In this study, the degradations of 2,3,4,5,6-pentabromotoluene (PBT), 2,3,4,5,6-pentabromoethyl benzene (PBEB), triclosan (TCS) and gemfibrozil (GFZ) in raw hospital wastewater were investigated with cerium (IV) oxide and titanium (IV) oxide nanoparticles considering the mechanisms of adsorption, photolysis, and photocatalysis with UV-C lamps. The effects of nano-CeO2 and nano-TiO2 concentrations, irradiation times, UV light powers and hospital wastewater pH on the photodegradation yields of micropollutants namely PBT, PBEB, TCS and GFZ were investigated throughout photocatalysis. The nano-TiO2 produced had an anatase phase with crystalline shape with a surface area of 205 m2 g−1 and an average size of 11.50 nm. The CeO2 nanoparticles had a spherical shape with a higher surface area (302 m2 g−1) than that of TiO2 and a lower average size (8.11 nm). It was found that the removals of PBT, PBEB, TCS and GFZ with adsorption (5.7%–17.1%) and photolysis (9.0%–15.9%) were not significant for both nanoparticles. The photodegradation of PBT (92%), PBEB (90%), TCS (97%) and GFZ (95%) with nano-CeO2 gave better results than nano-TiO2 (90%, 87%, 94% and 93% for PBT, PBEB, TCS and GFZ, respectively) under optimum experimental conditions (0.50 g L−1 nano-CeO2, 45 min irradiation time, 25 °C temperature, pH = 8.50, 210 W UV light power). Both nanoparticles were reused effectively after photo-removals of the micropollutants from the hospital wastewater. The lowest photodegradation yields were 80%, 78%, 75% and 74% for TCS, GFZ, PBT and PBEB, respectively, with nano-TiO2 after six sequential treatments. The lowest photodegradation yields were 86%, 83%, 80% and 79% for the same micropollutants, respectively, with nano-CeO2 after six sequential treatments. The cost to treat 1 m3 raw hospital wastewater were 8.70 € and 2.28 €, for the photocatalytic treatments with nano-TiO2 and nano-CeO2, respectively.

Micropollutants in hospital wastewater are directly discharged into the sewage system without treatment since the conventional sewage/urban treatment plants can only remove macropollutants such as biochemical oxygen demand, chemical oxygen demand (COD), nitrogen and phosphorus in Turkey. The micropollutants decrease the biological treatment efficiency if they are not treated and then discharged into the receiving environment (Petrie et al. 2014).

Gemfibrozil (GFZ) is a blood lipid regulator and is used to treat hypertriglyceridemia (Grenni et al. 2013). It is frequently detected in surface water and wastewater (Krkošek et al. 2011). GFZ is toxic to aquatic microorganisms and may also cause endocrine disruption in fish (Chen et al. 2017). It has been found in wastewater treatment plants and also in freshwaters at a concentration ranging from 0.07 to 0.51 μg L−1 (Fang et al. 2012). Triclosan (TCS) is a broad-spectrum antibacterial and antifungal agent (Xie et al. 2012) used in hospitals. It has been found in surface waters, wastewaters, sediments and sludges due its widespread usage. It has low acute and chronic toxicities (Tohidi & Cai 2017). The concentration of TCS has been found to range from 1.4 up to 133,000 ng L−1 in surface waters, and from 20 to 133,000 μg kg−1 dw in biosolids (Montaseri & Forbes 2016). 2,3,4,5,6-Pentabromoethyl benzene (PBEB) and 2,3,4,5,6-pentabromotoluene (PBT) are new brominated flame retardants (NBFRs) and have been used in order to prevent fires and in the formation of different types of polymers, resins and plastics (Cristale et al. 2012; Chen et al. 2012). NBFRs are of environmental concern because of their high lipophilicity, persistence and resistance to degradation (Barón et al. 2015). Their bioaccumulative and carcinogenic/mutagenic properties make them potentially dangerous for environmental health (Sun et al. 2015). PBEB was detected in samples of river sediments at concentrations ranging from 3.1 to 10.0 ng g−1 dry weight (Cristale et al. 2013). In river water, PBT concentrations range from 0.0003 to 0.0210 μg L−1 (Wu et al. 2010).

Nanoparticles with unique surface-active properties are beneficial in removing pharmaceuticals and micropollutants (Hu et al. 2012). Titanium (IV) oxide (TiO2) as a photocatalyst is widely applied to wastewater treatment due to its photocatalytic oxidation ability, and stable and non-toxic characteristics (Chen et al. 2017). Cerium (IV) oxide (CeO2) nanoparticles are widely used as a catalyst and as superconducting materials (Xu et al. 2017).

The treatability studies of GFZ, TCS, PBT and PBEB performed with nano-CeO2 and nano-TiO2 via photocatalytic processes are limited in recent literature. Chen et al. determined 61.8% GFZ removal with 0.05 g L−1 nano-TiO2 in a synthetic wastewater having a GFZ concentration of 2 mg L−1 under 350 W UV light power at a pH of 7.00 through 8.0 min (Chen et al. 2017). Son et al. reported 82% TCS photodegradation rate within 20 min with 0.1 g L−1 nano-TiO2 at a pH of 7.00 under 1.37 × 10−4 Einstein L−1 min−1 UV (Son et al. 2009). Yu et al. found 95% photocatalytic degradation for 9 mg L−1 TCS within 6 h using 100 mg L−1 nano-TiO2 at UV light power of 15 W (Yu et al. 2006). García et al. found 85% TCS photodegradation at 100 μg L−1 TCS by 0.335 g L−1 nano-TiO2 under solar light irradiation after 120 min (García et al. 2012). The physicochemical properties (hydrophobicity, solubility, volatilization) of micropollutants used in this study are shown in Table 1.

Table 1

Physicochemical properties of PBEB, PBT, TCS and GFZ (from US Environmental Protection Agency EPI Suite™ database, 2017)

IUPAC nameAbbreviationMolecular weight (g mol−1)Henry's Law constant, kH at 25 °C (atm m3 mol−1)Solubility in water, at 25 °C (g L−1)Dissociation constant, pKa (unitless)Octanol/water partition coefficient, log KowStructure
2,3,4,5,6-pentabromoethyl benzene (CAS No. 85-22-3) PBEB 500.65 1.12 × 10−4 2.895 × 10−7 N.A. 7.48  
2,3,4,5,6- pentabromotoluene (CAS No. 87-83-2) PBT 486.62 7.22 × 10−5 9.351 × 10−7 N.A. 6.99  
5-chloro-2-(2,4-dichloro-phenoxy)-phenol (CAS No. 3380-34-) TCS 289.55 2.13 × 10−8 4.621 × 10−3 7.9 4.76  
5-(2,5-dimethyl phenoxy)-2,2-dimethyl pentanoic acid (CAS No. 25812-30-0) GFZ 250.34 1.19 × 10−8 4.964 × 10−3 4.5 4.77  
IUPAC nameAbbreviationMolecular weight (g mol−1)Henry's Law constant, kH at 25 °C (atm m3 mol−1)Solubility in water, at 25 °C (g L−1)Dissociation constant, pKa (unitless)Octanol/water partition coefficient, log KowStructure
2,3,4,5,6-pentabromoethyl benzene (CAS No. 85-22-3) PBEB 500.65 1.12 × 10−4 2.895 × 10−7 N.A. 7.48  
2,3,4,5,6- pentabromotoluene (CAS No. 87-83-2) PBT 486.62 7.22 × 10−5 9.351 × 10−7 N.A. 6.99  
5-chloro-2-(2,4-dichloro-phenoxy)-phenol (CAS No. 3380-34-) TCS 289.55 2.13 × 10−8 4.621 × 10−3 7.9 4.76  
5-(2,5-dimethyl phenoxy)-2,2-dimethyl pentanoic acid (CAS No. 25812-30-0) GFZ 250.34 1.19 × 10−8 4.964 × 10−3 4.5 4.77  

N.A. Not available.

The novelty of the study is to treat some brominated and phenolic micropollutants present in raw hospital wastewater for the first time in Turkey by photocatalytic processes under laboratory conditions. The aim of this study was to evaluate the photocatalytic removals of PBT, PBEB, TCS and GFZ in a raw hospital wastewater using CeO2 and TiO2. Effects of irradiation times (15; 30; 45; 60; 90 min), nanoparticle concentrations (0.25; 0.50; 1.00; 1.50 g L−1), UV light powers (120; 210; 300 W) and raw hospital wastewater pH (4.00; 7.00; 8.50) on PBT, PBEB, TCS and GFZ photodegradation yields throughout photocatalysis at constant temperature (25 °C) were investigated using nano-CeO2 and nano-TiO2 generated under laboratory conditions. The recoveries of the nanoparticles and their costs were also investigated.

Source and characterization of raw hospital wastewater

Raw hospital wastewater was taken from the Dokuz Eylul University Hospital (Izmir, Turkey) sewer channel, which flows through the sewer channel without any treatment. The influent COD concentration of the raw hospital wastewater was 1,900 ± 20 mg L−1 while the influent PBT, PBEB, TCS and GFZ concentrations were 2.31 ± 0.003, 3.28 ± 0.004, 860.00 ± 0.05 and 53.30 ± 0.05 μg L−1, respectively. The pH, the temperature, the total nitrogen and total phosphorus concentrations were 8.50 ± 0.02, 18 ± 2 °C, 0.40 ± 0.01 and 1.00 ± 0.01 mg L−1, respectively.

Analytical procedure

Extraction of PBEB, PBT, TCS and GFZ from raw hospital wastewater

PBEB, PBT, TCS and GFZ were extracted the from raw hospital wastewater by solid-phase extraction method. Oasis HLB cartridges (200 mg) were used in the extraction experiments. PBEB and PBT extraction experiments were carried out according to Chokwe et al. (2015) while TCS and GFZ extraction experiments were carried out according to Diwan et al. (2010). Extraction recovery rates for PBEB, PBT, TCS and GFZ are shown in Table 2.

Table 2

The RT, m/z transitions, extraction and method recovery rates for PBEB, PBT, TCS and GFZ

CompoundRetention time (min)m/z transitions
Recovery study
Average extraction recovery (%)Average method recovery (%)
Ion 1Ion 2Ion 3Actual concentration (μg L−1)Peak area of the standardPeak area of the sampleExtraction recovery (%)Determined concentration (μg L−1)Method recovery (%)
PBEB 10.786 499.7 484.7 486.7 10 49,712 38,777 78.0 6.84 68.4 103.5 91.9 
50 264,915 259,219 97.9 45.71 91.4 
100 487,294 656,539 134.7 115.76 115.8 
PBT 10.550 485.7 487.7 406.7 10 84,792 64,637 76.2 7.03 70.3 103.3 94.8 
50 437,344 440,972 100.8 47.96 95.9 
100 818,197 1,087,394 132.9 118.27 118.3 
TCS 3.606 – – – 100 2916.30 4163.43 142.8 121.73 121.7 119.07 116.73 
500 14935.00 16405.40 109.8 576.92 115.4 
1000 29935.00 31301.40 104.6 1130.80 113.1 
GFZ 4.351 – – – 100 76.63 73.56 96.00 96.00 96.00 97.35 97.17 
500 563.53 110.45 98.00 490.00 98.00 
1,000 1008.81 98.92 98.06 975.00 97.50 
CompoundRetention time (min)m/z transitions
Recovery study
Average extraction recovery (%)Average method recovery (%)
Ion 1Ion 2Ion 3Actual concentration (μg L−1)Peak area of the standardPeak area of the sampleExtraction recovery (%)Determined concentration (μg L−1)Method recovery (%)
PBEB 10.786 499.7 484.7 486.7 10 49,712 38,777 78.0 6.84 68.4 103.5 91.9 
50 264,915 259,219 97.9 45.71 91.4 
100 487,294 656,539 134.7 115.76 115.8 
PBT 10.550 485.7 487.7 406.7 10 84,792 64,637 76.2 7.03 70.3 103.3 94.8 
50 437,344 440,972 100.8 47.96 95.9 
100 818,197 1,087,394 132.9 118.27 118.3 
TCS 3.606 – – – 100 2916.30 4163.43 142.8 121.73 121.7 119.07 116.73 
500 14935.00 16405.40 109.8 576.92 115.4 
1000 29935.00 31301.40 104.6 1130.80 113.1 
GFZ 4.351 – – – 100 76.63 73.56 96.00 96.00 96.00 97.35 97.17 
500 563.53 110.45 98.00 490.00 98.00 
1,000 1008.81 98.92 98.06 975.00 97.50 

PBEB, PBT, TCS and GFZ measurement in raw hospital wastewater

PBEB (CAS No. 85-22-3) and PBT (CAS No. 87-83-2) measurements were performed on a GC Agilent 7890A equipped with an Agilent 5975C inert mass selective detector. The gas chromatography–mass spectrometer (GC-MS) is equipped with a flame ionization detector and an HP-5MS capillary column with dimensions of 30.0 m (length), 0.250 mm (I.D.), and 0.25 μm (film thickness) from Agilent Technologies. The oven program was taken from Cristale et al. and developed as following: the oven was set at 40 °C (1 min), increased to 60 °C at 15 °C min−1 (8 min), to 220 °C at 10 °C min−1 and to 300 °C at 15 °C min−1 (8 min). Helium (purity 99.999%) was employed as carrier gas with a constant flow of 1.5 mL min−1. The standards were injected (2 μL) into the GC system in splitless mode, with a splitless time of 1.5 min. The injector, quadrupole and transfer line were set at 300, 150, and 280 °C respectively (Cristale et al. 2012). The retention times (RT), m/z transitions and method recovery rates for PBEB and PBT monitored by the GC-MS are shown in Table 2.

TCS (CAS No. 3380-34-5) and GFZ (CAS No. 25812-30-0) measurements were performed on an Agilent 1100 Series high-performance liquid chromatograph (HPLC) equipped with a diode array detector. A Thermo C18 column (5 μm, 250 mm × 4.6 mm, Thermo Scientific) was used and the injection volume of each sample was 20 μL. For TCS, the mobile phase consisted of pure acetonitrile and HPLC grade water (90:10 v:v) with a flow rate of 1.0 mL min−1 for a column oven temperature of 40 °C and retention time of 3.606 min (Maaroof & Uysal 2014), while for GFZ, the mobile phase consisted of pure acetonitrile, pure methanol and 0.05 M (pH = 7.00) NH4H2PO4 (60:30:10 v:v:v) with a flowrate of 0.8 mL min−1 for a column oven temperature of 20 °C and retention time of 4.351 min (Elsherif et al. 2013). The RT and method recovery rates of TCS and GFZ monitored by the HPLC are shown in Table 2.

Conventional-pollutants measurement methods

Conventional pollutants in hospital wastewater such as COD were measured according to Standard Methods (APHA/AWWA/WEF 2012). Total nitrogen and total phosphorus were measured with reagent kits in a Photometer Nova 60/Spectroquant. The pH and temperature were measured with WTW probes.

Preparation of nanoparticles and characterization

TiCl4 (0.92 mL) was dissolved in 35 mL de-ionized water under an ice water bath, and then 0.8 g of P123 as a structure-directing agent was added into the solution under magnet stirring (Tang et al. 2016). When the temperature was raised to 50 °C, the mixture solution was aged for another 12 h. Finally, the solution obtained was dried via lyophilization. The obtained solids were calcined at 300 °C for 2 h, and the final product was named as P0.5-TiO2. One millilitre of po1yoxyethylene octylphenol ether was mixed with 1 mL 1-hexanol and 2.5 mL cyclohexane at a mass ratio of 1:1:2.5, then 0.5 mol L−1 Ce(NO3)3·6H2O was added to obtain an emulsion (Wang et al. 2015). This mixture was aged at the reaction temperature of 300 °C for 3 h before being centrifuged to separate the solid, which was subsequently rinsed and dried to obtain the nano-CeO2.

Field emission scanning electron microscopy (SEM) with 20.0 kV was used to study the morphology of the synthesized TiO2 and CeO2 nanoparticles prepared at 300 °C calcination temperatures. Transmission electron microscopy (TEM) with a Zeiss 80.0 kV was used to analyze the estimated particle size of the TiO2 and CeO2. XRD analysis was used to identify the phase composition of the prepared TiO2 and CeO2 by means of an X-ray diffractometer with the D/max-2200 PC using a Cu Ka radiation source (λ = 1.541 Å) at 40.0 kV and 200 mA. The optical properties of absorption were measured by an ultraviolet–visible spectrophotometer (UV-Vis) and Fourier transform infrared (FT-IR) spectroscopy, Perkin Elmer Spectrum BX, using the KBr method. The average crystallite size of the samples was calculated by Scherrer's equation (Equation (1)) using TEM as illustrated below:
formula
(1)
where Φ is the crystal size, λ is the wavelength of the X-ray irradiation (0.154 nm), K is taken as 0.89, β is the peak width at half-maximum height and θ is the diffraction angle of the peak of the anatase phase. BET (Brunauer–Emmett–Teller) surface area analysis was used to establish the surface area, total pore volume and average pore diameter of the materials. The band gap energy was calculated according to Equation (2):
formula
(2)
where Eg is the band gap (eV) and λ (nm) is the wavelength of the absorption edge in the spectrum.

Photocatalytic studies

The photocatalytic experiments were carried out in a covered stainless steel system consisting of quartz glass reactors with a volume of 1,000 mL and UV light lamps with a 254 nm wavelength (895.0 mm × 26.0 mm, 30.0 W, 0.36 A, G13 Model, Osram). The light power required was provided by increasing or decreasing the UV lamp numbers. The distance between the UV lamps and quartz glass reactors was 10 cm (Figure 1).

Figure 1

The photocatalytic reactor system.

Figure 1

The photocatalytic reactor system.

Close modal

The control experiments were carried out under dark conditions for the study of the adsorption capacity of TiO2 and CeO2 nanoparticles. Raw hospital wastewater was first filtered through a 0.45 μm pore sized membrane to remove physical impurities. The nanoparticle was added to the filtered hospital wastewater in a quartz glass reactor and it was stirred at 25 °C. The control for photolysis was carried out under UV light without nano-TiO2 and nano-CeO2 to determine the photolysis capacity of the micropollutants. Raw hospital wastewater was filtered through a 0.45 μm pore sized membrane and then irradiated under UV in photocatalytic reactors without nanoparticles at 25 °C. All the experiments were performed in triplicate.

The effects of different nanoparticle concentrations (nano-TiO2 and nano-CeO2) (0.25; 0.50; 1.00; 1.50 g L−1), irradiation times (15; 30; 45; 60; 90 min), UV light powers (120; 210; 300 W) and hospital wastewater pH (4.00; 7.00; 8.50) on the PBT, PBEB, TCS and GFZ photodegradation yields throughout photocatalysis at constant temperature (25 °C) were investigated.

Recovery and reuse of nanoparticles

After the first utilization of nano-TiO2 and nano-CeO2 in the photocatalytic experiments, the nanoparticles were separated by filtering them through 0.45 μm pore sized membranes. Then, the nanoparticles were regenerated using ethanol with 0.10 mol L−1 HCl (Chowdhury & Balasubramanian 2014). The nanoparticles were dried under vacuum and they were reused for the second treatment process to treat TCS, GFZ, PBT and PBEB again. For each new treatment step the same procedure was applied to the nanoparticles. Six sequential treatment steps were investigated in order to detect the effect of reusability of the nanoparticles from the raw hospital wastewater after each treatment.

Statistical analysis

The correlation (R) between data was performed using Excel Microsoft 2010 while the significance between parameters was performed in analysis of variance (ANOVA) tests using an α-value of 0.05 with P (probability) values. All of the differences were considered significant at p < 0.05. The Kruskal–Wallis one way test was applied to calculate the significant difference of the operational conditions on the photodegradation yields. Kruskal–Wallis ANOVA χ2 values were calculated for P > 0.05(Quinn & Keough 2002).

Physicochemical properties of generated nano metal oxides under laboratory conditions

The SEM image in Figure 2(a) shows that the TiO2 nanoparticle samples exhibit a well-defined spherical structure after 300 °C calcination. The TEM figure shows that the as-prepared powder is completely crystalline, and consists entirely of a pure anatase phase (Figure 2(b)). The XRD analysis showed that the TiO2 nanoparticle originated completely from the pure anatase phase with a good crystallinity (Figure 2(c)). From the XRD pattern, it is clear that the material formed has a tetragonal crystal structure with lattice parameters a = b = 3.785 Å and c = 9.513 Å. The crystallite size of the TiO2 was in the range 10.5–11.5 nm and it had a unit cell volume of 136.27 Å (Table 3). The main peaks of anatase TiO2 appeared at 25°, 28°, 39°, 48°, 54°, 55°, 63°, 69°, 70° and 76° (Figure 2(c)). These peaks of anatase in TiO2 were in accordance with the peaks given by Hu et al. (2017) and Hassani et al. (2017) (at 25°, 28°, 38° and 48°, 54°, 55°, 63°, 69°, 70°, and 76°). The strongest and narrowest peaks were detected at 2θ = 25° for anatase. According to FT-IR analysis, the peaks of anatase TiO2 appeared at 3415.95, 2360.83, 1630.35 and 1386.86 cm−1 (Figure 2(d)). These peaks were similar to the findings of Hassani et al. (2017). In the FT-IR spectrum of the TiO2 nanoparticle, the bands appeared at 3415.95–2360.83 cm−1, which shows the bending vibration of OH molecules. A carbonyl peak appeared at around 1630.35 cm−1. The peak at 1386.86 cm−1 shows stretching vibrations of Ti-OH. The BET surface area of the synthesized TiO2 nanoparticle sample shows less aggregation and a smaller particle size, which is in accordance with the SEM and TEM analysis. The BET surface area of the TiO2 nanoparticle catalyst had an average pore diameter of 98.84 Å with a total pore volume of 0.402 cm3 g−1 while the surface area was 205 m2 g−1 (Table 3). The absorption spectrum of the TiO2 nanoparticle at a wavelength of 410 nm showed that it is a highly crystalline nanomaterial with a band gap value of 3.02 Ev due to the presence of a very fine anatase phase (Figure 2(e)).

Table 3

BET analysis results and band gap energies of nano-TiO2 and nano-CeO2

IUPAC nameSurface area (m2 g−1)Pore volume (cm3 g−1)Average size (nm)Pore diameter (Å)
TiO2 205 0.402 11.50 98.84 
Commercial TiO2 167 0.709 27.00 123.19 
CeO2 302 0.309 8.11 78.06 
Commercial CeO2 319 0.321 9.67 82.90 
Sample Band gap Eg (eV) 
TiO2 3.02 
CeO2 4.10 
IUPAC nameSurface area (m2 g−1)Pore volume (cm3 g−1)Average size (nm)Pore diameter (Å)
TiO2 205 0.402 11.50 98.84 
Commercial TiO2 167 0.709 27.00 123.19 
CeO2 302 0.309 8.11 78.06 
Commercial CeO2 319 0.321 9.67 82.90 
Sample Band gap Eg (eV) 
TiO2 3.02 
CeO2 4.10 
Figure 2

(a) SEM image, (b) TEM image, (c) XRD analysis and (d) FT-IR peaks of nano-TiO2 after 45 min irradiation time, at 0.50 g L-1 nano-TiO2 concentration, at 210 W UV light power.

Figure 2

(a) SEM image, (b) TEM image, (c) XRD analysis and (d) FT-IR peaks of nano-TiO2 after 45 min irradiation time, at 0.50 g L-1 nano-TiO2 concentration, at 210 W UV light power.

Close modal

SEM images showed that CeO2 nanoparticles exhibited spherical morphologies (Figure 3(a)). Figure 3(b) shows the TEM image of the sample synthesized at 165 °C displaying a particle size of 1–3 nm and strong agglomeration. The XRD patterns reveal that the prepared CeO2 nanoparticles had a cubic fluorite structure (Figure 3(c)). From the XRD peaks it is evident that the crystallites were oriented along 28°, 33°, 48°, 56°, 59° and 69°, corresponding to the cubic fluorite structure of CeO2. Similar peaks at 28°, 33°, 48° and 56° were reported by Phuruangrat et al. (2017) and Channei et al. (2017). The crystallite sizes calculated were around 4.5 and 6.5 nm (Table 2). The BET surface area of the CeO2 nanoparticle had an average pore diameter of 78.06 Å with a pore volume of 0.309 cm3 g−1 while the surface area was 302 m2 g−1 (Table 2). The FT-IR spectrum was recorded in the range of 650 to 4,000 cm−1 (Figure 3(d)). These FT-IR peaks are in accordance with the data obtained by Ravishankar et al. (2015) and Tambat et al. (2016) for CeO2 nanoparticles). These peaks are due to (Ce-O) metal–oxygen bond vibrations. The band gap of nano-CeO2 was higher than that of nano-TiO2 (Table 3). The prepared nano-CeO2 and nano-TiO2 exhibited better properties than the commercial nanoparticles (Table 3). Nano-CeO2 produced under laboratory conditions had a higher surface area than those of nano-TiO2 while the pore volume, the average size and pore diameter of nano-CeO2 was lower than nano-TiO2. This increased the photocatalytic properties of nano-CeO2. The UV–visible absorption spectral studies showed that the synthesized CeO2 nanoparticles had a strong absorption band at low wavelength near 380 nm, corresponded to a band gap energy of 3.26 eV (Figure 3(e)). In comparison with the UV–visible absorption spectrum of CeO2 nanoparticles reported in the literature (Farahmandjou et al. 2016), similar peaks in the spectrum located at around 400–700 nm were observed. The high band gap energy and low pore volume and pore diameter of nano-CeO2 compared to nano-TiO2 enhanced the UV light absorption of the CeO2 nanoparticle.

Figure 3

(a) SEM image, (b) TEM image, (c) XRD analysis and (d) FT-IR peaks of nano-CeO2 after 45 min irradiation time, at 0.50 g L−1 nano-CeO2 concentration, at 210 W UV light power.

Figure 3

(a) SEM image, (b) TEM image, (c) XRD analysis and (d) FT-IR peaks of nano-CeO2 after 45 min irradiation time, at 0.50 g L−1 nano-CeO2 concentration, at 210 W UV light power.

Close modal

Photocatalytic removal studies

Effect of adsorption alone and photolysis on the removals of GFZ, TCS, PBT and PBEB

During the adsorption studies of micropollutants on the surface of nano-TiO2 the maximum removal percentages of GFZ, TCS, PBT and PBEB via adsorption were 5.7%, 5.6%, 7.0% and 7.6%, respectively, at 0.50 g L−1 nano-TiO2 concentration after a 45 min stirring time at a pH of 8.50 and a temperature of 25 °C without UV (Figure 4(a)). The adsorption removal percentages of GFZ, TCS, PBT and PBEB were 9.0%, 6.0%, 13.4% and 17.1%, respectively, at 0.50 g L−1 nano-CeO2 concentration after a 45 min stirring time at the same operational conditions given above under dark experimental conditions (Figure 4(b)). It can be said that the adsorption of micropollutants on photocatalyst was not the primary mechanism for their removals. The adsorption yields were double for PBEB and PBT compared to GFZ and TCS. This can be explained by the physicochemical properties of PBEB (low solubility = 2.895 × 10−7 g L−1 and high adsorption capacity log Kow = 7.48) and PBT (low solubility = 9.351 × 10−7 g L−1 and high adsorption capacity log Kow = 6.99) compared to GFZ (low solubility = 4.964 × 10−3 g L−1 and high adsorption capacity log Kow = 4.77) and TCS (low solubility = 2.13 × 10−8 g L−1 and high adsorption capacity log Kow = 4.76) and by their structures (brominated groups in PBT and PBEB; phenolic groups in GFZ and TCS). ANOVA multiple regression analysis showed that there was a linear regression between adsorption yields and physicochemical properties of the studied compounds (R = 0.85) and this was not statistically significant (P = 0.21, α = 0.05). The Kruskal–Wallis one way test showed that the physicochemical properties had a significant effect on adsorption of all micropollutants (Kruskal–Wallis ANOVA, χ2 = 4.29, P < 0.05).

Figure 4

(a) The adsorption percentages of GFZ, TCS, PBT and PBEB at 0.50 g L−1 nano-TiO2. (b) The adsorption percentages of GFZ, TCS, PBT and PBEB at 0.50 g L−1 nano-CeO2. (c) The photolysis of GFZ, TCS, PBT and PBEB in the absence of nano-materials.

Figure 4

(a) The adsorption percentages of GFZ, TCS, PBT and PBEB at 0.50 g L−1 nano-TiO2. (b) The adsorption percentages of GFZ, TCS, PBT and PBEB at 0.50 g L−1 nano-CeO2. (c) The photolysis of GFZ, TCS, PBT and PBEB in the absence of nano-materials.

Close modal

The experiments were also performed using direct photolysis of micropollutants in the absence of nano-TiO2 and nano-CeO2 under 210 W UV at 254 nm. As a result, no obvious change in their concentrations within 45 min UV light irradiation was observed (9.0%, 10.0%, 14.2% and 15.9% removals for GFZ, TCS, PBT and PBEB, respectively) (Figure 4(c)). For sole UV exposure, it was found that micropollutants in hospital wastewater with phenol bonds (TCS and GFZ) or brominated bonds (PBT and PBEB) were not easily photolyzed by 254 nm at 210 W UV photons. During photolysis only the photons were generated during UV irradiation instead of generation of hydroxyl radicals (OH). The low removals of GFZ, TCS, PBT and PBEB in photolysis were expected to be achieved mostly by the low number of photons from UV irradiation (Son et al. 2009). Similar results were obtained for the direct UV photolysis of TCS and GFZ (Carlson et al. 2015). Stamatis et al. also found that the removal of 1.0 mg L−1 TCS with adsorption and photolysis was low (1.8% and 4.9%, respectively) with 550 mg L−1 TiO2 at an irradiation intensity of 700 W m−2 (Stamatis et al. 2014). The ANOVA multiple test statistic showed a linear regression between physicochemical properties and removals of micropollutants by photolysis (R = 0.84), although this regression was not significant (P = 0.15, α = 0.05). The Kruskal–Wallis one way test showed that photolysis had no significant effects on the removal of micropollutants (Kruskal–Wallis ANOVA, χ2 = 16.14, P > 0.05).

Photodegradation of GFZ, TCS, PBT and PBEB

Effect of irradiation time on the photodegradation of GFZ, TCS, PBT and PBEB. According to the previous studies given in the literature and from our preliminary studies it was decided to use a 0.25 g L−1 nano-TiO2 and nano-CeO2 concentration at a 120 W UV power. The photocatalytic experiments performed with nano-TiO2 showed that 45 min is the optimum irradiation time for the maximum photocatalytic removal of GFZ (80%), TCS (83%), PBT (72%) and PBEB (68%) at 25 °C at a pH of 7.00 (Figure 5(a)). Any further increase in irradiation time decreased all the micropollutants' photo-removals because of the deactivation of active sites by the deposition of by-products. Neves reported that the increasing of irradiation time from 1 h to 5 h positively affected the 0.01 g L−1 TCS removal (from 68% to 78%) at a pH value of 7.00 with 0.50 g L−1 TiO2 with UV (Neves 2014). It is important to note that the TCS photodegradation yields in our study were higher than the data obtained by Neves although they were obtained with half of the nanoparticle dose and four times the irradiation time. Also, the photocatalytic experiments performed with nano-CeO2 showed that 45 min was the optimum irradiation time for the maximum photocatalytic removals of GFZ (86%), TCS (88%), PBT (77%) and PBEB (75%) with 0.25 g L−1 nano-CeO2 concentration at 120 W power at 25 °C and at a pH of 7.00 (Figure 5(b)). It is widely accepted that UV excitation leads to the generation of electron () – hole () pairs on the nano-TiO2 and nano-CeO2 surface. These charges will react with water molecules and result in OH. The hydroxyl radicals oxidize the micropollutant molecules and cause the latter to decompose and mineralize to water and to carbon dioxide. ANOVA test statistics showed that a linear regression between irradiation time and micropollutant photo-removals was observed by increasing the irradiation time from 15 min to 60 min (R = 0.89) and this correlation was significant (P = 0.50). The Kruskal–Wallis one way test showed that the increased length of time up to 60 min had a significant effect on micropollutant removals (Kruskal–Wallis ANOVA, χ2 = 1.12, p < 0.05). A significant regression between photodegradation yields of micropollutants and irradiation time for higher than 45 min (60 and 90 min) (R = 0.23) was not observed and this correlation was not significant (P = 12.07, α = 0.05) (Kruskal–Wallis ANOVA, χ2 = 16.78, p < 0.05).

Figure 5

(a) Effect of time on the photodegradation of GFZ, TCS, PBT and PBEB at 0.25 g L−1 nano-TiO2 , at 120 W, 25 °C and at pH = 7.00. (b) Effect of time on the photodegradation of GFZ, TCS, PBT and PBEB at 0.25 g L−1 CeO2, at 120 W, 25 °C and at pH = 7.00. (c) Effect of nano-TiO2 on the photodegradation of GFZ, TCS, PBT and PBEB at 45 min, at 120 W, 25 °C and at pH = 7.00, (d) Effect of nano-CeO2 on the photodegradation of GFZ, TCS, PBT and PBEB at 45 min, at 120 W, at 25 °C and at pH = 7.00. (e) Effect of UV light power on the photodegradation of GFZ, TCS, PBT and PBEB at 45 min at 0.50 g L−1 nano-TiO2 at 25 °C and at pH = 7.00 (f) Effect of UV light power on the photodegradation of GFZ, TCS, PBT and PBEB at 45 min, at 0.50 g L−1 nano-CeO2 at 25 °C and at pH = 7.00. (g) Effect of pH on the photodegradation of GFZ, TCS, PBT and PBEB at 45 min, at 0.50 g L−1 nano-TiO2 at 210 W and at 25 °C. (h) Effect of pH on the photodegradation of GFZ, TCS, PBT and PBEB at 45 min at 0.50 g L−1 nano-CeO2 at 210 W and at 25 °C.

Figure 5

(a) Effect of time on the photodegradation of GFZ, TCS, PBT and PBEB at 0.25 g L−1 nano-TiO2 , at 120 W, 25 °C and at pH = 7.00. (b) Effect of time on the photodegradation of GFZ, TCS, PBT and PBEB at 0.25 g L−1 CeO2, at 120 W, 25 °C and at pH = 7.00. (c) Effect of nano-TiO2 on the photodegradation of GFZ, TCS, PBT and PBEB at 45 min, at 120 W, 25 °C and at pH = 7.00, (d) Effect of nano-CeO2 on the photodegradation of GFZ, TCS, PBT and PBEB at 45 min, at 120 W, at 25 °C and at pH = 7.00. (e) Effect of UV light power on the photodegradation of GFZ, TCS, PBT and PBEB at 45 min at 0.50 g L−1 nano-TiO2 at 25 °C and at pH = 7.00 (f) Effect of UV light power on the photodegradation of GFZ, TCS, PBT and PBEB at 45 min, at 0.50 g L−1 nano-CeO2 at 25 °C and at pH = 7.00. (g) Effect of pH on the photodegradation of GFZ, TCS, PBT and PBEB at 45 min, at 0.50 g L−1 nano-TiO2 at 210 W and at 25 °C. (h) Effect of pH on the photodegradation of GFZ, TCS, PBT and PBEB at 45 min at 0.50 g L−1 nano-CeO2 at 210 W and at 25 °C.

Close modal

The lower photodegradation yields in PBEB and PBT compared to GFZ and TCS can be attributed to the difficulties in separating the five brominated groups with methyl and ethyl during consecutive reductive debromination as reported by Lee & Kwon (2016). Their high log Kow and low solubilities and their hydrophobicities compared to GFZ and TCS resulted in low removals. It was reported that debromination is a main photochemical reaction pathway for NBFRs (Covaci et al. 2011). During photooxidation, intramolecular elimination of HBr from PBEB and PBT can subsequently generate lower brominated compounds via consecutive debromination (Covaci et al. 2011). Higher photodegradation yields of TCS were observed (87%) to produce several intermediates including a dioxin congener (2,8-dichlorodibenzodioxin) under UV with TiO2 via dechlorination as reported by Bianco et al. (2015). The separation of chlorinated and phenolic groups photodegraded the TCS to CO2 and water. Wang et al. (2017) found that TCS was photodegraded with a yield of 78% under 130 W UV power while the photodegradation was enhanced to 89% with 1–5 kGy gamma irradiation, resulting in the mineralization of TCS. When the OH addition attacked the GFZ, transient states of the OH adduct radicals were formed, resulting in the formation of hydroxylated by-products (e.g., GFZ-Pa), followed by further radical oxidation to form stable by-products (GFZ-Pb and -Pd). When the OH addition reaction took place on the ipso aromatic C atom of GFZ-Pa, this ipso-substitution led to stabilized carbon-centered radicals. In addition, O-dealkylation reactions caused either side chain cleavage of the parent compound, giving rise to the generation of GFZ-Pc/k (Chen et al. 2017). The photo-removals of PBT, PBEB, TCS and GFZ found in this study were higher than in the literature cited above.

Effects of nano-TiO2 and nano-CeO2 concentrations on the photodegradation of GFZ, TCS, PBT and PBEB

Photocatalytic removals of 84%, 88%, 78% and 75% were obtained for GFZ, TCS, PBT and PBEB with 0.50 g L−1 nano-TiO2 (Figure 5(c)). Increasing the amount of photocatalysts from 0.25 to 0.50 g L−1 resulted in an increase in the photodegradation efficiencies of GFZ, TCS, PBT and PBEB. Among the 0.25, 0.50, 1.00 and 1.50 g L−1 nanoparticle concentrations, 0.50 g L−1 was determined as the optimum nanoparticle concentration at 120 W power at 25 °C and at a pH of 7.00 for maximum photodegradation yields of micropollutants. The photodegradation yields found for GFZ and TCS were higher than for PBT and PBEB due to their physicochemical properties. Concentrations of nano-TiO2 and nano-CeO2 higher than 0.50 g L−1 slightly decreased the photodegradation yields of GFZ, TCS, PBT and PBEB. The increase in TiO2 loading from 0.25 g L−1 to 0.50 g L−1 provided more binding sites for substrate molecules to adsorb to the TiO2 surface. At low TiO2 loading low photodegradation yields were obtained for all the micropollutants since the number of active sites on the surface of the nanoparticle was low and partially exposed to the pollutants. With excessive nano-TiO2 and nano-CeO2 loadings (1.00 and 1.50 g L−1), it has been suggested that the TiO2 particles in the centre of an irradiating vessel may be shielded from the incident UV by other nano-TiO2 and nano-CeO2 particles, and thus particles in the center do not contribute significantly to the production of electron () – hole (h+VB) pairs (Bethi et al. 2016). At high TiO2 concentrations the contacting sites on the surface of the nanocomposite were not fully activated to bind to the micropollutants. Therefore, the photodegradation yields decreased. The optimum concentrations of nano-TiO2 and nano-CeO2 were useful both from the economic and the photodegradation efficiency standpoint, because an increase in opacity and light scattering of TiO2 particles occurred at high concentrations, leading to the decreased passage of UV irradiation through the suspension (Bethi et al. 2016). By increasing the catalyst concentration to an optimum level, the surface area was increased, leading to an increase in the production of reactive species. However, increased nanoparticle concentrations would also induce greater aggregation of the catalyst and decrease the total active surface area, thereby leading to a reduction in the photocatalytic treatment efficiency. The Kruskal–Wallis one way test showed that significant effects of nanoparticle concentration on the removal of micropollutants were observed as the nanoparticle concentration was raised from 0.25 to 0.50 g L−1 (Kruskal–Wallis ANOVA, χ2 = 1.07, p < 0.05). Any further increase of nano-TiO2 and nano-CeO2 concentrations (from 0.5 to 1.5 g L−1) did not have a significant effect on micropollutant removals (Kruskal–Wallis ANOVA, χ2 = 17.31, p > 0.05).

Corena studied the effect of TiO2 concentrations on the photocatalytic removal of GFZ at a pH of 5.00 for an initial GFZ concentration of 2 mg L−1 at 100 W UV light power. The increase in TiO2 concentration from 0.10 g L−1 to 1.50 g L−1 affected GFZ removal positively because greater surface area and more active sites became available for TiO2. However, an increase in TiO2 concentration from 1.50 g L−1 to 2.50 g L−1 decreased the ability of the UV radiation to reach the TiO2 surface because a portion of the UV light was reduced and scattered by the TiO2 nanoparticles (Corena 2015). A concentration of 0.50 g L−1 nano-CeO2 gave better results for the photocatalytic removal of GFZ (89%), TCS (92%), PBT (81%) and PBEB (79%) than nano-TiO2 with 0.50 g L−1 at 120 W power at 25 °C and at a pH of 7.00 (Figure 5(d)). This can be attributed to the higher specific surface area of nano-CeO2 (302 m2 g−1) with smaller particle sizes compared to TiO2 (205 m2 g−1). Balavi et al. showed that the photocatalytic activity characteristic of CeO2 results in it being a proper alternative instead of TiO2 and ZnO in photocatalytic removal of some colored wastewaters (Balavi et al. 2013).

During photodegradation of GFZ, TCS, PBT and PBEB, semiconductor nano-TiO2 and nano-CeO2 were excited by photons (hv) and produced pairs of electrons () and holes () in the conduction and valance bands, respectively. The photogenerated electrons () and holes () were able to react with O2 and OH- to generate OH. OH tended to attack the carbon atoms with the highest electron density (Yu et al. 2006). The steps given below removed the micropollutants from the raw hospital wastewater (Equations (3)–(24)).

  • (a)
    Generation of the electron () – hole () pairs:
    formula
    (3)
  • (b)
    Keeping the holes at the valance band:
    formula
    (4)
    formula
    (5)
  • (c)
    Electron () transfer to the conduction band:
    formula
    (6)
    formula
    (7)
    formula
    (8)
    formula
    (9)

  • (d)
    Oxidation of TCS: The phenol-type intermediates (benzoic acid and phenol) of TCS occurred during photocatalysis due to the formation of OH (Son et al. 2009). Photodegradation of TCS to CO2 and H2O occurred via OH according to the following equations (Equations (10)–(13)):
    formula
    (10)
    formula
    (11)
    formula
    (12)
    formula
    (13)
  • (e)
    Oxidation of GFZ: Photodegradation of GFZ to CO2 and H2O occurred via OH and O2 according to the following equations (Equations (14)–(18)) (Chen et al. 2017):
    formula
    (14)
    formula
    (15)
    formula
    (16)
    formula
    (17)
    formula
    (18)
  • (f)
    Mineralization of PBT and PBEB to CO2 and H2O via OH and O2 occurred via exciting the bromines according to the following equations, respectively (Equations (19)–(22)) (Chen et al. 2012):
    formula
    (19)
    formula
    (20)
    formula
    (21)
    formula
    (22)
  • (g)
    Conjunction of the charge conductors (electron () – hole () pairs get together and the reaction ends):
    formula
    (23)
    formula
    (24)
Effect of UV light power on the photodegradation of GFZ, TCS, PBT and PBEB

The photocatalytic reactor was operated with a 0.50 g L−1 nanoparticle (nano-TiO2 and nano-CeO2) concentration for 45 min UV irradiation at 25 °C and at a pH of 7.00 at increasing UV light powers (120, 210 and 300 W) in order to determine the optimum UV light power for the maximum micropollutant photodegradation. The micropollutant photodegradations were obtained as 84%, 88%, 78% and 75%, respectively, for 120 W UV light power after 45 min irradiation time at 0.50 g L−1 nano-TiO2 (Figure 5(e)). When the UV light power was increased to 210 W, the micropollutant photodegradation removals reached 90%, 92%, 88% and 83%, respectively, at the same irradiation time (Figure 5(e)). As the UV light power was increased to 300 W the photodegradation yields for GFZ, TCS, PBT and PBEB decreased slightly (Figure 5(e)). In this case, more hydroxyl radicals were available to attack the aromatic rings and the rate of removal increased. For the photocatalytic experiments carried out with nano-CeO2, the micropollutant photodegradation yields were obtained as 89%, 92%, 81% and 79%, respectively, at 120 W UV light power (Figure 5(f)). The micropollutant photodegradation yields increased significantly when the UV light power was increased to 210 W (Figure 5(f)). Finally when the UV light power was increased to 300 W the photodegradation yields for GFZ, TCS, PBT and PBEB decreased slightly (Figure 5(f)). The optimum UV light power for maximum micropollutant photo-removals was found to be 210 W at 0.50 g L−1 nano-TiO2 and nano-CeO2 concentrations after 45 min irradiation time at 25 °C at pH = 7.00. Kruskal–Wallis test statistics showed that the UV power significantly increased micropollutant removals, up to 210 W UV power (ANOVA, χ2 = 2.31, p < 0.05). UV power higher than 210 W decreased micropollutant removal ((ANOVA, χ2 = 14.88, p>0.05).

It should be pointed out that the photodegradation rates of micropollutants mentioned above increased when the UV light irradiation power was increased from 120 W up to 210 W. After 210 W UV power had been reached for a certain period of time, the margin of increase dropped. At this UV power, formation of long-lived intermediate by-products did not accumulate in the solution before their reaction with hydroxyl radical and photo-transformation into carbon dioxide and water (Bai & Acharya 2016). This is the reason for the high photodegradation yields of micropollutants found in this study. The removal yields of the micropollutants mentioned above during the photocatalysis process increased linearly when the UV light power was increased from 120 W to 210 W. The multiple regression analysis showed that a significant linear correlation between GFZ, TCS, PBT, and PBEB photodegradation yields and UV power up to 300 W (R = 0.86) and this regression was significant (P = 0.53, α = 0.05). As the UV power was increased to 300 W, the electron () – hole () recombination causes lower effects on the photo-removal yields since some metabolites accumulated and this decreased effective OH radical production, as reported by Caudillo-Flores et al. (2016). Ma et al. (2016) and Chen et al. (2017) found similar results for GFZ photodegradation under 230 W and 298 W powers. Chen et al. (2016a, 2016b) mentioned that for a maximum photodegradation kinetic constant an optimum UV power (267 W) was necessary. Covaci et al. (2011) and Ezechiás et al. (2014) mentioned that optimum UV powers (180–190 W) were necessary for the cleavage of brominated rings in PBEB and PBT from NBFRs since these organics have low volatility, low water solubility, and high log Kow levels.

Effect of pH on the photodegradation of GFZ, TCS, PBT and PBEB

In photocatalysis, pH can affect the reaction of OH radical production from H2O/OH and nano-TiO2. The photodegradation rates of GFZ, TCS, PBT and PBEB were researched under acidic (pH = 4.00), neutral (pH = 7.00) and alkaline pH (pH = 8.50) conditions after 45 min irradiation time at 210 W UV light power with a nanoparticle concentration of 0.50 g L−1 nano-TiO2 at 25 °C. Under acidic pH (4.00) GFZ, TCS, PBT and PBEB photodegradation yields were 47%, 50%, 40% and 35%, respectively (Figure 5(g)). The photocatalytic yields at neutral pH (7.00) increased (Figure 5(g)). Under alkaline conditions (8.50), GFZ, TCS, PBT and PBEB photodegradation yields reached the maximum (93%, 94%, 90% and 87%, respectively) under the same experimental conditions with the nano-TiO2 (Figure 5(g)).

The surface of the photocatalyst may become positively or negatively charged depending on the ambient pH. TiO2 has a point of zero charge at around pH = 6 (Son et al. 2009). The surface of the photocatalyst will be positively charged in solutions below this pH and negatively charged above. Therefore, the pH will impact the adsorption of micropollutants on the TiO2 surface depending on their chemical structure. In this study, it was found that the most efficient pH was 8.50 for this nanoparticle whereas more extreme pH conditions negatively affected the photodegradation yields. This is due to the chemical structure of GFZ and TCS at pH = 8.50. Both the phenoxy and chloro groups on chloro-phenol moieties in GFZ (dissociation constant pKa = 8.6) and dimethyl and phenoxy moieties (pKa = 6.1) in TCS are deprotonated with the overall charge being negative (Son et al. 2009). Therefore, because of negative surface charge of TiO2, maximal photodegradation occurs. Similarly, at low pH, both these functional groups are protonated, and hence the overall positive charge of GFZ and TCS results in low photodegradation yields as reported by Martínez-Zapata et al. (2013). Molinari et al. studied the effect of pH on the photocatalytic degradation of GFZ with UV/TiO2. They observed that the maximum GFZ photodegradation yield was at alkaline pH values since the TiO2 surface was charged negatively and the carboxylic group of GFZ was in the anionic form causing a repulsion phenomenon with the TiO2 nanoparticle (Molinari et al. 2008). The Kruskal–Wallis test showed that the effect of pH on micropollutant photodegradation yields was significant (ANOVA, χ2 = 2.03, p < 0.05).

The GFZ, TCS, PBT and PBEB photodegradation yields were lower (48%, 52%, 43% and 37%, respectively) in acidic pH (4.00) compared to neutral pH (7.00) conditions at 0.50 g L−1 CeO2 concentrations for the same operational conditions given for TiO2 (Figure 5(h)). The photocatalytic yields of GFZ, TCS, PBT and PBEB increased to 95%, 97%, 92% and 90%, respectively, at a pH of 8.50 (Figure 5(h)). The reason for the increase of the photodegradation yields of the four micropollutants at pH = 8.50 can be explained by the hydroxyl radicals being more readily produced at higher pH values. As a result, the larger amount of hydroxyl radicals produced increased the photocatalytic degradation removals, as reported by Chow et al. (2012). Son et al. reported that the decreasing of photocatalytic removals under acidic conditions is due to inhibition of hydroxyl radical generation (Son et al. 2009). Since the pH of the raw hospital wastewater was 8.50, no pH adjustment was needed to achieve maximum GFZ, TCS, PBT and PBEB photodegradation yields with nano-TiO2 and nano-CeO2. This would decrease the cost of chemicals used to adjust the pH of hospital wastewater.

Recovery and reuse of nano-TiO2 and nano-CeO2

In order to decrease the catalyst cost, recoveries of nano-TiO2 and nano-CeO2 were researched in this study. Six sequential treatment steps were investigated in order to determine possible reusability of the TiO2 and CeO2 nanoparticles. The TiO2 and CeO2 nanoparticles were reused at the optimum experimental conditions (45 min irradiation time, 0.50 g L−1 nanoparticle concentration, 210 W UV light power, pH = 8.50 at 25 °C). TCS, GFZ, PBT and PBEB removal efficiencies were taken into consideration to detect the reusability of nano-TiO2 and nano-CeO2 throughout the sequential six treatments. After the first treatment step with nano-TiO2, the maximum removal efficiencies of TCS, GFZ, PBT and PBEB were determined as 94%, 93%, 90% and 87%, respectively (Table 4). After the second treatment step, the removal efficiencies of TCS, GFZ, PBT and PBEB were determined as 92%, 91%, 90% and 86%, respectively (Table 4). The removal efficiencies of TCS, GFZ, PBT and PBEB were obtained as 88%, 88%, 85% and 81%, respectively, after the fifth treatment step. In the sixth treatment step, the removal efficiencies decreased to 80%, 78%, 75% and 74% for TCS, GFZ, PBT and PBEB, respectively (Table 4).

Table 4

The utilization of the nano-TiO2 and nano-CeO2 for six times for TCS, GFZ, PBT and PBEB removals under optimum experimental conditions (45 min, 0.50 g L−1 nano-CeO2, 210 W UV light power)

 TCSGFZPBTPBEB
 Removal efficiencies (%) with nano-TiO2 
1st treatment step 94 93 90 87 
2nd treatment step 92 91 90 86 
3rd treatment step 92 91 88 84 
4th treatment step 90 90 83 83 
5th treatment step 88 88 81 81 
6th treatment step 80 78 75 74 
 Removal efficiencies (%) with nano-CeO2 
1st treatment step 97 95 92 90 
2nd treatment step 96 93 91 90 
3rd treatment step 94 93 90 88 
4th treatment step 94 91 87 86 
5th treatment step 92 89 86 85 
6th treatment step 86 83 80 79 
 TCSGFZPBTPBEB
 Removal efficiencies (%) with nano-TiO2 
1st treatment step 94 93 90 87 
2nd treatment step 92 91 90 86 
3rd treatment step 92 91 88 84 
4th treatment step 90 90 83 83 
5th treatment step 88 88 81 81 
6th treatment step 80 78 75 74 
 Removal efficiencies (%) with nano-CeO2 
1st treatment step 97 95 92 90 
2nd treatment step 96 93 91 90 
3rd treatment step 94 93 90 88 
4th treatment step 94 91 87 86 
5th treatment step 92 89 86 85 
6th treatment step 86 83 80 79 

The maximum removal efficiencies of TCS, GFZ, PBT and PBEB were determined as 97%, 95%, 92% and 90%, respectively, after the first treatment step with nano-CeO2 (Table 4). After the second treatment step, removal efficiencies of TCS, GFZ, PBT and PBEB were obtained as 96%, 93%, 91% and 90%, respectively (Table 4). After the fifth treatment step, the removal efficiencies of TCS, GFZ, PBT and PBEB were determined as 92%, 89%, 86% and 85%, respectively. The removal efficiencies decreased to 86%, 83%, 80% and 79% for TCS, GFZ, PBT and PBEB, respectively after the sixth treatment step (Table 4). The results showed that the nano-TiO2 and nano-CeO2 can be used effectively for five sequential runs (the lowest removal efficiency was 81% with nano-TiO2 and it was 85% with nano-CeO2 for PBEB at the end of the fifth treatment step) for the photocatalytic treatment of TCS, GFZ, PBT and PBEB from raw hospital wastewater to decrease the cost of treatment.

Cost analysis

A cost analysis was carried out for the photocatalytic treatment of TCS, GFZ, PBT and PBEB from 1 m3 raw hospital wastewater at optimum experimental conditions (45 min irradiation time, 500 g L−1 nanoparticle concentration, 210 W UV light power, pH = 8.50 at 25 °C) with nano-TiO2 and nano-CeO2. The UV power required was provided using 20 UV lamps. Also, electricity consumption and nanoparticle synthesis costs were considered. The total cost of the photocatalytic treatment of TCS, GFZ, PBT and PBEB from 1 m3 raw hospital wastewater was found as 8.70 € and 2.28 € at the optimum experimental conditions with nano-TiO2 and nano-CeO2, respectively (Table 5). For both nanoparticles, 4 × 10−3 € was spent on 20 UV lamps and 9.5 × 10−2 € was determined as the investment cost. The chemical cost of synthesizing the TiO2 nanoparticle and electricity consumption cost were only 8.60 €, while the chemical cost to synthesize the CeO2 nanoparticle and electricity consumption cost were only 2.18 €. The cost analysis for the photodegradation of TCS, GFZ, PBT and PBEB from raw hospital wastewater showed that the photocatalytic removal with nano-CeO2 is cheaper compared to nano-TiO2. In this study, the main part of the cost consisted of the chemicals (8.57 € for nano-TiO2 synthesis and 2.15 € for nano-CeO2 synthesis). Conducting the studies in sunlight significantly decreased the cost of the study. Since these studies were performed during January and February in the laboratory, the power of the sun was low. Furthermore, a UV light simulator was not used. Preliminary studies showed that photocatalytic experiments' performed in sunlight decreased the total cost by 76%.

Table 5

Cost analysis for the photocatalytic treatment of TCS, GFZ, PBT and PBEB for 1 m3 raw hospital wastewater with nano-TiO2 and nano-CeO2 at the optimum experimental conditions (45 min, 500 g L−1 nanoparticle concentration, 210 W UV light power, pH = 8.50 at 25 °C)

Cost analysisPhotocatalytic treatment of TCS, GFZ, PBT and PBEB with nano-TiO2Photocatalytic treatment of TCS, GFZ, PBT and PBEB with nano-CeO2
UV 1 UV lamp: 0.2 × 10−3
20 UV lamp: 20 × (0.2 × 10−3 €) = 4 × 10−3 € 
1 UV lamp: 0.2 × 10−3
20 UV lamp: 20 × (0.2 × 10−3 €) = 4 × 10−3 € 
Photocatalytic reactor
Isolated cabin
Magnetic stirrer 
0.11 × 10−3 €/h → (0.11 × 10−3 €/h) × (0.75 h) =8.25 × 10−5
0.017 €/h → (0.017 €/h) × (0.75 h) =1.28 × 10−2
0.11 €/h → (0.11 €/h) × (0.75 h) =8.25 × 10−2 € 
0.11 × 10−3 €/h → (0.11 × 10−3 €/h) × (0.75 h) = 8.25 × 10−5
0.017 €/h → (0.017 €/h) × (0.75 h) = 1.28 × 10−2
0.11 €/h → (0.11 €/h) × (0.75 h) = 8.25 × 10−2 € 
Electricity consumption 1 UV lamp: 30 W power
1 kWh =1,000 W
45 min =0.75 h
1 kWh electricity: 0.06 €
(30 W/lamp) × (20 lamps) × (1 kWh/1,000 W) × (0.75 h) ×(0.06 €/1 kWh) =0.03 € 
1 UV lamp: 30 W power
1 kWh = 1,000 W
45 min = 0.75 h
1 kWh electricity: 0.06 €
(30 W/lamp) × (20 lamps) × (1 kWh/1,000 W) × (0.75 h) ×
(0.06 €/1 kWh) = 0.03 € 
Chemicals Titanium (IV) chloride (TiCl4) solution (500 mL):  45.92 €
Pluronics P123 (1 kg): 150.00 € 
Polyoxyethylene octylphenol ether (500 mL): 250.00 €
1-hexanol (500 mL): 75.00 €
Cyclohexane (500 mL): 75.00 €
Cerium (III) nitrate hexahydrate (Ce(NO3)3.6H2O)
(125 g): 337.50 € 
Nanoparticle synthesis For 500 g nano-TiO2:
92 mL TiCl4 solution: 8.45 €
0.80 g P123: 0.12 € 
For 500 g nano-CeO2:
1 mL polyoxyethylene octylphenol ether: 0.50 €
1 mL 1-hexanol: 0.15 €
1 mL cyclohexane: 0.15 €
0.50 g Ce(NO3)3.6H2O: 1.35 € 
Total cost 4 × 10−3 € + 8.25 × 10−5 € + 1.28 × 10−2 € + 8.25 × 10−2
+ 0.03 € + 8.45 € + 0.12 € = 8.70 € 
4 × 10−3 € + 8.25 × 10−5 € + 1.28 × 10−2 € + 8.25 × 10−2
+ 0.03 € + 0.50 € + 0.15 € + 0.15 € + 1.35 € = 2.28 € 
Cost analysisPhotocatalytic treatment of TCS, GFZ, PBT and PBEB with nano-TiO2Photocatalytic treatment of TCS, GFZ, PBT and PBEB with nano-CeO2
UV 1 UV lamp: 0.2 × 10−3
20 UV lamp: 20 × (0.2 × 10−3 €) = 4 × 10−3 € 
1 UV lamp: 0.2 × 10−3
20 UV lamp: 20 × (0.2 × 10−3 €) = 4 × 10−3 € 
Photocatalytic reactor
Isolated cabin
Magnetic stirrer 
0.11 × 10−3 €/h → (0.11 × 10−3 €/h) × (0.75 h) =8.25 × 10−5
0.017 €/h → (0.017 €/h) × (0.75 h) =1.28 × 10−2
0.11 €/h → (0.11 €/h) × (0.75 h) =8.25 × 10−2 € 
0.11 × 10−3 €/h → (0.11 × 10−3 €/h) × (0.75 h) = 8.25 × 10−5
0.017 €/h → (0.017 €/h) × (0.75 h) = 1.28 × 10−2
0.11 €/h → (0.11 €/h) × (0.75 h) = 8.25 × 10−2 € 
Electricity consumption 1 UV lamp: 30 W power
1 kWh =1,000 W
45 min =0.75 h
1 kWh electricity: 0.06 €
(30 W/lamp) × (20 lamps) × (1 kWh/1,000 W) × (0.75 h) ×(0.06 €/1 kWh) =0.03 € 
1 UV lamp: 30 W power
1 kWh = 1,000 W
45 min = 0.75 h
1 kWh electricity: 0.06 €
(30 W/lamp) × (20 lamps) × (1 kWh/1,000 W) × (0.75 h) ×
(0.06 €/1 kWh) = 0.03 € 
Chemicals Titanium (IV) chloride (TiCl4) solution (500 mL):  45.92 €
Pluronics P123 (1 kg): 150.00 € 
Polyoxyethylene octylphenol ether (500 mL): 250.00 €
1-hexanol (500 mL): 75.00 €
Cyclohexane (500 mL): 75.00 €
Cerium (III) nitrate hexahydrate (Ce(NO3)3.6H2O)
(125 g): 337.50 € 
Nanoparticle synthesis For 500 g nano-TiO2:
92 mL TiCl4 solution: 8.45 €
0.80 g P123: 0.12 € 
For 500 g nano-CeO2:
1 mL polyoxyethylene octylphenol ether: 0.50 €
1 mL 1-hexanol: 0.15 €
1 mL cyclohexane: 0.15 €
0.50 g Ce(NO3)3.6H2O: 1.35 € 
Total cost 4 × 10−3 € + 8.25 × 10−5 € + 1.28 × 10−2 € + 8.25 × 10−2
+ 0.03 € + 8.45 € + 0.12 € = 8.70 € 
4 × 10−3 € + 8.25 × 10−5 € + 1.28 × 10−2 € + 8.25 × 10−2
+ 0.03 € + 0.50 € + 0.15 € + 0.15 € + 1.35 € = 2.28 € 

Note: Recent euro exchange rate was used to convert Turkish liras to euro (1 TL = 0.302 €).

No recent literature was found concerning the photocatalytic cost of the micropollutants studied. However, some costs were compared with our data. Zhang et al. (2011) calculated that 99 € was needed to treat 500 m3 of wastewater containing 34 mg L-1 methyl orange dye using only 1.9 g L−1 TiO2. Fenoll et al. (2012) and Wandre et al. (2016) found that the cost of photodegrading 1 m3 of wastewater containing 12 mg L−1 TCS and 20 mg L−1 methyl orange was 102 € and 127 €, respectively, in the presence of 2 g L−1 nano-TiO2 and CeO2. These costs are higher than our cost of treating the raw wastewater containing the four micropollutants.

In this study the nano-TiO2 and nano-CeO2 were produced under laboratory conditions to photodegrade TCS, GFZ, PBT and PBEB from raw hospital wastewater for the first time in Turkey. XRD analysis showed that the TiO2 nanoparticle completely originated from the pure anatase phase with a tetragonal crystal structure. The XRD patterns of CeO2 showed that they had a cubic fluorite structure and exhibited strong agglomeration. The FT-IR spectrum of the TiO2 nanoparticle showed the bending vibration of OH molecules, the carbonyl peak and stretching vibrations of Ti-OH. The FT-IR spectrum peaks of CeO2 nanoparticles originated from the (Ce-O) metal–oxygen bond vibrations.

Under optimized operational conditions (45 min irradiation time, 0.50 g L−1 nanoparticle concentration, 210 W UV light power, pH = 8.50 at 25 °C) the nano-CeO2 had a higher photocatalytic activity (97% TCS, 95% GFZ, 92% PBT and 90% PBEB) than nano-TiO2 (94% TCS, 95% GFZ, 90% PBT and 87% PBEB) to remove these micropollutants from raw hospital wastewater since its pore volume, average size and pore diameter were low with high band gap energy. The slightly lower alkaline pH (8.50) maximized the photo-removals of micropollutants from the raw hospital wastewater. TiO2-photocatalytic efficiency decreased in acidic conditions by inhibiting the generation of OH radicals. The nano-CeO2 and nano-TiO2 used in this study were recovered effectively to be used again with photodegradation yields as high as 85–94% and 83–90%, respectively. The cost of treating the TCS, GFZ, PBT and PBEB from 1 m3 raw hospital wastewater was 2.28 € with CeO2 nanoparticles. The results of this study showed that TCS, GFZ, PBT and PBEB in raw hospital wastewaters can be effectively removed with photodegradation by the utilization of nano-CeO2 and nano-TiO2 generated under laboratory conditions for the first time in Turkey. This process could be used effectively in other pharmaceutical and hospital wastewater studies, both as pilot projects and as real treatment in the field.

The authors would like to thank TUBITAK for the PhD dissertation grant given to Gokce Guney on the Priority Section Subjects (Grant number 2211-C Prior Sections; Water; Prior, Specific and Micro Pollutants).

APHA/AWWA/WEF
2012
Standard Methods for the Examination of Water and Wastewater
, 22nd edn.
American Public Health Association/American Water Works Association/Water Environment Federation
,
Washington, DC
,
USA
.
Balavi
,
H.
,
Samadanian-Isfahani
,
S.
,
Mehrabani-Zeinabad
,
M.
&
Edrissi
,
M.
2013
Preparation and optimization of CeO2 nanoparticles and its application in photocatalytic degradation of reactive orange 16 dye
.
Powder Technology
249
,
549
555
.
Bethi
,
B.
,
Sonawane
,
S. H.
,
Bhanvase
,
B. A.
&
Gumfekar
,
S. P.
2016
Nanomaterials-based advanced oxidation processes for wastewater treatment: a review
.
Chemical Engineering and Processing
109
,
178
189
.
Bianco
,
A.
,
Fabbri
,
D.
,
Minella
,
M.
,
Brigante
,
M.
,
Mailhot
,
G.
,
Maurino
,
V.
,
Minero
,
C.
&
Vione
,
D.
2015
New insights into the environmental photochemistry of 5-chloro-2-(2,4-dichlorophenoxy)phenol (triclosan): reconsidering the importance of indirect photoreactions
.
Water Research
72
,
271
280
.
Caudillo-Flores
,
U.
,
Muñoz-Batista
,
M. J.
,
Ung-Medina
,
F.
,
Alonso-Núñez
,
G.
,
Kubacka
,
A.
,
Cortés
,
J. A.
&
Fernández-García
,
M.
2016
Effect of the anatase–rutile contact in gas phase toluene photodegradation quantum efficiency
.
Chemical Engineering Journal
299
,
393
402
.
Channei
,
D.
,
Nakaruk
,
A.
&
Phanichphant
,
S.
2017
Photocatalytic degradation of dye using CeO2/SCB composite catalysts
.
Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy
183
,
218
224
.
Chen
,
L.
,
Wang
,
Z.
,
Wang
,
Z.
&
Gu
,
X.
2016a
Influence of humic acid on the photolysis of triclosan in different dissociation forms
.
Water Air and Soil Pollution
227
(
9
).
DOI 10.1007/s11270-016-3024-7
.
Chen
,
L.
,
Wang
,
Z.
,
Cao
,
S.
,
Gu
,
X.
,
Gong
,
S.
&
Shen
,
Z.
2016b
Impact of humic acid on the photo-induced toxicity of triclosan
.
Water Air and Soil Pollution
227
(
7
).
DOI 10.1007/s11270-016-2940-x
.
Chen
,
P.
,
Wang
,
F.
,
Chen
,
Z. F.
,
Zhang
,
Q.
,
Su
,
Y.
,
Shen
,
L.
,
Yao
,
K.
,
Liu
,
Y.
,
Cai
,
Z.
,
Lv
,
W.
&
Liu
,
G.
2017
Study on the photocatalytic mechanism and detoxicity of gemfibrozilby a sunlight-driven TiO2/carbon dots photocatalyst: the significant roles of reactive oxygen species
.
Applied Catalysis B: Environmental
204
,
250
259
.
Chow
,
K. L.
,
Man
,
Y. B.
,
Zheng
,
J. S.
,
Liang
,
Y.
,
Tam
,
N. F. Y.
&
Wong
,
M. H.
2012
Characterizing the optimal operation of photocatalytic degradation of BDE-209 by nano-sized TiO2
.
Journal of Environmental Sciences
24
(
9
),
1670
1678
.
Chowdhury
,
S.
&
Balasubramanian
,
R.
2014
Recent advances in the use of graphene-family nanoadsorbents for removal of toxic pollutants from wastewater
.
Advances in Colloid and Interface Science
204
,
35
56
.
Corena
,
J.
2015
Heterogeneous Photocatalysis for the Treatment of Contaminants of Emerging Concern in Water
.
PhD thesis
,
Worcester Polytechnic Institute
,
Massachusetts
,
USA
.
Covaci
,
A.
,
Harrad
,
S.
,
Abdallah
,
M. A. E.
,
Ali
,
N.
,
Law
,
R. J.
,
Herzke
,
D.
&
De Wit
,
C. A.
2011
Novel brominated flame retardants: a review of their analysis, environmental fate and behaviour
.
Environment International
37
,
532
556
.
Cristale
,
J.
,
Katsoyiannis
,
A.
,
Sweetman
,
A. J.
&
Jones
,
K. C.
2013
Occurrence and risk assessment of organophosphorus and brominated flame retardants in the River Aire (UK)
.
Environmental Pollution
179
,
194
200
.
Diwan
,
V.
,
Tamhankar
,
A. J.
,
Khandal
,
R. K.
,
Sen
,
S.
,
Aggarwal
,
M.
,
Marothi
,
Y.
,
Iyer
,
R. V.
,
Sundblad-Tonderski
,
K.
&
Stalsby-Lundborg
,
C.
2010
Antibiotics and antibiotic-resistant bacteria in waters associated with a hospital in Ujjain, India
.
BMC Public Health
10
,
414
422
.
Elsherif
,
Z. A.
,
Abas
,
S. S.
,
Abdelwahab
,
M. H.
&
El-Weshahy
,
S.
2013
Determination of gemfibrozil and fenofibrate in pharmaceuticals in presence of their degradation products
.
International Journal of Pharmacy and Pharmaceutical Sciences
5
(
3
),
886
896
.
Ezechiáš
,
M.
,
Covino
,
S.
&
Cajthaml
,
T.
2014
Ecotoxicity and biodegradability of new brominated flame retardants: a review
.
Ecotoxicology and Environmental Safety
110
,
153
167
.
Fang
,
Y.
,
Karnjanapiboonwong
,
A.
,
Chase
,
D. A.
,
Wang
,
J.
,
Morse
,
A. N.
&
Anderson
,
T. A.
2012
Occurrence, fate, and persistence of gemfibrozil in water and soil
.
Environmental Toxicology and Chemistry
31
,
550
555
.
Farahmandjou
,
M.
,
Zarinkamar
,
M.
&
Firoozabadi
,
T. P.
2016
Synthesis of cerium oxide (CeO2) nanoparticles using simple CO-precipitation method
.
Revista Mexicana de Física
62
,
496
499
.
Fenoll
,
J.
,
Flores
,
P.
,
Hellín
,
P.
,
Martínez
,
C. M.
&
Navarro
,
S.
2012
Photodegradation of eight miscellaneous pesticides in drinking water after treatment with semiconductor materials under sunlight at pilot plant scale
.
Chemical Engineering Journal
204–206
,
54
64
.
García
,
A.
,
Delgado
,
L.
,
Torà
,
J. A.
,
Casals
,
E.
,
González
,
E.
,
Puntes
,
V.
,
Font
,
X.
,
Carrera
,
J.
&
Sánchez
,
A.
2012
Effect of cerium dioxide, titanium dioxide, silver, and gold nanoparticles on the activity of microbial communities intended in wastewater treatment
.
Journal of Hazardous Materials
199–200
,
64
72
.
Grenni
,
P.
,
Patrolecco
,
L.
,
Ademollo
,
N.
,
Tolomei
,
A.
&
Caracciolo
,
A. B.
2013
Degradation of gemfibrozil and naproxen in a river water ecosystem
.
Microchemical Journal
107
,
158
164
.
Hassani
,
A.
,
Khataee
,
A.
,
Karaca
,
S.
,
Karaca
,
C.
&
Gholami
,
P.
2017
Sonocatalytic degradation of ciprofloxacin using synthesized TiO2 nanoparticles on montmorillonite
.
Ultrasonics Sonochemistry
35
,
251
262
.
Hu
,
A.
,
Zhang
,
X.
,
Luong
,
D.
,
Oakes
,
K. D.
,
Servos
,
M. R.
,
Liang
,
R.
,
Kurdi
,
S.
,
Peng
,
P.
&
Zhou
,
Y.
2012
Adsorption and photocatalytic degradation kinetics of pharmaceuticals by TiO2 nanowires during water treatment
.
Waste and Biomass Valorization
.
DOI 10.1007/s12649-012-9142-6
.
Hu
,
X.
,
Zhu
,
Q.
,
Gu
,
Z.
,
Zhang
,
N.
,
Liu
,
N.
,
Stanislaus
,
M. S.
,
Li
,
D.
&
Yang
,
Y.
2017
Wastewater treatment by sonophotocatalysis using PEG modified TiO2 film in a circular photocatalytic-ultrasonic system
.
Ultrasonics Sonochemistry
36
,
301
308
.
Krkošek
,
W. H.
,
Koziar
,
S. A.
,
White
,
R. L.
&
Gagnon
,
G. A.
2011
Identification of reaction products from reactions of free chlorine with the lipid-regulator gemfibrozil
.
Water Research
45
,
1414
1422
.
Ma
,
J.
,
Lv
,
W.
,
Ping
,
C.
,
Lu
,
Y.
,
Wang
,
F.
,
Li
,
F.
,
Yao
,
K.
&
Liu
,
G.
2016
Photodegradation of gemfibrozil in aqueous solution under UV irradiation: kinetics, mechanism, toxicity, and degradation pathways
.
Environmental Science and Pollution Research
23
,
14294
14306
.
Maaroof
,
M.
&
Uysal
,
Y.
2014
Use the extended activated sludge process to remove triclosan (TCS) from wastewater treatment plant
.
KSU Journal of Engineering Sciences
17
,
20
26
.
Montaseri
,
H.
&
Forbes
,
P. B. C.
2016
A review of monitoring methods for triclosan and its occurrence in aquatic environments
.
Trends in Analytical Chemistry
85
,
221
231
.
Neves
,
A.
2014
Photochemical Degradation of Triclosan: A Comparison between Different Light Sources
.
PhD thesis
,
Faculdade De Ciencias E Tecnologia Universidade Nova De Lisboa
,
Lisbon
,
Portugal
.
Petrie
,
B.
,
McAdam
,
E. J.
,
Lester
,
J. N.
&
Cartmell
,
E.
2014
Obtaining process mass balances of pharmaceuticals and triclosan to determine their fate during wastewater treatment
.
Science of the Total Environment
497–498
,
553
560
.
Quinn
,
G. P.
&
Keough
,
M. J.
2002
Experimental Design and Data Analysis for Biologists
.
Cambridge University Press
,
New York
.
Ravishankar
,
T. N.
,
Ramakrishnappa
,
T.
,
Nagaraju
,
G.
&
Rajanaika
,
H.
2015
Synthesis and characterization of CeO2 nanoparticles via solution combustion method for photocatalytic and antibacterial activity studies
.
Chemistry Open.
DOI: 10.1002/open.201402046
.
Son
,
H. S.
,
Ko
,
G.
&
Zoh
,
K. D.
2009
Kinetics and mechanism of photolysis and TiO2 photocatalysis of triclosan
.
Journal of Hazardous Materials
166
,
954
960
.
Stamatis
,
N.
,
Antonopoulou
,
M.
,
Hela
,
D.
&
Ioannis Konstantinou
,
I.
2014
Photocatalytic degradation kinetics and mechanisms of antibacterial triclosan in aqueous TiO2 suspensions under simulated solar irradiation
.
Journal of Chemical Technology and Biotechnology
89
,
1145
1154
.
Sun
,
R. X.
,
Luo
,
X. J.
,
Tan
,
X. X.
,
Tang
,
B.
,
Li
,
Z. R.
&
Mai
,
B. X.
2015
An eight year (2005–2013) temporal trend of halogenated organic pollutants in fish from the Pearl River Estuary, South China
.
Marine Pollution Bulletin
93
,
61
67
.
Tambat
,
S.
,
Umale
,
S.
&
Sontakke
,
S.
2016
Photocatalytic degradation of Milling Yellow dye using sol–gel synthesized CeO2
.
Materials Research Bulletin
76
,
466
472
.
Tang
,
Y.
,
Yang
,
M.
,
Gao
,
H.
,
Li
,
J.
&
Wang
,
G.
2016
A facile approach for fabrication of TiO2 hierarchical nanostructures and their photocatalytic properties
.
Colloids and Surfaces A: Physicochemical and Engineering Aspects
508
,
184
191
.
Wandre
,
T. M.
,
Gaikwad
,
P. N.
,
Tapase
,
A. S.
,
Garadkar
,
K. M.
,
Vanalakar
,
S. A.
,
Lokhande
,
P. D.
,
Sasikala
,
R.
&
Hankare
,
P. P.
2016
Sol–gel synthesized TiO2–CeO2 nanocomposite: an efficient photocatalyst for degradation of methyl orange under sunlight
.
Journal of Materials Science: Materials in Electronics
27
(
1
),
825
833
.
Wang
,
H.
,
Li
,
Y.
,
Zhang
,
Y.
,
Chen
,
J.
,
Chu
,
G.
&
Shao
,
L.
2015
Preparation of CeO2 nano-support in a novel rotor–stator reactor and its use in Au-based catalyst for CO oxidation
.
Powder Technology
273
,
191
196
.
Wu
,
J. P.
,
Guan
,
Y. T.
,
Zhang
,
Y.
,
Luo
,
X. J.
,
Zhi
,
H.
,
Chen
,
S. J.
&
Mai
,
B. X.
2010
Trophodynamics of hexabromocyclododecanes and several other non-PBDE brominated flame retardants in a freshwater food web
.
Environmental Science & Technology
44
,
5490
5495
.
Yu
,
J. C.
,
Kwong
,
T. Y.
,
Luo
,
Q.
&
Cai
,
Z.
2006
Photocatalytic oxidation of triclosan
.
Chemosphere
65
,
390
399
.
Zhang
,
H.
,
Xu
,
P.
,
Du
,
G.
,
Chen
,
Z.
,
Oh
,
K.
,
Pan
,
D.
&
Jiao
,
Z.
2011
A facile one-step synthesis of TiO2/graphene composites for photodegradation of methyl orange
.
Nano Research
4
(
3
),
274
283
.