| Home | E-Submission | Sitemap | Contact Us |  
Environ Eng Res > Volume 30(4); 2025 > Article
Nguyen, Nguyen, Nguyen, Nguyen, Cao, and Van Pham: Combining montmorillonite mineral and SnO2/TiO2 photocatalysts for high-performance and cost-effective nitrogen oxide decomposition

Abstract

Development of cost-effective photocatalysts for environmental pollution treatment is trending in materials science research. In this study, montmorillonite (MMT) was combined with SnO2/TNTs photocatalysts for the efficient NO gas treatment. The MMT enhanced the photocatalytic efficiency for NO gas treatment. Notably, the 10% MMT/SnO2/TNTs photocatalyst exhibited a significant green product yield of 54.3% and achieved 10−4% of quantum efficiency. Moreover, the 10% MMT/SnO2/TNTs photocatalyst has high stability with less change of the NO photocatalytic degradation for five cycles. The 10% MMT/SnO2/TNTs photocatalyst also expressed oxidation and reduction reactions in the NO photocatalytic reactions. The incorporation of MMT has led to a noteworthy reduction in the production cost of the catalyst, making it a cost-effective solution. These findings emphasize the significance of integrating MMT to attain improved photocatalytic performance and cost efficiency in practical applications.

Graphical Abstract

/upload/thumbnails/eer-2024-268f9.gif

Introduction

In recent years, the issue of air pollution, specifically the presence of harmful gases such as nitrogen oxides (NOx), has become a significant concern worldwide [1, 2]. NOx emissions from various sources, including industrial processes and vehicle exhaust, contribute to the deterioration of air quality and pose serious health risks [3]. Therefore, developing effective and sustainable methods for NOx gas treatment is of utmost importance [4, 5].
Photocatalysis has emerged as a promising approach for air purification due to its ability to degrade pollutants under light irradiation [6, 7]. Semiconductor-based photocatalysts, such as titanium dioxide (TiO2), have shown great potential in removing NOx gases [2, 8]. However, the low efficiency and limited visible light response of TiO2 have prompted researchers to explore novel composite materials to enhance their photocatalytic performance [9, 10]. Recently, the SnO2/TiO2 nanotubes (SnO2/TNTs) composite has gained significant attention as a promising alternative to improve the photocatalytic efficiency of TiO2 for air purification applications [7]. The SnO2/TNTs composite has showcased outstanding effectiveness in eliminating NOx gases, which are significant air pollutants. Its heightened responsiveness to visible light and enhanced photocatalytic efficiency render SnO2/TNTs a promising contender for air purification, providing potential solutions for tackling environmental challenges. However, it is important to note that this material carries a higher cost than certain other materials.
Additionally, its limited thermal resistance and surface area, which is 12.4 m2/g [11] compared to other minerals, for instance, MMT’s 48 m2/g [12], can potentially impact the photocatalytic process. On the other hand, MMT, a type of layered clay mineral, exhibits remarkable properties that make it a valuable component in composite materials. One of its key features is its large surface area, which provides abundant active sites for adsorption and catalytic reactions [6, 13]. This characteristic allows MMT to capture and trap pollutants effectively. Moreover, MMT demonstrates high adsorption capacity thanks to its unique layered structure and interlayer spaces. These spaces can accommodate various gases, organic compounds, and heavy metal ions, effectively removing them from the surrounding environment [12]. This property makes MMT an excellent candidate for enhancing composite materials’ adsorption capabilities in air purification systems. Furthermore, the presence of metal ions within MMT significantly influences the photocatalytic activity of composite materials, particularly SnO2/TiO2. These metal ions can act as co-catalysts, promoting the separation of electron-hole pairs and enhancing the overall photocatalytic efficiency [14, 15]. The synergistic effect between MMT and the metal ions contributes to the improved performance of SnO2/TiO2 composites, making them more effective in degrading pollutants under light irradiation.
In this study, we focused on synthesizing the MMT/SnO2/TNTs materials and using various techniques such as XRD, FT-IR, EDX, BET, EPR, etc. to identify the characterization of materials. The BET analysis was conducted using a Gemini VII model 2390 BET analyzer. The adsorptive gas was nitrogen (N2), used at liquid nitrogen temperature (77°K). The measurement time for the 10% MMT/SnO2/TNTs sample was 3 hours and 10 minutes, while for the MMT sample, it was 4 hours and 30 minutes. Furthermore, the optimization of the NO photocatalytic efficiency of the composite materials in photocatalytic applications was also determined. The factors, including cost-effective production and photocatalysis, were prioritized during this research, aiming to synthesize a cost-effective photocatalyst for practical application.

Experimental Section

2.1. Chemicals and Materials

Merck’s 99.99% pure tin (IV) chloride pentahydrate (SnCl4.5H2O), hydrazine hydrate (N2H4.H2O), commercial titanium dioxide powder (TiO2, Merck, 99.99%), and sodium hydroxide (NaOH, Merck, 99%) were used in the experiment. In this study, the MMT used was sourced from the Binh Dinh province, Vietnam.

2.2. Preparation of Materials

The synthesis of SnO2/TNTs heterojunction and the preparation of TNTs and SnO2 NPs was carried out using a one-step hydrothermal method, following the procedures outlined in our previous publications [16]. The composite material was prepared using the ball milling method, combining MMT with SnO2/TNTs. A specified amount of MMT x (g) and 2 g of SnO2/TNTs were used to ensure a consistent ratio for composite formation. During ball milling, the mechanical forces led to the intimate mixing and interfacial contact between the two components. The repeated collisions and grinding actions resulted in the breakdown of particle agglomerates and the formation of a homogeneous mixture at the nanoscale, ensuring thorough dispersion of MMT within the SnO2/TNTs matrix. Composite samples with different compositions (5%, 10%, and 15% MMT content relative to SnO2/TNTs) were obtained by adjusting the amount of MMT added to the initial mixture.

2.3. Characterization of Materials

The characterization of the synthesized material involved the use of various techniques. X-ray diffraction (XRD) was conducted using a D8 ADVANCED instrument from Bruker AXS with Cu radiation to determine the material’s crystal structure. The Fourier Transform Infrared Spectroscopy (FT-IR) was performed using a Jasco FT-IR V-4700 spectrometer to analyze the vibrational modes of chemical bonds. Diffuse Reflectance Spectroscopy (DRS) was used to study the absorption wavelengths of the material within the UV-Vis range, employing a Jasco V-770 UV-VIS spectrophotometer. Electron Spin Resonance (ESR) was also employed to examine other material properties.

2.4. Evaluating the NO Photocatalytic Ability

The photocatalytic capability of the MMT/SnO2/TNTs material in treating NO gas was evaluated using an experimental setup as Scheme 1. The initial NO gas with 100 ppm of concentration was diluted by atmospheric air via a zero-air generator to achieve an inlet NO gas concentration of 500 ppb. The humidity within the reaction chamber was controlled using an airflow control system and a humidification system. The gas mixture was exposed to photoexcitation using a solar spectrum lamp and filter system, allowing the NO gas to react with the catalyst. The concentration of initial NO gas, the concentration change of NO and the formation of NO2 gas were identified by a NOx spectrometer system (Sabio, 6040).
For the preparation of the synthesized material, 0.2 g of the material was evenly spread on a 12 cm diameter petri dish using DI water as the medium. After undergoing a ten-minute ultrasonic bath, the material was dried at 60°C. The NO removal efficacy of the prepared sample was evaluated under visible light within a reaction chamber as in previous study [16]. Besides, the products of the photocatalytic reaction between semiconductors with NO gas were created via the green products of the NO photocatalytic, including nitrate (NO3), nitrite (HNO2), and nitric acid (HNO3), which are easily removable from the photocatalyst surface through washing with water. In contrast, nitrite (NO2) is regarded as an undesirable byproduct due to its toxicity. Therefore, the NO removal efficiency (η, %), NO2 conversion performance (η, %), and the conversion efficiency of NO into green products (ϕ, %) were calculated using Eq. (13), respectively.
(1)
η(%)=CNO0-CNOtCNO0×100
(2)
ψ(%)=CNO2t-CNO20CNO0×100
(3)
φ(%)=η-ψ
To compare the efficiency of using light for photocatalytic reactions, the quantum efficiency is calculated using Eq. (4)
(4)
ϕapp=NA0t(Ci-Cf)VtPhoton Plux×Irradiation Area×t×1000M×100
In Eq. (4), NA (mol−1) represents Avogadro’s constant, Vt (L min−1) is the flow rate of NO, and M (g mol−1) is the molar mass of NO. In the experiment, the photon flux was measured to be 2.72×1019 cm−2 min−1, and the irradiated area was 103.4 cm2 for a 12 cm diameter petri dish.
Furthermore, trapping investigations were performed to comprehend the photocatalytic mechanism and the impact of crucial factors, including the participation of holes, electrons, and redox reactions, on the elimination of NO through photocatalyst processes. These experiments involved using 0.2 g of the synthesized sample with a trapping agent mass equal to 1% of the catalyst mass. Potassium iodide (KI), potassium dichromate (K2Cr2O7), and terephthalic acid were employed as trapping agents to turn off photogenerated electrons (e), holes (h+), and hydroxyl radicals (•OH), respectively. Additionally, recycling experiments were carried out, where the photocatalyst dispersed on the petri dish was removed from the reaction chamber after each cycle. The photocatalytic degradation experiment was then repeated under the same conditions.

Results and Discussion

3.1. Material Characterizations

X-ray diffraction (XRD) analysis was conducted to examine the crystalline phase states and identify the characteristic bonds present in the SnO2/TiO2 and MMT/SnO2/TiO2 materials under investigation. The XRD analysis of the MMT/SnO2/TNTs composite material in Fig. 1a revealed the presence of three main phases: SiO2, SnO2, and TiO2. The TiO2 phase in the MMT/SnO2/TNTs sample exhibited peaks at 2θ values of 25.20°, 37.83°, 40.00°, 53.96°, 62.01°, and 71.03°, corresponding to the (101), (004), (200), (105), (211), and (220) crystallographic planes of TiO2. Additionally, the SnO2 phase displayed peaks at 2θ values of 21.48°, 25.01°, 30.60°, 36.18°, 41.76°, and 47.35°, corresponding to the (110), (101), (200), (211), (220), and (311) planes. The MMT clay component was characterized by SiO2 peaks at 2θ values of 22.02°, 26.64°, 31.48°, 36.67°, 42.25°, and 46.8°, corresponding to the (100), (011), (110), (102), (112), and (200) planes [17].
The FTIR spectrum was conducted to facilitate the observation of bond vibrations and the formation of SnO2/TNTs and MMT/SnO2/TNTs composite materials. The FTIR spectrum allows for the analysis of specific chemical bonds and the presence of functional groups in the region spanning wavenumbers from 4000 to 400 cm−1. The FTIR analysis in Fig. 1b confirmed the presence of specific bands in the MMT sample, such as the O-H stretching vibration of water at 3440 cm−1, the Si-O stretching vibration at 1630 cm−1, and the Si-O-Si bending vibration at 785 cm−1, indicating the presence of silica. In the FTIR spectra of the 5% MMT, 10% MMT, and 15% MMT samples, similar bands to those observed in the SiO2 sample were present, confirming the presence of silica. Additionally, evidence of SnO2 and TiO2 in the composite was provided by the Sn-O stretching vibration at 1050 cm−1 and the Ti-O stretching vibration at 880 cm−1. The 5% MMT sample exhibited a slightly weaker intensity of the SnO2 and TiO2 bands compared to the 10% MMT sample, indicating a lower concentration of SnO2 and TiO2.
X-ray photoelectron spectroscopy (XPS) was utilized to gain valuable insights into the chemical within the 10%MMT/SnO2/TNTs. Fig. 2a shows the XPS of MMT/SnO2/TNTs, including binding energy peaks of Si, Sn, Ti and O. The carbon is associated with unexpectedly adsorbed hydrocarbons. The high-resolution XPS spectrum of Sn 3d (Fig. 2b) displays two binding energy (BE) peaks at 494.1 and 484 eV, Sn is present in Sn (IV) state. Fig. 2c shows the high resolution XPS for Ti with two main peaks. In detail, the BEs of two peaks located at 456.55 eV for 2p3/2 and a weak peak at 462.3 eV for 2p1/2, corresponding to the Ti4+ oxidation state of TiO2. The deconvoluted HRXPS spectra for O 1s indicate two contributions for oxygen species as shown in Fig. 2d. One of which corresponds to Sn-O-Sn and the other Ti-O-Ti. These XPS evidence demonstrated the formation of SnO2/TNTs.
Fig. 3 displays the MMT clay containing Si4+, Na+, and K+ cations, which enhances the mechanical properties of the composite material [18], including tensile strength, compressive strength, hardness, and flexibility. Tin (Sn) is distributed evenly throughout the sample, while titanium (Ti) is concentrated in specific regions. Silicon (Si) is relatively evenly distributed, and sodium (Na) and potassium (K) are uniformly distributed.
The composition and content of the elements in the sample are as follows: Tin (Sn) constitutes 48.7%, titanium (Ti) constitutes 13.4%, silicon (Si) constitutes 13.4%, sodium (Na) constitutes 5.3%, potassium (K) constitutes 0.5%, and oxygen (O) constitutes 27.1%. The presence of these elements in MMT can enhance its electrical conductivity, corrosion resistance, mechanical strength, and catalytic efficiency [19]. In addition, the BET surface area of SnO2/TNTs and 10% MMT/SnO2/TNTs was determined.
In addition, the BET surface area of SnO2/TNTs and 10% MMT/SnO2/TNTs was determined. In detail, Fig. 4 and Table S1 compared the parameters for the two samples, MMT and 10% MMT/SnO2/TNTs. In detail, the presence of SnO2/TNTs significantly affected the surface area and pore size of the material. SnO2/TNTs not only penetrate the pores of MMT but also distribute on its surface, leading to a reduction in the sample’s surface area.
This is evident from the BET surface area of the 10% MMT/SnO2/TNTs sample, which is only 4.5948 m2/g, considerably lower than the original MMT sample’s 77.7942 m2/g. Additionally, the presence of SnO2/TNTs within and on the pore surfaces also reduces the average pore size from 18.3027 nm (adsorption) and 13.5354 nm (desorption) in MMT to 14.6424 nm and 8.3113 nm in the 10% MMT/SnO2/TNTs sample. Furthermore, the pore volume of the 10% MMT/SnO2/TNTs sample is also reduced compared to MMT, indicating that SnO2/TNTs have occupied part of the original pore volume of MMT.
The optical properties of semiconductor materials are crucial for their applications in various fields as they govern the interactions between light and matter. To gain a deeper understanding of these properties, the synthesized materials were analyzed using the diffuse reflectance spectroscopy (DRS) technique, as shown in Fig. 5a. The DRS analysis provided significant findings, revealing that the SnO2/TNTs material exhibited an absorption wavelength at 395 nm [17]. Interestingly, the incorporation of MMT into the SnO2/TNTs composite greatly enhanced its light absorption capacity. Among the samples synthesized with different radiation doses of MMT, the sample with a 10% MMT dose displayed the highest light absorption capability, particularly at a wavelength of 410 nm. This observation suggests that the addition of MMT positively influences the material’s ability to absorb light. Furthermore, an intriguing relationship was observed between the radiation dose of MMT and the mass percentage of the MMT integrated into the SnO2/TNTs material. As the radiation dose of MMT increased, the mass percentage of the MMT incorporated into the composite structure also increased. This increase in MMT content corresponded to a significant reduction in the material’s absorption wavelength. Increasing the dosage of the MMT resulted in a greater concentration of the MMT within the composite, which caused a decrease in the absorption wavelength.
Additionally, it was noted that the bandgap energy of the material decreased as the mass percentage of MMT increased in Fig. 5b. Specifically, the bandgap energy decreased from 3.2 eV for the SnO2/TNTs material to 3.12 eV for the sample with a 10% MMT dose [17]. This indicates that the incorporation of MMT not only improves light absorption but also affects the fundamental electronic properties of the material, resulting in a narrower bandgap. The application of DRS provided valuable insights into the optical properties of the synthesized materials. The inclusion of MMT in the SnO2/TNTs composite significantly enhanced its light absorption capacity, with the sample containing a 10% MMT dose exhibiting the highest absorption capability at a specific wavelength. Moreover, increasing the radiation dose of MMT led to a higher MMT content within the composite, resulting in a noteworthy decrease in the material’s absorption wavelength. These findings contribute to a comprehensive understanding of the optical properties of the synthesized materials and hold potential for various applications.

3.2. Photocatalytic Removal of NO Gas

Fig. 6a illustrates the NO gas removal efficiency of the SnO2/TNTs and MMT/SnO2/TNTs materials. It is evident that the photocatalytic efficiency of the material significantly increases during the initial 5 minutes of irradiation [20]. Subsequently, the efficiency stabilizes for the remaining duration of the experiment. Notably, the incorporation of MMT into the composite material significantly enhances its effectiveness. Moreover, the MMT sample shows negligible impact on NO gas decomposition, as depicted in Fig. 6a. Fig. 6b shows the SnO2/TNTs sample achieves a green product transformation efficiency of 34.2%. In contrast, the 10% MMT sample achieves a green product transformation efficiency of 54.3%, with a low NO2 conversion efficiency of 0.52%. Additionally, the 5% and 15% MMT samples also exhibit green product conversion efficiencies of 42.74% and 45.47%, respectively. The MMT acts as a promoter for the photocatalytic effect, acting as a matrix that carries the nano-sized particles of SnO2/TNTs, thanks to its components such as Si4+, Na+, and K+, which enhance the photocatalytic process. Furthermore, it is evident in Fig. 6c that both SnO2/TNTs and MMT-MMT-containing samples demonstrate higher quantum efficiencies compared to the base SnO2/TNTs sample, confirming the enhanced photocatalytic effectiveness of SnO2/TNTs with MMT. Notably, the quantum efficiency peaks at 10.35 (10−4%) for the 10% MMT sample, indicating that this sample exhibits the highest photocatalytic effectiveness for removing NO gas from the atmosphere.
To insight into the photocatalytic mechanism and understand the impact of key factors such as electron-hole pairs and redox reactions on the efficiency of the photocatalytic reaction, a trap experiment was conducted. The experiment aimed to capture and analyze the main factors, namely h+, e, and •OH, and their respective influence on the photocatalytic process. To trap these factors, KI, K2Cr2O7, and terephthalic acid (TPA) were utilized.
Fig. 6d illustrates the photocatalytic removal efficiency of NO in the presence of KI, K2Cr2O7, and terephthalic acid. The results demonstrate that electrons (e) and holes (h+) play a crucial role in driving the photocatalytic reaction, whereas •OH radicals have a lesser contribution. This is evident from the observed decrease in NO removal efficiency when KI, K2Cr2O7, and terephthalic acid are introduced to the reaction. Specifically, the NO removal efficiency decreases by 17.71% with the addition of KI, 5.51% with K2Cr2O7, and 44.76% with terephthalic acid. These findings highlight the significance of electrons and holes in the photocatalytic process, underscoring their primary involvement in the desired reaction. The limited contribution of •OH radicals suggests that their presence has a relatively minor impact on the overall efficiency of NO removal in this particular system. Further investigations could delve into these factors’ underlying mechanisms and interactions of these factors to gain a deeper understanding of the photocatalytic reaction and optimize its performance.
To evaluate the stability and practical applicability of the 10% MMT material, photocatalytic ability was evaluated via recycling tests under visible light conditions. Fig. 7a shows the NO gas photocatalytic removal efficiency of the material under visible light irradiation. It is noteworthy that the material exhibited a remarkable consistency in its performance, as there was no significant decrease in its efficiency after undergoing 5 consecutive cycles. The initial efficiency was measured at 54.82%, and even after the 5th cycle, it remained high at 53.60%.
These findings indicate the promising stability and durability of the 10% MMT material as a photocatalyst for NO gas treatment under visible light. The negligible decrease in its efficiency over multiple cycles suggests its potential for practical application in real-world scenarios where continuous, long-term performance is required. Further research and testing could be conducted to validate and optimize the performance of the 10% MMT material, exploring its suitability for large-scale applications and investigating any underlying mechanisms contributing to its sustained photocatalytic efficiency. Furthermore, the FTIR analysis conducted after undergoing five cycles of reuse reveals a striking resemblance between the spectra, affirming the structural consistency of the photocatalyst despite prolonged exposure to wide-spectrum light in Fig. 7b. As a result, these outcomes provide encouraging evidence regarding the material’s durability when employed in practical settings.
To provide additional evidence of the participation of MMT and SnO2/TNTs nanocomposite in the generation of reactive oxygen species (ROS) during photocatalytic reactions, an EPR technique was employed to monitor the signal of these species under varying conditions. The results obtained are presented in Fig. S1a, b showcasing the outcomes obtained from the EPR measurements. In the dark conditions, the results of the investigation revealed no detectable signals of reactive species such as •OH and •O2 radicals. This indicates that, without illumination, the generation of these radicals is negligible. However, when the MMT and SnO2/TNTs nanocomposite was exposed to light in the visible range, distinct and strong signals from reactive species were detected. This observation unequivocally demonstrates that the heterojunction formed by the nanocomposite is effectively activated by the incident light, leading to the formation of reactive species. The capability of the 10% MMT/SnO2/TNTs nanocomposite to facilitate both oxidation and reduction reactions is noteworthy. This property enables spatial distributions of oxidation, indicating that the composite material possesses the ability to promote simultaneous electron transfer processes, involving oxidation and reduction reactions.

3.3. Photocatalytic Mechanism of NO Gas

Based on the trapping results and ESR results, the photocatalytic mechanism of MMT/SnO2/TNTs for NO photocatalytic removal is indicated in Fig. 8. The MMT/SnO2/TNTs heterojunction allows both reduction and oxidation reactions, which leads to the distribution of reduction sites (SnO2 part) and oxidation sites (TNTs) via charge transfer interface. When SnO2 and TNTs combined, the photogenerated electrons from high energy level will transfer to low energy level and establishing the internal electric field formed at the contact surface of the two semiconductors as a result [21]. At the same time, under the internal electric field, the holes in TNTs and electrons in SnO2 transfer and recombine. Consequently, the residual electrons in the valence band (VB) of SnO2 and the conduction band (CB) of TNTs are utilized in the photocatalytic process, exhibiting significant redox potential values. As a result, it is able for the efficient harnessing of the generated electrons (e) and holes (h+) to enhance redox capability and effectively remove NOx gases as Eq. (511) as follows.
(5)
MMT/SnO2/TNTs+hvSnO2(h++e-)+TNTs (h++e-)
(6)
SnO2(h++e-)+TNTs (h++e-)SnO2(h+)+TNTs (e-)
(7)
TNTs (e-)+O2O2-
(8)
O2-+NONO3-
(9)
SnO2(h+)+OH-OH
(10)
OH+NOHNO2
(11)
HNO2+OHNO2+H2O

Conclusion

In this study, we successfully synthesized the MMT mineral combined SnO2/TNTs photocatalysts for a good NO gas removal. The presence of metal ions in MMT was also highlighted, serving as co-catalysts to enhance the overall photocatalytic performance of the MMT/SnO2/TNTs material. The 10% MMT/SnO2/TNTs photocatalyst exhibited a high green product yield of 54.3% and high stability. The 10% MMT/SnO2/TNTs photocatalyst also expressed both oxidation and reduction reactions in the NO photocatalytic reactions. The incorporation of MMT has led to a noteworthy reduction in the production cost of the catalyst, making it a cost-effective solution.

Supplementary Information

Acknowledgments

The authors sincerely thank you for the support from HUTECH University, Vietnam.

Notes

Author Contributions

P.H.N: Investigation, Writing – Original draft; H.C.N: Formal analysis; X.Q.N and K.T.N: Formal analysis; T.M.C: Formal analysis; V.V.P: Writing – Review & Editing, Supervisor.

Conflict of interest statement

On behalf of all authors, the corresponding author states that there is no conflict of interest.

References

1. Yaomin J, María CV, Christian K. Bioprocesses for the removal of nitrogen oxides from polluted air. J. Chem. Technol. Biotechnol. 2005;80(5)483–494. https://doi.org/10.1002/jctb.1260
crossref

2. Schneider T, Grant L. Air pollution by nitrogen oxides. 2013. Elsevier;


3. Ute L, Silke G, Xaver B. Effects of nitrogen dioxide on human health: systematic review of experimental and epidemiological studies conducted between 2002 and 2006. Int. J. Hyg. Environ. Health. 2009;212(3)271–287. https://doi.org/10.1016/j.ijheh.2008.06.003
crossref pmid

4. Fatemeh G, Martin T, Zahra G, Mohammadtaghi V. Technologies for the nitrogen oxides reduction from flue gas: A review. Sci. Total Environ. 2020;714:136712. https://doi.org/10.1016/j.scitotenv.2020.136712
crossref pmid

5. Nguyen HP, Cao MT, Nguyen TT, Pham VV. Improving photocatalytic oxidation of semiconductor (TiO2, SnO2, ZnO)/CNTs for NOx removal. J. Ind. Eng. Chem. 2023;127:321–330. https://doi.org/10.1016/j.jiec.2023.07.017
crossref

6. Oh YJ, Choi G, Choy YB, et al. Aripiprazole Montmorillonite: A New Organic–Inorganic Nanohybrid Material for Biomedical Applications. Chem. Eur. J. 2013;19(15)4869–4875. https://doi.org/10.1002/chem.201203384
crossref pmid

7. Sharma S, Kumar R, Raizada P, et al. An overview on recent progress in photocatalytic air purification: Metal-based and metal-free photocatalysis. Environ. Res. 2022;113995. https://doi.org/10.1016/j.envres.2022.113995
crossref pmid

8. Sun P, Han S, Liu J, et al. Introducing oxygen vacancies in TiO2 lattice through trivalent iron to enhance the photocatalytic removal of indoor NO. Int. J. Miner. Metall. Mater. 2023;30(10)2025–2035. https://doi.org/10.1007/s12613-023-2611-z
crossref

9. Duan Y, Lou J, Zhou S, et al. TiO2-supported Ag nanoclusters with enhanced visible light activity for the photocatalytic removal of NO. Appl. Catal. B. 2018;234:206–212. https://doi.org/10.1016/j.apcatb.2018.04.041
crossref

10. Cao J, Hhasegawa T, Asakura Y, et al. Synthesis of crystal-phase and color tunable mixed anion co-doped titanium oxides and their controllable photocatalytic activity. Int. J. Miner. Metall. Mater. 2023;30(10)2036–2043. https://doi.org/10.1007/s12613-022-2573-6
crossref

11. Chen FL, Chung HW, Zong NO. Degradation of 4-chlorophenol in TiO2, WO3, SnO2, TiO2/WO3 and TiO2/SnO2 systems. J. Hazard. Mater. 2008;154(1–3)1033–1039. https://doi.org/10.1016/j.jhazmat.2007.11.010
crossref pmid

12. Fatimah I, Wang S, Wulandari D. ZnO/montmorillonite for photocatalytic and photochemical degradation of methylene blue. Appl. Clay. Sci. 2011;53(4)553–560. https://doi.org/10.1016/j.clay.2011.05.001
crossref

13. Davidson EE, Nanfe RP, Napoleon W. Synthesis of a copper (II) oxide–montmorillonite composite for lead removal. Int J Miner Metall Mater H13-019-1788-7. https://doi.org/10.1007/s12613-019-1788-7
crossref

14. Muhammad T, NorAishah SA. Photocatalytic reduction of carbon dioxide with water vapors over montmorillonite modified TiO2 nanocomposites. Appl. Catal. B. 2013;142:512–522. https://doi.org/10.1016/j.apcatb.2013.05.054
crossref

15. An T, Chen J, Li G, et al. Characterization and the photocatalytic activity of TiO2 immobilized hydrophobic montmorillonite photocatalysts: Degradation of decabromodiphenyl ether (BDE 209). Catal. 2008;139(1)69–76. https://doi.org/10.1016/j.cattod.2008.08.024
crossref

16. Tran HH, Bui DP, Kang F, et al. SnO2/TiO2 nanotube heterojunction: The first investigation of NO degradation by visible light-driven photocatalysis. Chemosphere. 2019;215:323–332. https://doi.org/10.1016/j.chemosphere.2018.10.033
crossref pmid

17. Sim LC, Ng KW, Ibrahim S, Saravanan P. Synthesis, features and solar-light-driven photocatalytic activity of TiO2 nanotube arrays loaded with SnO2 . J. Nanosci. Nanotechnol. 2014;14(9)7001–7009. https://doi.org/10.1166/jnn.2014.8931
crossref pmid

18. Yu J, Zeng X, Wu S, Liu G. Preparation and properties of montmorillonite modified asphalts. Mater. Sci. Eng. A. 2007;447(1–2)233–238. https://doi.org/10.1016/j.msea.2006.10.037
crossref

19. Pannak P, Songsasen A, Foytong W, Kidkhunthod P, Sirisaksoontorn W. Homogeneous distribution of nanosized ZnO in montmorillonite clay sheets for the photocatalytic enhancement in degradation of Rhodamine B. Res. Chem. Intermed. 2018;44:6861–6875. https://doi.org/10.1007/s11164-018-3526-6
crossref

20. Shi X, Wang P, Wang L, et al. Change in photocatalytic NO removal mechanisms of ultrathin BiOBr/BiOI via NO3 adsorption. Appl. Catal. B. 2019;243:322–329. https://doi.org/10.1016/j.apcatb.2018.10.037
crossref

21. Huang H, Zhao S, Yang Y, et al. Axially wrinkled tubular SnO2/TiO2 heterostructures for effective degradation of organic pollutants. Mater. Sci. Semicond. Process. 2022;152:107065. https://doi.org/10.1016/j.mssp.2022.107065
crossref

Fig. 1
XRD patterns (a), FTIR spectrum (b) of SnO2/TNTs and MMT/SnO2/TNTs.
/upload/thumbnails/eer-2024-268f1.gif
Fig. 2
XPS of MMT/SnO2/TNTs (a) and HRXPS of Sn3d (b), Ti2p (c) and O1s (d).
/upload/thumbnails/eer-2024-268f2.gif
Fig. 3
SEM images of 10% MMT (a), EDX mapping was performed to analyze the distribution of Sn (b), Ti (c), O (d), Si (e), and K (f) elements, EDX spectrum of the 10% MMT sample (g).
/upload/thumbnails/eer-2024-268f3.gif
Fig. 4
Nitrogen adsorption-desorption isotherms of the MMT and 10% MMT/SnO2/TNTs photocatalysts
/upload/thumbnails/eer-2024-268f4.gif
Fig. 5
DRS spectra (a), Kubelka-Munk (b) of SnO2/TNTs and MMT/SnO2/TNTs.
/upload/thumbnails/eer-2024-268f5.gif
Fig. 6
Photocatalytic NO gas removal efficiency (a), NO conversion efficiency into clean products and NO into NO2 (b), quantum efficiency (c), photocatalytic ability over scavengers (d) of SnO2/TNTs and MMT/SnO2/TNTs materials.
/upload/thumbnails/eer-2024-268f6.gif
Fig. 7
The stability and efficiency of the material after five cycles of reuse (a), as well as the FTIR spectra of the initial 10% MMT material and after five cycles of use
/upload/thumbnails/eer-2024-268f7.gif
Fig. 8
NO photocatalytic degradation mechanism of the MMT/SnO2/TNTs heterojunction.
/upload/thumbnails/eer-2024-268f8.gif
TOOLS
PDF Links  PDF Links
PubReader  PubReader
Full text via DOI  Full text via DOI
Download Citation  Download Citation
Supplement  Supplement
  Print
Share:      
METRICS
0
Crossref
0
Scopus
840
View
26
Download
Editorial Office
464 Cheongpa-ro, #726, Jung-gu, Seoul 04510, Republic of Korea
FAX : +82-2-383-9654   E-mail : eer@kosenv.or.kr

Copyright© Korean Society of Environmental Engineers.        Developed in M2PI
About |  Browse Articles |  Current Issue |  For Authors and Reviewers