Next Article in Journal
Testing Evapotranspiration Estimates Based on MODIS Satellite Data in the Assessment of the Groundwater Recharge of Karst Aquifers in Southern Italy
Next Article in Special Issue
Microbial Detoxification of Dimethoate and Methomyl Residues in Aqueous Media
Previous Article in Journal
Flow Resistance in Open Channel Due to Vegetation at Reach Scale: A Review
Previous Article in Special Issue
The Horizontal Distribution of Siliceous Planktonic Radiolarian Community in the Eastern Indian Ocean
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Succession of Microbial Community in a Small Water Body within the Alluvial Aquifer of a Large River

1
Department of Biology, Faculty of Science, University of Zagreb, Rooseveltov trg 6, HR-10000 Zagreb, Croatia
2
Department of Hydrogeology and Engineering Geology, Croatian Geological Survey, Sachsova 2, HR-10000 Zagreb, Croatia
3
Ruđer Bošković Institute, Bijenička cesta 54, HR-10000 Zagreb, Croatia
4
Center of Excellence for Science and Technology-Integrating Mediterranean Region (STIM), HR-21000 Split, Croatia
5
Evolutionary Genetics Laboratory (eGL), Agricultural Faculty, Ankara University, Ankara TR-06135, Turkey
6
Department of Soil Amelioration, Faculty of Agriculture, University of Zagreb, HR-10000 Zagreb, Croatia
*
Author to whom correspondence should be addressed.
Water 2021, 13(2), 115; https://doi.org/10.3390/w13020115
Submission received: 25 November 2020 / Revised: 29 December 2020 / Accepted: 30 December 2020 / Published: 6 January 2021
(This article belongs to the Special Issue Microbial Communities in Water Environments: Dynamics and Interaction)

Abstract

:
Nitrogen is one of the essential elements limiting growth in aquatic environments. Being primarily of anthropogenic origin, it exerts negative impacts on freshwater ecosystems. The present study was carried out at the nitrate-vulnerable zone within the alluvial aquifer of the large lowland Drava River. The main aim was to investigate the ecosystem’s functionality by characterizing the bacterial and phytoplankton diversity of a small inactive gravel pit by using interdisciplinary approaches. The phytoplankton community was investigated via traditional microscopy analyses and environmental DNA (eDNA) metabarcoding, while the bacterial community was investigated by a molecular approach (eDNA). Variations in the algal and bacterial community structure indicated a strong correlation with nitrogen compounds. Summer samples were characterized by a high abundance of bloom-forming Cyanobacteria. Following the cyanobacterial breakdown in the colder winter period, Bacillariophyceae and Actinobacteriota became dominant groups. Changes in microbial composition indicated a strong correlation between N forms and algal and bacterial communities. According to the nitrogen dynamics in the alluvial aquifer, we emphasize the importance of small water bodies as potential buffer zones to anthropogenic nitrogen pressures and sentinels of the disturbances displayed as algal blooms within larger freshwater systems.

1. Introduction

Various aspects of nutrient dynamics in freshwater ecosystems are of paramount importance for understanding how the productivity of surface waters is controlled and provide the opportunity to analyse the current and future impacts of anthropogenic activities on freshwater ecosystems. In such environments, a large part of the primary production may depend on the recycling of nutrients such as nitrogen compounds [1]. Nitrogen is an essential element that often limits growth in aquatic ecosystems, and a key compound in many biochemical processes that are important for life, but can be harmful in high concentrations [2,3,4]. Nowadays, anthropogenic activities such as fertilizer synthesis and its widespread application on arable areas, as well as the burning of fossil fuels, significantly increase the N fluxes across different environmental compartments [3,4,5]. Its environmental effects on aquatic ecosystems include acidification, anthropogenic eutrophication, degradation of water quality, biodiversity loss, and increased greenhouse gas emission [6,7,8]. Nitrate (NO3) pollution is causing negative impacts on groundwater and surface water resources with its primary anthropogenic origin [9,10]. An elevated concentration of nitrates is associated with diffuse and point sources such as domestic or industrial wastewaters, atmospheric deposition, and animal farming waste. However, most environmental problems related to nitrate are linked to intensive agriculture production [11], as the nitrogen is used to promote crop growth [12,13]. Alluvial groundwater is particularly vulnerable to nitrate leaching from agricultural soils, since agricultural land is characterized by the presence of shallow groundwater and fertile soil suitable for farming [14,15].
The composition of the microbial community depends on environmental conditions that may affect the ecosystem’s function [16,17,18,19], as they drive the various processes of recycling, dynamics, and assimilation of nitrogen compounds in freshwater habitats [18,19,20]. The availability of certain N forms in freshwater habitats influences the composition of the phytoplankton community, increases its productivity, and causes harmful algal blooms [6,21,22]. Bloom-forming species encompass a variety of eukaryotic algae but also Cyanobacteria, a prokaryotic algal group closely related to problematic freshwater nuisances. Cyanobacteria are extremely adaptive and competitive organisms with a long evolutionary history, which endowed them with an array of physiological, morphological, and ecological adaptations to survive in a wide variety of environmental conditions [6,23]. Many species of Cyanobacteria are capable of surviving and even thriving in extremely inhabitable conditions, tolerating desiccation, high temperatures, extreme pH, high salinity, and pesticides, thus illustrating their capacity to acclimate in different kinds of habitats [24]. They are the only planktonic group capable of utilizing atmospheric nitrogen via biological N2 fixation, and, as such, can circumvent N-limited conditions [25,26]. Cyanobacterial genera capable of diazotrophy retain a competitive advantage over other phytoplankton groups. The ability of some Cyanobacteria to form potentially toxic surface blooms has drawn much attention from the general public [21,27,28]. Worldwide, fewer than 30 species that cause a real nuisance. It is still difficult to generalize their ecological requirements, as they can be ubiquistic, specifically preferring eutrophic conditions [29]. Anthropogenic eutrophication is recognized as a global environmental problem in terms of both freshwater biodiversity loss and harmful algal blooms due to the presence of toxins [30,31]. However, the impact of eutrophication may differ in large rivers and lakes, when compared to smaller water bodies, such as streams, ponds, bogs, and small lakes, as well as groundwater [32]. Generally, the small lowland water bodies support naturally high concentrations of nutrients and range from eutrophic to hypertrophic [33,34]. Despite high nutrient conditions, small lowland water bodies collectively support a very diverse and ofttimes unique biodiversity, often richer than the one found in big rivers or lakes [35]. The consequences of eutrophication on the biodiversity of small water bodies are poorly understood and have yet to be fully explored.
One of the systems characterized by high nitrogen inputs in Croatia is the alluvial aquifer of Drava, the second longest river in Croatia. The aquifer has dozens of small lotic and lentic ecosystems, which play a potentially important role as biogeochemical reactors in nitrogen buffering and recycling. The regime and quality of these small water bodies are under heavy anthropogenic pressure, mainly due to agriculture [36]. To investigate the role of these small water bodies the nitrogen recycling in the Drava River alluvial area, we have selected a small inactive gravel pit. By employing interdisciplinary approaches, we aim to characterize the ecosystem’s functionality, emphasizing bacterial and phytoplankton diversity and its effects on nitrogen recycling along the hydrological transport pathways.

2. Materials and Methods

2.1. Study Area

The study area, situated in the Drava River valley, upstream of the town of Varaždin (NW Croatia), belongs to the Black sea catchment and covers an area of approximately 200 km2 (Figure 1). On the NW side, the alluvial aquifer is adjoined by the Varaždin Lake, an artificial reservoir of the hydroelectric power plant Varaždin, of which the Drava River watercourse constitutes a natural border in the NE direction. The Plitvica Stream flows at the S-SE edge of the study area, while in the middle there are several active and inactive gravel pits. All gravel pits represent exposed groundwater and vulnerable areas where the contamination of groundwater can occur faster from surface contaminants. In some cases, inactive (abandoned) gravel pits are used as waste (industrial or urban) disposal sites and are becoming a threat to groundwater quality [37,38,39].
The study area is densely populated, with industrial and intensive agricultural production. The most common type of crops grown are corn, cabbage, potatoes, and vegetables. Extensive poultry farming is present, especially the fattening of chickens, quails, and pheasants, and the breeding of hens [40]. The Varaždin pumping site, one of two in the area, was shut down due to high nitrogen concentrations in groundwater caused by significant anthropogenic activities [41]. Nevertheless, demands for drinking and industrial water rise because of the growing production in the area. The gravel pit in the village of Šijanec was chosen because of its inactivity and accessibility. It is a small pit covering an area of approximately 12,000 m2 (Figure 1).

2.2. Sampling and Laboratory Analysis

Phytoplankton and bacterial samples were taken monthly from June 2017 until March 2018 on the deepest point of the gravel pit using the vertical Hydro-Bios water sampler (Hydro-Bios Apparatebau GmbH, Altenholz, Germany). Samples for chemical analysis were taken simultaneously with biological samples and transported in a portable freezer at 4 °C to the laboratory for further analysis. Before sampling, physico-chemical parameters of water, including electrical conductivity (EC), pH, temperature (T), and dissolved oxygen concentration (DO), were measured with a portable WTW Multi 3630 multimeter (Xylem Analytics, Weilheim, Germany). Water transparency (ZSD) was estimated using a Secchi Disc. Total alkalinity was measured by titration with 1.6 N H2SO4 using phenolphthalein and bromocresol green-methyl red as indicators. Dissolved cations (Na+, K+, Mg2+ and Ca2+) and anions (SO42−, NO3, Cl) were analyzed by ion chromatography using a Dionex ICS-6000 (Thermo Fisher Scientific, Waltham, MA, USA), while NH4+, NO2, and PO43−-P were measured spectrophotometrically. Dissolved inorganic and organic carbon (DIC, DOC) and total inorganic and organic carbon (TIC, TOC) were analysed using a HACH QbD1200 TOC analyser (Hach Company, Frederick, MD, USA). The analytical precision of the measurements of cations and anions, indicated by the ionic balance error (IBE), was computed on the basis of ions expressed in meqL−1. The IBE value was observed to be within a limit of <±5% [42,43]. The PHREEQC geochemical code [44] was used to calculate the charge balance and pCO2 pressure and to study the saturation state of the mineral phases. The samples for Chlorophyll a were filtered on Whatman GF/F glass filters (Whatman International Ltd, Kent, UK), extracted in 96% ethanol, and measured for Chl a using a UV–VIS spectrophotometer according to compliance monitoring standards [45].

2.3. Microbial Community Analysis

Phytoplankton samples were collected for both morphological and molecular analyses. Samples for morphological analysis were fixed with acid Lugol solution and stored in 250 mL volume glass bottles in the dark at 4 °C. The morphological analysis included a qualitative and quantitative community characterization according to the Utermöhl method [46], using a Zeiss Axiovert 200 inverted microscope (Carl Zeiss, Oberkochen, Germany).
Samples for molecular analysis were collected in sterile plastic bottles and preserved on ice during the transport to the laboratory. In the laboratory, they were immediately filtered on Nucleopore Track-Etch membrane filters (25 mm diameter, 0.4 μm pore size; Whatman International Ltd, Kent, UK) and stored at −20 °C until further processing. DNA was extracted from the frozen filters using DNeasy PowerWaterKit (Qiagen, Hilden, Germany). The manufacturer’s instructions were followed with a slight modification in the final step, where 60 μL of sterile DNA-free PCR Grade water was added instead of Qiagen’s C6 Solution. The quality of the extracted DNA was assessed with the Shimadzu BioSpec-nano spectrophotometer (Shimadzu Corporation, Kyoto, Japan).

2.4. PCR of the Phytoplankton (Eukaryotic) Community

The hypervariable V9-region of the SSU rRNA gene (ca. 150 base pairs) was amplified from environmental DNA using the universal eukaryotic primer pair [47,48]. The forward and reverse primers were 1391F (5′-GTAC ACACCGCCCGTC-3′) and EukB (5′-TGATCCTTCTGCAGGTTCACCTAC-3′), designed by Amaral-Zettler and colleagues [49]. Polymerase chain reactions (PCR) contained 1 U of Hot Start Taq DNA Polymerase (New England Biolabs, Ipswich, MA, USA), and for V9 amplification an initial activation step at 95 °C was employed for 5 min, followed by 30 three-step cycles consisting of 94 °C for 30 s, 57 °C for 45 s, and 72 °C for 1 min, followed by a final 2 min extension at 72 °C. PCR products were assessed by visualizing on a 1% agarose gel. Sequencing libraries were constructed using the NEB Next® Ultra™ DNA Library Prep Kit for Illumina (New England Biolabs, Ipswich, MA, USA). Libraries were sequenced on an Illumina MiSeq platform, generating 250-bp paired-end reads (SeqIT GmbH & Co. KG, Kaiserslautern, Germany).

2.5. PCR for the Bacterial Community

The V3-4 region of bacterial rRNA genes was amplified using the forward primer 341 F 5′-CCTACGGGNGGCWGCAG-3′ and reverse primer 805 R 5′-GACTACHVGGGTATCTAATCC-3′ [50]. PCR and sequencing were performed in the LGC Genomics GmbH laboratory (Berlin, Germany). The libraries were sequenced on an Illumina MiSeq platform, generating 300-bp paired-end reads.

2.6. Bioinformatic Analysis of the Phytoplankton (Eukaryotic) Community

Sequence reads were analysed using the programs implemented in the OBITools package, as described in De Barba et al. [51]. The quality of the reads was assessed using FastQC. Paired-end reads were aligned using Illumina paired-end, and alignments with quality scores >32 were kept. The aligned data set was demultiplexed using the ngsfilter command, which identified primers and tags and assigned the sequences to each sample. For dereplication, we used the obiuniq command for clustering together strictly identical sequences and keeping the information about their distribution among samples. Sequences shorter than 10 bp, or containing ambiguous nucleotides, or with occurrence lower or equal to 10 were excluded using the obigrep command. The obiclean command was then run to assign the status of “head”, “internal”, or “singleton” to each sequence within a PCR product. All sequences labeled “internal”, corresponding to PCR errors, were discarded. Finally, the taxonomic assignment of OTUs was performed with the ecotag command, combining two reference databases, filtered according to target taxa from NCBI taxonomy and the EMBL database, after running the ecoPCR program [52,53]. Only sequences with a 98% match to the reference sequence were kept. Single-read OTUs were removed from the samples to avoid potential false positives. The final filtered file with taxonomically assigned OTUs of eukaryotic algae groups was used as a basis for all downstream analyses. The DNA sequencing reaction on two samples (September and October 2017) did not yield valid results. A list of commands with related parameters are presented in the Supplementary Materials (S1). Raw demultiplexed reads were deposited at the ENA’s Sequence Read Archive and are publicly available under project number PRJEB40961.

2.7. Bioinformatic Analysis of the Bacterial Community

The quality of the reads was assessed using FastQC. Paired-end reads were quality-trimmed using the bbduk function and merged using the bbmerge function of the BBMap package 34.48 (Lawrence Berkeley National Lab., Berkeley, CA, USA) [54]. Merged reads were quality-filtered using QIIME v1.8.0 [55]. Reads with exact barcodes and primers, unambiguous nucleotides, and a minimum length of 250 base pairs were retained. A Chimera check was done using UCHIME [56]. Non-chimeric reads were clustered with SWARM v3.0.0 [57] with default settings into Operational Taxonomic Units (OTUs). The bacterial reads were blasted against the SILVA database release 138 (Max Planck Institute for Marine Microbiology and Jacobs University, Bremen, Germany) using blastn (BLAST v2.9.0) [58]. Nontarget OTUs (chloroplasts, mitochondria), as well as singletons and doubletons, were excluded. The resulting OTUs were filtered by the quality of the blast results (≥98% identity). The DNA sequencing reaction on two samples (February and March 2018) did not yield valid results. Raw demultiplexed reads were deposited at the ENA’s Sequence Read Archive and are publicly available under project number PRJEB40962.

2.8. Statistical Analysis

Statistical analyses were conducted in R v. 4.0.2 [59] using the program packages “vegan”, “fossil”, “factoextra”, “devtools”, and “ggbiplot”, as well as “ggplot2” for all graphical representations. To access the comparability of morphological and molecular methodologies in the phytoplankton community, the taxa lists derived from both approaches were compared with regard to the presence or absence of taxa and community composition. The bacterial community composition was analysed by using the molecular approach. The results for downstream analysis were combined into a single data set for each approach and for each community. The molecular results were transformed into relative abundances to normalize the OTU database [60]. Biomass data obtained by microscopy for the phytoplankton community were transformed following the logarithmic scale [61].
The Shannon, Simpson, and richness indices were calculated for both approaches and both communities as measures of alpha diversity using program packages “Vegan v. 2.5.6” [62].
To test the statistical significance of the environmental parameters and which parameter was singled out depending on the month studied, a principal component analysis (PCA) was performed using the R package “vegan” [62].
Canonical correspondence analyses (CCA) were performed on phytoplankton morphological data and bacterial molecular data to estimate variance in environmental variables for both communities. ANOVA test was applied to test the statistical significance of all axes, and forward selections were used to evaluate the importance of each variable. The logarithm function was used to transform environmental parameters and both community datasets for statistical analysis.

3. Results

3.1. Analysis of Environmental Parameters

The environmental variables of the investigated gravel pit are indicated in Table 1. The highest value of nitrates (NO3) concentration was measured in March (38.4 mg L1) and the lowest in June (0.62 mg L−1), whereas the maximum concentrations of ammonium (NH4+) and nitrites (NO2) were recorded in July (2.75 mg L−1 and 0.17 mg L−1, respectively). The highest value of pH was detected in August (9.48), indicating that the water was alkaline. The highest temperature (T) value was recorded in June (24.2 °C) and the lowest in February (1.2 °C), respectively. The maximum value of dissolved oxygen was measured in August (17.1 mg L−1), and the minimum value was in July (7.1 mg L−1). For electrical conductivity (EC) and bicarbonates (HCO3), the maximum values were recorded in March (497 μS cm−1 and 249 mg L−1, respectively) and the minimum ones in October (252 μS cm−1 and 107 mg L−1, respectively). Silicon dioxide (SiO2) concentrations were high during the warmer period, and the maximum was in September (26.8 mg L−1). During the colder period the concentrations were much lower, and the minimum was in December (7.2 mg L−1). In addition, a change was also observed in the concentrations of calcium, bicarbonates, and logpCO2 pressure (Table 1), with the lowest concentrations detected in the summer period and the highest concentrations during winter.

PCA Analysis

The principal component analysis (PCA) performed for the 24 environmental variables explained 71.1% of the total variance on the first two PC axes. The overall strength of correlations between the samples and environmental parameters are summarized in Table 2. The most important variables for the PC1 axis were TIC, Ca2+, and EC (intra-set correlations: 0.271, 0.269, and 0.268, respectively). Regarding the PC2 axis, NH4+ and DO (intra-set correlations: −0.355 and 0.407, respectively) were the variables that weighted most for the ordination. PCA arranged the samples into three groups (Figure 2): the first group consisted of samples from a warmer period of investigation (June, August, September, and October), the second group included a sample from July, while the third one comprised all samples from the colder period of investigation (November, December, January, February, and March).

3.2. Phytoplankton Succession

A total of 38 phytoplankton species were recorded by a morphological approach within the 10 samples collected during the investigated period. A total of 47,130 reads clustered into 88 OTUs were detected in the remaining eight samples. OTUs were taxonomically assigned into 30 eukaryotic algal taxa.
Phytoplankton abundance ranged between 1.9 × 106 cells L−1 in January 2018 to 2.8 × 108 cells L−1 in August 2017. Phytoplankton biomass ranged from 1.05 mg L−1 in January 2018 to 27.85 mg L−1 in August 2017, related to the cyanobacterial bloom of Microcystis spp. (Figure 3a). The chlorophyll a value fluctuated from 8.37 μg L−1 (March 2018) to 75.3 μg L−1 (July 2017). Concurrently, the highest peak of chlorophyll a concentration was not recorded during the cyanobacterial bloom in August 2017, but instead in July 2017, during the proliferation of green algae, predominantly Scenedesmus quadricauda (Turpin) Brébisson (Figure 3b).
As inferred from the morphological approach, the alpha diversity in the richness, Shannon, and Simpson indices of phytoplankton varied considerably during the investigated period. The maximum value of species richness was recorded in June 2017, while the minimum was noted in February 2018. The Shannon index values ranged from the maximum in October 2017 to the minimum in March 2018. The maximum value of the Simpson index was also noted in October 2017, but the minimum was recorded in June 2017 (Figure 4a). The richness, Shannon, and Simpson indices inferred from the molecular approach did not show the same pattern. The highest value of richness index was recorded in January 2018, while the lowest richness was noted in August 2017. The minimum and maximum values of the Shannon and Simpson indices were reported in July 2017 and June 2017, respectively (Figure 4b).
According to the frequency and biomass, the most dominant species were: filamentous cyanobacterium Limnothrix redekei (Goor) Meffert, colonial clathrate cyanobacteria Microcystis aeruginosa (Kützing) Kützing and M. wesenbergii (Komárek) Komárek ex Komárek, centric diatom Aulacoseira muzzanensis (F.Meister) Krammer, pennate diatoms Ulnaria ulna (Nitzsch) Compère and Ulnaria sp. (Kützing) Compère, cryptophyte Cryptomonas sp. Ehrenberg, dinoflagellate Peridinium sp. Ehrenberg, and colonial chlorophyte Scenedesmus quadricauda (Turpin) Brébisson.
A CCA was performed for the phytoplankton samples, and nine constrained environmental variables (Figure 5) indicated eigenvalues for the first two axes of 0.4328 and 0.3210, respectively, explaining 38.5% of the total variance on the first two axes. A Pearson environment-species permutation for the two significant axes indicated a significant correlation between abiotic constrained values and phytoplankton functional variables. According to the ANOVA permutation test, the ordination of both axes for environmental variables was statistically significant (p < 0.05). Canonical coefficients and intra-set correlations on the phytoplankton samples showed that NO3 and EC were the most important variables for the ordination axis 1 (intra-set correlation coefficients: 0.6332 and 0.5467, respectively). Regarding axis 2, NO2 and NH4+ (intra-set correlations −0.8155 and −0.6837, respectively) were the variables that weighted most for the ordination. At the positive end of both axes, phytoplankton samples were associated with EC, DO, HCO3, and NO3. At the negative end of both axes, phytoplankton samples were associated with pH, T, SiO2, NH4+, and NO2. Considering the environmental pressure to phytoplankton, the CCA analysis showed the separation of samples into three groups (Figure 5). The first group, comprised of summer and autumn samples (July to October 2017), correlated with high concentrations of NH4+ (2.75 mg L−1), NO2 (0.173 mg L−1), and SiO2 (26.8 mg L−1) and high values of pH (9.48) and T (24 °C). The most common species of the group were cyanobacteria M. aeruginosa and M. wesenbergii, green alga Scenedesmus quadricauda and cryptophyte Cryptomonas sp. According to the morphological approach, the samples collected in July were characterized by the highest Chl a concentration, the dominance of S. quadricauda, and the highest concentration of NH4+ (2.75 mg L−1). A pronounced increase in the total phytoplankton biomass was recorded in August as a result of Microcystis spp. bloom. M. aeruginosa and Cryptomonas sp. were the descriptive species of the phytoplankton community in September and October, with the continuing decrease of the phytoplankton biomass. The molecular approach did not confirm the same composition, but rather detected the cyanobacterial predominance and a higher contribution of OTUs taxonomically assigned to the Cryptomonas genera. According to the morphological approach, the second group, consisting of an outlying sample from June 2017, was characterized by the predominance of filamentous cyanobacterium Limnotrix redekei (Goor) Meffert and centric diatom Aulacoseira muzzanensis (Meister) Krammer. The molecular approach confirmed the dominance of OTUs taxonomically assigned to the Aulacoseira genera as well, but also detected a high number of reads of OTUs taxonomically assigned to genus Parvodinium. The third group, composed of winter samples (November 2017 to March 2018), correlated with low T and higher concentrations of NO3 (38.4 mg L−1) and HCO3 (249 mg L−1) and with mostly constant values of DO (11.9 to 14.7 mg L−1) and EC (497 μS cm−1). As confirmed by both morphological and molecular approaches, the most dominant species/taxonomically assigned OTUs during the colder period were pennate diatoms (Fragilariaceae) of the genus Ulnaria. Dinoflagellate Peridinium sp. was recorded only by the morphological approach as the subdominant species in November and January. The higher contribution of the genus Dinobryon was confirmed by both analyses during February and March.
As confirmed by both the morphological and molecular approaches, the most dominant species during the colder period were pennate diatoms (Fragilariaceae) of the genus Ulnaria with dinoflagellate Peridinium sp. as the subdominant species.

3.3. Bacterial Community Composition

For the eight samples, a total of 582,900 reads clustered into 1743 OTUs were recorded. Sequence reads were taxonomically assigned into 52 bacterial phyla. The bacterial composition showed a succession of differences between months. The variations of alpha diversity in the rarefied richness, Shannon, and Simpson indices are shown in Figure 6. The highest bacterial richness was recorded in July 2017, and the lowest in October 2017. The Shannon and Simpson indices showed similar results, with the highest values in January 2018 and the lowest in August 2017. Comparing the composition of bacteria and eukaryotic algae inferred from the molecular approach, low values in richness and both indices during the summer period were noted.
According to the percentage of classified OTUs, dominant bacterial phyla were Planctomycetota (22%), Cyanobacteria (64%), Bacteroidota (11%), Actinobacteriota (45%), and Proteobacteria (56%).
The CCA analysis performed on the bacterial community and seven constrained environmental variables (Figure 7) indicated eigenvalues for the first two axes of 0.7131 and 0.6563, respectively, explaining 47.2% of the total variance on the first two axes. A Pearson environment-bacterial community permutation for the two significant axes indicated a significant correlation between abiotic constrained values and bacterial functional variables. According to the ANOVA permutation test, the ordination of both axes for environmental variables was statistically significant (p < 0.05). Canonical coefficients and intra-set correlations carried out on the bacterial community samples showed that pH and DO were the most important variables for the ordination axis 1 (intra-set correlation coefficients: 0.6140 and 0.3770, respectively). Regarding axis 2, NO2 and NH4+ (intra-set correlations: 0.9229 and 0.8706, respectively) were the variables that weighted most for ordination. At the positive end of both axes, the bacterial community’s samples were associated with T, pH, and NO2. At the negative end of both axes, the bacterial community’s samples were associated with EC and HCO3. Based on the position of the samples related to the CCA1 axis, the bacterial community separated into two groups (Figure 7). The first group, comprised of summer samples (June to September 2017), correlated with a high concentration of NH4+ and NO2, and high values of pH and T. According to the percentage of classified OTUs, the most common phyla in the first group were Cyanobacteria, with the most dominant families being Pseudanabaenaceae (June 2017), Microcystaceae (July 2017), and Synechococcaceae and Microcystaceae (August 2017). This was also confirmed by a morphological approach, except for the picocyanobacterium of the genus Cyanobium, which is hard to detect under light microscopy. In June, Planctomycetota were codominant with Cyanobacteria. The sample from July was singled out as a result of the codominance of Cyanobacteria and Bacteroidota, which correlated with the concentration of NH4+. The second group, comprising samples from the colder months (October 2017 to January 2018), correlated with low T and higher concentrations of HCO3, DO, and EC. Actinobacteriota was the dominant phylum in October 2017 and January 2018, whilst samples from November and December were dominated by Proteobacteria.

4. Discussion

Data on ecology and the importance of small water bodies in alluvial lowlands are still quite scarce, as those systems are not included in the national strategies for the protection of water resources. Within larger freshwater systems, small water bodies act as biochemical reactors because of their potential for supporting high metabolic rates that are often paired with naturally high concentrations of nutrients and trophic conditions [33,34,63]. Their role is modulating nutrient retention and recycling along the hydrological pathways [63]. Even though usually related to eutrophic or hypertrophic conditions, small standing water bodies collectively support a very diverse biodiversity, often with species-richer communities more adapted to eutrophication conditions and to a broad range of physical and chemical conditions than the communities in larger water bodies [38,64]. Eutrophication has been described as a major stressor for the freshwater biodiversity of both large and small water bodies [65,66]. Rosset et al. [32] suggested that the eutrophication management of lowland small water bodies should be regulated differently than for larger freshwater systems, with the conservation efforts focused on the protection of small water bodies representing a mosaic of trophic conditions (and acting as hosts of regional biodiversity).
Seasonal changes with complex dynamic phases govern the high rates of biodiversity in small lowland water bodies altered by high anthropogenic pressure and climate-related impacts, such as the Šijanec gravel pit [67,68]. Previous studies on the Drava River lowland did not consider the importance of small water bodies within the whole alluvial system [69,70]. Due to hydromorphological characteristics of the catchment, Šijanec receives a high nitrogen input from the Drava River aquifer via groundwater recharging [41]. The high concentration of nitrates in the Drava River groundwater system, with an average of 60.9 mg L−1, was observed by Marković et al. [41]. Nitrogen compounds can easily percolate through the soil into the groundwater either from direct terrestrial runoff or with rainfall or irrigation water [71]. A higher nitrogen concentration was noted during the colder seasons as a consequence of the recorded decrease in the phytoplankton biomass and abundance, as well as the rise of the groundwater level due to an increase in precipitation [72,73], whereas nitrogen concentrations in the groundwater dropped during the warmer periods following the decrease of precipitation, as confirmed by the PCA analysis. Since nitrogen fertilizers are widely used in agriculture to increase crop production, the cropping practices and soil texture have been found to influence the extent of nitrate leaching [74].
The nutrient-based indication of eutrophic conditions in the gravel pit was further supported by high values of phytoplankton biomass and bacterial density throughout the investigated period, especially during the summer months, when the lowest NO3 concentration (June) and the highest concentrations of NH4+ and NO2 (July) were detected. This occurrence of elevated nitrogen is presumably occurring as an effect of the cropping season, e.g., fertilizer application through intense irrigation [75]. This is particularly true for nitrates, which are normally assumed not to be absorbed by soil particles and are therefore easily leached, in which case the nitrate distribution should follow the wetting front [76,77]. As found by Paredes et al. [78], high intra- and interannual hydrological fluctuations influenced nitrate occurrence in freshwater streams and ponds, with the main source of nitrate linked to agricultural practices and the use of both organic and synthetic fertilizers. Since it can be rapidly oxygenized, the concentration of NO2 is usually deficient [79], as was exhibited in this study. Most of the nitrogen uptake in shallow eutrophic systems is the form of nitrate, which has a positive effect on the growth of phytoplankton biomass [80,81]. A sample from June had the lowest concentrations of NO3, TN, TIC, and logpCO2, but the highest temperature and Si values. These conditions were characterized by the dominance of cyanobacteria Limnotrix redekei and centric diatom Aulacoseira muzzanensis. Limnotrix redekei, a species characteristic of eutrophic shallow water bodies used for recreation and fishing, often shows domination in spring and summer periods with co-occurrence of centric diatoms [82]. This species is known for its ability to adapt to low-light, cold conditions and capability to overwinter in considerable densities under the long-term ice- and snow-cover [83,84,85,86]. Aulacoseira muzzanensis, a species adapted to live in turbid and nutrient-rich waters [87], was described in a hyper-eutrophic lake (Lago di Muzzano) located in the Tessin region of Switzerland [88], which suffers from periodic Microcystis algal blooms [89,90,91,92]. During lower light conditions, both taxa can occur concomitantly with the Microcystis species in quantities capable of eliminating other phytoplankton taxa [93]. A sample from July in the grouped composed of summer and autumn samples was characterized by maximum concentrations of nitrites and ammonium. This suggested enhanced growth conditions for specific algal groups under a higher NH4+ supply, which was also evident by a recorded high concentration of the chlorophyll a. Certain phytoplankton taxa, such as cyanobacteria and chlorophytes, prefer a high supply of energetically favorable NH4+ [94], as was also noted in our investigation. These conditions can also inhibit NO3 uptake for other taxa, such as large diatoms [94,95,96]. The most abundant chlorophyte was Scenedesmus quadricauda, a small coenobium-forming and ammonium-tolerant species [97,98,99] with higher uptake abilities for ammonium under nitrogen limitation than species of the genus Microcystis and with competitive superiority in the large-pulse, low-frequency nutrient recharging [100], as was present in our study area. With regard to size, small algae can be more competitive than larger species, because they have high surface area to volume ratios, resulting in greater specific uptake and growth under low nutrient concentrations, when even slight NH4+ additions can be enough to promote the growth of small algae [101]. Secondly, the most dominant species during the summer period was Microcystis spp., with the culmination of dominance (surface bloom) in August, which was connected with the maximum of phytoplankton biomass and number of cells per liter. Microcystis is a non-N2-fixing cyanobacteria, which dominates in highly eutrophicated, stratified ponds, rivers, or lakes receiving elevated N loadings, since its growth is dependent on nitrogen supplies [6]. A strong bloom of Microcystis spp. was associated with the warmer period, low nitrogen concentrations, and the subsequent water level drawdown [102]. Cyanobacteria have the ability to adapt to different environments by adjusting their light harvesting and carbon fixation mechanisms. Adversely, a high rate of photosynthesis induced by the Microcystis bloom can considerably reduce dissolved CO2 concentrations and drive up the pH value [103]. Furthermore, Microcystis favors more alkaline conditions as a competitive advantage over other phytoplankton groups [104,105]. Along with the Microcystis bloom, the molecular approach detected the cyanobacterium of the genus Cyanobium, which was dominant in August. Cyanobium is a picocyanobacterial genus with a presumably significant role in the functioning of aquatic ecosystems, but rather hard to detect microscopically due to its size and taxonomical obscurity. Both Microcystis and Cyanobium genera are less demanding on nutrients and generally demonstrate summer peaks when the concentrations of nitrogen compounds are usually lower [106]. As detected by the molecular approach, the phylum Planctomycetota was subdominant in the bacterial community in June. Members of the phylum Planctomycetota have been found in a variety of environments, including freshwater [107], and are known for their role in the anaerobic oxidation of ammonium (anammox), as part of the biogeochemical nitrogen cycle [108,109,110,111,112]. In oxygen-limited systems, such as biofilm aggregates, the planctomycete anammox bacteria live closely associated with aerobic ammonium oxidizers. The aerobic ammonium-oxidizing bacteria consume oxygen at the outside of the biofilm, thus keeping the inside anoxic for the anammox bacteria. Together, they create conditions in which they can convert ammonium directly into dinitrogen gas. Anammox bacteria can contribute significantly to the loss of fixed nitrogen in both natural and anthropogenic-influenced ecosystems [110,112,113]. The members of the phylum Bacteroidota recorded in July are typical for freshwater environments [114]. Their dominance correlated with OTUs taxonomically assigned in eukaryotic biflagellate cryptophytes from Cryptomonas genera, and with chlorophyte Scenedesmus quadricauda, as recorded by the morphological approach. The increase in their abundance correlated with higher algal concentrations, presumably due to their ability to establish a mutualistic relationship on the algal cell surface [115]. High densities of the Cryptomonas genus can occur following the period of nutrient enrichment [6]. During the colder period, the composition of the phytoplankton switched from cyanobacteria to a diatom-dominated community characterized by the genus Ulnaria, whose winter blooms require both biogenic silica for the formation of their outer cell wall structures (frustules) and lower basic pH conditions, as not to corrode them [116] and decrease the growth rate [117]. This resulted in a threefold drop in the SiO2 concentration which was presumably consumed in the building of their frustules [118,119]. Diatoms are characterized as effective nitrate utilizers with high preferences to NO3 uptake [120]. Also subdominant in the colder period, together with diatoms, were dinoflagellate species, which were reported to have significant NO3 uptake rates [121]. A dense bloom of colonial chrysophyte Dinobryon spp. characterized the phytoplankton community in February and March. As adaptations to the lower temperature, the winter blooms of Dinobryon could indicate enhanced nutrient loading [6], but also the ability to obtain nutrients from bacteria by mixotrophy [122,123]. Members of Proteobacteria, as the most dominant bacterial phylum in the samples from November and December, are ubiquitous in freshwater environments, specifically in eutrophic conditions with high phytoplankton numbers [114,124]. Actinobacteriota were present throughout the investigated period, with increased abundance in October and January during the low abundance of cyanobacteria. This is plausibly correlated with the sensitivity of Actinobacteriota to conditions prevailing during the cyanobacterial blooms, such as high organic matter, inorganic nutrient availability, and high temperatures, under which the highly streamlined actinobacterial cells cannot compete [125]. Actinobacteriota can thrive under oligotrophic conditions due to their small size, high surface area-to-volume ratio, and enhanced capacity for efficient nutrient acquisition through high-affinity, broad-specific uptake systems [126]. Another important role of Actinobacteriota in freshwater habitats is connected to the proton-pumping rhodopsins (actinorhodopsins), thus revealing a photoheterotrophic lifestyle [127]. All these traits suggest that Actinobacteriota might serve as sentinels of impending ecological damage and have the potential to become standards of ecological freshwater quality [125].
Some of the OTUs were taxonomically assigned to species usually hard to detect under the microscope. These species included colonial chrysophyte Uroglenopsis americana (G.N.Calkins) Lemmermann, thecate dinoflagellate Asulcocephalium miricentonis Kazuya Takahashi, Moestrup & M. Iwataki and small green algae Actinochloris sphaerica Korschikov, Meyerella planktonica Fawley & K. P. Fawley, Wislouchiella planctonica Skvortzov and Chloromonas subdivisa (Pascher & Jahoda) Gerloff & Ettl. They were all detected in the winter samples, except for Asulcocephalium miricentonis and Wislouchiella planctonica, which were detected in June. Besides type locality, no further ecological data were available on Asulcocephalium miricentonis, a species described in a temperate freshwater artificial pond in northeastern Japan [128]. Therefore, this record presents a contribution to the ecological conditions in which the species likely occurs. Wislouchiella planctonica is associated with man-made reservoirs and lentic freshwater habitats with eutrophic conditions [129,130], which is in line with our findings. Uroglenopsis americana was detected in February together with Dinobryon spp., due to its ability to compete for nutrients during the colder period, unlike algal species in eutrophic conditions [131]. The picoplanktonic species Meyerella planktonica is a major component of aquatic systems and a significant primary producer regularly occurring during winter [132,133]. Some of the species from the Chloromonas genera were found in snow samples [134], as was the case with the species Chloromonas subdivisa detected during snowy winter conditions in Šijanec. Actinochloris sphaerica is a cosmopolitan species mostly recorded in soil cultures and puddles [135]. Since very scarce information is available on all these species, the presented results also contribute to elucidating their ecological preferences.
Both approaches showed variations in diversity richness. Based on the morphological approach, the maximum value of alpha diversity was recorded in June, presumably due to the favorable conditions for algal growth. The minimum value of alpha diversity was detected in February, due to the lower number of algal species, whereas by using the molecular approach, higher values of alpha diversity were detected in the colder period, which can be associated with two possible causes: (1) when cell abundances of specific taxa in the water sample drop below a specific threshold, they can still be detectable with the molecular approach, but may not be found by microcopy; and (2) the resting stages of some algal species cannot be identified and assigned correctly by microscopy, but might be more easily recorded by the molecular approach [136]. Moreover, the lowest value of alpha diversity detected in August with the molecular approach could be correlated with the cyanobacterial bloom. Some cyanobacteria are known to produce toxins and cyclic peptides which can inhibit regulatory enzymes in eukaryotic cells, thus causing PCR inhibition [137,138]. Similar variations were detected in bacterial alpha diversity, with a maximum in July and low values throughout August, September, and October. This finding also corresponds to cyanobacterial bloom, which can inhibit the stabilization of microbial diversity [139]. Surprisingly, the morphological approach, as a basic descriptive method, succeeded in recognizing a higher microalgal diversity in Šijanec than the molecular approach, which is commonly considered a more powerful detector tool [140]. Events such as cyanobacterial blooms or discrepancies in the various DNA extraction methods can also have a discernible impact on the certainty of the community analysis via the molecular approach. Eukaryotic algae have a large range of cell wall structures, thus imposing challenges to the unbiased, uniform, and universal extraction of nucleic acids from such communities [138]. In spite of this, the molecular approach proved far more effective in discerning small-sized eukaryotic algae and cyanobacterial taxa, which are generally hard to detect with microscopy due to the scarcity of taxonomic knowledge and limitations of resolving power. Molecular methods showed that they can be successfully used to complement the morphological approach for assessing the microbial community’s structure [138,141], especially in these kinds of eutrophic environments.

5. Conclusions

Despite the size, small water bodies like the Šijanec gravel pit play a key role as buffer zones within alluvial areas of large rivers. Due to the hydromorphological characteristics of the catchment, Šijanec receives a high nitrate input directly from the groundwater recharging. Nitrogen compounds likely control the phytoplankton biomass, thereby influencing the complete microbial community’s structure. The integration of morphological and molecular approaches facilitates the comprehensive assessment of the microbial community’s structure. Interdisciplinary approaches can be successfully used to elucidate the ecological preferences of microbial species and the prediction and prevention of algal blooms. The study emphasizes the importance of small water bodies in maintaining the state of water ecosystems and stresses the need for their enlistment in biomonitoring actions.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4441/13/2/115/s1. Document S1: includes the list of commands with related parameters used in the OBITools package.

Author Contributions

Formal analysis, investigation, data curation, writing—Original draft preparation, visualization, A.K.; chemical analysis, investigation, writing—Review and editing, T.M.; investigation, validation, writing—Review and editing, P.Ž.; data curation, writing—Review and editing, K.K.; description of the study area, investigation, writing—Review and editing, I.K.; methodology, investigation, writing—Review and editing, S.O.; software, writing—Review and editing, E.K.; validation, writing—Review and editing, V.F.; conceptualization, investigation, supervision, validation, writing—Review and editing M.G.U. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Croatian Scientific Foundation (HRZZ), grant number HRZZ-IP-2016-06-5356, and by the COST DNAqua-Net grant (CA15219–“eDNA metabarcoding in providing assessment of microbial eukaryotic biodiversity in small standing-water ecosystem”). SO and KK were partially supported by the project STIM – REI, Contract Number: KK.01.1.1.01.0003., funded by the European Union through the European Regional Development Fund—the Operational Programme Competitiveness and Cohesion 2014–2020 (KK.01.1.1.01).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in: Kulaš, A.; Marković, T.; Žutinić, P.; Kajan, K.; Karlović, I.; Orlić, S.; Keskin, E.; Filipović, V.; Gligora Udovič, M. Succession of Microbial Community in a Small Water Body within the Alluvial Aquifer of a Large River Water. Water 2021, 13, 115. https://doi.org/10.3390/w13020115.

Acknowledgments

We thank the evolutionary Genetics Laboratory (eGL) research team for their support and help in bioinformatics analysis.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Thingstad, T.; Sakshaug, E. Control of phytoplankton growth in nutrient recycling ecosystems. Theory and terminology. Mar. Ecol. Prog. Ser. 1990, 63, 261–272. [Google Scholar] [CrossRef]
  2. Chan, J.L.; Mantzoros, C.S. Role of leptin in energy-deprivation states: Normal human physiology and clinical implications for hypothalamic amenorrhoea and anorexia nervosa. Lancet 2005, 366, 74–85. [Google Scholar] [CrossRef]
  3. Green, P.; Vörösmarty, C.J.; Meybeck, M.; Galloway, J.N.; Peterson, B.J.; Boyer, E.W. Pre-industrial and contemporary fluxes of nitrogen through rivers: A global assessment based on typology. Biogeochemistry 2004, 68, 71–105. [Google Scholar] [CrossRef]
  4. Li, Z.G.; Lin, B.-L.; Sagisaka, M.; Yang, P.; Wu, W. Global-scale modelling of potential changes in terrestrial nitrogen cycle from a growing nitrogen deposition. Proc. Environ. Sci. 2012, 13, 1057–1068. [Google Scholar] [CrossRef] [Green Version]
  5. Galloway, J.N.; Cowling, E.B.; Seitzinger, S.; Socolow, R.H. Reactive nitrogen: Too much of a good thing? Ambio 2002, 31, 60–63. [Google Scholar] [CrossRef]
  6. Paerl, H.W.; Fulton, R.S.; Moisander, P.H.; Dyble, J. Harmful freshwater algal blooms, with an emphasis on Cyanobacteria. Sci. World J. 2001, 1, 76–113. [Google Scholar] [CrossRef]
  7. Erisman, J.W.; Galloway, J.; Seitzinger, S.; Bleeker, A.; Butterbach-Bahl, K. Reactive nitrogen in the environment and its effect on climate change. Curr. Opin. Environ. Sustain. 2011, 3, 281–290. [Google Scholar] [CrossRef] [Green Version]
  8. Dupas, R.; Delmas, M.; Dorioz, J.-M.; Garnier, J.; Moatar, F.; Gascuel-Odoux, C. Assessing the impact of agricultural pressures on N and P loads and eutrophication risk. Ecol. Indic. 2015, 48, 396–407. [Google Scholar] [CrossRef]
  9. Erisman, J.W.; Galloway, J.N.; Seitzinger, S.; Bleeker, A.; Dise, N.B.; Petrescu, A.M.R.; Leach, A.M.; de Vries, W. Consequences of human modification of the global nitrogen cycle. Philos. Trans. R. Soc. B Biol. Sci. 2013, 368, 20130116. [Google Scholar] [CrossRef] [Green Version]
  10. Mekonnen, M.M.; Hoekstra, A.Y. Global gray water footprint and water pollution levels related to anthropogenic nitrogen loads to fresh water. Environ. Sci. Technol. 2015, 49, 12860–12868. [Google Scholar] [CrossRef]
  11. le Doux, E.; Gomez, E.; Monget, J.; Viavattene, C.; Viennot, P.; Ducharne, A.; Benoit, M.; Mignolet, C.; Schott, C.; Mary, B. Agriculture and groundwater nitrate contamination in the Seine basin. The STICS–MODCOU modelling chain. Sci. Total Environ. 2007, 375, 33–47. [Google Scholar] [CrossRef] [PubMed]
  12. Mengel, K.; Hütsch, B.; Kane, Y. Nitrogen fertilizer application rates on cereal crops according to available mineral and organic soil nitrogen. Eur. J. Agron. 2006, 24, 343–348. [Google Scholar] [CrossRef]
  13. Sete, P.B.; Comin, J.J.; Ciotta, M.N.; Salume, J.A.; Thewes, F.; Brackmann, A.; Toselli, M.; Nava, G.; Rozane, D.-E.; Loss, A.; et al. Nitrogen fertilization affects yield and fruit quality in pear. Sci. Hortic. 2019, 258, 108782. [Google Scholar] [CrossRef]
  14. Jégo, G.; Sánchez-Pérez, J.M.; Justes, E. Predicting soil water and mineral nitrogen contents with the STICS model for estimating nitrate leaching under agricultural fields. Agric. Water Manag. 2012, 107, 54–65. [Google Scholar] [CrossRef] [Green Version]
  15. Bernard-Jannin, L.; Sun, X.; Teissier, S.; Sauvage, S.; Sánchez-Pérez, J.M. Spatio-temporal analysis of factors controlling nitrate dynamics and potential denitrification hot spots and hot moments in groundwater of an alluvial floodplain. Ecol. Eng. 2017, 103, 372–384. [Google Scholar] [CrossRef] [Green Version]
  16. Singer, G.; Besemer, K.; Schmitt-Kopplin, P.; Hödl, I.; Battin, T.J. Physical heterogeneity increases biofilm resource use and its molecular diversity in stream mesocosms. PLoS ONE 2010, 5, e9988. [Google Scholar] [CrossRef]
  17. Cardinale, B.J.; Duffy, J.E.; Gonzalez, A.; Hooper, D.U.; Perrings, C.; Venail, P.; Narwani, A.; Mace, G.M.; Tilman, D.; Wardle, D.A.; et al. Biodiversity loss and its impact on humanity. Nat. Cell Biol. 2012, 486, 59–67. [Google Scholar] [CrossRef]
  18. Besemer, K.; Singer, G.; Quince, C.; Bertuzzo, E.; Sloan, W.; Battin, T.J. Headewaters are critical reservoirs of microbial diversity for fluvial networks. Proc. R. Soc. B 2013, 280, 20131760. [Google Scholar] [CrossRef]
  19. Widder, S.; Besemer, K.; Singer, G.A.; Ceola, S.; Bertuzzo, E.; Quince, C.; Sloan, W.T.; Rinaldo, A.; Battin, T.J. Fluvial network organization imprints on microbial co-occurrence networks. Proc. Natl. Acad. Sci. USA 2014, 111, 12799–12804. [Google Scholar] [CrossRef] [Green Version]
  20. Peterson, B.J.; Wollheim, W.M.; Mulholland, P.J.; Webster, J.R.; Meyer, J.L.; Tank, J.L.; Martí, E.; Bowden, W.B.; Valett, H.M.; Hershey, A.E.; et al. Control of nitrogen export from watersheds by headwater streams. Science 2001, 292, 86–90. [Google Scholar] [CrossRef]
  21. Paerl, H.W. Mitigating harmful cyanobacterial blooms in a human- and climatically impacted world. Life 2014, 4, 988–1012. [Google Scholar] [CrossRef] [Green Version]
  22. Pal, R.; Choudhury, A.K. An Introduction to Phytoplanktons: Diversity and Ecology; Springer Science and Business Media LLC: Berlin/Heidelberg, Germany, 2014; pp. 25–52. [Google Scholar]
  23. Herrero, A.; Muro-Pastor, A.M.; Flores, E. Nitrogen control in Cyanobacteria. J. Bacteriol. 2001, 183, 411–425. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Stal, L. Cyanobacteria: Diversity and versatility. In Algae and Cyanobacteria in Extreme Environments: Cellular Origin, Life in Extreme Habitats and Astrobiology; Sedbach, J., Ed.; Springer: Dordrecht, The Netherlands, 2007; Volume 11, pp. 659–680. [Google Scholar]
  25. Sellner, K.G. Physiology, ecology, and toxic properties of marine cyanobacteria blooms. Limnol. Oceanogr. 1997, 42, 1089–1104. [Google Scholar] [CrossRef] [Green Version]
  26. Hemida, M.; Ohyam, T. Nitrogen fixing Cyanobacteria: Future prospect. Nitrogen Fixat. 2014, 23–48. [Google Scholar] [CrossRef] [Green Version]
  27. Bakker, E.S.; Hilt, S. Impact of water-level fluctuations on cyanobacterial blooms: Options for management. Aquat. Ecol. 2016, 50, 485–498. [Google Scholar] [CrossRef] [Green Version]
  28. Hamilton, L.C. Public awareness of the scientific consensus on climate. SAGE Open 2016, 6. [Google Scholar] [CrossRef]
  29. van Gremberghe, I.; Leliaert, F.; Mergeay, J.; Vanormelingen, P.; van der Gucht, K.; Debeer, A.-E.; Lacerot, G.; de Meester, L.; Vyverman, W. Lack of phylogeographic structure in the freshwater Cyanobacterium Microcystis aeruginosa suggests global dispersal. PLoS ONE 2011, 6, e19561. [Google Scholar] [CrossRef] [Green Version]
  30. Gijzen, H.J.; Mulder, A. The nitrogen cycle out of balance. Water 2001, 21, 38–40. [Google Scholar]
  31. Moss, B.; Barker, T.; Stephen, D.; Williams, A.E.; Balayla, D.J.; Beklioglu, M.; de Carvalho, L.P.S. Consequences of reduced nutrient loading on a lake system in a lowland catchment: Deviations from the norm? Freshw. Biol. 2005, 50, 1687–1705. [Google Scholar] [CrossRef]
  32. Rosset, V.; Angélibert, S.; Arthaud, F.; Bornette, G.; Robin, J.; Wezel, A.; Vallod, D.; Oertli, B. Is eutrophication really a major impairment for small waterbody biodiversity? J. Appl. Ecol. 2014, 51, 415–425. [Google Scholar] [CrossRef]
  33. Menetrey, N.; Sager, L.; Oertli, B.; Lachavanne, J.-B. Looking for metrics to assess the trophic state of ponds. Macroinvertebrates and amphibians. Aquat. Conserv. Mar. Freshw. Ecosyst. 2005, 15, 653–664. [Google Scholar] [CrossRef]
  34. Søndergaard, M.; Jeppesen, E.; Jensen, J.P.; Amsinck, S.L. Water Framework Directive: Ecological classification of Danish lakes. J. Appl. Ecol. 2005, 42, 616–629. [Google Scholar] [CrossRef]
  35. Williams, P.; Whitfield, M.; Biggs, J.; Bray, S.; Fox, G.; Nicolet, P.; Sear, D. Comparative biodiversity of rivers, streams, ditches and ponds in an agricultural landscape in Southern England. Biol. Conserv. 2004, 115, 329–341. [Google Scholar] [CrossRef]
  36. Gvozdić, V.; Brana, J.; Puntarić, D.; Vidosavljević, D.; Roland, D. Changes in the Lower Drava River water quality parameters over 24 years. Arch. Ind. Hyg. Toxicol. 2011, 62, 325–333. [Google Scholar] [CrossRef] [Green Version]
  37. Villanueva, J.; Rosell, A.; Grimalt, J.O.; Navarro, A.; Rosell-Melé, A. Chemical characterization of polycyclic aromatic hydrocarbon mixtures in uncontrolled hazardous waste dumps. Chemosphere 1991, 22, 317–326. [Google Scholar] [CrossRef]
  38. Navarro, A.; Carbonell, M. Assessment of groundwater contamination caused by uncontrolled dumping in old gravel quarries in the Besòs aquifers (Barcelona, Spain). Environ. Geochem. Health 2007, 30, 273–289. [Google Scholar] [CrossRef]
  39. Preziosi, E.; Frollini, E.; Zoppini, A.; Ghergo, S.; Melita, M.; Parrone, D.; Rossi, D.; Amalfitano, S. Disentangling natural and anthropogenic impacts on groundwater by hydrogeochemical, isotopic and microbiological data: Hints from a municipal solid waste landfill. Waste Manag. 2019, 84, 245–255. [Google Scholar] [CrossRef]
  40. Ondrasek, G.; Begić, H.B.; Zovko, M.; Filipović, L.; Meriño-Gergichevich, C.; Savić, R.; Rengel, Z. Biogeochemistry of soil organic matter in agroecosystems & environmental implications. Sci. Total Environ. 2019, 658, 1559–1573. [Google Scholar] [CrossRef]
  41. Marković, T.; Karlović, I.; Tadić, M.P.; Larva, O. Application of stable water isotopes to improve conceptual model of alluvial aquifer in the Varaždin area. Water 2020, 12, 379. [Google Scholar] [CrossRef] [Green Version]
  42. Mandel, S.; Shiftan, Z.L. Groundwater Resources: Investigation and Development, 1st ed.; Academic Press: New York, NY, USA, 1980; pp. 180–200. [Google Scholar]
  43. Domenico, P.A.; Schwartz, F.W. Physical and Chemical Hydrogeology, 2nd ed.; John Willey and Sons: New York, NY, USA, 1990. [Google Scholar]
  44. Parkhurst, D.L.; Appelo, C. User’s guide to PHREEQC (Version 2): A computer program for speciation, batch-reaction, one-dimensional transport, and inverse geochemical calculations. Water Res. Investig. Rep. 1999, 99, 4259. [Google Scholar]
  45. ISO/TC 147 Water Quality. Available online: https://www.iso.org/committee/52834/x/catalogue/ (accessed on 16 April 2018).
  46. Utermöhl, H. Zur vervollkomnung der quantitativen phytoplankton-methodik. Mitteilungen Internationale Vereiningung für Theoretische und Angewandte. Limnologie 1958, 9, 1–38. [Google Scholar]
  47. Stock, A.; Jürgens, K.; Bunge, J.; Stoeck, T. Protistan diversity in suboxic and anoxic waters of the Gotland Deep (Baltic Sea) as revealed by 18S rRNA clone libraries. Aquat. Microb. Ecol. 2009, 55, 267–284. [Google Scholar] [CrossRef]
  48. Stoeck, T.; Bass, D.; Nebel, M.; Christen, R.; Jones, M.D.M.; Breiner, H.-W.; Richards, T.A. Multiple marker parallel tag environmental DNA sequencing reveals a highly complex eukaryotic community in marine anoxic water. Mol. Ecol. 2010, 19, 21–31. [Google Scholar] [CrossRef] [PubMed]
  49. Amaral-Zettler, L.; McCliment, E.A.; Ducklow, H.W.; Huse, S.M. A method for studying protistan diversity using massively parallel sequencing of V9 hypervariable regions of small-subunit ribosomal RNA genes. PLoS ONE 2009, 4, e6372. [Google Scholar] [CrossRef]
  50. Herlemann, D.P.R.; Labrenz, M.; Jürgens, K.; Bertilsson, S.; Waniek, J.J.; Andersson, A.F. Transitions in bacterial communities along the 2000 km salinity gradient of the Baltic Sea. ISME J. 2011, 5, 1571–1579. [Google Scholar] [CrossRef] [Green Version]
  51. de Barba, M.; Miquel, C.; Boyer, F.; Mercier, C.; Rioux, D.; Coissac, E.; Taberlet, P. DNA metabarcoding multiplexing and validation of data accuracy for diet assessment: Application to omnivorous diet. Mol. Ecol. Resour. 2013, 14, 306–323. [Google Scholar] [CrossRef] [PubMed]
  52. Bellemain, E.; Carlsen, T.; Brochmann, C.; Coissac, E.; Taberlet, P.; Kauserud, H. ITS as an environmental DNA barcode for fungi: An in silico approach reveals potential PCR biases. BMC Microbiol. 2010, 10, 189. [Google Scholar] [CrossRef] [Green Version]
  53. Ficetola, G.F.; Coissac, E.; Zundel, S.; Riaz, T.; Shehzad, W.; Bessière, J.; Taberlet, P.; Pompanon, F. An in-silico approach for the evaluation of DNA barcodes. BMC Genom. 2010, 11, 434. [Google Scholar] [CrossRef] [Green Version]
  54. BBMap. Available online: https://sourceforge.net/projects/bbmap/ (accessed on 19 February 2019).
  55. Caporaso, J.G.; Kuczynski, J.; Stombaugh, J.; Bittinger, K.; Bushman, F.D.; Costello, E.K.; Fierer, N.; Peña, A.G.; Goodrich, J.K.; Gordon, J.I.; et al. QIIME allows analysis of high-throughput community sequencing data. Nat. Methods 2010, 7, 335–336. [Google Scholar] [CrossRef] [Green Version]
  56. Edgar, R.C.; Haas, B.J.; Clemente, J.C.; Quince, C.; Knight, R. UCHIME improves sensitivity and speed of chimera detection. Bioinformatics 2011, 27, 2194–2200. [Google Scholar] [CrossRef] [Green Version]
  57. Mahé, F.; Rognes, T.; Quince, C.; de Vargas, C.; Dunthorn, M. Swarm v2: Highly scalable and high-resolution amplicon clustering. PeerJ 2015, 3, e1420. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Altschul, S.F.; Gish, W.; Miller, W.; Myers, E.W.; Lipman, D.J. Basic local alignment search tool. J. Mol. Biol. 1990, 215, 403–410. [Google Scholar] [CrossRef]
  59. R Core Team. R: A Language and Environment for Statistical Computing; R Core Team: Vienna, Austria, 2020. [Google Scholar]
  60. Tapolczai, K.; Vasselon, V.; Lefrançois, E.; Stenger-Kovács, C.; Padisák, J.; Rimet, F. The impact of OTU sequence similarity threshold on diatom-based bioassessment: A case study of the rivers of Mayotte (France, Indian Ocean). Ecol. Evol. 2019, 9, 166–179. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Anderson, M.J. Distance-based tests for homogeneity of multivariate dispersions. Biometrics 2005, 62, 245–253. [Google Scholar] [CrossRef] [PubMed]
  62. Oksanen, J.; Blanchet, F.G.; Friendly, M.; Kindt, R.; Legendre, P.; McGlinn, D.; Minchin, P.R.; O’Hara, R.B.; Simpson, G.L.; Solymos, P.; et al. Vegan: Community Ecology Package. R Packag. Vers. 22-1. 2015, 2, 1–2. [Google Scholar]
  63. Bolpagni, R.; Poikane, S.; Laini, A.; Bagella, S.; Bartoli, M.; Cantonati, M. Ecological and conservation value of small standing-water ecosystems: A systematic review of current knowledge and future challenges. Water 2019, 11, 402. [Google Scholar] [CrossRef] [Green Version]
  64. Céréghino, R.; Boix, D.; Cauchie, H.-M.; Martens, K.; Oertli, B. The ecological role of ponds in a changing world. Hydrobiologia 2014, 723, 1–6. [Google Scholar] [CrossRef] [Green Version]
  65. Brönmark, C.; Hansson, L.-A. Environmental issues in lakes and ponds: Current state and perspectives. Environ. Conserv. 2002, 29, 290–307. [Google Scholar] [CrossRef] [Green Version]
  66. Hering, D.; Haidekker, A.; Schmidt-Kloiber, A.; Barker, T.; Buisson, L.; Graf, W.; Grenouillet, G.; Lorenz, A.; Sandin, L.; Stendera, S. Monitoring the responses of freshwater ecosystems to climate change. Clim. Chang. Impacts Freshw. Ecosyst. 2010, 84–118. [Google Scholar] [CrossRef]
  67. Boix, D.; Biggs, J.; Céréghino, R.; Hull, A.P.; Kalettka, T.; Oertli, B. Pond research and management in Europe: “Small is Beautiful”. Hydrobiologia 2012, 689, 1–9. [Google Scholar] [CrossRef] [Green Version]
  68. Bagella, S.; Gascón, S.; Filigheddu, R.; Cogoni, A.; Boix, D. Mediterranean temporary ponds: New challenges from a neglected habitat. Hydrobiologia 2016, 782, 1–10. [Google Scholar] [CrossRef]
  69. Stanković, I.; Vlahović, T.; Udovič, M.G.; Várbíró, G.; Borics, G. Phytoplankton functional and morpho-functional approach in large floodplain rivers. Hydrobiologia 2012, 698, 217–231. [Google Scholar] [CrossRef]
  70. Kovács, A.D.; Tóth, G.; Lóczy, D. Water Quality of the Lower Drava River; Springer Science and Business Media LLC: Berlin/Heidelberg, Germany, 2018; pp. 231–245. [Google Scholar]
  71. Gao, K.; Helbling, E.; Häder, D.; Hutchins, D. Responses of marine primary producers to interactions between ocean acidification, solar radiation, and warming. Mar. Ecol. Prog. Ser. 2012, 470, 167–189. [Google Scholar] [CrossRef] [Green Version]
  72. van Verseveld, W.J.; McDonnell, J.J.; Lajtha, K. The role of hillslope hydrology in controlling nutrient loss. J. Hydrol. 2009, 367, 177–187. [Google Scholar] [CrossRef]
  73. Jiang, R.; Woli, K.P.; Kuramochi, K.; Hayakawa, A.; Shimizu, M.; Hatano, R. Hydrological process controls on nitrogen export during storm events in an agricultural watershed. Soil Sci. Plant Nutr. 2010, 56, 72–85. [Google Scholar] [CrossRef] [Green Version]
  74. Serio, F.; Miglietta, P.P.; Lamastra, L.; Ficocelli, S.; Intini, F.; de Leo, F.; de Donno, A. Groundwater nitrate contamination and agricultural land use: A grey water footprint perspective in Southern Apulia Region (Italy). Sci. Total Environ. 2018, 645, 1425–1431. [Google Scholar] [CrossRef] [PubMed]
  75. Filipovic, V.; Kodešová, R.; Petosic, D. Experimental and mathematical modeling of water regime and nitrate dynamics on zero tension plate lysimeters in soil influenced by high groundwater table. Nutr. Cycl. Agroecosyst. 2012, 95, 23–42. [Google Scholar] [CrossRef]
  76. Filipović, V.; Romić, D.; Romić, M.; Borošić, J.; Filipović, L.; Mallmann, F.J.K.; Robinson, D.A. Plastic mulch and nitrogen fertigation in growing vegetables modify soil temperature, water and nitrate dynamics: Experimental results and a modeling study. Agric. Water Manag. 2016, 176, 100–110. [Google Scholar] [CrossRef] [Green Version]
  77. Hardie, M.; Ridges, J.; Swarts, N.; Close, D.C. Drip irrigation wetting patterns and nitrate distribution: Comparison between electrical resistivity (ERI), dye tracer, and 2D soil–water modelling approaches. Irrig. Sci. 2018, 36, 97–110. [Google Scholar] [CrossRef]
  78. Paredes, I.; Otero, N.; Soler, A.; Green, A.J.; Soto, D.X. Agricultural and urban delivered nitrate pollution input to Mediterranean temporary freshwaters. Agric. Ecosyst. Environ. 2020, 294, 106859. [Google Scholar] [CrossRef]
  79. Wetzel, R.G. Limnology: Lake and River Ecosystems, 3rd ed.; Academic Press: San Diego, CA, USA, 2001; pp. 239–288. [Google Scholar]
  80. Kanda, F.; Yagi, E.; Fukuda, M.; Nakajima, K.; Ohta, T.; Nakata, O. Elucidation of chemical compounds responsible for foot malodour. Br. J. Dermatol. 1990, 122, 771–776. [Google Scholar] [CrossRef]
  81. Strom, S. Novel interactions between phytoplankton and microzooplankton: Their influence on the coupling between growth and grazing rates in the sea. Hydrobiologia 2002, 480, 41–54. [Google Scholar] [CrossRef]
  82. Budzyńska, A.; Gołdyn, R.; Zagajewski, P.; Dondajewska, R.; Kowalczewska-Madura, K. The dynamics of a Planktothrix agardhii population in a shallow dimictic lake. Oceanol. Hydrobiol. Stud. 2009, 38, 1–6. [Google Scholar]
  83. Meffert, M.F. Planktic unsheathed filaments (Cyanophyceae) with polar and central gas-vacuoles. II. Biology population dynamics and biotopes of Limnothrix redekei (Van Goor) Meffert. Arch. Hydrobiol. 1989, 116, 257–282. [Google Scholar]
  84. Trifonova, I.S. Ecology and Succession of Lake Phytoplankton; Nauka Press: Leningrad, Russia, 1990. [Google Scholar]
  85. Wiedner, C.; Nixdorf, B. Success of chrysophytes, cryptophytes and dinoflagellates over blue green (cyanobacteria) during an extreme winter (1995/96) in eutrophic shallow lakes. Hydrobiologia 1998, 369, 229–235. [Google Scholar] [CrossRef]
  86. Romanov, R.E. Limnothrix redeckei (Van Goor) Meffert (Cyanoprocaryota) in the potamoplankton. Int. J. Algae 2007, 9, 105–116. [Google Scholar] [CrossRef]
  87. Licursi, M.; Sierra, M.V.; Gómez, N. Diatom assemblages from a turbid coastal plain estuary: Río de la Plata (South America). J. Mar. Syst. 2006, 62, 35–45. [Google Scholar] [CrossRef]
  88. Meister, F. Die Kieselalgen der Schweiz. In Beiträge zur Kryptogamenflora der Schweiz; Wyss, K.J., Ed.; K.J. Wyss: Bern, Switzerland, 1912; Volume 4, p. 254. [Google Scholar]
  89. Bottinelli, M. Fioriture di Cianobatteri della specie Microcystis wesenbergii nel Lago di Muzzano. Ph.D. Thesis, Tesi Sperimentale di Laurea in Scienze Naturali, Universita degli Studi di Pavia, Pavia, Italy, 1999; p. 172. [Google Scholar]
  90. Bottinelli, M.; Tonolla, M.; Forlani, G.; Crivelli, C.; Sanangelantoni, A.M.; e Peduzzi, R. Fioriture di Cianobatteri della specie Microcystis wesenbergii nel Lago di Muzzano (Svizzera). Boll. Della Soc. Tic. Di Sci. Nat. 2000, 88, 53–61. [Google Scholar]
  91. Isenburg, C.; Loizeau, J.L.; Tonolla, M.; Peduzzi, R. Aspetti limnologici e microbiologici del Laghetto di Muzzano (TI). Boll. Soc. Tic. Sci. Nat. 2000, 88, 41–51. [Google Scholar]
  92. Pedrotta, T. Analisi qualitativa degli immissari e dell’emissario del Laghetto di Muzzano. Tesi di Laurea in AGRN. Front. Ecol. Evolut. 2009, 3, 70. [Google Scholar]
  93. Reynolds, C. Ecology of Phytoplankton; Cambridge University Press: Cambridge, UK, 2006. [Google Scholar]
  94. Glibert, P.M.; Wilkerson, F.P.; Dugdale, R.C.; Raven, J.A.; Dupont, C.L.; Leavitt, P.R.; Parker, A.E.; Burkholder, J.M.; Kana, T.M. Pluses and minuses of ammonium and nitrate uptake and assimilation by phytoplankton and implications for productivity and community composition, with emphasis on nitrogen-enriched conditions. Limnol. Oceanogr. 2016, 61, 165–197. [Google Scholar] [CrossRef]
  95. Eppley, R.W.; Coatsworth, J.L.; Solórzano, L. Studies of nitrate reductase in marine phytoplankton. Limnol. Oceanogr. 1969, 14, 194–205. [Google Scholar] [CrossRef]
  96. Sommer, U. Disturbance-diversity relationships in two lakes of similar nutrient chemistry but contrasting disturbance regimes. Hydrobiologia 1993, 249, 59–65. [Google Scholar] [CrossRef]
  97. Miyazaki, T.; Honjo, Y.; Ichimura, S. Applicability of the stable isotope method using 13C and 15N simultaneously to the estimation of carbon and nitrogen assimilation in a eutrophic, freshwater lake, Lake Nakanuma, Japan. Arch. Hydrobiol. 1985, 102, 355–365. [Google Scholar]
  98. Karya, N.; van der Steen, N.; Lens, P. Photo-oxygenation to support nitrification in an algal–bacterial consortium treating artificial wastewater. Bioresour. Technol. 2013, 134, 244–250. [Google Scholar] [CrossRef]
  99. Hou, X.L.; Ma, D.L.; Yang, K.F.; Wang, R.; Rong, D.L.; Yin, X.T. The physiological responses of one algae species Scenedesmus quadricauda to ammonium and alanine. IOP Conf. Ser. Earth Environ. Sci. 2019, 344, 012167. [Google Scholar] [CrossRef]
  100. Watanabe, T.; Miyazaki, T. Maximum ammonium uptake rates of Scenedesmus quadricauda (Chlorophyta) and Microcystis novacekii (Cyanobacteria) grown under nitrogen limitation and implications for competition. J. Phycol. 1996, 32, 243–249. [Google Scholar] [CrossRef]
  101. Trommer, G.; Poxleitner, M.; Stibor, H. Responses of lake phytoplankton communities to changing inorganic nitrogen supply forms. Aquat. Sci. 2020, 82, 22. [Google Scholar] [CrossRef] [Green Version]
  102. Barone, R.; Naselli-Flores, L. Distribution and seasonal dynamics of cryptomonads in Sicilian water bodies. Hydrobiologia 2003, 502, 325–329. [Google Scholar] [CrossRef]
  103. Paerl, H.W. Mitigating toxic planktonic cyanobacterial blooms in aquatic ecosystems facing increasing anthropogenic and climatic pressures. Toxins 2018, 10, 76. [Google Scholar] [CrossRef] [Green Version]
  104. Hutchinson, G.E. The paradox of the plankton. Am. Nat. 2002, 95, 137–145. [Google Scholar] [CrossRef] [Green Version]
  105. Wilhelm, S.W.; Bullerjahn, G.S.; McKay, R.M.L. The complicated and confusing ecology of Microcystis blooms. mBio 2020, 11, 00529-20. [Google Scholar] [CrossRef] [PubMed]
  106. Belykh, O.I.; Dmitrieva, O.A.; Gladkikh, A.S.; Sorokovikova, E.G. Identification of toxigenic Cyanobacteria of the genus Microcystis in the Curonian Lagoon (Baltic Sea). Oceanology 2013, 53, 71–79. [Google Scholar] [CrossRef]
  107. Fuerst, J.A. Planctomycetes—new models for microbial cells and activities. Microb. Res. 2017, 1–27. [Google Scholar] [CrossRef]
  108. Strous, M.; Kuenen, J.G.; Jetten, M.S.M. Key physiology of anaerobic ammonium oxidation. Appl. Environ. Microbiol. 1999, 65, 3248–3250. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Jetten, M.S.; Wagner, M.; Fuerst, J.; van Loosdrecht, M.; Kuenen, G.; Strous, M. Microbiology and application of the anaerobic ammonium oxidation (‘anammox’) process. Curr. Opin. Biotechnol. 2001, 12, 283–288. [Google Scholar] [CrossRef]
  110. Jetten, M.S.; Sliekers, O.; Kuypers, M.M.M.; Dalsgaard, T.K.; van Niftrik, L.; Cirpus, I.; van de Pas-Schoonen, K.; Lavik, G.; Thamdrup, B.; le Paslier, D.; et al. Anaerobic ammonium oxidation by marine and freshwater planctomycete-like bacteria. Appl. Microbiol. Biotechnol. 2003, 63, 107–114. [Google Scholar] [CrossRef]
  111. Schmid, M.; Schmitz-Esser, S.; Jetten, M.; Wagner, M. 16S-23S rDNA intergenic spacer and 23S rDNA of anaerobic ammonium-oxidizing bacteria: Implications for phylogeny and in situ detection . Environ. Microbiol. 2001, 3, 450–459. [Google Scholar] [CrossRef]
  112. Kuypers, M.M.M.; Sliekers, A.O.; Lavik, G.; Schmid, M.; Jørgensen, B.B.; Kuenen, J.G.; Damsté, J.S.S.; Strous, M.; Jetten, M.S.M. Anaerobic ammonium oxidation by anammox bacteria in the Black Sea. Nat. Cell Biol. 2003, 422, 608–611. [Google Scholar] [CrossRef]
  113. van de Graaf, A.A.; de Bruijn, P.; Robertson, L.A.; Jetten, M.S.M.; Kuenen, J.G. Metabolic pathway of anaerobic ammonium oxidation on the basis of 15N studies in a fluidized bed reactor. Microbiology 1997, 143, 2415–2421. [Google Scholar] [CrossRef] [Green Version]
  114. Newton, P.N.; Amin, A.A.; E Bird, C.; Passmore, P.; Dukes, G.; Tomson, G.; Simons, B.; Bate, R.; Guerin, P.J.; White, N.J. The primacy of public health considerations in defining poor quality medicines. PLoS Med. 2011, 8, e1001139. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Russo, D.A.; Couto, N.; Beckerman, A.P.; Pandhal, J. A metaproteomic analysis of the response of a freshwater microbial community under nutrient enrichment. Front. Microbiol. 2016, 7, 1172. [Google Scholar] [CrossRef] [PubMed]
  116. Martin-Jezequel, V.; Hildebrand, M.; Brzezinski, M.A. Silicon metabolism in diatoms: Implications for growth. J. Phycol. 2000, 36, 821–840. [Google Scholar] [CrossRef]
  117. Lundholm, N.; Moestrup, Ø.; Kotaki, Y.; Hoef-Emden, K.; Scholin, C.; Miller, P. Inter- and intraspecific variation of the Pseudo-nitzschia delicatissima complex (Bacillariophyceae) illustrated by rRNA probes, morphological data and phylogenetic analyses. J. Phycol. 2006, 42, 464–481. [Google Scholar] [CrossRef]
  118. Nelson, D.M.; Dortch, Q. Silicic acid depletion in the plume of the Mississippi River and limitation of Si availability to diatoms in the northern Gulf of Mexico: Evidence from kinetic studies in spring and summer. Mar. Ecol. Prog. Ser. 1996, 136, 163–178. [Google Scholar] [CrossRef] [Green Version]
  119. Gibson, C.E.; Wang, G.; Foy, R.H. Silica and diatom growth in Lough Neagh: The importance of internal recycling. Freshw. Biol. 2000, 45, 285–293. [Google Scholar] [CrossRef]
  120. Glibert, P.M.; Wilkerson, F.P.; Dugdale, R.C.; Parker, A.E.; Alexander, J.A.; Blaser, S.; Murasko, S. Phytoplankton communities from San Francisco Bay Delta respond differently to oxidized and reduced nitrogen substrates-even under conditions that would otherwise suggest nitrogen sufficiency. Front. Mar. Sci. 2014, 1, 17. [Google Scholar] [CrossRef] [Green Version]
  121. Dortch, Q. The interaction between ammonium and nitrate uptake in phytoplankton. Mar. Ecol. Prog. Ser. 1990, 61, 183–201. [Google Scholar] [CrossRef]
  122. Kamjunke, N.; Henrichs, T.; Gaedke, U. Phosphorus gain by bacterivory promotes the mixotrophic flagellate Dinobryon spp. during re-oligotrophication. J. Plankton Res. 2006, 29, 39–46. [Google Scholar] [CrossRef] [Green Version]
  123. Žutinić, P.; Gligora Udovič, M.; Kralj Borojević, K.; Plenković-Moraj, A.; Padisák, J. Morpho-functional classifications of phytoplankton assemblages of two deep karstic lakes. Hydrobiologia 2014, 740, 147–166. [Google Scholar] [CrossRef] [Green Version]
  124. Eiler, A.; Bertilsson, S. Flavobacteria blooms in four eutrophic lakes: Linking population dynamics of freshwater bacterioplankton to resource availability. Appl. Environ. Microbiol. 2007, 73, 3511–3518. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Ghai, R.; Mizuno, C.M.; Picazo, A.; Camacho, A.; Rodriguez-Valera, F. Key roles for freshwater Actinobacteria revealed by deep metagenomic sequencing. Mol. Ecol. 2014, 23, 6073–6090. [Google Scholar] [CrossRef] [PubMed]
  126. Lauro, F.M.; McDougald, D.; Thomas, T.; Williams, T.J.; Egan, S.; Rice, S.; DeMaere, M.Z.; Ting, L.; Ertan, H.; Johnson, J.; et al. The genomic basis of trophic strategy in marine bacteria. Proc. Natl. Acad. Sci. USA 2009, 106, 15527–15533. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Sharma, A.K.; Sommerfeld, K.; Bullerjahn, G.S.; Matteson, A.R.; Wilhelm, S.W.; Jezbera, J.; Brandt, U.; Doolittle, W.F.; Hahn, M.W. Actinorhodopsin genes discovered in diverse freshwater habitats and among cultivated freshwater Actinobacteria. ISME J. 2009, 3, 726–737. [Google Scholar] [CrossRef] [PubMed]
  128. Takahashi, K.; Moestrup, Ø.; Jordan, R.W.; Iwataki, M. Two new freshwater Woloszynskioids Asulcocephalium miricentonis gen. et sp. nov. and Leiocephalium pseudosanguineum gen. et sp. nov. (Suessiaceae, Dinophyceae) lacking an apical furrow apparatus. Protist 2015, 166, 638–658. [Google Scholar] [CrossRef] [PubMed]
  129. Reveal, J.L. Wislouchiella planctonica Skvortz. (Chlorophyta, Volvocales), a new algal record for Nevada. Great Basin Nat. 1969, 29, 3–4. [Google Scholar]
  130. Broady, P.A.; Flint, E.A.; Nelson, W.A.; Cassie Cooper, V.; de Winton, M.D.; Novis, P.M. Phylum Chlorophyta and Charophyta: Green algae. In New Zealand Inventory of Biodiversity, Kingdoms Bacteria, Protozoa, Chromista, Plantae, Fungi; Gordon, D.P., Ed.; Canterbury University Press: Christchurch, New Zealand, 2012; Volume 3, pp. 347–381. [Google Scholar]
  131. Felip, M.; Sattler, B.; Psenner, R.; Catalan, J. Highly active microbial communities in the ice and snow cover of high mountain lakes. Appl. Environ. Microbiol. 1995, 61, 2394. [Google Scholar] [CrossRef] [Green Version]
  132. Callieri, C.; Stockner, J.G. Freshwater autotrophic picoplankton: A review. J. Limnol. 2002, 61, 1–14. [Google Scholar] [CrossRef] [Green Version]
  133. Fawley, M.W.; Fawley, K.P.; Buchheim, M.A. Molecular diversity among communities of freshwater microchlorophytes. Microb. Ecol. 2004, 48, 489–499. [Google Scholar] [CrossRef]
  134. Muramoto, K.; Nakada, T.; Shitara, T.; Hara, Y.; Nozaki, H. Re-examination of the snow algal species Chloromonas miwae (Fukushima) Muramoto et al., comb. nov. (Volvocales, Chlorophyceae) from Japan, based on molecular phylogeny and cultured material. Eur. J. Phycol. 2010, 45, 27–37. [Google Scholar] [CrossRef] [Green Version]
  135. Mikhailyuk, T.; Glaser, K.; Tsarenko, P.; Demchenko, E.; Karsten, U. Composition of biological soil crusts from sand dunes of the Baltic Sea coast in the context of an integrative approach to the taxonomy of microalgae and cyanobacteria. Eur. J. Phycol. 2019, 54, 263–290. [Google Scholar] [CrossRef]
  136. Medinger, R.; Nolte, V.; Pandey, R.V.; Jost, S.; Ottenwälder, B.; Schlötterer, C.; Boenigk, J. Diversity in a hidden world: Potential and limitation of next-generation sequencing for surveys of molecular diversity of eukaryotic microorganisms. Mol. Ecol. 2010, 19, 32–40. [Google Scholar] [CrossRef] [Green Version]
  137. Mackintosh, C.; Beattie, K.A.; Klumpp, S.; Cohen, P.; Codd, G.A. Cyanobacterial microcystin-LR is a potent and specific inhibitor of protein phosphatases 1 and 2A from both mammals and higher plants. FEBS Lett. 1990, 264, 187–192. [Google Scholar] [CrossRef] [Green Version]
  138. Eland, L.E.; Davenport, R.J.; Mota, C. Evaluation of DNA extraction methods for freshwater eukaryotic microalgae. Water Res. 2012, 46, 5355–5364. [Google Scholar] [CrossRef] [PubMed]
  139. Rashidan, K.K.; Bird, D.F. Role of predatory bacteria in the termination of a cyanobacterial bloom. Microb. Ecol. 2001, 41, 97–105. [Google Scholar] [CrossRef] [PubMed]
  140. Ghosh, S.; Love, N.G. Application of rbcL based molecular diversity analysis to algae in wastewater treatment plants. Bioresour. Technol. 2011, 102, 3619–3622. [Google Scholar] [CrossRef]
  141. van Elsas, J.D.; Boersma, F.G.H. A review of molecular methods to study the microbiota of soil and the mycosphere. Eur. J. Soil Biol. 2011, 47, 77–87. [Google Scholar] [CrossRef]
Figure 1. The geographical position of the study area with the location of the sampling area—Šijanec gravel pit—and indication of the groundwater flow direction.
Figure 1. The geographical position of the study area with the location of the sampling area—Šijanec gravel pit—and indication of the groundwater flow direction.
Water 13 00115 g001
Figure 2. Principal Component Analysis (PCA) ordination of the environmental variables during the investigated period.
Figure 2. Principal Component Analysis (PCA) ordination of the environmental variables during the investigated period.
Water 13 00115 g002
Figure 3. Phytoplankton: (a) total biomass and abundance; (b) chlorophyll a concentration during the investigated period.
Figure 3. Phytoplankton: (a) total biomass and abundance; (b) chlorophyll a concentration during the investigated period.
Water 13 00115 g003
Figure 4. Alpha diversity of the richness, Shannon, and Simpson indices of the phytoplankton community inferred from the morphological and molecular approach during the investigated period ((a) = morphological approach, (b) = molecular approach).
Figure 4. Alpha diversity of the richness, Shannon, and Simpson indices of the phytoplankton community inferred from the morphological and molecular approach during the investigated period ((a) = morphological approach, (b) = molecular approach).
Water 13 00115 g004
Figure 5. Ordination diagram of the canonical correspondence analysis (CCA) of the phytoplankton community inferred from the morphological approach in correlation to environmental variables during the investigated period.
Figure 5. Ordination diagram of the canonical correspondence analysis (CCA) of the phytoplankton community inferred from the morphological approach in correlation to environmental variables during the investigated period.
Water 13 00115 g005
Figure 6. Alpha diversity of the richness, Shannon, and Simpson indices of the bacterial community inferred from the molecular approach during the investigated period.
Figure 6. Alpha diversity of the richness, Shannon, and Simpson indices of the bacterial community inferred from the molecular approach during the investigated period.
Water 13 00115 g006
Figure 7. Ordination diagram of the canonical correspondence analysis (CCA) of the bacterial community inferred from the molecular approach in correlation with environmental variables during the investigated period.
Figure 7. Ordination diagram of the canonical correspondence analysis (CCA) of the bacterial community inferred from the molecular approach in correlation with environmental variables during the investigated period.
Water 13 00115 g007
Table 1. Ranges of environmental variables in the Šijanec gravel pit during the investigated period.
Table 1. Ranges of environmental variables in the Šijanec gravel pit during the investigated period.
VariableMinMaxMeanMedSD
T (°C)1.224.213.011.59.1
EC (μS cm−1)25249734731699
pH7.839.488.498.230.60
DO (mg L−1)7.117.112.512.72.9
logpCO2−4.69−2.54−3.44−3.150.74
HCO3 (mg L−1)10724915913854
PO43−-P (mg L−1)0.010.320.100.050.11
TN (mg L−1)0.2810.154.574.453.54
NH4+ (mg L−1)0.012.750.420.090.84
NO2 (mg L−1)0.050.170.080.070.04
NO3 (mg L−1)0.638.415.311.114.0
TIC (mg L−1)18.4429.3623.7623.064.62
DIC (mg L−1)14.5027.6821.6120.194.33
TOC (mg L−1)7.2024.6616.6116.256.68
DOC (mg L−1)6.1320.2214.0813.825.57
Ca2+ (mg L−1)20.066.138.231.718.3
Mg2+ (mg L−1)15.220.017.416.71.7
Na+ (mg L−1)6.917.412.213.33.7
K+ (mg L−1)0.91.61.11.10.2
Cl (mg L−1)11.824.717.517.34.2
SO42− (mg L−1)16.033.124.324.85.6
SiO2 (mg L−1)7.226.815.613.06.2
ZSD (m)0.1250.50.280.250.111
SICalcite−0.21.10.50.50.4
Table 2. Summary statistics for the first two axes of PCA performed on the environmental variables during the investigated period.
Table 2. Summary statistics for the first two axes of PCA performed on the environmental variables during the investigated period.
PCA AxisPC1PC2
Standard deviation3.5312.145
Proportion of variance (%)51.919.2
Cumulative proportion (%)51.971.1
Eigenvalues
T−0.246−0.094
EC0.268−0.012
pH−0.2540.173
DO0.0300.407
logpCO20.256−0.184
HCO30.237−0.045
PO43−-P−0.008−0.126
TN0.265−0.021
NH4+−0.089−0.355
NO2−0.154−0.294
NO30.2660.059
TIC0.271−0.067
DIC0.193−0.054
TOC−0.260−0.028
DOC−0.249−0.106
Ca2+0.269−0.0003
Mg2+−0.205−0.103
Na+−0.0060.377
K+−0.182−0.263
Cl0.0410.368
SO42−−0.1560.285
SiO2−0.2080.200
ZSD (Secchi)−0.079−0.103
SICalcite−0.2050.127
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kulaš, A.; Marković, T.; Žutinić, P.; Kajan, K.; Karlović, I.; Orlić, S.; Keskin, E.; Filipović, V.; Gligora Udovič, M. Succession of Microbial Community in a Small Water Body within the Alluvial Aquifer of a Large River. Water 2021, 13, 115. https://doi.org/10.3390/w13020115

AMA Style

Kulaš A, Marković T, Žutinić P, Kajan K, Karlović I, Orlić S, Keskin E, Filipović V, Gligora Udovič M. Succession of Microbial Community in a Small Water Body within the Alluvial Aquifer of a Large River. Water. 2021; 13(2):115. https://doi.org/10.3390/w13020115

Chicago/Turabian Style

Kulaš, Antonija, Tamara Marković, Petar Žutinić, Katarina Kajan, Igor Karlović, Sandi Orlić, Emre Keskin, Vilim Filipović, and Marija Gligora Udovič. 2021. "Succession of Microbial Community in a Small Water Body within the Alluvial Aquifer of a Large River" Water 13, no. 2: 115. https://doi.org/10.3390/w13020115

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