Next Article in Journal
Comparison of the Mineral and Nutraceutical Profiles of Elephant Garlic (Allium ampeloprasum L.) Grown in Organic and Conventional Fields of Valdichiana, a Traditional Cultivation Area of Tuscany, Italy
Previous Article in Journal
Influence of Light Conditions on Microalgae Growth and Content of Lipids, Carotenoids, and Fatty Acid Composition
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Bioactive Metabolites Produced by Cyanobacteria for Growth Adaptation and Their Pharmacological Properties

by
Pavitra Nandagopal
1,
Anthony Nyangson Steven
2,
Liong-Wai Chan
1,
Zaidah Rahmat
1,3,
Haryati Jamaluddin
1 and
Nur Izzati Mohd Noh
1,*
1
Department of Biosciences, Faculty of Science, Universiti Teknologi Malaysia, Skudai 81310, Malaysia
2
Department of Chemistry, Faculty of Science, Universiti Teknologi Malaysia, Skudai 81310, Malaysia
3
Institute of Bioproduct Development, Universiti Teknologi Malaysia, Skudai 81310, Malaysia
*
Author to whom correspondence should be addressed.
Biology 2021, 10(10), 1061; https://doi.org/10.3390/biology10101061
Submission received: 6 September 2021 / Revised: 10 October 2021 / Accepted: 14 October 2021 / Published: 18 October 2021
(This article belongs to the Section Microbiology)

Abstract

:

Simple Summary

Cyanobacteria are known as oxygenic microorganisms are able to release oxygen as a byproduct during photosynthesis. Rapidly changing environmental conditions require cyanobacteria to have dynamic adaptation strategies. They synthesize bioactive metabolites that are responsible for protection against harmful environmental conditions and to colonize in various habitats. This review focuses on the roles of bioactive metabolites for cyanobacterial survival and also discusses the bioactivities of these compounds for the treatment of numerous diseases.

Abstract

Cyanobacteria are the most abundant oxygenic photosynthetic organisms inhabiting various ecosystems on earth. As with all other photosynthetic organisms, cyanobacteria release oxygen as a byproduct during photosynthesis. In fact, some cyanobacterial species are involved in the global nitrogen cycles by fixing atmospheric nitrogen. Environmental factors influence the dynamic, physiological characteristics, and metabolic profiles of cyanobacteria, which results in their great adaptation ability to survive in diverse ecosystems. The evolution of these primitive bacteria resulted from the unique settings of photosynthetic machineries and the production of bioactive compounds. Specifically, bioactive compounds play roles as regulators to provide protection against extrinsic factors and act as intracellular signaling molecules to promote colonization. In addition to the roles of bioactive metabolites as indole alkaloids, terpenoids, mycosporine-like amino acids, non-ribosomal peptides, polyketides, ribosomal peptides, phenolic acid, flavonoids, vitamins, and antimetabolites for cyanobacterial survival in numerous habitats, which is the focus of this review, the bioactivities of these compounds for the treatment of various diseases are also discussed.

Graphical Abstract

1. Introduction

Cyanobacteria are photosynthetic microorganisms that possess various cellular strategies and physiological capacities to facilitate their adaptations for colonization in diverse environments on Earth. As a result, these photosynthetic microbes can exist in marine, terrestrial, and freshwater habitats. Furthermore, cyanobacteria are the most versatile ancient microorganisms that can thrive in extreme environments such as deserts, polar environments, geothermal springs, hypersaline lakes, and soils with high metal concentrations. They can be classified according to their ability to grow in high pH (alkaliphiles), beneath rock (endolithics), in high salinity (halophiles), under low nutrients (oligotrophics), in low (psychrophiles) or high (thermophiles) temperatures, and under high radiation levels (radiophiles) (Table 1).
Through centuries of evolution, cyanobacteria developed various sophisticated molecular, physiological, and metabolic characteristics for thriving in their habitats. The main aim of this paper is to review the roles of bioactive metabolites in addition to molecular machineries and physiological characteristics in ensuring the adaptation of cyanobacteria to environmental conditions. It is noteworthy that numerous studies have identified the potentials of cyanobacteria for modern drug discovery due to the bioactivities exhibited by cyanobacterial metabolites.

2. Adaptation Strategies of Cyanobacteria

2.1. Physiological Adaptation

Cyanobacteria possess the capacity to switch from one mode of metabolic approach to another. Most cyanobacteria conduct oxygenic photosynthetic mode. However, some can switch to anoxygenic photosynthetic mode [29]. For example, a filamentous mat of Leptolyngbya sp. and a cyanobacterial community of Planktothrix sp. and Annamia sp. dominating the sulfidic water column can conduct anoxygenic photosynthesis using sulfide as an electron donor [30,31]. Moreover, some cyanobacteria can carry out the fermentation processes under anoxic conditions and in the dark [32]. On the other hand, many cyanobacteria species form heterocysts, the cells that carry out atmospheric nitrogen fixation, especially during nitrogen deprivation [33]. This uniquely differentiated cell results in the dispersion of cyanobacterial genera in various ecosystems; for example, Anabaena and Trichodesmium inhabit open oceans, thermal springs and freshwaters [34,35], whereas Leptolynbya grows in geothermal springs, hot deserts, and surface crusts of semi-deserts [36,37].
The high adaptability of cyanobacteria to high temperature environments might be related to their photosynthetic machinery acclimation throughout many years of evolution. Previous studies have identified that light-harvesting phycobilisome (PBS) and photosystem II (PSII) are the main components that contribute to the survival of thermophilic cyanobacteria. Synechococcuss A/B clade, Mastigocladus laminosus, Synechococcus lividus and Synechococcus vulcanus have developed PBS with a greater thermostability during the evolutionary divergence [38]. The rigidity of the phycocyanin complex is important in achieving PBS thermostability [39]. On the other hand, D1 protein and PsbU, the key subunits of PSII, provide stability to PSII from denaturation at a high temperature [40,41]. Previous studies have also reported that filamentous cyanobacteria are thermostability related to metabolic mechanisms, which enables them to survive at high temperatures [42,43].
Additionally, some marine planktonic cyanobacteria, for example, Synechococcus sp. PCC 7942, exhibit DNA repair mechanisms, including detoxifying enzymes and pigments [44] and UV-absorbing sunscreen molecules [45] to release the damage caused by UV radiation and to protect them from harmful radiation pollutants [46]. Many planktonic cyanobacteria possess gas vesicles for position adjustment in the water column. Cyanobacteria use these gas-filled structures in connection to different environmental stimuli such as photic, gravitational, chemical and thermal to find a suitable niche [47].

2.2. Cellular Morphological Adaptation

Cyanobacteria exist in different morphological structures: unicellular in a single cell or colony with or without mucilaginous envelope, unbranched filamentous with single or multiple trichomes with or without sheath, and branched filamentous [48]. Furthermore, cyanobacteria have been subdivided into five subsections according to their morphological characteristics, subsection I (unicellular), subsection II (unicellular with baeocytes), subsection III (unbranched filamentous without heterocyst), subsection IV (false-branched or unbranched filamentous with heterocysts) and subsection V (branched filamentous with heterocysts) [49]. Although the forms are not habitat dependent, these physical characteristics might have contributed to cyanobacterial evolutionary adaptations. The earliest cyanobacterial genera are unicellular with sheath living in freshwater habitats [50]. The sheath enables benthic or sessile cyanobacteria such as Gloeocapsa, Synechococcus, Prochlorococcus and Aphanothece to form epilithic/endolithic biofilms in water bodies. This thick protective layer also provides high irradiance and UV light defense to the cells. Remarkably, some unicellular cyanobacteria possess thin firm colorless sheath for adaptations in extreme habitats, such as Chroococcidiopsis, which can be found in thermal and mineral springs, alkaline hypersaline swamps and hot or Antarctic deserts [51] as well as Chroococcus, which can be found in thermal springs and calcite speleotherms [52,53]. In contrast, the solitary cells or small groups of Halothece that inhabit coastal salty habitats lack mucilaginous envelope [54]. Noteworthy, most filamentous cyanobacteria produce extracellular sheath as a method of adaptation, especially to water level fluctuation and high solute concentration, by providing a microenvironment for trichomes. For example, Microcoleus, Trichocoleus, Oscillatoria and some Schizothrix covered by thick sheath grow in saline soil crusts, semi-desert regions, soil crusts of desert and polar environments [35,55,56]. However, some mat-forming Schizothrix and Oscillatoria enveloped by firm and thin sheaths inhabit diverse aquatic environments, freshwater, marine environments, thermal springs and polar water bodies [55]. Moreover, heterocystous cyanobacteria, such as Anabaena and Trichormus, have filaments without sheath or gelatinous envelope, which are necessary to allow more light penetration into the cells. In the case of cyanobacterial symbionts (cyanobionts), the absence of sheath or gelatinous envelope is important to enhance nitrogen and carbon transfer between the symbiotic partners [55].

2.3. Bioactive Metabolites for Cyanobacterial Adaptations and Their Pharmacological Properties

2.3.1. Indole Alkaloids

Alkaloids are ubiquitous in plants, bacteria, fungi and animals. In plants, alkaloids are produced as secondary metabolites in response to biotic or abiotic stresses [57]. Indole alkaloids are one of the alkaloid classes consisting of one indole structural moiety and are known for their bioactivities [58]. Many studies were conducted on the pharmacological properties of indole alkaloids from plants, fungi and animals, for example, antitumor activities showed by vinblastine and vincristine from Catharanthus roseus [59] and UV protective function by pityriacitrin from yeast Malassezia furfur [60], as well as the anti-inflammatory effect produced by conicamin from tunicate [61], lepadiformines A and B from ascidian [62], manzamine and carteramine A from sponges [63,64], and ascidiathiazones A and B from ascidan [65,66].
A diverse class of indole alkaloids synthesized by cyanobacteria have been reported as the bioactive secondary metabolites that possess pharmacological and biological properties (Table 2).
Thus far, 80 variants of indole alkaloids have been identified exclusively produced by the genera Westiella, Westiellopsis, Fischerella and Hapalosiphon (belonging to subsection V formerly order Stigonematales) [79,80,81]. The variants belong to nine different groups based on their carbon skeletons (Figure 1). Hapalindoles are the largest group of alkaloid indoles produced by cyanobacteria, which make up Group 1 (tetracyclic hapalindoles) and Group 2 (tricyclic hapalindoles). Furthermore, the hapalindoles are the precursors for the other groups, so-called hapalindole-type alkaloids: the hapalindolinones (Group 3), the ambiguines (Group 4 and Group 5), fischambiguines (Group 6), fischerindoles (Group 7) and welwitindolinones (Group 8 and Group 9). Such an extensive list of indole alkaloids suggests the important roles of these secondary metabolites in ensuring the survival of the cyanobacterial genera. Although most of the indole alkaloids are identified from the terrestrial and freshwater cyanobacteria in the genera Westiella, Westiellopsis, Fischerella and Hapalosiphon, it is tempting to speculate that these secondary metabolites are also produced by these genera inhabiting the other ecosystems, especially by those thriving in the extreme environments. Remarkably, the hapalindole family appears to be inherited vertically and thus, suggests the inheritance of hapalindole biosynthetic genes within the Subsection V [73].
Interestingly, scytonemin, an indole alkaloid UV-filtering pigment, is predominantly produced by cyanobacteria [82,83]. A recent study reported that an unculturable Halothece produced scytonemin in response to UV-A radiation at the driest Salar Grande, Atacama Desert [84]. In addition, scytonemin found in the terrestrial Lyngbya sp. CU2555, showed high resistance toward UV-B and heat, thusprotecting the cells against harsh environmental conditions. This strong oxidizing agent not only functions as a photoprotective compound against harmful UV radiation but also provides protection against deleterious short-wavelength radiation [85]. Notably, the accumulation of scytonemin in the unicellular Chroococcidiopsis-like cyanobacterial isolate from an epilithic desert crust occurred due to the increase in both temperature and photooxidative conditions together with UV-A exposure. However, the increased salt concentration under UV-A radiance blocked the production of scytonemin [86].
Furthermore, β–carboline, another indole alkaloid compound that is widely distributed in plants, animals and human tissues [87], has also been detected in cyanobacteria. It is noteworthythat norharmane (9H-pyrido(3,4-b) indole) excreted by Nodularia harveyana exhibited high algicidal activity [88]. In addition, this indole alkaloid can highly inhibit Gram-positive bacteria and moderately control Gram-negative bacteria and yeast [89]. Moreover, nostocarboline from Nostoc 78–12A could be used as an acetylcholinesterase inhibitor for the treatment of neuronal diseases [90].

2.3.2. Terpenoids

Terpenoids (or isoprenoids) are the largest group of bioactive compounds with more than 55,000 compounds discovered to date [91]. They can be classified as hemiterpenoid, monoterpenoids, sesquiterpenoids, diterpenoids, sesterterpenoids, triterpenoids (steroids) and tetraterpenoids (carotenoids) based on the number of isoprene units (Figure 2) [92].
In plants, terpenoids are involved in primary growth and development, defense against predators, and endophytic fungi or bacteria, as well as the attraction of pollinators [94]. These odorous metabolites are also synthesized by many bacteria including cyanobacteria that produce earthy odors in soil and water resources [95,96]. A range of terpenoids havebeen found in cyanobacteria and are known for their essential roles in ensuring cyanobacterial survival in a vast environment, as well as their importance as medicines, pigments and flavors.
Terpenoids identified from the halophilic Cylindrospermum muscicola HPUSD12 and Phormidium sp. HPUSD13, are suggested to play roles in providing protection against free radical oxidative damage to the cells that might be caused by the high-salinity condition of the Drang salt mine in India [97].
Sesquiterpenoid geosmin, a sesquiterpene without an isopropyl group, can be produced by several freshwater cyanobacteria, such as the filamentous Calothrix PCC 7507 [98] and the unicellular Synechocystis sp. PCC 6803 [99]. Furthermore, the heterologous expression of the sesquiterpene synthase gene from Nostoc punctiforme PCC 73102 and Nostoc sp. PCC 7120 in Escherichia coli suggests the production of sesquiterpenoids by this versatile species that inhabits various aquatic and terrestrial ecosystems [100]. This terpenoid group regulates the signaling defense activities of the cyanobacteria in response to the environmental stimuli. Interestingly, sesquiterpenoids produced by the marine Oscillatoria spongeliae are suggested to be responsible for the symbiotic interaction with the tropical marine sponge [101].
Triterpenoids, such as 2-methylhopanoids (2-MeBHPs), were discovered in a significant quantity in both laboratory cyanobacterial cultures and natural cyanobacterium-dominated microbial mats [102,103]. 2-MeBHPs is an example of pentacyclic triterpenoids, which play the role of biomarkers for modern cyanobacteria in some environmental settings [103]. Moreover, 2-MeBHP promotes osmotic, pH stress and freezing/thawing resistance in cyanobacteria to ensure their survival in dessert soil crusts, hot springs, hypersaline lake, Antarctic water, and Artic soil [100]. Apparently, the deletion of hpnP, the gene coding for the hopanoids protein responsible in C-2 methylation, caused a decrease in osmotic and pH stress tolerance by Nostoc punctiforme ATCC 29133S [104].
Tetraterpenoids, also known as carotenoids, are ubiquitousin most photosynthetic organisms and are essential for light-harvesting and energy dissipation during the photosynthesis process [105,106]. β-carotene, zeaxanthin, and echinenone are common carotenoids produced by cyanobacteria. The freshwater Aphanothece microscopica Nägeli (RSMan92) [107] and both the marine Cyanobium sp. LEGE 06113 [108] and Trichodesmium sp. [109] are also excellent sources of these terpenes. Carotenoids are lipophilic secondary metabolites from the isoprenoid pathway that are necessary to facilitate cyanobacteria against direct UV light exposure and photooxidative damage while conducting photosynthesis. In particular, it was suggested that echinenone and zeaxanthin protect PSII against singlet oxygen [110]. In addition, Pseudanabaena sp. CCNU1, inhabiting the arid and exposed region in Chetimari, Niger, produces carotenoids, specifically, β-carotene, echinenone, canthaxanthin, zeaxanthin, synechoxanthin, and three myxoxanthophyll derivatives, as a method of protection against UV-B radiation [111]. Moreover, carotenoids also play important roles in the survival of endolithic cyanobacteria, Chroococcidiopsis sp. at the Atacama Desert [112].
Recent studies have reported the biological activities of cyanobacterial carotenoids for the treatmentof various diseases. Several cyanobacteria strains, including the freshwater Alkalinema aff. pantanalense LEGE15481, Cyanobium gracile LEGE12431, Cuspidothrix issatschenkoi LEGE03282, the terrestrial Nodosilinea (Leptolyngbya) antarctica LEGE13457 and the marine Leptolyngbya-like sp. LEGE13412, have been characterized for their high content of carotenoids [113]. Remarkably, carotenoids and their derivatives extracted from terrestrial and marine cyanobacteria showed high superoxide anion radical (O2•−) scavenging and anti-inflammatory effects that enabled the treatment of psoriasis [113]. A high number of carotenoids was also detected in Cyanobium sp. LEGE 07175 and Tychonema sp. LEGE 07175 [114]. Both cyanobacterial extracts showed strong antiaging effects by inhibiting hyaluronidase, the enzyme that stimulates the depolymerization of hyaluronic acid under oxidative stress [114]. Although the information on the biochemical mechanisms of carotenoids in cell apoptosis and proliferation is still scarce, their antioxidantcapacity might be a contributing factor in anticancer and anti-aging effects.
Moreover, scytoscalarol (also known as an antimicrobial sesterterpene) from Scytonema sp. (UTEX 1163) culture showed growth inhibition against Bacillus anthracis, Staphylococcus aureus, Escherichia coli, Candida albicans, and Mycobacterium tuberculosis [115]. In addition, cybastacines A and B found in Nostoc sp. BEA-0956 also showed antibacterial activities against some clinical pathogenic bacteria [116]. Remarkably, scytonemides A and B extracted from the freshwater Scytonema hofmannii (UTEX 1834) havebeen identified and can function as an anticancer agent through the inhibition of 20S proteasome, the catalytic core of the proteasome complex that catalyzes the degradation of regulatory proteins [117].

2.3.3. Mycosporine-Like Amino Acids (MAAs)

Mycosporine-like amino acids (MAAs) are UV-absorbing compounds are involved in the evolution of organisms living in environments with high exposure to sunlight, such as cyanobacteria, microalgae, fungi, seaweeds, corals, and lichens [118,119,120,121]. These small water-soluble compounds (generally <400 Da) provide protection against ultraviolet radiation (UVR) exposure. In fact, the production of MAAs is induced by UV radiation and osmotic stresses; however, the mechanism remains poorly understood [122]. On the other hand, these photoprotective compounds also exhibit antioxidant activity [123,124,125].
Synechococcales, Chroococcales, Oscillatoriales, and Nostocales are efficient producers of MAAs for adaptations [118,126]. For example, a study found that MAAs produced by the benthic filamentous cyanobacteria in the Alpine Lake Gossenköllesee function as a protective shield against UVR. Despite a very low turnover of the synthesis of MAAs by the benthic filamentous cyanobacteria, especially during the ice-free season, these secondary metabolites reduce the transmission of UVR wavelengths received by the cyanobacteria in the clear Alpine lakes [127]. MAAs also provide protection against the damaging effects of solar UVR to the cyanobacterial mats in the Arctic [128]. Moreover, the MAA mys gene cluster in the filamentous Chlorogloeopsis fritschii PCC 6912 was up-regulated when simultaneously exposed to both UV and far-red lights. Thissuggests that MAAs may be involved in photon dissipation and thermodynamic optimization, which are important in regulating the heat from affecting the climate [129,130].
Recently, an usual mycosporine-glycine-alanine (MGA), an MAA derivative, was found in Sphaerospermopsis torques-reginae ITEP-024 [123]. Inhabiting freshwater with low salinity, this heterocystous filamentous cyanobacterium also produces the imino-mycosporines, shinorine and porphyra-334, as an acclamatory response to UV exposure [123]. Other than providing protection against UV light, rear mycosprorine-2-glycine (M2G), isolated from the halotolerant Aphanothece halophytica, possesses biological functions, such as free radical scavenging [131], oxidative stress protection [132], and osmoregulation [133], as well as inhibition of collagenase activity and protein glycation [134].
Interestingly, MAAs are attractive to cosmetic industries as the active ingredients for sunscreens and anti-aging products due to their characteristics [119,135]. For example, a recent in vitro study identified the protection activity of human keratinocytes against UV radiation by a novel MAA (13-O-β-galactosyl-porphyra-334) from Nostoc sphaericum [136]. This novel MAA also possesses radical scavenging activity that can reduce the damage caused by ROS in order to prevent photoaging. Nevertheless, MAAs have yet to be exploited for industrial production, with only a few edible products currently available. For example, both Helioguard 365® and Helionori® extracted from the red seaweed Porphyra umbilicalis [137,138] have been used as ingredients for the production of sunscreens. The in vivo study using Helioguard 365® from the red seaweed showed improvements in skin firmness and smoothness [138], whereas Helionori® offered protection against DNA damage due to UV radiation [139]. Nevertheless, these products are able to provide maximal protection in the UVA range but only allow minimum protection in the more damaging UVB range [125].

2.3.4. Non-Ribosomal Peptides and Polyketides

Biosynthesis of non-ribosomal peptides (NRPs) and polyketides (PKs) are catalyzed by non-ribosomal peptide synthases (NRPS) and polyketide synthases (PKS), respectively [140]. Notably, the gene clusters of NRPS and PKS were more frequently found in bacteria, including cyanobacteria, than archaea and eukarya [141]. The gene clusters have been found in the genera Lyngbya, Microcystis, Planktothrix, Nodularia, Nostoc, Pleurocapsa, and Anabaena [141,142]. Pleurocapsa and Nostoc species are the most common cyanobacteria producing NRP/PK [141]. However, cyanobacteria with genomes fewer than 3 Mbp, such as Prochlorococcus marinus SS120 and Synechococcus sp. WH8109, might not possess these clusters due to the extra metabolic burdens [141,143,144].
The biosynthesis of peptides and polyketides in microorganisms is a unique modular pathway regulated by NRPS and PKS. NRPS comprises modules, each of which integrates proteinogenic amino acids with non-proteinogenic amino acids, fatty acids, carbohydrates and other building blocks into peptide chains [142]. These peptide-synthesizing enzymes accept approximately 300 proteinogenic and nonproteinogenic substrates during the biosynthesis of non-ribosomal peptides [145]. In bacteria, PKS Type I are widely found to be responsible for polyketide chain elongation, processing and termination [146]. These modular enzymes are involved in the recognition, activation and condensation of coenzyme A (CoA) derivatives as the building blocks [141,146]. Not only are the secondary metabolites produced through this unique natural combinatorial biosynthetic pathway are important for growth, symbiotic interactions and protection against biotic and abiotic stresses, but they also possess various therapeutic activities (Table 3). Some of the compounds have been applied for the treatment of various acute and chronic diseases.

2.3.5. Ribosomal Peptides

Ribosomal peptides (RPs) are peptide chains of proteinogenic amino acids that can be found on ribosomes. In contrast to the biosynthesis of NRPs, only 20 proteinogenic amino acids are used as the building blocks during the biosynthesis of RPs [145]. There are three major RP families: cyanobactin, microviridins, and lantipeptides. Cyanobactin is diversely present in symbiotic and planktonic cyanobacteria [166]. Microviridins are the largest RPs, consisting of between 12 and 20 amino acids, which have been classified into four classes (Group I–IV) and are found in freshwater cyanobacteria [167]. Lantipeptides can be produced by four different classes (Class I–IV) of lantipeptidase in the cytosol of producing strains. Few studies have been conducted on cyanobacteria producing lantipeptides. However, comparative genomic analyses revealed that the marine Prochlorococcus and Synechococcus possess Class II lentipeptidase, ProcM [168,169]. The heterologous expression of procM and procA genes identified that ProcM can catalyze the dehydration and cyclization of all 29 different ProcA precursor peptides to produce the lantipeptides called prochlorosins [170]. Such an efficient biosynthetic pathway for generating prochlorosins with structural diversity is necessary for Prochlorococcus strains MIT9313 and MIT9303, as well as Synechococcus strain RS9916, which has a small genome size. Remarkably, RPs exhibit biological activities that could be used as natural drugs in the future (Table 4).
It is unclear on how the NRPs, PKs and RPs are involved in the survival of cyanobacteria; however, these posttranslational modified compounds are well known for their antibacterial properties, which could be important when in competition with other microbial species in the ecosystem [170]. For example, under poor nutrient conditions in hot springs, cyanobacteria and other bacterial classes, such as Deinococci, Alphaproteobacteria, Ignavibacteria, and Betaproteobacteria [180], may competefor organic and non-organic matters, as well as space for either exposure to sunlight or cover from direct sunlight. Additionally, benthic or sessile cyanobacteria may produce NRPs, PKs, and RPs for cell signaling in order to form epilithic/endolithic biofilms in water bodies, saline soil crusts or soil crusts of desert and polar environments. Moreover, the oligopeptides produced by cyanobacteria could be crucial for other organisms, such as eukaryotic algae, sponges, and plants, used as precursors for their metabolic pathways.

2.3.6. Phenolic Acids

Phenolic acids consist of one carboxyl group and one or more hydroxyl groups joined to the aromatic ring. These secondary metabolites are one of the largest groups of phenolic compounds. Phenolic acids are represented by hydrocinnamic acid, hydrobenzoic acid, phenylacetic acid and phenylpropionic acid derivatives from the shikimate pathway [181,182]. Phenolic acids produced by photosynthetic organisms are necessary for protection against oxidative damage that might be caused by reactive oxygen species (ROS) and the hydroxyl radical (OH).
In cyanobacteria, the accumulation of phenolic acids ensures the tolerance and adaptability of these photosynthetic microbes to various environmental stresses, which can cause the deposit of free radicals in cells, as well as chemical damage to deoxyribose and DNA. Notably, the accumulation of gallic acid, caffeic acid, chlorogenic, ferulic acid, and vanillic acid was detected when a high concentration of NaCI was supplied into the cultures of Plectonema boryanum, Haplosiphon intricatus, Anabaena doliolum, and Oscillatoria acuta [183], and thus suggesting that these phenolic acids play roles in the scavenging of free radicals under salt stress conditions. A recent study reported that the abundance of phenolic compounds in response to both cold and hot shocks might have stimulated the antioxidant capacity in halotolerant Halothece sp. PCC 7418 [184]. It is noteworthythat the synergistic effect of phenolic acids and other antioxidative compounds (flavonoids, MAAs and phycobiliproteins) is necessary for the response.
As with other phenolic compounds, many studies identified that plant phenolic acids also manifest antimicrobial [185] and antiviral properties [186]. Due to their ability to reduce oxidative damage or stress in cells, phenolic acids such as gentisic acid, gallic acid and syringic acid exhibitgood recovery in heart failure [187], memory loss [188] and wound healing [189], respectively. Previous studies have also reported that ferulic acids can produce skin whitening and anti-wrinkle effects [190] whereas gallic acid has anti-aging properties [191]. These therapeutic effects might also be produced by cyanobacterial phenolic acids due to their abilities to detoxify ROS and scavenge free radicals [183,184].

2.3.7. Flavonoids

Flavonoids are polyphenolic compounds that are widely distributed in plants [192]. These secondary metabolites can be classified into different subclasses: chalcones, flavanols, flavanones, flavones, isoflavones, flavonols, and anthocyanins. Remarkably, plant flavonoids exhibit antioxidant, anticancer, antiviral and anti-inflammatory properties [193,194].
In addition to phenolic acids, flavonoids are also antioxidants that are important for the survival of cyanobacteria. These antioxidative molecules, particularly quercertin and lutin, might facilitate Plectonema boryanum, Haplosiphon intricatus, Anabaena doliolum, and Oscillatoria acuta in salt acclimation mechanisms [183]. On the other hand, the chromatographic analysis identified that the thermophilic Leptolyngbyba sp. produces a high amount of luteolin-7-glucoside and naringenin [195], which could protect the cells from oxidative damage due to high temperatures. Additionally, naringenin not only plays a role as a strong free radical scavenger but also affects the growth and physiological functions of the halophilic Spirulina platensis and Arthrospira maxima and the freshwater Anabaena sp. by altering the cell wall and cellular membrane permeability [196]. These features are crucial to allow the secretion of exopolysaccharides (EPS) onto the surface of cyanobacterial cells for protection against unfavorable environmental conditions [196]. The oxidative power of the total flavonoids produced by cyanobacterial strains [183,184,195,197,198] suggests that these strong antioxidative compounds might also have pharmacological potentials similar to the plants flavonoids, such as nephroprotective [199], neuroprotective [200], anticancer [201], and antiatherosclerotic properties [202,203].

2.3.8. Vitamins

Vitamins are commonly synthesized by photosynthetic organisms. Vitamin B: B1 (thiamin), B2 (riboflavin), B3 (niacin), B5 (pantothenic acid), B6 (pyridoxine), B7 (biotin), B9 (folic acid), B12 (cobalamin), and C are water-soluble compounds, whereas vitamin A, D, E, and K are lipid-soluble compounds. Plants produce vitamin A, B, C, E, and K in most organ parts to alleviate the effects of environmental stresses [204]. However, not all plants produce all vitamins and in fact, vitamin D and K, as well as some types of vitamin B are scarcely present [205]. By contrast, microalgae including cyanobacteria can produce vitamin D, K and B12, which are not present in higher plants [205].
Arthrospira maxima, Anabaena cylindrica, and Synechococcus sp. displayed high contents of β-carotene (pro-vitamin A) [206,207]. Remarkably, these cyanobacteria produce much higher β-carotene than some fruits, such as carrots and oranges [205]. Similarly to in plants, this pro-vitamin A carotenoid compound produced by the cyanobacteria possesses a great oxidative efficacy against ROS, which is important for photooxidative protection. It is noteworthy that the efficacy of O2•− scavenging by β-carotene is greater than vitamin E and C [205].
Cyanobacteria are the major sources of B vitamins in marine and freshwater ecosystems. These water-soluble vitamins secreted by some cyanobacteria into the water bodies are important nutrients for other aquatic organisms [206,208,209]. Additionally, B vitamins might also be necessary in the metabolic pathways of cyanobacteria [205], as a high content of B2, B5 and B6 was detected in the freshwater Anabaena cylindrica [206] whereas, a marine Anabaena cylindrica was found to produce a high amount of B12. On the other hand, chromatographic analysis detected a high content of B2, B3, B9 and B12 in dried biomass of commercial Arthrospira maxima and Arthrospira platensis [210]. Notably, vitamin B also plays a role in cyanobacterial adaptation to environmental stresses. Nodularia spumegina accumulates B1 in response to salinity and temperature stresses [208]. Together with β-carotene, B1 ensures that this planktonic cyanobacterium is able to survive under high UV radiation [208].
Other than the higher plants, Anabaena cylindrica is also an excellent source of vitamin C [206]. This well-known antioxidant compound may provide protection against oxidative compounds in cyanobacteria.
A very low amount of vitamin D was detected in Arthrospira sp. [211]. As characterized in other algae species, this lipid-soluble vitamin might be important to ease the damage or degradation of cell membranes in cyanobacteria caused by UV radiation.
A high amount of vitamin E has been found in Nostoc sp. PCC 7120, Synechocystis sp. PCC 6803, Anabaena cylindrica, Synechococcus sp. PCC 7942 and Arthrospira maxima [206,207,212,213]. Remarkably, the vitamin E content is higher in these cyanobacteria compared to some common food sources [206]. The production of vitamin E is necessary for protection against photooxidative damage to PSII [214]. The accumulation of α-tocopherol was stimulated by the light intensity when Synechocystis sp. PCC 6803 was grown under photoautotrophic conditions [212]. Additionally, vitamin E also facilitates cyanobacteria to survive an emerging nutrient limitation. A study showed that Arthrospira maxima, Nostoc sp. PCC 7120 and Synechocystis sp. PCC 6803 produce low amount of vitamin E under optimum nitrogen availability [213]. However, Arthrospira sp. and Oscillatoria sp. synthesize high amounts of vitamin E in response to nitrogen deficiency at their logarithmic growth phase. Moreover, it is known that microalgae also produce vitamin E in response to nutrient limitation [215]. It is noteworthy that the production of vitamin E in microalgae is also a reaction in response to oxidative stress caused by metals [216,217]. This same antioxidative response could also happen in cyanobacteria, although so far, no study has been reported in these photosynthetic bacteria.
It was proposed that the marine Anabaena cylindrica possess a higher content of vitamin K1 than spinach and parsley [218]. Conversely, Spirulina sp. CS-785 produces a low amount of K1. Phylloquinone (vitamin K1) is synthesized by most cyanobacteria such as Anabaena variabilis, Mastigocladus laminosus, Nostoc muscorum, Prochlorococcus sp., Anacystis nidulans and Synechocystis sp. PCC 6803 [219,220,221]. Phylloquinone not only acts as an one-electron carrier at the A1 binding site of PSI, but it also provides protection against growth damage at high light intensity [222]. Similar to vitamin K1, menaquinone (vitamin K2) acts as a secondary electron acceptor of PS1 in Gloeobacter violaceus and Synechococcus sp. PCC 7002 [223,224]. Although phylloquinone and menaquinone exhibit structural similarity, the latter compound was absent in Synechococcus sp. PCC 7002, and two enzymes involved in its biosynthesis were missing in Gloeobacter violaceus.
Humans obtain vitamins through their diet. Vegetables, fruit, fish, and meat are great sources of vitamins. Currently, vitamin deficiencies occuring in humans are treated with synthetic vitamin analogs. For example, intramuscular injections or oral vitamin B12 therapy are the most common treatments for patients with vitamin B12 deficiency [225]. Those with vitamin D deficiency can be treated with oral ergocalciferol (vitamin D2) [226]. For adults with vitamin C deficiency and scurvy signs, oral ascorbic acid followed by a nutritious diet are always recommended [227,228]. Moreover, vitamin K1 can be administered via intramuscular injection or orally for people with vitamin K deficiency [229]. For adults with deficiency of vitamin A, vitamin A palmitate in oil is the most common method of treatment [230,231].
To date, only Spirulina (Arthrospira sp.) has been made available for consumption by humans as a supplementary diet. Indeed, several in vivo studies have showed the health benefits of well-characterized Spirulina, a rich source of vitamin B12, β-carotene and vitamin E. For example, Spirulina can improve bone strength and stiffness due to vitamin B12 deficiency [232], prevent ulcer formation [233] and recover blood retinol status [234].

2.3.9. Antimetabolites

Antimetabolites are small molecules that inhibit the biosynthetic pathway by binding to the active site of the target molecule. An unusual deoxy sugar, 7-deoxy-D-altro-2-heptulose (7-deoxysedoheptulose, 7dSh) obtained from the Synechococcus elongatus PCC 7942 culture supernatant was identified to show biological activity against several wild type organisms, specifically, Anabaena variabilis, Saccharomyces cerevisiae and Arabidopsis thaliana [235]. In vitro analysis suggested that this antimetabolite mimics 3-deoxy-D-arabino-heptulosonate 7-phosphate (DAHP), the substrate of 3-dehydroquinate (DHQ) synthase. The binding of 7dSH on DHQ synthase leads to the inhibition of the enzyme and consequently blocksthe shikimate pathway [235]. Additionally, a recent study detected the accumulation of 5-deoxyribose (5dR) and then 7dSh in the Synechococcus elongatus PCC 7942 culture supernatant under elevated CO2 conditions [236]. The formation of 5dR was reported to be derived from the 5-deoxyadenosine (5dAdo) salvage pathway as a detoxification strategy in order to protect the radical S-adenosyl-l-methionine (SAM) enzymes from feedback inhibition [236]. In fact, 5dR is continuously imported and exported by the cells and serves as a precursor for 7dSH, which is metabolized by transketolase activity when a relatively high extracellular 5dR concentration is reached [236]. This unique biosynthesis pathway strategy enables cyanobacteria to survive a niche competition by inhibiting the growth of other microalgae or bacteriaand is especially crucial for the unicellular cyanobacteria with small genome sizes and fewer plasmids [236].
Another cyanobacterial antimetabolite, a nonprotein amino acid β-methylamino-L-alanine (BMAA), may be involved in the nitrogen metabolism of cyanobacteria in order for the nitrogen-fixing microbes to survive under nutrient deprivation. Previous studies have suggested that the production of BMAA is correlated with nitrogen starvation under both natural and culture conditions, which results in the inhibition of the nitrogen assimilation pathway [237,238,239,240]. In turn, the concentration of BMAA was declined when a nitrogen source was added to the nitrogen-starved Microcystis PCC 7806 culture [241]. Although very little is known regarding the biological function of BMAA in cyanobacteria, Synechococcus sp. TAU-MAC 0499, Synechocystis PCC 6803 and Anabaena sp. PCC 7120 have been found to rapidly import exogenous BMAA [237,238,240]. BMAA was suggested to impair the activity of glutamine synthetase-glutamine-oxoglutarate aminotransferase (GS-GOGAT), the sequentially functioning enzymes that are involved in nitrogen assimilation, in the non-BMAA producer Synechococcus sp. TAU-MAC 0499 and the BMAA producer Synechocystis PCC 6803 [237,242]. Specifically, BMAA competes with glutamine to bind to GOGAT and acts as an inactivating factor [242]. On the other hand, nitrogenase activity was inhibited in Anabaena sp. PCC 7120 culture supplied with exogenous BMAA and the cyanobacterial growth was retarded by forming chlorotic cells. Similarly, the growth of BMAA producer Synechocystis PCC 6803 was arrested and the formation of chlorotic cells increased in the presence of exogenous BMAA [238]. Chlorotic is a dormant state of cyanobacterial cells in order to prolong their survival period under nitrogen starvation [243]. On the contrary, exogenous BMAA does not affect the physiology of Synechococcus sp. TAU-MAC 0499 [237]. A recent study reported a different response of growth retardation by the non-BMAA producers Microcystis aeruginosa (FACHB-836 and 905) [244] compared to Synechococcus sp. TAU-MAC 0499, whereas BMAA has no negative impacts on the BMAA producers Anabaena sp. FACHB-418 and Microcystis wesenbergii FACHB-908 [244]. To date, little is known about the effects of BMAA on cyanobacteria. However, it is tempting to speculate that cyanobacteria synthesize BMAA as a response to nutrient-limited conditions by either eliminating the competitors or forming dormant cells [237,238,244,245]. Additionally, the fact that a low concentration of bound BMAA was detected in the non-nitrogen-fixing Microcystis wesenbergii (FACHB-908) and Synechocystis (FACHB-898) suggested that BMAA may be involved in the formation of cyanobacterial proteins [244].
Remarkably, antibacterial, antifungal, and herbicidal properties exhibited by the unusual deoxy sugar 7dSh [235] suggest its applications in various fields including agriculture, water management, veterinary medicine, and human medicine. On the other hand, further confirmation on the biological function of BMAA in cyanobacteria could provide solutions to control cyanobacterial bloom and subsequently overcome its neurotoxic effects on humans, which are associated with amyotrophic lateral sclerosis, Parkinson’s disease, and Alzheimer’s disease [246].

3. Conclusions

Pharmacological effects exhibited by plant natural bioactive metabolites have led to their numerous applications in the treatment of serious and chronic diseases. In plants, bioactive metabolites are typically produced in low amounts as secondary metabolites. Therefore, a large amount of plant resources is required for traditional extraction methods of these compounds to obtain the industrial yield. However, these methods are known to be unsustainable. For decades, important discoveries of the biological activities possessed by the bioactive metabolites produced by cyanobacteria has attracted attention for modern therapy. These oxygenic photosynthetic microbes produce bioactive metabolites as a response to environmental stresses. Some of the compounds are produced in abundance to release the stress effects. Moreover, the emergence of synthetic biology tools has allowed the combinatorial synthesis of plant-derived biosynthetic genes involved in metabolic pathways to be heterologously expressed in cyanobacteria. In fact, the capacity to express P450, an enzyme involved in secondary metabolite production in plants, is beneficial in themetabolic engineering of cyanobacteria for the heterologous expression of the plant bioactive metabolic pathway. Together with this synthetic biology approach, the advancement in bioprocess engineering can produce natural bioactive compounds using sustainable approaches in order to meet industrial demands in the future.

Author Contributions

Writing—original draft preparation, P.N. and N.I.M.N.; Conceptualization, N.I.M.N. and A.N.S.; Writing—editing and proofreading, Z.R. and H.J.; resources, P.N., N.I.M.N. and L.-W.C. All authors have read and agreed to the published version of the manuscript.

Funding

We acknowledge funding for graduate research scheme to P.N by Ministry of Higher Ed ucation Malaysia, grant number FRGS/1/2019/STG03/UTM/02/4 awarded to N.I.M.N.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cho, S.M.; Jeoung, S.C.; Song, J.-Y.; Kupriyanova, E.V.; Pronina, N.A.; Lee, B.-W.; Jo, S.-W.; Park, B.-S.; Choi, S.-B.; Song, J.-J. Genomic survey and biochemical analysis of recombinant candidate cyanobacteriochromes reveals enrichment for near UV/violet sensors in the halotolerant and alkaliphilic cyanobacterium Microcoleus IPPAS B353. J. Biol. Chem. 2015, 290, 28502–28514. [Google Scholar] [CrossRef] [Green Version]
  2. Herbert, R.A.; Codd, G.A. Microbes in Extreme Environments; Academic Press: London, UK, 1986. [Google Scholar]
  3. Shih, P.M.; Wu, D.; Latifi, A.; Axen, S.D.; Fewer, D.P.; Talla, E.; Calteau, A.; Cai, F.; De Marsac, N.T.; Rippka, R. Improving the coverage of the cyanobacterial phylum using diversity-driven genome sequencing. Proc. Natl. Acad. Sci. USA 2013, 110, 1053–1058. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Bhaya, D.; Grossman, A.R.; Steunou, A.-S.; Khuri, N.; Cohan, F.M.; Hamamura, N.; Melendrez, M.C.; Bateson, M.M.; Ward, D.M.; Heidelberg, J.F. Population level functional diversity in a microbial community revealed by comparative genomic and metagenomic analyses. ISME J. 2007, 1, 703–713. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Cheevadhanarak, S.; Paithoonrangsarid, K.; Prommeenate, P.; Kaewngam, W.; Musigkain, A.; Tragoonrung, S.; Tabata, S.; Kaneko, T.; Chaijaruwanich, J.; Sangsrakru, D. Draft genome sequence of Arthrospira platensis C1 (PCC 9438). Stand. Genom. Sci. 2012, 6, 43–53. [Google Scholar] [CrossRef] [Green Version]
  6. Fujisawa, T.; Narikawa, R.; Okamoto, S.; Ehira, S.; Yoshimura, H.; Suzuki, I.; Masuda, T.; Mochimaru, M.; Takaichi, S.; Awai, K. Genomic structure of an economically important cyanobacterium, Arthrospira (Spirulina) platensis NIES-39. DNA Res. 2010, 17, 85–103. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Lefort, F.; Calmin, G.; Crovadore, J.; Falquet, J.; Hurni, J.-P.; Osteras, M.; Haldemann, F.; Farinelli, L. Whole-genome shotgun sequence of Arthrospira platensis strain Paraca, a cultivated and edible cyanobacterium. Genome Announc. 2014, 2, e00751-14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Carrieri, D.; Ananyev, G.; Lenz, O.; Bryant, D.A.; Dismukes, G.C. Contribution of a sodium ion gradient to energy conservation during fermentation in the cyanobacterium Arthrospira (Spirulina) maxima CS-328. Appl. Environ. Microbiol. 2011, 77, 7185–7194. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Janssen, P.J.; Morin, N.; Mergeay, M.; Leroy, B.; Wattiez, R.; Vallaeys, T.; Waleron, K.; Waleron, M.; Wilmotte, A.; Quillardet, P. Genome sequence of the edible cyanobacterium Arthrospira sp. PCC 8005. J. Bacteriol. 2010, 192, 2465–2466. [Google Scholar] [CrossRef] [Green Version]
  10. Dong, S.; Chen, J.; Wang, S.; Wu, Y.; Hou, H.; Li, M.; Yan, C. Draft genome sequence of cyanobacteria Arthrospira sp. TJSD091 isolated from seaside wetland. Mar. Genom. 2015, 24, 197–198. [Google Scholar] [CrossRef]
  11. Hirooka, S.; Hirose, Y.; Kanesaki, Y.; Higuchi, S.; Fujiwara, T.; Onuma, R.; Era, A.; Ohbayashi, R.; Uzuka, A.; Nozaki, H. Acidophilic green algal genome provides insights into adaptation to an acidic environment. Proc. Natl. Acad. Sci. USA 2017, 114, E8304–E8313. [Google Scholar] [CrossRef] [Green Version]
  12. Gross, W. Ecophysiology of algae living in highly acidic environments. Hydrobiologia 2000, 433, 31–37. [Google Scholar] [CrossRef]
  13. Ferris, M.J.; Sheehan, K.B.; Kuhl, M.; Cooksey, K.; Wigglesworth-Cooksey, B.; Harvey, R.; Henson, J.M. Algal species and light microenvironment in a low-pH, geothermal microbial mat community. Appl. Environ. Microbiol. 2005, 71, 7164–7171. [Google Scholar] [CrossRef] [Green Version]
  14. Khomutovska, N.; de Los Ríos, A.; Jasser, I. Diversity and Colonization Strategies of Endolithic Cyanobacteria in the Cold Mountain Desert of Pamir. Microorganisms 2021, 9, 6. [Google Scholar] [CrossRef] [PubMed]
  15. Ramos, V.; Castelo-Branco, R.; Leao, P.N.; Martins, J.; Carvalhal-Gomes, S.; Sobrinho da Silva, F.; Mendonca Filho, J.G.; Vasconcelos, V.M. Cyanobacterial diversity in microbial mats from the hypersaline lagoon system of Araruama, Brazil: An in-depth polyphasic study. Front. Microbiol. 2017, 8, 1233. [Google Scholar] [CrossRef] [Green Version]
  16. Sterner, R.W.; Reinl, K.L.; Lafrancois, B.M.; Brovold, S.; Miller, T.R. A first assessment of cyanobacterial blooms in oligotrophic Lake Superior. Limnol. Oceanogr. 2020, 65, 2984–2998. [Google Scholar] [CrossRef]
  17. Reinl, K.L.; Sterner, R.W.; Lafrancois, B.M.; Brovold, S. Fluvial seeding of cyanobacterial blooms in oligotrophic Lake Superior. Harmful Algae 2020, 100, 101941. [Google Scholar] [CrossRef] [PubMed]
  18. Nadeau, T.L.; Castenholz, R.W. Characterization of psychrophilic oscillatorians (cyanobacteria) from Antarctic meltwater ponds. J. Phycol. 2000, 36, 914–923. [Google Scholar] [CrossRef]
  19. Singh, S.M.; Elster, J. Cyanobacteria in Antarctic lake environments. In Algae and Cyanobacteria in Extreme Environments. Cellular Origin, Life in Extreme Habitats and Astrobiology; Seckbach, J., Ed.; Springer: Dordrecht, The Netherlands, 2007; Volume 11, pp. 303–320. [Google Scholar]
  20. Thangaraj, B.; Rajasekar, D.P.; Vijayaraghavan, R.; Garlapati, D.; Devanesan, A.A.; Lakshmanan, U.; Dharmar, P. Cytomorphological and nitrogen metabolic enzyme analysis of psychrophilic and mesophilic Nostoc sp.: A comparative outlook. 3 Biotech 2017, 7, 107. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Pedersen, D.; Miller, S.R. Photosynthetic temperature adaptation during niche diversification of the thermophilic cyanobacterium Synechococcus A/B clade. ISME J. 2017, 11, 1053–1057. [Google Scholar] [CrossRef] [PubMed]
  22. Maeda, K.; Tamura, J.; Okuda, Y.; Narikawa, R.; Midorikawa, T.; Ikeuchi, M. Genetic identification of factors for extracellular cellulose accumulation in the thermophilic cyanobacterium Thermo Synechococcus vulcanus: Proposal of a novel tripartite secretion system. Mol. Microbiol. 2018, 109, 121–134. [Google Scholar] [CrossRef] [Green Version]
  23. Tyagi, S.; Singh, R.K.; Tiwari, S.P. Anti-enterococcal and anti-oxidative potential of a thermophilic cyanobacterium, Leptolyngbya sp. HNBGU 003. Saudi J. Biol. Sci. 2021, 28, 4022–4028. [Google Scholar] [CrossRef]
  24. Karatay, S.E.; Dönmez, G.; Aksu, Z. Effective biosorption of phenol by the thermophilic cyanobacterium Phormidium sp. Water Sci. Technol. 2017, 76, 3190–3194. [Google Scholar] [CrossRef] [PubMed]
  25. El-Mohsnawy, E.; Abu-Khudir, R. A highly purified C-phycocyanin from thermophilic cyanobacterium Thermo Synechococcus elongatus and its cytotoxic activity assessment using an in vitro cell-based approach. J. Taibah Univ. Sci. 2020, 14, 1218–1225. [Google Scholar] [CrossRef]
  26. Ahmed, O.M.; Mahmoud, A.M.; Abdel-Moneim, A.; Ashour, M.B. Antidiabetic effects of hesperidin and naringin in type 2 diabetic rats. Diabetol. Croat. 2012, 41, 53–67. [Google Scholar]
  27. Zhu, Z.; Fu, F.; Qu, P.; Mak, E.W.K.; Jiang, H.; Zhang, R.; Zhu, Z.; Gao, K.; Hutchins, D.A. Interactions between ultraviolet radiation exposure and phosphorus limitation in the marine nitrogen-fixing cyanobacteria Trichodesmium and Crocosphaera. Limnol. Oceanogr. 2020, 65, 363–376. [Google Scholar] [CrossRef]
  28. Song, W.; Zhao, C.; Zhang, D.; Mu, S.; Pan, X. Different resistance to UV-B radiation of extracellular polymeric substances of two cyanobacteria from contrasting habitats. Front. Microbiol. 2016, 7, 1208. [Google Scholar] [CrossRef] [Green Version]
  29. Cohen, Y.; Jørgensen, B.B.; Revsbech, N.P.; Poplawski, R. Adaptation to hydrogen sulfide of oxygenic and anoxygenic photosynthesis among cyanobacteria. Appl. Environ. Microbiol. 1986, 51, 398–407. [Google Scholar] [CrossRef] [Green Version]
  30. Hamilton, T.L.; Klatt, J.M.; De Beer, D.; Macalady, J.L. Cyanobacterial photosynthesis under sulfidic conditions: Insights from the isolate Leptolyngbya sp. strain hensonii. ISME J. 2018, 12, 568–584. [Google Scholar] [CrossRef] [Green Version]
  31. Klatt, J.M.; Gomez-Saez, G.V.; Meyer, S.; Ristova, P.P.; Yilmaz, P.; Granitsiotis, M.S.; Macalady, J.L.; Lavik, G.; Polerecky, L.; Bühring, S.I. Versatile cyanobacteria control the timing and extent of sulfide production in a Proterozoic analog microbial mat. ISME J. 2020, 14, 3024–3037. [Google Scholar] [CrossRef]
  32. Stal, L.J.; Moezelaar, R. Fermentation in cyanobacteria. FEMS Microbiol. Rev. 1997, 21, 179–211. [Google Scholar] [CrossRef] [Green Version]
  33. Capone, D.G.; Burns, J.A.; Montoya, J.P.; Subramaniam, A.; Mahaffey, C.; Gunderson, T.; Michaels, A.F.; Carpenter, E.J. Nitrogen fixation by Trichodesmium spp.: An important source of new nitrogen to the tropical and subtropical North Atlantic Ocean. Glob. Biogeochem. Cycles 2005, 19. [Google Scholar] [CrossRef]
  34. Herrero, A.; Stavans, J.; Flores, E. The multicellular nature of filamentous heterocyst-forming cyanobacteria. FEMS Microbiol. Rev. 2016, 40, 831–854. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Mehda, S.; Muñoz-Martín, M.; Oustani, M.; Hamdi-Aïssa, B.; Perona, E.; Mateo, P. Microenvironmental Conditions Drive the Differential Cyanobacterial Community Composition of Biocrusts from the Sahara Desert. Microorganisms 2021, 9, 487. [Google Scholar] [CrossRef] [PubMed]
  36. Pushkareva, E.; Pessi, I.S.; Wilmotte, A.; Elster, J. Cyanobacterial community composition in Arctic soil crusts at different stages of development. FEMS Microbiol. Ecol. 2015, 91, fiv143. [Google Scholar] [CrossRef] [PubMed]
  37. Amarouche-Yala, S.; Benouadah, A.; López-García, P. Morphological and phylogenetic diversity of thermophilic cyanobacteria in Algerian hot springs. Extremophiles 2014, 18, 1035–1047. [Google Scholar] [CrossRef]
  38. Eisenberg, I.; Caycedo-Soler, F.; Harris, D.; Yochelis, S.; Huelga, S.F.; Plenio, M.B.; Adir, N.; Keren, N.; Paltiel, Y. Regulating the energy flow in a cyanobacterial light-harvesting antenna complex. J. Phys. Chem. B 2017, 121, 1240–1247. [Google Scholar] [CrossRef] [Green Version]
  39. Adir, N.; Dobrovetsky, Y.; Lerner, N. Structure of C-phycocyanin from the thermophilic cyanobacterium Synechococcus vulcanus at 2.5 Å: Structural implications for thermal stability in phycobilisome assembly. J. Mol. Biol. 2001, 313, 71–81. [Google Scholar] [CrossRef] [Green Version]
  40. Komenda, J. Role of two forms of the D1 protein in the recovery from photoinhibition of photosystem II in the cyanobacterium Synechococcus PCC 7942. Biochim. Biophys. Acta (BBA)-Bioenerg. 2000, 1457, 243–252. [Google Scholar] [CrossRef] [Green Version]
  41. Nishiyama, Y.; Los, D.A.; Hayashi, H.; Murata, N. Thermal protection of the oxygen-evolving machinery by PsbU, an extrinsic protein of photosystem II, in Synechococcus species PCC 7002. Plant Physiol. 1997, 115, 1473–1480. [Google Scholar] [CrossRef] [Green Version]
  42. Prihantini, N.B.; Fitrianti, A.N.; Sjamsuridzal, W.; Yokota, A. Growth temperature of hot springs filamentous cyanobacteria in artificial media. AIP Conf. Proc. 2020, 2242, 050012. [Google Scholar]
  43. Strunecký, O.; Kopejtka, K.; Goecke, F.; Tomasch, J.; Lukavský, J.; Neori, A.; Kahl, S.; Pieper, D.H.; Pilarski, P.; Kaftan, D. High diversity of thermophilic cyanobacteria in Rupite hot spring identified by microscopy, cultivation, single-cell PCR and amplicon sequencing. Extremophiles 2019, 23, 35–48. [Google Scholar] [CrossRef]
  44. Mittler, R.; Tel-or, E. Oxidative stress responses in the unicellular cyanobacterium Synechococcus PCC 7942. Free Radic. Res. Commun. 1991, 13, 845–850. [Google Scholar] [CrossRef]
  45. Ehling-Schulz, M.; Bilger, W.; Scherer, S. UV-B-induced synthesis of photoprotective pigments and extracellular polysaccharides in the terrestrial cyanobacterium Nostoc commune. J. Bacteriol. 1997, 179, 1940–1945. [Google Scholar] [CrossRef] [Green Version]
  46. Mloszewska, A.M.; Cole, D.B.; Planavsky, N.J.; Kappler, A.; Whitford, D.S.; Owttrim, G.W.; Konhauser, K.O. UV radiation limited the expansion of cyanobacteria in early marine photic environments. Nat. Commun. 2018, 9, 3088. [Google Scholar] [CrossRef]
  47. Mur, R.; Skulberg, O.M.; Utkilen, H. Cyanobacteria in the Environment. In Toxic Cyanobacteria in Water: A Guide to Their Public Health Consequences, Monitoring, and Management; Chorus, I., Bartram, J., Eds.; World Health Organization, Routledge: London, UK, 1999. [Google Scholar]
  48. Thajuddin, N.; Subramanian, G. Cyanobacterial biodiversity and potential applications in biotechnology. Curr. Sci. 2005, 89, 47–57. [Google Scholar]
  49. Rippka, R.; Deruelles, J.; Waterbury, J.B.; Herdman, M.; Stanier, R.Y. Generic assignments, strain histories and properties of pure cultures of cyanobacteria. Microbiology 1979, 111, 1–61. [Google Scholar] [CrossRef] [Green Version]
  50. Uyeda, J.C.; Harmon, L.J.; Blank, C.E. A comprehensive study of cyanobacterial morphological and ecological evolutionary dynamics through deep geologic time. PLoS ONE 2016, 11, e0162539. [Google Scholar] [CrossRef]
  51. Tiwari, G.L. On the morphology and life-history of a new species of Chroococcidiopsis Geitler (Chroococcales). Hydrobiologia 1972, 40, 177–182. [Google Scholar] [CrossRef]
  52. Demoulin, C.F.; Lara, Y.J.; Cornet, L.; François, C.; Baurain, D.; Wilmotte, A.; Javaux, E.J. Cyanobacteria evolution: Insight from the fossil record. Free Radic. Biol. Med. 2019, 140, 206–223. [Google Scholar] [CrossRef] [PubMed]
  53. Roldán, M.; Ramírez, M.; Del Campo, J.; Hernández-Mariné, M.; Komárek, J. Chalicogloea cavernicola gen. nov., sp. nov. (Chroococcales, Cyanobacteria), from low-light aerophytic environments: Combined molecular, phenotypic and ecological criteria. Int. J. Syst. Evol. Microbiol. 2013, 63, 2326–2333. [Google Scholar] [CrossRef] [PubMed]
  54. Margheri, M.C.; Ventura, S.; Kaštovský, J.; Komárek, J. The taxonomic validation of the cyanobacterial genus Halothece. Phycologia 2008, 47, 477–486. [Google Scholar] [CrossRef]
  55. Berrendero, E.; Valiente, E.F.; Perona, E.; Gómez, C.L.; Loza, V.; Muñoz-Martín, M.Á.; Mateo, P. Nitrogen fixation in a non-heterocystous cyanobacterial mat from a mountain river. Sci. Rep. 2016, 6, 30920. [Google Scholar] [CrossRef] [PubMed]
  56. Zhang, X.-J.; Feng, J.; Wang, G.-H.; Xie, S.-L. A morphological and phylogenetic study of a filamentous cyanobacterium, Microcoleus vaginatus, associated with the moss Mnium cuspidatum. Symbiosis 2014, 64, 43–51. [Google Scholar] [CrossRef]
  57. Taha, H.S.; El Bahr, M.K.; Seif, E.L.N.M.M. In Vitro Studies on Egyption Catharanthus roseus (L.). Ii. Effect of Biotic and Abiotic Stress on Indole Alkaloids Production. J. Appl. Sci. Res. 2009, 3, 3137–3144. [Google Scholar]
  58. Singh, B.; A Sharma, R.; K Vyas, G. Antimicrobial, Antineoplastic and Cytotoxic Activities of Indole Alkaloids from Tabernaemontana divaricata (L.) R. Br. Curr. Pharm. Anal. 2011, 7, 125–132. [Google Scholar] [CrossRef]
  59. El-Sayed, M.; Verpoorte, R. Catharanthus terpenoid indole alkaloids: Biosynthesis and regulation. Phytochem. Rev. 2007, 6, 277–305. [Google Scholar] [CrossRef] [Green Version]
  60. Mayser, P.; Schäfer, U.; Krämer, H.-J.; Irlinger, B.; Steglich, W. Pityriacitrin–an ultraviolet-absorbing indole alkaloid from the yeast Malassezia furfur. Arch. Dermatol. Res. 2002, 294, 131–134. [Google Scholar] [CrossRef]
  61. Aiello, A.; Borrelli, F.; Capasso, R.; Fattorusso, E.; Luciano, P.; Menna, M. Conicamin, a novel histamine antagonist from the mediterranean tunicate Aplidium conicum. Bioorganic Med. Chem. Lett. 2003, 13, 4481–4483. [Google Scholar] [CrossRef] [PubMed]
  62. Sauviat, M.-P.; Vercauteren, J.; Grimaud, N.; Jugé, M.; Nabil, M.; Petit, J.-Y.; Biard, J.-F. Sensitivity of cardiac background inward rectifying K+ outward current (I K1) to the alkaloids lepadiformines A, B, and C. J. Nat. Prod. 2006, 69, 558–562. [Google Scholar] [CrossRef]
  63. Kobayashi, H.; Kitamura, K.; Nagai, K.; Nakao, Y.; Fusetani, N.; van Soest, R.W.M.; Matsunaga, S. Carteramine A, an inhibitor of neutrophil chemotaxis, from the marine sponge Stylissa carteri. Tetrahedron Lett. 2007, 48, 2127–2129. [Google Scholar] [CrossRef]
  64. Sayed, K.A.E.; Khalil, A.A.; Yousaf, M.; Labadie, G.; Kumar, G.M.; Franzblau, S.G.; Mayer, A.M.S.; Avery, M.A.; Hamann, M.T. Semisynthetic studies on the manzamine alkaloids. J. Nat. Prod. 2008, 71, 300–308. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Pearce, A.N.; Chia, E.W.; Berridge, M.V.; Clark, G.R.; Harper, J.L.; Larsen, L.; Maas, E.W.; Page, M.J.; Perry, N.B.; Webb, V.L. Anti-inflammatory thiazine alkaloids isolated from the New Zealand ascidian Aplidium sp.: Inhibitors of the neutrophil respiratory burst in a model of gouty arthritis. J. Nat. Prod. 2007, 70, 936–940. [Google Scholar] [CrossRef] [PubMed]
  66. Pearce, A.N.; Chia, E.W.; Berridge, M.V.; Maas, E.W.; Page, M.J.; Webb, V.L.; Harper, J.L.; Copp, B.R. E/Z-rubrolide O, an anti-inflammatory halogenated furanone from the New Zealand ascidian Synoicum n. sp. J. Nat. Prod. 2007, 70, 111–113. [Google Scholar] [CrossRef] [PubMed]
  67. Chilczuk, T.; Steinborn, C.; Breinlinger, S.; Zimmermann-Klemd, A.M.; Huber, R.; Enke, H.; Enke, D.; Niedermeyer, T.H.J.; Gründemann, C. Hapalindoles from the cyanobacterium Hapalosiphon sp. inhibit T cell proliferation. Planta Med. 2020, 86, 96–103. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Moore, R.E.; Cheuk, C.; Yang, X.Q.G.; Patterson, G.M.L.; Bonjouklian, R.; Smitka, T.A.; Mynderse, J.S.; Foster, R.S.; Jones, N.D.; Swartzendruber, J.K. Hapalindoles, antibacterial and antimycotic alkaloids from the cyanophyte Hapalosiphon fontinalis. J. Org. Chem. 1987, 52, 1036–1043. [Google Scholar] [CrossRef]
  69. Moore, R.E.; Cheuk, C.; Patterson, G.M.L. Hapalindoles: New alkaloids from the blue-green alga Hapalosiphon fontinalis. J. Am. Chem. Soc. 1984, 106, 6456–6457. [Google Scholar] [CrossRef]
  70. Kim, H.; Lantvit, D.; Hwang, C.H.; Kroll, D.J.; Swanson, S.M.; Franzblau, S.G.; Orjala, J. Indole alkaloids from two cultured cyanobacteria, Westiellopsis sp. and Fischerella muscicola. Bioorganic Med. Chem. 2012, 20, 5290–5295. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Hillwig, M.L.; Zhu, Q.; Liu, X. Biosynthesis of ambiguine indole alkaloids in cyanobacterium Fischerella ambigua. ACS Chem. Biol. 2014, 9, 372–377. [Google Scholar] [CrossRef]
  72. Hillwig, M.L.; Fuhrman, H.A.; Ittiamornkul, K.; Sevco, T.J.; Kwak, D.H.; Liu, X. Identification and characterization of a welwitindolinone alkaloid biosynthetic gene cluster in the stigonematalean cyanobacterium Hapalosiphon welwitschii. ChemBioChem 2014, 15, 665–669. [Google Scholar] [CrossRef] [Green Version]
  73. Micallef, M.L.; Sharma, D.; Bunn, B.M.; Gerwick, L.; Viswanathan, R.; Moffitt, M.C. Comparative analysis of hapalindole, ambiguine and welwitindolinone gene clusters and reconstitution of indole-isonitrile biosynthesis from cyanobacteria. BMC Microbiol. 2014, 14, 1–18. [Google Scholar] [CrossRef] [Green Version]
  74. Smitka, T.A.; Bonjouklian, R.; Doolin, L.; Jones, N.D.; Deeter, J.B.; Yoshida, W.Y.; Prinsep, M.R.; Moore, R.E.; Patterson, G.M.L. Ambiguine isonitriles, fungicidal hapalindole-type alkaloids from three genera of blue-green algae belonging to the Stigonemataceae. J. Org. Chem. 1992, 57, 857–861. [Google Scholar] [CrossRef]
  75. Park, A.; Moore, R.E.; Patterson, G.M.L. Fischerindole L, a new isonitrile from the terrestrial blue-green alga Fischerella muscicola. Tetrahedron Lett. 1992, 33, 3257–3260. [Google Scholar] [CrossRef]
  76. Mo, S.; Krunic, A.; Santarsiero, B.D.; Franzblau, S.G.; Orjala, J. Hapalindole-related alkaloids from the cultured cyanobacterium Fischerella ambigua. Phytochemistry 2010, 71, 2116–2123. [Google Scholar] [CrossRef] [Green Version]
  77. Jimenez, J.I.; Huber, U.; Moore, R.E.; Patterson, G.M.L. Oxidized welwitindolinones from terrestrial fischerella spp. J. Nat. Prod. 1999, 62, 569–572. [Google Scholar] [CrossRef]
  78. Stratmann, K.; Moore, R.E.; Bonjouklian, R.; Deeter, J.B.; Patterson, G.M.L.; Shaffer, S.; Smith, C.D.; Smitka, T.A. Welwitindolinones, unusual alkaloids from the blue-green algae Hapalosiphon welwitschii and Westiella intricata. Relationship to fischerindoles and hapalinodoles. J. Am. Chem. Soc. 1994, 116, 9935–9942. [Google Scholar] [CrossRef]
  79. Richter, J.M.; Ishihara, Y.; Masuda, T.; Whitefield, B.W.; Llamas, T.; Pohjakallio, A.; Baran, P.S. Enantiospecific total synthesis of the hapalindoles, fischerindoles, and welwitindolinones via a redox economic approach. J. Am. Chem. Soc. 2008, 130, 17938–17954. [Google Scholar] [CrossRef] [Green Version]
  80. Demay, J.; Bernard, C.; Reinhardt, A.; Marie, B. Natural products from cyanobacteria: Focus on beneficial activities. Mar. Drugs 2019, 17, 320. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Walton, K.; Berry, J.P. Indole alkaloids of the Stigonematales (Cyanophyta): Chemical diversity, biosynthesis and biological activity. Mar. Drugs 2016, 14, 73. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Castenholz, R.W.; Garcia-Pichel, F. Cyanobacterial responses to UV radiation. In Ecology of Cyanobacteria II; Whitton, B.A., Ed.; Springer: Dordrecht, The Netherlands, 2012; pp. 481–499. [Google Scholar]
  83. Nägeli, C. Gattungen Einzelliger Algen: Physiologisch und Systematisch Bearbeitet; Friedrich Schulthess: Zürich, Switzerland, 1849. [Google Scholar]
  84. Orellana, G.; Gómez-Silva, B.; Urrutia, M.; Galetović, A. UV-A Irradiation Increases Scytonemin Biosynthesis in Cyanobacteria Inhabiting Halites at Salar Grande, Atacama Desert. Microorganisms 2020, 8, 1690. [Google Scholar] [CrossRef]
  85. 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] [Green Version]
  86. Dillon, J.G.; Tatsumi, C.M.; Tandingan, P.G.; Castenholz, R.W. Effect of environmental factors on the synthesis of scytonemin, a UV-screening pigment, in a cyanobacterium (Chroococcidiopsis sp.). Arch. Microbiol. 2002, 177, 322–331. [Google Scholar] [CrossRef]
  87. Cao, R.; Peng, W.; Wang, Z.; Xu, A. β-Carboline alkaloids: Biochemical and pharmacological functions. Curr. Med. Chem. 2007, 14, 479–500. [Google Scholar] [CrossRef]
  88. Volk, R.-B. Screening of microalgal culture media for the presence of algicidal compounds and isolation and identification of two bioactive metabolites, excreted by the cyanobacteria Nostoc insulare and Nodularia harveyana. J. Appl. Phycol. 2005, 17, 339–347. [Google Scholar] [CrossRef]
  89. Volk, R.-B.; Furkert, F.H. Antialgal, antibacterial and antifungal activity of two metabolites produced and excreted by cyanobacteria during growth. Microbiol. Res. 2006, 161, 180–186. [Google Scholar] [CrossRef]
  90. Becher, P.G.; Baumann, H.I.; Gademann, K.; Jüttner, F. The cyanobacterial alkaloid nostocarboline: An inhibitor of acetylcholinesterase and trypsin. J. Appl. Phycol. 2009, 21, 103–110. [Google Scholar] [CrossRef] [Green Version]
  91. Breitmaier, E. Terpenes: Importance, general structure, and biosynthesis. Terpenes Flavors Fragr. Pharmaca Pheromones 2006, 1, 1–3. [Google Scholar]
  92. Kandi, S.; Godishala, V.; Rao, P.; Ramana, K.V. Biomedical significance of terpenes: An insight. Biomedicine 2015, 3, 8–10. [Google Scholar]
  93. Abdallah, I.I.; Quax, W.J. A Glimpse into the Biosynthesis of Terpenoids. In Proceedings of the International Conference on Natural Resources and Life Sciences (2016), Surabaya, Indonesia, 20–21 October 2016; KnE Life Sciences: Dubai, United Arab Emirates, 2017; pp. 81–98. [Google Scholar]
  94. Gershenzon, J.; Dudareva, N. The function of terpene natural products in the natural world. Nat. Chem. Biol. 2007, 3, 408–414. [Google Scholar] [CrossRef] [PubMed]
  95. Dittmann, E.; Gugger, M.; Sivonen, K.; Fewer, D.P. Natural product biosynthetic diversity and comparative genomics of the cyanobacteria. Trends Microbiol. 2015, 23, 642–652. [Google Scholar] [CrossRef] [PubMed]
  96. Yamada, Y.; Kuzuyama, T.; Komatsu, M.; Shin-Ya, K.; Omura, S.; Cane, D.E.; Ikeda, H. Terpene synthases are widely distributed in bacteria. Proc. Natl. Acad. Sci. USA 2015, 112, 857–862. [Google Scholar] [CrossRef] [Green Version]
  97. Devi, S.; Rani, N.; Sagar, A. GC-MS Analysis and antioxidant activity of two species of cyanobacteria isolated from Drang salt mine of district Mandi, Himachal Pradesh, India. Plant Arch. 2020, 20, 7505–7510. [Google Scholar]
  98. Höckelmann, C.; Becher, P.G.; von Reuss, S.H.; Jüttner, F. Sesquiterpenes of the geosmin-producing cyanobacterium Calothrix PCC 7507 and their toxicity to invertebrates. Z. Nat. C 2009, 64, 49–55. [Google Scholar] [CrossRef] [Green Version]
  99. Dienst, D.; Wichmann, J.; Mantovani, O.; Rodrigues, J.S.; Lindberg, P. High density cultivation for efficient sesquiterpenoid biosynthesis in Synechocystis sp. PCC 6803. Sci. Rep. 2020, 10, 5932. [Google Scholar] [CrossRef] [Green Version]
  100. Agger, S.A.; Lopez-Gallego, F.; Hoye, T.R.; Schmidt-Dannert, C. Identification of sesquiterpene synthases from Nostoc punctiforme PCC 73102 and Nostoc sp. strain PCC 7120. J. Bacteriol. 2008, 190, 6084–6096. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Unson, M.D.; Faulkner, D.J. Cyanobacterial symbiont biosynthesis of chlorinated metabolites from Dysidea herbacea (Porifera). Experientia 1993, 49, 349–353. [Google Scholar] [CrossRef]
  102. Jahnke, L.L.; Embaye, T.; Hope, J.; Turk, K.A.; Van Zuilen, M.; Des Marais, D.J.; Farmer, J.D.; Summons, R.E. Lipid biomarker and carbon isotopic signatures for stromatolite-forming, microbial mat communities and Phormidium cultures from Yellowstone National Park. Geobiology 2004, 2, 31–47. [Google Scholar] [CrossRef]
  103. Summons, R.E.; Jahnke, L.L.; Hope, J.M.; Logan, G.A. 2-Methylhopanoids as biomarkers for cyanobacterial oxygenic photosynthesis. Nature 1999, 400, 554–557. [Google Scholar] [CrossRef] [PubMed]
  104. Garby, T.J.; Matys, E.D.; Ongley, S.E.; Salih, A.; Larkum, A.W.D.; Walter, M.R.; Summons, R.E.; Neilan, B.A. Lack of methylated hopanoids renders the cyanobacterium Nostoc punctiforme sensitive to osmotic and pH stress. Appl. Environ. Microbiol. 2017, 83, e00777-17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Hirschberg, J.; Chamovitz, D. Carotenoids in cyanobacteria. In The Molecular Biology of Cyanobacteria. Advances in Photosynthesis, 1st ed.; Bryant, D.A., Ed.; Springer: Dordrecht, Netherlands, 1994; Volume 1, pp. 559–579. [Google Scholar]
  106. Merhan, O. The biochemistry and antioxidant properties of carotenoids. Carotenoids 2017, 5, 51. [Google Scholar]
  107. Patias, L.D.; Fernandes, A.S.; Petry, F.C.; Mercadante, A.Z.; Jacob-Lopes, E.; Zepka, L.Q. Carotenoid profile of three microalgae/cyanobacteria species with peroxyl radical scavenger capacity. Food Res. Int. 2017, 100, 260–266. [Google Scholar] [CrossRef] [PubMed]
  108. Pagels, F.; Salvaterra, D.; Amaro, H.M.; Lopes, G.; Sousa-Pinto, I.; Vasconcelos, V.; Guedes, A.C. Bioactive potential of Cyanobium sp. pigment-rich extracts. J. Appl. Phycol. 2020, 32, 3031–3040. [Google Scholar] [CrossRef]
  109. Kelman, D.; Ben-Amotz, A.; Berman-Frank, I. Carotenoids provide the major antioxidant defence in the globally significant N2-fixing marine cyanobacterium Trichodesmium. Environ. Microbiol. 2009, 11, 1897–1908. [Google Scholar] [CrossRef] [PubMed]
  110. Kusama, Y.; Inoue, S.; Jimbo, H.; Takaichi, S.; Sonoike, K.; Hihara, Y.; Nishiyama, Y. Zeaxanthin and echinenone protect the repair of photosystem II from inhibition by singlet oxygen in Synechocystis sp. PCC 6803. Plant Cell Physiol. 2015, 56, 906–916. [Google Scholar] [CrossRef] [PubMed]
  111. Boucar, M.C.M.; Shen, L.-Q.; Wang, K.; Zhang, Z.-C.; Qiu, B.-S. UV-B irradiation enhances the production of unique mycosporine-like amino acids and carotenoids in the subaerial cyanobacterium Pseudanabaena sp. CCNU1. Eur. J. Phycol. 2021, 56, 316–323. [Google Scholar] [CrossRef]
  112. Vítek, P.; Ascaso, C.; Artieda, O.; Casero, M.C.; Wierzchos, J. Discovery of carotenoid red-shift in endolithic cyanobacteria from the Atacama Desert. Sci. Rep. 2017, 7, 11116. [Google Scholar] [CrossRef] [PubMed]
  113. Lopes, G.; Clarinha, D.; Vasconcelos, V. Carotenoids from cyanobacteria: A biotechnological approach for the topical treatment of psoriasis. Microorganisms 2020, 8, 302. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Morone, J.; Lopes, G.; Preto, M.; Vasconcelos, V.; Martins, R. Exploitation of Filamentous and Picoplanktonic Cyanobacteria for Cosmetic Applications: Potential to Improve Skin Structure and Preserve Dermal Matrix Components. Mar. Drugs 2020, 18, 486. [Google Scholar] [CrossRef]
  115. Mo, S.; Krunic, A.; Pegan, S.D.; Franzblau, S.G.; Orjala, J. An antimicrobial guanidine-bearing sesterterpene from the cultured cyanobacterium Scytonema sp. J. Nat. Prod. 2009, 72, 2043–2045. [Google Scholar] [CrossRef] [Green Version]
  116. Cabanillas, A.H.; Tena Pérez, V.C.; Maderuelo Corral, S.; Rosero Valencia, D.F.; Martel Quintana, A.; Ortega Doménech, M.; Rumbero Sánchez, A.n. Cybastacines A and B: Antibiotic sesterterpenes from a Nostoc sp. cyanobacterium. J. Nat. Prod. 2018, 81, 410–413. [Google Scholar] [CrossRef]
  117. Krunic, A.; Vallat, A.; Mo, S.; Lantvit, D.D.; Swanson, S.M.; Orjala, J. Scytonemides A and B, cyclic peptides with 20S proteasome inhibitory activity from the cultured cyanobacterium Scytonema hofmanii. J. Nat. Prod. 2010, 73, 1927–1932. [Google Scholar] [CrossRef] [Green Version]
  118. Geraldes, V.; Jacinavicius, F.R.; Genuário, D.B.; Pinto, E. Identification and distribution of mycosporine-like amino acids in Brazilian cyanobacteria using ultrahigh-performance liquid chromatography with diode array detection coupled to quadrupole time-of-flight mass spectrometry. Rapid Commun. Mass Spectrom. 2020, 34, e8634. [Google Scholar] [CrossRef]
  119. 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] [Green Version]
  120. 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] [Green Version]
  121. Shukla, V.; Kumari, R.; Patel, D.K.; Upreti, D.K. Characterization of the diversity of mycosporine-like amino acids in lichens from high altitude region of Himalaya. Amino Acids 2016, 48, 129–136. [Google Scholar] [CrossRef]
  122. Pathak, J.; Ahmed, H.; Singh, S.P.; Häder, D.-P.; Sinha, R.P. Genetic regulation of scytonemin and mycosporine-like amino acids (MAAs) biosynthesis in cyanobacteria. Plant Gene 2019, 17, 100172. [Google Scholar] [CrossRef]
  123. Geraldes, V.; de Medeiros, L.S.; Lima, S.T.; Alvarenga, D.O.; Gacesa, R.; Long, P.F.; Fiore, M.F.; Pinto, E. Genetic and biochemical evidence for redundant pathways leading to mycosporine-like amino acid biosynthesis in the cyanobacterium Sphaerospermopsis torques-reginae ITEP-024. Harmful Algae 2020, 35, 177–187. [Google Scholar]
  124. 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]
  125. 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]
  126. 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]
  127. Werner, N.; Orfanoudaki, M.; Hartmann, A.; Ganzera, M.; Sommaruga, R. Low temporal dynamics of mycosporine-like amino acids in benthic cyanobacteria from an alpine lake. Freshwat. Biol. 2021, 66, 169–176. [Google Scholar] [CrossRef] [PubMed]
  128. Mueller, D.R.; Vincent, W.F.; Bonilla, S.; Laurion, I. Extremotrophs, extremophiles and broadband pigmentation strategies in a high arctic ice shelf ecosystem. FEMS Microbiol. Ecol. 2005, 53, 73–87. [Google Scholar] [CrossRef]
  129. Llewellyn, C.A.; Greig, C.; Silkina, A.; Kultschar, B.; Hitchings, M.D.; Farnham, G. Mycosporine-like amino acid and aromatic amino acid transcriptome response to UV and far-red light in the cyanobacterium Chlorogloeopsis fritschii PCC 6912. Sci. Rep. 2020, 10, 20638. [Google Scholar] [CrossRef]
  130. Couradeau, E.; Karaoz, U.; Lim, H.C.; Da Rocha, U.N.; Northen, T.; Brodie, E.; Garcia-Pichel, F. Bacteria increase arid-land soil surface temperature through the production of sunscreens. Nat. Commun. 2016, 7, 10373. [Google Scholar] [CrossRef] [Green Version]
  131. Nishida, Y.; Kumagai, Y.; Michiba, S.; Yasui, H.; Kishimura, H. Efficient extraction and antioxidant capacity of mycosporine-like amino acids from red alga Dulse Palmaria palmata in Japan. Mar. Drugs 2020, 18, 502. [Google Scholar] [CrossRef] [PubMed]
  132. 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]
  133. Patipong, T.; Hibino, T.; Waditee-Sirisattha, R.; Kageyama, H. Efficient bioproduction of mycosporine-2-glycine, which functions as potential osmoprotectant, using Escherichia coli cells. Nat. Prod. Commun. 2017, 12. [Google Scholar] [CrossRef] [Green Version]
  134. Tarasuntisuk, S.; Patipong, T.; Hibino, T.; Waditee-Sirisattha, R.; Kageyama, H. Inhibitory effects of mycosporine-2-glycine isolated from a halotolerant cyanobacterium on protein glycation and collagenase activity. Lett. Appl. Microbiol. 2018, 67, 314–320. [Google Scholar] [CrossRef]
  135. Kageyama, H.; Waditee-Sirisattha, R. Mycosporine-like amino acids as multifunctional secondary metabolites in cyanobacteria: From biochemical to application aspects. Stud. Nat. Prod. Chem. 2018, 59, 153–194. [Google Scholar]
  136. Ishihara, K.; Watanabe, R.; Uchida, H.; Suzuki, T.; Yamashita, M.; Takenaka, H.; Nazifi, E.; Matsugo, S.; Yamaba, M.; Sakamoto, T. Novel glycosylated mycosporine-like amino acid, 13-O-(β-galactosyl)-porphyra-334, from the edible cyanobacterium Nostoc sphaericum-protective activity on human keratinocytes from UV light. J. Photochem. Photobiol. B Biol. 2017, 172, 102–108. [Google Scholar] [CrossRef]
  137. Soule, T.; Garcia-Pichel, F. Ultraviolet photoprotective compounds from cyanobacteria in biomedical applications. Cyanobacteria Econ. Perspect. 2014, 119–143. [Google Scholar] [CrossRef] [Green Version]
  138. Daniel, S.; Cornelia, S.; Fred, Z. UV-A sunscreen from red algae for protection against premature skin aging. Cosmet Toilet. Manuf. Worldw. 2004, 139–143. [Google Scholar]
  139. Andre, G.; Pellegrini, M.; Pellegrini, L. Algal Extracts Containing Amino Acid Analogs of Mycosporin Are Useful as Dermatological Protecting Agents against Ultraviolet Radiation; Institut National De La Propriete Industrielle: Courbevoie, France, 2001; pp. 1–22. [Google Scholar]
  140. Maurya, S.K.; Mishra, R. Importance of bioinformatics in genome Mining of Cyanobacteria for production of bioactive compounds. In Cyanobacteria, 1st ed.; Academic Press: London, UK, 2019; pp. 477–506. [Google Scholar]
  141. Wang, H.; Fewer, D.P.; Holm, L.; Rouhiainen, L.; Sivonen, K. Atlas of nonribosomal peptide and polyketide biosynthetic pathways reveals common occurrence of nonmodular enzymes. Proc. Natl. Acad. Sci. USA 2014, 111, 9259–9264. [Google Scholar] [CrossRef] [Green Version]
  142. Welker, M.; Von Döhren, H. Cyanobacterial peptides—nature’s own combinatorial biosynthesis. FEMS Microbiol. Rev. 2006, 30, 530–563. [Google Scholar] [CrossRef] [Green Version]
  143. Larsson, J.; Nylander, J.A.; Bergman, B. Genome fluctuations in cyanobacteria reflect evolutionary, developmental and adaptive traits. BMC Evol. Biol. 2011, 11, 187. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Dufresne, A.; Salanoubat, M.; Partensky, F.; Artiguenave, F.; Axmann, I.M.; Barbe, V.; Duprat, S.; Galperin, M.Y.; Koonin, E.V.; Le Gall, F. Genome sequence of the cyanobacterium Prochlorococcus marinus SS120, a nearly minimal oxyphototrophic genome. Proc. Natl. Acad. Sci. USA 2003, 100, 10020–10025. [Google Scholar] [CrossRef] [Green Version]
  145. Kehr, J.-C.; Picchi, D.G.; Dittmann, E. Natural product biosyntheses in cyanobacteria: A treasure trove of unique enzymes. Beilstein J. Org. Chem. 2011, 7, 1622–1635. [Google Scholar] [CrossRef] [Green Version]
  146. Dutta, S.; Whicher, J.R.; Hansen, D.A.; Hale, W.A.; Chemler, J.A.; Congdon, G.R.; Narayan, A.R.H.; Håkansson, K.; Sherman, D.H.; Smith, J.L. Structure of a modular polyketide synthase. Nature 2014, 510, 512–517. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Nishizawa, T.; Asayama, M.; Fujii, K.; Harada, K.-I.; Shirai, M. Genetic analysis of the peptide synthetase genes for a cyclic heptapeptide microcystin in Microcystis spp. J. Biochem. 1999, 126, 520–529. [Google Scholar] [CrossRef] [PubMed]
  148. Nishizawa, T.; Ueda, A.; Asayama, M.; Fujii, K.; Harada, K.-I.; Ochi, K.; Shirai, M. Polyketide synthase gene coupled to the peptide synthetase module involved in the biosynthesis of the cyclic heptapeptide microcystin. J. Biochem. 2000, 127, 779–789. [Google Scholar] [CrossRef] [Green Version]
  149. Tillett, D.; Dittmann, E.; Erhard, M.; Von Döhren, H.; Börner, T.; Neilan, B.A. Structural organization of microcystin biosynthesis in Microcystis aeruginosa PCC 7806: An integrated peptide–polyketide synthetase system. Chem. Biol. 2000, 7, 753–764. [Google Scholar] [CrossRef] [Green Version]
  150. Ishida, K.; Christiansen, G.; Yoshida, W.Y.; Kurmayer, R.; Welker, M.; Valls, N.; Bonjoch, J.; Hertweck, C.; Börner, T.; Hemscheidt, T. Biosynthesis and structure of aeruginoside 126A and 126B, cyanobacterial peptide glycosides bearing a 2-carboxy-6-hydroxyoctahydroindole moiety. Chem. Biol. 2007, 14, 565–576. [Google Scholar] [CrossRef] [Green Version]
  151. Mihali, T.K.; Kellmann, R.; Muenchhoff, J.; Barrow, K.D.; Neilan, B.A. Characterization of the gene cluster responsible for cylindrospermopsin biosynthesis. Appl. Environ. Microbiol. 2008, 74, 716–722. [Google Scholar] [CrossRef] [Green Version]
  152. Rouhiainen, L.; Jokela, J.; Fewer, D.P.; Urmann, M.; Sivonen, K. Two alternative starter modules for the non-ribosomal biosynthesis of specific anabaenopeptin variants in Anabaena (Cyanobacteria). Chem. Biol. 2010, 17, 265–273. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Tidgewell, K.; Engene, N.; Byrum, T.; Media, J.; Doi, T.; Valeriote, F.A.; Gerwick, W.H. Evolved diversification of a modular natural product pathway: Apratoxins F and G, two cytotoxic cyclic depsipeptides from a Palmyra collection of Lyngbya bouillonii. ChemBioChem 2010, 11, 1458–1466. [Google Scholar] [CrossRef] [PubMed]
  154. Grindberg, R.V.; Ishoey, T.; Brinza, D.; Esquenazi, E.; Coates, R.C.; Liu, W.-T.; Gerwick, L.; Dorrestein, P.C.; Pevzner, P.; Lasken, R. Single cell genome amplification accelerates identification of the apratoxin biosynthetic pathway from a complex microbial assemblage. PLoS ONE 2011, 6, e18565. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Edwards, D.J.; Gerwick, W.H. Lyngbyatoxin biosynthesis: Sequence of biosynthetic gene cluster and identification of a novel aromatic prenyltransferase. J. Am. Chem. Soc. 2004, 126, 11432–11433. [Google Scholar] [CrossRef] [PubMed]
  156. Ramaswamy, A.V.; Sorrels, C.M.; Gerwick, W.H. Cloning and biochemical characterization of the hectochlorin biosynthetic gene cluster from the marine cyanobacterium Lyngbya majuscula. J. Nat. Prod. 2007, 70, 1977–1986. [Google Scholar] [CrossRef] [PubMed]
  157. Chang, Z.; Flatt, P.; Gerwick, W.H.; Nguyen, V.-A.; Willis, C.L.; Sherman, D.H. The barbamide biosynthetic gene cluster: A novel marine cyanobacterial system of mixed polyketide synthase (PKS)-non-ribosomal peptide synthetase (NRPS) origin involving an unusual trichloroleucyl starter unit. Gene 2002, 296, 235–247. [Google Scholar] [CrossRef]
  158. Chang, Z.; Sitachitta, N.; Rossi, J.V.; Roberts, M.A.; Flatt, P.M.; Jia, J.; Sherman, D.H.; Gerwick, W.H. Biosynthetic Pathway and Gene Cluster Analysis of Curacin A, an Antitubulin Natural Product from the Tropical Marine Cyanobacterium Lyngbya m ajuscula. J. Nat. Prod. 2004, 67, 1356–1367. [Google Scholar] [CrossRef]
  159. Edwards, D.J.; Marquez, B.L.; Nogle, L.M.; McPhail, K.; Goeger, D.E.; Roberts, M.A.; Gerwick, W.H. Structure and biosynthesis of the jamaicamides, new mixed polyketide-peptide neurotoxins from the marine cyanobacterium Lyngbya majuscula. Chem. Biol. 2004, 11, 817–833. [Google Scholar] [CrossRef] [Green Version]
  160. Hoffmann, D.; Hevel, J.M.; Moore, R.E.; Moore, B.S. Sequence analysis and biochemical characterization of the nostopeptolide A biosynthetic gene cluster from Nostoc sp. GSV224. Gene 2003, 311, 171–180. [Google Scholar] [CrossRef]
  161. Becker, J.E.; Moore, R.E.; Moore, B.S. Cloning, sequencing, and biochemical characterization of the nostocyclopeptide biosynthetic gene cluster: Molecular basis for imine macrocyclization. Gene 2004, 325, 35–42. [Google Scholar] [CrossRef]
  162. Magarvey, N.A.; Beck, Z.Q.; Golakoti, T.; Ding, Y.; Huber, U.; Hemscheidt, T.K.; Abelson, D.; Moore, R.E.; Sherman, D.H. Biosynthetic characterization and chemoenzymatic assembly of the cryptophycins. Potent anticancer agents from Nostoc cyanobionts. ACS Chem. Biol. 2006, 1, 766–779. [Google Scholar] [CrossRef]
  163. Mareš, J.; Hájek, J.; Urajová, P.; Kopecký, J.; Hrouzek, P. A hybrid non-ribosomal peptide/polyketide synthetase containing fatty-acyl ligase (FAAL) synthesizes the β-amino fatty acid lipopeptides puwainaphycins in the Cyanobacterium Cylindrospermum alatosporum. PLoS ONE 2014, 9, e111904. [Google Scholar] [CrossRef] [Green Version]
  164. Gupta, V.; Vyas, D. Antimicrobial effect of a cyclic peptide Nostophycin isolated from wastewater cyanobacteria, Nostoc calcicola. Curr. Bot. 2021, 12, 94–101. [Google Scholar] [CrossRef]
  165. Moffitt, M.C.; Neilan, B.A. Characterization of the nodularin synthetase gene cluster and proposed theory of the evolution of cyanobacterial hepatotoxins. Appl. Environ. Microbiol. 2004, 70, 6353–6362. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Sivonen, K.; Leikoski, N.; Fewer, D.P.; Jokela, J. Cyanobactins—ribosomal cyclic peptides produced by cyanobacteria. Appl. Microbiol. Biotechnol. 2010, 86, 1213–1225. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. do Amaral, S.C.; Monteiro, P.R.; Neto, J.d.S.P.; Serra, G.M.; Gonçalves, E.C.; Xavier, L.P.; Santos, A.V. Current knowledge on microviridin from cyanobacteria. Mar. Drugs 2021, 19, 17. [Google Scholar] [CrossRef]
  168. Cubillos-Ruiz, A.; Berta-Thompson, J.W.; Becker, J.W.; Van Der Donk, W.A.; Chisholm, S.W. Evolutionary radiation of lanthipeptides in marine cyanobacteria. Proc. Natl. Acad. Sci. USA 2017, 114, E5424–E5433. [Google Scholar] [CrossRef] [Green Version]
  169. Knerr, P.J.; van der Donk, W.A. Discovery, biosynthesis, and engineering of lantipeptides. Annu. Rev. Biochem. 2012, 81, 479–505. [Google Scholar] [CrossRef]
  170. Li, B.; Sher, D.; Kelly, L.; Shi, Y.; Huang, K.; Knerr, P.J.; Joewono, I.; Rusch, D.; Chisholm, S.W.; Van Der Donk, W.A. Catalytic promiscuity in the biosynthesis of cyclic peptide secondary metabolites in planktonic marine cyanobacteria. Proc. Natl. Acad. Sci. USA 2010, 107, 10430–10435. [Google Scholar] [CrossRef] [Green Version]
  171. Ziemert, N.; Ishida, K.; Quillardet, P.; Bouchier, C.; Hertweck, C.; de Marsac, N.T.; Dittmann, E. Microcyclamide biosynthesis in two strains of Microcystis aeruginosa: From structure to genes and vice versa. Appl. Environ. Microbiol. 2008, 74, 1791–1797. [Google Scholar] [CrossRef] [Green Version]
  172. Portmann, C.; Blom, J.F.; Gademann, K.; Jüttner, F. Aerucyclamides A and B: Isolation and synthesis of toxic ribosomal heterocyclic peptides from the cyanobacterium Microcystis aeruginosa PCC 7806. J. Nat. Prod. 2008, 71, 1193–1196. [Google Scholar] [CrossRef] [PubMed]
  173. Portmann, C.; Blom, J.F.; Kaiser, M.; Brun, R.; Jüttner, F.; Gademann, K. Isolation of aerucyclamides C and D and structure revision of microcyclamide 7806A: Heterocyclic ribosomal peptides from Microcystis aeruginosa PCC 7806 and their antiparasite evaluation. J. Nat. Prod. 2008, 71, 1891–1896. [Google Scholar] [CrossRef]
  174. Ogino, J.; Moore, R.E.; Patterson, G.M.L.; Smith, C.D. Dendroamides, new cyclic hexapeptides from a blue-green alga. Multidrug-resistance reversing activity of dendroamide A. J. Nat. Prod. 1996, 59, 581–586. [Google Scholar] [CrossRef] [PubMed]
  175. Sudek, S.; Haygood, M.G.; Youssef, D.T.A.; Schmidt, E.W. Structure of trichamide, a cyclic peptide from the bloom-forming cyanobacterium Trichodesmium erythraeum, predicted from the genome sequence. Appl. Environ. Microbiol. 2006, 72, 4382–4387. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Schmidt, E.W.; Nelson, J.T.; Rasko, D.A.; Sudek, S.; Eisen, J.A.; Haygood, M.G.; Ravel, J. Patellamide A and C biosynthesis by a microcin-like pathway in Prochloron didemni, the cyanobacterial symbiont of Lissoclinum patella. Proc. Natl. Acad. Sci. USA 2005, 102, 7315–7320. [Google Scholar] [CrossRef] [Green Version]
  177. Leikoski, N.; Fewer, D.P.; Jokela, J.; Wahlsten, M.; Rouhiainen, L.; Sivonen, K. Highly diverse cyanobactins in strains of the genus Anabaena. Appl. Environ. Microbiol. 2010, 76, 701–709. [Google Scholar] [CrossRef] [Green Version]
  178. Okino, T.; Matsuda, H.; Murakami, M.; Yamaguchi, K. New microviridins, elastase inhibitors from the blue-green alga Microcystis aeruginosa. Tetrahedron 1995, 51, 10679–10686. [Google Scholar] [CrossRef]
  179. Tang, W.; Van Der Donk, W.A. Structural characterization of four prochlorosins: A novel class of lantipeptides produced by planktonic marine cyanobacteria. Biochemistry 2012, 51, 4271–4279. [Google Scholar] [CrossRef]
  180. Schuler, C.G.; Havig, J.R.; Hamilton, T.L. Hot spring microbial community composition, morphology, and carbon fixation: Implications for interpreting the ancient rock record. Front. Earth Sci. 2017, 5, 97. [Google Scholar] [CrossRef] [Green Version]
  181. Teodoro, G.R.; Ellepola, K.; Seneviratne, C.J.; Koga-Ito, C.Y. Potential use of phenolic acids as anti-Candida agents: A review. Front. Microbiol. 2015, 6, 1420. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Kumar, N.; Goel, N. Phenolic acids: Natural versatile molecules with promising therapeutic applications. Biotechnol. Rep. 2019, 24, e00370. [Google Scholar] [CrossRef] [PubMed]
  183. Singh, D.P.; Prabha, R.; Meena, K.K.; Sharma, L.; Sharma, A.K. Induced accumulation of polyphenolics and flavonoids in cyanobacteria under salt stress protects organisms through enhanced antioxidant activity. Am. J. Plant Sci. 2014, 2014, 43916. [Google Scholar] [CrossRef]
  184. Patipong, T.; Hibino, T.; Waditee-Sirisattha, R.; Kageyama, H. Induction of antioxidative activity and antioxidant molecules in the halotolerant cyanobacterium Halothece sp. PCC 7418 by temperature shift. Nat. Prod. Commun. 2019, 14, 1934578X19865680. [Google Scholar] [CrossRef] [Green Version]
  185. Monroe, M.B.B.; Easley, A.D.; Grant, K.; Fletcher, G.K.; Boyer, C.; Maitland, D.J. Multifunctional shape-memory polymer foams with bio-inspired antimicrobials. ChemPhysChem 2018, 19, 1999–2008. [Google Scholar] [CrossRef]
  186. Li, R.; Narita, R.; Nishimura, H.; Marumoto, S.; Yamamoto, S.P.; Ouda, R.; Yatagai, M.; Fujita, T.; Watanabe, T. Antiviral activity of phenolic derivatives in pyroligneous acid from hardwood, softwood, and bamboo. ACS Sustain. Chem. Eng. 2018, 6, 119–126. [Google Scholar] [CrossRef]
  187. Sun, S.; Kee, H.J.; Ryu, Y.; Choi, S.Y.; Kim, G.R.; Kim, H.-S.; Kee, S.-J.; Jeong, M.H. Gentisic acid prevents the transition from pressure overload-induced cardiac hypertrophy to heart failure. Sci. Rep. 2019, 9, 3018. [Google Scholar] [CrossRef] [Green Version]
  188. Mori, T.; Koyama, N.; Yokoo, T.; Segawa, T.; Maeda, M.; Sawmiller, D.; Tan, J.; Town, T. Gallic acid is a dual α/β-secretase modulator that reverses cognitive impairment and remediates pathology in Alzheimer mice. J. Biol. Chem. 2020, 295, 16251–16266. [Google Scholar] [CrossRef]
  189. Ren, J.; Yang, M.; Xu, F.; Chen, J.; Ma, S. Acceleration of wound healing activity with syringic acid in streptozotocin induced diabetic rats. Life Sci. 2019, 233, 116728. [Google Scholar] [CrossRef]
  190. Park, H.-J.; Cho, J.-H.; Hong, S.-H.; Kim, D.-H.; Jung, H.-Y.; Kang, I.-K.; Cho, Y.-J. Whitening and anti-wrinkle activities of ferulic acid isolated from Tetragonia tetragonioides in B16F10 melanoma and CCD-986sk fibroblast cells. J. Nat. Med. 2018, 72, 127–135. [Google Scholar] [CrossRef] [PubMed]
  191. Monteiro e Silva, S.A.; Calixto, G.M.F.; Cajado, J.; De Carvalho, P.C.A.; Rodero, C.F.; Chorilli, M.; Leonardi, G.R. Gallic acid-loaded gel formulation combats skin oxidative stress: Development, characterization and ex vivo biological assays. Polymers 2017, 9, 391. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Singh, B.; Kumar, A.; Malik, A.K. Flavonoids biosynthesis in plants and its further analysis by capillary electrophoresis. Electrophoresis 2017, 38, 820–832. [Google Scholar] [CrossRef] [PubMed]
  193. Ruiz-Cruz, S.; Chaparro-Hernández, S.; Hernández-Ruiz, K.L.; Cira-Chávez, L.A.; Estrada-Alvarado, M.I.; Ortega, L.E.G.; Mata, M.A.L. Flavonoids: Important biocompounds in food. In Flavonoids: From Biosynthesis to Human Health; Justino, J.G., Ed.; IntechOpen: London, UK, 2017; pp. 353–369. [Google Scholar]
  194. Wang, T.-Y.; Li, Q.; Bi, K.-S. Bioactive flavonoids in medicinal plants: Structure, activity and biological fate. Asian J. Pharm. Sci. 2018, 13, 12–23. [Google Scholar] [CrossRef] [PubMed]
  195. Trabelsi, L.; Mnari, A.; Abdel-Daim, M.M.; Abid-Essafi, S.; Aleya, L. Therapeutic properties in Tunisian hot springs: First evidence of phenolic compounds in the cyanobacterium Leptolyngbya sp. biomass, capsular polysaccharides and releasing polysaccharides. BMC Complementary Altern. Med. 2016, 16, 515. [Google Scholar] [CrossRef] [Green Version]
  196. Żyszka, B.; Anioł, M.; Lipok, J. Modulation of the growth and metabolic response of cyanobacteria by the multifaceted activity of naringenin. PLoS ONE 2017, 12, e0177631. [Google Scholar] [CrossRef]
  197. Jerez-Martel, I.; García-Poza, S.; Rodríguez-Martel, G.; Rico, M.; Afonso-Olivares, C.; Gómez-Pinchetti, J.L. Phenolic profile and antioxidant activity of crude extracts from microalgae and cyanobacteria strains. J. Food Qual. 2017, 2017, 2924508. [Google Scholar] [CrossRef] [Green Version]
  198. Mallick, N.; Mohn, F.H. Reactive oxygen species: Response of algal cells. J. Plant Physiol. 2000, 157, 183–193. [Google Scholar] [CrossRef]
  199. Hernández-Aquino, E.; Muriel, P. Beneficial effects of naringenin in liver diseases: Molecular mechanisms. World J. Gastroenterol. 2018, 24, 1679. [Google Scholar] [CrossRef]
  200. Sugumar, M.; Sevanan, M.; Sekar, S. Neuroprotective effect of naringenin against MPTP-induced oxidative stress. Int. J. Neurosci. 2019, 129, 534–539. [Google Scholar] [CrossRef]
  201. Choi, J.; Lee, D.-H.; Jang, H.; Park, S.-Y.; Seol, J.-W. Naringenin exerts anticancer effects by inducing tumor cell death and inhibiting angiogenesis in malignant melanoma. Int. J. Med. Sci. 2020, 17, 3049. [Google Scholar] [CrossRef]
  202. Mulvihill, E.E.; Assini, J.M.; Sutherland, B.G.; DiMattia, A.S.; Khami, M.; Koppes, J.B.; Sawyez, C.G.; Whitman, S.C.; Huff, M.W. Naringenin decreases progression of atherosclerosis by improving dyslipidemia in high-fat–fed low-density lipoprotein receptor–null mice. Atertio. Thromb. Vasc. Biol. 2010, 30, 742–748. [Google Scholar] [CrossRef] [Green Version]
  203. Assini, J.M.; Mulvihill, E.E.; Huff, M.W. Citrus flavonoids and lipid metabolism. Curr. Opin. Lipidol. 2013, 24, 34–40. [Google Scholar] [CrossRef] [PubMed]
  204. Asensi-Fabado, M.A.; Munné-Bosch, S. Vitamins in plants: Occurrence, biosynthesis and antioxidant function. Trends Plant Sci. 2010, 15, 582–592. [Google Scholar] [CrossRef] [PubMed]
  205. Del Mondo, A.; Smerilli, A.; Sané, E.; Sansone, C.; Brunet, C. Challenging microalgal vitamins for human health. Microb. Cell Factories 2020, 19, 201. [Google Scholar] [CrossRef]
  206. Aaronson, S.; Dhawale, S.W.; Patni, N.J.; DeAngelis, B.; Frank, O.; Baker, H. The cell content and secretion of water-soluble vitamins by several freshwater algae. Arch. Microbiol. 1977, 112, 57–59. [Google Scholar] [CrossRef]
  207. Santiago-Morales, I.S.; Trujillo-Valle, L.; Márquez-Rocha, F.J.; Hernández, J.F.L. Tocopherols, phycocyanin and superoxide dismutase from microalgae: As potential food antioxidants. Appl. Food Biotechnol. 2018, 5, 19–27. [Google Scholar]
  208. Sylvander, P.; Häubner, N.; Snoeijs, P. The thiamine content of phytoplankton cells is affected by abiotic stress and growth rate. Microb. Ecol. 2013, 65, 566–577. [Google Scholar] [CrossRef] [PubMed]
  209. Helliwell, K.E.; Lawrence, A.D.; Holzer, A.; Kudahl, U.J.; Sasso, S.; Kräutler, B.; Scanlan, D.J.; Warren, M.J.; Smith, A.G. Cyanobacteria and eukaryotic algae use different chemical variants of vitamin B12. Curr. Biol. 2016, 26, 999–1008. [Google Scholar] [CrossRef] [Green Version]
  210. Edelmann, M.; Aalto, S.; Chamlagain, B.; Kariluoto, S.; Piironen, V. Riboflavin, niacin, folate and vitamin B12 in commercial microalgae powders. J. Food Compos. Anal. 2019, 82, 103226. [Google Scholar] [CrossRef]
  211. Ljubic, A.; Jacobsen, C.; Holdt, S.L.; Jakobsen, J. Microalgae Nannochloropsis oceanica as a future new natural source of vitamin D3. Food Chem. 2020, 320, 126627. [Google Scholar] [CrossRef] [PubMed]
  212. Backasch, N.; Schulz-Friedrich, R.; Appel, J. Influences on tocopherol biosynthesis in the cyanobacterium Synechocystis sp. PCC 6803. J. Plant Physiol. 2005, 162, 758–766. [Google Scholar] [CrossRef] [PubMed]
  213. Mudimu, O.; Koopmann, I.K.; Rybalka, N.; Friedl, T.; Schulz, R.; Bilger, W. Screening of microalgae and cyanobacteria strains for α-tocopherol content at different growth phases and the influence of nitrate reduction on α-tocopherol production. J. Appl. Phycol. 2017, 29, 2867–2875. [Google Scholar] [CrossRef]
  214. Krieger-Liszkay, A.; Trebst, A. Tocopherol is the scavenger of singlet oxygen produced by the triplet states of chlorophyll in the PSII reaction centre. J. Exp. Bot. 2006, 57, 1677–1684. [Google Scholar] [CrossRef]
  215. Goiris, K.; Van Colen, W.; Wilches, I.; León-Tamariz, F.; De Cooman, L.; Muylaert, K. Impact of nutrient stress on antioxidant production in three species of microalgae. Algal Res. 2015, 7, 51–57. [Google Scholar] [CrossRef]
  216. Hamed, S.M.; Selim, S.; Klöck, G.; AbdElgawad, H. Sensitivity of two green microalgae to copper stress: Growth, oxidative and antioxidants analyses. Ecotoxicol. Environ. Saf. 2017, 144, 19–25. [Google Scholar] [CrossRef]
  217. Strejckova, A.; Dvorak, M.; Klejdus, B.; Krystofova, O.; Hedbavny, J.; Adam, V.; Huska, D. The strong reaction of simple phenolic acids during oxidative stress caused by nickel, cadmium and copper in the microalga Scenedesmus quadricauda. New Biotechnol. 2019, 48, 66–75. [Google Scholar] [CrossRef]
  218. Tarento, T.D.C.; McClure, D.D.; Vasiljevski, E.; Schindeler, A.; Dehghani, F.; Kavanagh, J.M. Microalgae as a source of vitamin K1. Algal Res. 2018, 36, 77–87. [Google Scholar] [CrossRef]
  219. Collins, M.D.; Jones, D. Distribution of isoprenoid quinone structural types in bacteria and their taxonomic implication. Microbiol. Rev. 1981, 45, 316–354. [Google Scholar] [CrossRef]
  220. Sakuragi, Y.; Bryant, D.A. Genetic manipulation of quinone biosynthesis in cyanobacteria. In Photosystem I: The Light-Driven Plastocyanin: Ferredoxin Oxidoreductase; Golbeck, J.H., Ed.; Springer: Dordrecht, The Netherlands, 2006; pp. 205–222. [Google Scholar]
  221. Joliot, P.; Joliot, A.; Johnson, G. Cyclic Electron Transfer around Photosystem I; Springer: Berlin/Heidelberg, Germany, 2006; pp. 639–656. [Google Scholar]
  222. Widhalm, J.R.; van Oostende, C.; Furt, F.; Basset, G.J.C. A dedicated thioesterase of the Hotdog-fold family is required for the biosynthesis of the naphthoquinone ring of vitamin K1. Proc. Natl. Acad. Sci. USA 2009, 106, 5599–5603. [Google Scholar] [CrossRef] [Green Version]
  223. Mimuro, M.; Tsuchiya, T.; Inoue, H.; Sakuragi, Y.; Itoh, Y.; Gotoh, T.; Miyashita, H.; Bryant, D.A.; Kobayashi, M. The secondary electron acceptor of photosystem I in Gloeobacter violaceus PCC 7421 is menaquinone-4 that is synthesized by a unique but unknown pathway. FEBS Lett. 2005, 579, 3493–3496. [Google Scholar] [CrossRef] [Green Version]
  224. Sakuragi, Y.; Zybailov, B.; Shen, G.; Bryant, D.A.; Golbeck, J.H.; Diner, B.A.; Karygina, I.; Pushkar, Y.; Stehlik, D. Recruitment of a foreign quinone into the A1 site of photosystem I: Characterization of a menB rubA double deletion mutant in Synechococcus sp. PCC 7002 devoid of FX, FA, and FB and containing plastoquinone or exchanged 9, 10-anthraquinone. J. Biol. Chem. 2005, 280, 12371–12381. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Langan, R.C.; Goodbred, A.J. Vitamin B12 deficiency: Recognition and management. Am. Fam. Physician 2017, 96, 384–389. [Google Scholar] [PubMed]
  226. Bordelon, P.; Ghetu, M.V.; Langan, R.C. Recognition and management of vitamin D deficiency. Am. Fam. Physician 2009, 80, 841–846. [Google Scholar] [PubMed]
  227. Maxfield, L.; Crane, J.S. Vitamin C Deficiency (Scurvy); StatPearls; StatPearls Publishing: Treasure Island, FL, USA, 2020. [Google Scholar]
  228. Golriz, F.; Donnelly, L.F.; Devaraj, S.; Krishnamurthy, R. Modern American scurvy—experience with vitamin C deficiency at a large children’s hospital. Pediatric Radiol. 2017, 47, 214–220. [Google Scholar] [CrossRef]
  229. Eden, R.E.; Coviello, J.M. Vitamin K Deficiency; StatPearls; StatPearls Publishing: Treasure Island, FL, USA, 2019. [Google Scholar]
  230. Kishimoto, N.; Hayashi, T.; Mizobuchi, K.; Kubota, M.; Nakano, T. Vitamin A deficiency after prolonged intake of an unbalanced diet in a Japanese hemodialysis patient. Doc. Ophthalmol. 2021, 143, 85–91. [Google Scholar] [CrossRef]
  231. Cordeiro, A.; Bento, C.; de Matos, A.C.; Ramalho, A. Vitamin A deficiency is associated with body mass index and body adiposity in women with recommended intake of vitamin A. Nutr. Hosp. Organo Of. Soc. Española Nutr. Parenter. Enter. 2018, 35, 1072–1078. [Google Scholar] [CrossRef] [Green Version]
  232. Ekeuku, S.O.; Chong, P.N.; Chan, H.K.; Mohamed, N.; Froemming, G.R.A.; Okechukwu, P.N. Spirulina supplementation improves bone structural strength and stiffness in streptozocin-induced diabetic rats. J. Tradit. Complementary Med. 2021, in press. [Google Scholar] [CrossRef]
  233. Anantharajappa, K.; Dharmesh, S.M.; Ravi, S. Gastro-protective potentials of Spirulina: Role of vitamin B 12. J. Food Sci. Technol. 2020, 57, 745–753. [Google Scholar] [CrossRef]
  234. Soudy, I.D.; Minet-Quinard, R.; Mahamat, A.D.; Ngoua, H.F.; Izzedine, A.A.; Tidjani, A.; Ngo Bum, E.; Lambert, C.; Pereira, B.; Desjeux, J.-F. Vitamin A status in healthy women eating traditionally prepared spirulina (Dihé) in the Chad Lake area. PLoS ONE 2018, 13, e0191887. [Google Scholar] [CrossRef] [Green Version]
  235. Brilisauer, K.; Rapp, J.; Rath, P.; Schöllhorn, A.; Bleul, L.; Weiß, E.; Stahl, M.; Grond, S.; Forchhammer, K. Cyanobacterial antimetabolite 7-deoxy-sedoheptulose blocks the shikimate pathway to inhibit the growth of prototrophic organisms. Nat. Commun. 2019, 10, 545. [Google Scholar] [CrossRef] [PubMed]
  236. Rapp, J.; Rath, P.; Kilian, J.; Brilisauer, K.; Grond, S.; Forchhammer, K. A bioactive molecule made by unusual salvage of radical SAM enzyme byproduct 5-deoxyadenosine blurs the boundary of primary and secondary metabolism. J. Biol. Chem. 2021, 296, 100621. [Google Scholar] [CrossRef] [PubMed]
  237. Vergou, Y.; Touraki, M.; Paraskevopoulou, A.; Triantis, T.M.; Hiskia, A.; Gkelis, S. β-Ν-Methylamino-L-alanine interferes with nitrogen assimilation in the cyanobacterium, non-BMAA producer, Synechococcus sp. TAU-MAC 0499. Toxicon 2020, 185, 147–155. [Google Scholar] [CrossRef] [PubMed]
  238. Downing, S.; van de Venter, M.; Downing, T.G. The effect of exogenous β-N-methylamino-L-alanine on the growth of Synechocystis PCC 6803. Microb. Ecol. 2012, 63, 149–156. [Google Scholar] [CrossRef] [PubMed]
  239. Scott, L.; Downing, S.; Phelan, R.; Downing, T. Environmental modulation of microcystin and β-N-methylamino-l-alanine as a function of nitrogen availability. Toxicon 2014, 87, 1–5. [Google Scholar] [CrossRef]
  240. Berntzon, L.; Erasmie, S.; Celepli, N.; Eriksson, J.; Rasmussen, U.; Bergman, B. BMAA inhibits nitrogen fixation in the cyanobacterium Nostoc sp. PCC 7120. Mar. Drugs 2013, 11, 3091–3108. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  241. Downing, S.; Banack, S.; Metcalf, J.; Cox, P.; Downing, T. Nitrogen starvation of cyanobacteria results in the production of β-N-methylamino-L-alanine. Toxicon 2011, 58, 187–194. [Google Scholar] [CrossRef] [PubMed]
  242. Downing, S.; Downing, T.G. The metabolism of the non-proteinogenic amino acid β-N-methylamino-L-alanine (BMAA) in the cyanobacterium Synechocystis PCC 6803. Toxicon 2016, 115, 41–48. [Google Scholar] [CrossRef]
  243. Forchhammer, K.; Schwarz, R. Nitrogen chlorosis in unicellular cyanobacteria–a developmental program for surviving nitrogen deprivation. Environ. Microbiol. 2019, 21, 1173–1184. [Google Scholar] [CrossRef] [Green Version]
  244. Yan, B.; Liu, Z.; Huang, R.; Xu, Y.; Liu, D.; Wang, W.; Zhao, Z.; Cui, F.; Shi, W. Impact factors on the production of β-methylamino-L-alanine (BMAA) by cyanobacteria. Chemosphere 2020, 243, 125355. [Google Scholar] [CrossRef]
  245. Cox, P.A.; Banack, S.A.; Murch, S.J.; Rasmussen, U.; Tien, G.; Bidigare, R.R.; Metcalf, J.S.; Morrison, L.F.; Codd, G.A.; Bergman, B. Diverse taxa of cyanobacteria produce β-N-methylamino-L-alanine, a neurotoxic amino acid. Proc. Natl. Acad. Sci. USA 2005, 102, 5074–5078. [Google Scholar] [CrossRef] [Green Version]
  246. Cox, P.A.; Banack, S.A.; Murch, S.J. Biomagnification of cyanobacterial neurotoxins and neurodegenerative disease among the Chamorro people of Guam. Proc. Natl. Acad. Sci. USA 2003, 100, 13380–13383. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Classification of indole alkaloids from Stigonematales order of cyanobacteria [81].
Figure 1. Classification of indole alkaloids from Stigonematales order of cyanobacteria [81].
Biology 10 01061 g001
Figure 2. Classes of terpenoids based on their isoprene units [93].
Figure 2. Classes of terpenoids based on their isoprene units [93].
Biology 10 01061 g002
Table 1. Different classes of cyanobacterial species based on their physiological characteristics.
Table 1. Different classes of cyanobacterial species based on their physiological characteristics.
ClassHabitatsCyanobacteriaReferences
AlkaliphilesHypersaline swamps, alkaline-saline lake or ponds, hot spring, alkaline hot spring, alkaline-saline volcanic lake, soda desertsMicrocoleus sp., Pleurocapsa sp., Synechococcus sp., Cyanobacterium sp., Spirulina subsalsa, Spirulina platensis, Spirulina maxima, and Arthrospira sp.[1,2,3,4,5,6,7,8,9,10]
AcidophilesSulfuric pools and acid mine drainageCyanobacteria cannot survive under this condition.[11,12,13]
EndolithicRocks, granites and quartzites in desert, freshwaterChroococcidiopsis-like cyanobacterium[14]
Halophilic Hypersaline lakes, coastal hypersaline lagoons, saline
springs, salt flats and ponds
Synechococcus sp., Leptolyngbya sp, Nodosilinea sp., and Geitlerinema sp[15]
OligotrophicsCoastal regions of marine and freshwaterDolichospermum lemmermanii[16,17]
Psychrophilic Alpines and polar regionsNostoc sp., Leptolyngba sp., Oscillatoria sp. and Phormidium sp.[18,19,20]
Thermophilic Thermal springs and
soil crusts of deserted area
Synechococcus sp., Thermosynechococcus vulcanus, Leptolyngbya sp., Thermosynechococcus elongatus, and
Phormidium sp.
[21,22,23,24,25]
RadiophilesMarine, freshwater and desert Synechocystis sp., Chroococcus minutus, Leptolyngbya sp., Trichodesmium and Crocosphaera[26,27,28]
Table 2. List of cyanobacteria producing diverse class of indole alkaloids.
Table 2. List of cyanobacteria producing diverse class of indole alkaloids.
Cyanobacteria speciesHabitatCompoundsBioactivitiesReferences
Hapalosiphon sp. CBT1235TerrestrialHapalindolesInhibit T Cell Proliferation[67]
Hapalosiphon fontinalisSoilHapalindolesAntibacterial and antimycotic[68]
Hapalosiphon fontinalisSoilHapalindolesAntialgal[69]
Westiellopsis sp. (SAG 20.93) and Fischerella muscicola (UTEX LB1829)Freshwater and terrestrialHapalindolesAntibacterial[70]
Fischerella ambigua UTEX1903TerrestrialAmbiguineUnknown[71]
Hapalosiphon welwitschii UTEX B1830FreshwaterWelwitindolinoneUnknown[72]
Westiella intricata UH strain HT-29-1FreshwaterWelwitindolinoneUnknown[73]
Fischerella ambigua (UTEX 1903),Westiellopsis prolifica and Hapalosiphon hibernicus BZ-3-1TerrestrialAmbiguine IsonitrilesFungicidal[74]
Fischerella muscicolaTerrestrialFischerindoleAntifungal[75]
Fischerella ambigua (UTEX 1903) TerrestrialFischambiguines and ambiguinesAntibacterial[76]
Fischerella sp. TerrestrialWelwitindolinonesMulti-drug resistance reversing activity[77]
Hapalosiphon welwitschia and Westiella intricataSoilWelwitindolinonesMulti-drug resistance re versing activity and insecticidal activity[78]
Table 3. Bioactivities of non-ribosomal peptides (NRPS) and polyketides (PKS) produced by cyanobacteria.
Table 3. Bioactivities of non-ribosomal peptides (NRPS) and polyketides (PKS) produced by cyanobacteria.
Cyanobacteria speciesHabitat CompoundsBioactivitiesReferences
Microcystis aeruginosaFreshwaterMicrocystinsInhibit eukaryotic types 1 and 2A phosphatases, cytoskeletal collapse, massive hepatic bleeding, potential tumor promoters and carcinogens[147,148,149]
Planktothrix agardhii NIVA-CYA 126FreshwaterAeruginosin Inhibit serine proteases[150]
Cylindrospermopsis raciborskii, Aphanizomenon ovalisporum and Aphanizomenon flos-aquaeFreshwaterCylindrospermopsin Cytotoxic, neurotoxic effects and carcinogen [151]
Anabaena sp. 90FreshwaterAnabaenopeptin Inhibit proteases[152]
Lyngbya bouilloniiMarineApratoxin Reversible inhibition of several cancer-associated receptors[153,154]
Lyngbya majusculaMarineLyngbyatoxin Potent skin irritant[155]
Lyngbya majuscule JHBMarineHectochlorin Antifungal and anticancer activity [156]
Lyngbya majuscule 19LMarineBarbamide Anti-molluscidal [157]
Lyngbya majuscule 19LMarineCuracin A Antiproliferative and cytotoxic activities[158]
Lyngbya majuscule JHBMarineJamaicamide Block sodium-channel [159]
Nostoc sp. GSV 224TerrestrialNostopeptolide No cytotoxic, antifungal and inhibit protease activities[160]
Nostoc sp. ATCC 53789TerrestrialNostocyclopeptide Antitoxin activity [161]
Nostoc sp. ATCC 53789TerrestrialCryptophycinsTubulin-destabilizing compound [162]
Cylindrospermum alatosporum CCALA 988TerrestrialPuwainaphycinsCytotoxic [163]
Nostoc calcicolaWastewaterNostophycinAntibacterial and antifungal [164]
Nodularia spumigena NSOR10FreshwaterNodularinInhibits phosphatase type 1 and 2A, cytoskeletal collapse, massive hepatic bleeding, potential tumor promoters and carcinogens[165]
Table 4. Bioactivities of ribosomal peptides (RPs) produced by cyanobacteria.
Table 4. Bioactivities of ribosomal peptides (RPs) produced by cyanobacteria.
Cyanobacteria SpeciesHabitat CompoundsBioactivitiesReferences
Cyanobactin
Microcystis aeruginosaFreshwaterAerucyclamide A, B, C and DCytotoxic and antimalarial[171,172,173]
Stigonema dendroideumTerrestrialDendroamide AMultidrug-resistance reversing activity[174]
Trichodesmium erythraeumMarineTrichamideNo biological effects found [175]
Prochloron didemnid (symbioant)MarinePatellamide A and CCytotoxic [176]
Anabaena sp. 90FreshwaterAnacyclamideCytotoxic [177]
Microcystis aeruginosa PCC 7806FreshwaterMicrocyclamideNo biological effects found[171]
Microviridin
Microcystis aeruginosa NIES-298 FreshwaterMicroviridios B and CInhibits elastase [178]
Lantipeptides
Prochlorococcus MIT9313MarineProchlorosins Bacteriocidal and act as signaling molecules [168,170,179]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Nandagopal, P.; Steven, A.N.; Chan, L.-W.; Rahmat, Z.; Jamaluddin, H.; Mohd Noh, N.I. Bioactive Metabolites Produced by Cyanobacteria for Growth Adaptation and Their Pharmacological Properties. Biology 2021, 10, 1061. https://doi.org/10.3390/biology10101061

AMA Style

Nandagopal P, Steven AN, Chan L-W, Rahmat Z, Jamaluddin H, Mohd Noh NI. Bioactive Metabolites Produced by Cyanobacteria for Growth Adaptation and Their Pharmacological Properties. Biology. 2021; 10(10):1061. https://doi.org/10.3390/biology10101061

Chicago/Turabian Style

Nandagopal, Pavitra, Anthony Nyangson Steven, Liong-Wai Chan, Zaidah Rahmat, Haryati Jamaluddin, and Nur Izzati Mohd Noh. 2021. "Bioactive Metabolites Produced by Cyanobacteria for Growth Adaptation and Their Pharmacological Properties" Biology 10, no. 10: 1061. https://doi.org/10.3390/biology10101061

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