Next Article in Journal
Atmospheric Carbonyl Compounds in the Central Taklimakan Desert in Summertime: Ambient Levels, Composition and Sources
Next Article in Special Issue
Effect of Pore Structure on CO2 Adsorption Performance for ZnCl2/FeCl3/H2O(g) Co-Activated Walnut Shell-Based Biochar
Previous Article in Journal
The Effect of Metro Construction on the Air Quality in the Railway Transport System of Sydney, Australia
Previous Article in Special Issue
Amine-Modified Biochar for the Efficient Adsorption of Carbon Dioxide in Flue Gas
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Sr1-xKxFeO3 Perovskite Catalysts with Enhanced RWGS Reactivity for CO2 Hydrogenation to Light Olefins

1
Key Laboratory of Clean Chemical Engineering in Universities of Shandong, College of Chemical Engineering, Qingdao University of Science & Technology, Qingdao 266042, China
2
State Key Laboratory of High-Efficiency Utilization of Coal and Green Chemical Engineering, Ningxia University, Yinchuan 750021, China
*
Authors to whom correspondence should be addressed.
Atmosphere 2022, 13(5), 760; https://doi.org/10.3390/atmos13050760
Submission received: 11 April 2022 / Revised: 6 May 2022 / Accepted: 6 May 2022 / Published: 8 May 2022
(This article belongs to the Special Issue CO2 Sequestration, Capture and Utilization)

Abstract

:
The catalytic hydrogenation of CO2 to light olefins (C2–C4) is among the most practical approaches to CO2 utilization as an essential industrial feedstock. To achieve a highly dispersed active site and enhance the reactivity of the reverse water–gas shift (RWGS) reaction, ABO3-type perovskite catalysts Sr1-xKxFeO3 with favorable thermal stability and redox activity are reported in this work. The role of K-substitution in the structure–performance relationship of the catalysts was investigated. It indicated that K-substitution expedited the oxygen-releasing process of the SrFeO3 and facilitated the synchronous formation of active-phase Fe3O4 for the reverse water–gas shift (RWGS) reaction and Fe5C2 for the Fischer–Tropsch synthesis (FTS). At the optimal substitution amount, the conversion of CO2 and the selectivity of light olefins achieved 30.82% and 29.61%, respectively. Moreover, the selectivity of CO was up to 45.57% even when H2/CO2=4 due to CO2-splitting reactions over the reduced Sr2Fe2O5. In addition, the reversibility of perovskite catalysts ensured the high dispersion of the active-phase Fe3O4 and Fe5C2 in the SrCO3 phase. As the rate-determining step of the CO2 hydrogenation reaction to light olefins over Sr1-xKxFeO3 perovskite catalysts, FTS should be further tailored by partial substitution of the B site. In sum, the perovskite-derived catalyst investigated in this work provided a new idea for the rational design of a catalyst for CO2 hydrogenation to produce light olefins.

1. Introduction

As a major greenhouse gas, atmospheric carbon dioxide has been increasing in concentration in the last two decades, causing severe global warming and extreme climates [1]. Efficient utilization of carbon dioxide to obtain high-value-added chemicals can alleviate both the environmental problems and the overdependence on other carbon resources [2]. The hydrogenation of carbon dioxide to C2–C4 light olefins, which are basic building blocks in chemical industries, has attracted extensive attention [3,4,5]. The main limitation of CO2 conversion is its high thermodynamic stability, thus requiring catalysts and energy input to drive the transformation.
CO2 hydrogenation to olefins mainly follows a CO2-modified Fischer–Tropsch synthesis (CO2-FTS) route or a methanol-mediated (MeOH) route, but is often limited by low selectivity and catalyst deactivation [4,5,6]. The CO2-FTS synthesis route usually consists of two consecutive catalytic reactions using Fe-based catalysts. Firstly, a reverse water–gas shift (RWGS) reaction converts CO2 to CO with a Fe3O4 phase. Subsequently, an FTS reaction causes CO hydrogenation to hydrocarbons, catalyzed by active iron carbide phase of Fe5C2 [7,8]. Because RWGS is related to the formation of the CO intermediate, while FTS determines the yield and distribution of the olefin products, both reactions should be controlled carefully.
To overcome low CO2 conversion and Anderson–Schulz–Flory (ASF) distribution caused by bulk Fe-based catalysts, various modification strategies [4] were adopted by adding electron promotors (e.g., alkali metals) and structure promotors (e.g., copper, zinc, and cobalt), as well as supports (e.g., Al2O3, SiO2, or MOFs). As efficient electron donors, the alkali metals Na and K favor the adsorption of CO2 onto Fe oxides, and facilitate the generation of the Fe5C2 phase [9,10,11]; however, excessive additions likely reduce the activity of CO2 hydrogenation by decreasing the total surface area and creating a hydrogen-lean environment [12]. The formed metallic Cu nanoparticles with CuO-doping benefits Fe-based catalyst activity for the RWGS [13,14], while ZnO addition suppresses the sintering of Fe oxides by generating ZnFe2O4 spinel [15], improving the catalyst’s stability [16]. Inert support such as Al2O3 and SiO2 can promote the dispersion of Fe species and suppress the aggregation of active iron particles due to the strong interactions between active phases and the support [17,18]. However, the active particles’ aggregation of the supported catalysts is still inevitable due to a lack of confinement [19].
Complicated metal oxides with superstructures, such as ABO3-type perovskites and AB2O4-type spinel, could confine transition metal ions’ A sites and B sites in the specific crystallite, therefore leading to a higher dispersion of the related metal ions at the atomic level over the supported metals [20,21]. Meanwhile, their activity can be adjusted flexibly using the substituted A site and B site with a transition metal [21]. On the one hand, these complicated metal oxides are well-known heterogeneous catalysts for redox reactions. As oxygen carriers, perovskite materials are applied widely in chemical-looping processes due to their excellent redox performances [22,23]. Perovskite materials including CaMnO3, SrFeO3, and LaFeO3 exhibited outstanding properties in CO2/H2O splitting to generate CO/H2 during chemical-looping dry/steam reforming [24,25,26,27]. On the other hand, these complicated metal oxides are prominent precursors in the preparation of metal nanoparticles for RWGS and FTS, thanks to the merits of the highly dispersed metal active site and its interaction with supports [28,29,30,31,32]. As for RWGS, a perovskite-derived catalyst provides oxygen vacancies for efficient CO2 adsorption and activation [28,29]. As for FTS, Ao et al. [30] achieved higher alcohol synthesis from syngas with a perovskite-derived trimetallic Co-Ni-Cu catalyst. Liu et al. [31] employed spinel ZnAl2O4-supported iron–catalyst for FTS, and gained an admirable light olefin selectivity of 64.4%. Ma et al. [32] reported a K/LaFeMnO3 perovskite-derived catalyst for FTS with a light olefin fraction of 54%. As for CO2 hydrogenation, Utsis et al. [33] adopted an Fe–Al–O spinel and an Fe–Ba–hexaaluminate as the catalyst precursor for CO2 hydrogenation, and yielded a high productivity of C5+ hydrocarbons. Recently, Xu et al. [34] fabricated a ternary spinel-type ZnCoxFe2-xO4 catalyst for an unprecedentedly high iron time yield for CO2 conversion of 29.1 μmol CO2·gFe−1·s−1. However, to the best of our knowledge, a perovskite-derived iron catalyst for CO2 hydrogenation to light olefins has been seldom reported.
To this end, a perovskite-type complicated metal oxide, SrFeO3, was firstly chosen as the precursor for CO2 hydrogenation to light olefins. SrFeO3 possesses the virtues of a relatively low cost without expressive elements, a high stability, and easily tunable chemical properties [35,36,37]. According to a report by Marek et al. [36], SrFeO3 could be reversibly reduced to SrO and Fe, resulting in the release of oxygen at a chemical potential suitable for splitting CO2/H2O. Its deep reduction in a CO2 atmosphere was susceptible to generating SrCO3; however, it could be reserved by increasing the temperature or lowering the CO2 partial pressure. These results highlighted the redox performance and structural stability of the SrFeO3 in a wide temperature range. Considering that potassium is a vital promoter that improves the reaction rate and increases the selectivity of olefins [38], the substitution of the A-site with potassium is a potential strategy to tailor the catalytic activity of SrFeO3 for both RWGS and FTS.
In the current work, a series of catalysts with perovskite-type Sr1-xKxFeO3(x = 0–0.6) as a precursor were prepared by using the sol-gel method. Characterizations including H2-TPR, XRD, XPS, and SEM demonstrated that K-substitution could lead to the formation of active-phase Fe3O4 for RWGS and Fe5C2 for FTS. The reversible ABO3-type perovskite structure ensured the high dispersion and stability of the active sites. Furthermore, the redox reaction of intermediate Sr2Fe2O5 gave rise to the dynamic regulation of CO2 by either splitting CO2 or forming SrCO3, thereby markedly promoting CO generation for FTS. The results shed light on the application of perovskite-type Sr1-xKxFeO3 in CO2 hydrogenation to light olefins.

2. Materials and Methods

2.1. Perovskite Preparation

The Sr1-xKxFeO3 perovskites (x = 0, 0.1, 0.2, 0.3, 0.4, 0.5, 0.6) were prepared by using the sol-gel method. The detailed procedures were as follows. Firstly, analytical reagents Sr(NO3)2, KNO3, and Fe(NO3)3·9H2O at the desired stoichiometric ratios; i.e., (1−x):x:1 (x = 0, 0.1, 0.2, 0.3, 0.4, 0.5, 0.6), were dissolved in deionized water and continuously stirred at 80 °C. Subsequently, citric acid and ethylene glycol with a molar ratio to metal ions of 1.5 was added into the solution to form a viscous gel. Then, the gel was dried at 100 °C for 24 h in an oven, followed by calcination at 400 °C for 2 h and 800 °C in air. The obtained particles were ground and sieved into 40~80 meshes for use. The perovskite samples of Sr1-xKxFeO3 with different K contents were respectively designated as SKFx.

2.2. Catalyst Characterization

The as-prepared and used catalysts were characterized by an X-ray diffractometer (XRD, Rigaku D-MAX 2500/PC) equipped with a Cu Κα radiation source (λ = 0.15406 nm). The 2θ range was from 10° to 80° at a scan rate of 5°/min in ambient conditions. The operation current and voltage were 150 mA and 40 kV, respectively. X-ray photoelectron spectra (XPS) were collected using a Thermo Fisher (Waltham, MA, USA) 250XI XPS spectrometer at the monochromatic Al Ka radiation and calibrated using the C peak at 284.6 eV. The deconvolution of the XPS spectra was carried out through the XPS PEAK 41 program with Gaussian functions by subtracting a Shirley background. The microscopic morphology of the catalyst particles was examined by TESCAN MIRA4 (Brno, Czech Republic) scanning electron microscopy (SEM) operated at an accelerating voltage of 0.2–30 keV. Energy-dispersive X-ray spectroscopy (EDS) mapping was examined using an Xplore facility with an operation voltage of 200 eV. H2 temperature-programmed reduction (H2-TPR) experiments were carried out on AutoChemiⅡ 2920 chemical adsorption instrument (Micromeritics Instrument Corporation, GA, USA), with a flow of 10 vol % H2 in argon (30 mL/min) from 100 to 900 °C at a constant heating rate of 10 °C/min. The CO2 temperature-programmed desorption (CO2-TPD) experiments were implemented on a Chembet Pulsar (Billing, MT, USA) chemical adsorption instrument. The sample (0.1 g) was first degassed at 400 ℃ for 1 h under an argon flow. After cooling to ambient temperature, CO2-TPD curves were obtained by heating up to 900 °C. Thermogravimetric analysis (TGA) was conducted by using NETZSCH STA (Selb, Germany) 409PC equipment under a 100 mL/min N2 or O2 (20%)/N2 atmosphere at a heating rate of 20 k/min. The specific surface area was obtained through the BET method.

2.3. Catalytic Performance Testing

The evaluation setup for the CO2 hydrogenation to olefins is shown in Figure 1. An appropriate amount of catalyst (2.0 g, 40–80 mesh) was loaded in the middle of the fixed-bed reactor with an inner diameter of 20 mm, and a small amount of quartz wool was filled at both ends to prevent the catalyst from being blown out. Before the activity test, the reactor was purged with argon for 30 min to remove the residual air, and then reduced to a flow of 10% H2/Argon at 450 °C for 2 h. At 350 °C, the reactant gas of H2/CO2 was fed into the reactor at 1 MPa and a GHSV of 7200 mL·gcat−1·h−1. The gaseous products with N2 as the internal standard were analyzed offline by using gas chromatography (PANNA A91 PLUS). CO, CH4, CO2, and N2 were detected with a TCD detector with a TDX-01 packed column, while the C1–C4 hydrocarbons were analyzed with an FID detector with a CP–Al2O3/KCl capillary column.
When the reaction was maintained continuously at least 4 h of time on stream, the CO2 conversion and product selectivities were determined using the following Equations (1)–(4):
X ( CO 2 ) = ( CO 2 ) in ( CO 2 ) out ( CO 2 ) in × 100 %
S ( CO ) = ( CO ) out ( CO 2 ) in ( CO 2 ) out × 100 %
S ( C 2 = C 4 = ) = ( C 2 = C 4 = ) out ( CO 2 ) in ( CO 2 ) out × 100 %
S ( C 5 + ) = 1 S ( CH 4 ) S ( CO ) S ( C 2 C 4 )

3. Results and Discussion

3.1. Characterization of As-Prepared Catalysts

The XRD patterns of the as-prepared SKFx (x = 0–0.6) catalysts are presented in Figure 2. All the SKFx (x = 0–0.6) catalysts displayed a typical diffraction peak of a cubic perovskite structure at 2θ = 32–34°, which was consistent with the literature reports [22,39]. There were no any other crystalline phases for the potassium-promoted catalysts, indicating that potassium successfully entered the lattice of the perovskite. Nevertheless, with an increasing K/Sr molar ratio, characteristic diffractions of the perovskite phase of all the K-substituting catalysts shifted gradually toward higher diffraction angles. This indicated that the entry of K into the A-site of the perovskites led to the changes in the lattice parameters. Since the valence of the K ion was lower and the ionic radius of K was larger than that of Sr, the intensities of the characteristic diffraction peaks over the modified perovskite decreased with an increase in the K substitution. In addition, the diffraction peak at 2θ = 46–48° was gradually divided into two peaks, and the degree of lattice distortion gradually increased. Moreover, SEM and EDS-mapping characterizations showed a high dispersion of all ions (Figure S1).
SrFeO3 is generally regarded as an oxygen-storage material due to its splendid redox performance. The O2-deficient state exhibits potential CO2/H2O splitting to generate CO/H2. Therefore, the O2-releasing behaviors of the as-prepared SKFx (x = 0–0.6) catalysts were investigated by thermogravimetric analysis. As shown in Figure 3, all as-prepared SKFx (x = 0–0.6) catalysts could release O2 at the evaluated temperature under an inert atmosphere. As for SKF0.0; i.e., SrFeO3, three O2-releasing phases were observed with an increase in the temperature that corresponded to three crystal structures. The calculated mass loss meant that the three crystal structures included Sr8Fe8O23 (SrFeO2.875) below 300 °C, Sr4Fe4O11 (SrFeO2.75) at 300–600 °C, and Sr2Fe2O5 (SrFeO2.5) above 600 °C, which agreed with the experimental results in the literature [40]. However, further decomposition of the brownmillerite structure Sr2Fe2O5 into SrO and Fe was not observed. According to the reported literature [22,36], only under a reducing atmosphere could the deep reduction of SrO and Fe occur. As for SKFx (x = 0.2, 0.4 and 0.6), the substitution of Sr with K significantly enhanced the release of O2, thereby promoting the generation of SrO and Fe. At 300–600 °C, the O2-releasing molar ratio approached 30 %, while this value was just 8.3 % for SrFeO3. The oxygen vacancy resulted from O2 releasing was conductive to the adsorption and conversion of CO2 into CO; i.e., RWGS. At the temperature of CO2 hydrogenation to light olefins, the SKF0.4 catalyst displayed the highest oxygen vacancy. In addition, the final mass loss increased gradually with the increase in the substitution amount of K. The O2-releasing molar ratio was close to 60% for the SKF0.6 catalyst, which suggested the total release of oxygen, and that the ultimate products were SrO, K2O, and Fe.
To reveal the redox behaviors of the SKFx catalysts in depth, H2-TPR experiments were carried out. As depicted in Figure 4, there were three distinct reduction peaks for all samples, in the vicinity of 450 °C, 600 °C, and 800 °C, respectively. They represented the reduction of SrFeO3 to Sr2Fe2O5, the further reduction of Sr2Fe2O5, and the formation of SrO and Fe, which were in line with the reported literature [22,36]. Combined with the aforementioned TG results, the reducing gas H2 accelerated the release behavior of oxygen, and achieves the deep reduction to SrO and Fe, even at 800 °C. According to [9], alkali metals act as electron promoters to inhibit the chemisorption of H2 on the catalyst surface, restricting the reduction of metal oxide ions. Thus, the first reduction peak of the SKFx (x = 0.2, 0.4, 0.6) catalysts was extended to a higher temperature. Meanwhile, the consumption of H2 decreased when is the temperature was lower than 450 °C. Nevertheless, among the three K-substituting catalysts, the SKF0.4 exhibited the maximum consumption of H2 in the temperature range to conduct CO2 hydrogenation (200–450 °C).
To obtain insights into the basic site distribution of the SKFx (x = 0, 0.2, 0.4) catalysts, CO2-TPD experiments were conducted. As shown in Figure 5, there were two basic sites on the surface of the SrFeO3 at 450 °C and 600 °C. However, the K-substituting SKF0.2 catalyst had a sharp desorption peak at 665 °C, while the SKF0.4 catalyst had two distinct sharp desorption peaks at 697 °C and 855 °C. Generally, the desorption peaks below 600 °C reflected the basic site distribution. The desorption peaks of the K-substituting catalysts at higher temperatures showed that the reactions between the K-substituting catalysts and CO2 took place. According to the following experiments and those in [22,41], we know that the reaction 5 between the generated Sr2Fe2O5 and CO2 occurred. The CO2 desorption peak at 665 °C for SKF0.2 and 697 °C for SKF0.4 were attributed to the reverse reaction 5, while the CO2 desorption peak at 855 °C was ascribed to the decomposition of SrCO3 to generate SrO and CO2. It should be pointed out that the K-substituting catalysts still possessed certain basic sites.
Sr2Fe2O5 + 2CO2= 2SrCO3 + Fe2O3

3.2. Catalytic Performance of SKFx Catalysts

To investigate the effect of K-substituting content on the catalytic performances, CO2 catalytic hydrogenation experiments using SKFx catalysts were carried out in a fixed-bed reactor. The CO2 conversion and product distribution using the SKFx catalysts are indicated in Figure 6. When SrFeO3 was used as the catalyst, the CO selectivity was up to 83.92%; however, few FTS products were generated. The main reason was a lack of the active iron carbide phase of Fe5C2 due to the confinement of the perovskite structure to B-site Fe ions. In addition, the high CO selectivity was ascribed to the outstanding redox performance of reduced SrFeO3; i.e., Sr2Fe2O5, which promoted the dissociation of CO2 into CO. With the substitution of Sr with K at the A-site, the distortion of the lattice gradually increased, and the confinement of the perovskite structure to Fe was weakened. Consequently, the active phases, such as Fe3O4 and Fe2C5, were prone to generation, followed by the increases in the CO2 conversion and FTS products. On the other hand, the alkalinity of the catalyst surface increased with the increasing substitution of K, which could promote CO2 adsorption and inhibit H2 adsorption indirectly [42], thereby promoting the accession of light-olefin selectivity. However, lower K substitution reduced the amount of active phases generated. More K substitution took the edge off both the CO2 conversion and the selectivity of light olefins, which could be attributed to the covering of the active site [12]. The optimal substitution amount was 0.4. As for the SKF0.4 catalyst, the highest light-olefin selectivity of 29.61% and CO2 conversion of 30.82% were achieved.
The composition of the SKFx catalysts after 4 h of reaction was detected using XRD analysis. As depicted in Figure 7, the SrFeO3 catalyst comprised an oxygen-deficient brownmillerite structure, Sr2Fe2O5, without any other phases. However, after the substitution of Sr with K at the A-site, the brownmillerite structure Sr2Fe2O5 disappeared. Instead, the Sr-containing phase became SrCO3, while the Fe-containing phase was composed of Fe3O4, Fe5C2, and Fe2O3, among which Fe3O4 and Fe5C2 were acknowledged as the active phases for RWGS and FTS, respectively. According to Marek’s research [36], only under a reduction atmosphere and an elevated CO2 pressure could SrFeO3 carbonate to SrCO3. Furthermore, they argued that SrCO3 could be removed by oxidation or avoided entirely at higher temperatures (>850 °C). Obviously, the XRD analysis indicated that K substitution made SrFeO3 more susceptible to carbonation. The Fe2O3 phase was accompanied by a carbonation process, referring to reaction (5). The generated carbonate weakened the confinement of the perovskite structure to iron ions, thereby promoting the generation of Fe3O4 and Fe5C2. As for the SKF0.6 catalyst, the other iron phase, Fe3C, was detected, which was attributed to the deep carbonization. This result could explain the excellent catalytic performance of the SKF0.4 over the SKF0.6 in CO2 hydrogenation. Despite the SrFeO3 being converted to SrCO3 and an iron-containing phase, a high dispersion of the active phase still could be achieved. First, the SKFx catalysts could regenerate under a high-temperature and oxidation atmosphere according to the reversible reaction (5), which will be further elaborated in the following section. Second, the elements were confined within the specific structure, thereby retaining its high dispersion. The surface topographies of the SrFO3 and SKF0.4 catalysts and their element distributions are shown in Figure 8. It is clear that the surface of both fresh SrFO3 and SKF0.4 catalysts were smooth and porous, and without any agglomeration. Compared with the SrFO3, the porosity of the fresh SKF0.4 catalyst decreased; however, it increased after the catalytic reaction. This could be attributed to their specific surface areas. This value was 6.9932 m2·g−1 for the fresh SrFO3 and 3.8366 m2·g−1 for the fresh SKF0.4, respectively; while it was 5.849 m2·g−1 for the reacted SKF0.4. Furthermore, elements including Fe, Sr, K, and O were dispersed evenly on the perovskite structure, as shown in their EDS mappings in Figure 8D–H. Meanwhile, the uniform distribution of the C element, which existed in the form of FeCx, was also observed.

3.3. Redox Performance of Catalysts

Aside from the catalytic effect of the active phase Fe3O4 and Fe5C2 on RWGS and FTS, the inherent redox properties of the SrFO3 and Sr2Fe2O5 played significant roles in the CO2-splitting reaction and the H2-oxidation reaction. Thus, the change in valence states and the oxidation behavior of the reacted SKFx were investigated.
As illustrated in Figure 9, the Fe2p spectra of both the fresh SrFeO3 and SKF0.4 had double peaks and two small satellite peaks, indicating the different chemical valence states of Fe [43]. The binding energies of 709.2 eV, 710.4 eV, and 711 eV corresponded to Fe2+, Fe3+, and Fe4+, respectively. The deconvolution of the O 1s spectra involved three peaks: lattice oxygen (Olat) at 528 eV, chemisorbed oxygen (Ocar) at 529 eV, and physisorbed oxygen (Oads) at 530.5 eV [44]. The component contents calculated from the integrated area of the subpeaks are tabulated in Table 1. As for the fresh catalysts, in light of the electroneutrality principle, the substitution of Sr with K increased the content of Fe4+, which had a stronger reducibility. The increase in Fe4+ might have been responsible for the formation of iron carbide. The decreases in Olat and Ocar and the increase in Oads represented the rising oxygen vacancies and oxygen mobility, which benefited the redox performance and the formation of iron carbide during the CO2 hydrogenation reaction.
The XPS spectra of both the SrFeO3 and SKF0.4 after the reaction were analyzed (Figure 10). As for the SrFeO3 catalyst, the increasing peak area of the Fe2+ and hydroxyl oxygen Ocar indicated that partial Fe4+ conversion into Fe3+ and Fe2+ occurred under the reductive atmosphere, which corresponded to the generation of Sr2Fe2O5 (Figure 7). As for the reacted SKF0.4, three types of oxygen existed on its surface, among which the peaks at 529.1, 530.9, and 532.6 eV corresponded to the binding energies of the Fe3O4, Fe2O3, and SrCO3 respectively. Within the binding-energy region of the Fe 2p3/2 core level, the peaks located at 720.15 eV (Fe 2p1/2) and 707.3 eV (Fe 2p3/2) were attributed to the FexCy species. The dominance of the broad peak centered around 710 eV was caused by the overlapping signals of Fe3+ and Fe2+, which suggested that the Fe3O4 species was present as the main component on the surface of these Fe catalysts.
To conduct the regeneration of the perovskite structure, the oxidation behaviors of the SKFx (x = 0–0.6) catalysts after reaction were investigated using thermogravimetric analysis under an O2 atmosphere. As shown in Figure 11, all the used SKFx (x = 0–0.6) catalysts underwent three stages: a mass-loss stage I at around 300 °C, followed by a mass-gain stage II, and then a sharp mass-loss stage III above 600 °C. Compared with the constant mass of the SrFeO3, stage I was attributed to the oxidation of FeCx. The maximum mass loss of SKF0.4 manifested the highest amount of FeCx forms in the catalytic process. The following mass-gain stage II resulted from the oxygen-obtaining process of the oxygen-deficient perovskite. This suggested the excellent regeneration of the used catalyst to its perovskite structure. It was surprising to observe the occurrence of O2 release, even under the O2 atmosphere. As distinguished from the mass-loss behavior of the fresh catalysts under an inert atmosphere, the sharp mass loss herein was caused by the reaction (5) between the SrCO3 and Fe2O3 to generate Sr2Fe2O5 and CO2. Meanwhile, the decomposition of Sr2Fe2O5 occurred due to K-substituting promotions. This highlighted that the gain of oxygen and loss of K-substituting catalysts verified the remarkable stability of the perovskite structure.

3.4. Probable Reaction Mechanism over K-Substituting Catalysts

To determine the influence factor and probable reaction mechanism of the SKF0.4 catalyst, different reaction conditions, including the reaction atmosphere and the H/C molar ratio, were examined. The CO2/CO conversion and product distributions using SKF0.4 are tabulated in Table 2. When using CO2 (25%) as the reaction gas, the CO2 conversion was 6.54%, with a CO selectivity of 86.22%. On the one hand, the CO2-splitting reaction to generate CO occurred when using the reduced catalyst existing in the form of Sr2Fe2O5 (Figure 12). On the other hand, CO2 was deeply reduced to form FeCx, accompanied by the production of SrCO3. The absence of the H2 reduction gas gave rise to the deep carbonization of iron, resulting in the formation of Fe7C3. The XRD patterns (Figure 12) of the catalysts after the reaction confirmed this inference.
To unveil the catalytic properties in FTS, the reactions using the SKF0.4 catalyst with H2 and CO as the reaction gas were separately investigated. We obtained 86.36% CO conversion, but only 22.57% CH4 and 22.1% light olefins. The selectivity of CO2 up to 35.02% indicated the intense WGS reaction. Furthermore, the selectivity of light olefins had a limited increase with the increase in H2/CO. As the XRD showed, the active phases of Fe5C2 and Fe3O4 were still generated with H2 and CO as the reaction gas, the same as that under the CO2 atmosphere. Obviously, the K-substituted SrFeO3 enhanced the generation of the active phase for both the RWGS and FTS in the presence of CO or CO2. With H2 and CO2 as the reaction gas, a high CO2 partial pressure restricted the WGS, resulting in higher light olefins (up to 29.61%). It should be pointed out that the CO selectivity was as high as 45.57%, even when the H2/CO2 molar ratio was equal to 4. Except for the catalysis of Fe3O4 on the RWGS, the redox properties of the SKFx catalysts elevated the conversion of CO2 to CO. In addition, the FTS was the rate-determining step of the CO2 hydrogenation reaction to light olefins with the SKFx perovskite catalysts, which should be enhanced in the future.
An ABO3-type perovskite SrFeO3 could be stepwise reduced to Sr2Fe2O5, and further to SrO and Fe, on the premise of being completely reversible, and oxidized to its perovskite structure. The H2O/CO2 splitting over SrO and Fe could produce H2/CO via the reversible reaction (6) and (7). However, the realization of a deep reduction required a high temperature. The A-site substitution of Sr with K increased the lattice distortion, thereby decreasing the confinement of ions within the crystal structure. The increasing oxygen vacancy improved the adsorption of CO2 on catalysts, furthering the generation of SrCO3 and Fe2O3 via reaction (5). Generally, the formation of CO in the RWGS reaction has two controversial paths; i.e., a redox mechanism with the dissociation of adsorbed CO2 as the rate-determining step, and the intermediates mechanism of carbonate and formate species [44]. In this work, apart from the catalytic properties of the acknowledged Fe3O4, both the redox effect of Sr2Fe2O5 on the CO2 splitting and the formation of SrCO3 carbonates were responsible for the enhanced RWGS. In addition, the reversible reaction (5) could regulate the CO2 equilibrium to accommodate the FTS. The probable reaction mechanism of CO2 hydrogenation to light olefins using K-substituting SKF0.4 catalysts is displayed in Figure 13. After being reduced by H2, the O2-deficient crystal structure Sr2Fe2O5 was formed. Under the impact of K-substitution, Sr2Fe2O5 reacted with CO2 to generate SrCO3 and Fe2O3, and then produced the active-phase Fe3O4 and Fe5C2. Furthermore, the redox properties of Sr2Fe2O5 and the carbonate SrCO3 promoted the RWGS. In addition, the reversible reaction (5) ensured the structural stability and high dispersion of the active phase. In sum, the SKFx perovskites exhibited immense potential for CO2 hydrogenation to light olefins.
2SrO + 2Fe + 3CO2 = Sr2Fe2O5 + 3CO
2SrO + 2Fe + 3H2O = Sr2Fe2O5 + 3H2

4. Conclusions

In light of their favorable thermal stabilities and redox activities, ABO3-type perovskite-derived catalysts of Sr1-xKxFeO3 with a highly dispersed active site and enhanced RWGS reactivities were reported in this work. The following conclusions were drawn.
(1)
K-substitution using SKFx perovskite catalysts could promote the production of active-phase Fe3O4 for the RWGS and Fe5C2 for the FTS. SKF0.4 displayed the optimal CO2 conversion of 30.82% and light olefin selectivity of 29.61%.
(2)
Apart from the catalytic properties of the acknowledged Fe3O4, both the redox effect of the Sr2Fe2O5 on the CO2 splitting and the formation of SrCO3 carbonates were responsible for the enhanced RWGS.
(3)
The reversible reaction Sr2Fe2O5 + 2CO2 = 2SrCO3 + Fe2O3 ensured the structural stability and high dispersion of the active phase. As the rate-determining step of the CO2 hydrogenation reaction to light olefins using SKFx perovskite catalysts, the FTS should be enhanced in the future by partially substituting the B-site of perovskite-type catalysts.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/atmos13050760/s1, Figure S1: SEM and EDS mapping of fresh SKF0.0 and SKF0.4.

Author Contributions

Conceptualization, Y.H. and Y.L.; Methodology, X.W.; Validation, X.W. and M.C.; Investigation, Y.H.; Resources, Y.H.; Data Curation, X.G.; Writing—Original Draft Preparation, Y.H.; Writing—Review and Editing, Y.L.; Visualization, Y.H.; Supervision, Q.G.; Project Administration, Y.L. and Q.G.; Funding Acquisition, Q.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Joint Funds of the National Natural Science Foundation of China (U20A20124), the Shandong Provincial Natural Science Foundation (ZR2020MB144), and the State Key Laboratory of High-Efficiency Utilization of Coal and Green Chemical Engineering (2022-K3).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Porosoff, M.D.; Yan, B.; Chen, J.G. Catalytic reduction of CO2 by H2 for synthesis of CO, methanol and hydrocarbons: Challenges and opportunities. Energy Environ. Sci. 2016, 9, 62–73. [Google Scholar] [CrossRef]
  2. Tackett, B.M.; Gomez, E.; Chen, J.G. Net reduction of CO2 via its thermocatalytic and electrocatalytic transformation reactions in standard and hybrid processes. Nat. Catal. 2019, 2, 381–386. [Google Scholar] [CrossRef]
  3. Ojelade, O.A.; Zaman, S.F. A review on CO2 hydrogenation to lower olefins: Understanding the structure-property relationships in heterogeneous catalytic systems. J. CO2 Util. 2021, 47, 101506. [Google Scholar] [CrossRef]
  4. Wang, D.; Xie, Z.; Porosoff, M.D.; Chen, J.G. Recent advances in carbon dioxide hydrogenation to produce olefins and aromatics. Chem 2021, 7, 2277–2311. [Google Scholar] [CrossRef]
  5. Atsbha, T.A.; Yoon, T.; Park, S.; Lee, C.-J. A review on the catalytic conversion of CO2 using H2 for synthesis of CO, methanol, and hydrocarbons. J. CO2 Util. 2021, 44, 101413. [Google Scholar] [CrossRef]
  6. Numpilai, T.; Cheng, C.K.; Limtrakul, J.; Witoon, T. Recent advances in light olefins production from catalytichydrogenation of carbon dioxide. Process. Saf. Environ. 2021, 151, 401–427. [Google Scholar] [CrossRef]
  7. Li, Z.; Wang, J.; Qu, Y.; Liu, H.; Tang, C.; Miao, S.; Feng, Z.; An, H.; Li, C. Highly selective conversion of carbon dioxide to lower olefins. ACS Catal. 2017, 7, 8544–8548. [Google Scholar] [CrossRef]
  8. Centi, G.; Quadrelli, E.A.; Perathoner, S. Catalysis for CO2 conversion: A key technology for rapid introduction of renewable energy in the value chain of chemical industries. Energy Environ. Sci. 2013, 6, 1711–1731. [Google Scholar] [CrossRef]
  9. Boreriboon, N.; Jiang, X.; Song, C.; Prasassarakich, P. Higher hydrocarbons synthesis from CO2 hydrogenation over K- and La-promoted Fe–Cu/TiO2 catalysts. Top. Catal. 2018, 61, 1551–1562. [Google Scholar] [CrossRef]
  10. Wei, J.; Sun, J.; Wen, Z.; Fang, C.; Ge, Q.; Xu, H. New insights into the effect of sodium on Fe3O4-based nanocatalysts for CO2 hydrogenation to light olefins. Catal. Sci. Technol. 2016, 6, 4786–4793. [Google Scholar] [CrossRef]
  11. Liu, R.; Ma, Z.; Sears, J.D.; Juneau, M.; Neidig, M.L.; Porosoff, M.D. Identifying correlations in Fischer-Tropsch synthesis and CO2 hydrogenation over Fe-based ZSM-5 catalysts. J. CO2 Util. 2020, 41, 101290. [Google Scholar] [CrossRef]
  12. Numpilai, T.; Chanlek, N.; Poo-Arporn, Y.; Cheng, C.K.; Siri-Nguan, N.; Sornchamni, T.; Chareonpanich, M.; Kongkachuichay, P.; Yigit, N.; Rupprechter, G. Tuning interactions of surface-adsorbed species over Fe-Co/K-Al2O3 catalyst by different K contents: Selective CO2 hydrogenation to light olefins. ChemCatChem 2020, 12, 3306–3320. [Google Scholar] [CrossRef]
  13. Zhu, M.; Tian, P.; Ford, M.E.; Chen, J.; Xu, J.; Han, Y.; Wachs, I.E. Nature of reactive oxygen intermediates on copper-promoted iron-chromium oxide catalysts during CO2 activation. ACS Catal. 2020, 10, 7857–7863. [Google Scholar] [CrossRef]
  14. Witoon, T.; Numpilai, T.; Dolsiririttigul, N.; Chanlek, N.; Poo-arporn, Y.; Cheng, C.K.; Ayodele, B.V.; Chareonpanich, M.; Limtrakul, J. Enhanced activity and stability of SO42−/ZrO2 by addition of Cu combined with CuZnOZrO2 for direct synthesis of dimethyl ether from CO2 hydrogenation. Int. J. Hydrogen Energy 2022, in press. [Google Scholar] [CrossRef]
  15. Zhang, J.; Lu, S.; Su, X.; Fan, S.; Ma, Q.; Zhao, T. Selective formation of light olefins from CO2 hydrogenation over Fe–Zn– K catalysts. J. CO2 Util. 2015, 12, 95–100. [Google Scholar] [CrossRef]
  16. Witoon, T.; Chaipraditgul, N.; Numpilai, T.; Lapkeatseree, V.; Ayodele, B.V.; Cheng, C.K.; Siri-Nguan, N.; Sornchamni, T.; Limtrakul, J. Highly active Fe-Co-Zn/K-Al2O3 catalysts for CO2 hydrogenation to light olefins. Chem. Eng. Sci. 2021, 233, 116428. [Google Scholar] [CrossRef]
  17. Numpilai, T.; Chanlek, N.; Poo-Arporn, Y.; Wannapaiboon, S.; Cheng, C.K.; Siri-Nguan, N.; Sornchamni, T.; Kongkachuichay, P.; Chareonpanich, M.; Rupprechter, G.; et al. Pore size effects on physicochemical properties of Fe-Co/K-Al2O3 catalysts and their catalytic activity in CO2 hydrogenation to light olefins. Appl. Surf. Sci. 2019, 483, 581–592. [Google Scholar] [CrossRef]
  18. Rafati, M.; Wang, L.; Shahbazi, A. Effect of silica and alumina promoters on coprecipitated Fe–Cu–K based catalysts for the enhancement of CO2 utilization during Fischer–Tropsch synthesis. J. CO2 Util. 2015, 12, 34–42. [Google Scholar] [CrossRef] [Green Version]
  19. Price, C.A.H.; Reina, T.R.; Liu, J. Engineering heterogenous catalysts for chemical CO2 utilization: Lessons from thermal catalysis and advantages of yolk@shell structured nanoreactors. J. Energy Chem. 2021, 57, 304–324. [Google Scholar] [CrossRef]
  20. Ren, J.; Mebrahtu, C.; van Koppen, L.; Martinovic, F.; Hofmann, J.P.; Hensen, E.J.M.; Palkovits, R. Enhanced CO2 methanation activity over La2-xCexNiO4 perovskite-derived catalysts: Understanding the structure-performance relationships. Chem. Eng. J. 2021, 426, 131760. [Google Scholar] [CrossRef]
  21. Yang, Q.; Liu, G.; Liu, Y. Perovskite-Type Oxides as the Catalyst Precursors for Preparing Supported Metallic Nanocatalysts: A Review. Ind. Eng. Chem. Res. 2018, 57, 1–17. [Google Scholar] [CrossRef]
  22. Lau, C.Y.; Dunstan, M.T.; Hu, W.; Grey, C.P.; Scott, S.A. Large scale in silico screening of materials for carbon capture through chemical looping. Energy Environ. Sci. 2017, 10, 818. [Google Scholar] [CrossRef] [Green Version]
  23. Zhu, X.; Imtiaz, Q.; Donat, F.; Müller, C.R.; Li, F. Chemical looping beyond combustion—A perspective. Energy Environ. Sci. 2020, 13, 772–804. [Google Scholar] [CrossRef] [Green Version]
  24. Pour, N.M.; Azimi, G.; Leion, H.; Rydén, M.; Lyngfelt, A. Production and examination of oxygen-carrier materials based on manganese ores and Ca(OH)2 in chemical looping with oxygen uncoupling. AIChE J. 2014, 60, 645–656. [Google Scholar] [CrossRef] [Green Version]
  25. Ezbiri, M.; Allen, K.M.; Gàlvez, M.E.; Michalsky, R.; Steinfeld, A. Design principles of perovskites for thermochemical oxygen separation. ChemSusChem 2015, 8, 1966–1971. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Mihai, O.; Chen, D.; Holmen, A. Catalytic consequence of oxygen of lanthanum ferrite perovskite in chemical looping reforming of methane. Ind. Eng. Chem. Res. 2011, 50, 2613–2621. [Google Scholar] [CrossRef]
  27. Liu, Y.; Liu, H.; Wang, X.; Ji, X.; Wei, J.; Zhang, J. Orthogonal Preparation of SrFeO3-δ Nanocomposites as Effective Oxygen Transfer Agents for Chemical-Looping Steam Methane Reforming. Energy Fuels 2021, 35, 17848–17860. [Google Scholar] [CrossRef]
  28. Lindenthal, L.; Popovic, J.; Rameshan, R.; Huber, J.; Schrenk, F.; Ruh, T.; Nenning, A.; Loffler, S.; Opitz, A.K.; Rameshan, C. Novel perovskite catalysts for CO2 utilization—Exsolution enhanced reverse water-gas shift activity. Appl. Catal. B Environ. 2021, 292, 120183. [Google Scholar] [CrossRef]
  29. Daza, Y.A.; Kent, R.A.; Yung, M.M.; Kuhn, J.N. Carbon Dioxide Conversion by Reverse Water−Gas Shift Chemical Looping on Perovskite-Type Oxides. Ind. Eng. Chem. Res. 2014, 53, 5828–5837. [Google Scholar] [CrossRef]
  30. Ao, M.; Pham, G.H.; Sage, V.; Pareek, V.; Liu, S. Perovskite-derived trimetallic Co-Ni-Cu catalyst for higher alcohol synthesis from syngas. Fuel Process. Technol. 2019, 193, 141–148. [Google Scholar] [CrossRef]
  31. Liu, Z.; Jia, G.; Zhao, C.; Xing, Y. Selective Iron Catalysts for Direct Fischer−Tropsch Synthesis to Light Olefins. Ind. Eng. Chem. Res. 2021, 60, 6137–6146. [Google Scholar] [CrossRef]
  32. Ma, L.; Gao, X.; Ma, J.; Hu, X.; Zhang, J.; Guo, Q. K/LaFeMnO3 Perovskite-Type Oxide Catalyst for the Production of C2–C4 Olefins via CO Hydrogenation. Catal. Lett. 2022, 152, 1451–1460. [Google Scholar] [CrossRef]
  33. Utsis, N.; Vidruk-Nehemya, R.; Landau, M.V.; Herskowitz, M. Novel bifunctional catalysts based on crystalline multi-oxide matrices containing iron ions for CO2 hydrogenation to liquid fuels and chemicals. Faraday Discuss. 2016, 188, 545. [Google Scholar] [CrossRef] [PubMed]
  34. Xu, Q.; Xu, X.; Fan, G.; Yang, L.; Li, F. Unveiling the roles of Fe-Co interactions over ternary spinel-type ZnCoxFe2−xO4 catalysts for highly efficient CO2 hydrogenation to produce light olefins. J. Catal. 2021, 400, 355–366. [Google Scholar] [CrossRef]
  35. Yu, W.; Wang, X.; Liu, Y.; Wei, J.; Zhang, J. Effect of Composition on the Redox Performance of Strontium Ferrite Nanocomposite. Energy Fuels 2020, 34, 8644–8652. [Google Scholar] [CrossRef]
  36. Marek, E.; Hu, W.; Gaultoisd, M.; Grey, C.; Scott, S.A. The use of strontium ferrite in chemical looping systems. Appl. Energy 2018, 223, 369–382. [Google Scholar] [CrossRef]
  37. Wang, X.; Du, X.; Yu, W.; Zhang, J.; Wei, J. Coproduction of Hydrogen and Methanol from Methane by Chemical Looping Reforming. Ind. Eng. Chem. Res. 2019, 58, 10296–10306. [Google Scholar] [CrossRef]
  38. Meiri, N.; Dinburg, Y.; Amoyal, M.; Koukouliev, V.; Nehemya, R.V.; Landau, M.V.; Herskowitz, M. Novel process and catalytic materials for converting CO2 and H2 containing mixtures to liquid fuels and chemicals. Faraday Discuss. 2015, 183, 197. [Google Scholar] [CrossRef]
  39. Takeda, Y.; Kanno, K.; Takada, T.; Yamamoto, O.; Takano, M.; Nakayama, N.; Bando, Y. Phase relation in the oxygen nonstoichiometric system, SrFeOx (2.5 ≤ x ≤ 3.0). J. Solid State Chem. 1986, 63, 237–249. [Google Scholar] [CrossRef]
  40. Ikeda, H.; Nikata, S.; Hirakawa, E.; Tsuchida, A.; Miura, N. Oxygen sorption/desorption behavior and crystal structural change for SrFeO3-δ. Chem. Eng. Sci. 2016, 147, 166–172. [Google Scholar] [CrossRef]
  41. Gallagher, P.K.; Sanders, J.P.; Woodward, P.M.; Lokuhewa, I.N. Reactions within the system, 2SrCO3–Fe2O3 Part I. CO2 atmosphere, J. Therm. Anal. Calorim. 2005, 80, 217–223. [Google Scholar] [CrossRef]
  42. Ngantsoue-Hoc, W.; Zhang, Y.; O’Brien, R.; Luo, M.; Davis, B.H. Fischer-Tropsch synthesis: Activity and selectivity for Group I alkali promoted iron-based catalysts. Appl. Catal. A-Gen. 2002, 236, 77–89. [Google Scholar] [CrossRef]
  43. Tian, X.; Zheng, C.; Zhao, H. Ce-modified SrFeO3-δ for ethane oxidative dehydrogenation coupled with CO2 splitting via a chemical looping scheme. Appl. Catal. B Environ. 2022, 303, 120894. [Google Scholar] [CrossRef]
  44. Zhang, Z.; Liu, Y.; Jia, L.; Sun, C.; Chen, B.; Liu, R.; Tan, Y.; Tu, W. Effects of the reducing gas atmosphere on performance of FeCeNa catalyst for the hydrogenation of CO2 to olefins. Chem. Eng. J. 2022, 428, 131388. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of reactor for carbon dioxide hydrogenation to prepare light olefins: (a) pressure gauge; (b) mass flowmeter; (c) preheater; (d) temperature controller; (e) thermocouple and temperature display; (f) silica wool; (g) catalyst; (h) gas–liquid separator; (i) gas chromatograph; (j) computer; (k) fixed-bed reactor.
Figure 1. Schematic diagram of reactor for carbon dioxide hydrogenation to prepare light olefins: (a) pressure gauge; (b) mass flowmeter; (c) preheater; (d) temperature controller; (e) thermocouple and temperature display; (f) silica wool; (g) catalyst; (h) gas–liquid separator; (i) gas chromatograph; (j) computer; (k) fixed-bed reactor.
Atmosphere 13 00760 g001
Figure 2. XRD patterns of the as-prepared SKFx (x = 0–0.6) catalysts: (A) full pattern; (B) at 2θ = 46–48°.
Figure 2. XRD patterns of the as-prepared SKFx (x = 0–0.6) catalysts: (A) full pattern; (B) at 2θ = 46–48°.
Atmosphere 13 00760 g002
Figure 3. Thermogravimetric analysis of the SKFx (x = 0, 0.2, 0.4, 0.6) catalysts from ambient temperature to 900 °C under the inert atmosphere.
Figure 3. Thermogravimetric analysis of the SKFx (x = 0, 0.2, 0.4, 0.6) catalysts from ambient temperature to 900 °C under the inert atmosphere.
Atmosphere 13 00760 g003
Figure 4. H2-TPR profiles of SKFx (x = 0, 0.2, 0.4, 0.6) perovskite catalysts.
Figure 4. H2-TPR profiles of SKFx (x = 0, 0.2, 0.4, 0.6) perovskite catalysts.
Atmosphere 13 00760 g004
Figure 5. CO2-TPD profiles of SKFx (x = 0, 0.2, 0.4, 0.6) perovskite catalysts.
Figure 5. CO2-TPD profiles of SKFx (x = 0, 0.2, 0.4, 0.6) perovskite catalysts.
Atmosphere 13 00760 g005
Figure 6. CO2 conversion and product distribution over SKFx catalysts (reaction conditions: H2/CO2 = 3/1, 350 ℃, 1.0 MPa, GHSV = 7200 mL·gcat·−1·h−1).
Figure 6. CO2 conversion and product distribution over SKFx catalysts (reaction conditions: H2/CO2 = 3/1, 350 ℃, 1.0 MPa, GHSV = 7200 mL·gcat·−1·h−1).
Atmosphere 13 00760 g006
Figure 7. XRD patterns for the SKFx catalysts after reaction (H2/CO2 = 3/1, 350 °C, 1.0 MPa, GHSV = 7200 mL·gcat·−1·h−1).
Figure 7. XRD patterns for the SKFx catalysts after reaction (H2/CO2 = 3/1, 350 °C, 1.0 MPa, GHSV = 7200 mL·gcat·−1·h−1).
Atmosphere 13 00760 g007
Figure 8. SEM patterns of perovskite catalysts: (A) SKF0.0 (fresh); (B) SKF0.4 (fresh); (C) SKF0.4 (after reaction). (DH) EDS mapping.
Figure 8. SEM patterns of perovskite catalysts: (A) SKF0.0 (fresh); (B) SKF0.4 (fresh); (C) SKF0.4 (after reaction). (DH) EDS mapping.
Atmosphere 13 00760 g008
Figure 9. O 1s and Fe 2p XPS spectra and deconvolution of fresh samples: (a,c) SKF0.0; (b,d) SKF0.4.
Figure 9. O 1s and Fe 2p XPS spectra and deconvolution of fresh samples: (a,c) SKF0.0; (b,d) SKF0.4.
Atmosphere 13 00760 g009
Figure 10. O 1s and Fe 2p XPS spectra and deconvolution of spend samples: (a,c) SKF0.0; (b,d) SKF0.4 (H2/CO2 = 3/1, 350 °C, 1.0 MPa, GHSV = 7200 mL·gcat·−1·h−1).
Figure 10. O 1s and Fe 2p XPS spectra and deconvolution of spend samples: (a,c) SKF0.0; (b,d) SKF0.4 (H2/CO2 = 3/1, 350 °C, 1.0 MPa, GHSV = 7200 mL·gcat·−1·h−1).
Atmosphere 13 00760 g010
Figure 11. Thermogravimetric analysis of the used SKFx (x = 0, 0.2, 0.4, 0.6) catalysts when heated to 900 °C under 20% O2 atmosphere.
Figure 11. Thermogravimetric analysis of the used SKFx (x = 0, 0.2, 0.4, 0.6) catalysts when heated to 900 °C under 20% O2 atmosphere.
Atmosphere 13 00760 g011
Figure 12. XRD patterns for SKF0.4 catalysts after reduction under different conditions (350 °C, 1.0 MPa).
Figure 12. XRD patterns for SKF0.4 catalysts after reduction under different conditions (350 °C, 1.0 MPa).
Atmosphere 13 00760 g012
Figure 13. Mechanism of CO2 hydrogenation on SKFx perovskite to generate products.
Figure 13. Mechanism of CO2 hydrogenation on SKFx perovskite to generate products.
Atmosphere 13 00760 g013
Table 1. Percentage of O1s and Fe2p3/2 of fresh perovskites.
Table 1. Percentage of O1s and Fe2p3/2 of fresh perovskites.
SampleFe2p3/2O1s
Fe2+/%Fe3+/%Fe4+/%Olat/%Ocar/%Oads/%
SKF0.0 (fresh)20.4153.8525.7420.8121.6357.56
SKF0.4 (fresh)22.2444.633.1615.3614.7469.9
SKF0.0 (used)28.0249.2322.7515.3132.6441.35
Table 2. CO2/CO conversion and product distribution using SKF0.4 under different conditions (350 °C, 1.0 MPa).
Table 2. CO2/CO conversion and product distribution using SKF0.4 under different conditions (350 °C, 1.0 MPa).
ConditionsH/CConversion (%)Selectivity (%)
CO2COCOCO2CH4 C 2 0 C 4 0 C 2 = C 4 = S/C5+
CO2+N2-6.54-86.22-----
CO2+H2116.89-68.56-10.392.1917.141.72
228.74-61.52-11.214.0415.257.98
330.82-46.68-13.965.4029.614.35
431.54-45.57-19.307.3425.911.88
CO+H22-86.36-35.0222.572.9622.1017.35
3-89.57-41.4221.753.8224.698.32
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hou, Y.; Wang, X.; Chen, M.; Gao, X.; Liu, Y.; Guo, Q. Sr1-xKxFeO3 Perovskite Catalysts with Enhanced RWGS Reactivity for CO2 Hydrogenation to Light Olefins. Atmosphere 2022, 13, 760. https://doi.org/10.3390/atmos13050760

AMA Style

Hou Y, Wang X, Chen M, Gao X, Liu Y, Guo Q. Sr1-xKxFeO3 Perovskite Catalysts with Enhanced RWGS Reactivity for CO2 Hydrogenation to Light Olefins. Atmosphere. 2022; 13(5):760. https://doi.org/10.3390/atmos13050760

Chicago/Turabian Style

Hou, Yuanhao, Xinyu Wang, Ming Chen, Xiangyu Gao, Yongzhuo Liu, and Qingjie Guo. 2022. "Sr1-xKxFeO3 Perovskite Catalysts with Enhanced RWGS Reactivity for CO2 Hydrogenation to Light Olefins" Atmosphere 13, no. 5: 760. https://doi.org/10.3390/atmos13050760

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