Next Article in Journal
The Effects of Expected Benefits on Image, Desire, and Behavioral Intentions in the Field of Drone Food Delivery Services after the Outbreak of COVID-19
Next Article in Special Issue
Impacts of Climatic Variation and Human Activity on Runoff in Western China
Previous Article in Journal
Looking for Answers to Food Loss and Waste Management in Spain from a Holistic Nutritional and Economic Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ti/RuO2-IrO2-SnO2 Anode for Electrochemical Degradation of Pollutants in Pharmaceutical Wastewater: Optimization and Degradation Performances

1
Key Laboratory of Yellow River Water Environment in Gansu Province, Lanzhou Jiaotong University, Lanzhou 730070, China
2
School of Environmental and Municipal Engineering, Lanzhou Jiaotong University, Lanzhou 730070, China
*
Author to whom correspondence should be addressed.
Sustainability 2021, 13(1), 126; https://doi.org/10.3390/su13010126
Submission received: 17 November 2020 / Revised: 18 December 2020 / Accepted: 21 December 2020 / Published: 24 December 2020

Abstract

:
Electrochemical oxidation technology is an effective technique to treat high-concentration wastewater, which can directly oxidize refractory pollutants into simple inorganic compounds such as H2O and CO2. In this work, two-dimensionally stable anodes, Ti/RuO2-IrO2-SnO2, have been developed in order to degrade organic pollutants from pharmaceutical wastewater. Characterization by scanning electron microscopy (SEM), energy-dispersive X-ray spectroscopy (EDS), and X-ray diffraction (XRD) showed that the oxide coating was successfully fabricated on the Ti plate surface. Electrocatalytic oxidation conditions of high concentration pharmaceutical wastewater was discussed and optimized, and the best results showed that the COD removal rate was 95.92% with the energy consumption was 58.09 kW·h/kgCOD under the electrode distance of 3 cm, current density of 8 mA/cm2, initial pH of 2, and air flow of 18 L/min.

1. Introduction

In the past several decades, pharmaceutical industry has played an important role in social development, but it also produces a lot of environmental pollution. The pharmaceutical industry wastewater is highly complex and contains a large number of toxic and poorly biodegradable compounds. Due to the unique properties of the pharmaceutical industry, it is recognized as one of the difficult high-concentration organic wastewater [1,2]. Traditional treatment methods (such as adsorption, cohesion, filtration, ion exchange, etc.) are not enough to completely purify the mineralized pharmaceutical wastewater, and the post-treatment procedures are too complicated, which limits the application in actual engineering [3,4]. In recent years, many researchers have devoted themselves to the development of advanced oxidation methods to treat wastewater, mainly including chemical oxidation, Fenton oxidation, wet air oxidation, supercritical water oxidation, photochemical oxidation and electrochemical oxidation [5,6,7,8,9]. Among many advanced oxidation methods, electrocatalytic oxidation technology, as the green chemical technology, can degrade complex organic pollutants in wastewater into non-toxic small molecular organic compounds or carbon dioxide and water through mineralization of various organic pollutants [10,11,12,13,14,15]. A large number of studies have shown that electrocatalytic oxidation technology is one of the effective methods for the treatment of high concentration pharmaceutical wastewater, but the technology is still in the initial stage [16].
In the electrocatalytic oxidation technology, the electrode is the key factor to complete the electrochemical reaction and improve the current efficiency. Different electrode materials can change the electrochemical reaction rate by orders of magnitude. Therefore, choosing the appropriate electrode material is an effective way to improve the efficiency of the electrochemical reaction. Since the development of the dimensionally stable anodes (DSA) electrode in the 1960s, it has good stability and catalytic activity and quickly gained popularity and has been widely used. The DSA electrode is a kind of metal oxide (such as Ti/SnO2) coating electrode with excellent electrocatalytic activity. This electrode has the advantages of high oxygen evolution potential, low cost, and easy operation, but also has low stability and short life.
Nowadays, to prolong the electrode life, researchers usually doped noble metal oxides (such as RuO2 and IrO2) with strong corrosion resistance and high stability into the titanium substrate anode coating [17,18]. For instance, Jiang et al. used a nanocrystal Ti/IrO2 coating electrode to degrade TNT red water, the results show that the COD removal rate was 68.5% after 30 h [19]. Zhang et al. utilized the Ti/SnO2-RuO2 composite electrode to degrade industrial gallic acid effluent, and the COD removal rate was 76.0% after 6 h [20]. Pathiraja et al. developed the Ti/IrO2 and Ti/IrO2–SnO2 electrodes to degrade chlorpyrifos. The chemical oxygen demand (COD) results revealed that the as-prepared Ti/IrO2-SnO2 electrode has 78.2% COD removal efficiency in comparison to 65.1% of COD removal efficiency obtained with Ti/IrO2 electrode after the electrolysis time of 6 h. Additionally, the instantaneous current efficiency value is always high in Ti/IrO2–SnO2 electrode during the period of reaction under similar conditions than the Ti/IrO2 electrode [21]. Obviously, the doping of noble metal oxides is beneficial to improve the electrochemical performance and the stability and the electrode life.
Hence, in this work, we used the ternary oxide composite Ti/RuO2-IrO2-SnO2 as the anode for electrocatalytic oxidation of pharmaceutical wastewater for the first time. The effects of electrode distance, current density, initial pH, and air-flow on the electrocatalytic degradation performance of organic from pharmaceutical wastewater were systematically investigated. The optimal reaction conditions were determined and the energy consumption of the reaction was investigated. In addition, the SEM, XRD, and EDS techniques are employed to analyze the surface morphology, structure, and composition of Ti/RuO2-IrO2-SnO2 anode before and after the electrocatalytic reaction.

2. Materials and Methods

2.1. Reagent and Materials

All chemicals were purchased commercially and can be used without further purification. Iridium hexahydrate (H2IrCl6·6H2O, AR), Hydrochloric acid (HCl, AR), Sodium hydroxide (NaOH, AR), Oxalic acid (C2H2O4, AR) were purchased from Tianjin Damao Chemical Co., Ltd. (Tianjin, China). Tin (IV) chloride (SnCl4·5H2O, AR) was purchased from Shanghai Zhongqin Chemical Reagent Co., Ltd. (Shanghai, China). Ruthenium (III) chloride (RuCl3·xH2O, AR), Citric acid (C6H8O7, AR), Ethylene glycol ((CH2OH)2, AR) were purchased from Tianjin Yongda Chemical Reagent Co., Ltd. (Tianjin, China). Deionized water with a specific resistance of 18.2 MΩ·cm was obtained by reverse osmosis, by ion–exchange and filtration. Ti plate was obtained from Baoji Changli Special Metal Co., Ltd. (Baoji, China). The pharmaceutical wastewater is taken from the Chinese medicine extraction workshop of Gansu Lanyao Pharmaceutical Co., Ltd. (Lanzhou, China), the composition of pharmaceutical wastewater is shown in Table 1.

2.2. Characterization

The surface morphology of the electrodes was observed by SEM (ULTRA plus, Carl Zeiss AG, Jena, German) and quantitative analyze trace element on the electrode surface by EDS (ULTRA plus, Carl Zeiss AG, Jena, German). Under 30 kw, 60 kV and 60 mA Cu Kα (λ = 1.5406 Å), the phase formation on the electrode was investigated via powder X-ray diffraction (XRD) technology on X’pert pro X-Ray Diffractometer. Diffraction patterns in the 20 to 80° regions were recorded at a rate of 5°/min. Track box furnace (KSL1400X, Kejing, Hefei, China) was used for the sintering of anodic oxide coating. Reactor power supply was provided by DC stabilized power supply (APS3005S-3D, Gratten, Nanjing, China). The pH of raw water was measured by PHS-3C pH meter (Yoke Instrument, Shanghai, China). The content of COD in water samples before and after reaction was measured by UV-vis spectrophotometer (UV-2800, UNICO, Shanghai, China).

2.3. Preparation of the Ti/RuO2-IrO2-SnO2 Electrode

The Ti/RuO2-IrO2-SnO2 electrode was prepared by a simple thermal decomposition method according to the previous literatures [22,23]. Typically, Ti plates with dimensions of 250 × 250 × 2 mm were used as the substrates. The Ti plates was etched in 20% boiling oxalic acid for 2 h, washed with deionized water and stored in ethanol solution. This pre-treatment served to remove the surface oxide layer and improve the coating adhesion of Ti plates. Ruthenium chloride, chloroiridic acid, and tin tetrachloride were used as the precursor materials. Using the Pechini method, the polymeric precursor solution was prepared by mixing citric acid (CA) and ethylene glycol (EG) at 60–70 °C. A certain molar ratio of Ru-Ir-Sn precursor was added to the polymer solution, and it was kept at 60–70 °C for 30 min to dissolve. The obtained precursor solution was evenly coated on both sides of the Ti substrate with a brush, dried at 80 °C for 8 min and then sintered in track box furnace at 300 °C for 10 min. The above steps were repeated until the precursor solution was used up. Finally, the sample was put into track box furnace and annealed at 550 °C for 1 h [24].

2.4. Experimental Device

The experimental reactor consisted of a power supply, Ti/RuO2-IrO2-SnO2 anode, Ti cathode, aerator, and an electrocatalytic cell as shown in Figure 1. The electrocatalytic oxidation cell is made of quartz glass with a size of 250 × 160 × 280 mm. The built-in anode is a self-made Ti/RuO2-IrO2-SnO2 electrode plate. The cathode is a Ti plate purchased from Baoji Changli Special Metal Co., Ltd. The electrodes were connected to power supply to install the reactor. The reactor is an improved parallel structure reactor with a shared-anode and double electrolytic cells, that is, it is composed of one anode plate and two cathode plates. The effective area of the anode plate immersed in solution is 250 × 250 × 2 mm. This device can double the treatment capacity of organic pollutants without increasing the area of the electrolytic cell, and effectively improve the utilization efficiency of the anode plate.

2.5. Experimental Procedure and Analytical Methods

The COD value in the system was determined by the fast confined catalytic digestion-spectrophotomete. The detection limit of this method for COD is 50~1000 mg/L [25]. The water samples in this experiment were measured after dilution. Generally, experiments were proceeding three times in parallel for each process, and the final data was taking the average values.
The electrocatalytic performance of the reactor was evaluated according to the COD removal rate and energy consumption. Determination of COD using fast closed digestion method. The COD removal rate and energy consumption were calculated using the following formula [26]:
COD removal rate (%) = (COD0 − COD1)/COD0·100
Energy consumption (EC) = U·I·T/(COD0 − COD1)/V
where COD0 is the COD value of raw pharmaceutical wastewater, mg/L; COD1 is the COD value of the treated pharmaceutical wastewater, mg/L; U is the reaction voltage, V; I is the reaction current, A; T is the reaction time, h; V is the volume of pharmaceutical wastewater, L.

3. Results and Discussion

3.1. Characterization of the Electrodes

The stability and lifetime of the electrode is a significant factor affecting the capability of the electrocatalytic oxidation reactor. The configuration of surface and constituent elements of electrodes are often served as precise and straightforward indicators for evaluating electrode stability [27]. As shown in Figure 2, the surface morphology of electrode was studied by SEM and EDS. In Figure 2a, the surface of the titanium electrode is generally flat with a few small particles. In Figure 2b, the anode (Ti/RuO2-IrO2-SnO2) was inerratic with many cracks that is representative “cracked clay” characteristic of electrodes produced by thermal decomposition, which may be caused by the high-temperature calcination of the precursor coating of the metal oxide formed on the surface of the Ti plates. The calcination of the electrode surface not only facilitates the formation of oxide coatings, but also increases the electrochemically active surface area [28].
The EDS spectrum of the coating on the RuO2-IrO2-SnO2 anode showed that Sn, Ru, Ir, Ti, C and O were on the electrode surface (as shown in Figure 3). This indicated that RuO2, IrO2, and SnO2 were successfully coated on the Ti plates. The C signal comes from the presence of adventitious carbon on the electrode surface. The observed titanium signal may be due to the inability of the oxide coating to completely cover the surface of the titanium substrate. In addition, we obtain the percentage of element content on the electrode surface through EDS analysis (as shown in Table 2). The weight percentages of Ti, Ru, Ir and Sn are 27.37, 8.71, 4.56 and 7.47%, respectively. After calculation, it is found that there are 17.16% of RuO2, 0.33% of IrO2, and 14.19% SnO2 on the surface of the Ti plate, respectively. The high concentration of ruthenium oxide is located in the electrode, which not only provides more active sites for electrocatalytic oxidation, but also enhances the stability and life of the electrode.
Furthermore, the crystal structure and chemical composition of the electrode was also evaluated by the XRD technique, and given in Figure 4. Obviously, the identified phases were RuO2 (JCPDS: 40-1290), IrO2 (JCPDS: 15-0870), SnO2 (JCPDS: 70-4177), Ti (JCPDS: 44-1288) and TiO2 (JCPDS: 01-1292, 04-0477). Except for the above diffraction peaks, there are no other obvious diffraction peaks in the XRD pattern. For Ti/RuO2-IrO2-SnO2 electrode plate, the diffraction angle 2θ is 28.009°, 35.050°, 54.245°,66.989 corresponding to the (110), (101), (211), (112)crystal planes of RuO2, the diffraction angle 2θ is 34.714°, 40.062°, 50.024° corresponding to the (101), (200), (211) crystal planes of IrO2, the diffraction angle 2θ is 26.601°, 37.974°, 54.790° corresponding to the (110), (200), (220) crystal planes of SnO2. The Ti diffraction angle 2θ is 38.481°, 55.541°, corresponding to crystal planes (110) and (200), the rutile diffraction angle 2θ is 27.506°, 76.082°, corresponding to crystal planes (110) and (202), the anatase diffraction angle 2θ is 68.594°, 70.357°, corresponding to crystal planes (116) and (220). This indicates that there are thin layers of RuO2, IrO2, SnO2 and TiO2 on the electrode.

3.2. Performance of Electrocatalytic Reactor

3.2.1. Effect of Electrode Distance

For the electrocatalytic oxidation, it was of interest to investigate the effect of electrode distance. At a current density of 1 mA/cm2, air flow of 12 L/min, pH of 7, as shown in Figure 5, When the electrode distance increased from 3 to 5 cm, the COD removal rate dropped by 10.2%. The experimental results indicated that the electrocatalytic performance with a plate spacing of 3 cm higher than a plate spacing of 5 cm. The reason was that when the distance between the plates is 5 cm, the distance between the two plates is too large, and the ion transmission distance between the two plates becomes longer, which increases the mass transfer resistance and weakens the mass transfer effect, thereby resulting in reduced electrochemical reaction efficiency [29,30]. Enhancing the electrode distance requires increasing the battery voltage output to maintain a constant current density. Under high voltage, not only energy consumption increase, but side reactions such as the release of hydrogen and oxygen will occur on the surface of the anode and cathode, which will reduce the efficiency of organic degradation [31,32].
In addition, Sun et al. found that the electrocatalytic oxidation performance is significantly reduced, the operating cost is increased with the distance between the anode and cathode decreases (especially less than 3 cm). It may be because the distance between the two electrodes is small, which makes it difficult for the solution between the plates to exchange with the solution outside the plates, which leads to polarization of wastewater concentration. This situation will affect the mass transfer efficiency, resulting in a decrease in the degradation rate [33]. At the same time, when the distance between the electrodes is very close, the surface of the electrode cannot be supplemented with oxygen, resulting in a reduction in the generation efficiency of OH and the degradation efficiency of pollutants. Therefore, the optimal electrode distance is 3 cm.

3.2.2. Effect of Current Density

Current density affects electrocatalytic performance by affecting electron transfer in oxidation reactions, and proper current density is critical for electrocatalytic removal of COD in wastewater. In Figure 6a, the degradation rate of COD increases obviously when the current density gradually increases in the range of 1–8 mA/cm2. When the current density was 8 mA/cm2, the COD removal rate was 90.52%, with the energy consumption of 81.71(kW·h)/kg COD. As the current density reached 10 mA/cm2, the COD removal rate was 90.48%.
The current density has a significant impact on the COD removal effect of the electrocatalytic system, and directly affects the current transmission rate and treatment effect in the electrochemical reaction process. As shown in Figure 6b, in the process of enhancing the current density from 1 to 8 mA/cm2, the degradation efficiency showed the simultaneously increasing trend. With the increase in current density, the rate and intensity of charge transfer also increase, thereby promoting the formation of active intermediates and increasing the degradation efficiency of pollutant in wastewater [34,35]. It may be that the side reactions also occur on the electrode surface under too high current density, and the concentration of the active group decreases, thereby reducing the efficiency of reaction electrolysis and increasing energy consumption. At the same time, the tank voltage is too high, the bypass current and short-circuit current increase significantly, the temperature of the wastewater rises, and part of the electrical energy is converted into heat. Therefore, the best current density is 8 mA/cm2.

3.2.3. Effect of Air Flow

In general, the degradation efficiency of organic pollutants is affected by the reactive oxygen radicals in the catalytic oxidation reaction system, and the formation of reactive oxygen radicals (e.g., superoxide radical, hydroxyl radical) is related to the oxygen of reaction system, as shown in the Equations (3) and (4) [36,37].
O2 + e → •O2−
O2 + 2H+ + 2e → H2O2 → 2•OH
Obviously, a systematic investigation of the oxygen (or air) concentration in the reaction system is of great significance for improving the performance of electrocatalytic oxidation. Figure 7 shows the law of COD removal efficiency and energy consumption in wastewater when the air flow rate is increased from 6 L/min to 30 L/min. With the increase in air flow rate, the removal rate of COD in wastewater is increasing because the concentration of H2O2 produced by the system increases. The increase in air velocity increases the dissolved oxygen in the water, increases the cathode efficiency, and produces more H2O2, which promotes an increase in the removal rate of wastewater COD. When the rate reaches 18 L/min, the best COD removal rate is 94.06%, and the energy consumption is 74.63 (kW·h)/kgCOD. When the air velocity is higher than 18 L/min, the wastewater COD removal rate does not increase significantly with the increase in the air velocity, indicating that when the air velocity is higher than 18 L/min, the wastewater COD removal rate is mainly controlled by the electrochemical kinetics that produces H2O2. In addition, increasing the oxygen/air concentration in the catalytic reaction system promotes the formation of active oxygen species, thereby improving the removal efficiency of COD [38]. Excessively increasing the air flow makes it difficult for the organic pollutants in the wastewater to be effectively adsorbed on the electrode surface and react with active oxygen species, which causes the COD removal rate to gradually decrease [39,40]. The results further imply that the electrocatalytic degradation rate is the largest when the aeration amount is 18 L/min.

3.2.4. Effect of Initial pH

The initial pH value in the electrocatalytic reaction determines the reaction process and the direction of the chemical reaction, and the rate of hydrogen peroxide or hydroxyl radical at the cathode is affected by the initial pH (as shown in Figure 8). At electrode distance of 3 cm, the current density of 8 mA/cm2, air flow of 18 L/min, the COD removal rate increased with the decreasing pH as shown in Figure 8. The removal rate of COD was 91.45% at pH 7, and the energy consumption was 53.93(kW·h)/kgCOD. In this process, hydroxyl radicals play an important role in the removal of hard-degradable pollutants [41]. Under the dual action of the electric field and the DSA electrode, a large amount of ·OH is generated in the reactor [42,43]. Compared with pH 7, the COD removal rate was 95.92% and the energy consumption was to 58.09(kW·h)/kgCOD at pH 2. It may be because H2O2 is more easily produced in the reaction system at the lower pH (O2 +2H+ + 2e → H2O2), thereby further increasing the degradation efficiency [44].

3.3. Characterization of the Electrode after Reaction

To further demonstrate the Ti/RuO2-IrO2-SnO2 electrode has excellent electrocatalytic ability and stability, we checked the surface morphology of the Ti/RuO2-IrO2-SnO2 electrode after long-term operation by SEM and EDS technology, and the results are given in Figure 9 and Figure 10. On the whole, the surface of the anode and cathode electrodes has undergone significant changes for a long time. For the cathode (Figure 9a), the smoothness of the surface of the Ti plate is reduced, and the surface has a coral-like structure. This may be due to the accumulation of organic molecules during the reaction. For the anode (Figure 9b), the surface smoothness of the Ti/RuO2-IrO2-SnO2 electrode also decreases, and a large number of granular structures are accumulated on the surface and cracks, with a particle diameter of about 300 nm. This may be due to the adsorption of a large amount of particulate organic on the surface of the oxide coating of the anode during the long-term reaction.
To verify this, we conducted EDS analysis for the anode material Ti/RuO2-IrO2-SnO2, and the results are shown in Figure 10. Obviously, Ti, Sn, Ru, Ir and other metal elements still exist on the electrode surface. In addition, compared with fresh electrodes (see Figure 3), the content of C element on the electrode surface greatly increases after a long time of reaction, which is caused by the adsorption of a large number of organic pollutants in the wastewater on the electrode surface. In addition, according to the results of the weight percentage of each element (as shown in Table 3), the carbon content has increased from 9.86 to 54.85%, which further proves that our inference is reasonable. The above results further suggested that the Ti/RuO2-IrO2-SnO2 electrode exhibited excellent activity and stability for the electrocatalytic oxidation of pharmaceutical wastewater during the long-term reaction.

4. Conclusions

In summary, we use a simple thermal decomposition method to prepare an oxide coating (RuO2-IrO2-SnO2) on the Ti substrate as an anode material for the electrocatalytic oxidation of pollutants in pharmaceutical wastewater. Experimental results show that the ternary oxide composite coating electrode has an excellent performance in the electrocatalytic oxidation of pharmaceutical wastewater. COD removal efficiency and energy consumption were used as evaluation indexes, and various parameters (such as electrode distance, current density, air flow, and pH) were optimized. The optimal reaction conditions are that the electrode distance is 3 cm, the current density is 8 mA/cm2, the air flow rate is 18 L/min, and the pH is 2. Under these conditions for 24 h of operation, the COD removal rate was 95.92%, and the energy consumption was 58.09 kW·h/kgCOD. No other pollutants are generated during the reaction, the pollutant concentration of the effluent is greatly reduced, and there is no secondary pollution. It is an environment-friendly, efficient and stable method for the degradation of organic pollutants in various industrial wastewaters.
But compared with the traditional water treatment technology, the operating cost of the electrocatalytic oxidation method has a certain increase, but with the development of new electrocatalytic materials, the production cost will be greatly reduced. The TiO2 used in this article also reduces the cost of the electrode plate to a certain extent. This work helps to understand the good oxidative degradation performance of the electrocatalytic oxidation system on high-concentration wastewater, the feasibility of the electrocatalytic oxidation method used in actual industrial wastewater treatment is explored, which provides certain technical support for the practical application of this technology.

Author Contributions

Conceptualization, G.Z.; formal analysis, X.H.; data curation, X.H. and J.M.; writing—original draft preparation, X.H.; writing—review and editing, T.Z.; funding acquisition, F.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China, grant number No. 51768031, National Key Research and Development Project, grant number (2019YFD1100103), Tianyou innovation team of Lanzhou Jiaotong University, grant number (TY202005) and the Youth Science Foundation of Lanzhou Jiaotong University, grant number (No. 2019012).

Informed Consent Statement

Informed consent was obtained from all subjects involved in the study.

Data Availability Statement

The data presented in this study are available in [supplementary material].

Acknowledgments

This research was supported by the financial contribution from Key laboratory of Yellow River Water Environment in Gansu Province and School of Environmental and Municipal Engineering of Lanzhou Jiaotong University. This work was supported by the National Natural Science Foundation of China (No. 51768031), National Key Research and Development Project(2019YFD1100103), Tianyou innovation team of Lanzhou Jiaotong University (TY202005), the Youth Science Foundation of Lanzhou Jiaotong University (No. 2019012).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

DSADimensionally Stable Anodes
CODChemical Oxygen Demand
ECEnergy Consumption

References

  1. Lucas, D.; Barcelo, D.; Rodriguezmozaz, S. Removal of pharmaceuticals from wastewater by fungal treatment and reduction of hazard quotients. Sci. Total Environ. 2016, 571, 909–915. [Google Scholar] [CrossRef] [PubMed]
  2. Zhao, F.; Ju, F.; Huang, K.; Mao, Y.; Zhang, X.X.; Ren, H.; Zhang, T. Comprehensive insights into the key components of bacterial assemblages in pharmaceutical wastewater treatment plants. Sci. Total Environ. 2019, 651, 2148–2215. [Google Scholar] [CrossRef] [PubMed]
  3. Zhang, J.F.; Chang, V.W.C.; Giannis, A.; Wang, J.Y. Removal of cytostatic drugs from aquatic environment: A review. Sci. Total Environ. 2013, 445, 281–298. [Google Scholar] [CrossRef] [PubMed]
  4. Wang, X.H.; Lin, A.Y.C. Is the phototransformation of pharmaceuticals a natural purification process that decreases ecological and human health risks? Environ. Pollut. 2014, 186, 203–215. [Google Scholar] [CrossRef] [PubMed]
  5. Sarkar, S.; Bhattacharjee, C.; Sarkar, S. Studies on the performance of annular photo reactor (APR) for pharmaceutical wastewater treatment. J. Water Process Eng. 2017, 19, 26–34. [Google Scholar] [CrossRef]
  6. Mirzaei, A.; Chen, Z.; Haghighat, F.; Yerushalmi, L. Removal of pharmaceuticals from water by homo/heterogonous Fenton-type processes—A review. Chemosphere 2017, 174, 665–688. [Google Scholar] [CrossRef]
  7. Zhan, J.; Li, Z.; Yu, G.; Pan, X.; Wang, J.; Zhu, W.; Han, X.; Wang, Y. Enhanced treatment of pharmaceutical wastewater by combining three-dimensional electrochemical process with ozonation to in situ regenerate granular activated carbon particle electrodes. Sep. Purif. Technol. 2019, 208, 12–18. [Google Scholar] [CrossRef]
  8. Li, L.; Jiang, L.; Yang, L.; Li, J.; Lu, N.; Qu, J. Optimization of Degradation Kinetics towards O-CP in H3PW12O40/TiO2 Photoelectrocatalytic System. Sustainability 2019, 11, 3551. [Google Scholar] [CrossRef] [Green Version]
  9. Chinh, V.D.; Broggi, A.; Di Palma, L.; Scarsella, M.; Speranza, G.; Vilardi, G.; Thang, P.N. XPS spectra analysis of Ti2+, Ti3+ ions and dye photodegradation evaluation of titania-silica mixed oxide nanoparticles. J. Electron. Mater. 2018, 47, 2215–2224. [Google Scholar] [CrossRef]
  10. Meiramkulova, K.; Jakupova, Z.; Orynbekov, D.; Tashenov, E.; Kydyrbekova, A.; Mkilima, T.; Inglezakis, V.J. Evaluation of electrochemical methods for poultry slaughterhouse wastewater treatment. Sustainability 2020, 12, 5110. [Google Scholar] [CrossRef]
  11. An, H.; Zhu, B.; Li, J.; Zhou, J.; Wang, S.; Zhang, S.; Wu, S.; Huang, W. Synthesis and characterization of thermally stable nanotubular TiO2 and its photocatalytic activity. J. Phys. Chem. C 2008, 112, 18772–18775. [Google Scholar] [CrossRef]
  12. Sohn, Y.S.; Smith, Y.R.; Misra, M.; Subramanian, V. Electrochemically assisted photocatalytic degradation of methyl orange using anodized titanium dioxide nanotubes. Appl. Catal. B Environ. 2008, 84, 372–378. [Google Scholar] [CrossRef]
  13. Kai, H.; Ishibashi, Y.; Mori, T.; Ishibashi, H.; Kawaguchi, I.; Ohwaki, H.; Takemasa, T.; Arizono, K. Decolorization and Estrogenic Activity of Colored Livestock Wastewater after Electrolysis Treatment. J. Mater. Cycles Waste Manag. 2010, 12, 128–135. [Google Scholar] [CrossRef]
  14. Liu, J.; Huang, L.; Li, S.; Feng, Y. Deep treatment efficiency of electrochemical catalytic oxidation for secondary effluent pharmaceutic. Chin. J. Environ. Eng. 2016, 10, 6269–6274. [Google Scholar]
  15. Fang, C.; Megharaj, M.; Naidu, R. Electrochemical advanced oxidation processes (EAOP) to degrade per- and polyfluoroalkyl substances (PFASs). J. Adv. Oxid. Technol. 2017, 20. [Google Scholar] [CrossRef]
  16. Yan, H.; Bai, Z.; Chao, S.; Cui, Q.; Niu, L.; Yang, L.; Qiao, J.; Jiang, K. Effects of additives on palladium nanocrystals supported on multiwalled carbon nanotubes and their electrocatalytic properties toward formic acid oxidation. Ionics 2014, 20, 259–268. [Google Scholar] [CrossRef]
  17. Guerrini, E.; Colombo, A.; Trasatti, S. Surface modification of RuO2 electrodes by laser irradiation and ion implantation: Evidence of electrocatalytic effects. J. Chem. Sci. 2009, 121, 639. [Google Scholar] [CrossRef]
  18. Nandi, S.; Nair, A.S.; Pathak, B. First principles investigation on the applicability of ruthenium as a potential ORR catalyst. J. Chem. Sci. 2020, 132, 132. [Google Scholar] [CrossRef]
  19. Jiang, N.; Wang, Y.; Zhao, Q.; Ye, Z. Application of Ti/IrO2 electrode in the electrochemical oxidation of the TNT red water. Environ. Pollut. 2020, 259, 113801. [Google Scholar] [CrossRef]
  20. Zhang, Y.; He, P.; Jia, L.; Zhang, T.; Liu, H.; Wang, S.; Zhou, S. Dimensionally stable Ti/SnO2-RuO2 composite electrode based highly efficient electrocatalytic degradation of industrial gallic acid effluent. Chemosphere 2019, 224, 707–715. [Google Scholar] [CrossRef]
  21. Pathiraja, G.C.; Jayathilaka, P.B.; Weerakkody, C.; Karunarathne, P.; Nanayakkara, N. Comparison study of dimensionally stable anodes for degradation of chlorpyrifos in water. Curr. Sci. 2014, 107, 219–226. [Google Scholar]
  22. Aguilar, Z.G.; Coreño, O.; Salazar, M.; Sirés, I.; Brillas, E.; Nava, J.L. Ti| Ir–Sn–Sb oxide anode: Service life and role of the acid sites content during water oxidation to hydroxyl radicals. J. Electroanal. Chem. 2018, 820, 82–88. [Google Scholar] [CrossRef]
  23. Zhang, Y.B.; Ma, Q.Q.; Feng, K.K.; Guo, J.; Wei, X.L.; Shao, Y.Q.; Zhuang, J.H.; Lin, T.S. Effects of microstructure and electrochemical properties of Ti/IrO2–SnO2–Ta2O5 as anodes on binder-free asymmetric supercapacitors with Ti/RuO2–NiO as cathodes. Ceram. Int. 2020, 46, 17640–17650. [Google Scholar] [CrossRef]
  24. Xin, Y.L.; Xu, L.K.; Wang, J.; Li, X. Effect of sintering temperature on microstructure and electrocatalytic properties of Ti/IrO2-Ta2O5 Anodes by pechini method. Rare Met. Mater. Eng. 2010, 39, 1903–1907. [Google Scholar]
  25. National Environmental Protection Agency. Water and Waste Water Monitoring and Analysis Method, 4th ed.; China Environmental Press: Beijing, China, 2002. [Google Scholar]
  26. Sun, W.; Ma, B.; Sun, Y.; Zheng, H.; Ma, G. Electrochemical degradation of tetracycline by γ-Al2O3-Bi-(Sn/Sb) three-dimensional particle electrode. Desalin. Water Treat. 2017, 98, 152–160. [Google Scholar] [CrossRef]
  27. Mukimin, A.; Vistanty, H.; Zen, N.; Purwanto, A.; Wicaksono, K.A. Performance of bioequalization-electrocatalytic integrated method for pollutants removal of hand-drawn batik wastewater. J. Water Process Eng. 2018, 21, 77–83. [Google Scholar] [CrossRef]
  28. Wang, H.; Wang, J.; Bo, G.; Wu, S.; Luo, L. Degradation of pollutants in polluted river water using Ti/IrO2-Ta2O5 coating electrode and evaluation of electrode characteristics. J. Clean. Prod. 2020, 273, 123019. [Google Scholar] [CrossRef]
  29. Lv, G.; Wu, D.; Fu, R. Performance of carbon aerogels particle electrodes for the aqueous phase electro-catalytic oxidation of simulated phenol wastewaters. J. Hazard. Mater. 2009, 165, 961–966. [Google Scholar] [CrossRef]
  30. Tijani, J.O.; Fatoba, O.O.; Madzivire, G.; Petrik, L.F. A review of combined advanced oxidation technologies for the removal of organic pollutants from water. Water Air Soil Pollut. 2014, 225, 2102. [Google Scholar] [CrossRef] [Green Version]
  31. Zhang, C.; Jiang, Y.; Li, Y.; Hu, Z.; Zhou, L.; Zhou, M. Three-dimensional electrochemical process for wastewater treatment: A general review. Chem. Eng. J. 2013, 228, 455–467. [Google Scholar] [CrossRef]
  32. Raschitor, A.; Llanos, J.; Canizares, P.; Rodrigo, M.A. (Novel integrated electrodialysis/electro-oxidation process for the efficient degradation of 2, 4-dichlorophenoxyacetic acid. Chemosphere 2017, 182, 85–89. [Google Scholar] [CrossRef] [PubMed]
  33. Sun, Y.; Zhu, S.; Sun, W.; Zheng, H. Degradation of high-chemical oxygen demand concentration pesticide wastewater by 3D electrocatalytic oxidation. J. Environ. Chem. Eng. 2019, 7, 103276. [Google Scholar] [CrossRef]
  34. Ventura, A.; Jacquet, G.; Bermond, A.; Camel, V. Electrochemical generation of the Fenton’s reagent: Application to atrazine degradation. Water Res. 2002, 36, 3517–3522. [Google Scholar] [CrossRef]
  35. Pangarkar, B.L.; Deshmukh, S.K.; Thorat, P.V. Multi-effect air gap membrane distillation process for pesticide wastewater treatment. Membr. Water Treat. 2017, 8, 529–541. [Google Scholar]
  36. Weisz, A.D.; Rodenas, L.G.; Morando, P.J.; Regazzoni, A.E.; Blesa, M.A. FTIR study of the adsorption of single pollutants and mixtures of pollutants onto titanium dioxide in water: Oxalic and salicylic acids. Catal. Today 2002, 76, 103–112. [Google Scholar] [CrossRef]
  37. Bi, Q. Investigation on Preparation and Properties of Electrode Materials for Electrochemical Oxidation of Organic Wastewater. Ph.D. Thesis, Xi’an University of Architecture and Technology, Xi’an, China, 2014. [Google Scholar]
  38. Wu, X.; Yang, X.; Wu, D.; Fu, R. Feasibility study of using carbon aerogel as particle electrodes for decoloration of RBRX dye solution in a three-dimensional electrode reactor. Chem. Eng. J. 2008, 138, 47–54. [Google Scholar] [CrossRef]
  39. Xiong, Y.; He, C.; Karlsson, H.T.; Zhu, X. Performance of three-phase three-dimensional electrode reactor for the reduction of COD in simulated wastewater-containing phenol. Chemosphere 2003, 50, 131–136. [Google Scholar] [CrossRef]
  40. Yan, L.; Ma, H.; Wang, B.; Wang, Y.; Chen, Y. Electrochemical treatment of petroleum refinery wastewater with three-dimensional multi-phase electrode. Desalination 2011, 276, 397–402. [Google Scholar] [CrossRef]
  41. Horáková, M.; Klementová, Š.; Kříž, P.; Balakrishna, S.K.; Špatenka, P.; Golovko, O.; Hájková, P.; Exnar, P. The synergistic effect of advanced oxidation processes to eliminate resistant chemical compounds. Surf. Coat. Technol. 2014, 241, 154–158. [Google Scholar] [CrossRef]
  42. Feng, Y.J.; Li, X.Y. Electro-catalytic oxidation of phenol on several metal-oxide electrodes in aqueous solution. Water Res. 2003, 37, 2399–2407. [Google Scholar] [CrossRef]
  43. García-Mancha, N.; Monsalvo, V.M.; Puyol, D.; Rodriguez, J.J.; Mohedano, A.F. Enhanced anaerobic degradability of highly polluted pesticides-bearing wastewater under thermophilic conditions. J. Hazard. Mater. 2017, 339, 320–329. [Google Scholar] [CrossRef] [PubMed]
  44. Vuorilehto, K.; Tamminen, A. Application of a solid ion-exchange electrolyte in three-dimensional electrodes. J. Appl. Electrochem. 1997, 27, 749–755. [Google Scholar] [CrossRef]
Figure 1. Experimental device.
Figure 1. Experimental device.
Sustainability 13 00126 g001
Figure 2. SEM images of (a) Ti and (b) Ti/RuO2-IrO2-SnO2 electrode.
Figure 2. SEM images of (a) Ti and (b) Ti/RuO2-IrO2-SnO2 electrode.
Sustainability 13 00126 g002
Figure 3. EDS spectra of Ti/RuO2-IrO2-SnO2 electrode.
Figure 3. EDS spectra of Ti/RuO2-IrO2-SnO2 electrode.
Sustainability 13 00126 g003
Figure 4. XRD patterns of Ti/RuO2-IrO2-SnO2 electrode.
Figure 4. XRD patterns of Ti/RuO2-IrO2-SnO2 electrode.
Sustainability 13 00126 g004
Figure 5. (a) Effect of electrode distance and (b) energy consumption at the current density of 1 mA/cm2, air flow of 12 L/min, pH of 7.
Figure 5. (a) Effect of electrode distance and (b) energy consumption at the current density of 1 mA/cm2, air flow of 12 L/min, pH of 7.
Sustainability 13 00126 g005
Figure 6. (a) Effect of current density and (b) energy consumption at the electrode distance of 3 cm, air flow of 12 L/min, pH of 7.
Figure 6. (a) Effect of current density and (b) energy consumption at the electrode distance of 3 cm, air flow of 12 L/min, pH of 7.
Sustainability 13 00126 g006
Figure 7. (a) Effect of air flow and (b) energy consumption at the electrode distance of 3 cm, current density of 8 mA/cm2, pH of 7.
Figure 7. (a) Effect of air flow and (b) energy consumption at the electrode distance of 3 cm, current density of 8 mA/cm2, pH of 7.
Sustainability 13 00126 g007
Figure 8. (a) Effect of initial pH and (b) energy consumption at the electrode distance of 3 cm, current density of 8 mA/cm2, air flow of 18 L/min.
Figure 8. (a) Effect of initial pH and (b) energy consumption at the electrode distance of 3 cm, current density of 8 mA/cm2, air flow of 18 L/min.
Sustainability 13 00126 g008
Figure 9. SEM images of (a) Ti plate and (b) Ti/RuO2-IrO2-SnO2 electrode after reaction.
Figure 9. SEM images of (a) Ti plate and (b) Ti/RuO2-IrO2-SnO2 electrode after reaction.
Sustainability 13 00126 g009
Figure 10. EDS of Ti/RuO2-IrO2-SnO2 electrode after reaction.
Figure 10. EDS of Ti/RuO2-IrO2-SnO2 electrode after reaction.
Sustainability 13 00126 g010
Table 1. The composition of pharmaceutical wastewater.
Table 1. The composition of pharmaceutical wastewater.
CompoundMolecular FormulaConcentrationUnit
1TetrahydrofuranC4H8O303.438mg/L
2ethyl acetateC4H8O24179.921mg/L
3ethanolC2H6O18,016.768mg/L
Table 2. Element content of Ti/RuO2-IrO2-SnO2 electrode.
Table 2. Element content of Ti/RuO2-IrO2-SnO2 electrode.
ElementCOTiRuIrSn
wt%9.8642.0327.378.714.567.47
Table 3. Element content of Ti/RuO2-IrO2-SnO2 electrode after reaction.
Table 3. Element content of Ti/RuO2-IrO2-SnO2 electrode after reaction.
ElementCOTiRuIrSn
wt%54.8522.6515.521.632.752.60
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, G.; Huang, X.; Ma, J.; Wu, F.; Zhou, T. Ti/RuO2-IrO2-SnO2 Anode for Electrochemical Degradation of Pollutants in Pharmaceutical Wastewater: Optimization and Degradation Performances. Sustainability 2021, 13, 126. https://doi.org/10.3390/su13010126

AMA Style

Zhang G, Huang X, Ma J, Wu F, Zhou T. Ti/RuO2-IrO2-SnO2 Anode for Electrochemical Degradation of Pollutants in Pharmaceutical Wastewater: Optimization and Degradation Performances. Sustainability. 2021; 13(1):126. https://doi.org/10.3390/su13010126

Chicago/Turabian Style

Zhang, Guozhen, Xingxing Huang, Jinye Ma, Fuping Wu, and Tianhong Zhou. 2021. "Ti/RuO2-IrO2-SnO2 Anode for Electrochemical Degradation of Pollutants in Pharmaceutical Wastewater: Optimization and Degradation Performances" Sustainability 13, no. 1: 126. https://doi.org/10.3390/su13010126

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