Next Article in Journal
Cloning and Characterization of a Novel Endo-Type Metal-Independent Alginate Lyase from the Marine Bacteria Vibrio sp. Ni1
Previous Article in Journal
Therapeutic Potential of Marine Bioactive Peptides against Human Immunodeficiency Virus: Recent Evidence, Challenges, and Future Trends
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Genome Mining as an Alternative Way for Screening the Marine Organisms for Their Potential to Produce UV-Absorbing Mycosporine-like Amino Acid

1
Faculty of Health, Southern Cross University, Gold Coast, QLD 4225, Australia
2
Marine Ecology Research Centre, Southern Cross University, Lismore, NSW 2480, Australia
Mar. Drugs 2022, 20(8), 478; https://doi.org/10.3390/md20080478
Submission received: 15 June 2022 / Revised: 21 July 2022 / Accepted: 21 July 2022 / Published: 26 July 2022

Abstract

:
Mycosporine-like amino acids (MAAs) are small molecules with robust ultraviolet (UV)-absorbing capacities and a huge potential to be used as an environmentally friendly natural sunscreen. MAAs, temperature, and light-stable compounds demonstrate powerful photoprotective capacities and the ability to capture light in the UV-A and UV-B ranges without the production of damaging free radicals. The biotechnological uses of these secondary metabolites have been often limited by the small quantities restored from natural resources, variation in MAA expression profiles, and limited success in heterologous expression systems. Overcoming these obstacles requires a better understanding of MAA biosynthesis and its regulatory processes. MAAs are produced to a certain extent via a four-enzyme pathway, including genes encoding enzymes dehydroquinate synthase, enzyme O-methyltransferase, adenosine triphosphate grasp, and a nonribosomal peptide synthetase. However, there are substantial genetic discrepancies in the MAA genetic pathway in different species, suggesting further complexity of this pathway that is yet to be fully explored. In recent years, the application of genome-mining approaches allowed the identification of biosynthetic gene clusters (BGCs) that resulted in the discovery of many new compounds from unconventional sources. This review explores the use of novel genomics tools for linking BGCs and secondary metabolites based on the available omics data, including MAAs, and evaluates the potential of using novel genome-mining tools to reveal a cryptic potential for new bioproduct screening approaches and unrevealing new MAA producers.

1. Introduction

Mycosporine-like amino acids (MAAs) are natural sunscreens and the most common group of secondary metabolites among aquatic organisms protecting them from harmful ultraviolet radiation (UVR) [1]. MAAs are discovered in various marine and freshwater species, including micro and macroalgae, cyanobacteria, and aquatic animals ranging from invertebrates to vertebrates [2,3,4,5,6,7,8,9]. These small hydrophilic compounds (<400 Da) are characterised by a core molecule made of cyclohexenone or cyclohexenimine rings accompanied by additional radical groups [10,11,12]. The combinations of different radical groups with the core molecule resulted in the identification of more than 30 different MAAs in nature [9,12,13,14]. MAAs are characterised by different UV properties and additional bioactivities (see Table 1) [2,11,15,16].
In marine and freshwater organisms, UV protection occurs at multiple levels (Figure 1). Beyond UV-absorbing compounds conducting the UV screening, there are also additional mechanisms employed against damaging UVR, including quenching (i.e., enzymatic and non-enzymatic) and repair mechanisms (e.g., DNA repair) [30]. The synthesis of different UV-absorbing MAAs provides the main mechanism for efficient UV photoprotection (280–400 nm) [11]. MAAs also demonstrate additional pharmacologically relevant properties such as antioxidative, anti-inflammatory, and antiaging capacities [31]. In addition, MAAs were found to play a role in osmotic regulation and in cellular protection during drought periods [11,14,32,33,34]. MAAs can absorb light in the range of UV-A (315–400 nm), which makes ~95% of UV energy that penetrates the atmosphere and UV-B (280–315 nm) [12,35,36]. Prolonged sun exposure, specifically UVR, results in the production of free radicals and oxidative stress leading to potential damage to cellular structures and DNA [37]. MAA secondary metabolites reduce oxidative cellular damage due to their antioxidative properties and ability to scavenge free radicals [31,38]. The MAA UV-protective plasticity occurs due to the ability to produce various MAAs and to increase their quantities during the summer season under higher levels of UVR [31,36].
Antioxidants, isolated from natural resources or synthetically produced, are commonly used as bioactive compounds in modern medicine due to their ability to decrease the number of free radicals in cells and tissues [39]. MAA antioxidative properties are demonstrated via their ability to scavenge reactive oxygen species (ROS) and suppress singlet oxygen-induced DNA damage [21,31]. MAAs, such as porphyra-334 and shinorine, were found to increase the expression of NrF2 genes from the cytoprotective KeaP1/NrF2 pathway [21]. This pathway is important for synthesising proteins involved in maintaining cellular homeostasis and protection from oxidative stress [40]. In addition, MAAs accumulated in the cytoplasm of cyanobacteria play a role in osmotic regulation and drought stress protection [33]. MAAs also provide cellular protection from salt stress [41], while at the same time, they are used as a source of nitrogen, which is important for daily photosynthetic processes [42].
Beyond their UV-protective and antioxidative properties, MAAs demonstrate additional biotechnological potentials, including anti-inflammatory, anti-proliferative, and antiaging properties (Table 1) [14], that could be further explored for more generation of environmentally friendly sunscreen with advanced capacity for skin cancer prevention [10,43]. However, MAAs have been used in very limited ways in several cosmetic applications as natural agents for sun protection [44]. For example, porphyra-334 and shinorine isolated from red alga Porphyra umbilicalis have been used in the commercial product Helioguard® 365, which was promoted for containing natural UV-screening compounds [45]. Helionori® is another commercial product that protects from UVR and contains MAAs palythine, porphyra-334, and shinorine [44].
MAA synthesis occurring via the shikimate pathway [12] and/or the pentose phosphate pathway [46] was also analysed using genome-mining approaches [47,48]. The genes from the MAA pathway (called the mys cluster) were first isolated from the cyanobacterium Anabaena variabilis and contained four genes: dehydroquinate synthase (DHQS), O-methyltransferase (O-MT), adenosine triphosphate (ATP) grasp, and nonribosomalpeptide synthetase (NRPS) [46]. The variability in genes encoding enzymes important for the MAA biosynthesis pathways was confirmed using genomics data encompassing over 300 cyanobacteria [48] and microalgae Symbiodiniaceae [49]. Consequently, there are still substantial knowledge gaps in understanding MAA biosynthesis, which needs to be resolved and is important for enabling the use of MAAs in biotechnology, especially when applying heterologous expression systems [50]. Therefore, further genomic analyses are needed to fully elucidate differences in MAA biosynthetic gene clusters (BGCs) in various MAA-synthesis-capable species. This review discusses ways to improve MAA biosynthetic knowledge, including the application of genome-mining screening approaches in MAA discoveries.

2. Traditional vs. New Ways of Natural Products Discovery

The emerging treatments of currently uncured diseases have mainly focused on exploring and utilising novel natural products (NPs) [51]. Therefore, the discovery of new NPs has been driven by the constant need for new compounds with supreme and improved bioactive properties, sustainable in vitro production, and a high yield of bioactive compounds [52,53,54]. These sustainability aspects of targeted NPs have been critical when assessing the biotechnological capacity of the bioactive compounds for potential pharmaceutical and industry applications [40,55]. Traditionally, the discovery pathways to the novel NPs were driven by the bioactivity-guided pathway that begins with the extraction step, followed by the isolation and purification of the molecule of interest (Figure 2). During these isolation and purification steps, the assessment of biological activities is happening in parallel, as the presence or absence of other molecules or external conditions may impact the activity of the targeted NP. These bioactivity screening options, which are often performed using different phenotypic assays, are implemented to confirm the presence of bioactivity in newly discovered NP [56,57,58]. However, the molecular mechanism of action for these NPs is often unclear; therefore, the application of specific phenotypic assays for revealing NPs actions (i.e., proteome profiling and other assays) is useful at this discovery stage [40,59]. Another obstacle in the traditional NP discovery pathway includes the purification efficiency for a targeted molecule and the availability of appropriate tools for chemical characterisation [40]. Furthermore, after successful NP isolation, characterisation, and confirmation of the beneficial bioactivity, additional limitations preventing its use in biotechnological applications may include the limited availability of natural resources [60] or culturing or mass production capacities [61,62]. Sustainable use of natural resources is currently on the list of priorities for policymakers worldwide to preserve ecosystems and protect biodiversity [63]. Anthropogenic pollution from chemicals such as agricultural pesticides, industrial chemicals, and air and water pollutants has been recognised as a global problem and an increased risk to humanity and environmental health [64,65]. Current issues important for environmental protection and preserving natural biodiversity could be partially resolved by the use of environmentally friendly, natural compounds in biotechnological applications [38,66].
During the last decade, there has been a shift towards new genomics-based NP discoveries that include applying next-generation sequencing data and new bioinformatics tools [55,67]. With a significant increase in the number of sequenced organisms provided within publicly available databases, genome-mining approaches can be used for unravelling the cryptic biosynthetic potential in species previously unidentified for the synthesis of targeted molecules and, consequently, for novel drug discoveries [68].
Recently, applying the traditional isolation and screening approach to red algae resulted in the identification of 23 potential MAAs in 40 different seaweed species from Brittany [69]. However, for an efficient application of MAAs in biotechnology, the key obstacles to be overcome include: (i) limited availability of MAAs from natural resources; (ii) the lack of clear understanding of the MAA biosynthesis; and (iii) inadequate success in MAA production in heterologous expression systems [43,44]. The widespread industry use of MAA was in some ways prevented due to limited success in heterologous MAA expression and also the low extraction yields from natural sources. The potential solution was investigated by the application of chemical synthesis, which resulted in the generation of synthetic MAA analogues [70]. However, these synthetic compounds only provide a very narrow range of pharmacological and UV-absorbing properties compared to the properties found in MAAs isolated from nature. Consequently, the existing limitations prompted new research directions via the application of big data and using novel omics approaches.

3. From Genes to Biotechnological Solutions

By applying genomic and phylogenetics methodologies, the variability in BGCs for the synthesis of various secondary metabolites has been revealed in different species [71]. Novel genes such as Cytochrome P450 and haemogobin were unexpectedly recovered in transcriptomics data [72,73], with heat shock proteins demonstrating differences in genetic polymorphism and differential gene expression profiles (Figure 3) [74,75]. Structural analyses and protein modelling from discovered BGC allowed better anticipation of protein function and gene origin, like in the case of MAA biosynthetic clusters [49]. The MAA BGC were identified in the cyanobacterium, including MAA gene analogous to some other analysed species in their omics data [46]. Similarly, the genes from the mys cluster encoding enzymes for the MAA biosynthetic pathway [46] were identified in the transcriptomics data of symbiotic dinoflagellates, the Symbiodiniaceae, including genes and their homologs of dehydroquinate synthase, adenosine triphosphate (ATP) grasp, O-methyltransferase; nonribosomal peptide synthetase (NRPS)-like genes [49]. The microbial origin of these mys genes in symbiotic dinoflagellates was confirmed via phylogenetic analyses (Figure 3) [49]. These types of analyses applied the forward genetic approach, starting from the phenotype (i.e., established presence of targeted NPs) towards sequence analyses. In contrast, the novel reverse genetic approaches go the opposite way by analysing sequence data and predicting the synthesis capacity [51]. These new approaches for assessing genomics sequence data allowed further discoveries [76] and improved further by applying different algorithms and machine learning tools for BCG discovery [77].
One of the suggested ways to overcome the issue of limited bioavailability of natural resources includes culturing of certain organisms such as microalgae [53]. These photosynthetic microorganisms are easy to be cultivated in vitro, and using cyanobacterial farming can be applied for massive and sustainable production of desired compounds and energy while decreasing CO2 emission [80]. Biotechnological processes using in vitro cultures of Escherichia coli are still applied as heterologous expression systems for the manufacturing of numerous biopharmaceuticals [81] and recombinant proteins [82]. The application of heterologous expression systems in recombinant biotechnology is performed via the introduction of foreign biosynthetic gene clusters into the host genome of E. coli, yeast, or other suitable organisms [82,83]. Recombinant proteins are synthesised in different organisms, including bacteria, yeast, and other heterologous expression systems [50,82,83,84]. In E. coli, human genes were successfully expressed and even manipulated via DNA family shuffling methodology (i.e., molecular evolution in vitro) and analysed for the presence of improved catalytic properties in new mutant genes created [58,78,79], and evaluated in the screening process (Figure 3d).
The biosynthetic gene cluster encoding the synthesis of UV-absorbing MAAs was also successfully isolated from the cyanobacterium Anabaena variabilis ATCC 29413 and expressed in bacteria [46]. Specifically, in vitro, the synthesis of primary MAA, shinorine, was successfully implemented in E. coli after introducing the cyanobacterial 4-gene mys cluster [46]. MAAs were also heterologously expressed in E. coli using a 5-gene cluster (mylA–E) isolated from cyanobacterium Cylindrospermum stagnale and resulted in the synthesis of mycosporine-lysine and the new MAA mycosporine-ornithine [50]. An MAA gene cluster isolated from Nostoc linckia resulted in the production of MAA precursor 4-deoxygadusol, then four MAAs (i.e., mycosporine-glycine, porphyra-334, shinorine, mycosporine-glycine-alanine, and palythine) [84].
MAAs molecules as secondary metabolites that are well recognised for their role in molecular interactions, ecological function, and other protective roles are characterised by the low yield in natural resources [14]. The application of omics technologies opens the window for the simultaneous discovery of multiple compounds by unrevealing the genetic signature and hidden biosynthetic potential of certain organisms [10]. Application of various approaches, including metabolomic, transcriptomic, proteomic, even metatranscriptomics, metagenomics, and other multi-omics, speed up the NP discovery rate [55,67,85]. Metagenomics or meta-transcriptomics data, containing multiple species sequences, often host and related-symbionts, such as reef-building corals and their microbial symbionts, provide insight into the symbiosis functionality and response to changing environmental conditions [86,87,88,89].
Various omics approaches were recently reviewed for MAAs and other NPs regarding the use of next-generation of sequencing (NGS) data, assessment of biosynthetic pathways, phylogenetic analyses, and liquid chromatography-tandem mass spectrometry data [10,76,90]. Therefore, heterologous expression systems were employed in an attempt to produce certain types of MAAs [46]. Although the expression of specific MAAs was successful in E. coli and other systems, there was still limited success as only certain types of MAAs were heterologous expressions produced in vitro [46,47]. Cyanobacterial MAAs were heterologously expressed as a result of five gene cluster (mylA to mylE) in E. coli, resulting in the production of mycosporine-lysine and mycosporine-ornithine [50]. Using a bioinformatic approach, the MAA BCG was identified in Nostoc and expressed in E. coli, producing direct MAA precursor 4-deoxygadusol and five MAAs (mycosporine-glycine, porphyra-334, shinorine, mycosporine-glycine-alanine and palythine [84]). The MAAs BGC from two Actinomycetales species were found to be homologous to cyanobacterial BGC and, when heterologously expressed in the host Streptomyces, new mycosporine-glycine-alanine was produced [47]. Interestingly, when Actinosynnema mirum was maintained in culture, lack of MAA production indicated the cryptic state of BGC, which sometimes may prevent detection of MAAs. The cryptic BGCs present the clear challenge to discover the full potential of analysed organisms as activation clues may be missing, where the advantage of genome mining may solve and provide insights for new secondary metabolite discoveries [76].

4. Genome-Mining Tools

Analyses of genomic sequences provided a hidden glance into organisms’ potential for producing bioactive NPs. Modern strategies in genome mining include in silico methods for BGC identification to facilitate the discovery of novel NPs [91]. For genome mining, different types of sequence data (e.g., genomics, transcriptomics, metabolomics, proteomics, epigenomics, and multi-omics) [10,40,54] and numerous bioinformatics tools are applied [67,92]. During this process, the starting point is the accessibility of omics data from publicly available databases such as GenBank at the National Center for Biotechnology Information (NCBI) [93] and other integrated databases resources such as the International Nucleotide Sequence Database Collaboration (INSDC; http://www.insdc.org/, accessed on 12 June 2022) [94]. The majority of genome-mining strategies targeted the specific BGC encoding enzyme important for NP synthesis [76] using various computational tools. The overview of commonly used bioinformatics tools, especially applicable for microbial BGCs, is provided in Table 2.
The original targeted-based genome-mining approaches applied reference alignment using a basic local alignment search tool known as BLAST [116]. Later, rule-based algorithms with the comprehensive pipeline improved the identification of BGCs for secondary metabolites from bacterial and fungal sequence data using AntiSMASH (http://antismash.secondarymetabolites.org; accessed on 12 June 2022) [95] with additional updated versions for improved detection of related BGCs [96,97,98,99,100]. Another useful computational tool is PRISM, which uses microbial genomic data and identifies BGCs for nonribosomal peptides and some polyketides [105,106,107]. This genome-guided prediction tool includes the complete chemical structure for all currently used bacteria-driven antibiotics and also forecasts possible NPs based on cryptic BGCs [107]. Both AntiSMAH and PRISM apply multiple sequence alignment-based profile Hidden Markov Model (HMM) [77]. ClusterFinder uses the machine learning approach and HMM for the systematic identification of BGCs [117]. However, this computational tool has limited detection of the higher-order properties of BGCs (e.g., position dependency effects) [77,118]. The RiPPER genome-mining tool is used for the discovery of BGCs encoding specialised microbial metabolites, ribosomally synthesised, and post-translationally modified peptides (RiPPs) [110]. Due to a lack of shared pattern, in silico prediction of RiPP BGCs was performed by identification of the co-existence of a specific precursor peptide and RiPP tailoring enzymes, but this is possible only for the identification of already discovered RiPP families.
Further development of genome mining applied the use of a deep-learning approach, including the use of recurrent neural networks (RNNs) and the protein family database (http://pfam.xfam.org; accessed on 12 June 2022)) [119]. Another deep-learning strategy (DeepBGC) applied the BGC prediction algorithm to further improve the identification of de novo bacterial gene clusters (https://github.com/Merck/deepbgc; accessed on 12 June 2022) [77]. Open sources for the prediction of BGC, such as Prodigal (http://compbio.ornl.gov/prodigal/; accessed on 12 June 2022) and automated annotation, helped in reducing false-positive BGCs identification [120]. Genome mining search for novel antibiotics, Antibiotic-Resistant Target Seeker (ARTS), is available at https://arts.ziemertlab.com (accessed on 12 June 2022) [121]. Recently, a new machine learning approach, GECCO (GEne Cluster prediction with COnditional random fields; https://gecco.embl.de; accessed on 12 June 2022), allowed much higher identification of de novo BGCs from metagenomics data, confirming the important link for prediction between protein domain and secondary metabolites [91]. An overview of different bioinformatics tools applied for genome mining and useful in the discovery of BCGs and related secondary metabolites are presented in Table 2. All these different genome-mining approaches allowed the unexpected discovery of important genes or gene clusters within unforeseen organisms.
The discovery of MAA biosynthetic capacities has been traditionally guided by phenotype or conformation of MAA profiles after chemistry-guided extraction (Figure 2). Follow-up genomic tools have been used via the application of traditional BLAST genome-mining approaches [48,49] for a better understanding of MAA biosynthetic pathways [46,47]. However, genotype-driven MAA discoveries via the application of diverse genomics tools such as the one presented in Table 2 are yet to be efficiently utilised. Using various genome-mining tools incorporating machine learning [117] and a deep-learning approach [119] will allow genotype-driven MAA discoveries or the discovery of a cryptic potential for MAA biosynthesis from currently unidentified resources. Therefore, applying advanced genome-mining tools could help overcome current biotechnological obstacles limiting cost-effective MAA industry applications by extension of available natural resources and a better understanding of regulatory mechanisms involved in MAA synthesis.

5. Conclusions

MAAs are secondary metabolites with promising potential to be used as marine-derived sunscreens due to their supreme UV-absorbing capacity and additional pharmacologically beneficial properties. The application of modern bioinformatics tools in genome mining provides exciting opportunities for the characterisation of the MAA biosynthetic pathways, its expression patterns and regulatory mechanisms, which could overcome current existing obstacles in scaling up MAA production in vivo and in vitro. Furthermore, new/old MAA compounds can be rediscovered in completely new resources, providing ecologically friendly alternatives for preserving natural biodiversity. Applying genome mining will provide an alternative way of screening marine and other organisms for their potential to produce MAA. Consequently, this will unlock biotechnological potential for the use of MAAs in cosmetic, pharmaceutical, and other industries, including producing environmentally friendly organic sunscreens with therapeutic properties.

Funding

This research received no external funding.

Acknowledgments

The author would like to thank the three anonymous reviewers, as well as Isidora Skrlin, for their critical reviews of this paper.

Conflicts of Interest

The author declare no conflict of interest.

References

  1. Garcia-Pichel, F.; Wingard, C.E.; Castenholz, R.W. Evidence Regarding the UV Sunscreen Role of a Mycosporine-Like Compound in the Cyanobacterium Gloeocapsa sp. Appl. Environ. Microbiol. 1993, 59, 170–176. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Rosic, N.N.; Braun, C.; Kvaskoff, D. Extraction and Analysis of Mycosporine-Like Amino Acids in Marine Algae. In Natural Products from Marine Algae; Methods in Molecular Biology; Stengel, D., Connan, S., Eds.; Humana Press: New York, NY, USA, 2015; Volume 1308. [Google Scholar] [CrossRef]
  3. 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]
  4. Dunlap, W.C.; Malcolm, S.J. Ultraviolet radiation-absorbing mycosporine-like amino acids in coral reef organisms: A biological and environmental perspective. J. Phycol. 1998, 34, 418–430. [Google Scholar] [CrossRef]
  5. Rosic, N.N.; Dove, S. Mycosporine-Like Amino Acids from Coral Dinoflagellates. Appl. Environ. Microbiol. 2011, 77, 8478–8486. [Google Scholar] [CrossRef] [Green Version]
  6. Figueroa, F.L. Mycosporine-Like Amino Acids from Marine Resource. Mar. Drugs 2021, 19, 18. [Google Scholar] [CrossRef] [PubMed]
  7. 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] [Green Version]
  8. Rastogi, R.P.; Richa; 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]
  9. 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]
  10. Rosic, N.N. Recent advances in the discovery of novel marine natural products and mycosporine-like amino acid UV-absorbing compounds. Appl. Microbiol. Biotechnol. 2021, 105, 7053–7067. [Google Scholar] [CrossRef] [PubMed]
  11. 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. [Google Scholar]
  12. Shick, J.M.; Dunlap, W.C. Mycosporine-Like Amino Acids and Related Gadusols: Biosynthesis, Accumulation, and UV-Protective Functions in Aquatic Organisms. Annu. Rev. Physiol. 2002, 64, 223–262. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. 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]
  14. Geraldes, V.; Pinto, E. Mycosporine-Like Amino Acids (MAAs): Biology, Chemistry and Identification Features. Pharmaceuticals 2021, 14, 63. [Google Scholar] [CrossRef]
  15. Rastogi, R.P.; Incharoensakdi, A. Characterization of UV-screening compounds, mycosporine-like amino acids, and scytonemin in the cyanobacterium Lyngbya sp. CU2555. FEMS Microbiol. Ecol. 2014, 87, 244–256. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Geraldes, V.; de Medeiros, L.S.; Jacinavicius, F.R.; Long, P.F.; Pinto, E. Development and validation of a rapid LC-MS/MS method for the quantification of mycosporines and mycosporine-like amino acids (MAAs) from cyanobacteria. Algal Res. 2020, 46, 101796. [Google Scholar] [CrossRef]
  17. 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]
  18. 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]
  19. 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] [Green Version]
  20. 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. 2008, 21, 161–169. [Google Scholar] [CrossRef]
  21. 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]
  22. 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]
  23. Choi, S.-Y.; Lee, S.Y.; Kim, H.G.; Jeong, J.C.; Batara, D.C.; Kim, S.-H.; Cho, J.-Y. Shinorine and porphyra-334 isolated from laver (Porphyra dentata) inhibit adipogenesis in 3T3-L1 cells. Food Sci. Biotechnol. 2022, 31, 617–625. [Google Scholar] [CrossRef]
  24. 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] [Green Version]
  25. 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]
  26. 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. 2018, 662, 33–39. [Google Scholar] [CrossRef] [PubMed]
  27. Lawrence, K.P.; Gacesa, R.; Long, P.; Young, A. Molecular photoprotection of human keratinocytes in vitro by the naturally occurring mycosporine-like amino acid palythine. Br. J. Dermatol. 2017, 178, 1353–1363. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Athukorala, Y.; Trang, S.; Kwok, C.; Yuan, Y.V. Antiproliferative and Antioxidant Activities and Mycosporine-Like Amino Acid Profiles of Wild-Harvested and Cultivated Edible Canadian Marine Red Macroalgae. Molecules 2016, 21, 119. [Google Scholar] [CrossRef] [Green Version]
  29. 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 Medica 2015, 81, 813–820. [Google Scholar] [CrossRef] [Green Version]
  30. 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]
  31. 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]
  32. 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]
  33. 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]
  34. Rosic, N. Molecular Mechanisms of Stress Tolerance in Cyanobacteria. In Ecophysiology and Biochemistry of Cyanobacteria; Rastogi, R.P., Ed.; Springer Nature: Singapore, 2021; p. 131. [Google Scholar]
  35. 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]
  36. 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. In Ecological Studies in the Antarctic Sea Ice Zone; Springer: Berlin/Heidelberg, Germany, 2002; pp. 237–245. [Google Scholar] [CrossRef]
  37. Bertram, C.; Hass, R. Cellular responses to reactive oxygen species-induced DNA damage and aging. Biol. Chem. 2008, 389, 211–220. [Google Scholar] [CrossRef]
  38. Rosic, N.N. Mycosporine-Like Amino Acids: Making the Foundation for Organic Personalised Sunscreens. Mar. Drugs 2019, 17, 638. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Koltover, V.K. Antioxidant biomedicine: From free radical chemistry to systems biology mechanisms. Russ. Chem. Bull. 2010, 59, 37–42. [Google Scholar] [CrossRef]
  40. Atanasov, A.G.; Zotchev, S.B.; Dirsch, V.M.; Supuran, C.T. Natural products in drug discovery: Advances and opportunities. Nat. Rev. Drug Discov. 2021, 20, 200–216. [Google Scholar] [CrossRef]
  41. 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]
  42. 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]
  43. Singh, A.; Čížková, M.; Bišová, K.; Vítová, M. Exploring Mycosporine-Like Amino Acids (MAAs) as Safe and Natural Protective Agents against UV-Induced Skin Damage. Antioxidants 2021, 10, 683. [Google Scholar] [CrossRef] [PubMed]
  44. 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]
  45. Cardozo, K.H.M.; Guaratini, T.; Barros, M.P.; Falcão, V.R.; Tonon, A.P.; Lopes, N.P.; Campos, S.; Torres, M.A.; Souza, A.O.; Colepicolo, P.; et al. Metabolites from algae with economical impact. Comp. Biochem. Physiol. Part C Toxicol. Pharmacol. 2007, 146, 60–78. [Google Scholar] [CrossRef] [PubMed]
  46. 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]
  47. 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]
  48. 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. Env. Microbiol. 2019, 21, 702–715. [Google Scholar] [CrossRef] [PubMed]
  49. 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]
  50. Katoch, M.; Mazmouz, R.; Chau, R.; Pearson, L.A.; Pickford, R.; Neilan, B.A. Heterologous Production of Cyanobacterial Mycosporine-Like Amino Acids Mycosporine-Ornithine and Mycosporine-Lysine in Escherichia coli. Appl. Environ. Microbiol. 2016, 82, 6167–6173. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Milshteyn, A.; Schneider, J.S.; Brady, S.F. Mining the Metabiome: Identifying Novel Natural Products from Microbial Communities. Chem. Biol. 2014, 21, 1211–1223. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Ghareeb, M.A.; Tammam, M.A.; El-Demerdash, A.; Atanasov, A.G. Insights about clinically approved and Preclinically investigated marine natural products. Curr. Res. Biotechnol. 2020, 2, 88–102. [Google Scholar] [CrossRef]
  53. Sánchez-Bayo, A.; Morales, V.; Rodríguez, R.; Vicente, G.; Bautista, L.F. Cultivation of Microalgae and Cyanobacteria: Effect of Operating Conditions on Growth and Biomass Composition. Molecules 2020, 25, 2834. [Google Scholar] [CrossRef]
  54. Tewari, D.; Atanasov, A.G.; Semwal, P.; Wang, D. Natural products and their applications. Curr. Res. Biotechnol. 2021, 3, 82–83. [Google Scholar] [CrossRef]
  55. Carroll, A.R.; Copp, B.R.; Davis, R.A.; Keyzers, R.A.; Prinsep, M.R. Marine Natural Products. Nat. Prod. Rep. 2020, 37. [Google Scholar]
  56. Gallo, C.; Barra, G.; Saponaro, M.; Manzo, E.; Fioretto, L.; Ziaco, M.; Nuzzo, G.; D’Ippolito, G.; Palma, R.; Fontana, A. A New Bioassay Platform Design for the Discovery of Small Molecules with Anticancer Immunotherapeutic Activity. Mar. Drugs 2020, 18, 604. [Google Scholar] [CrossRef] [PubMed]
  57. Salvador-Reyes, L.A.; Luesch, H. Biological targets and mechanisms of action of natural products from marine cyanobacteria. Nat. Prod. Rep. 2015, 32, 478–503. [Google Scholar] [CrossRef] [Green Version]
  58. Rosic, N.N. DNA Shuffling of Cytochromes P450 for Indigoid Pigment Production. In Cytochrome P450 Protocols; Humana Press: Totowa, NJ, USA, 2013; Volume 987, pp. 205–224. [Google Scholar] [CrossRef]
  59. Gregori-Puigjané, E.; Setola, V.; Hert, J.; Crews, B.A.; Irwin, J.J.; Lounkine, E.; Marnett, L.; Roth, B.L.; Shoichet, B.K. Identifying mechanism-of-action targets for drugs and probes. Proc. Natl. Acad. Sci. USA 2012, 109, 11178–11183. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Neumayer, E. Scarce or Abundant? The Economics of Natural Resource Availability. J. Econ. Surv. 2002, 14, 307–335. [Google Scholar] [CrossRef] [Green Version]
  61. Fabris, M.; Abbriano, R.; Pernice, M.; Sutherland, D.L.; Commault, A.S.; Hall, C.C.; Labeeuw, L.; McCauley, J.I.; Kuzhiumparambil, U.; Ray, P.; et al. Emerging Technologies in Algal Biotechnology: Toward the Establishment of a Sustainable, Algae-Based Bioeconomy. Front. Plant Sci. 2020, 11, 279. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Stewart, E.J. Growing Unculturable Bacteria. J. Bacteriol. 2012, 194, 4151–4160. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Lampert, A. Over-exploitation of natural resources is followed by inevitable declines in economic growth and discount rate. Nat. Commun. 2019, 10, 1419. [Google Scholar] [CrossRef] [PubMed]
  64. Naidu, R.; Biswas, B.; Willett, I.R.; Cribb, J.; Singh, B.K.; Nathanail, C.P.; Coulon, F.; Semple, K.T.; Jones, K.C.; Barclay, A.; et al. Chemical pollution: A growing peril and potential catastrophic risk to humanity. Environ. Int. 2021, 156, 106616. [Google Scholar] [CrossRef]
  65. Rosic, N.; Bradbury, J.; Lee, M.; Baltrotsky, K.; Grace, S. The impact of pesticides on local waterways: A scoping review and method for identifying pesticides in local usage. Environ. Sci. Policy 2020, 106, 12–21. [Google Scholar] [CrossRef]
  66. Kulak, V.; Longboat, S.; Brunet, N.D.; Shukla, M.; Saxena, P. In Vitro Technology in Plant Conservation: Relevance to Biocultural Diversity. Plants 2022, 11, 503. [Google Scholar] [CrossRef] [PubMed]
  67. Kenshole, E.; Herisse, M.; Michael, M.; Pidot, S.J. Natural product discovery through microbial genome mining. Curr. Opin. Chem. Biol. 2020, 60, 47–54. [Google Scholar] [CrossRef]
  68. Bachmann, B.O.; Van Lanen, S.G.; Baltz, R.H. Microbial genome mining for accelerated natural products discovery: Is a renaissance in the making? J. Ind. Microbiol. Biotechnol. 2014, 41, 175–184. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Lalegerie, F.; Lajili, S.; Bedoux, G.; Taupin, L.; Stiger-Pouvreau, V.; Connan, S. Photo-protective 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] [Green Version]
  70. Losantos, R.; Funes-Ardoiz, I.; Aguilera, J.; Herrera-Ceballos, E.; García-Iriepa, C.; Campos, P.J.; Sampedro, D. Rational Design and Synthesis of Efficient Sunscreens to Boost the Solar Protection Factor. Angew. Chem. Int. Ed. 2017, 56, 2632–2635. [Google Scholar] [CrossRef]
  71. Belknap, K.C.; Park, C.J.; Barth, B.M.; Andam, C.P. Genome mining of biosynthetic and chemotherapeutic gene clusters in Streptomyces bacteria. Sci. Rep. 2020, 10, 2003. [Google Scholar] [CrossRef] [PubMed]
  72. Rosic, N.N.; Pernice, M.; Dunn, S.; Dove, S.; Hoegh-Guldberg, O. Differential Regulation by Heat Stress of Novel Cytochrome P450 Genes from the Dinoflagellate Symbionts of Reef-Building Corals. Appl. Environ. Microbiol. 2010, 76, 2823–2829. [Google Scholar] [CrossRef] [Green Version]
  73. Rosic, N.N.; Leggat, W.; Kaniewska, P.; Dove, S.; Hoegh-Guldberg, O. New-old hemoglobin-like proteins of symbiotic dinoflagellates. Ecol. Evol. 2013, 3, 822–834. [Google Scholar] [CrossRef] [PubMed]
  74. Rosic, N.N.; Pernice, M.; Dove, S.; Dunn, S.; Hoegh-Guldberg, O. Gene expression profiles of cytosolic heat shock proteins Hsp70 and Hsp90 from symbiotic dinoflagellates in response to thermal stress: Possible implications for coral bleaching. Cell Stress Chaperon 2010, 16, 69–80. [Google Scholar] [CrossRef] [Green Version]
  75. Rosic, N.N.; Pernice, M.; Rodriguez-Lanetty, M.; Hoegh-Guldberg, O. Validation of Housekeeping Genes for Gene Expression Studies in Symbiodinium Exposed to Thermal and Light Stress. Mar. Biotechnol. 2010, 13, 355–365. [Google Scholar] [CrossRef] [PubMed]
  76. Scherlach, K.; Hertweck, C. Mining and unearthing hidden biosynthetic potential. Nat. Commun. 2021, 12, 3864. [Google Scholar] [CrossRef] [PubMed]
  77. Hannigan, G.D.; Příhoda, D.; Palicka, A.; Soukup, J.; Klempir, O.; Rampula, L.; Durcak, J.; Wurst, M.; Kotowski, J.; Chang, D.; et al. A deep learning genome-mining strategy for biosynthetic gene cluster prediction. Nucleic Acids Res. 2019, 47, e110. [Google Scholar] [CrossRef] [PubMed]
  78. Rosic, N.N. Versatile capacity of shuffled cytochrome P450s for dye production. Appl. Microbiol. Biotechnol. 2009, 82, 203–210. [Google Scholar] [CrossRef]
  79. Rosic, N.N.; Huang, W.; Johnston, W.A.; DeVoss, J.J.; Gillam, E.M. Extending the diversity of cytochrome P450 enzymes by DNA family shuffling. Gene 2007, 395, 40–48. [Google Scholar] [CrossRef] [PubMed]
  80. Pathak, J.; Rajneesh; Maurya, P.K.; Singh, S.P.; Häder, D.-P.; Sinha, R.P. Cyanobacterial Farming for Environment Friendly Sustainable Agriculture Practices: Innovations and Perspectives. Front. Environ. Sci. 2018, 6, 7640128. [Google Scholar] [CrossRef]
  81. Kopp, J.; Slouka, C.; Spadiut, O.; Herwig, C. The Rocky Road from Fed-Batch to Continuous Processing With E. coli. Front. Bioeng. Biotechnol. 2019, 7, 328. [Google Scholar] [CrossRef] [PubMed]
  82. Rosano, G.L.; Ceccarelli, E.A. Recombinant protein expression in Escherichia coli: Advances and challenges. Front. Microbiol. 2014, 5, 172. [Google Scholar] [CrossRef] [Green Version]
  83. Porro, D.; Sauer, M.; Branduardi, P.; Mattanovich, D. Recombinant Protein Production in Yeasts. Mol. Biotechnol. 2005, 31, 245–260. [Google Scholar] [CrossRef]
  84. Chen, M.; Rubin, G.M.; Jiang, G.; Raad, Z.; Ding, Y. Biosynthesis and Heterologous Production of Mycosporine-Like Amino Acid Palythines. J. Org. Chem. 2021, 86, 11160–11168. [Google Scholar] [CrossRef]
  85. Ziemert, N.; Alanjary, M.; Weber, T. The evolution of genome mining in microbes—A review. Nat. Prod. Rep. 2016, 33, 988–1005. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Kaniewska, P.; Chan, C.-K.K.; Kline, D.; Ling, E.Y.S.; Rosic, N.; Edwards, D.; Hoegh-Guldberg, O.; Dove, S. Transcriptomic Changes in Coral Holobionts Provide Insights into Physiological Challenges of Future Climate and Ocean Change. PLoS ONE 2015, 10, e0139223. [Google Scholar] [CrossRef] [PubMed]
  87. Rosic, N.; Kaniewska, P.; Chan, C.-K.K.; Ling, E.Y.S.; Edwards, D.; Dove, S.; Hoegh-Guldberg, O. Early transcriptional changes in the reef-building coral Acropora aspera in response to thermal and nutrient stress. BMC Genom. 2014, 15, 1052. [Google Scholar] [CrossRef] [Green Version]
  88. Sharpton, T.J. An introduction to the analysis of shotgun metagenomic data. Front. Plant Sci. 2014, 5, 209. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Voolstra, C.R.; Miller, D.J.; A Ragan, M.; Hoffmann, A.; Hoegh-Guldberg, O.; Bourne, D.; Ball, E.; Ying, H.; Foret, S.; Takahashi, S.; et al. The ReFuGe 2020 Consortium—using “omics” approaches to explore the adaptability and resilience of coral holobionts to environmental change. Front. Mar. Sci. 2015, 2, 1. [Google Scholar] [CrossRef] [Green Version]
  90. Williams, A.N.; Sorout, N.; Cameron, A.J.; Stavrinides, J. The Integration of Genome Mining, Comparative Genomics, and Functional Genetics for Biosynthetic Gene Cluster Identification. Front. Genet. 2020, 11, 600116. [Google Scholar] [CrossRef] [PubMed]
  91. Carroll, L.M.; Larralde, M.; Fleck, J.S.; Ponnudurai, R.; Milanese, A.; Cappio, E.; Zeller, G. Accurate de novo identification of biosynthetic gene clusters with GECCO. bioRxiv 2021. [Google Scholar] [CrossRef]
  92. Medema, M.H.; Fischbach, M.A. Computational approaches to natural product discovery. Nat. Chem. Biol. 2015, 11, 639–648. [Google Scholar] [CrossRef] [PubMed]
  93. Sayers, E.W.; Agarwala, R.; Bolton, E.E.; Brister, J.R.; Canese, K.; Clark, K.; Connor, R.; Fiorini, N.; Funk, K.; Hefferon, T.; et al. Database resources of the National Center for Biotechnology Information. Nucleic Acids Res. 2010, 39, D23–D28. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Cochrane, G.; Karsch-Mizrachi, I.; Takagi, T.; Sequence Database Collaboration, I.N. The international nucleotide sequence database collaboration. Nucleic Acids Res. 2018, 46, D48–D50. [Google Scholar]
  95. Medema, M.H.; Blin, K.; Cimermancic, P.; De Jager, V.; Zakrzewski, P.; Fischbach, M.A.; Weber, T.; Takano, E.; Breitling, R. antiSMASH: Rapid identification, annotation and analysis of secondary metabolite biosynthesis gene clusters in bacterial and fungal genome sequences. Nucleic Acids Res. 2011, 39, W339–W346. [Google Scholar] [CrossRef] [PubMed]
  96. Blin, K.; Medema, M.H.; Kazempour, D.; Fischbach, M.A.; Breitling, R.; Takano, E.; Weber, T. antiSMASH 2.0—A versatile platform for genome mining of secondary metabolite producers. Nucleic Acids Res. 2013, 41, W204–W212. [Google Scholar] [CrossRef] [Green Version]
  97. Blin, K.; Shaw, S.; Steinke, K.; Villebro, R.; Ziemert, N.; Lee, S.Y.; Medema, M.H.; Weber, T. antiSMASH 5.0: Updates to the secondary metabolite genome mining pipeline. Nucleic Acids Res. 2019, 47, W81–W87. [Google Scholar] [CrossRef] [Green Version]
  98. Blin, K.; Wolf, T.; Chevrette, M.G.; Lu, X.; Schwalen, C.J.; Kautsar, S.A.; Suarez Duran, H.G.; de los Santos, E.L.C.; Kim, H.U.; Nave, M.; et al. antiSMASH 4.0—improvements in chemistry prediction and gene cluster boundary identification. Nucleic Acids Res. 2017, 45, W36–W41. [Google Scholar] [CrossRef]
  99. Weber, T.; Blin, K.; Duddela, S.; Krug, D.; Kim, H.U.; Bruccoleri, R.; Lee, S.Y.; Fischbach, M.A.; Müller, R.; Wohlleben, W.; et al. AntiSMASH 3.0-A comprehensive resource for the genome mining of biosynthetic gene clusters. Nucleic Acids Res. 2015, 43, W237–W243. [Google Scholar] [CrossRef] [Green Version]
  100. Blin, K.; Shaw, S.; Kloosterman, A.M.; Charlop-Powers, Z.; van Wezel, G.P.; Medema, M.H.; Weber, T. antiSMASH 6.0: Improving cluster detection and comparison capabilities. Nucleic Acids Res. 2021, 49, W29–W35. [Google Scholar] [CrossRef] [PubMed]
  101. De Jong, A.; van Hijum, S.A.; Bijlsma, J.J.; Kok, J.; Kuipers, O.P. BAGEL: A web-based bacteriocin genome mining tool. Nucleic Acids Res. 2006, 34, W273–W279. [Google Scholar] [CrossRef]
  102. De Jong, A.; van Heel, A.J.; Kok, J.; Kuipers, O.P. BAGEL2: Mining for bacteriocins in genomic data. Nucleic Acids Res. 2010, 38, W647–W651. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Van Heel, A.J.; de Jong, A.; Montalban-Lopez, M.; Kok, J.; Kuipers, O.P. BAGEL3: Automated identification of genes encoding bacteriocins and (non-)bactericidal posttranslationally modified peptides. Nucleic Acids Res. 2013, 41, W448–W453. [Google Scholar] [CrossRef]
  104. Van Heel, A.J.; de Jong, A.; Song, C.; Viel, J.H.; Kok, J.; Kuipers, O.P. BAGEL4: A user-friendly web server to thoroughly mine RiPPs and bacteriocins. Nucleic Acids Res. 2018, 46, W278–W281. [Google Scholar] [CrossRef] [PubMed]
  105. Skinnider, M.; DeJong, C.A.; Rees, P.N.; Johnston, C.W.; Li, H.; Webster, A.L.H.; Wyatt, M.A.; Magarvey, N.A. Genomes to natural products PRediction Informatics for Secondary Metabolomes (PRISM). Nucleic Acids Res. 2015, 43, 9645–9662. [Google Scholar] [CrossRef] [Green Version]
  106. Skinnider, M.A.; Merwin, N.J.; Johnston, C.W.; Magarvey, N.A. PRISM 3: Expanded prediction of natural product chemical structures from microbial genomes. Nucleic Acids Res. 2017, 45, W49–W54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Skinnider, M.A.; Johnston, C.W.; Gunabalasingam, M.; Merwin, N.J.; Kieliszek, A.M.; MacLellan, R.J.; Li, H.; Ranieri, M.R.M.; Webster, A.L.H.; Cao, M.P.T.; et al. Comprehensive prediction of secondary metabolite structure and biological activity from microbial genome sequences. Nat. Commun. 2020, 11, 6058. [Google Scholar] [CrossRef] [PubMed]
  108. Weber, T.; Rausch, C.; Lopez, P.; Hoof, I.; Gaykova, V.; Huson, D.; Wohlleben, W. CLUSEAN: A computer-based framework for the automated analysis of bacterial secondary metabolite biosynthetic gene clusters. J. Biotechnol. 2009, 140, 13–17. [Google Scholar] [CrossRef] [PubMed]
  109. Agrawal, P.; Khater, S.; Gupta, M.; Sain, N.; Mohanty, D. RiPPMiner: A bioinformatics resource for deciphering chemical structures of RiPPs based on prediction of cleavage and cross-links. Nucleic Acids Res. 2017, 45, W80–W88. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Santos-Aberturas, J.; Chandra, G.; Frattaruolo, L.; Lacret, R.; Pham, T.H.; Vior, N.M.; Eyles, T.H.; Truman, A.W. Uncovering the unexplored diversity of thioamidated ribosomal peptides in Actinobacteria using the RiPPER genome mining tool. Nucleic Acids Res. 2019, 47, 4624–4637. [Google Scholar] [CrossRef]
  111. Moffat, A.D.; Santos-Aberturas, J.; Chandra, G.; Truman, A.W. A User Guide for the Identification of New RiPP Biosynthetic Gene Clusters Using a RiPPER-Based Workflow. In Antimicrobial Therapies: Methods and Protocols; Barreiro, C., Barredo, J.-L., Eds.; Springer: New York, NY, USA, 2021; p. 227. [Google Scholar]
  112. Tietz, J.I.; Schwalen, C.J.; Patel, P.S.; Maxson, T.; Blair, P.M.; Tai, H.-C.; Zakai, U.I.; Mitchell, D.A. A new genome-mining tool redefines the lasso peptide biosynthetic landscape. Nat. Chem. Biol. 2017, 13, 470–478. [Google Scholar] [CrossRef] [PubMed]
  113. Navarro-Muñoz, J.C.; Selem-Mojica, N.; Mullowney, M.W.; Kautsar, S.A.; Tryon, J.H.; Parkinson, E.I.; De Los Santos, E.L.C.; Yeong, M.; Cruz-Morales, P.; Abubucker, S.; et al. A computational framework to explore large-scale biosynthetic diversity. Nat. Chem. Biol. 2020, 16, 60–68. [Google Scholar] [CrossRef]
  114. Kautsar, S.A.; Suarez Duran, H.G.; Blin, K.; Osbourn, A.; Medema, M.H. plantiSMASH: Automated identification, annotation and expression analysis of plant biosynthetic gene clusters. Nucleic Acids Res. 2017, 45, W55–W63. [Google Scholar] [CrossRef] [Green Version]
  115. Töpfer, N.; Fuchs, L.-M.; Aharoni, A. The PhytoClust tool for metabolic gene clusters discovery in plant genomes. Nucleic Acids Res. 2017, 45, 7049–7063. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Altschul, S.F.; Gish, W.; Miller, W.; Myers, E.W.; Lipman, D.J. Basic local alignment search tool. J. Mol. Biol. 1990, 215, 403–410. [Google Scholar] [CrossRef]
  117. Cimermancic, P.; Medema, M.H.; Claesen, J.; Kurita, K.; Brown, L.C.W.; Mavrommatis, K.; Pati, A.; Godfrey, P.A.; Koehrsen, M.; Clardy, J.; et al. Insights into Secondary Metabolism from a Global Analysis of Prokaryotic Biosynthetic Gene Clusters. Cell 2014, 158, 412–421. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Yoon, B.-J. Hidden Markov Models and their Applications in Biological Sequence Analysis. Curr. Genom. 2009, 10, 402–415. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Finn, R.D.; Coggill, P.; Eberhardt, R.Y.; Eddy, S.R.; Mistry, J.; Mitchell, A.L.; Potter, S.C.; Punta, M.; Qureshi, M.; Sangrador-Vegas, A.; et al. The Pfam protein families database: Towards a more sustainable future. Nucleic Acids Res. 2016, 44, D279–D285. [Google Scholar] [CrossRef]
  120. Hyatt, D.; Chen, G.-L.; Locascio, P.F.; Land, M.L.; Larimer, F.W.; Hauser, L.J. Prodigal: Prokaryotic gene recognition and translation initiation site identification. BMC Bioinform. 2010, 11, 119. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Alanjary, M.; Kronmiller, B.; Adamek, M.; Blin, K.; Weber, T.; Huson, D.H.; Philmus, B.; Ziemert, N. The Antibiotic Resistant Target Seeker (ARTS), an exploration engine for antibiotic cluster prioritization and novel drug target discovery. Nucleic Acids Res. 2017, 45, W42–W48. [Google Scholar] [CrossRef] [PubMed]
Figure 1. UV protection in aquatic species occurs at multiple levels, including UV-absorbing compounds (i.e., MAAs and other pigments) plus quenching mechanisms (i.e., enzymatic and non-enzymatic) and repair mechanisms (e.g., DNA repair).
Figure 1. UV protection in aquatic species occurs at multiple levels, including UV-absorbing compounds (i.e., MAAs and other pigments) plus quenching mechanisms (i.e., enzymatic and non-enzymatic) and repair mechanisms (e.g., DNA repair).
Marinedrugs 20 00478 g001
Figure 2. Traditional pathway in discovering natural products (NP) with bioactive properties, including challenges.
Figure 2. Traditional pathway in discovering natural products (NP) with bioactive properties, including challenges.
Marinedrugs 20 00478 g002
Figure 3. Genome mining and applications of omics data: (a) sequence analysis of haemoglobin-like genes obtained from microalgal transcriptomics data [73]; (b) phylogenetic analysis of MAA biosynthesis genes from the omics data [49]; (c) protein modelling of targeted haemoglobin protein from the omics data [73]; (d) functional analyses after gene sequence manipulation via molecular evolutionary in vitro and the creation of new cytochrome P450 enzymes [58,78,79].
Figure 3. Genome mining and applications of omics data: (a) sequence analysis of haemoglobin-like genes obtained from microalgal transcriptomics data [73]; (b) phylogenetic analysis of MAA biosynthesis genes from the omics data [49]; (c) protein modelling of targeted haemoglobin protein from the omics data [73]; (d) functional analyses after gene sequence manipulation via molecular evolutionary in vitro and the creation of new cytochrome P450 enzymes [58,78,79].
Marinedrugs 20 00478 g003
Table 1. MAA precursor and primary MAAs, including their chemical structure, bioactivities, and UV-absorbing maximum.
Table 1. MAA precursor and primary MAAs, including their chemical structure, bioactivities, and UV-absorbing maximum.
MAA Name (Molecular Formula; Mw)Chemical StructureKey Features (ʎmax)/Bioactive Properties [Reference]
4-deoxygadusol
(C8H12O5; 188 g/mol)
Marinedrugs 20 00478 i001Direct MAA precursor
UV-absorbing property (268 nm)
Mycosporine-glycine
(C10H15NO6; 245 g/mol)
Marinedrugs 20 00478 i002UV-absorbing property (310 nm)
Antioxidative [17,18,19,20]
Anti-inflammatory [18]
Antiaging [18]
Shinorine
(C13H20N2O8; 332 g/mol)
Marinedrugs 20 00478 i003UV-absorbing property (333 nm)
Antioxidative [17,19,21]
Anti-inflammatory [18,22]
Antiaging [18]
Anti-adipogenic [23]
Porphyra-334
(C14H22N2O8; 346 g/mol)
Marinedrugs 20 00478 i004UV-absorbing property (334 nm)
Antioxidative [17,19,21]
Anti-inflammatory [22]
Antiaging [24]
Anti-adipogenic [23]
Mycosporine-2-glycine
(C12H18N2O7; 302 g/mol)
Marinedrugs 20 00478 i005UV-absorbing property (332 nm)
Antioxidative [19,25,26]
Anti-inflammatory [26]
Antiaging [26]
Palythine
(C13H20N2O5; 284 g/mol)
Marinedrugs 20 00478 i006UV-absorbing property (320 nm)
Antioxidative [25,27]
Anti-proliferative [28]
Antiaging [29]
Table 2. Genome-mining tools for identification and analyses of biosynthetic gene clusters (BGCs) relevant to the synthesis of secondary metabolites and other natural products (NPs).
Table 2. Genome-mining tools for identification and analyses of biosynthetic gene clusters (BGCs) relevant to the synthesis of secondary metabolites and other natural products (NPs).
Software Name
(Key Features)
Website AvailabilityApplication and ImprovementsReference
AntiSMASH
(ANTIbiotics & Secondary Metabolite Analysis Shell; BCGs discovery in bacteria and fungi genome sequences)
Bacteria: antiSMASH bacterial version (secondarymetabolites.org/, accessed on 12 June 2022)
Fungi: antiSMASH fungal version
(secondarymetabolites.org, accessed on 12 June 2022)
Release of software[95]
Improved versions (2–5) [96,97,98,99]
The latest version (6) with improved BCGs detection [100]
BAGEL
(Automated identification of genes encoding ribosomally synthesised and post-translationally modified peptides -RiPPs)
Bacteria: http://bagel.molgenrug.nl
http://bagel4.molgenrug.nl/, accessed on 12 June 2022
Release of software[101]
BAGEL2[102]
BAGEL3[103]
BAGEL4[104]
PRISM
(PRediction Informatics for Secondary Metabolomes; prediction of chemical structures of NPs)
Microbe genomes: http://magarveylab.ca/prism/, accessed on 12 June 2022Release of software[105]
PRISM 3[106]
PRISM 4[107]
CLUSEAN
(CLUster SEquence Analyzer; bacterial secondary metabolite BCGs, automated analyses; Bioperl-based annotation pipeline)
Bacteria: https://bitbucket.org/tilmweber/clusean, accessed on 12 June 2022Release of software[108]
RiPPMiner
(Automated Prediction of BGCs and Crosslinked Chemical Structures of -RiPPs)
http://www.nii.ac.in/rippminer.html, accessed on 12 June 2022Release of software[109]
RiPPER
(for detection of BGCs of– RiPPs)
Actinobacteria: streptomyces/ripdock—Docker Image|Docker HubRelease of software
(Specific application for thioamidated ribosomal peptides)
Updated version
[110]
[111]
RODEO
(Rapid ORF Description and Evaluation Online; for detection of BGCs for RiPPs)
http://www.ripp.rodeo/, accessed on 12 June 2022Release of tool
(AntiSMASH combined with Pfam * domain prediction)
[112]
BiG-SCAPE
(The Biosynthetic Gene Similarity Clustering and Prospecting Engine; build sequence similarity report for new BGCs; using metabolomic data)
Multigenomes: BiG-SCAPE CORASON|July 2018 (secondarymetabolites.org, accessed on 12 June 2022)Release of software[113]
plantiSMASH
(antiSMASH for plant genomes)
Plant genomes: http://plantismash.secondarymetabolites.org, accessed on 12 June 2022Release of tool[114]
PhytoClust
(discovery of Metabolic Gene Clusters (MGCs) in plant genomes)
http://phytoclust.weizmann.ac.il/, accessed on 12 June 2022Release of tool[115]
* Pfam: Home page (xfam.org, accessed on 12 June 2022).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rosic, N. Genome Mining as an Alternative Way for Screening the Marine Organisms for Their Potential to Produce UV-Absorbing Mycosporine-like Amino Acid. Mar. Drugs 2022, 20, 478. https://doi.org/10.3390/md20080478

AMA Style

Rosic N. Genome Mining as an Alternative Way for Screening the Marine Organisms for Their Potential to Produce UV-Absorbing Mycosporine-like Amino Acid. Marine Drugs. 2022; 20(8):478. https://doi.org/10.3390/md20080478

Chicago/Turabian Style

Rosic, Nedeljka. 2022. "Genome Mining as an Alternative Way for Screening the Marine Organisms for Their Potential to Produce UV-Absorbing Mycosporine-like Amino Acid" Marine Drugs 20, no. 8: 478. https://doi.org/10.3390/md20080478

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