Next Article in Journal
The Effects of Oxygen Functional Groups on Graphene Oxide on the Efficient Adsorption of Radioactive Iodine
Previous Article in Journal
The Influence of the Drawing Process on the Mechanical Properties of TRIP Steel Wires with 0.4% C Content
Previous Article in Special Issue
Complex Catalytic Materials Based on the Perovskite-Type Structure for Energy and Environmental Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Particle Size and Crystal Surface of CeO2 on the Catalytic Combustion of Benzene

Zhejiang Key Laboratory of Alternative Technologies for Fine Chemicals Process, Shaoxing University, Shaoxing 312000, China
*
Author to whom correspondence should be addressed.
Materials 2020, 13(24), 5768; https://doi.org/10.3390/ma13245768
Submission received: 11 November 2020 / Revised: 12 December 2020 / Accepted: 12 December 2020 / Published: 17 December 2020
(This article belongs to the Special Issue Catalysts for Energy and Environmental Applications)

Abstract

:
In this study, three kinds of CeO2 were synthesized, and supported PdOx (x = 0,1) catalysts were prepared for benzene catalytic combustion. The samples were characterized by XRD, N2 adsorption/desorption, HRTEM, XPS and H2-TPR. The results show that three kinds of CeO2 with different structures can be formed by different preparation methods. This is mainly reflected in the differences in pore structure, particle size and crystal plane. CeO2-DC obtained from directly calcined Ce(NO3)3·6H2O had the largest pore volume and pore diameter and smallest particle size. CeO2-DC was mainly exposed to the (200) plane. Combined with the results of the ability test, it could be concluded that when Pd2+ and Pd0 exist at the same time, the activity increases with an increase in the proportion of Pd2+. Meanwhile, the structure of CeO2 affects the formation of oxygen vacancies, thereby affecting the adsorption and degradation of benzene. This article reveals that the particle size, crystal planes, oxygen vacancies and proportion of Pd2+ have a great impact on the catalytic combustion of benzene and allow a more comprehensive understanding of the structure–activity relationship, which can guide us to design high-efficiency catalysts targeted to obtain suitable CeO2-based catalysts for the catalytic combustion of benzene.

1. Introduction

Volatile organic compounds (VOCs) are a main component of air pollution and have been increasing rapidly in recent years. Most VOCs not only harm the environment but also threaten human health due to their toxicity and carcinogenicity [1,2,3]. Aromatic hydrocarbons, as a kind of VOCs, are considered to cause great harm to the environment and are usually toxic and carcinogenic [4,5]. Benzene is a kind of carcinogen, which can be determined to be carcinogenic. Toluene, ethylbenzene and xylene are all possible carcinogens [6]. Among the aromatic hydrocarbons, benzene, toluene, ethylbenzene and xylene (BTEX) constitute the majority of the total industrial emissions [7]. Therefore, it is very urgent to develop an effective and suitable method to reduce the total amount of VOCs. The main methods for dealing with VOCs include adsorption, membrane separation, absorption and oxidative approaches (thermal incineration, biological degradation, photocatalytic decomposition, nonthermal plasma oxidation and catalytic combustion) [8,9,10,11,12,13]. Catalytic combustion has become the most widely used technology because of its characteristics of high catalytic efficiency, low energy consumption and low secondary pollution [14,15,16].
Catalytic combustion is a typical gas–solid catalytic reaction. In the reaction, VOCs need to be adsorbed on the surface of the catalyst, and then a catalytic degradation reaction is carried out. Supports with a high specific surface area (SBET) and rich pore structure can improve the adsorption property and the catalytic performance of VOCs [8]. Thus, it is important to identify excellent supports for the catalytic combustion of VOCs. Common supports include γ-Al2O3, cordierite, molecular sieves and pillared clays [17,18,19,20,21]. These supports have large SBET and pore volume (VP) but have no activity or low activity. Due to the low activity of the support, only the proportion, type and loading method of the active component can be improved, which limits the improvement of catalyst performance in the catalytic combustion of VOCs. In recent years, some researchers have found that using active supports can further promote the performance of catalysts [22,23,24,25]. For example, Jablonska et al. synthesized a new HY zeolite that can completely catalyze methanol combustion at 500 °C. After loading Pd, it can completely convert methanol below 150 °C, which is greatly improved compared with Pd/Al2O3 [25]. Li et al. synthesized a type of CoAl mixed oxide that can completely catalyze toluene conversion at 345 °C. After loading Pt, its activity is greatly improved, and toluene can be completely burned at 230 °C [26]. Therefore, the use of active supports instead of traditional supports provides a development direction for future research.
Compared with other active supports, rare-earth elements have many advantages in catalytic materials, such as increasing oxygen storage and releasing capacity, adjusting catalytic active sites and improving the dispersion of active components and the thermal stability of catalysts. Therefore, rare-earth oxides, as active supports, have been widely used in VOCs catalytic combustion reactions [24,27]. Wu et al. synthesized a series of LaCo composite oxides for the catalytic combustion of toluene. By introducing La, the valence state of Co can be adjusted, and the formation of oxygen vacancies can be promoted, which causes the catalyst to have good catalytic performance for toluene combustion [28]. Wu et al. prepared LaFeO3 perovskite oxide for the catalytic combustion of toluene, which showed high catalytic activity [29]. Cerium oxide (CeO2), one of the most significant rare-earth oxides, has attracted much attention in recent decades. Many researchers have applied CeO2 as a catalyst in the catalytic combustion of VOCs [27,30,31,32]. Chen et al. synthesized a Ce metal–organic frameworks (MOF) catalyst which showed good catalytic activity for toluene oxidation [33]. Wang et al. synthesized a honeycomb CeO2 catalyst. It had great differences in morphology and reducibility and exhibited high activity in the oxidation of o-xylene [24]. However, CeO2 catalysts cannot achieve ideal catalytic performance due to the limitation of structure and physicochemical properties. Therefore, it is still necessary to use CeO2 as a support to prepare catalysts. For combustion of light hydrocarbons, several works have pointed out that noble metals as active components exhibited excellent performance of catalytic combustion. Guo et al. successfully prepared a 3D ordered mesoporous CeO2 catalyst loaded with biogenic Pd nanoparticles for the catalytic combustion of benzene. The experimental results show that the catalyst has good catalytic activity and can achieve 90% catalytic conversion at 187 °C [34]. Abbasi et al. prepared nanostructured Pt/Al2O3–CeO2 catalysts for catalytic combustion of VOCs, and they exhibited high catalytic activity [35]. Lei et al. synthesized a Pd/CeO2 catalyst for methane catalytic combustion. The experimental results show that the catalytic activity of the catalyst is greatly improved after loading Pd, which can completely catalyze combustion at 360 °C [31]. Nowadays, many researchers focus on the selection of active components and the role of rare-earth oxides as promoters. However, there is little research on the effects of pore structure, particle size, crystal surface and other physical indicators on the performance of catalysts and the structure–activity relationship.
In this paper, three kinds of CeO2 with different physicochemical properties were prepared and used to prepare Pd/CeO2 catalysts. Then, the influence of these properties on the catalytic combustion of benzene was discussed. XRD, N2 adsorption/desorption, HRTEM, XPS and H2-TPR tests were used to systematically analyze the structure–activity relationship.

2. Materials and Methods

2.1. Support Preparation

CeO2 (particle size: 20–50 nm, 99.5% purity) was purchased from Aladdin and used as received, which is denoted CeO2-P. Ce(NO3)3·6H2O (99.5% purity, Aladdin) was directly calcined at 500 °C for 2 h to obtain CeO2, which is called CeO2-DC. Another CeO2 was made by thermal decomposition of Ce metal‑organic frameworks (MOF), which is named CeO2-MOF. The preparation method of Ce-MOF has been described in our previous articles [36].

2.2. Catalyst Preparation

The catalysts loaded with Pd nanoparticles (0.2 wt.%) were synthesized by high-temperature liquid reduction [37]. First, 2.0 g of CeO2 and 100 mL of ethylene glycol (AR, 98%, Aladdin, Xi’an, China) were added to a three-necked flask with stirring for 30 min, and then 0.4 mL of H2PdCl4 (10 mg/mL) was added into the three-necked flask and stirred for 12 h. Then, the pH was controlled at 11.0. The reaction was heated to 165 °C for 3 h, and then the temperature dropped to room temperature. The product was washed with distilled water twice, then washed with ethanol (AR, 99.7%, Aladdin, Xi’an, China) once and then dried for 12 h at 60 °C. After that, the product was calcined in a muffle furnace (Shanghai Pudong Yuexinxue Instrument Factory, Shanghai, China) at 500 °C for 2 h to obtain a 0.2% Pd/CeO2 catalyst. Pd/CeO2-P, Pd/CeO2-DC and Pd/CeO2-MOF were obtained by the above method using CeO2-P, CeO2-DC and CeO2-MOF, respectively.

2.3. Catalytic Performance Evaluation

The activity of the catalysts was evaluated on a WFS-3010 microreactor (Xianquan, Tianjin, China) with a benzene concentration of 1000 ppm and a space velocity of 20,000 h−1. An amount of 200 mg catalyst (40–60 mesh) was loaded in the quartz reactor and pretreated at 400 °C in air for 0.5 h. After that, 1000 ppm benzene was added for the activity test. Online gas chromatography (Shimadzu, GC-14C, Kyoto, Japan) with a hydrogen flame ionization detector (FID) was used to detect the concentration of the organic substance entering the chromatograph. The chromatographic working conditions were as follows. The temperatures of the column and vaporization chamber were 80 and 120 °C, respectively. An N2000 online chromatography workstation (Saizhi Technology Co., Ltd, Hangzhou, China) was used to record data. The durability of catalysts was tested under the same working conditions of the activity evaluation, and the catalytic time was as long as 100 h.

2.4. Catalyst Characterization

The structure of catalysts and supports was characterized using XRD (PANalytical, Almelo, The Netherlands, Cu Kα radiation, 40 kV, 40 mA, speed of 0.02°/s). The particle size of samples was calculated from the XRD data using the Scherrer formula (Equation (1)):
D = K Υ B cos θ
where:
  • D = Average thickness of particle perpendicular to crystal plane (nm);
  • K = Scherrer constant;
  • γ = X-ray wavelength (0.154 nm);
  • B = Full width at half maxima (FWHM);
  • θ = Diffraction angle.
A Tristar II 3020 apparatus (Micromeritics Company, Norcross, GA, USA) was used to determine the nitrogen adsorption and desorption isotherms at −195.8 °C. The Brunauer-Emmett-Teller (BET) model was used to measure the SBET. Meanwhile, the Barrett-Joyner-Halenda (BJH) method was used to calculate the pore size and average pore diameter. Before the adsorption process, the samples were kept under vacuum at 200 °C for 4 h.
The morphology of the supports and the catalysts was obtained using high-resolution transmission electron microscopy (HRTEM, JEOL-2010F, Tokyo, Japan) operated at 200 kV. The chemical composition of the Pd/CeO2 catalyst was detected using energy-dispersive X-ray spectroscopy (EDS) with an Oxford INCA instrument (OXFORD instruments, Abingdon, UK).
The valence states and the elemental proportions of samples were analyzed by X-ray photoelectron spectroscopy (XPS) (Thermo Fisher Scientific, Waltham, MA, USA) on a Thermo ESCALAB 250 with Al Kα (hν = 1486.8 eV) as the excitation source. C1s was used as the internal reference to calibrate electron energies.
Temperature-programmed reduction (H2-TPR) experiments were carried out in a CHEMBET-3000 apparatus prior to the measurement, and 200 mg catalyst was put into the quartz reactor and pretreated in air at 300 °C for 0.5 h. After being cooled to room temperature, the catalyst was raised from room temperature to 900 °C under H2 flow (5 vol.% in Ar, 40 mL/min) at a heating rate of 15 °C/min. A gas chromatograph (TCD) (Shimadzu, GC-14C, Kyoto, Japan) was used to record and analyze data.

3. Results and Discussion

3.1. Catalytic Activity Evaluation

The catalytic activity of the CeO2 supports and Pd/CeO2 catalysts is shown in Figure 1. The results show that the performance of three different CeO2 samples was obviously different, which indicated that the ability of CeO2 was closely related to the preparation method. The catalytic combustion activity of benzene followed the sequence CeO2-DC > CeO2-MOF > CeO2-P. The conversion rate of CeO2-DC for benzene reached 91.7% at 320 °C. After loading Pd, the catalytic combustion ability of benzene was significantly improved. All the catalysts showed better catalytic ability than single CeO2, indicating that the active component of PdOx plays an important role in benzene catalytic combustion. Catalytic activity decreases in the following order: Pd/CeO2-DC > Pd/CeO2-MOF > Pd/CeO2-P. The complete conversion temperatures of benzene using Pd/CeO2-MOF and Pd/CeO2-P were 280 and 300 °C, respectively. Pd/CeO2-DC showed the best performance of catalytic combustion, which could completely convert benzene at 260 °C. The catalytic activity data of the samples and related catalysts for benzene catalytic combustion are compared in Table S1. Although the Pd/CeO2 catalyst synthesized by Guo et al. can achieve 90% benzene catalytic combustion at 187 °C, they used a higher content of Pd (2 wt.%) than Pd/CeO2-DC [34]. Therefore, considering the comprehensive cost, Pd/CeO2-DC has high application value.

3.2. Durability Test

The temperature of the conversion rate at 52% was selected for the life test of the catalyst. Figure 2 shows the results of the durability test of Pd/CeO2-DC at 190 °C. After 100 h of reaction, the activity of the catalyst remained at approximately 49%, with no obvious decrease. It exhibited exceptional stable catalytic activity. Industrial chlorine-containing VOCs and water were widely distributed in waste gases. Figure 2 shows that when 3 vol.% water was added, its catalytic activity decreased because of competitive adsorption [38]. After introducing chlorobenzene at 100 ppm into the system, the catalytic activity further decreased. However, after water vapor and chlorobenzene were removed, the activity of the catalyst returned to its original level. The results show that the Pd/CeO2-DC catalyst had a certain ability to resist chlorine poisoning that may be due to the interaction between PdOx and CeO2-DC, which can timely release the combustion products of chlorobenzene and prevent chlorine poisoning. Meanwhile, it exhibited good resistance to moisture conditions. In addition, the catalyst maintained its catalytic activity over a long reaction time of 100 h.

3.3. XRD Analysis

Figure 3 (XRD patterns of supports and catalysts) shows that the patterns of samples have characteristic diffraction peaks of CeO2. All samples exhibited several peaks at 28.7°, 33.2°, 47.7°, 56.6°, 59.4°, 69.7°, 77.1° and 79.5°, pointing to the (111), (200), (220), (311), (222), (400), (331) and (420) crystal planes, respectively (JCPDS No. 34-0394) [26]. The XRD spectra of the precursor Ce-MOF are shown in Figure S1. From Figure 3, it can be found that the intensity and FWHM of the diffraction peak are different and decrease in the following order: CeO2-P > CeO2-MOF > CeO2-DC. According to the Scherrer formula, the average particle size of the samples can be calculated as follows: CeO2-P (34 nm) > CeO2-MOF (11 nm) > CeO2-DC (8 nm). This result indicates that CeO2-DC prepared by direct calcining had the smallest particle size. As shown in Figure 3, it can be concluded that the crystallinity of CeO2-P is the highest and that of CeO2-DC is the lowest. Combined with their particle size, it can be inferred that the smaller the particle size of CeO2, the poorer the crystallinity. It is well known that the lattice defects of CeO2 increase with decreasing particle size, so the smaller particle size of CeO2 will lead to the formation of surface lattice defects and the generation of reactive oxygen species [33]. However, no diffraction peak for the PdOx phase could be found, which may be attributed to the low loading content and the high dispersion of Pd.

3.4. HRTEM Analysis

Figure 4 shows the morphology and crystal plane structure of CeO2 supports and their Pd-based catalysts. Figure 4a–c show that all CeO2 synthesized by different preparation methods have the morphology of nanoparticles and show an irregular shape. The particle sizes of CeO2-P, CeO2-DC and CeO2-MOF were 30–60, 3–10 and 10–30 nm, respectively. It was found that the particle size of CeO2-DC was significantly smaller than that of CeO2-P and CeO2-MOF, which was consistent with the results of XRD analysis. The fine structure of three different Pd/CeO2 was observed when the resolution was further improved. We can observe the lattice fringes of PdOx and CeO2, as shown in Figure 4d–f. The lattice fringe spacing was calculated by Image J digital micrograph software. For Pd/CeO2-P, the lattice fringe spacing is 0.215 nm, corresponding to the PdO (110) crystal plane (JCPDS No. 41-1107), and CeO2-P mainly exposes the (311) and (200) crystal planes. For Pd/CeO2-DC and Pd/CeO2-MOF, the lattice fringe spacing is 0.225 nm, corresponding to the Pd (111) crystal plane (JCPDS No. 46-1043); CeO2-DC mainly exposes (200) crystal planes with a lattice fringe spacing of 0.270 nm; and CeO2-MOF has two crystal plane spacings of 0.191 and 0.270 nm, corresponding to the (220) and (200) crystal planes. The exposed crystal surface may affect the formation of oxygen vacancies, and different Pd/CeO2 should have different chemical properties [33]. Through catalytic activity testing, we can assume that the (200) crystal plane of CeO2 may play an important role in benzene catalytic combustion. The morphology of the recovered catalyst is shown in Figure S2. It can be found that the morphology of the catalyst has no obvious change after use, indicating that the catalyst has good stability.
In addition, the EDS spectrum of Pd/CeO2-DC had Pd, O and Ce signals, as shown in Figure 4g, indicating that the active components of PdOx have been successfully loaded and highly dispersed in CeO2-DC and the particle size of PdOx nanoparticles is about 2–3 nm.

3.5. N2 Adsorption/Desorption Analysis

The pore size distribution and SBET of the catalyst have a great influence on the activity of catalytic combustion. It can be measured by N2 adsorption/desorption, and the results are shown in Figure 5. It can be clearly seen in Figure 5a that all the samples showed a type IV isotherm, which indicates that there are mesoporous structures in the materials [39,40]. CeO2-P and CeO2-MOF showed a H3 hysteresis loop appearing at P/P0 = 0.4–1.0, indicating that there are slit-shaped pores in the sample [41]. Additionally, CeO2-DC exhibited a H1 hysteresis loop, indicating that there is a cylindrical order in the sample. The pore size distribution curve is shown in Figure 5b. From the pore size distribution curves, it can be clearly found that the average pore size of CeO2-DC is approximately 11.0–12.7 nm and the VP is 0.23 cm3/g. Meanwhile, the pore size does not change after loading PdOx nanoparticles, but the VP decreases slightly. Table 1 summarizes the physical properties of samples according to the N2 adsorption/desorption isotherms. From Table 1, we can conclude that the SBET of three different CeO2 is arranged as follows: CeO2-MOF > CeO2-DC > CeO2-P. After the introduction of PdOx nanoparticles, VP decreased in varying degrees, and some catalysts formed a microporous structure, which may be due to the blockage of the pore channels by PdOx nanoparticles. However, the SBET of CeO2-DC increases after loading with PdOx nanoparticles, which may be due to the fact that PdOx nanoparticles are mainly distributed on the surface of CeO2-DC. The results of EDS in Figure 4g also show that PdOx is mainly distributed on the surface of CeO2-DC and highly dispersed. Pd/CeO2-DC has the highest SBET (80.4 m2/g) and the largest VP (0.21 cm3/g) among all catalysts. It is well known that higher SBET and VP can provide more active sites and promote catalytic activity. The results are consistent with the experimental results of the catalytic combustion performance.

3.6. XPS Analysis

The valence states of the elements on the surfaces of materials can be investigated using the XPS technique. Figure 6a shows the Pd 3d spectra of the catalysts. It was reported that the complex spectrum of Pd 3d can be decomposed into four peaks associated with the two spin orbitals. The peaks of 339.3–342.9 and 335.3–337.1 eV can be allocated to Pd 3d3/2 and Pd 3d5/2, respectively [30]. It is obvious that Pd0 and Pd2+ were detected in the prepared Pd/CeO2-DC and Pd/CeO2-MOF catalysts, but only Pd2+ was detected in the prepared Pd/CeO2-P catalyst. The percentage of Pd2+ species in the catalyst was obtained by calculating the fitted area of Pd2+/(Pd2+ + Pd0). The concentration of Pd2+ in catalysts is reduced in the following order: Pd/CeO2-P (100%) > Pd/CeO2-DC (71.1%) > Pd/CeO2-MOF (57.1%). Many researchers found that Pd2+ plays an important role in hydrocarbon oxidation, indicating that Pd2+ is more active than Pd0 in the reaction [42]. However, it is worth noting that compared with other catalysts, Pd/CeO2-P exhibited relatively low activity, indicating that single Pd2+ has a negative effect on benzene catalytic combustion [43].
Figure 6b shows the Ce 3d spectra of the catalysts. Ce mainly exists in the form of Ce3+ and Ce4+ [44]. According to the equation of Ce3+/(Ce3+ + Ce4+), the ratio of Ce3+ on the catalyst was calculated based on the peak area. To compare the concentration of Ce3+ on the supports and catalysts, the concentration of Ce3+ on the CeO2 support was calculated. The results show that the proportion of Ce3+/(Ce3+ + Ce4+) on supports is reduced in the following order: CeO2-DC (22.5%) > CeO2-MOF (21.9%) > CeO2-P (17.8%). After loading PdOx, the concentration of Ce3+ on the CeO2 surface increased, which indicates that there is a strong interaction between PdOx and CeO2. Pd/CeO2-DC had the highest proportion of Ce3+/(Ce3+ + Ce4+). The proportion of Ce3+/(Ce3+ + Ce4+) decreased in the order: Pd/CeO2-DC (24.8%) > Pd/CeO2-MOF (23.1%) > Pd/CeO2-P (18.1%). This was consistent with the higher activity of Pd/CeO2-DC. The corresponding results are listed in Table 2. It has been reported that Ce3+ can promote the formation of oxygen vacancies [35]. Therefore, a higher ratio of Ce3+ means more oxygen vacancies on the surface of the catalyst, which can promote the interaction between PdOx and CeO2, which is crucial for redox performance and catalytic activity. It is well known that CeO2 has good oxygen storage performance and oxygen in the CeO2 phase can be moved. Due to the strong interaction between PdOx and CeO2, Pd/CeO2-DC can transfer oxygen from CeO2 to PdOx, resulting in the appearance of more Ce3+ and Pd2+. Combined with the data of activity evaluation, it can be found that the difference in activity of Pd/CeO2 catalysts with the same preparation method and the same Pd loading amount is mainly due to the different structure of CeO2. CeO2 prepared by different methods has different particle sizes and exposed crystal planes, which leads to different binding abilities between PdOx and CeO2, resulting in the change of the valence state of Pd and CeO2 on the catalyst surface, thus affecting the catalytic activity of the catalysts.

3.7. H2-TPR Analysis

To study the redox ability of different CeO2 and their Pd/CeO2 catalysts, H2-TPR was used. As shown in Figure 7, CeO2 exhibits one or two reduction peaks at temperatures below 600 °C, which are attributed to the reduction of CeO2 surface oxygen and subsurface oxygen. When the temperature is higher than 800 °C, the CeO2 lattice oxygen is reduced. The total hydrogen consumption follows the sequence CeO2-DC > CeO2-MOF > CeO2-P. The results show that CeO2-DC has the best redox capacity, which is consistent with the catalytic activity. After loading PdOx, a new peak appeared near 100 °C, and the peak disappeared between 423 and 523 °C, which indicates that there is a strong interaction between PdOx and CeO2 which promotes the reduction of CeO2. The peak at low temperature is caused by the reduction of PdOx and the co-reduction of oxygen adsorbed on the CeO2 surface. Obviously, although Pd/CeO2-DC shows a small acromion below 100 °C, neither Pd/CeO2-P nor Pd/CeO2-MOF has an acromion, which can be attributed to the strong binding with Pd-O-Ce, resulting in a higher PdOx reduction temperature. Therefore, the first reduction temperature of Pd/CeO2-DC is the lowest (70 °C); that is, the PdOx species on CeO2-DC is relatively easy to reduce, which leads to maintaining the palladium metal state and catalytic oxidation of benzene activity. It has been reported that the reducibility of catalysts is closely related to their catalytic activity, and catalysts with lower reduction temperatures usually show higher catalytic activity.

4. Conclusions

In this study, three kinds of CeO2 were obtained by purchase, calcining of Ce(NO3)3·6H2O and thermal decomposition of Ce-MOF, which exhibited different pore structures, particle sizes and crystal planes. CeO2-DC had the largest pore volume and the smallest particle size, which promotes the formation of oxygen vacancies. Meanwhile, the Pd/CeO2 catalysts were synthesized by high-temperature liquid reduction, and Pd/CeO2-DC showed the best catalytic performance, excellent durability and resistance to poisoning in the catalytic combustion of benzene due to the good properties of CeO2-DC, which could completely combust benzene at 260 °C. Moreover, the research indicates that the difference in binding ability between the exposed surfaces of CeO2 and Pd can affect catalytic activity and that CeO2 with the (200) crystal plane may play an important role in the catalytic combustion of benzene. Furthermore, the increase in the Pd2+ proportion can promote the catalytic activity when Pd0 and Pd2+ exist at the same time. These characteristics indicate that the catalyst has great potential in industrial applications. However, we must realize that although this catalyst has shown high catalytic activity in the laboratory, the industrial environment is more complex and changeable. Therefore, further research in environments with high concentrations and multiple mixed VOCs is urgently required.

Supplementary Materials

The following are available online at https://www.mdpi.com/1996-1944/13/24/5768/s1, Figure S1: The XRD diffraction pattern of the precursor Ce-MOF-120 of CeO2-MOF, Figure S2: The morphology of recovered Pd/CeO2-DC catalyst, Table S1: The catalytic activity data of the samples and related catalysts for benzene catalytic combustion.

Author Contributions

Data curation, Z.W., Z.C. and J.Z.; formal analysis, Z.C.; investigation, J.Z.; writing—original draft, Z.W.; writing—review and editing, S.Z.; conceptualization, S.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This paper was supported by the National Natural Science Foundation of China (No. 21577094), Zhejiang Public Welfare Technology Research Project (No. LGG19B070003) and the Foundation of Science and Technology of Shaoxing City (No. 2018C10019).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Oztürk, N.; Ergenekon, P.; Seçkin, G.O.; Bayır, S. Spatial Distribution and Temporal Trends of VOCs in a Highly Industrialized Town in Turkey. Bull. Environ. Contam. Toxicol. 2015, 94, 653–660. [Google Scholar] [CrossRef] [PubMed]
  2. Sofuoğlu, S.C.; Aslan, G.; Inal, F.; Sofuoglu, A. An Assessment of Indoor Air Concentrations and Health Risks of Volatile Organic Compounds in Three Primary Schools. Int. J. Hyg. Environ. Health 2011, 214, 36–46. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Montero-Montoya, R.; López-Vargas, R.; Arellano-Aguilar, O. Volatile Organic Compounds in Air: Sources, Distribution, Exposure and Associated Illnesses in Children. Ann. Glob. Health 2018, 84, 225–238. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Kim, S.C.; Shim, W.G. Catalytic Combustion of VOCs over a Series of Manganese Oxide Catalysts. Appl. Catal. B Environ. 2010, 98, 180–185. [Google Scholar] [CrossRef]
  5. Liu, F.F.; Peng, C.; Ng, J.C. BTEX In Vitro Exposure Tool Using Human Lung Cells: Trips and Gains. Chemosphere 2015, 128, 321–326. [Google Scholar] [CrossRef]
  6. Lim, S.K.; Shin, H.S.; Yoon, K.S.; Kwack, S.J.; Um, Y.M.; Hyeon, J.H.; Kwak, H.M.; Kim, J.Y.; Kim, T.H.; Kim, Y.J.; et al. Risk Assessment of Volatile Organic Compounds Benzene, Toluene, Ethylbenzene, and Xylene (BTEX) in Consumer Products. J. Toxicol. Environ. Health Part A 2014, 77, 1502–1521. [Google Scholar] [CrossRef]
  7. Liaud, C.; Nguyen, T.N.; Nasreddine, R.; Le Calvé, S. Experimental Performances Study of a Transportable GC-PID and Two Thermo-Desorption Based Methods Coupled to FID and MS Detection to Assess BTEX Exposure at Sub-ppb Level in Air. Talanta 2014, 127, 33–42. [Google Scholar] [CrossRef]
  8. Shafiei, M.; Alivand, M.S.; Rashidi, A.; Samimi, A.; Mohebbi-Kalhori, D. Synthesis and Adsorption Performance of a Modified Micro-Mesoporous MIL-101(Cr) for VOCs Removal at Ambient Conditions. Chem. Eng. J. 2018, 341, 164–174. [Google Scholar] [CrossRef]
  9. Dosa, M.; Piumetti, M.; Bensaid, S.; Andana, T.; Galletti, C.; Fino, D.; Russo, N. Photocatalytic Abatement of Volatile Organic Compounds by TiO2 Nanoparticles Doped with Either Phosphorous or Zirconium. Materials 2019, 12, 2121. [Google Scholar] [CrossRef] [Green Version]
  10. Li, J.; Zhang, H.; Ying, D.; Wang, Y.; Sun, T.; Jia, J. In Plasma Catalytic Oxidation of Toluene Using Monolith CuO Foam as a Catalyst in a Wedged High Voltage Electrode Dielectric Barrier Discharge Reactor: Influence of Reaction Parameters and Byproduct Control. Int. J. Environ. Res. Public Health 2019, 16, 711. [Google Scholar] [CrossRef] [Green Version]
  11. Shamskar, K.R.; Rashidi, A.; Azar, P.A.; Yousefi, M.; Baniyaghoob, S. Synthesis of Graphene by in situ Catalytic Chemical Vapor Deposition of Reed as a Carbon Source for VOC Adsorption. Environ. Sci. Pollut. Res. 2019, 26, 3643–3650. [Google Scholar] [CrossRef] [PubMed]
  12. Ghosh, M.; Liu, J.; Chuang, S.S.C.; Jana, S.C. Fabrication of Hierarchical V2O5 Nanorods on TiO2 Nanofibers and Their Enhanced Photocatalytic Activity under Visible Light. ChemCatChem 2018, 10, 3305–3318. [Google Scholar] [CrossRef]
  13. Li, J.-J.; Yu, E.-Q.; Cai, S.-C.; Chen, X.; Chen, J.; Jia, H.; Xu, Y.-J. Noble Metal Free, CeO2/LaMnO3 Hybrid Achieving Efficient Photo-Thermal Catalytic Decomposition of Volatile Organic Compounds under IR Light. Appl. Catal. B Environ. 2019, 240, 141–152. [Google Scholar] [CrossRef]
  14. Zhang, Z.X.; Jiang, Z.; Shangguan, W.F. Low-Temperature Catalysis for VOCs Removal in Technology and Application: A State-of-the-Art Review. Catal. Today 2016, 264, 270–278. [Google Scholar] [CrossRef]
  15. Shah, K.W.; Li, W. A Review on Catalytic Nanomaterials for Volatile Organic Compounds VOC Removal and Their Applications for Healthy Buildings. Nanomaterials 2019, 9, 910. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Alifanti, M.; Florea, M.; Parvulescu, V.I. Ceria-Based Oxides as Supports for LaCoO3 Perovskite; Catalysts for Total Oxidation of VOC. Appl. Catal. B Environ. 2007, 70, 400–405. [Google Scholar] [CrossRef]
  17. Silva, B.; Figueiredo, H.; Santos, V.P.; Pereira, M.F.; Figueiredo, J.L.; Lewandowska, A.E.; Bañares, M.A.; Neves, I.C.; Tavares, T. Reutilization of Cr-Y Zeolite Obtained by Biosorption in the Catalytic Oxidation of Volatile Organic Compounds. J. Hazard. Mater. 2011, 192, 545–553. [Google Scholar] [CrossRef] [Green Version]
  18. Tsoncheva, T.; Ivanova, L.; Rosenholm, J.; Linden, M. Cobalt Oxide Species Supported on SBA-15, KIT-5 and KIT-6 Mesoporous Silicas for Ethyl Acetate Total Oxidation. Appl. Catal. B Environ. 2009, 89, 365–374. [Google Scholar] [CrossRef]
  19. Zhao, W.; Liu, Y.H.; Wang, L.N.; Chu, J.L.; Qu, J.K.; Hao, Z.P.; Qi, T. Catalytic Combustion of Benzene on the Pd/Nanosize Al-HMS. Microporous Mesoporous Mater. 2011, 138, 215–220. [Google Scholar] [CrossRef]
  20. Zhao, S.; Li, K.; Jiang, S.; Lib, J. Pd–Co Based Spinel Oxides Derived from Pd Nanoparticles Immobilized on Layered Double Hydroxides for Toluene Combustion. Appl. Catal. B Environ. 2016, 181, 236–248. [Google Scholar] [CrossRef]
  21. Huang, Q.; Zuo, S.; Zhou, R. Catalytic Performance of Pillared Interlayered Clays (PILCs) Supported CrCe Catalysts for Deep Oxidation of Nitrogen-Containing VOCs. Appl. Catal. B Environ. 2010, 95, 327–334. [Google Scholar] [CrossRef]
  22. Deng, W.; Tang, Q.; Huang, S.; Zhang, L.; Jia, Z.; Guo, L. Low Temperature Catalytic Combustion of Chlorobenzene over Cobalt Based Mixed Oxides Derived from Layered Double Hydroxides. Appl. Catal. B Environ. 2020, 278, 119336. [Google Scholar] [CrossRef]
  23. Huang, H.; Gu, Y.F.; Zhao, J.; Wang, X.Y. Catalytic Combustion of Chlorobenzene over VOx/CeO2 Catalysts. J. Catal. 2015, 326, 54–68. [Google Scholar] [CrossRef]
  24. Wang, H.; Liu, M.; Guo, S.; Wang, Y.; Han, X.; Bai, Y. Efficient Oxidation of O-Xylene over CeO2 Catalyst Prepared from a Ce-MOF Template: The Promotion of K+ Embedding Substitution. Mol. Catal. 2017, 436, 120–127. [Google Scholar] [CrossRef]
  25. Jabłońska, M.; Król, A.; Kukulska-Zając, E.; Tarach, K.; Girman, V.; Chmielarz, L.; Góra-Marek, K. Zeolites Y Modified with Palladium as Effective Catalysts for Low-Temperature Methanol Incineration. Appl. Catal. B Environ. 2015, 166, 353–365. [Google Scholar] [CrossRef]
  26. Li, P.; He, C.; Cheng, J.; Ma, C.Y.; Dou, B.J.; Hao, Z. Catalytic Oxidation of Toluene over Pd/Co3AlO Catalysts Derived from Hydrotalcite-Like Compounds: Effects of Preparation Methods. Appl. Catal. B Environ. 2011, 101, 570–579. [Google Scholar] [CrossRef]
  27. Zeng, Y.; Wang, Y.; Song, F.; Zhang, S.; Zhong, Q. The Effect of CuO Loading on Different Method Prepared CeO2 Catalyst for Toluene Oxidation. Sci. Total. Environ. 2020, 712, 135635. [Google Scholar] [CrossRef]
  28. Wu, M.; Chen, S.; Soomro, A.; Ma, S.; Zhu, M.; Hua, X.; Xiang, W. Investigation of Synergistic Effects and High Performance of La-Co Composite Oxides for Toluene Catalytic Oxidation at Low Temperature. Environ. Sci. Pollut. Res. Int. 2019, 26, 12123–12135. [Google Scholar] [CrossRef]
  29. Wu, M.; Chen, S.; Xiang, W. Oxygen Vacancy Induced Performance Enhancement of Toluene Catalytic Oxidation Using LaFeO3 Perovskite Oxides. Chem. Eng. J. 2020, 387, 124101. [Google Scholar] [CrossRef]
  30. Yan, S.; Deng, J.; Zhao, M.; Cheng, T.; Wang, J.; Jiao, Y.; Chen, Y. Preparation of Ce0.5Zr0.5O2-Al2O3 with High-Temperature Sintering Resistance and Its Supported Pd-Only Three-Way Catalyst. J. Mater. Sci. 2019, 54, 2796–2813. [Google Scholar] [CrossRef]
  31. Lei, Y.; Li, W.; Liu, Q.; Lin, Q.; Zheng, X.; Huang, Q.; Guan, S.; Wang, X.; Wang, C.; Li, F. Typical Crystal Face Effects of Different Morphology Ceria on the Activity of Pd/CeO2 Catalysts for Lean Methane Combustion. Fuel 2018, 233, 10–20. [Google Scholar] [CrossRef]
  32. Tan, H.; Wang, J.; Yu, S.; Zhou, K. Support Morphology-Dependent Catalytic Activity of Pd/CeO2 for Formaldehyde Oxidation. Environ. Sci. Technol. 2015, 49, 8675–8682. [Google Scholar] [CrossRef] [PubMed]
  33. Chen, X.; Yu, E.; Cai, S.; Jia, H.; Chen, J.; Liang, P. In Situ Pyrolysis of Ce-MOF to Prepare CeO2 Catalyst with Obviously Improved Catalytic Performance for Toluene Combustion. Chem. Eng. J. 2018, 344, 469–479. [Google Scholar] [CrossRef]
  34. Guo, Y.; Gao, Y.; Li, X.; Zhuang, G.; Wang, K.; Zhuang, G.; Sun, D.; Zheng, Y.; Li, Q. Catalytic Benzene Oxidation by Biogenic Pd Nanoparticles over 3D-Ordered Mesoporous CeO2. Chem. Eng. J. 2019, 362, 41–52. [Google Scholar] [CrossRef]
  35. Abbasi, Z.; Haghighi, M.; Fatehifar, E.; Saedy, S. Synthesis and Physicochemical Characterizations of Nanostructured Pt/Al2O3-CeO2 Catalysts for Total Oxidation of VOCs. J. Hazard. Mater. 2011, 186, 1445–1454. [Google Scholar] [CrossRef]
  36. Zheng, J.; Wang, Z.; Chen, Z.; Zuo, S. Mechanism of CeO2 Synthesized by Thermal Decomposition of Ce-MOF and Its Performance of Benzene Catalytic Combustion. J. Rare Earths 2020. [Google Scholar] [CrossRef]
  37. Chen, Z.; Li, J.; Yang, P.; Cheng, Z.; Li, J.; Zuo, S. Ce-modified Mesoporous γ-Al2O3 Supported Pd-Pt Nanoparticle Catalysts and Their Structure-Function Relationship in Complete Benzene Oxidation. Chem. Eng. J. 2019, 356, 255–261. [Google Scholar] [CrossRef]
  38. Schmal, M.; Aranda, D.A.G.; Noronha, F.B.; Guimaraes, A.L.; Monteiro, R.D.S. Oxidation and Reduction Effects of Propane-Oxygen on Pd-Chlorine/Alumina Catalysts. Catal. Lett. 2000, 64, 163–169. [Google Scholar] [CrossRef]
  39. Saikia, H.; Hazarika, K.K.; Chutia, B.; Choudhury, B.; Bharali, P. A Simple Chemical Route toward High Surface Area CeO2 Nanoparticles Displaying Remarkable Radical Scavenging Activity. ChemistrySelect 2017, 2, 3369–3375. [Google Scholar] [CrossRef]
  40. Ambroz, F.; Macdonald, T.J.; Martis, V.; Parkin, I.P. Evaluation of the BET Theory for the Characterization of Meso and Microporous MOFs. Small Methods 2018, 2, 2. [Google Scholar] [CrossRef] [Green Version]
  41. Ghosh, M.; Lohrasbi, M.; Chuang, S.S.C.; Jana, S.C. Mesoporous Titanium Dioxide Nanofibers with a Significantly Enhanced Photocatalytic Activity. ChemCatChem 2016, 8, 2525–2535. [Google Scholar] [CrossRef]
  42. Shim, W.-G.; Lee, J.-W.; Kim, S.C. Analysis of Catalytic Oxidation of Aromatic Hydrocarbons over Supported Palladium Catalyst with Different Pretreatments Based on Heterogeneous Adsorption Properties. Appl. Catal. B Environ. 2008, 84, 133–141. [Google Scholar] [CrossRef]
  43. Okumura, K.; Kobayashi, T.; Tanaka, H.; Niwa, M. Toluene Combustion over Palladium Supported on Various Metal Oxide Supports. Appl. Catal. B Environ. 2003, 44, 325–331. [Google Scholar] [CrossRef]
  44. Bêche, E.; Charvin, P.; Perarnau, D.; Abanades, S.; Flamant, G. Ce 3d XPS Investigation of Cerium Oxides and Mixed Cerium Oxide (CexTiyOz). Surf. Interface Anal. 2008, 40, 264–267. [Google Scholar] [CrossRef]
Figure 1. Effect of different CeO2 and Pd/CeO2 on catalytic combustion of benzene.
Figure 1. Effect of different CeO2 and Pd/CeO2 on catalytic combustion of benzene.
Materials 13 05768 g001
Figure 2. Durability test for catalytic combustion of benzene over the 0.2% Pd/CeO2-DC catalyst.
Figure 2. Durability test for catalytic combustion of benzene over the 0.2% Pd/CeO2-DC catalyst.
Materials 13 05768 g002
Figure 3. XRD patterns of the samples.
Figure 3. XRD patterns of the samples.
Materials 13 05768 g003
Figure 4. HRTEM images: (a) CeO2-P; (b) CeO2-DC; (c) CeO2-MOF; (d) Pd/CeO2-P; (e) Pd/CeO2-DC; (f) Pd/CeO2-MOF; (g) EDS mapping images of Pd/CeO2-DC (The red, green and yellow dots represent O, Ce and Pd elements, respectively).
Figure 4. HRTEM images: (a) CeO2-P; (b) CeO2-DC; (c) CeO2-MOF; (d) Pd/CeO2-P; (e) Pd/CeO2-DC; (f) Pd/CeO2-MOF; (g) EDS mapping images of Pd/CeO2-DC (The red, green and yellow dots represent O, Ce and Pd elements, respectively).
Materials 13 05768 g004aMaterials 13 05768 g004b
Figure 5. Textural properties of the samples: (a) N2 adsorption/desorption isotherms; (b) pore size distributions.
Figure 5. Textural properties of the samples: (a) N2 adsorption/desorption isotherms; (b) pore size distributions.
Materials 13 05768 g005
Figure 6. XPS spectrum: (a) Pd 3d; (b) Ce 3d.
Figure 6. XPS spectrum: (a) Pd 3d; (b) Ce 3d.
Materials 13 05768 g006aMaterials 13 05768 g006b
Figure 7. H2-TPR curves of the samples.
Figure 7. H2-TPR curves of the samples.
Materials 13 05768 g007
Table 1. Characteristic data of the samples.
Table 1. Characteristic data of the samples.
SamplesSBETa (m2/g)Amesb (m2/g)VPc (cm3/g)Vmicd (cm3/g)
CeO2-P10.710.70.044-
CeO2-DC66.366.30.23-
CeO2-MOF90.590.50.14-
Pd/CeO2-P7.75.90.0390.00074
Pd/CeO2-DC80.477.30.210.00059
Pd/CeO2-MOF73.873.80.13-
a The BET specific surface area. b Calculated by the BJH method. c Total pore volume estimated at P/P0 = 0.99. d Micropore volume calculated by the t-plot method.
Table 2. Summary table of XPS characterization result of catalysts.
Table 2. Summary table of XPS characterization result of catalysts.
Catalyst Pd 3d (eV)Ce 3d (eV)Pd2+/(Pd2+ + Pd0) (%)Ce3+/(Ce3+ + Ce4+)
(%)
Pd0Pd2+Ce3+Ce4+
Pd/CeO2-P3d3/2-342.4883.7, 887.3882.2, 889.3, 897.810018.1
3d5/2-336.9898.6, 901.9900.7, 907.5, 916.6
Pd/CeO2-DC3d3/2340.8342.9884.0, 887.0 882.3, 889.2, 897.971.124.8
3d5/2335337.1898.7, 902.2900.8, 907.4, 916.5
Pd/CeO2-MOF3d3/2339.3342.2883.3, 884.8 882.3, 888.7, 897.957.123.1
3d5/2335.3336.8898.7, 902.4900.8, 907.4, 916.5
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, Z.; Chen, Z.; Zheng, J.; Zuo, S. Effect of Particle Size and Crystal Surface of CeO2 on the Catalytic Combustion of Benzene. Materials 2020, 13, 5768. https://doi.org/10.3390/ma13245768

AMA Style

Wang Z, Chen Z, Zheng J, Zuo S. Effect of Particle Size and Crystal Surface of CeO2 on the Catalytic Combustion of Benzene. Materials. 2020; 13(24):5768. https://doi.org/10.3390/ma13245768

Chicago/Turabian Style

Wang, Zhuo, Zhu Chen, Jie Zheng, and Shufeng Zuo. 2020. "Effect of Particle Size and Crystal Surface of CeO2 on the Catalytic Combustion of Benzene" Materials 13, no. 24: 5768. https://doi.org/10.3390/ma13245768

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