Next Article in Journal
Photoreactive Carbon Dots Modified g-C3N4 for Effective Photooxidation of Bisphenol-A under Visible Light Irradiation
Previous Article in Journal
Construction of Novel Z-Scheme g-C3N4/AgBr-Ag Composite for Efficient Photocatalytic Degradation of Organic Pollutants under Visible Light
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of Cu-Doped TiO2 Nanocatalyst for the Enhanced Photocatalytic Degradation and Mineralization of Gabapentin under UVA/LED Irradiation: Characterization and Photocatalytic Activity

by
Roghieh Ahmadiasl
1,
Gholamreza Moussavi
2,*,
Sakine Shekoohiyan
2 and
Fatemeh Razavian
1
1
Department of Environmental Science and Engineering, West Tehran Branch, Islamic Azad University, Tehran P.O. Box 1468763785, Iran
2
Department of Environmental Health Engineering, Tarbiat Modares University, Tehran P.O. Box 331-14115, Iran
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(11), 1310; https://doi.org/10.3390/catal12111310
Submission received: 18 September 2022 / Revised: 4 October 2022 / Accepted: 21 October 2022 / Published: 25 October 2022
(This article belongs to the Section Environmental Catalysis)

Abstract

:
Light-harvesting of titanium oxide (TiO2) was enhanced by copper (Cu) doping, and its performance was evaluated by gabapentin (GBP) degradation under UVA-LED irradiation. The morphology and structure of TiO2 and Cu-TiO2 were characterized using XRD, FTIR, FE-SEM, EDX, TEM, PL, DRS, and BET analysis. The complete degradation of 10 mg/L GBP was obtained in the developed photocatalytic process under the optimal conditions: catalyst loading, 0.4 g/L; pH solution, 8; and reaction time, 20 min. The reactive species trapping was studied to identify the degradation mechanism in this system. Among the water matrix experiments, phosphate (PO43−) anion indicated an inverse effect in increasing efficiency. Finally, the main intermediates generation during the GBP degradation was investigated based on LC-MS analysis, and a decomposition pathway was proposed. Accordingly, doping TiO2 with Cu resulted in the development of a UVA-activated photocatalyst for efficiently degrading and mineralizing GBP as a model of a pharmaceutical compound.

1. Introduction

Nowadays, emerging pollutants are widely consumed, but despite their beneficial effects, their adverse efficacy on human health and the environment cannot be denied [1]. The pharmaceutical industries produce a large number of medicinal materials for the prevention and therapy of human illnesses, and thus water resources are polluted by industrial wastewater. Additionally, the disposal of these therapeutic materials from the human body into the sewage will be another way of entering these micro-pollutants into aqueous environments [2,3]. Gabapentin (GBP) is widely used as an analgesic for kidney disorders and for treating neuropathic pain and epilepsy [4,5]. GBP has been detected at concentrations of up to 10 ηg/L, along with other drugs in raw wastewater, but the noteworthy point is that its destruction does not take place entirely, and no-metabolite structure can be dispersed in various solutions [6,7]. Additionally, this drug is ranked as an emerging micro-pollutant with adverse effects on human health [6,8]. Hence, the main issue is developing an effective method for treating contaminated water with different pharmaceutical pollutants. The selection of the heterogeneous photocatalytic oxidation (PCO) method based on the advanced oxidation technique is an attractive technology that has been applied in the degradation of various recalcitrant organic pollutants [9,10]. In the PCO process, photo-generated intermediates, such as electron/hole pairs, can enhance photoreaction under light irradiation. Many PCO processes have been studied to find effective photocatalysts for the decomposition of a wide range of contaminants [11]. Among the various semiconductors as photocatalysts, TiO2 has been most highlighted due to its remarkable properties, including great oxidation potential, chemical durability, nontoxicity, a high proportion of surface to volume, and economical trait [12,13]. Nevertheless, the high recombination rate of photo-induced charge carriers and poor ionic and electrical transmission are the significant intrinsic disadvantages of unmodified TiO2 [14]. As a solution, doping semiconductors with transition metal and nonmetal or coupling with other materials have been proposed in previous studies [2,15,16,17,18]. Doping TiO2 with metals has shown high photocatalytic activity that enhanced the separation capability of e/h+ pairs and reduced the band gap energy compared to pure TiO2 [14,19]. For instance, the application of copper (Cu2O, Cu2+, CuO, etc.) as a dopant to improve the photocatalytic activity of TiO2 has been efficiently practiced for the degradation of contaminants and is applicable due to its cost-effectiveness and abundance [20,21]. It has been reported that when a semiconductor is doped with copper, it can create an extra energy level, leading to increasing photocatalytic activity [22,23]. Moreover, Vaiano et al. reported that when Cu was introduced to ZnO, the band gap energy decreased from 3.2 to 2.92 eV, resulting in decreasing the recombination rate of e-/h+ pairs, thus the photocatalytic performance improved [24].
In this context, the Cu-TiO2 material was prepared and characterized as an improved form of TiO2 catalyst for the degradation and mineralization of Gabapentin under UVA irradiation emitted from LEDs. LEDs are environmentally benign, have a long life span, and provide flexibility in designing the photoreactor; therefore, they are considered an appropriate source of UV photons to drive the PCO [25]. Accordingly, the effect of the main experimental parameters of pH, catalyst concentration, water anions, radical scavengers, and reaction time was investigated on the degradation of Gabapentin. A pathway was also proposed for the degradation of Gabapentin in the developed photocatalytic process operated under optimal conditions.

2. Results and Discussion

2.1. Characterization of As-Prepared TiO2 and Cu-TiO2

The crystal structure of plain TiO2 and Cu-TiO2 products are shown in Figure 1a using XRD analysis. The characteristic peaks of (101), (004), (200), (105), (211), (204), (116), (220), and (215) planes are identified in the XRD pattern of both TiO2 and Cu-TiO2 products that can be attributed to the anatase phase of TiO2 based on (JCPDS Card No. 73–1764). Additionally, the diffraction peaks of (101), (110), (111), and (220) planes are observed in the XRD pattern of the Cu-TiO2 materials, which corresponds to the rutile phase of the TiO2 structure (JCPDS Card No. 78–1510). On the other hand, Cu-TiO2 with mixed phases (anatase and rutile) was prepared through the designed synthesizing method. Any characteristic peak is observed in the XRD pattern of Cu-TiO2 product for CuO, suggesting that Cu might be doped in the lattice of the TiO2 [26]. After Cu doping, the peak broadening in the XRD pattern qualitatively described the change in particle size and lattice expansion [27]. The crystallite size of TiO2 and Cu-TiO2 was estimated from the XRD pattern achieved using the Scherrer equation. The crystallite size of pure TiO2 and Cu-TiO2 was 29.3 and 45.6 nm. This matter indicates that introducing Cu into TiO2 increased the crystallite size of Cu-TiO2. The literature review [27,28] reported that the crystallite size of Cu-TiO2 was slightly increased due to the presence of Cu2+ and CuO species. The different ionic radii of Cu and Ti ions caused lattice distortion and strain field, changing the size of particles of Cu-TiO2.
The FTIR spectra of TiO2 and Cu-TiO2 are displayed in Figure 1b. The FTIR spectrum of the catalyst shows absorption peaks between 400 and 600 cm−1, which is ascribed to Ti-O-Ti stretching. In addition, other peaks of about 1623 and 2347 cm−1 are related to C = C and C = O functional groups, respectively. The medium peaks at about 3485 and 3500 cm−1 confirm the presence of OH bands in the sample structure due to surface adsorb water and hydroxyl agent. The vibration bands are observed between 1000 to 1250 cm−1 after Cu-doped TiO2 indicated lattice vibration of TiO2 [29,30,31,32].
The surface morphology of the as-prepared TiO2 and Cu-TiO2 materials was investigated by FESEM, and the elemental distribution was evaluated based on EDX and mapping analyses; the results are presented in Figure 2a–f for Cu-TiO2 and in Figure S1 for TiO2. FESEM images confirmed the almost spherical shape of the Cu-TiO2 product with uniform size distribution. Therefore, doping with Cu affected the morphology of TiO2. Norris et al. [33] reported that the photocatalyst’s morphology, size, and electronic structure could be modified by doping with copper. The EDX analysis demonstrated the presence of considerable percentages of Ti, O, and Cu in the structure of the as-prepared Cu-TiO2 catalyst. In addition, the distribution of elements with colored dots in the mapping images (Figure 2d–f) confirms the presence of Cu in the Cu-TiO2 product. The EDX mapping shows that the Cu-TiO2 seems to be homogenous. To further examine the morphology of the prepared materials, the TEM analysis was conducted, and images of TiO2 and Cu-TiO2 can be observed in Figure 2c. The average particle sizes of TiO2 and Cu-TiO2 determined using Digimizer software based on the TEM images were around 35 and 47 nm, respectively. The results confirmed that doping TiO2 with Cu increases the size of the particles. Tasbihi et al. [34] reported that Cu-TiO2 was fully crystallized at lower doping concentrations, but at higher concentrations of Cu (15 wt.%), the prepared catalyst had amorphous phases.
BET analysis estimated the surface areas of prepared TiO2 and Cu-TiO2 to be 52.1 and 9.9 m2/g, respectively (Figure 3). It can be related to changes in the morphology and crystalline structure of the TiO2 after doping with Cu [30]. This result agreed with Tasbihi et al. [34] and Bensouici et al. [22], who reported a decrease in the surface area of TiO2 after doping with Cu. From the pore size distribution curve, Figure 3a,b (inset), and IUPAK categorization, both catalysts’ isotherms were observed to be type IV and mesoporous structures [35].
The optical properties of pure TiO2 and Cu-TiO2 products were measured using the UV-Vis DRS analyses, and the results are given in Figure 4. Based on Figure 4a, the role of Cu doping on the spectra profile of Cu-TiO2 is visible by shifting the absorption edge of TiO2 to the higher wavelengths upon doping with Cu. It can be related to the surface plasmon resonance properties of copper [36], resulting in increased light absorption potential of the product and therefore in the accelerated generation of electron-hole pairs under light irradiation. Moreover, the band gap energy (Eg) of TiO2 and Cu-TiO2 was determined using the Tauc/David–Matt method [37]. According to Figure 4b, the Eg values of TiO2 and Cu-TiO2 were estimated to be about 3.2 and 2.9 eV, respectively. The results indicate that doping TiO2 with Cu could decrease its band gap energy, reconfirming the higher light absorption potential and thus the photocatalytic activity in the Cu-TiO2 material compared to that in the plain TiO2. Tasbihi et al. [34] observed that pure TiO2 exhibited Eg of 3.05 eV, while the Eg of Cu/TiO2 was lower due to the presence of Cu2+ ions in the structure of TiO2 lattice. Additionally, they [34] reported that low Cu concentration (0.2 wt.%) could not decrease the Eg of Cu-TiO2. Sahu and Biswas [27] reported that the Eg of TiO2 decreased from 3.31 to 2.51 eV with increasing the Cu concentration at the highest dopant concentration (15 wt.%). This change in Eg was related to incorporating Cu2+ ions into the TiO2 structure and forming CuO as a layer on the TiO2 particle surface [27]. The change in the optical absorption of Cu-TiO2 was attributed to the defect centers created by the replacement of Ti4+ by Cu2+ atoms in the TiO2 crystal lattice. A previous study [38] showed that doping TiO2 with aliovalent ions could change the local lattice symmetry and defect characteristics, changing the absorption properties. In Cu-TiO2, when Cu+2 ions are located inside or on the surface sites of TiO2, a rearrangement of the neighbor atoms to compensate for the charge deficiency is caused by lattice deformation. The lattice deformation influences the electronic structure of Cu-TiO2 and shifts the Eg. Therefore, these Cu-doped materials can be applied for diverse photocatalytic applications, which have been reported in numerous other studies.
The PL spectra of the pure TiO2 and Cu-TiO2 samples were investigated to comprehend the separation of e/h+ pairs and the charge–carrier transfer efficiency. Figure 5 shows a significant peak at around 388 nm in pure TiO2, which is assigned to the band-to-band recombination because it is near-band-edge luminescence. The peak in PL spectra of Cu-TiO2 is observed at 376 nm, but with a drop in intensity. This intensity reduction refers to the inhibition of the recombination of the photo-generated electron from the conduction band (CB) to the valence band (VB) of TiO2. The higher PL peak intensity in pure TiO2 refers to the reduction of the lifetime of the electron-hole pairs due to the fast recombination rate. Reda et al. [39] reported the highest reduction in PL intensity of doped-TiO2 with Cu and N, revealing the examined samples’ efficacy in reducing photo-generated electron-hole recombination compared with pure TiO2. The intensity of the light absorption of TiO2 was also decreased upon doping with Cu, implying that the rate of electron-hole recombination decreased in the doped product. It might result in the formation of higher reactive radical species and, thus, higher degradation performance in the developed photocatalytic process [40]. Therefore, it can be concluded that doping with copper significantly modified the properties of TiO2, which increased its photoactivity [41].

2.2. The Photocatalytic Activity of Cu-Doped TiO2

2.2.1. Effect of Initial pH Solution

The solution pH is an essential factor effective in GBP removal in the photocatalytic process. Moreover, the solution pH affects photocatalytic activity through two mechanisms. Firstly, the solution pH influences the extent of pollutant ionization and, therefore, the ionizable compounds’ adsorption on the photocatalyst’s surface. On the other hand, the ionizable compound’s absorption rate could control the photocatalytic degradation amount [42]. The removal of pollutant compounds affected by pH in photocatalytic processes depends on the solution pH, compound pKa, and catalyst pHpzc. The interaction between these factors promotes or inhibits the removal rate of the desired compound. Additionally, the literature review showed that change in the solution pH had altered the adsorption capacity of pollutants on the photocatalyst surfaces [43]. The experiment was evaluated under various pH levels (4–9), while the GBP concentration, catalyst dosage, and reaction time were considered to be 10 mg/L, 0.5 g/L, and 20 min, respectively. As shown in Figure 6a, the complete photo-activity was obtained at pH 8 and 9 while it was reduced at acidic pH. Additionally, it is interesting that in all pH values, the GPB degradation was higher than GBP adsorption. GBP is a compound with an isoelectric point at pH 7.14 and has two pKa values of 3.68 for the carboxylic acid and 10.7 for the primary amine [44]. On the other hand, the pHpzc value of prepared Cu-TiO2 was estimated at about 7.32, which depicts that the catalyst surface had a negative charge higher than pHpzc. As a result, at pH 8 and 9, electrostatic attraction between GBP molecules (positive) and Cu-TiO2 photocatalyst (negative) enhanced the destruction efficiency. Therefore, pH 8 was selected for the subsequent reaction. Nevertheless, at a pH value lower than 7, electrostatic repulsion is boosted due to the positive charge of GBP and as-made catalyst, resulting in low photocatalytic activity [45]. Ahmad and Yasin [46] investigated the Cu/TiO2/bentonite composite for the photocatalytic degradation of deltamethrin. The results show that a maximum degradation rate was observed at pH 12. They concluded that modifying pH is an effective strategy to improve photocatalytic performance. Sharma et al. [47] observed that the degradation rate of tetracycline with Cu2O-TiO2 nanotube was complete under acidic, neutral, and alkaline conditions, and it is related to the adsorption of pollutants on the surface catalyst.

2.2.2. Effect of Cu-TiO2 Concentration

Catalyst concentration directly influences the degradation efficiency. Figure 7 depicts GBP degradation under reaction conditions of pH = 8, reaction time = 20 min and pollutant concentration 10 mg/L with different catalyst dosages (0.1–0.5 g/L) in Cu-TiO2/UVA process. Without the Cu-TiO2, the removal efficiency was about 21%, but after the addition of the catalyst, the degradation rate was enhanced. In general, the degradation efficiency of GBP is enhanced with an increased concentration of Cu-TiO2. As can be seen, removal efficiencies were obtained to be 49.8, 56.2, 69.3, 100, and 100% for 0.1, 0.2, 0.3, 0.4, and 0.5 g/L of the as-prepared catalyst, respectively. These results prove that the degradation yield enhanced with increasing Cu-TiO2 loading to 0.4 g/L. Therefore, the optimal concentration of Cu-TiO2 for the following experiments was chosen as 0.4 g/L.
The number of active sites on the catalyst surface increased, which, in turn, boosted the formation of main reactive species, such as electron-hole pairs or radical hydroxyl during the photo process. Additionally, catalyst concentrations higher than 0.5 g/L tend to accumulate in a system where the penetration of light decreases, resulting in no change in the removal efficiency [48]. It is worth saying that adsorption was evaluated along with GBP photo-degradation, and increasing catalyst loading reduced the adsorption rate, indicating the role of light on the Cu-TiO2 activation [49]. Reda et al. [38] reported that 0.14 g/L of Cu-TiO2 was the optimal concentration for dye removal. By increasing the amount of photocatalyst up to the optimal level, the reaction rate due to more active sites was increased. Additionally, when the Cu-TiO2 dosage was increased beyond the optimal concentration, the aggregation of the suspended particles took place, and this phenomenon reduced the amount of light reaching the active sites of the catalyst, and subsequently, the rate of degradation decreased.

2.2.3. Catalytic Activity and Mechanism of GBP Degradation in the Cu-TiO2/UVA Process

The catalytic activity of TiO2 and Cu-TiO2 materials were studied after selecting the optimal conditions, and the results are presented in Figure 8a. As-made Cu-TiO2 catalyst exhibited poor adsorption toward GBP molecules under dark conditions. When no catalyst was introduced into the reactor, the GBP removal was increased by 21% under light irradiation. In addition, the pure TiO2 indicated poor degradation efficiency of about 41% under the UVA-LED system compared to Cu-TiO2. By utilizing the as-prepared Cu-TiO2 catalyst, the degradation of GBP molecules significantly improved and complete removal was achieved. In addition, the kinetics of the removal of GBP in the selected processes was evaluated based on the pseudo-first-order (PFO) reaction model [50], and the plots are
l n C 0 C t = k o b s t
in which C0 and Ct are the initial and residual concentrations of GBP at time t (min) of the reaction, and kobs is the observed PFO rate constant values for each initial concentration that was obtained from the slope of plotting ln (C0/Ct) vs. reaction time.
The PFO rate constants of GBP removal in the UVA, TiO2/UVA, and Cu-TiO2/UVA processes were 0.004, 0.029, and 0.122 min−1, respectively. It is seen that the degradation by direct photolysis was inefficient for the removal of GBP under the selected conditions. In addition, the removal rate of GBP in the photocatalytic process with Cu-TiO2 was over four times that with plain TiO2, implying that doping TiO2 considerably improved its photocatalytic activity for the removal of GBP. The improvement of photocatalytic activity of TiO2 upon doping with Cu can be related to the decrease in band gap and the recombination rate of the electron/hole. Indeed, Cu dopant as a free electron trap could reduce the recombination rate of electron-hole pairs, improve the light-harvesting capacity, and accelerate the photo-activity of modified TiO2, which is consistent with DRS and PL analysis. Additionally, in the presence of an appropriate light source, electrons are excited from the VB to the CB of TiO2 to create VB holes (positive charge carriers) and CB electrons (negative charge carriers). By scavenging the oxidizing species, these electron-hole pairs produce different reactive oxygen (ROS), further degrading the GBP on the surface of the Cu-TiO2 [47].
To understand the mechanisms involved in the removal of GBP in the Cu-TiO2/UVA process, the GPB removal was compared in the absence and presence of several scavengers, and the related results are reported in Figure 9. The radical scavengers were tert-butanol (TBA) as the scavenger of HO , p-benzoquinone (pBQ) as the scavenger of HO and O 2 , sodium azide as the scavenger of HO and 1O2, and oxalate as the scavenger of h+ [34]. As seen in Figure 9, the removal of GBP was less affected in the presence of TBA, pBQ, and sodium azide, but it was strongly suppressed in the presence of oxalate. It indicates that photo-generated holes played the main role in GBP oxidation, which could be quenched by water molecules or oxygen atoms to enhance the formation of hydroxyl and superoxide radicals, respectively [51].
Regarding the FTIR curve given above, the presence of the functional group (TiO2-O-Cu) confirmed the CuO formation that we expected: the CuO-TiO2 heterojunction led to the best performance of modified TiO2 [13,51]. Therefore, the low band gap of CuO (1.7 eV) compared to TiO2 (3.2 eV) is a reason for driving photo-generated electrons from the CB of metallic copper to TiO2, and at the same time, photo-induced holes jumped from the VB TiO2 to VB CuO [13]. On the other hand, the synergistic role between anatase–rutile heterophase in the as-prepared catalyst induced the separation of electron–hole pairs accelerated for the greater transfer of photo agents in the CB and VB of Cu-TiO2 nanoparticles (Equations (2)–(5)) [30].
Photocatalyst   ( Cu-TiO 2 ) + h ʋ     Photocatalyst   ( e C B + h V B + )
e C B + O 2   O 2
h V B + + H 2 O   H O
GBP + h V B + / H O / O 2     degradation   products
Sharma et al. [47] reported that the main reactive species involved in tetracycline degradation in the photocatalytic process using Cu2O-TiO2 was the superoxide radical (O2•−), followed by HO and valence band holes (VBh+). The effect of radical scavengers on carbamazepine (CBZ) degradation was investigated in the CuWO4-TiO2 process [52]. The results showed that the mechanism for the degradation of CBZ was the surface charge process driven by HO , O2•−, the generation of VBh+, and CBe−. Under an appropriate light source, the excitation process can generate holes in the VB and electrons in the CB accompanied by redox reactions such as Cu+2/Cu+. The generated electrons of the CB of TiO2 drift into the CB of CuWO4. This phenomenon suppresses electron-hole recombination and enhances CBZ degradation. Besides electron-hole generation, ROS production, including O2•− and HO , becomes available to drive the photodegradation activity [52].

2.2.4. Effect of Water Matrix on GBP Photo-Degradation

To investigate the as-prepared catalyst in practical applications, the presence of common water anions, such as chloride (Cl), nitrate (NO3), carbonate (CO32−), phosphate (PO43−), and sulfate (SO42−), were evaluated on the GBP decomposition in the Cu-TiO2/UVA process. Most studies have reported that anions have an inhibitory role in the photocatalytic processes due to (1) the competition between the anions and the target pollutant toward the available active sites on the catalyst surface; (2) the radicals’ scavenging properties of the anions, which caused the formation of ionic radicals such as Cl•−, NO3•−, and HCO3•− with less degradation potential; and (3) the hole scavenging proprieties of anions [52]. Therefore, the experiment was conducted with no anion sample. As can be seen in Figure 10, the addition of PO43− and SO42− decreased the efficiency to 49 and 91%, respectively, under conditions of pH = 8, catalyst concentration = 0.4 g/L, GBP concentration = 10 mg/L, and anions = 1 mM, during 20 min. Based on the results shown in Figure 10, a negligible negative impact was observed after adding Cl, NO3, and CO32− compared with the control sample on the GBP removal in the Cu-TiO2/UVA photocatalytic process, indicating no considerable interaction between these anions, and the reactive species contributed in removing GBP under the selected conditions. PO43− and SO42− at the higher levels had a limiting effect on GBP removal in the Cu-TiO2/UVA photocatalytic process. It can be related to the interaction of these anions with the surface of the catalyst and occupying the active sites of the catalyst, resulting in a decrease in the GBP degradation rate (Equations (6)–(8)) [53,54]. The introduction of SO42− decreased CBZ degradation, and a possible reason was the adherence of SO42− to the surface of TiO2 via Van der Waals forces and hydrogen bonds and perhaps the displacement of the surface hydroxyl group of TiO2 via the ligand exchange mechanism. Additionally, SO42−, due to the double charge, has a high adsorption ability onto the surface of the catalyst and a high scavenging affinity to HO , and these reasons decreased the GBP degradation rate as compared to the other anions [52].
P O 4 3 + H 2 O   H P O 4 2 + O H
H P O 4 2 + O H   O H + H 2 + H P O 4
S O 4 2 + h V B +   S O 4

2.2.5. Mineralization and Pathway Study of GBP Degradation in the Cu-TiO2/UVA Process

To evaluate the mineralization efficiency of GBP molecules in the Cu-TiO2/UVA photocatalytic process, the TOC concentration was measured when the process was operated under optimum conditions. According to the obtained result (Figure 11a), the removal of TOC increased with the reaction time and reached 63% at the reaction time of 60 min. In addition, the PFO removal rate of TOC under the selected condition was determined to be 0.016 min−1 (Figure 11b). These results clearly indicate that the developed process efficiently mineralized GBP. The high degree of TOC removal at 60 min can be demonstrated that the GBP molecules were changed to simple materials, which might be completed by continuing the reaction time. The mineralization of adsorbed GBP on Cu-TiO2 is possible either by direct oxidation of GBP on the catalyst surface or by produced hydroxyl radicals (Equation (9)).
H O + R H R + H 2 O C O 2 + H 2 O + m i n e r a l   a c i d s
RH and R are organic substances and organic radicals, respectively. Evgenidou et al. [55] examined the mineralization of a mixture of eight antibiotics using Cu-TiO2 photocatalysts. Against complete degradation of antibiotic compounds, the reduction in TOC after 30 min was not more than 20%, indicating the formation of intermediate products. There is a need for a long time in photocatalyst processes to achieve complete mineralization, approving the formation of recalcitrant intermediates. In their study, defluorination, the release of nitrite and nitrate ions, and desulfurization are the main steps in the mineralization of antibiotic compounds.
The difference between GBP degradation and mineralization rate proved the formation of some intermediates identified in the reaction solution. For this purpose, the degradation intermediates of the GBP were conducted by LC-MS spectra under the optimal condition at 60 min. Accordingly, four transformation intermediates were measured based on the m/z values, and the degradation pathway was proposed in Figure 12. Gabapentin is unstable and can be spontaneously converted to gabapentin lactam (m/z 154) during dehydration [56]. Then, the deoxygenation of the C-O bond resulted in forming an intermediate with m/z 128. The recorded fragment of an intermediate, such as cyclohexane (m/z 84) and isopropyl radical (m/z 43), is related to the presence of reactive species ( h VB + / HO / O 2 ) that destroyed the N-C, C-C, and N-H bonds to low molar mass. Finally, simple products, such as CO 2 and NH 4 + , were observed in this reaction, confirming the effective mineralization of GBP in the photocatalytic process with an as-made Cu-TiO2 catalyst. Dal Bello et al. [57] investigated GBP degradation in the presence of a TiO2 photocatalyst. The degradation of GBP (m/z 172) consisted of several stages. Firstly, the loss of a water molecule at m/z 154; secondly, simultaneous loss of NH3 (m/z 137) and carbon monoxide (m/z 126); and finally, losses of water and carbon monoxide.

3. Materials and Methods

3.1. Chemicals

Gabapentin (molecular formula: C9H17NO2) was obtained from Local Pharmaceutical. Titanium isopropoxide (C12H28O4Ti), isopropyl alcohol (CH3O8), acetic acid (CH3COOH), copper acetate Cu(CH3COO)2, potassium dihydrogen phosphate (KH2PO4), dipotassium dihydrogen phosphate (K2HPO4), HPLC grade water and acetonitrile were purchased from the local market. In this work, all solutions were supplied with deionized water.

3.2. Synthesis and Characterization of Cu-TiO2 Nanoparticles

First, 10 mL of isopropyl alcohol and 3 mL of acetic acid were mixed in a glass container, and then 2.5 mL of titanium isopropoxide as a precursor was added dropwise into the above solution. The reaction was carried out at an ambient temperature and stirred at 80 rpm for 30 min. Then 2 mL of distilled water was added drop by drop to the mixture under the hood. Finally, after 3 to 5 min mixing, the obtained gel was dried at 100 °C and calcined at 500 °C for 2 h under an air atmosphere. To prepare the Cu-doped TiO2, 0.319 g copper acetate was dissolved in 10 mL of distilled water; the obtained solution was added dropwise to the titanium dioxide gel under mixing, and the prepared gel was calcined under conditions similar to the plain TiO2.
The crystalline phase of the prepared catalyst was detected based on the X-ray diffraction (XRD) analysis in the 2θ scanning range from 10 to 80°, and Cu as anode material with ʎ radiation of 1.54 Å. The surface functional groups were identified by a Fourier transform infrared spectroscopy (FTIR) in the frequency range of 400–4000 cm−1. The surface structure and elemental distribution of as-the prepared catalyst were determined by field emission scanning electron microscopy (FE-SEM) coupled with energy-dispersive X-ray spectroscopy (EDX). The actual size of Cu-TiO2 was obtained with transmission electron microscopy (TEM) using Philips CM30. The specific surface area was characterized by Brunauer–Emmett–Teller (BET) in the condition of nitrogen adsorption/desorption at 77 °k. Energy band gap and photoelectronic quality were evaluated by differential reflectance spectroscopy (DRS) and photoluminescence (PL) methods, respectively. Finally, the identification of the intermediate through the gabapentin photodegradation was conducted by liquid chromatography-mass spectrometry (LC-MS) analysis.

3.3. Experimental and Analytical Procedures

The photodegradation efficiency of the as-prepared catalyst for gabapentin removal was examined under UVA-LED irradiation in a batch-quartz photoreactor with an inner diameter and volume of 10 cm and 50 mL, respectively. A setup equipped with 18 LEDs (3 W, SUN LED, Seoul Semiconductor, Korea) and emitted UVA photons (365 nm) was applied as a light source. Generally, a given range of catalysts was dispersed into the reactor with GBP solution (10 mg/L in all tests) at the target pH, stirred at an interval determined times for degradation of Gabapentin by Cu-TiO2 under UVA-LED irradiation. Then 5 mL of sample was filtered through a 0.2-micron syringe filter and the residual concentration of the contaminant was monitored by high-performance liquid chromatography (HPLC, Eclipse Plus C18 column; 3.5 µm, 4.6 × 100 mm, Agilent) with UV detection at 210 nm. The mobile phase consisting of phosphate buffer (20 mM, pH = 6.2) and 25% acetonitrile, was conducted at a 1 mL/min flow rate [5]. The removal efficiency of the GBP was calculated using the following equation:
Removal   efficiency   ( % ) = C 0 C t C 0 × 100
which C0 and Ct are the initial and final concentrations of GBP, respectively. All the experiments were conducted in duplicate, and the mean of the results was reported. The standard deviation was within ±5%.

4. Conclusions

Doping Cu-TiO2 was successfully performed by the sol-gel approach for GBP decomposition under the UVA-LED system. Based on the DRS technique, the band gap of the as-made catalyst was calculated at about 1.48 eV, and the pL results indicated a low recombination rate of charge carriers compared to pure TiO2 nanoparticles. The Cu-TiO2 particles/UVA-LED system indicated high GBP removal efficiency with a reaction rate constant of about 0.122 min−1 under optimal conditions. In addition to investigating the common parameters (catalyst concentration, solution pH, and reaction time), the effect of scavengers and water matrix were also checked on the trend of GBP removal under the photocatalytic process. The order of the negative effect of the water matrix on GBP degradation was PO43− > SO42−, and other anions did not affect the reaction. Scavenger experiments revealed that the h+ and HO had a significant role in the reaction mechanism of the Cu-TiO2 particles/UVA-LED system. Finally, the mineralization rate by TOC measurement was obtained at about 63% at 60 min reaction time, and the identified intermediates suggested that the pollutant converted to simple compounds. As a result, the as-made Cu-TiO2 nanocomposite showed high photocatalytic properties under UVA irradiation; thus, the Cu-TiO2/UVA process was confirmed under the bench-scale condition as a promising and efficient method for the removal of pharmaceutical compounds from the contaminated water and wastewater.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal12111310/s1, Figure S1: FE-SEM (a), EDX (b), Elemental mapping (c,d), and TEM (e) of TiO2.

Author Contributions

R.A.: Investigation, Writing—original draft, Writing—review & editing, Visualization. G.M., S.S. and F.R.: Conceptualization, Methodology, Resources, Writing—original draft, Writing—review & editing, Supervision, Project administration. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data is available on request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Choi, Y.; Lee, J.H.; Kim, K.; Mun, H.; Park, N.; Jeon, J. Identification, quantification, and prioritization of new emerging pollutants in domestic and industrial effluents, Korea: Application of LC-HRMS based suspect and non-target screening. J. Hazard. Mater. 2021, 402, 123706. [Google Scholar] [CrossRef] [PubMed]
  2. Cheshmeh Khavar, A.H.; Moussavi, G.; Mahjoub, A.R. The preparation of TiO2@rGO nanocomposite efficiently activated with UVA/LED and H2O2 for high rate oxidation of acetaminophen: Catalyst characterization and acetaminophen degradation and mineralization. Appl. Surf. Sci. 2018, 440, 963–973. [Google Scholar] [CrossRef]
  3. Akbari, S.; Moussavi, G.; Giannakis, S. Efficient photocatalytic degradation of ciprofloxacin under UVA-LED, using S, N-doped MgO nanoparticles: Synthesis, parametrization and mechanistic interpretation. J. Mol. Liq. 2021, 324, 114831. [Google Scholar] [CrossRef]
  4. Goswami, A.; Jiang, J.Q.; Petri, M. Non-parametric regression analysis of diuron and gabapentin degradation in lake Constance water by ozonation and their toxicity assessment. Water 2019, 11, 852. [Google Scholar] [CrossRef] [Green Version]
  5. Ragham, P.K.; Chandrasekhar, K.B. Development and validation of a stability-indicating RP-HPL C-CAD method for gabapentin and its related impurities in presence of degradation products. J. Pharm. Biomed. Anal. 2016, 125, 122–129. [Google Scholar] [CrossRef] [PubMed]
  6. Goswami, A.; Jiang, J.Q.; Petri, M. Treatability of five micro-pollutants using modified Fenton reaction catalysed by zero-valent iron powder (Fe (0)). J. Environ. Chem. Eng. 2021, 9, 105393. [Google Scholar] [CrossRef]
  7. Vymazal, J.; Březinová, T.D.; Koželuh, M.; Kule, L. Occurrence and removal of pharmaceuticals in four full-scale constructed wetlands in the Czech Republic–the first year of monitoring. Ecol. Eng. 2017, 98, 354–364. [Google Scholar] [CrossRef]
  8. Fundneider, T.; Alonso, V.A.; Wick, A.; Albrecht, D.; Lackner, S. Implications of biological activated carbon filters for micropollutant removal in wastewater treatment. Water Res. 2021, 189, 116588. [Google Scholar] [CrossRef]
  9. Kohantorabi, M.; Moussavi, G.; Mohammadi, S.; Oulego, P.; Giannakis, S. Photocatalytic activation of peroxymonosulfate (PMS) by novel mesoporous Ag/ZnO@ NiFe2O4 nanorods, inducing radical-mediated acetaminophen degradation under UVA irradiation. Chemosphere 2021, 277, 130271. [Google Scholar] [CrossRef]
  10. Chen, L.; Tang, J.; Song, L.N.; Chen, P.; He, J.; Au, C.T.; Yin, S.F. Heterogeneous photocatalysis for selective oxidation of alcohols and hydrocarbons. Appl. Catal. B Environ. 2019, 242, 379–388. [Google Scholar] [CrossRef]
  11. Kohantorabi, M.; Moussavi, G.; Oulego, P.; Giannakis, S. Radical-based degradation of sulfamethoxazole via UVA/PMS-assisted photocatalysis, driven by magnetically separable Fe3O4@ CeO2@ BiOI nanospheres. Sep. Purif. Technol. 2021, 267, 118665. [Google Scholar] [CrossRef]
  12. Thambiliyagodage, C.; Mirihana, S. Photocatalytic activity of Fe and Cu co-doped TiO2 nanoparticles under visible light. J. Sol-Gel Sci. Technol. 2021, 99, 109–121. [Google Scholar] [CrossRef]
  13. Thambiliyagodage, C.; Usgodaarachchi, L. Photocatalytic activity of N, Fe and Cu co-doped TiO2 nanoparticles under sunlight. Curr. Res. Green Sustain. Chem. 2021, 4, 100186. [Google Scholar] [CrossRef]
  14. Cheshmeh Khavar, A.H.; Moussavi, G.; Mahjoub, A.R.; Satari, M. Facile preparation of multi-doped TiO2/rGO cross-linked 3D aerogel (GaNF@TGA) nancomposite as an efficient visible-light activated catalyst for photocatalytic oxidation and detoxification of atrazine. Sol. Energy 2018, 173, 848–860. [Google Scholar] [CrossRef]
  15. Tang, T.; Yin, Z.; Chen, J.; Zhang, S.; Sheng, W.; Wei, W.; Xiao, Y.; Shi, Q.; Cao, S. Novel pn heterojunction Bi2O3/Ti3+-TiO2 photocatalyst enables the complete removal of tetracyclines under visible light. Chem. Eng. J. 2021, 417, 128058. [Google Scholar] [CrossRef]
  16. Gao, D.; Liu, W.; Xu, Y.; Wang, P.; Fan, J.; Yu, H. Core-shell Ag@ Ni cocatalyst on the TiO2 photocatalyst: One-step photoinduced deposition and its improved H2-evolution activity. Appl. Catal. B Environ. 2020, 260, 118190. [Google Scholar] [CrossRef]
  17. Adnan, M.; Muhd Julkapli, N.; Amir, M.; Maamor, A. Effect on different TiO2 photocatalyst supports on photodecolorization of synthetic dyes: A review. Int. J. Environ. Sci. Technol. 2019, 16, 547–566. [Google Scholar] [CrossRef]
  18. Sun, Q.; Lv, K.; Zhang, Z.; Li, M.; Li, B. Effect of contact interface between TiO2 and g-C3N4 on the photoreactivity of g-C3N4/TiO2 photocatalyst:(0 0 1) vs (1 0 1) facets of TiO2. Appl. Catal. B Environ. 2015, 164, 420–427. [Google Scholar]
  19. Jaiswal, R.; Bharambe, J.; Patel, N.; Dashora, A.; Kothari, D.; Miotello, A. Copper and Nitrogen co-doped TiO2 photocatalyst with enhanced optical absorption and catalytic activity. Appl. Catal. B Environ. 2015, 168, 333–341. [Google Scholar] [CrossRef]
  20. Cheng, C.; Fang, W.H.; Long, R.; Prezhdo, O.V. Water splitting with a single-atom Cu/TiO2 photocatalyst: Atomistic origin of high efficiency and proposed enhancement by spin selection. J. Am. Chem. Sos. 2021, 1, 550–559. [Google Scholar] [CrossRef]
  21. Čižmar, T.; Panžić, I.; Capan, I.; Gajović, A. Nanostructured TiO2 photocatalyst modified with Cu for improved imidacloprid degradation. Appl. Surf. Sci. 2021, 569, 151026. [Google Scholar] [CrossRef]
  22. Bensouici, F.; Bououdina, M.; Dakhel, A.; Tala-Ighil, R.; Tounane, M.; Iratni, A.; Souier, T.; Liu, S.; Cai, W. Optical, structural and photocatalysis properties of Cu-doped TiO2 thin films. Appl. Surf. Sci. 2017, 395, 110–116. [Google Scholar] [CrossRef]
  23. Raguram, T.; Rajni, K. Synthesis and analysing the structural, optical, morphological, photocatalytic and magnetic properties of TiO2 and doped (Ni and Cu) TiO2 nanoparticles by sol–gel technique. Appl. Phys. A 2019, 125, 288. [Google Scholar] [CrossRef]
  24. Vaiano, V.; Iervolino, G.; Rizzo, L. Cu-doped ZnO as efficient photocatalyst for the oxidation of arsenite to arsenate under visible light. Appl. Catal. B Environ. 2018, 238, 471–479. [Google Scholar] [CrossRef]
  25. Hossaini, H.; Moussavi, G.; Farrokhi, M. The investigation of the LED-activated FeFNS-TiO2 nanocatalyst for photocatalytic degradation and mineralization of organophosphate pesticides in water. Water Res. 2014, 59, 130–144. [Google Scholar] [CrossRef] [PubMed]
  26. Lin, C.-J.; Yang, W.-T. Ordered mesostructured Cu-doped TiO2 spheres as active visible-light-driven photocatalysts for degradation of paracetamol. Chem. Eng. J. 2014, 237, 131–137. [Google Scholar] [CrossRef]
  27. Sahu, M.; Biswas, P. Single-step processing of copper-doped titania nanomaterials in a flame aerosol reactor. Nanoscale Res. Lett. 2011, 6, 441. [Google Scholar] [CrossRef] [Green Version]
  28. Colón, G.; Maicu, M.; Hidalgo, M.S.; Navío, J. Cu-doped TiO2 systems with improved photocatalytic activity. Appl. Catal. B Environ. 2006, 67, 41–51. [Google Scholar] [CrossRef]
  29. Sagadevan, S.; Vennila, S.; Singh, P.; Lett, J.A.; Oh, W.C.; Paiman, S.; Mohammad, F.; Al-Lohedan, H.A.; Fatimal, I.; Shahid, M.M.; et al. Exploration of the antibacterial capacity and ethanol sensing ability of Cu-TiO2 nanoparticles. J. Exp. Nanosci. 2020, 15, 337–349. [Google Scholar] [CrossRef]
  30. Ullattil, S.G.; Zavašnik, J.; Maver, K.; Finšgar, M.; Novak Tušar, N.; Pintar, A. Defective Grey TiO2 with Minuscule Anatase–Rutile Heterophase Junctions for Hydroxyl Radicals Formation in a Visible Light-Triggered Photocatalysis. Catalysts 2021, 11, 1500. [Google Scholar] [CrossRef]
  31. Din, M.I.; Arshad, F.; Rani, A.; Aihetasham, A.; Mukhtar, M.; Mehmood, H. Single step green synthesis of stable copper oxide nanoparticles as efficient photo catalyst material. Biomed. Mater. 2017, 9, 41–48. [Google Scholar]
  32. Rajamannan, B.; Mugundan, S.; Viruthagiri, G.; Praveen, P.; Shanmugam, N. Linear and nonlinear optical studies of bare and copper doped TiO2 nanoparticles via sol gel technique. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2014, 118, 651–656. [Google Scholar] [CrossRef] [PubMed]
  33. Norris, D.J.; Efros, A.L.; Erwin, S.C. Doped nanocrystals. Science 2008, 319, 1776–1779. [Google Scholar] [CrossRef] [PubMed]
  34. Tasbihi, M.; Kočí, K.; Troppová, I.; Edelmannová, M.; Reli, M.; Čapek, L.; Schomäcker, R. Photocatalytic reduction of carbon dioxide over Cu/TiO2 photocatalysts. Environ. Sci. Pollut. Res. 2018, 25, 34903–34911. [Google Scholar] [CrossRef] [PubMed]
  35. Akbari, S.; Moussavi, G.; Decker, J.; Marin, M.L.; Bosca, F.; Giannakis, S. Superior visible light-mediated catalytic activity of a novel N-doped, Fe3O4-incorporating MgO nanosheet in presence of PMS: Imidacloprid degradation and implications on simultaneous bacterial inactivation. Appl. Catal. B Environ. 2022, 317, 121732. [Google Scholar] [CrossRef]
  36. Mathew, S.; Ganguly, P.; Rhatigan, S.; Kumaravel, V.; Byrne, C.; Hinder, S.J.; Bartllet, J.; Nolan, M.; Pillai, S. Cu-doped TiO2: Visible light assisted photocatalytic antimicrobial activity. Appl. Sci. 2018, 8, 2067. [Google Scholar] [CrossRef] [Green Version]
  37. Kohantorabi, M.; Giannakis, S.; Moussavi, G.; Bensimon, M.; Gholami, M.R.; Pulgarin, C. An innovative, highly stable Ag/ZIF-67@ GO nanocomposite with exceptional peroxymonosulfate (PMS) activation efficacy, for the destruction of chemical and microbiological contaminants under visible light. J. Hazard. Mater. 2021, 413, 125308. [Google Scholar] [CrossRef]
  38. Li, L.; Liu, J.; Su, Y.; Li, G.; Chen, X.; Qiu, X.; Yan, T. Surface doping for photocatalytic purposes: Relations between particle size, surface modifications, and photoactivity of SnO2: Zn2+ nanocrystals. Nanotechnology 2009, 20, 155706. [Google Scholar] [CrossRef]
  39. Reda, S.; Khairy, M.; Mousa, M. Photocatalytic activity of nitrogen and copper doped TiO2 nanoparticles prepared by microwave-assisted sol-gel process. Arab. J. Chem. 2020, 13, 86–95. [Google Scholar] [CrossRef]
  40. Etacheri, V.; Di Valentin, C.; Schneider, J.; Bahnemann, D.; Pillai, S.C. Visible-light activation of TiO2 photocatalysts: Advances in theory and experiments. J. Photochem. Photobiol. C Photochem. Rev. 2015, 25, 1–29. [Google Scholar] [CrossRef] [Green Version]
  41. Park, H.S.; Kim, D.H.; Kim, S.J.; Lee, K.S. The photocatalytic activity of 2.5 wt% Cu-doped TiO2 nano powders synthesized by mechanical alloying. J. Alloys Compd. 2006, 415, 51–55. [Google Scholar] [CrossRef]
  42. Apiwong-ngarm, K.; Pongwan, P.; Inceesungvorn, B.; Phanichphant, S.; Wetchakun, K.; Wetchakun, N. Photocatalytic activities of Fe–Cu/TiO2 on the mineralization of oxalic acid and formic acid under visible light irradiation. Powder Technol. 2014, 266, 447–455. [Google Scholar] [CrossRef]
  43. Varma, K.S.; Tayade, R.J.; Shah, K.J.; Joshi, P.A.; Shukla, A.D.; Gandhi, V.G. Photocatalytic degradation of pharmaceutical and pesticide compounds (PPCs) using doped TiO2 nanomaterials: A review. Water Energy Nexus 2020, 3, 46–61. [Google Scholar] [CrossRef]
  44. Ciavarella, A.B.; Gupta, A.; Sayeed, V.A.; Khan, M.A.; Faustino, P.J. Development and application of a validated HPLC method for the determination of gabapentin and its major degradation impurity in drug products. J. Pharm. Biomed. Anal. 2007, 43, 1647–1653. [Google Scholar] [CrossRef] [PubMed]
  45. Chen, M.; Yao, J.; Huang, Y.; Gong, H.; Chu, W. Enhanced photocatalytic degradation of ciprofloxacin over Bi2O3/(BiO)2CO3 heterojunctions: Efficiency, kinetics, pathways, mechanisms and toxicity evaluation. Chem. Eng. J. 2018, 334, 453–461. [Google Scholar] [CrossRef]
  46. Ahmad, S.; Yasin, A. Photocatalytic degradation of deltamethrin by using Cu/TiO2/bentonite composite. Arab. J. Chem. 2020, 13, 8481–8488. [Google Scholar] [CrossRef]
  47. Sharma, M.; Mandal, M.K.; Pandey, S.; Kumar, R.; Dubey, K.K. Visible-Light-Driven Photocatalytic Degradation of Tetracycline Using Heterostructured Cu2O–TiO2 Nanotubes, Kinetics, and Toxicity Evaluation of Degraded Products on Cell Lines. ACS Omega 2022, 7, 33572–33586. [Google Scholar] [CrossRef]
  48. Golshan, M.; Kakavandi, B.; Ahmadi, M.; Azizi, M. Photocatalytic activation of peroxymonosulfate by TiO2 anchored on cupper ferrite (TiO2@CuFe2O4) into 2, 4-D degradation: Process feasibility, mechanism and pathway. J. Hazard. Mater. 2018, 359, 325–337. [Google Scholar] [CrossRef]
  49. Sarafraz, M.; Sadeghi, M.; Yazdanbakhsh, A.; Amini, M.M.; Sadani, M.; Eslami, A. Enhanced photocatalytic degradation of ciprofloxacin by black Ti3+/N-TiO2 under visible LED light irradiation: Kinetic, energy consumption, degradation pathway, and toxicity assessment. Process Saf. Environ. Prot. 2020, 137, 261–272. [Google Scholar] [CrossRef]
  50. Akbari, S.; Ghanbari, F.; Moradi, M. Bisphenol A degradation in aqueous solutions by electrogenerated ferrous ion activated ozone, hydrogen peroxide and persulfate: Applying low current density for oxidation mechanism. Chem. Eng. J. 2016, 294, 298–307. [Google Scholar] [CrossRef]
  51. Yang, X.J.; Shu, W.; Sun, H.m.; Wang, X.B.; Lian, J.S. Preparation and photocatalytic performance of Cu-doped TiO2 nanoparticles. Trans. Nonferrous Met. Soc. China 2015, 25, 504–509. [Google Scholar] [CrossRef]
  52. Anucha, C.B.; Altin, I.; Bacaksız, E.; Kucukomeroglu, T.; Belay, M.H.; Stathopoulos, V.N. Enhanced photocatalytic activity of CuWO4 doped TiO2 photocatalyst towards carbamazepine removal under UV irradiation. Separations 2021, 8, 25. [Google Scholar] [CrossRef]
  53. Wang, Q.; Rao, P.; Li, G.; Dong, L.; Zhang, X.; Shao, Y.; Gao, N.; Chu, W.; Xu, B.; An, N.; et al. Degradation of imidacloprid by UV-activated persulfate and peroxymonosulfate processes: Kinetics, impact of key factors and degradation pathway. Ecotoxicol. Environ. Saf. 2020, 187, 109779. [Google Scholar] [CrossRef]
  54. Kohantorabi, M.; Moussavi, G.; Giannakis, S. A review of the innovations in metal-and carbon-based catalysts explored for heterogeneous peroxymonosulfate (PMS) activation, with focus on radical vs. non-radical degradation pathways of organic contaminants. Chem. Eng. J. 2021, 411, 127957. [Google Scholar] [CrossRef]
  55. Evgenidou, E.; Chatzisalata, Z.; Tsevis, A.; Bourikas, K.; Torounidou, P.; Sergelidis, D.; Koltsakidou, A.; Lambropoulou, D.A. Photocatalytic degradation of a mixture of eight antibiotics using Cu-modified TiO2 photocatalysts: Kinetics, mineralization, antimicrobial activity elimination and disinfection. J. Environ. Chem. Eng. 2021, 9, 105295. [Google Scholar] [CrossRef]
  56. Herrmann, M.; Menz, J.; Olsson, O.; Kümmerer, K. Identification of phototransformation products of the antiepileptic drug gabapentin: Biodegradability and initial assessment of toxicity. Water Res. 2015, 85, 11–21. [Google Scholar] [CrossRef] [PubMed]
  57. Dal Bello, F.; Medana, C.; Zorzi, M.; Kuck, B.; Fabbri, D.; Calza, P. Liquid chromatography/mass spectrometry analytical determination of gabapentin transformation products by heterogeneous photocatalysis and environmental evaluation. Rapid Commun. Mass Spectrom. 2020, 34, e8925. [Google Scholar] [CrossRef]
Figure 1. XRD images (a) and FTIR patterns (b) of pure TiO2 and as-made Cu-TiO2.
Figure 1. XRD images (a) and FTIR patterns (b) of pure TiO2 and as-made Cu-TiO2.
Catalysts 12 01310 g001
Figure 2. FE-SEM (a), EDX (b), TEM (c), and elemental mapping (df) of Cu-TiO2.
Figure 2. FE-SEM (a), EDX (b), TEM (c), and elemental mapping (df) of Cu-TiO2.
Catalysts 12 01310 g002
Figure 3. BET analysis of TiO2 (a) and Cu-TiO2 (b) and the BJH plots (inset).
Figure 3. BET analysis of TiO2 (a) and Cu-TiO2 (b) and the BJH plots (inset).
Catalysts 12 01310 g003
Figure 4. DRS (a) and Tauc’s plots analysis (b) of as-made plain TiO2 and Cu-TiO2 materials.
Figure 4. DRS (a) and Tauc’s plots analysis (b) of as-made plain TiO2 and Cu-TiO2 materials.
Catalysts 12 01310 g004aCatalysts 12 01310 g004b
Figure 5. Photoluminescence of as-made plain TiO2 and Cu-TiO2 materials.
Figure 5. Photoluminescence of as-made plain TiO2 and Cu-TiO2 materials.
Catalysts 12 01310 g005
Figure 6. Effect of solution pH on GBP degradation in the Cu-TiO2-UVA system (GBP = 10 mg/L, catalyst = 0.5 g/L, reaction time = 20 min).
Figure 6. Effect of solution pH on GBP degradation in the Cu-TiO2-UVA system (GBP = 10 mg/L, catalyst = 0.5 g/L, reaction time = 20 min).
Catalysts 12 01310 g006
Figure 7. Effect of catalyst concentration on GBP degradation in the Cu-TiO2-UVA system (GBP = 10 mg/L, pH = 8, reaction time = 20 min).
Figure 7. Effect of catalyst concentration on GBP degradation in the Cu-TiO2-UVA system (GBP = 10 mg/L, pH = 8, reaction time = 20 min).
Catalysts 12 01310 g007
Figure 8. The degradation percentage (a) and rate (b) of GBP by various processes (GBP = 10 mg/L, pH = 8, catalyst = 0.4 g/L, reaction time = 0–20 min).
Figure 8. The degradation percentage (a) and rate (b) of GBP by various processes (GBP = 10 mg/L, pH = 8, catalyst = 0.4 g/L, reaction time = 0–20 min).
Catalysts 12 01310 g008
Figure 9. Effect of scavengers on GBP degradation in the Cu-TiO2/UVA process (GBP = 10 mg/L, pH = 8, catalyst = 0.4 g/L, scavenger (when added) = 1 M, reaction time = 20 min).
Figure 9. Effect of scavengers on GBP degradation in the Cu-TiO2/UVA process (GBP = 10 mg/L, pH = 8, catalyst = 0.4 g/L, scavenger (when added) = 1 M, reaction time = 20 min).
Catalysts 12 01310 g009
Figure 10. Effect of different water anions on GBP degradation in the Cu-TiO2/UVA process (GBP = 10 mg/L, pH = 8, catalyst = 0.4 g/L, anion = 1 mM, reaction time = 20 min).
Figure 10. Effect of different water anions on GBP degradation in the Cu-TiO2/UVA process (GBP = 10 mg/L, pH = 8, catalyst = 0.4 g/L, anion = 1 mM, reaction time = 20 min).
Catalysts 12 01310 g010
Figure 11. Mineralization percentage (a) and rate (b) of GBP in the Cu-TiO2/VUA process (GBP = 10 mg/L, pH = 8, catalyst = 0.4 g/L).
Figure 11. Mineralization percentage (a) and rate (b) of GBP in the Cu-TiO2/VUA process (GBP = 10 mg/L, pH = 8, catalyst = 0.4 g/L).
Catalysts 12 01310 g011
Figure 12. Proposed pathway for GBP degradation in the Cu-TiO2/UVA process (GPB = 10 mg/L, pH = 8, catalyst = 0.4 g/L, reaction time = 60 min).
Figure 12. Proposed pathway for GBP degradation in the Cu-TiO2/UVA process (GPB = 10 mg/L, pH = 8, catalyst = 0.4 g/L, reaction time = 60 min).
Catalysts 12 01310 g012
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ahmadiasl, R.; Moussavi, G.; Shekoohiyan, S.; Razavian, F. Synthesis of Cu-Doped TiO2 Nanocatalyst for the Enhanced Photocatalytic Degradation and Mineralization of Gabapentin under UVA/LED Irradiation: Characterization and Photocatalytic Activity. Catalysts 2022, 12, 1310. https://doi.org/10.3390/catal12111310

AMA Style

Ahmadiasl R, Moussavi G, Shekoohiyan S, Razavian F. Synthesis of Cu-Doped TiO2 Nanocatalyst for the Enhanced Photocatalytic Degradation and Mineralization of Gabapentin under UVA/LED Irradiation: Characterization and Photocatalytic Activity. Catalysts. 2022; 12(11):1310. https://doi.org/10.3390/catal12111310

Chicago/Turabian Style

Ahmadiasl, Roghieh, Gholamreza Moussavi, Sakine Shekoohiyan, and Fatemeh Razavian. 2022. "Synthesis of Cu-Doped TiO2 Nanocatalyst for the Enhanced Photocatalytic Degradation and Mineralization of Gabapentin under UVA/LED Irradiation: Characterization and Photocatalytic Activity" Catalysts 12, no. 11: 1310. https://doi.org/10.3390/catal12111310

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