Next Article in Journal
A New Cross-Flow Type Turbine for Ultra-Low Head in Streams and Channels
Next Article in Special Issue
Assessment of Existing Fate and Transport Models for Predicting Antibiotic Degradation and Transport in the Aquatic Environment: A Review
Previous Article in Journal
A Case Study: Groundwater Level Forecasting of the Gyorae Area in Actual Practice on Jeju Island Using Deep-Learning Technique
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Impact of Antibiotic Pollution on the Bacterial Population within Surface Water with Special Focus on Mountain Rivers

by
Klaudia Kulik
1,
Anna Lenart-Boroń
1,* and
Kinga Wyrzykowska
2
1
Department of Microbiology and Biomonitoring, Faculty of Agriculture and Economics, University of Agriculture in Kraków, Adam Mickiewicz Ave. 24/28, 30-059 Krakow, Poland
2
Diagnostyka S.A. Medical Microbiological Laboratory, Na Skarpie 66, 31-913 Krakow, Poland
*
Author to whom correspondence should be addressed.
Water 2023, 15(5), 975; https://doi.org/10.3390/w15050975
Submission received: 30 January 2023 / Revised: 23 February 2023 / Accepted: 1 March 2023 / Published: 3 March 2023

Abstract

:
Environmental aquatic pollution with antibiotics is a global challenge that affects even pristine mountain environments. Monitoring the concentration of antibiotics in water is critical to water resource management. In this review, we present the sources and degradation routes of antibiotics polluting surface waters, with particular focus on mountain environments and pristine areas. This pollution is strongly related to anthropopressure resulting from intensive tourism. An important aspect of the threat to the environment is water containing antibiotics at sub-inhibitory concentrations, which affects bacterial populations. Antibiotics are ecological factors driving microbial evolution by changing the bacterial community composition, inhibiting or promoting their ecological functions, and enriching and maintaining drug resistance. We paid attention to the stability of antibiotics and their half-lives in water related to biotic and abiotic degradation, which results from the structures of molecules and environmental conditions. Wastewater treatment combined with advanced treatment techniques significantly increase the efficiency of antibiotic removal from wastewater. Modern methods of wastewater treatment are crucial in reducing the supply of antibiotics to aquatic environments and enhancing the possibility of economic and safe reuse of wastewater for technical purposes. We provide a perspective on current research investigating antibiotic emergence in mountain areas and identify knowledge gaps in this field.

1. Introduction

Antibiotics are bactericidal and bacteriostatic agents used to treat bacterial infections, thereby providing a solution in the treatment of many diseases. Antimicrobial substances are also used for non-medical purposes, such as livestock, poultry, and fish growth stimulation. Remarkably, more antibiotics are used in the US for animal growth promotion than in human medicine [1]. Approximately 24.6 million pounds of antibiotics are used each year in livestock farming [2]. Animal husbandry utilizes more antibiotics than human therapeutic applications of these drugs. Importantly, the increased consumption of antibiotics has led to the exposure of aquatic ecosystems to contamination with these substances. This is because most antibiotics are only partially metabolized by the target organism; therefore, their residues (30–90% of ingested antibiotic doses) are excreted in urine and feces to reach wastewater treatment plants (WWTPs) [3]. Wastewater contaminated with antibiotics undergoes treatment in treatment plants but complete removal of these compounds is impossible in conventional systems [4]. WWTPs are designed to reduce the pollutant load in the majority of urban and rural wastewater, but they are not effective in reducing the loads of antibiotics and antibiotic resistance genes (ARGs) [5]. Furthermore, WWTP effluents are discharged into surface water and the sludge can be used as manure fertilizer. Antibiotics are a problem around the world due to their frequent widespread use in large amounts, and they often become an inappropriate therapeutic pathway. Antibiotics are regarded as “pseudo-persistent” contaminants because of their continuous introduction into ecosystems—their entry rate into the environment is higher than their rate of elimination [4].
The presence of antibiotic residues in surface waters has been reported to influence the compositions and functions of microbial communities. Antibiotics dissolved in surface waters reach sub-inhibitory concentrations that affect the microbial ecology by increasing the mutation rate, causing horizontal drug resistance gene transfer, and driving the selection of antibiotic resistant bacteria [6]. Antibiotic use in human medicine, agriculture, aquaculture, and veterinary medicine puts a huge selective pressure on microbial communities [6]. If the selective pressure is high enough, acquisition of resistance becomes necessary for bacterial survival. Long-term exposure to sub-inhibitory concentrations of antibiotics in the aquatic environment could be the main factor responsible for the generation of drug resistance and transference of drug resistance genes [7,8].
Antimicrobial resistance determinants are commonly found in wastewater [5], thus posing a threat to the environment resulting from the possibility of drug resistance spread. Overuse and misuse of antibiotics contribute to the evolution of antibiotic-resistant bacteria (ARB) and antibiotic resistance genes (ARGs). The pollution of rivers by ARB and ARGs increases the risk of antibiotic resistance transmission from the environmental resistome to humans [9] and also affects the ecological balance of the aquatic environment [10]. Antibiotic resistance genes are recognized as emerging micropollutants with environmental persistence. Antibiotic resistance genes are present in both extracellular (eARGs) and intracellular (iARGs) forms in the environment [11]. However, eARGs play a crucial role in the spread of antibiotic resistance in the environment thanks to their ability of being absorbed on soil and sediment particles, thus persisting longer in water environments [12]. Horizontal gene transfer (HGT) enables the transport of ARGs by mobile genetic elements between bacterial cells. It allows dissemination of ARGs from commensal and environmental bacterial species to pathogenic species [13]. Water is a major pathway of dissemination of bacteria and ARGs between different environmental compartments. Human exposure could occur through contaminated recreational areas as well drinking and recreational water [14]. The problem of surface water contamination is very common [15,16,17,18], even in river waters in natural reserves that are expected to be clean [19].
The next-generation sequencing (NGS) technique has recently become the most frequently applied method for examining microbial community structure in water environments [20,21,22,23]. The total bacterial community plays an important role in aquatic ecosystems [23]. The NGS method allows researchers to determine the presence of all species of bacteria present in the studied environment, even non-cultivated ones. Therefore, the NGS method could become a useful tool for water quality monitoring purposes, including the development of bioindicators for sewage pollution and microbial source tracking [21].
Contamination of the aquatic environment with antibiotics is of interest to scientists, as evidenced by numerous scientific publications presenting the results of research investigating the concentrations of antibiotics in waters, their impact on the environment, and the risks resulting from this phenomenon. It should be noted that mountain regions are the sources of river systems that provide fresh water to more than half of the human population [24]. Human wellbeing depends on mountain resources because they provide clean water and harbor rich biodiversity. Water from mountain rivers is exploited for many purposes, including drinking water sources, irrigation, recreation, and snowmaking on ski slopes, which is common in these areas to ensure the operation of ski stations. In many countries, due to highly developed tourist infrastructures and accommodation facilities, mountain areas are characterized by a large amount of generated sewage. This is due to the growing number of residents in peak tourist seasons, which results in significant overload and decreased efficiency of local treatment plants [25]. The above described key function of mountain river systems may be disturbed by the impact of human pressure in this environment. Surface water contamination in mountain areas, particularly protected ones, and most importantly the contamination of mountain river waters, is one of the greatest issues of water management. The reason for this is because mountain river waters are particularly sensitive to both natural and anthropogenic changes due to their cleanliness.
For the above mentioned reasons, the aim of this review was to analyze the sources and stability of antibiotics in water, including the rivers of protected mountain areas. The review presents the anthropogenic pressure in mountain regions, which is mainly related to tourism, that contributes to the contamination of water systems with antibiotics. This review also presents the impact of this pollution on the composition of microbial populations and the development of drug resistance. Most importantly, this paper summarizes and compiles the most up-to-date research concerning issues related to the antibiotic contamination of surface water, with particular focus on mountain and pristine areas, and identifies knowledge gaps and ways to fill them within the discussed scope.
In this regard, we present a systematic review of the latest literature on antibiotic contamination of the aquatic environment, with special emphasis on mountain areas. The search strategy to prepare this research review included browsing through the following databases: Scopus, Science Direct, PubMed, Tylor, and Google Scholar. Descriptors such as antibiotics, antibiotic agents, drug-resistant bacteria, bacterial community structure, mountain water pollution, alpine rivers, mountain springs, pristine environment, tourism traffic, and their combinations were used to search for information on mountain water pollution with antibiotics. All science databases were searched for papers published in the English language and were not restricted to any specific region in order to find studies conducted worldwide. The period of time that was set for searching articles covered the years 2003–2023. We also used two articles published in 1985 and 1999 that dealt with the level of excretion of antibiotics by humans and animals. After the application of the above search strategy, very few studies (7 articles) were obtained that examined the occurrence of antibiotics precisely in mountain rivers. This indicates a substantial knowledge gap that needs to be filled or is being investigated with respect to studies focused on the mountain environment. A significant number of studies present the results on surface water pollution in urban areas, often resulting from the presence of sewage treatment plants and the resulting consequences for microorganism populations and humans. Considering the above, a literature review was created using the available information on the contamination of mountain waters with antibiotics in relation to the general contamination of surface waters by these substances.

2. Sources of Antibiotics in Mountain Rivers

The quality of water in mountain regions is shaped by many natural and anthropogenic factors. In the most pristine regions, including national parks and natural valuable protected areas, the natural factors include variable weather conditions, surface runoff, soil leaching, and snowmelt water [26,27]. Along the course of rivers, water pollution increases, which is strongly related to the influence of anthropopressure-related factors, including illegal discharge of sewage from households (human and animals feces), surface runoff carrying natural fertilizers from agricultural fields, and wastewater inflow from WWTPs increased by tourist traffic-related sewage inflow [26,28].
In many cases, anthropogenic pressure in mountain regions is related to tourist traffic contributing to increased wastewater inflow, thus contributing to the presence of antibiotics and bacterial contaminants in mountain rivers [25]. Constantly developing winter tourism can pollute the environment and affect the ecology of microorganisms in mountain aquatic ecosystems in various ways. Mountain hiking is one of the causes of the pollution of rivers with antimicrobial agents. Mountain areas attract tourists due to their unspoiled nature, hiking trails for mountain trekking, and areas for active recreation. One of the reasons for the growing number of tourists is the constant development of winter sports centers with snow-covered ski slopes, which also offer activities outside the winter season, such as downhill skiing, as well as facilities offering thermal pools with geothermal water (thermal spas). Excluding the Alpine countries and the United States, Poland is at the forefront of the development of ski infrastructure [29]. In the winter season of 2014/2015, one of the several ski stations located in the Białka river valley (Podhale, Southern Poland) was visited by approximately 318,000–342,000 people [30]. The same ski resort offered 6500 beds in 2012, while this number increased to 13,200 beds in 2020. Moreover, the transport capacity of ski lifts in 2020 was 32,975 people per hour. At the same time, the number of inhabitants in this mountain village was 2249 people as of December 2019 [31]. The presented numbers of tourists compared to the number of inhabitants for only one mountain village in Poland show a possible negative impact of tourism on the pollution of mountain rivers. Ski resorts exert a significant pressure on adjacent natural areas, which is often more important than the impact of general tourist activity located further from the ski resorts. Most importantly, the greatest negative pressure is exerted on water quality [32]. Samples of snow from the Sudety Mountains in Poland, collected in places with different levels of anthropogenic pressure, indicated that the presence of humans could affect the composition of the microbial resistome in snow [33]. In Central Europe, the most attractive countries in terms of winter sports and mountain hiking are Austria, Slovakia, the Czech Republic, and Poland [34]. Ski resort owners in all of these countries must comply with the legal aspects of environmental protection. Many ski resorts and mountain hiking areas in these countries are located on territories included in Natura 2000, the international network of protected areas covering Europe’s most valuable and threatened habitats [35,36]. There is a conflict between the constant development of mountain tourism and keeping the environment and rivers unchanged [37]. The Tatra Mountains, located on the border of Poland and Slovakia, are among the most frequently visited regions in these countries. The Tatra National Park (TNP) offers 275 km of hiking trails, 6 bicycle routes, and 8 tourist shelters with the availability of gastronomy and accommodation. The number of tourists visiting the TNP is growing year by year, from 2,926,012 tourists in 2014 to 4,600,025 tourists in 2021. On average, the TNP is visited by 3 million tourists each year [38]. Such great interest in mountain tourism is a burden on the environment. Mountain shelters located in the TNP are not connected to the sewerage system. They only have biological treatment plants or septic tanks, from which waste must be transported to treatment plants outside the TNP. According to an interview with TNP authorities (February 2020) [39], approximately 30 cubic meters of wastewater are generated per day in the mountain shelter on Morskie Oko Lake (one of the most recognizable of the Tatra lakes), which is visited by approximately 10,000 people per day in peak season. Such large amounts of wastewater produced in the national park have a significant impact on the pollution of mountain streams, with antibiotics and drug resistance determinants flowing down to the rivers from which water is drawn for a number of purposes, including snowmaking or irrigation of green areas. Lenart-Boroń et al. [40] examined the occurrence of antibiotics in the groundwater of pristine areas in the Tatra National Park and in waters of one of the mountain rivers. None of the tested antibiotics were detected in one of the groundwater sampling sites in the TNP area (located at 1600 m a.s.l.), but the second groundwater sampling site, still located in the TNP but much closer to households and located below mountain shelters, was contaminated with antibiotics at the following mean concentrations: erythromycin (0.89 ng/L), ofloxacin (0.27 ng/L), clindamycin (0.36 ng/L), vancomycin (2.99 ng/L), trimethoprim (0.29 ng/L), and sulfamethoxazole (0.20 ng/L). The total concentration of antibiotics at one groundwater site was 27.92 ng/L. River water in the TNP pristine area contained a total of 34.02 ng/L of antibiotics, including erythromycin, oxytetracycline, clindamycin, and vancomycin. The above-described considerable intensity of mountain tourism influences changes in the natural environment in a number of ways. Among them, increased water consumption and the production of wastewater are the main problems [24,28,32,41,42]. Antibiotics are continuously discharged into wastewater after their metabolism in target organisms via excretion in urine and feces from human and animal bodies. Importantly, most antimicrobials are not completely metabolized (the range of metabolism is 10–90%, Table 1) [43]; therefore, both the parent compounds and degradation products reach wastewater treatment plants. In addition, some metabolites, such as acetic conjugates of sulphonamides, can revert back to their parent compounds during manure storage [44].
The metabolism rate of these compounds varies for humans and animals depending on the antibiotic class. These pharmaceuticals may reach aquatic environments by effluents from WWTPs to rivers and groundwater, as well as leachate from unsealed sewage systems and manure and/or sewage storage tanks [52]. During the wastewater treatment process, bacteria are continuously mixed with sub-inhibitory concentrations of antibiotics, which creates suitable conditions for the development of drug resistance, and then released with antibiotic residues into water environments [53]. The removal efficiency of antibiotics and antibiotic resistance determinants from wastewater by conventional WWTPs is insufficient. WWTPs are not specifically designed to completely reduce levels of antibiotics and ARGs. Removal of different antibiotics occurs in different steps of the wastewater treatment process, and the effectiveness of their removal varies among antibiotics [5]. Removal efficiency depends on the physical and chemical properties of antibiotics and on the treatment process conditions. Blair et al. (2015) [54] evaluated the maximum concentrations of pharmaceuticals within conventional activated sludge treatment processes. The results showed the presence of nine antibiotics: ampicillin (160 ng   L 1 ), ciprofloxacin (2200 ng   L 1 ), enrofloxacin (34 ng   L 1 ), norfloxacin (140 ng   L 1 ), ofloxacin (2100 ng   L 1 ), penicillin G (30 ng   L 1 ), penicillin V (86 ng   L 1 ), sulfamethoxazole (7400 ng   L 1 ), and trimethoprim (570 ng   L 1 ). The removal efficiency for all antibiotics except ampicillin was negative in this research. WWTPs have been recognized as hotspots (as the main sources of antibiotics) for the release of antibiotics into the aquatic environment. In order to remove organic pollution from wastewater, activated sludge systems are sometimes combined with chemical additions [55]. However, previous studies have concluded that conventional secondary treatment processes are unable to completely remove antibiotics from wastewater [56,57]. Advanced wastewater treatment techniques, such as membrane processes, activated carbon adsorption, and UV radiation, may increase the percentage of antibiotic removal from wastewater and are presented later in this review. WWTPs have emerged as significant sources of antibiotics to mountain aquatic environments due to the contamination of WWTP influent by antibiotics from human and animal excreta, as well as improper disposal of antibiotics and agriculture runoff. The risk of contamination of surface water with antibiotics, drug resistance genes, and drug-resistant bacteria also results from the utilization of sewage sludge in agriculture as an organic fertilizer [58]. Sewage sludge, contaminated with antibiotics and drug resistance determinants, flows from the soil into rivers along with surface runoff. The most common antibiotics found in sewage sludge are fluoroquinolones, sulfonamides, and tetracyclines, the concentrations of which were measured at μg kg−1 [59,60,61,62].
Another significant source of antimicrobials in rivers, including mountain rivers, is their use in animal husbandry for therapeutic and preventive purposes. Mountain agriculture is dominated by livestock production based on grazing. Veterinary antibiotic usage is related to the treatment of infective diseases in animals. The use of antimicrobial agents in animal husbandry ensures the welfare and health of the animals. However, the use of antibiotics is extended to the whole livestock flock in order to limit pathogen spread, thus uninfected animals also take doses of antibiotics [63]. This is referred to as metaphylaxis—short-term antibiotic treatment of animal groups without disease symptoms that had contact with infected animals [64]. This action involves observation of a livestock flock and administration of high doses of antibiotics before clinical symptoms occur in order to counteract the effects of infection. In contrast, antibiotics can also be used for disease prevention (prophylaxis). This includes antibiotic administration in water and food for farm animals in low doses for longer periods of time. During this period, the risk of infection still exists [65]. Metaphylaxis and prophylaxis are common practices in livestock and poultry production to prevent whole livestock mortality and minimize losses, but they have boosted antibiotic consumption. From an epidemiological point of view, the preventive administration of antibiotics increases the risk of drug-resistant bacteria development in the livestock herd and significantly influences the contamination of the environment with antibiotics and drug resistance determinants. In addition, selection for antibiotic-resistant strains can be widespread in the environment via animal feces, thereby enhancing environmental drug resistance [66]. Residues of antibiotics and ARB are usually found in livestock and poultry manure and in waste from livestock companies, resulting in persistent environmental pollution [67]. Animal manure studies have proven the presence of various classes of antibiotics excreted in feces, for instance: enrofloxacin in broiler chicken feces (74% of orally applied enrofloxacin was excreted as the parent compound) [68], oxytetracyclines present in dairy cow feces (20% of injected oxytetracycline was detected in manure samples) [69], and sulfonamides in pig excreta (excretions of four sulfonamides reached 36–87%) [70]. Stored animal manure often reaches soil and surface water with runoff water after rain or due to leaks in manure tanks. Livestock manure is also used as a fertilizer to enrich the soil before growing crops. Mountain areas, in addition to their environmental and cultural functions, also have an agricultural function as they have abundant arable fields, meadows, and pastures. In sustainable and organic farming, the use of manure as a source of organic matter to improve soil quality is a common practice. However, manure is also a source of antibiotic residues, which can adsorb on soil particles, enter plant tissues, and end up in the food chain. There is a risk of enhanced antimicrobial resistance as a result of consumption of vegetables grown on manure [71,72]. Manure widely applied to agricultural lands as fertilizer has enriched the abundance of some ARGs (ermA, ermB, blaOXA-1, qnrS, and oqxA) in agricultural soil [73]. Antibiotic residues and drug resistance determinants in soil fertilized with manure enter rivers with surface runoff, thus polluting the aquatic environment. Active forms of antibiotics occurring in manure can act as a selective pressure and contribute to dissemination of antimicrobial resistance. Livestock animals are a constant link in the spread of ARGs and antibiotics in the aquatic environment because they are continuously exposed to large amounts of antibiotics. Livestock farming can be one of the main sources of antibiotics in rivers due to the excretion of incompletely metabolized antibiotics in animal feces and their further dissemination into the environment [67].
Mountain rivers provide water for the production of artificial snow to ensure snow cover on ski slopes in the winter season and for irrigation of green areas in the summer season. Mountain river water is also used by households to irrigate their crops. The use of water contaminated with antibiotics, drug resistance genes, or antibiotic-resistant bacteria results in further transmission of these micropollutants into the environment, thereby increasing the risk of spreading drug resistance and endangering public health (Figure 1).
Research on the microbiological pollution of the two largest rivers of Podhale (Tatra Mountains, Southern Poland), which are the source of water for snow production on ski slopes, conducted over two ‘high ski seasons’ demonstrated the bacterial pollution of both the river and fresh artificial snow produced from these waters. Bacterial indicators of fecal contamination (coliforms, E. coli, and E. faecalis) were observed in two out of seven samples of artificial snow produced from polluted water. Furthermore, E. coli strains presented the ESBL (extended-spectrum beta-lactamases) resistance mechanism and contained the blaTEM gene in their genome, as confirmed by the PCR method [74]. These observations indicated that the antibiotic-resistant bacterial strains were able to survive the production of artificial snow and that users of the ski slopes were exposed to direct contact with drug-resistant bacteria. Sanchez-Cid et al. [33] evaluated the bacterial community composition and microbial resistome in natural snow samples from mountain tourist attractions and forest areas in the Sudety Mountains, Poland. ARG reads from metagenomic sequencing of snow-derived DNA using the MiSeq System (Illumina, San Diego, CA, USA) were grouped by antibiotic class and their average abundance was compared between the studied sites. The sites with larger surrounding forest areas showed a higher ARG abundance. Several genes showed higher abundance in samples from paths with the highest human activity. These genes determined the resistance to: aminoglycosides (aadA17 gene), tetracycline (tetX gene), rifamycin (rphB gene), fosfomycin (fosA5 gene), beta-lactams (rm3 and LRA-13 genes), and the multidrug resistance gene meI [33]. The results suggested that anthropogenic activities could have a direct impact on the composition of the antibiotic resistome in snow. Furthermore, Segawa et al. [75] observed the prevalence of antibiotic resistance genes in glacier environments (snow and ice samples). ARGs, of both clinical (aac (3), blaIMP) and agricultural (strA and tetW) origin, were detected. These researchers indicated that ARGs in such pristine environments can be transferred by airborne bacteria and migratory birds. Research shows that determinants of drug resistance in the form of ARGs and antibiotics can be stable in the water environment and penetrate into other environmental compartments, such as snow or ice. In Yang and Carlson’s [76] study of river antibiotic contamination in Colorado, the only site in which no antibiotics were detected was a pristine mountain site upstream of urban or agricultural areas. Conversely, all five monitored tetracyclines were detected at a site that had undergone both urban and agricultural impacts at concentrations ranging from 0.08 to 0.30 mg/L.
Another essential route of further transmission of antibiotics in the mountain environment is the irrigation of fields with antibiotic-contaminated water [77]. Irrigating crops and green areas with antibiotic-contaminated water leads to crop contamination and dissemination of drug resistance genes [78]. Antibiotics from irrigation water can accumulate in the edible parts of plants or grasses on which livestock feed. Plants irrigated with antibiotic-contaminated water increase the threat of adaptive resistance selection of the gut microbiome. The amounts of antibiotics found in the environment are considered as trace contaminants, nevertheless, they have a very significant impact on the environment [25]. Although the concentrations of antibiotic residues in water environments range from ng/L to μg/L [79], the continuous discharge and persistence of these contaminants at sub-inhibitory concentrations may cause changes in bacterial communities and stimulate the development of drug resistance. The transference of drug resistance genes from environmental bacterial strains to human pathogens is a major threat to public health. Water-polluting antibiotics cause the development of antimicrobial resistance among microorganisms, hence their presence in the environment is of critical importance to public health.

3. Stability of Antimicrobial Agents in Water Environments

Antibiotic degradation rates are important for predicting their environmental exposure and impact on bacterial populations. Antibiotics dissolved in water undergo physicochemical modifications caused by biotic and abiotic factors that affect their structural stability. The following processes affect the stability of antibiotics in surface water: hydrolysis, photolysis, sorption, and biological degradation [80,81,82,83,84]. The occurrence of these processes depends on environmental conditions such as sunlight, water temperature, the abundance of microorganisms, water chemical composition, sediment properties, and organic matter content. Predicting the degradation pathways of antibiotics is essential for assessing their fate in the environment. The natural degradation pathways of antibiotics in water environments are presented in Figure 2.
An important factor affecting environmental contamination with parent antibiotics is their duration in water while they still have bactericidal properties before they decompose into transformation products. The half-life of an antibiotic in water is an important factor in determining its concentration in an aquatic sample. The literature data showing half-lives of parent antibiotic compounds in aquatic environments are summarized in Table 2.
Photocatalytic degradation and hydrolysis are two of the main abiotic pathways of antibiotic degradation in aquatic environments [80,81,82]. The degradation of antibiotics in water depends on the pH value and temperature, which are the most important parameters affecting hydrolysis rates. These rates typically increase when the temperature increases. Additionally, aqueous compounds such as metals and organic matter can catalyze the hydrolysis reaction. Hydrolysis is the main degradation pathway in aquatic environments without abundant microbial populations, such as rivers and streams, but biodegradation pathways are predominant in wastewater where microbial populations are much more abundant than in surface waters [80]. Fang et al. [82] reported that the pH value had a significant effect on the elimination of enrofloxacin in aquaculture water. Among the tested pH values of 5, 7, and 9, the enrofloxacin removal rate was highest at pH 5 and 25 °C. The half-life for enrofloxacin in water ranged from 3.34 to 6.75 days. In addition, the main degradation product of enrofloxacin is ciprofloxacin, an antibiotic from the fluoroquinolone class [90,92]. On the other hand, Mitchell et al. [80] determined that the hydrolysis of β-lactam antibiotics, ampicillin, cephalothin, and cefoxitin, was most effective at pH 9 and 25 °C, while antibiotic hydrolysis was independent of pH within pH range 4 to 8. The half-lives for cephalothin, cefoxitin, and ampicillin under these conditions were 1.4, 6.6, and 6.7 days, respectively. The authors also established that the hydrolysis rate constants were approximately 20 times higher at 50 °C than at 25 °C. These results confirmed that hydrolysis rates were highly temperature dependent. A study on oxytetracycline hydrolysis showed 72.7% degradation at pH 6.91 after 6 days of incubation at 25 °C. On the other hand, oxytetracycline incubation at pH 3.09 at the same temperature resulted in the hydrolysis of 10.5% of oxytetracycline [81]. The presented research results showed that the rate of hydrolysis depends on the chemical structure within various classes of antibiotics, temperature, and pH value. Antibiotics have to be stable under the acidic conditions in patients’ stomachs. Aquatic environmental conditions are around pH 7 at a temperature below 24 °C. Under such conditions, antibiotics can undergo hydrolysis. β-lactam antibiotics are subject to the hydrolytic cleavage of the β-lactam ring, especially under alkaline conditions [83]. Hydrolysis of the β-lactam ring in the antibiotic molecule causes the loss of bactericidal properties. Research results indicated that the predicted β-lactam antibiotic hydrolysis under ambient pH and temperature conditions and their degradation occurred within a few weeks in most surface waters [80] or even a few hours in ultrapure water [83]. Importantly, antibiotics are continuously released into water systems, which could result in their constant occurrence in the aquatic environment.
The second essential pathway for the degradation of antibiotics in water is photolysis. Degradation of antibiotics can occur through direct photolysis, which is caused by direct absorption of solar light, or indirect photolysis, which involves natural photosensitizers like nitrate and humic acid suspended in water. Under solar radiation, these constituents can generate excited compounds such as hydroxyl radicals and singlet oxygen [88,93]. In addition, organic matter dissolved in water is characterized by high mobility and can promote the solubility of organic pollutants in surface water [88]. Photolysis rates vary along with season, time of day, and water depth [87]. The results obtained by Wei et al. [87] in their study on the photo-transformation of tetracycline and sulfonamide antibiotics in surface water indicated that the decomposition of antibiotics could be dominated by direct photo transformation in summer. However, the half-lives of sulfonamide antibiotics could reach 1 month in the winter season. Sulfonamides are very sensitive to seasonal variations of UVB solar radiation. In winter, sulfonamide photo-transformation is replaced by sedimentation, adsorption, and biodegradation [87]. On the other hand, tetracycline antibiotics are unstable upon exposure to natural sunlight because the main absorption peak of these antibiotics (i.e., 365 nm) is well overlapped with the solar spectrum in the range of 290 to 420 nm [87]. Therefore, photolysis is the main pathway of tetracycline degradation in surface water. Direct photolytic degradation of antibiotics occurs mainly in the upper layer of surface water [85]. Aminopenicillins (amoxicillin, aminopenicillin) degrade faster under simulated sunlight conditions than penicillins (piperacillin, penicillin) [85,94]. Photolytic degradation of antibiotics in river water is much slower compared to that in ultrapure water due to organic matter absorbing the radiation and water turbidity [85]. Jiang et al. [83] determined that the half-lives of four cephalosporins (cefradine, cefuroxime, ceftriaxone, and cefepime) in surface water in the dark ranged from 2.7 to 18.7 days, while the half-lives of the cephalosporins decreased significantly to 2.2–5.0 days under exposure to simulated sunlight. The researchers concluded that direct photolysis was the primary process involved in the degradation of cephalosporins in the surface water of a lake. Photolytic degradation of antibiotics dissolved in water is particularly intense in summer when the intensity and exposure of sunlight is the strongest of the whole year. Therefore, due to low temperatures and less sunlight exposure, antibiotics are more stable in the water of mountain rivers in winter. Because of the extended stability of antibiotics in rivers, these compounds can enter other environmental compartments along with river water. One of such pathways is the production of artificial snow from river water in order to provide snow on ski slopes.
Antibiotics can also be retained in the aquatic environment by sorption with organic matter such as humic substances and organic carbon. Hydrogen bonds stabilize antibiotics on the surface of organic molecules [95]. Sorption of antimicrobial agents to the mineral components of the river sediment might protect these compounds against microbial degradation and thus prolong their half-lives in water [83]. For instance, oxolinic acid (quinolone group antibiotic) is very stable in the aquatic environment (9 days in water and 48 to 300 days in sediment) and its long persistence in water is related to adsorption onto sediments [90]. Antibiotics with high adsorption coefficients may undergo repeated adsorption and desorption in the aquatic environment. An even greater ecological threat to the environment is the deposition of antibiotics in river water adsorbed on solid particles, such as micropollutants, which include microplastics occurring in the aquatic environment [96,97]. Antibiotics and microplastics are two classes of emerging pollutants with negative impact to the aquatic environment. Microplastics have different adsorption capacities for organic pollutants, including antibiotics, due to different surface characteristics, pore size distributions, and various degrees of crystallization [96]. Adsorption of antibiotics on microplastics could result in their long-range transport and increase their exposure to aquatic environments. The main sources of both antibiotics and microplastics in the aquatic environment are wastewater treatment plants, the effluents of which are point sources of these micropollutants in rivers [98]. Li et al.’s (2018) results indicated that polyamide particles have high adsorption for amoxicillin, tetracycline, and ciprofloxacin because of their highly developed pore structure [96]. Therefore, polyamide micropollutants can serve as carriers of antibiotics in the aquatic environment [96]. According to Wang et al. [97], increased water salinity could reduce the adsorption of antibiotics onto microplastics. Microplastics can concentrate more antibiotics and ARGs in fresh river water than in seawater. In addition, the amount of sulfamerazine, chloramphenicol, and tylosin adsorbed on polyethylene microplastics in river water was twice that in sea water [97]. Adsorption of antimicrobial compounds with slow-sinking organic particles enables the spread of these substances across long distances in the aquatic environment and protects these compounds against rapid degradation.
Antibiotics found in surface waters can be transformed by microorganisms through biological degradation pathways. It is possible to distinguish microorganisms capable of modifying (biotransformation), cleaving (biodegradation), or mineralizing (subsistence) antibiotics [84]. Some bacterial strains, called antibiotrophs, have the potential to use antibiotics as sources of carbon and energy and survive as antibiotic-resistant strains in environments with antibiotics as the sole carbon source [99]. Moreover, these strains often possess transmissibility of virulence and pathogenicity. Cha and Carlson [100], in their research on the biodegradation of veterinary antibiotics in lagoon waters, established that antibiotic biodegradation rates were faster under aerobic conditions. Additionally, biodegradation depended on ambient temperature, with elevated temperature (20 °C) increasing the rate of decomposition and lower temperature (4 °C) reducing the biodegradation rate. The lower efficiency of antibiotic biodegradation under anaerobic conditions was also observed by Jiang et al. [83] in their study on the biodegradation of cephalosporins in lake water. Additionally, these researchers established that biodegradation played a minimal role in cephalosporin decomposition, regardless of a diverse bacterial community. Yang et al. [101] revealed the roles of 24 bacterial genera in a microbial community involved in the aerobic and anaerobic degradation of amoxicillin, tetracyclines, and sulfonamides in wastewater sludge. Pseudomonas sp. strains present the greatest aerobic degradation capability. On the other hand, bacterial strains of Bacillus sp. and Clostridium sp. show the greatest anaerobic antibiotic-degradation capability. Microbial degradation shows promising prospects in the removal of sulfamethoxazole. Ammonia-oxidizing bacteria and sulfate-reducing bacteria present good removal capacity of sulfamethoxazole. However, low concentration of sulfamethoxazole could be insufficiently bioavailable for environmental bacterial strains [102]. Environmental conditions, such as temperature, pH, antibiotic concentration, and additional carbon sources, can affect the degradation of these micropollutants. Biodegradation pathways of antibiotics prevail in environments with abundant and diverse microbial populations e.g., in wastewater [80]. In river waters with lower abundances of microorganisms, the biodegradation of antibiotics occurs at a lower level. However, biodegradation may predominate when hydrolysis and photolysis do not show much intensity, which depends on environmental conditions such as temperature, pH value, and the availability of sunlight.

4. Effect of Sub-Inhibitory Concentrations of Antibiotics on Bacterial Populations

Antibiotics entering aqueous environments as a result of anthropopressure could potentially affect the communities of microorganisms. They can be regarded as an ecological factor driving microbial evolution by changing the structures of microbial communities, inhibiting or promoting their ecological functions, and affecting drug resistance mechanisms [103]. The impact of antibiotics on the aquatic ecosystem is related to their concentrations, bioavailability, exposure time, and the addition of substrates, e.g., metals [103]. Antibiotic-induced changes in the ecological functions of the aquatic environment include the nitrogen transformation process, e.g., oxytetracycline inhibits the nitrification process in surface water [104]; however, in some cases, increased nitrification activity has been observed when bacteria are exposed to antibiotics [105]. Moreover, Fountoulakis et al. [106] reported that antibiotics could inhibit the methanogenesis process. In their research, sulfamethoxazole and ofloxacin mildly inhibited the anaerobic digestion process of methanogens. In other studies, Córdova-Kreylos and Scow [107], based on phospholipid fatty acid analysis, discovered that the broad-spectrum antibiotic ciprofloxacin favored the presence of sulfate-reducing bacteria and Gram-negative bacteria, such as Desulfovibrio, Desulfobulbus, and Desulfobacter, while reducing the number of Gram-positive bacteria in salt marsh sediment. In this research, ciprofloxacin was capable of modifying the bacterial community structure at concentrations as low as 20 μg mL 1 in anaerobic sediments. Importantly, despite the fact that the sorption of antibiotics on sediments is estimated at 80–90%, studies have shown that antibiotics at low bioavailability are still capable of modifying the microbial community.
Antibiotics polluting the natural environment do not reach the high therapeutic concentrations that inhibit the growth of bacteria (∼1 mg   mL 1 ) [108]. However, they are widely distributed at low concentrations ( ng μ g   L 1 ) [52,109] without reaching the minimum inhibitory concentration (MIC). The level of an antibiotic that is below the MIC concentration is referred to as sub-MIC (sub-minimum inhibitory concentration), or sub-inhibitory concentration in the literature. These levels are not considered as lethal concentrations, but they still affect individual cells of bacteria or their populations in various ways. The continuous increase in the prevalence of sub-inhibitory levels of antibiotics in the environment is a key aspect of the current problem of widespread drug resistance worldwide. Sub-inhibitory concentrations of antibiotics are increasingly found in many aquatic environments, such as sewage and sludge, rivers, lakes, and even drinking water and water in pristine environments [20,25,110,111,112,113]. Aquatic environmental concentrations of antibiotics reaching from ng/L to μg/L are generally too low to inhibit bacterial activity, but environmentally relevant concentrations of antibiotics could enhance bacterial communication and transcriptional regulation. Sub-MIC concentrations of antibiotics found in the natural environment are essential to enriching and maintaining drug resistance among bacteria. At sub-inhibitory concentrations of antibiotics, bacteria do not die but their growth is slowed down. Resistance mutations caused by sub-MIC antibiotic concentrations require much less adaptation energy than mutations caused by MIC antibiotic concentrations. Therefore, mutations that incur less fitness cost could be more competitive and enriched in the microbial population [114]. There is also the minimum selective concentration (MSC) of an antibiotic, meaning the lowest antibiotic concentration that is required to select for growth of the resistant mutant [114]. In the study by Gullberg et al. [114] exploring very low concentrations of antibiotics that select resistant bacteria, the MSC value for tetracycline was 15 ng/mL and it was 1/100 of the MIC value of the susceptible wild-type strain. For streptomycin, the MSC value (~1 μg/mL) was 1/4 of the MIC value of the susceptible strain. These results suggest that low antibiotic concentrations have a significant impact on maintaining resistance in the environment. Antibiotics at sub-inhibitory concentrations can function as signaling molecules between cells of the same or different species [115], which has an important role in the evolution of antibiotic resistance [116]. It has been shown that antibiotics at sub-inhibitory concentrations can modulate DNA transcription [117]. The genes affected by sub-inhibitory antibiotics include genes that confer antibiotic resistance, stimulate bacterial adhesion, increase biofilm formation, and regulate mutation frequency [117]. The low doses of antibiotics may favor and sustain antibiotic resistance genes in the environment [118]. Sub-inhibitory concentrations of antibiotics affect bacterial physiology, causing mutagenesis, virulence, biofilm formation, and horizontal gene transfer (HGT) recombination. These doses can cause the occurrence of de novo resistant bacteria or enrichment of preexisting antibiotic resistant bacteria [116]. Different classes of antibiotics at sub-inhibitory concentrations affect bacterial species by inducing biofilm formation, which is a serious global health concern because it increases their ability to tolerate antibiotics, thereby enabling their survival and development [119]. Sub-inhibitory levels of imipenem, a β-lactam antibiotic, cause biofilm formation in Pseudomonas aeruginosa species, one of the major opportunistic pathogens [120]. A similar effect is caused by tobramycin, an aminoglycoside antibiotic [121], and norfloxacin, a fluoroquinolone antibiotic [122], which both induce biofilm formation in P. aeruginosa at sub-inhibitory concentrations. β-lactam antibiotics below MIC concentrations induce biofilm formation in Escherichia coli by inducing colanic acid synthesis, which is involved in adhesion to surfaces in this species [123]. Bacterial biofilm formation triggered by low doses of antibiotics in water favors the colonization of surfaces such as bottom sediments and solid particles found in rivers or soil. The function of antibiotics as signaling molecules also has the effect of promoting horizontal gene exchange in microbial ecosystems [115]. Antibiotics induce a bacterial SOS response to DNA damage, which regulates the horizontal transfer of integrative and conjugative elements encoding bacterial virulence, antibiotic resistance, and variety of other properties of bacterial metabolism [115]. The SOS system is a set of co-regulated genes that is extensive in bacteria and promotes cell survival by repairing damaged genomes [116]. The SOS response can increase the rate of mutation occurrence in genes conferring antibiotic resistance and increase the acquisition of antibiotic resistance [124]. SOS-inducing antibiotics include fluoroquinolones, the mode of action of which involves interaction with two target enzymes: DNA gyrase and topoisomerase IV. Resistance to fluroquinolone antibiotics is mainly caused by point mutations in the quinolone resistance-determining region of gyrase and topoisomerase genes [125]. DNA destructive agents such as fluoroquinolones could increase HGT more than 300-fold [126]. HGT plays a crucial role in environmental dissemination of ARGs and is common in aquatic environments. This mechanism allows pathogenic bacteria to acquire antimicrobial resistance genes from the environmental gene pool (i.e., the environmental resistome) [127] by conjugation, transformation, or transduction [128]. The phenomenon of mixing drug-resistant bacteria of anthropogenic origin with environmental strains occurring in the aquatic environment increases the risk of the emergence of new antibiotic-resistant strains through HGT [129]. A second particularly essential ARG transfer element is an integron—a genetic assembly platform that can encode ARGs. The crucial integron gene encoding integrase (Int1) is induced by the SOS response. Importantly, antibiotics such as quinolone, β-lactams, and trimethoprim can induce the SOS response in bacterial cells [128]. SOS-dependent mutagenesis and horizontal gene transfer are essential factors that determine environmental antibiotic resistance and enhance the environment resistome. Antibiotics at low concentrations also act as signaling molecules that can regulate the homeostasis of microbial communities in the environment.
The presence of sub-inhibitory concentrations of antimicrobial compounds in waters causes great concern about their harmful effects on microbial community composition [130]. Direct effects of antibiotics on microbial populations might affect their abundance and species richness [131]. Antibiotics can negatively impact microbial populations involved in key ecosystem functions. Thus, they reduce biodiversity, which is crucial for maintaining the correctness of biological processes in ecosystems [130]. Importantly, sub-inhibitory levels of antibiotics can reduce bacterial community diversity by increasing the variance in fitness among taxa [132]. In a study by Cairns et al. [132] on the effect of low concentrations of antibiotics on an experimental microbial community of 62 strains, sub-inhibitory concentrations of antibiotics were found to reduce bacterial community diversity. In addition, the diversity-reducing effect of antibiotics was lost in the presence of spatial structures (biofilms) that protected bacterial cells from the effects of pharmaceuticals. The authors suggested that determining the appropriate ecological factors (biotic and abiotic) has a significant impact on understanding the effect of sub-inhibitory concentrations of antibiotics on bacterial community composition. Zhao et al. [133] showed that levofloxacin (LEV) and oxytetracycline (OTC) at 5 μg/L sharply changed the freshwater microbial community structure at the genus level in the microcosm system without affecting the alpha diversity of the bacterial community. After seven days of LEV and OTC exposure, the relative abundance of Proteobacteria significantly increased, while that of Bacteroidetes significantly decreased at the phylum level in both treated groups. At the genus level, the abundance of Flavobacteria and Emticicia decreased while that of Pseudomonas significantly increased in the two treated groups. After 14 days of exposure to LEV and OTC, the microbial composition significantly changed at the genus level compared to the control. Changes at the genus level differed between the LEV-treated and OTC-treated groups. Flavobacteria (significantly lower after 7 days in both groups) and Niveispirilla (dominant in OTC-group after 14 days) were significantly affected by exposure to LEV and OTC antibiotics. The varying impact of the tested antibiotics on the bacterial community suggested that different bacteria in the community are sensitive to different pollutants. It should be emphasized that antibiotic residues in aquatic environments prevail as a mixture of all types of antibiotics, not as a single drug. Therefore, it can be assumed that their effects on bacterial populations will be much greater. Another example of antibiotic impact on the bacterial community composition in water is the research by Waiser et al. [134], in which the specific effects of erythromycin, trimethoprim, and clindamycin on the aquatic bacterial community composition and function of biofilm growth were reported. Erythromycin used at a concentration of 4 mg/L resulted in a bacterial community diversity in cultured biofilms that was always different from the control. Biofilm thickness and bacterial biomass were decreased after erythromycin treatment of the bacterial population. The negative impact of the antibiotics on carbon utilization was also detected. Microbial communities are potentially excellent indicators of changes in ecosystem balance, which can be disturbed by antibiotic contamination. Microorganisms play a crucial role in organic matter biodegradation and biogeochemical nutrient cycling [134]. This is particularly applicable to mountain river ecosystems, which are characterized as having much higher quality that rivers running in urbanized areas with strongly developed industry. Many rivers take their sources in the mountains and these rivers provide drinking water resources. Therefore, single changes such as antibiotic contamination can significantly disturb the stability of these bacterial communities. On the one hand, mountain areas can be characterized by significant biodiversity, while on the other hand, the compositions of bacterial populations can be unique and very sensitive to natural and anthropogenic changes [20].

5. Antibiotic Removal Processes from Water and Wastewater

Removal of antibiotics contained in wastewater (of human and animal origin) is a key aspect that could reduce the contamination of the aquatic environment (surface water and groundwater) with antibiotics. Antibiotic contamination of water creates direct and indirect routes through which antibiotics then enter human organisms. The direct route includes drinking contaminated water, while indirect routes include using contaminated water to irrigate crops that humans and livestock animals then eat or to water livestock animals that are then eaten by humans [52]. From an environmental point of view, antibiotic residues can influence microbial populations by affecting their physiological functions or can lead to the disappearance of key environmental groups of microorganisms. The problem is that conventional WWTPs are not properly prepared to remove pharmaceuticals from raw wastewater using primary treatment methods [135]. Additionally, undegraded antibiotics can adsorb onto sewage sludge in biological treatment plants. Arun et al. (2020) reported 64 ng/g of roxithromycin in the concentrated sludge of a WWTP in China [136]. This is why it is so important to reduce the amounts of antibiotics released into the environment by developing advanced wastewater treatment techniques.
Conventional WWTPs generally use primary (mechanical treatment: filtration and sedimentation) and secondary treatment processes (biological processes to remove organic matter using aerobic or anaerobic systems). The most commonly used biological method is conventional activated sludge. Membrane bioreactors (MBR) are less common [137], probably because of their high operational costs related to maintaining sustainable filtration conditions and high energy consumption [138]. The MBR process comprises aerobic and anaerobic methods, combining modern membrane filtration technology and biological degradation by active sludge. The main advantage of MBR is the high quality of the treated water suitable for its reuse [139]. Membrane bioreactors contain micro- or ultra-filtration membranes ranging from 0.04 to 0.4 μm [140], resulting in significant improvements in the microbial quality of the produced effluent by removal of a wide range of microorganisms by size exclusion [138]. Research on a pilot-scale MBR [141] showed the following percentages of antibiotic elimination from the influent from a Swiss hospital: 51% for ciprofloxacin, 47% for norfloxacin, <60% for erythromycin, 7% for sulfamethoxazole, and 96% for trimethoprim. Xiao et al. [142] showed the reduction of sulfamethoxazole and trimethoprim from wastewater using an anaerobic membrane bioreactor (AnMBR) at levels of 67.8 ± 13.9% and 94.2 ± 5.5%, respectively. High removal of ampicillin (94.4%) was also achieved in MBR treatment [143]. On the other hand, the activated sludge process removed 82% of ampicillin and the disinfection process eliminated 91% of ampicillin [144] in a municipal wastewater treatment plant. MBR sewage treatment is distinguished by the increased quality of the effluent compared to conventional activated sludge systems. MBR is a compact process and its advantages are exploited in ski resorts, hotels, and trailer parks [138]. MBR effluent could be reused for technical applications, such as irrigation, snow production on ski slopes, or other non-potable water industrial applications, and not only directly discharged into the environment. Reusing water is a huge advantage in the midst of the current worldwide problem of water scarcity and water management regulations. This enables reducing the consumption of water resources and eliminating or decreasing concentrations of emerging pollutants, such as antibiotics, introduced into the environment. More developed techniques combine MBR with advanced water treatment, such as activated carbon, UV-irradiation, post-ozonation, or reverse osmosis [138,142,145]. Xiao et al. [142] achieved the elimination of sulfamethoxazole from 67.8 ± 13.9% to 95.5 ± 4.6% after adding 1 g/L of powdered activated carbon to the MBR bioreactor. Alacabey [146] achieved over 99% success in the removal of ciprofloxacin from aqueous systems with activated carbon obtained from pumpkin seed shells. Van der Waals interactions are the dominant mechanisms of organic compound removal (including antibiotics) in the activated carbon adsorption system. Carbon-based materials (activated carbon, carbon nanotubes, and graphene) are highly effective adsorbents for water-polluting antibiotics due to their large specific surface area, high porosity, and high reaction activity [147]. The efficiency of removing various compounds depends on the properties of the adsorbent (e.g., surface polarity and porosity) and the characteristics of the compound itself (e.g., size, charge, and hydrophobicity) [137]. Liu et al. [145] tested a combination of membrane bioreactor with ultraviolet/chlorine (MBR-UV/Cl2) to treat surface water polluted with pharmaceutical personal care products and antibiotics. The average removal efficiencies of selected antibiotics from surface water in the MBR and MBR-UV/Cl2 processes are presented in Figure 3, after supplying 200 ng/L antibiotic standards (based on [145]).
In the presented research [145], the sole MBR system was clearly ineffective in the removal of sulfamethoxazole. However, the MBR permeate subjected to additional treatment by UV irradiation and chlorination (chlorine concentration at 3 mg/L) effectively reduced 95.6% of the sulfamethoxazole contained in the polluted surface water. The UV/Cl2 process evidently increased the antibiotic removal efficiency. Similar results of sulfamethoxazole removal (100 μg/L) were obtained by Tambosi et al. [148]. The study concerned the treatment of wastewater from a municipal treatment plant in Germany. In the MBR process, the permeate was subjected to treatment with advanced techniques such as H2O2/UV, H2O2/Fe2+ (Fenton), H2O2/Fe2+/UV (photo-Fenton), UV radiation, and ozone (O3). The results showed that the elimination efficiency of sulfamethoxazole in the MBR process after 30 minutes of treatment was 64%. However, the advanced oxidation processes (H2O2/UV, H2O2/Fe2+/UV), UV radiation, and ozonation removed 100% of the sulfamethoxazole contained in the MBR permeates. Sulfamethoxazole was very sensitive to all of the applied UV treatment steps. On the other hand, the Fenton process was completely ineffective in the elimination of this compound (elimination rate: 0%). In the same research, the trimethoprim elimination rate by the MBR process was 94% after 30 minutes of treatment. The steps including UV treatment were less effective in trimethoprim elimination: H2O2/UV (7%), H2O2/Fe2+/UV (20%), UV radiation (5%), and Fenton process (10%). Ozone treatment was the most effective process that led to an elimination of 100% sulfamethoxazole and trimethoprim in the MBR permeate. High removal efficiency was also achieved for tetracycline antibiotics in anoxic/aerobic-MBR treatment for artificial wastewater containing antibiotics. The removal efficiencies after 60 days of solid retention times for tetracycline, chlortetracycline, and oxytetracycline were 93.6%, 82.9%, and 88.6%, respectively [143]. Tetracycline is one of the most frequently detected antibiotics in wastewater [149]. Dolar et al. [150] achieved excellent results in an integrated pilot scale membrane bioreactor coupled with reverse osmosis (MBR–RO) for municipal wastewater treatment. In the case of macrolide antibiotics, the MBR treatment showed removal rates of 75% for azithromycin and 87% for erythromycin. Sulfamethoxazole was partially removed (69%) with MBR treatment, whereas the removal of ofloxacin with the MBR system was ineffective in this research. Overall antibiotic removal rates with the RO membrane (pore size range < 0.001 μm) [137] were greater than 99% for each compound tested. Experimental results by Dolar et al. [150] showed that the removal performance of antibiotics was significantly higher in MBR combined with the RO system, which effectively removed low-molecular-weight pharmaceutical compounds.
Recently, metal-organic frameworks (MOFs), a multi-dimensional material held together by bonding between metal atoms and organic ligands, have been shown to be effective in treating wastewater with antibiotic residues [151,152]. MOFs exhibit desirable characteristics, such as large surface area and pore volume, hierarchical structures, biocompatibility, non-toxicity, regeneration capabilities, and tunable pore size and functional groups, that are suitable for wastewater treatment processes [152,153,154]. MOFs can maintain their structures in water conditions [155]. Applying MOFs in WWTPs can significantly improve treatment efficiency. MOFs can be applied in wastewater treatment by conducting adsorption, filtration, and degradation [156], including catalytic degradation of antibiotics by immobilized enzymes [153]. Zhou et al. [157] investigated the detection and removal of tetracycline solution (0.1 mM) in water with a luminescent MOF, resulting in 56% of this antibiotic being removed after 30 minutes. Dong et al. [151] showed photocatalytic decomposition of oxytetracycline with a stable 8-connected Cd(II) MOF as a photocatalyst. Metal-organic frameworks are considered as relevant materials for the adsorption and removal of emerging pollutants, such as antibiotics, in wastewater.
Zeolites can also be used as useful materials in the treatment of antibiotics in sewage. Zeolites are sorption materials which—if appropriately developed and selectively functionalized—can retain antibiotic residues in wastewater treatment systems [158]. The hydrophobicity of zeolites is a beneficial property that facilitates the adsorption of antibiotics in water solutions. High silica-zeolites almost completely (>90%) removed sulfonamide antibiotics from water [159,160]. Natural and modified minerals are also employed in the processes of antibiotic elimination from water. They have unique properties, including high specific surface area, low cost, availability, and good removal efficiency [161]. Natural colemanite mineral (mesoporous material) was used as an adsorbent material for the removal of four common fluoroquinolones from surface water and wastewater samples. Batch adsorption experiments resulted in the following antibiotic elimination: 81.9%, 78.4%, 80.3%, and 79.7% of the initial amounts for ofloxacin, norfloxacin, ciprofloxacin, and enrofloxacin, respectively [162].
A variety of research has reported that MBR systems applying ultrafiltration are partially successful in the removal of antimicrobial agents from wastewater [142,145,148,150]. Advanced techniques combined with MBR treatment, in particular UV radiation, activated carbon, advanced oxidation processes (AOPs), and adsorption methods (MOFs, zeolites, natural materials) (Figure 4), have the potential to be developed as effective technologies in treating wastewater and surface water polluted with antibiotics. The MBR process yields treated effluent of a quality that allows it to be discharged into sensitive water bodies (such as mountain rivers) or reclaimed in a variety of ways (including the production of artificial snow). Due to its compact size, it can be used for municipal sewage treatment and can be successfully applied in places where saving space is an important trait (such as mountain hotels, shelters, etc.). It overcomes the drawbacks of conventional activated sludge processes, including large space requirements for secondary clarifiers and production of excess sludge [138,163]. The application of advanced wastewater treatment methods offers great prospects for the economical reuse of wastewater free of micropollutants such as antibiotics. Reducing the content of antibiotics in wastewater will decrease the supply of these compounds to the environment after the discharge of treated wastewater. This will reduce the risk of developing drug resistance among microorganisms in the environment and other adverse environmental effects.

6. Conclusions

Antibiotics polluting the environment are recognized as emerging micropollutants affecting microbial populations. Water is the main dissemination pathway of antibiotics and drug resistance determinants between various environmental compartments. The rate of antibiotics entering the aquatic environment is higher than their rate of elimination. Long-term exposure to sub-inhibitory concentrations of antibiotics (ng/L-μg/L) in waters is the main driver of changes in the genomes of microorganisms, thus resulting in the emergence of drug resistance and exchange of drug resistance genes by HGT. Antibiotics (acting as signaling molecules) are the ecological factor driving the evolution of bacteria by interfering with their ecological functions and compositions of bacterial communities. This causes the reduction of bacterial biodiversity responsible for the proper occurrence of biological processes in ecosystems. Antibiotics at low concentrations and bioavailability are capable of modifying bacterial communities and affect transcriptional regulation, thereby causing drug resistant mutations. Microorganisms evolve in response to emerging factors, such as antibiotics, in their environment. This is particularly evident in sensitive environments such as pristine mountain ecosystems where rivers can be exposed to strong anthropogenic factors closely related to tourism, agriculture, and animal husbandry.
The research results published so far mainly concern urban environments affected by antibiotic contamination and changes resulting thereof. However, few studies show the impact of antibiotic pollution in pristine mountain environments that are particularly sensitive to changes and under strong anthropogenic pressure resulting, among other reasons, from expansive mountain tourism. Contamination of the mountain environment with antibiotics results from intensive tourist traffic, which causes overloading of wastewater treatment plants with sewage containing micropollutants. Only partial metabolism of antibiotics in human and animal bodies after their administration is responsible for their excretion in unchanged form at levels of 5–85% and their supply in parent form to wastewater treatment plants. Conventional wastewater treatment plants using primary and secondary wastewater treatment processes are not adapted to effectively remove these micropollutants. Therefore, antibiotics end up in surface waters along with treated sewage. In mountainous areas, surface water is used to produce snow on ski slopes and irrigate crops and green areas, as well as for recreational purposes by tourists. Additional wastewater treatment with more advanced methods (MBR, UV radiation, activated carbon, membrane techniques, oxidation processes, MOFs) gives the possibility of significant or complete removal of antimicrobial agents from wastewater. The prospect of using advanced wastewater treatment techniques gives the possibility of safely reusing wastewater for technical purposes, reducing the amount of antibiotics released into the environment, and providing an economical solution for the use of water resources.
Contamination of mountain waters with antibiotics is already present in the upper river courses of high-mountain national parks under protection. Mountain shelters, which are not equipped with sewage systems, are also sources of antibiotic contamination. There is a conflict between maintaining the pristine mountain environment and the continuous development of mountain tourism. The main threat to public health is the development of drug resistance and possible transfer of ARGs from environmental strains to clinical strains. With the continuous supply of sub-inhibitory concentrations of antibiotics in the environment affecting changes in the genomes of microorganisms, there may be a risk of a link between environmental and clinical drug resistance. On the other hand, changes in the biodiversity and composition of microbial populations that are responsible for important ecological functions in the ecosystem pose a threat to the environment. For this reason, monitoring the contamination of surface waters with antimicrobial agents is an important aspect. Contamination of surface waters with antibiotics is particularly harmful in the mountain environment. This is due to the fact that mountain water supplies are a valuable natural resource found mostly in pristine and protected areas where they give rise to rivers and constitute a reservoir of drinking water in every country. Protected environments, which are a valuable source of biodiversity, should be taken care of. Antibiotics are a type of micropollutant that is not routinely tested. Therefore, monitoring concentrations of antibiotics in waters is crucial for maintaining the quality of water resources for human use and the microbiological biodiversity within water ecosystems.
An urgent knowledge gap is the limited understanding of where and at what stage critical changes occur that affect the emergence of antibiotic resistance in bacteria and changes in bacterial population composition. Moreover, these gaps in knowledge are particularly relevant to the pristine mountain environment, where drinking water resources take their origin. Understanding the role of antibiotic contamination in the environment is important in terms of environmental changes and public health implications.

Author Contributions

Writing—original draft preparation, K.K.; writing—review and editing, A.L.-B. and K.W.; visualization, K.K. and K.W.; supervision, A.L.-B. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Meek, R.W.; Vyas, H.; Piddock, L.J.V. Nonmedical Uses of Antibiotics: Time to Restrict Their Use? PLOS Biol. 2015, 13, e1002266. [Google Scholar] [CrossRef] [PubMed]
  2. Van, T.T.H.; Yidana, Z.; Smooker, P.M.; Coloe, P.J. Antibiotic Use in Food Animals Worldwide, with a Focus on Africa: Pluses and Minuses. J. Glob. Antimicrob. Resist. 2020, 20, 170–177. [Google Scholar] [CrossRef] [PubMed]
  3. Du, L.; Liu, W. Occurrence, Fate, and Ecotoxicity of Antibiotics in Agro-Ecosystems. A Review. Agron. Sustain. Dev. 2012, 32, 309–327. [Google Scholar] [CrossRef] [Green Version]
  4. Polianciuc, S.I.; Gurzău, A.E.; Kiss, B.; Ștefan, M.G.; Loghin, F. Antibiotics in the Environment: Causes and Consequences. Med. Pharm. Rep. 2020, 93, 231. [Google Scholar] [CrossRef] [PubMed]
  5. Chow, L.K.M.; Ghaly, T.M.; Gillings, M.R. A Survey of Sub-Inhibitory Concentrations of Antibiotics in the Environment. J. Environ. Sci. 2021, 99, 21–27. [Google Scholar] [CrossRef]
  6. Wang, M.; Shen, W.; Yan, L.; Wang, X.-H.; Xu, H. Stepwise Impact of Urban Wastewater Treatment on the Bacterial Community Structure, Antibiotic Contents, and Prevalence of Antimicrobial Resistance. Environ. Pollut. 2017, 231, 1578–1585. [Google Scholar] [CrossRef]
  7. Ching, C.; Orubu, E.S.F.; Sutradhar, I.; Wirtz, V.J.; Boucher, H.W.; Zaman, M.H. Bacterial Antibiotic Resistance Development and Mutagenesis Following Exposure to Subinhibitory Concentrations of Fluoroquinolones in Vitro: A Systematic Review of the Literature. JAC-Antimicrob. Resist. 2020, 2, dlaa068. [Google Scholar] [CrossRef]
  8. Atterby, C.; Nykvist, M.; Lustig, U.; Andersson, D.I.; Järhult, J.D.; Sandegren, L. Selection of Resistant Bacteria in Mallards Exposed to Subinhibitory Concentrations of Ciprofloxacin in Their Water Environment. Antimicrob. Agents Chemother. 2021, 65, e01858-20. [Google Scholar] [CrossRef]
  9. Manaia, C.M. Assessing the Risk of Antibiotic Resistance Transmission from the Environment to Humans: Non-Direct Proportionality between Abundance and Risk. Trends Microbiol. 2017, 25, 173–181. [Google Scholar] [CrossRef] [Green Version]
  10. Michael-Kordatou, I.; Karaolia, P.; Fatta-Kassinos, D. The Role of Operating Parameters and Oxidative Damage Mechanisms of Advanced Chemical Oxidation Processes in the Combat against Antibiotic-Resistant Bacteria and Resistance Genes Present in Urban Wastewater. Water Res. 2018, 129, 208–230. [Google Scholar] [CrossRef]
  11. Hao, H.; Shi, D.; Yang, D.; Yang, Z.; Qiu, Z.; Liu, W.; Shen, Z.; Yin, J.; Wang, H.; Li, J.; et al. Profiling of Intracellular and Extracellular Antibiotic Resistance Genes in Tap Water. J. Hazard. Mater. 2019, 365, 340–345. [Google Scholar] [CrossRef]
  12. Zarei-Baygi, A.; Smith, A.L. Intracellular versus Extracellular Antibiotic Resistance Genes in the Environment: Prevalence, Horizontal Transfer, and Mitigation Strategies. Bioresour. Technol. 2021, 319, 124181. [Google Scholar] [CrossRef]
  13. von Wintersdorff, C.J.H.; Penders, J.; van Niekerk, J.M.; Mills, N.D.; Majumder, S.; van Alphen, L.B.; Savelkoul, P.H.M.; Wolffs, P.F.G. Dissemination of Antimicrobial Resistance in Microbial Ecosystems through Horizontal Gene Transfer. Front. Microbiol. 2016, 7, 173. [Google Scholar] [CrossRef] [Green Version]
  14. Huijbers, P.M.C.; Blaak, H.; de Jong, M.C.M.; Graat, E.A.M.; Vandenbroucke-Grauls, C.M.J.E.; de Roda Husman, A.M. Role of the Environment in the Transmission of Antimicrobial Resistance to Humans: A Review. Environ. Sci. Technol. 2015, 49, 11993–12004. [Google Scholar] [CrossRef]
  15. Anh, H.Q.; Le, T.P.Q.; Da Le, N.; Lu, X.X.; Duong, T.T.; Garnier, J.; Rochelle-Newall, E.; Zhang, S.; Oh, N.-H.; Oeurng, C.; et al. Antibiotics in Surface Water of East and Southeast Asian Countries: A Focused Review on Contamination Status, Pollution Sources, Potential Risks, and Future Perspectives. Sci. Total Environ. 2021, 764, 142865. [Google Scholar] [CrossRef]
  16. Sta Ana, K.M.; Madriaga, J.; Espino, M.P. β-Lactam Antibiotics and Antibiotic Resistance in Asian Lakes and Rivers: An Overview of Contamination, Sources and Detection Methods. Environ. Pollut. 2021, 275, 116624. [Google Scholar] [CrossRef]
  17. Mittal, P.; Prasoodanan PK, V.; Dhakan, D.B.; Kumar, S.; Sharma, V.K. Metagenome of a Polluted River Reveals a Reservoir of Metabolic and Antibiotic Resistance Genes. Environ. Microbiome 2019, 14, 5. [Google Scholar] [CrossRef] [Green Version]
  18. Koniuszewska, I.; Korzeniewska, E.; Harnisz, M.; Kiedrzyńska, E.; Kiedrzyński, M.; Czatzkowska, M.; Jarosiewicz, P.; Zalewski, M. The Occurrence of Antibiotic-Resistance Genes in the Pilica River, Poland. Ecohydrol. Hydrobiol. 2020, 20, 1–11. [Google Scholar] [CrossRef]
  19. Harnisz, M.; Gołaś, I.; Pietruk, M. Tetracycline-Resistant Bacteria as Indicators of Antimicrobial Resistance in Protected Waters—The Example of the Drwęca River Nature Reserve (Poland). Ecol. Indic. 2011, 11, 663–668. [Google Scholar] [CrossRef]
  20. Lenart-Boroń, A.M.; Boroń, P.M.; Prajsnar, J.A.; Guzik, M.W.; Żelazny, M.S.; Pufelska, M.D.; Chmiel, M.J. COVID-19 Lockdown Shows How Much Natural Mountain Regions Are Affected by Heavy Tourism. Sci. Total Environ. 2022, 806, 151355. [Google Scholar] [CrossRef]
  21. Tan, B.; Ng, C.; Nshimyimana, J.P.; Loh, L.L.; Gin, K.Y.-H.; Thompson, J.R. Next-Generation Sequencing (NGS) for Assessment of Microbial Water Quality: Current Progress, Challenges, and Future Opportunities. Front. Microbiol. 2015, 6, 1027. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Yergeau, E.; Lawrence, J.R.; Sanschagrin, S.; Waiser, M.J.; Korber, D.R.; Greer, C.W. Next-Generation Sequencing of Microbial Communities in the Athabasca River and Its Tributaries in Relation to Oil Sands Mining Activities. Appl. Environ. Microbiol. 2012, 78, 7626–7637. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Nimnoi, P.; Pongsilp, N. Marine Bacterial Communities in the Upper Gulf of Thailand Assessed by Illumina Next-Generation Sequencing Platform. BMC Microbiol. 2020, 20, 19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Viviroli, D.; Weingartner, R. The Hydrological Significance of Mountains: From Regional to Global Scale. Hydrol. Earth Syst. Sci. 2004, 8, 1017–1030. [Google Scholar] [CrossRef] [Green Version]
  25. Lenart-Boroń, A.; Prajsnar, J.; Guzik, M.; Boroń, P.; Chmiel, M. How Much of Antibiotics Can Enter Surface Water with Treated Wastewater and How It Affects the Resistance of Waterborne Bacteria: A Case Study of the Białka River Sewage Treatment Plant. Environ. Res. 2020, 191, 110037. [Google Scholar] [CrossRef]
  26. Lenart-Boroń, A.; Wolanin, A.; Jelonkiewicz, Ł.; Chmielewska-Błotnicka, D.; Żelazny, M. Spatiotemporal Variability in Microbiological Water Quality of the Białka River and Its Relation to the Selected Physicochemical Parameters of Water. Water. Air. Soil Pollut. 2016, 227, 22. [Google Scholar] [CrossRef]
  27. Mokhtar, M.B.; Aris, A.Z.; Abdullah, M.H.; Yusoff, M.K.; Abdullah, M.P.; Idris, A.R.; Raja Uzir, R.I. A Pristine Environment and Water Quality in Perspective: Maliau Basin, Borneo’s Mysterious World. Water Environ. J. 2009, 23, 219–228. [Google Scholar] [CrossRef]
  28. Bojarczuk, A.; Jelonkiewicz, Ł.; Lenart-Boroń, A. The Effect of Anthropogenic and Natural Factors on the Prevalence of Physicochemical Parameters of Water and Bacterial Water Quality Indicators along the River Białka, Southern Poland. Environ. Sci. Pollut. Res. 2018, 25, 10102–10114. [Google Scholar] [CrossRef] [Green Version]
  29. Vanat, L. International Report on Snow & Mountain Tourism—Overview of the Key Industry Figures for Ski Resorts. 2020. Available online: https://www.vanat.ch/RM-world-report-2020.pdf (accessed on 10 December 2022).
  30. Krzesiwo, K. Ocena wielkości ruchu turystycznego w ośrodku narciarskim kotelnica białczańska w sezonie zimowym 2014/2015. Pr. Geogr. 2016, 24, 47–70. [Google Scholar] [CrossRef]
  31. Krzesiwo, K. Ocena sytuacji rozwojowej i funkcjonalnej stacji narciarskich—Przykład polskich Karpat. Stud. Ind. Geogr. Comm. Pol. Geogr. Soc. 2021, 35, 259–276. [Google Scholar] [CrossRef]
  32. Pickering, C.M.; Harrington, J.; Worboys, G. Environmental Impacts of Tourism on the Australian Alps Protected Areas. Mt. Res. Dev. 2003, 23, 9. [Google Scholar] [CrossRef] [Green Version]
  33. Sanchez-Cid, C.; Keuschnig, C.; Torzewski, K.; Stachnik, Ł.; Kępski, D.; Luks, B.; Nawrot, A.; Niedzielski, P.; Vogel, T.M.; Larose, C. Environmental and Anthropogenic Factors Shape the Snow Microbiome and Antibiotic Resistome. Front. Microbiol. 2022, 13, 918622. [Google Scholar] [CrossRef]
  34. Krzesiwo, K.; Ziółkowska-Weiss, K.; Żemła, M. The Attractiveness of Selected Central European Countries for Winter Sports and Mountain Hiking. Turyzm/Tourism 2018, 28, 35–40. [Google Scholar] [CrossRef] [Green Version]
  35. Opermanis, O.; MacSharry, B.; Aunins, A.; Sipkova, Z. Connectedness and Connectivity of the Natura 2000 Network of Protected Areas across Country Borders in the European Union. Biol. Conserv. 2012, 153, 227–238. [Google Scholar] [CrossRef]
  36. Pietrzyk-Kaszyńska, A.; Cent, J.; Grodzińska-Jurczak, M.; Szymańska, M. Factors Influencing Perception of Protected Areas—The Case of Natura 2000 in Polish Carpathian Communities. J. Nat. Conserv. 2012, 20, 284–292. [Google Scholar] [CrossRef]
  37. Holden, A. Winter Tourism and the Environment in Conflict: The Case of Cairngorm, Scotland. Int. J. Tour. Res. 2000, 2, 247–260. [Google Scholar] [CrossRef]
  38. Tatra National Park, Poland Sale of Admission Tickets to the Tatra National Park—Statistics. 2022. Available online: https://tpn.pl/zwiedzaj/turystyka/statystyka (accessed on 10 December 2022).
  39. Now the Environment The Tatra National Park Wants to Channel All Shelters-an Interview. 2020. Available online: https://www.teraz-srodowisko.pl/aktualnosci/tatrzanski-park-narodowy-chce-skanalizowac-wszystkie-schroniska-8205.html (accessed on 10 December 2022).
  40. Lenart-Boroń, A.; Prajsnar, J.; Guzik, M.; Boroń, P.; Grad, B.; Żelazny, M. Antibiotics in Groundwater and River Water of Białka—A Pristine Mountain River. Appl. Sci. 2022, 12, 12743. [Google Scholar] [CrossRef]
  41. Kangas, K.; Vuori, K.-M.; Määttä-Juntunen, H.; Siikamäki, P. Impacts of Ski Resorts on Water Quality of Boreal Lakes: A Case Study in Northern Finland. Boreal Env. Res. 2012, 17, 313–325. [Google Scholar]
  42. Senetra, A.; Dynowski, P.; Cieślak, I.; Źróbek-Sokolnik, A. An Evaluation of the Impact of Hiking Tourism on the Ecological Status of Alpine Lakes—A Case Study of the Valley of Dolina Pięciu Stawów Polskich in the Tatra Mountains. Sustainability 2020, 12, 2963. [Google Scholar] [CrossRef] [Green Version]
  43. Kovalakova, P.; Cizmas, L.; McDonald, T.J.; Marsalek, B.; Feng, M.; Sharma, V.K. Occurrence and Toxicity of Antibiotics in the Aquatic Environment: A Review. Chemosphere 2020, 251, 126351. [Google Scholar] [CrossRef]
  44. Boxall, A.B.A.; Blackwell, P.; Cavallo, R.; Kay, P.; Tolls, J. The Sorption and Transport of a Sulphonamide Antibiotic in Soil Systems. Toxicol. Lett. 2002, 131, 19–28. [Google Scholar] [CrossRef] [PubMed]
  45. Medication Datasheet: Levofloxacin 500 Mg Aurovitas Pharma Poland; Aurovitas Pharma: Warsaw, Poland, 2022.
  46. Varoquaux, O.; Lajoie, D.; Gobert, C.; Cordonnier, P.; Ducreuzet, C.; Pays, M.; Advenier, C. Pharmacokinetics of the Trimethoprim-Sulphamethoxazole Combination in the Elderly. Br. J. Clin. Pharmacol. 1985, 20, 575–581. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Medication Datasheet: METRONIDAZOL I.V. BRAUN 5 Mg/Ml. Regist. Med. Prod. B. Braun Melsungen AG, Germany. Available online: https://www.bbraun.co.id/en/products/b/metronidazole-b-braun.html (accessed on 15 December 2022).
  48. Leung, H.W.; Minh, T.B.; Murphy, M.B.; Lam, J.C.W.; So, M.K.; Martin, M.; Lam, P.K.S.; Richardson, B.J. Distribution, Fate and Risk Assessment of Antibiotics in Sewage Treatment Plants in Hong Kong, South China. Environ. Int. 2012, 42, 1–9. [Google Scholar] [CrossRef] [PubMed]
  49. Regitano, J.B.; Leal, R.M.P. Comportamento e impacto ambiental de antibióticos usados na produção animal brasileira. Rev. Bras. Ciênc. Solo 2010, 34, 601–616. [Google Scholar] [CrossRef]
  50. Kolz, A.C.; Moorman, T.B.; Ong, S.K.; Scoggin, K.D.; Douglass, E.A. Degradation and Metabolite Production of Tylosin in Anaerobic and Aerobic Swine-Manure Lagoons. Water Environ. Res. 2005, 77, 49–56. [Google Scholar] [CrossRef] [Green Version]
  51. Montforts, M.H.M.M.; Kalf, D.F.; van Vlaardingen, P.L.A.; Linders, J.B.H.J. The Exposure Assessment for Veterinary Medicinal Products. Sci. Total Environ. 1999, 225, 119–133. [Google Scholar] [CrossRef]
  52. Carvalho, I.T.; Santos, L. Antibiotics in the Aquatic Environments: A Review of the European Scenario. Environ. Int. 2016, 94, 736–757. [Google Scholar] [CrossRef]
  53. Pulicharla, R.; Brar, S.K.; Drogui, P.; Verma, M.; Surampalli, R.Y. Removal Processes of Antibiotics in Waters and Wastewaters: Crucial Link to Physical-Chemical Properties and Degradation. J. Hazard. Toxic Radioact. Waste 2015, 19, 04015008. [Google Scholar] [CrossRef]
  54. Blair, B.; Nikolaus, A.; Hedman, C.; Klaper, R.; Grundl, T. Evaluating the Degradation, Sorption, and Negative Mass Balances of Pharmaceuticals and Personal Care Products during Wastewater Treatment. Chemosphere 2015, 134, 395–401. [Google Scholar] [CrossRef]
  55. Burch, K.D.; Han, B.; Pichtel, J.; Zubkov, T. Removal Efficiency of Commonly Prescribed Antibiotics via Tertiary Wastewater Treatment. Environ. Sci. Pollut. Res. 2019, 26, 6301–6310. [Google Scholar] [CrossRef]
  56. Wang, J.; Wang, S. Removal of Pharmaceuticals and Personal Care Products (PPCPs) from Wastewater: A Review. J. Environ. Manag. 2016, 182, 620–640. [Google Scholar] [CrossRef]
  57. Verlicchi, P.; Al Aukidy, M.; Zambello, E. Occurrence of Pharmaceutical Compounds in Urban Wastewater: Removal, Mass Load and Environmental Risk after a Secondary Treatment—A Review. Sci. Total Environ. 2012, 429, 123–155. [Google Scholar] [CrossRef]
  58. Buta, M.; Hubeny, J.; Zieliński, W.; Harnisz, M.; Korzeniewska, E. Sewage Sludge in Agriculture—The Effects of Selected Chemical Pollutants and Emerging Genetic Resistance Determinants on the Quality of Soil and Crops—A Review. Ecotoxicol. Environ. Saf. 2021, 214, 112070. [Google Scholar] [CrossRef]
  59. McClellan, K.; Halden, R.U. Pharmaceuticals and Personal Care Products in Archived U.S. Biosolids from the 2001 EPA National Sewage Sludge Survey. Water Res. 2010, 44, 658–668. [Google Scholar] [CrossRef] [Green Version]
  60. An, J.; Chen, H.; Wei, S.; Gu, J. Antibiotic Contamination in Animal Manure, Soil, and Sewage Sludge in Shenyang, Northeast China. Environ. Earth Sci. 2015, 74, 5077–5086. [Google Scholar] [CrossRef]
  61. Cheng, M.; Wu, L.; Huang, Y.; Luo, Y.; Christie, P. Total Concentrations of Heavy Metals and Occurrence of Antibiotics in Sewage Sludges from Cities throughout China. J. Soils Sediments 2014, 14, 1123–1135. [Google Scholar] [CrossRef]
  62. Hörsing, M.; Ledin, A.; Grabic, R.; Fick, J.; Tysklind, M.; la Jansen, J.C.; Andersen, H.R. Determination of Sorption of Seventy-Five Pharmaceuticals in Sewage Sludge. Water Res. 2011, 45, 4470–4482. [Google Scholar] [CrossRef] [Green Version]
  63. Economou, V.; Gousia, P. Agriculture and Food Animals as a Source of Antimicrobial-Resistant Bacteria. Infect. Drug Resist. 2015, 8, 49–91. [Google Scholar] [CrossRef] [Green Version]
  64. Baptiste, K.E. Do Antimicrobial Mass Medications Work? A Systematic Review and Meta-Analysis of Randomized Clinical Trials Investigating Antimicrobial Prophylaxis or Metaphylaxis against Naturally Occurring Bovine Respiratory Disease. Pathog. Dis. 2017, 29, ftx083. [Google Scholar] [CrossRef] [Green Version]
  65. Hosain, M.Z.; Kabir, S.M.L.; Kamal, M.M. Antimicrobial Uses for Livestock Production in Developing Countries. Vet. World 2021, 14, 210–221. [Google Scholar] [CrossRef]
  66. Spielmeyer, A. Occurrence and Fate of Antibiotics in Manure during Manure Treatments: A Short Review. Sustain. Chem. Pharm. 2018, 9, 76–86. [Google Scholar] [CrossRef]
  67. Zhang, X.; Li, Y.; Liu, B.; Wang, J.; Feng, C.; Gao, M.; Wang, L. Prevalence of Veterinary Antibiotics and Antibiotic-Resistant Escherichia Coli in the Surface Water of a Livestock Production Region in Northern China. PLoS ONE 2014, 9, e111026. [Google Scholar] [CrossRef] [PubMed]
  68. Slana, M.; Pahor, V.; Cvitkovič Maričič, L.; Sollner-Dolenc, M. Excretion Pattern of Enrofloxacin after Oral Treatment of Chicken Broilers. J. Vet. Pharmacol. Ther. 2014, 37, 611–614. [Google Scholar] [CrossRef] [PubMed]
  69. Ince, B.; Coban, H.; Turker, G.; Ertekin, E.; Ince, O. Effect of Oxytetracycline on Biogas Production and Active Microbial Populations during Batch Anaerobic Digestion of Cow Manure. Bioprocess Biosyst. Eng. 2013, 36, 541–546. [Google Scholar] [CrossRef]
  70. Qiu, J.; Zhao, T.; Liu, Q.; He, J.; He, D.; Wu, G.; Li, Y.; Jiang, C.; Xu, Z. Residual Veterinary Antibiotics in Pig Excreta after Oral Administration of Sulfonamides. Environ. Geochem. Health 2016, 38, 549–556. [Google Scholar] [CrossRef]
  71. Wei, R.; He, T.; Zhang, S.; Zhu, L.; Shang, B.; Li, Z.; Wang, R. Occurrence of Seventeen Veterinary Antibiotics and Resistant Bacterias in Manure-Fertilized Vegetable Farm Soil in Four Provinces of China. Chemosphere 2019, 215, 234–240. [Google Scholar] [CrossRef]
  72. Kumar, K.; Gupta, S.C.; Baidoo, S.K.; Chander, Y.; Rosen, C.J. Antibiotic Uptake by Plants from Soil Fertilized with Animal Manure. J. Environ. Qual. 2005, 34, 2082–2085. [Google Scholar] [CrossRef] [Green Version]
  73. Laconi, A.; Mughini-Gras, L.; Tolosi, R.; Grilli, G.; Trocino, A.; Carraro, L.; Di Cesare, F.; Cagnardi, P.; Piccirillo, A. Microbial Community Composition and Antimicrobial Resistance in Agricultural Soils Fertilized with Livestock Manure from Conventional Farming in Northern Italy. Sci. Total Environ. 2021, 760, 143404. [Google Scholar] [CrossRef]
  74. Lenart-Boroń, A.; Prajsnar, J.; Boroń, P. Survival and Antibiotic Resistance of Bacteria in Artificial Snow Produced from Contaminated Water. Water Environ. Res. 2017, 89, 2059–2069. [Google Scholar] [CrossRef]
  75. Segawa, T.; Takeuchi, N.; Rivera, A.; Yamada, A.; Yoshimura, Y.; Barcaza, G.; Shinbori, K.; Motoyama, H.; Kohshima, S.; Ushida, K. Distribution of Antibiotic Resistance Genes in Glacier Environments: Antibiotic Resistance Genes in Snow and Ice. Environ. Microbiol. Rep. 2013, 5, 127–134. [Google Scholar] [CrossRef]
  76. Yang, S.; Carlson, K. Evolution of Antibiotic Occurrence in a River through Pristine, Urban and Agricultural Landscapes. Water Res. 2003, 37, 4645–4656. [Google Scholar] [CrossRef]
  77. Jalloul, G.; Keniar, I.; Tehrani, A.; Boyadjian, C. Antibiotics Contaminated Irrigation Water: An Overview on Its Impact on Edible Crops and Visible Light Active Titania as Potential Photocatalysts for Irrigation Water Treatment. Front. Environ. Sci. 2021, 9, 767963. [Google Scholar] [CrossRef]
  78. Pan, M.; Chu, L.M. Fate of Antibiotics in Soil and Their Uptake by Edible Crops. Sci. Total Environ. 2017, 599–600, 500–512. [Google Scholar] [CrossRef]
  79. Yang, Q.; Gao, Y.; Ke, J.; Show, P.L.; Ge, Y.; Liu, Y.; Guo, R.; Chen, J. Antibiotics: An Overview on the Environmental Occurrence, Toxicity, Degradation, and Removal Methods. Bioengineered 2021, 12, 7376–7416. [Google Scholar] [CrossRef]
  80. Mitchell, S.M. PH and Temperature Effects on the Hydrolysis of Three β-Lactam Antibiotics: Ampicillin, Cefalotin and Cefoxitin. Sci. Total Environ. 2014, 9, 547–555. [Google Scholar] [CrossRef]
  81. Xuan, R.; Arisi, L.; Wang, Q.; Yates, S.R.; Biswas, K.C. Hydrolysis and Photolysis of Oxytetracycline in Aqueous Solution. J. Environ. Sci. Health Part B 2009, 45, 73–81. [Google Scholar] [CrossRef]
  82. Fang, L.; Zhou, Y.; Huang, Z.; Yang, G.; Li, T.; Song, C.; Chen, J. Dynamic Elimination of Enrofloxacin Under Varying Temperature and PH in Aquaculture Water: An Orthogonal Study. Bull. Environ. Contam. Toxicol. 2021, 106, 866–872. [Google Scholar] [CrossRef]
  83. Jiang, M.; Wang, L.; Ji, R. Biotic and Abiotic Degradation of Four Cephalosporin Antibiotics in a Lake Surface Water and Sediment. Chemosphere 2010, 80, 1399–1405. [Google Scholar] [CrossRef]
  84. Reis, A.C.; Kolvenbach, B.A.; Nunes, O.C.; Corvini, P.F.X. Biodegradation of Antibiotics: The New Resistance Determinants—Part I. New Biotechnol. 2020, 54, 34–51. [Google Scholar] [CrossRef]
  85. Timm, A.; Borowska, E.; Majewsky, M.; Merel, S.; Zwiener, C.; Bräse, S.; Horn, H. Photolysis of Four Β-lactam Antibiotics under Simulated Environmental Conditions: Degradation, Transformation Products and Antibacterial Activity. Sci. Total Environ. 2019, 651, 1605–1612. [Google Scholar] [CrossRef]
  86. Liu, X.; Lv, K.; Deng, C.; Yu, Z.; Shi, J.; Johnson, A.C. Persistence and Migration of Tetracycline, Sulfonamide, Fluoroquinolone, and Macrolide Antibiotics in Streams Using a Simulated Hydrodynamic System. Environ. Pollut. 2019, 252, 1532–1538. [Google Scholar] [CrossRef] [PubMed]
  87. Wei, C.; Li, X.; Xie, Y.; Wang, X. Direct Photo Transformation of Tetracycline and Sulfanomide Group Antibiotics in Surface Water: Kinetics, Toxicity and Site Modeling. Sci. Total Environ. 2019, 686, 1–9. [Google Scholar] [CrossRef] [PubMed]
  88. Andreozzi, R.; Raffaele, M.; Nicklas, P. Pharmaceuticals in STP Effluents and Their Solar Photodegradation in Aquatic Environment. Chemosphere 2003, 50, 1319–1330. [Google Scholar] [CrossRef] [PubMed]
  89. Lin, Y.-C.; Hsiao, K.-W.; Lin, A.Y.-C. Photolytic Degradation of Ciprofloxacin in Solid and Aqueous Environments: Kinetics, Phototransformation Pathways, and Byproducts. Environ. Sci. Pollut. Res. 2018, 25, 2303–2312. [Google Scholar] [CrossRef]
  90. Kwon, J.-W. Environmental Impact Assessment of Veterinary Drug on Fish Aquaculture for Food Safety: Environmental Impact Assessment on Veterinary Drugs for Aquaculture. Drug Test. Anal. 2016, 8, 556–564. [Google Scholar] [CrossRef] [Green Version]
  91. Han, S.; Li, X.; Huang, H.; Wang, T.; Wang, Z.; Fu, X.; Zhou, Z.; Du, P.; Li, X. Simultaneous Determination of Seven Antibiotics and Five of Their Metabolites in Municipal Wastewater and Evaluation of Their Stability under Laboratory Conditions. Int. J. Env. Res. Public Health 2021, 14, 10640. [Google Scholar] [CrossRef]
  92. Martín, B.S.; Cornejo, J.; Iragüen, D.; Hidalgo, H.; Anadón, A. Depletion Study of Enrofloxacin and Its Metabolite Ciprofloxacin in Edible Tissues and Feathers of White Leghorn Hens by Liquid Chromatography Coupled with Tandem Mass Spectrometry. J. Food Prot. 2007, 70, 1952–1957. [Google Scholar] [CrossRef]
  93. Jiao, S.; Zheng, S.; Yin, D.; Wang, L.; Chen, L. Aqueous Photolysis of Tetracycline and Toxicity of Photolytic Products to Luminescent Bacteria. Chemosphere 2008, 73, 377–382. [Google Scholar] [CrossRef]
  94. He, X.; Mezyk, S.P.; Michael, I.; Fatta-Kassinos, D.; Dionysiou, D.D. Degradation Kinetics and Mechanism of β-Lactam Antibiotics by the Activation of H2O2 and Na2S2O8 under UV-254 nm Irradiation. J. Hazard. Mater. 2014, 279, 375–383. [Google Scholar] [CrossRef]
  95. Pils, J.R.V.; Laird, D.A. Sorption of Tetracycline and Chlortetracycline on K- and Ca-Saturated Soil Clays, Humic Substances, and Clay−Humic Complexes. Environ. Sci. Technol. 2007, 41, 1928–1933. [Google Scholar] [CrossRef]
  96. Li, J.; Zhang, K.; Zhang, H. Adsorption of Antibiotics on Microplastics. Environ. Pollut. 2018, 237, 460–467. [Google Scholar] [CrossRef]
  97. Wang, S.; Xue, N.; Li, W.; Zhang, D.; Pan, X.; Luo, Y. Selectively Enrichment of Antibiotics and ARGs by Microplastics in River, Estuary and Marine Waters. Sci. Total Environ. 2020, 708, 134594. [Google Scholar] [CrossRef]
  98. McCormick, A.R.; Hoellein, T.J.; London, M.G.; Hittie, J.; Scott, J.W.; Kelly, J.J. Microplastic in Surface Waters of Urban Rivers: Concentration, Sources, and Associated Bacterial Assemblages. Ecosphere 2016, 7, e01556. [Google Scholar] [CrossRef]
  99. Woappi, Y.; Gabani, P.; Singh, A.; Singh, O.V. Antibiotrophs: The Complexity of Antibiotic-Subsisting and Antibiotic-Resistant Microorganisms. Crit. Rev. Microbiol. 2016, 42, 17–30. [Google Scholar] [CrossRef]
  100. Cha, J.; Carlson, K.H. Biodegradation of Veterinary Antibiotics in Lagoon Waters. Process Saf. Environ. Prot. 2019, 127, 306–313. [Google Scholar] [CrossRef]
  101. Yang, C.-W.; Liu, C.; Chang, B.-V. Biodegradation of Amoxicillin, Tetracyclines and Sulfonamides in Wastewater Sludge. Water 2020, 12, 2147. [Google Scholar] [CrossRef]
  102. Wang, J.; Wang, S. Microbial Degradation of Sulfamethoxazole in the Environment. Appl. Microbiol. Biotechnol. 2018, 102, 3573–3582. [Google Scholar] [CrossRef]
  103. Ding, C.; He, J. Effect of Antibiotics in the Environment on Microbial Populations. Appl. Microbiol. Biotechnol. 2010, 87, 925–941. [Google Scholar] [CrossRef]
  104. Klaver, A.L.; Matthews, R.A. Effects of Oxytetracycline on Nitrification in a Model Aquatic System. Aquaculture 1994, 123, 237–247. [Google Scholar] [CrossRef]
  105. Halling-Sørensen, B. Inhibition of Aerobic Growth and Nitrification of Bacteria in Sewage Sludge by Antibacterial Agents. Arch. Environ. Contam. Toxicol. 2001, 40, 451–460. [Google Scholar] [CrossRef]
  106. Fountoulakis, M.; Drillia, P.; Stamatelatou, K.; Lyberatos, G. Toxic Effect of Pharmaceuticals on Methanogenesis. Water Sci. Technol. 2004, 50, 335–340. [Google Scholar] [CrossRef] [PubMed]
  107. Córdova-Kreylos, A.L.; Scow, K.M. Effects of Ciprofloxacin on Salt Marsh Sediment Microbial Communities. ISME J. 2007, 1, 585–595. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Özel Duygan, B.D.; Gaille, C.; Fenner, K.; van der Meer, J.R. Assessing Antibiotics Biodegradation and Effects at Sub-Inhibitory Concentrations by Quantitative Microbial Community Deconvolution. Front. Environ. Sci. 2021, 9, 737247. [Google Scholar] [CrossRef]
  109. Kümmerer, K. Antibiotics in the Aquatic Environment—A Review—Part II. Chemosphere 2009, 75, 435–441. [Google Scholar] [CrossRef]
  110. Baquero, F.; Martínez, J.-L.; Cantón, R. Antibiotics and Antibiotic Resistance in Water Environments. Curr. Opin. Biotechnol. 2008, 19, 260–265. [Google Scholar] [CrossRef]
  111. Khan, G.A.; Berglund, B.; Khan, K.M.; Lindgren, P.-E.; Fick, J. Occurrence and Abundance of Antibiotics and Resistance Genes in Rivers, Canal and near Drug Formulation Facilities—A Study in Pakistan. PLoS ONE 2013, 8, e62712. [Google Scholar] [CrossRef]
  112. Fram, M.S.; Belitz, K. Occurrence and Concentrations of Pharmaceutical Compounds in Groundwater Used for Public Drinking-Water Supply in California. Sci. Total Environ. 2011, 409, 3409–3417. [Google Scholar] [CrossRef] [Green Version]
  113. Jiang, L.; Hu, X.; Xu, T.; Zhang, H.; Sheng, D.; Yin, D. Prevalence of Antibiotic Resistance Genes and Their Relationship with Antibiotics in the Huangpu River and the Drinking Water Sources, Shanghai, China. Sci. Total Environ. 2013, 458–460, 267–272. [Google Scholar] [CrossRef]
  114. Gullberg, E.; Cao, S.; Berg, O.G.; Ilbäck, C.; Sandegren, L.; Hughes, D.; Andersson, D.I. Selection of Resistant Bacteria at Very Low Antibiotic Concentrations. PLoS Pathog. 2011, 7, e1002158. [Google Scholar] [CrossRef] [Green Version]
  115. Aminov, R.I. The Role of Antibiotics and Antibiotic Resistance in Nature. Environ. Microbiol. 2009, 11, 2970–2988. [Google Scholar] [CrossRef]
  116. Andersson, D.I.; Hughes, D. Microbiological Effects of Sublethal Levels of Antibiotics. Nat. Rev. Microbiol. 2014, 12, 465–478. [Google Scholar] [CrossRef]
  117. Davies, J.; Spiegelman, G.B.; Yim, G. The World of Subinhibitory Antibiotic Concentrations. Curr. Opin. Microbiol. 2006, 9, 445–453. [Google Scholar] [CrossRef]
  118. Kümmerer, K. Antibiotics in the Aquatic Environment—A Review—Part I. Chemosphere 2009, 75, 417–434. [Google Scholar] [CrossRef]
  119. Sharma, D.; Misba, L.; Khan, A.U. Antibiotics versus Biofilm: An Emerging Battleground in Microbial Communities. Antimicrob. Resist. Infect. Control 2019, 8, 76. [Google Scholar] [CrossRef] [Green Version]
  120. Bagge, N.; Schuster, M.; Hentzer, M.; Ciofu, O.; Givskov, M.; Greenberg, E.P.; Høiby, N. Pseudomonas Aeruginosa Biofilms Exposed to Imipenem Exhibit Changes in Global Gene Expression and β-Lactamase and Alginate Production. Antimicrob. Agents Chemother. 2004, 48, 1175–1187. [Google Scholar] [CrossRef] [Green Version]
  121. Hoffman, L.R.; D’Argenio, D.A.; MacCoss, M.J.; Zhang, Z.; Jones, R.A.; Miller, S.I. Aminoglycoside Antibiotics Induce Bacterial Biofilm Formation. Nature 2005, 436, 1171–1175. [Google Scholar] [CrossRef]
  122. Linares, J.F.; Gustafsson, I.; Baquero, F.; Martinez, J.L. Antibiotics as Intermicrobial Signaling Agents Instead of Weapons. Proc. Natl. Acad. Sci. USA 2006, 103, 19484–19489. [Google Scholar] [CrossRef] [Green Version]
  123. Sailer, F.C.; Meberg, B.M.; Young, K.D. Î2-Lactam Induction of Colanic Acid Gene Expression in Escherichia Coli. FEMS Microbiol. Lett. 2003, 226, 245–249. [Google Scholar] [CrossRef] [Green Version]
  124. Crane, J.K.; Alvarado, C.L.; Sutton, M.D. Role of the SOS Response in the Generation of Antibiotic Resistance In Vivo. Antimicrob. Agents Chemother. 2021, 65, e00013-21. [Google Scholar] [CrossRef]
  125. Qin, T.-T.; Kang, H.-Q.; Ma, P.; Li, P.-P.; Huang, L.-Y.; Gu, B. SOS Response and Its Regulation on the Fluoroquinolone Resistance. Ann. Transl. Med. 2015, 3, 17. [Google Scholar]
  126. Beaber, J.W.; Hochhut, B.; Waldor, M.K. SOS Response Promotes Horizontal Dissemination of Antibiotic Resistance Genes. Nature 2004, 427, 72–74. [Google Scholar] [CrossRef] [PubMed]
  127. Crofts, T.S.; Gasparrini, A.J.; Dantas, G. Next-Generation Approaches to Understand and Combat the Antibiotic Resistome. Nat. Rev. Microbiol. 2017, 15, 422–434. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Berglund, B. Environmental Dissemination of Antibiotic Resistance Genes and Correlation to Anthropogenic Contamination with Antibiotics. Infect. Ecol. Epidemiol. 2015, 5, 28564. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Wellington, E.M.; Boxall, A.B.; Cross, P.; Feil, E.J.; Gaze, W.H.; Hawkey, P.M.; Johnson-Rollings, A.S.; Jones, D.L.; Lee, N.M.; Otten, W.; et al. The Role of the Natural Environment in the Emergence of Antibiotic Resistance in Gram-Negative Bacteria. Lancet Infect. Dis. 2013, 13, 155–165. [Google Scholar] [CrossRef]
  130. Grenni, P.; Ancona, V.; Barra Caracciolo, A. Ecological Effects of Antibiotics on Natural Ecosystems: A Review. Microchem. J. 2018, 136, 25–39. [Google Scholar] [CrossRef]
  131. Felis, E.; Kalka, J.; Sochacki, A.; Kowalska, K.; Bajkacz, S.; Harnisz, M.; Korzeniewska, E. Antimicrobial Pharmaceuticals in the Aquatic Environment—Occurrence and Environmental Implications. Eur. J. Pharmacol. 2020, 866, 172813. [Google Scholar] [CrossRef]
  132. Cairns, J.; Ruokolainen, L.; Hultman, J.; Tamminen, M.; Virta, M.; Hiltunen, T. Ecology Determines How Low Antibiotic Concentration Impacts Community Composition and Horizontal Transfer of Resistance Genes. Commun. Biol. 2018, 1, 35. [Google Scholar] [CrossRef] [Green Version]
  133. Zhou, Z.; Zhang, Z.; Feng, L.; Zhang, J.; Li, Y.; Lu, T.; Qian, H. Adverse Effects of Levofloxacin and Oxytetracycline on Aquatic Microbial Communities. Sci. Total Environ. 2020, 734, 139499. [Google Scholar] [CrossRef]
  134. Waiser, M.J.; Swerhone, G.D.W.; Roy, J.; Tumber, V.; Lawrence, J.R. Effects of Erythromycin, Trimethoprim and Clindamycin on Attached Microbial Communities from an Effluent Dominated Prairie Stream. Ecotoxicol. Environ. Saf. 2016, 132, 31–39. [Google Scholar] [CrossRef]
  135. Ghoshdastidar, A.J.; Fox, S.; Tong, A.Z. The Presence of the Top Prescribed Pharmaceuticals in Treated Sewage Effluents and Receiving Waters in Southwest Nova Scotia, Canada. Environ. Sci. Pollut. Res. 2015, 22, 689–700. [Google Scholar] [CrossRef]
  136. Ni, B.-J.; Zeng, S.; Wei, W.; Dai, X.; Sun, J. Impact of Roxithromycin on Waste Activated Sludge Anaerobic Digestion: Methane Production, Carbon Transformation and Antibiotic Resistance Genes. Sci. Total Environ. 2020, 703, 134899. [Google Scholar] [CrossRef]
  137. Michael, I.; Rizzo, L.; McArdell, C.S.; Manaia, C.M.; Merlin, C.; Schwartz, T.; Dagot, C.; Fatta-Kassinos, D. Urban Wastewater Treatment Plants as Hotspots for the Release of Antibiotics in the Environment: A Review. Water Res. 2013, 47, 957–995. [Google Scholar] [CrossRef] [Green Version]
  138. Le-Clech, P. Membrane Bioreactors and Their Uses in Wastewater Treatments. Appl. Microbiol. Biotechnol. 2010, 88, 1253–1260. [Google Scholar] [CrossRef]
  139. Yang, W.; Cicek, N. Treatment of Swine Wastewater by Submerged Membrane Bioreactors with Consideration of Estrogenic Activity Removal. Desalination 2008, 231, 200–208. [Google Scholar] [CrossRef]
  140. Marti, E.; Monclús, H.; Jofre, J.; Rodriguez-Roda, I.; Comas, J.; Balcázar, J.L. Removal of Microbial Indicators from Municipal Wastewater by a Membrane Bioreactor (MBR). Bioresour. Technol. 2011, 102, 5004–5009. [Google Scholar] [CrossRef]
  141. Kovalova, L.; Siegrist, H.; Singer, H.; Wittmer, A.; McArdell, C.S. Hospital Wastewater Treatment by Membrane Bioreactor: Performance and Efficiency for Organic Micropollutant Elimination. Environ. Sci. Technol. 2012, 46, 1536–1545. [Google Scholar] [CrossRef] [Green Version]
  142. Xiao, Y.; Yaohari, H.; De Araujo, C.; Sze, C.C.; Stuckey, D.C. Removal of Selected Pharmaceuticals in an Anaerobic Membrane Bioreactor (AnMBR) with/without Powdered Activated Carbon (PAC). Chem. Eng. J. 2017, 321, 335–345. [Google Scholar] [CrossRef]
  143. Xia, S.; Jia, R.; Feng, F.; Xie, K.; Li, H.; Jing, D.; Xu, X. Effect of Solids Retention Time on Antibiotics Removal Performance and Microbial Communities in an A/O-MBR Process. Bioresour. Technol. 2012, 106, 36–43. [Google Scholar] [CrossRef]
  144. Li, B.; Zhang, T. Mass Flows and Removal of Antibiotics in Two Municipal Wastewater Treatment Plants. Chemosphere 2011, 83, 1284–1289. [Google Scholar] [CrossRef]
  145. Liu, D.; Song, K.; Xie, G.; Li, L. MBR-UV/Cl2 System in Treating Polluted Surface Water with Typical PPCP Contamination. Sci. Rep. 2020, 10, 8835. [Google Scholar] [CrossRef]
  146. Alacabey, İ. Antibiotic Removal from the Aquatic Environment with Activated Carbon Produced from Pumpkin Seeds. Molecules 2022, 27, 1380. [Google Scholar] [CrossRef] [PubMed]
  147. Yu, F.; Li, Y.; Han, S.; Ma, J. Adsorptive Removal of Antibiotics from Aqueous Solution Using Carbon Materials. Chemosphere 2016, 153, 365–385. [Google Scholar] [CrossRef] [PubMed]
  148. Tambosi, J.L.; de Sena, R.F.; Gebhardt, W.; Moreira, R.F.P.M.; José, H.J.; Schröder, H.F. Physicochemical and Advanced Oxidation Processes—A Comparison of Elimination Results of Antibiotic Compounds Following an MBR Treatment. Ozone Sci. Eng. 2009, 31, 428–435. [Google Scholar] [CrossRef]
  149. Watkinson, A.J.; Murby, E.J.; Kolpin, D.W.; Costanzo, S.D. The Occurrence of Antibiotics in an Urban Watershed: From Wastewater to Drinking Water. Sci. Total Environ. 2009, 407, 2711–2723. [Google Scholar] [CrossRef]
  150. Dolar, D.; Gros, M.; Rodriguez-Mozaz, S.; Moreno, J.; Comas, J.; Rodriguez-Roda, I.; Barceló, D. Removal of Emerging Contaminants from Municipal Wastewater with an Integrated Membrane System, MBR–RO. J. Hazard. Mater. 2012, 239–240, 64–69. [Google Scholar] [CrossRef]
  151. Dong, X.; Li, Y.; Li, D.; Liao, D.; Quin, T.; Prakash, O.; Kumar, A.; Liu, J. A New 3D 8-Connected Cd(Ii) MOF as a Potent Photocatalyst for Oxytetracycline Antibiotic Degradation. CrystEngComm 2022, 24, 6933–6943. [Google Scholar] [CrossRef]
  152. Naseer, M.N.; Jaafar, J.; Junoh, H.; Zaidi, A.A.; Kumar, M.; Alqahtany, A.; Jamil, R.; Alyami, S.H.; Aldossary, N.A. Metal-Organic Frameworks for Wastewater Decontamination: Discovering Intellectual Structure and Research Trends. Materials 2022, 15, 5053. [Google Scholar] [CrossRef]
  153. Saddique, Z.; Imran, M.; Javaid, A.; Rizvi, N.B.; Akhtar, M.N.; Iqbal, H.M.N.; Bilal, M. Enzyme-Linked Metal Organic Frameworks for Biocatalytic Degradation of Antibiotics. Catal. Lett. 2023. [Google Scholar] [CrossRef]
  154. Nosakhare Amenaghawon, A.; Lewis Anyalewechi, C.; Uyi Osazuwa, O.; Agbovhimen Elimian, E.; Oshiokhai Eshiemogie, S.; Kayode Oyefolu, P.; Septya Kusuma, H. A Comprehensive Review of Recent Advances in the Synthesis and Application of Metal-Organic Frameworks (MOFs) for the Adsorptive Sequestration of Pollutants from Wastewater. Sep. Purif. Technol. 2023, 311, 123246. [Google Scholar] [CrossRef]
  155. Liu, X.; Shan, Y.; Zhang, S.; Kong, Q.; Pang, H. Application of Metal Organic Framework in Wastewater Treatment. Green Energy Environ. 2022, in press. [Google Scholar] [CrossRef]
  156. Li, J.; Wang, H.; Yuan, X.; Zhang, J.; Chew, J.W. Metal-Organic Framework Membranes for Wastewater Treatment and Water Regeneration. Coord. Chem. Rev. 2020, 404, 213116. [Google Scholar] [CrossRef]
  157. Zhou, Y.; Yang, Q.; Zhang, D.; Gan, N.; Li, Q.; Cuan, J. Detection and Removal of Antibiotic Tetracycline in Water with a Highly Stable Luminescent MOF. Sens. Actuators B Chem. 2018, 262, 137–143. [Google Scholar] [CrossRef]
  158. Grela, A.; Kuc, J.; Bajda, T. A Review on the Application of Zeolites and Mesoporous Silica Materials in the Removal of Non-Steroidal Anti-Inflammatory Drugs and Antibiotics from Water. Materials 2021, 14, 4994. [Google Scholar] [CrossRef]
  159. Blasioli, S.; Martucci, A.; Paul, G.; Gigli, L.; Cossi, M.; Johnston, C.T.; Marchese, L.; Braschi, I. Removal of Sulfamethoxazole Sulfonamide Antibiotic from Water by High Silica Zeolites: A Study of the Involved Host–Guest Interactions by a Combined Structural, Spectroscopic, and Computational Approach. J. Colloid Interface Sci. 2014, 419, 148–159. [Google Scholar] [CrossRef]
  160. Zuo, X.; Qian, C.; Ma, S.; Xiong, J.; He, J.; Chen, Z. Removal of Sulfonamide Antibiotics from Water by High-Silica ZSM-5. Water Sci. Technol. 2019, 80, 507–516. [Google Scholar] [CrossRef]
  161. Hacıosmanoğlu, G.G.; Mejías, C.; Bueno, J.M.; Santos, J.L.; Aparicio, I.; Alonso, E. Antibiotic Adsorption by Natural and Modified Clay Minerals as Designer Adsorbents for Wastewater Treatment: A Comprehensive Review. J. Environ. Manag. 2022, 317, 115397. [Google Scholar] [CrossRef]
  162. Hacıosmanoğlu, G.G.; Arenas, M.; Mejías, C.; Martín, J.; Santos, J.L.; Aparicio, I.; Alonso, E. Adsorption of Fluoroquinolone Antibiotics from Water and Wastewater by Colemanite. Int. J. Environ. Res. Public. Health 2023, 20, 2646. [Google Scholar] [CrossRef]
  163. Iorhemen, O.; Hamza, R.; Tay, J. Membrane Bioreactor (MBR) Technology for Wastewater Treatment and Reclamation: Membrane Fouling. Membranes 2016, 6, 33. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Dissemination routes of antibiotics and drug resistance determinants in the mountain environment.
Figure 1. Dissemination routes of antibiotics and drug resistance determinants in the mountain environment.
Water 15 00975 g001
Figure 2. Natural degradation pathways of antibiotics in water environments.
Figure 2. Natural degradation pathways of antibiotics in water environments.
Water 15 00975 g002
Figure 3. Antibiotic removal efficiency by MBR and MBR-UV/Cl2 processes [143]. SMZ—sulfamethoxazole, TC—tetracycline, OTC—oxytetracycline, CIP—ciprofloxacin, OFX—ofloxacin, ERY—erythromycin, ROX—roxithromycin.
Figure 3. Antibiotic removal efficiency by MBR and MBR-UV/Cl2 processes [143]. SMZ—sulfamethoxazole, TC—tetracycline, OTC—oxytetracycline, CIP—ciprofloxacin, OFX—ofloxacin, ERY—erythromycin, ROX—roxithromycin.
Water 15 00975 g003
Figure 4. Scheme of advanced antibiotic removal techniques from wastewater.
Figure 4. Scheme of advanced antibiotic removal techniques from wastewater.
Water 15 00975 g004
Table 1. Levels of antibiotics excreted in unchanged form after administration to humans (h) or animals (a).
Table 1. Levels of antibiotics excreted in unchanged form after administration to humans (h) or animals (a).
AntibioticClassExcretion LevelReference
LevofloxacinFluoroquinolones85% (h)[45]
SulfamethoxazoleSulfonamides12% (h)[46]
TrimethoprimNitroimidazoles60% (h)[46]
MetronidazoleNitroimidazoles60–80% (h)[47]
ErythromycinMacrolides5% (h)[48]
OfloxacinFluoroquinolones80% (h)[48]
TetracyclineTetracyclines80% (h)[49]
TylosinMacrolides40% (a)[50]
OxytetracyclineTetracyclines21% (a)[51]
ChlortetracyclineTetracyclines17–75% (a)[51]
Table 2. Half-lives ( T 1 / 2 ) in days (d) or hours (h) of antibiotics in different water samples under aerobic conditions and daylight.
Table 2. Half-lives ( T 1 / 2 ) in days (d) or hours (h) of antibiotics in different water samples under aerobic conditions and daylight.
Chemical GroupsCompoundSample TypeTemp. [°C]T1/2Reference
CephalosporinCefradine 1st
Cefuroxime 2nd
Ceftriaxone 3rd
Cefepime 4th
lake water25 ± 36.3 d
3.1 d
18.7 d
2.7 d
[83]
(Amino)penicillinAmoxicillin
Ampicillin
Penicillin V
Piperacillin
ultrapure water19 ± 0.5 3.32 ± 0.61 h
3.89 ± 0.43 h
4.37 ± 0.22 h
6.99 ± 0.45 h
[85]
TetracyclineTetracycline
Oxytetracycline
Chlortetracycline
river water
river water
surface water
25 ± 14.15 d
1.82 d
3.35 h

[86]
[87]
SulfonamideSulfamethoxazolesurface water
STP effluents
river water
25 ± 1
Winter
25 ± 1
14.22 h
2.4 d
17.8 d
[87]
[88]
[86]
Sulfamethazinesurface water
river water
25 ± 11.3 d
17.3 d
[87]
[86]
FluoroquinolonesEnrofloxacinsurface water
river water
25 ± 13.34–6.75 d
8.78 d
[82]
[86]
Ciprofloxacindeionized water
kaolinite suspension
river water
19± 1
19± 1
25 ± 1
0.33 h
1.2 h
5.33 d
[89]
[89]
[86]
OfloxacinSTP effluents
river water
winter
25 ± 1
10.6 d
11.1 d
[88]
[86]
Norfloxacinriver water25 ± 15.64 d[86]
MacrolidesErythromycinsea water
river water
18± 2
25 ± 1
11.11 d
4.22 d
[90]
[86]
Roxithromycinwastewater
river water
4
25 ± 1
2.9 d
2.76 d
[91]
[86]
Clarithromycin
Azithromycin
wastewater42.9 d
4.8 d
[91]
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

Kulik, K.; Lenart-Boroń, A.; Wyrzykowska, K. Impact of Antibiotic Pollution on the Bacterial Population within Surface Water with Special Focus on Mountain Rivers. Water 2023, 15, 975. https://doi.org/10.3390/w15050975

AMA Style

Kulik K, Lenart-Boroń A, Wyrzykowska K. Impact of Antibiotic Pollution on the Bacterial Population within Surface Water with Special Focus on Mountain Rivers. Water. 2023; 15(5):975. https://doi.org/10.3390/w15050975

Chicago/Turabian Style

Kulik, Klaudia, Anna Lenart-Boroń, and Kinga Wyrzykowska. 2023. "Impact of Antibiotic Pollution on the Bacterial Population within Surface Water with Special Focus on Mountain Rivers" Water 15, no. 5: 975. https://doi.org/10.3390/w15050975

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