Next Article in Journal
What Explains Education Disparities in Screening Mammography in the United States? A Comparison with The Netherlands
Next Article in Special Issue
Biofouling Mitigation by Chloramination during Forward Osmosis Filtration of Wastewater
Previous Article in Journal
Reducing Stress and Promoting Social Integration of International Medical Students through a Tandem Program: Results of a Prospective-Program Evaluation
Previous Article in Special Issue
Study of Hybrid PVA/MA/TEOS Pervaporation Membrane and Evaluation of Energy Requirement for Desalination by Pervaporation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Removal Characteristics of N-Nitrosamines and Their Precursors by Pilot-Scale Integrated Membrane Systems for Water Reuse

Research Center for Environmental Quality Management, Kyoto University, Shiga 520-0811, Japan
*
Author to whom correspondence should be addressed.
Int. J. Environ. Res. Public Health 2018, 15(9), 1960; https://0-doi-org.brum.beds.ac.uk/10.3390/ijerph15091960
Submission received: 24 July 2018 / Revised: 4 September 2018 / Accepted: 5 September 2018 / Published: 7 September 2018
(This article belongs to the Special Issue Membrane Technologies for Water and Wastewater Treatment)

Abstract

:
This study investigated the removal characteristics of N-Nitrosamines and their precursors at three pilot-scale water reclamation plants. These plants applies different integrated membrane systems: (1) microfiltration (MF)/nanofiltration (NF)/reverse osmosis (RO) membrane; (2) sand filtration/three-stage RO; and (3) ultrafiltration (UF)/NF and UF/RO. Variable removal of N-Nitrosodimethylamine (NDMA) by the RO processes could be attributed to membrane fouling and the feed water temperature. The effect of membrane fouling on N-Nitrosamine removal was extensively evaluated at one of the plants by conducting one month of operation and chemical cleaning of the RO element. Membrane fouling enhanced N-Nitrosamine removal by the pilot-scale RO process. This finding contributes to better understanding of the variable removal of NDMA by RO processes. This study also investigated the removal characteristics of N-Nitrosamine precursors. The NF and RO processes greatly reduced NDMA formation potential (FP), but the UF process had little effect. The contributions of MF, NF, and RO processes for reducing FPs of NDMA, N-Nitrosopyrrolidine and N-Nitrosodiethylamine were different, suggesting different size distributions of their precursors.

Graphical Abstract

1. Introduction

In response to increasing indirect and direct potable reuse of reclaimed water, the monitoring and control of micropollutant concentrations are increasingly important. In many potable reuse schemes, reverse osmosis (RO) membranes are key components due to their high removal performance for inorganic salts and trace organic chemicals [1,2,3]. However, residual trace organic chemicals in RO permeate have been reported [4,5,6]. Among the chemicals, N-Nitrosamines are of key concern for potable reuse, and further research and monitoring of them are recommended [4,7,8]. These N-Nitrosamines include N-Nitrosodimethylamine (NDMA), N-Nitrosomethylethylamine (NMEA), N-Nitrosopyrrolidine (NPYR), N-Nitrosodiethylamine (NDEA), N-Nitrosopiperidine (NPIP), N-Nitrosomorpholine (NMOR), N-Nitrosodipropylamine (NDPA), and N-Nitrosodi-n-butylamine (NDBA). Most of these N-Nitrosamines are probable carcinogens [9], and their frequent occurrence in raw and secondary-treated wastewater [10,11,12] and reclaimed water [7,13,14,15,16,17] has been reported. For the augmentation of drinking water sources, the California Department of Public Health sets a drinking water notification level of 10 ng/L for NDMA and NDEA [18], and the Australian Guidelines for Water Recycling sets guideline values for NDMA (10 ng/L), NDEA (10 ng/L), and NMOR (1 ng/L) [19].
N-Nitrosamines, in particular NDMA, are of great concern in potable water reuse because they occur ubiquitously in treated wastewater and readily permeate RO membranes. Due to the low and variable removal of NDMA by RO membranes, additional processes such as UV processes and UV-based advanced oxidation processes are installed to remove NDMA from RO permeate, which results in a high energy cost. The removal characteristics of N-Nitrosamines by RO membranes have been extensively investigated at laboratory-scale [20,21,22,23]. Since N-Nitrosamines are hydrophilic and non-ionized at the typical environmental pH range of 6–8, their removal by RO membranes is governed mainly by steric hindrance. A laboratory-scale study showed increased removal of N-Nitrosamines in order of increasing molecular weight of N-Nitrosamines [21]. The previous study also found a significant decrease in the removal of N-Nitrosamines with increasing feed water temperature, and a discernible decrease in NDMA removal with decreasing feed water pH and with increasing feed water ionic strength. An increase in the removal of N-Nitrosamines with increasing permeate flux has also been observed [21]. In addition, RO membrane characteristics influence the removal of N-Nitrosamines [22]. Furthermore, membrane fouling can enhance NDMA removal by RO membranes as a result of enhanced size exclusion [23].
These laboratory-scale studies show the removal characteristics of N-Nitrosamines by RO membranes. However, pilot-scale data to assess the effects of realistic operating conditions on N-Nitrosamines removal are scarce. In addition, the previous pilot-scale studies have focused extensively on NDMA and evaluation for other N-Nitrosamines are still limited. Fujioka et al. assessed the removal of eight N-Nitrosamines in three full-scale RO plants and reported that discrepancies in N-Nitrosamine removal data between laboratory- and full-scale studies were occurred probably due to differences in water recovery and operating conditions such as temperature, membrane fouling, and hydraulic conditions [14]. The authors assessed the effects of operating conditions on N-Nitrosamines removal. However, no study evaluated the impact of membrane fouling on the removal of N-Nitrosamines by RO processes at pilot-scale.
N-Nitrosamines have also attracted attention as disinfection by-products (DBPs), which are formed during the disinfection of biologically treated wastewater with chlorine or chloramines [24,25,26] as well as alternative disinfectants such as chlorine dioxide (ClO2) and ozone (O3) [27]. NDMA can be formed from compounds which can release a secondary amine, including pharmaceuticals [28], pesticides [29], cationic polymers and ion exchange resins [30], and quaternary amines used in toiletries [31]. Although some precursors convert to NDMA with molar yields of 90% [28], specific precursors responsible for significant NDMA formation have not yet been identified. N-Nitrosamines and their precursors are both present in domestic and industrial wastewater. Among domestic wastewater, laundry water has been reported as the most significant source of N-Nitrosamines and their precursors, followed by shower water and urine, while less contributions came from bathroom washbasin and kitchen sink waters [32]. Loadings of N-Nitrosamines and their precursors from industrial discharges are generally site-specific (e.g., anti-yellowing agents found in a wastewater in Japan [33,34]).
In water reclamation processes using RO process, disinfectants are often added upstream of the RO process to inhibit biological fouling, and to the final water before distribution. Since polyamide RO membranes are susceptible to free-chlorine, chloramines are often used for mitigating biofouling on RO membranes. Since N-Nitrosamines are likely to be formed during chloramination, the addition of the disinfectants leads to the formation of N-Nitrosamines during membrane processes [35,36,37] and distribution [36]. To control N-Nitrosamine concentrations in reclaimed water, it is therefore important to understand the fate of their precursors during membrane treatment processes.
The fate of NDMA precursors in integrated membrane processes has been studied mainly at microfiltration (MF)–RO treatment plants [16,37,38] and ultrafiltration (UF)–RO treatment plants [35,38]. Even though the fate of NDMA precursors in nanofiltration (NF) process has been studied at bench scale, pilot-scale data to assess the fate of NDMA precursors by NF process are very scarce. Miyashita et al. investigated the removal of N-Nitrosamine precursors (i.e., dimethylamine, methylethylamine, diethylamine, and dipropylamine) in a bench-scale NF treatment and reported >98% removal of all four precursors [20]. However, evaluation of the removal efficiency of specific precursors by membrane processes is insufficient to evaluate the efficiency of removal of N-Nitrosamine precursors, because not all precursors have been identified. Mamo et al. assessed the fate of NDMA precursors in pilot-scale membrane bioreactor (MBR)–NF processes and found high rate of reduction of NDMA formation potential (FP) (>90%) by NF process [17]. The previous study has evaluated the fate NDMA precursors by a pilot-scale NF process, but assessment of the fate of other N-Nitrosamine precursors is still limited.
This study aimed: (1) to investigate the occurrence and fate of eight N-Nitrosamines and their FPs at three pilot-scale water reclamation plants using MF, UF, NF and RO processes, (2) to explore realistic operating conditions that contribute to the variability of the removal of N-Nitrosamines by pilot-scale RO processes; and (3) to extensively assess the effect of membrane fouling on the removal of N-Nitrosamines by a pilot-scale RO process by performing chemical cleaning of the RO element.

2. Materials and Methods

2.1. Site Description

Samples were collected at three pilot-scale water reclamation plants (A, B, C) in Japan receiving secondary effluent derived from municipal wastewater treatment plants. Plant A applies an MF-NF-RO system following an anaerobic–oxic activated sludge process (Figure 1, Table 1). Plant B uses a sand filtration (SF) and three-stage RO system following a conventional activated sludge process (Figure 1, Table 1). Plant C applies UF and two parallel NF and RO systems following a conventional activated sludge process (Figure 1, Table 1). The membrane systems at the three plants were operated at a constant permeate flux condition. The operating conditions of the membrane processes are shown in Table 4 (Section 3.3). To mitigate biofouling of membranes, sodium hypochlorite was continuously added to the MF feed water at plant A and to the SF feed water at plant B. The added chlorine was converted to chloramines (Table 3) by the reaction with ammonia present in the secondary effluent (data not shown) at these plants, at which nitrification was not performed during the secondary treatment. Polyaluminum chloride (PACl) was added as a coagulant prior to the MF and SF processes. To mitigate biofouling at plant C, 2,2-dibromo-3-nitrilopropionamide was periodically added prior to the NF and RO processes.

2.2. Sampling Protocol

Grab samples were taken across the water reclamation treatment trains (Figure 1). All samples were collected in amber glass bottles and stored in darkness at 4 °C until analysis. Chlorine concentrations were measured with chlorine test kits (Pocket Colorimeter II, Hach, Loveland, CO, USA) at plants A and B, where sodium hypochlorite was added upstream of the MF and SF processes (Figure 1). To quench residual chlorine and stop the formation of N-Nitrosamines, sodium thiosulfate was added to all samples except secondary effluent to give a final concentration of 10 mg/L. Water temperature, pH, and conductivity were measured on site with a multi-function water quality meter (U-52G, Horiba, Kyoto, Japan). Operating pressures and flows were monitored within the plants. Sampling campaigns were conducted at plant A from January 2013 to August 2014 (n = 3), at plant B from December 2012 to December 2014 (n = 3), and at plant C from June 2013 to January 2014 (n = 5).

2.3. Chemical Cleaning of RO Element

To assess the effect of RO membrane fouling on removal of N-Nitrosamines, RO feed and permeate samples at plant C were collected at different stages of membrane fouling development. The RO process was operated at constant permeate flux condition of 21 L/m2 h. Samplings were conducted once a week from November to December 2015. Since feed water temperature was not so fluctuated during the sampling period (27–30 °C), TMP values monitored at each sampling campaign were used as an indicator of membrane fouling development. After 1 month of operation, the RO element was chemically cleaned as follows: (1) circulate cleaning solutions (below) in the RO system for 1 h; (2) soak the element in cleaning solution for 1 h; and (3) flush sufficient RO permeate to displace remaining cleaning solution. The chemical cleaning was performed using 0.03% sodium dodecyl sulfate (SDS) with sodium hydroxide (pH 11), 2% citric acid (pH 2), and sodium hydroxide (pH 11). The changes in TMP were monitored after each cleaning to evaluate the effectiveness of the chemical cleanings for removing membrane foulants. RO feed and permeate samples were collected before and after the cleaning to evaluate the effect of chemical cleaning on membrane fouling and removal of N-Nitrosamines.

2.4. Analytical Techniques

2.4.1. N-Nitrosamines

Eight N-Nitrosamines—NDMA, NMEA, NPYR, NDEA, NPIP, NMOR, NDPA, and NDBA—were targeted in this study. These N-Nitrosamines have molecular weights in the range of 74 to 158 g/mol (Table 2). An analytical method previously developed for the determination of N-Nitrosamines in wastewater was employed [25]. This method uses solid-phase extraction, gas chromatography and analysis by tandem mass spectrometry. Eight deuterated N-Nitrosamines—NDMA-d6, NMEA-d3, NPYR-d8, NDEA-d10, NPIP-d10, NMOR-d8, NDPA-d14, and NDBA-d18—were used as surrogate. These deuterated chemicals were obtained from CDN Isotopes (Pointe-Claire, QC, Canada) and a stock solution was prepared in pure methanol at 1 mg/L of each deuterated N-Nitrosamines. After spiking surrogate solution into each sample, N-Nitrosamines were extracted with Sep-Pak NH-2 and AC-2 cartridges (Waters, MA, USA) at a flow rate of 10 mL/min. After the AC-2 cartridges were dried, the analytes were eluted from the cartridges with 2 mL dichloromethane (Wako Pure Chemical Industries, Tokyo, Japan) and were concentrated under a nitrogen gas stream. After adding dichloromethane solution and deuterated toluene (toluene-d8) stock solution (Supelco, Bellefonte, PA, USA), used as an injection internal standard, into the eluents, N-Nitrosamine concentrations were quantified using Varian 450 Series gas chromatograph coupled with a Varian 300 series tandem mass spectrometer. The limits of detection and quantification of N-Nitrosamines are shown in Table S1.

2.4.2. Formation Potentials

N-Nitrosamine precursors were evaluated as formation potentials as described [24] with minor modifications [11]. In brief, 20 mM chloramine stock solution (1400 mg Cl2/L) was prepared before each experiment because of its tendency to auto-decompose at high concentrations. First, the sample pH was adjusted to 7 with 10 mM phosphate buffer. Next, chloramination was performed in a 1-L amber glass bottle by adding 20 mM chloramine stock solution (100 mL) to samples (900 mL). The samples were then stirred for 10 days at room temperature on a shaker (NR-80, Taitec, Saitama, Japan). FP was defined as the increment of N-Nitrosamine concentration during 10 days and calculated by subtracting background N-Nitrosamine concentrations from the final concentration.

2.5. Calculations

N-Nitrosamine removal rate ( R N ) and FP reduction rate ( R FP ) in each process was calculated using the following equations:
R N   [ % ] = ( 1 C p C f ) × 100
R FP   [ % ] = ( 1 F P p F P f ) × 100
where C p and C f are N-Nitrosamine concentrations in the permeate and feed, respectively; and F P p and F P f are formation potentials in the permeate and feed, respectively.

3. Results and Discussion

3.1. Occurrence of N-Nitrosamines

NDMA and NMOR were prevalent N-Nitrosamines, as reported in other studies [7,13,14]. NDMA was detected in all secondary effluent samples at 9 to 41 ng/L at plant A, 11 to 20 ng/L at plant B, and 8 to 70 ng/L at plant C (Figure 2). NMOR was present in some secondary effluent samples at 50 to 358 ng/L at plant A, N.D. (not detected) to 8 ng/L at plant B, and N.D. to 4 ng/L at plant C (Figure 2). Along with NDMA and NMOR, NDBA was detected in secondary effluent at plant B in the second sampling campaign at 11 ng/L. NMEA, NPYR, NDEA, NPIP, and NDPA were not detected at any sampling campaigns.
NDMA was detected in all RO permeate samples at 12 to 22 ng/L at plant A, 2 to 8 ng/L at plant B, and 5 to 10 ng/L at plant C (Figure 2). Due to the low NDMA removal by the RO process, NDMA concentrations in RO permeate were slightly higher than the guideline value of 10 ng/L [19] at plant A, but were lower than the guideline value at plants B and C. NMOR was also detected in RO permeate samples at plants A and B. RO permeate samples at plant A had high concentrations of NMOR (7–71 ng/L), which was much higher than the guideline value of 1 ng/L established for potable reuse [19]. High NMOR concentrations in RO permeate have been reported in previous studies (5–30 ng/L [10]; 177–475 ng/L [14]). These results indicate that both NMOR and NDMA can pose risks for potable reuse and should be monitored at water reclamation plants.

3.2. Occurrence of Formation Potentials

NDMA FP was found most frequently at the three plants (Figure 3). The NDMA FP level in secondary effluent ranged from 67 to 232 ng/L at plant A, from 70 to 242 ng/L at plant B, and from 236 to 482 ng/L at plant C. These NDMA FP levels were in general lower than values reported from the U.S. (200–2000 ng/L [12]) and Australia (300–1020 ng/L [35]). In addition to NDMA FP, NPYR FP (101 ng/L), NDEA FP (41 ng/L), and NMOR FP (4 ng/L) were present in secondary effluent at plant A at the second sampling campaign (Figure 3a).
The presence of N-Nitrosamine FPs could lead to the formation of N-Nitrosamines during the membrane processes due to the reaction between precursors and disinfectants. Chlorines added upstream of the MF process at plant A and of the SF process at plant B (Figure 1) were converted to chloramines by reaction with ammonium in the secondary effluent (Table 3). In addition to the N-Nitrosamine FP levels and disinfectant concentrations, the contact time between disinfectant and water is an important factor affecting N-Nitrosamines formation. Plants A and B were pilot-scale plants and the contact times were less than 2 h.
To evaluate the formation of N-Nitrosamines during the membrane processes, mass loadings of N-Nitrosamines across the treatment train at plant A were calculated. There was no increase in NDMA mass loadings during the NF and RO processes (Figure S1). This result indicates that NDMA was not formed during these processes despite the presence of NDMA precursors and chloramines, which could be resulted from the relatively low chloramine concentrations (Table 3) and short contact time between water and chloramines during the NF and RO processes (<30 min). On the other hand, NDMA mass loadings increased during the MF process in the second sampling campaign, indicating NDMA formation (Figure S1). At the second sampling campaign at plant A, relatively high chloramine concentration (1.2 mg/L) was observed, which could explain the NDMA formation. Since NDMA was not well removed by the RO process, some NDMA formed during the MF process may pass through the RO membrane and increase NDMA concentration in the final product water. The chlorine concentration added needs to be optimized to avoid the formation of excessive NDMA in the treatment process. NDEA, NPYR, and NMOR were not formed across the treatment train despite the presence of FPs.
NDMA FP was also found in some RO permeate samples at 2 to 22 ng/L at plant A, N.D. to 89 ng/L at plant B, and N.D. to 29 ng/L at plant C (Figure 3). These values are comparable to values reported from the U.S. (12–59 ng/L [12]) and Australia (6 ng/L [35]). Even though NDMA FPs in RO permeate were lower than those in secondary effluent, NDMA precursors remaining in RO permeate might reduce RO permeate quality. For example, a previous study reported an increase in NDMA concentration from 13 ng/L to 24 ng/L during 24 h chloramination of RO permeate [36]. In assessing NDMA risks associated with the use of reclaimed water, not only NDMA but also NDMA FP in RO permeate need to be considered.

3.3. Removal of N-Nitrosamines

The three integrated membrane systems provided variable removal performance for the detected three N-Nitrosamines (Figure 4). Although the RO processes removed NMOR and NDBA with high efficiency, their removal of NDMA varied significantly, even at the same plant (Figure 5). The different removal rates of the three N-Nitrosamines can be attributed to their molecular weight because N-Nitrosamines removal by RO membranes is governed mainly by steric hindrance.

3.3.1. Plant A

The MF process exhibited limited removal for NDMA (<8%) and NMOR (<20%) (Figure 4a). The small NDMA molecule easily passes through the MF process, and the removal level was comparable to a previous report of <10% [39]. The NF process gave low to moderate removal rate for efficiency for NDMA (13–41%) and NMOR (16–47%), respectively (Figure 4a). The rates of NDMA removal in our study (13–41%) were slightly higher than those reported in previous studies (<10% [22]; <20% [20]). Since enhanced NDMA removal by fouled NF membranes was reported in a previous study [40], the relatively high NDMA removal observed in this study might relate to membrane fouling.
The difference in removal efficiency between NDMA and NMOR was noticeable during the RO process (Figure 4a). While the removal of NDMA was <19%, that of NMOR was as high as 81%. This result is in accordance with a previous study reporting that the removal of N-Nitrosamines by RO membranes increased in order of increasing molecular weight [21]. Thus, molecular size can significantly affect removal by RO processes. Surprisingly, the RO process exhibited lower removal rate for NDMA (<19%) than the NF process (19–41%) at plant A. Since the RO process exhibited low removal rates for conductivity (Figure 5d), the quite low NDMA removal can be attributed to damage of the RO membrane rather than to membrane characteristics or operational conditions. The RO membrane applied at plant A is a polyamide membrane, which is susceptible to free chlorine attack. Chlorine can destruct polymer structure of RO membranes and cause an increase in free volume and flexibility of the polymer chain [41]. Even though sodium hypochlorite added to the MF feed water at plant A (Figure 1) was mostly converted to chloramines (Table 3), the residual free chlorine might have caused the RO membrane damage. Since N-Nitrosamines removal by RO membranes are governed mainly by steric hindrance effect, the enlargement of free-volume hole-size of damaged membrane can decrease the removal of N-Nitrosamines, in particular the removal of low molecular weight N-Nitrosamines (i.e., NDMA).

3.3.2. Plant B

As expected, removal of NDMA, NMOR, and NDBA by sand filtration was limited, indicating that these N-Nitrosamines were too small to remove by sand filtration (Figure 4b). Plant B applies a three-stage RO process following the sand filtration. The removal of N-Nitrosamines by the three RO stages ranged from 49 to 88% of NDMA, from 59 to 97% of NMOR, and >97% of NDBA. The rates of removal increased in order of increasing molecular weight, as found at plant A and in other studies [20,21]. An increase in NDMA removal from the first to the third RO stage was observed (Figure 5a), which might be attributed to membrane fouling. According to laboratory-scale studies, increased NDMA removal is caused by decreasing feed water temperature, increasing permeate flux, and formation of membrane fouling [21,23]. Since the feed water temperature was the same in each RO stage (Table 4), feed water temperature was not the reason for the increased NDMA removal. Although the permeate flux of each RO stage was not monitored, permeate flux would decrease with advancing stage because each stage received the concentrate of the previous stage, and the increased salt concentration (Table 4) would result in increased osmotic pressure. A decrease in permeate flux would result in a decrease in NDMA removal [21], but this is the opposite trend observed in our study. Thus, changes in permeate flux through the three-stage RO process is not the main reason for the increased NDMA removal. On the other hand, RO membrane fouling would be more severe with advancing RO stage because each RO stage received concentrate of the previous RO stage. Enhanced removal of neutral and hydrophilic solutes by fouled RO membranes have been found in several studies [23,42,43]. The previous work suggested that membrane foulants caused a shrinking of membrane pores and prevented solute permeation through the fouled RO membranes, which resulted in increased solute removal. Based on the results from the previous studies, RO membrane fouling could attribute to the increased NDMA removal observed in this study. Water reclamation plants often apply multiple-stage RO systems to increase water recovery rate. Due to the effect of membrane fouling, variable NDMA removal can be provided in multiple-stage RO systems.

3.3.3. Plant C

Removal of NDMA and NMOR by the UF process was limited (Figure 4c). The NF process removed <32% of NDMA and 70% of NMOR. The RO process removed 67 to 96% of NDMA and >99% of NMOR. Since the RO process was operated with stable water recovery rate (75%) and permeate flux (11 L/m2/h), the variable NDMA removal by the RO process would be due to differences in RO feed water characteristics or to RO membrane fouling (Table 4). NDMA removal had moderate correlations with feed water temperature (R2 = 0.49; Figure 6a) and transmembrane pressure (TMP) (R2 = 0.33; Figure 6c). The TMP values were compensated with feed water temperature. An increase in NDMA removal was observed with increasing TMP (Figure 6c), with the only exception in the fifth sampling campaign. This result suggests that membrane fouling is a key factor enhancing NDMA removal by the RO process. The effect of membrane fouling on N-Nitrosamines removal by RO process was extensively evaluated by performing 1 month of operation and chemical cleaning of the RO element (discussed in Section 3.4). In the fifth sampling campaign, high NDMA removal was observed regardless of the relatively low TMP value (Figure 6c). The high NDMA removal might be attributed to feed water temperature. According to previous study, decreased feed water temperature resulted in an increase in NDMA removal by RO membranes [21]. One possible reason for this trend is a decrease in NDMA permeability through the RO membrane, which are resulted from increased water viscosity and decreased solute diffusivity. Another possible reason is a shrink of free-volume hole-size of the RO membrane. Decreased feed water temperature can affect the polymer structure of RO membranes and decrease free-volume hole-size [44]. For these reasons, feed water temperature could be the reason for the high NDMA removal.

3.4. Impact of RO Membrane Fouling on N-Nitrosamines Removal

To assess the impact of RO membrane fouling on N-Nitrosamines removal, samplings were conducted at plant C every week for 1 month. During 1 month of operation, TMP increased from 0.60 to 0.79 MPa, indicating the development of RO membrane fouling (Figure 7a). At the same time, NDMA removal increased from 71 to 99% and NMOR removal increased from 92 to 99% (Figure 7a). After the 1 month of operation, chemical cleaning was conducted using (1) 0.03% SDS and sodium hydroxide (pH 11), (2) 2% citric acid (pH 2), and (3) sodium hydroxide (pH 11). When the RO element was cleaned with SDS and sodium hydroxide, TMP value was kept at 0.79 MPa (Figure 7a), indicating the cleaning solution had no impacts on removing membrane foulants. On the other hand, TMP value was decreased from 0.79 to 0.68 MPa by the cleaning with citric acid, and further decreased to 0.62 MPa with sodium hydroxide (Figure 7a), indicating partial reduction of membrane fouling. Since citric acid and sodium hydroxide solutions are effective to remove inorganic and organic foulants, respectively, the reduction of membrane fouling by these chemical solutions suggests the presence of inorganic and organic foulants on the RO element.
After performing these chemical cleanings, NDMA removal decreased from 99 to 77% and NMOR removal decreased from 99 to 94%. NDMA removal had a strong correlation with TMP (R2 = 0.99, n = 5; Figure 7b). Feed water temperature and permeate flux, which have been reported to affect NDMA removal by RO membranes, did not fluctuate enough to affect NDMA removal during the sampling campaigns (Table 5). These results indicate that membrane fouling enhanced the removal of these N-Nitrosamines. In a laboratory-scale study, membrane fouling increased the removal of micropollutants by RO membrane by enhancing the effect of size exclusion [23]. Our data indicate that RO membrane fouling is one of the most important factors affecting N-Nitrosamines removal, even at pilot scale.

3.5. Rate of Reduction of N-Nitrosamine FPs

The overall reduction rates of NDMA FP achieved was more than 75% at the three plants (Figure 8a). Even though NDMA was not sufficiently removed by the NF and RO processes, precursors of NDMA were well removed by these processes. This result suggests that most precursors were ionized or were large enough to be removed by the NF and RO processes. However, the reduction rates of NDMA FP by the NF processes were lower than values (>98%) reported in a previous study which investigated the removal of four NDMA precursors (dimethylamine, ethylmethylamine, diethylamine, and dipropylamine) using a bench-scale NF process [20]. That study attributed the high rates of removal to electrostatic interactions between the negatively charged NF membrane and the positively charged precursors. Therefore, the lower rate of reduction by the NF processes observed in our study suggests that some NDMA precursors were present in their uncharged forms and were thus difficult to remove by the NF processes. For example, dimethylsulfamide and dimethylformamide are small uncharged NDMA precursors which can likely pass through NF processes [38]. Further research to identify uncharged NDMA precursors and their efficiencies of removal by water reclamation processes is necessary.
While NF and RO processes at the three plants removed high proportions of NDMA FP, the MF and SF processes removed widely variable proportions: <59% by MF at plant A and <67% by SF at plant B (Figure 8a). The relatively high reduction rates by the MF and SF processes could be attributed to the effect of coagulant added upstream. Wang et al. reported that coagulation preferentially removed NDMA precursors with molecular weight >30 kDa and hydrophobic fraction of a biologically treated wastewater [45]. Charge neutralization of negatively charged colloids by coagulant might be attributed to adsorption of the NDMA precursors onto some colloids or particulate compounds, which could be the reason for the reduction of NDMA FP due to the addition of coagulant. On the other hand, the UF process at plant C removed little NDMA FP (Figure 8a). These results suggest that NDMA precursors were present mainly as dissolved compounds at plant C, but partially as colloidal or particulate compounds at plants A and B.
FPs of NPYR, NDEA, and NMOR were detected only at plant A. Since NMOR FP was low (Figure 3a), and it was difficult to clearly evaluate reduction across the treatment train, the efficiency of the integrated membrane processes was evaluated only for FPs of NPYR and NDEA. These FPs were effectively reduced (>89%) by the MF-NF-RO system (Figure 8b). The membrane processes were effective at removing these N-Nitrosamine precursors. However, the efficiencies of removal of the N-Nitrosamine FPs differed among processes, indicating different size distributions of these N-Nitrosamine precursors. For example, the contribution of each membrane process to the overall rates of reduction of NPYR FP were 64% by MF, 1% by NF, and 36% by RO, and those of NDEA FP were negligible by MF, 82% by NF, and 7% by RO (Figure 8b). Most NPYR precursors were removed by the MF process, and the rest were removed mostly by the RO process. This result suggests that most NPYR precursors were present as colloidal or particulate compounds and rest were present as dissolved compounds, which cannot be removed well by the NF process. On the other hand, NDEA precursors were removed not by the MF process, but by the NF and RO processes. This result suggests that NDEA precursors would be predominantly present as dissolved compounds.

4. Conclusions

This study investigated the removal characteristics of N-Nitrosamines and their precursors at three pilot-scale membrane systems for water reclamation. The main reasons for the variable NDMA removal by the RO process at a plant could be membrane fouling and feed water temperature. The effect of membrane fouling on N-Nitrosamines removal by an RO process was extensively evaluated at one of the plant by conducting 1 month of operation and chemical cleaning of the RO element. The extensive evaluation at a plant revealed that membrane fouling enhanced N-Nitrosamines removal by the pilot-scale RO process. The finding contributes to better understanding of the variable NDMA removal by RO processes. The three-stage RO process at a plant showed increased NDMA removals as progress of the RO stages, which might be attributed to the effect of membrane fouling. Since RO membranes in multi-stage RO system would have different membrane fouling or permeate flux conditions, these differences might lead to variable NDMA removals by multi-stage RO processes.
This study also investigated the removal characteristics of N-Nitrosamine precursors at the three plants. NDMA FP was predominant at all three plants. The NF and RO processes were effective to reduce NDMA FP. However, remaining NDMA FP in their permeates implies the presence of uncharged precursors that are not easily removed by the membrane processes. FPs of NPYR, NDEA, and NMOR were present at a plant which used MF-NF-RO system. Each membrane process reduced FPs of NPYR and NDEA with different efficiencies, suggesting different size distributions of their precursors. In designing new water reclamation plants, it is important to investigate which kinds of N-Nitrosamine FPs are present in the feed water.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/1660-4601/15/9/1960/s1, Figure S1: Formation rates of (a) NDMA and (b) NMOR across the membrane treatment at plant A, Table S1: Limit of detection (LOD) and limit of quantification (LOQ) for eight N-Nitrosamines.

Author Contributions

H.T. (Hiroaki Tanaka) conceived and designed the study. H.T. (Haruka Takeuchi), N.Y. and N.N. collected samples and analyzed the results. H.T. (Haruka Takeuchi) wrote the first draft of the article. All authors read and approved the final version of the manuscript.

Funding

This work was funded by the Core Research for Evolutional Science and Technology (CREST) Grant for ‘Innovative Technologies and Systems for Research on Sustainable Water Use’ by the Japan Science and Technology Agency (JST). The authors acknowledge the staff in the water reclamation plants for supporting the samplings.

Conflicts of interest

The authors declare no conflicts of interest.

References

  1. Drewes, J.E.; Amy, G.; Bellona, C.; Filteau, G. Comparing nanofiltration and reverse osmosis for treating recycled water. Water Environ. Res. Found. 2008, 100, 102–116. [Google Scholar]
  2. Alturki, A.A.; Tadkaew, N.; McDonald, J.A.; Khan, S.J.; Price, W.E.; Nghiem, L.D. Combining MBR and NF/RO membrane filtration for the removal of trace organics in indirect potable water reuse applications. J. Memb. Sci. 2010, 365, 206–215. [Google Scholar] [CrossRef]
  3. Gerrity, D.; Pecson, B.; Trussell, R.S.; Trussell, R.R. Potable reuse treatment trains throughout the world. J. Water Supply Res. Technol. 2013, 62, 321–338. [Google Scholar] [CrossRef]
  4. Linge, K.L.; Blair, P.; Busetti, F.; Rodriguez, C.; Heitz, A. Chemicals in reverse osmosis-treated wastewater: Occurrence, health risk, and contribution to residual dissolved organic carbon. J. Water Supply Res. Technol. 2012, 61, 494–505. [Google Scholar] [CrossRef]
  5. Agus, E.; Zhang, L.; Sedlak, D.L. A framework for identifying characteristic odor compounds in municipal wastewater effluent. Water Res. 2012, 46, 5970–5980. [Google Scholar] [CrossRef] [PubMed]
  6. Bellona, C.; Drewes, J.E.; Xu, P.; Amy, G. Factors affecting the rejection of organic solutes during NF/RO treatment—A literature review. Water Res. 2004, 38, 2795–2809. [Google Scholar] [CrossRef] [PubMed]
  7. Zeng, T.; Plewa, M.J.; Mitch, W.A. N-Nitrosamines and halogenated disinfection byproducts in U.S. Full Advanced Treatment trains for potable reuse. Water Res. 2016, 101, 176–186. [Google Scholar] [CrossRef] [PubMed]
  8. USEPA. Contaminant Candidate List 4—CCL4; Environmental Protection Agency: Washington, DC, USA, 2016.
  9. USEPA. Integrated Risk Information System (IRIS); Environmental Protection Agency: Washington, DC, USA, 1993.
  10. Krauss, M.; Longree, P.; Dorusch, F.; Ort, C.; Hollender, J. Occurrence and removal of N-Nitrosamines in wastewater treatment plants. Water Res. 2009, 43, 4381–4391. [Google Scholar] [CrossRef] [PubMed]
  11. Yoon, S.; Nakada, N.; Tanaka, H. Occurrence and fate of N-Nitrosamines and their formation potential in three wastewater treatment plants in Japan. Water Sci. Technol. 2013, 68, 2118–2126. [Google Scholar] [CrossRef] [PubMed]
  12. Sedlak, D.L.; Deeb, R.A.; Hawley, E.L.; Mitch, W.A.; Durbin, T.D.; Mowbray, S.; Carr, S. Sources and Fate of Nitrosodimethylamine and Its Precursors in Municipal Wastewater Treatment Plants. Water Environ. Res. 2005, 77, 32–39. [Google Scholar] [CrossRef] [PubMed]
  13. Gerrity, D.; Pisarenko, A.N.; Marti, E.; Trenholm, R.A.; Gerringer, F.; Reungoat, J.; Dickenson, E. Nitrosamines in pilot-scale and full-scale wastewater treatment plants with ozonation. Water Res. 2015, 72, 251–261. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Fujioka, T.; Khan, S.J.; McDonald, J.A.; Roux, A.; Poussade, Y.; Drewes, J.E.; Nghiem, L.D. N-Nitrosamine rejection by reverse osmosis membranes: A full-scale study. Water Res. 2013, 47, 6141–6148. [Google Scholar] [CrossRef] [PubMed]
  15. Farré, M.J.; Döderer, K.; Keller, J.; Poussade, Y.; Mueller, J.; Gernjak, W. N-Nitrosodimethylamine (NDMA) in Purified Recycled Water Urban Water Security Research Alliance Technical Report No. 34; Urban Water Security Research Alliance: City Ease, Australia, 2010. [Google Scholar]
  16. Sgroi, M.; Vagliasindi, F.G.A.; Snyder, S.A.; Roccaro, P. N-Nitrosodimethylamine (NDMA) and its precursors in water and wastewater: A review on formation and removal. Chemosphere 2018, 191, 685–703. [Google Scholar] [CrossRef] [PubMed]
  17. Mamo, J.; Insa, S.; Monclús, H.; Rodríguez-Roda, I.; Comas, J.; Barceló, D.; Farré, M.J. Fate of NDMA precursors through an MBR-NF pilot plant for urban wastewater reclamation and the effect of changing aeration conditions. Water Res. 2016, 102, 383–393. [Google Scholar] [CrossRef] [PubMed]
  18. NDMA and Other Nitrosamines Drinking Water Issues. Available online: https://www.waterboards.ca.gov/drinking_water/certlic/drinkingwater/NDMA.html (accessed on 7 July 2018).
  19. Environment Protection and Heritage Council; National Health and Medical Research Council; Natural Resource Management Ministerial Council. Australian Guidelines for Water Recycling: Managing Health and Environmental Risks (Phase 2): Augmentation of Drinking Water Supplies. Available online: https://www.nhmrc.gov.au/guidelines-publications/eh56 (accessed on 7 July 2018).
  20. Miyashita, Y.; Park, S.-H.; Hyung, H.; Huang, C.-H.; Kim, J.-H. Removal of nitrosamines and their precursors by nanofiltration and reverse osmosis membranes. J. Environ. Eng. 2009, 135, 788–795. [Google Scholar] [CrossRef]
  21. Fujioka, T.; Nghiem, L.D.; Khan, S.J.; McDonald, J.A.; Poussade, Y.; Drewes, J.E. Effects of feed solution characteristics on the rejection of N-Nitrosamines by reverse osmosis membranes. J. Memb. Sci. 2012, 409, 66–74. [Google Scholar] [CrossRef]
  22. Fujioka, T.; Khan, S.J.; McDonald, J.A.; Roux, A.; Poussade, Y.; Drewes, J.E.; Nghiem, L.D. N-Nitrosamine rejection by nanofiltration and reverse osmosis membranes: The importance of membrane characteristics. Desalination 2013, 316, 67–75. [Google Scholar] [CrossRef]
  23. Fujioka, T.; Khan, S.J.; McDonald, J.A.; Henderson, R.K.; Poussade, Y.; Drewes, J.E.; Nghiem, L.D. Effects of membrane fouling on N-Nitrosamine rejection by nanofiltration and reverse osmosis membranes. J. Memb. Sci. 2013, 427, 311–319. [Google Scholar] [CrossRef]
  24. Mitch, W.A.; Gerecke, A.C.; Sedlak, D.L. A N-Nitrosodimethylamine (NDMA) precursor analysis for chlorination of water and wastewater. Water Res. 2003, 37, 3733–3741. [Google Scholar] [CrossRef]
  25. Yoon, S.; Nakada, N.; Tanaka, H. Occurrence and removal of NDMA and NDMA formation potential in wastewater treatment plants. J. Hazard. Mater. 2011, 190, 897–902. [Google Scholar] [CrossRef] [PubMed]
  26. Mccurry, D.L.; Ishida, K.P.; Oelker, G.L.; Mitch, W.A. Reverse osmosis shifts chloramine speciation causing re-formation of NDMA during potable reuse of wastewater. Environ. Sci. Technol. 2017, 41, 8589–8596. [Google Scholar] [CrossRef] [PubMed]
  27. Krasner, S.W.; Mitch, W.A.; McCurry, D.L.; Hanigan, D.; Westerhoff, P. Formation, precursors, control, and occurrence of nitrosamines in drinking water: A review. Water Res. 2013, 47, 4433–4450. [Google Scholar] [CrossRef] [PubMed]
  28. Shen, R.; Andrews, S.A. Demonstration of 20 pharmaceuticals and personal care products (PPCPs) as nitrosamine precursors during chloramine disinfection. Water Res. 2011, 45, 944–952. [Google Scholar] [CrossRef] [PubMed]
  29. Chen, W.H.; Young, T.M. NDMA formation during chlorination and chloramination of aqueous diuron solutions. Environ. Sci. Technol. 2008, 42, 1072–1077. [Google Scholar] [CrossRef] [PubMed]
  30. Kohut, S.A.; Andrews, K.D. Polyelectrolyte age and N-Nitrosodimethylamine formation in drinking water treatment. Water Qual. Res. J. Can. 2003, 38, 719–735. [Google Scholar] [CrossRef]
  31. Kemper, J.M.; Walse, S.S.; Mitch, W.A. Quaternary amines as Nitrosamine precursors: A role for consumer products? Environ. Sci. Technol. 2010, 44, 1224–1231. [Google Scholar] [CrossRef] [PubMed]
  32. Zeng, T.; Mitch, W.A. Contribution of N-Nitrosamines and their precursors to domestic sewage by greywaters and blackwaters. Environ. Sci. Technol. 2015, 49, 13158–13167. [Google Scholar] [CrossRef] [PubMed]
  33. Kosaka, K.; Asami, M.; Konn, Y.; Oya, M.; Kunikane, S. Identification of antiyellowing agents as precursors of N-Nitrosodimethylamine production on ozonation from sewage treatment plant influent. Environ. Sci. Technol. 2009, 43, 5236–5241. [Google Scholar] [CrossRef] [PubMed]
  34. Kosaka, K.; Asami, M.; Ohkubo, K.; Iwamoto, T.; Tanaka, Y.; Koshino, H.; Echigo, S.; Akiba, M. Identification of a new N-Nitrosodimethylamine precursor in sewage containing industrial effluents. Environ. Sci. Technol. 2014, 48, 11243–11250. [Google Scholar] [CrossRef] [PubMed]
  35. Farré, M.J.; Döderer, K.; Hearn, L.; Poussade, Y.; Keller, J.; Gernjak, W. Understanding the operational parameters affecting NDMA formation at Advanced Water Treatment Plants. J. Hazard. Mater. 2011, 185, 1575–1581. [Google Scholar] [CrossRef] [PubMed]
  36. Sgroi, M.; Roccaro, P.; Oelker, G.L.; Snyder, S.A. N-Nitrosodimethylamine (NDMA) formation at an indirect potable reuse facility. Water Res. 2015, 70, 174–183. [Google Scholar] [CrossRef] [PubMed]
  37. Fujioka, T.; Khan, S.J.; Poussade, Y.; Drewes, J.E.; Nghiem, L.D. N-Nitrosamine removal by reverse osmosis for indirect potable water reuse—A critical review based on observations from laboratory-, pilot- and full-scale studies. Sep. Purif. Technol. 2012, 98, 503–515. [Google Scholar] [CrossRef]
  38. Mitch, W.A.; Sedlak, D.L. Characterization and fate of N-Nitrosodimethylamine precursors in municipal wastewater treatment plants. Environ. Sci. Technol. 2004, 38, 1445–1454. [Google Scholar] [CrossRef] [PubMed]
  39. Plumlee, M.H.; López-Mesas, M.; Heidlberger, A.; Ishida, K.P.; Reinhard, M. N-Nitrosodimethylamine (NDMA) removal by reverse osmosis and UV treatment and analysis via LC-MS/MS. Water Res. 2008, 42, 347–355. [Google Scholar] [CrossRef] [PubMed]
  40. Chon, K.; Kim, S.H.; Cho, J. Removal of N-Nitrosamines in a membrane bioreactor and nanofiltration hybrid system for municipal wastewater reclamation: Process efficiency and mechanisms. Bioresour. Technol. 2015, 190, 499–507. [Google Scholar] [CrossRef] [PubMed]
  41. Gohil, J.M.; Suresh, A.K. Chlorine attack on reverse osmosis membranes: Mechanisms and mitigation strategies. J. Memb. Sci. 2017, 541, 108–126. [Google Scholar] [CrossRef]
  42. Xu, P.; Drewes, J.E.; Kim, T.U.; Bellona, C.; Amy, G. Effect of membrane fouling on transport of organic contaminants in NF/RO membrane applications. J. Memb. Sci. 2006, 279, 165–175. [Google Scholar] [CrossRef]
  43. Altalyan, H.N.; Jones, B.; Bradd, J.; Nghiem, L.D.; Alyazichi, Y.M. Removal of volatile organic compounds (VOCs) from groundwater by reverse osmosis and nanofiltration. J. Water Process Eng. 2016, 9, 9–21. [Google Scholar] [CrossRef] [Green Version]
  44. Sharma, R.R.; Agrawal, R.; Chellam, S. Temperature effects on sieving characteristics of thin-film composite nanofiltration membranes: Pore size distributions and transport parameters. J. Memb. Sci. 2003, 223, 69–87. [Google Scholar] [CrossRef]
  45. Wang, M.; Meng, Y.; Ma, D.; Wang, Y.; Li, F.; Xu, X.; Xia, C.; Gao, B. Integration of coagulation and adsorption for removal of N-Nitrosodimethylamine (NDMA) precursors from biologically treated municipal wastewater. Environ. Sci. Pollut. Res. 2017, 24, 12426–12436. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Process flows and sampling points in the water reclamation plants.
Figure 1. Process flows and sampling points in the water reclamation plants.
Ijerph 15 01960 g001
Figure 2. Concentrations of (a) NDMA and (b) NMOR detected across the treatment trains at plants A, B, and C (means ± standard deviation (SD)). N.D.: not detected.
Figure 2. Concentrations of (a) NDMA and (b) NMOR detected across the treatment trains at plants A, B, and C (means ± standard deviation (SD)). N.D.: not detected.
Ijerph 15 01960 g002
Figure 3. (a) N-Nitrosamine FPs found across the treatment trains at plant A, and (b,c) NDMA FPs found at plants B and C (means ± SD). N.D.: not detected.
Figure 3. (a) N-Nitrosamine FPs found across the treatment trains at plant A, and (b,c) NDMA FPs found at plants B and C (means ± SD). N.D.: not detected.
Ijerph 15 01960 g003
Figure 4. N-Nitrosamines removal by each membrane process at (a) plant A, (b) plant B, and (c) plant C (means ± SD). Error bars are not shown for the compounds detected once.
Figure 4. N-Nitrosamines removal by each membrane process at (a) plant A, (b) plant B, and (c) plant C (means ± SD). Error bars are not shown for the compounds detected once.
Ijerph 15 01960 g004
Figure 5. Removal of (a) NDMA, (b) NMOR, (c) NDBA, and (d) conductivity by RO processes at plants A, B, and C (means ± minimum and maximum). N.A.: not available. Error bars are not shown for the compounds detected once.
Figure 5. Removal of (a) NDMA, (b) NMOR, (c) NDBA, and (d) conductivity by RO processes at plants A, B, and C (means ± minimum and maximum). N.A.: not available. Error bars are not shown for the compounds detected once.
Ijerph 15 01960 g005
Figure 6. Correlations between NDMA removal and (a) feed water temperature, (b) conductivity, and (c) TMP. TMP values compensated with feed water temperature are shown.
Figure 6. Correlations between NDMA removal and (a) feed water temperature, (b) conductivity, and (c) TMP. TMP values compensated with feed water temperature are shown.
Ijerph 15 01960 g006
Figure 7. (a) Relations between TMP and removal of NDMA, NMOR, and conductivity before and after chemical cleaning (n = 1), and (b) correlation between TMP and removal of the two N-Nitrosamines.
Figure 7. (a) Relations between TMP and removal of NDMA, NMOR, and conductivity before and after chemical cleaning (n = 1), and (b) correlation between TMP and removal of the two N-Nitrosamines.
Ijerph 15 01960 g007
Figure 8. Overall rates of reduction of (a) NDMA FP at plants A, B, and C and of (b) FPs of NPYR and NDEA at plant A.
Figure 8. Overall rates of reduction of (a) NDMA FP at plants A, B, and C and of (b) FPs of NPYR and NDEA at plant A.
Ijerph 15 01960 g008
Table 1. Characteristics of membranes used at plants A, B, and C.
Table 1. Characteristics of membranes used at plants A, B, and C.
PlantMembraneMembrane MaterialTypePore Size a (µm)NaCl Rejection aElement SizeNumber of Elements Per VesselNumber of Vessels Per UnitRecovery Rate (%)
AMFCeramicCylindrical0.1N.A.N.A.N.A.N.A.100
NFPoly-vinyl alcohol polyamideSpiralN.A.92%4 inch1165
ROFully aromatic polyamideSpiralN.A.99.5%4 inch1180
BROAromatic composite polyamideSpiralN.A.99.7%4 inch1767
CUFPoly-vinylidene fluorideHollow fiber0.01N.A.N.A.N.A.N.A.100
NFPiperazine polyamideSpiralN.A.60%4 inch1175
ROAromatic composite polyamideSpiralN.A.99.7%4 inch1175
N.A.: not available. a Pore size and NaCl rejection values as specified by the manufacturers.
Table 2. Physicochemical properties of eight N-Nitrosamines.
Table 2. Physicochemical properties of eight N-Nitrosamines.
CompoundNDMANMEANPYRNDEANPIPNMORNDPANDBA
Molecular weight (g/mol) a74.0888.11100.12102.14114.15116.12130.19158.25
p K a a3.223.423.303.323.303.143.303.30
L o g K o w a0.080.410.390.750.81−0.321.052.56
a Calculated in MarvinSketch software (ChemAxon, Budapest, Hungary).
Table 3. Concentration of chloramines across the treatment trains of plant A and plant B.
Table 3. Concentration of chloramines across the treatment trains of plant A and plant B.
Sampling CampaignPlant APlant B
Concentration (mg/L)Concentration (mg/L)
NF Perm.NF Conc.RO Perm.RO Conc.Sand Filter Eff.1st RO Perm.1st RO Conc.2nd RO Perm.2nd RO Conc.3rd RO Perm.3rd RO Conc.
1st0.330.290.200.311.030.410.650.300.810.380.81
2nd1.181.151.181.200.980.190.35<0.020.40<0.020.46
3rd0.030.09<0.020.023.240.420.670.210.390.130.33
Table 4. Operating conditions and feed water qualities of the NF and RO processes.
Table 4. Operating conditions and feed water qualities of the NF and RO processes.
PlantSampling CampaignMembrane ProcessPermeate Flux (L/m² h)TMP (MPa) aFeed Temperature (°C)Feed pHFeed Conductivity (µS/cm)
A1stNF24.80.4116.18.1815
RO24.80.4816.27.9579
2ndNF24.80.3816.57.8765
RO24.80.4816.77.1515
3rdNF24.80.3429.27.0867
RO24.80.4429.57.1577
B1st1st-stage RON.A.N.A.18.35.3582
2nd-stage RON.A.N.A.18.95.9739
3rd-stage RON.A.N.A.19.35.2880
2nd1st-stage RON.A.N.A.25.36.51230
2nd-stage RON.A.N.A.25.75.11290
3rd-stage RON.A.N.A.26.06.51830
3rd1st-stage RON.A.N.A.25.35.9739
2nd-stage RON.A.N.A.25.76.51230
3rd-stage RON.A.N.A.26.06.51830
C1stRO11.30.5430.97.41830
2ndRO11.30.5731.7N.A.1377
3rdRO11.30.5832.6N.A.1465
4thRO12.80.7430.87.42520
5thRO11.30.7124.77.41570
N.A.: not available. a TMP values not compensated with feed water temperature.
Table 5. Operating conditions and feed water qualities of the RO process.
Table 5. Operating conditions and feed water qualities of the RO process.
Sampling CampaignMembrane ProcessWater Recovery (%)Permeate Flux (L/m²h)TMP (MPa) aFeed Temperature (°C)Feed Conductivity (µS/cm)Feed TOC (mg/L)
1stRO5721.50.6029.9160013.9
2ndRO5621.00.6728.4156016.0
3rdRO5520.70.7027.4167018.0
4thRO5721.20.7927.3165014.6
5thRO5621.10.6426.6173014.1
a TMP values not compensated with feed water temperature.

Share and Cite

MDPI and ACS Style

Takeuchi, H.; Yamashita, N.; Nakada, N.; Tanaka, H. Removal Characteristics of N-Nitrosamines and Their Precursors by Pilot-Scale Integrated Membrane Systems for Water Reuse. Int. J. Environ. Res. Public Health 2018, 15, 1960. https://0-doi-org.brum.beds.ac.uk/10.3390/ijerph15091960

AMA Style

Takeuchi H, Yamashita N, Nakada N, Tanaka H. Removal Characteristics of N-Nitrosamines and Their Precursors by Pilot-Scale Integrated Membrane Systems for Water Reuse. International Journal of Environmental Research and Public Health. 2018; 15(9):1960. https://0-doi-org.brum.beds.ac.uk/10.3390/ijerph15091960

Chicago/Turabian Style

Takeuchi, Haruka, Naoyuki Yamashita, Norihide Nakada, and Hiroaki Tanaka. 2018. "Removal Characteristics of N-Nitrosamines and Their Precursors by Pilot-Scale Integrated Membrane Systems for Water Reuse" International Journal of Environmental Research and Public Health 15, no. 9: 1960. https://0-doi-org.brum.beds.ac.uk/10.3390/ijerph15091960

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