Next Article in Journal
Parameterization for Modeling Blue–Green Infrastructures in Urban Settings Using SWMM-UrbanEVA
Previous Article in Journal
An Improved ResNet-Based Algorithm for Crack Detection of Concrete Dams Using Dynamic Knowledge Distillation
Previous Article in Special Issue
Assessing the Sustainability of Photodegradation and Photocatalysis for Wastewater Reuse in an Agricultural Resilience Context
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Surface Modification of TiO2/g-C3N4 Electrode with N, P Codoped CQDs for Photoelectrocatalytic Degradation of 1,4-Dioxane

1
Guangdong Provincial Key Laboratory of Environmental Pollution Control and Remediation Technology, School of Environmental Science and Engineering, Sun Yat-sen University, Guangzhou 510006, China
2
Institute of Environmental and Ecological Engineering, Guangdong University of Technology, Guangzhou 510006, China
*
Author to whom correspondence should be addressed.
Water 2023, 15(15), 2837; https://doi.org/10.3390/w15152837
Submission received: 17 June 2023 / Revised: 27 July 2023 / Accepted: 4 August 2023 / Published: 6 August 2023

Abstract

:
The aim of this study was to synthesize N, P codoped CQDs modifying TiO2/g-C3N4 nanorod array (i.e., N, P-CQD/TCN NA) photoanodes for the degradation of 1,4-dioxane (1,4-D) and to explore the possibility of the photoelectrocatalytic (PEC) process in wastewater treatment. With the characterization of N, P-CQD/TCN NA anodes, 1,4-D degradation and pesticide wastewater was tested in the PEC cell, respectively. Under a bias voltage of 1.2 V and visible light, the current density of the N, P-CQD/TCN NAs was much higher than that of the CQD/TCN NAs (0.15 vs. 0.11 mA/cm2). The removal of 1,4-D reached 97% in the PEC cell within 6 h. The high performance of the N, P-CQD/TCN NA anodes could be attributed to the efficient charge separation, narrowed energy gap, and high upconverted PL properties. The C4 and C6 positions of 1,4-D were the preferential sites for the nucleophilic attack to form intermediates. The COD removal in the pesticide wastewater was kept stable at ~80% in a five-cycle operation using the PEC cell with the N, P-CQD/TCN NA photoelectric anodes. The results from this study should provide a promising way to develop novel photoelectric catalysts and to expand PEC application in wastewater treatment.

1. Introduction

1,4-dioxane (1,4-D) is an emerging organic pollutant with strong hydrophilicity and high stability in aqueous environments [1,2]. 1,4-D has been extensively detected in urban sewage systems, groundwater, and surface water worldwide [3,4,5]. The 1,4-D concentration in polluted groundwater can reach 10 mg/L in California [6,7]. The average effluent concentration of 1,4-D reached 62.3 ± 36.0 mg/L in a domestic sewage treatment plant with biological denitrification [8]. 1,4-D may cause serious damage in humans and environmental problems and has been classified a suspected carcinogen and an endocrine-disrupting chemical by reputable organizations such as the Agency for Toxic Substances and Disease Registry and the International Agency for Research on Cancer [9,10]. The concentration of 1,4-D in drinking water is recommended to be below 50 μg/L by the World Health Organization [1]. Due to a low Henry’s law constant (4.80 × 10−6 atm m3/mol at 25 °C) and high solubility, 1,4-D is difficult to remove using coagulation and adsorption [1,2]. 1,4-D removal using conventional chlorine treatment may produce chlorinated byproducts that are more toxic than itself [10]. Therefore, it is necessary to develop an innovative approach for the effective and rapid removal of 1,4-D in aqueous environments.
The advanced oxidation process (AOP) is an effective and feasible way to remove 1,4-D from various water sources by producing highly oxidative radicals such as hydroxyl radicals (OH·) [11,12,13]. Among the different AOPs, the photoelectrocatalytic (PEC) process has attracted much attention in recent years [14,15,16]. In the PEC process, a specific bias voltage is applied to the photoanode, facilitating efficient electron–hole pair separation and enhancing the degradation of organic pollutants [17]. The PEC process can efficiently remove organic pollutants because of its photoelectric–synergistic effects [18,19,20]. The catalytic capability of the TiO2 photoanode has been significantly improved using graphitic carbon nitride (g-C3N4) modification and heterojunction construction [21]. With an in situ growth method, TiO2 nanocrystals were successfully synthesized on the surface of g-C3N4 nanosheets, resulting in the fabrication of a TiO2/g-C3N4 (TCN) heterojunction. The formation of this heterojunction led to a synergistic effect, significantly enhancing photocatalytic activity under visible light irradiation [22]. Under visible light illumination, the photocurrent of TiO2 nanorod arrays modified with g-C3N4 led to a remarkable tenfold increase in photocurrent compared to TiO2 nanorod arrays [23]. Moreover, carbon-based quantum dots (CQDs), serving as electron reservoirs and transporters to shuttle the electrons between two semiconductors, can be used to suppress the electron–hole recombination in the TCN heterojunction [24,25]. Our previous results show that TiO2/g-C3N4/CQD nanorod arrays (denoted as TCNC NAs) produced a higher photocurrent than TiO2/g-C3N4 nanorod arrays (denoted as TCN NAs) at a 1.2 V bias voltage under the irradiation of visible light, and 1,4-D removal was 50% higher than that with TCN NAs (77.9% vs. 52.5%) [17]. However, the removal of 1,4-D in the PEC process still needs to be improved to realize the efficient degradation of 1,4-D.
CQDs can be doped by nonmetals such as N, P, and S to promote electron mobility and reaction sites [26,27,28]. The photocatalyst composed of P-, S-doped g-C3N4/2D TiO2 had a Z-scheme heterojunction architecture, which promoted efficient interfacial charge transfer between the TiO2 nanosheets and P, S-g-C3N4, facilitating the effective separation of photogenerated charge carriers [28]. Within 15 min, the methylene blue dye was completely degraded using N and P codoped carbon quantum dots (N, P-CQDs) as the catalyst. The degradation process exhibited a high quantum yield of 8.45%, indicating efficient conversion of the dye molecules into nontoxic products [29]. The heterostructures of TCN modified with nonmetal-doped CQDs may provide a feasible way to enhance the activity of photoanodes in the PEC process. Moreover, although various photoelectric catalysts have been synthesized and shown a high removal efficiency on refractory contaminants in artificial wastewater, only a few of them were used to treat real wastewater [26,27,28]. One reason may be that real wastewater usually contains various inorganic and organic compounds and could significantly decrease the ability of the photocatalyst. It is necessary to test the novel photoanode of the PEC cell in real wastewater treatment to clarify its potential application in the future.
The objective of this study was to synthesize N, P codoped CQDs modifying TCN nanorod array (i.e., N, P-CQD/TCN NA) photoanodes for 1,4-D degradation and to explore the possibility of the PEC process in wastewater treatment. The morphology, structure, and photoelectric properties of the photoelectrodes were characterized. The activity of N, P-CQD/TCN NAs was tested using 1,4-D degradation, with determination of the predominant active species. The degradation pathways of 1,4-D were proposed according to ion chromatography (IC) and density functional theory (DFT) calculation. Real pesticide wastewater containing various refractory organics were selected to test the stability of N, P-CQD/TCN NAs and the possibility of the PEC process in wastewater treatment.

2. Materials and Methods

2.1. Synthesis of TiO2 NA and TCN NA Photoelectrodes

Fluorine-doped tin oxide (FTO) with a diameter of 34 mm and sheet resistance of 7 Ω/square was obtained from Luoyang Shangzhuo Technology Co., Ltd., Luoyang, China. FTO was pretreated, as previously described by Su et al. [17]. Titanium butoxide (99%) was procured from Tianjin Damao Chemical Reagent Co., Ltd., Tianjin, China. Melamine (99%) was obtained from Macklin Biochemical Co., Ltd., Shanghai, China. The remaining reagents utilized in this study were of analytical grade quality. The TiO2 NAs were prepared based on a hydrothermal method [17,30]. Concentrated hydrochloric acid (36~38%) and ultrapure water were mixed in a ratio of 1:1 (v/v). Subsequently, 1.32 mL titanium butoxide was added into 80 mL mixture and stirred for 5 min. All the solution was transferred to a Teflon autoclave, with an FTO glass horizontally placed in the autoclave for 5 h for hydrothermal treatment at 150 °C. After the autoclave cooled down, the FTO glass was taken off and washed using ultrapure water. The dried FTO deposited with the TiO2 NRA was further heated in air at 550 °C for 3 h. The TCN NA photoelectrode was prepared using the vapor deposition method, as previously reported [17]. Two grams of melamine was carefully positioned in a covered alumina crucible with the TiO2 NAs placed horizontally above. After heating at 550 °C for 3 h, the conductive surface of TiO2 NAs changed from white to yellow. TCN NA photoelectrode was successfully synthesized with the g-C3N4 depositing on the TiO2 NAs.

2.2. Synthesis of N, P-CQD/TCN NA Photoanodes

The N, P codoped CQDs were synthesized based on a thermal treatment [17,29]. In brief, 0.21 g citric acid and 0.148 g O-phosphorylethanolamine were added to a Teflon-lined autoclave and heated at 200 °C for 12 h. After cooling to room temperature, the obtained solid was ultrasonically dissolved in the water and then centrifuged at 10,000 rpm for 15 min. The collected solution was dialyzed in ultrapure water using 500 Da cellulose ester dialysis membrane. A N, P codoped CQD solution was obtained and collected for next step. An undoped CQD solution was also prepared using the same procedure above without adding O-phosphorylethanolamine. N, P-CQD/TCN NA photoanodes were prepared through immerging TCN NA photoelectrode in the N, P codoped CQD solution and drying at 60 °C for 10 h. CQD/TCN NA photoanodes as the controls were also prepared with the undoped CQD solution above, using the same method.

2.3. PEC Experiments for 1,4-D Degradation

The PEC properties of the as-prepared photoanodes were characterized using an electrochemical workstation (CHI1000C, Shanghai Chenhua Instrument Co., LTD., Shanghai, China) [17]. For the three-electrode system, 0.1 M Na2SO4 was selected as the electrolyte. In the PEC tests, the photoanode served as the working electrode, the saturated calomel electrode (SCE) was used as the reference electrode, and a platinum wire acted as the counter electrode. An LED visible light source (≥400 nm) and a power intensity of 80–85 mW/cm2 were employed for illumination in the PEC tests. According to our pretests, a bias voltage of 1.2 V was applied to the working electrodes.
The degradation of 1,4-D was tested in the PEC cell with TiO2 NA, TCN NA, CQD/TCN NA, and N, P-CQD/TCN NA photoanodes, respectively [17]. The PEC cell was composed of a plexiglass container, with an effective volume of 20 mL, and the synthesized photoanode and a platinum wire were selected as the anode and cathode, respectively. The bias voltage applied on the photoanode of the PEC cell was 1.2 V throughout all the tests. The electrolyte consisted of 100 mg/L 1,4-D and 0.1 M Na2SO4. The PEC cell without visible light irradiation or without bias voltage was also tested as the controls to explore 1,4-D degradation, respectively.

2.4. PEC Experiments for Pesticide Wastewater

Pesticide wastewater samples were collected from a pesticide pharmaceutical factory located in Jiangmen City, Guangdong Province, China. The raw pesticide wastewater was pretreated using coagulation, sedimentation, and 0.45 μm microfiltration sequentially. The pretreated pesticide wastewater was characterized as follows: pH 6.8 ± 0.5, COD 350 ± 30 mg/L, and conductivity 5.3 ± 0.2 mS/cm. The performance of PEC cell with the photoanode of N,P-CQD/TCN NAs was conducted using the pretreated pesticide wastewater under a bias voltage of 1.2 V and pH of 9.

2.5. Analyses and Calculations

2.5.1. Analysis Methods

Scanning electron microscopy (SEM, Hitachi Su8010, Tokyo, Japan) and transmission electron microscopy (TEM, JEM-2100HR, JEOL, Japan) were used to characterize the morphology and interlayer structure of the photoanode [17]. The crystal structures of the photoanode were analyzed using X-ray diffraction (XRD, PAN-alytical, Almelo, The Netherland) with Cu Kα radiation (λ = 0.15406 nm) [17]. The optical properties of the photoanode were analyzed using UV-vis diffuse reflectance spectroscopy (DRS) using a PE lambda 750S instrument (Waltham, MA, USA) [17]. X-ray photoelectron spectroscopy (XPS) with monochromatic Al Kα radiation (1486.6 eV) from a Thermo Fisher Scientific instrument (Sunnyvale, CA, USA) was employed to ascertain the chemical composition of the photoanode [17]. Photoluminescence (PL) measurements were conducted using a fluorescence spectrophotometer (FLS1000, Edinburgh Analytical Instruments Ltd., Edinburgh, UK) with an excitation wavelength of 340 nm [17].
The concentration of 1,4-D was determined using high-performance liquid chromatography (HPLC, C, P230II, Dalian Yilite analytical instruments, Dalian, China). A Zorbax Eclipse XDB-C18 column (4.6 mm × 150 mm × 5 μm) was employed [17]. A mobile phase consisting of a mixture of acetonitrile and water (phosphate buffer, pH 3.0) in a volumetric ratio of 20:80 was employed in the HPLC system. The mobile phase was kept at a flow rate of 1 mL/min. The UV detector was operated at a wavelength of 190 nm, and the temperature of the column was maintained at 30 °C.

2.5.2. Reactive Species Determination

The reactive species, including O2·, 1O2, and ·OH, produced in the photoanode were detected and identified using electron spin resonance (ESR, Bruker A300, Karlsruher, Germany) spectrometry coupled with 5,5-Dimethyl-1-Pyrroline-N-Oxide (DMPO) as a spin trapping agent [17]. The main parameters of ESR spectrometer are included as follows: a center field of 3510 G, microwave frequency of 100 kHz, microwave power of 2.03 mW, and sweep time of 1 min.

2.5.3. Identified Organic Pollutants in the Pesticide Wastewater

Gas chromatography coupled with mass spectroscopy (GCMS, Thermo-Ultra Trace GC-DSQ, Thermo Fisher Scientific Inc., Waltham, MA, USA) were used to detect the organic pollutants in the pretreated pesticide wastewater [31]. A TG-5SILMS column (30 m × 0.25 mm × 0.25 μm, Thermo Scientific) was used with helium gas as carrier, operating with a flowrate of 20 mL/min. The temperature program was set up as follows: it rose up from 60 °C to 310 °C at a rate of 10 °C/min, holding at 310 °C for 10 min. The injection volume was 50 μL. The injector and ion source temperatures were maintained at 270 °C and 250 °C, respectively [31].

2.5.4. Three-Dimensional Fluorescence Spectroscopy Analysis

Samples were collected from the effluent of PEC cell at different times and were measured using three-dimensional fluorescence excitation–emission matrix (EEM) spectroscopy (F-4500, Hitachi, Tokyo, Japan). The excitation and emission wavelengths were in ranges from 200 to 500 nm and 250 to 650 nm, with a scanning interval of 5 nm, respectively. The scanning speed was 24,000 nm/min [32].

2.5.5. Calculations

1,4-D removal was calculated using Equation (1):
R = ( C 0 - C t ) C 0 × 100 %
where R is 1,4-D removal, C0 and Ct are 1,4-D concentration at 0, and t is time, respectively (mg/L).
Density functional theory (DFT) calculations were carried out to investigate the 1,4-D degradation using Gaussian 09 package [33,34]. The structure optimization of 1,4-D and its intermediates was finished using B3LYP/6-311G+(d, p) calculation. The electrophilic (f+), nucleophilic (f), and radical attacks (f0) related to condensed Fukui function were calculated to predict the reactive sites on 1,4-D as follows [33,34]:
Electrophilic   attack :   f + = q N - q N + 1
Nucleophilic   attack :   f - = q N - 1 - q N
Radical   attack :   f 0 = ( q N - 1 - q N + 1 ) 2
where q is the atom charge in the relevant state.

3. Results and Discussion

3.1. Morphology and Optical Properties of the Photoanodes

The XRD patterns showed the interlayer structures of the TiO2 NAs, TCN NAs, CQD/TCN Nas, and N, P-CQD/TCN NAs (Figure S1). The peaks of 36.09°, 41.17°, 62.69°, and 69.78° were attributed to the (101), (111), (002), (112) crystal planes and rutile (JCPD No. 21-1276) of TiO2 [35]. g-C3N4 doping enhanced the diffraction peaks of TiO2 NAs at 36.09° and 62.69°. N, P-CQD doping exhibited no apparent changes in the XRD patterns of TCN NAs, probably due to the small amount of N, P-CQDs.
The morphology of the top and cross-section of the TiO2 NA photoelectrodes is shown in Figure 1a. TiO2 NAs line up in uniform thickness, with a length of ~1.6 μm. The g-C3N4 layer was placed in the gap between the top and the surroundings of the TiO2 NAs. The morphology of the N, P-CQD/TiO2 NAs was almost the same as that of the TiO2 NAs (Figure 1b). The TEM results illustrate the dispersion of N, P-CQDs in ultrapure water, exhibiting an average diameter of approximately 6 nm (Figure 1c). The EDS analysis confirms that N, P-CQD/TCN NAs were deposited on the FTO substrate, which mainly consisted of the elements O, Ti, C, and N (Figure 1d).
The XPS analysis indicates that the surface composition of elements in the TCN NA photoanodes mainly included C, N, O, and Ti (Figure 2a). The atomic proportion of the phosphorus element in the N, P-CQD/TCN NA photoelectrodes increased from 0% to 1.42%. The high-resolution spectra of the C 1s region exhibited the presence of N-C=C and sp2-hybridized C-C bonds, according to the peaks observed at 288.2 eV and 284.9 eV, respectively (Figure 2b). Therefore, it indicates that N, P-CQDs were successfully doped on TCN NAs [36].
The high-resolution N 1s spectra (Figure 2c) show three distinct peaks observed at 401.18 eV, 400.23 eV, and 398.78 eV, which were attributed to specific nitrogen functionalities. The peak at 401.18 eV corresponded to amino functional groups (C-N-H), the peak at 400.23 eV indicated tertiary nitrogen bonds (N-(C)3), and the peak at 398.78 eV represented sp2-hybridized nitrogen (C=N-C) [37,38]. The O 1s high-resolution spectra of N, P-CQD/TCN NAs were deconvoluted into three different contributions at 532.18 eV, 531.13 eV, and 529.78 eV, which was attributed to H-O bond, P=O bond, and Ti-O-Ti linkages, respectively (Figure 2d) [39,40]. Two peaks centered at the binding energies of 464.18 eV and 458.53 eV, which corresponded to Ti 2p1/2 and Ti 2p3/2 and indicated the presence of Ti4+ in the photoanode (Figure 2e) [41]. In the high-resolution spectra of P 2p (Figure 2f), the peak at 133.43 eV could be attributed to the phosphorus from the P-O bond on the surface of N, P-CQD/TCN NAs [27].
The DRS results show that TiO2 NAs displayed basal absorption bands at 408.41 nm, which was consistent with the previously reported results (Figure 3a) [42]. Compared with TiO2 NAs, TCN NAs demonstrated a significant red shift and pronounced absorption in the visible region. The absorption of visible light was further enhanced following the incorporation of N, P-CQDs [43]. The fluorescence spectra of N, P-CQDs were used to distinguish their upconversion feature from others (Figure 3b). In the inset of Figure 3b, N, P-CQDs are well-dispersed in the water and emit brown light under sunlight irradiation and blue light in UV irradiation, respectively. N, P-CQDs exhibited excitation by long-wavelength light ranging from 600 nm to 900 nm, and their upconverted emissions were notably concentrated within the spectral range of 350 nm to 650 nm. Near-infrared light and visible light could be converted into shortwave light with higher energy through the N, P-CQDs.
The fluorescence intensity of N, P-CQD/TCN NAs was significantly lower compared to that of TiO2 NAs, TCN Nas, and CQD/TCN NAs, respectively, as shown in Figure 3c. The doping of N, P-CQDs may promote charge transfer and inhibit the recombination rate of the electron–hole in the N, P-CQD/TCN NA photoanodes, probably due to the band gap bending of the oxygen, nitrogen, and phosphorus groups [44,45,46]. The time-resolved PL spectra of the samples were employed to investigate the charge transfer characteristics of the photoanodes, as shown in Figure 3d and Table S1. The lifetime of N, P-CQD/TCN NAs (5.28 ns) was the longest among all the photoanodes, indicating that the electron–hole pairs of N, P-CQD/TCN NAs had much time for 1,4–D degradation.

3.2. Photoelectrochemical Property of the N, P-CQD/TCN NA Photoanodes

The transient photocurrent responses of different photoelectrodes were tested in this study (Figure 4). Under a bias voltage of 1.2 V, no current response was detected from the N, P-CQD/TCN NA photoanodes in the absence of visible light irradiation. Under visible light irradiation, the N, P-CQD/TCN NA photoanodes exhibited a stable current density of 0.15 mA/cm2, which was significantly higher than that of TiO2 NAs (0.024 mA/cm2), TCN NAs (0.07 mA/cm2), and CQD/TCN NAs (0.11 mA/cm2), respectively. The high current density of the N, P-CQD/TCN NA photoanodes could be ascribed to the enhancement in optical absorption and the improvement in the separation of photogenerated electron–hole pairs. The reactive species produced by the N, P-CQD/TCN NA photoanodes were determined (Figure 5). Without visible light irradiation (i.e., dark conditions), no apparent signal intensities of O2·, 1O2, and ·OH were observed after adding the DMPO trapping agent. After 10 min of visible light irradiation, the signal intensities of O2·, 1O2, and OH produced by the N, P-CQD/TCN NA photoanodes were all higher than that of the TCN NA photoelectrodes.

3.3. Degradation of 1,4-D in the PEC Cells

1,4-D could not be degraded in the photolysis (visible light irradiation) and electrolysis (a bias voltage of 1.2 V) processes after a 6 h operation without any catalyst (Figure 6). 1,4-D adsorption on the N, P-CQD/TCN NA photoanodes was less than 1.6% of the 1,4-D removal in the PEC cell according to the adsorption pretest. Under visible light irradiation and a bias voltage of 1.2 V, the N, P-CQD/TCN NA photoanodes could remove 97.0% of 1,4-D within 6 h, which was 22.1% higher than that with CQD/TCN NA photoanodes. It indicates the synergistic effect of N-, P-doped CQDs material on enhancing 1,4-D degradation. The mineralization of 1,4-D within 6 h reached 85% according to the additional TOC measurement. The degradation kinetic calculation shows that the 1,4-D degradation by the N, P-CQD/TCN NA photoanodes could be described by the pseudo-first-order kinetic equation (−ln(Ct/C0) = 0.5374 t, R = 0.9842). The intermediates of 1,4-D degradation were identified as some small molecule acids, such as oxalic acid, formic acid, and acetic acid, using ion chromatography (Figure 6b). Although 1,4-D was not completely mineralized in the PEC cell, easily degradable organics, including small molecule acids, indicated that the toxicity of 1,4-D may be greatly decreased. Formic acid concentrations increased from 0 to 30.14 mg/L within 6 h. The acetic acid and oxalic acid concentrations increased slowly and reached 10.1 mg/L and 6.12mg/L within 6h, respectively. The small molecule acids in 1,4-D degradation are nontoxic and easy to biodegrade in natural environments. Therefore, the PEC process with CQD/TCN NA photoanodes is useful for 1,4-D removal, with a lower toxic risk than other AOPs, such as conventional chlorine treatment.
Based on the DFT, the Fukui function was calculated to predict the reaction site for electrophilic, nucleophilic, and radical attacks to further clarify the degradation pathways of 1,4-D. The optimized structure of 1,4-D and the Fukui function calculations show that the C4 and C6 sites were surrounded by the isosurface of the Fukui Index, indicating that they were vulnerable to nucleophilic attack (Figure S2 and Table S2) [33,34]. The condensed Fukui Index distribution (f0) of C4 (0.064537) and C6 (0.064533) was lower than O1 (0.11822) and O2 (0.118219), indicating that the oxygen atoms of O1 and O2 were more susceptible to free radical attack [33,34]. Based on the intermediates identified through measurements and theoretical calculations, the degradation pathways of 1,4-D are proposed and illustrated in Figure 7. The intermediates, including methoxyacetaldehyde, acetaldehyde, 2-hydroxyethyl formate, and ethane-1,2-diyl diformate, were produced by the nucleophilic attack on the C4 and C6 sites of 1,4-D. The calculations of the condensed Fukui function isosurfaces of these intermediates show that acetic acid, formaldehyde, and glycolic acid were produced by free radicals attacking these intermediates. Finally, small molecular acids were mineralized into CO2 and H2O.
Photogenerated electron–hole pairs were excited by the CQD/TCN NAs Z-scheme heterojunction under visible light irradiation (Equation (5)) [47,48]. The N and P doped on the CQD/TCN NA heterojunction inhibited the recombination of the electron–hole pairs and then enhanced Equations (6) and (7). The photogenerated holes directly oxidized OH/H2O to ·OH. The photogenerated electrons from the CQDs to the TiO2 conduction band were utilized by O2 with O2· production and could be sequentially transferred into 1O2, H2O2, and ·OH. 1,4-D degraded by the reactive species of ·OH [49]. Nevertheless, N, P-CQD doping exhibited extraordinary upconverted PL properties, improving the production of reactive species such as O2·, 1O2, and ·OH, resulting in high 1,4-D removal.
N , P C Q D s / T C N   N A + h v N , P C Q D s / T C N   N A ( e + h + )
e + O 2 O 2
h + + H 2 O O H + H +
O H + 1,4 D C O 2 + H 2 O

3.4. Pesticide Wastewater Treatment in the PEC Cell

The GCMS results show that various refractory organics were identified in the pesticide wastewater, including phytol, ametryn, 1,4-dibutyl benzene-1,4-dicarboxylate, and bis(2-ethylhexyl) phthalate, etc. (Figure S3 and Table S3). A stable and high COD removal was achieved in the PEC cell within five repeated operations (Figure 8). The COD removal was kept at ~80.0% with the repeated usage of the photoelectric anode within 60 min, indicating that the PEC cell could efficiently treat the pesticide wastewater. The spectra of the EEM showed that the pesticide wastewater had a peak (λEX/λEM = 290–310 nm/390–410 nm), which could be mainly attributed to humic-like compounds. The strength and range of the peak in the spectra of EEM became weak and small with long-term operation in the PEC cell (Figure 9). Therefore, the photoelectric anodes of N, P-CQD/TCN NAs showed high catalytic oxidation on various organic pollutants in the pesticide wastewater. Nevertheless, our PEC cell would have promising potential in the application of pesticide wastewater treatment.

4. Conclusions

N, P-CQDs were used as a decoration material for TiO2/g-C3N4 film electrodes in this study. Compared to TiO2 NAs, TCN Nas, and CQD/TCN NAs, N, P-CQD/TCN NAs significantly enhanced PEC activity for the degradation of 1,4-D under visible light irradiation. The maximum current density of N, P-CQD/TCN NAs was higher than that of CQD/TCN NAs (0.15 vs. 0.11 mA/cm2). Under visible light and a bias voltage of 1.2 V, 1,4-D removal reached 97.0% on the N, P-CQD/TCN NA photoanodes within 6h. The modification of N, P-CQDs on the TCN NA photoanodes promoted the production of O2•−, 1O2, and ·OH active species. The high performance of N, P-CQD/TCN NAs could be explained by the efficient charge separation, narrowed energy gap, and high upconverted PL properties. The nucleophilic attack at the C4 and C6 positions of 1,4-D is found to be the preferential reaction pathway, resulting in the formation of intermediates such as oxalic acid, formic acid, and acetic acid. The N, P-CQD/TCN NA photoanodes had a stably catalytic ability in pesticide wastewater, with 80% COD removal in five repeated operations. Our results may provide a novel approach for improving the performance of the photoelectric catalysts and expanding PEC application in wastewater treatment.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/w15152837/s1, Table S1: The luminescence decay results of TiO2 NAs, TCN NAs, CQD/TCN Nas, and N, P-CQD/TCN NAs; Table S2: Condensed Fukui Index distribution on 1,4-D; Table S3: Typical organic pollutants in the pretreated pesticide wastewater identified using GC-MS; Figure S1: XRD results of TiO2 NAs, TCN NAs, CQD/TCN NAs and N, P-CQD/TCN NAs; Figure S2: Structural formula of 1,4-D optimized using DFT; Figure S3: Mass spectrometry results of the GCMS analysis on pretreated pesticide wastewater.

Author Contributions

Y.S.: Conceptualization, Formal analysis, Data curation, Writing—original draft; Y.Y.: Methodology, Formal analysis; S.L.: Methodology, Technical supervision; Y.L.: Technical supervision, Conceptualization, Validation; H.L.: Conceptualization, Funding acquisition; G.L.: Supervision, Writing—original draft preparation, Funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This work was partially supported by grants from the National Natural Science Foundation of China (No. 51978676, 42077286) and the Natural Science Foundation of Guangdong Province (No. 2022A1515011017).

Data Availability Statement

The authors confirm that the data supporting the findings of this study are available within the article.

Conflicts of Interest

The authors declare no conflict of interest. The funders played no role in the study design, data collection, analysis, interpretation, writing of the manuscript, or decision to publish the results.

References

  1. Tang, Y.; Mao, X. Recent advances in 1,4-dioxane removal technologies for water and wastewater treatment. Water 2023, 15, 1535. [Google Scholar] [CrossRef]
  2. Shukla, T.L.; Duranceau, S.J. Comparing hydrogen peroxide and sodium perborate ultraviolet advanced oxidation processes for 1,4-dioxane removal from tertiary wastewater effluent. Water 2023, 15, 1364. [Google Scholar] [CrossRef]
  3. Kang, Y.G.; Yoon, H.; Lee, W.; Kim, E.J.; Chang, Y.-S. Comparative study of peroxide oxidants activated by nZVI: Removal of 1,4-Dioxane and arsenic(III) in contaminated waters. Chem. Eng. J. 2018, 334, 2511–2519. [Google Scholar] [CrossRef]
  4. Li, W.; Xu, E.; Schlenk, D.; Liu, H. Cyto- and geno-toxicity of 1,4-dioxane and its transformation products during ultraviolet-driven advanced oxidation processes. Environ. Sci. Water Res. Technol. 2018, 4, 1213–1218. [Google Scholar] [CrossRef]
  5. Sei, K.; Kakinoki, T.; Inoue, D.; Soda, S.; Fujita, M.; Ike, M. Evaluation of the biodegradation potential of 1,4-dioxane in river, soil and activated sludge samples. Biodegradation 2010, 21, 585–591. [Google Scholar] [CrossRef] [PubMed]
  6. Feng, Y.; Li, H.; Lin, L.; Kong, L.; Li, X.Y.; Wu, D.; Zhao, H.; Shih, K. Degradation of 1,4-dioxane via controlled generation of radicals by pyrite-activated oxidants: Synergistic effects, role of disulfides, and activation sites. Chem. Eng. J. 2018, 336, 416–426. [Google Scholar] [CrossRef]
  7. Adamson, D.T.; Anderson, R.H.; Mahendra, S.; Newell, C.J. Evidence of 1,4-dioxane attenuation at groundwater sites contaminated with chlorinated solvents and 1,4-dioxane. Environ. Sci. Technol. 2015, 49, 6510–6518. [Google Scholar] [CrossRef]
  8. Stepien, D.K.; Diehl, P.; Helm, J.; Thoms, A.; Puttmann, W. Fate of 1,4-dioxane in the aquatic environment: From sewage to drinking water. Water Res. 2014, 48, 406–419. [Google Scholar] [CrossRef]
  9. Simonich, S.M.; Sun, P.; Casteel, K.; Dyer, S.; Wernery, D.; Garber, K.; Carr, G.; Federle, T. Probabilistic analysis of risks to us drinking water intakes from 1,4-dioxane in domestic wastewater treatment plant effluents. Integr. Environ. Assess. Manag. 2013, 9, 554–559. [Google Scholar] [CrossRef]
  10. Kikani, M.; Satasiya, G.V.; Sahoo, T.P.; Kumar, P.S.; Kumar, M.A. Remedial strategies for abating 1,4-dioxane pollution-special emphasis on diverse biotechnological interventions. Environ. Res. 2022, 214, 113939. [Google Scholar] [CrossRef]
  11. Hu, B.; Wang, Y.; Hu, C.; Zhou, X. Design, fabrication and high efficient visible-light assisted photoelectric-synergistic performance of 3-D mesoporous DSA electrodes. Mater. Des. 2016, 91, 201–210. [Google Scholar] [CrossRef]
  12. Liu, Z.; Zhang, X.; Nishimoto, S.; Jin, M.; Tryk, D.A.; Murakami, T.; Fujishima, A. Highly ordered TiO2 nanotube arrays with controllable length for photoelectrocatalytic degradation of phenol. J. Phys. Chem. C 2008, 112, 253–259. [Google Scholar] [CrossRef]
  13. Pan, X.; Yang, P.; Nan, H.; Yang, L.; Chen, H.; Zhao, X. Preparation and enhanced visible-light photoelectrocatalytic activity of ternary TiO2ZnO/RGO nanocomposites. Electrochim. Acta 2018, 261, 284–288. [Google Scholar] [CrossRef]
  14. Zhang, H.; Chen, G.; Bahnemann, D.W. Photoelectrocatalytic materials for environmental applications. J. Mater. Chem. 2009, 19, 5089–5121. [Google Scholar] [CrossRef]
  15. Xie, Z.; Liu, X.; Wang, W.; Wang, X.; Liu, C.; Xie, Q.; Li, Z.; Zhang, Z. Enhanced photoelectrochemical and photocatalytic performance of TiO2 nanorod arrays/CdS quantum dots by coating TiO2 through atomic layer deposition. Nano Energy 2015, 11, 400–408. [Google Scholar] [CrossRef]
  16. Chen, P.; Wang, F.; Chen, Z.F.; Zhang, Q.; Su, Y.; Shen, L.; Yao, K.; Liu, Y.; Cai, Z.; Lv, W.; et al. Study on the photocatalytic mechanism and detoxicity of gemfibrozil by a sunlight-driven TiO2/carbon dots photocatalyst: The significant roles of reactive oxygen species. Appl. Catal. B Environ. 2017, 204, 250–259. [Google Scholar] [CrossRef]
  17. Su, Y.; Liu, G.; Zeng, C.; Lu, Y.; Luo, H.; Zhang, R. Carbon quantum dots-decorated TiO2/g-C3N4 film electrode as a photoanode with improved photoelectrocatalytic performance for 1,4-dioxane degradation. Chemosphere 2020, 251, 126381. [Google Scholar] [CrossRef]
  18. Li, P.; Zhao, G.; Li, M.; Cao, T.; Cui, X.; Li, D. Design and high efficient photoelectric-synergistic catalytic oxidation activity of 2D macroporous SNO2/1D TiO2 nanotubes. Appl. Catal. B Environ. 2012, 111–112, 578–585. [Google Scholar] [CrossRef]
  19. Xiao, Y.; Tian, G.; Chen, Y.; Zhang, X.; Fu, H.; Fu, H. Exceptional visible-light photoelectrocatalytic activity of In2O3/In2S3/CdS ternary stereoscopic porous heterostructure film for the degradation of persistent 4-fluoro-3-methylphenol. Appl. Catal. B Environ. 2018, 225, 477–486. [Google Scholar] [CrossRef]
  20. Garcia-Segura, S.; Brillas, E. Applied photoelectrocatalysis on the degradation of organic pollutants in wastewaters. J. Photoch. Photobio. C Photochem. Rev. 2017, 31, 1–35. [Google Scholar] [CrossRef]
  21. Daghrir, R.; Drogui, P.; Robert, D. Modified TiO2 for environmental photocatalytic applications: A review. Ind. Eng. Chem. Res. 2013, 52, 3581–3599. [Google Scholar] [CrossRef]
  22. Zhang, G.; Zhang, T.; Li, B.; Jiang, S.; Zhang, X.; Hai, L.; Chen, X.; Wu, W. An ingenious strategy of preparing TiO2/g-C3N4 heterojunction photocatalyst: In situ growth of TiO2 nanocrystals on g-C3N4 nanosheets via impregnation-calcination method. Appl. Surf. Sci. 2018, 433, 963–974. [Google Scholar] [CrossRef]
  23. Wang, J.; Zhang, W.-D. Modification of TiO2 nanorod arrays by graphite-like C3N4 with high visible light photoelectrochemical activity. Electrochim. Acta 2012, 71, 10–16. [Google Scholar] [CrossRef]
  24. Hu, Z.; Xie, X.; Li, S.; Song, M.; Liang, G.; Zhao, J.; Wang, Z. Rational construct CQDs/BiOCOOH/uCN photocatalyst with excellent photocatalytic performance for degradation of sulfathiazole. Chem. Eng. J. 2021, 404, 126541. [Google Scholar] [CrossRef]
  25. Fu, Y.; Zeng, G.; Lai, C.; Huang, D.; Qin, L.; Yi, H.; Liu, X.; Zhang, M.; Li, B.; Liu, S.; et al. Hybrid architectures based on noble metals and carbon-based dots nanomaterials: A review of recent progress in synthesis and applications. Chem. Eng. J. 2020, 399, 125743. [Google Scholar] [CrossRef]
  26. Kim, M.G.; Jo, W.-K. Visible-light-activated N-doped CQDs/g-C3N4/Bi2WO6 nanocomposites with different component arrangements for the promoted degradation of hazardous vapors. J. Mater. Sci. Technol. 2020, 40, 168–175. [Google Scholar] [CrossRef]
  27. Chen, Z.; Li, X.; Xu, Q.; Tao, Z.; Yao, F.; Huang, X.; Wu, Y.; Wang, D.; Jiang, P.; Yang, Q. Three-dimensional network space Ag3PO4/NP-CQDs/rGH for enhanced organic pollutant photodegradation: Synergetic photocatalysis activity/stability and effect of real water quality parameters. Chem. Eng. J. 2020, 390, 124454. [Google Scholar] [CrossRef]
  28. Shi, R.; Li, Z.; Yu, H.; Shang, L.; Zhou, C.; Waterhouse, G.I.N.; Wu, L.-Z.; Zhang, T. Effect of nitrogen doping level on the performance of N-doped carbon quantum dot/TiO2 composites for photocatalytic hydrogen evolution. ChemSusChem 2017, 10, 4650–4656. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Guo, Y.; Cao, F.; Li, Y. Solid phase synthesis of nitrogen and phosphor co-doped carbon quantum dots for sensing Fe3+ and the enhanced photocatalytic degradation of dyes. Sens. Actuat. B Chem. 2018, 255, 1105–1111. [Google Scholar] [CrossRef]
  30. Liu, B.; Aydil, E.S. Growth of oriented single-crystalline rutile TiO2 nanorods on transparent conducting substrates for dye-sensitized solar cells. J. Am. Chem. Soc. 2009, 131, 3985–3990. [Google Scholar] [CrossRef]
  31. Lin, S.; Lu, Y.; Ye, B.; Zeng, C.; Liu, G.; Li, J.; Luo, H.; Zhang, R. Pesticide wastewater treatment using the combination of the microbial electrolysis desalination and chemical-production cell and Fenton process. Front. Environ. Sci. Eng. 2020, 14, 12. [Google Scholar] [CrossRef]
  32. Lan, J.; Ren, Y.; Lu, Y.; Liu, G.; Luo, H.; Zhang, R. Combined microbial desalination and chemical-production cell with Fenton process for treatment of electroplating wastewater nanofiltration concentrate. Chem. Eng. J. 2019, 359, 1139–1149. [Google Scholar] [CrossRef]
  33. Dong, Y.; Wang, X.; Sun, H.; Zhang, H.; Zhao, X.; Wang, L. Construction of a 0D/3D AgI/MOF-808 photocatalyst with a one-photon excitation pathway for enhancing the degradation of tetracycline hydrochloride: Mechanism, degradation pathway and DFT calculations. Chem. Eng. J. 2023, 460, 141842. [Google Scholar] [CrossRef]
  34. Liu, W.; Li, Y.; Liu, F.; Jiang, W.; Zhang, D.; Liang, J. Visible-light-driven photocatalytic degradation of diclofenac by carbon quantum dots modified porous g-C3N4: Mechanisms, degradation pathway and DFT calculation. Water Res. 2019, 151, 8–19. [Google Scholar] [CrossRef]
  35. Guo, T.-L.; Li, J.-G.; Ping, D.-H.; Sun, X.; Sakka, Y. Controlled photocatalytic growth of ag nanocrystals on brookite and rutile and their sers performance. ACS Appl. Mater. Interfaces 2014, 6, 236–243. [Google Scholar] [CrossRef] [PubMed]
  36. Ge, L.; Han, C. Synthesis of MWNTs/g-C3N4 composite photocatalysts with efficient visible light photocatalytic hydrogen evolution activity. Appl. Catal. B Environ. 2012, 117–118, 268–274. [Google Scholar] [CrossRef]
  37. Xiang, Q.; Yu, J.; Jaroniec, M. Preparation and enhanced visible-light photocatalytic H2-production activity of graphene/C3N4 composites. J. Phys. Chem. C 2011, 115, 7355–7363. [Google Scholar] [CrossRef]
  38. Yang, D.; Li, L.; Xiao, G.; Zhang, S. Steering charge kinetics in metal-free g-C3N4/melem hybrid photocatalysts for highly efficient visible-light-driven hydrogen evolution. Appl. Surf. Sci. 2020, 510, 145345. [Google Scholar] [CrossRef]
  39. Hu, K.; Li, R.; Ye, C.; Wang, A.; Wei, W.; Hu, D.; Qiu, R.; Yan, K. Facile synthesis of Z-scheme composite of TiO2 nanorod/g-C3N4 nanosheet efficient for photocatalytic degradation of ciprofloxacin. J. Clean. Prod. 2020, 253, 120055. [Google Scholar] [CrossRef]
  40. Singh, V.K.; Singh, V.; Yadav, P.K.; Chandra, S.; Bano, D.; Kumar, V.; Koch, B.; Talat, M.; Hasan, S.H. Bright-blue-emission nitrogen and phosphorus-doped carbon quantum dots as a promising nanoprobe for detection of Cr(vi) and ascorbic acid in pure aqueous solution and in living cells. New J. Chem. 2018, 42, 12990–12997. [Google Scholar] [CrossRef]
  41. Yu, Z.; Li, Y.; Qu, J.; Zheng, R.; Cairney, J.M.; Zhang, J.; Zhu, M.; Khan, A.; Li, W. Enhanced photoelectrochemical water-splitting performance with a hierarchical heterostructure: Co3O4 nanodots anchored TiO2@P-C3N4 core-shell nanorod arrays. Chem. Eng. J. 2021, 404, 126458. [Google Scholar] [CrossRef]
  42. Dhandole, L.K.; Mahadik, M.A.; Chung, H.-S.; Chae, W.-S.; Cho, M.; Jang, J.S. CdIn2S4 chalcogenide/TiO2 nanorod heterostructured photoanode: An advanced material for photoelectrochemical applications. Appl. Surf. Sci. 2019, 490, 18–29. [Google Scholar] [CrossRef]
  43. Li, D.; Huang, J.; Li, R.; Chen, P.; Chen, D.; Cai, M.; Liu, H.; Feng, Y.; Lv, W.; Liu, G. Synthesis of a carbon dots modified g-C3N4/SnO2 Z-scheme photocatalyst with superior photocatalytic activity for PPCPs degradation under visible light irradiation. J. Hazard. Mater. 2021, 401, 123257. [Google Scholar] [CrossRef]
  44. Wang, F.; Chen, P.; Feng, Y.; Xie, Z.; Liu, Y.; Su, Y.; Zhang, Q.; Wang, Y.; Yao, K.; Lv, W.; et al. Facile synthesis of N-doped carbon dots/g-C3N4 photocatalyst with enhanced visible-light photocatalytic activity for the degradation of indomethacin. Appl. Catal. B Environ. 2017, 207, 103–113. [Google Scholar] [CrossRef]
  45. Hu, S.; Tian, R.; Dong, Y.; Yang, J.; Liu, J.; Chang, Q. Modulation and effects of surface groups on photoluminescence and photocatalytic activity of carbon dots. Nanoscale 2013, 5, 11665–11671. [Google Scholar] [CrossRef] [PubMed]
  46. Song, S.; Wu, K.; Wu, H.; Guo, J.; Zhang, L. Multi-shelled ZnO decorated with nitrogen and phosphorus co-doped carbon quantum dots: Synthesis and enhanced photodegradation activity of methylene blue in aqueous solutions. RSC Adv. 2019, 9, 7362–7374. [Google Scholar] [CrossRef] [PubMed]
  47. Kakavandi, B.; Zehtab, S.M.; Ahmadi, M.; Naderi, P.; Roccaro, J.; Bedia, M.; Hasham, F.R.; Rezaei, K. Spinel cobalt ferrite-based porous activated carbon in conjunction with UV light irradiation for boosting peroxymonosulfate oxidation of bisphenol A. J. Environ. Manag. 2023, 342, 118242. [Google Scholar] [CrossRef] [PubMed]
  48. Tian, N.; Giannakis, S.; Akbarzadeh, L.; Hasanvandian, F.; Dehghanifard, E.; Kakavandi, B. Improved catalytic performance of ZnO via coupling with CoFe2O4 and carbon nanotubes: A new, photocatalysis-mediated peroxymonosulfate activation system, applied towards Cefixime degradation. J. Environ. Manag. 2023, 329, 117022. [Google Scholar] [CrossRef]
  49. Wei, Z.; Liang, F.; Liu, Y.; Luo, W.; Wang, J.; Yao, W.; Zhu, Y. Photoelectrocatalytic degradation of phenol-containing wastewater by TiO2/g-C3N4 hybrid heterostructure thin film. Appl. Catal. B Environ. 2017, 201, 600–606. [Google Scholar] [CrossRef] [Green Version]
Figure 1. (a) SEM images of TiO2 Nas. Inset: cross-section view of TiO2 NAs; (b) SEM images of N, P-CQD/TCN NAs in top view and cross-section view; (c) TEM analysis of CQDs; (d) EDS spectrum of N, P-CQD/TCN NAs.
Figure 1. (a) SEM images of TiO2 Nas. Inset: cross-section view of TiO2 NAs; (b) SEM images of N, P-CQD/TCN NAs in top view and cross-section view; (c) TEM analysis of CQDs; (d) EDS spectrum of N, P-CQD/TCN NAs.
Water 15 02837 g001
Figure 2. (a) XPS survey spectra of TCN NAs and N, P-CQD/TCN NAs; high-resolution spectra of (b) C 1s, (c) N 1s, (d) O 1s; (e) Ti 2p of TCN NAs and N, P-CQD/TCN NAs; (f) high-resolution spectra of P 2p of N, P-CQD/TCN NAs.
Figure 2. (a) XPS survey spectra of TCN NAs and N, P-CQD/TCN NAs; high-resolution spectra of (b) C 1s, (c) N 1s, (d) O 1s; (e) Ti 2p of TCN NAs and N, P-CQD/TCN NAs; (f) high-resolution spectra of P 2p of N, P-CQD/TCN NAs.
Water 15 02837 g002
Figure 3. (a) UV–vis diffuse reflectance spectra of the different photoelectrodes. Inset: the corresponding Kubelka–Munk-transformed reflectance spectra; (b) up-converted PL spectra of N, P-CQDs. Inset: N, P-CQDs dispersed in water illuminated under daylight and UV light; (c) photoluminescence spectra of the different photoelectrodes; (d) fluorescence decay of the different photoelectrodes.
Figure 3. (a) UV–vis diffuse reflectance spectra of the different photoelectrodes. Inset: the corresponding Kubelka–Munk-transformed reflectance spectra; (b) up-converted PL spectra of N, P-CQDs. Inset: N, P-CQDs dispersed in water illuminated under daylight and UV light; (c) photoluminescence spectra of the different photoelectrodes; (d) fluorescence decay of the different photoelectrodes.
Water 15 02837 g003
Figure 4. The chronoamperometry plots of the different photoelectrodes under a bias voltage of 1.2 V.
Figure 4. The chronoamperometry plots of the different photoelectrodes under a bias voltage of 1.2 V.
Water 15 02837 g004
Figure 5. ESR spectra of DMPO-O2· radical adducts, DMPO-·OH radical adducts, and DMPO-1O2 radical adducts under visible light irradiation.
Figure 5. ESR spectra of DMPO-O2· radical adducts, DMPO-·OH radical adducts, and DMPO-1O2 radical adducts under visible light irradiation.
Water 15 02837 g005
Figure 6. (a) Degradation of 1,4-D in the PEC cell with different anodes under visible light irradiation and bias voltage of 1.2 V, and (b) concentrations of oxalic acid, formic acid, and acetic acid in the PEC cell with the initial concentration of 100 mg/L 1,4-D.
Figure 6. (a) Degradation of 1,4-D in the PEC cell with different anodes under visible light irradiation and bias voltage of 1.2 V, and (b) concentrations of oxalic acid, formic acid, and acetic acid in the PEC cell with the initial concentration of 100 mg/L 1,4-D.
Water 15 02837 g006
Figure 7. The proposed degradation pathway of 1,4-D in the PEC cell with the N, P-CQD/TCN NA anodes.
Figure 7. The proposed degradation pathway of 1,4-D in the PEC cell with the N, P-CQD/TCN NA anodes.
Water 15 02837 g007
Figure 8. Change in COD in the pesticide wastewater treatment within 5-repeated-cycle operation in the PEC cell with the N, P-CQD/TCN NA anodes under visible light irradiation (initial conditions: pH 6.8 ± 0.5, COD 350 ± 30 mg/L, conductivity 5.3 ± 0.2 mS/cm, and a bias voltage of 1.2 V).
Figure 8. Change in COD in the pesticide wastewater treatment within 5-repeated-cycle operation in the PEC cell with the N, P-CQD/TCN NA anodes under visible light irradiation (initial conditions: pH 6.8 ± 0.5, COD 350 ± 30 mg/L, conductivity 5.3 ± 0.2 mS/cm, and a bias voltage of 1.2 V).
Water 15 02837 g008
Figure 9. The spectra of EEM measurements on the pesticide wastewater treated in the PEC cell with the N, P-CQD/TCN NA anodes within (a) 0 min, (b) 20 min, (c) 40 min, and (d) 60 min (initial conditions: pH 6.8 ± 0.5, COD 350 ± 30 mg/L, conductivity 5.3 ± 0.2 mS/cm, and a bias voltage of 1.2 V).
Figure 9. The spectra of EEM measurements on the pesticide wastewater treated in the PEC cell with the N, P-CQD/TCN NA anodes within (a) 0 min, (b) 20 min, (c) 40 min, and (d) 60 min (initial conditions: pH 6.8 ± 0.5, COD 350 ± 30 mg/L, conductivity 5.3 ± 0.2 mS/cm, and a bias voltage of 1.2 V).
Water 15 02837 g009
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Su, Y.; Ye, Y.; Lin, S.; Lu, Y.; Luo, H.; Liu, G. Surface Modification of TiO2/g-C3N4 Electrode with N, P Codoped CQDs for Photoelectrocatalytic Degradation of 1,4-Dioxane. Water 2023, 15, 2837. https://doi.org/10.3390/w15152837

AMA Style

Su Y, Ye Y, Lin S, Lu Y, Luo H, Liu G. Surface Modification of TiO2/g-C3N4 Electrode with N, P Codoped CQDs for Photoelectrocatalytic Degradation of 1,4-Dioxane. Water. 2023; 15(15):2837. https://doi.org/10.3390/w15152837

Chicago/Turabian Style

Su, Yuehan, Yongbei Ye, Songwei Lin, Yaobin Lu, Haiping Luo, and Guangli Liu. 2023. "Surface Modification of TiO2/g-C3N4 Electrode with N, P Codoped CQDs for Photoelectrocatalytic Degradation of 1,4-Dioxane" Water 15, no. 15: 2837. https://doi.org/10.3390/w15152837

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