Open access peer-reviewed chapter

The Efficiency of Wastewater Treatment Plants for the Removal of Antibiotics

Written By

Raed S. Al-Wasify, Majid M. Alruwaili, Fahad S. Aljohani, Shimaa R. Hamed and Samar Ragab

Submitted: 24 May 2023 Reviewed: 29 May 2023 Published: 23 August 2023

DOI: 10.5772/intechopen.111999

From the Edited Volume

Water Purification - Present and Future

Edited by Magdy M.M. Elnashar and Selcan Karakuş

Chapter metrics overview

84 Chapter Downloads

View Full Metrics

Abstract

Undoubtedly domestic Wastewater Treatment Plants (WWTPs) are not designed for the removal of some pollutants such as antibiotics. This chapter summarizes the occurrence and fate of six groups of the most widely used antibiotics (β-lactams, sulfonamides, quinolones, tetracyclines, macrolides, and others) in domestic WWTPs. The literature showed that the six groups of antibiotics have been frequently detected during wastewater treatment train (influent, primary treatment, secondary treatment, tertiary treatment, effluent, and sludge treatment) of domestic WWTPs. Also, it was clear that the main removal routes of antibiotics during sewage treatment of domestic WWTPs were adsorption, biodegradation, membrane filtration, and disinfection. Domestic WWTPs cannot remove most of the antibiotics which finally enter the environment through treated effluent and sludge.

Keywords

  • antibiotics
  • adsorption
  • biodegradation
  • domestic wastewater
  • organic pollutants
  • wastewater treatment plants

1. Introduction

Nowadays, environmental researchers have extended their focus beyond classic environmental contaminants such as pesticides, Polychlorinated Biphenyls (PCBs), and dioxins [1, 2]. Antibiotics are one of the new serious environmental contaminants. Antibiotics, a group of pharmaceuticals used as a medicine and growth promoter for both humans and animals, are considered a new serious environmental contaminant due to their continuous input into the environment and persistent presence [3, 4, 5].

Antibiotics are chemicals classified depending on their nature into three main categories: natural, semi-synthetic, and synthetic. Moreover, antibiotics can be classified depending on their mode of action into two categories: bactericidal (kill microorganisms) and bacteriostatic (impede microbial growth).

In developing countries, there is an increase in the consumption of antibiotics, without any prescription from physicians, to cure the different kinds of diseases that originate from improper general hygiene and poor sanitation systems. Also, antibiotics are used widely in animal farming for the protection of animal health to maintain the high demand for animal products [6].

Boyles et al. [7] reported that the high consumption of antibiotics by humans (households, hospitals, and industry) and in veterinary results in the increasing release of unchanged active ingredients and partially metabolized antibiotics into the sewer system (directly or indirectly) which consequently reach into the domestic wastewater treatment plants [8, 9, 10].

Conventional wastewater treatment plants cannot completely remove antibiotics and these antibiotics will finally contaminate the environment through effluent or sludge [11, 12]. Therefore, local wastewater treatment plants act as one of the main pathways for antibiotics to transfer into the environment [13].

Although antibiotics residues in water are very low (ng/L to μg/L), they still draw the researchers’ attention in the whole world since these antibiotics are the main source for the occurrence and transfer of Antibiotic-Resistant Genes (ARG) and Antibiotic-Resistant Bacteria (ARB) which have serious impacts on the environment [14, 15].

During last years, simultaneous detection of trace concentrations of antibiotics in wastewater and sludge samples is no longer difficult as a result of the invention of new detection methods such as liquid chromatography-mass spectrometry, solid-phase extraction, and ultra-performance liquid chromatography-mass spectrometry with the rapid development of analytical methods such as Solid Phase Extraction (SPE), High-Performance Liquid Chromatography Mass Spectrometry (HPLC-MS/MS) and Ultra-Performance Liquid Chromatography Mass Spectrometry (UPLC-MS/MS) [16, 17].

The occurrence of antibiotics in water environments such as groundwater and surface water was summarized by previous studies [18, 19, 20, 21, 22] as well as sediments, sludge, and soil [19]. Nevertheless, most of these studies focused only on the occurrence of antibiotics in different environments with little focus on the fate of antibiotics. For example, summarizing the removal of antibiotics in sediment, water, and soil environments, instead of the elimination of antibiotics in wastewater treatment plants [21]. Also, the elimination of Personal Care Products (PPCPs) and pharmaceuticals via biodegradation and other pathways in wastewater treatment plants, with limited content on antibiotics elimination [23].

Thus, in this review, we summarize the data and information on the occurrence and fate of antibiotics in wastewater treatment plants (WWTPs) to provide the overall profile of antibiotics concentrations in influent, treatment stages, sludge, and effluent of wastewater treatment plants, and to understand the elimination routes and fate of antibiotics in WWTPs.

Advertisement

2. Occurrence of antibiotics in the aquatic environment

Domestic wastewater treatment plants receive most of the used antibiotics through the sewer network, while the rest of the antibiotics are dumped directly into rivers and streams or escape as leachate from landfills. Figure 1 summarizes the introduction pathways of antibiotics into the aquatic environment.

Figure 1.

Pathways of antibiotics in the aquatic environment, WWTP: Wastewater treatment plant, DWTP: Drinking water treatment plant.

Advertisement

3. Occurrence of antibiotics in wastewater treatment plants (WWTPs)

Antibiotics can be classified using different ways such as their chemical structure. According to the chemical structure, there are 12 different classes of antibiotics (Table 1) such as β- lactams, aminoglycosides, macrolides, glycopeptides, oxazolidinones, sulfonamides, quinolones (fluoroquinolones), polymyxins, tetracyclines, streptogramins, and others such as chloramphenicol, thiamphenicol, lincomycin, trimethoprim, and clindamycin [5, 8]. Table 1 summarizes the chemical structure of the different 12 classes of antibiotics and their mode of action, mechanism of action as well as discovery dates.

Table 1.

Different classes of antibiotics and their mode of action and mechanism of action.

In this review, Tables 27 summarize data about the occurrence of the most common antibiotics (6 classes) in influent and effluent samples of wastewater treatment plants. These antibiotics include β-lactams, quinolones, sulfonamides, macrolides, tetracyclines, and others, while Table 8 summarizes the occurrence of these antibiotics in sludge generated from wastewater treatment plants.

AntibioticConcentration (ng/L)Sampling typeTreatment technology
InfluentEffluent
Penicillin G29.0NAGrabPrimary treatment
ND-10ND-300Composite
Penicillin V20–13,800ND–2000GrabPrimary treatment
Amoxicillin190–280ND–3024 h FPCMembrane filtration/Reverse osmosis
1400–6940ND-50Composite
AmpicillinNA7GrabActivated sludge
OxacillinNA<2024 h FPCActivated sludge + chlorination
CloxacillinND–320NDGrabActivated sludge
ND-4600ND-700Composite
Cefalexin1200980GrabActivated sludge + denitrification and nitrification
NA1110–141024 h FPCChemically enhanced primary treatment
2800–64,000ND-250Composite
Cefotaxime24<12GrabActivated sludge + denitrification and nitrification
Cefaclor500–980ND–60GrabActivated sludge
500–6150ND-1800Composite
CefradineNA12GrabActivated sludge
CefazolinNA27GrabActivated sludge

Table 2.

Concentrations of β-lactams in wastewater*.

ND: Not detected, NA: Not analyzed, FPC: Flow proportional composite.


AntibioticConcentration (ng/L)Sampling typeTreatment technology
InfluentEffluent
Sulfamethoxazole360–500270–320GrabActivated sludge
255–333NDGrabMembrane filtration/Reverse osmosis
<80–674<80–3041-week FPCActivated sludge + denitrification
39031048 h composite
230–750211–86024 h FPCActivated sludge + sand filtration
N4-acetylsulfamethoxazole850–1600<20–18072 h FPCActivated sludge + sand filtration
NA690–2200Grab
Sulfathiazole2–40ND–5CompositeActivated sludge
SulfasalazineND–60ND–10GrabActivated sludge
ND-1004–150Composite
Sulphapyridine60–15040–35024 h FPCActivated sludge + sand filtration
NA12424 h composite
NA85–8827 h FPCActivated sludge + sand filtration
SulfacetamideNA64–15124 h FPCActivated sludge + chlorination
110–210ND24 h compositeActivated sludge
Sulfacholoropyridazine<30–476<30–149GrabActivated sludge
Sulfadiazine3.71.9–3.8GrabActivated sludge
SulfisoxazoleNA0.13GrabActivated sludge
Sulfadimethoxine<10–213<10 − 70GrabActivated sludge
SulfamethizoleND–710ND–10Grab
Sulfamerazine29–7312–42Grab
SulfanilamideNA26GrabActivated sludge
SulfaguanidineNA2Grab

Table 3.

Concentrations of sulfonamides in wastewater*.

ND: Not detected, NA: Not analyzed, FPC: Flow proportional composite.


AntibioticConcentration (ng/L)Sampling typeTreatment technology
InfluentEffluent
Norfloxacin11085GrabActivated sludge + denitrification and nitrification
3398524 h compositeActivated sludge
72–1747–371-week FPCActivated sludge + denitrification
NA250–54624 h FPCChemically enhanced primary treatment
Ciprofloxacin3800–4600640–720GrabActivated sludge
2106024 h compositeActivated sludge
113–3007–321-week FPCActivated sludge + denitrification
Enrofloxacin10–10010GrabActivated sludge
ND-402–50Composite
Nalidixic acidND–20055GrabActivated sludge
Ofloxacin13741GrabActivated sludge + chlorination
ND-200ND-450Composite
47011048 h composite
NA740–122024 h FPCChemically enhanced primary treatment
LomefloxacinNA<45GrabActivated sludge + chlorination + sand filtration
FlumequineNA15GrabActivated sludge
Pipemidic acidNA70Grab
541224 h composite
Flerofloxacin285.824 h compositeActivated sludge
Lomefloxacin981724 h composite
Gatifloxacin1115624 h composite
Moxifloxacin441724 h composite

Table 4.

Concentrations of quinolones in wastewater*.

ND: Not detected, NA: Not analyzed.


AntibioticConcentration (ng/L)Sampling typeTreatment technology
InfluentEffluent
Tetracycline240–79050–16024 h compositeActivated sludge
9618024 h FPCActivated sludge + denitrification and nitrification
NA683–1420GrabPrimary treatment
OxytetracyclineNA<5.0–10024 h FPCChemically enhanced primary treatment
NDND–20GrabActivated sludge
Chlortetracycline10–35NDGrabMembrane filtration/Reverse osmosis
2706024 h FPCActivated sludge + chlorination
Demeclocycline50NDGrab
DoxycyclineND–65ND–40GrabActivated sludge
20–65010–150Composite
<64–2210<64–9151-week FPCActivated sludge + denitrification

Table 5.

Concentrations of tetracyclines in wastewater*.

ND: Not detected, NA: Not analyzed, FPC: Flow proportional composite.


AntibioticConcentration (ng/L)Sampling typeTreatment technology
InfluentEffluent
Erythromycin-H2O810850GrabPrimary treatment
470520GrabActivated sludge +DN
NA10524 h composite
60–19060–11024 h FPCActivated sludge + sand filtration
NA1310–4330GrabPrimary treatment
Erythromycin71–141145–29024 h FPCTrickling filters + activated sludge + UV
83062024 h FPCActivated sludge + denitrification
TylosinND–55ND–65GrabActivated sludge
20–401–5GrabMembrane filtration/Reverse osmosis
ND-603–3400Composite
11506024 h composite
RoxithromycinND–18ND–100GrabActivated sludge
140–17510–15GrabMembrane filtration/Reverse osmosis
10–4010–3024 h FPCActivated sludge + sand filtration
753524 h compositeActivated sludge + denitrification
Oleandomycin20–1905–30GrabMembrane filtration/Reverse osmosis
ND-5ND-150Composite
Azithromycin6–539–28GrabOxidation ditch + chlorination
90–38080–40024 h FPCActivated sludge + sand filtration
ClarithromycinNA172GrabActivated sludge
NA57–32824 h FPCActivated sludge + sand filtration
NA220–32927 h FPCActivated sludge + sand filtration

Table 6.

Concentrations of macrolides in wastewater*.

ND: Not detected, NA: Not analyzed, FPC: Flow proportional composite.


AntibioticConcentration (ng/L)Sampling typeTreatment technology
InfluentEffluent
Trimethoprim213–300218–322GrabTrickling filters + activated sludge + UV
79002400GrabActivated sludge + sand filtration
NA68–8172 h FPCActivated sludge + sand filtration
210–44020–31024 h FPCActivated sludge + sand filtration
140–130066–7001-week FPCActivated sludge + denitrification
Chloramphenicol311724 h FPCOxidation ditch + UV
NA92–1050GrabPrimary treatment
NA5024 h FPCActivated sludge
ThiamphenicolNA138GrabActivated sludge
Clindamycin2–55GrabActivated sludge
20–605–70Composite
1–10ND–5GrabMembrane filtration/Reverse osmosis
Lincomycin60–8050–60GrabActivated sludge
20–5003–30Composite

Table 7.

Concentrations of others in wastewater*.

ND: Not detected, FPC: Flow proportional composite.


AntibioticConcentration (ng/L)Treatment technologyCountry or region
InfluentEffluent
Sulfonamides
Sulphapyridine281Activated sludge + sand filtrationSwitzerland
Sulfamethoxazole20NAOxidation ditch + UVChina
Sulfadimidine31NAOxidation ditch + UVChina
Sulfadimethoxine<2.0–8.1NAUSA
Sulfisoxazole<4.1–21.9NAUSA
Macrolides
Roxithromycin61–131NAActivated sludge + denitrification and nitrificationSwitzerland and Germany
Azithromycin52–1582.3Activated sludge + denitrification and nitrificationSwitzerland and Germany
Clarithromycin27–630.8Activated sludge + denitrification and nitrificationSwitzerland and Germany
Erythromycin-H2O76NAActivated sludge + chlorinationChina
Quinolones
Ofloxacin886NAOxidation ditch + UVChina
Norfloxacin301NAActivated sludge + chlorinationChina
CiprofloxacinNA1400–4800Activated sludge + denitrification and nitrificationSweden
Tetracyclines
Tetracycline<7.5–15.8NAUSA
Chlortetracycline<6.9–14.7NAUSA
DoxycyclineNA1300–1500Activated sludge + denitrification and nitrificationSweden
Others
Clindamycin3.7–15.4NAUSA
Trimethoprim21–1330.1Activated sludge + denitrification and nitrificationSwitzerland and Germany

Table 8.

Concentrations of antibiotics in sludge*.

NA: Not analyzed.


Figure 2 shows the typical structure and the treatment train of a domestic wastewater treatment plant. The treatment train begins consists of three stages including primary, secondary, and tertiary. The primary treatment stage contains preliminary treatment units (screens and sand & grit removal), an oil and grease removal unit, and primary settling tanks (not common in most WWTPs). The secondary treatment stage contains a biological treatment unit (aeration tank) using different methods such as activated sludge (AS) and trickling filters (TF), followed by secondary settling tanks (secondary clarifiers). The tertiary treatment stage contains a sand filtration unit followed by a disinfection unit using different technologies such as chlorination, ozonation, and ultraviolet (UV), and finally nutrient removal unit (for treated effluent which contains high concentrations of phosphorus and nitrogen).

Figure 2.

Typical structure of domestic wastewater treatment plant.

3.1 Presence of antibiotics in influent and effluent samples of WWTPs

3.1.1 β-Lactams

β-lactams are a group of antibiotics characterized by the presence of the β-lactam ring. The β-lactams ring is the main structure that gives the antibacterial activity for β-lactams. The variation in pharmacological properties depends on the differentiation in the side chains. β-lactams include two subclasses which are cephalosporins and penicillins [24].

The β-lactam antibiotics industry has annual sales of about $15 billion [25]; this makes up 65% of the total antibiotics market [26]. 6-Aminopenicillanic acid is the precursor to produce β-lactam antibiotics, which can be produced using free and immobilized penicillin G acylase [27]. Immobilized enzymes are preferred over the free enzyme for many reasons including product-free enzymes as reported by Elnashar [28].

In the samples collected from WWTPs (Table 2), some penicillins were detected such as ampicillin, penicillin G & V, amoxicillin, cloxacillin, and oxacillin. Penicillin V showed the highest concentrations (2000–2013,800 ng/L) in influent and effluent wastewater samples [29]. For cephalosporins, six types were detected including cefotaxime, cefalexin, cefaclor, cloxacillin, cefazolin, and cephradine [5, 30]. Generally, β-lactams, especially penicillins were rarely detected in domestic wastewater though they are used in great amounts [14, 31]. This may be attributed to the unstable nature of the β-lactam ring. The β-lactam ring can be broken by a popular bacterial enzyme group called β-lactamases [22] or can be broken through the chemical hydrolysis process [32].

3.1.2 Sulfonamides

Sulfonamides are used commonly since 1968 and consist of a large group of broad-spectrum antibiotics [5, 13, 33]. In WWTPs, 16 types of sulfonamides have been detected (Table 3). The most detected sulfonamides were sulfamethoxazole (5597–6000 ng/L) followed by sulfamethazine, sulphapyridine, and sulfadiazine, sequentially. N4-acetylsulfamethoxazole and many other N4-acetylated sulfonamides, dominant human metabolites of sulfonamides, were also detected in WWTPs and it was discovered that these metabolites can be transformed into their parent compounds [11, 34, 35].

3.1.3 Quinolones (fluoroquinolones)

Quinolones are a class of antibiotics characterized by the presence of quinolone branches as their basic structure. Fluoroquinolones are a subclass of quinolones that contains a fluorine-substituted central ring. Nalidixic acid was the first quinolone antibiotic discovered in the 1960s followed by newly developed four generations of quinolones [3, 35]. As a result of the universal extensive usage of quinolones, all generations of these antibiotics were detected worldwide in WWTPs as shown in Table 4. The most commonly detected quinolones in WWTPs were ciprofloxacin, norfloxacin, and ofloxacin [8, 12, 36]. The highest concentrations detected in influent and effluent samples were 4600 and 7870 ng/L for ciprofloxacin and ofloxacin, respectively [11, 37, 38].

3.1.4 Tetracyclines

Tetracyclines are composed of eight antibiotics (semisynthetic and natural) which inhibit the synthesis of bacterial proteins. Tetracyclines are widely used for human use, the poultry industry, and animal agriculture [39]. In WWTPs, five types of tetracyclines were detected in influent and effluent samples and tetracycline was the most detected one as shown in Table 5 [5, 14]. Doxycycline showed the highest concentration (2210 ng/L) in influent samples [40], while tetracycline showed the highest concentration (1420 ng/L) in effluent samples [37]. Tetracyclines were discovered in sludge samples with no or little biodegradation. Generally, as shown in Table 8, tetracyclines have a relatively rare presence in WWTPs because tetracyclines are used rarely by humans [41, 42].

3.1.5 Macrolides

Macrolides are a group of antibiotics characterized by the presence of a lactone ring that is substituted with alkyl, hydroxyl, and ketone groups, which inhibit the synthesis of bacterial proteins and are usually used as substitutes for penicillin [1, 34]. In WWTPs, six types of macrolides and one metabolite of erythromycin (erythromycin-H2O) were detected in influent and effluent samples. The most frequently detected macrolide was erythromycin-H2O followed by roxithromycin, clarithromycin, azithromycin, and tylosin in sequence as shown in Table 6. The lowest detected macrolides in sequence were erythromycin-H2O, roxithromycin, clarithromycin, azithromycin, and tylosin. Erythromycin-H2O showed the highest concentrations in the influent (10,025 mg/L) and in the effluent (4330 ng/L) wastewater samples [37].

3.1.6 Others

In this review, the other category of antibiotics consists of five different types including chloramphenicol, trimethoprim, lincomycin, clindamycin, and thiamphenicol. All these other antibiotics were detected in WWTPs, and trimethoprim was the most abundant and widely distributed one as shown in Table 7. Trimethoprim showed the highest concentrations in influent (7900 ng/L) and in effluent (3052 ng/L) samples, while Chloramphenicol and thiamphenicol showed the lowest concentrations (<4 to 1050 ng/L) in WWTPs [14, 41].

3.1.7 Summary

Despite that β-lactams are the most consumed antibiotic by humans, they were not detected frequently due to the unstable nature of β-lactams [21]. Many reasons affecting the significant variation of antibiotics’ concentrations in wastewater influents such as:

  1. the consumption pattern of antibiotics, which may be different in different countries and even in the same country there is a great variation in consumption patterns of antibiotics [21].

  2. different sampling procedures; grab/composite sampling results in a great variation in antibiotic concentrations since pharmaceutical loads reached WWTPs are different during the day [43].

  3. hourly and seasonal fluctuation; seasonal variation of antibiotic consumption was high between summer and winter since antibiotic consumption is two times higher than consumption in summer which affects the concentrations of antibiotics in wastewater influent samples [44, 45]. Also, during winter, antibiotics in influent samples are diluted by rain [11, 46].

  4. wastewater treatment plant scale, which means that higher concentrations of antibiotics in influent samples occurred when WWTP serves a low population and vice versa. Additionally, the concentrations of antibiotics varied in effluent depending on the different sampling methods and the applied wastewater treatment technology [5, 8].

3.2 Presence of antibiotics in sludge samples of WWTP

The analysis process for antibiotics and their transformations in sludge samples is a challenge since antibiotics in sludge have a low detectability rate as well as a low extraction pattern [47]. Therefore, the studies on the detection and presence of antibiotics in sludge are much less than those on wastewater [48]. The results showed the presence of five different classes of antibiotics in both activated and digested sludge samples as shown in Table 8.

The most abundant antibiotics were ciprofloxacin (4.8 mg/kg) in digested sludge [40] and norfloxacin (2.7 mg/kg) in activated sludge [14, 49]. Quinolones were detected mostly at the level mg/kg, while most tetracyclines, macrolides, sulfonamides, and others were detected in lower levels (μg/kg) [11]. Despite the wide usage of β-lactams antibiotics in veterinary and human medicine, all sludge samples all over the world showed no presence of β-lactams antibiotics which is mainly attributed to their poor adsorption onto sludge and their unstable characteristics [37].

Advertisement

4. Antibiotics transformation and fate in WWTPs

4.1 Antibiotics removal pathways in WWTPs

Antibiotics in WWTPs can be removed through major pathways including biodegradation, adsorption, membrane separation, and disinfection. Furthermore, there are other pathways for the removal of antibiotics including photolysis, volatilization, and hydrolysis which were eliminated since they have an inconsiderable role in the reduction of antibiotics from WWTPs. For example, β- lactams antibiotics are not stable because of the presence of β- lactam ring which easily can be hydrolyzed. Thus, β-lactams will be hydrolyzed before reaching the WWTPs.

Additionally, some researchers reported that β- lactams have a relatively long half-life due to hydrolysis at neutral pH values (same as in WWTPs) such as more than 5 days for amoxicillin [50] and 52 h for meropenem [51]. Moreover, although β-lactams can be hydrolyzed in WWTP, the contribution of the hydrolysis process for the removal of antibiotics is useless because the wastewater treatment process has a relatively short hydraulic retention time (8–20 h).

In addition to that, there are some antibiotics such as amoxicillin which were degraded by sunlight-photolysis or UV-photolysis [50], macrolides [52], quinolones [53], and tetracyclines [54, 55]. However, this degradation process has a minor significance because wastewater has high concentrations of suspended solids which inhibit the deep penetration of sunlight or UV [5, 8]. Besides, the effect of the photolysis process can be neglected since the hydraulic retention time of WWTPs is much lower than the half-life of most antibiotics in wastewater.

4.2 Antibiotics transformation and fate in conventional WWTP

4.2.1 Primary treatment units

The primary treatment stage in WWTPs consists mainly of screens and primary settling tanks. In some WWTPs, some coagulant chemicals are added in primary treatment units such as ferric ion salts, aluminum salts, or polymers called CEPT (Chemically Enhanced Primary Treatment) [56]. Many previous studies reported that the primary treatment stage in WWTPs has no significant removal for different types of antibiotics including clarithromycin, sulfamethoxazole, ofloxacin, cefalexin, erythromycin, amoxicillin, clindamycin, and cefaclor [13].

Nevertheless, other studies reported that the chemically enhanced primary treatment (CEPT) process can significantly improve the removal efficiencies of some antibiotics such as norfloxacin (67.7%), ofloxacin (55.2%), erythromycin (44.8%), sulfamethoxazole (64.0%), and roxithromycin (76.3%). This effect is due to the destroying effect of coagulants on the chemical chains of some antibiotics [11].

4.2.2 Biological treatment units

Clearly, in biological treatment units, biodegradation and adsorption processes are the main pathways for the transformation of antibiotics in WWTPs. According to the classification of antibiotics, their transformation and fate in biological treatment units can be summarized as the following:

4.2.2.1 β-Lactams

Despite that β-lactams are the most consumed antibiotics for humans and animals, they have not been detected frequently in WWTPs [14, 57], thus there was not much-published data about the fate and transformation of β-lactams in WWTPs. Junker et al. [57] studied the fate of some 14C-labeled antibiotics (benzylpenicillin and ceftriaxone) in the activated sludge process. The results showed that ceftriaxone was not totally mineralized, whereas only about 25% of benzylpenicillin was mineralized. The same results were obtained using biodegradability tests (closed bottle tests method; CBT) at much higher concentrations of β-lactams antibiotics since ceftriaxone was kept unchanged whereas benzylpenicillin was biodegraded up to 27% [58].

The differences in biodegradability between β-lactams antibiotics may be due to the differences in their chemical structures because of diverse side chains [24, 59]. Andreozzi et al. [50] performed a standard batch experiment to study the fate of amoxicillin in the activated sludge process. The results proved that the adsorption and biodegradation processes were responsible for the removal and transformation of amoxicillin.

4.2.2.2 Sulfonamides

Most researchers who studied sulfonamides in WWTPs focused on sulfamethoxazole and N4-acetylsulfamethoxazole (a metabolite of sulfamethoxazole). Batt et al. [41] and Pérez et al. [60] found that sulfonamides were biodegraded to a certain degree (low removal efficiency) in the wastewater biological treatment stage. Sulfonamides were removed during the biological treatment process with an average removal efficiency of 25% [13, 38], as well as sulfamethoxazole showed poor removal efficiency of 20% [42, 61].

Also, some studies reported the resistance of sulfonamides to different treatment processes during wastewater treatment [11, 12, 62]. Nevertheless, some other researchers reported the relatively high removal efficiency of sulfamethoxazole like 55% [34], 56% [63], 66% [64], 67% [65], and 74% [63].

The significant variation in removal efficiencies of sulfonamides during the biological treatment process can be attributed to the following reasons: first, the transformation of some metabolites such as N4-acetylsulfamethoxazole to the parent molecule (sulfamethoxazole) in the influent. Second, the mentioned removal efficiencies depended on grab or composite samples (24 h), which cannot reflect the whole treatment process [12, 66].

Thus, to avoid the previous limitations, some researchers used well-controlled laboratory reactors for studying the fate of sulfonamides and their removal pathway during the activated sludge treatment process [8, 67]. The biodegradation process of three sulfonamides at low concentrations (20 μg/L) using activated sludge reactors was studied by Pérez et al. [60] and they reported that the biodegradation process was so efficient and was able to remove the three sulfonamides for 3 days. Less than 26% of the initial antibiotics’ concentrations were present by the third day, whereas by the tenth day, the removal efficiency increased up to 93%. In addition, it was reported that some microorganisms can utilize sulfamethoxazole as a carbon and/or nitrogen source after 3 days lag phase [68]. Despite these studies proving that sulfonamides can be biodegraded, the biodegradation process takes a long time than the usual hydraulic retention time of the biological treatment process at WWTPs.

4.2.2.3 Quinolones

Batt et al. [41] and Xu et al. [12] reported that adsorption followed by biodegradation are the main pathways for the removal of quinolones during biological treatment stages at WWTPs. The removal efficiencies for norfloxacin, ofloxacin, and ciprofloxacin were 87–100%, 75–77%, and 85%, respectively [69, 70]. The adsorption mechanism of quinolones by sludge depends on electrostatic interactions between particles rather than hydrophobic forces [21, 42].

4.2.2.4 Tetracyclines

In the biological treatment process at WWTPs, adsorption is considered the main mechanism for the removal of tetracyclines [41, 71, 72]. Tetracycline (10 μg/L) was removed (up to >95%) rapidly through an adsorption mechanism during 6 h inside activated sludge units. Also, two lab-scale Sequencing Batch Reactors (SBR) were utilized to stimulate the activated sludge process (biological treatment), in these SBRs, the effect of SRT (Sludge Retention Time) and HRT (Hydraulic Retention Time) on transformation and fate of tetracycline were studied (66). The results showed that the removal efficiency of tetracycline in phase 1 (SRT = 10 days; HRT = 24 h) was 86.4 ± 8.7% and phase 2 (SRT = 10 days; HRT = 7.4 h) was 85.1 ± 5.4%, while in phase 3 (SRT = 3 days; HRT = 7.4 h) was 78.4 ± 7.1%. In phase 3, it was clear that the removal efficiency of tetracycline decreased by a reduction in SRT, which indicated that more tetracycline can be adsorbed by old sludge. In addition to that, it was reported that ferrous chloride could enhance the removal of tetracycline through precipitation due to the strong complexation between tetracyclines and ferrous ions [41].

4.2.2.5 Macrolides

All previous studies have indicated that all macrolides were not significantly eliminated, even at low concentrations, during the biological treatment process at WWTPs [10, 67].

4.2.2.6 Trimethoprim

Many studies indicated that trimethoprim was not adsorbed during the Activated Sludge (AS) process [41, 60]. The studies also proved that trimethoprim cannot easily biodegrade during AS with a short sludge retention time [11, 57]. However, Pérez et al. [60] reported that trimethoprim was completely degraded by nitrifying activated sludge with long SRT within 3 days. The nitrifying bacteria present in the nitrifying activated sludge are responsible for trimethoprim degradation since it was noticed that when the activity of nitrifying bacteria is inhibited, the elimination efficiency of trimethoprim decreased from 70 to 25% [42, 73].

4.2.3 Digestion tank

A little number of studies were carried out to study the transformation and fate of antibiotics in digestion tanks because the digestion process is not applied at most of the WWTPs all over the world. Zhang and Bing [5] conducted a two-stage anaerobic sludge digester (SRT = 30 days) to study the fate of fluoroquinolones (norfloxacin and ciprofloxacin) and they reported the stability of these fluoroquinolones. The same results were reported by Lindberg et al. [66] that ciprofloxacin and norfloxacin showed no significant removal under mesophilic sludge digesters (38°C). In contrast, Du et al. [14] utilized anaerobic mesophilic sludge digesters to study the stability of trimethoprim and sulfamethoxazole and they reported the instability of trimethoprim and sulfamethoxazole.

This may be due to that both trimethoprim and sulfamethoxazole have no significant amounts in digested sludge while their concentrations are high in activated sludge. Gartiser et al. [74] studied the biodegradation of nine antibiotics under an anaerobic digestion process (35 ± 1°C) and they reported that the biodegradation process was inefficient for all nine antibiotics except in the case of benzylpenicillin which was biodegraded after 40 days lag phase.

4.3 Antibiotics transformation and fate in advanced treatment processes

4.3.1 Filtration

To improve the quality of treated effluent, some wastewater treatment plants apply advanced treatment units such as Membrane Filtration (MF) or Sand Filtration (SF). During sand filtration, the removal efficiencies of trimethoprim and clarithromycin were 60 and 15%, respectively [8, 34]. Nakada et al. [75] also reported the same results for trimethoprim since the removal efficiency was 55.2% in the sand filtration process. However, sulfamethoxazole and sulphapyridine showed lower removal efficiencies of 26.9 and 14.6%, respectively after the sand filtration process.

Moreover, clarithromycin, azithromycin, roxithromycin, and erythromycin-H2O showed no removal at all during the sand filtration process which may be due to the presence of highly diverse and effective biofilm on the SF particles [75]. Watkinson et al. [38] reported that about 43 and 94% of total antibiotics were removed by the microfiltration process and reverse osmosis, respectively via eliminating the particles that adsorbed antibiotics on them. Also, the Nanofiltration (NF) process increased the removal efficiencies of antibiotics up to more than 95% for antibiotics such as tetracyclines [42, 76].

4.3.2 Disinfection

Table 9 summarizes the transformation and fate of eight classes of antibiotics after disinfection units at WWTPs. The disinfectants reacted fast with the antibiotics and the removal occurred after 1–27 min. In most studies, ozonation was applied for disinfection purposes. The ozonation oxidized antibiotics either directly by ozone (O3) or by hydroxyl radicals (•OH). Hydroxyl radicals are generated due to the decay of ozone. Each ozone and hydroxyl radicals have different oxidation mechanisms since O3 is selective and usually attacks the special functional groups (such as the aromatic structure or a C〓C double bond), whereas hydroxyl radicals are non-selective and react with many types of moieties [75].

AntibioticDisinfectantpHPerformance
Contact time (min)Removal efficiency (%)
β-lactams
Penicillin GOzone7.71–70
CefalexinOzone7.71>99
Sulfonamides
SulfamethoxazoleFree available chlorine7.31−95
SulphapyridineOzone7.952793.9
Fluoroquinolones
CiprofloxacinFree available chlorine7.71>99
EnrofloxacinOzone7.71>99
Macrolides
AzithromycinOzone7.952792.6
Erythromycin-H2OOzone7.952788.7
ClarithromycinOzone7.218>76
TylosinOzone7.71>99
RoxithromycinOzone7–7.58-40-4>90–99
Tetracyclines
TetracyclineOzone7.71>99
AmikacinOzone7.71−25
VancomycinOzone7.71>99
Others
TrimethoprimFree available chlorine.8.110−100
Ozone7.71>99
LincomycinOzone5.52100

Table 9.

Transformation and fate results of antibiotics in the disinfection process.

In addition, Cha et al. [77] studied the ozonation of 14 antibiotics and found that only 4 antibiotics including cephalexin, penicillin G, N(4)-acetylsulfamethoxazole, and amikacin were oxidized by hydroxyl radicals and the other 10 antibiotics reacted mainly with O3. Moreover, by using other disinfectants such as hypochlorous acid (HOCl), some antibiotics such as trimethoprim showed no degradation, while sulfamethoxazole was degraded after a reaction with HOCl [78]. Liu et al. [10] reported that there was a significant variation in reaction rates of combined chlorine and free available chlorine with antibiotics in wastewater.

Advertisement

5. Conclusion

This review provides insight into the occurrence and fate of antibiotics in domestic wastewater treatment plants. Data was collected about the occurrence of the most widely six groups of antibiotics used for human cure including β-lactams, sulfonamides, quinolones, tetracyclines, macrolides, and others in wastewater and sludge samples of wastewater treatment plants. All previously mentioned groups of antibiotics were detected in wastewater and sludge samples with varied concentrations during the different treatment stages. It was clear that most of the wastewater treatment plants do not have the ability to fully remove these antibiotics. The main removal mechanisms of these antibiotics were biodegradation, adsorption, membrane filtration, and disinfection.

References

  1. 1. Gorito AM, Ribeiro AR, Almeida CMR, Silva AMT. A review on the application of constructed wetlands for the removal of priority substances and contaminants of emerging concern listed in recently launched EU legislation. Environmental Pollution. 2017;227:428-443
  2. 2. Hernández F, Sancho JV, Ibáñez M, Guerrero C. Antibiotic residue determination in environmental waters by LC-MS. Trac-Trends in Analytical Chemistry. 2007;26(6):466-485
  3. 3. Amarasiri M, Sano D, Suzuki S. Understanding human health risks caused by antibiotic resistant bacteria (ARB) and antibiotic resistance genes (ARG) in water environments: Current knowledge and questions to be answered. Critical Reviews in Environmental Science and Technology. 2020;50(19):2016-2059
  4. 4. Hernando MD, Mezcua M, Fernández-Alba AR, Barceló D. Environmental risk assessment of pharmaceutical residues in wastewater effluents, surface waters and sediments. Talanta. 2006;69(2):334-342
  5. 5. Zhang T, Bing L. Occurrence, transformation, and fate of antibiotics in municipal wastewater treatment plants. Critical Reviews in Environmental Science and Technology. 2011;41(11):951-998
  6. 6. Van Boeckel TP, Brower C, Gilbert M, Grenfell BT, Levin SA, Robinson TP, et al. Global trends in antimicrobial use in food animals. Proceedings of the National Academy of Sciences. 2015;112:5649-5654
  7. 7. Boyles T, Naicker V, Rawoot N, Raubenheimer P, Eick B, Mendelson M. Sustained reduction in antibiotic consumption in a south African public sector hospital: Four-year outcomes from the Groote Schuur hospital antibiotic stewardship programme. South African Medical Journal. 2017;107:115-118
  8. 8. Al-Isawi R, Ray S, Scholz M. Comparative study of domestic wastewater treatment by mature vertical-flow constructed wetlands and artificial ponds. Ecological Engineering. 2017;100:8-18
  9. 9. Giger W, Alder AC, Golet EM, Kohler HPE, McArdell CS, Molnar E, et al. Occurrence and fate of antibiotics as trace contaminants in wastewaters, sewage sludges, and surface waters. Chimia. 2003;57(9):485-491
  10. 10. Liu L, Li J, Fan H, Huang X, Wei H, Liu C. Fate of antibiotics from swine wastewater in constructed wetlands with different flow configurations. International Biodeterioration & Biodegradation. 2019;140:119-125
  11. 11. Bilgehan N, Taylan D, Serdar K. Behaviour and removal of ciprofloxacin and sulfamethoxazole antibiotics in three different types of full-scale wastewater treatment plants: A comparative study. Water, Air, & Soil Pollution. 2021;232:127
  12. 12. Xu WH, Zhang G, Zou SC, Li XD, Liu YC. Determination of selected antibiotics in the Victoria harbour and the Pearl River, South China using high-performance liquid chromatography-electrospray ionization tandem mass spectrometry. Environmental Pollution. 2007;145(3):672-679
  13. 13. Deng Y, Li B, Zhang T. Bacteria that make a meal of sulfonamide antibiotics: Blind spots and emerging opportunities. Environmental Science & Technology. 2018;52(7):3854-3868
  14. 14. Du J, Hongxia Z, Sisi L, Huaijun X, Yan W, Jingwen C. Antibiotics in the coastal water of the south yellow sea in China: Occurrence, distribution and ecological risks. Science of the Total Environment. 2017;595:521-527
  15. 15. Zhang T, Zhang M, Zhang XX, Fang HHP. Tetracycline resistance genes and tetracycline resistant lactose-fermenting enterobacteriaceae in activated sludge of sewage treatment plants. Environmental Science & Technology. 2009;43(10):3455-3460
  16. 16. Chang XS, Meyer MT, Liu XY, Zhao Q, Chen H, Chen J, et al. Determination of antibiotics in sewage from hospitals, nursery and slaughterhouse, wastewater treatment plant and source water in Chongqing region of three gorge reservoir in China. Environmental Pollution. 2010;158(5):1444-1450
  17. 17. Plósz BG, Leknes H, Liltved H, Thomas KV. Diurnal variations in the occurrence and the fate of hormones and antibiotics in activated sludge wastewater treatment in Oslo. Norway. Science of The Total Environment. 2010;408(8):1915-1924
  18. 18. Baquero F, Martínez JL, Cantón R. Antibiotics and antibiotic resistance in water environments. Current Opinion in Biotechnology. 2008;19(3):260-265
  19. 19. Díaz-Cruz MS, Barceló D. LC-MS2 trace analysis of antimicrobials in water, sediment and soil. TrAC Trends in Analytical Chemistry. 2005;24(7):645-657
  20. 20. Kemper N. Veterinary antibiotics in the aquatic and terrestrial environment. Ecological Indicators. 2008;8(1):1-13
  21. 21. Kümmerer K. Antibiotics in the aquatic environment: A review. Part I. Chemosphere. 2009;75(4):417-434
  22. 22. Sukul P, Spiteller M. Fluoroquinolone antibiotics in the environment. Reviews of Environmental Contamination and Toxicology. 2007;191:131-162
  23. 23. Onesios KM, Yu JT, Bouwer EJ. Biodegradation and removal of pharmaceuticals and personal care products in treatment systems: A review. Biodegradation. 2009;20:441-466
  24. 24. Dodd MC, Buffle MO, Von Gunten U. Oxidation of antibacterial molecules by aqueous ozone: Moiety-specific reaction kinetics and application to ozone-based wastewater treatment. Environmental Science & Technology. 2006;40(6):1969-1977
  25. 25. Elnashar MM, Wahba MI, Amin MA, El-Diwany A. Application of Plackett–Burman screening design to the modeling of grafted alginate–carrageenan beads for the immobilization of penicillin G acylase. Journal of Applied Polymers Science. 2014;131:40295. DOI: 10.1002/app.40295
  26. 26. Elnashar MM, Mohamed EH, Ghada EA. Grafted carrageenan gel disks and beads with polyethylenimine and glutaraldehyde for covalent immobilization of penicillin G acylase. Journal of Colloid Science and Biotechnology. 2013;2(1):27-33
  27. 27. Elnashar MM, Yassin AM, Kahil T. Novel thermally and mechanically stable hydrogel for enzyme immobilization of penicillin G acylase via covalent technique. Journal of Applied Polymer Sciences. 2008;109:4105-4111
  28. 28. Elnashar MM. Immobilized molecules using biomaterials and nanobiotechnology. Journal of Biomaterials and Nanobiotechnology. 2010;1:61-76. DOI: 10.4236/JBNB.2010.11008
  29. 29. Watkinson AJ, Murby EJ, Kolpin DW, Costanzo SD. The occurrence of antibiotics in an urban watershed: From wastewater to drinking water. Science of the Total Environment. 2009;407(8):2711-2723
  30. 30. Lin AYC, Yu TH, Lin CF. Pharmaceutical contamination in residential, industrial, and agricultural waste streams: Risk to aqueous environments in Taiwan. Chemosphere. 2008;74(1):131-141
  31. 31. Bailón-Pérez MI, García-Campaña AM, Cruces-Blanco C, Iruela MD. Trace determination of β-lactam antibiotics in environmental aqueous samples using off-line and on-line preconcentration in capillary electrophoresis. Journal of Chromatography A. 2008;1185(2):273-280
  32. 32. Hirsch R, Ternes T, Haberer K, Kratz KL. Occurrence of antibiotics in the aquatic environment. Science of the Total Environment. 1999;225(1–2):109-118
  33. 33. Littlefield NA, Sheldon WG, Allen R, Gaylor DW. Chronic toxicity/carcinogenicity studies of sulphamethazine in Fischer 344/N rats: Two-generation exposure. Food and Chemical Toxicology. 1990;28(3):157-167
  34. 34. Göbel A, Thomsen A, McArdell CS, Joss A, Giger W. Occurrence and sorption behavior of sulfonamides, macrolides, and trimethoprim in activated sludge treatment. Environmental Science and Technology. 2005;39(11):3981-3989
  35. 35. Xiao Y, Chang H, Jia A, Hu JY. Trace analysis of quinolone and fluoroquinolone antibiotics from wastewaters by liquid chromatography electrospray tandem mass spectrometry. Journal of Chromatography A. 2008;1214(1–2):100-108
  36. 36. Seifrtová M, Pena A, Lino CM, Solich P. Determination of fluoroquinolone antibiotics in hospital and municipal wastewaters in Coimbra by liquid chromatography with a monolithic column and fluorescence detection. Analytical and Bioanalytical Chemistry. 2008;391(3):799-805
  37. 37. Minh TB, Leung HW, Loi IH, Chan WH, So MK, Mao JQ, et al. Antibiotics in the Hong Kong metropolitan area: Ubiquitous distribution and fate in Victoria Harbor. Marine Pollution Bulletin. 2009;58(7):1052-1062
  38. 38. Watkinson AJ, Murby EJ, Costanzo SD. Removal of antibiotics in conventional and advanced wastewater treatment: Implications for environmental discharge and wastewater recycling. Water Research. 2007;41(18):4164-4176
  39. 39. Yang SW, Cha JM, Carlson K. Simultaneous extraction and analysis of 11 tetracycline and sulfonamide antibiotics in influent and effluent domestic wastewater by solid-phase extraction and liquid chromatography-electro spray ionization tandem mass spectrometry. Journal of Chromatography A. 2005;1097(1–2):40-53
  40. 40. Lindberg RH, Wennberg P, Johansson MI, Tysklind M, Andersson BAV. Screening of human antibiotic substances and determination of weekly mass flows in five sewage treatment plants in Sweden. Environmental Science and Technology. 2005;39(10):3421-3429
  41. 41. Batt AL, Kim S, Aga DS. Comparison of the occurrence of antibiotics in four full-scale wastewater treatment plants with varying designs and operations. Chemosphere. 2007;68(3):428-435
  42. 42. Bergeron S, Boopathy R, Nathaniel R, Corbin A, LaFleur G. Presence of antibiotic resistant bacteria and antibiotic resistance genes in raw source water and treated drinking water. International Biodeterioration & Biodegradation. 2015;102:370-374
  43. 43. Joss A, Keller E, Alder AC, Göbel A, McArdell CS, Ternes T, et al. Removal of pharmaceuticals and fragrances in biological wastewater treatment. Water Research. 2005;39(14):3139-3152
  44. 44. Elseviers MM, Ferech M, Vander Stichele RH, Goossens H. The ESAC project group. Antibiotic use in ambulatory care in Europe (ESAC data 1997–2002): Trends, regional differences and seasonal fluctuations. Pharmacoepidemiology and Drug Safety. 2007;16(1):115-123
  45. 45. McArdell CS, Molnar E, Suter MJF, Giger W. Occurrence and fate of macrolide antibiotics in wastewater treatment plants and in the Glatt Valley watershed, Switzerland. Environmental Science and Technology. 2003;37(24):5479-5486
  46. 46. Li B, Zhang T, Xu ZY, Fang HHP. Rapid analysis of 21 antibiotics of multiple classes in municipal wastewater using ultra performance liquid chromatography-tandem mass spectrometry. Analytica Chimica Acta. 2009;645(1–2):64-72
  47. 47. O’Connor S, Aga DS. Analysis of tetracycline antibiotics in soil: Advances in extraction, clean-up, and quantification. TrAC Trends in Analytical Chemistry. 2007;26(6):456-465
  48. 48. Kim SC, Carlson K. Quantification of human and veterinary antibiotics in water and sediment using SPE/LC/MS/MS. Analytical and Bioanalytical Chemistry. 2007;387(4):1301-1315
  49. 49. Golet EM, Xifra I, Siegrist H, Alder AC, Giger W. Environmental exposure assessment of fluoroquinolone antibacterial agents from sewage to soil. Environmental Science and Technology. 2003;37(15):3243-3249
  50. 50. Andreozzi R, Caprio V, Ciniglia C, De Champdoré M, Lo Giudice R, Marotta R, et al. Antibiotics in the environment: Occurrence in Italian STPs, fate, and preliminary assessment on algal toxicity of amoxicillin. Environmental Science and Technology. 2004;38(24):6832-6838
  51. 51. Al-Ahmad A, Daschner FD, Kümmerer K. Biodegradability of cefotiam, ciprofloxacin, meropenem, penicillin G, and sulfamethoxazole and inhibition of wastewater bacteria. Archives of Environmental Contamination and Toxicology. 1999;37(2):158-163
  52. 52. Vione D, Feitosa-Felizzola J, Minero C, Chiron S. Photo transformation of selected human-used macrolides in surface water: Kinetics, model predictions and degradation pathways. Water Research. 2009;43(7):1-9
  53. 53. Burhenne J, Ludwig M, Nikoloudis P, Spiteller M. Photolytic degradation of fluoroquinolone carboxylic acids in aqueous solution. I. Primary photoproducts and half-lives. Environmental Science and Pollution Research. 1997;4(1):10-15
  54. 54. Jiao SJ, Meng SR, Yin DQ, Wang LH, Chen LY. Aqueous oxytetracycline degradation and the toxicity change of degradation compounds in photoirradiation process. Journal of Environmental Sciences-China. 2008a;20(7):806-813
  55. 55. Jiao SJ, Zheng SR, Yin DQ, Wang LH, Chen LY. Aqueous photolysis of tetracycline and toxicity of photolytic products to luminescent bacteria. Chemosphere. 2008b;73(3):377-382
  56. 56. Sui Q, Huang J, Deng SB, Yu G, Fan Q. Occurrence and removal of pharmaceuticals, caffeine and DEET in wastewater treatment plants of Beijing, China. Water Research. 2010;44(2):417-426
  57. 57. Junker T, Alexy R, Knacker T, Kümmerer K. Biodegradability of 14C-labeled antibiotics in a modified laboratory scale sewage treatment plant at environmentally relevant concentrations. Environmental Science and Technology. 2006;40(1):318-324
  58. 58. Alexy R, Kümpel T, Kümmerer K. Assessment of degradation of 18 antibiotics in the closed bottle test. Chemosphere. 2004;57(6):505-512
  59. 59. Kümmerer K, Al-Ahmad A. Biodegradability of the antitumor agents 5-fluorouracil, cytarabine, and gemcitabine: Impact of the chemical structure and synergistic toxicity with hospital effluent. Acta Hydrochimica et Hydrobiolgica. 1997;25(4):166-172
  60. 60. Pérez S, Eichhorn P, Aga DS. Evaluating the biodegradability of sulfamethazine, sulfamethoxazole, sulfathiazole, and trimethoprim at different stages of sewage treatment. Environmental Toxicology and Chemistry. 2005;24(6):1361-1367
  61. 61. Brown KD, Kulis J, Thomson B, Chapman TH, Mawhinney DB. Occurrence of antibiotics in hospital, residential, and dairy, effluent, municipal wastewater, and the Rio Grande in New Mexico. Science of the Total Environment. 2006;366(2–3):772-783
  62. 62. Miao XS, Bishay F, Chen M, Metcalfe CD. Occurrence of antimicrobials in the final effluents of wastewater treatment plants in Canada. Environmental Science and Technology. 2004;38(13):3533-3541
  63. 63. Radjenovic J, Petrovic M, Barceló D. Analysis of pharmaceuticals in wastewater and removal using a membrane bioreactor. Analytical and Bioanalytical Chemistry. 2007;387(4):1365-1377
  64. 64. Clara M, Strenn B, Gans O, Martinez E, Kreuzinger N, Kroiss H. Removal of selected pharmaceuticals, fragrances and endocrine disrupting compounds in a membrane bioreactor and conventional wastewater treatment plants. Water Research. 2005;39(19):4797-4807
  65. 65. Carballa M, Omil F, Lema JM, Llompart M, García-Jares C, Rodríguez I, et al. Behavior of pharmaceuticals, cosmetics and hormones in a sewage treatment plant. Water Research. 2004;38(12):2918-2926
  66. 66. Lindberg RH, Olofsson U, Rendahl P, Johansson MI, Tysklind M, Andersson BAV. Behavior of fluoroquinolones and trimethoprim during mechanical, chemical, and active sludge treatment of sewage water and digestion of sludge. Environmental Science and Technology. 2006;40(3):1042-1048
  67. 67. Joss A, Zabczynski S, Göbel A, Hoffmann B, Löffler D, McArdell CS, et al. Biological degradation of pharmaceuticals in municipal wastewater treatment: Proposing a classification scheme. Water Research. 2006;40(8):1686-1696
  68. 68. Drillia P, Dokianakis SN, Fountoulakis MS, Kornaros M, Stamatelatou K, Lyberatos G. On the occasional biodegradation of pharmaceuticals in the activated sludge process: The example of the antibiotic sulfamethoxazole. Journal of Hazardous Materials. 2005;122(3):259-265
  69. 69. Radjenovic J, Petrovic M, Barceló D. Fate and distribution of pharmaceuticals in wastewater and sewage sludge of the conventional activated sludge (CAS) and advanced membrane bioreactor (MBR) treatment. Water Research. 2009;43(3):831-841
  70. 70. Vieno NM, Tuhkanen T, Kronberg L. Analysis of neutral and basic pharmaceuticals in sewage treatment plants and in recipient rivers using solid phase extraction and liquid chromatography-tandem mass spectrometry detection. Journal of Chromatography A. 2006;1134(1–2):101-111
  71. 71. Kim S, Eichhorn P, Jensen JN, Weber AS, Aga DS. Removal of antibiotics in wastewater: Effect of hydraulic and solid retention times on the fate of tetracycline in the activated sludge process. Environmental Science and Technology. 2005;39(15):5816-5823
  72. 72. Li B, Zhang T. Biodegradation and adsorption of antibiotics the activated sludge process. Environmental Science and Technology. 2010;44(9):3468-3473
  73. 73. Batt AL, Kim S, Aga DS. Enhanced biodegradation of iopromide and trimethoprim in nitrifying activated sludge. Environmental Science and Technology. 2006;40(23):7367-7373
  74. 74. Gartiser S, Urich E, Alexy R, Kümmerer K. Anaerobic inhibition and biodegradation of antibiotics in ISO test schemes. Chemosphere. 2007;66(10):1839-1848
  75. 75. Nakada N, Shinohara H, Murata A, Kiri K, Managaki S, Sato N, et al. Removal of selected pharmaceuticals and personal care products (PPCPs) and endocrine-disrupting chemicals (EDCs) during sand filtration and ozonation at a municipal sewage treatment plant. Water Research. 2007;41(19):4373-4382
  76. 76. Koyuncu I, Arikan OA, Wiesner MRC. Rice, removal of hormones and antibiotics by nanofiltration membranes. Journal of Membrane Science. 2008;309(1–2):94-101
  77. 77. Cha JM, Yang S, Carlson KH. Trace determination of β-lactam antibiotics in surface water and urban wastewater using liquid chromatography combined with electrospray tandem mass spectrometry. Journal of Chromatography A. 2006;1115(1–2):46-57
  78. 78. Dodd MC, Huang CH. Transformation of the antibacterial agent sulfamethoxazole in reactions with chlorine: Kinetics mechanisms, and pathways. Environmental Science and Technology. 2004;38(21):5607-5615

Written By

Raed S. Al-Wasify, Majid M. Alruwaili, Fahad S. Aljohani, Shimaa R. Hamed and Samar Ragab

Submitted: 24 May 2023 Reviewed: 29 May 2023 Published: 23 August 2023