Next Article in Journal
Heteroatom (N, S) Co-Doped CNTs in the Phenol Oxidation by Catalytic Wet Air Oxidation
Previous Article in Journal
Electrodeposition of Fe-Complexes on Oxide Surfaces for Efficient OER Catalysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Background Water Matrices on Pharmaceutical and Personal Care Product Removal by UV-LED/TiO2

1
Centre for Advanced Materials Joining, Department of Mechanical and Mechatronics Engineering, University of Waterloo, Waterloo, ON N2L3G1, Canada
2
Department of Biology, University of Waterloo, Waterloo, ON N2L3G1, Canada
3
Waterloo Institute of Nanotechnology, University of Waterloo, Waterloo, ON N2L3G1, Canada
4
Civil and Environmental Engineering, University of Alberta, Edmonton, AB T6G1H9, Canada
*
Authors to whom correspondence should be addressed.
Submission received: 30 March 2021 / Revised: 27 April 2021 / Accepted: 27 April 2021 / Published: 30 April 2021
(This article belongs to the Special Issue Photocatalysts in Wastewater Treatment)

Abstract

:
In this study, we evaluated the effectiveness of UV-LED-irradiated TiO2 in removing 24 commonly detected PPCPs in two water matrices (municipal wastewater effluent and Suwannee River NOM–synthetic water) and compared their performance with that of ultrapure water. Relatively fast removal kinetics were observed for 29% and 12% of the PPCPs in ultrapure water and synthetic surface water, respectively (kapp of 1–2 min−1). However, they all remained recalcitrant to photocatalysis when using wastewater effluent as the background matrix (kapp < 0.1 min−1). We also observed that the pH-corrected octanol/water partition coefficient (log Dow) correlated well with PPCP degradation rate constants in ultrapure water, whereas molecular weight was strongly associated with the rate constants in both synthetic surface water and wastewater. The electrical energy per order (EEO) values calculated at the end of the experiments suggest that UV-LED/P25 can be an energy-efficient method for water treatment applications (2.96, 4.77, and 16.36 kW h m−3 in ultrapure water, synthetic surface water, and wastewater effluents, respectively). Although TiO2 photocatalysis is a promising approach in removing PPCPs, our results indicate that additional challenges need to be overcome for PPCPs in more complex water matrices, including an assessment of photocatalytic removal under different background water matrices.

Graphical Abstract

1. Introduction

Currently existing water and wastewater treatment technologies can only partially remove emerging organic contaminants such as pharmaceuticals and personal care products (PPCPs). Therefore, innovative water treatment solutions are required for both water and wastewater purification [1,2]. Heterogenous photocatalysis, such as the use of titanium dioxide (TiO2) nanomaterials, has been widely studied as a secondary and tertiary treatment option for the removal of PPCPs that remain recalcitrant [3,4]. TiO2 can produce highly reactive and unselective oxidant species such as hydroxyl radicals (OH) upon UV irradiation, which can oxidize and mineralize organic contaminants at high reaction rates [5].
A total of 80 studies evaluated here, detailed in Table S1 in the Supplementary Materials (SI), including those from our lab [6,7,8,9,10], have intensively investigated the effect of parameters such as photocatalyst type and concentration, light intensity, pH, and organic and inorganic constituents in the background water (i.e., dissolved organic matter, suspended solids, and alkalinity) on PPCP degradation using TiO2 treatment. Although these studies have explored the photocatalytic removals of PPCPs in mixtures, approximately 70% of them investigated TiO2 photocatalytic processes in a pure water matrix and only 30% of these studies have used surface water and wastewater effluents.
Industrial or municipal applications require careful consideration of the source water and effluent makeup as they are a complex mixture of chemicals that can influence the efficiency of the photocatalyst. For instance, inorganic ions such as NH4+, NO3, and HCO3 simultaneously change the TiO2 surface charge and adsorption of microcontaminants [11]. These major water constituents can further behave as scavengers or donors of the oxidative species produced during photocatalytic reactions [11]. Humic substances account for most of the natural organic matter (NOM) in the surface waters [12]. They may act as photosensitizers by generating additional oxidative species, which can then contribute to the indirect photodegradation of organic contaminants [13]. On the other hand, they may also scavenge the OH produced during photocatalysis, slowing down the overall degradation of the target contaminant [14]. In addition, most of the studies restricted their experiments to a limited number of PPCPs (approximately three compounds on average, as shown in Table S1; please see Table S2 for the abbreviation expansions). However, pharmaceuticals are very structurally diverse and, therefore, have highly varied physicochemical properties, theoretically leading to large differences in removal kinetics under different conditions [9].
In this study, the photocatalytic degradation of 24 PPCPs (nominal concentration of 2 μg L−1) in three different water matrices (ultrapure water; synthetic surface water using Suwannee River NOM; and the Waterloo Region wastewater effluent in Ontario, Canada) were investigated using UV-LED/P25. We further hypothesized a relationship between PPCP removal kinetics and their physicochemical parameters (isoelectric charge, pH-corrected octanol/water partition coefficient, solubility, and molecular weight). This can be useful in future photocatalytic optimization studies. Finally, we assessed the energy efficiency of UV-LED/P25 to remove PPCPs in different background matrices.

2. Results and Discussion

2.1. Adsorption and Photolysis Control

The PPCP removal through adsorption during the equilibration period (i.e., 60 min dark treatment before irradiation starts) was calculated by dividing the difference between initial and final concentrations during the dark period by the initial concentration (Equation (S1) in SI). Although most compounds showed minimal removal (0–10%) in three water matrices via adsorption (Figure 1), removal of o-hydroxy atorvastatin (63%), p-hydroxy atorvastatin (43%), bisphenol (2%), diclofenac (18%), and monensin (15%) were significantly different in ultrapure water vs. synthetic surface water and wastewater (one-way ANOVA, α = 0.05, p-values in Table S3). Water characteristics such as pH can have an impact on PPCP removal [15]. The acidity and alkalinity of the water matrix change the isoelectric point of both the photocatalyst and the organic compounds and can influence the electrostatic interaction between them [16]. TiOH, TiOH2+, and TiO are the primary TiO2 surface functional groups and their distribution is determined by the pH at the point with zero charge (pHzpc). At pH < pHzpc, TiOH2+ is the predominant species, while at pH > pHzpc, the P25 surface becomes negative through reactions with hydroxyl compounds that produce TiO [17]. At pH conditions above the dissociation constant (pKa) of organic compounds, the functional groups become negatively charged and will likely facilitate interactions with oppositely charged photocatalysts [18]. As illustrated in Figure 2, charged PPCPs were attracted to oppositely charged P25 nanoparticles but were repelled from a surface that is similarly charged. The P25 surface is slightly positive in ultrapure water (pHzpc for P25 is 5.5 or 6 at pHUPW = 6) [19], while it is more negative in wastewater effluents and synthetic surface water (pHWWE = 7 and pHSSW = 8), as shown in Table 1. Therefore, the interaction between anionic compounds such as atorvastatin, monensin, and diclofenac is favored in ultrapure water.
Preferential adsorption of synthetic surface water or wastewater constituents can also occupy the adsorption sites for microcontaminants, reducing the adsorption capacity of TiO2 [20]. It was also observed in our study that the overall removal during the dark period was higher in ultrapure water compared to surface water and wastewater (23%, 10%, and 1%, respectively). Since most studies have explored photocatalysis in ultrapure water, photocatalyst-related adsorption (Figure 2) may not be achieved in complex water matrices due to additional water constituents.
The results of photolysis treatments in ultrapure water and wastewater effluents indicated that there was no statistical difference in removal under UV irradiation alone except for atorvastatin metabolites (o-hydroxy atorvastatin and p-hydroxy atorvastatin) and diclofenac (Figure S1 and Table S3). This result is not surprising as some of these compounds have been found to be susceptible to photodegradation [7]. However, their removals via photolysis remained lower compared to photocatalysis, as described subsequently below.

2.2. Removal of PPCPs during Photocatalytic Treatment

2.2.1. Effect of Water Matrices

To investigate the effects of background water matrices on PPCP photocatalytic degradation, the rate constant (kapp) of PPCPs in water matrices were calculated using the Langmuir–Hinshelwood kinetics model (Equation (S2) in the SI). Of the 24 compounds, 7 of the PPCPs were removed quickly in ultrapure water including bisphenol A, naproxen, atorvastatin, and its metabolites. However, only three were removed at this rate in surface water (i.e., atorvastatin and its metabolites), and all of them showed little to no removal in wastewater effluent (Figure 3). The rate constants of caffeine and carbamazepine were not impacted by any water matrices (Table S3), suggesting that they are either recalcitrant to treatment or the background matrices simply do not affect their behaviour.
As expected, poor removal was observed with the increasing complexity of the background matrix (Table 1). Reactive oxygen species (ROS) (i.e., OH, O2, and hVB+) produced by UV-LED/P25 photocatalysis reactions allow for the degradation of the wide range of organic contaminants, including PPCP [21]. However, parallel reactions between the oxidizing agents and water constituents, such as Cl, NO3, organic carbon, and carbonate ions (HCO3 and CO32−), may consume the oxidizing species required for the degradation of contaminants [20]. For instance, the presence of Cl ions in the water matrix can have multiple effects on photocatalysis: (i) when TiOH2+ is the main functional group (low pH), Cl ions can compete with organic contaminants for active sites on the catalyst surface and can decrease the reaction rate [22]; (ii) in the presence of oxygen molecules, Cl ions compete for scavenging photoinduced electrons and block the chain reactions of OH [23]; and (iii) possible reactions between Cl ions and free radicals can impede the degradation process [24]. In our experiments, the amounts of dissolved chloride were 40 and 490 mg L−1 in synthetic surface water and wastewater effluents (Table 1) and a range of 100–1000 mg L−1 was found to produce some of these effects [22,24,25,26].
Carbonate ions in wastewater effluents and surface waters (normally found in the ranges of 38–575 and 61–183 mg L−1, respectively) have been reported to have contrasting outcomes in TiO2 photocatalysis, i.e., by either inhibiting or enhancing the degradation of PPCPs [11,27,28,29]. For instance, HCO3- ions may scavenge ROS and decrease the photocatalytic efficiency, or react with them and produce carbonate radicals (CO3) [30]. CO3 is an oxidant (E0 = 1.78 at pH = 7) with high selectivity toward organic compounds (with a second-order rate constant 102–109 M−1 s−1) [31,32], and it can preferentially react with electron-rich compounds such as aromatic anilines, amino acids, and sulfur or nitrogen-containing compounds such as oxytetracycline and sulfamethoxazole through electron transfer of hydrogen abstraction [31,33,34,35].
The average PPCP rate constants in synthetic surface water were 10 times higher than that in wastewater effluents. This result may be related to the photosensitization effect of NOM. As indicated by prior studies, NOM can also have two contrasting effects on TiO2 photocatalysis [36,37,38]. It can either inhibit the photodegradation by scavenging the photoinduced holes, attenuate the light, compete for active sites, deactivate the photocatalyst surface [39], or act as precursor for reactive radical species in water matrices [40]. Humic substances contribute to photocatalytic degradation of organic compounds by producing reactive oxidant such as solvated electrons (eaq), singlet oxygen (1O2), OH, O2, and H2O2 [41]. Doll and Frimmel, in 2003, related the high degradation rate of carbamazepine in Hohloh Lake NOM to NOM-induced reactions [40].
TiO2 photocatalysis can increase the biodegradability of persistent PPCPs. In most cases, complete degradation of the parent compounds does not correspond directly to the mineralization rate, generating more stable by-products during the treatment. TOC reduction indicates breaking larger molecules into smaller ones and the decomposition of aromatic compounds, which can then increase the aliphatic hydrocarbon fraction. A high aliphatic fraction suggests a larger breakdown of organic compounds by OH [42,43]. In our experiments, we observed 26% and 30% reductions in TOC during 120 min photocatalytic reaction times in wastewater effluent and surface water, respectively. Limited removal of the background TOC was observed in previous studies as well (less than 20%) [44,45]. However, UV absorbance at λ = 254 nm (77% in synthetic surface water) suggested that the parent TOC molecules were oxidized, forming by-products that can also scavenge OH [46].
It should be noted that OH oxidation is not the only pathway for degrading PPCPs via TiO2 photocatalysis. Electrons and holes on the TiO2 surface can initiate PPCP degradation, alternatively referred to as the single electron transfer (SET) pathway. Both OH oxidation and SET pathways contribute to ring-opening reactions, but SET is attributed to the ring-opening reactions of hydroxylated aromatic compounds, particularly orthodihyroxylated ones. OH participates in OH substitution, alkyl group degradation, and ring hydroxylation [47,48,49]. Understanding the mechanistic pathway is beyond the scope of the current study since it involves comprehensive analysis of by-product formation during photocatalytic reactions.

2.2.2. Effect of Physicochemical Characteristics

Physicochemical properties such as charge, molecular weight, hydrophobicity, and molecular structure can affect photocatalytic degradation. Previous studies, including ours, show that the chemical structure of PPCPs can be correlated with their photocatalytic degradation [6,9]. Arlos et al. [9] studied the degradation of 16 PPCPs using TiO2 immobilized on porous support in ultrapure water under UV-LED exposure and indicated that ionic interactions can determine PPCP degradation. Cationic and anionic PPCPs interact well with oppositely charged membranes, while neutral compounds were found to be recalcitrant under any treatment. Liang et al. [6] studied the kinetic rate constants of degradation of the 18 PPCPs with different net charges, molecular weights, and solubilities on the PPCPs’ degradation rates and concluded that charge was the main factor determining the degradation rate constant among the other three parameters. Other studies focused on the correlation between chemical bonding and photodegradation under UV irradiation (photolysis) and suggested that chemicals with amide bonds (R–CO–NR2) are poorly degraded via TiO2 photocatalysis [50,51] whereas PPCPs with amine bonds (RxNH3−x) such as sulfamethoxazole showed higher degradation rate constants [50]. In another study, varying rate constants for trace organics with amine groups have been observed, but further concluded that this is not the main bond affected by UV irradiation during photocatalytic reaction [52]. In our study, we did not find consistent relationships between the PPCPs’ functional groups or the number of aromatic rings with degradation rate constants in any of the matrices (Table S4). However, our analysis showed that log Dow is well correlated with TiO2 photocatalysis in ultrapure water but that this was not the case for synthetic and wastewaters (Figure 4; Tables S4–S6). Instead, molecular weight may be more predictive of TiO2 photocatalysis in these background water matrices. Our result was unexpected, and we recognize that additional studies on several other matrices are needed to validate this finding.

2.3. Electrical Energy per Order Required for Removal

Energy efficiency of UV-based reactions is commonly reported using electrical energy per order (EEO), which is defined as the kilowatt hours of electrical energy needed to decrease contaminant concentration by the first order of magnitude in 1 m3 of polluted water [53] (Equation (S3) in the SI).
On average, we found that the energy required to remove the mixture of PPCPs in wastewater effluents (16.36 kW h m−3) was higher than that for synthetic surface water and ultrapure water (4.77 and 2.96 kW h m−3, respectively) (Figure 5). The energy required for removing individual PPCPs in ultrapure water and synthetic surface water was not significantly different except for atorvastatin, bisphenol A, naproxen, and sulfamethoxazole (p-values in Table S8). Among our target PPCPs, atrazine consumed the highest energy required to remove in wastewater effluents (76.37 kW h m−3). It was completely removed in ultrapure water and synthetic surface water in 2 h treatment, but complete removal of atrazine in wastewater effluents required prolonged photocatalytic treatment. In synthetic surface water, removal of bisphenol A required the highest EEO value of 32.84 kW h m−3. Bisphenol A has been reported to degrade at a higher rate in acidic conditions (pH < 6.5). The higher pH in synthetic surface water (pHSSW = 8) may affect the photocatalytic degradation of BPA, increasing the EEO value [54].
It should be noted that EEO values for PPCP removal in this study are comparatively lower than that for previous publications, suggesting that UV-LED/TiO2 is an energy efficient process. For instance, the EEO values of 50–70 kW h m−3 were reported by Paredes et al. [55] for degradation of gemfibrozil, carbamazepine, metoprolol, trimethoprim, and warfarin in secondary wastewater effluents using P25 and a low-pressure UV lamp (40 W). Poulopoulos et al. [56] also reported the EEO values of 121–250 kW h m−3 for degradations of phenol, 2-chlorophenol, 2,4 dichlorophenol, 2,4,6, trichorophenol, and 4-nitrophenol in synthetic wastewater using P25 and a UV lamp (6 W).

3. Materials and Methods

3.1. Water Matrices

Three different water matrices were used as the test matrix for the experiments: (i) ultrapure water (UPW) (resistivity of 18.2 mΩ·cm at 25 °C, Milli-Q Advantage A10 Water Purification System, EMD Millipore, Burlington, MA, USA); (ii) synthetic surface water (SSW) that was produced in the lab using Suwannee River NOM reverse osmosis isolate (International Humic Substances Society, City and County of Denver, CO, USA), alginic acid, and inorganic ions, published previously in our study [7,46] (purchased from Sigma Aldrich, St. Louis, MO, USA; purity ≥ 99%); and (iii) wastewater effluent (WWE) collected from the Waterloo Regional Wastewater Treatment Plant, stored in ice, and then filtered using a 2 μm glass fiber filter (Whatman, Maidstone, UK) to remove large suspended particulate matter. Maxxam Analytics (Mississauga, ON, Canada) analyzed the wastewater effluent characteristics including total organic carbon (TOC), conductivity, dissolved chloride, ammonia, nitrite, and nitrate. The TOC of the synthetic surface water was measured in filtered (polyether sulfone membrane filters; 0.45 μm) samples using a Shimadzu TOC-L total organic carbon analyzer supplied with 680 °C combustion catalytic oxidation, NDIR detection method, and a Shimadzu ASI-L autosampler.

3.2. Photocatalytic Degradation of PPCPs

Commercially available TiO2 (P25 AeroxideTM) was used as the model photocatalyst nanomaterial. Twenty-four pharmaceutical and personal care products with different properties and applications were selected as target compounds for the photocatalytic experiments (Table S9). PPCPs and isotopically labeled standards were purchased from different suppliers as detailed in Table S10, dissolved in methanol stock solution (1 g L−1), and stored at −20 °C.
Three glass beakers covered with aluminum foil were spiked from PPCP stock solution to create a nominal concentration of 2 µg L−1 of each compound. The compounds in methanol were first evaporated to dryness to reduce carrier solution concentration by Arlos et al. [10] and then 300 mL of the desired water matrix (ultrapure, synthetic surface water, or wastewater effluent) was added to the beakers followed by 30 mg of P25 and stirred at 600 rpm. The experimental setup was similar to previous work by Arlos et al. [9]. Briefly, the multi-positioned magnetic stir plate was placed 10.5 cm below the in-house assembled UV-LED beam (θbeam = 4 cm, λ = 365 nm) setup. The rated power of UV-LED lights was 2060 at 1000 mA (measured using Mouser Electronics, Mansfield, TX, Canada). The average power output of the UV-LEDs was 0.390 mW cm−2 determined by Thorlabs power and energy meter (PM100-USB). A 60 min equilibration period in the absence of UV light was used followed by 120 min UV-LED light irradiation. Aliquots of beakers (4.5 mL) were taken at several time points from −60 to 120 min.
The PPCP degradation experiments were conducted under six conditions with three replicates per treatment: (i) dark test in ultrapure water; (ii) photolysis in ultrapure water; (iii) photolysis in wastewater effluents; and P25 photocatalysis in (iv) ultrapure water, (v) wastewater effluents, and (vi) synthetic surface water. Sample preparation and analysis are explained in detail by Arlos et al. [9]. Data processing and principal component analysis were completed using Origin Pro (version 8). Statistical analysis was performed using one-way ANOVA, Tukey tests, and α = 0.05 Sigma Plot (version 13).

4. Conclusions

This study focused on the impact of the background water matrix on TiO2 photocatalytic treatment of 24 target PPCPs. Specifically, the photocatalytic degradation of PPCPs were compared in ultrapure water, synthetic surface water, and wastewater effluents. The results showed a substantial decrease in PPCP removal kinetics in synthetic surface water and wastewater effluent matrices. Out of the 24 PPCPs, 29% and 12% were removed quickly in ultrapure water and surface water, while all of the PPCPs were removed poorly (<0.1 min−1) in wastewater effluents. In addition, caffeine and carbamazepine removal kinetics via photocatalysis were not influenced by the background water matrix. Low rate constants of PPCPs in complex water matrices highlight the importance of the background water on photocatalytic degradation, which is usually neglected by many previous studies (we examined 80 in total). Furthermore, 26% and 30% TOC reductions in wastewater effluents and surface water were observed, suggesting that complete degradation did not occur during the 120 min irradiation.
Among the PPCP physicochemical characteristics that we evaluated, the pH-corrected n-octanol/water partition coefficient (log Dow) was correlated with the PPCP degradation rate constants in ultrapure water, and molecular weight was strongly associated with the rate constants in synthetic surface water and wastewater. As expected, the average electrical energy consumption for removing PPCPs in wastewater effluents were higher than that for synthetic water and ultrapure water (16.36 vs. 4.77 and 2.96 kW h m−3, respectively). Low EEO values in our study suggest that UV-LED-irradiated TiO2 is a promising energy-efficient technology for PPCP degradation in water.
Finally, since there is an increasing trend in the synthesis of novel nanomaterials for photocatalytic applications, we highly recommend that researchers closely investigate the impact of water matrices, especially during the “proof-of-concept” stage of nanomaterials development for the technology, to become more applicable in the future.

Supplementary Materials

The supplementary materials are available online at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/catal11050576/s1. Figure S1: Removal of targeted PPCPs in ultrapure water (UPW), wastewater effluent (WWE), and synthetic surface water (SSW) under dark (120 min dark irradiation using P25), photolysis (120 min under UV-LED irradiation without P25) and photocatalysis (120 min under UV-LED irradiation and using P25) treatments; Figure S2: Removal of targeted PPCPs in ultrapure water (UPW), wastewater effluent (WWE), and synthetic surface water (SSW) under dark (120 min dark irradiation using P25), photolysis (120 min under UV-LED irradiation without P25) and photocatalysis (120 min under UV-LED irradiation and using P25) treatments. Table S1: Studies in photocatalytic degradation of pharmaceuticals and personal care compounds in different water matrices; Table S2: Pharmaceuticals and personal care products (PPCP) and their abbreviations; Table S3: p-values for One-Way ANOVA tests (SigmaPlot, α = 0.05) during 60 min equilibrium period, photolysis and photocatalysis.Tukey tests (multiple comparisons) were conducted when a statistical significance was detected with overall statistical significance level of 0.05; Table S4: PPCPs types and number of bonds; Table S5: Correlation coefficient in PCA matrix; Table S6: Extracted Eigenvectors in three water matrices; Table S7: Total organic carbon for wastewater effluent upon collection from the Waterloo treatment plant and after photocatalytic tests were performed on the matrix; Table S8: p-values for One-Way ANOVA tests (SigmaPlot, α = 0.05) for EEO. Tukey tests (multiple comparisons) were conducted when a statistical significance was detected with overall statistical significance level of 0.05; Table S9: Chemical and physical properties of targeted PPCPs in three water matrices and their concentrations UPW = ultrapure water; SSW = synthetic surface water; WWE = wastewater effluent; Table S10. List of PPCPs, isotopically labeled standard and suppliers.

Author Contributions

Data curation, I.J.-Z.; Formal analysis, N.S.; Funding acquisition, N.Z. and M.S.; Methodology, L.B. and R.L.; Supervision, N.Z., M.S. and M.A.; Writing–review & editing, A.F. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Canada Research Chair (CRC) in water-quality protection and advanced material joining and processing, Natural Sciences and Engineering Research Council (NSERC) of Canada, Strategic Project Grant (STPGP 430654-12). The authors would like to acknowledge the assistance provided by Schwartz-Resiman Foundation under the Waterloo-Technion Research Co-operation Program.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, L.L.; Xiong, Z. Pillaring Chemically Exfoliated Graphene Oxide with Carbon Nanotubes for Photocatalytic Degradation of Dyes under Visible Light Irradiation. ACS Nano 2010, 4, 7030–7036. [Google Scholar] [CrossRef] [PubMed]
  2. Ranjit, K.T.; Viswanathan, B. Synthesis, Characterization and Photocatalytic Properties of Iron-Doped TiO2 Catalysts. J. Photochem. Photobiol. A Chem. 1997, 108, 79–84. [Google Scholar] [CrossRef]
  3. Klavarioti, M.; Mantzavinos, D.; Kassinos, D. Removal of Residual Pharmaceuticals from Aqueous Systems by Advanced Oxidation Processes. Environ. Int. 2009, 35, 402–417. [Google Scholar] [CrossRef] [PubMed]
  4. Gomes, J.; Lincho, J.; Domingues, E.; Quinta-Ferreira, R.M.; Martins, R.C. N–TiO2 Photocatalysts: A Review of Their Characteristics and Capacity for Emerging Contaminants Removal. Water 2019, 11, 373. [Google Scholar] [CrossRef] [Green Version]
  5. Malato, S.; Fernández-Ibáñez, P.; Maldonado, M.I.; Blanco, J.; Gernjak, W. Decontamination and Disinfection of Water by Solar Photocatalysis: Recent Overview and Trends. Catal. Today 2009, 147, 1–59. [Google Scholar] [CrossRef]
  6. Liang, R.; Van Leuwen, J.C.; Bragg, L.M.; Arlos, M.J.; Fong, L.C.M.L.C.; Schneider, O.M.; Jaciw-Zurakowsky, I.; Fattahi, A.; Rathod, S.; Peng, P. Utilizing UV-LED Pulse Width Modulation on TiO2 Advanced Oxidation Processes to Enhance the Decomposition Efficiency of Pharmaceutical Micropollutants. Chem. Eng. J. 2019, 361, 439–449. [Google Scholar] [CrossRef]
  7. Fattahi, A.; Arlos, M.J.; Bragg, L.M.; Kowalczyk, S.; Liang, R.; Schneider, O.M.; Zhou, N.; Servos, M.R. Photodecomposition of Pharmaceuticals and Personal Care Products Using P25 Modified with Ag Nanoparticles in the Presence of Natural Organic Matter. Sci. Total Environ. 2020, 752, 142000. [Google Scholar] [CrossRef] [PubMed]
  8. Arlos, M.J.; Liang, R.; Hatat-Fraile, M.M.; Bragg, L.M.; Zhou, N.Y.; Servos, M.R.; Andrews, S.A. Photocatalytic Decomposition of Selected Estrogens and Their Estrogenic Activity by UV-LED Irradiated TiO2 Immobilized on Porous Titanium Sheets via Thermal-Chemical Oxidation. J. Hazard. Mater. 2016, 318, 541–550. [Google Scholar] [CrossRef]
  9. Arlos, M.J.; Hatat-Fraile, M.M.; Liang, R.; Bragg, L.M.; Zhou, N.Y.; Andrews, S.A.; Servos, M.R. Photocatalytic Decomposition of Organic Micropollutants Using Immobilized TiO2 Having Different Isoelectric Points. Water Res. 2016, 101, 351–361. [Google Scholar] [CrossRef]
  10. Arlos, M.J.; Liang, R.; Fong, L.C.M.; Zhou, N.Y.; Ptacek, C.J.; Andrews, S.A.; Servos, M.R. Influence of Methanol When Used as a Water-Miscible Carrier of Pharmaceuticals in TiO2 Photocatalytic Degradation Experiments. J. Environ. Chem. Eng. 2017, 5, 4497–4504. [Google Scholar] [CrossRef]
  11. Choi, J.; Lee, H.; Choi, Y.; Kim, S.; Lee, S.; Lee, S.; Choi, W.; Lee, J. Heterogeneous Photocatalytic Treatment of Pharmaceutical Micropollutants: Effects of Wastewater Effluent Matrix and Catalyst Modifications. Appl. Catal. B Environ. 2014, 147, 8–16. [Google Scholar] [CrossRef]
  12. Autin, O.; Hart, J.; Jarvis, P.; MacAdam, J.; Parsons, S.A.; Jefferson, B. The Impact of Background Organic Matter and Alkalinity on the Degradation of the Pesticide Metaldehyde by Two Advanced Oxidation Processes: UV/H2O2 and UV/TiO2. Water Res. 2013, 47, 2041–2049. [Google Scholar] [CrossRef] [PubMed]
  13. Zepp, R.G.; Baughman, G.L.; Schlotzhauer, P.F. Comparison of Photochemical Behavior of Various Humic Substances in Water: I. Sunlight Induced Reactions of Aquatic Pollutants Photosensitized by Humic Substances. Chemosphere 1981, 10, 109–117. [Google Scholar] [CrossRef]
  14. Pereira, V.J.; Linden, K.G.; Weinberg, H.S. Evaluation of UV Irradiation for Photolytic and Oxidative Degradation of Pharmaceutical Compounds in Water. Water Res. 2007, 41, 4413–4423. [Google Scholar] [CrossRef]
  15. Luo, Y.; Guo, W.; Ngo, H.H.; Nghiem, L.D.; Hai, F.I.; Zhang, J.; Liang, S.; Wang, X.C. A Review on the Occurrence of Micropollutants in the Aquatic Environment and Their Fate and Removal during Wastewater Treatment. Sci. Total Environ. 2014, 473, 619–641. [Google Scholar] [CrossRef] [PubMed]
  16. Haroune, L.; Salaun, M.; Ménard, A.; Legault, C.Y.; Bellenger, J.-P. Photocatalytic Degradation of Carbamazepine and Three Derivatives Using TiO2 and ZnO: Effect of PH, Ionic Strength, and Natural Organic Matter. Sci. Total Environ. 2014, 475, 16–22. [Google Scholar] [CrossRef]
  17. Guillard, C.; Puzenat, E.; Lachheb, H.; Houas, A.; Herrmann, J.-M. Why Inorganic Salts Decrease the TiO2 Photocatalytic Efficiency. Int. J. Photoenergy 2005, 7, 641208. [Google Scholar] [CrossRef] [Green Version]
  18. Schäfer, A.I.; Akanyeti, I.; Semião, A.J.C. Micropollutant Sorption to Membrane Polymers: A Review of Mechanisms for Estrogens. Adv. Colloid Interface Sci. 2011, 164, 100–117. [Google Scholar] [CrossRef]
  19. Liang, R.; Li Chun Fong, L.C.M.; Arlos, M.J.; Van Leeuwen, J.; Shahnam, E.; Peng, P.; Servos, M.R.; Zhou, Y.N. Photocatalytic Degradation Using One-Dimensional TiO2 and Ag-TiO2 Nanobelts under UV-LED Controlled Periodic Illumination. J. Environ. Chem. Eng. 2017, 5, 4365–4373. [Google Scholar] [CrossRef]
  20. Mena, E.; Rey, A.; Beltrán, F.J. TiO2 Photocatalytic Oxidation of a Mixture of Emerging Contaminants: A Kinetic Study Independent of Radiation Absorption Based on the Direct-Indirect Model. Chem. Eng. J. 2018, 339, 369–380. [Google Scholar] [CrossRef]
  21. Ribao, P.; Corredor, J.; Rivero, M.J.; Ortiz, I. Role of Reactive Oxygen Species on the Activity of Noble Metal-Doped TiO2 Photocatalysts. J. Hazard. Mater. 2019, 372, 45–51. [Google Scholar] [CrossRef] [PubMed]
  22. Piscopo, A.; Robert, D.; Weber, J.V. Influence of PH and Chloride Anion on the Photocatalytic Degradation of Organic Compounds: Part I. Effect on the Benzamide and Para-Hydroxybenzoic Acid in TiO2 Aqueous Solution. Appl. Catal. B Environ. 2001, 35, 117–124. [Google Scholar] [CrossRef]
  23. Wiszniowski, J.; Robert, D.; Surmacz-Gorska, J.; Miksch, K.; Weber, J.V. Photocatalytic Mineralization of Humic Acids with TiO2: Effect of PH, Sulfate and Chloride Anions. Int. J. Photoenergy 2003, 5, 69–74. [Google Scholar] [CrossRef] [Green Version]
  24. Abdullah, M.; Low, G.K.C.; Matthews, R.W. Effects of Common Inorganic Anions on Rates of Photocatalytic Oxidation of Organic Carbon over Illuminated Titanium Dioxide. J. Phys. Chem. 1990, 94, 6820–6825. [Google Scholar] [CrossRef]
  25. Chen, H.Y.; Zahraa, O.; Bouchy, M. Inhibition of the Adsorption and Photocatalytic Degradation of an Organic Contaminant in an Aqueous Suspension of TiO2 by Inorganic Ions. J. Photochem. Photobiol. A Chem. 1997, 108, 37–44. [Google Scholar] [CrossRef]
  26. Wang, K.-H.; Hsieh, Y.-H.; Wu, C.-H.; Chang, C.-Y. The PH and Anion Effects on the Heterogeneous Photocatalytic Degradation of O-Methylbenzoic Acid in TiO2 Aqueous Suspension. Chemosphere 2000, 40, 389–394. [Google Scholar] [CrossRef]
  27. Hem, J.D. Study and Interpretation of the Chemical Characteristics of Natural Water; Department of the Interior, US Geological Survey: Washington, DC, USA, 1985; Volume 2254. [Google Scholar]
  28. Brezonik, P.L.; Fulkerson-Brekken, J. Nitrate-Induced Photolysis in Natural Waters: Controls on Concentrations of Hydroxyl Radical Photo-Intermediates by Natural Scavenging Agents. Environ. Sci. Technol. 1998, 32, 3004–3010. [Google Scholar] [CrossRef]
  29. Cavalcante, R.P.; Dantas, R.F.; Wender, H.; Bayarri, B.; González, O.; Giménez, J.; Esplugas, S.; Machulek, A., Jr. Photocatalytic Treatment of Metoprolol with B-Doped TiO2: Effect of Water Matrix, Toxicological Evaluation and Identification of Intermediates. Appl. Catal. B Environ. 2015, 176, 173–182. [Google Scholar] [CrossRef]
  30. Buxton, G.V.; Greenstock, C.L.; Helman, W.P.; Ross, A.B. Critical Review of Rate Constants for Reactions of Hydrated Electrons, Hydrogen Atoms and Hydroxyl Radicals (⋅OH/⋅O− in Aqueous Solution. J. Phys. Chem. Ref. Data 1988, 17, 513–886. [Google Scholar] [CrossRef] [Green Version]
  31. Liu, Y.; He, X.; Duan, X.; Fu, Y.; Dionysiou, D.D. Photochemical Degradation of Oxytetracycline: Influence of PH and Role of Carbonate Radical. Chem. Eng. J. 2015, 276, 113–121. [Google Scholar] [CrossRef]
  32. Lai, W.W.-P.; Hsu, M.-H.; Lin, A.Y.-C. The Role of Bicarbonate Anions in Methotrexate Degradation via UV/TiO2: Mechanisms, Reactivity and Increased Toxicity. Water Res. 2017, 112, 157–166. [Google Scholar] [CrossRef] [PubMed]
  33. Hu, L.; Flanders, P.M.; Miller, P.L.; Strathmann, T.J. Oxidation of Sulfamethoxazole and Related Antimicrobial Agents by TiO2 Photocatalysis. Water Res. 2007, 41, 2612–2626. [Google Scholar] [CrossRef] [PubMed]
  34. Canonica, S.; Kohn, T.; Mac, M.; Real, F.J.; Wirz, J.; von Gunten, U. Photosensitizer Method to Determine Rate Constants for the Reaction of Carbonate Radical with Organic Compounds. Environ. Sci. Technol. 2005, 39, 9182–9188. [Google Scholar] [CrossRef]
  35. Zhang, G.; He, X.; Nadagouda, M.N.; O’Shea, K.E.; Dionysiou, D.D. The Effect of Basic PH and Carbonate Ion on the Mechanism of Photocatalytic Destruction of Cylindrospermopsin. Water Res. 2015, 73, 353–361. [Google Scholar] [CrossRef]
  36. Liu, S.; Lim, M.; Fabris, R.; Chow, C.; Drikas, M.; Amal, R. TiO2 Photocatalysis of Natural Organic Matter in Surface Water: Impact on Trihalomethane and Haloacetic Acid Formation Potential. Environ. Sci. Technol. 2008, 42, 6218–6223. [Google Scholar] [CrossRef] [PubMed]
  37. Ndlangamandla, N.G.; Kuvarega, A.T.; Msagati, T.A.M.; Mamba, B.B.; Nkambule, T.T.I. A Novel Photodegradation Approach for the Efficient Removal of Natural Organic Matter (NOM) from Water. Phys. Chem. Earth 2018, 106, 97–106. [Google Scholar] [CrossRef]
  38. Gora, S.; Sokolowski, A.; Hatat-Fraile, M.; Liang, R.; Zhou, Y.N.; Andrews, S. Solar Photocatalysis with Modified TiO2 Photocatalysts: Effects on NOM and Disinfection Byproduct Formation Potential. Environ. Sci. Water Res. Technol. 2018, 4, 1361–1376. [Google Scholar] [CrossRef]
  39. Garbin, J.R.; Milori, D.M.B.P.; Simoes, M.L.; Da Silva, W.T.L.; Neto, L.M. Influence of Humic Substances on the Photolysis of Aqueous Pesticide Residues. Chemosphere 2007, 66, 1692–1698. [Google Scholar] [CrossRef]
  40. Doll, T.E.; Frimmel, F.H. Fate of Pharmaceuticals—Photodegradation by Simulated Solar UV-Light. Chemosphere 2003, 52, 1757–1769. [Google Scholar] [CrossRef]
  41. Carlos, L.; Mártire, D.O.; Gonzalez, M.C.; Gomis, J.; Bernabeu, A.; Amat, A.M.; Arques, A. Photochemical Fate of a Mixture of Emerging Pollutants in the Presence of Humic Substances. Water Res. 2012, 46, 4732–4740. [Google Scholar] [CrossRef]
  42. Hoffmann, M.R.; Martin, S.T.; Choi, W.; Bahnemann, D.W. Environmental Applications of Semiconductor Photocatalysis. Chem. Rev. 1995, 95, 69–96. [Google Scholar] [CrossRef]
  43. Maeng, S.K.; Cho, K.; Jeong, B.; Lee, J.; Lee, Y.; Lee, C.; Choi, K.J.; Hong, S.W. Substrate-Immobilized Electrospun TiO2 Nanofibers for Photocatalytic Degradation of Pharmaceuticals: The Effects of PH and Dissolved Organic Matter Characteristics. Water Res. 2015, 86, 25–34. [Google Scholar] [CrossRef] [PubMed]
  44. Chen, H.W.; Ku, Y.; Irawan, A. Photodecomposition of O-Cresol by UV-LED/TiO2 Process with Controlled Periodic Illumination. Chemosphere 2007, 69, 184–190. [Google Scholar] [CrossRef] [PubMed]
  45. Fattahi, A.; Arlos, M.J.; Bragg, L.M.; Liang, R.; Zhou, N.; Servos, M.R. Degradation of Natural Organic Matter Using Ag-P25 Photocatalyst under Continuous and Periodic Irradiation of 405 and 365 Nm UV-LEDs. J. Environ. Chem. Eng. 2021, 9, 104844. [Google Scholar] [CrossRef]
  46. Lamsal, R.; Walsh, M.E.; Gagnon, G.A. Comparison of Advanced Oxidation Processes for the Removal of Natural Organic Matter. Water Res. 2011, 45, 3263–3269. [Google Scholar] [CrossRef]
  47. Madras, G. Kinetics of Simultaneous Photocatalytic Degradation of Phenolic Compounds and Reduction of Metal Ions with Nano-TiO2. Environ. Sci. Technol. 2008, 42, 913–919. [Google Scholar]
  48. Ganiyu, S.O.; Van Hullebusch, E.D.; Cretin, M.; Esposito, G.; Oturan, M.A. Coupling of Membrane Filtration and Advanced Oxidation Processes for Removal of Pharmaceutical Residues: A Critical Review. Sep. Purif. Technol. 2015, 156, 891–914. [Google Scholar] [CrossRef]
  49. Lu, M. Photocatalysis and Water Purification: From Fundamentals to Recent Applications; John Wiley & Sons: Hoboken, NJ, USA, 2013. [Google Scholar]
  50. Hu, A.; Zhang, X.; Luong, D.; Oakes, K.D.; Servos, M.R.; Liang, R.; Kurdi, S.; Peng, P.; Zhou, Y. Adsorption and Photocatalytic Degradation Kinetics of Pharmaceuticals by TiO2 Nanowires during Water Treatment. Waste Biomass Valoriz. 2012, 3, 443–449. [Google Scholar] [CrossRef]
  51. Kim, I.; Tanaka, H. Photodegradation Characteristics of PPCPs in Water with UV Treatment. Environ. Int. 2009, 35, 793–802. [Google Scholar] [CrossRef]
  52. Fernández, R.L.; McDonald, J.A.; Khan, S.J.; Le-Clech, P. Removal of Pharmaceuticals and Endocrine Disrupting Chemicals by a Submerged Membrane Photocatalysis Reactor (MPR). Sep. Purif. Technol. 2014, 127, 131–139. [Google Scholar] [CrossRef]
  53. Cater, S.R.; Stefan, M.I.; Bolton, J.R.; Safarzadeh-Amiri, A. UV/H2O2 Treatment of Methyl Tert-Butyl Ether in Contaminated Waters. Environ. Sci. Technol. 2000, 34, 659–662. [Google Scholar] [CrossRef]
  54. Abdelraheem, W.H.M.; Patil, M.K.; Nadagouda, M.N.; Dionysiou, D.D. Hydrothermal Synthesis of Photoactive Nitrogen-and Boron-Codoped TiO2 Nanoparticles for the Treatment of Bisphenol A in Wastewater: Synthesis, Photocatalytic Activity, Degradation Byproducts and Reaction Pathways. Appl. Catal. B Environ. 2019, 241, 598–611. [Google Scholar] [CrossRef]
  55. Paredes, L.; Murgolo, S.; Dzinun, H.; Othman, M.H.D.; Ismail, A.F.; Carballa, M.; Mascolo, G. Application of Immobilized TiO2 on PVDF Dual Layer Hollow Fibre Membrane to Improve the Photocatalytic Removal of Pharmaceuticals in Different Water Matrices. Appl. Catal. B Environ. 2019, 240, 9–18. [Google Scholar] [CrossRef]
  56. Poulopoulos, S.G.; Yerkinova, A.; Ulykbanova, G.; Inglezakis, V.J. Photocatalytic Treatment of Organic Pollutants in a Synthetic Wastewater Using UV Light and Combinations of TiO2, H2O2 and Fe (III). PLoS ONE 2019, 14, e0216745. [Google Scholar] [CrossRef]
Figure 1. Removal of PPCP in ultrapure water (UPW), synthetic surface water (SSW), and wastewater effluent (WWE) during the equilibration period (60 min in the dark and at room temperature). Negative values are within the 15% acceptable experimental error.
Figure 1. Removal of PPCP in ultrapure water (UPW), synthetic surface water (SSW), and wastewater effluent (WWE) during the equilibration period (60 min in the dark and at room temperature). Negative values are within the 15% acceptable experimental error.
Catalysts 11 00576 g001
Figure 2. TiO2 functional groups and their adsorption tendency toward PPCPs in acidic and basic environments. In acidic conditions, negatively charged PPCPs tend to sorb to TiOH2+ (e.g., atorvastatin), whereas in basic conditions, positively charged PPCPs tend to interact with TiO. See Figure 1 for the adsorption results of other target compounds.
Figure 2. TiO2 functional groups and their adsorption tendency toward PPCPs in acidic and basic environments. In acidic conditions, negatively charged PPCPs tend to sorb to TiOH2+ (e.g., atorvastatin), whereas in basic conditions, positively charged PPCPs tend to interact with TiO. See Figure 1 for the adsorption results of other target compounds.
Catalysts 11 00576 g002
Figure 3. Pseudo first order rate constants for all target compounds during 120 min P25 photocatalysis in ultrapure water (UPW), synthetic surface water (SSW), and wastewater effluent (WWE). Inset magnifies the PPCPs with low kapp.
Figure 3. Pseudo first order rate constants for all target compounds during 120 min P25 photocatalysis in ultrapure water (UPW), synthetic surface water (SSW), and wastewater effluent (WWE). Inset magnifies the PPCPs with low kapp.
Catalysts 11 00576 g003
Figure 4. Principal component analysis (PCA) biplot showing the multivariate variation among 24 PPCPs in terms of physicochemical properties (net charge, weight, solubility, and log Dow) in (a) ultrapure water (b) synthetic surface water (c) wastewater effluents. Blue vectors indicate the direction and strength of each variable (physicochemical properties) and their correlation to kapp (green vector).
Figure 4. Principal component analysis (PCA) biplot showing the multivariate variation among 24 PPCPs in terms of physicochemical properties (net charge, weight, solubility, and log Dow) in (a) ultrapure water (b) synthetic surface water (c) wastewater effluents. Blue vectors indicate the direction and strength of each variable (physicochemical properties) and their correlation to kapp (green vector).
Catalysts 11 00576 g004
Figure 5. Values of electrical energy per order (EEO) required for removing PPCPs in ultrapure water (UPW), synthetic surface water (SSW), and wastewater effluent (WWE).
Figure 5. Values of electrical energy per order (EEO) required for removing PPCPs in ultrapure water (UPW), synthetic surface water (SSW), and wastewater effluent (WWE).
Catalysts 11 00576 g005
Table 1. Average water-quality characteristics of synthetic surface water and wastewater effluent.
Table 1. Average water-quality characteristics of synthetic surface water and wastewater effluent.
ParametersUnitsSynthetic Surface WaterWastewater Effluent
pH 8–8.57–7.5
TOCmg L−16.129.37
Dissolved Chloridemg L−140490
Hardnessmg L−1 as CaCO3114N/A
Alkalinitymg L−1 as CaCO3117N/A
Nitratemg L−1336.53
Nitritemg L−1N/A0.53
Sulfatemg L−1241N/A
N/A: Not available.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Fattahi, A.; Jaciw-Zurakowsky, I.; Srikanthan, N.; Bragg, L.; Liang, R.; Zhou, N.; Servos, M.; Arlos, M. Effect of Background Water Matrices on Pharmaceutical and Personal Care Product Removal by UV-LED/TiO2. Catalysts 2021, 11, 576. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11050576

AMA Style

Fattahi A, Jaciw-Zurakowsky I, Srikanthan N, Bragg L, Liang R, Zhou N, Servos M, Arlos M. Effect of Background Water Matrices on Pharmaceutical and Personal Care Product Removal by UV-LED/TiO2. Catalysts. 2021; 11(5):576. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11050576

Chicago/Turabian Style

Fattahi, Azar, Ivana Jaciw-Zurakowsky, Nivetha Srikanthan, Leslie Bragg, Robert Liang, Norman Zhou, Mark Servos, and Maricor Arlos. 2021. "Effect of Background Water Matrices on Pharmaceutical and Personal Care Product Removal by UV-LED/TiO2" Catalysts 11, no. 5: 576. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11050576

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