Next Article in Journal
Impact of Social Support on Depressive Symptoms among Postgraduates during the Coronavirus Disease 2019 Pandemic: A Moderated Mediation Analysis
Previous Article in Journal
A Novel Social Distancing Approach for Limiting the Number of Vehicles in Smart Buildings Using LiFi Hybrid-Network
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Water Quality, Heavy Metals, and Antifungal Susceptibility to Fluconazole of Yeasts from Water Systems

by
Luz Dary Caicedo-Bejarano
1,*,
Lizeth Stefania Osorio-Vanegas
2,
Mauricio Ramírez-Castrillón
2,
Jorge Enrique Castillo
3,
Carlos Andrés Martínez-Garay
1 and
Mónica Chávez-Vivas
4
1
Research Group in Mycology (GIM/CICBA), Facultad de Ciencias Básicas, Universidad Santiago de Cali, Santiago de Cali 760035, Colombia
2
Department of Biochemical Engineering, Faculty of Engineering and Design, Universidad Icesi, Santiago de Cali 760031, Colombia
3
Grupo de Investigación en Electroquímica y Ambiente (GIEMA), Facultad de Ciencias Básicas, Universidad Santiago de Cali, Santiago de Cali 760035, Colombia
4
Grupo de Investigación GIMMEIN, Programa de Medicina, Facultad de Salud, Universidad Libre Seccional Cali, Santiago de Cali 760031, Colombia
*
Author to whom correspondence should be addressed.
Int. J. Environ. Res. Public Health 2023, 20(4), 3428; https://doi.org/10.3390/ijerph20043428
Submission received: 28 December 2022 / Revised: 22 January 2023 / Accepted: 23 January 2023 / Published: 15 February 2023
(This article belongs to the Topic Microorganisms in Aquatic Environments)

Abstract

:
Aquatic environments could be reservoirs of pathogenic yeasts with acquired antifungal resistance. The susceptibility to antifungal agents of yeasts present in the wastewater and natural waters of the city of Cali was evaluated. Samples were taken from two types of water: drinking water (Meléndez River, drinking water treatment plant “Puerto Mallarino” in the Cauca River) and wastewater (South Channel of the Cauca River, “Cañaveralejo-PTAR” wastewater treatment plant). Physico-chemical parameters, heavy metal concentration, and yeast levels were determined using standard procedures. Yeasts were identified using API 20 C AUX (BioMérieux) and sequence analysis of the ITS1-5.8S-ITS2 and D1/D2 regions of the large subunit of the ribosome. Susceptibility assays against fluconazole and amphotericin B using the minimum inhibitory concentration (MIC) test were determined using the microdilution method. The influence of physico-chemical parameters and heavy metals was established using principal component analysis (PCA). Yeast counts were higher at WWTP “PTAR” and lower at Melendez River, as expected. A total of 14 genera and 21 yeast species was identified, and the genus Candida was present at all locations. Susceptibility tests showed a 32.7% resistance profile to fluconazole in the order DWTP “Puerto Mallarino = WWTP “PTAR” > South Channel “Navarro”. There were significant differences (p < 0.05) in the physico-chemical parameters/concentration of heavy metals and yeast levels between the aquatic systems under study. A positive association was observed between yeast levels and total dissolved solids, nitrate levels, and Cr at the “PTAR” WWTP; conductivity, Zn, and Cu in the South Channel; and the presence of Pb in the “Puerto Mallarino” DWTP. Rhodotorula mucilaginosa, Candida albicans, and Candida sp. 1 were influenced by Cr and Cd, and Diutina catelunata was influenced by Fe (p < 0.05). The water systems explored in this study showed different yeast levels and susceptibility profiles, and, therefore, possible genetic differences among populations of the same species, and different physico-chemical and heavy metals concentrations, which were probably modulating the antifungal-resistant yeasts. All these aquatic systems discharge their content into the Cauca River. We highlight the importance to further investigate if these resistant communities continue to other locations in the second largest river of Colombia and to determine the risk posed to humans and animals.

1. Introduction

Yeasts constitute an important component in the microbial community of aquatic environments and are comparable to other microorganisms such as bacteria and protozoa [1,2]. Aquatic environments and sediments are important habitats to study microorganism diversity, and yeasts can be easily found in these ecosystems, presenting phenotypic plasticity and adaptations to tolerate salinity, environmental temperature, oxygen saturation, and acidity [3,4,5]. In this sense, several pathogenic or opportunistic yeasts can be isolated from lakes [6,7,8,9], rivers [7,9,10], drinking water [7,11], and wastewater [1,4,5,12]. They are susceptible to acquiring resistance to medical antifungals, such as fluconazole, which is the most widely used agent in the clinic as primary treatment [13], and triazole fungicides, which are frequently used in agriculture [14], while some of them have intrinsic resistance, such as the species Candida krusei, Rhodotorula mucilaginosa, and Meyerozyma guilliermondii [15,16]. Furthermore, yeasts can acquire secondary resistance due to previous contact with antifungals in patients, or low drug activity in tissue or blood associated with immunosuppressed patients [17,18]. The increase in fungal infections has been accentuated by antifungal drug resistance being a serious public health problem. Reports of yeast strains resistant to commonly used antifungal drugs in aquatic environments are becoming more frequent, so these environments can present potential health risks to people who use contaminated water. For example, Medeiros et al. [9] found isolates of Candida spp. resistant to itraconazole and fluconazole in the lakes and rivers of southeastern Brazil. Milanezi et al. [4] conducted a study in the Rio Grande do Sul, Brazil, and found a high frequency of Candida spp. and Debaryomyces spp. resistant to azole antifungals and amphotericin B. Several studies reported the isolation of pathogenic yeasts resistant to antifungal agents from lakes, lagoons, rivers, and wastewater treatment plants (WWTP), mainly in South America [6,8] and South Africa [1,7,10,12]. In Colombia, the intensive use of azole antifungals without medical prescription, or their use as a prophylactic in immunosuppressed patients, increased the number of cases of resistant yeasts, mainly of C. albicans [19] or the C. parapsilosis complex [20]. The approach to this problem has focused mainly on studies from the clinical point of view [19,20,21]. Even though a high load of antifungals in the environment has been documented, more than 80% are azoles. This can be explained to the wide use in clinical therapy and the use of fungicides due to its low cost and systemic action, which allows both the prevention and treatment of fungal diseases, long-lasting stability in the environment and broad antifungal spectrum. [14,17,20,21,22]. Several studies around the world have revealed high levels of antifungal azoles in the aqueous phases of WWTP effluents, after their use in homes, hospitals, or agricultural industries [22,23], and a low elimination rate of these antifungals has been reported [24,25]. At the same time, some yeasts isolated from contaminated water have the potential in biotechnological processes to produce ethanol, biodiesel, and carotenoid pigments, among others [26]. Therefore, the objective of this study was to determine yeast levels in aquatic environments in the city of Cali, Colombia, and their fluconazole and amphotericin b susceptibilities. Furthermore, we aimed to establish their associations with physico-chemical parameters and heavy metal concentrations, with the purpose of driving further research and public policy interventions to control the quality of aquatic systems and strengthen the response to fungal infections and antifungal resistance.

2. Materials and Methods

2.1. Study Area and Sampling

The study was performed in Cali (at coordinates 3°27′00″ N 76°32′00″ W), a city located in southwestern Colombia. The weather of the city is warm, and in the course of the year, the temperature generally varies from 19 °C to 30 °C. Water samples were collected from four water systems: Melendez River (south of the city), the intake from the drinking water treatment plant (DWTP) “Puerto Mallarino” (intake is located on the Cauca River and supplies 76% of drinking water to the city), and wastewater samples were taken from a rainwater channel in Navarro, called “South Channel” (this channel also receives wastewater from a part of the community in the south of the city), and the municipal WWTP (best known as “PTAR”), which discharges its water into the Cauca River, the main river of the city and the second biggest river in Colombia. A schematic illustration of the sampling sites and how they are connected is shown in Figure 1. For fungal isolation, the water samples were taken in triplicate using a plastic container tied to a rigid support, in a sterile 500 mL amber glass container, protected with an aluminum seal and/or Teflon cap, fully filled in January 2018 (dry season) and May 2018 (wet season). The samples were taken at a depth of approximately 0.5 m and as far as possible from the shore, trying not to disturb the bottom and avoiding backwaters or stagnant areas, and were immediately sent to the laboratory in a cooler to be processed and to start the analysis within six hours of taking the sample. One liter (in triplicate) of collected water samples was divided into two equal portions of 500 mL, one of which was stabilized with nitric acid for the determination of heavy metals, while the other was used for the analysis of physico-chemical parameters; samples were kept at 4 °C for no more than two days before being analyzed [27].

2.2. Yeast Isolation and API20C Tests

Serial dilutions (10−1–10−3) were made with peptone water (Oxoid Ltd., Hampshire, United Kingdom). Equal volumes (200 μL) from each dilution were spread over the surface of Dicloran Rose Bengal Chloramphenicol (DRBC) (Merck, Darmstadt, Germany) and CHRomagar Candida™ agar plates (BBL International Inc., London, UK). All the plates were incubated for 48 h at 30 °C. Colony counts were performed on individual plates. Representative yeast colonies were selected and grouped by morphotype, isolated, and conserved using Saboraud Dextrose Agar (Merck) immersed in sterile mineral oil and cryopreserved in glycerol 30% (v/v). A macroscopic evaluation of colonies was performed with a stereoscope (Celestron Lab’s S10-60 Stereo, Celestron, LLC, Torrance, CA, USA) taking account of the shape, edge, surface, appearance, elevation, brightness, and consistency. All the yeasts were evaluated for sugar assimilation using the API 20C AUX system (BioMérieux, Marcy l’Etoile, France) according to the manufacturer’s instructions. Several carbohydrates were evaluated, including D-Glucose, Glycerol, 2-keto-Gluconate calcium, L-Arabinose, D-Xylose, Adonitol, Xylitol, D-Galactose, Inositol, D-Sorbitol, Methyl-D-Glucopyranoside, N-Acetyl-Glucosamine, D-Lactose (Bovine), D-Maltose, D-Sucrose, D-Trehalose, D-Melezitose and D-Raffinose [28]. The galleries were incubated at 30 °C for 72 h in an airtight box containing a small volume of water to create a humid atmosphere. The observation of yeast growth was considered positive. The negative control contained no carbon source, and the positive control contained glucose. The carbohydrate assimilation profile obtained for each tested isolate was interpreted using ApiwebTM software (BioMérieux, reference: 40011).

2.3. Molecular Identification

For DNA extraction, each yeast liquid culture was centrifuged for 2 min at 12,000× g and the supernatant was discarded. DNA was extracted using Kit GeneJet (Thermo Fisher, Waltham, MA, USA) and resuspended in 50 μL of TE (10 mM Tris, 1 mM EDTA, pH 7.4). DNA was visualized on a 1% (w/v) agarose gel and its concentration and purity were determined using a Nanodrop 2000 spectrophotometer (Thermo Scientific, Waltham, MA, USA). The ITS1-5.8-S-ITS2 (Internal transcribed spacer, ITS) and D1/D2 domain of the large subunit of the ribosome (LSU) regions were amplified using the polymerase chain reaction (PCR) procedure in a final volume of 25 μL. A total of 5 μL of genomic DNA (approximately 1 ng/μL) was taken and resuspended in 20 μL of the PCR mixture comprising 0.4 μL primer Forward 20 pmol/μL, 0.4 μL primer Reverse 20 pmol/uL, 1 μL 1 mM dNTPs, 1.5 mL MgCl2 50 mM 0.2 μL sterile milli-Q H2O, 0.6 μL dimethyl sulfoxide (DMSO), 0.2 μL of Taq Polymerase (5 U/μL) (Bioline, London, United Kingdom), and 2.5 μL of 10X Buffer. We used ITS5/ITS4 primers for the ITS region and NL1/NL4 primers for the LSU region. PCR products were analyzed using 1.5% (w/v) agarose gel electrophoresis in 1X TAE buffer (tris base, acetic acid, EDTA, distilled water), and run at 100 V for 1 h. To visualize band migration, the gel was stained with SybrGreen (LONZA) and observed under UV light. A 100-bp or 1-kb ladder (Gibco BRL, Burlington, ON, Canada) was used to estimate amplicon size. Amplification products were purified and sequenced on an ABI 3130 Genetic Analyzer (Applied Biosystems, Waltham, MA, USA) at CorpoGen company. Subsequently, the sequences were edited, assembled, and compared in Genbank and Mycobank databases using the Basic Local Alignment Search Tool (BLAST) algorithm. A 98.41% or 99.5% threshold was used for identification at the taxonomic level of species for ITS and LSU, respectively [29].

2.4. Antifungal Susceptibility Tests to Fluconazole and Amphotericin B

The antifungal susceptibility test was performed using the broth microdilution technique, following the M27-A3 guidelines from the Clinical and Laboratory Standards Institute (CLSI) [30]. Amphotericin B (AMPB, Sigma Chemical Co., St. Louis, MO, USA) was dissolved in dimethyl sulfoxide and fluconazole (FCZ, Pfizer Central Research) was suspended in sterile water. Microtiter plates with 96 round-bottom wells were prepared with a range of final concentrations from 0.03 to 16 μg/mL for Amphotericin B and from 0.125 to 64 μg/mL for FCZ. Serial two-fold dilutions of the various drugs were prepared in RPMI 1640 medium (with L-glutamine, without bicarbonate; Sigma Chemical Co.) and buffered to pH 7.0 using 0.165 M [N-morpholino]propane-sulfonic acid solution (MOPS; Sigma Chemical Co.). The yeast suspension was made at a McFarland concentration of 0.5 (adjusted to final concentrations of approx. 2.5 × 103 CFU/mL). The inoculated plates were incubated for 24–48 h at 35 °C. Strains C. krusei ATCC 6258 and C. parapsilosis ATCC 22,019 were used as controls to detect any abnormalities and deactivation of the antifungal. To determine the MIC, the concentration that produced a reduction in yeast growth of ≥50% compared to control growth without FCZ and total absence of growth for AMPB after 48 h of incubation was considered. Interpretations of the results were based on breakpoints provided by M57S and M27M44S of CLSI [30,31]. Isolates were classified as sensitive (S), dose-dependent sensitive (DDS), and resistant (R).

2.5. Analysis of Physico-Chemical Parameters and Heavy Metals

Physico-chemical parameter analyses (temperature, pH, turbidity, and electrical conductivity) were determined in situ using a 350 multi-parameter probe (Merck, Germany). Water samples were transported on ice to the laboratory to determine salinity, solid particles, dissolved oxygen, total phosphorus, and nitrites according to the standard methods described in the current National Drinking Water Quality Standard (GB5749). The concentrations of the seven heavy metals iron (Fe), copper (Cu), lead (Pb), cadmium (Cd), zinc (Zn), silver (Ag), and chromium (Cr) were determined using a flame atomic absorption spectrophotometer (Model ZEEnit 700P, Analytik Jena, Germany).

2.6. Statistical Analysis

Correlation tests were performed to investigate the relationship between physicochemical parameters, heavy metal concentrations, and fluconazole susceptibility values for the different yeast species. Principal component analysis (PCA) was performed to determine patterns and relationships in the data. Data output was provided in the form of correlation biplots using the packages FactoMineR and ggrepel for R v4.1.1. The correlation significance level was set at p < 0.05. A comparison between the different sampling points for each variable was performed using Kolmogorov–Smirnov tests (p < 0.05).

3. Results

Yeast counts were different for each sample when cultured on a DRBC medium. There were no significant differences in the number of yeasts at each sampling site during the dry and wet seasons. A higher concentration of yeasts was identified in the intake wastewater from the “PTAR” WWTP (2.4 × 105 CFU/mL), while the Melendez River had a lower concentration (<1.0 × 103 CFU/mL). The “Puerto Mallarino” DWTP and South Channel presented intermediate concentrations of yeasts (7.7 × 104 CFU/mL and 3.6 × 104 CFU/mL, respectively). In the present study, we recorded a total of 73 yeast isolates from all sources. Sixty-eight yeast isolates were identified, belonging to 14 genera and 21 species; five isolates were not identified (Table S1). The genus Candida represented most of the yeast isolates. It was interesting that elevated levels of species belonging to Candida (8 × 103 CFU/mL) were also determined in the samples from the “PTAR” WWTP. Chromagar Candida medium showed colonies of three morphotypes: C. albicans, C. tropicalis, and Pichia kudriavzevii (syn. C. krusei). Among the yeast species identified, Rhodotorula mucilaginosa (13.7%), Candida sp. 1 (8.2%), and C. tropicalis (8.2%) were the most frequent (Table S1, Figure 2). However, several isolates were not identifiable to the taxonomic level of species according to the criterium established by Vu et al. [29] using sequence analysis. In this sense, six strains marked as Candida sp. 1 were closely related to C. intermedia, while the strain Candida sp. P46 (Candida sp. 2) was closely related to C. pseudolambica. The strain Naganishia sp. PM20 was related to N. diffluens and two strains of the genus Pichia were related to P. fermentans, with 94.87% of sequence identity compared to the type strain of the species.
The antifungal susceptibility was assessed in 49 yeast isolates after 24–48 h of growth. The reading of the appearance of the biomass was carried out with the help of an inverted mirror. The minimum inhibitory concentration (MIC) breakpoints for FCZ for C. albicans, C. parapsilosis, and C. tropicalis isolates with MIC ≤ 2 µg/mL were considered susceptible, those with MIC 4 µg/mL were considered susceptible dose-dependent (DDS), and those with MIC ≥ 8 µg/mL were considered resistant; C. glabrata isolates with MIC ≤ 32 µg/mL were considered DDS, while isolates with MIC ≥ 64 µg/mL were considered resistant. All P. kudriavzevii (C. krusei) and Rh. mucilaginosa isolates were considered resistant regardless of the MIC value. In the case of amphotericin B, yeasts with an epidemiological cutoff value MIC ≤ 2 µg/mL were considered susceptible.
Yeast isolates with sensitivity to FCZ corresponded to 49%, followed by resistant isolates (32.7%), and dose-dependent sensitive isolates (18.4%). All yeast strains were sensitive to AMPB (Table 1). Inside each aquatic system, the ratios of susceptibilities were different for FCZ. In the South Channel, the ratio of sensitivity was higher (59.1%) compared to other susceptibilities, with the least resistance to FCZ (27.3%). On the other hand, the ratio of sensitivity from the “Puerto Mallarino” DWTP was 36.4% and the percentage of resistant yeasts was 36.4%. This resistant ratio was similar at the “PTAR” WWTP (35.7%). We found variations in sensitivity to FCZ among C. tropicalis isolates according to the site sampled; for example, at South Channel, there were sensitive and DDS strains, but at “PTAR” WWTP we found sensitive and resistant strains (i.e., P4). A similar trend was observed for C. albicans. For example, strains CS55 (South Channel) and P1 (WWTP) showed an MIC of 64 and 8 μg/mL, respectively. However, strains P13B (WWTP) and PM22 (DWTP) were DDS and sensitive, respectively, suggesting variability among populations of this species.
In the present study, PCA was used to establish associations between physico-chemical parameters, the concentrations of heavy metals, and yeast levels from different sampling sites. Table 2 shows the correlation between heavy metals and physico-chemical parameters of water samples and the MIC of FCZ-resistant yeasts. High correlation values (R2 > 0.5, p < 0.05) were obtained for the ratio of iron to T. mucoides and D. catenulate; zinc to C. parapsilosis and D. catenulata; iron to C. parapsilosis; Cd to C. albicans, Rh. mucilaginosa, and Candida sp 1; phosphates to D. hansenii; nitrates to Rh. mucilaginosa; pH to C. tropicalis and D. hansenii; conductivity to C. parapsilosis; and temperature to C. parapsilosis and C. albicans. Figure 3 depicts the distribution of the physico-chemical parameters and heavy metal concentration variables formed by the first two axes, which explained 81.0% of the total variance. A positive association was observed between yeast levels with total dissolved solids and nitrate levels in the WWTP, and with conductivity in the South Channel. The correlation between the levels of antifungal-resistant yeasts and the presence of heavy metals varied between the sampling sites. Yeast levels of C. albicans, C. parapsilosis, Candida sp. 1, P. fermentans, T. coremiiforme, P. laurentii, and G. candidum were influenced by the presence of Pb in the “Puerto Mallarino” DWTP. Furthermore, C. albicans, C. tropicalis, C. krusei, C. aaseri, Candida sp. 2, P. fermentans, S. cerevisiae, P. laurentii, and D. hansenii were influenced by Cr in the “PTAR” WWTP. Finally, C. parapsilosis, C. tropicalis, C. krusei, C. albicans, Candida sp. 1, P. fermentans, D. catenulata, Rh. mucilaginosa, Cr. neoformans, V. humicola, and G. candidum were influenced by Zn and Cu in the South Channel. We also found significant correlations between specific heavy metals and some specific fluconazole-resistant yeast species.

4. Discussion

Aquatic environments can become polluted because of various anthropogenic activities, such as agricultural, industrial, and medical waste production, the expansion of urban communities without adequate sanitation infrastructure, and poor wastewater management [1,10,12]. One of the main indicators of water quality is microorganisms, the main concern being water intended for human consumption and recreational activities, without neglecting surface water, groundwater, and wastewater [32]. Some studies have indicated that the increased contamination of aquatic environments was significantly positively correlated with the relative abundance of yeasts; however, few studies have been conducted to assess the quality of aquatic ecosystems contaminated with yeasts [33,34]. In this study, yeasts were detected at all sampling sites, including the intake of the “Puerto Mallarino” DWTP. At this plant, the order of the process carried out to treat the water is as follows: collection, sand removal, application of activated carbon, pre-chlorination, coagulation, flocculation, sedimentation, and filtration. In addition, there is post-chlorination and chemical stabilization with lime. However, the use of chlorine alone does not guarantee the absence of pathogenic fungi in drinking or recreational water. In fact, Ma et al. showed that there were non-significant changes in fungal community structure observed before and after the initiation of treatment of a hospital hot water system treated with monochloramine in situ [35].
This DWTP follows the parameters established in Colombia, according to decree 1575 of 2007 [36] and resolution 2115 of 2007 [37], that determine the physico-chemical and microbiological quality control of drinking water. The physico-chemical analysis includes the parameters of pH, color, turbidity, nitrate, fluoride, and residual chlorine, among others. Microbiological standards include the absence of pathogenic microorganisms and fecal bacteria (Escherichia coli and Enterococcus spp.), and the determination of the presence of Giardia sp. and Cryptosporidium spp. It should be noted that fungi are not included as potential water contaminants in these documents. The possible reasons for this may be the lack of knowledge of the fungal load in the water, the use of divergent culture methods, the heterogeneous mechanisms of pathogenicity of the fungi, and, consequently, the low number of reports connecting the presence of fungi in the tap water and the appearance of diseases in humans as stated by Kauffmann-Lacroix et al. [38]. Our results suggest that regulations about microbiological procedures should be updated and yeast counting included in routine analyses.
The number of yeasts was significantly higher in the “PTAR” WWTP than Melendez River. The variability of microbial counts found in the present study could be explained by the different levels of water quality of these environments, as they receive different kinds and levels of pollutants in the form of wastewater, industrial effluents, and others. However, the number of heterotrophic yeasts in the Melendez River (1.0 × 103 CFU/mL) was higher than that reported in other rivers around the world, such as Rio Doce, Brazil (from 10 to 466 CFU/mL) [9], some rivers in South Africa (Mooi River (from 0.5 to 9 CFU/mL) [11] and Eersterivier River catchment (1.0 × 102 CFU/mL)) [39], the River Danube in the area of Bratislava (100–210 CFU/mL) [40], and a lake in Patagonia, Argentina (22–141 CFU/mL) [41]. However, at the Tagus estuary, Portugal, a higher level of yeasts was reported (5.3–3272 CFU/mL) [42]. We highlight that the threshold of environmental quality in the eutrophic ecosystem was above 1.0 CFU/mL) [33].
An important aspect was the isolation of C. albicans, C. krusei, and C. parapsilosis, which are related to a high burden of fungal infections in the healthcare environment in Colombia [21]. The highest number of the genus Candida (8 × 103 CFU/mL) was counted at the “PTAR” WWTP. Our results were consistent with the findings of Assress et al. [12], who also found that Candida spp. were the principal species in WWTPs in Gauteng, South Africa. The results demonstrated the attractiveness of using yeasts as a microbiological indicator of organic pollution in aquatic ecosystems [12]. Regarding pigmented yeasts, we found only yeasts from the genus Rhodotorula, with high carotenoid production in almost all isolates [26]. Rh. mucilaginosa was found only at the “PTAR” WWTP and the South Channel in high numbers. These yeasts are associated with skin diseases, and mucosal and invasive fungal infections, especially in immunosuppressed patients, and they present a special metabolic activity in highly eutrophicated waters or severely contaminated municipal wastewaters, which is indicative of their association with anthropogenic activity [43]. Studies on the evaluation of the risk of infection by different species of environmental yeasts has been little described in Colombia. For example, Rhodotorula spp. have been isolated from the nails of patients with superficial mycosis in Antioquia, Colombia [44]. Cr. neoformans is mainly associated with the droppings of various bird species, especially pigeons, but is also isolated from bark, tree trunk hollows, and decaying wood [45], and has been reported in wastewater in previous studies [46]. Cr. neoformans causes opportunistic infections that usually affect the central nervous system of generally immunosuppressed patients, with high mortality in patients who do not receive treatment [47]. In Colombia, the main challenge to treat infections caused by this yeast is resistance to fluconazole [48]. M. guilliermondii is a globally distributed opportunistic pathogen that lives in various habitats. It exists on human skin and in the surface microbiota of mucous membranes. It can cause serious fungal infections such as candidemia [49]. It is a yeast considered a promising species in the field of biotechnology, especially in the biocontrol of moldy bacteria during the storage of fruits and vegetables [50]. The presence of these yeasts is understandable, since city wastewater effluents containing antimicrobials, antimicrobial resistance-carrying microbes, and antimicrobial resistance genes are discharged at these two sites. Among the most reported opportunistic yeasts we found M. guilliermondii, C. lusitaniae, C. tropicalis, P. laurentii, Rh. glutinis, and Rh. mucilaginosa in surface waters [11,39].
The main fungal pathogens in humans are Candida species, which develop resistance to triazoles and echinocandins. However, resistance to AMPB (Polyene) is extremely rare, despite its use for more than 60 years as a monotherapy, mainly for invasive fungal infections [51,52]. For this reason, it is interesting to know whether amphotericin B-resistant yeast strains can be found in the environment. The results obtained with the MIC assay showed that all strains of the different yeast species (including Candida species) were found to be sensitive to AMPB, with values at or below 2 μg/mL, confirming that resistance to this antifungal is rare in the assessed yeasts, and suggesting a low interaction of this antifungal with yeast communities in this location. This was not the case for FCZ, where 32.7% of yeast species were resistant to FCZ. These results were consistent with results obtained in previous studies focusing on the isolation of antifungal-resistant environmental yeast. Brilhante et al. reported azole resistance in Candida spp. isolated from Catú Lake, Ceará, Brazil in a high number [53]. Similarly, Medeiros et al. reported in 2008 that 50% of the yeasts isolated from water samples of the lakes and rivers of the Rio Doce, Brazil showed resistance to itraconazole and less resistance to FCZ [54]. However, an evaluation carried out in 2012 showed significant growth of yeast resistant to FCZ [9]. The yeasts isolated at the South Channel (Navarro) showed the highest percentage of resistance (37.5%) to FCZ, possibly associated with the high load of contaminants that this canal receives along the south of the city.
Table 1 illustrates the pathogenic species C. tropicalis, C. albicans, C. parapsilosis, D. catelunata, Rh. mucilaginosa, and D. hansenii, showing the highest MIC values to FCZ (64 μg/mL). Similarly, antifungal resistance to FCZ has been reported in Candida species from tropical freshwater environments in Brazil and China [9,53,54,55]. The presence of environmental yeasts with high resistance to FCZ is probably due to the fact that this azole is the most widely used antifungal in humans and animals, and it is among the most reported drugs in hospital and home wastewater [56,57]. In addition, there are fungicides routinely used in agriculture that share the action mechanism of azoles, which can generate cross-resistance in yeasts found in these aquatic environments [56,57,58]. In Asia, some studies reported a significant increase in resistant C. tropicalis isolated from poultry, and this has been associated with the extensive use of azoles in agriculture [56,59,60]. Brilhante et al. suggested that azole resistance in the Candida strains recovered from aquatic environments would be influenced by the activity of efflux pumps [53]. Several chemical compounds present in the environment, when interacting with microorganisms, would trigger the expression of genes that lead to these efflux pumps or other proteins involved in resistance to antifungals [56,57,58].
Rhodotorula, Candida, Pichia, and Trichosporon are yeasts typical of strongly eutrophic waters [61], and their presence indicates that aquatic systems are significantly polluted by industrial and municipal wastewater, where they are able to metabolize aromatic substances and heavy metals and have been considered as good indicators of pollution [1,8,33,62]. It is not surprising to find these yeasts in the environment; however, in a report by the WHO, which included a list of fungal priority pathogens, it recommended carrying out this type of study in order to provide epidemiological data that allow the establishment of local priorities in relation to the dynamics of fungal pathogens and the prevalence of resistance to reduce the impact of drug resistance worldwide [63].
The yeast species compositions in the aquatic systems studied were dissimilar and were conditioned by the type of water, but also by the processing conditions. Yeast culture is an effective way to study their diversity, as many more yeast species, especially some possible new species, were recorded and isolated using the culture method. We found an association between yeast levels with certain physico-chemical parameters, such as total dissolved solids, levels of nitrites, and conductivity in the “PTAR” WWTP and South Channel. These results can be explained by the high load of waste that these two places receive, especially due to domestic discharges and agricultural activities that cause an increase in the concentration of organic matter and salts that come from fertilizers. Bafico demonstrated in his study that pollution from urban sources strongly influences the periphytic community structure and dynamics of Lake Nahuel Huapi (Patagonia, Argentina), with greater biomass and cell densities at highly contaminated sites [41,64]. In addition, the highest number of fluconazole-resistant yeast isolates was detected at these two sites (31.3% and 37.5%, respectively). The observed spatial variation could be attributed to site-specific anthropogenic activities, such as eutrophication, the influx of domestic and industrial waste, and run-off from agricultural settings. Effluents from the Melendez River, rainwater at the South Channel (Navarro), and the WWTP (PTAR) discharge into the Cauca River. In addition, the vicinity of the river presents different levels of human interactions (e.g., illegal mining, steel making, informal settlements, sewage and raw pharmaceuticals discharge, and poultry, agricultural, and industrial waste), which influence yeast counts and the species present, and the activation or acquisition of antifungal resistance mechanisms.
The yeast count obtained during the dry and wet season did not present significant differences at each sampling site, with the reported number corresponding to the average of the two sampling moments. Colombia, being a tropical country, experiences changes in climatic conditions that are conditioned by two phenomena, the “La Niña” phenomenon (there can be rain all year with some sunny days) and the “El Niño” phenomenon (there can be dry days all year with some rainy days). This means that temperature and relative humidity conditions each year are not very variable. Precipitation in January and May 2018 was 50–100 mm and 100–150 mm, respectively, with average temperatures for the two months between 20 and 28 °C [65]. The yeast count obtained during the dry and wet season did not present significant differences at each sampling site, with the reported number corresponding to the average of the two sampling moments. For this reason, the data from the two sampling moments were averaged. The results differed from those found by Steffen et al., who found significant differences in site-specific yeast concentrations in the dry summer months, when concentrations increased significantly during the rainy months [39], and from the results of Ortíz-Vera et al., who found greater differences in community composition (relative abundance of fungal phyla) in the dry than in the rainy season [66].
On the other hand, heavy metal pollution in aquatic systems is a global problem originating from increased industrialization and urbanization, since they are accumulative, toxic, and carcinogenic in water bodies and biota. As seen from the result of the PCA, there was a significantly positive correlation between levels of Candida spp., P. fermentans, among others, and heavy metals, especially Cr, Zn, Cu, and Pb. There is little open literature on the effect of metal levels on the diversity and structure of the fungal community in aquatic environments. Assress et al. found that the distribution of classes such as Pezizomycetes, Lecanoromycetes, Agaricostilbomycetes, Schizosaccharomycetes, and Dothideomycetes was influenced by Mg and Zn concentrations [12]. Meanwhile, members of classes Eurotiomycetes, Exobasidiomycetes, Orbiliomycetes, Glomeromycetes, Saccharomycetes and Leotiomycetes were correlated with Ni and Mn levels, and Fe was the main environmental parameter influencing the fungal community belonging to classes Agaricomycetes, Pucchinomycetes, Atractiellomycetes, Sordariomycetes, and Archaeorhizomycetes in WWTPs located in Gauteng Province, South Africa.
Some studies have shown that the content of organic matter, calcium, and amorphous phase can increase the sensitivity to heavy metal contamination and the persistence of organic pollutants in different matrices (water, air, and soil), and interfere or inhibit some enzymes and metabolic processes [67].
It is also established that the organic matter content is known to positively influence the cation exchange capacity, buffering capacity, and retention of heavy metals; for example, some physico-chemical factors, such as pH and oxidation reduction potential, can influence the increased toxicity and genotoxicity of some yeasts to heavy metals [68,69].
It has been shown that the presence of heavy metals in aquatic environments influences the structure of bacterial communities resistant to antibiotics and the permanence over long periods of time of genes resistant to antibiotics by forming stable complexes of antibiotics and metal ions [70]. Dickinson et al. demonstrated the co-selection of heavy metals and antibiotics in the environment [71]. The prevalence and persistence of resistant genes to antibiotics and heavy metals have been demonstrated in WWTPs [72]. High levels of antibiotic resistant genes are reported to be related to high concentrations of Cr [73], Hg, and Zn [74]. Rajasekar et al. found that Cd, Pb, and Cu had a significant positive correlation with the sul2 and strB genes, conferring resistance to sulfonamides and streptomycin, respectively [75]. This would explain the results of our study, where it was found that Rh. mucilaginosa, C. albicans, and Candida sp. 1 were positively associated with the presence of Cd, and D. catelunata was associated with Fe.
Saccharomyces cerevisiae and some pathogenic yeasts, such as C. glabrata and C. albicans, have been reported to have ATP-binding cassette (ABC) transporters, which have evolved to play critical roles in adapting to environmental challenges such as heavy metal stress and multidrug resistance (PDR). For example, Yap1, Yap2, and Yap8 transporters regulate the response to oxidative and heavy metal stress by inducing the expression of the yeast activator protein YCF1, the pleiotropic drug resistance elements PDR5, and the multidrug transporter involved in multidrug resistance and singlet oxygen species resistance (SNQ2) in response to a variety of toxic metals. Yeast ABC transporters play an important role in plasma membrane homeostasis, the site of action of antifungals, such as polyenes and azoles, and are involved in resistance to different drugs, which could explain the correlation between the presence of some heavy metals and the susceptibility to fluconazole of some yeasts isolated in this study [76,77]. In relation to the presence of heavy metals in soils, Suillus luteus uses different mechanisms, such as the exclusion and regulation of heavy metals through transporters located in the cell membrane and chelating agents, which are released to the outside of the cell wall, through storage, and the trapped ions are stored or transported in vesicles and expelled from the cells, in the same way that oxidizing agents relieve oxidative stress caused by reactive oxygen species through reducing agents such as cytochromes p450 (CYP450) [78]. CYP450 exists in a large number of organisms, including bacteria, fungi, and mammals, catalyzing a variety of oxidation reactions that allow the detoxification of xenobiotics, such as heavy metals, the metabolism of drugs, such as antifungals, and the biosynthesis of steroids, such as ergosterol [79]. One of these antifungals, fluconazole, acts on ergosterol biosynthesis in fungal cells by inhibiting a CYP450-dependent fungal enzyme, lanosterol 14α-sterol demethylase. Resistance to fluconazole can occur through the increased expression of the ERG11 gene encoding the drug target enzyme, sterol 14α-demethylase (Erg11p), or by mutations in Erg11p that result in a reduced affinity for fluconazole and an overexpression of efflux pumps that transport fluconazole out of the cell [80].
However, more extensive studies are needed to determine the influence of physico-chemical parameters, heavy metals, and other contaminating compounds in yeasts resistant to Cd antifungals in water systems.

5. Study Limitations

Our study was based on cultivated yeasts, which limited the diversity analysis, so it will be important in subsequent studies to carry out metagenomic shotgun and metatranscriptome experiments that would allow us to further explore the mycodiversity and functional genes that are being expressed in water systems.
It is also important to include other antifungals in the antimicrobial susceptibility analysis, and to perform an evaluation of the virulence profile, such as the production of hydrolytic enzymes, and the adhesion capacity and formation of biofilms, which would allow us to clarify the possible route of transmission of these pathogens to humans.

6. Conclusions

We found opportunistic yeast pathogens, such as strain-specific azole-resistant C. albicans, C. parapsilosis, C. tropicalis, and Candida sp. 1 (closely related to C. intermedia), in different aquatic environments, including a WWTP and DWTP. Melendez River was an exception. These results are disturbing since this drug is frequently used as a prophylactic treatment in HIV-positive patients. We emphasized the importance of carrying out studies focusing on the isolation of antifungal-resistant environmental yeasts to investigate their participation in the deterioration of aquatic environments.
The results of the correlation analysis showed a general spatial variation In the characteristics of the wastewater samples. Yeast counts and some species considered opportunistic pathogens, such as C. albicans, Rh. mucilaginosa, and D. catelunata, showed an association with parameters and chemical compounds indicative of the presence of sewage or contaminated water, which act as selection agents for pathogenic yeasts.
The results indicated that the aquatic systems of the city of Cali are a reservoir of FCZ-resistant yeasts and a potential source of invasive fungal infections, which is important for the “One Health” approach, and, like other ecosystem services, these must be protected and managed in a sustainable manner.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijerph20043428/s1, Table S1: Yeast strains were analyzed in this study. All samples were identified with the biochemical tests API 20C and were confirmed using sequence analysis. Yeast strains labeled with “CS” were isolated from the South Channel, Navarro; yeast strains labeled with “P” were isolated from WWTP “PTAR”; yeast strains labeled with “PM” were isolated from DWTP “Puerto Mallarino”, Cauca River; and yeast strains labeled with “M” were isolated from Melendez River. ND: not determined. ITS = Internal Transcribed Spacers; LSU = D1/D2 domain of the Large Subunit of the Ribosome. References [26,81] are cited in the Supplementary Materials.

Author Contributions

Conceptualization, L.D.C.-B. and M.C.-V.; data curation, M.R.-C. and M.C.-V.; formal analysis, L.D.C.-B., M.R.-C., J.E.C., C.A.M.-G. and M.C.-V.; funding acquisition, L.D.C.-B.; investigation, L.D.C.-B., L.S.O.-V., M.R.-C., J.E.C., C.A.M.-G. and M.C.-V.; methodology, L.D.C.-B., L.S.O.-V., M.R.-C., J.E.C. and M.C.-V.; project administration, L.D.C.-B.; validation, M.C.-V.; visualization, M.R.-C.; writing—original draft, L.D.C.-B., L.S.O.-V. and M.R.-C.; writing—review and editing, L.D.C.-B., M.R.-C., C.A.M.-G. and M.C.-V. All authors have read and agreed to the published version of the manuscript.

Funding

This research has been funded by Dirección General de Investigaciones of Universidad Santiago de Cali under call No. 02-2023 and grant number 934-621115-B5.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
WWTPWastewater Treatment Plant
DWTPDrinking Water Treatment Plant
PTARPlanta de Tratamiento de Aguas Residuales
ITSInternal Transcribed Spacer 1, gene rDNA 5.8S and Internal Transcribed Spacer 2
LSUD1/D2 Domain of the Large Subunit of the Ribosome 26S
CFUColony Forming Units
MICMinimal Inhibitory Concentration
FCZFluconazole
AMPBAmphotericin B
PCAPrincipal Component Analysis

References

  1. Monapathi, M.; Bezuidenhout, C.; Rhode, O. Physico-chemical parameters and culturable yeast diversity in surface water: A consequence of pollution. Water SA 2020, 46, 593–601. [Google Scholar] [CrossRef]
  2. Yurkov, A.; Pozo, M.I. Yeasts community composition and structure. In Yeasts in Natural Ecosystems: Ecology; Buzzini, P., Lachance, M.-A., Yurkov, A., Eds.; Springer Nature: Cham, Switzerland, 2017; pp. 73–100. [Google Scholar]
  3. Silva-Bedoya, L.M.; Ramírez-Castrillón, M.; Osorio-Cadavid, E. Yeast diversity associated to sediments and water from two colombian artificial lakes. Braz. J. Microbiol. 2014, 45, 135–142. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Milanezi, A.C.M.; Witusk, J.P.D.; van der Sand, S.T. Antifungal susceptibility of yeasts isolated from anthropogenic watershed. An. Acad. Bras. Cienc. 2019, 91, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Tran, N.H.; Reinhard, M.; Gin, K.Y. Occurrence and fate of emerging contaminants in municipal wastewater treatment plants from different geographical regions—A review. Water Res. 2018, 133, 182–207. [Google Scholar] [CrossRef] [PubMed]
  6. Pagani, D.M.; Heidrich, D.; Tormente, F.; Milani, G.; Jank, L.; Haig, N.H.; Valente, P.; Scroferneker, M.L. High MICs for antifungal agents in yeasts from an anthropized lagoon in South America. Microbiol. Res. 2022, 262, 127083. [Google Scholar] [CrossRef]
  7. Monapathi, M.; Horn, S.; Vogt, T.; van Wyk, D.; Mienie, C.; Ezeokoli, O.T.; Coertze, R.; Rhode, O.; Bezuidenhout, C.C. Antifungal agents, yeast abundance and diversity in surface water: Potential risks to water users. Chemosphere 2021, 274, 129718. [Google Scholar] [CrossRef]
  8. Brandao, L.R.; Medeiros, A.O.; Duarte, M.C.; Barbosa, A.C.; Rosa, C.A. Diversity and antifungal susceptibility of yeasts isolated by multiple-tube fermentation from three freshwater lakes in Brazil. J. Water Health 2010, 8, 279–289. [Google Scholar] [CrossRef] [Green Version]
  9. Medeiros, A.O.; Missagia, B.S.; Brandão, L.R.; Callisto, M.; Barbosa, F.A.R.; Rosa, C.A. Water quality and diversity of teasts from tropical lakes and rivers from the Rio Doce Basin in Southeastern Brazil. Braz. J. Microbiol. 2012, 43, 1582–1594. [Google Scholar] [CrossRef] [Green Version]
  10. Monapathi, M.E.; Bezuidenhout, C.C.; Rhode, O.H.J. Water Quality and Antifungal susceptibility of opportunistic yeast pathogens from rivers. Water Sci. Technol. 2017, 75, 1319–1331. [Google Scholar] [CrossRef]
  11. Van Wyk, D.A.B.; Bezuidenhout, C.C.; Rhode, O.H.J. Diversity and characteristics of yeasts from water sources in the North West Province, South Africa. Water Sci. Technol. 2012, 12, 422–430. [Google Scholar] [CrossRef]
  12. Assress, H.A.; Selvarajan, R.; Nyoni, H.; Ntushelo, K.; Mamba, B.B.; Msagati, T.A.M. Diversity, Co-occurrence and implications of fungal communities in wastewater treatment plants. Sci. Rep. 2019, 9, 14056. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Nucci, M.; Queiroz-Telles, F.; Alvarado-Matute, T.; Tiraboschi, I.N.; Cortes, J.; Zurita, J.; Guzman-Blanco, M.; Santolaya, M.E.; Thompson, L.; Sifuentes-Osornio, J.; et al. Epidemiology of candidemia in Latin America: A laboratory-based survey. PLoS ONE 2013, 8, e59373. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Castelo-Branco, D.; Lockhart, S.R.; Chen, Y.; Santos, D.A.; Hagen, F.; Hawkins, N.J.; Lavergne, R.; Meis, J.F.; Pape, L.P.; Rocha, M.F.G.; et al. Collateral consequences of agricultural fungicides on pathogenic yeasts: A one health perspective to tackle azole resistance. Mycoses 2022, 65, 303–311. [Google Scholar] [CrossRef] [PubMed]
  15. Spader, T.B.; Ramírez-Castrillón, M.; Valente, P.; Alves, S.H.; Severo, L.C. In vitro interactions of amphotericin B combined with non-antifungal agents against Rhodotorula mucilaginosa Strains. Mycopathologia 2019, 184, 35–43. [Google Scholar] [CrossRef] [PubMed]
  16. Desnos-Ollivier, M.; Lortholary, O.; Bretagne, S.; Dromer, F. Azole susceptibility profiles of more than 9000 clinical yeast isolates belonging to 40 common and rare species. Antimicrob. Agents Chemother. 2021, 65, e02615-20. [Google Scholar] [CrossRef]
  17. Polvi, E.J.; Li, X.; O´Meara, T.R.; Leach, M.D.; Cowen, L.E. Opportunistic yeast pathogens: Reservoirs, virulence mechanisms, and therapeutic strategies. Cell. Mol. Life Sci. 2015, 72, 2261–2287. [Google Scholar] [CrossRef]
  18. López-Ávila, K.; Dzul-Rosado, K.R.; Lugo-Caballero, C.; Arias-León, J.J.; Zavala-Castro, J.E. Mecanismos de resistencia antifúngica de los azoles en Candida albicans. Una revisión. Rev. Bioméd. 2016, 27, 127–136. [Google Scholar] [CrossRef] [Green Version]
  19. Castro, L.Á.; Álvarez, M.I.; Martínez, E. Candida en la cavidad oral de pacientes con VIH en Cali, Colombia: Determinación de especies y sensibilidad al fluconazol. Iatreia 2015, 28, 368–377. [Google Scholar] [CrossRef] [Green Version]
  20. Demirci-Duarte, S.; Arikan-Akdagli, S.; Gülmez, D. Species distribution, azole resistance and related molecular mechanisms in invasive Candida parapsilosis complex isolates: Increase in fluconazole resistance in 21 years. Mycoses 2021, 64, 823–830. [Google Scholar] [CrossRef]
  21. Alvarez-Moreno, C.A.; Cortes, J.A.; Denning, D.W. Burden of fungal infections in Colombia. J. Fungi 2018, 4, 41. [Google Scholar] [CrossRef]
  22. Chen, Z.-F.; Ying, G.-G. Occurrence, fate and ecological risk of five typical azole fungicides as therapeutic and personal care products in the environment: A review. Environ. Int. 2015, 84, 142–153. [Google Scholar] [CrossRef] [PubMed]
  23. Peng, X.; Huang, Q.; Zhang, K.; Yu, Y.; Wang, Z.; Wang, C. Distribution, behavior and fate of azole antifungals during mechanical, biological, and chemical treatments in sewage treatment plants in China. Sci. Total Environ. 2012, 426, 311–317. [Google Scholar] [CrossRef] [PubMed]
  24. Assress, H.A.; Nyoni, H.; Mamba, B.B.; Msagati, T.A.M. Occurrence and risk assessment of azole antifungal drugs in water and wastewater. Ecotoxicol. Environ. Saf. 2020, 187, 109868. [Google Scholar] [CrossRef] [PubMed]
  25. Huang, Q.; Wang, Z.; Wang, C.; Peng, X. Chiral profiling of azole antifungals in municipal wastewater and recipient rivers of the Pearl River Delta, China. Environ. Sci. Pollut. Res. 2013, 20, 8890–8899. [Google Scholar] [CrossRef] [PubMed]
  26. Villota, C.S.V.; Vargas, S.A.F.; Martínez, G.C.A.; Caicedo, B.L.D.; Osorio, C.E.; Ramírez-Castrillón, M. Producción de carotenoides en levaduras nativas aisladas de sistemas acuáticos en Cali, Colombia. Rev. Asoc. Colomb. Cien. Biol. 2020, 1, 103–114. [Google Scholar] [CrossRef]
  27. Uasapud Garcia, J.D. Instructivo de Toma y Preservación de Muestras Sedimentos y Agua Superficial para la Red de Monitoreo de Calidad del IDEAM 2020, 21. Available online: http://sgi.ideam.gov.co (accessed on 9 June 2018).
  28. Otero Silva, J.; Candido, R.C. Avaliação do sistema API20C AUX na identificação de leveduras de interesse clínico. Rev. Soc. Bras. Med. Trop. 2005, 38, 261–263. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Vu, D.; Groenewald, M.; Szöke, S.; Cardinali, G.; Eberhardt, U.; Stielow, B.; de Vries, M.; Verkleij, G.J.M.; Crous, P.W.; Boekhout, T.; et al. DNA barcoding analysis of more than 9 000 yeast isolates contributes to quantitative thresholds for yeast species and genera delimitation. Stud. Mycol. 2016, 85, 91–105. [Google Scholar] [CrossRef]
  30. Clinical and Laboratory Standards Institute (CLSI). Perfomance standards for antifungal susceptibility testing of yeasts. In CLSI Supplement M27M44S; Clinical and Laboratory Standards Institute: Wayne, PA, USA, 2022; p. 40. [Google Scholar]
  31. Clinical and Laboratory Standards Institute (CLSI). Epidemiological cutoff values for antifungal susceptibility testing. In CLSI Supplement M57S; Clinical and Laboratory Standards Institute: Wayne, PA, USA, 2022; p. 40. [Google Scholar]
  32. Libkind, D.; Buzzini, P.; Turchetti, B.; Rosa, C.A. Yeasts in continental and seawater. In Yeasts in Natural Ecosystems: Diversity; Buzzini, P., Lachance, M.-A., Yurkov, A., Eds.; Springer: Cham, Switzerland, 2017; pp. 1–61. ISBN 9783319626833. [Google Scholar]
  33. Montanari, L.B.; Garcia Sartori, F.; Montrazi Ribeiro, D.B.; Leandro, L.F.; Pires, R.H.; Melhem, M.d.S.C.; de Mello, C.A.; Gomes Martins, C.H. Yeast isolation and identification in water used in a brazilian hemodialysis unit by classic microbiological techniques and Raman spectroscopy. J. Water Health 2018, 16, 311–320. [Google Scholar] [CrossRef] [Green Version]
  34. Hagler, A.N. Yeasts as indicators of environmental quality. In Biodiversity and Ecophysiology of Yeasts; Péter, G., Rosa, C., Eds.; Springer: Berlin/Heidelberg, Germany, 2006; pp. 515–532. [Google Scholar]
  35. Ma, X.; Baron, J.L.; Vikram, A.; Stout, J.E.; Bibby, K. Fungal Diversity and presence of potentially pathogenic fungi in a hospital hot water system treated with on-site monochloramine. Water Res. 2015, 71, 197–206. [Google Scholar] [CrossRef]
  36. Departamento Administrativo de la Función Pública, Colombia. Decreto 1575 de 2007. Sistema para la Protección y Control de la Calidad del Agua para Consumo Humano. Bogota, Colombia. 9 May 2007. Available online: https://www.funcionpublica.gov.co/eva/gestornormativo/norma.php?i=30007 (accessed on 21 January 2023).
  37. Ministerio de la Protección Social, Colombia; Ministerio de Ambiente, Vivienda y Desarrollo Territorial, Colombia. Resolución 2115 de 2007 Caracteristicas, Instrumentos Basicos y Frecuencias del Sistema de Control y Vigilancia para la Calidad del Agua para Consumo Humano. Bogota, Colombia. 22 June 2007; p. 23. Available online: https://www.minsalud.gov.co/sites/rid/Lists/BibliotecaDigital/RIDE/DE/DIJ/Resolución_2115_de_2007.pdf (accessed on 21 January 2023).
  38. Kauffmann–Lacroix, C.; Costa, D.; Imbert, C. Fungi, Water Supply and Biofilms. In Fungal Biofilms and Related Infections; Imbert, C., Ed.; Springer: Cham, Switzerland, 2016; Volume 931, pp. 49–61. [Google Scholar]
  39. Christa Steffen, H.; Smith, K.; van Deventer, C.; Weiskerger, C.; Bosch, C.; Brandão, J.; Wolfaardt, G.; Botha, A. Health risk posed by direct ingestion of yeasts from polluted river water. Water Res. 2023, 231, 119599. [Google Scholar] [CrossRef]
  40. Sláviková, E.; Vadkertiová, R. Seasonal Occurrence of yeasts and yeast-like organisms in the River Danube. Antonie Van Leeuwenhoek 1997, 72, 77–80. [Google Scholar] [CrossRef] [PubMed]
  41. Brandão, L.R.; Libkind, D.; Vaz, A.B.M.; Espírito Santo, L.C.; Moliné, M.; de García, V.; van Broock, M.; Rosa, C.A. Yeasts from an oligotrophic lake in Patagonia (Argentina): Diversity, distribution and synthesis of photoprotective compounds and extracellular enzymes. FEMS Microbiol. Ecol. 2011, 76, 1–13. [Google Scholar] [CrossRef] [PubMed]
  42. de Almeida, J.M.G.C.F. Yeast community survey in the Tagus estuary. FEMS Microbiol. Ecol. 2005, 53, 295–303. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Dominguez Cespedes, L.; Cespedes Fonseca, Y.M.; Rojas Sanchez, Y.V. Rhodotorula an emergent pathogen. J. Exp. Clin. Microbiol. 2022, 6, 51–54. [Google Scholar] [CrossRef]
  44. Mejía-Arango, M.A.; Santa-Vélez, C.; Manuela, C.-S.; Vélez, L.M.; Colmenares, L.M.; Restrepo-Jaramillo, B.N.; Cardona-Castro, N. Estudio etiológico y epidemiológico de las micosis cutáneas en un laboratorio de referencia-Antioquia-Colombia. Rev. CES Med. 2013, 27, 7–19. [Google Scholar]
  45. Rathore, S.S.; Sathiyamoorthy, J.; Lalitha, C.; Ramakrishnan, J. A holistic review on Cryptococcus neoformans. Microb. Pathog. 2022, 166, 105521. [Google Scholar] [CrossRef]
  46. Hamed, S. Isolation and identification of yeasts along wastewater treatment lines at Zagazig plant. Ann. Agric. Sci. Moshtohor 2016, 54, 77–84. [Google Scholar] [CrossRef]
  47. Rajasingham, R.; Govender, N.P.; Jordan, A.; Loyse, A.; Shroufi, A.; Denning, D.W.; Meya, D.B.; Chiller, T.M.; Boulware, D.R. The global burden of HIV-associated cryptococcal infection in adults in 2020: A modelling analysis. Lancet Infect. Dis. 2022, 22, 1748–1755. [Google Scholar] [CrossRef]
  48. Melendres, J.; Carvajal-Valencia, S.; Escandon, P.; Firacative, C. S8.5d Cryptococcus neoformans and Cryptococcus gattii clinical isolates from Colombia develop heteroresistance to fluconazole at high concentrations. Med. Mycol. 2022, 60, 2022. [Google Scholar] [CrossRef]
  49. Cebeci Güler, N.; Tosun, İ.; Aydin, F. The identification of Meyerozyma guilliermondii from blood cultures and surveillance samples in a university hospital in Northeast Turkey: A ten-year survey. J. Mycol. Med. 2017, 27, 506–513. [Google Scholar] [CrossRef]
  50. Zhang, D.; Spadaro, D.; Garibaldi, A.; Gullino, M.L. Potential biocontrol activity of a strain of Pichia guilliermondii against grey mold of apples and its possible modes of action. Biol. Control 2011, 57, 193–201. [Google Scholar] [CrossRef]
  51. Bohner, F.; Papp, C.; Gácser, A. The effect of antifungal resistance development on the virulence of Candida species. FEMS Yeast Res. 2022, 22, 1–10. [Google Scholar] [CrossRef] [PubMed]
  52. Tevyashova, A.; Efimova, S.; Alexandrov, A.; Omelchuk, O.; Ghazy, E.; Bychkova, E.; Zatonsky, G.; Grammatikova, N.; Dezhenkova, L.; Solovieva, S.; et al. Semisynthetic amides of amphotericin B and nynstatin A1: A comparative study of in vitro activity/toxicity ratio in relation to selectivity to ergosterol membranes. Antibiotics 2023, 12, 151–170. [Google Scholar] [CrossRef]
  53. Brilhante, R.S.N.; Paiva, M.A.N.; Sampaio, C.M.S.; Castelo-Branco, D.S.C.M.; Teixeira, C.E.C.; de Alencar, L.P.; Bandeira, T.J.P.G.; Monteiro, A.J.; Cordeiro, R.A.; Pereira-Neto, W.A.; et al. Azole resistance in Candida spp. isolated from Catú Lake, Ceará, Brazil: An efflux-pump-mediated mechanism. Braz. J. Microbiol. 2016, 47, 33–38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Medeiros, A.O.; Kohler, L.M.; Hamdan, J.S.; Missagia, B.S.; Barbosa, F.A.R.; Rosa, C.A. Diversity and antifungal susceptibility of yeasts from tropical freshwater environments in Southeastern Brazil. Water Res. 2008, 42, 3921–3929. [Google Scholar] [CrossRef]
  55. Liu, Y.; Chen, Z.; Li, J.; Zhu, Z.; Pang, S.; Xu, J.; Wu, J. Extensive diversity and prevalent fluconazole resistance among environmental yeasts from tropical China. Genes 2022, 13, 444–467. [Google Scholar] [CrossRef] [PubMed]
  56. Lima, R.; Ribeiro, F.C.; Colombo, A.L.; de Almeida, J.N., Jr. The emerging threat antifungal-resistant Candida tropicalis in humans, animals, and environment. Front. Fungal Biol. 2022, 3, 957021. [Google Scholar] [CrossRef]
  57. Paul, S.; Shaw, D.; Joshi, H.; Singh, S.; Chakrabarti, A.; Rudramurthy, S.M.; Ghosh, A.K. Mechanisms of azole antifungal resistance in clinical isolates of Candida tropicalis. PLoS ONE 2022, 17, e0269721. [Google Scholar] [CrossRef]
  58. Forastiero, A.; Mesa-Arango, A.C.; Alastruey-Izquierdo, A.; Alcazar-Fuoli, L.; Bernal-Martinez, L.; Pelaez, T.; Lopez, J.F.; Grimalt, J.O.; Gomez-Lopez, A.; Cuesta, I.; et al. Candida tropicalis antifungal cross-resistance is related to different azole target (Erg11p) modifications. Antimicrob. Agents Chemother. 2013, 57, 4769–4781. [Google Scholar] [CrossRef] [Green Version]
  59. Cordeiro, R.d.A.; de Oliveira, J.S.; Castelo-Branco, S.C.M.; Cordeiro Teixeira, C.E.; de Farias Marques, F.J.; Bittencourt, P.V.; Carvalho, V.L.; Gomes Bandeira, T.d.J.P.; Nogueira Brilhante, R.S.; Bezerra Moreira, J.L.; et al. Candida tropicalis isolates obtained from veterinary sources show resistance to azoles and produce virulence factors. Med. Mycol. 2015, 53, 145–152. [Google Scholar] [CrossRef] [Green Version]
  60. Subramanya, S.H.; Sharan, N.K.; Baral, B.P.; Hamal, D.; Nayak, N.; Prakash, P.Y.; Sathian, B.; Bairy, I.; Gokhale, S. Diversity, in-vitro virulence traits and antifungal susceptibility pattern of gastrointestinal yeast flora of healthy poultry, Gallus Gallus Domesticus. BMC Microbiol. 2017, 17, 113–125. [Google Scholar] [CrossRef] [PubMed]
  61. Boguslawska-Was, E.; Dabrowski, W. The seasonal variability of yeasts and yeast-like organisms in water and bottom sediment of the Szczecin lagoon. Int. J. Hyg. Environ. Health 2001, 203, 451–458. [Google Scholar] [CrossRef] [PubMed]
  62. Pietryczuk, A.; Cudowski, A.; Hauschild, T. Effect of trophic status in lakes on fungal species diversity and abundance. Ecotoxicol. Environ. Saf. 2014, 109, 32–37. [Google Scholar] [CrossRef]
  63. World Health Organization W.H.O. Fungal Priority Pathogens List to Guide Research, Development and Public Health Action, 1st ed.; World Health Organization: Geneva, Switzerland, 2022. [Google Scholar]
  64. Baffico, G.D. Variations in the periphytic community structure and dynamics of lake Nahuel Huapi (Patagonia, Argentina). Hydrobiologia 2001, 455, 79–85. [Google Scholar] [CrossRef]
  65. IDEAM Grupo de Climatología y Agrometereología. Mapas de Precipitación Promedio en Colombia. Available online: http://atlas.ideam.gov.co/visorAtlasClimatologico.html (accessed on 21 January 2023).
  66. Ortiz-Vera, M.P.; Olchanheski, L.R.; da Silva, E.G.; de Lima, F.R.; Rada Martinez, L.R.d.P.; Zanoli Sato, M.I.; Jaffé, R.; Alves, R.; Ichiwaki, S.; Padilla, G.; et al. Influence of water quality on diversity and composition of fungal communities in a tropical river. Sci. Rep. 2018, 8, 1–9. [Google Scholar] [CrossRef] [PubMed]
  67. Olaniran, A.O.; Balgobind, A.; Pillay, B. Bioavailability of heavy metals in soil: Impact on microbial biodegradation of organic compounds and possible improvement strategies. Int. J. Mol. Sci. 2013, 14, 10197–10228. [Google Scholar] [CrossRef] [Green Version]
  68. Shukla, P.K.; Misra, P.; Maurice, N.; Ramteke, P.W. Heavy metal toxicity and possible functional aspects of microbial diversity in heavy metal-contaminated sites. In Microbial Genomics in Sustainable Agroecosystems; Tripathi, V., Kumar, P., Tripathi, P., Kishore, A., Kamle, M., Eds.; Springer Nature: Singapore, 2019; pp. 255–317. ISBN 9789813298606. [Google Scholar]
  69. Varol, M.; Ustaoğlu, F.; Tokatlı, C. Ecological risks and controlling factors of trace elements in sediments of Dam Lakes in the Black Sea region (Turkey). Environ. Res. 2022, 205, 112478. [Google Scholar] [CrossRef]
  70. Wang, Q.; He, X.; Xiong, H.; Chen, Y.; Huang, L. Structure, Mechanism, and Toxicity in Antibiotics Metal Complexation: Recent Advances and Perspectives. Sci. Total Environ. 2022, 848, 1–14. [Google Scholar] [CrossRef]
  71. Dickinson, A.W.; Power, A.; Hansen, M.G.; Brandt, K.K.; Piliposian, G.; Appleby, P.; O’Neill, P.A.; Jones, R.T.; Sierocinski, P.; Koskella, B.; et al. Heavy Metal Pollution and Co-Selection for Antibiotic Resistance: A Microbial Palaeontology Approach. Environ. Int. 2019, 132, 105117. [Google Scholar] [CrossRef]
  72. Gupta, S.K.; Shin, H.; Han, D.; Hur, H.G.; Unno, T. Metagenomic Analysis Reveals the Prevalence and Persistence of Antibiotic- and Heavy Metal-Resistance Genes in Wastewater Treatment Plant. J. Microbiol. 2018, 56, 408–415. [Google Scholar] [CrossRef]
  73. Jia, J.; Guan, Y.; Cheng, M.; Chen, H.; He, J.; Wang, S.; Wang, Z. Occurrence and Distribution of Antibiotics and Antibiotic Resistance Genes in Ba River, China. Sci. Total Environ. 2018, 642, 1136–1144. [Google Scholar] [CrossRef] [PubMed]
  74. Seiler, C.; Berendonk, T.U. Heavy Metal Driven Co-Selection of Antibiotic Resistance in Soil and Water Bodies Impacted by Agriculture and Aquaculture. Front. Microbiol. 2012, 3, 1–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Rajasekar, A.; Qiu, M.; Wang, B.; Murava, R.T.; Norgbey, E. Relationship between Water Quality, Heavy Metals and Antibiotic Resistance Genes among Three Freshwater Lakes. Environ. Monit. Assess. 2022, 194, 64–73. [Google Scholar] [CrossRef]
  76. Buechel, E.R.; Pinkett, H.W. Transcription Factors and ABC Transporters: From Pleiotropic Drug Resistance to Cellular Signaling in Yeast. FEBS Lett. 2020, 594, 3943–3964. [Google Scholar] [CrossRef]
  77. Hamdoun, A.; Hellmich, U.A.; Szakacs, G.; Kuchler, K. The Incredible Diversity of Structures and Functions of ABC Transporters. FEBS Lett. 2021, 595, 671–674. [Google Scholar] [CrossRef] [PubMed]
  78. Bazzicalupo, A.L.; Ruytinx, J.; Ke, Y.H.; Coninx, L.; Colpaert, J.V.; Nguyen, N.H.; Vilgalys, R.; Branco, S. Fungal Heavy Metal Adaptation through Single Nucleotide Polymorphisms and Copy-Number Variation. Mol. Ecol. 2020, 29, 4157–4169. [Google Scholar] [CrossRef] [PubMed]
  79. Mi, L.; Wang, Z.; Yang, W.; Huang, C.; Zhou, B.; Hu, Y.; Liu, S. Cytochromes P450 in Biosensing and Biosynthesis Applications: Recent Progress and Future Perspectives. TrAC Trends Anal. Chem. 2023, 158, 116791. [Google Scholar] [CrossRef]
  80. Dunkel, N.; Blaß, J.; Rogers, P.D.; Morschhäuser, J. Mutations in the Multi-Drug Resistance Regulator MRR1, Followed by Loss of Heterozygosity, Are the Main Cause of MDR1 Overexpression in Fluconazole-Resistant Candida albicans Strains. Mol. Microbiol. 2008, 69, 827–840. [Google Scholar] [CrossRef]
  81. Ramirez-Castrillon, M.; Barona-Colorado, A.; Bados-Lopez, M.C.; Bolaños-Burbano, D. Chapter 7: Diversity of Environmental Yeasts of Colombia: A Systematic Review. In Catalogue of Fungi of Colombia, 1st ed.; de Almeida, R.F., Lücking, R., Vasco-Palacios, A., Gaya, E., Diazgranados, M., Eds.; Royal Botanic Gardens,: Kew, UK; Instituto de Investigación de Recursos Biológicos Alexander von Humboldt: Bogotá, Colombia, 2022; pp. 92–103. [Google Scholar]
Figure 1. Location of the sampling stations. South Channel: rainwater canal; DWTP: drinking water treatment plant “Puerto Mallarino”, Cauca River; WWTP: wastewater treatment plant “PTAR”; and Melendez River. The map on the right was modified from www.4UMaps.com (accessed on 21 October 2022) under the license Data CC BY-SA by OpenStreetMap.
Figure 1. Location of the sampling stations. South Channel: rainwater canal; DWTP: drinking water treatment plant “Puerto Mallarino”, Cauca River; WWTP: wastewater treatment plant “PTAR”; and Melendez River. The map on the right was modified from www.4UMaps.com (accessed on 21 October 2022) under the license Data CC BY-SA by OpenStreetMap.
Ijerph 20 03428 g001
Figure 2. Absolute frequency of yeast species found in different aquatic systems in Cali, Colombia. ND = not determined.
Figure 2. Absolute frequency of yeast species found in different aquatic systems in Cali, Colombia. ND = not determined.
Ijerph 20 03428 g002
Figure 3. Principal component analysis correlation biplots of the interaction of physico-chemical parameters, heavy metal concentrations, yeast levels, and the association of the various sampling points (C = rainwater canal “South Channel”; PM = DWTP “Puerto Mallarino”; P = WWTP “PTAR”).
Figure 3. Principal component analysis correlation biplots of the interaction of physico-chemical parameters, heavy metal concentrations, yeast levels, and the association of the various sampling points (C = rainwater canal “South Channel”; PM = DWTP “Puerto Mallarino”; P = WWTP “PTAR”).
Ijerph 20 03428 g003
Table 1. Minimal inhibitory concentrations (MIC) of yeasts associated with water systems in Cali against Fluconazole (FCZ) and Amphotericin B (AMPB).
Table 1. Minimal inhibitory concentrations (MIC) of yeasts associated with water systems in Cali against Fluconazole (FCZ) and Amphotericin B (AMPB).
StrainSpeciesSusceptibility to FluconazoleFCZ MIC (μg/mL)AMPB MIC (μg/mL)
CS11Rh. paludigenaSensitive22
CS14NDSensitive22
CS15V. humicolaSensitive0.50.25
CS16C. parapsilosisSensitive20.5
CS17D. catenulataSensitive0.51
CS18D. catenulataSensitive0.50.5
CS19C. tropicalisDDS42
CS1AC. glabrataDDS40.5
CS1BRh. mucilaginosaResistant *12
CS21P. fermentansDDS42
CS22NDSensitive0.252
CS23Cr. neoformansSensitive0.1252
CS24ACr. neoformansSensitive21
CS4T. mucoidesSensitive22
CS45P. kluyveriSensitive0.50.5
CS51AD. catenulataResistant82
CS51BC. tropicalisResistant162
CS55C. albicansResistant642
CS7C. tropicalisSensitive12
CS7BG. candidumResistant22
CS7CNDResistant642
CS9Candida sp. 1Sensitive0.251
M2NDSensitive22
M12NDResistant642
P1C. albicansResistant81
P13BC. albicansDDS41
P14C. aaseriSensitive0.50.5
P16AC. tropicalisSensitive22
P16BP. laurentiiSensitive0.52
P20AD. hanseniiDDS42
P22P. kudriavzevii (C. krusei)Resistant *41
P24AD. hanseniiResistant640.5
P24BRh. mucilaginosaResistant642
P3S. cerevisiaeDDS42
P3AHanseniaspora pseudoguilliermondiiSensitive12
P4C. tropicalisResistant640.5
P46Candida sp. 2Sensitive0.252
P9AP. fermentansSensitive22
PM14G. candidumSensitive0.51
PM15Pichia sp.DDS42
PM18T. coremiiformeSensitive20.5
PM19P. fermentansDDS42
PM20Naganishia sp.Sensitive0.1252
PM22C. albicansSensitive0.52
PM24Candida sp. 1DDS42
PM4AT. mucoidesResistant82
PM4BP. laurentiiResistant162
PM54Candida sp. 1Resistant640.5
PM54AC. parapsilosisResistant162
* Due to intrinsic resistance, these strains were marked as resistant, despite the observed MIC. Yeast strains labeled with “CS” were isolated from the South Channel, Navarro; yeast strains labeled with “P” were isolated from the “PTAR” WWTP; yeast strains labeled with “PM” were isolated from the “Puerto Mallarino” DWTP, Cauca River; and yeast strains labeled with “M” were isolated from the Melendez River. ND: not determined. DDS: dose-dependent Sensitive.
Table 2. Correlations between physico-chemical parameters, metal concentrations, and MIC of fluconazole-resistant yeasts.
Table 2. Correlations between physico-chemical parameters, metal concentrations, and MIC of fluconazole-resistant yeasts.
Heavy Metals/
Parameters
CpCtCaPkRhmTmGcDhPlDcCsp1
Fe0.4690.0900.2510.192−0.8160.5430.352−0.1350.4060.8160.333
Ag0.0200.4040.2490.0950.000−0.089−0.0270.456−0.188−0.8160.333
Zn0.5060.2690.4700.095−0.816−0.1340.0270.0260.2950.816−0.333
Cu0.5060.0000.2490.190−0.816−0.1340.188−0.1340.188−0.816−0.666
Cd−0.292−0.3240.8050.0240.8160.2080.000−0.194−0.388−0.8160.912
Cr−0.1790.1730.0000.0790.500−0.198−0.2080.505−0.148−1.000−0.182
Pb−0.0660.1720.3310.2390.0000.1720.1470.000−0.1470.0000.333
Mg0.3030.2240.3590.190−0.816−0.0440.0800.0800.1340.000−0.333
TP0.0600.4490.304−0.0470.000−0.269−0.2410.510−0.134−0.8160.000
Turbidity0.2630.2240.3590.191−0.8160.0000.0800.0800.0800.0000.000
TDS0.183−0.022−0.700−0.024−0.816−0.135−0.027−0.0540.1350.000−0.666
Nitrate−0.060−0.045−0.415−0.4290.816−0.674−0.4030.134−0.188−0.816−1.000
NT0.1010.3140.3040.000−0.8160.000−0.1340.2410.0260.0000.333
pH0.0670.5220.4640.2370.000−0.1240.0290.5340.0290.000−0.235
COND0.6040.0740.4020.342−0.8160.3220.386−0.1780.3260.000−0.235
TEMP0.4690.3720.5260.289−0.8160.0240.2070.1780.1480.000−0.235
Cp: Candida parapsilosis,Ct: C. tropicalis, Ca: C. albicans, Rhm: Rhodotorula mucilaginosa, Pk: Pichia kudriavzevii (C. krusei), Tm: Trichosporon mucoides, Gc: Geotrichum candidum, Dh: Debaryomyces hansenii, Pl: Papiliotrema laurentii, Dc: Diutina catenulata, Csp1: Candida sp. 1. TP = total phosphorus, TDS = total dissolved solids, NH4 = nitrate, TN = total nitrogen, COND = conductivity, TEMP = temperature. The values in bold present a statistically significant (p < 0.05) positive or negative correlation between variables.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Caicedo-Bejarano, L.D.; Osorio-Vanegas, L.S.; Ramírez-Castrillón, M.; Castillo, J.E.; Martínez-Garay, C.A.; Chávez-Vivas, M. Water Quality, Heavy Metals, and Antifungal Susceptibility to Fluconazole of Yeasts from Water Systems. Int. J. Environ. Res. Public Health 2023, 20, 3428. https://doi.org/10.3390/ijerph20043428

AMA Style

Caicedo-Bejarano LD, Osorio-Vanegas LS, Ramírez-Castrillón M, Castillo JE, Martínez-Garay CA, Chávez-Vivas M. Water Quality, Heavy Metals, and Antifungal Susceptibility to Fluconazole of Yeasts from Water Systems. International Journal of Environmental Research and Public Health. 2023; 20(4):3428. https://doi.org/10.3390/ijerph20043428

Chicago/Turabian Style

Caicedo-Bejarano, Luz Dary, Lizeth Stefania Osorio-Vanegas, Mauricio Ramírez-Castrillón, Jorge Enrique Castillo, Carlos Andrés Martínez-Garay, and Mónica Chávez-Vivas. 2023. "Water Quality, Heavy Metals, and Antifungal Susceptibility to Fluconazole of Yeasts from Water Systems" International Journal of Environmental Research and Public Health 20, no. 4: 3428. https://doi.org/10.3390/ijerph20043428

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