Next Article in Journal
Controlled Synthesis of Ag-SnO2/α-Fe2O3 Nanocomposites for Improving Visible-Light Catalytic Activities of Pollutant Degradation and CO2 Reduction
Previous Article in Journal
Metal-Doped Mesoporous MnO2-CeO2 Catalysts for Low-Temperature Pre-Oxidation of NO to NO2 in Fast SCR Process
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Non-Thermal Plasma Incorporated with Cu-Mn/γ-Al2O3 for Mixed Benzene Series VOCs’ Degradation

1
School of Chemical Engineering, China University of Petroleum (East China), Qingdao 266580, China
2
Key Laboratory of Petroleum and Petrochemical Pollution Control and Treatment, Ministry of Science and Technology, Beijing 102200, China
3
CAS Key Laboratory of Separation Science for Analytical Chemistry, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, Dalian 116023, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Submission received: 12 February 2023 / Revised: 28 March 2023 / Accepted: 30 March 2023 / Published: 3 April 2023

Abstract

:
In this work, a coaxial dielectric barrier discharge (DBD) reactor was constructed to degrade the mixture of toluene and o-xylene, two typical benzene series. The Cu-MnO2/γ-Al2O3 series catalysts prepared by redox and impregnation methods were filled into the plasma device to degrade VOCs synergistically and explore the degradation effect. The experimental results showed that the introduction of a Cu-doped MnO2 catalyst significantly improved the pollutants’ removal efficiency and CO2 selectivity, and greatly inhibited the formation of by-products. Among them, Cu0.15Mn/γ-Al2O3 showed the highest removal efficiency (toluene was 100% and o-xylene was 100%), and the best CO2 selectivity (92.73%). The XRD, BET, XPS and SEM results confirmed that the synergistic effect between Cu and Mn in the Cu-Mn solid solution could promote the amount and reducibility of the surface active oxygen species, which improved the catalytic performance. Finally, the toluene and o-xylene decomposition pathways in the NTP catalytic system were speculated according to the detected organic matter. This work provides a theoretical and experimental basis for the application of DBD-catalyzed hybrid benzene series VOCs.

Graphical Abstract

1. Introduction

The emission of volatile organic compounds (VOCs) not only harms human health, but also causes environmental problems such as haze, photochemical smog, and ozone depletion [1]. Among VOCs, the benzene series are a typical kind of VOC, which are usually used as organic solvents in industrial production due to their good stability and indissolubility with most organics. However, the benzene series can cause air pollution and be toxic to living organisms [2]. Many benzene series have pungent odors, causing discomfort to the human body. They also cause photochemical smog in cities, producing secondary pollution. Technologies such as adsorption, absorption, condensation, membrane separation, catalytic oxidation/combustion, biological treatment, photocatalytic oxidation, and non-thermal plasma have been applied to the degradation of VOCs [3]. Photocatalytic oxidation technology and NTP technology are two promising technologies. Among them, non-thermal plasma (NTP) technology has attracted considerable attention due to its advantages of simple operation, high oxidation degradation activity, and low cost [4,5]. However, the formation of by-products and low energy efficiency relatively limit the application of NTP [6]. Jiang et al. [7] detected intermediate products, including HCOOH, CO, CO2, H2O, N2O, and O3, in the experimental process of studying the pulsed-modulated dielectric barrier discharge (DBD) degradation of toluene. Chang et al. proved that O3 was an inevitable by-product when gas containing O2 was used as the carrier gas, and that the O3 concentration increased with the increase in the specific input energy (SIE) [8]. Ye et al. achieved a removal rate of 100% when degrading benzene containing gas, but the mineralization rate was not 100% [9]. It was found that aerosols were generated during the discharge process, and phenolic substances were deposited on the inner wall of the DBD tubes to form brown solids. Fan et al. [10] found that large amounts of O3 and trace amounts of NO2 were produced when dealing with low concentrations of p-xylene. Compared with other traditional treatment technologies, the non-thermal plasma degradation of VOCs has certain advantages, but it has disadvantages such as a low energy efficiency and undesirable by-products, such as NOX and O3, which inevitably remain in the exhaust gas. To overcome these shortcomings, researchers have found that combining plasma with catalytic technology can not only significantly improve energy efficiency and selectivity, but also greatly reduce the production of by-products. Plasma co-catalytic technology combines the advantages of plasma and heterogeneous catalytic technology, and has the advantages of a high VOC emission reduction efficiency, good product selectivity, less by-product generation, and high energy efficiency. It is a low-cost and energy-saving method for VOC removal.
The addition of the catalyst can effectively improve the efficiency of the system, especially for the degradation of lower concentrations of VOCs [11,12,13]. On the one hand, compared with the plasma alone, the catalyst can help the system produce active species (free radicals and various ions) with a lower input power. In turn, plasma can provide the thermal energy required for catalysis. On the other hand, catalysts can provide abundant oxygen vacancies and further generate oxygen-active species by utilizing O3 and gaseous O2 in the system. In addition, it also improves the selectivity of CO2 and inhibits the generation of O3 [14]. The catalyst is generally composed of support and active components. Common supports include Al2O3, SiO2, molecular sieve, diatomite, activated carbon, etc. These catalysts are usually characterized by a large specific surface area, high mechanical strength, high temperature resistance, and low fluid resistance, and their shapes are generally honeycomb, spherical, lumpy, and other shapes. The support can increase the interaction time between VOCs and catalysts, thus facilitating the degradation of VOCs [15]. γ-Al2O3 is widely used as a support due to its low price, convenient use, high reactivity, and chemical stability in VOCs’ degradation [15]. Wu et al. [16] load NiO on Al2O3 support to remove toluene, which has a higher dispersion than SBA-15, TiO2 support. Therefore, the catalytic effect was better. Dang et al. [17] used a DBD reactor filled with catalysts M/13X-γ-Al2O3 (M: Ag, Ce, Mn, and Co) to evaluate the removal effect of toluene, and the results showed that Ag/13X-γ-Al2O3 had the highest penetration and the highest mineralization rate (about 68%) at 20 kV. Jiang et al. [18] loaded the transition metals of Mn, Co, Fe, and Cu onto SiO2, ZSM-5-300, ZSM-5-38, and γ-Al2O3 and filled them into the DBD reactor to degrade toluene. The catalyst supported by γ-Al2O3 showed the best performance in toluene decomposition and the highest toluene decomposition was 97% when the MnO2 loaded on γ-Al2O3 was 1 wt%.
Transition metal oxides have been widely used as catalysts for VOCs’ degradation in plasma catalytic systems because of their low price and effective catalytic decomposition of gaseous pollutants [19,20]. MnO2 has oxygen vacancies and a valence state, which can promote the degradation of VOCs and improve COX selectivity. Meanwhile, MnO2 was effective for ozone removal [21,22]. Ge et al. [23] used a combination of DBD and MnO2 catalyst to degrade benzene and were able to significantly increase the COX selectivity from 38% to 80% with almost no O3 generated. Wang et al. [24] found that the MnOX catalysts prepared with different precursors had different o-xylene removal performances in the plasma discharge region, and this difference was mainly caused by the presence of Mn4+ and lattice oxygen on the catalyst. More Mn4+ and lattice oxygen were conducive to the degradation of o-xylene. Yang [25] et al. prepared catalyst three-dimensional hollow urchin α-MnO2 and combined it with the PPC (post-plasma catalysis) system to degrade toluene. The maximum toluene degradation rate can reach 100%, and the energy efficiency was improved by 64% compared with NTP alone. Zheng [26] et al. combined solar irradiation and PPC to degrade toluene over a MnO2/GFF (bifunctional graphene fin foam) catalyst. When the SIE was 350 J/L, the degradation rate of toluene reached 93% and the selectivity of CO2 reached 83%, which significantly reduced the energy consumption of the plasma catalytic gas purification process. Trinh et al. [27] reported that ceramics loaded with MnO2 exhibited a better catalytic performance for acetone decomposition than ZnO no matter if in the plasma discharge area or after the discharge. However, pure MnO2 catalysts usually require more energy in the degradation process of VOCs and may lead to the formation of other by-products, which makes it difficult to achieve the ideal catalytic degradation effect. Therefore, it is necessary to enhance the interaction of supports [28]. Roberto [29] reviewed the application of bimetallic catalysts in VOC oxidation in terms of the catalytic activity and physical and chemical properties. It was found that the bimetallic catalyst had a great improvement in properties compared with a nonmetallic catalyst due to the synergistic effect. In recent years, Cu-doped MnO2 catalysts have shown broad application prospects with high surface areas and redox performances [21,30]. A study proved that a Cu-O-Mn interface could be formed by introducing copper into manganese oxide, which improves the reducibility of Cu-OMS [31]. Therefore, a Cu-doped MnO2 catalyst can be considered to deal with benzene series VOCs.
Most of the current studies focus on a single target and ignore the interaction between the different components in the degradation process. In addition, toluene and o-xylene are two typical benzene VOCs, usually from the petrochemical, coating, automobile manufacturing, packaging, and printing industries. In this work, a coaxial DBD reactor was built to degrade mixed gases of toluene and o-xylene. A series of Cu-Mn/γ-Al2O3 catalysts with various Cu/Mn molar ratios were prepared and filled in the NTP reactor to co-degrade benzene series. The removal efficiency, COX selectivity, catalyst characterization, and energy efficiency were studied and discussed deeply. Organic matter during the degradation of toluene and o-xylene in the NTP catalytic system was detected by GC-MS and the possible decomposition pathways were speculated.

2. Results and Discussion

2.1. Catalysts Characterization

The XRD patterns of MnO2, Cu0.05Mn, Cu0.15Mn, Cu0.25Mn, and Cu0.5Mn are shown in Figure 1. In total, 5 main diffraction peaks can be seen in the XRD diffraction patterns of the pure MnO2 in the figure, centered at 2θ = 28.7°, 37.5°, 42.9°, 46.2°, and 67.4°, respectively, corresponding to standard MnO2 (PDF-72-1984). According to the standard card, they correspond to the (110), (101), (111), (210), and (310) crystal planes, respectively. After Cu was gradually doped, the primary crystal structure of MnO2 was unchanged and no visible diffraction peaks of CuO appeared, indicating the high dispersion of Cu species in MnO2 [14]. As the amount of Cu increased, a slight peak shift (shaded parts) can be seen until the molar ratio of Cu-to-Mn reached 0.15. Since the ionic radius of octahedral manganese ion in MnO2 was much smaller than that of Cu2+,the latter can easily replace Mn4+ in MnO2 [32]. However, only a small amount of Cu2+ can enter the MnO2 lattice to form a solid solution and lead to lattice expansion.
The SEM photos of Cu-doped MnO2 catalysts can be seen in Figure 2. As can be seen from the scan image magnified 5000 times, the pure MnO2 and Cu-doped MnO2 both showed similar morphologies, with rough and uneven surfaces, and a part of the larger particle loaded, which could easily fall off due to collision and friction in the later stage [33]. After comparative analysis, small particulate matter appeared on the original MnO2/γ-Al2O3 pattern after copper doping. As can be seen from the scanning figure magnified 20,000 times, in the pore structure of the γ-Al2O3 particles, Cu entered the pore in the form of small pellets and honeycomb and was loaded well inside. Because MnO2 is doped into the γ-Al2O3 support, it is judged that the gray area is alumina and the white area is MnO2. Most of the MnO2 (white) was embedded in γ-Al2O3 (gray) as particles. The surface structure of the support did not change greatly after loading.
Figure 3 shows the N2 adsorption/desorption isotherms of the catalysts. It can be seen that each catalyst had H1 hysteresis loops [34]. According to the results listed in Table 1, the catalysts prepared by a redox precipitation method were typical mesoporous material in the range of 2–50 nm. The N2 physisorption curve belongs to the type IV isotherm. The γ-Al2O3 support pellets selected in this study had the largest specific surface area of 322.560 m2·g−1. When MnO2 and Cu were loaded onto the catalyst, the specific surface area decreased, indicating that the catalyst was better coated on the surface of the support. Compared with other catalysts, Cu0.15Mn had the largest pore volume of 0.316 cm3·g−1 and the smallest pore size of 4.614 nm. The small pore size and large pore volume meant that there was more space for micro reactions, which was more conducive to the production of reactive species by micro discharge. Although the main mesoporous structure of MnO2 would not be affected by a small amount of Cu introduction, the pore volume can be changed effectively.
The XPS scanning spectra of O1s, Mn2p, and Cu2p are shown in Figure 4a–c, respectively. According to the deconvoluted O1s XPS spectra in Figure 4a, all catalysts showed 2 distinct sub-bands at around 532.18 eV and 529.88 eV, belonging to the surface labile oxygen (Oads) and lattice oxygen (Olat), respectively [35,36]. The redox cycle between Mn4+/Mn3+ and Cu2+/Cu+ consumes lattice oxygen and generates oxygen vacancies. The existence of oxygen vacancies is conducive to the formation of Oads on the catalyst surface, which has a higher mobility than Olat and is conducive to the catalytic oxidation reaction. The depleted oxygen vacancy can be filled by the decomposition of oxygen and O3 in the plasma gas phase [37].
Figure 4b shows the deconvoluted Mn2p XPS results of the Cu-doped MnO2 catalysts with a range of 635~660 eV. There were two peaks at 643.38 eV and 642.38 eV, which belong to the Mn4+ and Mn3+ species, respectively [38]. Similarly, there were also two sub-peaks of the Cu2p spectrum, as shown in Figure 4c. The region at 933.58 eV can correspond to the surface Cu2+, and the region near 931.78 eV was assigned to the Cu+ species, respectively [39].

2.2. The Degradation Efficiency of Toluene and o-Xylene

The removal efficiency of toluene and o-xylene in NTP alone and the NTP-catalysis system have been studied as a function of SIE under the same conditions. It can be seen from Figure 5 that the degradation efficiency of toluene and o-xylene increased with the increase in SIE, regardless of whether the catalyst was used. This was because, with the increase in SIE, the number of high-energy electrons and active free radicals generated in the reactor increased, and the probability of each VOC molecule sharing high-energy electrons and active free radicals increased, thus leading to the improvement of VOCs’ degradation efficiency. In addition, the degradation efficiency of o-xylene was higher than that of toluene at the same energy density. For example, when SIE was 5.72 kJ·L−1, the degradation efficiency of 700 mg·m−3 o-xylene was as high as 82.05%, while for toluene with the same concentration, the degradation efficiency was only 53.17%. This meant that the degradation efficiency of VOCs was related to VOC types and their chemical structure. On the one hand, the reason for the degradation was that the chemical bonds within VOCs were broken by the energetic electrons and active particles. In the case of toluene and o-xylene, the main fractures were C-H (methyl), C-H (benzene ring), C-C (methyl with benzene ring), and C-C (benzene ring). The chemical bond energy and molecular structure of the VOC molecules determined their degradation efficiency during NTP degradation [40,41]. The C-H bond energy on methyl (3.7 eV) is smaller than that on the benzene ring (4.3 eV), and the C-C bond energy between the methyl and benzene ring (4.4 eV) is also smaller than that on the benzene ring (5.0~5.3 eV). There are two methyl groups in o-xylene; therefore, when the number of energetic electrons and active radicals in the discharge area was constant, C-H (methyl) and C-C (methyl and benzene ring) on o-xylene molecules were more likely to collide with these active particles. Therefore, the o-xylene was more easily degraded compared with toluene. In toluene, on the other hand, the superchoke effect between the methyl and benzene rings [42] brought the electrons of the C-H bond on methyl closer to H, so that methyl was easily oxidized. In addition, the electron cloud density of benzene ring increased under the influence of the methyl group, and it was also prone to substitution reaction [43]. O-xylene contains two methyl groups, and the superconjugate effect between the methyl group and the benzene ring was more obvious. The electron cloud density of the benzene ring increased further, and the substitution reaction was more likely to occur, which caused the o-xylene to be degraded more easily.
In order to investigate the interaction between toluene and o-xylene in the degradation process, a mixture of 900 mg·m−3 toluene and 900 mg·m−3 o-xylene was prepared to study the degradation efficiency. It can be seen from the Figure 5c, in the whole range of energy density, that the degradation efficiency of toluene and o-xylene after mixing is higher than that of a single component with the same concentration. The degradation of single component VOCs can produce reactive intermediates, such as aldehydes and peroxides, which can improve the degradation efficiency of VOCs to a certain extent. More of this reactive substance is produced in the mixture degradation than in the degradation of a single component, and it is easier to form reaction intermediates/free radicals. Therefore, the degradation of the toluene and o-xylene mixture can improve the utilization of the active substance produced in the DBD reactor. Some scholars have also observed that VOCs are easily activated in mixtures [44,45]. Therefore, the degradation efficiency of toluene and o-xylene in the mixture is higher.
As shown in Figure 5a, when toluene was degraded under plasma alone, the conversion rate was 53.17% at the SIE of 5.72 kJ·L−1. After adding Cu-doped MnO2 catalysts to the plasma system, the removal efficiency of toluene was obviously improved, following the order of Cu0.15Mn > Cu0.05Mn > Cu0.25Mn > Cu0.5Mn > MnO2. The results indicated that the catalysts could improve the removal rate of toluene, and the catalytic effect of catalyst was related to the amount of the Cu load in the catalysts. Among these catalysts, Cu0.15Mn had the best performance, achieving 100% conversion at 5.72 kJ·L−1 SIE, which was about 2 times higher than that of separated plasma. As shown in Figure 5b, the Cu-doped MnO2 catalysts in the plasma reactor also greatly promoted the degradation of o-xylene. Similar to toluene, Cu0.15Mn still showed the highest removal efficiency (~100%). MnO2 had a special pore structure, high thermal stability, and various in valence states of Mnx+ (x = 2, 3 and 4). These unique properties gave MnO2 a good catalytic activity [46]. The solid solution effect between Cu and Mn in the catalyst may be the reason why the degradation efficiency was obviously improved with different amounts of Cu doping. According to the magnitude of the peak shift in the XRD pattern, the MnO2 structure may be affected by Cu, which may cause a charge imbalance and unsaturated chemical bonds on the catalyst, leading to the formation of more oxygen vacancies. Gaseous oxygen could fill the oxygen vacancy on the catalyst surface and generate reactive oxygen species [39,47]. Surface reactive oxygen species could be regarded as the main active substances in the plasma catalytic oxidation of toluene and o-xylene. A comparison of the degradation efficiency for this and other studies is shown in Table S2.
It also can be seen from the Figure that the degradation efficiency does not improve with the increase in Cu doping, which may be due to the following two reasons. First, the performance of catalyst was related to the dispersion state of the active component Cu on the support. As the load increased, the number of active centers of catalytic reaction exposed in the plasma space increased, and the oxidation rate of toluene gradually accelerated. When the loading exceeded a certain level, the Cu aggregated into larger particles, and the active components could be distributed on the support in the form of multilayer accumulation or clusters. The number of catalytic reaction centers did not increase with the excessive increase in the loading, but the number of reaction centers exposed to the plasma space decreased, which caused a decrease in the catalytic activity. Therefore, the catalytic effect of Cu0.05Mn and Cu0.5Mn was slightly worse than that of the Cu0.15Mn catalyst. Second, the pore size affects the residence time of the reaction; a suitable pore size was conducive to reactants into the active site of the catalyst for a reaction. A small pore size and large pore volume cause more space for small reactions, which was more conducive for micro discharges to produce reactive species and participate in the oxidation reaction of VOCs [48]. According to the BET characterization results, compared with other catalysts, Cu0.15Mn had the smallest pore size of 4.614 nm and the largest pore volume of 0.316 cm3·g−1. In conclusion, the catalyst Cu0.15Mn had the best degradation performance.

2.3. The COX Selectivity

As can be seen from the Figure 6, with the increase in SIE, the selectivity of CO2 increased and CO decreased in both the NTP alone system and the NTP-catalytic system. High specific input energy may lead to more electrons with high energy. Furthermore, high-energy electrons are also conducive to the production of active substances that completely oxidize toluene and o-xylene into CO2 [47], and part of CO would be oxidized into CO2, so the selectivity of CO is reduced while the CO2 selectivity is increased. Among the projects studied, Cu0.15Mn produced the highest CO2 selectivity (~92.73%) and the lowest CO selectivity (~3.57%) at 5.72 kJ·L−1. In contrast, a single plasma produced the lowest CO2 selectivity (~52.08%) and the highest CO (~12.20%). This is because the addition of a catalyst in the discharge gap increased the average field strength and the corresponding average electron energy, resulting in a surface discharge on the catalyst particles, which induced the effective interaction between plasma and catalyst to activate CO2, leading to the enhancement of the CO2 conversion. The Cu0.15Mn catalyst had more space for micro reactions, which was conducive to micro discharge, generated more active species, and made more intermediate products converted to CO2 and H2O.

2.4. Inhibition Effect of the NO2 and O3 Generation

NO2 and O3 are the main by-products in the plasma degradation process. The concentrations of NO2 and O3 produced by the degradation of VOCs in NTP alone and NTP-catalyst systems were determined by gas chromatography and indigo sodium disulfonate spectrophotometry from 0.51 to 5.72 kJ·L−1. Since NO was easily oxidized to NO2 by a large amount of ozone after discharge, the main NOX species was NO2 [15]. It can be seen from Figure 7a that the NO2 concentration increased with the increase in SIE. The main reason was that with the increase in SIE, the number of electrons, ions, excited molecules, and free radicals formed increased, and electrons with energy exceeding 5.1 eV or 9.18 eV increased, which led to the ionization and excitation of the N2, O2, and H2O molecules in the carrier gas, resulting in the generation of some primary active substances, such as O·, OH·, H·, N2*, and O3. As shown in Equations (1)–(4) [49,50]:
e + O 2 e + O D   1 + O P   3
e + N 2 N · + N · + e
e + H 2 O H · + OH · + e
O · + O 2 + M O 3 + M
Among them, the N· and O· generated NO2 reacted with O2 and H2O [51,52], as shown in Equations (5)–(9):
N · + O 2 NO + O ·
N · + H 2 O NO + H 2
NO + O · NO 2
O · + NO 2 NO 3
O · + NO 3 O 2 + NO 2
OH· reacted with NO and NO2 to form HNO2 and HNO3, which in turn reacted with OH· to form NO2:
OH · + NO HNO 2
OH · + NO 2 HNO 3
OH · + HNO 2 H 2 O + NO 2
OH · + HNO 3 H 2 O + NO 3
O3 was formed by the reaction of O2 and O·, and participated in the formation reaction of NO2, as shown in Equation (14):
O 3 + NO NO 2 + O 2
Therefore, the formation rate of NO2 increased with the increase in SIE. The combination of NTP and the catalyst could obviously inhibit the formation of NO2, and Cu0.5Mn had the best effect on the inhibition of NO2 emission. When SIE increased from 0.51 kJ·L−1 to 5.72 kJ·L−1, the NO2 concentration increased from 211 ppm to 394 ppm mg·m−3, compared with NTP alone, which reduced the concentration of NO2 by 36%, indicating that the Cu-doped MnO2 was also beneficial to inhibit NOX formation, and the inhibition effect was related to the amount of Cu doping. It was preliminarily speculated that the reducibility of Cu+ can directly reduce the original NO2 to N2. Compared with other catalysts, the effect of Cu0.5Mn was the best.
As can be seen from Figure 7b, in the whole energy density region, with the increase of SIE, the O3 concentration first increased and then decreased. The reason for this was that when the SIE was within 4.42 kJ·L−1, the number of free radicals and high-energy electrons increased with the electric field strength, and these substances collided or reacted with the air molecules to produce more atomic oxygen, leading to a greater chance of colliding with oxygen, which caused the concentration of O3 to increase [53]. Furthermore, when the SIE increased to more than 4.42 kJ·L−1, the temperature increased in the discharge area and the ultraviolet light emitted by a large number of excited particles when they transitioned to ground state particles, which caused O3 to break bonds or degrade.
Compared with NTP alone, the Cu-doped MnO2 catalyst could effectively reduce O3 emission. When SIE was 5.72 kJ·L−1, the O3 concentration decreased from 125.6 mg·m−3 to 55.8 mg·m−3. Different catalysts inhibited O3 generation in the order of Cu0.15Mn > Cu0.25Mn > Cu0.5Mn > Cu0.05Mn > MnO2. Cu0.15Mn had the highest O3 inhibition, and the concentration of O3 was about 56.6%, lower than that with NTP alone. This suggests that the catalyst Cu0.15Mn had more space for micro reactions. The active species generated by micro discharge decomposed O3 on the surface of Cu0.15Mn and generated reactive oxygen atoms, which participated in the oxidation process of VOCs.

2.5. Intermediate Product and Mechanism

2.5.1. Intermediate Product Analysis

Some by-products are produced in the non-thermal plasma catalysis system. The experimental results are shown in Figure S1, and the by-products are listed in Table S1. In this study, the plasma device input SIE was 5.72 kJ·L−1, the initial concentration of toluene and o-xylene was 700 mg·m−3, and the flow rate was 1 L·min−1. The exhaust gas produced after degradation was studied and analyzed by gas chromatography mass spectrometry (GC-MS). In the plasma reactor filled with five different Cu-doped MnO2 catalysts, the by-products were different. The detected 8 open benzene ring by-products and 15 benzene ring containing by-products are shown in Table 2.

2.5.2. Mechanism Analysis

The process of VOC treatment by non-thermal plasma co-catalysis was mainly composed of two parts, one was the reaction in the gas phase and the other was the reaction on the surface of the catalyst.
The reaction in the gas phase mainly had three processes: high-energy electrons collided with gas molecules to produce active substances, energetic electrons and active particles reacted with VOCs molecules, and unstable active particles such as free radicals further reacted with the small molecules of organic matter after decomposition.
(1) High-energy electrons collided with gas molecules to produce active substances: High-energy electrons collided with N2, O2, CO2, and H2O in the carrier gas at room temperature and a large number of O·, OH·, H·, N·, ·HO2, O3 and other active particles were generated. These active particles and free radicals promoted the oxidative decomposition reaction of VOC molecules, as shown in Equations (1)–(3).
(2) Energetic electrons and active particles reacted with VOCs molecules: In the process of non-thermal plasma discharge, a large number of high-energy electrons would be generated [53]. The high-energy electrons generated would react with the VOC molecules. The excitation and dissociation of VOC molecules depended on the energy of the high-energy electrons and the bond energy of chemical bonds in the molecule. When the energy of high-energy electrons was greater than the bond energy of the chemical bonds in VOC molecules, the chemical bonds broke, accompanied by unstable intermediates. The unstable intermediates either collided with the energetic electrons again or combined with other free radicals to form other stable substances [54].
The bond energies of different chemical bonds in VOCs and the reactions that would occur when they collided directly with high-energy electrons are shown in Table 2. According to Table 2, the lower the bond energy, the more likely the bond was to break in direct collision with high-energy electrons. The C-H bond energy on methyl was relatively small, and high-energy electrons could easily destroy this bond to form H· and phenylmethyl [55,56]. Secondly, the C-H bond on the benzene ring was broken [55]. Next, the C-C bond between methyl and benzene ring was broken to form phenyl and methyl [56]. Finally, the C-C bond on the benzene ring was broken and two small molecule hydrocarbons were generated [54]. Jiang [18] found that the degradation process of toluene in the plasma system loaded with 1 wt%Mn/γ-Al2O3 was as follows: toluene → benzyl → benzyl alcohol → benzaldehyde → benzyl acid → carbonates and bicarbonates → CO2 and H2O. However, because the structure of benzene ring was relatively stable, carbon atoms were bonded with the π bonds, which made it difficult to break or the energy required to break the C-C (benzene ring) bonds was high. Therefore, it was difficult to make a ring-opening reaction happen by pure electron collision as the reaction depended more on the oxidation decomposition of O3 and various active particle free radicals in the reaction system.
The reaction rate constants of electrons with toluene are greater than those of electrons with OH·, O·, and O3. The contribution of electrons to toluene degradation is very low. Electrons react with toluene by ionization, dissociation, and excitation of O2 and N2 to produce reactive oxygen species and reactive nitrogen species [37]. The excited nitrogen plays a very important role. Robby [57]’s results suggested that the N2(A3 U + ) species play the most important role in the destruction of ethylene. The direct destruction of ethylene by electron impact was negligible. In a previous study, Zhang [58] simulated the removal of formaldehyde under the N2 atmosphere and air atmosphere, respectively. It was found that N2(A3 U + ) can obviously remove formaldehyde in the N2 atmosphere, and it was removed by the collision of N2(A3 U + ) with formaldehyde. However, the removal of HCHO in the air atmosphere was determined by O· and OH· radicals, and N2(A3 U + ) mainly acted to increase the concentration of O· and OH· radicals in the system.
The unstable intermediates (such as C6H4·CH3 and C6H3·CH3CH3) generated by the high-energy electrons reacted with electrons again are shown in Equations (15)–(18):
C 6 H 4 · CH 3 + e C 3 H 2 · C 3 H 2 · + CH 3 · / C 6 H 4 · CH 2 · + H ·
C 6 H 3 · ( CH 3 ) 2 + e C 3 H 2 · C 3 H · C 3 H + CH 3 ·
C 6 H 3 · ( CH 3 ) 2 + e C 6 H 3 · CH 2 · CH 3 + H ·
C 6 H 3 · ( CH 3 ) 2 + e C 6 H 3 · CH 2 · CH 2 + 2 H ·
(3) The unstable active particles continued to react with the decomposed intermediates C6H5CH2·, C6H4CH2·CH2·, C6H4·CH3, C6H3·CH3CH3, C6H5·, and CH3·, as shown in Figure 8. In other words, VOCs were ultimately decomposed into CO2, H2O, and other environmentally pollution-free substances through various pathways under the action of free radicals O·, OH·, H·, N·, and ·HO2.
Table 2. The reaction of breaking different chemical bonds in toluene and o-xylene [56].
Table 2. The reaction of breaking different chemical bonds in toluene and o-xylene [56].
Chemical BondsEnergy/eVBond Breaking Reactions That Occurred after Collision
C-H (Methyl)3.7C7H8 + e → C6H5CH2· + H·
C8H10 + e → C7H7CH2· + H·/C8H10 + e → C6H4CH2·CH2· + 2H·
C-H (Benzene ring)4.3C7H8 + e → C6H4·CH3 + H·
C8H10 + e → C6H3·CH3CH3 + H·
C-C (Between methyl and benzene)4.4C7H8 + e → C6H5· + CH3·
C8H10 + e → C6H4·CH3 + CH3·/C8H10 + e → C3H2·C3H2· + 2CH3·
C-C (Benzene ring)5.0–5.3C7H8 + e → C3H4· + C4H4·
C8H10 + e → C4H6· + C4H4·
O·, OH·, and other free radicals determined the degree of oxidation in the formation of VOCs and small molecular organics. However, considering the limited number of active free radicals, some small molecules of organic matter would not be completely oxidized into CO2 and H2O, thus inhibiting the degradation of VOCs. Therefore, the removal efficiency of VOCs could be improved by introducing a Cu-doped MnO2 catalyst.
After the catalyst was introduced, toluene, o-xylene, and organic intermediates in the gas phase could be adsorbed on the surface of the catalyst. The gas phase products after the degradation of toluene and o-xylene were analyzed. Figure S1 shows that the types of by-products were reduced compared with the types of intermediate by-products generated by plasma degradation alone. This showed that the introduction of a Cu-doped MnO2 catalyst could transform more intermediate by-products into small molecule inorganic substances and improve the removal efficiency of VOCs. The synergistic effect between non-thermal plasma and a Cu-doped MnO2 catalyst was mainly reflected in the following aspects:
(4) Under the action of the applied electric field, a large number of active particles produced by the discharge, such as high-energy electrons, ions, excited atoms, and free radicals, impacted the surface of the catalyst, polarized the catalyst particles, and caused the secondary emission of electrons. The enhanced field region was formed on the surface of the catalyst, and the unreacted molecules in the discharge region were further excited into reactive free radicals to participate in the reaction [59].
(5) The oxidation of toluene and o-xylene molecules on mangan-based catalysts was carried out by the Mars-van Krevelen mechanism [60]. In the catalyst filling zone, toluene and o-xylene were adsorbed on the catalyst surface. There are two types of sources of surface reactive oxygen species. One is the decomposition of O3 and the collision between the plasma and the oxygen adsorbed by the catalyst [61]. The other is supplemented by the REDOX cycles of the Cu and Mn cations (Mn4+/Mn3+ and Cu2+/Cu+) in the Cu-Mn/γ-Al2O3 catalysts. The REDOX cycle between Mn4+/Mn3+ and Cu2+/Cu+ consumes lattice oxygen to create oxygen vacancies that can be filled by the decomposition of oxygen and O3 in the plasma gas phase [62].
(6) A large number of active particles generated by the discharge had abundant energies, which could reduce the activation energy required by part of the catalyst reaction and speed up the oxidation reaction rate.
(7) The catalyst selectively reacted with some of the by-products produced in the reaction process, further converted them into simple small molecule safety substances, and then inhibited the production of toxic by-products.
The mechanism of the co-catalytic treatment of toluene and o-xylene by non-thermal plasma was shown in Figure 9. Figure 9a depicted the main reactions of toluene and o-xylene in the gas phase, and Figure 9b depicted the main reactions of toluene and o-xylene on the surface of the catalyst.

3. Experimental Instruments and Methods

3.1. Catalysts Preparation

The Cu-doped MnO2 catalysts were prepared by the KMnO4 redox precipitation method [63]. γ-Al2O3, with a diameter of 2–3 mm, was used as the support of the catalyst. Typically, 17.82 g of MnCl2·4H2O and the desired amount of CuSO4·5H2O (1.875 g, 5.625 g, 9.375 g, 18.75 g) were first dissolved in the 100 mL of DI H2O. Then, 100 mL of the KMnO4 solution (9.48 g) was added dropwise into the above solution with vigorous stirring. After being stirred for 24 h, the pre-dried γ-Al2O3 support was immersed in the solution and stirred for 6 h, and then washed and filtered alternately by DI water and ethanol. The obtained precipitates were dried in an oven at 120 °C for 12 h. The dried samples were further calcined in air at 300 °C for 2 h with the heating rate of 5 °C·min−1. The catalysts were abbreviated as CuXMn for ease of reference, where x represents the molar ratio of Cu to Mn. In this experiment, the catalysts with different molar ratios were labeled with Cu0.05Mn, Cu0.15Mn, Cu0.25Mn, and Cu0.5Mn, respectively. In addition, the pure MnO2 catalyst was also synthesized with a similar method for comparison.

3.2. Catalysts Characterization

The specific surface area, pore size, and pore volume of the catalysts were measured using N2 adsorption-desorption isotherm (Mike Tristar 2020 surface area analyzer, Atlanta, GA, USA). The X-ray diffraction (XRD) pattern of the catalyst samples was performed on Brook D8 Advance with Cu-Kα radiation in the scan range of 10°~90° and scan rate of 2°·min−1. The surface microtopography of the samples were characterized with a Hitachi S4800 JSM-7500F scanning electron microscope (Tokyo, Japan). The X-ray photoelectron spectroscopy (XPS) was measured with the Thermo ESCALAB 250Xi equipment (Waltham, MA, USA) with Al Ka X-ray radiation. All binding energies were corrected by standard carbon (C1s, 284.8 eV).

3.3. Experimental System

As shown in Figure 10, the experimental system was mainly composed of VOC generator systems, a DBD reactor, and an exhaust gas detection and analysis system. The total gas flow was controlled to 1 L·min−1 by setting the flow rates of the two syringe pumps to 0.5 L·min−1, respectively. The concentration of toluene and o-xylene was maintained at 700 mg·m−3 by controlling the temperature and air flow rate of the vaporization chamber.
The structure of the coaxial dielectric barrier discharge plasma reactor filled with catalyst is shown in Figure 11. The inner electrode is a stainless-steel cylinder in the middle of the reactor, and the blocking medium is a quartz tube coaxial with an inner electrode (inner diameter: 20 mm). The outer side of the quartz tube is covered with a layer of stainless-steel net as the outer electrode (inner diameter: 25 mm) [15]. The discharge area of the reactor is the gap between the outer surface of the inner electrode and the inner surface of the quartz. The discharge gap is 3 mm and the length of discharge zone is 150 mm [6]. The DBD reactor is powered by an AC high voltage power supply (CTP-2000 K, Nanjing Suman Plasma Technology, Nanjing, China). We adjusted the voltage and current by adjusting the knob of the high voltage supply. The digital storage oscilloscope (TBS1052B) was used to measure the input power and applied voltage [15].

3.4. Test and Analysis Methods

(1) The toluene and o-xylene concentrations were adsorbed on an activated carbon adsorption tube, desorbed by CS2, and then they were measured by gas chromatography (Agilent Technologies 7890B, Santa Clara, CA, USA), which used a hydrogen flame ionization detector (FID) and a DB-FFAP capillary column. The conditions were as follows: the oven temperature was 40 °C, the detector temperature was 250 °C, and the inlet temperature was 150 °C [6].
The degradation efficiencies η1 and η2 were used to evaluate the treatment effect of toluene and o-xylene, respectively:
η 1 = [ C 7 H 8 ] 0 [ C 7 H 8 ] 1 [ C 7 H 8 ] 0 × 100 %
η 2 = [ C 8 H 10 ] 0 [ C 8 H 10 ] 1 [ C 8 H 10 ] 0 × 100 %
where [ C 7 H 8 ] 0 ,   [ C 7 H 8 ] 1 and [ C 8 H 10 ] 0 ,   [ C 8 H 10 ] 1 represented the mass concentrations (mg·m−3) of toluene and o-xylene before and after the reaction, respectively.
(2) The concentrations of CO2 and CO at the outlet of the reactor were determined by gas chromatograph equipped with a thermal conductivity detector (TCD). The chromatographic column was an HQ-packed column, the detection temperature was 80 °C in the column box, and the front detector was 150 °C. The CO2 and CO concentrations were calibrated by standard gas.
The selectivity of CO2, CO, and COX were:
S CO 2 = 2 [ CO 2 ] 8 { [ C 8 H 10 ] 0 [ C 8 H 10 ] 1 } + 7 { [ C 7 H 8 ] 0 [ C 7 H 8 ] 1 } × 100 %
S CO = 2 CO 8 { [ C 8 H 10 ] 0 [ C 8 H 10 ] 1 } + 7 { [ C 7 H 8 ] 0 [ C 7 H 8 ] 1 } × 100 %
S CO x = S CO 2 + S CO
where CO 2 ,   CO were the concentration of CO2 and CO in the outlet gas, respectively.
(3) The Lissajous graph method was used to calculate the discharge power of the plasma. Figure 12 shows the Lissajous graph. The discharge power, P, of the DBD reactor in one discharge cycle can be obtained by integrating the area of the Lissajous graph [64]:
P c o n s u m e = f × C × S
where f was the frequency, C was the capacitance at 0.47 µF, and S was the area of the Lissajous figure.
(4) The specific input energy (SIE, kJ·L−1) was used to indicate the power consumed by the plasma when processing a unit flow of gas:
SIE = P c o n s u m e Q × 60 × 10 3   KJ · L 1
where Q was the flow rate of the total gas (L/min).
(5) The comprehensive energy efficiency performance of DBD was evaluated by the energy efficiency (ηE, g·kWh−1), which represented the removal quality of pollutants per unit energy:
η E 1 = 3.6 × [ C 7 H 8 ] 0 [ C 7 H 8 ] 1 SED × 1000   g · kWh 1
η E 2 = 3.6 × [ C 8 H 10 ] 0 [ C 8 H 10 ] 1 SED × 1000   g · kWh 1
(6) The content of O3 was determined by the spectrophotometric method of indigo disulfonate; the concentration of NO2 at the outlet of the reactor was determined by gas chromatograph equipped with a thermal conductivity detector (TCD) under the same conditions as (2).
(7) The intermediate products were analyzed by GC-MS (Agilent Technologies 7890B GC/5977B MSD). The column was an HP-5 capillary column (30.00 m in length, 0.32 mm in diameter and 0.25 µm in thickness). The detection conditions were as follows: the oven temperature was 30 °C, kept for 5 min. The subsequent program was raised to 120 °C, kept for 2 min at the heating rate of 3 °C·min−1. The sample inlet temperature was 200 °C; the MS conditions were as follows: ion source temperature was 230 °C, quadrupole temperature was 150 °C, and mass-to-charge ratio scanning range was 35~270 amu.

4. Conclusions

In this paper, the effects of the Cu-doped MnO2 catalyst on the toluene and o-xylene catalytic reactions in non-thermal plasma were investigated, and the VOCs’ removal efficiency, COX selectivity, and by-product formation were compared. It was found that the removal efficiency of VOCs and the CO2 selectivity were significantly improved by introducing the catalyst into the plasma treatment process. When the catalyst Cu0.15Mn was packed into the discharge region, this plasma-catalytic system showed the highest removal efficiency (toluene was 100% and o-xylene was 100%), the best CO2 selectivity (92.73%), and the strongest by-product inhibition rate (36% for NO2 and 56.6% for O3) at 5.72 kJ·L−1. It was concluded that the excellent performance of the Cu-doped MnO2 catalyst was related to the formation of the Cu-Mn solid solution and surface reactive oxygen species. Finally, the possible degradation pathways of VOCs in plasma and plasma catalysis were proposed and discussed with the help of GC-MS. The results can provide a theoretical basis for the study on the performance of VOC degradation by NTP and have certain guiding significance for the application and mechanism research of VOC degradation by NTP.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/catal13040695/s1, Figure S1: The mass spectrum of the exhaust gas of the plasma-catalytic device; Table S1: Types of main by-products; Table S2: Comparison of degradation efficiency for this and other studies. References [10,24,65,66,67] cited in Supplementary Materials.

Author Contributions

Conceptualization, F.H.; Methodology, F.H.; Validation, P.S., X.L. and Y.Z. (Yupeng Zhang); Investigation, Y.Z. (Yifan Zhu); Resources, L.H. and F.H.; Data curation, Y.Z. (Yifan Zhu) and C.J.; Writing—original draft, Y.Z. (Yifan Zhu); Writing—review and editing, D.L.; Supervision, F.L. and L.H. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the National Nature Science Foundation of China (21505156); the Shandong Nature Science Foundation of China (ZR2020MA104); the Fundamental Research Funds for the Central Universities (19CX02039A).

Data Availability Statement

The data presented in this study are available in article and Supplementary Material.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shao, P.; An, J.; Xin, J.; Wu, F.; Wang, J.; Ji, D.; Wang, Y. Source apportionment of VOCs and the contribution to photochemical ozone formation during summer in the typical industrial area in the Yangtze River Delta, China. Atmos. Res. 2016, 176, 64–74. [Google Scholar] [CrossRef]
  2. Vandenbroucke, A.M.; Morent, R.; De Geyter, N.; Leys, C. Non-thermal plasmas for non-catalytic and catalytic VOC abatement. J. Hazard. Mater. 2011, 195, 30–54. [Google Scholar] [CrossRef] [PubMed]
  3. Feng, C.Y.; Zhao, J.K. Technology Status of Removing Volatile Organic Compounds by Corona Discharge. Sichuan Environ. 2003, 2, 2–5. [Google Scholar]
  4. Liang, C.J.; Li, K.W. Kinetic characterization of plasma-enhanced catalysis of high-concentration volatile organic compounds over mullite supported perovskite catalysts. J. Electrost. 2018, 96, 134–143. [Google Scholar] [CrossRef]
  5. Huang, Q.; Fang, C. Degradation of 3,3′,4,4′-tetrachlorobiphenyl (PCB77) by dielectric barrier discharge (DBD) non-thermal plasma: Degradation mechanism and toxicity evaluation. Sci. Total Environ. 2020, 739, 139926. [Google Scholar] [CrossRef]
  6. Han, F.; Li, M.; Zhong, H.; Li, T.; Li, D.; Zhou, S.; Guo, W.; Liu, F. Product analysis and mechanism of toluene degradation by low temperature plasma with single dielectric barrier discharge. J. Saudi Chem. Soc. 2020, 24, 673–682. [Google Scholar]
  7. Jiang, N.; Zhao, Y.; Shang, K.; Lu, N.; Li, J.; Wu, Y. Degradation of toluene by pulse-modulated multistage DBD plasma: Key parameters optimization through response surface methodology (RSM) and degradation pathway analysis. J. Hazard. Mater. 2020, 393, 122365. [Google Scholar] [CrossRef]
  8. Chang, T.; Shen, Z.; Huang, Y.; Lu, J.; Ren, D.; Sun, J.; Cao, J.; Liu, H. Post-plasma-catalytic removal of toluene using MnO2-Co3O4 catalysts and their synergistic mechanism. Chem. Eng. J. 2018, 348, 15–25. [Google Scholar] [CrossRef] [Green Version]
  9. Ye, Z.; Zhang, Y.; Li, P.; Yang, L.; Zhang, R.; Hou, H. Feasibility of destruction of gaseous benzene with dielectric barrier discharge. J. Hazard. Mater. 2008, 156, 356–364. [Google Scholar] [CrossRef]
  10. Fan, X.; Zhu, T.L.; Wang, M.Y.; Li, X.M. Removal of low-concentration BTX in air using a combined plasma catalysis system. Chemosphere 2009, 75, 1301–1306. [Google Scholar] [CrossRef]
  11. Karatum, O.; Deshusses, M.A. A comparative study of dilute VOCs treatment in a non-thermal plasma reactor. Chem. Eng. J. 2016, 294, 308–315. [Google Scholar] [CrossRef]
  12. Chen, J.X.; Pan, K.L.; Yu, S.J.; Yen, S.Y.; Chang, M.B. Combined fast selective reduction using Mn-based catalysts and nonthermal plasma for NOx removal. Environ. Sci. Pollut. Res. 2017, 24, 21496–21508. [Google Scholar] [CrossRef]
  13. Zhu, X.; Zhang, S.; Yang, Y.; Zheng, C.; Zhou, J.; Gao, X.; Tu, X. Enhanced performance for plasma-catalytic oxidation of ethyl acetate over La1−xCexCoO3+δ catalysts. Appl. Catal. B Environ. 2017, 213, 97–105. [Google Scholar] [CrossRef]
  14. Zeng, X.; Li, B.; Liu, R.; LI, X.; Zhu, T. Investigation of promotion effect of Cu doped MnO2 catalysts on ketone-type VOCs degradation in a one-stage plasma-catalysis system. Chem. Eng. J. 2020, 384, 123362. [Google Scholar] [CrossRef]
  15. Li, M.; Li, D.; Zhang, Z.; Ji, C.; Zhou, S.; Guo, W.; Zhao, C.; Liu, F.; Han, F. Study on the performance and mechanism of degradation of toluene with non-thermal plasmas synergized supported TiO2/γ-Al2O3 catalyst. J. Environ. Chem. Eng. 2021, 9, 105529. [Google Scholar] [CrossRef]
  16. Wu, J.; Xia, Q.; Wang, H.; Li, Z. Catalytic performance of plasma catalysis system with nickel oxide catalysts on different supports for toluene removal: Effect of water vapor. Appl. Catal. B Environ. 2014, 156–157, 265–272. [Google Scholar] [CrossRef]
  17. Dang, X.; Li, S.; Yu, X.; Guo, H.; Qin, C.; Cao, L. Kinetic characterization of adsorbed toluene removal involving hybrid material catalysts—[M/13X-γ-Al2O3 (M: Ag, Ce, Mn, and Co)] in a sequential non-thermal plasma system. Chem. Eng. Res. Des. 2020, 155, 80–87. [Google Scholar] [CrossRef]
  18. Jiang, B.; Xu, K.; Li, J.; Lu, H.; Fei, X.; Yao, X.; Yao, S.; Wu, Z. Effect of supports on plasma catalytic decomposition of toluene using in situ plasma DRIFTS. J. Hazard. Mater. 2021, 405, 124203. [Google Scholar] [CrossRef]
  19. Trinh, Q.H.; Mok, Y.S. Environmental plasma-catalysis for the energy-efficient treatment of volatile organic compounds. Korean J. Chem. Eng. 2016, 33, 735–748. [Google Scholar] [CrossRef]
  20. Li, Y.; Fan, Z.; Shi, J.; Liu, Z.; Shangguan, W. Post plasma-catalysis for VOCs degradation over different phase structure MnO2 catalysts. Chem. Eng. J. 2014, 241, 251–258. [Google Scholar] [CrossRef]
  21. Ma, J.; Wang, C.; He, H. Transition metal doped cryptomelane-type manganese oxide catalysts for ozone decomposition. Appl. Catal. B Environ. 2017, 201, 503–510. [Google Scholar] [CrossRef]
  22. Dhandapani, B.; Oyama, S.T. Gas phase ozone decomposition catalysts. Appl. Catal. B Environ. 1997, 11, 129–166. [Google Scholar] [CrossRef]
  23. Ge, H.; Hu, D.; Li, X.; Tian, Y.; Chen, Z.; Zhu, Y. Removal of low-concentration benzene in indoor air with plasma-MnO2 catalysis system. J. Electrost. 2015, 76, 216–221. [Google Scholar] [CrossRef]
  24. Wang, L.; He, H.; Zhang, C.; Wang, Y.; Zhang, B. Effects of precursors for manganese-loaded gamma-Al2O3 catalysts on plasma-catalytic removal of o-xylene. Chem. Eng. J. 2016, 288, 406–413. [Google Scholar] [CrossRef]
  25. Yang, S.; Yang, H.; Yang, J.; Qi, H.; Kong, J.; Bo, Z.; Li, X.; Yan, J.; Cen, K.; Tu, X. Three-dimensional hollow urchin α-MnO2 for enhanced catalytic activity towards toluene decomposition in post-plasma catalysis. Chem. Eng. J. 2020, 402, 126154. [Google Scholar] [CrossRef]
  26. Bo, Z.; Yang, S.; Kong, J.; Zhu, J.; Wang, Y.; Yang, H.; Li, X.; Yan, J.; Cen, K.; Tu, X. Solar-enhanced plasma-catalytic oxidation of toluene over a bifunctional graphene fin foam decorated with nanofin-like MnO2. ACS Catal. 2020, 10, 4420–4432. [Google Scholar] [CrossRef] [Green Version]
  27. Trinh, H.Q.; Mok, Y.S. Plasma-catalytic oxidation of acetone in annular porous monolithic ceramic-supported catalysts. Chem. Eng. J. 2014, 251, 199–206. [Google Scholar] [CrossRef]
  28. Luo, Y.; Zuo, J.; Feng, X.; Qian, Q.; Zheng, Y.; Lin, D.; Huang, B.; Chen, Q. Good interaction between well dispersed Pt and LaCoO3 nanorods achieved rapid Co3+/Co2+ redox cycle for total propane oxidation. Chem. Eng. J. 2018, 357, 395–403. [Google Scholar] [CrossRef]
  29. Fiorenza, R. Bimetallic catalysts for volatile organic compound oxidation. Catalysts 2020, 10, 661. [Google Scholar] [CrossRef]
  30. Li, Q.; Yin, L.; Li, Z.; Wang, X.; Qi, Y.; Ma, J. Copper doped hollow structured manganese oxide mesocrystals with controlled phase structure and morphology as anode materials for lithium ion battery with improved electrochemical performance. ACS Appl. Mater. Interfaces 2013, 5, 10975–10984. [Google Scholar] [CrossRef]
  31. Yang, Y.; Huang, J.; Zhang, S.; Wang, S.; Deng, S.; Wang, B.; Yu, G. Catalytic removal of gaseous HCBz on Cu doped OMS: Effect of Cu location on catalytic performance. Appl. Catal. B Environ. 2014, 150–151, 167–178. [Google Scholar]
  32. John, R.E.; Chandran, A.; George, J.; Jose, A.; Jose, G.; Jose, J.; Unnikrishnan, N.V.; Thomas, M.; George, K.C. High temperature ferroelectric behaviour in α-MnO2 nanorods realised through enriched oxygen vacancy induced non-stoichiometry. Phys. Chem. Chem. Phys. PCCP 2017, 19, 28756–28771. [Google Scholar] [CrossRef]
  33. Li, J. Study on Catalytic Oxidation of Toluene by Dielectric Barrier Discharge. Master’s Thesis, Beijing University of Chemical Technology, Beijing, China, 2017. [Google Scholar]
  34. Sing, K.S.W. Reporting physisorption data for gas/solid systems with special reference to the determination of surface area and porosity (Recommendations 1984). Pure Appl. Chem. 1985, 57, 603–619. [Google Scholar] [CrossRef]
  35. Gu, T.; Liu, Y.; Weng, X.; Wang, H.; Wu, Z. The enhanced performance of ceria with surface sulfation for selective catalytic reduction of NO by NH3. Catal. Commun. 2010, 12, 310–313. [Google Scholar] [CrossRef]
  36. Wu, Z.; Sheng, Z.; Liu, Y.; Wang, H.; Mo, J. Deactivation mechanism of PtOx/TiO2 photocatalyst towards the oxidation of NO in gas phase. J. Hazard. Mater. 2011, 185, 1053–1058. [Google Scholar] [CrossRef]
  37. Jiang, N.; Zhao, Y.; Qiu, C.; Shang, K.; Lu, N.; Li, J.; Wu, Y.; Zhang, Y. Enhanced catalytic performance of CoOx-CeO2 for synergetic degradation of toluene in multistage sliding plasma system through response surface methodology (RSM). Appl. Catal. B Environ. 2019, 259, 118061. [Google Scholar] [CrossRef]
  38. Arandiyan, H.; Dai, H.; Deng, J.; Liu, Y.; Bai, B.; Wang, Y.; Li, X.; Xie, S.; Li, J. Three-dimensionally ordered macroporous La0.6Sr0.4MnO3 with high surface areas: Active catalysts for the combustion of methane. J. Catal. 2013, 307, 327–339. [Google Scholar]
  39. Zhu, Z.; Lu, G.; Zhang, Z.; Guo, Y.; Guo, Y.; Wang, Y. Highly Active and Stable Co3O4/ZSM-5 Catalyst for Propane Oxidation: Effect of the Preparation Method. ACS Catal. 2013, 3, 1154–1164. [Google Scholar] [CrossRef]
  40. Yamamoto, T. VOC decomposition by nonthermal plasma processing—A new approach. J. Electrost. 1997, 42, 227–238. [Google Scholar] [CrossRef]
  41. Ogata, A.; Ito, D.; Mizuno, K.; Kushiyama, S.; Gal, A.; Yamamoto, T. Effect of coexisting components on aromatic decomposition in a packed-bed plasma reactor. Appl. Catal. Gen. 2002, 236, 9–15. [Google Scholar] [CrossRef]
  42. Li, Y.J.; Zhong, Z.H.; Ma, C.; Shi, J.L.; Shi, Q. Effects of methyl side chains on the preparation of carbon microspheres by chemical vapor deposition of benzene. Carbon Tech. 2015, 34, 20–23. [Google Scholar] [CrossRef]
  43. Dai, S.W. Experimental study on degradation of benzene series by Ce modified Mn-based activated carbon fiber coupled with low-temperature plasma. Master’s Thesis, Donghua University, Shanghai, China, 2021. [Google Scholar]
  44. Papenmeier, D.M.; Rossin, J.A. Catalytic oxidation of dichloromethane, chloroform, and theirbinary mixtures over a platinum alumina catalyst. Ind. Eng. Chem. Res. 1994, 33, 3094–3103. [Google Scholar] [CrossRef]
  45. Burgos, N.; Paulis, M.; Antxustegi, M.M.; Montes, M. Deep oxidation of VOC mixtures with platinum supported on Al2O3/Al monoliths. Appl. Catal. B Environ. 2002, 38, 251–258. [Google Scholar] [CrossRef]
  46. Yu, W.X.; Cheng, G.; Cui, X.E.; Zeng, X.H.; Lan, B.; Li, Y.F.; Zhang, C.Y.; Sun, M.; Yu, L. Preparation of ɛ-MnO2 with different morphologies and their catalytic combustion of toluene. Chem. Ind. Eng. Prog. 2021, 40, 5547–5553. (In Chinese) [Google Scholar]
  47. Solsona, B.; García, T.; Sanchis, R.; Soriano, M.D.; Moreno, M.; Rodríguez-Castellón, E.; Agouram, S.; Dejoz, A.; Nieto, J.M.L. Total oxidation of VOCs on mesoporous iron oxide catalysts: Soft chemistry route versus hard template method. Chem. Eng. J. 2016, 290, 273–281. [Google Scholar] [CrossRef]
  48. Sun, L. Degradation of Toluene Gas by NiO/γ-Al2O3 Catalyst with Low Temperature Plasma Technology. Master’s Thesis, East China University of Science and Technology, Shanghai, China, 2019. [Google Scholar]
  49. Lu, M.; Yang, W.; Yu, C.; Liu, Q.; Ye, D. Plasma-catalytic oxidation of toluene on Ag modified FeOx/SBA-15. Aerosol Air Qual. Res. 2020, 20, 193–202. [Google Scholar] [CrossRef] [Green Version]
  50. Dahiru, U.H.; Saleem, F.; Zhang, K.; Harvey, A. Plasma-assisted removal of methanol in N2, dry and humidified air using a dielectric barrier discharge (DBD) reactor. RSC Adv. 2022, 12, 10997–11007. [Google Scholar] [CrossRef]
  51. Guo, Y.; Liao, X.; Fu, M.; Huang, H.; Ye, D. Toluene decomposition performance and NOx by-product formation during a DBD-catalyst process. J. Environ. Sci. 2015, 28, 187–194. [Google Scholar] [CrossRef]
  52. Shahna, F.G.; Bahrami, A.; Alimohammadi, I.; Yarahmadi, R.; Jaleh, B.; Gandomi, M.; Ebrahimi, H.; Abedi, K.A. Chlorobenzene degeradation by non-thermal plasma combined with EG-TiO2/ZnO as a photocatalyst: Effect of photocatalyst on CO2 selectivity and byproducts reduction. J. Hazard. Mater. 2017, 324, 544–553. [Google Scholar] [CrossRef]
  53. Shang, K.; Wang, M.; Peng, B.; Yarahmadi, R.; Jaleh, B.; Gandomi, M.; Ebrahimi, H.; Abedi, K.A.-D. Characterization of a novel volume-surface DBD reactor: Discharge characteristics, ozone production and benzene degradation. J. Phys. D Appl. Phys. 2019, 53, 65201. [Google Scholar] [CrossRef]
  54. Xiao, G.; Xu, W.; Wu, R.; Ni, M.; Du, C.; Gao, X.; Luo, Z.; Cen, K. Non-Thermal Plasmas for VOCs Abatement. Plasma Chem. Plasma Process. 2014, 34, 1033–1065. [Google Scholar] [CrossRef]
  55. Huang, H.; Ye, D.; Leung, D.Y.; Feng, F.; Guan, X. Byproducts and pathways of toluene destruction via plasma-catalysis. J. Mol. Catal. A Chem. 2011, 336, 87–93. [Google Scholar] [CrossRef]
  56. Xu, D.; Sheng, L.; Wang, H.; Sun, Y.; Zhang, X.; Mi, J. Study of magnetically enhanced corona pre-charger. J. Electrost. 2007, 65, 101–106. [Google Scholar] [CrossRef]
  57. Aerts, R.; Tu, X.; De Bie, C.; Whitehead, J.C.; Bogaerts, A. An investigation into the dominant reactions for ethylene destruction in non-thermal atmospheric plasmas. Plasma Process. Polym. 2012, 9, 994–1000. [Google Scholar] [CrossRef]
  58. Zhang, J.; Lv, F.G.; Xu, Y.; Yang, X.F.; Zhu, A.M. Chemical kinetics Simulation of formaldehyde removal by dielectric barrier discharge. J. Acta Phys. Chim. Sin. 2007, 9, 1425–1431. [Google Scholar]
  59. Zhang, L. Study on Purification of Toluene Waste Gas by Low Temperature Plasma Technology. Master’s Thesis, Shandong University of Science and Technology, Qingdao, China, 2013. [Google Scholar]
  60. Zhu, X.; Liu, S.; Cai, Y.; Gao, X.; Zhou, J.; Zheng, C.; Tu, X. Post-plasma catalytic removal of methanol over Mn-Ce catalysts in an atmospheric dielectric barrier discharge. Appl. Catal. B Environ. 2016, 183, 124–132. [Google Scholar] [CrossRef] [Green Version]
  61. Yao, X.; Zhang, J.; Liang, X.; Long, C. Niobium doping enhanced catalytic performance of Mn/MCM-41 for toluene degradation in the NTP-catalysis system. Chemosphere 2019, 230, 479–487. [Google Scholar] [CrossRef]
  62. Li, S.; Yu, X.; Dang, X.; Wang, P.; Meng, X.; Wang, Q.; Hou, H. Double dielectric barrier discharge incorporated with CeO2-Co3O4/γ-Al2O3 catalyst for toluene abatement by a sequential adsorption–discharge plasma catalytic process. J. Clean. Prod. 2022, 340, 130774. [Google Scholar] [CrossRef]
  63. Njagi, E.C.; Chen, C.H.; Genuino, H.; Galindo, H.; Huang, H.; Suib, S.L. Total oxidation of CO at ambient temperature using copper manganese oxide catalysts prepared by a redox method. Appl. Catal. B Environ. 2010, 99, 103–110. [Google Scholar] [CrossRef]
  64. Yao, C.K. Experimental Study on Degradation of VOCs by DBD Synergistic and Photocatalytic. Master’s Thesis, Xi’an University of Technology, Xi’an, China, 2018. [Google Scholar]
  65. Yang, J.; Liu, S.; He, T.; Nengzi, L.-C.; Wang, Y.; Su, L.; Cao, J.; Ji, W.; Yuan, C.; Geng, M. Degradation of toluene by DBD plasma-catalytic method with MnxCoyCezOn catalysts: Characterization of catalyst, catalytic activity and continuous test. J. Environ. Chem. Eng. 2021, 9, 106361. [Google Scholar] [CrossRef]
  66. Liu, J.; Liu, X.; Chen, J.; Li, X.; Ma, T.; Zhong, F. Investigation of ZrMnFe/sepiolite catalysts on toluene degradation in a one-stage plasma-catalysis system. Catalysts 2021, 11, 828. [Google Scholar] [CrossRef]
  67. Hossain, M.M.; Mok, Y.S.; Nguyen, D.B.; Kim, S.-j.; Kim, Y.J.; Lee, J.H.; Heo, I. Nonthermal plasma in practical-scale honey-comb catalysts for the removal of toluene. J. Hazard. Mater. 2021, 404, 123958. [Google Scholar] [CrossRef]
Figure 1. XRD curves of Cu-doped MnO2 catalysts.
Figure 1. XRD curves of Cu-doped MnO2 catalysts.
Catalysts 13 00695 g001
Figure 2. SEM characterization of Cu-doped MnO2 catalysts. (a) MnO2/γ-Al2O3, (b) Cu0.05Mn/γ-Al2O3, (c) Cu0.15Mn/γ-Al2O3, (d) Cu0.25Mn/γ-Al2O3, (e) Cu0.5Mn/γ-Al2O3.
Figure 2. SEM characterization of Cu-doped MnO2 catalysts. (a) MnO2/γ-Al2O3, (b) Cu0.05Mn/γ-Al2O3, (c) Cu0.15Mn/γ-Al2O3, (d) Cu0.25Mn/γ-Al2O3, (e) Cu0.5Mn/γ-Al2O3.
Catalysts 13 00695 g002aCatalysts 13 00695 g002b
Figure 3. BET characterization analysis diagram of Cu-doped MnO2 catalysts. (a) N2 adsorption-desorption isotherm of the catalysts, (b) pore size distribution measurement diagrams.
Figure 3. BET characterization analysis diagram of Cu-doped MnO2 catalysts. (a) N2 adsorption-desorption isotherm of the catalysts, (b) pore size distribution measurement diagrams.
Catalysts 13 00695 g003
Figure 4. XPS curves of Cu-doped MnO2 catalysts: (a) O1s; (b) Mn2p; (c) Cu2p.
Figure 4. XPS curves of Cu-doped MnO2 catalysts: (a) O1s; (b) Mn2p; (c) Cu2p.
Catalysts 13 00695 g004
Figure 5. Degradation efficiency of toluene and o-xylene with different catalysts and mixed gas without catalyst. (a) Toluene, (b) o-xylene, (c) mixed gas without catalyst.
Figure 5. Degradation efficiency of toluene and o-xylene with different catalysts and mixed gas without catalyst. (a) Toluene, (b) o-xylene, (c) mixed gas without catalyst.
Catalysts 13 00695 g005
Figure 6. COX selectivity with different catalysts. (a) CO2 and (b) CO.
Figure 6. COX selectivity with different catalysts. (a) CO2 and (b) CO.
Catalysts 13 00695 g006
Figure 7. The relationship between by-product concentration and SIE with different catalysts. (a) NO2 and (b) O3.
Figure 7. The relationship between by-product concentration and SIE with different catalysts. (a) NO2 and (b) O3.
Catalysts 13 00695 g007
Figure 8. Schematic diagram of reaction between decomposed intermediates and active particles.
Figure 8. Schematic diagram of reaction between decomposed intermediates and active particles.
Catalysts 13 00695 g008
Figure 9. Degradation pathway of toluene and o-xylene under plasma catalysis system. (a) The reaction of VOCs in the gas phase (b) The reaction of VOCs on the surface of the catalyst.
Figure 9. Degradation pathway of toluene and o-xylene under plasma catalysis system. (a) The reaction of VOCs in the gas phase (b) The reaction of VOCs on the surface of the catalyst.
Catalysts 13 00695 g009
Figure 10. Schematic diagram of the experimental system.
Figure 10. Schematic diagram of the experimental system.
Catalysts 13 00695 g010
Figure 11. Structure diagram of a coaxial dielectric barrier discharge non-thermal plasma reactor.
Figure 11. Structure diagram of a coaxial dielectric barrier discharge non-thermal plasma reactor.
Catalysts 13 00695 g011
Figure 12. Lissajous graphic.
Figure 12. Lissajous graphic.
Catalysts 13 00695 g012
Table 1. Comparison of specific surface area, average pore diameter, and pore volume of different catalysts.
Table 1. Comparison of specific surface area, average pore diameter, and pore volume of different catalysts.
Catalyst TypeSBET (m2·g−1)Average Pore Size (nm)Pore Volume (cm3·g−1)
γ-Al2O3322.5604.7770.314
MnO2/γ-Al2O3295.5805.0930.297
Cu0.05Mn/γ-Al2O3304.8944.7600.289
Cu0.15Mn/γ-Al2O3303.4574.6140.316
Cu0.25Mn/γ-Al2O3294.6154.6250.277
Cu0.5Mn/γ-Al2O3234.7264.9550.233
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhu, Y.; Li, D.; Ji, C.; Si, P.; Liu, X.; Zhang, Y.; Liu, F.; Hua, L.; Han, F. Non-Thermal Plasma Incorporated with Cu-Mn/γ-Al2O3 for Mixed Benzene Series VOCs’ Degradation. Catalysts 2023, 13, 695. https://0-doi-org.brum.beds.ac.uk/10.3390/catal13040695

AMA Style

Zhu Y, Li D, Ji C, Si P, Liu X, Zhang Y, Liu F, Hua L, Han F. Non-Thermal Plasma Incorporated with Cu-Mn/γ-Al2O3 for Mixed Benzene Series VOCs’ Degradation. Catalysts. 2023; 13(4):695. https://0-doi-org.brum.beds.ac.uk/10.3390/catal13040695

Chicago/Turabian Style

Zhu, Yifan, Dandan Li, Chunjie Ji, Peizhuang Si, Xiaolin Liu, Yupeng Zhang, Fang Liu, Lei Hua, and Fenglei Han. 2023. "Non-Thermal Plasma Incorporated with Cu-Mn/γ-Al2O3 for Mixed Benzene Series VOCs’ Degradation" Catalysts 13, no. 4: 695. https://0-doi-org.brum.beds.ac.uk/10.3390/catal13040695

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