Next Article in Journal
Preparation of Metal Oxides Containing ppm Levels of Pd as Catalysts for the Reduction of Nitroarene and Evaluation of Their Catalytic Activity by the Fluorescence-Based High-Throughput Screening Method
Next Article in Special Issue
Biomorphic Fibrous TiO2 Photocatalyst Obtained by Hydrothermal Impregnation of Short Flax Fibers with Titanium Polyhydroxocomplexes
Previous Article in Journal
Ruthenium Catalysts Templated on Mesoporous MCM-41 Type Silica and Natural Clay Nanotubes for Hydrogenation of Benzene to Cyclohexane
Previous Article in Special Issue
Shape-Selective Mesoscale Nanoarchitectures: Preparation and Photocatalytic Performance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Photocatalytic Oxidative Degradation of Carbamazepine by TiO2 Irradiated by UV Light Emitting Diode

1
Institute of Innovational Education Research, School of Transportation and Environment, Shenzhen Institute of Information Technology, Shenzhen 518172, China
2
Shenzhen Water Affairs (Group) Co. Ltd., Shenzhen 518033, China
3
Department of Building and Environmental Engineering, Shenzhen Polytechnic, Shenzhen 518055, China
4
State Key Laboratory of Separation Membranes and Membrane Processes, School of Environmental and Chemical Engineering, Tiangong University, Tianjin 300387, China
*
Authors to whom correspondence should be addressed.
Catalysts 2020, 10(5), 540; https://doi.org/10.3390/catal10050540
Submission received: 5 April 2020 / Revised: 8 May 2020 / Accepted: 9 May 2020 / Published: 13 May 2020
(This article belongs to the Special Issue Recent Advances in TiO2 Photocatalysts)

Abstract

:
Here, ultraviolet light-emitting diodes (UV-LED) combined with TiO2 was used to investigate the feasibility of carbamazepine (CBZ) degradation. The effects of various factors, like crystal form of the catalyst (anatase, rutile, and mixed phase), mass concentration of TiO2, wavelength and irradiation intensity of the UV-LED light source, pH of the reaction system, and coexisting anions and cations, on the photocatalytic degradation of CBZ were studied. The mixed-phase (2.8 g/L) showed the best degradation efficiency at 365 nm among three kinds of TiO2, wherein CBZ (21.1 µM) was completely oxidized within 1 h. The results of batch experiments showed that: (i) CBZ degradation efficiency under UV-LED light at 365 nm was higher than 275 nm, due to stronger penetrability of 365 nm light in solution. (ii) The degradation efficiency increased with increase in irradiation intensity and pH, whereas it decreased with increase in initial CBZ concentration. (iii) The optimal amount of mixed-phase TiO2 catalyst was 2.8 g/L and excessive catalyst decreased the rate. (iv) The co-existence of CO32−, HCO3, and Fe3+ ions in water significantly accelerated the degradation rate of photocatalytic CBZ, whereas Cu2+ ions strongly inhibited the degradation process of CBZ. ·OH was found to be the main active species in the UV-LED photocatalytic degradation of CBZ. UV-LED is more environmentally friendly, energy efficient, and safer, whereas commercial TiO2 is economical and readily available. Therefore, this study provides a practically viable reference method for the degradation of pharmaceuticals and personal care products (PPCPs).

1. Introduction

In recent years, pharmaceuticals and personal care products (PPCPs) have gained more attention, due to the manufacturing and application of pesticides and medicines for treating humans and livestock [1]. PPCPs and its metabolites can be found in water bodies during regular inspections in aquatic environments conducted world-wide. Since, most PPCPs have very stable and complex molecular structures they cannot be easily captured and degraded by organisms. This increases the environmental pollution and is potentially harmful to human health [2,3]. At the same time, the presence of trace amounts of chemical pollutants in the aquatic environment may lead to the antibiotic-resistance in organisms and the generation of resistant genes, posing a serious threat to the entire ecosystem [4,5,6]. Therefore, PPCP must be completely eliminated from the natural water bodies to reduce its threat to all biomes.
Carbamazepine (CBZ, C15H12N2O), a dibenzazepine derivative, has molecular structure similar to that of tricyclic antidepressants. At present, it is widely used to treat trigeminal neuralgia, epilepsy, and mental illness [7]. Table 1 enlists the molecular structure and solubility of CBZ. Due to its stable structure, the removal efficiency of CBZ is generally less than 10% by traditional sewage treatment methods like coagulation/flocculation, chlorination, and microbial degradation [8]. The rest of it (major portion) is discharged into different aqueous environments either as a raw drug or its degradation intermediates. It is now commonly found in sewage, surface water, groundwater, and even drinking water. Its residues can trigger homologous estrogen activity and cause reproductive toxicity [9,10]. Hence, there is an urgent need to develop an environmentally friendly, college-level method to degrade or remove CBZ from the water bodies.
Various advanced oxidation process (AOP) methods are remediation techniques, effective in the removal of PPCPs. They include treatment methods using UV radiations in combination with H2O2, ozone, or titanium dioxide [11,12,13,14]. Most AOP methods can effectively remove CBZ residues (93–100%) from water [15,16,17,18]. One kind of AOP, semiconductor based photocatalytic oxidation technology, has the characteristics of environmental protection and shows great application prospects in the treatment of organic pollutants. Hydroxyl radicals have extremely strong oxidizing ability and, similar to most of the macromolecular organic pollutants having complex structures, can be oxidized and decomposed into smaller molecular weight and non-toxic organic compounds, or even completely decomposed into inorganic substances. Nano-semiconductor TiO2 particles are endowed with characteristics like green environmental protection, high chemical stability, large specific surface area, and low cost. They have become increasingly popular as photocatalytic semiconductor materials, because of their ability to photocatalytically decompose water [19]. Many previous studies by scholars have shown that TiO2 photocatalysis by ultraviolet irradiation is an efficient and environment-friendly technique for removing organic antibiotic compounds from the aquatic environment. However, the main limitation is the source of UV light in the practical applications of photocatalysis. The most commonly used source, a low and medium pressure mercury lamp, has many short-comings, such as fragility, bulkiness, serious heat radiations, high energy consumption, short life (500–2000 h), and so on [20]. In addition, mercury ultraviolet lamps emit light of only a fixed wavelength in a very narrow range and also have a risk of pollution due to mercury leakage [21].
In recent years, UV-LED technology has been widely explored for its applications, due to its unique advantages such as environmental protection, long service life, and short warm-up time. LED lights require low energy input and each LED lamp bead is very small, making it flexible and easy to install and transport [22,23]. In addition, the emission wavelength of UV-LEDs can be easily adjusted according to the material composition of the lamp beads. Today, UV-LED is a more powerful narrow-band light emitting device with shorter relaxation times [24]. Compared with traditional ultraviolet light, ultraviolet LED technology is more suitable for wastewater treatment, s as it is easy to regulate. For example, Cai et al. used UV-A type LED light to irradiate TiO2 nanoparticles to photo-catalyze antibiotics in wastewater, and their research showed that continuous UV-LED/TiO2 photocatalytic technology could be applied to 100 ppb concentration of sulfamethoxazole/methoxybenzyl above 320 nm. The degradation rate of pyridine was >90% [25]. Liang et al. found that an UV-LED/TiO2 photocatalytic reaction system, using periodic lighting control, was extremely effective for the degradation and removal of PPCP. At the same time, the use of porous titanium dioxide substrate significantly reduced the energy requirements of the reaction system [26]. It was also found that under certain experimental conditions, the UV-LED/TiO2 process could effectively decompose methylene blue (MB) in water. This indicated that UV-LED is a radiation source that effectively stimulates the photocatalytic action of TiO2, and this technology has a high potential for photodegradation [27]. In all, the UV-LED/TiO2 system is a valuable technology, especially for wastewater treatment containing PPCP.
At present, there is only limited research on photocatalytic degradation of CBZ, using a more energy-saving and environmentally safe UV-LED source and an economical and readily available commercial TiO2, which inspired us to explore the feasibility of this study. Herein, single-factor experiments were conducted to study the effects of crystal forms of nano-TiO2, the amount of TiO2, the wavelength of UV-LED and its irradiation intensity, CBZ concentration in water, pH of aqueous solution, and coexisting anions and cations common in water, on the degradation efficiency of CBZ. Our research demonstrated the feasibility of this UV LED combined with TiO2 photodegradation technology and explored the best operating conditions in actual water treatment project.

2. Results and Discussion

2.1. Characterization of Three Different Crystal Forms of TiO2 Particles

Scanning electron microscopy (SEM) was employed to characterize the morphologies of the three different phases of TiO2. Figure 1a–c showed the particle sizes of TiO2 nanoparticles, viz. anatase, rutile, and mixed phases, to be 20–100, 50–100, and 20–100 nm, respectively. The UV-Vis diffuse reflectance spectra (UV-Vis DRS, Figure 1d) showed that all the three phases of TiO2 nanoparticles could absorb light below 420 nm. The N2 adsorption-desorption tests were used to calculate the surface areas of the three crystal forms of TiO2. According to Table 2 and Figure 1e, the surface areas of the anatase, rutile, and mixed phases were 78.7, 32.2, and 102.6 m2/g, respectively, by Brunner−Emmet−Teller (BET). X-ray diffraction patterns of three forms of TiO2 nanoparticles can be seen in Figure 1f. Anatase and rutile matched with their respective standard cards (JCPDS 89-4921) and (JCPDS 21-1276). The equation, X = 1/(1 + 0.8IA/IR), was used to calculate the rutile content in the mixed phase, where X was the mass fraction of the rutile phase. IA and IR in the equation were the 2θ values of the characteristic peaks of anatase and rutile forms, which were 25.3° and 27.4°, respectively. Herein, the ratio of anatase to rutile forms in the mixed crystal TiO2 form was about 83:17, which was similar to the common commercial P25 (80:20).

2.2. Control Experiments of CBZ Degradation

The control experiments for CBZ degradation were also conducted. Figure 2a shows that the CBZ degradation failed to proceed in the absence of any of the following: illumination or the mixed-phase TiO2 photocatalyst. Owing to the adsorption of CBZ on the surface of TiO2, its concentration decreased by 5%, after reaching adsorption equilibrium. The degradation rate (ηt) of CBZ was calculated according to Equation (1) (Where ηt is the conversion rate (percentage) at a given time t, C0 is the initial concentration of the reagent, and Ct is the residual concentration at a given time t). The concentration of CBZ decreased sharply under irradiation of UV-LED light for 60 min in the presence of mixed-phase TiO2, wherein it was almost completely degraded (CBZ degradation rate reached 98.5%). Since, the valence band potential (3.1 eV vs. NHE) was more positive than H2O/·OH (2.74 eV vs. NHE), the TiO2 photocatalyst excited by UV-LED, could produce highly oxidized photogenerated holes and hydroxyl radicals that could oxidize the CBZ molecules into H2O, CO2, and other non-toxic small molecules. It is speculated that the CBZ photocatalytic degradation step may be a common photocatalytic reaction, as shown in reaction formulae 1–4. In addition, the CBZ degradation process was consistent with the pseudo-first-order kinetic models with a degree of fit R2 > 0.99 (Figure 2b). These results indicated that the photocatalytic method of UV-LED combined with TiO2 photocatalyst is a technology that can be implemented to effectively remove CBZ molecules from an aqueous environment.
ηt = (C0Ct)/C0·100%
TiO 2 + UV - LED e CB + h VB +
H 2 O / OH + h VB + OH + H +
CBZ + h VB + / OH H 2 O + CO 2 + products

2.3. Effect of UV-LED Wavelength and Its Intensity

Figure 3 shows the effects of two different UV-LED emission wavelengths on CBZ degradation. The CBZ photocatalytic degradation rate (98.5%) using 365 nm irradiation was significantly higher than that of 275 nm (36.78%). The degradation rate constant at 365 nm irradiation was 0.067 min−1, whereas at 275 nm it was only 0.007 min−1. Generally, the energy of the shorter-wavelength ultraviolet rays is higher than longer-wavelength, however, its penetrating power is weaker [27,28] Therefore, we speculate that the difference in degradation effects could be due to the differences in penetrating power. Compared with 275 nm irradiations, the use of 365 nm UV-LED source is a more cost-effective, energy-saving, and efficient method for photocatalytic oxidation of CBZ.
The radiation intensity of UV-LED is one of the important factors affecting the photocatalytic reaction, and it directly determines the number of photo-generated electrons and holes in the photocatalytic system. The effects of UV radiations of different intensities (220, 332, 437, 551, 663, and 774 μW/cm2) on the degradation rate of CBZ at a fixed wavelength of 365 nm, were studied. Obviously, as shown in Figure 4a, the degradation rate increased with increase in light intensity. In particular, CBZ could be completely degraded at an irradiation intensity of 774 μW/cm2 within 60 min. However, the degradation rate of CBZ at a light intensity of 220 μW/cm2 within 60 min was only 57.23%. When the light intensity was changed from 220 μW/cm2 to 774 μW/cm2, the reaction rate constant for photocatalytic degradation showed a 6.7-fold increase (Figure 4b).

2.4. Effect of Crystalline Form of Nano-TiO2

The photocatalytic performances of the three crystalline forms of TiO2 on CBZ degradation were studied. The degradation curves and pseudo first-order kinetic models are shown in Figure 5. The degradation rate of anatase form (87.78%) was significantly higher than the rutile form (80.29%), within 60 min. The mixed-phase TiO2 showed the highest rate and could degrade 98.5% of CBZ within 60 min. At the same time, the reaction rate constant of mixed-phase was as high as 0.067 min−1, which was nearly 2 and 3 times higher than those of anatase and rutile forms, respectively. Previous studies had demonstrated that the photogenerated carriers could easily transfer at the interface between the anatase and rutile forms, which promoted charge separation efficiency and photocatalytic activity [29].

2.5. Effect of Initial CBZ Concentration

CBZ solutions of four different concentrations (10.6, 21.2, 31.7, and 42.3 µM) were used to evaluate the effect of initial concentration of CBZ on degradation. As shown in Figure 6, when the CBZ concentration in the original aqueous solution increased from 10.6 to 42.3 µM, the degradation rate decreased slightly. However, the rate constant decreased sharply from 0.084 to 0.040 min−1. The reason for this phenomenon was that when the initial CBZ concentration was high, a higher number of CBZ molecules competed for the photogenerated holes and hydroxyl radicals in the system. The reduction in the main active material for photocatalytic reaction reduced the degradation rate of CBZ.

2.6. Effect of Nano-TiO2 Dosage

The effect of amount of photocatalyst on CBZ degradation was investigated by changing the mass concentration of initial TiO2 (mixed crystal form) in the range of 0.8–3.6 g/L. Figure 7 shows the degradation curves of CBZ with different dosages of TiO2. When the amount of TiO2 was 0.8 g/L, the degradation rate reached 87.78% within 60 min. The degradation rate gradually reached its maximum value (98.5%) when 2.8 g/L of TiO2 was used (Figure 7a). Further, as the TiO2 dosage was further increased from 2.8 to 3.6 g/L, the degradation efficiency decreased. High concentration of catalyst increases light scattering and reduces light penetration, which decreases the utilization of light. These factors were responsible for the decrease in degradation activity. CBZ underwent fastest degradation, with a rate constant of 0.058 min−1, when the TiO2 mass concentration was 2.4 g/L (Figure 7b).

2.7. Effect of pH

During the entire process of photocatalytic oxidation, the pH of reaction system also greatly influences the degradation efficiency of the target compounds, which in turn is mainly affected by the generation of hydroxyl radicals [30]. Figure 8 shows the degradation of CBZ under different pH conditions (3.0, 5.0, 7.0, 9.0, and 11.0). As the pH increased from 3.0 to 11.0, the degradation rate increased gradually. The reaction rate was the fastest when the pH was 11 and its rate constant reached 0.110 min−1, which was about 4 times than that at pH = 3.0. The ·OH species was mainly formed from OH, and the production efficiency of OH was much higher than that of H2O to produce ∙OH. Therefore, an aqueous alkaline environment was more conducive for the degradation of CBZ.

2.8. Influence of Co-Existing Ions in the Aquatic Environment

Coexistence of different anions and cations in water has different effects on the photocatalytic degradation of CBZ [31]. The concentration of a fixed coexisting anion and cation in the CBZ degradation reaction system was 1 mmol/L, keeping other conditions unchanged (C(mixed-phase TiO2) = 2.0 g/L; C(CBZ)0 = 18.8 µM; operating temperature = 20 °C; wavelength of UV-LED = 365 nm; I0 = 774 µW/cm2; and operating pH = 7.6). The effects of common anions and cations in water on the degradation of CBZ were investigated. As shown in Figure 9, the presence of Cl in the system had little effect on the photodegradation process of CBZ, whereas NO3 had a slight inhibitory effect. The SO42− promoted the degradation to certain extent, while the presence of CO32− and HCO3 significantly promoted the photocatalytic degradation of CBZ. When a large number of CO32− and HCO3 coexisted in the solution, the solution was alkaline. An increase in the number of OH ions was useful for the formation of ·OH, which improved the photocatalytic activity of the system. Hu et al. [32] reported that CO32− and HCO3 promoted the photocatalytic degradation of sulfamethoxazole, which was consistent with the results of this study.
The effects of four commonly coexisting cations on the degradation of CBZ in natural water are shown in Figure 10. The presence of Ca2+, Mg2+, and Cu2+ ions inhibited the photocatalytic degradation to a certain extent and the order of their inhibition effects was: Cu2+ > Ca2+ > Mg2+. The Fe3+ ions had a positive effect, and 100% CBZ degradation was observed in less than 20 min. The photocatalytic degradation of CBZ promoted by Fe3+ ions could be attributed to the fact that Fe3+ traps photo-generated electrons generated by TiO2, thereby effectively reducing the chance of electron-hole recombination. In addition, the addition of Fe3+ ions could form a Fenton system, thereby increasing the number of free active·OH species in the reaction system, which significantly increased the degradation capacity and degradation rate of CBZ.

2.9. Influence of Free Radical Scavenging

The ·OH radical is considered to be the main active species in TiO2 photocatalytic degradation reaction [33]. By adding different concentrations of t-butanol (TBA) as the OH scavenger to the reaction system, the effect of number of active ·OH radicals on the photocatalytic degradation was studied. As shown in Figure 11, the presence of TBA significantly inhibited the photocatalytic degradation of CBZ. With increase in TBA concentration in the reaction system, the inhibitory effect was enhanced. When the TBA concentration in the system was 10 mM, the degradation efficiency was 19.91% after 60 min of UV-LED irradiation. As the TBA concentration increased to 50 mM, the degradation efficiency was only 6.58%. These experiments also show that OH species are indeed the main active species in CBZ degradation.

3. Materials and Experimental Methods

3.1. Materials

Carbamazepine (≥98%, Shandong Xiya Chemical Co. Ltd., Linyi, Shandong, China) A stock mother liquor of CBZ, with a concentration of 211 µmol/L, was prepared and then serially diluted with ultrapure water, according to the concentration required for each experiment. Acetonitrile (HPLC grade purity, Merck KGaA, Darmstadt, Germany). HCl (0.1 mol/L), NaOH (0.01 mol/L), and phosphate buffer (0.01 mol/L) solutions were used to adjust the pH of the reaction system. The water used was ultra-pure and was prepared using Millipore Milli-Q ultra-pure water system, with a resistivity of 18.2 MΩ·cm. Three commercial nano TiO2 catalysts having different crystal forms: pure anatase, mixed-phase, and rutile forms (>98%, Shandong Xiya Chemical Co. Ltd., Linyi, Shandong, China). All other reagents were from Damao Chemical Reagent Factory (Analytical grade, Tianjin, China).

3.2. Test Device

The photocatalytic degradations were performed in a 500 mL open plexiglass container. In order to prevent light energy, UV-LED light, and heat energy, a layer of aluminum foil was tightly wrapped around the outer wall of the reactor. The UV-LED tube was placed in a cylindrical radiation window having a diameter of about 5 cm, which was made of quartz. The light source was composed of two self-made UV-LED units, and the working wavelengths of the two units were 275 and 365 nm. Each unit contained 40 LED chips (manufactured by Shenzhen Julan Technology Co. Ltd., China). The output of 275 nm was about 2.8 mW with a current of 40 mA and the output of 365 nm was about 700 mW with a current of 500 mA. The fluence rate was measured in all experiments using a potassium ferrioxalate actinometer [34]. A low-temperature constant temperature circulating water tank was used to circulate cooling water on the outer wall of the reactor in order to control the temperature during the operation at 20 ± 0.5 °C. The intensity of the radiations irradiated into the reactor was measured using an IL-2400 irradiation photometer (International Light, USA). The diagram of UV-LED/TiO2 photocatalytic device used in this study is shown in Figure 12.

3.3. Analytical Method

In the UV-LED/TiO2 photocatalytic experiment, 2 mL of the sample suspension was withdrawn at regular time intervals, and the suspension was filtered through a 0.22 µm syringe filter (Tianjin Jinteng Experimental Equipment Co. Ltd., Tianjin, China). The concentration of CBZ in the test sample was measured by HPLC (Waters e2695) using a Symmetry C18 column [35]. The reaction conditions were as follows: the mobile phase consisted of 60% acetonitrile and 40% pure water. The injection volume was 100 μL, the flow rate was 1 mL/min, the column temperature was maintained at 35 °C, and the UV-LED wavelength of CBZ was below 286 nm.

4. Conclusions

The main conclusions of this study were as follows:
(1) UV-LED irradiated TiO2 photocatalytic technology could effectively degrade CBZ in water, and the degradation rate of CBZ reached 98.5% after 60 min of reaction. The photocatalytic reaction process fit the pseudo-first-order kinetic model.
(2) Photocatalytic degradation of CBZ under LED wavelength of 365 nm was more effective than 275 nm, which could be due to the stronger permeability of 365 nm light in solution.
(3) Increasing the irradiation intensity of UV-LED could lead to increase the number of effective photons absorbed on the surface of TiO2. Hence, it could significantly promote the degradation of CBZ.
(4) The photocatalytic activities of different crystalline forms of TiO2 varied significantly and were in order of: mixed-phase > anatase > rutile.
(5) The optimal amount of mixed-phase TiO2 catalyst was 2.8 g/L, at which the CBZ degradation efficiency was the highest and excessive catalyst decreased the rate.
(6) The CBZ degradation efficiency increased with increase in pH.
(7) The CBZ degradation efficiency decreased with the increase in initial CBZ concentration in the aqueous solution.
(8) The effects of SO42−, NO3, Cl, Ca2+, and Mg2+ on the photocatalytic degradation of CBZ were negligible, and the presence of CO32−, HCO3, and Fe3+ ions in water significantly promoted CBZ degradation.
(9) Hydroxyl radical is the main active species in the degradation of CBZ by UV-LED photocatalytic reaction system.

Author Contributions

Data Curation, Writing—Original draft preparation, Z.R. and Y.F.; Methodology, Software. J.S.; Visualization, Investigation J.S. and Y.F.; Conceptualization, Methodology, Software, C.M.; Funding Acquisition, C.M. and Z.R. Supervision, Z.R., C.M. and S.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the International Cooperation Project of Shenzhen (GJHZ20180416164721073), the Science and Technology Project of SZIIT (SZIIT2019KJ006, 2017GKTSCX065), the National Science Foundation of China (51508383), and the Natural Science Foundation of Tianjin Province (18JCQNJC09000).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ellis, J.B. Pharmaceutical and personal care products (PPCPs) in urban receiving waters. Environ. Pollut. 2006, 144, 184–189. [Google Scholar] [CrossRef]
  2. Jacobs, H.E.; Haarhoff, J. Prioritisation of parameters influencing residential water use and wastewater flow. J. Water Supply Res. Technol. Aqua 2007, 56, 495–514. [Google Scholar] [CrossRef]
  3. Liu, J.L.; Wong, M.H. Pharmaceuticals and personal care products (PPCPs): A review on environmental contamination in China. Environ. Int. 2013, 59, 208–224. [Google Scholar] [CrossRef] [PubMed]
  4. Hofmann, R.; Amiri, F.; Wilson, S.; Garvey, E.; Metcalfe, C.; Ishida, C.; Lin, K. Comparing methods to remove emerging contaminants and disinfection by-product precursors at pilot scale. J. Water Supply Res. Technol. Aqua 2011, 60, 425–433. [Google Scholar] [CrossRef]
  5. Kwon, J.W.; Rodriguez, J.M. Occurrence and removal of selected pharmaceuticals and personal care products in three wastewater-treatment plants. Arch. Environ. Contam. Toxicol. 2014, 66, 538–548. [Google Scholar] [CrossRef] [PubMed]
  6. Xie, R.; Meng, X.; Sun, P.; Niu, J.; Jiang, W.; Bottomley, L.; Li, D.; Chen, Y.; Crittenden, J. Electrochemical oxidation of ofloxacin using a TiO2-based SnO2-Sb/polytetrafluoroethylene resin-PbO2 electrode: Reaction kinetics and mass transfer impact. Appl. Catal. B Environ. 2017, 203, 515–525. [Google Scholar] [CrossRef] [Green Version]
  7. Prajapati, S.T.; Gohel, M.C.; Patel, L.D. Studies to enhance dissolution properties of carbamazepine. Indian J. Pharm. Sci. 2007, 69, 427–430. [Google Scholar] [CrossRef] [Green Version]
  8. Snyder, S.A.; Wert, E.C.; Lei, H.; Westerhoff, P.; Yoon, Y. Removal of EDCs and Pharmaceuticals in Drinking and Reuse Treatment Processes. AWWA Res. Found. 2007. [Google Scholar]
  9. Wu, C.X.; Spongberg, A.L.; Witter, J.D.; Fang, M.; Czajkowski, K.P. Uptake of pharmaceutical and personal care products by soybean plants from soils applied with biosolids and irrigated with contaminated water. Environ. Sci. Technol. 2010, 44, 6157–6161. [Google Scholar] [CrossRef]
  10. Wang, S.Z.; Wang, J.L. Carbamazepine degradation by gamma irradiation coupled to biological treatment. J. Hazard. Mater. 2017, 321, 639–646. [Google Scholar] [CrossRef]
  11. Ali, I.; Al-Othman, Z.A.; Sanagi, M.M. Green synthesis of iron nano-impregnated adsorbent for fast removal of fluoride from water. J. Mol. Liq. 2015, 211, 457–465. [Google Scholar] [CrossRef]
  12. Aziz, A.R.A.; Asaithambi, P.; Daud, W.M.A.B.W. Combination of electrocoagulation with advanced oxidation processes for the treatment of distillery industrial effluent. Process Saf. Environ. Prot. 2016, 99, 227–235. [Google Scholar] [CrossRef]
  13. Gümüş, D.; Akbal, F. Comparison of Fenton and electro-Fenton processes for oxidation of phenol. Process Saf. Environ. Prot. 2016, 103, 252–258. [Google Scholar] [CrossRef]
  14. Popescu, M.; Rosales, E.; Sandu, C.; Meijide, J.; Pazos, M.; Lazar, G.; Sanromán, M.A. Soil flushing and simultaneous degradation of organic pollutants in soils by electrokinetic-Fenton treatment. Process Saf. Environ. Prot. 2017, 108, 99–107. [Google Scholar] [CrossRef]
  15. Vogna, D.; Marotta, R.; Andreozzi, R.; Napolitano, A.; d’Ischia, M. Kinetic and chemical assessment of the UV/H2O2 treatment of antiepileptic drug carbamazepine. Chemosphere 2004, 54, 497–505. [Google Scholar] [CrossRef]
  16. Horovitz, I.; Avisar, D.; Baker, M.A.; Grilli, R.; Lozzi, L.; Camillo, D.D.; Mamane, H. Carbamazepine degradation using a N-doped TiO2 coated photocatalytic membrane reactor: Influence of physical parameters. J. Hazard. Mater. 2016, 310, 98–107. [Google Scholar] [CrossRef] [Green Version]
  17. Wang, W.; Lu, Y.B.; Luo, H.P.; Liu, G.L.; Zhang, R.D.; Jin, S. A microbial electro-fenton cell for removing carbamazepine in wastewater with electricity output. Water Res. 2018, 139, 58–65. [Google Scholar] [CrossRef]
  18. Jin, Y.Y.; Wang, X.N.; Sun, S.P.; Dong, W.B.; Wu, Z.X.; Bian, G.Q.; Wu, W.D.; Wu, D.; Chen, X.D. Hydroxyl and sulfate radicals formation in UVA/FeIII-NTA/S2O82− system: Mechanism and effectiveness in carbamazepine degradation at initial neutral Ph. Chem. Eng. J. 2019, 368, 541–552. [Google Scholar] [CrossRef]
  19. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  20. Ye, Y.; Feng, Y.; Bruning, H.; Yntema, D.; Rijnaarts, H. Photocatalytic degradation of metoprolol by TiO2 nanotube arrays and UV-LED: Effects of catalyst properties, operational parameters, commonly present water constituents, and photo-induced reactive species. Appl. Catal. B Environ. 2018, 220, 171–181. [Google Scholar] [CrossRef]
  21. Kneissl, M.; Seong, T.Y.; Han, J.; Amano, H. The emergence and prospects of deep-ultraviolet light-emitting diode technologies. Nat. Photonics 2019, 13, 233–244. [Google Scholar] [CrossRef]
  22. Song, K.; Mohseni, M.; Taghipour, F. Application of ultraviolet light-emitting diodes (UV-LEDs) for water disinfection: A review. Water Res. 2016, 94, 341–349. [Google Scholar] [CrossRef]
  23. Matafonova, G.; Batoev, V. Recent advances in application of UV light-emitting diodes for degrading organic pollutants in water through advanced oxidation processes: A review. Water Res. 2018, 132, 177–189. [Google Scholar] [CrossRef] [PubMed]
  24. Korovin, E.; Selishchev, D.; Besov, A.; Kozlov, D. UV-LED TiO2 photocatalytic oxidation of acetone vapor: Effect of high frequency controlled periodic illumination. Appl. Catal. B Environ. 2015, 163, 143–149. [Google Scholar] [CrossRef]
  25. Cai, Q.; Hu, J. Effect of UVA/LED/TiO2 photocatalysis treated sulfamethoxazole and trimethoprim containing wastewater on antibiotic resistance development in sequencing batch reactors. Water Res. 2018, 140, 251–260. [Google Scholar] [CrossRef] [PubMed]
  26. Liang, R.; Van Leuwen, J.C.; Bragg, L.M.; Arlos, M.J.; Li, C.F.; Schneider, O.M.; Jaciw-Zurakowsky, I.; Fattahi, A.; Rathod, S.; Peng, P.; et al. Utilizing UV-LED pulse width modulation on TiO2 advanced oxidation processes to enhance the decomposition efficiency of pharmaceutical micropollutants. Chem. Eng. J. 2019, 361, 439–449. [Google Scholar] [CrossRef]
  27. Dai, K.; Lu, L.; Dawson, G. Development of UV-LED/TiO2 Device and Their Application for Photocatalytic Degradation of Methylene Blue. J. Mater. Eng. Perform. 2013, 22, 1035–1040. [Google Scholar] [CrossRef]
  28. D’Orazio, J.; Jarrett, S.; Amaro-Ortiz, A.; Scott, T. UV radiation and the skin. Int. J. Mol. Sci. 2013, 14, 12222–12248. [Google Scholar] [CrossRef] [Green Version]
  29. Heo, S.; Hwang, H.S.; Jeong, Y.; Na, K. Skin protection efficacy from UV irradiation and skin penetration property of polysaccharide-benzophenone conjugates as a sunscreen agent. Carbohydr. Polym. 2018, 195, 534–541. [Google Scholar] [CrossRef]
  30. Deiana, C.; Fois, E.; Coluccia, S.; Martra, G. Surface structure of TiO2 P25 nanoparticles: Infrared study of hydroxy groups on coordinative defect sites. J. Phys. Chem. C 2010, 114, 21531–21538. [Google Scholar] [CrossRef]
  31. Pereira, J.H.O.S.; Reis, A.C.; Queirós, D.; Nunes, O.C.; Borges, M.T.; Vilar, V.P.; Boaventura, R.A.R. Insights into solar TiO2-assisted photocatalytic oxidation of two antibiotics employed in aquatic animal production, oxolinic acid and oxytetracycline. Sci. Total Environ. 2013, 463–464, 274–283. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, L.; Guo, J.; Dang, J.; Huang, X.; Chen, S.; Guan, W. Comparison of the photocatalytic performance of TiO2/AC and TiO2/CNT nanocomposites for methyl orange photodegradation. Water Sci. Technol. 2018, 78, 1082–1093. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Hu, L.; Flanders, P.M.; Miller, P.L.; Strathmann, T.J. Oxidation of sulfamethoxazole and related antimicrobial agents by TiO2 photocatalysis. Water Res. 2007, 41, 2612–2626. [Google Scholar] [CrossRef] [PubMed]
  34. Al-Mamun, M.R.; Kader, S.; Islam, M.S.; Khan, M.Z.H. Photocatalytic activity improvement and application of UV-TiO2 photocatalysis in textile wastewater treatment: A review. J. Environ. Chem. Eng. 2019, 7, 103248. [Google Scholar] [CrossRef]
  35. Bolton, J.R.; Stefan, M.I.; Shaw, P.S.; Lykke, K.R. Determination of the quantum yields of the potassium ferrioxalate and potassium iodide–iodate actinometers and a method for the calibration of radiometer detectors. J. Photochem. Photobiol. A Chem. 2011, 222, 166–169. [Google Scholar] [CrossRef]
Figure 1. SEM images, UV-Vis graphs, and XRD patterns of anatase, mixed phase, and rutile forms of TiO2 nanoparticles. (ac) SEM images, (d) UV-Vis DRS, (e) N2 adsorption-desorption isotherms, and (f) XRD patterns.
Figure 1. SEM images, UV-Vis graphs, and XRD patterns of anatase, mixed phase, and rutile forms of TiO2 nanoparticles. (ac) SEM images, (d) UV-Vis DRS, (e) N2 adsorption-desorption isotherms, and (f) XRD patterns.
Catalysts 10 00540 g001
Figure 2. (a) CBZ degradation control experiments (b) The pseudo-first-order kinetic models of CBZ degradation. C(CBZ)0 = 18.8 µM; operating temperature = 20 °C; wavelength of UV-LED = 365 nm; I0 = 774 µW/cm2; and operating pH = 7.6.
Figure 2. (a) CBZ degradation control experiments (b) The pseudo-first-order kinetic models of CBZ degradation. C(CBZ)0 = 18.8 µM; operating temperature = 20 °C; wavelength of UV-LED = 365 nm; I0 = 774 µW/cm2; and operating pH = 7.6.
Catalysts 10 00540 g002
Figure 3. (a) Degradation of CBZ by UV-LED/TiO2 using different wavelengths of UV-LED. (b) The pseudo-first-order kinetic models of CBZ degradation by UV-LED/TiO2 using different wavelengths of UV-LED. C(CBZ)0 = 18.8 µM; operating temperature = 20 °C; wavelength of UV-LED: 365 nm; I0 = 774 µW/cm2, operating pH = 7.6; and C (mixed-phase TiO2) = 2.0 g/L.
Figure 3. (a) Degradation of CBZ by UV-LED/TiO2 using different wavelengths of UV-LED. (b) The pseudo-first-order kinetic models of CBZ degradation by UV-LED/TiO2 using different wavelengths of UV-LED. C(CBZ)0 = 18.8 µM; operating temperature = 20 °C; wavelength of UV-LED: 365 nm; I0 = 774 µW/cm2, operating pH = 7.6; and C (mixed-phase TiO2) = 2.0 g/L.
Catalysts 10 00540 g003
Figure 4. (a) Degradation of CBZ by UV-LED/TiO2 with different light intensities. (b) The pseudo-first-order kinetic models of CBZ degradation by UV-LED/TiO2 with different light intensities. C(CBZ)0 = 18.8 µM; operating temperature = 20 °C; wavelength of UV-LED = 365 nm; operating pH = 7.6; and C (mixed-phase TiO2) = 2.0 g/L.
Figure 4. (a) Degradation of CBZ by UV-LED/TiO2 with different light intensities. (b) The pseudo-first-order kinetic models of CBZ degradation by UV-LED/TiO2 with different light intensities. C(CBZ)0 = 18.8 µM; operating temperature = 20 °C; wavelength of UV-LED = 365 nm; operating pH = 7.6; and C (mixed-phase TiO2) = 2.0 g/L.
Catalysts 10 00540 g004
Figure 5. (a) Degradation of CBZ by UV-LED/TiO2 using different crystal forms of nano-TiO2 photocatalyst. (b) The pseudo-first-order kinetic models of CBZ degradation by UV-LED/TiO2 using different crystal forms of nano-TiO2 photocatalyst. C(CBZ)0 = 18.8 µM; operating temperature = 20 °C; wavelength of UV-LED = 365 nm; operating pH = 7.6; and C (mixed-phase TiO2) = 2.0 g/L.
Figure 5. (a) Degradation of CBZ by UV-LED/TiO2 using different crystal forms of nano-TiO2 photocatalyst. (b) The pseudo-first-order kinetic models of CBZ degradation by UV-LED/TiO2 using different crystal forms of nano-TiO2 photocatalyst. C(CBZ)0 = 18.8 µM; operating temperature = 20 °C; wavelength of UV-LED = 365 nm; operating pH = 7.6; and C (mixed-phase TiO2) = 2.0 g/L.
Catalysts 10 00540 g005
Figure 6. (a) Degradation of CBZ by UV-LED/TiO2 for different initial concentrations of CBZ (b) The pseudo-first-order kinetic models of CBZ degradation by UV-LED/TiO2 for different initial concentrations of CBZ. Operating temperature = 20 °C; wavelength of UV-LED = 365 nm; operating pH = 7.6; and C (mixed-phase TiO2) = 2.0 g/L.
Figure 6. (a) Degradation of CBZ by UV-LED/TiO2 for different initial concentrations of CBZ (b) The pseudo-first-order kinetic models of CBZ degradation by UV-LED/TiO2 for different initial concentrations of CBZ. Operating temperature = 20 °C; wavelength of UV-LED = 365 nm; operating pH = 7.6; and C (mixed-phase TiO2) = 2.0 g/L.
Catalysts 10 00540 g006
Figure 7. (a) Degradation of CBZ by UV-LED/TiO2 using different concentrations of nano-TiO2 (b) The pseudo-first-order kinetic models of CBZ degradation by UV-LED/TiO2 using different concentrations of nano-TiO2. C(CBZ)0 = 18.8 µM; operating temperature = 20 °C; wavelength of UV-LED = 365 nm; I0 = 774 µW/cm2: and operating pH = 7.6.
Figure 7. (a) Degradation of CBZ by UV-LED/TiO2 using different concentrations of nano-TiO2 (b) The pseudo-first-order kinetic models of CBZ degradation by UV-LED/TiO2 using different concentrations of nano-TiO2. C(CBZ)0 = 18.8 µM; operating temperature = 20 °C; wavelength of UV-LED = 365 nm; I0 = 774 µW/cm2: and operating pH = 7.6.
Catalysts 10 00540 g007
Figure 8. (a) Degradation of CBZ by UV-LED/TiO2 under different pH conditions (b) The pseudo-first-order kinetic models of CBZ degradation by UV-LED/TiO2 under different pH conditions. C(CBZ)0 = 18.8 µM; operating temperature = 20 °C; and wavelength of UV-LED = 365 nm.
Figure 8. (a) Degradation of CBZ by UV-LED/TiO2 under different pH conditions (b) The pseudo-first-order kinetic models of CBZ degradation by UV-LED/TiO2 under different pH conditions. C(CBZ)0 = 18.8 µM; operating temperature = 20 °C; and wavelength of UV-LED = 365 nm.
Catalysts 10 00540 g008
Figure 9. (a) Degradation of CBZ by UV-LED/TiO2 in presence of different co-existing anions (b) The pseudo-first-order kinetic models of CBZ degradation by UV-LED/TiO2 in presence of different co-existing anions.
Figure 9. (a) Degradation of CBZ by UV-LED/TiO2 in presence of different co-existing anions (b) The pseudo-first-order kinetic models of CBZ degradation by UV-LED/TiO2 in presence of different co-existing anions.
Catalysts 10 00540 g009
Figure 10. (a) Degradation of CBZ by UV-LED/TiO2 with different co-existing cations. (b) The pseudo-first-order kinetic models of CBZ degradation by UV-LED/TiO2 with different co-existing cations.
Figure 10. (a) Degradation of CBZ by UV-LED/TiO2 with different co-existing cations. (b) The pseudo-first-order kinetic models of CBZ degradation by UV-LED/TiO2 with different co-existing cations.
Catalysts 10 00540 g010
Figure 11. (a) Degradation of CBZ by UV-LED/TiO2 using different concentrations of TBA (b) The pseudo-first-order kinetic models of CBZ degradation by UV-LED/TiO2 using different concentrations of TBA.
Figure 11. (a) Degradation of CBZ by UV-LED/TiO2 using different concentrations of TBA (b) The pseudo-first-order kinetic models of CBZ degradation by UV-LED/TiO2 using different concentrations of TBA.
Catalysts 10 00540 g011
Figure 12. The diagrammatic representation of UV-LED/TiO2 photocatalytic device.
Figure 12. The diagrammatic representation of UV-LED/TiO2 photocatalytic device.
Catalysts 10 00540 g012
Table 1. Molecular structure and solubility of carbamazepine (CBZ).
Table 1. Molecular structure and solubility of carbamazepine (CBZ).
Chemical FormulaMolecular Weight/(g·mol−1)Solubility (20 °C)/(µg·mL−1)Molecular Structure
C15H12N2O236.2736.5 Catalysts 10 00540 i001
Table 2. Characteristics of nano-TiO2 nanoparticles.
Table 2. Characteristics of nano-TiO2 nanoparticles.
Crystalline PhaseCompositionParticle Size (nm)BET Surface Area (m2/g)
Anatase100% anatase20–10078.7
Rutile100% rutile50–10032.2
Mixed 83% anatase + 17% rutile20–100102.6

Share and Cite

MDPI and ACS Style

Ran, Z.; Fang, Y.; Sun, J.; Ma, C.; Li, S. Photocatalytic Oxidative Degradation of Carbamazepine by TiO2 Irradiated by UV Light Emitting Diode. Catalysts 2020, 10, 540. https://doi.org/10.3390/catal10050540

AMA Style

Ran Z, Fang Y, Sun J, Ma C, Li S. Photocatalytic Oxidative Degradation of Carbamazepine by TiO2 Irradiated by UV Light Emitting Diode. Catalysts. 2020; 10(5):540. https://doi.org/10.3390/catal10050540

Chicago/Turabian Style

Ran, Zhilin, Yuanhang Fang, Jian Sun, Cong Ma, and Shaofeng Li. 2020. "Photocatalytic Oxidative Degradation of Carbamazepine by TiO2 Irradiated by UV Light Emitting Diode" Catalysts 10, no. 5: 540. https://doi.org/10.3390/catal10050540

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