Open access peer-reviewed chapter

Biodiesel Production as a Renewable Resource for the Potential Displacement of the Petroleum Diesel

Written By

Ifeanyichukwu Edeh

Submitted: 08 April 2019 Reviewed: 25 May 2020 Published: 06 July 2020

DOI: 10.5772/intechopen.93013

From the Edited Volume

Biorefinery Concepts, Energy and Products

Edited by Venko Beschkov

Chapter metrics overview

952 Chapter Downloads

View Full Metrics

Abstract

In the quest to comply with the Intergovernmental Panel on Climate Change (IPCC) on reducing the global temperature to 1.5–2.0°C as a measure to minimize the climate change caused by the emission of greenhouse gases from the combustion of fossil fuels, and the need for replacement of these fossil fuels, which are also generally believed to be depleting, biodiesel is being studied as a potential replacement for the conventional petroleum diesel. This fuel among other desired properties is renewable, biodegradable, sustainable, and emits less particles. It also contains no amount of sulfur, in addition to possessing most of the good characteristics of petroleum diesel. At the moment, more than 95% of biodiesel produced globally is obtained from vegetable oil feedstocks, which are usually very expensive and thus, without tax waiver and subsidy, makes biodiesel non-competitive with the petroleum diesel. Based on this, non-edible feedstocks are being investigated. Although, their oil yield is low, studies are carried out to ensure efficient extraction. The economics of the process is considered to determine the most economic variables that impact the profitability of biodiesel production. This chapter deals with the biodiesel classification, feedstocks, lipid/oil extraction, biodiesel production methods and the economic aspect of the process.

Keywords

  • biodiesel
  • renewable
  • feedstocks
  • extraction
  • properties
  • technologies
  • economics

1. Introduction

The need to substitute the conventional petroleum diesel with a renewable alternative, one that is sustainable and environmentally friendly, has driven various investigators over a decade now to research on the potentials of biodiesel [1]. This has risen due to depletion of fossil fuels and emission of greenhouse gases such as CO2 and methane upon combustion, which causes climate change, the result of which is the rise in the global temperature above the nominal margin of 2°C with the potential to extinct over 1 million species [2, 3]. Other adverse effects of this global temperature rise also known as global warming include receding of glaciers, rise in sea level and loss of biodiversity [4]. However, biodiesel is a renewable fuel produced from the reaction between triacylglycerol or fatty acid with alcohol in the presence of a catalyst [5]. The fuel exists as liquid and consists of mono-alkyl esters of long-chain fatty acids with similar characteristics as the conventional petroleum diesel, making it a potential substitute [6]. Biodiesel is biodegradable, sustainable, and nontoxic, and has less impact on the environment. The shortcomings of biodiesel include low energy density, relatively high production cost and poor cold flow [7]. The global production capacity of biodiesel is envisaged to reach 12 billion gallons by 2020 with Brazil, United States of America, Malaysia, Argentina, Netherlands, Spain, Philippines, Belgium, Indonesia and Germany meeting more than 80% of the world demand [8, 9]. In 2016, the biodiesel produced globally were contributed mostly by USA and Brazil (see Table 1). Larger proportions of which are consumed by countries such as USA, Brazil, Germany, Indonesia and France [2]. Countries like US, China and India are currently experiencing a great growth in the biodiesel market with their respective governments planning to replace about 15% of the conventional diesel with biodiesel by 2020.

Country Biodiesel production/billion liters
USA 5.5
Brazil 3.8
Germany 3
Indonesia 3
Argentina 3
France 1.5
Thailand 1.4
Spain 1.1
Belgium 0.5
Colombia 0.5
Canada 0.4
China 0.3
India 0
Singapore 0

Table 1.

Countries with top biodiesel production in 2016.

Advertisement

2. Classification of biodiesel

Biodiesel can be classified into three types based on the kind of feedstocks used in its production [10]. These are first-, second- and third-generation biodiesels.

2.1 The first-generation biodiesel

This type of biodiesel is produced using edible vegetable oils. These oils are discussed in the next section. Biodiesel produced from these oils usually has the following disadvantages [11, 12, 13, 14]:

  1. Poor storage

  2. Oxidation stability

  3. High feedstock cost, up to 60−80% of biodiesel production cost

  4. Low heating value

  5. Higher NOx emission compared to the conventional diesel fuel

  6. Loss of biodiversity

2.2 Second-generation biodiesel

In order to minimize the over dependency on the edible vegetable oils feedstocks in biodiesel production, alternative sources from non-edible oils are explored. Biodiesel produced from this type of oils is known as second-generation biodiesel. The quality and yield obtained are similar to that from edible oils [15]. Lignocellulosic biomass (LCB) is also being considered as an alternative feedstock to edible oil in biodiesel production probably because it is suspected to promote faster production, less labour, more season and climate flexibility, easier scale-up, and potential economic advantage [16]. This biomass can be derived from food crops, non-food/energy crops, forest residue and industrial process residues (see Table 2). But, the most predominant is agricultural crop residues [17]. Although, some of the LCB resources might not be suitable for energy production, probably due to their wide dispersal or low bulk density, which makes energy recovery, transport and storage expensive [18]. Generally, the production of biodiesel from lignocellulosic biomass is hampered due to lack of economically feasible technologies [18].

Food crops Non-food/energy crops Forest residue Industrial process residues
*Rice straw *Cardoon (Cynara cardunculus, L.) *Tree residue (twigs, leaves, bark, and roots) *Rice husk
*Wheat straw *Giant reed (Arundo donax L.) *Wood processing residues (sawmill off-cuts and sawdust) *Rice bran
*Sugarcane tops *Salix *Recycled wood (from demolition of buildings, pallets, and packing crates) *Sugarcane bagasse
*Maize stalks millet *Jute stalks *Coconut husks
*Groundnut stalks *Willow *Maize husks
*Corn straw *Poplar *Groundnut husks
*Soybean residue *Eucalyptus
*Residue from vegetables *Miscanthus
*Residence from pulses *Reed canary grass
*Switch grass
*Hemp

Table 2.

Sources of lignocellulosic biomass for biodiesel production [17].

2.3 Third-generation biodiesel

This is produced from micro-and macro-species including algae [11]. Third-generation biodiesel is discussed further in Section 3.2.

2.4 Fourth-generation biodiesel and speculations

This can be produced from feedstocks that possess the capability of being genetically modified, accumulate large quantity of biomass, and can be utilized in photo-biological solar cells with the ability to convert solar energy directly to usable biodiesel. Example of such feedstocks is algal species. This concept focuses on producing biodiesel in addition to developing a means of trapping and storing CO2. The method of producing this energy is similar to that of the second-generation biofuels, except that CO2 is arrested at each stage of the production using techniques such as oxy-fuel combustion. The CO2 trapped is stored in saline aquifers, gas fields or old oils through the method known as geo-sequestration. The process has the capacity to trap carbon inclusively making it ‘carbon negative’ as opposed to ‘carbon neutral’ [2, 19].

Advertisement

3. Feedstock for biodiesel production

These include edible and non-edible oils, and are presented below.

3.1 Edible oils for biodiesel production

At the moment, over 95% of biodiesel globally is produced from edible vegetable oils. The commonly used of these oils are palm oil, soybean, coconut oil, rapeseed and sunflower due to their availability [12, 13, 14, 15, 19, 20]. Rapeseed oil, sunflower oil, palm oil and soybean oil are used in Europe, Malaysia, Indonesia, Philippians and US, respectively to produce biodiesel [21]. There is no doubt that the use of these feedstocks for biodiesel production competes with their need for human consumption and some other applications, the disadvantage of which is insecurity, high cost of production and potential depletion of ecological resources due to some agricultural practices. Biodiesel produced from these oils usually has the disadvantages highlighted in Section 2.1 [12, 13, 14, 21]. The sources of edible oil and the respective yield of oil are presented in Table 3.

S/N Source Yield
1. Rapeseed (Brassica oilseed) 38–46
2. Coconut 63–65
3. Soybean 15–20
4. Palm 30–60
5. Sunflower 25–35

Table 3.

Sources of edible oil used in biodiesel production [11].

3.2 Non-edible oils for biodiesel production

Non-edible oils are cultivated on lands requiring minimum attention and as such are less expensive compared to edible oils [22]. These oils include jatropha, karanja, polanga, cotton seed, Simmondsia chinensis (jojoba), tobacco, neem, linseed, rice bran oil, microalgae, mahua, waste cooking oil, animal fats, activated sludge lipid and rubber seed oils, and are used for biodiesel production [5, 23, 24, 25, 26, 27, 28], see Figure 1. The work done by Bankovic-lli et al. revealed that jatropha, karanja, mahua and castor are the most commonly sourced non-edible oils for biodiesel production [22]. The methyl esters of these oils can be blended with edible oils such as palm oil to produce an alternative to the conventional diesel fuels, which conforms to the standards of US ASTM D 6751 and European EN 14214 [32].

Figure 1.

Sources of non-edible oils used in biodiesel production [29, 30, 31].

3.2.1 Advantage of non-edible oils

These include

  1. Could be a replacement for edible oils in biodiesel production [32, 33, 34]

  2. Contain toxic materials, which make them unsuitable for human consumption [35]

  3. Naturally available [33]

  4. Inexpensive as they are planted in wastelands and no intensive care needed [22].

3.2.2 Examples of non-edible oil used for biodiesel production

Some of these non-edible oils are discussed in more detail below:

3.2.2.1 Waste cooking oil (WCO)

This oil maybe yellow or brown grease obtained from palm, canola, corn, sunflower and other edible oils. Usually, it is ubiquitous and inexpensive, making it ideal for biodiesel production. In recent times, some researchers have demonstrated that biodiesel can be produced from WCO by pyrolysis and transesterification methods. The latter method is preferred due its low cost and simplicity [1]. The performance of the process is usually measured in terms of yield and it depends on factors such as catalyst, catalyst loading, temperature, time and methanol-to-oil molar ratio (see Table 4).

S/N Catalyst Reaction condition Biodiesel yield (%) References
1. Calcined chicken manure Catalyst loading 7.5wt%, temperature 65°C and methanol-to-oil molar ratio 1:15 90 [36]
2. Chicken manure biochar Temperature 350°C 95 [37]
3. CsPW-CB Catalyst loading 2 wt%, methanol-to-oil molar ratio 11:1, temperature 70°C and time 2.5 h 95.1 [38]
4. KOH Catalyst loading 1 wt%, methanol-to-oil molar ratio 1:3, temperature 60°C and time 0.8 h 94 [39, 40]
5. Titanium iso-propoxide (TiO2) + graphene oxide (GO) Catalyst loading 1.5 wt%, methanol-to-oil molar ratio 1:12, temperature 65°C and time 3 h 98 [41]
6. Calcium diglyceroxide Catalyst loading 1.03 wt%, methanol-to-oil molar ratio 7.46:1, temperature 62°C and time 0.4 h 94.86 [40]
7. KOH Catalyst loading 1.5 wt%, methanol-to-oil molar ratio 7:1, temperature 60°C and time 1.5 h 92 [42]
8. KOH Catalyst loading 1.16 wt%, methanol-to-oil molar ratio 9.4:1, temperature 62.4°C and time 2 h 98.26 [43]
9. CaO/MgO Catalyst loading 6 wt%, methanol-to-oil molar ratio 1:15, temperature 90°C and time 2 h 96.47 [44]
10. CaO Catalyst loading 5 wt%, methanol-to-oil molar ratio 20:1, temperature 65°C and time 4 h 96.74 [45]
11. BaSnO3 Catalyst loading 6 wt%, methanol-to-oil molar ratio 10:1, temperature 90°C and time 2 h 96 [46]
12. Sulphamic acid Catalyst loading 1 wt%, methanol-to-oil molar ratio 10:1, temperature 110°C and time 2 h 95.6 [47]
13. Fusion waste chicken and fish bones Catalyst loading 1.98 wt%, methanol-to-oil molar ratio 10:1, temperature 65°C and time 1.5 h 89.5 [48]
14. Biomass fly ash Catalyst loading 10 wt%, methanol-to-oil molar ratio 9:1, temperature 60°C and time 3 h 95 [49]
15. Kettle limescale Catalyst loading 8.87 wt%, methanol-to-oil molar ratio 1.7:3, temperature 61.7°C and time 0.25 h 93.41 [43]
16. Calcium oxide (CaO) nano-catalyst Catalyst loading 1 wt%, methanol-to-oil molar ratio 8:1, temperature 50°C, time 1.5 h and particle size 29 nm 96 [50]

Table 4.

Dependence of the yield of biodiesel from WCO on reaction parameters.

3.2.2.2 Algae

The use of algae in biofuel production is gaining traction globally, especially as it is considered to be safer, non-competitive and made up of microorganisms with precocious growth. These organisms are aquatic and may be unicellular or multicellular with over 300,000 species. The number is greater than plant species and the organisms exhibit varying compositions, but are faced with higher cost of production. Also, more complexity of processes and technology are required for cultivation compared to plants [51]. Algae grow naturally in open ponds and can be cultivated through tubular photobioreactors. The former is the oldest method involving a simple and inexpensive process, compared to the latter, which enjoys high productivity rate, less maturity time and the capacity to selectively produced high lipid content using desirable algae species. Algae contain lipids, carbohydrates and complex oil depending on their species [52, 53, 54, 55]. The lipid content ranges from 20 to 80% depending on the various species. Some species such as Tribonema, Ulothrix and Euglena are considered to possess high lipid content and have great potential for biodiesel and kerosene production [56]. Efforts are being made to increase the lipid content of algae by modifying the algae genome in charge of nitrogen assimilation. This process could double the lipid content thereby increasing the potential of the commercial production of biodiesel. However, the production of algae is independent of season and it is characterized by an exponential growth rate with capacity to double their biomass in about 3.5h [51]. This ensures a relative abundance of algae on earth surface.

Algae are not edible and using them as a feedstock for biodiesel production poses no threat to food production. They have the capability to convert carbon dioxide to biofuels and oleochemical products [51]. The remaining biomass can be converted into useful chemicals to generate more revenue to ameliorate the high economic cost of the process.

3.2.2.3 Biodiesel production with algae as feedstock

Due to high lipid content and availability, several investigators have explored the potential of algae as a feedstock for biofuel production. This usually begins by selecting algae species with high lipid yield and very good fatty acid composition as shown in Figure 2. The desirable algae species for the production of biodiesel is usually selected based on growth rate, degree of survival and physicochemical properties and fatty acid composition.

Figure 2.

Processes involved in applying algae as feedstock for the production of biodiesel at a small scale or experimental level [56].

The typical properties of biodiesel algae oil compared with standards and biodiesel from other sources are presented in Table 5. Applying such biodiesel in an internal combustion engine usually consumes more fuel and has less thermal efficiency than petroleum diesel. This may be due to its physicochemical properties such as higher density and viscosity, lower calorific value and cetane number. The effect of this problem can be minimized by blending it with petroleum diesel (up to 30%) [57, 58, 59, 60, 61, 62, 63, 64, 65]. The presence of excess oxygen molecule in the algae biofuel ensures that complete combustion is attained, thereby eliminating the emission of undesirable substances such as hydrocarbons and carbon monoxide. But, NOx emission like biodiesel from other sources is high and can be reduced by the addition of n-butanol to the blends [63].

S/N Properties ASTM
6751-12
EN
14,214
Diesel Algae oil Palm oil Jatropha Karanja
1 Calorific value
(kJ/kg)
43,000 40,072 37,800 39,000 39,200
2 Density (kg/L) 086–0.90 0.86–0.90 0.84 0.912 0.850 0.940 0.874
3 Viscosity @ 40°C (mm2/s) 1.9–6.0 3.5–5.0 2.64 5.06 4.32 4.8 5.21
4 Cetane number >47 >51 53.3 46.5 55 50 50
5 Flash point (°C) 100–170 >120 71 145 167 135 100
6 Acid value (mg KOH/g) <0.5 <0.5 0.0 0.14 0.24 0.4 0.43
7 Oxidation stability @ 110°C 3.0 >6.0 6.76 10.3 3.2
8 Oil yield (L/ha) 58,000 5950 1892 2590

Table 5.

Comparison between the fuel properties of algae oil and the petroleum diesel [51].

S/N is serial number.

3.2.2.4 Tea seed oil

This is one of the cheapest vegetable oils with an average price of US $514 per ton. It is composed of predominantly unsaturated fatty acids with lower pour point, making it suitable for biodiesel production [64]. The characteristics of biodiesel from tea seed oil share some resemblances with those from vegetable oil, but it has lower pour point of −5 °C and is less viscous than biodiesel from palm oil, cotton seed oil and peanut oil [1, 64, 66]. Like in algae biodiesel, the application of tea seed oil biodiesel in internal consumption engine requires more fuel consumption and causes high emissions of CO and CO2. To solve these problems, hydrogen is usually added to the petroleum diesel and biodiesel blends, thereby improving the performance characteristics of the engine. This gain is possible since there is absence of carbon atoms in the chemical structure. But, the disadvantage is increased NOx emission [65].

3.2.2.5 Activated sludge

This sludge is a residue from the secondary/biological section of wastewater treatment plant and composed predominantly of microorganisms [67]. It is being investigated as a feedstock for biodiesel production probably due to its availability, lipid content and possibility of obtaining it without any cost implication [5, 68, 69, 70, 71, 72]. The lipid/oil content is relatively low and various researchers have investigated the potential of increasing the yield using different methods to ensure its adaptability as a substrate for biodiesel production. Notable among them are Edeh et al., who using the combination of subcritical water technology and optimization increased the lipid yield from 7.4 (wt./wt.)% to 41.0 (wt./wt.)% [28]. The predominant fatty acid in activated sludge is palmitic acid [27]. Researchers have shown that activated sludge can be used as a feedstock for biodiesel production. But, due to low yield of 3–6 wt%, (dry cell weight), which is below the minimum of 10 wt.% (dry cell weight) required for biodiesel to have an economic advantage over the conventional petroleum diesel, this feedstock is still unattractive [60, 73, 74]. Another problem is variation in the composition of fatty acids, which depends on the source and composition of wastewater and season of collection, which affect the quality and yield of the biodiesel [5, 75].

Other non-edible oils used in the production of biodiesel are presented in Table 6.

S/N Source Characteristics Yield (wt.%) Fatty acid composition References
1. Karanja Grown in Southeast Asia, flowers 3–4 years after planting while matures 4–7 years later, a single tree yields 9–90 kg of seeds 25–40 Oleic (44.5–71.3%), linoleic (10.8–18.3%) and stearic acids (2.4–8.9%) [76, 77, 78, 79]
2. Mahua Grown in Indian forest, produces 20–200 kg of seeds annually per tree depending on maturity, starts to bear seeds after 10 years of planting and continues up to 60 years 35–50 Oleic (41–51%), stearic (20.0–25.1%), palmitic acid (16.0–28.2%) and linoleic acids (8.9–18.3%) [51, 80, 81, 82]
3. Cotton Grown for cotton fiber in China, United States and Europe, the seeds contain non-glycerides such as gossypol, phospholipids, sterols, resins, carbohydrates and related pigments 17–25 Linoleic (55.2–55.5%), palmitic (11.67–20.1%) and oleic acids (19.2–23.26%) [83, 84, 85]
4. Neem Can grow in different kinds of soils such as saline, clay, dry, shallow, alkaline and stony in Asian countries including India, Malaysia and Indonesia. It matures after 15 years and has a life span of 150–200 years 20–30 Linoleic (6–16%), oleic (25–54%) and stearic (9–24%) acids [17, 51, 85, 86, 87]
5. Tobacco Grown in countries such as Turkey, Macedonia and North America for leaf collection 35–49 Linoleic acid (69.49–75.58%) [88, 89, 90, 91]
6. Rubber Forest-based tree largely grown in Malaysia, India, Thailand and Indonesia 50–60 Linoleic (39.6–40.5%), oleic (17–24.6%) and linolenic acid (16.3–26%) [20, 92, 93, 94]
7. Jatropha Grown in arid, semi-arid and tropical regions, such as United States, Brazil, Bolivia and Mexico. Produces seeds after 12 months of planting, attain optimum productivity by 5 years and has a life span of up to 30 years 20–60 Linoleic (31.4–43.2%), oleic acid (34.3–44.7%), stearic (7.1–7.4%) and palmitic (13.6–15.1%) acids [95, 96, 97, 98]

Table 6.

Non-edible oils from the seeds of their respective trees used for biodiesel production.

3.2.3 Fuel properties of biodiesel produced from various non-edible oils

These properties depend on the fatty acid and chemical composition of the non-edible oils. The fuel properties of biodiesel can be measured by using different standards including ASTM D6751 and EN 14214. The most essential properties used in assessing the suitability of biodiesel as fuel include density, flash point, cloud point, pour point, calorific value and cetane point (see Table 7). The standards for measuring each property are presented in Table 8 [99].

Non-edible oil Density at 40°C (kg/m3) Viscosity at 40°C (mm2/s) Flash point (°C) Cloud point (°C) Pour point (°C) Cetane number Calorific value (MJ/kg) References
Karanja (Pongamia pinnata L.) 876–890 4.37–9.60 163–187 13–15 −3 to 5.1 52–58 36–38 [15, 77, 99, 100, 101]
Polanga (Calophyllum inophyllum) 888.6–910 4–5.34 151–170 13.2–14 4.3 57.3 39.25–41.3 [15, 102]
Mahua (Madhuca indica) 904–916 3.98–5.8 127–129 3–5 1–6 51–52 39.4–39.91 [15, 103, 104, 105, 106]
Rubber seed oil (Hevea brasiliensis) 860–881 5.81–5.96 130–140 4–5 –8 37–49 36.5–41.07 [92, 107, 108, 109]
Cotton seed 874–911 4–4.9 210–243 1.7 −10 to −15 41.2–59.5 39.5–40.1 [110, 111, 112]
Jojoba oil (Simmondsia chinensis) 863–866 19.2–25.4 61–75 6–16 −6 to 6 63.5 42.76–47.38 [113, 114, 115, 116]
Tobacco oil (Nicotiana tabacum) 860–888.5 3.5–4.23 152–165.4 −12 49–51.6 38.43–39.81 [89, 90, 117]
Neem (Azadirachta) 912–965 20.5–48.5 34 51 33.7–39.5 [87, 110, 118, 119]
Linseed oil (Linum usitatissimum) 865–950 16.2–36.6 108 1.7 −4 to −18 28–35 37.7–39.8 [110, 120, 121]
Jatropha (Jatropha curcas L.) 864–880 3.7–5.8 163–238 5 46–55 38.5–42 [122, 123]
Diesel 816–840 2.5–5.7 50–98 −10 to −5 −20 to 5 45–55 42–45.9 [124, 125, 126]

Table 7.

Properties of diesel fuel and those of biodiesel produced from non-edible oils.

S/N Property Characteristics Standard References
1. Density Higher than the diesel ASTM D1298 and EN ISO 3675 [127]
2. Kinematic viscosity High viscosity causes poor fuel flow resulting in delayed combustion ASTM D445 and EN ISO 3104 [128]
3. Flash point Measures the flammability hazard of a substance. At flash point, if the source of ignition is removed, vapor ceases to burn ASTM D93 and EN ISO 3697 [127]
4. Cetane number (CN) Measures the ignition quality of fuel in a power diesel engine. Higher CN causes shorter ignition delay. Biodiesel has higher CN due to its longer fatty acid carbon chains ASTM D613 and EN ISO 5165 [127]
5. Cloud point (CP) Higher CP than diesel ASTM D2500 [129, 130]
6. Pour point (PP) Higher PP than diesel ASTM D97 [129, 130]
7. Calorific value (HHV)* Measures the heat content of a fuel. Biodiesel has lower calorific value than diesel due to its higher oxygen content ASTM D2015 [107, 131, 132]

Table 8.

Standards for measuring properties of biodiesel [98].

Advertisement

4. Lipid/oil extraction methods

In most cases oils are extracted from the oil-bearing biomass, for example oil seeds prior to use in biodiesel production. The methods used in oil extraction include solvent extraction, critical fluid extraction, mechanical extraction, enzymatic oil extraction, microwave-assisted extraction (MAE) and ultrasound-assisted extraction (UAE). They are discussed below:

4.1 Solvent extraction

This extraction method utilizes organic solvents extract lipid/oil from the oil-bearing biomass. The organic solvents used include: hexane, chloroform, ethyl ether, petroleum ether, toluene, methanol, ethanol and acetone [5]. The solvents can also be combined together depending on their polarity to achieve higher yield of oil, for instance, chloroform and methanol, hexane and ethanol, dichloromethane and methanol [23, 133]. The properties that influence the selection of a particular solvent for oil extraction are polarity, volatility, non-miscibility with water, safety, boiling point, environmental factors, absence of toxic or reactive impurities, ability to form two phases with for easy separation, capacity to extract a large range of lipid classes and cost of the solvent [134, 135].

The solvent extraction methods used in the laboratory include Soxhlet, Folch, and Bligh and Dyer methods. Soxhlet method is preferred due to the following advantages: it is easy to use, does not require filtration and inexpensive, and it ensures higher oil extraction, supports simultaneous and parallel extraction. Despite these advantages, its demerits include requirement of high volume of solvents, health and environmental risks, long extraction time, potential to thermally degrade sample and difficulty to automate due to selectivity issues [136]. The Soxhlet extraction is influenced by the following factors: temperature, sample preparation, extraction time, high solvent-to-sample ratio, type and the volume of solvent [137].

Soxhlet extraction is carried out by heating the distillation/boiling flask containing the organic solvent to its boiling point (see Figure 3). The vapor produced passes through the tube to the condenser where it is condensed and the liquid formed trickles down to the thimble containing the sample. The soluble part of the sample is dissolved by this liquid and the process continues until the liquid marked is reached. The solubilized sample is aspirated to the distillation/boiling flask and the process continues until the predetermined number of cycle or extraction time is attained [138].

Figure 3.

Soxhlet apparatus.

4.2 Folch method

This method was developed by Folch et al. [139]. The method utilizes a combination of organic solvents: chloroform and methanol in a ratio of 2:1 (v/v) for lipid/oil extraction. It is usually used for extracting and quantifying total lipids [140].

4.3 Bligh and dyer method

This method has some similarities with the Folch method in terms of the solvent system and function. The method uses combined chloroform and methanol in a ratio of 1:2 (v/v) in converse to the Folch method to extract lipid/oil from samples. With this ratio, the Bligh and Dyer method is more economical than the Folch method [140].

4.4 Critical fluid extraction

This involves the use of supercritical or subcritical fluids in oil extraction. These are discussed below:

4.4.1 Supercritical fluids (SCFs)

These are fluids with critical temperature and pressure above their critical points. For example, above the critical point of CO2 (31.1°C and 7.38 MPa) and that of water (374°C and 22.1 MPa), supercritical fluids exist [141, 142]. Supercritical fluids usually have high density, which increases their solubilization, and low viscosity, which enhances their mass transfer rate [143]. SCFs have the advantages of low operating cost; high product quality; ability to combine some operation units into one and to selectively extract certain lipids at different operating conditions of temperature, pressure, and time. The advantage of this is a reduction in cost and extraction time [143, 144]. The disadvantage of SCFs is that they required the use of high-pressure vessels which are usually expensive. A brief discussion on supercritical CO2 and supercritical water is presented below.

4.4.1.1 Supercritical CO2 extraction

This lipid extraction method uses CO2 as the supercritical fluid probably because it is cheap, non-toxic, non-explosive and non-flammable and possesses high purity and low critical temperature [141, 143]. The low critical temperature makes it the most suitable method for the extraction of thermal labile substances such as lipid/oil as the original properties of the materials are protected [143]. Supercritical CO2 is usually used to extract non-polar lipids but due to the introduction of co-solvents such as methanol, ethanol and water, it could recover polar lipids [143]. For instance, Hanif et al. increased the yield of phospholipid fatty acids (PLFAs) from 0.5 to 7.28 nmol/mg using methanol (10%, v/v) as a co-solvent [145].

4.4.1.2 Supercritical water extraction

Water is used as a supercritical fluid here. Supercritical water possesses liquid and gaseous properties including diffusivity, density and heat transfer, which can be manipulated through temperature and pressure to achieve an efficient extraction. For instance, a low-density supercritical water can be used to extract non-polar substances, and due to low dissolution it will not be effective in extracting ionic substances. At high temperature, it can dissolve organic substances, gases and salts due to its decrease in dielectric constant [146]. Supercritical water extraction has been used by Gungoren et al. for oil recovery and products distribution from sewage sludge at temperatures between 350 and 450°C and pressures of between 21.5 and 30 MPa [147].

4.4.2 Subcritical water extraction

Subcritical water as shown in Figure 4 is water at temperatures between its boiling point (Tb), 100°C and its critical point of 374°C with pressure sufficient to maintain water in the liquid state. Within this temperature range, water behaves like organic solvents due to decrease in its dielectric constant. At low temperature, subcritical water can extract both polar and ionic substances while at temperatures close to the critical temperature, extraction of non-polar substances is possible by the interaction with these substances and reduction in the binding forces [148, 149, 150]. Subcritical water has been demonstrated to be useful in decontaminating soil, removing polyhydroxyalkanoates (PAH), hydrocarbons and metals and extracting variety of natural products [151]. It has also been used to increase the lipid yield of activated sludge [34, 35].

Figure 4.

Phase diagram.

4.5 Enzymatic oil extraction

This method uses the right enzymes in extracting oil from the oil-bearing biomass and it is environmentally friendly as there is no emission of volatile organic matter [2]. The disadvantages include: relatively high cost of enzyme production, prolonged incubation periods and requirement of de-emulsification during the downstream processing (DSP) [151, 152]. Some of these problems such as high cost of enzyme production can be minimized using enzyme immobilization, which helps to reduce enzyme losses, although, could reduce reaction rate due to steric hindrance. While others like de-emulsification during DSP can be made easier through the use of affinity chromatography and perfusion chromatography [2].

4.6 Mechanical oil extraction

Oil is extracted using a manual ram press or an engine-driven screw press. With the manual ram press extracting up to 60–65% and engine-driven press recovering 68–80% of the oil content of the feedstocks, respectively. Usually, the oil extract undergoes filtration and degumming as a way of obtaining a more refined oil. Mechanical extraction is inefficient in extracting oil from seeds, which it was not designed for, although, this problem can be solved by using pretreatment methods such as cooking of the seeds and using at least double passes in the screw press. This could give rise to up to 91% yield of oil [2, 151].

4.7 Microwave-assisted extraction (MAE)

This extraction method uses microwave oven in the extraction process. It has been utilized in extracting values from plant materials [153]. The method requires transferring heat through direct contact to the polar solvent and/to the target substance. This can be controlled through ionic conduction and dipole rotation, which occurs simultaneously. Comparing MAE with the conventional extraction method, the latter requires longer time and resources while the former supports high yield of extraction with lesser volume of solvents and controllable heating process [154]. MAE also emits smaller amount of CO2 and consumes lesser quantity of energy compared to the conventional extraction methods. The disadvantages are that the process is accompanied with the presence of solid residue, which limits heat and mass transfer, and the extraction using non-polar solvents or extracting non-polar substances is greatly affected [2, 151].

4.8 Ultrasound-assisted extraction (UAE)

This involves submerging the feedstocks usually of plant origin in a polar solvent (e.g. water) or non-polar solvent (e.g. ethanol) and subjecting the resulting mixture to an ultrasonic vibration. The vibration is made up of sound waves at the range of 18 kHz–100 MHz. This sound wave in the solvent enhances the biomass (flowers, seeds, leaves, etc.) solubilization resulting in the release of values such as oils entrapped within them, thereby increasing yield of the valuable materials. UAE has a very fast extraction rate and high efficiency, but could denature the structure of the extracted substance, for example, oil due to prolonged exposure to ultrasound. Also, it requires the use of large volume of solvent and repetition of the process in order to achieve an efficient extraction. This thus impacts on the operating cost of the entire process [155, 156, 157].

Advertisement

5. Biodiesel production

5.1 Methods of biodiesel production

According to Rezania et al., there are four commonly used methods for biodiesel production [1]. These are explained below:

5.1.1 Pyrolysis

This involves preheating of vegetable oil or animal fat at a temperature of 300–1300°C in the presence of catalyst and absence of oxygen [2]. This may result in product possessing desirable properties such as low viscosity, high cetane number, low amount of sulfur and water content, and standard corrosion values [158].

5.1.2 Microemulsions

These are clear, thermodynamically stable, isotropic liquid mixtures of oil, water, surfactant, mostly in combination of cosurfactant [159]. This method using ethanol has been used with soybean as feedstock to produce biodiesel with similar properties as No. 2 diesel. These properties include cetane number and viscosity [81, 151].

5.1.3 Blending

This is also known as dilution and it is simplest and oldest method used in biodiesel production. It involves the blending of preheated vegetable oil or animal fats with the conventional petroleum diesel in a ratio of 10–40% (w/w) [160].

5.1.4 Transesterification/esterification

This involves the reaction between triglyceride from vegetable oil or animal fat with alcohol usually methanol in the presence of catalyst such as acidic, basic or enzymatic catalyst [161]. When methanol is used, the reaction is called methanolysis while it is called ethanolysis if ethanol is used as the alcohol. The schematic diagram representing the processes involved in biodiesel production via transesterification is shown in Figure 5. Transesterification of triglyceride to biodiesel (alkyl ester) and glycerol as the by-product is illustrated in Figure 6. The reaction mechanism involves the conversion of triglyceride (TG) to diglyceride (DG) followed by monoglyceride (MG) and then to a free glyceride. Each step is catalyzed by alkoxide, for instance methoxide when methanol is used as the alcohol [163]. The reaction mechanism is presented in Figure 7.

Figure 5.

Flowchart of biodiesel (FAME) production through transesterification [156].

Figure 6.

Production of biodiesel through a transesterification of triglyceride [162]. (Where R!, R!! and R!!! are carbon chain of fatty acids and R is the alkyl group of the alcohol, which could be methyl or ethyl when methanol or ethanol is used respectively).

Figure 7.

Reaction mechanism (chain reaction) of the transesterification of triglyceride to biodiesel (fatty acid methyl acid-FAME) [2].

Similarly, esterification as a method of producing biodiesel involves a reaction between fatty acid and alcohol in the presence of catalyst (see Figure 8).

Figure 8.

Esterification of free fatty acid to methyl ester and water [164]. (Where R is the carbon chain of fatty acid and R1 is the alkyl group of the alcohol, which could be methyl assuming that methanol is used as the alcohol).

Both transesterification and esterification can occur simultaneously in the same process. This is most suitable for feedstocks with high free fatty acid and water content. The feedstock is firstly esterified using the acidic catalyst before transesterification by the alkali catalyst [2]. The performance of these reactions is measured in terms of yield.

Transesterification is the most commonly used method in biodiesel production probably due to its simplicity and low cost [165]. It can be carried out in situ using the oil-bearing biomass or ex situ directly with the oil extracted from the biomass-bearing oil. Some researchers have demonstrated the application of in situ transesterification of oil-bearing biomass to biodiesel. For instance, Mondala et al. investigated the production of biodiesel from municipal primary and secondary sludge (activated sludge) through in situ transesterification reaction [154]. On the other hand, numerous works have been conducted using lipid extracted from oil-bearing biomass (ex situ) to produce biodiesel. For example, Siddiquee and Rohani worked on the production of biodiesel via the methanolysis of lipids extracted from the primary and secondary sludge [133].

5.2 Factors affecting the production of biodiesel

These include catalysts type, reactor type, temperature, agitation speed, solvent type, alcohol-to-oil ratio, residence time and nature of feedstock (water content, quantity of free fatty acid and esterifiable substances present in the feedstock) [133]. The catalyst type and nature of the feedstock are the most influential factors as they determine the cost of the production of biodiesel [11]. High free fatty acid and water content can cause low yield of biodiesel production due to soap formation via saponification reaction [166, 167].

Advertisement

6. Economic aspects

Researchers have posited that biodiesel is currently not competitive with the conventional petroleum diesel due to higher production cost despite numerous advantages [168]. This can be influenced by the type of raw materials, selling price of the by-product, labour and operation cost, catalyst and the reaction type [1]. The average production cost for biodiesel and diesel fuel is $0.50 and $0.35 per liter, respectively [169]. The price for producing biodiesel can be estimated using Eq. (1)

Production price for biodiesel = Operating costs $ / yr Byproduct credit $ / yr Product yr kg / yr E1

The cost of biodiesel production can be reduced by increasing yield using improved technologies, reducing capital investment cost and reducing the raw materials cost [168, 170, 171, 172, 173].

6.1 Factors that influence the cost of biodiesel production

6.1.1 Alternative raw materials

This involves the use of cheaper feedstocks including wastes from oils, fats and non-edible crops in order to reduce the unit cost of producing biodiesel [28, 174]. The major drawbacks to using these feedstocks are high free fatty acid (FFA) and water content with the capacity to reduce the yield and quality of the biodiesel [9, 12, 22, 175]. The effect of this can be reduced by using multiple chemical processes with the tendency to increase the overall production cost [176]. For instance, using alkali to catalyze the transesterification reaction may require feedstock pretreatment, product separation and purification, thereby rendering the entire process uneconomical due to additional cost incurred [177]. However, acid catalysts are most suitable for the conversion of WCO with high FFA and water content to biodiesel. But, the disadvantages of this are that the reaction is very slow, requires more alcohol and large volume of reactor, and the acid used may corrode equipment, causing them to break down [178]. The use of acid catalyst may also increase the production cost. Some of these problems may be solved using supercritical fluid. The process does not need catalyst, it is faster and may require large volume of alcohol, high temperature and pressure giving rise to a considerable cost implication [179, 180].

The use of cheap and low-cost feedstock may affect the quality of the biodiesel, although, this can be improved. For example, poor cold properties can be improved using additives, although not without some cost implications. Despite the potential of cheap and low-cost feedstock to reduce the production cost of biodiesel, due to high level of impurity, it may require pretreatment prior to use, product purification due to poor quality and, thus, have some cost implications.

6.1.2 Effects of technologies

Technologies used in biodiesel production to a large extent impact the cost of production. Some of these technologies require more unit operations than the other, which influences energy utilization and number of equipment [181]. For instance, the use of catalytic distillation (CD) process is more economical than conventional reactor as capital and production costs are reduced. This is possible due to reduction in the number of equipment, for example plug flow reactor and flash separation units, which are essential when using the conventional reactors are not needed [182].

Alkali catalyst technologies are preferred for producing biodiesel, especially heterogeneous catalyst technology using neat vegetable oil. The reason being that it requires less unit operation and number of equipment. It is also faster and cheaper and can easily be recovered. An example of such catalysts is calcium oxide [181, 183, 184, 185]. For use with high free fatty acid and water content feedstocks, alkali catalyst will cause such problems such as soap formation, which reduces the yield of biodiesel (Figure 9). The soap can gel at room temperature causing the production of semisolid mass instead of biodiesel and can cause difficulty in purifying glycerol [186]. Thus, when considering waste oils such as waste cooking oil with high free fatty acid and water content, acid catalyst technologies are the best option with the aim of reducing the overall production cost. The cost can be reduced because acid catalysts are less corrosive, easy to separate, can be reused and do not require additional washing steps. This will help to produce high-quality products in terms of biodiesel and glycerols [187, 188].

Figure 9.

Soap formation during the transesterification of triglyceride to produce biodiesel.

Alternatively, enzyme and supercritical technologies can be used to process feedstock with high free fatty acid and water content to biodiesel, although they are more expensive than acid-catalysed technologies [173, 189]. Enzyme-catalysed transesterification is a slow process and takes longer time, and the soluble enzymes are not reusable except if immobilized enzyme is used. These disadvantages impact negatively on the cost of production [190]. On the other hand, supercritical technologies do not require the use of catalyst and encourage the production of by-product glycerol with high purity [192].

Generally, technologies like feedstock and catalyst influence the overall cost of biodiesel production (Table 9).

Production technology type Capacity Feedstock Production cost $/ton References
KOH-catalyzed transesterification with methanol 8000 ton/yr Waste cooking oil 868,60 [173]
H2SO4-catalyzed transesterification with methanol Waste cooking oil 750,38
Lipase (Novozym-435) -catalyzed transesterification Waste cooking oil 1047,97
Alkali catalyst process Batch process with a production capacity of 1000 tons Palm oil 1166,67 [191]
Soluble lipase catalyst process Palm oil 7821,37
Immobilzed lipase catalyst process Palm oil 2414,63
Homogeneous H2SO4-catalyzed and using purchased feedstock Continuous reactor operating at 30°C Microalgae oil 620 [182]
Homogeneous H2SO4-catalyzed and using self-produced feedstock from recycled glycerol Microalgae oil 580
Homogeneous KOH catalyst and hot water purification process Batch process with a production capacity of 1452 Waste cooking oil 921 [193]
Homogeneous KOH catalyst and vacuum FAME distillation process Waste cooking oil 984
Heterogeneous CaO catalyst and hot water purification process Waste cooking oil 911
Heterogeneous CaO catalyst and vacuum FAME distillation process Waste cooking oil 969
Homogeneous KOH catalyst and hot water purification process Batch mode with a production capacity of 7260 tons/year Waste cooking oil 598 [193]
Homogeneous KOH catalyst and vacuum FAME distillation process Waste cooking oil 641
Heterogeneous CaO catalyst and hot water purification process Waste cooking oil 584
Heterogeneous CaO catalyst and vacuum FAME distillation process Waste cooking oil 622

Table 9.

Dependence of biodiesel production cost on technologies [168].

6.1.3 Effect of alternative catalysts

The effect of alternative catalysts in the production of biodiesel can been seen in the reduction of production cost as supported by some of their characteristics such as being inexpensive, reusability and high catalytic potential. Examples of such catalysts are obtained from shells from egg, coconut, mussel, scallop and crustacean [183, 190, 194, 195, 196].

Generally, catalysts used in catalysing the transesterification reaction leading to the production of biodiesel may be either homogeneous or heterogeneous. The choice of which to use is dependent on the free fatty acid and water content composition of the feedstock. Usually, heterogeneous catalysts unlike homogeneous catalysts are used to catalyse reactions involving feedstock with high free fatty acid and water content as they can be reused, require less products separation and purification steps, and possess the capacity to enable the production of pure by-products such as glycerol. Although, these advantages have some cost implications, heterogenous catalysts remain the best choice for biodiesel production unit cost reduction [168, 187, 197].

6.2 Profitability of biodiesel production

This is a measure of the amount of profit that can be obtained from an investment in biodiesel production. The profit is usually calculated from the difference between the income obtained from the sales of the products and the expenses incurred. Profitability of biodiesel can be determined using such economic parameters as net present value, break-even price of biodiesel, after-tax internal rate of return, gross margin [168].

6.2.1 Factors affecting the profitability of biodiesel production

6.2.1.1 Market variables

These include income variables such as biodiesel and glycerol and outcome variables, which are feedstock, catalyst, alcohol and washing water. Studies have shown that the major market variable that influences the profitability of biodiesel production is the cost of feedstock due to large quantity required, and others are selling price of biodiesel and glycerol, while outcome variables such as catalyst and washing water have less effect because less quantities are required [162, 198, 199].

6.2.1.2 Production scale

This is another factor affecting the profitability of biodiesel production. The higher the production scale, the lower the unit production cost of biodiesel, see Figure 10. This can be seen from the work of Van Kasteren et al. who compared three biodiesel production processes via supercritical method [201]. The results obtained show increase in profitability of biodiesel at high production scale compared to low production scale. The result was corroborated by the study conducted by You et al. on the effect of production scales 8000, 30,000 and 100,000 annually on the feasibility of biodiesel production from food grade soybean oil using NaOH-catalyzed transesterification [202]. This conclusion was reached as production scale of 100,000 gave higher net annual profit after taxes (NNP) and after-tax of return (ARR), and lower biodiesel break-even price (BBP) compared to other production scales.

Figure 10.

Effect of plant capacity/production scale on unit production cost [200].

Advertisement

7. Conclusions

Commercial quantity of biodiesel is currently being produced from edible vegetable oils with the global production capacity envisaged to reach 12 billion gallons by 2020 and countries such as Brazil, United States of America, Malaysia, Argentina, Netherlands, Spain, Philippines, Belgium, Indonesia and Germany meeting more than 80% of the world demand. The problem with this type of biodiesel includes poor storage, oxidation stability, high feedstock cost, low heating value and higher NOx emission. The implication of these is that biodiesel is not competitive with the conventional petroleum diesel.

Researchers have suggested the utilization of non-edible oils as a way of minimizing cost since feedstocks consume up to 80% in biodiesel production. But, the problem with this is the presence of high free fatty acid (FFA) and water content, which reduces the yield and quality of the biodiesel. This can be reduced through the use of multiple chemical processes, although there is a tendency to increase the overall production cost.

Generally, the cost of biodiesel production is influenced by factors such as raw materials, technologies and catalyst. The raw material and catalyst cost can be reduced using alternatives to these factors while improved technologies could help to minimize the production cost.

The profitability of biodiesel can be determined using economic parameters such as net present value, break-even price of biodiesel, after tax internal rate of returns, and gross margin. These parameters are influenced by market variables and production scale.

Advertisement

Nomenclature

IPCC

Intergovernmental Panel on Climate Change

ASTM

American Society of Testing Materials

WCO

waste cooking oil

CN

cetane number

CP

cloud point

PP

pour point

HHV

high heating value

MAE

microwave-assisted extraction

UAE

ultrasound-assisted extraction

SCF

supercritical fluid

PLFA

phospholipid fatty acids

PAH

polyhydroxyalkanoate

Tb

boiling point

Tc

critical temperature

Pc

critical pressure

DSP

downstream processing

FAME

fatty acid methyl ester

TG

triglyceride

DG

diglyceride

MG

monoglyceride

FFA

free fatty acid

CD

catalytic distillation

NNP

net annual profit after taxes

ARR

after-tax of return

BBP

biodiesel break-even price

LCB

lignocellulosic biomass

References

  1. 1. Rezania S, Oryani B, Park J, Hashemi B, Yadav KK, Kwon EE, et al. Review on transesterification of non-edible sources for biodiesel production with a focus on economic aspects, fuel properties and by-product applications. Energy Conversion and Management. 2019;201:112155. DOI: 10.1016/j.enconman.2019.112155
  2. 2. Karmakar B, Halder G. Progress and future of biodiesel synthesis: Advancements in oil extraction and conversion technologies. Energy Conversion and Management. 2019;182:307-339
  3. 3. Parry M. Millions at Risk. Norwich: School of Environmental Sciences, University of East Anglia; 2001
  4. 4. Gullison RE, Frumhoff PC, Canadell JG, Field CB, Nepstad DC, Hayhoe K, et al. Tropical forests and climate policy. Science. 2007;316:985-986
  5. 5. Edeh I. A Critical Evaluation of the Utility of Subcritical water to Support the Production of Biodiesel and Renewable diesel from the Lipid fraction of Activated sludge [thesis]. University of Birmingham; 2016
  6. 6. ASTM (American Society for Testing and Materials), D 6751-09 – Standard Specification for Biodiesel Fuel (B100) Blend Stock for Distillate Fuels, in “Annual Book of ASTM Standards”, V. 05.04, ASTM International, West Conshohocken, PA; 2009
  7. 7. Wang WC, Thapaliya N, Campos A, et al. Hydrocarbon fuels from vegetable oils via hydrolysis and thermo-catalytic decarboxylation. Fuel. 2012;95:622-629
  8. 8. Ambat I, Srivastava V, Sillanpää M. Recent advancement in biodiesel production methodologies using various feedstock: A review. Renewable and Sustainable Energy Reviews. 2018;90:356-369
  9. 9. Balat M. Potential alternatives to edible oils for biodiesel production–A review of current work. Energy Conversion and Management. 2011;52:1479-1492
  10. 10. Živković SB, Veljković MV, Banković-Ilić IB, Krstić IM, Konstantinović SS, Ilić SB, et al. Technological, technical, economic, environmental, social, human health risk, toxicological and policy considerations of biodiesel production and use. Renewable and Sustainable Energy Reviews. 2017;79:222-247
  11. 11. Baskar G, Aiswarya R. Trends in catalytic production of biodiesel from various feedstocks. Renewable and Sustainable Energy Reviews. 2016;57:496-504
  12. 12. Gui MM, Lee KT, Bhatia S. Feasibility of edible oil vs. non-edible oil vs. waste edible oil as biodiesel feedstock. Energy. 2008;33:1646-1653
  13. 13. Leung DYC, Wu X, Leung MKH. A review on biodiesel production using catalyzed transesterification. Applied Energy. 2010;87:1083-1095
  14. 14. Pinzi S, Dorado MP. Feedstock for advanced biodiesel production. In: Luque A, Melero JA, editors. Advances in Biodiesel Production. United Kingdom: Woodhead Publishing Series in Energy; 2012. pp. 69-90
  15. 15. Pinzi S, Garcia IL, Gimenez FJL, Castro MDL, Dorado G, Dorado MP. The ideal vegetable oil-based biodiesel composition: A review of social, economical and technical implications. Energy & Fuels. 2009;23:2325-2341
  16. 16. Yousuf A. Biodiesel from lignocellulosic biomass-prospects and challenges. Waste Management. 2012;32:2061-2067
  17. 17. Hossain AK, Badr O. Prospects of renewable energy utilization for electricity generation in Bangladesh. Renewable and Sustainable Energy Reviews. 2007;11:1617-1649
  18. 18. Zuccaro G, Pirozzi D, Yousuf A. Lignocellulosic biomass to biodiesel. In: Yousuf A, Pirozzi D, Sannino F, editors. Lignocellulosic Biomass to Liquid Biofuels. United Kingdom: Academic Press; 2020. pp. 127-167
  19. 19. Schmetz E, Ackiewicz M, Tomlinson G, White C, Gray D. Increasing Security and Reducing Carbon Emissions of the U.S. Transportation Sector: A Transformational Role for Coal with Biomass. United States: National Energy Technology Laboratory; 2007
  20. 20. Abbaszaadeh A, Ghobadian B, Omidkhah MR, Najafi G. Current biodiesel production technologies: A comparative review. Energy Conversion and Management. 2012;63:138-148
  21. 21. Korbitz W. Biodiesel production in Europe and North America: An encouraging Prospect. Renewable Energy. 1999;16:1078-1083
  22. 22. Banković-Ilić IB, Stamenković OS, Veljković VB. Biodiesel production from nonedible plant oils. Renewable and Sustainable Energy Reviews. 2012;16:3621-3647
  23. 23. Agarwal AK, Rajamanoharan K. Biofuels (alcohols and biodiesel) applications as fuels for internal combustion engines. Progress in Energy and Combustion Science. 2007;33:233-271
  24. 24. Moser BR. Biodiesel production, properties, and feedstocks. In: Tomes D, Lakshmanan P, Songstad D, editors. Biofuels. New York, NY: Springer; 2011. pp. 285-347
  25. 25. Kibazohi O, Sangwan RS. Vegetable oil production potential from Jatropha curcas, croton megalocarpus, Aleurites moluccana, Moringa oleifera and Pachira glabra: Assessment of renewable energy resources for bio-energy production in Africa. Biomass and Bioenergy. 2011;35:1352-1356
  26. 26. Edeh I. Activated Sludge for Bioenergy and Oleochemical Production. Germany: LAP LAMBERT Academic publishing; 2019. p. 304
  27. 27. Edeh I, Overton T, Bowra S. Evaluation of the efficacy of subcritical water to enhance the lipid fraction from activated sludge for biodiesel and oleochemical production. Journal of Food Process Engineering. 2019a;42:1-9. DOI: 10.1111/jfpe.13070
  28. 28. Edeh I, Overton T, Bowra S. Optimization of subcritical water-mediated lipid extraction from activated sludge for biodiesel production. Biofuels. 2019b. DOI: 10.1080/17597269.2018.1558839
  29. 29. Kumar A, Sharma S. Potential non-edible oil resources as biodiesel feedstock: An Indian perspective. Renewable and Sustainable Energy Reviews. 2011;15:1791-1800
  30. 30. Chhetri AB, Tango MS, Budge SM, Watts KC, Islam MR. Non-edible plant oils as new sources for biodiesel production. International Journal of Molecular Sciences. 2008;9:169-180
  31. 31. Atabani AE, Silitonga AS, Badruddin IA, Mahlia TMI, Masjuki HH, Mekhilef S. A comprehensive review on biodiesel as an alternative energy resource and its characteristics. Renewable and Sustainable Energy Reviews. 2012;16:2070-2093
  32. 32. Silitonga A, Masjuki H, Mahlia T, Ong H, Chong W, Boosroh M. Overview properties of biodiesel diesel blends from edible and non-edible feedstock. Renewable and Sustainable Energy Reviews. 2013;22:346-360
  33. 33. Syers JK, Wood D, Thongbai P. The Proceedings of the International Technical Workshop on the Feasibility of Non-edible Oil Seed Crops for Biofuel Production. Chiang Rai, Thailand: Mae Fah Luang University; 2007
  34. 34. Vedharaj S, Vallinayagam R, Yang W, Chou S, Chua K, Lee P. Experimental investigation of kapok Ceiba pentandra oil biodiesel as an alternate fuel for diesel engine. Energy Conversion and Management. 2013;75:773-779
  35. 35. Ahmad A, Yasin N, Derek C, Lim J. Microalgae as a sustainable energy source for biodiesel production: A review. Renewable and Sustainable Energy Reviews. 2011;15:584-593
  36. 36. Maneerung T, Kawi S, Dai Y, Wang C-H. Sustainable biodiesel production via transesterification of waste cooking oil by using CaO catalysts prepared from chicken manure. Energy Conversion and Management. 2016;123:487-497
  37. 37. Jung J-M, Oh J-I, Baek K, Lee J, Kwon EE. Biodiesel production from waste cooking oil using biochar derived from chicken manure as a porous media and catalyst. Energy Conversion and Management. 2018;165:628-633
  38. 38. Li L, Zou C, Zhou L, Lin L. Cucurbituril-protected Cs2. 5H0. 5PW12O40 for optimized biodiesel production from waste cooking oil. Renewable Energy. 2017;107:14-22
  39. 39. Sadaf S, Iqbal J, Ullah I, Bhatti HN, Nouren S, Nisar J, et al. Biodiesel production from waste cooking oil: An efficient technique to convert waste into biodiesel. Sustainable Cities and Society. 2018;41:220-226
  40. 40. Gupta AR, Rathod VK. Calcium diglyceroxide catalyzed biodiesel production from waste cooking oil in the presence of microwave: Optimization and kinetic studies. Renewable Energy. 2018;121:757-767
  41. 41. Borah MJ, Devi A, Saikia RA, Deka D. Biodiesel production from waste cooking oil catalyzed by in-situ decorated TiO2 on reduced graphene oxide nanocomposite. Energy. 2018;158:881-889
  42. 42. Satyanarayana PA, Oleti RK, Uppalapati S, Sridevi V. A comparative study on characterization of used cooking oil and mustard oil for biodiesel production: Engine performance. Materials Today: Proceedings. 2018;5:18187-18201
  43. 43. Aghel B, Mohadesi M, Ansari A, Maleki M. Pilot-scale production of biodiesel from waste cooking oil using kettle limescale as a heterogeneous catalyst. Renewable Energy. 2019;142:207-214
  44. 44. Rabie AM, Shaban M, Abukhadra MR, Hosny R, Ahmed SA, Negm NA. Diatomite supported by CaO/MgO nanocomposite as heterogeneous catalyst for biodiesel production from waste cooking oil. Journal of Molecular Liquids. 2019;279:224-231
  45. 45. Borah MJ, Das A, Das V, Bhuyan N, Deka D. Transesterification of waste cooking oil for biodiesel production catalyzed by Zn substituted waste egg shell derived CaO nanocatalyst. Fuel. 2019;242:345-354
  46. 46. Farooq M, Ramli A, Naeem A, Noman M, Shah LA, Khattak NS, et al. A green route for biodiesel production from waste cooking oil over base heterogeneous catalyst. International Journal of Energy Research. 2019;43. DOI: 10.1002/er.4646
  47. 47. Gao Y, Chen Y, Gu J, Xin Z, Sun S. Butyl-biodiesel production from waste cooking oil: Kinetics, fuel properties and emission performance. Fuel. 2019;236:1489-1495
  48. 48. Tan YH, Abdullah MO, Kansedo J, Mubarak NM, San Chan Y, Nolasco-Hipolito C. Biodiesel production from used cooking oil using green solid catalyst derived from calcined fusion waste chicken and fish bones. Renewable Energy. 2019;139:696-706
  49. 49. Vargas EM, Neves MC, Tarelho LAC, Nunes MI. Solid catalysts obtained from wastes for FAME production using mixtures of refined palm oil and waste cooking oils. Renewable Energy. 2019;136:873-883
  50. 50. Degfie TA, Mamo TT, Mekonnen YS. Optimized biodiesel production from waste cooking oil (CaO) Nano-catalyst. Scientific Reports. 2019;9:18982. DOI: 10.1038/s41598-019-55403-4
  51. 51. Joshi MP, Thipse SS. An evaluation of algae biofuel as the next generation alternative fuel and its effects on engine characteristics: A review. International Journal of Mechanical and Production Engineering Research and Development (IJMPERD). 2019;9:435-440
  52. 52. Demirbas A. Tea seed upgrading facilities and economic assessment of biodiesel production from tea seed oil. Energy Conversion and Management. 2010;51:2595-2599
  53. 53. Kleinova A et al. Biofuels from algae. Procedia Engineering. 2012;42:231-238
  54. 54. Singh J, Gu S. Commercialization potential of microalgae for biofuels production. Renewable and Sustainable Energy Reviews. 2010;14:2596-2610
  55. 55. Schlagermann P et al. Composition of algal oil and its potential as biofuel. Journal of Combustion. 2012;2012:1-14
  56. 56. Khan S, Siddique R, Sajjad W, Nabi G, Hayat KM, Duan P, et al. Biodiesel production from algae to overcome the energy crisis. HAYATI Journal of Biosciences. 2017;24:163-167
  57. 57. Haik Y, Selim M, Abdulrehman T. Combustion of algae oil methyl ester in an indirect injection diesel engine. Energy. 2011;36:1827-1835
  58. 58. Hariram V, Kumar MG. The effect of injection timing on combustion, performance and emission parameters with AOME blends as a fuel for compression ignition engine. Research Journal of Applied Sciences. 2012;7:510-519
  59. 59. Tuccar G, Ozgur T, Yasar A, Aydin K. Experimental investigation of engine performance and emission characteristics of a diesel engine using blends containing microalgae biodiesel, n-butanol and diesel fuel. International Conference on Geological and Environmental Sciences. 2014;73:25-29
  60. 60. Jayaprabakar J et al. Experimental investigation on the performance and emission characteristics of a CI engine with rice bran and microalgae biodiesel blends. Journal of Chemical and Pharmaceutical Sciences. 2015;7:19-22
  61. 61. Velappan R, Sivaprakasam S. Investigation of single cylinder diesel engine using biodiesel from marine algae. International Journal of Innovative Science Engineering and Technology (IJISET). 2014;1:399-403
  62. 62. Rajesh S, Kulkarni BM, Kumarappa S, Shanmukhappa S. Investigations on fuel properties of ternary mixture of ethanol, bio diesel from acid oil and petroleum diesel to evaluate alternate fuel for diesel engine. International Journal of Research in Engineering and Technology. 2014;2:181-188
  63. 63. Umesh K, Pravin V, Rajagopal K. An approach (performance score) for experimental analysis of exhaust manifold of multi-cylinder SI engine to determine optimum geometry for recreational and commercial vehicles. International Journal of Automobile Engineering Research and Development (IJAuERD). 2014;4:2277-4785
  64. 64. Serin H, Ozcanli M, Gokce MK, Tuccar G. Biodiesel production from tea seed (camellia sinensis) oil and its blends with diesel fuel. International Journal of Green Energy. 2013;10:370-377
  65. 65. Serin H, Akar NY. The performance and emissions of a diesel engine fueled with tea seed (camellia sinensis) oil biodiesel-diesel fuel blends. International Journal of Green Energy. 2014;11:292-301
  66. 66. Demirbas A. Oil from tea seed by supercritical fluid extraction. Energy Sources, Part A: Recovery, Utilization, and Environmental Effects. 2009;31:217-222
  67. 67. Edeh I. Evaluation of the lipid composition in activated sludge biomass for biodiesel and oleochemical production by thin layer chromatography (TLC). International Journal of Trend in Research and Development. 2018;5:40-45
  68. 68. Borchardt JA, Cleland JK, Redman GO. Viruses and Trace Contaminants in Water and Wastewater. Ann Arbor: Ann Arbor Science Publ; 1977
  69. 69. Chipasa KB, Mdrzycka K. Characterization of the fate of lipids in activated sludge. Journal of Environmental Sciences. 2008;5:536-542
  70. 70. Go LC, Fortela DLB, Revellame MZ, Chirdon W, Holmes W, Hernandez R. Biodiesel chemical and energy recovered from waste microbial matrices. Current Opinion in Chemical Engineering. 2019;26:65-71. DOI: 10.1016/j.coche.2019.08.005
  71. 71. Revellame E, Hernandez R, French W, Holmes W, Alley E. Biodiesel from activated sludge through in situ transesterification. Journal of Chemical Technology and Biotechnology. 2010;85:614-620. DOI: 10.1002/jctb.2317.16
  72. 72. Oladejo J, Shi K, Luo X, Yang G, Wu T. A review of sludge-to-energy recovery methods. Energies. 2019;12:1-38. DOI: 10.3390/en12010060
  73. 73. Revellame E, Hernandez R, French W, Holmes W, Alley E, Callahan R II. Production of biodiesel from wet activated sludge. Journal of Chemical Technology and Biotechnology. 2011;86:61-68
  74. 74. Fortela DL, Hernandez R, Zappi M, French TW, Bajpai A, Chistoserdov A, et al. Microbial lipid accumulation capability of activated sludge feeding on short chain fatty acids as carbon sources through fed-batch cultivation. Journal of Bioprocessing and Biotechniques. 2016;6:1-9. DOI: 10.4172/2155-9821.1000275
  75. 75. Revellame ED, Hernandez R, French WT, Holmes WE, Forks A, Callahan R II. Lipid-enhancement of activated sludges obtained from conventional activated sludge and oxidation ditch processes. Bioresource Technology. 2013;148:487-493. DOI: 10.1016/j.biortech.2013.08.158
  76. 76. Balat M, Balat H. Progress in biodiesel processing. Applied Energy. 2010;87:1815-1835
  77. 77. Karmee SK, Chadha A. Preparation of biodiesel from crude oil of Pongamia pinnata. Bioresource Technology. 2005;96:1425-1429
  78. 78. Scott PT, Pregelj L, Chen N, Hadler JS, Djordjevic MA, Gresshoff PM. Pongamia pinnata: An untapped resource for the biofuels industry of the future. Bioenergy Research. 2008;1:2-11
  79. 79. Bajpai S, Sahoo PK, Das LM. Feasibility of blending karanja vegetable oil in petro-diesel and utilization in a direct injection diesel engine. Fuel. 2009;88:705-711
  80. 80. Jena PC, Raheman H, Prasanna Kumar G, Machavaram R. Biodiesel production from mixture of mahua and simarouba oils with high free fatty acids. Biomass and Bioenergy. 2010;34:1108-1116
  81. 81. Singh SP, Singh D. Biodiesel production through the use of different sources and characterization of oils and their esters as the substitute of diesel: A review. Renewable and Sustainable Energy Reviews. 2010;14:200-216
  82. 82. Puhan S, Nagarajan G, Vedaraman N, Ramabramhmam B. Mahua oil (Madhuca indica oil) derivatives as a renewable fuel for diesel engine systems in India: A performance and emissions comparative study. International Journal of Green Energy. 2007;4:89-104
  83. 83. Fontaras G, Tzamkiozis T, Hatziemmanouil E, Samaras Z. Experimental study on the potential application of cottonseed oil–diesel blends as fuels for automotive diesel engines. Process Safety and Environment Protection. 2007;85:396-403
  84. 84. Nabi MN, Rahman MM, Akhter MS. Biodiesel from cotton seed oil and its effect on engine performance and exhaust emissions. Applied Thermal Engineering. 2009;29:2265-2270
  85. 85. Ilkilic C, Yucesu H. The use of cottonseed oil methyl ester on a diesel engine. Energy Sources, Part A: Recovery, Utilization, and Environmental Effects. 2008;30:742-753
  86. 86. Sanford SD, White JM, Shah PS, Wee C, Valverde MA, Meier GR. Feedstock and biodiesel characteristics report. Renewable Energy Group. 2009;416
  87. 87. Ali MH, Mashud M, Rubel MR, Ahmad RH. Biodiesel from neem oil as an alternative fuel for diesel engine. Process Engineering. 2013;56:625-630
  88. 88. Bryan RM. Biodiesel production, properties and feedstocks. In Vitro Cellular & Developmental Biology – Plant. 2009;45:229-266
  89. 89. Usta N. An experimental study on performance and exhaust emissions of a diesel engine fuelled with tobacco seed oil methyl ester. Energy Conversion and Management. 2005;46:2373-2386
  90. 90. Usta N, Aydogan B, Con A, Uguzdogan E, Ozkal S. Properties and quality verification of biodiesel produced from tobacco seed oil. Energy Conversion and Management. 2011;52:2031-2039
  91. 91. Giannelos P, Zannikos F, Stournas S, Lois E, Anastopoulos G. Tobacco seed oil as an alternative diesel fuel: Physical and chemical properties. Industrial Crops and Products. 2002;16:1-9
  92. 92. Ahmad J, Yusup S, Bokhari A, Kamil RNM. Study of fuel properties of rubber seed oil-based biodiesel. Energy Conversion and Management. 2014;78:266-275
  93. 93. Gimbun J, Ali S, Kanwal CCSC, Shah LA, Ghazali NHM, Cheng CK, et al. Biodiesel production from rubber seed oil using activated cement clinker as catalyst. Process Engineering. 2013;53:13-19
  94. 94. Gill P, Soni SK, Kundu K, Srivastava S. Effect of blends of rubber seed oil on engine performance and emissions. Fuel. 2011;88:738-743
  95. 95. Mohibbe Azam M, Waris A, Nahar NM. Prospects and potential of fatty acid methyl esters of some non-traditional seed oils for use as biodiesel in India. Biomass and Bioenergy. 2005;29:293-302
  96. 96. Achten WMJ, Verchot L, Franken YJ, Mathijs E, Singh VP, Aerts R, et al. Jatropha bio-diesel production and use. Biomass and Bioenergy. 2008;32:1063-1084
  97. 97. Manickam M, Kadambamattam M, Abraham M. Combustion characteristics and optimization of neat biodiesel on high speed common rail diesel engine powered SUV. SAE Technical paper 2009-01-2786; 2009
  98. 98. Ashraful AM, Masjuki HH, Kalam MA, Rizwanul Fattah IM, Imtenan S, Shahir SA, et al. Production and comparison of fuel properties, engine performance and emission characteristics of biodiesel from various non-edible vegetable oils: A review. Energy Conversion and Management. 2014;80:202-228
  99. 99. Sahoo P, Das L. Process optimization for biodiesel production from Jatropha, Karanja and Polanga oils. Fuel. 2009;88:1588-1594
  100. 100. Sureshkumar K, Velraj R, Ganesan R. Performance and exhaust emission characteristics of a CI engine fueled with Pongamia pinnata methyl ester (PPME) and its blends with diesel. Renewable Energy. 2008;33:2294-2302
  101. 101. Raheman H, Phadatare A. Diesel engine emissions and performance from blends of karanja methyl ester and diesel. Biomass and Bioenergy. 2004;27:393-397
  102. 102. Hegde AK, Rao KS. Performance and emission study of 4S CI engine using calophyllum inophyllum biodiesel with additives. International Journal on Theoretical and Applied Research in Mechanical Engineering. 2012;1:2319-3182
  103. 103. Saravanan N, Nagarajan G, Puhan S. Experimental investigation on a DI diesel engine fuelled with Madhuca indica ester and diesel blend. Biomass and Bioenergy. 2010;34:838-843
  104. 104. Puhan S, Vedaraman N, Ram BV, Sankarnarayanan G, Jeychandran K. Mahua oil (Madhuca indica seed oil) methyl ester as biodiesel-preparation and emission characteristics. Biomass and Bioenergy. 2005;28:87-89
  105. 105. Raheman H, Ghadge SV. Performance of compression ignition engine with mahua (Madhuca indica) biodiesel. Fuel. 2007;86:2568-2573
  106. 106. Godiganur S, Suryanarayana Murthy CH, Reddy RP. 6BTA 5.9 G2-1 Cummins engine performance and emission tests using methyl ester mahua (Madhuca indica) oil/diesel blends. Renewable Energy. 2009;34:2172-2177
  107. 107. Ramadhas AS, Jayaraj S, Muraleedharan C. Biodiesel production from high FFA rubber seed oil. Fuel. 2005;84:335-340
  108. 108. Ikwuagwu O, Ononogbu I, Njoku O. Production of biodiesel using rubber [Hevea brasiliensis (Kunth. Muell.)] seed oil. Industrial Crops and Products. 2000;12:57-62
  109. 109. Edwin Geo V, Nagarajan G, Nagalingam B. Studies on dual fuel operation of rubber seed oil and its bio-diesel with hydrogen as the inducted fuel. International Journal of Hydrogen Energy. 2008;33:6357-6367
  110. 110. No SY. Inedible vegetable oils and their derivatives for alternative diesel fuels in CI engines: A review. Renewable and Sustainable Energy Reviews. 2011;15:131-149
  111. 111. Rakopoulos C, Antonopoulos K, Rakopoulos D, Hountalas D, Giakoumis E. Comparative performance and emissions study of a direct injection diesel engine using blends of diesel fuel with vegetable oils or bio-diesels of various origins. Energy Conversion and Management. 2006;47:3272-3287
  112. 112. Altin R, Cetinkaya S, Yucesu H. The potential of using vegetable oil fuels as fuel for diesel engines. Energy Conversion and Management. 2001;42:529-538
  113. 113. Huzayyin A, Bawady A, Rady M, Dawood A. Experimental evaluation of diesel engine performance and emission using blends of jojoba oil and diesel fuel. Energy Conversion and Management. 2004;45:2093-2112
  114. 114. Shehata M, Razek S. Experimental investigation of diesel engine performance and emission characteristics using jojoba/diesel blend and sunflower oil. Fuel. 2011;90:886-897
  115. 115. Panwar NL, Shrirame HS, Rathore NS, Jindal S, Kurchania AK. Performance evaluation of a diesel engine fueled with methyl ester of castor seed oil. Applied Thermal Engineering. 2010;30:245-249
  116. 116. Ghadge SV, Raheman H. Process optimization for biodiesel production from mahua (Madhuca indica) oil using response surface methodology. Bioresource Technology. 2006;97:379-384
  117. 117. Veljkovic V, Lakicevic S, Stamenkovic O, Todorovic Z, Lazic M. Biodiesel production from tobacco (Nicotiana tabacum L.) seed oil with a high content of free fatty acids. Fuel. 2006;85:2671-2675
  118. 118. Prasad L, Agrawal A. Experimental investigation of performance of diesel engine working on diesel and neem oil blends. Carbon. 2012;86:78-92
  119. 119. Anbumani K, Singh AP. Performance of mustard and neem oil blends with diesel fuel in CI engine. Carbon. 2006;86:78-92
  120. 120. Puhan S, Saravanan N, Nagarajan G, Vedaraman N. Effect of biodiesel unsaturated fatty acid on combustion characteristics of a DI compression ignition engine. Biomass and Bioenergy. 2010;34:1079-1088
  121. 121. Puhan S, Jegan R, Balasubbramanian K, Nagarajan G. Effect of injection pressure on performance, emission and combustion characteristics of high linolenic linseed oil methyl ester in a DI diesel engine. Renewable Energy. 2009;34:1227-1233
  122. 122. Ganapathy T, Gakkhar RP, Murugesan K. Influence of injection timing on performance, combustion and emission characteristics of Jatropha biodiesel engine. Applied Energy. 2011;88:4376-4386
  123. 123. Chauhan BS, Kumar N, Jun YD, Lee KB. Performance and emission study of preheated Jatropha oil on medium capacity diesel engine. Energy. 2010;35:2484-2492
  124. 124. Silitonga A, Masjuki H, Mahlia T, Ong HC, Chong W. Experimental study on performance and exhaust emissions of a diesel engine fuelled with Ceiba pentandra biodiesel blends. Energy Conversion and Management. 2013;76:828-836
  125. 125. Palash S, Masjuki H, Kalam M, Masum B, Sanjid A, Abedin M. State of the art of NOx mitigation technologies and their effect on the performance and emission characteristics of biodiesel-fueled compression ignition engines. Energy Conversion and Management. 2013;76:400-420
  126. 126. Palash S, Kalam M, Masjuki H, Arbab M, Masum B, Sanjid A. Impacts of NOx reducing antioxidant additive on performance and emissions of a multi cylinder diesel engine fueled with Jatropha biodiesel blends. Energy Conversion and Management. 2014;77:577-585
  127. 127. Masjuki HH. Biofuel Engine: A New Challenge. Kuala Lumpur: International & Corporate relation office, University of Malaya; 2010. pp. 1-56
  128. 128. Shahabuddin M, Kalam M, Masjuki H, Bhuiya M, Mofijur M. An experimental investigation into biodiesel stability by means of oxidation and property determination. Energy. 2012;44:616-622
  129. 129. Fernando S, Karra P, Hernandez R, Jha SK. Effect of incompletely converted soybean oil on biodiesel quality. Energy. 2007;32:844-851
  130. 130. Sanford S, White J, Shah P, Wee C, Valverde M, Meier G. Feedstock and biodiesel characteristics report. 2011. Available from: http://www.regfuel.com/pdfs/FeedstockandBiodieselCharacteristicsReport.pdf
  131. 131. Sivaramakrishnan K. Determination of higher heating value of biodiesels. International Journal of Engineering, Science and Technology (IJEST). 2011;3(11):7981-7987
  132. 132. Nwafor O. The effect of elevated fuel inlet temperature on performance of diesel engine running on neat vegetable oil at constant speed conditions. Renewable Energy. 2003;28:171-181
  133. 133. Siddiquee MN, Rohani S. Experimental analysis of lipid extraction and biodiesel production from wastewater sludge. Fuel Processing Technology. 2011;92:2241-2251
  134. 134. Boocock DGB, Konar SK, Leung A, et al. Fuels and chemicals from sewage sludge: 1. The solvent extraction and composition of a lipid from raw sewage sludge. Fuel. 1992;71:1283-1289
  135. 135. Hara A, Radin NS. Lipid extraction of tissues with a low-toxicity solvent. Analytical Biochemistry. 1978;90:420-426
  136. 136. Naude Y, De Beer WHJ, Jooste S, et al. Comparison of supercritical fluid extraction and soxhlet extraction for the determination of DDT, DDD and DDE in sediment. Water SA. 1998;24:3
  137. 137. Luthria DL. Oil Extraction and Analysis: Critical Issues and Comparative Studies. United States of America: AOCS; 2019. p. 288
  138. 138. Luque de Castro MD, Garcia-Ayuso LE. Soxhlet extraction of solid materials: An outdated technique with a promising innovative future. Analytica Chimica Acta. 1998;369:1-10
  139. 139. Folch J, Lees M, Stanley GH. A simple method for the isolation and purification of total lipids from animal tissues. Journal of Biological Chemistry. 1957;226(1):497-509
  140. 140. Iverson J, Lang SLC, Cooper MH. Comparison of the Bligh and dyer and Folch methods for total lipid determiation i a broad range of marine tissue. Paper no. L8731 in Lipids. 2001;36:1283-1287
  141. 141. Arai Y, Sako T, Takebayashi Y. Supercritical Fluids: Molecular Interactions, Physical Properties and New Applications. London: Springer; 2002
  142. 142. Sahena F, Zaidul ISM, Jinap S, et al. Application of supercritical CO2 in lipid extraction- a review. Journal of Food Engineering. 2009;95:240-253
  143. 143. Sovova H, Stateva RP. Supercritical fluid extraction from vegetable materials. Reviews in Chemical Engineering. 2011;27:79-156
  144. 144. Xu Z, Godber JS. Comparison of supercritical fluid and solvent extraction methods in extracting Gammer - oryzanol from rice bran. Journal of the American Oil Chemists' Society. 2000;77:547-551
  145. 145. Hanif M, Atsuta Y, Fujie K, et al. Supercritical fluid extraction of microbial phospholipid fatty acids from activated sludge. Journal of Chromatography A. 2010;1217(43):6704-6708
  146. 146. Kritzer P. Corrosion in high-temperature and supercritical water and aqueous solutions: A review. Journal of Supercritical Fluids. 2004;29:1-29
  147. 147. Gungoren T, Saglam M, Yuksel M, et al. Near-critical and supercritical fluid extraction of industrial sewage sludge. Industrial and Engineering Chemistry Research. 2007;6:1051-1057
  148. 148. Brunner G, Misch B, Firus A, et al. Cleaning of soil with supercritical water and supercritical carbon dioxide. In: Stegmann R, Brunner G, Calmano W, Matz G, editors. Treatment of Soil-Fundamentals, Analysis. Springer: Applications. Berlin; 2001
  149. 149. Kruse A, Dinjus E. Hot compressed water as reaction medium and product: Properties and synthesis reaction. Journal of Supercritical Fluids. 2007;39:362-380
  150. 150. Kulkarni AK, Daneshvarhosseini S, Yoshida H. Effective recovery of pure aluminum from waste composite laminates by sub- and super-critical water. The Journal of Supercritical Fluids. 2011;55:992-997
  151. 151. Mahanta P, Shrivastava A. Technology development of bio-diesel as an energy alternative. 2011. Available from: http://www.newagepublishers.com/sample chapter/001305.pdf
  152. 152. Lamsal BP, Johnson LA. Separating oil from aqueous extraction fractions of soybean. Journal of the American Oil Chemists' Society. 2007;84:785-792
  153. 153. Golmakani MT, Rezaei K. Comparison of microwave-assisted hydrodistillation with the traditional hydrodistillation method in the extraction of essential oils from Thymus vulgaris L. Food Chemistry. 2008;109:925-930
  154. 154. Mandal V, Mohan Y, Hemalatha S. Microwave-assisted extraction-an innovative and promising extraction tool for medicinal plant research. Pharmacognosy Reviews. 2007;1(1):7-18
  155. 155. Karim Assami DP. Ultrasound-induced intensification and selective extraction of essential oil from Carum carvi L. seeds. Chemical Engineering and Processing Process Intensification. 2012;62:99-105
  156. 156. Bhaskaracharya RK, Kentish S, Ashokkumar M. Selected applications of ultrasonic in food processing. Food Engineering Reviews. 2009;1:31-49
  157. 157. Sereshti H, Rohanifar A, Bakhtiari S, Samadi S. Bifunctional ultrasound assisted extraction and determination of Elettaria cardamomum Maton essential oil. Journal of Chromatography. A. 2012;1238:46-53. DOI: 10.1016/j.chroma.2012.03.061
  158. 158. Sharma YC, Singh B, Upadhyay SN. Advancements in development and characterization of biodiesel: A review. Fuel. 2008;87:2355-2373
  159. 159. Kogan A, Garti N. Microemulsions as transdermal drug delivery vehicles. Advances in Colloid and Interface Science. 2006;123–126:369-385
  160. 160. Ma F, Hanna MA. Biodiesel production: A review. Bioresource Technology. 1999;70:1-15
  161. 161. Hajjari M, Tabatabaei M, Aghbashlo M, Ghanavati H. A review on the prospects of sustainable biodiesel production: A global scenario with an emphasis on waste-oil biodiesel utilization. Renewable and Sustainable Energy Reviews. 2017;72:445-464
  162. 162. Schuchardt U, Sercheli R, Vargas RM. Transesterification of vegetable oils: A review. Journal of the Brazilian Chemical Society. 1998;9:199-210
  163. 163. Gerpen JHV, He B. Biodiesel production and property. In: Crocker M, editor. Themochemical Conversion of Biomass to Liquid Fuels and Chemicals. United Kingdom: RSC Energy and Environment Series No. 1; 2010. pp. 382-415
  164. 164. Marchetti JM, Errazu AF. Esterification of free fatty acids using sulfuric acid as catalyst in the presence of triglycerides. Biomass and Bioenergy. 2008;32:892-895
  165. 165. Mahmudul HM, Hagos FY, Mamat R, Adam AA, Ishak WFW, Alenezi R. Production, characterization and performance of biodiesel as an alternative fuel in diesel engines–A review. Renewable and Sustainable Energy Reviews. 2017;72:497-509
  166. 166. Abdeshahian P, Dashti MG, Kalil MS, et al. Production of biofuel using biomass as a sustainable biological resources. Biotechnology. 2010;9:274-282
  167. 167. Sharma A, Kodgire P, Kachhwaha SS, Raghavendra HB, Thakkar K. Application of microwave energy for biodiesel production using waste cooking oil. Materials Today: Proceedings. 2018;5:23064-23075
  168. 168. Gebremariam SN, Marchetti JM. Economics of biodiesel production: Review. Energy Conversion and Management. 2018;168:74-84
  169. 169. Atabani AE, Silitonga AS, Ong HC, Mahlia TMI, Masjuki HH, Irfan AB, et al. Non-edible vegetable oils: A critical evaluation of oil extraction, fatty acid compositions, biodiesel production, characteristics, engine performance and emissions production. Renewable and Sustainable Energy Reviews. 2013;18:211-245
  170. 170. Colombo K, Ender L, Barros AAC. The study of biodiesel production using CaO as a heterogeneous catalytic reaction. Egyptian Journal of Petroleum. 2017;26:341-349
  171. 171. Olkiewicz M, Torres CM, Jiménez L, Font J, Bengoa C. Scale-up and economic analysis of biodiesel production from municipal primary sewage sludge. Bioresource Technology. 2016;214:122-131
  172. 172. Reşitoğlu İA, Keskin A, Gürü M. The optimization of the esterification reaction in biodiesel production from trap grease. Energy Sources, Part A: Recovery, Utilization, and Environmental Effects. 2012;34:1238-1248
  173. 173. Karmee SK, Patria RD, Lin CSK. Techno-economic evaluation of biodiesel production from waste cooking oil—A case study of Hong Kong. International Journal of Molecular Sciences. 2015;16:4362-4371
  174. 174. Zhang Y, Dube M, McLean D, Kates M. Biodiesel production from waste cooking oil: 2. Economic assessment and sensitivity analysis. Bioresource Technology. 2003;90:229-240
  175. 175. Kumar A, Sharma S. An evaluation of multipurpose oil seed crop for industrial uses (Jatropha curcas L.): A review. Industrial Crops and Products. 2008;28:1-10
  176. 176. Sahoo P, Das L, Babu M, Arora P, Singh V, Kumar N, et al. Comparative evaluation of performance and emission characteristics of jatropha, karanja and polanga based biodiesel as fuel in a tractor engine. Fuel. 2009;88:1698-1707
  177. 177. Chen K-S, Lin Y-C, Hsu K-H, Wang H-K. Improving biodiesel yields from waste cooking oil by using sodium methoxide and a microwave heating system. Energy. 2012;38:151-156
  178. 178. Canakci M, Sanli H. Biodiesel production from various feedstocks and their effects on the fuel properties. Journal of Industrial Microbiology & Biotechnology. 2008;35:431-441
  179. 179. Lee J-S, Saka S. Biodiesel production by heterogeneous catalysts and supercritical technologies. Bioresource Technology. 2010;101:7191-7200
  180. 180. Reyero I, Arzamendi G, Gandía LM. Heterogenization of the biodiesel synthesis catalysis: CaO and novel calcium compounds as transesterification catalysts. Chemical Engineering Research and Design. 2014;92:1519-1530
  181. 181. Ojolo SJ, Ogunsina BS, Adelaja AO, Ogbonnaya M. Study of an effective technique for the production of biodiesel. Journal of Emerging Trends in Engineering and Applied Sciences. 2011;2:79-86
  182. 182. Brunet R, Carrasco D, Muñoz E, Guillén-Gosálbez G, Katakis I, Jiménez L. Economic and environmental evaluation of microalgae biodiesel production using process simulation tools. In: Symposium on Computer Aided Process Engineering. 2012. p. 20
  183. 183. Wei Z, Xu C, Li B. Application of waste eggshell as low-cost solid catalyst for biodiesel production. Bioresource Technology. 2009;100:2883-2885
  184. 184. Kouzu M, Hidaka J-S. Transesterification of vegetable oil into biodiesel catalyzed by CaO: A review. Fuel. 2012;93:1-12
  185. 185. Boey P-L, Maniam GP, Hamid SA. Performance of calcium oxide as a heterogeneous catalyst in biodiesel production: A review. Chemical Engineering Journal. 2011;168:15-22
  186. 186. Kargbo DM. Biodiesel production from municipal sewage sludge. Energy Fuels Reviews. 2010;24:2791-2794
  187. 187. Lam MK, Lee KT, Mohamed AR. Homogeneous, heterogeneous and enzymatic catalysis for transesterification of high free fatty acid oil (waste cooking oil) to biodiesel: A review. Biotechnology Advances. 2010;28:500-518
  188. 188. Zheng S, Kates M, Dube MA, McLean DD. Acid-catalyzed production of biodiesel from waste frying oil. Biomass and Bioenergy. 2006;30:267-272
  189. 189. West AH, Posarac D, Ellis N. Assessment of four biodiesel production processes using HYSYS. Bioresource Technology. 2008;99:6587-6601
  190. 190. Hidayat A, Rochmadi Wijaya K, Nurdiawati A, Kurniawan W, Hinode H, et al. Esterification of palm fatty acid distillate with high amount of free fatty acids using coconut shell char-based catalyst. Energy Procedia. 2015;75:969-974
  191. 191. Jegannathan KR, Eng-Seng C, Ravindra P. Economic assessment of biodiesel production: Comparison of alkali and biocatalyst processes. Renewable and Sustainable Energy Reviews. 2011;15:745-751
  192. 192. Demirbas A. Biodiesel production from vegetable oils by supercritical methanol. Journal of Scientific and Industrial Research. 2005;64:858-865
  193. 193. Sakai T, Kawashima A, Koshikawa T. Economic assessment of batch biodiesel production processes using homogeneous and heterogeneous alkali catalysts. Bioresource Technology. 2009;100:3268-3276
  194. 194. Sánchez M, Marchetti JM, El Boulifi N, Aracil J, Martínez M. Kinetics of jojoba oil methanolysis using a waste from fish industry as catalyst. Chemical Engineering Journal. 2015;262:640-647
  195. 195. Buasri A, Worawanitchaphong P, Trongyong S, Loryuenyong V. Utilization of scallop waste shell for biodiesel production from palm oil – Optimization using Taguchi method. APCBEE Procedia. 2014;8:216-221
  196. 196. Correia LM, Campelo NS, Albuquerque RF, Cavalcante CL, Cecilia JA, RodríguezCastellón E, et al. Calcium/chitosan spheres as catalyst for biodiesel production. Polymer International. 2015;64:242-249
  197. 197. Avhad MR, Marchetti JM. A review on recent advancement in catalytic materials for biodiesel production. Renewable and Sustainable Energy Reviews. 2015;50:696-718
  198. 198. Mulugetta Y. Evaluating the economics of biodiesel in Africa. Renewable and Sustainable Energy Reviews. 2009;13:1592-1598
  199. 199. Marchetti JM. Influence of economic variables on a supercritical biodiesel production process. Energy Conversion and Management. 2013;75:658-663
  200. 200. Apostolakou AA, Kookos IK, Marazioti C, Angelopoulos KC. Techno-economic analysis of a biodiesel production process from vegetable oils. Fuel Processing Technology. 2009;90:1023-1031
  201. 201. Van Kasteren J, Nisworo A. A process model to estimate the cost of industrial scale biodiesel production from waste cooking oil by supercritical transesterification. Resources, Conservation and Recycling. 2007;50:442-458
  202. 202. You Y-D, Shie J-L, Chang C-Y, Huang S-H, Pai C-Y, Yu Y-H, et al. Economic cost analysis of biodiesel production: Case in soybean oil. Energy & Fuels. 2008;22:182

Written By

Ifeanyichukwu Edeh

Submitted: 08 April 2019 Reviewed: 25 May 2020 Published: 06 July 2020