Next Article in Journal
A Nickel/Organoboron-Catalyzed Coupling of Aryl Bromides with Sodium Sulfinates: The Synthesis of Sulfones under Visible Light
Previous Article in Journal
Suitable Stereoscopic Configuration of Electrolyte Additive Enabling Highly Reversible and High—Rate Zn Anodes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Praseodymium-Doped Cr2O3 Prepared by In Situ Pyrolysis of MIL-101(Cr) for Highly Efficient Catalytic Oxidation of 1,2-Dichloroethane

School of Biological and Environmental Engineering, Nanjing University of Science & Technology, Nanjing 210094, China
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(14), 3417; https://doi.org/10.3390/molecules29143417 (registering DOI)
Submission received: 13 May 2024 / Revised: 19 June 2024 / Accepted: 19 July 2024 / Published: 21 July 2024

Abstract

:
The development of economical catalysts that exhibit both high activity and durability for chlorinated volatile organic compounds (CVOCs) elimination remains a challenge. The oxidizing and acidic sites play a crucial role in the oxidation process of CVOCs; herein, praseodymium (Pr) was introduced into CrOx catalysts via in situ pyrolysis of MIL-101(Cr). With the decomposition of the ligand, a mixed micro-mesoporous structure was formed within the M-Cr catalyst, thereby reducing the contact resistance between catalyst active sites and the 1,2-dichloroethane molecule. Moreover, the synergistic interaction between chromium and praseodymium facilitates Oβ species and acidic sites, significantly enhancing the low-temperature catalytic performance and durability of the M-PrCr catalyst for 1,2-dichloroethane (1,2-DCE) oxidation. The M-30PrCr catalyst possess enhanced active oxygen sites and acid sites, thereby exhibiting the highest catalytic activity and stability. This study may provide a novel and promising strategy for practical applications in the elimination of 1,2-DCE.

1. Introduction

In the past few decades, the emissions of volatile organic compounds (VOCs) have increased dramatically with the rapid development of the industrial economy [1]. Among these compounds, chlorinated volatile organic compounds (CVOCs) have been associated with the generation of photochemical smog pollution and teratogenic effects due to their high chemical stability and severe toxicity [2,3,4,5,6]. One notorious member of the CVOC family is 1,2-dichloroethane (1,2-DCE), which is commonly found in various industries such as paint manufacturing, pharmaceutical chemicals production, and petrochemical operations [7]. Therefore, there is an urgent need to develop cost-effective methods for removing 1,2-DCE. Among the different treatment technologies available, catalytic oxidation is considered a highly efficient and low-temperature operation technology for eliminating CVOCs. However, developing economical catalysts that exhibit both high activity and durability remains a challenge [8]. Firstly, CVOCs present a greater challenge in achieving complete oxidation compared to other VOCs and require higher temperatures for their catalytic degradation [9]. Secondly, the catalyst is susceptible to surface adsorption of chlorine species and carbon deposition, which can potentially lead to its poisoning [10,11]. Therefore, it is crucial to develop exceptionally efficient catalysts that demonstrate superior activity, selectivity (towards CO2 and HCl), and stability in order to facilitate practical applications in catalytic oxidation degradation.
Although noble metal (e.g., Pt, Pd, and Ru) catalysts exhibit excellent deep oxidation activity in the destruction of CVOCs, their vulnerability to toxic deactivation, sintering, and high prices have strictly limited practical application [12,13,14,15]. In contrast, transition metal oxide catalysts are more promising and have drawn more attention due to their low cost, superior activity, and stability after modification [16]. Among transition metal catalysts, chromium-based catalysts showed excellent catalytic performance mainly due to the enhanced surface lattice oxygen mobility and strong acidity [17,18,19]. Zhang et al. loaded different transition metals (Cr, Mn, Fe, Ni, and Cu) on a TiO2 support for 1,2-DCE oxidation and found that the introduction of Cr showed the best catalytic performance [20]. Separately, rare Earth elements have also received extensive attention due to their unique bonding characteristics and special physical and chemical properties [21]. Particularly, praseodymium oxides showed great catalytic ability owing to their highest oxygen mobility among all undoped rare Earth oxides [21,22,23]. Doping transition metals is an effective strategy to enhance the degradation activity of the catalyst as well as the ability to inhibit chlorine poisoning [16]. It is therefore reasonable to hypothesize that a catalyst combining the advantageous properties of both Cr2O3 and Pr6O11 could exhibit exceptional efficiency in the deep degradation of CVOCs. Metal organic frameworks (MOFs) are crystalline materials composed of metal ions or clusters coordinated with organic ligands. MOF materials have been applied in various systems for CVOCs treatment, including air filters, gas masks, and industrial scrubbers. Their high capacity and selectivity suit indoor and outdoor air purification applications. However, MOF materials still lack effective application in high-concentration CVOCs removal.
In this work, aiming to develop highly efficient catalysts for 1,2-DCE oxidation, a series of porous M-PrCrOx catalysts derived from MIL-101(Cr) were prepared with different Pr/Cr molar ratios. Compared with the conventional coprecipitation method, the pyrolysis strategy provides higher surface areas and good diffusion properties [24], resulting in better activity. The synergy effects between Pr and Cr elements in PrCrOx catalysts during 1,2-DCE destruction were investigated by evaluating the physicochemical properties (structural properties, surface composition and element status, reducibility, and surface acidity), catalytic activity, and toxic resistance. It was found that M-PrCr exhibited good catalytic performance and stability, which proved that praseodymium doping and in situ pyrolysis of MOFs are promising strategies to develop effective transition metal oxide catalysts for CVOCs elimination.

2. Results and Discussion

2.1. Structural Properties

As shown in Figure S2, MOF-Cr and MOF-30PrCr have decomposed into M-Cr and M-30PrCr, respectively. Thermogravimetric analysis (TGA) (Figure S1) was applied to further confirm the pyrolysis of MOF-Cr, and the structure of MOF precursor was completely collapsed at around 380 °C, suggesting that the calcination temperature was enough to decompose the MIL-101(Cr) and eventually formed Cr2O3. The crystal structures of the PrCrOx catalysts were investigated by XRD and the patterns of all samples are presented in Figure 1. Several typical characteristic peaks were observed at 24.49°, 33.60°, 36.20°, 39.75°, 41.48°, 50.22°, 54.85°, 63.45°, and 65.10, as well as 72.94° on the PrCrOx catalysts, corresponding to the (0 1 2), (1 0 4), (1 1 0), (0 0 6), (1 1 3), (0 2 4), (1 1 6), (2 1 4), (3 0 0), and (1 0 10) of the rhombohedral phase Cr2O3 (PDF#38-1479) [3], respectively. It is evident that all samples displayed similar XRD peaks and showed a good match with the XRD pattern of the standard Cr2O3 sample, suggesting the successful synthesis of the Cr2O3 catalyst through MIL-101(Cr) pyrolysis [24]. Separately, P-Pr prepared by precipitation method possessed a cubic phase praseodymium oxide structure (PDF#42-1121) [25]. However, no diffraction peaks of praseodymium species can be observed on the M-PrCr samples, which was attributed to the good dispersion of Pr species in the PrCrOx catalysts. The EDS results also confirmed the Pr element was uniformly dispersed on the M-30PrCr catalysts. According to Scherrer equation, the intensity and width of XRD diffraction peaks are related to the crystal size of the sample [26]. Compared with P-Cr, the diffraction peaks of M-Cr are notably wider, suggesting that pyrolysis strategy provided smaller crystal size of Cr2O3 [27]. These nano-scaled Cr2O3 produced more lattice defects and higher surface area of materials [24]. Moreover, the doping of Pr further reduced the crystal size of M-PrCr catalysts, which can provide abundant active sites for the catalytic oxidation. Additionally, P-30PrCr obtained by the co-precipitation method presented very low crystallinity (Figure S2), which might be due to the significant gap in solubility product between Cr(OH)3 (Ksp = 6.3 × 10−31) and Pr(OH)3 (Ksp = 1.5 × 10−20). The XRD findings present preliminary evidence that the choice of preparation method significantly influences the characteristics of Cr2O3 crystal structure, encompassing surface area, lattice defects, crystal size, and other pertinent factors. These factors are crucial for attaining optimal catalytic performance in VOCs oxidation [28,29].
The N2 adsorption-desorption isotherms and corresponding pore-size distribution curves for all samples are presented in Figure 2 and Figure 3, respectively. As shown in Figure 3, the prepared catalysts exhibited characteristic IV type adsorption isotherms accompanied by H3 hysteresis loops within the relative pressure range of 0.5–1.0, indicating the presence of a mesoporous structure in the samples. As illustrated in Figure 3, all catalysts synthesized through pyrolysis exhibited a combination of microporous and mesoporous structures, with predominant pore sizes located within the range of 0.5–5 nm. Compared with P-Cr, M-Cr derived from the MOF-Cr precursor exhibited obviously promoted porosity, which improved the diffusion of the 1,2-dichloroethane molecule and reduced the contact resistance between catalyst active sites and 1,2-dichloroethane molecule. With the introducing of praseodymium, the pore size as well as the pore volume of M-PrCr catalysts were gradually enlarged compared with M-Cr catalyst. Separately, the samples synthesized by the precipitation method exhibited bigger irregular mesopore structures and the pore sizes were mainly located at 10–40 nm. The relevant information regarding the BET surface areas, pore volumes, average pore diameters, and catalytic performance of each catalyst can be accessed from Table 1. The decomposition of the ligand led to the formation of numerous irregular pores, resulting in a significant enhancement in the specific surface area (SBET) of M-Cr up to 135.24 m2/g. This value is considerably higher than that observed for P-Cr (57.99 m2/g). After the introduction of praseodymium (Pr) doping, the specific surface area (SBET value) exhibited a significant increase, with M-30PrCr catalyst demonstrating the highest value at 699.31 m2/g. This can be attributed to the moderating effect of praseodymium on the decomposition of terephthalic acid within MOF-Cr, leading to the gradual formation of a more porous structure during pyrolysis. This is further supported by the observed progressive enlargement in average pore size as Pr doping concentration increases on M-PrCr catalysts. After calcination, the formation of PrOx within the CrOx lattice led to the development of novel pore structures, consequently resulting in a gradual augmentation of pore volume. However, the excessive doping of praseodymium blocked certain pore structures, resulting in a reduction in both specific surface area and pore volume. Separately, the coprecipitation method yielded only a slight improvement in specific surface area and pore volume for P-30PrCr compared to P-Cr. It is widely acknowledged that an increased pore volume and higher specific surface area can have beneficial effects on catalytic activity when it comes to oxidizing volatile organic compounds [30].
The catalyst structure is vital to catalytic performance. Hence, morphology and element distribution of the catalysts were investigated by SEM and EDS mapping tests. As shown in Figure 4c,g, catalysts prepared with precipitation method displayed irregular granule structure with porosity and looseness. However, small particles of P-Cr were agglomerated to large plates after Pr doping, which blocked the pore structure and reduced the specific area of the catalyst. As shown in Figure 4a, MIL-101(Cr) precursor with octahedral structure was prepared by fluorine-free hydrothermal method. After being calcined at 400 °C for 4 h, the smooth surface of MIL-101(Cr) became rather rough, which is due to the terephthalic acid of MIL-101(Cr) precursor being pyrolyzed, leaving octahedral porous Cr2O3 nanosatellites being noted as M-Cr. With the doping of Pr, the octahedral structure of M-20PrCr and M-30PrCr was partially altered, presenting smaller particle sizes and higher specific surface area. The highest specific surface area was found in M-30PrCr; the partially damaged octahedral structure enhanced the specific area, which promoted contact with CVOCs, thus benefiting the catalytic performance. Additionally, the excessive doping of Pr hindered the formation of MOF precursors, resulting in the irregular blocky structure of M-40PrCr, which significantly reduced the specific area of the catalyst. The elimination of CVOCs can be influenced by the specific area of the catalyst, as it has the potential to affect the accessibility of catalytic active sites. These findings are consistent with the results obtained from XRD and N2 adsorption/desorption analyses.
As shown in Figure 4i–l, the surface element distribution of the M-30PrCr catalyst was determined through EDS analysis. Randomly selected regions were scanned and subjected to EDS analysis, revealing a uniform dispersion of Cr, Pr, and O elements on the catalyst surface. This can be attributed to the presence of an amorphous crystal structure observed in the XRD patterns. Additionally, the molar ratio of Pr/Cr (0.096) obtained from EDS analysis was found to be significantly lower than the theoretical value (0.30). This observation suggests that Pr has been incorporated into the lattice of CrOx, facilitating interaction between Pr and Cr. Consequently, it is inferred that doping with PrOx may hinder the crystallization process of CrOx. Ultimately, by employing pyrolysis method along with appropriate proportions of Pr doping, modifications can be made to both structure and morphology of catalysts leading to developed catalytic activity. Notably, among all tested catalysts, M-30PrCr demonstrated superior performance in this regard.

2.2. Surface Composition and Element Status

The surface elemental compositions, chemical valence states, and oxygen species of the prepared catalysts were analyzed using XPS technique. Corresponding results are presented in Figure 5, which displays the XPS spectra for Cr 2p, O 1s, and Pr 3d signals.
Figure 5b shows the spectra of O 1s, which is decomposed into three peaks located at 530.2, 531.8, and 533.2 eV. Among them, the peak observed at about 530.2 eV corresponds to lattice oxygen (Oα.), primarily originating from Cr and Pr. The peaks located at around 531.8 eV can be attributed to the presence of surface-adsorbed oxygen, indicating the existence of Oβ species. Meanwhile, the peak detected at 533.0–533.4 eV corresponds to carbonates and chemisorbed water (Oγ) on the catalyst’s surface [18,31,32]. The Oβ/(Oα + Oβ) ratio was determined by calculating the areas of Oα and Oβ, as presented in Table 2. The prepared catalysts exhibited the following order of Oβ/(Oα + Oβ) ratios: M-30PrCr (0.81) > M-40PrCr (0.79) > M-20PrCr (0.74) > P-Cr (0.45) > M-Cr (0.32) > P-30PrCr (0.30). The introduction of Pr alters the valence state of Cr, leading to a significant increase in the proportion of Oβ. In general, Oβ demonstrates higher activity compared to Oα due to its relatively enhanced mobility and reduced surface bonding energy [33]. The high concentration of Oβ in M-30PrCr contributes to its exceptional catalytic activity at low temperatures.
The Pr 3d spectra of M-PrCr and P-PrCr catalysts are decomposed into five peaks, denoted as v and u, corresponding to the energy levels of Pr 3d3/2 and Pr 3d5/2, respectively. As depicted in Figure S5, the peaks v′ (933.1 eV) and u′ (928.4 eV) are attributed to surface Pr3+ species while the remaining peaks are assigned to surface Pr4+ species [34,35,36,37,38]. The results demonstrate the presence of Pr3+ and Pr4+ in a mixed valence state on the P-30PrCr surface, indicating the formation of oxygen vacancies [39]. However, due to the significant dispersion of Pr, the characteristic peak associated with Pr cannot be observed on the surface of the M-PrCr catalyst.
As shown in Figure 5a, The Cr 2p spectra of the catalysts are resolved into four peaks. The Cr 2p3/2 bond energies of Cr3+ and Cr6+ are approximately located at 576.4 eV and 578.7 eV, and the bands located at 586.2 eV and 588.2 eV are attributed to the Cr3+ and Cr6+ of the Cr 2p1/2 spectra, respectively [31,40,41], revealing the coexistence of Cr3+ and Cr6+ on the catalyst surface. The ratios of Cr6+/Cr3+ are 0.97, 2.16, 1.23, 1.48, 1.41, and 4.07 for M-Cr, P-Cr, M-20PrCr, M-30PrCr, M-40PrCr, and P-30PrCr, respectively. With the addition of Pr, the proportion of Cr6+ initially increases and then decreases. The appropriate doping of Pr can effectively enhance the content of Cr6+. The results presented above demonstrate that the interaction between Pr6O11 and Cr2O3 enhances the presence of Cr6+ and OLatt on the catalyst surface, which collectively contribute to an improved catalytic performance in 1,2-DCE combustion. Based on the XPS analysis findings, it can be concluded that M-30PrCr exhibits superior catalytic activity for 1,2-DCE combustion, aligning with the conclusions drawn from catalytic performance evaluation in Section 3.4.

2.3. Surface Redox Properties and Acid Properties

The reducibility at low temperatures is a crucial factor in determining the catalytic efficiency of metal oxide catalysts [18]. To evaluate the redox property of the PrCrOx catalysts, H2-TPR technique was conducted over a temperature range spanning from 50 to 900 °C. The reduction peaks in the ranges of <250 °C, 250–400 °C and 400–650 °C could be assigned to the reduction of physically absorbed oxygen, surface chemical oxygen, and lattice oxygen species [42,43]. As shown in Figure 6, the reduction peak of P-Pr appeared at 501 °C, which represents the reduction of the surface Pr6O11 [35,44]. Compared to P-Cr catalyst, the M-Cr exhibited significantly enhanced reducibility. The peaks observed at 331 °C and 443 °C were attributed to the reduction of Cr6+ to Cr3+ and the subsequent reduction of Cr3+ to metallic Cr, respectively [18,32]. Compared with M-Cr, the reduction peak of M-PrCr catalyst is significantly enhanced, indicating that the introduction of Pr has greatly developed the reducibility of the catalyst. Furthermore, with an increase in the doping amount of praseodymium, there is a noticeable shift of the reduction peaks towards a lower temperature range. This observation suggests that the introduction of an optimal concentration of Pr significantly enhances the mobility of oxygen species. The quantification of H2 consumption was conducted to evaluate the reducibility of catalysts at temperatures lower than 700 °C, as indicated in Table 2. The amount of H2 consumption (<700 °C) exhibited a descending trend as follows: M-40PrCr (28.37 mmol⋅g−1) > M-30PrCr (27.24 mmol⋅g−1) > P-Pr (14.19 mmol⋅g−1) > M-20PrCr (10.61 mmol⋅g−1) > M-Cr (7.10 mmol⋅g−1)> P-Cr (5.79 mmol⋅g−1), demonstrating that M-40PrCr exhibited the highest amount of H2 consumption. These findings suggest that the M-30PrCr catalyst exhibits improved reducibility, facilitating the efficient transport of oxygen species and initiation of reactants [45].
The NH3-TPD technique was utilized to assess the acidity of the prepared catalysts, and Figure 7 illustrates the corresponding results. Desorption peaks corresponding to NH3 absorption on acid sites categorized as weak, medium, and strong are observed at temperatures below 300 °C, between 300–600 °C, and above 600 °C respectively [20,31,33]. In general, it is widely acknowledged that deep oxidation efficiency for CVOCs is higher on strong acid sites compared to weak acid sites [40]. The quantities of acid sites were calculated and the results are presented in Table 2, and the order of medium and strong acid sites is as follows: M-40PrCr (763) > M-30PrCr (722) > M-20PrCr (697) > M-Cr (578) > P-Pr (251) > P-Cr (74). The acidity of the M-Cr catalyst, prepared by pyrolysis method, is significantly stronger than that of P-Cr catalyst, prepared by co-precipitation method. Furthermore, the introduction of praseodymium doping resulted in a shift towards higher temperatures for desorption peaks, indicating a significant enhancement in the acidity of M-PrCr catalysts. It is widely recognized that the presence of strong acid sites enhances the dissociative adsorption of chlorine from CVOCs [18,32,45]. Apparently, the incorporation of praseodymium species significantly enhances the abundance of potent acidic sites in PrCrOx catalysts, thereby facilitating the release of chlorine from 1,2-DCE [46].

2.4. Catalytic Performance and Structure-Activity Relationship

Catalysts with varying molar ratios of Pr/Cr were utilized for the catalytic oxidation of 1,2-dichloroethane. The relationship between temperature and Pr/Cr on 1,2-dichloroethane conversion is presented in Figure 8a. In general, the conversion efficiency of 1,2-DCE increased with the temperature rise and presented a S-shaped curve, suggesting that high temperature is conducive to 1,2-DCE elimination. The T50 (the temperature of 50% 1,2-DCE conversion) of P-Cr, M-Cr, P-30PrCr, M-20PrCr, M-40PrCr and M-30PrCr catalysts is 371 °C, 320 °C, 302 °C, 274 °C, 271 °C and 255 °C, respectively. The corresponding T90 (the temperature of 90% 1,2-DCE conversion) for M-20PrCr, M-40PrCr and M-30PrCr is 312 °C, 298 °C and 286 °C, respectively. The T50 of P-Pr was above 400 °C, indicating that Cr2O3 provided the main active sites. It is obviously seen that P-30PrCr presented lower T50 than pure P-Cr and P-Pr, which indicates that the interaction between Cr2O3 and Pr6O11 can improve the catalytic ability for 1,2-dichloroethane oxidation. Notably, M-Cr derived from MIL-101(Cr) exhibited higher catalytic performance than P-Cr, indicating the porous structure generated by pyrolysis method promoted the contact between CVOCs and catalyst. Moreover, the catalytic performance of M-PrCr catalysts (M-20PrCr, M-30PrCr and M-40PrCr) are higher than M-Cr and P-30PrCr in 1,2-DCE oxidation, and M-30PrCr presented the highest activity. These results could conclude that the doping of Pr significantly promoted the activity of the M-PrCr catalysts for 1,2-DCE oxidation. This might be attributing to the synergistic effect between chromium and praseodymium, causing the distortion of chromium lattice and the improved oxidation performance of 1,2-DCE.
Due to their exceptional chemical stability, CVOCs are difficult to completely oxidize, leading to the formation of highly toxic intermediates such as dioxins [8]. Therefore, in practical applications, complete mineralization is even more important than a lower conversion temperature. As demonstrated in Figure 8b, the mineralization rates exhibit a similar trend to the conversion rates observed on the catalyst. It is shown that the M-30PrCr catalyst doped with Pr presented excellent oxidation ability, and complete oxidation of 1,2-DCE was achieved at about 360 °C. Compared to the co-precipitation method, the pyrolysis method offers a higher abundance of low-temperature reduction sites for CrOx, enhances the presence of active oxygen species on the catalyst surface, thus benefits the complete oxidation of 1,2-DCE. Moreover, the introduction of Pr generated new oxygen vacancies and significantly improved the acidity of the catalysts, thereby further facilitated the elimination of 1,2-DCE.
To further investigate the reaction mechanism, the Arrhenius plots with linear fitting was presented in Figure 8c. From these plots, the apparent activation energy (Ea) for the oxidation of 1,2-DCE was determined using the Arrhenius equation [47]. The order trend of calculated Ea is P-Pr (74.9 kJ/mol) > P-Cr (69.4 kJ/mol) > M-Cr (65.9 kJ/mol) > P-30PrCr (62.5 kJ/mol) > M-20PrCr (45.2 kJ/mol) > M-40PrCr (43.1 kJ/mol) > M-30PrCr (40.8 kJ/mol). Clearly, praseodymium-doped catalysts presented lower activation energy than pristine P-Cr, which was agreeing well with the results of activity tests. Compared with P-30PrCr, M-30PrCr prepared by pyrolysis method presented lower activation energy, indicating that the pyrolysis method facilitated the interaction between chromium and praseodymium. These results confirmed the pyrolysis method and the introduction of praseodymium species significantly enhanced the oxidation ability of CrOx catalysts.

2.5. Durability Test

To investigate the durability of PrCrOx catalysts, an on-stream reaction experiment was carried out at 290 °C. As shown in Figure 9, M-30PrCr exhibited remarkable durability for 1,2-DCE oxidation over a prolonged period of 100 h. In the initial 48-h period, the conversion of 1,2-DCE over M-30PrCr decreased from 95% to 82%, attributed to the accumulation of chlorine species and carbon. Separately, the conversion over M-Cr decreased from 44% to 36% and the conversion over P-PrCr decreased from 37% to 26%. Additionally, the influence of H2O on the catalytic activity was also studied. After the introduction of 5 vol.% water vaper, the conversion of 1,2-DCE over M-30PrCr gradually dropped to 78%, followed by a further decrease to 70% with the introduction of 10 vol.% water vapor. A similar situation was observed on the M-Cr catalyst, where the conversion decreased from 23% to 20% with the increase of water vapor. The observed phenomenon can be attributed to the competitive adsorption of water on the active sites of the catalysts. Interestingly, when H2O is removed, the conversion of 1,2-DCE over M-30PrCr and P-30PrCr was recovered to a higher level than that after 48h. This might be due to the fact that the presence of water vapor developed hydroxyl groups on the catalyst surface, facilitating the desorption of Cl species and thereby mitigating catalyst deactivation [48]. However, only partial conversion over pristine M-Cr was recovered after the removal of water vapor, indicating pristine Cr2O3 is vulnerable to Cl poisoning. The durability test results show that Pr doping can effectively enhance the Cl resistance of CrOx catalysts, among which M-30PrCr has excellent durability against the effects of 1,2-DCE and H2O.

2.6. Characterization of Used Catalysts

To gain further insights into the reaction behavior of the PrCrOx catalysts during 1,2-DCE degradation, a comprehensive analysis including XRD, SEM, EDS and H2-TPR was conducted on the used catalysts. As shown in Figure 10, the X-ray diffraction patterns of both M-Cr and M-30PrCr exhibit the appearance of distinct broad peaks at an angle of 2θ = 16°. This phenomenon suggests that carbon species were accumulating on the catalyst surface during the durability assessment. It is obvious that the new peak of used M-30PrCr is much weaker than that of used M-Cr, indicating that the presence of Pr can effectively inhibit carbon deposition. Separately, the characteristic peaks of the used catalysts were enhanced compared with that of the fresh catalysts, which may be due to the sintering of the catalyst. Figures S3 and S4 depict the SEM and EDS mapping images, respectively, while Table S1 presents the corresponding results. After the durability tests, the deposition of C and Cl species was found on the surface of both M-Cr and M-30PrCr catalysts. However, compared with M-Cr, the accumulation of C and Cl species on the surface of M-30PrCr catalyst was significantly reduced, indicating that Pr doping can inhibit catalyst poisoning. The H2-TPR curves are presented in Figure 11. After the durability test, the reduction peaks located in the high-temperature area almost completely disappeared, and the reduction peaks in the low-temperature area also weakened significantly. This phenomenon illustrates that the surface chemical oxygen of the catalysts played a crucial role in CVOCs oxidation. Compared with M-Cr, the reduction peaks of M-Pr are better preserved, indicating that the introduction of Pr is helpful for the migration of oxygen species, thus enhancing the lifetime of the catalyst.

3. Materials and Methods

3.1. Chemicals and Materials

The praseodymium nitrate hexahydrate (Pr(NO3)3·6H2O), chromic nitrate nonahydrate (Cr(NO3)3·9H2O), and terephthalic acid (H2BDC) were purchased from Aladdin Biochemical Technology Co., Ltd., Shanghai, China. The 1,2-dichloroethane (DCE), ethyl alcohol, nitric acid, and sodium hydroxide were purchased from Sinopharm Chemical Reagent Co., Ltd., Shanghai, China. All reagents were analytical grade and used without further purification.

3.2. Catalyst Preparation

3.2.1. Synthesis of MIL-101(Cr)

The synthesis of MIL-101(Cr) was achieved using a hydrothermal method without the use of fluorine [49]. In this procedure, 4 g of Cr(NO3)3·9H2O and 1.66 g of terephthalic acid were dissolved in 48 mL of deionized water and stirred at room temperature for 1 h. Subsequently, a modulator consisting of approximately 1 mL nitric acid (~60%) was added to the solution, followed by ultrasonication for 1 h. The reaction took place in a Teflon-lined autoclave at a temperature of 210 °C for a duration of 8 h. After cooling, the resulting product was separated through centrifugation and subjected to heating in ethanol (100 mL) for an extended period of time (24 h). Finally, the obtained green powder was collected via centrifugation and dried at a temperature of 80 °C over a period lasting 12 h.

3.2.2. Synthesis of M-Cr and M-xPrCr

As shown in Figure 12, the M-Cr and M-xPrCr catalysts were synthesized via in situ pyrolysis of the MOF-Cr precursor. Typically, the sacrificial precursor MIL-101(Cr) was calcined at 400 °C under an air atmosphere for 4 h with a heating rate of 5 °C·min−1. The obtained porous Cr2O3 was marked as M-Cr. Praseodymium-doped Cr2O3 was prepared with a similar method. Typically, a certain mass of Pr(NO3)3·6H2O (theoretical addition amounts are determined by the atom molar ratio of Pr/Cr, e.g., M-20PrCr requires about 0.87 g of Pr(NO3)3·6H2O), which was dissolved into the precursor solution, and other conditions were the same as the synthesis of MIL-101(Cr). The MIL-101(Cr) material doped with praseodymium was subjected to calcination at 400 °C for 4 h in the presence of air. The resulting product, denoted as M-xPrCr, represents praseodymium-doped Cr2O3, where x/100 corresponds to the molar ratio of Pr/Cr atoms.

3.2.3. Synthesis of P-Cr and P-xPrCr

For comparative purposes, P-Cr and P-xPrCr were synthesized using a conventional precipitation technique. Taking the example of P-30PrCr, a solution was prepared by dissolving 4 g of Cr(NO3)3·9H2O and 1.31 g of Pr(NO3)3·6H2O in 50 mL of deionized water. Subsequently, under vigorous stirring for 2 h, a slow addition of 100 mL of 1 M NaOH solution took place. The resulting precipitate was collected through centrifugation and subjected to three rounds of water washing. Finally, the precipitate underwent drying at a temperature of 80 °C for a duration of 12 h, followed by calcination at an air atmosphere with temperatures maintained at 400 °C for 4 h. The praseodymium-doped Cr2O3 obtained from this process was denoted as P-xPrCr, where x/100 represents the atom molar ratio between Pr and Cr.

3.3. Catalyst Characterization

The catalysts synthesized in this study underwent a systematic characterization process involving techniques such as X-ray diffraction (XRD), scanning electron microscopy (SEM), energy dispersive spectroscopy (EDS), N2 adsorption/desorption, X-ray photoelectron spectroscopy (XPS), H2 temperature-programmed reduction (H2-TPR), and temperature-programmed desorption of ammonia (NH3-TPD). The detailed procedures and parameters for these tests can be found in the Supplementary Materials.

3.4. Catalytic Test

The performance of catalytic combustion for 1,2-dichloroethane was evaluated in a fixed-bed reactor consisting of a quartz tube with an inner diameter of 4 mm under standard atmospheric pressure. In each experiment, 100 mg of catalysts (40–60 mesh) were immobilized using quartz wool, resulting in a gas hourly space velocity (GHSV) of 20,000 mL·g−1·h−1 and a total flow rate of 33.33 mL·min−1. The liquid reactants (1,2-dichloroethane) were introduced into the feed stream by passing dried air through a saturator maintained at a temperature of 6 °C. The feed stream was then diluted with dried air, generating a feeding flow containing 1000 ppm reactant and composed of 21% O2/79% N2. Flow rates were regulated using online mass flowmeters. Concentrations of 1,2-DCE and chlorinated by-products were determined via analysis on an online gas chromatograph (GC3420A, Beifen Ruili Co., Ltd., Beijing, China) equipped with both flame ionization detector (FID) and electron capture detector (ECD) for quantitative assessment of organic compounds. Furthermore, the outlet CO2 underwent reduction by hydrogen within a methanation furnace before being detected by the same chromatograph system. A K-type thermocouple was inserted into the catalyst bed to monitor its temperature during catalysis. Data of catalytic oxidation tests were collected after achieving steady state conditions for at least thirty minutes.
The 1,2-dichloroethane conversion rate (XDCE, %), CO2 yield (CO2yield, %), and organic byproduct yield (Yb, %) were calculated as follows:
X D C E = ( D C E ) i n ( D C E ) o u t ( D C E ) i n × 100 %
C O 2 y i e l d = ( C O 2 ) o u t 2 × ( D C E ) i n × 100 %
Y b = X D C E C O 2 y i e l d
The reaction rate of 1,2- dichloroethane was calculated as follows:
r = Q × X D C E m
where r is the reaction rate (mol·g−1·s−1), Q is the molar flow of 1,2-dichloroethane (mol·s−1), m is the mass of catalyst (g).
Activation energy was calculated via the antiderivative of Arrhenius equation:
l n r = l n A E a R T
where E a is activation energy (kJ·mol−1), T is reaction temperature (K), A is a preexponential factor, and R is the molar gas constant (8.314 J·mol−1·K−1).
The turnover frequency (TOF) was obtained by the following equation:
T O F = Q × X D C E n C r
where n C r is the amount of Cr element (mol) in the catalyst.
The stability test for 1,2-dichloroethane was performed at 290 °C for a 120 h continuous test under the same conditions as the catalytic combustion test, unless otherwise indicated.

4. Conclusions

In this work, the M-PrCr catalysts were prepared via the in situ pyrolysis of MIL-101(Cr) for 1,2-dichloroethane oxidation. Compared to conventional coprecipitation method, PrCrOx catalysts prepared via pyrolysis strategy exhibited much higher specific surface area and pore volume, which improved the diffusion of 1,2-dichloroethane molecule. In addition, the introduction of Pr-doping effectively increased the number of acidic sites and enhanced the mobility of oxygen species. Consequently, this significantly improved the performance at low temperatures and prolonged the durability of M-PrCr catalysts. M-30PrCr exhibited the highest catalytic performance (T50 = 255 °C, T90 = 286 °C) and CO2 selectivity. Separately, the M-30PrCr catalyst presented exceptional durability against 1,2-DCE and H2O during 120 h continues test. In summary, conclusions can be drowned from this study that pyrolysis of MOF strategy and appropriate doping of Pr species can significantly promote the oxidation ability of the catalyst, which provide promising strategy for practical applications in the elimination of 1,2-DCE.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29143417/s1, Figure S1: TGA analysis of MIL-101(Cr); Figure S2: XRD patterns of samples: (a) MOF-Cr; (b) MOF-30PrCr; (c) M-Cr; (d) M-30PrCr; (e) P-30PrCr; Figure S3: The EDS characterization result for M-30PrCr (total graph); Figure S4: SEM images and EDS mapping results of used catalysts: (a) used M-Cr, (b) used M-30PrCr; Figure S5: The EDS characterization result for (a) used M-Cr and (b) used M-30PrCr (total graph); Figure S6: The XPS Pr 3d spectra for P-30PrCr; Table S1: Surface element distribution of used catalysts.

Author Contributions

Conceptualization, P.Z. and S.C.; methodology, P.Z. and S.C.; validation, Z.H. and S.C.; software: P.Z.; formal analysis, P.Z.; investigation, P.Z.; resources, Z.H. and S.C.; data curation, P.Z. and S.C.; writing—original draft preparation, P.Z.; writing—review and editing, Z.H. and S.C.; visualization, P.Z.; supervision, S.C.; project administration, S.C.; funding acquisition, S.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Fundamental Research Funds for the Central Universities (30920021114), China.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data will be made available on request.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Li, N.; Gaillard, F. Catalytic combustion of toluene over electrochemically promoted Ag catalyst. Appl. Catal. B Environ. 2009, 88, 152–159. [Google Scholar] [CrossRef]
  2. Yang, P.; Fan, S.; Chen, Z.; Bao, G.; Zuo, S.; Qi, C. Synthesis of Nb2O5 based solid superacid materials for catalytic combustion of chlorinated VOCs. Appl. Catal. B Environ. 2018, 239, 114–124. [Google Scholar] [CrossRef]
  3. Zhang, X.; Liu, Y.; Deng, J.; Zhao, X.; Zhang, K.; Yang, J.; Han, Z.; Jiang, X.; Dai, H. Three-dimensionally ordered macroporous Cr2O3−CeO2: High-performance catalysts for the oxidative removal of trichloroethylene. Catal. Today 2020, 339, 200–209. [Google Scholar] [CrossRef]
  4. Wan, J.; Yang, P.; Guo, X.; Zhou, R. Elimination of 1,2-dichloroethane over (Ce,Cr)O2/Nb2O5 catalysts: Synergistic performance between oxidizing ability and acidity. Chin. J. Catal. 2019, 40, 1100–1108. [Google Scholar] [CrossRef]
  5. Pan, C.; Wang, W.; Zhang, Y.; Nam, J.C.; Wu, F.; You, Z.; Xu, J.; Li, J. Porous graphitized carbon-supported FeOCl as a bifunctional adsorbent-catalyst for the wet peroxide oxidation of chlorinated volatile organic compounds: Effect of mesopores and mechanistic study. Appl. Catal. B Environ. 2023, 330, 122659. [Google Scholar] [CrossRef]
  6. Mustafa, M.F.; Liu, Y.; Duan, Z.; Guo, H.; Xu, S.; Wang, H.; Lu, W. Volatile compounds emission and health risk assessment during composting of organic fraction of municipal solid waste. J. Hazard. Mater. 2017, 327, 35–43. [Google Scholar] [CrossRef] [PubMed]
  7. Huang, B.; Lei, C.; Wei, C.; Zeng, G. Chlorinated volatile organic compounds (Cl-VOCs) in environment—Sources, potential human health impacts, and current remediation technologies. Environ. Int. 2014, 71, 118–138. [Google Scholar] [CrossRef] [PubMed]
  8. Lin, F.; Zhang, Z.; Li, N.; Yan, B.; He, C.; Hao, Z.; Chen, G. How to achieve complete elimination of Cl-VOCs: A critical review on byproducts formation and inhibition strategies during catalytic oxidation. Chem. Eng. J. 2021, 404, 126534. [Google Scholar] [CrossRef]
  9. Sun, P.; Wang, W.; Dai, X.; Weng, X.; Wu, Z. Mechanism study on catalytic oxidation of chlorobenzene over MnxCe1−xO2/H-ZSM5 catalysts under dry and humid conditions. Appl. Catal. B Environ. 2016, 198, 389–397. [Google Scholar] [CrossRef]
  10. Yang, P.; Yang, S.; Shi, Z.; Meng, Z.; Zhou, R. Deep oxidation of chlorinated VOCs over CeO2-based transition metal mixed oxide catalysts. Appl. Catal. B Environ. 2015, 162, 227–235. [Google Scholar] [CrossRef]
  11. Yang, P.; Shi, Z.; Yang, S.; Zhou, R. High catalytic performances of CeO2–CrOx catalysts for chlorinated VOCs elimination. Chem. Eng. Sci. 2015, 126, 361–369. [Google Scholar] [CrossRef]
  12. Chen, Z.; Mao, J.; Zhou, R. Preparation of size-controlled Pt supported on Al2O3 nanocatalysts for deep catalytic oxidation of benzene at lower temperature. Appl. Surf. Sci. 2019, 465, 15–22. [Google Scholar] [CrossRef]
  13. Guo, Y.; Gao, Y.; Li, X.; Zhuang, G.; Wang, K.; Zheng, Y.; Sun, D.; Huang, J.; Li, Q. Catalytic benzene oxidation by biogenic Pd nanoparticles over 3D-ordered mesoporous CeO2. Chem. Eng. J. 2019, 362, 41–52. [Google Scholar] [CrossRef]
  14. Topka, P.; Delaigle, R.; Kaluža, L.; Gaigneaux, E.M. Performance of platinum and gold catalysts supported on ceria-zirconia mixed oxide in the oxidation of chlorobenzene. Catal. Today 2015, 253, 172–177. [Google Scholar] [CrossRef]
  15. Shi, Y.; Wang, J.; Zhou, R. Pt-support interaction and nanoparticle size effect in Pt/CeO2-TiO2 catalysts for low temperature VOCs removal. Chemosphere 2021, 265, 129127. [Google Scholar] [CrossRef] [PubMed]
  16. Jia, H.; Xing, Y.; Zhang, L.; Zhang, W.; Wang, J.; Zhang, H.; Su, W. Progress of catalytic oxidation of typical chlorined volatile organic compounds (CVOCs): A review. Sci. Total Environ. 2023, 865, 161063. [Google Scholar] [CrossRef]
  17. Sun, W.; Gong, B.; Pan, J.; Wang, Y.; Xia, H.; Zhang, H.; Dai, Q.; Wang, L.; Wang, X. Catalytic combustion of CVOCs over Cr Ti-oxide catalysts. J. Catal. 2020, 391, 132–144. [Google Scholar] [CrossRef]
  18. Feng, X.; Tian, M.; He, C.; Li, L.; Shi, J.-W.; Yu, Y.; Cheng, J. Yolk-shell-like mesoporous CoCrOx with superior activity and chlorine resistance in dichloromethane destruction. Appl. Catal. B Environ. 2020, 264, 118493. [Google Scholar] [CrossRef]
  19. Ye, N.; Li, Y.; Yang, Z.; Zheng, J.; Zuo, S. Rare earth modified kaolin-based Cr2O3 catalysts for catalytic combustion of chlorobenzene. Appl. Catal. A Gen. 2019, 579, 44–51. [Google Scholar] [CrossRef]
  20. Zhang, X.; Liu, Y.; Deng, J.; Jing, L.; Yu, X.; Han, Z.; Dai, H. Effect of transition metal oxide doping on catalytic activity of titania for the oxidation of 1,2-dichloroethane. Catal. Today 2021, 375, 623–634. [Google Scholar] [CrossRef]
  21. Wang, L.; Yu, X.; Wei, Y.; Liu, J.; Zhao, Z. Research advances of rare earth catalysts for catalytic purification of vehicle exhausts—Commemorating the 100th anniversary of the birth of Academician Guangxian Xu. J. Rare Earths 2021, 39, 1151–1180. [Google Scholar] [CrossRef]
  22. Borchert, P.S.Y.; Wilhelm, M.; Borchert, H.; Baumer, M. Nanostructured Praseodymium Oxide: Preparation, Structure, and Catalytic Properties. J. Phys. Chem. C 2008, 112, 3054–3063. [Google Scholar] [CrossRef]
  23. Sonström, P.; Birkenstock, J.; Borchert, Y.; Schilinsky, L.; Behrend, P.; Gries, K.; Müller, K.; Rosenauer, A.; Bäumer, M. Nanostructured Praseodymium Oxide: Correlation Between Phase Transitions and Catalytic Activity. ChemCatChem 2010, 2, 694–704. [Google Scholar] [CrossRef]
  24. Chen, X.; Chen, X.; Cai, S.; Chen, J.; Xu, W.; Jia, H.; Chen, J. Catalytic combustion of toluene over mesoporous Cr2O3-supported platinum catalysts prepared by in situ pyrolysis of MOFs. Chem. Eng. J. 2018, 334, 768–779. [Google Scholar] [CrossRef]
  25. Li, H.; Pan, Q.; Liu, J.; Liu, W.; Li, Q.; Wang, L.; Wang, Z. Synthesis and catalytic properties of praseodymium oxide (Pr6O11) nanorods for diesel soot oxidation. J. Environ. Chem. Eng. 2023, 11, 109152. [Google Scholar] [CrossRef]
  26. Zhou, G.; He, X.; Liu, S.; Xie, H.; Fu, M. Phenyl VOCs catalytic combustion on supported CoMn/AC oxide catalyst. J. Ind. Eng. Chem. 2015, 21, 932–941. [Google Scholar] [CrossRef]
  27. Liu, X.; Wang, J.; Zeng, J.; Wang, X.; Zhu, T. Catalytic oxidation of toluene over a porous Co3O4-supported ruthenium catalyst. RSC Adv. 2015, 5, 52066–52071. [Google Scholar] [CrossRef]
  28. Santos, V.P.; Soares, O.S.G.P.; Bakker, J.J.W.; Pereira, M.F.R.; Órfão, J.J.M.; Gascon, J.; Kapteijn, F.; Figueiredo, J.L. Structural and chemical disorder of cryptomelane promoted by alkali doping: Influence on catalytic properties. J. Catal. 2012, 293, 165–174. [Google Scholar] [CrossRef]
  29. Zhang, C.; Wang, C.; Zhan, W.; Guo, Y.; Guo, Y.; Lu, G.; Baylet, A.; Giroir-Fendler, A. Catalytic oxidation of vinyl chloride emission over LaMnO3 and LaB0.2Mn0.8O3 (B=Co, Ni, Fe) catalysts. Appl. Catal. B Environ. 2013, 129, 509–516. [Google Scholar] [CrossRef]
  30. Tang, W.; Wu, X.; Chen, Y. Co-templating synthesis of mesoporous hollow silica spheres and their application in catalytic oxidation with low Pt loading. Mater. Lett. 2016, 168, 111–115. [Google Scholar] [CrossRef]
  31. Huang, Y.; Tian, M.; Jiang, Z.; Ma, M.; Chen, C.; Xu, H.; Zhang, J.; Albilali, R.; He, C. Inserting Cr2O3 dramatically promotes RuO2/TiO2 catalyst for low-temperature 1,2-dichloroethane deep destruction: Catalytic performance and synergy mechanism. Appl. Catal. B Environ. 2022, 304, 121002. [Google Scholar] [CrossRef]
  32. Tian, M.; Guo, X.; Dong, R.; Guo, Z.; Shi, J.; Yu, Y.; Cheng, M.; Albilali, R.; He, C. Insight into the boosted catalytic performance and chlorine resistance of nanosphere-like meso-macroporous CrOx/MnCo3Ox for 1,2-dichloroethane destruction. Appl. Catal. B Environ. 2019, 259, 118018. [Google Scholar] [CrossRef]
  33. Liu, L.; Xu, K.; Su, S.; He, L.; Qing, M.; Chi, H.; Liu, T.; Hu, S.; Wang, Y.; Xiang, J. Efficient Sm modified Mn/TiO2 catalysts for selective catalytic reduction of NO with NH3 at low temperature. Appl. Catal. A Gen. 2020, 592, 117413. [Google Scholar] [CrossRef]
  34. Yu, C.; Huang, B.; Dong, L.; Chen, F.; Yang, Y.; Fan, Y.; Yang, Y.; Liu, X.; Wang, X. Effect of Pr/Ce addition on the catalytic performance and SO2 resistance of highly dispersed MnOx/SAPO-34 catalyst for NH3-SCR at low temperature. Chem. Eng. J. 2017, 316, 1059–1068. [Google Scholar] [CrossRef]
  35. Deng, Y.; Tian, P.; Liu, S.; He, H.; Wang, Y.; Ouyang, L.; Yuan, S. Enhanced catalytic performance of atomically dispersed Pd on Pr-doped CeO2 nanorod in CO oxidation. J. Hazard. Mater. 2022, 426, 127793. [Google Scholar] [CrossRef]
  36. Gu, Q.; Wang, L.; Wang, Y.; Li, X. Effect of praseodymium substitution on La1−xPrxMnO3 (x = 0–0.4) perovskites and catalytic activity for NO oxidation. J. Phys. Chem. Solids 2019, 133, 52–58. [Google Scholar] [CrossRef]
  37. Zhang, Q.; Zhao, W.; Su, H.; Sun, X.; Sun, L.; Zhao, L.; Qi, C. The advantage of praseodymium doped TiO2 composite as the support of Au catalysts for CO oxidation in CO2-rich atmosphere. Appl. Catal. A Gen. 2023, 661, 119261. [Google Scholar] [CrossRef]
  38. Tankov, I.; Pawelec, B.; Arishtirova, K.; Damyanova, S. Structure and surface properties of praseodymium modified alumina. Appl. Surf. Sci. 2011, 258, 278–284. [Google Scholar] [CrossRef]
  39. Jin, Q.; Shen, Y.; Zhu, S.; Liu, Q.; Li, X.; Yan, W. Effect of praseodymium additive on CeO2(ZrO2)/TiO2 for selective catalytic reduction of NO by NH3. J. Rare Earths 2016, 34, 1111–1120. [Google Scholar] [CrossRef]
  40. Yang, P.; Zuo, S.; Shi, Z.; Tao, F.; Zhou, R. Elimination of 1,2-dichloroethane over (Ce,Cr)xO2/MOy catalysts (M = Ti, V, Nb, Mo, W and La). Appl. Catal. B Environ. 2016, 191, 53–61. [Google Scholar] [CrossRef]
  41. Ye, N.; Zheng, J.; Xie, K.; Jiang, B.; Zuo, S. Performance of mesoporous CeO2-Cr2O3 mixed metal oxides applied to benzene catalytic combustion. J. Rare Earths 2023, 41, 889–895. [Google Scholar] [CrossRef]
  42. Wu, Z.; Deng, J.; Xie, S.; Yang, H.; Zhao, X.; Zhang, K.; Lin, H.; Dai, H.; Guo, G. Mesoporous Cr2O3-supported Au–Pd nanoparticles: High-performance catalysts for the oxidation of toluene. Microporous Mesoporous Mater. 2016, 224, 311–322. [Google Scholar] [CrossRef]
  43. Storaro, L.; Ganzerla, R.; Lenarda, M.; Zanoni, R.; Lopez, A.J.; Olivera-Pastor, P.; Castellh, E.R. Catalytic behavior of chromia and chromium-doped alumina pillared clay materials for the vapor phase deep oxidation of chlorinated hydrocarbons. J. Mol. Catal. A Chem. 1997, 115, 329–338. [Google Scholar] [CrossRef]
  44. Popescu, I.; Martínez-Munuera, J.C.; García-García, A.; Marcu, I.-C. Insights into the relationship between the catalytic oxidation performances of Ce-Pr mixed oxides and their semiconductive and redox properties. Appl. Catal. A Gen. 2019, 578, 30–39. [Google Scholar] [CrossRef]
  45. Zhao, H.; Dong, F.; Han, W.; Tang, Z. Study of Morphology-Dependent and Crystal-Plane Effects of CeMnOx Catalysts for 1,2-Dichlorobenzene Catalytic Elimination. Ind. Eng. Chem. Res. 2019, 58, 18055–18064. [Google Scholar] [CrossRef]
  46. Cao, S.; Wang, H.; Yu, F.; Shi, M.; Chen, S.; Weng, X.; Liu, Y.; Wu, Z. Catalyst performance and mechanism of catalytic combustion of dichloromethane (CH2Cl2) over Ce doped TiO2. J. Colloid Interface Sci. 2016, 463, 233–241. [Google Scholar] [CrossRef] [PubMed]
  47. Crapse, J.; Pappireddi, N.; Gupta, M.; Shvartsman, S.Y.; Wieschaus, E.; Wuhr, M. Evaluating the Arrhenius equation for developmental processes. Mol. Syst. Biol. 2021, 17, e9895. [Google Scholar] [CrossRef]
  48. Duan, X.; Zhao, T.; Yang, Z.; Niu, B.; Li, G.; Li, B.; Zhang, Z.; Cheng, J.; Hao, Z. Oxygen activation-boosted manganese oxide with unique interface for chlorobenzene oxidation: Unveiling the roles and dynamic variation of active oxygen species in heterogeneous catalytic oxidation process. Appl. Catal. B Environ. 2023, 331, 122719. [Google Scholar] [CrossRef]
  49. Rico-Barragán, A.A.; Álvarez, J.R.; Hernández-Fernández, E.; Rodríguez-Hernández, J.; Garza-Navarro, M.A.; Dávila-Guzmán, N.E. Green synthesis of metal-organic framework MIL-101(Cr)—An assessment by quantitative green chemistry metrics. Polyhedron 2022, 225, 116052. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of samples: (a) M-Cr; (b) P-Cr; (c) M-20PrCr; (d) M-30PrCr; (e) M-40PrCr; (f) P-Pr.
Figure 1. XRD patterns of samples: (a) M-Cr; (b) P-Cr; (c) M-20PrCr; (d) M-30PrCr; (e) M-40PrCr; (f) P-Pr.
Molecules 29 03417 g001
Figure 2. Nitrogen adsorption-desorption isotherms of samples: (a) M-Cr; (b) P-Cr; (c) M-20PrCr; (d) M-30PrCr; (e) M-40PrCr; (f) P-Pr; (g) P-30PrCr.
Figure 2. Nitrogen adsorption-desorption isotherms of samples: (a) M-Cr; (b) P-Cr; (c) M-20PrCr; (d) M-30PrCr; (e) M-40PrCr; (f) P-Pr; (g) P-30PrCr.
Molecules 29 03417 g002
Figure 3. Pore-size distributions of samples prepared via precipitation method and pyrolysis method.
Figure 3. Pore-size distributions of samples prepared via precipitation method and pyrolysis method.
Molecules 29 03417 g003
Figure 4. SEM images of the PrCr catalysts with different Pr/Cr ratios (ah): (a) MIL-101(Cr), (b) M-Cr, (c) P-Cr, (d) M-20PrCr, (e) M-30PrCr, (f) M-40PrCr, (g) P-Pr, (h) P-30PrCr, and EDS mapping result for M-30PrCr: (i) selected section, (j) Cr (Lα1), (k) Pr (Mα), and (l) O (Kα1).
Figure 4. SEM images of the PrCr catalysts with different Pr/Cr ratios (ah): (a) MIL-101(Cr), (b) M-Cr, (c) P-Cr, (d) M-20PrCr, (e) M-30PrCr, (f) M-40PrCr, (g) P-Pr, (h) P-30PrCr, and EDS mapping result for M-30PrCr: (i) selected section, (j) Cr (Lα1), (k) Pr (Mα), and (l) O (Kα1).
Molecules 29 03417 g004
Figure 5. The XPS spectra of the catalysts: (a) Cr 2p; (b) O 1s.
Figure 5. The XPS spectra of the catalysts: (a) Cr 2p; (b) O 1s.
Molecules 29 03417 g005
Figure 6. H2-TPR curves of praseodymium chromium composite oxides catalysts: (a) M-Cr, (b) P-Cr, (c) P-Pr, (d) M-20PrCr, (e) M-30PrCr, (f) M-40PrCr.
Figure 6. H2-TPR curves of praseodymium chromium composite oxides catalysts: (a) M-Cr, (b) P-Cr, (c) P-Pr, (d) M-20PrCr, (e) M-30PrCr, (f) M-40PrCr.
Molecules 29 03417 g006
Figure 7. NH3-TPD curves of praseodymium chromium composite oxides catalysts: (a) M-Cr, (b) P-Cr, (c) P-Pr, (d) M-20PrCr, (e) M-30PrCr, (f) M-40PrCr.
Figure 7. NH3-TPD curves of praseodymium chromium composite oxides catalysts: (a) M-Cr, (b) P-Cr, (c) P-Pr, (d) M-20PrCr, (e) M-30PrCr, (f) M-40PrCr.
Molecules 29 03417 g007
Figure 8. The catalytic performance for 1,2-DCE oxidation over the prepared catalysts: (a) Conversion rate, (b) Complete oxidation rate, (c) Apparent activation energy.
Figure 8. The catalytic performance for 1,2-DCE oxidation over the prepared catalysts: (a) Conversion rate, (b) Complete oxidation rate, (c) Apparent activation energy.
Molecules 29 03417 g008
Figure 9. Durability test for catalytic oxidation of 1,2-DCE over PrCrOx catalysts at 290 °C.
Figure 9. Durability test for catalytic oxidation of 1,2-DCE over PrCrOx catalysts at 290 °C.
Molecules 29 03417 g009
Figure 10. XRD patterns of used catalysts: (a) M-Cr; (b) used M-Cr; (c) M-30PrCr; (d) used M-30PrCr.
Figure 10. XRD patterns of used catalysts: (a) M-Cr; (b) used M-Cr; (c) M-30PrCr; (d) used M-30PrCr.
Molecules 29 03417 g010
Figure 11. H2-TPR curves of used catalysts: (a) M-Cr, (b) used M-Cr, (c) M-30PrCr, (d) used M-30PrCr.
Figure 11. H2-TPR curves of used catalysts: (a) M-Cr, (b) used M-Cr, (c) M-30PrCr, (d) used M-30PrCr.
Molecules 29 03417 g011
Figure 12. Synthetic process of M-xPrCr catalysts via in situ pyrolysis method.
Figure 12. Synthetic process of M-xPrCr catalysts via in situ pyrolysis method.
Molecules 29 03417 g012
Table 1. Structural properties and catalytic performance of synthesized catalysts.
Table 1. Structural properties and catalytic performance of synthesized catalysts.
SamplesSBET a
(m2/g)
Pore Volume b (cm3/g)Average Pore Size c (nm) T50 d
(°C)
T90 d
(°C)
Cr
Content e (%)
Pr
Content e
(%)
M-Cr135.240.360.86320>400--
M-20PrCr634.420.370.7130231216.2483.53
M-30PrCr699.310.490.7525528623.4576.41
M-40PrCr648.520.440.7327129828.2971.58
P-30PrCr44.530.3720.24274>40016.9682.99
P-Cr57.990.201.16371>400--
P-Pr19.840.2125.18>400>400--
a BET specific surface area. b Total pore volume estimated at p/p0 = 0.99. c BJH pore diameter calculated from the absorption branch. d Temperatures at which 50% and 90% conversions of 1,2-dichloroethane occur. e The elemental content was quantified using ICP-MS.
Table 2. Textural property of prepared catalysts.
Table 2. Textural property of prepared catalysts.
SampleMolar Ratio aH2 Consumption b (mmol·g−1)Acid Sites
Oβ/(Oα+Oβ)Cr6+/Cr3+ Pwa cPma dPsa e
M-Cr0.320.977.109565513
P-Cr0.452.165.79514331
P-Pr- f-14.1913019259
M-20PrCr0.741.2310.61213112585
M-30PrCr0.811.4827.24387122600
M-40PrCr0.791.4128.37363137626
P-30PrCr0.304.07----
a Oα and Oβ are the surface adsorbed oxygen species and lattice oxygen, respectively; b H2 consumption amount below 700 °C over all samples; Desorption peak areas of NH3 absorbed on c weak acid sites, d medium acid sites and e strong acid sites; f The data that are irrelevant or have not been validated are indicated with a hyphen (-).
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

Zhu, P.; Hu, Z.; Chen, S. Praseodymium-Doped Cr2O3 Prepared by In Situ Pyrolysis of MIL-101(Cr) for Highly Efficient Catalytic Oxidation of 1,2-Dichloroethane. Molecules 2024, 29, 3417. https://doi.org/10.3390/molecules29143417

AMA Style

Zhu P, Hu Z, Chen S. Praseodymium-Doped Cr2O3 Prepared by In Situ Pyrolysis of MIL-101(Cr) for Highly Efficient Catalytic Oxidation of 1,2-Dichloroethane. Molecules. 2024; 29(14):3417. https://doi.org/10.3390/molecules29143417

Chicago/Turabian Style

Zhu, Pengfei, Zhaoxia Hu, and Shouwen Chen. 2024. "Praseodymium-Doped Cr2O3 Prepared by In Situ Pyrolysis of MIL-101(Cr) for Highly Efficient Catalytic Oxidation of 1,2-Dichloroethane" Molecules 29, no. 14: 3417. https://doi.org/10.3390/molecules29143417

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop