Daniyal
Kiani
,
Sagar
Sourav
,
Israel E.
Wachs
* and
Jonas
Baltrusaitis
*
Department of Chemical and Biomolecular Engineering, Lehigh University, B336 Iacocca Hall, 111 Research Drive, Bethlehem, PA 18015, USA. E-mail: job314@lehigh.edu; iew0@lehigh.edu
First published on 5th October 2021
The experimentally validated computational models developed herein, for the first time, show that Mn-promotion does not enhance the activity of the surface Na2WO4 catalytic active sites for CH4 heterolytic dissociation during OCM. Contrary to previous understanding, it is demonstrated that Mn-promotion poisons the surface WO4 catalytic active sites resulting in surface WO5 sites with retarded kinetics for C–H scission. On the other hand, dimeric Mn2O5 surface sites, identified and studied via ab initio molecular dynamics and thermodynamics, were found to be more efficient in activating CH4 than the poisoned surface WO5 sites or the original WO4 sites. However, the surface reaction intermediates formed from CH4 activation over the Mn2O5 surface sites are more stable than those formed over the Na2WO4 surface sites. The higher stability of the surface intermediates makes their desorption unfavorable, increasing the likelihood of over-oxidation to COx, in agreement with the experimental findings in the literature on Mn-promoted catalysts. Consequently, the Mn-promoter does not appear to have an essential positive role in synergistically tuning the structure of the Na2WO4 surface sites towards CH4 activation but can yield MnOx surface sites that activate CH4 faster than Na2WO4 surface sites, but unselectively.
Dispersed phase catalysts offer key advantages over catalysts containing a mixture of dispersed and crystalline phases,16,20–23 since dispersed phase catalysts consist of atomically dispersed sites with well-defined molecular and electronic structure that can be tuned via the addition of promoters. Such control of the nature and the structure of catalytic active sites remains challenging in the presence of a crystalline phase in supported mixed phase catalysts, which exhibit an ensemble distribution of the catalytic active sites distributed in the various phases, hence complicating the structure–activity analysis. This difference in dispersed vs. mixed phase catalysts is illustrated in Fig. 1 using silica (SiO2) supported metal oxide (MOx) catalysts, such as those used for OCM, with well-defined active sites found in dispersed phase catalysts allowing systematic analysis for structure–activity relationships, and supported mixed oxide phase catalysts resulting in an variety of various types of sites.
The structure–activity relationships for the state-of-the-art supported Mn–Na2WO4/SiO2 tri-metal oxide OCM catalyst have remained elusive not only due to the demanding reaction conditions, but also due to the mixed-oxide phase nature of this catalyst.24 While dispersed phase tri-metal oxide catalysts, i.e., containing all three metal oxides (W, Na, Mn), have not been studied to date, the structure–activity relationships of active sites in dispersed phase bi-metal oxide catalysts (Na–WOx/SiO2, Mn–WOx/SiO2) for OCM have been recently reported.22,25 It was shown that the distorted surface WO4 sites are present at elevated temperatures (673 K) on the amorphous SiO2 surface in unpromoted, supported WOx/SiO2 catalysts.22 Upon Na-promotion, the extent of distortion of the WO4 surface site diminishes.22,25 The dispersed-phase Na-coordinated WO4 surface sites (Na–WO4) with Na/W molar ratio less than 2 were shown to be significantly more selective to C2 products than the unpromoted, surface WO4 sites and the corresponding mixed-oxide phase catalysts containing crystalline phase Na2WO4 (Na/W ∼ 2).22,25 Moreover, the promotion with Mn i.e Mn–WOx/SiO2, did not have a significant impact on the molecular structure and OCM C2 selectivity of the WO4 surface sites in comparison to unpromoted WO4 surface sites.25 Thus, the role of Mn for the OCM catalytic reaction remains unclear.
Older studies of mixed-phase tri-metal oxide catalysts under/near OCM reaction conditions revealed that Mn3+ reduces to Mn2+ during CH4 activation, which can then be re-oxidized to Mn3+ during the second half of the OCM reaction redox cycle.26 Based on the proposed Mn3+ → Mn2+ → Mn3+ redox cycle, multiple reaction models suggesting Mn sites as the catalytic active site during OCM have been put forth in the literature.26–29 Note that the observed redox cycle, however, was purely based on transitions observed in the crystalline phases of manganese oxide (Mn2O3 and Mn3O4) and the redox behavior of the dispersed phase manganese oxide surface sites (MnOx) in the catalyst are still not known. More recently, an operando synchrotron μ-XRF/XRD/absorption-computed tomography investigation of the mixed-phase tri-metal oxide catalyst found that during phase transformations occurring in the crystalline manganese oxide phases (Mn2O3, Mn7SiO12 and MnWO4), product distribution did not change. Hence, it was concluded that such manganese oxide phases did not have a critical role in the catalytic OCM reaction.30 It was further concluded that dispersed phase MnOx surface sites present on the SiO2 support in the mixed-phase tri-metal oxide catalyst, which bulk X-ray based techniques such as XRD, XAS, XRF cannot effectively monitor, may be involved in the OCM catalytic cycle.30
Herein, we provide the first in silico framework validated with in situ spectroscopy measurements to study the structure–activity relationship of metal oxide sites in model tri-metal oxide OCM catalysts containing Mn, W, Na metal oxides on a SiO2 support. In particular, the nature and structure of the tri-metal oxide surface sites (Mn–Na2WO4/SiO2) was elucidated and contrasted with the bi-metal oxide surface sites (Na2WO4) via ab initio molecular dynamics (AIMD) and periodic density functional theory (DFT). Na2WO4 surface sites are used as the base for this comparison as these surface sites were shown to be the selective active sites for CH4 activation during OCM.19,22,25 The DFT predicted molecular and electronic structures of the dispersed phase surface sites in bi- and tri-metal oxide OCM catalyst prior to CH4 activation were also validated by in situ Raman and in situ UV-Vis diffused reflectance (UV-VisDR) spectroscopies at dehydrated oxidative conditions (400 °C, 10% O2/inert), respectively. Intrinsic kinetics of CH4 dissociation over surface sites in dispersed phase bi-metal oxide and tri-metal oxide case were calculated and contrasted. Results herein provide molecular-level insights on the nature of surface sites, and their activity towards heterolytic CH4 dissociation during OCM, providing structure–function relationships for this OCM catalyst.
Vibrational modes have been calculated for the selected surface species within the harmonic approximation. Only the tungsten active center and its 1st and 2nd neighbors (O–Si and OH groups) were considered in the Hessian matrix. This matrix was computed by the finite difference method followed by a diagonalization procedure. The eigenvalues of the resulting matrix led to frequency values. The assignment of the vibrational modes was done by inspection of the corresponding eigenvectors. Static (NSW = 0, IBRION = −1) self-consistent calculations were performed to compute DOS (ISMEAR = 0, Sigma = 0.02), and the resulting vasprun.xml files were visualized using P4Vasp33,34 to extract local and total density of states (LDOS, total DOS).
Ab initio molecular dynamics (AIMD) calculations were performed using Anderson thermostat (1700 K, 2 fs steps) to provide sufficient kinetic energy to the simulated catalyst surface for the atoms to rearrange so the global minima for the total energy can be calculated. The antiferromagnetic configuration in dimeric Mn2Ox cases was consistently lower in energy than ferromagnetic configurations, so all dimeric Mn2Ox studied here were treated as antiferromagnetic. Typically, ∼8000 fs AIMD simulations resulted in 3–4 lowest energy structures, which were then extracted and further optimized by periodic DFT before subsequent calculations were conducted for frequencies, Bader35 charges and transition states (from nudged elastic band (NEB)36,37) on the lowest energy structure amongst the 3–4 selected from AIMD. Ab initio thermodynamics (AITD) were calculated according to the framework described in the recent literature.13,38–40 Specifically, the relative stability of various possible surface MnaOb clusters were computed in reference to the bulk α-Mn2O3 phase (unit cell: 32Mn atoms, 48O atoms),41 which is typically reported to be stable under OCM relevant conditions (T = 1000 K, PO2 = 0.33–0.1).42 The Mn2O3 atom positions and cell parameters were relaxed with high precision and cutoff energy of 520 eV to minimize any Pulay stress. Given the unit cell of α-Mn2O3 containing 32Mn and 48O atoms, the general governing stoichiometric equation can be written as (1)
(1) |
(2) |
In this method, the PV contributions of solids were neglected,13,38–40 and the Gibbs free energies of solids were approximated as their respective electronic energies computed by DFT with thermal corrections in accordance with the methodology described in the literature,43 as shown in eqn (3)
μ(solid) = EDFT0K(solid) + Gvib(T). | (3) |
The chemical potential of the gas phase O2 depends on the temperature (T) and the corresponding partial pressure (P). At arbitrary T and P, μO2 can then be written as:
(4) |
For example, using the general eqn (1), the Mn1O1–Na2WO4/SiO2 system was calculated as:
(5) |
(6) |
Using eqn (1), the comprehensive MnaOb chemical potential landscape was computed for various mono-atomic (Mn, MnO, MnO2, MnO3) and di-atomic (Mn2O, Mn2O2, Mn2O3, Mn2O4, Mn2O5 compositions to compare their respective thermodynamic stability given OCM relevant conditions of 1000 K and PO2/P° of 0.1–0.33. A higher degree of oligomerization for surface MnaOb compositions was not considered due to the complexity of the possible structures in tri- and tetra-atomic cases such as straight 2D-chain vs. 3D-clusters.
Lastly, intrinsic rate constant of C–H dissociation in CH4 was calculated using the DFT-computed E0 values of transition states identified via NEB method, according to equation:
(7) |
Fig. 2 Individual heat maps of computed Gibbs free energy of formation for (MnaOb–Na2WO4)/SiO2 as a function of T (300–1300 K) and PO2/P = 0–1 (ref. 10) for (a) Mn, (b) MnO, (c) MnO2, (d) MnO3, (e) Mn2O, (f) Mn2O2, (g) Mn2O3, (h) Mn2O4, (i) Mn2O5. The dashed rectangle in (i) signifies the OCM relevant conditions. |
In a comparison shown in Fig. 3, the entire catalyst structure space of MnaOb–Na2WO4/SiO2 was compared to ascertain the most thermodynamically stable stoichiometry of surface MnaOb sites in MnaOb–Na2WO4/SiO2 catalysts under OCM relevant PO2/P° of 0.2–0.33, and temperature of 1000 K. Under these conditions, the formation of Mn2O5 dimeric sites is evidently thermodynamically favorable (i.e. most exergonic) with reference to bulk Mn2O3, especially towards PO2/P° values ∼0.30–0.33 that represent the OCM reaction stoichiometry i.e. CH4:O2 = 2:1. For instance, as the PO2/P° values increase from 0.25 to 0.35, the ΔGform of Mn2O5 sites decreases from −0.002 eV (−0.20 kJ mol−1) to −0.031 eV (−3.00 kJ mol−1). Note that since the Gibbs energy difference is so small between MnO2 and Mn2O5, a minor presence of MnO2 sites cannot be precluded under OCM conditions in the catalyst. However, for simplicity, the Mn2O5–Na2WO4/SiO2 catalyst was chosen for further study presented in the next sections. Specifically, the molecular and electronic structure of the surface sites in Mn2O5–Na2WO4/SiO2 catalyst are calculated and compared against the base case of unpromoted Na2WO4/SiO2 to ascertain the effect of Mn promotion. The base case i.e. Na2WO4/SiO2 contains Na-coordinated WO4 surface sites that have been shown to be the selective active sites for OCM in such bi-metal oxide catalysts.22,25 Moreover, the molecular and electronic structures of DFT-optimized structures are also validated against experimental in situ characterization data to ensure agreement. Lastly, CH4 activation pathways and intrinsic kinetics are also compared in each case to elucidate the role of surface Mn sites in OCM.
Fundamental frequencies for each case were computed and contrasted with experimental in situ Raman spectra of representative dehydrated catalysts to validate the computational model against experimentally observed structures, as shown in Fig. 4b and c. In each case, the computed fundamental modes are assigned in the color-coded tables in Fig. 4b and c and the solid bars are the DFT-calculated frequencies. In both cases, general agreement is observed between computed frequencies and experimentally measured Raman bands, validating the DFT-optimized structural models. From this comparison, it is inferred that WO bonds are the shortest, given they vibrate in the 945–1000 cm−1 range. Likewise, support-grafting bonds like W–O–Si and Mn–O–Si are significantly longer, given they vibrate in the 800–900 cm−1 range. Lastly, bridging bonds like Mn–O–Mn and Mn–O–W are the longest (weakest), since they vibrate at 616 and 761 cm−1, respectively. It should be noted that we infer a small presence of poorly crystalline nanoclusters/nanoparticles of MnWO4 and Mn-contaminated WO3 in the experimental Raman spectrum of the tri-metal oxide catalyst shown in Fig. 4c, which exhibit Raman bands in the same region as dispersed phase sites. However, since operando/in situ characterization studies in the literature suggest that crystalline MnWO4 is not critical to OCM,29,30 and because the nature (size, shape, surface structure, defect density) of these nanoparticles is not known presently, the nanoparticles have been excluded from this computational model and only the dispersed phase surface sites were studied for structure–function relationships towards CH4 activation. Interested readers are directed to a recent, fully experimental study, where further detailed experimental characterization including UV-VisDRS, Raman at 400 °C, Raman during OCM at 900 °C, TPSR, etc. of several model OCM catalysts including the two catalysts compared in Fig. 4b and c can be found.48
Fig. 5 Structural visualization of (a) Na2WO4/SiO2 and (b) Mn2O5–Na2WO4/SiO2 with calculated oxidation states mentioned next to the transition metal atoms in corresponding colors (Mn: red, W: pink). Oxidation states were calculated using a calibrated-fit for Bader charges, shown in Fig. S2 in the ESI section.† Calculated Bader charges for the surface sites in each case are summarized in table (c). (d) In situ dehydrated UV-Vis DRS of Na–WOx/SiO2(red) and Mn–Na–WOx/SiO2(green) catalysts, with Eg values and LMCT positions marked. |
The Eg value after Mn-promotion of Na–WOx sites is lower than the Eg value in the Na–WOx case (Fig. 5d), which can be further understood via density of states (DOS) analysis of DFT-optimized structures. Total DOS plots are shown in Fig. 6a and b for Na2WO4/SiO2 and Mn2O5–Na2WO4/SiO2, respectively. For the Na2WO4/SiO2 case, the valence band is primarily composed of occupied O 2p and W 5d states, while the conductance band of empty W 5d states. The energy gap between the top of the valence edge and the bottom of the conduction edge is ∼1.7 eV. However, in the Mn2O5–Na2WO4/SiO2 case, the valence and conductance bands are primarily composed of Mn 3d states, with some contribution from occupied O 2p states in the valence band. In this case, the lowest part of the conductance band lowers due to low-lying unoccupied Mn 3d states, reducing the bandgap to ∼0 eV. The DOS analysis shows that Mn-addition fundamentally changes the electronic structure of the surface sites by introducing low-lying vacant Mn 3d states above the Fermi level, which will have strong implications for the catalytic CH4 activation. Elsewhere, it has been found that low-lying empty d states in a transition metal like Pt can act an acceptor for σ-donation from the C–H bond of CH4 and lead to facile C–H scission.52 Thus, it suffices to conclude at this point that experimental in situ UV-Vis DRS and computational DOS analysis are in qualitative agreement evidencing that Mn-promotion of Na2WO4 surface sites will lower the Eg values due to the introduction of low-lying empty Mn 3d states above the Fermi level, that Mn is present in its +3 oxidation state, and that W is present in its fully oxidized +6 state.
CH4 + MOx → M–O–CH3 (methoxy) + M–OH (hydroxy) | (8) |
CH4 + MOx → M–O–CH3 + M–H (hydride) | (9) |
CH4 + MOx → M–CH3 (methyl) + M–OH | (10) |
Fig. 7 (a) Reaction coordinate vs. ΔEDFT (eV) and (b) calculated intrinsic rate constants (log-scale) for CH4 dissociation for the three possible pathways over the Na2WO4/SiO2 catalyst. Arrow indicates the dominant (fastest) pathway. (c) Structural depictions of the initial, TS and final states during CH4 dissociation over Na2WO4/SiO2 catalysts yielding methyl-hydroxy (CH3, OH) intermediates. Pictorial summary for other pathways can be found in Table S3.† |
CH4 dissociation over Na2WO4 surface sites was studied as a function of the possible pathways listed above, and as a function of various active oxygen atoms. The results are summarized in Fig. 7a–c, highlighting the activation barriers associated with various possible pathways over the Na2WO4/SiO2 catalyst. According to the computed energy barriers summarized in Fig. 7a, the highest energy barrier was for the pathway yielding OCH3, OH surface intermediates (3.4 eV), indicating that CH4 dissociation turnover frequency for this pathway will be lower than other pathways with smaller transition state barriers. The two pathways with OCH3, H intermediates formation differ in terms of the oxygen atom of the WO4 where C–H scission occurs, and consequently have slightly varying TS barriers (1.6, 1.9 eV). Lastly, the pathway yielding CH3, OH surface intermediates exhibited the lowest energy barrier of 1.1 eV, suggesting that this is the dominant reaction pathway for CH4 dissociation over Na-coordinated WO4 surface sites in the Na2WO4/SiO2 catalysts. The resulting intrinsic rate constants, calculated according to eqn (7) are plotted in Fig. 7b, suggesting that the intrinsic rate constant for the CH3, OH pathway is 2–3 orders of magnitude higher than other pathways making this the dominant pathway. Lastly, the initial, transition state (TS) and final structures for CH4 activation over Na2WO4 surface sites via the dominant reaction pathway are shown in Fig. 7c, showing the formation of W–CH3 and W–O–H intermediates. Importantly, these results show that Na2WO4 surface sites can effectively dissociate CH4 in the absence of Mn-promoter in agreement with a recently published report,22 as opposed to Lunsford19 reaction model where Na–O–Mn sites are assumed to be the critical active sites for CH4 dissociation.
Local DOS (LDOS) analysis was conducted to elucidate the changes in electronic structure of initial versus the TS structure during CH4 dissociation over the Na2WO4 surface sites. LDOS for W 5d states and O 2p states of the initial Na2WO4 structure are shown in Fig. 8a. The valence band (below the Fermi level, EF) is comprised largely of occupied 2p states from the double-bonded oxygen i.e. WO, while the conductance band is comprised largely of unoccupied W 5d states with a minor contribution from unoccupied O 2p states of WO. Upon CH4 dissociation, the lowest part of the conductance band lowers in energy as the unoccupied O 2p states get filled. Moreover, significant overlap of occupied states is observed for W 5d and C 2p states near the top of the valence edge at ca. −1.7 eV, signifying W–C bond formation as in the W–CH3 surface intermediate. Lastly, occupied H 1s states and occupied O 2p states also overlap in the lower part of the valence edge, confirming O–H bond formation i.e. W–O–H intermediate. The Bader charge for W barely decreases; it changes from +2.51 in initial structure to +2.48 in the TS structure, indicating minimal reduction of the W center. The partial reduction of W center instead of a full reduction to +5 or to +4 might be the reason that reduced W centers could not be observed during OCM in SQUID-EPR experimentally.26 Interestingly, a similar CH4 dissociation pathway yielding CH3, OH surface intermediates was also proposed for La2O3-based catalysts, where La–CH3 and La–O–H intermediates formed upon heterolytic dissociation of CH4.54 Likewise, CH4 dissociation over Al2O3 was also shown to proceed via Al–CH3 and Al–O–H formation.55
Next, intrinsic kinetics of CH4 dissociation over Mn2O5–Na2WO4/SiO2 catalyst were studied to ascertain if CH4 still dissociated over the WOx like it did in the case of Na2WO4/SiO2 catalyst. Once again, the color-coded pathways signify the unique surface intermediates produced. In this case however, CH4 dissociation was studied as a function of the moiety since three distinct moieties are present in the surface site i.e. WO5 moiety, W–O–Mn moiety, and Mn2O5 moiety. The motivation behind studying CH4 dissociation as a function of moieties is and to ascertain the affect of Mn2O5 promotion on the reaction pathway and intrinsic kinetics. If, CH4 dissociates preferentially over the Mn2O5 moiety, it would indicate that the effect of Mn-promotion in OCM catalysts is not to tune the WOx surface site but in fact to generate a second active site altogether. On the other hand, if it still activates over WO5 moiety, the role of Mn2O5 would be inferred as a structural/chemical promoter.
As shown in Fig. 9a, CH4 dissociation becomes energetically unfavorable over the WO5 moiety i.e. 2–4 eV TS barriers, in stark contrast to facile CH4 dissociation observed over WO4 site in Na2WO4/SiO2 without Mn2O5 promotion (1.1 eV). The higher TS barriers in the case of WO5 moiety in Mn2O5–Na2WO4/SiO2 catalyst in comparison to the pseudo-Td WO4 moiety in Na2WO4/SiO2 catalyst are likely due to the loss of Td geometry of the WOx center upon Mn2O5 promotion, where the Td WO4 serves as a reaction site. The Td geometry of the WO4 is often regarded as a critical requirement of the selective active site in OCM in various experimental studies,19,22,25,56–60 and that notion finds support from the DFT results herein.
Fig. 9 Reaction coordinate vs. ΔEDFT (eV) over (a) WO5 moiety, (b) W–O–Mn moiety, (c) Mn2O5 moiety in Mn2O5–Na2WO4/SiO2 catalyst. (d) Calculated intrinsic rate constants (log-scale) for CH4 dissociation for the possible pathways in Mn2O5–Na2WO4/SiO2 catalyst. Arrow indicates the dominant (lowest barrier) pathway. All numbered pathways 1–9 are described in ESI Table S4.† (e) Visualization of the dominant (lowest barrier) pathway for CH4 dissociation over the Mn2O5–Na2WO4/SiO2 catalyst, yielding methoxy–hydroxy (OCH3, OH) surface intermediates. |
On the other hand, CH4 can also dissociate over the W–O–Mn and Mn2O5 moieties via various pathways with TS barriers in the 1–5 eV range, shown in Fig. 9b and c. Overall, however, the lowest barrier pathway was the one yielding methoxy–hydroxy (OCH3, OH) intermediates, over the Mn2O5 moiety, with a TS barrier of 0.72 eV, which is 0.3 eV lower than the lowest TS barrier pathway for CH4 activation over Na2WO4/SiO2 in the absence of Mn-promoter. Fig. 9d compares the calculated intrinsic rate constants for CH4 dissociation across all moieties via various pathways studied, showing that the lowest TS barrier pathway indeed exhibits a rate constant orders of magnitude higher than other cases and hence is the dominant pathway. Lastly, initial, TS, and final structures for the dominant reaction pathway in this case are summarized in Fig. 9e, evidencing the involvement of Mn–O and Mn–O–Mn bonds in CH4 dissociation to yield Mn–O–CH3 and (Mn)2–O–H surface intermediates.
To ascertain the electronic structural dynamics of the surface sites during CH4 dissociation, LDOS analysis (Mn 3d, O 2p, C 2p, H 1s states) was conducted for the initial and TS structures of Mn2O5–Na2WO4/SiO2, shown in Fig. 10a and b, respectively. Initially, the valence edge is comprised of a mixture of occupied Mn 3d states and O 2p states from Mn–O–Mn and Mn–O bonds. The conductance edge is primarily made up of unoccupied Mn 3d states with some contribution from empty O 2p states. Upon CH4 dissociation and TS formation, the LDOS changes drastically. The upper part of the valence band lowers in energy, and occupied, spin-up Mn 3d states appear at −2.0 eV, which were not present in the initial structure. Moreover, overlapping O 2p and C 2p states as well as the O 2p and H 1s states confirm formation of Mn–O–CH3 and (Mn)2–O–H intermediates. In this case, the occupied C 2p states overlap with occupied O 2p states of Mn–O towards the bottom of the valence edge (−7.0 eV). Lastly, Mn Bader charge decreases from +1.63 to +1.55 from initial to TS structure, indicating reduction of the Mn centers from ∼+3 to ∼+2 in agreement with experimental literature.26,61
Fig. 11 Comparison of calculated intrinsic rate constants for dominant (lowest barrier) pathways for CH4 dissociation over the Na2WO4/SiO2 (green) and Mn2O5–Na2WO4/SiO2 (red) catalysts. |
The faster intrinsic kinetics of heterolytic CH4 dissociation over the Mn2O5 sites versus the Na2WO4 surface sites are not surprising. In fact, DFT calculations have been used to correlate the C–H bond activation energy to the surface reducibility (oxygen vacancy formation energy, work function) in the literature.63 A linear correlation was found between the C–H activation energy and the oxygen vacancy formation energy of pure/doped metal-oxides, making surface reducibility a descriptor for predicting catalyst activity and selectivity for CH4 dissociation in OCM.63 Given that Mn-oxides are more reducible than W-oxides, the CH4 dissociation over surface Mn2O5 sites is expected to be faster than Na2WO4. Moreover, according to this correlation between C–H activation barrier and reducibility of the surface metal oxide site, one of the roles of Na-promotion is to increase the reducibility of surface WOx sites, as evidenced by lowering of peak temperature values during H2-TPR with Na-addition, reported in a recent study on Na–WOx/SiO2 model catalysts.22
It is known that MnOx-based catalysts form excellent partial and total oxidation catalysts owing to their enhanced reducibility. Examples of Mn-oxide based catalysts used for oxidation reactions include: Mn2O3 for complete oxidation of methane,61 Mn-oxide–CeO2 catalysts for formaldehyde oxidation,66 Spinel CoMn2O4 for toluene oxidation,67 mesoporous Mn-oxide catalysts for water oxidation,68 alkane oxidation over Mn-perovskites,69 CeO2-supported Mn-oxide for propane oxidation,70 In fact, analogous to the CH4 dissociation pathway over Mn2O5 moiety envisioned in the present study, operando FTIR spectroscopy elsewhere71 confirmed the formation of Mn–OH surface species by abstraction of hydrogen atoms by nucleophilic oxygen atoms (Mn–O–Mn) from propane during propane oxidation, providing support for the DFT modelling results and chemical insights presented herein.
Recent state-of-the-art understanding regarding the role of Mn- and Na-promoters in Mn2O3–Na2WO4/SiO2 OCM catalysts puts chemical insights provided in the present work in perspective. Specifically, it was recently shown via experimental CH4 + O2 temperature-programmed-surface-reaction of well-defined single site catalysts that Na2WO4 surface sites could effectively and selectively activate CH4 in absence of any Mn-promoter to form C2 products, making the role of Mn-promoter unclear.22 Further evidence of the dispersed phase Na–WO4 surface sites being the critical OCM active sites was provided in recent experimental studies48,72 where temporal analysis of products (TAP) reactor studies in conjunction with steady-state OCM reaction studies of Na2WO4/SiO2 catalysts (without Mn-oxide) demonstrated that the dispersed surface Na–WO4 sites were responsible for selectively activating CH4 to yield C2 and CO products, while molten Na2WO4 phase was found to be mainly responsible for the over-oxidation of CH4 to CO2, and oxidative dehydrogenation of C2H6 to C2H4.72 Likewise, Mn-containing phases including MnOx surface oligomers, MnWO4 and Mn–WO3 nanoparticles in Mn–Na–WOx/SiO2 tri-metal oxide model OCM catalyst were found to be primarily spectating during OCM under differential reaction conditions and when such Mn phases are present in small populations.48 An older model study had previously shown that Na-coordinated WO4 surface sites were more selective but slightly less active than Mn-coordinated WOx and uncoordinated WOx sites during OCM in a series of model bi-metal oxide OCM catalysts.25 Lastly, via operando synchrotron μ-XRF/XRD/absorption-computed tomography, it was shown that crystalline manganese oxide phases including Mn2O3, Mn7SiO12 and MnWO4 are not required components to yield an active catalyst,6 while no conclusion could be reached regarding the dispersed phase MnaOb surface sites that can be present on the SiO2.30
The current study, utilizing in situ spectroscopy, ab initio molecular dynamics, ab initio thermodynamics, and periodic-DFT modelling bridges the knowledge gap present in the literature and provides the following critical mechanistic insights, which suggest that Mn-promotion in fact is likely not essential for a yielding a selective-active OCM catalyst:
(1) CH4 heterolytically dissociates effectively over the Na2WO4 surface sites in absence of Mn-promoter, with a TS barrier of 1.07 eV forming W–CH3 and W–O–H surface intermediates. The surface intermediate formation is endothermic by 0.60 eV, indicative of their instability and likelihood of facile desorption.
(2) CH4 heterolytically dissociates over the WO5 moiety significantly slower when the Mn2O5–Na2WO4/SiO2 catalyst is studied due to structural changes incurred upon Mn2O5 coordination to the WO4 site forming the WO5. This shows that Mn-promotion can infact poison the otherwise pseudo-Td WO4 surface active sites, retarding the kinetics of CH4 dissociation over them.
(3) CH4 dissociates over the Mn2O5 moiety with a TS barrier of 0.72 eV forming Mn–O–CH3 and (Mn)2–O–H surface intermediates. The surface intermediates formation is appreciably exothermic (−0.48 eV), indicative of the high stability of the formed intermediates.
(4) The significantly higher stability of surface intermediates formed when CH4 dissociates over the Mn2O5 moiety in Mn2O5–Na2WO4/SiO2 catalyst versus the intermediates formed upon CH4 dissociation over the Na2WO4/SiO2 signifies the likelihood of the Mn–O–CH3, (Mn)2–O–H surface intermediates over-oxidizing to COx, in agreement with the higher CH4 conversion but lower C2 selectivity reported in the OCM literature for MnOx/SiO2 catalysts.64 Therefore, the selective-active sites for OCM are the Na-coordinated WO4 surface sites present in Na2WO4/SiO2 catalysts without Mn-promotion. Mn promotion can create a second set of Mn-based surface sites (like the surface Mn2O5 sites) that are more active for CH4 dissociation than the WO4-based sites, but less selective to C2 products due to the unfavorable desorption of the reaction intermediates formed over Mn2O5.
(5) Computational results herein also suggest that CH4 activation via the MvK mechanism where the oxygen species involved in the reaction originate in the solid phase catalytic active site is quite possible. The DFT-calculated intrinsic barriers of the MvK type reaction pathways are in general agreement with the experimental apparent activation energy values reported in the literature, providing qualitative support for the possibility of the MvK mechanism prevailing during catalytic OCM. However, we note that only the heterolytic CH4 activation pathway is studied herein and the homolytic CH4 activation pathway over these surface sites will be addressed in a future study. Literature studies have shown that both heterolytic and homolytic activation of CH4 occurs during OCM, where both pathways yield CH3 radicals in the gas-phase.73,74 Typically, homolytic pathways exhibit higher energy barriers than the heterolytic pathway55 due to the stabilization of the CH3 fragment from surface interactions in the heterolytic pathway.
Future investigations on the role of surface MnOx sites during CH4 activation in well-defined catalysts should include CH4 + O2 temperature-programmed surface reaction, steady state performance tests, isotope-switch experiments in temporal analysis of products (TAP) reactor, and computational analysis of the homolytic pathway for CH4 activation over the catalyst models generated in this study.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/d1sc02174e |
This journal is © The Royal Society of Chemistry 2021 |