Next Article in Journal
The First Step in Standardizing an Artificial Aging Protocol for Dental Composites—Evaluation of Basic Protocols
Next Article in Special Issue
Natural Formulations Based on Olea europaea L. Fruit Extract for the Topical Treatment of HSV-1 Infections
Previous Article in Journal
Screening of Sourdough Starter Strains and Improvements in the Quality of Whole Wheat Steamed Bread
Previous Article in Special Issue
Extraction, Encapsulation into Lipid Vesicular Systems, and Biological Activity of Rosa canina L. Bioactive Compounds for Dermocosmetic Use
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Microalgae Bioactive Compounds to Topical Applications Products—A Review

by
Manuel Martínez-Ruiz
1,
Carlos Alberto Martínez-González
1,
Dong-Hyun Kim
1,
Berenice Santiesteban-Romero
1,
Humberto Reyes-Pardo
1,
Karen Rocio Villaseñor-Zepeda
1,
Edgar Ricardo Meléndez-Sánchez
1,
Diana Ramírez-Gamboa
1,
Ana Laura Díaz-Zamorano
1,
Juan Eduardo Sosa-Hernández
1,
Karina G. Coronado-Apodaca
1,
Ana María Gámez-Méndez
2,
Hafiz M. N. Iqbal
1,* and
Roberto Parra-Saldivar
1,*
1
School of Engineering and Sciences, Tecnológico de Monterrey, Monterrey 64849, Mexico
2
Basic Sciences Department, Universidad de Monterrey, San Pedro Garza García 66238, Mexico
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(11), 3512; https://doi.org/10.3390/molecules27113512
Submission received: 2 February 2022 / Revised: 21 May 2022 / Accepted: 24 May 2022 / Published: 30 May 2022
(This article belongs to the Special Issue New Trends in Skin Care: Topical Delivery of Cosmeceutical Molecules)

Abstract

:
Microalgae are complex photosynthetic organisms found in marine and freshwater environments that produce valuable metabolites. Microalgae-derived metabolites have gained remarkable attention in different industrial biotechnological processes and pharmaceutical and cosmetic industries due to their multiple properties, including antioxidant, anti-aging, anti-cancer, phycoimmunomodulatory, anti-inflammatory, and antimicrobial activities. These properties are recognized as promising components for state-of-the-art cosmetics and cosmeceutical formulations. Efforts are being made to develop natural, non-toxic, and environmentally friendly products that replace synthetic products. This review summarizes some potential cosmeceutical applications of microalgae-derived biomolecules, their mechanisms of action, and extraction methods.

1. Introduction

Cosmetics are products designed to improve the appearance of the skin without affecting its function and structure due to the active ingredients contained in the product [1,2]. On the other hand, the cosmetic industry uses the word “cosmeceuticals” to refer to cosmetic products with ingredients that have medicinal or drug-like benefits [1,3]. The vast majority of cosmeceutical products are lotions or creams for topical use. Although the terms may seem confusing, and cosmeceuticals are not formally recognized by the United States Food and Drug Administration (US-FDA) or the European Union, dermatologists prescribe 30–40% of these products [4]. Furthermore, the global natural cosmetics market showed 10–11% annual growth from 2015 to 2019 [5]. While botanical cosmeceuticals are in demand and display numerous benefits, the potential of microalgae as an alternative source for cosmeceutical products has been extensively studied [6,7,8].
Algae are a diverse group of photosynthetic eukaryotic microorganisms that possess different structures and forms and can be divided into either macro or microalgae based on their size. Many microalgae species have been identified as having various biochemical characteristics associated with nutritional benefits and human health.
Microalgae synthesize an extensive diversity of compounds from different metabolic pathways such as amino acids, fatty acids, lycopene, polysaccharides, steroids, carotenoids, lectins, polyketides, toxins, etc. Some of these are shown in Figure 1.
Microalgae produce and secrete valuable metabolites as a result of being constantly exposed to several stress conditions, such as high or low temperature, high salinity, osmotic pressure, photo-oxidation, and ultraviolet radiation [9]. Within the group of organisms with photosynthetic capacity are microalgae and cyanobacteria, microorganisms classified as eukaryotes and prokaryotes, respectively [10]. Both present great biodiversity of species, ranging between 80,000 and 200,000 [11]. Microalgae and cyanobacteria are responsible for 32% of photosynthetic activity in the world; therefore, they play an important role in various ecosystems [12]. The biotechnological interest in microalgae and cyanobacteria has focused on their content of valuable metabolites that can be utilized to develop a wide range of different applications in fields such as medicine, industry, energy, agriculture, food, and others. Stress factors increase or decrease the production of different convenient compounds; for this reason, scientific research focuses on defining the best microalgae cultivation conditions to obtain the maximum industrial production [13]. Furthermore, it is widely recognized that microalgae metabolites exhibit a wide range of biological activities such as ultraviolet (UV) absorbing, antioxidant, anti-aging, anti-blemish, anti-inflammatory, and antimicrobial properties [1,14]. Additionally, microalgae capture CO2 more efficiently than trees, reducing the greenhouse effect [15].
Some of the advantages of microalgae include a higher annual photon-to-biomass conversion efficiency in comparison with plants, as reported in previous studies; microalgae can growth in wastewater as treatment for water bioremediation [16,17,18,19,20,21]. Even though it is possible to use wastewater to produce microalgal biomass and obtain bioactive compounds, their use for topical applications or human consumption still represents a great challenge due to the diversity of contaminants present in wastewater, such as pathogens and heavy metals, despite pre-treatment steps such as sterilization or filtration to remove bacteria and fungi and recovery of heavy metals using coagulation–flocculation methods [22]. Pre-treatment also helps to reduce the concentration of nutrients and the turbidity of the water to allow the penetration of light necessary for microalgal growth [23,24]. Furthermore, methods of extraction and purification of compounds of interest are being investigated for their safe application in cosmetic and nutraceutical products [25,26].
Currently, some microalgae-derived products are available on the market, and the biotechnology of microalgae is attracting more attention. Labels such as ©DSM use Nannochloropsis oculata and Dunaliella salina extract in some of their products, such as PEPHA®-TIGHT CB or PEPHA®-CTIVE CB, to provide skin-tightening effects and stimulate cell proliferation, respectively [27]. However, an ever-increasing effort is being invested in the search for new environmentally friendly cosmetic products. All of the diverse and interesting features qualify microalgae metabolites as an alternative feedstock or potential source of biomolecules of commercial interest.
The present review describes the accumulated knowledge about the foremost groups of cosmeceutical compounds from microalgae metabolites and putative drug delivery methods using microalgae-based biotechnology.

2. Immunomodulatory Activity of Compounds from Microalgae

Microalgae and cyanobacteria serve as a source of bioactive compounds that present immunomodulatory activity, such as amino acids [28,29,30], peptides, pigment-protein complex, and exopolysaccharides, the latter known to possess anti-inflammatory, antimicrobial, and antiviral properties [1,31,32,33,34]. The major sugars involved in the composition of polysaccharides are xylose, glucose, mannose, and galactose [35]. Specifically, sulfated polysaccharides have been associated with blood coagulation, antilipidemic activity, and mainly immunomodulatory activity [31]. Multiple studies have reported anti-inflammatory properties from Arthrospira/Spirulina, Chlorella vulgaris, Chlorella pyrenoidosa, Isochrysis, Pleurochrysis carterae, Dunaliella, Porphyridium purpureum, and Rhodosorus marinus [29,35,36,37,38,39].
Immunomodulatory agents refer to compounds that increase (immunostimulators) or decrease (immunosuppressants) the response of the immune system. In cosmetology, topical immunomodulators are used to regulate the local immune response of the skin to treat a wide range of skin diseases [40,41]. Moreover, topical immunomodulators are easier to apply and safer for longer periods in comparison with systemic immunomodulators [40,41,42].
Anti-inflammatory, antitumoral, antimicrobial, and antiviral properties from a wide variety of compounds are directly linked to immunomodulator activities as they work synergistically with the immune system to elicit an anti-inflammatory response against damaged tissue or act against external microorganisms, interfering with their growth and activating apoptotic cell death. All of these properties have made microalgae metabolites feasible for use as immunotherapy to treat some of these disorders, as mentioned before [28,29,30].
The immune system coordinates its response via multiple signaling pathways, and these responses when unregulated contribute to the pathogenesis of chronic inflammatory skin diseases [43]. The immunomodulator mechanisms include cytokines, interferons, interleukins, and tumor necrosis factors, secreted primarily by macrophages, lymphocytes, and keratinocytes in the epidermis [28,44]. Additionally, various factors such as uncontrolled responses to pathogens, toxic substances and reactive oxygen species (ROS)-mediated oxidative stress promote the release of pro-inflammatory mediators [28,45]. Accordingly, immunomodulatory agents are used to inhibit inflammatory responses and fight diseases.
Immunomodulatory compounds such as exopolysaccharides can interact with cell surface receptors such as the Dectin-1 receptor, Toll-like receptors (TLRs), and scavenger receptors in immune cells such as macrophages, dendritic cells, neutrophils, and NK cells [4,46]. Exopolysaccharides and sulfated polysaccharides can bind to these cell surface receptors and induce signaling cascades that activate macrophage, NK cell and T/B lymphocyte activity, phagocytosis, and cytokine secretion [46,47,48,49].
In vitro and in vivo studies have shown the immunomodulatory and anti-inflammatory potential of microalgae-derived bioactive compounds and extracts [50,51,52]. In vitro studies of sulfated polysaccharides from red microalgae Porphyridium showed that the polysaccharide inhibited the movement of polymorphonuclear leukocytes, producing an anti-inflammatory effect, while in vivo studies showed that the topical application of the polysaccharides in humans inhibited the development of erythema [51]. Furthermore, immunostimulatory activity has been observed from in vitro assays involving microalgal material such as Spirulina platensis extracts, revealing an increase in human dermal fibroblast cell proliferation and enhanced wound area closure rates [52]. Table 1 lists some of the immunomodulatory compounds obtained from microalga and their mechanisms of action, reported extraction methods and culture conditions.
Studies are currently being conducted with the aim of identifying microalgal compounds responsible for immunomodulatory activities, including mainly the antimicrobial, antidiabetic, anticancer, and anti-inflammatory activities. Some of these studies are described briefly below.
The antibacterial activity of microalgae active compounds has been tested largely for the widespread microbial resistance to antibiotics in a search for new treatments against pathogens. Guzmán et al. [53] identified compounds produced by Tetraselmis suecica with antimicrobial properties, finding that an elution of 40% acetonitrile had the highest protein concentration and antibacterial activity tested against three Gram-negative and four Gram-positive bacterial strains. Likewise, in 2020 Alsenani et al. [54] screened 14 microalgae and two cyanobacterial strains to determine their antimicrobial capabilities. The fatty acid methyl esters (FAME) species found in the extracts of I. galbana, Scenedesmus sp. NT8c and Chlorella sp. FN1 showed high inhibitory activity against the growth of all six Gram-positive bacteria tested. Similarly, a study by Mukherjee et al. [55] with the hexane and chloroform extracts from Scenedesmus obliquus, tested in both Gram-positive and Gram-negative bacteria, showed the highest antibacterial activity with a minimum inhibitory concentration (MIC) of 15.6–125 μg·mL−1. In the same line with previous works, Cepas et al. [56] utilized lipids from over 600 microalgae and cyanobacteria species to determine their activities as antimicrobials and antibiofilms. The authors found that the extracts were effective against one of the three bacterial strains tested, whereas the assayed extracts from the methanol and ethyl acetate fraction reached an 80% biofilm inhibition. In addition, Potocki et al. [57] analyzed the water and ethanol extracts from Planktochlorella nurekis against three strains of Gram-positive and two strains of Gram-negative bacteria. They concluded that lauric acid, myristic acid, and stearic acid from the extracts had a high impact on the growth of Gram-negative bacteria. In addition, monounsaturated (MUFA) and polyunsaturated fatty acids (PUFA) are responsible for the modulation of Gram-positive bacteria.
Another immunomodulatory property of microalgae compounds is their anticancer activity. For instance, Marrez et al. [58] determined the antimicrobial and cytotoxicity activities of crude and fractions from the extract of Scenedesmus obliquus. The authors concluded that the microalgae diethyl ether extract had the major activity against bacteria, fungi, and three cancer cell lines, conferring to the microalgae with great potential for use in this field. Correspondingly, Peraman and Nachimuthu [59] evaluated 10 microalgal species for the production of fucoxanthin, and the extracts were tested to determine their antibacterial, antifungal, and antioxidant activities. The microalgae Dunaliella salina showed the best activity against Gram-negative bacteria and fungi, whereas for the antioxidant activity, eight of the 10 microalgal strains showed more than 50% inhibition by 2,2′-azino-bis (3-ethylbenzothiazoline-6-sulfonic acid) ABTS scavenging activity. Lastly, Lauritano et al. [60] screened 32 microalgal strains and analyzed the extracts to determine their antioxidant, anti-inflammatory, anticancer, anti-diabetes, antibacterial, and anti-biofilm activities. The authors assessed that from the microalgal species tested, three showed anti-inflammatory activity, one species had anticancer activity, two presented antibacterial activity, and the other two hindered the formation of biofilm.

3. Antioxidants

Antioxidants can be defined as “natural or synthetic substances that may prevent or delay oxidative cell damage caused by physiological oxidants having distinctly positive reduction potentials, covering ROS/reactive nitrogen species (RNS) and free radicals” [61]. This is one of the most common causes of oxidative damage due to UV exposure [62]. Lipids, nucleic acids, and proteins are ROS, RNS, and reactive sulfur species (RSS) targets [63]. UV-A damages DNA indirectly by generating ROS-like radical singlet oxygen (1O2) and/or hydrogen peroxide among others, causing DNA mutations during the replication process such as the guanine–thiamine (G–T) transversion, while UV-B directly damages DNA strands, generating cyclobutane pyrimidine dimers (CPDs) by covalently bonding two neighboring pyrimidines; their accumulation over time can produce an interruption of DNA replication and transcription, disturbing the function of the damaged cell [62,64]. Over the last two decades, studies have been conducted to find and obtain antioxidants from natural sources in order to replace synthetic sources to treat oxidative damage. Thus, various studies have focused on the protective effect of enzymatic extracts from microalgae against DNA damage induced by oxidative stress [65,66,67,68].
Antioxidants produced by microalgae are substances with high nutritional value, considering that these microorganisms have a greater ability to produce them compared to those obtained from plant-derived sources [69]. They can produce multiple components for a single species, and even if their production yield is lower than those obtained synthetically, nowadays, the generation of bioactive molecules from microalgae is expected to surpass synthetic sources, given that their production is renewable and sustainable [70]. Some of the antioxidants that can be obtained from microalgae are chlorophyll, vitamins, flavonoids, polyphenols, sterols, and carotenoids [15,63]. Table 2 lists a variety of microalgae-derived products with antioxidant properties. Carotenoids, β-carotene, chlorophyll a, chlorophyll b, and xanthophylls were the main photosynthetic pigments identified in green microalgae such as Spirogyra neglecta and Microspora indica [71]. Fucoxanthin, a bioactive microalgal compound, can be found in several species of diatoms (golden-brown microalgae) and has remarkable antioxidant, anti-obesity, anti-diabetic, anti-cancer, anti-inflammatory, anti-hypertensive, and anti-osteoporotic properties. Fucoxanthin is an important carotenoid for human health that is marketed at USD 30,000 per gram and is considered a valued nutraceutical product or functional food to prevent or help treat different diseases [72,73,74]. Recent research on the use of diatoms rich in carotenoids has been extended to space missions where resources are limited and the need to implement CO2 recycling systems provides an opportunity to produce O2 and food based on microalgal biomass as an alternative to health supplementation for humans in these conditions [69,75,76]. Additionally, in vitro experiments on human fibroblasts have reported that even small doses of infrared radiation can produce free radicals, affect collagen and elastin expression, and upregulate metalloproteinases (MMPs), representing an area of opportunity for the development of alternative supplementation to obtain antioxidants [77].
Microalgae can produce a large variety of vitamins, including, vitamins A, B1, B2, B6, B12, C and E. Minerals can also be obtained from microalgae, such as potassium, iron, magnesium, calcium, and iodine, in addition to the high protein content with complete essential amino acids [69,78]. Dunaliella tertiolecta synthesizes vitamin B12, B2, E, and beta carotene [79]. Tetraselmis suecica produces vitamin C, a ROS scavenger, and protects against lipid peroxidation via lipid hydroperoxyl radical reduction [80]. extract is rich in carotenoids (xanthophylls, lutein, violaxanthin, neoxanthin, T. suecica antheraxanthin and loroaxanthin esters), and it has shown excellent antioxidant, antiproliferative, and repairing activity in human lung cancer cell lines [81]. Pistelli et al. previously reported that Skeletonema marinoi, Cyclotella cryptica, and Nannochloropsis oceanica coculture improved bioactive compound richness as vitamins and biomass augmented the antioxidant or chemoprotective activity [82]. The Skeletonema marinoi strain has also been studied for ovothiol biosynthesis, an efficient thiohistidine compound in the scavenging of radicals and peroxides and which has only been identified in clams [83]. Astaxanthin (a red secondary carotenoid) is denominated as a super antioxidant, and most of its production is carried out synthetically (95%); while the microalgae Haematococcus pluvialis is the natural source for this excellent compound. It is present in other organisms such as plants, yeast, bacteria, seafood and other microalgae, but it is in H. pluvialis that the 3S,3′S stereoisomer is produced, and it is the most profitable of the three ((3S, 3′S); (3R, 3′S), and (3R, 3′R)) and 65 times more powerful than vitamin C and 54 times stronger than β-carotene in antioxidant activity, highlighting that the synthetic form is 20 times less effective in antioxidant quality than the natural form [84]. Glutathione scavenges electrophilic and oxidant species either in a direct way or through enzymatic catalysis: (i) by directly quenching reactive hydroxyl free radicals, other oxygen-centered free radicals, and radical centers on DNA, and (ii) as the co-substrate of glutathione peroxidase, allowing peroxide reduction [85]. However, glutathione also plays an oxidant role to a lesser extent, during GSH catabolism [85]. Mycosporine-like amino acids (MAAs) have the ability to scavenge ROS and can be of significant in the scavenging of free radicals induced by UV radiation [86].
Antioxidants are synthesized by microalgae commonly as a response to environmental stress, mainly to avoid oxidative stress [63]. The microalgae-derived antioxidants can be used in cosmeceutical applications such as moisturizers and sunscreens to prevent and treat multiple skin conditions from photoaging to skin cancer [87,88,89].
Table 2. List of compounds with antioxidant activity obtained from microalgae.
Table 2. List of compounds with antioxidant activity obtained from microalgae.
Type of CompoundMechanism of ActionMicroalgae SpeciesCulture ConditionsExtractionReference
Amino acidsROS attenuationChlamydomonas hedleyi
(Chlorophyta)
F/2 medium, 22 ± 1 °C,
80 μmol·m−2·s−1 light,
16:8 D:L cycle
rotary shaker + bubbled with CO2-enriched (1%) air.
Digestion of dry biomass with aqueous methanol (20% v/v) at 45 °C.[90]
PolysaccharidesAttenuation of free radicals, hydroxyl radicals, and ROS Arthrospira platensis
(Cyanobacteria)
Zarrouk medium, 32 ± 1 °C,
100 μmol·m−2·s−1 light, stirred with CO2-enriched (1%) air.
Tangential ultrafiltration cell—30 Kda membrane[91]
Free radical scavengingOdontella aurita
(Bacillariophyta)
Modified L1 medium (6–18 mM), 25 ± 2 °C,
100 and 300 μmol·m−2·s−1 light,
sparging with air + 1% CO2.
Digestion of dry biomass at 60 °C withsulfuric acid [92]
Sulfated polysaccharidesROS attenuationPorphyridium sp.
(Rhodophyta)
Seawater medium, 24 ± 3 °C,
150 μmol·m−2·s−1 light,
bubbled with air + 1–3% CO2.
Culture centrifugation (17,000× g, 20 min) and supernatant filtrate in dialysis tube (⌀ 2.3 cm—MW cutoff 8000)[93]
Extracellular polysaccharidesFree radical scavengingRhodella reticulata
(Rhodophyta)
Kock medium, 25 °C,
92 μmol·m−2·s−1 light.
Ethanol extraction [94]
Pigments, peptides and vitaminsROS attenuationSkeletonema marinoi
(Ochrophyta)
F/2 enriched medium, 20 °C,
150 μmol·m−2·s−1 light,
12:12 D:L cycle.
Mechanical grounding with absolute methanol[95]
PigmentsFree radical scavengingDunaliella salina
(Chlorophyta)
Johnson medium with artificial seawater (35 g L−1), pH 7.5,
100 μmol·m−2·s−1 light,
12:12 D:L cycle, bubbled with air 2 L min−1.
Sonication with methanol and filtration Fluoropore PTFE 0.2-μm[96]
ROS attenuationChlorella vulgaris
(Chlorophyta)
BG-11 medium (without and with nitrogen starvation + NaCl addition 30%), 25 °C,
150 and 1000 μmol·m−2·s−1 light,
bubbled with air.
Homogenization with acetone and supernatant filtration with Na2SO4[97]
Free radical scavengingAnabaena vaginicola
(Cyanobacteria)
BG-11 medium without nitrates,
400 μmol·m−2·s−1 light, 25 ± 2 °C,
12:12 D:L cycle
Freezing of biomass dry at −20 °C with methanol + ultrasonication [98]
No extraction methodology reported. D:L (dark:light) cycle.

4. Photoprotectors

As mentioned before, microalgae possess mechanisms against environmental challenges such as activation of several photo/dark repair mechanisms, antioxidant systems, UVR avoidance, DNA repair, and cell protection by producing UV photoprotective compounds such as mycosporine-like amino acids (MAAs), carotenoids, polyamines, and scytonemin [99].
Ectoine is a compatible solute that serves as a protective substance by acting as an osmolyte and thus helps organisms survive extreme osmotic stress, for example, with high salt content [100]. Ectoine from Halomonas elongate extracted with low salt concentration solution is an adsorbent of ultraviolet light [101,102]. Mycosporine-like amino acids that also absorb ultraviolet light can be extracted from dry biomass employing methanol to assist the extraction and chloroform to eliminate pigments from Scytonema cf. crispum [103].
The variation in different growth parameters has been associated with MAA composition in microalgae, due to improved production in direct correlation with sunlight exposure, or a variation in the MAAs’ accumulation according to the type of light [104]. As there are diverse molecular configurations and growth conditions of the harvested microorganisms, the spectrum of light that the photoprotectors can adsorb is also diverse. Photoprotection is defined as the decrease in UV radiation damage that can cause skin disease and the risk of skin cancer [105]. UV radiation comprises UVA (320–400 nm), UVB (290–320 nm), and UVC (200–290 nm). UVA causes indirect DNA damage via ROS production; it is also related to skin aging and pigmentation. UVB exposure enhances sunburns and DNA strand breaking; the pyrimidine dimer mutations associated with nonmelanoma skin cancer are also linked to UVB exposure. UVC, the UV radiation with higher energy, is completely absorbed by the ozone layer and does not represent a risk from its natural source [106].
Scientific research and new product development efforts in photoprotectors have been pursued. Some UV-resistant compounds have been investigated for their potential application in the generation of new products. Lutein, a compound that protects the skin from damage caused by UV rays, has been found in different microalgae such as C. protothecoides, Scenedesmus almeriensis, Muriellopsis sp., Neospongiococcus gelatinosum, Chlorococcum citriforme, C. zofingiensis, D. salina, and Galdieria sulphuraria [27]. The application of MAAs in the cosmetic field has also been studied previously since these compounds can absorb light between 309 and 362 nm and dissipate radiation as heat, protecting cells from mutation caused by UV-R and free radicals [107]. Some microalgae that have been reported to contain this compound include Anabaena spp., C. vulgaris, D. salina, Eutreptiella sp., Scenedesmus sp. and S. platensis. High accumulations of carotenoids in microalgae can also provide photosynthetic protection features [108]; this effect has been reported in the microalgae Nostoc sp., Eutreptiella sp., C. protothecoides, P. antarctica and P. glacialis.
There are not many reports on the application of the photoprotective properties of microalgae in cosmetic products such as sunscreens. Nevertheless, this is a finding that can generate substantial impact in the cosmetics market for UV skincare products. It has been reported that for some microalgae, MAAs also play a role as anti-desiccants that help the cell to overcome nocive effects from UV-B radiation; a study reported this behavior for the Leptolyngbya sp. cyanobacteria [109].
The study of microalgae and cyanobacteria responses against UV radiation in a combination of other conditions is an approach covered by many authors, some of them are described as follows. Singh et al. in 2020 evaluated the responses of Anabaena sp. in comparison to UVB radiation and exogenous ammonium chloride as a form of nitrogen supplementation. The study probed an interaction between NH4Cl supplementation with a protective effect, according to the photosynthetic activity, maximum quantum efficiency of PSII, and maximum electron transport rate. The MAAs also were accumulated in larger quantities in relation to NH4Cl supplementation [110]. On the other hand, C. vulgaris was evaluated for the capacity to resist different UV-B radiation intensities and times of exposure, regarding inhibition of growth by 50% at the maximum intensity and time evaluated. The study showed that the exposure of C. vulgaris to short-time periods or low-intensity levels enhanced the production of carotenoids [111]. Additionally, a comparative study of C. vulgaris, Microcoleus vaginatus, Nostoc and Scytonema javanicum distinguished the most competent microalgae with regard to UVB radiation due to ROS attenuation and their capability to repair photosystem II and DNA damages. Among the different species, Nostoc sp. was the strain capable of surviving the high levels of radiation [112].
Active compounds such as scytonemin along with MAAs are determinants that protect cyanobacteria from UV damage [113]. The biosynthesis mechanism of scytonemin has been recently elucidated in a model strain Nostoc punctiforme due to the promising antioxidant and UV protection properties that represent a potential for cosmetic and medical applications [114]. In a similar approach to Singh et al. 2020, Bennett and Soule 2022 evaluated osmotic stress and UVA and UVB exposure in the N. punctiforme strain to evaluate the expression of scytonemin genes. The study reported an increase in scytonemin gene expression in the presence of UVA, UVB and high light; however, the up-regulation of the genes did not reflect the scytonemin production [115]. Another study in 2021 was carried out to evaluate the physiological changes and scytonemin production due to UV radiation and salinity in cyanobacteria; as ROS increased, the protein and phycobiliprotein contents were reduced. The exposure to photosynthetically active radiation (UVA and UVB) and osmotic stress for at least 3 days induced scytonemin production in Scytonema sp. strain [116]. Orellana et al. in 2020 reported increases in scytonemin production in the desertic area of Atacama, showing that the indigenous endolithic cyanobacteria (Halothece) from Salar Grande suffered scytonemin reduction due to UV-A radiation under the simulated desertic conditions [117]. Table 3 shows a list of microalgae that have shown UV resistance or have been tested to exhibit UV photoprotective compounds.

5. Moisturizers, Regenerative and Other Activities

In the dermis, collagen has a dual-action mechanism providing building blocks for the formation of collagen and elastin fibers, while also binding to receptors present on fibroblasts to stimulate the production of collagen, elastin, and hyaluronan, which helps to maintain good skin appearance and elasticity and enhance its strengthening against harmful environmental factors [124]. The bioactive compounds obtained from microalgae sources have proven to possess properties like the ones presented in this paper, providing a range of applications that vary widely depending on the industrial and commercial sector of interest [125]. Expressly, bioactive compounds from microalgae are currently being incorporated into the cosmetic and cosmeceutical industry as they can improve and maintain the structure and function of the skin [1].
Exopolysaccharides from microalgae can be rheology modifiers, conditioners, moisturizing agents, healing agents, emulsifiers, and substitutes for hyaluronic acids. In addition, they can stimulate collagen synthesis and provide protective activities against enzymatic proteolysis [31,32,33,124]. Proteins found in microalgae used as food supplements can stimulate collagen synthesis, leading to a reduction in vascular imperfections and promoting tissue regeneration [1,33,126,127,128]. Additionally, some microalgae species are known to produce compounds that can act as substitutes for hyaluronic acids or protect against enzymatic proteolysis [1,31,32,33,126,127,128,129]. Table 4 shows some of the main molecules obtained from microalgae cultures that could be implemented as moisturizers and promote putative regenerative activity on the skin.
Furthermore, chlorophyll a and b can be used as dyeing agents and also as additive agents that can be used to mask odors in formulations [39,130]. β-carotene, phycoerythrin, phycocyanin, allophycocyanin, phycoerythrocyanin, astaxanthin, lutein, lycopene, and violaxanthin can be used as dyeing agents in the textile industry and in cosmetics as greener alternatives [1,130,131].
Maintaining correct skin moisturization is the first step to aiming for a strong defense mechanism against irritant agents and tensioactive materials. It has been clear that during the actual COVID-19 pandemic emergency, several alcohol-based products released as sanitizers also help to reduce the transmission of the virus. Nonetheless, those products often need to be supplemented with humectants, emollients, or moisturizers to avoid the effects of alcohol, which principally affect the epidermal layer, causing dehydration. Although dry skin does not represent a serious problem, it can lead to complications in the management of dermatological infections [132,133].
Table 4. List of compounds with moisturizing and regenerative activity obtained from microalgae.
Table 4. List of compounds with moisturizing and regenerative activity obtained from microalgae.
Type of CompoundMechanism of ActionMicroalgae SpeciesCulture ConditionsExtractionReference
CollagenStimulation of collagen synthesis in the skinC. vulgarisNRMethanol extraction [134]
Regeneration by stimulation of collagen synthesis in the skinS. platensisZarrouk’s medium, pH 9.8–10.0, 25 °C,
46 μmol m−2·s−1 light
Raw biomass[135]
Mycosporine-2-glycineCollagenase inhibition and glycation products inhibitionAphanothece halophyticaBG-11Methanol extraction and mechanical disruption[136]
PolysaccharidesMoisturizing agentsPediastrum duplexMIII medium, pH 7.9 ± 0.1
20 ± 0.5°C, 80 µmol m−2·s−1 light, 12:12 D:L
Methanol extraction[137]
Regulate water distribution in the skinCodium tomentosumNRNR[138]
Moisturizing agentsUndaria pinnatifaNRNR[138]
Moisturizing agentsDurvillea antarcticaNRNR[138]
Moisturizing agentsCladosiphon okamuranusNRNR[138]
Fat acidsDermal collagen content rescueS. rubescensNRHydrophilic hot extrusion[139]
Amino acidsIncreases expression of
Procollagen C Proteinase
C. hedleyiF/2, pH 8–9, 22 °C,
80 μmol m−2·s−1 light, 8:16 D:L cycle, 1% CO2
Raw biomass[91]
TrpA proteinCollagen-like proteinTrichodesmium erythraeumNRNR[11]
D:L = (dark:light) cycle. NR = Not Reported.

6. Novel Vehicles or Excipient Compounds

Some microalgae-derived molecules can serve as vehicles or excipients for other molecules of interest. Proteins from microalgae are molecules of commercial interest; they can be used in the food industry, pharmaceutical, and cosmetic sectors, or more specifically, in nutraceuticals and cosmeceuticals. They can be used as emulsifying, foaming, thickening, and gelling agents [15]; these functions are mainly explored in the food sector but could potentially be extrapolated to other sectors. C. vulgaris and Tetraselmis sp. have emulsifying and foaming properties. Thickening or gelling properties have also been reported for Arthrospira platensis, H. pluvialis and T. suecica proteins [140]. However, this application field is recent; therefore, not many studies have been conducted in this area.
Extracellular polysaccharides (EPSs) are secreted by various marine organisms, including microalgae. They represent a physical barrier protecting cells from harmful agents and environmental constraints and serve in different physiological processes including cell adhesion, cell interaction, and biofilm formation. The protective capacity of microalgae EPSs has been reported also in the environment, where diatoms delay the degradation of materials through the secretion of EPSs that allows adhesion to the material superfice and create a barrier between the material and the corrosive environment. This behavior can be extrapolated to the potential use of these biomolecules in mixtures with topical bioactive compounds for slow release into the skin or for the improvement of the skin protection capacity in commercial products. Other areas of interest for the use of EPSs are the pharmaceutical and medical industries, where properties such as viscosity, porosity, and absorption capacity of EPSs produced by microalgae and cyanobacteria can be applied for the development of different products such as drug delivery hydrogels, tissue engineering, occlusion devices, and bioadhesives for active release, among others [141]. One of the main strategies to improve EPS applications is the management of EPS rheological properties, which have been reported to depend on EPS concentration. In this sense, the characteristics can be controlled in order to allow their use in different applications as active texturizing agents, carriers, or as active protectors and enhancers [142]. Furthermore, EPSs have been reported as important emulsion stabilizers, flocculants, foam stabilizers, and hydrating agents (tested in cosmetics and pharmaceuticals) [143]. According to the literature reviewed, the applications of microalgae EPSs as vehicles or excipient compounds showed interesting characteristics to impact a wide range of options. Table 4 is lists compounds obtained from microalgae with applications in topical formulations.
Extracellular vesicles (EVs) are membrane-surrounded structures released in the extracellular milieu by every cell type. EVs contain proteins, lipids, nucleic acids, and other metabolites [144]. In addition, EVs possess many attributes of a drug delivery vehicle of interest to the cosmeceutical field. A newly discovered subtype of EVs derived from microalgae, named nanoalgosomes, are extracellular nano-objects from cultures of microalgal strains. Nanoalgosomes are novel membranous biogenic nanomaterials refined for the first time from a sustainable and renewable bioresource (microalgae), which can be used as a new natural delivery system for high-value microalgal substances (such as antioxidants, pigments, lipids, and complex carbohydrates), bioactive biological molecules and/or synthetic drugs. EV-like nanoparticles have been isolated from several microalgae strains such as T. chuii and D. tertiolecta [145], suggesting that other microalgae may contain EV-like nanoparticles and have a major impact on drug delivery in the cosmeceutical field.
Research has been conducted regarding the potential of porous silica-based particles for drug delivery applications due to their extended drug release profiles and high efficacy in delivering hydrophobic drugs [146]. These silica-based nanoparticles obtained from diatoms are used as drug delivery carriers due to their biodegradability, easy functionalization, low cost to obtain and maintain, and simple features compared to the synthetic ones, which make these agents proper alternatives to synthetic silica nanoparticles [147].
These findings demonstrate the potential of microalgae as excipients and vehicles for molecules or drugs, implying great potential for application and a breakthrough in the search for more natural alternatives in cosmeceuticals.

7. Perspectives

Currently, microalgae cultures as a source of high-value biomolecules have become more attractive due to their potential application in the development of technologies to mitigate greenhouse gas emissions, which agrees with the sustainable development goals of the United Nations. Nevertheless, most of the bibliography consulted microalgae biomass is cultivated using synthetic and controlled media. In this sense, future research should be focused on the evaluation and optimization of processes that allow the fixation of CO2 from different sources at the same time that microalgae produce molecules of interest. Actually, the obtention of molecules from microalgae biomass have been widely studied, and reported data showed a dependence of the produced metabolite profiles on the microalgae strains and different stress conditions (light intensity, temperature, pH, nutrients concentrations, and others). For future research, the authors identified a clear procedure for the obtention of biomolecules from microalgae that can be described in four main steps: (1) microalgae strain selection, (2) standardization of growth conditions, (3) biomolecule characterization, and (4) extraction and functionalization (Figure 2). The photobioreactor design, moreover, represents an area of opportunity to improve those processes, to guarantee the light incidence to the microalgae culture, the retention time for gasses, and small spaces for technology implementation in established industries. Finally, the extraction and purification process represents a challenge due to the different nature of those metabolites of interest and the characteristics of the different microalgae. Developing simpler extraction protocols is also an area for improvement. However, efforts should be oriented to the implementation of processes where microalgal compounds are more accessible and can be used directly from raw biomass or biomass that requires minimal processing.
Along with microalgae, macroalgae (not explored in this work) also represent an alternative source of molecules of interest to improve functionality and extend the time of action. Furthermore, biocompatibility analysis of those metabolites must be characterized to ensure their safe use as a cosmeceutical components.
Unfortunately, most of the biotechnology research only focuses on biomolecule extraction with certain putative activity from microalgae cultures, as shown by some of the data presented in this work. Nevertheless, microalgae can produce a wide variety of biomolecules that can be approached in different areas, so there is still work to be done to bring biodegradable microalgae-based products into general use.
The investigation of the presence of cosmeceutical compounds in microalgae has been widely reported; however, despite the existence of records on the extraction and evaluation of the potential of these compounds, the information regarding their application and distribution in the market remains scarce, which could be an indicator of the lack of scaling of this practice from a laboratory test to its commercial application. It is necessary to evaluate strategies that allow scaling the use of these biomolecules, and to identify barriers to this process. Is it the quality of the obtained biomolecules? Is it a scarce market for the product? Does misinformation about the advantages and disadvantages exist? Or, are there limitations on access to economic resources? Regardless of the answer, the next steps for the use of microalgae bioactive compounds should consider a market study that allows the materialization of these areas of opportunity.

8. Conclusions

The cosmeceutical industry has been expanding over the years and making an effort to use active ingredients from natural sources. Microalgae are a source of added-value functional compounds that can be ultimately a healthier option. Moreover, microalgae-derived metabolites can be obtained through low-cost, eco-friendly cultivation processes. The application of microalgae as a source of bioactive molecules has been successfully evaluated by different authors. In the process of taking advantage of the potential of microalgae-derived compounds to develop and improve a new generation of cosmeceutical products, it is important to consider the microalgae strain, growth conditions, biomass characterization, and protocols for extraction and functionalization. Those cosmeceutical products enriched with photoprotectors, antioxidants, immunomodulators, and moisturizers/regenerative molecules extracted from different microalgae species suggest an improvement in the potential of different cosmeceutical product characteristics with novel advantages in terms of environmental impact.
Most of the actual studies develop microalgae growth under specific stress conditions to stimulate the production of the metabolites of interest for posterior extraction and use in the development and enrichment of cosmeceutical products. Nevertheless, the direct application of raw or dry microalgae in the process of product improvement is still scarce; hence, the use of microalgae as a source of bioactive molecules still has areas of opportunity for future research, focused on the method of approach and its application to commercial cosmeceutical products.

Author Contributions

Conceptualization, M.M.-R., C.A.M.-G. and D.-H.K.; methodology, H.R.-P., E.R.M.-S. and K.R.V.-Z.; investigation, C.A.M.-G., D.-H.K. and K.G.C.-A.; data curation, E.R.M.-S. and K.R.V.-Z.; writing—original draft preparation, M.M.-R., C.A.M.-G., D.-H.K., B.S.-R. and K.G.C.-A.; writing—review and editing, A.L.D.-Z., D.R.-G., K.R.V.-Z., E.R.M.-S. and A.M.G.-M.; supervision, J.E.S.-H. and A.M.G.-M.; project administration, R.P.-S. and H.M.N.I.; funding acquisition, R.P.-S. and H.M.N.I. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by Consejo Nacional de Ciencia y Tecnología (CONACYT) master fellowship No. 1104949 awarded to author B.S.-R. and partially supported under Sistema Nacional de Investigadores (SNI) program awarded to Manuel Martínez Ruiz (CVU: 418151), Juan Eduardo Sosa-Hernández (CVU: 375202), Roberto Parra-Saldívar (CVU: 35753) and Hafiz M.N. Iqbal (CVU: 735340).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Figures were created with BioRender.com.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mourelle, M.L.; Gómez, C.P.; Legido, J.L. The Potential Use of Marine Microalgae and Cyanobacteria in Cosmetics and Thalassotherapy. Cosmetics 2017, 4, 46. [Google Scholar] [CrossRef] [Green Version]
  2. Vermeer, B.J.; Gilchrest, B.A. Cosmeceuticals: A proposal for rational definition, evaluation, and regulation. Arch. Dermatol. 1996, 132, 337–340. [Google Scholar] [CrossRef] [PubMed]
  3. Kligman, A. The future of cosmeceuticals: An interview with Albert Kligman, MD, PhD. Interview by Zoe Diana Draelos. Dermatol. Surg. Off. Publ. Am. Soc. Dermatol. Surg. 2005, 31, 890–891. [Google Scholar]
  4. Pandey, A.; Jatana, G.K.; Sonthalia, S. Cosmeceuticals; StatPearls Publishing: Treasure Island, FL, USA, 2021. [Google Scholar]
  5. What is the Demand for Natural Ingredients for Cosmetics on the European Market?|CBI. Available online: https://www.cbi.eu/market-information/natural-ingredients-cosmetics/what-demand (accessed on 23 January 2022).
  6. Spolaore, P.; Joannis-Cassan, C.; Duran, E.; Isambert, A. Commercial applications of microalgae. J. Biosci. Bioeng. 2006, 101, 87–96. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Chakdar, H.; Jadhav, S.D.; Dhar, D.W.; Pabbi, S. Potential applications of blue green algae. J. Sci. Ind. Res. 2012, 71, 13–20. [Google Scholar]
  8. Stolz, P.; Obermayer, B. Manufacturing microalgae for skin care. Cosmet. Toilet. 2005, 120, 99–106. [Google Scholar]
  9. Gauthier, M.R.; Senhorinho, G.N.A.; Scott, J.A. Microalgae under environmental stress as a source of antioxidants. Algal Res. 2020, 52, 102104. [Google Scholar] [CrossRef]
  10. 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]
  11. Wu, J.; Gu, X.; Yang, D.; Xu, S.; Wang, S.; Chen, X.; Wang, Z. Bioactive substances and potentiality of marine microalgae. Food Sci. Nutr. 2021, 9, 5279–5292. [Google Scholar] [CrossRef]
  12. Rath, B. Microalgal Bioremediation:Current Practices and Perspectives. J. Biochem. Technol. 2011, 3, 299–304. [Google Scholar]
  13. Patel, A.; Matsakas, L.; Rova, U.; Christakopoulos, P. A perspective on biotechnological applications of thermophilic microalgae and cyanobacteria. Bioresour. Technol. 2019, 278, 424–434. [Google Scholar] [CrossRef] [PubMed]
  14. Sartori, R.B.; Siqueira, S.F.; Maroneze, M.M.; Fagundes, M.B.; Wagner, R.; Zepka, L.Q.; Jacob-Lopes, E. Microalgal secondary metabolites: Effect of climatic variables, seasons, and photocycles on the biogeneration of volatile organic compounds (VOCs). J. Appl. Phycol. 2021, 33, 1457–1472. [Google Scholar] [CrossRef]
  15. Levasseur, W.; Perré, P.; Pozzobon, V. A review of high value-added molecules production by microalgae in light of the classification. Biotechnol. Adv. 2020, 41, 107545. [Google Scholar] [CrossRef]
  16. Al-Jabri, H.; Das, P.; Khan, S.; Thaher, M.; AbdulQuadir, M. Treatment of Wastewaters by Microalgae and the Potential Applications of the Produced Biomass—A Review. Water 2021, 13, 27. [Google Scholar] [CrossRef]
  17. Ummalyma, S.B.; Sahoo, D.; Pandey, A. Resource recovery through bioremediation of wastewaters and waste carbon by microalgae: A circular bioeconomy approach. Environ. Sci. Pollut. Res. 2021, 28, 58837–58856. [Google Scholar] [CrossRef] [PubMed]
  18. Khanzada, Z.T. Phosphorus removal from landfill leachate by microalgae. Biotechnol. Rep. 2020, 25, e00419. [Google Scholar] [CrossRef] [PubMed]
  19. Abdel-Raouf, N.; Al-Homaidan, A.A.; Ibraheem, I.B.M. Microalgae and wastewater treatment. Saudi J. Biol. Sci. 2012, 19, 257–275. [Google Scholar] [CrossRef] [Green Version]
  20. Wang, L.; Min, M.; Li, Y.; Chen, P.; Chen, Y.; Liu, Y.; Wang, Y.; Ruan, R. Cultivation of Green Algae Chlorella sp. in Different Wastewaters from Municipal Wastewater Treatment Plant. Appl. Biochem. Biotechnol. 2010, 162, 1174–1186. [Google Scholar] [CrossRef]
  21. Martínez-Ruiz, M.; Molina-Vázquez, A.; Santiesteban-Romero, B.; Reyes-Pardo, H.; Villaseñor-Zepeda, K.R.; Meléndez-Sánchez, E.R.; Araújo, R.G.; Sosa-Hernández, J.E.; Bilal, M.; Iqbal, H.M.N.; et al. Micro-algae assisted green bioremediation of water pollutants rich leachate and source products recovery. Environ. Pollut. 2022, 306, 119422. [Google Scholar] [CrossRef]
  22. Guldhe, A.; Kumari, S.; Ramanna, L.; Ramsundar, P.; Singh, P.; Rawat, I.; Bux, F. Prospects, recent advancements and challenges of different wastewater streams for microalgal cultivation. J. Environ. Manag. 2017, 203, 299–315. [Google Scholar] [CrossRef]
  23. Cho, S.; Lee, N.; Park, S.; Yu, J.; Luong, T.T.; Oh, Y.-K.; Lee, T. Microalgae cultivation for bioenergy production using wastewaters from a municipal WWTP as nutritional sources. Bioresour. Technol. 2013, 131, 515–520. [Google Scholar] [CrossRef] [PubMed]
  24. Zhou, W.; Min, M.; Li, Y.; Hu, B.; Ma, X.; Cheng, Y.; Liu, Y.; Chen, P.; Ruan, R. A hetero-photoautotrophic two-stage cultivation process to improve wastewater nutrient removal and enhance algal lipid accumulation. Bioresour. Technol. 2012, 110, 448–455. [Google Scholar] [CrossRef] [PubMed]
  25. Mat Aron, N.S.; Khoo, K.S.; Chew, K.W.; Veeramuthu, A.; Chang, J.-S.; Show, P.L. Microalgae cultivation in wastewater and potential processing strategies using solvent and membrane separation technologies. J. Water Process Eng. 2021, 39, 101701. [Google Scholar] [CrossRef]
  26. Tan, J.S.; Lee, S.Y.; Chew, K.W.; Lam, M.K.; Lim, J.W.; Ho, S.-H.; Show, P.L. A review on microalgae cultivation and harvesting, and their biomass extraction processing using ionic liquids. Bioengineered 2020, 11, 116–129. [Google Scholar] [CrossRef] [Green Version]
  27. Yarkent, Ç.; Gürlek, C.; Oncel, S.S. Potential of microalgal compounds in trending natural cosmetics: A review. Sustain. Chem. Pharm. 2020, 17. [Google Scholar] [CrossRef]
  28. Riccio, G.; Lauritano, C. Microalgae with Immunomodulatory Activities. Mar. Drugs 2019, 18, 2. [Google Scholar] [CrossRef] [Green Version]
  29. Tabarzad, M.; Atabaki, V.; Hosseinabadi, T. Anti-inflammatory Activity of Bioactive Compounds from Microalgae and Cyanobacteria by Focusing on the Mechanisms of Action. Mol. Biol. Rep. 2020, 47, 6193–6205. [Google Scholar] [CrossRef]
  30. Fu, W.; Nelson, D.R.; Yi, Z.; Xu, M.; Khraiwesh, B.; Jijakli, K.; Chaiboonchoe, A.; Alzahmi, A.; Al-Khairy, D.; Brynjolfsson, S.; et al. Bioactive Compounds From Microalgae: Current Development and Prospects. In Studies in Natural Products Chemistry; Rahman, A., Ed.; Elsevier: Amsterdam, The Netherlands, 2017; Volume 54, pp. 199–225. [Google Scholar] [CrossRef]
  31. Bellou, S.; Baeshen, M.N.; Elazzazy, A.M.; Aggeli, D.; Sayegh, F.; Aggelis, G. Microalgal lipids biochemistry and biotechnological perspectives. Biotechnol. Adv. 2014, 32, 1476–1493. [Google Scholar] [CrossRef]
  32. Martins, A.; Vieira, H.; Gaspar, H.; Santos, S. Marketed marine natural products in the pharmaceutical and cosmeceutical industries: Tips for success. Mar. Drugs 2014, 12, 1066–1101. [Google Scholar] [CrossRef] [Green Version]
  33. Berthon, J.-Y.; Kappes, R.N.; Bey, M.; Cadoret, J.-P.; Renimel, I.; Filaire, E. Marine algae as attractive source to skin care. Free Radic. Res. 2017, 51, 555–567. [Google Scholar] [CrossRef]
  34. Majtan, J.; Jesenak, M. β-Glucans: Multi-functional modulator of wound healing. Molecules 2018, 23, 806. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Raposo, M.F.D.J.; de Morais, R.M.S.C.; de Morais, A.M.M.B. Bioactivity and applications of sulphated polysaccharides from marine microalgae. Mar. Drugs 2013, 11, 233–252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Fujitani, N.; Sakaki, S.; Yamaguchi, Y.; Takenaka, H. Inhibitory effects of microalgae on the activation of hyaluronidase. J. Appl. Phycol. 2001, 13, 489–492. [Google Scholar] [CrossRef]
  37. Pulz, O.; Gross, W. Valuable products from biotechnology of microalgae. Appl. Microbiol. Biotechnol. 2004, 65, 635–648. [Google Scholar] [CrossRef]
  38. Guzmán, S.; Gato, A.; Calleja, J.M. Antiinflammatory, analgesic and free radical scavenging activities of the marine microalgae Chlorella stigmatophora and Phaeodactylum tricornutum. Phytother. Res. 2001, 15, 224–230. [Google Scholar] [CrossRef]
  39. Borowitzka, M.A. Microalgae as sources of pharmaceuticals and other biologically active compounds. J. Appl. Phycol. 1995, 7, 3–15. [Google Scholar] [CrossRef]
  40. Jovanović, M.; Golušin, Z. Nonsteroidal Topical Immunomodulators in Allergology and Dermatology. Biomed. Res. Int. 2016, 2016, 5185303. [Google Scholar] [CrossRef] [Green Version]
  41. Meykadeh, N.; Hengge, U.R. Topische immunmodulation in der dermatologie. Der Hautarzt 2003, 50, 641–662. [Google Scholar] [CrossRef]
  42. Hengge, U.R.; Benninghoff, B.; Ruzicka, T.; Goos, M. Topical immunomodulators--progress towards treating inflammation, infection, and cancer. Lancet Infect. Dis. 2001, 1, 189–198. [Google Scholar] [CrossRef]
  43. Pasparakis, M.; Haase, I.; Nestle, F.O. Mechanisms regulating skin immunity and inflammation. Nat. Rev. Immunol. 2014, 14, 289–301. [Google Scholar] [CrossRef]
  44. Sauder, D. The Role of Epidermal Cytokines in Inflammatory Skin Diseases. J. Investig. Dermatol. 1990, 95, 275–285. [Google Scholar] [CrossRef] [PubMed]
  45. Choo, W.-T.; Teoh, M.-L.; Phang, S.-M.; Convey, P.; Yap, W.-H.; Goh, B.-H.; Beardall, J. Microalgae as Potential Anti-Inflammatory Natural Product Against Human Inflammatory Skin Diseases. Front. Pharmacol. 2020, 11. [Google Scholar] [CrossRef] [PubMed]
  46. Kim, H.S.; Hong, J.T.; Kim, Y.; Han, S.-B. Stimulatory Effect of β-glucans on Immune Cells. Immune Netw. 2011, 11, 191–195. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Brown, G.D. Dectin-1: A signalling non-TLR pattern-recognition receptor. Nat. Rev. Immunol. 2006, 6, 33–43. [Google Scholar] [CrossRef]
  48. Huang, L.; Shen, M.; Morris, G.A.; Xie, J. Sulfated polysaccharides: Immunomodulation and signaling mechanisms. Trends Food Sci. Technol. 2019, 92, 1–11. [Google Scholar] [CrossRef]
  49. Sun, L.; Zhao, Y. The biological role of dectin-1 in immune response. Int. Rev. Immunol. 2007, 26, 349–364. [Google Scholar] [CrossRef]
  50. Miguel, S.P.; Ribeiro, M.P.; Otero, A.; Coutinho, P. Application of microalgae and microalgal bioactive compounds in skin regeneration. Algal Res. 2021, 58, 102395. [Google Scholar] [CrossRef]
  51. Matsui, M.S.; Muizzuddin, N.; Arad, S.; Marenus, K. Sulfated polysaccharides from red microalgae have antiinflammatory properties In Vitro and In Vivo. Appl. Biochem. Biotechnol. 2003, 104, 13–22. [Google Scholar] [CrossRef]
  52. Syarina, P.N.A.; Karthivashan, G.; Abas, F.; Arulselvan, P.; Fakurazi, S. Wound healing potential of Spirulina platensis extracts on human dermal fibroblast cells. EXCLI J. 2015, 14, 385–393. [Google Scholar] [CrossRef]
  53. Guzmán, F.; Wong, G.; Román, T.; Cárdenas, C.; Alvárez, C.; Schmitt, P.; Albericio, F.; Rojas, V. Identification of Antimicrobial Peptides from the Microalgae Tetraselmis suecica (Kylin) Butcher and Bactericidal Activity Improvement. Mar. Drugs 2019, 17, 453. [Google Scholar] [CrossRef] [Green Version]
  54. Alsenani, F.; Tupally, K.R.; Chua, E.T.; Eltanahy, E.; Alsufyani, H.; Parekh, H.S.; Schenk, P.M. Evaluation of microalgae and cyanobacteria as potential sources of antimicrobial compounds. Saudi Pharm. J. 2020, 28, 1834–1841. [Google Scholar] [CrossRef] [PubMed]
  55. Mukherjee, C.; Suryawanshi, P.G.; Kalita, M.C.; Deka, D.; Aranda, D.A.G.; Goud, V.V. Polarity-wise successive solvent extraction of Scenedesmus obliquus biomass and characterization of the crude extracts for broad-spectrum antibacterial activity. Biomass Convers. Biorefin. 2022. [Google Scholar] [CrossRef]
  56. Cepas, V.; Gutiérrez-Del-Río, I.; López, Y.; Redondo-Blanco, S.; Gabasa, Y.; Iglesias, M.J.; Soengas, R.; Fernández-Lorenzo, A.; López-Ibáñez, S.; Villar, C.J.; et al. Microalgae and Cyanobacteria Strains as Producers of Lipids with Antibacterial and Antibiofilm Activity. Mar. Drugs 2021, 19, 675. [Google Scholar] [CrossRef] [PubMed]
  57. Potocki, L.; Oklejewicz, B.; Kuna, E.; Szpyrka, E.; Duda, M.; Zuczek, J. Application of Green Algal Planktochlorella nurekis Biomasses to Modulate Growth of Selected Microbial Species. Molecules 2021, 26, 4038. [Google Scholar] [CrossRef]
  58. Marrez, D.A.; Naguib, M.M.; Sultan, Y.; Higazy, A.M. Antimicrobial and anticancer activities of Scenedesmus obliquus metabolites. Heliyon 2019, 5, e01404. [Google Scholar] [CrossRef] [Green Version]
  59. Chester, K.; Zahiruddin, S.; Ahmad, A.; Khan, W.; Paliwal, S.; Ahmad, S. Bioautography-based identification of antioxidant me-tabolites of Solanum nigrum L. and exploration its hepatoprotective potential against D-galactosamine-induced hepatic fi-brosis in rats. Phcog. Mag. 2019, 15 (Suppl. S1), 104–110. [Google Scholar]
  60. Lauritano, C.; Andersen, J.H.; Hansen, E.; Albrigtsen, M.; Escalera, L.; Esposito, F.; Helland, K.; Hanssen, K.Ø.; Romano, G.; Ianora, A. Bioactivity Screening of Microalgae for Antioxidant, Anti-Inflammatory, Anticancer, Anti-Diabetes, and Antibacterial Activities. Front. Mar. Sci. 2016, 3, 68. [Google Scholar] [CrossRef] [Green Version]
  61. Apak, R.; Özyürek, M.; Güçlü, K.; Çapanoğlu, E. Antioxidant Activity/Capacity Measurement. 1. Classification, Physicochemical Principles, Mechanisms, and Electron Transfer (ET)-Based Assays. J. Agric. Food Chem. 2016, 64, 997–1027. [Google Scholar] [CrossRef]
  62. de Jager, T.L.; Cockrell, A.E.; Du Plessis, S.S. Ultraviolet Light Induced Generation of Reactive Oxygen Species. In Ultraviolet Light in Human Health, Diseases and Environment; Springer: Cham, Switzerland, 2017; Volume 996, pp. 15–23. [Google Scholar] [CrossRef]
  63. Sansone, C.; Brunet, C. Promises and Challenges of Microalgal Antioxidant Production. Antioxidants 2019, 8, 199. [Google Scholar] [CrossRef] [Green Version]
  64. Vechtomova, Y.L.; Telegina, T.A.; Buglak, A.A.; Kritsky, M.S. UV radiation in DNA damage and repair involving DNA-photolyases and cryptochromes. Biomedicines 2021, 9, 1564. [Google Scholar] [CrossRef]
  65. Goiris, K.; Muylaert, K.; Fraeye, I.; Foubert, I.; De Brabanter, J.; De Cooman, L. Antioxidant potential of microalgae in relation to their phenolic and carotenoid content. J. Appl. Phycol. 2012, 24, 1477–1486. [Google Scholar] [CrossRef]
  66. Lekshmi, S.; Saramma, A.V. Antioxidant activity of synechococcus sp. Nägeli isolated from cochin estuary, India. Indian J. Geo-Mar. Sci. 2018, 47, 2213–2216. [Google Scholar]
  67. Foo, S.C.; Yusoff, F.M.; Ismail, M.; Basri, M.; Yau, S.K.; Khong, N.M.H.; Chan, K.W.; Ebrahimi, M. Antioxidant capacities of fucoxanthin-producing algae as influenced by their carotenoid and phenolic contents. J. Biotechnol. 2017, 241, 175–183. [Google Scholar] [CrossRef] [PubMed]
  68. Jerez-Martel, I.; García-Poza, S.; Rodríguez-Martel, G.; Rico, M.; Afonso-Olivares, C.; Gómez-Pinchetti, J.L. Phenolic Profile and Antioxidant Activity of Crude Extracts from Microalgae and Cyanobacteria Strains. J. Food Qual. 2017, 2017, 2924508. [Google Scholar] [CrossRef] [Green Version]
  69. Koyande, A.K.; Chew, K.W.; Rambabu, K.; Tao, Y.; Chu, D.-T.; Show, P.-L. Microalgae: A potential alternative to health supplementation for humans. Food Sci. Hum. Wellness 2019, 8, 16–24. [Google Scholar] [CrossRef]
  70. Gouveia, L.; Batista, A.P.; Sousa, I.; Raymundo, A.; Bandarra, N.M. Microalgae in Novel Food Products; Nova Science Publishers: New York, NY, USA, 2009; ISBN 9781606920084. [Google Scholar]
  71. Aditi, N.; Jaishini, C.; Raisa, F.; Aamna, K.; Rupali, B. Photosynthetic pigments, lipids and phenolic compounds of three green algae isolated from freshwater ecosystem. J. Algal Biomass Util. 2020, 11, 68–83. [Google Scholar]
  72. Seth, K.; Kumar, A.; Rastogi, R.P.; Meena, M.; Vinayak, V. Harish Bioprospecting of fucoxanthin from diatoms—Challenges and perspectives. Algal Res. 2021, 60, 102475. [Google Scholar] [CrossRef]
  73. Su, J.; Guo, K.; Huang, M.; Liu, Y.; Zhang, J.; Sun, L.; Li, D.; Pang, K.-L.; Wang, G.; Chen, L.; et al. Fucoxanthin, a Marine Xanthophyll Isolated From Conticribra weissflogii ND-8: Preventive Anti-Inflammatory Effect in a Mouse Model of Sepsis. Front. Pharmacol. 2019, 10, 906. [Google Scholar] [CrossRef] [Green Version]
  74. Zarekarizi, A.; Hoffmann, L.; Burritt, D. Approaches for the sustainable production of fucoxanthin, a xanthophyll with potential health benefits. J. Appl. Phycol. 2019, 31, 281–299. [Google Scholar] [CrossRef]
  75. Vinayak, V. Algae as sustainable food in space missions. In Biomass, Biofuels, Biochemicals; ELSIVIER: Alpharetta, GA, USA, 2021; pp. 517–540. ISBN 9780323898553. [Google Scholar]
  76. Cycil, L.M.; Hausrath, E.M.; Ming, D.W.; Adcock, C.T.; Raymond, J.; Remias, D.; Ruemmele, W.P. Investigating the Growth of Algae Under Low Atmospheric Pressures for Potential Food and Oxygen Production on Mars. Front. Microbiol. 2021, 12, 199–225. [Google Scholar] [CrossRef]
  77. Aziz, J.; Shezali, H.; Radzi, Z.; Yahya, N.A.; Abu Kassim, N.H.; Czernuszka, J.; Rahman, M.T. Molecular Mechanisms of Stress-Responsive Changes in Collagen and Elastin Networks in Skin. Ski. Pharmacol. Physiol. 2016, 29, 190–203. [Google Scholar] [CrossRef] [PubMed]
  78. Becker, E.W. Microalgae for Human and Animal Nutrition. In Handbook of Microalgal Culture: Biotechnology and Applied Phycology; Wiley Online Library: Hoboken, NJ, USA, 2004; pp. 461–503. [Google Scholar]
  79. Galasso, C.; Gentile, A.; Orefice, I.; Ianora, A.; Bruno, A.; Noonan, D.M.; Sansone, C.; Albini, A.; Brunet, C. Microalgal Derivatives as Potential Nutraceutical and Food Supplements for Human Health: A Focus on Cancer Prevention and Interception. Nutrients 2019, 11, 1226. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Traber, M.G.; Stevens, J.F. Vitamins C and E: Beneficial effects from a mechanistic perspective. Free Radic. Biol. Med. 2011, 51, 1000–1013. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Sansone, C.; Galasso, C.; Orefice, I.; Nuzzo, G.; Luongo, E.; Cutignano, A.; Romano, G.; Brunet, C.; Fontana, A.; Esposito, F.; et al. The green microalga Tetraselmis suecica reduces oxidative stress and induces repairing mechanisms in human cells. Sci. Rep. 2017, 7, 41215. [Google Scholar] [CrossRef]
  82. Pistelli, L.; Del Mondo, A.; Smerilli, A.; Corato, F.; Piscitelli, C.; Pellone, P.; Carbone, D.A.; Sansone, C.; Brunet, C. Microalgal co-cultivation prospecting to modulate vitamin and bioactive compounds production. Antioxidants 2021, 10, 1360. [Google Scholar] [CrossRef] [PubMed]
  83. Milito, A.; Orefice, I.; Smerilli, A.; Castellano, I.; Napolitano, A.; Brunet, C.; Palumbo, A. Insights into the light response of skeletonema marinoi: Involvement of ovothiol. Mar. Drugs 2020, 18, 477. [Google Scholar] [CrossRef]
  84. 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. [Google Scholar] [CrossRef] [Green Version]
  85. Gaucher, C.; Boudier, A.; Bonetti, J.; Clarot, I.; Leroy, P.; Parent, M. Glutathione: Antioxidant properties dedicated to nanotechnologies. Antioxidants 2018, 7, 62. [Google Scholar] [CrossRef] [Green Version]
  86. Kageyama, H.; Waditee-Sirisattha, R. Antioxidative, anti-inflammatory, and anti-aging properties of mycosporine-like amino acids: Molecular and cellular mechanisms in the protection of skin-aging. Mar. Drugs 2019, 17, 222. [Google Scholar] [CrossRef] [Green Version]
  87. Guedes, A.C.; Amaro, H.M.; Malcata, F.X. Microalgae as sources of high added-value compounds—A brief review of recent work. Biotechnol. Prog. 2011, 27, 597–613. [Google Scholar] [CrossRef]
  88. Alvares, B.A.; Miola, A.C.; Schimitt, J.V.; Miot, H.A.; Abbade, L.P.F. Efficacy of sunscreen with photolyase or regular sunscreen associated with topical antioxidants in treating advanced photodamage and cutaneous field cancerization: A randomized clinical trial. An. Bras. Dermatol. 2022, 97, 157–165. [Google Scholar] [CrossRef] [PubMed]
  89. Petruk, G.; Del Giudice, R.; Rigano, M.M.; Monti, D.M. Antioxidants from plants protect against skin photoaging. Oxid. Med. Cell. Longev. 2018, 2018, 1454936. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Gouveia, L.; Raymundo, A.; Batista, A.P.; Sousa, I.; Empis, J. Chlorella vulgaris and Haematococcus pluvialis biomass as colouring and antioxidant in food emulsions. Eur. Food Res. Technol. 2006, 222, 362–367. [Google Scholar] [CrossRef] [Green Version]
  91. Suh, S.S.; Hwang, J.; Park, M.; Seo, H.H.; Kim, H.S.; Lee, J.H.; Moh, S.H.; Lee, T.K. Anti-inflammation activities of mycosporine-like amino acids (MAAs) in response to UV radiation suggest potential anti-skin aging activity. Mar. Drugs 2014, 12, 5174–5187. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Challouf, R.; Trabelsi, L.; Ben Dhieb, R.; El Abed, O.; Yahia, A.; Ghozzi, K.; Ben Ammar, J.; Omran, H.; Ben Ouada, H. Evaluation of cytotoxicity and biological activities in extracellular polysaccharides released by cyanobacterium Arthrospira platensis. Braz. Arch. Biol. Technol. 2011, 54, 831–838. [Google Scholar] [CrossRef]
  93. Xia, S.; Gao, B.; Li, A.; Xiong, J.; Ao, Z.; Zhang, C. Preliminary Characterization, Antioxidant Properties and Production of Chrysolaminarin from Marine Diatom Odontella aurita. Mar. Drugs 2014, 12, 4883–4897. [Google Scholar] [CrossRef] [Green Version]
  94. Tannin-Spitz, T.; Bergman, M.; Van-Moppes, D.; Grossman, S.; Arad, S. Antioxidant activity of the polysaccharide of the red microalga Porphyridium sp. J. Appl. Phycol. 2005, 17, 215–222. [Google Scholar] [CrossRef]
  95. Chen, B.; You, W.; Huang, J.; Yu, Y.; Chen, W. Isolation and antioxidant property of the extracellular polysaccharide from Rhodella reticulata. World J. Microbiol. Biotechnol. 2010, 26, 833–840. [Google Scholar] [CrossRef]
  96. Smerilli, A.; Balzano, S.; Maselli, M.; Blasio, M.; Orefice, I.; Galasso, C.; Sansone, C.; Brunet, C. Antioxidant and photoprotection networking in the coastal diatom skeletonema marinoi. Antioxidants 2019, 8, 154. [Google Scholar] [CrossRef] [Green Version]
  97. Colusse, G.A.; Mendes, C.R.B.; Duarte, M.E.R.; de Carvalho, J.C.; Noseda, M.D. Effects of different culture media on physiological features and laboratory scale production cost of Dunaliella salina. Biotechnol. Rep. 2020, 27. [Google Scholar] [CrossRef]
  98. Hashtroudi, M.S.; Shariatmadari, Z.; Riahi, H.; Ghassempour, A. Analysis of Anabaena vaginicola and Nostoc calcicola from Northern Iran, as rich sources of major carotenoids. Food Chem. 2013, 136, 1148–1153. [Google Scholar] [CrossRef] [PubMed]
  99. Rastogi, R.P.; Madamwar, D.; Nakamoto, H.; Incharoensakdi, A. Resilience and self-regulation processes of microalgae under UV radiation stress. J. Photochem. Photobiol. C Photochem. Rev. 2020, 43, 100322. [Google Scholar] [CrossRef]
  100. Sharma, A.; Singh, R.S.; Gupta, G.; Ahmad, T.; Kaur, B. Metabolic engineering of enzyme-regulated bioprocesses. In Biomass, Biofuels, Biochemicals: Advances in Enzyme Catalysis and Technologies; Elsivier: Alpharetta, GA, USA, 2019; pp. 293–323. [Google Scholar] [CrossRef]
  101. Putu Parwata, I.; Wahyuningrum, D.; Suhandono, S.; Hertadi, R. Production of ectoine by Halomonas elongata BK-AG25 using osmotic shock technique. IOP Conf. Ser. Earth Environ. Sci. 2018, 209. [Google Scholar] [CrossRef]
  102. Botta, C.; Di Giorgio, C.; Sabatier, A.S.; De Méo, M. Genotoxicity of visible light (400-800 nm) and photoprotection assessment of ectoin, l-ergothioneine and mannitol and four sunscreens. J. Photochem. Photobiol. B Biol. 2008, 91, 24–34. [Google Scholar] [CrossRef] [PubMed]
  103. D’Agostino, P.M.; Javalkote, V.S.; Mazmouz, R.; Pickford, R.; Puranik, P.R.; Neilan, B.A. Comparative profiling and discovery of novel glycosylated mycosporine-like amino acids in two strains of the cyanobacterium Scytonema cf. crispum. Appl. Environ. Microbiol. 2016, 82, 5951–5959. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Raj, S.; Kuniyil, A.M.; Sreenikethanam, A.; Gugulothu, P.; Jeyakumar, R.B.; Bajhaiya, A.K. Microalgae as a source of mycosporine-like amino acids (Maas); advances and future prospects. Int. J. Environ. Res. Public Health 2021, 18, 12402. [Google Scholar] [CrossRef] [PubMed]
  105. Gabros, S.; Nessel, T.; Zito, P. Sunscreens and Photoprotection. Available online: https://www.ncbi.nlm.nih.gov/books/NBK537164/ (accessed on 26 January 2022).
  106. Rai, R.; Shanmuga, S.C.; Srinivas, C. Update on photoprotection. Indian J. Dermatol. 2012, 57, 335–342. [Google Scholar] [CrossRef]
  107. Núñez-Pons, L.; Avila, C.; Romano, G.; Verde, C.; Giordano, D. UV-Protective Compounds in Marine Organisms from the Southern Ocean. Mar. Drugs 2018, 16, 336. [Google Scholar] [CrossRef] [Green Version]
  108. White, A.L.; Jahnke, L.S. Contrasting Effects of UV-A and UV-B on Photosynthesis and Photoprotection of β-carotene in two Dunaliella spp. Plant Cell Physiol. 2002, 43, 877–884. [Google Scholar] [CrossRef] [Green Version]
  109. Joshi, D.; Mohandass, C.; Dhale, M. Effect of UV-B Radiation and Desiccation Stress on Photoprotective Compounds Accumulation in Marine Leptolyngbya sp. Appl. Biochem. Biotechnol. 2018, 184, 35–47. [Google Scholar] [CrossRef]
  110. Singh, D.K.; Pathak, J.; Pandey, A.; Singh, V.; Ahmed, H.; Kumar, D.; Rajneesh; Sinha, R.P. Response of a rice-field cyanobacterium Anabaena sp. HKAR-7 upon exposure to ultraviolet-B radiation and ammonium chloride. Environ. Sustain. 2021, 4, 95–105. [Google Scholar] [CrossRef]
  111. Xiong, F.; Komenda, J.; Kopecky, J.; Nedbal, L. Strategies of ultraviolet-B protection in microscopic algae. Physiol. Plant. 1997, 100, 378–388. [Google Scholar] [CrossRef]
  112. Gao, Y.; Cui, Y.; Xiong, W.; Li, X.; Wu, Q. Effect of UV-C on Algal Evolution and Differences in Growth Rate, Pigmentation and Photosynthesis Between Prokaryotic and Eukaryotic Algae. Photochem. Photobiol. 2009, 85, 774–782. [Google Scholar] [CrossRef] [PubMed]
  113. Ganapathy, K.; Chidambaram, K.; Janarthanan, R.; Ramasamy, R. Effect of UV-B Radiation on Growth, Photosynthetic Activity andMetabolic Activities of Chlorella vulgaris. Res. Rev. J. Microbiol. Biotechnol. 2017, 6, 53–60. [Google Scholar]
  114. Hernando, M.; Schlossi, I.; Roy, S.; Ferreyra, G. Symposium-in-Print: UV Effects on Aquatic and Coastal Ecosystems Photoacclimation to Long-Term Ultraviolet Radiation Exposure of Natural Sub-Antarctic Phytoplankton Communities: Fixed Surface Incubations Versus Mixed Mesocosms. Photochem. Photobiol. 2006, 923–935. [Google Scholar] [CrossRef]
  115. Chen, L.; Deng, S.; De Philippis, R.; Tian, W.; Wu, H.; Wang, J. UV-B resistance as a criterion for the selection of desert microalgae to be utilized for inoculating desert soils. J. Appl. Phycol. 2013, 25, 1009–1015. [Google Scholar] [CrossRef]
  116. Fouqueray, M.; Mouget, J.-L.; Morant-Manceau, A.; Tremblin, G. Dynamics of short-term acclimation to UV radiation in marine diatoms. J. Photochem. Photobiol. B. 2007, 89, 1–8. [Google Scholar] [CrossRef]
  117. Helbling, E.W.; Gao, K.; Ai, H.; Ma, Z.; Villafañe, V.E. Differential responses of Nostoc sphaeroides and Arthrospira platensis to solar ultraviolet radiation exposure. J. Appl. Phycol. 2006, 18, 57–66. [Google Scholar] [CrossRef]
  118. Buma, A.; Wright, S.; Enden, R.; van de Poll, W.; Davidson, A. PAR acclimation and UVBR-induced DNA damage in Antarctic marine microalgae. Mar. Ecol. Prog. Ser. 2006, 315, 33–42. [Google Scholar] [CrossRef]
  119. Bennett, J.; Soule, T. Expression of Scytonemin Biosynthesis Genes under Alternative Stress Conditions in the Cyanobacterium Nostoc punctiforme. Microorganisms 2022, 10, 427. [Google Scholar] [CrossRef]
  120. Pathak, J.; Kumar, D.; Singh, D.K.; Ahmed, H.; Kannaujiya, V.K.; Sinha, R.P. Ultraviolet radiation and salinity-induced physiological changes and scytonemin induction in cyanobacteria isolated from diverse habitats. Biointerface Res. Appl. Chem. 2022, 12, 3590–3606. [Google Scholar] [CrossRef]
  121. Orellana, G.; Gómez-Silva, B.; Urrutia, M.; Galetović, A. UV-A Irradiation Increases Scytonemin Biosynthesis in Cyanobacteria Inhabiting Halites at Salar Grande, Atacama Desert. Microorganisms 2020, 8, 1690. [Google Scholar] [CrossRef] [PubMed]
  122. Sinha, R.P.; Pathak, J.; Rajneesh; Ahmed, H.; Pandey, A.; Singh, P.R.; Mishra, S.; Häder, D.-P. Cyanobacterial photoprotective compounds: Characterization and utilization in human welfare. In Natural Bioactive Compounds; Sinha, R.P., Häder, D.-P., Eds.; Academic Press: Cambridge, MA, USA, 2021; pp. 187–201. ISBN 9780128206553. [Google Scholar]
  123. Gao, X.; Jing, X.; Liu, X.; Lindblad, P. Biotechnological Production of the Sunscreen Pigment Scytonemin in Cyanobacteria: Progress and Strategy. Mar. Drugs 2021, 19, 129. [Google Scholar] [CrossRef] [PubMed]
  124. Sibilla, S.; Godfrey, M.; Brewer, S.; Budh-Raja, A.; Genovese, L. An overview of the beneficial effects of hydrolysed collagen as a nutraceutical on skin properties: Scientific background and clinical studies. Open Nutraceuticals J. 2015, 8, 29–42. [Google Scholar] [CrossRef] [Green Version]
  125. Perin, G.; Bellan, A.; Bernardi, A.; Bezzo, F.; Morosinotto, T. The potential of quantitative models to improve microalgae photosynthetic efficiency. Physiol. Plant. 2019, 166, 380–391. [Google Scholar] [CrossRef] [PubMed]
  126. Dixon, C.; Wilken, L.R. Green microalgae biomolecule separations and recovery. Bioresour. Bioprocess. 2018, 5, 14. [Google Scholar] [CrossRef] [Green Version]
  127. Sheih, I.-C.; Wu, T.-K.; Fang, T.J. Antioxidant properties of a new antioxidative peptide from algae protein waste hydrolysate in different oxidation systems. Bioresour. Technol. 2009, 100, 3419–3425. [Google Scholar] [CrossRef]
  128. Yaakob, Z.; Ali, E.; Zainal, A.; Mohamad, M.; Takriff, M.S. An overview: Biomolecules from microalgae for animal feed and aquaculture. J. Biol. Res. 2014, 21, 6. [Google Scholar] [CrossRef] [Green Version]
  129. de la Noue, J.; de Pauw, N. The potential of microalgal biotechnology: A review of production and uses of microalgae. Biotechnol. Adv. 1988, 6, 725–770. [Google Scholar] [CrossRef]
  130. Odjadjare, E.C.; Mutanda, T.; Olaniran, A.O. Potential biotechnological application of microalgae: A critical review. Crit. Rev. Biotechnol. 2017, 37, 37–52. [Google Scholar] [CrossRef]
  131. Hamed, I. The Evolution and Versatility of Microalgal Biotechnology: A Review. Compr. Rev. Food Sci. Food Saf. 2016, 15, 1104–1123. [Google Scholar] [CrossRef] [PubMed]
  132. Gold, N.A.; Mirza, T.M.; Avva, U. Alcohol Sanitizer; StatPearls Publishing: Treasure Island, FL, USA, 2022. [Google Scholar]
  133. Hoang, H.T.; Van Tran, V.; Bui, V.K.H.; Kwon, O.-H.; Moon, J.-Y.; Lee, Y.-C. Novel moisturized and antimicrobial hand gel based on zinc-aminoclay and Opuntia humifusa extract. Sci. Rep. 2021, 11, 17821. [Google Scholar] [CrossRef] [PubMed]
  134. Bernaerts, T.M.M.; Gheysen, L.; Foubert, I.; Hendrickx, M.E.; Van Loey, A.M. The potential of microalgae and their biopolymers as structuring ingredients in food: A review. Biotechnol. Adv. 2019, 37, 107419. [Google Scholar] [CrossRef] [PubMed]
  135. Rossi, F.; De Philippis, R. Exocellular Polysaccharides in Microalgae and Cyanobacteria: Chemical Features, Role and Enzymes and Genes Involved in Their Biosynthesis. In The Physiology of Microalgae; Springer International Publishing: Berlin/Heidelberg, Germany, 2016; pp. 565–590. [Google Scholar]
  136. Pierre, G.; Delattre, C.; Dubessay, P.; Jubeau, S.; Vialleix, C.; Cadoret, J.-P.; Probert, I.; Michaud, P. What Is in Store for EPS Microalgae in the Next Decade? Molecules 2019, 24, 4296. [Google Scholar] [CrossRef] [Green Version]
  137. Jain, R.; Raghukumar, S.; Tharanathan, R.; Bhosle, N. Extracellular Polysaccharide Production by Thraustochytrid Protists. Mar. Biotechnol. 2005, 7, 184–192. [Google Scholar] [CrossRef]
  138. Urbanelli, L.; Buratta, S.; Tancini, B.; Sagini, K.; Delo, F.; Porcellati, S.; Emiliani, C. The Role of Extracellular Vesicles in Viral Infection and Transmission. Vaccines 2019, 7, 102. [Google Scholar] [CrossRef] [Green Version]
  139. Adamo, G.; Fierli, D.; Romancino, D.P.; Picciotto, S.; Barone, M.E.; Aranyos, A.; Božič, D.; Morsbach, S.; Raccosta, S.; Stanly, C.; et al. Nanoalgosomes: Introducing extracellular vesicles produced by microalgae. J. Extracell. Vesicles 2021, 10, e12081. [Google Scholar] [CrossRef]
  140. Delalat, B.; Sheppard, V.C.; Rasi Ghaemi, S.; Rao, S.; Prestidge, C.A.; McPhee, G.; Rogers, M.-L.; Donoghue, J.F.; Pillay, V.; Johns, T.G.; et al. Targeted drug delivery using genetically engineered diatom biosilica. Nat. Commun. 2015, 6, 8791. [Google Scholar] [CrossRef] [Green Version]
  141. Khavari, F.; Saidijam, M.; Taheri, M.; Nouri, F. Microalgae: Therapeutic potentials and applications. Mol. Biol. Rep. 2021, 48, 4757–4765. [Google Scholar] [CrossRef]
  142. Enamala, M.K.; Enamala, S.; Chavali, M.; Donepudi, J.; Yadavalli, R.; Kolapalli, B.; Aradhyula, T.V.; Velpuri, J.; Kuppam, C. Production of biofuels from microalgae-A review on cultivation, harvesting, lipid extraction, and numerous applications of microalgae. Renew. Sustain. Energy Rev. 2018, 94, 49–68. [Google Scholar] [CrossRef]
  143. Gunes, S.; Tamburaci, S.; Dalay, M.C.; Gurhan, I.D. In vitro evaluation of spirulina platensis extract incorporated skin cream with its wound healing and antioxidant activities. Pharm. Biol. 2017, 55, 1824–1832. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Tarasuntisuk, S.; Patipong, T.; Hibino, T.; Waditee-Sirisattha, R.; Kageyama, H. Inhibitory effects of mycosporine-2-glycine isolated from a halotolerant cyanobacterium on protein glycation and collagenase activity. Lett. Appl. Microbiol. 2018, 67, 314–320. [Google Scholar] [CrossRef] [PubMed]
  145. Eigemann, F.; Vanormelingen, P.; Hilt, S. Sensitivity of the Green Alga Pediastrum duplex Meyen to Allelochemicals Is Strain-Specific and Not Related to Co-Occurrence with Allelopathic Macrophytes. PLoS ONE 2013, 8, e78463. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Wang, H.M.D.; Chen, C.C.; Huynh, P.; Chang, J.S. Exploring the potential of using algae in cosmetics. Bioresour. Technol. 2015, 184, 355–362. [Google Scholar] [CrossRef] [PubMed]
  147. Campiche, R.; Sandau, P.; Kurth, E.; Massironi, M.; Imfeld, D.; Schuetz, R. Protective effects of an extract of the freshwater microalga Scenedesmus rubescens on UV-irradiated skin cells. Int. J. Cosmet. Sci. 2018, 40, 187–192. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Molecular structure from microalgae-derived compounds used for multiple biotechnological applications.
Figure 1. Molecular structure from microalgae-derived compounds used for multiple biotechnological applications.
Molecules 27 03512 g001
Figure 2. Microalgae-derived biomolecule production. Four main steps are required to obtain biomolecules from microalgae biomass: (1) microalgae selection, (2) culture standardization, (3) purification, and (4) functionalization. Considerations must be made for maintaining sustainability and efficacy. Created with BioRender.com.
Figure 2. Microalgae-derived biomolecule production. Four main steps are required to obtain biomolecules from microalgae biomass: (1) microalgae selection, (2) culture standardization, (3) purification, and (4) functionalization. Considerations must be made for maintaining sustainability and efficacy. Created with BioRender.com.
Molecules 27 03512 g002
Table 1. List of compounds with immunomodulatory activity obtained from microalgae.
Table 1. List of compounds with immunomodulatory activity obtained from microalgae.
Type of CompoundMechanism of ActionMicroalgae SpeciesCulture ConditionsExtractionReference
PeptidesAntimicrobial activity through interaction with negatively charged membranesT. suecica
(Chlorophyta)
F/2 medium, 21 ± 0.5 °C and 40 W LightAcid extraction and reverse phase column chromatography separation.[53]
Fatty acid methyl esters (FAME), lipids and carotenoidsInterference in biosynthesis of bacterial fatty acidsVarious microalgae speciesF/2 medium for marine microalgae and Bold’s
basal medium (BBM) for freshwater microalgae
Three extraction systems with different solvents[54]
Fatty acids and pigmentsInterference in biosynthesis of bacterial fatty acidsS. obliquus-Polarity-wise successive solvent extractions[55]
LipidsMembrane permeabilizationS. brasiliensis, E. acutiformis and Sphaerospermopsis sp.M7 medium for microalgae,
Z8 medium for cyanobacteria
Extraction with different solvents[56]
Fatty acidsInterference in biosynthesis of bacterial fatty acidsP. nurekisMedium with various macro- and microelements at pH < 6, 30 °C and 16,000 LUXExtraction with water and ethanol[57]
Fatty acids and phenolic compoundsInterference in biosynthesis of bacterial fatty acidsS. obliquusBG-11 medium Extraction with organic solvents and sonication[58]
PigmentsFree radical scavengingVarious microalgae species Guillard’s F/2
medium, 22 ± 2 °C, 14:10 D:L cycle, 60–65 µ·E·m−2·s−1light for 16 days
Ethanol extraction [59]
Microalgae extractVarious mechanisms tested32 microalgae speciesGuillard’s F/2
medium, 19 °C, 12:12 D:L cycle, and 100 µmo·m−2·s−1 Light
Sonication and acetone extraction [60]
Table 3. List of different microalgae with photoprotection features.
Table 3. List of different microalgae with photoprotection features.
Microalgae SpeciesUV TestIntensity UnitResistance FactorType of StudyCulture ConditionsReference
Anabaena spp.
(Cyanobacteria)
UVB1 W·m−2 for 4 h/dayScytonemin, MAAsExposure to UVB-RBG-11 medium, NH4Cl as N-source, 28 ± 2 °C[110]
Characium terrestre
(Chlorophyta)
UVBNRSporopolleninExposure to extreme UVB irradianceMedium reported by Zachleder et al., 26 °C.[118]
C. protothecoides
(Chlorophyta)
UVC0.01 to 0.20 W·m−2LuteinExposure to different UVC irradiancesBG-11 medium,
25 ± 1 °C
[119]
C. vulgaris
(Chlorophyta)
UVB1 to 5 W·m−2Sporopollenin, Scytonemin, MAAsExposure to different UVB irradiancesBold Basal medium,
25 ± 1 °C
[111]
D.salina
(Chlorophyta)
UVA
UVB
110 mmol·m−2·s−1 for UVASporopollenin, Scytonemin, MAAsExposure to UVA-R and UVB-RMedium with NaCl,
26 °C
[108]
Eutreptiella sp.
(Euglenozoa)
UVB280 to 320 nmXanthophylls, MAAsTested under fixed lightF/2 medium, 10 °C[120]
Nostoc sp.
(Cyanobacteria)
UVB312 nmCarotenoids, ScytoneminPhotosynthetic activity essayBG-11 medium, 25 °C[112]
O. aurita
(Bacillariophyta)
UVA
UVB
110 kJ·m−2D1 protein, activation of antioxidant enzymesExposure to UVA-R and UVB-RArtificial seawater reported by Harrison et al. [121]
N. sphaeroides
(Cyanobacteria)
UVA320 nmNot identifiedExposure to UVA irradianceBG-11 medium, 23 °C[122]
P. antarctica, P. glacialis
(Rhodophyta)
UVB280 to 400 nmMAAs, XanthophyllsAcclimation to photosynthetically active radiationGP5 medium,
1.0 ± 0.5 °C
[123]
S. platensis
(Cyanobacteria)
UVA320 nmSporopollenin, Scytonemin, MAAsExposure to UVA irradianceBG-11 medium, 23 °C[122]
N. punctiformeUVA
UVC
100 µm photon m−2s−1ScytoneminExposure to UVA, UVB and osmotic stressAA/4 medium[115]
Scytonema sp.UVA
UVB
6.5 Wm−2
0.56 Wm−2
MAAs and antioxidant enzymesPhotosynthetically active radiation and salinityBG-11 medium[116]
HalotheceUVA3.6 Wm−2ScytoneminExposure to UVA-RHalites[117]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Martínez-Ruiz, M.; Martínez-González, C.A.; Kim, D.-H.; Santiesteban-Romero, B.; Reyes-Pardo, H.; Villaseñor-Zepeda, K.R.; Meléndez-Sánchez, E.R.; Ramírez-Gamboa, D.; Díaz-Zamorano, A.L.; Sosa-Hernández, J.E.; et al. Microalgae Bioactive Compounds to Topical Applications Products—A Review. Molecules 2022, 27, 3512. https://doi.org/10.3390/molecules27113512

AMA Style

Martínez-Ruiz M, Martínez-González CA, Kim D-H, Santiesteban-Romero B, Reyes-Pardo H, Villaseñor-Zepeda KR, Meléndez-Sánchez ER, Ramírez-Gamboa D, Díaz-Zamorano AL, Sosa-Hernández JE, et al. Microalgae Bioactive Compounds to Topical Applications Products—A Review. Molecules. 2022; 27(11):3512. https://doi.org/10.3390/molecules27113512

Chicago/Turabian Style

Martínez-Ruiz, Manuel, Carlos Alberto Martínez-González, Dong-Hyun Kim, Berenice Santiesteban-Romero, Humberto Reyes-Pardo, Karen Rocio Villaseñor-Zepeda, Edgar Ricardo Meléndez-Sánchez, Diana Ramírez-Gamboa, Ana Laura Díaz-Zamorano, Juan Eduardo Sosa-Hernández, and et al. 2022. "Microalgae Bioactive Compounds to Topical Applications Products—A Review" Molecules 27, no. 11: 3512. https://doi.org/10.3390/molecules27113512

Article Metrics

Back to TopTop