Next Article in Journal
Natural Occurrence of Regulated and Emerging Mycotoxins in Wheat Grains and Assessment of the Risks from Dietary Mycotoxins Exposure in China
Next Article in Special Issue
Regulation of Clostridial Toxin Gene Expression: A Pasteurian Tradition
Previous Article in Journal
Induced Expression of the Acinetobacter sp. Oxa Gene in Lactobacillus acidophilus and Its Increased ZEN Degradation Stability by Immobilization
Previous Article in Special Issue
From Bacterial Toxin to Therapeutic Agent: The Unexpected Fate of Mycolactone
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Cyanotoxins and Other Bioactive Compounds from the Pasteur Cultures of Cyanobacteria (PCC)

Institut Pasteur, Université Paris Cité, Collection of Cyanobacteria, 75015 Paris, France
*
Author to whom correspondence should be addressed.
Toxins 2023, 15(6), 388; https://doi.org/10.3390/toxins15060388
Submission received: 12 May 2023 / Revised: 5 June 2023 / Accepted: 7 June 2023 / Published: 9 June 2023

Abstract

:
In tribute to the bicentenary of the birth of Louis Pasteur, this report focuses on cyanotoxins, other natural products and bioactive compounds of cyanobacteria, a phylum of Gram-negative bacteria capable of carrying out oxygenic photosynthesis. These microbes have contributed to changes in the geochemistry and the biology of Earth as we know it today. Furthermore, some bloom-forming cyanobacterial species are also well known for their capacity to produce cyanotoxins. This phylum is preserved in live cultures of pure, monoclonal strains in the Pasteur Cultures of Cyanobacteria (PCC) collection. The collection has been used to classify organisms within the Cyanobacteria of the bacterial kingdom and to investigate several characteristics of these bacteria, such as their ultrastructure, gas vacuoles and complementary chromatic adaptation. Thanks to the ease of obtaining genetic and further genomic sequences, the diversity of the PCC strains has made it possible to reveal some main cyanotoxins and to highlight several genetic loci dedicated to completely unknown natural products. It is the multidisciplinary collaboration of microbiologists, biochemists and chemists and the use of the pure strains of this collection that has allowed the study of several biosynthetic pathways from genetic origins to the structures of natural products and, eventually, their bioactivity.
Key Contribution: The Pasteur Cultures of Cyanobacteria collection has helped scientists for decades to reveal the potential of cyanotoxins and natural products.

1. Introduction

Cyanobacteria represent a monophyletic lineage of Gram-negative oxygenic photosynthetic bacteria [1]. This phylum has inhabited the Earth for 2.8 billion years, contributing to changes in the geochemistry and the biology of the globe [2]. They are ubiquitous and found in diverse ecological niches, from aquatic ecosystems such as lakes, rivers and oceans to deserts, Polar Regions, caves and even in symbiosis with other organisms, such as fungi, to form lichens, for example. Cyanobacteria are also the ancestors of chloroplasts that are found in most plants and algae, while recurrent examples of various cyanobacterial morphotypes are found associated with the leaves or in the roots of plants. Although cyanobacteria are famous for the oxygenation of the Earth [2,3], they are also sadly famous for the toxic blooms they can massively develop in marine and fresh waters all around the globe [4]. The field of cyanotoxins and other cyanobacterial bioactive compounds has greatly extended over the last century, notably through chemical analysis and compound structure elucidation, and due to the availability of enough material for toxicological studies. Despite 1100 cyanobacterial natural products (NPs) discovered by these approaches [5], we still know little about them. In this review, we will first introduce the diversity of the cyanobacterial phylum; secondly, we will present the repository dedicated to this phylum at the Institut Pasteur, and finally, we will discuss how to exploit the cyanobacterial biobank to reveal novel compounds, corresponding pathways, new enzymes and even intriguing chemistry. We will end this overview with examples of bioactivity studies on cyanotoxins and NPs.

2. The Phylum Cyanobacteria, Diversity in Terms of Morphology and Genome

The phylum of Cyanobacteria containing all bacteria capable of performing oxygenic photosynthesis [1] presents a wide breadth of habitats and, thus, of ecology, physiology and morphology. Their most simple morphotypes can be unicellular, as single cells, e.g., Prochlorococcus, Chisolm et al. (1992), that colonize the oceans [6], or as colonies of single cells embedded in mucilage, e.g., Microcystis, Kützing ex Lemmermann (1907), that form toxic blooms in lakes [7]. The colonies of unicellular cyanobacteria can be tightly organized in one layer of arranged cells of Merispomedia, Meyen (1839), or can appear as cell aggregations of various sizes surrounded by mucilaginous envelopes such as the tiny Chroococus, Nägeli (1849), and the large Gloeocapsa, Kützing (1843). More complex morphotypes of single cells in a colony are the Pleurocapsa, Thuret in Hauck (1885), and other baeocytous cyanobacteria that proliferate in desert environments. These cyanobacteria are extremely resistant to desiccation; containing baeocytes that will revive a novel colony when the environmental conditions are more favorable for the growth of these bacteria. Filamentous cyanobacteria also have a more organized cell division which is perpendicular to the growing axis of the filament, such as the solitary trichomes found in toxic freshwater blooms such as Planktothrix, Gaget et al. (2015) or the several trichomes embedded in a common sheath such as Hydrocoleum, Kützing ex Gomont (1892). The filamentous heterocystous cyanobacteria, exemplified by Byssus flos-aquae, Linnaeus (1753), are even more complex; this taxa is currently invalid and has been replaced by Aphanizomenon flos-aquae, Ralfs ex Bornet and Flahault (1886). This morphotype bears differentiated cells such as akinetes, used to revive a novel filament after harsh conditions, and heterocytes, used to fix atmospheric nitrogen. Aphanizomenon is also a toxic bloom-forming cyanobacterium found in lakes and rivers. Finally, the most complex morphotypes, and often larger than other cyanobacterial morphologies, are the filamentous cyanobacteria with differentiated cells (heterocytes and akinetes) and true ramifications, surrounded or not by mucilage and sheaths. The best ubiquitous representative of this cyanobacterial morphotype is Fischerella (Bornet and Flahault) Gomont, 1895. This summarizes at a glance the morphological diversity of Cyanobacteria that can be seen with the naked eye or using a light microscope. A glimpse into the morphologies encountered in the cyanobacterial phylum can be found in Figure 1; a more detailed view of the morphological diversity, through a botanical approach with camera lucida drawings, can be found in the book series Süsswasserflora von Mitteleuropa on Cyanoprokaryota [8,9,10]. Moreover, a detailed view of the morphological diversity through a bacteriological approach with photography of the various morphologies and electron microscopy photographs of the ultrastructural arrangement within the cyanobacterial cells can be found in the Bergey’s Manual of Systematic Bacteriology [11].
In response to the morphological diversity within the phylum of Cyanobacteria, their genomes are also extremely diverse. The publicly available genomes present sizes ranging from 1.5 to 15 Mb and a GC content from 30 to 68 %. The smallest genome size corresponds to the picocyanobacterial genus of Prochlorococcus, while the largest ones were found in filamentous cyanobacteria with differentiated cells and ramifications. On the contrary, the highest and lowest GC contents were reported in the genus Synechococcus, Nägeli (1849), while the filamentous cyanobacteria with differentiated cells and with or without branching presented a GC content centered on approximately 42%. The comparison of the morphologies and ultrastructural data with genomic sequence data did not show any solid groupings corresponding to the different morphotypes; however, a correlation between the ultrastructures and genes coding for cellular inclusions was identified [12].

3. Collection of Cyanobacteria and the Pasteur Cultures of Cyanobacteria (PCC)

Several collections of cyanobacteria have been constituted in universities and institutions around the world, as cyanobacterial blooms occur everywhere on Earth. In addition, a few major collections often conserved these bacteria along with algal isolates, protists and other bacteria (for example, NIES-MCC in Japan; UTEX and ATCC in the USA; GCC, NFMC and NCCS in India; CCAP in Scotland, CCALA in Czech Republic; BCCM-ULC in Belgium; NIVA in Norway; UHCC in Finland; SAG in Germany; PCC, PMC, TCC and RCC in France). The collection of cyanobacteria currently present at the Institut Pasteur arrived from the University of California, Berkeley, with Pr. Roger Stanier (1916–1982), who worked on the classification of the so-called blue-green algae at that time and included them among the bacterial classification as the Cyanobacteria [13]. Pr. Stanier led a research team named Microbial Physiology Unit, which focused on various subjects such as pigments, photosynthesis, gas vacuoles, fatty acids and taxonomy, upon of these 150 pure cyanobacterial strains. From the same laboratory, the isolation of a purple cyanobacterium without thylakoids, internal membranes on which the light-harvesting complexes sit, led to the description of the first representative of the basal clade of the phylum Cyanobacteria, Gloeobacter violaceaus, Rippka, Waterbury and Cohen-Bazire (1974), strain PCC 7421 [14]. After a short period (1982–1988) during which Dr. Germaine Cohen-Bazire Stanier (1920–2001) was heading this team and working on the ultrastructure of cyanobacteria [15,16], Pr. Nicole Tandeau de Marsac (1944–2020) was named head of a novel team called the Unit of Cyanobacteria from 1988–2009. This team maintained the cyanobacterial collection of the Institut Pasteur, which had greatly expanded since its arrival in France through the effort of several researchers and visitors. She and collaborators worked extensively on a few model PCC strains to thoroughly investigate the phycobilisome and the photoregulation, the cellular differentiation of hormogonium, the gas vesicle genes, and the phosphorylation of the signal transducer PII [17,18,19,20,21]. Pr. Tandeau de Marsac also revealed the complementary chromatic adaptation [22,23,24]. Towards the end of her research career, she majorly focused on the toxic cyanobacterial isolate Microcystis aeruginosa, PCC 7806 [7,25,26,27]. In July 2009, Dr. Muriel Gugger was appointed head of the Collection of Cyanobacteria with the mission to maintain, distribute and valorize the Pasteur Cultures of Cyanobacteria collection. The PCC collection is also used as a reference for this phylum and to exemplify the bacterial classification in Bergey’s Manual [11].
Since Stanier’s time, several collaborators and visitors cooperated with Rosemarie Rippka (1944–today), the curator of the PCC collection until June 2009, to carry out research [28,29,30] and, in the meantime, bred up to 800 axenic living PCC strains isolated from all over the world (Figure 2). From the 150 strains that originally arrived with Stanier, 102 are still maintained at the PCC, with the emblematic strain Synechococcus elongatus, PCC 6301, isolated in 1956 from the USA, rendered monoclonal and axenic by 1963 and maintained alive since then as well as entered in cryopreservation stocks several times as all other PCC strains [31]. In 1973 and 1989, strong efforts generated about 100 cyanobacterial monoclonal purified isolates incorporated in the collection, but over time, several isolates were lost or stopped growing. The purification of cyanobacteria is a long delicate process that can take several years, as exemplified by the two years required to obtain an axenic monoclonal culture of Prochlorococcus marinus, PCC 9511 [32], which vanished in 2013 due to the replacement of the incubator in which it was maintained. The last strains incorporated in the collection also belong to the genus Synechococcus; two come from India and are closely related to the PCC 6301 strain but represent another species, and one comes from Singapore described as a transformable strain with a fast-growing capacity under high light [31,33,34,35]. Since 2006, the PCC collection was integrated into the Biological Resource Center of the Institut Pasteur (CRBIP) along with the Collection of Bacteria of the Institut Pasteur (CIP) and the fungal collection (former CMIP) and joined later under the same umbrella of biobank and quality management by the Collection Nationale de Cultures de Microorganismes (CNCM) and the Integrated Collections for Adaptive Research in Biomedicine (ICAReB-Biobank). Today, the PCC collection contains around 800 monoclonal axenic cultures, 600 maintained alive in liquid or solid cultures, and all of them cryopreserved. The maintenance of the collection takes up to 4500 transfers, along with 4500 purity tests per year. While the PCC collection represents a breadth of the cyanobacterial phylum, the study of their genomes of diverse PCC strains revealed a plethora of gene clusters for NPs [3,36].

4. Natural Products of Cyanobacteria—From Toxins to Novel Compounds

In the 1870s, a report documented the occurrence of cattle poisonings from Australian lakes [37]. Over 100 years later, dog deaths became a vivid subject connected to cyanotoxins from cyanobacterial developments on the surface or benthos of water ecosystems in North America and Europe, as well as in South Africa and New Zealand [38,39,40,41,42,43,44]. More recently, a human fatality during a renal dialysis treatment in Brazil and a fatality in bald eagles in the southeastern USA were demonstrated to be due to freshwater cyanobacterial occurrences [45,46]. Several wildlife intoxications or deaths have also been associated with cyanotoxins, for example, with flamingos and most probably with elephants in Africa and with fishes in Canada [47,48,49]. The above examples clearly demonstrate that for the last 145 years, the recurrent problem of cyanobacterial blooms in fresh and marine water bodies has presented a threat to animals and humans. Moreover, the cyanotoxins released by bloom-forming cyanobacteria into water bodies used for drinking water create a global public health issue. The World Health Organization has developed guidance values for the most common cyanotoxins present in recreational water and drinking water [50]. In the USA, the simultaneous incident of animal and human disease around one lake with blooms of toxic cyanobacteria has been documented and has called for the development of a proactive relationship between the healthcare system and veterinarians to protect human health [51,52]. In the field of cyanotoxins, the monoclonal and axenic PCC strains have been useful since 1988. For example, the Microcystis aeruginosa PCC 7820 was used to monitor the hepatotoxic effect of microcystin-LR on mice and rat liver damages and pulmonary emboli leading to acute toxicities and death [53]. In particular, two strains have made it possible to discover the genetic bases of three cyanotoxins: Microcystis aeruginosa PCC 7806 for the discovery of the microcystin biosynthetic gene cluster, and Kamptonema sp. PCC 6506 to reveal anatoxin-a and related compounds as well as cylindrospermopsins [54,55,56]. Indeed, a recent bibliographic survey (27th April 2023, in Pubmed: Cyanobacteria AND PCC AND Toxin) reveals about 150 publications dedicated to cyanotoxin discovery, cyanotoxin effects (larvicidal, antifungal, …), toxin-antitoxin and treatment against toxic cyanobacteria based on PCC strains.
Cyanobacteria do not only produce toxins; they contain a real diversity in terms of natural substances. In order to have greater visibility of this diversity, we undertook the sequencing of the genomes of the living axenic strains preserved in the PCC collection. First, we obtained the genomes of 54 PCC strains selected on their morphology, their ecology and their physiology to better represent the breadth of cyanobacterial phylum. Combined with genomic data from 72 publicly available strains, a phylogenetic tree based on 31 genes conserved in Bacteria was constructed to reflect the evolution and relationship between these organisms. In parallel, a systematic analysis of gene clusters coding for NPs and toxins was undertaken on this dataset with the search for ribosomally synthesized and post-translationally modified peptides (RiPPs), non-ribosomal peptides synthetases (NRPS) and polyketide synthases (PKSs) [3]. This analysis showed the presence of these three classes of metabolites, with cis and trans AT-PKS, peptides from both ribosomal and non-ribosomal pathways, and terpenes throughout the phylum represented by these 126 genomes. In line with this work, a focus on the NRPS and PKS in the same dataset made it possible to highlight 452 biosynthetic gene clusters (BGCs) distributed into 286 cluster families based on the similarity of the modules of the NRPS and PKS and the length of the regions compared [36]. Interestingly, one-fourth of the BGCs were hybrids of NRPS and PKS, mostly distributed in the late branches of the cyanobacterial phylogeny, whereas the early branches contained mainly PKS. In addition, 80% of these cluster families did not correspond to any known NPs, giving an idea of the scope of the investigative work, which could be devolved by chemists and biochemists.
Based on the above findings and the availability of cyanobacterial strains potentially producing unknown NPs in the PCC collection, collaborations with various chemist and biochemist colleagues helped reveal the NPs derived from these unattributed BGCs. For the BGCs smaller than 15 kb, a cloning strategy performed in a heterologous host was more straightforward to discover compounds of interest and to find them further in the cyanobacterium of interest, such as the schizokinen-like siderophore of Leptolyngbya sp., PCC 7376 [57]. However, most of the NRPS-PKS BGCs were larger than 20 kb, and they could not be produced with this approach. For the investigations of these large BGCs, we sometimes cultivated liters of biomass of pure cyanobacterium, potentially producing the desired metabolites in our laboratory conditions. Through numerous collaborations, we have discovered more than 20 novel NPs and/or their BGC, thanks to the cyanobacteria in the PCC collection and the collaborators, as well as other researchers with PCC strains (Table 1).
Finally, in the investigation course of these NPs by a multidisciplinary consortium of chemists, biochemists and microbiologists, we were fascinated to find novel enzymes and unknown chemistry from the metabolisms of the cyanobacterial strains of the PCC, examples of which are discussed hereafter. First, the search for the proteusins of the cyanobacteria and the radical S-adenosyl methionine epimerase (rSAM) of this pathway revealed regioselective D-configured amino acids into peptidic NPs. For this, an ingenious methodology was developed to understand how these rSAMs work to irreversibly insert multiple D-amino acids in the peptides from the strains Kamptonema sp. PCC 6506, Pleurocapsa sp. PCC 7319, and Anabaena variabilis ATCC 29413 [58,59]. Secondly, the same talented chemists working with the genomic data of Pleurocapsa sp. PCC 7319 discovered non-canonical protein splicing via a post-translational excision of a tyramine equivalent, leading to an α-keto-β-amino amide [60]. Third, in the RiPPs family of chemically diverse cyanobactins, the BGC is highly conserved and thus, the genes coding the enzymes of these pathways are named consistently from A to F. Nevertheless, several F enzymes enlarged the prenyltransferase family, with two of these enzymes in the muscoride pathway acting differently by introducing a regioselective prenylation on the amino acid termini of the produced linear cyanobactin in Nostoc sp., PCC 7906 [61], or with another prenyltransferase which places a forward-prenyl on a threonine residue in the cyclic cyanobactin tolypamide of Tolypothrix sp. PCC 7601 [62]. Finally, the strain Lyngbya sp. PCC 8601 and two other cyanobacterial strains were used to uncover a suite of post-translational modifying bacterial enzymes that install single or multiple strained cyclophane macrocycles. As the cyclophane natural products are found in fungi, plants and bacteria; this enzyme family is widely distributed in nature [63].

5. From Molecules to Bioactivity

From the discovery of new molecules or their genetic heritage to the knowledge of their activity, the path is not straightforward. Initially, the origin of a toxic event was sought before finding the cyanotoxin responsible for it. Several reviews described the potential of cyanobacterial compounds to become drug products such as anticancer agents and antibiotics, for example [64,65,66,67]. Chemists have found more than 1100 NPs from cyanobacteria, but less than 20% of them are associated with a biosynthesis pathway [5]. In these chemical studies, it has often been attempted to find bioactivity associated with it by means of conventional screening techniques, in particular, to find a therapeutic potential. However, almost half of the compounds were not tested or detected in any bioactivity assay. More recently, a review on the bioactivity of NPs of cyanobacteria found 1630 unique molecules, classified into 260 families of metabolites [68]. Importantly, most of the compounds were not tested for their bioactivities. This is because bioactivity testing requires different knowledge and specialty than that needed to discover the compounds or the genetic data that encodes them. In addition, as the characterisation of a compound will often require the extraction and collection of grams of it from the biomass of the producing organism, it is often for lack of material that the bioactivity test cannot be carried out or confirmed.
Table 1. Natural products (NPs) and their biosynthetic gene clusters (BGC) discovered based on the monoclonal and axenic strains of the collection PCC.
Table 1. Natural products (NPs) and their biosynthetic gene clusters (BGC) discovered based on the monoclonal and axenic strains of the collection PCC.
NPs or Their BGCsType of NPsGene ClusterStrainReference
MicrocyclamideRiPPs, cyanobactinsmca, 13 kbMicrocystis aeruginosa PCC 7806[26]
ViridisamideRiPPs, 1st linear cyanobactinsVariation
from pat *
Oscillatoria viridis PCC 7112[69]
Aeruginosamide B and C3Microcystis aeruginosa PCC 9432
MuscorideRiPPs, linear cyanobactinmus, 12.7 kb12 Nostoc strains, in which 6 PCC strains[61]
TolypamideRiPPs, cyanobactinstol, 10.4 kbTolypothrix sp. PCC 7601[62]
GeosminRiPPs, sesquiterpenegeosmin synthaseNostoc spp. PCC 7310 and PCC 7120, Kamptonema sp. PCC 6506[70,71]
Merosterol A and B
+ isomer
RiPPs, meroterpenemst, 29 kbScytonema sp. PCC 10023[72]
CyclophanesRiPPs, cyclopetide alkaloidslsc, 2.6 kbLyngbya sp. PCC 8106
and various other strains
[63]
LandornamidesRiPPs, proteusinsosp, 12 kbKamptonema sp. PCC 6506 and 6 other Kamptonema PCC strains[73]
KamptornamideRiPPs, 1st ribosomal fatty-acylated lipo- petides, selidamidesksp, 6.3 kbKamptonema sp. PCC 6506 and
Nostoc punctiforme PCC 73102
[74]
Microguanidine amide Aeruginoguanidine BGCNRPSagd, 34 kb11 Microcystis,
in which 7 PCC strains
[75]
Hassallidin ENRPShas, 48 kbPlanktothrix serta PCC 8927[76]
CyanopeptolinNRPSoci *, 31.5 kbMicrocystis aeruginosa PCC 7806
Scytonema hofmanni PCC 7110
[77,78]
ScyptolinsNRPS Scytonema hofmanni PCC 7110[79]
Anatoxin-a and
dihydroanatoxin
PKS, alkaloidana, 20 kbKamptonema sp. PCC 6506,
Cylindrospermum sp. PCC 7417
and 13 other PCC strains
[36,55]
CylindrospermopsinsPKS, alkaloidcyr, 42 kbKamptonema sp. PCC 6506[56]
Luminaolide BTrans AT-PKSlum, 99 kbPlanktothrix paucivesiculata PCC 9631[80,81]
Tolytoxin BGCtto, >100 kbScytonema sp. PCC 10023
Tolytoxin, Scytophycintto, 92.8 kbPlanktothrix sp. PCC 11201
LeptolyngbyalideTrans AT-PKSlept, 96.7 kbLeptolyngbya sp. PCC 7375[82]
Alkene and alkanesPKS, Hydrocarbonols 10 kbSynechococcus sp. PCC 7002
16 unicellular PCC strains
[83,84,85]
Heterocyte glycolipidsPKS, polyunsaturated fatty acidCF1 *Nostoc sp. PCC 7120, 18 Nostocales strains and Microchaete sp. PCC 7126[86,87,88]
MicrocystinNRPS-PKS Hybridmcy, 55 kbMicrocystis aeruginosa PCC 7806, Fischerella sp. PCC 9339[36,54]
NostopeptolideNRPS-PKS Hybridpks2, 62.7 kbNostoc puctiforme ATCC29133/PCC 73102[89,90]
AranazolesNRPS-PKS Hybridarz, 43 kbFischerella sp. PCC 9339[91]
* Indicates: pat from Prochoron, a cyanobacterial symbiont of a tunicate [92]; oci from diverse cyanobacteria [93]; and CF1 correspond to the cluster in Nostoc sp. PCC 7120 [36,88].
The study of biological activity can be directed to help human interest, for a pharmacological application or a biotechnological development. Tolytoxin has the potential to be used for therapeutic application, as this cyanobacterial macrolide inhibits actin filament dynamics and was proposed as a potential anti-cancer drug [94,95,96]. However, this molecule was also proved to be extremely toxic at nM concentrations and to induce cell death [94,97]. As we recently described tolytoxin producers and the tolytoxin biosynthetic gene cluster from PCC pure strains [80], we found several other PCC strains capable of producing this molecule [81,98]. We revisited the activity of tolytoxin in human cells from neuronal and epithelial origins with the goal of reducing disease transmission by tunneling nanotubes mainly constituting of actin [98]. In this experiment, with the two strains we used, we noticed a strong decrease in tolytoxin dose needed (3 and 15 nM) to obtain an inhibitory effect without setting off the toxic side effects previously observed in cells. During the isolation of pure tolytoxin from Planktothrix serta PCC 8926 and Scytonema sp. PCC 10023, we noticed that a fatty acid was extracted along with the tolytoxin almost until the end of our extraction procedure. To perform this experiment and extract enough purified tolytoxin, we worked in a chemist’s laboratory, thanks to J. Piel’s team and R. Ueoka in particular, without whom we would have missed this trace of contamination. For the activity of tolytoxin, we collaborated with specialists in the cells and nanotubes to be tested, thanks to C Zurzolo’s team and A. Dilsizoglu-Senol in particular [98]. In addition, the odorous volatile compound geosmin is also of concern for human health with its biological activity. Using the geosmin-producing strain Kamptonema sp. PCC 6506, and the non-producing strain Leptolyngbya sp. PCC 8913 isolated from a lake colonized by mosquitoes in the south of France [99], we collaborated with researchers working on insect olfaction at Lund University (Sweden) to reveal the attraction of the Aedes aegypti mosquito for this compound and check if this odour is an indicator of egg-laying site for this insect [70].
The study of biological activity can also be investigated to learn more about its need for producing cyanobacteria. Two clear examples of useful NPs for producing cyanobacteria have been reported. The first one is the production of heterocyte glycolipids by a PKS cluster [88]. When a vegetative cell differentiates into a future heterocyte, the nascent cell only becomes active in fixing nitrogen when a layer of glycolipids covers it. This layer prevents oxygen, produced by adjacent cells, to enter into the heterocyte and to inhibit the nitrogenase. This mechanism must be tightly regulated and programmed because heterocytes can only survive 3 to 4 days before being replaced by another heterocyte resulting from the differentiation of a vegetative cell. The second example is from the product of a PKS cluster, the nostopeptolide that governs the cellular differentiation of a symbiotic Nostoc [89,90]. This example also indicates a clear scheduling of the production of the natural product at the time needed by the producing organism. Finally, the last example of the need of the producer to produce certain molecules at a certain time can be seen through the study of the toxic-bloom forming Microcystis aeruginosa PCC 7806. With an ingenious culture system consisting of two compartments separated by a filter, which allows compounds but not cells to pass, Briand and his collaborators demonstrated that this strain produced certain NPs in the medium only by sharing it with another strain of Microcystis such as PCC 9432 or Planktothrix agardhii PCC 7805 [100,101]. Thus, Microcystis aeruginosa PCC 7806 produces these NPs when it detects another cyanobacterium in its environment. This allelopathic research deserves further study because it illustrates a very controlled production of these metabolites beyond the genetic potential of the producer. It can also lead to the discovery of so-called cryptic NPs.

6. Concluding Remarks

In conclusion, the Pasteur Cultures of Cyanobacteria collection has been a living biobank and a research tool since its creation at the Institut Pasteur. The status of these strains has allowed research in the global scientific community. Within the framework of cyanobacterial toxins, the strains of the PCC collection led to the discovery of cyanotoxins and NPs. While several cyanotoxins were already structurally known, the pure strains maintained at the Institut Pasteur for 52 years have made it possible to discover the genetic origins of these toxins, intriguing enzymes, even unprecedented chemistry, and certain bioactivities. The genomics of the strains of the PCC collection highlights the wide diversity of NPs that we are still fully investigating.

Author Contributions

Writing—original draft preparation, M.G., A.B. and T.L.; writing—review and editing, M.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We thank Aniket Saraf, Collection of Cyanobacteria at the Institut Pasteur, for helpful comments. We also are grateful to our genomic and chemical collaborators for helping to valorize the collection. We acknowledge former researchers from the laboratories Unité de Physiologie Microbienne and Unité des Cyanobacteries who have preserved and developed the PCC. The Institut Pasteur funds the PCC collection and our research activities.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Garcia-Pichel, F.; Zehr, J.P.; Bhattacharya, D.; Pakrasi, H.B. What’s in a name? The case of cyanobacteria. J. Phycol. 2020, 56, 1–5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Schirrmeister, B.E.; Gugger, M.; Donoghue, P.C. Cyanobacteria and the Great Oxidation Event: Evidence from genes and fossils. Palaeontology 2015, 58, 769–785. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Shih, P.M.; Wu, D.; Latifi, A.; Axen, S.D.; Fewer, D.P.; Talla, E.; Calteau, A.; Cai, F.; Tandeau de Marsac, N.; Rippka, R.; et al. Improving the coverage of the cyanobacterial phylum using diversity-driven genome sequencing. Proc. Natl. Acad. Sci. USA 2013, 110, 1053–1058. [Google Scholar] [CrossRef] [Green Version]
  4. Huisman, J.; Codd, G.A.; Paerl, H.W.; Ibelings, B.W.; Verspagen, J.M.H.; Visser, P.M. Cyanobacterial blooms. Nat. Rev. Microbiol. 2018, 16, 471–483. [Google Scholar] [CrossRef]
  5. 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]
  6. Partensky, F.; Hess, W.R.; Vaulot, D. Prochlorococcus, a marine photosynthetic prokaryote of global significance. Microbiol. Mol. Biol. Rev. 1999, 63, 106–127. [Google Scholar] [CrossRef] [Green Version]
  7. Humbert, J.F.; Barbe, V.; Latifi, A.; Gugger, M.; Calteau, A.; Coursin, T.; Lajus, A.; Castelli, V.; Oztas, S.; Samson, G.; et al. A tribute to disorder in the genome of the bloom-forming freshwater cyanobacterium Microcystis aeruginosa. PLoS ONE 2013, 8, e70747. [Google Scholar] [CrossRef] [PubMed]
  8. Komarek, J.; Anagnostidis, K. Cyanoprokaryota I (Chroococales); Pascher, A., Ettl, H., Gärtner, G., Heyning, H., Mollenhauer, D., Eds.; Gustav Fischer: Jena, Germany, 1998; Volume 19/1. [Google Scholar]
  9. Komarek, J.; Anagnostidis, K. Cyanoprokaryota II (Oscillatoriales); Büdel, B., Gärtner, G., Krietniz, L., Schagerl, M., Eds.; Spektrum Akademischer Verlag: Jena, Germany, 2005; Volume 19/2. [Google Scholar]
  10. Komarek, J. Cyanoprokaryota III (Heterocytous Genera); Springer Spektrum: Berlin/Heidelberg, Germany, 2013; Volume 19/3. [Google Scholar]
  11. Castenholz, R.W.; Phylum, B.X. Cyanobacteria, Oxygenic Photosynthetic Bacteria. In Bergey’s Manual of Systematic Bacteriology, 2nd ed.; Boone, D.R., Castenholz, R.W., Garrity, G.M., Eds.; Volume 1—The Archaea and the Deeply Branching and Phototrophic Bacteria; Springer: New York, NY, USA, 2001; pp. 473–487. [Google Scholar]
  12. Gonzalez-Esquer, C.R.; Smarda, J.; Rippka, R.; Axen, S.D.; Guglielmi, G.; Gugger, M.; Kerfeld, C.A. Cyanobacterial ultrastructure in light of genomic sequence data. Photosynth. Res. 2016, 129, 147–157. [Google Scholar] [CrossRef]
  13. Stanier, R.Y.; Sistrom, W.R.; Hansen, T.A.; Whitton, B.A.; Castenholz, R.W.; Pfennig, N.; Gorlenko, V.N.; Kondratieva, E.N.; Eimhjellen, K.E.; Whittenbury, R.; et al. Proposal to place the nomemclature of the cyanobacteria (blue-green algae) under the rules of the International Code of Nomenclature of Bacteria. Int. J. Syst. Bacteriol. 1978, 28, 335–336. [Google Scholar] [CrossRef] [Green Version]
  14. Rippka, R.; Waterbury, J.; Cohen-Bazire, G. A cyanobacterium which lacks thylakoids. Arch. Microbiol. 1974, 100, 419–436. [Google Scholar] [CrossRef]
  15. Guglielmi, G.; Cohen-Bazire, G. Structure et distribution des pores et des perforations de l’enveloppe de peptidoglycane chez quelques cyanobactéries. Protistologica 1982, 18, 151–165. [Google Scholar]
  16. Guglielmi, G.; Cohen-Bazire, G. Etude taxonomique d’un genre de cyanobactérie oscillatoriacée: Le genre Pseudanabaena Lauterborn. I. Etude ultrastructurale. Protistologica 1984, 20, 377–391. [Google Scholar]
  17. Campbell, D.; Houmard, J.; Tandeau de Marsac, N. Electron transport regulates cellular differentiation in the filamentous cyanobacterium Calothrix. Plant Cell 1993, 5, 451–463. [Google Scholar] [CrossRef] [PubMed]
  18. Damerval, T.; Castets, A.M.; Guglielmi, G.; Houmard, J.; Tandeau de Marsac, N. Occurrence and distribution of gas vesicle genes among cyanobacteria. J. Bacteriol. 1989, 171, 1445–1452. [Google Scholar] [CrossRef] [Green Version]
  19. Forchhammer, K.; Tandeau de Marsac, N. Phosphorylation of the PII protein (glnB gene product) in the cyanobacterium Synechococcus sp. strain PCC 7942: Analysis of in vitro kinase activity. J. Bacteriol. 1995, 177, 5812–5817. [Google Scholar] [CrossRef] [Green Version]
  20. Herdman, M.; Coursin, T.; Rippka, R.; Houmard, J.; Tandeau de Marsac, N. A new appraisal of the prokaryotic origin of eukaryotic phytochromes. J. Mol. Evol. 2000, 51, 205–213. [Google Scholar] [CrossRef]
  21. Tandeau de Marsac, N.; Mazel, D.; Damerval, T.; Guglielmi, G.; Capuano, V.; Houmard, J. Photoregulation of gene expression in the filamentous cyanobacterium Calothrix sp. PCC 7601: Light-harvesting complexes and cell differentiation. Photosynth. Res. 1988, 18, 99–132. [Google Scholar] [CrossRef]
  22. Mazel, D.; Houmard, J.; Tandeau de Marsac, N. A multigene family in Calothrix sp. PCC 7601 encodes phycocyanin, the major component of the cyanobacterial light-harvesting antenna. Mol. Gen. Genet. 1988, 211, 296–304. [Google Scholar] [CrossRef]
  23. Damerval, T.; Guglielmi, G.; Houmard, J.; Tandeau de Marsac, N. Hormogonium differentiation in the cyanobacterium Calothrix: A photoregulated developmental process. Plant Cell 1991, 3, 191–201. [Google Scholar] [CrossRef]
  24. Liotenberg, S.; Campbell, D.; Rippka, R.; Houmard, J.; Tandeau de Marsac, N. Effect of the nitrogen source on phycobiliprotein synthesis and cell reserves in a chromatically adapting filamentous cyanobacterium. Microbiology 1996, 142, 611–622. [Google Scholar] [CrossRef] [Green Version]
  25. Frangeul, L.; Quillardet, P.; Castets, A.M.; Humbert, J.F.; Matthijs, H.C.; Cortez, D.; Tolonen, A.; Zhang, C.C.; Gribaldo, S.; Kehr, J.C.; et al. Highly plastic genome of Microcystis aeruginosa PCC 7806, a ubiquitous toxic freshwater cyanobacterium. BMC Genom. 2008, 9, 274. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Ziemert, N.; Ishida, K.; Quillardet, P.; Bouchier, C.; Hertweck, C.; Tandeau de Marsac, N.; 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] [PubMed] [Green Version]
  27. Zilliges, Y.; Kehr, J.C.; Mikkat, S.; Bouchier, C.; Tandeau de Marsac, N.; Börner, T.; Dittmann, E. An extracellular glycoprotein is implicated in cell-cell contacts in the toxic cyanobacterium Microcystis aeruginosa PCC 7806. J. Bacteriol. 2008, 190, 2871–2879. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Rippka, R.; Deruelles, J.; Waterbury, J.B.; Herdman, M.; Stanier, R.Y. Generic assignments, strain histories and properties of pure cultures of cyanobacteria. J. Gen. Microbiol. 1979, 111, 1–61. [Google Scholar] [CrossRef] [Green Version]
  29. Rippka, R.; Herdman, M. Pasteur Culture Collection of Cyanobacterial Strains in Axenic Culture. In Catalogue & Taxonomic Handbook; Institut Pasteur: Paris, France, 1992; pp. 1–103. [Google Scholar]
  30. Palinska, K.A.; Jahns, T.; Rippka, R.; Tandeau de Marsac, N. Prochlorococcus marinus strain PCC 9511, a picoplanktonic cyanobacterium, synthesizes the smallest urease. Microbiology 2000, 146, 3099–3107. [Google Scholar] [CrossRef] [Green Version]
  31. Adomako, M.; Ernst, D.; Simkovsky, R.; Chao, Y.Y.; Wang, J.; Fang, M.; Bouchier, C.; Lopez-Igual, R.; Mazel, D.; Gugger, M.; et al. Comparative genomics of Synechococcus elongatus explains the phenotypic diversity of the strains. mBio 2022, 13, e0086222. [Google Scholar] [CrossRef]
  32. Rippka, R.; Coursin, T.; Hess, W.; Lichtlé, C.; Scanlan, D.J.; Palinska, K.A.; Iteman, I.; Partensky, F.; Houmard, J.; Herdman, M. Prochlorococcus marinus Chisholm et al. 1992, subsp. nov. pastoris, strain PCC 9511, the first axenic chlorophyll a2/b2-containing cyanobacterium (Oxyphotobacteria). Int. J. Syst. Evol. Microbiol 2000, 50, 1833–1847. [Google Scholar] [CrossRef] [Green Version]
  33. Włodarczyk, A.; Selão, T.T.; Norling, B.; Nixon, P.J. Newly discovered Synechococcus sp. PCC 11901 is a robust cyanobacterial strain for high biomass production. Commun. Biol. 2020, 3, 215. [Google Scholar] [CrossRef]
  34. Jaiswal, D.; Sengupta, A.; Sohoni, S.; Sengupta, S.; Phadnavis, A.G.; Pakrasi, H.B.; Wangikar, P.P. Genome features and biochemical characteristics of a robust, fast growing and naturally transformable cyanobacterium Synechococcus elongatus PCC 11801 isolated from India. Sci. Rep. 2018, 8, 16632. [Google Scholar] [CrossRef]
  35. Jaiswal, D.; Sengupta, A.; Sengupta, S.; Madhu, S.; Pakrasi, H.B.; Wangikar, P.P. A novel cyanobacterium Synechococcus elongatus PCC 11802 has distinct genomic and metabolomic characteristics compared to its neighbor PCC 11801. Sci. Rep. 2020, 10, 191. [Google Scholar] [CrossRef] [Green Version]
  36. Calteau, A.; Fewer, D.P.; Latifi, A.; Coursin, T.; Laurent, T.; Jokela, J.; Kerfeld, C.A.; Sivonen, K.; Piel, J.; Gugger, M. Phylum-wide comparative genomics unravel the diversity of secondary metabolism in Cyanobacteria. BMC Genom. 2014, 15, 977. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Francis, G. Poisonous Australian Lake. Nature 1878, 18, 11–12. [Google Scholar] [CrossRef] [Green Version]
  38. Edwards, C.; Beattie, K.A.; Scrimgeour, C.M.; Codd, G.A. Identification of anatoxin-A in benthic cyanobacteria (blue-green algae) and in associated dog poisonings at Loch Insh, Scotland. Toxicon 1992, 30, 1165–1175. [Google Scholar] [CrossRef] [PubMed]
  39. Gugger, M.; Lenoir, S.; Berger, C.; Ledreux, A.; Druart, J.C.; Humbert, J.F.; Guette, C.; Bernard, C. First report in a river in France of the benthic cyanobacterium Phormidium favosum producing anatoxin-a associated with dog neurotoxicosis. Toxicon 2005, 45, 919–928. [Google Scholar] [CrossRef]
  40. Harding, W.R.; Rowe, N.; Wessels, J.C.; Beattie, K.A.; Codd, G.A. Death of a dog attributed to the cyanobacterial (blue-green algal) hepatotoxin nodularin in South Africa. J. S. Afr. Vet. Assoc. 1995, 66, 256–259. [Google Scholar]
  41. Kelly, L.T.; Bouma-Gregson, K.; Puddick, J.; Fadness, R.; Ryan, K.G.; Davis, T.W.; Wood, S.A. Multiple cyanotoxin congeners produced by sub-dominant cyanobacterial taxa in riverine cyanobacterial and algal mats. PLoS ONE 2019, 14, e0220422. [Google Scholar] [CrossRef] [Green Version]
  42. Mahmood, N.A.; Carmichael, W.W.; Pfahler, D. Anticholinesterase poisonings in dogs from a cyanobacterial (blue-green algae) bloom dominated by Anabaena flos-aquae. Am. J. Vet. Res. 1988, 49, 500–503. [Google Scholar]
  43. Turner, A.D.; Turner, F.R.I.; White, M.; Hartnell, D.; Crompton, C.G.; Bates, N.; Egginton, J.; Branscombe, L.; Lewis, A.M.; Maskrey, B.H. Confirmation Using Triple Quadrupole and High-Resolution Mass Spectrometry of a Fatal Canine Neurotoxicosis following Exposure to Anatoxins at an Inland Reservoir. Toxins 2022, 14, 804. [Google Scholar] [CrossRef]
  44. Wood, S.A.; Heath, M.W.; Holland, P.T.; Munday, R.; McGregor, G.B.; Ryan, K.G. Identification of a benthic microcystin-producing filamentous cyanobacterium (Oscillatoriales) associated with a dog poisoning in New Zealand. Toxicon 2010, 55, 897–903. [Google Scholar] [CrossRef]
  45. Azevedo, S.M.; Carmichael, W.W.; Jochimsen, E.M.; Rinehart, K.L.; Lau, S.; Shaw, G.R.; Eaglesham, G.K. Human intoxication by microcystins during renal dialysis treatment in Caruaru-Brazil. Toxicology 2002, 181–182, 441–446. [Google Scholar] [CrossRef]
  46. Breinlinger, S.; Phillips, T.J.; Haram, B.N.; Mareš, J.; Martínez Yerena, J.A.; Hrouzek, P.; Sobotka, R.; Henderson, W.M.; Schmieder, P.; Williams, S.M.; et al. Hunting the eagle killer: A cyanobacterial neurotoxin causes vacuolar myelinopathy. Science 2021, 371, eaax9050. [Google Scholar] [CrossRef] [PubMed]
  47. Krienitz, L.; Ballot, A.; Kotut, K.; Wiegand, C.; Pütz, S.; Metcalf, J.S.; Codd, G.A.; Pflugmacher, S. Contribution of hot spring cyanobacteria to the mysterious deaths of Lesser Flamingos at Lake Bogoria, Kenya. FEMS Microbiol. Ecol. 2003, 43, 141–148. [Google Scholar] [CrossRef] [PubMed]
  48. Skafi, M.; Duy, S.V.; Munoz, G.; Dinh, Q.T.; Simon, D.F.; Juneau, P.; Sauvé, S. Occurrence of microcystins, anabaenopeptins and other cyanotoxins in sh from a freshwater wildlife reserve impacted by harmful cyanobacterial blooms. Toxicon 2021, 194, 44–52. [Google Scholar] [CrossRef] [PubMed]
  49. Veerman, J.; Kumar, A.; Mishra, D.R. Exceptional landscape-wide cyanobacteria bloom in Okavango Delta, Botswana in 2020 coincided with a mass elephant die-off event. Harmful Algae 2022, 111, 102145. [Google Scholar] [CrossRef]
  50. Chorus, I.; Welker, M. Toxic Cyanobacteria in Water: A Guide to Their Pubic Health Consequences, Monitoring and Management, 2nd ed.; Welker, I.C.M., Ed.; CRC Press: Boca Raton, FL, USA; on behalf of the World Health Organization: Geneva, Switzerland, 2021. [Google Scholar]
  51. Trevino-Garrison, I.; DeMent, J.; Ahmed, F.S.; Haines-Lieber, P.; Langer, T.; Ménager, H.; Neff, J.; van der Merwe, D.; Carney, E. Human illnesses and animal deaths associated with freshwater harmful algal blooms-Kansas. Toxins 2015, 7, 353–366. [Google Scholar] [CrossRef] [Green Version]
  52. Hilborn, E.D.; Beasley, V.R. One health and cyanobacteria in freshwater systems: Animal illnesses and deaths are sentinel events for human health risks. Toxins 2015, 7, 1374–1395. [Google Scholar] [CrossRef] [Green Version]
  53. Theiss, W.C.; Carmichael, W.W.; Wyman, J.; Bruner, R. Blood pressure and hepatocellular effects of the cyclic heptapeptide toxin produced by the freshwater cyanobacterium (blue-green alga) Microcystis aeruginosa strain PCC-7820. Toxicon 1988, 26, 603–613. [Google Scholar] [CrossRef]
  54. Tillett, D.; Dittmann, E.; Erhard, M.; von Döhren, H.; Börner, T.; Neilan, B.A. Structural organization of microcystin biosynthesis in Microcystis aeruginosa PCC7806: An integrated peptide-polyketide synthetase system. Chem. Biol. 2000, 7, 753–764. [Google Scholar] [CrossRef] [Green Version]
  55. Méjean, A.; Mann, S.; Maldiney, T.; Vassiliadis, G.; Lequin, O.; Ploux, O. Evidence that biosynthesis of the neurotoxic alkaloids anatoxin-a and homoanatoxin-a in the cyanobacterium Oscillatoria PCC 6506 occurs on a modular polyketide synthase initiated by L-proline. J. Am. Chem. Soc. 2009, 131, 7512–7513. [Google Scholar] [CrossRef]
  56. Mazmouz, R.; Chapuis-Hugon, F.; Mann, S.; Pichon, V.; Méjean, A.; Ploux, O. Biosynthesis of cylindrospermopsin and 7-epicylindrospermopsin in Oscillatoria sp. strain PCC 6506: Identification of the cyr gene cluster and toxin analysis. Appl. Environ. Microbiol. 2010, 76, 4943–4949. [Google Scholar] [CrossRef] [Green Version]
  57. Wang, S.; Pearson, L.A.; Mazmouz, R.; Liu, T.; Neilan, B.A. Heterologous expression and biochemical analysis reveal a schizokinen-based siderophore pathway in Leptolyngbya (Cyanobacteria). Appl. Environ. Microbiol. 2022, 88, e0237321. [Google Scholar] [CrossRef] [PubMed]
  58. Morinaka, B.I.; Verest, M.; Freeman, M.F.; Gugger, M.; Piel, J. An orthogonal D2 O-based induction system that provides insights into d-amino acid pattern formation by radical S-adenosylmethionine peptide epimerases. Angew. Chem. Int. Ed. 2017, 56, 762–766. [Google Scholar] [CrossRef]
  59. Morinaka, B.I.; Vagstad, A.L.; Helf, M.J.; Gugger, M.; Kegler, C.; Freeman, M.F.; Bode, H.B.; Piel, J. Radical S-adenosyl methionine epimerases: Regioselective introduction of diverse D-amino acid patterns into peptide natural products. Angew. Chem. Int. Ed. 2014, 53, 8503–8507. [Google Scholar] [CrossRef] [PubMed]
  60. Morinaka, B.I.; Lakis, E.; Verest, M.; Helf, M.J.; Scalvenzi, T.; Vagstad, A.L.; Sims, J.; Sunagawa, S.; Gugger, M.; Piel, J. Natural noncanonical protein splicing yields products with diverse β-amino acid residues. Science 2018, 359, 779–782. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Mattila, A.; Andsten, R.M.; Jumppanen, M.; Assante, M.; Jokela, J.; Wahlsten, M.; Mikula, K.M.; Sigindere, C.; Kwak, D.H.; Gugger, M.; et al. Biosynthesis of the bis-prenylated alkaloids muscoride A and B. ACS Chem. Biol. 2019, 14, 2683–2690. [Google Scholar] [CrossRef] [PubMed]
  62. Purushothaman, M.; Sarkar, S.; Morita, M.; Gugger, M.; Schmidt, E.W.; Morinaka, B.I. Genome-mining-based discovery of the cyclic peptide tolypamide and tolF, a Ser/Thr forward O-prenyltransferase. Angew. Chem. Int. Ed. 2021, 60, 8460–8465. [Google Scholar] [CrossRef]
  63. Nguyen, T.Q.N.; Tooh, Y.W.; Sugiyama, R.; Nguyen, T.P.D.; Purushothaman, M.; Leow, L.C.; Hanif, K.; Yong, R.H.S.; Agatha, I.; Winnerdy, F.R.; et al. Post-translational formation of strained cyclophanes in bacteria. Nat. Chem. 2020, 12, 1042–1053. [Google Scholar] [CrossRef]
  64. Chlipala, G.E.; Mo, S.; Orjala, J. Chemodiversity in freshwater and terrestrial cyanobacteria—A source for drug discovery. Curr. Drug Targets 2011, 12, 1654–1673. [Google Scholar] [CrossRef]
  65. Gerwick, W.H.; Fenner, A.M. Drug discovery from marine microbes. Microb. Ecol. 2013, 65, 800–806. [Google Scholar]
  66. Vijayakumar, S.; Menakha, M. Pharmaceutical applications of cyanobacteria-A review. J. Acute Med. 2015, 5, 15–23. [Google Scholar] [CrossRef] [Green Version]
  67. Ali Shah, S.A.; Akhter, N.; Auckloo, B.N.; Khan, I.; Lu, Y.; Wang, K.; Wu, B.; Guo, Y.W. Structural diversity, biological properties and applications of natural products from cyanobacteria. A review. Mar. Drugs 2017, 15, 354. [Google Scholar] [CrossRef] [Green Version]
  68. 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]
  69. Leikoski, N.; Liu, L.; Jokela, J.; Wahlsten, M.; Gugger, M.; Calteau, A.; Permi, P.; Kerfeld, C.A.; Sivonen, K.; Fewer, D.P. Genome mining expands the chemical diversity of the cyanobactin family to include highly modified linear peptides. Chem. Biol. 2013, 20, 1033–1043. [Google Scholar] [CrossRef] [PubMed]
  70. Melo, N.; Wolff, G.H.; Costa-da-Silva, A.L.; Arribas, R.; Triana, M.F.; Gugger, M.; Riffell, J.A.; DeGennaro, M.; Stensmyr, M.C. Geosmin attracts Aedes aegypti mosquitoes to oviposition sites. Curr. Biol. 2020, 30, 127–134. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. 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]
  72. Moosmann, P.; Ueoka, R.; Grauso, L.; Mangoni, A.; Morinaka, B.I.; Gugger, M.; Piel, J. Cyanobacterial ent-sterol-like natural products from a deviated ubiquinone pathway. Angew. Chem. Int. Ed. 2017, 56, 4987–4990. [Google Scholar] [CrossRef] [PubMed]
  73. Bösch, N.M.; Borsa, M.; Greczmiel, U.; Morinaka, B.I.; Gugger, M.; Oxenius, A.; Vagstad, A.L.; Piel, J. Landornamides: Antiviral ornithine-containing ribosomal peptides discovered through genome mining. Angew. Chem. Int. Ed. 2020, 59, 11763–11768. [Google Scholar] [CrossRef]
  74. Hubrich, F.; Bösch, N.M.; Chepkirui, C.; Morinaka, B.I.; Rust, M.; Gugger, M.; Robinson, S.L.; Vagstad, A.L.; Piel, J. Ribosomally derived lipopeptides containing distinct fatty acyl moieties. Proc. Natl. Acad. Sci. USA 2022, 119, e2113120119. [Google Scholar] [CrossRef]
  75. Pancrace, C.; Ishida, K.; Briand, E.; Pichi, D.G.; Weiz, A.R.; Guljamow, A.; Scalvenzi, T.; Sassoon, N.; Hertweck, C.; Dittmann, E.; et al. Unique biosynthetic pathway in bloom-forming cyanobacterial genus Microcystis jointly assembles cytotoxic aeruginoguanidines and microguanidines. ACS Chem. Biol. 2019, 14, 67–75. [Google Scholar] [CrossRef] [Green Version]
  76. Pancrace, C.; Jokela, J.; Sassoon, N.; Ganneau, C.; Desnos-Ollivier, M.; Wahlsten, M.; Humisto, A.; Calteau, A.; Bay, S.; Fewer, D.P.; et al. Rearranged biosynthetic gene cluster and synthesis of hassallidin E in Planktothrix serta PCC 8927. ACS Chem. Biol. 2017, 12, 1796–1804. [Google Scholar] [CrossRef] [Green Version]
  77. Martin, C.; Oberer, L.; Buschdtt, M.; Weckesser, J. Cyanopeptolins, new depsipeptides from the cyanobacterium Microcystis sp. PCC 7806. J. Antibiot. 1993, 46, 1550–1556. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Matern, U.; Oberer, L.; Erhard, M.; Herdmand, M.; Weckesser, J. Hofmannolin, a cyanopeptolin from Scytonema hofmanni PCC 7110. Phytochemistry 2003, 64, 1061–1067. [Google Scholar] [CrossRef] [PubMed]
  79. Matern, U.; Oberer, L.; Falchetto, R.A.; Erhard, M.; König, W.A.; Herdman, M.; Weckesser, J. Scyptolin A and B, cyclic depsipeptides from axenic cultures of Scytonema hofmanni PCC 7110. Phytochemistry 2001, 58, 1087–1095. [Google Scholar] [CrossRef] [PubMed]
  80. Ueoka, R.; Uria, A.R.; Reiter, S.; Mori, T.; Karbaum, P.; Peters, E.E.; Helfrich, E.J.; Morinaka, B.I.; Gugger, M.; Takeyama, H.; et al. Metabolic and evolutionary origin of actin-binding polyketides from diverse organisms. Nat. Chem. Biol. 2015, 11, 705–712. [Google Scholar] [CrossRef]
  81. Pancrace, C.; Barny, M.A.; Ueoka, R.; Calteau, A.; Scalvenzi, T.; Pédron, J.; Barbe, V.; Piel, J.; Humbert, J.F.; Gugger, M. Insights into the Planktothrix genus: Genomic and metabolic comparison of benthic and planktic strains. Sci. Rep. 2017, 7, 41181. [Google Scholar] [CrossRef] [Green Version]
  82. Helfrich, E.J.N.; Ueoka, R.; Dolev, A.; Rust, M.; Meoded, R.A.; Bhushan, A.; Califano, G.; Costa, R.; Gugger, M.; Steinbeck, C.; et al. Automated structure prediction of trans-acyltransferase polyketide synthase products. Nat. Chem. Biol. 2019, 15, 813–821. [Google Scholar] [CrossRef]
  83. Brito, Ã.; Vieira, J.; Vieira, C.P.; Zhu, T.; Leao, P.N.; Ramos, V.; Lu, X.; Vasconcelos, V.M.; Gugger, M.; Tamagnini, P. Comparative genomics discloses the uniqueness and the biosynthetic potential of the marine cyanobacterium Hyella patelloides. Front. Microbiol. 2020, 11, 1527. [Google Scholar] [CrossRef]
  84. Coates, R.C.; Podell, S.; Korobeynikov, A.; Lapidus, A.; Pevzner, P.; Sherman, D.H.; Allen, E.E.; Gerwick, L.; Gerwick, W.H. Characterization of cyanobacterial hydrocarbon composition and distribution of biosynthetic pathways. PLoS ONE 2014, 9, e85140. [Google Scholar] [CrossRef] [Green Version]
  85. Mendez-Perez, D.; Begemann, M.B.; Pfleger, B.F. Modular synthase-encoding gene involved in α-olefin biosynthesis in Synechococcus sp. strain PCC 7002. Appl. Environ. Microbiol. 2011, 77, 4264–4267. [Google Scholar] [CrossRef] [Green Version]
  86. Bauersachs, T.; Gugger, M.; Schwark, L. Heterocyte glycolipid diketones: A novel type of biomarker in the N2-fixing heterocytous cyanobacterium Microchaete sp. Org. Geochem. 2020, 141, 103976. [Google Scholar] [CrossRef]
  87. Bauersachs, T.; Miller, S.R.; Gugger, M.; Mudimu, O.; Friedl, T.; Schwark, L. Heterocyte glycolipids indicate polyphyly of stigonematalean cyanobacteria. Phytochemistry 2019, 166, 112059. [Google Scholar] [CrossRef] [PubMed]
  88. Fan, Q.; Huang, G.; Lechno-Yossef, S.; Wolk, C.P.; Kaneko, T.; Tabata, S. Clustered genes required for synthesis and deposition of envelope glycolipids in Anabaena sp. strain PCC 7120. Mol. Microbiol. 2005, 58, 227–243. [Google Scholar] [CrossRef] [PubMed]
  89. Liaimer, A.; Jenke-Kodama, H.; Ishida, K.; Hinrichs, K.; Stangeland, J.; Hertweck, C.; Dittmann, E. A polyketide interferes with cellular differentiation in the symbiotic cyanobacterium Nostoc punctiforme. Environ. Microbiol. Rep. 2011, 3, 550–558. [Google Scholar] [CrossRef]
  90. Liaimer, A.; Helfrich, E.J.N.; Hinrichs, K.; Guljamow, A.; Ishida, K.; Hertweck, C.; Dittmann, E. Nostopeptolide plays a governing role during cellular differentitaion of the symbiotic cyanobacterium Nostoc punctiforme. Proc. Natl. Acad. Sci. USA 2015, 112, 1862–1867. [Google Scholar] [CrossRef] [Green Version]
  91. Moosmann, P.; Ueoka, R.; Gugger, M.; Piel, J. Aranazoles: Extensively chlorinated nonribosomal peptide-polyketide hybrids from the cyanobacterium Fischerella sp. PCC 9339. Org. Lett. 2018, 20, 5238–5241. [Google Scholar] [CrossRef] [PubMed]
  92. 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]
  93. Rounge, T.B.; Rohrlack, T.; Tooming-Klunderud, A.; Kristensen, T.; Jakobsen, K.S. Comparison of cyanopeptolin genes in Planktothrix, Microcystis, and Anabaena strains: Evidence for independent evolution within each genus. Appl. Environ. Microbiol. 2007, 73, 7322–7330. [Google Scholar] [CrossRef] [Green Version]
  94. Patterson, G.M.L.; Smith, C.D.; Kimura, L.H.; Britton, B.A.; Carmeli, S. Action of tolytoxin on cell morphology, cytoskeletal organization, and actin polymerization. Cell Motil. Cytoskelet. 1993, 24, 39–48. [Google Scholar] [CrossRef]
  95. Zhang, X.; Minale, L.; Zampella, A.; Smith, C.D. Microfilament depletion and circumvention of multiple drug resistance by sphinxolides. Cancer Res. 1997, 57, 3751–3758. [Google Scholar]
  96. Smith, C.D.; Carmeli, S.; Moore, R.E.; Patterson, G.M. Scytophycins, novel microfilament-depolymerizing agents which circumvent P-glycoprotein-mediated multidrug resistance. Cancer Res. 1993, 53, 1343–1347. [Google Scholar]
  97. Patterson, G.M.; Carmeli, S. Biological effects of tolytoxin (6-hydroxy-7-O-methyl-scytophycin b), a potent bioactive metabolite from cyanobacteria. Arch. Microbiol. 1992, 157, 406–410. [Google Scholar] [CrossRef] [PubMed]
  98. Dilsizoglu Senol, A.; Pepe, A.; Grudina, C.; Sassoon, N.; Ueoka, R.; Bousset, L.; Melki, R.; Piel, J.; Gugger, M.; Zurzolo, C. Effect of tolytoxin on tunneling nanotube formation and function. Sci. Rep. 2019, 9, 5741. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Thiery, I.; Nicolas, L.; Rippka, R.; Tandeau de Marsac, N. Selection of cyanobacteria isolated from mosquito breeding sites as a potential food source for mosquito larvae. Appl. Environ. Microbiol. 1991, 57, 1354–1359. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Briand, E.; Reubrecht, S.; Mondeguer, F.; Sibat, M.; Hess, P.; Amzil, Z.; Bormans, M. Chemically mediated interactions between Microcystis and Planktothrix: Impact on their growth, morphology and metabolic profiles. Environ. Microbiol. 2019, 21, 1552–1566. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Briand, E.; Bormans, M.; Gugger, M.; Dorrestein, P.C.; Gerwick, W.H. Changes in secondary metabolic profiles of Microcystis aeruginosa strains in response to intraspecific interactions. Environ. Microbiol. 2016, 18, 384–400. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Morphologies of cyanobacteria. Unicellular morphotypes are on the top row: from tiny single cells of Cyanobium sp. PCC 7001 in dark field (a), to flat square colonies of Merismopedia, from an environmental sample (b) to the baeocytous former Stanieria sp. PCC 7301 (c). Filamentous morphotypes on the bottom row: from Planktothrix agardhii PCC 10110 with bright gas vacuoles (d) to Aphanizomenon flos aquae PCC 7905 in dark field with gas vacuoles and barrel-shaped heterocyte (e), to Fischerella sp. in the late stage of purification (f). All scale bars represent 5 µm.
Figure 1. Morphologies of cyanobacteria. Unicellular morphotypes are on the top row: from tiny single cells of Cyanobium sp. PCC 7001 in dark field (a), to flat square colonies of Merismopedia, from an environmental sample (b) to the baeocytous former Stanieria sp. PCC 7301 (c). Filamentous morphotypes on the bottom row: from Planktothrix agardhii PCC 10110 with bright gas vacuoles (d) to Aphanizomenon flos aquae PCC 7905 in dark field with gas vacuoles and barrel-shaped heterocyte (e), to Fischerella sp. in the late stage of purification (f). All scale bars represent 5 µm.
Toxins 15 00388 g001
Figure 2. Number of axenic monoclonal cyanobacterial strains of the PCC obtained over the years 1963–2019 (in white), cultures still preserved today (2023, in black), and total number of PCC strains (black line and second vertical axis).
Figure 2. Number of axenic monoclonal cyanobacterial strains of the PCC obtained over the years 1963–2019 (in white), cultures still preserved today (2023, in black), and total number of PCC strains (black line and second vertical axis).
Toxins 15 00388 g002
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gugger, M.; Boullié, A.; Laurent, T. Cyanotoxins and Other Bioactive Compounds from the Pasteur Cultures of Cyanobacteria (PCC). Toxins 2023, 15, 388. https://doi.org/10.3390/toxins15060388

AMA Style

Gugger M, Boullié A, Laurent T. Cyanotoxins and Other Bioactive Compounds from the Pasteur Cultures of Cyanobacteria (PCC). Toxins. 2023; 15(6):388. https://doi.org/10.3390/toxins15060388

Chicago/Turabian Style

Gugger, Muriel, Anne Boullié, and Thierry Laurent. 2023. "Cyanotoxins and Other Bioactive Compounds from the Pasteur Cultures of Cyanobacteria (PCC)" Toxins 15, no. 6: 388. https://doi.org/10.3390/toxins15060388

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