Next Article in Journal
Accumulation of Metals in the Environment and Grazing Livestock near A Mongolian Mining Area
Next Article in Special Issue
A Review of Common Cyanotoxins and Their Effects on Fish
Previous Article in Journal
Disposition of Aerosols of Isothiazolinone-Biocides: BIT, MIT and OIT
Previous Article in Special Issue
(Eco)Toxicology of Cyanobacteria and Cyanotoxins: From Environmental Dynamics to Adverse Effects
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Genetic Diversity of Microcystin Producers (Cyanobacteria) and Microcystin Congeners in Aquatic Resources across Africa: A Review Paper

by
Mathias Ahii Chia
1,*,
Ilu Ameh
2,3,
Korie Chibuike George
2,3,
Emmanuel Oluwadare Balogun
2,
Suwebat Ayanronke Akinyemi
4 and
Adriana Sturion Lorenzi
5
1
Department of Botany, Ahmadu Bello University, Zaria 810107, Nigeria
2
Department of Biochemistry, Ahmadu Bello University, Zaria 810107, Nigeria
3
African Centre of Excellence for Neglected Tropical Diseases and Forensic Biotechnology, Ahmadu Bello University, Zaria 810107, Nigeria
4
Department of Plant Biology, Osun State University, Osogbo 210001, Nigeria
5
Department of Cellular Biology, Institute of Biological Sciences, University of Brasília—UnB, Brasília 70910-900, Brazil
*
Author to whom correspondence should be addressed.
Toxics 2022, 10(12), 772; https://doi.org/10.3390/toxics10120772
Submission received: 29 October 2022 / Revised: 30 November 2022 / Accepted: 6 December 2022 / Published: 10 December 2022

Abstract

:
Microcystins are produced by multifaceted organisms called cyanobacteria, which are integral to Africa’s freshwater environments. The excessive proliferation of cyanobacteria caused by rising temperature and eutrophication leads to the production and release of copious amounts of microcystins, requiring critical management and control approaches to prevent the adverse environmental and public health problems associated with these bioactive metabolites. Despite hypotheses reported to explain the phylogeography and mechanisms responsible for cyanobacterial blooms in aquatic water bodies, many aspects are scarcely understood in Africa due to the paucity of investigations and lack of uniformity of experimental methods. Due to a lack of information and large-scale studies, cyanobacteria occurrence and genetic diversity are seldom reported in African aquatic ecosystems. This review covers the diversity and geographical distribution of potential microcystin-producing and non-microcystin-producing cyanobacterial taxa in Africa. Molecular analyses using housekeeping genes (e.g., 16S rRNA, ITS, rpoC1, etc.) revealed significant sequence divergence across several cyanobacterial strains from East, North, West, and South Africa, but the lack of uniformity in molecular markers employed made continent-wise phylogenetic comparisons impossible. Planktothrix agardhii, Microcystis aeruginosa, and Cylindrospermopsis raciborskii (presently known as Raphidiopsis raciborskii) were the most commonly reported genera. Potential microcystin (MCs)-producing cyanobacteria were detected using mcy genes, and several microcystin congeners were recorded. Studying cyanobacteria species from the African continent is urgent to effectively safeguard public and environmental health because more than 80% of the continent has no data on these important microorganisms and their bioactive secondary metabolites.

1. Introduction

The 3.5 billion-year-old phylum of photosynthetic prokaryotes known as cyanobacteria can be found in various habitats, such as desert rocks, freshwater bodies, hydrothermal vents, marine ecosystems, and Antarctic lakes [1]. Eutrophication, increased solar irradiation, and high temperatures from global warming promote the excessive proliferation of harmful cyanobacteria in aquatic systems [2]. Some cyanobacterial strains synthesize bioactive secondary metabolites (e.g., cyanotoxins), which could adversely affect human and ecosystem health [3]. For example, animal and human mortality have been recorded following exposure to significant doses of some cyanotoxins [4]. Cyanotoxins vary in toxicity and chemical structure, falling into three major chemical structures: lipopolysaccharides, cyclic peptides, and alkaloids [5].
There is a lack of information on the occurrence of cyanobacteria blooms in African aquatic resources. A UNESCO survey found that, except in Ethiopia, Kenya, Morocco, and Southern Africa, knowledge of the events of cyanobacterial blooms in Africa is almost non-existent [6,7]. In addition, very little has been reported about the underlining morphology, genetic characterization, and geographical origin among cyanobacterial strains across Africa. This knowledge gap limits our understanding of cyanobacteria distribution, diversity, toxin production, and the nature of their congeners, emphasizing the risk of exposure via contamination of potable water resources. To comprehend cyanobacteria’s phylogeny, geographic origin, and distribution in Africa, an evolutionary reconstruction of previous relationships using genetic material (DNA and RNA) or protein sequences, is necessary to close this information gap. Numerous phylogenetic-based scientific hypotheses contend that speciation, extinction, and dispersal can account for the development of biogeographic patterns [8].
In freshwater and brackish waterbodies worldwide, different cyanobacteria synthesize a class of small bioactive peptides known as microcystins [9]. Microcystins are cyclic heptapeptides with the same unusual amino acid known as ADDA (3-amino-9methoxy-2,6,8-trimethyl-10-phenyldeca-4,6-dienoic acid) that forms part of the general cyclo-(-d-Ala-l-X-d-MeAsp-l-Z-Adda-d-Glu-Mdha) structure (Figure 1). The d-MeAsp is d-erythro-b-methylaspartic acid, Mdha is N-methyldehydroalanine, and X and Z are variable l-amino acids [10]. Changes in the variable amino acids X and Z result in over 300 structural microcystin congeners (Figure 2) [11]. Microcystin congeners are named using the one-letter abbreviation for the amino acids substituted at the X and Z positions, respectively (Figure 1). This can be seen on the depicted analogs/congeners (A) Microcystin-AR, (B) Microcystin-LF, (C) Microcystin-LR, (D) Microcystin –LY (Figure 2), and on any of the identified congeners. Despite their environmental importance, microcystin distribution, isoforms, and levels have been scantily investigated in African water bodies [12]. This review aims to highlight existing information regarding cyanobacteria’s morphological, genetic, and chemical diversity in water bodies across the East, West, North, and South of Africa. The reviewed data cover field samples and isolated strains morphologically examined and subjected to a DNA-based molecular analysis to reveal genetic variation and geographical divergences or relatedness of cyanobacteria. In addition, we provide a compilation of microcystin detection (molecular and chemical methods) and detected concentrations in African water bodies as a platform for future studies.

2. Toxicosis of Microcystins in Africa

The bioactivity of microcystins is based on the inhibition of the catalytic subunits on serine-threonine protein phosphatases 1 and 2A due to the covalent binding of microcystins to these enzymes [14,15,16], which causes the hyper-phosphorylation of cytoskeleton-associated proteins in hepatocytes. This rapid breakdown of the sinusoidal architecture and cell attachment is followed by a disruption of cytoskeletal compartments and a subsequent rearrangement of filamentous actin [17,18]. Low doses of exposure to microcystin cause chronic inflammation, focal degeneration of hepatocytes, and metabolite accumulation, which may lead to progressive kidney and liver failure. Higher doses of microcystin result in immediate intrahepatic hemorrhage, liver necrosis, and hypovolemic shock. Other effects of microcystin exposure include alterations in mitochondrial function, alterations in intracellular calcium levels, free radical formation, and oxidative stress, all of which could contribute to apoptosis and hepatotoxicity. In addition to being hepatotoxic, microcystins also harm exposed organisms’ ability to reproduce and develop, and induce nephrotoxicity, cardiovascular disease, immunomodulation, endocrine disturbance, and gastrointestinal issues [19].
Although there are no documented human deaths due to microcystin poisoning in Africa, direct contact with high concentrations of microcystins has been reported to cause adverse reactions. These include eye and skin irritations, gastrointestinal distress, nausea, and human liver toxicity [20]. Hepatotoxicosis has been reported in domestic animals after exposure to microcystins [21]. Microcystins partly caused the mass mortalities of Lesser Flamingos in Tanzania and Kenya. This conclusion was made after a histopathological analysis of liver tissues [22,23]. Furthermore, several other deaths of wild animals, such as rhinoceroses, zebras, and wildebeests have been associated with a daily intake of MC-LR ranging from 754.29 to 2566.6 µg Kg−1 day−1 in South Africa [24].

3. The Taxonomic Diversity of African Cyanobacteria Using Genetic Markers

Most research on the genetic diversity of cyanobacteria in Africa has focused on toxin-producing versus non-toxin-producing strains that occur simultaneously or consecutively throughout bloom seasons [25,26]. In addition, the identification of cyanobacteria in Africa has traditionally been made microscopically based on morphological differences between field and isolated species/strains. A compilation of geographic and molecular information from cyanobacterial studies in Africa is summarized in Table 1.
Advances in molecular biology have facilitated the use of polyphasic approaches involving chemotaxonomic, phenotypic, and genotypic data to characterize cyanobacteria. Genotypic diversity through PCR (polymerase chain reaction) amplification of target sequences, cloning, and DNA sequencing of isolated strains and field samples has been poorly described in Africa [27,28,29]. The most commonly used genetic markers in Africa include 16S rRNA, PC-IGS, 16S-23S ITS1-L, 16S-23S ITS1-S, rpoB, rpoC1, and 16S-23S ITS. These molecular markers have been employed to understand the taxonomy and phylogeography of cyanobacteria worldwide [8,30].
Table 1. Compilation of geographic and molecular information from African cyanobacterial studies.
Table 1. Compilation of geographic and molecular information from African cyanobacterial studies.
Sub-RegionCountry/Sampling SitesIdentified/Isolated
Species
Number of Strains/Sequences AnalyzedMethod of IdentificationTaxonomic Genetic Marker and/or Molecular TechniqueMicrocystin Detection—Molecular Markers Microcystin Detection—Chemical AnalysisLocation of Closer Related spp. /Genotype/StrainReference
North AfricaWadi El-Natrun, EgyptSpirulina, Oscillatoria, MicrocystisNASEM, Light microscopy16S rDNA, SDS-PAGE28 aNANA[28]
TunisiaMicrocystis aeruginosa, Raphidiopsis raciborskii, Planktothrix agardhii27NA16S–23S rRNA ITS, rpoC1NANASouth Africa, Sweden, and USA strains.
Toxic Micro-HJ-02 strain
clustered with a French strain
[31]
BirM’cherga reservoir,
Tunisia
Raphidiopsis raciborskii4NA16S rRNA, rpoC1mcyA, mcyEHPLC and MALDI-TOFDistinct
from other African strains
[25]
East AfricaLake Bogori, KenyaArthrospira fusiformis, Anabaenopsis, Synechococcus, Arthrospira, Spirulina, Oscillatoria, LeptolyngbyaNALight microscopyMultilocus 454- sequencing (16S rRNA and (cpcBA-IGS, MetagenomicsNANANA[29,32]
Kenya, Tanzania, UgandaArthrospira18Width of
trichome, types of helix, and apical shape of
trichomes
16S-23S ITS (31 *), cpcBA-IGS (23 *)NANAIndia, Asia[33]
Kenya, UgandaMicrocystis
aeruginosa
17 (8)Light microscopyPC-IGS, ITS1 rDNA4 (3 MC producing)NAAustralia, Brazil[34]
UgandaRaphidiopsis raciborskii(44 **)Morphological traitsITS1, PC-IGS, nifH, rpoC1NANAGermany[35]
Sediments and plankton of saline-alkaline and freshwater lakes of KenyaAnabaenopsis, Umezakia, Arthrospira sp., Chroococcidiopsis sp.Nucleotide sequences (8) from DGGE bands: Nostocales (5), Pleurocapsales (1), Oscillatoriales (2)BLAST
search analysis
PCR-based denaturing gradient gel electrophoresis (DGGE) analysis of 16S rRNA gene fragmentsmcyE and ndaFNAKenya[33]
Ugandan freshwatersAnabaena,
Aphanocapsa, Chroococcus, Merismopedia, Microcystis, Planktolyngbia, and
Pseudanabaena
NAMorphological keysPC-IGSmcyB and mcyEHPLC-DAD, MALDI-TOF MS, Q-TOFUganda[36]
West AfricaRivers, lagoons, and coastal waters, NigeriaMicrocystis aeruginosa,
Planktothrix aghardii,
Anabaena sp.
NALight microscopyNAmcyEAbraxis microcystins /Nodularin Adda Kit (Abraxis Inc., Warminster, PA, USA)Southern Germany, Korea, Japan[12]
Freshwater lakes and reservoir, SenegalRaphidiopsis raciborskii, Raphidiopsis africana
Raphidiopsis sp.
18Light microscopy, the ability to differentiate
heterocysts when grown in a nitrogen-free medium
16S rRNA, ITS1-S, ITS1-L, rpoCl, nifH [10 *]NANAAmerica, Europe[27]
Southern AfricaHartbeespoort Dam, South AfricaSphaerospermopsis reniformis, Sphaerospermopsis aphanizomenoides,
Raphidiopsis curvispora, Raphidiopsis curvata, Raphidiopsis mediterrranea, Microcystis aeruginosa
27 (1 MC producing)Light microscopyNAmcyE
41 b (35) MC (1 d; MC-AnaR)
NATropical and
subtropical regions of Africa
[26]
Crusts Karoo and Nama
Karoo, South Africa
Consortia of cyanobacteria45Morphological features16S rRNA (140 *)103 a (11 d)NANA[37]
South AfricaMicrocystis aeruginosa23Light microscopyAFLPsNANAJapan, Europe, North America[38]
Theewaterskloof Dam, South AfricaAnabaena
ucrainica
NASEM16S rRNANA NANA[39]
Phakalane ponds, BotswanaMicrocystis novacekiiNAMicroscopic identification
BLAST
search analysis
16S rRNAmcyA, -B, -C, -D, -E and –GLC-ESI-MS and solid-phase extraction (SPE) stepsBotswana[40]
NA: not applicable; a Strain; b Variant; d Novel variant; MC: Microcystis; *: reference sequences; **: reference strains.

3.1. 16S Ribosomal RNA and 16S–23S rDNA Internal Transcribed Spacer (ITS)

The small subunit ribosomal RNA molecules of ribosomes, which are encoded by the 16S rRNA gene, are necessary for the process of translating genetic instructions into functional cell components via the translation of mRNA to proteins. Because it is a part of all self-replicating systems, ribosomal RNA is very easy to extract, and its slow rate of sequence change over time makes it possible to identify relationships between quite different species [41]. The 16S rRNA gene’s sequence was discovered to contain both conserved and changeable sections through a sequence analysis and structural modeling. Variable nucleotides appear at the surface of the ribosome, where it is hypothesized that nucleotide substitutions do not alter ribosome activity, in contrast to evolutionarily conserved nucleotides that appear toward the center of the ribosome at functional regions [42].
In this review, we retrieved 16S rRNA and ITS1 (Figure 3 and Figure 4) sequences deposited in the National Center for Biotechnology Information (NCBI) database that had a study or data on African cyanobacteria strains to generate phylogenetic trees using the NJ (neighbor-joining) algorithm: a widely used method for a long time for constructing phylogenetic trees. NJ is a greedy algorithm based on the distance between sequences that endeavors to minimize the sum of all the branch lengths of the resulting tree. It is out of the scope of our paper to cover the limitations or comparisons of methods employed for a DNA sequence analysis. Moreover, all the figures (Figure 3 and Figure 4) contain data from annotated sequences of cyanobacteria blooms and isolated strains in Africa and other continents in the NCBI database. There was a tendency for 16S rRNA sequences of isolated strains from the same country to group independently of the cyanobacterial genera (Figure 3). Furthermore, cyanobacteria samples collected from Kenya and Algeria formed subclusters with cultured Microcystis aeruginosa and uncultured cyanobacteria from Israel and China, respectively (Figure 3). Another subcluster was formed by cyanobacteria isolated from Tunisia, South Africa, Kenya, and Japan.
The combination of molecular markers increases the accuracy of the taxonomic classification of cyanobacteria. Fathalli et al. [25] reported using 16S rRNA and rpoC1 molecular markers to characterize four Raphidiopsis raciborskii strains isolated from Tunisia. The rpoC1 gene sequence was used to determine the phylogenetic connections among the strains, which were taxonomically identified using 16S rRNA sequences. The composite molecular markers revealed that Tunisian strains established a unique clade from other African strains included in the phylogenetic analysis. This result demonstrated that rpoC1 sequences can discriminate between close taxa—unfortunately, only a few studies have used this molecular marker in Africa. This strongly corroborates with previous findings showing that the rpoC1 gene sequence is extra selective at the species level compared to the 16S rRNA sequence [43]. Furthermore, the maximum likelihood (ML) phylogenetic tree of R. raciborskii worldwide strains inferred from combined sequences of 16S rRNA, ITS-L, ITS-S, and rpoC1 encompassing~2927 bp grouped into three well-supported distinct clusters: the European (I), African/American (II), and Asian/Australian (III) strains [31]. The African strains consisted of Tunisian strains. In a review article, Padisák [44] postulated that R. raciborskii originated in Africa. Over the years, it has spread to equatorial regions such as Indonesia and Central America, with a second radiation center in Australia, accounting for the more recent cyanobacteria invasion in the tropical, subtropical, and temperate regions. The distinct clustering of African–Australian R. raciborskii strains supports this evidence [27,45].
Using the Denaturing Gradient Gel Electrophoresis (DGGE) technique coupled with the 16S–23S rDNA internal transcribed spacer (ITS), van Gremberghe et al. [46] analyzed the local and regional genetic diversity as well as the structure of Microcystis populations in 30 man-made reservoirs from Tigray (Northern Ethiopia) in wet and dry seasons. Both regional and local Microcystis ITS diversity (number of distinct ITS types in the DGGE profile per sample) were relatively low, with several dense blooms containing only a single ITS type (EG07). In addition, the authors observed a geographic signature and a limited influence of local environmental conditions on the Microcystis population structure, represented by the presence or absence of ITS types in the DGGE profile per sample and their relative abundances. Members of the genus Microcystis are the most investigated cyanobacteria worldwide, and Africa is no exception. Research by Moreira et al. [47] suggested that Microcystis aeruginosa originated on the African continent, with further Europe-mediated global dispersal of the species.

3.2. Amplified Fragment Length Polymorphism

Amplified fragment length polymorphism (AFLP), a PCR-based fingerprinting technique, and SDS-PAGE (sodium dodecyl sulfate-polyacrylamide gel electrophoresis), a protein profile analysis technique, have been used to characterize cyanobacteria in Africa molecularly. The SDS-PAGE protein profile was used for the first time to analyze the genus Merismopedia in Egypt (North Africa) [48]. Oberholster et al. [38] employed AFLP fingerprinting to highlight the differences among 23 strains of geographically unrelated Microcystis isolated from South Africa, and the UPGMA (unweighted pair group method with arithmetic mean) hierarchical clustering method revealed four groups. The NIES strains of Japanese origin clustered in Group 1; the European strains clustered in Group 2; the South African strains (northern SA) formed the third distinct group; and the strains collected from the central and southern regions of SA clustered together with the US strains in Group 4. Their results revealed extensive evidence of the applicability of AFLP in cyanobacterial characterization, validating its discriminative power towards the differentiation of unrelated Microcystis strains of the same species (M. aeruginosa), as well as highlighting the potential of genomic restriction fingerprinting in evolutionary studies.

3.3. Omics—Metagenomics for Characterization of Cyanobacteria in Africa

Metagenomic studies of cyanobacteria are scarcely conducted in Africa. Using multilocus 454-amplicons sequencing, Dadheech et al. [32] used metagenomics to investigate the cyanobacterial community structure of Lake Bogoria, Kenya. A high degree of endemism was found for Oscillatoriales phylotypes (Leptolyngbya, Spirulina, Oscillatoria-like, and Planktothricoides) in the Bogoria hot springs. Different clades of Synechococcus represented Chroococcales. The pelagic zone and the sediments were inhabited by only a few taxa, dominated by Arthrospira and Anabaenopsis. Arthrospira, the main food for Lesser Flamingos, was detected in all three habitats (hot springs, pelagic zone, and sediment), indicating its resilience and critical role as a primary producer. We look forward to seeing additional metagenomics studies in Africa because taxonomic classification using this approach enables the establishment of the exact origin of species and cataloging or classifying various cyanobacterial groups inhabiting unique environments in Africa.

4. Detection of Potentially African Microcystin-Producing Cyanobacteria by Molecular Markers

As previously stated, microcystins are a family of cyclic peptide toxins produced by cyanobacteria that are widely responsible for the toxicosis and death of wild and domestic animals worldwide due to their hepatotoxic effects. Microcystins are assembled on megadalton enzyme complexes consisting of a combination of nonribosomal peptide synthetases (NRPS), type I polyketide synthases (PKS-I), hybrid NRPS/PKS-I, and tailoring enzymes [9,49,50,51]. The proposed functions of proteins encoded by the microcystins’ biosynthetic gene cluster (mcy) are summarized in Table 2.
A pairwise comparison of mcyE gene sequences of samples collected from coastal areas of Nigeria showed identities higher than 86% (query coverage > 96%) with toxic Microcystis strains [12]. Most Nigerian mcyE sequences formed distinct clusters from those of other countries in Africa and other continents (Figure 5). Additionally, mcyE sequences of samples collected from Kenya and Uganda formed a distinct subcluster from other African samples and those obtained from other continents. PCR amplification and sequencing of the mcyA gene in samples collected from Hartbeespoort Dam, South Africa, revealed the presence of toxin-producing strains of M. aeruginosa in the dam [53].
Five other mcy genes (mcyBCDEG) were also PCR amplified and analyzed during this study. Integrating phylogeny, geographic niche partitioning, and analysis of the mcy gene cluster composition in bloom-forming Planktothrix, Kurmayer et al. [54] analyzed the genotypic dispersal of 138 strains from Europe, Russia, North America, and East Africa. Although the strains had at least remnants of the mcy gene cluster, various phylogenetic lineages might have evolved and adapted to specific ecological niches. Douma et al. [55] reported the presence of potentially toxic Microcystis strains in two Middle Atlas Mountains’ natural lakes in Morocco for the first time. A whole-cell multiplex PCR detected mcy genes, which allowed the simultaneous amplification of DNA sequences corresponding to specific mcy regions.
Fathalli et al. [56] characterized 27 potentially toxic cyanobacteria strains, mainly of M. aeruginosa, C. raciborskii, and P. agardhii, from seven reservoirs in Northern and Central Tunisia. In most Microcystis isolates, six distinct segments of the microcystin synthetase mcy cluster (mcyA, -B, -C, -D, -E, and -G) were detected. The presence of a toxic strain of M. aeruginosa in the upper littoral zone of Lake Krugersdrift, South Africa, was accessed by detecting the mcyB and mcyD genes and confirmed through the use of ELISA [57]. This information is indexed in PubMed® from the National Center for Biotechnology Information (NCBI) database and represents the toxic cyanobacteria molecular studies conducted in Africa. Further studies from the African continent are summarized in Table 1, indicating that toxin-producing cyanobacteria in African waters are of a public and environmental health concern. Several studies could not be included in the phylogenetic tree for mcyE (Figure 5) because the sequence accession numbers were not provided.

5. Cyanobacterial Abundance and Microcystin Congeners Detected in African Waters

The abundance of cyanobacteria, microcystin congeners, and detected concentrations in several countries of the African continent are presented in Table 3. Here, we revised important aspects of cyanobacterial abundance and microcystin congeners recorded in African waters per region (East Africa, North Africa, South Africa, and West Africa, respectively).

5.1. East Africa

Using a polyphasic approach to characterize M. aeruginosa isolates from East African aquatic ecosystems, Haande et al. [34] reported ten chemotypes by a MALDI-TOF-MS oligopeptide analysis. Interestingly, only about 17% of the isolated strains produced microcystins, while most strains were found to produce the protease inhibitor oligopeptides aeruginosin and cyanopeptolin in freshwater and saline lakes.
The phytoplankton composition of Ugandan freshwater lakes is similar to that found in other East African lakes, where it is common to find cyanobacteria as the dominant group. Although Ugandan lakes have different trophic conditions ranging from mesotrophic to hypertrophic, these conditions, coupled with the warm climate of East Africa, facilitate the excessive proliferation of cyanobacteria. Anabaena spp. and Microcystis spp. are the most frequent and dominant species found in eutrophic water bodies, with total microcystins concentrations ranging from 0.03 to 144 fg cell−1 and 0.02 to 10 μg MC-LR eq./L [36,86,87,88] (Table 3). The key congeners of microcystins detected in Ugandan aquatic systems are MC-RR, [Asp3]-MC-RY, and [MeAsp3]-MC-RY. To date, Lake Saka in Uganda is the only water body with microcystins higher than 60 μg L−1 [88].
As a fast-growing economy, Ethiopia generates many domestic and industrial effluents that are changing the physical-chemical conditions and phytoplankton diversity of aquatic ecosystems in the country. Tracking seasonal changes in phytoplankton diversities has revealed the dominance of Microcystis aeruginosa, M. panniformis, Anabaena spiroides, and Cylindrospermopsis spp. in Ethiopian lakes, which highly correlates with the presence of microcystins (0.58–1547.28 μgL−1) [46,73,75,76,116]. The peak concentrations of microcystins in these water bodies are detected during the dry season around November to December of every year, and the main microcystins variants found in lentic ecosystems include MC-LR, MC-YR, MC-RR, MC-dmLR, and MC-LA (Table 3).

5.2. North Africa

Studies on cyanobacterial abundance and toxicity in the coastal regions are generally scarce in Africa, making the study by Douma et al. [70] very important. From 2004 to 2013, the authors investigated the diversity of cyanobacteria and the presence of microcystins in a Moroccan coastal lagoon, the Sidi Boughaba, and found that only Microcystis flos-aquae produced the microcystin variants MC-LR, MC-RR, and MC-WR (Table 3). Highly toxic microcystin-producing floating and benthic mats of Nostoc muscorum are formed from March to October yearly when temperatures are above 10 °C in the Oukaïmeden River [117]. The river is situated in the High-Atlas Mountains of Marrakech, Morocco, representing a critical environment in the country. Characterization of the blooms indicates the production of three microcystin variants, including MC-LR (Table 3).
A high diversity of microcystin variants has been recorded in blooms occurring in Algerian aquatic ecosystems (Table 3). The characterization of Microcystis spp. blooms from the Lake des Oiseaux shows the highest diversity of microcystins reported from any African aquatic ecosystem [58]. Twelve (12) out of twenty-one (21) putative congeners are fully characterized (MC-RR, MC-LR, MC-FR, MC-WR, MC-YR, MC-LA, MC-(H4)YR, MC-HilR, [Asp3]MC- rAba, [Glu(OCH3)6]MC-LR), [Asp3]MC-HarAba, and [Glu(OCH3)6]MC-FR) (Table 3). Among these congeners, the [Asp3]MC-HarAba and [Glu(OCH3)6]MC-FR are novel congeners found in Algeria. In addition to the high diversity of microcystin variants, Algerian water bodies present some of the highest microcystins found in African aquatic ecosystems (Table 3). Nasri et al. [60], for instance, measured concentrations as high as 26,163 μg L−1 in raw lake water, which was equivalent to 4590 μg g−1 DW (dry weight) bloom material. Detecting high microcystin concentrations associated with cyano-HABs (cyanobacterial harmful algal blooms) has led to several animal deaths in this country [59]. Bloom extract analyzed by Nasri et al. [59] confirmed that the microcystin variants MC-LR, MC-YR, and MC-RR contributed to the deaths of turtles (Emys orbicularis and Mauremys leprosa) at Lake Oubeira, Algeria. Autopsies of these terrapins revealed that the highest MC concentrations were detected in the liver, viscera, and muscle tissues.
Most aquatic ecosystems in Egypt, including the Nile River, irrigation canals, lakes, and fish ponds, are plagued with frequent and intense cyanobacterial harmful blooms (cyanoHABs) [118]. In this country, microcystin production has been primarily associated with Microcystis blooms. However, blooms containing microcystin-producing Osciallatoria spp. reveal microcystin levels ranging from 300 to 877 µg L−1 [62,118,119] (Table 3). As in other parts of Africa, MC-LR and MC-RR are the main variants detected in Egyptian waters, and the level of total microcystins approaches ca. 341 µg L−1 in the Nile River [118]. Moreover, the levels of microcystins in raw and treated water have frequently been found to exceed the 1.00 µg L−1 MC-LR limit set by the WHO [119].

5.3. South Africa

South Africa remains one of the regions with the highest investigations on cyanoHABs. In South Africa, most poisonings have been associated with M. aeruginosa, and the first case of cattle intoxication by Microcystis was reported as far back as 1927 [38]. In the following years, South African cyanobacteria have been characterized by several studies on ecology, toxicology, and water management implications [24,26,38,103,109,120].
Eguzozie et al. [103] demonstrated the diversity of microcystin congeners in floating scums of cyanobacteria at the Swartspruit River, South Africa. The main microcystin congeners detected were MC-YR, MC-(H4)-YR, MC-LR, (D-Asp3, Dha7) MC-RR, and MC-RR. Among the variants, MC-LR had the highest concentration (14.10 to 270.7 μg g−1 DW), while MC-YR had the lowest (0.15 to 72.28 μg g−1 DW).
There are few studies on cyanobacteria and their toxins in Mozambique compared to South Africa [99,100,101]. The country’s poor economic situation is partly responsible for the limited attention given to the presence of microcystins and other cyanobacterial toxins. Few studies have been conducted to determine the presence of microcystins in the public water supply systems of Pequenos Libombos Dam, Nhambavale Lake, and Chòkwé Irrigation Channel of Mozambique. Microcystins were detected in Nhambavale Lake [99,100] and Chòkwé Irrigation Channel [99], reaching ca. 0.10–7.89 μg L−1 and 2.1–159.4 ng g−1 DW, respectively. MC-LR, MC-YR, and MC-RR produced by Microcystis spp. were the most predominant congeners in the studied water bodies.
In Botswana, Phakalane Waste Water Secondary Maturation ponds and the Limpopo River are some water bodies in which microcystins have already been investigated and quantified [40]. MC-RR was highest (53.62 μg g−1 DW), followed by MC-LR (12.114 μg g−1 DW) in the water bodies. These concentrations differ from what has been reported in other parts of Africa, where MC-LR is frequently the most abundant among microcystin congeners, even considering different aquatic ecosystems.

5.4. West Africa

Although comparatively less information exists on the occurrence and distribution of toxic cyanobacterial blooms in West African aquatic ecosystems compared to other parts of Africa, several studies have been conducted to investigate the diversity of potential microcystin producers in the subregion. Kadiri et al. [12] investigated the presence and composition of cyanobacteria coupled with microcystin production in several coastal aquatic ecosystems of Southern Nigeria. Substantial biomasses of Microcystis aeruginosa and Planktothrix aghardii were found in almost all investigated aquatic ecosystems. Furthermore, Anabaena sp. was observed in some of the ecosystems. Microcystin concentrations ranging from 0.19 µg/L in the Eleme River to 7.75 µg/L in Lekki were reported in the study. In 2008, a survey of several aquatic ecosystems in Zaria, Northern Nigeria, recorded potential microcystin producers such as Anabaena spp. and Microcystis spp. [121]. Other cyanobacterial genera detected during the study included Spirulina spp., Gloetrichia spp., and Cylindrospermopsis spp. Although microcystin levels were below detection limits in some of the surveyed water bodies, some had total microcystins’ concentrations reaching 3.8 µg/L [104]. Another study of aquaculture ponds in Zaria found Microcystis spp., Nostoc, Planktothrix, and Anabaena [93]. In this study, total microcystins’ concentrations ranging from 0.6 to 5.89 µg/L were reported in some fishponds.
In Kano, a city in North West Nigeria and the capital of Kano State, microcystin congeners produced by M. aeruginosa isolated from 6 burrow pits were characterized monthly over three years, and four congeners were identified (MC-RR [24%], MC-YR [26%], MC-LR [48%], and MC-LF [2%]) [95]. They found variations in detected microcystin congeners, except for MC-LR, which was found in all samples. Chia and Kwaghe [104] reported variations in the total microcystin content of surface water supply reservoirs in Zaria. In most reservoirs, higher total microcystin concentrations (1.68–3.94 µg/L) were reported during the dry season than in the rainy season. In addition, there was a significant positive correlation between the biomass of Microcystis aeruginosa and Anabaena subcylindrica with particulate and dissolved microcystin concentrations [104].
In Ghana, a study conducted by Addico et al. [89] to identify the cyanobacteria in the Weija and Kpong reservoirs, which are two crucial freshwater sources in the Southern and Eastern parts of the capital Accra, respectively, identified four potentially cyanotoxin-producing genera: Anabaena flos-aquae, Cylindrospermopsis raciborskii, Microcystis aeruginosa, and Planktothrix agardhii. The phytoplankton biomass comprised 70–90% of these four genera at all water treatment stages. A microcystin analysis from both reservoirs revealed the presence of 6 different unidentified variants and MC-RR in the Weija and Kpong reservoirs, respectively. Microcystin concentrations ranged from 0.03 µg/L in the Kpong reservoir to 3.21 µg/L in the Weija reservoir [89].
Addico et al. [122] conducted an additional study to identify the toxins associated with a strain of Anabaena flos-aquae, one of the dominant species found in major Ghanaian freshwater reservoirs. They reported MC-RR as the main microcystin produced, with a concentration of 10.6 µg/g DW. Another study by Addico et al. [90] from water samples of the Brimsu and Kwanyarko reservoirs in the Central Ghana region indicated the presence of the congeners MC-LR, MC-YR, MC-RR, and MC-LA, with concentrations ranging from 0.1 µg/L to 0.79 µg/L at various stages of water treatment. This report was the first MC-LA in Ghana, and M. aeruginosa and Planktothrix agardhii were the major species found in the reservoirs.
The occurrence of potential microcystin producers has been scarcely documented in other West African countries. In South East Cote d’Ivoire, Kouassi et al. [123] conducted monthly investigations on the cyanobacterial flora of Adzope reservoir for a year. The authors reported Planktothrix rubescens and M. aeruginosa as predominant species throughout the study. Although microcystin production from either species was not characterized in the study, this revealed the presence of potentially toxin-producing species in the country.
The absence of studies in other African countries does not mean the lack of a rich diversity of cyanobacteria and their metabolites. A close look at phytoplankton ecological studies in several countries confirmed the presence of potential toxin-producing cyanobacteria and the frequent occurrence of cyanobacterial blooms. This absence of data on microcystin congeners and other cyanotoxins highlights a severe problem requiring research and the development of accessible and cost-effective means of evaluating the risk of cyanobacteria and communicating cyanobacterial-related information in the countries. The major limitation is the lack of skilled personnel and researchers in this critical environmental and public health science area.

6. Relationship between Genetic Characterization and Geographical Origin within Toxic Cyanobacteria Strains across Africa

Although DNA-based molecular techniques help reveal the presence of genes involved in toxin production, only RNA-based methods (e.g., RT-qPCR, reverse transcription polymerase chain reaction, and transcriptomic approaches) allow the detection of potential toxin-producing cyanobacteria that are actively transcribing the toxin genes into a toxin. Thus, the presence of toxin-producing genes only in cyanobacteria strains does not directly translate to a toxicological threat. Furthermore, toxin genes can be mutated, which prevents or diminishes their transcription and subsequent toxin production [124,125].
In a South African study conducted by Ballot et al. [26], the mcyE gene of the microcystin gene cluster was found in a microcystin-producing Microcystis strain (AB2011/53) but also in eight non-microcystin-producing Microcystis strains. Non-toxic cyanobacteria strains may result from significant gene deletion events induced by a transposable element [125]. Surprisingly, an unusual, isolated strain of Cylindrospermopsis raciborskii (CYL-BM-07) from Tunisia was identified as a non-microcystin producer, despite the positive results for two segments (mcyA and mcyE genes) of the microcystin synthetase mcy cluster [56]. This result was further confirmed by HPLC and MALDI-TOF analyses. Various mechanisms, including horizontal gene transfer (HGT), mutations, insertions, and deletions, have been proposed to explain non-toxin-producing cyanobacteria possessing parts of toxin-encoding gene clusters [125], which might have led to the loss of toxin production. As stated previously, a phylogenetic analysis of 16S rRNA, ITS-L, and ITS-S sequences indicated a high genotype similarity among African cyanobacterial strains and field samples compared to those from other regions of the world [31,56]. In addition, Tunisian strains clustered together, and their genetic variability differed considerably from different African strains [56]. However, among the four Tunisian C. raciborskii strains isolated from the Bir M’cherga Dam tested for the presence of the genes encoding for the hepatotoxin microcystin, only Cylindrospermopsis raciborskii strain CYL-BM-07 was positive for mcyA and mcyE genes. A BLAST search showed that the mcyA and mcyE genes of CYL-BM-07 were genetically more similar to homologous sequences found in Microcystis sp. than those found in other cyanobacteria genera [56], which might be associated with a possible HGT (horizontal gene transfer) mechanism. Therefore, phylogeny, genetic diversity, and geographical distribution play a vital role in determining the ecotoxicological risks that potentially microcystin-producing cyanobacteria pose to environmental and animal health and should be carefully evaluated.

7. Conclusions

Molecular evidence of strong genotypic divergences among West, South, East, and North African strains of cyanobacteria were extensively reviewed. These strains possessed distinct genotypes when 16S rRNA and ITS sequences were employed as molecular markers. Despite positive results of microcystin-producing genes (e.g., mcyA, -B, -C, -D, -E, and -G) in African strains, this does not necessarily translate to toxin production. It is, however, pertinent to note that the discussed methods were based mainly on the microscopy and Sanger sequencing of PCR products, which limit the knowledge of African cyanobacteria compared to cyanobacterial studies performed in other parts of the world. Whenever possible, metagenomic, metatranscriptomic, and metabolomic approaches should be employed to investigate cyanobacteria from different African regions.

Author Contributions

M.A.C.: Conceptualization, Writing—Original Draft, Writing—Review and Editing. I.A.: Writing—Original Draft, Writing—Review and Editing. K.C.G.: Writing—Original Draft, Writing—Review and Editing. E.O.B.: Writing—Original Draft, Writing—Review and Editing. S.A.A.: Writing—Original Draft, Writing—Review and Editing. A.S.L.: Writing—Original Draft, Writing—Review and Editing. 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

All reported data are referenced in the text.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Buratti, F.M.; Manganelli, M.; Vichi, S.; Stefanelli, M.; Scardala, S.; Testai, E.; Funari, E. Cyanotoxins: Producing Organisms, Occurrence, Toxicity, Mechanism of Action and Human Health Toxicological Risk Evaluation. Arch. Toxicol. 2017, 91, 1049–1130. [Google Scholar] [CrossRef] [PubMed]
  2. Paerl, H.W.; Huisman, J. Climate Change: A Catalyst for Global Expansion of Harmful Cyanobacterial Blooms. Environ. Microbiol. Rep. 2009, 1, 27–37. [Google Scholar] [CrossRef] [PubMed]
  3. Funari, E.; Testai, E. Human Health Risk Assessment Related to Cyanotoxins Exposure. Crit. Rev. Toxicol. 2008, 38, 97–125. [Google Scholar] [CrossRef] [PubMed]
  4. Puschner, B.; Hoff, B.; Tor, E.R. Diagnosis of Anatoxin-a Poisoning in Dogs from North America. J. Vet. Diagn. Investig. 2008, 20, 89–92. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Metcalf, J.S.; Codd, G.A. Cyanotoxins. In Ecology of Cyanobacteria II; Springer: Berlin/Heidelberg, Germany, 2012; pp. 651–675. [Google Scholar]
  6. Codd, G.A.; Azevedo, S.; Bagchi, M.D.; Burch, M.D.; Carmichael, W.W.; Harding, W.R.; Kaya, K.; Utkilen, H.C. CYANONET: A Global Network for Cyanobacterial Bloom and Toxin Risk Management: Initial Situation Assessment and Recommendations; UNESCO: Paris, France, 2005. [Google Scholar]
  7. Ndlela, L.L.; Oberholster, P.J.; Van Wyk, J.H.; Cheng, P.-H. An Overview of Cyanobacterial Bloom Occurrences and Research in Africa over the Last Decade. Harmful Algae 2016, 60, 11–26. [Google Scholar] [CrossRef]
  8. Moreira, C.; Vasconcelos, V.; Antunes, A. Phylogeny and Biogeography of Cyanobacteria and Their Produced Toxins. Mar. Drugs 2013, 11, 4350–4369. [Google Scholar] [CrossRef] [Green Version]
  9. Fewer, D.P.; Wahlsten, M.; Österholm, J.; Jokela, J.; Rouhiainen, L.; Kaasalainen, U.; Rikkinen, J.; Sivonen, K. The Genetic Basis for O-Acetylation of the Microcystin Toxin in Cyanobacteria. Chem. Biol. 2013, 20, 861–869. [Google Scholar] [CrossRef] [Green Version]
  10. Malik, J.K.; Bharti, V.K.; Rahal, A.; Kumar, D.; Gupta, R.C. Cyanobacterial (Blue-Green Algae) Toxins. In Handbook of Toxicology of Chemical Warfare Agents; Elsevier: Amsterdam, The Netherlands, 2020; pp. 467–478. [Google Scholar]
  11. Roy-Lachapelle, A.; Solliec, M.; Sauvé, S.; Gagnon, C. A Data-Independent Methodology for the Structural Characterization of Microcystins and Anabaenopeptins Leading to the Identification of Four New Congeners. Toxins 2019, 11, 619. [Google Scholar] [CrossRef] [Green Version]
  12. Kadiri, M.O.; Isagba, S.; Ogbebor, J.U.; Omoruyi, O.A.; Unusiotame-Owolagba, T.E.; Lorenzi, A.S.; Bittencourt-Oliveira, M.; Chia, M.A. The Presence of Microcystins in the Coastal Waters of Nigeria, from the Bights of Bonny and Benin, Gulf of Guinea. Environ. Sci. Pollut. Res. 2020, 27, 35284–35293. [Google Scholar] [CrossRef]
  13. Butler, N.; Carlisle, J.C.; Linville, R.; Washburn, B. Microcystins: A Brief Overview of Their Toxicity and Effects, with Special Reference to Fish, Wildlife, and Livestock; Ecotoxicology Program Integrated Risk Assessment Branch; Office of Environmental Health Hazard Assessment California Environmental Protection Agency: Sacramento, CA, USA, 2009; Volume 5. [Google Scholar]
  14. Hastie, C.J.; Borthwick, E.B.; Morrison, L.F.; Codd, G.A.; Cohen, P.T.W. Inhibition of Several Protein Phosphatases by a Non-Covalently Interacting Microcystin and a Novel Cyanobacterial Peptide, Nostocyclin. Biochim. Biophys. Acta (BBA)-Gen. Subj. 2005, 1726, 187–193. [Google Scholar] [CrossRef]
  15. Maynes, J.T.; Luu, H.A.; Cherney, M.M.; Andersen, R.J.; Williams, D.; Holmes, C.F.B.; James, M.N.G. Crystal Structures of Protein Phosphatase-1 Bound to Motuporin and Dihydromicrocystin-LA: Elucidation of the Mechanism of Enzyme Inhibition by Cyanobacterial Toxins. J. Mol. Biol. 2006, 356, 111–120. [Google Scholar] [CrossRef]
  16. Arman, T.; Clarke, J.D. Microcystin Toxicokinetics, Molecular Toxicology, and Pathophysiology in Preclinical Rodent Models and Humans. Toxins 2021, 13, 537. [Google Scholar] [CrossRef]
  17. Humbert, J.F. Toxins of Cyanobacteria. In Handbook of Toxicology of Chemical Warfare Agents; Academic Press: Cambridge, MA, USA, 2009. [Google Scholar] [CrossRef]
  18. Mrdjen, I.; Morse, M.A.; Ruch, R.J.; Knobloch, T.J.; Choudhary, S.; Weghorst, C.M.; Lee, J. Impact of Microcystin-LR on Liver Function Varies by Dose and Sex in Mice. Toxins 2018, 10, 435. [Google Scholar] [CrossRef] [Green Version]
  19. Chen, L.; Giesy, J.P.; Adamovsky, O.; Svirčev, Z.; Meriluoto, J.; Codd, G.A.; Mijovic, B.; Shi, T.; Tuo, X.; Li, S.-C. Challenges of Using Blooms of Microcystis Spp. in Animal Feeds: A Comprehensive Review of Nutritional, Toxicological and Microbial Health Evaluation. Sci. Total Environ. 2021, 764, 142319. [Google Scholar] [CrossRef]
  20. Van Halderen, A.; Harding, W.R.; Wessels, J.C.; Schneider, D.J.; Heine, E.W.P.; Van der Merwe, J.; Fourie, J.M. Cyanobacterial (Blue-Green Algae) Poisoning of Livestock in the Western Cape Province of South Africa. J. S. Afr. Vet. Assoc. 1995, 66, 260–264. [Google Scholar]
  21. Puschner, B.; Moore, C. Cyanobacteria. In Small Animal Toxicology, 3rd. ed; Elsevier Inc.: Amsterdam, The Netherlands, 2012; pp. 533–540. [Google Scholar]
  22. Krienitz, L.; Ballot, A.; Kotut, K.; Wiegand, C.; Pütz, S.; Metcalf, J.S.; Codd, G.A.; Stephan, P. 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]
  23. Miles, C.O.; Sandvik, M.; Nonga, H.E.; Rundberget, T.; Wilkins, A.L.; Rise, F.; Ballot, A. Identification of Microcystins in a Lake Victoria Cyanobacterial Bloom Using LC–MS with Thiol Derivatization. Toxicon 2013, 70, 21–31. [Google Scholar] [CrossRef]
  24. Masango, M.G.; Myburgh, J.G.; Labuschagne, L.; Govender, D.; Bengis, R.G.; Naicker, D. Assessment of Microcystis Bloom Toxicity Associated with Wildlife Mortality in the Kruger National Park, South Africa. J. Wildl. Dis. 2010, 46, 95–102. [Google Scholar] [CrossRef] [Green Version]
  25. Fathalli, A.; Jenhani, A.B.R.; Moreira, C.; Azevedo, J.; Welker, M.; Romdhane, M.; Antunes, A.; Vasconcelos, V. Genetic Variability of the Invasive Cyanobacteria Cylindrospermopsis Raciborskii from Bir M’cherga Reservoir (Tunisia). Arch. Microbiol. 2011, 193, 595–604. [Google Scholar] [CrossRef]
  26. Ballot, A.; Sandvik, M.; Rundberget, T.; Botha, C.J.; Miles, C.O. Diversity of Cyanobacteria and Cyanotoxins in Hartbeespoort Dam, South Africa. Mar. Freshw. Res. 2013, 65, 175–189. [Google Scholar] [CrossRef]
  27. Gugger, M.; Molica, R.; Le Berre, B.; Dufour, P.; Bernard, C.; Humbert, J.-F. Genetic Diversity of Cylindrospermopsis Strains (Cyanobacteria) Isolated from Four Continents. Appl. Environ. Microbiol. 2005, 71, 1097–1100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Hoballah, E.M.; Attallah, A.G.; Abd-El-Aal, S.K. Genetic Diversity of Some New Local Strains of Cyanobacteria Isolated from Wadi El-Natrun, Egypt. Int. J. Acad. Res. 2012, 4, 314–326. [Google Scholar]
  29. Krienitz, L.; Dadheech, P.K.; Fastner, J.; Kotut, K. The Rise of Potentially Toxin Producing Cyanobacteria in Lake Naivasha, Great African Rift Valley, Kenya. Harmful Algae 2013, 27, 42–51. [Google Scholar] [CrossRef]
  30. Konur, O. The Pioneering Research on the Cyanobacterial Photosystems and Photosynthesis. In Handbook of Algal Science, Technology and Medicine; Elsevier: Amsterdam, The Netherlands, 2020; pp. 231–243. [Google Scholar]
  31. Moreira, C.; Fathalli, A.; Vasconcelos, V.; Antunes, A. Genetic Diversity and Structure of the Invasive Toxic Cyanobacterium Cylindrospermopsis Raciborskii. Curr. Microbiol. 2011, 62, 1590–1595. [Google Scholar] [CrossRef] [PubMed]
  32. Dadheech, P.K.; Glöckner, G.; Casper, P.; Kotut, K.; Mazzoni, C.J.; Mbedi, S.; Krienitz, L. Cyanobacterial Diversity in the Hot Spring, Pelagic and Benthic Habitats of a Tropical Soda Lake. FEMS Microbiol. Ecol. 2013, 85, 389–401. [Google Scholar] [CrossRef] [Green Version]
  33. Dadheech, P.K.; Ballot, A.; Casper, P.; Kotut, K.; Novelo, E.; Lemma, B.; Pröschold, T.; Krienitz, L. Phylogenetic Relationship and Divergence among Planktonic Strains of Arthrospira (Oscillatoriales, Cyanobacteria) of African, Asian and American Origin Deduced by 16S–23S ITS and Phycocyanin Operon Sequences. Phycologia 2010, 49, 361–372. [Google Scholar] [CrossRef] [Green Version]
  34. Haande, S.; Ballot, A.; Rohrlack, T.; Fastner, J.; Wiedner, C.; Edvardsen, B. Diversity of Microcystis Aeruginosa Isolates (Chroococcales, Cyanobacteria) from East-African Water Bodies. Arch. Microbiol. 2007, 188, 15–25. [Google Scholar] [CrossRef]
  35. Haande, S.; Rohrlack, T.; Ballot, A.; Røberg, K.; Skulberg, R.; Beck, M.; Wiedner, C. Genetic Characterisation of Cylindrospermopsis Raciborskii (Nostocales, Cyanobacteria) Isolates from Africa and Europe. Harmful Algae 2008, 7, 692–701. [Google Scholar] [CrossRef]
  36. Okello, W.; Kurmayer, R. Seasonal Development of Cyanobacteria and Microcystin Production in Ugandan Freshwater Lakes. Lakes Reserv. Res. Manag. 2011, 16, 123–135. [Google Scholar] [CrossRef]
  37. Dojani, S.; Kauff, F.; Weber, B.; Büdel, B. Genotypic and Phenotypic Diversity of Cyanobacteria in Biological Soil Crusts of the Succulent Karoo and Nama Karoo of Southern Africa. Microb. Ecol. 2014, 67, 286–301. [Google Scholar] [CrossRef]
  38. Oberholster, P.J.; Botha, A.-M.; Cloete, T.E. An Overview of Toxic Freshwater Cyanobacteria in South Africa with Special Reference to Risk, Impact and Detection by Molecular Marker Tools. Biokemistri 2005, 17, 57–71. [Google Scholar]
  39. Oberholster, P.J.; Jappie, S.; Cheng, P.H.; Botha, A.M.; Matthews, M.W. First Report of an Anabaena Bory Strain Containing Microcystin-LR in a Freshwater Body in Africa First Report of an Anabaena Bory Strain Containing Microcystin-LR in a Freshwater Body in Africa. Afr. J. Aquat. Sci. 2015, 40, 37–41. [Google Scholar] [CrossRef]
  40. Mbukwa, E.; Msagati, T.A.; Mamba, B.B.; Boussiba, S.; Wepener, V.; Leu, S.; Kaye, Y. Toxic Microcystis Novacekii T20-3 from Phakalane Ponds, Botswana: PCR Amplifications of Microcystin Synthetase (Mcy) Genes, Extraction and LC-ESI-MS Identification of Microcystins. J. Environ. Anal. Toxicol. 2015, 525–2161. [Google Scholar]
  41. Woese, C.R.; Fox, G.E. Phylogenetic Structure of the Prokaryotic Domain: The Primary Kingdoms. Proc. Natl. Acad. Sci. USA 1977, 74, 5088–5090. [Google Scholar] [CrossRef] [Green Version]
  42. Wuyts, J.; Van de Peer, Y.; De Wachter, R. Distribution of Substitution Rates and Location of Insertion Sites in the Tertiary Structure of Ribosomal RNA. Nucleic Acids Res. 2001, 29, 5017–5028. [Google Scholar] [CrossRef] [Green Version]
  43. Wilson, K.M.; Schembri, M.A.; Baker, P.D.; Saint, C.P. Molecular Characterization of the Toxic Cyanobacterium Cylindrospermopsis Raciborskii and Design of a Species-Specific PCR. Appl. Environ. Microbiol. 2000, 66, 332–338. [Google Scholar] [CrossRef] [Green Version]
  44. Padisak, J. Cylindrospermopsis Raciborskii (Woloszynska) Seenayya et Subba Raju, an Expanding, Highly Adaptive Cyanobacterium: Worldwide Distribution and Review of Its Ecology. Arch. Für Hydrobiol. Suppl. Monogr. Beitrage 1997, 107, 563–593. [Google Scholar]
  45. Neilan, B.A.; Saker, M.L.; Fastner, J.; Törökné, A.; Burns, B.P. Phylogeography of the Invasive Cyanobacterium Cylindrospermopsis Raciborskii. Mol. Ecol. 2003, 12, 133–140. [Google Scholar] [CrossRef]
  46. van Gremberghe, I.; Van der Gucht, K.; Vanormelingen, P.; Asmelash, T.; Dejenie, T.; D’hondt, S.; Declerck, S.; De Meester, L.; Vyverman, W. Genetic Diversity of Microcystis Blooms (Cyanobacteria) in Recently Constructed Reservoirs in Tigray (Northern Ethiopia) Assessed by RDNA ITS. Aquat. Ecol. 2011, 45, 289–306. [Google Scholar] [CrossRef]
  47. Moreira, C.; Spillane, C.; Fathalli, A.; Vasconcelos, V.; Antunes, A. African Origin and Europe-Mediated Global Dispersal of the Cyanobacterium Microcystis Aeruginosa. Curr. Microbiol. 2014, 69, 628–633. [Google Scholar] [CrossRef] [PubMed]
  48. Palinska, K.A.; Liesack, W.; Rhiel, E.; Krumbein, W.E. Phenotype Variability of Identical Genotypes: The Need for a Combined Approach in Cyanobacterial Taxonomy Demonstrated on Merismopedia-like Isolates. Arch. Microbiol. 1996, 166, 224–233. [Google Scholar] [CrossRef] [PubMed]
  49. 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] [PubMed] [Green Version]
  50. Rouhiainen, L.; Vakkilainen, T.; Siemer, B.L.; Buikema, W.; Haselkorn, R.; Sivonen, K. Genes Coding for Hepatotoxic Heptapeptides (Microcystins) in the Cyanobacterium Anabaena Strain 90. Appl. Environ. Microbiol. 2004, 70, 686–692. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Christiansen, G.; Kurmayer, R.; Liu, Q.; Börner, T. Transposons Inactivate Biosynthesis of the Nonribosomal Peptide Microcystin in Naturally Occurring Planktothrix spp. Appl. Environ. Microbiol. 2006, 72, 117–123. [Google Scholar] [CrossRef] [PubMed]
  52. Heck, K.; Alvarenga, D.O.; Shishido, T.K.; Varani, A.M.; Dörr, F.A.; Pinto, E.; Rouhiainen, L.; Jokela, J.; Sivonen, K.; Fiore, M.F. Biosynthesis of Microcystin Hepatotoxins in the Cyanobacterial Genus Fischerella. Toxicon 2018, 141, 43–50. [Google Scholar] [CrossRef] [Green Version]
  53. Mbukwa, E.A.; Msagati, T.A.M.; Mamba, B.B. Quantitative Variations of Intracellular Microcystin-LR,-RR and-YR in Samples Collected from Four Locations in Hartbeespoort Dam in North West Province (South Africa) during the 2010/2011 Summer Season. Int. J. Environ. Res. Public Health 2012, 9, 3484–3505. [Google Scholar] [CrossRef] [Green Version]
  54. Kurmayer, R.; Blom, J.F.; Deng, L.; Pernthaler, J. Integrating Phylogeny, Geographic Niche Partitioning and Secondary Metabolite Synthesis in Bloom-Forming Planktothrix. ISME J. 2015, 9, 909–921. [Google Scholar] [CrossRef]
  55. Douma, M.; Ouahid, Y.; Loudiki, M.; Del Campo, F.F.; Oudra, B. The First Detection of Potentially Toxic Microcystis Strains in Two Middle Atlas Mountains Natural Lakes (Morocco). Environ. Monit. Assess. 2017, 189, 1–9. [Google Scholar] [CrossRef]
  56. Fathalli, A.; Jenhani, A.B.R.; Moreira, C.; Welker, M.; Romdhane, M.; Antunes, A.; Vasconcelos, V. Molecular and Phylogenetic Characterization of Potentially Toxic Cyanobacteria in Tunisian Freshwaters. Syst. Appl. Microbiol. 2011, 34, 303–310. [Google Scholar] [CrossRef]
  57. Oberholster, P.J.; Botha, A.-M.; Ashton, P.J. The Influence of a Toxic Cyanobacterial Bloom and Water Hydrology on Algal Populations and Macroinvertebrate Abundance in the Upper Littoral Zone of Lake Krugersdrift, South Africa. Ecotoxicology 2009, 18, 34–46. [Google Scholar] [CrossRef] [Green Version]
  58. Bouhaddada, R.; Nélieu, S.; Nasri, H.; Delarue, G.; Bouaïcha, N. High Diversity of Microcystins in a Microcystis Bloom from an Algerian Lake. Environ. Pollut. 2016, 216, 836–844. [Google Scholar] [CrossRef]
  59. Nasri, H.; El Herry, S.; Bouaïcha, N. First Reported Case of Turtle Deaths during a Toxic Microcystis Spp. Bloom in Lake Oubeira, Algeria. Ecotoxicol. Environ. Saf. 2008, 71, 535–544. [Google Scholar] [CrossRef]
  60. Nasri, A.-B.; Bouaïcha, N.; Fastner, J. First Report of a Microcystin-Containing Bloom of the Cyanobacteria Microcystis Spp. in Lake Oubeira, Eastern Algeria. Arch. Environ. Contam. Toxicol. 2004, 46, 197–202. [Google Scholar] [CrossRef]
  61. Nasri, H.; Bouaïcha, N.; Harche, M.K. A New Morphospecies of Microcystis Sp. Forming Bloom in the Cheffia Dam (Algeria): Seasonal Variation of Microcystin Concentrations in Raw Water and Their Removal in a Full-scale Treatment Plant. Environ. Toxicol. Int. J. 2007, 22, 347–356. [Google Scholar] [CrossRef]
  62. Brittain, S.; Mohamed, Z.A.; Wang, J.; Lehmann, V.K.B.; Carmichael, W.W.; Rinehart, K.L. Isolation and Characterization of Microcystins from a River Nile Strain of Oscillatoria Tenuis Agardh Ex Gomont. Toxicon 2000, 38, 1759–1771. [Google Scholar] [CrossRef]
  63. Mohamed, Z.A.; Carmichael, W.W. Seasonal Variation in Microcystin Levels of River Nile Water at Sohag City, Egypt. Int. J. Limnol. 2000, 36, 227–234. [Google Scholar] [CrossRef]
  64. Mohamed, Z.A. Accumulation of Cyanobacterial Hepatotoxins by Daphnia in Some Egyptian Irrigation Canals. Ecotoxicol. Environ. Saf. 2001, 50, 4–8. [Google Scholar] [CrossRef]
  65. Mohamed, Z.A.; Deyab, M.A.; Abou-Dobara, M.I.; El-Sayed, A.K.; El-Raghi, W.M. Occurrence of Cyanobacteria and Microcystin Toxins in Raw and Treated Waters of the Nile River, Egypt: Implication for Water Treatment and Human Health. Environ. Sci. Pollut. Res. 2015, 22, 11716–11727. [Google Scholar] [CrossRef]
  66. Mohamed, Z.A. First Report of Toxic Cylindrospermopsis Raciborskii and Raphidiopsis Mediterranea (Cyanoprokaryota) in Egyptian Fresh Waters. FEMS Microbiol. Ecol. 2007, 59, 749–761. [Google Scholar] [CrossRef] [Green Version]
  67. Amer, R.; Díez, B.; El-Shehawy, R. Diversity of Hepatotoxic Cyanobacteria in the Nile Delta, Egypt. J. Environ. Monit. 2009, 11, 126–133. [Google Scholar] [CrossRef]
  68. Oudra, B.; El Andaloussi, M.; Franca, S.; Barros, P.; Martins, R.; Oufdou, K.; Sbiyyaa, B.; Loudiki, M.; Mezrioui, N.; Vasconcelos, V. Harmful Cyanobacterial Toxic Blooms in Waste Stabilisation Ponds. Water Sci. Technol. 2000, 42, 179–186. [Google Scholar] [CrossRef]
  69. Oudra, B.; Loudiki, M.; Sbiyyaa, B.; Martins, R.; Vasconcelos, V.; Namikoshi, N. Isolation, Characterization and Quantification of Microcystins (Heptapeptides Hepatotoxins) in Microcystis Aeruginosa Dominated Bloom of Lalla Takerkoust Lake–Reservoir (Morocco). Toxicon 2001, 39, 1375–1381. [Google Scholar] [CrossRef] [PubMed]
  70. Douma, M.; Manaut, N.; Oudra, B.; Loudiki, M. First Report of Cyanobacterial Diversity and Microcystins in a Microcystis Strain from Sidi Boughaba, a Moroccan Coastal Lagoon. Afr. J. Aquat. Sci. 2016, 41, 445–452. [Google Scholar] [CrossRef]
  71. Hammou, H.A.; Latour, D.; Samoudi, S.; Mouhri, K.; Douma, M.; Robin, J.; Loudiki, M. Occurrence of the First Toxic Microcystis Bloom in a Recent Moroccan Reservoir. Water Resour. 2018, 45, 409–417. [Google Scholar] [CrossRef]
  72. El Herry-Allani, S.; Bouaïcha, N. Cyanobacterial Blooms in Dams: Environmental Factors, Toxins, Public Health, and Remedial Measures. In Dams: Structure, Performance and Safety Management; Nova Science Publishers, Inc.: Hauppauge, NY, USA, 2013; pp. 227–264. [Google Scholar]
  73. Willén, E.; Ahlgren, G.; Tilahun, G.; Spoof, L.; Neffling, M.-R.; Meriluoto, J. Cyanotoxin Production in Seven Ethiopian Rift Valley Lakes. Inl. Waters 2011, 1, 81–91. [Google Scholar] [CrossRef] [Green Version]
  74. Major, Y.; Kifle, D.; Spoof, L.; Meriluoto, J. Cyanobacteria and Microcystins in Koka Reservoir (Ethiopia). Environ. Sci. Pollut. Res. 2018, 25, 26861–26873. [Google Scholar] [CrossRef]
  75. Admas, A.; Agida, A.; Melese, S.; Genetu, A. Isolation and Characterization of Toxic Cyanobacterial Communities Distribution in Lake Tana, Amhara Regional State, Ethiopia. bioRxiv 2019. preprint. [Google Scholar] [CrossRef]
  76. Tilahun, S.; Kifle, D.; Zewde, T.W.; Johansen, J.A.; Demissie, T.B.; Hansen, J.H. Temporal Dynamics of Intra-and Extra-Cellular Microcystins Concentrations in Koka Reservoir (Ethiopia): Implications for Public Health Risk. Toxicon 2019, 168, 83–92. [Google Scholar] [CrossRef]
  77. Krienitz, L.; Ballot, A.; Wiegand, C.; Kotut, K.; Codd, G.A.; Pflugmacher, S. Cyanotoxin-Producing Bloom of Anabaena Flos-Aquae, Anabaena Discoidea and Microcystis aeruginosa (Cyanobacteria) in Nyanza Gulf of Lake Victoria, Kenya. Angew. Bot. 2002. [Google Scholar]
  78. Krienitz, L.; Ballot, A.; Casper, P.; Kotut, K.; Wiegand, C.; Pflugmacher, S. Cyanobacteria in Hot Springs of East Africa and Their Potential Toxicity. Algol. Stud. Hydrobiol. Suppl. Vol. 2005, 297–306. [Google Scholar] [CrossRef]
  79. Ballot, A.; Krienitz, L.; Kotut, K.; Wiegand, C.; Pflugmacher, S. Cyanobacteria and Cyanobacterial Toxins in the Alkaline Crater Lakes Sonachi and Simbi, Kenya. Harmful Algae 2005, 4, 139–150. [Google Scholar] [CrossRef]
  80. Kotut, K.; Ballot, A.; Wiegand, C.; Krienitz, L. Toxic Cyanobacteria at Nakuru Sewage Oxidation Ponds–A Potential Threat to Wildlife. Limnologica 2010, 40, 47–53. [Google Scholar] [CrossRef]
  81. Mwaura, F.; Koyo, A.O.; Zech, B. Cyanobacterial Blooms and the Presence of Cyanotoxins in Small High Altitude Tropical Headwater Reservoirs in Kenya. J. Water Health 2004, 2, 49–57. [Google Scholar] [CrossRef]
  82. Kaggwa, M.N.; Straubinger-Gansberger, N.; Schagerl, M. Cyanotoxins in Small Artificial Dams in Kenya Utilised for Cage Fish Farming–a Threat to Local People? Afr. J. Aquat. Sci. 2018, 43, 123–129. [Google Scholar] [CrossRef]
  83. Sekadende, B.C.; Lyimo, T.J.; Kurmayer, R. Microcystin Production by Cyanobacteria in the Mwanza Gulf (Lake Victoria, Tanzania). Hydrobiologia 2005, 543, 299–304. [Google Scholar] [CrossRef]
  84. Kimambo, O.N.; Gumbo, J.R.; Chikoore, H. The Occurrence of Cyanobacteria Blooms in Freshwater Ecosystems and Their Link with Hydro-Meteorological and Environmental Variations in Tanzania. Heliyon 2019, 5, e01312. [Google Scholar] [CrossRef] [Green Version]
  85. Green, J. Geographical Variation in Rotifers Associated with Microcystis Blooms. Hydrobiologia 2011, 662, 197–204. [Google Scholar] [CrossRef]
  86. Okello, W.; Ostermaier, V.; Portmann, C.; Gademann, K.; Kurmayer, R. Spatial Isolation Favours the Divergence in Microcystin Net Production by Microcystis in Ugandan Freshwater Lakes. Water Res. 2010, 44, 2803–2814. [Google Scholar] [CrossRef] [Green Version]
  87. Okello, W.; Portmann, C.; Erhard, M.; Gademann, K.; Kurmayer, R. Occurrence of Microcystin-producing Cyanobacteria in Ugandan Freshwater Habitats. Environ. Toxicol. 2010, 25, 367–380. [Google Scholar] [CrossRef] [Green Version]
  88. Poste, A.E.; Hecky, R.E.; Guildford, S.J. Phosphorus Enrichment and Carbon Depletion Contribute to High Microcystis Biomass and Microcystin Concentrations in Ugandan Lakes. Limnol. Oceanogr. 2013, 58, 1075–1088. [Google Scholar] [CrossRef]
  89. Addico, G.; Hardege, J.; Komarek, J.; Babica, P.; de Graft-Johnson, K.A.A. Cyanobacteria Species Identified in the Weija and Kpong Reservoirs, Ghana, and Their Implications for Drinking Water Quality with Respect to Microcystin. Afr. J. Mar. Sci. 2006, 28, 451–456. [Google Scholar] [CrossRef]
  90. Addico, G.N.D.; Lawton, L.; Edwards, C. Hepatotoxic-Microcystins in Two Drinking Water Reservoirs in the Central Region of Ghana. Hepatotoxic-Microcystins in Two Drinking. Toxicol. Forensic Med. Open J. 2017, 2, 1–11. [Google Scholar] [CrossRef]
  91. Akin-Oriola, G.A.; Anetekhai, M.A.; Oriola, A. Algal Blooms in Nigerian Waters: An Overview. Afr. J. Mar. Sci. 2006, 28, 219–224. [Google Scholar] [CrossRef]
  92. Odokuma, L.O.; Isirima, J.C. Distribution of Cyanotoxins in Aquatic Environments in the Niger Delta. Afr. J. Biotechnol. 2007, 6. [Google Scholar] [CrossRef] [Green Version]
  93. Chia, A.M.; Oniye, S.J.; Ladan, Z.; Lado, Z.; Pila, A.E.; Inekwe, V.U.; Mmerole, J.U. A Survey for the Presence of Microcystins in Aquaculture Ponds in Zaria, Northern-Nigeria: Possible Public Health Implication. Afr. J. Biotechnol. 2009, 8, 6282–6289. [Google Scholar] [CrossRef]
  94. Chia, M.A.; Kwaghe, M.J. Microcystins Contamination of Surface Water Supply Sources in Zaria-Nigeria. Environ. Monit. Assess. 2015, 187, 606. [Google Scholar] [CrossRef]
  95. Indabawa, I. Detection of Variants of Microcystin Produced by Microcystis Aeruginosa in Some Burrow Pits of Kano, Nigeria. Bayero J. Pure Appl. Sci. 2010, 2, 189–197. [Google Scholar] [CrossRef]
  96. Okechukwu I, O. Cyanobacteria Abundance and Its Relationship to Water Quality in the Mid-Cross River Floodplain, Nigeria. Rev. Biol. Trop. 2009, 57, 33–43. [Google Scholar] [CrossRef] [Green Version]
  97. Onyema, I.C. Phytoplankton Diversity and Succession in the Iyagbe Lagoon, Lagos. Eur. J. Sci. Res. 2010, 43, 61–74. [Google Scholar]
  98. Mussagy, A.; Annadotter, H.; Cronberg, G. An Experimental Study of Toxin Production in Arthrospira Fusiformis (Cyanophyceae) Isolated from African Waters. Toxicon 2006, 48, 1027–1034. [Google Scholar] [CrossRef]
  99. Pedro, O.; Correia, D.; Lie, E.; Skåre, J.U.; Leão, J.; Sandvik, M.; Berdal, K.G. Polymerase Chain Reaction (PCR) Detection of the Predominant Microcystin-Producing Genotype of Cyanobacteria in Mozambican Lakes. Afr. J. Biotechnol. 2011, 10, 19299–19308. [Google Scholar] [CrossRef] [Green Version]
  100. Pedro, O.; Rundberget, T.; Lie, E.; Correia, D.; Skaare, J.U.; Berdal, K.G.; Neves, L.; Sandvik, M. Occurrence of Microcystins in Freshwater Bodies in Southern Mozambique. J. Res. Environ. Sci. Toxicol. 2012, 1, 58–65. [Google Scholar]
  101. Pedro, O.; Lie, E.; Correia, D.; Neves, L.; Skaare, J.U.; Sandvik, M.; Berdal, K.G. Quantification of Microcystin-Producing Microcystis in Freshwater Bodies in the Southern Mozambique Using Quantitative Real Time Polymerase Chain Reaction. Afr. J. Biotechnol. 2013, 12, 4850–4857. [Google Scholar] [CrossRef] [Green Version]
  102. Magonono, M.; Oberholster, P.J.; Shonhai, A.; Makumire, S.; Gumbo, J.R. The Presence of Toxic and Non-Toxic Cyanobacteria in the Sediments of the Limpopo River Basin: Implications for Human Health. Toxins 2018, 10, 269. [Google Scholar] [CrossRef] [Green Version]
  103. Eguzozie, K.U.; Mavumengwana, V.; Nkosi, D.; Kayitesi, E.; Nnabuo-Eguzozie, E.C. Screening of Cyanobacterial Peptide Toxin, Microcystins in Hyperscum Water Samples from an Inland Sub Saharan Drinking Freshwater Reservoir. Bull. Environ. Contam. Toxicol. 2016, 97, 728–736. [Google Scholar] [CrossRef]
  104. Conradie, K.R.; Barnard, S. The Dynamics of Toxic Microcystis Strains and Microcystin Production in Two Hypertrofic South African Reservoirs. Harmful Algae 2012, 20, 1–10. [Google Scholar] [CrossRef]
  105. Van Ginkel, C.E.; Silberbauer, M.J.; Du Plessis, S.; Carelsen, C.I.C. Monitoring Microcystin Toxin and Chlorophyll in Five South African Impoundments. Int. Ver. Theor. Angew. Limnol. Verh. 2006, 29, 1611–1616. [Google Scholar] [CrossRef]
  106. Nchabeleng, T.; Cheng, P.; Oberholster, P.J.; Botha, A.M.; Smit, W.J.; Luus-Powell, W.J. Microcystin-LR Equivalent Concentrations in Fish Tissue during a Postbloom Microcystis Exposure in Loskop Dam, South Africa. Afr. J. Aquat. Sci. 2014, 39, 459–466. [Google Scholar] [CrossRef] [Green Version]
  107. Wicks, R.J.; Thiel, P.G. Environmental Factors Affecting the Production of Peptide Toxins in Floating Scums of the Cyanobacterium Microcystis Aeruginosa in a Hypertrophic African Reservoir. Environ. Sci. Technol. 1990, 24, 1413–1418. [Google Scholar] [CrossRef]
  108. Oberholster, P.J.; Botha, A.-M. Use of PCR Based Technologies for Risk Assessment of a Winter Cyanobacterial Bloom in Lake Midmar, South Africa. Afr. J. Biotechnol. 2007, 6. [Google Scholar] [CrossRef] [Green Version]
  109. Mokoena, M.M.; Mukhola, M.S. Current Effects of Cyanobacteria Toxin in Water Sources and Containers in the Hartbeespoort Dam Area, South Africa. Int. J. Environ. Res. Public Health 2019, 16, 4468. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Van Ginkel, C.E.; Conradie, B. Potentially Toxic Algal Incident in the Orange River, Northern Cape, 2000; Internal Report No. N/D801/12/DEQ/0800; Institute for Water Quality Studies, Department of Water Affairs and Forestry: Pretoria, South Africa, 2001. [Google Scholar]
  111. Oberholster, P.J.; Myburgh, J.G.; Govender, D.; Bengis, R.; Botha, A.-M. Identification of Toxigenic Microcystis Strains after Incidents of Wild Animal Mortalities in the Kruger National Park, South Africa. Ecotoxicol. Environ. Saf. 2009, 72, 1177–1182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Mhlanga, L.; Day, J.; Cronberg, G.; Chimbari, M.; Siziba, N.; Annadotter, H. Cyanobacteria and Cyanotoxins in the Source Water from Lake Chivero, Harare, Zimbabwe, and the Presence of Cyanotoxins in Drinking Water. Afr. J. Aquat. Sci. 2006, 31, 165–173. [Google Scholar] [CrossRef]
  113. Magadza, C.H.D. 2009. Re-Eutrophication and Pathenogenic Contamination of Lake Chivero. Integrated Lake Basin Management; International Lake Environment Committee: Shiga, Japan, 2008. [Google Scholar]
  114. Kunz, M.J. Effect of Large Dams in the Zambezi River Basin: Changes in Sediment, Carbon and Nutrient Fluxes. Ph.D. Thesis, Swiss Federal Institute of Technology, Zürich, Switzerland, 2011. [Google Scholar]
  115. Tendaupenyu, P. Nutrient Limitation of Phytoplankton in Five Impoundments on the Manyame River, Zimbabwe. Water SA 2012, 38, 97–104. [Google Scholar] [CrossRef] [Green Version]
  116. Mankiewicz-Boczek, J.; Gągała, I.; Jurczak, T.; Urbaniak, M.; Negussie, Y.Z.; Zalewski, M. Incidence of Microcystin-producing Cyanobacteria in Lake Tana, the Largest Waterbody in Ethiopia. Afr. J. Ecol. 2015, 53, 54–63. [Google Scholar] [CrossRef]
  117. Oudra, B.; Dadi-El Andaloussi, M.; Vasconcelos, V.M. Identification and Quantification of Microcystins from a Nostoc Muscorum Bloom Occurring in Oukaïmeden River (High-Atlas Mountains of Marrakech, Morocco). Environ. Monit. Assess. 2009, 149, 437–444. [Google Scholar] [CrossRef]
  118. Mohamed, Z.A. Harmful Cyanobacteria and Their Cyanotoxins in Egyptian Fresh Waters–State of Knowledge and Research Needs. Afr. J. Aquat. Sci. 2016, 41, 361–368. [Google Scholar] [CrossRef]
  119. Mohamed, Z.A. Breakthrough of Oscillatoria Limnetica and Microcystin Toxins into Drinking Water Treatment Plants-Examples from the Nile River, Egypt. Water SA 2016, 42, 161–165. [Google Scholar] [CrossRef] [Green Version]
  120. Mokoena, M.M.; Mukhola, M.S.; Okonkwo, O.J. Hazard Assessment of Microcystins from the Household’s Drinking Water. Appl. Ecol. Environ. Res. 2016, 14, 695–710. [Google Scholar] [CrossRef]
  121. Chia, A.M.; Abolude, D.S.; Ladan, Z.; Akanbi, O.; Kalaboms, A. The Presence of Microcystins in Aquatic Ecosystems in Northern Nigeria: Zaria as a Case Study. Res. J. Environ. Toxicol. 2009, 3, 170–178. [Google Scholar] [CrossRef] [Green Version]
  122. Addico, G.N.D.; Hardege, J.D.; Babica, P.; Kohoutek, J.; Johnson, K.A.A.D.G. Microcystin-RR Like Toxin Identified in the Cyano- Bacterium Anabaena Flos-Aquae Strain CCAP 1403/13B Culture. West Afr. J. Appl. Ecol. 2011, 19. [Google Scholar]
  123. Kouassi, B.A.T.; Adon, M.P.; Komoé, K.; Ouattara, A. Cyanobacteria from a Shallow Reservoir in Côte d’Ivoire UFR Biological Sciences, Peleforo Gon Coulibaly University, Korhogo, Côte d’Ivoire. J. Biodivers. Environ. Sci. 2015, 7, 136–149. [Google Scholar]
  124. Kaebernick, M.; Dittmann, E.; Börner, T.; Neilan, B.A. Multiple Alternate Transcripts Direct the Biosynthesis of Microcystin, a Cyanobacterial. Appl. Environ. Microbiol. 2002, 68, 449–455. [Google Scholar] [CrossRef] [Green Version]
  125. Christiansen, G.; Molitor, C.; Philmus, B.; Kurmayer, R. Nontoxic Strains of Cyanobacteria Are the Result of Major Gene Deletion Events Induced by a Transposable Element. Mol. Biol. Evol. 2008, 25, 1695–1704. [Google Scholar] [CrossRef]
Figure 1. Microcystin chemical structure generalized as cyclo-D-Ala1-X2-D-MeAsp3-Z4-Adda5-D-Glu6-Mdha7, where X and Z denote the highly variable L-amino acids present at the second and fourth positions, with possible different combinations of seven amino acids that can produce different microcystin variants (modified from Butler et al. [13]).
Figure 1. Microcystin chemical structure generalized as cyclo-D-Ala1-X2-D-MeAsp3-Z4-Adda5-D-Glu6-Mdha7, where X and Z denote the highly variable L-amino acids present at the second and fourth positions, with possible different combinations of seven amino acids that can produce different microcystin variants (modified from Butler et al. [13]).
Toxics 10 00772 g001
Figure 2. Microcystin congeners. (A) Microcystin-AR, (B) Microcystin-LF, (C) Microcystin-LR, (D) Microcystin –LY. Source: PubChem.
Figure 2. Microcystin congeners. (A) Microcystin-AR, (B) Microcystin-LF, (C) Microcystin-LR, (D) Microcystin –LY. Source: PubChem.
Toxics 10 00772 g002
Figure 3. Neighbor-joining phylogenetic tree of the 16S rDNA gene of bloom samples and isolated cyanobacteria from Africa and other continents. The bootstrap consensus tree was inferred from 1000 replicates, and the evolutionary distances were computed using the maximum composite likelihood method. All ambiguous positions were removed for each sequence pair (pairwise deletion option). There were a total of 1462 bp in the final dataset. Evolutionary analyses were conducted in MEGA X version 11 for macOS.
Figure 3. Neighbor-joining phylogenetic tree of the 16S rDNA gene of bloom samples and isolated cyanobacteria from Africa and other continents. The bootstrap consensus tree was inferred from 1000 replicates, and the evolutionary distances were computed using the maximum composite likelihood method. All ambiguous positions were removed for each sequence pair (pairwise deletion option). There were a total of 1462 bp in the final dataset. Evolutionary analyses were conducted in MEGA X version 11 for macOS.
Toxics 10 00772 g003
Figure 4. Phylogenetic tree constructed using the neighbor-joining method of the ITS1 sequences’ gene of cyanobacteria detected from studies on the continent. Branches corresponding to partitions reproduced in less than 50% of bootstrap replicates are collapsed. The percentage of replicate trees in which the associated taxa clustered together in the bootstrap test (1000 replicates) are shown next to the branches. The evolutionary distances were computed using the Kimura 2-parameter method and are in the units of the number of base substitutions per site. This analysis involved 34 nucleotide sequences. All ambiguous positions were removed for each sequence pair (pairwise deletion option). There was a total of 539 bp in the final dataset. Evolutionary analyses were conducted in MEGA X version 11.
Figure 4. Phylogenetic tree constructed using the neighbor-joining method of the ITS1 sequences’ gene of cyanobacteria detected from studies on the continent. Branches corresponding to partitions reproduced in less than 50% of bootstrap replicates are collapsed. The percentage of replicate trees in which the associated taxa clustered together in the bootstrap test (1000 replicates) are shown next to the branches. The evolutionary distances were computed using the Kimura 2-parameter method and are in the units of the number of base substitutions per site. This analysis involved 34 nucleotide sequences. All ambiguous positions were removed for each sequence pair (pairwise deletion option). There was a total of 539 bp in the final dataset. Evolutionary analyses were conducted in MEGA X version 11.
Toxics 10 00772 g004
Figure 5. UPGMA phylogenetic tree of the mcyE gene of blooms and isolated cyanobacteria from Africa and other continents. The bootstrap consensus tree was inferred from 1000 replicates, and the evolutionary distances were computed using the maximum composite likelihood method. The analysis involved 54 nucleotide sequences. All ambiguous positions were removed for each sequence pair (pairwise deletion option). There was a total of 847 bp in the final dataset. Evolutionary analyses were conducted in MEGA X version 11 for macOS.
Figure 5. UPGMA phylogenetic tree of the mcyE gene of blooms and isolated cyanobacteria from Africa and other continents. The bootstrap consensus tree was inferred from 1000 replicates, and the evolutionary distances were computed using the maximum composite likelihood method. The analysis involved 54 nucleotide sequences. All ambiguous positions were removed for each sequence pair (pairwise deletion option). There was a total of 847 bp in the final dataset. Evolutionary analyses were conducted in MEGA X version 11 for macOS.
Toxics 10 00772 g005
Table 2. The mcy operon involved in the biosynthesis of microcystins by cyanobacteria.
Table 2. The mcy operon involved in the biosynthesis of microcystins by cyanobacteria.
GeneProteinFunctions
mcyH McyHABC transporter
mcyI McyI Putative dehydrogenase
mcyF McyF Amino acid racemase
mcyEMcyE NRPS-PKS (KS-AT-ACP-AMT-C-A-PCP-C)
mcyDMcyDPKS (KS-DH-CM-KR-ACP-KS-AT-DH-KR-ACP)
mcyGMcyGNRPS-PKS (A-PCP-KS-AT-CM-KR-ACP)
mcyAMcyANRPS (A-NMT-PCP-C-A-PCP-E)
mcyBMcyBNRPS (C-A-PCP-C-A-PCP)
mcyCMcyC NRPS (C-A-PCP-TE)
mcyJ McyJO-acetyltransferase
mcyA-C and mcyD-J, putative operons bidirectionally transcribed; NRPS, nonribosomal peptide synthetases; PKS, polyketide synthases; NRPS-PKS, hybrid NRPS/PKS; A-domain, substrate activation; NMT-domain, N-methylation; PCP, covalent intermediate tethering; C-domain, condensation of intermediates; E-domain, substrate epimerization; TE-domain, cyclization and release; O-acetyltransferase, substrate acetylation; ACP, covalent intermediate tethering; KR-domain, ketoreductase; CM-domain, C-methyltransferase; DH-domain, dehydratase; AT-domain, acyltransferase; KS-domain, ketosynthase; AMT-domain, transfer of amino group [9,49,52].
Table 3. Diversity of microcystin congeners recorded in different aquatic ecosystems in Africa.
Table 3. Diversity of microcystin congeners recorded in different aquatic ecosystems in Africa.
Sub-regionCountryType of Water BodyMajor Toxin-Producing SpeciesDetected CyanotoxinsConcentration of Detected CyanotoxinsDetection MethodReferences
North AfricaAlgeriaFreshwater lakesMicrocystis spp.MC-RR, MC-LR, MC-FR, MC-WR, MC-YR, MC-LA, MC-(H4)YR, MC-HilR, [Asp3]MCRAba, [Glu(OCH3)6]MC-LR, and [Glu(OCH3)6]MC-FR4590 μg g−1 DW total microcystinsLCMS/MS, HPLC[58,59,60,61]
EgyptNile river; irrigation canals; upper Egypt fishpondsMicrocystis aeruginosa,
Oscillatoria tenuis,
Anabaena subcylindrica,
Anabaena variables,
Nostoc spongiaeforme,
Oscillatoria limnetica,
Phormidium corium,
Planktothrix agardhii,
Anabaena variables,
Plectonema boryanum,
Raphidiopsis raciborskii
Total MCs, MC-LR, MC-RR, MC-YR300 to 877 µg L−1 total microcystinsHPLC, NMR, ESIMS, ESIMS-CID-MS, ELISA[62,63,64,65,66,67]
MoroccoReservoirs, ponds, waste stabilization ponds, and riversNostoc muscorumTotal MCs, MC-LR, MC-RR, and MC-WR229.4 μg/g−1 DWELISA, HPLC, FAB-MS[68,69,70,71]
TunisiaDams,
reservoirs
Microcystis spp.,
Oscillatoria tenuis,
M. aeruginosa, R. raciborskii, Planktothrix agardhii
MC-LR equivalent
MC-LR (MC-FR, MC-RR, MC-YR, and MCWR)
0.008–5.57 μg L−1GC/MS, PP2 inhibition assay[31,59,72]
East AfricaEthiopiaReservoirs and lakesMicrocystis spp.,
Microcystis aeruginosa,
Microcystis botrys,
Microcystis flos-aquae
Microcystis novacekii,
Microcystis panniformis,
Anabaena spp.,
Nostoc spp.,
Raphidiopsis spp.,
Raphidiopsis africana,
Raphidiopsis. Raciborskii
MC-LR, MC-YR, MC-RR, MC-dmLR, and MC-LA0.58- 1547.28 μg L−1 total microcystinsELISA, LC-ESI-MS-MRM, HPLC-DAD[46,73,74,75,76]
KenyaFreshwater lakes and saline lakesMicrocystis aeruginosa,
Arthrospira fusiformis,
Microcystis flos-aquae,
Phormidium terebriformis,
Oscillatoria willei,
Spirulina subsalsa
MC-LR;MC-RR; MC-LF; MC-YR1.6–19,800 μg g−1 DWHPLC-PDA, MALDI-TOF, ELISA[29,34,77,78,79,80,81,82]
TanzaniaFreshwater lakesMicrocystis sp. MC-RR0–1.0 μg L−1HPLC-DAD[83,84]
UgandaFreshwater lakesMicrocystis aeruginosa,
M. flos-aquae,
Anabaenopsis spp.,
Aphanizomenon sp.,
Anabaena sp.,
Raphidiopsis raciborskii
MC-RR, (Asp3) MC-RR, MC-YR, (Asp3)
MC-YR, MC-LR, MC-RY, (Asp3) MC-RY
0.02–10.00 μg MC-LR eq./L
0.03 to 144 fg cell−1 total microcystins
HPLC-DAD, LC-MS/MS,
MALDI-TOF MS
[34,35,36,85,86,87,88]
West AfricaGhanaFreshwater reservoirsMicrocystis aeruginosaMC-RR0.03–3.21 μg L−1 total microcystins, 0.1–0.79 μg L−1, 10.60 μg g−1 DW MC-RR HPLC-UV[89,90]
NigeriaLakes, reservoirs, lagoons, rivers, ponds, floodplainRaphidiopsis sp.,
Microcystis sp.,
Microcystis aeruginosa,
Microcystis flos-aquae,
Microcystis wesenbergii,
Lyngbya sp.
MCs, MC-LR,
MC-RR
0.19–7.75, 3.8, 1.68–3.94, 0.6–5.89 μg L−1 total microcystinsELISA, HPLC-DAD[12,91,92,93,94,95,96,97]
South Africa MozambiqueWastewater treatment ponds, dams, lakes, irrigation channelsArthrospira fusiformis,
Microcystis spp.
MC-LR, MC-YR, MC-RR0.10–- 7.89 μg L−1
2.1–159.4 ng g−1 DW
LCMS[98,99,100,101]
BotswanaPonds, rivers, and lakesMicrocystis novacekii,
Raphidiopsis raciborskii,
Phormidium spp.,
Planktothrix spp.
MC-RR, MC-YR, MC-LR, and MC-WR53.62 μg g−1 DW MC-RR; 12.114 μg g−1 DW MC-LRLC-MS/MS[102,103]
South AfricaReservoirs, lakes, and riversMicrocystis spp.,
Anabaena spp.,
Microcystis aeruginosa,
Microcystis flos-aquae,
Raphidiopsis raciborskii,
Phormidium spp.,
Planktothrix spp.
MC-LR, MC-YR, MC-RR, MC-(H4)YR and (D-Asp3, Dha7) MC-RR; MC-LA, MC-LF14.10 to 270.7 μg g−1 MC-LR;
0.15 to 72.28 μg g−1 MC-YR
HPLC-DAD, HPLC-UV, ELISA, LC-MS [20,24,26,39,53,57,102,103,104,105,106,107,108,109,110,111]
ZimbabweRivers and lakesMicrocystis spp.,
Raphidiopsis raciborskii,
Phormidium spp.,
Planktothrix spp.
MC-LR0.2–22.48 μg L−1ELISA, HPLC[102,112,113,114,115]
DW = dry weight.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chia, M.A.; Ameh, I.; George, K.C.; Balogun, E.O.; Akinyemi, S.A.; Lorenzi, A.S. Genetic Diversity of Microcystin Producers (Cyanobacteria) and Microcystin Congeners in Aquatic Resources across Africa: A Review Paper. Toxics 2022, 10, 772. https://doi.org/10.3390/toxics10120772

AMA Style

Chia MA, Ameh I, George KC, Balogun EO, Akinyemi SA, Lorenzi AS. Genetic Diversity of Microcystin Producers (Cyanobacteria) and Microcystin Congeners in Aquatic Resources across Africa: A Review Paper. Toxics. 2022; 10(12):772. https://doi.org/10.3390/toxics10120772

Chicago/Turabian Style

Chia, Mathias Ahii, Ilu Ameh, Korie Chibuike George, Emmanuel Oluwadare Balogun, Suwebat Ayanronke Akinyemi, and Adriana Sturion Lorenzi. 2022. "Genetic Diversity of Microcystin Producers (Cyanobacteria) and Microcystin Congeners in Aquatic Resources across Africa: A Review Paper" Toxics 10, no. 12: 772. https://doi.org/10.3390/toxics10120772

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