Next Article in Journal
Structural Construction of WO3 Nanorods as Anode Materials for Lithium-Ion Batteries to Improve Their Electrochemical Performance
Next Article in Special Issue
Oxidation of Methanol and Dichloromethane on TiO2-CeO2-CuO, TiO2-CeO2 and TiO2-CuO@VUKOPOR®A Ceramic Foams
Previous Article in Journal
Metallocene-Filled Single-Walled Carbon Nanotube Hybrids
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Comparative Study of α- and β-MnO2 on Methyl Mercaptan Decomposition: The Role of Oxygen Vacancies

1
Faculty of Environmental Science and Engineering, Kunming University of Science and Technology, Kunming 650500, China
2
The Innovation Team for Volatile Organic Compounds Pollutants Control and Resource Utilization of Yunnan Province, The Higher Educational Key Laboratory for Odorous Volatile Organic Compounds Pollutants Control of Yunnan Province, Kunming 650500, China
3
Faculty of Chemical Engineering, Kunming University of Science and Technology, Kunming 650500, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(4), 775; https://doi.org/10.3390/nano13040775
Submission received: 24 January 2023 / Revised: 5 February 2023 / Accepted: 15 February 2023 / Published: 19 February 2023
(This article belongs to the Special Issue Nanocatalysts for Air Purification)

Abstract

:
As a representative sulfur-containing volatile organic compounds (S-VOCs), CH3SH has attracted widespread attention due to its adverse environmental and health risks. The performance of Mn-based catalysts and the effect of their crystal structure on the CH3SH catalytic reaction have yet to be systematically investigated. In this paper, two different crystalline phases of tunneled MnO2 (α-MnO2 and β-MnO2) with the similar nanorod morphology were used to remove CH3SH, and their physicochemical properties were comprehensively studied using high-resolution transmission electron microscope (HRTEM) and electron paramagnetic resonance (EPR), H2-TPR, O2-TPD, Raman, and X-ray photoelectron spectroscopy (XPS) analysis. For the first time, we report that the specific reaction rate for α-MnO2 (0.029 mol g−1 h−1) was approximately 4.1 times higher than that of β-MnO2 (0.007 mol g−1 h−1). The as-synthesized α-MnO2 exhibited higher CH3SH catalytic activity towards CH3SH than that of β-MnO2, which can be ascribed to the additional oxygen vacancies, stronger surface oxygen migration ability, and better redox properties from α-MnO2. The oxygen vacancies on the catalyst surface provided the main active sites for the chemisorption of CH3SH, and the subsequent electron transfer led to the decomposition of CH3SH. The lattice oxygen on catalysts could be released during the reaction and thus participated in the further oxidation of sulfur-containing species. CH3SSCH3, S0, SO32−, and SO42− were identified as the main products of CH3SH conversion. This work offers a new understanding of the interface interaction mechanism between Mn-based catalysts and S-VOCs.

Graphical Abstract

1. Introduction

As a particular class of volatile organic compounds (VOCs), sulfur-containing volatile organic compounds (S-VOCs) can be converted into sulfate aerosols in the atmosphere through complex physicochemical reactions [1]. They can also react indirectly with NOx through photochemistry reactions, which are the crucial precursors for forming PM 2.5 and O3. Methyl mercaptan (CH3SH), a representative S-VOC, is considered as an important air odor pollutant, which is harmful to the ecosystem and human health owing to its severe toxicity and low olfactory threshold [2,3,4]. In previous studies, various methods have been employed to eliminate CH3SH, such as adsorption [5,6], biodegradation [7], photocatalytic oxidation [8], and catalytic oxidation [9,10]. However, these remediation technologies suffer from secondary pollution because of incomplete removal and high cost. Until now, catalytic decomposition has been regarded as the most promising strategy for removing S-VOCs due to its high purification efficiency, energy-saving nature, lack of additional additives (O2, H2, O3, etc.), and less secondary pollution [11,12].
Metal-based catalysts are widely used for the removal of VOCs on account of their superior catalytic performance. Among the various transition metal oxides, MnO2 is of great interest because of its low cost, low toxicity, environmental friendliness, and wide natural distribution [13,14]. Meanwhile, MnO2 has been extensively studied in heterogeneous catalysis due to its unique physicochemical properties (i.e., multivalent, reactive oxygen species, and polycrystalline nature) and is recognized as one of the most active catalysts for VOCs removal among transition metal oxides [15,16,17]. Nanostructured MnO2 possesses a rich structural flexibility, which adopts various crystallographic forms such as α-, β-, δ-, and γ-MnO2, depending upon the size of the tunnel [18,19]. These polymorphs of MnO2 include a one-dimensional chain-like tunnel (α-, β-, and γ-MnO2) and two-dimensional layer structures (δ-MnO2) based on different linkage ways of the basic octahedral molecular sieves [20,21].
Among the MnO2 polymorphs, α-MnO2 has one of the largest tunnel sizes (4.6 Å) consisting of 1D (1 × 1) and (2 × 2) channels, composed of double chains of edge-sharing [MnO6] octahedra, filled with alkali metal cations, NH4+ or H3O+, inside the 2 × 2 tunnels to stabilize the structure [16]. The pyrolusite-type β-MnO2 has a rutile-type structure with infinite [MnO6] octahedral chains that share opposing edges. Each chain is connected to four similar chain corners, forming the smallest tunnel structure (2.3 Å) of MnO2 polymorphs, consisting of 1D (1 × 1) and (1 × 1) channels [22,23]. It is generally accepted that catalysts with different crystal structures exhibit different catalytic efficiencies and reaction mechanisms for the reactants. For instance, Hayashi et al. evaluated the aerobic catalytic performance of six types of MnO2 (α-, β-, γ-, δ, λ, ε -phases) for the transformation of 5-hydroxymethylfurfural to 2,5-furandicarboxylic acid and concluded the best activity of β-MnO2 [19]. Chen et al. found that the α- and γ-MnO2 presented higher benzene oxidation activity than β- and δ-MnO2, whereas δ-MnO2 displayed the best in formaldehyde oxidation among all of the MnO2 materials [20]. This may be due to the varying oxygen species in different types of MnO2, which play distinct roles in the catalytic oxidation of formaldehyde and benzene. In addition, surface defects of manganese-based catalysts are regarded as an important determinant of their catalytic activity, and their formation and driving of catalytic reactions are often related to their surface oxygen species [24,25,26]. Yang et al. investigated the phase-activity relationship of MnO2 toward toluene catalytic oxidation. They proposed that the excellent catalytic performance of δ-MnO2 may be associated with the rich oxygen vacancy and the strong mobility of oxygen species [27]. Tian et al. prepared α-, β-, and ε- MnO2 for CO oxidation and found that β-MnO2 possesses the lowest energies for oxygen vacancy generation as well as excellent redox properties, thus exhibiting the best CO oxidation activity [13]. These studies highlight the importance of oxygen vacancies in VOC removal and illustrate that the concentration of oxygen vacancies in different crystal structures tends to dominate their catalytic activity. Therefore, it is necessary to understand the relationship between the catalytic activity of manganese oxide and its surface structure to provide a standard for the further modification of MnOx or other metal oxide catalysts. However, noticeable differences are presented in the surface morphology and crystal structure of different crystalline MnO2, which makes it difficult to clarify the contribution of oxygen vacancies to the catalytic reaction. For example, δ-MnO2 is a layered structure, whereas α-, β- and γ-MnO2 are common tunneling structures. Among them, γ-MnO2 is a spherical structure, and α- and β-MnO2 are similar nanorod-like structures. The differences in crystal structure and surface morphology can inherently lead to variations in catalyst surface properties, which can obscure the critical role of oxygen vacancies. Therefore, it is essential to reveal the effect of surface vacancies on catalytic reactions based on the same morphology. In addition, oxygen vacancy-mediated catalytic reactions may be accompanied by the migration and release of oxygen species and the generation of new oxygen vacancies, thus the transformation of these active surface species in catalytic reactions and their contribution to the removal of VOCs need further clarification.
Herein, we compared the removal efficiency of CH3SH by α-MnO2 and β-MnO2 with similar surface morphology but different crystal structures. Their physicochemical properties were subsequently characterized by various analysis techniques. The number of surface low valence Mn, oxygen vacancies and redox properties were studied regarding high-resolution transmission electron microscope (HRTEM) and electron paramagnetic resonance (EPR), X-ray photoelectron spectroscopy (XPS), H2-TPR and O2-TPD. The changes of catalyst surface species before and after the reaction were characterized by XPS, and the variation of intermediate species of CH3SH during the reaction were also monitored.

2. Materials and Methods

2.1. Chemical Reagents

α-MnO2 and β-MnO2 were synthesized through the hydrothermal synthesis method according to previous research [28]. Potassium permanganate (KMnO4, Chengdu Colon Chemicals Co., LTD, Chengdu, China), hydrated manganese sulfate (MnSO4·H2O, Aladdin Reagent Co., LTD, Shanghai, China) and ammonium persulfate ((NH4)2S2O8, Aladdin Reagent Co., LTD, Shanghai, China) were used without further purification.
Synthesis of α-MnO2: 0.1 M KMnO4 and 0.05 M MnSO4·H2O were dissolved in 70 mL deionized water and stirred for 30 min. The resulting solution was transferred to a 100 mL Teflon-lined autoclave and maintained at 160 °C for 12 h. After cooling to room temperature, the precipitate was centrifuged and washed with distilled water (700–1000 mL) three times. Finally, the precipitate was dried at 80 °C for 4 h and calcination at 360 °C for 2 h.
Synthesis of β-MnO2: 0.14 M MnSO4·H2O and 0.14 M (NH4)2S2O8 were dissolved in 70 mL deionized water and stirred for 30 min. The resulting solution was transferred to a 100 mL Teflon-lined autoclave and maintained at 140 °C for 12 h. After cooling to room temperature, the precipitate was centrifuged and washed with distilled water (700–1000 mL) three times. Finally, the precipitate was dried at 80 °C for 4 h and calcination at 360 °C for 2 h.

2.2. Catalyst Characterization

The refined test of X-ray powder diffraction (XRD) of the products was performed using a Bragg-Brentano-type powder diffractometer (Nihongo TTRIII, Tokyo City, Japan, operated at 40 kV and 200 mA, Cu Kα radiation, λ = 0.15418 nm). To investigate the Brunauer-Emmett-Teller (BET) surface areas, average pore diameters, and total pore volumes of the samples, N2 adsorption-desorption isotherms were determined using a NOVA 4200e Surface Area and Pore Size Analyzer. Electron paramagnetic resonance (EPR) signals were carried out on a Bruker A300 spectrometer (Saarbrucken, Germany) at 25 °C. XPS profiles were obtained with a Thermo Scientific K-Alpha spectrometer (Waltham, MA, USA). The binding energy (BE) values were calibrated using the C 1 s peak at 284.8 eV. The Raman spectra were recorded using a 514 nm laser excitation source with an integration time of 3 s and 30 accumulations (Raman, BX41, HOEIBA Scientific, Paris, France). Scanning electron microscopy (SEM, VEGA3SBH, Brno, Czech Republic) and high-resolution transmission electron microscopy (HRTEM, Talos F200X, Thermo Scientific, Waltham, MA, USA) were used to observe catalyst morphologies.
Hydrogen temperature-programmed reduction (H2-TPR) and oxygen temperature- programmed desorption (O2-TPD) experiments were performed on a FULI II 7970 gas chromatograph (Fuli Analytical Instrument Inc., Hangzhou, China) with a thermal conductivity detector (TCD). In H2-TPR experiments, 50 mg of the sample was placed in a quartz tube and pretreated in a gas flow of 10% H2/Ar (30 mL min−1) at 100 °C for 30 min to remove impurities. After the pretreatment process, the sample was reduced by 10% H2/Ar (30 mL min−1) from 100 to 800 °C with a heating rate of 10 °C/min. For O2-TPD analysis, 50 mg of sample was loaded on the quartz tube, heated to 105 °C and pretreated with He (30 mL min−1) for 30 min to remove surface adsorbed water, followed by cooling to 30 °C. Subsequently, the sample was adsorbed by 10% O2/He (30 mL min−1) at room temperature for 60 min, and then He (30 mL min−1) was used to purge the sample for 30 min to remove physically adsorbed O2 and stabilize the baseline. Subsequently, the temperature was ramped from 30 to 850 °C at 10 °C mL min−1.

2.3. Catalyst Activity Evaluation

The catalytic performance for the CH3SH decomposition was investigated in a fixed-bed quartz tube reactor (i.d. = 6 mm). 200 mg samples with the size of 40–60 meshes were loaded into the reactor. The reaction temperature was controlled and maintained for about 1 h at each designated temperature. The inlet CH3SH concentration was set at 5000 ppm, and the total flow rate was maintained at 30 mL min−1. The concentration of CH3SH was recorded by GC-9790 (FULI, China) equipped with a flame ionization detector (FID) and flame photometric detector (FPD), and the CH3SH conversion ratio was calculated as follows:
CH 3 SH   conversion = C i n C o u t C i n × 100 %
C i n   represents the inlet concentration of CH3SH and C o u t is the outlet concentration of CH3SH.
The reaction rates of CH3SH decomposition were determined in the kinetic regime at a CH3SH conversion lower than 20% at different temperatures; The reaction rate (rCH3SH; mol g−1 h−1) for CH3SH decomposition was calculated according to the following equations:
r C H 3 S H = C C H 3 S H × X C H 3 S H × F m c a t
r n o r m = C C H 3 S H × X C H 3 S H × F m c a t × S B E T
where the   C C H 3 S H   represents the initial methyl mercaptan concentration, F (mol·h−1) represents the total flow rate, X C H 3 S H   denotes CH3SH conversion, and S B E T (m2·g−1) represents the specific surface area of catalysts.
The turnover frequency (TOF, h−1) was calculated for different crystal types based on oxygen vacancy concentration for MnO2, and indicates the number of reactions of methyl mercaptan at each active site per unit of time, thus TOF was obtained using the following equation:
T O F h 1 = C C H 3 S H × X C H 3 S H × F m M n O 2 M M n O 2 × M n 2 + + M n 3 +
M M n O 2 (mol·g−1) is the molar mass of M n O 2 , and M n 2 + + M n 3 + derived from XPS data, which represent the concentration of the oxygen vacancies of MnO2 deduced from the obtained XPS spectra.

3. Results and Discussion

3.1. Structure and Morphology

XRD was used to determine the crystal structure of the prepared material. The XRD patterns of as-prepared MnO2 with various crystal types are shown in Figure 1. The diffraction peaks located at ~12.7°, ~18.1°, ~28.8°, and ~37.5° can be assigned to α-MnO2 (JCPDS card no. 44-0141) (Figure 1A), and the peaks at ~28.7°, ~37.2°, ~42.7° and ~56.4° can be ascribed to β-MnO2 (JCPDS card no. 24-0735) (Figure 1B) [29]. There was no apparent crystal transformation on these samples after 360 °C calcination. The sharp and strait peaks of β-MnO2 could indicate its great crystallization and large grain size; α-MnO2 presented wide bands with relatively lower crystallinity and smaller grain sizes. The above results indicate the successful obtaining of the two kinds of MnO2 with specific crystal phases.
The morphologies of MnO2 samples were characterized by a scanning electron microscopy (SEM), a transmission electron microscope (TEM), and a high-resolution TEM (HRTEM). As can be seen in Figure 2A, α-MnO2 showed a stacking-nanorod structure with an average length of about 330 nm. β-MnO2 (Figure 2F) also showed a typical rod shape with a diameter near 50 nm and a length of ~1.2 μm. TEM showed consistent results with the SEM that α-MnO2 (Figure 2B) and β-MnO2 (Figure 2G) had similar nanorod-like structures as previously reported [30]. The well-identified periodic lattice fringes of 6.9 Å can be clearly observed in Figure 2C, corresponding to the interplanar distance of (110) facet of α-MnO2. Figure 2H exhibited the lattice fringes of 3.1 Å, which match the interplanar distance of the (110) facet of β-MnO2 well. Compared with the β-MnO2 samples, α-MnO2 showed more blurry lattice fringes, representing poor crystallinity, which also agrees well with the XRD patterns. In addition, the presence of defects was further demonstrated using the inverse Fast Fourier Transform (FFT) pattern (Figure 2E,J). Significantly more lattice distortion can be clearly observed on the surface of α-MnO2 (Figure 2D) (highlighted by red ovals), thus leading to more defects than β-MnO2 [10]. Besides, severe blurring of the lattice fringes was also detected on α-MnO2 than β-MnO2. Lattice distortion can be caused by nearby point defects. Simultaneously, a defect layer will be formed once the defect concentration is high enough, resulting in a blurry lattice fringe in the HRTEM images [31]. Hence, the intrinsic defective structure of α-MnO2 was confirmed. Oxygen vacancies, as an important point defect in catalysts, play a prominent role in the catalytic reaction process, and the high oxygen vacancy concentration will result in a blurry lattice fringe, which can be reflected in the HRTEM images [32]. As shown in Figure 3, the EPR signal corresponding to g = 2.003 can be attributed to oxygen vacancies, and its signal intensity can represent the number of oxygen vacancies [27]. Therefore, more oxygen vacancies on α-MnO2 than β-MnO2 can be confirmed based on EPR, consistent with HRTEM analysis.
The BET surface areas (SBET), and pore volumes of the two catalysts are shown in Figure 4A,B. It is reported that the different structures assembled by MnO6 octahedra in MnO2 will affect the related surface areas and pore volumes. β-MnO2 presented relatively low specific surface areas (12.76 m2 g−1) and pore volumes (0.06 cm3 g−1), whereas α-MnO2 showed higher specific surface areas (34.59 m2 g−1) and pore volumes (0.13 cm3 g−1). Moreover, the nitrogen adsorption-desorption isotherms of α-MnO2 and β-MnO2 displayed a type IV curve with H3-type hysteresis loops, indicating that both samples were mesoporous structures [33].

3.2. Catalytic Performance

In order to explore the activity of two catalysts on sulfur-containing volatile organic pollutants (S-VOCs), methyl mercaptan (CH3SH) was chosen as the model S-VOCs, and the catalytic activities of α-MnO2 and β-MnO2 are shown in Figure 5A. Two MnO2 samples exhibited significantly different catalytic performance in CH3SH catalytic reaction. α-MnO2 (74%) exhibited significantly better catalyst activity than β-MnO2 (3%) at 30 °C. The decreases of CH3SH conversion for α-MnO2 at 50 °C may be due to the desorption of CH3SH on the catalyst. As the temperature increased, the conversion of CH3SH reached 100% at 100 °C with both catalysts. Furthermore, the reaction rates of α-MnO2 and β-MnO2 at 50 °C were calculated based on the activity experiments. As shown in Figure 5B, α-MnO2 showed the CH3SH reaction rate of 2.9 × 10−2 mol g−1 h−1, this being ~4.1 times higher than the rates measured for and β-MnO2 at 50 °C, which was consistent with the results for the catalytic activity.
It is well known that the specific surface area plays a critical role in catalytic reactions. To eliminate its influence, the reaction rates with surface area normalization were calculated at different temperatures based on the data from the activity experiments. The results of the normalized reaction rates (rnorm, mol m−2 h−1) of CH3SH decomposition are shown in Figure 6A. The normalized reaction rates for α-MnO2 were obviously higher than those of β-MnO2 at 30, 50, 60 and 80 °C, which suggested that reactivity was not governed by the specific surface area. Turnover frequency (TOF) is essential for studying the intrinsic reactivity of catalysts. In this work, the TOF (h−1) was calculated based on oxygen vacancy concentration, and the TOF value for the CH3SH catalytic decomposition was conducted at 50 °C with 0.01 g of catalyst and was calculated within a low CH3SH conversion (1 h of reaction, below 15.0%). As displayed in Figure 6B, the α-MnO2 showed the highest TOF value of 0.14 h−1, which was 1.8 times as that of the β-MnO2 (0.08 h−1), indicating that α-MnO2 has better catalytic performance for CH3SH.

3.3. Redox Capacity and Oxygen Species

To evaluate the reduction behaviors of MnO2 samples, H2-temperature-programmed reduction (TPR) was performed (Figure 7A). For α-MnO2, the peaks at 289 and 309 °C corresponded to the reduction of Mn4+ → Mn3+ and Mn3+ → Mn2+, respectively, and the peaks around 291 and 317 °C for β-MnO2 were attributed to Mn4+ → Mn3+ and Mn3+ → Mn2+, respectively [20,34]. The reduction temperature of α-MnO2 was lower than that of β-MnO2, indicating that the reduction of α-MnO2 is relatively faster. More importantly, the more remarkable reduction ability of α-MnO2 means easier deoxygenation during hydrogen treatment, suggesting that oxygen migration is more likely to occur on its surface. Therefore, α-MnO2 features stronger oxygen species mobility than β-MnO2.
O2-TPD was conducted further to explore the oxygen species of the MnO2 catalysts. Figure 7B shows three desorption peaks related to oxygen species that can be observed on the MnO2. The low-temperature peak below 400 °C was ascribed to the chemisorbed active oxygen species on the surface (O and O2) [35]. The desorption peaks at 400–650 °C and 700–850 °C were related to the release of subsurface and bulk lattice oxygen species (O2−), respectively [36,37]. The desorption of surface oxygen at low temperature (<400 °C) plays the primary role as reactive oxygen species participating in the catalytic reaction [38]. Moreover, the lower temperature of the surface oxygen desorption peak means better low-temperature mobility of oxygen species. As depicted in Figure 7B, α-MnO2 showed a lower temperature at 146 °C of the surface oxygen desorption peak than β-MnO2 at 367 °C, indicating the better low-temperature mobility of oxygen species, which is in agreement with H2-TPR.
More bonding properties were discussed through Raman spectra (Figure 8A). The peaks at 348 and 640 cm−1 corresponded to the Mn-O bending and the stretching vibration, respectively [39]. Significantly weaker and broader Raman peaks at around 640 cm−1 were detected for α-MnO2 than β-MnO2, suggesting lower crystallinity and more defects due to the lattice distortion [40]. To evaluate the strength of the Mn-O bond, the bond force constant (k) was calculated from Hooke’s law [41,42] using the following equation: ω =   1 2 Π c k μ , where ω is the Raman shift (cm−1), c is light velocity, and μ is the effective mass of the Mn-O bond. The calculated Mn-O force constant (k) is shown in the inset of Figure 8B. Thus, the Mn-O bond force constant of α-MnO2 (293 N/m) was smaller than that of β-MnO2 (296 N/m), implying the weaker Mn-O bond. The weaker Mn-O bond means easier migration of O and easier redox of Mn during the reaction, which is beneficial for catalytic reactions [25].

3.4. Identification of the Role of Oxygen Vacancies in CH3SH Degradation

The surface elemental composition and chemical state of these MnO2 samples were identified by XPS. The XPS spectra of Mn 2p3/2 of the samples are shown in Figure 9. The peaks corresponding to binding energies at 642.7, 641.7 and 640.6 eV can be attributed to Mn4+, Mn3+ and Mn2+, respectively [43,44]. It is noteworthy that the binding energy corresponding to different valences of Mn were slightly different in both MnO2 samples, indicating that crystal phase structure has a certain effect on the electron density of the MnO2 surface, which is related to the degree of charge imbalance, oxygen vacancies, as well as the relative content of Mn2+, Mn3+ and Mn4+ [45]. Specifically, the oxygen vacancy will be generated to maintain electrostatic balance with the increasing Mn2+ and Mn3+ proportion, and the proportion of low-valence Mn is generally regarded as an indicator of surface oxygen vacancies [44]. As shown in Table 1 and Figure 9A, the proportion of the low valence Mn (Mn2+ + Mn3+) showed α-MnO2 (41.62%) > β-MnO2 (37.74%). Besides, the average oxidation state (AOS) of MnO2 was calculated according to the formula of AOS = 8.956 − 1.126 ΔE [46], which was based on the size of Mn 3s multiple splitting (ΔE) in Mn 3s XPS spectra (Figure 10A). The AOS values of the Mn element in α-MnO2 and β-MnO2 were calculated to be 3.57 and 3.75, respectively. In previous studies, lower AOS of MnO2 was also able to indicate more surface oxygen vacancies [47]. Therefore, it can be inferred that α-MnO2 has a greater surface oxygen vacancy density than β-MnO2 (in agreement with HETEM and EPR).
Additionally, Mn2+-O and Mn3+-O bonds are weaker than Mn4+-O [48]. Thus, the higher ratio of low valence Mn (Mn2+ + Mn3+) endows MnO2 with a larger proportion of weaker M-O bonds on its surface, meaning the easier release of O to participate in the reaction. Figure 9A showed that α-MnO2 had more low valence Mn (Mn2+ + Mn3+) than β-MnO2, which also implies the easier release of surface oxygen species, consistent with the results of H2-TPR and O2-TPD. Comparison of the Mn 2p3/2 spectra before and after the reaction (Figure 9B) of MnO2 with CH3SH showed that the valence state of Mn in both samples changed obviously, suggesting the electron transfer during the reaction. For both samples, Mn4+ decreased, and Mn2+ and Mn3+ increased, proving that the high-valent Mn (IV) was reduced by gaining electrons during the reaction. After the reaction, the Mn2++Mn3+/Mn4+ of α-MnO2 increased by 1.51 and that of β-MnO2 by 1.39, and AOS decreased by 0.74 for α-MnO2 and that of β-MnO2 decreased by 0.5 (Figure 10B), testifying that α-MnO2 was reduced to a greater extent by gaining more electrons than β-MnO2, which can well match H2-TPR results. It is noteworthy that higher AOS usually indicates a stronger electron-gaining ability of the catalyst because of the presence of more high-valent atoms, however, α-MnO2 exhibited a stronger electron-gaining ability in the reaction with CH3SH, suggesting that oxygen vacancies play a more important role in catalyzing CH3SH comparing to the high-valent Mn. This may explain the fact that chemisorption is the rate-limiting step for electron transfer, and more surface oxygen vacancies provide more surface adsorption sites for CH3SH.
In addition, the reaction between CH3SH and the catalyst could change the electronic environment of the catalyst. During the reaction, as the ratio of Mn2+ and Mn3+ increased, weaker Mn-O bonds were continuously formed and broken, leading to deoxygenation and further generation of oxygen vacancies to maintain electrostatic equilibrium, which may provide new sites for the reaction, and these desorbed oxygen species may favor the catalytic oxidation of CH3SH as well.
The XPS spectra of O 1s of the samples are shown in Figure 11. Peaks with binding energies at 529–529.8, 530.9–532 and 533 eV in the XPS spectra of O 1s of the MnO2 samples (Figure 11) can be attributed to lattice oxygen (Olatt) and surface adsorption oxygen (Oads), and surface hydroxyl oxygen (OadsO-H), respectively [49,50,51]. The molar ratio of Oads/Olatt is shown in Table 2 and follows the order of α-MnO2 (0.65) > β-MnO2(0.35). Oads was generally considered the most reactive oxygen species in the catalytic reaction and capable of participating in the catalytic oxidation of VOCs in previous reports [52,53]. The oxygen species changes before and after the reaction are shown in Figure 11B and Table 2. The Olatt and Oads for both two materials decreased and increased, respectively, suggesting the migration of Olatt to form Oads during the reduction of Mn. Obviously, α-MnO2 formed more Oads after the reaction, which corresponds to its greater degree of reduction, also implying that α-MnO2 has a stronger catalytic capacity.

3.5. Product Detection during the Reaction

The main gas phase products were monitored quantitatively to better understand the reaction process of CH3SH over two different catalysts. As displayed in Figure 12, the decomposition of CH3SH at different temperatures corresponded to the production of CH3SSCH3. Meanwhile, the concentration of CH3SH during the reaction showed an excellent correlation with the concentration of CH3SSCH3, indicating that CH3SSCH3 was the main gas-phase product. At 150 °C, CH3SH was completely decomposed for both MnO2 catalysts, consistent with the thermodynamic theory that catalytic reactions proceed easier at higher temperatures. The yield of CH3SSCH3 gradually decreased when T > 100 °C, which may be due to the further catalytic oxidation of CH3SSCH3 at higher temperatures.
Figure 13 shows the changes of S 2p before and after the reaction, which was used to detect the solid-phase intermediates in the reaction process. No S species were detected on two catalysts before the reaction. In contrast, significant amounts of S species were detected on both samples after the reaction, indicating that some sulfur-containing products were adsorbed on the catalyst surface. As shown in Figure 13B, three peaks at 163.2, 167.9, and 169.2 eV corresponding to S0, SO32− (S4+), and SO42− (S6+), respectively [30,54], which all showed higher valence than S2- from CH3SH, indicating the oxidation of S during the reaction. It is worth noting that the catalytic experiments in CH3SH were performed under a nitrogen atmosphere, so it can be concluded that the O in the S-O species was mainly derived from the Oads of MnO2. This illustrated the Oads involvement in the catalytic reaction of CH3SH. SO32− and SO42− were mainly retained on the manganese dioxide surface in the form of MnSO3 and MnSO4, implying that chemisorption was a prerequisite for the decomposition of CH3SH on MnO2. Notably, reacted α-MnO2 showed a higher proportion of SO42− (S6+) (26.31%) than β-MnO2 (12.56%) (Table 3 and Figure 13B), suggesting a greater degree of S oxidation, which corresponds to a greater reduction of Mn4+ after the reaction (Table 1 and Figure 9B). Furthermore, SO42− requires more oxygen to be coordinated with S than SO32−, so a higher proportion of SO42− production requires more Oads to participate in the reaction. Correspondingly, the H2-TPR, O2-TPD, and XPS analysis demonstrated more Oads and better surface oxygen mobility for α-MnO2 than β-MnO2.
Based on the above experimental and characterization analysis, the catalytic mechanism of CH3SH by α-MnO2 and β-MnO2 can be inferred in Figure 14. CH3SH was first chemisorbed on the MnO2 surface and subsequently underwent a single electron transfer to form CH3S·, and then the two CH3S· were coupled to form CH3SSCH3. Based on the formation of S-S bonds, it is speculated that the single electron transfer occurs on S, suggesting that the chemisorption may be through the formation of Mn-S bonds. Moreover, more lattice oxygen was released during the reduction of Mn, which was involved in the further catalytic oxidation of S-containing species to produce SO32− and SO42−, and may further form new oxygen vacancies to support more active sites. Although β-MnO2 enjoys a higher AOS based on the proportion of high valence Mn, α-MnO2 showed better catalytic activity due to more oxygen vacancies and stronger oxygen mobility.

4. Conclusions

In this paper, MnO2 catalysts with similar morphology but different crystal structures (α-MnO2 and β-MnO2) were successfully prepared, and the effects of the physicochemical properties on the catalytic activities were systematically investigated. Both Mn-based catalysts showed significant removal of CH3SH at 150 °C, achieving complete conversion, whereas α-MnO2 exhibited significantly better catalytic activity compared to β-MnO2 at a lower temperature (T < 100 °C) under a GHSV of 9000 mL g−1 h−1. Coupled with O2-TPD, H2-TPR, Raman spectra, XPS, EPR, and HRTEM, it was demonstrated that α-MnO2 has more oxygen vacancies, stronger surface oxygen migration ability, and better redox properties, which can be favorable for CH3SH decomposition. The readily released lattice oxygen during the reaction promoted further oxidative decomposition of S-containing species. The intermediate products of the solid and gas phases were determined as CH3SSCH3 and the S0, SO32−, and SO42−, respectively. The catalytic mechanism was further proposed as the oxygen vacancies on MnO2 provided active sites for the adsorption of CH3SH, facilitating the electron transfer of MnO2 with CH3SH, and the oxygen species derived from the Mn surface were further involved in the CH3SH catalytic oxidation. The findings of this study are essential for broadening the application of Mn-based catalysts in the removal of S-VOCs and providing new insights into the mechanism of interfacial reactions between VOCs and metal-based catalysts.

Author Contributions

Conceptualization, J.L. (Jiangping Liu) and Y.L.; data curation, H.S., J.L. (Jiangping Liu), Y.H., T.A. and Y.L.; writing—original draft preparation, H. S.; writing—review and editing, J.L. (Jiangping Liu) and J.L. (Jichang Lu); investigation: C.G. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by National Natural Science Foundation of China (No. 42207127, 42030712, 21966018), Key Project of Natural Science Foundation of Yunnan Province (Grant No. 202101AS070026) and Applied Basic Research Foundation of Yunnan Province (Grant No. 202101BE070001-027, 2022J0069 and 202105AE160019).

Data Availability Statement

All date used to support the findings of this study are included within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Guo, C.; Zhu, J.; He, J.; Hu, L.; Li, D. Catalytic oxidation/photocatalytic degradation of ethyl mercaptan on α-MnO2@H4Nb6O17-NS nanocomposite. Vacuum 2020, 182, 109718. [Google Scholar] [CrossRef]
  2. Pétron, G.; Frost, G.; Miller, B.R.; Hirsch, A.I. Hydrocarbon emissions characterization in the Colorado Front Range: A pilot study. J. Geophys. Res. 2012, 117, D04304. [Google Scholar] [CrossRef]
  3. He, D.; Zhao, Y.; Yang, S.; Mei, Y.; Yu, J.; Liu, J.; Chen, D.; He, S.; Luo, Y. Enhancement of catalytic performance and resistance to carbonaceous deposit of lanthanum (La) doped HZSM-5 catalysts for decomposition of methyl mercaptan. Chem. Eng. J. 2018, 336, 579–586. [Google Scholar] [CrossRef]
  4. Lu, J.; Hao, H.; Zhang, L.; Xu, Z.; Zhong, L.; Zhao, Y.; He, D.; Liu, J.; Chen, D.; Pu, H.; et al. The investigation of the role of basic lanthanum (La) species on the improvement of catalytic activity and stability of HZSM-5 material for eliminating methanethiol-(CH3SH). Appl. Catal. B 2018, 237, 185–197. [Google Scholar] [CrossRef]
  5. Zhao, S.; Yi, H.; Tang, X.; Kang, D.; Gao, F.; Wang, J.; Huang, Y.; Yang, Z. Removal of volatile odorous organic compounds over NiAl mixed oxides at low temperature. J. Hazard. Mater. 2018, 344, 797. [Google Scholar] [CrossRef]
  6. Yi, H.; Tao, T.; Zhao, S.; Yu, Q.; Gao, F.; Zhou, Y.; Tang, X. Promoted adsorption of methyl mercaptan by γ-Al2O3 catalyst loaded with Cu/Mn. Environ. Technol. Inno. 2021, 21, 101349. [Google Scholar] [CrossRef]
  7. Roman, P.; Veltman, R.; Bijmans, F.M.M.; Keesman, K.J.; Janssen, A.J.H. Effect of methanethiol concentration on sulfur production in biological desulfurization systems under haloalkaline conditions. Environ. Sci. Technol. 2015, 49, 9212. [Google Scholar] [CrossRef]
  8. Hu, L.; He, H.; Xia, D.; Huang, Y.; Xu, J.; Li, H.; He, C.; Yang, W.; Shu, D.; Wong, P.K. Highly Efficient Performance and Conversion Pathway of Photocatalytic CH3SH Oxidation on Self-Stabilized Indirect Z-Scheme g-C3N4/I3-BiOI. ACS Appl. Mater. Inter. 2018, 10, 18693–18708. [Google Scholar] [CrossRef]
  9. He, C.; Liao, Y.; Chen, C.; Xia, D.; Wang, Y.; Tian, S.; Yang, J.; Shu, D. Realizing a redox-robust Ag/MnO2 catalyst for efficient wet catalytic ozonation of S-VOCs: Promotional role of Ag(0)/Ag(I)-Mn based redox shuttle. Appl. Cataly. B 2022, 303, 120881. [Google Scholar] [CrossRef]
  10. Lian, Q.; Hu, L.; Ma, D.; Jiao, Y.; Xia, D.; Huang, Y.; Tang, Z.; Qu, W.; Zhao, H.; He, C.; et al. Interstitial Atomic Bi Charge-Alternating Processor Boosts Twofold Molecular Oxygen Activation Enabling Rapid Catalytic Oxidation Reactions at Room Temperature. Advan. Func. Mater. 2022, 32, 2205054. [Google Scholar] [CrossRef]
  11. Liu, J.; He, D.; Chen, D.; Hao, H.; Yu, J.; Lu, J.; Liu, F.; Liu, P.; Zhao, Y.; Luo, Y. Promotional effects of rare-earth (La, Ce and Pr) modification over HZSM-5 for methyl mercaptan catalytic decomposition. J. Taiwan Ins.Chem. Eng. 2017, 80, 262–268. [Google Scholar] [CrossRef]
  12. Zhao, Y.; Chen, D.; Liu, J.; He, D.; Cao, X.; Han, C.; Lu, J.; Luo, Y. Tuning the metal-support interaction on chromium-based catalysts for catalytically eliminate methyl mercaptan: Anchored active chromium species through surface hydroxyl groups. Chem. Eng. J. 2020, 389, 124384. [Google Scholar] [CrossRef]
  13. Tian, F.-X.; Li, H.; Zhu, M.; Tu, W.; Lin, D.; Han, Y.-F. Effect of MnO2 Polymorphs’ Structure on Low-Temperature Catalytic Oxidation: Crystalline Controlled Oxygen Vacancy Formation. ACS Appl. Mater. Inter. 2022, 14, 18525–18538. [Google Scholar] [CrossRef]
  14. Huang, J.; Zhong, S.; Dai, Y.; Liu, C.-C.; Zhang, H. Effect of MnO2 Phase Structure on the Oxidative Reactivity toward Bisphenol A Degradation. Environ. Sci. Technol. 2018, 52, 11309–11318. [Google Scholar] [CrossRef]
  15. Zhang, K.; Han, X.; Hu, Z.; Zhang, X.; Tao, Z.; Chen, J. Nanostructured Mn-based oxides for electrochemical energy storage and conversion. Chem. Soc. Rev. 2015, 44, 699–728. [Google Scholar] [CrossRef]
  16. Yang, R.; Fan, Y.; Ye, R.; Tang, Y.; Cao, X.; Yin, Z.; Zeng, Z. MnO2-Based Materials for Environmental Applications. Advan. Mater. 2021, 33, 2004862. [Google Scholar] [CrossRef]
  17. Wang, Y.; Liu, K.; Wu, J.; Hu, Z.; Huang, L.; Zhou, J.; Ishihara, T.; Guo, L. Unveiling the Effects of Alkali Metal Ions Intercalated in Layered MnO2 for Formaldehyde Catalytic Oxidation. ACS Catal. 2020, 10, 10021–10031. [Google Scholar] [CrossRef]
  18. Sun, W.; Kitchaev, D.A.; Kramer, D.; Ceder, G. Non-equilibrium crystallization pathways of manganese oxides in aqueous solution. Nat. Commun. 2019, 10, 573. [Google Scholar] [CrossRef] [Green Version]
  19. Hayashi, E.; Yamaguchi, Y.; Kamata, K.; Tsunoda, N.; Kumagai, Y.; Oba, F.; Hara, M. Effect of MnO2 Crystal Structure on Aerobic Oxidation of 5-Hydroxymethylfurfural to 2,5-Furandicarboxylic Acid. Jo. Amer. Chem. Soc. 2019, 141, 890–900. [Google Scholar] [CrossRef]
  20. Chen, B.; Wu, B.; Yu, L.; Crocker, M.; Shi, C. Investigation into the Catalytic Roles of Various Oxygen Species over Different Crystal Phases of MnO2 for C6H6 and HCHO Oxidation. ACS Catal. 2020, 10, 6176–6187. [Google Scholar] [CrossRef]
  21. Shi, J.; Qi, T.; Sun, B.C.; Chu, G.W.; Chen, J.F. Catalytic oxidation of benzyl alcohol over MnO2: Structure-activity description and reaction mechanism. Chem. Eng. J. 2022, 440, 135802. [Google Scholar] [CrossRef]
  22. Jiao, F.; Bruce, P.G. Mesoporous Crystalline β-MnO2-a Reversible Positive Electrode for Rechargeable Lithium Batteries. Advan. Mater. 2007, 19, 657–660. [Google Scholar] [CrossRef]
  23. Hayashi, E.; Yamaguchi, Y.; Kita, Y.; Kamata, K.; Hara, M. One-pot aerobic oxidative sulfonamidation of aromatic thiols with ammonia by a dual-functional β-MnO2 nanocatalyst. Chem. Commun. 2020, 56, 2095–2098. [Google Scholar] [CrossRef] [PubMed]
  24. Li, H.; Shang, H.; Li, Y.; Cao, X.; Yang, Z.; Ai, Z.; Zhang, L. Interfacial Charging–Decharging Strategy for Efficient and Selective Aerobic NO Oxidation on Oxygen Vacancy. Environ. Sci. Technol. 2019, 53, 6964–6971. [Google Scholar] [CrossRef]
  25. Rong, S.; Zhang, P.; Liu, F.; Yang, Y. Engineering Crystal Facet of α-MnO2 Nanowire for Highly Efficient Catalytic Oxidation of Carcinogenic Airborne Formaldehyde. ACS Catal. 2018, 8, 3435–3446. [Google Scholar] [CrossRef]
  26. Wang, Y.; Liu, M.; Hu, C.; Xin, Y.; Ma, D.; Gao, M.; Xie, H. Enhanced MnO2/peroxymonosulfate activation for phthalic acid esters degradation: Regulation of oxygen vacancy. Chem. Eng. J. 2022, 433, 134048. [Google Scholar] [CrossRef]
  27. Yang, R.; Guo, Z.; Cai, L.; Zhu, R.; Fan, Y.; Zhang, Y.; Han, P.; Zhang, W.; Zhu, X.; Zhao, Q.; et al. Investigation into the Phase-Activity Relationship of MnO2 Nanomaterials toward Ozone-Assisted Catalytic Oxidation of Toluene. Small 2021, 17, 2103052. [Google Scholar] [CrossRef]
  28. Yang, W.; Su, Z.A.; 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 2020, 260, 118150. [Google Scholar] [CrossRef]
  29. Zhao, Z.; Li, G.; Sun, Y.; Li, N.; Zhang, Z.; Cheng, J.; Ma, C.; Hao, Z. The positive effect of water on acetaldehyde oxidation depended on the reaction temperature and MnO2 structure. Appl. Catal. B 2022, 303, 120886. [Google Scholar] [CrossRef]
  30. Zheng, X.; Zhang, G.; Yao, Z.; Zheng, Y.; Shen, L.; Liu, F.; Cao, Y.; Liang, S.; Xiao, Y.; Jiang, L. Engineering of crystal phase over porous MnO2 with 3D morphology for highly efficient elimination of H2S. J. Hazard. Mater. 2021, 411, 125180. [Google Scholar] [CrossRef]
  31. 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 2017, 209, 729–737. [Google Scholar] [CrossRef]
  32. 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 2018, 560, 195–205. [Google Scholar] [CrossRef]
  33. Yu, D.; Wang, L.; Zhang, C.; Peng, C.; Yu, X.; Fan, X.; Liu, B.; Li, K.; Li, Z.; Wei, Y.; et al. Alkali Metals and Cerium-Modified La-Co-Based Perovskite Catalysts: Facile Synthesis, Excellent Catalytic Performance, and Reaction Mechanisms for Soot Combustion. ACS Catal. 2022, 12, 15056–15075. [Google Scholar] [CrossRef]
  34. Guan, S.; Huang, Q.; Ma, J.; Li, W.; Ogunbiyi, A.T.; Zhou, Z.; Chen, K.; Zhang, Q. HCHO Removal by MnO2(x)-CeO2: Influence of the Synergistic Effect on the Catalytic Activity. Ind. Eng. Chem. R. 2020, 59, 596–608. [Google Scholar] [CrossRef]
  35. Chen, L.; Zhang, C.; Li, Y.; Chang, C.-R.; He, C.; Lu, Q.; Yu, Y.; Duan, P.; Zhang, Z.; Luque, R. Hierarchically Hollow MnO2@CeO2 Heterostructures for NO Oxidation: Remarkably Promoted Activity and SO2 Tolerance. ACS Catal. 2021, 11, 10988–10996. [Google Scholar] [CrossRef]
  36. Zheng, C.; Bao, S.; Mao, D.; Xu, Z.; Zheng, S. Insight into phase structure-dependent soot oxidation activity of K/MnO2 catalyst. J. Environ. Sci. 2023, 126, 668–682. [Google Scholar] [CrossRef]
  37. Wu, S.; Liu, H.; Huang, Z.; Xu, H.; Shen, W. O-vacancy-rich porous MnO2 nanosheets as highly efficient catalysts for propane catalytic oxidation. Appl. Catal. B 2022, 312, 121387. [Google Scholar] [CrossRef]
  38. Wan, J.; Zhou, L.; Deng, H.; Zhan, F.; Zhang, R. Oxidative degradation of sulfamethoxazole by different MnO2 nanocrystals in aqueous solution. J. Molecular Catal. A 2015, 407, 67–74. [Google Scholar] [CrossRef]
  39. Cheng, S.; Yang, L.; Chen, D.; Ji, X.; Jiang, Z.; Ding, D.; Liu, M. Phase evolution of an alpha MnO2-based electrode for pseudo-capacitors probed by in operando Raman spectroscopy. Nano Energy 2014, 9, 161–167. [Google Scholar] [CrossRef]
  40. Shan, C.; Zhang, Y.; Zhao, Q.; Fu, K.; Zheng, Y.; Han, R.; Liu, C.; Ji, N.; Wang, W.; Liu, Q. Acid Etching-Induced In Situ Growth of λ-MnO2 over CoMn Spinel for Low-Temperature Volatile Organic Compound Oxidation. Environ. Sci. Technol. 2022, 56, 10381–10390. [Google Scholar] [CrossRef]
  41. Chen, G.; Hong, D.; Xia, H.; Sun, W.; Shao, S.; Gong, B.; Wang, S.; Wu, J.; Wang, X.; Dai, Q. Amorphous and homogeneously Zr-doped MnOx with enhanced acid and redox properties for catalytic oxidation of 1,2-Dichloroethane. Chem. Eng. J. 2022, 428, 131067. [Google Scholar] [CrossRef]
  42. Wei, Y.J.; Yan, L.Y.; Wang, C.Z.; Xu, X.G.; Wu, F.; Chen, G. Effects of Ni Doping on [MnO6] Octahedron in LiMn2O4. J. Phy. Chem. B 2004, 108, 18547–18551. [Google Scholar] [CrossRef]
  43. Zhang, S.; Wang, H.; Si, H.; Jia, X.; Wang, Z.; Li, Q.; Kong, J.; Zhang, J. Novel Core–Shell (ε-MnO2/CeO2)@CeO2 Composite Catalyst with a Synergistic Effect for Efficient Formaldehyde Oxidation. ACS Appl. Mater. Inter. 2020, 12, 40285–40295. [Google Scholar] [CrossRef] [PubMed]
  44. Wu, P.; Dai, S.; Chen, G.; Zhao, S.; Xu, Z.; Fu, M.; Chen, P.; Chen, Q.; Jin, X.; Qiu, Y.; et al. Interfacial effects in hierarchically porous α-MnO2/Mn3O4 heterostructures promote photocatalytic oxidation activity. Appl. Catal. B 2020, 268, 118418. [Google Scholar] [CrossRef]
  45. Nawaz, F.; Cao, H.; Xie, Y.; Xiao, J.; Chen, Y.; Ghazi, Z.A. Selection of active phase of MnO2 for catalytic ozonation of 4-nitrophenol. Chemosphere 2017, 168, 1457–1466. [Google Scholar] [CrossRef] [PubMed]
  46. Ding, D.; Zhou, Y.; He, T.; Rong, S. Facet selectively exposed α-MnO2 for complete photocatalytic oxidation of carcinogenic HCHO at ambient temperature. Chem. Eng. J. 2022, 431, 133737. [Google Scholar] [CrossRef]
  47. Jia, J.; Zhang, P.; Chen, L. Catalytic decomposition of gaseous ozone over manganese dioxides with different crystal structures. Appl. Catal. B 2016, 189, 210–218. [Google Scholar] [CrossRef]
  48. Zhang, Y. Electronegativities of elements in valence states and their applications. 1. Electronegativities of elements in valence states. Inorg. Chem. 1982, 21, 3886–3889. [Google Scholar] [CrossRef]
  49. Liu, J.; Wang, H.; Wang, L.; Jian, P.; Yan, X. Phase-dependent catalytic performance of MnO2 for solvent-free oxidation of ethybenzene with molecular oxygen. Appl. Catal. B 2022, 305, 121050. [Google Scholar] [CrossRef]
  50. Wang, J.; Cheng, L.; An, W.; Xu, J.; Men, Y. Boosting soot combustion efficiencies over CuO-CeO2 catalysts with a 3DOM structure. Catal. Sci. Technol. 2016, 6, 7342–7350. [Google Scholar] [CrossRef]
  51. Liu, J.; Meng, R.; Li, J.; Jian, P.; Wang, L.; Jian, R. Achieving high-performance for catalytic epoxidation of styrene with uniform magnetically separable CoFe2O4 nanoparticles. Appl. Catal. B 2019, 254, 214–222. [Google Scholar] [CrossRef]
  52. He, T.; Shao, D.; Zeng, X.; Rong, S. Harvesting the vibration energy of α-MnO2 nanostructures for complete catalytic oxidation of carcinogenic airborne formaldehyde at ambient temperature. Chemosphere 2020, 261, 127778. [Google Scholar] [CrossRef] [PubMed]
  53. Setvín, M.; Aschauer, U.; Scheiber, P.; Li, Y.F.; Hou, W.; Schmid, M.; Selloni, A.; Diebold, U. Reaction of O2 with Subsurface Oxygen Vacancies on TiO2 Anatase (101). Science 2013, 341, 988–991. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Cao, X.; Lu, J.; Zheng, X.; He, D.; Zhu, W.; Zhao, Y.; Zhang, W.; Tian, R.; Luo, Y. Regulation of the reaction pathway to design the high sulfur/coke-tolerant Ce-based catalysts for decomposing sulfur-containing VOCs. Chem. Eng. J. 2022, 429, 132473. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of α-MnO2 (A) and β-MnO2 (B).
Figure 1. XRD patterns of α-MnO2 (A) and β-MnO2 (B).
Nanomaterials 13 00775 g001
Figure 2. SEM, TEM and HRTEM images of α-MnO2 (AD) and β-MnO2 (FI); The inverse FFT images for α-MnO2 (E) and β-MnO2 (J).
Figure 2. SEM, TEM and HRTEM images of α-MnO2 (AD) and β-MnO2 (FI); The inverse FFT images for α-MnO2 (E) and β-MnO2 (J).
Nanomaterials 13 00775 g002
Figure 3. EPR profiles of α-MnO2 and β-MnO2.
Figure 3. EPR profiles of α-MnO2 and β-MnO2.
Nanomaterials 13 00775 g003
Figure 4. N2 adsorption-desorption isotherm plots (A) and pore distributions (B) of α-MnO2 and β-MnO2.
Figure 4. N2 adsorption-desorption isotherm plots (A) and pore distributions (B) of α-MnO2 and β-MnO2.
Nanomaterials 13 00775 g004
Figure 5. CH3SH degradation over α-MnO2 and β-MnO2 catalysts (A). Reaction condition: 0.20 g catalysts, 5000 ppm CH3SH, total flow rate = 30 mL min−1, WHSV = 9000 mL g−1 ·h−1); The reaction rates (×10−2 mol g−1 h−1) at 50 °C for α-MnO2 and β-MnO2 (B).
Figure 5. CH3SH degradation over α-MnO2 and β-MnO2 catalysts (A). Reaction condition: 0.20 g catalysts, 5000 ppm CH3SH, total flow rate = 30 mL min−1, WHSV = 9000 mL g−1 ·h−1); The reaction rates (×10−2 mol g−1 h−1) at 50 °C for α-MnO2 and β-MnO2 (B).
Nanomaterials 13 00775 g005
Figure 6. The value of rnorm (×10−4, mol·m−2 h−1) from 30 to 80 °C for α-MnO2 and β-MnO2 (A); The value of TOF (h−1) at 50 °C for α-MnO2 and β-MnO2 (B). Reaction conditions: 0.01 g catalysts, 5000 ppm styrene, total flow rate = 30 mL min−1, GHSV = 9000 h−1.
Figure 6. The value of rnorm (×10−4, mol·m−2 h−1) from 30 to 80 °C for α-MnO2 and β-MnO2 (A); The value of TOF (h−1) at 50 °C for α-MnO2 and β-MnO2 (B). Reaction conditions: 0.01 g catalysts, 5000 ppm styrene, total flow rate = 30 mL min−1, GHSV = 9000 h−1.
Nanomaterials 13 00775 g006
Figure 7. H2-TPR (A) and O2-TPD (B) profiles α-MnO2 and β-MnO2.
Figure 7. H2-TPR (A) and O2-TPD (B) profiles α-MnO2 and β-MnO2.
Nanomaterials 13 00775 g007
Figure 8. Raman profiles (A) and the Mn-O force constant (B) of α-MnO2 and β-MnO2.
Figure 8. Raman profiles (A) and the Mn-O force constant (B) of α-MnO2 and β-MnO2.
Nanomaterials 13 00775 g008
Figure 9. XPS profiles of Mn 2p3/2 over the fresh (A) and spent (B) catalysts.
Figure 9. XPS profiles of Mn 2p3/2 over the fresh (A) and spent (B) catalysts.
Nanomaterials 13 00775 g009
Figure 10. XPS profiles of Mn 3s over the fresh (A) and spent (B) catalysts.
Figure 10. XPS profiles of Mn 3s over the fresh (A) and spent (B) catalysts.
Nanomaterials 13 00775 g010
Figure 11. XPS profiles of O 1s over the fresh (A) and spent (B) catalysts.
Figure 11. XPS profiles of O 1s over the fresh (A) and spent (B) catalysts.
Nanomaterials 13 00775 g011
Figure 12. The CH3SH (A) and CH3SSCH3 (B) concentration on α-MnO2 and β-MnO2.
Figure 12. The CH3SH (A) and CH3SSCH3 (B) concentration on α-MnO2 and β-MnO2.
Nanomaterials 13 00775 g012
Figure 13. XPS profiles of S 2p over the fresh (A) and spent (B) catalysts.
Figure 13. XPS profiles of S 2p over the fresh (A) and spent (B) catalysts.
Nanomaterials 13 00775 g013
Figure 14. Reaction mechanism of CH3SH decomposition over MnO2 catalysts.
Figure 14. Reaction mechanism of CH3SH decomposition over MnO2 catalysts.
Nanomaterials 13 00775 g014
Table 1. Mn 3s Mn 2p3/2 results before and after the reaction of B-MnO2 and C-MnO2.
Table 1. Mn 3s Mn 2p3/2 results before and after the reaction of B-MnO2 and C-MnO2.
CatalystsMn2+ (%)Mn3+ (%)Mn4+ (%)Mn2+ + Mn3+/
Mn4+ (%)
AOS 1
α-MnO2-Fresh11.3130.3158.380.713.57
β-MnO2-Fresh2.7135.0462.260.63.75
α-MnO2-Spent25.1943.7731.042.222.83
β-MnO2-Spent16.3946.9936.621.993.25
1 AOS = 8.956 − 1.126 × ΔE.
Table 2. O 1s results before and after the reaction of B-MnO2 and C-MnO2.
Table 2. O 1s results before and after the reaction of B-MnO2 and C-MnO2.
CatalystsOadsOH2OOlattOads/Olatt
α-MnO2-Fresh30.2223.5346.250.65
β-MnO2-Fresh20.6814.1065.220.32
α-MnO2-Spent55.2018.3926.412.09
β-MnO2-Spent39.7820.3439.871.00
Table 3. S 2p results before and after the reaction of B-MnO2 and C-MnO2.
Table 3. S 2p results before and after the reaction of B-MnO2 and C-MnO2.
CatalystsS0S4+S6+
α-MnO2-Fresh000
β-MnO2-Fresh000
α-MnO2-Spent12.9660.7326.31
β-MnO2-Spent13.3874.0612,56
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

Su, H.; Liu, J.; Hu, Y.; Ai, T.; Gong, C.; Lu, J.; Luo, Y. Comparative Study of α- and β-MnO2 on Methyl Mercaptan Decomposition: The Role of Oxygen Vacancies. Nanomaterials 2023, 13, 775. https://doi.org/10.3390/nano13040775

AMA Style

Su H, Liu J, Hu Y, Ai T, Gong C, Lu J, Luo Y. Comparative Study of α- and β-MnO2 on Methyl Mercaptan Decomposition: The Role of Oxygen Vacancies. Nanomaterials. 2023; 13(4):775. https://doi.org/10.3390/nano13040775

Chicago/Turabian Style

Su, Hong, Jiangping Liu, Yanan Hu, Tianhao Ai, Chenhao Gong, Jichang Lu, and Yongming Luo. 2023. "Comparative Study of α- and β-MnO2 on Methyl Mercaptan Decomposition: The Role of Oxygen Vacancies" Nanomaterials 13, no. 4: 775. https://doi.org/10.3390/nano13040775

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