Next Article in Journal
Efficient Synthesis of 3-Sulfonyl-2-sulfonylmethyl-2H-chromenes via Tandem Knoevenagel Condensation/Oxa-Michael Addition Protocol
Previous Article in Journal
Inorganic Salt Catalysed Hydrothermal Carbonisation (HTC) of Cellulose
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Oxidative Dehydrogenation of Ethane with CO2 over Mo/LDO Catalyst: The Active Species of Mo Controlled by LDO

School of Chemical Engineering, Northwest University, Xi’an 710069, China
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(5), 493; https://doi.org/10.3390/catal12050493
Submission received: 17 March 2022 / Revised: 17 April 2022 / Accepted: 26 April 2022 / Published: 28 April 2022

Abstract

:
A series of the layered double oxides supported molybdenum oxide catalysts were synthesized and evaluated in the oxidative dehydrogenation of ethane with CO2 (CO2-ODHE). The 22.3 wt% Mo/LDO catalyst delivered a 92.3%selectivity to ethylene and a 7.9% ethane conversion at relatively low temperatures. The molybdenum oxide catalysts were fully characterized by XRD, BET, SEM, TEM, UV–vis, Raman TG, and XPS. Isolated [MoO4]2− dominated on the surface of the fresh 12.5 wt% Mo/LDO catalyst. With the increase of the Mo content, the Mo species transformed from [MoO4]2− to [Mo7O24]6− and [Mo8O26]4 on the 22.3 wt% and 30.1 wt% Mo/LDO catalysts, respectively. The redox mechanism was proposed and three Mo species including [MoO4]2−, [Mo7O24]6−, and [Mo8O26]4− showed quite different functions in the CO2-ODHE reaction: [MoO4]2−, with tetrahedral structure, preferred the non-selective pathway; [Mo7O24]6, with an octahedral construction, promoted the selective pathway; and the existence of [Mo8O26]4− reduced the ability to activate ethane. This work provides detailed insights to further understand the relationship between structure–activity and the role of surface Mo species as well as their aggregation state in CO2-ODHE.

1. Introduction

After the shale gas revolution in North America, the verified huge natural gas reserve stimulated great interest in exploiting effective ways of utilizing it to produce value-added chemicals [1,2]. Ethane, as the second most abundant component (up to 15 vol%) of shale gas, can be used as feedstock to produce ethylene with a high selectivity in steam cracking [3]. However, this process is intensive in energy consumption while suffering from the coking formation and environmentally harmful gases emission due to the thermodynamic constraint [4,5]. Several alternative approaches that enhance the effectiveness and reduce emissions are being investigated. The oxidative dehydrogenation of ethane combined with CO2, which has an endothermic feature (C2H6 + CO2 → C2H4 + CO + H2O ∆H = + 177.6 KJ·mol−1) and suppresses the formation of coking, can be an attractive route for producing ethylene [6]. However, the thermodynamic energy of C-H bond breaking in ethane (423 kJ/mol) is less than that of the C-C bond (377 kJ/mol); the non-selective catalytic conversion of ethane often takes place during the reaction process (C2H6 + 2CO2→4CO + 3H2, C2H6→2C(s) + 3H2, C2H6 + H2→2CH4, H2 O + CO ↔ CO2 + H2, CO2 + C(s) ↔ 2CO), resulting in the formation of by-products [6].
Molybdenum-based oxide catalysts have been researched extensively because of their excellent activity in many dehydrogenation reactions [7,8]. The performance of Mo-supported catalytic systems strongly relates to the local structure and surface density of molybdenum species in CO2-ODHE. Christodoulakis et al. derived the relationship of structure–activity, finding that the variation trend of the Mo-O-Al bond number per Mo in surface MoOx species was consistent with the reaction rate per Mo [9]. Tsilomelekis et al. reported that terminal Mo=O sites participated in the non-selective reaction turnovers and the reaction routes followed a primarily selective pathway at high coverage [10]. The aggregation state of surface Mo species is the critical factor for this reaction. However, the effective active species in molybdenum oxide catalysts have not been determined yet. Thus, it would be more desirable to verify the effective active species to lay a foundation for the subsequent design and preparation of catalysts.
Layered double hydroxides (LDHs) are lamellar two-dimensional anionic clays composed of positively charged brucite-like layers [11]. [M2+1−xM3+x(OH)2](An−)x/n·mH2O is the general molecular formula of an LDH, in which M2+ and M3+ correspond to divalent and trivalent metal cations, respectively. An inorganic or organic anion inserts into the intercalation layers as charge compensation, written as An− [12]. The degree of aggregation can be regulated by anion exchange with different amounts of target active species. However, LDHs are easily gelatinized, which restricts its possibility for reuse [13]. The calcination of LDHs is usually adopted in the preparation of catalysts to improve its stability, resulting in layered double oxides (LDOs) as carrier [14]. Because of its unique physical and chemical properties, LDOs can be utilized as supported loading Mo to control the surface Mo species aggregation.
In this article, we first synthesized a series of Mo loaded LDHs by the anion-exchange method, then the target catalysts, Mo/LDO, were subsequently obtained by the calcination of precursors. We expect that the catalyst with the appropriate aggregation degree of Mo species can be prepared by the control of the LDO carrier and explore the most effective active species on this basis. The physicochemical properties of the Mo/LDO were characterized by BET, XRD, SEM, TEM, XPS, TG, UV–vis, and Raman spectroscopy, and the relationship between aggregation degree and the catalytic effect was demonstrated by these characterizations.

2. Results

2.1. Morphological Characteristics and Structural Properties

The powder XRD profiles of the as-prepared Mo/LDO catalysts and used Mo/LDO catalysts are demonstrated in Figure 1a,b, respectively. All the samples exhibited strong characteristic diffraction peaks related to the crystal structure of the cubic phase MgO (JCPDS card no. 45-0946), which were located at 43.8° and 63.0° and corresponded to the crystal planes of (200) and (220), respectively [15]. The minor XRD peak at 36° was attributed to (111) of MgO [16,17]. The crystallite size was estimated from the XRD patterns using the Scherrer equation. The crystallite size of the MgO phase followed the order of 30.1 wt% Mo/LDO (3.7 nm) < 22.3 wt% Mo/LDO (4.0 nm) < 12.5 wt% Mo/LDO (4.4 nm). The reflection planes such as (101), (112), (301), and (303) of CaMoO4 were well matched with the diffraction peaks at 2θ = 18.7°, 28.8°, 53.2°, and 58.2°, manifesting the appearance of CaMoO4 in all samples [18,19]. However, there were no diffraction peaks of aluminum oxide: manifesting aluminum oxide existed in the amorphous phase, in agreement with previous studies [20,21]. No diffraction peaks of carbon were observed because the coke formed during the reaction was considered amorphous carbon [22]. Moreover, the characteristic diffraction peaks of crystalline MoO3 located at 23.7°, 25.8°, and 27.4°did not appear, indicating the good dispersion of Mo species over the catalyst surface [23].
The nitrogen physisorption isotherms and corresponding pore size distribution curves of the synthesized catalysts are shown in Figure 2. According to the IUPAC classification, the curves of all samples displayed a type IV isotherm and contained the hysteresis in the high pressure region of P/P0 > 0.4, suggesting the obtained materials were mesoporous structural materials [24]. This structure was generated not only through the elimination of inter-layer CO32− and layer collapse, but also the elimination of pTOS in the synthesis process, which acted as a template and transformed into a porous mixed oxide [20]. As Mo loading increased, the nitrogen physisorption isotherms for 30.1 wt% Mo/LDO were type IV isotherms with an H4 type hysteresis loop, indicating that plate-like particles were agglomerated and resulted in slit-shaped mesopores [25]. Such a result suggests that the aggregation degree of the catalyst will increase significantly at a high Mo load.
As shown in Figure 2b, the pore size of the 12.5 wt% Mo/LDO showed a distribution in 5–15 nm with a maximum of about 8 nm. As the Mo contents increased, the maximum pore size shifted to 5 nm while the pore diameter distribution extended. For the 30.1 wt% Mo/LDO, the number of larger apertures decreased significantly and the maximum pore size shifted to a lower value, about 3 nm. Moreover, there was a small local peak with a maximum value at 10 nm on the 30.1 wt% Mo/LDO. The difference in pore size distribution implied the changes in specific surface area and pore volume.
According to the isotherms, the detailed textural properties were summarized in Table 1. There was a decrease in SBET, from 203.8 m2·g−1 for the 12.5 wt% Mo/LDO to 194.0 m2·g−1 for the 22.3 wt% Mo/LDO. Moreover, the specific surface area of the 30.1 wt% Mo/LDO declined significantly to 171.0 m2·g−1. The decrease of SBET was ascribed to the blocking of mesopores by the polymeric MoOx species and the collapse of the layer, particularly in the 30.1 wt% Mo/LDO. There was a pronounced decrease in the average pore diameter and total pore volume with the increase of Mo content. According to the results of BET, the mesoporous material was synthesized, and the aggregation degree of the catalyst increased with the elevation of Mo loading.
SEM and the elemental mapping were further employed to understand the nanostructures and the surface elements distribution of the catalysts. As shown in Figure 3, all the catalysts exhibited the layered structure of solid lamellar, similar to a previous finding [26]. However, there were distinct profiles between the series of Mo/LDO at the 1 µm level. The aggregation of catalysts appeared gradually, and a pronounced agglomeration was exhibited in the 30.1 wt% Mo/LDO, manifesting the collapse of the layer structure and the blocking of the pores.
The elements of Mg and Al were uniformly distributed on the surface of the catalysts, while the density of the Mo elements increased gradually with the elevation of the Mo load. According to the results of elemental mapping, the structure of the LDO formed successfully. and it could be used as a carrier to disperse Mo species aptly.
TEM images of the as-prepared catalysts are shown in Figure 4. The 2D sheet-like crystallites nanosheets with 10–30 nm appeared on all tested samples. Such nanostructures were composed of aggregated needle shapes or folded ribbon-like crystallites [27,28]. As the Mo loading increased, the aggregation of catalysts started, especially on the 30.1 wt% Mo/LDO (Figure 4f). The measured interplanar spacing was 0.21 nm, which matched well with the d-value of the (200) plane of MgO, as shown in Figure 4g,h [25]. The average particle size of MgO on the 22.3 wt% Mo/LDO was about 4.0 nm, which was consistent with the result of XRD. Moreover, the d spacing of 0.31 nm corresponded to the (112) plane of CaMoO4 [29].
The bonding information and molecular coordination environment of inorganic materials can be analyzed by UV–vis DRS. As shown in Figure 5a, there were absorption bands between 200 and 400 nm in the spectra, which originated from the ligand and metal charge transfer (CT) transitions (O2−→Mo6+) [30]. The absorption band of the 12.5 wt% Mo/LDO was mainly located at about 220–260 nm, which was ascribed to the isolated MoOx species with a tetrahedral structure and the terminal Mo=O bond of tetrahedral molybdate [31,32]. The absorption band at around 280 nm had been assigned to both characters as tetrahedrally coordinated species and octahedrally coordinated species [33,34]. Therefore, we attributed this band to the overlap of monomer and polymerized molybdate species ([MoO4]2−, [Mo7O24]6- and [Mo8O26]4−). Moreover, the bond located at 320 nm was assigned to the octahedral species with a Mo-O-Mo bridge bond [35]. The new absorption at 300 nm was detected on the 22.3 wt% Mo/LDO, which was characterized as the overlap of the absorption bands at 280 and 320 nm, indicating the formation of the octahedral species ([Mo7O24]6− and [Mo8O26]4−). The absorption band at 300 nm was more pronounced on the 30.1 wt% Mo/LDO, suggesting more octahedral molybdate species were formed by polymerizing the tetrahedral molybdate species. It could be deduced that introducing excessive Mo could result in a higher aggregation of Mo species.
The energy gap transformed from UV–vis DRS can be used to analyze the local structures of Mo oxides. There was an empirical linear correlation between the energy gap (E) and the average number of nearest Mo neighbors (NMo): NMo = 16–3.8 × E [36]. The average aggregation degree of Mo species increased as the energy gap decreased [37,38]. The energy gap of [MoO4]2− and [Mo7O24]6− was 4.4 and 3.3 eV, respectively [34,39]. In this work, the energy gap decreased from 3.40 on the 12.5 wt% Mo/LDO to 3.33 on the 22.3 wt% Mo/LDO, suggesting that the structure transformed from isolated [MoO4]2− to a polymerized molybdate species, [Mo7O24]6−. As the Mo content further increased to 30.1 wt%, the energy gap decreased to 3.28, indicating the formation of [Mo8O26]4−.
The adhibition of Raman spectroscopy relates to the status of transition metal oxides; this technique can be used to study Mo species on the surface of catalysts. Figure 6 displays the Raman spectra of fresh and used Mo/LDO catalysts.
As is shown in Figure 6a, three peaks were observed in the profiles of the fresh Mo/LDO catalysts. The peaks located at 912 and 980 cm−1 were related to the terminal Mo=O stretching vibration of highly dispersed [MoO4]2 [40,41]. These two peaks dominated on the surface of the 12.5 wt% Mo/LDO. As the Mo loading increased, the peak intensity of 912 and 980 cm−1 gradually decreased, inferring the disappearance of isolated [MoO4]2. The peak at 877 cm−1 was attributed to Mo-O-Mo stretching in [Mo7O24]6− and [Mo8O26]4− [40,42]. Oppositely, the peak located at around 877 cm−1 was more pronounced as the Mo loading elevated; this may be ascribed to the formation of polymerized molybdate species ([Mo7O24]6− and [Mo8O26]4−) on the 22.3 wt% and 30.1 wt% Mo/LDO. Moreover, the decrease of 877 cm−1 on the 30.1 wt% Mo/LDO was more significant, suggesting the higher aggregation degree of Mo species ([Mo8O26]4−).
After the reaction (Figure 6b), the peak at 912 cm−1 could still be observed on the used 12.5 wt% Mo/LDO, inferring the existence of [MoO4]2−. Simultaneously, the peak intensity of 877 cm−1 showed the same trend after the reaction, suggesting the existence of [Mo7O24]6− and [Mo8O26]4 on the 22.3 wt% Mo/LDO and 30.1 wt% Mo/LDO, respectively. The peak at 940 cm−1 which was assigned to the terminal Mo=O stretch of [Mo7O24]6 appeared on all used catalysts [43]. The peak at 940 cm−1 became relatively pronounced and the peak at 912 cm−1 declined over the 12.5 wt% Mo/LDO, indicating the transformation of [MoO4]2− to [Mo7O24]6−. The peak located at 1338 cm−1 related to amorphous carbon appeared after the reaction, and the peak intensity decreased significantly with the increase of Mo content [44]. No peaks were observed at 822 and 994 cm−1, suggesting the absence of crystalline MoO3 [45].
The above results revealed that the Mo species on the surface of catalysts varied with the Mo content. According to UV–vis, the surface Mo species transformed from isolated [MoO4]2− to [Mo7O24]6 and [Mo8O26]4− with the increase of Mo content. A substance will be reduced easier because of the higher density of states capable of accepting electrons [10]. Therefore, the catalyst with an appropriate aggregation degree of Mo species exhibited the highest selectivity.
The amount of carbon deposit formed during the reaction was verified by TG. As shown in Figure 7, the first weight loss in the temperature range of 303–473 K was attributed to the removal of physical absorbed water and other volatile compounds [46]. The second weight loss between 473 K and 713 K was ascribed to the burning of amorphous carbon [47,48]. The third weight loss at 713–873 K was due to the decomposition of hydrocarbons [48,49]. The deposition of graphitic carbon was usually above 873 K: no peak was observed above 873 K, indicating the absence of graphitic carbon. [50,51]. The amount of carbon deposit formed during the reaction followed the order of 12.5 wt% Mo/LDO (21.7%) > 22.3 wt% Mo/LDO (17.4%) > 30.1 wt% Mo/LDO (11.1%). According to the results of Raman and UV–vis, isolated [MoO4]2− dominated on the surface of the 12.5 wt% Mo/LDO, which produced the largest amount of carbon deposition during the reaction, indicating that the catalyst with the isolated [MoO4]2− species may promote the non-selective pathway. As Mo content increased, the Mo species transformed from isolated [MoO4]2− to [Mo7O24]6− and [Mo8O26]4− on the surface of the 22.3 wt% and 30.1wt% Mo/LDO, resulting in the gradual disappearance of isolated [MoO4]2−. Meanwhile, the formation of coke reduced as the Mo loading elevated, manifesting [Mo7O24]6− and [Mo8O26]4− that promoted the selective pathway in the reaction process.
The valence state change of the surface Mo species was investigated by XPS, and the spectra are shown in Figure 8. As shown in Figure 8a, all spectra of the as-prepared Mo/LDO catalysts displayed a binding energy peak at 232.6 eV with a satellite peak at 235.9 eV. This doublet was associated with Mo (VI), corresponding to Mo 3d5/2 and 3d3/2 peaks, respectively, suggesting that all Mo/LDO catalysts were fully oxidized and the existence of Mo VI species on the surface [52]. The peaks of Mo 3d broadened with the elevation of Mo content, manifesting the presence of more than one type of Mo (VI) species with different chemical characteristics [53]. Meanwhile, the intensity of both Mo 3d peaks increased with the increase of the Mo/(Mg + Al) ratio (as shown in Table 2), indicating the increase of Mo species on the surface of the samples.

2.2. Reaction Mechanism Analysis

After the reaction, part of the Mo (VI) species was reduced into lower states and new peaks showed up (Figure 8b). The Mo 3d5/2 binding energy peak appearing at 230.2 eV was attributed to Mo (IV) and the peak at 231.2 eV was related to Mo (V) [54,55]. The concentrations of Mo (VI), Mo (V), and Mo (IV) were 34.91%, 26.87%, and 38.22%, respectively. The Mo (VI) species on the surface of fresh catalyst would be partially transformed into low-valance species under the reaction conditions.
Moreover, as shown in Figure 8c, the residual carbon on the as-prepared catalyst might originate from the carbon-containing compounds during sample preparation, resulting in the existence of peaks in the fresh 22.3 wt% Mo/LDO [56]. The peak at 289.2 eV was attributed to the C=O bond in CO32−, which may formed due to the adsorbed carbon dioxide [57,58]. The peak located at 286.3 eV was ascribed to the function group of the C-O bond in hydrocarbon [59,60]. The higher intensity of the C-O bond after the reaction also indicated the formation of hydrocarbon. The peaks located at 285 and 284.5 eV were related to the sp3 C-C bond in amorphous carbon and the sp2 C-C bond in graphitic carbon; both peaks increased significantly after the reaction, demonstrating the formation of carbon deposition after the CO2-ODHE reaction [61,62].
According to the previous literature, the mechanism of the Mo-based catalysts in the CO2-ODHE reaction (as shown in Figure 9) was the redox mechanism [6]. Ethane adsorbed by interacting with the lattice oxygen sites of the dispersed MoOx; subsequently, the adjacent lattice oxygen achieved C-H bond rupture by extracting one of the hydrogens in ethane. After the initial C-H bond cleavage, the fragment -C2H5 eliminated the β-H to form ethylene. The leaving hydroxyl groups at the Mo center recombined and desorbed into H2O, leaving an oxygen site. CO2 was adsorbed as an oxidant at the adjacent O site to generate carbonate intermediates, and the O=C=O bond was broken, where O was added into the oxygen site of the reduced Mo species. The remaining CO* continued to adsorb on the original O site. Finally, the CO dissociated and the Mo site returned to its original state to complete the cycle.

3. Discussion

Figure 10 and Table 3 exhibit the catalytic performance of the Mo/LDO catalysts. The reactivity of the LDO was consistent with the blank test, manifesting that the LDO was not the active species for the reaction. As expected, the 22.3 wt% Mo/LDO delivered an excellent selectivity to ethylene (92.3%) at an approximately 7.9% ethane conversion in the stable stage. All the curves of ethane conversion exhibited a consistent trend that decreased first and stabilized subsequently. The labile stage was attributed to the coke deposition on the catalyst which primarily took place in the initial time-on-stream [63].
The initial formation rate of ethylene exhibited the same trend with the increase of Mo loading. The 22.3 wt% Mo/LDO catalyst reached the highest initial rate of ethylene formation, 15.4 µmol·h−1·g−1cat, while the ethylene formation rate of the 12.5 wt% Mo/LDO was slightly higher than that of the 30.1 wt% Mo/LDO catalyst, 10.6 and 10.6 µmol·h−1·g−1cat, respectively, indicating [MoO4]2− possessed a high activity for ethane activation. However, Raman and UV–vis revealed that isolated [MoO4]2 dominated on the surface of the fresh 12.5 wt% Mo/LDO catalyst, the C2H4 selectivity was only 83% in the initial stage, and side reactions were prone to occur. It has been reported that it is easier for the tetrahedral species to extract oxygen from the terminal Mo=O bonds so that ethane is over-oxidized to form carbon monoxide [64]. Therefore, the catalyst with the isolated [MoO4]2 species is prone to promote the non-selective pathway. With the increase of Mo content, the Mo species on the 22.3 wt% Mo/LDO catalyst transformed from [MoO4]2 to [Mo7O24]6−, while the C2H4 selectivity elevated from 83.0% to 92.5% in the initial stage, indicating [Mo7O24]6− promoted the selective pathway in the reaction process. With the formation of the polymerized molybdate species, [Mo8O26]4−, the initial formation rate of ethylene decreased significantly. The new species in a highly polymeric state reduced the ability to activate ethane and resulted in the reduction of the ethylene formation rate.
As shown in Figure 10c, the curve of the 12.5 wt% Mo/LDO in ethylene selectivity increased first and then reached a plateau. A significant change in selectivity could be observed during the initial period of the reaction due to the structural transformation, according to the previous literature [6]. The initial polymerization of Mo led to the formation of the polymer Mo oxo species, resulting in the improvement of the ethylene selectivity.
The variety of reaction performances indicated that [MoO4]2 was the most active species, whereas it was prone to promote the non-selective oxidation of ethane. The transformation in the active species increased the selectivity to ethylene because of the enhancement in the M=O bond. However, the reaction performance decreased significantly with the formation of [Mo8O26]4−, indicating it was not conducive to the reaction. It could be deduced that a suitable proportion of [MoO4]2− and [Mo7O24]6− species over the surface of the catalyst enabled it to maintain a high selectivity even under a high conversion.

4. Materials and Methods

4.1. Catalyst Preparation

All the materials were used without further purification. Mg(NO3)2·6H2O, Al(NO3)3·9H2O, NaOH, Na2CO3, and CaMoO4 were purchased from Macklin Biochemical Co., Ltd. (Shanghai, China). Para-toluene sulfonic acid (pTOS) and HNO3 were purchased from Aladdin Reagent Co., Ltd. (Shanghai, China).

4.1.1. Preparation of LDH

The hydrotalcite was synthesized by the co-precipitation method, which was reported with some modifications [12]. Briefly, the precursors Mg(NO3)2·6H2O and Al(NO3)3·9H2O were mingled in deionized water with a proper Mg/Al molar ratio. The basic solution consisting of NaOH and Na2CO3 was prepared to adjust the pH of the mixture. Subsequently, the solutions mentioned above were added into a flask by dropwise addition. Simultaneously, the pH value of the mixture solution was maintained at around 9.5 by controlling the titration rate of both solutions. The formative precipitates were aged in their mother liquor at 338 K for 12 h. Finally, the precipitates were filtered, rinsed with distilled water until the pH value was 7, and dried at 353 K overnight.

4.1.2. Preparation of Mo/LDO

The Mo/LDO catalysts were prepared by the anion exchange method [15]. First, Mg-Al/LDH was presented to the anion exchange at pH 4.5 (controlling the pH with HNO3) with pTOS (an organic swelling agent). After 1 h, the solution containing CaMoO4 with a molar ratio of Mg: Al: Mo: pTOS = 3: 1: x: 0.79 was dropped into the reaction mixture at a speed of 1 mL/min. The pH of the mixture was adjusted to 4.5, and the forming slurry was stirred for 24 h. Subsequently, the pH of the mixture was modulated to 10 by dropping the solution of NaOH. All the samples were aged at 338 K for 18 h, then filtered and washed with distilled water to pH = 7. The precipitation was dried at 353 K for 12 h and calcined in a muffle furnace at 873 K for 2 h to obtain the final catalyst.

4.2. Catalyst Characterizations

XRD analysis of the powdered samples was conducted on a D/220-PC diffractometer (Rigaku Corporation, Tokyo, Japan) with a Cu Kα radiation (λ = 0.15406 nm), which operated at 40 kV and 30 mA with a scan rate of 5°·min−1, in the range of 2θ from 10° to 75°. The Brunauer–Emmett–Teller (SBET), total pore volume (Vtotal) and Barrett–Joyner–Halenda pore diameter (DBJH) were measured on an ASAP-2460 analyzer (Micromeritics, Norcross, GA, USA). Each sample was evacuated at 150 °C for 5 h for outgassing before the surface area measurement. The morphology and structure of the catalysts were determined by scanning electron microscopy (Carl Zeiss SIGMA, Carl Zeiss AG, Oberkochen, Germany) equipped with elemental analysis and a transmission electron microscope (FEI Tecnai G2 F20, FEI Company, Hillsboro, OR, USA). Laser Raman spectra were collected by a LabRAM HR Evolution (HORIBA Scientific, Kyoto, Japan) with a 532 nm laser at ranges of 600−1500 cm−1. Diffuse reflectance UV–vis spectra were measured by a UV-3600 UV/vis spectrophotometer (Shimadzu Corporation, Kyoto, Japan). The spectra of the catalysts were recorded in the wavelength range of 200–600 nm using BaSO4 as the standard. The thermogravimetric-differential thermal analysis was measured by a HITACHI STA 7300 (Hitachi, Kyoto, Japan) to determine the formation of carbon deposition during the reaction. X-ray photoelectron spectroscopy (Thermo Scientific K-Alpha, Thermo Fisher Scientific, Waltham, MA, USA) was carried out to measure the surface elemental variation of the as-prepared Mo/LDO and used 22.3 wt% Mo/LDO.

4.3. Catalyst Test

The catalyst testing of the Mo/LDO catalysts were carried out in a continuous fixed-bed flow reactor (Yuanbang Electric Furnace Manufacturing Co., Longkou, China) at atmospheric pressure. Typically, 300 mg catalysts were put in the quartz tube for the catalyst test. Under N2 atmosphere (15 mL/min), the temperature of the reactor elevated to 873 K with a speed of 10 K/min. Then, the feedstock C2H6, CO2, and N2 (as a diluent gas) was introduced into the reactor with the ratio of 1: 1: 2 and maintained for 5 h. Utilizing a gas chromatograph (GC-2000, Shanghai Ramiin Instrument Co., Shanghai, China) equipped with a flame ionization detector (FID) detector, the feed and products were analyzed.
The conversion of ethane, the selectivity and yield of ethylene, and the initial formation rate of ethylene were calculated according to the following formula:
C 2 H 6   Conversion % = n C 2 H 6 , in n C 2 H 6 , out n C 2 H 6 , in   ×   100
C 2 H 4   Selectivity ( % ) = n C 2 H 4 , out n C 2 H 6 , in n C 2 H 6 , out   ×   100
C 2 H 4   Yield ( % ) = C 2 H 6   Conversion ( % )   ×   C 2 H 4   Selectivity ( % )
Initial   formation   rate   of   ethylene   ( μ mol · h 1 · g cat 1 ) = n C 2 H 6 ( in ,   In   Initial   Strage ) m cat ×   C 2 H 6   Conversion 100   ×   C 2 H 4   Selectivity 100
Carbon   balance = C in C out C in
where (C) denotes the number of moles of carbon. The maximum total carbon balance at the end of the run did not exceed 5%.

5. Conclusions

A series of LDO-supported molybdenum oxide catalysts were synthesized and characterized by XRD, BET, SEM, TEM, Raman, UV–vis, TG, and XPS. The catalytic performance was tested in the CO2-ODHE reaction. The 12.5 wt% Mo/LDO catalyst with isolated [MoO4]2− on the surface exhibited only a 4.4% ethane conversion and produced the largest amount of carbon deposition during the reaction. As the Mo content increased, the Mo species transformed from isolated [MoO4]2− to a polymerized molybdate species, [Mo7O24]6−, on the surface of the 22.3 wt% Mo/LDO, delivering an excellent selectivity to ethylene (92.3%) with a 7.9% ethane conversion. Due to the higher aggregation degree of Mo species on the 30.1 wt% Mo/LDO ([Mo8O26]4−), a decreased ethane conversion was found. The CO2-ODHE reaction over Mo/LDO catalysts in this work is supposed to follow the redox mechanism. The surface Mo species and their aggregation state could determine the reaction route. Isolated [MoO4]2− preferred the non-selective pathway, [Mo7O24]6− promoted the selective pathway, and the existence of [Mo8O26]4− reduced the ability to activate ethane.

Author Contributions

Conceptualization, S.L.; Data curation, G.S., Q.W., L.Y. and D.L.; funding acquisition, S.L.; investigation, G.S.; methodology, G.S. and Q.W.; project administration, S.L.; writing—original draft, G.S.; writing—review and editing, S.L., L.Y. and D.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the National Natural Science Foundation of China (21878244).

Data Availability Statement

Data are available from the corresponding author upon request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pan, Y.; Bhowmick, A.; Wu, W.; Zhang, Y.; Diao, Y.; Zheng, A.; Zhang, C.; Xie, R.; Liu, Z.; Meng, J. Titanium Silicalite-1 Nanosheet-Supported Platinum for Non-oxidative Ethane Dehydrogenation. ACS Catal. 2021, 11, 9970–9985. [Google Scholar] [CrossRef]
  2. Zhang, B.; Shan, B.; Zhao, Y.; Zhang, L. Review of Formation and Gas Characteristics in Shale Gas Reservoirs. Energies 2020, 13, 5427. [Google Scholar] [CrossRef]
  3. Dar, H.J.; Jakobsen, H.A.; Rout, K.R.; Jens, K.J.; Chen, D. Autothermal Gas-Phase Oxidative Dehydrogenation of Ethane to Ethylene at Atmospheric Pressure. Ind. Eng. Chem. Res. 2021, 60, 6784–6802. [Google Scholar] [CrossRef]
  4. Zhou, Y.; Lin, J.; Li, L.; Tian, M.; Li, X.; Pan, X.; Chen, Y.; Wang, X. Improving the selectivity of Ni-Al mixed oxides with isolated oxygen species for oxidative dehydrogenation of ethane with nitrous oxide. J. Catal. 2019, 377, 438–448. [Google Scholar] [CrossRef]
  5. Xie, Z.; Tian, D.; Xie, M.; Yang, S.-Z.; Xu, Y.; Rui, N.; Lee, J.H.; Senanayake, S.D.; Li, K.; Wang, H. Interfacial Active Sites for CO2 Assisted Selective Cleavage of C–C/C–H Bonds in Ethane. Chem 2020, 6, 1–14. [Google Scholar] [CrossRef]
  6. Yao, R.; Herrera, J.E.; Chen, L.; Chin, Y.-H.C. Generalized Mechanistic Framework for Ethane Dehydrogenation and Oxidative Dehydrogenation on Molybdenum Oxide Catalysts. ACS Catal. 2020, 10, 6952–6968. [Google Scholar] [CrossRef]
  7. Abello, M.C.; Gomez, M.F.; Ferretti, O. Mo/γ-Al2O3 catalysts for the oxidative dehydrogenation of propane.: Effect of Mo loading. Appl. Catal. A 2001, 207, 421–431. [Google Scholar] [CrossRef]
  8. Woods, M.P.; Mirkelamoglu, B.; Ozkan, U.S. Oxygen and Nitrous Oxide as Oxidants: Implications for Ethane Oxidative Dehydrogenation over SilicaTitania-Supported Molybdenum. J. Phys. Chem. C 2009, 113, 10112–10119. [Google Scholar] [CrossRef]
  9. Christodoulakis, A.; Heracleous, E.; Lemonidou, A.; Boghosian, S. An operando Raman study of structure and reactivity of alumina-supported molybdenum oxide catalysts for the oxidative dehydrogenation of ethane. J. Catal. 2006, 242, 16–25. [Google Scholar] [CrossRef]
  10. Tsilomelekis, G.; Boghosian, S. An operando Raman study of molecular structure and reactivity of molybdenum(VI) oxide supported on anatase for the oxidative dehydrogenation of ethane. Phys. Chem. Chem. Phys. 2012, 14, 2216–2228. [Google Scholar] [CrossRef]
  11. Pooresmaeil, M.; Behzadi Nia, S.; Namazi, H. Green encapsulation of LDH(Zn/Al)-5-Fu with carboxymethyl cellulose biopolymer; new nanovehicle for oral colorectal cancer treatment. Int. J. Biol. Macromol. 2019, 139, 994–1001. [Google Scholar] [CrossRef] [PubMed]
  12. Ahmadi-Kashani, M.; Dehghani, H.; Zarrabi, A. A biocompatible nanoplatform formed by MgAl-layered double hydroxide modified Mn3O4/N-graphene quantum dot conjugated-polyaniline for pH-triggered release of doxorubicin. Mat. Sci. Eng. C-Mater. 2020, 114, 111055. [Google Scholar] [CrossRef] [PubMed]
  13. Lu, Y.; Zhang, Z.; Xu, Y.; Liu, Q.; Qian, G. CaFeAl mixed oxide derived heterogeneous catalysts for transesterification of soybean oil to biodiesel. Bioresour. Technol. 2015, 190, 438–441. [Google Scholar] [CrossRef] [PubMed]
  14. Hou, B.; Du, Y.; Liu, X.; Ci, C.; Wu, X.; Xie, X. Tunable preparation of highly dispersed NixMn-LDO catalysts derived from NixMn-LDHs precursors and application in low-temperature NH3-SCR reactions. RSC Adv. 2019, 9, 24377–24385. [Google Scholar] [CrossRef] [Green Version]
  15. Li, P.; Yu, Y.; Huang, P.-P.; Liu, H.; Cao, C.-Y.; Song, W.-G. Core–shell structured MgAl-LDO@Al-MS hexagonal nanocomposite: An all inorganic acid–base bifunctional nanoreactor for one-pot cascade reactions. J. Mater. Chem. A 2014, 2, 339–344. [Google Scholar] [CrossRef]
  16. Zhang, Q.; Zhang, K.; Zhang, S.; Liu, Q.; Chen, L.; Li, X.; Wang, C.; Ma, L. Ga3+-stabilized Pt in PtSn-Mg(Ga)(Al)O catalyst for promoting ethane dehydrogenation. J. Catal. 2018, 368, 79–88. [Google Scholar] [CrossRef]
  17. Chen, Q.; Wang, J.; Li, F. Formation of Carbon Nanofibers from Supported Pt Catalysts through Catalytic Chemical Vapor Deposition from Acetylene. Ind. Eng. Chem. Res. 2011, 50, 9034–9042. [Google Scholar] [CrossRef]
  18. Ansari, A.A.; Parchur, A.K.; Alam, M.; Azzeer, A. Structural and photoluminescence properties of Tb-doped CaMoO4 nanoparticles with sequential surface coatings. Mater. Chem. Phys. 2014, 147, 715–721. [Google Scholar] [CrossRef]
  19. Ramakrishna, P.V. Cathodoluminescence properties of gadolinium-doped CaMoO4:Eu nanoparticles. Spectrochim. Acta. A Mol. Biomol. Spectrosc. 2015, 149, 312–316. [Google Scholar] [CrossRef]
  20. Liu, X.; Fan, B.; Gao, S.; Li, R. Transesterification of tributyrin with methanol over MgAl mixed oxides derived from MgAl hydrotalcites synthesized in the presence of glucose. Fuel Process. Technol. 2013, 106, 761–768. [Google Scholar] [CrossRef]
  21. Yuan, X.; Niu, J.; Lv, Y.; Jing, Q.; Li, L. Ultrahigh-capacity and fast-rate removal of graphene oxide by calcined MgAl layered double hydroxide. Appl. Clay Sci. 2018, 156, 61–68. [Google Scholar] [CrossRef]
  22. Deng, S.; Li, S.; Li, H.; Zhang, Y.; Deng, S. Oxidative Dehydrogenation of Ethane to Ethylene with CO2 over Fe-Cr/ZrO2 Catalysts. Ind. Eng. Chem. Res. 2009, 16, 7561–7566. [Google Scholar] [CrossRef]
  23. Solsona, B.; Dejoz, A.; Garcia, T.; Concepción, P.; Nieto, J.; Vázquez, M.; Navarro, M.T. Molybdenum–vanadium supported on mesoporous alumina catalysts for the oxidative dehydrogenation of ethane. Catal. Today 2006, 117, 228–233. [Google Scholar] [CrossRef]
  24. Wu, X.; Feng, Y.; Du, Y.; Liu, X.; Zou, C.; Li, Z. Enhancing DeNOx performance of CoMnAl mixed metal oxides in low-temperature NH3-SCR by optimizing layered double hydroxides (LDHs) precursor template. Appl. Surf. Sci. 2019, 467–468, 802–810. [Google Scholar] [CrossRef]
  25. Mohd Sidek, H.B.; Jo, Y.K.; Kim, I.Y.; Hwang, S.-J. Stabilization of Layered Double Oxide in Hybrid Matrix of Graphene and Layered Metal Oxide Nanosheets: An Effective Way To Explore Efficient CO2 Adsorbent. J. Phys. Chem. C 2016, 120, 23421–23429. [Google Scholar] [CrossRef]
  26. Li, X.; Tian, J.; Liu, H.; Hu, X.; Zhang, J.; Xia, C.; Chen, J.; Liu, H.; Huang, Z. Efficient Synthesis of 5-Amino-1-pentanol from Biomass-Derived Dihydropyran over Hydrotalcite-Based Ni–Mg3AlOx Catalysts. ACS Sust. Chem. Eng. 2020, 8, 6352–6362. [Google Scholar] [CrossRef]
  27. Tao, Q.; Zhu, J.; Frost, R.L.; Bostrom, T.E.; Wellard, R.M.; Wei, J.; Yuan, P.; He, H. Silylation of layered double hydroxides via a calcination-rehydration route. Langmuir 2010, 26, 2769–2773. [Google Scholar] [CrossRef]
  28. Cao, X.; Zhou, J.; Zhai, Z.; Li, S.; Yuan, G.; Qin, G. Synchronous Growth of Porous MgO and Half-Embedded Nano-Ru on a Mg Plate: A Monolithic Catalyst for Fast Hydrogen Production. ACS Sust. Chem. Eng. 2021, 9, 3616–3623. [Google Scholar] [CrossRef]
  29. Bhagwan, J.; Hussain, S.K.; Yu, J.S. Facile Hydrothermal Synthesis and Electrochemical Properties of CaMoO4 Nanoparticles for Aqueous Asymmetric Supercapacitors. ACS Sust. Chem. Eng. 2019, 14, 12340–12350. [Google Scholar] [CrossRef]
  30. Liu, L.-J.; Wang, Z.-M.; Lyu, Y.-J.; Zhang, J.-F.; Huang, Z.; Qi, T.; Si, Z.-B.; Yang, H.-Q.; Hu, C.-W. Catalytic mechanisms of oxygen-containing groups over vanadium active sites in an Al-MCM-41 framework for production of 2,5-diformylfuran from 5-hydroxymethylfurfural. Catal. Sci. Technol. 2020, 10, 278–290. [Google Scholar] [CrossRef]
  31. Xiong, G.; Feng, Z.; Li, J.; Yang, Q.; Ying, P.; Xin, Q.; Li, C. UV Resonance Raman Spectroscopic Studies on the Genesis of Highly Dispersed Surface Molybdate Species on γ-Alumina. J. Phys. Chem. B 2000, 104, 3581–3588. [Google Scholar] [CrossRef]
  32. Ji, Z.; Lv, H.; Pan, X.; Bao, X. Enhanced ethylene selectivity and stability of Mo/ZSM-5 upon modification with phosphorus in ethane dehydrogenation. J. Catal. 2018, 361, 94–104. [Google Scholar] [CrossRef]
  33. Che, S.; Sakamoto, Y.; Yoshitake, H.; Terasaki, O.; Tatsumi, T. Synthesis and Characterization of Mo/SBA-1 Cubic Mesoporous Molecular Sieves. J. Phys. Chem. B 2001, 14, 2433. [Google Scholar] [CrossRef]
  34. Zhang, L.; Chen, Z.; Zheng, S.; Cai, G.; Fu, W.; Tang, T.; He, M. Effect of the Co/Mo Ratio on the Morphology and Activity of the CoMo Catalyst Supported on MgO Nanosheets in Dibenzothiophene Hydrodesulfurization. Ind. Eng. Chem. Res. 2020, 59, 12338–12351. [Google Scholar] [CrossRef]
  35. Kong, L.; Li, J.; Zhao, Z.; Liu, Q.; Sun, Q.; Liu, J.; Wei, Y. Oxidative dehydrogenation of ethane to ethylene over Mo-incorporated mesoporous SBA-16 catalysts: The effect of MoO dispersion. Appl. Catal. A 2016, 510, 84–97. [Google Scholar] [CrossRef]
  36. Weber, R.S. Effect of Local Structure on the UV-Visible Absorption Edges of Molybdenum Oxide Clusters and Supported Molybdenum Oxides. J. Catal. 1995, 151, 470–474. [Google Scholar] [CrossRef]
  37. Nair, H.; Liszka, M.J.; Gatt, J.E.; Baertsch, C.D. Effects of Metal Oxide Domain Size, Dispersion, and Interaction in Mixed WOx/MoOx Catalysts Supported on Al2O3 for the Partial Oxidation of Ethanol to Acetaldehyde. J. Phys. Chem. C 2008, 112, 1612–1620. [Google Scholar] [CrossRef]
  38. Lee, S.; Zoican Loebick, C.; Pfefferle, L.D.; Haller, G.L. High-Temperature Stability of Cobalt Grafted on Low-Loading Incorporated Mo/MCM-41 Catalyst for Synthesis of Single-Walled Carbon Nanotubes. J. Phys. Chem. C 2011, 115, 1014–1024. [Google Scholar] [CrossRef]
  39. Tian, H.; Roberts, C.A.; Wachs, I.E. Molecular Structural Determination of Molybdena in Different Environments: Aqueous Solutions, Bulk Mixed Oxides, and Supported MoO3 Catalysts. J. Phys. Chem. C 2010, 114, 14110–14120. [Google Scholar] [CrossRef]
  40. Herrmann, J.M.; Villain, F.; Appel, L.G. Characterization of Mo–Sn–O system by means of Raman spectroscopy and electrical conductivity measurements. Appl. Catal. A 2003, 240, 177–182. [Google Scholar] [CrossRef]
  41. Wachs, I.E.; Roberts, C.A. Monitoring surface metal oxide catalytic active sites with Raman spectroscopy. Chem. Soc. Rev. 2010, 39, 5002–5017. [Google Scholar] [CrossRef]
  42. Savenko, D.Y.; Velieva, N.Y.; Svetlichnyi, V.A.; Vodyankina, O.V. The influence of the preparation method on catalytic properties of Mo–Fe–O/SiO2 catalysts in selective oxidation of 1,2-propanediol. Catal. Today 2020, 357, 399–408. [Google Scholar] [CrossRef]
  43. Gibson, E.K.; Zandbergen, M.W.; Jacques, S.D.M.; Biao, C.; Cernik, R.J.; O’Brien, M.G.; Di Michiel, M.; Weckhuysen, B.M.; Beale, A.M. Noninvasive Spatiotemporal Profiling of the Processes of Impregnation and Drying within Mo/Al2O3 Catalyst Bodies by a Combination of X-ray Absorption Tomography and Diagonal Offset Raman Spectroscopy. ACS Catal. 2013, 3, 339–347. [Google Scholar] [CrossRef]
  44. Bugrova, T.A.; Dutov, V.V.; Svetlichnyi, V.A.; Cortés Corberán, V.; Mamontov, G.V. Oxidative dehydrogenation of ethane with CO2 over CrOx catalysts supported on Al2O3, ZrO2, CeO2 and CexZr1-xO2. Catal. Today 2019, 333, 71–80. [Google Scholar] [CrossRef]
  45. Ou, J.Z.; Campbell, J.L.; Yao, D.; Wlodarski, W.; Kalantar-zadeh, K. In Situ Raman Spectroscopy of H2 Gas Interaction with Layered MoO3. J. Phys. Chem. C 2011, 115, 10757–10763. [Google Scholar] [CrossRef]
  46. Xi, Y.; Xiao, J.; Lin, X.; Yan, W.; Wang, C.; Liu, C. SiO2-Modified Pt/Al2O3 for Oxidative Dehydrogenation of Ethane: A Preparation Method for Improved Catalytic Stability, Ethylene Selectivity, and Coking Resistance. Ind. Eng. Chem. Res. 2018, 57, 10137–10147. [Google Scholar] [CrossRef]
  47. Fu, T.; Wang, Y.; Li, Z. Surface-Protection-Induced Controllable Restructuring of Pores and Acid Sites of the Nano-ZSM-5 Catalyst and Its Influence on the Catalytic Conversion of Methanol to Hydrocarbons. Langmuir 2020, 36, 3737–3749. [Google Scholar] [CrossRef]
  48. Shah, M.; Bordoloi, A.; Nayak, A.K.; Mondal, P. Experimental and Kinetic Studies of Methane Reforming with CO2 over a La-Doped Ni/Al2O3 Bimodal Catalyst. Ind. Eng. Chem. Res. 2021, 60, 18243–18255. [Google Scholar] [CrossRef]
  49. Liu, C.; He, Y.; Wei, L.; Zhang, Y.; Zhao, Y.; Hong, J.; Chen, S.; Wang, L.; Li, J. Hydrothermal Carbon-Coated TiO2 as Support for Co-Based Catalyst in Fischer–Tropsch Synthesis. ACS Catal. 2018, 8, 1591–1600. [Google Scholar] [CrossRef]
  50. Xiao, Z.; Hou, F.; Zhang, J.; Zheng, Q.; Xu, J.; Pan, L.; Wang, L.; Zou, J.; Zhang, X.; Li, G. Methane Dry Reforming by Ni-Cu Nanoalloys Anchored on Periclase-Phase MgAlOx Nanosheets for Enhanced Syngas Production. ACS Appl. Mater. Inter. 2021, 13, 48838–48854. [Google Scholar] [CrossRef]
  51. Wang, N.; Shen, K.; Huang, L.; Yu, X.; Qian, W.; Chu, W. Facile Route for Synthesizing Ordered Mesoporous Ni–Ce–Al Oxide Materials and Their Catalytic Performance for Methane Dry Reforming to Hydrogen and Syngas. ACS Catal. 2013, 3, 1638–1651. [Google Scholar] [CrossRef]
  52. Vo, T.K.; Kim, W.-S.; Kim, S.-S.; Yoo, K.S.; Kim, J. Facile synthesis of Mo/Al2O3–TiO2 catalysts using spray pyrolysis and their catalytic activity for hydrodeoxygenation. Energ Convers. Manag. 2018, 158, 92–102. [Google Scholar] [CrossRef]
  53. Lede, E.J.; Requejo, F.G.; Pawelec, B.; Fierro, J. XANES Mo L-edges and XPS study of Mo loaded in HY zeolite. J. Phys. Chem. B 2002, 106, 7824–7831. [Google Scholar] [CrossRef]
  54. Song, Y.; Sun, C.; Shen, W.; Lin, L. Hydrothermal post-synthesis of HZSM-5 zeolite to enhance the coke-resistance of Mo/HZSM-5 catalyst for methane dehydroaromatization reaction: Reconstruction of pore structure and modification of acidity. Appl. Catal. A 2007, 317, 266–274. [Google Scholar] [CrossRef]
  55. Julian, I.; Roedern, M.B.; Hueso, J.L.; Irusta, S.; Baden, A.K.; Mallada, R.; Davis, Z.; Santamaria, J. Supercritical solvothermal synthesis under reducing conditions to increase stability and durability of Mo/ZSM-5 catalysts in methane dehydroaromatization. Appl. Catal. B 2020, 263, 118360. [Google Scholar] [CrossRef]
  56. Liu, Y.; Shu, W.; Chen, K.; Peng, Z.; Chen, W. Enhanced Photothermocatalytic Synergetic Activity Toward Gaseous Benzene for Mo+C-Codoped Titanate Nanobelts. ACS Catal. 2012, 2, 2557–2565. [Google Scholar] [CrossRef]
  57. Liu, B.S.; Chang, R.Z.; Jiang, L.; Liu, W.; Au, C.T. Preparation and High Performance of La2O3−V2O5/MCM-41 Catalysts for Ethylbenzene Dehydrogenation in the Presence of CO2. J. Phys. Chem. C 2008, 112, 15490–15501. [Google Scholar] [CrossRef]
  58. Elkins, T.W.; Neumann, B.; Bäumer, M.; Hagelin-Weaver, H.E. Effects of Li Doping on MgO-Supported Sm2O3 and TbOx Catalysts in the Oxidative Coupling of Methane. ACS Catal. 2014, 4, 1972–1990. [Google Scholar] [CrossRef]
  59. Shu, J.; Adnot, A.; Grandjean, B. Bifunctional Behavior of Mo/HZSM-5 Catalysts in Methane Aromatization. Ind. Eng. Chem. Res. 1999, 38, 3860–3867. [Google Scholar] [CrossRef]
  60. Yang, J.; Zhao, Y.; Chang, L.; Zhang, J.; Zheng, C. Mercury Adsorption and Oxidation over Cobalt Oxide Loaded Magnetospheres Catalyst from Fly Ash in Oxyfuel Combustion Flue Gas. Environ. Sci. Technol. 2015, 49, 8210–8218. [Google Scholar] [CrossRef]
  61. Jiménez-González, C.; Gil-Calvo, M.; de Rivas, B.; González-Velasco, J.R.; Gutiérrez-Ortiz, J.I.; López-Fonseca, R. Oxidative Steam Reforming and Steam Reforming of Methane, Isooctane, and N-Tetradecane over an Alumina Supported Spinel-Derived Nickel Catalyst. Ind. Eng. Chem. Res. 2016, 55, 3920–3929. [Google Scholar] [CrossRef]
  62. Zhang, M.; Tan, X.; Zhang, T.; Han, Z.; Jiang, H. The deactivation of a ZnO doped ZrO2–SiO2 catalyst in the conversion of ethanol/acetaldehyde to 1,3-butadiene. RSC Adv. 2018, 8, 34069–34077. [Google Scholar] [CrossRef] [Green Version]
  63. Koirala, R.; Safonova, O.V.; Pratsinis, S.E.; Baiker, A. Effect of cobalt loading on structure and catalytic behavior of CoOx/SiO2 in CO2-assisted dehydrogenation of ethane. Appl. Catal. A 2018, 552, 77–85. [Google Scholar] [CrossRef]
  64. Heracleous, E.; Vakros, J.; Lemonidou, A.A.; Kordulis, C. Role of preparation parameters on the structure–selectivity properties of MoO3/Al2O3 catalysts for the oxidative dehydrogenation of ethane. Catal. Today 2004, 91–92, 289–292. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of (a) fresh Mo/LDO catalysts and (b) used Mo/LDO catalysts.
Figure 1. XRD patterns of (a) fresh Mo/LDO catalysts and (b) used Mo/LDO catalysts.
Catalysts 12 00493 g001
Figure 2. (a) N2 adsorption−desorption isotherms and (b) the corresponding pore size distribution of the as-prepared Mo/LDO catalysts.
Figure 2. (a) N2 adsorption−desorption isotherms and (b) the corresponding pore size distribution of the as-prepared Mo/LDO catalysts.
Catalysts 12 00493 g002
Figure 3. SEM images of fresh catalysts: (a,d) 12.5 wt% Mo/LDO; (b,e) 22.3 wt% Mo/LDO; (c,f) 30.1 wt% Mo/LDO and the elemental mapping: (g) 12.5 wt% Mo/LDO; (h) 22.3 wt% Mo/LDO; (i) 30.1 wt% Mo/LDO.
Figure 3. SEM images of fresh catalysts: (a,d) 12.5 wt% Mo/LDO; (b,e) 22.3 wt% Mo/LDO; (c,f) 30.1 wt% Mo/LDO and the elemental mapping: (g) 12.5 wt% Mo/LDO; (h) 22.3 wt% Mo/LDO; (i) 30.1 wt% Mo/LDO.
Catalysts 12 00493 g003
Figure 4. TEM images of fresh catalysts: (a,d) 12.5 wt% Mo/LDO; (b,e) 22.3 wt% Mo/LDO; (c,f) 30.1 wt% Mo/LDO; and images of (g) MgO and (h) CaMoO4 in 22.3 wt% Mo/LDO.
Figure 4. TEM images of fresh catalysts: (a,d) 12.5 wt% Mo/LDO; (b,e) 22.3 wt% Mo/LDO; (c,f) 30.1 wt% Mo/LDO; and images of (g) MgO and (h) CaMoO4 in 22.3 wt% Mo/LDO.
Catalysts 12 00493 g004
Figure 5. (a) UV–vis diffuse reflectance spectra of Mo/LDO catalysts and (b) Kubelka–Munk plots of the samples.
Figure 5. (a) UV–vis diffuse reflectance spectra of Mo/LDO catalysts and (b) Kubelka–Munk plots of the samples.
Catalysts 12 00493 g005
Figure 6. Raman profiles of the (a) fresh and (b) used Mo/LDO catalysts.
Figure 6. Raman profiles of the (a) fresh and (b) used Mo/LDO catalysts.
Catalysts 12 00493 g006
Figure 7. TG and DTG profiles of used Mo/LDO catalysts.
Figure 7. TG and DTG profiles of used Mo/LDO catalysts.
Catalysts 12 00493 g007
Figure 8. (a) Mo 3d XPS spectra of fresh Mo/LDO catalysts; (b) Mo 3d XPS spectra of used 22.3 wt% Mo/LDO catalyst; (c) C 1s XPS spectra of fresh and used 22.3 wt% Mo/LDO catalyst.
Figure 8. (a) Mo 3d XPS spectra of fresh Mo/LDO catalysts; (b) Mo 3d XPS spectra of used 22.3 wt% Mo/LDO catalyst; (c) C 1s XPS spectra of fresh and used 22.3 wt% Mo/LDO catalyst.
Catalysts 12 00493 g008
Figure 9. The reaction pathway of ODHE over the Mo/LDO catalyst.
Figure 9. The reaction pathway of ODHE over the Mo/LDO catalyst.
Catalysts 12 00493 g009
Figure 10. (a) Ethane conversion in the stable stage and the initial ethylene formation rate; (b) C2H6 conversion and (c) C2H4 selectivity as functions of the time on stream over Mo/LDO catalysts. Reaction conditions: T = 873 K, C2H6/CO2/N2 = 1: 1: 2, the total flow rate = 60 mL min−1.
Figure 10. (a) Ethane conversion in the stable stage and the initial ethylene formation rate; (b) C2H6 conversion and (c) C2H4 selectivity as functions of the time on stream over Mo/LDO catalysts. Reaction conditions: T = 873 K, C2H6/CO2/N2 = 1: 1: 2, the total flow rate = 60 mL min−1.
Catalysts 12 00493 g010
Table 1. Textural properties derived from the different Mo/LDO catalysts.
Table 1. Textural properties derived from the different Mo/LDO catalysts.
SampleSBET
(m2·g−1)
Average Pore Diameter (nm)Total Pore Volume (cm3·g−1)
12.5 wt% Mo/LDO203.85.60.29
22.3 wt% Mo/LDO194.05.90.29
30.1 wt% Mo/LDO171.04.00.17
Table 2. Binding energies (eV) and atomic ratios of Mo/(Mg + Al) determined by XPS.
Table 2. Binding energies (eV) and atomic ratios of Mo/(Mg + Al) determined by XPS.
SampleMo 3d5/2Mo 3d3/2Mo/(Mg + Al) (%)
12.5 wt% Mo/LDO232.6235.98.0
22.3 wt% Mo/LDO232.6235.913.8
30.1 wt% Mo/LDO232.6235.918.3
Table 3. The detail of catalytic performance over the Mo/LDO catalysts.
Table 3. The detail of catalytic performance over the Mo/LDO catalysts.
SampleConversion
(%)
Selectivity
(%)
Yield
(%)
C2H4 Formation Rate,
(µmol·h−1·g−1cat)
12.5 wt% Mo/LDO4.490.54.010.6
22.3 wt% Mo/LDO7.992.37.315.4
30.1 wt% Mo/LDO5.988.95.210.2
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Song, G.; Wang, Q.; Yang, L.; Liao, D.; Li, S. Oxidative Dehydrogenation of Ethane with CO2 over Mo/LDO Catalyst: The Active Species of Mo Controlled by LDO. Catalysts 2022, 12, 493. https://doi.org/10.3390/catal12050493

AMA Style

Song G, Wang Q, Yang L, Liao D, Li S. Oxidative Dehydrogenation of Ethane with CO2 over Mo/LDO Catalyst: The Active Species of Mo Controlled by LDO. Catalysts. 2022; 12(5):493. https://doi.org/10.3390/catal12050493

Chicago/Turabian Style

Song, Gengzhe, Qi Wang, Liang Yang, Duohua Liao, and Shuang Li. 2022. "Oxidative Dehydrogenation of Ethane with CO2 over Mo/LDO Catalyst: The Active Species of Mo Controlled by LDO" Catalysts 12, no. 5: 493. https://doi.org/10.3390/catal12050493

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