Open access peer-reviewed chapter - ONLINE FIRST

Third-Generation Biodiesel: Different Production Processes

Written By

Luciano Aguilera Vázquez, Sheila Genoveva Pérez Bravo, Nohra Violeta Gallardo Rivas, Ulises Páramo García, Ana Lidia Martínez Salazar, María Lucila Morales Rodríguez and María del Refugio Castañeda Chávez

Submitted: 26 September 2023 Reviewed: 28 September 2023 Published: 20 November 2023

DOI: 10.5772/intechopen.1003709

Microalgae IntechOpen
Microalgae Current and Potential Applications Edited by Sevcan Aydin

From the Edited Volume

Microalgae - Current and Potential Applications [Working Title]

Prof. Sevcan Aydin

Chapter metrics overview

43 Chapter Downloads

View Full Metrics

Abstract

The production of biodiesel from microalgae faces several problems to be solved, among them is the necessity of increasing their lipid content, optimizing the harvesting, and improving the conversion of lipids to bioenergy, therefore reducing the energy cost of the production process prior to its commercial launch. Research focused on optimizing the biodiesel production process known as transesterification has various objectives such as eliminating the biomass drying stage, unifying the extraction and transesterification stages, improving the reaction yield using supercritical conditions, providing heating with microwave and ultrasonic radiation, reusing enzymatic and heterogeneous catalysts, among others. This chapter aims to summarize the advances that have been achieved with the various operating conditions for the in situ, direct, and supercritical oil transesterification process of microalgae from the genera Chlorella, Scenedesmus, Spirulina, and Nannochloropsis.

Keywords

  • transesterification
  • microalgae
  • reaction conditions
  • yield
  • biodiesel

1. Introduction

Research and development on microalgae as a feedstock for third-generation biofuel have focused on increasing their production, this is due to their ability to synthesize and accumulate lipids [1]. The main interest of these studies is their doubling rate, which varies from 3.5 to 24 h, resulting in a high growth rate and high biomass yield in short periods [1, 2, 3].

The overall success of biodiesel production from microalgae begins with the selection of the right strain, which must exhibit optimal biomass production performance, as well as a high accumulation of triglycerides, and must adapt to extreme environments [4].

Obtaining lipids from microalgae begins with their growth, followed by harvesting, filtration, drying, and, finally, oil extraction [5]. Microalgal biomass production cost is estimated at $5.8 USD/kg, with an energy requirement of 33 MJ/kg dry biomass for lipid recovery [6].

Biodiesel has similar physicochemical properties to diesel, making it suitable for use in compression-ignition internal combustion engines; it is also considered to have a net contribution of carbon to the environment, high flash point, complete combustion with low sulfur emissions, and high cetane number [7]. Emissions from its combustion depend on the percentage used and the oil used for its combustion [8]. In addition, its calorific value is lower than that of diesel, being 39.45 MJ/kg versus 44.24 G/kg for diesel [9]. The biodiesel obtained from castor oil has an energy of 26.03 MJ/kg [10].

To optimize the production of biodiesel by means of transesterification, researchers have used different reaction conditions, as explained below.

Advertisement

2. Microalgae

Microalgae are unicellular photosynthetic eukaryotic microorganisms that inhabit freshwater and brackish water aquatic environments [11]. They produce fatty acids as precursors for the synthesis of several types of lipids, which vary between species: polar and neutral lipids, waxes, sterols, phospholipids, glycolipids, carotenoids, terpenes, carotenoids, tocopherols, and quinones, for the synthesis of proteins and cell growth [12], of which only neutral lipids are suitable for biodiesel production.

Chemical stimuli such as nutrient deficiency (nitrogen, phosphorus, sulfur, and silicon), salinity, and pH, as well as physical stimuli such as temperature and light intensity, influence the composition of lipids and fatty acids contained in microalgae [13, 14]. The C:N ratio is the main factor responsible for the synthesis and accumulation of intracellular lipids [15]; both elements are important in the photosynthetic metabolic pathway of microalgae [16]. Several strains of microalgae have been studied in different culture media, with the intention of increasing biomass productivity or lipid productivity; however, these studies have not focused on the utilization of sunlight and contaminated natural waters.

The fatty acids present in microalgae include linear molecules with a length of 12 to 22 carbon atoms in even numbers, both saturated and unsaturated. Unsaturated fatty acids vary in the number and position of double bonds in the main carbon chain, monounsaturated fatty acid contains only one double bond, while polyunsaturated fatty acids contain two or more double bonds [17]. The lipid contents of various species of microalgae have been reported to range from 11 to 63% on a dry weight basis and are listed in Table 1.

Microalgae% LipidsReference
Chlamydomonas reinhardtii21[18]
Chlorella vulgaris36–40[11]
Scenedesmus sp.40[12]
Chlorella sp.28–32[13]
Scenedesmus dimorphus31[14]
10.23[15]
Chlorella minutissima31[16]
Scenedesmus obliquus12–14[18]
Nannochloropsis salina50–55[17]

Table 1.

Percentage weight basis.

Among the fatty acids present in vegetable oils are the saturated palmitic (C16:0) and stearic (C18:0) fatty acids, the saturated fatty acids oleic (C18:1) and linoleic (C18:2), as well as the polyunsaturated fatty acids linolenic (C18:3), eicosapentaenoic (C20:5) and docohexanoic (C22:6) [19, 20, 21]. Microalgae with a high percentage of fatty acids, mainly those containing acids from C16 to C18, are suitable for the production of biodiesel [22].

Scenedesmus dimorphus contains fatty acids mainly from C16 to C18 and can also be obtained from C20 to C22 depending on the culture conditions. In BG-11 medium, it can accumulate from C14 to C20, the ones found in higher percentage were palmitic acid at 24.23%, linolenic acid at 44.01%, hexacosanoic acid (C26:0) at 19.12%, while when cultured in lactic acid-rich wastewater the lipid profile deviates to the accumulation of palmitic acid with 25.01% and oleic acid with 53.24% [23]. In another research where it is cultivated in municipal wastewater, the lipids obtained range from C15 to C18, predominantly palmitic acid with 21.6%, oleic acid with 23.3%, and linolenic acid with 24.01%, when the biomass is subjected to salt stress with NaCl at 5% P/V, the lipid profile changes, increasing the percentage of palmitic acid to 40.5% and stearic acid from 2.5 to 11.2%, while linolenic acid decreases to 4.16% [24]. In other words, the lipids of Scenedesmus dimorphus include palmitic, oleic, and linolenic acid in different percentages depending on the growth metabolism.

A comparative study of biomass and lipid content of Chlorella vulgaris reported higher biomass productivity with the use of mixotrophic growth (53%, 3 g/L) in comparison to autotrophic (20%, 0.8 g/L) and heterotrophic (16%, 0.2 g/L) growth, highlighting that microalgae can use both organic and inorganic carbon, nevertheless, in both cases, light is required for growth [25]. In another experiment, it was reported that Chlorella strains are capable of bio-sequestration of heavy metals such as Hg, Cd, Pb, Au, Ag, Cu, Cr (VI), and Ni [26]. The lipid profile of Chlorella vulgaris contains saturated and unsaturated fatty acids, specifically palmitic acid, linoleic acid, oleic acid, and stearic and linolenic acids. Generally, the acids present range from C16 to C18, with a higher percentage of palmitic acid [27, 28, 29, 30].

Chlorella minutissima presented 57.1% palmitic acid (C16:0), 5.92% stearic acid (C18:0), 26.3% oleic acid (C18:1) when grown in Guillard f/2 medium under 150 Lux illumination [31]. Meanwhile, Scenedesmus sp stores 25.35% of palmitic acid and 74.67% of oleic acid [32]. Finally, Spirulina platensis has been shown to store palmitic acid and oleic acid [33].

Different species of green microalgae and their growth metabolisms vary in lipid contents; however, it is observed that they tend to accumulate palmitic and oleic acid as main sources.

Advertisement

3. Conversion of biomass to bioenergy

Biodiesel is a chemically long-chain fatty acid monoalkyl ester, one of the processes to obtain it is transesterification, which consists of combining vegetable oil with alcohol, obtaining biodiesel and the value-added residue glycerol [27, 32, 34, 35].

Among the different alternatives for the production of biodiesel from microalgae as raw material, there is the conventional transesterification of oils, transesterification under supercritical operating conditions, direct transesterification of wet biomass, and in situ transesterification of dried microalgae, in addition to the effect of different types of catalysis.

Cao et al. investigated the possibility of avoiding the steps of dehydration of the biomass and extraction of its lipids by proposing a method of direct transesterification of wet biomass, observing a yield inversely proportional to the humidity, the yield of the reaction remains stable for humidities lower than 30%, the advantage of this alternative is the reduction of the energy requirements for the production of biodiesel [21]. Direct transesterification is a method that does not require dehydration of the biomass and extraction of lipids, in Table 2 it is observed that direct transesterification has been studied in supercritical conditions, in these conditions a catalyst is not required for the reaction, the reason for which the product does not require purification.

MicroalgaeOperating conditionsPerformanceReference
Transesterification under supercritical conditions
NannochloropsisAlgae moisture/methanol (W/V) 1:8, 250°C, 25 min.84.16%[36]
Nannochloropsis spWet algae/ethanol (W/V) 1:9, 25 min, 265°C, 80 Bar30.9%[17]
Chlorella sp20 gr algae mesh 40, 2 L/min CO2, 60°C, 20 MPa63.7%[37]
Schizochytrium limacinumBiomass 1.6% moisture: methanol 1:6, 210°C, 7.6 MPa, 60 min.84%[38]
Chlorella prothothecoidesMicroalgae oil: methanol 1:19, 320°C, 152 bar, 31 min90.8%[39]
Microalgae oil: ethanol 1:33, 340°C, 170 bar, 35 min87.8%
Chlorella spBiomass 75% moisture: methanol 1:10 P/V, 265°C, 50 min42.62%[40]
Nannochloropsis oculataBiomass 75% moisture: methanol 1:10 P/V, 265°C, 50 min21.67%
Spirulina platensisBiomass: methanol 1gr:8 ml, cosolvent hexane 4 ml: gr biomass 40% moisture, 244.8 °C, 6.61 MPa99.32%[41]
ChlorellaMicroalgae oil: methanol 1:30, 270°C, 8.1 MPa, 30 min97.1%[42]
Spirulina platensisMicroalgae oil: methanol 1:40, 0.003 g CO2/gr methanol, 300°C, 20 MPa, 30 min72%[43]
Microalgae oil: ethanol 1:40, 0.001 g CO2/gr methanol, 300°C, 20 MPa, 30 min68%
Schizochytrium limacinumBiomass: methanol 1:10 P/V, 260°C, 20 MPa, 20 min61.4%[44]
Biomass: methanol 1:10 P/V, 270°C, 20 MPa, 40 min>90%
Biomass: dimethyl carbonate 1:10 P/V, 270°C, 20 MPa, 30 min>50%
Biomass: methyl acetate 1:10 P/V, 270°C, 20 MPa, 30 min>40%
Schizochytrium limacinumBiomass: etanol 1:10 P/V, 320°C, 20 MPa, 50 min>95%[45]
Biomass: etanol 1:10 P/V, 380°C, 20 MPa, 60 min>60%
Direct transesterification
Chlorella vulgaris200 mg biomass with 5.3% moisture:3 ml methanol, H2SO4 (mol:80.8 mol), 60°C, 120 min, 200 rpm94.6%[23]
200 mg biomass with 5.3% moisture:3 ml methanol, KOH (mol:19.9 mol), 60°C, 120 min, 200 rpm70.5%
Chlorella vulgarisBiomass at 86–91% moisture: methanol 1:66.9, lipase Burkholderia 2028 U/g, 45°C, hexane: methanol 1:65, 500 rpm, 48 h58.21%[46]
Chlorella vulgarisBiomas moisture 40.6%, Biomass: methanol 1:9, biomass: IL (Phosphonium carboxylate) 1:8 W, 102.4°C, 4.6 h98.6 ± 1.82%[47]
In situ transesterification
Chlorella sp.Biomass: methanol 100 mg:12 ml, H2 SO4 3%, 65°C, 2 h45.26%[19]
Biomass: methanol 100 mg:12 ml, H2 SO4 4%, 65°C, 2 h53.19%
Biomass: methanol 100 mg:12 ml, H2 SO4 3.5%, 65°C, 2 h60.71%
Biomass: methanol 100 mg:12 ml, H3 PO3 3%, 65°C, 2 h39.12%
Biomass: methanol 100 mg:12 ml, H2 SO4 3% (E1), NaOH 1%W (E2), 65°C, 2 h43.92%
Spirulina platensisBiomass: methanol 1gr:5 ml, NaOH 0.5 gr/gr, 60°C, 3 h30%[20]
Biomass: methanol 1 g:5 ml, H2 SO4, 1 gr/gr, 60°C, 3 h79.5%
Biomass: methanol 1gr:5 ml, HCl, 1 gr/gr, 60°C, 3 h58.6%
Biomass: methanol 1 g:5 ml, HNO3 1 g/gr, 60°C, 3 h40.3%
Biomass: methanol 1 g:5 ml, H3 PO4, 1 gr/gr, 60°C, 3 h3.6%
Scenedesmus sp.0.5 gr algae, 5 ml DMC1/gr microalgae, 1% W/W water/microalgae, 50°C, 36 h, 180 rpm92%[12]
Chlorella pyrenoidosa4 ml methanol, 8 ml n-hexane, H2 SO4 0.5 M, 120°C, 180 min92.5%[21]
Nannochloropsis gaditana (11.1% saponifiable lipids LS and 75% moisture)Methanol/LS 171.1 mg/gr, 5% V/V acetyl chloride, 100°C, 105 min, 2.5 atm, 104 ml/gr hexane100%[22]
Nannocloropsis sp1gr biomass, lipid: methanol 1:200, KOH 2%W, methanol: hexane 1:0, 60°C, 2 h49.9%[48]
1gr biomass, lipid: methanol 1:200, KOH 2%W, methanol: hexane 1:1, 60°C, 2 h56.1%
1gr biomass, lipid: methanol 1:200, KOH 2%W, methanol: hexan 1:1.5, 60°C, 2 h48.43%
1gr biomass, lipid: methanol 1:300, KOH 2%W, methanol: hexan 1:1.5, 60°C, 2 h76.77%
1gr biomass, lipid: methanol 1:400, KOH 2%W, methanol: hexan 1:1.5, 60°C, 2 h78.87%
1gr biomass, lipid: methanol 1:400, KOH 2%W, methanol: hexane 1:1.5, 60°C, 4 h90.95%
Spirulina spBiomass: methanol 50gr:100 ml, KOH 2%W, biomass: cosolvent 1:200 chloroform: methanol, 180 W ultrasound, 15 min1.6% W Biomass 95.2% FAME content[49]
Chlorella spBiomass: methanol 1gr:120 ml, 10%W catalyst2, 80°C, 3 h33%[50]
Biomass: methanol 1gr:120 ml, 20%W catalyst2, 80°C, 3 h47%
Chlorella vulgaris, Scenedesmus bijurjos, Euglena sp, Oscillatoria quadripuntctulata, Microcystisaeruginosa and Chlamydomonas debaryanaBiomass: methanol 1gr:12 ml, 4%W Fe2O3-Al2O3, 6 h95.6%[51]
Chlorella pyrenoidosa1 gr biomass, methanol: CPME3: HCl 10:1.5:1, 80°C, 380 rpm, 150 min92%[52]
1 gr biomass, methanol: 2-MeTHF4: HCl 10:1.5:1, 80°C, 380 rpm, 150 min91%
1 gr biomass, methanol: chloroform: HCl 10:1.5:1, 80°C, 380 rpm, 150 min76%
Microwave-assisted transesterification
NannochloropsisDry biomass/methanol (W/V) 1:12, KOH 2%W, 6 min, 60–64°C80.1%[36]
Chlorella pyrenoidosaWet biomass (77%): methanol+ H2SO4 3% V, (2gr:6 ml), 70°C, 30 min.50%[24]
Wet biomass (77%): ethanol+ H2SO4 3% V, (2gr:6 ml), 70°C, 30 min.80.6%
Wet biomass (77%): isopropanol+ H2SO4 3% V, (2gr:6 ml), 70°C, 30 min.85.7%
Transesterification of oils with homogeneous catalyst
Chlorella vulgarisMethanol: oil 6:1, NaOH 1%, 55°C, 15 min.>98%[25]
Tetraselmis spMethanol: oil 6:1, 0.5 g NaOH, 65°C, 300 rpm, 3 hNot specified[26]
Chlorella vulgarisMethanol: oil 10:1, 1.5% H2SO4 V/V, 60°C, 40 min (E1)
Methanol: oil 10:1, KOH 1.5%, 60°C, 15 min microwave 700 W (E2)
84.01%[28]
Methanol: oil 10:1, 1.5% H2 SO4 V/V, 60°C, 40 min (E1)
Methanol: oil 10:1, KOH 1.5%, 60°C, 3 h (E2)
83.23%
Chlorella vulgaris38 mg oil:3 ml methanol, H2SO4 (mol:50.7 mol), 60°C, 120 min, 200 rpm96.6%[23]
38 mg oil:3 ml methanol, KOH (mol:2.5 mol), 60°C, 120 min, 200 rpm90.4%
Chlorella spMethanol: oil 30:1, H2SO4 10%P, 4 h87.5%[53]
Ethanol: oil 30:1, H2SO4 10%P, 4 h75.4%
Transesterification of oils with heterogeneous catalyst
Nannochloropsis oculataMethanol: oil 6:1, 80% CaO/Al2O3 2% W/W, 50°C, 4 h, 1100 rpm23[29]
Methanol: oil 30:1, 80% CaO/Al2O3 2% W/W, 50°C, 4 h, 1100 rpm97.5%
Tetraselmis spIsooctane: oil 5 ml:10 ml, 1gr immobilized Candida rugosa, 60 ml methanol, 40°C, 180 rpm, 10 hNot specified[26]
Nannochloropsis oculataMethanol: oil 15:1, 3.5% W/W ZnO: Mn+2 encapsulated in PEG5, 60°C, 4 h87.5%[30]
Scenedesmus sp.Methanol: oil 11:1, goat bone nanocatalyst 2% P, 60°C, 1500 rpm, 3 h92%[31]
Chlorella sp., Scenedesmus sp. Synechocystiss sp., Spirulina sp.Methanol: oil 20:1, 5% P sulfonated peanut shell biochar, 65°C, 4 h92%[33]
Chlorella vulgarisMethanol: oil in hexane (5 mg: ml) 0.15 ml:0.5 ml, 8.6 mg ROL/MNP-AP-GA6, 45°C, 24 h69.8%[54]

Table 2.

Third-generation biodiesel production processes.

Dimethylcarbonate.


LiOH-pumice.


Cyclopentyl methyl ether.


2-Methyltetrahydrofuran.


Polyethylene glycol.


Rhizopus oryzae lipase (ROL)-Superparamagnetic nanoparticles with 3-aminopropyl triethylenesilane-gutaraldehyde (MNP-AP-GA).


Sivaramakrishnan et al. investigated the direct transesterification of the microalgae Scenedesmus sp. using dimethyl carbonate (DMC) and lipase (Lipozyme-Novozyme CAL-B from Candida antarctica, Novozymes Denmark) immobilized on Celite (commercial brand of diatoms), with 1% water for 36 h at 50°C, stirring at 180 rpm obtaining a yield of 92%, before the synthesis, a process of cell disruption is carried out with ultrasound, the use of lipase in this process showed a yield of 90% up to the fifth cycle, addition, this process generates glycerol carbonate as a by-product [12]. he most common methylating agent is methanol, as a by-product of the transesterification reaction, glycerin is obtained, which is oversupplied in the market, consequently, its price has decreased, the use of different methylating agents such as dimethyl carbonate, and methyl acetate produced as a by-product glycerol dicarbonate and triacetin [44].

Another alternative method considers the use of supercritical methanol coupled to oil extraction with CO2 under supercritical conditions, for the case of biodiesel production from Chlorella sp. the advantage of this process is the utilization of the extraction residues in the production of drugs because carbohydrates, trace elements, and minerals are not damaged in the process and can be easily separated from CO2 by decompression [37]. Supercritical transesterification of microalgae oils offers better yields than when using dry or wet biomass, on the other hand, the use of cosolvents improves the process.

Patil et al. compared the yield of a simultaneous microwave-assisted extraction and transesterification reaction under supercritical conditions of the microalgae Nannocloropsis obtaining yields of 80.13 and 84.15% respectively, they conclude that microwave radiation is more effective in cell lysis, due to that the reaction time is shorter, the process in supercritical conditions consumes 600 KJ while the assisted transesterification consumes only 254.5 KJ, that is, this alternative is an easy, fast and efficient process with a lower energy requirement [36].

Biodiesel is influenced by the physicochemical characteristics of the feedstock from which it is produced. Because of this relationship, microalgae oil should contain long-chain fatty acids with a low degree of unsaturation, preferably palmitoleic (16:1), oleic (18:1), and myristic (14:1) acids, thus allowing the reduction of toxic emissions, improving the cetane number and oxidative stability, without affecting lubricity, viscosity, and cloud point. Common methyl esters in biodiesel are methyl palmitate, methyl stearate, methyl oleate, methyl linoleate, methyl linolenate, and methyl laureate [55].

Advertisement

4. Conclusions

The production of third-generation biodiesel has focused on the variation of operating conditions to increase yield. Direct transesterification has the advantage of converting wet biomass directly avoiding the drying and extraction stages, as well as transesterification under supercritical conditions, another advantage of the latter is that it does not require the use of a catalyst, consequently, the biodiesel does not need purification. Both processes have yields inversely proportional to the amount of moisture in the biomass, the use of a cosolvent increases the reaction yield.

Microwave and ultrasound-assisted transesterification have yields greater than 75% in shorter reaction times, which is attributed to the higher mass transfer. The transesterification of oils with homogeneous and heterogeneous catalysis presents yields higher than 87%, the advantage of heterogeneous catalysts is their reusability, easy recovery, and elimination of the biodiesel purification stage.

Among the variables that affect the reaction yield is that a high temperature favors mass transfer, and an excess of alcohol favors the formation of products, however, it must be optimized to avoid yield losses due to the dissolution of biodiesel in alcohol, likewise an excess of catalyst can reduce the reaction yield.

Advertisement

Acknowledgments

We are grateful for the support of the National Council of Science and Technology (CONACYT), for the scholarship granted to Sheila Genoveva Pérez Bravo to study the Doctorate in Engineering Sciences within the doctoral program of the Instituto Tecnológico de Ciudad Madero, included in the PNPC, under agreement 21542.

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. Zhu LD, Li ZH, Guo DB, Huang F, Nugroho Y, Xia K. Cultivation of chlorella sp. with livestock waste compost for lipid production. Bioresource Technology. 2017;223:296-300. DOI: 10.1016/j.biortech.2016.09.094
  2. 2. Chisti Y, Mata TM, Martins AA, Caetano NS. Microalgae for biodiesel production and other applications: A review. Biotechnology Advances. 2007;25(1):294-306. DOI: 10.1016/j.biotechadv.2007.02.001
  3. 3. Alam F, Mobin S, Chowdhury H. Third generation biofuel from algae. Procedia Engineering. 2015;105:763-768. DOI: 10.1016/j.proeng.2015.05.068
  4. 4. Marisol M-H, Suastes-Acosta SL, Lozano-Ramírez Cruz RPMC. Lipidic profile of Lagerheinia Sp. isolated of Idustrial. Avances en Ciencias e Ingeniería. 2018;9(1):25-33
  5. 5. 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(4):163-167. DOI: 10.1016/j.hjb.2017.10.003
  6. 6. Castillo OSS, Torres-Badajoz SGG, Núñez-Colín CAA, Peña-Caballero V, Herrera Méndez CHH, Rodríguez-Núnez JRR. Biodiesel production from microalgae: Progress and biotechnological prospects | Producción de biodiésel a partir de microalgas: Avances y perspectivas biotecnológicas. Hidrobiologica. 2017;27(3):337-352
  7. 7. Malani RS, Shinde V, Ayachit S, Goyal A, Moholkar VS. Ultrasound–assisted biodiesel production using Heterogeneous Base catalyst and mixed non–edible oils. Ultrasonics Sonochemistry. 2018;52:232-243. DOI: 10.1016/j.ultsonch.2018.11.021
  8. 8. 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(May):222-247. DOI: 10.1016/j.rser.2017.05.048
  9. 9. Sharma T, Chauhan RS. Comparative transcriptomics reveals molecular components associated with differential lipid accumulation between microalgal sp., Scenedesmus dimorphus and Scenedesmus quadricauda. Algal. Research. 2016;19:109-122. DOI: 10.1016/j.algal.2016.07.020
  10. 10. Naveenkumar R, Baskar G. Process optimization, green chemistry balance and technoeconomic analysis of biodiesel production from castor oil using heterogeneous nanocatalyst. Bioresource Technology. 2021;320:124347. DOI: 10.1016/j.biortech.2020.124347
  11. 11. Trivedi T, Jain D, Mulla NSS, Mamatha SS, Damare SR, Sreepada RA, et al. Improvement in biomass, lipid production and biodiesel properties of a euryhaline Chlorella vulgaris NIOCCV on mixotrophic cultivation in wastewater from a fish processing plant. Renewable Energy. 2019;139(3):326-335. DOI: 10.1016/j.renene.2019.02.065
  12. 12. Sivaramakrishnan R, Incharoensakdi A. Utilization of microalgae feedstock for concomitant production of bioethanol and biodiesel. Fuel. 2018;217(January):458-466. DOI: 10.1016/j.fuel.2017.12.119
  13. 13. Chisti Y. Biodiesel from microalgae beats bioethanol. Biotechnology Advances. 2008;26(3):126-131. DOI: 10.1016/j.biotechadv.2007.02.001
  14. 14. Shen Y, Pei Z, Yuan W, Mao E. Effect of nitrogen and extraction method on algae lipid yield. International Journal of Agricultural & Biological Engineering. 2009;2(1):51-57
  15. 15. Bordoloi N, Narzari R, Sut D, Saikia R, Chutia RS, Kataki R. Characterization of bio-oil and its sub-fractions from pyrolysis of Scenedesmus dimorphus. Renewable Energy. 2016;98:245-253. DOI: 10.1016/j.renene.2016.03.081
  16. 16. Illman AMM, Scragg AHH, Shales SWW. Increase in chlorella strains calorific values when grown in low nitrogen medium. Enzyme and Microbial Technology. 2000;27:631-635
  17. 17. Patil PD, Reddy H, Muppaneni T, Schaub T, Holguin FO, Cooke P, et al. In situ ethyl ester production from wet algal biomass under microwave-mediated supercritical ethanol conditions. Bio/Technology. 2013;139:308-315. DOI: 10.1016/j.biortech.2013.04.045
  18. 18. Spolaore P, Joannis-cassan C, Duran E, Isambert A, De GL, Paris EC. Commercial applications of microalgae. Journal of Bioscience and Bioengineering. 2006;101(2):87-96. DOI: 10.1263/jbb.101.87
  19. 19. Mathimani T, Uma L, Prabaharan D. Homogeneous acid catalysed transesterification of marine microalga chlorella sp. BDUG 91771 lipid—An efficient biodiesel yield and its characterization. Renewable Energy. 2015;81:523-533. DOI: 10.1016/j.renene.2015.03.059
  20. 20. Sumprasit N, Wagle N, Glanpracha N, Annachhatre AP. Biodiesel and biogas recovery from Spirulina platensis. International Biodeterioration & Biodegradation. 2017;119:196-204. DOI: 10.1016/j.ibiod.2016.11.006
  21. 21. Cao H, Zhang Z, Wu X, Miao X. Direct biodiesel production from wet microalgae biomass of Chlorella pyrenoidosa through in situ transesterification. BioMed Research International. 2013;2013:1-6. DOI: 10.1155/2013/930686
  22. 22. Macías-Sánchez MD, Robles-Medina A, Hita-Peña E, Jiménez-Callejón MJ, Estéban-Cerdán L, González-Moreno PA, et al. Biodiesel production from wet microalgal biomass by direct transesterification. Fuel. 2015;150:14-20. DOI: 10.1016/j.fuel.2015.01.106
  23. 23. Nguyen TT, Uemura Y, Lam MK, Mansor N, Lim JW. Revealing the effect of reaction parameters towards alkyl group distribution in in-situ transesterification of Chlorella vulgaris. Energy Conversion and Management. 2019;185(December 2018):223-231. DOI: 10.1016/j.enconman.2019.01.113
  24. 24. Huang R, Cheng J, Qiu Y, Li T, Zhou J, Cen K. Using renewable ethanol and isopropanol for lipid transesterification in wet microalgae cells to produce biodiesel with low crystallization temperature. Energy Conversion and Management. 2015;105:791-797. DOI: 10.1016/j.enconman.2015.08.036
  25. 25. Xaaldi Kalhor A, Mohammadi Nassab AD, Abedi E, Bahrami A, Movafeghi A. Biodiesel production in crude oil contaminated environment using Chlorella vulgaris. Bioresource Technology. 2016;222:190-194. DOI: 10.1016/j.biortech.2016.09.113
  26. 26. Teo CL, Jamaluddin H, Zain NAM, Idris A. Biodiesel production via lipase catalysed transesterification of microalgae lipids from Tetraselmis sp. Renewable Energy. 2014;68:1-5. DOI: 10.1016/j.renene.2014.01.027
  27. 27. Canakci M, Van GJ. Biodiesel production from oils and fats with high free fatty acids. American Society of Agricultural Engineers. 2001;44(6):1429-1436. DOI: 10.1016/j.fuel.2013.09.020
  28. 28. Sharma AK, Sahoo PK, Singhal S, Joshi G. Exploration of upstream and downstream process for microwave assisted sustainable biodiesel production from microalgae Chlorella vulgaris. Bioresource Technology. 2016;216:793-800. DOI: 10.1016/j.biortech.2016.06.013
  29. 29. Umdu ES, Tuncer M, Seker E. Transesterification of Nannochloropsis oculata microalga’s lipid to biodiesel on Al2O3 supported CaO and MgO catalysts. Bioresource Technology. 2009;100(11):2828-2831. DOI: 10.1016/j.biortech.2008.12.027
  30. 30. Raj JVA, Bharathiraja B, Arokiyaraj S, Iyyappan J, Praveen Kumar R. Biodiesel production from microalgae Nannochloropsis oculata using heterogeneous poly ethylene glycol (PEG) encapsulated ZnOMn2+ nanocatalyst. Bioresource Technology. 2019;282:348-352. DOI: 10.1016/j.biortech.2019.03.030
  31. 31. Mamo TT, Mekonnen YS. Microwave-assisted biodiesel production from microalgae, Scenedesmus species, using goat bone–made Nano-catalyst. Applied Biochemistry and Biotechnology. 2020;190(4):1147-1162. DOI: 10.1007/s12010-019-03149-0
  32. 32. Berrios M, Skelton RLL. Comparison of purification methods for biodiesel. Chemical Engineering Journal. 2008;144(3):459-465. DOI: 10.1016/j.cej.2008.07.019
  33. 33. Behera B, Selvam SM, Dey B, Balasubramanian P. Algal biodiesel production with engineered biochar as a heterogeneous solid acid catalyst. Bioresource Technology. 2020;310:1223392. DOI: 10.1016/j.biortech.2020.123392
  34. 34. Hupp AM, Perron J, Roques N, Crandall J, Ramos S, Rohrback B. Analysis of biodiesel-diesel blends using ultrafast gas chromatography (UFGC) and chemometric methods: Extending ASTM D7798 to biodiesel. Fuel. 2018;231(May):264-270. DOI: 10.1016/j.fuel.2018.05.102
  35. 35. Tran DT, Chang JS, Lee DJ. Recent insights into continuous-flow biodiesel production via catalytic and non-catalytic transesterification processes. Applied Energy. 2016;185:376-409. DOI: 10.1016/j.apenergy.2016.11.006
  36. 36. Patil PD, Gude VG, Mannarswamy A, Cooke P, Nirmalakhandan N, Lammers P, et al. Comparison of direct transesterification of algal biomass under supercritical methanol and microwave irradiation conditions. Fuel. 2012;97:822-831. DOI: 10.1016/j.fuel.2012.02.037
  37. 37. Zhou D, Qiao B, Li G, Xue S, Yin J. Continuous production of biodiesel from microalgae by extraction coupling with transesterification under supercritical conditions. Bioresource Technology. 2017;238:609-615. DOI: 10.1016/j.biortech.2017.04.097
  38. 38. Bi Z, He BB, McDonald AG. Biodiesel production from green microalgae schizochytrium limacinum via in situ transesterification. Energy and Fuels. 2015;29(8):5018-5027. DOI: 10.1021/acs.energyfuels.5b00559
  39. 39. Nan Y, Liu J, Lin R, Tavlarides LL. Production of biodiesel from microalgae oil (Chlorella protothecoides) by non-catalytic transesterification in supercritical methanol and ethanol: Process optimization. Journal of Supercritical Fluids. 2015;97:174-182. DOI: 10.1016/j.supflu.2014.08.025
  40. 40. Mohamadzadeh Shirazi H, Karimi-Sabet J, Ghotbi C. Biodiesel production from spirulina microalgae feedstock using direct transesterification near supercritical methanol condition. Bioresource Technology. 2017;239:378-386. DOI: 10.1016/j.biortech.2017.04.073
  41. 41. Karimi Jazzar S, Quesada-Medina J, Olivares-Carrillo P, Marzouki MN, Acién-Fernández FG, Fernández-Sevilla JM, et al. A whole biodiesel conversion process combining isolation, cultivation and in situ supercritical methanol transesterification of native microalgae. Bioresource Technology. 2015;190:281-288. DOI: 10.1016/j.biortech.2015.04.097
  42. 42. Srivastava G, Paul AK, Goud V. Optimization of non-catalytic transesterification of microalgae oil to biodiesel under supercritical methanol condition. Energy Conversion and Management. 2018;156(October 2017):269-278. DOI: 10.1016/j.enconman.2017.10.093
  43. 43. Tobar M, Núñez GA. Supercritical transesterification of microalgae triglycerides for biodiesel production: Effect of alcohol type and co-solvent. Journal of Supercritical Fluids. 2018;137(November 2017):50-56
  44. 44. Mani Rathnam V, Madras G. Conversion of Shizochitrium limacinum microalgae to biodiesel by non-catalytic transesterification using various supercritical fluids. Bioresource Technology. 2019;288(May):121538. DOI: 10.1016/j.biortech.2019.121538
  45. 45. Mani Rathnam V, Modak JM, Madras G. Non-catalytic transesterification of dry microalgae to fatty acid ethyl esters using supercritical ethanol and ethyl acetate. Fuel. 2020;275(October 2019):117998. DOI: 10.1016/j.fuel.2020.117998
  46. 46. Tran DT, Chen CL, Chang JS. Effect of solvents and oil content on direct transesterification of wet oil-bearing microalgal biomass of Chlorella vulgaris ESP-31 for biodiesel synthesis using immobilized lipase as the biocatalyst. Bioresource Technology. 2013;135:213-221. DOI: 10.1016/j.biortech.2012.09.101
  47. 47. Malekghasemi S, Kariminia HR, Plechkova NK, Ward VCA. Direct transesterification of wet microalgae to biodiesel using phosphonium carboxylate ionic liquid catalysts. Biomass and Bioenergy. 2021;150(May):106126. DOI: 10.1016/j.biombioe.2021.106126
  48. 48. Dianursanti RP, Wijanarko A. Utilization of n-hexane as co-solvent to increase biodiesel yield on direct transesterification reaction from marine microalgae. Procedia environmental. Science. 2015;23(Ictcred 2014):412-420. DOI: 10.1016/j.proenv.2015.01.059
  49. 49. Martínez N, Callejas N, Morais EG, Vieira Costa JA, Jachmanián I, Vieitez I. Obtaining biodiesel from microalgae oil using ultrasound-assisted in-situ alkaline transesterification. Fuel. 2017;202:512-519
  50. 50. De Luna MDG, Doliente LMT, Ido AL, Chung TW. In situ transesterification of chlorella sp. microalgae using LiOH-pumice catalyst. Journal environmental. Chemical Engineering. 2017;5(3):2830-2835. DOI: 10.1016/j.jece.2017.05.006
  51. 51. Kazemifard S, Nayebzadeh H, Saghatoleslami N, Safakish E. Application of magnetic alumina-ferric oxide nanocatalyst supported by KOH for in-situ transesterification of microalgae cultivated in wastewater medium. Biomass and Bioenergy. 2019;129(October 2018):105338. DOI: 10.1016/j.biombioe.2019.105338
  52. 52. de Jesus SS, Ferreira GF, Moreira LS, Filho RM. Biodiesel production from microalgae by direct transesterification using green solvents. Renewable Energy. 2020;160:1283-1294
  53. 53. Lemões JS, Alves Sobrinho RCM, Farias SP, De Moura RR, Primel EG, Abreu PC, et al. Sustainable production of biodiesel from microalgae by direct transesterification. Sustainable Chemistry and Pharmacy. 2016;3:33-38. DOI: 10.1016/j.scp.2016.01.002
  54. 54. Nematian T, Salehi Z, Shakeri A. Conversion of bio-oil extracted from Chlorella vulgaris micro algae to biodiesel via modified superparamagnetic nano-biocatalyst. Renewable Energy. 2020;146:1796-1804. DOI: 10.1016/j.renene.2019.08.048
  55. 55. Knothe G. Calidad del combustible biodiésel y la norma ASTM. Palmas. 2010;31(II):162-171

Written By

Luciano Aguilera Vázquez, Sheila Genoveva Pérez Bravo, Nohra Violeta Gallardo Rivas, Ulises Páramo García, Ana Lidia Martínez Salazar, María Lucila Morales Rodríguez and María del Refugio Castañeda Chávez

Submitted: 26 September 2023 Reviewed: 28 September 2023 Published: 20 November 2023