Next Article in Journal
The Effect of H2O2 Pretreatment on TiO2-Supported Ruthenium Catalysts for the Gas Phase Catalytic Combustion of Dichloromethane (CH2Cl2)
Previous Article in Journal
Recent Advances in the Synthesis of Chiral Tetrahydroisoquinolines via Asymmetric Reduction
Previous Article in Special Issue
Magnetic Ternary Hybrid Composites as an Efficient Photocatalyst for Degradation of Acid Orange 7 Dye
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Catalytic Ozonation of Formaldehyde with an Oxygen-Vacancy-Rich MnOx/γ-Al2O3 Catalyst at Room Temperature

1
State Key Laboratory of Clean Energy Utilization, Zhejiang University, Hangzhou 310027, China
2
College of Ocean Science and Engineering, Shanghai Maritime University, Shanghai 201306, China
*
Author to whom correspondence should be addressed.
Catalysts 2024, 14(12), 885; https://doi.org/10.3390/catal14120885
Submission received: 4 November 2024 / Revised: 26 November 2024 / Accepted: 1 December 2024 / Published: 3 December 2024
(This article belongs to the Special Issue Novel Catalysts for Environmental Catalysis)

Abstract

:
Formaldehyde (HCHO) is known as one of the important indoor organic pollutants. How to remove and decompose the low concentration of formaldehyde at room temperature is important for indoor environments. Catalytic ozonation is an efficient method to thoroughly remove HCHO at room temperature, with high efficiency and few byproducts. A series of MnOx/γ-Al2O3 catalysts were prepared in this work via the impregnation method and treated with different reagents (acid, alkali, and H2O2) to evaluate their catalytic activity for HCHO removal. The results showed that MnAl-II (acid treatment) performed well in activity tests, reaching a nearly 100% HCHO conversion at an O3/HCHO of 2.0 and attaining a CO2 selectivity of above 95% at an O3/HCHO of 3.0 at 30 °C, with almost no ozone residual existing. The larger specific surface area, abundant oxygen vacancies, and higher number of acid sites contributed to the excellent performance of MnAl-II. Stability and H2O resistance tests of MnAl-II were also conducted. To reveal the intermediate product formation and further investigate the reaction mechanism of HCHO ozonation, in-situ DRIFTS measurement was carried out combined with DFT calculations.

Graphical Abstract

1. Introduction

With the heavy use of building and decorative materials, formaldehyde (HCHO), which commonly exists in paints, adhesives, furniture, electrical appliances, etc., has become a major pollutant of indoor air worldwide [1,2,3]. As a colorless but irritating gas, HCHO does great harm to human health, with quite strong teratogenicity and carcinogenicity. Long-term exposure to even low concentrations of HCHO will cause respiratory problems and permanently damage the lungs [4]. High concentrations of HCHO can cause breathing difficulties and even death [4]. Thus, it is of great significance to develop methods to efficiently remove HCHO and improve indoor air quality.
Many methods have been applied for the elimination of HCHO, such as ventilation, adsorption, botanical purification, photocatalytic oxidation, biological degradation, catalytic oxidation, and catalytic ozonation [5,6]. Among them, the catalytic ozonation method, which can achieve the complete oxidation of VOCs (volatile organic compounds) into harmless CO2 and H2O without secondary pollution at low temperatures, is the most attractive and effective in HCHO treatment. The strong oxidizability of ozone (O3) is due to the free radicals (·O2, ·O2−, ·OH) generated after activation, which can efficiently break C–O, C–C, and C–H bonds for the deep oxidation of VOCs [7]. Catalysts can reduce the activation energy and accelerate the reaction rate, thus the selection of catalyst materials is vital for the reaction process [8]. Noble metal catalysts (Pt, Pd, Rh, Au, etc.) always exhibit a good catalytic performance; however, their rarity and high cost severely impede their large-scale application [9,10,11,12]. Consequently, it is very necessary to find alternatives which have a low price and high activity.
The cost-effective transition metals have attracted extensive attention from researchers, due to their diverse oxidation states, favorable stability, and excellent performance in catalytic reactions [13]. Among them, manganese oxides (MnOx), which have multiple valence states, a diverse structure, and superior redox properties, have been widely explored in pollutant degradation. Typically, Sekine et al. [14] tested the HCHO removal efficiencies of different metal oxides (Pd, Co, Ce, Fe, Mn, etc.), and MnO2 appeared to offer the best catalytic performance suitable for the catalytic oxidation of HCHO, with no release of hazardous byproducts. Zhang et al. [15] synthesized a series of rod-like MnO2 samples with various crystal structures (α-, β-, γ-MnO2) via a hydrothermal process, and α-MnO2 exhibited the highest HCHO conversion and CO2 yield with ozone assistance at room temperature. The increase in O3 can significantly facilitate the mineralization of HCHO, attributed to the reactive oxygen species produced by O3 decomposition.
In addition, the activity of Mn-based catalysts is determined by many factors, such as doped metals, support types, acid-alkali treatment, and the preparation methods used [16]. Different types of support always greatly influence the properties of catalysts, including the surface area, lattice defects, and metal dispersion. Oyama et al. [17] loaded MnOx onto Al2O3 and SiO2 for acetone catalytic ozonation, and the Al2O3-supported catalyst acquired higher activity and turnover frequencies (TOFs) than SiO2. Shao et al. [6] prepared supported Mn-based catalysts over γ-Al2O3, SiO2, and TiO2 for toluene catalytic ozonation. They found that Mn3+ and surface oxygen species played a positive role in catalytic activity and 5% Mn/γ-Al2O3 achieved the highest toluene conversion and ozone decomposition efficiency. M. Ghavami et al. [18] synthesized a MnOx/Al2O3-P catalyst through the polyol method, which showed higher toluene conversion and CO2 selectivity than the impregnated catalyst, due to its smaller manganese particle size, better dispersion, and greater surface area. Yang et al. [19] treated a copper-manganese spinel with acid and alkali for toluene catalytic combustion. They discovered that both acid and alkali treatment could produce a large number of surface oxygen vacancies and appropriate acidity, which greatly improved the catalytic activity and stability of catalysts.
Hence, a series of MnOx/γ-Al2O3 catalysts were prepared and treated with CH3COOH, NH3·H2O, and H2O2 solutions to research the catalytic ozonation of HCHO at room temperature. The HCHO conversion efficiency and CO2 selectivity of the catalysts were evaluated during the activity tests. The properties of the catalysts, such as their crystal structures, textual properties, oxygen species, and acidity, were investigated by various characterization methods. The H2O resistance experiment was also tested to explore the effect of humidity. Moreover, to study the reaction mechanism of HCHO catalytic ozonation, the process of intermediate formation was conducted through in situ DRIFTs analysis. The DFT calculations were carried out as well for further discussion of the catalysts’ properties.

2. Results and Discussion

2.1. Catalytic Ozonation of HCHO

To study the catalytic ozonation performance of the prepared MnAl catalysts, activity tests were conducted and the results are shown in Figure 1a. The ratio of O3/HCHO was set from 0.5 to 3.0, and the theoretical equivalent of O3/HCHO for complete oxidation is 2, as presented in Equation (1). Apparently, the conversion of HCHO improved with the increase of the O3/HCHO ratio, suggesting that more O3 input could promote the process of catalytic reaction. MnAl-II showed the best HCHO catalytic activity, which achieved nearly 100% HCHO conversion at an O3/HCHO ratio of 2.0 and reached above 95% HCHO conversion at an O3/HCHO ratio of 1.5. In general, the O3/HCHO ratio corresponding to 90% HCHO conversion (O90) increased in the following order: MnAl-II (1.3) < MnAl-IV (1.5) < MnAl-I (1.7) < MnAl-III (2.0). Across the whole range of the ozone input, there was almost no O3 residual for all the catalysts. A comparison of the catalyst researched in this work with previously published catalysts for HCHO ozonation is summarized in Table 1.
The incomplete oxidation of VOCs usually generates undesirable intermediates which are environmentally hazardous. To achieve green emission, CO2 and H2O are expected to be the final products at the outlet. As shown in Figure 1b, it was obvious that the CO2 selectivity increased almost linearly with the molar ratio of O3/HCHO. With the optimal catalytic activity, MnAl-II also attained the highest CO2 selectivity. When the ratio of O3/HCHO was 3, MnAl-II obtained above 95% CO2 selectivity, and the order for the CO2 selectivity decreased as follows: MnAl-II > MnAl-I > MnAl-IV > MnAl-III.
HCHO + (2 − x)O3 → (1 − x)CO2 + xCO + (2 − x)O2 + H2O

2.2. Catalyst Characterization

2.2.1. Crystalline Structures

As shown in Figure 2, the XRD patterns presented the crystalline structures of the synthesized catalysts, which had similar characteristic diffraction peaks. The peaks at 37.6°, 45.8°, and 66.8° were ascribed to the (3 1 1), (4 0 0), and (4 4 0) planes of γ-Al2O3 (PDF#29-0063), while the peaks appearing at 37.3° and 42.8° should be ascribed to the (1 0 1) and (1 1 1) planes of α-MnO2 (PDF#24-0735) [6]. Obviously, the intensity of the MnAl-II catalyst was weaker than the other three catalysts, indicating its better metal dispersion and smaller metal particle size.
The SEM images of these MnAl catalysts exhibit different morphologies and microstructures, as depicted in Figure 3. MnAl-I was composed of oval particles, with a relatively larger size, while the other three catalysts exhibited bulk-like structures with a rough surface and an irregular structure. The surface micropore structure of the catalysts might facilitate the formation of defects, promoting the adsorption of VOCs and the further ozonation process [23,24].

2.2.2. Textual Properties

The textural properties of the prepared catalysts were characterized by the N2 adsorption–desorption isotherms and the distribution curves of pore size (Figure S1). Referring to the IUPAC classification, the catalysts exhibited type IV isotherms with an H3 hysteresis loop, revealing irregular mesoporous structures [25]. As tabulated in Table 2, the surface area, pore volume, and pore diameter were summarized by BET and BJH methods. The catalysts had a similar pore size and their BET surface area decreased in the following order: MnAl-II (188.3 m2·g−1) > MnAl-I (183.5 m2·g−1) > MnAl-IV (174.5 m2·g−1) > MnAl-III (169.4 m2·g−1), corresponding to the sequence of CO2 selectivity. According to the reports, a higher BET surface area could provide more effective active sites and promote the diffusion effect of gaseous reactants [26]. With full contact between the VOCs and the catalytic active sites, the deep oxidation was further enhanced, therefore increasing the selectivity of CO2 [27].

2.2.3. Surface Properties

To evaluate the surface properties of the catalysts, the deconvoluted results of XPS spectra are presented in Figure 4, and the contents of the Mn and O species with different oxidation states are tabulated in Table 3. As shown in Figure 4a, the Mn 2p3/2 spectrum consisted of two peaks at ~642 eV and 643 eV, ascribed to the Mn3+ and Mn4+ species, respectively. Clearly, Mn3+ was the main Mn species, and the order of the Mn3+/Mn4+ molar ratio was as follows: MnAl-II (1.65) > MnAl-I (1.64) > MnAl-IV (1.57) > MnAl-III (1.36). In general, Mn3+ usually triggers the Jahn–Teller distortion, and to preserve the electrostatic balance, oxygen vacancies are accordingly generated in the process (Equation (2), Vo refers to a oxygen vacancy). Moreover, as reactive sites, more oxygen vacancies are beneficial for O3 decomposition and HCHO ozonation [28,29].
4Mn4+ + O2− → 4Mn4+ + 2e/Vo + 1/2O2 → 2Mn4+ + 2Mn3+ + Vo + 1/2O2
The O 1s spectra of the prepared catalysts are shown in Figure 4b, and the curves exhibited two characteristic peaks at ~530 and 531 eV, corresponding to lattice oxygen species (Ola) and surface-adsorbed oxygen species (Oad), respectively. The Oad/Ola molar ratio decreased in the following sequence: MnAl-II (6.30) > MnAl-IV (4.65) > MnAl-I (4.18) > MnAl-III (3.67). With a higher mobility, Oad facilitated the formation of oxygen vacancies, providing efficient active sites for oxygen adsorption, ozone decomposition, and HCHO ozonation [30].
Among all the catalysts, MnAl-II possessed the highest proportion of Mn3+ and Oad, generating abundant oxygen vacancies and promoting the catalytic activity. As for MnAl-III, its Mn3+ and Oad peaks slightly shifted to a higher binding energy, indicating its higher oxygen-escaping energy, which was bad for oxygen vacancies formation and oxygen species mobility, fitting with its HCHO conversion, which was the lowest [16]. Overall, the XPS results were roughly consistent with the catalytic efficiency, illustrating that oxygen vacancies may be an important factor for the catalytic ozonation of HCHO.

2.2.4. Temperature-Programmed Studies

The redox ability of the synthesized catalysts was revealed by the H2-TPR profiles, as shown in Figure 5a. The catalysts exhibited typical reduction peaks, which could be assigned to the successive reduction process of MnO2 (IV) → Mn2O3 (III) → Mn3O4 (II and III) → MnO (II) [31]. By contrast, MnAl-II appeared at a lower H2 reduction temperature, at 236 °C, showing an excellent reducibility at a low temperature [7]. As the report goes, a higher H2 uptake amount always contributes to a better redox ability [32]. However, the amount of H2 uptake (Table 4) increased in the following order: MnAl-I (4.03 mmol·gcat−1) < MnAl-II (4.05 mmol·gcat−1) < MnAl-IV (5.15 mmol·gcat−1) < MnAl-III (5.28 mmol·gcat−1), inconsistent with the results of the removal efficiency. It demonstrated that there may be no positive correlation between the H2 uptake amount and catalytic ozonation activity.
To research the properties of the oxygen species, the O2-TPD profiles are shown in Figure 5b. As the desorption temperature goes from low to high, the oxygen species are classified as chemical-adsorbed oxygen (<250 °C), surface lattice oxygen (250~650 °C), and bulk lattice oxygen (>650 °C) [33]. Compared to the other three catalysts, the desorption peak of MnAl-II at 426 °C was relatively lower, indicating its superior surface lattice oxygen mobility. As tabulated in Table 4, the total O2 desorption amount decreased in the following order: MnAl-II (3.75 a.u.·gcat−1) > MnAl-IV (3.68 a.u.·gcat−1) > MnAl-I (3.60 a.u.·gcat−1) > MnAl-III (3.36 a.u.·gcat−1). This is fitting with the results of O 1s. With the highest HCHO conversion efficiency, MnAl-II also possessed better oxygen mobility and a higher oxygen desorption amount, confirming that abundant active oxygen species were crucial for the catalytic ozonation of HCHO [34].
Surface acidity is important for the catalytic ozonation of VOCs, affecting adsorption, degradation, and desorption processes [35]. Herein, the NH3-TPD experiments were conducted and the obtained profiles are shown in Figure 5c. According to differences in the acidity strength, the desorption peaks below 280 °C were assigned to weak acidity, while the peaks at 280~450 °C and 450~600 °C were ascribed to moderate and strong acidity, respectively [6]. All the catalysts showed only weak and strong acid sites, corresponding to Lewis and Brönsted acid sites, respectively. Generally, Lewis acid sites are conducive to the cleavage of C–C bonds and the deep oxidation of VOCs, while Brönsted acid sites favor the adsorption of VOCs and the desorption of products [36,37]. With the highest total acid sites (3.29 mmol·gcat−1), MnAl-II performed well in the activity tests. Furthermore, the amount of MnAl-III in the total acid sites was 3.28 mmol·gcat−1, which was more than the MnAl-I catalyst. It suggested that both acid treatment and alkaline treatment contribute to the formation of acid sites, as report [19] recorded. However, the catalytic efficiency of MnAl-III was the worst; this could be because the activity was determined by many factors, not only the acidity.

2.3. Effect of H2O

As water vapor commonly exists in indoor air, it may influence catalytic performance and activity. Thus, a long-term H2O resistance test was carried out on MnAl-II to study the effect of H2O on its excellent performance in the activity tests. The reaction temperature was 30 °C, and the molar ratio of O3/HCHO was set as 2.5. As depicted in Figure 6, in the stability test, the conversion of HCHO was maintained at ~99% for 7 h, demonstrating the favorable stability of MnAl-II. When RH = 50%, a slight drop in HCHO conversion was caused, falling from ~99% to ~98% after 2 h. When the H2O was stopped, the conversion of HCHO gradually recovered to ~99%. Subsequently, when RH = 100%, the HCHO conversion decreased to ~97% and then recovered to ~98% after stopping H2O. Obviously, the effect of H2O was negligible, and the slight deactivation was fortunately reversible. Moreover, the presence of H2O promoted the selectivity of CO2; this may be because the generated OH active radicals were beneficial to the complete oxidation of HCHO to CO2 [38].

2.4. In Situ DRIFTS Measurement

To study the transformation of intermediate species, in situ DRIFTS experiments were performed to reveal the possible reaction mechanism of HCHO ozonation over the MnAl-II catalyst. Initially, the process of HCHO adsorption at 30 °C (with N2 as a background gas) was investigated, as recorded in Figure 7a. Notably, the peak located at 2896 cm−1 could be ascribed to the C–H stretching of adsorbed HCHO, while the bands at 1431, 1136, and 1097 cm−1 corresponded to dioxymethylene (DOM) species [39,40]. The absorbance bands centered at 1266, 1360, 1575, and 1593 cm−1 were related to the symmetric and asymmetric stretching vibration of formate species [41]. Additionally, the peaks at 1663 and 3703 cm−1, which could be designated to adsorbed water and hydroxyl groups on the catalyst surface, decreased to negative peaks, possibly due to the consumption of OH species during the adsorption process [42,43].
When the adsorption of CH2O was saturated, O3 was introduced to the reaction gas (Figure 7b). Meanwhile, the absorption peak of monodentate formate species (1593 cm−1) was converted to bidentate formate species (1575 cm−1) and declined with time. Moreover, the peaks corresponding to DOM species (1431, 1136, 1097 cm−1) and HCHO (2896 cm−1) disappeared, while a peak related to bicarbonate species (1421 cm−1) emerged [44]. A negative peak at 1663 cm−1 was not observed, indicating the generation of H2O during the mineralization of formate species. Additionally, the peaks at 2876 and 2396 cm−1 could be assigned to the hydrocarbon vibrations of formate species and carbon oxides, respectively [41].

2.5. Theoretical Calculation and Reaction Mechanism

Vital to the O3 decomposition and ozonation process, the effect of oxygen vacancies on the catalytic reaction was further researched through theoretical calculations. As the main reaction sites for O3, the calculations were carried out on a α-MnO2 (1 0 0) surface. As depicted in Figure 8, the adsorption energy of an O3 molecule on α-MnO2 with an oxygen vacancy (Eads = −1.59 eV) was lower than the original one (Eads = −1.38 eV). Generally, a lower adsorption energy illustrated a stronger interaction between O3 and the α-MnO2 surface, meaning that oxygen vacancies facilitated the adsorption of O3 [45]. Next, with the elongation of the O–O bond, O3 was decomposed into *O and O2. For α-MnO2 with an oxygen vacancy, the energy change for O3 decomposition was −2.20 eV, proving to be an exothermic reaction, which was thermodynamically favorable and spontaneously occurred. As for the original α-MnO2, the required energy change to overcome the energy barrier was 0.50 eV, limiting the reaction rate. Therefore, the generation of oxygen vacancies greatly enhanced the adsorption and decomposition of O3, accelerating the speed of catalytic ozonation.
To summarize, a possible reaction mechanism of HCHO catalytic ozonation was proposed, as presented in Figure 9. Firstly, HCHO was adsorbed on OH species through hydrogen bonds, while O3 was adsorbed on the α-MnO2 surface. With the promotion of oxygen vacancies, O3 was activated and broken down into oxygen and active oxygen radicals (O*). Next, the adsorbed HCHO was oxidized to DOM species and subsequently transformed into formate species under the action of O* radicals. Then, formate species were deeply oxidized into labile carbonic acid and rapidly resolved into CO2 and H2O, which are harmless to the environment. Simultaneously, the active sites and free radicals (O*, OH, etc.) were regenerated. Overall, the reaction pathway followed the decomposition route as follows: HCHO→DOM→formate→carbonic acid→CO2 and H2O.

3. Materials and Methods

3.1. Catalyst Preparation

The Mn-based catalysts were prepared by the equivalent-volume impregnation method, supported by activated γ-Al2O3. Typically, a 1.1 mL Mn(CH3COO)2·4H2O (Aladdin, Shanghai, China) solution with a concentration of 139.8 g/L was evenly impregnated onto 1 g γ-Al2O3 (Aladdin, China, 40~60 meshes). After being treated by an ultrasonic wave for 1 h and placed for 23 h, the mixtures were dried at 100 °C for 8 h and calcined at 500 °C for 3 h in air. Then a 5%Mn/γ-Al2O3 catalyst was obtained, sieved to 40–60 meshes again, and labeled as MnAl-I. Analogously, MnAl-II, MnAl-III, and MnAl-IV were synthesized by the same method. The difference was that the γ-Al2O3 samples were pretreated by 0.5 mol/L CH3COOH (Sinopharm, Shanghai, China), NH3·H2O (Sinopharm, China), and H2O2 (30 wt%, Sinopharm, China) solutions, respectively. The Mn content of the above MnAl catalysts were measured by ICP-OES, and the results were similar to the theoretical mass fraction (~5%).

3.2. Catalytic Activity

The experimental setup schematic of formaldehyde (HCHO) catalytic ozonation is depicted in Figure S2, containing a gas mixing part, a catalytic reaction part, and measuring instruments. The simulated flue gas, containing HCHO (60 ppm/balance N2), N2 (99.999%) and O2 (99.999%), was fixed at 200 mL/min and supplied by cylinder gases (Jingong Gas Co., Ltd., Hangzhou, China. Each gas was regulated by an individual mass flow controller (MFC S500, HIRIBA METRON Co., Ltd., Irvine, CA, USA) and the initial concentration of HCHO and O2 were 30 ppm and 10 vol%, respectively. Ozone was generated by a dielectric barrier discharge reactor (VMUS-1S, AZCO Industries., Ltd., Vancouver, BC, Canada) and the produced O3/O2 mixtures were divided into two streams. One was injected into the mixing tank and the other flowed into a wide-range ozone analyzer (BMT-964BT, OSTI, Inc., Monterey, CA, USA, 0~20 g/Nm3, ±0.1 g/Nm3) to control the concentration of ozone. The water vapor was carried by N2 through the bubbling method. About 0.15 g of the catalyst was mixed with the appropriate amount of SiO2 (Sinopharm, 40~60 mesh), corresponding to a gas hour space velocity (GHSV) of 50,000 h−1, and then placed in the middle of a quartz tube, with a thermal couple to monitor the reaction temperature in the furnace. After the catalytic reaction, the exhaust gas flowed into a FTIR gas analyzer (Gasmet FTIR DX4000, Vantaa, Finland) and a gas chromatograph (GC9790II, Zhejiang Fuli Co., Ltd., Taizhou, China) to measure the concentration of HCHO and CO/CO2, while the residual O3 was detected by a low-range ozone analyzer (BMT-932-1, OSTI Inc.; 0~100 ppm, ±0.1 ppm).
The catalytic activity was evaluated by the conversion efficiencies of HCHO and O3 and the selectivity of CO and CO2, which were calculated using the following equations
[ H C H O ] c o n v . = [ C O ] o u t l e t + [ C O 2 ] o u t l e t [ H C H O ] i n i t i a l × 100 %
[ O 3 ] c o n v . = [ O 3 ] i n i t i a l [ O 3 ] o u t l e t [ O 3 ] i n i t i a l × 100 %
[ C O 2 ] s e l e c . = [ C O 2 ] o u t l e t [ C O ] o u t l e t + [ C O 2 ] o u t l e t × 100 %
where [HCHO]initial and [O3]initial are the initial concentrations of HCHO and O3 in ppm, respectively. [CO]outlet, [CO2]outlet, and [O3]outlet are the outlet concentrations of CO, CO2, and O3 in ppm, respectively.

3.3. Catalyst Characterization

The elemental contents of the catalysts was determined by inductively coupled plasma optical emission spectroscopy (ICP-OES, Agilent 5800, Santa Clara, CA, USA). X-ray powder diffraction (XRD) patterns were obtained by an X-ray diffractometer (PANalytical B.V., Almelo, The Netherlands) using Cu Kα radiation (λ = 0.15406 nm) at 40 kV and 40 mA, from 10 to 80° (2θ, diffraction angle). The surface morphology of the samples was observed by a field emission scanning electron microscope (FE-SEM, Zeiss Sigma 300, Jena, Germany). The surface area, pore volume, and pore diameter were measured by N2 adsorption–desorption isotherms and the BJH method using an automatic specific surface area and porosity analyzer (Micromeritics ASAP 2460, Norcross, GA, USA) at 77 K. The valence states of surface elements on the catalysts were carried out by an X-ray photoelectron spectrometer (XPS, Thermo Scientific K-Alpha, Waltham, MA, USA), and the binding energies of all elements were referenced to the C 1s peak at 284.8 eV. The O2 temperature-programmed desorption (O2-TPD), NH3 temperature-programmed desorption (NH3-TPD), and H2 temperature-programmed reduction (H2-TPR) measurements were conducted on a chemical adsorption analyzer (AutoChem II 2920, Micromeritics, Norcross, GA, USA). The reaction mechanism of the catalytic ozonation process was studied using an in situ diffuse reflectance infrared Fourier transform spectrometer (in situ DRIFTS, Thermo Scientific Nicolet iS50, Waltham, MA, USA) equipped with an MCT/A detector and a reaction chamber (Harrick, Pleasantville, NY, USA). All the in situ spectra were recorded at a resolution of 4 cm−1 and within a range of 600~4000 cm−1.

3.4. DFT Calculation

All the calculations were implemented in the CASTEP package using Material Studio software v2019 on the basis of density functional theory (DFT) [46]. The exchange–correlation interaction was processed with the generalized gradient approximation (GGA) using the Perdew–Burke–Enzerhof (PBE) function. Based on plane-wave theory, a cutoff energy of 650 eV was adopted. To further improve the accuracy, a DFT+U correction was applied to Mn 3d and the optimized Hubbard U = 1.6 eV was used in this work [47,48]. The convergence criteria of the self-consistent field (SCF) and absolute energy were set as 1.0 × 10−6 eV/atom and 1.0 × 10−5 eV/atom, respectively. A 3 × 3 × 9 Brillouin zone of the unit cell was employed.
After geometry optimization, the lattice parameters of bulk α-MnO2 (a = b = 9.81 Å and c = 2.89 Å) were close to the experimental data (a = b = 9.75 Å and c = 2.86 Å), validating the accuracy of the calculation method used in this work. For its stability and reactivity, the (1 0 0) surface of α-MnO2 was cleaved and a 2 × 3 supercell was built, as shown in Figure S3 [49]. A vacuum layer of 16 Å was constructed along the z-axis to avoid the next periodic slab effects. The adsorption energy (Eads) of adsorbates on the surface of the catalyst were calculated using the following equation
Eads = E(adsorbate-catalyst) − (Ecatalyst + Eadsorbate)
where E(adsorbate-catalyst) is the total energy of the catalyst–adsorbate system, Ecatalyst is the total energy of the catalyst, and Eadsorbate is the total energy of the gas–phase molecule.

4. Conclusions

Herein, a series of MnOx/γ-Al2O3 catalysts were synthesized by the equivalent-volume impregnation method to study the catalytic ozonation of HCHO at room temperature. The MnAl-II (treated by acid solution) exhibited the highest catalytic activity, reaching an HCHO conversion above 95% at an O3/HCHO ratio of 1.5, and achieved nearly 100% HCHO conversion with above 95% CO2 selectivity at an O3/HCHO ratio of 3.0 at 30 °C. The better metal dispersion, smaller particle size, and surface micropore structure contributed to the larger surface area of MnAl-II, while the correspondingly generated active sites promoted the deep oxidation of HCHO. Furthermore, its higher Mn3+ and Oad content facilitated the formation of oxygen vacancies, accelerating the process of O3 decomposition and HCHO ozonation. Moreover, the XPS results were roughly consistent with the order of catalytic efficiency, suggesting that oxygen vacancies may be an important factor for catalytic ozonation progress. Also, the plentiful acid sites and active oxygen species were crucial to the excellent performance of MnAl-II. In the stability and H2O resistance tests, MnAl-II also performed well, maintaining over 95% HCHO conversion when RH = 100%. A possible reaction mechanism of HCHO ozonation was proposed based on in-situ DRIFTS measurement (HCHO→DOM→formate→carbonic acid→CO2 and H2O), while the important role of oxygen vacancies in ozone adsorption and decomposition was confirmed by DFT calculations.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/catal14120885/s1. Figure S1: N2 adsorption–desorption isotherms and distribution curves of the pore size of the synthesized catalysts. Figure S2: Schematic of the experimental set-up. Figure S3: The 2 × 3 supercell schematic structure model of α-MnO2. (a) Front view and (b) side view.

Author Contributions

Y.S.: investigation, conceptualization, methodology, validation, writing—original draft. Y.Z. (Yiwei Zhang): writing—review and editing, conceptualization. B.H.: investigation, data curation, visualization. Y.H.: project administration, resources, methodology. W.W.: conceptualization, supervision, project administration. Y.Z. (Yanqun Zhu): writing—review and editing. Z.W.: writing—review and editing, methodology, supervision, funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the “Pioneer” and “Leading Goose” R&D Program of Zhejiang (2023C03126) and the National Natural Science Foundation of China (52125605).

Data Availability Statement

Data will be made available on request.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Liang, W.; Yang, S.; Yang, X. Long-term formaldehyde emissions from medium-density fiberboard in a full-scale experimental room: Emission characteristics and the effects of temperature and humidity. Environ. Sci. Technol. 2015, 49, 10349–10356. [Google Scholar] [CrossRef] [PubMed]
  2. Jiang, C.; Li, D.; Zhang, P.; Li, J.; Wang, J.; Yu, J. Formaldehyde and volatile organic compound (VOC) emissions from particleboard: Identification of odorous compounds and effects of heat treatment. Build. Environ. 2017, 117, 118–126. [Google Scholar] [CrossRef]
  3. Salthammer, T. Formaldehyde sources, formaldehyde concentrations and air exchange rates in European housings. Build. Environ. 2019, 150, 219–232. [Google Scholar] [CrossRef]
  4. Ji, J.; Lu, X.; Chen, C.; He, M.; Huang, H. Potassium-modulated δ-MnO2 as robust catalysts for formaldehyde oxidation at room temperature. Appl. Catal. B Environ. 2020, 260, 118210. [Google Scholar] [CrossRef]
  5. Robert, B.; Nallathambi, G. Indoor formaldehyde removal by catalytic oxidation, adsorption and nanofibrous membranes: A review. Environ. Chem. Lett. 2021, 19, 2551–2579. [Google Scholar] [CrossRef]
  6. Shao, J.; Lin, F.; Wang, Z.; Liu, P.; Tang, H.; He, Y.; Cen, K. Low temperature catalytic ozonation of toluene in flue gas over Mn-based catalysts: Effect of support property and SO2/water vapor addition. Appl. Catal. B Environ. 2020, 266, 118662. [Google Scholar] [CrossRef]
  7. He, C.; Wang, Y.; Li, Z.; Huang, Y.; Liao, Y.; Xia, D.; Lee, S. Facet engineered α-MnO2 for efficient catalytic ozonation of odor CH3SH: Oxygen vacancy-induced active centers and catalytic mechanism. Environ. Sci. Technol. 2020, 54, 12771–12783. [Google Scholar] [CrossRef]
  8. Ge, Y.; Fu, K.; Zhao, Q.; Ji, N.; Song, C.; Ma, D.; Liu, Q. Performance study of modified Pt catalysts for the complete oxidation of acetone. Chem. Eng. Sci. 2019, 206, 499–506. [Google Scholar] [CrossRef]
  9. Zhang, C.; He, H.; Tanaka, K.I. Perfect catalytic oxidation of formaldehyde over a Pt/TiO2 catalyst at room temperature. Catal. Commun. 2005, 6, 211–214. [Google Scholar] [CrossRef]
  10. Zhang, C.; He, H.; Tanaka, K.I. Catalytic performance and mechanism of a Pt/TiO2 catalyst for the oxidation of formaldehyde at room temperature. Appl. Catal. B Environ. 2006, 65, 37–43. [Google Scholar] [CrossRef]
  11. Huang, H.; Leung, D.Y. Complete oxidation of formaldehyde at room temperature using TiO2 supported metallic Pd nanoparticles. ACS Catal. 2011, 1, 348–354. [Google Scholar] [CrossRef]
  12. Yu, X.; He, J.; Wang, D.; Hu, Y.; Tian, H.; He, Z. Facile controlled synthesis of Pt/MnO2 nanostructured catalysts and their catalytic performance for oxidative decomposition of formaldehyde. J. Phys. Chem. C 2012, 116, 851–860. [Google Scholar] [CrossRef]
  13. 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]
  14. Sekine, Y. Oxidative decomposition of formaldehyde by metal oxides at room temperature. Atmos. Environ. 2002, 36, 5543–5547. [Google Scholar] [CrossRef]
  15. Zhang, Y.; Shi, J.; Fang, W.; Chen, M.; Zhang, Z.; Jiang, Z.; Shangguan, W.; Einaga, H. Simultaneous catalytic elimination of formaldehyde and ozone over one-dimensional rod-like manganese dioxide at ambient temperature. J. Chem. Technol. Biotechnol. 2019, 94, 2305–2317. [Google Scholar] [CrossRef]
  16. Xiang, L.; Lin, F.; Cai, B.; Li, G.; Zhang, L.; Wang, Z.; Yan, B.; Wang, Y.; Chen, G. Catalytic ozonation of CH2Cl2 over hollow urchin-like MnO2 with regulation of active oxygen by catalyst modification and ozone promotion. J. Hazard. Mater. 2022, 436, 129217. [Google Scholar] [CrossRef]
  17. Xi, Y.; Reed, C.; Lee, Y.K.; Oyama, S.T. Acetone oxidation using ozone on manganese oxide catalysts. J. Phys. Chem. B 2005, 109, 17587–17596. [Google Scholar] [CrossRef]
  18. Ghavami, M.; Soltan, J.; Chen, N.J.C.L. Synthesis of MnOx/Al2O3 catalyst by Polyol method and its application in room temperature ozonation of toluene in air. Catal. Lett. 2021, 151, 1418–1432. [Google Scholar] [CrossRef]
  19. Yang, Y.; Si, W.; Peng, Y.; Chen, J.; Wang, Y.; Chen, D.; Tian, Z.; Wang, J.; Li, J. Oxygen vacancy engineering on copper-manganese spinel surface for enhancing toluene catalytic combustion: A comparative study of acid treatment and alkali treatment. Appl. Catal. B Environ. 2024, 340, 123142. [Google Scholar] [CrossRef]
  20. Zhao, D.Z.; Shi, C.; Li, X.S.; Zhu, A.M.; Jang, B.W.L. Enhanced effect of water vapor on complete oxidation of formaldehyde in air with ozone over MnOx catalysts at room temperature. J. Hazard. Mater. 2012, 239, 362–369. [Google Scholar] [CrossRef]
  21. Wang, H.; Huang, Z.; Jiang, Z.; Jiang, Z.; Zhang, Y.; Zhang, Z.; Shangguan, W. Trifunctional C@ MnO catalyst for enhanced stable simultaneously catalytic removal of formaldehyde and ozone. ACS Catal. 2018, 8, 3164–3180. [Google Scholar] [CrossRef]
  22. Li, J.W.; Pan, K.L.; Yu, S.J.; Yan, S.Y.; Chang, M.B. Removal of formaldehyde over MnxCe1−xO2 catalysts: Thermal catalytic oxidation versus ozone catalytic oxidation. J. Environ. Sci. 2014, 26, 2546–2553. [Google Scholar] [CrossRef] [PubMed]
  23. Wang, Y.; Wang, G.; Deng, W.; Han, J.; Qin, L.; Zhao, B.; Guo, L.; Xing, F. Study on the structure-activity relationship of Fe-Mn oxide catalysts for chlorobenzene catalytic combustion. Chem. Eng. J. 2020, 395, 125172. [Google Scholar] [CrossRef]
  24. Zhang, Z.; Xiang, L.; Lin, F.; Wang, Z.; Yan, B.; Chen, G. Catalytic deep degradation of Cl-VOCs with the assistance of ozone at low temperature over MnO2 catalysts. Chem. Eng. J. 2021, 426, 130814. [Google Scholar] [CrossRef]
  25. Esmaeilirad, M.; Zabihi, M.; Shayegan, J.; Khorasheh, F. Oxidation of toluene in humid air by metal oxides supported on γ-alumina. J. Hazard. Mater. 2017, 333, 293–307. [Google Scholar] [CrossRef]
  26. Liu, P.; Tang, H.; Shao, J.; He, Y.; Zhu, Y.; Alegria, E.C.; Wang, Z.; Pombeiro, A.J. Catalytic ozonation of multi-VOCs mixtures over Cr-based bimetallic oxides catalysts from simulated flue gas: Effects of NO/SO2/H2O. Chemosphere 2023, 340, 139851. [Google Scholar] [CrossRef]
  27. Liu, P.; Chen, L.; Tang, H.; Shao, J.; Lin, F.; He, Y.; Zhu, Y.; Wang, Z. Low temperature ozonation of acetone by transition metals derived catalysts: Activity and sulfur/water resistance. Catalysts 2022, 12, 1090. [Google Scholar] [CrossRef]
  28. Jia, J.; Zhang, P.; Chen, L. Catalytic decomposition of gaseous ozone over manganese dioxides with different crystal structures. Appl. Catal. B Environ. 2016, 189, 210–218. [Google Scholar] [CrossRef]
  29. Liu, Y.; Yang, W.; Zhang, P.; Zhang, J. Nitric acid-treated birnessite-type MnO2: An efficient and hydrophobic material for humid ozone decomposition. Appl. Surf. Sci. 2018, 442, 640–649. [Google Scholar] [CrossRef]
  30. Zhang, Z.; Lin, F.; Xiang, L.; Yu, H.; Wang, Z.; Yan, B.; Chen, G. Synergistic effect for simultaneously catalytic ozonation of chlorobenzene and NO over MnCoOx catalysts: Byproducts formation under practical conditions. Chem. Eng. J. 2022, 427, 130929. [Google Scholar] [CrossRef]
  31. Shao, J.; Lin, F.; Huang, Y.; Wang, Z.; Li, Y.; Chen, G.; Cen, K. MnOx fabrication with rational design of morphology for enhanced activity in NO oxidation and SO2 resistance. Appl. Surf. Sci. 2020, 503, 144064. [Google Scholar] [CrossRef]
  32. Yang, J.; Huang, Y.; Chen, Y.W.; Xia, D.; Mou, C.Y.; Hu, L.; Zeng, J.; He, C.; Wong, P.K.; Zhu, H.Y. Active site-directed tandem catalysis on CuO/VO-MnO2 for efficient and stable catalytic ozonation of S-VOCs under mild condition. Nano Today 2020, 35, 100944. [Google Scholar] [CrossRef]
  33. Dong, L.; Tang, Y.; Li, B.; Zhou, L.; Gong, F.; He, H.; Sun, B.; Tang, C.; Gao, F.; Dong, L. Influence of molar ratio and calcination temperature on the properties of TixSn1−xO2 supporting copper oxide for CO oxidation. Appl. Catal. B Environ. 2016, 180, 451–462. [Google Scholar] [CrossRef]
  34. Fernandes, A.; Gągol, M.; Makoś, P.; Khan, J.A.; Boczkaj, G. Integrated photocatalytic advanced oxidation system (TiO2/UV/O3/H2O2) for degradation of volatile organic compounds. Sep. Purif. Technol. 2019, 224, 1–14. [Google Scholar] [CrossRef]
  35. Sun, P.; Wang, W.; Dai, X.; Weng, X.; Wu, Z. Mechanism study on catalytic oxidation of chlorobenzene over MnxCe1−xO2/H-ZSM5 catalysts under dry and humid conditions. Appl. Catal. B Environ. 2016, 198, 389–397. [Google Scholar] [CrossRef]
  36. Weng, X.; Sun, P.; Long, Y.; Meng, Q.; Wu, Z. Catalytic oxidation of chlorobenzene over MnxCe1−xO2/HZSM-5 catalysts: A study with practical implications. Environ. Sci. Technol. 2017, 51, 8057–8066. [Google Scholar] [CrossRef]
  37. Tian, M.; Guo, X.; Dong, R.; Guo, Z.; Shi, J.; Yu, Y.; Cheng, M.; Albilali, R.; He, C. Insight into the boosted catalytic performance and chlorine resistance of nanosphere-like meso-macroporous CrOx/MnCo3Ox for 1, 2-dichloroethane destruction. Appl. Catal. B Environ. 2019, 259, 118018. [Google Scholar] [CrossRef]
  38. Tang, H.; Wang, Z.; Shao, J.; Lin, F.; Liu, P.; He, Y.; Zhu, Y. Catalytic decomposition of residual ozone over cactus-like MnO2 nanosphere: Synergistic mechanism and SO2/H2O interference. ACS Omega 2022, 7, 9818–9833. [Google Scholar] [CrossRef]
  39. Sun, S.; Ding, J.; Bao, J.; Gao, C.; Qi, Z.; Li, C. Photocatalytic oxidation of gaseous formaldehyde on TiO2: An in situ DRIFTS study. Catal. Lett. 2010, 137, 239–246. [Google Scholar] [CrossRef]
  40. Chen, B.B.; Zhu, X.B.; Crocker, M.; Wang, Y.; Shi, C. FeOx-supported gold catalysts for catalytic removal of formaldehyde at room temperature. Appl. Catal. B Environ. 2014, 154, 73–81. [Google Scholar] [CrossRef]
  41. Zhu, L.; Wang, J.; Rong, S.; Wang, H.; Zhang, P. Cerium modified birnessite-type MnO2 for gaseous formaldehyde oxidation at low temperature. Appl. Catal. B Environ. 2017, 211, 212–221. [Google Scholar] [CrossRef]
  42. Soria, J.; Sanz, J.; Sobrados, I.; Coronado, J.M.; Hernández-Alonso, M.D.; Fresno, F. Water-hydroxyl interactions on small anatase nanoparticles prepared by the hydrothermal route. J. Phys. Chem. C 2010, 114, 16534–16540. [Google Scholar] [CrossRef]
  43. Hernández-Alonso, M.D.; García-Rodríguez, S.; Suárez, S.; Portela, R.; Sánchez, B.; Coronado, J.M. Operando DRIFTS study of the role of hydroxyls groups in trichloroethylene photo-oxidation over titanate and TiO2 nanostructures. Catal. Today 2013, 206, 32–39. [Google Scholar] [CrossRef]
  44. Yamamoto, M.; Yoshida, T.; Yamamoto, N.; Nomoto, T.; Yamamoto, Y.; Yagi, S.; Yoshida, H. Photocatalytic reduction of CO2 with water promoted by Ag clusters in Ag/Ga2O3 photocatalysts. J. Mater. Chem. A 2015, 3, 16810–16816. [Google Scholar] [CrossRef]
  45. Cheng, Y.; Chen, Z.; Yan, P.; Shen, J.; Kang, J.; Wang, S.; Duan, X. Synergistic catalytic ozonation mediated by dual active sites of oxygen vacancies and defects in biomass-derived composites for long-lasting water decontamination. ACS Catal. 2024, 14, 4040–4052. [Google Scholar] [CrossRef]
  46. Segall, M.D.; Lindan, P.J.; Probert, M.A.; Pickard, C.J.; Hasnip, P.J.; Clark, S.J.; Payne, M.C. First-principles simulation: Ideas, illustrations and the CASTEP code. J. Phys. Condens. Matter 2002, 14, 2717. [Google Scholar] [CrossRef]
  47. Luo, J.; Meng, X.; Crittenden, J.; Qu, J.; Hu, C.; Liu, H.; Peng, P. Arsenic adsorption on α-MnO2 nanofibers and the significance of (1 0 0) facet as compared with (1 1 0). Chem. Eng. J. 2018, 331, 492–500. [Google Scholar] [CrossRef]
  48. Li, G.; Zhao, P.; Zheng, H.; Yang, L.; Lu, S.; Peng, P. Research on the removal mechanism of antimony on α-MnO2 nanorod in aqueous solution: DFT+ U method. J. Hazard. Mater. 2018, 354, 8–16. [Google Scholar] [CrossRef]
  49. Luo, J.; Hu, C.; Meng, X.; Crittenden, J.; Qu, J.; Peng, P. Antimony removal from aqueous solution using novel α-MnO2 nanofibers: Equilibrium, kinetic, and density functional theory studies. ACS Sustain. Chem. Eng. 2017, 5, 2255–2264. [Google Scholar] [CrossRef]
Figure 1. (a) HCHO conversion and (b) CO2 selectivity during the catalytic ozonation of HCHO in the synthesized catalysts. (HCHO initial concentration: 30 ppm, T = 30 °C).
Figure 1. (a) HCHO conversion and (b) CO2 selectivity during the catalytic ozonation of HCHO in the synthesized catalysts. (HCHO initial concentration: 30 ppm, T = 30 °C).
Catalysts 14 00885 g001
Figure 2. XRD patterns of the prepared catalysts.
Figure 2. XRD patterns of the prepared catalysts.
Catalysts 14 00885 g002
Figure 3. SEM images of (a1,a2) MnAl-I, (b1,b2) MnAl-II, (c1,c2) MnAl-III, and (d1,d2) MnAl-IV.
Figure 3. SEM images of (a1,a2) MnAl-I, (b1,b2) MnAl-II, (c1,c2) MnAl-III, and (d1,d2) MnAl-IV.
Catalysts 14 00885 g003
Figure 4. XPS spectra of (a) Mn 2p and (b) O 1s for the synthesized catalysts.
Figure 4. XPS spectra of (a) Mn 2p and (b) O 1s for the synthesized catalysts.
Catalysts 14 00885 g004
Figure 5. H2-TPR, O2-TPD, and NH3-TPD profiles of the synthesized catalysts: (a) H2-TPR; (b) O2-TPD; (c) NH3-TPD.
Figure 5. H2-TPR, O2-TPD, and NH3-TPD profiles of the synthesized catalysts: (a) H2-TPR; (b) O2-TPD; (c) NH3-TPD.
Catalysts 14 00885 g005
Figure 6. Effect of H2O on HCHO conversion efficiency and outlet CO2 concentration in the MnAl-II catalyst.
Figure 6. Effect of H2O on HCHO conversion efficiency and outlet CO2 concentration in the MnAl-II catalyst.
Catalysts 14 00885 g006
Figure 7. In situ DRIFTS measurement for the MnAl-II catalyst at 30 °C with different test conditions: (a) 30 ppm HCHO + N2; (b) 30 ppm HCHO + 90 ppm O3 + N2.
Figure 7. In situ DRIFTS measurement for the MnAl-II catalyst at 30 °C with different test conditions: (a) 30 ppm HCHO + N2; (b) 30 ppm HCHO + 90 ppm O3 + N2.
Catalysts 14 00885 g007
Figure 8. Relative energies of different coordinates and the interfacial reaction pathways of O3 on the α-MnO2 (1 0 0) surface. The purple atoms referred to Mn and the red atoms referred to O.
Figure 8. Relative energies of different coordinates and the interfacial reaction pathways of O3 on the α-MnO2 (1 0 0) surface. The purple atoms referred to Mn and the red atoms referred to O.
Catalysts 14 00885 g008
Figure 9. Proposed mechanism for the catalytic ozonation of HCHO in the MnAl catalysts.
Figure 9. Proposed mechanism for the catalytic ozonation of HCHO in the MnAl catalysts.
Catalysts 14 00885 g009
Table 1. Comparison of catalysts for HCHO catalytic ozonation.
Table 1. Comparison of catalysts for HCHO catalytic ozonation.
CatalystReaction
Mixture
GHSV
(h−1)
Temperature (°C)Conversion (%)Reference
MnOxHCHO: 82 ppm200,00025~100[20]
Ozone: 720 ppm
α-MnO2HCHO: 60 ppm/Ambient
temperature
80[15]
Ozone: 230 ppm
MnO@CHCHO: 60 ppm/30100[21]
Ozone: 180 ppm
Mn0.5Ce0.5O2HCHO: 61 ppm10,00025~83[22]
Ozone: 203 ppm
MnAl-IIHCHO: 30 ppm50,00030100This work
Ozone: 60 ppm
Table 2. Textual properties of the prepared catalysts.
Table 2. Textual properties of the prepared catalysts.
CatalystBET Surface Area
/m2·g−1
Pore Volume a
/cm3·g−1
Average Pore
Diameter b/nm
MnAl-I183.50.356.0
MnAl-II188.30.345.9
MnAl-III169.40.325.9
MnAl-IV174.50.335.9
a BJH desorption cumulative volume of pores. b BJH desorption average pore diameter.
Table 3. XPS results for the synthesized catalysts.
Table 3. XPS results for the synthesized catalysts.
CatalystsMn 2p3/2O 1sMn3+/Mn4+Oad/Ola
Mn4+ (%)Mn3+ (%)Ola (%)Oad (%)
MnAl-I37.962.119.380.71.644.18
MnAl-II37.762.313.786.31.656.30
MnAl-III42.357.721.478.61.363.67
MnAl-IV38.961.117.782.31.574.65
Table 4. Amount of O2, NH3 desorption, and H2 uptake in the synthesized catalysts.
Table 4. Amount of O2, NH3 desorption, and H2 uptake in the synthesized catalysts.
CatalystH2 Uptake
/(mmol·gcat−1)
O2 Desorption
/(a.u.·gcat−1)
NH3 Desorption
/(mmol·gcat−1)
MnAl-I4.033.603.26
MnAl-II4.053.753.29
MnAl-III5.283.363.28
MnAl-IV5.153.683.22
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

Sun, Y.; Zhang, Y.; Hou, B.; He, Y.; Weng, W.; Zhu, Y.; Wang, Z. Catalytic Ozonation of Formaldehyde with an Oxygen-Vacancy-Rich MnOx/γ-Al2O3 Catalyst at Room Temperature. Catalysts 2024, 14, 885. https://doi.org/10.3390/catal14120885

AMA Style

Sun Y, Zhang Y, Hou B, He Y, Weng W, Zhu Y, Wang Z. Catalytic Ozonation of Formaldehyde with an Oxygen-Vacancy-Rich MnOx/γ-Al2O3 Catalyst at Room Temperature. Catalysts. 2024; 14(12):885. https://doi.org/10.3390/catal14120885

Chicago/Turabian Style

Sun, Yulin, Yiwei Zhang, Baoqing Hou, Yong He, Wubin Weng, Yanqun Zhu, and Zhihua Wang. 2024. "Catalytic Ozonation of Formaldehyde with an Oxygen-Vacancy-Rich MnOx/γ-Al2O3 Catalyst at Room Temperature" Catalysts 14, no. 12: 885. https://doi.org/10.3390/catal14120885

APA Style

Sun, Y., Zhang, Y., Hou, B., He, Y., Weng, W., Zhu, Y., & Wang, Z. (2024). Catalytic Ozonation of Formaldehyde with an Oxygen-Vacancy-Rich MnOx/γ-Al2O3 Catalyst at Room Temperature. Catalysts, 14(12), 885. https://doi.org/10.3390/catal14120885

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