| Home | E-Submission | Sitemap | Contact Us |  
Environ Eng Res > Volume 28(5); 2023 > Article
Wei, Sha, and Wang: Study on Treatment of Basic Yellow 28 dye Wastewater by Micro-nano Bubble Ozone Catalytic Oxidation

Abstract

Mn loaded AC (Mn/AC), Cu loaded AC (Cu/AC), and Fe loaded AC (Fe/AC) catalysts were prepared by the impregnation-calcination method, which can be used to degrade Basic Yellow 28 dyes in simulated printing and dyeing wastewater for catalytic ozonation. The catalyst preparation and reaction conditions were optimized with the degradation efficiency of chemical oxygen demand (COD) and the absorbance of some organic compounds under 254 nm UV light (UV254) as indicators. The surface structure, specific surface area, and active component distribution of the catalysts were analyzed by scanning electron microscopy (SEM), nitrogen adsorption, and X-ray diffraction (XRD) characterization. Results indicated that Mn was the best active component, and a catalyst dosage of 3.5 g/L, an initial pH of 9.4, and an initial temperature of 31.7°C were the best reaction parameters to degrade Basic Yellow 28 dyes. After five recycling experiments, the catalysts could retain a relatively stable catalytic effect. Furthermore, this study can be informative for the sustainable development of future catalysts.

1. Introduction

Ozone oxidation is a typical advanced oxidation method. The ozone-activated advanced oxidation process often generates ·OH, O2 and 1O2 [1, 2]. Although O3 has a high reduction potential (E=2.07V) [1], oxidation by ozone alone is highly selective and it only attacks the nucleophilic sites of the aromatic ring due to ozone’s dipolar, electrophilic, and nucleophilic nature [1, 3]. In recent years, the heterogeneous catalytic ozonation has been proposed to overcome these problems. The use of catalysts in this technology can increase the generation of ·OH during the ozonation process [2]. This technology can minimize the loss of active components of the catalyst and ensure higher stability and no secondary pollution for the catalyst [4]. O3 can produce ·OH with a redox potential of 2.80 V in the presence of catalysts. ·OH is an ion with a strong oxidizing property and is effective in degrading organic matters [5].
Activated carbon, molecular sieves, zeolites, and Al2O3 have been used as supporters for catalytic ozonation [4]. Activated carbon has become one of the most widely used supporters in catalytic ozonation due to its easy availability, large surface area, high stability, and good adsorption performance [6, 7]. Several metal oxides have been investigated as the active components for ozonation such as Mn, Fe and Cu [810]. Mn and Fe exist in multiple valence states, which offer excellent catalytic properties [1113]. Cu is also a common and effective precursor in catalytic applications [14, 15]. During the catalytic ozonation process, catalysts can accelerate the generation of free active radicals by ozone and enhance the kinetics of pollutant conversion by ozone [1618]. Many previous studies have indicated that compared with ozonation alone, metal oxides supported on activated carbon can enhance the pollutant removal efficiency [1922].
Furthermore, the commercial application of ozone has been limited due to its small specific surface area, low dissolved ozone concentration and low partial pressure [23]. A gas-liquid mixing pump can be used to disperse ozone into numerous micro-nano bubbles (MNBs) to enhance the mass transfer rate of ozone [23, 24]. Unlike the conventional fine-bubble aeration, the micro-nano bubble-aerated systems (MNBAs) possess more advantages such as higher O3 stability, higher gas-liquid mass transfer efficiency, and larger specific surface areas [25]. MNBs can increase the mass transfer efficiency of ozone significantly and thus enhance the treatment efficiency [26, 27]. In addition, the bubbles generated by MNBAs could remain stable in water over a period of days or months owing to the high zeta potential [28, 29], and then promote the proliferation of reactive oxygen species (ROS), such as hydroxyl radicals (·OH), and superoxide radical anion ( O2-·), through ozone decomposition [30, 31]. Studies have shown that the O3/MBAs process removes organic matters at a higher efficiency than ozonation alone, which can be ascribed to accelerated decomposition of ozone and enhanced generation of hydroxyl radicals during the collapse of ozone microbubbles [32, 33]. Therefore, the MNBA technology can be widely used in catalytic ozonation. However, very few studies have focused on the combination of the MNBA technology and heterogeneous catalytic ozonation along with the use of metal-loaded activated carbon as catalysts [34].
In this work, we combined O3/MNBA technology and catalysts to degrade Basic Yellow 28 dye which are typical cationic dyes with benzene ring and unsaturated structure and are widely used in industry [35]. Then we optimized the reaction conditions in order to provide a reference for the efficient application of the catalysts. We also investigated the stability of the catalysts during the reaction process.

2. Experimental

2.1. Materials

The Mn(NO3)2, Cu(NO3)2, Fe(NO3)2, and activated carbon used for the preparation of catalysts were purchased from Shenyang Shenglongfu Experimental Equipment Co., Ltd. and Shenyang Dongxing Reagent Factory; the Basic Yellow 28 dyes used for the preparation of experimentally simulated wastewater were obtained from Hangzhou Advance Technology Co. Ltd. Additional information on other chemicals and reagents is shown in Table S1. All chemicals and reagents were analytically pure. The water used throughout the whole work was deionized.

2.2. Preparation of Catalysts

The catalysts were prepared by the impregnation-calcination method. First, 6–8 mesh granular activated carbon was washed with pure water 2–3 times. Next, a certain amount of pretreated granular activated carbon was dipped into the impregnation solution. After being stirred for 6 h at room temperature, the impregnated activated carbon was first steamed dried in a water bath at 80 °C and then dried in an oven at 120 °C. Finally, the dried activated carbon was roasted in a muffle furnace at a specific temperature for a while to obtain the catalysts.

2.3. Characterization of Catalysts

The surface crystal structure of the catalysts was analyzed by an X-ray diffractometer (XRD-7000, Shimadzu, Japan) using Cu K α radiation. Scanning electron microscopy (SEM) images of the catalysts were obtained through a high-energy electron beam using a cold field emission scanning electron microscope (S-4800, HITACHI, Japan). The pore structure and specific surface area of the catalysts were measured by a Brunauer-Emmett-Teller (BET) physical adsorption instrument (Autosorb-IQ, Conta, USA). Nitrogen was used as the test gas for adsorption. The samples were dried and degassed under vacuum at 300 °C before testing.

2.4. Catalytic Ozonation Experiment

Pressurized dissolved gas was used to produce MNBs. The experimental setup of catalytic ozonation is schematically illustrated in Fig. S1. An O3 generator (Shijiazhuang Ouneng General Technology Co., Ltd.) was used to generate O3. A gas-liquid mixing pump (model 20QY-1, power 0.55 kW–7.5kW, gas-liquid volume ratio: 11%) mixed the sucked wastewater and O3, pressed the mixture into a dissolved gas tank through a connecting pipe and then into a reactor through an aeration disk. This process generated MNBs 20–30 μm in diameter [36]. The experimental reactor was a plastic tank with an effective working volume of 20 L. The catalysts were fixed in the middle of the reactor. At 15-min intervals, a water sample was taken from the sampling port at the bottom of the reactor. During the experiment, the residual O3 was quenched by 10 % KI.

2.5. Analytical Methods

The pH of the wastewater was determined by a composite glass electrode (SX711, San-Xin Instrumentation Factory, Shanghai). The COD of the wastewater was measured using the rapid closed catalytic digestion method. The concentration of O3 in the solution was measured using the sodium indigo disulfonate (IDS) spectrophotometry method. The point of zero charge (pHpzc) of the catalysts was determined by the salt titration method [37]. The content of the metal in the solution was measured by ICP-MS. A UV-visible spectrophotometer (UV-5500, Yuan Analysis Instrument Co., Ltd., Shanghai) was used to determine the absorbance of water samples at 440 nm, which is the maximum absorption wavelength of the Basic Yellow 28 dyes. UV254 is a parameter representing the amount of carbon source in wastewater, which directly reflects the ability of ozone to mineralize organic pollutants.

3. Results and Discussion

3.1. Comparation of Different Aeration Systems

The experiment was intended to discuss the performance of O3-Micro-nano Bubbles (O3-MNBs) and O3-Ordinary Bubbles (O3-OBs) concretely. The ozone inlet flowrate and concentration were set at 3 L/min and 10.3 mg/L, respectively, to serve as the reaction conditions in the subsequent experiments.
The effects of O3-MNBs and O3-OBs on the liquid phase ozone concentration and ozone mass transfer coefficient were investigated. We assumed that the dissolution of ozone in water can be described by a mass transfer coefficient equation of gas in solution (based on Eqs (1)(2)) [38]:
(1)
dCdt=Kla(Cs-C)
(2)
-Klat=ln(Cs-C)
where Kla, Cs, and Ct are the ozone mass transfer coefficient, ozone saturated concentration, and bulk concentration in water, respectively. The results are shown in Fig. 1a and b. Clearly, the O3-MNBs exhibited a higher liquid phase ozone concentration and a higher ozone mass transfer efficiency than the O3-OBs. According to Young–Laplace equation, MNBs have a higher internal pressure than ordinary bubbles, leading to the dissolution of more ozone in water [28, 39]. As described in a previous study, more ozone can be dissolved into water through the bubble interface because of the larger total surface area of the MNBs [40]. The ozone mass transfer coefficient of the O3-MNBs was 1.77 times higher than that of the O3-OBs, which agreed with the conclusion in the literature [41, 42]. Studies have shown that the gas-liquid transfer efficiency of ozone was effectively improved due to the self-pressurization effect of MNBs [40, 43]. It can be concluded from Fig. 1c that the efficiency of COD removal by O3/MNBAs was approximately 3.1 and 1.7 times higher than that by MNBAs alone and by ozonation alone, respectively. This could be attributed to the fact that O3/MNBAs produced a higher concentration of ·OH in solution, which corresponded to a higher COD removal rate. Nam et al. demonstrated that microbubble ozonation performed particularly well in benzo pyrene degradation due to the higher ·OH concentration [44]. Therefore, the O3/MNBA technology was applied in this work.

3.2. Comparation of Different Catalyst Systems

Different roasting temperatures (400 °C, 500 °C, and 600 °C), roasting times (2 h, 3 h, and 4 h), and impregnation solution concentrations (0.5 mol/L, 1.0 mol/L, and 2.0 mol/L) were set to optimize the preparation conditions of the catalysts. To degrade 20 L of simulated dye wastewater with a concentration of 100 mg/L, 2 L/min ozone dosage, 10.47 mg/L ozone concentration, and 1 g/L catalyst dosage or activated carbon were used as the reaction conditions.
As shown in Table 1, the optimal preparation conditions for Mn/GAC, Cu/GAC, and Fe/GAC were 500 °C, 500 °C, and 400 °C roasting times, 3 h, 4 h, and 4 h roasting times, and 0.5 mol/L, 0.5 mol/L, and 1.0 mol/L impregnation solution concentrations, respectively. According to the COD removal results, the Mn/GAC catalyst showed a better treatment performance than the other two catalysts.

3.3. Characterization of Catalysts

The three catalysts with the best removal performance were characterized and analyzed to observe the surface morphology and crystal structure. The SEM images of the unmodified activated carbon and the catalysts are provided in Fig. S2a, b, c, and d. Unlike the unmodified activated carbon (Fig. S2a), the surface of the Mn-nitrate-modified activated carbon (Fig. S2b) was covered by lamellar particles, which may be the typical morphology of MnO2 as reported in the literature [45, 46]. From Table S2, it can be known that the activated carbon showed a decrease in its specific surface area, pore volume, and pore size after being loaded with three metals. In addition, the pore structure may change and recombine after calcination [47]. This helps to ensure the successful loading of metal oxides on the surface and in the pore channels of the activated carbon [48]. It is worth noting that the activated carbon modified by iron nitrate showed a sharp decrease in its specific surface area, pore volume, and pore size. This could be attributed to the fact that the iron oxide particles were mainly distributed on the surface of the supporters in the form of metal clusters (Fig. S2d) [47, 49]. After modification by manganese nitrate and copper nitrate, the activated carbon did not show apparent change in its specific surface area, but its COD removal rate changed significantly.
Fig. 2 shows the effects of different reaction systems on the COD removal efficiency. When activated carbon was used as an adsorbent, the COD removal efficiency was only 2.4%. This suggested that activated carbon offered very limited adsorption for the removal of contaminants in the system. Compared with ozonation alone, adding activated carbon to the ozonation system could increase the efficiency of COD removal by 16.11 %, implying that activated carbon could promote the transformation of aqueous ozone into more[50]. The figure also provides a clear illustration of the performance of Mn/AC, Cu/AC, and Fe/AC catalysts in the removal of COD from dye wastewater. During the reaction process, all these catalysts showed good performance in COD treatment, with Mn/AC performing the best of all. Mn is a particular transition metal oxide with valence states from +2 to +7, compared to Cu or Fe with a maximum formal oxidation state of +4 or +3. This advantage promotes the participation of manganese in the electron transfer reaction with O3 during the catalytic cycle [51]. After data analysis and catalyst characterization, Mn/AC was finally selected as the optimal catalyst.
The XRD patterns of activated carbon and Mn/AC are presented in Fig. S3. The unmodified activated carbon exhibited characteristic diffraction peaks of SiO2 and CaCO3, which indicated that the activated carbon contained a small amount of SiO2 and CaCO3 impurities [52]. After modification by manganese nitrate, characteristic diffraction peaks at 12.80°, 18.10°, and 32.88° were observed, which might be the crystalline phases of α-MnO2 [53]. The intensity of the diffraction peaks of the activated carbon became weaker, indicating that the diffraction peaks of the activated carbon decreased after loading of manganese onto the activated carbon. It has also been demonstrated that MnO2 [54] has the best catalytic treatment performance among the various valence states of manganese oxides.

3.4. Optimization of Process Parameters

Given its good catalytic performance, Mn/AC was used in this part of the experiment, and the operating conditions were optimized by investigating the basic parameters (catalyst dosage, pH, and reaction temperature).

3.4.1. Effect of catalyst dosage

Both Brønsted and Lewis acid sites are considered to be the active centers for catalytic ozonation [55, 56]. Water molecules can be adsorbed on manganese oxides to combine with Lewis acid sites, and then dissociate into protonated surface hydroxyl groups. Lots of research has shown that neutral hydroxyl groups and protonated hydroxyl groups are the active sites for catalytic ozonation [57]. Hence, the catalyst dosage affects the number of active sites indirectly. The effect of catalyst dosage in the range of 1 to 5 g/L on the COD and UV254 removal rates of the Mn/AC system was investigated. The ozone dosage, pH, reaction temperature, and continuous aeration time were set at 2 L/min, 7.1, 25 °C, and 120 min, respectively. As shown in Fig. 3, the COD and UV254 removal rates increased significantly when the catalyst dosage increased from 1.0 g/L to 3.0 g/L. This suggested that the addition of the catalyst enhanced the degradation efficiency of Basic Yellow 28 dyes. This may be explained as follows. As the catalyst dosage increased, the total specific surface area and the number of active sites of the catalyst were increased as well. This could facilitate the reaction with O3 to promote the generation of ·OH [58, 59], thus improving the COD and UV254 removal rates. From the stabilized COD and UV254 removal rates, it can be deduced that excessive catalyst dosage hindered the generation of ·OH, which converted into without catalytic ability. Besides, as the number of O3 molecules per unit volume of the solution was fixed, increasing the catalyst dosage failed to make the O3 in the bubbles come into sufficient contact with the active sites on the catalyst surface [41, 60]. It can also be seen that the COD and UV254 removal rates were not linear related with the increase of catalyst dosage, but there existed a critical value [60]. Therefore, from the perspectives of both cost efficiency and treatment performance, 3.0 g/L was used as the optimal catalyst dosage in the subsequent experiments.

3.4.2. Effect of initial pH of solution

The performance of heterogeneous catalytic oxidation is greatly affected by environmental factors, and as reported in the literature, it is particularly affected by the pH of the solution. On the one hand, the pH of the solution affects the oxidation mode and the decomposition rate of O3 [57]. For instance, based on Eqs.(3)(8) [61], OH in the solution will decompose O3. Nevertheless, excessive OH will inhibit the generation of ·OH.
(3)
O3+OH-O2+HO2-
(4)
HO2-+O3HO2·+O3-·
(5)
HO2·O2-·+H+
(6)
O2-·+O3O3-+O2
(7)
O3-O2+O-·
(8)
O-·+H2OOH-+·OH
On the other hand, the pH of the solution affects the charged state of the catalyst surface, which can be expressed by the metal oxide proton balance equations [62] (Eqs.(9)(10)).
(9)
MeOH+H+MeOH2+(pH<pHpzc)
(10)
MeOH+OH-MeO-+H2O (pH>pHpzc)
We investigated how the initial pH values 5.0, 7.1, 9.4, and 10.8 affect the COD and UV254 removal rates of the Mn/AC/O3 system. The ozone dosage, catalyst dosage, reaction temperature, and continuous aeration time were set at 2 L/min, 3.0 g/L, 25 °C, and 120 min, respectively. As shown in Fig. 4, when the pH of the solution increased in the range of 5.0 to 9.4, it could increase the rate of COD removal from dye wastewater. Therefore, it could be concluded that the increase of the pH in this range could increase the concentration of in the solution, which promoted the generation of ·OH. After 120 min of continuous aeration, when the initial pH was 9.4, the removal efficiencies of COD and UV254 reached 73.6% and 79.3%, respectively, compared with the initial pH of 5.0, COD and UV254 removal rates soared about 12.9% and 9.8%, respectively. First of all, the pH at the potential of zero-point charge (pHpzc) of the catalyst was found to be approximately 6.54 lower than the pH of 9.4 in the solution. In this case, the catalyst surface was negatively charged and showed good catalytic performance [63]. This was conducive to the adsorption of the Basic Yellow 28 dyes with a positive charge. As reported in the literature, MnOx/SBA-15 can adsorb organic matters to combine with the hydroxyl groups on the catalyst surface to facilitate further oxidization and decomposition by ·OH in the solution [64]. It can be seen from Fig. 5, the catalysts were more stable under alkaline conditions and retained more manganese. The manganese caused the formation of oxygen vacancies, which adsorbed more pollutants to accelerate the ozonation process [65, 66]. Apart from this, ·OH could also facilitate the decomposition of O3 to produce more ·OH. However, it is worth noting that when the pH exceeded 9.4, the COD and UV254 removal rates declined sharply. This could be attributed to that excessive OH consumed the ·OH (Eq(8)) and competed with the catalyst to adsorb O3, which inhibited the activity of the catalyst and thus affected the COD and UV254 removal rates. The above-mentioned results demonstrated that the pH of the solution had a significant effect on the removal performance. It could affect both the decomposition rate of O3 and the charged state of the catalyst surface. Moreover, it could also affect the surface properties of metal oxides covered by surface hydroxyl groups [67, 68]. The COD and UV254 degradation efficiencies obtained under alkaline conditions were much higher than under acidic conditions. Given these findings, the initial pH of 9.4 was used in the subsequent experiments.

3.4.3. Effect of initial temperature of solution

We investigated how the initial solution temperatures 12.5°C, 23. 2°C, and 31.7°C affect the COD and UV254 removal rates of the Mn/AC/O3 system. The subsequent experiment was carried out under the optimum reaction conditions. From Fig. 6, it can be observed that COD and UV254 degradation efficiencies increased slowly with the increase of reaction temperature. The reaction temperature affected the catalytic ozonation process from two perspectives: the activation energy of the reaction and the solubility of O3 in the solution. The gradually increased reaction temperature could cause the catalysts to reduce the activation energy of the reaction. This could accelerate the chemical reaction between the catalysts, O3, and pollutants. This could also promote the decomposition of O3 molecules in the dye wastewater, which in turn accelerated the generation of ·OH to degrade organic matters. However, with the increase in the initial temperature of the dye wastewater, the solubility of O3 in the wastewater would decrease [6971]. This could hinder the conversion of O3 from the gas phase to the liquid phase and affect the concentration of O3 and ·OH, thus failing to facilitate the COD and UV254 degradation efficiencies. It was also found that the removal rates varied little among the three solution temperatures, indicating that the initial temperature of the solution had little effect on the ozone-catalyzed oxidation system. However, the experimental results suggested that the initial temperature of the solution had a positive effect on the catalytic ozonation for the degradation of Basic Yellow 28 dyes. In other words, the increase in reaction temperature played a greater role in the increase in reaction rate than in the decrease in the solubility of O3.

3.5. The Stability of Catalysts

To evaluate the stability of the Mn/AC catalyst, it was reused five times under the same conditions. As shown in Fig. 7, as the catalyst was used for an increased number of times, the COD and UV254 removal efficiency decreased, but to a small extent. After the catalyst was used for the fifth time, the leaching percentage of Mn was only 1.4%, which could have little effect on the activity of the catalyst. These results demonstrated that the Mn/AC catalyst remained stable and effective during the reaction process.

4. Conclusion

In this study, three catalysts were prepared by the impregnation-calcination method under optimized conditions. The Mn/AC catalyst obtained under the optimal conditions (500 °C roasting temperature, 3h roasting time, and 0.5 mol/L impregnation solution concentration) achieved the best COD removal efficiency. The characterization of the catalysts revealed that the Mn/AC catalyst had better surface characteristics and pore structure than the other two. Our investigation showed that different reaction conditions, i.e., catalyst dosage, initial pH of the solution, and initial temperature of solution, had different effects on the treatment performance of Basic Yellow 28 dye wastewater. During the reaction process, it was demonstrated that the increasing catalyst dosage could enhance the active sites in the system. For a good balance between the removal rates of COD and UV254, 3.0 g/L was determined as the optimal catalyst dosage. An appropriate alkalinity could not only promote the decomposition of O3 but also improve the surface characteristics of the catalysts. Accordingly, the initial pH 9.4 was beneficial to promoting catalytic ozonation. Besides, the rising reaction temperature played a positive role in catalytic ozonation. A possible area of future research would be to investigate methods to improve the COD and UV254 degradation efficiencies by loading multiple active components onto the supporter. Furthermore, it is also important to go beyond laboratory models to explore the practical functionality and applicability of the catalysts.

Supplementary Information

Acknowledgements

We acknowledge the financial support provided by the Major Science and Technology Project on Water Pollution Control and Treatment (2018ZX0760100401) and the Liaoning Provincial Department of Education Project (lnjc202011).

Notes

Conflict-of-Interest Statement

The authors declare no competing interests.

Author Contributions

W.R.N. (student) conducted the experiments. S.H.W. (student) conducted the experiments, wrote and revised the manuscript. W.J. (associate professor) revised the manuscript.

Reference

1. Liu HB, Gao Y, Wang J, Pan JW, Gao BY, Yue QY. Catalytic ozonation performance and mechanism of Mn-CeOx@γ-Al2O3/O3 in the treatment of sulfate-containing hypersaline antibiotic wastewater. Sci. Total Environ. 2022;807:150867. https://doi.org/10.1016/j.scitotenv.2021.150867.
crossref

2. Wang JL, Wang SZ. Reactive species in advanced oxidation processes: Formation, identification and reaction mechanism. Chem. Eng. J. 2020;401:126158. https://doi.org/10.1016/j.cej.2020.126158.
crossref

3. Stock NL, Peller J, Vinodgopal K, Kamat PV. Combinative sonolysis and photocatalysis for textile dye degradation. Environ. Sci. Technol. 2000;34:1747–1750. https://doi.org/10.1021/es991231c.
crossref

4. Jiang HB, Zhang R, Hao JL, et al. Design, preparation, characterization, and application of MnxCu1−xOy/γ-Al2O3 catalysts in ozonation to achieve simultaneous organic carbon and nitrogen removal in pyridine wastewater. Sci. Total Environ. 2021;774:145189. https://doi.org/10.1016/j.scitotenv.2021.145189.
crossref pmid

5. Martins RC, Quinta-Ferreira RM. Remediation of phenolic wastewaters by advanced oxidation processes (AOPs) at ambient conditions: comparative studies. Chem. Eng. Sci. 2011;66:3243–3250. https://doi.org/10.1016/j.ces.2011.02.023
crossref

6. Ahmadi M, Kakavandi B, Jaafarzadeh N, Babaei AA. Catalytic ozonation of high saline petrochemical wastewater using PAC@FeIIFe2IIIO4: optimization, mechanisms and biodegradability studies. Sep. Purif. Technol. 2017;177:293–303. https://doi.org/10.1016/j.seppur.2017.01.008.
crossref

7. Rekhate CV, Srivastava J. Recent advances in ozone-based advanced oxidation processes for treatment of wastewater-A review. Chem. Eng. J. Adv. 2020;3:100031. https://doi.org/10.1016/j.ceja.2020.100031.
crossref

8. Cheng J, Xie YR, Wei Y, et al. Degradation of tetracycline hydrochloride in aqueous via combined dielectric barrier discharge plasma and Fe-Mn doped AC. Chemosphere. 2022;286:131841. https://doi.org/10.1016/j.chemosphere.2021.131841.
crossref pmid

9. Hu EL, Wu XB, Shang SM, Tao XM, Jiang SX, Gan L. Catalytic ozonation of simulated textile dyeing wastewater using mesoporous carbon aerogel supported copper oxide catalyst. J. Clean Prod. 2016;112:4710–4718. https:/doi.org/10.1016/j.jclepro.2015.06.127.
crossref

10. Nawrocki J. Catalytic ozonation in water: controversies and questions. Discussion paper. Appl. Catal. B. 2013;142:465–471. https://doi.org/10.1016/j.apcatb.2013.05.061.
crossref

11. Wu ZW, Zhang GQ, Zhang RY, Yang FL. Insights into mechanism of catalytic ozonation over practicable mesoporous Mn-CeOx/γ-Al2O3 catalysts. Ind. Eng. Chem. Res. 2018;57:1943–1953. https:/doi.org/10.1021/acs.iecr.7b04516.
crossref

12. Sui MH, Liu J, Sheng L. Mesoporous material supported manganese oxides (MnOx/MCM-41) catalytic ozonation of nitrobenzene in water. Appl. Catal. B. 2011;106:195–203. https://doi.org/10.1016/j.apcatb.2011.05.025.
crossref

13. Tian SQ, Qi JY, Wang YP, Liu YL, Wang L, Ma J. Heterogeneous catalytic ozonation of atrazine with Mn-loaded and Fe-loaded biochar. Water Res. 2021;193:116860. https://doi.org/10.1016/j.watres.2021.116860.
crossref pmid

14. de Oliveira SB, Barbosa DP, de Melo Monteiro AP, Rabelo D, do Carmo Rangel M. Evaluation of copper supported on polymeric spherical activated carbon in the ethylbenzene dehydrogenation. Catal. Today. 2008;133:92–98. https://doi.org/10.1016/j.cattod.2007.12.040.
crossref

15. Pimentel M, Oturan N, Dezotti M, Oturan MA. Phenol degradation by advanced electrochemical oxidation process electro-Fenton using a carbon felt cathode. Appl. Catal. B. 2008;83:140–149. https://doi.org/10.1016/j.apcatb.2008.02.011.
crossref

16. Huang YX, Cui CC, Zhang DF, Li L, Pan D. Heterogeneous catalytic ozonation of dibutyl phthalate in aqueous solution in the presence of iron-loaded activated carbon. Chemosphere. 2015;119:295–301. https://doi.org/10.1016/j.chemosphere.2014.06.060.
crossref pmid

17. Abdedayem A, Guiza M, Ouederni A. Copper supported on porous activated carbon obtained by wetness impregnation: effect of preparation conditions on the ozonation catalyst's characteristics. C. R. Chim. 2015;18:100–109. https://doi.org/10.1016/j.crci.2014.07.011.
crossref

18. Rao YF, Luo HJ, Wei CH, Luo LF. Catalytic ozonation of phenol and oxalic acid with copper-loaded activated carbon. J. Cent. South Univ. 2010;17:300–306. https://doi.org/10.1007/s11771-010-0046-y.
crossref

19. Ma J, Sui MH, Chen ZL, Wang LN. Degradation of refractory organic pollutants by catalytic ozonation—activated carbon and Mn-loaded activated carbon as catalysts. Ozone Sci. Eng. 2004;26:3–10. https://doi.org/10.1080/01919510490426027.
crossref

20. Wu GP, Jeong TS, Won CH, Cui LZ. Comparison of catalytic ozonation of phenol by activated carbon and manganese-supported activated carbon prepared from brewing yeast. Korean J. Chem. Eng. 2010;27:168–173. https://doi.org/10.1007/s11814-009-0337-x.
crossref

21. Huang F, Luo MH, Cui LZ, Wu GP. Catalytic ozonation of methylene blue in aqueous solution by loading transition metal (Co/Cu/Fe/Mn) on carbon. Korean J. Chem. Eng. 2015;32:268–273. https://doi.org/10.1007/s11814-014-0238-5.
crossref

22. Chen CM, Chen HS, Guo X, Guo SH, Yan GX. Advanced ozone treatment of heavy oil refining wastewater by activated carbon supported iron oxide. J. Ind. Eng. Chem. 2014;20:2782–2791. https://doi.org/10.1016/j.jiec.2013.11.007.
crossref

23. Quan XJ, Luo D, Wu J, Li RH, Cheng W, Ge SP. Ozonation of acid red 18 wastewater using O3/Ca(OH)2 system in a micro bubble gas-liquid reactor. J. Environ. Chem. Eng. 2017;5:283–291. https://doi.org/10.1016/j.jece.2016.12.007.
crossref

24. Khuntia S, Majumder SK, Ghosh P. Catalytic ozonation of dye in a microbubble system: hydroxyl radical contribution and effect of salt. J. Environ. Chem. Eng. 2016;4:2250–2258. https://doi.org/10.1016/j.jece.2016.04.005.
crossref

25. Rojviroon O, Rojviroon T. Photocatalytic process augmented with micro/nano bubble aeration for enhanced degradation of synthetic dyes in wastewater. Water Resour. Ind. 2022;27:100169. https://doi.org/10.1016/j.wri.2021.100169.
crossref

26. Khuntia S, Majumder SK, Ghosh P. Removal of ammonia from water by ozone microbubbles. Ind. Eng. Chem. Res. 2013;52:318–326. https://doi.org/10.1021/ie302212p.
crossref

27. Hu LM, Xia ZR. Application of ozone micro-nano-bubbles to groundwater remediation. J. Hazard. Mater. 2018;342:446–453. https://doi.org/10.1016/j.jhazmat.2017.08.030.
crossref pmid

28. Wu JJ, Zhang KJ, Cen C, Wu XG, Mao RY, Zheng YY. Role of bulk nanobubbles in removing organic pollutants in wastewater treatment. AMB. Express. 2021;11:1–13. https://doi.org/10.1186/s13568-021-01254-0.
crossref pmid pmc

29. Agarwal A, Ng WJ, Liu Y. Principle and applications of microbubble and nanobubble technology for water treatment. Chemosphere. 2011;84:1175–1180. https://doi.org/10.1016/j.chemosphere.2011.05.054.
crossref pmid

30. Takahashi M, Chiba K, Li P. Free-radical generation from collapsing microbubbles in the absence of a dynamic stimulus. J. Phys. Chem. B. 2007;111:1343–1347. https://doi.org/10.1021/jp0669254.
crossref pmid

31. Liu S, Oshita S, Kawabata S, Makino Y, Yoshimoto T. Identification of ROS produced by nanobubbles and their positive and negative effects on vegetable seed germination. Langmuir. 2016;32:11295–11302. https://doi.org/10.1021/acs.langmuir.6b01621.
crossref pmid

32. Xia ZR, Hu LM. Treatment of organics contaminated wastewater by ozone micro-nano-bubbles. Water. 2018;11:55. https://doi.org/10.3390/w11010055.
crossref

33. Zheng TL, Zhang T, Wang QH, et al. Advanced treatment of acrylic fiber manufacturing wastewater with a combined microbubble-ozonation/ultraviolet irradiation process. RSC Adv. 2015;5:77601–77609. https://doi.org/10.1039/c5ra14575a.
crossref

34. Zhang J, Huang GQ, Liu C, Zhang RN, Chen XX, Zhang L. Synergistic effect of microbubbles and activated carbon on the ozonation treatment of synthetic dyeing wastewater. Sep. Purif. Technol. 2018;201:10–18. https://doi.org/10.1016/j.seppur.2018.02.003.
crossref

35. Lorimer JP, Mason TJ, Plattes M, Phull SS, Walton DJ. Degradation of dye effluent. Pure Appl. Chem. 2001;73:1957–1968. https://doi.org/10.1351/pac200173121957.
crossref

36. Nanyuan Pump Industry Corporation. Self-priming gas-liquid mixing pump [Internet]. c2019. [cited 20 October 2022]. Available from: https://www.nyp-pump.cn/product/557.html


37. Mustafa S, Dilara B, Nargis K, Naeem A, Shahida P. Surface properties of the mixed oxides of iron and silica. Colloids. Surf. A. Physicochem. Eng. Asp. 2002;205:273–282. https://doi.org/10.1016/s0927-7757(02)00025-0
crossref

38. Garcia-Ochoa F, Gomez E. Bioreactor scale-up and oxygen transfer rate in microbial processes: an overview. Biotechnol. Adv. 2009;27:153–176. https://doi.org/10.1016/j.biotechadv.2008.10.006.
crossref pmid

39. Xu Z, Mochida K, Naito T, Yasuda K. Effects of operational conditions on 1, 4-dioxane degradation by combined use of ultrasound and ozone microbubbles. Jpn. J. Appl. Phys. 2012;51:07GD08. https://doi.org/10.1143/jjap.51.07gd08.
crossref

40. Li HZ, Hu LM, Song DJ, Lin F. Characteristics of micro-nano bubbles and potential application in groundwater bioremediation. Water Environ. Res. 2014;86:844–851. https://doi.org/10.2175/106143014x14062131177953.
crossref pmid

41. Chu LM, Xing XH, Yu AF, Zhou YN, Sun XL, Jurcik B. Enhanced ozonation of simulated dyestuff wastewater by microbubbles. Chemosphere. 2007;68:1854–1860. https://doi.org/10.1016/j.chemosphere.2007.03.014.
crossref pmid

42. Liu S, Wang QH, Zhai XD, Huang QF, Huang PK. Improved Pretreatment (Coagulation-Floatation and Ozonation) of Younger Landfill Leachate by Microbubbles. Water Environ. Res. 2010;82:657–665. https://doi.org/10.2175/106143010x12609736966522.
pmid

43. Kim S, Kim H, Han M, Kim T. Generation of sub-micron (nano) bubbles and characterization of their fundamental properties. Environ. Eng. Res. 2019;24:382–388. https://doi.org/10.4491/eer.2018.210.


44. Nam G, Mohamed MM, Jung J. Enhanced degradation of benzo [a] pyrene and toxicity reduction by microbubble ozonation. Environ. Technol. 2021;42:1853–1860. https://doi.org/10.1080/09593330.2019.1683077.
crossref pmid

45. Saroyan H, Kyzas GZ, Deliyanni EA. Effective dye degradation by graphene oxide supported manganese oxide. Processes. 2019;7:40. https://doi.org/10.3390/pr7010040.
crossref

46. Wang YX, Xie YB, Sun HQ, Xiao JD, Cao HB, Wang SB. 2D/2D nano-hybrids of γ-MnO2 on reduced graphene oxide for catalytic ozonation and coupling peroxymonosulfate activation. J. Hazard. Mater. 2016;301:56–64. https://doi.org/10.1016/j.jhazmat.2015.08.031.
crossref pmid

47. Lee CG, Javed H, Zhang DN, et al. Porous electrospun fibers embedding TiO2 for adsorption and photocatalytic degradation of water pollutants. Environ. Sci. Technol. 2018;52:4285–4293. https://doi.org/10.1021/acs.est.7b06508.
crossref pmid

48. Glaze WH, Kang JW, Chapin DH. The Chemistry of Water Treatment Processes Involving Ozone, Hydrogen Peroxide and Ultraviolet Radiation. Ozone Sci. Eng. 2008;9:335–352. https://doi.org/10.1080/01919518708552148.
crossref

49. Mosallanejad S, Dlugogorski BZ, Kennedy EM, Stockenhuber M. On the chemistry of iron oxide supported on γ-alumina and silica catalysts. ACS omega. 2018;3:5362–5374. https://doi.org/10.1021/acsomega.8b00201.
crossref pmid pmc

50. Li LS, Zhu WP, Zhang PY, Zhang QY, Zhang ZL. AC/O3-BAC processes for removing refractory and hazardous pollutants in raw water. J. Hazard. Mater. 2006;135:129–133. https://doi.org/10.1016/j.jhazmat.2005.11.045.
crossref pmid

51. Roshani B, McMaster I, Rezaei E, Soltan J. Catalytic ozonation of benzotriazole over alumina supported transition metal oxide catalysts in water. Sep. Purif. Technol. 2014;135:158–164. https://doi.org/10.1016/j.seppur.2014.08.011.
crossref

52. Zheng QF, Wang ZM, Chen BG, Liu GF, Zhao J. Analysis of XRD Spectral Structure and Carbonization of the Biochar Preparation. Spectrosc. spect. anal. 2016;36:3355–3359. https://doi.org/10.3964/j.issn.1000-0593(2016)10-3355-05
crossref

53. Li PC, Hu CC, Lee TC, Chang WS, Wang TH. Synthesis and characterization of carbon black/manganese oxide air cathodes for zinc-air batteries. J. Power Sources. 2014;269:88–97. https://doi.org/10.1016/j.jpowsour.2014.06.108.
crossref

54. Hinokuma S, Shimanoe H, Kawabata Y, Kiritoshi S, Araki K, Machida M. Supported and unsupported manganese oxides for catalytic ammonia combustion. Catal. Commun. 2018;105:48–51. https://doi.org/10.1016/j.catcom.2017.11.009.
crossref

55. Beltrán FJ, Rivas FJ, Montero-de-Espinosa R. Catalytic ozonation of oxalic acid in an aqueous TiO2 slurry reactor. Appl. Catal. B: Environ. 2002;39:221–231. https://doi.org/10.1016/s0926-3373(02)00102-9
crossref

56. Ernst M, Lurot F, Schrotter JC. Catalytic ozonation of refractory organic model compounds in aqueous solution by aluminum oxide. Appl. Catal. B: Environ. 2004;47:15–25. https://doi.org/10.1016/s0926-3373(03)00290-x
crossref

57. Zhao L, Ma J, Sun ZZ, Zhai XD. Mechanism of influence of initial pH on the degradation of nitrobenzene in aqueous solution by ceramic honeycomb catalytic ozonation. Environ. Sci. Technol. 2008;42:4002–4007. https://doi.org/10.1021/es702926q.
crossref pmid

58. Bai ZY, Yang Q, Wang JL. Fe3O4/multi-walled carbon nanotubes as an efficient catalyst for catalytic ozonation of p-hydroxybenzoic acid. Int. J. Environ. Sci. Technol. 2016;13:483–492. https://doi.org/10.1007/s13762-015-0881-3.
crossref

59. Liu XH, Li HP, Fang Y, Yang ZG. Heterogeneous catalytic ozonation of sulfamethazine in aqueous solution using maghe-mite-supported manganese oxides. Sep. Purif. Technol. 2021;274:118945. https://doi.org/10.1016/j.seppur.2021.118945.
crossref

60. Song JS, Ma NW, Chen WQ, Chen JM, Dai QZ. Insights into mechanism of catalytic ozonation of cinnamyl alcohol over core-shell Fe3O4@ SiO2@ La2O3 catalyst. Sep. Purif. Technol. 2022;282:119969. https://doi.org/10.1016/j.seppur.2021.119969.
crossref

61. Radhakrishnan R, Oyama ST, Chen JG, Asakura K. Electron transfer effects in ozone decomposition on supported manganese oxide. J. Phys. Chem. B. 2001;105:4245–4253. https://doi.org/10.1021/jp003246z.
crossref

62. Zhao L, Sun ZZ, Ma J, Liu HL. Enhancement mechanism of heterogeneous catalytic ozonation by cordierite-supported copper for the degradation of nitrobenzene in aqueous solution. Environ. Sci. 2009;43:2047–2053. https://doi.org/10.1021/es803125h.
crossref

63. Psaltou S, Kaprara E, Triantafyllidis K, Mitrakas M, Zouboulis A. Heterogeneous catalytic ozonation: The significant contribution of PZC value and wettability of the catalysts. J. Environ. Chem. Eng. 2021;9:106173. https://doi.org/10.1016/j.jece.2021.106173.
crossref

64. Sun QQ, Li LS, Yan HH, Hong XT, Hui KS, Pan ZQ. Influence of the surface hydroxyl groups of MnOx/SBA-15 on heterogeneous catalytic ozonation of oxalic acid. Chem. Eng. J. 2014;242:348–356. https://doi.org/10.1016/j.cej.2013.12.097.
crossref

65. Li JR, Chen JS, Yu YK, He C. Fe-Mn-Ce/ceramic powder composite catalyst for highly volatile elemental mercury removal in simulated coal-fired flue gas. J. Ind. Eng. Chem. 2015;25:352–358. https://doi.org/10.1016/j.jiec.2014.11.015.
crossref

66. Zhang JW, Guo Q, Wu WL, et al. Preparation of Fe-MnOx/AC by high gravity method for heterogeneous catalytic ozonation of phenolic wastewater. Chem. Eng. Sci. 2022;255:117667. https://doi.org/10.1016/j.ces.2022.117667.
crossref

67. Ma J, Sui MH, Zhang T, Guan CY. Effect of pH on MnOx/GAC catalyzed ozonation for degradation of nitrobenzene. Water Res. 2005;39:779–786. https://doi.org/10.1016/j.watres.2004.11.020.
crossref pmid

68. Liu RX, Tang HX. Oxidative decolorization of direct light red F3B dye at natural manganese mineral surface. Water Res. 2000;34:4029–4035. https://doi.org/10.1016/s0043-1354(00)00166-4
crossref

69. Zhao L, Ma J, Sun ZZ, Zhai XD. Catalytic ozonation for the degradation of nitrobenzene in aqueous solution by ceramic honeycomb-supported manganese. Appl. Catal. B: Environ. 2008;83:256–264. https://doi.org/10.1016/j.apcatb.2008.02.009.
crossref

70. Nöthe T, Fahlenkamp H, Sonntag CV. Ozonation of wastewater: rate of ozone consumption and hydroxyl radical yield. Environ. Sci. Technol. 2009;43:5990–5995. https://doi.org/10.1021/es900825f.
crossref

71. Rischbieter E, Stein H, Schumpe A. Ozone solubilities in water and aqueous salt solutions. J. Chem. Eng. Data. 2000;45:338–340. https://doi.org/10.1021/je990263c.
crossref

Fig. 1
Effect of different aeration processes (ozone dosage at 2 L/min, pH at 7.1, reaction temperature at 25 °C, and continuous aeration time at 10 min).
/upload/thumbnails/eer-2022-606f1.gif
Fig. 2
COD removal rates of different catalysts
/upload/thumbnails/eer-2022-606f2.gif
Fig. 3
Effect of different catalyst dosages on (a) COD removal rate and (b) UV254 removal rate (ozone dosage at 2 L/min, pH at 7.1, reaction temperature at 25 °C, and continuous aeration time at 120 min).
/upload/thumbnails/eer-2022-606f3.gif
Fig. 4
Effect of different initial pH values of the solution on (a) COD removal rate and (b) UV254 removal rate (ozone dosage at 2 L/min, catalyst dosage at 3.0 g/L, reaction temperature at 25 °C, and continuous aeration time at 120 min).
/upload/thumbnails/eer-2022-606f4.gif
Fig. 5
Effect of different initial pH values of the solution (ozone dosage at 2 L/min, catalyst dosage at 3.0 g/L, reaction temperature at 25 °C, and continuous aeration time at 120 min).
/upload/thumbnails/eer-2022-606f5.gif
Fig. 6
Effect of different initial temperatures of the solution on (a) COD removal rate and (b) UV254 removal rate (ozone dosage at 2 L/min, catalyst dosage at 3.0 g/L, pH at 7.1, and continuous aeration time at 120 min).
/upload/thumbnails/eer-2022-606f6.gif
Fig. 7
The stability of catalysts (ozone dosage at 2 L/min, catalyst dosage at 3.0 g/L, pH at 9.4, and continuous aeration time at 120 min).
/upload/thumbnails/eer-2022-606f7.gif
Table 1
COD removal rates of catalysts prepared under different preparation conditions.
Mn/GAC Cu/GAC Fe/GAC

Roasting temperature (°C) Roasting time (h) Impregnation solution concentration (mol/L) COD removal rate (%) Roasting temperature (°C) Roasting time (h) Impregnation solution concentration (mol/L) COD removal rate (%) Roasting temperature (°C) Roasting time (h) Impregnation solution concentration (mol/L) COD removal rate (%)
400 2 0.5 50.90% 400 2 0.5 51.20% 400 2 0.5 51.50%
500 2 0.5 58.90% 500 2 0.5 53.70% 500 2 0.5 49%
600 2 0.5 48.70% 600 2 0.5 49.60% 600 2 0.5 48.00%
500 2 0.5 50.90% 500 2 0.5 53.70% 400 2 0.5 51.50%
500 3 0.5 63.20% 500 3 0.5 54.60% 400 3 0.5 52.30%
500 4 0.5 63.90% 500 4 0.5 57.30% 400 4 0.5 54.60%
500 3 1 64.30% 500 4 1 53.40% 400 4 1 56.20%
500 3 2 57.80% 500 4 2 51.80% 400 4 2 53.50%
Editorial Office
464 Cheongpa-ro, #726, Jung-gu, Seoul 04510, Republic of Korea
FAX : +82-2-383-9654   E-mail : eer@kosenv.or.kr

Copyright© Korean Society of Environmental Engineers.        Developed in M2PI
About |  Browse Articles |  Current Issue |  For Authors and Reviewers