Next Article in Journal
Single-Layer MoS2-MoO3-x Heterojunction Nanosheets with Simultaneous Photoluminescence and Co-Photocatalytic Features
Previous Article in Journal
Microcystis@TiO2 Nanoparticles for Photocatalytic Reduction Reactions: Nitrogen Fixation and Hydrogen Evolution
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synergistic Effect of Neighboring Fe and Cu Cation Sites Boosts FenCum-BEA Activity for the Continuous Direct Oxidation of Methane to Methanol

1
State Key Laboratory of Chemical Resource Engineering, Beijing University of Chemical Technology, Beijing 100029, China
2
Faculty of Environment and Life, Beijing University of Technology, Beijing 100124, China
3
Jiangsu Key Laboratory of Advanced Catalytic Materials and Technology, School of Petrochemical Engineering, Changzhou University, Changzhou 213164, China
*
Author to whom correspondence should be addressed.
Submission received: 5 November 2021 / Revised: 22 November 2021 / Accepted: 25 November 2021 / Published: 27 November 2021
(This article belongs to the Topic Catalysis for Sustainable Chemistry and Energy)

Abstract

:
Direct oxidation of methane to methanol (DMTM), constituting a major challenge for C1 chemistry, has aroused significant interest. The present work reports the synergistic effect of neighboring [Fe]--[Cu] cations, which can significantly boost the CH3OH productivity (100.9 and 41.9 → 259.1 μmol∙g−1cat∙h−1) and selectivity (0.28 and 17.6% → 71.7%) of the best performing Fe0.6%Cu0.68%-BEA (relative to monomeric Fe1.28%- and Cu1.28%-BEA) during the continuous H2O-mediated N2O DMTM. The combined experimental (in situ FTIR, D2O isotopic tracer technique) and theoretical (DFT, ab initio molecular dynamics (AIMD)) studies reveal deeper mechanistic insights that the synergistic effect of [Fe]--[Cu] can not only significantly favor active O production (ΔG = 0.18 eV), but also efficiently motivate the reaction following a H2O proton-transfer route (ΔG = 0.07 eV), eventually strikingly promoting CH3OH productivity/selectivity. Generally, the proposed strategy by employing the synergistic effect of bimetallic cations to modify DMTM activity would substantially favor other highly efficient catalyst designs.

Graphical Abstract

1. Introduction

Methane, being a primary component of natural gas, constitutes the main feedstock during methanol production, which traditionally follows the syngas route, needing high temperature and pressure [1]. By contrast, the direct oxidation of methane to methanol (DMTM) operated under mild conditions is greatly appealing for CH3OH production; however, this process suffers from low selectivity and productivity, due to the extremely high C-H bond energy (413 kJ∙mol−1) of CH4 but relatively low thermostability of CH3OH [2,3,4]. Although it is still a major challenge to make breakthroughs for the DMTM, the development of suitable catalysts and reaction systems to promote CH3OH selectivity and productivity has never ceased.
Methane monooxygenases (MMOs), as a type of methanotrophic bacteria, can readily transform CH4 into CH3OH at room temperature, the active site of which consists of iron or copper [5]. To mimic its catalytic behavior, the Fe- and Cu-exchanged zeolites have been extensively investigated for the DMTM [6,7,8,9,10,11,12,13,14,15,16], wherein the specific active site motifs significantly affect the DMTM activity. For example, Jeong et al. [7] recently reported a continuous DMTM by passing O2 over Cu-MOR, finding that the binuclear [Cu-O-Cu]2+ site possessed much higher activity than that of the mononuclear [Cu]+ site. To further improve the DMTM activity, the strategy of introducing another transition metal into the Cu- or Fe-exchanged zeolite to modify the active motif has also been proposed [8,9,10,11,12,13,14]. The group of Hutchings [8,9,10] performed pioneering works finding that the Cu species can dramatically promote the CH3OH selectivity (>92%) and productivity of the Fe/Cu-ZSM-5 zeolite catalyst through efficiently inhibiting the overoxidation of CH3OH (into formic acid and CO2) during the H2O2 DMTM. Recently, Yu et al. [11] further found that the Cu species of Fe/Cu-ZSM can also facilitate the generation of OH- radicals which thereby can quickly react with -CH3 to produce CH3OH. Meanwhile, some theoretical simulations have also been conducted for the DMTM over a series of zeolitic bimetal active centers of [Cu-O-M]2+ and [Cu-O2-M]2+ (M = Pd, Pt, Fe, Co, Ni, Zn, Au and Ag) [12,13,14], aiming at favoring the rational design of highly efficient bimetallic ion-exchanged zeolite catalysts.
In addition to the active site motifs, the utilized oxidant as well as H2O also significantly affect the zeolitic DMTM activity and efficiency. In comparison to O2 and H2O2, N2O is a promising oxidant for the DMTM due to the much more ready formation of highly active oxygen species (named αO, T < 250 °C) after the in situ interaction with the metal modified zeolite (M-Z) catalyst [17]. Additionally, the in situ generation of active oxygen (αO) at low temperatures demonstrates N2O as being a promising oxidant for the continuous DMTM, which can strikingly promote DMTM efficiency relative to that of the widely reported stepwise DMTM. Moreover, differing from H2O2, which is a type of expensive oxidant, N2O can be largely released from the tail gas of the adipic acid industry (effluent N2O content > 35 vol%) [18], which can effectively solve the source problem of N2O for the N2O DMTM. In light of these advantages, several studies on N2O DMTM operated under moderate temperatures have recently been reported, achieving the superior DMTM activity relative to that of O2 DMTM [19,20,21,22,23,24,25]. As for H2O, it has long been recognized as a solvent to sweep out CH3OH during the DMTM [6,26] until recently, when Bokhoven et al. [27,28] and Liu et al. [29] discovered that H2O could directly participate in the DMTM, working as an oxidant and directing the reaction pathway. Inspired by these literature reports, a proton-transfer reaction of H2O can be found in our previous work [24], pronouncedly the enhancing of CH3OH production (16.8 (absence of H2O) → 242.9 μmol∙g−1cat∙h−1) and the selectivity (3.1 → 71.6%) of Cu0.6%-BETA (BEA) zeolite during the continuous N2O DMTM.
In the present work, to further improve the catalytic behavior of Cu0.6%-BEA, a series of bimetallic FenCum-BEA zeolites with different Fe/Cu ratios (see Table S1) was prepared and evaluated for the H2O-mediated continuous N2O DMTM, based on the acknowledgement that the Fe-modified zeolite is much more favorable for N2O dissociation [17], as well as the assumption that the bimetallic Fe and Cu can provide some unique activity through the synergetic effect. In particular, a strikingly promoted activity can be observed for the best-performing sample of Fe0.6%Cu0.68%-BEA. Specifically, higher CH3OH productivity (259.1 μmol∙g−1cat∙h−1) can be achieved at a much lower reaction temperature (T = 270 °C) relative to that of Cu0.6%-BEA (242.9 μmol∙g−1cat∙h−1, T = 320 °C), while in comparison to the monomeric Fe1.28%- or Cu1.28%-BEA with the same metal loadings, Fe0.6%Cu0.68%-BEA also exhibits overwhelming superiority in both CH3OH productivity (259.1 versus 100.9 and 41.9 μmol∙g−1cat∙h−1) and CH3OH selectivity (71.7% versus 0.28 and 17.6%). These can be closely correlated with the synergetic effect of the neighboring [Fe]--[Cu] cations, which can efficiently favor the active O generation as well as the H2O-mediated proton-transfer reaction. This specific mechanistic insight was illustrated based on the combined experimental (in situ DRIFTS and TPD-MS) and theoretical (DFT and ab initio dynamics (AIMD)) investigations. Overall, the present work systematically investigated reported the synergistic effect of the neighboring [Fe]--[Cu] cations over the FenCum-BEA zeolite, eventually leading to boosted H2O-mediated N2O DMTM activity, the study of which would substantially contribute to other highly efficient zeolite catalyst designs for the continuous DMTM.

2. Result and Discussion

2.1. Physicochemical Property Characterizations

Table S1 lists the physiochemical properties of the investigated FenCum-BEA samples (n and m represent the metal loadings), which were prepared by the tradition ion-exchange method. Briefly, the H-BEA (5g) was mixed with certain amounts (see Table S2) of Fe(NO3)3∙9H2O and Cu(NO3)2∙5H2O and thereby solved by 200 mL deionized water under stirring for 4 h at 100 °C. After being washed, dried (100 °C, 24 h) and calcined (550 °C, 4 h), the final product can be obtained. As noted, for the purpose of comparison the physiochemical properties of the monomeric Fe1.28%- and Cu1.28%-BEA, being prepared by the incipient-wetness impregnation method and with a metal loading amount of 1.28 wt.% (see details in Supplementary Materials), was also listed in Table S1. The small crystallite size (~17–18 mm) associated with a high Brunner−Emmet−Teller (BET) surface area (>500 m2∙g−1) indicates the nano-zeolite catalysts of the prepared FenCum-BEA and H-BEA. The related XRD patterns confirm the characteristic BEA crystal structures of the prepared samples, with no CuOx or FexOy diffraction patterns being observed (see Figure S2a). However, the coexistence of well-dispersed metal oxides cannot be totally ruled out due to their minor amounts, being below the XRD detection limit.
To further identify the chemical states of the loaded Fe and Cu species over the FenCum-BEA samples, the H2 temperature program reduction (H2-TPR) was thereby conducted, with the results being profiled in Figure 1a. As can be seen, there are four types of reduction peak in total, as marked by orange, purple, dark green and blue, which can be assigned to be the reduction of CuO, Fe3+ ion, Cu2+ ions and FexOy, respectively [18,24,30]. The species occupations are further depicted in Figure 1b (see details in Table S3), indicating that the Cu2+ cations constitute the major species over Fe0.38%Cu0.74%, while both the Cu2+ and Fe3+ cations can be observed for Fe0.6%Cu0.68%-BEA (0.56 and 0.40%) and Fe1.0%Cu0.5%-BEA (0.37 and 0.28%), although large amounts of FexOy species (0.74%, Figure 1b) can also coexist with metal cations over Fe1.0%Cu0.5%-BEA. Noteworthily, the Cu2+ cations’ contents (green columns) were higher than those of the Fe3+ cations (purple columns), even for the sample of Fe1.0%Cu0.5%-BEA (0.37 versus 0.28%, Figure 1b) with much higher Fe loadings than those of Cu. This finding reveals that the Cu2+ cations prefer to occupy the Brönsted acid ion-exchange site compared to Fe3+ cations, eventually resulting in FexOy as the major Fe species, especially for the sample of Fe1.0%Cu0.5%-BEA with the highest Fe loadings. The valance states of the loaded Cu and Fe species of FenCum-BEA were further characterized by XPS, as shown in Figure 1c,d. The main peak located at 933.5 (Cu 2p3/2), being associated with a shake-up peak at 944.5 eV (Figure 1c), indicates the characteristic Cu2+ valance state of the surface Cu species [18], while the two main peaks located at 712.3 (Fe 2p3/2) and 725.7 eV (Fe 2p1/2), being associated with the satellite peaks at 718.2 and 732.9 eV (Figure 1d), indicate the 3+ valance state of the surface Fe species [31]. For comparison, H2-TPR and XPS were also conducted on the monomeric Fe1.28% and Cu1.28%-BEA (see Figure 1a and Figure S3a,b), which indicate that both the CuO (0.52%) and Cu2+ (0.76%) cations can be formed over Cu1.28%-BEA. The large reduction domain of Fe3+ species over Fe-BEA-1.28% (250–600 °C of Figure 1a) can be related to the reduction of Fe3+ cations or oxo-cations into Fe2+,18 which can be further confirmed by the H2 consumption quantitation (see Table S3). As noted, in comparison to that of Fe-BEA-1.28%, the Fe3+ cation reduction peaks became narrower for the scenario of FenCum-BEA samples, which can be related to the much lower Fe3+ cation amounts resulting from the much stronger ion-exchange ability of Cu2+ cations at the Brönsted acid site (as stated above).

2.2. Activity Measurement

2.2.1. Activity Comparison with Monomeric Fe- and Cu-BEA

With a basic understanding of the loaded metallic species, the continuous N2O DMTM activity measurements were further conducted over the FenCum-BEA and Fe1.28%- Cu1.28%-BEA samples, as shown in Figure 2a–d and Figure S4, being associated with the CH4 conversion and product selectivities being listed in Table S4. As can be seen, much higher CH3OH productivities (Figure 2a) and selectivities (Figure 2b & Table S4) can be observed for the samples of FenCum-BEA relative to those of the monomeric Fe1.28%- and Cu1.28%-BEA. Especially for the best-performed sample of Fe0.6%Cu0.68%-BEA, it displays a boosted CH3OH productivity (259.1 μmol∙g−1cat∙h−1, Figure 2a), being respectively of 2.6 and 6.2 times high than those of Fe1.28%- and Cu1.28%-BEA (100.9 and 41.9 μmol∙g−1cat∙h−1) with the similar total metal loadings. This finding implies that there probably exists some synergetic effect between the loaded Fe and Cu species, which can efficiently promote the H2O-mediated N2O DMTM activity of Fe0.6%Cu0.68%-BEA. This will be further discussed later based on the combined experimental and theoretical studies. As for the monomeric Fe1.28%-BEA, large amounts of CO2 (selectivity of 97.2%, see Figure 2b) can be formed, being accompanied with the N2O conversion of 99.1% and CH4 conversion of 47.8%, which indicates the overoxidation of CH4 probably due to the strong oxidation property of the generated αO over the Fe cation site; additionally, the Cu1.28%-BEA displays the lowest CH3OH productivity (41.9 μmol∙g−1cat∙h−1), which can be related to the lower reaction efficiency to generate αO of the Cu cations relative to that of Fe as well as the coexistence of large amounts of CuOx species (0.52%, see Table S3) that can readily lead to the overoxidation of CH3OH into CO2 (22.6%, Figure 2b).
In addition to that, it can be found that the simultaneous introduction of 10 vol% H2O into the reaction system can pronouncedly promote the CH3OH productivity (134.8 μmol∙g−1cat∙h−1 → 259.1 μmol∙g−1cat∙h−1, Figure 2a) and CH3OH selectivity (14.4 → 71.7%, Figure 2b) as well as the long-term stability (passing through 25 h’s test, Figure 2d) of the best-performed Fe0.6%Cu0.68%-BEA. This finding gives us a clue that there probably also exits the proton-transfer reaction over the Fe0.6%Cu0.68%-BEA, wherein the H2O molecule could directly participate in the reaction through a proton-transfer route, as proposed in our previous work over the Cu0.6%-BEA zeolite catalyst [24], eventually pronouncedly promoting the CH3OH productivity & selectivity. On the contrary, a large amount of coke can be generated during H2O-absence N2O DMTM over Fe0.6Cu0.68-BEA (67.3%, see Figure 2b), which can be closely related to the undesired side reaction of the accumulated CH3- radicals being generated during CH4 activation over the active site. Meanwhile, it is also worth noting that the Fe0.6%Cu0.68%-BEA can achieve higher CH3OH productivity under a much lower reaction temperature (259.1 μmol∙g−1cat∙h−1, T = 270 °C) than that of Cu0.6%-BEA (242.5 μmol∙g−1cat∙h−1, T = 320 °C) of our previous report [24], which can well verify the viability of the proposed strategy by modifying active site motifs to improve the N2O DMTM activity of Cu-BEA.

2.2.2. Activity Comparison among FenCum-BEA

The diverse catalytic activities of the FenCum-BEA samples can be closely correlated with the chemical states as well as the correlated synergetic effect of the loaded Fe and Cu species. As is well known, the metal cations constitute the active sites for N2O dissociation [17], which are responsible for the production of αO for the scenario of the present reaction system, while the metal oxides readily lead to the overoxidation of CH3OH into CO2 [24]. In this regard, we can deduce that the highest CH3OH productivity of Fe0.6%Cu0.68%-BEA can be closely correlated with its higher Fe and Cu metal cation amounts (0.56 and 0.40%, Figure 1b), which greatly favor the active O production as well as the reaction synergies between these bimetal cations. Similar findings can also be derived for the sample of Fe1.0%Cu0.5%-BEA; however, the relatively lower metal cation amounts (0.37 and 0.28%, Figure 1b) combined with the larger amounts of coexisting FexOy (0.74%, Figure 1b) constitute two major factors leading to the relatively lower CH3OH productivity compared to that of Fe0.6%Cu0.68%-BEA (223.6 versus 259.1 μmol∙g−1cat∙h−1, Figure 2a). As for the sample of Fe0.38%Cu0.74%-BEA, possessing the highest Cu cation amount (0.73%, Figure 1b), it exhibits the lowest CH3OH productivity (108.6 μmol∙g−1cat∙h−1) among the FenCum-BEA samples; this can be closely related to it having the lowest amount of Fe cations (0.08%, Figure 1b), which significantly limits the reaction synergies between the loaded Cu and Fe cations. This finding can also support the proposed assumption that the synergetic effect of the loaded Fe and Cu species (the metal cations) play a vital role in the H2O-mediated N2O DMTM.

2.3. Evolution of the Bimetal Active Center of [Fe-O-Cu] and Thermodynamics Stability Analysis

2.3.1. NO In Situ FTIR

The in situ FTIR utilizing NO as the probe molecule (NO in situ FTIR) was employed to explore the active-site structure evolutions under the reaction condition of the present work over the best-performed Fe0.6%Cu0.68%-BEA. As shown in Figure 3a, after the He (>99.999 vol%, T = 500 °C for 1 h) pretreatment (wine line), two broad v(NO) bands centered at 1873 and 1864 cm−1, being associated with characteristic v(NO) bands belonging to the monomeric metal cation sites including 1962, 1937 and 1913 cm−1 of [Cu]2+ and 1840, 1818, and 1805 cm−1 of reduced-cation sites ([Cu]+ or [Fe]2+) [32,33,34], can be clearly observed. Interestingly, after the further N2O (30 vol% in He, T = 250 °C for 1 h) treatment, an obvious enhancement for the bands at 1873 and 1864 cm−1 (green line) can be clearly observed, wherein the band of 1864 cm−1 was merged as a shoulder of 1873 cm−1, being accompanied with the evident decreasing of v(NO) over the monomeric cation sites, especially for the 1840, 1818, and 1805 cm−1 of reduced metal cation sites.
After carefully analyzing this experimental phenomenon as well as scanning the literature reports [32,33,34], we would assign the major bands at 1873 and 1864 cm−1 to be v(NO) over the evolved bimetal [Fe-O-Cu] site, being in the form of [NO-Fe-O-Cu] and [Fe-O-Cu-NO]. Specifically, after the He pretreatment, part of the monomeric Fe3+ and Cu2+ cations would be auto-reduced by displaying the characteristic v(NO) bands at 1840, 1818 and 1805 cm−1; concurrently, parts of the neighboring Fe3+ and Cu2+ cations would prefer forming the bimetallic [Fe-O-Cu] through the dihydroxylation by displaying tow-obvious NO adsorption bands at 1873 and 1864 cm−1 (see Figure 3a). The subsequent N2O treatment would further favor [Fe-O-Cu] formation through the interaction with reduced neighboring Fe and Cu cations. The free energy barrier of this step was simulated to be only 0.18 eV at T = 270 °C (as will be further discussed later). In this regard, the further enhancements of v(NO) at 1873 and 1864 cm−1 ([Fe-O-Cu] site) being accompanied with the simultaneous decreasing of v(NO), especially at the reduced metal cation sites of (1840, 1818, 1805 cm−1), can be observed in Figure 3a. To exclude the possibility of generations of large amounts of [Cu-O-Cu] or [Fe-O-Fe] species, the UV-vis and ab initio thermodynamics analyses have also been conducted (See Figure S5a,b): (i) no obvious characteristic UV-vis band belonging to [Cu-O-Cu] (~440 nm) can be observed in Figure S5a, although the same formation Gibbs free energy (ΔfG = −7.62 eV) can be achieved for the [Cu-O-Cu] and [Fe-O-Fe] site (see Figure S5b); (ii) the [Fe-O-Cu] exhibits much lower formation Gibbs free energy than that of [Fe-O-Fe] (−7.62 versus 7.15 eV, see Figure S5b), which indicates that it would be much more preferable to form [Fe-O-Cu] under the reaction condition of the present work due to the higher thermodynamic stabilities; (iii) the split two obvious v(NO) bands at 1873 and 1864 cm−1 (see Figure 3a of He pretreatment) can also give good evidence of the formation of [Fe-O-Cu]2+ species due to the fact that the [Fe-O-Fe] can only show up one type of characteristic v(NO) band around 1875 cm−1 [32].
In light of the above statement, we can deduce that the [Fe-O-Cu] would constitute the major active site during the N2O DMTM over Fe0.6%Cu0.68%-BEA by displaying much more extensive band intensities at 1873 and 1864 cm−1 than those of 1840, 1818 and 1805 cm−1 after the N2O pretreatment (see Figure 3a). To make comparisons, the NO in situ FTIR was also conducted over Fe0.38%Cu0.74%- and Fe1.0%Cu0.5%-BEA (being after in situ N2O pretreatment), as shown in Figure 3b. The characteristic v(NO) bands over the bimetallic [Fe-O-Cu] site (1873 and 1864 cm−1) can also be clearly observed for these two samples, the band area of which were however much lower than that of Fe0.6%Cu0.68%-BEA, as quantified in Figure 3c. This finding gives us a clue that the much higher N2O DMTM activity of Fe0.6%Cu0.68%-BEA (Figure 2a) can be closely correlated with its higher amounts of evolved [Fe-O-Cu] active species in comparison to other two samples.

2.3.2. H2-TPR after N2O and O2 Pretreatment

The H2-TPR was further conducted over the N2O and O2 (taken as a reference) pretreated Fe0.6%Cu0.68%-BEA to confirm the evolved [Fe-O-Cu] site. Similarly, a pretreatment strategy to that of NO in situ FTIR was applied, wherein the sample were initially pretreated by He at T = 500 °C for 1 h before the subsequent N2O (30 vol% in He, T = 250 °C for 1 h) and O2 (15 vol% in He, T = 500 °C for 1 h) pretreatments; and to obtain the better signals, the mass spectrometer (MS) was utilized to monitor the H2 (m/e = 2) consumptions. As noted, the O2 pretreatment temperature of 500 °C was chosen according to literature reports [26] that it commonly needs a high temperature (T > 500 °C) to generate active O species during DMTM over the zeolitic catalyst. As shown in Figure 3d, three types of H2 reduction peaks can be clearly observed, which can be respectively corresponding to the reduction of FexOy (~500 °C, blue), [Fe-O-Cu] (650 °C, purple) and the Fe2+/Cu+ cations (900 °C, dare green). Obviously, in comparison with the O2-pretreated Fe0.6%Cu0.68%-BEA, much higher amounts of [Fe-O-Cu] species (purple peak) can be formed over the N2O-pretreated Fe0.6%Cu0.68%-BEA, which also constitute the major active species being in good agreement with the NO in situ FTIR result of Figure 3a. On the contrary, only trace [Fe-O-Cu] can be formed over the Fe0.6%Cu0.68%-BEA due to the extremely low dissociation activity of O2. This finding is in good agreement with the significantly lower O2 DMTM activity (~1 μmol∙g−1cat∙h−1 of CH3OH, see Figure S6) relative to that of N2O DMTM. Therefore, the evolution of the [Fe-O-Cu] site through the in situ interaction with N2O at low temperature (T = 250 °C) can be safely confirmed based on the H2-TPR of Figure 3d over the Fe0.6%Cu0.68%-BEA, which also constitutes the major active sites during N2O DMTM over Fe0.6%Cu0.68%-BEA.

2.3.3. Ab Initio Thermodynamics (AIT) Analysis

In this section, DFT-based ab initio thermodynamics (AIT) analysis was further conducted to evaluate the thermostability of the [Fe-O-Cu] site under the reaction conditions of the present work (10 vol% H2O, 30 vol% N2O and 1atm), as shown in Figure 4b,c, wherein the other active-site motifs, including the monomeric [Cu], [Cu-O], [CuOH] and neighboring bimetallic [Fe]--[Cu] sites, were also taken into account. Obviously, the [Fe-O-Cu] site exhibited dramatically lower ΔG values than the other active sites, indicating the stronger thermodynamic stability of the evolved [Fe-O-Cu] site. As noted, the much higher thermodynamic stability of [Fe-O-Cu] relative to that of [Fe]--[Cu] indicates that it is also thermodynamically favorable to evolve into [Fe-O-Cu] through the in situ interaction of [Fe]--[Cu] with N2O under the reaction conditions of the present work.

2.4. Mechanistic Insight into Synergetic Effect of Neighboring Fe and Cu Cations during H2O-Mediated N2O DMTM

2.4.1. Experimental Mechanism Study

Based on the above studies, the active-site motif evolutions have been clarified. In this section, the H2O-mediated N2O DMTM reaction mechanism was further investigated by in situ FTIR and D2O isotopic tracer techniques to explore the reaction synergy of the loaded neighboring Cu and Fe cations over the best-performing Fe0.6%Cu0.64%-BEA.
(a)
In situ FTIR
As shown in Figure 5a, three types of v(CHx) vibration frequencies can be clearly observed after introducing the reactant mixtures (N2O, CH4 and H2O) into the system at T = 270 °C, corresponding to the v(CH) of CH4 (3015 cm−1), the v(CH3) of generated CH3OH (2963, 2853 cm−1) and the v(CH3) of the metal cation site (2926 cm−1, [M-CH3]), respectively [24]. As noted, except for the band at 2926 cm−1 belonging to the v(CH3) of [M-CH3], no obvious bands related to the v(CH3) of methoxy groups (CH3O, 2980, 2869 and 2823 cm−1) [27] can be observed, which implies that the H2O-mediated N2O DMTM follows the radical mechanism in which the CH4 is activated into the radical of CH3- and OH- at the evolved active site of [Fe-O-Cu] of Fe0.6%Cu0.64%-BEA. To further verify this assumption, in situ FTIR by introducing the CH4 (2 vol% in He) into the N2O-pretreated Fe0.6%Cu0.64%-BEA was further conducted, wherein the H2O was not fed into the system to avoid its influence on v(OH). As shown in Figure 5c, the band at 3675, being related to the v(OH) at metal cations site (α-site), can be clearly observed, which can solidly verify the deduction that the CH4 activation would follow the radical mechanism over Fe0.6%Cu0.64%-BEA, as illustrated by Equations (1) and (2), wherein the Ma and Mb in Equation (2) represent the metal cations of Cu or Fe, respectively (the specific states will be determined by the DFT, as will be further discussed later).
N 2 O + [ Cu ] - - [ Fe ] - Z [ Cu - O - Fe ] - Z + N 2
C H 4 + [ Cu - O - Fe ] - Z [ M a - C H 3 ] - - [ M b - OH ] - Z
Therefore, based on the in situ FTIR, we can deduce that the neighboring [Fe]--[Cu] would initially interact with N2O, generating [Fe-O-Cu] (α-site), which thereby favors CH4 activation following a radical mechanism.
(b)
D2O isotopic tracer experiment
According to our previous study [24], the H2O molecules could directly participate in N2O DMTM through a proton-transfer route after the activation of CH4 at the evolved [Cu-O-Cu] site of Cu0.6%-BEA. In light of that, the temperature-programmed surface reaction (TPSR-) MS-based D2O isotopic tracer technique was also employed in the present work to explore whether the proton-transfer reaction of H2O also occurs during the H2O-mediated N2O DMTM over the Fe0.6%Cu0.68%-BEA. As shown in Figure 6a, the obviously emerged peaks of CH3OD (33) and D3O+ (33), accompanied by the simultaneous increase/decline in the signals of HOD (19)/[D2O (20)], respectively, provide us with solid evidence that the D2O can also participate in the N2O DMTM through a proton-transfer route after CH4 activation over the evolved [Fe-O-Cu] site of Fe0.6%Cu0.68%-BEA (see Equation (3)).
[ M a - C H 3 ] - - [ M b - OH ] - Z + 2 D 2 O D 3 O + Proton   trasfer C H 3 OD + HOD + D 2 O + [ M a ] - - [ M b ]
To make comparisons, the TPSR-MS without D2O addition was also conducted over Fe0.6%Cu0.68%-BEA, as shown in Figure 6b. It is observed that the signals of CH3OD (33), D3O+ (22), HOD (19) and D2O (22) can also be detected, especially for D2O (22), emerging as a small peak. In fact, this can be related to the overoxidation of CH4 (99.999%), which contains small isotopic abundance of the atomic D, to produce minor D2O that can further participate in the N2O DMTM reaction following the proton-transfer route, eventually generating the detected signals of CH3OD (33), HOD (19) and D3O+ (22) in Figure 6b. Therefore, this finding can also support the H2O proton reaction during N2O DMTM over Fe0.6%Cu0.68%-BEA.
Figure 3c further compares the peak areas of the signals of CH3OH (31) and CH3OD (33) derived from Figure 3a,b, which demonstrates that the addition of D2O would lead to N2O DMTM following the proton-transfer route by producing much higher amounts of CH3OD (33) relative to that of CH3OH (31). Small amounts of CH3OH can also be generated during the D2O-mediated N2O DMTM, which may be related to the N2O DMTM over the monomeric cation sites, where no proton-transfer reactions can occur [24]. In addition to that, herein, we suggest that the D2O direct-reaction route may also exist during the D2O-mediated N2O DMTM, wherein one D2O molecule can dissociatively react with the [CuCH3]--[FeOH] to generate [CuCH3OD]--[FeOHD] (see Equation (4)). This reaction route will be discussed later based on the DFT and with the results being further compared with that of the H2O proton-transfer route.
D - O a D + [ CuCH 3 ] - - [ FeOH ] [ CuCH 3 O a D ] - - [ FeOHD ]

2.4.2. Theoretical Mechanism Simulations by DFT and AIMD

To shed deeper mechanistic insight into reaction synergy of the neighboring Fe and Cu cations of Fe0.6%Cu0.68%-BEA, both the DFT and AIMD were employed in this part to simulate the reaction mechanism based on the constructed FeCu-BEA model with the neighboring [Fe]--[Cu] as the active site (auto-reduced form after He pretreatment, see Figure S7). In total, three types of reaction mechanisms were proposed, including the H2O absence mechanism (taken for comparison) and the H2O proton-transfer mechanism as well as the H2O direct-reaction mechanism, as illustrated in detail in in Scheme 1. The derived reaction energy diagrams are depicted in Figure 7a and Figure S8a,b, as stated in detail below.
(a)
N2O dissociation and CH4 activation steps
As shown in Scheme 1, the same reaction routes are followed during the initial N2O dissociation and CH4 activation steps (marked by black arrow) for the three proposed types of mechanisms. Specifically, the N2O molecule readily interacts with the neighboring [Fe]--[Cu] site to generate the α-site of [Fe-O-Cu] by overcoming an extremely low Gibbs free energy barrier of ΔG = 0.18 eV (see TS-I of Figure 7a), which is much lower than those derived from Cu-BEA of our previous work (ΔG = 0.91 and 1.31 eV of di-copper [Cu]--[Cu] site and monomeric [Cu] site, respectively) [24]. This finding demonstrates that the synergetic effect of the neighboring [Fe]--[Cu] could significantly promote N2O dissociation to generate the active O, which constitutes the major reason leading to the much higher CH3OH productivity and lower operation temperature of Fe0.6%Cu0.68%-BEA relative to that of Cu0.6%-BEA of our previous work. After that, the CH4 can be subsequently activated into CH3- and OH- (III→IV, Figure 7a); a similar reaction route of III→TSII has also been reported by the literature [35,36], following a radical mechanism as revealed by the in situ FTIR (Figure 5a–d), the free energy barrier of which is calculated to be 0.92 eV (TS-II of Figure 7a). Noteworthily, according to the DFT calculations (Figure S9a,b), the CH4 is both kinetically and thermodynamically much more favorable to be activated into the [Fe-OH]--[Cu-CH3] state relative to the scenario of the [Fe-CH3]--[Cu-OH] state.
(b)
Further reaction of [Fe-OH]--[Cu-CH3] in the absence and presence of H2O
H2O absence mechanism. Three types of routes were proposed for the further reaction of [Fe-OH]--[Cu-CH3]. As shown in Scheme 1 and Figure S8a, the H2O absence route would follow the widely reported rebound mechanism, wherein the radical of CH3- could migrate from the [Cu]2+ site to interact with OH-, producing [Fe(CH3OH)] (IV→V) by overcoming a relatively high free energy barrier of 1.21 eV (TS-III). After crossing another energy barrier of 1.12 eV (VI), the CH3OH can be eventually desorbed.
H2O proton-transfer mechanism. Based on the above in situ FTIR (Figure 5a–d) and D2O isotopic transfer (Figure 6a,b) studies, the AIMD-based metadynamic simulations were employed to explore the H2O proton-transfer mechanism (see Movie S1, T = 270 °C, P = 1 atm). As shown in Figure 7a, two H2O molecules can be initially inserted between the generated radicals of [FeOH]--[CuCH3] (model V), which is much more thermodynamically favorable than the scenario of the H2O direct-reaction mechanism with one H2O molecule being adsorbed on the [Cu]2+ site (as will be further discussed later). After that, one H2O molecule would dissociatively interact with the neighboring [CuCH3] and H2O molecule to produce [CuCH3OH]+ and H3O+, respectively (as detected by means of the D2O isotopic transfer technique in Figure 6a). Subsequently, one H+ could quickly migrate from the H3O+ to [FeOH], generating [FeOH2]+ (adsorbed H2O), and the H3O+ would be restored to H2O. Finally, the adsorbed CH3OH and H2O ([CuCH3OH]--[FeOH2]) could be barrierlessly desorbed to regenerate the [Fe]--[Cu] active site. Noteworthily, an extremely low energy barrier of 0.07 eV can be observed during the H2O proton-transfer process, which is comparable to the value of 0.03 and 0.05 eV for the scenario of Cu0.6%-BEA presented in our previous work [24]. This finding validates the suggestion that the synergistic effect of the neighboring [Fe]--[Cu] site can also favor the H2O proton-transfer reaction to pronouncedly promote CH3OH productivity over Fe0.6%Cu0.68%-BEA, as observed in Figure 2a. The Gibbs free energy surface is further depicted in Figure 7b,c as a function of the coordinate number of CNa(O-H) and CNb(O-H) (see Figure S1a), which clearly displays the reaction energy variations from one energy minimum to another minimum.
H2O direct-reaction mechanism. As stated above, the H2O direct-reaction route may also occur during the H2O-mediated DMTM over Fe0.6%Cu0.68%-BEA. In this regard, the H2O direct-reaction mechanism was further simulated by DFT. As shown in Figure S8b, initially, one H2O site forms [H2OCuCH3]+ (VII). Thereby, the H2O can be readily dissociated from (H-aOaH) to react with the [FeOH] and [Cu-CH3], producing [FeOH2]--[aHaOCuCH3] (ΔG = 0.24 eV, TS-III). After the rebounding of CH3- and aOaH- by crossing a free energy barrier of ΔG = 1.14 eV (TS-V), [CuCH3aOaH] can be formed. With the assistance of H2O (being desorbed from [FeOH2]), the CH3aOaH can be barrierlessly released from the [Cu]2+ site.
(c)
Reaction mechanism comparisons
In comparison to the H2O absence mechanism, the slightly lower barrier of the radical rebound step (1.14 (TSIV, marked blue) versus 1.21 eV (TSIII, marked green) (see inserted energy diagram of Figure 7a) combined with the barrierless desorption of CH3OH indicates that H2O could also favor CH3OH production through the direct-reaction route. However, the H2O-mediated N2O DMTM would generally follow the H2O proton-transfer route due to the extremely low energy barrier of 0.07 eV. To make quantitative comparisons, microkinetic modeling was further conducted for these three types of mechanisms, with the results being profiled in Figure S10a–d. As can be seen (Figure S10a), the overall reaction rates were predicted to be 3.64 × 109 s−1 (H2O proton-transfer route), 1.05 × 104 s−1 (H2O direct-reaction route) and 7.64 × 10−10 s−1 (H2O absence route), respectively, which quantitatively verified that the significant promotion effect of H2O during N2O DMTM is generally through the H2O proton-transfer route, being favored by the synergistic effect of the neighboring [Fe]--[Cu] site. It is worth noting that the DFT-calculated reaction rate of the H2O-mediated N2O DMTM was several orders of magnitude higher than that of the H2O-absence N2O DMTM; however, it was only two times higher for the scenario of the experimental result of Figure 2a (CH3OH productivity). This discrepancy is related to the utilized DFT calculation method during the microkinetic modeling, especially for the pre-exponential factor calculations derived by the frequency calculations. However, it does not affect the parallel comparisons of the DFT-based microkinetic modeling results.

2.4.3. Illustration of the Synergistic Effect of Neighboring Fe and Cu Cations

Based on the above experimental and theoretical mechanism studies, we can obtain a comprehensive understanding of the synergistic effect of the neighboring [Fe]--[Cu] cations during the H2O-mediated N2O DMTM. On the one hand, it can efficiently promote N2O dissociation to generate active O ([Fe-O-Cu]) by extensively reducing the N2O dissociation barrier from 0.91 eV (the neighboring [Cu]--[Cu] site [24]) to 0.18 eV. On the other hand, it can also motivate the reaction following the H2O proton-transfer route by crossing an extremely low free energy barrier of 0.07 eV, thereby pronouncedly enhancing CH3OH productivity/selectivity, as well as long-term stability (reducing carbon depositions). To provide deeper insights, from the electronic structure point of view, into the synergistic effect of neighboring [Fe]--[Cu] on N2O dissociation to generate active O, the theoretical simulations including electronic density difference, the Bader charge analysis and density of state (DOS) were further conducted (see Figure 8a–d) based on the optimized N2O-adsorption model and DFT, wherein the neighboring [Cu]--[Cu] was also taken into account for comparison.
As can be seen (Figure 8a,b), much higher amounts of charge transfers occur after the adsorption of N2O onto the the [Fe]--[Cu] site, wherein the bimetal cations of [Fe]--[Cu], especially for the cationic [Fe], acting as the charge donors, donate much greater charges (0.2 and 0.44) to the N2O molecule relative to that of the neighboring [Cu]--[Cu] cations (0.13 and 0.04, see Figure 8c). This finding indicates that the neighboring [Fe]--[Cu] site can exert much a stronger electric field effect on the adsorbed N2O, which is greatly favorable for the pre-activation of N2O and subsequent O-N2 bond fracture to produce active O. As observed (Figure S11a,b), much more extensive structure distortion can be clearly observed for the adsorbed N2O molecule onto the [Fe]--[Cu] site: ∠N-N-O bending from 180° to 129.7° versus (180° → 175°) of the [Cu]--[Cu] site. The stronger charge transfer effect (or electric filed effect) of the [Fe]--[Cu] site can be explained by the total DOS (Figure 8d) as well as partial DOS (PDOS) results (Figure S12a–h), which indicates that the [Fe]--[Cu] site not only possesses much lower electron-transfer band gaps but also can reduce the band gaps of the N and O atoms, N2O (0.77 versus 1.83 eV), eventually significantly promoting the charge transfers between the [Fe]--[Cu] active site and the N2O molecule. In light of the above studies, we can conclude that the much more efficient synergistic effect of the [Fe]--[Cu] active site (relative to that of [Cu]--[Cu] site) during the N2O dissociation step can be interpreted by its stronger electric field effect, leading to the significantly enhanced electron transfer between the bimetal active site and the N2O molecule due to the lowered band gaps of bimetal cations as well as the N and O atoms of the N2O molecule.

3. Materials and Methods

3.1. Experimental Methods

3.1.1. Catalyst Preparation

The commercial H-BEA with the Si/Al of 12.5 (mole ratio) was purchased from Tianjin Nankai University Catalyst Co., Ltd. of China, based on which a series of FenCum-BEA zeolites (Table S1) were prepared by means of the traditional wet ion-exchange method. For the purpose of comparison, the monomeric Fe1.28%- and Cu1.28%-BEA, with the total metal loading amounts being the same as those of Fe0.6%Cu0.68%-BEA (1.28 wt.%), were prepared by means of an incipient-wetness impregnation method. More details regarding the catalytic preparations are stated in the Supplementary Materials.

3.1.2. Catalyst Characterizations

The characterizations of X-ray diffraction (XRD), element content (determined by inductively coupled plasma (ICP), X-ray photoelectron spectroscopy (XPS), N2 adsorption/desorption, H2-temperature programmed reduction (H2-TPR), UV-vis diffuse-reflectance, in situ FTIR and D2O isotopic tracer were conducted as described in detail in our previous work [24]—alternatively, please see details in the Supplementary Materials.

3.1.3. Activity Measurement

The continuous N2O DMTM in both scenarios of the presence and absence of H2O (10 vol%) was conducted over a fixed bed reactor (O.D. 11 mm, I.D. 8 mm, L485 mm) under a total flow rate of 100 mL∙ min−1 [GHSV (gas hourly space velocity) = 12,000 h−1]. The reaction products were analyzed by a gas chromatograph (SHIMADZU GC-2014) being equipped with the thermal conductivity detector (TCD) and flame ionization detector (FID). To prevent condensation, the gas line from the point of liquid injection to then GC unit was heated (T = 200 °C) by resistive heating tape. More details regarding activity measurement can be found in the Supplementary Materials or our previous work [24].

3.2. Computational Methods

3.2.1. Constructed Model

The model of FenCum-BEA with the neighboring [Fe]--[Cu] bimetal active site was constructed based on the BEA structure (a = 12.632, b = 12.632 and c = 9.421 Å) from the database of IZA [37], wherein the framework Al was located at the T2 and T5 site according to our previous work [24], and the Si atoms were set to be fixed during the structure optimization and transition state (TS) calculations to keep the structure of BEA.

3.2.2. Computational Method

The periodic density functional theory (DFT) analysis was performed based on the Vienna ab-initio simulation package (VASP) [38], which was employed in the present work to simulate the N2O DMTM mechanism (in the presence and absence of H2O) and conduct microkinetic analysis, electronic structure analysis (including electronic density difference, Bader charge and density of state (DOS)) and ab initio thermodynamics analysis (evaluating the thermostabilities of active-site motif and generating radical intermediate). The ab initio molecular dynamics (AIMD) analysis was conducted based on CP2K code [39] to explore the H2O proton-transfer mechanism, wherein the AIMD-based metadynamic simulations were conducted utilizing two types of collective variables, the coordination numbers (CN) of [CNaO-H] and [CNbO-H] (as illustrated in Figure S1a). The detailed method descriptions of DFT and CP2K can be found in the Supplementary Materials or our previous work [24].

4. Conclusions

The present work systematically investigated the H2O-mediated continuous N2O DMTM over a series of FenCum-BEA zeolites based on combined experimental and theoretical approaches. The strikingly promoted CH3OH productivity (259.1 μmol∙g−1cat∙h−1) and selectivity (71.7%), as well as long-term reaction stability, can be achieved over the best-performed sample of Fe0.6%Cu0.68%-BEA, which is closely correlated with the synergistic effect of the loaded neighboring Fe and Cu cations. On the one hand, it is much more favorable for N2O dissociation to generate the αO (bridge O of [Fe-O-Cu]2+, ΔG = 0.18 eV at T = 270 °C) due to the exerted strong electric field effect. On the other hand, it can motivate the N2O DMTM reaction following a H2O-mediated proton-transfer route to produce CH3OH by crossing a much lower energy barrier (ΔG = 0.07 eV), eventually pronouncedly enhancing CH3OH production and desorption as well as the catalytic reaction’s long-term stability (through efficiently reducing carbon depositions). Generally, the present work reported the synergistic effect of neighboring Fe and Cu cations over the FenCm-BEA zeolite, which can efficiently promote its CH3OH productivity/selectivity during H2O-mediated N2O DMTM. Thereby, employing the synergistic effect of bimetallic cations to modify zeolitic activity constitutes a promising strategy for a highly efficient catalyst design for the DMTM.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/catal11121444/s1: Figure S1: (a) Illustration of selected coordination number (CN) during the AMID-based metadynamic simulations: CNa(O-H) (red dash line) and CNb(O-OH) (green dash line) respectively represent the O-H bond of H2O molecules; CNc(O-H) (blue arrow) and CNd(O-H) (pink arrow) respectively represent the to be generate O-H bond of H3O+ and H2O (over the Fe site); (b) variations of derived CN numbers indicating (O-H) bond factures the H2O molecule to follow the H2O-proton-transfer reaction at t < 50 fs, Figure S2: (a) X-ray diffraction patters and (b) isothermal adsorption curves (derived by N2 adsorption/desorption) of FenCum-BEA and monomeric Fe1.28%- and Cu1.28%-BEA; (c) TEM-EDX mapping results of Fe0.6%Cu0.68%-BEA-0.6% (the Fe and Cu species are uniformly distributed on the catalyst), Figure S3: XPS characterization results of (a) Fe1.28%- and (b) Cu1.28%-BEA (the metal cations constitute major metallic species). Figure S4: Activity measurement results of N2O-DMTM in both presence and absence of H2O over the prepared bimetallic FenCum-BEA and monomeric Fe1.28%-, Cu1.28%-BEA samples after 5 h’s reactions: (a) CH3OH productivity; (b) product selectivity; (c) CH4 conversion; (d) N2O conversion; reaction condition N2O:CH4:H2O:He = 30:15:10:45 (presence of H2O); and O2:CH4:He = 30:15:55 (absence of H2O); GHSV = 12,000 h−1 and T = 270 °C, Figure S5: (a) UV-vis spectra of Fe0.6%Cu0.68%-BEA being respectively pretreated by N2O at T = 200, 250 and 300 °C; no obvious band emerged at 440 nm belong to the characteristic peak of [Cu-O-Cu]; (b) thermodynamic stabilities comparisons of diverse bimetallic active sites, including [Cu-O-Cu], [Fe-O-Cu] and [Fe-O-Fe] based on the DFT-based ab initio thermodynamic analysis by taking the reaction condition of present work into account: 10 vol% H2O, 30 vol% N2O, and total pressure of 1 atm; obviously, the [Fe-O-Cu] & [Cu-O-Cu] possessed much lower formation Gibbs free energy than that of [Fe-O-Fe]2+ site, indicating higher thermodynamic stability; as noted, the data of ΔG for [Cu-O-Cu]-Z was covered by those of [Fe-O-Cu]-Z, Figure S6: O2-DMTM activity measurement result over Fe0.6%Cu0.68%-BEA with the reaction condition being similar to that of N2O-DMTM:O2:CH4:H2O:He = 30:15:10:45, GHSV = 12,000 h-1 and T = 270 °C, Figure S7: Optimized periodic FeCu-BEA model with neighboring [Fe]--[Cu] as the active site, Figure S8: (a) DFT simulated H2O absence mechanism during N2O-DMTM over construed FenCum-BEA model with neighboring [Fe]--[Cu] as the active site, (b) DFT simulated H2O direct reaction mechanism during N2O-DMTM over construed FenCum-BEA model with the neighboring [Fe]--[Cu] as active site with the derived Gibbs free energy being compared with that of H2O absence mechanism (see inserted energy diagram), Figure S9: (a) kinetic comparisons of radical evolutions during CH4 activation step; (b) thermostability comparisons of generated intermediate radicals in [FeCH3]--[CuOH] and [FeOH]--[CuCH3] state based on ab initio thermodynamics taking the reaction conditions of 10 vol% H2O, 30 vol% N2O and 1 atm into account, Figure S10: Microkinetic modeling results based on DFT at T = 270 °C: (a) overall reaction rate of different reaction mechanisms; surface coverage variations as a function of reaction time t of (b) H2O absence route; (c) H2O direct reaction route; and (d) H2O proton transfer reaction route. Figure S11: Optimized N2O-adsorption model of with (a) neighboring [Fe]--[Cu] site and (b) neighboring [Cu]--[Cu] site; Si (yellow), O (red), N (blue), Al (pink), Cu (orange) and Fe (purple), Figure S12: Partial density of sate (PDOS) of bimetal cations and N, O atoms of N2O based on N2O-adsorption models; (a–d) [Feb], [Cua], atomic O and end N of N2O over neighboring [Fe]--[Cu] model; (e–h) [Cub], [Cua], atomic O and end N of N2O over neighboring [Cu]--[Cu] model, Figure S13: Reaction condition determination based on the N2O-DMTM measurement over Fe0.6%Cu0.68%-BEA: (a) reaction temperature (T); (b) introduced H2O content, Figure S14: N2O-DMTM activity measurement result over FenCum-BEA and monomeric Fe1.28%-, Cu1.28%-BEA in presence (a–b) and absence (c–d) of H2O for 5 h: (a,c) CH3OH productivity; (b,d) N2O conversion, Figure S15: Thermogravimetry (TG) and differential thermogravimetric analysis (DTG) curves of the Fe0.6%Cu0.68%-BEA-0.6% samples being after 25 h’s long-term test in absence (a) and presence (b) of 10vol% H2O, Figure S16: 29Si MAS NMR of Fe0.6%Cu0.68%-BEA being respectively before and after the 5 h’s reaction (H2O-mediated N2O-DMTM), Table S1: Physicochemical property of investigated samples, Table S2: Consumption of metal salts for catalysts preparation, Table S3: Loaded Fe and Cu species quantifications based on H2-TPR, Table S4: Methane conversion and products selectivity of investigated samples, Table S5: (a) Elementary steps and kinetic parameters involved in microkinetic modeling for the H2O absence mechanism, (b) Elementary steps and kinetic parameters involved in microkinetic modeling for the H2O direct reaction mechanism, (c) Elementary steps and kinetic parameters involved in microkinetic modeling for the H2O-proton-transer mechanism. Movie S1: AIMD-based metadynamics simulation of H2O-mediated proton transfer route during N2O-DMTM over FenCum-BEA.

Author Contributions

Conceptualization, N.L.; methodology, Y.L.; validation, C.D.; investigation, Y.L.; resources, B.C.; data curation, R.X.; writing—original draft preparation, Y.L.; writing—review and editing, N.L.; visualization, G.Y.; supervision, B.C.; project administration, N.W.; funding acquisition, J.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Key Research and Development Program of China (Projects: 2017YFC0210905) and National Natural Science Foundation of China (No. 21906003).

Data Availability Statement

Data is available upon request to the corresponding authors.

Acknowledgments

We acknowledge the final support from the National Key Research and Development Program of China (Projects: 2017YFC0210905) and National Natural Science Foundation of China (No. 21906003).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Agarwal, N.; Freakley, S.J.; McVicker, R.U.; Althahban, S.M.; Dimitratos, N.; He, Q.; Morgan, D.J.; Jenkins, R.L.; Willock, D.J.; Taylor, S.H.; et al. Aqueous Au-Pd colloids catalyze selective CH4 oxidation to CH3OH with O2 under mild conditions. Science 2017, 358, 223–227. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Luo, L.; Luo, J.; Li, H.; Ren, F.; Zhang, Y.; Liu, A.; Li, W.-X.; Zeng, J. Water enables mild oxidation of methane to methanol on gold single-atom catalysts. Nat. Commun. 2021, 12, 1218. [Google Scholar] [CrossRef] [PubMed]
  3. Ravi, M.; Ranocchiari, M.; Van Bokhoven, J.A. The Direct Catalytic Oxidation of Methane to Methanol-A Critical Assessment. Angew. Chem. Int. Ed. 2017, 56, 16464–16483. [Google Scholar] [CrossRef]
  4. Li, M.; Shan, J.; Giannakakis, G.; Ouyang, M.; Cao, S.; Lee, S.; Allard, L.F.; Flytzani-Stephanopoulos, M. Single-step selective oxidation of methane to methanol in the aqueous phase on iridium-based catalysts. Appl. Catal. B Environ. 2021, 292, 120124. [Google Scholar] [CrossRef]
  5. Kulkarni, A.R.; Zhao, Z.-J.; Siahrostami, S.; Nørskov, J.K.; Studt, F. Monocopper Active Site for Partial Methane Oxidation in Cu-Exchanged 8MR Zeolites. ACS Catal. 2016, 6, 6531–6536. [Google Scholar] [CrossRef]
  6. Dinh, K.T.; Sullivan, M.; Serna, P.; Meyer, R.J.; Dincă, M.; Román-Leshkov, Y. Viewpoint on the Partial Oxidation of Methane to Methanol Using Cu- and Fe-Exchanged Zeolites. ACS Catal. 2018, 8, 8306–8313. [Google Scholar] [CrossRef] [Green Version]
  7. Jeong, Y.R.; Jung, H.; Kang, J.; Han, J.W.; Park, E.D. Continuous Synthesis of Methanol from Methane and Steam over Copper-Mordenite. ACS Catal. 2021, 11, 1065–1070. [Google Scholar] [CrossRef]
  8. Hammond, C.; Forde, M.; Ab Rahim, M.H.; Thetford, A.; He, Q.; Jenkins, R.L.; Dimitratos, N.; Lopez-Sanchez, J.A.; Dummer, N.; Murphy, D.; et al. Direct Catalytic Conversion of Methane to Methanol in an Aqueous Medium by using Copper-Promoted Fe-ZSM-5. Angew. Chem. Int. Ed. 2012, 51, 5129–5133. [Google Scholar] [CrossRef]
  9. Hammond, C.; Dimitratos, N.; Jenkins, R.L.; Lopez-Sanchez, J.A.; Kondrat, S.A.; ab Rahim, M.H.; Forde, M.M.; Thetford, A.; Taylor, S.H.; Hagen, H.; et al. Elucidation and Evolution of the Active Component within Cu/Fe/ZSM-5 for Catalytic Methane Oxidation: From Synthesis to Catalysis. ACS Catal. 2013, 3, 689–699. [Google Scholar] [CrossRef]
  10. Xu, J.; Armstrong, R.D.; Shaw, G.; Dummer, N.; Freakley, S.; Taylor, S.H.; Hutchings, G.J. Continuous selective oxidation of methane to methanol over Cu- and Fe-modified ZSM-5 catalysts in a flow reactor. Catal. Today 2016, 270, 93–100. [Google Scholar] [CrossRef] [Green Version]
  11. Yu, T.; Li, Z.; Lin, L.; Chu, S.; Su, Y.; Song, W.; Wang, A.; Weckhuysen, B.M.; Luo, W. Highly Selective Oxidation of Methane into Methanol over Cu-Promoted Monomeric Fe/ZSM-5. ACS Catal. 2021, 11, 6684–6691. [Google Scholar] [CrossRef]
  12. Wang, L.; Li, Z.; Wang, Z.; Chen, X.; Song, W.; Zhao, Z.; Wei, Y.; Zhang, X. Hetero-Metallic Active Sites in Omega (MAZ) Zeolite-Catalyzed Methane Partial Oxidation: A DFT Study. Ind. Eng. Chem. Res. 2021, 60, 2400–2409. [Google Scholar] [CrossRef]
  13. Dandu, N.K.; Adeyiga, O.; Panthi, D.; Bird, S.A.; Odoh, S.O. Performance of density functional theory for describing hetero-metallic active-site motifs for methane-to-methanol conversion in metal-exchanged zeolites. J. Comput. Chem. 2018, 39, 2667–2678. [Google Scholar] [CrossRef] [PubMed]
  14. Wang, G.; Huang, L.; Chen, W.; Zhou, J.; Zheng, A. Rationally designing mixed Cu–(μ-O)–M (M = Cu, Ag, Zn, Au) centers over zeolite materials with high catalytic activity towards methane activation. Phys. Chem. Chem. Phys. 2018, 20, 26522–26531. [Google Scholar] [CrossRef] [PubMed]
  15. Sazama, P.; Moravkova, J.; Sklenak, S.; Vondrova, A.; Tabor, E.; Sadovska, G.; Pilar, R. Effect of the Nuclearity and Coordination of Cu and Fe Sites in β Zeolites on the Oxidation of Hydrocarbons. ACS Catal. 2020, 10, 3984–4002. [Google Scholar] [CrossRef]
  16. Tang, X.; Wang, L.; Yang, B.; Fei, C.; Yao, T.; Liu, W.; Lou, Y.; Dai, Q.; Cai, Y.; Cao, X.-M.; et al. Direct oxidation of methane to oxygenates on supported single Cu atom catalyst. Appl. Catal. B Environ. 2021, 285, 119827. [Google Scholar] [CrossRef]
  17. Zhang, R.; Liu, N.; Lei, Z.; Chen, B. Selective Transformation of Various Nitrogen-Containing Exhaust Gases toward N2 over Zeolite Catalysts. Chem. Rev. 2016, 116, 3658–3721. [Google Scholar] [CrossRef] [PubMed]
  18. 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]
  19. Zhao, G.; Benhelal, E.; Adesina, A.A.; Kennedy, E.M.; Stockenhuber, M. Comparison of Direct, Selective Oxidation of Methane by N2O over Fe-ZSM-5, Fe-Beta, and Fe-FER Catalysts. J. Phys. Chem. C 2019, 123, 27436–27447. [Google Scholar] [CrossRef]
  20. Zhao, G.; Adesina, A.A.; Kennedy, E.M.; Stockenhuber, M. Formation of Surface Oxygen Species and the Conversion of Methane to Value-Added Products with N2O as Oxidant over Fe-Ferrierite Catalysts. ACS Catal. 2020, 10, 1406–1416. [Google Scholar] [CrossRef]
  21. Parfenov, M.V.; Starokon, E.V.; Pirutko, L.V.; Panov, G.I. Quasicatalytic and catalytic oxidation of methane to methanol by nitrous oxide over FeZSM-5 zeolite. J. Catal. 2014, 318, 14–21. [Google Scholar] [CrossRef]
  22. Fan, L.; Cheng, D.-G.; Song, L.; Chen, F.; Zhan, X. Direct conversion of CH4 to oxy-organics by N2O using freeze-drying FeZSM-5. Chem. Eng. J. 2019, 369, 522–528. [Google Scholar] [CrossRef]
  23. Ipek, B.; Lobo, R.F. Catalytic conversion of methane to methanol on Cu-SSZ-13 using N2O as oxidant. Chem. Commun. 2016, 52, 13401–13404. [Google Scholar] [CrossRef]
  24. 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. Int. Ed. 2021, 60, 16634–16640. [Google Scholar] [CrossRef] [PubMed]
  25. Memioglu, O.; Ipek, B. A potential catalyst for continuous methane partial oxidation to methanol using N2O: Cu-SSZ-39. Chem. Commun. 2020, 57, 1364–1367. [Google Scholar] [CrossRef]
  26. Narsimhan, K.; Iyoki, K.; Dinh, K.; Román-Leshkov, Y. Catalytic Oxidation of Methane into Methanol over Copper-Exchanged Zeolites with Oxygen at Low Temperature. ACS Cent. Sci. 2016, 2, 424–429. [Google Scholar] [CrossRef] [Green Version]
  27. Sushkevich, V.L.; Palagin, D.; Ranocchiari, M.; Van Bokhoven, J.A. Selective anaerobic oxidation of methane enables direct synthesis of methanol. Science 2017, 356, 523–527. [Google Scholar] [CrossRef] [PubMed]
  28. Palagin, D.; Sushkevich, V.L.; van Bokhoven, J.A. Water Molecules Facilitate Hydrogen Release in Anaerobic Oxidation of Methane to Methanol over Cu/Mordenite. ACS Catal. 2019, 9, 10365–10374. [Google Scholar] [CrossRef]
  29. Liu, Z.; Huang, E.; Orozco, I.; Liao, W.; Palomino, R.M.; Rui, N.; Duchoň, T.; Nemšák, S.; Grinter, D.C.; Mahapatra, M.; et al. Water-promoted interfacial pathways in methane oxidation to methanol on a CeO 2 -Cu 2 O catalyst. Science 2020, 368, 513–517. [Google Scholar] [CrossRef] [PubMed]
  30. Wu, Y.; Zhang, H.; Yan, Y. High efficiency of phenol oxidation in a structured fixed bed over Cu-ZSM-5/PSSF prepared by ion-exchanged method. Chem. Eng. J. 2020, 380, 122466. [Google Scholar] [CrossRef]
  31. Xia, Y.; Zhan, W.; Guo, Y.; Guo, Y.; Lu, G. Fe-Beta zeolite for selective catalytic reduction of NOx with NH3: Influence of Fe content. Chin. J. Catal. 2016, 37, 2069–2078. [Google Scholar] [CrossRef]
  32. Xiao, P.; Osuga, R.; Wang, Y.; Kondo, J.N.; Yokoi, T. Bimetallic Fe–Cu/beta zeolite catalysts for direct hydroxylation of benzene to phenol: Effect of the sequence of ion exchange for Fe and Cu cations. Catal. Sci. Technol. 2020, 10, 6977–6986. [Google Scholar] [CrossRef]
  33. Gao, F.; Kollar, M.; Kukkadapu, R.K.; Washton, N.M.; Wang, Y.; Szanyi, J.; Peden, C.H. Fe/SSZ-13 as an NH3-SCR catalyst: A reaction kinetics and FTIR/Mössbauer spectroscopic study. Appl. Catal. B Environ. 2015, 164, 407–419. [Google Scholar] [CrossRef] [Green Version]
  34. Berlier, G.; Lamberti, C.; Rivallan, M.; Mul, G. Characterization of Fe sites in Fe-zeolites by FTIR spectroscopy of adsorbed NO: Are the spectra obtained in static vacuum and dynamic flow set-ups comparable? Phys. Chem. Chem. Phys. 2010, 12, 358–364. [Google Scholar] [CrossRef]
  35. He, M.; Zhang, J.; Sun, X.-L.; Chen, B.-H.; Wang, Y.-G. Theoretical Study on Methane Oxidation Catalyzed by Fe/ZSM-5: The Significant Role of Water on Binuclear Iron Active Sites. J. Phys. Chem. C 2016, 120, 27422–27429. [Google Scholar] [CrossRef]
  36. Mahyuddin, M.H.; Staykov, A.; Shiota, Y.; Yoshizawa, K. Direct Conversion of Methane to Methanol by Metal-Exchanged ZSM-5 Zeolite (Metal = Fe, Co, Ni, Cu). ACS Catal. 2016, 6, 8321–8331. [Google Scholar] [CrossRef]
  37. Available online: http://www.iza-structure.org/databases/ (accessed on 5 November 2021).
  38. Kresse, G.; Furthmüller, 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]
  39. Hutter, J.; Iannuzzi, M.; Schiffmann, F.; VandeVondele, J. cp2k: Atomistic simulations of condensed matter systems. WIREs Comput. Mol. Sci. 2014, 4, 15–25. [Google Scholar] [CrossRef] [Green Version]
Figure 1. (a) H2-TPR profiles of the investigated FenCum-BEA samples; (b) Fe, Cu species quantification based on the H2 consumption of H2-TPR; XPS of Cu 2p; (c) and Fe 2p; (d) of the FenCum-BEA samples.
Figure 1. (a) H2-TPR profiles of the investigated FenCum-BEA samples; (b) Fe, Cu species quantification based on the H2 consumption of H2-TPR; XPS of Cu 2p; (c) and Fe 2p; (d) of the FenCum-BEA samples.
Catalysts 11 01444 g001
Figure 2. Activity measurement results of H2O-mediated continuous N2O DMTM over FenCum-BEA and Cu1.28%- and Fe1.28%-BEA; as noted, for the purpose of better description, the activity measurement results for the sample of Fe0.6%Cu0.68%-BEA evaluated in both scenarios of presence (aFe0.6%Cu0.68%) and absence of H2O (bFe0.6%Cu0.68%) are depicted in Figure 2: (a) CH3OH productivity derived as an average value after a 5 h reaction; (b) product selectivity; the selectivities for the Fe0.6%Cu0.68%-BEA samples were derived from 25 h of activity measurement; (c) reactant conversion of CH4 and N2O; (d) 25 h long-term activity test; reaction conditions: N2O:CH4:H2O:He = 30:15:10:45, GHSV = 12,000 h−1, T = 270 °C.
Figure 2. Activity measurement results of H2O-mediated continuous N2O DMTM over FenCum-BEA and Cu1.28%- and Fe1.28%-BEA; as noted, for the purpose of better description, the activity measurement results for the sample of Fe0.6%Cu0.68%-BEA evaluated in both scenarios of presence (aFe0.6%Cu0.68%) and absence of H2O (bFe0.6%Cu0.68%) are depicted in Figure 2: (a) CH3OH productivity derived as an average value after a 5 h reaction; (b) product selectivity; the selectivities for the Fe0.6%Cu0.68%-BEA samples were derived from 25 h of activity measurement; (c) reactant conversion of CH4 and N2O; (d) 25 h long-term activity test; reaction conditions: N2O:CH4:H2O:He = 30:15:10:45, GHSV = 12,000 h−1, T = 270 °C.
Catalysts 11 01444 g002
Figure 3. (a) In situ FTIR with NO as the probe molecule over Fe0.6%Cu0.68%-BEA after the in situ pretreatment by He (>99.999 vol%, 40 mL∙min−1 at T = 500 °C for 1 h) and N2O (30 vol% in He, 40 mL∙min−1 at T = 250 °C for 1 h); as noted, during the N2O pretreatment, the sample was also initially pretreated with He at 500 °C for 1 h and subsequently pretreated with N2O; (b) NO in situ FTIR of the FenCum-BEA samples after the N2O pretreatment; (c) peak area comparisons of 1873 and 1864 cm−1 in panel (b) for the FenCum-BEA samples; (d) H2-TPR of Fe0.6%Cu0.68%-BEA after the O2 and N2O pretreatment; this specific pretreatment strategy is similar to that utilized during NO in situ FTIR.
Figure 3. (a) In situ FTIR with NO as the probe molecule over Fe0.6%Cu0.68%-BEA after the in situ pretreatment by He (>99.999 vol%, 40 mL∙min−1 at T = 500 °C for 1 h) and N2O (30 vol% in He, 40 mL∙min−1 at T = 250 °C for 1 h); as noted, during the N2O pretreatment, the sample was also initially pretreated with He at 500 °C for 1 h and subsequently pretreated with N2O; (b) NO in situ FTIR of the FenCum-BEA samples after the N2O pretreatment; (c) peak area comparisons of 1873 and 1864 cm−1 in panel (b) for the FenCum-BEA samples; (d) H2-TPR of Fe0.6%Cu0.68%-BEA after the O2 and N2O pretreatment; this specific pretreatment strategy is similar to that utilized during NO in situ FTIR.
Catalysts 11 01444 g003
Figure 4. (a) Optimized structure models with diverse active-site motifs on the BEA zeolite; Z represents the BEA zeolite framework; the Al atom was located at the T5 site for the monomeric active site [18]; the T2 and T5 sites were chosen as the neighboring framework Al site based on previous studies [24]; (b) Gibbs-free energy of formation (ΔG) as a function of ΔμN2O and ΔμH2O (chemical potentials) for different active motifs at T = 270 °C; (c) ΔG as a function of temperature under N2O DMTM reaction condition of 10 vol% H2O, 30 vol% N2O and 1 atm pressure.
Figure 4. (a) Optimized structure models with diverse active-site motifs on the BEA zeolite; Z represents the BEA zeolite framework; the Al atom was located at the T5 site for the monomeric active site [18]; the T2 and T5 sites were chosen as the neighboring framework Al site based on previous studies [24]; (b) Gibbs-free energy of formation (ΔG) as a function of ΔμN2O and ΔμH2O (chemical potentials) for different active motifs at T = 270 °C; (c) ΔG as a function of temperature under N2O DMTM reaction condition of 10 vol% H2O, 30 vol% N2O and 1 atm pressure.
Catalysts 11 01444 g004
Figure 5. In situ FTIR spectra derived under different conditions over the Fe0.60%Cu0.68%-BEA: (a,b) co-feeding experiment of N2O (2 vol%, in He), CH4 (2 vol%, in He) and H2O (bubbled by the mixed flow of 40 mL∙min−1) at a time interval of 0.5 min (T = 270 °C), wherein the sample was pretreated in situ for 1 h at T = 500 °C under vacuum conditions; (c,d) after the vacuum (T = 500 °C, 1 h) and subsequent N2O (2 vol% in He, T = 270 °C of 1 h) pretreatment, the IR signal begins to be monitored along with the interaction with CH4 (2 vol%, in He).
Figure 5. In situ FTIR spectra derived under different conditions over the Fe0.60%Cu0.68%-BEA: (a,b) co-feeding experiment of N2O (2 vol%, in He), CH4 (2 vol%, in He) and H2O (bubbled by the mixed flow of 40 mL∙min−1) at a time interval of 0.5 min (T = 270 °C), wherein the sample was pretreated in situ for 1 h at T = 500 °C under vacuum conditions; (c,d) after the vacuum (T = 500 °C, 1 h) and subsequent N2O (2 vol% in He, T = 270 °C of 1 h) pretreatment, the IR signal begins to be monitored along with the interaction with CH4 (2 vol%, in He).
Catalysts 11 01444 g005
Figure 6. D2O isotopic tracer study based on temperature-programmed surface reaction (TPSR) over the Fe0.60%Cu0.68%-BEA. (a) D2O-mediated N2O DMTM and (b) N2O DMTM in the absence of D2O; as noted, the signal contributions of CH3OD (33), D3O+ (22), HOD (19) and D2O (22) from panel (b) (N2O DMTM without D2O) have been subtracted from panel (a); (c) peak area comparisons of CH3OH (31) and CH3OD (33) derived in panels (a,b).
Figure 6. D2O isotopic tracer study based on temperature-programmed surface reaction (TPSR) over the Fe0.60%Cu0.68%-BEA. (a) D2O-mediated N2O DMTM and (b) N2O DMTM in the absence of D2O; as noted, the signal contributions of CH3OD (33), D3O+ (22), HOD (19) and D2O (22) from panel (b) (N2O DMTM without D2O) have been subtracted from panel (a); (c) peak area comparisons of CH3OH (31) and CH3OD (33) derived in panels (a,b).
Catalysts 11 01444 g006
Figure 7. H2O-mediated N2O DMTM mechanism simulation results of the H2O proton-transfer route over Fe0.6%Cu0.68%-BEA with the neighboring bimetallic cation active site of [Fe]+--[Cu]+, wherein the initial N2O dissociation and CH4 activation steps of (I→IV) were simulated by DFT, while the H2O proton-transfer route (V→VI) was simulated by AIMD; (a) derived free energy diagram along with the reaction coordinate (with the unadsorbed reactant and desorbed product being taken into account to obtain the net change of the reaction energy); metadynamics-computed free energy surface in a (b) 2D and (c) 3D view based on the AIMD, wherein the minimal free energy route is shown by gray balls in panel (b).
Figure 7. H2O-mediated N2O DMTM mechanism simulation results of the H2O proton-transfer route over Fe0.6%Cu0.68%-BEA with the neighboring bimetallic cation active site of [Fe]+--[Cu]+, wherein the initial N2O dissociation and CH4 activation steps of (I→IV) were simulated by DFT, while the H2O proton-transfer route (V→VI) was simulated by AIMD; (a) derived free energy diagram along with the reaction coordinate (with the unadsorbed reactant and desorbed product being taken into account to obtain the net change of the reaction energy); metadynamics-computed free energy surface in a (b) 2D and (c) 3D view based on the AIMD, wherein the minimal free energy route is shown by gray balls in panel (b).
Catalysts 11 01444 g007
Scheme 1. Schematic diagram of proposed N2O DMTM reaction mechanism over the best-performing sample of Fe0.6%Cu0.68%-BEA in the presence and absence of H2O; specifically, after the initial N2O dissociation and CH4 activation steps (black line), three types of routes were proposed, namely: (i) in the absence of H2O (blue line); (ii) the H2O proton-transfer route (red line); and (iii) the H2O direct-reaction route. The chemical structure was labeled according to constructed models depicted in the energy diagrams of Figure 7 and Figure S8a,b.
Scheme 1. Schematic diagram of proposed N2O DMTM reaction mechanism over the best-performing sample of Fe0.6%Cu0.68%-BEA in the presence and absence of H2O; specifically, after the initial N2O dissociation and CH4 activation steps (black line), three types of routes were proposed, namely: (i) in the absence of H2O (blue line); (ii) the H2O proton-transfer route (red line); and (iii) the H2O direct-reaction route. The chemical structure was labeled according to constructed models depicted in the energy diagrams of Figure 7 and Figure S8a,b.
Catalysts 11 01444 sch001
Figure 8. (a,b) Electronic density difference upon adsorption of N2O onto FeCu-BEA and Cu-BEA, respectively, with neighboring [Fe]--[Cu] and [Cu]--[Cu] as the active site; the yellow and blue colors represent electron density increases and decreases, respectively; the metal cations are labeled by the subscript characters a and b to mark differences; (c) comparisons of transferred charges after adsorption of N2O for the neighboring metal cations based on the Bader charge analysis; (d) total DOS comparisons for the models depicted in panels (a,b); Si (yellow), O (red), N (blue), Al (pink), Cu (orange) and Fe (purple).
Figure 8. (a,b) Electronic density difference upon adsorption of N2O onto FeCu-BEA and Cu-BEA, respectively, with neighboring [Fe]--[Cu] and [Cu]--[Cu] as the active site; the yellow and blue colors represent electron density increases and decreases, respectively; the metal cations are labeled by the subscript characters a and b to mark differences; (c) comparisons of transferred charges after adsorption of N2O for the neighboring metal cations based on the Bader charge analysis; (d) total DOS comparisons for the models depicted in panels (a,b); Si (yellow), O (red), N (blue), Al (pink), Cu (orange) and Fe (purple).
Catalysts 11 01444 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

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. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11121444

AMA Style

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(12):1444. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11121444

Chicago/Turabian Style

Li, Yan, Ning Liu, Chengna Dai, Ruinian Xu, Gangqiang Yu, Ning Wang, Jie Zhang, and Biaohua Chen. 2021. "Synergistic Effect of Neighboring Fe and Cu Cation Sites Boosts FenCum-BEA Activity for the Continuous Direct Oxidation of Methane to Methanol" Catalysts 11, no. 12: 1444. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11121444

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