Next Article in Journal
The Purification and Biochemical Characterization of a Weissella cibaria F1 Derived β-Mannanase for Its Use in the Preparation of Konjac Oligo-Glucomannan with Immunomodulatory Properties
Next Article in Special Issue
The Biorefinery of the Marine Microalga Crypthecodinium cohnii as a Strategy to Valorize Microalgal Oil Fractions
Previous Article in Journal
Conversion of Syngas from Entrained Flow Gasification of Biogenic Residues with Clostridium carboxidivorans and Clostridium autoethanogenum
Previous Article in Special Issue
Effects of Fermented Seaweed Fertilizer Treatment on Paddy Amino Acid Content and Rhizosphere Microbiome Community
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Marine Microalgae Production: Highlighting Human Health Products from Microalgae in View of the Coronavirus Pandemic (COVID-19)

by
Mohamed Ashour
1,* and
Ashraf M. M. Omran
2
1
National Institute of Oceanography and Fisheries (NIOF), Cairo 11516, Egypt
2
Sustainable Development of Aquatic Organisms Foundation (SDAO), Alexandria 21533, Egypt
*
Author to whom correspondence should be addressed.
Fermentation 2022, 8(9), 466; https://doi.org/10.3390/fermentation8090466
Submission received: 30 July 2022 / Revised: 1 September 2022 / Accepted: 16 September 2022 / Published: 18 September 2022
(This article belongs to the Special Issue Marine-Based Biorefinery: A Path Forward to a Sustainable Future)

Abstract

:
Blue biotechnology can greatly help solve some of the most serious social problems due to its wide biodiversity, which includes marine environments. Microalgae are important resources for human needs as an alternative to terrestrial plants because of their rich biodiversity, rapid growth, and product contributions in many fields. The production scheme for microalgae biomass mainly consists of two processes: (I) the Build-Up process and (II) the Pull-Down process. The Build-Up process consists of (1) the super strain concept and (2) cultivation aspects. The Pull-Down process includes (1) harvesting and (2) drying algal biomass. In some cases, such as the manufacture of algal products, the (3) extraction of bioactive compounds is included. Microalgae have a wide range of commercial applications, such as in aquaculture, biofertilizer, bioenergy, pharmaceuticals, and functional foods, which have several industrial and academic applications around the world. The efficiency and success of biomedical products derived from microalgal biomass or its metabolites mainly depend on the technologies used in the cultivation, harvesting, drying, and extraction of microalgae bioactive molecules. The current review focuses on recent advanced technologies that enhance microalgae biomass within microalgae production schemes. Moreover, the current work highlights marine drugs and human health products derived from microalgae that can improve human immunity and reduce viral activities, especially COVID-19.

1. Introduction

Recently, aquatic industries have been developed successfully [1,2], as well as aquatic products (aquatic plants and aquatic animals), which have great benefits for human health, immunity, and nutrition [3,4]. Aquatic plants (microalgae and seaweed) are photosynthetic organisms that consume carbon, nutrients, and solar energy and convert them into organic compounds, such as proteins, lipids, carbohydrates, and pigments [5], which may be transformed into several bioproducts. The production of aquatic plant biomass is an important branch of aquaculture activities.
Aquatic plant biomass has a wide range of organic materials of commercial interest, such as antioxidants, biopolymers, pigments, polysaccharides, and biopeptides [6]. Aquatic plant biomass has received attention at the commercial and academic levels due to its multiple applications in several fields, including food supplements [4,7,8], pharmaceuticals [3,9], cosmetics [7,10,11], aquaculture [2,12,13,14,15,16,17], biofuel [2,18,19], and biofertilizers [20,21,22,23,24]. Moreover, algae biomass can produce many bioenergy sources, such as biomethane, bioethanol, biodiesel, biohydrogen, and bio-oil [25,26,27,28,29,30]. However, concerning the environmental applications of algae biomass, a wide range of polluted wastewater types, including sewage, industrial, agriculture, and aquaculture effluents, can also be treated through the phytoremediation mechanism [31,32,33,34,35,36,37,38]. Furthermore, algae biomass can effectively reduce highly toxic compounds to less biologically toxic compounds, producing a variety of high-value products [35,36,37,39,40].
According to the statistics report of the Food and Agriculture Organization (FAO) [41], in 2003, the biomass of Arthrospira (Spirulina) produced in China only was 16,483 tons (Ts). The global biomass production of microalgae was 93,756 Ts in 2010 and 87,000 Ts in 2018. In 2019, the total global production of microalgae biomass was 56,465 Ts, which was produced by the top ten producers in the world (China 97.16%, Chile 1.6%, France 0.37%, Greece 0.25%, Tunisia 0.25%, Burkina Faso 0.25%, Central African Republic 0.09%, Chad 0.04%, Bulgaria 0.005%, and Spain 0.003%). FAO statistics reported that because of their commercial attitude, they do not capture the true production of microalgae in some countries, including the United States of America, Australia, Czechia, Iceland, India, Italy, Japan, Malaysia, and Myanmar [42]. In 2019, FAO statistics also reported that commercial microalgae biomass contributes less than 0.2% of the global production of aquatic biomass (microalgae and seaweed). Arthrospira (Spirulina) comprises 96.56% of global microalgae biomass production, while the remaining (0.44%) comes from four species of green microalgae: Haematococcus pluvialis (242 Ts, 0.429%), Chlorella vulgaris (4.77 Ts, 0.008%), Tetraselmis sp. (1.45 Ts, 0.003%), and Dunaliella salina (0.22 Ts, 0.0004%) [41].
From laboratory scale to commercial production, biomass from Arthrospira, Chlorella, Haematococcus, and Nannochloropsis has been produced globally in many countries to serve several industries, including biofertilizers, animal feeds, aqua-feed, and human food supplements [43]. In terms of algae biomass volume, Arthrospira, followed by Chlorella, is the most productive species [44]. These two species are the most promising microalgae for energy production, including biogas, bioethanol, and biodiesel. Some species, such as Tetraselmis sp., Isochrysis sp., Dunaliella sp., Phaeodactylum sp., and Scenedesmus sp., also have high industrial importance, while some species, such as Chaetoceros sp., Thalassiosira sp., and Acutodesmus sp., are of low industrial importance [45,46].
The main advantages of marine microalgae are greater than the advantages of freshwater microalgae or terrestrial plants. There is no competition between marine microalgae and freshwater microalgae or terrestrial plants in the utilization of freshwater sources or arable land. However, freshwater, and marine microalgae are more efficient in carbon sequestration and growth rates than terrestrial plants [47,48].
Microalgae biomass production technologies have unique characteristics that make them more suitable for the production of certain species and for commercial purposes. Their land requirements, construction, operational expenses, technical development, and environmental parameter management vary according to the technological method used [49]. Although microalgae biomass has shown environmental benefits, it also has serious consequences related to rising energy, water supply, wastewater management, land use, and the risk of microbial contamination [50,51,52]. Microalgae biomass production techniques that are cost-effective, environmentally friendly, and practical are urgently needed on a large scale [53]. Despite the widespread applications of microalgae, there are many obstacles to the development and promotion of biomass production technologies [54]. The microalgae biomass production scheme mainly consists of two processes: (I) the Build-Up process and (II) the Pull-Down process. Generally, the Build-Up process consists of (1) the super strain concept and (2) cultivation, while the Pull-Down process consists of (1) harvesting, (2) drying, and (3) extraction [55].
Microalgae biomass production technologies have huge potential for innovation, by using low-cost techniques that expand and revolutionize algae industries, especially for human health products. Currently, the coronavirus epidemic (COVID-19) remains the most significant health concern in the world, and it is critical to discover effective drugs to stop this epidemic from killing hundreds of thousands of people [56]. To combat COVID-19, scientists, professionals, and global health organizations have been working together to develop and implement quick diagnoses, reliable vaccinations, and treatment approaches [57]. Some microalgae-derived bioactive compounds and metabolites can successfully treat many diseases, while others can improve human immunity and decrease viral activities, especially COVID-19 [57,58,59]. The effectiveness of these suggested bioactive molecules obtained from microalgae biomass in combating viral diseases varies depending on the technology used for biomass production, which mainly consists of two stages: (I) the Build-Up process and (II) the Pull-Down process.
Here, the current review aimed to provide an overview of the recent advanced technologies that enhance the microalgae biomass within the microalgal biomass production scheme, starting with strain selection, genetic optimization, culture systems, growth conditions, nutrient limitations, and harvesting, drying, and extraction techniques. Moreover, the current work focuses on marine drugs and human health products from microalgae that can improve human immunity and reduce viral activities, including COVID-19.

2. Build-Up Process

2.1. Super-Strain Concept

The greatest challenge in the algae biomass industry is finding suitable strains for industrial applications. To achieve the “super strain concept”, several steps should be performed, such as algal strain selection, morphological and molecular identification, optimum growth condition determination, growth curve determination, and high biomass yield production under different culture conditions [60].

2.1.1. Strain Selection

According to algaebase.org [61], there are more than 150,000 species of microalgae around the world; however, only a few species have been studied in terms of their beneficial applications [55,60,62]. There are more than 37 microalgae culture collections (collections, centers, and seeds banks) around the world, mostly in the USA and the EU. Microalgae species differ in their biochemical compositions; therefore, algae collections isolate and identify several strains for their purposes. To reduce the culture period of microalgae, cultured strains should be robust and resistant to unsuitable stress conditions (abiotic conditions), such as light, temperature, and salinity, as well as resistant to biotic stress conditions, such as herbivores and pathogens [63]. To select the best-performing microalgae strains, some standard criteria are required, including (1) high biomass production rates, (2) pathogen resistance, (3) optimization to growth media, (4) a high rate of CO2 capture and exchange, (5) adaptation to several environmental culture conditions, and (6) enhancing performance characteristics through the applications of genetic engineering, breeding, and genome editing [60]. Recently, many studies have attempted to enhance the biomass yield of the required molecules, mainly lipids, carbohydrates, proteins, and other valuable bioactive compounds, under various stressful conditions at laboratory levels. However, it is important to develop a microalgae culture of effluent wastewater to produce high value-added products at mass production levels [55].

2.1.2. Strain Transformation

Microalgae are a viable option for genetic engineering because they are unicellular and have a simple genetic structure with quick reproduction capabilities [55]. Microalgae’s bioactive compounds and metabolites can be enhanced by genetic engineering techniques to produce them more reliably, as well as by increasing their volume and concentration. These techniques are applied to increase bioactive compound and metabolite production for the transformation of the substrate into the target product, especially in the field of human health products [64].
Recently, genetic modification techniques have been successfully used in the microalgae industry due to their potential to improve microalgae biomass production [65]. Important results have been reported in microalgae genomics, including the development of efficient gene-transformation systems, engineering, and biomass optimization [60]. The microalgae cell wall is the key step in achieving the transformation process. As a result, protoplast or enzyme-treated cells are used in microalgae transformation methods [7]. In general, these techniques include biotic and abiotic stresses (Figure 1), such as transferring genes isolated from one species to another to produce breeding strains with desirable commercial characteristics, such as excessive tolerance to light and heat stress, resistance to herbivores and pathogens, and the ability to compete against opportunistic organisms or express biosynthetic pathways toward more profitable strains [66].
Chlamydomonas reinhardtii, a plant model organism, is the most extensively studied microalgae strain for genetic modification [67]. This species has been successfully utilized as a source of enzymes and feed additives [68,69], nutrient supplements, such as selenium [70], gut-active biologics [71], antibodies and immunotoxins [72], and vaccine subunits [73]. However, this species is not used in aquaculture. A. platensis, C. calcitrans, C. vulgaris, and Nannochloropsis sp. are the major microalgae species that have been extensively studied to produce genetically modified species as aquaculture feed, improving their nutritional value and increasing the market value of their products [74,75,76,77,78].
Prasad et al. [79] demonstrated stable Agrobacterium-mediated DNA transfer in the nuclear genomes of two strains of haptophytes (I. galbana and Isochrysis sp.). First, Isochrysis sp. and I. galbana were co-cultivated with the A. tumefaciens strain LBA 4404 carrying pCAMBIA 1380-pds-L504R to evaluate their resistance to the herbicide norflurazon and the antibiotic hygromycin. This work shows that, in the case of 200 M acetosyringone, Isochrysis sp. co-cultivated with Agrobacterium resulted in a huge number of norflurazon-resistant cells, whereas I. galbana did the same in the case of 100 M acetosyringone. Despite the lack of pre-treatment before co-cultivation with Agrobacterium, the vector could still enter Isochrysis and penetrate the cell membrane’s calcium-rich barrier. Prasad [80] confirmed that the metabolic engineering of Pavlova lutheri was successively performed using the Agrobacterium-mediated nuclear transformation protocol.
Genetically modified microalgae can be used to produce other recombinant proteins. Kim et al. [81] reported that Flounder fry that fed on transformed Chlorella ellipsoidea had a 25% increase in growth after 30 days of feeding, demonstrating that the recombinant protein can be transferred up the food chain while maintaining function. Successful transgenic studies have been conducted in Dunalella salina and Haematococcus pluvialis to enhance carotenoid accumulation. Vaccines against white spot disease that affect a variety of crustaceans have been effectively developed by Dunaliella salina [73,82]. Fayyaz et al. [65] confirmed that microalgal pigment production is increased by transformed cartenogenic genes, such as bkt, psy, chyB, and pds, while microalgae lipid content may also increase by the overexpression of enzymes, such as LPAT, DGAT2, GPAT, and MAT.
Way-Rong et al. [83] demonstrated that in Chlorella sorokiniana and Chlorella vulgaris, the overexpression of the enzyme carbonic anhydrase is a potential method for effectively capturing excess CO2. Transgenic microalgae species of Chlorella sorokiniana and Chlorella vulgaris with an exogenous MlCA gene showed an increased biomass, protein percentage, and lipid accumulation. The same findings were reported by Wei et al. [84], who concluded that the biomass productivity of Nannochloropsis oceanica was improved by the overexpression of a nuclear-encoded, cbbX-homologous, candidate RuBisCO activase (g1915 or ‘nNoRca-like’) that is transcriptionally stimulated at air-level CO2. However, the microalgal transformation of either nuclear or chloroplast genomes is achievable (Figure 2) using several methods, including Agrobacterium-mediated methods, electroporation, enzyme-mediated methods, silicon carbide whiskers, glass beads, and microprojectile bombardment [55]. Table 1 details the history of the most commercially important microalgae strains that have been genetically transformed using the respective strategies.

2.2. Cultivation Aspects

Microalgae biomass is an attractive, renewable, and sustainable source for many bio-products, such as drugs, pharmaceuticals, and food supplements [96]. The biochemical composition and biomass productivity of microalgae vary depending on many factors, including cultivation systems, environmental growth conditions, nutrient availability and limitations, the interaction of microalgae and bacteria, and the type of cultured species [97]. Any changes in cultivation aspects lead to a change in the biochemical composition of microalgae, which may significantly affect the productivity of microalgae biomass. Advances in the development of microalgae biomass require appropriate conditions. However, the percentages of microalgae biochemical compositions are positively or negatively affected by these parameters [98]. Carbohydrates contribute 15–40% of the biomass of microalgae, which is a direct result of CO2 capture via the Calvin cycle. Carbohydrates exist in several forms, such as polysaccharides (found in cell walls), starch (found in plastids), glycogen, cellulose, and other similar substances [99]. Microalgae are a rich source of protein and contribute 30–70% of the algal biomass, having nutritional benefits higher than traditional protein sources, such as eggs, meat, milk, legumes, and soybeans [53]. During the exponential growth phase, microalgae tend to accumulate a lot of protein [100]. Lipids are the most important components of microalgae, providing about 5–65% of the algal biomass, and they have higher nutritional benefits than other traditional sources [55,60].

2.2.1. Cultivation Systems

The cultivation system is the most significant key that controls microalgae biomass productivity. Globally, two microalgae culture systems have been implemented: an open pond (OP) raceway and a closed pond (photobioreactors, PBR). These types are effective in producing high-value biomass from microalgae, and they also help reduce environmental stress by sequestering CO2 [101].
Based on the literature, it is difficult to evaluate the performance of different outdoor systems due to their different locations and operating systems [102]. In the open pond culture system, microalgae utilize atmospheric CO2 and sunlight. Circular ponds, inclined systems, and raceway ponds are the most popular open pond forms for microalgae cultivation [103]. Photobioreactors operate under controlled conditions. In terms of pH, medium agitation, and light intensity, the regulated growth environment of the PBR ensures consistency. After a complicated chemical process, H2O and CO2 are converted into algal biomass rich in oil content in the PBR system [104].
However, some authors have indicated that the hybrid system is a third culture system that uses both OP and PPR [105,106]. In general, a PBR system is more cost-effective than an OP system in terms of its operation, maintenance, and energy consumption. However, in terms of pH, medium agitation, and light intensity, the PBR has a greater control over biomass production than the open pond method. As a result, the microalgae biomass productivity of PBR is much higher [105].
Carotenoids have increased demands and a variety of commercial applications, especially in pharmaceuticals, with great potential as antioxidants. Prieto et al. [107] studied the effect of the culture system (tubular PBR or OP) and feeding regime (batch and semi-continuous regime) on carotenoid productivity and β-carotene abundance of Dunaliella salina. The highest carotenoid productivity (10% DW) and β-carotene abundance (90% of the total carotenoids) were obtained using a PBR culture system, rather than OP. Because of its potential application in pharmaceuticals, drugs, and human nutrition sectors, astaxanthin has a high antiviral and antioxidant capacity that is widely employed in several industries. Table 2 shows a comparison between the two types of microalgae cultivation systems.
According to Li et al. [109], natural astaxanthin may be less expensive than synthetic astaxanthin, using recently developed low-cost PBR and OP methodologies. According to a comprehensive financial analysis, Haematococcus biomass and astaxanthin production have lower costs than USD 18 and 718 kg−1, respectively. Regarding the economic and environmental aspects, the use of wastewater effluent as an alternative culture medium provides the nutrients needed for microalgae production, in addition to an increase in biomass and astaxanthin production [110,111].

2.2.2. Stress via Environmental Growth Conditions

Temperature is one of the main factors controlling microalgae biomass. If the temperature changes outside of the optimal range, it may cause the overgrowth, inhibition, or death of microalgae cells [105]. In many cases, higher temperatures enhance biomass production, especially carbohydrates, which may be related to the participation of the thermophilic enzyme system in sucrose generation [112], while lower temperatures generally reduce carboxylase activities [113]. Most microalgae species can survive at an ideal temperature of 20–30 °C.
Covarrubias et al. [114] reported that thermophilic algae, such as Anacystis nidulans and Chaetoceros, can survive at temperatures of up to 40 °C [114]. In the culture of microalgae under uncontrolled conditions, such as their cultivation in an open pond, the surface temperature may be higher than 40 °C, and a loss of culture water will be obtained. This significant problem can be dealt with when using controlled culture conditions, such as indoor cultivation or PBR. If indoor cultivation and/or PBR are conducted in hot areas, an additional cooling unit must be provided. Therefore, it makes the culture expensive. As a result, the development of biomass production is limited [105].
Ferro et al. [115] observed that microalgae biomass tends to be higher in summer than winter, while the cold-stress tolerance of Desmodesmus sp. and Scenedesmus sp. was higher when grown in Nordic culture conditions [115]. Temperature has been used as a stress therapy to stimulate the production of beneficial metabolites. C. vulgaris cultivation at 25 °C produced higher carbohydrate and lipid levels than cultivation at 30 °C [116]. Most studies have focused on lipid accumulation under high temperatures, and there are few studies on the effect of temperature on carbohydrate and protein accumulation; thus, more studies on this topic are still needed [55].
Another important factor controlling microalgae biomass production is radiation, which is a primary energy source for photosynthesis and has a direct impact on the growth rate, cell composition, and CO2 fixation rate [7]. Muhammad et al. [55] demonstrated a well-documented linear relationship between biomass productivity and carbohydrate content with increasing radiation intensity, although the effect of light intensity varies by species. Increasing radiation above the optimal level may affect photosynthesis, while biomass saturation reduces light penetration due to self-shading. Gifuni et al. [117] reported that a lower light intensity (less than 275 μmol m2 s−1) may result in decreased carbohydrate production over time. The key enzyme for carbohydrate production (called phosphoglucomutase) may be regulated by radiation.
De Farias Silva et. al. [118] found that under high light intensity (252–364 μmol m−2 s−1) and high temperature (28 °C), the biomass of Pseudoneochloris marina (260 mg L−1 day−1) improved. However, these conditions did not appear optimal for protein, lipids, and pigment production. In contrast, under low light intensity (140 μmol m−2 s−1) and low temperature (20 °C), the maximum protein content (236 mg g−1) was observed. To protect cells from stress, as well as in response to high light intensity, microalgae tend to synthesize high-energy compounds, such as carbohydrates and lipids [55].
Both the quality and quantity of radiation are important keys affecting the performance of the photosynthetic process [119]. García-Cubero et al. [120] demonstrated that, in the upper layer of a dense culture, while photoinhibition increased, cells in the lower layers may already be light-deprived when light penetrates the culture, resulting in a decreased biomass production. In the case of dense cultures, while photoinhibition may occur in the upper layer, cells in the lower layers may be light-deprived due to light attenuation as they penetrate the culture, resulting in a reduced biomass production [120]. As reported by El-Khouly et al. [121], LED light is an emerging and cost-effective technology in microalgae production due to its lower power consumption, longer life, lower heat dissipation, smaller mass and size, less heat generation when supplying light, longer life span, especially in comparison to fluorescent lamps, and higher conversion efficiency. Mohsenpour and Willoughby [122] observed that red light enhanced the biomass production of both C. vulgaris and G. membranacea (0.135 g L−1 day−1 and 0.184 g L−1 day−1, respectively). An enhanced growth rate was observed in N. salina cultured in PBR using a white LED (400 to 780 nm) by achieving a maximum growth rate of 150 μmol photons m−2 s−1 at 0.521 day−1. Using blue light, Nannochloropsis sp. had the highest specific growth rate, at 0.64 day−1 [123]. The effects of radiation on microalgae biomass productivity and carbohydrate content, as the most affected compound, have been extensively studied, as presented in Table 3.

2.2.3. Stress via Nutrient Availability and/or Limitations

Nitrogen, potassium, and phosphorus (macronutrients) have a primary effect on the growth of microalgae; however, nutrient limitation or availability (starvation) is an applied strategy to enhance the metabolite composition and biomass production of microalgae [55]. During the Calvin cycle, nitrogen and phosphorus limitations can convert fixed carbon, directing it toward the production of non-nitrogenous compounds, such as lipids and carbohydrates, especially polysaccharides, starch, and glycogen. When Chlorella was grown under nitrogen conversion from sufficient conditions to nitrogen-starved conditions, carbohydrate content increased from 12 to 54%, while proteins decreased to 20% from 60% [129]. Similarly, when Chlorella sp. was cultured under phosphorus limitation conditions, the carbohydrate content increased from 10 to 60%, while the protein content decreased from 57% to 7% [130].
Microalgae culture medium should be cost-effective, develop and meet the requirements of microalgae cells, and be simple to produce. In marine hatcheries, the production cost of microalgae as livefeeds is about 30% of the total cost of larval production, which is a constraint to marine hatchery development. Therefore, forms of agricultural fertilizers should be used in medium preparation rather than laboratory-grade products. However, nutrient ratios (N/P ratios) are an important key that enhances the biochemical composition and biomass production of microalgae [131,132,133].
Ashour and Kamel [133] reported that the cost of cultured media prepared from agricultural fertilizers (commercial grade of nitric acid and/or ammonium sulfate with the same N/P ratio and concentration of F/2 standard medium) was about 1/37:1/39 times lower than F/2 culture medium, respectively, enhancing the lipid content and biomass productivity of Nannochloropsis oceanica. Another interesting point to discuss is the source and form of carbon. Carbon is important for maintaining the culture of microalgae. Sucrose (as a carbon source) is required for cell function, growth, and development, energy storage, and stress absorption. HCO3 and CO2, inorganic carbon forms, are important for photosynthesis, CO2 capture, and bioproduct accumulation [55]. Table 4 shows the effects of nutrient limitations (macro and micronutrients) on microalgae biomass production (mg L−1), carbohydrate production (mg L−1), and metabolite improvement (%) because of the application of a nutrient limitation strategy.
The most popular carbon source is sodium bicarbonate, which increases the medium pH, improves the availability of dissolved inorganic carbon, and stimulates cell growth and the generation of energy-rich molecules. However, some algae are vulnerable to high shifts in pH, which negatively affect biomass productivity [139,140]. Table 5 shows the effect of different carbon sources on the biomass productivity of microalgae, as well as on the improvement of protein and carbohydrate percentages.
The application of microalgae in aquaculture wastewater treatment is valuable because this process is very cost-effective and enables nitrogen and phosphorus to be recycled and to produce biomass or utilize it as fertilizer. In addition, in many cases, it is not important to provide an external source of carbon to remove these elements [142]. Integrating aquaculture wastewater treatment with microalgae production in the recirculation aquaculture system (RAS) not only develops the environment but also increases economic benefits [143]. Microalgae biomass generated through this technology can be used in cosmetics, pharmaceuticals, animal feeds, and biofuels, including those derived from lipids collected throughout the process [144]. Under most conditions, the supernatant of aquaculture wastewater has a lower nutrient content than sludge from the bottom of the aquaculture system [145]. Table 6 shows the nutrient content of aquaculture wastewater produced from different cultured animals.
Environmentally, microalgae cells, especially those of Scenedesmus abundans and  Chlorella minutissima [152], are a sustainable and environmentally friendly resource that can treat several types of industrial wastewater [145]. The aquaculture effluent contains many pollutants that may harm the environment. However, algal cells can utilize these chemicals and convert them into valuable biomass. Only 20–30% of the nitrogen in fish food is absorbed or consumed by fish, while the rest is returned to the water [153]. As a result, in addition to suspended particles, wastewater contains a considerable quantity of nitrogen and phosphorus, which can harm the ecosystem [154]. According to the literature, Table 7 shows the biomass productivity of several microalgal species cultured in wastewater.

2.2.4. Microalgae–Bacteria Interaction

Recently, controlling the interactions between microalgae and bacteria has received global attention, which may improve the efficiency of biomass production from microalgae and their related bioactive compounds. However, studies of the interaction between microalgae and bacteria have shown a significant influence of parasitic relationships on algal growth [161]. The evaluation of the effect of bacteria on algae biomass cannot be fully understood if taken individually. Traditionally, in algal cultures, bacteria are known to be pollutants. Therefore, the development of the axenic biomass production of microalgae has always been a priority [1].
Interactions between microalgae and bacteria have the potential to increase algal biomass production with biomolecules relevant to commercial aquaculture interests, such as lipids and carbohydrates. In this context, all typical bacterial properties (motility, chemotaxis, type IV secretion systems, quorum sensing systems, and the production of growth promoters) may be relevant in interactions with microalgae and may affect their biomass [162].
Interactions between bacteria and microalgae are very complicated. Currently, there is little information about the chemical nature of a handful of common mediator molecules, such as nutrients, that control the relationship between microalgae and bacteria [163]. The main mediator molecules regulating the interaction between microalgae and bacteria were identified as amino acids and vitamins. Effective control of these chemical reactions has been recommended as a useful strategy to increase biomass yield and reduce microalgae production costs. Despite this, only a limited amount of information is provided at the molecular level [164].
With a better understanding of the molecular control of microalgae–bacteria interactions with sequenced organisms, particular algal–bacterial systems may be driven to achieve the desired results [1]. Table 8 shows examples of microalgae–bacteria interactions that promote algal growth and biomass production, reduce production costs, and improve the accumulation of valuable bioactive compounds, such as fatty acids, lipids, starch, and carbohydrates. These examples show bacteria’s importance in improving carbon storage in microalgae, which is especially beneficial in algal biomass production [164].

3. Pull-Down Process

The Pull-Down process includes (1) harvesting and (2) drying algal biomass (wet weight). In the case of the manufacture of algal products, the (3) extraction of bioactive compounds is included [175].

3.1. Harvesting Technologies

Harvesting (dewatering) is the separation of algae biomass from the culture column using several harvesting techniques. The harvesting of microalgae is a significant challenge facing microalgae cultivation, which is attributed to the small cell size (5–20 μm), low biomass concentrations (0.2–1.0 g L−1 in OP and 2–9 g L−1 in PBR), their similar density to water (1.08–1.13 g mL−1), and a negative charge on their cells. The cost of the harvesting process contributes to 20–30% of the total cost of biomass production.
Harvested biomass contains a lot of moisture, which reduces the quality of biomass within a few hours at room temperature. However, the selected harvesting technique impacts the quality of microalgae biomass [7]. As reported by Mathimani and Mallick [176], microalgae harvesting techniques are (1) physical techniques, which include centrifugation, filtration, gravity sedimentation, and flotation; (2) chemical techniques (flocculation methods), including inorganic and organic flocculants; (3) biological techniques, including autoflocculation and bioflocculation [177]. Harvesting techniques may be used alone or in combination with other techniques. However, there is no all-purpose harvesting approach that can handle all types of microalgae suspensions in terms of cost, energy consumption, and target products.
Among all the harvesting methods, centrifugation is considered the most widely and fastest-used harvesting technology [55]. A study by Levine and Fleurence [178] concluded that centrifugation technology works based on the density difference, which offers more advantages over other technologies, such as chemical-free biomass, biological-free biomass, no toxicity, 100% complete recovery efficiency, short duration, and high biomass quality. These advantages make the centrifugation technique the main technology used in the food products, cosmetics, and pharmaceutical industries. Although centrifugation has been reported as a very effective technique, it is highly energy consuming and expensive when applied on a commercial scale [179]. There are several types of centrifugation equipment, including the nozzle discharge centrifuge, continuous flow centrifuge, spiral plate centrifuge, self-cleaning disc stack centrifuge, and decanter bowl centrifuge [179].
Recently, harvesting edible fungi, either fungal spores or fungal pellets, has become the most recent sustainable, cost-effective, and highly efficient harvesting technique [177]. The co-cultivation of microalgae with fungal spores has been studied by Zhou et al. [180], who reported that using fungi pellets could harvest more than 95% of microalgae biomass in a short time (about 1.5 h), resulting in a great performance of the microalgae biomass [181]. In aquaculture, the selected fungi strain should have a high safety level and should not contain toxic components. Some fungi species, such as Aspergillus sp., Penicillium sp., and Monascus sp., which are classified as food-grade, have a safety level as a value-added compound, while the beneficial impacts of these fungi on aquatic animals have not yet received much attention from aquaculture scientists [177,181]. Interestingly, the Revolving Algal Biofilm technique (RAB technology), which was designed to grow microalgae on a film, was supposed to be a promising harvesting technology to simplify harvesting and wastewater treatment and increase land-use efficiency [182]. In this technology, when compared to traditional harvesting technologies, RAB technology is more efficient, cost-effective, and environmentally friendly for harvesting the biomass attached to films via a scraper [177,182,183,184].

3.2. Drying Technologies

After harvesting microalgae biomass, the harvested biomass contains a high water content that must be drained before further treatments of the biomass, depending on the type of process and the products that will be extracted in the next step. Neves et al. [185] stated that the value of microalgal biomass may be lost depending on the drying technique applied. The drying process changes the characteristics of microalgae cells. In the food industry, the nutritional quality of dried microalgae may decrease due to the loss of pigments, proteins, carbohydrates, lipids, and other valuable components. In general, pigments are most affected by drying techniques and drying conditions, such as the temperature level, time exposed, and oxygen [185]. There are several drying methods, including solar, cross-flow air-drying, oven and microwave drying, and spray- and freeze-drying [7]. Solar drying is the most environmentally friendly drying technique, but it is time-consuming and requires a massive surface area [186]. Spray-drying is assumed to be an effective technique for high-value-added products. This technique can produce green or green dark microalgae powder due to the use of spray-drying at the correct temperature [187].
Interestingly, spray-dried microalgae biomass can also be used for pharmaceutical products and human food supplements. There are two different spray-drying techniques: spray-drying and drum-drying. The second method is recommended for pharmaceutical and human food consumption products due to its better digestibility, lower investment, and less energy [188]. Freeze-drying is a widely applied drying technique used in the pharmaceutical and food industries since the cell constituents are well protected without cell wall disruption using this technique [187].
Among all drying techniques concerning maximizing the productivity of microalgae biomass, spray-drying is the most promising technology used to extract high-value products. However, the spray-drying technique is costly and may damage the biomass pigment content [189]. In the future, more attention should be given to improving the spray drying-technique so that it can be applied to all microalgae products. As described by Ruiz-Dominguez et al. [190], Muriellopsis sp. contains low amounts of astaxanthin, zeaxanthin, and violaxanthin, while having a huge amount of lutein. However, these pigments have received global attention and extensively contribute to pharmacological, antiviral, food supplement, nutraceutical, and several biotechnological applications. Ruiz-Domínguez et al. [190] compared the effect of spray-drying and freeze-drying techniques on the biomass yield of Muriellopsis sp. and found that the freeze-drying technique increased the lutein content and recovery by 0.3–2.5-fold when compared to the spray-drying technique. They concluded that the freeze-drying technique is a promising technique for lutein production, particularly for pharmacological, antiviral, food supplement, and nutraceutical applications [190]. Similar findings were reported by Stramarkou et al. [191] on Chlorella vulgaris. The advantages and disadvantages of the most common drying techniques are shown in Table 9.

3.3. Extraction Technologies

In recent years, high-value bioproducts derived from microalgae biomass, especially in the fields of pharmaceuticals and human food supplements, have received global interest. Therefore, choosing the correct extraction procedure is important to consider. The extraction of bioactive compounds from microalgae biomass is no less important than the harvesting or drying techniques. The extraction technique is one of the most important keys to determining the quality of the target product, especially in the fields of nutrition and medicine. While extracting the biomolecules from microalgae biomass, it is necessary to use a completely biocompatible technology that does not change or affect the bioactivity of the extracted molecule [192].
Based on the microalgae’s target product and the nature of the bioactive compounds, the extraction method is determined. Several extraction procedures may be used to disrupt the cell wall [193]. Currently, traditional extraction procedures include organic solvents, such as methanol, acetone, chloroform, and diethyl ether, which have been used many times [194]. Onay et al. [195] reported that conventional organic solvent extraction procedures include Folch [196], Bligh and Dyer [197], and Soxhlet [198], while some assisted lipid extraction techniques include cell disruption, homogenization, microwave, ultrasonication, glass bead, and lyophilization-assisted methods [195].
Several published reports have demonstrated the efficiency of one-step or assisted extraction techniques for lipid extraction from microalgae. Koberg et al. [199] used a one-step biodiesel extraction process using a microwave and ultrasonication directly from Nannochloropsis. The direct transesterification of Cryptococcus curvatus was determined using the microwave irradiation technique of Cui et al. [200]. Cheng et al. [201] investigated the direct transesterification of Chlorella pyrenoidosa using a microwave-assisted lipid extraction method, compared to the two-step process of biodiesel extraction. They reported that for Chlorella pyrenoidosa, the one-step technique demonstrated a six-fold greater biomass than the two-step procedure [201]. Table 10 shows the most important microalgae extraction techniques, cell extraction, and cell rupture mechanisms [202].
New generations of green extraction technologies have been developed. These technologies do not require the involvement of toxic solvents. Global interest has been conducted in green extraction technologies that enhance the yield of extracted molecules while minimizing environmental impacts [218]. These environmentally friendly technologies require bioactive products to be obtained, thus reducing environmental impact, in agreement with various green chemistry principles. In addition, a minimization of extraction time and yield (productivity) has been obtained [40]. In recent years, the demand for safer, greener, and more natural bioproducts that do not require toxic solvents has increased. For Botryococcus sp., Chlorella vulgaris, and Scenedesmus sp., Lee et al. [219] reported that microwaves were the most effective technique used in oil extraction.
According to Lee et al. [202], green extraction techniques are divided into mechanical (such as high-pressure homogenization, rotor-stator homogenization, bead milling, grinding with mortar and pestle, ultrasonication, hydrodynamic cavitation, and screw expeller pressing), physical (such as repeated freeze–thaw, osmotic shock, explosive decompression, pulsed electric field, microwave, and autoclave), and chemical (such as the use of acids or alkalis, surfactant, and enzyme) techniques.
Alam et al. [7] demonstrated that, currently, many bioproducts are extracted from marine microalgae at a commercial scale, including for human nutrition, phycobilins, animal and aquatic feed, PUFA, sterols, β-carotene, vitamins, stable isotope biochemicals, polysaccharides, bioactive molecules of antimicrobials, antibacterial, antiviral, anti-inflammatory, and anticancer drugs. Most recently, microalgae extracts have been strongly recommended as an anti-pro-inflammatory agent, as well as a blocker, and inhibitor of cytokine storm and Tumor Necrosis Factor-α (TNF-α), suggesting that microalgae extracts are an attractive source against many diseases, including COVID-19 [220]. The most notable microalgae species rich in astaxanthin and carotenoids are Haematococcus pluvialis [221] and Dunaliella salina [222,223]. Species rich in omega-3, PUFA, and fatty acids include Phaeodactylum tricornutum [224], Porphyridium cruentum [225], Crypthecodinium cohnii [226], and Nannochloropsis spp. [227].
Microalgal extracts, such as polysaccharides, are extensively used in food items as thickeners and stabilizers [228]. As reported by Rahman [229], in 2010, over 75% of the global market and manufacturing of microalgae were utilized in human health food (mainly EPA and DHA). In the USA, microalgae-derived DHA is found and marketed in more than 99% of all baby foods [230]. Taufiqurrahmi et al. [231] demonstrated that phycoerythrin and phycocyanin, red and blue pigment–protein complexes, respectively, extracted from Arthrospira platensis, were extensively used in the food industry as natural coloring agents [231]. Zanella and Alam [232] reported that some microalgae strains were significant suppliers of specific compounds that accumulate in huge quantities, particularly when grown under ideal environmental conditions.

4. Microalgae Derivatives against the COVID-19 Cytokine Storm

Several studies have suggested that a key factor in COVID-19 patients’ deaths is Acute Respiratory Distress Syndrome (ARDS). Cytokine storm syndrome causes ARDS. ARDS is the main cause of death in COVID patients [233]. Intensive care units (ICUs) are necessary for ARDS patients if mechanical ventilation is needed [234]. Pro-inflammatory cytokines (IL-1, IL-6, and NFT-α), as well as chemokines (CCL2, CCL3, CXCL10, and CXCL9), are produced during a cytokine storm in great quantities, leading to an overactive immune system and Acute Lung Damage (ALI) in a short time [235,236]. The drugs considered for the treatment of COVID-19 may operate through one of two mechanisms: (1) slowing the rate of viral reproduction or (2) symptom suppression by anti-inflammatory therapy [237].
Patients with COVID-19 often suffer immune damage caused by a reactive cytokine storm as a result of hyperactive inflammatory responses that culminate in cytokine release syndrome [238,239]. Tzachor et al. [59] reported that as a part of the cytokine storm (CS), the overflow of TNF-α results in destabilized endothelial cell networks, which cause damage to the vascular barrier, capillary damage, diffuse alveolar damage (DAD), apoptotic cell death, and multi-organ failure [59]. Since the pandemic started, TNF-blockers have demonstrated promising results in the treatment and mitigation of severe sickness [240]. The increased production of TNF-α has a significant role in disrupting the lung endothelial and epithelial barriers, which may cause ARDS [241].
Interestingly, several studies reported that introducing microalgae derivatives to COVID-19 patients may reduce the cytokine storm and have anti-TNF-α effects, preventing ARDS and ALI [56,59,242,243,244,245], as shown in Figure 3.
Microalgae are of commercial importance due to their structural, functional, and nutritional importance, as well as their high yield of proteins, lipids, carbohydrates, and other bioactive compounds [97]. According to the literature, microalgae derivatives have many beneficial immunomodulatory effects on humans [246]. Microalgae are one of the richest sources of natural bioactive compounds that have antioxidant, anti-inflammatory, antimicrobial, and antiviral activities [247]. Therefore, microalgae can scavenge free radicals; thus, the damage caused by free radicals is reduced either in vitro or in vivo [248,249]. These natural free radical scavengers include carbohydrates (polysaccharide, monosaccharide, oligosaccharide, carrageenan, alginate, etc.), proteins (amino acids, peptides, DNA, and RNA), lipids (fatty acids, saturated fatty acids, unsaturated fatty acids, phospholipids, polyunsaturated fatty acids, omega-3, AA, EPA, DHA, etc.), pigments (phycobiliprotein, phycocyanin, phycoerythrin, β-carotenes, lutein, zeaxanthin, astaxanthin, chlorophyll, violaxanthin, etc.), vitamins, and minerals [250].
However, many types of free radical scavengers are found in different origins of microalgae and are well known as antioxidant compounds, such as phenolics, polyphenolics, flavonoids, terpenoids, sterols, anthraquinones, stilbenes, phlorotannin, and hydroxycinnamic compounds [193]. For the biomedical applications of microalgae carotenoids, the biological activities of β-carotene, astaxanthin, and phycocyanin are widely recognized as being candidates for use as natural antioxidants, as well as antiviral and anti-inflammatory substrates [251]. The natural superfood supplement Arthrospira is well known as a powerful natural source of pigments, such as phycocyanin, phycobiliproteins, β-carotenes, lutein, and astaxanthin [16,252,253,254], amino acids and polysaccharides [255] that have a long history of several biomedical applications [256]. The success of any of the microalgae biomedical products depends on which technologies are applied in the cultivation, harvesting, drying, and extraction of the active biomolecules of these health products. In the context of combating COVID-19 using the extract of the blue–green algae Arthrospira platensis, Tzachor et al. [59] investigated the effect of several doses of A. platensis extracts, as a novel approach, as TNF-blockers and TNF-α inhibitors. A. platensis (strain UTEX 3086) was grown in flat panel airlift photobioreactors (PBR, 180 L) using Zarrouk medium [257], under a control temperature (31 ± 2 °C), pH (10.8 ± 0.2 °C), and filtered air conditions (with a flow of 0.5 vvm). In this study, the only different cultural condition was cultural illumination, which was performed in two irradiations: (1) Solar Spirulina: full-range solar irradiation (750 μmol m2 s−1) and (2) LED Spirulina: red, blue, and UV at photosynthetic irradiation of 750 μmol m2 s−1. The two different cultures (Solar and LED) were physically extracted using water and freeze–thawing cellular disruption, as previously described by Chu et al. [220]. In this study, macrophages and monocytes, which were stimulated by pathogenic stimulator lipopolysaccharides (LPSs), were exposed to several doses of A. platensis extracts from both Solar and LED. Photosynthetically active Arthrospira (LED Spirulina) inhibited the secretion of TNF-α by 70% and 40% from LPS-activated macrophage and monocyte cells, respectively. The findings of this study suggest that the advanced technologies applied in the production of the bioactive compounds of Arthrospira platensis, in both their culture (including the PBR system with controlled conditions of temperature, pH, and filtered air using Zarrouk medium and LED photosynthetic illumination) and extraction (physically extracted using water and freeze–thawing cellular disruption), are strongly recommended, as these molecules may serve as blockers and inhibitors of TNF-α and act as anti-NFT-α, suggesting that the extracts of blue–green microalga A. platensis may be attractive for combating COVID-19 [220].

5. Conclusions and Future Perspectives

Microalgae have a wide range of commercial applications, such as aquaculture, biofertilizer, biofuel, cosmetics, functional foods, and pharmaceuticals, receiving more global attention both industrially and academically. The success of the biomedical products of microalgae biomass, as well as their derivatives or metabolites, mainly depends on the technologies used in the cultivation, harvesting, and drying processes, as well as the extraction of bioactive molecules. The extract of the blue–green microalgae A. platensis may be an attractive source for combating COVID-19, depending on the techniques used for culturing, harvesting, drying, and extracting its bioactive substrates. The production scheme for microalgae biomass mainly consists of two processes: (I) the Build-Up process and (II) the Pull-Down process. The Build-Up process consists of (1) the super strain concept and (2) cultivation aspects. The Pull-Down process includes (1) harvesting and (2) drying algal biomass (wet weight). In some cases, such as the manufacture of algal bioproducts, the (3) extraction of bioactive compounds is included. To achieve the “super strain concept”, the selection of the microalgal strain must (1) have all primary requirements of the quality and quantity of bioactive molecules, especially concerning its lipid content and profile, (2) be able to be genetically engineered, and (3) be able to produce bioenergy and valuable co-products that enhance economic profitability. Economically, scientists are searching for microalgae strains that have been genetically engineered to increase growth rate, biomass yield, lipid content, and high-value co-products. Genetic improvement techniques have been widely applied to revolutionize the microalgae cultivation industry. Considering these observations, there is a need for innovations, solutions, and technologies that enhance and advance the commercial viability of algal biomass production and its technologies. Closed photobioreactor systems (PBR) are better than open pond systems (OP) in terms of their operation, maintenance, and biomass productivity, and they produce a greater control over biomass production than open pond systems. Temperature and light intensity must be adjusted to promote the accumulation of targeted bioactive compounds. Microalgae culture medium should be cost-effective, promote rapid development, meet the requirements of microalgae cells, and be easy to produce. LED light is a cost-effective technology for producing microalgae. Microalgal–bacteria interactions promote algal growth and biomass production, reduce production costs, and improve the accumulation of valuable bioactive compounds. The centrifugation harvesting technique is considered the most widely harvested technology. Recently, Revolving Algal Biofilm (RAB technology) has been proposed as a promising technology to simplify harvesting and wastewater treatment and increase land-use efficiency. Among all drying techniques concerning maximizing the productivity of microalgae biomass, spray-drying is the most promising technology used to extract high-value products. In the future, more attention should be given to the improvement of spray-drying and the development of the devices and methods applied to all types of microalgae products. Globally, the most frequently extracted molecule from microalgae is oil. Green extracting technologies are being developed that do not require the involvement of toxic solvents.

Author Contributions

Conceptualization, M.A. and A.M.M.O.; methodology, M.A. and A.M.M.O.; validation, M.A. and A.M.M.O.; formal analysis, M.A. and A.M.M.O.; investigation, M.A. and A.M.M.O.; resources, M.A. and A.M.M.O.; writing—original draft preparation, M.A.; writing—review and editing, M.A.; visualization, M.A. and A.M.M.O.; supervision, M.A. and A.M.M.O.; project administration, M.A. and A.M.M.O.; funding acquisition, M.A. and A.M.M.O. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the authors themselves.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Abbas, E.M.; Ali, F.S.; Desouky, M.G.; Ashour, M.; El-Shafei, A.; Maaty, M.M.; Sharawy, Z.Z. Novel Comprehensive Molecular and Ecological Study Introducing Coastal Mud Shrimp (Solenocera Crassicornis) Recorded at the Gulf of Suez, Egypt. J. Mar. Sci. Eng. 2020, 9, 9. [Google Scholar] [CrossRef]
  2. Ashour, M.; Mabrouk, M.M.; Abo-Taleb, H.A.; Sharawy, Z.Z.; Ayoub, H.F.; Van Doan, H.; Davies, S.J.; El-Haroun, E.; Goda, A.M.S.A. A liquid seaweed extract (TAM®) improves aqueous rearing environment, diversity of zooplankton community, whilst enhancing growth and immune response of Nile tilapia, Oreochromis niloticus, challenged by Aeromonas hydrophila. Aquaculture 2021, 543, 736915. [Google Scholar] [CrossRef]
  3. El-Shenody, R.A.; Ashour, M.; Ghobara, M.M.E. Evaluating the chemical composition and antioxidant activity of three Egyptian seaweeds: Dictyota dichotoma, Turbinaria decurrens, and Laurencia obtusa. Braz. J. Food Technol. 2019, 22. [Google Scholar] [CrossRef]
  4. Elshobary, M.E.; El-Shenody, R.A.; Ashour, M.; Zabed, H.M.; Qi, X. Antimicrobial and antioxidant characterization of bioactive components from Chlorococcum minutum. Food Biosci. 2020, 35, 100567. [Google Scholar] [CrossRef]
  5. Elshobary, M.E.; Osman, M.E.; Abo-Shady, A.M.; Komatsu, E.; Perreault, H.; Sorensen, J.; Piercey-Normore, M.D. Algal carbohydrates affect polyketide synthesis of the lichen-forming fungus Cladonia rangiferina. Mycologia 2016, 108, 646–656. [Google Scholar] [CrossRef]
  6. Valdovinos-García, E.M.; Petriz-Prieto, M.A.; Olán-Acosta, M.d.l.Á.; Barajas-Fernández, J.; Guzmán-López, A.; Bravo-Sánchez, M.G. Production of microalgal biomass in photobioreactors as feedstock for bioenergy and other uses: A techno-economic study of harvesting stage. Appl. Sci. 2021, 11, 4386. [Google Scholar] [CrossRef]
  7. Alam, M.A.; Xu, J.-L.; Wang, Z. Microalgae Biotechnology for Food, Health and High Value Products; Springer: Berlin/Heidelberg, Germany, 2020. [Google Scholar]
  8. Santhakumaran, P.; Ayyappan, S.M.; Ray, J.G. Nutraceutical applications of twenty-five species of rapid-growing green-microalgae as indicated by their antibacterial, antioxidant and mineral content. Algal Res. 2020, 47, 101878. [Google Scholar] [CrossRef]
  9. El-Shouny, W.; Sharaf, M.; Abomohra, A.; Abo-Eleneen, M. Production enhancement of some valuable compounds of Arthrospira Platensis. J. Basic Environ. Sci. 2015, 2, 74–83. [Google Scholar]
  10. Barra, L.; Chandrasekaran, R.; Corato, F.; Brunet, C. The challenge of ecophysiological biodiversity for biotechnological applications of marine microalgae. Mar. Drugs 2014, 12, 1641–1675. [Google Scholar] [CrossRef]
  11. Kamani, M.H.; Eş, I.; Lorenzo, J.M.; Remize, F.; Roselló-Soto, E.; Barba, F.J.; Clark, J.; Khaneghah, A.M. Advances in plant materials, food by-products, and algae conversion into biofuels: Use of environmentally friendly technologies. Green Chem. 2019, 21, 3213–3231. [Google Scholar] [CrossRef]
  12. Abualnaja, K.M.; Alprol, A.E.; Abu-Saied, M.A.; Mansour, A.T.; Ashour, M. Studying the Adsorptive Behavior of Poly(Acrylonitrile-co-Styrene) and Carbon Nanotubes (Nanocomposites) Impregnated with Adsorbent Materials towards Methyl Orange Dye. Nanomaterials 2021, 11, 1144. [Google Scholar] [CrossRef] [PubMed]
  13. El-Sayed, H.S.; Ibrahim, H.A.; Beltagy, E.A.; Khairy, H.M. Effects of short term feeding of some marine microalgae on the microbial profile associated with Dicentrarchus labrax post larvae. Egypt. J. Aquat. Res. 2014, 40, 251–260. [Google Scholar] [CrossRef]
  14. Sharawy, Z.Z.; Ashour, M.; Abbas, E.; Ashry, O.; Helal, M.; Nazmi, H.; Kelany, M.; Kamel, A.; Hassaan, M.; Rossi, W.; et al. Effects of dietary marine microalgae, Tetraselmis suecica, on production, gene expression, protein markers and bacterial count of Pacific white shrimp Litopenaeus vannamei. Aquac. Res. 2020, 51, 2216–2228. [Google Scholar] [CrossRef]
  15. Sharawy, Z.Z.; Ashour, M.; Labena, A.; Alsaqufi, A.; Mansour, A.; Abbas, E. Effects of dietary Arthrospira platensis nanoparticles on growth performance, feed utilization, and growth-related gene expression of Pacific white shrimp, Litopenaeus vannamei. Aquaculture 2022, 551, 737905. [Google Scholar] [CrossRef]
  16. Mansour, A.T.; Ashour, M.; Abbas, E.M.; Alsaqufi, A.S.; Kelany, M.S.; El-Sawy, M.A.; Sharawy, Z.Z. Growth Performance, Immune-Related and Antioxidant Genes Expression, and Gut Bacterial Abundance of Pacific White Leg Shrimp, Litopenaeus vannamei, Dietary Supplemented With Natural Astaxanthin. Front. Physiol. 2022, 13. [Google Scholar] [CrossRef]
  17. Abdelrhman, A.M.; Ashour, M.; Al-Zahaby, M.A.; Sharawy, Z.Z.; Nazmi, H.; Zaki, M.A.; Ahmed, N.H.; Ahmed, S.R.; El-Haroun, E.; Van Doan, H. Effect of polysaccharides derived from brown macroalgae Sargassum dentifolium on growth performance, serum biochemical, digestive histology and enzyme activity of hybrid red tilapia. Aquac. Rep. 2022, 25, 101212. [Google Scholar] [CrossRef]
  18. Abomohra, A.E.-F.; Almutairi, A.W. A close-loop integrated approach for microalgae cultivation and efficient utilization of agar-free seaweed residues for enhanced biofuel recovery. Bioresour. Technol. 2020, 317, 124027. [Google Scholar] [CrossRef]
  19. Ashour, M.; Mabrouk, M.M.; Ayoub, H.F.; El-Feky, M.; Zaki, S.Z.; Hoseinifar, S.H.; Rossi, W.; Van Doan, H.; El-Haroun, E.; Goda, A.M.A.S. Effect of dietary seaweed extract supplementation on growth, feed utilization, hematological indices, and non-specific immunity of Nile Tilapia, Oreochromis niloticus challenged with Aeromonas hydrophila. J. Appl. Phycol. 2020, 32, 3467–3479. [Google Scholar] [CrossRef]
  20. Ashour, M.; El-Shafei, A.A.; Khairy, H.M.; Abd-Elkader, D.Y.; Mattar, M.A.; Alataway, A.; Hassan, S.M. Effect of Pterocladia capillacea Seaweed Extracts on Growth Parameters and Biochemical Constituents of Jew’s Mallow. Agronomy 2020, 10, 420. [Google Scholar] [CrossRef]
  21. Ashour, M.; Hassan, S.M.; Elshobary, M.E.; Ammar, G.A.G.; Gaber, A.; Alsanie, W.F.; Mansour, A.T.; El-Shenody, R. Impact of Commercial Seaweed Liquid Extract (TAM®) Biostimulant and Its Bioactive Molecules on Growth and Antioxidant Activities of Hot Pepper (Capsicum annuum). Plants 2021, 10. [Google Scholar] [CrossRef]
  22. Hassan, S.M.; Ashour, M.; Sakai, N.; Zhang, L.; Hassanien, H.A.; Gaber, A.; Ammar, G. Impact of Seaweed Liquid Extract Biostimulant on Growth, Yield, and Chemical Composition of Cucumber (Cucumis sativus). Agriculture 2021, 11, 320. [Google Scholar] [CrossRef]
  23. Hassan, S.M.; Ashour, M.; Soliman, A. Anticancer Activity, Antioxidant Activity, Mineral Contents, Vegetative and Yield of Eruca sativa Using Foliar Application of Autoclaved Cellular Extract of Spirulina platensis Extract, Comparing to NPK Fertilizers. J. Plant Prod. 2017, 8, 529–536. [Google Scholar] [CrossRef]
  24. Hassan, S.M.; Ashour, M.; Soliman, A.A.F.; Hassanien, H.A.; Alsanie, W.F.; Gaber, A.; Elshobary, M.E. The Potential of a New Commercial Seaweed Extract in Stimulating Morpho-Agronomic and Bioactive Properties of Eruca vesicaria (L.) Cav. Sustainability 2021, 13, 4485. [Google Scholar] [CrossRef]
  25. Abomohra, A.E.-F.; Elsayed, M.; Esakkimuthu, S.; El-Sheekh, M.; Hanelt, D. Potential of fat, oil and grease (FOG) for biodiesel production: A critical review on the recent progress and future perspectives. Prog. Energy Combust. Sci. 2020, 81, 100868. [Google Scholar] [CrossRef]
  26. Abomohra, A.E.-F.; Zheng, X.; Wang, Q.; Huang, J.; Ebaid, R. Enhancement of biodiesel yield and characteristics through in-situ solvo-thermal co-transesterification of wet microalgae with spent coffee grounds. Bioresour. Technol. 2020, 323, 124640. [Google Scholar] [CrossRef]
  27. Abomohra, A.E.F.; El-Sheekh, M.; Hanelt, D. Pilot cultivation of the chlorophyte microalga Scenedesmus obliquus as a promising feedstock for biofuel. Biomass Bioenergy 2014, 64, 237–244. [Google Scholar] [CrossRef]
  28. Abomohra, A.E.-F.; Elshobary, M. Biodiesel, bioethanol, and biobutanol production from microalgae. In Microalgae Biotechnology for Development of Biofuel and Wastewater Treatment; Springer: Singapore, 2019; pp. 293–321. [Google Scholar]
  29. Grigorenko, A.; Kostyukevich, Y.I.; Chernova, N.; Kiseleva, S.; Kiseleva, E.; Popel, O.; Vladimirov, G.; Nikolaev, E.; Kumar, V.; Vlaskin, M. Hydrothermal liquefaction of Arthrospira platensis for bio-oil production and study of chemical composition for bio-oil and its gasoline fraction. Russ. J. Appl. Chem. 2019, 92, 1480–1486. [Google Scholar] [CrossRef]
  30. Vlaskin, M.S.; Grigorenko, A.V.; Chernova, N.I.; Kiseleva, S.V.; Lipatova, I.A.; Popel, O.S.; DombrovsKy, L.A. The hydrothermal liquefaction as a promising procedure for microalgae-to-biofuel conversion: A general review and some thermophysical problems to be solved. High Temp. High Press. 2019, 48, 309–351. [Google Scholar] [CrossRef]
  31. Alprol, A.E.; Ashour, M.; Mansour, A.T.; Alzahrani, O.M.; Mahmoud, S.F.; Gharib, S.M. Assessment of Water Quality and Phytoplankton Structure of Eight Alexandria Beaches, Southeastern Mediterranean Sea, Egypt. J. Mar. Sci. Eng. 2021, 9, 1328. [Google Scholar] [CrossRef]
  32. Alprol, A.E.; El-Metwally, M.; Amer, A. Sargassum latifolium as eco-friendly materials for treatment of toxic nickel (II) and lead (II) ions from aqueous solution. Egypt. J. Aquat. Biol. Fish. 2019, 23, 285–299. [Google Scholar] [CrossRef]
  33. Alprol, A.E.; Heneash, A.M.M.; Soliman, A.M.; Ashour, M.; Alsanie, W.F.; Gaber, A.; Mansour, A.T. Assessment of Water Quality, Eutrophication, and Zooplankton Community in Lake Burullus, Egypt. Diversity 2021, 13, 268. [Google Scholar] [CrossRef]
  34. Al-Rub, F.A.; El-Naas, M.; Ashour, I.; Al-Marzouqi, M. Biosorption of copper on Chlorella vulgaris from single, binary and ternary metal aqueous solutions. Process Biochem. 2006, 41, 457–464. [Google Scholar] [CrossRef]
  35. Mansour, A.T.; Alprol, A.E.; Abualnaja, K.M.; El-Beltagi, H.S.; Ramadan, K.M.A.; Ashour, M. Dried Brown Seaweed’s Phytoremediation Potential for Methylene Blue Dye Removal from Aquatic Environments. Polymers 2022, 14, 1375. [Google Scholar] [CrossRef] [PubMed]
  36. Mansour, A.T.; Alprol, A.E.; Abualnaja, K.M.; El-Beltagi, H.S.; Ramadan, K.M.A.; Ashour, M. The Using of Nanoparticles of Microalgae in Remediation of Toxic Dye from Industrial Wastewater: Kinetic and Isotherm Studies. Materials 2022, 15, 3922. [Google Scholar] [CrossRef] [PubMed]
  37. Mansour, A.T.; Alprol, A.E.; Ashour, M.; Ramadan, K.M.; Alhajji, A.H.; Abualnaja, K.M. Do Red Seaweed Nanoparticles Enhance Bioremediation Capacity of Toxic Dyes from Aqueous Solution? Gels 2022, 8, 310. [Google Scholar] [CrossRef] [PubMed]
  38. Mansour, A.T.; Alprol, A.E.; Khedawy, M.; Abualnaja, K.M.; Shalaby, T.A.; Rayan, G.; Ramadan, K.M.; Ashour, M. Green Synthesis of Zinc Oxide Nanoparticles Using Red Seaweed for the Elimination of Organic Toxic Dye from an Aqueous Solution. Materials 2022, 15, 5169. [Google Scholar] [CrossRef]
  39. Vega, J.; Álvarez-Gómez, F.; Güenaga, L.; Figueroa, F.L.; Gómez-Pinchetti, J.L. Antioxidant activity of extracts from marine macroalgae, wild-collected and cultivated, in an integrated multi-trophic aquaculture system. Aquaculture 2020, 522, 735088. [Google Scholar] [CrossRef]
  40. Ardiles, P.; Cerezal-Mezquita, P.; Salinas-Fuentes, F.; Órdenes, D.; Renato, G.; Ruiz-Domínguez, M.C. Biochemical Composition and Phycoerythrin Extraction from Red Microalgae: A Comparative Study Using Green Extraction Technologies. Processes 2020, 8, 1628. [Google Scholar] [CrossRef]
  41. Cai, J.; Lovatelli, A.; Aguilar-Manjarrez, J.; Cornish, L.; Dabbadie, L.; Desrochers, A.; Diffey, S.; Garrido Gamarro, E.; Geehan, J.; Hurtado, A. Seaweeds and Microalgae: AnOverview for Unlocking Their Potential in Global Aquaculture Development; FAO: Rome, Italy, 2021. [Google Scholar]
  42. World Fisheries and Aquaculture. 2020. Available online: https://www.aqua.cl/wp-content/uploads/2020/06/FAO-SOFIA-2020-Brief.pdf (accessed on 29 July 2022).
  43. Larson, S.; Stoeckl, N.; Fachry, M.E.; Mustafa, M.D.; Lapong, I.; Purnomo, A.H.; Rimmer, M.A.; Paul, N.A. Women’s well-being and household benefits from seaweed farming in Indonesia. Aquaculture 2021, 530, 735711. [Google Scholar] [CrossRef]
  44. Vigani, M.; Parisi, C.; Rodríguez-Cerezo, E.; Barbosa, M.J.; Sijtsma, L.; Ploeg, M.; Enzing, C. Food and feed products from micro-algae: Market opportunities and challenges for the EU. Trends Food Sci. Technol. 2015, 42, 81–92. [Google Scholar] [CrossRef]
  45. Callegari, A.; Bolognesi, S.; Cecconet, D.; Capodaglio, A.G. Production technologies, current role, and future prospects of biofuels feedstocks: A state-of-the-art review. Crit. Rev. Environ. Sci. Technol. 2020, 50, 384–436. [Google Scholar] [CrossRef]
  46. Walsh, M.J.; Van Doren, L.G.; Sills, D.L.; Archibald, I.; Beal, C.M.; Lei, X.G.; Huntley, M.E.; Johnson, Z.; Greene, C.H. Algal food and fuel coproduction can mitigate greenhouse gas emissions while improving land and water-use efficiency. Environ. Res. Lett. 2016, 11, 114006. [Google Scholar] [CrossRef]
  47. Zaky, A.S. Introducing a Marine Biorefinery System for the Integrated Production of Biofuels, High-Value-Chemicals, and Co-Products: A Path Forward to a Sustainable Future. Processes 2021, 9, 1841. [Google Scholar] [CrossRef]
  48. Zaky, A.S.; Kumar, S.; Welfle, A.J. Integrated Approaches and Future Perspectives. In Waste-to-Energy; Springer: Berlin/Heidelberg, Germany, 2022; pp. 613–651. [Google Scholar]
  49. Zhu, L. Biorefinery as a promising approach to promote microalgae industry: An innovative framework. Renew. Sustain. Energy Rev. 2015, 41, 1376–1384. [Google Scholar] [CrossRef]
  50. Béchet, Q.; Plouviez, M.; Chambonnière, P.; Guieysse, B. Environmental impacts of full-scale algae cultivation. In Microalgae-Based Biofuels and Bioproducts; Elsevier: Amsterdam, The Netherlands, 2017; pp. 505–525. [Google Scholar]
  51. Mayers, J.J.; Ekman Nilsson, A.; Svensson, E.; Albers, E. Integrating microalgal production with industrial outputs—Reducing process inputs and quantifying the benefits. Ind. Biotechnol. 2016, 12, 219–234. [Google Scholar] [CrossRef]
  52. Usher, P.K.; Ross, A.B.; Camargo-Valero, M.A.; Tomlin, A.S.; Gale, W.F. An overview of the potential environmental impacts of large-scale microalgae cultivation. Biofuels 2014, 5, 331–349. [Google Scholar] [CrossRef]
  53. Mansour, A.T.; Ashour, M.; Alprol, A.E.; Alsaqufi, A.S. Aquatic Plants and Aquatic Animals in the Context of Sustainability: Cultivation Techniques, Integration, and Blue Revolution. Sustainability 2022, 14, 3257. [Google Scholar] [CrossRef]
  54. Araújo, R.; Vázquez Calderón, F.; Sánchez López, J.; Azevedo, I.C.; Bruhn, A.; Fluch, S.; Garcia Tasende, M.; Ghaderiardakani, F.; Ilmjärv, T.; Laurans, M. Current status of the algae production industry in Europe: An emerging sector of the Blue Bioeconomy. Front. Mar. Sci. 2021, 7, 1247. [Google Scholar] [CrossRef]
  55. Muhammad, G.; Alam, M.; Xiong, W.; Lv, Y.; Xu, J.-L. Microalgae biomass production: An overview of dynamic operational methods. In Microalgae Biotechnology for Food, Health and High Value Products; Springer: Singapore, 2020; pp. 415–432. [Google Scholar]
  56. Talukdar, J.; Dasgupta, S.; Nagle, V.; Bhadra, B. COVID-19: Potential of microalgae derived natural astaxanthin as adjunctive supplement in alleviating cytokine storm. SSRN. 2020. Available online: https://papers.ssrn.com/sol3/papers.cfm?abstract_id=3579738 (accessed on 29 July 2022).
  57. Alam, M.A.; Parra-Saldivar, R.; Bilal, M.; Afroze, C.A.; Ahmed, M.N.; Iqbal, H.M.; Xu, J. Algae-derived bioactive molecules for the potential treatment of SARS-CoV-2. Molecules 2021, 26, 2134. [Google Scholar] [CrossRef]
  58. Ziyaei, K.; Ataie, Z.; Mokhtari, M.; Adrah, K.; Daneshmehr, M.A. An insight to the therapeutic potential of algae-derived sulfated polysaccharides and polyunsaturated fatty acids: Focusing on the COVID-19. Int. J. Biol. Macromol. 2022, 209, 244–257. [Google Scholar] [CrossRef]
  59. Tzachor, A.; Rozen, O.; Khatib, S.; Jensen, S.; Avni, D. Photosynthetically controlled spirulina, but not solar spirulina, inhibits TNF-α secretion: Potential implications for COVID-19-related cytokine storm therapy. Mar. Biotechnol. 2021, 23, 149–155. [Google Scholar] [CrossRef] [PubMed]
  60. Rajvanshi, M.; Sayre, R. Recent advances in algal biomass production. In Biotechnological Applications of Biomass; Intech Open: London, UK, 2020. [Google Scholar]
  61. Guiry, M.; Guiry, G. AlgaeBase. World-Wide Electronic Publication; National University of Ireland: Galway, Ireland, 2013. [Google Scholar]
  62. Yew, G.Y.; Lee, S.Y.; Show, P.L.; Tao, Y.; Law, C.L.; Nguyen, T.T.C.; Chang, J.-S. Recent advances in algae biodiesel production: From upstream cultivation to downstream processing. Bioresour. Technol. Rep. 2019, 7, 100227. [Google Scholar] [CrossRef]
  63. Neofotis, P.; Huang, A.; Sury, K.; Chang, W.; Joseph, F.; Gabr, A.; Twary, S.; Qiu, W.; Holguin, O.; Polle, J.E. Characterization and classification of highly productive microalgae strains discovered for biofuel and bioproduct generation. Algal Res. 2016, 15, 164–178. [Google Scholar] [CrossRef]
  64. Saini, D.K.; Chakdar, H.; Pabbi, S.; Shukla, P. Enhancing production of microalgal biopigments through metabolic and genetic engineering. Crit. Rev. Food Sci. Nutr. 2020, 60, 391–405. [Google Scholar] [CrossRef] [PubMed]
  65. Fayyaz, M.; Chew, K.W.; Show, P.L.; Ling, T.C.; Ng, I.-S.; Chang, J.-S. Genetic engineering of microalgae for enhanced biorefinery capabilities. Biotechnol. Adv. 2020, 43, 107554. [Google Scholar] [CrossRef]
  66. Benedetti, M.; Vecchi, V.; Barera, S.; Dall’Osto, L. Biomass from microalgae: The potential of domestication towards sustainable biofactories. Microb. Cell Factories 2018, 17, 173. [Google Scholar] [CrossRef] [PubMed]
  67. Rasala, B.A.; Mayfield, S.P. Photosynthetic biomanufacturing in green algae; production of recombinant proteins for industrial, nutritional, and medical uses. Photosynth. Res. 2015, 123, 227–239. [Google Scholar] [CrossRef] [PubMed]
  68. Rasala, B.A.; Lee, P.A.; Shen, Z.; Briggs, S.P.; Mendez, M.; Mayfield, S.P. Robust expression and secretion of Xylanase1 in Chlamydomonas reinhardtii by fusion to a selection gene and processing with the FMDV 2A peptide. PLoS ONE 2012, 7, e43349. [Google Scholar] [CrossRef]
  69. Georgianna, D.R.; Hannon, M.J.; Marcuschi, M.; Wu, S.; Botsch, K.; Lewis, A.J.; Hyun, J.; Mendez, M.; Mayfield, S.P. Production of recombinant enzymes in the marine alga Dunaliella tertiolecta. Algal Res. 2013, 2, 2–9. [Google Scholar] [CrossRef]
  70. Hou, Q.; Qiu, S.; Liu, Q.; Tian, J.; Hu, Z.; Ni, J. Selenoprotein-transgenic Chlamydomonas reinhardtii. Nutrients 2013, 5, 624–636. [Google Scholar] [CrossRef]
  71. Manuell, A.L.; Beligni, M.V.; Elder, J.H.; Siefker, D.T.; Tran, M.; Weber, A.; McDonald, T.L.; Mayfield, S.P. Robust expression of a bioactive mammalian protein in Chlamydomonas chloroplast. Plant Biotechnol. J. 2007, 5, 402–412. [Google Scholar] [CrossRef] [PubMed]
  72. Rasala, B.A.; Muto, M.; Lee, P.A.; Jager, M.; Cardoso, R.M.; Behnke, C.A.; Kirk, P.; Hokanson, C.A.; Crea, R.; Mendez, M. Production of therapeutic proteins in algae, analysis of expression of seven human proteins in the chloroplast of Chlamydomonas reinhardtii. Plant Biotechnol. J. 2010, 8, 719–733. [Google Scholar] [CrossRef] [PubMed]
  73. Feng, S.; Feng, W.; Zhao, L.; Gu, H.; Li, Q.; Shi, K.; Guo, S.; Zhang, N. Preparation of transgenic Dunaliella salina for immunization against white spot syndrome virus in crayfish. Arch. Virol. 2014, 159, 519–525. [Google Scholar] [CrossRef]
  74. Cha, T.-S.; Chen, C.-F.; Yee, W.; Aziz, A.; Loh, S.-H. Cinnamic acid, coumarin and vanillin: Alternative phenolic compounds for efficient Agrobacterium-mediated transformation of the unicellular green alga, Nannochloropsis sp. J. Microbiol. Methods 2011, 84, 430–434. [Google Scholar] [CrossRef]
  75. He, Y.; Peng, H.; Liu, J.; Chen, F.; Zhou, Y.; Ma, X.; Chen, H.; Wang, K. Chlorella sp. transgenic with Scy-hepc enhancing the survival of Sparus macrocephalus and hybrid grouper challenged with Aeromonas hydrophila. Fish Shellfish Immunol. 2018, 73, 22–29. [Google Scholar] [CrossRef]
  76. Lau, C.C.; Loh, S.H.; Aziz, A.; San Cha, T. Effects of disrupted omega-3 desaturase gene construct on fatty acid composition and expression of four fatty acid biosynthetic genes in transgenic Chlorella vulgaris. Algal Res. 2017, 26, 143–152. [Google Scholar] [CrossRef]
  77. Liu, J.; Sun, Z.; Gerken, H.; Huang, J.; Jiang, Y.; Chen, F. Genetic engineering of the green alga Chlorella zofingiensis: A modified norflurazon-resistant phytoene desaturase gene as a dominant selectable marker. Appl. Microbiol. Biotechnol. 2014, 98, 5069–5079. [Google Scholar] [CrossRef]
  78. Ma, X.; Pan, K.; Zhang, L.; Zhu, B.; Yang, G.; Zhang, X. Genetic transformation of Nannochloropsis oculata with a bacterial phleomycin resistance gene as dominant selective marker. J. Ocean. Univ. China 2016, 15, 351–356. [Google Scholar] [CrossRef]
  79. Prasad, B.; Vadakedath, N.; Jeong, H.-J.; General, T.; Cho, M.-G.; Lein, W. Agrobacterium tumefaciens-mediated genetic transformation of haptophytes (Isochrysis species). Appl. Microbiol. Biotechnol. 2014, 98, 8629–8639. [Google Scholar] [CrossRef]
  80. Prasad, B. Agrobacterium-mediated Nuclear Transformation of Haptophyte and Rhodophyte Species. Ph.D. Thesis, University of Erlangen-Nuremberg, Erlangen, Germany, 2017. [Google Scholar]
  81. Kim, D.-H.; Kim, Y.T.; Cho, J.J.; Bae, J.-H.; Hur, S.-B.; Hwang, I.; Choi, T.-J. Stable integration and functional expression of flounder growth hormone gene in transformed microalga, Chlorella ellipsoidea. Mar. Biotechnol. 2002, 4, 63–73. [Google Scholar] [CrossRef]
  82. Liu, J.; Gerken, H.; Huang, J.; Chen, F. Engineering of an endogenous phytoene desaturase gene as a dominant selectable marker for Chlamydomonas reinhardtii transformation and enhanced biosynthesis of carotenoids. Process Biochem. 2013, 48, 788–795. [Google Scholar] [CrossRef]
  83. Lin, W.-R.; Lai, Y.-C.; Sung, P.-K.; Tan, S.-I.; Chang, C.-H.; Chen, C.-Y.; Chang, J.-S.; Ng, I.-S. Enhancing carbon capture and lipid accumulation by genetic carbonic anhydrase in microalgae. J. Taiwan Inst. Chem. Eng. 2018, 93, 131–141. [Google Scholar] [CrossRef]
  84. Wei, L.; Wang, Q.; Xin, Y.; Lu, Y.; Xu, J. Enhancing photosynthetic biomass productivity of industrial oleaginous microalgae by overexpression of RuBisCO activase. Algal Res. 2017, 27, 366–375. [Google Scholar] [CrossRef]
  85. Endo, H.; Yoshida, M.; Uji, T.; Saga, N.; Inoue, K.; Nagasawa, H. Stable nuclear transformation system for the coccolithophorid alga Pleurochrysis carterae. Sci. Rep. 2016, 6, 22252. [Google Scholar] [CrossRef] [PubMed]
  86. Chai, X.-J.; Chen, H.-X.; Xu, W.-Q.; Xu, Y.-W. Expression of soybean Kunitz trypsin inhibitor gene SKTI in Dunaliella salina. J. Appl. Phycol. 2013, 25, 139–144. [Google Scholar] [CrossRef]
  87. Yang, Y.; Chai, Z.; Wang, Q.; Chen, W.; He, Z.; Jiang, S. Cultivation of seaweed Gracilaria in Chinese coastal waters and its contribution to environmental improvements. Algal Res. 2015, 9, 236–244. [Google Scholar] [CrossRef]
  88. Te, M.R.; Miller, D.J. Genetic transformation of dinoflagellates (Amphidinium and Symbiodinium): Expression of GUS in microalgae using heterologous promoter constructs. Plant, J. 1998, 13, 427–435. [Google Scholar] [CrossRef]
  89. Dunahay, T. Transformation of Chlamydomonas reinhardtii with silicon carbide whiskers. Biotechniques 1993, 15, 452–455, 457. [Google Scholar] [PubMed]
  90. Dong, B.; Cheng, R.-Q.; Liu, Q.-Y.; Wang, J.; Fan, Z.-C. Multimer of the antimicrobial peptide Mytichitin-A expressed in Chlamydomonas reinhardtii exerts a broader antibacterial spectrum and increased potency. J. Biosci. Bioeng. 2018, 125, 175–179. [Google Scholar] [CrossRef]
  91. Sizova, I.; Greiner, A.; Awasthi, M.; Kateriya, S.; Hegemann, P. Nuclear gene targeting in Chlamydomonas using engineered zinc-finger nucleases. Plant J. 2013, 73, 873–882. [Google Scholar] [CrossRef]
  92. Cui, Y.; Jiang, P.; Wang, J.; Li, F.; Chen, Y.; Zheng, G.; Qin, S. Genetic transformation of Platymonas (Tetraselmis) subcordiformis (Prasinophyceae, Chlorophyta) using particle bombardment and glass-bead agitation. Chin. J. Oceanol. Limnol. 2012, 30, 471–475. [Google Scholar] [CrossRef]
  93. Kathiresan, S.; Chandrashekar, A.; Ravishankar, G.; Sarada, R. Regulation of astaxanthin and its intermediates through cloning and genetic transformation of β-carotene ketolase in Haematococcus pluvialis. J. Biotechnol. 2015, 196, 33–41. [Google Scholar] [CrossRef] [PubMed]
  94. Hempel, F.; Lau, J.; Klingl, A.; Maier, U.G. Algae as protein factories: Expression of a human antibody and the respective antigen in the diatom Phaeodactylum tricornutum. PLoS ONE 2011, 6, e28424. [Google Scholar] [CrossRef]
  95. Miyagawa, Y.; Tamoi, M.; Shigeoka, S. Overexpression of a cyanobacterial fructose-1, 6-/sedoheptulose-1, 7-bisphosphatase in tobacco enhances photosynthesis and growth. Nat. Biotechnol. 2001, 19, 965–969. [Google Scholar] [CrossRef] [PubMed]
  96. Jaiswal, K.K.; Banerjee, I.; Singh, D.; Sajwan, P.; Chhetri, V. Ecological stress stimulus to improve microalgae biofuel generation: A review. Octa. J. Biosci. 2020, 8, 48–54. [Google Scholar]
  97. Madkour, F.F.; Kamil, A.E.-W.; Nasr, H.S. Production and nutritive value of Spirulina platensis in reduced cost media. Egypt. J. Aquat. Res. 2012, 38, 51–57. [Google Scholar] [CrossRef]
  98. González-González, L.M.; de-Bashan, L.E. Toward the enhancement of microalgal metabolite production through microalgae–bacteria consortia. Biology 2021, 10, 282. [Google Scholar] [CrossRef]
  99. 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, 36. [Google Scholar] [CrossRef]
  100. de Carvalho, J.C.; Sydney, E.B.; Tessari, L.F.A.; Soccol, C.R. Culture media for mass production of microalgae. In Biofuels from Algae; Elsevier: Amsterdam, The Netherlands, 2019; pp. 33–50. [Google Scholar]
  101. De Godos, I.; Mendoza, J.; Acién, F.; Molina, E.; Banks, C.; Heaven, S.; Rogalla, F. Evaluation of carbon dioxide mass transfer in raceway reactors for microalgae culture using flue gases. Bioresour. Technol. 2014, 153, 307–314. [Google Scholar] [CrossRef]
  102. Zerrouki, D.; Henni, A. Outdoor Microalgae Cultivation for Wastewater Treatment. In Application of Microalgae in Wastewater Treatment; Springer: Berlin/Heidelberg, Germany, 2019; pp. 81–99. [Google Scholar]
  103. Kumar, K.; Mishra, S.K.; Shrivastav, A.; Park, M.S.; Yang, J.-W. Recent trends in the mass cultivation of algae in raceway ponds. Renew. Sustain. Energy Rev. 2015, 51, 875–885. [Google Scholar] [CrossRef]
  104. Al-lwayzy, S.H.; Yusaf, T.; Al-Juboori, R.A. Biofuels from the fresh water microalgae Chlorella vulgaris (FWM-CV) for diesel engines. Energies 2014, 7, 1829–1851. [Google Scholar] [CrossRef]
  105. Rahul, M.; Sundaramahalingam, M.; Shivamthi, C.; Shyam, K.; Varalkshmi, P.; Karthikumar, S.; Kanimozhi, J.; Vinoth, K.; Sabarathinam, S.; Ganesh, M. Insights about sustainable biodiesel production from microalgae biomass: A review. Int. J. Energy Res. 2021, 45, 17028–17056. [Google Scholar] [CrossRef]
  106. Vickers, N.J. Animal communication: When i’m calling you, will you answer too? Curr. Biol. 2017, 27, R713–R715. [Google Scholar] [CrossRef] [PubMed]
  107. Prieto, A.; Cañavate, J.P.; García-González, M. Assessment of carotenoid production by Dunaliella salina in different culture systems and operation regimes. J. Biotechnol. 2011, 151, 180–185. [Google Scholar] [CrossRef] [PubMed]
  108. Bošnjaković, M.; Sinaga, N. The perspective of large-scale production of algae biodiesel. Appl. Sci. 2020, 10, 8181. [Google Scholar] [CrossRef]
  109. Li, J.; Zhu, D.; Niu, J.; Shen, S.; Wang, G. An economic assessment of astaxanthin production by large scale cultivation of Haematococcus pluvialis. Biotechnol. Adv. 2011, 29, 568–574. [Google Scholar] [CrossRef]
  110. Chandra, R.; Iqbal, H.M.; 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]
  111. Oslan, S.N.H.; Oslan, S.N.; Mohamad, R.; Tan, J.S.; Yusoff, A.H.; Matanjun, P.; Mokhtar, R.A.M.; Shapawi, R.; Huda, N. Bioprocess Strategy of Haematococcus lacustris for Biomass and Astaxanthin Production Keys to Commercialization: Perspective and Future Direction. Fermentation 2022, 8, 179. [Google Scholar] [CrossRef]
  112. Barati, B.; Gan, S.-Y.; Lim, P.-E.; Beardall, J.; Phang, S.-M. Green algal molecular responses to temperature stress. Acta Physiol. Plant. 2019, 41, 26. [Google Scholar] [CrossRef]
  113. Ras, M.; Steyer, J.-P.; Bernard, O. Temperature effect on microalgae: A crucial factor for outdoor production. Rev. Environ. Sci. Bio Technol. 2013, 12, 153–164. [Google Scholar] [CrossRef]
  114. Covarrubias, Y.; Cantoral-Uriza, E.A.; Casas-Flores, J.S.; García-Meza, J.V. Thermophile mats of microalgae growing on the woody structure of a cooling tower of a thermoelectric power plant in Central Mexico. Rev. Mex. Biodivers. 2016, 87, 277–287. [Google Scholar] [CrossRef]
  115. Ferro, L.; Gorzsás, A.; Gentili, F.G.; Funk, C. Subarctic microalgal strains treat wastewater and produce biomass at low temperature and short photoperiod. Algal Res. 2018, 35, 160–167. [Google Scholar] [CrossRef]
  116. Converti, A.; Casazza, A.A.; Ortiz, E.Y.; Perego, P.; Del Borghi, M. Effect of temperature and nitrogen concentration on the growth and lipid content of Nannochloropsis oculata and Chlorella vulgaris for biodiesel production. Chem. Eng. Process. Process Intensif. 2009, 48, 1146–1151. [Google Scholar] [CrossRef]
  117. Ho, S.-H.; Ye, X.; Hasunuma, T.; Chang, J.-S.; Kondo, A. Perspectives on engineering strategies for improving biofuel production from microalgae—A critical review. Biotechnol. Adv. 2014, 32, 1448–1459. [Google Scholar] [CrossRef] [PubMed]
  118. Gonçalves, C.F.; Menegol, T.; Rech, R. Biochemical composition of green microalgae Pseudoneochloris marina grown under different temperature and light conditions. Biocatal. Agric. Biotechnol. 2019, 18, 101032. [Google Scholar] [CrossRef]
  119. El-Khouly, M.E.; El-Mohsnawy, E.; Fukuzumi, S. Solar energy conversion: From natural to artificial photosynthesis. J. Photochem. Photobiol. C Photochem. Rev. 2017, 31, 36–83. [Google Scholar] [CrossRef]
  120. Amini Khoeyi, Z.; Seyfabadi, J.; Ramezanpour, Z. Effect of light intensity and photoperiod on biomass and fatty acid composition of the microalgae, Chlorella vulgaris. Aquac. Int. 2012, 20, 41–49. [Google Scholar] [CrossRef]
  121. Sforza, E.; Barbera, E.; Bertucco, A. Improving the photoconversion efficiency: An integrated photovoltaic-photobioreactor system for microalgal cultivation. Algal Res. 2015, 10, 202–209. [Google Scholar] [CrossRef]
  122. Mohsenpour, S.F.; Willoughby, N. Luminescent photobioreactor design for improved algal growth and photosynthetic pigment production through spectral conversion of light. Bioresour. Technol. 2013, 142, 147–153. [Google Scholar] [CrossRef]
  123. Das, P.; Lei, W.; Aziz, S.S.; Obbard, J.P. Enhanced algae growth in both phototrophic and mixotrophic culture under blue light. Bioresour. Technol. 2011, 102, 3883–3887. [Google Scholar] [CrossRef]
  124. Toledo-Cervantes, A.; Solórzano, G.G.; Campos, J.E.; Martínez-García, M.; Morales, M. Characterization of Scenedesmus obtusiusculus AT-UAM for high-energy molecules accumulation: Deeper insight into biotechnological potential of strains of the same species. Biotechnol. Rep. 2018, 17, 16–23. [Google Scholar] [CrossRef] [PubMed]
  125. Gifuni, I.; Olivieri, G.; Pollio, A.; Marzocchella, A. Identification of an industrial microalgal strain for starch production in biorefinery context: The effect of nitrogen and carbon concentration on starch accumulation. New Biotechnol. 2018, 41, 46–54. [Google Scholar] [CrossRef] [PubMed]
  126. de Farias Silva, C.E.; Sforza, E.; Bertucco, A. Stability of carbohydrate production in continuous microalgal cultivation under nitrogen limitation: Effect of irradiation regime and intensity on Tetradesmus obliquus. J. Appl. Phycol. 2018, 30, 261–270. [Google Scholar] [CrossRef]
  127. Coronel, C.D.; Do Nascimento, M.; Curatti, L. Effect of matching microalgal strains origin and regional weather condition on biomass productivity in environmental photobioreactors. Bioresour. Technol. Rep. 2019, 5, 104–112. [Google Scholar] [CrossRef]
  128. García-Cubero, R.; Cabanelas, I.T.D.; Sijtsma, L.; Kleinegris, D.M.; Barbosa, M.J. Production of exopolysaccharide by Botryococcus braunii CCALA 778 under laboratory simulated Mediterranean climate conditions. Algal Res. 2018, 29, 330–336. [Google Scholar] [CrossRef]
  129. de Farias Silva, C.E.; Barbera, E.; Bertucco, A. Biorefinery as a promising approach to promote ethanol industry from microalgae and cyanobacteria. In Bioethanol Production from Food Crops; Elsevier: Amsterdam, The Netherlands, 2019; pp. 343–359. [Google Scholar]
  130. Samiee-Zafarghandi, R.; Karimi-Sabet, J.; Abdoli, M.A.; Karbassi, A. Increasing microalgal carbohydrate content for hydrothermal gasification purposes. Renew. Energy 2018, 116, 710–719. [Google Scholar] [CrossRef]
  131. Selvakumar, P.; Umadevi, K. Mass cultivation of marine micro alga Nannochloropsis gaditana KF410818 isolated from Visakhapatnam offshore and fatty acid profile analysis for biodiesel production. J. Algal Biomass Util. 2014, 5, 28–37. [Google Scholar]
  132. Wu, L.F.; Chen, P.C.; Lee, C.M. The effects of nitrogen sources and temperature on cell growth and lipid accumulation of microalgae. Int. Biodeterior. Biodegrad. 2013, 85, 506–510. [Google Scholar] [CrossRef]
  133. Ashour, M.; Kamel, A. Enhance growth and biochemical composition of Nannochloropsis oceanica, cultured under nutrient limitation, using commercial agricultural fertilizers. J. Mar. Sci. Res. Dev 2017, 7, 233. [Google Scholar]
  134. Lai, Y.H.; Puspanadan, S.; Lee, C.K. Nutritional optimization of Arthrospira platensis for starch and total carbohydrates production. Biotechnol. Prog. 2019, 35, e2798. [Google Scholar] [CrossRef]
  135. Hanifzadeh, M.; Garcia, E.C.; Viamajala, S. Production of lipid and carbohydrate from microalgae without compromising biomass productivities: Role of Ca and Mg. Renew. Energy 2018, 127, 989–997. [Google Scholar] [CrossRef]
  136. Mathiot, C.; Ponge, P.; Gallard, B.; Sassi, J.-F.; Delrue, F.; Le Moigne, N. Microalgae starch-based bioplastics: Screening of ten strains and plasticization of unfractionated microalgae by extrusion. Carbohydr. Polym. 2019, 208, 142–151. [Google Scholar] [CrossRef] [PubMed]
  137. Dammak, M.; Hadrich, B.; Miladi, R.; Barkallah, M.; Hentati, F.; Hachicha, R.; Laroche, C.; Michaud, P.; Fendri, I.; Abdelkafi, S. Effects of nutritional conditions on growth and biochemical composition of Tetraselmis sp. Lipids Health Dis. 2017, 16, 41. [Google Scholar] [CrossRef] [PubMed]
  138. Hernández-García, A.; Velásquez-Orta, S.B.; Novelo, E.; Yáñez-Noguez, I.; Monje-Ramírez, I.; Ledesma, M.T.O. Wastewater-leachate treatment by microalgae: Biomass, carbohydrate and lipid production. Ecotoxicol. Environ. Saf. 2019, 174, 435–444. [Google Scholar] [CrossRef] [PubMed]
  139. Choi, Y.Y.; Hong, M.-E.; Chang, W.S.; Sim, S.J. Autotrophic biodiesel production from the thermotolerant microalga Chlorella sorokiniana by enhancing the carbon availability with temperature adjustment. Biotechnol. Bioprocess Eng. 2019, 24, 223–231. [Google Scholar] [CrossRef]
  140. Pancha, I.; Chokshi, K.; Ghosh, T.; Paliwal, C.; Maurya, R.; Mishra, S. Bicarbonate supplementation enhanced biofuel production potential as well as nutritional stress mitigation in the microalgae Scenedesmus sp. CCNM 1077. Bioresour. Technol. 2015, 193, 315–323. [Google Scholar] [CrossRef]
  141. Varshney, P.; Beardall, J.; Bhattacharya, S.; Wangikar, P.P. Isolation and biochemical characterisation of two thermophilic green algal species-Asterarcys quadricellulare and Chlorella sorokiniana, which are tolerant to high levels of carbon dioxide and nitric oxide. Algal Res. 2018, 30, 28–37. [Google Scholar] [CrossRef]
  142. Sengupta, S.; Nawaz, T.; Beaudry, J. Nitrogen and phosphorus recovery from wastewater. Curr. Pollut. Rep. 2015, 1, 155–166. [Google Scholar] [CrossRef]
  143. Walsh, M.J.; Van Doren, L.G.; Shete, N.; Prakash, A.; Salim, U. Financial tradeoffs of energy and food uses of algal biomass under stochastic conditions. Appl. Energy 2018, 210, 591–603. [Google Scholar] [CrossRef]
  144. Lerat, Y.; Cornish, M.; Critchley, A.T. Applications of algal biomass in global food and feed markets: From traditional usage to the potential for functional products. Blue Biotechnol. Prod. Use Mar. Mol. 2018, 1, 143–189. [Google Scholar]
  145. Hawrot-Paw, M.; Koniuszy, A.; Gałczyńska, M.; Zając, G.; Szyszlak-Bargłowicz, J. Production of microalgal biomass using aquaculture wastewater as growth medium. Water 2019, 12, 106. [Google Scholar] [CrossRef]
  146. Sfez, S.; Van Den Hende, S.; Taelman, S.E.; De Meester, S.; Dewulf, J. Environmental sustainability assessment of a microalgae raceway pond treating aquaculture wastewater: From up-scaling to system integration. Bioresour. Technol. 2015, 190, 321–331. [Google Scholar] [CrossRef] [PubMed]
  147. Boopathy, R.; Fontenot, Q.; Kilgen, M.B. Biological treatment of sludge from a recirculating aquaculture system using a sequencing batch reactor. J. World Aquac. Soc. 2005, 36, 542–545. [Google Scholar] [CrossRef]
  148. Boopathy, R.; Bonvillain, C.; Fontenot, Q.; Kilgen, M. Biological treatment of low-salinity shrimp aquaculture wastewater using sequencing batch reactor. Int. Biodeterior. Biodegrad. 2007, 59, 16–19. [Google Scholar] [CrossRef]
  149. Schulz, C.; Gelbrecht, J.; Rennert, B. Constructed wetlands with free water surface for treatment of aquaculture effluents. J. Appl. Ichthyol. 2004, 20, 64–70. [Google Scholar] [CrossRef]
  150. Mirzoyan, N.; Parnes, S.; Singer, A.; Tal, Y.; Sowers, K.; Gross, A. Quality of brackish aquaculture sludge and its suitability for anaerobic digestion and methane production in an upflow anaerobic sludge blanket (UASB) reactor. Aquaculture 2008, 279, 35–41. [Google Scholar] [CrossRef]
  151. Suhr, K.I.; Pedersen, L.-F.; Nielsen, J.L. End-of-pipe single-sludge denitrification in pilot-scale recirculating aquaculture systems. Aquac. Eng. 2014, 62, 28–35. [Google Scholar] [CrossRef]
  152. Kumar, V.; Gururani, P.; Parveen, A.; Verma, M.; Kim, H.; Vlaskin, M.; Grigorenko, A.V.; Rindin, K.G. Dairy Industry wastewater and stormwater energy valorization: Effect of wastewater nutrients on microalgae-yeast biomass. Biomass Convers. Biorefinery 2022, 1–10. [Google Scholar] [CrossRef]
  153. Yogev, U.; Sowers, K.R.; Mozes, N.; Gross, A. Nitrogen and carbon balance in a novel near-zero water exchange saline recirculating aquaculture system. Aquaculture 2017, 467, 118–126. [Google Scholar] [CrossRef]
  154. Barbu, M.; Ceangă, E.; Caraman, S. Water quality modeling and control in recirculating aquaculture systems. Urban Agric. 2018, 2, 64. [Google Scholar]
  155. Pittman, J.K.; Dean, A.P.; Osundeko, O. The potential of sustainable algal biofuel production using wastewater resources. Bioresour. Technol. 2011, 102, 17–25. [Google Scholar] [CrossRef] [PubMed]
  156. Rawat, I.; Kumar, R.R.; Mutanda, T.; Bux, F. Dual role of microalgae: Phycoremediation of domestic wastewater and biomass production for sustainable biofuels production. Appl. Energy 2011, 88, 3411–3424. [Google Scholar] [CrossRef]
  157. Wu, J.-Y.; Lay, C.-H.; Chen, C.-C.; Wu, S.-Y. Lipid accumulating microalgae cultivation in textile wastewater: Environmental parameters optimization. J. Taiwan Inst. Chem. Eng. 2017, 79, 1–6. [Google Scholar] [CrossRef]
  158. El-Kassas, H.Y.; Mohamed, L.A. Bioremediation of the textile waste effluent by Chlorella vulgaris. Egypt. J. Aquat. Res. 2014, 40, 301–308. [Google Scholar] [CrossRef]
  159. Gopalakrishnan, V.; Ramamurthy, D. Dyeing industry effluent system as lipid production medium of Neochloris sp. for biodiesel feedstock preparation. BioMed Res. Int. 2014, 2014, 529560. [Google Scholar] [CrossRef]
  160. Parikh, A.; Madamwar, D. Textile dye decolorization using cyanobacteria. Biotechnol. Lett. 2005, 27, 323–326. [Google Scholar] [CrossRef]
  161. Verma, K.; Kumar, P.K.; Krishna, S.V.; Himabindu, V. Phycoremediation of Sewage-Contaminated Lake Water Using Microalgae–Bacteria Co-Culture. Water Air Soil Pollut. 2020, 231, 299. [Google Scholar] [CrossRef]
  162. Luo, H.; Moran, M.A. Evolutionary ecology of the marine Roseobacter clade. Microbiol. Mol. Biol. Rev. 2014, 78, 573–587. [Google Scholar] [CrossRef]
  163. Paddock, M.B.; Fernández-Bayo, J.D.; VanderGheynst, J.S. The effect of the microalgae-bacteria microbiome on wastewater treatment and biomass production. Appl. Microbiol. Biotechnol. 2020, 104, 893–905. [Google Scholar] [CrossRef]
  164. Fuentes, J.L.; Garbayo, I.; Cuaresma, M.; Montero, Z.; González-del-Valle, M.; Vílchez, C. Impact of microalgae-bacteria interactions on the production of algal biomass and associated compounds. Mar. Drugs 2016, 14, 100. [Google Scholar] [CrossRef]
  165. Peng, H.; de-Bashan, L.E.; Bashan, Y.; Higgins, B.T. Indole-3-acetic acid from Azosprillum brasilense promotes growth in green algae at the expense of energy storage products. Algal Res. 2020, 47, 101845. [Google Scholar] [CrossRef]
  166. Sandhya, S.; Vijayan, K. Symbiotic association among marine microalgae and bacterial flora: A study with special reference to commercially important Isochrysis galbana culture. J. Appl. Phycol. 2019, 31, 2259–2266. [Google Scholar] [CrossRef]
  167. Xue, L.; Shang, H.; Ma, P.; Wang, X.; He, X.; Niu, J.; Wu, J. Analysis of growth and lipid production characteristics of Chlorella vulgaris in artificially constructed consortia with symbiotic bacteria. J. Basic Microbiol. 2018, 58, 358–367. [Google Scholar] [CrossRef] [PubMed]
  168. Park, J.; Park, B.S.; Wang, P.; Patidar, S.K.; Kim, J.H.; Kim, S.-H.; Han, M.-S. Phycospheric native bacteria Pelagibaca bermudensis and Stappia sp. ameliorate biomass productivity of Tetraselmis striata (KCTC1432BP) in co-cultivation system through mutualistic interaction. Front. Plant Sci. 2017, 8, 289. [Google Scholar] [CrossRef]
  169. Higgins, B.T.; VanderGheynst, J.S. Effects of Escherichia coli on mixotrophic growth of Chlorella minutissima and production of biofuel precursors. PLoS ONE 2014, 9, e96807. [Google Scholar] [CrossRef]
  170. Paul, C.; Mausz, M.A.; Pohnert, G. A co-culturing/metabolomics approach to investigate chemically mediated interactions of planktonic organisms reveals influence of bacteria on diatom metabolism. Metabolomics 2013, 9, 349–359. [Google Scholar] [CrossRef]
  171. Do Nascimento, M.; de los Angeles Dublan, M.; Ortiz-Marquez, J.C.F.; Curatti, L. High lipid productivity of an Ankistrodesmus–Rhizobium artificial consortium. Bioresour. Technol. 2013, 146, 400–407. [Google Scholar] [CrossRef]
  172. Le Chevanton, M.; Garnier, M.; Bougaran, G.; Schreiber, N.; Lukomska, E.; Bérard, J.-B.; Fouilland, E.; Bernard, O.; Cadoret, J.-P. Screening and selection of growth-promoting bacteria for Dunaliella cultures. Algal Res. 2013, 2, 212–222. [Google Scholar] [CrossRef]
  173. Choix, F.J.; De-Bashan, L.E.; Bashan, Y. Enhanced accumulation of starch and total carbohydrates in alginate-immobilized Chlorella spp. induced by Azospirillum brasilense: II. Heterotrophic conditions. Enzym. Microb. Technol. 2012, 51, 300–309. [Google Scholar] [CrossRef]
  174. Rivas, M.O.; Vargas, P.; Riquelme, C.E. Interactions of Botryococcus braunii cultures with bacterial biofilms. Microb. Ecol. 2010, 60, 628–635. [Google Scholar] [CrossRef]
  175. Butler, T.; Golan, Y. Astaxanthin production from microalgae. In Microalgae Biotechnology for Food, Health and High Value Products; Springer: Berlin/Heidelberg, Germany, 2020; pp. 175–242. [Google Scholar]
  176. Mathimani, T.; Mallick, N. A comprehensive review on harvesting of microalgae for biodiesel—Key challenges and future directions. Renew. Sustain. Energy Rev. 2018, 91, 1103–1120. [Google Scholar] [CrossRef]
  177. Han, P.; Lu, Q.; Fan, L.; Zhou, W. A review on the use of microalgae for sustainable aquaculture. Appl. Sci. 2019, 9, 2377. [Google Scholar] [CrossRef]
  178. Levine, I.; Fleurence, J. Microalgae in Health and Disease Prevention; Academic Press: Cambridge, MA, USA, 2018. [Google Scholar]
  179. Alam, M.A.; Wang, Z. Microalgae Biotechnology for Development of Biofuel and Wastewater Treatment; Springer: Berlin/Heidelberg, Germany, 2019. [Google Scholar]
  180. Zhou, W.; Min, M.; Hu, B.; Ma, X.; Liu, Y.; Wang, Q.; Shi, J.; Chen, P.; Ruan, R. Filamentous fungi assisted bio-flocculation: A novel alternative technique for harvesting heterotrophic and autotrophic microalgal cells. Sep. Purif. Technol. 2013, 107, 158–165. [Google Scholar] [CrossRef]
  181. Chen, J.; Leng, L.; Ye, C.; Lu, Q.; Addy, M.; Wang, J.; Liu, J.; Chen, P.; Ruan, R.; Zhou, W. A comparative study between fungal pellet-and spore-assisted microalgae harvesting methods for algae bioflocculation. Bioresour. Technol. 2018, 259, 181–190. [Google Scholar] [CrossRef]
  182. Liu, H.; Lu, Q.; Wang, Q.; Liu, W.; Wei, Q.; Ren, H.; Ming, C.; Min, M.; Chen, P.; Ruan, R. Isolation of a bacterial strain, Acinetobacter sp. from centrate wastewater and study of its cooperation with algae in nutrients removal. Bioresour. Technol. 2017, 235, 59–69. [Google Scholar] [CrossRef]
  183. Naaz, F.; Bhattacharya, A.; Pant, K.K.; Malik, A. Impact of heavy metal laden algal biomass on hydrothermal liquefaction and biorefinery approach. Process Saf. Environ. Prot. 2021, 145, 141–149. [Google Scholar] [CrossRef]
  184. Zhang, Q.; Yu, Z.; Zhu, L.; Ye, T.; Zuo, J.; Li, X.; Xiao, B.; Jin, S. Vertical-algal-biofilm enhanced raceway pond for cost-effective wastewater treatment and value-added products production. Water Res. 2018, 139, 144–157. [Google Scholar] [CrossRef]
  185. de Farias Neves, F.; Demarco, M.; Tribuzi, G. Drying and quality of microalgal powders for human alimentation. In Microalgae-from Physiology to Application; IntechOpen: London, UK, 2019. [Google Scholar]
  186. Brennan, L.; Owende, P. Biofuels from microalgae—A review of technologies for production, processing, and extractions of biofuels and co-products. Renew. Sustain. Energy Rev. 2010, 14, 557–577. [Google Scholar] [CrossRef]
  187. Chen, C.-L.; Chang, J.-S.; Lee, D.-J. Dewatering and drying methods for microalgae. Dry. Technol. 2015, 33, 443–454. [Google Scholar] [CrossRef]
  188. Kishor, R.; Purchase, D.; Saratale, G.D.; Saratale, R.G.; Ferreira, L.F.R.; Bilal, M.; Chandra, R.; Bharagava, R.N. Ecotoxicological and health concerns of persistent coloring pollutants of textile industry wastewater and treatment approaches for environmental safety. J. Environ. Chem. Eng. 2021, 9, 105012. [Google Scholar] [CrossRef]
  189. Tan, X.B.; Lam, M.K.; Uemura, Y.; Lim, J.W.; Wong, C.Y.; Lee, K.T. Cultivation of microalgae for biodiesel production: A review on upstream and downstream processing. Chin. J. Chem. Eng. 2018, 26, 17–30. [Google Scholar] [CrossRef]
  190. Ruiz-Domínguez, M.C.; Marticorena, P.; Sepúlveda, C.; Salinas, F.; Cerezal, P.; Riquelme, C. Effect of drying methods on lutein content and recovery by supercritical extraction from the microalga Muriellopsis sp.(MCH35) cultivated in the arid north of Chile. Mar. Drugs 2020, 18, 528. [Google Scholar] [CrossRef] [PubMed]
  191. Stramarkou, M.; Papadaki, S.; Kyriakopoulou, K.; Krokida, M. Effect of drying and extraction conditions on the recovery of bioactive compounds from Chlorella vulgaris. J. Appl. Phycol. 2017, 29, 2947–2960. [Google Scholar] [CrossRef]
  192. Saini, R.K.; Keum, Y.-S. Carotenoid extraction methods: A review of recent developments. Food Chem. 2018, 240, 90–103. [Google Scholar] [CrossRef]
  193. Monteiro, M.; Santos, R.; Iglesias, P.; Couto, A.; Serra, C.; Gouvinhas, I.; Barros, A.; Oliva-Teles, A.; Enes, P.; Díaz-Rosales, P. Effect of extraction method and solvent system on the phenolic content and antioxidant activity of selected macro-and microalgae extracts. J. Appl. Phycol. 2020, 32, 349–362. [Google Scholar] [CrossRef]
  194. Zhang, R.; Parniakov, O.; Grimi, N.; Lebovka, N.; Marchal, L.; Vorobiev, E. Emerging techniques for cell disruption and extraction of valuable bio-molecules of microalgae Nannochloropsis sp. Bioprocess Biosyst. Eng. 2019, 42, 173–186. [Google Scholar] [CrossRef]
  195. Onay, M.; Sonmez, C.; Oktem, H.A.; Yucel, M. Evaluation of various extraction techniques for efficient lipid recovery from thermo-resistant microalgae, Hindakia, Scenedesmus and Micractinium species—Comparison of lipid extraction methods from microalgae. Am. J. Anal. Chem. 2016, 7, 141. [Google Scholar] [CrossRef]
  196. Folch, J.; Lees, M.; Sloane Stanley, G.H. A simple method for the isolation and purification of total lipids from animal tissues. J. Biol. Chem. 1957, 226, 497–509. [Google Scholar] [CrossRef]
  197. Bligh, E.G.; Dyer, W.J. A rapid method of total lipid extraction and purification. Can. J. Biochem. Physiol. 1959, 37, 911–917. [Google Scholar] [CrossRef]
  198. De Castro, M.L.; Priego-Capote, F. Soxhlet extraction: Past and present panacea. J. Chromatogr. A 2010, 1217, 2383–2389. [Google Scholar] [CrossRef]
  199. Koberg, M.; Cohen, M.; Ben-Amotz, A.; Gedanken, A. Bio-diesel production directly from the microalgae biomass of Nannochloropsis by microwave and ultrasound radiation. Bioresour. Technol. 2011, 102, 4265–4269. [Google Scholar] [CrossRef]
  200. Cui, Y.; Liang, Y. Direct transesterification of wet Cryptococcus curvatus cells to biodiesel through use of microwave irradiation. Appl. Energy 2014, 119, 438–444. [Google Scholar] [CrossRef]
  201. Cheng, J.; Yu, T.; Li, T.; Zhou, J.; Cen, K. Using wet microalgae for direct biodiesel production via microwave irradiation. Bioresour. Technol. 2013, 131, 531–535. [Google Scholar] [CrossRef] [PubMed]
  202. Lee, S.Y.; Khoiroh, I.; Vo, D.-V.N.; Senthil Kumar, P.; Show, P.L. Techniques of lipid extraction from microalgae for biofuel production: A review. Environ. Chem. Lett. 2021, 19, 231–251. [Google Scholar] [CrossRef]
  203. Xie, Y.; Ho, S.-H.; Chen, C.-N.N.; Chen, C.-Y.; Jing, K.; Ng, I.-S.; Chen, J.; Chang, J.-S.; Lu, Y. Disruption of thermo-tolerant Desmodesmus sp. F51 in high pressure homogenization as a prelude to carotenoids extraction. Biochem. Eng. J. 2016, 109, 243–251. [Google Scholar] [CrossRef]
  204. Wang, R.; Wang, Z.; Zhang, J. Science of Marine Shellfish Culture; Ocean University of China: Qingdao, China, 2008. [Google Scholar]
  205. Rivera, E.C.; Montalescot, V.; Viau, M.; Drouin, D.; Bourseau, P.; Frappart, M.; Monteux, C.; Couallier, E. Mechanical cell disruption of Parachlorella kessleri microalgae: Impact on lipid fraction composition. Bioresour. Technol. 2018, 256, 77–85. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Zheng, H.; Yin, J.; Gao, Z.; Huang, H.; Ji, X.; Dou, C. Disruption of Chlorella vulgaris cells for the release of biodiesel-producing lipids: A comparison of grinding, ultrasonication, bead milling, enzymatic lysis, and microwaves. Appl. Biochem. Biotechnol. 2011, 164, 1215–1224. [Google Scholar] [CrossRef]
  207. Greenly, J.M.; Tester, J.W. Ultrasonic cavitation for disruption of microalgae. Bioresour. Technol. 2015, 184, 276–279. [Google Scholar] [CrossRef]
  208. Lee, A.K.; Lewis, D.M.; Ashman, P.J. Microalgal cell disruption by hydrodynamic cavitation for the production of biofuels. J. Appl. Phycol. 2015, 27, 1881–1889. [Google Scholar] [CrossRef]
  209. Khedkar, S.V.; Chavan, Y.; Bhagat, S. Extraction of oil from algae by solvent extraction and oil expeller method. Sadguru Publ. 2011, 9, 1746–1750. [Google Scholar]
  210. Thoisen, C.; Hansen, B.W.; Nielsen, S.L. A simple and fast method for extraction and quantification of cryptophyte phycoerythrin. MethodsX 2017, 4, 209–213. [Google Scholar] [CrossRef] [PubMed]
  211. González-González, L.M.; Astals, S.; Pratt, S.; Jensen, P.D.; Schenk, P.M. Impact of osmotic shock pre-treatment on microalgae lipid extraction and subsequent methane production. Bioresour. Technol. Rep. 2019, 7, 100214. [Google Scholar] [CrossRef]
  212. Günerken, E.; D’hondt, E.; Eppink, M.H.; Wijffels, R.H.; Elst, K. Disruption of microalgae with a novel continuous explosive decompression device. Algal Res. 2019, 39, 101376. [Google Scholar] [CrossRef]
  213. Gim, G.H.; Kim, S.W. Optimization of cell disruption and transesterification of lipids from Botryococcus braunii LB572. Biotechnol. Bioprocess Eng. 2018, 23, 550–556. [Google Scholar] [CrossRef]
  214. Sheng, J.; Vannela, R.; Rittmann, B. Disruption of Synechocystis PCC 6803 for lipid extraction. Water Sci. Technol. 2012, 65, 567–573. [Google Scholar] [CrossRef]
  215. Harun, R.; Danquah, M.K. Influence of acid pre-treatment on microalgal biomass for bioethanol production. Process Biochem. 2011, 46, 304–309. [Google Scholar] [CrossRef]
  216. Sert, B.Ş.; İnan, B.; Özçimen, D. Effect of chemical pre-treatments on bioethanol production from Chlorella minutissima. Acta Chim. Slov. 2018, 65, 160–165. [Google Scholar] [CrossRef] [Green Version]
  217. Zuorro, A.; Maffei, G.; Lavecchia, R. Optimization of enzyme-assisted lipid extraction from Nannochloropsis microalgae. J. Taiwan Inst. Chem. Eng. 2016, 67, 106–114. [Google Scholar] [CrossRef]
  218. Armenta, S.; Garrigues, S.; Esteve-Turrillas, F.A.; de la Guardia, M. Green extraction techniques in green analytical chemistry. TrAC Trends Anal. Chem. 2019, 116, 248–253. [Google Scholar] [CrossRef]
  219. Lee, J.-Y.; Yoo, C.; Jun, S.-Y.; Ahn, C.-Y.; Oh, H.-M. Comparison of several methods for effective lipid extraction from microalgae. Bioresour. Technol. 2010, 101, S75–S77. [Google Scholar] [CrossRef]
  220. Chu, W.-L.; Lim, Y.-W.; Radhakrishnan, A.K.; Lim, P.-E. Protective effect of aqueous extract from Spirulina platensis against cell death induced by free radicals. BMC Complement. Altern. Med. 2010, 10, 53. [Google Scholar] [CrossRef] [PubMed]
  221. Guerin, M.; Huntley, M.E.; Olaizola, M. Haematococcus astaxanthin: Applications for human health and nutrition. Trends Biotechnol. 2003, 21, 210–216. [Google Scholar] [CrossRef]
  222. Jin, E.S.; Melis, A. Microalgal biotechnology: Carotenoid production by the green algae Dunaliella salina. Biotechnol. Bioprocess Eng. 2003, 8, 331–337. [Google Scholar] [CrossRef]
  223. Del Campo, J.A.; García-González, M.; Guerrero, M.G. Outdoor cultivation of microalgae for carotenoid production: Current state and perspectives. Appl. Microbiol. Biotechnol. 2007, 74, 1163–1174. [Google Scholar] [CrossRef]
  224. Reis, A.; Gouveia, L.; Veloso, V.; Fernandes, H.L.; Empis, J.; Novais, J.M. Eicosapentaenoic acid-rich biomass production by the microalga Phaeodactylum tricornutum in a continuous-flow reactor. Bioresour. Technol. 1996, 55, 83–88. [Google Scholar] [CrossRef]
  225. Asgharpour, M.; Rodgers, B.; Hestekin, J.A. Eicosapentaenoic acid from Porphyridium cruentum: Increasing growth and productivity of microalgae for pharmaceutical products. Energies 2015, 8, 10487–10503. [Google Scholar] [CrossRef]
  226. Mendes, A.; Reis, A.; Vasconcelos, R.; Guerra, P.; Lopes da Silva, T. Crypthecodinium cohnii with emphasis on DHA production: A review. J. Appl. Phycol. 2009, 21, 199–214. [Google Scholar] [CrossRef]
  227. Forján Lozano, E.; Garbayo Nores, I.; Casal Bejarano, C.; Vílchez Lobato, C. Enhancement of carotenoid production in Nannochloropsis by phosphate and sulphur limitation. In Communicating Current Research and Educational Topics and Trends in Applied Microbiology; Formatex: Badajoz, Spain, 2007. [Google Scholar]
  228. Mišurcová, L.; Škrovánková, S.; Samek, D.; Ambrožová, J.; Machů, L. Health benefits of algal polysaccharides in human nutrition. Adv. Food Nutr. Res. 2012, 66, 75–145. [Google Scholar]
  229. Rahman, K.M. Food and high value products from microalgae: Market opportunities and challenges. In Microalgae Biotechnology for Food, Health and High Value Products; Springer: Singapore, 2020; pp. 3–27. [Google Scholar]
  230. Eckelberry, R. Algae—Food or chemical grade. Algae Ind. Mag. 2011. Available online: http://www.algaeindustrymagazine.com/algae-business-algae-food-or-chemical-grade/ (accessed on 29 July 2022).
  231. Taufiqurrahmi, N.; Religia, P.; Mulyani, G.; Suryana, D.; Tanjung, F.; Arifin, Y. Phycocyanin extraction in Spirulina produced using agricultural waste. Proc. IOP Conf. Ser. Mater. Sci. Eng. 2017, 296, 012097. [Google Scholar] [CrossRef]
  232. Zanella, L.; Alam, M. Extracts and bioactives from microalgae (sensu stricto): Opportunities and challenges for a new generation of cosmetics. In Microalgae Biotechnology for Food, Health and High Value Products; Springer: Singapore, 2020; pp. 295–349. [Google Scholar]
  233. Mehta, P.; McAuley, D.F.; Brown, M.; Sanchez, E.; Tattersall, R.S.; Manson, J.J. COVID-19: Consider cytokine storm syndromes and immunosuppression. Lancet 2020, 395, 1033–1034. [Google Scholar] [CrossRef]
  234. Mittermaier, M.; Pickerodt, P.; Kurth, F.; de Jarcy, L.B.; Uhrig, A.; Garcia, C.; Machleidt, F.; Pergantis, P.; Weber, S.; Li, Y. Evaluation of PEEP and prone positioning in early COVID-19 ARDS. EClinicalMedicine 2020, 28, 100579. [Google Scholar] [CrossRef] [PubMed]
  235. Khavari, F.; Saidijam, M.; Taheri, M.; Nouri, F. Microalgae: Therapeutic potentials and applications. Mol. Biol. Rep. 2021, 48, 4757–4765. [Google Scholar] [CrossRef] [PubMed]
  236. Singh, S.; Dwivedi, V.; Sanyal, D.; Dasgupta, S. Therapeutic and nutritional potential of Spirulina in combating COVID-19 infection. AIJR Preprints 2020, 49, 2–8. [Google Scholar]
  237. Bao, J.; Li, C.; Zhang, K.; Kang, H.; Chen, W.; Gu, B. Comparative analysis of laboratory indexes of severe and non-severe patients infected with COVID-19. Clin. Chim. Acta 2020, 509, 180–194. [Google Scholar] [CrossRef] [PubMed]
  238. Zhou, F.; Yu, T.; Du, R.; Fan, G.; Liu, Y.; Liu, Z.; Xiang, J.; Wang, Y.; Song, B.; Gu, X. Clinical course and risk factors for mortality of adult inpatients with COVID-19 in Wuhan, China: A retrospective cohort study. Lancet 2020, 395, 1054–1062. [Google Scholar] [CrossRef]
  239. Lu, L.; Zhang, H.; Zhan, M.; Jiang, J.; Yin, H.; Dauphars, D. Preventing mortality in COVID-19 patients: Which cytokine to target in a raging storm? Front. Cell. Dev. Biol. 2020, 8, 677. [Google Scholar] [CrossRef]
  240. Ye, Q.; Wang, B.; Mao, J. The pathogenesis and treatment of theCytokine Storm’in COVID-19. J. Infect. 2020, 80, 607–613. [Google Scholar] [CrossRef]
  241. Shimizu, M. Clinical features of cytokine storm syndrome. In Cytokine Storm Syndrome; Springer: Berlin/Heidelberg, Germany, 2019; pp. 31–41. [Google Scholar]
  242. Wan Afifudeen, C.L.; Teh, K.Y.; Cha, T.S. Bioprospecting of microalgae metabolites against cytokine storm syndrome during COVID-19. Mol. Biol. Rep. 2021, 49, 1475–1490. [Google Scholar] [CrossRef]
  243. Moldoveanu, B.; Otmishi, P.; Jani, P.; Walker, J.; Sarmiento, X.; Guardiola, J.; Saad, M.; Yu, J. Inflammatory mechanisms in the lung. J. Inflamm. Res. 2009, 2, 1. [Google Scholar]
  244. Chia, W.Y.; Kok, H.; Chew, K.W.; Low, S.S.; Show, P.L. Can algae contribute to the war with Covid-19? Bioengineered 2021, 12, 1226–1237. [Google Scholar] [CrossRef]
  245. Barrantes, M.G.; Espinoza, O.G.; Montero, K.N. Microalgae-based approaches to overcome the effects of the COVID-19 pandemic. Tecnol. Marcha 2022, 35, 84–93. [Google Scholar]
  246. Norambuena, F.; Hermon, K.; Skrzypczyk, V.; Emery, J.A.; Sharon, Y.; Beard, A.; Turchini, G.M. Algae in fish feed: Performances and fatty acid metabolism in juvenile Atlantic salmon. PLoS ONE 2015, 10, e0124042. [Google Scholar] [CrossRef] [PubMed]
  247. Michalak, I.; Chojnacka, K. Algae as production systems of bioactive compounds. Eng. Life Sci. 2015, 15, 160–176. [Google Scholar] [CrossRef]
  248. Yuan, Y.; Zhang, J.; Fan, J.; Clark, J.; Shen, P.; Li, Y.; Zhang, C. Microwave assisted extraction of phenolic compounds from four economic brown macroalgae species and evaluation of their antioxidant activities and inhibitory effects on α-amylase, α-glucosidase, pancreatic lipase and tyrosinase. Food Res. Int. 2018, 113, 288–297. [Google Scholar] [CrossRef]
  249. Hamed, I.; Özogul, F.; Özogul, Y.; Regenstein, J.M. Marine bioactive compounds and their health benefits: A review. Compr. Rev. Food Sci. Food Saf. 2015, 14, 446–465. [Google Scholar] [CrossRef]
  250. Yaakob, Z.; Ali, E.; Zainal, A.; Mohamad, M.; Takriff, M.S. An overview: Biomolecules from microalgae for animal feed and aquaculture. J. Biol. Res. Thessalon. 2014, 21, 6. [Google Scholar] [CrossRef]
  251. Wang, J.; Hu, X.; Chen, J.; Wang, T.; Huang, X.; Chen, G. The Extraction of β-Carotene from Microalgae for Testing Their Health Benefits. Foods 2022, 11, 502. [Google Scholar] [CrossRef]
  252. Jung, F.; Krüger-Genge, A.; Waldeck, P.; Küpper, J.-H. Spirulina platensis, a super food? J. Cell. Biotechnol. 2019, 5, 43–54. [Google Scholar] [CrossRef]
  253. Mathur, M. Bioactive Molecules of: A Food Supplement. In Bioactive Molecules in Food; Springer: Berlin/Heidelberg, Germany, 2019; pp. 1621–1642. [Google Scholar]
  254. Morais, M.G.d.; Vaz, B.d.S.; Morais, E.G.d.; Costa, J.A.V. Biological effects of Spirulina (Arthrospira) biopolymers and biomass in the development of nanostructured scaffolds. BioMed Res. Int. 2014, 2014, 762705. [Google Scholar] [CrossRef]
  255. de Sousa e Silva, A.; de Magalhaes, W.T.; Moreira, L.M.; Rocha, M.V.P.; Bastos, A.K.P. Microwave-assisted extraction of polysaccharides from Arthrospira (Spirulina) platensis using the concept of green chemistry. Algal Res. 2018, 35, 178–184. [Google Scholar] [CrossRef]
  256. Nuhu, A.A. Spirulina (Arthrospira): An important source of nutritional and medicinal compounds. J. Mar. Biol. 2013, 2013, 325636. [Google Scholar] [CrossRef] [Green Version]
  257. Rajasekaran, C.; Ajeesh, C.M.; Balaji, S.; Shalini, M.; Ramamoorthy, S.; Ranjan, D.; Fulzele, D.P.; Kalaivani, T. Effect of modified Zarrouk’s medium on growth of different Spirulina strains. Walailak J. Sci. Technol. 2016, 13, 67–75. [Google Scholar]
Figure 1. Schematic diagram illustrating the desired implementation of genetic modification traits toward higher production of microalgae biomass.
Figure 1. Schematic diagram illustrating the desired implementation of genetic modification traits toward higher production of microalgae biomass.
Fermentation 08 00466 g001
Figure 2. Scheme of the natural processes in microalgae genetic modifications.
Figure 2. Scheme of the natural processes in microalgae genetic modifications.
Fermentation 08 00466 g002
Figure 3. Impact of microalgae-derived products and the Coronavirus (COVID-19) pandemic.
Figure 3. Impact of microalgae-derived products and the Coronavirus (COVID-19) pandemic.
Fermentation 08 00466 g003
Table 1. Transformations of some microalgae species.
Table 1. Transformations of some microalgae species.
Transformation MethodsMicroalgae SpeciesRefs.
PEG mediatedPleurochrysis carterae[85]
Dunaliella salina[86]
Enzyme-MediatedChlorella vulgaris[87]
Silicon carbide fibersAmphidinium sp.[88]
Chlamydomonas reinhardtii[89]
ElectroporationChlamydomonas reinhardtii[90]
Chlorella sp.[75]
Nannochloropsis oculata[78]
Glass beadsDunaliella salina[73]
Chlamydomonas reinhardtii[91]
Platymonas subcordiformis[92]
Agrobacterium-mediatedChlorella vulgaris[76]
Haematococcus pluvialis[93]
Nannochloropsis sp.[74]
Microprojectile bombardmentChlorella zofingiensis[77]
Phaeodactylum tricornutum[94]
Chaetoceros sp.[95]
Table 2. Comparison of the two types of microalgae cultivation systems [108].
Table 2. Comparison of the two types of microalgae cultivation systems [108].
PropertiesOpen Pond CultivationClosed Pond Cultivation
Required spaceHighLow
Water lossVery highLow
Biomass qualityVariableVariable and able to increase
Biomass concentration (g L−1)Low (0.1 and 0.5)High (0.5 and 8)
Efficiency of light utilizationLowExcellent
TemperatureVariableControlled
CO2 lossHigh (depend on pond depth)Low (controlled)
ContaminationHighLow
CleaningNoneRequired
Process controlLimitedPossible
Weather dependenceHighLow
Start-up1.5–2 Months0.5–1 Month
Capital expensesHighhigher
Operating costsLowHigh
Efficiency of harvestingLowHigh
Table 3. Effects of different radiations on microalgae biomass and carbohydrate production, with metabolite improvement percentages.
Table 3. Effects of different radiations on microalgae biomass and carbohydrate production, with metabolite improvement percentages.
Irradiance Stress (μmol m−2 s−1)Microalgal StrainBiomass Production (mg L−1)Carbohydrate Productivity (mg L−1 Day−1)Metabolite Improvement (%)Refs.
890S. obtusiusculus97828031[124]
300C. sorokiniana2800170-[125]
650Tetradesmus obliquus170080030[126]
310Desmodesmus sp.2380-13.4[127]
2000B. braunii1300900-[128]
Table 4. Effect of nutrient limitations on microalgae biomass productivity and metabolite improvement.
Table 4. Effect of nutrient limitations on microalgae biomass productivity and metabolite improvement.
Nutrient Limitation/StarvationMicroalgal StrainCarbohydrate Productions
(mg L−1)
Metabolite Improvements (%)Biomass Productions (mg L−1)Refs.
NitrogenA. platensis4.39.36192[134]
NitrogenM. aeruginosa-202.25 × 107 cell mL−1[135]
PhosphorusA. platensis6.3159.7195[134]
SulfurC. reinhardtii507051-[136]
vitamins, N, P, and metalTetraselmis sp.420 mg g−11305720[137]
Calcium and magnesiumC. sorokiniana45050-[135]
Multiple nutrientsDesmodesmus sp.400 mg g−1641950[138]
Table 5. Effect of different carbon sources on the biomass production and metabolite content of microalgae.
Table 5. Effect of different carbon sources on the biomass production and metabolite content of microalgae.
Stress ConditionsMicroalgal
Strain
Biomass
Production (mg L−1)
Stress on Carbohydrate Content (%)Stress on Protein
Content (%)
Refs.
BeforeAfterBeforeAfter
0.9 g L−1 NaHCO3Scenedesmus sp.28.3218.5314749.5[140]
5% PentoseC. minutissima6032.558.515.514[129]
5% CO2A. quadricellulare900 mg L−131712014[141]
5% CO2C. sorokiniana960 mg L−130.2532410[141]
Table 6. Nutrient content of several aquaculture wastewaters *.
Table 6. Nutrient content of several aquaculture wastewaters *.
Animal TypeTN
(mg L−1)
NH3-N
(mg L−1)
TP
(mg L−1)
COD
(mg L−1)
TS
(g L−1)
Refs.
Shrimp36190NA1321NA[146]
Shrimp>36584NA1593NA[147]
Shrimp>395102NA120113[148]
Rainbow trout1.20.270.217.60.01[149]
Crucian carp61>0.7NANA[150]
nd1110.07NA20NA[151]
nd7785038434920[150]
* TN: total nitrogen, NH3-N: ammonia-nitrogen; TP: total phosphorus, COD: chemical oxygen demand, and TS: total solids. NA: not detected.
Table 7. Biomass productivity of microalgal species cultured in wastewater.
Table 7. Biomass productivity of microalgal species cultured in wastewater.
Microalgae SpeciesWastewater TypeBiomass Productivity
(mg dw L−1 d−1)
Refs.
B. brauniiMunicipal wastewater345.6[155,156]
S. obliquusMunicipal wastewater26[155,156]
P. carteraeIndustrial wastewater33[155,156]
B. brauniiIndustrial wastewater34[155,156]
C. pyrenoidosaIndustrial wastewater8.114[155,156]
C. saccharophila Industrial wastewater23[155,156]
D. tertiolectaIndustrial wastewater28[155,156]
Chlorella sp.Industrial wastewater0.00005[157]
C. vulgarisIndustrial wastewater0.0019[158]
Chlorella sp.Agricultural wastewater81.4[155,156]
Neochloris sp.Industrial wastewater0.109[159]
C. minutusSynthetic wastewater2.04[160]
G. pleurocapsoidesSynthetic wastewater3.34[160]
Table 8. Microalgae enhancements as a result of co-cultivation with bacteria.
Table 8. Microalgae enhancements as a result of co-cultivation with bacteria.
Microalgal StrainBacterium StrainInteractions
(Microalgae Enhancement)
Refs.
C. sorokinianabrasilense11% increase in cell density[165]
A. protothecoidesA. brasilense90% increase in cell density[165]
I. galbanaAlteromonas sp.52% increase in cell density[166]
I. galbanaLabrenzia sp.71% increase in cell density[166]
P. tricornutumStappia sp.72%, 144%, and 172% increase in growth, chlorophylls, and pigment, respectively[166]
C. vulgarisS. smaltophilia18%, 20%, and 22% increase in productivity, growth rate, and biomass, respectively[167]
T. striataP. bermudensis200% increase in biomass productivity[168]
C. minutissimaEscherichia coli700% increase in biomass productivity[169]
T. pseudonanaD. shibae35% increase in cell density[170]
Ankistrodesmus sp.Rhizobium sp.29% increase in dry weight[171]
Dunaliella sp.Muricauda sp7% increase in cell biovolume[172]
C. sorokinianabrasilense40% and 35% increase in cell density and growth rate, respectively[173]
C. vulgarisbrasilense16% and 11% increase in cell density and growth rate, respectively[173]
B. brauniiRhizobium sp.55% increase in optical density[174]
Table 9. Advantages and disadvantages of the most common microalgae biomass-drying technologies [55].
Table 9. Advantages and disadvantages of the most common microalgae biomass-drying technologies [55].
MethodAdvantagesDisadvantages
SolarSustainable and no energy consumptionDependence on the weather
SprayFast and economical method, suitable for algae production for human consumptionDegradation in the quality, operational cost
FreezeHighly energy intensiveApplicable for small scale
Oven dryingLess energy intensiveSuitable for small scale
Crossflow air-dryingEconomical and fast dryingEnergy cost
IncineratorAlgal biomass burning can be avoidedHigh cost and complicated
Table 10. Techniques for microalgae cell extract and cell rupture mechanisms.
Table 10. Techniques for microalgae cell extract and cell rupture mechanisms.
CategoryTechniqueMechanismRefs.
MechanicalHigh-pressure homogenizationWhen cells are pushed to flow via a narrow valve under tremendous pressure, they break.[203]
Rotor–stator homogenizationCells are disrupted by the shearing between a fixed outer stator and a rapid-spinning inner rotor when they are drawn into a long shaft[204]
Bead milling Contact of cells with agitated beads crushes them.[205]
Grinding with mortar and pestleCells are crushed when they are cut between two hard surfaces, such as a stationary mortar and a rotating pestle. Before the grinding procedure, the sample is frozen by submerging it in liquid nitrogen.[206]
UltrasonicationThe cavitation, which is created by high-frequency sound waves, causes the cells to explode[207]
Hydrodynamic cavitationCells break down by the cavitation caused by a rapid shift in pressure[208]
Screw expeller pressingWhen the dried cells are pressed through a barrel-like chamber, a large volume of them is crushed. Direct oil extraction is possible with this technique.[209]
PhysicalRepeated freeze-thawProduction of intracellular ice crystals during the freezing process and cell expansion during the thawing process both cause cell disruption.[210]
Osmotic shockWhen fluid comes into cells rapidly during a sudden osmotic transition, internal pressure builds up in the cells, causing the cells to burst.[211]
Explosive decompressionGas bubbles escape out from the cells in point holes when pressure is released suddenly[212]
Pulsed electric field (PEF)The effect of electroporation, caused by strong electric fields, causes cells to lysis[177]
MicrowaveThe intrusion of water vapor within the cells causes the cells to be disturbed. Microwaves create a fast fluctuating electric field that leads to the development of heat due to the frictional forces by inert-molecular movement.[213]
Thermolysis (autoclave) When cells are heated to 121 °C for 30 min, they are lysed. Heat is transferred from the outside to the inside of the cells via the cell membrane.[214]
ChemicalAlkalis or acidsThe cell membrane is solubilized through saponification with an alkali or acid[215]
Detergent(surfactant)Detergent chemicals cause cellular disruption by dissolving cell membrane proteins.[216]
Enzyme EThrough the process of enzymatic hydrolysis, the enzyme digests the cell wall.[217]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ashour, M.; Omran, A.M.M. Recent Advances in Marine Microalgae Production: Highlighting Human Health Products from Microalgae in View of the Coronavirus Pandemic (COVID-19). Fermentation 2022, 8, 466. https://doi.org/10.3390/fermentation8090466

AMA Style

Ashour M, Omran AMM. Recent Advances in Marine Microalgae Production: Highlighting Human Health Products from Microalgae in View of the Coronavirus Pandemic (COVID-19). Fermentation. 2022; 8(9):466. https://doi.org/10.3390/fermentation8090466

Chicago/Turabian Style

Ashour, Mohamed, and Ashraf M. M. Omran. 2022. "Recent Advances in Marine Microalgae Production: Highlighting Human Health Products from Microalgae in View of the Coronavirus Pandemic (COVID-19)" Fermentation 8, no. 9: 466. https://doi.org/10.3390/fermentation8090466

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