Next Article in Journal
Strategies to Prepare Chitin and Chitosan-Based Bioactive Structures Aided by Deep Eutectic Solvents: A Review
Previous Article in Journal
Green Synthesis of Manganese-Cobalt Oxyhydroxide Nanocomposite as Electrocatalyst for Enhanced Oxygen Evolution Reaction in Alkaline Medium
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

CO2 Oxidative Dehydrogenation of Propane to Olefin over Cr-M (M = Zr, La, Fe) Based Zeolite Catalyst

College of Environmental Science and Engineering, Beijing University of Technology, Beijing 100124, China
*
Author to whom correspondence should be addressed.
Submission received: 12 May 2024 / Revised: 29 May 2024 / Accepted: 4 June 2024 / Published: 7 June 2024
(This article belongs to the Special Issue Zeolites and Zeolite-Based Catalysis)

Abstract

:
CO2 oxidative dehydrogenation of propane (CO2-ODHP), being not only favorable for olefin production but also beneficial for CO2 emission control, has recently attracted great attention. Here, a series of single metal (Cr) and bimetal (Zr, La, Fe) modified ZSM-5 zeolites were prepared via an impregnation method. It was found that the bimetal modified ZSM-5 possessed much higher C3H8 and CO2 conversion than that of monometallic modified Cr3%-ZSM-5 (Cr3%-Z5), especially for Cr3%Zr2%-ZSM-5 (Cr3%Zr2%-Z5), which displayed the highest activity (65.4%) and olefin yield (1.65 × 103 μmol· g c a t 1 h−1). Various characterizations were performed, including XRD, N2 adsorption-desorption, H2-TPR, Raman, XPS, HAAD-STEM, and TEM. It was revealed that Zr not only favored an improvement in the redox ability of Cr, but also contributed to the surface dispersion of loaded Cr species, constituting two major reasons explaining the superior activity of Cr3%Zr2%-Z5. To further improve CO2-ODHP catalytic behavior, a series of Cr3%-ZSM-5@SBA-15-n composite zeolite catalysts with diverse (ZSM-5/SBA-15) mass ratios were prepared (Cr3%-ZS-n, n = 0.5, 2, 6, 16), which screened out an optimum mass ratio of six. Based on this, the Cr3%Zr2%-ZS-6 compound was further prepared, and it eventually achieved even higher CO2-ODHP activity (76.9%) and olefin yield (1.72 × 103 μmol· g c a t 1 h−1). Finally, the CO2-ODHP reaction mechanism was further investigated using in situ FTIR, and it was found that the reaction followed the Mars–van Krevelen mechanism, wherein CO2 participated in the reaction through generation of polydentate carbonates. The Cr6+ constituted as the active site, which was reduced to Cr3+ after the dihydrogen reaction, and was then further oxidized into Cr6+ by CO2, forming polydentate carbonates, and thus cycling the reactive species Cr6+. Additionally, assisted by a Brönsted acid site (favoring breaking of the C-C bond), C2H4 and CH4 were produced.

Graphical Abstract

1. Introduction

Low-carbon olefins, especially ethylene and propylene, can be used as basic raw materials in the synthesis of polymeric materials and fine chemicals, which play an important role in the development of the country’s economy [1,2,3,4]. However, given the scarcity of oil resources and rising oil prices, the technology for producing olefins from oil needs further innovation. In recent years, the global demand for olefins has gradually increased [5]. Compared with the traditional olefin production process of cracking and splitting petroleum products, the use of low-carbon alkanes (e.g., propane) to produce olefins has the advantages of low production costs, high yields, and efficient use of resources. Olefin production from low-carbon alkanes can be divided into two types: direct alkane to olefin production, and alkane oxidation to olefin. The oxidative dehydrogenation of alkanes is an exothermic reaction that is not limited by thermodynamics [6,7]. By utilizing CO2 as an oxidant, the conversion rate of alkane, as well as the selectivity of olefins, can be improved [4,8,9]. Moreover, it also favors the reduction of associated CO2 emissions, which is a typical greenhouse gas that contributes to the global warming problem.
The oxidative dehydrogenation of alkanes (e.g., propane and ethane) by CO2 (CO2-ODHP) has been widely investigated, including through the use of Pt [10,11], Cr [12], Co [13], Ga [14,15], Mo [2,16], and B-based [17] catalysts. In comparison to other catalysts, Cr-based catalysts have the advantages of low cost and good reactivity [12]. However, problems such as carbon accumulation and easy accumulation of Cr particles occur during the catalytic reaction. The introduction of auxiliary metals into the catalyst and the selection of suitable supports to improve the dispersion of Cr species constitute two approaches to develop highly efficient Cr-based catalysts. For example, taking advantage of the large surface area and ordered pore structure of ZSM-5, with the pore size of 0.56 × 0.51 nm being larger than the molecule dynamics diameter (0.47 nm) of propane molecules, Wang et al. [18] investigated propane dehydrogenation over Cr loaded ZSM-5 (Cr-ZSM-5) zeolite, the Brönsted acid sites of which have also been reported to be favorable for the adsorption and reaction of propane. Igonina et al. [19] impregnated 5 wt.% Cr on MCM-41 using the first impregnation method, which resulted in 20% conversion of C3H8 and 76% selectivity of C3H6 at 650 °C. Singh et al. [20] used ZrO2 as a support loaded with 2.5 wt.% Cr and achieved 16% conversion of C3H8 and 86.6% selectivity of C3H6 at 550 °C. Zhang et al. [21] introduced 5 wt.% Sn into an Al2O3 catalyst loaded with 3 wt.% Cr in order to improve the stability of the catalyst. The initial conversion of C3H8 in the catalyst was 40%, and after 10 h of reaction, the conversion of C3H8 was stable at 32%. Moreover, one type of composite molecular sieve of ZSM-5@SBA-15 was synthesized by adding ZSM-5 into the mother liquor of SBA-15 and was then applied for propane dehydrogenation [22,23], displaying superior catalytic activity in comparison to that of the single ZSM-5 molecular sieve.
Inspired by the above literature, a series of Cr-based ZSM-5 catalysts with different additives (Zr, La, Fe) were prepared by the impregnation method, with the aim of further improving the intrinsic CO2-ODHP activity of Cr-ZSM-5. Additionally, in order to further improve the reaction performance, a series composite molecular sieve catalyst (Cr3%-ZSM-5@SBA-15 and Cr3%Zr2%-ZSM-5@SAB-15) was further prepared based on the hydrothermal synthesis method, in which the specific mass ratios of ZSM-5 to SBA-15 (Cr3%-ZS-n, n = 0.5, 2, 6, 16, representing the mass ratio) was detailed and adjusted, eventually achieving a best mass ratio of 6:1 and showing C3H8 conversion of 76.9% and an olefin yield of (1.72 × 103 μmol· g c a t 1 ·h−1) from Cr3%Zr2%-ZS-6. The structure–activity relationship was further illustrated by a series of characterizations (XRD, N2 adsorption-desorption, H2-TPR, Raman, XPS, HAAD-STEM, and TEM). The present work has developed an efficient Cr3%Zr2%-Z5 catalyst for the CO2-driven oxidative dehydrogenation of propane (ODHP). Moreover, a Cr3%Zr2%-ZS-6 composite zeolite catalyst with promoted ODHP activity has been further developed. The present work favors the development of other highly efficient CO2-ODHP catalyst designs.

2. Results and Discussion

2.1. Characterization of Catalysts

X-ray diffraction (XRD) was used to characterize the crystal structure of the synthesized samples. Figure 1a displays the large angle XRD characterization results of the synthesized catalyst samples. As can be seen, both the [Cr3%-Z5, Cr3%M2%-Z5 (M = La, Zr, Fe)] and composite catalyst samples (Cr3%-ZS-n and Cr3%Zr2%-ZS-6) display the characteristic MFI crystal structure at the 2θ of 8.0°, 8.9°, and 23.7° [24,25,26]. The loaded metal species did not significantly affect the associated crystal structure during the synthesis process, and no obvious peaks of metal oxide species, such as CrO3 or Cr2O3, were observed, probably due to their good dispersion over the zeolite samples [27]. Additionally, the diffraction peaks of MFI became sharp as the mass ratio of mZSM-5/mSBA-15 increased for the Cr3%-ZS-n composite. Figure 1b displays the small-angle XRD of the composite catalysts, where the characteristic diffraction patterns of SBA-15 at the 2θ of 0.8°, 1.5°, and 1.8°, corresponding to the crystalline facets of (100), (110), and (200), respectively, can be observed [27,28,29]. This finding indicates the successful synthesis of SBA-15. As noted, the intensity of the diffraction pattern of SBA-15 gradually decreased as the mass ratio of mZSM-5/mSBA-15 increased. Figure 1c displays the N2 adsorption-desorption curves of Cr3%-ZS-6, which displayed a typical IV curve [23], further verifying the successful synthesis of composite zeolite catalysts with mesoporous structures. The specific surface areas of the synthesized samples were characterized using the N2 adsorption-desorption method, as listed in Table 1. The addition of the second metallic element of M (M = Zr, La, Fe) did not seriously affect the specific surface area of Cr3%M2%-Z5 in comparison to that of Cr3%-Z5 (367.7 m2 g−1). However, much higher surface areas can be achieved for the composite zeolite catalysts (Cr3%-ZS-n) relative to those of Cr3%-ZSM-5 and Cr3%M2%-Z5, especially for Cr3%-ZS-6, which showed the highest value of 505.7 m2/g. This result indicates that encapsulation by SBA-15 can efficiently improve specific surface area, which would be favorable for the CO2-ODHP reaction.
H2-TPR was conducted to characterize the specific chemical states of the loaded Cr and M (Zr, La, Fe) species over the synthesized catalyst samples (see Figure 1d). For Cr3%-Z5, the broad peak centered at T of 300–400 °C can be attributed to the reduction of CrO3 species, and the weak reduction peak at T of 400–500 °C can be attributed to the reduction of Cr6+ [30,31]. As for the Cr3%M2%-Z5 samples, in addition to the abovementioned reduction peaks of CrO3 and Cr6+, the additional reduction peaks of the (Zr, La, Fe) species could also be clearly identified. Specifically, for the Cr3%Fe2%-ZSM-5 sample, the reduction peak around 400 °C was related to the reduction of Fe2O3 to Fe3O4; the peak around 500 °C was attributed to the reduction of Fe3O4 to FeO; and the peak around 600 °C was ascribed to the reduction of FeO to Fe [32]. As for Cr3%Zr2%-Z5 and Cr3%La2%-Z5, the additional reduction peaks at T of 300–400 °C and 400–500 °C were related to the reductions of ZrO2 and La2O3, respectively [33,34]. As noted above, after the introduction of Zr, the reduction peaks of Cr (CrO3 → Cr2O3) were shifted to lower temperatures (450 → 290 °C), and also developed a larger reduction peak area. This was closely related to the fact that ZrO2, which is a P-type semiconductor, formed a strong interaction with the active metal Cr, favoring the reduction of CrO3, which is conducive to CO2-ODHP. Similarly, in the Cr3%La2%-Z5 and Cr3%Fe2%-Z5 scenarios, it is evident that the presence of La and Fe additives also promoted the reduction of CrO3 to Cr2O3, although to a lesser extent compared to Zr. This constitutes one of the major reasons leading to the lower CO2-ODOH activity of Cr3%Zr2%-Z5 than that observed for Cr3%La2%-Z5 and Cr3%Fe2%-Z5 (as will be stated later). In addition to this, for the composite zeolite catalyst Cr3%Zr2%-ZS-6, it was found that, interestingly, the reduction peaks of CrO3 and Cr6+ further shifted to lower temperatures (450 → 200 °C) due to encapsulation by SBA-15. This favored the dispersion of the Cr species. This is also one of the main reasons for the further improvement in the CO2-ODH activity in this sample compared to that of Cr3%Zr2%-Z5.
Raman spectroscopy was further used to characterize the loaded Cr species over the synthesized samples, as shown in Figure 1e, where the peak centered at 550 cm−1 can be attributed to the Cr-O vibration of Cr2O3 species over Cr3%-Z5. In addition to the peak at 550 cm−1, the Raman peaks corresponding to Cr2O72− (350–400 cm−1 of Cr3%Zr2%-Z5), Cr3O102− (850 cm−1 of Cr3%La2%-Z5) and Cr4O132− (900 cm−1 of Cr3%Zr2%-Z5) appeared after the introduction of Zr and La [35]. Only one Raman peak of Cr2O72− (350–400 cm−1) emerged over the composite Cr3%-ZS-6. The above results indicate that the addition of the second metal, as well as encapsulation by SBA-15 (generating composite catalyst), could significantly affect the valence state of CrOx. According to a previous report in the literature [36], the formed Cr2O72− and Cr4O132− species would be more active for the ODHP reaction relative to that of Cr2O3. Thus, the results there (Figure 1) give a further indication of the much higher CO2-ODHP of Cr3%Zr2%-Z5 and Cr3%-ZS-6 relative to that of Cr3%-Z5. As noted above, no characteristic Raman peaks of CrOx species could be found over Cr3%Fe2%-Z5, which was probably related to the tendency of FeOx to react with oxygen in air to form an oxide film, which eventually interferes with Raman spectroscopic analysis.
The valence states of the active metal Cr over the synthesized samples were further analyzed by XPS (Figure 1f), with the derived Cr6+/(Cr6++Cr3+) ratios being listed in Table 1. As can be seen, the Cr 2p1/2 for the investigated samples could be split into three peaks, of which the characteristic peaks with electron binding energies at 575.0 eV were attributed to Cr3+, and the two other peaks with electron binding energies of 577.2 and 579.0 eV were attributed to Cr6+ [37]. As listed in Table 1, the Cr6+/(Cr6++Cr3+) ratios of bimetallic modified ZSM-5 samples (68.2–73.9%) were much higher than that of monometallic Cr3%/ZSM-5 (46.9%), indicating that the second metal additives favored the formation of Cr6+ species, which have much higher activity than that of Cr3+. In this regard, the addition of the second metal promoted the Cr6+ ratio relative to the single-metal-promoted Cr3%-Z5, which is another important reason leading to the much higher CO2-ODHP activity. As noted, no XPS signal of Cr species was observed over Cr3%-ZS-6, which could be related to the encapsulation of Cr3%-ZSM-5 by SBA-15, leading to the Cr being undetectable.
Figure 2 shows the HAADF-STEM, HRTEM, and EDS-mapping results of Cr3%-Z5 (Figure 2a–c), Cr3%Zr2%-Z5 (Figure 2d–f) and Cr3%Zr2%-ZS-6 (Figure 2g–i), which provide information on the crystalline structure and surface distribution of the active metal Cr. As compared in Figure 2a,d,g, the Cr agglomerates can be clearly observed over Cr3%-Z5; however, they cannot be seen over the samples of Cr3%Zr2%-Z5 and Cr3%Zr2%-ZS-6. This finding is well supported by the EDS-mapping results shown in Figure 2c,f,i, displaying the aggregated Cr nanoparticles over Cr3%-Z5 and the monodispersed Cr over Cr3%Zr2%-Z5 and Cr3%Zr2%-ZS-6. This finding gives us a clue that the addition of Zr (Figure S1), as well as encapsulation by SBA-15, could favor the dispersion of loaded Cr species. In addition, the obvious lattice stripes at 0.375 nm belonging to the (151) plane of ZSM-5 [38] can be clearly observed for the investigated samples (Figure 2b,e,h), and the regular pore structure of SBA-15 [29], with a clear interface between SBA-15 and ZSM-5, can also be observed in Figure 2h, verifying the successful synthesis of the Cr3%Zr2%-ZS-6 composite.

2.2. Reaction Mechanism Investigation

In situ FTIR was employed to investigate the CO2-ODHP reaction mechanism over the best-performing sample of Cr3%Zr2%-ZS-6 at T of 250–500 °C, as shown in Figure 3. Several intermediate species with IR vibration peaks centered at 1700–1350 cm−1 were identified, including the methyl radical [CH3*] in the range of 1400–1370 cm−1 [34], polydentate carbonate species around 1468 cm−1 [39], H2O around 1640 cm−1, and olefins with the telescopic vibrational peak of C=C around 1650 cm−1 [40]. As noted, the other peaks in the range of 3600–3750 cm−1 belonged to hydroxyl species [OH*], and the 1501 cm−1 peak was attributed to be the telescopic vibrational peak of the Brönsted acid site [41]. Based on the above derived intermediate information, as well as reports from the literature [36], it can be proposed that CO2-ODHP would follow the Mars–van Krevelen mechanism [42]. First, CO2 is adsorbed on the support at a low temperature (T of 250 °C), forming polydentate carbonates (1468 cm−1). The reaction from propane to olefin then begins with an increase in temperature (T > 300 °C), where the polydentate carbonates participate in the reaction with a gradual decrease in the IR signal (1470 cm−1) and a simultaneous gradual increase in the IR signal of olefin (vC=C of 1649 cm−1) and H2O (vH2O of 1625 cm−1). As reported [43], Cr6+ is the activation site for the propane dehydrogenation reaction, and the C-H bond broken to produce radicals of C3H7* (1370–1400 cm−1, Figure 3) and H* is reported to be the rate-determining step (RDS). In addition to C3H6, C2H4 and CH4 were also observed during activity measurement. This can be attributed to the strong C-C bond breaking effect of the Brönsted acid site (around 1540 cm−1) of the ZSM-5 support. The interaction of C3H8 with a Brönsted acid site leads to the breaking of the C-C bond, generating radicals of ethyl (C2H5*) and (CH3*) methyl (1400–1370 cm−1, Figure 3). Subsequently, the broken C-H bond of C2H5* leads to the formation of C2H4 and the dissociated proton H*, which further interacts with the radical of CH3* to produce CH4. Based on the above discussion, we can conclude that the Cr6+ is the activation site for propane dehydrogenation, where CO2 participates in the reaction as polydentate carbonates. Assisted by the Brönsted acid site (favoring C-C bond breaking), C2H4 and CH4 are produced. As noted, after the dehydrogenation reaction, Cr6+ is reduced to Cr3+, which is then further oxidized to Cr6+ by CO2, forming polydentate carbonates and thus cycling the reactive species Cr6+. The CO2/C3H8 adsorption over ZrO2 [111] and CrO3 [110] was further simulated by DFT to further illustrate the role played by Zr during CO2-ODHP. The results are shown in Tables S1 and S2. As can be seen, the adsorption energies of CO2 and C3H8 on the ZrO2 were −0.14 eV and −0.03 eV, and those of CrO3 were −0.09 and −0.06 eV, respectively. Obviously, the ZrO2 exhibited a much lower CO2 adsorption energy (−0.14 eV) than that of CrO3 (−0.09 eV). This finding indicates that the introduced Zr also functioned as the active site during CO2-ODHP, favoring CO2 adsorption over the Cr-based ZSM-5 zeolite.

2.3. Catalytic Activity of Cr-Based Catalysts

Figure 4a,b display the propane conversion, CO2 conversion, olefin selectivity, and olefin yield of the synthesized samples of Cr3%M2%-Z5 (M = Zr, La, Fe), where the Cr loading was 3 wt.% and the additive metal loading was 2 wt.%. The proportion of each product (C3H6, C2H4, C2H6, CH4) is depicted in Figure S2a. As can be seen, adding additional metals, such as Zr, La, and Fe, increased C3H8 conversion and olefin yield; this was accompanied by a slight decrease in olefin selectivity (62.9 → 58.6%). Among the samples, Cr3%Zr2%-Z5 displayed the best CO2-ODHP catalytic behavior, achieving C3H8 conversion of 65.4% and an olefin yield of 1.65 × 103 μmol· g c a t 1 ·h−1. Figure 4c,d display the activity measurement results for the composite catalysts of Cr3%-ZS-n (n = 0.5, 2, 6, 16) and Cr3%Zr2%-ZS-6, wherein the ZSM-5: SBA-15 ratio of 6:1 led to much higher C3H8 conversion (73.7%) and olefin yield (1.69 × 103 μmol· g c a t 1 ·h−1) than those of the other catalysts. The distribution of each element (C3H6, C2H4, C2H6, CH4) in the product is shown in Figure S2b, showing a slight decrease in olefin selectivity up to 52.0%. Moreover, the C3H8 conversion and olefin yield was further improved to 76.9% and 1.72 × 103 μmol· g c a t 1 ·h−1, respectively, over the bimetallic composite molecular sieve of Cr3%Zr2%-ZS-6 (mass ratio of 6.0) due to the superior intrinsic CO2-ODHP activity of Cr3%Zr2%-Z5 compared to that of Cr3%-Z5 (Figure 4a,b). It can be seen from Figure S3 that the activity of the catalysts (Cr3%-Z5, Cr3%Zr2%-Z5, and Cr3%Zr2%-ZS-6) decreased slightly after 180 min of reaction time.

3. Material and Methods

3.1. Preparation of Cr-Based ZSM-5 Zeolite Catalyst

Approximately 0.476 g of Cr(NO3)3·9H2O and a certain amount of nitrate compounds of the additive metals were weighed into a 250 mL tomato-shaped bottle. Then, 50 mL of deionized water and 2 g of ZSM-5 carrier with an Si:Al ratio of 50 were added and stirred in a water bath at 90 °C for 4 h. The tomato-shaped bottle containing the sample was mounted on a spinner to remove water, and was thereafter placed in an oven at 60 °C and dried overnight. Finally, the sample was calcined in a muffle furnace at 550 °C under air atmosphere for 6 h to obtain the Cr3%-Z5 and Cr3%M2%-Z5 (M = Zr, La, Fe) samples.

3.2. Preparation of Cr-Based ZSM-5@SBA-15 Composite Zeolite Catalyst

SBA-15 was prepared by using triblock copolymer P123 as the structure-directing agent and tetraethyl silicate TEOs as the silicon source. First, 0.326 g of P123 was dissolved in 50 mL of 1.6 mol/L hydrochloric acid and stirred for 15 min in a water bath at 40 °C. Then, 0.75 mL of tetraethyl silicate was added and stirred continuously for 40 min to form the mother solution of SBA-15. A certain amount of Cr(NO3)3·9H2O and ZSM-5 molecular sieves (Si:Al = 50) were added to the mother solution of SBA-15 according to the mass ratios of ZSM-5 to SBA-15 of 0.5, 2, 6, and 16, respectively, and stirred for 24 h. After that, the samples were placed in a crystallization tank at 100 °C for 24 h. Then, after washing three times with water and alcohol, centrifugation, drying (60 °C for 12 h), and calcination (550 °C for 6 h), the Cr3%-ZS-n (n = 0.5, 2, 6, 16) was obtained. A similar procedure was also applied during Cr3%Zr2%-ZS-6 preparation, utilizing Zr(NO3)3·5H2O as the Zr precursor.

3.3. Characterizations

X-ray diffraction (XRD) was performed on a D8 Advance instrument (Karlsruhe, Germany) to analyze the crystalline phase structure of the catalyst samples using a Cu Kα radiation source in the range of 5–60°. Specific surface area analysis (BET) was carried out on an Autosorb iQ (Boynton Beach, FL, USA) adsorption apparatus to analyze the specific surface area of the catalyst samples, and N2 adsorption and desorption curves were obtained using an Autosorb iQ adsorption apparatus. The samples were degassed at 300 °C for 6 h. H2 programmed temperature rising reduction (H2-TPR) was carried out on a fully automated multicurrent apparatus, the tp-5080-B (Tianjin, Beijing), and was used to analyze the composition of the metal oxide species of the samples. A 0.1 g sample was first pretreated at 100 °C for 10 min, then cooled to room temperature to level the baseline, and finally heated from 100 to 900 °C at a rate of 10 °C/min. Raman spectroscopy (Raman) was performed on a LabRam HR Evolution Raman spectrometer (Paris, France) in the range of 100 to 1200 cm−1 with a laser wavelength of 532 nm. X-ray photoelectron spectroscopy (XPS) was performed on a Thermo Scientific K-Alpha X-ray photoelectron spectrometer, Waltham, MA, USA, to analyze the material composition of the samples. Transmission electron microscopy (TEM) was performed using a JEOL JEM-F200 transmission electron microscope (Tokyo, Japan) to observe the structural compositions of the catalyst samples, which were first sonicated in ethanol for 60 min and then added drop-wise to a copper grid to prepare the samples for measurement. In situ Fourier infrared spectroscopy (FTIR) was carried out using a BRUKER (Billerica, MA, USA) TENSOR II infrared spectrometer and was used to measure the functional groups and chemical compositions etc. produced during the reaction. Infrared variable temperature experiments were conducted, in which the samples to be measured were pressed and first pretreated with H2 at 500 °C for 60 min, before being cooled to room temperature. Then, a mixed gas comprising 5% C3H8 and 5% CO2, balanced by He (total flow rate of 20 mL/min), was introduced into the sample under temperatures of 250, 300, 350, 400, 450, and 500 °C for 30 min, with the IR signal being collected each minute.

3.4. Catalytic Test

The CO2-ODHP activity measurement was carried out in a continuous flow fixed bed reactor under atmospheric pressure and T of 600 °C. The volume ratio of the feed gas was set to be C3H8:CO2:He = 2:2:16. The gas hourly velocity (GHSV) was 12,000 h−1. As noted, before the reaction, 0.1 g of catalyst (40–60 mesh) was initially pretreated using 5% H2 in He for 60 min at 600 °C. The composition of the reacted gas, as well as the product, were analyzed on-line using an SP-7890 gas chromatograph with a TCD detector (Shimadzu, Kyoto, Japan). The propane conversion, CO2 conversion, olefin selectivity, and olefin yield were calculated by Equations (1)–(4), as shown below:
X C 3 H 8 = C 3 H 8 ( i n ) C 3 H 8 ( o u t ) C 3 H 8 ( i n ) × 100 %
X C O 2 = C O 2 ( i n ) C O 2 ( o u t ) C O 2 ( i n ) × 100 %
S o l e f i n s = 3 C 3 H 6 + 2 C 2 H 4 3 C 3 H 6 + 2 C 2 H 6 + 2 C 2 H 4 + C H 4 × 100 %
Y i e l d   o f   o l e f i n s = P P m × F i n × 60 × 273 × 10 4 T × M c a t × 22.4 × 10 3
where Fin represents the flow rate of gases introduced (C3H8 and CO2); ΔPPm represents the increase in propylene and ethylene; T represents reaction temperature; and Mcat represents the mass of catalyst.

4. Conclusions

In conclusion, a series of Cr-based zeolite catalysts were prepared for CO2-ODHP with the aim of providing a type of candidate catalyst that not only favors propane conversion to olefin, but also contributes to CO2 emission control. It was found that the addition of a second metal (Zr, La, or Fe) enhanced CO2-ODHP activity relative to that of single metal Cr promoted ZSM-5 (Cr3%-Z5), with Zr exhibiting the most pronounced effect. Characterizations using H2-TPR, XPS, Raman spectroscopy, and HAAD-STEM suggested that the addition of Zr improved the redox ability of Cr, increased the Cr6+ species ratio (which acted as active sites for the reaction), and favored the dispersion of Cr species over the ZSM-5 support, which together constituted three major factors leading to the superior CO2-ODHP activity of Cr3%Zr2%-Z5. In addition, one zeolite composite, in which Cr3%Zr2%-Z5 was encapsulated by SBA-15 (mCr3%-ZSM-5/mSBA-15 mass ratio of 6), was prepared using the hydrothermal synthesis method, which further promoted CO2-ODHP activity due to the large surface area of SBA-15 favoring the dispersion of active Cr species. Finally, the CO2-ODHP reaction mechanism was investigated using in situ FTIR. It was found that the reaction followed the Mars–van Krevelen mechanism, wherein the CO2 was initially adsorbed over the Cr3%Zr2%-ZS-6, forming polydentate carbonates, nand then participated in the reaction through reacting with dissociated H* from C3H8 to produce H2O and C3H6. Additionally, assisted by the Brönsted acid site (favoring C-C bond breaking), C2H4 and CH4 were generated. In general, the present work investigated CO2-ODHP over a series of Cr-based zeolite catalysts, based on which one Cr3%Zr2%-ZS-6 candidate composite catalyst was screened out. The present work favors the development of CO2-ODHP techniques.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/catal14060370/s1, Figure S1. EDS-Mapping image of Zr in catalyst Cr3%Zr2%-Z5. Figure S2. Selectivity of the synthesized samples of (a) Cr3%-Z5, Cr3%M2%-Z5 (M = La, Zr, Fe) and (b) Cr3%-ZS-n (n = 0.5, 2, 6, 16) as well as Cr3%Zr2%-ZS-6 (mass ratio of 6). Figure S3. Conversion of propane with catalysts Cr3%-Z5, Cr3%Zr2%-Z5 and Cr3%Zr2%-ZS-6 (reaction time: 180 min). Table S1. Model diagram of CO2 and C3H8 adsorption over CrO3 [110] and ZrO2 [111]. Table S2. Adsorption energy of CO2 and C3H8 over CrO3 [110] and ZrO2 [111]. Ref. [44] is cited in Supplementary Materials.

Author Contributions

Conceptualization, B.C. and N.L.; methodology, M.X.; validation, C.D. and M.X.; resources, B.C. and N.L.; writing, original draft preparation, M.X.; writing, review and editing, N.L.; supervision, B.C. and N.L.; project administration, B.C. and N.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (No. 22178011, 22176006).

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Conflicts of Interest

The authors declare no competing interests.

References

  1. Atanga, M.A.; Rezaei, F.; Jawad, A.; Fitch, M.; Rownaghi, A.A. Oxidative dehydrogenation of propane to propylene with carbon dioxide. Appl. Catal. B Environ. 2018, 220, 429–445. [Google Scholar] [CrossRef]
  2. Bagheri, S.; Julkapli, N.M. Mo3VOx catalyst in biomass conversion: A review in structural evolution and reaction pathways. Int. J. Hydrogen Energy. 2017, 42, 2116–2126. [Google Scholar] [CrossRef]
  3. Li, C.; Wang, G. Dehydrogenation of light alkanes to mono-olefins. Chem. Soc. Rev. 2021, 50, 4359–4381. [Google Scholar] [CrossRef]
  4. Yuan, Y.; Porter, W.N.; Chen, J.G. Comparison of direct and CO2-oxidative dehydrogenation of propane. Trends. Chem. 2023, 5, 840–852. [Google Scholar] [CrossRef]
  5. Shou, H.; Dasari, P.R.; Broekhuis, R.R.; Mondal, A.; Patel, A.; Nguyen, C.; Fickel, D.W. Cracking of Butane on a Pt/H-ZSM-5 Catalyst in the Presence of Hydrogen. Ind. Eng. Chem. Res. 2022, 61, 4763–4773. [Google Scholar] [CrossRef]
  6. Hu, Z.-P.; Yang, D.; Wang, Z.; Yuan, Z.-Y. State-of-the-art catalysts for direct dehydrogenation of propane to propylene. Chin. J. Catal. 2019, 40, 1233–1254. [Google Scholar] [CrossRef]
  7. Qu, Y.; Li, G.; Zhao, T.; Zhang, Z.; Douthwaite, M.; Zhang, J.; Hao, Z. Low-temperature direct dehydrogenation of propane over binary oxide catalysts: Insights into geometric effects and active sites. ACS Sustain. Chem. Eng. 2021, 9, 12755–12765. [Google Scholar] [CrossRef]
  8. Mukherjee, D.; Park, S.-E.; Reddy, B.M. CO2 as a soft oxidant for oxidative dehydrogenation reaction: An eco benign process for industry. J. CO2 Util. 2016, 16, 301–312. [Google Scholar] [CrossRef]
  9. Otroshchenko, T.; Jiang, G.; Kondratenko, V.A.; Rodemerck, U.; Kondratenko, E.V. Current status and perspectives in oxidative, non-oxidative and CO2-mediated dehydrogenation of propane and isobutane over metal oxide catalysts. Chem. Soc. Rev. 2021, 50, 473–527. [Google Scholar] [CrossRef]
  10. Lian, Z.; Si, C.; Jan, F.; Zhi, S.; Li, B. Coke deposition on Pt-based catalysts in propane direct dehydrogenation: Kinetics, suppression, andl eimination. ACS Catal. 2021, 11, 9279–9292. [Google Scholar] [CrossRef]
  11. Zhu, J.; Wang, T.; Xu, X.; Xiao, P.; Li, J. Pt nanoparticles supported on SBA-15: Synthesis, characterization and applications in heterogeneous catalysis. Appl. Catal. B Environ. 2013, 130, 197–217. [Google Scholar] [CrossRef]
  12. Dai, Y.; Gao, X.; Wang, Q.; Wan, X.; Zhou, C.; Yang, Y. Recent progress in heterogeneous metal and metal oxide catalysts for direct dehydrogenation of ethane and propane. Chem. Soc. Rev. 2021, 50, 5590–5630. [Google Scholar] [CrossRef] [PubMed]
  13. Barsan, M.M.; Thyrion, F.C. Kinetic study of oxidative dehydrogenation of propane over Ni-Co molybdate catalyst. Catal. Today 2003, 81, 159–170. [Google Scholar] [CrossRef]
  14. Liu, D.; Cao, L.; Zhang, G.; Zhao, L.; Gao, J.; Xu, C. Catalytic conversion of light alkanes to aromatics by metal-containing HZSM-5 zeolite catalysts-A review. Fuel Process. Technol. 2021, 216, 106770. [Google Scholar] [CrossRef]
  15. Yuan, Y.; Zhao, Z.; Lobo, R.F.; Xu, B. Site diversity and mechanism of metal-exchanged zeolite catalyzed non-oxidative propane dehydrogenation. Adv. Sci. 2023, 10, 2207756. [Google Scholar] [CrossRef] [PubMed]
  16. Madeira, L.M.; Portela, M.F.; Mazzocchia, C. Nickel molybdate catalysts and their use in the selective oxidation of hydrocarbons. Catal. Rev. Sci. Eng. 2004, 46, 53–110. [Google Scholar] [CrossRef]
  17. Love, A.M.; Cendejas, M.C.; Thomas, B.; McDermott, W.P.; Uchupalanun, P.; Kruszynski, C.; Burt, S.P.; Agbi, T.; Rossini, A.J.; Hermans, I. Synthesis and characterization of silica-supported boron oxide catalysts for the oxidative dehydrogenation of propane. J. Phys. Chem. C 2019, 123, 27000–27011. [Google Scholar] [CrossRef]
  18. Silva, B.; Figueiredo, H.; Soares, O.S.G.P.; Pereira, M.F.R.; Figueiredo, J.L.; Lewandowska, A.E.; Banares, M.A.; Neves, I.C.; Tavares, T. Evaluation of ion exchange-modified Y and ZSM5 zeolites in Cr(VI) biosorption and catalytic oxidation of ethyl acetate. Appl. Catal. B Environ. 2012, 117, 406–413. [Google Scholar] [CrossRef]
  19. Igonina, M.; Tedeeva, M.; Kalmykov, K.; Kapustin, G.; Nissenbaum, V.; Mishin, I.; Pribytkov, P.; Dunaev, S.; Kustov, L.; Kustov, A. Properties of CrOx/MCM-41 and its catalytic activity in the reaction of propane dehydrogenation in the presence of CO2. Catalysts 2023, 13, 906. [Google Scholar] [CrossRef]
  20. Singh, R.; Nayak, S.C.; Singh, R.; Deo, G. O2 and CO2 assisted oxidative dehydrogenation of propane using ZrO2 supported vanadium and chromium oxide catalysts. Catal. Today 2024, 432, 114617. [Google Scholar] [CrossRef]
  21. Zhang, L.; Chen, K.; Chen, H.; Han, X.; Liu, C.; Qiao, L.; Wu, W.; Yang, B. Elucidating the promoting advantages and fundamentals for their creation in Sn-modified commercial CrOx/Al2O3 catalyst for propane dehydrogenation. Chem. Eng. J. 2024, 483, 149366. [Google Scholar] [CrossRef]
  22. Razavian, M.; Fatemi, S. Synthesis and application of ZSM-5/SAPO-34 and SAPO-34/ZSM-5 composite systems for propylene yield enhancement in propane dehydrogenation process. Micropor. Mesopor. Mat. 2015, 201, 176–189. [Google Scholar] [CrossRef]
  23. Zhang, L.; Zhao, Y.; Dai, H.; He, H.; Au, C.T. A comparative investigation on the properties of Cr-SBA-15 and CrOx/SBA-15. Catal. Today 2008, 131, 42–54. [Google Scholar] [CrossRef]
  24. Ji, F.; Li, C.; Liu, Y.; Liu, P. Heterogeneous activation of peroxymonosulfate by Cu/ZSM5 for decolorization of Rhodamine B. Sep. Purif. Technol. 2014, 135, 1–6. [Google Scholar] [CrossRef]
  25. Kim, M.-S.; Lee, D.-W.; Chung, S.-H.; Hong, Y.-K.; Lee, S.H.; Oh, S.-H.; Cho, I.-H.; Lee, K.-Y. Oxidation of ammonia to nitrogen over Pt/Fe/ZSM5 catalyst: Influence of catalyst support on the low temperature activity. J. Hazard. Mater. 2012, 237, 153–160. [Google Scholar] [CrossRef] [PubMed]
  26. Saba, T.; Estephane, J.; El Khoury, B.; El Khoury, M.; Khazma, M.; El Zakhem, H.; Aouad, S. Biodiesel production from refined sunflower vegetable oil over KOH/ZSM5 catalysts. Renew. Energy 2016, 90, 301–306. [Google Scholar] [CrossRef]
  27. Tian, J.; Guan, J.; Xu, M.; Qian, S.; Ma, K.; Wan, S.; Zhang, Z.; Xiong, H.; Wang, S.; Wang, Y.; et al. Repairing vacancy defects for stabilization of high surface area hexagonal boron nitride under harsh conditions. Chem. Eng. J. 2023, 475, 146015. [Google Scholar] [CrossRef]
  28. Aboul-Enein, A.A.; Soliman, F.S.; Betiha, M.A. Co-production of hydrogen and carbon nanomaterials using NiCu/SBA15 catalysts by pyrolysis of a wax by-product: Effect of Ni-Cu loading on the catalytic activity. Int. J. Hydrogen Energy 2019, 44, 31104–31120. [Google Scholar] [CrossRef]
  29. Yang, W.; Liu, H.; Li, Y.; Wu, H.; He, D. CO2 reforming of methane to syngas over highly-stable Ni/SBA-15 catalysts prepared by P123-assisted method. Int. J. Hydrogen Energy 2016, 41, 1513–1523. [Google Scholar] [CrossRef]
  30. Al-Awadi, A.S.; El-Toni, A.M.; Al-Zahrani, S.M.; Abasaeed, A.E.; Alhoshan, M.; Khan, A.; Labis, J.P.; Al-Fatesh, A. Role of TiO2 nanoparticle modification of Cr/MCM41 catalyst to enhance Cr-support interaction for oxidative dehydrogenation of ethane with carbon dioxide. Appl. Catal. A Gen. 2019, 584, 117114. [Google Scholar]
  31. Zhang, Q.-H.; Yan, F.; Xia, W.; Liu, C. Study of the coordinative nature of alkylaluminum modified Phillips CrOx/SiO2 catalyst by multinucleai solid-state NMR. Pet. Sci. 2013, 10, 577–583. [Google Scholar] [CrossRef]
  32. Zhang, S.; Zhou, C.-A.; Wang, S.; Qin, Z.; Shu, G.; Wang, C.; Song, L.; Zheng, L.; Wei, X.; Ma, K.; et al. Facilitating CO2 dissociation via Fe doping on supported vanadium oxides for intensified oxidative dehydrogenation of propane. Chem. Eng. J. 2024, 481, 148231. [Google Scholar] [CrossRef]
  33. Ma, Z.; Xu, W.; Wang, Q.; Zhou, Q.; Zhou, J. Highly effective microwave catalytic oxidative dehydrogenation of propane by CO2 over V-La-doped dendritic mesoporous silica-based microwave catalysts. Chem. Eng. J. 2022, 435, 135081. [Google Scholar] [CrossRef]
  34. Zhang, Y.; Yu, Y.; Wang, R.; Dai, Y.; Bao, L.; Li, M.; Zhang, Y.; Liu, Q.; Xiong, D.; Wu, Q.; et al. Identifying the active site and structure-activity relationship in propane dehydrogenation over Ga2O3/ZrO2 catalysts. J. Catal. 2023, 428, 115208. [Google Scholar] [CrossRef]
  35. Shao, H.; Wang, X.; Gu, X.; Wang, D.; Jiang, T.; Guo, X. Improved catalytic performance of CrOx catalysts supported on foamed Sn-modified alumina for propane dehydrogenation. Micropor. Mesopor. Mater. 2021, 311, 110684. [Google Scholar] [CrossRef]
  36. Wang, J.; Song, Y.-H.; Liu, Z.-T.; Liu, Z.-W. Active and selective nature of supported CrOx for the oxidative dehydrogenation of propane with carbon dioxide. Appl. Catal. B Environ. 2021, 297, 120400. [Google Scholar] [CrossRef]
  37. Sun, G.; Huang, Q.; Li, H.; Liu, H.; Zhang, Z.; Wang, X.; Wang, Q.; Wang, J. Different supports-supported Cr-based catalysts for oxidative dehydrogenation of isobutane with CO2. Chin. J. Catal. 2011, 32, 1424–1429. [Google Scholar] [CrossRef]
  38. Scire, S.; Fiorenza, R.; Gulino, A.; Cristaldi, A.; Riccobene, P.M. Selective oxidation of CO in H2-rich stream over ZSM5 zeolites supported Ru catalysts: An investigation on the role of the support and the Ru particle size. Appl. Catal. A Gen. 2016, 520, 82–91. [Google Scholar]
  39. Zhang, S.; Ying, M.; Yu, J.; Zhan, W.; Wang, L.; Guo, Y.; Guo, Y. NixAl1O2-δ mesoporous catalysts for dry reforming of methane: The special role of NiAl2O4 spinel phase and its reaction mechanism. Appl. Catal. B Environ. 2021, 291, 120074. [Google Scholar] [CrossRef]
  40. Karina Mathisen, D.G.N.; Andrew, M. Beale, Manuel Sanchez-Sanchez, Gopinathan Sankar, Wim Bras, and Serge Nikitenko. The selective catalytic reduction of NOx with CH4 on Mn-ZSM5: A comparison with Co-ZSM5 and Cu-ZSM5. Appl. Catal. B Environ. 2007, 111, 3130–3138. [Google Scholar]
  41. Liu, Z.; Zhang, F.; Rui, N.; Li, X.; Lin, L.; Betancourt, L.E.; Su, D.; Xu, W.; Cen, J.; Attenkofer, K.; et al. Highly active ceria-supported Ru catalyst for the dry reforming of methane: In situ identification of Ruδ+–Ce3+ interactions for enhanced conversion. ACS Catal. 2019, 9, 3349–3359. [Google Scholar] [CrossRef]
  42. Shishido, T.; Shimamura, K.; Teramura, K.; Tanaka, T. Role of CO2 in dehydrogenation of propane over Cr-based catalysts. Catal. Today 2012, 185, 151–156. [Google Scholar] [CrossRef]
  43. Sokolov, S.; Stoyanova, M.; Rodemerck, U.; Linke, D.; Kondratenko, E.V. Comparative study of propane dehydrogenation over V-, Cr-, and Pt-based catalysts: Time on-stream behavior and origins of deactivation. J. Catal. 2012, 293, 67–75. [Google Scholar] [CrossRef]
  44. Kresse, G.; Furthmuller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
Figure 1. The characterization results of prepared Cr-based zeolite samples: (a) small-angle XRD, (b) conventional angular XRD, (c) N2 physisorption-desorption isotherms of Cr3%-ZS-6 (mZSM-5/mSBA-15 mass ratio of 6), (d) H2-TPR (temperature-programmed reduction) profiles, (e) Raman spectrum, and (f) XPS spectra. As noted, the ZSM-5 was abbreviated as Z5, and the ZSM-5@SBA-15 was abbreviated as ZS.
Figure 1. The characterization results of prepared Cr-based zeolite samples: (a) small-angle XRD, (b) conventional angular XRD, (c) N2 physisorption-desorption isotherms of Cr3%-ZS-6 (mZSM-5/mSBA-15 mass ratio of 6), (d) H2-TPR (temperature-programmed reduction) profiles, (e) Raman spectrum, and (f) XPS spectra. As noted, the ZSM-5 was abbreviated as Z5, and the ZSM-5@SBA-15 was abbreviated as ZS.
Catalysts 14 00370 g001
Figure 2. HAADF-STEM, HRTEM, and EDS-mapping images of (ac) Cr3%-Z5, (df) Cr3%Zr2%-Z5, and (gi) Cr3%Zr2%-ZS-6.
Figure 2. HAADF-STEM, HRTEM, and EDS-mapping images of (ac) Cr3%-Z5, (df) Cr3%Zr2%-Z5, and (gi) Cr3%Zr2%-ZS-6.
Catalysts 14 00370 g002
Figure 3. In situ FTIR over Cr3%Zr2%-ZS-6. Reaction conditions: the sample was first pretreated at 500 °C under an H2 atmosphere for 1 h. Then, the temperature was lowered to room temperature, followed by the simultaneous introduction of a mixed gas comprising 5% C3H8 and 5% CO2, balanced by He (total flow rate of 20 mL/min) and introduced into the sample under 250, 300, 350, 400, 450, and 500 °C for 30 min.
Figure 3. In situ FTIR over Cr3%Zr2%-ZS-6. Reaction conditions: the sample was first pretreated at 500 °C under an H2 atmosphere for 1 h. Then, the temperature was lowered to room temperature, followed by the simultaneous introduction of a mixed gas comprising 5% C3H8 and 5% CO2, balanced by He (total flow rate of 20 mL/min) and introduced into the sample under 250, 300, 350, 400, 450, and 500 °C for 30 min.
Catalysts 14 00370 g003
Figure 4. Catalytic performance evaluation of the synthesized samples of (a,b) Cr3%-Z5, Cr3%M2%-Z5 (M = La, Zr, Fe) and (c,d) Cr3%-ZS-n (n = 0.5, 2, 6, 16) as well as Cr3%Zr2%-ZS-6 (mass ratio of 6), including propane and CO2 conversion and olefin selectivity and yield. Reduction conditions: T = 600 °C, GHSV = 12,000 h−1, C3H8:CO2:He = 2:2:16. As noted, ZSM-5 was abbreviated as Z5, and ZSM-5@SBA-15 was abbreviated as ZS.
Figure 4. Catalytic performance evaluation of the synthesized samples of (a,b) Cr3%-Z5, Cr3%M2%-Z5 (M = La, Zr, Fe) and (c,d) Cr3%-ZS-n (n = 0.5, 2, 6, 16) as well as Cr3%Zr2%-ZS-6 (mass ratio of 6), including propane and CO2 conversion and olefin selectivity and yield. Reduction conditions: T = 600 °C, GHSV = 12,000 h−1, C3H8:CO2:He = 2:2:16. As noted, ZSM-5 was abbreviated as Z5, and ZSM-5@SBA-15 was abbreviated as ZS.
Catalysts 14 00370 g004
Table 1. Specific surface area and Cr6+ ratio of the prepared Cr-based zeolite catalysts.
Table 1. Specific surface area and Cr6+ ratio of the prepared Cr-based zeolite catalysts.
SampleSBET a, m2/gCr6+/(Cr6+ + Cr3+)
Cr3%-Z5367.746.9%
Cr3%Zr2%-Z5324.271.1%
Cr3%La2%-Z5363.268.2%
Cr3%Fe2%-Z5376.273.9%
Cr3%-ZS-0.5456.2
Cr3%-ZS-2469.8
Cr3%-ZS-6505.7
Cr3%-ZS-16403.0
a SBET: BET specific surface area.
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

Xing, M.; Liu, N.; Dai, C.; Chen, B. CO2 Oxidative Dehydrogenation of Propane to Olefin over Cr-M (M = Zr, La, Fe) Based Zeolite Catalyst. Catalysts 2024, 14, 370. https://0-doi-org.brum.beds.ac.uk/10.3390/catal14060370

AMA Style

Xing M, Liu N, Dai C, Chen B. CO2 Oxidative Dehydrogenation of Propane to Olefin over Cr-M (M = Zr, La, Fe) Based Zeolite Catalyst. Catalysts. 2024; 14(6):370. https://0-doi-org.brum.beds.ac.uk/10.3390/catal14060370

Chicago/Turabian Style

Xing, Mingqiao, Ning Liu, Chengna Dai, and Biaohua Chen. 2024. "CO2 Oxidative Dehydrogenation of Propane to Olefin over Cr-M (M = Zr, La, Fe) Based Zeolite Catalyst" Catalysts 14, no. 6: 370. https://0-doi-org.brum.beds.ac.uk/10.3390/catal14060370

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