Next Article in Journal
Zinc Iodide-Metal Chloride-Organic Base: An Efficient Catalytic System for Synthesis of Cyclic Carbonates from Carbon Dioxide and Epoxides under Ambient Conditions
Next Article in Special Issue
Effect of Different Zinc Species on Mn-Ce/CuX Catalyst for Low-Temperature NH3-SCR Reaction: Comparison of ZnCl2, Zn(NO3)2, ZnSO4 and ZnCO3
Previous Article in Journal
Recent Progress in Pd-Catalyzed Tandem Processes
Previous Article in Special Issue
Research and Application Development of Catalytic Redox Technology for Zeolite-Based Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mechanistic Insight into the Propane Oxidation Dehydrogenation by N2O over Cu-BEA Zeolite with Diverse Active Site Structures

Faculty of Environment and Life, Beijing University of Technology, Beijing 100124, China
*
Author to whom correspondence should be addressed.
Catalysts 2023, 13(8), 1212; https://doi.org/10.3390/catal13081212
Submission received: 19 July 2023 / Revised: 13 August 2023 / Accepted: 14 August 2023 / Published: 15 August 2023

Abstract

:
The present work theoretically investigated propane oxidation dehydrogenation by utilizing N2O as an oxidant (N2O-ODHP) over Cu-BEA with three different types of active site, including monomeric Cu ([Cu]+), dimeric Cu ([Cu−Cu]2+), and distant monomeric Cu sites ([Cu]+—[Cu]+). Energetically, we calculated that the monomeric [Cu]+ is favorable for the αH dehydrogenation step (∆E = 0.05 eV), which, however, suffers from high barriers of N2O dissociation and βH dehydrogenation steps of 1.40 and 1.94 eV, respectively. Although the dimeric [Cu−Cu]2+ site with a Cu—Cu distance of 4.91 Å is much more favorable for N2O dissociation (0.95 eV), it still needs to overcome an extremely high barrier (∆E = 2.15 eV) for βH dehydrogenation. Interestingly, the distant [Cu]+—[Cu]+ site with the Cu—Cu distance of 5.82 Å exhibits low energy barriers for N2O dissociation (0.89 eV) and ODHP steps (0.01 and 0.33 eV) due to the synergistic effect of distant [Cu]+. The microkinetic analyses quantitatively verified the superior activity of the distant [Cu]+—[Cu]+ site with a reaction rate being eight to nine orders of magnitude higher than those of the monomeric and the dimeric Cu sites, and this is related to its ready charge-transfer ability, as shown by the partial Density of State (PDOS) analysis and the static charge differential density analysis in this study. Generally, the present work proposes that the distance between the [Cu]+ sites plays a significant and important role in N2O-ODHP over the Cu-based zeolite catalyst and modulates Cu—Cu distance, and this constitutes a promising strategy for highly-efficient Cu-zeolite catalyst design for N2O-ODHP.

Graphical Abstract

1. Introduction

Propylene is one of the most important organic raw materials, and it can be used to synthesize petrochemical products, such as polyurethane, polypropylene, acetone, acrylonitrile, polyacrylonitrile, and propylene oxide. The traditional commercial production of propylene is mainly through fluid catalytic cracking (FCC) and steam cracking (SC) of petroleum by-products, such as naphtha and light diesel oil [1,2]. However, relatively low selectivity as well as limited resources cannot meet the growing demand of propylene, thereby making it highly desirable to develop some efficient and economical methods to produce propylene [3,4,5]. The massive exploitation and use of shale gas has increased the production of low-carbon alkanes, and propane has thus become a cheap chemical raw material. Converting abundant propane into propylene is not only an important topic in the field of the petrochemical industry but is also a research hotspot in the field of heterogeneous catalysis [4,5], and it is a promising means of meeting the huge demand of the propylene market [2,6].
The catalytic dehydrogenation of propane techniques include the direct dehydrogenation of propane (PDH) and the oxidative dehydrogenation of propane (ODHP). The direct PDH has been industrialized: one example is the Catofin process using chromium aluminum oxide as a catalyst by the Lummus Company [7,8], and another is the Oleflex process from UOP utilizing platinum-based catalysts [7,8]. However, the direct PDH is a reversible and strongly-endothermic reaction that is limited by thermodynamic equilibrium. Moreover, C−C bond breaking to produce methane/ethylene is much more likely to occur at high temperatures, which can further lower propylene selectivity [7,8]. Due to these shortcomings, researchers have conducted studies on the catalytic dehydrogenation of propane under aerobic conditions. Recently, it has been reported that utilizing N2O as the oxidant is much more effective than O2 for the ODHP [9,10,11]. For example, Bulanek et al. [9] have reported that the N2O-ODHP is much more efficient relative to that of O2-ODHP by displaying its higher propane conversion rate and its propylene selectivity. Katerinas et al. [10] have also reported that the selectivity of propylene increased from 33.6% to 68% when using N2O as the oxidant, which can also improve the conversion rate of propane [12].
At the current stage, researchers are still trying to find the suitable catalyst for the N2O-ODHP, and the zeolite catalyst is notable here due to its excellent N2O dissociation activity, which generates αO [13]. The BEA zeolite, with a unique twelve-ring structure, possesses better N2O dissociation activity than those of Y, MFI, FAU, and MOR [14,15]. Sobalikseta al. [16] have reported that Cu (II) in BEA is the active site of N2O decomposition. In our previous works, the catalytic dissociation of N2O [17] as well as the N2O oxidation of methane into methanol [18] were also investigated over the Cu-BEA, and it was found that both the monomeric and the dimeric Cu sites can function as the active sites for N2O dissociation to generate αO. In the present work, the N2O-ODHP was theoretically investigated over the Cu-BEA zeolite with diverse active site structures, including monomeric [Cu]+, dimeric [Cu−Cu]2+, and distant [Cu]+—[Cu]+ sites. The Mars van Krevelen mechanism is related to the reaction between the reactant and the lattice oxygen of the oxidation catalyst [19,20]. The first step is the oxidation of the reductant by the lattice oxygen of the catalyst that generates the product, with a simultaneous formation of oxygen vacancy. The second step is the regeneration of the catalytic active site through the dissociated oxygen in order to replenish oxygen vacancies. Being similar to such a mechanism, in the present work, the αO functions as the active site, and it is utilized to oxidize C3H8 into C3H6 and H2O. The αO would thereby be further regenerated through the reoxidation by N2O. The specific reaction mechanisms were well illustrated by DFT, and the microkinetic modeling was further conducted in order to quantitatively compare the reaction rates of the diverse active sites. Generally, the present work aims to shed a deeper mechanistic light on the active-site motif structural effect on N2O-ODHP and, moreover, emphasize that modulating the Cu—Cu distance would constitute a promising method for a highly efficient Cu-based zeolite catalyst design for the N2O-ODHP.

2. Result and Discussion

2.1. N2O-ODHP Mechanism Simulation over Diverse ACTIVE Site

Three types of Cu-BEA models with different active centers (Figure S1a,c) were constructed for the N2O-ODHP mechanism simulations, which comprised three steps: (i) N2O dissociation to produce αO with the simultaneous release of N2, and the dehydrogenation of (ii) the αH and (ii) the βH of C3H8. The derived energy diagrams along with a different reaction route (Routes A–C) over these active sites are depicted in Figure 1a–d, and based on this, we conducted an in-depth analysis of the specific reaction pathways and the evolution of the structures intermediately generated, and we then compared the derived energy barrier in order to determine the optimal active center for N2O-ODHP.

2.1.1. N2O-ODHP over Monomeric [Cu]+ Site of Route A

(a) Dissociation of N2O to form αO (Reaction Step A1). In this part, the N2O-ODHP over the monomeric [Cu]+ site of Cu-BEA (noted as Za−Cu) was simulated by DFT. Firstly, the N2O molecule can be adsorbed over the Za−Cu site through the O end with a bond length of 2.00 Å. After overcoming a relatively high energy barrier of 1.40 eV (TSA1; Figure 1a), the αO can be generated. As noted, compared with the structure of the N2O-adsorption state, the bond of [CuO]−N was elongated from 1.22 to 1.89 Å, and the bond angle of Cu−O−N2 bent from 121.4 to 124.0° (Figure 2a,b). Such structural distortion can be related to the pre-activation effects of the monomeric [Cu]+. The αO of [Cu−O]+ can be formed, as shown in Figure 2c, with a [CuO]−N bond of 3.76 Å.
(b) Propane dehydrogenation of αH (Reaction Step A2). Reaction Route A2 describes the propane dehydrogenation of αH over Za−Cu−O, wherein there exists one bond fracture of H−C3H7 and two bond formations of [Cu−O]−H and [Cu−O]−C3H7 (the αO being connected with the subtracted H and the radical of C3H7-). As shown in Figure 2d,e, the distance of O−αH slightly shrunk from 2.26 to 1.84 Å, and the bond of C−αH extended from 1.11 to 1.13 Å after the adsorption of C3H8 (∆E = −0.06 eV, Figure 1a). The αH dehydrogenation occurred by crossing the energy barrier of 0.05 eV through the TSA2, which can be characterized by the Cu−O−C bond angle of 100.6° and the Cu−O−H bond angle of 97.5°. After that, the αH can be subtracted from C3H8, forming Za−Cu−OH with the OH bond of 0.97 Å (Figure 2e). In comparison to 1.49 eV of αH dehydrogenation over the CeO2(111) [21], it would be easier for αH dehydrogenation to occur over the monomeric [Cu]+ site of Cu-BEA.
(c) Propane dehydrogenation of βH to form propylene (Reaction Step A3). The dehydrogenation of βH occurred in Route A3 through another transition state of TSA3 with a greatly higher energy barrier of 1.94 eV (Figure 1a), wherein the βH can migrate from C3H7 to Za−Cu−OH, finally generating the C3H6 and the H2O. The distance of the [CuO]−βH shrunk from 2.62 to 1.72 Å, and the bond of C−αH extended from 1.11 to 1.28 Å and formed the Za−Cu−H2O−C3H8 that can be seen in Figure 2f,g. Therefore, according to the above DFT energy calculations, we can derive that the monomeric [Cu]+ site possessing a relatively high barrier for N2O-ODHP, especially for the βH dehydrogenation step, is not potentially active for the N2O-ODHP.

2.1.2. N2O-ODHP over Dimeric [Cu−Cu]2+ Site of Route B

(a) Dissociation of N2O to form αO (Reaction Step B1). In this part, the N2O-ODHP was simulated over the dimeric [Cu−Cu]2+ site (noted as Zb−Cu). As shown in Figure 3a, the N2O could be absorbed over Zb−Cu through both its O and N end, with a Cu−ON2 bond of 1.95 Å, a Cu−N2O bond of 1.80 Å, and an N−N−O bond angle of 175.0°. Much stronger structural distortion can be observed for the adsorbed N2O relative to that of the N2O being adsorbed over the monomeric [Cu]+ site, and this is closely related to the strong synergistic effect of the dimeric [Cu−Cu]2+. Moreover, due to such a synergistic effect, the N2O can be readily (∆E = 0.95 eV, Figure 1b) dissociated to generate the αO ([Cu−O−Cu]2+, Figure 3c) and the N2 through the TS1B. Further comparing the structures of the adsorption state (Figure 3a) and the TS1B (Figure 3b), the bond length of Cu−ON2 shrunk from 1.95 to 1.75 Å, while the N2−O bond enlarged from 1.79 to 1.80 Å, and the bond angle of N−N−O decreased from 112.9 to 118.9°.
(b) Propane dehydrogenation of αH (Reaction step B2). The C3H8 can be initially adsorbed over Zb−Cu−O−Cu (Figure 3d). Subsequently, the αH would migrate from C3H8 to the αO through the TS2B (Figure 3e), which is characterized by the Cua−O bond of 1.80 Å, the Cub−O bond of 1.79 Å, and the Cu−O−Cu bond angle of 143.2°. This crosses a low energy barrier of 0.30 eV (Figure 1b). After that, an intermediate structure of Zb−Cu2−C3H7−OH (Figure 3f) with a Cua−OH bond of 1.92 Å, a Cub−C bond of 2.11 Å, and a C−C−C bond angle of 115.9° can be formed, wherein the radical of the C3H7- can be inserted into the Zb−Cu−O−Cu site.
(c) Propane dehydrogenation of βH to form propylene (Reaction Step B3). After Reaction Step B2, the βH further migrates to the αO of the Zb−Cu2−C3H7−OH site, producing C3H6 and H2O (Figure 3h), which, however, needs to overcome a significantly high barrier of 2.15 eV (Figure 1b) that is comparable to the value of 1.94 eV (Figure 1b) for the scenario of the monomeric [Cu]+ site. The TS3B (Figure 3g) can be characterized by a Cua−O bond of 1.82 Å, an O−βH bond of 1.77Å, a C−βH of 1.15 Å, and a Cu−O−αH bond angle of 103.9°. As noted, the Cub site forms a bridge with H and C in TS3B (Cub−H of 1.71 Å and Cub−C of 2.11 Å). Subsequently, the C−H bond would expand from 1.71 to 1.77 Å, eventually leading to the C−H bond being inviable in Figure 3h. Eventually, the C3H6 can be produced after the C−βH bond breaking and the Cub−C bond formation that is associated with the formation of H2O. As noted, such a high barrier of 2.15 eV also indicates that the dimeric Cu site of Zb−Cu is not suitable for the N2O-ODHP.

2.1.3. N2O-ODHP over Distant [Cu]+—[Cu]+ Site of Route C

(a) Dissociation of N2O to form αO (Reaction step C1). The N2O-ODHP was further simulated over the distant [Cu]+—[Cu]+ site noted as Zc−Cu and with the Cu—Cu distance of 5.82 Å.
As shown in Figure 4a, given that it is similar to that of the dimeric [Cu−Cu]2+ site, the N2O molecule can also be adsorbed over Zc−Cu through both its O and N end, with a bond length of 1.94 and 1.81 Å and a N−N−O bond angle of 173.8º. Relatively stronger structural distortion can be observed for the adsorbed N2O over the Zc−Cu site in comparison to that of the Zb−Cu site, which indicates a stronger synergistic effect of the Zc−Cu site compared to the Zb−Cu for N2O preactivation. A similar finding can also be observed for the N2O dissociation step to generate αO, wherein the Zc−Cu exhibits a relatively lower N2O-dissociation energy barrier (0.89 eV, Figure 1c) than that of the [Cu−Cu]2+ dimeric site (Zb−Cu of 0.95 eV, Figure 1b), which is due to this type of synergistic effect. The TSC1 (Figure 4b) can be characterized by a Cua−O bond of the 1.81 Å, an N−N−O bond angle of 127.6°, and a Zc−Cu-O bond angle of 147.9°. Most importantly, the Zc−Cu site evolved into the motif structure of [Cu−O]+—[Cu]+ (Figure 4c, Cu−O bond of 1.71 Å), which contains two distant monomeric Cu sites that are greatly favorable to the further dehydrogenation of both the αH and the βH of the C3H8 molecule, as will be detailed below.
(b) Propane dehydrogenation of αH (Reaction step C2). The C3H8 can be adsorbed over the [Cu]+ site of [Cu−O]+—[Cu]+ through the C end, forming a C−Cu bond of 2.13 Å and a C3H8 (C−C−C) bond angle of 112.2°. As noted, much stronger structural distortion of C3H8 can also be observed over the Zc−Cu site than that of the Zb−Cu site, which eventually leads to an extremely low barrier of 0.01 eV (Figure 1c) during the αH dehydrogenation through a transition state of TS2C (Figure 4e), and this is characterized by a Cu−C bond of 2.14 Å, a Cu−O bond of 1.72 Å, a [Cu−O]−αH bond of 2.11 Å, and a C-C-C bond angle of 112.6°. Finally, an intermediate structure of Zc−Cu2−C3H7−OH (Figure 4f) can be formed with the Cua−OH bond of 1.78 Å. As noted, it is interesting to see that the generated C3H7- radical was well inserted between the distant [Cu−OH]+—[Cu]+ site, forming, respectively, a Cua−C bond of 1.98 Å and a Cub−C bond of 2.16 Å. This would be greatly favorable for βH dehydrogenation by taking advantage of the synergistic effect of the distant [Cu−OH]+—[Cu]+ site, as stated below in Reaction Step C3.
(c) Propane dehydrogenation of βH to form propylene (Reaction step C3). In this step, the βH would migrate from the C3H7- to the [Cua−OH]+ site, generating the adsorbed H2O and C3H6. Being totally different from the scenarios of both the monomeric and the dimeric Cu active sites (∆E = 1.94 and 2.15 eV, respectively; Figure 1a,b), the βH can be readily dehydrogenated from C3H7- by crossing a significantly lower energy barrier of 0.33 eV (Figure 1c) over the Zc−Cu site due to its strong synergistic effect. The TS3C can be characterized by a Cu-OH bond of 1.78 Å, a βH-O bond of 2.62 Å, and a C−C−C bond angle of 99.9° (Figure 4g). Carefully analyzing the motif structure of Zb−Cu2-C3H7-OH (Figure 3f) and Zc−Cu2-C3H7-OH (Figure 4f), one can find that although the radical of C3H7- can be inserted between both the Zb−Cu and the Zc−Cu site, the specific adsorption mode was different for the two. The C3H7- was adsorbed over the Zb−Cu site through the O−Ca and the Cub−Cc bond, and was adsorbed through the Cua−Ca and the Cub−Cb bond over the Zc−Cu site. In this regard, the C3H7- has to break up the two bonds of O−Ca and βH−Cb in order to generate the C3H6 and H2O during the transition state of the TS3B (as seen in Figure 3g), whereas, on the contrary, the C3H7- only needs to break up one βH−Cb bond during the transition state of the TS3C (seen in Figure 4g).
In the light of the above statements, we can therefore note that the Zc−Cu site possessing the lowest energy barrier for N2O dissociation (0.89 eV) as well as αH (0.01 eV) and βH (0.33 eV) dehydrogenation (Figure 1c,d), relative to those of the Zb−Cu and the Za−Cu sites (especially for the βH dehydrogenation step), is the most active site for N2O-ODHP. Moreover, this finding also shows that modulating the Cu—Cu distance may constitute a promising strategy for highly-efficient zeolite-based N2O-ODHP catalyst design.

2.2. Microkinetic Modeling

Based on the above DFT simulations and transition state theory, microkinetic modeling was further conducted to explore the reaction dynamics (through the intermediate surface coverage variations) and to determine and compare the reaction rate of the rate of the determining step (RDS) over the three different Cu active sites. The N2O-ODHP reaction can be described by five elementary steps, given that it is associated with the kinetic equations listed in Table 1. The calculated kinetic parameters, including reaction rate constant, pre-exponential factor, and specific forward and reverse reaction rate, were listed in Table S1.
(a) Microkinetic modeling over Za−Cu. Figure 5a displays the intermediate coverage variations along with reaction time (t) over the Za−Cu (T = 823 K). Initially, the unoccupied active site coverage (θv) would decrease from 1 to 0.8 ML as it is associated with the increase of adsorbed N2O (θN2O) from 0 to 0.2 ML. This process corresponds with the N2O dissociation step that generates αO, wherein the adsorbed N2O (θN2O) constitutes the major active-site covered species over the Za−Cu due to the relatively high energy barrier of R2 (1.40 eV, Figure 1a, N2O dissociation to generate αO). Along with the reaction, both θv and θN2O would quickly decrease to 0 ML as they are both accompanied with the rapid growth of the coverage of the intermediate of propanol (θC3H7-OH), which finally reaches the equilibrium. This finding indicates that after the formation of αO, it would quickly participate in the ODHP reaction in order to generate the intermediate of propanol, and the intermediate of propanol (θC3H7-OH) constitutes the major active-site covered species, which indicates that the R5 (∆E = 1.94 eV, Figure 1a) constitutes the RDS during N2O-ODHP over the Za−Cu site, leading to the accumulation of C3H7−OH over the active site.
(b) Microkinetic modeling over Zb−Cu. Figure 5b displays the variations of the intermediate coverages during the N2O-ODHP over the Zb−Cu site. Being similar to that of Za−Cu site, the adsorbed N2O would initially cover the active site by displaying the θN2O of 0.3 ML. However, the αO, given that it is in the form of [Cu−O−Cu]2+, would shortly occupy the active site due to the relatively lower N2O dissociation barrier (0.95 versus 1.40 eV of R2) and higher αH dehydrogenation barrier (0.3 versus 0.05 eV of R4) than that of Za−Cu site, which leads to the short accumulation of the αO over the active site. Along with the further reaction (reaching equilibrium), the active site would be eventually covered by the propanol (θC3H7-OH), which is similar to the scenario of the Za−Cu site due to the extremely high barrier of R5 (2.15 eV). Thus, similar to that of Za−Cu, the R5 would constitute the RDS during the N2O-ODHP over the Zb−Cu site.
(c) Microkinetic modeling over Zc−Cu. The surface coverage of the reactant as well as the generated intermediates during N2O-ODHP over the Zc−Cu site were both depicted in Figure 5c. Being totally different from the scenarios of the Za−Cu and the Zb−Cu sites, the N2O constitutes the major active-site covered species over the Zc−Cu site at T = 823 K, wherein the θN2O initially increases up to 1 ML and then decreases to a stable value of above 0.9 ML due to another intermediate αO (θo = 0.1 ML) after the reaction, which reaches the equilibrium. This finding indicates that the N2O dissociation step (R2) would constitute the RDS during the N2O-ODHP over the Zc−Cu site. This finding correlates well with the highest energy barrier of 0.89 eV of R2 during N2O-ODHP (see Figure 1c).
(d) Reaction rate comparisons. Figure 5d displays the forward reaction rate comparisons of each elementary step during N2O-ODHP over the different active sites of Za−Cu, Zb−Cu, and Zc−Cu. Obviously, the Zc−Cu site displays much higher reaction rates than those of the Za−Cu and the Zb−Cu. Moreover, in the net reaction rate (NRR) comparisons, which were further depicted in Figure S1, the Zc−Cu displays an NRR of 1.25+E8 mol∙m−3∙s−1, and it is five and six orders of magnitude higher, respectively, than those of the Za−Cu (120.57 mol∙m−3∙s−1) and the Zb−Cu (49.21 mol∙m−3∙s−1) sites. These findings quantitatively verify the superior activity of the Zc−Cu site relative to those of the Za−Cu and the Zb−Cu sites.

2.3. Static Charge Difference PDOS and Analyses

To further illustrate the superior activity of the Zc−Cu, the static charge difference and the partial density of state (PDOS) analyses were further conducted based on TS3 (corresponding to the βH dehydrogenation step). As shown in Figure 6a–d, large amounts of charge transfers occurred during the βH dehydrogenation, and the Cu of Zc−Cu provided more charges relative to those of the Za−Cu and the Zb−Cu (Figure 6d), which can be closely related to the smallest band gap between the Cu and C of C3H7-, as shown by the PDOS analyses of Figure 6e–g (4.97 versus 5.82 and 5.29 eV). This finding indicates that the Zc−Cu would exhibit a stronger electric field effect on C3H7-, and that it is thereby greatly favorable for βH dehydrogenation.

3. N2O-ODHP Activity Measurement

As is well known, in addition to the [Cu]+ cations, the CuOx species can also exist over Cu-modified zeolite catalyst (Cu-Zeolite) and shed more light on the activity behaviors of these different Cu species. The 1%Cu-BEA and 1%CuO-SiO2 were prepared by the impregnation method (the metal loading of 1wt.%), and they were further evaluated for the N2O-ODHP. In addition, the Fe-modified zeolites (Fe-Zeolite) have also been reported to possess excellent N2O dissociation activity in order to produce αO [17], and to make a comparison with the Cu modified zeolite, the 1%Fe-BEA, and 1%Fe-ZSM-5 were also prepared by the impregnation method (metal loading of 1 wt.%) and evaluated by N2O-ODHP. The specific preparation method is stated in detail in the Supporting Information section of this paper. The activity measurement results, including the C3H8 and N2O conversions as well as the product selectivity, were profiled in Figure 7a–c. As can be seen there, the 1%Cu-BEA displays a higher C3H8 conversion (31.5%) and a higher C3H6 selectivity (74.5%) than the other catalyst samples, especially in comparison with that of 1%CuO-SiO2, displaying a C3H8 conversion of 3.5% and a C3H6 selectivity of 43.5%. This finding indicates that the CuO species would not constitute the major active species for the N2O-ODHP.
As further shown by the N2O conversion of Figure 7b, the 1%CuO-SiO2 displays a much lower N2O conversion (27.2%) than the 1%Cu-BEA (69.4%). This finding shows that the lower N2O dissociation activity of CuO species probably constitutes one of the major reactions that leads to the extremely low N2O-ODHP activity of the 1%CuO-SiO2. Conversely, the 1%Cu-BEA possessing active Cu cations for N2O dissociation that generate αO exhibits a much higher level of N2O-ODHP activity relative to the 1%CuO-SiO2. As has also been reported on a theoretical level [22], the CuO exhibits a much high energy barrier (2.71 eV) for N2O dissociation, which indicates that it is very difficult to decompose N2O and produce α-O over CuO.
As for the samples of 1%Fe-BEA and 1%Fe-ZSM-5, the nearly complete N2O conversion (~100%) can be achieved due to the superior N2O dissociation activity of Fe cations than those of the Cu cations [23], although this does lead to the ready overoxidation of C3H8 into COx (CO and CO2 of 64.1 and 46.0%, respectively; see Figure 7c). This finding shows that the Fe cations are active for N2O dissociation, although they suffer from the overoxidation of C3H8. The Cu-based zeolites would therefore probably be much more suitable for the N2O-ODHP relative to that of the Fe-based zeolite catalyst. However, we would also like to emphasize that such works still need further investigation. We would also like to note that the TPR and UV-vis correlate with the active-site-structure in the theoretical calculation. This will be further studied in our research in the future, which will focus on the influence of a diversely structured topologized zeolite (MFI, FER, MOR, and BEA) on a specific structure, as well as on the location of [Cu]+ cations and how they are related to N2O-ODHP catalytic behaviors, both experimentally and theoretically.

4. Computational Modeling and Methodology

Density functional theory (DFT) adopts the Vienna ab-initio simulation package (VASP). The Projection Enhanced Wave (PAW) method utilizes the interaction between electrons and the core, and it uses Generalized Gradient Approximation (GGA) and Perdew Burke Ernzerhof (PBE) functions to achieve electron exchange correlation [24]. In the process of geometric optimization, the convergence achieved at the energy difference is 10−5 eV, with an ion relaxation convergence standard of 0.05 eV/Å. The energy cutoff of the set plane wave is 400 eV. The K point of the Brillouin region is set to 2 × 2 × 1 in the calculation of the structure, and it is set to 4 × 4 × 2 in the partial wave density of states. In the International Zeolite Association (IZA) database, the cellular model data for BEA is A = 12, B = 12.632, and C = 9.421 Å [25]. The calculation of the transition state (TS) uses the climbing image light pushing elastic bond (CI-NEB) and the dimer method. Four points are inserted between the initial state and the final state, and the saddle points corresponding to the transition state are located in order to find the lowest energy path. The TS state is only identified on one imaginary frequency [25,26,27,28,29]. The method of micro-dynamic modeling is included in the Supporting Information section of this paper.

5. Conclusions

The present work theoretically investigates N2O-ODHP over Cu-BEA with three types of Cu sites––monomeric [Cu]+ (Za−Cu), dimeric [Cu−Cu]2+ (Zb−Cu), and distant [Cu]+—[Cu]+ (Zc−Cu). The Za−Cu is beneficial for αH dehydrogenation (0.05 eV), but it requires a high energy barrier in N2O dissociation and βH dehydrogenation (1.40 and 1.94 eV) to be overcome. The Zb−Cu, with a Cu—Cu distance of 4.91 Å, is suitable for the N2O dissociation step (0.95 eV), but it is highly resistant to the βH hydrogenation step because it displays an extremely high barrier of 2.15 eV. Being contrary to the scenarios of the Za−Cu and the Zb−Cu, the Zc−Cu site with the Cu—Cu distance of 5.82 Å is not only favorable for N2O dissociation (0.89 eV) but also greatly active for the ODHP steps of αH (0.01 eV) and βH (0.33 eV) dehydrogenation. The microkinetic modeling further showed that the Zc−Cu exhibits a five to six orders of magnitude higher net reaction rate than those of the Za−Cu and the Zb−Cu sites. This is closely correlated with the specific structure of the Zc−Cu, which possesses a much stronger electric field effect on the C3H8 molecule due to the synergic effect of Cu, as this possesses the smallest band gaps and they are favorable to the charge transfers between the Cu active site and the C3H8 molecule. Generally, we ultimately propose that modulating the Cu active site distance (Cu—Cu) probably constitutes a promising strategy for highly-efficient Cu-zeolite design for the N2O-ODHP. Additionally, we would like to mention that it is not possible for there to be only one type of Z−Cu site (Cu cation site)] experimentally, and other types of Cu site (such as [Cu3O3]2+, which is proposed by Lercher et al. [30] for methane direct oxidation to methanol) may also be active for the N2O-ODHP, which would be a good direction for further study of this subject.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal13081212/s1, Figure S1. Optimized Cu-BEA models with diverse Cu active site motifs (a) Za-Cu; (b) Zb-Cu; (c) Zc-Cu site; Si (yellow), O (red), N (blue), Al (pink), Cu (orange); Scheme S1. Schematic reaction for N2O-ODHP over Za-Cu (reaction Route A), Zb-Cu (reaction Route B) and Zc-Cu (reaction Route C), respectively; Table S1. Micro-dynamics parameters of the reaction steps over Za-Cu, Zb-Cu and Zc-Cu site; Figure S2. The comparison of net reaction rate over Schematic reaction for N2O-ODHP over Za-Cu, Zb-Cu and Zc-Cu site at 823 K [17,18,25,31,32,33,34,35,36].

Author Contributions

Methodology, N.W.; validation, C.D.; investigation, R.W.; resources, B.C.; data curation, R.X.; writing—original draft preparation, R.W.; writing—review and editing, N.L.; visualization, G.Y.; supervision, B.C. 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 and 22176006).

Data Availability Statement

Data is available upon request to the corresponding authors.

Acknowledgments

We acknowledge the final support from National Natural Science Foundation of China (No. 22178011 and 22176006).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pérez-Ramírez, J.; Gallardo-Llamas, A. Framework composition effects on the performance of steam-activated FeMFI zeolites in the N2O-mediated propane oxidative dehydrogenation to propylene. J. Phys. Chem. B 2005, 109, 20529–20538. [Google Scholar] [CrossRef] [PubMed]
  2. Kondratenko, E.; Cherian, M.; Baerns, M.; Su, D.; Schlogl, R.; Wang, X.; Wachs, I. Oxidative dehydrogenation of propane over V/MCM-41 catalysts: Comparison of O2 and N2O as oxidants. J. Catal. 2005, 234, 131–142. [Google Scholar] [CrossRef] [Green Version]
  3. Sanchezgalofre, O.; Segura, Y.; Perezramirez, J. Deactivation and regeneration of iron-containing MFI zeolites in propane oxidative dehydrogenation by N2O. J. Catal. 2007, 249, 123–133. [Google Scholar] [CrossRef]
  4. Wei, W.; Moulijn, J.A.; Mul, G. FAPO and Fe-TUD-1: Promising catalysts for N2O mediated selective oxidation of propane. J. Catal. 2009, 262, 1–8. [Google Scholar] [CrossRef]
  5. Kondratenko, E.V.; Brückner, A. On the nature and reactivity of active oxygen species formed from O2 and N2O on VOx/MCM-41 used for oxidative dehydrogenation of propane. J. Catal. 2010, 274, 111–116. [Google Scholar] [CrossRef]
  6. Orlyk, S.; Kyriienko, P.; Kapran, A.; Chedryk, V.; Balakin, D.; Gurgul, J.; Zimowska, M.; Millot, Y.; Dzwigaj, S. CO2-assisted dehydrogenation of propane to propene over Zn-BEA zeolites: Impact of acid–base characteristics on catalytic performance. Catalysts 2023, 13, 681. [Google Scholar] [CrossRef]
  7. Vogt, E.T.; Weckhuysen, B.M. Fluid catalytic cracking: Recent developments on the grand old lady of zeolite catalysis. Chem. Soc. Rev. 2015, 44, 7342–7370. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Chen, J.Q.; Bozzano, A.; Glover, B.; Fuglerud, T.; Kvisle, S. Recent advancements in ethylene and propylene production using the UOP/Hydro MTO process. Catal. Today 2005, 106, 103–107. [Google Scholar] [CrossRef]
  9. Bulánek, R.; Wichterlová, B.; Novoveská, K.; Kreibich, V. Oxidation of propane with oxygen and/or nitrous oxide over Fe-ZSM-5 with low iron concentrations. Appl. Catal. A 2004, 264, 13–22. [Google Scholar] [CrossRef]
  10. Novoveská, K.; Bulánek, R.; Wichterlová, B. Oxidation of propane with oxygen, nitrous oxide and oxygen/nitrous oxide mixture over Co- and Fe-zeolites. Catal. Today 2005, 100, 315–319. [Google Scholar] [CrossRef]
  11. Wu, G.; Hao, Y.; Zhang, N.; Guan, N.; Li, L.; Grünert, W. Oxidative dehydrogenation of propane with nitrous oxide over Fe–O–Al species occluded in ZSM-5: Reaction and deactivation mechanisms. Microporous Mesoporous Mater. 2014, 198, 82–91. [Google Scholar] [CrossRef]
  12. Jiang, X.; Sharma, L.; Fung, V.; Park, S.J.; Jones, C.W.; Sumpter, B.G.; Baltrusaitis, J.; Wu, Z. Oxidative dehydrogenation of propane to propylene with soft oxidants via heterogeneous catalysis. ACS Catal. 2021, 11, 2182–2234. [Google Scholar] [CrossRef]
  13. Patet, R.E.; Koehle, M.; Lobo, R.F.; Caratzoulas, S.; Vlachos, D.G. General acid-type catalysis in the dehydrative aromatization of furans to aromatics in H-[Al]-BEA, H-[Fe]-BEA, H-[Ga]-BEA, and H-[B]-BEA zeolites. J. Phys. Chem. C 2017, 121, 13666–13679. [Google Scholar] [CrossRef]
  14. Chalupka, K.; Thomas, C.; Millot, Y.; Averseng, F.; Dzwigaj, S. Mononuclear pseudo-tetrahedral V species of VSiBEA zeolite as the active sites of the selective oxidative dehydrogenation of propane. J. Catal. 2013, 305, 46–55. [Google Scholar] [CrossRef]
  15. Mauvezin, M.; Delahay, G.; Kißlich, F.; Coq, B.; Kieger, S. Catalytic reduction of N2O by NH3 in presence of oxygen using Fe-exchanged zeolites. Catal. Lett. 1999, 62, 41–44. [Google Scholar] [CrossRef]
  16. Sobalik, Z.; Sazama, P.; Dedecek, J.; Wichterlová, B. Critical evaluation of the role of the distribution of Al atoms in the framework for the activity of metallo-zeolites in redox N2O/NOx reactions. Appl. Catal. A 2014, 474, 178–185. [Google Scholar] [CrossRef]
  17. Liu, N.; Zhang, R.; Chen, B.; Li, Y.; Li, Y. Comparative study on the direct decomposition of nitrous oxide over M (Fe, Co, Cu)–BEA zeolites. J. Catal. 2012, 294, 99–112. [Google Scholar] [CrossRef]
  18. Xu, R.; Liu, N.; Dai, C.; Li, Y.; Zhang, J.; Wu, B.; Yu, G.; Chen, B. H2O-built proton transfer bridge enhances continuous methane oxidation to methanol over Cu-BEA zeolite. Angew. Chem. 2021, 60, 16634–16640. [Google Scholar] [CrossRef]
  19. Widmann, D.; Behm, R.J. Dynamic surface composition in a Mars-van Krevelen type reaction: CO oxidation on Au/TiO2. J. Catal. 2018, 357, 263–273. [Google Scholar] [CrossRef]
  20. Wang, C.; Gu, X.-K.; Yan, H.; Lin, Y.; Li, J.; Liu, D.; Li, W.-X.; Lu, J. Water-mediated Mars–Van Krevelen mechanism for CO oxidation on ceria-supported single-atom Pt1 catalyst. ACS Catal. 2016, 7, 887–891. [Google Scholar] [CrossRef]
  21. Jan, F.; Lian, Z.; Zhi, S.; Yang, M.; Si, C.; Li, B. Revealing the role of HBr in propane dehydrogenation on CeO2(111) via DFT-based microkinetic simulation. Phys. Chem. Chem. Phys. 2022, 24, 9718–9726. [Google Scholar] [CrossRef] [PubMed]
  22. Suo, W.; Sun, S.; Liu, N.; Li, X.; Wang, Y. The adsorption and dissociation of N2O on CuO(111) surface: The effect of surface structures. Surf. Sci. 2020, 696, 121596. [Google Scholar] [CrossRef]
  23. Li, Y.; Liu, N.; Dai, C.; Xu, R.; Yu, G.; Wang, N.; Zhang, J.; Chen, B. Synergistic Effect of neighboring Fe and Cu cation sites boosts FenCum-BEA activity for the continuous direct oxidation of Methane to Methanol. Catalysts 2021, 11, 1444. [Google Scholar] [CrossRef]
  24. Jianwen, Z.; Zhifeng, H.; Ziqian, Y.; Meijuan, L.; Fei, C.; Qiang, S. The influence of alkaline earth elements on electronic properties of α-Si3N4 via DFT calculation. J. Wuhan Univ. Technol. Mater. Sci. Ed. 2020, 35, 863–871. [Google Scholar]
  25. Zhou, Z.; Qin, B.; Li, S.; Sun, Y. A DFT-based microkinetic study on methanol synthesis from CO2 hydrogenation over the In2O3 catalyst. Phys. Chem. Chem. Phys. 2021, 23, 1888–1895. [Google Scholar] [CrossRef] [PubMed]
  26. Montejo-Valencia, B.D.; Curet-Arana, M.C. Periodic DFT study of the opening of fructose and glucose rings and the further conversion of fructose to trioses catalyzed by M-BEA (M = Sn, Ti, Zr, or Hf). J. Phys. Chem. C 2019, 123, 3532–3540. [Google Scholar] [CrossRef]
  27. Mahyuddin, M.H.; Staykov, A.; Shiota, Y.; Yoshizawa, K. Direct conversion of methane to methanol by metal-exchanged ZSM5 zeolite (Metal = Fe, Co, Ni, Cu). ACS Catal. 2016, 6, 8321–8331. [Google Scholar] [CrossRef]
  28. Liu, N.; Zhang, R.; Li, Y.; Chen, B. Local electric field effect of TMI (Fe, Co, Cu)-BEA on N2O direct dissociation. J. Phys. Chem. C 2014, 118, 10944–10956. [Google Scholar] [CrossRef]
  29. Konsolakis, M. Recent advances on nitrous oxide (N2O) decomposition over non-noble-metal oxide catalysts: Catalytic performance, mechanistic considerations, and surface chemistry aspects. ACS Catal. 2015, 5, 6397–6421. [Google Scholar] [CrossRef]
  30. Ikuno, T.; Grundner, S.; Jentys, A.; Li, G.; Pidko, E.A.; Fulton, J.L.; Sanchez-Sanchez, M.; Lercher, J.A. Formation of Active Cu-oxo Clusters for Methane Oxidation in Cu-Exchanged Mordenite. J. Phys. Chem. C 2019, 123, 8759–8769. [Google Scholar] [CrossRef]
  31. Liu, N.; Yuan, X.; Zhang, R.; Xu, R.; Li, Y. Mechanistic insight into selective catalytic combustion of acrylonitrile (C2H3CN): NCO formation and its further transformation towards N2. Phys. Chem. Chem. Phys. 2017, 19, 7971–7979. [Google Scholar] [CrossRef] [PubMed]
  32. Liu, N.; Yuan, X.; Zhang, R.; Li, Y.; Chen, B. Mechanistic insight into selective catalytic combustion of HCN over Cu-BEA: Influence of different active center structures. Phys. Chem. Chem. Phys. 2017, 19, 23960–23970. [Google Scholar] [CrossRef] [PubMed]
  33. Dixit, M.; Baruah, R.; Parikh, D.; Sharma, S.; Bhargav, A. Autothermal reforming of methane on rhodium catalysts: Microkinetic analysis for model reduction. Comput. Chem. Eng. 2016, 89, 149–157. [Google Scholar] [CrossRef] [Green Version]
  34. Ke, C.; Lin, Z. Density Functional Theory Based Micro- and Macro-Kinetic Studies of Ni-Catalyzed Methanol Steam Reforming. Catalysts 2020, 10, 349. [Google Scholar] [CrossRef] [Green Version]
  35. Gokhale, A.A.; Kandoi, S.; Greeley, J.P.; Mavrikakis, M.; Dumesic, J.A. Molecular-level descriptions of surface chemistry in kinetic models using density functional theory. Chem. Eng. Sci. 2004, 59, 4679–4691. [Google Scholar] [CrossRef]
  36. Cai, Q.-X.; Wang, J.-G.; Wang, Y.-G.; Mei, D. Mechanistic insights into the structure-dependent selectivity of catalytic furfural conversion on platinum catalysts. AIChE J. 2015, 61, 3812–3824. [Google Scholar] [CrossRef]
Figure 1. Energy diagram of N2O-ODHP over Cu-BEA with different active sites of (a) Za−Cu (reaction Route A); (b) Zb−Cu (reaction Route B; (c) Zc−Cu (reaction Route C); and (d) energy barrier comparisons. Si (yellow), O (red), N (blue), Al (pink), Cu (orange).
Figure 1. Energy diagram of N2O-ODHP over Cu-BEA with different active sites of (a) Za−Cu (reaction Route A); (b) Zb−Cu (reaction Route B; (c) Zc−Cu (reaction Route C); and (d) energy barrier comparisons. Si (yellow), O (red), N (blue), Al (pink), Cu (orange).
Catalysts 13 01212 g001
Figure 2. Optimized model of Cu-BEA with Za−Cu site for the N2O-ODHP: (a) Za−Cu+N2O (adsorption fo N2O); (b) TS1A; (c) Za−Cu−O+N2; (d) Za−Cu−O+C3H8 (adsorption of C3H8); (e) TS2A; (f) Za−Cu−OH−C3H7; (g) TS3A; (h) Za−Cu−H2O−C3H6. Si (yellow), O (red), N (blue), Al (pink), Cu (orange).
Figure 2. Optimized model of Cu-BEA with Za−Cu site for the N2O-ODHP: (a) Za−Cu+N2O (adsorption fo N2O); (b) TS1A; (c) Za−Cu−O+N2; (d) Za−Cu−O+C3H8 (adsorption of C3H8); (e) TS2A; (f) Za−Cu−OH−C3H7; (g) TS3A; (h) Za−Cu−H2O−C3H6. Si (yellow), O (red), N (blue), Al (pink), Cu (orange).
Catalysts 13 01212 g002
Figure 3. Optimized model of Cu-BEA with Zb−Cu site for the N2O-ODHP: (a) Zb−Cu2+N2O (adsorption fo N2O); (b) TS1B; (c) Zb−Cu−O−Cu+N2; (d) Zb−Cu−O−Cu+C3H8 (adsorption of C3H8); (e) TS2B; (f) Zb−Cu2−OH−C3H7; (g) TS3B; (h) Zb−Cu2−C3H6−H2O. Si (yellow), O (red), N (blue), Al (pink), Cu (orange).
Figure 3. Optimized model of Cu-BEA with Zb−Cu site for the N2O-ODHP: (a) Zb−Cu2+N2O (adsorption fo N2O); (b) TS1B; (c) Zb−Cu−O−Cu+N2; (d) Zb−Cu−O−Cu+C3H8 (adsorption of C3H8); (e) TS2B; (f) Zb−Cu2−OH−C3H7; (g) TS3B; (h) Zb−Cu2−C3H6−H2O. Si (yellow), O (red), N (blue), Al (pink), Cu (orange).
Catalysts 13 01212 g003
Figure 4. Optimized model of Cu-BEA with Zc−Cu site for the N2O-ODHP: (a) Zc−Cu2+N2O (adsorption fo N2O); (b) TS1C; (c) Zc−Cu−O−Cu+N2; (d) Zc−Cu2−O+C3H8 (adsorption of C3H8); (e) TS2C; (f) Zc−Cu2−C3H7−OH; (g) TS3C; (h) Zc−Cu2−C3H6−H2O. Si (yellow), O (red), N (blue), Al (pink), Cu (orange).
Figure 4. Optimized model of Cu-BEA with Zc−Cu site for the N2O-ODHP: (a) Zc−Cu2+N2O (adsorption fo N2O); (b) TS1C; (c) Zc−Cu−O−Cu+N2; (d) Zc−Cu2−O+C3H8 (adsorption of C3H8); (e) TS2C; (f) Zc−Cu2−C3H7−OH; (g) TS3C; (h) Zc−Cu2−C3H6−H2O. Si (yellow), O (red), N (blue), Al (pink), Cu (orange).
Catalysts 13 01212 g004
Figure 5. Microkinetic modeling results: the surface coverage variations along with reaction time (t) over the (a) Za−Cu, (b) Zb−Cu, and (c) Zc−Cu sites of Cu-BEA at 823 K, and (d) the forward reaction rate comparisons.
Figure 5. Microkinetic modeling results: the surface coverage variations along with reaction time (t) over the (a) Za−Cu, (b) Zb−Cu, and (c) Zc−Cu sites of Cu-BEA at 823 K, and (d) the forward reaction rate comparisons.
Catalysts 13 01212 g005
Figure 6. The static charge difference (ac) and partial density of state (PDOS) analyses (eg) of TS3 models for Za−, Zb−, and Zc−Cu-BEA; (d) bader charges analysis results; (e) partial density of state (PDOS) of atomic Cu and C over (e) Za−Cu; (f) Zb−Cu; and (g) Zc−Cu site. Yellow and blue colors represent the increase and decrease in electron density, respectively. Si (yellow), O (red), N (blue), Al (pink), Cu (orange).
Figure 6. The static charge difference (ac) and partial density of state (PDOS) analyses (eg) of TS3 models for Za−, Zb−, and Zc−Cu-BEA; (d) bader charges analysis results; (e) partial density of state (PDOS) of atomic Cu and C over (e) Za−Cu; (f) Zb−Cu; and (g) Zc−Cu site. Yellow and blue colors represent the increase and decrease in electron density, respectively. Si (yellow), O (red), N (blue), Al (pink), Cu (orange).
Catalysts 13 01212 g006
Figure 7. Activity measurement of N2O-ODHP over 1%Cu-BEA (orange), 1%Fe-BEA (red), 1%Fe-ZSM-5 (dark blue), and 1%CuO-SiO2 (green): (a) C3H8 conversion; (b) N2O conversion; (c) product selectivity of C2H6 (light green), CH4 (light purple), CO (light blue), CO2 (blue), C2H4 (light pink), and C3H6 (light red). GHSV = 12,000 h−1, C3H8:N2O:He = 10:10:80, T = 550 °C.
Figure 7. Activity measurement of N2O-ODHP over 1%Cu-BEA (orange), 1%Fe-BEA (red), 1%Fe-ZSM-5 (dark blue), and 1%CuO-SiO2 (green): (a) C3H8 conversion; (b) N2O conversion; (c) product selectivity of C2H6 (light green), CH4 (light purple), CO (light blue), CO2 (blue), C2H4 (light pink), and C3H6 (light red). GHSV = 12,000 h−1, C3H8:N2O:He = 10:10:80, T = 550 °C.
Catalysts 13 01212 g007
Table 1. Elementary steps of micro-dynamics and the equations of reaction rate for the N2O-ODHP over Cu-BEA.
Table 1. Elementary steps of micro-dynamics and the equations of reaction rate for the N2O-ODHP over Cu-BEA.
StepElementary StepsReaction Rate Equations
R1Z−Cu−N2O(g) Z−Cu−N2Or1 = k1PN2Oθvk-1θN2O
R2Z−Cu−N2O → Z−Cu−O+N2(g)r2 = k2θN2O
R3Z−Cu−O+C3H8(g) Z−Cu−O−C3H8r3 = k3PC3H8θOk-3θO-C3H8
R4Z−Cu−O−C3H8 Z−Cu−OH−C3H7r4 = k4θO-C3H8k-4θC3H7-OH
R5Z−Cu−OH−C3H7 Z−Cu−H2O−C3H6r5 = k5θC3H7-OHk-5θC3H6-H2O
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

Wu, R.; Liu, N.; Dai, C.; Xu, R.; Yu, G.; Wang, N.; Chen, B. Mechanistic Insight into the Propane Oxidation Dehydrogenation by N2O over Cu-BEA Zeolite with Diverse Active Site Structures. Catalysts 2023, 13, 1212. https://doi.org/10.3390/catal13081212

AMA Style

Wu R, Liu N, Dai C, Xu R, Yu G, Wang N, Chen B. Mechanistic Insight into the Propane Oxidation Dehydrogenation by N2O over Cu-BEA Zeolite with Diverse Active Site Structures. Catalysts. 2023; 13(8):1212. https://doi.org/10.3390/catal13081212

Chicago/Turabian Style

Wu, Ruiqi, Ning Liu, Chengna Dai, Ruinian Xu, Gangqiang Yu, Ning Wang, and Biaohua Chen. 2023. "Mechanistic Insight into the Propane Oxidation Dehydrogenation by N2O over Cu-BEA Zeolite with Diverse Active Site Structures" Catalysts 13, no. 8: 1212. https://doi.org/10.3390/catal13081212

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