Next Article in Journal
Study on the Formaldehyde Oxidation Reaction of Acid-Treated Manganese Dioxide Nanorod Catalysts
Next Article in Special Issue
The Influence of Cerium to Manganese Ratio and Preparation Method on the Activity of Ceria-Manganese Mixed Metal Oxide Catalysts for VOC Total Oxidation
Previous Article in Journal
Advances in Catalytic C–F Bond Activation and Transformation of Aromatic Fluorides
Previous Article in Special Issue
Protection Effect of Ammonia on CeNbTi NH3-SCR Catalyst from SO2 Poisoning
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Catalytic Degradation of Toluene over MnO2/LaMnO3: Effect of Phase Type of MnO2 on Activity

1
State Key Laboratory of Electrical Insulation and Power Equipment, Center of Nanomaterials for Renewable Energy, School of Electrical Engineering, Xi’an Jiaotong University, Xi’an 710049, China
2
School of Chemistry and Chemical Engineering, Nantong University, 9, Seyuan Road, Nantong 226019, China
3
School of Environmental Science and Engineering, Shanghai Jiao Tong University, 800, Dong Chuan Road, Shanghai 200240, China
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(12), 1666; https://doi.org/10.3390/catal12121666
Submission received: 13 November 2022 / Revised: 10 December 2022 / Accepted: 13 December 2022 / Published: 18 December 2022

Abstract

:
Series of α, β, γ, δ type MnO2 supported on LaMnO3 perovskite was developed by a one-pot synthesis route. Compared with α-MnO2, β-MnO2, γ-MnO2, δ-MnO2 and LaMnO3 oxides, all MnO2/LaMnO3 showed promotional catalytic performance for toluene degradation. Among them, α-MnO2/LaMnO3 holds the best active and mineralization efficiency. By the analysis of N2 adsorption-desorption, XPS and H2-TPR, it can be inferred that the improved activity should be ascribed to the higher proportion of lattice oxygen, better low-temperature reducibility and larger specific surface area. Besides, the byproducts from the low-temperature reaction of toluene oxidation were detected by a TD/GC-MS, confirming the presence of the intermediates. Combined with the in-situ DRIFTS, the catalytic degradation path of toluene oxidation has also been discussed in depth.

1. Introduction

Volatile organic compounds (VOCs) are harmful pollutants released into the environment resulting from a variety of commercial, industrial, and domestic practices [1,2,3]. Among them, toluene is a ubiquitous aromatic VOC and an important contributor to the formation of ozone (O3) as well as a secondary organic aerosol (SOA) [2,4]. Meanwhile, toluene is easily absorbed by the skin and mucous membranes, which has devastating effects on human organs and metabolic systems [5]. Many traditional technologies, such as condensation, membrane separation, absorption, thermal incineration, and catalytic oxidation, have been utilized to control VOCs [6,7,8], wherein catalytic oxidation has been believed to be one of the most effective strategies, which could completely convert pollutants into harmfulness final products, carbon dioxide and water [9]. Catalysts are the key factor in catalytic oxidation technology. Generally, noble metal and non-precious metal catalysts are the two major types of materials applied in toluene catalytic combustion. However, noble metals, such as gold, palladium, platinum, ruthenium, etc., are easily restricted by their high cost, sintering, and poisoning in industrial applications [10]. On the contrary, non-noble mixed metal oxides have been considered universal candidates due to their wide source, low price, relatively high activity, and good anti-toxicity [11,12].
Metal oxides/perovskites (MOx/ABO3) are a typical class of non-noble mixed metal oxide compounds, especially Mn-based MnO2/AMnO3, which have been widely used in the fields of denitration, desulfurization and VOCs oxidation [13,14,15]. Si et al. synthesized a γ-MnO2/LaMnO3 catalyst by selectively removing surficial La cation from LaMnO3 and obtained a superior catalytic performance on toluene oxidation [16]. Yang et al. studied the LaMnO3 perovskite assembled by δ-MnO2 through a gunpowder-like combustion method and pointed out that the boosted active phase-carrier interplays, improved redox abilities led to good performance for toluene oxidation [17]. We also synthesized a γ-MnO2/SmMnO3 material by in-situ acid etching of SmMnO3 and found that the composite of γ-MnO2 was successfully exposed on the surface of SmMnO3, which improved the physicochemical properties, thus promoted toluene oxidation capacity [18]. Kim et al. [19] paid attention to different manganese valent oxides Mn3O4, Mn2O3, and MnO2 in the catalytic combustion of benzene and toluene. Figueredo et al. [20] also made a comparison of different manganese oxides Mn3O4, Mn2O3, and their biphasic compound Mn3O4/Mn2O3 material on the catalytic evaluation of ethene and propene. However, little attention is paid to digging out the effect of different phase types of manganese dioxide supported on perovskites, as well as the degree of activating lattice oxygen and catalytic activity. Meanwhile, a simple modulation process of MnO2 on the perovskite surface is also particularly important due to the practical point of view. Besides, the catalytic mechanism of catalysts and the degradation path of toluene still require great efforts to clarify.
Based on the above challenges, herein we provide a relatively simple synthesis route for MnO2/LaMnO3 preparation, in which the phase type (α, β, δ, and γ) of MnO2 can be directly adjusted by controlling the proportion of manganese source precursors. Meanwhile, the activation degree of lattice oxygen on the surface could also be controlled by adjusting the surficial phase type of MnO2 of LaMnO3. Combined with the characterizations of XRD, SEM, N2 physisorption, XPS, H2-TPR, etc., the influence of the phases of MnO2 in LaMnO3 on the physical and chemical properties and the catalytic activities for toluene degradation were explored, as well as the internal relationship of structure and catalytic behavior was also in-depth discussion. Finally, the types of byproducts by the low-temperature catalytic oxidation of toluene and the degradation path of toluene oxidation were discussed.

2. Results and Discussion

2.1. Morphology and Crystal Phase Structure

The crystal structures of samples were characterized by XRD. As shown in Figure S1, the diffraction peaks at 2θ = 22.94°, 32.78°, 38.47°, 40.16°, 40.64°, 46.84°, 52.62°, 52.96°, 58.10°, 58.70°, 68.02, 68.70 and 78.10°, corresponding to the (012), (110), (113), (202), (006), (024), (122), (116), (214), (018), (220), (208), and (128) planes of LaMnO3 (PDF# 50-0298) [21]. Meanwhile, as illustrated in Figure 1, it was found that the new peaks of α-MO/LMO (Figure 1a) at 2θ = 28.84° (310) and 37.5° (211), β-MO/LMO (Figure 1b) at 2θ = 28.68° (110), 37.32° (101), and 42.8° (111), γ-MO/LMO (Figure 1c) at 2θ = 22.4° (120), 34.5° (031), 37.14° (131), and 57.4° (160), and δ-MO/LMO (Figure 1d) at 2θ = 12.68° (001), and 25.2° (002), corresponded to the standard cards of PDF# 44-0141 (α-MnO2) [22], PDF# 24-0735 (β-MnO2) [23], PDF# 80-1098 (δ-MnO2) [24], and PDF# 14-0644 (γ-MnO2) [18], respectively, showing that the different phases types (α, β, δ, and γ) MnO2 were successfully combined with LaMnO3 perovskite. In addition, the full width at half maximum (FWHM) of the characteristic peaks at 2θ = 32.78° are 0.38° for LMO, 0.37° for α-MO/LMO, 0.31° for β-MO/LMO, 0.29° for δ-MO/LMO and 0.28° for γ-MO/LMO. Obviously, the FWHM at 2θ = 32.78° of α-MO/LMO is closer to that of LMO, indicating that the composite degree of α-MnO2 on the surface of LaMnO3 is better, which is beneficial to inhibit the further agglomeration of LaMnO3 during the hydrothermal reaction.
Figure 2a–j shows the SEM images of LMO, α-MO/LMO, β-MO/LMO, δ-MO/LMO and γ-MO/LMO. As can be seen in Figure 2a,b, the morphology of the LMO sample exhibits a bulk structure with large pore sizes, indicating that the high-temperature preparation process caused the agglomeration of the LMO catalyst. When the α-MnO2 was combined with LaMnO3, the composite showed an obvious porous structure (Figure 2c), and the surface of the LaMnO3 framework was covered with a furry substance (Figure 2d), which should be the α-MnO2 with stacking-nanoneedle structure [23]. These changes may imply an increase in the specific surface area of the catalyst, which is conducive to the contact between toluene molecules and the catalyst, promoting the catalytic oxidation of toluene. In addition, it can be observed from the SEM images of β-MO/LMO in Figure 2e,f that the morphology of β-MO/LMO is disorganized and a small amount of sheet-like structures are mixed in the sample, indicating that the presence of β-MnO2 in the composite [23,25]. Similarly, the large-scale flower-like structures are also found in Figure 2g,h, confirming the formation of δ-MnO2 in the δ-MO/LMO sample [26,27]. However, when γ-MnO2 and LaMnO3 were combined, lots of ca. 50 nm nano-spherical particles and large-sized agglomerates appeared in the SEM image (Figure 2i,j). Combined with the analysis of XRD, it can be inferred that the nano-spherical particles should belong to γ-MnO2.

2.2. Catalytic Performance

Figure 3a demonstrates the toluene oxidation over LMO, α-MO/LMO, β-MO/LMO, δ-MO/LMO, and γ-MO/LMO. Before testing the properties of all samples, the blank experiment was performed without a catalyst, and the conversion was not found (<320 °C), indicating that there was no occurrence of homogeneous reactions under the adopted reaction [18,28]. The catalytic activities of all samples hoisted with the reaction temperature and the heightened sequence were as follows: LMO < γ-MO/LMO < δ-MO/LMO < β-MO/LMO < α-MO/LMO. Obviously, after loading manganese dioxide, the performance of new catalysts improved. Moreover, it can be seen from Figure S2 that the activity of each composite was also higher than that of the corresponding pure MnO2. In order to further analyze the catalytic activities of all catalysts, the T90, and T50 values (reaction temperature vs. conversion of 50%, and 90%) were used to further assess their performance. As listed in Table 1, the T90, and T50 values were 260, and 237 °C for α-MO/LMO, 289, and 246 °C for β-MO/LMO, 294, and 255 °C for δ-MO/LMO, 316, and 261 °C for γ-MO/LMO, >320, and 300 °C for LMO, 295, and 252 °C for α-MO, > 320, and 302 °C for β-MnO2, > 320, and 289 °C for δ-MO, and > 320, and 286 °C for γ-MO, respectively. Obviously, α-MO/LMO holds lower T90, and T50 values, indicating a better performance for toluene oxidation.
Since the specific surface areas (SBET) are quite different (shown in Table 2) and thermal effect may also exist at higher DCE conversion, the apparent reaction rates of these catalysts are calculated at low temperatures (210 °C) to reflect their inherent catalytic activities. As listed in Table 1, the values decrease in the sequence of α-MO/LMO > β-MO/LMO > LMO > δ-MO/LMO > γ-MO/LMO, which is basically consistent with the order for conversion efficiency, while LMO showed a higher apparent reaction rate than that of δ- and γ- phased samples.
To better evaluate the destruction ability of prepared catalysts, the mineralization efficiency of toluene oxidation was also measured. As shown in Figure 3b and Table 1, the mineralization temperatures of these catalysts for toluene are higher than that of conversion, where the variation trend of CO2 yield was similar to that of conversion. Besides, α-MO/LMO sample still maintains the best behavior in CO2 generation capacity.

2.3. Textual Structure and Surface Area

The N2 adsorption-desorption isotherms of LMO, α-MO/LMO, β-MO/LMO, δ-MO/LMO and γ-MO/LMO are shown in Figure 4a, and the specific surface areas are listed in Table 2. As displayed in Figure 4a, the isotherms of LMO, α-MO/LMO, β-MO/LMO, δ-MO/LMO and γ-MO/LMO presented a typical IUPAC type II pattern with non-uniform slit pores (H3 type) of the hysteresis loop, signifying the existence of macroporous and mesoporous structures [11,18,29]. In addition, it can also be observed from the data listed in Table 2 that the specific surface areas of LMO, α-MO/LMO, β-MO/LMO, δ-MO/LMO, and γ-MO/LMO were 14.0, 42.3, 40.0, 39.6, and 35.6 m2·g−1, respectively. Compared with pure perovskite, the specific surface areas of various phase types of MnO2/LaMnO3 were all increased. The high specific surface area is beneficial to promote the oxidation of toluene molecules on the surface of the catalyst [29,30]. Namely, the α-MO/LMO catalyst with the highest specific surface area may result in better activity [31].

2.4. Mn Oxidized State and Oxygen Species

The Mn-oxidized state and oxygen species on the surface of LMO, α-MO/LMO, β-MO/LMO, δ-MO/LMO, and γ-MO/LMO were obtained using XPS. The Mn 2p3/2, O 1s XPS spectra of all catalysts are shown in Figure 5a,b and the quantitative analysis is presented in Table 2. As displayed in Figure 5a, the components divided from the Mn 2p3/2 XPS spectra at 643.58, 641.78, and 640.46 eV corresponded to Mn4+, Mn3+, and Mn2+, respectively [11,32,33,34]. Compared with that of pure LMO perovskite, the Mn 2p binding energy of MO/LMO is relatively reduced, suggesting that the density of the extranuclear electron cloud of the Mn element is increased. In fact, the asymmetric charge density distribution of Mn showed a significant increase in the hybridization of Mn 3d and O 2p orbitals and shortening of Mn-O bond length [21], which may be helpful to activate the lattice oxygen and improve the reducibility of the catalysts [12]. The increase in the molar ratio of Mn4+/Mn3+ in the order of LMO (0.38) < γ-MO/LMO (0.40) < δ-MO/LMO (0.42) < β-MO/LMO (0.46) < α-MO/LMO (0.52), which is consistent with the order of their activities. The increase in the proportion of high-valent manganese implies the lattice oxygen on the surface is more easily activated. The activated lattice oxygen can migrate to the oxygen vacancies and then transform into overflowed adsorbed oxygen species easily to oxidize toluene [21]. Consequently, the α-MO/LMO with the high molar ratio of Mn4+/Mn3+ has a higher toluene oxidation potential.
Furthermore, the oxygen species on the surface of all catalysts were determined by XPS. As shown in Figure 5b, the peaks of O 1s XPS spectra located at 529.1, 530.9, and 532.4 eV are attributed to the lattice oxygen (Olatt), surface adsorbed oxygen (Oads) and surface carbonate (CO32−) or hydroxide (OH), respectively [18,23,28]. In contrast to that of Mn 2p, the O 1s binding energy of MO/LMO is relatively elevated, explaining that the density of the extranuclear electron cloud of O element cut down, facilitating the activation of lattice oxygen. As listed in Table 2, the order of the Olatt/Oads molar ratio enhanced in the arrangement of LMO (1.21) < γ-MO/LMO (1.89) < δ-MO/LMO (2.12) < β-MO/LMO (2.39) < α-MO/LMO (2.48), which is consistent with the increasing sequences of Mn4+/Mn3+ ratio and toluene conversion. The results testified that the composite of MnO2 and LaMnO3 promoted the enhancement of the activated lattice oxygen, and the surface of α-MO/LMO possessed the most Olatt, which induced the best catalytic oxidation performance of toluene.

2.5. Reducibility

The reducibility of LMO, α-MO/LMO, β-MO/LMO, δ-MO/LMO, and γ-MO/LMO was evaluated by H2-TPR. As shown in Figure 4b, the peaks for LMO at 378 °C are attributed to the restituting of Mn4+, and the peak at 512 °C was assigned to the single-electron reduction of Mn3+ located in a coordination-unsaturated microenvironment, and the peak at 864 °C is classified to the retrieval of the left Mn3+ [18,24,26]. For α-MO/LMO, the peaks at 266 and 288 °C should correspond to the reduction of surface Mn4+ and deep Mn4+; the peaks at 353 and 512 °C belong to the reduction of surface Mn3+ and deep Mn3+; the peaks at 810 °C is pertain to the lowering of the left Mn3+ in LMO. Analogously, the peaks of β-MO/LMO at 279 and 329 °C belong to the surface Mn4+ and deep Mn4+, 416 and 510 °C corresponding to the reduction of surface Mn3+ and deep Mn3+, 808 °C corresponding to the left Mn3+ in LMO; the peaks of δ-MO/LMO at 272 °C corresponding to the reduction of Mn4+, 349 and 522 °C corresponding to the reduction of surface Mn3+ and deep Mn3+, 828 °C corresponding to the left Mn3+ in LMO; the peaks of γ-MO/LMO at 280 and 320 °C belong to the surface Mn4+ and deep Mn4+, 408 and 525 °C corresponding to the reduction of surface Mn3+ and deep Mn3+, 836 °C corresponding to the left Mn3+ in LMO. Obviously, α-MO/LMO has better reducibility at low temperatures. The H2 consumption of each catalyst was calculated, and it can be confirmed from Table 2 that the order of hydrogen consumption between 50 and 450 °C is as follows: LMO (2.03 mmol·g−1) < γ-MO/LMO (2.44 mmol·g−1) < δ-MO/LMO (3.03 mmol·g−1) < β-MO/LMO (3.51 mmol·g−1) < α-MO/LMO (4.09 mmol·g−1). Namely, the α-MO/LMO hold the high H2-uptake, and the more H2-uptake represents a higher valence of Mn [22,26], which indirectly shows that α-MO/LMO has more activated lattice oxygen on the surface, leading to the better toluene oxidation. Combined with the results of H2-TPR, XPS, BET and their activities, α-MO/LMO has a higher proportion of lattice oxygen, better low-temperature reducibility and larger surface area in comparison with other catalysts, which leads to the best catalytic activity.

2.6. Possible Degradation Mechanism

The rare content of exhaust byproducts was identified by the thermal desorption/gas chromatograph mass spectrometer instrument. As shown in Figure 6, trace byproducts from the T50 condition of α-MO/LMO oxidizing toluene illustrated some identifiable organic byproducts, such as benzene, benzoic acid, benzaldehyde, 1,3-diethenyl-benzene, acetic acid, benzyl alcohol, phenol, 4-phenyl-3-buten-2-one, octanoic acid, cyclohexasiloxane, tridecane, nonanoic acid, n-decanoic acid, benzophenone, and benzyl benzoate. In fact, toluene oxidation involves the surface active oxygen species excited by the heating [5,31,35]. Furthermore, the intermediates for the catalytic oxidation of toluene over α-MO/LMO were obtained and displayed in Figure 7. A series of structural changes in the surface of α-MO/LMO occurred. With the temperature change, the peak at 3070 cm−1 belongs to the C-H of vinyl stretching vibration [36]. The peaks at 1596, 1533 and 1419 cm−1 pertained to the C-C and C=O stretching vibration and carbonate species [37,38]. Moreover, the vibration modes at 1288 cm−1 and 1038 cm−1 belong to vs(COO) stretching vibration [39]. The results showed that toluene was oxidized to some carbonyl-containing species. Even though a higher degradation efficiency was found for α-MO/LMO catalysts and only trace organic byproducts were detectable in the outgas, some of the products would cause more toxicity to human health or the environment [40], even in a very small quantity. The comprehensive assessment of these trace byproducts would be our future research direction to help the choice of new-dedicated catalysts for the VOC oxidation technique.
Based on the characterization of GC-MS and in-situ DRIFTS, the toluene oxidation over α-MO/LMO should adhere to the Mars-van Krevelen mechanism [41,42]. The active oxygen species on the surface of α-MO/LMO include adsorbed oxygen (Oads) and lattice oxygen (Olatt). At a temperature, the activated Oads and Olatt can be transformed into each other. Meantime, the consumed active oxygen can be repaired by the interaction of O2 molecules in the atmosphere with oxygen vacancies. On account of a relatively low temperature, the number of excited reactive oxygen species on the surface of α-MO/LMO was limited, resulting in insufficient toluene oxidation, which can bring about a series of organic byproducts [18,21,43]. The toluene was decomposed and may undergo a series of consequent reactions [18,24]: toluene → benzyl alcohol → benzaldehyde → benzoic acid → benzene → acetic acid → CO2. Wherefore, it is necessary for toluene oxidation to further heighten the reaction temperature to manufacture more active oxygen species.

3. Experimental Section

3.1. Synthesis of Catalysts

LaMnO3 perovskite was first prepared via direct calcination without adding water as solvent of metal salts and citric acid, which was similar to the preparation we previously reported [44]. α-MnO2/LaMnO3, β-MnO2/LaMnO3, δ-MnO2/LaMnO3, γ-MnO2/LaMnO3 (recorded as α-MO/LMO, β-MO/LMO, δ-MO/LMO, γ-MO/LMO) were synthesized by a simple one-step method, with the same molar ratio of MnO2 to LaMnO3 is 3:7. The typical synthesis procedure of α-MO/LMO was as follows: 0.35 g of KMnO4 and 0.15 g of MnSO4·H2O was dissolved in 60 mL deionized water. After forming an emulsion-like, 2 g of LaMnO3 powders were added to the above-mixed liquid and stirred for 30 min, and then transferred to a 100 mL PTFE kettle for hydrothermal reaction at 160 °C for 12 h. Ultimately, the acquired precipitates were calcined at 400 °C for 4 h.
For β-MO/LMO, 0.528 g of MnSO4·H2O and 0.712 g of (NH4)2S2O8 were added to 60 mL of deionized water. Other steps are similar to that of α-MO/LMO. For δ-MO/LMO, the addition amounts of KMnO4 and MnSO4·H2O were 0.42 g and 0.78 g, respectively, and other steps were similar to that of α-MO/LMO. For γ-MO/LMO, 2 g of LaMnO3, 0.3169 g of MnSO4·H2O, and 1 mL of concentrated H2SO4 were added to 60 mL deionized water and agitated at 80 °C for 30 min, then added 0.948 g of KMnO4 to keep stirring for 24 h. The acquired precipitate was calcined at 400 °C for 4 h.

3.2. Characterization

The prepared samples were characterized by various measurement techniques, X-ray diffraction (XRD, Shimadzu, Tokyo, Japan), Scanning electron microscope (SEM, Bruker, Berlin, Germany), X-ray photoelectron spectra (XPS, Perkin-Elmer, Waltham, MA, USA), and H2 temperature-programmed reduction (H2-TPR, Micromeritics, Norcross, GA, USA). All the character processes and instrument parameters are detailed in the Supplementary Material document.

3.3. Catalytic Evaluation

The catalytic oxidation of toluene over all samples was estimated in a successive flow fixed bed quartz micro-reactor (I.D. = 0.006 m). 0.20 g of samples (80–120 mesh) were applied to investigate their properties. The air stream with approximately 1000 ppm of toluene crossed over the bed with a total flow rate of 40 mL/min, with a weight hourly space velocity (WHSV) of 12,000 mL/(g h−1). The inset and outset concentration of toluene was detected by the GC with a flame ionization detector (FID).
The conversion of toluene is calculated by the following equations:
X = C in C out C in × 100 %
where X is the conversion efficiency of toluene, and Cin and Cout are the toluene concentrations in the inlet and outlet gas streams, respectively.
Meantime, the yield of CO2, is resulted from Equation (2):
Y = C C O 2 6 C in × 100 %
where Y is the yield of CO2, CCO2 is the outlet CO2 concentration.
For reaction products, the concentration of CO2 from the outlet stream was measured on the GC-FID equipped with a nickel converting unit. Trace organic byproducts were trapped in a Carbopack B tube (Camsco, Huston, TX, USA) and then detected on a Gas Chromatograph Mass Spectrometer (Agilent 7890B-7000C, St. Clara, CA, USA) coupled with the Thermal Desorption (Markes, Unity-xr, London, UK) instrument. The reaction intermediate was monitored by Fourier transform spectroscopy (DRIFTS, Thermo Nicolet 6700, Waltham, MA, USA), and the experiment and instrument information are shown in Supplementary Material.

4. Conclusions

Manganese dioxides with different phases types (α, β, δ, and γ) were successfully composited with LaMnO3 perovskite by a simple one-pot synthesis route. Combined with the characterization results of SEM, N2 physisorption, XPS and H2-TPR, the upgraded catalytic activity of the most active α-MnO2/LaMnO3 should be attributed to its high content of lattice oxygen, better low-temperature reducibility and larger surface area. The activation degree of lattice oxygen on the surface of MnO2/LaMnO3 can be controlled by adjusting the phase-type of MnO2. In addition, the byproducts from the low-temperature oxidation of toluene over α-MnO2/LaMnO3 mainly include benzyl alcohol, benzaldehyde, benzoic acid, benzene, acetic acid, etc. This study also provides a method to control the crystal phase of manganese dioxide on the surface of perovskite and a demonstration to reveal the structure-activity relationship.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal12121666/s1, Figure S1: XRD patterns of LMO; Figure S2: Toluene oxidation over of (a) α-MO and α-MO/LMO, (b) β-MO and β-MO/LMO, (c) δ-MO and δ-MO/LMO, (d) γ-MO and γ-MO/LMO; Text S1: Chemicals and materials; Text S2: Characterization; Text S3: Preparation of LaMnO3; Text S4: Preparation of MnO2 with different crystal phases; Text S5: Normalized reaction rate equation [45,46,47].

Author Contributions

L.L. (Lu Li): Investigation, Data curation, Visualization, Writing—original draft, Funding acquisition; Y.L. and J.L.: Software, Supervision, writing—review & editing; B.Z.: Investigation, Validation, Methodology; M.G. and L.L. (Lizhong Liu): Resources, Funding acquisition, Project administration, Supervision, writing—review & editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science Fundamental Research Project of Jiangsu Colleges and Universities of China (Grant No. 22KJB610022) and Jiangsu Provincial Key Research and Development Program (Grant No. BE2022767).

Data Availability Statement

Not applicable.

Acknowledgments

The authors of this work appreciate the technical support from the Instrumental Analysis Center of Xi’an Jiaotong University and Nantong University Analysis & Testing center.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, X.; Sun, Y.; Li, M.; Zhang, W.; Zhu, Y. Excellent catalytic oxidation performance on toluene and benzene over OMS-2 with a hierarchical porous structure synthesized by a one-pot facile method: Modifying surface properties by introducing different amounts of K. Catal. Sci. Technol. 2022, 12, 2872–2886. [Google Scholar] [CrossRef]
  2. Liu, B.; Ji, J.; Zhang, B.; Huang, W.; Gan, Y.; Leung, D.Y.C.; Huang, H. Catalytic ozonation of VOCs at low temperature: A comprehensive review. J. Hazard. Mater. 2022, 422, 126847. [Google Scholar] [CrossRef] [PubMed]
  3. Li, S.; Wang, D.; Wu, X.; Chen, Y. Recent advance on VOCs oxidation over layered double hydroxides derived mixed metal oxides. Chin. J. Catal. 2020, 41, 550–560. [Google Scholar] [CrossRef]
  4. Gao, Z.; Wang, J.; Muhammad, Y.; Hu, P.; Hu, Y.; Chu, Z.; Zhao, Z.; Zhao, Z. Hydrophobic shell structured NH2-MILl(Ti)-125@mesoporous carbon composite via confined growth strategy for ultra-high selective adsorption of toluene under highly humid environment. Chem. Eng. J. 2022, 432, 134340. [Google Scholar] [CrossRef]
  5. Guo, Y.; Wen, M.; Li, G.; An, T. Recent advances in VOC elimination by catalytic oxidation technology onto various nanoparticles catalysts: A critical review. Appl. Catal. B Environ. 2021, 281, 119447. [Google Scholar] [CrossRef]
  6. Wu, P.; Jin, X.; Qiu, Y.; Ye, D. Recent progress of thermocatalytic and photo/thermocatalytic oxidation for VOCs purification over manganese-based oxide catalysts. Environ. Sci. Technol. 2021, 55, 4268–4286. [Google Scholar] [CrossRef]
  7. Muir, B.; Sobczyk, M.; Bajda, T. Fundamental features of mesoporous functional materials influencing the efficiency of removal of VOCs from aqueous systems: A review. Sci. Total Environ. 2021, 784, 147121. [Google Scholar] [CrossRef]
  8. Topuz, F.; Abdulhamid, M.A.; Hardian, R.; Holtzl, T.; Szekely, G. Nanofibrous membranes comprising intrinsically microporous polyimides with embedded metal–organic frameworks for capturing volatile organic compounds. J. Hazard. Mater. 2022, 424, 127347. [Google Scholar] [CrossRef]
  9. Feng, Y.; Wang, C.; Wang, C.; Huang, H.; Hsi, H.-C.; Duan, E.; Liu, Y.; Guo, G.; Dai, H.; Deng, J. Catalytic stability enhancement for pollutant removal via balancing lattice oxygen mobility and VOCs adsorption. J. Hazard. Mater. 2022, 424, 127337. [Google Scholar] [CrossRef]
  10. Li, J.; Zhang, M.; Elimian, E.A.; Lv, X.; Chen, J.; Jia, H. Convergent ambient sunlight-powered multifunctional catalysis for toluene abatement over in situ exsolution of Mn3O4 on perovskite parent. Chem. Eng. J. 2021, 412, 128560. [Google Scholar] [CrossRef]
  11. Guo, M.; Liu, L.; Gu, J.; Zhang, H.; Min, X.; Liang, J.; Jia, J.; Li, K.; Sun, T. Catalytic performance improvement of volatile organic compounds oxidation over MnOx and GdMnO3 composite oxides from spent lithium-ion batteries: Effect of acid treatment. Chin. J. Chem. Eng. 2021, 34, 278–288. [Google Scholar] [CrossRef]
  12. Liu, L.; Liu, R.; Xu, T.; Zhang, Q.; Tan, Y.; Zhang, Q.; Ding, J.; Tang, Y. Enhanced catalytic oxidation of chlorobenzene over MnO2 grafted in situ by rare earth oxide: Surface doping induces lattice oxygen activation. Inorg. Chem. 2020, 59, 14407–14414. [Google Scholar] [CrossRef] [PubMed]
  13. He, C.; Cheng, J.; Zhang, X.; Douthwaite, M.; Pattisson, S.; Hao, Z. Recent advances in the catalytic oxidation of volatile organic compounds: A review based on pollutant sorts and sources. Chem. Rev. 2019, 119, 4471–4568. [Google Scholar] [CrossRef] [PubMed]
  14. Royer, S.; Duprez, D.; Can, F.; Courtois, X.; Batiot-Dupeyrat, C.; Laassiri, S.; Alamdari, H. Perovskites as substitutes of noble metals for heterogeneous catalysis: Dream or reality. Chem. Rev. 2014, 114, 10292–10368. [Google Scholar] [CrossRef] [PubMed]
  15. Zhu, J.; Li, H.; Zhong, L.; Xiao, P.; Xu, X.; Yang, X.; Zhao, Z.; Li, J. Perovskite oxides: Preparation, characterizations, and applications in heterogeneous catalysis. ACS Catal. 2014, 4, 2917–2940. [Google Scholar] [CrossRef]
  16. Si, W.; Wang, Y.; Zhao, S.; Hu, F.; Li, J. A facile method for in situ preparation of the MnO2/LaMnO3 catalyst for the removal of toluene. Environ. Sci. Technol. 2016, 50, 4572–4578. [Google Scholar] [CrossRef]
  17. Yang, J.; Li, L.; Yang, X.; Song, S.; Li, J.; Jing, F.; Chu, W. Enhanced catalytic performances of in situ-assembled LaMnO3/δ-MnO2 hetero-structures for toluene combustion. Catal. Today 2019, 327, 19–27. [Google Scholar] [CrossRef]
  18. Liu, L.; Li, J.; Zhang, H.; Li, L.; Zhou, P.; Meng, X.; Guo, M.; Jia, J.; Sun, T. In situ fabrication of highly active γ-MnO2/SmMnO3 catalyst for deep catalytic oxidation of gaseous benzene, ethylbenzene, toluene, and o-xylene. J. Hazard. Mater. 2019, 362, 178–186. [Google Scholar] [CrossRef]
  19. Figueredo, M.J.M.; Cocuzza, C.; Bensaid, S.; Fino, D.; Piumetti, M.; Russo, N. Catalytic abatement of volatile organic compounds and soot over manganese oxide catalysts. Materials 2021, 14, 4534. [Google Scholar] [CrossRef]
  20. Kim, C.S.; Shim, G.W. Catalytic combustion of VOCs over a series of manganese oxide catalysts. Appl. Catal. B Envion. 2010, 98, 180–185. [Google Scholar] [CrossRef]
  21. Liu, X.; Mi, J.; Shi, L.; Liu, H.; Liu, J.; Ding, Y.; Shi, J.; He, M.; Wang, Z.; Xiong, S.; et al. In situ modulation of a-site vacancies in LaMnO3.15 perovskite for surface lattice oxygen activation and boosted redox reactions. Angew. Chem. Int. Edit. 2021, 60, 26747–26754. [Google Scholar] [CrossRef] [PubMed]
  22. Min, X.; Guo, M.; Li, K.; Gu, J.; Guo, X.; Xue, Y.; Liang, J.; Hu, S.; Jia, J.; Sun, T. Enhancement of toluene removal over α@δ-MnO2 composites prepared via one-pot by modifying the molar ratio of KMnO4 to MnSO4·H2O. Appl. Surf. Sci. 2021, 568, 150972. [Google Scholar] [CrossRef]
  23. Yang, W.; Su, Z.; Xu, Z.; Yang, W.; Peng, Y.; Li, J. Comparative study of α-, β-, γ- and δ-MnO2 on toluene oxidation: Oxygen vacancies and reaction intermediates. Appl. Catal. B Environ. 2020, 260, 118150. [Google Scholar] [CrossRef]
  24. Liu, L.; Liu, J.; Guo, M. One-pot synthesis of dual-phase manganese dioxide for toluene removal: Effect of crystal phase blending level on oxygen species and activity. J. Environ. Chem. Eng. 2022, 10, 107448. [Google Scholar] [CrossRef]
  25. Li, K.; Chen, C.; Zhang, H.; Hu, X.; Sun, T.; Jia, J. Effects of phase structure of MnO2 and morphology of δ-MnO2 on toluene catalytic oxidation. Appl. Surf. Sci. 2019, 496, 143662. [Google Scholar] [CrossRef]
  26. Huang, N.; Qu, Z.; Dong, C.; Qin, Y.; Duan, X. Superior performance of α@β-MnO2 for the toluene oxidation: Active interface and oxygen vacancy. Appl. Catal. A-Gen. 2018, 560, 195–205. [Google Scholar] [CrossRef]
  27. Wang, D.; Wang, L.; Liang, G.; Li, H.; Liu, Z.; Tang, Z.; Liang, J.; Zhi, C. A superior δ-MnO2 cathode and a self-healing Zn-δ-MnO2 battery. ACS Nano 2019, 13, 10643–10652. [Google Scholar] [CrossRef]
  28. Liu, L.; Zhang, H.; Jia, J.; Sun, T.; Sun, M. Direct molten polymerization synthesis of highly active samarium manganese perovskites with different morphologies for VOC removal. Inorg. Chem. 2018, 57, 8451–8457. [Google Scholar] [CrossRef]
  29. Liu, L.; Zhou, B.; Liu, Y.; Liu, J.; Hu, L.; Tang, Y.; Wang, M. In-situ regulation of acid sites on Mn-based perovskite@mullite composite for promoting catalytic oxidation of chlorobenzene. J. Colloid. Interf. Sci. 2021, 606, 1866–1873. [Google Scholar] [CrossRef]
  30. Ding, J.; Liu, L.; Xue, J.; Zhou, Z.; He, G.; Chen, H. Low-temperature preparation of magnetically separable Fe3O4@CuO-RGO core-shell heterojunctions for high-performance removal of organic dye under visible light. J. Alloys Compd. 2016, 688, 649–656. [Google Scholar] [CrossRef]
  31. Liu, R.; Zhou, B.; Liu, L.; Zhang, Y.; Chen, Y.; Zhang, Q.; Yang, M.; Hu, L.; Wang, M.; Tang, Y. Enhanced catalytic oxidation of VOCs over porous Mn-based mullite synthesized by in-situ dismutation. J. Colloid. Interf. Sci. 2021, 585, 302–311. [Google Scholar] [CrossRef] [PubMed]
  32. Si, W.; Wang, Y.; Peng, Y.; Li, J. Selective dissolution of A-site cations in ABO3 perovskites: A new path to high-performance catalysts. Angew. Chem. Int. Edit. 2015, 54, 7954. [Google Scholar] [CrossRef] [PubMed]
  33. Guo, M.; Wang, X.; Liu, L.; Min, X.; Hu, X.; Guo, W.; Zhu, N.; Jia, J.; Sun, T.; Li, K. Recovery of cathode materials from spent lithium-ion batteries and their application in preparing multi-metal oxides for the removal of oxygenated VOCs: Effect of synthetic methods. Env. Res. 2021, 193, 110563. [Google Scholar] [CrossRef] [PubMed]
  34. Guo, M.; Li, K.; Zhang, H.; Min, X.; Liang, J.; Hu, X.; Guo, W.; Jia, J.; Sun, T. Promotional removal of oxygenated VOC over manganese-based multi oxides from spent lithium-ions manganate batteries: Modification with Fe, Bi and Ce dopants. Sci. Total Environ. 2020, 740, 139951. [Google Scholar] [CrossRef]
  35. Chen, J.; Chen, X.; Yan, D.; Jiang, M.; Xu, W.; Yu, H.; Jia, H. A facile strategy of enhancing interaction between cerium and manganese oxides for catalytic removal of gaseous organic contaminants. Appl. Catal. B Environ. 2019, 250, 396–407. [Google Scholar] [CrossRef]
  36. Li, X.; Zhu, Z.; Zhao, Q.; Wang, L. Photocatalytic degradation of gaseous toluene over ZnAl2O4 prepared by different methods: A comparative study. J. Hazard. Mater. 2011, 186, 2089–2096. [Google Scholar] [CrossRef]
  37. Chen, J.; Chen, X.; Xu, W.; Xu, Z.; Chen, J.; Jia, H.; Chen, J. Hydrolysis driving redox reaction to synthesize Mn-Fe binary oxides as highly active catalysts for the removal of toluene. Chem. Eng. J. 2017, 330, 281–293. [Google Scholar] [CrossRef]
  38. Gao, J.; Tong, X.; Li, X.; Miao, H.; Xu, J. The efficient liquid-phase oxidation of aromatic hydrocarbons by molecular oxygen in the presence of MnCO3. J. Chem. Technol. Biot. 2007, 82, 620–625. [Google Scholar] [CrossRef]
  39. Pan, H.; Jian, Y.; Chen, C.; He, C.; Hao, Z.; Shen, Z.; Liu, H. Sphere-shaped Mn3O4 catalyst with remarkable low-temperature activity for methyl–ethyl–ketone combustion. Environ. Sci. Technol. 2017, 51, 6288–6297. [Google Scholar] [CrossRef] [Green Version]
  40. Gennequin, C.; Kouassi, S.; Tidahy, L.; Cousin, R.; Lamonier, J.F.; Garcon, G.; Shirali, P.; Cazier, F.; Aboukais, A.; Siffert, S. Co-Mg-Al oxides issued of hydrotalcite precursors for total oxidation of volatile organic compounds. Identification and toxicological impact of the by-products. Comptes Rendus Chim. 2010, 13, 494–501. [Google Scholar] [CrossRef]
  41. Deng, J.; Zhang, L.; Dai, H.; He, H.; Au, C.T. Strontium-doped lanthanum cobaltite and manganite: Highly active catalysts for toluene complete oxidation. Ind. Eng. Chem. Res. 2008, 47, 8175–8183. [Google Scholar] [CrossRef]
  42. Liu, Y.; Dai, H.; Deng, J.; Du, Y.; Li, X.; Zhao, Z.; Wang, Y.; Gao, B.; Yang, H.; Guo, G. In situ poly(methyl methacrylate)-templating generation and excellent catalytic performance of MnOx/3DOMLaMnO3 for the combustion of toluene and methanol. Appl. Catal. B Environ. 2013, 140, 493–505. [Google Scholar] [CrossRef]
  43. Liu, L.; Sun, J.; Ding, J.; Zhang, Y.; Sun, T.; Jia, J. Highly active Mn3–xFexO4 spinel with defects for toluene mineralization: Insights into regulation of the oxygen vacancy and active metals. Inorg. Chem. 2019, 58, 13241–13249. [Google Scholar] [CrossRef] [PubMed]
  44. Liu, L.; Sun, J.; Ding, J.; Zhang, Y.; Jia, J.; Sun, T. Catalytic oxidation of VOCs over SmMnO3 perovskites: Catalyst synthesis, change mechanism of active species, and degradation path of toluene. Inorg. Chem. 2019, 58, 14275–14283. [Google Scholar] [CrossRef] [PubMed]
  45. Zhao, J.; Xi, W.; Tu, C.; Dai, Q.; Wang, X. Catalytic oxidation of chlorinated VOCs over Ru/TixSn1-x catalysts. Appl. Catal. B 2020, 263, 118237. [Google Scholar] [CrossRef]
  46. Schmid, D.; Micic, V.; Laumann, S.; Hofmann, T. Measuring the reactivity of commercially available zero-valent iron nanoparticles used for environmental remediation with iopromide. J. Contam. Hydrol. 2015, 181, 36–45. [Google Scholar] [CrossRef]
  47. Yang, P.; Fan, S.; Chen, Z.; Bao, G.; Zuo, S.; Qi, C. Synthesis of Nb2O5 based solid superacid materials for catalytic combustion of chlorinated VOCs. Appl. Catal. B 2018, 239, 114–124. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of (a) α-MO/LMO, α-MnO2 and corresponding standard PDF card diffraction profile; (b) β-MO/LMO, β-MnO2 and corresponding standard PDF card diffraction profile; (c) γ-MO/LMO γ-MnO2 and corresponding standard PDF card diffraction profile; and (d) δ-MO/LMO, δ-MnO2 and corresponding standard PDF card diffraction profile.
Figure 1. XRD patterns of (a) α-MO/LMO, α-MnO2 and corresponding standard PDF card diffraction profile; (b) β-MO/LMO, β-MnO2 and corresponding standard PDF card diffraction profile; (c) γ-MO/LMO γ-MnO2 and corresponding standard PDF card diffraction profile; and (d) δ-MO/LMO, δ-MnO2 and corresponding standard PDF card diffraction profile.
Catalysts 12 01666 g001
Figure 2. SEM images of (a,b) LMO, (c,d) α-MO/LMO, (e,f) β-MO/LMO, (g,h) δ-MO/LMO, and (i,j) γ-MO/LMO.
Figure 2. SEM images of (a,b) LMO, (c,d) α-MO/LMO, (e,f) β-MO/LMO, (g,h) δ-MO/LMO, and (i,j) γ-MO/LMO.
Catalysts 12 01666 g002
Figure 3. (a) Toluene conversion vs. temperature and (b) CO2 yield over LMO, α-MO/LMO, β-MO/LMO, δ-MO/LMO, and γ-MO/LMO.
Figure 3. (a) Toluene conversion vs. temperature and (b) CO2 yield over LMO, α-MO/LMO, β-MO/LMO, δ-MO/LMO, and γ-MO/LMO.
Catalysts 12 01666 g003
Figure 4. (a) N2 adsorption-desorption isotherms and (b) H2-TPR profiles of LMO, α-MO/LMO, β-MO/LMO, δ-MO/LMO and γ-MO/LMO.
Figure 4. (a) N2 adsorption-desorption isotherms and (b) H2-TPR profiles of LMO, α-MO/LMO, β-MO/LMO, δ-MO/LMO and γ-MO/LMO.
Catalysts 12 01666 g004
Figure 5. (a) Mn 2p3/2, and (b) O 1s XPS spectra of LMO, α-MO/LMO, β-MO/LMO, δ-MO/LMO and γ-MO/LMO.
Figure 5. (a) Mn 2p3/2, and (b) O 1s XPS spectra of LMO, α-MO/LMO, β-MO/LMO, δ-MO/LMO and γ-MO/LMO.
Catalysts 12 01666 g005
Figure 6. GC-MS spectrum of toluene oxidation over α-MO/LMO.
Figure 6. GC-MS spectrum of toluene oxidation over α-MO/LMO.
Catalysts 12 01666 g006
Figure 7. In-situ DRIFTS result of toluene oxidation over α-MO/LMO.
Figure 7. In-situ DRIFTS result of toluene oxidation over α-MO/LMO.
Catalysts 12 01666 g007
Table 1. Catalytic performance for toluene oxidation over various catalysts.
Table 1. Catalytic performance for toluene oxidation over various catalysts.
SamplesConversionCO2 Yieldr
T90 (°C)T50 (°C)T90 (°C)T50 (°C)(10−10 mol m−2 s−1)
α-MO/LMO2602372692413.13
β-MO/LMO2892462932502.70
δ-MO/LMO2942552992591.85
γ-MO/LMO3162613202641.73
LMO>320300>3203092.62
α-MO295252///
β-MO>320302///
δ-MO>320289///
γ-MO>320286///
Table 2. Surface element compositions, surface area and reducibility of LMO, α-MO/LMO, β-MO/LMO, δ-MO/LMO and γ-MO/LMO.
Table 2. Surface element compositions, surface area and reducibility of LMO, α-MO/LMO, β-MO/LMO, δ-MO/LMO and γ-MO/LMO.
CatalystsSBET/
m2·g−1
Molar Ratios of Surface Elements by XPSH2 uptake/mmol·g−1
(50–450 °C)
Mn4+/Mn3+Olatt/Oads
LMO14.00.381.212.03
α-MO/LMO42.30.522.484.09
β-MO/LMO40.00.462.393.51
δ-MO/LMO39.60.422.123.03
γ-MO/LMO35.60.401.892.44
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, L.; Liu, Y.; Liu, J.; Zhou, B.; Guo, M.; Liu, L. Catalytic Degradation of Toluene over MnO2/LaMnO3: Effect of Phase Type of MnO2 on Activity. Catalysts 2022, 12, 1666. https://doi.org/10.3390/catal12121666

AMA Style

Li L, Liu Y, Liu J, Zhou B, Guo M, Liu L. Catalytic Degradation of Toluene over MnO2/LaMnO3: Effect of Phase Type of MnO2 on Activity. Catalysts. 2022; 12(12):1666. https://doi.org/10.3390/catal12121666

Chicago/Turabian Style

Li, Lu, Yuwei Liu, Jingyin Liu, Bing Zhou, Mingming Guo, and Lizhong Liu. 2022. "Catalytic Degradation of Toluene over MnO2/LaMnO3: Effect of Phase Type of MnO2 on Activity" Catalysts 12, no. 12: 1666. https://doi.org/10.3390/catal12121666

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