Next Article in Journal
The Assessment of Green Water Based on the SWAT Model: A Case Study in the Hai River Basin, China
Previous Article in Journal
Recent Glacier Mass Balance and Area Changes from DEMs and Landsat Images in Upper Reach of Shule River Basin, Northeastern Edge of Tibetan Plateau during 2000 to 2015
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Disinfection Methods for Swimming Pool Water: Byproduct Formation and Control

1
Water Treatment and Management Consultancy, 2289 ED Rijswijk, The Netherlands
2
IHE Delft, Institute for Water Education, 2611 AX Delft, The Netherlands
3
Department of Water Management, Faculty of Civil Engineering and Geosciences, Delft University of Technology, 2600 GA Delft, The Netherlands
4
Strategic Centre, Waternet, 1096 AC Amsterdam, The Netherlands
*
Author to whom correspondence should be addressed.
Water 2018, 10(6), 797; https://doi.org/10.3390/w10060797
Submission received: 29 April 2018 / Revised: 10 June 2018 / Accepted: 12 June 2018 / Published: 16 June 2018
(This article belongs to the Section Water Quality and Contamination)

Abstract

:
This paper presents a comprehensive and critical comparison of 10 disinfection methods of swimming pool water: chlorination, electrochemically generated mixed oxidants (EGMO), ultraviolet (UV) irradiation, UV/chlorine, UV/hydrogen peroxide (H2O2), UV/H2O2/chlorine, ozone (O3)/chlorine, O3/H2O2/chlorine, O3/UV and O3/UV/chlorine for the formation, control and elimination of potentially toxic disinfection byproducts (DBPs): trihalomethanes (THMs), haloacetic acids (HAAs), haloacetonitriles (HANs), trihaloacetaldehydes (THAs) and chloramines (CAMs). The statistical comparison is carried out using data on 32 swimming pools accumulated from the reviewed studies. The results indicate that O3/UV and O3/UV/chlorine are the most promising methods, as the concentration of the studied DBPs (THMs and HANs) with these methods was reduced considerably compared with chlorination, EGMO, UV irradiation, UV/chlorine and O3/chlorine. However, the concentration of the studied DBPs including HAAs and CAMs remained much higher with O3/chlorine compared with the limits set by the WHO for drinking water quality. Moreover, the enhancement in the formation of THMs, HANs and CH with UV/chlorine compared with UV irradiation and the increase in the level of HANs with O3/UV/chlorine compared with O3/UV indicate the complexity of the combined processes, which should be optimized to control the toxicity and improve the quality of swimming pool water.

1. Introduction

Chlorination methods for the disinfection of swimming pool water lead to the formation of potentially toxic disinfection byproducts (DBPs) such as trihalomethanes (THMs), haloacetic acids (HAAs), haloacetonitriles (HANs), trihaloacetaldehyde (THAs) and chloramines (CAMs) (Table A1 and Table A2: given as Appendix A) [1,2,3,4]. With this awareness, the demand for alternatives to chlorine disinfection has increased. The alternative methods aimed at improving the quality of swimming pool water are: electrochemically generated mixed oxidants (EGMO) [3,5,6,7,8], ultraviolet (UV) irradiation [9,10,11,12], UV-based advanced oxidation processes (UV-based AOPs) such as UV/H2O2 [13], ozone (O3) [14,15,16], and ozone-based AOPs such as O3/H2O2 and O3/UV [17,18,19]. The use of O3 and UV irradiation for the disinfection of swimming pool water has been adopted in some cases, though together with chlorine for the provision of a residual disinfectant [20]. However, chlorination methods for disinfection are widely used in practice [21,22]. AOPs are still in the research and development phase for the disinfection of swimming pool water [22].
The efficiency of the ozonation process can be increased by combining O3 with H2O2 and UV such as O3/H2O2 and O3/UV. Although substantial research has been done on the application of O3/H2O2 and O3/UV processes for the disinfection of drinking water [23,24,25,26,27,28], a limited number of studies is available on the disinfection of swimming pool water. The use of O3/H2O2 and O3/UV is reported by Glauner et al. [17], and O3/UV by Kristensen et al. [18] and Cheema et al. [19]. Similarly, significant research has been done on the use of UV-based AOPs (UV/H2O2) for the disinfection of drinking water [23,25,26,27,28,29,30], but in swimming pool water only by Spiliotopoulou et al. [13]. Moreover, the studies conducted on the use of O3/H2O2, O3/UV and UV/H2O2 for the treatment of swimming pool water only investigated few DBPs such as THMs and HANs.
It is recognized that the conditions for the treatment of drinking water are completely different from the treatment of swimming pool water. Compared with drinking water, swimming pool water DBPs have their own distinct characteristics due to the different nature of organic precursors [20,31,32] and continuous loading of dissolved organic carbon (DOC) and dissolved organic nitrogen (DON), which are released by swimmers. All these factors add an additional complication to the disinfection and toxicological safety of swimming pool water [1]. For instance, the reported level of total organic carbon (TOC) in tap water (used as source water) and swimming pool water was in the range of 0.3–1.4 and 0.5–7.0 mg·L−1, respectively. Similarly, the level of nitrate–nitrogen (NO3–N) was higher in swimming pool water (6.6–23.8 mg·L−1) compared with tap water (1.1–1.9 mg·L−1) [3]. Consistently, the reported level of total nitrogen (TN) in drinking water was lower (0.1–0.3 mg·L−1) compared with swimming pool water (3.6–12.3 mg·L−1) [7,33]. The temperature of swimming pool water is generally higher (25–35 °C) compared with drinking water used as source water (1.0–23 °C) [34], which is an important parameter to restrain the continuous anthropogenic pollutant release [32] and increases the demand for higher doses of disinfectants due to their higher rates of decay, for instance, in the case of chlorination [35,36]. The reported level of free residual chlorine (FRC) in tap water and swimming pool water was in the range of 0.03–0.57 and 0.24–1.4 mg·L−1, respectively. Analogous to that the pH of tap water (6.8–7.8) was lower than swimming pool water (7.6–8.2) [3]. Thus, the studies of drinking water disinfection cannot be replicated to optimize the treatment conditions of swimming pool water.
In the last few years research has been increasing on alternative methods for the disinfection of swimming pool water, as is evident from several published studies. Although the comparative analysis of different disinfection methods within studies is available, it has not been done between the studies. For instance, a comparison between chlorination, EGMO and O3/chlorine [3,6]; between chlorination and EGMO [5,7,8]; between chlorination, UV irradiation and UV/chlorine [10,37]; between chlorination and UV/chlorine [9,11,12]; between chlorination, UV irradiation, UV/chlorine, UV/H2O2 and UV/H2O2/chlorine [13]; between chlorination and O3/chlorine [14,15,16,38]; between UV irradiation and O3/UV [18]; between chlorination, O3/chlorine, O3/H2O2/chlorine and O3/UV/chlorine [17]; and between chlorination, UV/chlorine, O3/chlorine, O3/UV and O3/UV/chlorine [19]. A comprehensive and critical review of performance and a comparison of all the disinfection methods is lacking. Moreover, all of the studies reported in this review considered different DBPs (see Appendix A: Table A3, Table A4, Table A5, Table A6 and Table A7). A thorough analysis is lacking for making sound inferences about the effects of disinfection methods on certain types of DBPs. Therefore, more studies are needed to conduct a synthesis of the recent developments of disinfection methods used for disinfection of swimming pool water and draw informed conclusions about the performance potential of different disinfection methods for the formation and control of DBPs.
The main objective of this paper is to study the effects of different disinfection methods on the formation, control and elimination of DBPs, and to evaluate the differences between distinct methods.
The treatment performance of 10 disinfection methods is analysed in this paper for the formation and/or elimination of DBPs such as THMs, HAAs, HANs, THAs and CAMs. The examined disinfection methods are: chlorination, EGMO, UV irradiation, UV/chlorine, UV/H2O2, UV/H2O2/chlorine, O3/chlorine, O3/H2O2/chlorine, O3/UV and O3/UV/chlorine.
The research focused on the following specific questions:
  • What are the concentrations of DBPs with different methods of disinfection?
  • How do the tested disinfection methods affect the formation and/or elimination of DBPs?
  • Is the performance of different disinfection methods statistically different for the formation of DBPs?
  • Which disinfection method is best to improve the quality of swimming pool water?
The presented results contribute to the scientific knowledge on the use of various disinfection methods to combat the toxicity and improve the quality of swimming pool water. In addition, future research needs are also outlined in this paper.

2. Methods of Disinfection

Different methods used for the disinfection of swimming pool water are summarized in Table 1.

2.1. Chlorination

Chlorination is the most commonly used method of disinfection for the prevention of waterborne diseases and inactivation of pathogenic microorganisms in swimming pools [20]. The chemicals used for chlorination are given in Table 1. However, the use of chlorination in pools has many drawbacks, such as the presence of chlorine-resistant microorganisms such as Cryptosporidium parvum and Giardia lambia [39] and the formation of potentially toxic DBPs [1,40]. Past research identified more than 100 DBPs in pool water samples [16,36]. Among the known carbonaceous DBPs (C-DBPs), the most common are THMs, HAAs, THAs; among the nitrogenous DBPs (N-DBPs), the most common are HANs and CAMs [1,2,3,4]. The formation and distribution of DBPs depends on several factors such as water source, bromide ion (Br) concentration, chlorine dose and FRC, TOC, TN, temperature and pH [4,15,22,34,41]. The level of FRC is of major concern [15,35,36], as it plays a vital role in the formation of DBPs. It is recommended to maintain the FRC in the range of 0.8–2.0 and 0.8–3.0 mg·L−1 in the case of indoor and outdoor pools, respectively [34].

2.2. EGMO

In the EGMO technique an electric current (240–400 V) is passed through a salt brine solution (3000–6000 mg·L−1) to produce oxidants. The primary oxidant produced is chlorine in the form of hypochlorous acid (HOCl) [42]. It has been suggested that oxidants other than chlorine, such as O3, chlorine dioxide, H2O2 and hydroxyl radical (OH), are also produced [42]. However, several studies have demonstrated the drawbacks that chlorine in the form of HOCl is the primary oxidant produced and other oxidants have not been detected at measureable levels [3,43,44,45]. Dowd [46] found that EGMO produced the same levels of THMs in model and real waters as free chlorine. Considering that, with EGMO disinfection the toxicity concerns are similar to those of typical chlorine disinfection. The overall disinfectant dosage of EGMO technique is also quantified in terms of FRC [42,47,48].

2.3. UV Irradiation

UV irradiation is commonly applied as a secondary disinfection process in chlorinated pools, and has been found to be very effective for the control of Cryptosporidium pavrum and Giardia lamblia [49,50]. The UV-based processes are cost-competitive with chlorine to improve the quality of swimming pool water and air [9,51]. The use of UV irradiation in the form of low-pressure UV (LPUV) and medium-pressure UV (MPUV) with and without post-chlorination has been explored by some recent studies [10,11,12,18]. Furthermore, the effect of different doses of UV irradiation was also investigated to control the formation of DBPs [10,12,19]. The available research reported the formation of N-DBPs (HANs) with UV irradiation and UV/chlorine [10,11,12,13,19].

2.4. UV-Based AOPs

The oxidizing ability of the UV/H2O2 process may be attributed to the formation of OH produced by UV irradiation of H2O2 as shown in Equation (1):
H2O2 + hν → 2OH
This process requires a relatively high dose of H2O2 and much longer UV irradiation time compared with the O3/UV process [52]. However, the dosage of H2O2 in the UV/H2O2 process needs to be optimized. An excess of H2O2 showed a strongly negative effect as a radical scavenger because of its ability to react with the OH produced during the decomposition process and form a less reactive HO2, as shown in Equations (2) and (3). On the other hand, a low concentration of H2O2 results in insufficient formation of OH and leads to a slower oxidation rate [52].
H2O2 + OH → H2O + HO2
HO2 + OH → H2O + O2
The effectiveness of the UV/H2O2 process with and without post-chlorination for the formation and/or control of DBPs (THMs and HANs) in swimming pool water has been investigated by Spiliotopoulou et al. [13].

2.5. Ozonation

The use of O3 has increased for disinfection due to its high oxidation potential (E0 = 2.07 V) compared to chlorine (E0 = 1.36 V) (Table 2). However, O3 as a residual disinfectant is unsuitable in swimming pool water, as it readily vaporizes and decomposes. In addition, it is toxic and heavier than air, which leads to adverse health effects [20]. Due to a lack of residuals and the relatively high dose requirement, O3 disinfection is usually followed by deozonation before water enters the pool, and the addition of chlorine (O3/chlorine) [14,15,16,17].
O3 reacts with a variety of organic and inorganic compounds in an aqueous solution as molecular O3 or through the formation of OH induced by O3 decomposition. Despite its effectiveness in treating several organic compounds, the accumulation of refractory compounds is of major concern as it impedes the mineralization of the organic matter present in water. Some compounds are found to be refractory to the ozonation process [53,54,55]. Therefore, O3 is combined with other oxidants to improve the efficiency.

2.6. Ozone-Based AOPs

The efficiency of the ozonation process was increased by combining O3 with UV irradiation and H2O2 such as O3/UV and O3/H2O2 [17,18,19]. By O3/UV and O3/H2O2 processes, OH was generated and accelerated to give it a higher oxidation potential (E0 = 2.86 V) than O3 (E0 = 2.07 V) (Table 2). OH attacks most organic compounds promptly and non-selectively. Thus, the degradation of organic compounds is facilitated by both O3 and OH. Consequently, the elimination of a variety of very persistent substances can be achieved [52].
Photolysis of O3 in the presence of water generates OH as shown in Equation (4):
O3 + hν + H2O → 2OH + O2
The decomposition of organic matter may proceed in three different ways: by O3, by direct UV photolysis, and by photolysis of O3, which generates OH.
The practicality of the O3/UV process with and without post-chlorination for the formation and/or control of THMs and HANs in swimming pool water has been investigated by Kristensen et al. [18] and Cheema et al. [19].
The interaction between O3 and H2O2 leads to the generation of OH, as shown in Equation (5):
H2O2 + 2O3 → 2OH + 3O2
The feasibility of the O3/H2O2 process with and without post-chlorination for the formation and/or control of THMs in swimming pool water has been investigated by Glauner et al. [17].

3. Methodology

The comparative performance and impact analysis is mainly based on the information compiled from the available studies so far, which used different methods for disinfection of swimming pool water. These were 18 studies published in peer-reviewed journals with different methods of disinfection from nine countries. Some studies only reported descriptive statistics (means, standard deviations and ranges), and these values carry the accumulated effect of many pools. Therefore, the studies only reporting descriptive statistics could not be used to conduct statistical analysis to draw informed comparison. Consequently, individual pools were selected from all the available studies where such records were provided. In total, 32 pools were selected from the 14 published research studies with different methods for disinfection. The novel database compiled to conduct this analysis is given as Table A3, Table A4, Table A5, Table A6 and Table A7. The treatment performance of 10 disinfection methods was analysed in this study. These were chlorination, EGMO, UV irradiation, UV/chlorine, UV/H2O2, UV/H2O2/chlorine, O3/chlorine, O3/H2O2/chlorine, O3/UV and O3/UV/chlorine. Different parameters such as pool location (indoor and outdoor), disinfectant dose, method of detection, temperature, pH, TOC, TN, THMs, HAAs, HANs, THAs and CAMs were also presented for the comparison of different disinfection methods. These parameters were gathered from the reviewed studies or estimated using the information available in those studies.
Firstly, the effects of disinfection methods on the formation and/or elimination of DBPs are discussed based on all the available studies. Then, statistical analyses are carried out and descriptive statistics are computed for some parameters where adequate data were available based on 32 individual pools. Finally, the comparison among different methods is done with one-way ANOVA for the significance and z-Test for the comparison of means.

4. Results and Discussion

4.1. Effect of Disinfection Methods on DBPs Formation

The effects of different methods are discussed on the most documented species of different DBPs.

4.1.1. Effect on THMs

According to the study by Lee et al. [6], the concentration of TCM with chlorination, EGMO and O3/chlorine was 41, 27 and 29 μg·L−1, respectively. However, the levels of brominated THMs, such as BDCM, DBCM and TBM, were higher with EGMO (9.8, 9.1 and 19 μg·L−1, respectively) compared with chlorination (3.0, 0.5 and <0.2 μg·L−1, respectively) and O3/chlorine (2.4, 0.2 and <0.2 μg·L−1, respectively). The high levels of brominated THMs are attributed to the presence of Br from NaCl salt used in the EGMO process, which is mostly from seawater and contains Br. However, Kanan [7] observed the high level of TCM (207 μg·L−1) with EGMO compared with chlorination (119 μg·L−1) (Table A3).
In another comparative study of chlorination, EGMO and O3/chlorine the concentration of TCM was 21, 15 and 7.0 μg·L−1, respectively, indicating the decrease in the formation of TCM with O3/chlorine [3]. In some other studies, it is reported that the concentration of different categories of DBPs in swimming pool water disinfected with O3/chlorine was lower than the pool disinfected with chlorine. This indicates the efficacy of O3/chlorine due to its high oxidation potential compared with chlorination (Table 2); thus the degradation of persistent organic substances is possible [14,15]. For instance, Hang et al. [15] reported a higher concentration of TCM, BDCM and DBCM (220, 202 and 3.8 μg·L−1, respectively) with chlorination compared with 141, 106 and 2.0 μg·L−1, respectively in the case of O3/chlorine (Table A3).
In the study of different doses of O3 in clean pool water (TOC 1.5 mg·L−1), the formation of TCM increased from 25 to 60 μg·L−1 with the first three subsequent doses (2 mg·L−1 of O3 each) and then with further addition decreased to 45 μg·L−1, but still remained higher compared with chlorination (25 μg·L−1) [14] (Table A3). Daiber et al. [16] also reported almost twice the level of TCM with O3/chlorine (31 μg·L−1) compared with chlorination (17 μg·L−1). Chlorination of polluted pool water (TOC 2.4 mg·L−1) resulted in elevated levels of TCM (35 μg·L−1) compared with initial concentrations (5 μg·L−1) before chlorination. With the addition of O3 (0.7–3.4 mg·L−1), the concentration of TCM decreased from 35 to 18 μg·L−1 and with increasing O3 dosage (100 mg·L−1) it decreased to 7.0 μg·L−1. The increase in the formation of TCM with initial dose of O3 in the clean pool water is likely due to decomposition of O3 to OH and radical oxidation of fraction of DOC present in the pool water, which has low reactivity towards chlorine. The radical oxidation of that fraction of DOC enhances its reactivity to chlorine, and consequently the formation of TCM. The decrease in the concentration of TCM with a fourth and fifth dose is due to mineralization of that fraction of DOC with increasing oxidation [14]. Cheema et al. [19] observed a similar trend in the formation of THMs with different doses of O3, as reported by Hansen et al. [14] (Table A3).
The previous research gives limited and conflicting information regarding the effects of UV-based treatment on water chemistry in the presence of residual chlorine and the occurrence of THMs in swimming pool water. For example, Beyer et al. [9] reported the lower concentration of THMs with the use of UV/chlorine. On the other hand, Cassan et al. [37] observed a higher concentration of THMs with UV irradiation compared with chlorination and UV/chlorine. In further contrast, Kristensen et al. [18] observed no effect on THM concentrations in a swimming pool treated with chlorination and UV irradiation (Table A3). Some recent studies reported that THMs were not formed directly by exposure to UV irradiation but in the post-chlorination (UV/chlorine) [10,12,13,19]. This might be due to the fact that the reactivity of organic precursors present in the pool water towards chlorine increases with UV treatment. Consequently, the activated organic matter reacts with chlorine and THMs are formed. For instance, the concentration of TCM decreased from 16 μg·L−1 with chlorination to 13 μg·L−1 with UV irradiation, but increased with UV/chlorine to 65 μg·L−1 [13] (Table A3).
Furthermore, the comparison between LPUV and MPUV irradiation without post-chlorination demonstrates the efficiency of MPUV irradiation for decreasing the concentration of TCM [18] and with post-chlorination suggests the effectiveness of LPUV irradiation [10,11]. For instance, the concentration of TCM was 75 μg·L−1 with chlorination and 60 and 52 μg·L−1 with MPUV/chlorine and LPUV/chlorine, respectively [11] (Table A3). The increase in the levels of TCM was attributed to the production of chlorine radical (Cl) through the photolysis of FRC, which further reacted with organic matter to form TCM. On the other hand, the concentration of DBCM and TBM was less with MPUV irradiation [37]. This lower concentration of brominated species of THMs was due to their broad absorption band with a maximum at 220 nm; MPUV lamps are more efficient for their photo-degradation compared with LPUV lamps [56] (cited in [10]). The progressive transformation of TBM into TCM and BDCM by substitution of bromine atom to chlorine atom might increase the concentration of TCM with MPUV irradiation [37].
Moreover, the effect of different doses of UV irradiation followed by chlorine indicated that the level of THMs increased with increasing the dose of UV irradiation [10,12,19]. For instance, the reported concentration of TCM was 43 μg·L−1 with chlorination and 78 and 100 μg·L−1 with the UV dose of 2.35 and 4.7 J·cm−2, respectively [10]. Similar to the prior study, the concentration of TCM shows the increase from 12 μg·L−1 with chlorination to 32 μg·L−1 with UV/chlorine (2.1 J·cm−2), which remained higher with the UV dose up to 19 J·cm−2 but decreased (10 μg·L−1) with the UV dose of 47.5 J·cm−2 [19] (Table A3). This decrease in the level of TCM with increasing UV dose might be due to the decrease in the DOC level by oxidation at a very high UV dose and resulted in a lower amount of organic precursors available for reaction [19].
The efficacy of UV-based AOPs, such as UV/H2O2 with and without post-chlorination, was studied by Spiliotopoulou et al. [13]. The study results show that the concentration of THM was lower with UV/H2O2 compared with UV/chlorine. For example, the concentration of TCM was 65 and 9.1 μg·L−1 with UV/chlorine and UV/H2O2, respectively. However, the concentration of TCM increased with post-chlorination (UV/H2O2/chlorine) to 36 μg·L−1 (Table A3).
The effectiveness of ozone-based AOPs was first studied by Glauner et al. [17]. They reported the effective removal of TOC with O3/UV and O3/H2O2 compared with ozonation. The contact time of 3 min between oxidant and pool water in AOPs was sufficient for the increased elimination efficiency of TOC compared with 20 min contact time in case of ozonation. However, ozonation showed a decreased THMs formation potential due to selective oxidation compared with OH reactions in AOPs, which are non-specific and produce activated compounds suitable for THMs formation in the post-chlorination step. As a result, the formation potential of THMs increased with O3/UV/chlorine and O3/H2O2/chlorine processes compared with O3/chlorine. Nevertheless, the concentration of THMs is not reported with different methods of disinfection. On the other hand, Kristensen et al. [18] observed the decrease in concentration of TCM from 53 μg·L−1 with UV irradiation to 35 μg·L−1 with O3/UV. Analogous to the previous research, Cheema et al. [19] observed the positive effect of O3/UV on the decrease of THMs compared with separate UV irradiation and ozonation processes. The concentration of TCM was 35, 25, 5.0 μg·L−1 with UV irradiation, O3/chlorine and O3/UV, respectively. Still, the formation of TCM was increased (13 μg·L−1) in the post-chlorination step (O3/UV/chlorine) [19] (Table A3).

4.1.2. Effect on HAAs

In the comparative studies of chlorination, EGMO and O3/chlorine, the concentration of TCAA was 156, 97 and 17 μg·L−1, respectively [3]. Conversely, Kanan [7] and Yeh et al. [8] observed that the concentration of DCAA and TCAA was higher in EGMO disinfected pools compared with the chlorinated pools. For instance, the concentration of DCAA in chlorinated and EGMO disinfected indoor pools was 1233 μg·L−1 and 2400 μg·L−1, respectively, and the concentration of TCAA in chlorinated and EGMO disinfected indoor pools was 1153 μg·L−1 and 2600 μg·L−1, respectively [8] (Table A4).
In another study, the concentration of BCAA in chlorinated pools and O3/chlorine disinfected pools was 874 and 424 μg·L−1, respectively. The concentration of all species of HAAs in chlorinated pools was much higher compared with O3/chlorine disinfected pools [15]. This indicates that with O3/chlorine the decomposition of persistent organic substances is achievable. Contrarily, Daiber et al. [16] reported a very high level of DCAA and TCAA with O3/chlorine (343 and 1865 μg·L−1, respectively) compared with chlorination (89 and 65 μg·L−1, respectively) (Table A4).
UV exposure followed by post-chlorination did not significantly affect the formation of HAAs. The formation of DCAA and TCAA remained almost stable with chlorination and different doses of UV irradiation [10]. The concentration of DCAA was 196 μg·L−1 with chlorination and 201 and 197 μg·L−1 with the UV dose of 2.35 and 4.7 J·cm−2, respectively. The concentration of TCAA was 118 μg·L−1 with chlorination and 119 and 106 μg·L−1 with the UV dose of 2.35 and 4.7 J·cm−2, respectively (Table A4). UV irradiation followed by post-chlorination significantly reduced the level of DBAA and TBAA compared with chlorination, though not with different doses of UV irradiation [12], which indicated that brominated HAAs are photolysed by UV treatment [57]. The level of DBAA was 80 μg·L−1 with chlorination and 15 and 20 μg·L−1 with the UV dose of 2.35 and 4.7 J·cm−2, respectively. The level of TBAA was 45 μg·L−1 with chlorination and 2.5 and 5.0 μg·L−1 with the UV dose of 2.35 and 4.7 J·cm−2, respectively [12] (Table A4).
Unfortunately, the concentration of HAAs was not reported in the available studies with the use of UV-based AOPs (UV/H2O2 and UV/H2O2/chlorine) and ozone-based AOPs (O3/H2O2/chlorine, O3/UV and O3/UV/chlorine).

4.1.3. Effect on HANs

Lee et al. [3] reported that the concentration of DCAN with chlorination, EGMO and O3/chlorine, was 3.9, 3.8 and 1.3 μg·L−1, respectively. The concentrations of brominated HANs were much higher in EGMO disinfected pools compared with chlorinated pools. For instance, the concentration of BCAN and DBAN was 3.5 and 2.6 μg·L−1, respectively with EGMO, and 0.8 and 0.5 μg·L−1, respectively with chlorination, which might be due to the presence of Br from NaCl salt used in the EGMO process as salt used is mostly from seawater and contains Br. As reported by Lee et al. [3], Kanan [7] observed a similar trend in the formation of DCAN and BCAN with EGMO and chlorination. Analogous to the prior research, Hang et al. [15] also reported a higher concentration of DCAN (9.2 μg·L−1) with chlorination compared with (5.3 μg·L−1) O3/chlorine (Table A5).
In a study of different doses of O3 in the clean pool water (TOC 1.5 mg·L−1), the formation of DCAN increased from 2.2 to 4.2 μg·L−1 with the five subsequent doses (2 mg·L−1 of O3 each) [14] (Table A5). Daiber et al. [16] also reported a higher level of DCAN with O3/chlorine (14 μg·L−1) compared with chlorination (9.4 μg·L−1). Chlorination of polluted pool water (TOC 2.4 mg·L−1) resulted in elevated levels of DCAN (7.0 μg·L−1) compared with initial concentrations (2.5 μg·L−1) before chlorination. With the addition of O3 (0.7–3.4 mg·L−1), the concentration of DCAN decreased from 7.0 to 4.0 μg·L−1; with increasing O3 dosage (100 mg·L−1), it decreased to 2.5 μg·L−1, which was equal to the initial concentration [14]. Cheema et al. [19] observed a similar trend in the formation of DCAN with different doses of O3 as reported by Hansen et al. [14] for clean pool water. The concentration of DCAN and BCAN increased from 1.8 and 0.2 μg·L−1 to 2.5 and 0.4 μg·L−1, respectively (Table A5).
The available research reported the formation of HANs with UV irradiation and UV/chlorine [10,11,12,13,19]. For instance, Spiliotopoulou et al. [13] observed that the concentration of DCAN increased from 1.9 μg·L−1 with chlorination to 4.9 μg·L−1 with UV irradiation, which further increased to 5.4 μg·L−1 with UV/chorine (Table A5). The use of LPUV and MPVU irradiation followed by chlorination showed that the concentrations of DCAN (8.0 μg·L−1) with chlorination and LPUV/chlorine (7.7 μg·L−1) were not significantly different. However, the concentration of DCAN with MPUV/chorine increased to 10 μg·L−1 [11]. Contrarily, a significant increase in the level of DCAN from 8.0 μg·L−1 with chlorination to 15 μg·L−1 with LPUV/chlorine was observed by Cimetiere and De Laat [10] (Table A5).
The concentration of DCAN and DBAN was higher with UV irradiation but remained almost stable when increasing the UV dose [10,12,19]. A significant increase in the level of DCAN was observed from 8.0 μg·L−1 with chlorination to 15 and 17 μg·L−1 with the UV dose of 2.35 and 4.7 J·cm−2, respectively, showing that the level remained almost similar with different doses of UV irradiation [10]. Similarly, the concentration of DCAN increased from 1.8 μg·L−1 with chlorination to 4.0 μg·L−1 with UV/chlorine (2.1 J·cm−2), which remained stable at 4.0 μg·L−1 with the UV dose of 47.5 J·cm−2 [19]. The level of DBAN significantly increased from 12 μg·L−1 with chlorination to 28 and 35 μg·L−1 with the UV dose of 2.35 and 4.7 J·cm−2, respectively [12] (Table A5).
The use of the UV/H2O2 process with and without post-chlorination contributed to a negligible decrease in the concentration of HANs compared with UV/chlorine [13]. The study results showed that the concentration of DCAN was 5.4, 3.6 and 3.1 μg·L−1 with UV/chlorine, UV/H2O2, and UV/H2O2/chlorine, respectively (Table A5).
The level of DCAN was lower with O3/UV (3.0 μg·L−1) compared with UV irradiation (4.5 μg·L−1) but higher compared with O3/chlorine (2.5 μg·L−1). This might be due to the fact that UV irradiation increased chlorine demand, which favours the formation of DCAN. The chlorine demand is subsequently removed with the ozonation of the UV treated pool water. This leads to a decrease in the level of DCAN. However, the level of DCAN increased in the post-chlorination step (O3/UV/chlorine) to 3.2 μg·L−1 compared with O3/UV (3.0 μg·L−1) [19] (Table A5).

4.1.4. Effect on THAs

The concentration of CH with chlorination, EGMO and O3/chlorine was 17, 10 and 3.6 μg·L−1, respectively [3]. The lower level of CH with O3/chlorine (101 μg·L−1) compared with chlorination (165 μg·L−1) was also observed by Daiber et al. [16] (Table A6). However, the formation of CH increased with UV/chlorine compared with chlorination that increased further with increasing the dose of UV irradiation [10]. For instance, the concentration of CH increased from 237 μg·L−1 with chlorination to 261 and 275 μg·L−1 with UV dose of 2.35 and 4.7 J·cm−2, respectively (Table A6). This indicates that the use of O3/chlorine for disinfection of swimming pool water is effective to overcome this possibly carcinogenic DBP compared with chlorination, EGMO and UV/chlorine.
The studies conducted to control the formation of toxic DBPs in swimming pool water with the use of UV-based AOPs (UV/H2O2 and UV/H2O2/chlorine) and ozone-based AOPs (O3/H2O2/chlorine, O3/UV and O3/UV/chlorine) did not explore the concentration of CH.

4.1.5. Effect on CAMs

The concentration of TCAM remained unchanged with chlorination (370 μg·L−1) and MPUV/chlorine (350 μg·L−1). After the installation of LPUV/chlorine, the concentration of TCAM increased (450 μg·L−1) compared with chlorination [11] (Table A7). This might be due to fast re-formation of TCAM from the reaction of photolysis products of TCAM with chlorine, and also a higher concentration of FRC with the application of LPUV/chlorine (free chlorine promotes TCAM formation). On the other hand, while comparing LPUV and MPUV irradiation, the concentration of TCAM was lower after the inclusion of MPUV irradiation than after inclusion of LPUV irradiation. This indicates the photodecay efficiency of CAMs using MPUV irradiation compared with LPUV irradiation at the same dose [11,51].
The concentration of CAMs is not significantly reduced with the use of O3/chlorine compared with chlorination in swimming pools [38], although O3 is a strong oxidizing agent and should oxidize CAMs to reduce the concentration. The concentration of tCAMs was 1470 and 1310 μg·L−1 in chlorinated and O3/chlorine disinfected pools, respectively [38] (Table A7).
The concentration of CAMs was not examined by the studies that used UV-based AOPs (UV/H2O2 and UV/H2O2/chlorine) and ozone-based AOPs (O3/H2O2/chlorine, O3/UV and O3/UV/chlorine).

4.2. Comparison of Different Disinfection Methods for DBPs Formation

The results of ANOVA and z-Test for comparison of means are shown in Table 3, indicating the statistical significance or non-significance of observed difference among different methods of disinfection. The quantitative differences among different methods for the studied water quality parameters are discussed below and are substantiated by the corresponding figures.
As seen in Table 3, only a few DBP species could be examined (TCM, DCAA, TCAA, DCAN and CH), as sufficient data were not available in other cases. For similar reasons, the comparison of presented species was not possible with all the disinfection methods.
THMs: The formation of TCM with EGMO was much higher (mean and standard deviation: 85 ± 56 μg·L−1) compared with chlorination (44 ± 55 μg·L−1) as well as UV irradiation (38 ± 23 μg·L−1) and UV/chlorine (47 ± 37 μg·L−1), although only significantly different from UV irradiation (Figure 1 and Table 3). Similarly, the formation of TCM with O3/chlorine (39 ± 32 μg·L−1) is less than with chlorination, UV/chlorine and EGMO, but only significantly different from EGMO (Figure 1 and Table 3). However, the level of TCM with O3/UV/chlorine (16 ± 4.1 μg·L−1) is significantly reduced compared with chlorination, EGMO, UV irradiation, UV/chlorine and O3/chlorine (Figure 1 and Table 3). The level of TCM with UV/H2O2, UV/H2O2/chlorine and O3/UV was 11 ± 2.8, 31 ± 7.8 and 20 ± 21 μg·L−1, respectively (Figure 1).
HAAs: The concentration of DCAA and TCAA with EGMO is more than twice (1373 ± 1908 and 816 ± 805 μg·L−1, respectively) compared with chlorination (619 ± 633 and 470 ± 553 μg·L−1, respectively), though means are not significantly different (Figure 2 and Table 3). The concentration of DCAA and TCAA with UV/chlorine was 218 ± 109 and 124 ± 42 μg·L−1, respectively, which is significantly lower compared with EGMO (Figure 2 and Table 3). However, the concentration of DCAA and TCAA with O3/chlorine was 272 ± 101 and 943 ± 1305 μg·L−1, respectively, which is comparable with UV/chlorine in case of DCAA (Figure 2). The level of HAAs with UV/H2O2, UV/H2O2/chlorine, O3/H2O2/chlorine, O3/UV and O3/UV/chlorine methods is not reported in the literature. Furthermore, the level of TCAA reduced with UV/chlorine and met the guidelines set by WHO for drinking water quality [58], but the level of DCAA remained much higher (Table A2). Therefore, to control the formation of DCAA, which is possibly a human carcinogen [59] (Table A2), the application of UV-based AOPs (UV/H2O2 and UV/H2O2/chlorine) and ozone-based AOPs (O3/H2O2/chlorine, O3/UV and O3/UV/chlorine) needs to be investigated.
HANs: The average level of DCAN with EGMO is much higher (17 ± 13 μg·L−1) compared with chlorination (7.9 ± 6.9 μg·L−1) as well as UV/chlorine (9.9 ± 8.3 μg·L−1) though did not exhibit a significant difference (Figure 3 and Table 3). Similarly, the mean value of DCAN is almost halved with O3/chlorine (4.0 ± 2.9 μg·L−1) compared with chlorination but not significantly different (Figure 3 and Table 3). Furthermore, the level of DCAN with O3/UV (3.0 μg·L−1) and O3/UV/chlorine (3.6 ± 0.5 μg·L−1) is considerably reduced, although increased in the post-chlorination step but is significantly lower compared with chlorination (Figure 3 and Table 3). Moreover, the level of DCAN with EGMO and UV/chlorine is significantly different compared with O3/chlorine and O3/UV/chlorine; however, O3/chlorine and O3/UV/chlorine do not demonstrate a significant difference (Figure 3 and Table 3). Based on the limited evidence, the level of DCAN was 3.8 ± 1.6, 2.9 ± 1.0 and 2.6 ± 0.7 μg·L−1 with UV irradiation, UV/H2O2 and UV/H2O2/chlorine, respectively (Figure 3).
THAs: The mean concentration of CH with UV/chlorine (309 ± 130 μg·L−1) is higher compared with chlorination (253 ± 139 μg·L−1), though not significantly different (Figure 4 and Table 3). Based on the limited evidence, the average level of CH with O3/chlorine is 101 μg·L−1. The level of CH with UV/H2O2, UV/H2O2/chlorine, O3/H2O2/chlorine, O3/UV and O3/UV/chlorine methods is not reported in the literature. The level of CH is much higher compared with the guidelines set by WHO for drinking water quality [58] (Table A2). To control the formation of CH, which is possibly a human carcinogen [59] (Table A2), the application of UV-based AOPs (UV/H2O2 and UV/H2O2/chlorine) and ozone-based AOPs (O3/H2O2/chlorine, O3/UV and O3/UV/chlorine) needs to be investigated.
CAMs: The concentration of TCAM with UV/chlorine (400 ± 71 μg·L−1) is almost equivalent with chlorination (370 μg·L−1) (Figure 5). However, the concentration of TCAM with UV/H2O2, UV/H2O2/chlorine, O3/chlorine, O3/H2O2/chlorine, O3/UV and O3/UV/chlorine methods is not reported in the literature. The level of tCAMs with chlorination and O3/chlorine is also comparable with the values of 1745 ± 389 and 1600 ± 410 μg·L−1, respectively. Furthermore, the level of tCAMs remained much higher than the guidelines set by WHO for swimming pool water [20] (Table A2). Therefore, to control the formation of tCAMs, which is a very toxic DBP, the application of UV-based AOPs (UV/H2O2 and UV/H2O2/chlorine) and ozone-based AOPs (O3/H2O2/chlorine, O3/UV and O3/UV/chlorine) need to be investigated.
Based on our review, the developments in disinfection methods and corresponding concentrations of DBPs in swimming pool water are summarized in Figure 6. This could serve as a quick guide for a scientific reference and the application of emerging disinfection methods to control the formation of toxic DBPs.

4.3. Major Governing Factors for DBPs Formation

Many studies on chlorinated pools have shown that the formation and distribution of DBPs depends on several factors such as source water, Br concentration, chlorine dose and FRC, TOC, TN, temperature and pH [4,7,12,15,22,34,41,60].
The higher temperature leads to more sweat production, which is an organic precursor from anthropogenic inputs [22]. The organic precursors can be natural organic matter (NOM) from the source water used to fill the pool and many other anthropogenic inputs such as sweat, urine, lotions, cosmetics, sunscreens and soap residuals [2,3,20,22,61] as well as skin lipids [32]. Since the higher temperature accelerates the consumption of FRC, higher doses of chlorine are required to ensure FRC in swimming pool water [35,36]. The formation of DBPs is correlated with FRC, therefore, the higher temperature promoted the formation of DBPs [7,33,34,62].
TOC and TN are other dominant factors influencing the formation of DBPs [3,63]. In swimming pool water the continuous loading of DOC and DON is obtained from swimmers. Chu and Nieuwenhuijsen [64] observed that DOC was significantly increased with the number of swimmers, ranging from 3.3 to 13 mg·L−1. In a study of two outdoor swimming pools, it was estimated that on average 1.09 g DOC per person is brought into swimming pool water [65] (cited in [61]). Manasfi et al. [4] attributed that the higher level of DBPs in freshwater pools was due to more bathers compared with seawater pools. However, Peng et al. [61] concluded that the introduction of anthropogenic pollutants and consequent DBPs formation in swimming pool water cannot be predicted simply from the number of visitors. The actual DBPs formation can be estimated with the content of organic matter in the pool water. Thus, DOC is proved to be a suitable parameter to predict THMs production [61]. DON leads to the formation of N-DBPs (HANs and CAMs) [21,36,66]. Some amino acids (nitrogen-containing compounds) such as histidine present in sweat and urine favour the formation of HAAs (C-DBPs) during chlorination [7,67]. The highest level of HAAs could be due to their less volatile nature compared with other DBPs (e.g., THMs) [3]. Therefore, HAAs are more likely to remain in the pool water after their formation [21,22]. On the other hand, turbulence caused by the movement of swimmers could influence the release of volatile DBPs (e.g., THMs) into the air [68,69]. In addition, the very high concentrations of HAAs are likely due to bather organic loads in swimming pool water, which tend to preferentially form HAAs rather than THMs. Furthermore, HAAs are highly soluble in water and do not degrade in the presence of high FRC [33].
Unlike temperature and TOC, which have positive correlations with most of the studied DBPs, pH has both positive and negative relationships with DBPs. This mixed impact makes pH management a more complicated task. However, most of the research indicated that at pH < 7.0 the formation of THMs decreases; the formation of HAAs remains constant but the level of HANs increases, and the level of TCAM is drastically increased [62].
The type of source water also plays a major role for the formation of DBPs [21]. For instance, seawater contains higher levels of Br compared with freshwater (tap, surface or ground) and leads to the formation of brominated species of DBPs [4,12], which are more toxic compared with their chlorinated analogues [16,70].
This review highlighted the formation of DBPs with different methods of disinfection, thus indicating the occurrence of DBPs in swimming pool water. Different parameters such as pool location (indoor and outdoor), type of source water, disinfectant dose, method of detection, temperature, pH, TOC and TN were also presented for the comparison of different disinfection methods. The values of these parameters, given in Table A3, Table A4, Table A5, Table A6 and Table A7, were comparable in most cases when different technologies were compared while using same or similar swimming pools. For instance, the study conducted by Hang et al. [15] reported values of temperature and pH for chlorinated pool and O3/chlorine pool were 28 °C and 7.1–7.4, respectively; both pools were indoor, and used the tap water as source water. Similarly, the values of temperature, pH and TOC in chlorinated, UV irradiation and UV/chlorine disinfected pools were 27–28 °C, 7.2 and 1.8–1.9 mg·L−1, respectively in case of study by Cassan et al. [37]. Analogous to that, Cimetiere and De Laat [10] reported the similar level of temperature, pH and TOC in indoor chlorinated pools as well as with different dosage of UV irradiation for disinfection of indoor swimming pools (Table A1). Furthermore, for chlorination, UV irradiation, UV/chlorine, UV/H2O2 and UV/H2O2/chlorine the reported values of temperature, pH and TOC, and type of source water were same [13]. Therefore, a comparative evaluation of 10 disinfection methods carried out in this review could provide reliable assessment of their performance, as many other governing factors were not much different in individual case studies.

4.4. Advantages and Disadvantages of Disinfection Methods for Swimming Pool Water

The critical examination of the reviewed literature revealed that each methods has advantages as well as drawbacks. Thus, a thorough understanding should be established to advance research and make informed decisions on application of these methods. Based on the reviewed studies, some advantages and disadvantages of the studied methods are synthesized in Table 4.

5. Future Research Needs

  • In case of EGMO disinfection, the toxicity concerns are similar as in case of chlorination because both often results in harmful DBPs, which in many cases could be much higher than the guidelines set by WHO for drinking water quality. Since these methods are considered cost-effective, further research is needed to improve these methods.
  • The level of HAAs, THAs and CAMs with UV/H2O2, UV/H2O2/chlorine, O3/H2O2/chlorine O3/UV and O3/UV/chlorine is not reported in the literature. Furthermore, the level of TCAA reduced with UV/chlorine and met the guidelines set by WHO for drinking water quality, but the level of DCAA and CH remained much higher. Therefore, to control the formation of DCAA and CH, which are possible human carcinogens, and CAMs, which are very toxic, the application of UV-based AOPs (UV/H2O2 and UV/H2O2/chlorine), O3/chlorine and ozone-based AOPs (O3/UV and O3/UV/chlorine) need exploration.
  • The enhancement in the formation of THMs, HANs and CH in the post-chlorination step (UV/chlorine) compared with UV irradiation, the increase in THMs with UV/H2O2/chlorine compared with UV/H2O2 and the increase in the level of HANs with O3/UV/chlorine compared with O3/UV, indicate the complexity of the combined process, which needs to be optimized.
  • The application of O3/H2O2 is only tested for THMs and even the levels of THMs are not presented in the literature. Therefore, the future applications of this method should investigate the formation and/or elimination of toxic DBPs.
  • The optimization in the post-chlorination step and the application of UV-based AOPs (UV/H2O2), O3/chlorine and ozone-based AOPs (O3/H2O2 and O3/UV) for the disinfection of swimming pool water need further attention to control the toxicity and enhance the quality of swimming pool water.

6. Conclusions

Chlorination of swimming pool water leads to the formation of potentially toxic DBPs including THMs, HAAs, HANs, THAs and CAMs. With this knowledge, the demand for alternatives to chlorine disinfection has increased. Alternative methods tested for their potential to improve the quality of swimming pool water are: EGMO, UV irradiation, UV-based AOPs (UV/H2O2), O3 and ozone-based AOPs (O3/H2O2 and O3/UV). However, these methods are still in the research and development phase. The specific conclusions drawn from this research are:
  • In the case of THMs and HANs, the level of TCM and DCAN is almost twice with EGMO compared with chlorination. Similarly, the level of TCM and DCAN is higher with UV/chlorine compared with chlorination, UV irradiation, O3/chlorine, O3/UV and O3/UV/chlorine. Although the level of TCM is significantly reduced with O3/UV/chlorine compared with chlorination, UV irradiation, UV/chlorine, O3/chlorine and O3/UV, the level of DCAN increased in the post-chlorination step (O3/UV/chlorine) compared with O3/UV process.
  • The concentration of DCAA and TCAA (representatives of HAAs) with EGMO is more than twice compared with chlorination. However, the level of DCAA and TCAA reduced more than half with UV/chlorine compared with chlorination. Furthermore, the concentration of DCAA with O3/chlorine was much less than with chlorination. On the other hand, the level of CH (THAs) with UV/chlorine and chlorination is not significantly different.
  • The comparative studies for the formation of CAMs with different methods of disinfection are very limited in number. Thus, based on limited evidence, the formation of TCAM with UV/chlorine is comparable with chlorination. Similarly, the level of tCAMs with chlorination and O3/chlorine is not much different.
  • The comparative studies of LPUV and MPUV irradiation without post-chlorination indicated the effectiveness of MPUV to reduce the formation of THMs but with post-chlorination the efficacy of LPUV/chlorine was emphasized. On the other hand, to reduce the level of HANs the efficiency of LPUV/chlorine and MPUV/chlorine is not clear because of contradictory information.
  • The ozone-based AOPs, O3/UV and O3/UV/chlorine, significantly reduced THMs and HANs, and, thus, stand out as the most promising disinfection methods, though more research is needed to validate this hypothesis, especially related to effects on other DBPs that have not yet been studied by these methods.

Author Contributions

H.I. developed the concept of the paper, conducted review and statistical analysis, and wrote the manuscript. I.M. contributed to concept development, assisted in literature compilation and provided comments on the written manuscript. J.P.v.d.H. further improved the concept, structure, contents and writing of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Table A1. Types and species of DBPs studied in this review.
Table A1. Types and species of DBPs studied in this review.
DBPs TypeDBPs SpeciesAbbreviationChemical Formula
Trihalomethanes
(THMs)
Trichloromethane (chloroform)TCMCHCl3
BromodichloromethaneBDCMCHBrCl2
DibromochloromethaneDBCMCHBr2Cl
Tribromomethane (bromoform)TBMCHBr3
Haloacetic acids
(HAAs)
Monochloroacetic acidMCAACH2ClCOOH
Dichloroacetic acidDCAACHCl2COOH
Trichloroacetic acidTCAACCl3COOH
Monobromoacetic acidMBAACH2BrCOOH
Dibromoacetic acidDBAACHBr2COOH
Bromochloroacetic acidBCAACHBrClCOOH
Bromodichloroacetic acidBDCAACBrCl2COOH
Dibromochloroacetic acidDBCAACBr2ClCOOH
Tribromoacetic acidTBAACBr3COOH
Haloacetonitriles
(HANs)
DichloroacetonitrileDCANCHCl2CN
TrichloroacetonitrileTCANCCl3CN
BromochloroacetonitrileBCANCHBrClCN
DibromoacetonitrileDBANCHBr2CN
ChloroacetonitrileCANCH2ClCN
BromoacetonitrileBANCH2BrCN
Trihaloacetaldehydes
(THAs)
Chloral hydrateCHCCl3CH(OH)2 or C2H3Cl3O2
Chloramines
(CAMs)
MonochloramineMCAMNH2Cl
DichloramineDCAMNHCl2
TrichloramineTCAMNCl3
Table A2. Carcinogenic group classification and WHO guidelines for DBPs.
Table A2. Carcinogenic group classification and WHO guidelines for DBPs.
CompoundCarcinogenic Group US EPA, IRIS [59]WHO Guidelines-Upper Limits (μg·L−1) WHO [20,58] *
ChloroformB2300
BromodichloromethaneB260
DibromochloromethaneC100
BromoformB2100
Total trihalomethanes100
Monochloracetic acid20
Dichloroacetic acidB250
Trichloroacetic acidB2200
DichloroacetonitrileD20
DibromoacetonitrileD70
Chloral hydrateC10
Total chloramines<200
Note: Group B2: Probable human carcinogen (sufficient data from animal studies), Group C: Possible human carcinogen; Group D: Not classifiable as to human carcinogenicity; * The guideline value of tCAM is for swimming pool water, other parameters have only drinking water reference.
Table A3. Concentration of THMs with different disinfection methods.
Table A3. Concentration of THMs with different disinfection methods.
Country/Pool IDPool Type/Source WaterT (°C)pHDisinfectant DoseTOC/TN (mg·L−1)Concentration of THMs (μg·L−1)Method of DetectionAuthor
TCMBDCMDBCMTBMtTHMs
Chlorination
NANA/NANANA**NA/NA 38NABeyer et al. [9]
FranceIndoor/Tap277.21.71.8/NA275.13.11.036GC-MSCassan et al. [37]
U.S./S2WIndoor/Tap317.43.011/4.71193.0<1.0 122GC-ECDKanan [7]
France/BLAIndoor/NA257.61.085.5/NA556.81.60.764HS-GC-MSCimetiere and De Laat [10]
France/GANIndoor/NA257.62.72.8/NA283.40.90.733HS-GC-MSCimetiere and De Laat [10]
France/BELIndoor/NA257.42.25.1/NA615.41.20.769HS-GC-MSCimetiere and De Laat [10]
Denmark/LyngbyNA/Ground267.10.41.6/NA161.1<0.6<0.6 PAT-GC-MSSpiliotopoulou et al. [13]
Denmark/Gladsaxe mainNA/Ground267.20.62.1/NA130.8<0.6<0.6 PAT-GC-MSSpiliotopoulou et al. [13]
Denmark/Gladsaxe hotNA/Ground347.21.42.1/NA475.20.7<0.6 PAT-GC-MSSpiliotopoulou et al. [13]
U.S./3-Public PoolIndoor/TapNANANANA/NA179.05.0<1.0 GC-MSDaiber et al. [16]
China/BIndoor/Tap287.13.430 a/NA2202023.828 GC-ECDHang et al. [15]
Denmark/CleanNA/Ground277.31.71.5 a/NA256.0<0.6<0.6 PAT-GC-MSHansen et al. [14]
Denmark/PollutedNA/Synthetic277.31.02.4 a/NA356.02.0 PAT-GC-MSHansen et al. [14]
U.S.Indoor/Tap287.52.5NA/NA75 1.51.3 MIMSAfifi and Blatchley III [11]
France/Pool-1Indoor/Sea328.10.62.7/NA2.5 2.525 LLE-GC-ECDCheema et al. [12]
France/Pool-2Indoor/Sea308.10.53.1/NA2.5 3.030 LLE-GC-ECDCheema et al. [12]
France/Pool-3Indoor/Sea338.20.93.9/NA< 0.8 2.525 LLE-GC-ECDCheema et al. [12]
Denmark/Gladsaxe mainNA/Ground267.21.01.7 a/NA122.20.4 PAT-GC-MSCheema et al. [19]
EGMO
U.S./S4CIndoor/Tap277.44.07.1/9.1371.0<1.0 38GC-ECDKanan [7]
U.S./S4WIndoor/Tap307.43.06.5/3.2493.0<1.0 53GC-ECDKanan [7]
U.S./S6LIndoor/Tap297.62.07.9/4.8723.0<1.0 76GC-ECDKanan [7]
U.S./S6TIndoor/Tap347.43.07.7/2.572124.01.090GC-ECDKanan [7]
U.S./S15LIndoor/Tap277.72.83.8/0.8386.01.0 45GC-ECDKanan [7]
U.S./S15TIndoor/Tap327.81.27.5/1.21215.01.0 127GC-ECDKanan [7]
U.S./S17LIndoor/Tap287.63.57.3/4.182112.0 95GC-ECDKanan [7]
U.S./S17TIndoor/Tap337.43.524/4.42076.0<1.0 213GC-ECDKanan [7]
UV irradiation
FranceIndoor/Tap277.20.1451.9/NA76162.70.595GC-MSCassan et al. [37]
Denmark/Gladsaxe hotNA/Ground32NA1.8 *NA/NA34 34MIMSKristensen et al. [18]
Denmark/Gladsaxe hotNA/Ground32NA1.8 **NA/NA37 37MIMSKristensen et al. [18]
Denmark/Gladsaxe hotNA/Ground32NA3.6 ***NA/NA53 53MIMSKristensen et al. [18]
Denmark/LyngbyNA/Ground267.14.21.6/NA130.7<0.6<0.6 PAT-GC-MSSpiliotopoulou et al. [13]
Denmark/Gladsaxe mainNA/Ground267.24.22.1/NA170.9<0.6<0.6 PAT-GC-MSSpiliotopoulou et al. [13]
UV/chlorine
NANA/NANANANANA/NA 21NABeyer et al. [9]
FranceIndoor/Tap287.20.145/2.11.8/NA58152.20.575GC-MSCassan et al. [37]
France/BLAIndoor/NA257.62.35/2.55.5/NA85152.60.6103HS-GC-MSCimetiere and De Laat [10]
France/GANIndoor/NA257.62.35/1.62.8/NA64132.00.580HS-GC-MSCimetiere and De Laat [10]
France/BELIndoor/NA257.42.35/2.35.1/NA97162.9 118HS-GC-MSCimetiere and De Laat [10]
France/BLAIndoor/NA257.64.7/2.25.5/NA101183.30.4122HS-GC-MSCimetiere and De Laat [10]
France/GANIndoor/NA257.64.7/1.52.8/NA89171.90.4108HS-GC-MSCimetiere and De Laat [10]
France/BELIndoor/NA257.44.7/2.35.1/NA120213.2 144HS-GC-MSCimetiere and De Laat [10]
Denmark/LyngbyNA/Ground267.14.2/2.01.6/NA65111.8<0.6 PAT-GC-MSSpiliotopoulou et al. [13]
Denmark/Gladsaxe mainNA/Ground267.24.2/2.02.1/NA352.6<0.6<0.6 PAT-GC-MSSpiliotopoulou et al. [13]
Denmark/Gladsaxe hotNA/Ground347.24.2/2.02.1/NA878.80.9<0.6 PAT-GC-MSSpiliotopoulou et al. [13]
U.S.Indoor/Tap277.60.06 **/3.0NA/NA60 2.50.5 MIMSAfifi and Blatchley III [11]
U.S.Indoor/Tap278.00.06 */3.1NA/NA52 2.51.5 MIMSAfifi and Blatchley III [11]
France/Pool-1Indoor/Sea328.12.35/0.62.7/NA2.5 4.045 LLE-GC-ECDCheema et al. [12]
France/Pool-2Indoor/Sea308.12.35/0.53.1/NA3.5 4.040 LLE-GC-ECDCheema et al. [12]
France/Pool-3Indoor/Sea338.22.35/0.93.9/NA2.0 4.045 LLE-GC-ECDCheema et al. [12]
France/Pool-1Indoor/Sea328.14.7/0.62.7/NA5.0 5.558 LLE-GC-ECDCheema et al. [12]
France/Pool-2Indoor/Sea308.14.7/0.53.1/NA1.0 4.555 LLE-GC-ECDCheema et al. [12]
France/Pool-3Indoor/Sea338.24.7/0.93.9/NA4.5 5.560 LLE-GC-ECDCheema et al. [12]
Denmark/Gladsaxe mainNA/Ground267.22.1/1.01.9 a/NA325.51.0 PAT-GC-MSCheema et al. [19]
Denmark/Gladsaxe mainNA/Ground267.24.2/1.01.8 a/NA316.01.3 PAT-GC-MSCheema et al. [19]
Denmark/Gladsaxe mainNA/Ground267.29.5/1.01.7 a/NA356.21.5 PAT-GC-MSCheema et al. [19]
Denmark/Gladsaxe mainNA/Ground267.219/1.01.7 a/NA307.21.8 PAT-GC-MSCheema et al. [19]
Denmark/Gladsaxe mainNA/Ground267.247.5/1.01.2 a/NA105.81.6 PAT-GC-MSCheema et al. [19]
UV/H2O2
Denmark/LyngbyNA/Ground267.114.3/1.01.6/NA9.1<0.6<0.6<0.6 PAT-GC-MSSpiliotopoulou et al. [13]
Denmark/Gladsaxe mainNA/Ground267.216/1.02.1/NA13<0.6<0.6<0.6 PAT-GC-MSSpiliotopoulou et al. [13]
UV/H2O2/chlorine
Denmark/LyngbyNA/Ground267.114.3/1.0/2.01.6/NA36177.10.8 PAT-GC-MSSpiliotopoulou et al. [13]
Denmark/Gladsaxe mainNA/Ground267.216/1.0/2.02.1/NA253.10.6<0.6 PAT-GC-MSSpiliotopoulou et al. [13]
O3/chlorine
U.S./6-Public PoolOutdoor/TapNANANA/NANA/NA31<1.0<1.0<1.0 GC-MSDaiber et al. [16]
China/AIndoor/Tap287.4NA/2.113 a/NA1411062.047 GC-ECDHang et al. [15]
Denmark/CleanNA/Ground277.32.0/1.71.5 a/NA427.0<0.6<0.6 PAT-GC-MSHansen et al. [14]
Denmark/CleanNA/Ground277.34.0/1.71.5 a/NA607.0<0.6<0.6 PAT-GC-MSHansen et al. [14]
Denmark/CleanNA/Ground277.36.0/1.71.5 a/NA598.0<0.6<0.6 PAT-GC-MSHansen et al. [14]
Denmark/CleanNA/Ground277.38.0/1.71.5 a/NA457.0<0.6<0.6 PAT-GC-MSHansen et al. [14]
Denmark/CleanNA/Ground277.310/1.71.5 a/NA457.0<0.6<0.6 PAT-GC-MSHansen et al. [14]
Denmark/PollutedNA/Synthetic277.30.7/1.02.4 a/NA256.22.5 PAT-GC-MSHansen et al. [14]
Denmark/PollutedNA/Synthetic277.31.2/1.02.4 a/NA257.02.7 PAT-GC-MSHansen et al. [14]
Denmark/PollutedNA/Synthetic277.31.8/1.02.4 a/NA227.02.6 PAT-GC-MSHansen et al. [14]
Denmark/PollutedNA/Synthetic277.33.4/1.02.4 a/NA185.82.2 PAT-GC-MSHansen et al. [14]
Denmark/PollutedNA/Synthetic277.3100/1.02.4 a/NA7.01.0<0.6 PAT-GC-MSHansen et al. [14]
Denmark/Gladsaxe mainNA/Ground267.21.0/1.01.8 a/NA171.20.5 PAT-GC-MSCheema et al. [19]
Denmark/Gladsaxe mainNA/Ground267.22.0/1.01.7 a/NA211.70.5 PAT-GC-MSCheema et al. [19]
Denmark/Gladsaxe mainNA/Ground267.24.0/1.01.7 a/NA251.60.4 PAT-GC-MSCheema et al. [19]
O3/UV
Denmark/Gladsaxe hotNA/Ground32NANA/3.6 ***NA/NA35 35MIMSKristensen et al. [18]
Denmark/Gladsaxe mainNA/Ground267.27.0/9.5NA/NA5.01.00.1 PAT-GC-MSCheema et al. [19]
O3/UV/chlorine
Denmark/Gladsaxe mainNA/Ground267.21.0/9.5/1.01.5 a/NA214.01.4 PAT-GC-MSCheema et al. [19]
Denmark/Gladsaxe mainNA/Ground267.22.0/9.5/1.01.5 a/NA203.01.3 PAT-GC-MSCheema et al. [19]
Denmark/Gladsaxe mainNA/Ground267.24.0/9.5/1.01.4 a/NA152.50.7 PAT-GC-MSCheema et al. [19]
Denmark/Gladsaxe mainNA/Ground267.27.0/9.5/1.01.3 a/NA131.00.3 PAT-GC-MSCheema et al. [19]
Denmark/Gladsaxe mainNA/Ground267.210/9.5/1.01.3 a/NA120.10.1 PAT-GC-MSCheema et al. [19]
Note: Trihalomethanes (THMs); Chloroform (TCM); Bromodichloromethane (BDCM); Dibromochloromethane (DBCM), Bromoform (TBM); Total trihalomethanes (tTHMs); Temperature (T); Total organic carbon (TOC); Total nitrogen (TN); Electrochemically generated mixed oxidants (EGMO); Ultraviolet irradiation (UV); Ozone (O3); Gas chromatography (GC); Mass spectrometry (MS); Headspace (HS); Purge and trap (PAT); Electron capture detection (ECD); Membrane introduction mass spectrometry (MIMS); Not available (NA); Bold values are above the guideline values set by WHO for drinking water quality [58]; THMs concentration with low, medium, and combined low and medium pressure UV, respectively (*, **, ***); Organic carbon is reported as dissolved organic carbon (a); Unit for disinfectant dose: Free residual chlorine (mg·L−1); UV (J·cm−2); O3 (mg·L−1).
Table A4. Concentration of HAAs with different methods of disinfection.
Table A4. Concentration of HAAs with different methods of disinfection.
Country/Pool IDPool Type/Source WaterT (°C)pHDisinfectant DoseTOC/TN (mg·L−1)Concentration of HAAs (μg·L−1)Method of DetectionAuthor
MCAADCAATCAAMBAADBAABCAABDCAADBCAATBAAtHAAs
Chlorination
U.S./S2WIndoor/Tap317.43.011/4.7 8967181.0<1.01436<1.0 1665GC-ECDKanan [7]
France/BLAIndoor/NA257.61.085.5/NA7.8168183 <1.08.0 367LLE-GC-MSCimetiere and De Laat [10]
France/GANIndoor/NA257.62.72.8/NA6.012877 <1.07.0 218LLE-GC-MSCimetiere and De Laat [10]
France/BELIndoor/NA257.42.25.1/NA19361136 <1.016 531LLE-GC-MSCimetiere and De Laat [10]
Australia/5Covered outdoor/Tap267.55.0NA/NA33760870<0.5<0.5<0.58<0.5 GC-ECDYeh et al. [8]
Australia/6Indoor/Tap327.40.9NA/NA<0.5770460<0.5<0.5<0.5<0.5<0.5 GC-ECDYeh et al. [8]
Australia/7Indoor/Tap337.53.4NA/NA12021001700<0.5<0.5<0.511<0.5 GC-ECDYeh et al. [8]
U.S./3-Public PoolIndoor/TapNANANANA/NA9.489659.61937146.1<4.0 GC-ECDDaiber et al. [16]
China/BIndoor/Tap287.13.430 a/NA10296212.3<0.4874<0.4<0.4<0.4 GC-ECDHang et al. [15]
France/Pool-1Indoor/Sea328.10.62.7/NA 452.0 5.032 GC-ECDCheema et al. [12]
France/Pool-2Indoor/Sea308.10.53.1/NA 453.5 4.025 GC-ECDCheema et al. [12]
France/Pool-3Indoor/Sea338.20.93.9/NA 806.5 7.045 GC-ECDCheema et al. [12]
EGMO
U.S./S4CIndoor/Tap277.44.07.1/9.1 688789<1.01.010273.0 1518GC-ECDKanan [7]
U.S./S4WIndoor/Tap307.43.06.5/3.2 812411.0<1.04.0605.0 392GC-ECDKanan [7]
U.S./S6LIndoor/Tap297.62.07.9/4.8 9285521.02.030483.0 1563GC-ECDKanan [7]
U.S./S6TIndoor/Tap347.43.07.7/2.5 10333511.04.036367.0 1468GC-ECDKanan [7]
U.S./S15LIndoor/Tap277.72.83.8/0.8 1151022.05.0214714 306GC-ECDKanan [7]
U.S./S15TIndoor/Tap327.81.27.5/1.2 6901832.06.031452.0 960GC-ECDKanan [7]
U.S./S17LIndoor/Tap287.63.57.3/4.1 5042885.02510611032 1070GC-ECDKanan [7]
U.S./S17TIndoor/Tap337.43.524/4.4 678719254.016176934.0 9005GC-ECDKanan [7]
Australia/5Indoor/Tap327.51.0NA/NA11024002600<0.5<0.5<0.516<0.5 GC-ECDYeh et al. [8]
Australia/6Outdoor/Tap277.42.6NA/NA40480650<0.5<0.5<0.58.0<0.5 GC-ECDYeh et al. [8]
Australia/7Outdoor/Tap287.55.3NA/NA6414001300<0.5<0.5<0.58.0<0.5 GC-ECDYeh et al. [8]
UV/chlorine
France/BLAIndoor/NA257.62.35/2.55.5/NA10158187 < 1.05.0 360LLE-GC-MSCimetiere and De Laat [10]
France/GANIndoor/NA257.62.35/1.62.8/NA8.013675 <1.05.1 224LLE-GC-MSCimetiere and De Laat [10]
France/BELIndoor/NA257.42.35/2.35.1/NA21372139 <1.0 531LLE-GC-MSCimetiere and De Laat [10]
France/BLAIndoor/NA257.64.7/2.25.5/NA8.7147144 <1.07.3 307LLE-GC-MSCimetiere and De Laat [10]
France/GANIndoor/NA257.64.7/1.52.8/NA9.014979 <1.06.0 240LLE-GC-MSCimetiere and De Laat [10]
France/BELIndoor/NA257.44.7/2.35.1/NA22344122 <1.08.6 496LLE-GC-MSCimetiere and De Laat [10]
France/Pool-1Indoor/Sea328.12.35/0.62.7/NA 102.5 3.012 GC-ECDCheema et al. [12]
France/Pool-2Indoor/Sea308.12.35/0.53.1/NA 153.5 1.53.0 GC-ECDCheema et al. [12]
France/Pool-3Indoor/Sea338.22.35/0.93.9/NA 154.5 1.52.5 GC-ECDCheema et al. [12]
France/Pool-1Indoor/Sea328.14.7/0.62.7/NA 103.5 1.53.0 GC-ECDCheema et al. [12]
France/Pool-2Indoor/Sea308.14.7/0.53.1/NA 102.0 1.42.5 GC-ECDCheema et al. [12]
France/Pool-3Indoor/Sea338.24.7/0.93.9/NA 206.5 1.85.0 GC-ECDCheema et al. [12]
O3/chlorine
U.S./6-Public PoolOutdoor/TapNANANANA/NA313431865<1.0<1.02.612<2.0<4.0 GC-ECDDaiber et al. [16]
China/AIndoor/Tap287.42.113 a/NA412002016<0.4425< 0.41.28.1 GC-ECDHang et al. [15]
Note: Haloacetic acids (HAAs); Monochloroacetic acid (MCAA); Dichloroacetic acid (DCAA); Trichloroacetic acid (TCAA); Monobromoacetic acid (MBAA); Dibromoacetic acid (DBAA); Bromochloroacetic acid (BCAA); Bromodichloroacetic acid (BDCAA); Dibromochloroacetic acid (DBCAA); Tribromoacetic acid (TBAA); Total haloacetic acid (tHAAs); Temperature (T); Total organic carbon (TOC); Total nitrogen (TN); Electrochemically generated mixed oxidants (EGMO); Ultraviolet irradiation (UV); Ozone (O3); Liquid-liquid extraction (LLE); Gas chromatography (GC); Mass spectrometry(MS); Electron capture detection (ECD); Not available (NA); Bold values are above the guideline values set by WHO for drinking water quality [58]; Organic carbon is reported as dissolved organic carbon (a); Units for disinfectant dose: Free residual chlorine (mg·L−1); UV (J·cm−2); O3 (mg·L−1).
Table A5. Concentration of HANs with different disinfection methods.
Table A5. Concentration of HANs with different disinfection methods.
Country/Pool IDPool Type/Source WaterT (°C)pHDisinfectant DoseTOC/TN (mg·L−1)Concentration of HANs (μg·L−1)Method of DetectionAuthor
DCANDBANBCANTCANCANBANtHANs
Chlorination
U.S./S2WIndoor/Tap317.43.011/4.722 1.0 1.0 25GC-ECDKanan [7]
France/BLAIndoor/NA257.61.085.5/NA190.63.80.06 23LLE-GC-MSCimetiere and De Laat [10]
France/GANIndoor/NA257.62.72.8/NA3.00.00.90.0 4.0LLE-GC-MSCimetiere and De Laat [10]
France/BELIndoor/NA257.42.25.1/NA150.05.70.0 21LLE-GC-MSCimetiere and De Laat [10]
Denmark/LyngbyNA/Ground267.10.41.6/NA1.9 PAT-GC-MSSpiliotopoulou et al. [13]
Denmark/Gladsaxe mainNA/Ground267.20.62.1/NA1.4 PAT-GC-MSSpiliotopoulou et al. [13]
Denmark/Gladsaxe hotNA/Ground347.21.42.1/NA2.6 PAT-GC-MSSpiliotopoulou et al. [13]
U.S./3-Public PoolIndoor/TapNANANANA/NA9.4<1.07.4<1.0 GC-MSDaiber et al. [16]
China/BIndoor/Tap287.13.430 a/NA9.2<0.343.2<0.46 GC-ECDHang et al. [15]
Denmark/CleanNA/Ground277.31.71.5 a/NA2.2 PAT-GC-MSHansen et al. [14]
Denmark/PollutedNA/Synthetic277.31.02.4 a/NA7.0 PAT-GC-MSHansen et al. [14]
U.S.Indoor/Tap287.52.5NA/NA8.0 MIMSAfifi and Blatchley III [11]
France/Pool-1Indoor/Sea328.10.62.7/NA 121.0 LLE-GC-ECDCheema et al. [12]
France/Pool-2Indoor/Sea308.10.53.1/NA 9.01.5 LLE-GC-ECDCheema et al. [12]
France/Pool-3Indoor/Sea338.20.93.9/NA 121.0 LLE-GC-ECDCheema et al. [12]
Denmark/Gladsaxe mainNA/Ground267.21.01.7 a/NA1.8 0.2 PAT-GC-MSCheema et al. [16]
EGMO
U.S./S4CIndoor/Tap277.44.07.1/9.115 1.0 1.0 16GC-ECDKanan [7]
U.S./S4WIndoor/Tap307.43.06.5/3.221 2.0 1.0 24GC-ECDKanan [7]
U.S./S6LIndoor/Tap297.62.07.9/4.812 1.0 1.0 15GC-ECDKanan [7]
U.S./S6TIndoor/Tap347.43.07.7/2.5132.04.0 1.0 21GC-ECDKanan [7]
U.S./S15LIndoor/Tap277.72.83.8/0.87.01.02.0 1.0 11GC-ECDKanan [7]
U.S./S15TIndoor/Tap327.81.27.5/1.24.0 1.0 6GC-ECDKanan [7]
U.S./S17LIndoor/Tap287.63.57.3/4.1161.03.0 1.01.022GC-ECDKanan [7]
U.S./S17TIndoor/Tap337.43.524/4.447 2.01.03.0 53GC-ECDKanan [7]
UV irradiation
Denmark/LyngbyNA/Ground267.14.21.6/NA4.9 PAT-GC-MSSpiliotopoulou et al. [13]
Denmark/Gladsaxe mainNA/Ground267.24.22.1/NA2.7 PAT-GC-MSSpiliotopoulou et al. [13]
UV/chlorine
France/BLAIndoor/NA257.62.35/2.55.5/NA290.45.60.06 35LLE-GC-MSCimetiere and De Laat [10]
France/GANIndoor/NA257.62.35/1.62.8/NA8.00.02.60.05 11LLE-GC-MSCimetiere and De Laat [10]
France/BELIndoor/NA257.42.35/2.35.1/NA140.04.70.0 18LLE-GC-MSCimetiere and De Laat [10]
France/BLAIndoor/NA257.64.7/2.25.5/NA290.26.40.08 36LLE-GC-MSCimetiere and De Laat [10]
France/GANIndoor/NA257.64.7/1.52.8/NA110.03.00.05 14LLE-GC-MSCimetiere and De Laat [10]
France/BELIndoor/NA257.44.7/2.35.1/NA150.05.70.0 21LLE-GC-MSCimetiere and De Laat [10]
Denmark/LyngbyNA/Ground267.14.2/2.01.6/NA5.4 PAT-GC-MSSpiliotopoulou et al. [13]
Denmark/Gladsaxe mainNA/Ground267.24.2/2.02.1/NA2.3 PAT-GC-MSSpiliotopoulou et al. [13]
Denmark/Gladsaxe hotNA/Ground347.24.2/2.02.1/NA5.6 PAT-GC-MSSpiliotopoulou et al. [13]
U.S.Indoor/Tap277.60.06 **/3.0NA/NA10 MIMSAfifi and Blatchley III [11]
U.S.Indoor/Tap278.00.06 */3.1NA/NA7.7 MIMSAfifi and Blatchley III [11]
France/Pool-1Indoor/Sea328.12.35/0.62.7/NA 173.5 LLE-GC-ECDCheema et al. [12]
France/Pool-2Indoor/Sea308.12.35/0.53.1/NA 172.6 LLE-GC-ECDCheema et al. [12]
France/Pool-3Indoor/Sea338.22.35/0.93.9/NA 283.4 LLE-GC-ECDCheema et al. [12]
France/Pool-1Indoor/Sea328.14.7/0.62.7/NA 192.8 LLE-GC-ECDCheema et al. [12]
France/Pool-2Indoor/Sea308.14.7/0.53.1/NA 153.0 LLE-GC-ECDCheema et al. [12]
France/Pool-3Indoor/Sea338.24.7/0.93.9/NA 353.5 LLE-GC-ECDCheema et al. [12]
Denmark/Gladsaxe mainNA/Ground267.22.1/1.01.9 a/NA4.0 0.7 PAT-GC-MSCheema et al. [19]
Denmark/Gladsaxe mainNA/Ground267.24.2/1.01.8 a/NA4.5 0.7 PAT-GC-MSCheema et al. [19]
Denmark/Gladsaxe mainNA/Ground267.29.5/1.01.7 a/NA4.5 0.7 PAT-GC-MSCheema et al. [19]
Denmark/Gladsaxe mainNA/Ground267.219/1.01.7 a/NA4.6 0.6 PAT-GC-MSCheema et al. [19]
Denmark/Gladsaxe mainNA/Ground267.247.5/1.01.2 a/NA4.0 0.5 PAT-GC-MSCheema et al. [19]
UV/H2O2
Denmark/LyngbyNA/Ground267.114.3/1.01.6/NA3.6 PAT-GC-MSSpiliotopoulou et al. [13]
Denmark/Gladsaxe mainNA/Ground267.216/1.02.1/NA2.2 PAT-GC-MSSpiliotopoulou et al. [13]
UV/H2O2/chlorine
Denmark/LyngbyNA/Ground267.114.3/1.0/2.01.6/NA3.1 PAT-GC-MSSpiliotopoulou et al. [13]
Denmark/Gladsaxe mainNA/Ground267.216/1.0/2.02.1/NA2.1 PAT-GC-MSSpiliotopoulou et al. [13]
O3/chlorine
U.S./6-Public PoolOutdoor/TapNANANANA/NA14<1.0<1.0<1.0 GC-MSDaiber et al. [16]
China/AIndoor/Tap287.42.113 a/NA5.3<0.34<0.28<0.46 GC-ECDHang et al. [15]
Denmark/CleanNA/Ground277.32.0/1.71.5 a/NA2.7 PAT-GC-MSHansen et al. [14]
Denmark/CleanNA/Ground277.34.0/1.71.5 a/NA2.7 PAT-GC-MSHansen et al. [14]
Denmark/CleanNA/Ground277.36.0/1.71.5 a/NA3.5 PAT-GC-MSHansen et al. [14]
Denmark/CleanNA/Ground277.38.0/1.71.5 a/NA3.2 PAT-GC-MSHansen et al. [14]
Denmark/CleanNA/Ground277.310/1.71.5 a/NA4.2 PAT-GC-MSHansen et al. [14]
Denmark/PollutedNA/Synthetic277.30.7/1.02.4 a/NA4.5 PAT-GC-MSHansen et al. [14]
Denmark/PollutedNA/Synthetic277.31.2/1.02.4 a/NA4.0 PAT-GC-MSHansen et al. [14]
Denmark/PollutedNA/Synthetic277.31.8/1.02.4 a/NA4.0 PAT-GC-MSHansen et al. [14]
Denmark/PollutedNA/Synthetic277.33.4/1.02.4 a/NA4.0 PAT-GC-MSHansen et al. [14]
Denmark/PollutedNA/Synthetic277.3100/1.02.4 a/NA2.5 PAT-GC-MSHansen et al. [14]
Denmark/Gladsaxe mainNA/Ground267.21.0/1.01.8 a/NA1.8 0.3 PAT-GC-MSCheema et al. [19]
Denmark/Gladsaxe mainNA/Ground267.22.0/1.01.7 a/NA1.8 0.3 PAT-GC-MSCheema et al. [19]
Denmark/Gladsaxe mainNA/Ground267.24.0/1.01.7 a/NA2.5 0.4 PAT-GC-MSCheema et al. [19]
O3/UV
Denmark/Gladsaxe mainNA/Ground267.27.0/9.5NA/NA3.0 0.1 PAT-GC-MSCheema et al. [19]
O3/UV/chlorine
Denmark/Gladsaxe mainNA/Ground267.21.0/9.5/1.01.5 a/NA4.2 0.6 PAT-GC-MSCheema et al. [19]
Denmark/Gladsaxe mainNA/Ground267.22.0/9.5/1.01.5 a/NA4.0 0.5 PAT-GC-MSCheema et al. [19]
Denmark/Gladsaxe mainNA/Ground267.24.0/9.5/1.01.4 a/NA3.5 0.3 PAT-GC-MSCheema et al. [19]
Denmark/Gladsaxe mainNA/Ground267.27.0/9.5/1.01.3 a/NA3.2 0.3 PAT-GC-MSCheema et al. [19]
Denmark/Gladsaxe mainNA/Ground267.210/9.5/1.01.3 a/NA3.0 0.2 PAT-GC-MSCheema et al. [19]
Note: Haloacetonitlies (HANs); Dichloroacetonitrile (DCAN); Dibromoacetonitrile (DBAN); Bromochloroacetonitrile (DBAN); Trichloroacetonitrile (TCAN); Chloroacetonitrile (CAN); Bromoacetonitrile (BAN); Total haloacetonitlies (tHANs); Temperature (T); Total organic carbon (TOC); Total nitrogen (TN); Electrochemically generated mixed oxidants (EGMO); Ultraviolet irradiation (UV); Ozone (O3); Liquid–liquid extraction (LLE); Gas chromatography (GC); Mass spectrometry (MS); Purge and trap (PAT); Electron capture detection (ECD); Membrane introduction mass spectrometry (MIMS); Not available (NA); Bold values are above the guideline values set by WHO for drinking water quality [58]; HANs concentration with low and medium pressure UV, respectively (*, **); Organic carbon is reported as dissolved organic carbon (a). Units for disinfectant dose: Free residual chlorine (mg·L−1); UV (J·cm−2); O3 (mg·L−1).
Table A6. Concentration of THAs with different methods of disinfection.
Table A6. Concentration of THAs with different methods of disinfection.
Country/Pool IDPool Type/Source WaterT (°C)pHDisinfectant DoseTOC/TN (mg·L−1)CH (μg·L−1)Method of DetectionAuthor
Chlorination
France/BLAIndoor/NA257.61.085.5/NA363LLE-GC-MSCimetiere and De Laat [10]
France/GANIndoor/NA257.62.72.8/NA104LLE-GC-MSCimetiere and De Laat [10]
France/BELIndoor/NA257.42.25.1/NA378LLE-GC-MSCimetiere and De Laat [10]
U.S./3-Public PoolIndoor/TapNANANANA/NA165GC-MSDaiber et al. [16]
UV/chlorine
France/BLAIndoor/NA257.62.35/2.55.5/NA367LLE-GC-MSCimetiere and De Laat [10]
France/GANIndoor/NA257.62.35/1.62.8/NA136LLE-GC-MSCimetiere and De Laat [10]
France/BELIndoor/NA257.42.35/2.35.1/NA404LLE-GC-MSCimetiere and De Laat [10]
France/BLAIndoor/NA257.64.7/2.25.5/NA399LLE-GC-MSCimetiere and De Laat [10]
France/GANIndoor/NA257.64.7/1.52.8/NA150LLE-GC-MSCimetiere and De Laat [10]
France/BELIndoor/NA257.44.7/2.35.1/NA400LLE-GC-MSCimetiere and De Laat [10]
O3/chlorine
U.S./6-Public PoolOutdoor/TapNANANA/NANA/NA101GC-MSDaiber et al. [16]
Note: Trihaloacetaldehyde (THA); Chloral hydrate (CH); Temperature (T); Total organic carbon (TOC); Total nitrogen (TN); Ultraviolet irradiation (UV); ); Ozone (O3); Liquid–liquid extraction (LLE); Gas chromatography (GC); Mass spectrometry (MS); Not available (NA); Bold values are above the guideline values set by WHO for drinking water quality [58]; Units for disinfectant dose: Free residual chlorine (mg·L−1); UV (J·cm−2); O3 (mg·L−1).
Table A7. Concentration of CAMs with different methods of disinfection.
Table A7. Concentration of CAMs with different methods of disinfection.
Country/Pool IDPool Type/Source WaterT (°C) pHDisinfectant DoseTOC/TN (mg·L−1)Concentration of CAMs (μgL−1)Method of DetectionAuthor
MCAMDCAMTCAMtCAMs
Cholrination
BurnabyIndoor/NANANANANA/NA 1470DPD-KIMah and Heacock [38]
BurnabyWhirlpool/NANANANANA/NA 2020DPD-KIMah and Heacock [38]
U.S.Indoor/Tap287.52.5NA/NA8261370 MIMSAfifi and Blatchley III [11]
UV/chlorine
U.S.Indoor/Tap277.60.06 **/3.0NA/NA5537350 MIMSAfifi and Blatchley III [11]
U.S.Indoor/Tap278.00.06 */3.1NA/NA7553450 MIMSAfifi and Blatchley III [11]
O3 /chlorine
BurnabyIndoor/NANANANA/NANA/NA 1310DPDMah and Heacock [38]
BurnabyWhirlpool/NANANANA/NANA/NA 1890DPDMah and Heacock [38]
Note: Chloramines (CAMs); Monochloramine (MCAM); Dichloramine (DCAM); Trichloramine (TCAM); Total chloramines (tCAMs); Temperature (T); Total organic carbon (TOC); Total nitrogen (TN); Ultraviolet irradiation (UV); Ozone (O3); Diethyl-p-phenylenediamine (DPD); Colorimetric method (KI); Membrane introduction mass spectrometry (MIMS); Not available (NA); Bold values are above the guideline values set by WHO for swimming pool water quality [20]; Concentration of CAMs with low and medium pressure UV, respectively (*, **); Units for disinfectant dose: Free residual chlorine (mg·L−1); UV (J·cm−2); O3 (mg·L−1).

References

  1. Zwiener, C.; Richardson, S.D.; De Marini, D.M.; Grummt, T.; Glauner, T.; Frimmel, F.H. Drowning in disinfection byproducts? Assessing swimming pool water. Environ. Sci. Technol. 2007, 41, 363–372. [Google Scholar] [CrossRef] [PubMed]
  2. Weaver, W.A.; Li, J.; Wen, Y.; Johnston, J.; Blatchley, M.R.; Blatchley, E.R., III. Volatile disinfection by-product analysis from chlorinated indoor swimming pools. Water Res. 2009, 43, 3308–3318. [Google Scholar] [CrossRef] [PubMed]
  3. Lee, J.; Jun, M.-J.; Lee, M.-H.; Lee, M.-H.; Eom, S.-W.; Zoh, K.-D. Production of various disinfection byproducts in indoor swimming pool waters treated with different disinfection methods. Int. J. Hyg. Environ. Health 2010, 213, 465–474. [Google Scholar] [CrossRef] [PubMed]
  4. Manasfi, T.; Méo, M.D.; Coulomb, B.; Giorgio, C.D.; Boudenne, J.-L. Identification of disinfection by-products in freshwater and seawater swimming pools and evaluation of genotoxicity. Environ. Int. 2016, 88, 94–102. [Google Scholar] [CrossRef] [PubMed]
  5. Jacobs, J.; Spaan, S.; Van Rooy, G.; Meliefste, C.; Zaat, V.; Rooyackers, J.; Heederik, D. Exposure to trichloramine and respiratory symptoms in indoor swimming pool workers. Eur. Respir. J. 2007, 29, 690–698. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Lee, J.; Ha, K.-T.; Zoh, K.-D. Characteristics of trihalomethane (THM) production and associated health risk assessment in swimming pool waters treated with different disinfection methods. Sci. Total Environ. 2009, 407, 1990–1997. [Google Scholar] [CrossRef] [PubMed]
  7. Kanan, A. Occurrence and Formation of Disinfection by-Products in Indoor Swimming Pools Water. Ph.D. Thesis, Clemson University, Clemson, SC, USA, 2010. [Google Scholar]
  8. Yeh, R.Y.; Farré, M.J.; Stalter, D.; Tang, J.Y.; Molendijk, J.; Escher, B.I. Bioanalytical and chemical evaluation of disinfection by-products in swimming pool water. Water Res. 2014, 59, 172–184. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Beyer, A.; Worner, H.; van Lierop, R. The Use of UV for Destruction of Combined Chlorine; Version 1.0; Wallace & Tiernan: The Netherlands, 2004; Available online: https://www.pwtag.org.uk/reference/ (accessed on 29 September 2017).
  10. Cimetiere, N.; De Laat, J. Effects of UV-dechloramination of swimming pool water on the formation of disinfection by-products: A lab-scale study. Microchem. J. 2014, 112, 34–41. [Google Scholar] [CrossRef] [Green Version]
  11. Afifi, M.Z.; Blatchley, E.R., III. Effects of UV-based treatment on volatile disinfection byproducts in a chlorinated, indoor swimming pool. Water Res. 2016, 105, 167–177. [Google Scholar] [CrossRef] [PubMed]
  12. Cheema, W.A.; Manasfi, T.; Kaarsholm, K.M.S.; Andersen, H.R.; Boudenne, J.-L. Effect of medium-pressure UV-lamp treatment on disinfection by-products in chlorinated seawater swimming pool. Sci. Total Environ. 2017, 599–600, 910–917. [Google Scholar] [CrossRef] [PubMed]
  13. Spiliotopoulou, A.; Hansen, K.M.S.; Andersen, H.R. Secondary formation of disinfection by-products by UV treatment of swimming pool water. Sci. Total Environ. 2015, 520, 96–105. [Google Scholar] [CrossRef] [PubMed]
  14. Hansen, K.M.S.; Spiliotopoulou, A.; Cheema, W.A.; Andersen, H.R. Effect of ozonation of swimming pool water on formation of volatile disinfection by-products—A laboratory study. Chem. Eng. J. 2016, 289, 277–285. [Google Scholar] [CrossRef] [Green Version]
  15. Hang, C.; Zhang, B.; Gong, T.; Xian, Q. Occurrence and health risk assessment of halogenated disinfection byproducts in indoor swimming pool water. Sci. Total Environ. 2016, 543, 425–431. [Google Scholar] [CrossRef] [PubMed]
  16. Daiber, E.J.; DeMarini, D.M.; Ravuri, S.A.; Liberatore, H.K.; Cuthbertson, A.A.; Thompson-Klemish, A.; Byer, J.D.; Schmid, J.E.; Afifi, M.Z.; Blatchley, E.R., III; et al. Progressive increase in disinfection byproducts and mutagenicity from source to tap to swimming pool and spa water: Impact of human inputs. Environ. Sci. Technol. 2016, 50, 6652–6662. [Google Scholar] [CrossRef] [PubMed]
  17. Glauner, T.; Kunz, F.; Zwiener, C.; Frimmel, F.H. Elimination of swimming pool water disinfection byproducts with advanced oxidation processes (AOPs). Acta Hydrochim. Hydrobiol. 2005, 33, 585–594. [Google Scholar] [CrossRef]
  18. Kristensen, G.H.; Klausen, M.M.; Andersen, H.R.; Erdinger, L.; Lauritsen, F.R.; Arvin, E.; Albrechtsen, H.-J. Full scale test of UV-based water treatment technologies at Gladsaxe Sport centre—With and without advanced oxidation mechanisms. In Proceedings of the Third International Swimming Pool and Spa Conference, London, UK, 17–20 March 2009. [Google Scholar]
  19. Cheema, W.A.; Kaarsholm, K.M.S.; Andersen, H.R. Combined UV treatment and ozonation for the removal of by-product precursors in swimming pool water. Water Res. 2017, 110, 141–149. [Google Scholar] [CrossRef] [PubMed]
  20. World Health Organization (WHO). Guidelines for Safe Recreational Water Environments; Swimming Pools and Similar Environements; WHO Press: Geneva, Switzerland, 2006; Volume 2, Available online: http://apps.who.int/iris/bitstream/10665/43336/1/9241546808_eng.pdf (accessed on 29 September 2017).
  21. Chowdhury, S.; Alhooshani, K.; Karanfil, T. Disinfection by-products in swimming pool: Occurrences, implications and future needs. Water Res. 2014, 53, 68–109. [Google Scholar] [CrossRef] [PubMed]
  22. Teo, T.L.L.; Coleman, H.M.; Khan, S.J. Chemical contaminants in swimming pools: Occurrence, implications and control. Environ. Int. 2015, 76, 16–31. [Google Scholar] [CrossRef] [PubMed]
  23. Wang, K.; Guo, J.; Yang, M.; Junji, H.; Deng, R. Decomposition of two haloacetic acids in water using UV radiation, ozone and advanced oxidation processes. J. Hazard. Mater. 2009, 162, 1243–1248. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Xu, B.; Chen, Z.; Qi, F.; Ma, J.; Wu, F. Comparison of N-nitrosodiethylamine degradation in water by UV irradiation and UV/O3: Efficiency, product and mechanism. J. Hazard. Mater. 2010, 179, 976–982. [Google Scholar] [CrossRef] [PubMed]
  25. Scheideler, J.; Lekkerkerker-Teunissen, K.; Knol, T.; Ried, A.; Verberk, J.; van Dijk, H. Combination of O3/H2O2 and uv for multiple barrier micropollutant treatment and bromate formation control-An economic attractive option. Water Pract. Technol. 2011, 6. [Google Scholar] [CrossRef]
  26. Jin, X.; Peldszus, S.; Huck, P.M. Reaction kinetics of selected micropollutants in ozonation and advanced oxidation processes. Water Res. 2012, 46, 6519–6530. [Google Scholar] [CrossRef] [PubMed]
  27. Lekkerkerker-Teunissen, K.; Knol, A.H.; Van Altena, L.P.; Houtman, C.J.; Verberk, J.Q.J.C.; Van Dijk, J.C. Serial ozone/peroxide/low pressure UV treatment for synergistic and effective organic micropollutant conversion. Sep. Purif. Technol. 2012, 100, 22–29. [Google Scholar] [CrossRef]
  28. Medellin-Castillo, N.A.; Ocampo-Perez, R.; Leyva-Ramos, R.; Sanchez-Polo, M.; Rivera-Utrilla, J.; Mendez-Diaz, J.D. Removal of diethyl phthalate from water solution by adsorption, photo-oxidation, ozonation and advanced oxidation process (UV/H2O2, O3/H2O2 and O3/activated carbon). Sci. Total Environ. 2013, 442, 26–35. [Google Scholar] [CrossRef] [PubMed]
  29. Kruithof, J.C.; Kamp, P.C.; Martijn, B.J. UV/H2O2 treatment: A practical solution for organic contaminant control and primary disinfection. Ozone Sci. Eng. 2007, 29, 273–280. [Google Scholar] [CrossRef]
  30. Hansen, K.M.S.; Andersen, H.R. Energy effectiveness of direct UV and UV/H2O2 treatment of estrogenic chemicals in biologically treated sewage. Int. J. Photoenergy 2012, 2012, 270320. [Google Scholar] [CrossRef]
  31. Kim, H.; Shim, J.; Lee, S. Formation of disinfection by-products in chlorinated swimming pool water. Chemosphere 2002, 46, 123–130. [Google Scholar] [CrossRef]
  32. Keuten, M.; Peters, M.; Daanen, H.; de Kreuk, M.; Rietveld, L.; van Dijk, J. Quantification of continual anthropogenic pollutants released in swimming pools. Water Res. 2014, 53, 259–270. [Google Scholar] [CrossRef] [PubMed]
  33. Kanan, A.; Karanfil, T. Formation of disinfection by-products in indoor swimming pool water: The contribution from filling water natural organic matter and swimmer body fluids. Water Res. 2011, 45, 926–932. [Google Scholar] [CrossRef] [PubMed]
  34. Simard, S.; Tardif, R.; Rodriguez, M.J. Variability of chlorination by-product occurrence in water of indoor and outdoor swimming pools. Water Res. 2013, 47, 1763–1772. [Google Scholar] [CrossRef] [PubMed]
  35. Weisel, C.P.; Richardson, S.D.; Nemery, B.; Aggazzotti, G.; Baraldi, E.; Blatchley, E.R.; Blount, B.C.; Carlsen, K.H.; Eggleston, P.A.; Frimmel, F.H.; et al. Childhood asthma and environmental exposures at swimming pools: State of the science and research recommendations. Environ. Health Perspect. 2009, 117, 500–507. [Google Scholar] [CrossRef] [PubMed]
  36. Richardson, S.D.; DeMarini, D.M.; Kogevinas, M.; Fernandez, P.; Marco, E.; Lourencetti, C.; Balleste, C.; Heederik, D.; Meliefste, K.; McKague, A.B.; et al. What’s in the pool? A comprehensive identification of disinfection by products and assessment of mutagenicity of chlorinated and brominated swimming pool water. Environ. Health Perspect. 2010, 118, 1523–1530. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Cassan, D.; Mercier, B.; Castex, F.; Rambaud, A. Effects of medium-pressure UV lamps radiation on water quality in a chlorinated indoor swimming pool. Chemosphere 2006, 62, 1507–1513. [Google Scholar] [CrossRef] [PubMed]
  38. Mah, D.J.; Heacock, H. The effectiveness of ozone-chlorine treatment for reducing chloramine concentration compared to chlorine treatment in swimming pools and whirlpools. BCIT Environ. Health J. 2014. Available online: http://www.ncceh.ca/sites/default/files/BCIT-Mah-2014.pdf (accessed on 10 September 2017).
  39. Korich, D.; Mead, J.R.; Madore, M.S.; Sinclair, N.A.; Sterling, C.R. Effects of ozone, chlorine dioxide, chlorine, and monochloramine on Cryptosporidium parvum oocyst viability. Appl. Environ. Microbiol. 1990, 56, 1423–1428. [Google Scholar] [PubMed]
  40. Glauner, T.; Waldmann, P.; Frimmel, F.H.; Zwiener, C. Swimming pool water fractionation and geno-toxicological characterization of organic constituents. Water Res. 2005, 39, 4494–4502. [Google Scholar] [CrossRef] [PubMed]
  41. Ilyas, H.; Masih, I.; van der Hoek, J.P. An exploration of disinfection by-products formation and governing factors in chlorinated swimming pool water. J. Water Health. (under review).
  42. USACHPPM. Electrochemically Generated Oxidant Disinfection in the Use of Individual Water Purification Devices. 2006. Available online: www.dtic.mil/cgi-bin/GetTRDoc?AD=ADA453956 (accessed on 27 September 2017).[Green Version]
  43. Patermarakis, G.; Fountoukidis, E. Disinfection of water by electrochemical treatment. Water Res. 1990, 24, 1491–1496. [Google Scholar] [CrossRef]
  44. Drees, K.P.; Abbaszadegan, M.; Maier, R.M. Comparative electrochemical inactivation of bacteria and bacteriophage. Water Res. 2003, 37, 2291–2300. [Google Scholar] [CrossRef] [Green Version]
  45. Kerwick, M.I.; Reddy, S.M.; Chamberlain, A.H.L.; Holt, D.M. Electrochemical disinfection, an environmentally acceptable method of drinking water disinfection? Electrochim. Acta 2005, 50, 5270–5277. [Google Scholar] [CrossRef]
  46. Dowd, M. Assessment of THM formation with MIOX. Master’s Thesis, University of North Carolina, Chapel Hill, SC, USA, 1994. [Google Scholar]
  47. Nakajima, N.; Nakano, T.; Harada, F.; Taniguchi, H.; Yokoyama, I.; Hirose, J.; Sano, K. Evaluation of disinfective potential of reactivated free chlorine in pooled tap water by electrolysis. J. Microbiol. Methods 2004, 57, 163–173. [Google Scholar] [CrossRef] [PubMed]
  48. Alvarez-Uriate, J.I.; Iriarte-Velasco, U.; Chimeno-Alanis, N.; Gonzalez-Velasco, J.R. The effect of mixed oxidants and powdered activated carbon on the removal of natural organic matter. J. Hazard. Mater. 2010, 181, 426–431. [Google Scholar] [CrossRef] [PubMed]
  49. Craik, S.A.; Weldon, D.; Finch, G.R.; Bolton, J.R.; Belosevic, M. Inactivation of Cryptosporidium parvum oocysts using medium- and low-pressure ultraviolet radiation. Water Res. 2001, 35, 1387–1398. [Google Scholar] [CrossRef]
  50. Hijnen, W.A.M.; Beerendonk, E.F.; Medema, G.J. Inactivation credit of UV radiation for viruses, bacteria and protozoan (oo)cysts in water: A review. Water Res. 2006, 40, 3–22. [Google Scholar] [CrossRef] [PubMed]
  51. Li, J.; Blatchley, E.R., III. UV photodegradation of inorganic chloramines. Environ. Sci. Technol. 2009, 43, 60–65. [Google Scholar] [CrossRef] [PubMed]
  52. Muruganandham, M.; Suri, R.P.S.; Jafari, S.; Sillanpää, M.; Lee, G.-J.; Wu, J.J.; Swaminathan, M. Recent Developments in Homogeneous Advanced Oxidation Processes for Water and Wastewater Treatment. Review Article. Int. J. Photoenergy 2014, 2014, 821674. [Google Scholar] [CrossRef]
  53. Beltran, F.J.; Garcıa-Araya, J.F.; Giraldez, I. Gallic acid water ozonation using activated carbon. Appl. Catal. B 2006, 63, 249–259. [Google Scholar] [CrossRef]
  54. Wu, J.J.; Chen, S.H.; Muruganandham, M. Catalytic ozonation of oxalic acid using carbon-free rice husk ash catalysts. Ind. Eng. Chem. Res. 2008, 47, 2919–2925. [Google Scholar] [CrossRef]
  55. Wu, J.J.; Muruganandham, M.; Chang, L.T.; Chen, S.H. Oxidation of propylene glycol methyl ether acetate using ozone-based advanced oxidation processes. Ozone Sci. Eng. 2008, 30, 332–338. [Google Scholar] [CrossRef]
  56. Nicole, I.; De Laat, J.; Dore, M.; Duguet, J.P.; Suty, H. Etude de la degradation des trihalomethanes en milieu aqueux dilue par irradiation UV—Determination du rendement quantique de photolyse a 2537 nm 1. Environ. Technol. 1991, 12, 21–31. [Google Scholar] [CrossRef]
  57. Jo, C.H.; Dietrich, A.M.; Tanko, J.M. Simultaneous degradation of disinfection byproducts and earthy-musty odorants by the UV/H2O2 advanced oxidation process. Water Res. 2011, 45, 2507–2516. [Google Scholar] [CrossRef] [PubMed]
  58. World Health Organization (WHO). Guidelines for Drinking-water Quality, 4th ed.; Incorporating the First Addendum: Geneva, Switzerland, 2017; Available online: http://apps.who.int/iris/bitstream/10665/254637/1/9789241549950-eng.pdf (accessed on 29 September 2017).
  59. US EPA. Integrated Risk Information System (IRIS). Available online: https://cfpub.epa.gov/ncea/iris2/atoz.cfm (accessed on 15 September 2017).
  60. Manasfi, T.; Temime-Roussel, B.; Coulomb, B.; Vassalo, L.; Boudenne, J.-L. Occurrence of brominated disinfection by-products in the air and water of chlorinated seawater swimming pools. Int. J. Hyg. Environ. Health 2017, 220, 583–590. [Google Scholar] [CrossRef] [PubMed]
  61. Peng, D.; Saravia, F.; Abbt-Braun, G.; Horn, H. Occurrence and simulation of trihalomethanes in swimming pool water: A simple prediction method based on DOC and mass balance. Water Res. 2016, 88, 634–642. [Google Scholar] [CrossRef] [PubMed]
  62. Hansen, K.M.S.; Willach, S.; Antoniou, G.M.; Mosbæk, H.; Albrechtsen, H.J.; Andersen, H.R. Effect of pH on the formation of disinfection byproducts in swimming pool water—Is less THM better? Water Res. 2012, 46, 6399–6409. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Parinet, J.; Tabaries, S.; Coulomb, B.; Vassalo, L.; Boudenne, J.L. Exposure levels to brominated compounds in seawater swimming pools treated with chlorine. Water Res. 2012, 46, 828–836. [Google Scholar] [CrossRef] [PubMed]
  64. Chu, H.; Nieuwenhuijsen, M. Distribution and determinants of trihalomethane concentrations in indoor swimming pools. Occup. Environ. Med. 2002, 59, 243–247. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Glauner, T.N. Veröffentlichungen des Lehrstuhles für Wasserchemie und der DVGW-Forschungsstelle am Engler-Bunte-Institut des Karlsruher Institutes für Technologie. 2007, 45, 97–100. (In German) [Google Scholar]
  66. Florentin, A.; Hautemanière, A.; Hartemann, P. Health effects of disinfection by-products in chlorinated swimming pools. Int. J. Hyg. Environ. Health 2011, 214, 461–469. [Google Scholar] [CrossRef] [PubMed]
  67. Hong, H.C.; Wong, M.H.; Liang, Y. Amino acids as precursors of trihalomethane and haloacetic acid formation during chlorination. Arch. Environ. Contam. Toxicol. 2009, 56, 638–645. [Google Scholar] [CrossRef] [PubMed]
  68. Aggazzotti, G.; Fantuzzi, G.; Righi, E.; Predieri, G. Environmental and biological monitoring of chloroform in indoor swimming pools. J. Chromatogr. A 1995, 710, 181–190. [Google Scholar] [CrossRef]
  69. Aggazzotti, G.; Fantuzzi, G.; Righi, E.; Predieri, G. Blood and breath analyses as biological indicators of exposure to trihalomethanes in indoor swimming pools. Sci. Total Environ. 1998, 217, 155–163. [Google Scholar] [CrossRef]
  70. Manasfi, T.; Méo, M.D.; Giorgio, C.D.; Coulomb, B.; Boudenne, J.-L. Assessing the genotoxicity of two commonly occurring byproducts of water disinfection: Chloral hydrate and bromal hydrate. Mutat. Res./Genet. Toxicol. Environ. Mutagen. 2017, 813, 37–44. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Mean and standard deviation of THMs with different methods of disinfection. Note: “a” shows that chlorination, EGMO, UV irradiation, UV/chlorine and ozone/chlorine are not significantly different from each other; “b” shows that EGMO is significantly different from UV irradiation and ozone/chlorine; “c” shows that UV irradiation, UV/chlorine and ozone/chlorine are not significantly different from each other; “d” shows that ozone/UV/chlorine is significantly different from chlorination, EGMO, UV irradiation, UV/chlorine and ozone/chlorine at α = 0.05 (p < 0.05); Statistical analysis to compare the means in case of UV/hydrogen peroxide, UV/hydrogen peroxide/chlorine and ozone/UV was not carried out because of fewer data points.
Figure 1. Mean and standard deviation of THMs with different methods of disinfection. Note: “a” shows that chlorination, EGMO, UV irradiation, UV/chlorine and ozone/chlorine are not significantly different from each other; “b” shows that EGMO is significantly different from UV irradiation and ozone/chlorine; “c” shows that UV irradiation, UV/chlorine and ozone/chlorine are not significantly different from each other; “d” shows that ozone/UV/chlorine is significantly different from chlorination, EGMO, UV irradiation, UV/chlorine and ozone/chlorine at α = 0.05 (p < 0.05); Statistical analysis to compare the means in case of UV/hydrogen peroxide, UV/hydrogen peroxide/chlorine and ozone/UV was not carried out because of fewer data points.
Water 10 00797 g001
Figure 2. Mean and standard deviation of HAAs with different methods of disinfection. Note: “a” shows that chlorination, EGMO and UV/chlorine are not significantly different from each other; ”b“ shows that UV/chlorine is significantly different from EGMO at α = 0.05 (p < 0.05). Statistical analysis to compare the means in the case of ozone/chlorine was not carried out because of fewer data points.
Figure 2. Mean and standard deviation of HAAs with different methods of disinfection. Note: “a” shows that chlorination, EGMO and UV/chlorine are not significantly different from each other; ”b“ shows that UV/chlorine is significantly different from EGMO at α = 0.05 (p < 0.05). Statistical analysis to compare the means in the case of ozone/chlorine was not carried out because of fewer data points.
Water 10 00797 g002
Figure 3. Mean and standard deviation of HANs with different methods of disinfection. Note: “a” shows that chlorination is not significantly different from EGMO, UV/chlorine and ozone/chlorine; “b” shows that EGMO and UV/chlorine are significantly different from ozone/chlorine; “c” shows that ozone/chlorine and ozone/UV/chlorine are not significantly different from each other; “d” shows that ozone/UV/chlorine is significantly different from chlorination, EGMO and UV/chlorine at α = 0.05 (p < 0.05). Statistical analysis to compare the means in case of UV irradiation, UV/hydrogen peroxide, UV/hydrogen peroxide/chlorine and ozone/UV was not carried out because of fewer data points.
Figure 3. Mean and standard deviation of HANs with different methods of disinfection. Note: “a” shows that chlorination is not significantly different from EGMO, UV/chlorine and ozone/chlorine; “b” shows that EGMO and UV/chlorine are significantly different from ozone/chlorine; “c” shows that ozone/chlorine and ozone/UV/chlorine are not significantly different from each other; “d” shows that ozone/UV/chlorine is significantly different from chlorination, EGMO and UV/chlorine at α = 0.05 (p < 0.05). Statistical analysis to compare the means in case of UV irradiation, UV/hydrogen peroxide, UV/hydrogen peroxide/chlorine and ozone/UV was not carried out because of fewer data points.
Water 10 00797 g003
Figure 4. Mean and standard deviation of THAs with different methods of disinfection. Note: “a” shows that chlorination and UV/chlorine are not significantly different at α = 0.05 (p < 0.05). Statistical analysis to compare the means in the case of ozone/chlorine was not carried out because of fewer data points.
Figure 4. Mean and standard deviation of THAs with different methods of disinfection. Note: “a” shows that chlorination and UV/chlorine are not significantly different at α = 0.05 (p < 0.05). Statistical analysis to compare the means in the case of ozone/chlorine was not carried out because of fewer data points.
Water 10 00797 g004
Figure 5. Mean and standard deviation of CAMs with different methods of disinfection. Note: a statistical analysis to compare the means was not carried out because of fewer data points.
Figure 5. Mean and standard deviation of CAMs with different methods of disinfection. Note: a statistical analysis to compare the means was not carried out because of fewer data points.
Water 10 00797 g005
Figure 6. A graphical summary of concentrations of DBPs with different disinfection methods examined in this study.
Figure 6. A graphical summary of concentrations of DBPs with different disinfection methods examined in this study.
Water 10 00797 g006
Table 1. Methods used for the disinfection of swimming pool water.
Table 1. Methods used for the disinfection of swimming pool water.
Disinfection MethodChemicals Used for Disinfection
ChlorinationChlorine gas, calcium/sodium hypochlorite, Dichloroisocyanorates
Electrochemically generated mixed oxidantsElectrolysis of salt brine solution → Chlorine in the form of hypochlorous acid is the primary oxidant
Ultraviolet (UV) irradiationUV irradiation without post-chlorination
UV irradiation/chlorineUV irradiation with post-chlorination (chlorine gas, calcium/sodium hypochlorite)
UV irradiation/hydrogen peroxideUV irradiation with hydrogen peroxide
UV irradiation/hydrogen peroxide/chlorineUV irradiation with hydrogen peroxide and post-chlorination (chlorine gas)
Ozone/chlorineOzone with post-chlorination (chlorine gas, sodium hypochlorite)
Ozone/hydrogen peroxide/chlorineOzone with hydrogen peroxide and post-chlorination (chlorine gas)
Ozone/UV irradiationOzone with UV irradiation
Ozone/UV irradiation/chlorineOzone with UV irradiation and post-chlorination (chlorine gas)
Note: The methods and chemicals indicated are used in the studies reviewed in this paper.
Table 2. The oxidation potential of some reactive species.
Table 2. The oxidation potential of some reactive species.
SubstanceOxidation Potential (V)
Hydroxyl radical (OH)2.86
Ozone molecule (O3)2.07
Hydrogen peroxide (H2O2)1.78
Chlorine (Cl2)1.36
Chlorine dioxide (ClO2)1.27
Adopted from Muruganandham et al. [52].
Table 3. The results (p-values) of one-way ANOVA and z-test for comparison of means.
Table 3. The results (p-values) of one-way ANOVA and z-test for comparison of means.
ParameterTCMDCAATCAADCANCH
ANOVA results
0.100.200.110.0040.53
z-test results
Chlorination VS EGMO0.090.220.260.08NA
Chlorination VS UV irradiation0.71NANANANA
Chlorination VS UV/chlorine0.890.060.060.470.51
Chlorination VS O3/chlorine0.72NANA0.06NA
Chlorination VS O3/UV/chlorine0.03NANA0.03NA
EGMO VS UV irradiation0.04NANANANA
EGMO VS UV/chlorine0.070.040.0040.17NA
EGMO VS O3/chlorine0.03NANA0.01NA
EGMO VS O3/UV/chlorine0.001NANA0.005NA
UV irradiation VS UV/chlorine0.51NANANANA
UV irradiation VS O3/chlorine0.97NANANANA
UV irradiation VS O3/UV/chlorine0.02NANANANA
UV/chlorine VS O3/chlorine0.50NANA0.01NA
UV/chlorine VS O3/UV/chlorine0.0001NANA0.002NA
O3/chlorine VS O3/UV/chlorine0.01NANA0.56NA
Note: Bold values indicate significant difference at α = 0.05 (p < 0.05) for ANOVA and z-test results; NA: Not Available.
Table 4. Summary of advantages and disadvantages of the disinfection methods for swimming pool water.
Table 4. Summary of advantages and disadvantages of the disinfection methods for swimming pool water.
Disinfection MethodAdvantages [Reference]Disadvantages [Reference]
ChlorinationProvides rapid and long-lasting disinfection effects [22].Presence of chlorine-resistant microorganisms such as Cryptosporidium parvum and Giardia lambia [39]. Formation of potentially toxic carbonaceous and nitrogenous-based DBPs [1,2,3,4,40].
Electrochemically generated mixed oxidants Chlorine in the form of HOCl is the primary oxidant [43,44]. Toxicity concerns are similar, as in the case of typical chlorine disinfection [46].
Ultraviolet (UV) irradiationEffective for the control of microorganisms such as Cryptosporidium parvum and Giardia lambia [49,50]. Cost-competitive with chlorine to improve the quality of swimming pool water and air [9,51].The formation of nitrogenous-based DBPs (HANs) [10,11,12,13].
UV irradiation/chlorine The formation of nitrogenous-based DBPs (HANs) [10,11,12,13,19]. The post-chlorination leads to the formation of carbonaceous DBPs (THMs) [10,12,13,19].
UV irradiation/hydrogen peroxideDecrease in the formation of carbonaceous DBPs (THMs) [13].The formation of nitrogenous-based DBPs (HANs) [13].
UV irradiation/hydrogen peroxide/chlorine With this process post-chlorination is required, which leads to the formation of carbonaceous DBPs (THMs) [13].
Ozone/chlorine Ozone as a residual disinfectant is unsuitable in swimming pool water, because it readily vaporizes and decomposes. In addition, it is toxic and heavier than air, which leads to adverse health effects [20]. Lack of residuals and the relatively high doses required, ozone disinfection is usually followed by deozonation before water enters the pool and sequential disinfection by chlorine is required [3,6,14,15,16,17,19].
Ozone/hydrogen peroxide/chlorineThe effective removal of TOC with ozonation. The contact time of 3 min between oxidant and pool water was sufficient for the increased elimination efficiency of TOC compared with 20 min contact time in case of ozonation [17].Residual disinfection with chlorine is needed [17,19].
OH reactions in AOPs are non-specific and produce activated compounds suitable for THMs formation in the post-chlorination step. Therefore, the formation potential of THMs increased [17]
Ozone/UV irradiationDecrease in THMs and HANs formation [19]
Ozone/UV irradiation/chlorineThe effective removal of TOC with ozonation. The contact time of 3 min between oxidant and pool water was sufficient for the increased elimination efficiency of TOC compared with 20 min contact time in case of ozonation [17]Residual disinfection with chlorine is needed [17,19].
The increase in the level of THMs and HANs with ozone/UV/chlorine compared with ozone/UV was observed, indicating the complication of post-chlorination [19]. OH reactions in AOPs are non-specific and produce activated compounds suitable for THMs formation in the post-chlorination step. Therefore, the formation potential of THMs increased [17].
Note: Disinfection byproducts (DBPs); Hypochlorous acid (HOCl); Trihalomethanes (THMs); Haloacetonitriles (HANs); Hydroxyl radical (OH); Advanced oxidation processes (AOPs); Total organic carbon (TOC). The advantages and disadvantages of the disinfection methods indicated in the studies reviewed in this paper.

Share and Cite

MDPI and ACS Style

Ilyas, H.; Masih, I.; Van der Hoek, J.P. Disinfection Methods for Swimming Pool Water: Byproduct Formation and Control. Water 2018, 10, 797. https://doi.org/10.3390/w10060797

AMA Style

Ilyas H, Masih I, Van der Hoek JP. Disinfection Methods for Swimming Pool Water: Byproduct Formation and Control. Water. 2018; 10(6):797. https://doi.org/10.3390/w10060797

Chicago/Turabian Style

Ilyas, Huma, Ilyas Masih, and Jan Peter Van der Hoek. 2018. "Disinfection Methods for Swimming Pool Water: Byproduct Formation and Control" Water 10, no. 6: 797. https://doi.org/10.3390/w10060797

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