Next Article in Journal
Study on the Catalytic Oxidation Modification Effect of Heavy Oil at Low Temperature under the Action of Different Ligand Ferric-Based Systems
Previous Article in Journal
Platinum-Modified Mixed Metal Oxide Electrodes for Efficient Chloralkaline-Based Energy Storage
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Templated Synthesis of Cu2S Hollow Structures for Highly Active Ozone Decomposition

1
Navy Submarine Academy, Qingdao 266199, China
2
State Key Laboratory of Mesoscience and Engineering, Institute of Process Engineering, Chinese Academy of Sciences, Beijing 100190, China
3
School of Chemical Engineering, University of Chinese Academy of Sciences, Beijing 100049, China
*
Authors to whom correspondence should be addressed.
Catalysts 2024, 14(2), 153; https://doi.org/10.3390/catal14020153
Submission received: 11 January 2024 / Revised: 7 February 2024 / Accepted: 7 February 2024 / Published: 19 February 2024
(This article belongs to the Section Catalytic Materials)

Abstract

:
Nowadays, it is highly desired to develop highly active and humidity-resistive ozone decomposition catalysts to eliminate the ozone contaminant, one of the primary pollutants in the air. In this work, a series of Cu2S hollow structured materials were rapidly synthesized using different structured Cu2O templates. The Cu2S from porous Cu2O showed the highest ozone catalytic decomposition efficiency of >95% to 400 ppm ozone with a weight hourly space velocity of 480,000 cm3·g−1·h−1 in dry air. Importantly, the conversion remained >85% in a high relative humidity of 90%. The mechanism was explored by diffusive reflectance infrared spectroscopy which showed the decomposition intermediate of O22−, and X-ray photoelectron spectroscopy revealed the dual active site of both Cu and S. The EPR and UPS characterization results also explained the superiority of porous Cu2S catalysts from the material itself. All these results show the effective decomposition of ozone by Cu2S, especially in harsh environments, promising for active ozone elimination.

Graphical Abstract

1. Introduction

Among the various air pollutants that humans are currently confronted with, near-surface ozone (O3) pollution presents a particularly challenging issue [1]. This type of pollution is commonly found in large urban areas and is closely linked to the presence of volatile organic compounds (VOCs) and nitrogen oxides (NOx), which pose significant health risks [2,3]. O3 is currently recognized as the second most harmful air pollutant and is the primary air pollutant in many urban areas during the summer. Consequently, there is a widespread and pressing need for O3 treatment technology, both in outdoor environments and enclosed spaces. The most effective ozone treatment technology currently relies on precious metal catalysts such as Au and Pd, which offer excellent performance but lack cost-effectiveness [4,5]. In contrast, transition metal oxides such as MnOx and Cu2O are more favorable due to their high efficiency and lower cost for O3 decomposition [6,7,8].
It is important to note that transition metal oxides typically exhibit high activity in decomposing O3 in dry air, but their performance is hindered by competitive adsorption with water vapor, resulting in reduced activity. Additionally, O3 pollution is often accompanied by more acidic gas pollutants such as SO2 and NO2. For example, Ma et al. [9] showed that although SO2 and NO2 pollution in China’s atmosphere has been significantly reduced in recent years, their total amount is still higher than O3. Mukta et al. [10] reported that the concentration of NO2 is always higher than O3 in Gazipur in Bangladesh. The study by Stevens et al. [11] also supports the considerable amount of acidic gases such as NO2 and SO2 in the atmosphere of Europe and North America. Therefore, in addition to the interference caused by humidity, metal oxide catalysts are also susceptible to combining with acidic substances, which can hasten the deactivation of the catalysts.
Copper sulfide (Cu2S), a narrow bandgap p-type semiconductor, has been extensively researched and utilized in advanced fields such as photocatalysis and high-temperature superconductivity due to its distinctive crystal structure and band gap [12,13,14,15,16]. Notably, Cu2S demonstrates exceptional acid resistance owing to its extremely low solubility (Ksp = 2 × 10−47) [17,18]. This insolubility provides additional advantages to sulfides in creating complex morphologies, which are essential for catalysts, particularly certain hollow structures and microstructures. Chun-Hong Kuo et al. [19] reported that Cu2O nanoparticles can be gradually transformed into Cu2S by 0.2 M Na2S solution in 360 s. After being transformed into Cu2S, the nanoparticles form a hollow cage structure based on their original cubic morphology. For example, Lei Ran et al. [20] reported a synthesis process of double-walled heterostructured Cu2-xSe/Cu7S4 nano boxes used as a material for quantum dot sensitized solar cells. The complex morphology of this catalyst was synthesized using a simple Cu2O template. Cu2S materials have been also widely studied and applied in the field of hydrogen energy [21]. They are used as catalysts for the hydrogen evolution reaction (HER) in electrochemical water splitting, which is a key process for producing clean hydrogen fuel [22]. Copper sulfide materials have shown promising catalytic activity and stability for HER, making them potential candidates for efficient and cost-effective hydrogen production [23]. Although the application of Cu2S materials in CO2 conversion, HER, and solar cell materials has attracted increasing attention from researchers, there are currently no reports on the application of Cu2S in O3 decomposition. In a previous study, p-type semiconductor showed relatively higher O3 decomposition activity than the n-type counterparts [24]. As a typical p-type semiconductor, Cu2S has more Cu vacancies, which is advantageous in the catalyzed decomposition of ozone [25,26,27,28]. The performance of Cu2S in ozone catalytic decomposition is worth exploring, because its convenient morphology construction and stable composition are rare advantages.
In this study, hollow structures of Cu2S were fabricated using a Cu2O sacrificial template method and employed as a catalyst for ozone decomposition. The resulting catalyst exhibited high activity and demonstrated strong resistance to relative humidity, indicating potential for effective ozone decomposition in the atmosphere.

2. Results and Discussion

The Cu2O sacrificial template was synthesized according to the experimental Section 3.1. Cu2S was synthesized using Cu2O with porous (Cu2O-P), spherical (Cu2O-P), and cubic (Cu2O-C) morphologies as sacrificial templates, which are named Cu2S porous (Cu2S-P), Cu2S sphere (Cu2S-S), and Cu2S cube (Cu2S-C). The catalytic performance of Cu2S for ozone decomposition is superior to its corresponding Cu2O template. The performance of Cu2S varies significantly among the three morphologies, with Cu2S-P exhibiting substantially better performance (Figure 1a). More than 95% of 400 ppm O3 can be decomposed at a space velocity of WHSV 480,000 cm3·g−1·h−1 in dry conditions. To verify whether Cu2S is a catalyst or a simple chemically absorbing agent for O3, a durability test was conducted on Cu2S-P in Figure 1b. After working continuously for 18 h, the cumulative amount of O3 processed was 3.86 mmol, which was tens of times higher than the amount of Cu2S-P substance (50 mg, ~0.31 mmol). After dealing with much more O3 than itself, the O3 decomposition activity still exceeds 90%, which proves the catalytic process of ozone decomposition by Cu2S, rather than a simple oxidation/reduction process of Cu2S and O3. At the same time, the catalytic decomposition performance of Cu2S in high-humidity environments was investigated. As shown in Figure 1c, due to the competitive adsorption of water molecules, high humidity has a particular impact on its catalytic performance. However, the conversion still maintains over 85% performance under conditions of 480,000 cm3·g−1·h−1 and humidity greater than 90%, showing the good humidity resistance of this catalyst.
To investigate the relationship between structure and performance further, XRD characterization was performed on three types of Cu2S. In Figure 2, the characteristic peak intensity in Cu2S-P and Cu2S-S is relatively weak, with most Cu2S exhibiting an amorphous structure. Cu2S-C particles crystallize better, and the three types of Cu2S conform with the calcite alpha low in the standard card (JCPDS Ref. code 00-009-0328). High crystallinity means fewer defects, which is usually detrimental to the performance of heterogeneous catalysts, and may be the reason for the poor performance of Cu2S-C.
To further understand the specific morphologies of Cu2S samples and their templates, FESEM and TEM techniques were used as shown in Figure 3. Through SEM images, all three Cu2O templates exhibit different morphologies: the sphere template has fine particles on the surface (Figure 3a), the cube template reveals varying sizes (Figure 3b), and the porous templates are formed by the accumulation of many small particles (Figure 3c). Due to the much lower solubility product of Cu2S in water compared to Cu2O [19], Cu2O dispersed in water is prone to generate more insoluble Cu2S with S2−, and the morphology of the developed product is correlated with the sacrificial template (Cu2O) [29]. However, due to differences in mass transfer rates, internal substances quickly transfer to the surface through mass transfer channels and form voids inside, as shown in Figure 3d,e, which is described in the literature as the Kirkendall effect [30,31,32]. The significant contrast difference between the edge and center of particles in TEM images indicates that Cu2S-C and Cu2S-S have hollow structures. However, the tiny particles in Figure 3f were not observed to have apparent hollow structures, and it is speculated that the nanoscale effect of the stacked small particles is enhanced, thereby overcoming the interface resistance and fusing. The average diameter of Cu2S-C and Cu2S-S particles slightly increased (by about 10% compared to the template), which can be attributed to the hollowing of the internal structure; the diameter of Cu2S-P particles increased significantly (about 40%) due to the agglomeration of some particles during sulfurization. All these processes can be clearly seen in the schematic view in Figure 4: Taking spherical Cu2O particles as an example, when spherical particles are suspended in water solution, the ions (Cu+) in the solution are attracted by the surface charge of the particles, forming an electrical double layer. This electrical double layer leads to an increase in the solution concentration near the particle surface, creating a concentration gradient in the solution. Due to the concentration gradient generating osmotic pressure, solvent molecules accumulate near the particle surface, further increasing the solution concentration near the particle surface. The particle core transports Cu to the surface through several mass transfer channels, gradually forming voids inside. This process represents the Kirkendall effect in spherical particles. This effect is equally applicable in both cubic and porous structures.
Another important property of the catalyst is the specific surface area and surface pore structure. The characterization results of three types of Cu2S are shown in Table 1. According to the BET characterization results, Cu2S-P has the largest specific surface area and the smallest average pore size of about 5 nm. Although studies [33,34] have shown that specific surface area is not a decisive indicator of the performance of ozone catalysts, Cu2S-P has a larger specific surface area and pore capacity compared to other structures, which can improve its contact efficiency with ozone gas flow and is a favorable factor in situations where surface chemical properties are similar.
Although the sample with the best performance also has the largest specific surface area, overall, the difference in specific surface area among the three morphologies of Cu2S particles is not significant. It seems difficult to explain the significant performance differences, especially in Cu2S-P. Therefore, EPR characterization was employed to measure the sulfur vacancies in three different morphologies of Cu2S. Figure 5 shows that Cu2S-P has prominent sulfur vacancies [35], which are not present in other samples. This can be attributed to the generation of S vacancies caused by the nanoscale effect of small particles during mass transfer. More sulfur vacancies mean more defects and crystal structure imbalances, which are usually beneficial for catalytic reactions. It is generally believed in the literature that these sulfur vacancies are active sites [36,37].
The work function (Φ) has an important impact on the performance of semiconductor catalysts, which can affect their electron transfer, reaction activity, and photocatalytic performance, thereby affecting the efficiency and effectiveness of catalysts in catalytic reactions [38]. The photoelectron spectroscopy method obtains the escape work of materials by measuring the non-elastic secondary electron cutoff edge. According to the basic energy relationship of photoelectric emission, the energy interval from the cutoff edge of inelastic scattering secondary electrons to the vacuum level is the energy of photons [39]. Therefore, the work function is calculated as follows:
Φ = hν − (EFermi − Ecutoff)
Figure 6 shows that there are significant differences in the work functions of the three morphologies of Cu2S materials. Although the work function is usually not significantly correlated with the catalytic performance of the material, the lower work function of the Cu2S-P (3.04 eV) material means that electrons have higher transfer efficiency in the system, which is more favorable for the adsorption and desorption of ozone molecules. This also explains that Cu2S-S has a lower specific surface area and similar pore capacity, while its catalytic performance is not inferior to Cu2S-C. Although we have explained in the discussion of the data results in Table 1 that the internal surface areas of hollow spheres and hollow cubes may make it difficult for them to participate in catalytic reactions, there is no difference in the adsorption/desorption data of mesoporous performance, and the difference in their work functions may be one of the determining factors.
To reveal the process of ozone conversion on the catalyst surface, in situ FTIR spectroscopy was employed to detect changes in surface functional groups after exposure to ozone. In Figure 7a, the introduction of ozone produced a weak O22− [40] peak at a wavenumber of 756 cm−1, and in Figure 7b, the peak weakens after stopping the ozone for a period, which is similar to the transition metal oxides in the literature [41,42], which can be attributed to the classical reaction process of ozone on sulfide surfaces. Due to the widespread presence of ppm-level water vapor in gas cylinders, its characteristic peaks gradually increase over time. In addition, signals generated by S–O (1060 cm−1) and S=O (1350–1420 cm−1) stretching vibrations [43] were observed on the surface of Cu2S and did not disappear after the introduction of ozone was stopped. This indicates that during the surface transformation process of ozone, a layer of sulfate salt is formed on the surface of Cu2S. The literature indicates that a surface sulfate layer helps maintain the stability of catalyst particles [44]. This result is consistent with Figure 5. The S vacancy on the surface is beneficial for O3 adsorption, and the adsorbed O atom fills the vacancy, stabilizing the surface chemical structure and facilitating further participation in surface reactions.
XPS characterization was performed on Cu2S-P before and after the reaction to study the mechanism of the catalytic decomposition of ozone. In Figure 8a, the S 2p 1/2 and S 2p 2/3 peaks at ~162 eV can be attributed to the S–Cu bond [45], and peaks at ~169 eV can be attributed to the S–O bond [46]. After continuous operation for 18 h, the S–O bond in Cu2S-P was enhanced, and the Cu II peak was clearly observed in the shoulder peak of the dominant Cu I peak (Figure 8b). The enhancement of S–O bonds indicates that the surface S atoms are also effective active sites besides the commonly believed active sites of Cu atoms and defects. Secondly, after 18 h of continuous catalysis of O3, there is a certain degree of oxidation on the catalyst surface, but the initial composition is still dominant. This indicates that the durability of sulfide-catalytic materials and oxide-catalyzed ozone is similar. In some previous studies [47], the same situation is commonly observed in metastable transition metal oxides. Since the test sample for XPS is Cu2S-P, this may also be due to a large number of S vacancies on the surface being filled by O atom, which has a high electronegativity and is prone to further electron valence change in Cu. Combined with the surface depth limitation of XPS testing, this phenomenon is more pronounced on the surface.
In addition, Raman spectroscopy characterization was performed on Cu2S-P before and after use, and the results are shown in Figure 9. According to the literature [48,49,50], the Raman peaks at 268 cm−1 and 472 cm−1 are attributed to Cu2S, while the Raman peak at 284 cm−1 is attributed to Cu2O. Consistent with the XPS characterization results, Raman spectroscopy shows that although there is a weak CuO signal in the catalyst after 18 h of use, Cu2S is still the main component, further proving the catalyst’s stability.
Based on the above characterization and test results, it can be inferred that Cu2S, a p-type semiconductor similar to Cu2O, has a similar catalytic principle for ozone decomposition. However, the active center has been replaced by metal atoms and oxygen defects in transition metal oxides with metal and surface sulfur atoms. The catalytic process and intermediate products, as proposed by Oyama et al. [24], are described in Equations (1)–(3).
O3 + active site (Cu, S)* O2 + O
O3 + O + * O2 + O22−
O22−  O2 + 2*
The charged intermediate products in the process of ozone decomposition have strong oxidizing properties, which can quickly oxidize some surfaces and transform them into CuO. This heterostructure is not entirely unfavorable for ozone decomposition, as shown in Figure 1c, where the performance improves in the first few hours and then slightly decreases. This is very consistent with the performance of cuprous oxide catalytic materials, and all these results show the promising prospect of the Cu2S catalyst for effective ozone decomposition.

3. Experimental

3.1. Synthesis of Catalysts

The Cu2S catalyst was prepared using the sacrificial template method. Referring to the synthesis principle in the literature [33,51,52], three types of Cu2O templates were synthesized using different methods: cubic, porous, and spherical. The specific methods are as follows.
Cubic Cu2O: 80 mmol (20 g) CuSO4·5H2O (Macklin reagent, AR, Jining, China) was dissolved in 250 mL distilled water. Then, 100 mL NaOH (Macklin reagent, AR, Jining, China) solution (4 mol/L) was added dropwise. After further stirring the blue suspension for 30 min, 112 mL ascorbic acid (AA) (MERYER CHEMICAL, 99%, Shanghai, China) aqueous solution (1 mol/L) was added dropwise within approximately 5 min. The solution was then stirred until the suspension gradually turned orange red. The precipitate was separated by centrifugation, washed three times with water and ethanol, and then dried in an 80 °C oven for 8 h.
Spherical Cu2O: 20 mmol (3.98 g) Cu(OAC)2·H2O (Macklin reagent, AR, Jining, China) was dissolved in 150 mL distilled water. After it was completely dissolved, 10 mL NaOH solution (4 mo1/L) was added and stirred for 30 min. Then, 20 mL AA aqueous solution (1 mol/L) was added dropwise. The precipitate was obtained by centrifuge, washed three times with water and ethanol, and then dried in an 80 °C oven for 8 h.
Porous Cu2O: 20 mmol (5 g) CuSO4·5H2O was dissolved in 150 mL distilled water. After it was completely dissolved, 10 mL NaOH solution (4 mol/L) was added. After further stirring the blue suspension for 30 min, 3.5 g AA powder was added directly. After stirring for 5 min, the sediment was separated by centrifuge, washed three times with water and ethanol, and then dried in an 80 °C oven for 8 h.
Cu2S hollow structures were synthesized by vulcanizing the three Cu2O templates. They were added to a Na2S·9H2O (Aladdin Reagent Co., Ltd., 99%, Shanghai, China) aqueous solution at a concentration of 2M with a S:Cu molar ratio of 1:2. The suspension was stirred for 15 min to react, and then the powders were obtained by separation, washing, and drying as the porous Cu2O sample.

3.2. Characterization of the Catalyst

The catalyst’s crystal structure was analyzed using an X-ray diffractometer (XRD, X’pert Pro System) manufactured by Panalytical, Eindhoven The Netherlands, operating at 40 kV and 40 mA, with Cu Kα radiation (wavelength of 0.154 nm). The scanning range was 5–90° at a speed of 10° min−1. The obtained results were converted to xrdml format using PowDll software V 2.97, and then imported into XpertHighscore software Version plus 3 for analysis and compared with standard cards in the database. Microscopy analysis was conducted using an ultra-high resolution field emission scanning electron microscope (FESEM, JSM-7800, JEOL, Tokyo, Japan). The Cu2O/Cu2S were fixed onto the sample stage using conductive tape and the sample morphology was captured under an accelerated voltage of 15 kV after spray-gold treatment. The focus and brightness were adjusted, then the image was recorded at a magnification of 50,000 to 100,000. The transmission electron microscope (TEM) photos and high-resolution transmission electron microscopy (HRTEM) were conducted using a JEOL JEM-2100F transmission electron microscope (JEOL, Nagoya, Japan) at an accelerating voltage of 200 kV. The catalyst samples were dispersed in ethanol and loaded onto copper grids covered with microgrid carbon films. The Cu2S products’ specific surface areas were analyzed using the Brumaire–Emmett–Teller (BET) method with a Surface Area Analyzer by Micromeritics (ASAP2460, Beijing, China) at a temperature of liquid nitrogen (−196 °C) with N2 gas as the adsorbate. Prior to the analysis, the samples were dried at 120 °C for 4 h and then degassed at 150 °C for 1 h. The pore size distributions were determined from the desorption branches of the isotherms based on the Barrett–Joyner–Halenda (BJH) theory. Surface chemical bonds and chemical states of the catalysts were characterized by X-ray photoelectron spectroscopy (XPS, ESCALAB 250XI, Thermo Fisher, Waltham, MA, USA) using a monochromatic Al Kα X-ray source (1486 eV) with a beam size of 200 μm. Charge compensation was achieved by dual-beam charge neutralization, and the binding energy was corrected by setting the binding energy of the hydrocarbon C 1s feature to 284.8 eV. The electron paramagnetic resonance (EPR) signals of radicals’ spin trapped were carried out on a Bruker EMXplus-6/1 (Munich, Germany) spectrometer by spin-trap reagents of DMPO and TEMP for S vacancy. The ultraviolet photoelectron spectroscopy (UPS) testing instrument was the ESCALab 250Xi multi energy electron spectrometer produced by ThermoScientific (Waltham, MA, USA). During UPS testing, a He I ultraviolet light source (h ≈ 21.22 eV) was selected, and the vacuum in the analysis room was about 3 × 10−6 Pa. The bias voltage was set at −5 eV in the experiment. An appropriate energy analyzer was selected for energy and spectral scanning range, and the corresponding secondary electron energy distribution curve was recorded. The measurement of the secondary electron cutoff edge (Ecutoff) and Fermi edge (EFermi) involved cleaning the surface of the semiconductor sample with an Ar+ beam and correcting it with an Au standard sample before testing. Raman spectroscopy was analyzed at LabRAM HR Evolution (Horiba, Kyoto, Japan). The measurement span was from 1500 to 200 cm−1 at room temperature. The 633 nm line of the laser was used as the excitation source, with the capability of supplying 250 mW. Intermediate products and surface adsorption groups on the Cu2S catalyst were characterized via FTIR-DRIFTS (Diffuse Reflectance Infrared Fourier Transform Spectroscopy), the infrared spectrometer was made by BRUKER (INVENIO, Ettlingen, Germany). The sample was mixed with equal mass KBr powder in an in situ cell, and the measured wavelength range was from 400 to 4000 cm−1. After the sample was placed in the gas cell, the entire experimental section was heated to 150 °C under nitrogen protection for 2 h, and then cooled to room temperature to drive away moisture and other adsorbed gases. Before testing, the sample was first purged with dry nitrogen gas for 1 h, followed by oxygen as the background. Afterwards, the ultraviolet ozone generation device was activated in the oxygen gas path, and the ozone concentration was maintained at approximately 30–50 ppm at a flow rate of 100 sccm. Stopping the introduction of ozone refers to shutting down the ozone generator and continuing to blow the sample with the original flow rate of O2/N2 gas.

3.3. Ozone Decomposition Test

The O3 decomposition performance was tested in a U-shape quartz tube reactor (diameter 5.5 mm) at 25 °C with 50 mg (40–60 mesh) Cu2S mixed with 450 mg quartz sand. The entire reactor was placed in a water bath constant temperature device (SC-20A, Doosi, Shanghai, China) to ensure that the temperature was not affected by the laboratory environment or thermal effects generated during the reaction.
The overall weight hourly space velocity (WHSV) remained at 480,000 cm3·g−1·h−1. Ozone was generated at concentrations from 200 to 400 ppm by a commercial ozone generator (COM-AD-01-OEM, ANSEROS COMPANY, Anshan, China), and the inlet and outlet ozone concentrations were analyzed by an ozone monitor (model 106MH, 2B Technologies, Denver, CO, USA). Ozone was generated from pure oxygen and mixed with air in a mixing bottle to ensure stable concentration. The ozone conversion was calculated as:
Conv. of O3 = 100% × (O3 inlet − O3 outlet)/O3 inlet
The moisture was produced by bubbling water with the airflow; in the bubbling device, it can be ensured that the relative humidity is greater than 90. The relative humidity (RH) was measured by a humidity and temperature sensor meter (center 310 RS-232, TES, Xinbei, China).

4. Conclusions

Using the Kirkendall effect, Cu2S hollow structured materials with different morphologies were rapidly synthesized from different shaped Cu2O templates. The catalytic effect of Cu2S material on the ozone decomposition process was confirmed by combining actual test results and material characterization before and after the reaction. Its catalytic efficiency of 400 ppm ozone can exceed 95% at WHSV of 480,000 cm3·g−1·h−1 under dry conditions, one of the highest in the literature. The DRIFT results showed the intermediate O22−, showing the catalytic reaction mechanism. Furthermore, sulfur can be also active in ozone decomposition as illustrated by XPS, which has broadened the application prospects in the catalytic decomposition of ozone.

Author Contributions

Conceptualization, methodology: Q.Z., F.X. and G.M.; formal analysis, investigation, data curation, writing—original draft preparation: Y.J., Y.X. and X.Z.; writing—review and editing, resources, supervision: F.X. and G.M.; software: X.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was financially supported by a research fund of State Key Laboratory of Mesoscience and Engineering (MESO-23-A06).

Data Availability Statement

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

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Brunekreef, B.; Holgate, S.T. Air pollution and health. Lancet 2002, 360, 1233–1242. [Google Scholar] [CrossRef]
  2. Taylan, O.; Alkabaa, A.S.; Alamoudi, M.; Basahel, A.; Balubaid, M.; Andejany, M.; Alidrisi, H. Air Quality Modeling for Sustainable Clean Environment Using ANFIS and Machine Learning Approaches. Atmosphere 2021, 12, 713. [Google Scholar] [CrossRef]
  3. Li, Y.; Yin, S.; Yu, S.; Bai, L.; Wang, X.; Lu, X.; Ma, S. Characteristics of ozone pollution and the sensitivity to precursors during early summer in central plain, China. J. Environ. Sci. 2021, 99, 354–368. [Google Scholar] [CrossRef] [PubMed]
  4. Touati, H.; Mehri, A.; Karouia, F.; Richard, F.; Batiot-Dupeyrat, C.; Daniele, S.; Clacens, J.-M. Low-Temperature O3 Decomposition over Pd-TiO2 Hybrid Catalysts. Catalysts 2022, 12, 448. [Google Scholar] [CrossRef]
  5. Hao, Z.; Cheng, D.; Yun, G.; Liang, Y. Supported gold catalysts used for ozone decomposition and simultaneous elimination of ozone and carbon monoxide at ambient temperature. Appl. Catal. B-Environ. 2001, 33, 217–222. [Google Scholar] [CrossRef]
  6. Gong, S.; Wu, X.; Zhang, J.; Han, N.; Chen, Y. Facile solution synthesis of Cu2O–CuO–Cu(OH)2 hierarchical nanostructures for effective catalytic ozone decomposition. Cryst. Eng. Comm. 2018, 20, 3096–3104. [Google Scholar] [CrossRef]
  7. Deng, H.; Kang, S.; Ma, J.; Wang, L.; Zhang, C.; He, H. Role of Structural Defects in MnOx Promoted by Ag Doping in the Catalytic Combustion of Volatile Organic Compounds and Ambient Decomposition of O3. Environ. Sci. Technol. 2019, 53, 10871–10879. [Google Scholar] [CrossRef]
  8. Wang, A.; Zhang, L.; Guan, J.; Wang, X.; Ma, G.; Fan, G.; Wang, H.; Han, N.; Chen, Y. Highly efficient ozone elimination by metal doped ultra-fine Cu2O nanoparticles. J. Environ. Sci. 2023, 134, 108–116. [Google Scholar] [CrossRef]
  9. Chu, B.; Ma, Q.; Liu, J.; Ma, J.; Zhang, P.; Chen, T.; Feng, Q.; Wang, C.; Yang, N.; Ma, H.; et al. Air Pollutant Correlations in China: Secondary Air Pollutant Responses to NOx and SO2 Control. Environ. Sci. Technol. Lett. 2020, 7, 695–700. [Google Scholar] [CrossRef]
  10. Mukta, A.; Hoque, M.; Sarker, E.; Hossain, N.; Biswas, K. Seasonal variations of gaseous air pollutants (SO2, NO2, O3, CO) and particulates (PM2.5, PM10) in Gazipur: An industrial city in Bangladesh. Adv. Environ. Technol. 2020, 6, 195–209. [Google Scholar] [CrossRef]
  11. Stevens, C.J.; Bell, J.N.B.; Brimblecombe, P.; Clark, C.M.; Dise, N.B.; Fowler, D.; Lovett, G.M.; Wolseley, P.A. The impact of air pollution on terrestrial managed and natural vegetation. Philos. T R Soc. A. 2020, 378, 20190317. [Google Scholar] [CrossRef]
  12. Ranjith, K.S.; Ranjith Kumar, D.; Huh, Y.S.; Han, Y.-K.; Uyar, T.; Rajendra Kumar, R.T. Promotional Effect of Cu2S–ZnS Nanograins as a Shell Layer on ZnO Nanorod Arrays for Boosting Visible Light Photocatalytic H2 Evolution. J. Phys. Chem. C. 2020, 124, 3610–3620. [Google Scholar] [CrossRef]
  13. Ansari, M.Z.; Faraz, M.; Munjal, S.; Kumar, V.; Khare, N. Highly dispersible and uniform size Cu2ZnSnS4 nanoparticles for photocatalytic application. Adv. Powder. Technol. 2017, 28, 2402–2409. [Google Scholar] [CrossRef]
  14. Wang, S.; Huang, Q.; Wen, X.; Li, X.-y.; Yang, S. Thermal oxidation of Cu2S nanowires: A template method for the fabrication of mesoscopic CuxO (x = 1,2) wires. Phys. Chem. Chem. Phys. 2002, 4, 3425–3429. [Google Scholar] [CrossRef]
  15. Zhu, S.; Wu, W.; Li, Z.; Luo, J. First-order transition in LK-99 containing Cu2S. Matter 2023, 6, 4401–4407. [Google Scholar] [CrossRef]
  16. Ahmed, H.S.; Mohammed, R.Y. The Effect of Deposition Parameters on Morphological and Optical Properties of Cu2S Thin Films Grown by Chemical Bath Deposition Technique. Photonics 2022, 9, 161. [Google Scholar] [CrossRef]
  17. Chen, T.-W.; Rajaji, U.; Chen, S.-M.; Govindasamy, M.; Paul Selvin, S.S.; Manavalan, S.; Arumugam, R. Sonochemical synthesis of graphene oxide sheets supported Cu2S nanodots for high sensitive electrochemical determination of caffeic acid in red wine and soft drinks. Compos. Part B Eng. 2019, 158, 419–427. [Google Scholar] [CrossRef]
  18. Honarnezhad, R.; Fathinia, M.; Khataee, A. Mechanical production and sonocatalytic application of Cu2S nanoparticles for degradation of isopropylxanthic acid: Kinetic modeling via white and black box methods. J. Mol. Liq. 2019, 287, 110899. [Google Scholar] [CrossRef]
  19. Kuo, H.; Chu, T.; Song, F.; Huang, H. Cu2O Nanocrystal-Templated Growth of Cu2S Nanocages with Encapsulated Au Nanoparticles and In-Situ Transmission X-ray Microscopy Study. Adv. Funct. Mater. 2011, 21, 792–797. [Google Scholar] [CrossRef]
  20. Ran, L.; Yin, L. Double-walled heterostructured Cu2−xSe/Cu7S4 nanoboxes with enhanced electrocatalytic activity for quantum dot sensitized solar cells. Cryst. Eng. Comm. 2017, 19, 5640–5652. [Google Scholar] [CrossRef]
  21. Yang, D.; Cao, L.; Huang, J.; Liu, Q.; Li, G.; He, D.; Wang, J.; Feng, L. Vanadium-doped hierarchical Cu2S nanowall arrays assembled by nanowires on copper foam as an efficient electrocatalyst for hydrogen evolution reaction. Scripta Mater. 2021, 196, 113756. [Google Scholar] [CrossRef]
  22. Yu, X.; Pan, L.; Son, K.; Mayer, T.; Zhang, D.; Hagfeldt, A.; Luo, J.; Grätzel, M. Solution-Processed Cu2S Photocathodes for Photoelectrochemical Water Splitting. ACS Energy Lett. 2018, 3, 760–766. [Google Scholar] [CrossRef]
  23. Li, Z.; Zhang, Z. Tetrafunctional Cu2S thin layers on Cu2O nanowires for efficient photoelectrochemical water splitting. Nano Res. 2018, 11, 1530–1540. [Google Scholar] [CrossRef]
  24. Dhandapani, B.; Oyama, S.T. Gas phase ozone decomposition catalysts. Appl. Catal. B Environ. 1997, 11, 129–166. [Google Scholar] [CrossRef]
  25. Mousavi-Kamazani, M.; Zarghami, Z.; Salavati-Niasari, M. Facile and Novel Chemical Synthesis, Characterization, and Formation Mechanism of Copper Sulfide (Cu2S, Cu2S/CuS, CuS) Nanostructures for Increasing the Efficiency of Solar Cells. J. Phys. Chem. C 2016, 120, 2096–2108. [Google Scholar] [CrossRef]
  26. Mosali, V.S.S.; Puxty, G.; Horne, M.D.; Bond, A.M.; Zhang, J. Selective electrochemical methanation of carbon dioxide using a sulphide derived CuZn catalyst. Electrochim. Acta 2024, 475, 143628. [Google Scholar] [CrossRef]
  27. Mosali, V.S.S.; Zhang, X.; Liang, Y.; Li, L.; Puxty, G.; Horne, M.D.; Brajter-Toth, A.; Bond, A.M.; Zhang, J. CdS-Enhanced Ethanol Selectivity in Electrocatalytic CO2 Reduction at Sulfide-Derived Cu−Cd. ChemSusChem 2021, 14, 2924–2934. [Google Scholar] [CrossRef] [PubMed]
  28. Van Thang, B.; Tung, H.T.; Phuc, D.H.; Nguyen, T.P.; Van Man, T.; Vinh, L.Q. High-efficiency quantum dot sensitized solar cells based on flexible rGO-Cu2S electrodes compared with PbS, CuS, Cu2S CEs. Sol. Energy Mater. Sol. Cells 2023, 250, 112042. [Google Scholar] [CrossRef]
  29. Ma, R.; Stegemeier, J.; Levard, C.; Dale, J.G.; Noack, C.W.; Yang, T.; Brown, G.E.; Lowry, G.V. Sulfidation of copper oxide nanoparticles and properties of resulting copper sulfide. Environ. Sci. Nano 2014, 1, 347–357. [Google Scholar] [CrossRef]
  30. Minguez-Bacho, I.; Courté, M.; Fan, H.J.; Fichou, D. Conformal Cu2S-coated Cu2O nanostructures grown by ion exchange reaction and their photoelectrochemical properties. Nanotechnology 2015, 26, 185401. [Google Scholar] [CrossRef]
  31. Gusak, A.M.; Tu, K.N. Interaction between the Kirkendall effect and the inverse Kirkendall effect in nanoscale particles. Acta Mater. 2009, 57, 3367–3373. [Google Scholar] [CrossRef]
  32. Rice, K.P.; Paterson, A.S.; Stoykovich, M.P. Nanoscale Kirkendall Effect and Oxidation Kinetics in Copper Nanocrystals Characterized by Real-Time, In Situ Optical Spectroscopy. Part. Part. Syst. Char. 2015, 32, 373–380. [Google Scholar] [CrossRef]
  33. Gong, S.; Wang, A.; Zhang, J.; Guan, J.; Han, N.; Chen, Y. Gram-scale synthesis of ultra-fine Cu2O for highly efficient ozone decomposition. RSC Adv. 2020, 10, 5212–5219. [Google Scholar] [CrossRef] [PubMed]
  34. Gong, S.Y.; Chen, J.Y.; Wu, X.F.; Han, N.; Chen, Y.F. In-situ synthesis of Cu2O/reduced graphene oxide composite as effective catalyst for ozone decomposition. Catal. Commun. 2018, 106, 25–29. [Google Scholar] [CrossRef]
  35. Zhang, H.; Deng, L.; Chen, J. How MoS2 assisted sulfur vacancies featured Cu2S in hollow Cu2S@MoS2 nanoboxes to activate H2O2 for efficient sulfadiazine degradation? Chem. Eng. J. 2022, 9, 10370. [Google Scholar] [CrossRef]
  36. Zhang, H.; Zeng, C.; Wu, H.; Yang, H.; Deng, L.; Zhou, L. ZIF-8 assisted synthesis of magnetic core-shell Fe3O4@CuS nanoparticles for efficient sulfadiazine degradation via H2O2 activation: Performance and mechanism. J. Colloid Interface Sci. 2021, 594, 502–512. [Google Scholar] [CrossRef]
  37. Zhang, H.; Zhou, C.; Zeng, H.; Deng, L.; Shi, Z. Can Cu2ZnSnS4 nanoparticles be used as heterogeneous catalysts for sulfadiazine degradation? J. Hazard. Mater. 2020, 395, 122613. [Google Scholar] [CrossRef]
  38. Wójcik, S.; Grzybek, G.; Stelmachowski, P.; Sojka, Z.; Kotarba, A. Bulk, Surface and Interface Promotion of Co3O4 for the Low-Temperature N2O Decomposition Catalysis. Catalysts 2020, 10, 41. [Google Scholar] [CrossRef]
  39. Cho, H.; Joo, H.; Kim, H.; Kim, J.-E.; Kang, K.-S.; Jung, H.; Yoon, J. Enhanced Photoelectrochemical Activity of TiO2 Nanotubes Decorated with Lanthanide Ions for Hydrogen Production. Catalysts 2022, 12, 866. [Google Scholar] [CrossRef]
  40. Zhang, L.; Huo, F.; Wang, A.; Chai, S.; Guan, J.; Fan, G.; Yang, W.; Ma, G.; Han, N.; Chen, Y. Coordination-Controlled Catalytic Activity of Cobalt Oxides for Ozone Decomposition. Inorg. Chem. 2023, 62, 9178–9189. [Google Scholar] [CrossRef]
  41. Li, W.; Gibbs, G.V.; Oyama, S.T. Mechanism of Ozone Decomposition on a Manganese Oxide Catalyst. 1. In Situ Raman Spectroscopy and Ab Initio Molecular Orbital Calculations. J. Am. Chem. Soc. 1998, 120, 9041–9046. [Google Scholar] [CrossRef]
  42. Li, X.; Ma, J.; Zhang, C.; Zhang, R.; He, H. Detrimental role of residual surface acid ions on ozone decomposition over Ce-modified γ-MnO2 under humid conditions. J. Environ. Sci. 2020, 91, 43–53. [Google Scholar] [CrossRef] [PubMed]
  43. Pietrogiacomi, D.; Sannino, D.; Magliano, A.; Ciambelli, P.; Tuti, S.; Indovina, V. The catalytic activity of CuSO4/ZrO2 for the selective catalytic reduction of NOx with NH3 in the presence of excess O2. Appl. Catal. B Environ. 2002, 36, 217–230. [Google Scholar] [CrossRef]
  44. Yang, S.; Chen, S.; Mosconi, E.; Fang, Y.; Xiao, X.; Wang, C.; Zhou, Y.; Yu, Z.; Zhao, J.; Gao, Y.; et al. Stabilizing halide perovskite surfaces for solar cell operation with wide-bandgap lead oxysalts. Science 2019, 365, 473–478. [Google Scholar] [CrossRef] [PubMed]
  45. Wahlqvist, M.; Shchukarev, A. XPS spectra and electronic structure of Group IA sulfates. J. Electron. Spectrosc. 2007, 156–158, 310–314. [Google Scholar] [CrossRef]
  46. Smirnov, M.Y.; Kalinkin, A.V.; Pashis, A.V.; Sorokin, A.M.; Noskov, A.S.; Kharas, K.C.; Bukhtiyarov, V.I. Interaction of Al2O3 and CeO2 Surfaces with SO2 and SO2 + O2 Studied by X-ray Photoelectron Spectroscopy. J. Phys. Chem. B. 2005, 109, 11712–11719. [Google Scholar] [CrossRef]
  47. Ma, G.; Tang, W.; Wang, A.; Zhang, L.; Guan, J.; Han, N.; Chen, Y. Heterojunctioned CuO/Cu2O catalyst for highly efficient ozone removal. J. Environ. Sci. 2023, 125, 340–348. [Google Scholar] [CrossRef]
  48. Huang, H.; Li, F.; Wang, H.; Zheng, X. The size controlled synthesis of Cu2S/P25 hetero junction solar-energy-materials and their applications in photocatalytic degradation of dyes. RSC Adv. 2017, 7, 50056–50063. [Google Scholar] [CrossRef]
  49. Fu, S.-Y.; Chang, H.-H.; Hsu, Y.-K.; Lin, Y.-G. Facile synthesis of Cu2S nanoarchitectures in application of surface enhanced Raman scattering. In Nanophotonic Materials XI; SPIE: Bellingham, WA, USA, 2014; pp. 25–28. [Google Scholar]
  50. Mai, H.E.; Fang, P.; Xie, G.Q.; Xie, Y.L.; Luo, M.F. Characterization of CuO Species in CuO/CeO2-Al2O3 Catalysts by In-situ XRD, Raman Spectroscopy and TPR. Acta Phys. Chim. Sin. 2005, 21, 997–1000. [Google Scholar] [CrossRef]
  51. Liu, J.; Gao, Z.; Han, H.; Wu, D.; Xu, F.; Wang, H.; Jiang, K. Mesoporous Cu2O submicro-spheres, facile synthesis and the selective adsorption properties. Chem. Eng. J. 2012, 185–186, 151–159. [Google Scholar] [CrossRef]
  52. Zhang, Y.; Wang, D.; Zhang, X.; Qu, F. Template-Free Synthesis of Porous Cu2O Nanospheres at Room Temperature and Investigation on Their Adsorption Property. J. Nanomater. 2013, 2013, 378919. [Google Scholar] [CrossRef]
Figure 1. Cu2S catalytic performance testing. (a) Cu2S with different morphologies for ozone decomposition at room temperature, 480,000 cm3·g−1·h−1 (50 mg catalyst, 400 sccm airflow), and 400 ppm dry O3; (b) durability testing of CuS-P at room temperature, 480,000 cm3·g−1·L−1 (50 mg catalyst, 400 sccm), and 200 ppm dry O3; (c) performance of Cu2S-P on catalytic decomposition of 200 ppm O3 at 480,000 cm3·g−1·h−1 and high humidity (>90% RH). All catalysts were loaded into a U-type reactor using a mixture of 50 mg granulated particles and 450 mg quartz sand of similar size. Using only 500 mg quartz sand as a blank control test showed no catalytic activity for ozone decomposition.
Figure 1. Cu2S catalytic performance testing. (a) Cu2S with different morphologies for ozone decomposition at room temperature, 480,000 cm3·g−1·h−1 (50 mg catalyst, 400 sccm airflow), and 400 ppm dry O3; (b) durability testing of CuS-P at room temperature, 480,000 cm3·g−1·L−1 (50 mg catalyst, 400 sccm), and 200 ppm dry O3; (c) performance of Cu2S-P on catalytic decomposition of 200 ppm O3 at 480,000 cm3·g−1·h−1 and high humidity (>90% RH). All catalysts were loaded into a U-type reactor using a mixture of 50 mg granulated particles and 450 mg quartz sand of similar size. Using only 500 mg quartz sand as a blank control test showed no catalytic activity for ozone decomposition.
Catalysts 14 00153 g001
Figure 2. XRD patterns of Cu2S samples.
Figure 2. XRD patterns of Cu2S samples.
Catalysts 14 00153 g002
Figure 3. SEM and TEM image of the Cu2S and template. (a) SEM of Cu2O sphere template, (b) SEM of Cu2O cube template, (c) SEM of Cu2O porous template, (d) TEM image of Cu2S-S, (e) TEM image of Cu2S-C, and (f) TEM image of Cu2S-P.
Figure 3. SEM and TEM image of the Cu2S and template. (a) SEM of Cu2O sphere template, (b) SEM of Cu2O cube template, (c) SEM of Cu2O porous template, (d) TEM image of Cu2S-S, (e) TEM image of Cu2S-C, and (f) TEM image of Cu2S-P.
Catalysts 14 00153 g003
Figure 4. Schematic diagram of structural evolution process: the internal Cu2O is rapidly transferred to the surface through several mass transfer channels, forming Cu2S hollow structures.
Figure 4. Schematic diagram of structural evolution process: the internal Cu2O is rapidly transferred to the surface through several mass transfer channels, forming Cu2S hollow structures.
Catalysts 14 00153 g004
Figure 5. EPR signal of S vacancies in different samples.
Figure 5. EPR signal of S vacancies in different samples.
Catalysts 14 00153 g005
Figure 6. UV photoelectron spectroscopy of samples with different morphologies. The instrument automatically calibrated the Fermi edge, and then corrected the abscissa using the emitted photon kinetic energy and bias voltage. The intercept can be equivalent to the work function (Φ).
Figure 6. UV photoelectron spectroscopy of samples with different morphologies. The instrument automatically calibrated the Fermi edge, and then corrected the abscissa using the emitted photon kinetic energy and bias voltage. The intercept can be equivalent to the work function (Φ).
Catalysts 14 00153 g006
Figure 7. In situ FTIR spectroscopy characterization of Cu2S in (a) O2/ozone mixed inlet and (b) stopped ozone airflow.
Figure 7. In situ FTIR spectroscopy characterization of Cu2S in (a) O2/ozone mixed inlet and (b) stopped ozone airflow.
Catalysts 14 00153 g007
Figure 8. XPS characterization of Cu2S-P deconvolution peak before and after reaction. (a) S 2p scan and (b) Cu 2p scan.
Figure 8. XPS characterization of Cu2S-P deconvolution peak before and after reaction. (a) S 2p scan and (b) Cu 2p scan.
Catalysts 14 00153 g008
Figure 9. Raman spectra of Cu2S-P, before and after the ozone decomposition.
Figure 9. Raman spectra of Cu2S-P, before and after the ozone decomposition.
Catalysts 14 00153 g009
Table 1. Specific surface area and pore structure parameters of Cu2S.
Table 1. Specific surface area and pore structure parameters of Cu2S.
SampleBET Surface Area (m2/g)Pore Size (Å)Pore Volume (cm3/g)
Cu2S-C11.4871.310.02047
Cu2S-S7.31111.750.02043
Cu2S-P14.6759.250.02174
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

Jiang, Y.; Xu, Y.; Zhang, Q.; Zhao, X.; Xiao, F.; Wang, X.; Ma, G. Templated Synthesis of Cu2S Hollow Structures for Highly Active Ozone Decomposition. Catalysts 2024, 14, 153. https://doi.org/10.3390/catal14020153

AMA Style

Jiang Y, Xu Y, Zhang Q, Zhao X, Xiao F, Wang X, Ma G. Templated Synthesis of Cu2S Hollow Structures for Highly Active Ozone Decomposition. Catalysts. 2024; 14(2):153. https://doi.org/10.3390/catal14020153

Chicago/Turabian Style

Jiang, Yishan, Ying Xu, Qichao Zhang, Xin Zhao, Feng Xiao, Xinbo Wang, and Guojun Ma. 2024. "Templated Synthesis of Cu2S Hollow Structures for Highly Active Ozone Decomposition" Catalysts 14, no. 2: 153. https://doi.org/10.3390/catal14020153

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