Next Article in Journal
Five Bonds to Carbon through Tri-Coordination in Al3C3/0
Next Article in Special Issue
Au Single Metal Atom for Carbon Dioxide Reduction Reaction
Previous Article in Journal
Exploiting Supramolecular Synthons in Cocrystals of Two Racetams with 4-Hydroxybenzoic Acid and 4-Hydroxybenzamide Coformers
Previous Article in Special Issue
Hydrogen Evolution upon Ammonia Borane Solvolysis: Comparison between the Hydrolysis and Methanolysis Reactions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Probing Low-Temperature OCM Performance over a Dual-Domain Catalyst Bed

Faculty of Chemical Engineering, Technion-Israel Institute of Technology, Haifa 3200003, Israel
*
Author to whom correspondence should be addressed.
Chemistry 2023, 5(2), 1101-1112; https://doi.org/10.3390/chemistry5020075
Submission received: 15 March 2023 / Revised: 4 May 2023 / Accepted: 5 May 2023 / Published: 8 May 2023

Abstract

:
The Mn-Na2WO4/SiO2 catalyst is regarded as the most promising catalyst for the oxidative coupling of methane (OCM). Despite its remarkable performance, the Mn-Na2WO4/SiO2 catalyst requires a high reaction temperature (>750 °C) to show significant activity, a temperature regime that simultaneously causes quick deactivation. In the current work, we show that the benefits of this catalyst can be leveraged even at lower reaction temperatures by a using a stacked catalyst bed, which includes also a small amount of 5% La2O3/MgO on-top- of the Mn-Na2WO4/SiO2 catalyst. The simple stacking of the two catalysts provides >7-fold higher activity and ~1.4-fold higher C2 yield at 705 °C compared to Mn-Na2WO4/SiO2 and La2O3/MgO, respectively. We specifically show that the enhanced OCM performance is associated with synergistic interactions between the two catalyst domains and study their origin.

Graphical Abstract

1. Introduction

The oxidative coupling of methane (OCM) reaction was identified by Keller and Bhasin in 1983 as a promising route to convert methane directly into ethane and ethylene [1]. If this process can be accomplished in an effective small-scale and distributed way, it can be used to convert methane from natural gas, shale gas [2], biogas [3], and landfills into important value-added products such as polymers. However, due to the symmetric CH4 and the strong C-H bonds, a high reaction temperature is usually required to promote OCM (>750 °C) [1]. The high reaction temperature also promotes the over-oxidation of methane into COx (CO and CO2) and affects the longevity of the catalyst over prolonged exposure [4]. Despite extensive efforts, a catalyst of industrially applicable yield and high stability has not yet been found.
The Mn-Na2WO4/SiO2 (MnNaWSi) catalyst has been and is still regarded as the most promising catalyst for this reaction, with early reports showing CH4 conversion ~25% with a C2 selectivity of ~80% (>800 °C) over long periods of time (~500 h) [5,6]. Despite the remarkable performance of the MnNaWSi catalyst, more recent reports show that structural changes in the multi-metal oxide material occur under reaction conditions [6,7,8,9,10,11,12,13,14]. Hayek et al. have shown that these structural changes also occur in the presence of additional dopants and result in strong deactivation, which is masked by the use of large amounts of catalyst [15]. Distinctly, it is the melting of crystalline Na2WO4 at 698 °C that is believed to impose instability in the catalyst structure due to its strong interaction with the silica support and the MnOx phase as well as its evaporation (~1000 °C) at hotspots in the catalyst bed [8,16]. An operating temperature below or close to 700 °C should in principle limit those structural changes and help suppress the over-oxidation of the desired C2 products into CO2 [17].
It is generally accepted that the activation of the reaction involves the cleavage of the C-H bond to form methyl radicals (CH3•). The CH3• then couples to form ethane in the gas phase near the catalyst’s surface [4,14,16,18]. It has been shown that the production rate of methyl radicals is strongly correlated to the C2 yield [19,20,21]. Ideally, a good OCM catalyst would be able to generate methyl radicals at low temperatures (below 700 °C) and allow those to desorb and couple in the gas phase [14,20,21]. Hence, it is the ability of the MnNaWSi catalyst to generate and release methyl radicals that enable its remarkably high C2 selectivity [14]. However, the MnNaWSi catalyst is only able to effectively activate methane above 750–800 °C, which makes the produced C2 products highly susceptible to further oxidation to unwanted COx.
Lansford and coworkers showed that lanthanum oxide (La2O3) is active for OCM and is able to generate CH3• even at temperatures as low as ~500 °C [20,22]. However, the C2 selectivity of the La2O3 is relatively low compared to the MnNaWSi catalyst, which limits its applicability on its own. Hence, the combination of La2O3 and the MnNaWSi catalyst should produce a superior low-temperature OCM catalyst. Wu et al. reported using La to dope the MnNaWSi catalyst to achieve enhanced dispersion of the Mn and W on the silica support [23], along with a significant improvement in C2 selectivity and activity in OCM. Whereas Ghose et al. and Ismagilov et al., showed only a marginal enhancement in C2 selectivity, following a similar catalyst doping procedure [24,25]. Neither of them reported low temperatures reaction performance. Nevertheless, at present, there is a consensus in the OCM literature that supports the high potential of La2O3 to reduce OCM reaction temperature [22,26,27,28,29]. Recently, Jaroenpanon et al. showed that Mn-Na2WO4 supported on La2O3, rather than on the commonly used SiO2 support, provided a C2 yield of 5~10% between 500 and 650 °C [30]. Zou et al. reported a ~32% C2 selectivity and a C2 yield of ~10% at 550–700 °C over a catalyst bed composed of a 1:1 (%wt) mechanical mixture of La2O3 and Na2WO4/SiO2 (NaWSi) [31]. However, the economic benefit of using the rare earth La2O3 as a bulk support material for OCM is questionable [9,14,15]. In this study, we evaluate the OCM performance over MnNaWSi mechanically mixed with a small amount (~0.6 wt.%) of La2O3 nanoparticles supported on MgO (La/MgO). The small amount of La is shown to provide significant OCM enhancement with the MgO support mitigating the deactivation by limiting the direct interaction of the Na2WO4 melt with the La2O3. We demonstrate that the La/MgO to MnNaWSi weight ratio in the catalyst bed has a strong effect on catalytic performance. We further compared the effect of mixing the two catalysts vs. stacking the La/MgO catalyst on-top of the MnNaWSi. The effect of the above configuration on catalytic performance and catalyst deactivation at low-temperature reaction is discussed in detail.

2. Experimental Procedures

2.1. Catalyst Preparation

Mn2O3-Na2WO4/SiO2 (MnNaWSi). The Mn2O3-Na2WO4/SiO2 was synthesized following a multi-step incipient wetness impregnation (IWI) as described in detail elsewhere [12,15]. Prior to synthesis, fumed silica of 7 nm primary particle size (Sigma-Aldrich, St. Louis, MO, USA) was dried in an oven at 140 °C for 12 h. Firstly, 1.4 mL of aqueous solution (miliQ water) of Mn(Ac)2·4H2O (Sigma-Aldrich) with an appropriate concentration was dropwise added to 1 g of silica while vortex mixing. After impregnation, the powder was sufficiently mixed and then dried in the oven at 140 °C overnight. After drying, the powder was calcined to obtain Mn2O3/SiO2. The calcination was conducted in a tubular furnace, under air flow of ~60 mL·min−1 at 110 °C for 2 h and then at 500 °C for 3 h. The ramping rate was set to 2 °C per minute. Next, 1.4 mL solution of Na2WO4·2H2O (Alfa Aesar, Haverhill, MA, USA) in miliQ water with appropriate concentration was impregnated in the same way on the obtained Mn2O3/SiO2. After drying in the oven at 140 °C overnight, the catalyst was calcined in a tubular furnace, under air flow of ~60 mL·min−1 at 115 °C for 2 h and then at 850 °C for 5 h. The ramping rate was set to 5 °C per minute.
La/MgO. The La/MgO was synthesized by IWI. Prior to synthesis, the MgO of 325 mesh particle size (Strem Chemicals, Inc., Newburyport, MA, USA) was dried in an oven at 140 °C for 12 h. An amount of 1 mL of aqueous solution (miliQ water) of La(NO3)3·6H2O (Strem Chemicals, Inc.) with an appropriate concentration was dropwise added to 1 g of MgO while vortex mixing. After impregnation, the powder was sufficiently mixed and then dried in the oven at 140 °C overnight. After drying, the powder was calcined to obtain La/MgO. The calcination was conducted in a tubular furnace, under air flow of ~60 mL·min−1 at 800 °C for 4 h. The ramping rate was set to 5 °C per minute.
La2O3. The La2O3 was synthesized by the thermal decomposition of La(NO3)3·6H2O (Strem Chemicals, Inc.) at 115 °C for 2 h and 700 °C for 4 h. The ramping rate was set at 5 °C per minute.

2.2. Characterization Methods

The loading of the metals in each sample was determined by Inductively Coupled Plasma-Optical Emission Spectroscopy (ICP-OES) using a Thermo Scientific (iCAP 6000) spectrometer. Prior to analysis, each sample was dissolved in HNO3 and HCl in ¼ volumetric mixture at 70 °C overnight and then diluted with miliQ water. Crystalline components of the catalysts were determined by Powder X-ray Diffraction (PXRD). Patterns were recorded at ambient conditions on a Rigaku SmartLab diffractometer using Cu Kα radiation (λ = 1.54 Å) with 3° min−1 scan speed and 0.01° angle step. The catalyst morphology was evaluated before and after the reaction using High-Resolution Scanning Electron Microscopy (HR-SEM). The analysis was performed on an Ultraplus (Carl Zeiss, Heidelberg, Germany) Microscope. Everhart–Thornley and Backscattering detectors were used. For analysis, the sample powder was sprinkled on a carbon film before imaging.

2.3. Catalytic Testing

The catalysts were tested using a single-pass fixed-bed flow reactor made of quartz with an inner diameter of 7 mm. In each run of reaction, 50 mg of catalyst was loaded and packed inside the reactor. Two pieces of quartz wool were loaded and packed on both ends of the catalyst bed like a sandwich configuration to stabilize the reactor bed. Prior to catalyst loading, the sample was sieved and collected within 60–100 mesh. The reactor bed was packed to be 5 mm long. As La2O3 and MgO support were denser than SiO2, some amount of inert quartz sand (60–100 mesh) was added and mixed with the reactor bed to reach a consistent bed size. The temperature of the quartz reactor was controlled by a temperature-controlled split tubular furnace (Vecstar, Eurotherm, Suisse). The feed gas temperature was monitored by an external K-type thermocouple housed in a quartz capillary tube. The thermocouple was placed above the catalyst bed inside the reactor. It is noted that the measured furnace temperature was consistently higher by ~10 °C; hence, for convenience, we note only the furnace temperature. The feed gas composition and flow rate were controlled by mass flow controllers (Brooks SLA5850). Typically, the feed gas mixture had a total flow of 50 mL min−1 STP with composition of 4:1:1:4 CH4:O2:N2:Ar, which resulted in GHSV (gas hourly space velocity) of 60,000 mL h−1 g−1 (unless otherwise stated upon total flow change). The absolute reactor pressure was kept constant at ~1.3 bar using a back-pressure controller (Brooks SLA5820). The reaction-produced water was separated in an ice trap placed at the outlet of the reactor. The outlet stream of the reactor was analyzed by a Clarus 580 GC (PerkinElmer) equipped with TCD and FID detectors. The CH4 and O2 conversions were calculated according to Equation (1) (i denotes CH4 or O2), while C2 selectivity and C2 yield were calculated according to Equation (2) (Fi is molar flow rate) and Equation (3). To consider the change in the molar flow following the reaction, N2 was used as an internal standard. The mass balance was ensured to close with >95%.
Conversion % = F i , in F i , out F i , in · 100
C 2   Selectivity % = 2 · F C 2 H 4 + F C 2 H 6 F CH 4 , in F C H 4 , out · 100
C 2   yield ( % ) = CH 4   Conversion ( % ) · C 2   Selectivity ( % ) / 100
The stacking and mixing of different catalytic beds were prepared as follows:
Stacked Beds (s). The stacking (s) beds La2O3-MnNaWSi and La/MgO-MnNaWSi were prepared by packing the La-containing catalysts (La2O3 or La/MgO) on top of the already packed MnNaWSi catalyst bed. Different bed ratios (x) between the La-containing bed and the entire bed (50 mg) were studied, denoted as La2O3-MnNaWSi_xs and La/MgO-MnNaWSi_xs (x = 0.1, 0.2, or 0.5), respectively. The catalyst bed size for all tests was kept constant at length of 5 mm. Since La2O3 and La/MgO have higher density than the MnNaWSi catalyst, the MnNaWSi catalyst bed was diluted with inert quartz sand (60–100 mesh) to maintain a consistent bed size of 5 mm. The quartz sand was mixed with MnNaWSi bed instead of La catalyst beds to minimize the limitation of •CH3 transport to the lower bed. Prior to the preparation of catalyst bed, all particles were sieved and collected in between 60 and 100 mesh.
Mixed Beds (m). The mechanically mixed (m) beds of La2O3-MnNaWSi and La/MgO-MnNaWSi were prepared by mixing the La-containing catalysts (La2O3 or La/MgO) with MnNaWSi. Prior to mixing, all particles were sieved and collected in between 60 and 100 mesh. Different weight ratios (x) between the La-containing catalysts and the entire bed (50 mg) were studied, denoted as La2O3-MnNaWSi_xm and La/MgO-MnNaWSi_xm (x = 0.1, 0.2 and 0.8), respectively. The catalyst beds for all tests were kept constant at mass of 50 mg and length of 5 mm. Since La2O3 and La/MgO have higher density than the MnNaWSi catalyst, the MnNaWSi catalyst bed was diluted with inert quartz sand (60–100 mesh) to maintain a consistent bed size of 5 mm.

3. Results and Discussion

The catalysts in this study include La2O3, MgO-supported La2O3 (La/MgO), and Mn-Na2WO4-SiO2 (MnNaWSi) or a combination thereof. We examined two types of catalyst bed packing: (1) stacking (s), namely the La-containing catalyst (La2O3 or La/MgO) placed on top of the MnNaWSi catalyst bed, and (2) mixed (m), namely a uniform mechanical mix of La-containing catalyst (La2O3 or La/MgO) and the MnNaWSi. For clarity, the nomenclature of all the reactor bed combination tested was summarized in Table 1. Different loadings of the La-containing bed (La catalyst mass fraction in bed) were studied, as shown in Table 1. Each catalyst alone, La2O3, La/MgO, or MnNaWSi was tested separately as controls.

3.1. Catalyst Characterization

The loadings of the active metals in the La/MgO and the MnNaWSi catalysts were measured by ICP-OES, and the results are summarized in Table S1. The crystal phases identified for MnNaWSi were cubic Na2WO4, tetragonal α-cristobalite, and tetragonal Mn7O8SiO4, consistent with the literature of Mn2O3-Na2WO4-SiO2, as shown in Figure S1 [10,15,32,33]. The crystal phases of La/MgO were cubic MgO and hexagonal La(OH)3, as shown in Figure S1a. Notably, the La2O3 phase is prone to react with the moisture in the air. Since the catalyst samples were exposed to ambient air containing water moisture, the spectra of the La-containing samples also showed the formation of La(OH)3, as shown in Figure S1b in Supplementary Materials. To validate the formation of La2O3, we analyzed a dedicated fresh catalyst sample, which was analyzed immediately as prepared using PXRD, as shown in Figure S1c. Previous work reported that at around 500 °C under air, the La(OH)3 phase transforms into La2O3 [34]. Hence, it is assumed that because of the in-situ calcination, the reaction temperature (>650 °C), and the oxidative conditions of OCM, the La in the active catalyst bed was initially present as La2O3 in both the La/MgO and La2O3 catalysts.

3.2. Catalytic Performance

The reaction onset temperature was ~700 °C for the MnNaWSi catalyst and <650 °C for the La/MgO catalyst in Figure 1a. Interestingly, by simply stacking a small amount of the La/MgO catalyst (10 wt.% of the bed) over the MnNaWSi catalyst (La/MgO-MnNaWSi_0.1s), the reaction started already below 650 °C, as shown in Figure 1a. Even more interesting is the higher C2 selectivity and C2 yield obtained by the La/MgO-MnNaWSi_0.1s, as shown in Figure 1c,d.
The low activity of the MnNaWSi catalyst below 750 °C is consistent with previously published observations [35,36,37]. The drastic increase in activity of the MnNaWSi above 750 °C was previously associated with its ability to desorb molecular oxygen [37], facilitated by the interactions between the MnOx and the Na2WO4 phases [10,38]. Gordienko et al. showed that the onset temperature for O2 desorption from the MnNaWSi catalyst was 700 °C and reached a maximum of around 750–800 °C [37]. Using operando Raman and XRD-CT, Werny and Matras et al. demonstrated that at temperatures > 650 °C, the Mn7O8SiO4 phase starts losing its crystallinity due to reaction with the molting Na2WO4 and the simultaneous formation of MnWO4 [10,38]. These coincide with the strong deactivation of the MnNaWSi catalyst, as was previously addressed by Lunsford and coworkers [39] and later on emphasized by Hayek et al. [15]. Hence, from the perspective of the longevity of the MnNaWSi catalyst, it is beneficial to run the OCM reaction at a temperature lower than the melting temperature of Na2WO4 (698 °C).
We found that the CH4 conversion over the La/MgO catalyst increased only slightly (~16–21%) as compared to the more significant increase in MnNaWSi (1–16%), but the C2 selectivity of La/MgO did increase with temperature, as shown in Figure 1a,c, respectively. The La/MgO-MnNaWSi_0.1s catalyst showed C2 selectivity greater or similar to that of the MnNaWSi, which comprises 90 wt.% of the catalyst bed, as shown in Figure 1c. However, its conversion was comparable to that of the La/MgO catalyst, which comprises only 10 wt.% of the La/MgO-MnNaWSi_0.1s catalyst bed. Specifically, at 725 °C, we find that the CH4 and O2 conversions of the La/MgO-MnNaWSi_0.1s were the same as those of the La/MgO while the selectivity of the former was higher by 14% and even slightly higher than that of the MnNaWSi. The reaction results at 725 °C with time on stream for 7 h demonstrate that the performance of all three catalysts was stable, with the C2 yield being consistently highest over the La/MgO-MnNaWSi_0.1s, as shown in Figure 2d.
The benefit of mixing bulk La2O3 with Na2WO4/SiO2 (NaWSi) at a 50:50 wt.% ratio was previously identified by Zou et al. [31]. Their results at ~725 °C show similar performance between the La2O3 and the La2O3 mixed with Na2WO4/SiO2, which indicates that the La2O3 phase at that temperature range was governing the catalytic performance. We, therefore, tested OCM using unsupported La2O3 stacked on MnNaWSi at a 10:90 wt.% (La-MnNaWSi_0.1s) at 725 °C and compared it to the performance of the La/MgO-MnNaWSi_0.1s catalyst, as shown in Figure S2. Given that MgO gives a very low C2 yield (Figure S2), it can be expected that the La-MnNaWSi_0.1s, having ~30 fold more La2O3, would give significantly higher activity and C2 yield compared to the La/MgO-MnNaWSi_0.1s catalyst. Interestingly, it was observed that over the relatively short reaction time (7 h), the C2 yield of the former started dropping while the latter remained fairly stable (Figure S2d). The most evident difference between the two catalysts was in the sharper decrease in O2 conversion of the La/MgO-MnNaWSi_0.1s compared to La-MnNaWSi_0.1s. We associate the leveling in the C2 yield in the former with the decrease in O2 gas phase concentration, which maintains a more stable C2 selectivity in the former. Speculating that the direct interaction of the La2O3 with the Na2WO4 is promoting the deactivation, we analyzed the spent La2O3-NaWSi_0.8m using PXRD, as shown in Figure S3a. From the XRD data, we noticed that the Na2WO4 disappeared completely (after 14 h on stream), the peaks associated with the La(OH)3 phase diminished, and a new inactive NaLa(WO4)2 phase emerged, as shown in Figure S3. These peaks were found to be mild in the La/MgO-MnNaWSi_0.1m catalyst, as shown in Figure S4. The formation of this phase was previously reported by Matras et al. using a La-containing membrane reactor [40]. These results clearly show the instability of the La2O3 when in direct contact with the Na2WO4. Presumably, in the work of Zou et al., the effect of Na2WO4-induced deactivation was not evident, which was likely masked by the use of much larger amounts of catalysts, as we previously reported for the MnNaWSi catalyst [15]. Importantly, no such interactions were evidenced for MgO and Na2WO4. To further monitor the deactivation process, we examined the HR-SEM images of bulk La2O3, La/MgO, and MnNaWSi, as shown in Figure S5. The images clearly show that the spent La/MgO in the presence of the MnNaWSi phase undergoes sintering. Notably, the XRD data in Figures S3 and S4 show that the La-MnNaWSi_0.1s is more prone to deactivate compared to the La/MgO-MnNaWSi_0.1s, which highlights the benefit of the MgO support.
To examine the effect of residence time on catalytic performance, we tested the La/MgO-MnNaWSi_0.1s catalyst under several GHSV values at a temperature below the melting temperature of Na2WO4 (675 °C), which also means that the MnNaWSi phase is inactive, and at 725 °C, where both catalyst domains are active, as shown in Figure 3. Interestingly, at 725 °C, increasing the residence time (lower GHSV) did not significantly affect the methane conversion, C2 selectivity, or C2 yield, whereas the O2 conversion showed a consistent decrease. It can therefore be concluded that the trend seen in CH4 conversion at 725 °C is attributed to active site saturation at all tested GHSV values.
At 675 °C, we find that the C2 selectivity and C2 yield increased with GHSV. The CH4 conversion remained similar at 20 and 40 L/gcat/h and increased slightly at 60 and 80 L/gcat/h. Moreover, the C2 selectivity was found to increase monotonically while the O2 conversion decreased with the increase in GHSV. It is noted that the MnNaWSi catalyst is largely inactive at 675 °C (Figure 1a). To better understand the catalytic behavior of the stacked catalysts below the melting temperature of Na2WO4, we tested the bare La/MgO and MnNaWSi at the same GHSV values, keeping the stacked structure of the catalyst bed and its size the same while replacing one of the domains with quartz particles, as shown in Figure S6. Consistently, we find that in all the tested GHSV values, the activity, selectivity, and yield of the La/MgO-MnNaWSi_0.1s were much higher than the respective levels in the stacked single component bed of La/MgO and MnNaWSi. The stacked MnNaWSi alone was found to mildly consume both methane and O2 with no measurable formation of C2 products. We find that decrease in O2 conversion of the La/MgO-MnNaWSi_0.1s catalyst as a function of GHSV is correlated to that of the La/MgO. In contrast, the stacked La/MgO showed a decrease in C2 selectivity, whereas the La/MgO-MnNaWSi_0.1s showed an increase in C2 selectivity and yield with the increase in GHSV. It is postulated that the high reactivity of the La/MgO phase leads to the depletion of oxygen in the upper part of the bed. Hence, the generated methyl radicals and ethane are channeled to the bottom part of the catalyst bed, which, as previously reported [14,41,42] and shown here in Figure S6, affords the higher C2 selectivity under lean oxygen conditions.
The trends above strongly indicate that the two catalyst domains “communicate” to affect the kinetic mechanism. This observation is further supported by looking at the change in C2H4/C2H6 ratio as a function of temperature, as shown in Figure 4. We observe that above 700 °C, when the MnNaWSi becomes significantly active, the ratio starts to deviate from that obtained by the La/MgO catalyst to align with that of the MnNaWSi at 750 °C.
To evaluate the effect of the transport of ethane and methyl radicals between the two catalyst phases, we tested the effect of catalytic bed composition with different mass ratio mechanical mixing (instead of stacking), as shown in Figure 5. Interestingly, the catalysts with the higher bed fractions of La/MgO (20 wt.% and 50 wt.%) showed higher and more stable oxygen consumption in conjunction with a proportionally lower C2 selectivity and, in turn, a lower C2 yield. Overall, the 10 wt.% stacked catalyst showed both the highest and most stable C2 selectivity and C2 yield with TOS. With the increase in the bed volume to 20 wt.% and 50 wt.% we find that the C2 selectivity drops. The drop in selectivity and, in turn, the C2 yield is attributed to the short lifetime of methyl radicals [43], which most likely terminated before traversing to the bottom domain of the catalytic bed. This observation is further supported by the likeliness of the catalytic performance between the La/MgO-MnNaWSi_0.5s and the La/MgO-only bed. Explicitly, in the two catalyst bed compositions, the La/MgO dominated the catalytic performance, and the MnNaWSi was barely active in the test. The mechanically mixed 10 wt.% catalyst bed (La/MgO-MnNaWSi_0.1m) demonstrated low catalytic performance, similar to that of the MnNaWSi, as shown in Figure 5. Moreover, the mixed catalyst shows a drop in the catalytic activity and selectivity with TOS, which highlights the continuous occurrence of La2O3 deactivation due to interaction with the molten Na2WO4 phase, as previously reported for other oxides as well [15].
Plotting the C2 yield vs. the ethylene/ethane ratio at 725 °C for all the catalysts in this work, we find that the efficient stacking of the catalysts promotes both a higher C2 yield and a higher olefin-to-paraffin ratio, as shown in Figure 6. Interestingly, we find that the increase in the La/MgO amount, in the La/MgO-MnNaWSi_xs (x = 0.1, 0.2, 0.5) catalyst bed is a correlated to a decrease in both the C2 yield and the C2H4/C2H6 ratio with the La/MgO-MnNaWSi_0.5s nearing those of the mixed-bed La/MgO-MnNaWSi_0.1m. The mechanical mixing of the two catalysts leads to relatively low C2 yield and low C2H4/C2H6 ratios. Both of these trends are attributed to the deactivation process, as discussed above. Under lean oxygen conditions, Yoon et al. demonstrated that the conversion of ethane to ethylene occurs over surface lattice oxygens in MnNaWSi [41]. This coincides well with the higher C2H4/C2H6 ratios measured for the stacked beds. In these cases, the depletion of gas-phase oxygen in the top part of the bed results in the promotion of OCM in the bottom part of the bed under lean oxygen conditions, as opposed to the joint consumption of oxygen in the mixed catalyst bed.

4. Conclusions

In this work, we demonstrate the promotion of OCM over a dual-domain catalyst bed composed of a La/MgO catalyst and the MnNaWSi catalyst. We show that the OCM mechanism is better promoted when the bed is composed of only 10 wt.% La/MgO stacked on top of 90 wt.% MnNaWSi (La/MgO-MnNaWSi_0.1s). Testing the latter bed composition at ~650 °C and 60 L/gcat/h, we obtain ~10% CH4 conversion and ~19% C2 selectivity, whereas the MnNaWSi alone is inactive, and the La/MgO gives only 13% selectivity. Increasing the temperature to ~705 °C, the La/MgO-MnNaWSi_0.1s catalyst is found to give ~six-fold higher activity compared to the MnNaWSi catalyst or similar compared to the La/MgO catalyst, while the selectivity of the La/MgO-MnNaWSi_0.1s was 41%, which is higher than both individual catalysts. We further find that increasing the portion of the La/MgO or mixing it with the MnNaWSi showed diminished OCM performance, which is attributed to the deactivation of the La2O3 by reaction with the Na2WO4. Finally, we show that the dual-domain catalyst bed is promoting OCM through a synergistic kinetic mechanism that is active even below 700 °C.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/chemistry5020075/s1, Figure S1: XRD patterns for the original catalysts, (a) La/MgO; (b)La2O3 (after exposure to ambient air); (c) La2O3(fresh); (d) MnNaWSi; Figure S2: Catalytic performance of La2O3 or La/MgO bed stacking over MnNaWSi bed at 725°C; Figure S3: XRD patterns: (a) Spent La2O3-NaWSi_0.8m bed reacted for 14 h at 675–725 °C; (b) Fresh NaWSi catalyst; (c) Fresh La2O3 catalyst; Figure S4: XRD patterns: (a) Spent La/MgO-MnNaWSi_0.1m bed, reacted for 14 h at 725–750 °C; (b) Fresh MnNaWSi catalyst; (c) Fresh La/MgO catalyst; Figure S5: HR-SEM images of fresh and spent catalysts: (a) fresh bulk La2O3; (b) Spent La2O3; (c) fresh La/MgO; (d) Spent La/MgO; (e) fresh MnNaWSi; (f) spent MnNaWSi; (g) spent La/MgO-MnNaWSi_0.1s; (h) spent La/MgO-MnNaWSi_0.1m; Figure S6: Catalytic performance over a range of GHSV (20-80 L g−1 h−1) Table S1: Compositions of different pure catalysts.; Table S2: Expected catalytic performance of stacked 10 wt% catalyst based on performance of single components and their mass composition.

Author Contributions

Conceptualization, B.H., J.W., D.S. and O.M.G.; software, B.H.; methodology, B.H.; validation, B.H. and O.M.G.; formal analysis, B.H., J.W. and D.S.; investigation, B.H.; resources, O.M.G.; data curation, B.H., J.W. and O.M.G.; writing—original draft preparation, B.H.; writing—review and editing, B.H., J.W., D.S. and O.M.G.; visualization, B.H.; supervision, O.M.G.; project administration, O.M.G.; funding acquisition, O.M.G. All authors have read and agreed to the published version of the manuscript.

Funding

The generous supported of the Binational Industrial Research and Development (BIRD) foundation (Grant number 716613) is gratefully acknowledged.

Data Availability Statement

Not applicable.

Acknowledgments

The financial support of the Grand Technion Energy Program and the Russell Berrie Nanotechnology Institute is also acknowledged. The author B.H. acknowledges the financial support of the Energean Excellence Scholarship.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Keller, G.E.; Bhasin, M.M. Synthesis of Ethylene via Oxidative Coupling of Methane. I. Determination of Active Catalysts. J. Catal. 1982, 73, 9–19. [Google Scholar] [CrossRef]
  2. Galadima, A.; Muraza, O. Revisiting the Oxidative Coupling of Methane to Ethylene in the Golden Period of Shale Gas: A Review. J. Ind. Eng. Chem. 2016, 37, 1–13. [Google Scholar] [CrossRef]
  3. Weiland, P. Biogas Production: Current State and Perspectives. In Applied Microbiology and Biotechnology; Springer: Berlin/Heidelberg, Germany, 2010; pp. 849–860. [Google Scholar] [CrossRef]
  4. Kiani, D.; Sourav, S.; Baltrusaitis, J.; Wachs, I.E. Oxidative Coupling of Methane (OCM) by SiO2-Supported Tungsten Oxide Catalysts Promoted with Mn and Na. ACS Catal. 2019, 9, 5912–5928. [Google Scholar] [CrossRef]
  5. Arndt, S.; Otremba, T.; Simon, U.; Yildiz, M.; Schubert, H.; Schomäcker, R. Mn-Na2WO4/SiO2 as Catalyst for the Oxidative Coupling of Methane. What Is Really Known? Appl. Catal. A Gen. 2012, 425–426, 53–61. [Google Scholar] [CrossRef]
  6. Xueping, F.; Shuben, L.; Jingzhi, L.; Yanlai, C. Oxidative Coupling of Methane on W-Mn Catalysts. J. Mol. Catal. 1992, 6, 427–433. [Google Scholar]
  7. Kiani, D.; Sourav, S.; Wachs, I.E.; Baltrusaitis, J. Synthesis and Molecular Structure of Model Silica-Supported Tungsten Oxide Catalysts for Oxidative Coupling of Methane (OCM). Catal. Sci. Technol. 2020, 10, 3334–3345. [Google Scholar] [CrossRef]
  8. Li, S.B. Oxidative Coupling of Methane over W-Mn/SiO2 Catalyst. Chin. J. Chem. 2001, 19, 16–21. [Google Scholar] [CrossRef]
  9. Wang, D.J.; Rosynek, M.P.; Lunsford, J.H. Oxidative Coupling of Methane over Oxide-Supported Sodium-Manganese Catalysts. J. Catal. 1995, 155, 390–402. [Google Scholar] [CrossRef]
  10. Werny, M.J.; Wang, Y.; Girgsdies, F.; Schlögl, R.; Trunschke, A. Fluctuating Storage of the Active Phase in a Mn-Na2WO4/SiO2 Catalyst for the Oxidative Coupling of Methane. Angew. Chem. 2020, 132, 15031–15036. [Google Scholar] [CrossRef]
  11. Sinev, M.; Ponomareva, E.; Sinev, I.; Lomonosov, V.; Gordienko, Y.; Fattakhova, Z.; Shashkin, D. Oxygen Pathways in Oxidative Coupling of Methane and Related Processes. Case Study: NaWMn/SiO2 Catalyst. Catal. Today 2019, 333, 36–46. [Google Scholar] [CrossRef]
  12. Hayek, N.S.; Lucas, N.S.; Warwar Damouny, C.; Gazit, O.M. Critical Surface Parameters for the Oxidative Coupling of Methane over the Mn-Na-W/SiO2 Catalyst. ACS Appl. Mater. Interfaces 2017, 9, 40404–40411. [Google Scholar] [CrossRef]
  13. Huang, B.; Hayek, N.S.; Sun, G.; Mottaghi-Tabar, S.; Simakov, D.S.A.; Gazit, O.M. Correlating Properties of the Mn2O3-Na2WO4/SiO2 Catalyst with Statistically Estimated Parameters for the Oxidative Coupling of Methane. Energy Fuels 2021, 35, 9589–9598. [Google Scholar] [CrossRef]
  14. Pak, S.; Qiu, P.; Lunsford, J.H. Elementary Reactions in the Oxidative Coupling of Methane over Mn/Na2WO4/SiO2 and Mn/Na2WO4/MgO Catalysts. J. Catal. 1998, 179, 222–230. [Google Scholar] [CrossRef]
  15. Hayek, N.S.; Khlief, G.J.; Horani, F.; Gazit, O.M. Effect of Reaction Conditions on the Oxidative Coupling of Methane over Doped MnOx-Na2WO4/SiO2 Catalyst. J. Catal. 2019, 376, 25–31. [Google Scholar] [CrossRef]
  16. Lomonosov, V.I.; Sinev, M.Y. Oxidative Coupling of Methane: Mechanism and Kinetics. Kinet. Catal. 2016, 57, 647–676. [Google Scholar] [CrossRef]
  17. Martin, G.A.; Mirodatos, C. Surface Chemistry in the Oxidative Coupling of Methane. Fuel Process. Technol. 1995, 42, 179–215. [Google Scholar] [CrossRef]
  18. Horn, R.; Schlögl, R. Methane Activation by Heterogeneous Catalysis. Catal. Lett. 2015, 145, 23–39. [Google Scholar] [CrossRef] [Green Version]
  19. Campbell, K.D.; Lunsford, J.H. Contribution of Gas-Phase Radical Coupling in the Catalytic Oxidation of Methane. J. Phys. Chem. 1988, 92, 5792–5796. [Google Scholar] [CrossRef]
  20. Tong, Y.; Rosynek, M.P.; Lunsford, J.H. Secondary Reactions of Methyl Radicals with Lanthanide Oxides: Their Role in the Selective Oxidation of Methane. J. Phys. Chem. 1989, 93, 2896–2898. [Google Scholar] [CrossRef]
  21. Zhou, Q.; Wang, Z.Q.; Li, Z.; Wang, J.; Xu, M.; Zou, S.; Yang, J.; Pan, Y.; Gong, X.Q.; Xiao, L.; et al. CH3•-Generating Capability as a Reactivity Descriptor for Metal Oxides in Oxidative Coupling of Methane. ACS Catal. 2021, 11, 14651–14659. [Google Scholar] [CrossRef]
  22. Lin, C.H.; Campbell, K.D.; Wang, J.X.; Lunsford, J.H. Oxidative Dimerization of Methane over Lanthanum Oxide. J. Phys. Chem. 1986, 90, 534–537. [Google Scholar] [CrossRef]
  23. Wu, J.; Zhang, H.; Qin, S.; Hu, C. La-Promoted Na2WO4/Mn/SiO2 Catalysts for the Oxidative Conversion of Methane Simultaneously to Ethylene and Carbon Monoxide. Appl. Catal. A Gen. 2007, 323, 126–134. [Google Scholar] [CrossRef]
  24. Ismagilov, I.Z.; Matus, E.V.; Vasil’Ev, S.D.; Kuznetsov, V.V.; Kerzhentsev, M.A.; Ismagilov, Z.R. Oxidative Condensation of Methane in the Presence of Modified MnNaW/SiO2 Catalysts. Kinet. Catal. 2015, 56, 456–465. [Google Scholar] [CrossRef]
  25. Ghose, R.; Hwang, H.T.; Varma, A. Oxidative Coupling of Methane Using Catalysts Synthesized by Solution Combustion Method: Catalyst Optimization and Kinetic Studies. Appl. Catal. A Gen. 2014, 472, 39–46. [Google Scholar] [CrossRef]
  26. Zhao, M.; Ke, S.; Wu, H.; Xia, W.; Wan, H. Flower-like Sr-La2O3 Microspheres with Hierarchically Porous Structures for Oxidative Coupling of Methane. Ind. Eng. Chem. Res. 2019, 58, 22847–22856. [Google Scholar] [CrossRef]
  27. Hou, Y.H.; Han, W.C.; Xia, W.S.; Wan, H.L. Structure Sensitivity of La2O2CO3 Catalysts in the Oxidative Coupling of Methane. ACS Catal. 2015, 5, 1663–1674. [Google Scholar] [CrossRef]
  28. Noon, D.; Seubsai, A.; Senkan, S. Oxidative Coupling of Methane by Nanofiber Catalysts. ChemCatChem 2013, 5, 146–149. [Google Scholar] [CrossRef]
  29. Song, J.; Sun, Y.; Ba, R.; Huang, S.; Zhao, Y.; Zhang, J.; Sun, Y.; Zhu, Y. Monodisperse Sr–La2O3 Hybrid Nanofibers for Oxidative Coupling of Methane to Synthesize C2 Hydrocarbons. Nanoscale 2015, 7, 2260–2264. [Google Scholar] [CrossRef]
  30. Jaroenpanon, K.; Tiyatha, W.; Chukeaw, T.; Sringam, S.; Witoon, T.; Wattanakit, C.; Chareonpanich, M.; Faungnawakij, K.; Seubsai, A. Synthesis of Na2WO4-MnxOy Supported on SiO2 or La2O3 as Fiber Catalysts by Electrospinning for Oxidative Coupling of Methane: Synthesis of Na2WO4-MnxOy Supported on SiO2 or La2O3 as Fiber Catalysts. Arab. J. Chem. 2022, 15, 103577. [Google Scholar] [CrossRef]
  31. Zou, S.; Li, Z.; Zhou, Q.; Pan, Y.; Yuan, W.; He, L.; Wang, S.; Wen, W.; Liu, J.; Wang, Y.; et al. Surface Coupling of Methyl Radicals for Efficient Low-Temperature Oxidative Coupling of Methane. Chin. J. Catal. 2021, 42, 1117–1125. [Google Scholar] [CrossRef]
  32. Rollins, B.; Grosjean, N.; Poulston, S. The Impact of the Presence of Various Relevant Gases during Aging of MnNa2WO4/SiO2 Catalyst on Its Performance for the Oxidative Coupling of Methane. Appl. Catal. A Gen. 2020, 598, 117606. [Google Scholar] [CrossRef]
  33. Hayek, N.S.; Lucas, N.S.; Warwar Damouny, C.; Gazit, O.M. A Comparative Study of Precursor Effect on Manganese Post-Synthetic Incorporation into the T-Sites of Dealuminated B-Zeolite. Microporous Mesoporous Mater. 2017, 244, 31–36. [Google Scholar] [CrossRef]
  34. Haibel, E.; Berendts, S.; Walter, D. Thermogravimetric and X-ray Diffraction Investigation on Carbonated Lanthanum Oxide and Lanthanum Hydroxide Formed in Humid CO2 Atmosphere. J. Therm. Anal. Calorim. 2018, 134, 261–267. [Google Scholar] [CrossRef]
  35. Sadjadi, S.; Jašo, S.; Godini, H.R.; Arndt, S.; Wollgarten, M.; Blume, R.; Görke, O.; Schomäcker, R.; Wozny, G.; Simon, U. Feasibility Study of the Mn–Na2WO4/SiO2 Catalytic System for the Oxidative Coupling of Methane in a Fluidized-Bed Reactor. Catal. Sci. Technol. 2015, 5, 942–952. [Google Scholar] [CrossRef]
  36. Hiyoshi, N.; Ikeda, T. Oxidative Coupling of Methane over Alkali Chloride–Mn–Na2WO4/SiO2 Catalysts: Promoting Effect of Molten Alkali Chloride. Fuel Process. Technol. 2015, 133, 29–34. [Google Scholar] [CrossRef]
  37. Gordienko, Y.; Usmanov, T.; Bychkov, V.; Lomonosov, V.; Fattakhova, Z.; Tulenin, Y.; Shashkin, D.; Sinev, M. Oxygen Availability and Catalytic Performance of NaWMn/SiO2 Mixed Oxide and Its Components in Oxidative Coupling of Methane. Catal. Today 2016, 278, 127–134. [Google Scholar] [CrossRef]
  38. Matras, D.; Vamvakeros, A.; Jacques, S.D.M.; Grosjean, N.; Rollins, B.; Poulston, S.; Stenning, G.B.G.; Godini, H.R.; Drnec, J.; Cernik, R.J.; et al. Effect of Thermal Treatment on the Stability of Na-Mn-W/SiO2catalyst for the Oxidative Coupling of Methane. Faraday Discuss. 2021, 229, 176–196. [Google Scholar] [CrossRef] [Green Version]
  39. Pak, S.; Lunsford, J.H. Thermal Effects during the Oxidative Coupling of Methane over Mn/Na2WO4/SiO2 and Mn/Na2WO4/MgO Catalysts. Appl. Catal. A Gen. 1998, 168, 131–137. [Google Scholar] [CrossRef]
  40. Vamvakeros, A.; Matras, D.; Jacques, S.D.M.; di Michiel, M.; Middelkoop, V.; Cong, P.; Price, S.W.T.; Bull, C.L.; Senecal, P.; Beale, A.M. Real-Time Tomographic Diffraction Imaging of Catalytic Membrane Reactors for the Oxidative Coupling of Methane. Catal. Today 2021, 364, 242–255. [Google Scholar] [CrossRef]
  41. Yoon, S.; Lim, S.; Choi, J.W.; Suh, D.J.; Song, K.H.; Ha, J.M. Study on the Unsteady State Oxidative Coupling of Methane: Effects of Oxygen Species from O2, Surface Lattice Oxygen, and CO2 on the C2+ Selectivity. RSC Adv. 2020, 10, 35889–35897. [Google Scholar] [CrossRef]
  42. Tiemersma, T.P.; Tuinier, M.J.; Gallucci, F.; Kuipers, J.A.M.; Annaland, M.V.S. A Kinetics Study for the Oxidative Coupling of Methane on a Mn/Na2WO4/SiO2 Catalyst. Appl. Catal. A Gen. 2012, 433–434, 96–108. [Google Scholar] [CrossRef]
  43. Jones, C.A.; Leonard, J.J.; Sofranko, J.A. The Oxidative Conversion of Methane to Higher Hydrocarbons over Alkali-Promoted MnSiO2. J. Catal. 1987, 103, 311–319. [Google Scholar] [CrossRef]
Figure 1. Temperature dependencies (650–750 °C) of the La/MgO, MnNaWSi catalysts, and the stacking experiment, La/MgO-MnNaWSi_0.1s. (a) CH4 conversion; (b) O2 conversion; (c) C2 selectivity; and (d) C2 yield. La/MgO-MnNaWSi_0.1s, ● La/MgO, and ▲ MnNaWSi. Stacking of La/MgO and MnNaWSi beds with 10 wt.% of La/MgO and 90 wt.% of MnNaWSi.
Figure 1. Temperature dependencies (650–750 °C) of the La/MgO, MnNaWSi catalysts, and the stacking experiment, La/MgO-MnNaWSi_0.1s. (a) CH4 conversion; (b) O2 conversion; (c) C2 selectivity; and (d) C2 yield. La/MgO-MnNaWSi_0.1s, ● La/MgO, and ▲ MnNaWSi. Stacking of La/MgO and MnNaWSi beds with 10 wt.% of La/MgO and 90 wt.% of MnNaWSi.
Chemistry 05 00075 g001
Figure 2. Catalytic performance at low temperature (725 °C) of single catalyst beds and composite bed. (a) CH4 conversion; (b) O2 conversion; (c) C2 selectivity; and (d) C2 yield. ● La/MgO, ▲ MnNaWSi, and La/MgO-MnNaWSi_0.1s.
Figure 2. Catalytic performance at low temperature (725 °C) of single catalyst beds and composite bed. (a) CH4 conversion; (b) O2 conversion; (c) C2 selectivity; and (d) C2 yield. ● La/MgO, ▲ MnNaWSi, and La/MgO-MnNaWSi_0.1s.
Chemistry 05 00075 g002
Figure 3. Analysis of La/MgO-MnNaWSi_0.1s at 725 °C and at 675 °C: (a) CH4 conversion; (b) O2 conversion; (c) C2 selectivity; and (d) C2 yield. Reaction conditions: 50 mg catalyst; catalyst bed 5 mm; molar ratio of CH4:O2:N2:Ar = 4:1:1:4, and P ~ 1.3 bar. Each point is an average of data collected over 4 h of reaction, with the error bars representing the standard deviation.
Figure 3. Analysis of La/MgO-MnNaWSi_0.1s at 725 °C and at 675 °C: (a) CH4 conversion; (b) O2 conversion; (c) C2 selectivity; and (d) C2 yield. Reaction conditions: 50 mg catalyst; catalyst bed 5 mm; molar ratio of CH4:O2:N2:Ar = 4:1:1:4, and P ~ 1.3 bar. Each point is an average of data collected over 4 h of reaction, with the error bars representing the standard deviation.
Chemistry 05 00075 g003
Figure 4. C2H4 to C2H6 ratio as a function of temperature (650–750 °C) for La/MgO-MnNaWSi_0.1s, ● La/MgO, and ▲ MnNaWSi.
Figure 4. C2H4 to C2H6 ratio as a function of temperature (650–750 °C) for La/MgO-MnNaWSi_0.1s, ● La/MgO, and ▲ MnNaWSi.
Chemistry 05 00075 g004
Figure 5. Catalytic performance with different La/MgO and MnNaWSi bed compositions at 725 °C. (a) CH4 conversion; (b) O2 conversion; (c) C2 selectivity; (d) C2 yield. La/MgO-MnNaWSi_0.1s, La/MgO-MnNaWSi_0.2s, La/MgO-MnNaWSi_0.5s, La/MgO-MnNaWSi_0.1m, ● La/MgO, ▲ MnNaWSi, and ◆ MgO.
Figure 5. Catalytic performance with different La/MgO and MnNaWSi bed compositions at 725 °C. (a) CH4 conversion; (b) O2 conversion; (c) C2 selectivity; (d) C2 yield. La/MgO-MnNaWSi_0.1s, La/MgO-MnNaWSi_0.2s, La/MgO-MnNaWSi_0.5s, La/MgO-MnNaWSi_0.1m, ● La/MgO, ▲ MnNaWSi, and ◆ MgO.
Chemistry 05 00075 g005
Figure 6. Pearson’s correlation between the C2 yield and the ratio of selectivity ratio of C2H4/C2H6 for all the catalytic tests at 725 °C. Mixed, Stacked, and ⬛ Single components.
Figure 6. Pearson’s correlation between the C2 yield and the ratio of selectivity ratio of C2H4/C2H6 for all the catalytic tests at 725 °C. Mixed, Stacked, and ⬛ Single components.
Chemistry 05 00075 g006
Table 1. Nomenclature and the description of the catalyst reactor bed, including its reactor bed configuration and the weight proportion of the La bed.
Table 1. Nomenclature and the description of the catalyst reactor bed, including its reactor bed configuration and the weight proportion of the La bed.
NameNomenclatureLa Catalyst Mass Fraction in Bed
La/MgO--
La2O3--
MnNaWSi--
La2O3 + MnNaWSiLa2O3-MnNaWSi_0.1s0.1
La/MgO + MnNaWSiLa/MgO-MnNaWSi_0.1s0.1
La/MgO-MnNaWSi_0.2s0.2
La/MgO-MnNaWSi_0.5s0.5
La/MgO-MnNaWSi_0.1m0.1
La/MgO-MnNaWSi_0.2m0.2
La/MgO-MnNaWSi_0.8m0.8
where ‘s’ and ‘m’ represent a stacked and mixed-bed configuration, respectively.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Huang, B.; Wang, J.; Shpasser, D.; Gazit, O.M. Probing Low-Temperature OCM Performance over a Dual-Domain Catalyst Bed. Chemistry 2023, 5, 1101-1112. https://doi.org/10.3390/chemistry5020075

AMA Style

Huang B, Wang J, Shpasser D, Gazit OM. Probing Low-Temperature OCM Performance over a Dual-Domain Catalyst Bed. Chemistry. 2023; 5(2):1101-1112. https://doi.org/10.3390/chemistry5020075

Chicago/Turabian Style

Huang, Baoting, Jin Wang, Dina Shpasser, and Oz M. Gazit. 2023. "Probing Low-Temperature OCM Performance over a Dual-Domain Catalyst Bed" Chemistry 5, no. 2: 1101-1112. https://doi.org/10.3390/chemistry5020075

Article Metrics

Back to TopTop