Next Article in Journal
g-C3N4-Stabilised Organic–Inorganic Halide Perovskites for Efficient Photocatalytic Selective Oxidation of Benzyl Alcohol
Next Article in Special Issue
Methane to Methanol through Heterogeneous Catalysis and Plasma Catalysis
Previous Article in Journal
Metal Oxide-Based Photocatalytic Paper: A Green Alternative for Environmental Remediation
Previous Article in Special Issue
Adsorption of VOCs Is a Key Step in Plasma-Catalyst Coupling: The Case of Acetone onto TiO2 vs. CeO2
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Removal of VOCs by Ozone: n-Alkane Oxidation under Mild Conditions

by
Alina I. Mytareva
*,
Igor S. Mashkovsky
,
Sergey A. Kanaev
,
Dmitriy A. Bokarev
,
Galina N. Baeva
,
Alexander V. Kazakov
and
Alexander Yu. Stakheev
N.D. Zelinsky Institute of Organic Chemistry Russian Academy of Sciences, Leninsky Prospect 47, 119991 Moscow, Russia
*
Author to whom correspondence should be addressed.
Catalysts 2021, 11(4), 506; https://doi.org/10.3390/catal11040506
Submission received: 31 March 2021 / Revised: 14 April 2021 / Accepted: 14 April 2021 / Published: 16 April 2021

Abstract

:
Volatile organic compounds (VOCs) have a negative effect on both humans and the environment; therefore, it is crucial to minimize their emission. The conventional solution is the catalytic oxidation of VOCs by air; however, in some cases this method requires relatively high temperatures. Thus, the oxidation of short-chain alkanes, which demonstrate the lowest reactivity among VOCs, starts at 250–350 °C. This research deals with the ozone catalytic oxidation (OZCO) of alkanes at temperatures as low as 25–200 °C using an alumina-supported manganese oxide catalyst. Our data demonstrate that oxidation can be significantly accelerated in the presence of a small amount of O3. In particular, it was found that n-C4H10 can be readily oxidized by an air/O3 mixture over the Mn/Al2O3 catalyst at temperatures as low as 25 °C. According to the characterization data (SEM-EDX, XRD, H2-TPR, and XPS) the superior catalytic performance of the Mn/Al2O3 catalyst in OZCO stems from a high concentration of Mn2O3 species and oxygen vacancies.

1. Introduction

Volatile organic compounds (VOCs) are a large group of chemical compounds with boiling points below 260 °C; they include alkanes, alkenes, aromatic hydrocarbons, aldehydes, ketones, alcohols, etc. [1]. VOCs are considered major air pollutants because of their significant contribution to the formation of photochemical smog, tropospheric O3 and secondary aerosols [2,3,4,5]. VOCs negatively affect human health because of their toxic, malodorous, mutagenic and carcinogenic nature [6]. The main anthropogenic sources of VOCs are vehicle exhaust fumes and emissions from chemical and power plants, petroleum refining, food and textile manufacturing, etc. [7]. Rapid urbanization and industrialization contribute to the growing emissions of VOCs into the environment. Therefore, the development of effective methods and materials for the abatement of VOCs is of significant importance.
Several technologies have been developed for VOC removal, including adsorption [8] and absorption [9] processes, thermal oxidation [1], photocatalytic decomposition [10,11,12], biological degradation [13], non-thermal plasma catalysis [14,15], catalytic combustion [7,16] and hybrid treatment [17,18]. However, insufficient selectivity and low energy efficiency limit their application. In recent years, ozone catalytic oxidation (OZCO) has attracted much attention due to its versatility in handling a range of organic emissions at relatively low operating temperatures [19,20,21].
Among various OZCO catalysts, supported MnOx demonstrates excellent VOC oxidation performance [22,23,24]. Einaga et al. compared M/Al2O3 and M/SiO2 (M = Mn, Fe, Co, Ni, Cu) according to their catalytic performances in OZCO [25,26]. It has been observed that Mn-containing catalysts are the most active for benzene and cyclohexane oxidation. Huang et al. investigated ozone catalytic oxidation of benzene over Mn, Co, Cu, Ni, Zn and Ce supported on ZSM-5 [27]. It has been shown that MnO2/ZSM-5 exhibits the best activity and CO2 selectivity (100% and 84.7%, respectively) at room temperature. Gopi et al. studied the OZCO of toluene over Mn, Ce, Cu, Ag and Co metal oxides supported on 13X zeolite [28]. It was revealed that the activity of the catalysts decreases in the following order: Mn/13X > Ce/13X > Cu/13X > Ag/13X > Co/13X. Chen et al. studied the ozonation of chlorobenzene over manganese supported on different carriers (Al2O3, TiO2, SiO2, CeO2, and ZrO2) to reveal the effect of carrier on the catalytic performance of MnOx in OZCO [29]. The authors found that Mn/Al2O3 exhibited the highest efficiency in chlorbenzene abatement.
The high efficiency of supported MnOx catalysts in the OZCO of VOCs can be attributed to their ability in terms of O3 decomposition [30]. It has been reported that highly reactive oxygen species such as atomic oxygen, peroxide, and OH radicals are formed (Equations (1)–(4)) on the surface of metal oxides as a result of O3 decomposition [30,31,32,33]. These species can completely oxidize adsorbed VOCs to CO2, even at room temperature (RT).
O3 + * → O2 + O*
O* + O3 → O2 + O2*
O2* → O2 + *
O* + H2O → 2OH*
Recently, it was shown that OZCO may be considered a promising solution for the abatement of methane, which is difficult to destroy compared to alkenes and aromatics [34,35]. Thus, the conversion of methane achieved by catalytic oxidation using an air/O3 mixture is almost four times higher than that using air only [35]. However, to the best of our knowledge, there are no available data on ozone catalytic oxidation of alkanes, which are the dominant VOCs generated by the petroleum industry [36,37]. Alkanes are also one of the major components among VOCs detected in vehicle exhaust [38]. Therefore, the aim of the present work was to investigate the possibility of ozone catalytic oxidation of n-butane (selected as a representative short-chain VOC molecule) at temperatures as low as 25–200 °C using an alumina-supported manganese oxide catalyst.

2. Results

2.1. Catalyst Characterization

Figure 1 shows the results of scanning electron microscopy (SEM) with energy-dispersive X-ray (EDX) spectroscopy analysis performed for the parent Al2O3 and the Mn/Al2O3 catalyst. The alumina (Figure 1a) has a relatively smooth surface. After loading 10 wt% Mn, the surface becomes rough with a number of defects (Figure 1b). The metal content is 9.4 wt% (Figure 1c), which is in a good agreement with data for inductively coupled plasma optical emission spectroscopy (ICP-OES) (see Section 4.1). The EDX mapping (Figure 1d) shows that Mn species are uniformly distributed on the catalyst surface.
X-ray powder diffraction (XRD) was used to identify the main MnOx phases of the Mn/Al2O3 catalyst. Figure 2 shows XRD patterns of the parent alumina and the catalyst. The characteristic reflections of γ-Al2O3 (2θ = 37°, 46° and 67°) are detected in both spectra, which is in a good agreement with the published literature [23,29]. In the Mn/γ-Al2O3 pattern, typical reflexes of MnO2 at 28.7°, 37.3°, 42.7° and 56.6° and Mn2O3 at 32.7° are observed, indicating that the two phases coexist [22,29]. The average particle sizes calculated from the diffraction peaks at 37.3° (MnO2) and 32.7° (Mn2O3) using the Debye-Scherrer equation are 19.0 nm and 9.0 nm, respectively.
The redox properties of Mn/Al2O3 were investigated using temperature programmed reduction (H2-TPR). A reduction profile composed of two peaks was observed (Figure 3a). The first peak, with a higher intensity at approximately 220–400 °C (Tmax = 325 °C), corresponds to the reduction of MnO2 to Mn2O3. The second peak, with Tmax = 431 °C, is attributed to the reduction of Mn2O3 to MnO [39].
To determine the valence states of the Mn atoms, the Mn/Al2O3 catalyst was examined by X-ray photoelectron spectroscopy (XPS). In the Mn 2p spectra (Figure 3b) of Mn/Al2O3, the peaks at 642.4, 641.8 and 640.8 eV were assigned to Mn4+, Mn3+ and traces of Mn2+, respectively [40,41,42]. The contents of Mn3+ and Mn4+ in the Mn/Al2O3 catalyst were approximately 69% and 31%, respectively.

2.2. Catalytic Activity

Three series of experiments were performed on the Mn/Al2O3 catalyst and the parent alumina: (a) n-C4H10 oxidation by air, (b) n-C4H10 oxidation by an air/O3 mixture (OZCO), and (c) O3 decomposition. In order to estimate the possible contribution of the gas-phase reactions, blank tests using an empty reactor were carried out.

2.2.1. n-C4H10 Oxidation

The blank oxidation reactions were carried out by air and air/O3 mixture. The results obtained are summarized in Figure 4a. Evidently, the gas-phase oxidation of n-C4H10 by air does not proceed, even at temperatures as high as 350 °C. Contrarily, OZCO starts at 125 °C, and 80–90% conversion of n-C4H10 is achieved at 290–350 °C.
Over the Mn/Al2O3 catalyst, n-C4H10 can be oxidized by air; however, oxidation starts only at 225–250 °C, and at 350°C the conversion does not exceed 20% (Figure 4a). In the course of OZCO, n-C4H10 is oxidized on the Mn/Al2O3 catalyst at 25 °C. Initially the conversion increases to 70% with the temperature and levels off at 85–200 °C (i.e., a plateau). At 200–250 °C, the conversion gradually decreases to ~55%; however, at 250–350 °C, it rises again to 85%.
It is known that the oxidation reaction may proceed not only on Mn-containing sites but also on the carrier surface. For this reason it was informative to evaluate the contribution of the parent Al2O3 (Figure 4a). The reaction between n-C4H10 and air does not proceed even at 350 °C. In the presence of O3 the alkane is oxidized at 100 °C; however, the conversion is much lower than on Mn/Al2O3. This observation allows us to conclude that the activity of alumina surface in OZCO is negligible compared to MnOx species.
Analysis of the reaction products reveals that n-C4H10 is completely oxidized by air over Mn/Al2O3; CO2 selectivity is almost 100% (Figure 4b). When n-C4H10 reacts with both air and O3 simultaneously, significant amounts of CO are formed. Thus, in the course of the gas-phase process and OZCO over the parent alumina, CO2 selectivity gradually decreases with the increase in temperature, and at 350 °C it becomes 66% and 44%, respectively. In contrast, when n-C4H10 is oxidized by the air/O3 mixture over the Mn/Al2O3 catalyst, CO mostly forms at relatively low temperatures (150–250 °C), and CO2 selectivity is ca. 70%. At 350 °C the selectivity rises to ca. 85%. Other byproducts as well as a carbon deposition on the catalyst surface were not detected, which is in good agreement with the high (>99%) carbon balance.

2.2.2. O3 Decomposition

Since the enhancement of the oxidation efficiency in the presence of O3 is attributable to a formation of highly reactive O atoms [29,32,33], catalytic O3 decomposition (over Mn/Al2O3 or Al2O3) as well as gas-phase decomposition (in the empty reactor) was also studied in detail. The results obtained are shown in Figure 5. The gas-phase O3 decomposition proceeds at relatively high temperatures (150–350 °C). In contrast, the reaction over the Mn/Al2O3 catalyst occurs at 25 °C, and the complete conversion of O3 is achieved at 250 °C. On the parent alumina, O3 decomposition starts only at temperatures above 100 °C.

2.2.3. The Effect of O3/n-C4H10 Ratio on the Oxidation Efficiency

The effect of O3 concentration on the n-C4H10 conversion on the Mn/Al2O3 catalyst was studied using feed gas containing different concentrations of O3 (0–800 ppm) while keeping the n-butane concentration constant (100 ppm). Thus, the O3/n-C4H10 ratio varies from 0 to 8. The results are summarized in Figure 6. The increase of n-butane conversion with O3 concentration is observed. The complete conversion is obtained at 125 °C using an O3/n-C4H10 ratio of eight.

3. Discussion

3.1. Structure of Mn/Al2O3 Catalyst

The characterization data suggest that the Mn/Al2O3 catalyst contains a mixture of manganese oxides. The XRD pattern (Figure 2) shows pronounced peaks of MnO2 along with a weak peak of Mn2O3, indicating that both oxide phases coexist in the catalyst [22,29]. The broadness of the MnO2 peaks allows us to suggest that the oxide particles have poor crystalline structure and may contain structural defects, including oxygen vacancies. Low intensity of the Mn2O3 peak indicates that the oxide phase is well dispersed on alumina surface.
The H2-TPR study of Mn/Al2O3 reveals an O/Mn stoichiometry of 1.71. This result agrees well with the XRD data, indicating the coexistence of MnO2 and Mn2O3 in the catalyst. Based on the assumption that all MnOx species are totally reduced to MnO, we estimate that the catalyst contains ca. 58% of Mn3+ and ca. 42% of Mn4+ species. Similar results were obtained by XPS, which shows that the surface of the Mn/Al2O3 catalyst is enriched with Mn3+ species.
The data obtained are in good agreement with previously reported results. Thus, it was shown repeatedly that the oxidation state of Mn species in Mn/Al2O3 catalysts with a low Mn loading (<10 wt%) is close to Mn3+, while at higher Mn loading (≥10 wt%) it tends to Mn4+ [22,26]. As reported in the literature [23,43], manganese in a low oxidation state is preferable for OZCO since it can transfer electrons to O3 more readily (Equations (5)–(7)), thus facilitating O3 decomposition.
O3 + Mnn+ → O2 + O + Mn(n+1)+
O3 + O + Mn(n+1)+ → O2 + O2 + Mn(n+1)+
O2 + Mn(n+1)+ → O2 + Mnn+
It is known that the structural defects of MnO2, such as oxygen vacancies, also favor OZCO due to their activity in O3 decomposition. According to the mechanism proposed in the literature [30,44,45,46,47], O3 molecules are adsorbed on the oxygen vacancies to form highly reactive oxygen species (O2−, O22−, O), which readily react with VOC molecules. Thus, it has been reported that the high catalytic activity of the Mn-Ag/HZSM-5 for toluene oxidation by ozone is attributed mainly to the abundant oxygen vacancies presented on the catalyst surface [48].

3.2. Catalytic Activity

It was shown that oxidation of n-C4H10 is significantly accelerated in the presence of small amounts of O3 (Figure 4a). The pronounced positive effect of O3 addition is observed in the both gas-phase and catalytic reactions. Thus, the oxidation of n-C4H10 by air in the gas-phase and on the parent alumina does not proceed, even at temperatures as high as 350 °C. On the Mn/Al2O3 catalyst, only a minor conversion is observed. O3 promotes oxidation in the gas-phase and on the alumina; the OZCO starts at 100–125 °C, and near-complete conversion is achieved at 300–350 °C. The effect of O3 is significantly more pronounced when the reaction proceeds on the Mn/Al2O3 catalyst. In this case, the light-off temperature of the n-C4H10 oxidation decreases from 250 °C to 25 °C.
It was found that the behavior of n-butane oxidation is correlated with the O3 decomposition process. Thus, it was observed that O3 readily decomposes over the Mn/Al2O3 catalyst at 25 °C, the same temperature the n-C4H10 oxidation by air/O3 mixture begins (Figure 4a and Figure 5, respectively). Similar parallels of light-off temperatures have been observed when O3 decomposition and n-C4H10 oxidation proceed in the gas-phase (at 150 °C) or over the alumina surface (at 100 °C). The results obtained are consistent with the published data [49]. Thus, it was found that at low temperatures the activation energy of benzene oxidation by O3 is similar to that of O3 decomposition, which suggests that O3 decomposition can be considered as the rate-determining step of the oxidation process.
Comparative analysis of the conversion profiles obtained during the n-C4H10 OZCO and O3 decomposition (Figure 4a and Figure 5) allows us to suggest that the alkane is oxidized over Mn/Al2O3 in two stages. In the first stage (25–250 °C) the reaction proceeds mainly over MnOx species, because the contribution of the alumina and the gas-phase reaction is negligible. The n-C4H10 is readily oxidized by O3; however, the maximum conversion does not exceed 70% at 85–200 °C. One of the possible reasons of the limitation is the consumption of active atomic oxygen species in the reaction with O3 (Equation (2)), which results in O2 formation [24,30,31]. Apparently, the contribution of this reaction becomes more pronounced at temperatures of 200–250 °C, and n-C4H10 conversion decreases to ~55%.
At 250–350 °C, n-C4H10 conversion rises again and reaches 85%. Comparison of the conversion profiles obtained in the gas-phase and over the Mn/Al2O3 catalyst allows us to suggest that at higher temperatures the gas-phase n-C4H10 oxidation by air/O3 mixture becomes the dominant catalytic pathway (Figure 4a). In addition, the n-C4H10 oxidation by air over Mn/Al2O3 may contribute at temperatures above 250 °C (Figure 4a). This fact is indicated by the higher CO2 selectivity as compared to the selectivity in the course of the gas-phase process and OZCO over the parent alumina (Figure 4b).
It has been shown that higher O3 concentration enhances the alkane oxidation efficiency (Figure 6). Thus, as the O3/n-C4H10 ratio increases from two to eight, the maximum n-C4H10 conversion obtained at low temperature (25–220 °C) rises from 43% to 100%. At the same time the plateau becomes wider, the decrease of conversion becomes less significant. According to the literature [24,50], both results are due to the increased availability of the highly reactive oxygen species for the oxidation of n-C4H10 over MnOx species.

4. Materials and Methods

4.1. Catalyst Preparation

The 10%Mn/Al2O3 catalyst was prepared by incipient wetness impregnation of commercial Al2O3 carrier (SASOL, Sandton, South Africa; SBET = 150 m2/g, calcined at 550 °C for 4 h) by an appropriate amount of an aqueous solution of manganese (II) nitrate tetrahydrate (≥97%, Sigma Aldrich, St. Louis, MO, USA). The catalyst was dried at room temperature overnight and calcined in air flow (550 °C, 3 h).
Mn content in the resultant Mn/Al2O3 catalyst was determined by inductively coupled plasma optical emission spectroscopy (ICP-OES) on a Perkin Elmer OPTIMA 2000 instrument (PerkinElmer Inc., Wellesley, MA, USA) and found to be 9.8 wt%.

4.2. Catalyst Characterization

The morphologies of the samples were studied using a Hitachi SU8000 field-emission scanning electron microscope (Hitachi, Tokyo, Japan). Before measurements the samples were mounted on a 25 mm aluminum specimen stub and fixed by conductive graphite adhesive tape. Images were acquired in secondary electron mode at a 2–20 kV accelerating voltage and at a working distance of 8–10 mm. The energy dispersive X-ray spectroscopy (EDX) of the samples was tested on the same SEM instrument equipped with an Oxford Instruments X-max EDX system (Oxford Instruments, Abingdon, OX, UK) to analyze the element distributions on the surface.
The crystallographic structure of the as-prepared catalysts was characterized by X-ray powder diffraction (XRD) using a D8 Advance diffractometer (Bruker AXS, Karlsruhe, Germany) with Bragg–Brentano geometry, Ni-filtered CuKα radiation, and LYNXEYE detector. The XRD patterns were recorded in the 2θ range 5–75° (scan rate 1.2°/min). Crystallographic parameters were calculated using the Rietan-FT software, which uses the Rietveld method.
H2-temperature programmed reduction (H2-TPR) was conducted on a USGA-101 instrument (Unisit, Moscow, Russia). The catalyst sample (100 mg) was pre-treated in Ar flow at 325 °C for 1 h and cooled down to RT. H2-TPR was carried out in a 5%H2/Ar mixture at a flow rate of 30 mL/min. The temperature was ramped linearly from RT to 820 °C at a heating rate of 10 °C/min. The data were processed using the Data Treatment software of the USGA-101 instrument. H2 consumption was quantified using CuO (99.999%, Sigma Aldrich, St. Louis, MO, USA) and NiO (99.99%, Sigma Aldrich, St. Louis, MO, USA) standards. The H2-TPR pattern of the Mn/Al2O3 catalyst was corrected using the H2-TPR profile of the parent alumina as a background.
The X-ray photoelectron spectroscopy (XPS) spectra of the samples were recorded on Axis Ultra DLD spectrometer (Kratos Analytical Limited, Manchester, UK) with a monochromatic Al Kα source (1486.6 eV, 150 W). The binding energies were calibrated with reference to the standard binding energy of contaminant carbon (C1s 284.5 eV).

4.3. Catalytic Tests

A scheme of the experimental setup is shown in Figure 7. N-butane was supplied from a 0.98% n-C4H10/N2 cylinder (Linde Gas Rus, Balashikha, Russia) balanced with N2. O3 was generated from air using a Triozon OG-A0.5 ozone generator (TriOzon, Moscow, Russia). The concentrations of n-butane and O3 were 100 and 450 ppm, respectively (unless otherwise stated). The total gas flow rate was 750 mL/min, with a corresponding gas hourly space velocity (GHSV) of 155.000 h−1. The gas mixture passed through a reactor (a quartz tube with an inner diameter of 10 mm) with the catalyst inside (0.2 g, fraction 0.4–1.0 mm).
The ozone concentration before and after the reactor was measured with an UV-100 ozone analyzer (Eco Sensors Division of KWJ Engineering Inc., Newark, CA, USA) in the range of 0–999 ppm with ±2% accuracy. For the correct analysis of the carbon-containing reaction products, an ozone trap with a saturated aqueous solution of KI was placed at the reactor outlet. An FTIR spectrometer Gasmet DX 4000 (Temet Instruments Oy, Helsinki, Finland) was used to detect concentrations of CO, CO2, and n-C4H10.
Catalyst activity (conversion) was studied as a function of temperature ranging from 25 to 350 °C (temperature ramp of 5 °C/min). The conversions (X) of n-C4H10 and O3 were calculated using the following equation:
X ( % ) = C i n C o u t C i n × 100 % ,
where Cin and Cout are the inlet and outlet concentrations of n-C4H10 and O3, respectively.
Since only COx (CO, CO2) were detected as the carbon-containing products of n-C4H10 oxidation, CO2 formation was selected as a criterion for evaluating the effectiveness of the reaction. The CO2 selectivity was calculated by the following equations:
C O 2 s e l e c t i v i t y ( % ) = C ( C O 2 ) o u t ( C ( C 4 H 10 ) i n C ( C 4 H 10 ) o u t ) × m × 100 % ,
where Cin and Cout are the inlet and outlet concentrations of n-C4H10 and CO2, respectively; m is the number of carbon atoms in the hydrocarbon (for n-C4H10, m = 4).
All carbon balances were in the range (100 ± 5)%.

5. Conclusions

The effect of O3 on the efficiency of alkane oxidation was studied in details. It was shown for the first time that it is possible to oxidize short-chain VOCs by air/O3 mixture at ambient temperatures. It was observed that n-C4H10 reacted readily with O3 over the Mn/Al2O3 catalyst at temperatures as low as 25 °C, and 80–100% conversion could be achieved in the temperature range 70–200 °C (O3/n-C4H10 = 8).
Comparison of the catalytic and characterization data (SEM-EDX, XRD, H2-TPR, and XPS) revealed that superior catalytic performance of the Mn/Al2O3 catalyst in OZCO was due to high concentration of Mn2O3 species and oxygen vacancies.
The data obtained indicated that the oxidation pathway depends on the reaction temperature. Thus, at low temperatures n-C4H10 was oxidized by active oxygen, which was formed via O3 decomposition over MnOx species, while at 250–350 °C the contributions of the gas-phase oxidation by air/O3 mixture and conventional oxidation by air over Mn-containing active centers became significant.
The findings of this study may contribute to the development of the efficient technologies for the low-temperature abatement of the short-chain hydrocarbons.

Author Contributions

Conceptualization, A.I.M. and A.Y.S.; investigation, D.A.B. and S.A.K.; data curation, I.S.M.; catalyst characterization, G.N.B. and A.V.K.; writing—original draft preparation, A.I.M.; writing—review and editing, A.Y.S. and D.A.B.; visualization, I.S.M.; supervision, A.Y.S. All authors have read and agreed to the published version of the manuscript.

Funding

Scientific Schools Development Program by Zelinsky Institute of Organic Chemistry is gratefully acknowledged.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

Authors gratefully acknowledge the Department of Structural Studies of N.D. Zelinsky Institute of Organic Chemistry for electron microscopy characterization.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Huang, H.; Xu, Y.; Feng, Q.; Leung, D.Y.C. Low temperature catalytic oxidation of volatile organic compounds: A review. Catal. Sci. Technol. 2015, 5, 2649–2669. [Google Scholar] [CrossRef]
  2. Berezina, E.; Moiseenko, K.; Skorokhod, A.; Pankratova, N.V.; Belikov, I.; Belousov, V.; Elansky, N.F. Impact of VOCs and NOx on ozone formation in Moscow. Atmosphere 2020, 11, 1262. [Google Scholar] [CrossRef]
  3. Juráň, S.; Grace, J.; Urban, O. Temporal changes in ozone concentrations and their impact on vegetation. Atmosphere 2021, 12, 82. [Google Scholar] [CrossRef]
  4. Juráň, S.; Edwards-Jonášová, M.; Cudlín, P.; Zapletal, M.; Šigut, L.; Grace, J.; Urban, O. Prediction of ozone effects on net ecosystem production of Norway spruce forest. iForest-Biogeosci. For. 2021, 11, 743–750. [Google Scholar] [CrossRef] [Green Version]
  5. Wang, H.; Wang, Q.; Gao, Y.; Zhou, M.; Jing, S.; Qiao, L.; Yuan, B.; Huang, D.; Huang, C.; Lou, S.; et al. Estimation of secondary organic aerosol formation during a photochemical smog episode in Shanghai, China. JGR Atmos. 2020, 125, e2019JD032033. [Google Scholar] [CrossRef]
  6. Okal, J.; Zawadzki, M. Catalytic combustion of butane on Ru/γ-Al2O3 catalysts. Appl. Catal. B Environ. 2009, 89, 22–32. [Google Scholar] [CrossRef]
  7. Kamal, M.S.; Razzak, S.A.; Hossain, M.M. Catalytic oxidation of volatile organic compounds (VOCs)—A review. Atmos. Environ. 2016, 140, 117–134. [Google Scholar] [CrossRef]
  8. Zhu, L.; Shen, D.; Luo, K.H. A critical review on VOCs adsorption by different porous materials: Species, mechanisms and modification methods. J. Hazard. Mater. 2020, 389, 122102. [Google Scholar] [CrossRef]
  9. Ozturk, B.; Yilmez, D. Absorptive removal of volatile organic compounds from flue gas streams. Process Saf. Environ. Prot. 2006, 84, 391–398. [Google Scholar] [CrossRef]
  10. Huang, Y.; Ho, S.S.H.; Lu, Y.; Niu, R.; Xu, L.; Cao, J.; Lee, S. Removal of indoor volatile organic compounds via photocatalytic oxidation: A short review and prospect. Molecules 2016, 21, 56. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Higashimoto, S.; Katsuura, K.; Yamamoto, M.; Takahashi, M. Photocatalytic activity for decomposition of volatile organic compound on Pt-WO3 enhanced by simple physical mixing with TiO2. Catal. Commun. 2019, 133, 105831. [Google Scholar] [CrossRef]
  12. Sansotera, M.; Geran Malek Kheyli, S.; Baggioli, A.; Bianchi, C.L.; Pedeferri, M.P.; Diamanti, M.V.; Navarrini, V. Absorption and photocatalytic degradation of VOCs by perfluorinated ionomeric coating with TiO2 nanopowders for air purification. Chem. Eng. J. 2019, 361, 885–896. [Google Scholar] [CrossRef] [Green Version]
  13. Zhang, S.; You, J.; Kennes, C.; Cheng, Z.; Ye, J.; Chen, D.; Chen, J.; Wang, L. Current advances of VOCs degradation by bioelectrochemical systems: A review. Chem. Eng. J. 2018, 334, 2625–2637. [Google Scholar] [CrossRef]
  14. Cools, P.; De Geyter, N.; Morent, R. Plasma-catalytic removal of VOCs. In Plasma Catalysis. Springer Series on Atomic, Optical, and Plasma Physics; Tu, X., Whitehead, J., Nozaki, T., Eds.; Springer: Cham, Switzerland, 2019; Volume 106, pp. 145–180. [Google Scholar] [CrossRef]
  15. Wang, B.; Xu, X.; Xu, W.; Wang, N.; Xiao, H.; Sun, Y.; Huang, H.; Yu, L.; Fu, M.; Wu, J.; et al. The Mechanism of non-thermal plasma catalysis on volatile organic compounds removal. Catal. Surv. Asia 2018, 22, 73–94. [Google Scholar] [CrossRef]
  16. Yang, C.; Miao, G.; Pi, Y.; Xia, Q.; Wu, J.; Li, Z.; Xiao, J. Abatement of various types of VOCs by adsorption/catalytic oxidation: A review. Chem. Eng. J. 2019, 370, 1128–1153. [Google Scholar] [CrossRef]
  17. Lee, J.E.; Ok, Y.S.; Tsang, D.C.W.; Song, J.; Jung, S.-C.; Park, Y.-K. Recent advances in volatile organic compounds abatement by catalysis and catalytic hybrid processes: A critical review. Sci. Total Environ. 2020, 719, 137405. [Google Scholar] [CrossRef]
  18. Shu, Y.; Xu, Y.; Huang, H.; Ji, J.; Liang, S.; Wu, M.; Leung, D.Y.C. Catalytic oxidation of VOCs over Mn/TiO2/activated carbon under 185 nm VUV irradiation. Chemosphere 2018, 208, 550–558. [Google Scholar] [CrossRef]
  19. Aghbolaghy, M.; Ghavami, M.; Soltan, J.; Chen, N. Effect of active metal loading on catalyst structure and performance in room temperature oxidation of acetone by ozone. J. Ind. Eng. Chem. 2019, 77, 118–127. [Google Scholar] [CrossRef]
  20. Fang, R.; Huang, H.; Huang, W.; Ji, J.; Feng, Q.; Shu, Y.; Zhan, Y.; Liu, G.; Xie, R. Influence of peracetic acid modification on the physicochemical properties of activated carbon and its performance in the ozone-catalytic oxidation of gaseous benzene. Appl. Surf. Sci. 2017, 420, 905–910. [Google Scholar] [CrossRef]
  21. Sun, W.; Gao, X.; Wu, B.; Ombrello, T. The effect of ozone addition on combustion: Kinetics and dynamics. Prog. Energy Combust. Sci. 2019, 73, 1–25. [Google Scholar] [CrossRef]
  22. Rezaei, E.; Soltan, J.; Chen, N. Catalytic oxidation of toluene by ozone over alumina supported manganese oxides: Effect of catalyst loading. Appl. Catal. B Environ. 2013, 136–137, 239–247. [Google Scholar] [CrossRef]
  23. Zhu, B.; Li, X.-S.; Sun, P.; Liu, J.-L.; Ma, X.-Y.; Zhu, X.; Zhu, A.-M. A novel process of ozone catalytic oxidation for low concentration formaldehyde removal. Chin. J. Catal. 2017, 38, 1759–1769. [Google Scholar] [CrossRef]
  24. Xi, Y.; Reed, C.; Lee, Y.-K.; Oyama, S.T. Acetone oxidation using ozone on manganese oxide catalysts. J. Phys. Chem. B 2005, 109, 17587–17596. [Google Scholar] [CrossRef] [PubMed]
  25. Einaga, H.; Futamura, S. Comparative study on the catalytic activities of alumina-supported metal oxides for oxidation of benzene and cyclohexane with ozone. React. Kinet. Catal. Lett. 2004, 81, 121–128. [Google Scholar] [CrossRef]
  26. Einaga, H.; Maeda, N.; Nagai, Y. Comparison of catalytic properties of supported metal oxides for benzene oxidation using ozone. Catal. Sci. Technol. 2015, 5, 3147–3158. [Google Scholar] [CrossRef]
  27. Huang, H.; Ye, X.; Huang, W.; Chen, J.; Xu, Y.; Wu, M.; Shao, Q.; Peng, Z.; Ou, G.; Shi, J.; et al. Ozone-catalytic oxidation of gaseous benzene over MnO2/ZSM-5 at ambient temperature: Catalytic deactivation and its suppression. Chem. Eng. J. 2015, 264, 24–31. [Google Scholar] [CrossRef]
  28. Gopi, T.; Swetha, G.; Chandra Shekar, S.; Krishna, R.; Ramakrishna, C.; Saini, B.; Rao, P.V.L. Ozone catalytic oxidation of toluene over 13X zeolite supported metal oxides and the effect of moisture on the catalytic process. Arab. J. Chem. 2019, 12, 4502–4513. [Google Scholar] [CrossRef] [Green Version]
  29. Chen, G.; Wang, Z.; Lin, F.; Zhang, Z.; Yu, H.; Yan, B. Comparative investigation on catalytic ozonation of VOCs in different types over supported MnOx catalysts. J. Hazard. Mater. 2020, 391, 122218. [Google Scholar] [CrossRef]
  30. Li, W.; Gibbs, G.V.; Oyama, S.T. Mechanism of ozone decomposition on a manganese oxide catalyst. 1. In situ Raman spectroscopy and Ab initio molecular orbital calculations. J. Am. Chem. Soc. 1998, 120, 9041–9046. [Google Scholar] [CrossRef]
  31. Li, W.; Oyama, S.T. Mechanism of ozone decomposition on a manganese oxide catalyst. 2. Steady-state and transient kinetic studies. J. Am. Chem. Soc. 1998, 120, 9047–9052. [Google Scholar] [CrossRef]
  32. Ramakrishna, C.; Shekar, S.C.; Gupta, A.K.; Saini, B.; Krishna, R.; Swetha, G.; Gopi, T. Degradation of diethyl sulfide vapors with manganese oxide catalysts supported on zeolite-13X: The influence of process parameters and mechanism in presence of ozone. J. Environ. Chem. Eng. 2017, 5, 1484–1493. [Google Scholar] [CrossRef]
  33. Sekiguchi, K.; Kurita, Y.; Sankoda, K.; Namiki, N.; Yasui, F.; Tamura, H. Ozone catalytic oxidation of gaseous toluene over MnO2-based ozone decomposition catalysts immobilized on a nonwoven fabric. Aerosol Air Qual. Res. 2017, 17, 2110–2118. [Google Scholar] [CrossRef] [Green Version]
  34. Wan, M.P.; Hui, K.S.; Chao, C.Y.H.; Kwong, C.W. Catalytic combustion of methane with ozone using Pd-exchanged zeolite X: Experimental investigation and kinetics model. Combust. Sci. Technol. 2010, 182, 1429–1445. [Google Scholar] [CrossRef]
  35. Hui, K.S.; Kwong, C.W.; Chao, C.Y.H. Methane emission abatement by Pd-ion-exchanged zeolite 13X with ozone. Energy Environ. Sci. 2010, 3, 1092–1098. [Google Scholar] [CrossRef]
  36. Chen, C.-L.; Fang, H.Y.; Shu, C.-M. Source location and characterization of volatile organic compound emissions at a petrochemical plant in Kaohsiung, Taiwan. J. Air Waste Manag. Assoc. 2005, 55, 1487–1497. [Google Scholar] [CrossRef]
  37. Ragothaman, A.; Anderson, W.A. Air quality impacts of petroleum refining and petrochemical industries. Environments 2017, 4, 66. [Google Scholar] [CrossRef] [Green Version]
  38. Deng, C.; Jin, Y.; Zhang, M.; Liu, X.; Yu, Z. Emission characteristics of VOCs from on-road vehicles in an urban tunnel in Eastern China and predictions for 2017–2026. Aerosol Air Qual. Res. 2018, 18, 3025–3034. [Google Scholar] [CrossRef]
  39. Ryu, H.W.; Song, M.Y.; Park, J.S.; Kim, J.M.; Jung, S.-C.; Song, J.-H.; Kim, B.-J.; Park, Y.-K. Removal of toluene using ozone at room temperature over mesoporous Mn/Al2O3 catalysts. Environ. Res. 2019, 172, 649–657. [Google Scholar] [CrossRef]
  40. Langell, M.A.; Hutchings, C.W.; Carson, G.A.; Nassir, M.H. High resolution electron energy loss spectroscopy of MnO(100) and oxidized MnO(100). J. Vac. Sci. Technol. A 1996, 14, 1656–1661. [Google Scholar] [CrossRef] [Green Version]
  41. Stranick, M.A. MnO2 by XPS. Surf. Sci. Spectra 1999, 6, 31–38. [Google Scholar] [CrossRef]
  42. Stranick, M.A. Mn2O3 by XPS. Surf. Sci. Spectra 1999, 6, 39–46. [Google Scholar] [CrossRef]
  43. Radhakrishnan, R.; Oyama, S.T.; Chen, J.G.; Asakura, K. Electron transfer effects in ozone decomposition on supported manganese oxide. J. Phys. Chem. B 2001, 105, 4245–4253. [Google Scholar] [CrossRef]
  44. Li, X.; Ma, J.; Yang, L.; He, G.; Zhang, C.; Zhang, R.; He, H. Oxygen vacancies induced by transition metal doping in γ-MnO2 for highly efficient ozone decomposition. Environ. Sci. Technol. 2018, 52, 12685–12696. [Google Scholar] [CrossRef] [PubMed]
  45. Zhu, G.; Zhu, J.; Jiang, W.; Zhang, Z.; Wang, J.; Zhu, Y.; Zhang, Q. Surface oxygen vacancy induced α-MnO2 nanofiber for highly efficient ozone elimination. Appl. Catal. B Environ. 2017, 209, 729–737. [Google Scholar] [CrossRef]
  46. Peng, B.; Bao, W.; Wei, L.; Zhang, R.; Wang, Z.; Wang, Z.; Wie, Y. Highly active OMS-2 for catalytic ozone decomposition under humid conditions. Pet. Sci. 2019, 16, 912–919. [Google Scholar] [CrossRef] [Green Version]
  47. Ding, Y.H.; Zhang, X.L.; Chen, L.; Wang, X.R.; Zhang, N.; Liu, Y.F.; Fang, Y.Z. Oxygen vacancies enabled enhancement of catalytic property of Al reduced anatase TiO2 in the decomposition of high concentration ozone. J. Solid State Chem. 2017, 250, 121–127. [Google Scholar] [CrossRef]
  48. Li, J.; Na, H.; Zeng, X.; Zhu, T.; Liu, Z. In situ DRIFTS investigation for the oxidation of toluene by ozone overMn/HZSM-5, Ag/HZSM-5 and Mn–Ag/HZSM-5. Appl. Surf. Sci. 2014, 311, 690–696. [Google Scholar] [CrossRef]
  49. Naydenov, A.; Mehandjiev, D. Complete oxidation of benzene on manganese dioxide by ozone. Appl. Catal. A Gen. 1993, 97, 17–22. [Google Scholar] [CrossRef]
  50. Kwong, C.W.; Chao, C.Y.H.; Hui, K.S.; Wan, M.P. Catalytic ozonation of toluene using zeolite and MCM-41 materials. Environ. Sci. Technol. 2008, 42, 8504–8509. [Google Scholar] [CrossRef]
Figure 1. SEM micrographs of (a) the parent alumina and (b) the Mn/Al2O3 catalyst; (c) EDX spectrum and the elemental distribution maps of (d) Mn, (e) O and (f) Al for the Mn/Al2O3 catalyst.
Figure 1. SEM micrographs of (a) the parent alumina and (b) the Mn/Al2O3 catalyst; (c) EDX spectrum and the elemental distribution maps of (d) Mn, (e) O and (f) Al for the Mn/Al2O3 catalyst.
Catalysts 11 00506 g001
Figure 2. XRD patterns of (a) the parent alumina and (b) the Mn/Al2O3 catalyst.
Figure 2. XRD patterns of (a) the parent alumina and (b) the Mn/Al2O3 catalyst.
Catalysts 11 00506 g002
Figure 3. (a) H2-TPR and (b) Mn 2p XPS spectra of the Mn/Al2O3 catalyst.
Figure 3. (a) H2-TPR and (b) Mn 2p XPS spectra of the Mn/Al2O3 catalyst.
Catalysts 11 00506 g003
Figure 4. The gas-phase and catalytic oxidation of n-C4H10 by air and air/O3 mixture: (a) n-C4H10 conversion and (b) CO2 selectivity. Conditions: 100 ppm n-C4H10, air or 450 ppm O3, balanced with N2; Ftotal = 750 mL/min (GHSV = 155.000 h−1).
Figure 4. The gas-phase and catalytic oxidation of n-C4H10 by air and air/O3 mixture: (a) n-C4H10 conversion and (b) CO2 selectivity. Conditions: 100 ppm n-C4H10, air or 450 ppm O3, balanced with N2; Ftotal = 750 mL/min (GHSV = 155.000 h−1).
Catalysts 11 00506 g004
Figure 5. The gas-phase and catalytic O3 decomposition. Conditions: 450 ppm O3, balanced with N2; Ftotal = 750 mL/min (GHSV = 155.000 h−1).
Figure 5. The gas-phase and catalytic O3 decomposition. Conditions: 450 ppm O3, balanced with N2; Ftotal = 750 mL/min (GHSV = 155.000 h−1).
Catalysts 11 00506 g005
Figure 6. The OZCO of n-C4H10 at different O3/n-C4H10 ratios. Conditions: 100 ppm n-C4H10, 200–800 ppm O3, balanced with N2; Ftotal = 750 mL/min (GHSV = 155.000 h−1).
Figure 6. The OZCO of n-C4H10 at different O3/n-C4H10 ratios. Conditions: 100 ppm n-C4H10, 200–800 ppm O3, balanced with N2; Ftotal = 750 mL/min (GHSV = 155.000 h−1).
Catalysts 11 00506 g006
Figure 7. Scheme of the experimental setup.
Figure 7. Scheme of the experimental setup.
Catalysts 11 00506 g007
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mytareva, A.I.; Mashkovsky, I.S.; Kanaev, S.A.; Bokarev, D.A.; Baeva, G.N.; Kazakov, A.V.; Stakheev, A.Y. Removal of VOCs by Ozone: n-Alkane Oxidation under Mild Conditions. Catalysts 2021, 11, 506. https://doi.org/10.3390/catal11040506

AMA Style

Mytareva AI, Mashkovsky IS, Kanaev SA, Bokarev DA, Baeva GN, Kazakov AV, Stakheev AY. Removal of VOCs by Ozone: n-Alkane Oxidation under Mild Conditions. Catalysts. 2021; 11(4):506. https://doi.org/10.3390/catal11040506

Chicago/Turabian Style

Mytareva, Alina I., Igor S. Mashkovsky, Sergey A. Kanaev, Dmitriy A. Bokarev, Galina N. Baeva, Alexander V. Kazakov, and Alexander Yu. Stakheev. 2021. "Removal of VOCs by Ozone: n-Alkane Oxidation under Mild Conditions" Catalysts 11, no. 4: 506. https://doi.org/10.3390/catal11040506

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