State of the art of biodiesel production processes: a review of the heterogeneous catalyst

A. M. Ruhul*a, M. A. Kalama, H. H. Masjuki*a, I. M. Rizwanul Fattahab, S. S. Rehama and M. M. Rasheda
aCentre for Energy Sciences, Department of Mechanical Engineering, Faculty of Engineering, University of Malaya, 50603 Kuala Lumpur, Malaysia. E-mail: ruhulamin07ruet@gmail.com; masjuki@um.edu.my; Fax: +60 3 79675317; Tel: +60 3 79674448
bSchool of Mechanical and Manufacturing Engineering, University of New South Wales, Kensington, Australia

Received 26th May 2015 , Accepted 23rd October 2015

First published on 30th October 2015


Abstract

A broadened focus on energy, the fast growing value of petroleum oil, harmful atmospheric emissions because of the evolution of greenhouse gases, natural contamination, and quick reduction approaches to obtain fossil fuels are critical factors in the search for alternative energy sources. The need for developing renewable energy sources with fewer environmental effects is increasing because of the problems caused by the extensive use of fossil fuels. Currently, creating energy from low-carbon origins and introducing eco-friendly modern technology are the main targets of researchers in the field. Biodiesel has been identified as an alternative renewable liquid fuel source that can be derived through thermal cracking, esterification and transesterification of different triglycerides. Among these processes, the most popular and convenient technique for biodiesel production is the transesterification of triglyceride with the help of a suitable alcohol and a catalyst. Many scientists have introduced different types of catalysts to optimize the reaction conditions and the biodiesel production yields. Catalyst selection involves the determination of the water content and the free fatty acids in the oil. Homogeneous base catalysts provide faster reaction rates than homogeneous acid catalysts. Recently researchers have paid attention to heterogeneous catalysts because of their high activity, high selectivity, catalyst recovery, reusability, easy separation from the products, and water tolerance properties. Biocatalysts present significant advantages in terms of environmental issues over conventional alkali-catalyzed processes. This review article focuses on various technologies used for biodiesel production, as well as the benefits and limitations of the different types of catalysts in the relevant production technologies. We also conduct a comparative study of biocatalysts and homogeneous and heterogeneous catalysts in biodiesel production technologies at the laboratory scale, as well as their industrial applications.


1. Introduction

The constantly growing expense of traditional fossil fuels and the related natural effects of their use are significant concerns worldwide. The use of fuel increases daily given the high demand for energy. This demand is mostly met by fossil fuels such as coal, petroleum oils, and natural gas. Renewable sources are the most preferred alternative energy sources. Energy can be extracted from fossil fuels economically as well as in large quantities. Thus, scientists worldwide are working to enhance the development of the fuel mileage and the emission quality of internal combustion (IC) engines. Researchers have also concentrated on investigating alternative energy sources that can be utilized in IC engines without demonstrative changes in vehicle design. Low carbon sources are preferred for the production of energy and represent an eco-friendly green technology.

Considering the increase in global population, additional resources are needed to deliver energy for human consumption. To fulfill additional demands, researchers are considering renewable energy. The term biodiesel implies the substitution of traditional energy sources with an inexhaustible liquid fuel, which can be obtained from triglycerides and supplements the additional requirements of conventional petroleum diesel.1 Biodiesel has recently become a popular research field because of its renewability, biodegradability, nontoxicity, and carbon neutrality. The transesterification process proceeds by adding triglyceride to methanol, ethanol, or any desirable alcohol, which is used to develop biodiesel.2–4 Biodiesel can also be produced domestically from vegetable oils, animal fats, micro and macro algal oil5 or used cooking oil. The biodiesel produced can be treated as a cleaner-burning substitute for conventional petroleum diesel. Biodiesel contains mono alkyl esters of long chain fatty acids6 as well as methyl esters. These methyl esters include plant seed oils, animal fats, or even waste cooking oils and are produced by transesterification with methanol. Acids, bases, and enzymes catalyze the transesterification reactions. Heterogeneous catalysts are promising materials for the synthesis of biodiesel from different feedstocks.7 Biodiesel can fulfill the requirements for additional energy. Biodiesel is an interesting product because of its natural advantages and production from renewable assets. Unfortunately, high costs and the constrained accessibility of fat and oil resources limit the wider use of this alternative energy source. Biodiesel cost may be viewed from two aspects: the cost of the raw materials (fats and oils) and the processing cost. The expense of the crude materials represents 60–75% of the cost of aggregate biodiesel fuel.8 Utilization of waste cooking oil may significantly reduce the cost of the raw materials but also reduces the fuel quality.9 Studies are expected to discover a less expensive approach to utilize used cooking oils to produce biodiesel fuel. A continuous transesterification procedure may allow the reduction of the production cost associated with biodiesel. Feedstock obtainability, types, conversion technology, catalyst use, and process cost contribute to the total biodiesel production expenses.5

This review provides knowledge on the different biodiesel production technologies using different catalytic and non-catalytic processes. The use of biocatalysts to synthesize biodiesel is also discussed, and a comparative study of the conventional biodiesel production processes is performed. In addition to the catalyst preparation, the roles and effects of the different categories of heterogeneous and homogeneous catalysts on the esterification process of biodiesel production are described. This study also describes the advantages of non-catalytic biodiesel production processes over catalytic processes.

2. Biodiesel and its production processes

Several techniques are available to obtain biodiesel from different feedstocks. For example, transesterification via radio frequency, microwave, ultra-sonication, alcohol reflex temperatures, and alcohol supercritical temperatures are possible. All biodiesel production techniques can be classified as pyrolysis or thermal cracking, esterification, or transesterification processes.

Petro diesel, widely known as diesel, is obtained through the fractional distillation of crude oil. Petro diesel contains hydrocarbon molecules that range in size from 8 to 21 carbon atoms. An ordinary petro diesel containing 16 carbon atoms is presented in Fig. 1(a). A petro diesel molecule is composed of a pure hydrocarbon, that is, a molecule containing only hydrogen and carbon, with no oxygen molecules. Thus, after proper burning in atmospheric air, only CO2 and H2O are released from this molecule. Sometimes, hydrogen sulfide (H2S) is also produced because of the presence of sulfur (S) in diesel. Typically, biodiesels contain long chain carbon molecules with hydrogen atoms, similar to petro diesel, with an additional ester functional group (–COOR). Biodiesel with 17 or 16 carbons with an ester group is illustrated in Fig. 1(b). Vegetable oil also typically contains long rows of carbon and hydrogen atoms with ester functional groups. Vegetable oil molecules are almost three times larger than normal diesel molecules. This large-sized structure is known as a triglyceride. The atomic size and structure of vegetable oil make it gel in cold weather, which means its direct use in engines is difficult. The ordinary atomic structure of vegetable oil is shown in Fig. 1(c).


image file: c5ra09862a-f1.tif
Fig. 1 Molecular structure of (a) petroleum diesel (b) biodiesel and (c) vegetable oil.

Triglycerides are initially reduced to diglycerides, which are then reduced to monoglycerides. Monoglycerides are finally reduced to fatty acid esters. The gradual reaction mechanism producing the monoglycerides from triglycerides10 or vegetable oils is shown in Fig. 2, where R represents an alkyl group; R1, R2, and R3 are the fatty acid chains; and k1, k2, k3, k4, k5, and k6 represent the catalysts.


image file: c5ra09862a-f2.tif
Fig. 2 Step by step triglyceride to monoglyceride production of vegetable oils.

Biodiesel can be produced in single or two-step reactions depending on the feedstock quality, for example, the water content and if they contain FFAs. Higher FFA content indicates a more acidic feedstock. Crude oil with a high acid value is first esterified with an acid catalyst and then transesterified with a suitable base catalyst. However, a large amount of wastewater is associated with this technique when a homogeneous catalyst used; thus, the process presents some harm to the environment. Laboratory-scale heterogeneous and bio-catalytic processes can minimize this problem. A flow diagram of the laboratory-scale or fixed-bed or single-step transesterification biodiesel production process is presented in Fig. 3(a). Researchers have attempted to utilize not only the main oil sources (such as seeds) but also the dry waste shell of the seeds. Other value-added products associated with biodiesel production have attracted research attention.


image file: c5ra09862a-f3.tif
Fig. 3 (a) Biodiesel production flow chart for laboratory-scale processes (single step). (b) Biodiesel production flow chart for industrial scale processes (continuous).

Maiti et al.11 introduced an integrated system where power was generated from dry jatropha seed shells through gasification. Then, utilizing the gasified producer gas, power was generated with the help of a producer gas engine. This electric power was partially utilized as an energy source in different steps of the integrated biodiesel production system from Jatropha curcas seeds. The relevant steps include screw pressing, oil refining, transesterification, glycerol purification, and soap making. The authors found that 8 h of continuous operation of the gasification process with a 64.8% efficiency can generate 10 kW of captive power with a 24.5% efficiency; here, the producer gas heating value was considered to be 5.2 MJ m−3 and the calorific value of the empty shells was 17.2 MJ kg−1. Ghosh et al.12 proposed an integrated process to produce oil-bearing Chlorella variabilis for lipid extrication utilizing a by-product of jatropha methyl ester production. Later, Ghosh et al.13 described an improved and integrated process to prepare fatty acid methyl ester (FAME) from whole seeds of J. curcas with the least energy use and zero effluent discharge; crude glycerol utilization was also integrated in this work. An integrated continuous Jatropha curcas biodiesel production process with its other value added product recovery techniques is represented in Fig. 3(b).

The industrially produced pure biodiesel is not exactly the same as petroleum diesel, with a small variation in some properties. The properties of petroleum diesel and the produced biodiesel according to the American Society for Testing and Materials and the European Standards are shown in Table 1.

Table 1 Comparison of the properties of biodiesel and petroleum diesel based on the ASTM and European Standards (EN)
Property Units Diesel Biodiesel (B100)
ASTM D975 EN 590 ASTM D6751 EN 14214
Density @ 15 °C (59 °F) kg m−3 850 835 800–900 860–900
Kinematic viscosity @ 40 °C mm2 s−1 1.3–4.1 3.5 1.9–6.0 3.5–5.0
Lower heating value kJ kg−1 43[thin space (1/6-em)]000
Cetane no. 40–55 53 48–65 Min. 51
Specific gravity @ 15 °C g cm−3 0.85 0.88
Carbon % mass 87 77
Hydrogen % mass 13 12
Oxygen, by dif. % mass 0 0 11
Sulfur content ppm 500 Max. 10 (mg kg−1) Max. 0.05 Max. 0.0010
Boiling point °C 180 to 340 315 to 350
Flash point °C 60 to 80 Min. 55 100 to 170 Min. 120
Cloud point °C −15 to 5 −5 −3 to 12
Pour point °C −35 to −15 −15 to 16
Cold flow plugging point °C Max. +5
Lubricity (HFRR) μm 300–600 Max. 300
Water content mg kg−1 Max. 500
Acid value mg KOH per g Max. 0.80 Max. 0.50


2.1. Prior to biodiesel production

2.1.1. Pyrolysis (thermal cracking) of oils. Pyrolysis is the shifting or transformation of one substance to another by employing heat. A catalyst is introduced into the process to minimize the conversion time. In other words, pyrolysis refers to the transformation of one organic material into another by thermal decomposition with the presence of a desired catalyst and the absence of air or oxygen.14 Vegetable oil, animal fats, natural fatty acids, or the methyl ester of fatty acids may be used as pyrolysis material. The transformation of vegetable oil and animal fat into a feedstock capable of producing biodiesel is a potential technology using a thermal cracking reaction. This innovation is particularly encouraging in areas where the hydro preparing industry is entrenched because this innovation is fundamentally the same as that of traditional petroleum refining.15 Dissimilar lower hydrocarbons are obtained by the pyrolysis of vegetable oil, which could be used as fuel. The properties of treated fuel made via pyrolytic syntheses and derived from vegetable oil are very similar to those of diesel fuel. Thus, numerous scientists have stated that this fuel is a suitable alternative for diesel.15–18 Based on its operating conditions, the pyrolysis process can be classified19 as conventional pyrolysis, fast pyrolysis, or flash pyrolysis. The mechanism of the thermal decomposition of triglycerides is prone to complexity because of the various possible reactions and chemical structures, which depend upon the reaction conditions or pathway. A simple systematic reaction mechanism of the thermal decomposition of triglycerides is presented in Fig. 4.18,20 Scientists have performed experiments on soybean, palm, and castor oils to determine the optimum distillation temperature, thereby obtaining fuel properties similar to those of petroleum-based fuels.19 Compared to the other conventional processes with respect to the yield, the equipment used for thermal cracking and pyrolysis is costly. By contrast, this process sometimes produces low-value materials and more gasoline than diesel fuel.21
image file: c5ra09862a-f4.tif
Fig. 4 The mechanism of the thermal decomposition of triglycerides.

Nevertheless, the pyrolysis process requires high temperatures ranging from approximately 300 °C to 500 °C and product characterization is difficult because of the differences in the reaction pathway and the products acquired from the reactants.14 Lima et al.18 employed zeolite as a catalyst during the pyrolysis of soybean oil and found that the reaction temperature was approximately 400 °C under N2 flow. The obtained products were olefins, paraffin, carboxylic acids, and aldehydes. Şensöz et al.22 inspected the effect of particle size on the pyrolysis of rapeseed. By changing the particle size of rapeseed over the range of 0.224–1.8 mm, they observed that the yields of the products are not dependent on the particle size. More than 30 compounds were detected from the pyrolysis of macauba fruit, and the amount of the main products diminished by increasing the pyrolysis temperature.23

2.2. Esterification and transesterification processes

An esterification reaction is one where an ester is produced from one or two other organic substances. The most common method to produce esters is by chemically reacting an organic acid (carboxylic acid) and an alcohol (methanol) with the help of an acid catalyst. The general esterification reaction is shown in Fig. 5, where R represents small alkyl groups and R1 denotes fatty acid chains.
image file: c5ra09862a-f5.tif
Fig. 5 General esterification reaction.

Esterification can be performed on vegetable oils or animal fats (triglycerides) with methanol or ethanol (short chain alcohols) to produce biodiesel, especially where considerable quantities of free fatty acids (FFAs) are present. These include byproducts of waste oils, non-edible oils, animal oils, and refined vegetable oils. Such oils possess considerable quantities of saturated fatty acids, specifically stearic acid (IUPAC name: octadecanoic acid), which contains 18 carbon atoms. In some cases, the homogenous acid-catalyzed reaction is not viable because it may produce corrosion and environmental problems. By contrast, heterogeneous reactions do not show corrosive behavior. In addition, heterogeneous reactions are easier to use for splitting products, diminishing wastewater quantities, and lowering process instrumentation, expenses, time, and environmental effects. Thus, the heterogeneous acid-catalyzed reaction is preferred for esterification reactions. Such catalysis plays a significant role in producing cleaner and more profitable biodiesel by esterification. Thus chemical processes employing heterogeneous catalysts are the most acceptable to researchers for creating biodiesel by esterification.24

Creating biodiesel from the transesterification reaction with the help of a catalyst is a highly favored method. The transesterification reaction for biodiesel production can be performed using various methods and is broadly described as the addition of an alcohol (generally methanol or ethanol) to lipids (vegetable oil, algal oil or animal fats) in the presence of a catalyst (acid or base).25 Outlines of the transesterification reactions for fatty acid methyl esters (FAMEs) and fatty acid ethyl esters (FAEEs) are shown in Fig. 6(a) and (b), respectively, where R1, R2, and R3 represent mixtures of long fatty acid chains.26


image file: c5ra09862a-f6.tif
Fig. 6 Chemistry of the transesterification reaction of triglycerides for biodiesel production: using (a) FAMEs, and (b) FAEEs.
2.2.1. Microwave-assisted transesterification technology. Microwave irradiation or microwave-assisted transesterification is a technique where an electromagnetic microwave is used to heat up the system. Electromagnetic wave frequencies ranging from 0.3 GHz to 300 GHz, which are relatively low in the electromagnetic spectrum, are used to produce energy.27 During microwave irradiation, the bonds are neither formed nor broken. However, the energy is rapidly transferred to the sample. A microwave is integrated into the process after the mixture (methanol, feedstock, and catalyst) chamber. Utilization of electromagnetic microwaves has drawn significant attention28–31 because these waves present some advantages over conventional heating in transesterification. A simple schematic of the microwave-assisted transesterification process29 is shown in Fig. 7. A faster and uniform heat distribution, higher yields of a cleaner product, less energy consumption, and the reduction of the catalyst to methanol ratio are among the attractive features of this technique.
image file: c5ra09862a-f7.tif
Fig. 7 Microwave-assisted transesterification process.
2.2.2. Ultrasonic technology. Ultrasonic technologies are typically used in a variety of biological and chemical reactions to enhance the yield within a shorter reaction time. Ultrasonic technologies are an effective method to enhance the mass transfer rate between immiscible liquid–liquid phases within a heterogeneous catalyst.32 Generally, the audible range of human beings lies between 16 and 18 kHz,33 whereas the ultrasonic sound range lies between 20 kHz and 100 MHz. The principle of ultrasound action in biodiesel production is primarily based on the emulsification of immiscible liquid reactants by micro-turbulence generated by the radial motion of cavitation bubbles. Molecules in the medium continuously vibrate and create cavities by compressing and stretching the molecular spacing of the medium, which is developed by a high-frequency sound wave. As a result, micro fine bubbles are formed by the sudden expansion and violent collapse, generating energy for chemical and mechanical effects.34 This process allows a short reaction time and high yield because of the cavitation of the liquid–liquid immiscible system.32 Maneechakr et al.35 stated that ultrasonic-assisted transesterification shortens the reaction time and minimizes the molar ratio of alcohol to oil, and reduces the energy consumption compared with the conventional mechanical stirring method. According to Lee et al.36 ultrasonic irradiation reduces reaction times by at least 30 min compared with the conventional methods and produces the highest biodiesel yields.

3. Catalysts used for biodiesel production

At present, biodiesel is produced using soybean oil in the U.S.; rapeseed, sunflower, or soybean oils in the EU; and palm oil in Southeast Asia. Food versus fuel concerns have prompted the investigation of non-consumable oil feedstocks. The top Asian countries in biodiesel generation, such as the Philippines and Malaysia, also exploit edible oil (coconut and palm oil, respectively). Other edible raw materials, such as sunflower, peanut, camelina and linseed oil, are also utilized by some developing countries. The breakdown of crude materials used for biodiesel generation worldwide is 84% rapeseed oil, 13% sunflower oil, 1% palm oil and 2% others. The principal crude material utilized in India is jatropha, a non-edible oil.18 Determination of which oil is used for biodiesel generation depends on its availability, characteristics, and price.17,18 Some edible and non-edible feedstocks for biodiesel production and their properties are listed in Tables 2 and 3.
Table 2 Feedstocks and their oil yield for biodiesel production
Crops Algae
Name Oil yield (%) Species Oil yield (%)
Babassu oil (Attalea speciosa) 60–70 Botryococcus bruanii 25–75
Borage oil (Borago officinalis) 20 Chlorella sp. 28–32
Camelina oil (C. sativa) 38–40 Crypthecodinium cohni 20
Castor oil (Ricinus communis) 45–50 Cylindrotheca sp. 16–37
Cuphea oil (Cuphea viscosissima) 25–43 Nitzschia sp. 45–47
Hemp oil (Cannabis sativa) 33 Phaeodactylum tricornutu 20–30
Jatropha oil (J. curcas) 45 Schizochytrium sp. 50–77
Jojoba oil (Simmondsia chinensis) 44 Tetraselmis suecica 15–23
Karanja oil (P. pinnata) 27–39 Isochrysis galbana 30–41
Linseed oil (Linum usitatissimum) 37–42 Pavlova lutheri 35.5
Neem oil (Azadirachta indica) 40–50 Nannochloropsis sp. 31–68


Table 3 Some edible and non-edible sources (feedstocks) for biodiesel production and their properties
  Feedstock Density, g cm−3 Flash point, °C Acid value, mg KOH per g Heating value, MJ kg−1 Viscosity, mm2 s−1 at 40 °C Cloud point, °C Peroxide value Pour point, °C Cetane no. References
Edible oils Soybean 0.910 254.0 0.20 39.60 32.60 2 44.5 −12.2 48.00 20,37–39
Rapeseed 0.910 246.0 2.92 39.70 37.00 30.2 35, 36 and 40
Sunflower 0.920 274.0 0.15 39.60 33.90 10.7 35, 36 and 40
Palm 0.920 267.0 0.10 39.90 36.00 12 13 61.15 35, 40 and 41
Peanut 0.900 271.0 3.00 39.80 39.60 82.7 35, 36 and 40
Corn 0.910 277.0 0.11 39.50 18.4 35 and 36
Cotton 0.910 234.0 39.50 64.8 35
Moringa 0.859 176.0 0.19 40.11 5.074 21 19 67.07 42 and 43
Calophyllum 0.877 162.5 0.30 39.51 5.538 12 13 57.30 42, 44 and 45
Coconut 0.860 118.5 0.11 38.30 3.144 1 −4 59.00 42 and 46
Aphanamixis polystachya 0.873 188.5 0.45 39.96 4.718 8 8 42, 45 and 47
Rice bran 0.868 174.5 0.59 39.96 5.366 0 −3 73.60 42, 45 and 48
Neem 0.868 120 0.65 39.81 3.700 9 2 48–53 45, 48 and 49
Sesame 0.884 208.5 0.30 39.99 4.399 1 1 50.48 50 and 51
Non edible oils Jatropha curcas 0.920 225.0 28.0 38.50 10 57.10 35 and 42
Pongamina pinnata 0.910 205.0 5.06 34.00 35
Palanga 0.900 221.0 44.0 39.25 35
Tallow 0.920 40.05 35
Poultry 0.900 39.40 35
Used cooking oil 0.900 2.50 35


Several catalysts are associated with biodiesel production technology. Based on previous reviews5,52–55 catalysts can be classified as homogeneous catalysts, heterogeneous catalysts, and biocatalysts; these catalyst types, including their sub-types are listed in Fig. 8.


image file: c5ra09862a-f8.tif
Fig. 8 Catalysts used in biodiesel production technology.

3.1. Homogeneous catalysts

Homogeneous catalysis involves a sequence of reactions that includes a catalyst from the same phase as the reactants. The term phase in this article refers to solid, liquid and gas. Most often, a homogeneous catalyst is dissolved or co-dissolved in the solvent with all the reactants. Sodium hydroxide (NaOH) or potassium hydroxide (KOH) are currently the most popular homogeneous catalysts for biodiesel production.56 However, some researchers suggest that these homogeneous base catalysts are only suitable for feedstocks with a low FFA content. If the FFA content is larger than 6 wt%, the base catalyst process is unsuitable for biodiesel synthesis.57 Thus, some scientists recommend that the FFA content should be less than 2 wt%.21,58–60 This homogeneous base catalyst is gradually gaining popularity in industrial biodiesel production for the following reasons:

(a) A low reaction temperature is needed to synthesize biodiesel at atmospheric pressure.

(b) A high biodiesel yield could be possible at optimal conditions.

(c) Widely available and economical.

(d) Handling the reactants and catalyst is easier than one solid or a combination.

Although homogeneous catalysts present several advantages, they also feature some drawbacks. The limitations of using a homogeneous catalyst for biodiesel synthesis are listed below:

(a) Water generation throughout the acid esterification retards the process. Proper care should be taken to remove water via evaporation or chemical drying, which adds expenses to biodiesel production.

(b) Although instances of reusing catalysts have been reported, this method is almost never witnessed if ever completed over a production scale because of the associated expenditures.

(c) Corrosive nature of the catalysts concerned. The preferred acid (H2SO4) and base (SMO) catalysts presently employed in biodiesel synthesis are corrosive and must be handled conservatively.

3.2. Heterogeneous catalysts

Catalysts with a phase or state different from that of the reactants are considered to be heterogeneous catalysts. A heterogeneous catalyst is often a practical material that regularly creates active sites with its reactants under the reaction atmosphere. The application of a heterogeneous catalyst will result in simpler and less costly separation processes, as well as additional capital and energy costs. The disadvantages of heterogeneous catalysts include elevated temperatures and higher oil to alcohol ratios than those required in the homogeneous catalytic procedure. Some of these catalysts have demonstrated good performance even under the reaction conditions employed for homogeneous catalysts.61,62 The separation, purification, and reusability of the catalyst are among the more attractive features of the heterogeneous catalytic process. Carbonates and hydrocarbonates of alkaline metals, alkaline metal oxides, alkaline metal hydroxide anionic resins, and basic zeolites may be used as heterogeneous catalysts.

Dossin et al.63 demonstrated the heterogeneous catalyst transesterification process employing MgO as a catalyst. About 100[thin space (1/6-em)]000 tons of biodiesel were generated every year by this process from a triolin feedstock. The remarkable heterogeneous procedure known as the Esterfip-H technology, which was developed by the Institute Français du Petrole (IFP) in 2006, delivers 160[thin space (1/6-em)]000 tons of biodiesel every year.64 The role and utilization of base and acid heterogeneous catalysts for biodiesel synthesis has been broadly described in the literature.

3.2.1. Acid heterogeneous catalysts for biodiesel production. Acid catalysts have a corrosive but less toxic effect and create few environmental problems.65 This type of catalyst is more popular for both esterification and transesterification reactions when producing biodiesel from low quality feedstock.66 Transesterification is catalyzed by Bronsted acids. The most ordinarily employed acid catalyst is H2SO4. Broadly, heterogeneous acid catalysts can be divided into two types: the low temperature type and the high temperature type;67 catalysts of these types show better activities at low and high temperatures, respectively. While these catalysts provide exceptional returns in alkyl esters, their rates of reaction are moderate, demanding reaction times longer than 3 h and temperatures higher than 100 °C to achieve complete transformation.68 The basic mechanism of an acid catalyst is shown in Fig. 9. Investigations using a heterogeneous acid catalyst to produce biodiesel are discussed in the next paragraph.
image file: c5ra09862a-f9.tif
Fig. 9 Basic reaction mechanism of an acid catalytic process.

3.2.1.1. ZrO2 as an acid heterogeneous catalyst. Zirconium dioxide (ZrO2), sometimes referred to as zirconia, is used as an acid heterogeneous catalyst by many researchers. This substance has a strong surface acidity. Thus, many investigators often choose ZrO2, sulfated ZrO2,69 and other modified metal oxides with sulfated ZrO2 as acid heterogeneous catalysts.70

According to Kiss et al.,71 sulfated zirconia (SO42−/ZrO2) shows the best performance as a heterogeneous catalyst for esterification among zeolites, ion exchange resins, and mixed metal oxides. Park et al.72 successfully performed the transesterification of vegetable oil to convert FFA to FAME using sulfated zirconia (SO42−/ZrO2) and tungstated zirconia (WO3/ZrO2). However, tungstated zirconia (WO3/ZrO2) requires a longer reaction time. Approximately 140 h of reaction time and a 75 °C reaction temperature are needed to achieve only 65% conversion.

Implementation of tungstated zirconia-alumina (WZA) and sulfated zirconia-alumina (SZA) was assessed by Furuta et al.73 during the transesterification of soybean oil with methanol. The reaction was performed under atmospheric pressure and at 200–300 °C with the help of a fixed bed reactor. The investigators showed that WZA has higher activity than SZA but did not elaborate on the reason behind this phenomenon.

Jitputti et al.74 utilized both sulfated zirconia (SO42−/ZrO2) and unsulfated zirconia (ZrO2) for the transesterification of palm kernel oil and crude coconut oil with the help of methanol. They reported 90.3% and 86.3% methyl ester yields from palm kernel oil and crude coconut oil, respectively, while using sulfated zirconia. Only a 64.5% palm kernel yield and a 49.3% crude coconut oil yield was possible in the case of unsulfated zirconia. These results clearly indicate that a slight change in the metal oxide surface activity is a major parameter influencing the yield of methyl ester.


3.2.1.2. Cation-exchange resins as heterogeneous catalysts. Many researchers have extensively used cation-exchange resins for biodiesel production at the laboratory scale; this resin is unsuitable for industrial applications. Liang et al.75 described an optimal operational condition to convert soybean oil to biodiesel with the help of chloroaluminate ([Et3NH]Cl–AlCl3). 5 g of soybean oil, 2.33 g of methanol, a reaction time of 9 h, and a reaction temperature of 70 °C were recorded as the optimal conditions for a 98.5% biodiesel yield. The main advantages of the use of such a catalyst are low cost, high yield of biodiesel production, reusability of catalyst, no need for saponification, and simplicity of operation.

Some investigators have attempted to use a poly(DVB) resin sulfated with H2SO4, Amberlyst-35 (Rohm & Haas),76 Amberlyst-15 (Rohm & Haas),76 Amberlyst 15 DRY77 and Nafion SAC-13 (ref. 78) as sulfonic acid ionic exchange resins. Limited functionality with a significantly high ratio of oil to alcohol is one of the major shortcomings of this type of catalyst. By using an organically functionalized acid catalyst, the key shortcomings of the aforementioned catalysts, including leaching and low surface areas, could be solved.79


3.2.1.3. Solid heteropoly acids as heterogeneous catalysts. Heteropoly acid (HPA) catalysts are another low temperature heterogeneous catalyst utilized by Chai et al.80 for high-quality biodiesel production. Narasimharao et al.81 studied insoluble HPA salts for solid acid esterification and transesterification processes; the general formula of these salts is y CsxH3−xPW12O40, where x varies from 0.9 to 3. A Cs2.5H0.5PW12O40 catalyst could be used in simultaneous esterification and transesterification reactions without losing activity and selectivity. Hamad et al.82 introduced H3PW12O40/SiO2 and Cs2HPW12O40 as heterogeneous HPAs for the transesterification of rapeseed oil with the help of ethanol. The strengths of the acid sites are measured in terms of the heating value.

Zhang et al.65 studied the maximum yield of FAMEs using Cs2.5H0.5PW12O40 as a heterogeneous HPA for microwave-assisted transesterification to produce biodiesel from yellow horn (Xanthoceras sorbifolia Bunge) oil. The purpose of this study was to optimize the reaction temperature, reaction time, methanol to oil molar ratio, amount of catalyst, and catalyst reusability cycle. An approximately 96.22% FAME yield was achieved under optimal conditions. A reaction temperature of 70 °C, reaction time of 10 min, methanol to oil molar ratio of 12[thin space (1/6-em)]:[thin space (1/6-em)]1, and catalyst loading of 1% wt were recorded as optimal conditions and the catalyst could be reused a minimum of nine times.


3.2.1.4. Zeolites as acid heterogeneous catalysts. Zeolites are crystalline solid structures made of silicon, aluminum, and oxygen that frame a structure with holes and channels inside where cations, water, and/or little atoms may dwell. Zeolites have unique properties as catalysts, including acidic strength and shape selectivity. The catalytic properties vary because of the porous structure. The internal crystal structure and surface properties of a zeolite can suit a wide assortment of cations, such as Na+, K+, Ca2+, and Mg2+ that ascribe to its simple nature. Approximately 191 exceptional zeolite frameworks have been identified.83 About 40 naturally occurring zeolite frameworks are known. An improved zeolite (using La) called zeolite beta (La/zeolite beta) was studied by Shu et al.84 as a heterogeneous acid catalyst for the methanolysis of soybean oil. This catalyst was synthesized using an ion exchange process through a suspension of zeolite beta in a lanthanum nitrate La(NO3)3 aqueous solution under high mixing at ambient temperature for 3 h and drying at 100 °C for 24 h. Then, the catalyst is calcined at 250 °C for 4 h. The author recorded only a 48.9 wt% triglyceride conversion using this catalyst. Karmee et al.85 studied Hβ-zeolite, montmorillonite K-10, and ZnO as heterogeneous catalysts for the transesterification of non-edible Pongammia pinnata under reaction conditions of a 1[thin space (1/6-em)]:[thin space (1/6-em)]10 oil to methanol ratio, 0.115 wt% catalytic loading, 120 °C temperature, and 24 h reaction time. Yields of 59%, 47%, and 83%, respectively, were recorded. Suppes et al.86 examined a biodiesel synthesis process where zeolites were used as the prospective acid heterogeneous catalyst. They used several types of zeolites and metals to examine the potential of zeolites during the transesterification of soybean oil possessing a 2.6 wt% FFA content. Xie et al. studied4 NaX zeolites (Si/Al = 1.23) as heterogeneous catalysts after improving the strength from <9.3 to 15.0–18.4. A yield of 85.6% was recorded under a 10% catalytic loading, 125 °C reaction temperature, and 2 h of reaction time. Ramos et al.87 used zeolites (mordenite, beta, and X) to convert crude sunflower oil to FAME and recorded a 93.5–95.1% FAME yield at 60 °C. However, the time required to prepare the catalyst was relatively long because the process required heating at 500 °C for 10 h, drying at 120 °C for 12 h, and calcining at 550 °C for 15 h.

Zeolites can sometimes be used as a base heterogeneous catalyst. Marchetti et al.88 used basic NaY zeolites and VOX over ultrastable acidic Y zeolites as heterogeneous catalysts for the conversion of FFA to FEME. At 300 °C, the conversion times of FFA oil to biodiesel using these zeolites were only 10 min and 50 min, respectively. During the reaction, the presence of H2O initially accelerated the reaction. However, when the reaction proceeded further, the H2O decelerated the reaction. Thus, the overall biodiesel conversion rate was hampered.

Though acid heterogeneous catalysts are promising to biodiesel production technology, they also have some drawbacks for the production process. The advantages and disadvantages of acid heterogeneous catalysts are discussed in Table 4.

Table 4 Benefits and limitations of acid heterogeneous catalysts for biodiesel production
Catalyst used Reaction parameters Benefits and limitations Ref.
Tungstated zirconia (WO3/ZrO2) Reaction time 140 h, reaction temperature 75 °C and yield 65% Long reaction times 66
Sulfated zirconia (SO42−/ZrO2) Palm kernel oil yield 90.3%, coconut oil yield 86.3% under optimal conditions Shorter reaction time than tungstated zirconia 66 and 79
Unsulfated zirconia (ZrO2) Palm kernel oil yield 64.5%, coconut oil yield 49.3% under optimal conditions Very poor biodiesel yield compared to others 79
Tungstated zircona-alumina (WZA) Atmospheric pressure, reaction temperature 200–300 °C, with fixed bed reactor Relatively higher activity than sulfated zirconia-alumina 67
Sulfated zirconia-alumina (SZA) Atmospheric pressure, reaction temperature 200–300 °C, with fixed bed reactor Relatively lower activity than tungstated zircona-alumina 67
Chloroaluminate ([Et3NH]Cl–AlCl3) Soybean oil yield 98.5%, reaction time 9 h, reaction temperature 70 °C Reaction time is quite high though yield is good 69
Heterogeneous HPA (Cs2.5H0.5PW12O40) Reaction temperature 70 °C, reaction time 10 min, 12[thin space (1/6-em)]:[thin space (1/6-em)]1 methanol to oil ratio, catalyst loading 1% wt from yellow horn Minimum time required with a high oil to alcohol ratio and the catalyst can be reused nine times 65
Sulfonic acid ionic exchange resins High oil to alcohol ratio is needed 76​–78


3.2.2. Base heterogeneous catalysts for biodiesel production. Currently, several solid base catalysts have been developed for biodiesel production, such as basic zeolites, alkaline earth metal oxides, and hydrotalcites. The most general catalysts, which are used as base heterogeneous catalysts, are the K/γ-Al2O3 catalyst,89 the HTiO2 hydrotalcite catalyst,90 Ca and Zn mixed oxides,91 the calcium oxide (CaO) catalyst,92,93 the MgO catalyst, Al2O3 supported CaO & MgO catalysts,3 the Li4SiO4 catalyst,94 the Na2SiO3 catalyst,95,96 alkaline earth metal oxides,97 KF/Ca–Al,98 basic zeolites, and alkali metal loaded alumina.99 In addition, alkaline earth steel oxides, particularly CaO, have attracted much attention owing to their relatively higher strength, significantly low solubility in methanol, and synthesis from affordable resources such as limestone and calcium hydroxide.100,101 Researchers recommend base heterogeneous-catalyzed transesterification for the following reasons: simplification of biodiesel production, simplification of the purification process, reduction in wastewater amount, lowering of the process and equipment costs, and reductions in environmental impact.67,102 In addition to the ease of catalyst restoration, the action of the heterogeneous alkali catalyst may resemble that of its homogeneous counterpart under similar running conditions.103 Transformation of the crude oil to biodiesel occurs in several steps. The systematic reaction mechanism of a simple base catalytic reaction is shown in Fig. 10.
image file: c5ra09862a-f10.tif
Fig. 10 The reaction mechanism of base catalysis process.104

3.2.2.1. CaO as a base heterogeneous catalyst. CaO is the most commonly exploited alkaline earth metal oxide for transesterification. FAME yields of nearly 98% were initially reported.105 The ability to reuse the catalyst is a major topic of concern. The modification of CaO to a compound with an organic metallic nature, for example, Ca(OCH3)2 and Ca(C3H7O3)2, allows investigation of the reutilization function. Collected works on biodiesel state that an approximately 93% biodiesel yield may be obtained from a 20-cycle reaction. 95% Ca(C3H7O3)2/CaCO3 has also been determined to be a capable heterogeneous catalyst with a reusability of 5 cycles and a good FAME yield.106

The mechanics of transesterification introduced by Lam et al.56 used CaO as a heterogeneous base catalyst. CaO was reacted with FFAs, and a certain amount of the catalyst was transformed into Ca soap by rejoining with the FFAs, causing limited catalyst recovery. As a basic standard of biodiesel, the concentration of the mineral matter should be less than 200 ppm. Kouzu et al.107 determined a Ca concentration of 3065 ppm in the reaction products, thus exceeding this standard.

Some investigators noted that soluble matter can be removed by CaO throughout transesterification. CaO slightly dissolves in methanol, thus transforming into soluble calcium diglyceroxide, where CaO reacts with glycerin during the transesterification of soybean oil with methanol.107​–109 Activated CaO is used to study the function of H2O and CO2 in the loss of catalytic performance in the presence of O2 during the transesterification of sunflower oils.110 In the above studies, CaO was quickly hydrated and carbonated in the air. Activated CaO was affected because of the surface activity, as well as the absorption of CO2 and H2O on the surface area. If CaO is heated to 700 °C to eliminate the carbonate groups from the surface, then the catalytic action of CaO might be restored. However, filtering or removing the catalyst from the product was noticeable in the transesterification reaction. Considering its low solubility in methanol, CaO results in a high basic strength and fewer ecological effects. Furthermore, CaO can be produced from economical resources such as limestone and Ca(OH)2.

KOH can be used for alternative of CaO. But the advantages of CaO over KOH include lower price, lower solubility, higher basicity, and easier handling. In actual practice, the reaction rate is unsatisfactory during the transesterification for relatively low activity.109,111 However, reusable activity can be enhanced by washing109 and improving the thermal activation treatment.105 Being nano sized, CaO provides effective catalyst activity because of its high surface area. A surface structure of a metal oxide is presented in Fig. 11.


image file: c5ra09862a-f11.tif
Fig. 11 Surface structure of metal oxides comparing its acidic and basic sites.

Biodiesel production from sunflower oil using a heterogeneous base catalyst (CaO) was conducted by Vujicic et al.112 Experiments were accomplished using a commercial bench stirred tank reactor. The experiment was conducted using unchanged factors such as a reactor volume of 2 dm3, a stirring speed of 200 rpm, a 6[thin space (1/6-em)]:[thin space (1/6-em)]1 methanol to oil ratio, and 1 wt% of CaO catalyst loading. Temperature (60–120 °C), pressure (1–15 bar), and reaction time (1.5–5.5 h) effected the ester yields. The optimal transformation to methyl ester (almost 91%) was observed at 100 °C, although a positive effect of pressure (up to 10 bar) was observed at 80 °C. The founding of basic sites depends on the catalyst activation in air, which occurs at 900 °C. Catalyst corpuscle coalescence occurred during the reaction, which provided a gum-like construction and meaningful catalyst deactivation.

CaO was produced using a simple and flexible technique113 to raise the activity and to improve the properties of the calcined calcites through the hydration–dehydration method. This process prepares CaO that is extremely fit for biodiesel production. Newly prepared CaO has a relatively higher surface area and more basic sites than CaO obtained from the decomposition of calcite. With calcined calcite, the methyl ester content was enhanced to 93.9 wt% from 75.5 wt%. This study, developed through new hydration, delivers different important information on the influence of water on the properties and activities of CaO, and ensures a thermal disintegration technique for calcined calcite.

Bai et al.114 established a novel morphology for producing CaO simply and at low cost. The morphology produced possesses high catalytic activity in the transesterification reaction for biodiesel. A CaO microsphere with tiny holes was obtained by calcining a spherical CaCO3 precursor prepared simply through the reaction of CaCl2 and Na2CO3. During the transesterification of soybean oil, the CaO microsphere was employed, which possesses outstanding catalytic capability. An approximately 98.72% FAME yield was obtained.


3.2.2.2. BaO as a heterogeneous catalyst. Mootabadi et al.97 conducted a palm oil transesterification experiment with the aid of alkaline earth metal oxide catalysts such as CaO, SrO, and BaO. The experiment was performed to observe the result of varying the reaction time (10–60 min), alcohol to palm oil molar ratio (3[thin space (1/6-em)]:[thin space (1/6-em)]1–15[thin space (1/6-em)]:[thin space (1/6-em)]1), and catalyst loading (0.5–3%) by creating 20 kHz ultrasonic cavitation and fluctuating ultrasonic amplitudes (25–100%). The natures of the catalysts were mostly dependent on their basic strengths. The activity ranking of the catalysts was CaO < SrO < BaO. In ideal circumstances, a 95% yield was reached with formal stirring within 60 minutes. Moreover, yields accomplished within 60 min for CaO, SrO, and BaO were 5.5–77.3%, 48.2–95.2% and 67.3–95.2%, respectively. Ultrasonic irradiation at a 50% amplitude was estimated as optimal, and the physical variations of the catalysts can be effectively explained after the ultrasonic-assisted reaction. The major reason for this activity drop for the recycled catalyst, dissolution, was investigated, especially for the BaO catalyst.
3.2.2.3. MgO as a base heterogeneous catalyst. Among the alkali earth materials, MgO is one of the widely used materials for transesterification.70 López et al.115 studied MgO used as a base heterogeneous catalyst where an 18% yield of feedstock was achieved at a calcination temperature of 600 °C. The low surface area of the catalyst means that the reaction takes almost 8 h. MgO is used to analyze the catalytic activity to produce biodiesel. An experiment by Di Serio et al.116,117 that used a 12[thin space (1/6-em)]:[thin space (1/6-em)]1 methanol to oil molar ratio and 5.0 wt% of the catalyst (MgO) gave a 92% biodiesel yield within 1 h. Another experiment stated118 that in a batch reactor, MgO worked efficiently and 500 tons of biodiesel production was achieved by transesterification at ambient temperature. In batch reactor biodiesel production, cost is minimal because of the temperature. Some researchers116,119 stated that in supercritical conditions (300 °C) and at a high methanol to oil molar ratio (39.6[thin space (1/6-em)]:[thin space (1/6-em)]1), a MgO catalyst gives a 91% FAME yield.
3.2.2.4. SrO as a base heterogeneous catalyst. In solid base catalysis, although Ca and Mg are the more extensively used alkaline earth materials, strontium oxide (SrO) has also been used in biodiesel production. Using CO2 temperature programmed desorption, pure SrO was confirmed to have the maximum basic strength,120 which is comparable to that of BaO (26.5 < H). In addition, SrO has an inferior surface area with respect to MgO and CaO.121 Only a few experiments have been conducted using SrO as a heterogeneous catalyst for the transesterification of biodiesel production.

Zabeti et al.122 discussed the appropriateness of using SrO as a catalyst for the transesterification reaction. Liu et al.123 found that in the reaction medium, SrO acts as an extremely active and soluble metal oxide. However, vegetable oil, methyl esters and SrO remain insoluble in methanol, thus SrO becomes a suitable heterogeneous catalyst for transesterification. After 30 min at 65 °C with an alcohol to oil molar ratio of 12 wt% and a 3 wt% catalyst loading, 90% yields of methyl esters were accomplished during the transesterification of soybean oil. However, the specific surface area of the catalyst was as low as 1.05 m2 g−1.

Salamatinia et al.124 investigated the heterogeneous transesterification of palm oil for biodiesel production using an ultrasonic process. Briefly, response surface methodology was applied to optimize the biodiesel production process with the help of two (BaO and SrO) alkaline earth metal oxide catalysts. To optimize production, four variable factors were considered. Reaction time (10–60 min), alcohol to oil molar ratio (3[thin space (1/6-em)]:[thin space (1/6-em)]1–15[thin space (1/6-em)]:[thin space (1/6-em)]1), catalyst loading (0.5–3.0 wt%), and ultrasonic amplitude (25–100%) were included as the optimized factors. The mathematical frameworks and the steps of the process were established. The frameworks were able to correctly predict the biodiesel yield with less than 5% error for the SrO and BaO catalysts. The high activity of a catalyst mostly depends on its basic strength. Ultrasound was found to significantly enhance the process by decreasing the reaction time by almost 50 min, and a 2.8 wt% catalyst loading was used for creating biodiesel yields of more than 95%. The best results were given by a 9[thin space (1/6-em)]:[thin space (1/6-em)]1 alcohol to oil ratio and a ∼70% and ∼80% ultrasonic amplitude for the BaO and SrO catalysts, respectively.


3.2.2.5. Boron group oxides (Al2O3) supported upon CaO and MgO. Boron group components, especially aluminum and Al2O3, are widely utilized stacked with different other metal oxides, halides, nitrates, and alloys.3,89,123,125 Boron and aluminum oxides are the most commonly used catalysts for mixed metal oxide production among the boron group elements. To achieve a decent yield of biodiesel production, different forms of oxides such as Al2O3, γ-Al2O3, and Al2O3 supported upon CaO and MgO are used as heterogeneous base catalysts. Xu et al.126 used a mesoporous polyoxometalate–tantalum pentoxide composite catalyst, H3PW12O40/Ta2O5, which was prepared using a one-step sol–gel hydrothermal method in the presence of a triblock copolymer surfactant.
3.2.2.6. Biodiesel production with carbon group catalysts. Formulating a carbon-based heterogeneous catalyst is simple, and is profitable for biodiesel production when used as a catalyst. Shu et al.127 used the sulfonation of carbonized vegetable oil asphalt as a solid catalyst to produce biodiesel via the transesterification of vegetable oil. This catalyst can be used in both the esterification and the transesterification processes, while waste vegetable oil with large amounts of FFAs is present in the feedstock. The maximum conversion of triglyceride and FFA reached 80.5% to 94.8% after 4.5 h at 220 °C when using a methanol to oil ratio of 16.8[thin space (1/6-em)]:[thin space (1/6-em)]1 and 0.2 wt% of catalyst.

Dehkhoda et al.128 studied the transesterification reaction of palm oil for biodiesel production, where KOH/AC was used as a heterogeneous catalyst. In their study, operating conditions were 70 °C, the molar ratio of alcohol to oil was 15[thin space (1/6-em)]:[thin space (1/6-em)]1, catalyst loading was 5 wt%, and reaction time was 15 h. An approximately 94% biodiesel yield was achieved, and the catalyst could be reused up to three times. Through the incineration of commercial grade sugar, a carbon catalyst was prepared and studied by Toda et al.129 Only a 50% oil to ester conversion could be achieved from these carbon structures at the end of its first cycle, although catalytic activity remained unchanged.

Lou et al.130 reported the arrangements of carbohydrate-derived catalysts from D-glucose, sucrose, cellulose, and different sorts of starches. The synergistic and textural properties of the arranged catalysts were examined at points of interest, and the starch-based catalyst was found to have the best reactant execution. The carbohydrate-derived catalysts displayed considerably higher synergistic effects for both esterification and transesterification in contrast with the two ordinary strong solid acid catalysts, sulfated zirconia (S-ZrO2) and niobic acid (Nb2O5·nH2O). The carbohydrate-derived catalysts also gave a significantly upgraded yield of methyl esters when converting waste cooking oils containing 27.8 wt% high FFAs to biodiesel. In addition, under the maximum response condition, the starch-based catalyst held a significantly high extent (approximately 93%) of its unique synergistic action after 50 cycles of progressive reuse and showed exceptional operational dependability. The carbohydrate-derived catalysts, particularly the starch-based catalyst, were exceptionally compelling, recyclable, eco-friendly, and apposite for the creation of biodiesel from waste oils containing high FFAs.

Faria et al.131 utilized tetramethyl guanidine on a silica gel surface132 as a base catalyst. 13C and 29Si atomic attractive reverberation points of interest are in concurrence with the proposed structure. The investigator stated that the biodiesel yield was approximately 86.73% and the reaction time was approximately 3 h. In addition, the catalyst recovery was complete; almost 62% after the 9th cycle of catalyst reuse.


3.2.2.7. Waste materials based heterogeneous catalysts. Mainly calcium-enriched waste materials are used as catalyst synthesis sources, with mollusc shell, eggshell, and bones being the most common. Researchers examined a number of systems to remove misused atoms and change useable catalysis with high cost viability this useable atoms recovery deals with high cost. CaO obtained from these waste materials could have potential for the biodiesel era. Boey et al.133 reported that waste shells, which is an issue of CaO, can be utilized as a catalyst for the transesterification of palm olein into methyl esters. Analysis showed that the fundamental segment of the shell was CaCO3, which changed into CaO when heated over 700 °C for 2 h. The economically produced catalyst affected the transesterification for biodiesel production, similar to laboratory CaO. Ideal conditions were found at a 0.5[thin space (1/6-em)]:[thin space (1/6-em)]1 methanol to oil mass ratio, a 5 wt% catalyst loading, a stirring speed of 500 rpm, and a reaction temperature of 65 °C. Reusability measurements established that the arranged catalyst could be recycled up to 11times. Factual examination was performed utilizing a central composite design to assess the effect of the parameters on biodiesel quality. Chakraborty et al.134 obtained CaO from waste eggshells, which was considered as a viable catalyst for transesterification at 65 °C, with an oil/alcohol ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]9, and a catalyst loading of 10 wt%. An approximately 97% to 98% FAME yield was achieved, and the catalyst could be reused up to 17 times. Viriya-empikul et al.135 studied the transesterification reaction of palm olein oil for biodiesel production, where waste eggshell, golden apples, and Meretrix venus were used as a waste base for a solid heterogeneous catalyst. In their study, operating conditions were 60 °C, the molecular ratio of alcohol to oil was 18[thin space (1/6-em)]:[thin space (1/6-em)]1, the catalyst loading was 10 wt%, and the reaction time was 1 h. The study produced 97%, 83%, and 78% biodiesel yields, respectively. Effective waste administration and a waste to vitality transformation can facilitate biodiesel generation utilizing eggshells.

Du et al.136 studied the biodiesel production from soybean oil where waste eggshells were used as a heterogeneous catalyst. Operating conditions were as follows: temperature was 70 °C, molar ratio of alcohol to oil was 6.9[thin space (1/6-em)]:[thin space (1/6-em)]1, catalyst loading was 5 wt%, and reaction time was 5 h. An approximately 97.73% biodiesel yield was achieved, and the catalyst could be reused up to 6 times. Alternate specialists, Nakatani et al.137 inspected the transesterification of soybean oil catalyzed by calcined shellfish shells.


3.2.2.8. Alkali metal inserted complexes as heterogeneous catalysts. Xie et al.4 studied the transesterification reaction of soybean oil for biodiesel production using NaX zeolites loaded with 10% KOH as a heterogeneous catalyst. In their study, the operating conditions were 65 °C, a molar ratio of alcohol to oil of 10[thin space (1/6-em)]:[thin space (1/6-em)]1, a catalyst loading of 3 wt%, and a reaction time of 8 h. An approximately 85.6% biodiesel yield was achieved. Fabbri et al.138 studied the transesterification reaction of soybean oil for biodiesel production using Na2PEG (300), a dimethyl carbonate, as a solid heterogeneous catalyst. In their study, the operating conditions were 70 °C, a molar ratio of alcohol to oil of 30[thin space (1/6-em)]:[thin space (1/6-em)]1, a catalyst loading of 6 wt%, and a reaction time of 5 h. An approximately 99% biodiesel yield was achieved.

Kondamudi et al.139 incorporated an extraordinary bifunctional catalyst Quntinite-3T (Q-3T) for biodiesel generation from waste vegetable oils, restaurant oil, and poultry fat. These oils picked up mechanical vitality compared with costly sustenance-based vegetable oils. This bifunctional heterogeneous catalyst simultaneously changes FFAs and triglycerides (TGs) into biodiesel. Q-3T is obtained from a sodium source (Na-Q-3T) and ammonium (NH4-Q-3T) sources using the sol–gel process and is characterized using X-ray powder diffraction (XRD), scanning electron microscopy (SEM), and high resolution transmission electron microscopy (HRTEM). The catalyst was tried for soy, canola, espresso, and waste vegetable oils with variable measures of FFAs (0–30 wt%). The catalyst effectively changed both FFAs and TGs in a solitary step bunch reactor.


3.2.2.9. Transition metal oxides and derivatives as heterogeneous catalysts. Many transition metals of the periodic table exist. Some transition metals and their oxides are chosen as base heterogeneous catalysts by researchers for biodiesel production. ZnO, TiO2,140 TiO2/SO42− and ZrO2/SO42−,119 ZnO, and ZrO2 (ref. 74, 141 and 142) are the most commonly used heterogeneous catalysts. The action of ZrO functionalized with TiO was likewise investigated by López et al.143 Nakagaki et al.144 engineered sodium molybdate (Na2MoO4) and utilized it as a heterogeneous catalyst for the methanolysis of soybean oil. Transesterification responses occurred under moderately gentle conditions: obliging low temperatures, short reaction times, and normal pressure. In their study, the produced methyl ester was characterized using reverse-phase and size-exclusion chromatography and hydrogen-nuclear magnetic resonance spectroscopy. Transesterification responses occurred under moderately gentle conditions: low temperatures, short reaction times and normal pressure. The methyl esters combined were described using opposite stage and size-exclusion chromatography and hydrogen-nuclear magnetic resonance spectroscopy. The transesterification response of the triglycerides with methanol was exceptionally effective with methyl ester yields higher than 95%. The molybdenum(VI) complex appeared to have high Lewis acidity and most certainly act on alcohol O–H bonds, prompting transition states with strongly nucleophilic characteristics. The catalyst was effortlessly recouped and after being washed, demonstrated recyclability for an alternate synergistic response with comparative movement.

Yoo et al.140 recommended ZnO as the best catalyst during the transesterification of rapeseed oil using transition metal oxides (ZnO, TiO2, and ZrO2) as heterogeneous catalysts because of their high activity. Operating conditions include a molar ratio of methanol to oil 40[thin space (1/6-em)]:[thin space (1/6-em)]1, a catalyst loading of 1 wt%, and a reaction time of 10 min. Brito et al.145 studied transesterification and esterification reactions to obtain methyl ester using two series of complexes with the general formula M(n-butoxide)4−x(maltolate)x, where M = Ti or Zr and x = 0–4, as catalysts. Mixes containing different ratios of maltolate and n-butoxide ligands were produced from the response of maltol (3-hydroxy-2-methyl-4-pyrone) and n-butoxide metal forerunners. All structures containing maltolate as ligand, were exceptionally productive as catalysts in esterification, mostly those containing zirconium. By contrast, these catalysts showed exceptionally poor activity for transesterification.

da Silva et al.146 studied the transesterification reaction of soybean and babassu oils for biodiesel production where Cu(II) and Co(II) were used as catalysts. The catalysts were characterized using infrared, atomic absorption, and thermogravimetry, and the biodiesels were characterized using infrared, NMR, CG, thermogravimetry, and physical chemistry analysis. In their study, the maximum adsorption values reported for the Cu(II) and Co(II) cations were 1.584 mg g−1 and 1.260 mg g−1, respectively, after 180 minutes. However, the conversion of oil to biodiesel was better when Co(II) was adsorbed on chitosan. Krohn et al.147 considered the generation of algal biodiesel from Dunaliella tertiolecta, Nannochloropsis oculata, wild freshwater microalga, and macroalga lipids utilizing a highly effective continuous catalytic Mcgyan® process. The heterogeneous catalytic procedure utilizes supercritical methanol and permeable titania microspheres in a settled bunk reactor to catalyze the synchronous transesterification and esterification of triacylglycerides and FFAs, separately, into biodiesel. Yields of up to 85% of alkyl esters were obtained from triglycerides and FFAs, as measured by a 300 MHz HNMR spectroscope.


3.2.2.10. Some mixed metal oxides and derivatives as heterogeneous catalysts. Xu et al.126 studied the transesterification reaction using KF/Zn(Al)O as a catalyst. The prepared KF/Zn(Al)O catalyst had an activity higher than that of Zn–Al hydrotalcites, such as Zn(Al)O, KF, KF/γ-Al2O3, and KF/ZnO compounds. This catalyst was especially viable when operating conditions were 65 °C, a molar ratio of methanol to oil of 6[thin space (1/6-em)]:[thin space (1/6-em)]1, KF/Zn(Al)O catalyst loading of 3 wt%, and a reaction time of 3 h. An approximately 95% biodiesel yield was achieved. The high action was credited to the establishment of new stages KF and KOH, and the effect of Zn(Al)O.

Wang et al.148 produced MgO–MgAl2O4 by utilizing γ-Al2O3, and this MgAl2O4 composite was used as heterogeneous catalyst. This gave a more noteworthy biodiesel yield compared with a MgO/MgAl2O4/γ-Al2O3 material with the comparable loading of magnesium arranged by an ordinary impregnation process. The improved catalytic action of the prior material could be attributed to its higher basicity, specific surface area, pore volume, and size. Wen et al.149 studied the transesterification of soybean oil with methanol for biodiesel production. Kalsilite (KAlSiO4) was used as a heterogeneous catalyst. A method called co-precipitation was introduced to produce kalsilite and add lithium for property improvement as well as catalytic performance improvement for the transesterification of soybean oil. Comparatively low catalytic activity was shown by kalsilite. However, a small amount of lithium nitrate added using the impregnation method significantly boosted the catalytic activity. With operating conditions of a reaction temperature of 120 °C, a 3.84 cSt kinematic viscosity, and 2.3 wt% of lithium modified kalsilite, an almost 100% biodiesel yield was achieved.


3.2.2.11. Hydrotalcite metal oxides as heterogeneous catalysts. Georgogianni et al.150 studied the transesterification of rapeseed oil for biodiesel production where an alkaline catalyst was used as either a homogeneous or heterogeneous catalyst. In this study, NaOH was used as a homogeneous catalyst, Mg MCM-41, Mg–Al hydrotalcite, and K+ impregnated zirconia were used as heterogeneous catalysts, and a 24 kHz ultrasonication frequency and 600 rpm stirring speed were maintained. Heterogeneous catalyst criteria depend on the porosity and surface basicity of the catalyst, and the catalytic activity depends on the basic strength of the catalyst. An approximately 97% biodiesel yield can be achieved with Mg–Al, showing the highest activity. Increasing the amount of potassium cations with the catalyst increases the activity of ZrO2 in transesterification. The involvement of the ultrasonication frequency enhanced the transesterification reaction rather than the involvement of mechanical stirring. By performing filtration, the heterogeneous catalyst can be easily separated from the reaction mixture after the reaction.

Glišić et al.151 studied various systems for developing FAME production at higher temperatures and pressures with or without catalysts. His aim was to find the optimum way to produce biodiesel with minimum energy consumption and improve life cycle energy efficiency. Energy consumption (MJ kg−1 FAME) depends on the degree of conversion of triglycerides. Energy consumption will be 25% lesser if the degree of conversion increased from 97 wt% to complete conversion. Other meaningful decreases in energy consumption might be gained at subcritical conditions by using an appropriate catalyst. This study exposed that if the heterogeneous process of biodiesel synthesis is acknowledged at subcritical conditions, then a further decrease in energy consumption is possible.

A comparative assessment, in terms of yield, reaction conditions and reusability of different promising heterogeneous catalysts utilized by several researchers has been listed in Table 5.

Table 5 Comparative assessment of promising heterogeneous catalysts with different feedstocks for biodiesel production
Catalyst name Feedstock Reaction conditions Yield (v/v%) Ref.
Temp. (°C) Time Loading in wt% Molar alcohol to oil ratio
Tungstated zirconia WO3/ZrO2 Vegetable oil 75 140 h 65 Kiss et al.71
Sulphated zirconia (SO42−/ZrO2) Palm kernel 90.3 Jitputti et al.74
Unsulfated zirconia (ZrO2) Coconut oil 86.3 Jitputti et al.74
Chloroaluminate ([Et3NH]Cl–AlCl3) Soybean oil 70 9 h 0.5[thin space (1/6-em)]:[thin space (1/6-em)]1 98.5 Liang et al.75
Cs2.5H0.5PW12O40 Yellow horn oil 70 10 min 1% 12[thin space (1/6-em)]:[thin space (1/6-em)]1 96.22 Zhang et al.65
La/zeolite beta Soybean oil 100 3 h     48.9 Shu et al.84
Hβ-zeolite Pongamia pinnata 120 24 h 0.12% 10[thin space (1/6-em)]:[thin space (1/6-em)]1 59 Karmee et al.85
Montmorillonite K-10 Pongamia pinnata 120 24 h 0.12% 11[thin space (1/6-em)]:[thin space (1/6-em)]1 47 Karmee et al.85
ZnO Pongamia pinnata 120 24 h 0.12% 12[thin space (1/6-em)]:[thin space (1/6-em)]1 83 Karmee et al.85
NaX zeolites (Si/Al = 1.23) Pongamia pinnata 125 2 h 10% 85.6 Xie et al.4
Zeolites Sunflower oil 60 93.5–95.1 Ramos et al.87
CaO Sunflower oil 100 1.5–5.5 h 1% 6[thin space (1/6-em)]:[thin space (1/6-em)]1 91 Vujicic et al.112
CaO Palm oil 1 h 0.5–5% 3[thin space (1/6-em)]:[thin space (1/6-em)]1–15[thin space (1/6-em)]:[thin space (1/6-em)]1 48.2–95.2 Mootabadi et al.97
SrO Palm oil 1 h 67.3–95.2 Mootabadi et al.97
BaO Palm oil 1 h 55–77.3 Mootabadi et al.97
MgO Vegetable oil 600 8 h 18 López et al.115
MgO Vegetable oil   5% 12[thin space (1/6-em)]:[thin space (1/6-em)]1 92 Di Serio et al.116
BaO & SrO Vegetable oil 10–60 min 0.5–3% 3[thin space (1/6-em)]:[thin space (1/6-em)]1–15[thin space (1/6-em)]:[thin space (1/6-em)]1 95% Salamatinia et al.124
SrO Vegetable oil 65 30 min 3% 12[thin space (1/6-em)]:[thin space (1/6-em)]1 90 Liu et al.123
KOH/AC Palm oil 70 15 h 5% 15[thin space (1/6-em)]:[thin space (1/6-em)]1 94 Dehkhoda et al.128
Silica gel Vegetable oil 3 h Faria et al.131
CaO from eggshell Palm olein 65   10% 9[thin space (1/6-em)]:[thin space (1/6-em)]1 97.98 Chakraborty et al.134
CaO from eggshell Palm olein 60 1 h 10% 18[thin space (1/6-em)]:[thin space (1/6-em)]1 97 Viriya-empikul et al.135
Golden apple Palm olein 60 1 h 10% 18[thin space (1/6-em)]:[thin space (1/6-em)]1 83 Viriya-empikul et al.135
Meretrix venus Palm olein 60 1 h 10% 18[thin space (1/6-em)]:[thin space (1/6-em)]1 78 Viriya-empikul et al.135
CaO from eggshell Soybean oil 70 5 h 5% 6.9[thin space (1/6-em)]:[thin space (1/6-em)]1 97.73 Du et al.136
NaX zeolites loaded with 10% KOH   65 8 h 3% 10[thin space (1/6-em)]:[thin space (1/6-em)]1 85.6 Xie et al.4
Dimethyl carbonate Soybean oil 70 5 h 6% 30[thin space (1/6-em)]:[thin space (1/6-em)]1 99 Fabbri et al.138
KF/Zn(Al)O Vegetable oil 65 3 h 3% 6[thin space (1/6-em)]:[thin space (1/6-em)]1 95 Wang et al.148
Kalsilite (KAlSiO4) Soybean oil 120 2.30% 100 Wen et al.149


3.3. Biocatalysts

Biocatalysts are catalysts obtained from a living organism and promote chemical reactions without being affected themselves. Biocatalysts are also called enzymes or enzyme catalysts. These catalysts initiate or modify nearly all biochemical reactions in living cells. Each enzyme possesses individual three-dimensional patterns that suit the shape of the reactants. Biocatalysts have recently become progressively critical in the examination of biodiesel generation. These catalysts are hypothesized to beat synthetic catalysts.152 The search for an environmentally friendly methodology for biodiesel synthesis was explored utilizing proteins as catalysts. Typically, the difficulties confronted with conventional catalysts such as feedstock pretreatment, catalyst elimination, wastewater treatment, and the requirements for large amounts of energy are addressed in enzyme-catalyzed transesterification reactions. Biocatalysis utilizes an accumulation of enzymes known as lipases generated through microorganisms, animals, and plants.153,154 Lipases could be separated from different sources of bacterial species, such as Pseudomonas fluorescens, Pseudomonas cepacia, Rhizomucor miehei, Rhizopus oryzae, Candida rugosa, Thermomyces lanuginosus, and Candida antarctica.155

Biocatalysts could be classified into a range of major varieties as follows. (a) Microbes: a microscopic organism such as yeast and other anaerobic bacteria, archaea, bacteria, fungi, viruses, and microbial mergers. (b) Lipases: the most widely used class of enzymes in organic synthesis, lipases are preferred widely because of their better stability compared with other biocatalysts. Lipase can be classified into extracellular and intracellular lipases.154 Extracellular lipases are mainly obtained from live-producing microorganism broth through purification. Major extracellular microorganisms are Mucor miehei, R. oryzae, C. antarctica, and P. cepacia. Intracellular lipases are present inside the cell or in the cell-producing wall. In most cases, intracellular lipases are found in the immobilized form. (c) Proteases: enzymes that break down proteins. Proteases can be found in animals, plants, bacteria, archaea, and viruses, for example, TEV protease and trypsinogen. (d) Cellulases: enzymes that break down cellulose. (e) Amylases: enzymes that break down starch into simple sugars.

Researchers choose enzyme transesterification because of the advantages over the chemical catalyzed transesterification process. Easy product removal, moderate process temperature (35–45 °C), zero by-products, and the reusability of the catalysts are the main features of using this type of catalyst. However, enzymatic reactions are insensitive to FFAs and the water content of the feedstock. Some investigations conducted by researchers using lipase-catalyzed transesterification of different feedstocks to produce biodiesel are presented in Table 6. Different types of catalysts used in the biodiesel production process are listed in Table 7.

Table 6 Transesterification of different feedstocks using biocatalysts
Enzyme Feedstock Alcohol Temperature (°C) Time (h) Yield (%) Ref.
IM B. cepacia lipase Palm oil Methanol 30 72 100 156
Lipozyme IM60 Tallow oil Primary alcohols 45 5 94.8–98.5 157
Lipozyme IM-20 Mowrah oil Alcohols (C4–C18) 60 6 86.8–99.2 158
E. aerogenes lipase Jatropha oil Methanol 55 48 94 159
Novozym® 435 Soybean oil Methyl acetate 40 14 92 160
Lipozyme IM-77 Soybean oil Methanol 92.2 161
R. oryzae lipase Plant oil Methanol 90 162
P. expansum lipase Corn oil Methanol 40 24 86 163
Cryptococcus spp. S-2 Rice bran oil Methanol 30 96 80 164
Pseudomonas lipase Sunflower oil Methanol 45 5 79 165
PS 30 lipase Palm oil Ethanol 40 8 72 166


Table 7 Relative advantages and disadvantages of different catalysts involved in biodiesel production technology
Catalyst type Advantages Disadvantages Ref.
Homogeneous • Effectively active on metal atoms • Hazardous for the environment compared to heterogeneous catalysts 56 and 167
• Very fast reaction rate • Hydroscopic nature (NaOH, KOH)
• Reaction can occur under mild conditions thus relatively less energy is required for esterification and transesterification • Low quality glycerol produced thus requires a lengthy distillation process for purification
• Relatively cheap and available (NaOH and KOH) • Homogeneous base catalysts are sensitive to the FFA content present in the oil
• Preferred method for low-grade feedstock • Soap forms in the case of higher (2 wt%) FFA content in the oil, thus the biodiesel conversion rate is reduced
• Sometimes esterification and transesterification occur instantaneously • Purification of biodiesel from the product is relatively difficult and requires a huge amount of water
  • Poisoning occurs when the catalyst is exposed to ambient air
• Acid homogeneous catalysts are very harmful, very corrosive to the reactor and pipeline and require careful handling
Heterogeneous • The separation of the glycerol and catalyst from biodiesel is much easier • Converts triglycerides at a relatively slower rate 56
• Not mixed with ethanol or methanol thus separation is easier • Complicated catalyst synthesis procedures leads to higher cost
• Because of large pore size, the diffusion problem is minimized • Effectively active only on surface atoms
• High catalytic stability against leaching and poisoning  
• Easy separation of the catalyst from the product
• Economic because of its reusable nature
Biocatalyst • Tolerate free fatty acids and water content • Could not be commercialized for the production of biodiesel due to the long residence time and high cost 19 and 56
• Easy purification of biodiesel and glycol • High cost
• Environmentally friendly and does not produces volatile organic compounds • Long process time due to very slow reaction rate
• High possibility to reuse and regenerate the catalyst • Sensitive to alcohol, normally methanol, that can deactivated the enzyme
• Only a simple purification step is required  
• Transesterification can be carried out at low reaction temperatures, even lower than those for a homogeneous base catalyst


4. Non-catalytic biodiesel production

Several techniques are involved in biodiesel synthesis aside from a catalytic response, such as the purification of esters, supercritical transesterification, BIOX co-solvent transesterification, and catalyst separation. Considering the advantages over catalytic biodiesel synthesis, as well as the environmental impact, supercritical transesterification, and BIOX co-solvent transesterification processes are receiving popularity.168

4.1. Supercritical alcohol transesterification

Biodiesel can be produced by the transesterification reaction under supercritical conditions in methanol (SC MeOH) without using any catalyst.104,169 Forward reaction rates are as high as 50–95%, and forward reactions occur within the initial 10 minutes.170,171 However, this reaction needs higher temperatures and pressures, approximately 250–400 °C and 1200 psi,172,173 respectively. For this reason, this reaction consumes more energy and increases the production cost. Supercritical methanol, ethanol, propanol, and butanol have been demonstrated to be involved in probably the best procedures for the transesterification of triglycerides without the involvement of a catalyst. Supercritical alcohol techniques are a non-catalytic method for biodiesel synthesis, where high pressures and high temperatures sustain the transesterification reaction.174 Initially, conversion is significantly fast because of the reaction conditions, which are high temperatures, high pressures, and high ratios of alcohol to oil. An approximately 50–95% conversion occurs within the first 10 minutes but temperature ranges are quite high (250–400 °C). During the transesterification of vegetable oil, an approximately 1[thin space (1/6-em)]:[thin space (1/6-em)]6 to 1[thin space (1/6-em)]:[thin space (1/6-em)]40 oil/alcohol molar ratio is used under supercritical alcohol conditions.109

The major difficulties or shortcomings of using supercritical alcohol are high temperatures (250–400 °C), high pressures (200–400 bar), and64 high alcohol to oil ratios,175 usually at 41[thin space (1/6-em)]:[thin space (1/6-em)]1, which increases the biodiesel production cost. In the supercritical alcohol transesterification process, the presence of small amounts of water does not hamper the oil to biodiesel conversion.104,176,177 Moreover, the presence of water accelerates the formation of the methyl ester and the esterification of FFA in one stage. The reaction took 4 min at 250–400 °C and 35–60 MPa. Iijima et al.178 suggested a supercritical condition where reaction conditions were a reaction temperature of 643–773 K, a reaction pressure of 20–60 MPa, and a reaction time of only 4–12 min for the production of biodiesel without a glycerin by-product.

The supercritical transesterification procedure involves high temperature and pressure conditions that enhance phase solubility, reduce mass transfer impediments, provide higher conversion rates, and provide easier steps for separation and purification. Table 8 presents the supercritical transesterification of vegetable oil with corresponding reaction conditions.179

Table 8 Reaction parameters and corresponding yields of supercritical alcohol processes
Reaction parameter Unit Methanol Ethanol 1-Propanol 1-Butanol 1-Octanol
Temperature °C 239.2 243.2 264.2 287.2 385
Pressure MPa 8.09 6.38 5.06 4.9 2.86
Yield after 10 min Mass% 98 79 81 80
Yield after 30 min Mass% 98 88 85 75


4.2. BIOX co-solvent transesterification

BIOX (co-solvent) transesterification is a new method proposed by Canadian professor David Boocock from the University of Toronto. Owing to the lower solubility of methanol in oil, the rate of transformation of the oil directly into the ester is quite slow. This problem could be minimized by introducing a co-solvent, which can mix with the methanol and oil. Tetrahydrofuran (THF) is a type of co-solvent with a boiling point very close to that of methanol; thus, it requires a very low operating temperature of 30 °C. The continuous BIOX co-solvent process takes less than 90 min near ambient temperature and at atmospheric pressure. FFA and triglycerides both converted in a single phase in two steps.180 By improving the solubility of the alcohol in triglyceride using the co-solvent, the slow reaction rate could be improved. The reaction time becomes 5–10 min and, aside from catalyst residues, is minimized to only one phase (ester or the glycerol phase).

Demirbas181 studied THF as a co-solvent with methanol to form a single phase. After finishing the reaction, the biodiesel glycerol phase was clean. The alcohol and THF co-solvent were both recovered in a single step. Nevertheless, because of the probable hazards and toxicity of the co-solvents, the solvent needs to be entirely removed from the glycerol phase and the biodiesel phase; in addition, the end products must be water-free.182 Using a co-solvent such as tetrahydrofuran as well as methyl tertiary butyl ether significantly accelerates methanolysis. In spite of this, similar to one-phase butanolysis, one-phase methanolysis initially displays a rapid development of the ester, but then drastically slows.182

The restoration of excessive alcohol is challenging in the case of the BIOX co-solvent approach because the boiling point for the THF co-solvent is quite close to that of methanol.183 The particular remarkable benefit of the BIOX co-solvent technique is that it employs inert, recoverable co-solvents in a single pass reaction that normally requires only seconds at ambient temperature and pressure, and no catalyst residues appear in both the biodiesel phase and the glycerol phase.183 This technique can be used with crude vegetable oils, waste cooking oils, and animal fats. Table 9 presents a comparative study between catalytic and non-catalytic transesterification biodiesel production processes.

Table 9 Advantages and disadvantages of using non-catalytic transesterification over conventional transesterificationa
  Advantages Disadvantages
a Ref. 180​–183
Non-catalytic method • Less water is produced as a by product and sometimes the presence of water accelerates the conversion rate • More energy is required by the reaction step especially in the heating step as a high power consumption is involved
• Simpler purification steps involved • High temperature and pressure required
• Simpler separation steps involved • High alcohol to oil ratio is needed
• High quality glycerin is generated as a by product • Relatively lower production yield than for the conventional method
• Environmentally friendly as a smaller amount of chemicals are used • Need to take more care over the production process as it involves a higher pressure and temperature
• Less time required  
• Low quality feedstock could be transformed easily into biodiesel
Conventional method • Lower power consumption regarding heating • Higher process cost
• Higher yield is possible • Greater time is required than for the non-conventional method
• Relatively lower temperature and pressure required • Cost involved with catalyst loading
  • Preparation of catalyst is quite complex


5. Conclusion

Among the several biodiesel synthesis processes from natural oils and fats, transesterification is currently the most attractive approach because transesterification is essentially a successive response. The motivation behind the methodology is to reduce the viscosity of the oil or fat. Although mixing oils and different solvents and microemulsions of vegetable oils reduces the viscosity, problems of engine performance (i.e., lubricating oil contamination, carbon decomposition) still exist. Involving specialized focal points in both materials science and reactor design is crucial if biodiesel is to remain a vital participant in the renewable energy sector in the 21st century.

The targets of this study were to review distinctive biodiesel generation techniques (both catalytic and non-catalytic) and the utilization of heterogeneous catalysts in biodiesel production to date. In view of the audit, the accompanying conclusions are:

• A base homogeneous catalyst process has a rapid reaction rate, high yield, needs mild reaction conditions, and has a lower energy consumption; it is insensitive to water content, easy to obtain, and has a low cost. However, this process is sensitive to FFA content (>2 wt%) in the oil and forms soap and glycerol as byproducts, thereby needing excess water during purification.

• Acid homogeneous catalysts are insensitive to both FFAs and water content in oil and are suitable for low-grade oil, which has a high acid value. However, some problems are associated with this catalyst compared with the heterogeneous catalyst, including a relatively slow reaction rate, corrosive nature, catalyst separation from the product, reusability of catalyst, and soap formation.

• A heterogeneous catalyst has some advantages over a homogeneous catalyst, including easy separation, simple recovery techniques, and reusability of the catalyst from the product. In addition, this catalyst has faster reaction rate (base) and milder reaction conditions are required compared to the homogeneous process. Sometimes, catalyst preparation is expensive and unavailable. This catalyst is still a long way from industrial applications because its assessment has only been completed in stirred batch reactors. Only a few studies have been conducted on persistent procedures utilizing pressed bed stream reactors.

• A relatively lower reaction temperature is needed for biocatalytic processes compared with that for both homogeneous acid and base catalytic processes. The major limitation is the preparation cost of the enzymes and the reaction rate, and this method is the slowest among the processes. Furthermore, the catalyst is more expensive than that for both the homogeneous acid and base catalytic processes. In addition, the low solubility of glycerin in biodiesel reduces the enzyme activity. Further development of the existing process, enzyme flexibility, and adaptability must be studied, which can lower the cost and improve the conversion rate.

• Non-catalytic supercritical alcohol procedures need harsh reaction conditions (temperature, pressure, and methanol to oil ratio). However, non-catalytic supercritical alcohol transesterification takes less time, and FFAs completely convert into the ester. In addition, the BIOX co-solvent process is a more acceptable process to minimize the low solubility problem of methanol in oils.

Nomenclature and abbreviations

ICInternal combustion
CICompression ignition
SISpark ignition
HCHydrocarbon
ASTMAmerican Society for Testing and Materials
ENEuropean Standard
FAMEFatty acid methyl ester
FAEEFatty acid ethyl ester
FFAFree fatty acid
SC MeOHSupercritical condition in presence of methanol
psiPound per square inch
HPAsHeteropolyacids
MATMicrowave-assisted transesterification
kHzKilo hertz
MJ kg−1Mega joule per kilogram
°CDegree celsius
RSMResponse surface methodology
TGsTriglycerides
DGsDiglycerides
MGsMonoglycerides
IUPACInternational Union of Pure and Applied Chemistry
XRDX-ray diffraction
SEMScanning electron microscopy
HR-TEMHigh resolution transmission electron microscopy
TEVTobacco etch virus
SMOSodium methylate

Acknowledgements

The authors would like to thank University of Malaya for financial support through High Impact Research grant titled: Clean Diesel Technology for Military and Civilian Transport Vehicles having grant number UM.C/HIR/MOHE/ENG/07 and Program Rakan Penyelidikan Universiti Malaya (PRPUM) (Grant no. CG054-2013).

References

  1. L. C. Meher, D. V. Sagar and S. N. Naik, Renewable Sustainable Energy Rev., 2006, 10, 248–268 CrossRef CAS.
  2. P. L. Boey, G. P. Maniam and S. A. Hamid, Bioresour. Technol., 2009, 100, 6362–6368 CrossRef CAS PubMed.
  3. E. S. Umdu, M. Tuncer and E. Seker, Bioresour. Technol., 2009, 100, 2828–2831 CrossRef CAS PubMed.
  4. W. Xie, X. Huang and H. Li, Bioresour. Technol., 2007, 98, 936–939 CrossRef CAS PubMed.
  5. K. Ramachandran, T. Suganya, N. Nagendra Gandhi and S. Renganathan, Renewable Sustainable Energy Rev., 2013, 22, 410–418 CrossRef CAS.
  6. M. I. Arbab, H. H. Masjuki, M. Varman, M. A. Kalam, S. Imtenan and H. Sajjad, Renewable Sustainable Energy Rev., 2013, 22, 133–147 CrossRef CAS.
  7. S. Basumatary, J. Chem. Pharm. Res., 2013, 5, 1–7 CAS.
  8. T. Krawczyk, Inform, 1996, 7, 801–829 Search PubMed.
  9. T. Murayama, Inform, 1994, 5, 1138–1145 Search PubMed.
  10. L. P. Christopher, K. Hemanathan and V. P. Zambare, Appl. Energy, 2014, 119, 497–520 CrossRef CAS.
  11. S. Maiti, P. Bapat, P. Das and P. K. Ghosh, Fuel, 2014, 121, 126–132 CrossRef CAS.
  12. P. K. Ghosh, S. C. P. Mishra, M. R. Gandhi, S. C. Upadhyay, S. K. Mishra, I. Pancha, A. V. Shrivastav, D. Jain, B. Shethia and S. Maiti, US Pat., US 8741628 B2, 2014.
  13. P. K. Ghosh, S. C. P. Mishra, M. R. Gandhi, S. C. Upadhyay, P. Paul, P. S. Anand, K. M. Popat, A. V. Shrivastav, S. K. Mishra and N. Ondhiya, US Pat., US 8956836 B2, 2015.
  14. N. Sonntag, Bailey’s industrial oil and fat products, 1979, vol. 1, pp. 99–176 Search PubMed.
  15. K. Maher and D. Bressler, Bioresour. Technol., 2007, 98, 2351–2368 CrossRef CAS PubMed.
  16. H. Lappi and R. Alén, J. Anal. Appl. Pyrolysis, 2009, 86, 274–280 CrossRef CAS.
  17. H. Lappi and R. Alén, J. Anal. Appl. Pyrolysis, 2011, 91, 154–158 CrossRef CAS.
  18. D. G. Lima, V. C. Soares, E. B. Ribeiro, D. A. Carvalho, É. C. Cardoso, F. C. Rassi, K. C. Mundim, J. C. Rubim and P. A. Suarez, J. Anal. Appl. Pyrolysis, 2004, 71, 987–996 CrossRef CAS.
  19. A. E. Atabani, A. S. Silitonga, H. C. Ong, T. M. I. Mahlia, H. H. Masjuki, I. A. Badruddin and H. Fayaz, Renewable Sustainable Energy Rev., 2013, 18, 211–245 CrossRef CAS.
  20. A. Schwab, G. Dykstra, E. Selke, S. Sorenson and E. Pryde, J. Am. Oil Chem. Soc., 1988, 65, 1781–1786 CrossRef CAS.
  21. F. Ma and M. A. Hanna, Bioresour. Technol., 1999, 70, 1–15 CrossRef CAS.
  22. S. Şensöz, D. Angın and S. Yorgun, Biomass Bioenergy, 2000, 19, 271–279 CrossRef.
  23. I. Fortes and P. Baugh, J. Anal. Appl. Pyrolysis, 2004, 72, 103–111 CrossRef CAS.
  24. W. Liu, P. Yin, X. Liu, S. Zhang and R. Qu, J. Ind. Eng. Chem., 2015, 21, 893–899 CrossRef CAS.
  25. D. Y. C. Leung, X. Wu and M. K. H. Leung, Appl. Energy, 2010, 87, 1083–1095 CrossRef CAS.
  26. Y. M. Park, J. Y. Lee, S. H. Chung, I. S. Park, S. Y. Lee, D. K. Kim, J. S. Lee and K. Y. Lee, Bioresour. Technol., 2010, 101, S59–S61 CrossRef CAS PubMed.
  27. N. Kapilan and B. D. Baykov, Pet. Coal, 2014, 56, 62–73 Search PubMed.
  28. W. K. Teng, G. C. Ngoh, R. Yusoff and M. K. Aroua, Chem. Eng. J., 2016, 284, 469–477 CrossRef CAS.
  29. M. A. M. Ali, R. M. Yunus, C. K. Cheng and J. Gimbun, RSC Adv., 2015, 5, 76743–76751 RSC.
  30. Y. Zu, S. Zhng, Y. Fu, W. Liu, Z. Liu, M. Luo and T. Efferth, Eur. Food Res. Technol., 2010, 229, 43–49 CrossRef.
  31. P. D. Patil, H. Reddy, T. Muppaneni, A. Mannarswamy, T. Schuab, F. O. Holguin, P. Lammers, N. Nirmalakhandan, P. Cooke and S. Deng, Green Chem., 2012, 14, 809–818 RSC.
  32. J. Ji, J. Wang, Y. Li, Y. Yu and Z. Xu, Ultrasonics, 2006, 44, e411–e414 CrossRef PubMed.
  33. A. P. Vyas, J. L. Verma and N. Subrahmanyam, Fuel, 2010, 89, 1–9 CrossRef CAS.
  34. J. A. Colucci, E. E. Borrero and F. Alape, J. Am. Oil Chem. Soc., 2005, 82, 525–530 CrossRef CAS.
  35. P. Maneechakr, J. Samerjit and S. Karnjanakom, RSC Adv., 2015, 5, 55252–55261 RSC.
  36. S. B. Lee, J. D. Lee and I. K. Hong, J. Ind. Eng. Chem., 2011, 17, 138–143 CrossRef CAS.
  37. D. Y. Leung, X. Wu and M. Leung, Appl. Energy, 2010, 87, 1083–1095 CrossRef CAS.
  38. C. Goering, M. Daugherty, A. Heakin, E. Pryde and A. Schwab, Am. Soc. Agric. Eng., Pap., 1982, 25, 5696175 Search PubMed.
  39. K. Varatharajan and M. Cheralathan, Fuel Process. Technol., 2013, 106, 526–532 CrossRef CAS.
  40. A. Demirbas, Prog. Energy Combust. Sci., 2005, 31, 466–487 CrossRef CAS.
  41. I. R. Fattah, H. Masjuki, M. Kalam, M. Mofijur and M. Abedin, Energy Convers. Manage., 2014, 79, 265–272 CrossRef.
  42. A. Atabani, T. Mahlia, H. Masjuki, I. A. Badruddin, H. W. Yussof, W. Chong and K. T. Lee, Energy, 2013, 58, 296–304 CrossRef CAS.
  43. U. Rashid, F. Anwar, B. R. Moser and G. Knothe, Bioresour. Technol., 2008, 99, 8175–8179 CrossRef CAS PubMed.
  44. S. D. Sanford, J. M. White, P. S. Shah, C. Wee, M. A. Valverde and G. R. Meier, Feedstock and biodiesel characteristics report, Renewable Energy Group, 2009, vol. 416, pp. 1–136 Search PubMed.
  45. A. Atabani, A. Silitonga, H. Ong, T. Mahlia, H. Masjuki, I. A. Badruddin and H. Fayaz, Renewable Sustainable Energy Rev., 2013, 18, 211–245 CrossRef CAS.
  46. N. Kumar and S. R. Chauhan, Renewable Sustainable Energy Rev., 2013, 21, 633–658 CrossRef CAS.
  47. S. Palash, H. Masjuki, M. Kalam, A. Atabani, I. R. Fattah and A. Sanjid, Energy Convers. Manage., 2015, 91, 149–157 CrossRef CAS.
  48. M. Wakil, M. Kalam, H. Masjuki, I. R. Fattah and B. Masum, RSC Adv., 2014, 4, 56984–56991 RSC.
  49. A. Karmakar, S. Karmakar and S. Mukherjee, Renewable Sustainable Energy Rev., 2012, 16, 1050–1060 CrossRef CAS.
  50. S. Ragit, S. Mohapatra and K. Kundu, J. Sci. Ind. Res., 2010, 69, 62–66 CAS.
  51. L. Danielski, C. Zetzl, H. Hense and G. Brunner, J. Supercrit. Fluids, 2005, 34, 133–141 CrossRef CAS.
  52. L. Meher, D. V. Sagar and S. Naik, Renewable Sustainable Energy Rev., 2006, 10, 248–268 CrossRef CAS.
  53. Z. Helwani, M. Othman, N. Aziz, J. Kim and W. Fernando, Appl. Catal., A, 2009, 363, 1–10 CrossRef CAS.
  54. P. Mohanty, K. K. Pant, S. N. Naik, J. Parikh, A. Hornung and J. N. Sahu, Renewable Sustainable Energy Rev., 2014, 38, 131–153 CrossRef CAS.
  55. Y. H. Tan, M. O. Abdullah and C. Nolasco-Hipolito, Renewable Sustainable Energy Rev., 2015, 47, 589–603 CrossRef CAS.
  56. M. K. Lam, K. T. Lee and A. R. Mohamed, Biotechnol. Adv., 2010, 28, 500–518 CrossRef CAS PubMed.
  57. E. Lotero, Y. Liu, D. E. Lopez, K. Suwannakarn, D. A. Bruce and J. G. Goodwin, Ind. Eng. Chem. Res., 2005, 44, 5353–5363 CrossRef CAS.
  58. A. S. Ramadhas, S. Jayaraj and C. Muraleedharan, Fuel, 2005, 84, 335–340 CrossRef CAS.
  59. A. K. Tiwari, A. Kumar and H. Raheman, Biomass Bioenergy, 2007, 31, 569–575 CrossRef.
  60. P. Sahoo, L. Das, M. Babu and S. Naik, Fuel, 2007, 86, 448–454 CrossRef CAS.
  61. M. Mittelbach and C. Remschmidt, Biodiesel:the comprehensive handbook, ed. M. Mittelbach, 2004 Search PubMed.
  62. M. Di Serio, R. Tesser, L. Pengmei and E. Santacesaria, Energy Fuels, 2007, 22, 207–217 CrossRef.
  63. T. F. Dossin, M.-F. Reyniers, R. J. Berger and G. B. Marin, Appl. Catal., B, 2006, 67, 136–148 CrossRef CAS.
  64. J. A. Melero, J. Iglesias and G. Morales, Green Chem., 2009, 11, 1285–1308 RSC.
  65. S. Zhang, Y. G. Zu, Y. J. Fu, M. Luo, D. Y. Zhang and T. Efferth, Bioresour. Technol., 2010, 101, 931–936 CrossRef CAS PubMed.
  66. J. Maçaira, A. Santana, F. Recasens and M. A. Larrayoz, Fuel, 2011, 90, 2280–2288 CrossRef.
  67. E. F. Aransiola, T. V. Ojumu, O. O. Oyekola, T. F. Madzimbamuto and D. I. O. Ikhu-Omoregbe, Biomass Bioenergy, 2014, 61, 276–297 CrossRef CAS.
  68. A. Gupta and A. Gupta, Journal of Scientific and Industrial Research, 2004, 63, 39–47 Search PubMed.
  69. H. Muthu, V. SathyaSelvabala, T. Varathachary, D. Kirupha Selvaraj, J. Nandagopal and S. Subramanian, Braz. J. Chem. Eng., 2010, 27, 601–608 CAS.
  70. A. P. S. Chouhan and A. K. Sarma, Renewable Sustainable Energy Rev., 2011, 15, 4378–4399 CrossRef CAS.
  71. A. A. Kiss, A. C. Dimian and G. Rothenberg, Adv. Synth. Catal., 2006, 348, 75–81 CrossRef CAS.
  72. Y. M. Park, D. W. Lee, D. K. Kim, J. S. Lee and K. Y. Lee, Catal. Today, 2008, 131, 238–243 CrossRef CAS.
  73. S. Furuta, H. Matsuhashi and K. Arata, Biomass Bioenergy, 2006, 30, 870–873 CrossRef CAS.
  74. J. Jitputti, B. Kitiyanan, P. Rangsunvigit, K. Bunyakiat, L. Attanatho and P. Jenvanitpanjakul, Chem. Eng. J., 2006, 116, 61–66 CrossRef CAS.
  75. X. Liang, G. Gong, H. Wu and J. Yang, Fuel, 2009, 88, 613–616 CrossRef CAS.
  76. S. M. de Rezende, M. de Castro Reis, M. G. Reid, P. L. Silva, F. M. Coutinho, R. A. d. S. San Gil and E. R. Lachter, Appl. Catal., A, 2008, 349, 198–203 CrossRef CAS.
  77. Y. Chang and C. Bae, Curr. Org. Synth., 2011, 8, 208–236 CrossRef CAS.
  78. X. Mo, E. Lotero, C. Lu, Y. Liu and J. G. Goodwin, Catal. Lett., 2008, 123, 1–6 CrossRef CAS.
  79. S. Yan, C. DiMaggio, S. Mohan, M. Kim, S. O. Salley and K. S. Ng, Top. Catal., 2010, 53, 721–736 CrossRef CAS.
  80. F. Chai, F. Cao, F. Zhai, Y. Chen, X. Wang and Z. Su, Adv. Synth. Catal., 2007, 349, 1057–1065 CrossRef CAS.
  81. K. Narasimharao, D. Brown, A. F. Lee, A. Newman, P. Siril, S. Tavener and K. Wilson, J. Catal., 2007, 248, 226–234 CrossRef CAS.
  82. B. Hamad, R. L. de Souza, G. Sapaly, M. C. Rocha, P. P. de Oliveira, W. Gonzalez, E. A. Sales and N. Essayem, Catal. Commun., 2008, 10, 92–97 CrossRef CAS.
  83. C. Baerlocher, L. B. McCusker and D. H. Olson, Atlas of zeolite framework types, Elsevier, 2007 Search PubMed.
  84. Q. Shu, B. Yang, H. Yuan, S. Qing and G. Zhu, Catal. Commun., 2007, 8, 2159–2165 CrossRef CAS.
  85. S. K. Karmee and A. Chadha, Bioresour. Technol., 2005, 96, 1425–1429 CrossRef CAS PubMed.
  86. G. J. Suppes, M. A. Dasari, E. J. Doskocil, P. J. Mankidy and M. J. Goff, Appl. Catal., A, 2004, 257, 213–223 CrossRef CAS.
  87. M. J. Ramos, A. Casas, L. Rodríguez, R. Romero and Á. Pérez, Appl. Catal., A, 2008, 346, 79–85 CrossRef CAS.
  88. J. Marchetti and A. Errazu, Fuel, 2008, 87, 3477–3480 CrossRef CAS.
  89. D. M. Alonso, R. Mariscal, R. Moreno-Tost, M. D. Z. Poves and M. L. Granados, Catal. Commun., 2007, 8, 2074–2080 CrossRef CAS.
  90. N. Barakos, S. Pasias and N. Papayannakos, Bioresour. Technol., 2008, 99, 5037–5042 CrossRef CAS PubMed.
  91. C. Ngamcharussrivichai, P. Totarat and K. Bunyakiat, Appl. Catal., A, 2008, 341, 77–85 CrossRef CAS.
  92. G. Peterson and W. Scarrah, J. Am. Oil Chem. Soc., 1984, 61, 1593–1597 CrossRef CAS.
  93. M. Verziu, S. M. Coman, R. Richards and V. I. Parvulescu, Catal. Today, 2011, 167, 64–70 CrossRef CAS.
  94. J. X. Wang, K. T. Chen, J. S. Wu, P. H. Wang, S. T. Huang and C. C. Chen, Fuel Process. Technol., 2012, 104, 167–173 CrossRef CAS.
  95. Y. D. Long, F. Guo, Z. Fang, X. F. Tian, L. Q. Jiang and F. Zhang, Bioresour. Technol., 2011, 102, 6884–6886 CrossRef CAS PubMed.
  96. Y. D. Long, Z. Fang, T. C. Su and Q. Yang, Appl. Energy, 2014, 113, 1819–1825 CrossRef CAS.
  97. H. Mootabadi, B. Salamatinia, S. Bhatia and A. Z. Abdullah, Fuel, 2010, 89, 1818–1825 CrossRef CAS.
  98. L. Gao, G. Teng, G. Xiao and R. Wei, Biomass Bioenergy, 2010, 34, 1283–1288 CrossRef CAS.
  99. N. Kondamudi, S. K. Mohapatra and M. Misra, Appl. Catal., A, 2011, 393, 36–43 CrossRef CAS.
  100. M. Zabeti, W. M. A. W. Daud and M. K. Aroua, Fuel Process. Technol., 2009, 90, 770–777 CrossRef CAS.
  101. M. Kouzu, T. Kasuno, M. Tajika, Y. Sugimoto, S. Yamanaka and J. Hidaka, Fuel, 2008, 87, 2798–2806 CrossRef CAS.
  102. J. Zhang, S. Chen, R. Yang and Y. Yan, Fuel, 2010, 89, 2939–2944 CrossRef CAS.
  103. H. J. Kim, B. S. Kang, M. J. Kim, Y. M. Park, D. K. Kim, J. S. Lee and K. Y. Lee, Catal. Today, 2004, 93, 315–320 CrossRef.
  104. S. Saka and D. Kusdiana, Fuel, 2001, 80, 225–231 CrossRef CAS.
  105. V. B. Veljković, O. S. Stamenković, Z. B. Todorović, M. L. Lazić and D. U. Skala, Fuel, 2009, 88, 1554–1562 CrossRef.
  106. R. Liu, X. Wang, X. Zhao and P. Feng, Carbon, 2008, 46, 1664–1669 CrossRef CAS.
  107. M. Kouzu, T. Kasuno, M. Tajika, S. Yamanaka and J. Hidaka, Appl. Catal., A, 2008, 334, 357–365 CrossRef CAS.
  108. S. Gryglewicz, Bioresour. Technol., 1999, 70, 249–253 CrossRef CAS.
  109. A. Kawashima, K. Matsubara and K. Honda, Bioresour. Technol., 2009, 100, 696–700 CrossRef CAS PubMed.
  110. M. L. Granados, M. Poves, D. M. Alonso, R. Mariscal, F. C. Galisteo, R. Moreno-Tost, J. Santamaría and J. Fierro, Appl. Catal., B, 2007, 73, 317–326 CrossRef CAS.
  111. G. Arzamendi, I. Campo, E. Arguiñarena, M. Sánchez, M. Montes and L. M. Gandía, Chem. Eng. J., 2007, 134, 123–130 CrossRef CAS.
  112. D. Vujicic, D. Comic, A. Zarubica, R. Micic and G. Boskovic, Fuel, 2010, 89, 2054–2061 CrossRef CAS.
  113. B. Yoosuk, P. Udomsap, B. Puttasawat and P. Krasae, Chem. Eng. J., 2010, 162, 135–141 CrossRef CAS.
  114. H. X. Bai, X. Z. Shen, X. H. Liu and S. Y. Liu, Trans. Nonferrous Met. Soc. China, 2009, 19, s674–s677 CrossRef CAS.
  115. D. E. López, J. G. Goodwin Jr, D. A. Bruce and E. Lotero, Appl. Catal., A, 2005, 295, 97–105 CrossRef.
  116. M. Di Serio, R. Tesser, M. Dimiccoli, F. Cammarota, M. Nastasi and E. Santacesaria, J. Mol. Catal. A: Chem., 2005, 239, 111–115 CrossRef CAS.
  117. M. Di Serio, R. Tesser, L. Pengmei and E. Santacesaria, Energy Fuels, 2008, 22, 207–217 CrossRef CAS.
  118. T. F. Dossin, M. F. Reyniers, R. J. Berger and G. B. Marin, Appl. Catal., B, 2006, 67, 136–148 CrossRef CAS.
  119. T. Tateno and T. Sasaki, US Pat., US 6818026 B2, 2004.
  120. M. Su, R. Yang and M. Li, Fuel, 2013, 103, 398–407 CrossRef CAS.
  121. A. F. Lee, J. A. Bennett, J. C. Manayil and K. Wilson, Chem. Soc. Rev., 2014, 43, 7887–7916 RSC.
  122. M. Zabeti, W. M. A. Wan Daud and M. K. Aroua, Fuel Process. Technol., 2009, 90, 770–777 CrossRef CAS.
  123. X. Liu, H. He, Y. Wang and S. Zhu, Catal. Commun., 2007, 8, 1107–1111 CrossRef CAS.
  124. B. Salamatinia, H. Mootabadi, S. Bhatia and A. Z. Abdullah, Fuel Process. Technol., 2010, 91, 441–448 CrossRef CAS.
  125. J. L. Shumaker, C. Crofcheck, S. A. Tackett, E. Santillan-Jimenez, T. Morgan, Y. Ji, M. Crocker and T. J. Toops, Appl. Catal., B, 2008, 82, 120–130 CrossRef CAS.
  126. L. Xu, X. Yang, X. Yu, Y. Guo and Maynurkader, Catal. Commun., 2008, 9, 1607–1611 CrossRef CAS.
  127. Q. Shu, J. Gao, Z. Nawaz, Y. Liao, D. Wang and J. Wang, Appl. Energy, 2010, 87, 2589–2596 CrossRef CAS.
  128. A. M. Dehkhoda, A. H. West and N. Ellis, Appl. Catal., A, 2010, 382, 197–204 CrossRef CAS.
  129. M. Toda, A. Takagaki, M. Okamura, J. N. Kondo, S. Hayashi, K. Domen and M. Hara, Nature, 2005, 438, 178 CrossRef CAS PubMed.
  130. W. Y. Lou, M. H. Zong and Z. Q. Duan, Bioresour. Technol., 2008, 99, 8752–8758 CrossRef CAS PubMed.
  131. E. A. Faria, H. F. Ramalho, J. S. Marques, P. A. Z. Suarez and A. G. S. Prado, Appl. Catal., A, 2008, 338, 72–78 CrossRef CAS.
  132. M. J. Climent, A. Corma and S. Iborra, RSC Adv., 2012, 2, 16–58 RSC.
  133. P.-L. Boey, G. P. Maniam and S. A. Hamid, Bioresour. Technol., 2009, 100, 6362–6368 CrossRef CAS PubMed.
  134. R. Chakraborty, S. Bepari and A. Banerjee, Bioresour. Technol., 2011, 102, 3610–3618 CrossRef CAS PubMed.
  135. N. Viriya-empikul, P. Krasae, B. Puttasawat, B. Yoosuk, N. Chollacoop and K. Faungnawakij, Bioresour. Technol., 2010, 101, 3765–3767 CrossRef CAS PubMed.
  136. W. Du, Y. Xu, D. Liu and J. Zeng, J. Mol. Catal. B: Enzym., 2004, 30, 125–129 CrossRef CAS.
  137. N. Nakatani, H. Takamori, K. Takeda and H. Sakugawa, Bioresour. Technol., 2009, 100, 1510–1513 CrossRef CAS PubMed.
  138. D. Fabbri, V. Bevoni, M. Notari and F. Rivetti, Fuel, 2007, 86, 690–697 CrossRef CAS.
  139. N. Kondamudi, S. K. Mohapatra and M. Misra, Appl. Catal., A, 2011, 393, 36–43 CrossRef CAS.
  140. S. J. Yoo, H. S. Lee, B. Veriansyah, J. Kim, J. D. Kim and Y. W. Lee, Bioresour. Technol., 2010, 101, 8686–8689 CrossRef CAS PubMed.
  141. X. Xiao, J. W. Tierney and I. Wender, Appl. Catal., A, 1999, 183, 209–219 CrossRef CAS.
  142. M. Stöcker, J. Mol. Catal., 1985, 29, 371–377 CrossRef.
  143. D. E. López, K. Suwannakarn, D. A. Bruce and J. G. Goodwin Jr, J. Catal., 2007, 247, 43–50 CrossRef.
  144. S. Nakagaki, A. Bail, V. C. d. Santos, V. H. R. d. Souza, H. Vrubel, F. S. Nunes and L. P. Ramos, Appl. Catal., A, 2008, 351, 267–274 CrossRef CAS.
  145. Y. C. Brito, V. M. Mello, C. C. S. Macedo, M. R. Meneghetti, P. A. Z. Suarez and S. M. P. Meneghetti, Appl. Catal., A, 2008, 351, 24–28 CrossRef CAS.
  146. R. B. da Silva, A. F. Lima Neto, L. S. Soares dos Santos, J. R. de Oliveira Lima, M. H. Chaves, J. R. dos Santos Jr, G. M. de Lima, E. M. de Moura and C. V. R. de Moura, Bioresour. Technol., 2008, 99, 6793–6798 CrossRef PubMed.
  147. B. J. Krohn, C. V. McNeff, B. Yan and D. Nowlan, Bioresour. Technol., 2011, 102, 94–100 CrossRef CAS PubMed.
  148. Y. Wang, H. Wu and M. Zong, Bioresour. Technol., 2008, 99, 7232–7237 CrossRef CAS PubMed.
  149. G. Wen, Z. Yan, M. Smith, P. Zhang and B. Wen, Fuel, 2010, 89, 2163–2165 CrossRef CAS.
  150. K. Georgogianni, A. Katsoulidis, P. Pomonis and M. Kontominas, Fuel Process. Technol., 2009, 90, 671–676 CrossRef CAS.
  151. S. Glišić, I. Lukic and D. Skala, Bioresour. Technol., 2009, 100, 6347–6354 CrossRef PubMed.
  152. S. Al-Zuhair, F. W. Ling and L. S. Jun, Process Biochem., 2007, 42, 951–960 CrossRef CAS.
  153. M. Szczęsna Antczak, A. Kubiak, T. Antczak and S. Bielecki, Renewable Energy, 2009, 34, 1185–1194 CrossRef.
  154. A. Gog, M. Roman, M. Toşa, C. Paizs and F. D. Irimie, Renewable Energy, 2012, 39, 10–16 CrossRef CAS.
  155. P. T. Vasudevan and M. Briggs, J. Ind. Microbiol. Biotechnol., 2008, 35, 421–430 CrossRef CAS PubMed.
  156. K. R. Jegannathan, L. Jun Yee, E. S. Chan and P. Ravindra, Fuel, 2010, 89, 2272–2277 CrossRef CAS.
  157. P. M. Nielsen, J. Brask and L. Fjerbaek, Eur. J. Lipid Sci. Technol., 2008, 110, 692–700 CrossRef CAS.
  158. B. De, D. Bhattacharyya and C. Bandhu, J. Am. Oil Chem. Soc., 1999, 76, 451–453 CrossRef CAS.
  159. A. Kumari, P. Mahapatra, V. K. Garlapati and R. Banerjee, Biotechnol. Biofuels, 2009, 2, 1–6 CrossRef PubMed.
  160. W. Du, Y. Xu, D. Liu and J. Zeng, J. Mol. Catal. B: Enzym., 2004, 30, 125–129 CrossRef CAS.
  161. C. J. Shieh, H. F. Liao and C. C. Lee, Bioresour. Technol., 2003, 88, 103–106 CrossRef CAS PubMed.
  162. K. Ban, M. Kaieda, T. Matsumoto, A. Kondo and H. Fukuda, Biochem. Eng. J., 2001, 8, 39–43 CrossRef CAS PubMed.
  163. R. C. Chang, S. J. Chou and J. F. Shaw, Biotechnol. Appl. Biochem., 1994, 19, 93–97 CAS.
  164. N. Kamini and H. Iefuji, Process Biochem., 2001, 37, 405–410 CrossRef CAS.
  165. M. Mittelbach, J. Am. Oil Chem. Soc., 1990, 67, 168–170 CrossRef CAS.
  166. R. Abigor, P. Uadia, T. Foglia, M. Haas, K. Jones, E. Okpefa, J. Obibuzor and M. Bafor, Biochem. Soc. Trans., 2000, 28, 979–981 CrossRef CAS PubMed.
  167. J. M. Marchetti, V. U. Miguel and A. F. Errazu, Renewable Sustainable Energy Rev., 2007, 11, 1300–1311 CrossRef CAS.
  168. C. D. Silva and J. V. Oliveira, Braz. J. Chem. Eng., 2014, 31, 271–285 CrossRef.
  169. A. Demirbaş, Energy Convers. Manage., 2002, 43, 2349–2356 CrossRef.
  170. L. Wang and J. Yang, Fuel, 2007, 86, 328–333 CrossRef CAS.
  171. A. Demirbas, Journal of Scientific and Industrial Research, 2005, 64, 858–865 CAS.
  172. S. Al-Zuhair, Biofuels, Bioprod. Biorefin., 2007, 1, 57–66 CrossRef CAS.
  173. J. van Gerpen, B. Shanks, R. Pruszko, D. Clements and G. Knothe, Biodiesel Production Technology, August 2002–January 2004 Search PubMed.
  174. K. T. Tan, K. T. Lee and A. R. Mohamed, Biomass Bioenergy, 2009, 33, 1096–1099 CrossRef CAS.
  175. M. Balat and H. Balat, Energy Convers. Manage., 2008, 49, 2727–2741 CrossRef CAS.
  176. D. Kusdiana and S. Saka, Fuel, 2001, 80, 693–698 CrossRef CAS.
  177. D. Kusdiana and S. Saka, Bioresour. Technol., 2004, 91, 289–295 CrossRef CAS PubMed.
  178. W. Iijima, Y. Kobayashi and K. Taniwaki, US Pat., US 20060288636 A1, 2004.
  179. C. M. Drapcho, N. P. Nhuan and T. H. Walker, Biofuels engineering process technology, McGraw-Hill New York, NY, USA, 2008 Search PubMed.
  180. http://www.bioxcorp.com/production-process/, vol, accessed March 2015.
  181. A. Demirbas, Biodiesel: a Realistic Fuel Alternative for Diesel Engines, 2008, pp. 161–173 Search PubMed.
  182. D. G. Boocock, S. K. Konar, V. Mao and H. Sidi, Biomass Bioenergy, 1996, 11, 43–50 CrossRef CAS.
  183. J. van Gerpen, B. Shanks, R. Pruszko, D. Clements and G. Knothe, National Renewable Energy Laboratory, Colorado, 2004, pp. 37–47 Search PubMed.

This journal is © The Royal Society of Chemistry 2015