Next Article in Journal
Asymmetric Dual Enamine Catalysis/Hydrogen Bonding Activation
Next Article in Special Issue
Metal–Organic Frameworks for Electrocatalytic CO2 Reduction into Formic Acid
Previous Article in Journal
Removal of Organic Sulfur Pollutants from Gasification Gases at Intermediate Temperature by Means of a Zinc–Nickel-Oxide Sorbent for Integration in Biofuel Production
Previous Article in Special Issue
Recent Mechanistic Understanding of Fischer-Tropsch Synthesis on Fe-Carbide
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Core-Shell Structured Na/Fe@Co Bimetallic Catalyst for Light-Hydrocarbon Synthesis from CO2 Hydrogenation

Department of Applied Chemistry, School of Engineering, University of Toyama, Gofuku 3190, Toyama 930-8555, Japan
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(7), 1090; https://doi.org/10.3390/catal13071090
Submission received: 12 June 2023 / Revised: 29 June 2023 / Accepted: 3 July 2023 / Published: 11 July 2023
(This article belongs to the Special Issue Catalysis for Selective Hydrogenation of CO and CO2)

Abstract

:
The direct CO2 Fischer–Tropsch synthesis (CO2-FTS) process has been proven as one of the indispensable and effective routes in CO2 utilization and transformation. Herein, we present a core-shell structured Na/Fe@Co bimetallic catalyst to boost CO2 conversion and light hydrocarbon (C2 to C4) selectivity, as well as inhibit the selectivity of CO. Compared to the Na/Fe catalyst, our Na/Fe@CoCo-3 catalyst enabled 50.3% CO2 conversion, 40.1% selectivity of light hydrocarbons (C2-C4) in all hydrocarbon products and a high olefin-to-paraffin ratio (O/P) of 7.5 at 330 °C and 3.0 MPa. Through the characterization analyses, the introduction of CoCo Prussian Blue Analog (CoCo PBA) not only increased the reducibility of iron oxide (Fe2O3 to Fe3O4), accelerated the formation of iron carbide (FexCy), but also adjusted the surface basic properties of catalysts. Moreover, the trace Co atoms acted as a second active center in the CO2-FTS process for heightening light hydrocarbon synthesis from CO hydrogenation. This work provides a novel core-shell structured bimetallistic catalyst system for light hydrocarbons, especially light olefin production from CO2 hydrogenation.

Graphical Abstract

1. Introduction

The massive carbon dioxide (CO2) emissions are the main cause of global warming, sea-level rising, and extreme weather events nowadays. Thus, the development of CO2 capture, usage and storage (CCUS) technologies has become an important research direction. Recently, the conversion of CO2 into high-value-added chemicals (such as methanol, ethanol, light hydrocarbons, gasoline range hydrocarbons, and aromatics) has attracted widespread attention [1,2,3]. To produce C2+ chemical products from CO2 hydrogenation, the direct CO2 Fischer–Tropsch synthesis (CO2-FTS) process and the methanol intermediate route have been proven as two indispensable and effective routes [4,5,6,7,8,9]. The methanol intermediate route involves methanol synthesis from CO2 and H2 by single or multiple metal-oxide catalysts and follows by using zeolites to catalyze methanol into dimethyl ether, light olefins, gasoline range hydrocarbons, light aromatics, etc. [5,10,11,12,13,14,15,16]. In comparison, the CO2-FTS would be a more promising strategy to achieve high CO2 conversion and low CO selectivity at more moderate reaction conditions [17,18,19].
Iron-based catalysts have been verified as efficient catalysts and are used widely in the CO2-FTS process [17,19,20,21]. During the reaction, two iron components have been reported as key roles in CO2 hydrogenation to hydrocarbons. Firstly, CO formed from CO2 and H2 via the reverse water gas shift (RWGS) reaction on the Fe3O4 sites, then, CO and H2 convert into hydrocarbon products on the FexCy sites by the FTS reaction. However, because of the Anderson–Schulz–Flory (ASF) distribution law and the thermodynamic conversion equilibrium of CO2, there is still a big challenge for the highly selective production of the target range hydrocarbon with high CO2 conversion [22,23]. Various studies have been adopted to regulate the activity and product selectivity over iron catalysts, such as introducing alkali promoters (such as Na, K), combining with acidic zeolites (including ZSM-5, MCM-22, Y, Beta), and co-feeding with other metal or metal-oxide components (Mn, Cu, Co, ZnO, MgO, Al2O3, etc.) [2,19,20,24,25,26] for production of light olefins, gasoline-range liquid fuel, and aromatics.
Light hydrocarbons (C2-C4) are very important and well-used intermediate materials and energy sources in the chemical industry and our daily life. Light olefins, including ethylene, propylene, and butene, are widely used in the production of various plastics, solvents, drugs, cosmetics, etc. Propane and butane are the main components of liquefied petroleum gas (LPG). However, light hydrocarbons are mainly produced in the process of petroleum refining [27,28]. A green and new light hydrocarbon formation route starting from sustainable raw materials such as CO2 is becoming particularly important. In our previous works, we prepared a series of metal-oxide and zeolite bifunctional catalysts for CO and CO2 hydrogenation to light hydrocarbons through both FTS and methanol intermediate routes [29,30]. Moreover, a series of hybrid catalysts with unique core-shell structures were developed for highly selective synthesis of target hydrocarbons from CO and CO2 hydrogenation [11,31,32,33,34,35,36].
Herein, we report a series of novel Na promoted Fe@Co bimetallic catalysts with different ratios of Co/Fe, named Na/Fe@CoCo-1, Na/Fe@CoCo-2, and Na/Fe@CoCo-3, for light hydrocarbon production via the CO2-FTS reaction process. Compared to the normal Na/Fe catalyst, the Na/Fe@CoCo-3 catalyst enabled a high C2-C4 selectivity of 40.1% (CO free) and a high olefin-to-paraffin ratio (O/P) of 7.5, with CO2 conversion as high as 50.3% at 330 °C, 3.0 Mpa. The characterization analyses indicated that with the introduction of CoCo Prussian Blue Analog (CoCo PBA), the reducibility of iron oxide (Fe2O3 to Fe3O4) was increased, the formation of iron carbide (FexCy) was accelerated, and the strong basic sites on the catalyst surface were weakened. Moreover, the Co nanoparticles also acted as a second active center for heightening light hydrocarbon synthesis in the FTS reaction.

2. Results and Discussion

2.1. Structural Characterization

The XRD analysis was utilized to investigate the crystalline structure of the Na/Fe@Co catalysts. The normal Na/Fe catalyst was employed as a reference catalyst in this study. As shown in Figure 1, the typical diffraction peaks of hematite (Fe2O3) at 24.3°, 33.2°, 25.4°, 40.7°, 49.5°, 53.9°, 57.5°, 62.3°, and 64.2°, correspond to (012), (104), (110), (113), (024), (116), (122), (214), and (300) planes. The prepared Na/Fe@CoCo-1, Na/Fe@CoCo-2, and Na/Fe@CoCo-3 exhibited weaker diffraction patterns of Fe2O3 than that of the Na/Fe [37,38]. We can see from Figure 1, the diffraction peak of (104) plane of Fe2O3 almost disappeared, and the other Fe2O3 peak shifted gradually into FeCo bimetallic-oxide species when the Co content increased. These results indicating that the existence of interaction between Fe and Co atoms, which significantly affected the iron oxide nanoparticle size and exposed crystal planes. Moreover, we did not observe the characteristic peaks of CoO or CoCo PBA (see Figure 3a) on these Na/Fe@Co catalysts. This suggests that these Na/Fe@Co samples had good Co dispersion and made the Fe2O3 species highly dispersed on the CoCo PBA core.
We also utilized XRD to analyze the catalysts after the CO2-FTS reaction, as shown in Figure 2. After the reduction and reaction process, all Fe2O3 crystal diffraction patterns were eliminated. Although four catalysts obtained inferior diffraction patterns of Fe3O4, it can be found that more FexCy species (group peaks between 42 and 47°) were formed on these Na/Fe@Co catalysts after CO2-FTS reaction, compared to that of Na/Fe. In addition, the FexCy species peaks intensities were followed as Na/Fe@CoCo-3 > Na/Fe@CoCo-2 > Na/Fe@CoCo-1 > Na@Fe, indicating more Co species boosted the in situ production of FexCy species during the CO2-FTS reaction process. The FexCy species are considered as an active center of the FTS reaction, and high FexCy species content would be an important effect for the CO2-FTS reaction.
Scanning electron microscopy (SEM) analysis was utilized to observe surface morphology for the pure CoCo PBA, Na/Fe, Na/Fe@CoCo-1, Na/Fe@CoCo-2, and Na/Fe@CoCo-3 catalysts. The result revealed that the pure CoCo PBA samples contained a regular cubic geometrical structure with a mean size of 120 nm, as seen in Figure 3b [39,40]. Figure 3c shows the Na/Fe sample exhibited random spherical nanoparticle geometry, and the particles are much smaller than that of CoCo PBA. The SEM images of the Na/Fe@CoCo-1, Na/Fe@CoCo-2, and Na/Fe@CoCo-3 were listed in Figure 3d–f, respectively. Uniformly attached iron oxide nanoparticles can be seen on the surface of these Na/Fe@Co samples, and no cubic geometrical structure CoCo PBA particles could be found. Interestingly, the CoCo PBA core was not based on one crystal, but a CoCo PBA cluster. As seen in Figure 3b, a number of CoCo PBA crystals were assembled as a cluster, and iron oxide nanoparticles were further attached to this cluster evenly and transformed into a core-shell such as Na/Fe@Co catalysts in size of a few microns. There was no significant difference in morphology between the Na/Fe@CoCo-1, Na/Fe@CoCo-2, and Na/Fe@CoCo-3 catalysts. Furthermore, we further conducted the SEM element mapping analysis of the Na/Fe@CoCo-3 catalyst as indicated in Figure 3g. Only Fe, O, and Na atoms signals appeared and were distributed on the catalyst surface evenly. These results indicating a core-shell structured catalyst with CoCo PBA crystal core covered with iron oxide nanoparticles shell was successfully designed and prepared.
In addition, we employed N2 physisorption to investigate the textural properties of the Na/Fe, Na/Fe@CoCo-1, Na/Fe@CoCo-2, and Na/Fe@CoCo-3 catalysts. As indicated in Figure 4 and Table 1, all four catalysts exhibit similar hysteresis loops for N2 adsorption-desorption isotherms. The normal Na/Fe sample showed higher BET surface area (SBET), total pore volume, and average pore diameter of 40 m2/g, 0.59 cm3/g, and 67 nm, respectively. Moreover, the SBET of the Na/Fe@Co catalysts decreases gradually from 30 m2/g to 28 m2/g to 16 m2/g, with increasing the CoCo PBA amount, demonstrating because of the existence of CoCo PBA, iron oxide nanoparticles became tighter during precipitation. Three Na/Fe@Co catalysts display similar total pore volume of approximately 0.22–0.24 cm3/g, while the Na/Fe@CoCo-3 manifests a higher average pore diameter of over 51 nm than that of Na/Fe@CoCo-1 (32 nm) and Na/Fe@CoCo-2 (33 nm). The different N2 physisorption behavior may have some unpredictable effects on the chemical properties and reactivity of the catalysts.
The H2-TPR results are utilized to compare the reducing properties of pure CoCo PBA and four catalysts, as shown in Figure 5. The Na/Fe sample exhibits three main peaks for its TPR profile under pure H2 atmosphere. The three peaks at around 370 °C, 460 °C, and 605 °C are assigned to the reduction of Fe2O3 to Fe3O4, Fe3O4 to FeO, and FeO to Fe, respectively [2,17]. Compared to the Na/Fe, after the introduction of CoCo BPA, the reduction peaks of Fe2O3 to Fe3O4, and Fe3O4 to FeO shifted to a lower temperature (lower than 400 °C). Importantly, the reduction peak area of Fe2O3 to Fe3O4 also increased with the increase in CoCo BPA amount, certifying more Fe2O3 species reduced under the first reduction temperature and easier for further reduction. These observations suggest that the introduction of CoCo BPA not only demonstrated the existence of interaction between Fe and Co species but also significantly improved the reducibility of the Fe2O3 species. The reduced ability of the Fe2O3 species could be a key factor affecting the CO2-FTS reaction performance. Furthermore, as seen in Figure 5, the reduction peaks of pure CoCo PBA and those marked with blue dash of Na/Fe@Co catalysts were considered to be the thermal decomposition of CoCo PBA at high temperatures, indicating it possesses good thermal stability at the reduction and reaction temperatures in this study, that of 400 °C and 330 °C, respectively [40,41].
To elucidate the CO2 adsorption properties, also known as the basic nature of the catalysts, we performed CO2-TPD analysis for the Na/Fe, Na/Fe@CoCo-1, Na/Fe@CoCo-2, and Na/Fe@CoCo-3 samples as in Figure 6. Before CO2-TPD analysis, all catalysts were reduced at 400 °C under pure H2 atmosphere for 2 h. The Na/Fe sample observed a large high-temperature broad desorption peak from 530 °C to 640 °C, which was mainly caused by the strong basic properties of the Na component impregnated on the catalyst [26,42]. On these Na/Fe@Co samples, besides a high-temperature desorption peak at approximately 570 °C, a supernumerary low-temperature desorption peak, at approximately 400 °C, appeared. Compared to the Na/Fe catalyst, the high-temperature CO2 adsorption peaks (>500 °C) of the Na/Fe@Co samples were significantly inhibited. In addition, both low-temperature and high-temperature desorption peaks were strengthened because of the introduction of CoCo PBA. This demonstrated that the Na/Fe@CoCo-3 sample generated the highest CO2 adsorption properties among these Na/Fe@Co catalysts. The unequal CO2 adsorption properties further indicate that they may have different performances for CO2 hydrogenation to light hydrocarbons in the CO2-FTS process.

2.2. CO2 Hydrogenation Performance

The CO2-FTS reaction performance of four catalysts were tested in a continuous fixed-bed reactor. The conversions of CO2 and the product distributions are summarized in Figure 7. In the comparison of the Na/Fe, these Na/Fe@Co catalysts enabled both high CO2 conversion and light hydrocarbon selectivity. In particular, the CO2 conversion of the Na/Fe@CoCo-3 increased from 39.3% to 50.3%, and the C2-C4 selectivity was enhanced from 28.2% to 40.1% in all hydrocarbon products. Moreover, the Na/Fe@CoCo-3 catalyst exhibited a lower CO selectivity (4.6%) than the other catalysts, especially that of the Na/Fe (8.5%). The CO by-product selectivity was inversely correlated with the amount of CoCo PBA added. Moreover, in these four samples, CO2 conversion and C2-C4 selectivity grew linearly, while higher hydrocarbon (C5+) selectivity decreased linearly from the Na/Fe to Na/Fe@CoCo-1 to Na/Fe@CoCo-2. With the further increasing CoCo PBA amount of the Na/Fe@CoCo-2 to Na/Fe@CoCo-3, the CO2-FTS performance of the catalyst further improved, but the growth slowed down. The reaction performance results strongly indicated that the Na/Fe@Co catalyst not only boosted CO2 conversion conspicuously but also largely increased the C2-C4 selectivity and inhibited the formation of the unwanted by-product, CO.
According to the characterization results, we demonstrated that the order of the formed FexCy amounts in the spent catalysts was followed as Na/Fe@CoCo-3 > Na/Fe@CoCo-2 > Na/Fe@CoCo-1 > Na/Fe. From the H2-TPR analysis results, the Fe2O3 reducibility proved that there should be more Fe3O4 species in the Na/Fe@Co catalysts than that in Na/Fe after the reduction and reaction process. The more active Fe3O4 species increased CO2 activation capacity in the RWGS reaction naturally. Moreover, as an active center of the FTS reaction, the FexCy species play a key role in converting CO to hydrocarbons. As a tandem reaction process, there is a delicate balance and interplay between the RWGS reaction and FTS reaction in the CO2-FTS reaction process. That is, the reaction rate of the second step reaction (FTS reaction) has a great influence on the reaction rate of the first step reaction (RWGS reaction) [17,20,21,25]. Therefore, more FexCy species would enhance the FTS reactivity and can further promote the reaction rate of the first step RGWS reaction in the CO2-FTS reaction, resulting in increased CO2 conversion from 39.3% to 44.7% to 49.6% to 50.3% for the Na/Fe, Na/Fe@CoCo-1, Na/Fe@CoCo-2, and Na/Fe@CoCo-3, respectively. That’s why the Na/Fe@CoCo-3 catalyst achieved higher CO2 conversion and lower CO selectivity at the same time than all the other Na/Fe@Co catalysts, including the Na/Fe catalyst [18,22,43,44].
In contrast, the Na/Fe@Co catalysts generated high selectivity of C2-C4 products and low selectivity of C5+ hydrocarbons in the CO2-FTS process. It has been proved that Fe-based catalysts modified by alkali metals such as Na or K are beneficial to improving the selectivity toward long-chain hydrocarbon products since alkali metals, as highly efficient electron donors, favor the adsorption and activation of CO2 [25,26,42,45,46]. In the CO2-TPD analysis results, the introduction of CoCo PBA decreased the strength of basic properties, with the same Na amount (1% wt. in all catalysts) as that of Na/Fe. The low basic properties of these Na/Fe@Co suppressed their C-C chain growth ability to produce long-chain (C5+) hydrocarbons. As seen in Figure 7, the C5+ hydrocarbon selectivity of the Na/Fe, Na/Fe@CoCo-1, Na/Fe@CoCo-2, and Na/Fe@CoCo-3 decreased from 59.7% to 46.2% to 44.8% to 43.5%. The olefin-to-paraffin ratio (O/P) of C2-C4 hydrocarbon products was also enhanced from 5.7 to 6.7 to 7.5 with the increased introduction of CoCo PBA. Nevertheless, Co species can also improve the hydrogenation ability in the CO2-FTS reaction, rising the C2-C4 selectivity [23]. Simultaneously, the existence of Co species maybe also the reason for the slightly increased methane production (13.5%) at such a high reaction temperature, compared to that of Na/Fe (7.1%). These results unveiled that the suitable basic properties and additional Co species of the Na/Fe@Co catalysts achieved their high C2-C4 products selectivity.

3. Materials and Methods

3.1. Catalyst Synthesis

All reagents used in this study were of A.R. grade and used as received without any additional purification. Iron (Ⅲ) nitrate nonahydrate (Fe(NO3)3·9H2O, 99.9%, FUJIFILM Wako Pure Chemical Corporation, Osaka, Japan), urea (assay NH2CONH2, 99.0%, FUJIFILM Wako Pure Chemical Corporation), potassium hexacyanocobaltate (Ⅲ) (K3[Co(CN)6], 98.0%, FUJIFILM Wako Pure Chemical Corporation), cobalt (Ⅱ) nitrate hexahydrate (Co(NO3)2·6H2O, 99.5%, FUJIFILM Wako Pure Chemical Corporation), sodium citrate (98%, FUJIFILM Wako Pure Chemical Corporation), and sodium nitrate (NaNO3, 99.0%, FUJIFILM Wako Pure Chemical Corporation).
Synthesis of CoCo PBA. The CoCo PAB particles were prepared by a facile solution-based ion-exchange/precipitation strategy. Typically, 3.6 mmol of Co(NO3)2·6H2O and 5.4 mmol of sodium citrate were added into 120 mL of deionized (DI) water and magnetically stirred to form transparent solution A. Another solution B was obtained by dissolving 2.4 mmol of potassium hexacyanocobaltate (III) (0.7975 g) into 120 mL of DI water. Next, solution B was added slowly into solution A with magnetic stirring, followed by stirring for 10 min and then aging for 48 h at ambient temperature. The resultant precipitate was collected by centrifugation and finally dried at 60 °C for 12 h.
Synthesis of Fe2O3 nanoparticles (Fe NPs). Typically, 45 mmol of Fe(NO3)3·9H2O and 450 mmol of urea were dissolved into 400 mL of DI water with vigorously stirring. Next, the mixture solution was heated from ambient temperature to 90 °C using a water bath and kept for 4 h. The resulting product was collected by suction filtration and dried at 60 °C. Then, the dried product was calcined at 400 °C for 4 h after being ground and denoted as Fe NPs.
Synthesis of Fe@CoCo PBA catalysts. Initially, CoCo PBA of x g (x = 1, 2, 3, each gram of CoCo PBA contains approximately 4.5 mmol of Co), 45 mmol of Fe(NO3)3·9H2O and 450 mmol of urea were fully dispersed in 400 mL of DI water with violent ultrasonication. Following the same preparation process of Fe NPs, the Fe@CoCo PBA samples were synthesized. The obtained Fe@CoCo PBA samples with different CoCo PBA content were named as Fe@CoCo-1, Fe@CoCo-2, and Fe@CoCo-3, respectively. Finally, Na (1% wt.) was introduced by a wet-impregnation method. An amount of 0.0739 g of NaNO3 was dissolved in 1 g of DI water and dipped then in 2 g of Fe-based catalyst. The resulting samples were named as Na/Fe, Na/Fe@CoCo-1, Na/Fe@CoCo-2, and Na/Fe@CoCo-3, respectively. Before the CO2 hydrogenation reaction, all the as-prepared samples were pressed into a 40–60 mesh particle size.

3.2. Catalyst Characterization

The crystal phase of both the fresh and spent catalysts was determined by conducting powder X-ray diffraction (XRD) analysis using a Rigaku Ultima Ⅳ-NS diffractometer. The X-ray source employed was Cu-Kα (λ = 0.15406 nm). The scanning range was set from 5 to 80 °C at a scanning speed of 2 °C per minute. To examine the morphologies and structures of the catalysts, a Field Emission Scanning Electron Microscope (FE-SEM) with Energy Dispersive Spectroscopy (EDS) capability was employed. The specific surface area and pore size distribution were calculated using an AUTOSORB-1 instrument at a temperature of liquid nitrogen. Prior to conducting N2 adsorption measurements, the samples were subjected to vacuum degassing at a temperature of 160 °C for a duration of 10 h.
To assess the fundamental characteristics of the catalysts, we conducted temperature-programmed desorption of carbon dioxide (CO2-TPD) experiments using a BELCAT II-T-SP instrument manufactured by Microtrac MRB, USA. The experimental procedure involved introducing 50 mg of catalyst into a quartz tube. Initially, the catalyst was purged with a flow of 5.0% H2 gas at a rate of 30 mL/min, while maintaining a temperature of 150 °C. Following that, the catalyst was subjected to reduction at 400 °C for 120 min in a pure H2 flow rate of 30 mL/min. Subsequently, the temperature was lowered to 80 °C, and the catalyst was flushed with a flow of helium gas for 120 min to eliminate any remaining gaseous H2. To perform the CO2-TPD, a constant flow of 5.0% CO2 mixed with argon (20 mL/min) was passed over the catalyst. The temperature was increased from 80 to 800 degrees Celsius at a heating rate of 10 °C per minute. The concentration of CO2 in the gas exiting the system was monitored using a thermal conductivity detector (TCD). The obtained peak area was then quantitatively calibrated using CO2 pulses.
To assess the reducibility of the catalysts, we employed the H2 temperature-programmed reduction (H2-TPR) technique. Initially, 50 mg of the catalyst was subjected to a pretreatment step, where it was heated at 400 °C for 1 h in a pure helium (He) atmosphere at a flow rate of 30 mL/min. Following this, the temperature was gradually lowered to 50 °C. Subsequently, the catalyst was subjected to a temperature ramp-up, starting from 50 °C and reaching 900 °C at a heating rate of 10 °C per minute. This ramp-up was performed under a flow of 5.0% H2 mixed with Ar at a flow rate of 30 mL/min. The consumption of H2 during the process was monitored and recorded using a TCD.

3.3. Catalytic Performance Tests

The CO2 hydrogenation reactions were performed in a fixed-bed reactor made of SUS316. The reactor was loaded with 0.25 g of catalyst, which was diluted with 1 g of quartz granules. Prior to the reaction, the catalyst underwent a reduction step. It was first subjected to a temperature of 400 °C for 6 h in a pure H2 flow with a flow rate of 60 mL/min, at atmospheric pressure. After the completion of the reduction reaction, the reactor was allowed to naturally cool down to 330 °C. Subsequently, a gas mixture containing H2, CO2, and Ar was introduced into the reactor. The composition of the gas mixture was as follows: H2 (71.96% v), CO2 (24.03% v), and Ar (4.01% v). The pressure of the system was set to 3 MPa. The gaseous products produced in the reactor were continuously monitored using two online gas chromatographs (GC). The first GC, equipped with a thermal conductivity detector (TCD), was used to detect Ar, CH4, CO, and CO2. The second GC, equipped with a flame ionization detector (FID), was employed to analyze the gaseous hydrocarbons. Furthermore, the liquid hydrocarbon products and oxygenates were collected using an ice trap containing tridecane as a solvent. These collected samples were subsequently analyzed using two offline GC equipped with FID for the liquid hydrocarbon products and oxygenates, respectively. If not indicated otherwise, the CO2 hydrogenation reactions were conducted under the conditions of H2/CO2 = 3 (with 4% Ar), 330 °C, 3 MPa and 4800 mL·h−1·gcat−1. The hydrocarbon distributions were calculated basing the total carbon moles with a C-mol%.
CO2 conversion (CO2 conv.) was calculated by Equation (1):
CO 2   conv .   ( C-mol% ) = CO 2   in CO 2   out CO 2   in × 100 %
where CO2 in and CO2 out represent the molar fraction of CO2 at the inlet and outlet, respectively.
CO selectivity (CO sel.) was calculated using Equation (2):
S CO   ( C-mol% ) = CO   out CO 2   in CO 2 out × 100 %
where COout represents the molar fraction of CO at the outlet.
The hydrocarbon selectivity (Ci sel.) was given according to Equation (3):
C i   sel .   ( C-mol% ) = M C i i - 1 n M C i + i = 1 n M C i   oxy . × 100 %
where MCi and MCi oxy. represent the carbon mole fractions of hydrocarbon i and oxygenate i at the outlet, respectively.

4. Conclusions

In conclusion, a series of Na/Fe@Co bimetallic catalysts with a core-shell structure, named Na/Fe@CoCo-1, Na/Fe@CoCo-2, and Na/Fe@CoCo-3, were successfully designed and prepared for the CO2-FTS reaction process. The CoCo PBA was introduced as a core catalyst to improve CO2 conversion and light hydrocarbon production. The characterization results demonstrated that these Na/Fe@Co catalysts possessed better reduction ability of iron oxide (Fe2O3 to Fe3O4), suitable surface basic properties, and CO2 adsorption properties. More importantly, the Na/Fe@Co catalysts significantly increased the in situ FexCy formation during the CO2-FTS reaction, largely boosting CO2 conversion and light hydrocarbon production. As a result, the Na/Fe@CoCo-3 catalyst achieved an improved selectivity of C2-C4 as high as 40.1% in all hydrocarbon products and a high olefin-to-paraffin ratio (O/P) of 7.5 with a high CO2 conversion of 50.3%. These results and observations confirmed that a well-designed core-shell Fe@Co bimetallic structure with a suitable basic property can enhance CO2 conversion capacity, light hydrocarbon formation, and inhibit CO selectivity in the CO2-FTS reaction process.

Author Contributions

The idea was conceived by N.T., G.Y. and Y.H.; Y.L. performed the experiments and drafted this paper under the supervision of Y.H.; K.F., C.W., X.S., W.G., X.G., S.Y. and G.Y. helped to collect and analyze some characterization data. The manuscript was revised and checked through the comments of all authors. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data are contained within this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhou, W.; Cheng, K.; Kang, J.; Zhou, C.; Subramanian, V.; Zhang, Q.; Wang, Y. New horizon in C1 chemistry: Breaking the selectivity limitation in transformation of syngas and hydrogenation of CO2 into hydrocarbon chemicals and fuels. Chem. Soc. Rev. 2019, 48, 3193–3228. [Google Scholar] [CrossRef] [PubMed]
  2. Zhao, H.; Guo, L.; Gao, W.; Chen, F.; Wu, X.; Wang, K.; He, Y.; Zhang, P.; Yang, G.; Tsubaki, N. Multi-Promoters Regulated Iron Catalyst with Well-Matching Reverse Water-Gas Shift and Chain Propagation for Boosting CO2 Hydrogenation. J. CO2 Util. 2021, 52, 101700. [Google Scholar] [CrossRef]
  3. Wang, Y.; Gao, X.; Wu, M.; Tsubaki, N. Thermocatalytic hydrogenation of CO2 into aromatics by tailor-made catalysts: Recent advancements and perspectives. EcoMat 2021, 3, 12080. [Google Scholar] [CrossRef]
  4. Song, C. CO2 conversion and utilization: An overview. In CO2 Conversion and Utilization; ACS Symposium Series; American Chemical Society: Washington, DC, USA, 2002; Volume 809, pp. 2–30. [Google Scholar]
  5. Li, F.; Thevenon, A.; Rosas-Hernández, A.; Wang, Z.; Li, Y.; Gabardo, C.M.; Ozden, A.; Dinh, C.T.; Li, J.; Wang, Y.; et al. Molecular tuning of CO2-to-ethylene conversion. Nature 2020, 577, 509–513. [Google Scholar] [CrossRef] [Green Version]
  6. Wei, J.; Yao, R.; Han, Y.; Ge, Q.; Sun, J. Towards the development of the emerging process of CO2 heterogenous hydrogenation into high-value unsaturated heavy hydrocarbons. Chem. Soc. Rev. 2021, 50, 10764–10805. [Google Scholar] [CrossRef]
  7. Li, Y.; Wang, M.; Liu, X.; Hu, C.; Xiao, D.; Ma, D. Catalytic Transformation of PET and CO2 into High-Value Chemicals. Angew. Chem. 2022, 134, e202117205. [Google Scholar] [CrossRef]
  8. Guo, L.; Guo, X.; He, Y.; Tsubaki, N. CO2 heterogeneous hydrogenation to carbon-based fuels: Recent key developments and perspectives. J. Mater. Chem. A 2023, 11, 11637–11669. [Google Scholar] [CrossRef]
  9. Wang, Y.; Wang, K.; Zhang, B.; Peng, X.; Gao, X.; Yang, G.; Hu, H.; Wu, M.; Tsubaki, N. Direct Conversion of CO2 to Ethanol Boosted by Intimacy-Sensitive Multifunctional Catalysts. ACS Catal. 2021, 11, 11742–11753. [Google Scholar] [CrossRef]
  10. Li, Z.; Qu, Y.; Wang, J.; Liu, H.; Li, M.; Miao, S.; Li, C. Highly Selective Conversion of Carbon Dioxide to Aromatics over Tandem Catalysts. Joule 2019, 3, 570–583. [Google Scholar] [CrossRef] [Green Version]
  11. Wang, Y.; Gao, W.; Kazumi, S.; Li, H.; Yang, G.; Tsubaki, N. Direct and Oriented Conversion of CO2 into Value-Added Aromatics. Chem. A Eur. J. 2019, 25, 5149–5153. [Google Scholar] [CrossRef]
  12. Hu, J.; Yu, L.; Deng, J.; Wang, Y.; Cheng, K.; Ma, C.; Zhang, Q.; Wen, W.; Yu, S.; Pan, Y.; et al. Sulfur vacancy-rich MoS2 as a catalyst for the hydrogenation of CO2 to methanol. Nat. Catal. 2021, 4, 242–250. [Google Scholar] [CrossRef]
  13. Zhao, H.; Yu, R.; Ma, S.; Xu, K.; Chen, Y.; Jiang, K.; Fang, Y.; Zhu, C.; Liu, X.; Tang, Y.; et al. The role of Cu1–O3 species in single-atom Cu/ZrO2 catalyst for CO2 hydrogenation. Nat. Catal. 2022, 5, 818–831. [Google Scholar] [CrossRef]
  14. Zhu, Q.; Zhou, H.; Wang, L.; Wang, L.; Wang, C.; Wang, H.; Fang, W.; He, M.; Wu, Q.; Xiao, F.-S. Enhanced CO2 utilization in dry reforming of methane achieved through nickel-mediated hydrogen spillover in zeolite crystals. Nat. Catal. 2022, 5, 1030–1037. [Google Scholar] [CrossRef]
  15. Chen, F.; Liang, J.; Wang, F.; Gao, W.; Kugue, Y.; He, Y.; Guo, X.; Yang, G.; Liu, G.; Wu, J.; et al. Alcohol Solvent Assisted Synthesis of Metallic and Metal Oxide Catalysts: As-Prepared Cu/ZnO/Al2O3 Catalysts for Low-Temperature Methanol Synthesis with an Ultrahigh Yield. ACS Catal. 2023, 13, 6169–6184. [Google Scholar] [CrossRef]
  16. Li, H.; Wang, L.; Xiao, F.-S. Silica-modulated Cu-ZnO-Al2O3 catalyst for efficient hydrogenation of CO2 to methanol. Catal. Today 2023, 418, 114051. [Google Scholar] [CrossRef]
  17. Wei, J.; Ge, Q.; Yao, R.; Wen, Z.; Fang, C.; Guo, L.; Xu, H.; Sun, J. Directly converting CO2 into a gasoline fuel. Nat. Commun. 2017, 8, 15174. [Google Scholar] [CrossRef] [Green Version]
  18. Liu, J.; Zhang, G.; Jiang, X.; Wang, J.; Song, C.; Guo, X. Insight into the role of Fe5C2 in CO2 catalytic hydrogenation to hydrocarbons. Catal. Today 2021, 371, 162–170. [Google Scholar] [CrossRef]
  19. Liang, J.; Guo, L.; Gao, W.; Wang, C.; Guo, X.; He, Y.; Yang, G.; Tsubaki, N. Direct Conversion of CO2 to Aromatics over K–Zn–Fe/ZSM-5 Catalysts via a Fischer–Tropsch Synthesis Pathway. Ind. Eng. Chem. Res. 2022, 61, 10336–10346. [Google Scholar] [CrossRef]
  20. Han, Y.; Fang, C.; Ji, X.; Wei, J.; Ge, Q.; Sun, J. Interfacing with Carbonaceous Potassium Promoters Boosts Catalytic CO2 Hydrogenation of Iron. ACS Catal. 2020, 10, 12098–12108. [Google Scholar] [CrossRef]
  21. Wang, C.; Fang, W.; Liu, Z.; Wang, L.; Liao, Z.; Yang, Y.; Li, H.; Liu, L.; Zhou, H.; Qin, X.; et al. Fischer–Tropsch synthesis to olefins boosted by MFI zeolite nanosheets. Nat. Nanotechnol. 2022, 17, 714–720. [Google Scholar] [CrossRef]
  22. Jiang, Y.; Wang, K.; Wang, Y.; Liu, Z.; Gao, X.; Zhang, J.; Ma, Q.; Fan, S.; Zhao, T.-S.; Yao, M. Recent advances in thermocatalytic hydrogenation of carbon dioxide to light olefins and liquid fuels via modified Fischer-Tropsch pathway. J. CO2 Util. 2023, 67, 102321. [Google Scholar] [CrossRef]
  23. Guo, L.; Gao, X.; Gao, W.; Wu, H.; Wang, X.; Sun, S.; Wei, Y.; Kugue, Y.; Guo, X.; Sun, J.; et al. High-yield production of liquid fuels in CO2 hydrogenation on a zeolite-free Fe-based catalyst. Chem. Sci. 2023, 14, 171–178. [Google Scholar] [CrossRef] [PubMed]
  24. Oschatz, M.; Krause, S.; Krans, N.A.; Hernández Mejía, C.; Kaskel, S.; De Jong, K.P. Influence of precursor porosity on sodium and sulfur promoted iron/carbon Fischer–Tropsch catalysts derived from metal–organic frameworks. Chem. Commun. 2017, 53, 10204–10207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Xu, Y.; Zhai, P.; Deng, Y.; Xie, J.; Liu, X.; Wang, S.; Ma, D. Highly Selective Olefin Production from CO2 Hydrogenation on Iron Catalysts: A Subtle Synergy between Manganese and Sodium Additives. Angew. Chem. 2020, 132, 21920–21928. [Google Scholar] [CrossRef]
  26. Yang, Q.; Kondratenko, V.A.; Petrov, S.A.; Doronkin, D.E.; Saraçi, E.; Lund, H.; Arinchtein, A.; Kraehnert, R.; Skrypnik, A.S.; Matvienko, A.A.; et al. Identifying Performance Descriptors in CO2 Hydrogenation over Iron-Based Catalysts Promoted with Alkali Metals. Angew. Chem. Int. Ed. 2022, 61, e202116517. [Google Scholar] [CrossRef]
  27. Mango, F.D. The light hydrocarbons in petroleum: A critical review. Org. Geochem. 1997, 26, 417–440. [Google Scholar] [CrossRef]
  28. Mango, F.D. The origin of light hydrocarbons. Geochim. Cosmochim. Acta 2000, 64, 1265–1277. [Google Scholar] [CrossRef]
  29. Tan, L.; Zhang, P.; Cui, Y.; Suzuki, Y.; Li, H.; Guo, L.; Yang, G.; Tsubaki, N. Direct CO2 hydrogenation to light olefins by suppressing CO by-product formation. Fuel Process. Technol. 2019, 196, 106174. [Google Scholar] [CrossRef]
  30. Li, H.; Zhang, P.; Guo, L.; He, Y.; Zeng, Y.; Thongkam, M.; Natakaranakul, J.; Kojima, T.; Reubroycharoen, P.; Vitidsant, T.; et al. A Well-Defined Core–Shell-Structured Capsule Catalyst for Direct Conversion of CO2 into Liquefied Petroleum Gas. ChemSusChem 2020, 13, 2060–2065. [Google Scholar] [CrossRef]
  31. Yang, G.; Xing, C.; Hirohama, W.; Jin, Y.; Zeng, C.; Suehiro, Y.; Wang, T.; Yoneyama, Y.; Tsubaki, N. Tandem catalytic synthesis of light isoparaffin from syngas via Fischer–Tropsch synthesis by newly developed core–shell-like zeolite capsule catalysts. Catal. Today 2013, 215, 29–35. [Google Scholar] [CrossRef]
  32. Bao, J.; He, J.; Zhang, Y.; Yoneyama, Y.; Tsubaki, N. A Core/Shell Catalyst Produces a Spatially Confined Effect and Shape Selectivity in a Consecutive Reaction. Angew. Chem. 2008, 120, 359–362. [Google Scholar] [CrossRef]
  33. Song, F.; Yong, X.; Wu, X.; Zhang, W.; Ma, Q.; Zhao, T.; Tan, M.; Guo, Z.; Zhao, H.; Yang, G.; et al. FeMn@HZSM-5 capsule catalyst for light olefins direct synthesis via Fischer-Tropsch synthesis: Studies on depressing the CO2 formation. Appl. Catal. B Environ. 2022, 300, 120713. [Google Scholar] [CrossRef]
  34. Zhang, L.; Gao, W.; Wang, F.; Wang, C.; Liang, J.; Guo, X.; He, Y.; Yang, G.; Tsubaki, N. Highly selective synthesis of light aromatics from CO2 by chromium-doped ZrO2 aerogels in tandem with HZSM-5@SiO2 catalyst. Appl. Catal. B Environ. 2023, 328, 122535. [Google Scholar] [CrossRef]
  35. Amoo, C.C.; Xing, C.; Tsubaki, N.; Sun, J. Tandem Reactions over Zeolite-Based Catalysts in Syngas Conversion. ACS Cent. Sci. 2022, 8, 1047–1062. [Google Scholar] [CrossRef]
  36. Gao, W.; Guo, L.; Wu, Q.; Wang, C.; Guo, X.; He, Y.; Zhang, P.; Yang, G.; Liu, G.; Wu, J.; et al. Capsule-like zeolite catalyst fabricated by solvent-free strategy for para-Xylene formation from CO2 hydrogenation. Appl. Catal. B Environ. 2022, 303, 120906. [Google Scholar] [CrossRef]
  37. Lemine, O.M. Microstructural characterisation of α-Fe2O3 nanoparticles using, XRD line profiles analysis, FE-SEM and FT-IR. Superlattices Microstruct. 2009, 45, 576–582. [Google Scholar] [CrossRef]
  38. Kashyap, S.J.; Sankannavar, R.; Madhu, G.M. Iron oxide (Fe2O3) synthesized via solution-combustion technique with varying fuel-to-oxidizer ratio: FT-IR, XRD, optical and dielectric characterization. Mater. Chem. Phys. 2022, 286, 126118. [Google Scholar] [CrossRef]
  39. Wang, Y.; Ma, J.; Wang, J.; Chen, S.; Wang, H.; Zhang, J. Interfacial Scaffolding Preparation of Hierarchical PBA-Based Derivative Electrocatalysts for Efficient Water Splitting. Adv. Energy Mater. 2019, 9, 1802939. [Google Scholar] [CrossRef]
  40. Wang, J.; Wang, B.; Liu, X.; Bai, J.; Wang, H.; Wang, G. Prussian blue analogs (PBA) derived porous bimetal (Mn, Fe) selenide with carbon nanotubes as anode materials for sodium and potassium ion batteries. Chem. Eng. J. 2020, 382, 123050. [Google Scholar] [CrossRef]
  41. Wu, X.; Ru, Y.; Bai, Y.; Zhang, G.; Shi, Y.; Pang, H. PBA composites and their derivatives in energy and environmental applications. Coord. Chem. Rev. 2022, 451, 214260. [Google Scholar] [CrossRef]
  42. Zhai, P.; Xu, C.; Gao, R.; Liu, X.; Li, M.; Li, W.; Fu, X.; Jia, C.; Xie, J.; Zhao, M.; et al. Highly Tunable Selectivity for Syngas-Derived Alkenes over Zinc and Sodium-Modulated Fe5C2 Catalyst. Angew. Chem. 2016, 128, 10056–10061. [Google Scholar] [CrossRef]
  43. Yang, C.; Zhao, H.; Hou, Y.; Ma, D. Fe5C2 Nanoparticles: A Facile Bromide-Induced Synthesis and as an Active Phase for Fischer–Tropsch Synthesis. J. Am. Chem. Soc. 2012, 134, 15814–15821. [Google Scholar] [CrossRef]
  44. Song, C.; Liu, X.; Xu, M.; Masi, D.; Wang, Y.; Deng, Y.; Zhang, M.; Qin, X.; Feng, K.; Yan, J.; et al. Photothermal Conversion of CO2 with Tunable Selectivity Using Fe-Based Catalysts: From Oxide to Carbide. ACS Catal. 2020, 10, 10364–10374. [Google Scholar] [CrossRef]
  45. Wang, L.-X.; Wang, L.; Xiao, F.-S. Tuning product selectivity in CO2 hydrogenation over metal-based catalysts. Chem. Sci. 2021, 12, 14660–14673. [Google Scholar] [CrossRef] [PubMed]
  46. Wang, C.; Zhang, J.-L.; Gao, X.-H.; Zhao, T.-S. Research progress on iron-based catalysts for CO2 hydrogenation to long-chain linear α-olefins. J. Fuel Chem. Technol. 2023, 51, 67–85. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction (XRD) patterns of the fresh Fe-based catalysts.
Figure 1. X-ray diffraction (XRD) patterns of the fresh Fe-based catalysts.
Catalysts 13 01090 g001
Figure 2. XRD patterns of the spent Fe-based catalysts.
Figure 2. XRD patterns of the spent Fe-based catalysts.
Catalysts 13 01090 g002
Figure 3. XRD pattern of CoCo PBA (a); Scanning electron microscopy (SEM) image of CoCo PBA (b), Na/Fe (c), Na/Fe@CoCo-1 (d), Na/Fe@CoCo-2 (e), and Na/Fe@CoCo-3 (f); and SEM element mapping analysis image of Na/Fe@CoCo-3 (g), the element mapping images of Fe, O, Co, Na, and C listed as (g1) to (g5), respectively.
Figure 3. XRD pattern of CoCo PBA (a); Scanning electron microscopy (SEM) image of CoCo PBA (b), Na/Fe (c), Na/Fe@CoCo-1 (d), Na/Fe@CoCo-2 (e), and Na/Fe@CoCo-3 (f); and SEM element mapping analysis image of Na/Fe@CoCo-3 (g), the element mapping images of Fe, O, Co, Na, and C listed as (g1) to (g5), respectively.
Catalysts 13 01090 g003
Figure 4. N2 adsorption-desorption isotherms for the Fe−based catalysts.
Figure 4. N2 adsorption-desorption isotherms for the Fe−based catalysts.
Catalysts 13 01090 g004
Figure 5. H2-TPR profiles for the Fe−based catalysts.
Figure 5. H2-TPR profiles for the Fe−based catalysts.
Catalysts 13 01090 g005
Figure 6. CO2-TPD profiles for the Fe−based catalysts.
Figure 6. CO2-TPD profiles for the Fe−based catalysts.
Catalysts 13 01090 g006
Figure 7. CO2 hydrogenation performance of the Fe−based catalysts. Reaction conditions: 330 °C, 3.0 MPa, feed gas (H2 (71.96 v%), CO2 (24.03 v%), and Ar (4.01 v%)), W/F = 5 g h/mol, catalyst 0.25 g, reaction time 6 h.
Figure 7. CO2 hydrogenation performance of the Fe−based catalysts. Reaction conditions: 330 °C, 3.0 MPa, feed gas (H2 (71.96 v%), CO2 (24.03 v%), and Ar (4.01 v%)), W/F = 5 g h/mol, catalyst 0.25 g, reaction time 6 h.
Catalysts 13 01090 g007
Table 1. The SBET, total pore volume, and average pore diameter of the fresh catalysts.
Table 1. The SBET, total pore volume, and average pore diameter of the fresh catalysts.
CatalystsSBET/(m2/g) aTotal Pore Volume/(cm3/g) bAverage Pore Diameter/(nm) b
Na/Fe400.5967
Na/Fe@CoCo-1300.2432
Na/Fe@CoCo-2280.2433
Na/Fe@CoCo-3160.2251
a The specific surface area was determined by Brunauer–Emmett–Teller (BET) method. b The total pore volume was determined by Barrett–Joyner–Halenda (BJH) method.
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

Li, Y.; He, Y.; Fujihara, K.; Wang, C.; Sun, X.; Gao, W.; Guo, X.; Yasuda, S.; Yang, G.; Tsubaki, N. A Core-Shell Structured Na/Fe@Co Bimetallic Catalyst for Light-Hydrocarbon Synthesis from CO2 Hydrogenation. Catalysts 2023, 13, 1090. https://doi.org/10.3390/catal13071090

AMA Style

Li Y, He Y, Fujihara K, Wang C, Sun X, Gao W, Guo X, Yasuda S, Yang G, Tsubaki N. A Core-Shell Structured Na/Fe@Co Bimetallic Catalyst for Light-Hydrocarbon Synthesis from CO2 Hydrogenation. Catalysts. 2023; 13(7):1090. https://doi.org/10.3390/catal13071090

Chicago/Turabian Style

Li, Yanbing, Yingluo He, Kensei Fujihara, Chengwei Wang, Xu Sun, Weizhe Gao, Xiaoyu Guo, Shuhei Yasuda, Guohui Yang, and Noritatsu Tsubaki. 2023. "A Core-Shell Structured Na/Fe@Co Bimetallic Catalyst for Light-Hydrocarbon Synthesis from CO2 Hydrogenation" Catalysts 13, no. 7: 1090. https://doi.org/10.3390/catal13071090

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