Next Article in Journal
Similar Response of a Range Expanding Dragonfly to Low- and High-Elevation Predators
Next Article in Special Issue
Nodosilinea hunanesis sp. nov. (Prochlorotrichaceae, Synechococcales) from a Freshwater Pond in China Based on a Polyphasic Approach
Previous Article in Journal
The Association of Waminoa with Reef Corals in Singapore and Its Impact on Putative Immune- and Stress-Response Genes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Taxonomic Discussion on Cyanobacterial Systematics at Family Level, with Special Regards to Phormidiaceae by Using the Strains of Chinese Newly Recorded Genera Ancylothrix and Potamolinea

1
College of Life and Environmental Sciences, Wenzhou University, Wenzhou 325035, China
2
Hangzhou Institute of Ecological and Environmental Sciences, Hangzhou 310000, China
3
College of Life Science, Henan Normal University, Xinxiang 453007, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Submission received: 22 March 2022 / Revised: 12 April 2022 / Accepted: 14 April 2022 / Published: 15 April 2022
(This article belongs to the Special Issue Diversity and Ecology of Algae in China)

Abstract

:
In the past decades, the taxonomic status of the cyanobacterial family Phormidiaceae has always been chaotic and problematic. In this study, filamentous cyanobacteria were investigated in the east of China, and twenty strains isolated from different locations of Zhejiang Province were characterized. Using the polyphasic approach combining morphological, molecular and phylogenetic features, these strains were grouped within the members of the genera Ancylothrix and Potamolinea, the newly recorded genera of cyanobacteria in China. Based on the collected taxonomic information of the family Phormidiaceae, cyanobacterial systematics at family level were further discussed. This study provided a simple and efficient example to perform the phylogenetic evaluation for the monophyly and rationality of currently used families of cyanobacteria by using the regional strains based on the polyphasic approach.

1. Introduction

The cyanobacterial taxonomic system has encountered rapid and radical revisions during the last decades [1,2,3,4]. Such revisions, through intensive usage of the polyphasic approach combining morphological, ecophysiological, biochemical and molecular characters, led to the establishment of a large number of new genera and species [5,6,7,8]. The updated cyanobacterial taxonomy summarized by Komárek et al. (2014) proposed an eight orders, 46 families and 202 genera system, in which the eight-ordered frame was obtained by phylogenomic analyses. This new system further presented a detailed and effective guideline to characterize the cyanobacterial genera and provided evaluation of the present status of all the 202 cyanobacterial genera mainly based on the availability of molecular data. However, the monophyly at the family level in this system was not discussed even though all the families were described in this system, and such an evaluation on the status of all the families are highly dependent on premised work on the cyanobacterial genera. In order to achieve the ideal taxonomic system of cyanobacteria at all categories, the polyphasic approach should also be used in the studies on the cyanobacterial families, especially some important ones.
The family Phormidiaceae, separated from the family Oscillatoriaceae, was created by Anagnostidis and Komárek (1988) in their monograph titled “modern approach to the classification system of cyanophytes 3-Oscillatoriales”. With Phormidium as the type genus, Phormidiaceae was further divided into three subfamilies as Phormidioideae, Microcoleoideae and Spirulinoideae, including many important genera such as Planktothrix, Trichodesmium, Microcoleus, Spirulina and Arthrospira. In “system of cyanoprokaryotes (cyanobacteria) state in 2004, Hoffman et al. (2005) indicated that the higher-level cyanobacterial classification system did not reflect the evolutionary history of the cyanobacteria and needed to be revised, and they retained the family Phormidiaceae in Oscillatoriales, Oscillatoriophycidae in their revised system. Ten years later, the family Phormidiaceae was not included in the updated taxonomic system of cyanobacteria proposed by Komárek et al. (2014) since the authors indicated that the type species of Phormidium (P. lucidum) was considered to correspond to the family Oscillatoriaceae. Later, several taxonomic studies with descriptions of new cyanobacterial genera/species clearly belonging to the family Phormidiaceae were published [9,10]. Moreover, many cyanobacterial studies still retained the name of Phormidiaceae [11,12,13,14]. In particular, AlgaeBase (https://www.algaebase.org/, accessed on 12 April 2022) currently recognizes the existence of the family Phormidiaceae with a total of five extant genera, including Ammassolinea [15], Ancylothrix [9], Cephalothrix [16], Potamolinea [10] and Pseudoscillatoria [17], with reliable molecular data, and the genus Phormidium was not included in Phormidiaceae in the AlgaeBase. Such conflicting points led to certain confusion about the existence or non-existence of the family Phormidiaceae. Therefore, it is much needed to evaluate the family Phormidiaceae with regards to its existence and phylogeny by using the polyphasic approach.
During an investigation on filamentous cyanobacteria in China, twenty strains from three cyanobacterial samples were isolated at different localities of Zhejiang Province. Phylogenetic analyses showed that these strains were very closely related to the genera Ancylothrix and Potamolinea, both of which were newly recorded genera of cyanobacteria in China. The present study focused on a taxonomic discussion of the cyanobacterial systematics at the family level, with the aim to explore the relationship among different families by using the strains studies in this study. The results and discussion are expected to help the evaluation of the existence of Phormidiaceae as a family in Oscillatoriales.

2. Materials and Methods

2.1. Cyanobacterial Collection and Cultivation

Twenty cyanobacterial strains studied in this work were isolated from two localities in Zhejiang Province of China. Detailed information about the habitat of studied strains is summarized in Table 1. Benthic mat samples near to riverbank and soil samples (Figure 1) were scraped using a tweezer or scraper and live samples were washed thoroughly with sterile water before isolation. A single trichome from the pretreated samples were isolated by lab-mad Pasteur pipette under a 40 times magnification dissecting microscope (Carl Zeiss STEMI 508, Jena, Germany) and then moved to 24-well plates containing sterilized liquid CT medium for the unialgal culture [18]. After two or three weeks, the blue-green and uncontaminated cultures, after checking, were transferred into screw-capped tubes containing 10 mL of the same medium. These strains were kept at 25 °C under a 12 h:12 h light–dark cycle with a photon flux density of 35 μmol m−2 s−1. These strains were archived in the Algae Culture Collection of the college of life and environmental science, Wenzhou University, Wenzhou, China, under the strain number in Table 1.

2.2. Morphological Observation

All the live cultures were examined with a LEICA DM2000 LED microscope (Leica, Germany). Microphotographs were produced using a LEICA DMC 5400 digital camera (Leica, Germany) photomicrographic system attached to the microscope and the images were analyzed using Leica Application Suite X 3.7.4 software. Measurements of unialgal cellular morphology were made from digital images taken under 1000 times magnification and the mean width of vegetative cells of more than 60 filaments were measured.

2.3. DNA Extraction and PCR Amplification

A certain amount of filament was collected and was washed three times using sterile phosphorus-free sterilized liquid CT medium to avoid contamination with other bacteria. Total genomic DNA from the strains were extracted using the modified cetyltrimethylammonium bromide (CTAB) method [19]. The PCR primers, PA [20] and B23S [21], were chosen for obtaining segments containing 16S ribosomal RNA gene and the associated 16S–23S internal transcribed spacer (ITS) region. The total PCR reaction volume of 50 μL comprised 1 μL of template DNA, each primer (10 μmol L−1) for 1 μL, 22 μL sterile water and 25 μL 2× Taq Plus Master Mix (Dye Plus) (Vazyme Biotech Co., Ltd., Nanjing, China). PCR amplification was performed using a SimpliAmpTM Thermal Cycler (Waltham, MA, USA) with a PCR profile of an initial denaturation at 95 °C for 5 min, 31 cycles of 30 s at 95 °C, 30 s at 55 °C, 2 min at 72 °C and then a final 10 min elongation step at 72 °C.
PCR products were purified using a TIANgel Midi Purification kit (Tiangen Biotech Co., Ltd., Beijing, China) and were then cloned using pClone007 Versatile Simple Vector Kit (Beijing Tsingke Biotech CO., Ltd., Beijing, China). The target gene fragment was inserted into Escherichia coli DH5α cells for replication, and then the inserted fragment was sequenced bidirectionally by using the standard primers M13F (5′-GTA-AAA-CGA-CGG-CCA-GT-3′) and M13R (5′-GTC-ATA-GCT-GTT-TCC-TG-3′). Clones including the target fragment were sequenced by the Wuhan Tianyi Huayu Gene Technology (Wuhan, China). All sequences obtained in this study were submitted to the NCBI GenBank database after being checked and under following accession numbers in Table 2 and Table 3.

2.4. Phylogenetic Analyses

The obtained 16S rRNA gene sequences of this study were compared by using the online tool nucleotide BLAST and the highly similarly sequences of 16S rRNA gene were downloaded from the NCBI GenBank database and used for phylogenetic analyses. All downloaded sequences and obtained sequences were aligned using MAFFT v7.463 [22] and cut both ends neatly by using BioEdit v7.0.9 [23]. Sequences of Gloeobaacter violaceus and Gloeobaacter kilaueensis were used as outgroups. After this, multiple sequence alignment finally formed a data matrix of 108 sequences with 1071 nucleotide sites. The nucleic acid substitution model (GTR + F + R4) was selected for the optimal maximum likelihood analysis (ML) and the other model (GTR + F + 4) was selected for the optimal Bayesian analysis (BI). Both BI and ML analysis were conducted on the basis of Akaike Information Criterion (AIC) in PhyloSuite v1.2.2 molecular phylogenetic platform [24]. The specific operation parameters of substitution models were individually estimated by IQ-TREE v2.1.3 [25,26] and MrBayes v3.2.7 [27]. For the ML analysis, a total of 1000 bootstrap replications were made under standard option. BI analysis comprised two parallel runs for 10,000,000 generations, sampling every 100th generation, in which the initial 25% of sampled data were discarded as burn-in. Neighbor-Joining (NJ) analyses were computed using the Kimura 2-parameter model with 1000 bootstrap replications in MEGA 11 [28]. Phylogenetic trees were visualized in FigTree v1.4.3 (http://tree.bio.ed.ac.uk/software/figtree/, accessed on 22 March 2022), and all obtained were edited by TreeView 1.6.6 software [29]. Genetic distances of similarity matrices of the 16S rRNA gene sequences were estimated using the Kimura 2-parameter model in MEGA 11 to calculate p-distance with pairwise deletion of gaps.

2.5. Analyses of 16S–23S Internal Transcribed Spacer (ITS)

The presence of tRNA gene sequences were tested with the tRNAscan-SE2.0 web server [30]. The conserved hypothetical secondary structures were constructed using S-fold web-based software [31] and adjusted in the Adobe illustrator 2020 software.

3. Results

3.1. Morphological Description

Diagnosis: The twenty strains isolated in this study were similar to Phormidium in morphological and ecological characteristics, and the phylogenetic analysis showed that these strains were close to the species of the genera Ancylothrix (WZU 0009, WZU 0011, WZU 0013, WZU 0014, WZU 0153, WZU 0154) and Potamolinea (WZU 0155–WZU 0168). According to their morphological and ecological differences, these strains were divided into three different species.

3.1.1. Phormidium-like sp. 1

Description: In the natural environment, thallus is macroscopic, with dark-green filaments forming benthic mats on submerged substrata. Filaments are bright-green or dark-green. Sheaths are very thin, colorless and hyaline. Trichomes are cylindrical, 4.6–7.1 μm wide, 3.3–6.0 μm long, shorter than wide to isodiametric, with a single filament growing, contracting slightly at the crosswalls, and the ends of filaments gradually become sharp and undirected left and right curved. Most of the apical cells are conical-rounded and narrowed, without calyptra (Figure 2). Reproduction is by fragmentation of trichomes at necridic cells. They are without heterocytes, aerotopes and akinetes.
Habitat: River benthos near to riverbank.
Reference strains: WZU 0009, WZU 0010, WZU 0013, WZU 0014.

3.1.2. Phormidium-like sp. 2

Description: Thallus is macroscopic forming bright-green compact biofilms on a xerophytic environment. Sheaths are very thin, colorless and hyaline. Through observation, sheaths are more obvious than Ancylothrix rivularis in laboratory culture. Filaments are olive-green or dark-green. Trichomes are cylindrical, 3.8–6.1μm wide and 2.9–6.4 μm long, shorter than wide to isodiametric, with a single filament growing, contracting slightly at the crosswalls and the ends of filaments gradually become sharp and undirected left and right curved. Most of the apical cells are conical-rounded and narrowed, a small number of apical cells are rounded, without calyptra (Figure 3). Reproduction is by fragmentation of trichomes at necridic cells. Heterocytes, aerotopes and akinetes were not observed under culture conditions.
Habitat: Attached on rock surfaces.
Reference strains: WZU 0153, WZU 0154.

3.1.3. Phormidium-like sp. 3

Description: Thallus is attached to the substrate, dark-green. Trichomes are densely entangled with each other. Sheaths are hyaline, colorless and thin. Trichomes are cylindrical and vegetative cells are usually cylindrical, 6.8–10.6 μm wide and 4.5–7.8 μm long, shorter than wide. Trichome has a single filament growing, contracting slightly at the crosswalls, and the ends of filaments gradually become sharp (Figure 4). Most of the apical cells are rounded, a small number of the apical cells are conical-rounded and narrowed, without calyptra. Reproduction is by the disintegration of trichomes at the necridic cells. Heterocytes, aerotopes and akinetes were not observed.
Habitat: At the bottom of the stream.
Reference strains: WZU 0155-WZU 0168.

3.2. Molecular and Phylogenetic Analyses

The phylogenetic analyses of the 16S rRNA gene confirmed that the studied strains were a well supported clade among genera Ancyclothrix and Potamolinea. In total, 108 sequences from the order Oscillatoriales were used based on NJ, ML and BI methods to construct the 16S rRNA gene phylogenetic trees (Figure 5) in order to better understand the strains from Phormidiaceae in the present and investigate their relationships in the order Oscillatoriales, including in the families Coleofasciculaceae, Desertifilaceae, Microcoleaceae and Oscillatoriaceae and all genus from Phormidiaceae. In this study, six strains (WZU 0009, 0011, 0013, 0014, 0153, 0154) with the strains of two species, Ancylothrix rivularis and Ancylothrix terrestris, were placed in a clade based on the phylogenetic tree with high bootstrap values of 99%/100%/1.00 supported by the NJ/ML/BI approaches. Pairwise p-distance analysis of the 16S rRNA gene can also be used as evidence in phylogenetic comparison [32]. The 16S rRNA gene similarities between Phormidium-like sp. 1 and Ancylothrix rivularis strains, and Phormidium-like sp. 2 and Ancylothrix terrestris strains were all >99%, higher than 97%—the threshold of bacterial genus classification (Table 4). Phylogenetic analyses placed the other fourteen WZU strains (0155–0168) with Potamolinea aerugineocaerulea in a robust clade; all the species from species Potamolinea aerugineocaerulea shared 98.13%–100% similarities, also higher than the threshold of bacterial species classification (Table 5). Therefore, based on phylogenetic analysis, we determined that Phormidium-like sp. 1–3 represent the three species Ancylothrix rivularis, Ancylothrix terrestris and Potamolinea aerugineocaerulea, respectively.

3.3. Analyses of ITS between 16S and 23S rRNA Gene and Secondary Structures

The 16S–23S ITS secondary structures were compared among Ancyclothrix and Potamolinea aerugineocaerulea in this study. The ITS region can be used as a tool to distinguish closely related organisms [33,34]. Specific information of the various parts of the 16S–23S ITS region is summarized in Table 2 and Table 3.
For genus Ancyclothrix, all sequences of the six strains contained both tRNAIle and tRNAAla. Compared with the original seven strains, the D1–D1′, Box–B and V3 helix (nt) of all thirteen strains exhibited eight (Figure 6), six (Figure 7) and seven (Figure 8) different types. For genus Potamolinea, all strains contained only tRNAIle without tRNAAla, and regardless of whether D1–D1′, Box–B or V3 helices, WZU strains showed their uniqueness (Figure 9).
The studied D1–D1′ helix showed conserved secondary structures in strains from the Ancylothrix rivularis clade, and variable secondary structures in strains from the Ancylothrix terrestris clade. WZU0009, 0011 and 0014 share the same structure (Figure 6a). For the Ancylothrix rivularis clade, the main difference lies in the terminal loop: the number of unpaired bases A and C have a slight difference (Figure 6a–c). For Ancylothrix terrestris clade, WZU 0153, 0154 and 10PC, and 13PC share the same structure (Figure 6e,f). 11PC and 12PC have a clear difference to the other four strains. (Figure 6g,h). For the Potamolinea aerugineocaerulea clade, the strains in this study showed a clear difference from the three original strains. WZU strains consisted of one 4-bp helix, one 5-bp helix, one small unidirectional bulge, one 2-bp helix, one 3:3 bp base bilateral bulge, one 3-bp helix, one 4:1 unidirectional bulge, one 5-bp helix and one 4-bp apical loop (Figure 9c).
The studied Box–B helix showed higher conserved secondary structures in strains from the Ancylothrix rivularis clade than the D1–D1′ helix, and WZU strains shared the same structure (Figure 7a); other original strains shared the structure also (Figure 7b). The secondary structures between the two types showed high similarity. There are slight differences in their respective base bilateral bulge. For the Ancylothrix rivularis clade, WZU strains share the same structure with 10PC and 13PC; the main difference lies in the terminal loop and the number of unpaired bases had a slight difference (Figure 7c,d). 11PC and 12PC have a clear difference from the other four strains. (Figure 7e,f). For the Potamolinea aerugineocaerulea clade, WZU strains had their own unique Box–B helix, consisting of one 5-bp helix, one 1:2 bp base bilateral bulge, one 3-bp helix, one small unidirectional bulge, one 7-bp helix and 15-bp unpaired bases in the terminal loop (Figure 9g).
The studied V3 helix showed the highest conserved secondary structures in strains from the Ancylothrix rivularis clade. Except for WZU 0009, all other strains in the Ancylothrix rivularis clade shared the same structure (Figure 8b); WZU 0009 had a one 2:3 bp base bilateral bulge (5′-GG-GAA-3′) (Figure 8a), while all other strains in the Ancylothrix rivularis clade at the same place had a one 2:3 bp base bilateral bulge (5′-GG-AAA-3′) (Figure 8b). For the Ancylothrix rivularis clade, WZU strains share the same structure with 10PC and 13PC (Figure 8c–e), whereas, in the terminal loop, the unpaired bases arrangement had a slight difference (C-U). For the Potamolinea aerugineocaerulea clade, WZU strains have their own unique V3 helix, consisting of one 2-bp helix, one 4:4 bp base bilateral bulge, one 8-bp helix, one 1:1 bp base bilateral bulge, one 4-bp helix, one 1:1 bp base bilateral bulge, one 5-bp helix and 4-bp unpaired bases in the terminal loop, which differ significantly from other strains in the Potamolinea aerugineocaerulea clade (Figure 9h–k).

4. Discussion

With the advent of phylogenetic analyses based on molecular sequence data, extensive revision and reconstruction of cyanobacterial classification have been performed in recent years. An ideal cyanobacterial taxonomic system in which all categories should be monophyletic has been repeatedly proposed [2,35,36]. As we described above, the order-level and genus-level classification in the newly revised system of cyanobacteria have been well documented [32]. However, phylogenetic elucidation and evaluation on the monophyly of cyanobacterial groups at a family level have barely been conducted, implying that taxonomic revision of the cyanobacterial families may be enforced in the future. In this study, we aimed to investigate the diversity of oscillatorean cyanobacteria in the east of China. Through isolating filamentous strains from mat-forming habitats, twenty strains with Phormidium-like morphologies were successfully obtained. Previous studies on the Phormidium-like group based on phenotypic and genotypic data, mostly using 16S rRNA gene sequences, have revealed a wide diversity and heterogeneity, leading to the establishment of several new genera such as Roseofilum [37], Ammassolinea [15] and Kamptonema [38]. DNA sequence-based comparison and phylogenetic analyses in this study revealed these twenty strains as belonging to the two newly proposed genera Ancylothrix and Potamolinea, respectively. Furthermore, the secondary structures of ITS also supported their taxonomic positions decided by the 16S rRNA gene sequences. Both Ancylothrix and Potamolinea were found in China for the first time, and they were added as new recorded genera of cyanobacteria in China. Ancylothrix was newly described by Martins et al. (2016) from samples in southern Brazil and was further found to contain two internal groups: one group represents a species with only freshwater occurrences and the other group with only aerophytic populations, both corresponding to Ancylothrix rivularis and A. terrestris, respectively. The finding of the genus Ancylothrix in the present study with two species and habitats was exactly the same as that by Martins et al. (2016), reflecting the common and stable features of the genus Ancylothrix [9]. Similarly, the genus Potamolinea was created by Martins and Branco (2016) from ten populations of bottom streams in Brazil, with further description of two species. Potamolinea aerugineocaerulea, found in the present study, was very similar to that described by Martin and Branco (2016) in ecological and phylogenetic aspects [10].
The modern taxonomic system of cyanobacteria is superior to the traditional systems because it more closely reflects the phylogeny. However, the delimitation of cyanobacterial family is still unclear. This is indeed the case for the family Phormidiaceae. Continuous changes along the history of Phormidiaceae, from the beginning of the family in 1988 to its disappearance in 2014 by Komárek et al. (2014), and the current retention of this family in many studies, have resulted in much confusion and uncertainty for the cyanobacterial taxonomic system. The genera Ammassolinea [15], Ancylothrix [9], Cephalothrix [16] and Potamolinea [10] were founded after Komárek et al. (2014), but the inventors of those new genera only considered the morphological similarity to the polyphyletic genus Phormidium without taking into account the fact of the absence of the family Phormidiaceae in Komárek et al. (2014). In order to evaluate the monophyly of the family Phormidiaceae, the phylogenetic tree, based on a large number of 16S rRNA gene sequences including those from type species of both Oscillatoria and Phormidium, was obtained (Figure 5). As shown in Figure 5, the strains of different Phormidium species including the type species, were shown to be diversely and sparsely located at different clades, indicating Phormidium as a largely heterogeneous genus. Furthermore, phylogenetic evaluation showed that the five extant genera within the family Phormidiaceae in AlgaeBase did not group together, and were diversely distributed at different clades, corresponding to Oscillatoriaceae, Microcoleaceae and Coleofasciculaceae, respectively (Figure 5). Obviously, the family Phormidiaceae currently used is far from the monophyly based on the phylogenetic evaluation. With the combination of the above results and analyses, the conclusion was drawn that the family Phormidiaceae should not exist in the current taxonomic system of cyanobacteria. The results and conclusion strongly support the taxonomic removal of the family Phormidiaceae in the revised taxonomic system by Komárek et al. (2014), clearly preventing further confusion in determining the taxonomic belonging of Phormidium-like cyanobacteria.
The taxonomic revision of cyanobacteria at genus and species levels based on the polyphasic approach have been largely performed and emphasized during the last decades, but there are fewer studies on taxonomic revisions at family level and elucidation for family delimitation. The present study provided a simple and efficient example to perform phylogenetic evaluation for the monophyly and rationality of currently accepted families of cyanobacteria by using the regional strains based on the polyphasic approach.

Author Contributions

Conceptualization, Y.C. and R.G.; methodology, Y.C. and R.G.; software, R.G. and P.X.; formal analysis, Y.C. and R.G.; investigation, L.S.; resources, Y.L. and H.Z.; data curation, Y.C. and R.G.; writing—original draft preparation, Y.C.; writing—review and editing, R.L.; visualization, Z.M.; funding acquisition, R.L. All authors have read and agreed to the published version of the manuscript.

Funding

This study is funded by the NSFC project (No. 31970219).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The 16S rRNA gene and 16S–23S ITS sequences are available in the GenBank database.

Acknowledgments

The first author thanks Guotao Wang and Wei Deng for physical assistance in the process of collecting samples.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. 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]
  2. Anagnostidis, K.; Komárek, J. Modern approach to the classification system of cyanophytes. 1–Introduction. Algol. Stud./Arch. Für Hydrobiol. 1985, 38–39, 291–302. [Google Scholar]
  3. Hoffmann, L.; Komárek, J.; Kaštovský, J. System of cyanoprokaryotes (cyanobacteria)—State in 2004. Algol. Stud. 2005, 117, 95–115. [Google Scholar] [CrossRef]
  4. Hoffmann, L.; Komárek, J.; Kaštovský, J. Proposal of cyanobacterial system—2004. In Süsswasserflora Von Mitteleuropa 19/2; Büdel, B., Krienitz, L., Gärtner, G., Schagerl, M., Eds.; Elsevier: Heidelberg, Germany, 2005; pp. 657–660. [Google Scholar]
  5. Fiore, M.F.; Sant’anna, C.L.; Azevedo, M.T.D.P.; Komárek, J.; Kaštovský, J.; Sulek, J.; Lorenzi, A.A.S. The cyanobacterial genus Brasilonema, gen. nov., a molecular and phenotypic evaluation. J. Phycol. 2007, 43, 789–798. [Google Scholar] [CrossRef]
  6. Bohunická, M.; Pietrasiak, N.; Johansen, J.R.; Gómez, E.B.; Hauer, T.; Gaysina, L.A.; Lukešová, A. Roholtiella, gen. nov.(Nostocales, Cyanobacteria)—A tapering and branching cyanobacteria of the family Nostocaceae. Phytotaxa 2015, 197, 84–103. [Google Scholar] [CrossRef] [Green Version]
  7. Genuário, D.B.; Vaz, M.G.M.V.; Hentschke, G.S.; Sant’anna, C.L.; Fiore, M.F. Halotia gen. nov., a phylogenetically and physiologically coherent cyanobacterial genus isolated from marine coastal environments. Int. J. Syst. Evol. Microbiol. 2015, 15, 663–675. [Google Scholar] [CrossRef] [Green Version]
  8. Sciuto, K.; Moro, I. Detection of the new cosmopolitan genus Thermoleptolyngbya (Cyanobacteria, Leptolyngbyaceae) using the 16S rRNA gene and 16S–23S ITS region. Mol. Phylogenetics Evol. 2016, 105, 15–35. [Google Scholar] [CrossRef]
  9. Martins, M.D.; Rigonato, J.; Taboga, S.R.; Branco, L.H.Z. Proposal of Ancylothrix gen. nov., a new genus of phormidiaceae (Cyanobacteria, Oscillatoriales) based on a polyphasic approach. Int. J. Syst. Evol. Microbiol. 2016, 66, 2396–2405. [Google Scholar] [CrossRef]
  10. Martins, M.D.; Branco, L.H.Z. Potamolinea gen. nov (Oscillatoriales, Cyanobacteria): A phylogenetically and ecologically coherent cyanobacterial genus. Int. J. Syst. Evol. Microbiol. 2016, 66, 3632–3641. [Google Scholar] [CrossRef]
  11. Borges, H.L.F.; Branco, L.H.Z.; Martins, M.D.; Lima, C.S.; Barbosa, P.T.; Lira, G.A.S.T.; Bittencourt-Oliveira, M.C.; Molica, R.J.R. Cyanotoxin production and phylogeny of benthic cyanobacterial strains isolated from the northeast of Brazil. Harmful Algae 2015, 43, 46–57. [Google Scholar] [CrossRef]
  12. Yap-Dejeto, L.G.; Batula, H.S. Bloom of Trichodesmium (Oscillatoriales, Phormidiaceae) and seasonality of potentially harmful phytoplankton in San Pedro Bay, Leyte, Philippines. Rev. De Biol. Trop. 2016, 64, 897–911. [Google Scholar]
  13. Te, S.H.; Tan, B.F.; Boo, C.Y.; Thompson, J.R.; Gin, K.Y.H. Genomics insights into production of 2-methylisoborneol and a putative cyanobactin by Planktothricoides sp. SR001. Stand. Genom. Sci. 2017, 12, 35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Rozwalak, P.; Podkowa, P.; Buda, J.; Niedzielski, P.; Zawierucha, K. Cryoconite—From minerals and organic matter to bioengineered sediments on glacier’s surfaces. Sci. Total Environ. 2022, 807, 150874. [Google Scholar] [CrossRef] [PubMed]
  15. Hašler, P.; Dvořák, P.; Poulíčková, A.; Casamatta, D.A. A novel genus Ammassolinea gen. nov. (Cyanobacteria) isolated from sub–tropical epipelic habitats. Fottea 2014, 14, 241–248. [Google Scholar] [CrossRef] [Green Version]
  16. Malone, C.F.d.S.; Rigonato, J.; Laughinghouse, H.D.; Schmidt, É.C.; Bouzon, Z.L.; Wilmotte, A.; Sant’Anna, C.L. Cephalothrix gen. nov. (Cyanobacteria): Towards an intraspecific phylogenetic evaluation by multilocus analyses. Int. J. Syst. Evol. Microbiol. 2015, 65, 2993–3007. [Google Scholar] [CrossRef]
  17. Rasoulouniriana, D.; Siboni, N.; Ben-Dov, E.; Kramarsky-Winter, E.; Loya, Y.; Kushmaro, A. Pseudoscillatoria coralii gen. nov., sp nov., a cyanobacterium associated with coral black band disease (BBD). Dis. Aquat. Org. 2009, 87, 91–96. [Google Scholar] [CrossRef]
  18. Ichimura, T. Isolation and culture methods of algae. Methods Phycol. Stud. 1979, 2, 294–305. [Google Scholar]
  19. Neilan, B.A.; Jacobs, D.; Goodman, A.E. Genetic diversity and phylogeny of toxic cyanobacteria determined by DNA polymorphisms within the phycocyanin locus. Appl. Environ. Microbiol. 1995, 61, 3875–3883. [Google Scholar] [CrossRef] [Green Version]
  20. Edwards, U.; Rogall, T.; Blöcker, H.; Emde, M.; Böttger, E.C. Isolation and direct complete nucleotide determination of entire genes. Characterization of a gene coding for 16S ribosomal RNA. Nucleic Acids Res. 1989, 17, 7843–7853. [Google Scholar] [CrossRef] [Green Version]
  21. Gkelis, S.; Rajaniemi, P.; Vardaka, E.; Moustaka–gouni, M.; Lanaras, T.; Sivonen, K. Limnothrix redekei (Van Goor) Meffert (Cyanobacteria) strains from Lake Kastoria, Greece from a separate phylogenetic group. Microb. Ecol. 2005, 49, 176–182. [Google Scholar] [CrossRef]
  22. Katoh, K.; Standley, D.M. MAFFT multiple sequence alignment software version 7: Improvements in performance and usability. Mol. Biol. Evol. 2013, 30, 772–780. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Hall, T.A. BioEdit: A User-Friendly Biological Sequence Alignment Editor and Analysis Program for Windows 95/98/NT. Nucleic Acids Symp. Ser. 1999, 41, 95–98. [Google Scholar]
  24. Zhang, D.; Gao, F.; Jakovlić, I.; Zou, H.; Zhang, J.; Li, W.X.; Wang, G.T. PhyloSuite: An integrated and scalable desktop platform for streamlined molecular sequence data management and evolutionary phylogenetics studies. Mol. Ecol. Resour. 2020, 20, 348–355. [Google Scholar] [CrossRef] [PubMed]
  25. Guindon, S.; Dufayard, J.-F.; Lefort, V.; Anisimova, M.; Hordijk, W.; Gascuel, O. New Algorithms and Methods to Estimate Maximum-Likelihood Phylogenies: Assessing the Performance of PhyML 3.0. Syst. Biol. 2010, 59, 307–321. [Google Scholar] [CrossRef] [Green Version]
  26. Trifinopoulos, J.; Nguyen, L.T.; von Haeseler, A.; Minh, B.Q. W-IQ-TREE: A fast online phylogenetic tool for maximum likelihood analysis. Nucleic Acids Res. 2016, 44, W232–W235. [Google Scholar] [CrossRef] [Green Version]
  27. Ronquist, F.; Teslenko, M.; van der Mark, P.; Ayres, D.L.; Darling, A.; Höhna, S.; Huelsenbeck, J.P. MrBayes 3.2: Efficient Bayesian Phylogenetic Inference and Model Choice Across a Large Model Space. Syst. Biol. 2012, 61, 539–542. [Google Scholar] [CrossRef] [Green Version]
  28. Tamura, K.; Stecher, G.; Kumar, S. MEGA11: Molecular Evolutionary Genetics Analysis Version 11. Mol. Biol. Evol. 2021, 38, 3022–3027. [Google Scholar] [CrossRef]
  29. Page, R.D.M. TreeView: An application to display phylogenetic trees on personal computers. Bioinformatics 1996, 12, 357–358. [Google Scholar] [CrossRef] [Green Version]
  30. Lowe, T.M.; Chan, P.P. tRNAscan-SE On-line: Integrating search and context for analysis of transfer RNA genes. Nucleic Acids Res. 2016, 44, W54–W57. [Google Scholar] [CrossRef]
  31. Rennie, W.; Kanoria, S.; Liu, C.; Mallick, B.; Long, D.; Wolenc, A.; Ding, Y. STarMirDB: A database of microRNA binding sites. RNA Biol. 2016, 13, 554–560. [Google Scholar] [CrossRef]
  32. Osorio-Santos, K.; Pietrasiak, N.; Bohunická, M.; Miscoe, L.H.; Kováčik, L.; Martin, M.P.; Johansen, J.R. Seven new species of Oculatella (Pseudanabaenales, Cyanobacteria): Taxonomically recognizing cryptic diversification. Eur. J. Phycol. 2014, 49, 450–470. [Google Scholar] [CrossRef] [Green Version]
  33. Caires, T.A.; Lyra, G.D.; Hentschke, G.S.; da Silva, A.M.S.; de Araujo, V.L.; Sant’Anna, C.L.; Nunes, J.M.D. Polyphasic delimitation of a filamentous marine genus, Capillus gen. nov. (Cyanobacteria, Oscillatoriaceae) with the description of two Brazilian species. Algae 2018, 33, 291–304. [Google Scholar] [CrossRef]
  34. Pietrasiak, N.; Mühlsteinová, R.; Siegesmund, M.A.; Johansen, J.R. Phylogenetic placement of Symplocastrum (Phormidiaceae, Cyanophyceae) with a new combination S. californicum and two new species: S. flechtnerae and S. torsivum. Phycologia 2014, 53, 529–541. [Google Scholar] [CrossRef]
  35. Komárek, J.; Kaštovský, J.; Mareš, J.; Johansen, J.R. Taxonomic classification of cyanoprokaryotes (Cyanobacterial genera) 2014, using a polyphasic approach. Preslia 2014, 86, 295–335. [Google Scholar]
  36. Miscoe, L.H.; Johansen, J.R.; Vaccarino, M.A.; Pietrasiak, N.; Sherwood, A.R. The diatom flora and cyanobacteria from caves on Kauai, Hawaii. II. Novel cyanobacteria from caves on Kauai, Hawaii. Bibl. Phycol. 2016, 123, 75–152. [Google Scholar]
  37. Casamatta, D.; Stanić, D.; Gantar, M.; Richardson, L.L. Characterization of Roseofilum reptotaenium (Oscillatoriales, Cyanobacteria) gen. et sp. nov. isolated from Caribbean black band disease. Phycologia 2012, 51, 489–499. [Google Scholar] [CrossRef]
  38. Strunecky, O.; Komárek, J.; Šmarda, J. Kamptonema (Microcoleaceae, Cyanobacteria), a new genus derived from the polyphyletic Phormidium on the basis of combined molecular and cytomorphological markers. Preslia 2014, 86, 193–207. [Google Scholar]
Figure 1. The habitats of studied strains. (a) Benthic mat samples near to riverbank including Phormidium-like sp. 1 strains. (b) Soil samples including Phormidium-like sp. 2 strains. (c) Benthic mat samples at the bottom of the streams including Phormidium-like sp. 3 strains. (d) Macroscopic view of all WZU strains.
Figure 1. The habitats of studied strains. (a) Benthic mat samples near to riverbank including Phormidium-like sp. 1 strains. (b) Soil samples including Phormidium-like sp. 2 strains. (c) Benthic mat samples at the bottom of the streams including Phormidium-like sp. 3 strains. (d) Macroscopic view of all WZU strains.
Diversity 14 00301 g001
Figure 2. Light microscopy of Phormidium-like sp. 1 strains: (af) The filaments are densely entangled with each other; (gj) different shapes of filaments, some apical cells rounded, some apical cells conical-rounded and narrowed; (k,l) trichomes slightly constricted at the crosswalls; (ms) details of the conical-rounded and narrowed apical cells; (t,u) details of the thin and colorless sheath. Scale bars: (a,b) = 50 μm, (cj) = 20 μm, (ku) = 10 μm.
Figure 2. Light microscopy of Phormidium-like sp. 1 strains: (af) The filaments are densely entangled with each other; (gj) different shapes of filaments, some apical cells rounded, some apical cells conical-rounded and narrowed; (k,l) trichomes slightly constricted at the crosswalls; (ms) details of the conical-rounded and narrowed apical cells; (t,u) details of the thin and colorless sheath. Scale bars: (a,b) = 50 μm, (cj) = 20 μm, (ku) = 10 μm.
Diversity 14 00301 g002
Figure 3. Light microscopy of Phormidium-like sp. 2 strains: (ac) Represent twisted or entwined filaments; (dg) trichomes wrapped in the sheath; (h,i) trichomes slightly constricted at the crosswalls; (jr) the details of various diversiform of apical cells. Scale bars: (ac) = 50 μm, (dr) = 10 μm.
Figure 3. Light microscopy of Phormidium-like sp. 2 strains: (ac) Represent twisted or entwined filaments; (dg) trichomes wrapped in the sheath; (h,i) trichomes slightly constricted at the crosswalls; (jr) the details of various diversiform of apical cells. Scale bars: (ac) = 50 μm, (dr) = 10 μm.
Diversity 14 00301 g003
Figure 4. Light microscopy of Phormidium-like sp. 3 strains: (ac) Numerous trichomes are densely entangled with each other; (d,e) detail of the sheath attached to filaments; (fn,pr,t) the details of various diversiform of apical cells; (o,s) detail of biconcave necridic cells. Scale bars: (a,b) = 50 μm, (c,e) = 20 μm, (d,ft) = 10 μm.
Figure 4. Light microscopy of Phormidium-like sp. 3 strains: (ac) Numerous trichomes are densely entangled with each other; (d,e) detail of the sheath attached to filaments; (fn,pr,t) the details of various diversiform of apical cells; (o,s) detail of biconcave necridic cells. Scale bars: (a,b) = 50 μm, (c,e) = 20 μm, (d,ft) = 10 μm.
Diversity 14 00301 g004
Figure 5. Bayesian inference (BI) phylogenetic tree of WZU strains based on 16S rRNA gene sequences. Bootstrap values greater than 50% are shown on the BI tree for NJ/ML methods and Bayesian posterior probabilities, clade A–E represent Potamolinea, Cephalothrix, Ancylothrix, Ammassolinea, Pseudoscillatoria. Bar, 0.03 substitutions per nucleotide position.
Figure 5. Bayesian inference (BI) phylogenetic tree of WZU strains based on 16S rRNA gene sequences. Bootstrap values greater than 50% are shown on the BI tree for NJ/ML methods and Bayesian posterior probabilities, clade A–E represent Potamolinea, Cephalothrix, Ancylothrix, Ammassolinea, Pseudoscillatoria. Bar, 0.03 substitutions per nucleotide position.
Diversity 14 00301 g005
Figure 6. D1–D1′ helix of Ancylothrix strains: (a) Ancylothrix rivularis WZU 0009, 0011, 0014; (b) Ancylothrix rivularis WZU 0013; (c) Ancylothrix rivularis 7PC, 8PC; (d) Ancylothrix rivularis 9PC; (e) Ancylothrix terrestris 10PC, WZU 0153, WZU 0154; (f) Ancylothrix terrestris 13PC; (g) Ancylothrix terrestris 11PC; (h) Ancylothrix terrestris 12PC.
Figure 6. D1–D1′ helix of Ancylothrix strains: (a) Ancylothrix rivularis WZU 0009, 0011, 0014; (b) Ancylothrix rivularis WZU 0013; (c) Ancylothrix rivularis 7PC, 8PC; (d) Ancylothrix rivularis 9PC; (e) Ancylothrix terrestris 10PC, WZU 0153, WZU 0154; (f) Ancylothrix terrestris 13PC; (g) Ancylothrix terrestris 11PC; (h) Ancylothrix terrestris 12PC.
Diversity 14 00301 g006
Figure 7. Box–B helix of Ancylothrix strains: (a) Ancylothrix rivularis WZU 0009, 0011, 0013, 0014; (b) Ancylothrix rivularis 7PC, 8PC, 9PC; (c) Ancylothrix terrestris 10PC, 13PC; (d) Ancylothrix terrestris WZU 0153, 0154; (e) Ancylothrix terrestris 11PC; (f) Ancylothrix terrestris 12PC.
Figure 7. Box–B helix of Ancylothrix strains: (a) Ancylothrix rivularis WZU 0009, 0011, 0013, 0014; (b) Ancylothrix rivularis 7PC, 8PC, 9PC; (c) Ancylothrix terrestris 10PC, 13PC; (d) Ancylothrix terrestris WZU 0153, 0154; (e) Ancylothrix terrestris 11PC; (f) Ancylothrix terrestris 12PC.
Diversity 14 00301 g007
Figure 8. V3 helix of Ancylothrix strains: (a) Ancylothrix rivularis WZU 0009; (b) Ancylothrix rivularis WZU 0011, 0013, 0014 and Ancylothrix rivularis 7PC, 8PC, 9PC; (c) Ancylothrix terrestris 10PC; (d) Ancylothrix terrestris 13PC; (e) Ancylothrix terrestris WZU 0153, 0154; (f) Ancylothrix terrestris 11PC; (g) Ancylothrix terrestris 12PC.
Figure 8. V3 helix of Ancylothrix strains: (a) Ancylothrix rivularis WZU 0009; (b) Ancylothrix rivularis WZU 0011, 0013, 0014 and Ancylothrix rivularis 7PC, 8PC, 9PC; (c) Ancylothrix terrestris 10PC; (d) Ancylothrix terrestris 13PC; (e) Ancylothrix terrestris WZU 0153, 0154; (f) Ancylothrix terrestris 11PC; (g) Ancylothrix terrestris 12PC.
Diversity 14 00301 g008
Figure 9. D1–D1’ helix of Potamolinea aerugineocaerulea: (a) Potamolinea aerugineocaerulea 1PC, 2PC; (b) Potamolinea aerugineocaerulea 38PC; (c) Potamolinea aerugineocaerulea WZU strains. Box–B helix of Potamolinea aerugineocaerulea: (d) Potamolinea aerugineocaerulea 1PC; (e) Potamolinea aerugineocaerulea 2PC; (f) Potamolinea aerugineocaerulea 38PC; (g) Potamolinea aerugineocaerulea WZU strains. V3 helix of Potamolinea aerugineocaerulea: (h) Potamolinea aerugineocaerulea 1PC; (i) Potamolinea aerugineocaerulea 2PC; (j) Potamolinea aerugineocaerulea 38PC; (k) Potamolinea aerugineocaerulea WZU strains.
Figure 9. D1–D1’ helix of Potamolinea aerugineocaerulea: (a) Potamolinea aerugineocaerulea 1PC, 2PC; (b) Potamolinea aerugineocaerulea 38PC; (c) Potamolinea aerugineocaerulea WZU strains. Box–B helix of Potamolinea aerugineocaerulea: (d) Potamolinea aerugineocaerulea 1PC; (e) Potamolinea aerugineocaerulea 2PC; (f) Potamolinea aerugineocaerulea 38PC; (g) Potamolinea aerugineocaerulea WZU strains. V3 helix of Potamolinea aerugineocaerulea: (h) Potamolinea aerugineocaerulea 1PC; (i) Potamolinea aerugineocaerulea 2PC; (j) Potamolinea aerugineocaerulea 38PC; (k) Potamolinea aerugineocaerulea WZU strains.
Diversity 14 00301 g009
Table 1. Locality and habitat where the cyanobacterial strains in this study were collected.
Table 1. Locality and habitat where the cyanobacterial strains in this study were collected.
StrainsHabitatLocalityLatitude (N)Longitude (E)
WZU 0009, 0011, 0013, 0014BenthosFeiyun River Basin27°66′82″120°05′64″
WZU 0153, 0154SoilDahongyan Scenic Area28°81′28″119°66′22″
WZU 0155-0168BenthosNiutou Mountain National Forest Park28°66′76″119°49′78″
Table 2. Analyses on ITS of 16S–23S region for Ancylothrix strains.
Table 2. Analyses on ITS of 16S–23S region for Ancylothrix strains.
StrainGenBankComplete ITS (nt)D1–D1′
Helix (nt)
tRNAIletRNAAlaBox–B
Helix (nt)
Box–A
Helix (nt)
D4V3
Helix (nt)
D5
WZU 0009OL74257544157++391274319
WZU 0011OL74257244157++391274319
WZU 0013OL74257344559++391274319
WZU 0014OL74257444157++401274319
7PCKT81919643757++401274319
8PCKT81919743857++401274319
9PCKT81919843656++401274319
10PCKT81919951554++381275422
11PCKT81920036463+-391271217
12PCKT81920140966+-331184416
13PCKT81920251654++381275421
WZU 0153OM23745351154++361275422
WZU 0154OM23745451154++361275422
Table 3. Analyses on ITS of 16S–23S region for Potamolinea aerugineocaerulea strains.
Table 3. Analyses on ITS of 16S–23S region for Potamolinea aerugineocaerulea strains.
StrainsGenBankComplete ITS (nt)D1–D1′
Helix (nt)
tRNAIletRNAAlaBox–B
Helix (nt)
Box–A
Helix (nt)
D4V3
Helix (nt)
D5
1PCKX00178637263+491273022
2PCKX00178737263+491273022
38PCKX00178837263+481273022
WZU 0155OM23743939461+491275422
WZU 0156OM23744039461+491275422
WZU 0157OM23744139461+491275422
WZU 0158OM23744239461+491275422
WZU 0159OM23744339461+491275422
WZU 0160OM23744439461+491275422
WZU 0161OM23744539461+491185422
WZU 0162OM23744639461+491275422
WZU 0163OM23744739461+491275422
WZU 0164OM23744839461+491275422
WZU 0165OM23744939461+491275422
WZU 0166OM23745039461+491275422
WZU 0167OM23745139461+491275422
WZU 0168OM23745239461+491275422
Table 4. Comparison of the 16S rRNA gene sequence similarity among all Ancylothrix strains.
Table 4. Comparison of the 16S rRNA gene sequence similarity among all Ancylothrix strains.
Strains123456789101112
1. WZU 0009
2. WZU 001199.74
3. WZU 001399.6699.57
4. WZU 001499.9199.8399.74
5. 7PC99.6699.5799.6699.74
6. 8PC99.5799.4899.5799.6699.91
7. 9PC99.7499.6699.7499.8399.9199.83
8. 10PC95.9495.8496.1396.0396.1296.0396.39
9. 11PC96.0395.9496.1396.1296.2296.1296.39100.00
10. 12PC95.8595.7695.8795.9496.0495.1496.1399.5799.65
11. 13PC95.5595.4695.7995.6495.7495.6596.0499.4899.4899.22
12. WZU 015395.7495.6695.7595.8496.1396.0396.0399.8399.8399.3999.31
13. WZU 015495.9395.8495.7596.0396.1296.0396.03100.00100.0099.5799.4899.83
Table 5. Comparison of the 16S rRNA gene sequence similarity among all Potamolinea aerugineocaerulea strains.
Table 5. Comparison of the 16S rRNA gene sequence similarity among all Potamolinea aerugineocaerulea strains.
Strains12345678910111213141516
1. P. aerugineocaerulea 1PC
2. P. aerugineocaerulea 2PC99.93
3. P. aerugineocaerulea 38PC99.4699.52
4. WZU 015598.4198.4898.69
5. WZU 015698.4198.4898.69100
6. WZU 015798.4198.4898.69100100
7. WZU 015898.4198.4898.69100100100
8. WZU 015998.4198.4898.69100100100100
9. WZU 016098.4198.4898.69100100100100100
10. WZU 016198.4198.4898..69100100100100100100
11. WZU 016298.4198.4898.69100100100100100100100
12. WZU 016398.4198.4898.69100100100100100100100100
13. WZU 016498.4198.4898.69100100100100100100100100100
14. WZU 016598.2098.2798.4999.5399.5399.5399.5399.53100100100100100
15. WZU 016698.1398.2098.4299.4699.4699.4699.4699.4699.4699.4699.4699.4699.4699.93
16. WZU 016798.2098.2798.4999.5399.5399.5399.5399.5399.5399.5399.5399.5399.5310099.93
17. WZU 016898.4198.4898.6910010010010010010010010010010099.5399.4699.53
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Cheng, Y.; Geng, R.; Shan, L.; Liu, Y.; Zhang, H.; Xiao, P.; Ma, Z.; Li, R. Taxonomic Discussion on Cyanobacterial Systematics at Family Level, with Special Regards to Phormidiaceae by Using the Strains of Chinese Newly Recorded Genera Ancylothrix and Potamolinea. Diversity 2022, 14, 301. https://0-doi-org.brum.beds.ac.uk/10.3390/d14040301

AMA Style

Cheng Y, Geng R, Shan L, Liu Y, Zhang H, Xiao P, Ma Z, Li R. Taxonomic Discussion on Cyanobacterial Systematics at Family Level, with Special Regards to Phormidiaceae by Using the Strains of Chinese Newly Recorded Genera Ancylothrix and Potamolinea. Diversity. 2022; 14(4):301. https://0-doi-org.brum.beds.ac.uk/10.3390/d14040301

Chicago/Turabian Style

Cheng, Yao, Ruozhen Geng, Liang Shan, Yang Liu, He Zhang, Peng Xiao, Zengling Ma, and Renhui Li. 2022. "Taxonomic Discussion on Cyanobacterial Systematics at Family Level, with Special Regards to Phormidiaceae by Using the Strains of Chinese Newly Recorded Genera Ancylothrix and Potamolinea" Diversity 14, no. 4: 301. https://0-doi-org.brum.beds.ac.uk/10.3390/d14040301

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