Next Article in Journal
Deciphering Molecular Mechanism of the Neuropharmacological Action of Fucosterol through Integrated System Pharmacology and In Silico Analysis
Next Article in Special Issue
Quantitative and Qualitative HPLC Analysis of Mycosporine-Like Amino Acids Extracted in Distilled Water for Cosmetical Uses in Four Rhodophyta
Previous Article in Journal
Anti-Obesity Effect of Diphlorethohydroxycarmalol Isolated from Brown Alga Ishige okamurae in High-Fat Diet-Induced Obese Mice
Previous Article in Special Issue
Bostrychines A–F, Six Novel Mycosporine-Like Amino-Acids and a Novel Betaine from the Red Alga Bostrychia scorpioides
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mycosporine-Like Amino Acids: Making the Foundation for Organic Personalised Sunscreens

by
Nedeljka N. Rosic
1,2
1
School of Health and Human Sciences, Southern Cross University, Southern Cross Drive, Bilinga, QLD 4225, Australia
2
Marine Ecology Research Centre, Southern Cross University, Lismore, Military Rd, East Lismore, NSW 2480, Australia
Mar. Drugs 2019, 17(11), 638; https://doi.org/10.3390/md17110638
Submission received: 27 October 2019 / Revised: 9 November 2019 / Accepted: 10 November 2019 / Published: 12 November 2019
(This article belongs to the Special Issue Mycosporine-Like Amino Acids from Marine Resource)

Abstract

:
The surface of the Earth is exposed to harmful ultraviolet radiation (UVR: 280–400 nm). Prolonged skin exposure to UVR results in DNA damage through oxidative stress due to the production of reactive oxygen species (ROS). Mycosporine-like amino acids (MAAs) are UV-absorbing compounds, found in many marine and freshwater organisms that have been of interest in use for skin protection. MAAs are involved in photoprotection from damaging UVR thanks to their ability to absorb light in both the UV-A (315–400 nm) and UV-B (280–315 nm) range without producing free radicals. In addition, by scavenging ROS, MAAs play an antioxidant role and suppress singlet oxygen-induced damage. Currently, there are over 30 different MAAs found in nature and they are characterised by different antioxidative and UV-absorbing capacities. Depending on the environmental conditions and UV level, up- or downregulation of genes from the MAA biosynthetic pathway results in seasonal fluctuation of the MAA content in aquatic species. This review will provide a summary of the MAA antioxidative and UV-absorbing features, including the genes involved in the MAA biosynthesis. Specifically, regulatory mechanisms involved in MAAs pathways will be evaluated for controlled MAA synthesis, advancing the potential use of MAAs in human skin protection.

1. Introduction

Due to a reduction in aerosols and cloud cover, the levels of ultraviolet radiation (UVR) reaching the Earth’s surface are predicted to increase during the 21st century [1]. Organisms have developed several photoprotective mechanisms to survive high levels of UVR. Different mitigation strategies are utilised by different species and often in combination including DNA repair systems, antioxidant activities, and the application of UV-absorbing compounds. Exposure to UVR results in reactive oxygen species (ROS) production, oxidative stress and DNA damage. The mitigation strategies to reduce the UV-induced damage include DNA repair mechanisms via processes such as photoreactivation, excision, and mismatch repair. In response to the generation of ROS, organisms start accumulating antioxidants to capture free radicals. In addition, as a part of the UVR interception response, organisms accumulate photoprotective compounds with UVR absorption capabilities such as mycosporine-like amino acids (MAAs) [2,3,4,5]. MAAs are found in a large number of aquatic species, including marine and freshwater organisms that have been exposed to high levels of damaging UVR [6]. The vast variety of species containing MAAs includes phytoplankton, cyanobacteria, fungi, macroalgae, microalgae, as well as animals coming from both aquatic and terrestrial ecosystems [7]. These ubiquitous and highly abundant secondary metabolites are found to have a critical photoprotective role in these aquatic species [8]. MAAs have been evolutionarily conserved in aquatic organisms [9], together with some other important photoprotective compounds such as scytonemin in cyanobacteria [10,11], carotenoid pigments in plants and many microorganisms [12,13].
In this review, the aim was (1) to provide a summary of the MAA UV-absorbing, antioxidative, anti-inflammatory; antiaging features; (2) to discuss the genes involved in MAA biosynthesis and (3) to evaluate the regulatory mechanisms involved in MAA synthesis. In particular, this review assesses current knowledge about the biosynthesis of MAAs and suggests possible directions for the evolving biotechnological potential of MAAs in human skin protection.

2. MAA Diversity

MAAs are a diverse group of colourless and hydrophilic compounds characterised by their small molecular mass (<400 Da). In their core, MAAs are composed of a cyclohexenone or a cyclohexenimine ring conjugated to an amino acid residue or its imino alcohol (Figure 1A) [14,15,16]. MAAs absorb mainly in the range of 310 to 362 nm and are characterised with high molar extinction coefficients (ε = 28,100 to 50,000 M−1 cm−1) [6,17]. There are over 30 different MAA compounds identified in nature [18,19] and probably more to be discovered with further development of novel high-throughput technologies. The examples of five major MAAs found in most aquatic species including their direct precursor, 4-deoxygadusol (4-DG), are presented in Figure 1B.
A huge variety of MAAs have been reported in cyanobacteria [20,21,22], red algae [23,24,25,26,27,28], fungi [29], green algae [30,31], dinoflagellates [32,33,34], invertebrates (e.g., sponges [15,35,36], corals [21,37,38], sea urchins [39,40]) and vertebrates such as fish [41,42,43,44]. However, not all species have all the MAAs. From all organisms exposed to high levels of UV radiation, cyanobacteria have been under an especially huge evolutionary pressure to survive damaging UV radiation. An overview of MAAs reported in this diverse taxonomic group provides an important summary of the organisms’ ability to adapt for survival [22]. Red algae (Rhodophyta) are characterised by the highest diversity and concentrations of MAAs [16,45]. In a new study, MAA profiles were established for 23 red algae, with the most abundant MAAs being shinorine, palythine, asterina-330; porphyra-334 [28]. Furthermore, six new MAAs were recently chemically characterised in the red alga Bostrychia scorpioides [46]. In addition, two new MAAs (LC (Lendenfeldia chondrode)-343 and mycosporine-ethanolamine) along with well-known asterina-330 and shinorine were recently isolated from the marine sponge Lendenfeldia chondrode [36].
MAAs were also reported in many other animals (e.g., fish and sea stars) that naturally lack the MAA biosynthetic pathway, but do acquire MAAs through their algal diet or symbiosis with algae or/and bacterial symbionts [6,7,42,47]. However, beyond the diet, de novo synthesis of gadusol (the MAA precursor) was reported in some fish [43,44] and corals [48]. A wide range of MAAs was reported in the coral Pocillopora capitate, including both primary and secondary MAAs: mycosporine-glycine, shinorine, porphyra-334, mycosporine-methylamine-serine, mycosporine-methylamine-threonine, palythine-serine, palythine, and palythine-threonine [49]. In holobionts like corals, the presence of their microbial symbionts may be an efficient way to develop and adapt to different UV conditions [6,32]. However, coral dinoflagellates in cultures had less diverse MAA profiles compared to when dinoflagellates were in symbiosis within the coral host [32,49].
Identification and characterisation of MAAs requires the use of multiple diverse chemistry methods and measurements. The overview of different methodologies applied for the identification and characterisation of some common MAAs are provided in a very comprehensive review by Carreto and Carigan [16]. Furthermore, different methods of the MAA extractions are used for the successful isolation of MAAs. The appropriate methods for MAA extractions are influenced by the organisms’ characteristics and the tissue types used for extraction [45]. For the characterisation of MAAs, the combination of different methods is required, usually including the use of high-performance liquid chromatography (HPLC) and mass spectrometry (Figure 2, [45]). Multiple methods including HPLC analysis, reverse-phase liquid chromatography-mass spectrometry (RPLC-MS) and hydrophilic interaction liquid chromatography (HILIC) were recently used for the discovery and characterisation of novel MAAs in cyanobacteria [50]. The importance of combining different methods of chemical analyses characterised by improved sensitivity was recently confirmed by Lagegerie et al. [27]. The initial analyses of MAAs in red macroalgae from the Brittany region in France resulted in the detection of 23 potential MAAs using HPLC; further analyses using liquid chromatography–mass spectrometry (LC-MS) confirmed that only six different types of MAAs (shinorine, palythine, asterina-330, porphyra-334, usurijene; palythene) were found in 40 species of red macroalgae [27]. Additional and more comprehensive characterisation of the MAA chemical structures can be done by a combination of nuclear magnetic resonance (NMR) and LC-MS analyses [51]. In addition, the combination of infrared (IR) spectroscopic analysis and gas chromatography (GC)-MS analysis is also used for improving the characterisation of MAAs [52].

2.1. MAA UV-Absorbing Features

In nature, various compounds are used for protection against the mutagenic effects of UV radiation. Dark pigment melanins found in humans and animals absorb the light in the UV and visible range. In humans, melanin protects skin by providing a physical barrier to UV and by absorbing 50%–75% of UVR [53]. Beyond UV absorbing properties, melanin pigments play a role as antioxidants [54]. In plants and many microorganisms, carotenoid pigments have a photoprotective role and are characterised by absorption in the range of 300–600 nm [13]. These pigments are important not only for the UV protection [4,17] but also in photosynthesis as an essential part of the photosynthetic apparatus [12,17].
Another well-known photoprotective compound is scytonemin, a small, yellow-brown, hydrophobic pigment that is found only in cyanobacteria [11]. Scytonemin absorbs in the UV-A wavelength region, with maximum absorption in purified form at 384 nm [55,56] and in vivo at 370 nm [20]. This pigment works not only as a potent UV protector but also as a powerful antioxidant [57] and has significant potential for sunscreen application in cosmetics [57,58]. Over a long period of history, cyanobacteria were exposed to extremely high UVR and consequently, these organisms developed additional protection of their DNA by applying two powerful UV-absorbing compounds, scytonemin and MAAs [10,59].
MAAs are the most common group of secondary metabolites found in aquatic species [17]. Marine and freshwater organisms utilising MAA compounds in UV protection include a variety of species from cyanobacteria, fungi, algae, to higher-order animals such as cnidaria, fishes, arthropods, mollusks, tunicates and echinoderms [7]. MAAs are involved in photoprotection due to their ability to absorb light in the range of UV-A (315–400 nm; making ~95 % of UV energy that penetrates the atmosphere) and UV-B (280–315 nm) without production of free radicals. The vast majority of different MAAs absorb within the UV-A range, like mycosporine-2-glycine, shinorine; porphyra-334 with ʎmax in the range of 332–334 nm (Figure 1B) [16]. Palythine, palythine-threonine, palythine-serine and palythinol have ʎmax at 320 nm, while palythine-serine sulfate and palythine-threonine sulfate have maximum absorbance at 321 nm. Usujirene, palythene; euhalothece-362 are characterised by the ability to absorb UV at higher UV-A wavelengths in the range of 357–362 nm [16].
MAAs containing cyclohexenone ring have maximum absorption within the UV-B range including mycosporine-glycine (ʎmax = 310 nm), mycosporine-taurine (ʎmax = 309 nm) and mycosporine-serine (ʎmax = 310 nm). On the other hand, MAA precursor 4-deoxygadusol is characterised by maximum absorption at 268 nm in acidic conditions and at 294 nm in basic environments [60]. Various amino acids in the MAA core result in different MAA profiles, which are characterised by different UV screening properties [60]. Therefore, only slight changes in the MAA structures will result in quick changes in the level of UV protection. This MAA property allows organisms to adjust to changeable UV conditions, which is important from the evolutionary perspective for survival under harsh and damaging UV radiation. Consequently, the accumulation of MAAs, like adaptable sunscreen, plays a critical role in the UV protection of marine organisms [8].

2.2. MAA Antioxidative Properties

Oxidative stress happens due to the production of ROS, which includes, in general, the following products: hydrogen peroxide (H2O2), hydroxyl radical (OH) and superoxide anion (O2•−). Prolonged exposure to sun radiation results in UV-induced oxidative stress. MAAs are ubiquitous metabolites that beyond their photoprotective role, also have a role as antioxidants [3,7,52]. MAAs are able to scavenge ROS and suppress singlet oxygen-induced damage [24,61,62,63]. An overview of the antioxidant properties of different MAAs has been provided in recent reviews [52,60]. In addition, a number of studies are summarised in Table 1 that assessed the antioxidative, anti-inflammatory and antiaging activities of individual MAAs in vitro.
Large differences in antioxidative capacities were reported for different MAAs. In vitro analyses of MAA antioxidative activities were done using different assays and it was apparent that modification in the external environment, such as the acidity or temperature, may increase their antioxidative properties [52]. Weak antioxidative activity was described for shinorine and porphyra-334 [24], although changes in the environment, like heat stress, resulted in an increase in the antioxidative properties for porphira-343 [67]. On the other hand, a highly abundant primary MAA mycosporine-glycine demonstrated effective antioxidant properties [24], as well as 4-deoxygadusol [16,72]. The antioxidant capacity of mycosporine-glycine was initially tested using peroxidation assay [24]. The concentration-dependent inhibition of lipid peroxidation was reported as a result of antioxidant action of this MAA [24]. The highest antioxidant activity was reported for mycosporine-glycine isolated from the marine lichen Lichina pygmaea at pH 8.5, which was eightfold higher than for ascorbic acid [61]. Furthermore, using the extracts of marine green alga, antioxidant properties were confirmed for mycosporine-glycine via the 2,2-diphenyl-1-picryhydrazyl (DPPH) assay, but not for porphyra-334 and shinorine [31]. On the contrary, Gacesa and colleagues [64] recently demonstrated antioxidant properties of porphyra-334 and shinorine using different free-radical quenching assays. The in vitro antioxidant activities were tested using the DPPH assay and were lower in two analysed MAAs compared to ascorbic acid [64]. The second assay, an oxygen radical absorption capacity (ORAC) assay, actually demonstrated substantial antioxidant activities for both MAAs [64]. Porphyra-334 and shinorine were also shown as the activator of the cytoprotective pathway demonstrating the potential for treating human degenerative diseases related to aging [64]. Specifically, the antioxidative activities of the two MAAs were related to the Keap1-Nrf2 pathway, which regulates cytoprotective cellular responses during oxidative stress. The Kelch-like ECH-associated protein 1 (Keap1) actin protein was found to detect changes in the redox status within the cell by controlling the activity of transcription nuclear factor erythroid 2-related factor 2 protein (Nrf2). Under oxidative stress, Nrf2 is activated due to detachment from Keap1 and Nrf2 was shown to regulate the expression of genes involved in the antioxidant response (antioxidant response element: ARE) and to play a role in oncogenesis [73]. In primary skin fibroblast cells, MAAs porphyra-334 and shinorine were able to provide protection from UVR-induced oxidative stress via the activation of the Keap1-Nrf2-ARE pathway and plus directly, by quenching free radicals [64]. Another MAA, mycosporine-2-glycine, in both in vivo and in vitro studies, demonstrated a high antioxidant activity that was equivalent to ascorbic acid [69]. To induce oxidative stress, the macrophage cells were exposed H2O2, while the presence of mycosporine-2-glycine resulted in downregulation of the expression of oxidative stress-induced genes such as Cu/Zn-superoxide dismutase 1 (Sod1) and catalase (Cat) [70].
The prevention of UV-induced damage and oxidative stress in epidermal skin cells involves the role of an endogenous defense system [74]. Antioxidant enzymes that play a critical role in this endogenous defense system include superoxide dismutase, catalase, glutathione peroxidase, glutathione reductase and thioredoxin oxidase enzymes. However, the cellular defense due to accumulated UV-absorbing compounds has been shown to be the first line of defense from oxidative stress [75]. In two scleractinian corals, the antioxidant capacities of MAAs were highly important as the protection coming from MAAs occurred before the action of antioxidative enzymes such as superoxide dismutase and catalase [75]. Thanks to the MAAs’ capacity to quench ROS and scavenge free radicals, UV induced damage is minimised in these vulnerable aquatic species. Consequently, the antioxidant defense mechanisms and ROS scavenging are critical for organisms’ survival [52].

2.3. MAA Anti-Inflammatory and Antiaging Properties

The ROS production happening due to UV damage may lead to inflammation and immune stress responses [76]. Free radicals are able to work as signalling molecules changing gene expression, leading to oxidative stress, protein oxidation and resulting in the activation of inflammatory processes through the activation of different cellular pathways [77]. Inflammatory processes induced by UV exposure are mainly regulated by nuclear factor kappa b (NF-κB) and include a number of signalling mediators such as nitric oxide (NO), inducible NO synthase (iNOS) tumor necrosis factor α (TNF- α), cyclooxygenase (COX-2) and cytokines (i.e., interleukins) [52].
The anti-inflammatory activity in the microalga MAA extracts was tested in vitro by Suh et al. [31]. The HaCAT cells (immortal human keratinocytes) were exposed to UV radiation and supplemented with an increased concentration of MAAs (0.03, 0.15, or 0.3 mM). Real-time qPCR was used to evaluate the changes in gene expression of the COX-2 gene that is found to be elevated in the case of tissue inflammation. MAAs have shown different anti-inflammatory properties, with porphyra-334 having no effect on COX-2 gene expression, while mycosporine-glycine and shinorine had the inhibitory effect on the expression of inflammation-related gene (i.e., COX-2 gene) [31]. Similarly, mycosporine-2-glycine reduced the transcription of genes critical for the inflammatory signalling processes, COX-2 and iNOS [70]. Anti-inflammatory properties of MAAs shinorine and porphyra-334 were tested in human myelomonocytic cells under inflammatory stimulation by lipopolysaccharide (LPS) [65]. Both MAAs stimulated NF-κB activity prior to LPS induction, while under LPS-induced conditions, shinorine increased the activity of transcription factor NF-κB in a dose-dependent manner. On the other hand, porphyra-334 reduced the activity of NF-κB and demonstrated anti-inflammatory action. The aqueous extracts of red algae Hydropuntia cornea and Gracilariopsis longissima containing the mixture of MAAs (palythine, asterina-330, shinorine, porphyra-334; palythinol) and other compounds were reported to actively induce the production of TNF-α and anti-inflammatory/pro-inflammatory cytokine interleukin-6 [78].
The skin aging process is happening as a result of collagen destruction and elastin content reduction [79]. Exposure to UV radiation increases the rate of skin aging via oxidative stress and DNA mutations [80,81,82]. As UV exposure leads to photoaging, the evaluation of MAAs antiaging property is also critical when assessing MAAs’ potential for use in cosmetics. Suh et al. [31] tested the expression of genes related to the skin aging processes (i.e., genes for procollagen c-endopeptidase enhancer and elastin). The mRNA levels of these UV-suppressed genes were elevated in the presence of all analysed MAAs. The UV-induced downregulation of another gene related to skin aging (involucrin) was suppressed in the presence of mycosporine-glycine and shinorine, but not porphyra-334. Additional studies also confirmed the potential of MAAs as anti-photoaging molecules [66,68,82]. Consequently, more studies are needed to extend the promising potential of MAAs for their application in skin-care products.

3. Fluctuation of the MAA Content in Aquatic Species

MAAs are multifunctional secondary metabolites that beyond the role in photoprotection and as antioxidants, also play a part in osmotic regulation, control of reproduction, as well as nitrogen reservoirs; an accessory light-harvesting pigment in photosynthesis [83]. In the soft coral species Lobophytum compactum, MAA accumulation was three times higher in eggs than in maternal tissue indicating MAA importance for larval survival [84]. MAA levels and diversity are affected by seasonal fluctuations [85]. UV radiation is a major factor influencing MAA accumulation and resulting in changes in organisms’ MAA profile [38]. Spectral variability and intensity were found to affect the synthesis of MAAs [17]. Blue light within photosynthetically active radiation (PAR) and UV-A lights stimulated the production of MAAs in free-living algae and Antarctic diatoms [86,87,88,89,90]. The positive effect of blue light and UV-A exposure was also reported in the red macroalga Chondrus crispus [91], while UV-B had a positive effect on the MAA level in Nodularia cyanobacteria [92]. In some corals, both UV levels and PAR were required for successful MAA production [93,94]. Host–microbe interactions are another important factor influencing the diversity of MAAs. Lower MAA levels and diversity were reported for symbiotic dinoflagellates when in cultures than were in symbiosis within the coral host [32,95].
Together with UVR, there are other factors that stimulate the production of MAAs such as variation in salinity conditions and changes in nutrient availability [84,96,97,98,99]. In the halotolerant cyanobacterium Aphanothece halophytica, the production of mycosporine-2-glycine was stimulated more by the high salinity condition than by UV-B stress. In addition, the upregulation of MAA biosynthetic genes was reported in parallel to the accumulation of mycosporine-2-glycine in response to salt stress [100]. When hypersalinity is present in the surrounding environment, it can result in cell dehydration and the production of ROS, leading to oxidative stress. Via the synthesis of MAAs under the salt stress condition, MAAs are considered to be osmotic solutes that are helping in osmotic regulation and reestablishing osmotic balance [83]. Furthermore, when aquatic species are challenged by cold environmental conditions, MAAs are found to act as osmotic protector [83].
Beyond salinity stress and UV radiation [83,98,100] other environmental stress factors were reported to influence MAA levels such as desiccation [101] and heat stress [6,75]. Seasonal changes in UV levels and temperature had a positive effect on MAA accumulation in phytoplankton and the copepod Cyclops abyssorum tatricus obtained from an alpine lake [102]. These different environmental stress factors contribute to oxidative stress and MAA accumulation. Consequently, the importance of MAAs and their role as antioxidants is strongly stimulated by changes in external environmental conditions [60].

4. Genes from the MAA Biosynthetic Pathway and Their Regulation

The shikimate pathway was the first pathway proposed to be responsible for MAA biosynthesis [6,94]. In this pathway, 3-dehydroquinate (DHQ) was used as a basis for the production of 4-deoxygadusol (4-DG), which is recognised as a direct precursor of MAAs, leading to the production of primary and secondary MAAs [8,96,103]. The second pathway was proposed by Balskus and Walsh [104] and suggested that synthesis of MAAs occurs via the pentose phosphate pathway and from another intermediate sedoheptulose 7-phosphate (SH 7-P) through the four-enzyme shinorine pathway. The role of this pathway in the synthesis of primary MAA shinorine in the cyanobacterium Anabaena variabilis ATCC 29413 was confirmed when the entire gene cluster (containing all four genes) was cloned and expressed in vitro using heterologous expression system in Escherichia coli. The 6.5-kb shinorine biosynthetic gene cluster included the following genes: dehydroquinate synthase (DHQS), O-methyltransferase (O-MT), adenosine triphosphate (ATP) grasp and a nonribosomalpeptide synthetase (NRPS) [104]. Via this pathway, 4-DG was initially produced thanks to the activity of DHQS and O-MT, followed by the second part of the pathway resulting in the generation of mycosporine-glycine and then shinorine. The same gene cluster has been also found in dinoflagellates [104], as well as in the sperm samples of the coral Acropora digitifera [48]. In coral dinoflagellates, three genes from the shinorine biosynthetic gene cluster were reported including two forms of O-MT gene, NRPS and ATP grasp homolog genes [9]. The lack of one gene from the shinorine biosynthetic pathway encoding DHQS reported for Symbiodiniaceae [9] may be due to limited sequence coverage, but could be potentially explained by the importance of host factors for the MAA biosynthesis [32,94]. Similar to coral dinoflagellate, the three-gene shinorine pathway was also reported in dinoflagellate Heterocapsa triquetra, in which 4-deoxygadusol (DDG) synthase-encoding gene was fused to the O-MT-encoding gene (Figure 3A) [4,105]. Furthermore, not all genes from the shinorine pathway (called the mys cluster) were identified in the 363 cyanobacterial genomes (Figure 3B) [106]. Huge genetic variability was observed among different cyanobacteria species, indicating that the formation of shinorine may happen due to the activity of different enzymes. The major genetic variability reported was in terms of presence or absence of two distinct enzymes NRPS (encoded by gene mysE; from Anabaena-type mys cluster) and D-Ala-D-Ala ligase (encoded by gene mysD from Nostoc-type mys cluster) in different cyanobacterial species [106]. Further variability was detected in terms of the presence of dehydrogenases or reductases and duplication of ATP-grasp genes [106,107]. Consequently, there are still considerable gaps in understanding the genetic diversity and regulatory mechanisms of the MAA biosynthetic pathways.

5. Potential Use of MAAs in Human Skin Protection

UV radiation can lead to substantial skin damage, particularly within epidermal cells, resulting in an increased incidence of skin cancer [108]. In humans, excessive exposure to UVR is associated with the development of over 95% of skin cancers [109]. The effect of UV radiation depends on the UV range, where high energy UVB usually leads to direct DNA damage and has a highly mutagenic and carcinogenic effect compared to lower energy UVA radiation [110]. Both UVA and UVB damage DNA directly and indirectly through oxidative stress [110]. A major mechanism leading to skin cancer occurs due to the UV-induced formation of cyclobutane pyrimidine dimers (CPDs) leading to DNA base damage [111]. The mechanisms of skin protection include mitigation strategies in reducing UV-induced damage, such as DNA repair mechanisms, the accumulation of antioxidants and UV-absorbing compounds [4]. In the human body, the pigment melanin helps in the protection from UV-induced damage due to its broad spectrum and its antioxidant activity [54]. However, there are two forms of melanin, the brown/black pigment eumelanin that plays a photoprotective role and the orange/yellow pigment pheomelanin that is considered to be photosensitizing [112]. Pheomelanin produces superoxide and nitric oxide resulting in the generation of ROS that can damage DNA [112]; as recently revealed, this damage even occurs in darkness, with increased CPD formation during the first 2–3 hours after UV exposure [113]. Also, in human skin, only 50%–75% of UVR is absorbed by melanin [53]. Subsequently, human skin needs additional protection via externally applied sunscreens. However, current chemical UVR protection is not adequate because these commercially available sunscreen products contain active ingredients that lack photostability, can produce free radicals leading to skin damage, irritation, skin aging and also can cause allergic reactions. Furthermore, the UV-filter compounds included in cosmetic products and used in the packing industry are disposed of, resulting in environmental pollution [114]. These organic compounds coming from wastewater reach groundwater and lead to the destruction of ecosystems such as coral reefs. Toxic effects of the sunscreens and specifically the UV-filter compound (oxybenzone, benzophenone-3), have been reported on marine life, particularly on tested coral planulae and cultured cells [115]. Consequently, biodegradable solutions, such as nature-based UV-filters like MAAs, should be used instead of the current chemical UV-filters and may help in the prevention of further environmental damage.
Interest in MAAs has been growing during the past decades. The use of MAAs for health products and cosmetics was recently reviewed by Chrapusta and colleagues [116]. MAAs are photostable, transparent compounds found in a wide range of aquatic species. As an old-new ecologically friendly option, MAAs show a stable photoprotective and antioxidative capacity against a range of UVR [117,118]. A number of research–based partnerships led to the generation of over 60 patents around the world [116]. The first application of MAAs, specifically shinorine, has been utilised in the product Helioguard®365 [119]. In vitro studies on the human keratinocyte cells confirmed the antiaging activity of the product when exposed to UV-A and stable activity of MAAs for a three-month period [119]. There are a few other products on the market that also utilise MAAs such as Helionori® and an increasing number of industry patents [116]. Still, more can be done in the market still dominated by chemical UV sunscreen products by enhancing industry–research collaborations to speed up the technological advancement of the MAA application as organic sunscreens.

6. Conclusions

Environmental pollution coming from the use of chemical UV filters is becoming an increasing problem that destabilizes the natural environment and contaminates groundwater. MAAs are excellent candidates for use in the cosmetic industry and for improved skin protection in an ecologically friendly way as water-soluble, colourless; highly diverse compounds. The MAAs’ ability to easily change form, as well as their UV properties and antioxidant capacities, present enormous biotechnological potential. The fact that MAA synthesis is adjustable in relation to changes in the environment, such as an increase in UV radiation, could be utilised further. Improving our understanding of the regulatory mechanisms and genetic diversity of MAA biosynthetic pathways will allow us to use MAAs in various biotechnological applications. Future research is needed to allow for better control of MAA biosynthesis and the production of adjustable, personalised natural sunscreens.

Funding

This research received no external funding.

Acknowledgments

The author would like to thank the two anonymous reviewers as well as Dr Jacinta Arellano and Isidora Skrlin for their critical reviews of this paper.

Conflicts of Interest

The author declares no conflicts of interest.

References

  1. Watanabe, S.; Sudo, K.; Nagashima, T.; Takemura, T.; Kawase, H.; Nozawa, T. Future projections of surface UV-B in a changing climate. J. Geophys. Res. Atmos. 2011, 116. [Google Scholar] [CrossRef]
  2. Reef, R.; Dunn, S.; Levy, O.; Dove, S.; Shemesh, E.; Bhckner, I.; Leggat, W.; Hoegh-Guldberg, O. Photoreactivation is the main repair pathway for UV-induced DNA damage in coral planulae. J. Exp. Biol. 2009, 212, 2760–2766. [Google Scholar] [CrossRef] [PubMed]
  3. Sinha, R.P.; Häder, D.P. UV-induced DNA damage and repair: A review. Photochem. Photobiol. Sci. 2002, 1, 225–236. [Google Scholar] [CrossRef] [PubMed]
  4. Korbee, N.; Teresa Mata, M.; Figueroa, F.L. Photoprotection mechanisms against ultraviolet radiation in Heterocapsa sp. (Dinophyceae) are influenced by nitrogen availability: Mycosporine-like amino acids vs. xanthophyll cycle. Limnol. Oceanogr. 2010, 55, 899–908. [Google Scholar] [CrossRef]
  5. Singh, S.P.; Klisch, M.; Sinha, R.P.; Häder, D.P. Genome mining of mycosporine-like amino acid (MAA) synthesizing and non-synthesizing cyanobacteria: A bioinformatics study. Genomics 2010, 95, 120–128. [Google Scholar] [CrossRef] [PubMed]
  6. Shick, J.M.; Dunlap, W.C. Mycosporine-like amino acids and related gadusols: Biosynthesis, accumulation; UV-protective functions in aquatic organisms. Annu. Rev. Physiol. 2002, 64, 223–262. [Google Scholar] [CrossRef] [PubMed]
  7. Sinha, R.P.; Singh, S.P.; Häder, D.P. Database on mycosporines and mycosporine-like amino acids (MAAs) in fungi, cyanobacteria, macroalgae, phytoplankton and animals. J. Photochem. Photobiol. B Biol. 2007, 89, 29–35. [Google Scholar] [CrossRef] [PubMed]
  8. Singh, S.P.; Kumari, S.; Rastogi, R.P.; Singh, K.L.; Sinha, R.P. Mycosporine-like amino acids (MAAs): Chemical structure, biosynthesis and significance as UV-absorbing/screening compounds. Indian J. Exp. Biol. 2008, 46, 7–17. [Google Scholar] [PubMed]
  9. Rosic, N.N. Phylogenetic analysis of genes involved in mycosporine-like amino acid biosynthesis in symbiotic dinoflagellates. Appl. Microbiol. Biotechnol. 2012, 94, 29–37. [Google Scholar] [CrossRef] [PubMed]
  10. Garcia-Pichel, F.; Castenholz, R.W. Occurrence of UV-absorbing, mycosporine-like compounds among cyanobacterial isolates and an estimate of their screening capacity. Appl. Environ. Microbiol. 1993, 59, 163–169. [Google Scholar] [PubMed]
  11. Garcia-Pichel, F.; Castenholz, R.W. Characterization and biological implications of scytonemin, a cyanobacterial sheath pigment. J. Phycol. 1991, 27, 395–409. [Google Scholar] [CrossRef]
  12. Klassen, J.L. Phylogenetic and evolutionary patterns in microbial carotenoid biosynthesis are revealed by comparative genomics. PLoS ONE 2010, 5, e11257. [Google Scholar] [CrossRef] [PubMed]
  13. Britton, G. Structure and properties of carotenoids in relation to function. FASEB J. 1995, 9, 1551–1558. [Google Scholar] [CrossRef] [PubMed]
  14. Dunlap, W.C.; Chalker, B.E. Identification and quantitation of near-UV absorbing compounds (S-320) in a hermatypic scleractinian. Coral Reefs 1986, 5, 155–159. [Google Scholar] [CrossRef]
  15. Nakamura, H.; Kobayashi, J.; Hirata, Y. Separation of mycosporine-like amino acids in marine organisms using reversed-phase high-performance liquid chromatography. J. Chromatogr. A 1982, 250, 113–118. [Google Scholar] [CrossRef]
  16. Carreto, J.I.; Carignan, M.O. Mycosporine-like amino acids: Relevant secondary metabolites. chemical and ecological aspects. Mar. Drugs 2011, 9, 387–446. [Google Scholar] [CrossRef] [PubMed]
  17. Rastogi, R.P.; Sinha, R.P.; Singh, S.P.; Häder, D.P. Photoprotective compounds from marine organisms. J. Ind. Microbiol. Biotechnol. 2010, 37, 537–558. [Google Scholar] [CrossRef] [PubMed]
  18. Bandaranayake, W.M. Mycosporines: Are they nature’s sunscreens? Nat. Prod. Rep. 1998, 15, 159–172. [Google Scholar] [CrossRef] [PubMed]
  19. Llewellyn, C.A.; Airs, R.L. Distribution and abundance of MAAs in 33 species of microalgae across 13 classes. Mar. Drugs 2010, 8, 1273–1291. [Google Scholar] [CrossRef] [PubMed]
  20. Sinha, R.P.; Häder, D.P. UV-protectants in cyanobacteria. Plant Sci. 2008, 174, 278–289. [Google Scholar] [CrossRef]
  21. Shibata, K. Pigments and a UV-absorbing substance in corals and a blue-green alga living in the great barrier reef. Plant Cell Physiol. 1969, 10, 325–335. [Google Scholar]
  22. Jain, S.; Prajapat, G.; Agrawal, A. A taxonomical study of mycosporine-like amino acids producing cyanobacteria. In Sunscreens: Source, Formulations, Efficacy and Recommendations; Rastogi, R.P., Ed.; Nova Science Publishers, Inc.: Hauppauge, NY, USA, 2018; pp. 131–151. [Google Scholar]
  23. Karsten, U.; Sawall, T.; Hanelt, D.; Bischof, K.; Figueroa, F.L.; Flores-Moya, A.; Wiencke, C. An inventory of UV-absorbing mycosporine-like amino acids in macroalgae from polar to warm-temperate regions. Bot. Mar. 1998, 41, 443–453. [Google Scholar] [CrossRef]
  24. Dunlap, W.C.; Yamamoto, Y. Small-molecule antioxidants in marine organisms: Antioxidant activity of mycosporine-glycine. Comp. Biochem. Physiol. Part B Biochem. Mol. Biol. 1995, 112, 105–114. [Google Scholar] [CrossRef]
  25. Figueroa, F.L.; Bueno, A.; Korbee, N.; Santos, R.; Mata, L.; Schuenhoff, A. Accumulation of mycosporine-like amino acids in asparagopsis armata grown in tanks with fishpond effluents of gilthead sea bream, Sparus aurata. J. World Aquac. Soc. 2008, 39, 692–699. [Google Scholar] [CrossRef]
  26. Figueroa, F.L.; Escassi, L.; Pérez-Rodríguez, E.; Korbee, N.; Giles, A.D.; Johnsen, G. Effects of short-term irradiation on photoinhibition and accumulation of mycosporine-like amino acids in sun and shade species of the red algal genus Porphyra. J. Photochem. Photobiol B Biol. 2003, 69, 21–30. [Google Scholar] [CrossRef]
  27. Lalegerie, F.; Lajili, S.; Bedoux, G.; Taupin, L.; Stiger-Pouvreau, V.; Connan, S. Photoprotective compounds in red macroalgae from Brittany: Considerable diversity in mycosporine-like amino acids (MAAs). Mar. Environ. Res. 2019, 147, 37–48. [Google Scholar] [CrossRef] [PubMed]
  28. Orfanoudaki, M.; Hartmann, A.; Karsten, U.; Ganzera, M. Chemical profiling of mycosporine-like amino acids in twenty-three red algal species. J. Phycol. 2019, 55, 393–403. [Google Scholar] [CrossRef] [PubMed]
  29. Favre-Bonvin, J.; Bernillon, J.; Salin, N.; Arpin, N. Biosynthesis of mycosporines: Mycosporine glutaminol in Trichothecium roseum. Phytochemistry 1987, 26, 2509–2514. [Google Scholar] [CrossRef]
  30. Figueroa, F.L.; Bonomi Barufi, J.; Malta, E.J.; Conde-Álvarez, R.; Nitschke, U.; Arenas, F.; Mata, M.; Connan, S.; Abreu, M.H.; Marquardt, R.; et al. Short-term effects of increasing CO2, nitrate and temperature on three mediterranean macroalgae: Biochemical composition. Aquat. Biol. 2014, 22, 177–193. [Google Scholar] [CrossRef]
  31. 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]
  32. Banaszak, A.T.; Barba Santos, M.G.; LaJeunesse, T.C.; Lesser, M.P. The distribution of mycosporine-like amino acids (MAAs) and the phylogenetic identity of symbiotic dinoflagellates in cnidarian hosts from the Mexican Caribbean. J. Exp. Mar. Biol. Ecol. 2006, 337, 131–146. [Google Scholar] [CrossRef]
  33. Vale, P. Effects of Light and Salinity Stresses in Production of Mycosporine-Like Amino Acids by Gymnodinium catenatum (Dinophyceae). Photochem. Photobiol. 2015, 91, 1112–1122. [Google Scholar] [CrossRef] [PubMed]
  34. Vale, P. Can Mycosporine-Like Amino Acids Act as Multifunctional Compounds in Gymnodinium catenatum (Dinophyceae)? Photochem. Photobiol. 2016, 92, 264–275. [Google Scholar] [CrossRef] [PubMed]
  35. Bandaranayake, W.M.; Bemis, J.E.; Bourne, D.J. Ultraviolet absorbing pigments from the marine sponge Dysidea herbacea: Isolation and structure of a new mycosporine. Comp. Biochem. Physiol. Part C Pharmacol. Toxicol. Endocrinol. 1996, 115, 281–286. [Google Scholar] [CrossRef]
  36. Oda, Y.; Zhang, Q.; Matsunaga, S.; Fujita, M.J.; Sakai, R. Two new mycosporine-like amino acids LC-343 and mycosporine-ethanolamine from the micronesian marine sponge Lendenfeldia chondrodes. Chem. Lett. 2017, 46, 1272–1274. [Google Scholar] [CrossRef]
  37. Ferrier-Pagès, C.; Richard, C.; Forcioli, D.; Allemand, D.; Pichon, M.; Shick, J.M. Effects of temperature and UV radiation increases on the photosynthetic efficiency in four scleractinian coral species. Biol. Bull. 2007, 213, 76–87. [Google Scholar] [CrossRef] [PubMed]
  38. Yakovleva, I.M.; Baird, A.H. Ontogenetic change in the abundance of mycosporine-like amino acids in non-zooxanthellate coral larvae. Coral Reefs 2005, 24, 443–452. [Google Scholar] [CrossRef]
  39. Adams, N.L.; Shick, J.M. Mycosporine-like amino acids prevent UVB-induced abnormalities during early development of the green sea urchin Strongylocentrotus droebachiensis. Mar. Biol. 2001, 138, 267–280. [Google Scholar] [CrossRef]
  40. Carroll, A.K.; Shick, J.M. Dietary accumulation of UV-absorbing mycosporine-like amino acids (MAAs) by the green sea urchin (Strongylocentrotus droebachiensis). Mar. Biol. 1996, 124, 561–569. [Google Scholar] [CrossRef]
  41. Eckes, M.; Dove, S.; Siebeck, U.E.; Grutter, A.S. Fish mucus versus parasitic gnathiid isopods as sources of energy and sunscreens for a cleaner fish. Coral Reefs 2015, 34, 823–833. [Google Scholar] [CrossRef]
  42. Newman, S.J.; Dunlap, W.C.; Nicol, S.; Ritz, D. Antarctic krill (Euphausia superba) acquire a UV-absorbing mycosporine-like amino acid from dietary algae. J. Exp. Mar. Biol. Ecol. 2000, 255, 93–110. [Google Scholar] [CrossRef]
  43. Osborn, A.R.; Almabruk, K.H.; Holzwarth, G.; Asamizu, S.; LaDu, J.; Kean, K.M.; Karplus, P.A.; Tanguay, R.L.; Bakalinsky, A.T.; Mahmud, T. De novo synthesis of a sunscreen compound in vertebrates. Elife 2015, 4, e05919. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Brotherton, C.A.; Balskus, E.P. Shedding light on sunscreen biosynthesis in zebrafish. eLife 2015, 4, e07961. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Rosic, N.N.; Braun, C.; Kvaskoff, D. Extraction and Analysis of Mycosporine-Like Amino Acids in Marine Algae. Methods Mol. Biol. 2015, 1308, 119–129. [Google Scholar] [PubMed]
  46. Orfanoudaki, M.; Hartmann, A.; Miladinovic, H.; Ngoc, H.N.; Karsten, U.; Ganzera, M. Bostrychines A–F, Six Novel Mycosporine-Like Amino-Acids and a Novel Betaine from the Red Alga Bostrychia scorpioides. Mar. Drugs 2019, 17, 356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Rastogi, R.P.; Incharoensakdi, A. UV radiation-induced biosynthesis, stability and antioxidant activity of mycosporine-like amino acids (MAAs) in a unicellular cyanobacterium Gloeocapsa sp. CU2556. J. Photochem. Photobiol. B Biol. 2014, 130, 287–292. [Google Scholar] [CrossRef] [PubMed]
  48. Shinzato, C.; Shoguchi, E.; Kawashima, T.; Hamada, M.; Hisata, K.; Tanaka, M.; Fujie, M.; Fujiwara, M.; Koyanagi, R.; Ikuta, T.; et al. Using the Acropora digitifera genome to understand coral responses to environmental change. Nature 2011, 476, 320–323. [Google Scholar] [CrossRef] [PubMed]
  49. Carignan, M.O.; Cardozo, K.H.M.; Oliveira-Silva, D.; Colepicolo, P.; Carreto, J.I. Palythine-threonine, a major novel mycosporine-like amino acid (MAA) isolated from the hermatypic coral Pocillopora capitata. J. Photochem. Photobiol. B Biol. 2009, 94, 191–200. [Google Scholar] [CrossRef] [PubMed]
  50. 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]
  51. Ngoennet, S.; Nishikawa, Y.; Hibino, T.; Waditee-Sirisattha, R.; Kageyama, H. A method for the isolation and characterization of mycosporine-like amino acids from cyanobacteria. Methods Protoc. 2018, 1, 46. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Kageyama, H.; Waditee-Sirisattha, R. Antioxidative, anti-inflammatory; 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] [PubMed] [Green Version]
  53. Kaidbey, K.H.; Agin, P.P.; Sayre, R.M.; Kligman, A.M. Photoprotection by melanin—A comparison of black and Caucasian skin. J. Am. Acad. Dermatol. 1979, 1, 249–260. [Google Scholar] [CrossRef]
  54. Brenner, M.; Hearing, V.J. The protective role of melanin against UV damage in human skin. Photochem. Photobiol. 2008, 84, 539–549. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Jain, S.; Prajapat, G.; Abrar, M.; Ledwani, L.; Singh, A.; Agrawal, A. Cyanobacteria as efficient producers of mycosporine-like amino acids. J. Basic Microbiol. 2017, 57, 715–727. [Google Scholar] [CrossRef] [PubMed]
  56. Rastogi, R.P.; Sinha, R.P.; Incharoensakdi, A. Partial characterization, UV-induction and photoprotective function of sunscreen pigment, scytonemin from Rivularia sp. HKAR-4. Chemosphere 2013, 93, 1874–1878. [Google Scholar] [CrossRef] [PubMed]
  57. Rastogi, R.P.; Sonani, R.R.; Madamwar, D. Cyanobacterial Sunscreen Scytonemin: Role in Photoprotection and Biomedical Research. Appl. Biochem. Biotechnol. 2015, 176, 1551–1563. [Google Scholar] [CrossRef] [PubMed]
  58. Pathak, J.; Pandey, A.; Maurya, P.K.; Rajneesh, R.; Sinha, R.P.; Singh, S.P. Cyanobacterial Secondary Metabolite Scytonemin: A Potential Photoprotective and Pharmaceutical Compound. Proc. Natl. Acad. Sci. India Sect. B Biol. Sci. 2019, 89, 1–15. [Google Scholar] [CrossRef]
  59. Garcia-Pichel, F.; Bebout, B.M. Penetration of ultraviolet radiation into shallow water sediments: High exposure for photosynthetic communities. Mar. Ecol. Prog. Ser. 1996, 131, 257–262. [Google Scholar] [CrossRef]
  60. Wada, N.; Sakamoto, T.; Matsugo, S. Mycosporine-like amino acids and their derivatives as natural antioxidants. Antioxidants 2015, 4, 603–646. [Google Scholar] [CrossRef] [PubMed]
  61. De La Coba, F.; Aguilera, J.; Figueroa, F.L.; De Gálvez, M.V.; Herrera, E. Antioxidant activity of mycosporine-like amino acids isolated from three red macroalgae and one marine lichen. J. Appl. Phycol. 2009, 21, 161–169. [Google Scholar] [CrossRef]
  62. Rastogi, R.P.; Richa; Singh, S.P.; Häder, D.P.; Sinha, R.P. Mycosporine-like amino acids profile and their activity under PAR and UVR in a hot-spring cyanobacterium Scytonema sp. HKAR-3. Aust. J. Bot. 2010, 58, 286–293. [Google Scholar] [CrossRef]
  63. Suh, H.J.; Lee, H.W.; Jung, J. Mycosporine Glycine Protects Biological Systems Against Photodynamic Damage by Quenching Singlet Oxygen with a High Efficiency. Photochem. Photobiol. 2003, 78, 109–113. [Google Scholar] [CrossRef]
  64. Gacesa, R.; Lawrence, K.P.; Georgakopoulos, N.D.; Yabe, K.; Dunlap, W.C.; Barlow, D.J.; Wells, G.; Young, A.R.; Long, P.F. The mycosporine-like amino acids porphyra-334 and shinorine are antioxidants and direct antagonists of Keap1-Nrf2 binding. Biochimie 2018, 154, 35–44. [Google Scholar] [CrossRef] [PubMed]
  65. Becker, K.; Hartmann, A.; Ganzera, M.; Fuchs, D.; Gostner, J.M. Immunomodulatory Effects of the Mycosporine-Like Amino Acids Shinorine and Porphyra-334. Mar. Drugs 2016, 14, 119. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Hartmann, A.; Gostner, J.; Fuchs, J.E.; Chaita, E.; Aligiannis, N.; Skaltsounis, L.; Ganzera, M. Inhibition of Collagenase by Mycosporine-like Amino Acids from Marine Sources. Planta Med. 2015, 81, 813–820. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Yoshiki, M.; Tsuge, K.; Tsuruta, Y.; Yoshimura, T.; Koganemaru, K.; Sumi, T.; Matsui, T.; Matsumoto, K. Production of new antioxidant compound from mycosporine-like amino acid, porphyra-334 by heat treatment. Food Chem. 2009, 113, 1127–1132. [Google Scholar] [CrossRef]
  68. Ryu, J.; Park, S.J.; Kim, I.H.; Choi, Y.H.; Nam, T.J. Protective effect of porphyra-334 on UVA-induced photoaging in human skin fibroblasts. Int. J. Mol. Med. 2014, 34, 796–803. [Google Scholar] [CrossRef] [PubMed]
  69. Cheewinthamrongrod, V.; Kageyama, H.; Palaga, T.; Takabe, T.; Waditee-Sirisattha, R. DNA damage protecting and free radical scavenging properties of mycosporine-2-glycine from the Dead Sea cyanobacterium in A375 human melanoma cell lines. J. Photochem. Photobiol. B Biol. 2016, 164, 289–295. [Google Scholar] [CrossRef] [PubMed]
  70. Tarasuntisuk, S.; Palaga, T.; Kageyama, H.; Waditee-Sirisattha, R. Mycosporine-2-glycine exerts anti-inflammatory and antioxidant effects in lipopolysaccharide (LPS)-stimulated RAW 264.7 macrophages. Arch. Biochem. Biophys. 2019, 662, 33–39. [Google Scholar] [CrossRef] [PubMed]
  71. Lawrence, K.P.; Gacesa, R.; Long, P.F.; Young, A.R. Molecular photoprotection of human keratinocytes in vitro by the naturally occurring mycosporine-like amino acid palythine. Br. J. Dermatol. 2018, 178, 1353–1363. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Dunlap, W.C.; Shick, J.M. Ultraviolet radiation-absorbing mycosporine-like amino acids in coral reef organisms: A biochemical and environmental perspective. J. Phycol. 1998, 34, 418–430. [Google Scholar] [CrossRef]
  73. Suzuki, T.; Yamamoto, M. Molecular basis of the Keap1–Nrf2 system. Free Radic. Biol. Med. 2015, 88, 93–100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Panich, U.; Sittithumcharee, G.; Rathviboon, N.; Jirawatnotai, S. Ultraviolet radiation-induced skin aging: The role of DNA damage and oxidative stress in epidermal stem cell damage mediated skin aging. Stem Cells Int. 2016, 2016. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Yakovleva, I.; Bhagooli, R.; Takemura, A.; Hidaka, M. Differential susceptibility to oxidative stress of two scleractinian corals: Antioxidant functioning of mycosporine-glycine. Comp. Biochem. Physiol. B Biochem. Mol. Biol. 2004, 139, 721–730. [Google Scholar] [CrossRef] [PubMed]
  76. Ma, Q. Role of Nrf2 in oxidative stress and toxicity. Annu. Rev. Pharmacol. Toxicol. 2013, 53, 401–426. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Siomek, A. NF-κB signaling pathway and free radical impact. Acta Biochim. Pol. 2012, 59, 323–331. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Álvarez-Gómez, F.; Korbee, N.; Casas-Arrojo, V.; Abdala-Díaz, R.T.; Figueroa, F.L. UV photoprotection, cytotoxicity and immunology capacity of red algae extracts. Molecules 2019, 24, 341. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Grimbaldeston, M.A.; Simpson, A.; Finlay-Jones, J.J.; Hart, P.H. The effect of ultraviolet radiation exposure on the prevalence of mast cells in human skin. Br. J. Dermatol. 2003, 148, 300–306. [Google Scholar] [CrossRef] [PubMed]
  80. Bhandari, R.; Sharma, P.K. Effect of UV-B and high visual radiation on photosynthesis in freshwater (Nostoc spongiaeforme) and marine (Phormidium corium) cyanobacteria. Indian J. Biochem. Biophys. 2007, 44, 231–239. [Google Scholar] [PubMed]
  81. Gao, K.; Xu, J. Effects of solar UV radiation on diurnal photosynthetic performance and growth of Gracilaria lemaneiformis (Rhodophyta). Eur. J. Phycol. 2008, 43, 297–307. [Google Scholar] [CrossRef]
  82. Rui, Y.; Zhaohui, Z.; Wenshan, S.; Bafang, L.; Hu, H. Protective effect of MAAs extracted from Porphyra tenera against UV irradiation-induced photoaging in mouse skin. J. Photochem. Photobiol. B Biol. 2019, 192, 26–33. [Google Scholar] [CrossRef] [PubMed]
  83. Oren, A.; Gunde-Cimerman, N. Mycosporines and mycosporine-like amino acids: UV protectants or multipurpose secondary metabolites? FEMS Microbiol. Lett. 2007, 269, 1–10. [Google Scholar] [CrossRef] [PubMed]
  84. Michalek-Wagner, K.; Willis, B.L. Impacts of bleaching on the soft coral Lobophytum compactum. Ii. Biochemical changes in adults and their eggs. Coral Reefs 2001, 19, 240–246. [Google Scholar] [CrossRef]
  85. Al-Utaibi, A.A.; Niaz, G.R.; Al-Lihaibi, S.S. Mycosporine-like amino acids in six scleractinian coral species. Oceanologia 2009, 51, 93–104. [Google Scholar] [CrossRef] [Green Version]
  86. Bohm, G.A.; Pfleiderer, W.; Boger, P.; Scherer, S. Structure of a novel oligosaccharide-mycosporine-amino acid ultraviolet A/B sunscreen pigment from the terrestrial cyanobacterium Nostoc commune. J. Biol. Chem. 1995, 270, 8536–8539. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Hernando, M.; Carreto, J.I.; Carignan, M.O.; Ferreyra, G.A.; Gross, C. Effects of solar radiation on growth and mycosporine-like amino acids content in Thalassiosira sp, an Antarctic diatom. Polar Biol. 2002, 25, 12–20. [Google Scholar]
  88. Laurion, I.; Roy, S. Growth and photoprotection in three dinoflagellates (including two strains of Alexandrium tamarense) and one diatom exposed to four weeks of natural and enhanced ultraviolet-b radiation. J. Phycol. 2009, 45, 16–33. [Google Scholar] [CrossRef] [PubMed]
  89. Riegger, L.; Robinson, D. Photoinduction of UV-absorbing compounds in Antarctic diatoms and Phaeocystis antarctica. Mar. Ecol. Prog. Ser. 1997, 160, 13–25. [Google Scholar] [CrossRef] [Green Version]
  90. Taira, H.; Aoki, S.; Yamanoha, B.; Taguchi, S. Daily variation in cellular content of UV-absorbing compounds mycosporine-like amino acids in the marine dinoflagellate Scrippsiella sweeneyae. J. Photochem. Photobiol. B Biol. 2004, 75, 145–155. [Google Scholar] [CrossRef]
  91. Franklin, L.A.; Kräbs, G.; Kuhlenkamp, R. Blue light and UV-A radiation control the synthesis of mycosporine-like amino acids in Chondrus crispus (Florideophyceae). J. Phycol. 2001, 37, 257–270. [Google Scholar] [CrossRef] [Green Version]
  92. Sinha, R.P.; Ambasht, N.K.; Sinha, J.P.; Klisch, M.; Häder, D.P. UV-B-induced synthesis of mycosporine-like amino acids in three strains of Nodularia (cyanobacteria). J. Photochem. Photobiol. B Biol. 2003, 71, 51–58. [Google Scholar] [CrossRef] [PubMed]
  93. Lesser, M.P. Depth-dependent photoacclimatization to solar ultraviolet radiation in the Caribbean coral Montastraea faveolata. Mar. Ecol. Prog. Ser. 2000, 192, 137–151. [Google Scholar] [CrossRef]
  94. Shick, J.M.; Romaine-Lioud, S.; Ferrier-Pagès, C.; Gattuso, J.P. Ultraviolet-B radiation stimulates shikimate pathway-dependent accumulation of mycosporine-like amino acids in the coral Stylophora pistillata despite decreases in its population of symbiotic dinoflagellates. Limnol. Oceanogr. 1999, 44, 1667–1682. [Google Scholar] [CrossRef] [Green Version]
  95. Banaszak, A.T.; Lesser, M.P. Effects of solar ultraviolet radiation on coral reef organisms. Photochem. Photobiol. Sci. 2009, 8, 1276–1294. [Google Scholar] [CrossRef] [PubMed]
  96. Portwich, A.; Garcia-Pichel, F. Biosynthetic pathway of mycosporines (mycosporine-like amino acids) in the cyanobacterium Chlorogloeopsis sp. strain PCC 6912. Phycologia 2003, 42, 384–392. [Google Scholar] [CrossRef]
  97. Rath, J.; Mandal, S.; Adhikary, S.P. Ecophysiology of the estuarine cyanobacterium Lyngbya aestuarii to varying salinity in vitro. Acta Physiol. Plant. 2014, 36, 409–419. [Google Scholar] [CrossRef]
  98. Singh, S.P.; Klisch, M.; Sinha, R.P.; Häder, D.P. Effects of abiotic stressors on synthesis of the mycosporine-like amino acid shinorine in the cyanobacterium Anabaena variabilis PCC 7937. Photochem. Photobiol. 2008, 84, 1500–1505. [Google Scholar] [CrossRef] [PubMed]
  99. Whitehead, K.; Hedges, J.I. Electrospray ionization tandem mass spectrometric and electron impact mass spectrometric characterization of mycosporine-like amino acids. Rapid Commun. Mass Spectrom. 2003, 17, 2133–2138. [Google Scholar] [CrossRef] [PubMed]
  100. Waditee-Sirisattha, R.; Kageyama, H.; Sopun, W.; Tanaka, Y.; Takabe, T. Identification and upregulation of biosynthetic genes required for accumulation of mycosporine-2-glycine under salt stress conditions in the halotolerant cyanobacterium Aphanothece halophytica. Appl. Environ. Microbiol. 2014, 80, 1763–1769. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Feng, Y.N.; Zhang, Z.C.; Feng, J.L.; Qiu, B.S. Effects of UV-B radiation and periodic desiccation on the morphogenesis of the edible terrestrial cyanobacterium Nostoc flagelliforme. Appl. Environ. Microbiol. 2012, 78, 7075–7081. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Tartarotti, B.; Sommaruga, R. Seasonal and ontogenetic changes of mycosporine-like amino acids in planktonic organisms from an alpine lake. Limnol. Oceanogr. 2006, 51, 1530–1541. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Rosic, N.N.; Dove, S. Mycosporine-like amino acids from coral dinoflagellates. Appl. Environ. Microbiol. 2011, 77, 8478–8486. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Balskus, E.P.; Walsh, C.T. The genetic and molecular basis for sunscreen biosynthesis in cyanobacteria. Science 2010, 329, 1653–1656. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Miyamoto, K.T.; Komatsu, M.; Ikeda, H. Discovery of gene cluster for mycosporine-like amino acid biosynthesis from Actinomycetales microorganisms and production of a novel mycosporine-like amino acid by heterologous expression. Appl. Environ. Microbiol. 2014, 80, 5028–5036. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. D’Agostino, P.M.; Woodhouse, J.N.; Liew, H.T.; Sehnal, L.; Pickford, R.; Wong, H.L.; Burns, B.P.; Neilan, B.A. Bioinformatic, phylogenetic and chemical analysis of the UV-absorbing compounds scytonemin and mycosporine-like amino acids from the microbial mat communities of Shark Bay, Australia. Environ. Microbiol. 2019, 21, 702–715. [Google Scholar] [CrossRef] [PubMed]
  107. Gao, Q.; Garcia-Pichel, F. An ATP-Grasp ligase involved in the last biosynthetic step of the iminomycosporine shinorine in Nostoc punctiforme ATCC 29133. J. Bacteriol. 2011, 193, 5923–5928. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Miller, A.J.; Mihm, M.C., Jr. Melanoma. N. Engl. J. Med. 2006, 355, 51–65. [Google Scholar] [CrossRef] [PubMed]
  109. Savoye, I.; Olsen, C.M.; Whiteman, D.C.; Bijon, A.; Wald, L.; Dartois, L.; Clavel-Chapelon, F.; Boutron-Ruault, M.C.; Kvaskoff, M. Patterns of Ultraviolet Radiation Exposure and Skin Cancer Risk: The E3N-SunExp Study. J. Epidemiol. 2018, 28, 27–33. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Ichihashi, M.; Ueda, M.; Budiyanto, A.; Bito, T.; Oka, M.; Fukunaga, M.; Tsuru, K.; Horikawa, T. UV-induced skin damage. Toxicology 2003, 189, 21–39. [Google Scholar] [CrossRef]
  111. Ikehata, H.; Ono, T. The mechanisms of UV mutagenesis. J. Radiat. Res. 2011, 52, 115–125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Napolitano, A.; Panzella, L.; Monfrecola, G.; d’Ischia, M. Pheomelanin-induced oxidative stress: Bright and dark chemistry bridging red hair phenotype and melanoma. Pigment. Cell Melanoma Res. 2014, 27, 721–733. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Premi, S.; Wallisch, S.; Mano, C.M.; Weiner, A.B.; Bacchiocchi, A.; Wakamatsu, K.; Bechara, E.J.; Halaban, R.; Douki, T.; Brash, D.E. Photochemistry. Chemiexcitation of melanin derivatives induces DNA photoproducts long after UV exposure. Science 2015, 347, 842–847. [Google Scholar] [CrossRef] [PubMed]
  114. Rykowska, I.; Wasiak, W. Research trends on emerging environment pollutants—A review. Open Chem. 2015, 13, 1353–1370. [Google Scholar] [CrossRef]
  115. Downs, C.A.; Kramarsky-Winter, E.; Segal, R.; Fauth, J.; Knutson, S.; Bronstein, O.; Ciner, F.R.; Jeger, R.; Lichtenfeld, Y.; Woodley, C.M.; et al. Toxicopathological Effects of the Sunscreen UV Filter, Oxybenzone (Benzophenone-3), on Coral Planulae and Cultured Primary Cells and Its Environmental Contamination in Hawaii and the U.S. Virgin Islands. Arch. Environ. Contam. Toxicol. 2016, 70, 265–288. [Google Scholar] [CrossRef] [PubMed]
  116. Chrapusta, E.; Kaminski, A.; Duchnik, K.; Bober, B.; Adamski, M.; Bialczyk, J. Mycosporine-Like Amino Acids: Potential health and beauty ingredients. Mar. Drugs 2017, 15, 326. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Lawrence, K.P.; Long, P.F.; Young, A.R. Mycosporine-like amino acids for skin photoprotection. Curr. Med. Chem. 2018, 25, 5512–5527. [Google Scholar] [CrossRef] [PubMed]
  118. Gröniger, A.; Häder, D.P. Stability of mycosporine-like amino acids. Recent Res. Dev. Photochem. Photobiol. 2000, 4, 247–252. [Google Scholar]
  119. Schmidt, E.W. An Enzymatic Route to Sunscreens. ChemBioChem 2011, 12, 363–365. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Chemical structures of mycosporine-like amino acids (MAAs): (A) MAA core composed of a cyclohexenone, a cyclohexenone, or cyclohexenimine ring conjugated to an amino acid residue or its imino alcohol; (B) MAA precursor 4-deoxygadusol, plus the primary MAAs found including mycosporine-glycine, mycosporine-2-glycine, shinorine, palythine and porphyra-334, including the maximum absorbance values.
Figure 1. Chemical structures of mycosporine-like amino acids (MAAs): (A) MAA core composed of a cyclohexenone, a cyclohexenone, or cyclohexenimine ring conjugated to an amino acid residue or its imino alcohol; (B) MAA precursor 4-deoxygadusol, plus the primary MAAs found including mycosporine-glycine, mycosporine-2-glycine, shinorine, palythine and porphyra-334, including the maximum absorbance values.
Marinedrugs 17 00638 g001
Figure 2. Chemical structures and masses (m/z) of commonly found MAAs in the red alga (indicated by #), Symbiodiniaceae (indicated by $) and the hermatypic coral Stylophora pistillata (indicated by *) as adapted from Rosic et al. [45].
Figure 2. Chemical structures and masses (m/z) of commonly found MAAs in the red alga (indicated by #), Symbiodiniaceae (indicated by $) and the hermatypic coral Stylophora pistillata (indicated by *) as adapted from Rosic et al. [45].
Marinedrugs 17 00638 g002
Figure 3. Gene clusters for mycosporines and MAAs biosynthesis. (A) The MAA gene clusters from eight different species including representatives from bacteria, algae, fungi; cnidarians (adapted from Miyamoto et al. [105]). (B) Gene clusters for MAA biosynthesis identified in different cyanobacterial species (adapted from reference [106]).
Figure 3. Gene clusters for mycosporines and MAAs biosynthesis. (A) The MAA gene clusters from eight different species including representatives from bacteria, algae, fungi; cnidarians (adapted from Miyamoto et al. [105]). (B) Gene clusters for MAA biosynthesis identified in different cyanobacterial species (adapted from reference [106]).
Marinedrugs 17 00638 g003
Table 1. Properties of main MAAs based on in vitro studies using cell cultures exposed to individual MAAs isolated from various species.
Table 1. Properties of main MAAs based on in vitro studies using cell cultures exposed to individual MAAs isolated from various species.
MAAsActivity:
UV-Absorbing
Activity:
Antioxidative
Activity: Anti- Inflammatory Activity: Antiaging Sources of MAAs and References
Mycosporine-glycineYes
ʎmax = 310 nm
Yes [24,31,51,61,63]Yes [31]Yes [31]Red alga Porphyra tenera [24]
Green alga Chlamydomonas hedleyi [31]
Ascidian Lissoclinum patella [63]
Marine lichen Lichina pygmaea [61]
ShinorineYes
ʎmax = 333 nm
Yes [24,51,64]

No [31]
Yes [31]

No [65]
Yes [31,66]Red alga Chondrus yendoi [64]
Red alga Porphyra sp. [65]
Green alga Chlamydomonas hedleyi [31]
Red algae Porphyra sp. & Palmaria palmate [66]
Porphyra-334Yes
ʎmax = 334 nm
Yes [24,51,64,67,68]
No [31]
Yes [65,68]

No [31]
Yes [31,66]

No [31]
Green alga Chlamydomonas hedleyi [31]
Red algae Porphyra sp. & Palmaria palmate [66]
Red alga Porphyra yezoensis [68]
Mycosporine-2-glycineYes
ʎmax = 332 nm
Yes [51,69,70]Yes [70]Yes [70]Cyanobacterium Aphanothece halophytica [51,69,70]
PalythineYes
ʎmax = 320 nm
Yes [71]

No [69]
-Yes [66]Red alga Chrondus yendoi [71]
Cyanobacterium Aphanothece halophytica [69]
Red algae Porphyra sp. & Palmaria palmate [66]
Cyanobacterium Aphanothece halophytica [69]

Share and Cite

MDPI and ACS Style

Rosic, N.N. Mycosporine-Like Amino Acids: Making the Foundation for Organic Personalised Sunscreens. Mar. Drugs 2019, 17, 638. https://doi.org/10.3390/md17110638

AMA Style

Rosic NN. Mycosporine-Like Amino Acids: Making the Foundation for Organic Personalised Sunscreens. Marine Drugs. 2019; 17(11):638. https://doi.org/10.3390/md17110638

Chicago/Turabian Style

Rosic, Nedeljka N. 2019. "Mycosporine-Like Amino Acids: Making the Foundation for Organic Personalised Sunscreens" Marine Drugs 17, no. 11: 638. https://doi.org/10.3390/md17110638

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

Article Metrics

Back to TopTop