Next Article in Journal
Modal Verification and Strength Analysis of Bladed Rotors of Turbine in Rated Working Conditions
Next Article in Special Issue
The Transformation by Catalysis of Prebiotic Chemical Systems to Useful Biochemicals: A Perspective Based on IR Spectroscopy of the Primary Chemicals: Solid-Phase and Water-Soluble Catalysts
Previous Article in Journal
Ecological Response in the Integrated Process of Biostimulation and Bioaugmentation of Diesel-Contaminated Soil
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Water and Sediment Bacterial Communities in a Small Mediterranean, Oxygen-Stratified, Saline Lake (Lake Alboraj, SE Spain)

by
Yolanda Espín
*,
Alfonso Menchén
,
José Luis Moreno
,
David Sanz
,
Manuel Álvarez-Ortí
,
José Antonio Fernández
and
Juan José Gómez-Alday
Group of Hydrogeology, Biotechnology and Natural Resources Laboratory, Institute for Regional Development, University of Castilla-La Mancha, 02071 Albacete, Spain
*
Author to whom correspondence should be addressed.
Appl. Sci. 2021, 11(14), 6309; https://doi.org/10.3390/app11146309
Submission received: 5 May 2021 / Revised: 8 June 2021 / Accepted: 2 July 2021 / Published: 8 July 2021

Abstract

:
Lake Alboraj, located in southeast Spain, was declared natural Microreserve and included into European Natura-2000 Network due to its contribution to environmental heritage. Unfortunately, the ecological status of the lake has changed dramatically, mainly due to the lowering of water table caused by groundwater abstractions for irrigation. It is a permanent small karstic lake whose surface has reduced in the last decades to nearly the third part of its historical water level. The water column shows a marked seasonal oxycline, that splits an aerobic upper layer (epilimnion) from an anaerobic layer below (hypolimnion). Sequencing 16S rRNA gene amplicons and applying chemical tools at epilimnion, hypolimnion and sediment, showed a clear gradient in the bacterial community structure, which support the co-existence of assimilatory and dissimilatory microbial mediated reactions. Results allows to infer that microbial stratification could provide various physical and chemical environments at different depths in the water column related to biogeochemical reactions providing N-S-C- recycling processes.

1. Introduction

In recent times, there is a growing interest in the study of the role of small water bodies in the ecology, biodiversity and provision of ecosystem services at a basin-scale [1]. These small sized lakes (less than 0.1 km2 area) are attracting the attention of current worldwide research due to the active global role on environment dynamics that they play [2]. At the same time, they are highly vulnerable to different temporal and spatial-scale anthropic pressures, as well as to climate change. As a result, small lakes suffer eutrophication processes, water acidification and changes in the water balances, which alter their natural behaviour. Such modifications of the natural behaviour are related to changes in nutrient’s dynamics, with the consequent influence in the amount of greenhouse gas-emissions and in the response of the system to water pollution.
Lake Alboraj is included in a Special Area of Conservation (SAC) of the European Natura 2000 Network (code ES4210011), due to ecologic reasons, since the area surrounding the lake maintains the presence of endemic halophytic flora growing on gypsum soils (the only known population of the endangered Cistaceae species Helianthemum polygonoides). The natural and cultural importance of these ecosystems makes their conservation a matter of environmental awareness by authorities, water users and environmental organisations. In this sense, a responsible use of natural resources is mandatory to guarantee sustainable conditions, which may be easily altered by anthropic pressures nowadays. This also generates a growing concern in the society because Alboraj “Microreserve” contributes to an important biodiversity heritage at local and national level [3]. Nevertheless, from the 80’s of the past century to the present, the ecological status of the lake has changed dramatically mainly due to changes in land uses produced in the hydrogeological boundary of the “Tobarra-Tedera-Pinilla” groundwater body, which formerly contributed to the formation and maintenance of the lake. The most noticeable modifications include the changes from dry-extensive to irrigated-intensive croplands facilitated by groundwater pumping, which has leaded to a lowering of the lake water table.
Even though some limnological approaches have been addressed on the Lake Alboraj [4,5,6,7,8], studies on stratification dynamics of the water column have recently started. Since 2017, a marked oxycline in the eutrophic lake has been observed [9,10], which can constitute a biogeochemical interface by providing aerobic and anaerobic conditions. The dissolved oxygen stratification patterns may originate different conditions in the environment, at different depths in the water column, that lead to different rates in nutrient recycling. Specifically, the turnover of nitrogen in the water column is highly relevant for assessing eutrophication processes in surface water bodies. In this line, the lake became eutrophic from an initial oligotrophic state due to the overgrowth of benthic cyanobacterial mats, which promoted a progressive burial of its submerged macrophyte meadows [10]. The multiple spatial variations in oxygen, N-species concentration, as well as land occupation, may act akin to dependent factors in the composition of bacterial communities and the processes they carry out [11]. Consequently, it affects the correct functioning of the aquatic ecosystem since it may be significantly altered by the activity of microorganisms. Despite bacteria can play an important role in the water column chemostratification, due to their close involvement in N-S-C speciation via redox interfaces, the study of bacterial populations through specific DNA applications in small lakes remain widely unaddressed [12]. Understanding environmental processes is the first step in forecasting lake evolution and, in the same way, a comprehension of the dynamics of stratification is required to prognosticate the past and the future of lakes [13].
The study of environmental DNA by culture-independent tools in Lake Alboraj represents a starting point for studying the microbial structure involved in the N-S-C cycles. In this sense, both the eutrophic character of the small lake and the persistence of an oxycline offer an opportunity for studying the bacterial structure changes across the water column in order to gain knowledge about the microorganisms involved in eutrophication processes as well as in greenhouse gas emissions (i.e., CO2, N2O, CH4). Thus, in this study, novel techniques of next generation sequencing (NGS) are applied to obtain information about the differences existing in the bacterial communities present in the sediment and at different depths of the water column, with different chemical characteristics due to the stratification of the lake. The results obtained will serve as a first step in the study of the biogeological processes that may be happening in these saline environments.

2. Materials and Methods

2.1. Study Area

Lake Alboraj is located in SE Spain, a karstic-tectonic sinkhole formed about 700,000 years ago [14], which sustains a permanent waterbody. The region is influenced by a semiarid meso-Mediterranean climate, with an historic average annual precipitation below 380 mm, with a fresh autumn (average autumn temperature of 6.7 °C) and a hot summer (average summer temperature of 24.3 °C). As freshwater habitat, this lake has been classified as type 3190 “Lakes of gypsum karst” within Natural Habitat Types of Community Interest [4,15]. According to the lithology characteristics, the lake is hyposaline, calcium-sulphated, with a historic conductivity ranging from 2240–2496 mg/L (Regional Decree 182/2000; Declaration of the Micro-reserve Lake Alboraj and Natural Resources Management Plan of Lake Alboraj). The small endorheic lake basin (120 ha) is in an agricultural landscape, where eutrophication, salinisation and lowering of the water level are the main anthropic impacts. Groundwater and surface water withdrawals have caused the shrinkage of the lake from a flooded area of 16,000 m2, in the early 80’s of the past century, to 5500 m2, in 2017 [16]. The lake shrinking in turn has caused a salinity increase up to 5760–7680 mg/L [9]. Nowadays, the lake water table reaches 100 m in diameter, 5000 m2 of surface and settle three deep dolines (5 m depth in its deepest doline) with shallow interdoline areas less than 2 m deep [10]. This geomorphological feature promotes the formation of a marked seasonal oxygen stratification pattern of the water column.
The water level fluctuates and very much depends on the precipitation regime, although groundwater inputs from a local shallow Pliocene aquifer formed in the surroundings prevent drying. From a bathymetry model, the maximum volume of stored water has been estimated in 0.018 hm3, corresponding to a laminar water area of 9500 m2 and a maximum depth of 6.5 m (unpublished data). No water withdrawal or surface sewage discharges occur in the lake.
Typically, when the oxygen stratification occurs, the lake shows three different layers: the epilimnion, which comprises the upper aerobic layer; the metalimnion, the next layer in the depth profile that comprises from the depth at which oxygen starts to decrease to the depth at which anoxia is reached, i.e., the oxycline; and the hypolimnion, extends from the end of metalimnion to the lake bottom, i.e., the anaerobic layer. The oxygen depletion coupled with a strong gradient in physico-chemical parameters in the metalimnion.

2.2. Water and Sediment Samples for Chemistry and Bacterial Analysis

Samples for chemistry and bacterial analysis were collected on March 13, 2019, in the center of deepest doline from a boat (38°34′56.60148″ N and −1°39′12.76319″ W).
One liter of water for bacterial analysis and two liters of water for chemical parameters were collected with a Rüttner bottle at 1 m (W1; aerobic epilimnion) and 4 m (W4; anaerobic hypolimnion) deep. Glass bottles were sterilised in autoclave 120 °C for 15 min and were pre-rinsed several times with water sample before collection. Afterwards, bottles were sealed with a screw cap, avoiding headspace, kept in a cooler box in darkness and transported to the laboratory. At the laboratory, water samples were filtered immediately thought nylon membranes with 0.45 µm pore diameter (Merck KGaA, Darmstadt, Germany) and stored in a cold room (4–6 °C) until further analysis in the next 48 h. Nylon membranes were frozen and kept at −20 °C to be used for further DNA analysis.
A sediment sample was collected at the bottom of deepest point of the lake (4.7 m deep at the sampling date). Sampling was performed by collecting the first 5 cm of the lake bottom sediments with a sediment corer. The sample, collected aseptically in a sterile 50 mL polypropylene tube, was sealed with a screw cap, kept in a cooler box in darkness and transported to the laboratory, for a time not exceeding 5 h.

2.3. Physico-Chemical and Pigment Profiling

Depth profiles of dissolved oxygen (DO), redox condition (Eh), pH, total dissolved solids (TDS) and temperature (T) were performed at 0.5-metre depth intervals from a boat on the same date as water and sediment sampling using a HQ40d portable Multi-Parameter Meter (Hach Company, Loveland, CO, USA). The accuracy of physico-chemical determinations was ±0.01 mg/L for DO, ±0.1 mV for Eh, ±0.002 for pH, ±0.5% for TDS and ±0.3 °C for T.
As a measure of phytoplankton biomass that could help in the interpretation of DNA analysis, pigment concentration profiling was also determined. The samples were taken with a Rüttner bottle at a 1-metre depth interval and transferred to 1 L opaque bottles and kept cold and dark in a cooler box. Phytoplankton chlorophyll a (algae and cyanobacteria) were determined by filtering 400 mL of lake water. Fiberglass filters (0.7 µm pore, 47 mm diameter) were frozen for 1 week to break algal cells. Pigment extraction was accomplished in a vial by adding 5 mL of 90% acetone and leaving it in the fridge at 4 °C at least 12 h up to a maximum incubation of 24 h. After performing spectrophotometry scanning, chlorophyll a was calculated by the trichromatic method [17].

2.4. Chemical Analysis

Alkalinity (HCO3 and CO32−) was determined by titration in the laboratory with a Methrom 855 robotic titrosampler (Methrom 855, Methrom, Switzerland); 50 mL subsamples was treated with acid titration (0.05 M HCl) with a limit of detection of 1 mg/L and an accuracy of ±0.2%. Dissolved Organic Carbon (DOC) and Total Dissolved Nitrogen (TN) concentration was determined by the high temperature combustion method using a Shimadzu TOC-V Analyzer and Total Nitrogen Measuring Unit. The aqueous concentration of Cl, SO42−, F and Br was determined by ionic chromatography (IC) using a 930 Compact IC Flex Oven/SeS/PP/Deg system coupled 944 Professional UV/VIS detector Vario. For Cl, SO42−, F and Br determinations, the limit of detection was 1 mg/L. The concentration of NO3 was determined by UV-vis spectrophotometry following the Griess reaction assay by cadmium reduction to nitrite (NO2) [18]. The detection limit of NO3 measurements was 0.3 mg/L. NH4+ concentration was determined by indophenol blue method [19], with a detection limit of 0.02 mg/L. Accuracy for NO3 and NH4+ determinations was ±1.0%. Sample treatments and chemical analyses were performed at the Biotechnology and Natural Resources laboratory (Institute for Regional Development, University of Castilla-La Mancha).

2.5. DNA Extraction and MiSeq Illumina Sequencing

Environmental DNA was extracted immediately after sampling to minimise significant DNA degradation. Prior to DNA extraction, water and sediment samples were aseptically homogenised by vortexing to prevent decantation or non-representative sampling. After homogenisation, water samples (W1 and W4) were filtered throughout nylon membranes with 0.45 µm pore diameter (Merck KGaA, Darmstadt, Germany), at a rate of 1 nylon membrane per 500 mL, approximately. For the DNA extraction from sediments, 0.5 g of sediment sample were used, while for the water samples, nylon membranes obtained after filtering were used. Environmental DNA extraction was performed using the NucleoSpin® Soil DNA extraction kit (Macherey-Nagel, Düren, Germany) according to the manufacturer’s instructions with slight adaptation for water samples. Nylon membranes were sterilely cut into small fragments by a scalpel, immersed into lysis buffer and vortexed. The resultant supernatant was taken to further proceed to DNA extraction following manufacturer’s instructions. Yield, quality and purity of extracted DNA was quantified by 1% agarose gel and spectrophotometer, using a multimode lector Cytation 5 with the specific software Gen5 (Biotek, VT, USA). DNA yield and quantitation were measured by absorbance at wavelengths 260, 280 and 320 nm. Purity ratio A260/280 was around 1.8 in water and sediment DNA. All extracted DNA samples were stored at −20 °C until further processing.
After DNA quality and quantity controls, amplicon sequencing (V3–V4 region) of 16S rRNA gene [20] was performed on DNA extracts on a MiSeq using v3 chemistry by Stabvida (Caparica, Portugal, Portugal). 16S amplicon PCR primers used in the library preparation was 5′-CCTACGGGNGGCWGCAG-3′ with the Illumina adapter 5′-TCGTCGGCAGCGTCAGATGTGTATAAGAGACAG-3′ for forward sequencing and 5′-GACTACHVGGGTATCTAATCC-3′ with the Illumina adapter 5′-GTCTCGTGGGCTCGGAGATGTGTATAAGAGACAG-3′ for reverse sequencing. Final taxonomic assignment was obtained after amplicon data processing and sequence alignment using QIIME 2™ v2018.6.0 [21] and the version 132 of SILVA database [22]. Generated sequences from epilimnion, hypolimnion and sediment were deposited in the GenBank database under the BioProject PRJNA734791 and BioSample accessions number SAMN19539535–SAMN19539537.

3. Results

3.1. Chemical and Ecological Changes in the Water Column

The lake showed a well-defined stratification pattern of the water column (Figure 1). DO profile defined three layers: epilimnion (0–2.5 m), oxycline (2.5–3.5 m) and hypolimnion (3.5–4.7 m). DO, T, TDS, Eh, pH and Chl a parameters were measured every 0.5 m up to 4.7 m deep, or 4.5 m in case of Chl a, while all other parameters were measured at epilimnion and hypolimnion layers. In epilimnion, DO concentrations ranged from 13.6 to 15.2 mg/L and temperature decrease from 12.8 to 11.4 °C. Eh, pH and Chl a slightly varied, ranging from −80 to −83.1 mV, 8.31 to 8.38 and 57.4 to 67.4 mg/L, respectively, while TDS gradually increase with depth from 7776 to 7814 mg/L. The metalimnion was defined by a sharp change in DO (oxycline), reaching values of 0.06 mg/L at 3.5 m, in the top of the hypolimnion. T and pH both decrease from 10.2 to 8.7 °C and 8.21 to 7.72, respectively, while Eh started increasing up to −46.2 mV and Chl a reaches it maximum value of 127.3 mg/L. In the hypolimnion layer, DO remained close to 0 values, as well as all other parameters hardly changed their values. At the lake bottom was obtained 8.5 °C for T, 7.59 for pH, −39.9 mV for Eh and 7930 mg/L for TDS.
The visual disappearance of the “Secchi disc” occurred at 0.9 m deep, denoting a turbid state of the lake. Regarding N-species, concentrations of NO3, TIN and TON denoted a marked decrease from the epilimnion to the hypolimnion, while NH4+ ranged up in the deeper layer. SO42−, Cl, F and Br also showed higher values at the epilimnion than hypolimnion (Table 1).

3.2. Comparative Analysis of the Microbial Structure

Water and sediment samples for microbial characterisation were taken in early spring coinciding with a period of a well-defined oxygen stratification pattern. The amplicon-based next generation sequencing analysis of the 16S rRNA V3-V4 region, carried out in water and sediment samples, provides a comprehensive screening of bacterial population. After denoising, from a total of 379,326 high-quality sequences representing the bacterial community, an average of 126,442 ± 9509 sequences was obtained per sample. Afterward, with a clustering threshold of 97% similarity level, a total of 682 non-singletons and unique features (OTUs) were identified with an average of 458 ± 132 OTUs per sample (Table 2). Regarding archaeal communities, the dataset reveals poorly enriched and characterised lineages. The sequences clustered in 8 OTU, at 97% similarity level, with very few reads per OTU. A minimum significative number of sequences must be clustered to be considered a valid OTU for classification purposes. For this reason, only the Bacteria domain was considered for community analysis.
The diversity of prokaryotic population was compared between samples by rarefaction curve analysis of the estimated OTUs. Rarefaction curves showed a plateau stage in all samples, achieving a good coverage of the sequencing analysis of the entire community (Supplemental Figure S1). OTU-based alpha diversity measurement, as Shannon diversity index, decreased in depth at the water column, reaching the highest value in the epilimnion (W1) and the lowest value in the hypolimnion (W4). Prokaryotic diversity in sediments was similar to epilimnion diversity (Table 2).
Microbial communities of the studied zones varied greatly in taxonomic composition (Figure 2 and Figure 3). The microbial structure associated showed different distribution based on the sample nature, even within depth-related water location. Proteobacteria phylum was widely present in all samples but differed in class composition either to water or sediment sample. Water samples were also represented by Cyanobacteria and Bacteroidetes phyla, whereas Actinobacteria and Firmicutes were characteristic from sediment phyla structure. The Proteobacterial Alphaproteobacteria and Gammaproteobacteria class, represented by Rhodobacteraceae and Burkholderiaceae, respectively, contributed mostly to the spatial distribution of Lake Alboraj W1 sample (upper water sample, 1 m). OTUs clustering with members of Hydrogenophilaceae and Saccharospirillaceae families, within Gammaproteobacteria class, led the spatial distribution of W4 (deepest water sample, 4 m). Oxyphotobacteria members, within Cyanobacteria, also comprise a representative relative abundance in both water samples depth. Finally, sequences clustering with Pseudomonadaceae, Burkholderiaceae (both Gammaproteobacteria), Propionibacteriaceae (Actinobacteria) families and the Bacilli class contributed mostly to the spatial distribution of Lake Alboraj sediment (S) sample.

3.2.1. Water Associated Bacterial Communities

Bacterial community structure in the water column was more diverse than in sediments; results also showed that there were not large differences between epilimnion and hypolimnion bacterial richness (Figure 3). The sequencing of W1 clustered into 22 phyla, 36 classes, 158 families and 155 genera, while W4 clustered a total of 23 phyla, 39 classes, 171 families and 160 genera. In both cases, more than one third of the genera were unknown or uncultured. Water samples shared >85% of the main phyla, comprising Proteobacteria (W1, 30%; W4, 51%), Cyanobacteria (W1, 27%; W4, 21%) and Bacteroidetes (W1, 27%; W4, 14%). In addition, with the presence of the phylum Verrucomicrobia, the spatial distribution of the bacterial population covered up to 96% in W1. On the other hand, W4 reached a 97% of total spatial distribution with Patescibacteria, Actinobacteria and Verrucomicrobia. Despite of the similarity in phyla distribution, the water strata become divergent at the family level.
Oxyphotobacteria, the only family within Cyanobacteria, comprised 27% of relative abundance in W1, decreasing to 21% in W4. The second most abundant family was Rubritaleaceae, within Verrucomicrobia, contributing 8% in W1. An environment of low oxygen availability such as W4 habitat drastically limits family development because of its mandatory aerobic chemoheterotrophic metabolism [23]. The Bacteroidia relatives clustered mainly in Saprospiraceae, Cyclobacteriaceae, Cryomorphaceae, Crocinitomicaceae and Flavobacteriaceae families, accounting a total of 23% over the bacterial population in W1 sample. It is a clade strongly dependent on the availability of organic matter, with complex organic carbon sources transformation and decomposition (i.e., Flavobacteriaceae), saprophytism of microorganisms—algae or bacteria—(i.e., Saprospiraceae) [24]. Moreover, the clade involves other families without any known function due to the difficulties in isolating their members (i.e., Cryomorphaceae). Members of the family Sphingomonadaceae also were present, although only represented a 2% of the total spatial bacterial distribution in W1 and was described as xenobiotic compounds transformation [25,26].
As regards W4 sample, Cyanobacteria continue to be among the most abundant taxa, representing 20% of total identified sequences. Contrary to W1 sample, Gammaproteobacteria members as well as Hydrogenophilaceae and Saccharospirillaceae families increased their representation accounting for 30%, 19% and 14%, respectively. Little is known about these families and their metabolic functions. Hydrogenophilaceae had been poorly described as nitrogen fixers by Calvin-Benson cycle and reduced-sulphur compounds oxidisers [27]. Unknown family from the recent proposed lineage Gracilibacteria (currently, included in Patescibacteria phylum) was identified in a 5% of relative abundance in W4. Alphaproteobacteria clade was comprised mainly by Terasakiellaceae, Rhodobacteraceae, Thalassobaculaceae and Caulobacteraceae. Members of the family Terasakiellaceae and Caulobacteraceae were documented to be nitrogen fixers [28] and denitrifiers through nitrate reducing pathway [29], respectively, and may be important for nitrogen natural regulation. Furthermore, Rhodobacteraceae and Desulfobulbaceae families had been reported to play a key role in sulphur cycle. Rhodobacteraceae acts as sulphur, sulphide, thiosulphate or sulphate oxidiser to obtain carbon sources [30], while Desulfobulbaceae is a well-known sulphate-reducing bacteria family [31]. Finally, some members of the family Microbacteriaceae (Actinobacteria) appeared in W4. This family has been found to be associated to saprophytic processes by degradation of organic matter. They are also associated to phosphorus extraction under aerobic and anaerobic conditions or recalcitrant xenobiotic degradation [32,33].

3.2.2. Sediment Associated Microbial Communities

Lake Alboraj sediment bacterial communities presented the lowest diversity regarding bacterial taxa. The sequencing clustered a total of 28 phyla, 58 classes, 94 orders, 117 families and 109 genera. Most of the genera were unknown or uncultured. At the phylum level, Proteobacteria (55%), Actinobacteria (19%) and Firmicutes (10%) comprised the overall structure, accounting for 84% of relative abundance (Figure 2). Within Proteobacteria classes, Gammaproteobacteria (48%) was the dominant, which was related to sulphur cycle processes. Moreover, Alphaproteobacteria and Deltaproteobacteria (4% and 3%, respectively), which are related to nitrate reduction and sulphur cycle, completed the main phylum. In addition to the Proteobacteria members, representative of the Actinobacteria phylum, were observed. Actinobacteria, the second phylum, was almost dominated by members of Actinobacteria class (17%). Firmicutes, the third most abundant phyla, was formed by Bacilli members (7%).
The most abundant family within the bacterial community (Figure 3) was Pseudomonadaceae, a Gammaproteobacteria with 26% of the total population. Pseudomonadaceae is closely reported to nitrogen cycling, including some species which have been reported to be able to perform aerobic or anaerobic denitrification processes commonly by nitrate-reduction or nitroaromatic compound degradation (i.e., fertilisers, pesticides) [34,35]. In addition, Burkholderiaceae family, with a 10% of relative abundance, has been reported to manage widespread adaptation to different ecological niches and can also perform very diverse metabolic functions (i.e., nitrogen fixation, aromatic compound catabolism) [36]. Other Gammaproteobacteria member with a 9% of relative abundance was Moraxellaceae family, which has saprophytic capacities. The members of this family have been described as active agents in decomposition and subsequent mineralisation of organic matter present in the environment [37]. In addition, although it is not a family related to nitrogen-cycle, some of its genera have been reported to be related to nitrate-reducing processes [38].
The Actinobacteria relatives present in Lake Alboraj sediment clustered mainly in the recent described Propionibacteriaceae family, accounting 16% of relative abundance of the total. Remarkably, its metabolic properties are monospecific, however, its metabolic functionality may change whether new species were included [39]. On the other hand, a well-known Alphaproteobacteria member was Rhodobacteraceae. This family is characterised by its metabolic multifunctionality on sulphur and carbon cycles, as well as the symbiotic association with other microorganisms [30], which makes it can be found widespread distributed throughout the lake. The noteworthy function of the family is the oxidation of sulphur, its simple compounds (sulphide, thiosulphate) or sulphate (in some cases) to obtain a carbon source [30]. Elemental sulphur is usually the end product of its metabolism. Another significant finding is the abundance of Flavobacteriaceae members, which is mainly particle-associated and strongly dependent on the availability of organic matter [24,40].

4. Discussion

To our knowledge, this study provides the first effort to investigate the bacterial community structure along depth at Lake Alboraj, an ecosystem which is subjected to self-regulation due to the forgotten human care and nature preserving. Results showed that the diversity of bacterial communities, according to the phylogenetic analysis, gives rise to different lineages along spatial location and habitat. Phylogenetic diversity indicated a longer evolutionary history in water samples, due to a phylogenetic distancing between the microorganisms that coexist in the environment [41,42,43]. Moreover, a non-metric multidimensional scaling (NMDS) analysis revealed clear differences among the bacterial community structure between water samples and sediment (Supplemental Figure S2). So, both phylogenetic diversity measurement and NMDS analysis suggest the coexistence of different bacteria community structures into the water column and water-sediment interface. Under an environment with conditions of low-oxygen and anoxia, microorganisms are expected to adapt their requirements towards metabolic strategies diversification to fill the expected oxygen-limiting niche. For that reason, the bacterial community changes in their distribution between the epilimnion and hypolimnion layers, and the sediment, coinciding with noticeable oxygen depletion in the water column. DO concentrations shift to lower concentrations in a few decimeters at 2.5 m water depth, describing a sharp oxycline (2.5–3.5 m) which was not linked to the thermocline (Figure 1). This result is not in line with the accepted consensus which states that in oxygen stratified lakes is a common process of the development of a thermocline [44]. The small dimension of the lake, a water column of about 4.7 m depth and an extension near to 5500 m2, may not promote the development of a temperature gradient during winter as well as an efficient oxygen exchange at the water-atmosphere interface. In fact, the existence of a complete anaerobic water column from late spring to late summer suggests that oxygen exchange with atmosphere can be limited to cold periods; during warm periods, a decrease in oxygen solubility matches with the augmentation of water temperature.
Regarding to processes involved in the N-cycle, a greater abundance of bacterial families was found in the sediment and hypolimnion layer (W4), compared to epilimnion (W1). Despite of the similarity in the uses of nitrogen compounds, the bacterial communities found in the sediment and W4 samples are dissimilar. Noticeably, the greater abundance of denitrifiers, such as members belonging to Pseudomonadaceae, Burkholderiaceae and Halanaerobiaceae, was detected in the sediment. In W4 sample, Caulobacteraceae and some members clustering in Rhodobacteraceae and Hydrogenophillaceae, were related to processes contributing to nitrogen turnover. Even though the presence of microorganisms identified as facultative aerobic denitrifiers were found in sample W1 (i.e., Burkholderiacceae, Rhodobacteriaceae), they have not such relative frequency such as in the deeper water layer and bottom lake sediment samples. This agrees with the fact that conventional biological nitrogen removal is achieved by two well-defined sequential stages, nitrification under aerobic conditions followed by denitrification under anoxic conditions. Therefore, in the oxic layer of the water column the denitrification is not such manageable by biological processes mediated by bacteria, which are commonly adapted to anaerobic conditions. Across the bacterial community clustered in W1, only Burkholderiaceae has been identified within the denitrification processes. Even though little is known about the metabolic capacities, this finding suggests that some bacteria are capable of change their metabolism in agreement with the environmental conditions and N-species availability [11]. The higher presence of denitrifier bacteria in the deeper zones of the water column, matches with a decrease in the concentration of nitrate (NO3), from 10.33 mg/L, in the epilimnion, to 4.70–0.60 mg/L, in the hypolimnion. The higher number of denitrification related families in sediments could be explained by low oxygen habitat and the availability of NO3 and organic energy sources, since NH4+ is rarely limiting in sediments [45,46]. In W4, members belonging to Saprospiraceae, Flavobacteriaceae, Crocinitomicaceae and Microbacteriaceae families, which can perform saprophytic processes linked to decomposition of organic matter and microbial biomass [24], have also been identified.
The production and consumption of NH4+ can be driven under different microbially mediated processes linked to gradients of oxygen [47,48,49]. The development of an anaerobic hypolimnion is typical in eutrophic lakes and the processes can involve the increase of nutrients, such as NH4+. At the water-sediment interface, processes such as ammonification can release NH4+ during organic matter decay. In this sense, it has been proposed that in the hypolimnion NH4+ was the major N species identified, and sediment release can contribute nearly 85% hypolimnetic NH4+ [47]. A decrease in NH4+ concentration is clearly detected from the hypolimnion (2.18 mg/L) to the epilimnion (0.09 mg/L). The nutrient availability found at hypolimnion may be stimulated by microbially-mediated processes at the water-sediment interface.
Dissimilatory nitrate reduction to ammonium (DNRA), may produce NH4+ under strictly anaerobic conditions [45]. A wide distribution group of sediment bacteria is capable to perform DNRA [50], which acts in Alboraj sediment as facultative anaerobes, Enterobacteriaceae, to aerobes, Pseudomonadaceae. Moreover, bacteria belonging to Desulfuromonadaceae family, which have been found to be associated to DNRA were also found in W4 sample. Bacterial structure results could support these findings where DNRA processes could take place in sediment by related members from Gammaproteobacteria and Deltaproteobacteria. Under anaerobic conditions, the oxidation of NH4+ can be driven under the presence of NO2 (anaerobic ammonium oxidation, anammox). According to present knowledge, anammox is carried out by a few anaerobic, autotrophic bacteria of a monophyletic clade in Planctomycetes [51]. Sediment gathered the conditions of anaerobic medium and substrate supplier necessary to keep families related to carried out anammox processes as Pirellulaceae, also poorly identified in W4 sample, within the phylum Planctomycetes. However, due to the limited information about factors controlling the processes in natural systems, knowledge about anammox in nature must continue as an ongoing effort to understand the distribution of the process that occur in the environment.
Noticeably, in the epilimnion, the nitrogen fixers are abundant. Oxygenic photoautotrophic cyanobacteria (Oxyphotobacterias) together with populations of Terasakiellaceae might be responsible for the primary productivity in the epilimnion. The proliferation of Cyanobacteria is strongly linked to biosphere’s environmental perturbations, including nutrient enrichment [52] and climatic changes (i.e., hydrologic changes, drought and precipitation events hardening) [53]. Their role as atmospheric nitrogen fixers, as well as their capacity of symbiotic cooperation with other microorganisms are well known. Nevertheless, cyanobacterial bloom results in a massive decomposition of dead cells which is associated to cyanotoxins release to the environment [53]. Whenever toxicity occurs, it can cause harmful effects to microorganisms, leading to an unbalanced development of those communities. Moreover, in epilimnion layer, NH4+ can be assimilated by biomass as well as transformed into NO3 by nitrifiers under aerobic conditions (aerobic ammonium oxidation, nitrification). Therefore, nitrification may explain the appearance of NO3 in the epilimnion. Nonetheless, in our case study, the importance of external inputs (i.e., discharge of polluted groundwater, surface runoff) in nutrient availability cannot either be confirmed or discarded.
Biogeochemical S-cycle can be closely linked to microbial redox processes in aquatic systems [54,55]. Sulphate is an important end-product during microbially-mediated sulphide oxidation processes but can also be derived from the dissolution of sulphate-bearing minerals, as gypsum. Gypsum has been identified in the Triassic evaporitic rocks outcropping in the study area as well as in detrital grains in the Holocene sediments in which the lake is located [56]. The difference in SO42− concentrations between the hypolimnion (6310 mg/L) and epilimnion (5426 mg/L) is important enough to dismiss the detection limit of the method. Nonetheless, a variety of sulphur related bacteria were identified in W1 and W4, which can perform coupled sulphate reduction–sulphide oxidation processes. At this interface, SO42− can be used during suphate-reducing bacteria processes to produce H2S, which is re-oxidised in the epilimnion to SO42− in the presence of oxygen. In this sense, S–cycle bacteria were better represented in anoxic hypolimnion layer, W4, than in the oxic epilimnion layer, W1. In W4 sample bacteria belonging to Desulfobulbaceae, Rhodobacteraceae, Crocinitomicaceae families, which include various sulphate–reducing bacteria [30,57] were found. High activity of sulphate-reducing bacteria is accompanied by production of H2S, which acts as an electron donor for the chemolithotropic sulphur–oxidising bacteria. Microorganisms within the families Rhodobacteraceae and Cyclobacteriaceae found in the epilimnion are known sulphate–oxidising bacteria [58,59], which may participate in the activity of re-oxidising the bottom–produced H2S to SO42− in the upper oxic layer. Interestingly, both, the family Methylomonadaceae, identified in W4, and Methylophagaceae, identified in the sediment sample, are associated with H2S oxidation processes and methane (CH4) production. The availability of dissolved organic carbon is noticeable in the water column, but water samples do not show an important difference in DOC concentrations at W4 (14.9 mg/L) and W1 (12.9 mg/L). The consumption of organic matter by heterotrophs may lead to the production of CO2 (g) which, in turn, can be reduced to CH4 (g), under anaerobic conditions. The intensity of DO consuming processes by heterotrophic microorganisms may lead to the development of an oxygen deficient zone of different thickness in the lake. Depending on the rate of oxygen consumption processes and the renewal of oxygen at the water-atmosphere interface, a well-defined oxycline can be developed, with different thickness, in a poor ventilated water column, in absence of thermal stratification.
The bacterial distribution found matched with changes in the observed dissolved oxygen concentration decrease, from epilimnion to hypolimnion, chlorophyll concentrations and the distribution of some dissolved chemicals. These results support the idea that microbial structure plays an important role in the distribution and fate of chemical species. The bacterial distribution in the bottom lake sediment and the stratified water column can promote the development of different interfaces in which redox conditions lead to the recycling of N, S and C-species.
Despite the importance of inherent interplay between native microorganisms and the environmental conditions in the development of a chemical stratification, there is still a paucity of evidence on what biochemical processes are being conducted by bacteria communities in Lake Alboraj. The interfaces, in turn, may offer the opportunity of learning about the distribution and fate of some pollutants, as well as deepening in the knowledge of the processes conducting to eutrophication of water bodies in which an oxycline is developed independently of the water volume stored and morphology of the lake.

5. Conclusions

The amplicon-based MiSeq Illumina sequencing allows obtaining a complete screening of the bacterial population inhabiting at different depths in the water column as well as in the sediment of Lake Alboraj. The composition of bacterial communities differs from the epilimnion to the hypolimnion, while the sediment holds a bacterial population different from those representing water samples. A greater abundance of bacteria families related to processes involved in the N-cycle was found in the sediment and hypolimnion layer (W4), although with different composition. A greater abundance of denitrifiers such as those belonging to families Caulobacteraceae, Rhodobacteraceae and Hydrogenophillaceae was detected in W4 samples, while in the sediment, members of the families Pseudomonadaceae and Burkholderiaceae and Halanaerobiaceae, related to processes that contribute to nitrogen turnover, were found. The higher presence of denitrifier bacteria in the deeper zones is consistent with the decrease in the concentration of NO3 detected. On the other hand, NH4+ concentration was higher in the hypolimnion. In this sense, a large group of bacteria capable to perform DNRA processes was found in the sediment, from facultative anaerobes such as Enterobacteriaceae, to aerobes such as Pseudomonadaceae or others akin to those from the family Desulfuromonadaceae.
As regards S-cycle, the results obtained showed the presence of a variety of sulphur related bacteria in the water samples, which can perform coupled sulphate reduction–sulphide oxidation processes. The anoxic hypolimnion layer showed a greater abundance of S-cycle related bacteria, including those belonging to the families Desulfobulbaceae, Rhodobacteraceae, Crocinitomicaceae, which can perform sulphate–reducing processes to produce H2S. This H2S may be re-oxidised in the presence of oxygen by known sulphate-oxidising bacteria such as Rhodobacteraceae and Cyclobacteriaceae found in the epilimnion.
This study has shed light on the bacterial community structure and its support to the co-existence of assimilatory and dissimilatory microbial mediated reactions. Differences in physico-chemical conditions in the water column are associated to a bacterial community structure adapted to the available nutrients. These dynamic biological processes might hold a key role in maintaining the ecological equilibrium of the aquatic ecosystem. Bacterial distribution points out that N-S-C biogeochemical cycles can be interconnected throughout interfaces: water-atmosphere, oxycline and water-sediment.
Future work is needed to understand the temporal changes in the structure of the bacterial communities through detailed monitoring that matches the changes in oxygen stratification patterns during the year. Detailed exploration of microbial community structure variation over a non-stratified situation and oxygen dynamic states, in addition to their function adjustment relationship, will be essential to deepen into the knowledge of the ecosystem development. Moreover, an exploration of the remaining microbial processes is still needed to understand the lake functioning and thus the resilience to its ongoing deteriorated ecological status.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/app11146309/s1, Figure S1: Observed OTUs alpha rarefaction curve, Figure S2: Non-metric multidimensional scaling (NMDS) analysis of the weighted UniFrac distance of water and sediment samples from Lake Alboraj.

Author Contributions

Conceptualisation, Y.E., M.Á.-O. and J.J.G.-A.; methodology, Y.E., A.M. and D.S.; formal analysis, Y.E. and A.M.; investigation, M.Á.-O., J.A.F. and J.J.G.-A.; resources, J.L.M.; data curation, Y.E., A.M., J.L.M. and D.S.; writing—original draft preparation, Y.E.; writing—review and editing, J.L.M., D.S., J.A.F., M.Á.-O. and J.J.G.-A.; funding acquisition, Y.E., D.S. and J.J.G.-A. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by a PhD grant (2020-PREDUCLM-14753) from the University of Castilla-La Mancha, a contract (2019-CON-9523) from the Spanish Government-FSE, and the research projects CGL2017-87216-C4-2-R from the National Research Program I+D+i from FEDER/Ministerio de Ciencia, Investigación y Universidades and SBPLY/17/180501/000296L from the Regional Research Program I+D+i of the Junta de Comunidades de Castilla-La Mancha.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kelly-Quinn, M.; Biggs, J.; von Fumetti, S. Preface: The importance of small water bodies. Hydrobiologia 2017, 793, 1–2. [Google Scholar] [CrossRef]
  2. 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]
  3. Cebrián, A. El patrimonio natural de Albacete como recurso turístico. Gran Tour Rev. Investig. Turísticas 2010, 2, 62–82. [Google Scholar]
  4. Camacho, A.; Borja, C.; Valero-Garcés, B.; Sahuquillo, M.; Cirujano, S.; Soria, J.M.; Rico, E.; De La Hera, A.; Santamans, A.C.; García De Domingo, A.; et al. 3190 Lagos y lagunas kársticas sobre yesos. In Bases Ecológicas Preliminares para la Conservación de los Tipos de Hábitat de Interés Comunitario en España; Ministerio de Medio Ambiente, y Medio Rural y Marino: Madrid, Spain, 2009. [Google Scholar]
  5. Boronat, L.; Miracle, M.R.; Armengol, X. Cladoceran assemblages in a mineralization gradient. Hydrobiologia 2001, 442, 75–88. [Google Scholar] [CrossRef]
  6. Millán, A.; Moreno, J.L.; Velasco, J. Estudio faunístico y ecológico de los coleópteros y heterópteros acuáticos de las lagunas y humedales de Albacete (Alboraj, Los Patos, Ojos de Villaverde, Ontalafia y Pétrola). Sabuco 2001, 1, 43–94. [Google Scholar]
  7. Soria, J.; Vicente, E.; Camacho, A.; Miracle, M.; Rico, E.; García-Avilés, J.; Rueda, J.; Armengol, X.; Roca, J. Estudio Limnológico de 28 Humedales de Castilla-La Mancha Como Base para la Elaboración del Plan de Ordenación de Recursos Naturales; Junta de Comunidades de Castilla-La Mancha: Valencia, Spain, 1998. [Google Scholar] [CrossRef]
  8. Cirujano, S. Flora y Vegetación de las Lagunas y Humedales de la Provincia de Albacete; Instituto de Estudios Albacetenses: Albacete, Spain, 1990; p. 52. ISBN 978-84-8713-617-7. [Google Scholar]
  9. Moreno, J.L.; Menchén, A.; Toledo, B.; Sanz, D.; Gómez-Alday, J.J. Formación de oxiclina estacional, anoxia y dinámica de nutrientes en una pequeña laguna kárstica del SE ibérico (Laguna de Alboraj, Albacete). In Proceedings of the Congress of the Iberian Association of Limnology & III Iberoamerican Congress of Limnology, Murcia, Spain, 26–29 October 2020. [Google Scholar]
  10. Moreno, J.L.; Ortega, J.F.; Moreno, M.A.; Ballesteros, R. Using Unmanned aerial vehicles (UAV) for lake management: Ecological status, lake state shifts and biogeochemical processes in a small Mediterranean lake. In Proceedings of the Congress of the Iberian Association of Limnology & III Iberoamerican Congress of Limnology, Murcia, Spain, 26–29 October 2020. [Google Scholar]
  11. Zeglin, L.H. Stream microbial diversity in response to environmental changes: Review and synthesis of existing research. Front. Microbiol. 2015, 6, 454. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Harper, L.R.; Buxton, A.S.; Rees, H.C.; Bruce, K.; Brys, R.; Halfmaerten, D.; Read, D.S.; Watson, H.V.; Sayer, C.D.; Jones, E.P.; et al. Prospects and challenges of environmental DNA (eDNA) monitoring in freshwater ponds. Hydrobiologia 2019, 826, 25–41. [Google Scholar] [CrossRef] [Green Version]
  13. Boehrer, B.; Schultze, M. Stratification of lakes. Rev. Geophys. 2008, 46. [Google Scholar] [CrossRef] [Green Version]
  14. Vegas, J.; Rodríguez-Pascua, M.; Perucha, M.A.; Pérez-López, R.; Israde-Alcántara, I.; Galán, L.; Mediato, J. Upper Holocene seismites in the Alboraj lake (I to IV centuries, Albacete, Spain). GeoTemas 2012, 13, 1543–1546. [Google Scholar]
  15. European Union. Council Directive 92/43/EEC of 21 May 1992 on the conservation of natural habitats and of wild fauna and flora. Off. J. Eur. Union 1992, 206, 7–50. [Google Scholar]
  16. Moreno, J.L.; Monteagudo, L.; Florin, M.; Chicote, A.; Ortega, J.F.; Ballesteros, R.; Moreno, M.A.; De Las Heras, J. Evolución histórica y reciente de la de laguna de Alboraj (Tobarra, Albacete): Bases para su estudio. In Proceedings of the III Jornadas Sobre el Medio Natural Albacetense, Albacete, Spain, 5–8 October 2016; Blanco, D., Fajardo, J., Ferrandis, P., Gómez, J., Picazo, J., Sanz, D., Valdés, A., Verde, A., Eds.; Instituto de Estudios Albacetenses: Albacete, Spain, 2018; pp. 73–81. ISBN 978-84-948930-0-1. [Google Scholar]
  17. Jeffrey, S.W.; Humphrey, G.F. New spectrophotometric equation for determining chlorophylls a, b, c1 and c2 in higher plants, algae and natural phytoplankton. Biochem. Plant Physiol. 1975, 167, 191–194. [Google Scholar] [CrossRef]
  18. Wood, E.D.; Armstrong, F.A.J.; Richards, F.A. Determination of nitrate in sea water by cadmium-copper reduction to nitrite. J. Mar. Biolog. Assoc. U. K. 1967, 47, 23–31. [Google Scholar] [CrossRef] [Green Version]
  19. Koroleff, F. Direct determination of ammonia in natural waters as indophenol blue. Int. Con. Explor. Sea 1969, 9, 152–155. [Google Scholar]
  20. Petrosino, J.F.; Highlander, S.; Luna, R.A.; Gibbs, R.A.; Versalovic, J. Metagenomic Pyrosequencing and Microbial Identification. Clin. Chem. 2009, 55, 856–866. [Google Scholar] [CrossRef] [PubMed]
  21. Bolyen, E.; Rideout, J.R.; Dillon, M.R.; Bokulich, N.A.; Abnet, C.C.; Al-Ghalith, G.A.; Alexander, H.; Alm, E.J.; Arumugam, M.; Asnicar, F.; et al. Reproducible, interactive, scalable and extensible microbiome data science using QIIME 2. Nat. Biotechnol. 2019, 37, 852–857. [Google Scholar] [CrossRef]
  22. Quast, C.; Pruesse, E.; Yilmaz, P.; Gerken, J.; Schweer, T.; Yarza, P.; Peplies, J.; Glöckner, F.O. The SILVA ribosomal RNA gene database project: Improved data processing and web-based tools. Nucleic Acids Res. 2013, 41, D590–D596. [Google Scholar] [CrossRef] [PubMed]
  23. Rosenberg, E. The Family Rubritaleaceae. In The Prokaryotes; Springer: Berlin, Germany, 2014; pp. 861–862. [Google Scholar] [CrossRef]
  24. Zárate, A.; Dorador, C.; Araya, R.; Guajardo, M.; Florez, J.Z.; Icaza, G.; Cornejo, D.; Valdés, J. Connectivity of bacterial assemblages along the Loa River in the Atacama Desert, Chile. PeerJ 2020, 8, e9927. [Google Scholar] [CrossRef]
  25. Oh, S.; Choi, D. Microbial Community Enhances Biodegradation of Bisphenol A Through Selection of Sphingomonadaceae. Microb. Ecol. 2018, 77, 631–639. [Google Scholar] [CrossRef]
  26. Basta, T.; Keck, A.; Klein, J.; Stolz, A. Detection and Characterization of Conjugative Degradative Plasmids in Xenobiotic-Degrading Sphingomonas Strains. J. Bacteriol. 2004, 186, 3862–3872. [Google Scholar] [CrossRef] [Green Version]
  27. Garrity, G.M.; Bell, J.A.; Lilburn, T. Hydrogenophilaceae fam. nov. In Bergey’s Manual of Systematics of Archaea and Bacteria; Bergey’s Manual Trust and John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2015; p. 1. [Google Scholar] [CrossRef]
  28. Weiler, B.A.; Verhoeven, J.T.P.; Dufour, S.C. Bacterial Communities in Tissues and Surficial Mucus of the Cold-Water Coral Paragorgia arborea. Front. Mar. Sci. 2018, 5. [Google Scholar] [CrossRef]
  29. Oh, Y.S.; Roh, D.H. Phenylobacterium muchangponense sp. nov., isolated from beach soil, and emended description of the genus Phenylobacterium. Int. J. Syst. Evol. Microbiol. 2011, 62, 977–983. [Google Scholar] [CrossRef]
  30. Pujalte, M.J.; Lucena, T.; Ruvira, M.A.; Arahal, D.R.; Macián, M.C. The Family Rhodobacteraceae. In The Prokaryotes; Rosenberg, E., DeLong, E.F., Lory, S., Stackebrandt, E., Thompson, F., Eds.; Springer: Berlin, Germany, 2014; pp. 439–512. [Google Scholar] [CrossRef]
  31. Klepac-Ceraj, V.; Bahr, M.; Crump, B.C.; Teske, A.P.; Hobbie, J.E.; Polz, M.F. High overall diversity and dominance of microdiverse relationships in salt marsh sulphate-reducing bacteria. Environ. Microbiol. 2004, 6, 686–698. [Google Scholar] [CrossRef]
  32. Gaytán, I.; Sánchez-Reyes, A.; Burelo, M.; Vargas-Suárez, M.; Liachko, I.; Press, M.; Sullivan, S.; Cruz-Gómez, M.J.; Loza-Tavera, H. Degradación del poliuretano recalcitrante y aditivos xenobióticos por una comunidad microbiana de vertederos seleccionada y su potencial biodegradativo revelado por análisis metagenómico basado en ligadura de proximidad. Front. Microbiol. 2019, 10, 2986. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Evtushenko, L.I.; Takeuchi, M. The Family Microbacteriaceae. In The Prokaryotes; Dworkin, M., Falkow, S., Rosenberg, E., Schleifer, K.H., Stackebrandt, E., Eds.; Springer: New York, NY, USA, 2006. [Google Scholar] [CrossRef]
  34. Ye, J.; Singh, A.; Ward, O.P. Biodegradation of nitroaromatics and other nitrogen-containing xenobiotics. World J. Microbiol. Biotechnol. 2004, 20, 117–135. [Google Scholar] [CrossRef]
  35. Rieger, P.G.; Meier, H.M.; Gerle, M.; Vogt, U.; Groth, T.; Knackmuss, H.J. Xenobiotics in the environment: Present and future strategies to obviate the problem of biological persistence. J. Biotechnol. 2002, 94, 101–123. [Google Scholar] [CrossRef]
  36. Compant, S.; Nowak, J.; Coenye, T.; Clément, C.; Ait Barka, E. Diversity and occurrence of Burkholderia spp. in the natural environment. FEMS Microbiol. Rev. 2008, 32, 607–626. [Google Scholar] [CrossRef] [Green Version]
  37. Shchegolkova, N.M.; Krasnov, G.S.; Belova, A.A.; Dmitriev, A.A.; Kharitonov, S.L.; Klimina, K.M.; Melnikova, N.V.; Kudryavtseva, A.V. Microbial Community Structure of Activated Sludge in Treatment Plants with Different Wastewater Compositions. Front. Microbiol. 2016, 7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Dong, Y.; Lin, H.; Zhang, X. Simultaneous Ammonia Nitrogen and Phosphorus Removal from Micro-Polluted Water by Biological Aerated Filters with Different Media. Water Air Soil Pollut. 2020, 231. [Google Scholar] [CrossRef]
  39. Stackebrandt, E. The Family Propionibacteriaceae: Genera other than Propionibacterium. Prokaryotes 2014, 725–741. [Google Scholar] [CrossRef]
  40. Newton, R.J.; Jones, S.E.; Eiler, A.; McMahon, K.D.; Bertilsson, S. A Guide to the Natural History of Freshwater Lake Bacteria. Microbiol. Mol. Biol. Rev. 2011, 75, 14–49. [Google Scholar] [CrossRef] [Green Version]
  41. Velásquez-Puentes, F.J.; Bacon, C.D. An introduction to phylogenetic community structure: A case study in Bolivian palms. Ecol. Boliv. 2016, 51, 126–140. [Google Scholar]
  42. Peterson, A.T. Conservatism of ecological niches in evolutionary time. Science 1999, 285, 1265–1267. [Google Scholar] [CrossRef] [PubMed]
  43. Faith, D.P. Conservation evaluation and phylogenetic diversity. Biol. Conserv. 1992, 61, 1–10. [Google Scholar] [CrossRef]
  44. Foley, B.; Jones, I.D.; Maberly, S.C.; Rippey, B. Long-term changes in oxygen depletion in a small temperate lake: Effects of climate change and eutrophication. Freshw. Biol. 2012, 57, 278–289. [Google Scholar] [CrossRef]
  45. Meyer, R.L.; Risgaard-Petersen, N.; Allen, D.E. Correlation between anammox activity and microscale distribution of nitrite in a subtropical mangrove sediment. Appl. Environ. Microbiol. 2005, 71, 6142–6149. [Google Scholar] [CrossRef] [Green Version]
  46. Risgaard-Petersen, N.; Nielsen, L.P.; Rysgaard, S.; Dalsgaard, T.; Meyer, R.L. Application of the isotope pairing technique in sediments where anammox and denitrification coexist. Limnol. Oceanogr. Meth. 2003, 1, 63–73. [Google Scholar] [CrossRef]
  47. Su, X.; He, Q.; Mao, Y.; Chen, Y.; Hu, Z. Dissolved oxygen stratification changes nitrogen speciation and transformation in a stratified lake. Environ. Sci. Pollut. Res. 2019, 26, 2898–2907. [Google Scholar] [CrossRef]
  48. Kirf, M.K.; Hans, R.; Holtappels, M.; Fischer, J.P.; Schubert, C.J.; Wehrli, B. Redox gradients at the low oxygen boundary of lakes. Aquat. Sci. 2015, 77, 81–93. [Google Scholar] [CrossRef]
  49. Degermendzhy, A.G.; Zadereev, E.S.; Rogozin, D.Y.; Prokopkin, I.G.; Barkhatov, Y.V.; Tolomeev, A.P.; Khromechek, E.B.; Janse, J.H.; Mooij, W.M.; Gulati, R.D. Vertical stratification of physical, chemical and biological components in two saline lakes Shira and Shunet (South Siberia, Russia). Aquat. Ecol. 2010, 44, 619–632. [Google Scholar] [CrossRef] [Green Version]
  50. Tiedje, J. Ecology of denitrification and dissimilatory nitrate reduction to ammonium. In Biology of Anaerobic Microorganisms; Zehnder, J.M., Ed.; Wiley: New York, NY, USA, 1988; pp. 179–244. [Google Scholar]
  51. Jetten, M.S.; Niftrik, L.V.; Strous, M.; Kartal, B.; Keltjens, J.T.; Op den Camp, H.J. Biochemistry and molecular biology of anammox bacteria. Crit. Rev. Biochem. Mol. Biol. 2009, 44, 65–84. [Google Scholar] [CrossRef] [PubMed]
  52. Wilhelm, S.W.; Farnsley, S.E.; LeCleir, G.R.; Layton, A.C.; Satchwell, M.F.; DeBruyn, J.M.; Boyer, G.L.; Zhu, G.; Paerl, H.W. The relationships between nutrients, cyanobacterial toxins and the microbial community in Taihu (Lake Tai), China. Harmful Algae 2011, 10, 207–215. [Google Scholar] [CrossRef]
  53. Paerl, H.W.; Paul, V.J. Climate change: Links to global expansion of harmful cyanobacteria. Water Res. 2012, 46, 1349–1363. [Google Scholar] [CrossRef] [PubMed]
  54. Jørgensen, B.B. Ecology of the sulphur cycle: Oxidative pathways in the sediment. In The Nitrogen and Sulphur Cycles; Cole, J.A., Ferguson, S.J., Eds.; Cambridge University Press: Cambridge, UK, 1988; pp. 31–63. [Google Scholar]
  55. Holmer, M.; Storkholm, P. Sulphate reduction and sulphur cycling in lake sediments: A review. Freshw. Biol. 2001, 46, 431–451. [Google Scholar] [CrossRef]
  56. Rodríguez-Pascua, M.A.; Bischoff, J.; Garduño-Monroy, V.H.; Pérez-López, R.; Giner-Robles, J.L.; Israde-Alcántara, I.; Calvo, J.P.; Williams, R.W. Estimation of the tectonic slip-rate from Quaternary lacustrine facies within the intraplate Albacete province (SE of Spain). Sediment. Geol. 2009, 222, 89–97. [Google Scholar] [CrossRef] [Green Version]
  57. Tarnovetskii, I.Y.; Merkel, A.Y.; Kanapatskii, T.A.; Ivanova, E.A.; Gulin, M.B.; Toshchannkov, S.; Pimenov, N.V. Decoupling between sulfate reduction and the anaerobic oxidation of methane in the shallow methane seep of the Black sea. FEMS Microbiol. Lett. 2018, 365. [Google Scholar] [CrossRef] [PubMed]
  58. Tourova, T.P.; Slobodova, N.V.; Bumazhkin, B.K.; Kolganova, T.V.; Muyzer, G.; Sorokin, D.Y. Analysis of community composition of sulfur-oxidizing bacteria in hypersaline and soda lakes using soxB as a functional molecular marker. FEMS Microbiol. Ecol. 2013, 84, 280–289. [Google Scholar] [CrossRef] [Green Version]
  59. AnilKumar, P.; Srinivas, T.N.R. The family Cyclobacteriaceae. In The Prokaryotes: Other Major Lineages of Bacteria and the Archaea; Springer: Berlin, Germany, 2014; pp. 551–575. [Google Scholar]
Figure 1. Physico-chemical properties of water column from Lake Alboraj were recorded during the spring oxycline state (13 March 2019).
Figure 1. Physico-chemical properties of water column from Lake Alboraj were recorded during the spring oxycline state (13 March 2019).
Applsci 11 06309 g001
Figure 2. Relative abundance of major bacterial phyla from Lake Alboraj water and sediment. Only phyla with a relative abundance >0.5% of total bacterial population were included.
Figure 2. Relative abundance of major bacterial phyla from Lake Alboraj water and sediment. Only phyla with a relative abundance >0.5% of total bacterial population were included.
Applsci 11 06309 g002
Figure 3. Relative abundance (%) of major families identified in water (1 m and 4 m) and sediment samples from Lake Alboraj. Only families with a relative abundance >1% of total bacterial population were included.
Figure 3. Relative abundance (%) of major families identified in water (1 m and 4 m) and sediment samples from Lake Alboraj. Only families with a relative abundance >1% of total bacterial population were included.
Applsci 11 06309 g003
Table 1. Chemical and limnological values from water samples from Lake Alboraj in sampling campaign. Depth is given in meters and concentration of ions, in mg/L. -: no data available.
Table 1. Chemical and limnological values from water samples from Lake Alboraj in sampling campaign. Depth is given in meters and concentration of ions, in mg/L. -: no data available.
EpilimnionHypolimnion
Depth144.5
NH4+0.042.20.06
NO310.30.64.7
TIN2.4 1.1
TON1.3 0.5
Ntotal3.73.201.6
Ptotal0.1 0.12
DOC12.914.9
HCO3379398-
SO42−63105426-
Cl13761187-
F8.77.5-
Br2.72.4-
Table 2. Number of sequences and OTUs generated and diversity estimators from MiSeq Illumina sequencing analysis.
Table 2. Number of sequences and OTUs generated and diversity estimators from MiSeq Illumina sequencing analysis.
SampleDepth (m)HabitatSequences
Generated
OTUs
Generated
ShannonPhylogenetic
Diversity
W11Water, epilimnion133,1675586.2451.89
W44Water, hypolimnion115,5635075.7551.00
S4.7Sediment130,5963086.1243.16
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Espín, Y.; Menchén, A.; Moreno, J.L.; Sanz, D.; Álvarez-Ortí, M.; Fernández, J.A.; Gómez-Alday, J.J. Water and Sediment Bacterial Communities in a Small Mediterranean, Oxygen-Stratified, Saline Lake (Lake Alboraj, SE Spain). Appl. Sci. 2021, 11, 6309. https://doi.org/10.3390/app11146309

AMA Style

Espín Y, Menchén A, Moreno JL, Sanz D, Álvarez-Ortí M, Fernández JA, Gómez-Alday JJ. Water and Sediment Bacterial Communities in a Small Mediterranean, Oxygen-Stratified, Saline Lake (Lake Alboraj, SE Spain). Applied Sciences. 2021; 11(14):6309. https://doi.org/10.3390/app11146309

Chicago/Turabian Style

Espín, Yolanda, Alfonso Menchén, José Luis Moreno, David Sanz, Manuel Álvarez-Ortí, José Antonio Fernández, and Juan José Gómez-Alday. 2021. "Water and Sediment Bacterial Communities in a Small Mediterranean, Oxygen-Stratified, Saline Lake (Lake Alboraj, SE Spain)" Applied Sciences 11, no. 14: 6309. https://doi.org/10.3390/app11146309

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