Next Article in Journal
Workforce Planning Framework for a Mobile Call Center Considering a Special Event
Previous Article in Journal
Life-Related Hazards of Materials Applied to Mg–S Batteries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Latest Expansions in Lipid Enhancement of Microalgae for Biodiesel Production: An Update

1
Department of Biotechnology, Sir J. C. Bose Technical Campus Bhimtal, Kumaun University, Nainital 263136, Uttarakhand, India
2
Department of Life Sciences, School of Basic Sciences and Research, Sharda University, Greater Noida 201310, Uttar Pradesh, India
3
Department of Materials Science and Engineering, University of Pennsylvania, Philadelphia, PA 19104, USA
4
Lam Research Corporation, Fremont, CA 94538, USA
5
Department of Life Sciences, Graphic Era Deemed to be University, Dehradun 248002, Uttarakhand, India
6
Biorefining and Advanced Materials Research Centre, SRUC (Scotland’s Rural College), Edinburgh EH9 3JG, UK
7
Department of Mechanical Engineering, School of Engineering, Shiv Nadar University, Noida 201314, Uttar Pradesh, India
8
School of Engineering, University of Petroleum & Energy Studies (UPES), Dehradun 248007, Uttarakhand, India
*
Authors to whom correspondence should be addressed.
Energies 2022, 15(4), 1550; https://doi.org/10.3390/en15041550
Submission received: 31 December 2021 / Revised: 27 January 2022 / Accepted: 14 February 2022 / Published: 19 February 2022

Abstract

:
Research progress on sustainable and renewable biofuel has gained motion over the years, not just due to the rapid reduction of dwindling fossil fuel supplies but also due to environmental and potential energy security issues as well. Intense interest in microalgae (photosynthetic microbes) as a promising feedstock for third-generation biofuels has grown over recent years. Fuels derived from algae are now considered sustainable biofuels that are promising, renewable, and clean. Therefore, selecting the robust species of microalgae with substantial features for quality biodiesel production is the first step in the way of biofuel production. A contemporary investigation is more focused on several strategies and techniques to achieve higher biomass and triglycerides in microalgae. The improvement in lipid enhancement in microalgae species by genetic manipulation approaches, such as metabolic or genetic alteration, and the use of nanotechnology are the most recent ways of improving the production of biomass and lipids. Hence, the current review collects up-to-date approaches for microalgae lipid increase and biodiesel generation. The strategies for high biomass and high lipid yield are discussed. Additionally, various pretreatment procedures that may aid in lipid harvesting efficiency and improve lipid recovery rate are described.

1. Introduction

Algae is the world’s largest photosynthetic group that contributes most of the carbon sequestration on the globe, converting greenhouse gases into carbohydrates and lipids. These photosynthetic microbes have received high attention as potential cell factories for fatty acids (FA) and carotenoid production. Microalgae oil is used as biodiesel and has significant advantages over vegetable oils. Biodiesel acquired from microalgae is sulfur-free and releases low hydrocarbon, CO, NOx [1], and Sox emissions in contrast to traditional petroleum diesel [2,3]. However, cultivation conditions, harvesting, and cost reduction is a key barrier to its practical commercialization [4]. Biofuel production from microalgae alone does not satisfy the economic feasibility. Hence, to improve the budget and reduce the cultivation cost, the source could be utilized in many ways, such as wastewater treatment, sewage treatment, CO2 sequestration [5]. Its co-products (protein, carbohydrates, pigments, vitamins, and antioxidants) could further be utilized in the pharmaceutical and nutraceutical industries [6,7].
Under positive growing conditions, microalgal species typically accumulate lipids between 10% and 30% of their dry weight. Some species of algae have been documented to yield greater amounts of lipids (56% in Nannochloris sp. 80% in Schizochytrium sp.). On the other hand, Chlorella sp. and Scenedesmus sp. have comparatively less lipid content but a greater growth level [8,9]. Concerning conditions required for optimum growth and lipid accumulation, algae strains have been reported to have contradictory behavior [10]. To obtain the cost-effective biodiesel cultivation of numerous low-lipid cells or a few high-lipid cells will not lead to the economically sustainable production of microalgae-derived biofuel [11]. Thus, appropriate strategies should be implemented to rectify these opposing traits so an ideal equilibrium between microalgae biomass and lipid content can be maintained [12,13]. Genetic engineering in microalgae offers a lot of possibilities to expand the procedure (Figure 1). Rapid advancements in the synthesis of DNA, tools and methods for genetic manipulation, and the accessibility of functioning genomes have expanded the potential for improved engineering in microalgae in recent years. Different environmental, nutritional, and physiological conditions have also been tried for microalgae cultivation to improve lipid production. Additionally, nanoparticles (NPs) have been widely used as an effective method to resolve barriers and technological limitations regarding the two stages [14]. The successful retrieval and recycling of NPs using economic and cost-effective technologies is a critical component of microalgae harvesting research. This review is a comprehensive study of basically two recent techniques, genetic engineering approaches and the application of nanoparticles for lipid enhancement, simultaneously using various pretreatment methods of lipid recovery to overcome bottlenecks of biodiesel production.

2. Genetic Engineering in Microalgae

Over the years, a balance has been pursued between increasing the lipid percent of microalgae by various methods while concurrently maintaining lipid productivity. Storage lipids in microalgae are usually neutral or triacylglycerides (TAGs) [15]. TAGs are biosynthesized in plastids, mitochondria, and the endomembrane and have esterified FA chains linked to the hydroxyl groups of the glycerol backbone. Specific genes that code for components of a metabolic pathway can be altered through genetic engineering in microalgae strains to improve the metabolites synthesis [16]. Methods such as zinc finger nucleases (ZFN), homologous recombination (HR), and transcription activator-like effector nucleases (TALEN) have been utilized to change the genetic makeup of eukaryotic cells [17,18]. However, putting these approaches into practice is time-consuming, difficult, and costly. TAG and FA synthesis involved a series of reactions driven by a variety of enzymes. Overexpression of these enzymes resulted in an increase in their function, which would positively promote lipid accumulation [19]. Genetically modified microalgae, such as Dunaliella salina, Chlamydomonas reinhardtii, and Pharodactylum tricornutum, were explored to boost FA synthesis, and consequently, lipid accumulation [20]. Some microalgae strains and overexpression of genes responsible for lipids biosynthesis are summarized in Table 1. Moreover, it has been opined that genetically engineered microalgae may change other biosynthetic pathways, which could produce toxicity and could affect other beneficial microbes and the environment. Therefore, before releasing genetically altered microalgae, it must be examined and approved by international committees [21]. A strategy such as biodiesel production in a closed photobioreactor could be promoted applied to avoid risk [22].

3. Regulation of Biosynthetic Pathways

The restriction of metabolic pathways that promote the storage of energy-rich compounds is another approach that leads to the increase in cellular lipid content [38]. In microalgae, starch and lipid synthesis share a common precursor [39]. When the biosynthesis of the starch pathway is blocked, carbon flux is diverted towards the lipid biosynthetic pathway, resulting in a rise in FA, consequently raising total FAs [40]. The ADP-glucose pyrophosphorylase or isoamylase genes, sta6 and sta7 mutants, were disrupted, respectively, in two separate starch-deficient C. reinhardtii strains [41,42]. During the period of N deprivation, these mutants accumulated higher quantities of TAGs [43]. The starchless mutant of Chlorella pyrenoidosa is reported to have higher polyunsaturated FAs [44].

3.1. Shift of Starch Pathway to Lipid Pathway

In photosynthetic cells, starch and triglycerides are the main carbon storage components, and lipid and starch production proceed via competing metabolic processes. Microalgae change their lipid biosynthetic pathways as a carbon and energy storage form in adverse conditions (stress) to accumulate higher TAGs [45,46]. Fan et al. [47] stated carbon availability is a critical metabolic component influencing lipid production and carbon segregation among starch and lipids. TAG content was found to be much greater (up to three times against control) in starch-deficient Dunaliella tertiolecta mutants after N-depletion [48]. The blocking of starch biosynthesis pathways increased lipid accumulation in Chlorella sp. and Chlamydomonas is perceived in various studies [43,44,49,50]. Though the metabolic process in microalgae that controls carbon splitting from starch to lipids is still unclear, investigations are underway harnessing the mechanism behind the shift of pathways. Ho et al. [51] investigated the molecular mechanisms underlying the shift from starch to lipid biosynthesis in Chlamydomonas sp. JSC4 and measured the upregulation in the expression of genes encoding enzymes for lipid synthesis Acetyl-CoA carboxylase (ACCase), pyruvate decarboxylase, acetyl-CoA-synthetase, acetaldehyde dehydrogenase, and genes involved in starch degradation (starch phosphorylases).

3.2. Overexpression of Gene/Enzymes Involved in the Lipid Biosynthesis Pathway

The main enzyme involved in the precursor formation and lipid synthesis are ACCase, ATP citrate lyase (ACL), glycerol-3-phosphate (GPAT), lysophosphatidic acid acyltransferase (LPAAT), phospholipid: diacylglycerols acyltransferase (PDAT), and acyl-Co-A: diacylglycerols acyltransferase (DGAT) [52,53,54]. The expression of these gene-encoding enzymes determines lipid content, and the regulation of these genes affects lipid content [55,56]. Additionally, enhanced lipid accumulation was documented due to the knockdown/overexpression of transcription factors that target the upregulation of lipid biosynthetic genes. Certain related work is also briefly summarized in Table 1. Acetyl-CoA carboxylase (ACCase) is a common and important enzyme responsible for the increase of the accumulation of lipid in microalgae. It is the first rate-limiting step in FA biosynthesis, which accelerates TAGs synthesis. In lipid synthesis, the conversion of Acetyl-CoA to malonyl-CoA is regulated by ACCase. The overexpression of the gene (ACCase) is one of the effectively applied techniques that can be used to improve FAs in microalgae. A previous study revealed that the overexpression of ACCase affected lipid accumulation less [57]. However, coordinated overexpression with the ACCase subunit (accD) of malic enzyme (ME) was able to raise the lipid productivity of Dunaliella salina [58]. Likewise, overexpression of the ME enhanced Phaeodactylum tricornutum lipid output by 2.5-fold without adversely affecting the growth rate [59]. Genes associated with the synthesis of lipids were knocked out and overexpressed in prior work to see how they affected lipid accumulation in microalgae. ACCase encoding the FA enzyme synthesis was first over-expressed in a diatom, Cyclotella cryptica, by Dunahay et al. in 1996 [60]. Chimeric plasmid vectors were used for random recombinant DNA integration, and copies of multiple genes were inserted [60]. Although ACCase overexpression resulted in a 2–3-fold increase in ACCase activity, it did not affect FA synthesis [61]. C. cryptica has been credited with considerable genetic potential for hyper-lipid production, as compared to related diatom Thalassiosira pseudonana. The genomic study revealed increased expression of critical lipogenesis genes and expansion of TAG biosynthesis enzymes significantly [62]. The initial report of ACCase overexpression in C. reinhardtii, performed through the insertion of an overexpression vector, demonstrated that this technique can result in higher ACCase activity with better FA production [63]. However, it has been shown that the upregulation of ACCase in conjunction with malic enzymes, which catalyze the conversion of malate to pyruvate, increases lipid accumulation in D. Salina [58]. In Phaeodactylum tricornutum (PtACC2) microalgae, ACCase modification raised ACCase activity by 3.3-fold and led to a 1.77-fold rise in lipids, reaching 40.8% dry biomass [64]. Diacylglycerol acyltransferase (DGAT) overexpression, which catalyzes the last stage of TAGs biosynthesis, led to lipid enhancement [37]. The expression of multifunctional enzymes phospholipase/lipase/acyltransferase was inhibited, increasing higher lipid storage without sacrificing the growth of T. pseudonana [65]. It is understood that transcriptional regulation can affect the metabolomics flow of the system because transcription factors in a metabolic pathway can mark numerous regulatory points. In N. gaditana, knocking down a single ZnCys transcriptional regulator caused a 2-fold increase in lipid content [66]. Moreover, silencing the cht7 gene, which encodes TAG lipase, resulted in a 10-fold rise in TAGs [67].

4. CRISPR/Cas9 (Clustered Regularly Interspaced Short Palindromic Repeats)

Recent advancements in gene-editing technologies, particularly CRISPR/Cas9, could result in gene alterations in commercially relevant microalgae strains (Figure 1). Genome editing involves modifying DNA in a sequence-specific manner via integrations, deletions, and insertions. In an experiment, gene manipulation was performed in C. vulgaris using the CRISPR/Cas9 method. For which a Cas9 fragment was constructed with engineered sgRNA in the omega-3 fatty acid desaturase (fad3) gene, resulting in a 46% (w/w) increase in lipid accumulation in the strain [68]. The Cas9 method was used to perform a target-specific knockout of the phospholipase A2 gene in Chlamydomonas reinhardtii. Subsequently, the mutants displayed an increased pool of diacylglycerol accompanied by a greater TAGs accumulation without extensively compensating the growth of the cells. Thus, the average lipid productivity of knockout mutants with phospholipase A2 increased up to 64.25% (80.92 g.L−1.d−1) [69]. CRISPR-Cas9 outperforms TALEN and ZFN because of its ease, adaptability, reduced price, and increased specificity [70]. Regardless of this, the most well-studied model of C. reinhardtii (freshwater green microalgae) had relatively little success [18,64,71]. In a recent study, a fragment of mGFP was transferred using Agrobacterium tumefaciens plasmid-mediated transfer to Chlorella vulgaris and C. sorokiniana FSP-E by electroporation, respectively. An increase of 67% fluorescence was observed against a wild-type strain by inverted fluorescence microscopy. Subsequently, a plasmid-containing Cas 9 fragment with sgRNA targeting the omega-3 fatty acid desaturase (fad3) gene was created. Higher lipid accumulation (46% w/w) in C. vulgaris was perceived and considered as the first successful gene manipulation in Chlorella [68].

5. Alteration of Fatty Acid Composition

Aside from the use of genetic engineering for lipid enhancement, it is also necessary to consider lipid quality in terms of aptness as a feedstock for fuel production. The length of the carbon chain and the degree of unsaturation (double bond) in acyl composition are equally responsible for the determination of biodiesel properties. Natural microalgae synthesize a broad range of fatty acids [72]. Biodiesel with shorter chain FAs (C10–C12) improves cold flow properties. Therefore, the isolation of genes encoding unique shorter-chain acyl-ACP thioesterases may be beneficial and of great importance for decreasing the chain length of fatty acyl groups in oil extracted from microalgae. The first positive example of such a gene alteration attempt was demonstrated in a diatom where two genes encoding the plant-based acyl-ACP thioesterase in P. tricornutum were over-expressed by researchers from the Colorado School of Mines (USA) to generate a medium FA chain in the oil fraction [73].

6. Role of Nanoparticles

6.1. In Lipid Induction

Nanotechnology is the science, engineering, and technology that is measured in nanoscale (100 nm or less) [74]. Various types of metallic nanoparticles (MNPs), ranging from 5 to 100 nm have been explored due to their diverse physical and chemical properties than the same metals on a macroscopic scale [75,76]. The exceptional physiochemical behavior of MNPs has allowed them to be approached in several ways, comprising the food industry, drug delivery, cosmetics, and synthesis of multifunctional biomaterials [77]. The ability of NPs to boost the gas–liquid mass transfer rate in fermentation is a relatively new application of NPs [78,79]. According to the concept, the presence of NPs boosts the mass transfer coefficient at the gas–liquid interface [80]. Hence, rising CO2 concentrations may alter growth rate and lipid stimulation in some microalgae. Few MNPs, such as Au, Ag, ZnO, CuO, Pd, Se, and FeO, are highly toxic to various organisms [81,82,83,84,85,86]. The toxic effect of NPs is also observed in microalgae, and it is connected with the reactive oxygen species (ROS) generation and the stimulation of oxidative stress, which is attained when the amount of NP reaches an effective level [82,83,84]. Some experts stated that when microalgae are encountered with sufficient doses of NPs, they can generate oxidative stress, and consequently, increase lipid synthesis in microalgae [77,87,88]. Recently, He et al. studied the impact of carbon nanotubes (CNT), α-Fe2O3 and MgO NPs were tested in Scenedesmus obliqus. It was noticed that exposure to 5 mgL−1 CNT, 5 mgL−1 α-Fe2O3, and 40 mgL−1 MgO NPs resulted in increased lipid up to 8.9%, 39.6%, and 18.5%, respectively. Moreover, when microalgae encountered high doses of NPs, a reduction in microalgae growth and lipid enhancement was observed, owing to the high amount of ROS that induced cell death [77]. Nanomaterials have several hundred times more surface area than their corresponding macroscale material weight. The surface area is not only significantly increased, but also the elasticity, persistence, strength, and electricity are improved. The utilization of nanomaterials could increase and potentially achieve lipid extraction efficiency without harming microalgae. In the transesterification process of lipids, nanomaterials, for example, CaO and MgO NPs, have been used as biocatalyst carriers or as heterogeneous catalysts [89]. Competitive inhibition, non-competitive inhibition, and denaturation are all ways that NPs might selectively inhibit enzyme activity [89] (Figure 2). As a result, fabricating NPs with specific characteristics to bind specific enzymes or proteins could allow for the regulation of their activity. To avoid non-specific binds and aid in recognition of specific enzymes or biomolecules, functionalizing NPs by altering their enormous surface area with organic molecules through covalent or non-covalent interactions has been suggested [90]. In this case, a substantial study is required to record and verify the molecular process of binding of a specific NP (with specified properties) with AGPase enzyme to facilitate, block, or inhibit the enzyme’s activity, hence inhibiting the starch manufacturing pathway [91]. This technique will aid in overcoming the lipid production bottleneck in microalgae, resulting in increased biofuel output.

6.2. Harvesting of Microalgae

Microalgae-based biodiesel manufacturing on a large scale is one way to address energy constraints [92]. The collection of microalgae cells is another obstacle that interrupts the commercial development of algae-based biodiesel. Thus, the establishment of an efficient method of harvesting is important for the urgent need to bring about a significant reduction of the operation cost. In the photobioreactor culture procedure, the magnetic NP powder is used in the microalgae cell suspension to flocculate the cells for the even dispersal of nutrients and light throughout the reactor (Figure 2). An additional well-known technique for enhancing cell suspension is immunomagnetic identification and manipulation of microalgae cells from NP. It is difficult to separate algae from a large volume of growing medium. Various harvesting techniques have been developed using an attached culture device, such as flotation, filtration, coagulation, flocculation, centrifugation, and scratching [93,94]. Although, because of their low concentration, these traditional harvesting processes do have some limitations. Therefore, creating an effective technology to extract small algae cells from highly diluted solutions remains a challenge [95]. The harvesting of Nannochloropsis sp. is more difficult due to its smaller diameter compared to other microalgae [96].
Another method, magnetic separation, has been now applied due to its benefits, such as easy process, energy efficiency, quick separation, and minimum operating costs [97]. Silica coating in magnetic NPs or cationic polyelectrolyte was tested to harvest marine and freshwater microalgae [95,98]. Chlorella ellipsoidea and Botryococcus braunii were collected from freshwater using bare Fe3O4 NPs [97]. The evaluation of the harvesting effectiveness of Nannochloropsis sp. using Fe3O4 NPs is of great importance in contrast to freshwater microalgae. Some essential process parameters for magnetic NPs for microalgal harvesting technologies, such as algal growth phase, harvesting temperature, and medium reusability, are yet unknown.

7. Pretreatment Methods

Pretreatment is an important step in restoring biomass composition for optimal biofuel production. It is crucial to disrupt the cell wall of microalgae to stimulate the release of inner substances, including lipids, proteins, carotenoids, and carbohydrates, into the medium [99]. Figure 3 depicts various pretreatment strategies for lipid extraction from microalgae. To solubilize the microalgae cell wall, a chemical pretreatment approach employs alkaline and acid with a heating range of 120–180 °C [100]. Whereas physical pressures (solid–liquid shear forces) are utilized in mechanical pretreatment to disturb the structure of cellulose by broadening the surface area of organic material in order to depolymerize the hemicellulose that includes the algal cell wall [101]. To evade the enzymatic hydrolysis stage, acid and alkaline are used to increase the breakdown of cellulose matrix, hemicellulose depolymerization, and starch hydrolysis. By lowering the starch crystallinity and size of starch polymers, this approach causes solvation and saponification processes, resulting in the creation of openings in the cell wall that promote the discharge of internal constituents [102].
Sert et al. [103] investigated the effects of concentration of the solution, duration of pretreatment, and temperature and discovered that 60 min of acid pretreatment (1 N H2SO4) at 100 °C yielded the highest bioethanol content (18.52%). This amount was three times higher than the alkaline pretreatment. The chemical pretreatment, on the other hand, is caustic, poisonous, and creates inhibitory chemicals that might bring contamination downstream [104]. The natural ability of enzymes and some microorganisms to break the constituent of the microalgae cell wall is exploited in enzyme-based treatment (biological pretreatment) [99]. By boosting the release of intracellular components, hydrolytic enzymes are used to hydrolyze the cell wall of microalgae, resulting in a quicker and more successful recovery [105]. Hydrolytic bacteria are also used in the pretreatment method; the algicidal capacity of these bacteria utilized for microalgae pretreatment is imperative. This is because the bacteria’s algicidal molecule causes autolysis in microalgae cells, resulting in the liberation of extracellular components [106]. Bai et al. [107] and Muoz et al. [108] stated that the pretreatment of C. vulgaris with Flammeovirga yaeyamensis and Bacillus thuringiensis increased the lipid recovery rate by 44.3% and 100%, respectively. Another study found that employing Bacillus licheniformis caused considerable cell wall breakdown in Chlorella sp. within 60 h [109]. In this condition, pretreatment with hydrolytic bacteria was found to be more successful than enzyme pretreatment, as enzymes drop their capability to function with time. The pure culture system, which is used to pretreat enzymes and bacteria, has several difficulties, including long pretreatment duration, pretreatment in open conditions, and the maintenance of pure culture [110]. In addition, to guarantee appropriate biomass degradation during the pretreatment process, microbial consortia should have cellulose and hemicellulose degradation capacity [111]. The microbial consortium’s synergistic metabolism caught some curiosity; thus, further investigation for bioprocessing technology is still needed.

8. Conclusions

Microalgae-based biodiesel is yet to be marketed, the reason being that the total cost of processing is twice that of fuels based on petroleum. For the growth of microalgae, culture maintenance, biomass production, lipid yield, extraction, and later conversion to biodiesel, each step needs high effort and strategy to get cost-effective biodiesel production compared to fossil fuels. The current review is focused on the promising metabolic engineering innovations that enable enhanced TAG production. The article outlines the genetic engineering methodologies, NPs applications, and several pretreatment methods that have been explored to increase TAGs synthesis in microalgae species, resulting in an economically viable energy production strategy. Where NP’s amendment triggers lipid production, on the other hand, pretreatment sustains the high lipid recovery rate. Thus, the concept of combining various technologies supporting biomass and lipid enhancement is a viable strategy for biodiesel production from microalgae.

Author Contributions

Writing—original draft preparation, review and editing, J.R.; Writing—review and editing, S.P., K.P., D.A. and M.P.; Conceptualization, supervision, project administration, V.P., V.K.T. and P.K.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research did not receive any specific grant from the funding agencies in the public, commercial, or not-for-profit sectors.

Institutional Review Board Statement

Not Applicable.

Informed Consent Statement

Not Applicable.

Data Availability Statement

Not Applicable.

Conflicts of Interest

The authors declare no competing interests with the work presented in the manuscript.

References

  1. Dębowski, M.; Michalski, R.; Zieliński, M.; Kazimierowicz, J. A Comparative Analysis of Emissions from a Compression–Ignition Engine Powered by Diesel, Rapeseed Biodiesel, and Biodiesel from Chlorella protothecoides Biomass Cultured under Different Conditions. Atmosphere 2021, 12, 1099. [Google Scholar] [CrossRef]
  2. Mwangi, J.K.; Lee, W.-J.; Tsai, J.-H.; Wu, T.S. Emission Reductions of Nitrogen Oxides, Particulate Matter and Polycyclic Aromatic Hydrocarbons by Using Microalgae Biodiesel, Butanol and Water in Diesel Engine. Aerosol Air Qual. Res. 2015, 15, 901–914. [Google Scholar] [CrossRef]
  3. Chew, K.W.; Yap, J.Y.; Show, P.L.; Suan, N.H.; Juan, J.C.; Ling, T.C.; Lee, D.-J.; Chang, J.-S. Microalgae biorefinery: High value products perspectives. Bioresour. Technol. 2017, 229, 53–62. [Google Scholar] [CrossRef] [PubMed]
  4. Khan, M.I.; Shin, J.H.; Kim, J.D.; Khan, M.I.; Shin, J.H.; Kim, J.D. The promising future of microalgae: Current status, challenges, and optimization of a sustainable and renewable industry for biofuels, feed, and other products. Microb. Cell Factories 2018, 17, 1–21. [Google Scholar] [CrossRef]
  5. Chandra, R.; Iqbal, H.M.N.; Vishal, G.; Lee, H.-S.; Nagra, S. Algal biorefinery: A sustainable approach to valorize algal-based biomass towards multiple product recovery. Bioresour. Technol. 2019, 278, 346–359. [Google Scholar] [CrossRef]
  6. Koyande, A.K.; Chew, K.W.; Rambabu, K.; Tao, Y.; Chu, D.T.; Show, P.L. Microalgae: A potential alternative to health supplementation for humans. Food Sci. Hum. Wellness 2019, 8, 16–24. [Google Scholar] [CrossRef]
  7. Dębowski, M.; Zieliński, M.; Kazimierowicz, J.; Kujawska, N.; Talbierz, S. Microalgae Cultivation Technologies as an Opportunity for Bioenergetic System Development—Advantages and Limitations. Sustainability 2020, 12, 9980. [Google Scholar] [CrossRef]
  8. Nayak, M.; Suh, W.I.; Chang, Y.K.; Lee, B. Exploration of two-stage cultivation strategies using nitrogen starvation to maximize the lipid productivity in Chlorella sp. HS2. Bioresour. Technol. 2019, 276, 110–118. [Google Scholar] [CrossRef]
  9. Shokravi, H.; Shokravi, Z.; Aziz, M.A.; Shokravi, H. The Fourth-Generation. In Fossil Free Fuels: Trends in Renewable Energy; Taylor & Francis Group: Oxford, UK, 2019; pp. 213–251. [Google Scholar]
  10. Arora, N.; Pienkos, P.T.; Pruthi, V.; Poluri, K.M.; Guarnieri, M.T. Leveraging algal omics to reveal potential targets for augmenting TAG accumulation. Biotechnol. Adv. 2018, 36, 1274–1292. [Google Scholar] [CrossRef]
  11. Hokravi, H.; Shokravi, Z.; Aziz, M.A.; Shokravi, H. Algal Biofuel: A Promising. In Fossil Free Fuels: Trends in Renewable Energy; Taylor & Francis Group: Oxford, UK, 2019; pp. 187–211. [Google Scholar]
  12. Chen, B.; Wan, C.; Mehmood, M.A.; Chang, J.-S.; Bai, F.; Zhao, X. Manipulating environmental stresses and stress tolerance of microalgae for enhanced production of lipids and value-added products—A review. Bioresour. Technol. 2017, 244, 1198–1206. [Google Scholar] [CrossRef]
  13. Ghosh, A.; Khanra, S.; Mondal, M.; Halder, G.; Tiwari, O.; Saini, S.; Bhowmick, T.K.; Gayen, K. Progress toward isolation of strains and genetically engineered strains of microalgae for production of biofuel and other value added chemicals: A review. Energy Convers. Manag. 2016, 113, 104–118. [Google Scholar] [CrossRef]
  14. Khan, I.; Saeed, K.; Khan, I. Nanoparticles: Properties, applications and toxicities. Arab. J. Chem. 2019, 12, 908–931. [Google Scholar] [CrossRef]
  15. Ördög, V.; Stirk, W.A.; Bálint, P.; Aremu, A.O.; Okem, A.; Lovász, C.; Molnár, Z.; van Staden, J. Effect of temperature and nitrogen concentration on lipid productivity and fatty acid composition in three Chlorella strains. Algal Res. 2016, 16, 141–149. [Google Scholar] [CrossRef]
  16. Gimpel, J.; A Specht, E.; Georgianna, D.R.; Mayfield, S.P. Advances in microalgae engineering and synthetic biology applications for biofuel production. Curr. Opin. Chem. Biol. 2013, 17, 489–495. [Google Scholar] [CrossRef]
  17. Jeon, S.; Lim, J.M.; Lee, H.G.; Shin, S.E.; Kang, N.K.; Park, Y.I.; Oh, H.M.; Jeong, W.J.; Jeong, B.; Chang, Y.K. Current status and perspectives of genome editing technology for microalgae. Biotechnol. Biofuels 2017, 10, 267. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Naduthodi, M.I.S.; Barbosa, M.J.; Van Der Oost, J. Progress of CRISPR-Cas Based Genome Editing in Photosynthetic Microbes. Biotechnol. J. 2018, 13, e1700591. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Tan, K.W.M.; Lee, Y.K. The dilemma for lipid productivity in green microalgae: Importance of substrate provision in improving oil yield without sacrificing growth. Biotechnol. Biofuels 2016, 9, 255. [Google Scholar] [CrossRef] [Green Version]
  20. Rathod, J.; Gade, R.M.; Rathod, D.R.; Dudhare, M. A Review on Molecular Tools of Microalgal Genetic Transformation and their Application for Overexpression of Different Genes. Int. J. Curr. Microbiol. Appl. Sci. 2017, 6, 3191–3207. [Google Scholar] [CrossRef]
  21. Beacham, T.A.; Sweet, J.B.; Allen, M.J. Large scale cultivation of genetically modified microalgae: A new era for environmental risk assessment. Algal Res. 2017, 25, 90–100. [Google Scholar] [CrossRef]
  22. Villarreal, J.V.; Burgués, C.; Rösch, C. Acceptability of genetically engineered algae biofuels in Europe: Opinions of experts and stakeholders. Biotechnol. Biofuels 2020, 13, 1–21. [Google Scholar] [CrossRef]
  23. Rengel, R.; Smith, R.T.; Haslam, R.; Sayanova, O.; Vila, M.; León, R. Overexpression of acetyl-CoA synthetase (ACS) enhances the biosynthesis of neutral lipids and starch in the green microalga Chlamydomonas reinhardtii. Algal Res. 2018, 31, 183–193. [Google Scholar] [CrossRef] [Green Version]
  24. Yang, B.; Liu, J.; Jiang, Y.; Chen, F. Chlorellaspecies as hosts for genetic engineering and expression of heterologous proteins: Progress, challenge and perspective. Biotechnol. J. 2016, 11, 1244–1261. [Google Scholar] [CrossRef] [PubMed]
  25. Wang, X.; Dong, H.-P.; Wei, W.; Balamurugan, S.; Yang, W.-D.; Liu, J.-S.; Li, H.-Y. Dual expression of plastidial GPAT1 and LPAT1 regulates triacylglycerol production and the fatty acid profile in Phaeodactylum tricornutum. Biotechnol. Biofuels 2018, 11, 1–14. [Google Scholar] [CrossRef] [PubMed]
  26. Yao, Y.; Lu, Y.; Peng, K.-T.; Huang, T.; Niu, Y.-F.; Xie, W.-H.; Yang, W.-D.; Liu, J.-S.; Li, H.-Y. Glycerol and neutral lipid production in the oleaginous marine diatom Phaeodactylum tricornutum promoted by overexpression of glycerol-3-phosphate dehydrogenase. Biotechnol. Biofuels 2014, 7, 110. [Google Scholar] [CrossRef] [Green Version]
  27. Xue, J.; Balamurugan, S.; Li, D.-W.; Liu, Y.-H.; Zeng, H.; Wang, L.; Yang, W.-D.; Liu, J.-S.; Li, H.-Y. Glucose-6-phosphate dehydrogenase as a target for highly efficient fatty acid biosynthesis in microalgae by enhancing NADPH supply. Metab. Eng. 2017, 41, 212–221. [Google Scholar] [CrossRef]
  28. Yan, J.; Kuang, Y.; Gui, X.; Han, X.; Yan, Y. Engineering a malic enzyme to enhance lipid accumulation in Chlorella protothecoides and direct production of biodiesel from the microalgal biomass. Biomass Bioenergy 2019, 122, 298–304. [Google Scholar] [CrossRef]
  29. Bajhaiya, A.; Dean, A.; Zeef, L.A.; Webster, R.E.; Pittman, J.K. PSR1 Is a Global Transcriptional Regulator of Phosphorus Deficiency Responses and Carbon Storage Metabolism in Chlamydomonas reinhardtii. Plant Physiol. 2016, 170, 1216–1234. [Google Scholar] [CrossRef] [Green Version]
  30. Kwon, S.; Kang, N.K.; Koh, H.G.; Shin, S.-E.; Lee, B.; Jeong, B.-R.; Chang, Y.K. Enhancement of biomass and lipid productivity by overexpression of a bZIP transcription factor in Nannochloropsis salina. Biotechnol. Bioeng. 2018, 115, 331–340. [Google Scholar] [CrossRef] [Green Version]
  31. Msanne, J.; Xu, D.; Konda, A.R.; Casas-Mollano, J.A.; Awada, T.; Cahoon, E.B.; Cerutti, H. Metabolic and gene expression changes triggered by nitrogen deprivation in the photoautotrophically grown microalgae Chlamydomonas reinhardtii and Coccomyxa sp. C-169. Phytochemistry 2012, 75, 50–59. [Google Scholar] [CrossRef] [Green Version]
  32. Hsieh, H.-J.; Su, C.-H.; Chien, L.-J. Accumulation of lipid production in Chlorella minutissima by triacylglycerol biosynthesis-related genes cloned from Saccharomyces cerevisiae and Yarrowia lipolytica. J. Microbiol. 2012, 50, 526–534. [Google Scholar] [CrossRef]
  33. Wei, H.; Shi, Y.; Ma, X.; Pan, Y.; Hu, H.; Li, Y.; Luo, M.; Gerken, H.; Liu, J. A type-I diacylglycerol acyltransferase modulates triacylglycerol biosynthesis and fatty acid composition in the oleaginous microalga, Nannochloropsis oceanica. Biotechnol. Biofuels 2017, 10, 1–18. [Google Scholar] [CrossRef] [PubMed]
  34. Fan, J.; Xu, H.; Li, Y. Transcriptome-based global analysis of gene expression in response to carbon dioxide deprivation in the green algae Chlorella pyrenoidosa. Algal Res. 2016, 16, 12–19. [Google Scholar] [CrossRef]
  35. Yamaoka, Y.; Achard, D.; Jang, S.; Legeret, B.; Kamisuki, S.; Ko, D.; Schulz-Raffelt, M.; Kim, Y.; Song, W.; Nishida, I.; et al. Identification of a Chlamydomonas plastidial 2-lysophosphatidic acid acyltransferase and its use to engineer microalgae with increased oil content. Plant Biotechnol. J. 2016, 14, 2158–2167. [Google Scholar] [CrossRef] [PubMed]
  36. Aratboni, H.A.; Rafiei, N.; Garcia-Granados, R.; Alemzadeh, A.; Morones-Ramírez, J.R. Biomass and lipid induction strategies in microalgae for biofuel production and other applications. Microb. Cell Fact. 2019, 18, 178. [Google Scholar] [CrossRef] [Green Version]
  37. Li, D.-W.; Cen, S.-Y.; Liu, Y.-H.; Balamurugan, S.; Zheng, X.-Y.; Alimujiang, A.; Yang, W.-D.; Liu, J.-S.; Li, H.-Y. A type 2 diacylglycerol acyltransferase accelerates the triacylglycerol biosynthesis in heterokont oleaginous microalga Nannochloropsis oceanica. J. Biotechnol. 2016, 229, 65–71. [Google Scholar] [CrossRef]
  38. Zhu, L.D.; Li, Z.H.; Hiltunen, E. Strategies for Lipid Production Improvement in Microalgae as a Biodiesel Feedstock. BioMed Res. Int. 2016, 2016, 8792548. [Google Scholar] [CrossRef] [Green Version]
  39. Li, Y.; Han, D.; Hu, G.; Dauvillée, D.; Sommerfeld, M.; Ball, S.; Hu, Q. Chlamydomonas starchless mutant defective in ADP-glucose pyrophosphorylase hyper-accumulates triacylglycerol. Metab. Eng. 2010, 12, 387–391. [Google Scholar] [CrossRef]
  40. Ahmad, I.; Sharma, A.; Daniell, H.; Kumar, S. Altered lipid composition and enhanced lipid production in green microalga by introduction of brassica diacylglycerol acyltransferase 2. Plant Biotechnol. J. 2015, 13, 540–550. [Google Scholar] [CrossRef] [Green Version]
  41. Posewitz, M.; King, P.; Smolinski, S.; Smith, R.D.; Ginley, A.; Ghirardi, M.; Seibert, M. Identification of genes required for hydrogenase activity in Chlamydomonas reinhardtii. Biochem. Soc. Trans. 2005, 33, 102–104. [Google Scholar] [CrossRef] [Green Version]
  42. Ball, S.G.; Deschamps, P. Starch Metabolism, The Chlamydomonas Sourcebook, Second Edition: Organellar and Metabolic Processes, Stern and Stern, 2nd ed.; Academic Press: Oxford, UK, 2009; pp. 1–40. ISBN 9780080919560. [Google Scholar]
  43. Wang, Z.T.; Ullrich, N.; Joo, S.; Waffenschmidt, S.; Goodenough, U. Algal lipid bodies: Stress induction, purification, and biochemical characterization in wild-type and starch-less Chlamydomonas reinhardtii. Eukaryot. Cell. 2009, 8, 1856–1868. [Google Scholar] [CrossRef] [Green Version]
  44. Ramazanov, A.; Ramazanov, Z. Isolation and characterization of a starchless mutant of Chlorella pyrenoidosa STL-PI with a high growth rate, and high protein and polyunsaturated fatty acid content. Phycol. Res. 2006, 54, 255–259. [Google Scholar] [CrossRef]
  45. Zhang, Y.-M.; Chen, H.; He, C.-L.; Wang, Q. Nitrogen Starvation Induced Oxidative Stress in an Oil-Producing Green Alga Chlorella sorokiniana C3. PLoS ONE 2013, 8, e69225. [Google Scholar] [CrossRef] [PubMed]
  46. Chen, H.; Zheng, Y.; Zhan, J.; He, C.; Wang, Q. Comparative metabolic profiling of the lipid-producing green microalga Chlorella reveals that nitrogen and carbon metabolic pathways contribute to lipid metabolism. Biotechnol. Biofuels 2017, 10, 1–20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Fan, J.; Yan, C.; Andre, C.; Shanklin, J.; Schwender, J.; Xu, C. Oil accumulation is controlled by carbon precursor supply for fatty acid synthesis in Chlamydomonas reinhardtii. Plant Cell Physiol. 2012, 53, 1380–1390. [Google Scholar] [CrossRef] [Green Version]
  48. Sirikhachornkit, A.; Vuttipongchaikij, S.; Suttangkakul, A.; Yokthongwattana, K.; Juntawong, P.; Pokethitiyook, P.; Kangvansaichol, K.; Meetam, M. Increasing the Triacylglycerol Content in Dunaliella tertiolecta through Isolation of Starch-Deficient Mutants. J. Microbiol. Biotechnol. 2016, 26, 854–866. [Google Scholar] [CrossRef]
  49. Li, Y.; Han, D.; Hu, G.; Sommerfeld, M.; Hu, Q. Inhibition of starch synthesis results in overproduction of lipids in Chlamydomonas reinhardtii. Biotechnol. Bioeng. 2010, 107, 258–268. [Google Scholar] [CrossRef]
  50. Work, V.; Radakovits, R.; Jinkerson, R.; Meuser, J.E.; Elliott, L.G.; Vinyard, D.; Laurens, L.M.L.; Dismukes, G.C.; Posewitz, M.C. Increased Lipid Accumulation in the Chlamydomonas reinhardtii sta7-10 Starchless Isoamylase Mutant and Increased Carbohydrate Synthesis in Complemented Strains. Eukaryot. Cell 2010, 9, 1251–1261. [Google Scholar] [CrossRef] [Green Version]
  51. Ho, S.H.; Nakanishi, A.; Kato, Y.; Yamasaki, H.; Chang, J.-S.; Misawa, N.; Hirose, Y.; Minagawa, J.; Hasunuma, T.; Kondo, A. Dynamic metabolic profiling together with transcription analysis reveals salinity-induced starch-to-lipid biosynthesis in alga Chlamydomonas sp. JSC4. Sci. Rep. 2017, 7, 45471. [Google Scholar] [CrossRef]
  52. Vieler, A.; Wu, G.; Tsai, C.-H.; Bullard, B.; Cornish, A.J.; Harvey, C.; Reca, I.-B.; Thornburg, C.; Achawanantakun, R.; Buehl, C.J.; et al. Genome, Functional Gene Annotation, and Nuclear Transformation of the Heterokont Oleaginous Alga Nannochloropsis oceanica CCMP1779. PLoS Genet. 2012, 8, e1003064. [Google Scholar] [CrossRef] [Green Version]
  53. Su, H.; Feng, J.; Lv, J.; Liu, Q.; Nan, F.; Liu, X.; Xie, S. Molecular mechanism of lipid accumulation and metabolism of oleaginous Chlorococcum sphacosum GD from soil under salt stress. Int. J. Mol. Sci. 2021, 22, 1304. [Google Scholar] [CrossRef]
  54. Li-Beisson, Y.; Thelen, J.J.; Fedosejevs, E.; Harwood, J.L. The lipid biochemistry of eukaryotic algae. Prog. Lipid Res. 2019, 74, 31–68. [Google Scholar] [CrossRef] [PubMed]
  55. Zienkiewicz, K.; Du, Z.-Y.; Ma, W.; Vollheyde, K.; Benning, C. Stress-induced neutral lipid biosynthesis in microalgae—Molecular, cellular and physiological insights. Biochim. Biophys. Acta BBA Mol. Cell Biol. Lipids 2016, 1861, 1269–1281. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Li-Beisson, Y.; Beisson, F.; Riekhof, W. Metabolism of acyl-lipids in Chlamydomonas reinhardtii. Plant J. 2015, 82, 504–522. [Google Scholar] [CrossRef] [PubMed]
  57. Chu, W.L. Strategies to enhance production of microalgal biomass and lipids for biofuel feedstock. Eur. J. Phycol. 2017, 52, 419–437. [Google Scholar] [CrossRef]
  58. Talebi, A.F.; Tohidfar, M.; Bagheri, A.; Lyon, S.R.; Salehi-Ashtiani, K.; Tabatabaei, M. Manipulation of carbon flux into fatty acid biosynthesis pathway in Dunaliella salina using AccD and ME genes to enhance lipid content and to improve produced biodiesel quality. Biofuel Res. J. 2014, 1, 91–97. [Google Scholar] [CrossRef]
  59. Xue, J.; Niu, Y.-F.; Huang, T.; Yang, W.-D.; Liu, J.-S.; Li, H.-Y. Genetic improvement of the microalga Phaeodactylum tricornutum for boosting neutral lipid accumulation. Metab. Eng. 2015, 27, 1–9. [Google Scholar] [CrossRef]
  60. Dunahay, T.G.; Jarvis, E.E.; Dais, S.S.; Roessler, P.G. Manipulation of Microalgal Lipid Production Using Genetic Engineering. Seventeenth Symp. Biotechnol. Fuels Chem. 1996, 57, 223–231. [Google Scholar] [CrossRef]
  61. Sheehan, J.; Dunahay, T.; Benemann, J.; Roessler, P. Look Back at the U.S. Department of Energy’s Aquatic Species Program: Biodiesel from Algae; Close-Out Report; National Renewable Energy Laboratory: Golden, CO, USA, 1998; Volume 328, pp. 1–294.
  62. Traller, J.C.; Cokus, S.J.; Lopez, D.A.; Gaidarenko, O.; Smith, S.R.; McCrow, J.P.; Gallaher, S.D.; Podell, S.; Thompson, M.; Cook, O.; et al. Genome and methylome of the oleaginous diatom Cyclotella cryptica reveal genetic flexibility toward a high lipid phenotype. Biotechnol. Biofuels 2016, 9, 1–20. [Google Scholar] [CrossRef] [Green Version]
  63. Chen, D.; Yuan, X.; Liang, L.; Liu, K.; Ye, H.; Liu, Z.; Liu, Y.; Huang, L.; He, W.; Chen, Y.; et al. Overexpression of acetyl-CoA carboxylase increases fatty acid production in the green alga Chlamydomonas reinhardtii. Biotechnol. Lett. 2019, 41, 1133–1145. [Google Scholar] [CrossRef]
  64. Li, D.W.; Xie, W.H.; Hao, T.B.; Cai, J.X.; Zhou, T.B.; Balamurugan, S.; Yang, W.D.; Liu, J.S.; Li, H.Y. Constitutive and Chloroplast Targeted Expression of Acetyl-CoA Carboxylase in Oleaginous Microalgae Elevates Fatty Acid Biosynthesis. Mar. Biotechnol. 2018, 20, 566–572. [Google Scholar] [CrossRef]
  65. Trentacoste, E.M.; Shrestha, R.P.; Smith, S.R.; Glé, C.; Hartmann, A.C.; Hildebrand, M.; Gerwick, W.H. Metabolic engineering of lipid catabolism increases microalgal lipid accumulation without compromising growth. Proc. Natl. Acad. Sci. USA 2013, 110, 19748–19753. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Ajjawi, I.; Verruto, J.; Aqui, M.; Soriaga, L.B.; Coppersmith, J.; Kwok, K.; Peach, L.; Orchard, E.; Kalb, R.; Xu, W.; et al. Lipid production in Nannochloropsis gaditana is doubled by decreasing expression of a single transcriptional regulator. Nat. Biotechnol. 2017, 35, 647–652. [Google Scholar] [CrossRef] [PubMed]
  67. Tsai, C.H.; Warakanont, J.; Takeuchi, T.; Sears, B.B.; Moellering, E.R.; Benning, C. The protein compromised hydrolysis of triacylglycerols 7 (CHT7) acts as a repressor of cellular quiescence in Chlamydomonas. Proc. Natl. Acad. Sci. USA 2014, 111, 15833–15838. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Lin, W.R.; Ng, I.S. Development of CRISPR/Cas9 system in Chlorella vulgaris FSP-E to enhance lipid accumulation. Enzym. Microb. Technol. 2020, 133, 109458. [Google Scholar] [CrossRef] [PubMed]
  69. Shin, Y.S.; Jeong, J.; Nguyen, T.H.T.; Kim, J.Y.H.; Jin, E.; Sim, S.J. Targeted knockout of phospholipase A2 to increase lipid productivity in Chlamydomonas reinhardtii for biodiesel production. Bioresour. Technol. 2019, 271, 368–374. [Google Scholar] [CrossRef]
  70. Shin, S.E.; Lim, J.M.; Koh, H.G.; Kim, E.K.; Kang, N.K.; Jeon, S.; Kwon, S.; Shin, W.-S.; Lee, B.; Hwangbo, K.; et al. CRISPR/Cas9-induced knockout and knock-in mutations in Chlamydomonas reinhardtii. Sci. Rep. 2016, 6, 27810. [Google Scholar] [CrossRef]
  71. Baek, K.; Lee, Y.; Nam, O.; Park, S.; Sim, S.J.; Jin, E. Introducing Dunaliella LIP promoter containing light-inducible motifs improves transgenic expression in Chlamydomonas reinhardtii. Biotechnol. J. 2016, 11, 384–392. [Google Scholar] [CrossRef]
  72. Harwood, J.L.; Guschina, I.A. The versatility of algae and their lipid metabolism. Biochimie 2009, 91, 679–684. [Google Scholar] [CrossRef]
  73. Radakovits, R.; Eduafo, P.M.; Posewitz, M.C. Genetic engineering of fatty acid chain length in Phaeodactylum tricornutum. Metab. Eng. 2011, 13, 89–95. [Google Scholar] [CrossRef]
  74. Richard, F.; Nigel, G.; James, M.Q.; Anthony, S. Nanoscience and Nanotechnol-Ogies: Opportunities and Uncertainties; The Royal Society and The Royal Academy of Engineering: London, UK, 2004; ISBN 0 85403 604 0. [Google Scholar]
  75. Dickinson, S.; Mientus, M.; Frey, D.; Amini-Hajibashi, A.; Ozturk, S.; Shaikh, F.; Sengupta, D.; El-Halwagi, M.M. A review of biodiesel production from microalgae. Clean Technol. Environ. Policy 2017, 19, 637–668. [Google Scholar] [CrossRef]
  76. Alishah, H.; Pourseyedi, S.; Ebrahimipour, S.Y.; Mahani, S.E.; Rafiei, N. Green synthesis of starch-mediated CuO nanoparticles: Preparation, characterization, antimicrobial activities and in vitro MTT assay against MCF-7 cell line. Rend. Lince 2017, 28, 65–71. [Google Scholar] [CrossRef]
  77. He, M.; Yan, Y.; Pei, F.; Wu, M.; Gebreluel, T.; Zou, S.; Wang, C. Improvement on lipid production by Scenedesmus obliquus triggered by low dose exposure to nanoparticles. Sci. Rep. 2017, 7, 1–12. [Google Scholar] [CrossRef] [PubMed]
  78. Kim, J.H.; Chung, Y.K. Copper nanoparticle-catalyzed borylation of alkyl bromides with an organodiboron compound. RSC Adv. 2014, 4, 39755–39758. [Google Scholar] [CrossRef]
  79. Zhu, H.; Shanks, B.H.; Choi, D.W.; Heindel, T.J. Effect of functionalized MCM41 nanoparticles on syngas fermentation. Biomass Bioenergy 2010, 34, 1624–1627. [Google Scholar] [CrossRef]
  80. Ruthiya, K.C. Mass Transfer and Hydrodynamics in Catalytic Slurry Reactors; Technische Universiteit Eindhoven: Eindhoven, The Netherlands, 2005. [Google Scholar] [CrossRef]
  81. Reichelt, K.; Hoffmann-Lücke, P.; Hartmann, B.; Weber, B.; Ley, J.; Krammer, G.; Swanepoel, K.; Engel, K.-H. Phytochemical characterization of South African bush tea (Athrixia phylicoides DC.). S. Afr. J. Bot. 2012, 83, 1–8. [Google Scholar] [CrossRef] [Green Version]
  82. Ji, J.; Long, Z.; Lin, D. Toxicity of oxide nanoparticles to the green algae Chlorella sp. Chem. Eng. J. 2011, 170, 525–530. [Google Scholar] [CrossRef]
  83. Aruoja, V.; Dubourguier, H.-C.; Kasemets, K.; Kahru, A. Toxicity of nanoparticles of CuO, ZnO and TiO2 to microalgae Pseudokirchneriella subcapitata. Sci. Total Environ. 2009, 407, 1461–1468. [Google Scholar] [CrossRef]
  84. Franklin, N.M.; Rogers, N.J.; Apte, S.C.; Batley, G.E.; Gadd, G.E.; Casey, P.S. Comparative Toxicity of Nanoparticulate ZnO, Bulk ZnO, and ZnCl2 to a Freshwater Microalga (Pseudokirchneriella subcapitata): The Importance of Particle Solubility. Environ. Sci. Technol. 2007, 41, 8484–8490. [Google Scholar] [CrossRef]
  85. Alishah, H.; Pourseyedi, S.; Mahani, S.E.; Ebrahimipour, S.Y. Extract-mediated synthesis of Ag@AgCl nanoparticles using Conium maculatum seeds: Characterization, antibacterial activity and cytotoxicity effect against MCF-7 cell line. RSC Adv. 2016, 6, 73197–73202. [Google Scholar] [CrossRef]
  86. Alishah, H.; Pour Seyedi, S.; Ebrahimipour, S.Y.; Mahani, S.E. A Green Approach for the Synthesis of Silver Nanoparticles Using Root Extract of Chelidonium majus: Characterization and Antibacterial Evaluation. J. Clust. Sci. 2016, 27, 421–429. [Google Scholar] [CrossRef]
  87. Kang, N.K.; Lee, B.; Choi, G.-G.; Moon, M.; Park, M.S.; Lim, J.; Yang, J.-W. Enhancing lipid productivity of Chlorella vulgaris using oxidative stress by TiO2 nanoparticles. Korean J. Chem. Eng. 2014, 31, 861–867. [Google Scholar] [CrossRef]
  88. Hu, Q.; Sommerfeld, M.; Jarvis, E.; Ghirardi, M.; Posewitz, M.; Seibert, M.; Darzins, A. Microalgal triacylglycerols as feedstocks for biofuel production: Perspectives and advances. Plant J. 2008, 54, 621–639. [Google Scholar] [CrossRef] [PubMed]
  89. Chouhan, A.S.; Sarma, A. Modern heterogeneous catalysts for biodiesel production: A comprehensive review. Renew. Sustain. Energy Rev. 2011, 15, 4378–4399. [Google Scholar] [CrossRef]
  90. Wu, Z.; Zhang, B.; Yan, B. Regulation of Enzyme Activity through Interactions with Nanoparticles. Int. J. Mol. Sci. 2009, 10, 4198–4209. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Sarkar, R.D.; Singh, H.B.; Kalita, M.C. Enhanced lipid accumulation in microalgae through nanoparticle-mediated approach, for biodiesel production: A mini-review. Heliyon 2021, 7, 08057. [Google Scholar] [CrossRef] [PubMed]
  92. Gao, F.; Peng, Y.-Y.; Li, C.; Yang, G.-J.; Deng, Y.-B.; Xue, B.; Guo, Y.-M. Simultaneous nutrient removal and biomass/lipid production by Chlorella sp. in seafood processing wastewater. Sci. Total Environ. 2018, 640-641, 943–953. [Google Scholar] [CrossRef]
  93. Zheng, H.; Gao, Z.; Yin, J.; Tang, X.; Ji, X.-J.; Huang, H. Harvesting of microalgae by flocculation with poly (γ-glutamic acid). Bioresour. Technol. 2012, 112, 212–220. [Google Scholar] [CrossRef]
  94. Johnson, M.B.; Wen, Z. Development of an attached microalgal growth system for biofuel production. Appl. Microbiol. Biotechnol. 2010, 85, 525–534. [Google Scholar] [CrossRef]
  95. Cerff, M.; Morweiser, M.; Dillschneider, R.; Michel, A.; Menzel, K.; Posten, C. Harvesting fresh water and marine algae by magnetic separation: Screening of separation parameters and high gradient magnetic filtration. Bioresour. Technol. 2012, 118, 289–295. [Google Scholar] [CrossRef] [PubMed]
  96. Ríos, S.D.; Salvadó, J.; Farriol, X.; Torras, C. Antifouling microfiltration strategies to harvest microalgae for biofuel. Bioresour. Technol. 2012, 119, 406–418. [Google Scholar] [CrossRef] [PubMed]
  97. Xu, L.; Guo, C.; Wang, F.; Zheng, S.; Liu, C.-Z. A simple and rapid harvesting method for microalgae by in situ magnetic separation. Bioresour. Technol. 2011, 102, 10047–10051. [Google Scholar] [CrossRef] [PubMed]
  98. Lim, J.K.; Chieh, D.C.J.; Jalak, S.A.; Toh, P.Y.; Yasin, N.H.M.; Ng, B.W.; Ahmad, A.L. Rapid Magnetophoretic Separation of Microalgae. Small 2012, 8, 1683–1692. [Google Scholar] [CrossRef]
  99. Galbe, M.; Wallberg, O. Pretreatment for biorefineries: A review of common methods for efficient utilisation of lignocellulosic materials. Biotechnol. Biofuels 2019, 12, 1–26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Khoo, C.G.; Dasan, Y.K.; Lam, M.K.; Lee, K.T. Algae biorefinery: Review on a broad spectrum of downstream processes and products. Bioresour. Technol. 2019, 292, 121964. [Google Scholar] [CrossRef] [PubMed]
  101. Dahunsi, S. Mechanical pretreatment of lignocelluloses for enhanced biogas production: Methane yield prediction from biomass structural components. Bioresour. Technol. 2019, 280, 18–26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Sui, Z.; Gizaw, Y.; BeMiller, J.N. Extraction of polysaccharides from a species of Chlorella. Carbohydr. Polym. 2012, 90, 1–7. [Google Scholar] [CrossRef] [PubMed]
  103. Sert, B.Ş.; Inan, B.; Ozcimen, D. Effect of Chemical Pre-treatments on Bioethanol Production from Chlorella minutissima. Acta Chim. Slov. 2018, 65, 160–165. [Google Scholar] [CrossRef] [Green Version]
  104. Khoo, K.S.; Chew, K.W.; Yew, G.Y.; Leong, W.H.; Chai, Y.H.; Show, P.L.; Chen, W.-H. Recent advances in downstream processing of microalgae lipid recovery for biofuel production. Bioresour. Technol. 2020, 304, 122996. [Google Scholar] [CrossRef]
  105. Kucharska, K.; Rybarczyk, P.; Hołowacz, I.; Łukajtis, R.; Glinka, M.; Kamiński, M. Pretreatment of lignocellulosic materials as substrates for fermentation processes. Molecules 2018, 23, 2937. [Google Scholar] [CrossRef] [Green Version]
  106. Demuez, M.; González-Fernández, C.; Ballesteros, M. Algicidal microorganisms and secreted algicides: New tools to induce microalgal cell disruption. Biotechnol. Adv. 2015, 33, 1615–1625. [Google Scholar] [CrossRef]
  107. Bai, M.-D.; Chen, C.-Y.; Lu, W.-C.; Wan, H.-P.; Ho, S.-H.; Chang, J.-S. Enhancing the oil extraction efficiency of Chlorella vulgaris with cell-disruptive pretreatment using active extracellular substances from Bacillus thuringiensis ITRI-G1. Biochem. Eng. J. 2015, 101, 185–190. [Google Scholar] [CrossRef]
  108. Muñoz, C.; Hidalgo, C.; Zapata, M.; Jeison, D.; Riquelme, C.; Rivas, M. Use of Cellulolytic Marine Bacteria for Enzymatic Pretreatment in Microalgal Biogas Production. Appl. Environ. Microbiol. 2014, 80, 4199–4206. [Google Scholar] [CrossRef] [Green Version]
  109. He, S.; Fan, X.; Katukuri, N.R.; Yuan, X.; Wang, F.; Guo, R.-B. Enhanced methane production from microalgal biomass by anaerobic bio-pretreatment. Bioresour. Technol. 2016, 204, 145–151. [Google Scholar] [CrossRef]
  110. Lin, Z.; Li, S.-J.; Sun, F.; Ba, D.-C.; Li, X.-C. Surface characteristics of a dental implant modified by low energy oxygen ion implantation. Surf. Coat. Technol. 2019, 365, 208–213. [Google Scholar] [CrossRef]
  111. Mishra, V.; Jana, A.K.; Jana, M.M.; Gupta, A. Improvement of selective lignin degradation in fungal pretreatment of sweet sorghum bagasse using synergistic CuSO4-syringic acid supplements. J. Environ. Manag. 2017, 193, 558–566. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Genetically modified microalgae improve lipid content.
Figure 1. Genetically modified microalgae improve lipid content.
Energies 15 01550 g001
Figure 2. Nanoparticles trigger lipid synthesis and harvesting efficiency.
Figure 2. Nanoparticles trigger lipid synthesis and harvesting efficiency.
Energies 15 01550 g002
Figure 3. Different pretreatment methods for high lipid recovery.
Figure 3. Different pretreatment methods for high lipid recovery.
Energies 15 01550 g003
Table 1. Overexpression of genes/enzymes resulted in lipid enhancement in microalgae species.
Table 1. Overexpression of genes/enzymes resulted in lipid enhancement in microalgae species.
S. No.Microalgae SpeciesGenes/EnzymesLipid EnhancementReferences
1.Chlamydomonas reinhardtiiACCase Overexpression2.4-fold increase in TAGs[23]
2.PhaeodactylumtricornutumG6PD Overexpression55.7% increase in lipid[24]
3.P. tricornutumGPAT1; LPAT1 Overexpression2.3-fold increase in TAGs in N-depletion[25]
4.P. tricornutumG3PDH Overexpression1.9-fold increase in neutral lipid with slight decline in growth[26]
5.P. tricornutumG6PD Overexpression2.7-fold increase in lipid content[27]
6.Chlorella protothecoidesME Overexpression2.8-fold increase in total lipid content[28]
7.C. reinhardtiiPSR1 OverexpressionIncrease in starch granules, decrease in neutral lipid content[29]
8.Nannochloropsis salinabZIP OverexpressionImprovement in growth and lipid [30]
9.C. reinhardtiiDGTA Overexpression Enhanced saturated fatty acids[31]
10.Chlorella minutissimaGPAT; LPAAT; DGAT Overexpression2-fold increase in lipid content[32]
11.N. oceanicaNoDGAT1A Overexpression2.4-fold increase in TAGs accumulation[33]
12.C. pyrenoidosaNAD(H) kinase Overexpression 1.6 times increase in lipid content[34]
13.C. reinhardtiiLPAAT Overexpression 20% increase in TAGs[35]
14.T. pseudonanaKnock-down of a multifunctional lipase/phospholipase/acetyltransferase enzyme2.4–3.3-fold higher lipids in contrast to wild-type[36]
15.Nannochloropsis oceanicaDGAT Overexpression69% increase in total lipids[37]
GPAT: Glycerol-3-phosphate aceyltransferase; LPAT: Lysophosphatidic aceyltransferase; DGAT: Diacylglycerol aceyltransferase; N: Nitrogen; ME: Malic enzyme; ACCase: Acetyl-CoA carboxylase; LPAAT: Lysophosphatidic acid acyltransferase; G3PDH: Glyceraldehyde-3-phosphate dehydrogenase; G6PD: Glucose-6-phosphate dehydrogenase and DGAT: acyl-Co-A: diacylglycerols acyltransferase.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rawat, J.; Gupta, P.K.; Pandit, S.; Priya, K.; Agarwal, D.; Pant, M.; Thakur, V.K.; Pande, V. Latest Expansions in Lipid Enhancement of Microalgae for Biodiesel Production: An Update. Energies 2022, 15, 1550. https://doi.org/10.3390/en15041550

AMA Style

Rawat J, Gupta PK, Pandit S, Priya K, Agarwal D, Pant M, Thakur VK, Pande V. Latest Expansions in Lipid Enhancement of Microalgae for Biodiesel Production: An Update. Energies. 2022; 15(4):1550. https://doi.org/10.3390/en15041550

Chicago/Turabian Style

Rawat, Jyoti, Piyush Kumar Gupta, Soumya Pandit, Kanu Priya, Daksh Agarwal, Manu Pant, Vijay Kumar Thakur, and Veena Pande. 2022. "Latest Expansions in Lipid Enhancement of Microalgae for Biodiesel Production: An Update" Energies 15, no. 4: 1550. https://doi.org/10.3390/en15041550

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

Article Metrics

Back to TopTop