Chem. Eng. 2022, 66, 30–42. [Google Scholar] [CrossRef]
  221. Tavares, G.R.; Goncalves, J.E.; dos Santos, W.D.; da Silva, C. Enzymatic interesterification of crambe oil assisted by ultrasound. Ind. Crops Prod. 2017, 97, 218–223. [Google Scholar] [CrossRef]
  222. Subhedar, P.B.; Gogate, P.R. Ultrasound assisted intensification of biodiesel production using enzymatic interesterification. Ultrason. Sonochemistry 2016, 29, 67–75. [Google Scholar] [CrossRef]
  223. Bokhari, A.; Yusup, S.; Chuah, L.F.; Klemeš, J.J.; Asif, S.; Ali, B.; Akbar, M.M.; Kamil, R.N.M. Pilot scale intensification of rubber seed (Hevea brasiliensis) oil via chemical interesterification using hydrodynamic cavitation technology. Bioresour. Technol. 2017, 242, 272–282. [Google Scholar] [CrossRef]
  224. Ketzer, F.; Celante, D.; de Castilhos, F. Catalytic performance and ultrasonic-assisted impregnation effects on WO3/USY zeolites in esterification of oleic acid with methyl acetate. Microporous Mesoporous Mater. 2020, 291, 109704. [Google Scholar] [CrossRef]
  225. Poppe, J.K.; Matte, C.R.; Fernandez-Lafuente, R.; Rodrigues, R.C.; Ayub, M.A.Z. Transesterification of waste frying oil and soybean oil by combi-lipases under ultrasound-assisted reactions. Appl. Biochem. Biotechnol. 2018, 186, 576–589. [Google Scholar] [CrossRef]
  226. Gusniah, A.; Veny, H.; Hamzah, F. Ultrasonic assisted enzymatic transesterification for biodiesel production. Ind. Eng. Chem. Res. 2018, 58, 581–589. [Google Scholar] [CrossRef]
  227. Pukale, D.D.; Maddikeri, G.L.; Gogate, P.R.; Pandit, A.B.; Pratap, A.P. Ultrasound assisted transesterification of waste cooking oil using heterogeneous solid catalyst. Ultrason. Sonochemistry 2015, 22, 278–286. [Google Scholar] [CrossRef]
  228. Kim, S.-J.; Jung, S.-M.; Park, Y.-C.; Park, K. Lipase catalyzed transesterification of soybean oil using ethyl acetate, an alternative acyl acceptor. Biotechnol. Bioprocess Eng. 2007, 12, 441. [Google Scholar] [CrossRef]
  229. Chuepeng, S.; Komintarachat, C. Interesterification optimization of waste cooking oil and ethyl acetate over homogeneous catalyst for biofuel production with engine validation. Appl. Energy 2018, 232, 728–739. [Google Scholar] [CrossRef]
  230. Jeong, G.-T.; Park, D.-H. Synthesis of rapeseed biodiesel using short-chained alkyl acetates as acyl acceptor. Appl. Biochem. Biotechnol. 2010, 161, 195–208. [Google Scholar] [CrossRef]
  231. Jazie, A.A.; Jaddan, R.I.; Al-Dawody, M.F.; Abed, S.A. Lipase Acrylic Resin Catalyzed Interesterification of Sewage Sludge in Micro Packed Bed Reactor: Box-Behnken Design. In Key Engineering Materials; Trans Tech Publications Ltd.: Freienbach, Switzerland, 2020; pp. 81–96. [Google Scholar]
  232. Akkarawatkhoosith, N.; Kaewchada, A.; Ngamcharussrivichai, C.; Jaree, A. Biodiesel production via Interesterification of palm oil and ethyl acetate using ion-exchange resin in a packed-bed reactor. BioEnergy Res. 2020, 13, 542–551. [Google Scholar] [CrossRef]
  233. Rahmat, N.; Abdullah, A.Z.; Mohamed, A.R. Recent progress on innovative and potential technologies for glycerol transformation into fuel additives: A critical review. Renew. Sustain. Energy Rev. 2010, 14, 987–1000. [Google Scholar] [CrossRef]
  234. Melero, J.A.; Vicente, G.; Morales, G.; Paniagua, M.; Bustamante, J. Oxygenated compounds derived from glycerol for biodiesel formulation: Influence on EN 14214 quality parameters. Fuel 2010, 89, 2011–2018. [Google Scholar] [CrossRef]
  235. Gonçalves, V.L.; Pinto, B.P.; Silva, J.C.; Mota, C.J. Acetylation of glycerol catalyzed by different solid acids. Catal. Today 2008, 133, 673–677. [Google Scholar] [CrossRef]
  236. Liao, X.; Zhu, Y.; Wang, S.-G.; Li, Y. Producing triacetylglycerol with glycerol by two steps: Esterification and acetylation. Fuel Processing Technol. 2009, 90, 988–993. [Google Scholar] [CrossRef]
  237. Rezayat, M.; Ghaziaskar, H.S. Continuous synthesis of glycerol acetates in supercritical carbon dioxide using Amberlyst 15®. Green Chem. 2009, 11, 710–715. [Google Scholar] [CrossRef]
  238. Testa, M.L.; La Parola, V.; Liotta, L.F.; Venezia, A.M. Screening of different solid acid catalysts for glycerol acetylation. J. Mol. Catal. A: Chem. 2013, 367, 69–76. [Google Scholar] [CrossRef]
  239. Dalla Costa, B.O.; Decolatti, H.P.; Legnoverde, M.S.; Querini, C.A. Influence of acidic properties of different solid acid catalysts for glycerol acetylation. Catal. Today 2017, 289, 222–230. [Google Scholar] [CrossRef]
  240. Huang, M.-Y.; Han, X.-X.; Hung, C.-T.; Lin, J.-C.; Wu, P.-H.; Wu, J.-C.; Liu, S.-B. Heteropolyacid-based ionic liquids as efficient homogeneous catalysts for acetylation of glycerol. J. Catal. 2014, 320, 42–51. [Google Scholar] [CrossRef]
  241. Khayoon, M.; Triwahyono, S.; Hameed, B.; Jalil, A. Improved production of fuel oxygenates via glycerol acetylation with acetic acid. Chem. Eng. J. 2014, 243, 473–484. [Google Scholar] [CrossRef]
  242. Betiha, M.; Hassan, H.M.; El-Sharkawy, E.; Al-Sabagh, A.; Menoufy, M.; Abdelmoniem, H.M. A new approach to polymer-supported phosphotungstic acid: Application for glycerol acetylation using robust sustainable acidic heterogeneous–homogenous catalyst. Appl. Catal. B Environ. 2016, 182, 15–25. [Google Scholar] [CrossRef]
  243. Sandesh, S.; Manjunathan, P.; Halgeri, A.B.; Shanbhag, G.V. Glycerol acetins: Fuel additive synthesis by acetylation and esterification of glycerol using cesium phosphotungstate catalyst. RSC Adv. 2015, 5, 104354–104362. [Google Scholar] [CrossRef]
  244. Venkateswara Rao, P.; Appa Rao, B. Performance and emission characteristics of diesel engine with COME-Triacetin additive blends as fuel. Int. J. Energy Environ. 2012, 3, 629–638. [Google Scholar]
  245. Verma, P.; Stevanovic, S.; Zare, A.; Dwivedi, G.; Chu Van, T.; Davidson, M.; Rainey, T.; Brown, R.J.; Ristovski, Z.D. An overview of the influence of biodiesel, alcohols, and various oxygenated additives on the particulate matter emissions from diesel engines. Energies 2019, 12, 1987. [Google Scholar] [CrossRef] [Green Version]
  246. Khounani, Z.; Hosseinzadeh-Bandbafha, H.; Moustakas, K.; Talebi, A.F.; Goli, S.A.H.; Rajaeifar, M.A.; Khoshnevisan, B.; Jouzani, G.S.; Peng, W.; Kim, K.-H. Environmental life cycle assessment of different biorefinery platforms valorizing olive wastes to biofuel, phosphate salts, natural antioxidant, and an oxygenated fuel additive (triacetin). J. Clean. Prod. 2021, 278, 123916. [Google Scholar] [CrossRef]
  247. Okoye, P.; Abdullah, A.; Hameed, B. A review on recent developments and progress in the kinetics and deactivation of catalytic acetylation of glycerol—A byproduct of biodiesel. Renew. Sustain. Energy Rev. 2017, 74, 387–401. [Google Scholar] [CrossRef]
  248. Casas, A.; Ruiz, J.R.; Ramos, M.A.J.S.; Pérez, A. Effects of triacetin on biodiesel quality. Energy Fuels 2010, 24, 4481–4489. [Google Scholar] [CrossRef]
  249. Venkateswara Rao, P.; Appa Rao, B.; Radhakrishna, D. Experimental analysis of DI diesel engine performance with blend fuels of oxygenated additive and COME biodiesel. Iran. (Iran.) J. Energy Environ. 2012, 3, 109–117. [Google Scholar]
  250. Zare, A.; Bodisco, T.A.; Nabi, M.N.; Hossain, F.M.; Ristovski, Z.D.; Brown, R.J. Engine performance during transient and steady-state operation with oxygenated fuels. Energy Fuels 2017, 31, 7510–7522. [Google Scholar] [CrossRef]
  251. Elias, R.C.; Senra, M.; Soh, L. Cold flow properties of fatty acid methyl ester blends with and without triacetin. Energy Fuels 2016, 30, 7400–7409. [Google Scholar] [CrossRef]
  252. Rao, P.V. Role of Triacetin additive in the performance of single cylinder DI diesel engine with COME biodiesel. Int. J. Adv. Eng. Res. Sci. 2018, 5, 264286. [Google Scholar]
  253. Tabatabaei, M.; Aghbashlo, M.; Najafi, B.; Hosseinzadeh-Bandbafha, H.; Ardabili, S.F.; Akbarian, E.; Khalife, E.; Mohammadi, P.; Rastegari, H.; Ghaziaskar, H.S. Environmental impact assessment of the mechanical shaft work produced in a diesel engine running on diesel/biodiesel blends containing glycerol-derived triacetin. J. Clean. Prod. 2019, 223, 466–486. [Google Scholar] [CrossRef]
  254. Rao, P.V.; Rao, B.A. Heat release rate, performance and vibration analysis of diesel engine operating with biodiesel-Triacetin additive blend fuels. Int. J. Automob. Eng. Res. Dev. 2018, 8, 11–22. [Google Scholar]
  255. Zare, A.; Bodisco, T.A.; Nabi, M.N.; Hossain, F.M.; Rahman, M.M.; Ristovski, Z.D.; Brown, R.J. The influence of oxygenated fuels on transient and steady-state engine emissions. Energy 2017, 121, 841–853. [Google Scholar] [CrossRef] [Green Version]
  256. Lapuerta, M.; González-García, I.; Céspedes, I.; Estévez, C.; Bayarri, N. Improvement of cold flow properties of a new biofuel derived from glycerol. Fuel 2019, 242, 794–803. [Google Scholar] [CrossRef]
  257. Senra, M.; McCartney, S.N.; Soh, L. The effect of bio-derived additives on fatty acid methyl esters for improved biodiesel cold flow properties. Fuel 2019, 242, 719–727. [Google Scholar] [CrossRef]
  258. Leng, L.; Li, W.; Li, H.; Jiang, S.; Zhou, W. Cold flow properties of biodiesel and the improvement methods: A review. Energy Fuels 2020, 34, 10364–10383. [Google Scholar] [CrossRef]
  259. Kampars, V.; Abelniece, Z.; Lazdovica, K.; Kampare, R. Interesterification of rapeseed oil with methyl acetate in the presence of potassium tert-butoxide solution in tetrahydrofuran. Renew. Energy 2020, 158, 668–674. [Google Scholar] [CrossRef]
  260. Ang, G.T.; Tan, K.T.; Lee, K.T. Recent development and economic analysis of glycerol-free processes via supercritical fluid transesterification for biodiesel production. Renew. Sustain. Energy Rev. 2014, 31, 61–70. [Google Scholar] [CrossRef]
  261. Kralisch, D.; Staffel, C.; Ott, D.; Bensaid, S.; Saracco, G.; Bellantoni, P.; Loeb, P. Process design accompanying life cycle management and risk analysis as a decision support tool for sustainable biodiesel production. Green Chem. 2013, 15, 463–477. [Google Scholar] [CrossRef]
  262. Qadeer, M.U.; Ayoub, M.; Komiyama, M.; Daulatzai, M.U.K.; Mukhtar, A.; Saqib, S.; Ullah, S.; Qyyum, M.A.; Asif, S.; Bokhari, A. Review of Biodiesel Synthesis Technologies, Current Trends, Yield Influencing Factors and Economical Analysis of Supercritical Process. J. Clean. Prod. 2021, 309, 127388. [Google Scholar] [CrossRef]
  263. Trentini, C.P.; de Mello, B.T.F.; Postaue, N.; Stevanato, N.; Cardozo-Filho, L.; da Silva, C. Interesterification of grease trap waste lipids using methyl acetate under supercritical conditions. J. Supercrit. Fluids 2020, 164, 104896. [Google Scholar] [CrossRef]
  264. Postaue, N.; de Mello, B.T.F.; Cardozo-Filho, L.; da Silva, C. Use of the Product from Low Pressure Extraction (Crambe Seed Oil and Methyl Acetate) for Synthesis of Methyl Esters and Triacetin Under Supercritical Conditions. Eur. J. Lipid Sci. Technol. 2020, 122, 2000004. [Google Scholar] [CrossRef]
  265. Visioli, L.J.; de Castilhos, F.; da Silva, C. Use of heterogeneous acid catalyst combined with pressurized conditions for esters production from macauba pulp oil and methyl acetate. J. Supercrit. Fluids 2019, 150, 65–74. [Google Scholar] [CrossRef]
  266. Esan, A.O.; Olabemiwo, O.M.; Smith, S.M.; Ganesan, S. A concise review on alternative route of biodiesel production via interesterification of different feedstocks. Int. J. Energy Res. 2021, 45, 12614–12637. [Google Scholar] [CrossRef]
  267. Esan, A.O.; Adeyemi, A.D.; Ganesan, S. A review on the recent application of dimethyl carbonate in sustainable biodiesel production. J. Clean. Prod. 2020, 257, 120561. [Google Scholar] [CrossRef]
  268. Jung, J.-M.; Oh, J.-I.; Kwon, D.; Park, Y.-K.; Zhang, M.; Lee, J.; Kwon, E.E. Synthesis of fatty acid methyl esters via non-catalytic transesterification of avocado oil with dimethyl carbonate. Energy Convers. Manag. 2019, 195, 1–6. [Google Scholar] [CrossRef]
  269. Fiorani, G.; Perosa, A.; Selva, M. Dimethyl carbonate: A versatile reagent for a sustainable valorization of renewables. Green Chem. 2018, 20, 288–322. [Google Scholar] [CrossRef]
  270. Medrano-García, J.; Javaloyes-Antón, J.; Vázquez, D.; Ruiz-Femenia, R.; Caballero, J. Alternative carbon dioxide utilization in dimethyl carbonate synthesis and comparison with current technologies. J. CO2 Util. 2021, 45, 101436. [Google Scholar] [CrossRef]
  271. Tan, H.-Z.; Wang, Z.-Q.; Xu, Z.-N.; Sun, J.; Xu, Y.-P.; Chen, Q.-S.; Chen, Y.; Guo, G.-C. Review on the synthesis of dimethyl carbonate. Catal. Today 2018, 316, 2–12. [Google Scholar] [CrossRef]
  272. Okoye, P.U.; Longoria, A.; Sebastian, P.; Wang, S.; Li, S.; Hameed, B. A review on recent trends in reactor systems and azeotrope separation strategies for catalytic conversion of biodiesel-derived glycerol. Sci. Total Environ. 2020, 719, 134595. [Google Scholar] [CrossRef] [PubMed]
  273. Ji, Y. Recent development of heterogeneous catalysis in the transesterification of glycerol to glycerol carbonate. Catalysts 2019, 9, 581. [Google Scholar] [CrossRef] [Green Version]
  274. Algoufi, Y.; Kabir, G.; Hameed, B. Synthesis of glycerol carbonate from biodiesel by-product glycerol over calcined dolomite. J. Taiwan Inst. Chem. Eng. 2017, 70, 179–187. [Google Scholar] [CrossRef]
  275. Algoufi, Y.; Akpan, U.; Kabir, G.; Asif, M.; Hameed, B. Upgrading of glycerol from biodiesel synthesis with dimethyl carbonate on reusable Sr–Al mixed oxide catalysts. Energy Convers. Manag. 2017, 138, 183–189. [Google Scholar] [CrossRef]
  276. Okoye, P.; Abdullah, A.; Hameed, B. Glycerol carbonate synthesis from glycerol and dimethyl carbonate using trisodium phosphate. J. Taiwan Inst. Chem. Eng. 2016, 68, 51–58. [Google Scholar] [CrossRef]
  277. Wan, Y.; Lei, Y.; Lan, G.; Liu, D.; Li, G.; Bai, R. Synthesis of glycerol carbonate from glycerol and dimethyl carbonate over DABCO embedded porous organic polymer as a bifunctional and robust catalyst. Appl. Catal. A: Gen. 2018, 562, 267–275. [Google Scholar] [CrossRef]
  278. Pradhan, G.; Sharma, Y.C. Studies on green synthesis of glycerol carbonate from waste cooking oil derived glycerol over an economically viable NiMgOx heterogeneous solid base catalyst. J. Clean. Prod. 2020, 264, 121258. [Google Scholar] [CrossRef]
  279. Pradhan, G.; Sharma, Y.C. Green synthesis of glycerol carbonate by transesterification of bio glycerol with dimethyl carbonate over Mg/ZnO: A highly efficient heterogeneous catalyst. Fuel 2021, 284, 118966. [Google Scholar] [CrossRef]
  280. Notari, M.; Rivetti, F. Use of A Mixture of Esters of Fatty Acids as Fuel or Solvent. U.S. Patent 7462206, 9 December 2008. [Google Scholar]
  281. Fabbri, D.; Bevoni, V.; Notari, M.; Rivetti, F. Properties of a potential biofuel obtained from soybean oil by transmethylation with dimethyl carbonate. Fuel 2007, 86, 690–697. [Google Scholar] [CrossRef]
  282. Islam, M.R.; Kurle, Y.M.; Gossage, J.L.; Benson, T.J. Kinetics of triazabicyclodecene-catalyzed canola oil conversion to glycerol-free biofuel using dimethyl carbonate. Energy Fuels 2013, 27, 1564–1569. [Google Scholar] [CrossRef]
  283. Zhang, L.; Sheng, B.; Xin, Z.; Liu, Q.; Sun, S. Kinetics of transesterification of palm oil and dimethyl carbonate for biodiesel production at the catalysis of heterogeneous base catalyst. Bioresour. Technol. 2010, 101, 8144–8150. [Google Scholar] [CrossRef]
  284. Gu, J.; Gao, Y.; Xu, X.; Wu, J.; Yu, L.; Xin, Z.; Sun, S. Biodiesel production from palm oil and mixed dimethyl/diethyl carbonate with controllable cold flow properties. Fuel 2018, 216, 781–786. [Google Scholar] [CrossRef]
  285. Dhawan, M.S.; Yadav, G.D. Insight into a catalytic process for simultaneous production of biodiesel and glycerol carbonate from triglycerides. Catal. Today 2018, 309, 161–171. [Google Scholar] [CrossRef]
  286. Al-Saadi, L.S.; Eze, V.C.; Harvey, A.P. Techno-economic analysis of processes for biodiesel production with integrated co-production of higher added value products from glycerol. Biofuels 2019, 9, 1033. [Google Scholar] [CrossRef]
  287. Kai, T.; Mak, G.L.; Wada, S.; Nakazato, T.; Takanashi, H.; Uemura, Y. Production of biodiesel fuel from canola oil with dimethyl carbonate using an active sodium methoxide catalyst prepared by crystallization. Bioresour. Technol. 2014, 163, 360–363. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  288. Syamsuddin, Y.; Murat, M.; Hameed, B. Transesterification of Jatropha oil with dimethyl carbonate to produce fatty acid methyl ester over reusable Ca–La–Al mixed-oxide catalyst. Energy Convers. Manag. 2015, 106, 1356–1361. [Google Scholar] [CrossRef]
  289. Al-Saadi, L.S.; Eze, V.C.; Harvey, A.P. Experimental determination of optimal conditions for reactive coupling of biodiesel production with in situ glycerol carbonate formation in a triglyceride transesterification process. Front. Chem. 2018, 6, 625. [Google Scholar] [CrossRef]
  290. Lee, J.; Jung, J.-M.; Oh, J.-I.; Ok, Y.S.; Lee, S.-R.; Kwon, E.E. Evaluating the effectiveness of various biochars as porous media for biodiesel synthesis via pseudo-catalytic transesterification. Bioresour. Technol. 2017, 231, 59–64. [Google Scholar] [CrossRef]
  291. Celante, D.; Schenkel, J.V.D.; de Castilhos, F. Biodiesel production from soybean oil and dimethyl carbonate catalyzed by potassium methoxide. Fuel 2018, 212, 101–107. [Google Scholar] [CrossRef]
  292. Tang, Y.; Ren, H.; Chang, F.; Gu, X.; Zhang, J. Nano KF/Al2O3 particles as an efficient catalyst for no-glycerol biodiesel production by coupling transesterification. RSC Adv. 2017, 7, 5694–5700. [Google Scholar] [CrossRef] [Green Version]
  293. Panchal, B.; Qin, S.; Wang, J.; Bian, K.; Tao, C. Biodiesel synthesis with iron oxide nano-catalyst catalyzed Pongamia pinnata seed oil and dimethyl carbonate. Am. J. Energy Eng 2018, 6, 21–28. [Google Scholar] [CrossRef]
  294. Syamsuddin, Y.; Murat, M.; Hameed, B. Synthesis of fatty acid methyl ester from the transesterification of high-and low-acid-content crude palm oil (Elaeis guineensis) and karanj oil (Pongamia pinnata) over a calcium–lanthanum–aluminum mixed-oxides catalyst. Bioresour. Technol. 2016, 214, 248–252. [Google Scholar] [CrossRef]
  295. Ilham, Z.; Saka, S. Production of biodiesel with glycerol carbonate by non-catalytic supercritical dimethyl carbonate. Lipid Technol. 2011, 23, 10–13. [Google Scholar] [CrossRef]
  296. Tan, K.T.; Lee, K.T. A review on supercritical fluids (SCF) technology in sustainable biodiesel production: Potential and challenges. Renew. Sustain. Energy Rev. 2011, 15, 2452–2456. [Google Scholar] [CrossRef]
  297. Ilham, Z.; Saka, S. Esterification of glycerol from biodiesel production to glycerol carbonate in non-catalytic supercritical dimethyl carbonate. SpringerPlus 2016, 5, 923. [Google Scholar] [CrossRef] [Green Version]
  298. Rathore, V.; Tyagi, S.; Newalkar, B.; Badoni, R. Glycerin-free synthesis of jatropha and pongamia biodiesel in supercritical dimethyl and diethyl carbonate. Ind. Eng. Chem. Res. 2014, 53, 10525–10533. [Google Scholar] [CrossRef]
  299. Sun, S.; Zhang, L.; Meng, X.; Xin, Z. Kinetic study on lipase catalyzed trans-esterification of palm oil and dimethyl carbonate for biodiesel production. J. Renew. Sustain. Energy 2013, 5, 33127. [Google Scholar] [CrossRef]
  300. Wang, Y.; Cao, X. Enzymatic synthesis of fatty acid ethyl esters by utilizing camellia oil soapstocks and diethyl carbonate. Bioresour. Technol. 2011, 102, 10173–10179. [Google Scholar] [CrossRef]
  301. Su, E.; Du, L.; Gong, X.; Wang, P. Lipase-Catalyzed Irreversible Transesterification of Jatropha curcas L. Seed Oil to Fatty Acid Esters: An Optimization Study. J. Am. Oil Chem. Soc. 2011, 88, 793–800. [Google Scholar] [CrossRef]
  302. Gharat, N.; Rathod, V.K. Ultrasound assisted enzyme catalyzed transesterification of waste cooking oil with dimethyl carbonate. Ultrason. Sonochemistry 2013, 20, 900–905. [Google Scholar] [CrossRef]
  303. Panadare, D.; Rathod, V. Microwave assisted enzymatic synthesis of biodiesel with waste cooking oil and dimethyl carbonate. J. Mol. Catal. B Enzym. 2016, 133, S518–S524. [Google Scholar] [CrossRef]
  304. Jo, Y.J.; Lee, O.K.; Lee, E.Y. Dimethyl carbonate-mediated lipid extraction and lipase-catalyzed in situ transesterification for simultaneous preparation of fatty acid methyl esters and glycerol carbonate from Chlorella sp. KR-1 biomass. Bioresour. Technol. 2014, 158, 105–110. [Google Scholar] [CrossRef]
  305. Kim, K.H.; Lee, E.Y. Simultaneous production of transformer insulating oil and value-added glycerol carbonates from soybean oil by lipase-catalyzed transesterification in dimethyl carbonate. Energies 2018, 11, 82. [Google Scholar] [CrossRef] [Green Version]
  306. Leão, R.A.; de Souza, S.P.; Nogueira, D.O.; Silva, G.M.; Silva, M.V.; Gutarra, M.L.; Miranda, L.S.; Castro, A.M.; Junior, I.I.; de Souza, R.O. Consecutive lipase immobilization and glycerol carbonate production under continuous-flow conditions. Catal. Sci. Technol. 2016, 6, 4743–4748. [Google Scholar] [CrossRef]
  307. Ishak, Z.I.; Sairi, N.A.; Alias, Y.; Aroua, M.K.T.; Yusoff, R. A review of ionic liquids as catalysts for transesterification reactions of biodiesel and glycerol carbonate production. Catal. Rev. 2017, 59, 44–93. [Google Scholar] [CrossRef]
  308. Fan, P.; Wang, J.; Xing, S.; Yang, L.; Yang, G.; Fu, J.; Miao, C.; Lv, P. Synthesis of glycerol-free biodiesel with dimethyl carbonate over sulfonated imidazolium ionic liquid. Energy Fuels 2017, 31, 4090–4095. [Google Scholar] [CrossRef]
  309. Samorì, C.; Basaglia, M.; Casella, S.; Favaro, L.; Galletti, P.; Giorgini, L.; Marchi, D.; Mazzocchetti, L.; Torri, C.; Tagliavini, E. Dimethyl carbonate and switchable anionic surfactants: Two effective tools for the extraction of polyhydroxyalkanoates from microbial biomass. Green Chem. 2015, 17, 1047–1056. [Google Scholar] [CrossRef]
  310. Abdalla, A.O.; Liu, D. Dimethyl carbonate as a promising oxygenated fuel for combustion: A review. Energies 2018, 11, 1552. [Google Scholar] [CrossRef] [Green Version]
  311. Kim, K.H.; Lee, E.Y. Environmentally-benign dimethyl carbonate-mediated production of chemicals and biofuels from renewable bio-oil. Energies 2017, 10, 1790. [Google Scholar] [CrossRef] [Green Version]
  312. Pyo, S.-H.; Park, J.H.; Chang, T.-S.; Hatti-Kaul, R. Dimethyl carbonate as a green chemical. Curr. Opin. Green Sustain. Chem. 2017, 5, 61–66. [Google Scholar] [CrossRef]
  313. Szőri, M.; Giri, B.R.; Wang, Z.; Dawood, A.E.; Viskolcz, B.; Farooq, A. Glycerol carbonate as a fuel additive for a sustainable future. Sustain. Energy Fuels 2018, 2, 2171–2178. [Google Scholar] [CrossRef] [Green Version]
  314. Kumar, B.R.; Saravanan, S. Partially premixed low temperature combustion using dimethyl carbonate (DMC) in a DI diesel engine for favorable smoke/NOx emissions. Fuel 2016, 180, 396–406. [Google Scholar] [CrossRef]
  315. Yang, J.; Jiang, Y.; Karavalakis, G.; Johnson, K.C.; Kumar, S.; Cocker, D.R., III; Durbin, T.D. Impacts of dimethyl carbonate blends on gaseous and particulate emissions from a heavy-duty diesel engine. Fuel 2016, 184, 681–688. [Google Scholar] [CrossRef] [Green Version]
  316. Kumar, B.R.; Saravanan, S.; Rana, D.; Nagendran, A. Combined effect of injection timing and exhaust gas recirculation (EGR) on performance and emissions of a DI diesel engine fuelled with next-generation advanced biofuel–diesel blends using response surface methodology. Energy Convers. Manag. 2016, 123, 470–486. [Google Scholar] [CrossRef]
  317. Srihari, S.; Thirumalini, S. Investigation on reduction of emission in PCCI-DI engine with biofuel blends. Renew. Energy 2017, 114, 1232–1237. [Google Scholar] [CrossRef]
  318. Khalife, E.; Tabatabaei, M.; Demirbas, A.; Aghbashlo, M. Impacts of additives on performance and emission characteristics of diesel engines during steady state operation. Prog. Energy Combust. Sci. 2017, 59, 32–78. [Google Scholar] [CrossRef]
  319. Nakamura, H.; Curran, H.J.; Córdoba, A.P.; Pitz, W.J.; Dagaut, P.; Togbé, C.; Sarathy, S.M.; Mehl, M.; Agudelo, J.R.; Bustamante, F. An experimental and modeling study of diethyl carbonate oxidation. Combust. Flame 2015, 162, 1395–1405. [Google Scholar] [CrossRef] [Green Version]
  320. Shahla, R.; Togbé, C.; Thion, S.; Timothée, R.; Lailliau, M.; Halter, F.; Chauveau, C.; Dayma, G.; Dagaut, P. Burning velocities and jet-stirred reactor oxidation of diethyl carbonate. Proc. Combust. Inst. 2017, 36, 553–560. [Google Scholar] [CrossRef]
  321. Sun, W.; Huang, C.; Tao, T.; Zhang, F.; Li, W.; Hansen, N.; Yang, B. Exploring the high-temperature kinetics of diethyl carbonate (DEC) under pyrolysis and flame conditions. Combust. Flame 2017, 181, 71–81. [Google Scholar] [CrossRef] [Green Version]
  322. Panchal, B.; Chang, T.; Kang, Y.; Qin, S.; Zhao, Q.; Wang, J.; Bian, K.; Sun, Y. Synthesis of polymer based catalyst: Optimization and kinetics modeling of the transesterification of Pistacia chinensis oil with diethyl carbonate using acidic ionic liquids. Fuel 2020, 276, 118121. [Google Scholar] [CrossRef]
  323. Panchal, B.; Zhu, Z.; Qin, S.; Chang, T.; Zhao, Q.; Sun, Y.; Zhao, C.; Wang, J.; Bian, K.; Rankhamb, S. The current state applications of ethyl carbonate with ionic liquid in sustainable biodiesel production: A review. Renew. Energy 2022, 181, 341–354. [Google Scholar] [CrossRef]
  324. Esan, A.O.; Olalere, O.A.; Gan, C.-Y.; Smith, S.M.; Ganesan, S. Synthesis of biodiesel from waste palm fatty acid distillate (PFAD) and dimethyl carbonate (DMC) via Taguchi optimisation method. Biomass Bioenergy 2021, 154, 106262. [Google Scholar] [CrossRef]
  325. Luna, D.; Bautista, F.M.; Caballero, V.; Campelo, J.M.; Marinas, J.M.; Romero, A.A. Method for Producing Biodiesel Using Porcine Pancreatic Lipase as an Enzymatic Catalyst. European Patent WO2008009772A1, 24 January 2008. [Google Scholar]
  326. Caballero, V.; Bautista, F.M.; Campelo, J.M.; Luna, D.; Marinas, J.M.; Romero, A.A.; Hidalgo, J.M.; Luque, R.; Macario, A.; Giordano, G. Sustainable preparation of a novel glycerol-free biofuel by using pig pancreatic lipase: Partial 1, 3-regiospecific alcoholysis of sunflower oil. Process Biochem. 2009, 44, 334–342. [Google Scholar] [CrossRef]
  327. Verdugo, C.; Luque, R.; Luna, D.; Hidalgo, J.M.; Posadillo, A.; Sancho, E.D.; Rodriguez, S.; Ferreira-Dias, S.; Bautista, F.; Romero, A.A. A comprehensive study of reaction parameters in the enzymatic production of novel biofuels integrating glycerol into their composition. Bioresour. Technol. 2010, 101, 6657–6662. [Google Scholar] [CrossRef]
  328. Luna, C.; Sancho, E.; Luna, D.; Caballero, V.; Calero, J.; Posadillo, A.; Verdugo, C.; Bautista, F.M.; Romero, A.A. Biofuel that keeps glycerol as monoglyceride by 1,3-selective ethanolysis with pig pancreatic lipase covalently immobilized on AlPO4 support. Energies 2013, 6, 3879–3900. [Google Scholar] [CrossRef] [Green Version]
  329. Luna, D.; Posadillo, A.; Caballero, V.; Verdugo, C.; Bautista, F.M.; Romero, A.A.; Sancho, E.D.; Luna, C.; Calero, J. New biofuel integrating glycerol into its composition through the use of covalent immobilized pig pancreatic lipase. Int. J. Mol. Sci. 2012, 13, 10091–10112. [Google Scholar] [CrossRef] [Green Version]
  330. Chen, X.; Li, Z.; Chun, Y.; Yang, F.; Xu, H.; Wu, X. Effect of the Formation of Diglycerides/Monoglycerides on the Kinetic Curve in Oil Transesterification with Methanol Catalyzed by Calcium Oxide. ACS Omega 2020, 5, 4646–4656. [Google Scholar] [CrossRef]
  331. Yaşar, F. Comparision of fuel properties of biodiesel fuels produced from different oils to determine the most suitable feedstock type. Fuel 2020, 264, 116817. [Google Scholar] [CrossRef]
  332. Sharma, H.O. Production of biodiesel: Industrial, economic and energy aspects: A review. Plant Arch. 2020, 20, 2058–2066. [Google Scholar]
  333. Yesilyurt, M.K.; Cesur, C.; Aslan, V.; Yilbasi, Z. The production of biodiesel from safflower (Carthamus tinctorius L.) oil as a potential feedstock and its usage in compression ignition engine: A comprehensive review. Renew. Sustain. Energy Rev. 2020, 119, 109574. [Google Scholar] [CrossRef]
  334. Singh, N.; Kaushal, R. Outcomes of advanced biodiesel with nanoparticle additives on performance of CI engines. Mater. Today: Proc. 2021, 44, 4612–4620. [Google Scholar] [CrossRef]
  335. Knothe, G.; Razon, L.F. Biodiesel fuels. Prog. Energy Combust. Sci. 2017, 58, 36–59. [Google Scholar] [CrossRef]
  336. Hosseinzadeh-Bandbafha, H.; Tabatabaei, M.; Aghbashlo, M.; Khanali, M.; Demirbas, A. A comprehensive review on the environmental impacts of diesel/biodiesel additives. Energy Convers. Manag. 2018, 174, 579–614. [Google Scholar] [CrossRef]
  337. Paryanto, I.; Prakoso, T.; Suyono, E.A.; Gozan, M. Determination of the upper limit of monoglyceride content in biodiesel for B30 implementation based on the measurement of the precipitate in a Biodiesel–Petrodiesel fuel blend (BXX). Fuel 2019, 258, 116104. [Google Scholar] [CrossRef]
  338. Tongroon, M.; Suebwong, A.; Kananont, M.; Aunchaisri, J.; Chollacoop, N. High quality jatropha biodiesel (H-FAME) and its application in a common rail diesel engine. Renew. Energy 2017, 113, 660–668. [Google Scholar] [CrossRef]
  339. Verdugo, C.; Luna, D.; Posadillo, A.; Sancho, E.D.; Rodríguez, S.; Bautista, F.; Luque, R.; Marinas, J.M.; Romero, A.A. Production of a new second generation biodiesel with a low cost lipase derived from Thermomyces lanuginosus: Optimization by response surface methodology. Catal. Today 2011, 167, 107–112. [Google Scholar] [CrossRef]
  340. Luna, C.; Verdugo, C.; Sancho, E.D.; Luna, D.; Calero, J.; Posadillo, A.; Bautista, F.M.; Romero, A.A. Production of a biodiesel-like biofuel without glycerol generation, by using Novozym 435, an immobilized Candida antarctica lipase. Bioresour. Bioprocess. 2014, 1, 11. [Google Scholar] [CrossRef] [Green Version]
  341. Luna, C.; Verdugo, C.; Sancho, E.D.; Luna, D.; Calero, J.; Posadillo, A.; Bautista, F.M.; Romero, A.A. Biocatalytic behaviour of immobilized Rhizopus oryzae lipase in the 1,3-selective ethanolysis of sunflower oil to obtain a biofuel similar to biodiesel. Molecules 2014, 19, 11419–11439. [Google Scholar] [CrossRef] [Green Version]
  342. Escobar-Niño, A.; Luna, C.; Luna, D.; Marcos, A.T.; Cánovas, D.; Mellado, E. Selection and characterization of biofuel-producing environmental bacteria isolated from vegetable oil-rich wastes. PLoS ONE 2014, 9, e104063. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  343. Calero, J.; Luna, D.; Sancho, E.D.; Luna, C.; Verdugo, C.; Posadillo, A.; Bautista, F.M.; Romero, A.A. Achievement of a biofuel-like biodiesel by regioselective transesterification of sunflower oil with mucor miehei lipase. New Biotechnol. 2014, 31, S95. [Google Scholar] [CrossRef]
  344. Luna, C.; Verdugo, C.; Sancho, E.D.; Luna, D.; Calero, J.; Posadillo, A.; Bautista, F.M.; Romero, A.A. Enzymatic production of biodiesel that avoids glycerol as byproduct, by using immobilized rhizopus oryzae lipase. New Biotechnol. 2014, 31, S94. [Google Scholar] [CrossRef]
  345. Calero, J.; Verdugo, C.; Luna, D.; Sancho, E.D.; Luna, C.; Posadillo, A.; Bautista, F.M.; Romero, A.A. Selective ethanolysis of sunflower oil with Lipozyme RM IM, an immobilized Rhizomucor miehei lipase, to obtain a biodiesel-like biofuel, which avoids glycerol production through the monoglyceride formation. New Biotechnol. 2014, 31, 596–601. [Google Scholar] [CrossRef] [PubMed]
  346. Luna, C.; Verdugo, C.; Sancho, E.D.; Luna, D.; Calero, J.; Posadillo, A.; Bautista, F.M.; Romero, A.A. A biofuel similar to biodiesel obtained by using a lipase from Rhizopus oryzae, optimized by response surface methodology. Energies 2014, 7, 3383–3399. [Google Scholar] [CrossRef] [Green Version]
  347. Luna, C.; Luna, D.; Bautista, F.M.; Estevez, R.; Calero, J.; Posadillo, A.; Romero, A.A.; Sancho, E.D. Application of Enzymatic Extracts from a CALB Standard Strain as Biocatalyst within the Context of Conventional Biodiesel Production Optimization. Molecules 2017, 22, 2025. [Google Scholar] [CrossRef] [Green Version]
  348. Calero, J.; Luna, D.; Luna, C.; Bautista, F.M.; Hurtado, B.; Romero, A.A.; Posadillo, A.; Estevez, R. Rhizomucor miehei Lipase Supported on Inorganic Solids, as Biocatalyst for the Synthesis of Biofuels: Improving the Experimental Conditions by Response Surface Methodology. Energies 2019, 12, 831. [Google Scholar] [CrossRef] [Green Version]
  349. Luna, C.; Gascón-Pérez, V.; López-Tenllado, F.J.; Bautista, F.M.; Verdugo-Escamilla, C.; Aguado-Deblas, L.; Calero, J.; Romero, A.A.; Luna, D.; Estévez, R. Enzymatic Production of Ecodiesel by Using a Commercial Lipase CALB, Immobilized by Physical Adsorption on Mesoporous Organosilica Materials. Catalysts 2021, 11, 1350. [Google Scholar] [CrossRef]
  350. Calero, J.; Cumplido, G.; Luna, D.; Sancho, E.D.; Luna, C.; Posadillo, A.; Bautista, F.M.; Romero, A.A.; Verdugo-Escamilla, C. Production of a biofuel that keeps the glycerol as a monoglyceride by using supported KF as heterogeneous catalyst. Energies 2014, 7, 3764–3780. [Google Scholar] [CrossRef] [Green Version]
  351. 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. [Google Scholar] [CrossRef]
  352. Calero, J.; Luna, D.; Luna, C.; Bautista, F.M.; Romero, A.A.; Posadillo, A.; Estevez, R. Optimization by response surface methodology of the reaction conditions in 1, 3-selective transesterification of sunflower oil, by using CaO as heterogeneous catalyst. Mol. Catal. 2020, 484, 110804. [Google Scholar] [CrossRef]
  353. Hurtado, B.; Posadillo, A.; Luna, D.; Bautista, F.; Hidalgo, J.; Luna, C.; Calero, J.; Romero, A.; Estevez, R. Synthesis, Performance and Emission Quality Assessment of Ecodiesel from Castor Oil in Diesel/Biofuel/Alcohol Triple Blends in a Diesel Engine. Catalysts 2019, 9, 40. [Google Scholar] [CrossRef] [Green Version]
  354. Markov, V.; Kamaltdinov, V.; Devyanin, S.; Sa, B.; Zherdev, A.; Furman, V. Investigation of the influence of different vegetable oils as a component of blended biofuel on performance and emission characteristics of a diesel engine for agricultural machinery and commercial vehicles. Resources 2021, 10, 74. [Google Scholar] [CrossRef]
  355. Che Mat, S.; Idroas, M.Y.; Teoh, Y.H.; Hamid, M.F. Physicochemical, performance, combustion and emission characteristics of melaleuca cajuputi oil-refined palm oil hybrid biofuel blend. Energies 2018, 11, 3146. [Google Scholar] [CrossRef] [Green Version]
  356. Che Mat, S.; Idroas, M.; Teoh, Y.; Hamid, M. Assessment of basic properties and thermal analysis of hybrid biofuel blend. Energy Sources Part A Recovery Util. Environ. Eff. 2019, 41, 2073–2082. [Google Scholar] [CrossRef]
  357. Mat, S.C.; Idroas, M.; Teoh, Y.; Hamid, M. Optimisation of viscosity and density of refined palm Oil-Melaleuca Cajuputi oil binary blends using mixture design method. Renew. Energy 2019, 133, 393–400. [Google Scholar] [CrossRef]
  358. Vallinayagam, R.; Vedharaj, S.; Yang, W.; Roberts, W.L.; Dibble, R.W. Feasibility of using less viscous and lower cetane (LVLC) fuels in a diesel engine: A review. Renew. Sustain. Energy Rev. 2015, 51, 1166–1190. [Google Scholar] [CrossRef]
  359. Shah, P.R.; Gaitonde, U.; Ganesh, A. Influence of soy-lecithin as bio-additive with straight vegetable oil on CI engine characteristics. Renew. Energy 2018, 115, 685–696. [Google Scholar] [CrossRef]
  360. Shah, P.R.; Ganesh, A. A comparative study on influence of fuel additives with edible and non-edible vegetable oil based on fuel characterization and engine characteristics of diesel engine. Appl. Therm. Eng. 2016, 102, 800–812. [Google Scholar] [CrossRef]
  361. Purushothaman, K.; Nagarajan, G. Performance, emission and combustion characteristics of a compression ignition engine operating on neat orange oil. Renew. Energy 2009, 34, 242–245. [Google Scholar] [CrossRef]
  362. Vallinayagam, R.; Vedharaj, S.; Yang, W.; Lee, P.; Chua, K.; Chou, S. Pine oil–biodiesel blends: A double biofuel strategy to completely eliminate the use of diesel in a diesel engine. Appl. Energy 2014, 130, 466–473. [Google Scholar] [CrossRef]
  363. Subramanian, T.; Varuvel, E.G.; Martin, L.J.; Beddhannan, N. Effect of lower and higher alcohol fuel synergies in biofuel blends and exhaust treatment system on emissions from CI engine. Environ. Sci. Pollut. Res. 2017, 24, 25103–25113. [Google Scholar] [CrossRef]
  364. Panneerselvam, N.; Ramesh, M.; Murugesan, A.; Vijayakumar, C.; Subramaniam, D.; Kumaravel, A. Effect on direct injection naturally aspirated diesel engine characteristics fuelled by pine oil, Ceiba pentandra methyl ester compared with diesel. Transp. Res. Part D Transp. Environ. 2016, 48, 225–234. [Google Scholar] [CrossRef]
  365. Thiyagarajan, S.; Geo, V.E.; Martin, L.J.; Nagalingam, B. Effects of Low Carbon Biofuel Blends with Karanja (Pongamia pinnata) Oil Methyl Ester in a Single Cylinder CI Engine on CO2 Emission and other Performance and Emission Characteristics. Nat. Environ. Pollut. Technol. 2016, 15, 1249–1256. [Google Scholar]
  366. Senthil, R.; Sivakumar, E.; Silambarasan, R. Effect of di ethyl ether on the performance and emission characteristics of a diesel engine using biodiesel–eucalyptus oil blends. RSC Adv. 2015, 5, 54019–54027. [Google Scholar] [CrossRef]
  367. Vallinayagam, R.; Vedharaj, S.; Yang, W.; Lee, P.; Chua, K.; Chou, S. Combustion performance and emission characteristics study of pine oil in a diesel engine. Energy 2013, 57, 344–351. [Google Scholar] [CrossRef]
  368. Tamilselvan, P.; Nallusamy, N. Performance, combustion and emission characteristics of a compression ignition engine operating on pine oil. Biofuels 2015, 6, 273–281. [Google Scholar] [CrossRef]
  369. Subramanian, T.; Varuvel, E.G.; Ganapathy, S.; Vedharaj, S.; Vallinayagam, R. Role of fuel additives on reduction of NO X emission from a diesel engine powered by camphor oil biofuel. Environ. Sci. Pollut. Res. 2018, 25, 15368–15377. [Google Scholar] [CrossRef]
  370. Mat, S.C.; Idroas, M.; Hamid, M.; Zainal, Z. Performance and emissions of straight vegetable oils and its blends as a fuel in diesel engine: A review. Renew. Sustain. Energy Rev. 2018, 82, 808–823. [Google Scholar] [CrossRef]
  371. Coughlin, B.; Hoxie, A. Combustion characteristics of ternary fuel Blends: Pentanol, butanol and vegetable oil. Fuel 2017, 196, 488–496. [Google Scholar] [CrossRef]
  372. Atmanli, A.; Ileri, E.; Yuksel, B.; Yilmaz, N. Extensive analyses of diesel–vegetable oil–n-butanol ternary blends in a diesel engine. Appl. Energy 2015, 145, 155–162. [Google Scholar] [CrossRef]
  373. Atmanlı, A.; Ileri, E.; Yüksel, B. Effects of higher ratios of n-butanol addition to diesel–vegetable oil blends on performance and exhaust emissions of a diesel engine. J. Energy Inst. 2015, 88, 209–220. [Google Scholar] [CrossRef]
  374. Zhu, L.; Xiao, Y.; Cheung, C.S.; Guan, C.; Huang, Z. Combustion, gaseous and particulate emission of a diesel engine fueled with n-pentanol (C5 alcohol) blended with waste cooking oil biodiesel. Appl. Therm. Eng. 2016, 102, 73–79. [Google Scholar] [CrossRef] [Green Version]
  375. Yilmaz, N.; Vigil, F.M. Potential use of a blend of diesel, biodiesel, alcohols and vegetable oil in compression ignition engines. Fuel 2014, 124, 168–172. [Google Scholar] [CrossRef]
  376. Atmanli, A. Effects of a cetane improver on fuel properties and engine characteristics of a diesel engine fueled with the blends of diesel, hazelnut oil and higher carbon alcohol. Fuel 2016, 172, 209–217. [Google Scholar] [CrossRef]
  377. Atmanli, A.; Ileri, E.; Yilmaz, N. Optimization of diesel–butanol–vegetable oil blend ratios based on engine operating parameters. Energy 2016, 96, 569–580. [Google Scholar] [CrossRef]
  378. Krishnamoorthy, V.; Dhanasekaran, R.; Rana, D.; Saravanan, S.; Kumar, B.R. A comparative assessment of ternary blends of three bio-alcohols with waste cooking oil and diesel for optimum emissions and performance in a CI engine using response surface methodology. Energy Convers. Manag. 2018, 156, 337–357. [Google Scholar] [CrossRef]
  379. Kumar, N.; Bansal, S.; Pali, H.S. Blending of Higher Alcohols with Vegetable Oil-Based Fuels for Use in Compression Ignition Engine; SAE Technical Paper; SAE International: Warrendale, PA, USA, 2015; pp. 1–958. [Google Scholar]
  380. Kommana, S.; Naik Banoth, B.; Radha Kadavakollu, K. Eucalyptus-palm kernel oil blends: A complete elimination of diesel in a 4-stroke VCR diesel engine. J. Combust. 2015, 2015, 182879. [Google Scholar] [CrossRef] [Green Version]
  381. Rakopoulos, D.; Rakopoulos, C.; Giakoumis, E.; Dimaratos, A. Studying combustion and cyclic irregularity of diethyl ether as supplement fuel in diesel engine. Fuel 2013, 109, 325–335. [Google Scholar] [CrossRef]
  382. Krishnamoorthi, M.; Malayalamurthi, R. A review on effect of diethyl ether additive on combustion, performance and emission characteristics of a diesel and biodiesel/vegetable oil fuelled engine. Adv. Nat. Appl. Sci. 2016, 10, 9–18. [Google Scholar]
  383. Kumar, A.; Rajan, K.; Kumar, K.S.; Maiyappan, K.; Rasheed, U.T. Green fuel utilization for diesel engine, combustion and emission analysis fuelled with CNSO diesel blends with Diethyl ether as additive. In IOP Conference Series: Materials Science and Engineering; IOP Publishing: Bristol, UK, 2017; p. 12013. [Google Scholar]
  384. Rakopoulos, D.C.; Rakopoulos, C.D.; Kyritsis, D.C. Butanol or DEE blends with either straight vegetable oil or biodiesel excluding fossil fuel: Comparative effects on diesel engine combustion attributes, cyclic variability and regulated emissions trade-off. Energy 2016, 115, 314–325. [Google Scholar] [CrossRef]
  385. Rakopoulos, D.C.; Rakopoulos, C.D.; Giakoumis, E.G. Impact of properties of vegetable oil, bio-diesel, ethanol and n-butanol on the combustion and emissions of turbocharged HDDI diesel engine operating under steady and transient conditions. Fuel 2015, 156, 1–19. [Google Scholar] [CrossRef]
  386. Krishnamoorthi, M.; Malayalamurthi, R. Experimental investigation on performance, emission behavior and exergy analysis of a variable compression ratio engine fueled with diesel-aegle marmelos oil-diethyl ether blends. Energy 2017, 128, 312–328. [Google Scholar] [CrossRef]
  387. Krishnamoorthi, M.; Malayalamurthi, R. Experimental investigation on the availability, performance, combustion and emission distinctiveness of bael oil/diesel/diethyl ether blends powered in a variable compression ratio diesel engine. Heat Mass Transf. 2018, 54, 2023–2044. [Google Scholar] [CrossRef]
  388. Aguado-Deblas, L.; Hidalgo-Carrillo, J.; Bautista, F.M.; Luna, D.; Luna, C.; Calero, J.; Posadillo, A.; Romero, A.A.; Estevez, R. Diethyl ether as an oxygenated additive for fossil diesel/vegetable oil blends: Evaluation of performance and emission quality of triple blends on a diesel engine. Energies 2020, 13, 1542. [Google Scholar] [CrossRef] [Green Version]
  389. Aguado-Deblas, L.; Hidalgo-Carrillo, J.; Bautista, F.M.; Luna, D.; Luna, C.; Calero, J.; Posadillo, A.; Romero, A.A.; Estevez, R. Acetone prospect as an additive to allow the use of castor and sunflower oils as drop-in biofuels in diesel/acetone/vegetable oil triple blends for application in diesel engines. Molecules 2020, 25, 2935. [Google Scholar] [CrossRef]
  390. Dhanarasu, M.; RameshKumar, K.; Maadeswaran, P. Effect of Acetone as an oxygenated additive with used sunflower oil biodiesel on performance, combustion and emission in diesel engine. Environ. Technol. 2021, 1–26. [Google Scholar] [CrossRef]
  391. Aguado-Deblas, L.; Estevez, R.; Hidalgo-Carrillo, J.; Bautista, F.M.; Luna, C.; Calero, J.; Posadillo, A.; Romero, A.A.; Luna, D. Outlook for direct use of sunflower and castor oils as biofuels in compression ignition Diesel engines, being part of diesel/ethyl acetate/straight vegetable oil triple blends. Energies 2020, 13, 4836. [Google Scholar] [CrossRef]
  392. Contino, F.; Foucher, F.; Mounaim-Rousselle, C.; Jeanmart, H. Experimental characterization of ethyl acetate, ethyl propionate, and ethyl butanoate in a homogeneous charge compression ignition engine. Energy Fuels 2011, 25, 998–1003. [Google Scholar] [CrossRef]
  393. Jones, R. Ethyl Acetate as Fuel or Fuel Additive. U.S. Patent US20110296744A1, 8 December 2011. [Google Scholar]
  394. Gangwar, J.N.; Saraswati, S.; Agarwal, S. Performance and emission improvement analysis of CI engine using various additive based diesel fuel. In Proceedings of the 2017 International Conference on Advances in Mechanical, Industrial, Automation and Management Systems (AMIAMS), Prayagraj, India, 3 February 2017; pp. 189–195. [Google Scholar]
  395. Wu, S.; Yang, H.; Hu, J.; Shen, D.; Zhang, H.; Xiao, R. The miscibility of hydrogenated bio-oil with diesel and its applicability test in diesel engine: A surrogate (ethylene glycol) study. Fuel Process. Technol. 2017, 161, 162–168. [Google Scholar] [CrossRef]
  396. Zhang, L.; Shen, C.; Liu, R. GC–MS and FT-IR analysis of the bio-oil with addition of ethyl acetate during storage. Front. Energy Res. 2014, 2, 3. [Google Scholar] [CrossRef]
  397. Ren, Y.; Huang, Z.; Miao, H.; Di, Y.; Jiang, D.; Zeng, K.; Liu, B.; Wang, X. Combustion and emissions of a DI diesel engine fuelled with diesel-oxygenate blends. Fuel 2008, 87, 2691–2697. [Google Scholar] [CrossRef]
  398. Arteconi, A.; Mazzarini, A.; Di Nicola, G. Emissions from ethers and organic carbonate fuel additives: A review. Water Air Soil Pollut. 2011, 221, 405–423. [Google Scholar] [CrossRef]
  399. Rao, P.V.; Ramesh, S.; Kumar, S.A. Effects of Oxygenated Additives with Diesel on the Performance of DI Diesel Engine. J. Energy Res. Rev. 2019, 2, 1–9. [Google Scholar] [CrossRef]
  400. Kozak, M.; Merkisz, J.; Bielaczyc, P.; Szczotka, A. The influence of oxygenated diesel fuels on a diesel vehicle PM/NO x emission trade-off. SAE Tech. Pap. 2009, 1, 2696. [Google Scholar] [CrossRef]
  401. Bridjesh, P.; Geetha, N. Effect of Diethyl Carbonate as Additive to Waste Plastic Oil on Performance and Emission of a Diesel Engine. Orient. J. Chem 2020, 36, 189–194. [Google Scholar] [CrossRef]
  402. Anugraha, R.; Tetrisyanda, R.; Altway, A.; Wibawa, G. The Effects of Diethyl Carbonate in Light Naphtha Blending to Utilize New Energy Resource. In IOP Conference Series: Materials Science and Engineering; IOP Publishing: Bristol, UK, 2021; p. 12057. [Google Scholar]
  403. Shukla, K.; Srivastava, V.C. Diethyl carbonate: Critical review of synthesis routes, catalysts used and engineering aspects. RSC Adv. 2016, 6, 32624–32645. [Google Scholar] [CrossRef]
  404. Aguado-Deblas, L.; Hidalgo-Carrillo, J.; Bautista, F.M.; Luna, C.; Calero, J.; Posadillo, A.; Romero, A.A.; Luna, D.; Estévez, R. Biofuels from diethyl carbonate and vegetable oils for use in triple blends with diesel fuel: Effect on performance and smoke emissions of a Diesel engine. Energies 2020, 13, 6584. [Google Scholar] [CrossRef]
  405. Venkanna, B.; Reddy, C.V. Performance, emission and combustion characteristics of DI diesel engine running on blends of honne oil/diesel fuel/kerosene/DMC. Int. J. Agric. Biol. Eng. 2011, 4, 48–57. [Google Scholar]
  406. Kasiraman, G.; Geo, V.E.; Nagalingam, B. Assessment of cashew nut shell oil as an alternate fuel for CI (Compression ignition) engines. Energy 2016, 101, 402–410. [Google Scholar] [CrossRef]
  407. Aguado-Deblas, L.; Hidalgo-Carrillo, J.; Bautista, F.M.; Luna, C.; Calero, J.; Posadillo, A.; Romero, A.A.; Luna, D.; Estévez, R. Evaluation of Dimethyl Carbonate as Alternative Biofuel. Performance and Smoke Emissions of a Diesel Engine Fueled with Diesel/Dimethyl Carbonate/Straight Vegetable Oil Triple Blends. Sustainability 2021, 13, 1749. [Google Scholar] [CrossRef]
  408. Sharan, P.; Bhaskaran, A.; Kalpana, S.; Ramesh, B. Effect of dimethyl carbonate on performance and emission characteristics of a diesel engine. Int. J. Curr. Res. Rev. 2018, 10, 116–121. [Google Scholar]
  409. Mei, D.; Hielscher, K.; Baar, R. Study on combustion process and emissions of a single-cylinder diesel engine fueled with DMC/diesel blend. J. Energy Eng. 2014, 140, 4013004. [Google Scholar] [CrossRef]
  410. Pan, M.; Qian, W.; Huang, R.; Tong, C.; Huang, H.; Xu, L.; Hao, B. Effects of dimethyl carbonate and 2-ethylhexyl nitrate on energy distribution, combustion and emissions in a diesel engine under different load conditions. Energy Convers. Manag. 2019, 199, 111985. [Google Scholar] [CrossRef]
  411. Lü, X.-c.; Yang, J.-g.; Zhang, W.-g.; Huang, Z. Improving the combustion and emissions of direct injection compression ignition engines using oxygenated fuel additives combined with a cetane number improver. Energy Fuels 2005, 19, 1879–1888. [Google Scholar] [CrossRef]
  412. Nayak, S.K.; Mishra, P.C. Application of neem biodiesel and dimethyl carbonate as alternative fuels. Energy Sources Part A Recovery Util. Environ. Eff. 2017, 39, 284–290. [Google Scholar] [CrossRef]
  413. Jayabal, R.; Thangavelu, L.; Subramani, S. Combined effect of oxygenated additives, injection timing and EGR on combustion, performance and emission characteristics of a CRDi diesel engine powered by sapota biodiesel/diesel blends. Fuel 2020, 276, 118020. [Google Scholar] [CrossRef]
  414. Lakshminarayanan, A.; Olsen, D.B.; Cabot, P.E. Performance and emission evaluation of triglyceride-gasoline blends in agricultural compression ignition engines. Appl. Eng. Agric. 2014, 30, 523–534. [Google Scholar]
  415. Estevez, R.; Aguado-Deblas, L.; Posadillo, A.; Hurtado, B.; Bautista, F.M.; Hidalgo, J.M.; Luna, C.; Calero, J.; Romero, A.A.; Luna, D. Performance and emission quality assessment in a diesel engine of straight castor and sunflower vegetable oils, in diesel/gasoline/oil triple blends. Energies 2019, 12, 2181. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Different alternative methods of transforming animal fats and vegetable oils into biofuels, avoiding the generation of glycerol, adapted with permission of Ref. [24]. Copyright, 2014 Elsevier.
Figure 1. Different alternative methods of transforming animal fats and vegetable oils into biofuels, avoiding the generation of glycerol, adapted with permission of Ref. [24]. Copyright, 2014 Elsevier.
Energies 15 03173 g001
Figure 2. Publications found in the Web of Science database by the keywords “biodiesel” separated by document types from year 2000 to 2021.
Figure 2. Publications found in the Web of Science database by the keywords “biodiesel” separated by document types from year 2000 to 2021.
Energies 15 03173 g002
Figure 3. Standard flowchart of an alkali transesterification process in a conventional biodiesel plant, reproduced with permission of Ref. [68]. Copyright 2019 Elsevier.
Figure 3. Standard flowchart of an alkali transesterification process in a conventional biodiesel plant, reproduced with permission of Ref. [68]. Copyright 2019 Elsevier.
Energies 15 03173 g003
Figure 4. Different chemicals obtained to valorize crude glycerol generated in the industrial production of biodiesel, reproduced with permission of Ref. [78]. Copyright 2020 Elsevier.
Figure 4. Different chemicals obtained to valorize crude glycerol generated in the industrial production of biodiesel, reproduced with permission of Ref. [78]. Copyright 2020 Elsevier.
Energies 15 03173 g004
Figure 5. Current scenario of green diesel production worldwide. Source: Futurebridge analysis; EC Reports.
Figure 5. Current scenario of green diesel production worldwide. Source: Futurebridge analysis; EC Reports.
Energies 15 03173 g005
Figure 6. General scheme for the pyrolysis process of triglycerides.
Figure 6. General scheme for the pyrolysis process of triglycerides.
Energies 15 03173 g006
Figure 7. General saturated fatty acid deoxygenation reaction steps under inert atmosphere over supported metal catalyst, adapted with permission of Ref. [102]. Copyright 2015, Elsevier.
Figure 7. General saturated fatty acid deoxygenation reaction steps under inert atmosphere over supported metal catalyst, adapted with permission of Ref. [102]. Copyright 2015, Elsevier.
Energies 15 03173 g007
Figure 8. General scheme conversion of triglycerides to hydrocarbons in the absence of hydrogen by using a fluid catalytic cracking (FCC) process involving hydrogen transfer reactions, “Reprinted (adapted) with permission from Shimada et al. Ind. Eng. Chem. Res. 2017, 56, 75–86. Copyright 2017 American Chemical Society [103].
Figure 8. General scheme conversion of triglycerides to hydrocarbons in the absence of hydrogen by using a fluid catalytic cracking (FCC) process involving hydrogen transfer reactions, “Reprinted (adapted) with permission from Shimada et al. Ind. Eng. Chem. Res. 2017, 56, 75–86. Copyright 2017 American Chemical Society [103].
Energies 15 03173 g008
Figure 9. Reaction pathways of triglycerides reactions over hydrotreating catalysts, reproduced with permission of Ref. [126]. Copyright, 2014 Elsevier.
Figure 9. Reaction pathways of triglycerides reactions over hydrotreating catalysts, reproduced with permission of Ref. [126]. Copyright, 2014 Elsevier.
Energies 15 03173 g009
Figure 10. Green diesel production process by hydrotreatment of vegetable oils in a biomass hydroprocessing plant. Adapted from [90].
Figure 10. Green diesel production process by hydrotreatment of vegetable oils in a biomass hydroprocessing plant. Adapted from [90].
Energies 15 03173 g010
Figure 11. A schematic view on the transformation of triglycerides into hydrocarbons under hydrotreating conditions.
Figure 11. A schematic view on the transformation of triglycerides into hydrocarbons under hydrotreating conditions.
Energies 15 03173 g011
Figure 12. Publications found in the Web of Science database by the keywords “green diesel” from year 2000 to 2022. (Blue triangle indicates the number of publication found with the keywords “biofuel in aviation”).
Figure 12. Publications found in the Web of Science database by the keywords “green diesel” from year 2000 to 2022. (Blue triangle indicates the number of publication found with the keywords “biofuel in aviation”).
Energies 15 03173 g012
Figure 13. General reaction scheme of Gliperol® production, obtained by interesterification reaction of a triglyceride molecule with methyl acetate, obtained by a conventional catalyst, producing a mixture of FAMEs and one molecule of glycerol triacetate (triacetin).
Figure 13. General reaction scheme of Gliperol® production, obtained by interesterification reaction of a triglyceride molecule with methyl acetate, obtained by a conventional catalyst, producing a mixture of FAMEs and one molecule of glycerol triacetate (triacetin).
Energies 15 03173 g013
Figure 14. Reaction scheme of biodiesel-like biofuel “DMC-BioDs”, patented by Polimeri Europa (Italy). Obtained by reacting oils with DMC under alkaline conditions, obtaining a mixture of 2 mol of FAMEs and 1 mol of FAGC.
Figure 14. Reaction scheme of biodiesel-like biofuel “DMC-BioDs”, patented by Polimeri Europa (Italy). Obtained by reacting oils with DMC under alkaline conditions, obtaining a mixture of 2 mol of FAMEs and 1 mol of FAGC.
Energies 15 03173 g014
Figure 15. Reaction scheme of biodiesel-like biofuel “Ecodiesel-100®”, obtained by enzymatic technology patented by the University of Cordoba (UCO).
Figure 15. Reaction scheme of biodiesel-like biofuel “Ecodiesel-100®”, obtained by enzymatic technology patented by the University of Cordoba (UCO).
Energies 15 03173 g015
Table 1. Comparison between green diesel, biodiesel, and fossil diesel.
Table 1. Comparison between green diesel, biodiesel, and fossil diesel.
PropertiesGreen DieselBiodiesel EN 14214Fossil Diesel
Cetane number75–9050–6540–55
Energy density, MJ/kg443843
Density, g/mL0.780.880.83–0.85
Cloud Point, °C−1020−5
Lubricity a>700-226–354
Energy Content, BTU/gal123 K118 K129
Sulfur<10 ppm<5 ppm<10 ppm
NOx emissions b−10 to 0+10Baseline
Viscosity, mm2/s2–42.9–111.9–4.1
Global warming, gCO2eq/MJ−7.3261.3579.93
Acidification, gSO2eq/MJ0.3960.70.547
Ozone layer depletion, mgCFC-11eq/MJ 0.0030.0060.012
a Measure by wear scare diameter (µm); b Percentage in comparison to fossil diesel. Data taken from [90,168,169,170].
Table 2. Summary sheet of the pros and cons of different existing methodologies for obtaining biodiesel-like biofuels, integrating the glycerol as a derivative in the same transesterification process. Data taken from references of Section 4.1, Section 4.2 and Section 4.3.
Table 2. Summary sheet of the pros and cons of different existing methodologies for obtaining biodiesel-like biofuels, integrating the glycerol as a derivative in the same transesterification process. Data taken from references of Section 4.1, Section 4.2 and Section 4.3.
Biodiesel EN 14214Biodiesel-like Biofuels
GliperolDMC-BioEcodiesel®
ReactiveMethanol or ethanolMethyl acetateMethyl carbonateMethanol or ethanol
CatalystNaOH or KOHAcid, basic, or lipasesBasic or lipasesLipases
Products3 FAME or 3 FAEEGlycerol triacetate + 3 FAMEFatty acid glycerol carbonateMonoglycerides + 2 FAEE
By-productsGlycerolNo wasteNo wasteNo Waste
Separation process and cleaningComplexNot neededNot neededNot needed
Investments facilitiesMediumLowLowLow
Free fatty acids and/or water in the starting oilFree fatty acids are transformed to soapsFree fatty acids are transformed to biofuelFree fatty acids are transformed to biofuelFree fatty acids are transformed to biofuel
Catalyst costLowLowLowLow
Environmental impactHigh. Alkaline and saline effluents are generated. Wastewater treatment is neededLowLowLow
Table 3. Summary sheet of the main technologies currently available to produce renewable liquid fuels from vegetable oils, able to operate correctly in current internal combustion engines, as well as in spark-ignition engines or/and aviation.
Table 3. Summary sheet of the main technologies currently available to produce renewable liquid fuels from vegetable oils, able to operate correctly in current internal combustion engines, as well as in spark-ignition engines or/and aviation.
Parameters for ComparisonType of Biofuel
BiodieselBiodiesel-like BiofuelGreen DieselLVLC Blended with Vegetable Oils
Atomic efficiency85%100%85%100%
By-products/waste generationDirty glycerol (15%)No wastesCO, CO2, and H2O (15%)No wastes
Cleaning processComplex, high-water consumptionNot neededNot neededNot needed
Cetane indexSlightly lower than dieselSlightly lower than dieselLike dieselSlightly lower than diesel
LubricityHighHighLowHigh
Industrial productionComplexSimpleSimpleVery simple
Environmental impactHighLowLowNone
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Estevez, R.; Aguado-Deblas, L.; López-Tenllado, F.J.; Luna, C.; Calero, J.; Romero, A.A.; Bautista, F.M.; Luna, D. Biodiesel Is Dead: Long Life to Advanced Biofuels—A Comprehensive Critical Review. Energies 2022, 15, 3173. https://doi.org/10.3390/en15093173

AMA Style

Estevez R, Aguado-Deblas L, López-Tenllado FJ, Luna C, Calero J, Romero AA, Bautista FM, Luna D. Biodiesel Is Dead: Long Life to Advanced Biofuels—A Comprehensive Critical Review. Energies. 2022; 15(9):3173. https://doi.org/10.3390/en15093173

Chicago/Turabian Style

Estevez, Rafael, Laura Aguado-Deblas, Francisco J. López-Tenllado, Carlos Luna, Juan Calero, Antonio A. Romero, Felipa M. Bautista, and Diego Luna. 2022. "Biodiesel Is Dead: Long Life to Advanced Biofuels—A Comprehensive Critical Review" Energies 15, no. 9: 3173. https://doi.org/10.3390/en15093173

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

Article Metrics

Back to TopTop