NC State
BioResources
Baala Harini, A., and Rajkumar, R. (2022). "Development of sustainable bioproducts from microalgae biomass: Current status and future perspectives," BioResources 17(4), 7285-7312.

Abstract

Population and pollution make notable contributions to introducing novel sophisticated techniques. From vehicles to industries, the release of CO2 into the atmosphere and wastewater into the running water streams are key concerns. On the other hand, the population is responsible for the rapid manufacturing of all commercial goods. Microalgae are the only answer accessible for the aforementioned difficulties. Similar to plants, microalgae need CO2 and light to thrive and produce a variety of bioproducts such as carbohydrates, protein, lipids, vitamins, sterols, pigments, and silica. Physical (light, temperature, CO2, and UV), chemical (nutrient addition or depletion), enzymatic, and metabolic pathway reconfiguration, as well as indoor or outdoor growing, are highly regarded among the several optimization strategies to make desired products. Wastewater pollution is rectified by growing microalgae in nutrient-rich organic water for their growth, which is used to accelerate bioproducts. This review considers the use of bioproducts in food, animal and aquatic feed, fertilizer, biofuel, medicinal and nutraceutical sectors. This paper also provides different optimization strategies, which include physical and chemical means of extraction methods for enhancing bioactive products. Challenges and future recommendations for enhancing target bioproducts are discussed to overcome environmental issues.


Download PDF

Full Article

Development of Sustainable Bioproducts from Microalgae Biomass: Current Status and Future Perspectives

Anandapadmanaban Baala Harini and Renganathan Rajkumar *

Population and pollution make notable contributions to introducing novel sophisticated techniques. From vehicles to industries, the release of CO2 into the atmosphere and wastewater into the running water streams are key concerns. On the other hand, the population is responsible for the rapid manufacturing of all commercial goods. Microalgae are the only answer accessible for the aforementioned difficulties. Similar to plants, microalgae need CO2 and light to thrive and produce a variety of bioproducts such as carbohydrates, protein, lipids, vitamins, sterols, pigments, and silica. Physical (light, temperature, CO2, and UV), chemical (nutrient addition or depletion), enzymatic, and metabolic pathway reconfiguration, as well as indoor or outdoor growing, are highly regarded among the several optimization strategies to make desired products. Wastewater pollution is rectified by growing microalgae in nutrient-rich organic water for their growth, which is used to accelerate bioproducts. This review considers the use of bioproducts in food, animal and aquatic feed, fertilizer, biofuel, medicinal and nutraceutical sectors. This paper also provides different optimization strategies, which include physical and chemical means of extraction methods for enhancing bioactive products. Challenges and future recommendations for enhancing target bioproducts are discussed to overcome environmental issues.

DOI: 10.15376/biores.17.4.Harini

Keywords: Microalgae; Biomass; Bioproducts; Genetic engineering; Wastewater

Contact information: Department of Environmental Sciences, School of Life Sciences, Bharathiar University, Coimbatore-641 046, Tamil Nadu, India;

* Corresponding author: micro_rajkumar@yahoo.co.in

INTRODUCTION

Population growth increases the demand for natural resources, which are employed to produce food, energy, and chemicals. Consequently, it greatly enhances the greenhouse gas emissions in the atmosphere, increasing global temperature (Hussain et al. 2021). However, the existing system of producing food and supporting products that cannot meet the current needs of humans. To combat the shortage of the existing and future world, it is imperative to obtain various alternatives or technological innovations that could ramp up production (Rahul et al. 2020). Microalgae is an appropriate answer to the aforementioned challenges because of its major advantages. It has a doubling time on average of 26 h, high productivity, and the capacity to withstand food and feed competition (Kawamura et al. 2021). Microalgae is ecologically susceptible to manipulation, grows in non-arable land, requires exceptionally small spaces for farming, harnesses greenhouse gases for its growth, and is able to grow in wastewater with minimal nutrients (Odjadjare et al. 2015).

Microalgae are broadly classified into Chrorophyceae (green algae), Cyanophyceae (blue-green algae), Xanthophyceae (yellow green algae), Chrysophyceae (golden algae), Bacillariophyceae (diatom), and others (Andrade et al. 2018; Pirbazari et al. 2019; Yu et al. 2017) on the basis of pigments. Microalgae are a complex range of marine and freshwater eukaryotes, including unicellular to colonial forms of tiny algae with sizes ranging from a few micrometers to a few millimeters (Sarpal et al. 2015). They have tremendous ecological plasticity in adapting to harsh conditions, such as high temperature, salinity, light, pH, moisture, and nutrients. Similar to plants, microalgae function as a primary producer reliant on the abundantly available light source and assimilate atmospheric carbon dioxide (CO2) to synthesize carbohydrates (18 to 46%), lipids (12 to 48%), proteins (18 to 46%), and other bioproducts (Tibbetts et al. 2014; Williams et al. 2019). Its metabolic process responds to the removal of 20% of CO2 from the atmosphere and 40% of CO2 from the ocean. According to assessments, about 1.83 kg of CO2 was used to produce 1 kg of algal biomass (Kumar et al. 2011; Li et al. 2013). Microalgae can synthesize a variety of goods including food, animal and aquatic feed, fertilizers, nutraceuticals, pharmaceuticals, cosmetics, and alternative bioenergy products (Khan et al. 2018). One of the main unprecedented increases in demand is fuel, as fossil fuels are nearly expended. To meet customer demands, crude oil is imported from numerous nations. It also increases the processing costs of producing gasoline and diesel substantially. Microalgae are commonly cultivated for lipid production, which accounts for almost 40% of their total biomass (Raja et al. 2018). The generation of a high proportion of lipids by microalgae paved the path for the centralization and restructuring of biodiesel synthesis via transesterification. The biodiesel produced from algal resources can be blended with conventional diesel or gasoline at various proposition to run motor vehicles, thereby reducing the consumption and cost of traditional fuel (Melvelle 2012).

The study of genetic architecture also stimulates the production of novel microalgal products via insertion, deletion, and translocation of genes. This genetic method of microalgae can help to accumulate the target products to improve economic feasibility (Fu et al. 2017). Although there have been many studies reporting on bioactive products, understanding their low-cost production at a commercial scale still needs extensive work. To overcome these issues, microalgae have numerous advantages when compared to other conventional resources (Odjadjare et al. 2015). Studies show that microalgae effectively utilize wastewater as a low-cost nutrient source for producing biomass and its target products. Notwithstanding the massive potential applications of microalgae, their improvement is hampered by a myriad of challenges. Therefore, independent research and novel low-cost technologies should be implemented for the enhancement of algal biorefinery on the commercial scale (Mehrabadi et al. 2016).

This review article describes the latest technological approaches in microalgal biorefinery for the production of high-value-added bioproducts. In addition, this study also addresses the current challenges and future work on wastewater treatment using microalgae and genetic approaches for enhanced production of various bioproducts.

BIOACTIVE PRODUCTS FROM MICROALGAE BIOMASS

Microalgae are capable of creating numerous bioactive products that act as feedstock for different products (Chew et al. 2017). Carbohydrates, proteins, lipids, cellulose, silica, pigments, vitamins, and sterols are essential for a healthy human diet and the treatment of numerous ailments. The residual goods encompass food for humans, biofuels, animal and aquatic feed, cosmetics, nutraceuticals, and pharmaceuticals (Mendoza et al. 2013; Prussi et al. 2014). Deep seawater (DSW) is used as a novel method for microalgae cultivation for fisheries, aquaculture, and medicine. Plant-growth-promoting bacteria (PGPB), when co-cultured, form a symbiotic relationship with the microalgae for better biomass production (Tan et al. 2015). The entire product will be renewable and harmless compared to chemically synthesized products. Different bioproducts produced from the microalgae are illustrated in Fig. 1.

Fig. 1. Illustration of various cultivation conditions used in microalgae to improve diverse bioproducts

Carbohydrates

Microalgal biomass contains a high amount of carbohydrates derived from the photosynthetic process. Carbohydrates are either located in the cell wall as structural components or in the plastids as reserve materials (Chen et al. 2013). Most of the microalgae possess starch as a reserve material, while cyanobacteria synthesize glycogen and sucrose as storage products (Gonzalez-Fernandez and Ballesteros 2012; Markou et al. 2013). The carbohydrate content in the microalgal cell may vary from species to species. Nevertheless, the composition of carbohydrates could be increased by various environmental stress conditions (Domozych et al. 2012).

Carbohydrates from microalgae are utilized for the development and manufacturing of biofuels including bioethanol, biobutanol, biomethane, and biohydrogen (Markou and Georgakakis 2012) (Table 1). These biofuels are employed in running high-performance with low-emission engines (Simas-Rodrigues et al. 2015; Quader and Ahmed 2017). Examples of fuels, such as biobutanol, are reported to be produced by Chlorella vulgaris and Spirulina platensis approximately 21% and 71% via fermentation, respectively (Tan et al. 2020). Synechocystis sp. produced bioethanol at a density of 0.186 g/g (Ashokkumar et al. 2019). Further, the production of bioethanol in Chlorella minutissima was achieved by increasing carbohydrate up to 60.5% (Menestrino et al. 2020). Scenedesmus obliquus produced biohydrogen at a volume of 68.9 mL/g (Singh et al. 2019). Cuellar-Bermudez et al. (2019) proved that Pseudanabaena sp. yielded 25.1 mL/g of biomethane. Other possibilities include food thickeners, painkillers, biodegradable materials, and functional foods (De Souza et al. 2020).

Table 1. Several Optimization Strategies to Manufacture a High Amount of Carbohydrates to Extract Numerous Bioproducts

Polysaccharides

Polysaccharides are also called carbohydrate polymers. They have intricate structures that differ (structurally and biochemically) among the different species of microorganisms. Xylose, glucose, galactose, mannose, and rhamnose are the major components in microalgal polysaccharides (Yi et al. 2021; Chanda et al. 2019; Bernaerts et al. 2018). In microalgae, polysaccharides are primarily formed as part of the cell wall (as structural polymers), involved in various metabolic functions (as energy storage polymers) (Yi et al. 2021; Markou and Georgakakis 2012), and also in cellular interaction and protection (as exopolysaccharide) (Morais et al. 2022; Prybylski et al. 2020).

Microalgae are able to synthesize polysaccharides under various stress conditions (Parwani et al. 2021; Colusse et al. 2021) such as temperature, light, salinity, and nutrient uptake etc. (Costa et al. 2021). Hence, standardization of these cultural conditions for microalgae cultivation are imperative to increase the production of polysaccharide (Colusse et al. 2021). Earlier reports showed that white light was more suitable for the production of polysaccharides at levels of about 0.10 and 0.14 g/L in Porphyridium sordidum and Porphyridium purpureum, respectively (Medina-Cabrera et al. 2020). Further, Chlorella vulgaris can accumulate about 32.7% of polysaccharides at a light intensity of 65 µmol m−2 s−1 at 28 °C (Gui et al. 2019). About 65 µmol m−2 s−1 of light was supplied in nutrient media with 1%w/v of glucose helps to produce 1.46 g/L of polysaccharide in Chlorella sp. (Cheirsilp et al. 2016). In addition to that, the nutrient media supplemented with high salinity (40g/L) was supported the polysaccharide production in Spirulina sp. by 1.02g/g of biomass (Chentir et al. 2017).

Polysaccharides derived from Anabaena sp. CCC 745 have a significant antioxidant and scavenging activity in food industries (Tiwari et al. 2019). It has been reported that polysaccharides from Nostoc sp., Phormidium sp. and Scytonema arcangeli can be used as soil-fixing agent in the agricultural fields (Park et al. 2017). On the other hand, polysaccharides of Spirulina platensis were used as feed for zebra fish growth and development (Rajasekar et al. 2019). Utilizing nanotechnology to create polysaccharide-based goods can be applied in the food, health and beauty industries (Morais et al. 2022). Moreover, the polysaccharides are being used in the biomedical field for antithrombotic, immunomodulatory, antitumor, anticoagulant, anti-inflammatory, antimutagenic, antiviral, and antioxidant activities, as reported by many authors (Xu et al. 2017; Moreira et al. 2022; Patil et al. 2018; Li et al. 2019).

Cellulose

Cellulose is the most abundant sustainable source on Earth. It has enormous potential for producing renewable fuels, bioplastics, and nanomaterials. Cellulose is a linear homopolymer consisting of repeating β-d-glucopyranosyl units connected by 1–4 glycosidic linkages in a diversified arrangement depending on the presence of crystallites and disordered amorphous regions. Cellulose is the major component in the cell wall of plants and algae (Popper et al. 2011). In general, the ratio of cellulose Iα/Iβ in algae was found to be 60/40, whereas in plants it was 25/75. Until now, four totally different sub-polymorphs of cellulose such as, cellulose (I), cellulose (II), cellulose (III), and cellulose (IV) have been reported. Hydrogen bond arrangements and polarity in constituting chains are varied between these sub-polymorphs. The most common form of crystalline cellulose (I) was observed in nature, whereas cellulose (II), (III), and (IV) have been synthesized by thermal and chemical treatments. Based on the chain alignment and interchain hydrogen bonding, cellulose allomorphs have distinctive geometries (Zanchetta et al. 2021). Regarding micro- and macroalgae, it was reported that microalgae have remarkable flexibility in terms of their mode of cultivation for cellulose production. However, only limited studies are available for increasing the cellulose content in microalgae biomass, which may be the focus of more attention in the future.

From a commercial standpoint, producing higher biomass in a short period of time is essential for the overall yield of cellulose. For instance, Chlorella is reported as a fast-growing microalga and consists of a fibrillar layer of polysaccharides that can be secured in certain cases through a resistant algaenan outer layer (Domozych et al. 2012; Kroger et al. 2018). Another report shows that the cell wall of Scenedesmus quadricauda is a trilaminar arrangement in which cellulosic and pectic layers can be distinguished and secured by an algaenan layer (Nemcova 2003). Further, the cell wall of Oedogonium bharuchae has two layers that are a mix of cellulose, pectins, and glycoproteins followed by a cellulose-free layer having both extensin and arabinogalactan proteins (Estevez et al. 2008). Similarly, the cell wall of Nannochloropsis has a special porous inner layer mixed with struts connecting the cell membrane to a porous cellulose-dependent layer covered by an algaenan outer layer (Scholz et al. 2014).

Furthermore, Nannochloropsis sp. is a well-studied microalga that is known for its great potential in cellulose extraction and production (Hamed et al. 2016; Lee et al. 2018). However, there is no sufficient information about other microalgal species, and hence, more attention is required to identify potential strains for cellulose production. A high concentration of cellulose was reported in Chlorella sp., Oocystis sp. (Aguirre and Bassi 2013; Kroger et al. 2018), Scenedesmus sp., Coelastrella sp., Chlorococcum sp., Selenastrum sp. (Yap et al. 2016; Kroger et al. 2018), Chaetosphaeridium sp., and Staurastrum sp., which garners increased interest for further study in detail. Table 2 illustrates the quantification of cellulose content in various microalgal strains. From these, the highest cellulose content (75 wt% dry weight basis) was observed in Nannochloropsis gaditana in their cell wall (Scholz et al. 2014). In contrast, low cellulose content (15.4 wt%) was reported in Chlorella pyrenoidosa (Northcote et al. 1960). The variation of cellulose content from species to species can be influenced by many cultural conditions. Currently, genetic engineering technologies are also opening the way to enhance biomass yields with desired products. It was proposed that the biorefinery approach can be used for the successful production of cellulose (Lee et al. 2018).

Table 2. Cellulose Content of the Microalgal Feedstock

Proteins

Proteins are produced from a variety of green vegetables and animal meat, but diatoms can generate them in greater quantities, which allows them to be sold competitively as tablets and pills. For instance, Tetraselmis suecica can produce a protein of approximately 12% using bead milling. In general, microalgae generated a protein with outstanding emulsification, frothing, and gelatin properties, a noteworthy breakthrough for the food sector (Garcia et al. 2018). Chlorella thermophile exhibited an increase of 36% in its protein content and was studied using artificial neural networks-genetic algorithm (ANN-GA) (Sarkar et al. 2022). In addition, Arthrospira platensis was discovered to account for 110% of the protein production from sugarcane bagasse in solid-state cultivation (Pelizer et al. 2015). Microalgal proteins are used in the cosmetic, medicinal, and animal/aquaculture feed industries (Caporgno and Mathys 2018). Chlorella has the capability of producing 62% protein content for cosmetics and nutraceutical production through enzyme hydrolysis (Pekkoh et al. 2021). Microalgae such as Botryococcus braunii, Chlorella protothecoides, Chlorella vulgaris, Neochloris oleoabundans, and Scenedesmus acutus can synthesize protein as a nutrition and dietary supplement of approximately 40%, 36.1%, 48.6%, 49.5%, and 53.2% protein content, respectively, using photobioreactors (PBRs) (Tibbetts et al. 2015; Baldisserotto et al. 2022). In comparison to another optimization method, enzyme hydrolysis helps to extract the maximum amount of protein present in the microalgal cell.

Lipids and Their Derivatives: Eicosapentaenoic Acid, Docosahexaenoic Acid, and Isoprenoids

Lipid productivity in microalgae can be classified into two parts: (i) 14 to 19 carbon atom chains used for the generation of biodiesel because of the unavailability of double bonds in the chain; and (ii) poly-unsaturated fatty acids (PUFAs) (Surendhiran et al. 2015). Furthermore, δ-3 fatty acids have many health benefits and can be found in a variety of food products that act against arthritis, asthma, cancer, cardiovascular disease (CVD), inflammatory disorders, depression, and schizophrenia by preventing the production of oxidative stress as antioxidants (Adarme-Vega et al. 2012). Lipids are also used as nutritional supplements and feed, antioxidants in beverages and functional foods, pills, capsules, and food additives in candies, gum, pasta, etc. (Chen et al. 2014).

The PUFAs, such as docosahexaenoic acid (DHA) and eicosapentaenoic acid (EPA), are necessary for the development of the human brain and the prevention of coronary heart disease (CHD). Additionally, because of their anti-carcinogenic, anti-thrombotic, anti-diabetic, and anti-obesity quality, as well as an immune modulator and in pregnant women for better fetus growth (Echeverria et al. 2017), they are commonly utilized in many health drinks. The production of EPA, C20:5 n-3 was 23.6% and DHA, C22:6 n-3 was 36.5% in Phaeodactylum tricomulum under the nutrient depletion (Yi et al. 2017). The quantity and composition of PUFAs are invariably influenced by growth phases and environmental factors. The lipid’s PUFA can also be turned into biodiesel by the transesterification method (Chen et al. 2012).

Microalgae can store a substantial amount of lipids with high biological activity. Phaeodactylum tricornutum constituted a neutral fraction containing greater than 60% triglycerides (TAGs), the primary component for biodiesel generation under outdoor conditions (Steinrucken et al. 2018). It was reported that Chlorella vulgaris, Scenedesmus sp., and Spirogyra sp., can accumulate 15 to 40% of lipids by their dry weight of biomass (Mata et al. 2009; Cai et al. 2013). In addition, this fraction of lipid accumulation might be augmented to 70 to 90% by modifying cultivation conditions, such as supplementing KNO3, as a nitrogen source (Gour et al. 2018). Addition of NaNO3 at a concentration of 18.75 mg/L to Nitzschia sp. produced a lipid content of around 60 mg/L (Harini et al. 2020). Carbon-nitrogen (C/N) ratio, pH, high salinity, temperature, and nitrogen depletion in the medium play a key role in boosting lipid formation (Kwak et al. 2016). Amphora coffeaeformis RR03 produced a lipid of about 67.15% cultured in an open raceway pond (Rajaram et al. 2018). In the semicontinuous mode in the open raceway pond (ORP) Chlorella vulgaris UTEX 26 was grown and produced the highest lipid of 6.1 g m-2 d-1 along with the periodic addition of NH4HCO3 and ammonia (Ramirez-Lopez et al. 2019). Nitrogen plays a major role in the synthesis of lipids. Excessive addition of nitrogen to the media helps improve the production of the lipid and the products (Table 3).

Table 3. Adaptation of Conditions to Promote Lipid Products for the Production of a Wide Range Bioproducts

Pigments

Pigments present in microalgae have a pivotal role in the development of a myriad of bioactive compounds in the form of secondary metabolites. Microalgae are used to produce feed, antibiotics, cosmetics, nutritional food, and economically efficient pigments. It is also used to treat cancers, neurological disorders, and eye ailments (Chew et al. 2017). Current research focuses on the exploitation of wastewater sources for pigment synthesis (McClure et al. 2018). Algal pigments include phycocyanin, lutein, fucoxanthin, β-carotene, diatoxanthin, diadinoxanthin, and astaxanthin (Sathasivam et al. 2017). Phycocyanin fluorescent blue-colored phycobiliprotein acts as an antioxidant and anti-inflammatory, is found in cosmetics, and it also helps treat liver, colon, lung, and breast cancers (Fernandez-Rojas et al. 2014; Kumar et al. 2014). Among its many health benefits are its antioxidant, anti-inflammatory, and hepatoprotective properties (Lima et al. 2018). It is also used in popsicles, chewing gum, confectionery, wasabi, dairy products, and soft drinks (Gattullo et al. 2012). Research has shown that extract of phycocyanin used for the production of biscuits along with which wheat flour has higher nutritional properties (Garcia et al. 2017). Spirulina platensis produced 159.9 mg/g of phycocyanin using a pulsed electric field, a sort of pretreatment method that increases the permeability of algal cells (Martinez et al. 2017). Ultrasound-assisted extraction (UAE) of phycocyanin from Spirulina platenis yielded around 13.6% of pigment (Hadiyanto and Suttrisnorhadi 2016). Phycoerythrin, a phycobiliprotein was purified and exploited as a fluorescent dye for Porphyridium marinum study and produced B-phycoerythrin of about 40 mg/g of dry cell weight (DCW) NaNO3 = 3.4 g/L with the light intensity of 70 μmol photons/m2/s and metal solution about 1.5 mL/L (Garcouch et al. 2018). Biomass from Porphyra sp. and Arthrospira sp. produced 8.32 mg/g and 8.18 mg/g of phycoerythrin, respectively using UAE (Ardiles et al. 2020).

Lutein is one of the two most abundant carotenoids in the human eye (macula and retina). Many individuals consider lutein to be “the eye vitamin.” Many microalgae can synthesize lutein pigment. For example, Chlorella minutissima MCC-27 contributes to the generation of 17.28 mg/L of lutein based on the study of ANN and particle swarm optimization (PSO) (Dineshkumar et al. 2015). In a two-stage fed-batch mixotrophy condition, Chlorella sorokiniana FZU60 produced 65.96 mg/L of lutein (Ma et al. 2020).

Isoprenoids constitute a significant and vital class of biomolecules. A collection of isoprenoids composes the various pigments. Fucoxanthin, β-carotene, diatoxanthin, and diadinoxanthin are regarded as isoprenoids in this context and found in Phaeodactylum tricornutum and Botryococcus braunii (Niehaus et al. 2011). Fucoxanthin has an allenic link, nine conjugated double bonds, a 5,6-monoepoxide, and several oxygenic functional groups. They possess numerous biological features, including antioxidant, anti-obesity, and anti-inflammatory qualities (Maeda 2015; Zhang et al. 2015). Fucoxanthinol, the deacetylated derivative, demonstrated potential in the therapy of numerous cancer cell types and antineoplastic activity (Martin 2015). Fucoxanthin is one of the main chlorophyll a/c complex compounds found predominantly in diatoms, where it functions as a light-harvesting pigment during photosynthesis and growth. For instance, Isochrysis sp. in better media after optimization produced 7.5 to 23.3 mg/g (Sun et al. 2019). For Nitzschia sp. under high silica, the concentration produced was 12 to 32.8 mg/g (Mao et al. 2020) and Tisochrysis lutea synthesized 2.1 to 79.4 mg/g under single-cell fluorescence (Gao et al. 2020). Phaeodactylum tricornutum produced approximately 59.2 mg/g of fucoxanthin. Compared to the synthesis of fucoxanthin in an open raceway pond, the growth of microalgae in a controlled environment, such as a PBR, yields a higher product yield (Quader and Ahmed 2017).

Carotenoids are isoprenoid structured lipophilic pigments that are found in non-photosynthetic organisms. They have strong antioxidant properties, thereby protecting the organisms from oxidative and free-radical stress. A range of 0.1 to 0.2% of total dry matter of microalgae may consist of carotenoids. It contains an abundance of colors, including yellow, orange, and red. Carotenoids consist of over 600 colors found in nature (Herrero et al. 2013). In the food and pharmaceutical industries, it has expanded applications to lessen the effects of smoking, hypertension, dyslipidemia, diabetes, cancer, cardiovascular disease, and atherosclerosis (Lobo et al. 2010; Herrero et al. 2013). According to reports, using the modified medium with different concentrations of NaCl, 10% to 14% of β-carotene was recovered from Dunaliella salina, which is beneficial for vision and the immune system (Sathasivam and Juntawong 2013).

Xanthophylls, specifically diatoxanthin and diadinoxanthin, are regarded as beneficial chemicals that are diatom-specific (Sathasivam et al. 2017). The article demonstrates that numerous diatom species manufacture these pigments efficiently. Mytilus coruscus produced approximately 133.97 mg/kg of diatoxanthin and 107.16 mg/kg of diadinoxanthin from its DCW by altering the acetylenic carotenoid pathway and 4-keto oxidative pathway (Li et al. 2022a). Isochrysis zhangjiangensis also generated 0.75 mg/g and 4.5 mg/g of diatoxanthin and diadinoxanthin, respectively. The researcher also found that increasing light induces the biosynthesis of fatty acids but reduces the formation of fucoxanthin, whereas, in intense light, the cycle of diadinoxanthin to diatoxanthin was also triggered (Li et al. 2022b). Haematococcus pluvialis produced approximately 4% of its DCW as astaxanthin using different cell disruption methods (Kim et al. 2022). It was observed that Oedocladium carolinianum produced 24.2 mgL-1day-1 of astaxanthin along with the production of lipids in the open and closed PBRs (Wang et al. 2022). Possessing medicinal properties, these pigments treat maladies such as diabetes, ageing, cancer, obesity, and stroke (Raposo et al. 2013; Lin et al. 2016). Based on previous studies, it can be assumed that a consistent light source can enhance the growth of microalgae, leading to a rise in product output. Depending on the different physical and chemical factors and the type of microalgae, the composition of the pigments varies in their biomass.

Silica

Silica forms the cell wall of diatoms, a characteristic feature in the family Bacillariophyceae. They overlap with valves designated epitheca and hypotheca, which mimic a petri dish structure interconnected by a silica girdle band. Utilizing a minuscule amount of energy, silica absorbs diatoms from the external environment via silica acid transporters. In addition, the consumption of silica from surface water to sediments is a benefit of diatoms. Because diatoms are the primary and dominant producers in aquatic ecosystems, they deposit a greater quantity of silica in deep water over time (Sun et al. 2017). The presence of silica causes the diatoms in diatomaceous earth to settle and settle over years. According to reports, diatomaceous earth could potentially act as an adjuvant for vaccination against chicken infections (Nazmi et al. 2017). This silica is primarily obtained from Coscinodiscus wailesii, Cyclotella sp., and Chaetoceros sp. (Esfandyari et al. 2020) and is taken advantage of in the biosensor field (antibody conjugation). In addition, it serves as a drug carrier for Nitzschia palea (Singh et al. 2020) and Thalassiosira weissflogii (Cicco et al. 2016).

Vitamins

Microalgae are capable of generating and accumulating a wide assortment of vitamins, including pro-vitamin A, some B vitamins (B1, B2, B3, B5, B6, B8, B9, and B12), vitamin C, and vitamin E (Galasso et al. 2019). Spirulina platensis, Isochyris galbana, T. suecia, and P. cruentum produced an abundance of vitamin E (tocopherols) about 120.5, 115.5, 159.8, and 184.7 µg/g in continuous cultivation (Lopez-Hernandez et al. 2020). Porphyridium cruentum generates large quantities of vitamin A of 0.75 mg/g in closed PBR (Santiago-Morales et al. 2018). Dunaliella salina synthesizes vitamin E, vitamin C, pyridoxine, nicotinic acid, thiamine, riboflavin, and biotin efficiently (Tafreshi and Shariati 2009). In contrast, Tarento et al. (2018) investigated the vitamin synthesis of several microalgae. Based on his study, 1 g of cylindrical Anabaena powder contains 64% of vitamin B12. Additionally, Spirulina has a maximum of 40.9 mg/g of vitamin B2. Approximately 0.24 mg/g of vitamin B3 was extensively synthesized by Chlorella sp.

Sterols

Phytosterols are sterols, such as sitosterol, campesterol, and stigmasterol, which have diverse pharmacological effects, including anticancer and anti-inflammatory, are exploited as food additives. Several microalgal strains produce phytosterols primarily utilized to reduce cholesterol levels (Luo et al. 2015). Chaetoceros sp. yielded phytosterols at a concentration of 27.7 mg/g DCW. In similarity, Pavlova lutheri produced 22 mg/g of total sterol (Ahmed and Schenk 2017). Pavlova lutheri, Tetraselmis sp. M8, and Nannochloropsis BR2 yield a range of 0.4 to 2.6% sterols (Ahmed et al. 2015; Santhosh et al. 2016) (Table 4).

Table 4. Bioactive Chemicals from Several Microalgae for the Synthesis of Non-Hazardous Renewable Products

GENETIC ENGINEERING OF MICROALGAE

Genetic engineering is the process of manipulating genes to mass-produce the desired product. The associated procedures are mostly utilized in chemistry, pharmacology, biochemistry, and biotechnology (Manuel et al. 2018). Important steps in genetic engineering investigation include transformation techniques and selection. Genetic engineering uses particle bombardment, glass beads, electroporation, agrobacterium-mediated transformations, direct gene editing, etc. for direct gene transfer (Ng et al. 2017).

The genomic sequences of Chlamydomonas reinhardtii and Phaeodactylum tricornutum were completely analyzed and sequenced. In microalgae, Chlamydomonas reinhardtii is the first model organism (Merchant et al. 2007). The comprehensive examination of the order of the genes prepares for the enhancement of the production of various chemicals via genetic modifications such as inducible promoters, regulatory elements, and the insertion or exclusion of genes. The transition of the gene into nuclear DNA results in either stable or temporary gene expression for the synthesis of various products (Kao and Ng 2017).

Through CRISPR-mediated phosphoenolpyruvate carboxylase regulation, Nannochloropsis sp. has become a new model organism for carbon sequestration and oil production with 94% stability over seven generations (Kao and Ng 2017). It has been reported that suppression of CrPEPC1 by substituting the CRISPRi gene in Chlamydomonas reinhardtii CC400 was used for the first time to effectively increase lipid synthesis (Johnson et al. 2016). As described by Jester et al. (2022), homologous recombination of the antibiotic-resistance (ABR) gene and the gene of interest (GOI) in the genomic DNA of Spirulina produced 15% of the targeted product, it can be administered orally without purification. The HMG reductase gene was targeted to alter the critical components that are produced by the diatoms. In Phaeodactylum tricornutum, the gene HMG reductase and IDI-SQS are targeted for the expression and overproduction of the triterpenoid biosynthesis pathway (D’Adamo et al. 2018).

Ribonuclease (RNA) interference using the PEPC1 gene in Chlamydomonas reinhardtii resulted in a 1.5-fold increase in lipid accumulation by electroporation (Ahmad et al. 2015). Modified NAB1 gene from Chlamydomonas reinhardtii (T541A, T676A) upregulated by glass bead transformation (Beckmann et al. 2009). In Haematococcus pluvialis, particle bombardment induced the expression of the modified PDS gene (L504R) to boost astaxanthin synthesis by 45% (Steinbrenner and Sandmann 2006). Electroporation and agrobacterium-mediated transformation of the gene by RNA interference, gene-editing employing ZFNs and CRISPR was tested in Chlamydomonas reinhardtii to determine the most effective gene transformation (Mini et al. 2018).

WASTEWATER AS A NUTRIENT SOURCE FOR PRODUCING BIOACTIVE COMPOUNDS

Microalgae, particularly diatoms, can proliferate under low-light circumstances, which may aid their growth in wastewater. The increase of numerous hazardous greenhouse gases, such as CO2, methane (CH4), nitrous oxide (N2O), and wastewater (Gimpel et al. 2013) commensurately accentuates several environmental dangers. To reduce the use of commercial nutrients, wastewater is increasingly and commonly used for algae cultivation (Ho et al. 2011). Because of its dual role in treating wastewater and creating biomass, it has attracted considerable interest (Rajkumar et al. 2022).

Consequently, microalgae play a vital role in a variety of industries for satisfying test demands without causing harm to the environment or posing a threat to human health. Phycoremediation is the utilization of microalgae in wastewater treatment (Phang et al. 2015). Because of their efficient cellular mechanisms and adaptive methodology, microalgae have the capacity to phycoremediate diverse forms of wastewater. They uptake macro-and micronutrients from wastewater to make biomass. When diatoms are grown in wastewater, their biomass can be used to produce a wide range of high-value chemicals with a broad range of applications in bioenergy, medicinal chemistry, food, and nutrition (Olguin 2012). Thus, wastewater can be utilized for the production of desirable items such as renewable fuels, food, fertilizer, pharmaceuticals, cosmeceuticals, PUFAs, and aquaculture. The oxygen produced by photosynthesis can support the growth of heterotrophic aerobic bacteria, hence accelerating the biodegradation of pollutants (Godos et al. 2010). Microalgae were cultivated using several forms of wastewater, including municipal, aquaculture, dairy, poultry, etc. (Fig. 2)

Approximately 116.2 mg/g of ethanol was produced from 1.4 g/L of biomass containing 38% carbohydrates, 15% proteins, and 22% lipids by growing in dairy effluent (Hemalatha et al. 2019). Tribonema sp. was cultivated in swine effluent and yielded a lipid concentration of approximately 42.4% (Cheng et al. 2020). In Desmodesmus sp. PW1, 29.4% of the generated lipids were derived from swine wastewater (Chen et al. 2020). Chlorella sorokiniana CY-1 cultivated in the wastewater of a palm oil mill yielded 14.43% lipids (Cheah et al. 2020).

Chlorella sp. helps anemic consumers boost their haemoglobin (Barrow and Shahidi 2007), whereas Azolla and Anabaena work as a biofertilizer to increase the nitrogen content of soil (Priyadarshani and Rath 2012). For cosmetics, Dunaliella salina strongly affects the energy metabolism of cells to promote their growth (Stolz and Obermayer 2013). Treatment of slaughterhouse wastewater by Chlorella salina for the utilization of nitrate and phosphate for the growth of the microalgae was performed in the open raceway pond. The biomass further used for producing different bioactive products (Habibi et al. 2018).

Fig. 2. Exploitation of microalgae for the biodegradation of toxic nutrients from wastewater from different streams to manufacture bioactive products and recycle water

CHALLENGES AND PROSPECTS

Because the products from the microalgae are exposed to the higher demands in various streams such as food, cosmetics, nutraceuticals, and synthesis of the renewable products occurs. Challenges occur when the production cost is comparatively higher than the product cost. Hence the need for alternatives is high before production at a low cost. The necessity of the compounds has increased, and therefore the best technologies have to be implemented for better results.

Existing systems for cultivation and post-harvest processing are unsustainable and unaffordable, as they consume around 40% of the total cost (Gifuni et al. 2019). Changes in geometry and fluid mixing pattern, better gas exchange, light penetration, and building material, all of which have advantages and disadvantages, must still be adjusted for production in PBRs. When operating PBRs, the majority of studies prioritize lighting patterns, fluid dynamics, cooling requirements, mixing efficiency, and mass transfer (Sirohi et al. 2022a).

There are certain downsides to the large-scale cultivation procedure. Therefore, the creation of open ponds for growth decreases costs and is simple to handle, but there are more variances such as pollution, irregular growth, predator contamination, temperature and light complications (Xu et al. 2009).

The creation of large-scale PBRs, such as tubes and containers, with minimal contamination risk would be more beneficial in the future (Wang et al. 2012). The PBRs provide numerous advantages over open systems, in terms of reduced pollution and the ability to cultivate monocultures of axenic algae. They provide more places before starting factors including pH, temperature, light, and CO2 concentration. In addition, water does not evaporate in PBRs. In PBRs, higher cell concentrations are also realistically possible (Sirohi et al. 2022b; Udayan et al. 2022). The construction cost of the PBRs is higher compared with the open raceway ponds. The production of the product has to be more continuous than batch culturing. For biomass collection and product recovery, several technologies must be updated for the effective collection of products. Environmental and economic studies are needed along with a life cycle assessment for a better yield of the products without loss.

CONCLUDING REMARKS

In comparison to a great number of other chemical and artificially manufactured goods, microalgae are considered as more important while avoiding many of the associated drawbacks. Scientists in multiple fields are studying the potential of microalgae in different products. Microalgae include a variety of species capable of utilizing environmentally harmful compounds, wastewater, and CO2. Treatment of wastewater is one of the most efficient strategies for the management and production of high-value compounds. Even while it serves as biomass for biofuels, health, cosmetics, and saves the environment from profoundly detrimental repercussions, biodiesel is less economically competitive than biohydrogen and biobutanol.

The high value-added bioproducts, such as astaxanthin, that were derived from microalgal sources have the potential to be applied in the pharmaceutical and nutraceutical industries. In comparison to first and third-generation fuels, intensive research has been conducted on fourth-generation fuels, which includes advanced low-cost technology and genetic modification for enhancing sustainable bioproducts. This has been done in an effort to improve the algal bioeconomy.

There is still a significant amount of cross-disciplinary research and development work that has to be done before more complex uses of algae may be implemented in any industry. To improve the process of acquiring the products, a wide variety of procedures and extraction methodologies need to be developed and put into practice.

ACKNOWLEDGMENTS

This study was financially supported by Science & Engineering Research Board (SERB), Department of Science and Technology, Government of India under grant number: YSS/2015/001217.

REFERENCES CITED

Abou-Shanab, R. A. I., Ji, M. K., Kim, H. C., Paeng, K. J., and Jeon, B. H. (2013). “Microalgal species growing on piggery wastewater as a valuable candidate for nutrient removal and biodiesel production,” J. Environ. Manag. 115, 257-264. DOI: 10.1016/j.jenvman.2012.11.022

Adarme-Vega, T. C., Lim, D. K., Timmins, M., Vernen, F., Li, Y., and Schenk, P. M. (2012). “Microalgal biofactories: A promising approach towards sustainable omega-3 fatty acid production,” Microb. Cell Factories 1, article no. 96. DOI: 10.1186/1475-2859-11-96

Aguirre, A. M., and Bassi, I. A. (2013). “Investigation of biomass concentration, lipid production, and cellulose content in Chlorella vulgaris cultures using response surface methodology,” Biotechnol. Bioeng. 110(8), 2114-2122. DOI: 10.1002/bit.24871

Ahmad, I., Sharma, A. K., Daniell, H., and Kumar, S. (2015). “Altered lipid composition and enhanced lipid production in green microalga by introduction of brassica diacylglycerol acyltransferase 2,” Plant Biotechnol. J. 13(4), 540-550. DOI: 10.1111/pbi.12278

Ahmed, F., and Schenk, P. M. (2017). “UV-C radiation increases sterol production in the microalga Pavlova lutheri,” Phytochem. 139, 25-32. DOI: 10.1016/j.phytochem.2017.04.002

Ahmed, F., Zhou, W., and Schenk, P. M. (2015). “Pavlova lutheri is a high-level producer of phytosterols,” Algal Res. 10, 210-217. DOI: 10.1016/j.algal.2015.05.013

Andrade, L., Batista, F., Lira, T., Barrozo, M., and Vieira, L. (2018). “Characterization and product formation during the catalytic and non-catalytic pyrolysis of the green microalgae Chlamydomonas reinhardtii,” Renew. Energy 119, 731-740. DOI: 10.1016/j.renene.2017.12.056

Ardiles, P., Cerezal-Mezquita, P., Salinas-Fuentes, F., Ordenes, D., Renato, G., and Ruiz-Domínguez, M. C. (2020). “Biochemical composition and phycoerythrin extraction from red microalgae: A comparative study using green extraction technologies,” Process 8(12), article ID 1628. DOI: 10.3390/pr8121628

Arora, N., and Philippidis, G. P. (2021). “Insights into the physiology of Chlorella vulgaris cultivated in sweet sorghum bagasse hydrolysate for sustainable algal biomass and lipid production,” Sci. Rep. 11, article no. 6779. DOI: 10.1038/s41598-021-86372-2

Ashokkumar, V., Chen, W. H., Ngamcharussrivichai, C., Agila, E., and Ani, F. N. (2019). “Potential of sustainable bioenergy production from Synechocystis sp. cultivated in wastewater at large scale – A low cost biorefinery approach,” Energy Convers. Manag. 186, 188-199. DOI: 10.1016/j.enconman.2019.02.056

Baldisserotto, C., Sabia, A., Giovanardi, M., Ferroni, L., Maglie, M., and Pancaldi, S. (2022). “Chlorophyta microalgae as dietary protein supplement: A comparative analysis of productivity related to photosynthesis,” J. Appl. Phycol. 34, 1323-1340. DOI: 10.1007/s10811-022-02724-z

Barrow, C., and Shahidi, F. (2007). Marine Nutraceuticals and Functional Foods, CRC Press, Boca Raton, FL, USA. DOI: 10.1201/9781420015812

Beckmann, J., Lehr, F., Finazzi, G., Hankamer, B., Posten, C., Wobbe, L., and Kruse, O. (2009). “Improvement of light to biomass conversion by de-regulation of light-harvesting protein translation in Chlamydomonas reinhardtii,” J. Biotechnol. 142(1), 70-77. DOI: 10.1016/j.jbiotec.2009.02.015

Bernaerts, T. M. M., Gheysen, L., Kyomugasho, C., Kermani, Z. J., Vandionant, S., Foubert, I., Hendrickx, M. E., and Loey, A. M. (2018). “Comparison of microalgal biomasses as functional food ingredients: Focus on the composition of cell wall related polysaccharides,” Algal Res. 32, 150-161. DOI: 10.1016/j.algal.2018.03.017

Braga, V. S., Moreira, J. B., Costa, J. A. V., and Morais, M. G. (2019). “Enhancement of the carbohydrate content in Spirulina by applying CO2, thermoelectric fly ashes and reduced nitrogen supply,” Int. J. Biol. Macromol. 123, 1241-1247. DOI: 10.1016/j.ijbiomac.2018.12.037

Breil, C., Vian, M. A., Zemb, T., Kunz, W., and Chemat, F. (2017). “Bligh and Dyer and Folch methods for solid–liquid–liquid extraction of lipids from microorganisms, comprehension of solvation mechanisms and towards substitution with alternative solvents,” Int. J. Mol. Sci. 18(4), article no. 708. DOI: 10.3390/ijms18040708

Cai, T., Park, S. Y., and Li, Y. (2013). “Nutrient recovery from wastewater streams by microalgae: Status and prospects,” Renew. Sustain. Energy Rev. 19, 360-369. DOI: 10.1016/j.rser.2012.11.030

Caporgno, M. P., and Mathys, A. (2018). “Trends in microalgae incorporation into innovative food products with potential health benefits,” Front. Nutr. 5, article no. 58. DOI: 10.3389/fnut.2018.00058

Chanda, M., Merghoub, N., and Arroussi, H. (2019). “Microalgae polysaccharides: The new sustainable bioactive products for the development of plant bio-stimulants?,” World J. Microbiol. Biotechnol. 35, 1-10. DOI: 10.1007/s11274-019-2745-3

Cheah, W. Y., Show, P. L., Yap, Y. J., Mohd Zaid, H. F., Lam, M. K., Lim, J. W., Ho, Y. C., and Tao, Y. (2020). “Enhancing microalga Chlorella sorokiniana CY-1 biomass and lipid production in palm oil mill effluent (POME) using novel-designed Photobioreactor,” Bioengineered 11(1), 61-69. DOI: 10.1080/21655979.2019.1704536

Cheirsilp, B., Mandik, Y. I., and Prasertsan, P. (2016). “Evaluation of optimal conditions for cultivation of marine Chlorella sp. as potential sources of lipids, exopolymeric substances and pigments,” Aquacult. Int. 24, 313-326. DOI: 10.1007/s10499-015-9927-2

Chen, Y., Huang, B., Chiang, T., and Tang, T. (2012). “Fuel properties of microalgae (Chlorella protothecoides) oil biodiesel and its blends with petroleum diesel,” Fuel 94, 270-273. DOI: 10.1016/j.fuel.2011.11.031

Chen, C. Y., Zhao, X. Q., Yen, H. W., Ho, S. H., Cheng, C. L., Lee, D. J., Bai, F. W., and Chang, J. S. (2013). “Microalgae-based carbohydrates for biofuel production,” Biochem. Eng. J. 78, 1-10. DOI: 10.1016/j.bej.2013.03.006

Chen, C. L., Change, J. S., Huang, C. C., Ho, K. C., Hsiao, P. X., and Wu, M. S. (2014). “A novel biodiesel production method consisting of oil extraction and transesterification from wet microalgae,” Energy Procedia 61, 1294-1297. DOI: 10.1016/j.egypro.2014.11.1084

Chen, Z., Shao, S., He, Y., Luo, Q., Zheng, M., Zheng, M., Chen, B. and Wang, M. (2020). “Nutrients removal from piggery wastewater coupled to lipid production by a newly isolated self-flocculating microalga Desmodesmus sp. PW1,” Bioresour. Technol. 302, article ID 122806. DOI: 10.1016/j.biortech.2020.122806

Cheng, P., Chen, D., Liu, W., Cobb, K., Zhou, N., Li, Y., Liu, H., Wang, Q., Chen, P., Zhou, C., et al. (2020). “Auto-flocculation microalgae species Tribonema sp. and Synechocystis sp. with T-IPL pretreatment to improve swine wastewater nutrient removal,” Sci. Total Environ. 725, article ID 138263. DOI: 10.1016/j.scitotenv.2020.138263

Chentir, I., Hamdi, M., Doumandji, A., HadjSadok, A., Ouadam H. B., Nasri, M., and Jridi, M. (2017). “Enhancement of extracellular polymeric substances (EPS) production in Spirulina (Arthrospira sp.) by two-step cultivation process and partial characterization of their polysaccharidic moiety,” Int. J. Biol. Macromol. 105(2), 1412-1420. DOI: 10.1016/j.ijbiomac.2017.07.009.

Chew, K. W., Yap, J. Y., Show, P. L., Suanc, N. H., Juan, J. C., Ling, T. C., Lee, D-J., and Chang, J-S. (2017). “Microalgae biorefinery: High value products perspectives,” Bioresour. Technol. 229, 53-62. DOI: 10.1016/j.biortech.2017.01.006

Cicco, S. R., Vona, D., Gristina, R., Sardella, E., Ragni, R., Presti, M., and Farinola, G. M. (2016). “Biosilica from living diatoms: Investigations on biocompatibility of bare and chemically modified Thalassiosira weissflogii silica shells,” Bioeng. 3
(4), article no. 35. DOI: 10.3390/bioengineering3040035

Colusse, G. A., Carneiro, J., Duarte, M. E. R., de Carvalho, J. C., and Noseda, M. D. (2022). “Advances in microalgal cell wall polysaccharides: A review focused on structure, production, and biological application,” Crit. Rev. Biotechnol. 42(4), 562-577. DOI: 10.1080/07388551.2021.1941750

Condor, B. E., de Luna, G. M. D., Abarca, R. R. M., Chang, Y.-H., Leong, Y. K., Chen, C.-Y., Chen, P.-T., Lee, D.-J., and Chang, J.-S. (2022). “Optimization and modeling of carbohydrate production in microalgae for use as feedstock in bioethanol fermentation,” Int. J. Energy Res. (Online), 1-13. DOI: 10.1002/er.7709

Costa, J. A. V., Lucas, B. F., Alvarenga, A. G. P., Moreira, J. B.. and de Morais, M. G. (2021). “Microalgae polysaccharides: An overview of production, characterization, and potential applications,” Polysaccharides 2, 759-772. DOI: 10.3390/polysaccharides2040046

Cuellar-Bermudez, S., Magdalena, J. A., Muylaert, K., and Gonzalez-Fernandez, C. (2019). “High methane yields in anaerobic digestion of the cyanobacterium Pseudanabaena sp.,” Algal Res. 44, article ID 101689. DOI: 10.1016/j.algal.2019.101689

Custodio, L., Justo, T., Silvestre, L., Barradas, A., Duarte, C. V., Pereira, H., Barreira, L., Rauter, A. P., Albericio, F., and Varelaa, J. (2012). “Microalgae of different phyla display antioxidant, metal chelating and acetylcholinesterase inhibitory activities,” Food Chem. 131(1), 134-140. DOI: 10.1016/j.foodchem.2011.08.047

D’Adamo, S., Visconte, G. S., Low, G., Szaub-Newton, J., Beacham, T., Landels, A., Allen, M. J., Spicer, A., and Matthijs, M. (2018). “Engineering the unicellular alga Phaeodactylum tricornutum for high-value plant triterpenoid production,” Plant Biotechnol. J. 17(1), 75-87. DOI: 10.1111/pbi.12948

De Souza, M. P., Sanchez-Barrio, A. S., Rizzetti, T. M., Benitez, L. B., Hoeltz, M., Schneider, R. C. S., and Neves, F. F. (2020). “Concepts and trends for extraction and application of microalgae carbohydrates,” Microalgae from Physiology to Application, M. Vitova (ed.), IntechOpen, London, UK, pp. 1-13. DOI: 10.5772/intechopen.83737

Dineshkumar, R., Dhanarajan, G., Dash, S. K., and Sen, R. (2015). “An advanced hybrid medium optimization strategy for the enhanced productivity of lutein in Chlorella minutissima,” Algal Res. 7, 24-32. DOI: 10.1016/j.algal.2014.11.010

Domozych, D. S., Ciancia, M., Fangel, J. U., Mikkelsen, M. D., Ulvskov, P., and Willats, W. G. T. (2012). “The cell walls of green algae: A journey through evolution and diversity,” Front. Plant Sci. 3, article no. 82. DOI: 10.3389/fpls.2012.00082

Dong, X., Han, B., Zhao, Y., Ding, W., and Yu, X. (2019). “Enhancing biomass, lipid production, and nutrient utilization of the microalga Monoraphidium sp. QLZ-3 in walnut shell extracts supplemented with carbon dioxide,” Bioresour. Technol. 287, article ID 121419. DOI: 10.1016/j.biortech.2019.121419

Echeverria, F., Valenzuela, R., Hernandez-Rodas, M. C., and Valenzuela, A. (2017). “Docosahexaenoic acid (DHA), a fundamental fatty acid for the brain: New dietary sources,” Prostaglandins Leukot. Essent. Fat. Acid. 124, 1-10. DOI: 10.1016/j.plefa.2017.08.001

Esfandyari, J., Shojaedin-Givi, B., Hashemzadeh, H., Mozafari-Nia, M., Vaezi, Z., and Naderi-Manesh, H. (2020). “Capture and detection of rare cancer cells in blood by intrinsic fluorescence of a novel functionalized diatom,” Photodiagn. Photodyn. Ther. 30, article ID 101753. DOI: 10.1016/j.pdpdt.2020.101753

Estevez, J. M., Leonardi, P. I., and Alberghina, J. S. (2008). “Cell wall carbohydrate epitopes in the green alga Oedogonium bharuchae f. minor (Oedogoniales, Chlorophyta),” J. Phycol. 44(5), 1257-1268. DOI: 10.1111/j.1529-8817.2008.00568.x

Fernandez-Rojas, B., Hernandez-Juarez, J., and Pedraza-Chaverri, J. (2014). “Nutraceutical properties of phycocyanin,” J. Funct. Foods 11, 375-392. DOI: 10.1016/j.jff.2014.10.011

Fu, W., Nelson, D. R., Yi, Z., Xu, M., Khraiwesh, B., Jijakli, K., Chaiboonchoe, A., Alzahmi, A., Al-Khairy, D., Brynjolfsson, S., et al. (2017). “Bioactive compounds from microalgae: Current development and prospects,” Stud. Nat. Prod. Chem. 54, 199-225. DOI: 10.1016/B978-0-444-63929-5.00006-1

Galasso, C., Gentile, A., Orefic, I., Ianora, A., Bruno, A., Noonan, D. M., Sansone, C., Albini, A., and Brunet, C. (2019). “Microalgal derivatives as potential nutraceutical and food supplements for human health: A focus on cancer prevention and interception,” Nutrients 11(6), article no. 1226. DOI: 10.3390/nu11061226

Gao, F., Teles, I., Ferrer-Ledo, N., Wijffels, R. H., and Barbosa, M. J. (2020). “Production and high throughput quantification of fucoxanthin and lipids in Tisochrysis lutea using single-cell fluorescence,” Bioresour. Technol. 318, article ID 124104. DOI: 10.1016/j.biortech.2020.124104

Garcia, J. L., Vicente, M., and Galan, B. (2017). “Microalgae, old sustainable food and fashion nutraceuticals,” Microb. Biotechnol. 10, 1017-1024. DOI: 10.1111/1751-7915.12800

Garcia, E. S., Leeuwen, J. A. A. V., Sajtsma, S. L., Broek, Eppink, M. H. M., Wijfells, R. H., and Berg, C. (2018). “Techno-functional properties of crude extracts from the green microalga Tetraselmis suecica,” J. Agric. Food Chem. 66(2), 7831-7838. DOI: 10.1021/acs.jafc.8b01884

Garcouch, N., Karkouch, I., Elleuch, J., Elkahoui, S., Michaud, P., Abdelkafi, S., Laroche, C., and Fendri, I. (2018). “Enhanced B-phycoerythrin production by the red microalgae Porphyridium marinum: A powerful agent in industrial applications,” Int. J. Biol. Macromol. 120, 2106-2114. DOI: 10.1016/j.ijbiomac.2018.09.037

Gattullo, C. E., Bahrs, H., Steinberg, C. E. W., and Loffredo, E. (2012). “Removal of bisphenol A by the freshwater green alga Monoraphidium braunii and the role of natural organic matter,” Sci. Total Environ. 416, 501-506. DOI: 10.1016/j.scitotenv.2011.11.033

Gifuni, I., Pollio, A., Marzocchella, A., and Olivieri, G. (2018). “New ultra-flat photobioreactor for intensive microalgal production: The effect of light irradiance,” Algal Res. 34, 134-142. DOI: 10.1016/j.algal.2018.07.014

Gifuni, I., Pollio, A., Safi, C., Marzocchella, A., and Olivieri, G. (2019). “Current bottlenecks and challenges of the microalgal biorefinery,” Trends Biotechnol. 37, 242-252. DOI: 10.1016/j.tibtech.2018.09.006

Gimpel, J. A., Specht, E. A., Georgianna, D. R., and Mayfield, S. P. (2013). “Advances in microalgae engineering and synthetic biology applications for biofuel production,” Curr. Opin. Chem. Biol. 17, 489-495. DOI: 10.1016/j.cbpa.2013.03.038

Godos, I., Vargas, V. A., Blanco, S., González, M. C. G., Soto, R., García-Encina, P. A., Becares, E., and Munoz, R. (2010). “A comparative evaluation of microalgae for the degradation of piggery wastewater under photosynthetic oxygenation,” Bioresource Technol. 101, 5150-5158. DOI: 10.1016/j.biortech.2010.02.010

Goncalves, C. F., Menegol, T., and Rech, R. (2019). “Biochemical composition of green microalgae Pseudoneochloris marina grown under different temperature and light conditions,” Biocatal. Agric. Biotechnol. 18, article ID 101032. DOI: 10.1016/j.bcab.2019.101032

Gonzalez-Fernandez, C., and Ballesteros, M. (2012). “Linking microalgae and cyanobacteria culture conditions and key-enzymes for carbohydrate accumulation,” Biotechnol. Adv. 300(6), 1655-1661. DOI: 10.1016/j.biotechadv.2012.07.003

Gour, R. S., Bairagi, M., Garlapati, V. K., and Kant, A. (2018). “Enhanced microalgal lipid production with media engineering of potassium nitrate as a nitrogen source,” Bioengineered 9, 98-107. DOI: 10.1080/21655979.2017.1316440

Gui, J., Tong, W., Huang, S., Liang, X., Fang, Z., Wang, W., and Zhang, Y. (2019). “Effects of Chlorella vulgaris polysaccharides accumulation on growth characteristics of Trachemys scripta elegans,” Int. J. Biol. Macromol. 141, 1304-1313. DOI: 10.1016/j.ijbiomac.2019.08.248

Gunnison, D., and Alexander, M. (1975). “Basis for the susceptibility of several algae to microbial decomposition,” Can. J. Microbiol. 21, 619-628. DOI: 10.1139/m75-089

Habibi, A., Teymouri, A., Amrei, H. D., and Shariati, F. P. (2018). “A novel open raceway pond design for microalgae growth and nutrients removal from treated slaughterhouse wastewater,” Pollution 4(1), 103-110. DOI: 10.22059/poll.2017.238894.299

Hadiyanto, H., and Suttrisnorhadi, S. (2016). “Response surface optimization of ultrasound assisted extraction (UAE) of phycocyanin from microalgae Spirulina platensis,” Emir. J. Food Agric. 28(4), 227-234. DOI: 10.9755/ejfa.2015-05-193

Hamed, I. (2016). “The evolution and versatility of microalgal biotechnology: A review,” Compr. Rev. Food Sci. Food Saf. 15(6), 1104-1123. DOI: 10.1111/1541-4337.12227

Hamed, S. B., Hamed, M. B. B., Kassouar, S., Ayad, A., and El, M. (2016). “Physicochemical analysis of cellulose from microalgae Nannochloropsis gaditana, African J. Biotechnol. 15, 1201-1207. DOI: 10.5897/ AJB2016.15321

Hanifzadeh, M. M., Garcia, E. C., and Viamajala, S. (2018). “Production of lipid and carbohydrate from microalgae without compromising biomass productivities: Role of Ca and Mg,” Renew. Energy 127, 989-997. DOI: 10.1016/j.renene.2018.05.012

Harini, A. B., Rajkumar, R., and Takriff, M. S. (2020). “Enhanced production of lipid as biofuel feedstock from the marine diatom Nitzschia sp. by optimizing cultural conditions,” BioResources 15(4), 7532-7550. DOI: 10.15376/biores.15.4.7532-7550

Hemalatha, M., Sravan, S., Min, B., and Mohan, V. (2019). “Microalgae-biorefinery with cascading resource recovery design associated to dairy wastewater treatment,” Bioresour. Technol. 284, 424-429. DOI: 10.1016/j.biortech.2019.03.106

Herrero, M., Mendiola, J., Plaza, M., and Ibanez, E. (2013). “Screening for bioactive compounds from algae,” in: Advanced Biofuels and Bioproducts, J. W. Lee (ed.), Springer, New York, NY, USA, pp. 833-872. DOI: 10.1007/978-1-4614-3348-4_35

Ho, L., Grasset, C., Hoefel, D., Dixon, M. B., Leusch, F. D. L., Newcombe, G., Saint, C. P., and Brookes, J. D. (2011). “Assessing granular media filtration for the removal of chemical contaminants from wastewater,” Water Res. 45(11), 3461-3472. DOI: 10.1016/j.watres.2011.04.005

Huo, S., Liu, J., Zhu, F., Basheer, S., Necas, D., Zhang, R., Li, K., Chen, D., Cheng, P., Cobb, K., et al. (2020). “Post treatment of swine anaerobic effluent by weak electric field following intermittent vacuum assisted adjustment of N:P ratio for oil-rich filamentous microalgae production,” Bioresour. Technol. 314, article ID 123718. DOI: 10.1016/j.biortech.2020.123718

Hussain, F., Shah, S. Z., Ahmad, H., Abubshait, S. A., Abubshait, H. A., Laref, A., Manikandan, A., Kusuma, H. S., and Iqbal, M. (2021). “Microalgae an ecofriendly and sustainable wastewater treatment option: Biomass application in biofuel and bio-fertilizer production. A review,” Renew. Sustain. Energy Rev. 137, article ID 110603. DOI: 10.1016/j.rser.2020.110603

Jester, B. W., Zhao, H., Gewe, M., Adame, T., Perruzza, L., Bolick, D. T., Agosti, J., Khuong, N., Kuestner, R., Gamble, C., et al. (2022). “Development of spirulina for the manufacture and oral delivery of protein therapeutics,” Nat. Biotechnol. 40, 956-964. DOI: 10.1038/s41587-022-01249-7

Johnson, T. J., Gibbons, J. L., Gu, L., Zhou, R., and Gibbons, W. R. (2016). “Molecular genetic improvements of cyanobacteria to enhance the industrial potential of the microbe: A review,” Biotechnol. Prog. 32(6), 1357-1371. DOI: 10.1002/btpr.2358

Kao, P., and Ng, I. (2017). “CRISPRi mediated phosphoenolpyruvate carboxylase regulation to enhance the production of lipid in Chlamydomonas reinhardtii,” Bioresour. Technol. 245, 1527-1537. DOI: 10.1016/j.biortech.2017.04.111

Kawamura, K., Nishikawa, S., Hirano, K., Ardianor, A., Nugroho, R. A., and Okada, S. (2021). “Large-scale screening of natural genetic resource in the hydrocarbon-producing microalga Botrycoccus braunii identified novel fast-growing strains,” Sci. Rep. 11, article no. 7638. DOI: 10.1038/s41598-021-86760-8

Khan, M. I., Shin, J. H., and Kim, J. D. (2018). “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 17, article no. 36. DOI: 10.1186/s12934-018-0879-x

Kim, B., Lee, S. Y., Narasimhan, A. L., Kim, S., and Oh, Y. (2022). “Cell disruption and astaxanthin extraction from Haematococcus pluvialis: Recent advances,” Bioresour. Technol. 343, article ID 126124. DOI: 10.1016/j.biortech.2021.126124

Koller, M., Muhr, A., and Braunegg, G. (2014). “Microalgae as versatile cellular factories for valued products,” Algal Res. 6, 52-63. DOI: 10.1016/j.algal.2014.09.002

Kroger, M., Klemm, M., and Nelles, M. (2018). “Hydrothermal disintegration and extraction of different microalgae species,” Energies 11(2), article no. 450. DOI: 10.3390/en11020450

Kumar, A., Sharma, S., Shah, E., Parikh, B., Patel, A., Dixit, G., Gupta, S., and Divecha, J. (2019). “Cultivation of Ascochloris sp. ADW007-enriched microalga in raw dairy wastewater for enhanced biomass and lipid productivity,” Int. J. Environ. Sci. Technol. 16, 943-954. DOI: 10.1007/s13762-018-1712-0

Kumar, D., Dhar, D. W., Pabbi, S., Kumar, N., and Walia, S. (2014). “Extraction and purification of C-phycocyanin from Spirulina platensis (CCC540),” Indian J. Plant Physiol. 19(2), 184-188. DOI: 10.1007/s40502-014-0094-7

Kumar, K., Dasgupta, C. N., Nayak, B. K., Lindblad, P., and Das, D. (2011). “Development of suitable photobioreactors for CO2 sequestration addressing global warming using green algae and cyanobacteria,” Bioresour. Technol. 102(8), 4945-4953. DOI: 10.1016/j.biortech.2011.01.054

Kwak, H. S., Kim, J. Y. H., Woo, H. M., Jin, E., Min, B. K., and Sim, S. J. (2016). “Synergistic effect of multiple stress conditions for improving microalgal lipid production,” Algal Res. 19, 215-224. DOI: 10.1016/j.algal.2016.09.003

Lee, H., Kim, K., Mun, S. C., Chang, Y. K., and Choi, S. Q. (2018). “A new method to produce cellulose nanofibrils from microalgae and the measurement of their mechanical strength,” Carbohydr. Polym. 180, 276-285. DOI: 10.1016/j.carbpol.2017.09.104

Li, J., Cai, C., Yang, C., Li, J., Sun, T., and Yu, G. (2019). “Recent advances in pharmaceutical potential of brown algal polysaccharides and their derivatives,” Curr. Pharm. Des. 25(11), 1290-1311. DOI: 10.2174/1381612825666190618143952

Li, S., Fang, K., Chen, S., Xu, J., Chen, H., and Chen, J. (2022a). “Carotenoid profiling of Mytilus coruscus with unialgal cultures: Insights into biosynthetic pathways in tissues,” Aquac. 547, article ID 737472. DOI: 10.1016/j.aquaculture.2021.737472

Li, S., Luo, S., and Guo, R. (2013). “Efficiency of CO2 fixation by microalgae in a closed raceway pond,” Bioresour. Technol. 136, 267-272. DOI: 10.1016/j.biortech.2013.03.025

Li, Y., Sun, H., Wang, Y., Yang, S., Wang, J., Wu, T., Lu, X., Chu, Y., and Chen, F. (2022b). “Integrated metabolic tools reveal carbon alternative in Isochrysis zhangjiangensis for fucoxanthin improvement,” Bioresour. Technol. 347, article ID 126401. DOI: 10.1016/j.biortech.2021.126401

Lima, G. M., Teixeira, P. C. M., Teixeira, C. M. L. L., Filocomo, D., and Lage, C. L. S. (2018). “Influence of spectral light quality on the pigment concentrations and biomass productivity of Arthrospira platensis,” Algal Res. 31, 157-166. DOI: 10.1016/J.ALGAL.2018.02.012

Lin, J., Huang, L., Yu, J., Xiang, S., Wang, J., Zhang, J., Yan, X., Cui, W., He, S., and Wang, Q. (2016). “Fucoxanthin, a marine carotenoid, reverses scopolamine-induced cognitive impairments in mice and inhibits acetylcholinesterase in vitro,” Mar. Drugs, 14, article no. 67. DOI: 10.3390/md14040067

Lobo, V., Patil, A., Phatak, A., and Chandra, N. (2010). “Free radicals, antioxidants and functional foods: Impact on human health,” Pharmacogn. Rev. 4(8), 118-126. DOI: 10.4103/0973-7847.70902

Lopez-Hernandez, J. F., Gracia-Alamilla, P., Plama-Ramirez, D., Alvarez-Gonzalez, C. A., Paredes-Rojas, J. C., and Marquez-Rocha, F. J. (2020). “Continuous microalgal cultivation for antioxidants production,” Molecules 25(18), article no. 4171. DOI: 10.3390/molecules25184171

Luo, X., Su, P., and Zhang, W. (2015). “Advances in microalgae-derived phytosterols for functional food and pharmaceutical applications,” Mar. Drugs 13(7), 4231-4254. DOI: 10.3390/md13074231

Ma, R., Zhang, Z., Ho, S., Ruan, C., Li, J., Xie, Y., Shi, X., Liu, L., and Chen, J. (2020). “Two-stage bioprocess for hyperproduction of lutein from microalga Chlorella sorokiniana FZU60: Effects of temperature, light intensity, and operation strategies,” Algal Res. 52, article ID 102119. DOI: 10.1016/j.algal.2020.102119

Maeda, H. (2015). “Nutraceutical effects of fucoxanthin for obesity and diabetes therapy: A review,” J. Oleo. Sci. 64(2), 125-132. DOI: 10.5650/jos.ess14226

Manuel, S., Gwendoline, D., Anne-Laure, F., Marie-Ange, T., Denis, J., and Fayza, D. (2018). “One step generation of multiple gene knock-outs in the diatom Phaeodactylum tricornutum by DNA-free genome editing,” Nat. Commun. 9(1), article ID 3924. DOI: 10.1038/s41467-018-06378-9

Mao, X., Chen, S. H. Y., Lu, X., Yu, J., and Liu, B. (2020). “High silicate concentration facilitates fucoxanthin and eicosapentaenoic acid (EPA) production under heterotrophic condition in the marine diatom Nitzschia laevis,” Algal Res. 52, article ID 102086. DOI: 10.1016/j.algal.2020.102086

Markou, G., and Georgakakis, D. (2012). “Microalgal carbohydrates: An overview of the factors influencing carbohydrates production, and of main bioconversion technologies for production of biofuels,” Appl. Microbiol. Biotechnol. 96(3), 631-645. DOI: 10.1007/s00253-012-4398-0

Markou, G., Angelidaki, I., and Georgakakis, D. (2013). “Carbohydrate-enriched cyanobacterial biomass as feedstock for bio-m4 ethane production through anaerobic digestion,” Fuel, 111, 872-879. DOI: 10.1016/j.fuel.2013.04.013

Martin, L. J. (2015). “Fucoxanthin and its metabolite fucoxanthinol in cancer prevention and treatment,” Mar. Drugs 13, 4784-4798. DOI: 10.3390/md13084784

Martinez, J. M., Luengo, E., Saldana, G., Alvarez, I., and Raso, J. (2017). “C-phycocyanin extraction assisted by pulsed electric field from Artrosphira platensis,” Food Res. Int. 99(3), 1042-1047. DOI: 10.1016/j.foodres.2016.09.029

Mata, T. M., Nio Martins, A. A., and Caetano, N. S. (2009). “Microalgae for biodiesel production and other applications: A review,” Renew. Sustain. Energy Rev. 14, 217-232. DOI: 10.1016/j.rser.2009.07.020

McClure, D. D., Luiz, A., Gerber, B., Barton, G. W., and Kavanagh, J. M. (2018). “An investigation into the effect of culture conditions on fucoxanthin production using the marine microalgae,” Algal Res. 29(2), 41-48. DOI: 10.1007/s12010-012-9602-2

Medina-Cabrera, E. V., Ruhmann, B., Schmid, J., and Sieber, V. (2020). “Optimization of growth and EPS production in two Porphyridum strains,” Bioresour. Technol. Rep. 11, article ID 100486. DOI: 10.1016/j.biteb.2020.100486

Mehrabadi, A., Craggs, R., and Farid, M. M. (2016). “Biodiesel production potential of wastewater treatment high rate algal pond biomass,” Bioresour. Technol. 221, 222-233. DOI: 10.1016/j.biortech.2016.09.028

Melvelle, J. (2012). Synthesis and Characterization of Biofuels, General Chemistry and Quantitative analysis, Lab at UC Berkeley College of Chemistry, Berkeley, CA, USA, pp. 1-14.

Mendoza, J. L., Granados, M. R., de Godos, I., Acien, F. G., Molina, E., Banks, C., and Heaven, S. (2013). “Fluid-dynamic characterization of real-scale raceway reactors for microalgae production,” Biomass Bioenerg. 54, 267-275. DOI: 10.1016/j.biombioe.2013.03.017

Menestrino, B. C., Pintos, T. H. C., Sala, L., Costa, J. A. V., and Santos, L. O. (2020). “Application of static magnetic fields on the mixotrophic culture of Chlorella minutissima for carbohydrate production,” Appl. Biochem. Biotechnol. 192(3), 822-830. DOI: 10.1007/s12010-020-03364-0

Merchant, S. S., Prochnik, S. E., Vallon, O., Harris, E. H., Karpowicz, S. J., Witman, G. B., Terry, A., Salamov, A., Fritz-Laylin, L. K., and Marechal-Drouard, L. (2007). “The Chlamydomonas genome reveals the evolution of key animal and plant functions,” Science 318, 245-250. DOI: 10.1126/science.1143609

Mini, P., Demurtas, O. C., Valentini, S., Pallara, P., Aprea, G., Ferrante, P., and Giuliano, G. (2018). “Agrobacterium-mediated and electroporation-mediated transformation of Chlamydomonas reinhardtii: A comparative study,” BMC Biotechnol. 18, article no. 11. DOI: 10.1186/s12896-018-0416-3

Morais, M. G., Santos, T. D., Moraes, L., Vaz, B. S., Morais, E. G., and Costa, J. A. (2022). “Exopolysaccharides from microalgae: Production in a biorefinery framework and potential applications,” Bioresour. Technol. Rep. 18, article ID 101006. DOI: 10.1016/j.biteb.2022.101006

Moreira, J. B., Vaz, B. D. S., Cardias, B. B., Cruz, C. G., Almeida, A. C. A. D., Costa, J. A. V., and Morais, M. G. D. (2022). “Microalgae polysaccharides: An alternative source for food production and sustainable agriculture,” Polysaccharides 3, 441-457. DOI: 10.3390/polysaccharides3020027

Mostafa, S. S. M. (2012). “Microalgal biotechnology: Prospects and applications,” Plant Sci. 12, 275-314. DOI: 10.5772/53694

Narchonai, G., Arutselvan, C., LewisOscar, F., and Thajuddin, N. (2020). “Enhancing starch accumulation/production in Chlorococcum humicola through sulphur limitation and 2,4- D treatment for butanol production,” Biotechnol. Reports 28, article ID e00528. DOI: 10.1016/j.btre.2020.e00528

Nazmi, A., Hauck, R., Davis, A., Hildebrand, M., Corbeil, L., and Gallardo, R. (2017). “Diatoms and diatomaceous earth as novel poultry vaccine adjuvants,” Poult. Sci. 96(2), 288-294. DOI: 10.3382/ps/pew250

Nemcova, Y. (2003). “Detection of cell wall structural polysaccharides by cellulase-gold and chitinase-gold complexes,” Czech Phycol. Olomouc. 3, 31-36.

Ng, I., Tan, S., Kao, P., Chang, Y., and Chang, J. (2017). “Recent developments on genetic engineering of microalgae for biofuels and bio-based chemicals,” Biotechnol. J. 12(10), article ID 1600644. DOI: 10.1002/biot.201600644

Niehaus, T. D., Okada, S., Devarenne, T. P., Watt, D. S., Sviripa, V., and Chappell, J. (2011). “Identification of unique mechanisms for triterpene biosynthesis in Botryococcus braunii,” PNAS 108(30), 12260-12265. DOI: 10.1073/pnas.1106222108

Northcote, D. H., Goulding, K. J., and Horne, R. W. (1960). “The chemical composition and structure of the cell wall of Hydrodictyon africanum Yaman,” Biochem. J. 77(3), 503-508. DOI: 10.1042/bj0770503

Odjadjare, E. C., Mutanda, T., and Olaniran A. O. (2015). “Potential biotechnological application of microalgae: A critical review,” Crit. Rev. Biotechnol. 37(1), 37-52. DOI: 10.3109/07388551.2015.1108956

Olguin, E. J. (2012). “Dual purpose microalgae-bacteria-based systems that treat wastewater and produce biodiesel and chemical products within a biorefinery, Biotechnol. Adv. 30, 1031-1046. DOI: 10.1016/j.biotechadv.2012.05.001

Park, H., Li, X. R., Zhao, Y., Jia, R. L. and Hur, S. (2017). “Rapid development of cyanobacterial crust in the field for combating desertification,” PLOS ONE 12(6), e0179903. DOI: 10.1371/journal.pone.0179903

Parwani, L., Bhatt, M., and Singh, J. (2021). “Potential biotechnological applications of cyanobacterial exopolysaccharides,” Braz. Arch. Biol. Technol. 64, e21200401. DOI: 10.1590/1678-4324-2021200401

Patil, N. P., Le, V., Sligar, A. D., Mei, L., Chavarria, D., Yang, E. Y., and Baker, A. B. (2018). “Algal polysaccharides as therapeutic agents for atherosclerosis,” Front. Cardiovasc. Med. 5, 153. DOI: 10.3389/fcvm.2018.00153

Pekkoh, J., Ruangrit, K. K., Pumas, C., Duangjan, K., Chaipoot, S., Phongphisutthinant, R., Jeerapan, I., Sawangrat, K., Pathomaree, W., and Srinuanpan, S. (2021). “Transforming microalgal Chlorella biomass into cosmetically and nutraceutically protein hydrolysates using high-efficiency enzymatic hydrolysis approach,” Biomass Conv. Bioref. (Online). DOI: 10.1007/s13399-021-01622-7

Pelizer, L. H., de Carvalho, J. C. M., and Moraes, I. O. (2015). “Protein production by Arthrospira (Spirulina) platensis in solid state cultivation using sugarcane bagasse as support,” Biotechnol. Rep. 5, 70-76. DOI: 10.1016/j.btre.2014.12.006

Phang, S. M., Chu, W. L., and Rabiei, R. (2015). “Phycoremediation,” in: The Algae World, Springer, Dordrecht, Netherlands, pp. 357–389.

Pirbazari, S., Norouzi, O., Kohansal, K., and Tavasoli, A. (2019). “Experimental studies on high-quality bio-oil production via pyrolysis of Azolla by the use of a three metallic/modified pyrochar catalyst,” Biores. Technol. 291, article ID 121802. DOI: 10.1016/j.biortech.2019.121802

Popper, Z. A., Michel, G., Herve, C., Domozych, D. S., Willats, W. G. T., Tuohy, M. G., Kloareg, B., and Stengel, D. B. (2011). “Evolution and diversity of plant cell walls: From algae to flowering plants,” Annu. Rev. Plant Biol. 62, 567-590. DOI: 10.1146/annurev-arplant-042110-103809

Priyadarshani, I., and Rath, B. (2012). “Commercial and industrial applications of micro algae – A review,” J. Algal Biomass Util. 3(4), 89-100.

Prussi, M., Buffi, M., Casini, D., Chiaramonti, D., Martelli, F., Carnevale, M., Tredici, M. R., and Rodolfi, L. (2014). “Experimental and numerical investigations of mixing in raceway ponds for algae cultivation,” Biomass Bioenerg. 67, 390-400. DOI: 10.1016/j.biombioe.2014.05.024

Prybylski, N., Toucheteau, C., El Alaoui, H., Bridiau, N., Maugard, T., Abdelkafi, S., Fendri, I., Delattre, C., Dubessay, P., Pierre, G., et al. (2020). “Bioactive polysaccharides from microalgae,” in: Handbook of Microalgae-Based Processes and Products: Fundamentals and Advances in Energy, Food, Feed, Fertilizer, and Bioactive Compounds, Academic Press, Cambridge, MA, USA, pp. 533-571.

Qu, W., Zhang, C., Zhang, Y., and Ho, S.-H. (2019). “Optimizing real swine wastewater treatment with maximum carbohydrate production by a newly isolated indigenous microalga Parachlorella kessleri QWY28,” Bioresour. Technol. 289, article ID 121702. DOI: 10.1016/j.biortech.2019.121702

Quader, M., and Ahmed, S. (2017). “Bioenergy with carbon capture and storage (BECCS): Future prospects of carbon negative technologies,” in: Clean Energy for Sustainable Development, M. G. Rasul, A. K. Azad, and S. C. Sharma (eds.), Academic Press, Cambridge, MA ,USA, pp. 91-140. DOI: 10.1016/B978-0-12-805423-9.00004-1

Rahul, S. M., Sundaramahalingam, M. A., Shivamthi, C. S., Kumar, R. S., Varalakshmi, P., Karthikumar, S., Kanimozhi, J., Kumar, R. V., Sabarathinam, S., Ganesh, M., et al. (2020). “Insights about sustainable biodiesel production from microalgae biomass: A review,” Int. J. Energy. Res. 45(12), 17028-17056. DOI: 10.1002/er.6138

Raja, R., Coelho, A., Hemaiswarya, S., Kumar, P., Carvalho, I. S., and Alagarsamya, A. (2018). “Applications of microalgal paste and powder as food and feed: An update using text mining tool,” Beni-Suef Univ. J. Basic Appl. Sci. 7(4), 740-747. DOI: 10.1016/j.bjbas.2018.10.004

Rajaram, M. G., Nagaraj, S., Manjunath, M., Boopathy, A. B., Kurinjimalar, C., Rengasamy, R., Jayakumar, T., Sheu, J., and Li, J. (2018). “Biofuel and biochemical analysis of Amphora coffeaeformis RR03, a novel marine diatom, cultivated in an open raceway pond,” Energies 11(6), article no. 1341. DOI: 10.3390/en11061341

Rajasekar, P., Palanisamy, S., Anjali, R., Vinosha, M., Elakkiya, M., Marudhupandi, T., Tabarsa, M., You, S., and Prabhu, N. M. (2019). “Isolation and structural characterization of sulfated polysaccharide from Spirulina platensis and its bioactive potential: In vitro antioxidant, antibacterial activity and Zebrafish growth and reproductive performance,” Int. J. Biol. Macromol. 141, 809-821. DOI: 10.1016/j.ijbiomac.2019.09.024.

Rajkumar, R., Takriff, M. S., and Veeramuthu, A. (2022). “Technical insights into carbon dioxide sequestration by microalgae: A biorefinery approach towards sustainable environment,” Biomass Conv. Bioref. (Online). DOI: 10.1007/s13399-022-02446-9

Ramírez-Lopez, C., Perales-Vela, H. V., and Ernández-Linares, L. (2019). “Biomass and lipid production from Chlorella vulgaris UTEX 26 cultivated in 2m3 raceway ponds under semicontinuous mode during the spring season,” Bioresour. Technol. 274, 252-260. DOI: 10.1016/j.biortech.2018.11.096

Raposo, M. F., Morais, R. M. S., and Morais, A. M. M. B. (2013). “Health applications of bioactive compounds from marine microalgae,” Life Sci. 95(15), 479-486. DOI: 10.1016/j.lfs.2013.08.002

Rempel, A., Biolchi, G. N., Antunes, A. C. F., Gutkoski, J. P., Treichel, H., and Colla, L. M. (2021). “Cultivation of microalgae in media added of emergent pollutants and effect on growth, chemical composition, and use of biomass to enzymatic hydrolysis,” Bioenergy Res. 14, 265-277. DOI: 10.1007/s12155-020-10177-w

Santhosh, S., Dhandapani, R., and Hemalatha, R. (2016). “Bioactive compounds from Microalgae and its different applications-a review,” Adv. Appl. Sci. Res. 7(4), 153-158.

Santiago-Morales, I. S., Trujillo-Valle, L., Marquez-Roch, F. J., and Lopez-Hernández, J. F. (2018). “Tocopherols, phycocyanin and superoxide dismutase from microalgae: As potential food antioxidants,” Appl. Food Biotechnol. 5(1), 19-27. DOI: 10.22037/afb.v5i1.17884

Sarkar, S., Mankad, J., Padhihar, N., Manna, M. S., Bhowmick, T. K., and Gayen, K. (2022). “Enhancement of growth and biomolecules (carbohydrates, proteins, and chlorophylls) of isolated Chlorella thermophila using optimization tools,” Prep. Biochem. Biotechnol. (Online), 1-17. DOI: 10.1080/10826068.2022.2033995

Sarpal, A. S., Teixeira, C. M. L. L., Silva, P. R. M., Monteiro, T. V., Itacolomy, J., Smarcaro, V., and Daroda, R. J. (2015). “NMR techniques for determination of lipid content in microalgal biomass and their use in monitoring the cultivation with biodiesel potential,” Appl. Microbiol. Biotechnol. 100(5), 2471-2485. DOI: 10.1007/s00253-015-7140-x

Sathasivam, R., and Juntawong, N. (2013). “Modified medium for enhanced growth of Dunaliella strains,” Int. J. Cur. Sci. 5, 67-73

Sathasivam, R., Radhakrishnan, R., Hashem, A., and Abd-Allah, E. F. (2017). “Microalgae metabolites: A rich source for food and medicine,” Saudi J. Biol. Sci. 26(4), 709-722. DOI: 10.1016/j.sjbs.2017.11.003

Scholz, M. J., Weiss, T. L., Jinkerson, R. E., Jing, J., Roth, R., Goodenough, U., Posewitz, M. C., and Gerken, H. G. (2014). “Ultrastructure and composition of the Nannochloropsis gaditana cell wall,” Eukaryot. Cell 13, 1450-1464. DOI: 10.1128/EC.00183-14

Simas-Rodrigues, C., Villela, H. D. M., Martins, A. P., Marques, L. G., Colepicolo, P., and Tonon, A. P. (2015). “Microalgae for economic applications: Advantages and perspectives for bioethanol,” J. Exp. Bot. 66(14), 4097-4108. DOI: 10.1093/jxb/erv130

Singh, H., Varanasi, J. L., Banerjee, S., and Das, D. (2019). “Production of carbohydrate enrich microalgal biomass as a bioenergy feedstock,” Energy 188, article ID 116039. DOI: 10.1016/j.energy.2019.116039

Singh, R., Khan, M. J., Rane, J., Gajbhiye, A., Vinayak, V., and Joshi, K. B. (2020). “Biofabrication of diatom surface by tyrosine-metal complexes: Smart microcontainers to inhibit bacterial growth,” Chem. Select 5, 3091-3097. DOI: 10.1002/slct.201904248

Sirohi, R., Choi, H., and Sim, S. J. (2022b). “Microalgal fuels: Promising energy reserves for the future,” Fuel 312, article ID 122841. DOI: 10.1016/j.fuel.2021.122841

Sirohi, R., Pandey, A. K., Ranganathan, P., Singh, S., Udayan, A., Awasthi, M. K., Hoang, A. T., Chilakamarry, C. R., Kim, S. H., and Sim, S. J. (2022a). “Design and applications of photobioreactors – A review,” Bioresour. Technol. 349, article ID 126858. DOI: 10.1016/j.biortech.2022.126858

Solis-Salinas, C. E., Patlan-Juarez, G., Okoye, P. U., Guillen-Garces, A., Sebastian, P. J., and Arias, D. M. (2021). “Long-term semi-continuous production of carbohydrate-enriched microalgae biomass cultivated in low-loaded domestic wastewater,” Sci. Tot. Env. 798, article ID 149227. DOI: 10.1016/j.scitotenv.2021.149227

Steinbrenner, J., and Sandmann, G. (2006). “Transformation of the green alga Haematococcus pluvialis with a phytoene desaturase for accelerated astaxanthin biosynthesis,” Appl. Environ. Microbiol. 72(12), 7477-7484. DOI: 10.1128/AEM.01461-06

Steinrucken, P., Prestegard, S. K., Vree, J. H., Storesund, J. E., Pree, B., Mjos, S. A., and Erga, S. R. (2018). “Comparing EPA production and fatty acid profiles of three Phaeodactylum tricornutum strains under western Norwegian climate conditions,” Algal Res. 30, 11-22. DOI: 10.1016/j.algal.2017.12.001

Stolz, P., and Obermayer, B. (2013). “Manufacturing microalgae for skin care,” Cosmetics and Toiletries 120, 99-106.

Sun, X. W., Zhang, Y. X., and Losic, D. (2017). “Diatom silica, an emerging biomaterial for energy conversion and storage,” J. Mater. Chem. 5, 8847-8859. DOI: 10.1039/C7TA02045G

Sun, X., Wang, X., and Liu, J. (2019). “Screening of Isochrysis strains for simultaneous production of docosahexaenoic acid and fucoxanthin,” Algal Res. 41, article ID 101545. DOI: 10.1016/j.algal.2019.101545

Surendhiran, D., and Sirajunnisa, A. R., and Vijay, M. (2015). “An alternative method for production of microalgal biodiesel using novel Bacillus lipase,” Biotechnology 5, 715-725. DOI: 10.1007/s13205-014-0271-4

Tafreshi, A. H., and Shariati, M. (2009). “Dunaliella biotechnology: Methods and applications,” J. Appl. Microbiol. 107(1), 14-35. DOI: 10.1111/j.1365-2672.2009.04153.x

Tan, C. H., Show, P. L., Chang, J-S., Ling, T. C., and Lan, J. C-W. (2015). “Novel approaches of producing bioenergies from microalgae: A recent review,” Biotechnol. Adv. 33, 1219-1227. DOI: 10.1016/j.biotechadv.2015.02.013

Tan, J. S., Leec, S. Y., Chewd, K. W., Lame, M. K., Lim, J. W., Hoh, S., and Show, P. L. (2020). “A review on microalgae cultivation and harvesting, and their biomass extraction processing using ionic liquids,” Bioengineered 11(1), 116-129. DOI: 10.1080/21655979.2020.1711626

Tarento, T. D. C., McClure, D. D., Vasiljevski, E., Schindeler, A., Dehghani, F., and Kavanagh, J. M. (2018). “Microalgae as a source of vitamin K1,” Algal Res. 36, 77-87. DOI: 10.1016/j.algal.2018.10.008

Tibbetts, S. M., Milley, J. E., and Lall, S. P. (2015). “Chemical composition and nutritional properties of freshwater and marine microalgal biomass cultured in photobioreactors,” J. Appl. Phycol. 27, 1109-1119. DOI: 10.1007/s10811-014-0428-x

Tiwari, O. N., Mondal, A., Bhunia, B., Bandyopadhyay, T. K., Jaladi, P., Oinam, G., and Indrama, T. (2019). “Purification, characterization and biotechnological potential of new exopolysaccharide polymers produced by cyanobacterium Anabaena sp. CCC 745,” Polymer 178, article no. 121695. DOI: 10.1016/j.polymer.2019.121695

Udayan, A., Sirohi, R., Sreekumar, N., Sang, B. I., and Sim, S. J. (2022). “Mass cultivation and harvesting of microalgal biomass: Current trends and future perspectives,” Bioresour. Technol. 344, article ID 126406. DOI: 10.1016/j.biortech.2021.126406

Ververis, C., Georghiou, K., Danielidis, D., Hatzinikolaou, D. G., Santas, P., Santas, R., and Corleti, V. (2007). “Cellulose, hemicelluloses, lignin and ash content of some organic materials and their suitability for use as paper pulp supplements,” Bioresour. Technol. 98, 296-301. DOI: 10.1016/j.biortech.2006.01.007

Wang, B., Lan, C. Q., and Horsman, M. (2012). “Closed photobioreactors for production of microalgal biomasses,” Biotechnol. Adv. 30, 904-912. DOI: 10.1016/j.biotechadv.2012.01.019

Wang, F., Gao, B., Su, M., Dai, C., Huang, L., and Zhang, C. (2019). “Integrated biorefinery strategy for tofu wastewater biotransformation and biomass valorization with the filamentous microalga Tribonema minus,” Bioresour. Technol. 292, article ID 121938. DOI: 10.1016/j.biortech.2019.121938

Wang, Y., Ho, S. H., Cheng, C. L., Nagarajan, D., Guo, W. Q., Lin, C., Li, S., Ren, N., and Chang, J. S. (2017). “Nutrients and COD removal of swine wastewater with an isolated microalgal strain Neochloris aquatica CL-M1 accumulating high carbohydrate content used for biobutanol production,” Bioresour. Technol. 242, 7-14. DOI: 10.1016/j.biortech.2017.03.122

Wang, Y., Jia, J., Chi, Q., Li, Y., Wang, H., Gong, Y., Liu, G., Hu, Z., Han, D., and Hu, Q. (2022). “Critical assessment of the filamentous green microalga Oedocladium carolinianum for astaxanthin and oil production,” Algal Res. 61, article ID 102599. DOI: 10.1016/j.algal.2021.102599

Williams, M., Kookana, R. S., Mehta, A., Yadav, S. K., Tailor, B. L., and Maheshwari, B. (2019). “Emerging contaminants in a river receiving untreated wastewater from an Indian urban centre,” Sci. Total Environ. 647, 1256-1265. DOI: 10.1016/j.scitotenv.2018.08.084

Xia, S., Gao, B., Li, A., Xiong, J., Ao, Z., and Zhang, C. (2014). “Preliminary characterization, antioxidant properties and production of chrysolaminarin from marine diatom Odontella aurita,” Mar. Drugs 12(9), 4883-4897. DOI: 10.3390/md12094883

Xu, L., Weathers, P. J., Xiong, X. R., and Liu, C. (2009). “Microalgal bioreactors: Challenges and opportunities,” Eng. Life Sci. 9, 178-189. DOI: 10.1002/elsc.200800111

Xu, S. Y., Huang, X., and Cheong, K. L. (2017). “Recent advances in marine algae polysaccharides: Isolation, structure, and activities,” Mar. Drugs 15(12), 388. DOI: 10.3390/md15120388.

Yap, B. H. J., Crawford, S. A., Dagastine, R. R., Scales, P. J., and Martin, G. J. O. (2016). “Nitrogen deprivation of microalgae: Effect on cell size, cell wall thickness, cell strength, and resistance to mechanical disruption,” J. Ind. Microbiol. Biotechnol. 43, 1671-1680. DOI: 10.1007/s10295-016-1848-1

Yi, Z., Su, Y., Brynjolfsson, S., Olafsdottir, K., and Fu, W. (2021). “Bioactive polysaccharides and their derivatives from microalgae: Biosynthesis, applications, and challenges,” Stud. Nat. Prod. Chem. 71, 67-85. DOI: 10.1016/B978-0-323-91095-8.00007-6

Yi, Z., Xu, M., Di, X., Brynjolfsson, S., and Fu, W. (2017). “Exploring valuable lipids in diatoms,” Front. Mar. Sci. 4, 17. DOI: 10.3389/fmars.2017.00017

Yu, K. L., Show, P. L., Ong, H. C., Ling, T. C., Chi-Wei Lan, J., Chen, W., and Chang, J. (2017). “Microalgae from wastewater treatment to biochar – Feedstock preparation and conversion technologies,” Energy Convers. Manag. 150, 1-13. DOI: 10.1016/j.enconman.2017.07.060

Zanchetta, E., Damergi, E., Patel, B., Borgmeyer, T., Pick, H., Pulgarin, A., and Ludwig, C. (2021). “Algal cellulose, production and potential use in plastics: Challenges and opportunities,” Algal Res. 56, article ID 102288. DOI: 10.1016/j.algal.2021.102288

Zhang, H., Tang, Y., Zhang, Y., Zhang, S., Qu, J., Wang, X., Kong, R., Han, C., and Liu, Z. (2015). “Fucoxanthin: A promising medicinal and nutritional ingredient,” J. Evid. Based Complementary Altern. 2015, article ID 723515 DOI: 10.1155/2015/723515

Zhang, W., Wang, F., Gao, B., Huang, L., and Zhang, C. (2018). “An integrated biorefinery process: Stepwise extraction of fucoxanthin, eicosapentaenoic acid and chrysolaminarin from the same Phaeodactylum tricornutum biomass,” Algal Res. 32, 193-200. DOI: 10.1016/j.algal.2018.04.002

Zhou, W., Wang, Z., Alam, M., Xu, J., Zhu, S., Yuan, Z., Huo, S., Guo, Y., Qin, L., and Ma, L. (2019). “Repeated utilization of ionic liquid to extract lipid from algal biomass,” Int. J. Polym. Sci. 2019, article ID 9209210. DOI: 10.1155/2019/9209210

Article submitted: September 16, 2022; Peer review completed: October 1, 2022; Revised version received and accepted October 19, 2022; Published: October 28, 2022.

DOI: 10.15376/biores.17.4.Harini