Next Article in Journal
Development of a New Route for the Immobilization of Unmodified Single-Stranded DNA on Chitosan Beads and Detection of Released Guanine after Hydrolysis
Next Article in Special Issue
Relationship between Protein Digestibility and the Proteolysis of Legume Proteins during Seed Germination
Previous Article in Journal
Study of the Counter Cation Effects on the Supramolecular Structure and Electronic Properties of a Dianionic Oxamate-Based {NiII2} Helicate
Previous Article in Special Issue
A Review on the Effect of Calcium Sequestering Salts on Casein Micelles: From Model Milk Protein Systems to Processed Cheese
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Extraction of Valuable Biomolecules from the Microalga Haematococcus pluvialis Assisted by Electrotechnologies

1
Université de technologie de Compiègne, UTC/ESCOM, TIMR (Transformations Intégrées de la Matière Renouvelable), 60200 Compiègne, France
2
CentraleSupélec, Laboratoire de Génie des Procédés et Matériaux, Centre Européen de Biotechnologie et de Bioéconomie (CEBB), Université Paris-Saclay, 3 Rue des Rouges Terres, 51110 Pomacle, France
3
Laboratory of Physical Chemistry of Disperse Minerals, F. D. Ovcharenko Institute of Biocolloidal Chemistry, NAS of Ukraine, 03142 Kyiv, Ukraine
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(5), 2089; https://doi.org/10.3390/molecules28052089
Submission received: 30 January 2023 / Revised: 20 February 2023 / Accepted: 21 February 2023 / Published: 23 February 2023
(This article belongs to the Special Issue Featured Review Papers in Food Chemistry)

Abstract

:
The freshwater microalga Haematococcus pluvialis is well known as the cell factory for natural astaxanthin, which composes up to 4–7% of its total dry weight. The bioaccumulation of astaxanthin in H. pluvialis cysts seems to be a very complex process that depends on different stress conditions during its cultivation. The red cysts of H. pluvialis develop thick and rigid cell walls under stress growing conditions. Thus, the biomolecule extraction requires general cell disruption technologies to reach a high recovery rate. This short review provides an analysis of the different steps in H. pluvialis’s up and downstream processing including cultivation and harvesting of biomass, cell disruption, extraction and purification techniques. Useful information on the structure of H. pluvialis’s cells, biomolecular composition and properties and the bioactivity of astaxanthin is collected. Special emphasis is given to the recent progress in application of different electrotechnologies during the growth stages and for assistance of the recovery of different biomolecules from H. pluvialis.

1. Introduction

In recent years, the up and downstream processing problem of H. pluvialis microalgae has attracted great attention [1,2,3,4,5]. The low-cost bio-production of natural astaxanthin using this microalga is still one of the market’s most concerning issues. It remains a challenge to biosynthesize natural astaxanthin using microalgae such as H. pluvialis on a large scale. The main reason for the high cost of natural astaxanthin is that the recovery of the final product necessitates additional manufacturing steps.
The cultivation of H. pluvialis for natural astaxanthin production (known as a king of antioxidants) involves two phases, green and red. During the second one, the red cysts or aplanospores accumulate a high amount of astaxanthin, which accounts for ≈80% of the total carotenoid content. Accumulation of astaxanthin is a rather complex process that typically includes specific abiotic stress applications such as nutrient starvation [6], pH [7], illumination stresses [8,9,10], temperature [11], irradiation or pressure stresses [12] and a mixture of different stresses [13]. Particularly, the synergetic combination of different factors allowed improving of astaxanthin accumulation [13,14].
Application of electrotreatments during the growth stages allows genetic transformation, inactivation of culture contaminants, improvement of growth kinetics, accumulation of bioactives and increasing of permeability of cell walls [15,16,17]. The cells of H. pluvialis are covered by extraordinarily thick, rigid and indigestible cell walls, and the availability of internal biomolecules is very limited [18,19]. Therefore, the selective recovery of these biomolecules, and particularly astaxanthin, from H. pluvialis requires application of rather complex cell disruption and purification techniques. The topic is a very hot issue in contemporary studies of extraction from various microalgae [19]. The processing of H. pluvialis using pulsed electrotechnologies (pulsed electric fields and (or) high-voltage electric discharges) is very attractive because with proper adaptation these techniques allow effective extraction and obtaining of high purity extracts [20]. However, the thick cell walls of H. pluvialis cysts can restrict the extraction efficiency and applications of pulsed electrotechnologies. Therefore, careful adaptation and optimization of such treatments is required [21]. Nowadays, the existing literature on the application of pulsed electrotechnologies for H. pluvialis is very limited.
In this review, the analysis of different steps of up and downstream processing in H. pluvialis biorefinery including cultivation and harvesting of biomass, cell disruption, extraction and purification techniques is presented. The useful information on the structure of H. pluvialis cells, biomolecular composition, properties and bioactivity of astaxanthin is collected. Special emphasis on the recent progress in the application of different pulsed electrotechnologies for the recovery of biomolecules from H. pluvialis is given.

2. Main Steps Haematococcus pluvialis Biorefinery

At both laboratory and industrial scales, many important steps are included in the processing biorefinery of H. pluvialis microalgae. The general overviews of these steps are presented in Figure 1.
Each of these steps may be crucial in determination of the quantity and quality of extracted bioactive molecules. In recent years, the different methods of H. pluvialis processing optimization are intensively discussed [1,2,3]. These methods include optimization of favorable and unfavorable growing conditions on cultivation stages and screening of optimized techniques for harvesting of biomass, cell disruption, extraction and purification processes.

3. Cultivation and Growth Stages of Haematococcus pluvialis, Effects of Electrotreatment

The freshwater microalga H. pluvialis was described for the first time in 1841 [22]. It has been reported that this microscopic object can occur in green and red forms. The green forms were observed in juvenile states. In cold seasons (winter, spring, and autumn), the green forms were mainly observed [23,24]. A detailed discussion of the early studies of H. pluvialis was presented in 1899 [25]. H. pluvialis occurs in different regions of Europe, Africa, and North America [26].

3.1. Cultivation and Growth Stages

It is still acknowledged that the green microalga H. pluvialis is recognized for its capacity to produce significant quantities of natural astaxanthin via the application of different stress conditions and its high specific growth rate 1.3 d−1 [27].
H. pluvialis astaxanthin production for the market is relatively new. The induction of astaxanthin bioaccumulation in H. pluvialis cells is carried out by subjecting the cells to stress growth conditions. In the last two decades, the two-stage strategy for cultivation has been widely adopted for both laboratory and large-scale cultures. The first stage involves growing H. pluvialis under favorable conditions for green biomass production (green stage); this is followed by a second stage (red stage) in which the astaxanthin biosynthesis is induced by changing the favorable growth conditions to unfavorable ones. The natural response to the harsh environmental conditions involves the inhibition of normal growth and the accumulation of astaxanthin for the purpose of protecting the photosynthetic apparatus from photo-oxidative damage by absorbing the excess light. The role of astaxanthin is important to protect cells by scavenging free radicals and reacting with reactive oxygen species. This role is to ensure the maintenance of photosynthesis during red-stage stress [28].
Obviously, the life cycle of H. pluvialis microalgae is complicated. The two-stage culturing implies different cell forms. This life cycle can be roughly divided into motile (bi-flagellated cells) and non-motile stages (aplanospores, astaxanthin-accumulating aplanospores, red cysts cells, sporangium and zoospores released from the sporangium) [19,29,30,31]. Each transition from the motile to the non-motile cell form involves remarkable changes in the cell wall structure.
Photosynthetic activity is highest in the initial vegetative green growth phase (days 1 to 12) [7]. In this green vegetative phase, the cells do not accumulate astaxanthin and their carotenoid pattern is mainly composed of lutein (75–80%) and β-carotene (10–20%) [32]. However, the astaxanthin accumulation is triggered in the second phase by specific stress application such as nitrogen starvation, oxidative stress, salinity stress, high temperature and high irradiation. These cited factors significantly affect the astaxanthin accumulation.
The recent literature on cultivation and growth conditions of H. pluvialis is rather vast. It includes discussions of the effects of temperature [11], pH [7], nutrient concentration (nitrogen, phosphorus and carbon starvation) [6,33] and illumination conditions [8,9,10,34,35,36]. The effects of light illumination on the growth of biomass and accumulation of bioastaxanthin were analyzed, and it was demonstrated that red light is more effective for increasing the biomass, whereas blue light supports the accumulation of bioastaxanthin [10]. A comprehensive examination of the effects of different photo-, mixo- and heterotrophic cultivation conditions on H. pluvialis cell germination has been recently presented [37].
The effects of temperature (20–30.5 °C) on the cell growth and accumulation of astaxanthin has been discussed [11]. It was demonstrated that increased temperature allows enhanced astaxanthin accumulation combined with nitrogen starvation stress. It was demonstrated that the photosynthetic properties are differentially modulated in response to nitrogen starvation/high-light-illumination stress and a correct balance between these stresses is required for efficient astaxanthin production [38]. Effects of nuclear irradiation and a high concentration of carbon dioxide CO2 on enhancing the growth rate and astaxanthin yield were demonstrated [12]. Advanced methods for the genetic engineering of H. pluvialis were discussed in [39].
The economical two-stage strategy with low light illumination in the initial phase (growth for biomass production) and a combination of high light illumination and elevated carbon dioxide levels (5 or 15%) in the second phase (astaxanthin accumulation) has been tested [8]. The applied procedures showed a significant increase (2–3 times) in the accumulation of astaxanthin. A two-step process was proposed to maximize algal growth and astaxanthin yield [13]. During the first step (biomass production), the different nitrogen sources were tested. During the second step (carotenogenesis induction), a mix of moderate stressors (mild light, nitrogen limitation, and the addition of sodium acetate) was used. The synergetic combination of different factors allowed the promotion of astaxanthin accumulation. The new sequential heterotrophy–dilution–photoinduction cultivation strategy was shown to be rather effective for production of astaxanthin from H. pluvialis [14]. The effects of continuous and interrupted (light/dark cycles) illumination conditions on the production of vegetative green cells and astaxanthin accumulation were investigated [40]. It was demonstrated that light/dark cycles are optimal to produce green cells, whereas continuous illumination is better for astaxanthin accumulation. The effects of fulvic acid on biomass growth and astaxanthin accumulation were investigated [41]. It was demonstrated that the astaxanthin content increased by 86.89% in 5 mg/L-treated cells.
Recently, the details of laboratory and commercial scale cultivation procedures in the production of natural astaxanthin derived from H. pluvialis have been discussed [42,43,44]. The effects of different cultivation conditions and strategies, technological innovations and types of cultivation systems and an economic assessment for astaxanthin production were also reviewed [45,46,47,48,49,50,51,52].

3.2. Electrotreatments during the Growth Stages

The effects of electrotreatments applied during the growth of H. pluvialis on cell viability, cell number density and accumulation of astaxanthin attract great attention. A general review of the application of different electrotechnologies for genetic transformation, inactivation of culture contaminants and improvement of growth kinetics of different species of microalgae is given in [15]. Application of electrotreatments of H. pluvialis during the different growing stages may reduce the strength and impermeability of cell walls and positively affect the accumulation of bioactives.
The effects of pulsed electric field (PEF) treatments with square wave and exponential decay wave pulses on cell viability during the growth stages were discussed in [53]. Nanosecond PEF treatment (with very short pulses of 25–50 nm and high electric field strength 40 kV/cm) has been applied to H. pluvialis cells during the growth stage [16]. It was demonstrated that such treatment can modulate the expression of the astaxanthin biosynthesis genes. PEF treatment increased both the cell mortality (by up to 20%) and astaxanthin accumulation (by 20–30%) from the initial values. Such results were explained by the expression of carotenoid-pathway-related genes.
Electrical treatment of H. pluvialis cells was studied in the bio-electrochemical chambers [54,55]. Such treatments enhanced the nitrogen consumption and chlorophyll biosynthesis and resulted in a considerable increase in electro-treated cell number density (by 20%) after cultivation for seven days [54]. However, such treatments did not significantly affect the intracellular astaxanthin content (10% increase). The special adaptation of electrotreatment parameters (electric strength and duration) allowed a significant increase in the astaxanthin content in electrostimulated H. pluvialis (by 36.9%) [55].
Application of electrical treatment by low-temperature plasma in H. pluvialis biotechnology has been also discussed. It was demonstrated that low-temperature plasma promotes the growth of H. pluvialis [56]. The plasma treatment is widely used as a safe and environmentally friendly sterilization technology [57] and may be applied during the growth stages as a strong oxidative stress and mutagenesis tool without damaging the microalgae cells [17]. Atmospheric pressure argon dielectric barrier discharge plasma was used for mutation induction in H. pluvialis [58]. The presence of enhanced astaxanthin accumulation in the mutants was observed that was attributed to genetic modification by the mechanism of carotenogenesis. It was indicated that plasma mutation is suitable for effective H. pluvialis breeding with promising industrial applications [59]. The plasma treatment effectively improved the growth of H. pluvialis and its astaxanthin accumulation; under optimal conditions the astaxanthin content was 51.96% higher than in the starting strain. The possible mechanisms of astaxanthin accumulation induced by plasma treatment were also discussed [60,61].

4. Structure of Cell and Biomolecular Composition of Haematococcus pluvialis

4.1. Cell Structure

Figure 2 gives a schematic presentation of the cell structure of H. pluvialis at the resistant cyst stage. On the final aplanospore stage of cyst formation, the rigid and resistant multilayered cell walls are formed [62,63,64,65,66,67]. The total size of the cells at this stage is approximately 20–30 µm [29]. The cell internal compartment contains cytoplasm starch granules, sub-cellular micro-compartments of pyrenoids and astaxanthin deposited in extra-plastidial oil globules (lipid droplets). This internal space is covered by lipid membrane with thickness of ≈5 nm.
The outside part of the cell is covered with an aplanospore cell wall (with total thinness of 2–3 µm). This wall is composed of an outer primary algaenan-based wall (a very stable sporopollenin-like wall), a secondary wall (composed of mannose and cellulose with homogeneous arrangement) and a tertiary wall (composed of mannose and cellulose with heterogeneous arrangement). Therefore, the cell is protected by complex, strong and rigid cell walls.
The three layered cell structure acts as a strong barrier to different external physical and chemical stresses and provides a high resistance against extreme environmental stress conditions [18,19]. Consequently, the cell wall developed is indigestible and reduces the bio-accessibility of the internal biomolecules such as astaxanthin. Moreover, the thickness of the H. pluvialis cell wall that is reached after the stress culturing stage is considered as an important bottleneck limiting the large-scale production of natural astaxanthin.

4.2. Biomolecular Composition

The chemical constituents of extracellular extracts of H. pluvialis were recently analyzed [68]. H. pluvialis cell composition includes ash, protein, carbohydrates, dietary fiber, carotenoids and lipids.
Figure 3 presents the relative percentage proportions of ash, protein, carbohydrates total dietary fiber, carotenoids and lipids in the initial biomass and extracts obtained using supercritical CO2 extraction at optimized conditions (pressure of 350 bar, temperature of 50 °C, and CO2 flow rate of 0.47 Kg/min) for the H. pluvialis red phase [69]. The total dietary fiber and protein were observed as the main intracellular constituents in the initial biomass. Carotenoids include astaxanthin (66–70%), β-carotene (3.5–6.5%) and lutein (27–28%). The content of proteins, lipids and carotenoids in extracts obtained by supercritical CO2 was higher in comparison to that of the biomass. An inverse situation was observed for the content of total dietary fibers (58% in the biomass compared with 26% in the extract).

4.2.1. Lipids, Proteins and Carbohydrates

The content of lipids, proteins and carbohydrates in the H. pluvialis cells strongly depends on the stage of cell evolution [29,43,70]. Lipid content in the H. pluvialis cells may be rather high during the green vegetative stage (≈20–25% of the total biomass). During the red-stage evolution, lipid content increases (≈32–37%) and can be manipulated by operation conditions and the stress imposed. The protein and carbohydrate contents also change with transition from the green phase to the red phase. The protein content is ≈29–45% in the green stage and it reduces up to 17–25% in the red stage. The carbohydrate content is ≈15–17% in the green stage and it increases up to 36–40% in the red stage.

4.2.2. Carotenoids

During the green stage, the other carotenoids in H. pluvialis are lutein (≈50–70% of total carotenoids), neoxanthin, β-carotene and violaxanthin (≈8–13% for each of them); there is no astaxanthin content [43,71]. During the red stage, the total carotenoid content is significantly increased (mainly due to the astaxanthin accumulation) and can reach more than 4% of the biomass dry weight. At this stage, astaxanthin is the main cell carotenoid, and the other carotenoids represent ≈5–20% of the total content.
Astaxanthin (the chemical IUPAC name is 3,3′-dihydroxy-β,β-carotene-4,4′-dione) is a liposoluble red dietary carotenoid. It has the molecular formula C40H52O4, a molecular weight of 596.86 g/mol, a boiling point of 774.04 °C, a density of 1.071 g/cm3 and a melting point of 182–183 °C. Astaxanthin contains 13 conjugated double bonds with two benzene rings and hydroxyl groups at the ends [72]. Due to the presence of hydroxyl groups, the molecule of astaxanthin is polar. At room temperature (25 °C), astaxanthin is practically insoluble in water, but it is highly soluble in different organic solvents (Table 1).
It was also demonstrated that the solubility of astaxanthin in acidic conditions was 10–20 times higher than in neutral and basic conditions [76].
Astaxanthin is rather unstable and can be degraded during the production process due to heating (even for a very short time), long-term storage, in acidic conditions under ultraviolet irradiation and in the presence of oxygen [77]. The effect of extraction and drying methods on the antioxidant activity of astaxanthin was discussed in [78]. The general overview of different chemical and biochemical properties of astaxanthin is presented in [43].
The microalga H. pluvialis can accumulate the optical pure isomer of astaxanthin (3S, 3′S), which has high antioxidant activity (in neutralization of destructive reactive oxygen species) that is approximately 20 times stronger the synthetic astaxanthin [79]. Several studies have shown that astaxanthin accumulated within the cysts of H. pluvialis exists in an esterified form, mainly as monoesters (70%), di-esters (25%) and free form (5%) with other occurring carotenoids [48]. The astaxanthin can be esterified on one or both of its hydroxyl groups with various fatty acids such as palmitic, stearic and oleic acids [80].
For the commercialization of natural astaxanthin, the free form is only approved for the market. This means the purification of final astaxanthin product of H. pluvialis needs to pass the additional step of de-esterification (known as saponification). Two methods exist for the de-esterification of extracted astaxanthin: de-esterification through enzymolysis using a cholesterol esterase and de-esterification through saponification in darkness at low or room temperature using an alkaline solution of NaOH or KOH in methanol. However, this step of de-esterification could induce a loss of carotenoids due to their degradation. Three factors could negatively affect the final pigments: treatment time, NaOH or KOH concentration and temperature [81].

5. Harvesting of Biomass

The harvesting procedure of H. pluvialis is a very important step before application of cell disruption technique. It consists of separation of the biomass from the bulk culture medium, which can be achieved using gravity sedimentation, centrifugation, filtration, flotation or flocculation techniques [82,83]. Gravity sedimentation or centrifugation are rather effective for the large cells of H. pluvialis and can be successfully employed. The filtration technique is based on the liquid biomass flowing through the membrane. Flotation is a gravity separation technique, and it involves attaching micro-sized air bubbles to microalgae cells. The flocculation technique is based on the aggregation of H. pluvialis cells caused by the addition of flocculants. A combination of these techniques can be useful to achieve a good harvest efficiency for H. pluvialis [84].

6. Cell Disruption, Extraction, Purification Techniques

The industrial perspective of the downstream processing of H. pluvialis and different steps for the optimization of bioactive recovery are widely discussed in the current literature [85,86].

6.1. Cell Disruption

Cell disruption can use different physical, chemical and biological techniques (Figure 1). Physical treatments including mechanical, thermal and different innovative methods are widely applied for cell disruption and separation of the value-added biocompounds [82,87,88] phycobiliproteins, carotenoids [89] and astaxanthin [19,43,90,91,92,93,94,95]. Traditional mechanical treatments use bead milling, grinding, high-pressure homogenization and hydraulic and screw pressing [96,97]. Generally, mechanical treatments are non-selective, non-toxic and suitable for large-scale production. Mechanical cell disruption requires high energy input. The thermal treatments use conductive, ohmic or radio frequency heating, autoclaving or application of freeze–thaw. These techniques are also non-selective and non-toxic. However, they are highly energy consuming and some of them may cause the thermal degradation of biomolecules. The new innovative methods assisted by electric and magnetic [98] fields, ultrasound [99,100,101] and microwaves [102,103] may be more selective and are non-toxic, but may require more sophisticated equipment and special adaptation for large-scale production.
Chemical treatments use acid, alkalis, detergent, salt, surfactant, nanoparticle, polymer, solvent, ionic liquid, oxidative and osmotic shock treatments [104,105,106,107]. These treatments do not require advanced equipment but are non-selective, toxic, less suitable for large-scale production and may add chemical contaminants to the extracted biomolecules.
Biological treatments use enzymes, viruses and algicidal bacteria [91,108]. They are relatively selective and less toxic compared with chemical techniques, but they are rather expensive, time consuming and less suitable for large-scale production.
All these treatments can be applied individually; however, for enhanced disruption and increased selectivity, combinations of different techniques may be useful.

6.2. Extraction

Once the cell walls are disrupted or weakened, the biomolecules with antioxidant and antimicrobial activity, and particularly astaxanthin, must be rapidly recovered from H. pluvialis microalgae to prevent their degradation. Popular methods of extraction are subcritical water extraction, supercritical fluid extraction and extraction with solvents and oils. Supercritical carbon dioxide (S-CO2) extraction is considered as the most efficient and has been widely used for industrial applications [71,109]. Subcritical water extraction at 200 °C was demonstrated to be a green and fast extraction method to obtain antioxidant and antimicrobial extracts in the red growing phase and permitted obtaining the high extraction yields of vitamin E and simple phenolics (up to 30% of dry weight) [110]. However, astaxanthin destruction is rather significant at high temperatures. In many works, the different modes of S-CO2 extraction at moderate critical temperatures and pressures were tested [111,112,113,114,115,116,117,118,119] (see [120] for a recent review). Supercritical CO2 is non-polar, non-toxic, not flammable and inexpensive. The solubility of astaxanthin in supercritical CO2 is rather high (≈10−2 g/L at 25 °C and 100 bar) [121]. In addition, the CO2 at a supercritical state has a high diffusion coefficient and a low viscosity that allows its rapid penetration into the porous matrix of H. pluvialis. In some works, CO2 modified by addition of ethanol [111,114] and ionic liquids [117] was used to enhance extraction efficiency. The economic assessments of the S-CO2 extraction of astaxanthin from H. pluvialis were analyzed in [119].
In a simple liquid–solvent extraction the biomolecule recovery is commonly assisted by safe solvents (ethanol, ethyl lactate and ethyl acetate) [122,123] with some limitations for the use of more toxic solvents (acetone, toluene, hexane, benzene, petroleum ether, methanol, dimethylaminocyclohexane and methyl-tert-butyl ether) [121].

6.3. Purification

The necessity of a subsequent purification step is mainly defined by the selected cell disruption and extraction methods. Typically, the purification methods are based on the application of electrophoresis, membrane separation (ultra- or electro-membrane filtration), ultracentrifugation or liquid multi-phase flotation (for a review, see, for example, [109]).

7. Processing of Haematococcus pluvialis by Pulsed Electrotechnologies

The modern, innovative cell disruption techniques based on pulsed electrotechnologies (pulsed electric field (PEF), moderate electric field and high-voltage electric discharges (HVED)) have been previously applied to assist the extraction of biomolecules from different types of microalgae [124,125,126,127]. These techniques typically require lower energy input, smaller amounts of solvent and shorter extraction times than traditional methods [20].
PEF treatment is based on the application of short-duration and high-voltage electric pulses. This treatment can cause perforation of the membrane and the formation of nanopores in them (the phenomenon of so-called electroporation). In turn, electroporation facilitates the solvent-mediated extraction of different intracellular biomolecules from the microalgae species. The effects of electroporation significantly depend on the parameters of PEF treatment such as pulse duration ti, number of pulses n, total time of PEF treatment tPEF = nti, electric field strength E, specific energy input W and other details of the PEF protocol. The pulses may be unipolar or bipolar and have an exponential or square-wave shape. The pulse duration ranges from several nanoseconds to several milliseconds, and electric strength ranges in the interval E = 100 V/cm–80 kV/m with application of moderate (100–200 V/cm), intermediate (typically under 1000 V/cm), high (1–20 kV/cm) and extra-high (above 20 kV/cm) electric fields dependent on the type of treated microalgae. At the moderate specific energy inputs (W = 0.5–5 kJ/kg), electroporation may be reversible, and it becomes irreversible at higher specific energy inputs (W = 50–200 kJ/kg). PEF-assisted extraction typically involves moderate specific energies below 20 kJ/kg. At the extra-high electric field strength, high-voltage electric discharges typically occur and are accompanied by different phenomena such as pressure shock waves, bubble cavitation, liquid turbulence, etc. HVED treatment can affect the integrity of the cell wall and enhance the extraction of bioactive compounds.
PEF has been frequently applied in previous studies and there are many recent reviews on applications of PEF for treatment of different biomolecules [87,109,128,129], antioxidants [19,130], pigments and carotenoids [131,132,133,134], phycobiliproteins (fluorescent proteins) [89], proteins and carbohydrates [135], polyphenols [136,137] and lipids [138]. Moreover, the efficiency of different electric-field-based techniques for enhanced microalgae biomolecule extractions (PEF, electrolysis, high-voltage electrical discharges and moderate electric fields) were recently compared [139]. In general, PEF-treatment can be considered as a suitable technique for extraction of the intracellular proteins and pigments from microalgae [140].
However, as far as we know, the data reported in the literature on the application of pulsed electrotechnologies for the recovery of biomolecules from H. pluvialis concerns just the extraction of proteins [141,142] and astaxanthin [92,143]. It was stated that the larger cells of H. pluvialis are more susceptible to electroporation [82]. However, the thick, rigid cell walls of H. pluvialis can lead to inefficient extraction of biomolecules even at very high PEF intensities and numbers of pulses [21].
In study of protein extraction, the H. pluvialis suspension was PEF treated in distilled water at E = 3–6 kV/cm (nine bipolar pulses with duration of 2 ms); this treatment was followed by a 24 h incubation period in salty buffer. Protein recovery was compared for the PEF-treated and untreated samples [141]. PEF treatment significantly increased the protein recovery (by 10–20 times), and the content of extracted proteins significantly depended on the temperature (4 and 20 °C) and the concentration of cells (105 cells/mL or 106 cells/mL). A post-pulse incubation step was found to be necessary to allow oozing. It was assumed that PEF treatment induces some alteration of the cell walls, increasing their porosity to allow leakage of cytoplasmic proteins. In other work, PEF treatment in water at E = 1 kV/cm allowed extraction of 46% of the total quantity of proteins (10.2 µg protein per mL of culture) [142]. The non-destructive electroextraction from H. pluvialis with cultures recovered after 72 h was observed. A biocompatibility index (fraction of surviving cells after PEF treatment) was estimated.
Rather contradictory results on the PEF-assisted astaxanthin extraction from H. pluvialis were obtained [92,143]. For example, a very low astaxanthin extraction efficiency (3.1–6.5%) was observed using PEF pretreatment in methanol (10–30 kV/cm, 30 min) and an additional 2 h incubation [92]. PEF treatment led to the lowest percentage of astaxanthin extractability compared with ultrasounds, high-pressure microfluidization, hydrochloric acid and ionic liquid techniques. For instance, high-pressure microfluidization led to the maximum extractability value of 91.4%. PEF treatment was considered as inefficient for the cell wall disintegration of H. pluvialis. In other work, the PEF treatment was applied to extract astaxanthin from the Nordic microalga H. pluvialis [143]. A high extraction efficiency (96% of the total carotenoid content) was observed using PEF treatment in water (10 pulses of 5 ms at 1 kV/cm and 20 °C). The value of around of E = 1 kV/cm was considered as a threshold for the effective extraction from this Nordic H. pluvialis strain. It was demonstrated that PEF treatment resulted in a 1.2-fold more efficient extraction compared with the classical disruption methods (bead-beating and thermal treatments). It was stated that efficient PEF-assisted extraction of carotenoids requires proper optimization of the incubation time after PEF application. The differences between results obtained in [92,143] may reflect the significant variations in the PEF-assisted extraction protocols and differences in the studied strains [19].

8. Bioactivity of Haematococcus pluvialis Extracts

The bioactivity of biocompounds extracted from H. pluvialis was intensively studied in the current literature [69,144,145,146,147,148]. Many of these studies are related with applications of natural bioastaxanthin. Figure 4 demonstrates that the natural astaxanthin has exceptionally high antioxidant activity compared with other known antioxidants. The biosynthetic (natural) astaxanthin produced by H. pluvialis has superior therapeutic properties and is highly preferable in comparison with astaxanthin produced by chemical synthesis from petrochemical sources in the form of a mixture of isomers [49]. Moreover, synthetic astaxanthin is not recommended for use as a human nutraceutical supplement [149].
The estimated cost of bioastaxanthin production from H. pluvialis is around USD3000–USD3600 kg−1 compared with the cost of production of synthetic astaxanthin, which is around USD1000 kg−1 [93]. Bioastaxanthin demonstrated significantly stronger (by order of magnitudes) bioactivity than astaxanthin synthesized from petrochemicals (see Figure 4).
The recent advancements in bioastaxanthin production from H. pluvialis and the multifunctional applications and world market for it were analyzed in [93,152,153,154]. Natural astaxanthin is more powerful and a stronger antioxidant for the neutralization of destructive reactive oxygen species found in the human body compared with all other antioxidants [26,29,49,77,151]. It is ≈6000 times more active than vitamin C, ≈55 times more active than coenzyme Q10 (CoQ10), 100 times more potent than vitamin E and five times more powerful than β-carotene for the neutralization of destructive reactive oxygen species found in the human body.
Currently, bioastaxanthin is approved by the FDA (United States Food and Drug Administration) as a health food supplement and food colorant for animal and fish feed [155]. Bioastaxanthin has been also approved as a dietary supplement for human consumption in many European countries, the USA and Japan. The recommended or approved doses of astaxanthin range between 2 and 24 mg/day [155].
Bioastaxanthin demonstrates many attractive anti-aging, anti-inflammatory, sun proofing and immune system boosting effects on living organisms. The possibilities and limitations of the use of astaxanthin in food technology were analyzed in [156]. The possible applications of natural astaxanthin also involve the pharmaceutical [144,145,157] and cosmetic [69,147,158] industries. The medical and human health applications [145,159] including anti-cancer, anti-diabetic and anti-atherosclerotic effects and the immune, neuroprotective and genetic engineering potential exerted by astaxanthin have been discussed in [26,145,159,160,161]. The different effects related to the impact of natural astaxanthin on oxidative stresses, aging (skin and brain aging), skin physiology, central nervous system functioning [77] and inflammation-associated diseases [162] were also discussed.

9. Conclusions

The freshwater microalga H. pluvialis has a large content of natural astaxanthin (bioastaxanthin), which comprises up to 4–7% of its total dry weight. Bioastaxanthin is very powerful and strong antioxidant for the neutralization of destructive reactive oxygen species found in the human body. However, the cells of H. pluvialis are covered by thick and rigid cell walls and extraction of intracellular biomolecules is not an easy task. Pulsed electrotechnologies, and particularly pulsed electric field (PEF), are very promising for the extraction of different biomolecules from H. pluvialis. However, the thick, rigid cell walls of H. pluvialis can effectively protect the cell membrane against electroporation and lead to inefficient extraction of biomolecules even under heavy PEF loads (high intensities and big number of pulses). PEF treatment is usually referred to as a mild treatment method that could not significantly affect the structure of cell walls. The PEF treatment of H. pluvialis requires careful optimization of all the PEF treatment parameters, selection of the type of organic solvent and combined application of different cell destruction techniques. Therefore, more careful future studies on the application of PEF in combination with other cell disruption techniques for H. pluvialis are needed.

Author Contributions

All the authors participated in the collection of data from the literature, analysis of data and drafting the manuscript. The final publication was prepared with contributions from all authors. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Nikolai Lebovka acknowledges the partial supports received under the KPKVK 6541230 program.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

PEFpulsed electric fields
HVED high-voltage electric discharges

References

  1. Zarei, Z.; Zamani, H. Biorefinery Potential of Microalga Haematococcus pluvialis to Produce Astaxanthin and Biodiesel under Nitrogen Deprivation. BioEnergy Res. 2022, 1–12. [Google Scholar] [CrossRef]
  2. Khoo, K.S.; Lee, S.Y.; Ooi, C.W.; Fu, X.; Miao, X.; Ling, T.C.; Show, P.L. Recent Advances in Biorefinery of Astaxanthin from Haematococcus pluvialis. Bioresour. Technol. 2019, 288, 121606. [Google Scholar] [CrossRef] [PubMed]
  3. Hosseini, A.; Jazini, M.; Mahdieh, M.; Karimi, K. Efficient Superantioxidant and Biofuel Production from Microalga Haematococcus pluvialis via a Biorefinery Approach. Bioresour. Technol. 2020, 306, 123100. [Google Scholar] [CrossRef]
  4. Papadaki, S.; Kyriakopoulou, K.; Krokida, M. Recovery and Encapsualtion of Bioactive Extracts from Haematococcus pluvialis and Phaedodactylum Tricornutum for Food Applications. IOSR Int. Organ. Sci. Res. J. Environ. Sci. Toxicol. Food Technol. 2017, 10, 53–58. [Google Scholar] [CrossRef]
  5. Nishshanka, G.K.S.H.; Liyanaarachchi, V.C.; Nimarshana, P.H.V.; Ariyadasa, T.U.; Chang, J.-S. Haematococcus pluvialis: A Potential Feedstock for Multiple-Product Biorefining. J. Clean. Prod. 2022, 344, 131103. [Google Scholar] [CrossRef]
  6. Bogoçlua, M.; Keçecilerb, C.; Çokcanb, C.; Özelb, C.; Tekerekc, B.S.; Yücelb, S. The Accumulation of Astaxanthin from Haematococcus pluvialis in Different Stress Factors. J. Indian Chem. Soc. 2019, 96, 1–6. [Google Scholar]
  7. Dos Santos, A.C.; Lombardi, A.T. Growth, Photosynthesis and Biochemical Composition of Haematococcus pluvialis </i> at Various PH. J. Algal Biomass Util. 2017, 8, 1–15. [Google Scholar]
  8. Christian, D.; Zhang, J.; Sawdon, A.J.; Peng, C.-A. Enhanced Astaxanthin Accumulation in Haematococcus pluvialis Using High Carbon Dioxide Concentration and Light Illumination. Bioresour. Technol. 2018, 256, 548–551. [Google Scholar] [CrossRef]
  9. Nishshanka, S.H.; Liyanaarachchi, V.C.; Premaratne, M.; Ariyadasa, T.U.; Nimarshana, V. Sustainable Cultivation of Haematococcus pluvialis for the Production of Natural Astaxanthin. In Proceedings of the Moratuwa Engineering Research Conference (MERCon), Moratuwa, Sri Lanka, 27–29 July 2021; pp. 297–302. [Google Scholar]
  10. Ahirwar, A.; Meignen, G.; Khan, M.J.; Sirotiya, V.; Scarsini, M.; Roux, S.; Marchand, J.; Schoefs, B.; Vinayak, V. Light Modulates Transcriptomic Dynamics Upregulating Astaxanthin Accumulation in Haematococcus: A Review. Bioresour. Technol. 2021, 340, 125707. [Google Scholar] [CrossRef]
  11. Giannelli, L.; Yamada, H.; Katsuda, T.; Yamaji, H. Effects of Temperature on the Astaxanthin Productivity and Light Harvesting Characteristics of the Green Alga Haematococcus pluvialis. J. Biosci. Bioeng. 2015, 119, 345–350. [Google Scholar] [CrossRef]
  12. Cheng, J.; Li, K.; Yang, Z.; Zhou, J.; Cen, K. Enhancing the Growth Rate and Astaxanthin Yield of Haematococcus pluvialis by Nuclear Irradiation and High Concentration of Carbon Dioxide Stress. Bioresour. Technol. 2016, 204, 49–54. [Google Scholar] [CrossRef] [PubMed]
  13. Rizzo, A.; Ross, M.E.; Norici, A.; Jesus, B. A Two-Step Process for Improved Biomass Production and Non-Destructive Astaxanthin and Carotenoids Accumulation in Haematococcus pluvialis. Appl. Sci. 2022, 12, 1261. [Google Scholar] [CrossRef]
  14. Wan, M.; Zhang, Z.; Wang, J.; Huang, J.; Fan, J.; Yu, A.; Wang, W.; Li, Y. Sequential Heterotrophy--Dilution--Photoinduction Cultivation of Haematococcus pluvialis for Efficient Production of Astaxanthin. Bioresour. Technol. 2015, 198, 557–563. [Google Scholar] [CrossRef] [PubMed]
  15. Geada, P.; Rodrigues, R.; Loureiro, L.; Pereira, R.; Fernandes, B.; Teixeira, J.A.; Vasconcelos, V.; Vicente, A.A. Electrotechnologies Applied to Microalgal Biotechnology--Applications, Techniques and Future Trends. Renew. Sustain. Energy Rev. 2018, 94, 656–668. [Google Scholar] [CrossRef] [Green Version]
  16. Bai, F.; Gusbeth, C.; Frey, W.; Nick, P. Nanosecond Pulsed Electric Fields Modulate the Expression of the Astaxanthin Biosynthesis Genes Psy, CrtR-b and Bkt 1 in Haematococcus pluvialis. Sci. Rep. 2020, 10, 1–15. [Google Scholar] [CrossRef]
  17. Nguyen, A.D.; Scarsini, M.; Poncin-Epaillard, F.; Noel, O.; Marchand, J.; Schoefs, B. Plasma Applications in Microalgal Biotechnology. In Agritech: Innovative Agriculture Using Microwaves and Plasmas. Thermal and Non-Thermal Processing; Horikoshi, S., Brodie, G., Takaki, K., Serpone, N., Eds.; Springer: Berlin/Heidelberg, Germany, 2022; pp. 327–349. [Google Scholar]
  18. Li, Q.; Zhao, Y.; Ding, W.; Han, B.; Geng, S.; Ning, D.; Ma, T.; Yu, X. Gamma-Aminobutyric Acid Facilitates the Simultaneous Production of Biomass, Astaxanthin and Lipids in Haematococcus pluvialis under Salinity and High-Light Stress Conditions. Bioresour. Technol. 2021, 320, 124418. [Google Scholar] [CrossRef]
  19. Kim, B.; Lee, S.Y.; Narasimhan, A.L.; Kim, S.; Oh, Y.-K. Cell Disruption and Astaxanthin Extraction from Haematococcus pluvialis: Recent Advances. Bioresour. Technol. 2022, 343, 126124. [Google Scholar] [CrossRef]
  20. Vorobiev, E.; Lebovka, N. Processing of Foods and Biomass Feedstocks by Pulsed Electric Energy; Springer: Cham, Switzerland, 2020; ISBN 978-3-030-40917-3. [Google Scholar]
  21. Safi, C.; Rodriguez, L.C.; Mulder, W.J.; Engelen-Smit, N.; Spekking, W.; den Broek, L.A.M.; Olivieri, G.; Sijtsma, L. Energy Consumption and Water-Soluble Protein Release by Cell Wall Disruption of Nannochloropsis Gaditana. Bioresour. Technol. 2017, 239, 204–210. [Google Scholar] [CrossRef]
  22. von Flotow, J. Beobachtungen Über Haematococcus pluvialis (Observations on Haematococcus pluvialis) Verhandlungen Der Kaiserlichen Leopoldinisch-Carolinischen Deutschen Akademie Der Naturforscher 12(Abt. 2): 413–606, 3 Pls. Verhandlungen der Kais. Leopoldinisch-Carolinischen Dtsch. Akad. der Naturforscher (Proceedings Imp. Leopoldine-Carolinic Ger. Acad. Sci. 1844, 12, 413–606. [Google Scholar]
  23. Kützing, F.T. Über Die Verwandlung Der Infusorien in Niedere Algenformen (On the Transformation of Infusoria into Lower Forms of Algae; Verlag von W. Kohne: Nordhausen, Germany, 1844. [Google Scholar]
  24. Focke, G.W. Physiologische Studien; C. Schünemann: Bremen, Germany, 1847. [Google Scholar]
  25. Hazen, T.E. The Life History of Sphaerella lacustris (Haematococcus pluvialis). Mem. Torrey Bot. Club 1899, 6, 211–246. [Google Scholar]
  26. Rather, A.H.; Singh, S.; Rao, R.; Choudhary, S. Haematococcus pluvialis Astaxanthin and Its Applications. In Advances in Biotechnology and Bioscience; Kumar, S., Ed.; AkiNik Publications: New Delhi, India, 2020; Volume 4, pp. 21–35. [Google Scholar]
  27. Boussiba, S. Carotenogenesis in the Green Alga Haematococcus pluvialis: Cellular Physiology and Stress Response. Physiol. Plant. 2000, 108, 111–117. [Google Scholar] [CrossRef]
  28. Grujić, V.J.; Todorović, B.; Kranvogl, R.; Ciringer, T.; Ambrožič-Dolinšek, J. Diversity and Content of Carotenoids and Other Pigments in the Transition from the Green to the Red Stage of Haematococcus pluvialis Microalgae Identified by HPLC-DAD and LC-QTOF-MS. Plants 2022, 11, 1026. [Google Scholar] [CrossRef] [PubMed]
  29. Shah, M.M.R.; Liang, Y.; Cheng, J.J.; Daroch, M. Astaxanthin-Producing Green Microalga Haematococcus pluvialis: From Single Cell to High Value Commercial Products. Front. Plant Sci. 2016, 7, 531. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Bauer, A.; Minceva, M. Techno-Economic Analysis of a New Downstream Process for the Production of Astaxanthin from the Microalgae Haematococcus pluvialis. Bioresour. Bioprocess. 2021, 8, 1–18. [Google Scholar] [CrossRef]
  31. Zhang, C.; Liu, J.; Zhang, L. Cell Cycles and Proliferation Patterns in Haematococcus pluvialis. Chin. J. Oceanol. Limnol. 2017, 35, 1205–1211. [Google Scholar] [CrossRef]
  32. Samorì, C.; Pezzolesi, L.; Galletti, P.; Semeraro, M.; Tagliavini, E. Extraction and Milking of Astaxanthin from Haematococcus pluvialis Cultures. Green Chem. 2019, 21, 3621–3628. [Google Scholar] [CrossRef]
  33. Nunes, M.; Vieira, A.A.H.; Pinto, E.; Carneiro, R.L.; Monteiro, A.C. Carotenogenesis in Haematococcus pluvialis Cells Induced by Light and Nutrient Stresses. Pesqui. Agropecuária Bras. 2013, 48, 825–832. [Google Scholar] [CrossRef]
  34. Xi, T.; Kim, D.G.; Roh, S.W.; Choi, J.-S.; Choi, Y.-E. Enhancement of Astaxanthin Production Using Haematococcus pluvialis with Novel LED Wavelength Shift Strategy. Appl. Microbiol. Biotechnol. 2016, 100, 6231–6238. [Google Scholar] [CrossRef]
  35. Niño-Castillo, C.M.; Rodríguez-Rivera, F.C.; Díaz, L.E.; Lancheros-Díaz, A.G. Evaluation of Cell Growth Conditions for the Astaxanthin Production as of Haematococcus pluvialis Microalgae. Nova 2017, 15, 19–31. [Google Scholar]
  36. Pham, K.T.; Nguyen, T.C.; Luong, T.H.; Dang, P.H.; Vu, D.C.; Do, T.N.; Phi, T.C.M.; Nguyen, D.B. Influence of Inoculum Size, CO2 Concentration and LEDs on the Growth of Green Microalgae Haematococcus Pluvialis Flotow. Vietnam J. Sci. Technol. Eng. 2018, 60, 59–65. [Google Scholar] [CrossRef] [Green Version]
  37. Bauer, A.; Minceva, M. Examination of Photo-, Mixo-, and Heterotrophic Cultivation Conditions on Haematococcus pluvialis Cyst Cell Germination. Appl. Sci. 2021, 11, 7201. [Google Scholar] [CrossRef]
  38. Scibilia, L.; Girolomoni, L.; Berteotti, S.; Alboresi, A.; Ballottari, M. Photosynthetic Response to Nitrogen Starvation and High Light in Haematococcus pluvialis. Algal Res. 2015, 12, 170–181. [Google Scholar] [CrossRef]
  39. Sharon-Gojman, R.; Maimon, E.; Leu, S.; Zarka, A.; Boussiba, S. Advanced Methods for Genetic Engineering of Haematococcus Pluvialis (Chlorophyceae, Volvocales). Algal Res. 2015, 10, 8–15. [Google Scholar] [CrossRef]
  40. Domínguez, A.; Pereira, S.; Otero, A. Does Haematococcus pluvialis Need to Sleep? Algal Res. 2019, 44, 101722. [Google Scholar] [CrossRef]
  41. Zhao, Y.; Shang, M.; Xu, J.-W.; Zhao, P.; Li, T.; Yu, X. Enhanced Astaxanthin Production from a Novel Strain of Haematococcus pluvialis Using Fulvic Acid. Process Biochem. 2015, 50, 2072–2077. [Google Scholar] [CrossRef]
  42. Panis, G.; Carreon, J.R. Commercial Astaxanthin Production Derived by Green Alga Haematococcus pluvialis: A Microalgae Process Model and a Techno-Economic Assessment All through Production Line. Algal Res. 2016, 18, 175–190. [Google Scholar] [CrossRef] [Green Version]
  43. Jannel, S.; Caro, Y.; Bermudes, M.; Petit, T. Novel Insights into the Biotechnological Production of Haematococcus pluvialis-Derived Astaxanthin: Advances and Key Challenges to Allow Its Industrial Use as Novel Food Ingredient. J. Mar. Sci. Eng. 2020, 8, 789. [Google Scholar] [CrossRef]
  44. Miyakawa, K. Commercial Production of Astaxanthin from the Green Alga Haematococcus pluvialis. In Carotenoids: Biosynthetic and Biofunctional Approaches; Misawa, N., Ed.; Springer: Berlin/Heidelberg, Germany, 2021; pp. 3–10. [Google Scholar]
  45. Jiang, S.; Tong, S. Advances in Astaxanthin Biosynthesis in Haematococcus pluvialis. Chin. J. Biotechnol. 2019, 35, 988–997. [Google Scholar] [CrossRef]
  46. 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]
  47. Oslan, S.N.H.; Shoparwe, N.F.; Yusoff, A.H.; Rahim, A.A.; Chang, C.S.; Tan, J.S.; Oslan, S.N.; Arumugam, K.; Ariff, A.B.; Sulaiman, A.Z.; et al. A Review on Haematococcus pluvialis Bioprocess Optimization of Green and Red Stage Culture Conditions for the Production of Natural Astaxanthin. Biomolecules 2021, 11, 256. [Google Scholar] [CrossRef]
  48. Oslan, S.N.H.; Tan, J.S.; Oslan, S.N.; Matanjun, P.; Mokhtar, R.A.M.; Shapawi, R.; Huda, N. Haematococcus pluvialis as a Potential Source of Astaxanthin with Diverse Applications in Industrial Sectors: Current Research and Future Directions. Molecules 2021, 26, 6470. [Google Scholar] [CrossRef] [PubMed]
  49. Ren, Y.; Deng, J.; Huang, J.; Wu, Z.; Yi, L.; Bi, Y.; Chen, F. Using Green Alga Haematococcus pluvialis for Astaxanthin and Lipid Co-Production: Advances and Outlook. Bioresour. Technol. 2021, 340, 125736. [Google Scholar] [CrossRef] [PubMed]
  50. Souza, R.L.M.; Viana, W.K.R.; Moreira, M.S.M.; Soares Filho, A.A.; Lisboa, V.; de Vidigal, R.C.A.B.; Catter, K.M.; Eloy, H.R.F.; Matias, J.F.N. Technological Innovations in the Cultivation of Haematococcus pluvialis Microalgae. Sist. Gestão 2020, 15, 223–234. [Google Scholar] [CrossRef]
  51. Liu, J.; van der Meer, J.P.; Zhang, L.; Zhang, Y. Cultivation of Haematococcus pluvialis for Astaxanthin Production. In Microalgal Production for Biomass and High-Value Products; Slocombe, S.P., Benemann, J.R., Eds.; CRC Press: Boca Raton, FL, USA, 2017; pp. 267–293. [Google Scholar]
  52. Zhu, X.; Meng, C.; Sun, F.; Wei, Z.; Chen, L.; Chen, W.; Tong, S.; Du, H.; Gao, J.; Ren, J. Sustainable Production of Astaxanthin in Microorganisms: The Past, Present, and Future. Crit. Rev. Food Sci. Nutr. 2022, 1–17. [Google Scholar] [CrossRef]
  53. Lu, S.; Chen, G.; Wei, D. Effect of Electroporation Conditions on Cell Viability of Haematococcus pluvialis. Mod. Food Sci. Technol. 2010, 26, 554–557. [Google Scholar]
  54. Kim, J.Y.; Lee, C.; Jeon, M.S.; Park, J.; Choi, Y.-E. Enhancement of Microalga Haematococcus pluvialis Growth and Astaxanthin Production by Electrical Treatment. Bioresour. Technol. 2018, 268, 815–819. [Google Scholar] [CrossRef]
  55. Fitriana, H.-N.; Lee, S.-Y.; Choi, S.-A.; Lee, J.-Y.; Kim, B.-L.; Lee, J.-S.; Oh, Y.-K. Electric Stimulation of Astaxanthin Biosynthesis in Haematococcus pluvialis. Appl. Sci. 2021, 11, 3348. [Google Scholar] [CrossRef]
  56. Li, L.; Chen, Z.; Acheampong, A.; Huang, Q. Low-Temperature Plasma Promotes Growth of Haematococcus pluvialis and Accumulation of Astaxanthin by Regulating Histone H3 Lysine 4 Tri-Methylation. Bioresour. Technol. 2022, 343, 126095. [Google Scholar] [CrossRef]
  57. Gao, L.; Shi, X.; Wu, X. Applications and Challenges of Low Temperature Plasma in Pharmaceutical Field. J. Pharm. Anal. 2021, 11, 28–36. [Google Scholar] [CrossRef]
  58. Liu, J.; Chen, J.; Chen, Z.; Qin, S.; Huang, Q. Isolation and Characterization of Astaxanthin-Hyperproducing Mutants of Haematococcus pluvialis (Chlorophyceae) Produced by Dielectric Barrier Discharge Plasma. Phycologia 2016, 55, 650–658. [Google Scholar] [CrossRef]
  59. Wu, X.; Liu, Z.; Jiang, Y. Mutation Breeding of Haematococcus pluvialis by Atmospheric and Room Temperature Plasma and Isolation of High-Producing Mutants. J. Food Saf. Qual. 2016, 7, 3781–3787. [Google Scholar]
  60. Chen, Z.; Li, L.; Zhang, H.; Huang, Q. Stimulation of Biomass and Astaxanthin Accumulation in Haematococcus pluvialis Using Low-Temperature Plasma (LTP). Bioresour. Technol. Rep. 2020, 9, 100385. [Google Scholar] [CrossRef]
  61. Chen, Z.; Chen, J.; Liu, J.; Li, L.; Qin, S.; Huang, Q. Transcriptomic and Metabolic Analysis of an Astaxanthin-Hyperproducing Haematococcus pluvialis Mutant Obtained by Low-Temperature Plasma (LTP) Mutagenesis under High Light Irradiation. Algal Res. 2020, 45, 101746. [Google Scholar] [CrossRef]
  62. Bowen, W.R. Ultrastructural Aspects of the Cell Boundary of Haematococcus pluvialis. Trans. Am. Microsc. Soc. 1967, 86, 36–43. [Google Scholar] [CrossRef]
  63. Hagen, C.; Siegmund, S.; Braune, W. Ultrastructural and Chemical Changes in the Cell Wall of Haematococcus pluvialis (Volvocales, Chlorophyta) during Aplanospore Formation. Eur. J. Phycol. 2002, 37, 217–226. [Google Scholar] [CrossRef]
  64. Damiani, M.C.; Leonardi, P.I.; Pieroni, O.I.; Cáceres, E.J. Ultrastructure of the Cyst Wall of Haematococcus pluvialis (Chlorophyceae): Wall Development and Behaviour during Cyst Germination. Phycologia 2006, 45, 616–623. [Google Scholar] [CrossRef]
  65. D’Hondt, E.; Martin-Juarez, J.; Bolado, S.; Kasperoviciene, J.; Koreiviene, J.; Sulcius, S.; Elst, K.; Bastiaens, L. Cell Disruption Technologies. In Microalgae-Based Biofuels and Bioproducts. From Feedstock Cultivation to End-products; Muñoz, R., Gonzalez-Fernandez, C., Eds.; Woodhead Publishing: Sawston, UK, 2017; pp. 133–154. [Google Scholar]
  66. Praveenkumar, R.; Lee, J.; Vijayan, D.; Lee, S.Y.; Lee, K.; Sim, S.J.; Hong, M.E.; Kim, Y.-E.; Oh, Y.-K. Morphological Change and Cell Disruption of Haematococcus pluvialis Cyst during High-Pressure Homogenization for Astaxanthin Recovery. Appl. Sci. 2020, 10, 513. [Google Scholar] [CrossRef] [Green Version]
  67. Xu, R.; Zhang, L.; Yu, W.; Liu, J. A Strategy for Interfering with the Formation of Thick Cell Walls in Haematococcus pluvialis by down-Regulating the Mannan Synthesis Pathway. Bioresour. Technol. 2022, 362, 127783. [Google Scholar] [CrossRef]
  68. Zhu, F.; Shi, B.; Li, Y.; Yang, H.; Sun, S.; Yao, S.; Yin, X.; Zhang, J. Chemical Constituents of Extracellular Extract of Haematococcus pluvialis. Chem. Nat. Compd. 2019, 55, 578–579. [Google Scholar] [CrossRef]
  69. Marino, T.; Iovine, A.; Casella, P.; Martino, M.; Chianese, S.; Larocca, V.; Musmarra, D.; Molino, A. From Haematococcus pluvialis Microalgae a Powerful Antioxidant for Cosmetic Applications. Chem. Eng. Trans. 2020, 79, 271–276. [Google Scholar] [CrossRef]
  70. Shahid, A.; Khan, F.; Ahmad, N.; Farooq, M.; Mehmood, M.A. Microalgal Carbohydrates and Proteins: Synthesis, Extraction, Applications, and Challenges. In Microalgae Biotechnology for Food, Health and High Value Products; Alam, M.A., Wang, Z., Eds.; Springer: Singapore, 2020; pp. 433–468. [Google Scholar]
  71. Jung, S.M.; Jeon, J.; Park, T.H.; Yoon, J.H.; Lee, K.S.; Shin, H.W. Enhancing Production-Extraction and Antioxidant Activity of Astaxanthin from Haematococcus pluvialis. J. Environ. Biol. 2019, 40, 924–931. [Google Scholar] [CrossRef]
  72. Li, X.; Wang, X.; Duan, C.; Yi, S.; Gao, Z.; Xiao, C.; Agathos, S.N.; Wang, G.; Li, J. Biotechnological Production of Astaxanthin from the Microalga Haematococcus pluvialis. Biotechnol. Adv. 2020, 43, 107602. [Google Scholar] [CrossRef] [PubMed]
  73. National Center for Biotechnology Information (2023). PubChem Compound Summary for CID 5281224, Astaxanthin. Available online: https://pubchem.ncbi.nlm.nih.gov/compound/Astaxanthin (accessed on 2 February 2023).
  74. Harith, Z. Production of Astaxanthin by Xanthophyllomyces Dendrorhous DSMZ 5626 Using Rapeseed Meal Hydrolysates as Substrate. Ph.D. Thesis, Department of Food and Nutritional Sciences, University of Reading, Reading, UK, 2019. [Google Scholar]
  75. Johnson, E.A.; An, G.-H. Astaxanthin from Microbial Sources. Crit. Rev. Biotechnol. 1991, 11, 297–326. [Google Scholar] [CrossRef]
  76. Kim, S.-Y.; Cho, E.-A.; Yoo, J.-M.; In, M.-J.; Chae, H.-J. Solubility and Storage Stability of Astaxanthin. Korean Soc. Biotechnol. Bioeng. J. 2008, 23, 546–550. [Google Scholar]
  77. Bjørklund, G.; Gasmi, A.; Lenchyk, L.; Shanaida, M.; Zafar, S.; Mujawdiya, P.K.; Lysiuk, R.; Antonyak, H.; Noor, S.; Akram, M.; et al. The Role of Astaxanthin as a Nutraceutical in Health and Age-Related Conditions. Molecules 2022, 27, 7167. [Google Scholar] [CrossRef]
  78. Zhao, X.; Zhang, X.; Fu, L.; Zhu, H.; Zhang, B. Effect of Extraction and Drying Methods on Antioxidant Activity of Astaxanthin from Haematococcus pluvialis. Food Bioprod. Process. 2016, 99, 197–203. [Google Scholar] [CrossRef]
  79. Lemoine, Y.; Schoefs, B. Secondary Ketocarotenoid Astaxanthin Biosynthesis in Algae: A Multifunctional Response to Stress. Photosynth. Res. 2010, 106, 155–177. [Google Scholar] [CrossRef]
  80. Cheng, X.; Shah, M. Bioextraction of Astaxanthin Adopting Varied Techniques and Downstream Processing Methodologies. In Global Perspectives on Astaxanthin: From Industrial Production to Food, Health, and Pharmaceutical Applications; Ravishankar, G.A., Rao, A.R., Eds.; Academic Press: Cambridge, MA, USA, 2021; pp. 313–339. [Google Scholar]
  81. Galarza, J.I.; Vega, B.O.A.; Villón, J.; Henríquez, V. Deesterification of Astaxanthin and Intermediate Esters from Haematococcus pluvialis Subjected to Stress. Biotechnol. Rep. 2019, 23, e00351. [Google Scholar] [CrossRef]
  82. Nitsos, C.; Filali, R.; Taidi, B.; Lemaire, J. Current and Novel Approaches to Downstream Processing of Microalgae: A Review. Biotechnol. Adv. 2020, 45, 107650. [Google Scholar] [CrossRef]
  83. Bharte, S.; Desai, K. Techniques for Harvesting, Cell Disruption and Lipid Extraction of Microalgae for Biofuel Production. Biofuels 2021, 12, 285–305. [Google Scholar] [CrossRef]
  84. Putri, D.S.; Sari, D.A.; Zuhdia, L.D. Flocculants Optimization in Harvesting Freshwater Microalgae Haematococcus pluvialis. J. Kim. Ris. Sci. J. Chem. Res. 2020, 5, 49–54. [Google Scholar] [CrossRef]
  85. Butler, T.O.; Guimarães, B. Industrial Perspective on Downstream Processing of Haematococcus pluvialis. In Global Perspectives on Astaxanthin: From Industrial Production to Food, Health, and Pharmaceutical Applications; Ravishankar, G.A., Ambati, R.R., Eds.; Academic Press: Cambridge, MA, USA, 2021; pp. 283–311. [Google Scholar]
  86. Koopmann, I.K.; Möller, S.; Elle, C.; Hindersin, S.; Kramer, A.; Labes, A. Optimization of Astaxanthin Recovery in the Downstream Process of Haematococcus Pluvialis. Foods 2022, 11, 1352. [Google Scholar] [CrossRef] [PubMed]
  87. Ahirwar, A.; Meignen, G.; Khan, M.; Khan, N.; Rai, A.; Schoefs, B.; Marchand, J.; Varjani, S.; Vinayak, V. Nanotechnological Approaches to Disrupt the Rigid Cell Walled Microalgae Grown in Wastewater for Value-Added Biocompounds: Commercial Applications, Challenges, and Breakthrough. Biomass Convers. Biorefinery 2021, 1–26. [Google Scholar] [CrossRef]
  88. Shpigelman, A.; Okun, Z. Efficient Production of Functional and Bioactive Compounds and Foods for Use in Food, Pharma, Cosmetic and Other Industries. In Nonthermal Processing in Agri-Food-Bio Sciences; Režek Jambrak, A., Ed.; Springer: Cham, Switzerland, 2022; pp. 623–637. [Google Scholar]
  89. Sarkarat, R.; Mohamadnia, S.; Tavakoli, O. Recent Advances in Non-Conventional Techniques for Extraction of Phycobiliproteins and Carotenoids from Microalgae. Braz. J. Chem. Eng. 2022, 1–22. [Google Scholar] [CrossRef]
  90. Kim, D.-Y.; Vijayan, D.; Praveenkumar, R.; Han, J.-I.; Lee, K.; Park, J.-Y.; Chang, W.-S.; Lee, J.-S.; Oh, Y.-K. Cell-Wall Disruption and Lipid/Astaxanthin Extraction from Microalgae: Chlorella and Haematococcus. Bioresour. Technol. 2016, 199, 300–310. [Google Scholar] [CrossRef]
  91. Machado Jr, F.R.S.; Trevisol, T.C.; Boschetto, D.L.; Burkert, J.F.M.; Ferreira, S.R.S.; Oliveira, J.V.; Burkert, C.A.V. Technological Process for Cell Disruption, Extraction and Encapsulation of Astaxanthin from Haematococcus pluvialis. J. Biotechnol. 2016, 218, 108–114. [Google Scholar] [CrossRef] [Green Version]
  92. Liu, Z.-W.; Zeng, X.-A.; Cheng, J.-H.; Liu, D.-B.; Aadil, R.M. The Efficiency and Comparison of Novel Techniques for Cell Wall Disruption in Astaxanthin Extraction from Haematococcus pluvialis. Int. J. Food Sci. Technol. 2018, 53, 2212–2219. [Google Scholar] [CrossRef]
  93. Butler, T.; Golan, Y. Astaxanthin Production from Microalgae. In Microalgae Biotechnology for Food, Health and High Value Products; Alam, M.A., Xu, J.-L., Wang, Z., Eds.; Springer: Berlin/Heidelberg, Germany, 2020; pp. 175–242. [Google Scholar]
  94. Tan, Y.; Ye, Z.; Wang, M.; Manzoor, M.F.; Aadil, R.M.; Tan, X.; Liu, Z. Comparison of Different Methods for Extracting the Astaxanthin from Haematococcus pluvialis: Chemical Composition and Biological Activity. Molecules 2021, 26, 3569. [Google Scholar] [CrossRef]
  95. Bauer, A.; Minceva, M. Direct Extraction of Astaxanthin from the Microalgae Haematococcus pluvialis Using Liquid--Liquid Chromatography. RSC Adv. 2019, 9, 22779–22789. [Google Scholar] [CrossRef] [Green Version]
  96. Bueno, M.; Gallego, R.; Chourio, A.M.; Ibáñez, E.; Herrero, M.; Saldaña, M.D.A. Green Ultra-High Pressure Extraction of Bioactive Compounds from Haematococcus pluvialis and Porphyridium Cruentum Microalgae. Innov. Food Sci. Emerg. Technol. 2020, 66, 102532. [Google Scholar] [CrossRef]
  97. Macias-Sánchez, M.D. High-Pressure Extraction of Astaxanthin from Haematococcus pluvialis. In Global Perspectives on Astaxanthin: From Industrial Production to Food, Health, and Pharmaceutical Applications; Elsevier: Amsterdam, The Netherlands, 2021; pp. 355–373. [Google Scholar]
  98. Zhao, X.; Fu, L.; Liu, D.; Zhu, H.; Wang, X.; Bi, Y. Magnetic-Field-Assisted Extraction of Astaxanthin from Haematococcus pluvialis. J. Food Process. Preserv. 2016, 40, 463–472. [Google Scholar] [CrossRef]
  99. Khoo, K.S.; Chew, K.W.; Yew, G.Y.; Manickam, S.; Ooi, C.W.; Show, P.L. Integrated Ultrasound-Assisted Liquid Biphasic Flotation for Efficient Extraction of Astaxanthin from Haematococcus pluvialis. Ultrason. Sonochem. 2020, 67, 105052. [Google Scholar] [CrossRef] [PubMed]
  100. Kim, S.-Y.; Cho, E.-A.; Yoo, J.-M.; In, M.-J.; Chae, H.-J. Extraction and Analysis of Astaxanthin from Haematococcus pluvialis Using Sonication. J. Korean Soc. Food Sci. Nutr. 2008, 37, 1363–1368. [Google Scholar] [CrossRef]
  101. Zou, T.-B.; Jia, Q.; Li, H.-W.; Wang, C.-X.; Wu, H.-F. Response Surface Methodology for Ultrasound-Assisted Extraction of Astaxanthin from Haematococcus pluvialis. Mar. Drugs 2013, 11, 1644–1655. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Plîngău, E.; Rudi, L.; Miscu, V.; Dencicov-Cristea, L. Application of Microwave Pretreatment Technique of Aplanospores of Haematococcus pluvialis for Astaxanthin Recovering. In Proceedings of the Microbial Biotechnology, Chișinău, Moldova, 11–12 October 2018; p. 91. [Google Scholar]
  103. Zhao, L.; Chen, G.; Zhao, G.; Hu, X. Optimization of Microwave-Assisted Extraction of Astaxanthin from Haematococcus pluvialis by Response Surface Methodology and Antioxidant Activities of the Extracts. Sep. Sci. Technol. 2009, 44, 243–262. [Google Scholar] [CrossRef]
  104. Desai, R.K.; Streefland, M.; Wijffels, R.H.; Eppink, M.H.M. Novel Astaxanthin Extraction from Haematococcus pluvialis Using Cell Permeabilising Ionic Liquids. Green Chem. 2016, 18, 1261–1267. [Google Scholar] [CrossRef]
  105. Tavanandi, H.A.; Devi, A.C.; Raghavarao, K. Sustainable Pre-Treatment Methods for Downstream Processing of Harvested Microalgae. In Sustainable Downstream Processing of Microalgae for Industrial Application; Gayen, K., Bhowmick, T.K., Maity, S.K., Eds.; CRC Press: Boca Raton, FL, USA, 2019; pp. 101–135. ISBN 9780429027970. [Google Scholar]
  106. Liu, Z.-W.; Yue, Z.; Zeng, X.-A.; Cheng, J.-H.; Aadil, R.M. Ionic Liquid as an Effective Solvent for Cell Wall Deconstructing through Astaxanthin Extraction from Haematococcus pluvialis. Int. J. Food Sci. Technol. 2019, 54, 583–590. [Google Scholar] [CrossRef]
  107. Choi, S.-A.; Oh, Y.-K.; Lee, J.; Sim, S.J.; Hong, M.E.; Park, J.-Y.; Kim, M.-S.; Kim, S.W.; Lee, J.-S. High-Efficiency Cell Disruption and Astaxanthin Recovery from Haematococcus pluvialis Cyst Cells Using Room-Temperature Imidazolium-Based Ionic Liquid/Water Mixtures. Bioresour. Technol. 2019, 274, 120–126. [Google Scholar] [CrossRef]
  108. In, M.-J.; Choi, J.-H.; Kim, S.-Y.; Chae, H.-J.; Kim, D.-H. Enhanced Extraction of Astaxanthin from Haematococcus pluvialis Using Enzyme Treatments. Appl. Biol. Chem. 2008, 51, 247–249. [Google Scholar]
  109. Corrêa, P.S.; Morais Júnior, W.G.; Martins, A.A.; Caetano, N.S.; Mata, T.M. Microalgae Biomolecules: Extraction, Separation and Purification Methods. Processes 2020, 9, 10. [Google Scholar] [CrossRef]
  110. Rodríguez-Meizoso, I.; Jaime, L.; Santoyo, S.; Señoráns, F.J.; Cifuentes, A.; Ibáñez, E. Subcritical Water Extraction and Characterization of Bioactive Compounds from Haematococcus pluvialis Microalga. J. Pharm. Biomed. Anal. 2010, 51, 456–463. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Bustamante, A.; Roberts, P.; Aravena, R.; Del Valle, J.M. Supercritical Extraction of Astaxanthin from H. Pluvialis Using Ethanol-Modified CO2. Experiments and Modeling. In Proceedings of the 11th International Conference of Eng Food, Athens, Greece, 20–26 May 2011; pp. 22–26. [Google Scholar]
  112. Reyes, F.A.; Muñoz, L.A.; Hansen, A.; del Valle, J.M. Water Relationships in Haematoccoccus Pluvialis and Their Effect in High-Pressure Agglomeration for Supercritical CO2 Extraction. J. Food Eng. 2015, 162, 18–24. [Google Scholar] [CrossRef]
  113. Di Sanzo, G.; Mehariya, S.; Martino, M.; Larocca, V.; Casella, P.; Chianese, S.; Musmarra, D.; Balducchi, R.; Molino, A. Supercritical Carbon Dioxide Extraction of Astaxanthin, Lutein, and Fatty Acids from Haematococcus pluvialis Microalgae. Mar. Drugs 2018, 16, 334. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Molino, A.; Mehariya, S.; Iovine, A.; Larocca, V.; Di Sanzo, G.; Martino, M.; Casella, P.; Chianese, S.; Musmarra, D. Extraction of Astaxanthin and Lutein from Microalga Haematococcus Pluvialis in the Red Phase Using CO2 Supercritical Fluid Extraction Technology with Ethanol as Co-Solvent. Mar. Drugs 2018, 16, 432. [Google Scholar] [CrossRef] [Green Version]
  115. Ruiz-Domínguez, M.C.; Espinosa, C.; Paredes, A.; Palma, J.; Jaime, C.; Vílchez, C.; Cerezal, P. Determining the Potential of Haematococcus pluvialis Oleoresin as a Rich Source of Antioxidants. Molecules 2019, 24, 4073. [Google Scholar] [CrossRef] [Green Version]
  116. Espinosa Álvarez, C.; Vardanega, R.; Salinas-Fuentes, F.; Palma Ramírez, J.; Bugueño Muñoz, W.; Jiménez-Rondón, D.; Meireles, M.A.A.; Cerezal Mezquita, P.; Ruiz-Domínguez, M.C. Effect of CO2 Flow Rate on the Extraction of Astaxanthin and Fatty Acids from Haematococcus Pluvialis Using Supercritical Fluid Technology. Molecules 2020, 25, 6044. [Google Scholar] [CrossRef]
  117. Khoo, K.S.; Ooi, C.W.; Chew, K.W.; Foo, S.C.; Lim, J.W.; Tao, Y.; Jiang, N.; Ho, S.-H.; Show, P.L. Permeabilization of Haematococcus Pluvialis and Solid-Liquid Extraction of Astaxanthin by CO2-Based Alkyl Carbamate Ionic Liquids. Chem. Eng. J. 2021, 411, 128510. [Google Scholar] [CrossRef]
  118. Aravena, R.I.; del Valle, J.M.; de la Fuente, J.C. Supercritical CO2 Extraction of Aqueous Suspensions of Disrupted Haematococcus Pluvialis Cysts. J. Supercrit. Fluids 2022, 181, 105392. [Google Scholar] [CrossRef]
  119. Vardanega, R.; Cerezal-Mezquita, P.; Veggi, P.C. Supercritical Fluid Extraction of Astaxanthin-Rich Extracts from Haematococcus pluvialis: Economic Assessment. Bioresour. Technol. 2022, 361, 127706. [Google Scholar] [CrossRef]
  120. Saini, K.C.; Yadav, D.S.; Mehariya, S.; Rathore, P.; Kumar, B.; Marino, T.; Leone, G.P.; Verma, P.; Musmarra, D.; Molino, A. Overview of Extraction of Astaxanthin from Haematococcus Pluvialis Using CO2 Supercritical Fluid Extraction Technology Vis-a-Vis Quality Demands. In Global Perspectives on Astaxanthin: From Industrial Production to Food, Health, and Pharmaceutical Applications; Ravishankar, G., Ambati, R., Eds.; Academic Press: London, UK, 2021; pp. 341–354. ISBN 9780128233054. [Google Scholar]
  121. Youn, H.-S.; Roh, M.-K.; Weber, A.; Wilkinson, G.T.; Chun, B.-S. Solubility of Astaxanthin in Supercritical Carbon Dioxide. Korean J. Chem. Eng. 2007, 24, 831–834. [Google Scholar] [CrossRef]
  122. Molino, A.; Rimauro, J.; Casella, P.; Cerbone, A.; Larocca, V.; Chianese, S.; Karatza, D.; Mehariya, S.; Ferraro, A.; Hristoforou, E. Extraction of Astaxanthin from Microalga Haematococcus pluvialis in Red Phase by Using Generally Recognized as Safe Solvents and Accelerated Extraction. J. Biotechnol. 2018, 283, 51–61. [Google Scholar] [CrossRef] [PubMed]
  123. Castillo, A.; Pereira, S.; Otero, A.; Garcia-Jares, C.; Lores, M. Multicomponent Bioactive Extract from Red Stage Haematococcus pluvialis Wet Paste: Avoiding the Drying Step and Toxic Solvents. J. Appl. Phycol. 2022, 34, 1537–1553. [Google Scholar] [CrossRef]
  124. Grimi, N.; Dubois, A.; Marchal, L.; Jubeau, S.; Lebovka, N.I.; Vorobiev, E. Selective Extraction from Microalgae Nannochloropsis Sp. Using Different Methods of Cell Disruption. Bioresour. Technol. 2014, 153, 254–259. [Google Scholar] [CrossRef] [PubMed]
  125. 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] [PubMed]
  126. Parniakov, O.; Barba, F.J.; Grimi, N.; Marchal, L.; Jubeau, S.; Lebovka, N.; Vorobiev, E. Pulsed Electric Field and PH Assisted Selective Extraction of Intracellular Components from Microalgae Nannochloropsis. Algal Res. 2015, 8, 128–134. [Google Scholar] [CrossRef]
  127. Parniakov, O.; Barba, F.J.; Grimi, N.; Marchal, L.; Jubeau, S.; Lebovka, N.; Vorobiev, E. Pulsed Electric Field Assisted Extraction of Nutritionally Valuable Compounds from Microalgae Nannochloropsis Spp. Using the Binary Mixture of Organic Solvents and Water. Innov. Food Sci. Emerg. Technol. 2015, 27, 79–85. [Google Scholar] [CrossRef]
  128. Akyil, S.; İlter, I.; Koç, M.; Ertekin, F. Recent Trends in Extraction Techniques for High Value Compounds from Algae as Food Additives. Turk. J. Agric. Sci. Technol. 2018, 6, 1008–1014. [Google Scholar] [CrossRef] [Green Version]
  129. Martínez, J.M.; Delso, C.; Álvarez, I.; Raso, J. Pulsed Electric Field-Assisted Extraction of Valuable Compounds from Microorganisms. Compr. Rev. Food Sci. Food Saf. 2020, 19, 530–552. [Google Scholar] [CrossRef]
  130. Donsì, F.; Ferrari, G.; Pataro, G. Emerging Technologies for the Clean Recovery of Antioxidants from Microalgae. In Microalgae; Galanakis, C.M., Ed.; Academic Press: Cambridge, MA, USA, 2021; pp. 173–205. [Google Scholar]
  131. Poojary, M.M.; Barba, F.J.; Aliakbarian, B.; Donsì, F.; Pataro, G.; Dias, D.A.; Juliano, P. Innovative Alternative Technologies to Extract Carotenoids from Microalgae and Seaweeds. Mar. Drugs 2016, 14, 214. [Google Scholar] [CrossRef] [Green Version]
  132. Amaro, H.M.; Sousa-Pinto, I.; Malcata, F.X.; Guedes, A.C. Microalgae as a Source of Pigments: Extraction and Purication Methods. In Marine Microorganisms; Nollet, L.M.L., Ed.; CRC Press: Boca Raton, FL, USA, 2016; pp. 113–142. [Google Scholar]
  133. Gogate, P.R.; Joshi, S.M. Process Intensification Aspects of Extraction of Pigments from Microalgae. In Pigments from Microalgae Handbook; Jacob-Lopes, E., Queiroz, M.I., Zepka, L.Q., Eds.; Springer Nature: Berlin/Heidelberg, Germany, 2020; pp. 309–324. [Google Scholar]
  134. Pagels, F.; Pereira, R.N.; Vicente, A.A.; Guedes, A.C. Extraction of Pigments from Microalgae and Cyanobacteria—A Review on Current Methodologies. Appl. Sci. 2021, 11, 5187. [Google Scholar] [CrossRef]
  135. Pereira, R.N.; Avelar, Z.; Pereira, S.G.; Rocha, C.M.R.; Teixeira, J.A. Pulsed Electric Fields for the Extraction of Proteins and Carbohydrates from Marine Resources. In Innovative and Emerging Technologies in the Bio-marine Food Sector; Garcia-Vaquero, M., Rajauria, G., Eds.; Academic Press: Cambridge, MA, USA, 2022; pp. 173–195. [Google Scholar]
  136. Lafarga, T.; Aguiló-Aguayo, I. Pulsed Electric Fields for the Extraction of Lipids, Pigments, and Polyphenols from Cultured Microalgae. In Innovative and Emerging Technologies in the Bio-marine Food Sector; Garcia-Vaquero, M., Rajauria, G., Eds.; Academic Press: Cambridge, MA, USA, 2022; pp. 197–221. [Google Scholar]
  137. Suwal, S.; Marciniak, A. Technologies for the Extraction, Separation and Purification of Polyphenols--A Review. Nepal J. Biotechnol. 2018, 6, 74–91. [Google Scholar] [CrossRef] [Green Version]
  138. Patel, A.; Mikes, F.; Matsakas, L. An Overview of Current Pretreatment Methods Used to Improve Lipid Extraction from Oleaginous Microorganisms. Molecules 2018, 23, 1562. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. bin Azmi, A.A.; Sankaran, R.; Show, P.L.; Ling, T.C.; Tao, Y.; Munawaroh, H.S.H.; San Kong, P.; Lee, D.-J.; Chang, J.-S. Current Application of Electrical Pre-Treatment for Enhanced Microalgal Biomolecules Extraction. Bioresour. Technol. 2020, 302, 122874. [Google Scholar] [CrossRef] [PubMed]
  140. Ahirwar, A.; Khan, M.J.; Sirotiya, V.; Mourya, M.; Rai, A.; Schoefs, B.; Marchand, J.; Varjani, S.; Vinayak, V. Pulsed Electric Field--Assisted Cell Permeabilization of Microalgae (Haematococcus pluvialis) for Milking of Value-Added Compounds. BioEnergy Res. 2022, 1–14. [Google Scholar] [CrossRef]
  141. Coustets, M.; Joubert-Durigneux, V.; Hérault, J.; Schoefs, B.; Blanckaert, V.; Garnier, J.-P.; Teissié, J. Optimization of Protein Electroextraction from Microalgae by a Flow Process. Bioelectrochemistry 2015, 103, 74–81. [Google Scholar] [CrossRef]
  142. Gateau, H.; Blanckaert, V.; Veidl, B.; Burlet-Schiltz, O.; Pichereaux, C.; Gargaros, A.; Marchand, J.; Schoefs, B. Application of Pulsed Electric Fields for the Biocompatible Extraction of Proteins from the Microalga Haematococcus pluvialis. Bioelectrochemistry 2021, 137, 107588. [Google Scholar] [CrossRef]
  143. Martínez, J.M.; Gojkovic, Z.; Ferro, L.; Maza, M.; Álvarez, I.; Raso, J.; Funk, C. Use of Pulsed Electric Field Permeabilization to Extract Astaxanthin from the Nordic Microalga Haematococcus pluvialis. Bioresour. Technol. 2019, 289, 121694. [Google Scholar] [CrossRef]
  144. Rodríguez-Sifuentes, L.; Marszalek, J.E.; Hernández-Carbajal, G.; Chuck-Hernández, C. Importance of Downstream Processing of Natural Astaxanthin for Pharmaceutical Application. Front. Chem. Eng. 2021, 2, 601483. [Google Scholar] [CrossRef]
  145. Yan, N.; Fan, C.; Chen, Y.; Hu, Z. The Potential for Microalgae as Bioreactors to Produce Pharmaceuticals. Int. J. Mol. Sci. 2016, 17, 962. [Google Scholar] [CrossRef] [Green Version]
  146. Yang, Y.; Hassan, S.H.A.; Awasthi, M.K.; Gajendran, B.; Sharma, M.; Ji, M.-K.; Salama, E.-S. The Recent Progress on the Bioactive Compounds from Algal Biomass for Human Health Applications. Food Biosci. 2022, 51, 102267. [Google Scholar] [CrossRef]
  147. Oliveira, C.; Coelho, C.; Teixeira, J.A.; Ferreira-Santos, P.; Botelho, C.M. Nanocarriers as Active Ingredients Enhancers in the Cosmetic Industry—The European and North America Regulation Challenges. Molecules 2022, 27, 1669. [Google Scholar] [CrossRef] [PubMed]
  148. Chakraborty, T.; Akhtar, N. Biofertilizers: Characteristic Features and Applications. In Biofertilizers: Study and Impact; Ahamed, I.M.I., Boddula, R., Rezakazemi, M., Eds.; Wiley Online Library: Hoboken, NJ, USA, 2021; pp. 429–489. [Google Scholar]
  149. Capelli, B.; Bagchi, D.; Cysewski, G.R. Synthetic Astaxanthin Is Significantly Inferior to Algal-Based Astaxanthin as an Antioxidant and May Not Be Suitable as a Human Nutraceutical Supplement. Nutrafoods 2013, 12, 145–152. [Google Scholar] [CrossRef]
  150. Yamashita, E. Extensive Bioactivity of Astaxanthin from Haematococcus pluvialis in Human. In Carotenoids: Biosynthetic and Biofunctional Approaches; Misawa, N., Ed.; Advances in Experimental Medicine and Biology; Springer: Singapore, 2021; Volume 1261, pp. 249–259. ISBN 978-981-15-7359-0. [Google Scholar]
  151. Higuera-Ciapara, I.; Felix-Valenzuela, L.; Goycoolea, F.M. Astaxanthin: A Review of Its Chemistry and Applications. Crit. Rev. Food Sci. Nutr. 2006, 46, 185–196. [Google Scholar] [CrossRef] [PubMed]
  152. Patel, A.K.; Tambat, V.S.; Chen, C.-W.; Chauhan, A.S.; Kumar, P.; Vadrale, A.P.; Huang, C.-Y.; Dong, C.-D.; Singhania, R.R. Recent Advancements in Astaxanthin Production from Microalgae: A Review. Bioresour. Technol. 2022, 154, 128030. [Google Scholar] [CrossRef]
  153. Mularczyk, M.; Michalak, I.; Marycz, K. Astaxanthin and Other Nutrients from Haematococcus pluvialis—Multifunctional Applications. Mar. Drugs 2020, 18, 459. [Google Scholar] [CrossRef]
  154. Mota, G.C.P.; de Moraes, L.B.S.; Oliveira, C.Y.B.; Oliveira, D.W.S.; de Abreu, J.L.; Dantas, D.M.M.; Gálvez, A.O. Astaxanthin from Haematococcus pluvialis: Processes, Applications, and Market. Prep. Biochem. Biotechnol. 2022, 52, 598–609. [Google Scholar] [CrossRef]
  155. Brendler, T.; Williamson, E.M. Astaxanthin: How Much Is Too Much? A Safety Review. Phyther. Res. 2019, 33, 3090–3111. [Google Scholar] [CrossRef]
  156. Stachowiak, B.; Szulc, P. Astaxanthin for the Food Industry. Molecules 2021, 26, 2666. [Google Scholar] [CrossRef]
  157. Patil, A.D.; Kasabe, P.J.; Dandge, P.B. Pharmaceutical and Nutraceutical Potential of Natural Bioactive Pigment: Astaxanthin. Nat. Prod. Bioprospect. 2022, 12, 1–27. [Google Scholar] [CrossRef]
  158. Kalasariya, H.S.; Yadav, V.K.; Yadav, K.K.; Tirth, V.; Algahtani, A.; Islam, S.; Gupta, N.; Jeon, B.-H. Seaweed-Based Molecules and Their Potential Biological Activities: An Eco-Sustainable Cosmetics. Molecules 2021, 26, 5313. [Google Scholar] [CrossRef]
  159. Kumar, S.; Kumar, R.; Diksha; Kumari, A.; Panwar, A. Astaxanthin: A Super Antioxidant from Microalgae and Its Therapeutic Potential. J. Basic Microbiol. 2022, 62, 1064–1082. [Google Scholar] [CrossRef] [PubMed]
  160. Zhang, L.; Wang, H. Multiple Mechanisms of Anti-Cancer Effects Exerted by Astaxanthin. Mar. Drugs 2015, 13, 4310–4330. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  161. Kishimoto, Y.; Yoshida, H.; Kondo, K. Potential Anti-Atherosclerotic Properties of Astaxanthin. Mar. Drugs 2016, 14, 35. [Google Scholar] [CrossRef] [PubMed]
  162. Chang, M.X.; Xiong, F. Astaxanthin and Its Effects in Inflammatory Responses and Inflammation-Associated Diseases: Recent Advances and Future Directions. Molecules 2020, 25, 5342. [Google Scholar] [CrossRef]
Figure 1. Overview of different steps in H. pluvialis microalgae processing biorefinery.
Figure 1. Overview of different steps in H. pluvialis microalgae processing biorefinery.
Molecules 28 02089 g001
Figure 2. Description of H. pluvialis cell structure at the resistant cyst stage. Based on the data collected in [62,63,64,65,66,67].
Figure 2. Description of H. pluvialis cell structure at the resistant cyst stage. Based on the data collected in [62,63,64,65,66,67].
Molecules 28 02089 g002
Figure 3. Initial biomass and supercritical CO2 extract relative percentage proportions for different biomolecular component compositions in the H. pluvialis red phase. Based on the data obtained in [69].
Figure 3. Initial biomass and supercritical CO2 extract relative percentage proportions for different biomolecular component compositions in the H. pluvialis red phase. Based on the data obtained in [69].
Molecules 28 02089 g003
Figure 4. Relative antioxidant activity of natural astaxanthin compared with other antioxidants in the neutralization of destructive reactive oxygen species found in the human body. Based on the data collected in [26,77,150,151].
Figure 4. Relative antioxidant activity of natural astaxanthin compared with other antioxidants in the neutralization of destructive reactive oxygen species found in the human body. Based on the data collected in [26,77,150,151].
Molecules 28 02089 g004
Table 1. Solubility of astaxanthin at room temperature (25 °C) in different organic solvents.
Table 1. Solubility of astaxanthin at room temperature (25 °C) in different organic solvents.
SolventSolubility (25 °C), References
Water≈10−12 g/L, [73]
Ethanol0.038 g/L, [74]
Methanol0.04 g/L, [74]
Acetone0.55 g/L, 0.2 g/L, [75]
Dimethyl sulfoxide1.64 g/L [74], 0.5 g/L [75]
Dimethyl sulfoxide: acetone mixture (1:1)2.03 g/L, [74]
Chloroform10 g/L, [75]
Dichloromethane30 g/L, [75]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gherabli, A.; Grimi, N.; Lemaire, J.; Vorobiev, E.; Lebovka, N. Extraction of Valuable Biomolecules from the Microalga Haematococcus pluvialis Assisted by Electrotechnologies. Molecules 2023, 28, 2089. https://doi.org/10.3390/molecules28052089

AMA Style

Gherabli A, Grimi N, Lemaire J, Vorobiev E, Lebovka N. Extraction of Valuable Biomolecules from the Microalga Haematococcus pluvialis Assisted by Electrotechnologies. Molecules. 2023; 28(5):2089. https://doi.org/10.3390/molecules28052089

Chicago/Turabian Style

Gherabli, Adila, Nabil Grimi, Julien Lemaire, Eugène Vorobiev, and Nikolai Lebovka. 2023. "Extraction of Valuable Biomolecules from the Microalga Haematococcus pluvialis Assisted by Electrotechnologies" Molecules 28, no. 5: 2089. https://doi.org/10.3390/molecules28052089

Article Metrics

Back to TopTop