Next Article in Journal
Tuning Electrochemical Hydrogen-Evolution Activity of CoMoO4 through Zn Incorporation
Next Article in Special Issue
Hydrogenation of CO2 to Valuable C2-C5 Hydrocarbons on Mn-Promoted High-Surface-Area Iron Catalysts
Previous Article in Journal
Integrated Nanostructures of TiO2/g-C3N4/Diatomite Based on Low-Grade Diatomite as Efficient Catalyst for Photocatalytic Degradation of Methylene Blue: Performance and Mechanism
Previous Article in Special Issue
A One-Pot Hydrothermal Preparation of High Loading Ni/La2O3 Catalyst for Efficient Hydrogenation of Cinnamaldehyde
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Ce Content on the Chemical Looping Oxidative Dehydrogenation of Propane to Propylene over a VOx-CeO2/γ-Al2O3 Oxygen Carrier

1
College of Chemical Engineering, Qingdao University of Science & Technology, Qingdao 266042, China
2
State Key Laboratory of High-Efficiency Coal Utilization and Green Chemical Engineering, College of Chemistry and Chemical Engineering, Ningxia University, Yinchuan 750021, China
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(5), 797; https://doi.org/10.3390/catal13050797
Submission received: 28 March 2023 / Revised: 19 April 2023 / Accepted: 23 April 2023 / Published: 24 April 2023
(This article belongs to the Special Issue Catalytic Transformation of Low-Carbon Resources)

Abstract

:
The chemical looping oxidative dehydrogenation of propane to propylene (CL-ODHP) replaces molecular oxygen with lattice oxygen (Olatt) in oxygen carriers. This method boosts propylene selectivity by avoiding the deep oxidation of propane. Herein, a series of 10V-XCe/Al oxygen carriers with different Ce contents were prepared to realize different VOx-CeOy interactions. The effect of the Ce content in 10V-XCe/Al oxygen carriers on the CL-ODHP reaction was studied and the optimal Ce content was determined. CeO2 prevents the outward diffusion and evolution of Olatt in VOx carriers to the adsorbed electrophilic oxygen species (Oelec), effectively inhibiting the loss of Olatt, improving the selectivity of propylene, and extending the lifetime and activity of the oxygen carriers. After characterizing and analyzing the oxygen carriers, it was found that 10V-3Ce/Al has the highest specific surface area, highest oxygen capacity, and lowest reducibility. The 10V-3Ce/Al also delivers the highest oxidative dehydrogenation performance. At 550 °C, the average propylene and COx selectivity values of 10V-3Ce/Al were 81.87% and 7.28%, respectively (vs. 62.79% and 25.64% respectively, for 10V/Al). It is demonstrated that 10V-3Ce/Al exhibits good cycle stability with no significant decrease in catalytic performance after 15 cycles. In situ diffuse-reflectance infrared Fourier-transform spectroscopy indicates that CL-ODHP on 10V-3Ce/Al undergoes the Mars-van Krevelen mechanism. The migration and evolution of Olatt in oxygen carriers is controlled by reasonably modifying the metal oxide interactions to improve propylene yield. This work will thus guide the subsequent development of novel and efficient CL-ODHP oxygen carriers.

Graphical Abstract

1. Introduction

Propylene is an important raw material for petrochemicals, and is mainly used to produce chemical materials such as polypropylene, propylene oxide, and acrylonitrile. The demand for propylene has increased with development in the downstream market. The current development in the shale gas industry has increased the availability of low-cost propane, making propylene production from propane appealing [1,2].
The industrial process, direct dehydrogenation of propane to propylene (DHP), is a strong heat-absorption process with restricted thermodynamic equilibrium. Effective propane conversion requires a high temperature, which increases the risk of problems such as catalyst carbon build-up. These inherent drawbacks limit the performance of catalytic dehydrogenation of propane [3,4]. In contrast, oxidative dehydrogenation of propane to propylene (ODHP) is an exothermic reaction not limited by thermodynamics. As ODHP can theoretically achieve high conversion rates at lower temperatures than DHP without carbon accumulation, it has attracted extensive research interest. However, in this process, O2 deeply oxidizes propane to COx, which reduces propylene production. The widespread adoption of ODHP is further limited by the requirement for expensive air-separation equipment. The safety of co-feeding propane and O2 raises additional concerns [5,6].
The shortcomings of propylene production technologies can be overcome by a new propylene production route called chemical looping oxidative dehydrogenation of propane to propylene (CL-ODHP), which includes oxidation and reduction reactions using oxygen carriers in two reactors. Propane is oxidized to propylene and water by lattice oxygens (Olatt) in the oxygen carrier. The water is separated from the reaction system via condensation, driving the reaction equilibrium in the positive direction and thus increasing the propylene yield. The reduced oxygen carrier enters the air reactor, where it is oxidized and regenerated before entering the dehydrogenation reactor to complete the oxidation and dehydrogenation cycles [7,8]. The CL-ODHP process confers the following advantages over other propylene production processes: (a) the oxygen carrier avoids direct contact between propane and O2, thereby eliminating explosion risks; (b) the process does not require an air separation device, reducing the investment and operating costs; (c) the reaction occurs without O2, improving the operability and temperature control; and (d) hydrogen combustion is indirect and flameless, thus reducing COx emissions.
However, the CL-ODHP process requires suitable oxygen carriers. Metal oxides have been reported as suitable oxygen carriers for chemical looping processes. Liu et al. [9] found that polymetallic oxide oxygen carriers have high CL-ODHP activity, and by regulating the ratio between metal oxides, the polymetallic oxide oxygen carriers can be made to release Olatt slowly, thus improving the selectivity of low-carbon olefins. After studying the CL-ODHP properties of different metal oxides (MoOx/γ-Al2O3, GaOx/γ-Al2O3, and VOx/γ-Al2O3) as oxygen carriers, Wu et al. [10] found that VOx/γ-Al2O3 oxygen carriers can conduct CL-ODHP reactions at a lower temperature and have higher CL-ODHP activity. Although VOx/γ-Al2O3 has subsequently attracted wide research interest, Jiang et al. [11] found that Olatt in VOx/γ-Al2O3 oxygen carriers are rapidly released and propane is deeply oxidized to COx, causing propylene selectivity to deteriorate. This selectivity can be improved by a metal-oxide promoter that regulates the Olatt release rate of the VOx/γ-Al2O3 oxygen carrier. According to previous research, CeO2 has excellent oxygen storage and release characteristics as well as redox properties, which can change the Olatt-transfer ability of adjacent metal oxides [12,13]. Hedun et al. [14] prepared FeCeOx as an oxygen carrier via the impregnation method. They reported that CeO2 inhibits the conversion rate of Olatt to Oelec in Fe-based oxygen carriers, thereby improving propylene selectivity. To similarly regulate the slow release of Olatt and improve propylene selectivity, we add CeO2 to VOx/γ-Al2O3 oxygen carriers and investigate the synergistic effect of VOx and CeO2 oxygen carriers on the CL-ODHP reaction. We also add bimetallic oxides to improve propylene selectivity. CeO2 increases the oxygen capacity of the carrier and controls the Olatt-release rate in the VOx/γ-Al2O3 oxygen carrier. To determine the optimum Ce content for increasing the propylene yield, we prepared a series of VOx (10 wt.%)-CeO2 (X wt.%)/γ-Al2O3 oxygen carriers (X = 0, 1, 3, 5, 7) for the CL-ODHP reaction and investigated the effect of Ce content on the physicochemical properties of the oxygen carriers. These properties were then related to the reactivity performance of CL-ODHP.

2. Oxygen Carrier Characterization

2.1. XRD Characterization Results

Figure 1 shows XRD profiles of the fresh 10V/Al, 10V-1Ce/Al, 10V-3Ce/Al, 10V-5Ce/Al, and 10V-7Ce/Al oxygen carriers. The diffraction peaks at 2θ = 45.56° and 67.03° are assigned to mesoporous γ-Al2O3 (JCPDS 10-0425) [15]. With increasing Ce content in the oxygen carrier, the peak intensity corresponding to the CeO2 phase becomes sharper and more intense. The diffraction peaks at 2θ = 28.89°, 32.38°, and 47.85° characterize the cubic crystal structure of the CeO2 phase, indicating that CeO2 exists in the cubic fluorite phase. The diffraction peaks of VOx species are absent, possibly because these species are highly dispersed on the surface of the γ-Al2O3 carrier, rendering them undetectable by the XRD instruments [16].

2.2. Analysis of Oxygen Carrier Morphology

Figure 2a–d shows the morphological characteristics of the oxygen carriers detected via SEM. All oxygen carriers appear as spherical particles with numerous surface folds. The surface smoothness of the 10V-XCe/Al oxygen carriers increases with increasing Ce content from 0 to 3 wt.%. When the Ce content reaches 7 wt.% (10V-7Ce/Al), the surface is rendered uneven by particle agglomeration.
The distributions of the Al, V, O, and Ce elements were investigated through mapping energy spectral analysis. Figure 2e,f shows the mapping images of the 10V/Al and 10V-3Ce/Al oxygen carriers, respectively. Each metal can be distinguished in the maps. Comparison of Figure 2e-III and Figure 2f-III shows that Ce introduction improves the uniformity of the V species through the sample. Moreover, Figure 2(e-II,f-II) shows that the introduction of Ce resulted in a denser distribution of O species through the samples, indicating that Ce increases the O content. As shown in Figure 2(f-IV), the Ce species is uniformly distributed on the surface of 10V-3Ce/Al.
Figure 3 shows the TEM of a fresh 10V-3Ce/Al oxygen carrier, where the polycrystalline material consists of small randomly oriented crystal structure domains. The lattice stripe spacing d = 0.31 nm and 0.26 nm correspond to the (111) crystal plane and the (200) crystal plane of CeO2, respectively [17]. The lattice stripe spacing d = 0.28 nm corresponds to the (301) crystal planes of V2O5 [18]. By Figure 2 and Figure 3, it is shown that CeO2 and V2O5 are uniformly dispersed on the surface of the oxygen carrier.

2.3. Pore-Structure Analysis of the Oxygen Carriers

Figure 4 shows the isothermal sorption and desorption curves of the oxygen carriers. The sorption and desorption curves of γ-Al2O3 and all oxygen carriers are type IV and exhibit H4-type hysteresis loops with low-pressure adsorption (P/P0 = 0–0.10, where P and P0 denote that P is the true pressure of the gas and P0 is the saturated vapor pressure of the gas at the measured temperature, respectively), indicating that the oxygen carriers contain few micropores. When P/P0 = 0.40, capillary coalescence occurs, resulting in a rapid increase in the adsorption amount [19].
Table 1 summarizes the pore-structure parameters of the oxygen carriers. γ-Al2O3 exhibits the highest specific surface area (206.87 m2/g), implying that it enhances the oxygen carrier’s performance most favorably. The specific surface area decreases to 119.95 m2/g after impregnating the VOx species. In the 10V-XCe/Al oxygen carriers, the pore structure parameters first increase and then decrease with increasing Ce content. The specific surface area, pore volume, and pore diameter are maximized in 10V-3Ce/Al (131.10 m2/g, 0.21 cm3/g, and 5.38 nm, respectively). These results indicate that adding a small amount of Ce to the 10V/Al oxygen carrier increases the specific surface area of the oxygen carrier, allowing more uniform dispersion of the active component on the carrier surface. Further, the enlarged pore size benefits propylene desorption [20].

2.4. Thermal Stabilities of the Oxygen Carriers

Figure 5 plots the TGA and differential thermogravimetric (DTG) curves of the 10V-XCe/Al samples (X = 0, 1, 3, 5, 7) as the temperature was increased from 50 °C to 800 °C in inert nitrogen gas. The DTG curves (Figure 5b) present a weight-loss peak at 90–120 °C, which probably represents the evaporation of physical water molecules from the sample surfaces [21]. The oxygen carrier with the highest cerium content (7 Ce) shows the least weight loss (Figure 5). The thermal stability of the oxygen carriers increases with increasing Ce loading, probably because the CeO2 phase inhibits γ-Al2O3 sintering and phase transition at high temperatures [22].

2.5. Reduction Properties of the Oxygen Carriers

The reducibility and oxygen-carrying capacities of the oxygen carriers were characterized using H2-TPR. The H2-TPR curves of all oxygen carriers are shown in Figure 6a. The 3Ce/Al oxygen carrier shows a reduction peak near 450 °C, which is attributed to the reduction of Ce4+, and the reduction peak of CeO2 phase overlaps with the reduction peak of VOx species [23]. The 10V/Al oxygen carrier shows an initial large reduction peak near 450 °C, attributed to the reduction of amorphous monomer or polymeric VOx species phases with oxidation states between V5+ and V4+ [24]. Therefore, in the temperature range of 400-550 °C, the H2 absorption is equal to the sum of H2 consumed during the reduction of the VOx phase and the CeO2 phase. Increasing the Ce loading increasingly shifts the position of the reduction peak from that of 10V/Al. Therefore, CeO2 changes the reducibility of the oxygen carrier. Moreover, the reducibility decreases gradually with temperature from left to right and its strength decreases in the order of 10V/Al > 10V-7Ce/Al > 10V-5Ce/Al > 10V-1Ce/Al > 10V-3Ce/Al. Gu [25] and Leon [26], who investigated the relationship between oxygen carrier reducibility and propylene selectivity, reported that lower reducibility usually corresponds to higher propylene selectivity.
The oxygen loadings in the oxygen carriers were calculated from the hydrogen consumptions of the carriers. Although the oxygen carrier conducts the H2-reduction process on its surface, the consumed Olatt is supplied from its interior [24]. As shown in Figure 6b, increasing the Ce content does not monotonically increase the amount of Olatt stored in the carrier; the oxygen loading is maximized in the carrier containing 3 wt.% Ce.

2.6. XPS Elemental Valence Analysis of the Oxygen Carriers

Figure 7a shows the Ce 3d XPS spectra of the oxygen carriers. Due to the splitting of the Ce 3d3/2 and Ce 3d5/2 main peaks and the different final states after relaxation, the XPS spectra of Ce species can be deconvoluted into six peaks. Peaks at 904.03 and 885.22 eV were attributed to Ce3+, and peaks at 907.04, 900.72, 898.17, and 881.74 eV were attributed to Ce4+ [27]. The Ce4+ species improve the electron-transfer channel and optimize the activity of the oxygen-carrier Olatt, thereby facilitating the dehydrogenation reaction [17]. Increasing the Ce loading in the oxygen carrier increases the total area of the Ce 3d peak. The Ce4+ to (Ce4+ + Ce3+) ratio is largest (57.32%) in 10V-3Ce/Al whereas the percentage of Ce4+ is smallest (37.64%) in 10V-7Ce/Al. The change in Ce4+/(Ce4+ + Ce3+) ratio follows the oxygen-loading trend of the oxygen carriers (Figure 6c). However, Olatt is present only in CeO2 and is absent in Ce3+ species [28]. When the mass fraction of Ce content in the oxygen carrier exceeds 3 wt.%, the Ce3+ content increases while the Ce4+ content decreases.
Figure 7b shows the XPS pattern of O 1s of fresh oxygen carriers, the peak at 530.02 eV corresponds to the adsorbed hydroxyl oxygen species (OOH), the peak at 531.17 eV corresponds to the lattice oxygen species (Olatt), and the peak at 532.39 eV corresponds to the electrophilic oxygen species (Oelec) [29]. As shown in Table 2, the content of the oxygen carrier Olatt increases significantly with increasing Ce loading, reaching a maximum of 10V-3Ce/Al. More Olatt promotes the selective oxidation of propane to propylene, thus improving propylene selectivity. The content of Oelec decreases with increasing Ce loading, and the highest concentration of Ce4+ is found in 10V-3Ce/Al oxygen carriers, where Ce4+ inhibits the conversion of Olatt to Oelec on the surface of the oxygen carriers, thus storing more Olatt for the carriers [27]. Figure 7e shows the O 1s XPS spectra before and after the reaction of 10V-3Ce/Al oxygenates, and the Olatt content of the oxygenate decreased significantly after the reaction. However, the rate of conversion of Olatt to Oelec was inhibited by the presence of Ce4+, and the oxidative dehydrogenation of propane also consumed Oelec, so the content of Oelec decreased after the reaction.
Figure 7c compares the XPS spectra of V 2p before and after the 10V-3Ce/Al reaction. The peaks at 515.35, 516.65, and 517.64 eV belong to V3+, V4+, and V5+ species, respectively [30]. In the spectrum of fresh 10V-3Ce/Al, the V3+ peak is absent because this species is oxidized during calcination in air. As shown in Table 3, the reaction significantly changes the content of V5+ species (the active species of ODHP [31]) in the oxygen carriers. After 15 min of reaction, 20.88% of the V5+ was completely converted to V3+. Figure 7d compares the Ce 3d XPS spectra before and after the 10V-3Ce/Al reaction. Note that Olatt exists in CeO2 [28]. As shown in Table 4, the Ce4+ content decreased from 63.27% before the reaction to 54.96% after the reaction; concomitantly, the Ce3+ content increased from 36.73% to 45.04% after the reaction. As the Olatt are slowly released, Ce4+ is not completely transformed into Ce3+. The slow release of Olatt accompanies the valence change in the Ce and V elements. The whole reaction is an oxidation-reduction process, in which the Olatt released by the oxygen carrier catalyzes the oxidation of propane to propylene.

3. CL-ODHP Performance of the Oxygen Carrier

3.1. In Situ DRIFT Investigation of the 10V-3Ce/Al Oxygen Carrier

To further elucidate the catalytic oxidation mechanism, the 10V-3Ce/Al oxygen carrier was subjected to in situ DRIFT experiments in an N2:C3H8 (v/v = 45:5) atmosphere. The vibration peaks at 1355, 1412, 1662, 2963, and 3189 cm−1 in the DRIFT spectra (Figure 8) are assigned to C-H bond stretching, C=O bond stretching, C=C bond stretching, variant Ce-O-V bonds, and V-OH bond stretching, respectively [7]. The peak at 1662 cm−1 is the absorption peak of CO2 [7]. Note that the area under the C=C bond peak is smaller at 600 °C than at 550 °C, indicating that more olefins are generated at 550 °C than at 600 °C. V5+ is the main activation site of the C-H bond, and in the presence of V5+, the propylidene C-H bond is broken and the propyl radical C3H7* and the hydrogen radical H* are adsorbed on the surface of the oxygen carrier [26]. When the temperature reaches 600 °C, the CO2 peak is significantly enhanced. At high temperatures (600 °C), the inhibition of Ce4+ is weakened, the rate of Olatt evolution to Oelec is enhanced, and the propyl radical C3H7* is deeply oxidized to COx, while the high temperature (600 °C) weakens the metal-to-metal interaction forces, leading to the disruption of the V-O-Ce bond [20,32]. For this reason, the activity of the oxygen carrier was tested at 550 °C in this study.

3.2. Effect of the Oxygen Carrier on Propane Conversion and Propylene Selectivity

Figure 9a compares the propane conversions of the 10V-XCe/Al oxygen carriers (X = 0, 1, 3, 5, 7). As CL-ODHP consumes Olatt from the oxygen carriers, the propane conversion decreases with increasing reaction time. The average propane conversion rates of 10V/Al, 10V-1Ce/Al, 10V-3Ce/Al, 10V-5Ce/Al, and 10V-7Ce/Al are 16.32%, 18.64%, 24.72%, 20.31%, and 23.16%, respectively, in the time range of 3-15 min. The results show that introducing CeO2 at 550 °C significantly improves propane conversion. Propane conversion is maximized at a Ce loading of 3 wt.%. This improvement can be explained by oxygen loading in the oxygen carrier combined with a structural pore change. Ce increases the oxygen carrying capacity of the oxygen carrier; in particular, Ce4+ inhibits the conversion and loss of Olatt and facilitates the oxidation of propane to propylene [17]. In addition, a small amount of Ce increases the specific surface area of the oxygen carrier, exposing more active sites on the surface of the oxygen carrier. The propane then reacts more fully.
Figure 9b compares the propylene selectivity of the 10V-XCe/Al oxygen carriers (X = 0, 1, 3, 5, 7). The propylene selectivity slowly decreases with increasing reaction time and is maximized at 83.42% at a Ce loading of 3 wt.%. The surface of the oxygen carrier hosts two main types of reactive oxygen species: Oelec, which oxidizes propane or propylene to COx by depth, and Olatt, which mainly reacts to produce propylene [33]. Previous XPS characterization showed that 10V-3Ce/Al contains more Ce4+ than the other samples. As Ce4+ inhibits the conversion of Olatt to Oelec, the 10V-3Ce/Al oxygen carrier retains a large number of Olatt and thus produces more propylene per unit time than the other oxygen carriers.
Figure 9c compares the COx selectivity of the 10V-XCe/Al oxygen carriers (X = 0, 1, 3, 5, 7). The 10V/Al oxygen carrier exhibits the highest COx selectivity (23.01%): the average COx selectivity values of 10V-1Ce/Al, 10V-3Ce/Al, 10V-5Ce/Al, 10V-7Ce/Al, and 3Ce/Al are 9.71%, 7.06%, 15.22%, 19.56%, and 52.42%, respectively. When the 3Ce/Al catalyst carrier (which lacks the VOx component) reacts with propane, the Olatt species rapidly evolves into Oelec and propane is directly and deeply oxidized to COx, basically bypassing the ODHP effect [34].
Figure 9d compares the propylene yield of the 10V-XCe/Al oxygen carriers (X = 0, 1, 3, 5, 7). The average yields of propylene for 10V/Al, 10V-1Ce/Al, 10V-3Ce/Al, 10V-5Ce/Al, and 10V-7Ce/Al were 10.53%, 13.89%, 19.36%, 14.78%, and 15.23%. The propylene yield of 10V-3Ce/Al was improved by 83.86% compared to 10V/Al oxygenate carrier. From BET, it is known that 10V-3Ce/Al has the largest specific surface area, which provides more active for sites for CL-ODHP and facilitates the adequate reaction of propane. From H2-TPR, 10V-3Ce/Al has the lowest reducibility. According to XPS, 10V-3Ce/Al has the highest Olatt capacity and Ce4+ content. Ce4+ inhibits the evolution of Olatt on the surface of the oxygen carrier towards Oelec, weakening the rate of propane being deeply oxidized to COx. Therefore, the average yield of propylene from 10V-3Ce/Al oxygen carriers is the highest.

3.3. CL-ODHP Cycling of the 10V/Al and 10V-3Ce/Al Oxygen Carriers

In this experiment, two oxygen carriers, 10V/Al and 10V-3Ce/Al, were cycled through 15 reaction regenerations at 550 °C. After 15 cycles, the propane conversion, propylene selectivity, and COx selectivity of the 10V-3Ce/Al oxygen carrier were 22.41%, 81.74%, and 6.88%, respectively (Figure 10), almost unchanged from those of the first dehydrogenation. The cyclic oxidative-dehydrogenation results confirmed the high reactivity and stability of the 10V-3Ce/Al oxygen carrier.
Figure 11a XRD patterns of fresh oxygen carriers, Figure 11b XRD patterns of oxygen carriers after 15 cycles of CL-ODHP. No diffraction peaks of VOx species were observed before and after the cycle, and the absence of VOx species peaks could be due to the formation of highly dispersed amorphous VOx phases or the formation of VOx microcrystals undetectable by XRD [16]. The peak intensity of the CeO2 phase of the 10V-3Ce/Al oxygen carrier did not decrease significantly after cycling, indicating that the CeO2 phase was still stably dispersed on the surface of the oxygen carrier after 15 cycles of CL-ODHP.
Figure 12 shows the TEM of the 10V10V-3Ce/Al 3Ce/Al oxygenate before and after cycling. After 15 cycles of CL-ODHP, there is no sintering on the surface of the 10V-3Ce/Al oxygenate, indicating that the 10V10V-3Ce/Al 3Ce/Al oxygenates have good cycling stability. The lattice stripe spacing d = 0.31 nm corresponds to the (111) crystal plane of CeO2. The lattice stripe spacing d = 0.28 nm corresponds to the (301) crystal planes of V2O5. V2O5 phase and CeO2 phase are stable distributions on the surface of the oxygen carrier, indicating that the 10V-3Ce/Al oxygen carrier has good cycle stability.

4. Mechanism of CL-ODHP

The CL-ODHP follows the Mars-van Krevelen mechanism. The CL-ODHP mechanism was proposed by combining in situ DRIFT and XPS characterization. As shown in Figure 13, the first step of CL-ODHP is the adsorption of propane on the surface of the 10V-3Ce/Al oxygen carrier and the breaking of the propane methylene C-H bond. V5+ is the dehydrogenation activation site [16]. The rate-limiting step is breakage of the C-H bond of propane to form the propyl radical C3H7* and the hydrogen radical H*. Based on changes in XPS Olatt content before and after oxygen carrier reaction combined with previous in situ DRIFTS characterization, VOx then releases Olatt to selectively oxidize C3H7* and H* to C3H6 and H2O, respectively. CeO2 uniformly dispersed on the surface of the oxygen carrier and controls the release and evolution rate of Olatt, thus avoiding the deep oxidation of C3H7*. However, at high temperatures (600 °C), CeO2 loses its inhibition ability and Olatt rapidly evolves into Oelec, enabling the deep oxidation of C3H7* to CO2 [20]. Similar results were obtained by Gong et al. and Zuo et al. Gong et al. [7] investigated the application of VOx-TiO2 oxygen carriers in CL-ODHP. V5+ is the main C-H bond activation site, and the propyl radical C3H7* and the hydrogen radical H* are generated by the breakage of the propane methylene bond under the activation of V5+. Zuo et al. [20] investigated the application of CeO2-modified 17.5 Cr/Al oxygen carriers in CL-ODHP. CeO2 increased the amount of Olatt and enhanced the adsorption of propyl radicals, while Ce4+ modulated the rate of Olatt evolution to Oelec and propyl radicals were further oxidized to propylene.

5. Experiment

5.1. Preparation of the Oxygen Carriers

Ce (NO3)·6H2O, NH4VO3, and H2C2O4·2H2O were purchased from Sinopharm Group Chemical Reagent Co., Ltd., (Shanghai, China). Ce (NO3)·6H2O, NH4VO3, and H2C2O4·2H2O were weighed and simultaneously dissolved in deionized water. During this process, the Ce content was adjusted to 0, 1, 3, 5, or 7 wt.%, and the mass ratio of oxalic acid to ammonium metavanadate was fixed at 2:1. The solution was then ultrasonicated for 30 min to ensure complete mixing of the constituents. After immersing the γ-Al2O3 carrier, the solution was allowed to stand at room temperature for 12 h and then dried in an oven at 90 °C for 12 h. The dried oxygen carrier was calcined in a 600 °C muffle furnace for 4 h (heating rate: 3 °C/min). Finally, the oxygen carrier was ground and sieved. The prepared samples were expressed as 10V-XCe/Al (VOx content = 10 wt.%; CeOy content, X = 0, 1, 3, 5, and 7 wt.%).

5.2. Oxygen-Carrier Performance Tests

Figure 14 shows the CL-ODHP reactor for investigating the activities of the oxygen carriers. The mass of the oxygen carriers was 0.50 g. The temperature was increased to 550 °C at 10 °C/min in an N2 atmosphere. The N2 was then purged for 20 min before passing propane through the reactor. The flow rates of the N2 and C3H8 feed gases were 45 and 5 mL/min, respectively. After 3 min of reaction, the reaction gas was collected from the bottom of the quartz tube into a gas-sampling bag and then pumped into a gas chromatograph (A91 Plus Panna, Guangzhou Baiwei Instrument Technology Co., Ltd., Guangzhou, China) for analysis.
The propane conversion, propylene selectivity, COx selectivity, and propylene yield were respectively calculated as follows:
Propane conversion rate
X C 3 H 8 = N C 3 H 6   + 2 3 N C 2 + 1 3 N C 1 N C 3 + 2 3 N C 2 + 1 3 N C 1 × 100 %
Propylene selectivity
S C 3 H 6 =   N C 3 H 6 N C 3 H 6 + 2 3 N C 2 + 1 3 N C 1 × 100 %
COx selectivity
S CO x = N CO x N C 3 H 6 + 2 3 N C 2 + 1 3 N C 1 × 100 %
Propylene yield
Y C 3 H 6 = X C 3 H 8 × S C 3 H 6
here, N represents the molar concentration of each substance at the outlet, C1 denotes CH4, CO, and CO2, C2 denotes C2H4 and C2H6, and C3 denotes C3H6 and C3H8.

5.3. Oxygen Carrier Characterization

The structural and physical phases in the oxygen carrier were analyzed via X-ray diffraction (XRD) using a DX-2007 Advance X-ray diffractometer (Haoyuan Corporation, Dandong, China) with Cu Kα as the X-ray emission source, a tube voltage of 40 KV, and a wide angle-scanning range (5–90° in 5° increments). The structural parameters of the oxygen carriers were characterized using a Micromeritics TriStar 3000 physisorption instrument (Norcross, GA, USA) and the specific surface areas and pore-size distributions were calculated via Brunauer-Emmett-Teller and Barrett-Joyner-Halenda computational modeling, respectively. The sample to be tested was first degassed at 200 °C under vacuum for 3 h. The surface morphological characteristics of the oxygen carriers were observed under a scanning electron microscope (SEM) (JSM-IT100; JOEL Corporation, Tokyo, Japan) operated at 20 kV.
H2 temperature-programmed reduction (H2-TPR) was tested with a Micromeritics Autochen 2920 chemisorption instrument (Mack Corporation, USA). For this test, 50 mg of the sample was weighed and treated in a He atmosphere at 200 °C for 1 h. The treated sample was cooled to room temperature and the atmosphere was switched to 8% H2/Ar, then purged until stabilization was achieved. After stabilization, the sample was heated to 900 °C and the H2-TPR curve was obtained by recording the hydrogen consumption signal with a thermal conductivity detector (TCD). The elemental valence states of the oxygen carriers were characterized using X-ray electron spectroscopy (XPS) (Kratos Analytical, Manchester, UK) with monochromatic AlKα (hv = 1486.65 eV) as the excitation source. The sample to be tested was pressed and pretreated under the reaction conditions before loading it into the test bag. The thermal stabilities of the oxygen carriers were analyzed via thermogravimetric analysis (TGA) using an STA-409PC instrument (Netz, Selb, Germany) under an N2 atmosphere. The temperature was increased from the programmed 30 °C to 800 °C and the tail-gas concentration was detected using TCD.
In situ diffuse-reflectance infrared Fourier-transform spectroscopy (DRIFTS) data were collected using a DRIFTS instrument equipped with a mercury-cadmium-telluride detector and an in situ reaction cell (Thermo Fisher Nicolet, Waltham, MA, USA). The resolution of the data collection was 4 cm−1. Prior to analysis, the samples were dried at 600 °C for 30 min under an N2 (45 mL/min) atmosphere. The background spectra were then collected at the required temperature while the temperature of the in situ reaction tank naturally declined to 30 °C. Subsequently, the oxygen carrier absorbed N2: C3H8 (v/v = 45/5) gas mixture for 20 min. During measurements, the temperature was ramped at 3 °C/min. Each temperature point was stabilized for 10 min.

6. Conclusions

In this study, a series of 10V-XCe/Al oxygen carriers (X = 0, 1, 3, 5, 7) for CL-ODHP was prepared through an impregnation technique. The effects of Ce content on the catalytic activity and physicochemical properties of 10V-XCe/Al were investigated. The main findings are summarized below.
  • Introducing CeO2 increases the Olatt capacity of VOx/γ-Al2O3 oxygen carriers and prolongs their catalytic oxidation activity. A certain amount of CeO2 is uniformly dispersed on the surface of VOx/γ-Al2O3 oxygen carriers, inhibiting the migration of Olatt species to electrophilic oxygen species, thus improving propylene selectivity.
  • Among the prepared samples, 10V-3Ce/Al exhibits the highest specific surface area, the highest oxygen capacity, and the lowest reducibility. The 10V-3Ce/Al oxygen carrier also delivers the highest oxidative dehydrogenation performance.
  • In the VOx/γ-Al2O3 oxygen carrier with the CeO2 additive (10V-3Ce/Al), the propylene selectivity improves to 83.42% (from 62.59% in 10V/Al) and the COx selectivity reduces to 6.88% (from 22.76% in 10V/Al).

Author Contributions

Conceptualization, F.Q.; data curation, F.Q.; formal analysis, F.Q., M.W. and T.G.; investigation, F.Q. and Q.G.; methodology, F.Q., T.G. and M.W.; software, F.Q. and M.N.; validation, F.Q., Q.G. and M.W.; visualization, F.Q., T.G., M.N. and Y.L.; writing—original draft, F.Q.; writing—review & editing, F.Q., T.G., Q.G. and M.W.; project administration, Q.G.; resources, Q.G.; supervision, Q.G. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Joint Funds of the National Natural Science Foundation of China (U20A20124), Shandong provincial natural science foundation (ZR2020MB144), the State Key Laboratory of High-Efficiency Utilization of Coal and Green Chemical Engineering (2022-K3), and the Natural Science Foundation Project of Ningxia (2022AAC01001).

Data Availability Statement

Data sharing is not applicable to this article.

Acknowledgments

Thanks is expressed for the platform provided by Qingdao University of Science and Technology.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chen, S.; Chang, X.; Sun, G.; Zhang, T.; Xu, Y.; Wang, Y.; Pei, C.; Gong, J. Propane dehydrogenation: Oxygen carrier development, new chemistry, and emerging technologies. Chem. Soc. Rev. 2021, 50, 3315–3354. [Google Scholar] [CrossRef]
  2. Monai, M.; Gambino, M.; Wannakao, S.; Weckhuysen, B.M. Propane to olefins tandem catalysis: A selective route towards light olefins production. Chem. Soc. Rev. 2021, 50, 11503–11529. [Google Scholar] [CrossRef]
  3. Dai, Y.; Gao, X.; Wang, Q.; Wan, X.; Zhou, C.; Yang, Y. Recent progress in heterogeneous metal and metal oxide oxygen carriers for direct dehydrogenation of ethane and propane. Chem. Soc. Rev. 2021, 50, 5590–5630. [Google Scholar] [CrossRef] [PubMed]
  4. Hu, Z.P.; Yang, D.; Wang, Z.; Yuan, Z.H. State-of-the-art oxygen carriers for direct dehydrogenation of propane to propylene. Chin. J. Catal. 2019, 40, 1233–1254. [Google Scholar] [CrossRef]
  5. Jiang, X.; Sharma, L.; Fung, V.; Park, S.; Jones, C.; Sumpter, B.; Baltrusaitis, J.; Wu, L. Oxidative dehydrogenation of propane to propylene with soft oxidants via heterogeneous catalysis. ACS Catal. 2021, 11, 2182–2234. [Google Scholar] [CrossRef]
  6. Li, H.; Zhang, J.; Wu, P.; Xun, S.; Jiang, W.; Zhang, M.; Zhu, W.; Li, H. O2 activation and oxidative dehydrogenation of propane on hexagonal boron nitride: Mechanism revisited. J. Phys. Chem. C 2019, 123, 2256–2266. [Google Scholar] [CrossRef]
  7. Chen, S.; Pei, C.; Chang, X.; Zhao, Z.; Mu, R.; Xu, Y.; Gong, J. Coverage-Dependent Behaviors of Vanadium Oxides for Chemical Looping Oxidative Dehydrogenation. Angew. Chem. Int. Ed. 2020, 59, 22072–22079. [Google Scholar] [CrossRef]
  8. Chen, S.; Zeng, L.; Mu, R.; Xiong, C.; Zhao, Z.; Zhao, C.; Pei, C.; Peng, L.; Luo, J.; Fan, L.; et al. Modulating Lattice Oxygen in Dual-Functional Mo-V-O Mixed Oxides for Chemical Looping Oxidative Dehydrogenation. J. Am. Chem. Soc. 2019, 141, 18653–18657. [Google Scholar] [CrossRef] [PubMed]
  9. Liu, J.; Li, F. Mixed oxides as multi-functional reaction media for chemical looping catalysis. Chem. Commun. 2022, 59, 10–28. [Google Scholar] [CrossRef]
  10. Wu, T.W.; Yu, Q.B.; Roghair, I.; Wang, K.; Annaland, M.S. Chemical looping oxidative dehydrogenation of propane: A comparative study of Ga-based, Mo-based, V-based oxygen carriers. Chem. Eng. Process. 2020, 157, 1076–1085. [Google Scholar] [CrossRef]
  11. Jiang, X.; Zhang, X.; Purdy, S.C.; He, Y.; Huang, Z.; You, R.; Wei, Z.; Meyer, H.M.; Yang, J.; Pan, Y.; et al. Multiple promotional effects of vanadium oxide on boron nitride for oxidative dehydrogenation of propane. JACS Au 2022, 2, 1096–1104. [Google Scholar] [CrossRef]
  12. Wang, J.; Wan, C.; Cheng, D.; Chen, F.; Zhang, X. Unveiling the morphology dependence of ceria nanocrystals for boosting low-temperature cyclohexane oxidative dehydrogenation. Appl. Surf. Sci. 2021, 565, 15060–15071. [Google Scholar] [CrossRef]
  13. Prathap, C.; Ramana, K.V.V.; Venkata, R.M.; Nagaiah, P.; David, R.B. Promotional role of ceria in CeO2/MgAl2O4 spinel oxygen carriers in CO2 assisted selective oxidative dehydrogenation of ethylbenzene to styrene. J. Ind. Eng. Chem. 2019, 79, 97–105. [Google Scholar]
  14. Wang, H.; Tsilomelekis, G. Catalytic performance and stability of Fe-doped CeO2 in propane oxidative dehydrogenation using carbon dioxide as an oxidant. Catal. Sci. Technol. 2020, 10, 4362–4372. [Google Scholar] [CrossRef]
  15. Shang, X.; Wang, X.; Nie, W.; Guo, X.; Zou, X.; Ding, W.; Lu, X. Facile strategy for synthesis of mesoporous crystalline γ-alumina by partially hydrolyzing aluminum nitrate solution. J. Mater. Chem. 2012, 22, 23806–23814. [Google Scholar] [CrossRef]
  16. Khan, M.Y.; Adamu, S.; Lucky, R.A.; Hossain, M.M. Oxidative dehydrogenation of n-butane to C4 olefins using lattice oxygen of VOx/Ce-meso-Al2O3 under gas-phase oxygenfree conditions. Energy Fuels 2020, 34, 7410–7421. [Google Scholar] [CrossRef]
  17. Liu, J.; Hao, M.; Chen, C.; Du, K.; Zhou, Q.; Zou, S.; Fan, J. Chlorinating CeO2 at surface oxygen vacancies to promote their selectivity in oxidative dehydrogenation of propane to propene. Appl. Surf. Sci. 2020, 528, 147025. [Google Scholar] [CrossRef]
  18. Kazerooni, H.; Towfighi Darian, J.; Mortazavi, Y.; Asadi, R. Titania-supported vanadium oxide synthesis by atomic layer deposition and its application for low-temperature oxidative dehydrogenation of propane. Catal. Lett. 2020, 150, 2807–2822. [Google Scholar] [CrossRef]
  19. Fu, S.; Fang, Q.; Li, A.; Li, Z.; Han, J.; Dang, X.; Han, W. Accurate characterization of full pore size distribution of tight sandstones by low-temperature nitrogen gas adsorption and high-pressure mercury intrusion combination method. Energy Sci. Eng. 2020, 9, 80–100. [Google Scholar] [CrossRef]
  20. Zuo, C.; Wu, M.; Guo, Q. The effect of the Ce content on the oxidative dehydrogenation of propane over CrO-CeO2/γ-Al2O3 oxygen carriers. Chin. J. Chem. Eng. 2020, 28, 3035–3043. [Google Scholar] [CrossRef]
  21. Nasution, P.S.; Jung, J.W.; Oh, K.; Koh, H.L. Coke combustion kinetics of spent Pt-Sn/Al2O3 oxygen carriers in propane dehydrogenation. Korean J. Chem. Eng. 2020, 37, 1490–1497. [Google Scholar] [CrossRef]
  22. Xie, S.; Wang, Z.; Tan, W.; Zhu, Y.; Collier, S.; Ma, L.; Ehrlich, S.N.; Xu, P.; Yan, Y.; Xu, T.; et al. Highly Active and Stable Palladium Oxygen carriers on Novel Ceria-Alumina Supports for Efficient Oxidation of Carbon Monoxide and Hydrocarbons. Environ. Sci. Technol. 2021, 55, 7624–7633. [Google Scholar] [CrossRef]
  23. Khan, M.Y.; Al-Ghamdi, S.; Razzak, S.A.; Hossain, M.M. Fluidized bed oxidative dehydrogenation of ethane to ethylene over VOx/Ce-γAl2O3 catalysts: Reduction kinetics and catalyst activity. Mol. Catal. 2017, 443, 78–91. [Google Scholar] [CrossRef]
  24. Hu, P.; Lang, W.Z.; Yan, X.; Chen, F.X.; Guo, Y.J. Vanadium-doped porous silica materials with high catalytic activity and stability for propane dehydrogenation reaction. Appl. Catal. A-Gen. 2018, 553, 65–73. [Google Scholar] [CrossRef]
  25. Gu, J.; Zang, Y.; Gao, F. Study on preparation of vanadium oxide modified heteroatom B-Beta molecular sieve oxygen carrier and its catalytic performance on propane dehydrogenation. J. Light Ind. 2020, 35, 35–46. [Google Scholar]
  26. Schumacher, L.; Hess, C. The active role of the support in propane ODH over VOx/CeO2 oxygen carriers studied using multiple operando spectroscopies. J. Catal. 2021, 398, 29–43. [Google Scholar] [CrossRef]
  27. Maslakov, K.I.; Teterin, Y.A.; Popel, A.J.; Teterin, A.Y.; Ivanov, K.E.; Kalmykov, S.N.; Petrov, V.G.; Farnan, I. XPS study of ion irradiated and unirradiated CeO2 bulk and thin film samples. Appl. Surf. Sci. 2018, 448, 154–162. [Google Scholar] [CrossRef]
  28. Kovacevic, M.; Agarwal, S.; Mojet, B.L.; Lefferts, L. The effects of morphology of cerium oxide oxygen carriers for dehydrogenation of ethylbenzene to styrene. Appl. Catal. A Gen. 2015, 505, 354–364. [Google Scholar] [CrossRef]
  29. Shan, Y.; Zhao, W.; Zhao, S.; Wang, X.; Sun, H.; Yu, W.; Ding, J.; Feng, X.; Chen, D. Effects of alumina phases on the structure and performance of VOx/Al2O3 oxygen carriers in non-oxidative propane dehydrogenation. Mol. Catal. 2021, 504, 11466–11475. [Google Scholar]
  30. Wang, Y.; Chen, S.; Sun, J.; Xie, Y.; Zhao, Z.; Gong, J. Roles of V-O sites for non-oxidative propane dehydrogenation over supported vanadium oxides. Sci. China Mater. 2022, 603, 1062–1070. [Google Scholar] [CrossRef]
  31. Kaichev, V.V.; Chesalov, Y.A.; Saraev, A.A.; Tsapina, A.M. A Mechanistic study of dehydrogenation of propane over vanadia-titania oxygen carriers. J. Phys. Chem. C 2019, 123, 19668–19680. [Google Scholar] [CrossRef]
  32. Huang, C.; Wang, Z.; Gong, X. Activity and selectivity of propane oxidative dehydrogenation over VO3/CeO2 (111) catalysts: A density functional theory study. Chin. J. Catal. 2018, 39, 1520–1526. [Google Scholar] [CrossRef]
  33. Zeeshan, M.; Chang, Q.Y.; Zhang, J.; Hu, P.; Sui, Z.; Chen, D.; Zhu, Y. Effects of oxygen vacancy and Pt doping on the catalytic performance of CeO2 in propane dehydrogenation: A first-principles study. Chin. J. Chen. 2021, 39, 2391–2402. [Google Scholar] [CrossRef]
  34. Zuo, C.; Wu, M.; Guo, Q. CeO2-CrOy-γ-Al2O3 redox oxygen carrier for the oxidative dehydrogenation of propane to propylene. Can. J. Chem. Eng. 2020, 99, 235–250. [Google Scholar] [CrossRef]
Figure 1. XRD spectra of the 10V-XCe/Al oxygen carriers (X = 0, 1, 3, 5, 7).
Figure 1. XRD spectra of the 10V-XCe/Al oxygen carriers (X = 0, 1, 3, 5, 7).
Catalysts 13 00797 g001
Figure 2. SEM images of (a) 10V/Al, (b) 10V-1Ce/Al, (c) 10V-3Ce/Al, and (d) 10V-7Ce/Al; elemental maps of (e) 10V/Al and (f) 10V-3Ce/Al (I: aluminum; II: oxygen; III: vanadium; IV: cerium).
Figure 2. SEM images of (a) 10V/Al, (b) 10V-1Ce/Al, (c) 10V-3Ce/Al, and (d) 10V-7Ce/Al; elemental maps of (e) 10V/Al and (f) 10V-3Ce/Al (I: aluminum; II: oxygen; III: vanadium; IV: cerium).
Catalysts 13 00797 g002
Figure 3. TEM images of 10V-3Ce/Al oxygen carrier.
Figure 3. TEM images of 10V-3Ce/Al oxygen carrier.
Catalysts 13 00797 g003
Figure 4. N2 adsorption-desorption isotherms of γ-Al2O3 and the 10V-XCe/Al oxygen carriers.
Figure 4. N2 adsorption-desorption isotherms of γ-Al2O3 and the 10V-XCe/Al oxygen carriers.
Catalysts 13 00797 g004
Figure 5. (a) Thermogravimetric and (b) differential thermogravimetric curves of the oxygen carriers.
Figure 5. (a) Thermogravimetric and (b) differential thermogravimetric curves of the oxygen carriers.
Catalysts 13 00797 g005
Figure 6. (a) H2-temperature programmed reduction spectra and (b) oxygen carrying capacities of the oxygen carriers; (c) ratios of Ce4+ content to (Ce4+ + Ce3+) content and oxygen carrying capacities in the oxygen carriers with different Ce contents.
Figure 6. (a) H2-temperature programmed reduction spectra and (b) oxygen carrying capacities of the oxygen carriers; (c) ratios of Ce4+ content to (Ce4+ + Ce3+) content and oxygen carrying capacities in the oxygen carriers with different Ce contents.
Catalysts 13 00797 g006
Figure 7. (a) Ce 3d XPS spectra and (b) O 1s XPS spectra of the oxygen carriers; (c) V 2p XPS spectra, (d) Ce 3d XPS spectra, and (e) O1s XPS spectra before and after the 10V-3Ce/Al reaction.
Figure 7. (a) Ce 3d XPS spectra and (b) O 1s XPS spectra of the oxygen carriers; (c) V 2p XPS spectra, (d) Ce 3d XPS spectra, and (e) O1s XPS spectra before and after the 10V-3Ce/Al reaction.
Catalysts 13 00797 g007
Figure 8. In situ diffuse-reflectance infrared Fourier-transform spectra of the 10V-3Ce/Al oxygen carrier at different temperatures (350–600 °C).
Figure 8. In situ diffuse-reflectance infrared Fourier-transform spectra of the 10V-3Ce/Al oxygen carrier at different temperatures (350–600 °C).
Catalysts 13 00797 g008
Figure 9. Effects of reaction time on (a) propane conversions, (b) propylene selectivity, (c) COx selectivity of the oxygen carriers, and (d) propylene yield.
Figure 9. Effects of reaction time on (a) propane conversions, (b) propylene selectivity, (c) COx selectivity of the oxygen carriers, and (d) propylene yield.
Catalysts 13 00797 g009
Figure 10. Effects of oxygen carrier cycling on (a) propane conversions and (b) propylene selectivity and COx selectivity of 10V/Al (black) and 10V-3Ce/Al (red).
Figure 10. Effects of oxygen carrier cycling on (a) propane conversions and (b) propylene selectivity and COx selectivity of 10V/Al (black) and 10V-3Ce/Al (red).
Catalysts 13 00797 g010
Figure 11. XRD spectra of (a) fresh 10V/Al and 10V10V-3Ce/Al 3Ce/Al oxygen carrier, (b) 10V/Al and 10V-3Ce/Al oxygen carrier after 15 CL-ODHP cycles.
Figure 11. XRD spectra of (a) fresh 10V/Al and 10V10V-3Ce/Al 3Ce/Al oxygen carrier, (b) 10V/Al and 10V-3Ce/Al oxygen carrier after 15 CL-ODHP cycles.
Catalysts 13 00797 g011
Figure 12. TEM images of (a,b) fresh 10V-3Ce/Al oxygen carrier, (c,d) 10V-3Ce/Al oxygen carrier after 15 CL-ODHP cycles.
Figure 12. TEM images of (a,b) fresh 10V-3Ce/Al oxygen carrier, (c,d) 10V-3Ce/Al oxygen carrier after 15 CL-ODHP cycles.
Catalysts 13 00797 g012
Figure 13. Mechanism of CL-ODHP on the 10V-3Ce/Al oxygen carrier.
Figure 13. Mechanism of CL-ODHP on the 10V-3Ce/Al oxygen carrier.
Catalysts 13 00797 g013
Figure 14. Schematic of the CL-ODHP setup.
Figure 14. Schematic of the CL-ODHP setup.
Catalysts 13 00797 g014
Table 1. Structural properties of γ-Al2O3 and the 10V-XCe/Al oxygen carriers.
Table 1. Structural properties of γ-Al2O3 and the 10V-XCe/Al oxygen carriers.
SamplesSBET (m2/g)Vp (cm3/g)Dp (nm)
γ-Al2O3206.870.253.25
10V-Al119.950.184.64
10V-1Ce/Al124.910.194.95
10V-3Ce/Al131.100.215.38
10V-5Ce/Al104.290.164.30
10V-7Ce/Al90.130.113.75
Table 2. Relative content of oxygen species on the surface of fresh oxygen carriers.
Table 2. Relative content of oxygen species on the surface of fresh oxygen carriers.
Oxygen CarrierOlatt/(OOH + Olatt + Oelec)Oelec/(OOH + Olatt + Oelec)
10V/Al31.45%27.78%
10V-1Ce/Al37.63%22.01%
10V-3Ce/Al45.38%19.03%
10V-5Ce/Al35.61%23.51%
10V-7Ce/Al33.77%26.08%
Table 3. Valence changes in the V element before and after the 10V-3Ce/Al reaction.
Table 3. Valence changes in the V element before and after the 10V-3Ce/Al reaction.
The Valence State of Element VV5+V4+V3+
Before reaction41.92%58.08%0
After Reaction21.04%37.76%41.20
Table 4. Valence changes in the Ce element before and after the 10V-3Ce/Al reaction.
Table 4. Valence changes in the Ce element before and after the 10V-3Ce/Al reaction.
The Valence State of Element CeCe4+Ce3+
Before reaction63.27%36.73%
After reaction54.96%45.04%
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

Qiang, F.; Guo, T.; Nie, M.; Liu, Y.; Wu, M.; Guo, Q. Effect of Ce Content on the Chemical Looping Oxidative Dehydrogenation of Propane to Propylene over a VOx-CeO2/γ-Al2O3 Oxygen Carrier. Catalysts 2023, 13, 797. https://doi.org/10.3390/catal13050797

AMA Style

Qiang F, Guo T, Nie M, Liu Y, Wu M, Guo Q. Effect of Ce Content on the Chemical Looping Oxidative Dehydrogenation of Propane to Propylene over a VOx-CeO2/γ-Al2O3 Oxygen Carrier. Catalysts. 2023; 13(5):797. https://doi.org/10.3390/catal13050797

Chicago/Turabian Style

Qiang, Fangyuan, Tuo Guo, Mengdong Nie, Yongzhuo Liu, Man Wu, and Qingjie Guo. 2023. "Effect of Ce Content on the Chemical Looping Oxidative Dehydrogenation of Propane to Propylene over a VOx-CeO2/γ-Al2O3 Oxygen Carrier" Catalysts 13, no. 5: 797. https://doi.org/10.3390/catal13050797

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