Next Article in Journal
One-Pot Solvent-Free Synthesis of N,N-Bis(2-Hydroxyethyl) Alkylamide from Triglycerides Using Zinc-Doped Calcium Oxide Nanospheroids as a Heterogeneous Catalyst
Next Article in Special Issue
Selective CO Hydrogenation Over Bimetallic Co-Fe Catalysts for the Production of Light Paraffin Hydrocarbons (C2–C4): Effect of Space Velocity, Reaction Pressure and Temperature
Previous Article in Journal
Recent Advances in Industrial Sulfur Tolerant Water Gas Shift Catalysts for Syngas Hydrogen Enrichment: Application of Lean (Low) Steam/Gas Ratio
Previous Article in Special Issue
Bimetallic Iron–Cobalt Catalysts and Their Applications in Energy-Related Electrochemical Reactions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Reverse Water-Gas Shift Iron Catalyst Derived from Magnetite

Center for Catalytic Science and Technology, Department of Chemical and Biomolecular Engineering, University of Delaware, Newark, DE 19716, USA
*
Author to whom correspondence should be addressed.
Catalysts 2019, 9(9), 773; https://doi.org/10.3390/catal9090773
Submission received: 19 August 2019 / Revised: 29 August 2019 / Accepted: 30 August 2019 / Published: 14 September 2019
(This article belongs to the Special Issue Iron and Cobalt Catalysts)

Abstract

:
The catalytic properties of unsupported iron oxides, specifically magnetite (Fe3O4), were investigated for the reverse water-gas shift (RWGS) reaction at temperatures between 723 K and 773 K and atmospheric pressure. This catalyst exhibited a fast catalytic CO formation rate (35.1 mmol h−1 gcat.−1), high turnover frequency (0.180 s−1), high CO selectivity (>99%), and high stability (753 K, 45000 cm3h−1gcat.−1) under a 1:1 H2 to CO2 ratio. Reaction rates over the Fe3O4 catalyst displayed a strong dependence on H2 partial pressure (reaction order of ~0.8) and a weaker dependence on CO2 partial pressure (reaction order of 0.33) under an equimolar flow of both reactants. X-ray powder diffraction patterns and XPS spectra reveal that the bulk composition and structure of the post-reaction catalyst was formed mostly of metallic Fe and Fe3C, while the surface contained Fe2+, Fe3+, metallic Fe and Fe3C. Catalyst tests on pure Fe3C (iron carbide) suggest that Fe3C is not an effective catalyst for this reaction at the conditions investigated. Gas-switching experiments (CO2 or H2) indicated that a redox mechanism is the predominant reaction pathway.

1. Introduction

Today’s anthropogenic emissions of carbon dioxide to the atmosphere amount to about 35,000 Tg per year, and the greenhouse effect of these accumulated emissions has been recognized as an alarming hazard to the well-being of modern societies. Although multiple approaches have been considered to mitigate these emissions, recently more emphasis has been placed on the potential synergy of carbon capture and the utilization of these large outflows of carbon dioxide. Among several possibilities, a recent National Academies of Science and Engineering report [1] on CO2 waste gas utilization highlights the conversion of CO2 by hydrogenation into CO and water—the reverse water-gas shift (RWGS) reaction, Equation (1)— as critical, and points to the need for improved catalysts with high stability and durability.
The RWGS reaction can be part of a two-step hydrogenation process for the conversion of CO2 to valuable products. First, CO2 is reduced to CO via the RWGS reaction, and second, CO can be converted to either hydrocarbons via the Fischer-Tropsch (FT) process or methanol via CAMERE (CArbon dioxide hydrogenation to form MEthanol via a Reverse WGS reaction) process [2]. The RWGS reaction is an endothermic reaction (ΔH°298 K = 41.2 kJ mol−1), and thus is thermodynamically favorable at higher temperatures.
C O 2 + H 2 C O + H 2 O
Noble metals such as Pt [3,4,5] and Pd [6,7], and other various metals such as Cu [8,9,10,11], Ni [12,13] and Fe [14] supported on oxides were reported to be active for the production of CO. Among them, Cu-based materials have been widely studied, and thus have also been investigated in many instances for the RWGS reaction. For example, Cu-Ni/Al2O3 [15], Cu/ZnO [16], Cu-Zn/Al2O3 [16] and Cu/SiO2 promoted with potassium [17] have all shown good RWGS activity. However, Cu materials tend to deactivate by sintering (frittage) at high temperatures (T > 773 K), which are required for high RWGS activity. For high temperature applications, iron can be added as a thermal stabilizer: Chen et al. [10] showed that adding small amounts of iron to 10% Cu/SiO2 resulted in stable RWGS activity for 120 h at 873 K and atmospheric pressure, while non-promoted 10% Cu/SiO2 deactivated rapidly.
Iron oxides (FexOy) are often used industrially for FT synthesis (473 K–623 K, 1 MPa) [18,19] and the high-temperature (623 K–723 K) WGS reaction [20,21,22]. In FT synthesis, alkalized, iron-based catalysts are combined with Cu for reduction promotion. Schulz et al. [18,23] showed that the working FT catalysts contain several iron carbide phases and elemental carbon formed after the hydrogen reduction period. Iron oxides and the metallic iron are still present in the catalyst composition, but iron carbide phases are identified as active sites [23]. The RWGS and WGS reactions are often carried out in conjunction with FT synthesis at a higher temperature regime on iron catalysts, and iron oxide or oxidic amorphous iron phases are known as the active phases for WGS and RWGS [24,25,26].
Extensive research on iron-based catalysts has been reported mainly on the WGS reaction over decades [20]. Chromium is a structural promoter that helps prevent the iron from sintering at high temperature. A more recent survey on Cr-free, Fe-based WGS catalysts shows the current strong interest in this topic [27]. However, the studies on RWGS reactions over iron-based catalysts are much less frequent. Fishman et al. [28] synthesized hematite nano-sheets to obtain a 28% CO2 conversion at 783 K, and hematite nanowires to obtain a 50% CO2 conversion at a very high temperature of 1023 K. Hematite was reduced to magnetite during the reaction. The catalytic behavior over time on stream and the stability of the CO production were, however, not investigated. Fe nanoparticles have also shown good stability and activity (35% CO2 conversion, >85% CO selectivity) in RWGS by Kim et al. [29], yet no kinetic parameters were determined, and the mechanism was not discussed.
Two principal mechanisms of the WGS (or RWGS) reaction have been investigated extensively: The “redox mechanism” and the “associative” mechanism [30,31]. Different catalysts may lead to a different reaction pathway. The redox mechanism was suggested to be active for the WGS reaction over iron catalysts promoted with chromium [32]. A distinguishing feature of the redox mechanism is that products can be generated in the absence of both reactants. The catalyst is first reduced by the adsorbed H2, and is subsequently oxidized by CO2 (in RWGS) or H2O (in WGS). The associative mechanism was proposed to be dominant in the WGS reaction over iron oxide catalysts [33]. In this mechanism, both reactants are adsorbed on the catalyst surface at the same time to create products. Several carbon-containing intermediates, including formate, carbonate, carbonyl and carboxyl species, have been proposed. In a previous report in alumina-supported iron catalysts [14], we showed that the redox mechanism is the only pathway for RWGS over Fe/γ-Al2O3, and the predominant pathway over Fe-K/γ-Al2O3. The addition of the potassium promoter activates a secondary pathway for CO formation, which is probably the associative pathway.
In the present report, unsupported Fe3O4-derived catalyst is identified as a highly active, selective and stable catalyst for the reverse water-gas shift reaction at temperatures between 723 K and 753 K. The characterization of surface composition, bulk properties, and the evaluation of the CO production specific rate showed that the working catalyst is constructed during the H2-activation and the period of reaction conditions. Quantitative gas-switching experiments in combination with isotopic switching experiments allowed the redox and associative reaction pathway to be differentiated. The catalysts appear to be highly stable under the reaction conditions investigated.

2. Results and Discussion

The catalytic CO formation rates on the Fe3O4 catalyst with various H2 to CO2 ratios (Figure 1) show that after an induction period of ~120 min, the catalyst produced CO at a steady rate of 35.1 mmol h−1 gcat. −1 at 753 K with 12.5% CO2 conversion. The selectivity to CO was greater than 99% under equimolar CO2 and H2. After 950 min, the partial pressure of CO2 was raised to 60 kPa, while the partial pressure of H2 was kept constant. The rate increased to 54.6 mmol h−1 gcat.−1 with 4.4% CO2 conversion. Deactivation also occurred during this period: Starting at a deactivation rate of 3.71 mmol h−1 gcat.−1 per h, this rate gradually decreased to 0.23 mmol h−1 gcat.−1 per h. When the partial pressure of H2 was 60 kPa, and the CO2 was switched back to 15 kPa, the CO formation rate increased first to 91.3 mmol h−1 gcat.−1 and gradually stabilized to a value of 95.3 mmol h−1 gcat.−1 with 33.7% CO2 conversion. The final rate of the catalyst reactivation was about 0.50 mmol h−1 gcat.−1 per h. It should be noted that the differential condition (see Equation (8)) was used to determine reaction rates, and preferred for the investigation of the kinetic properties of our materials. Under higher concentration of H2 (15 kPa CO2 + 60 kPa H2), high CO2 conversion (>12%) could lead to small errors in the estimation of the reaction rate and the reaction order. The trend observed in Figure 1, however, should not be affected by this approximation.
With 60 kPa of H2 and 15 kPa of CO2 in the feed, a small amount of CH4—the only side-product—was produced at the rate of 1.35 mmol h−1 gcat.−1, reducing the CO selectivity from near 100% to 98.6%. Methane production implies that C–H bond formation is facilitated at higher partial pressure of H2. There was no further C–C chain growth under this reaction condition, indicating that the FT synthesis was not active over the working catalyst. The effect of H2 partial pressure on the CO formation rate was much higher than the effect of CO2, implying a higher reaction order on H2 than CO2. The catalyst showed overall high stability in 1300 min, and the final reactivation rate in excess H2 was higher than the deactivation rate under the excess CO2 condition. This indicates that this Fe3O4-derived catalyst is easy to regenerate in a very short period (<100 min), making the catalyst attractive in industrial use for long-term application.
The activation of the catalyst was carried out by a pretreatment under reducing conditions (H2 gas). The bulk structure of the catalyst after the pretreatment can be identified. Based on the body-centered cubic (BCC) structure of α-Fe (JCPDS PDF 00-006-0696) after the pretreatment (Figure S1), the surface density of Fe atoms on the Fe(110) surface can be calculated as 1.297 × 1019 Fe atoms m−2. Assuming all Fe atoms on Fe(110) were active sites, the observed CO formation rates can be converted to turnover frequency (TOF). This is reported in Figure 1 based on the measured CO formation rates, atomic surface density, and BET surface area (2.52 m2/g for Fe3O4) of the pristine catalyst. The turnover frequency of this catalyst under the equimolar condition was as high as 0.18 s−1 (Ptot = 1 bar, T = 753 K, Ftot. = 75 sccm, GHSV = 4.5 × 104 cm3 h−1 gcat.−1).
The stable reaction rates observed after the initial break-in period at 753 K allow for the determination of kinetic parameters without having to model deactivation profiles. The kinetic parameters, including reaction orders with respect to CO2 and H2, and measured activation energies (Emeas) over Fe3O4 under near equimolar CO2 and H2 (~1:1), and in H2 excess (2:1, 4:1, and 9:1)—see Table 1— indicate that CO formation rates have a higher dependence on H2 partial pressure (order of ~0.8) than CO2 partial pressure (order of ~0.33) under equimolar composition. In general, rate orders depend on reaction conditions: Increasing the H2 partial pressure increases the order on CO2 to 0.39 and decreases the rate order on H2 to 0.72. At a ratio of H2:CO2 near 4:1, the reaction orders still show the same trend: An increasing dependence on CO2 (order of 0.43) and decreasing dependence on H2 (order of 0.31). In a high excess of H2 (H2:CO2 = ~9:1), the reaction rate over Fe3O4 was of the order 1.30 with respect to CO2, and was independent of H2 pressure. The activation energies (Emeas) also depend on the H2:CO2 partial pressure ratios; that is, different reaction pathways may occur under these conditions. This behavior is not unique to iron catalysts. Similar reaction orders were also observed by Ginés et al. [34] in the same regime of PH2/PCO2 < 3 (CO2 order ≈ 0.3, H2 order ≈ 0.8) on the CuO/ZnO/Al2O3 catalyst, and by Kim et al. [3] on Pt/Al2O3 catalysts (CO2 order = 0.32, H2 order = 0.70). It was also suggested by Ginés et al. [34] that different reaction pathways should be existed for PH2/PCO2 < 3 and PH2/PCO2 > 3 regions.
Figure 2 presents the temperature-programmed reduction (TPR) profiles of (a) the fresh Fe3O4 sample and (b) the post-reaction Fe3O4 sample. In Figure 2a, the fresh Fe3O4 was heated to 753 K in a hydrogen atmosphere, kept for 2 h at these conditions, and then heated to 1,073 K at the rate of 5 K/min. The small peak observed in the TPR trace at about 563 K is assigned to an impurity of hematite present in the initial sample of magnetite (Fe3O4), but not detected in the XRD pattern, as shown in the report by Jozwiak et al. [35]. The following broader and asymmetric peak suggests a two-step reduction process that has been previously postulated in literature [36], as the following: (1)   F e 3 O 4 H 2 F e O and (2) F e O H 2 F e 0 . These two steps can be deconvoluted into two peaks located at ~ 688 K and 773 K in the TPR traces. After the 2 h reduction period at 773 K, there was no further H2 consumption at higher temperatures. That is, the sample, after the reduction pretreatment used in our activation protocol, has been converted into metallic iron. This result is also consistent with the XRD pattern in Figure S1, which shows that α-Fe was the crystal formed after the reduction pretreatment of the Fe3O4 sample in the microreactor.
As soon as CO2 was fed into the reactor, the surface of the catalyst was partially oxidized. This is known from the results of Figure 2b that illustrates the presence of two significant peaks in the TPR profile for the post-reaction catalysts (at 630 K and 780 K, respectively). The location of these two predominant peaks shows good agreement with the results of Figure 2a and results reported elsewhere [35,37]. There is at least a two-step reduction at ~630 K and 780 K, implying the coexistence of different oxidation states of iron on the post-reaction sample due to the partial oxidation from CO2. For this post-reaction sample, the H2 concentration in the effluent stream decreased by only a small amount (less than 1%), suggesting that the consumption of the H2 feed would not affect the reduction rate during the TPR reaction.
Figure 3 displays the diffraction pattern of fresh Fe3O4 and the catalyst after the RWGS reaction (post-reaction Fe3O4). The 2θ degree peak positions in fresh Fe3O4 were 30.15°, 35.45°, 37.15°,43.15°, 53.50°, 56.95° and 62.55°, which are all consistent with magnetite (JCPDS PDF 01-071-6336). The post-reaction Fe3O4 shows a very different XRD pattern: This pattern was composed of metallic iron (α-Fe, 44.67°, shown in the inset of Figure 3), iron carbide (Fe3C), and a small peak of FeOX (35.47°).
Iron oxides can be converted directly into carbides in a reducing and carburizing atmosphere [38], and the carbon source of the Fe3C production can be either from impurities in the fresh Fe3O4 sample or due to reaction with the product CO, after the RWGS reaction as indicated by Equation (2) [39]:
3 F e + 2 C O F e 3 C + C O 2
The bulk composition of the catalyst after the reaction is also consistent with the TPR results in Figure 2a,b. During TPR, the amount of H2 consumption of magnetite relative to the amount of H2 consumption after the reaction was 12:1, therefore the iron species in the catalyst has been changed into a more reduced chemical state after the pretreatment and RWGS reaction. The reduction was mainly caused by the pretreatment, while the following CO2/H2 reaction shifted the metallic iron back to a slightly more oxidized state; a combination of iron carbide, metallic iron and some iron oxides.
Besides bulk property information obtained from XRD, XPS analyses were conducted to characterize the surface composition of the initial fresh Fe3O4 and the change of the catalyst after the RWGS reaction. Figure 4 shows the XPS spectra, peak deconvolutions and the fitting envelopes for the Fe 2p3/2 spectra of Fe3O4 and post-reaction Fe3O4.
Atomic percent contributions are calculated from the fitted peaks of Fe 2p3/2 due to its larger intensity (Area of Fe 2p3/2:Fe 2p1/2 = 2:1). The Fe 2p3/2 spectra were fitted over the range of 705–722 eV. The spectra between 716–722 eV were not shown in the figure for clarity. In this range, there were only small Fe3+ satellite peaks located at 719.2 eV and 719.4 eV for both the fresh and post-reaction Fe3O4 samples, respectively, although the area of the satellite peaks was still included in the corresponding components when calculating the relative atomic percentage. The fits, including the binding energy, full width at half maximum (FWHM), and the relative iron composition, are summarized in Table 2. The fitted XPS spectrum of fresh Fe3O4 was composed of doublets for Fe2+ at 709.8 eV and Fe3+ at 711.3 eV. Fe3O4 has an inverse spinel structure which can be written as Fe3+TET[Fe2+Fe3+]OCTO4, with one Fe3+ on a tetrahedral site, and Fe2+ and the other Fe3+ distributed on octahedral sites. Therefore, the theoretical relative composition of Fe2+/Fe3+ is 0.5, which is close to the area fitted and the calculated relative composition in our fresh Fe3O4 (Fe2+:Fe3+ = 34.8/65.2 = 0.53). The Fe3+ peak has a larger FWHM than Fe2+. This is as expected because the electronic configuration of Fe2+ is 3d6, while that of Fe3+ is 3d5, that is, Fe2+ will have a longer life time compared to Fe3+; and therefore the FWHM of the Fe2+ peak should be smaller than the Fe3+ peak [40]. Additionally, the Fe3+ peaks can be attributed to two different structures, octahedral Fe3+ and tetrahedral Fe3+, a factor that will also lead to broader Fe3+ peaks.
After the RWGS reaction, the spectrum was fitted using four different components corresponding to metallic Fe (706.7 eV), Fe3C (707.9 eV), Fe2+ (709.8 eV) and Fe3+ (711.2 eV). The peak locations of Fe2+ and Fe3+ were the same or very close to the fresh sample, indicating that there was only a small surface charging effect with the flood gun on. The binding energies of the components are in agreement with literature results [38,40,41,42]. In terms of the atomic percentages, the overall peak area was re-allocated to a more reduced regime after the RWGS reaction. The Fe2+ decreased from 34.8% to 27.5%, the Fe3+ decreased from 65.2% to 43.9%, while there were two components formed: Fe (12.2%) and Fe3C (16.4%). The shift of the spectra was due to the H2 reduction pretreatment before operating the RWGS reaction, and the flow of H2/CO2 reactants through the system would balance each other to make the catalyst partially oxidized or reduced. Though the sample surface could be oxidized by the air during the transportation from the reactor to the XPS analysis chamber, the result from XPS still can confirm the reduction of the surface during the reaction, since new crystal structures such as Fe and Fe3C are detected by XRD.
The XPS analyses indicate that the active catalyst consisted of a mixture of metallic Fe, Fe3C, Fe2+ and Fe3+; however, it cannot establish the relative contributions of these components to the observed rate of the RWGS reaction. To determine if the iron carbide formed in our reaction can catalyze the RWGS reaction, reaction rates over pure Fe3C were measured (Figure 5). In Figure 5a, Fe3C showed an initial CO formation rate of 26.0 mmol h−1 g−1, but dropped 48% to near 13.5 mmol h−1 g−1 in 10 min, and then further down to 5.7 mmol h−1 g−1 in 160 min at 753 K. This is clearly different from the properties of the Fe3O4-derived catalyst in Figure 1, since the Fe3O4-derived catalyst displayed high stability for at least 1300 min. To evaluate the fast deactivation shown in Figure 5a and remove the initial reduction effect during the ramping by hydrogen, another measurement was carried out with respect to temperature, (see Figure 5b). This experiment was conducted flowing a CO2/H2/He gas mixture in the same relative concentration used in the standard activity test during the ramping procedure from room temperature to 773 K. At 573 K, the material did not catalyze the formation of CO; however, when the temperature was increased to 623 K, the catalyst immediately showed catalytic rates in the range of 2.60 mmol h−1 g−1 to 2.99 mmol h−1 g−1. After 1 h of the reaction, the rate remained at quasi-steady state at this low temperature (623 K).
The CO formation rates were 6.60 mmol h−1 g−1, 9.80 mmol h−1 g−1, and 9.16 mmol h−1 g−1, as the temperature was increased to 673 K, 723 K and 773 K, respectively. The formation rate at 773 K was not higher than that at the lower temperature because of rapid deactivation at this temperature. Faster deactivation rates were observed at higher temperatures: The average deactivation rates were 1.36 mmol h−1 g−1 per h, 2.57 mmol h−1 g−1 per h and 3.50 mmol h−1 g−1 per h.
After reacting at 773 K, the temperature was reduced to 723 K to monitor the reaction rate and compare to the previous value. Much lower rates (2.55 mmol h−1 g−1) were observed than in the previous measurement at the same temperature (723 K), indicating that there was an irreversible change in the catalyst structure or composition. At higher temperatures, the reverse reaction of Equation (2), where Fe3C reacts with CO2 and forms Fe and CO, is more favorable than the Fe3C formation [39]. Therefore, the initial CO formation rate was probably due to the formation of CO from decomposition, but quickly dropped since it is harder to convert the metallic iron back to Fe3C at higher temperatures. The iron carbide catalyst only showed steady CO production at 623 K. This explains why the deactivation at 723 K over Fe3C is very different from the steady-state magnetite catalyst reported in Figure 1, which showed a very stable CO formation rate at 753 K. This observation suggests that the operating temperature of RWGS in this study was not an environment conducive for a stable iron carbide for CO production. In addition, the iron carbide (Fe5C2 or Fe3C) is normally considered to be the active phase of iron for hydrocarbon production [43,44], and iron oxide is the active phase for WGS and RWGS [25]. Several reports have suggested that the stability of the iron catalyst in either FT synthesis [45] or RWGS [29] can be related to an iron carbide layer. Davis [45] suggested that catalyst composition and reaction condition will define the existence of the pseudo-equilibrium layer of iron carbide to ensure a very slow deactivation condition. Kim et al. [29] concluded that the stability of the catalyst could have originated from migration of C and O into the catalyst bulk, forming iron oxide and iron carbide, which likely prevented the nanoparticles on the surface from agglomerating. Based on the XPS results and Fe3C catalytic tests, the iron carbide of the working catalyst is less likely to be the main active site for CO production, but is an important species to provide stability in the overall catalytic performance.
Gas-switching experiments, in which H2 and CO2 are flown on and off, were used to distinguish and quantify contributions from redox and associative reaction pathways [3,14,46]. In the simplest form of the redox mechanism, gas-phase CO2 adsorbs on a reduced site to form CO and an oxidized site (Equation (3)), which can then be reduced by gas phase H2 to reform the reduced site (Equation (4)). The simplest redox cycle can be described as follows:
C O 2 ( g ) + s r e d . C O ( g ) + O · s
H 2 + O · s H 2 O ( g ) + s r e d .
A simplified associative pathway can be described generally by Equation (5). CO2 and H2 adsorb on the catalyst surface to form a carbon-containing intermediate (i.e. formate, carbonate, or bicarbonate), which then decomposes in the presence of H2 to form CO and H2O.
C O 2 ( g ) + H 2 ( g ) C O O H · s + H · s   C O · s + H 2 O · s
CO and H2O were the main products formed during gas-switching experiment (Figure 6). In the first three cycles, CO was formed when switching from H2 to CO2, and a very small amount of CO was formed when switching from CO2 to H2 at 30 min, 70 min and 110 min, respectively. When the catalyst was purged 20 min with helium before switching from CO2 to H2, CO was not formed, and H2O was produced at 195 min. Water was formed when switching from H2 to CO2 and when switching from CO2 to H2. After flowing H2 and purging the reactor with He for 20 min, only a negligible amount of H2O was formed upon the admission of CO2 (at 150 min).
In Figure 6, the fact that CO was formed when the reduced form of the Fe3O4 catalyst was contacted with CO2, (even after the purge with He to decrease the concentration of any surface H2), is evidence of a redox pathway. During the first 125 min of gas-switching experiments, H2O was produced during flows of only CO2 or only H2. This differs from what is expected in the traditional redox cycle, in which H2O is only produced during the H2 feeding period (see Equation (4)). However, after flowing H2 and purging the reactor with He for 20 min, the admission of CO2 only produced negligible amounts of water. Table 3 summarizes the estimated initial rates of CO production on the Fe3O4 catalyst during each segment of the gas-switching experiments. The CO production rates were calculated from the initial slopes of the concentration vs. time data in Figure 6, essentially modeling the system as a batch reactor (Equation (6)).
d C A d t = r A
It is observed (Table 3) that the rate after switch from H2 to CO2 fluctuated between 2.74 µmol L−1 s−1 gcat.−1 and 2.98 µmol L−1 s−1 gcat.−1 in the first three periods of CO2 admission, and it decreased after the He purge. The rate after switching from CO2 to H2 decreased in the first three periods, and it was zero (with no CO produced) during the last admission of CO2 after the He purge. The decrease of the CO initial rate after switching from CO2 to H2, especially when equal to zero after the purge, raises doubts about the existence of residual CO2 during the first three admissions of H2 in the switching experiment. As a control, when the gas was switched from CO2 to H2, the CO2 gas did not exit from the surface very quickly (see Figure S2). Therefore, a small amount of CO can be produced by the residual CO2 with the available reduced sites; evidence of this interpretation in the detection of very small peaks after the H2 admissions (Figure 6). The negligible CO production (relative time = 34 min, 74 min, and 114 min, respectively) after the H2 admissions should not be considered evidence of the associative mechanism. In summary, the CO formation upon switching from H2 to CO2 is evidence consistent with the redox mechanism, while the small contribution of CO production upon switching from CO2 to H2 was suppressed by the confirmation of the helium purge. Thus, from the gas-switching experiment, only the redox pathway is active on our Fe3O4-derived catalyst.
Isotopic experiments were conducted to gain insight into the mechanism of the reaction. The isotopic C18O2 to the CO2 switching experiment is shown in Figure 7. Here it can be seen that C18O (m/z = 30) formed and CO (m/z = 28) decreased when the gas (CO2/H2) was switched to C18O2/H2. CO can be only formed from CO2, and not from the lattice oxygen.
The gas-switching experiments with CO2 and H2 led us to conclude that a redox pathway is active on the Fe3O4-derived catalyst. A model that can present a redox reaction pathway for this catalyst is given in Scheme 1. It includes the adsorption of both of the reactants, CO2 and H2. In the surface redox mechanism, the dissociation of CO2 at the catalyst surface (step 2 in Scheme 1) is known to be the RDS [8,30,34]. Evidence for H2 dissociation (step 5 in Scheme 1) was observed when H2/D2 mixtures were fed to the catalyst in the presence of CO2 (see Figure S3).
HD formation was observed to occur quickly, since the amount of CO2 to CO conversion decreased on the same time scale when switching the concentration from H2/D2 (7.5 kPa/7.5 kPa) to H2 (7.5 kPa), indicating that H2 dissociation is reversible and not rate limiting.
An additional gas switching experiment was conducted to monitor the H2O production (Figure S4). The amount of H2O produced during H2 flow periods was consistent between each cycle (Figure S4a), that is, the adsorbed O·s species formed upon CO2 reduction are stable at these reaction conditions. However, the amount of H2O produced during the period of CO2 flow decreased as the purge time in helium increased. After only a five min purge, the amount of H2O produced during the period of CO2 flow was much greater than that produced following a 20 min purge in helium, and the rate fitted from the initial slope of this region dropped dramatically (see Table S1). This suggests that H* atoms from the catalyst surface appeared to desorb (as H2) during the He purge. This was a slow process, because even following a 20 min purge, there were enough H* atoms on the sample to form small amounts of H2O when CO2 was administered. Both gas-switching experiments (Figure 6 and Figure S4) suggest that the redox mechanism should be the dominant reaction pathway for this Fe3O4-derived catalyst during the CO2 hydrogenation at our reaction conditions (753 K, 1 atm).

3. Materials and Methods

3.1. Materials

Magnetite (Fe3O4, 99.99%, 50–100 nm particle size, Sigma-Aldrich, St. Louis, MO, USA) powder was pressed by hydraulic press and sieved (mesh 40 to mesh 60) to obtain particle sizes within the range of 250–425 μm before the reaction. Iron carbide (Fe3C, 99.5%, American Element, Los Angeles, CA, USA) granules were crushed to a powder using a ball-mill, pressed using a hydraulic press, and sieved within the range of 250–425 μm. The gases used were CO2 (Grade 5.0, Keen, Wilmington, DE, USA), H2 (UHP, Matheson, Basking Ridge, NJ, USA), Helium (Grade 5.0, Keen, Wilmington, DE, USA), C18O2 (95 atom% 18O, Sigma-Aldrich, St. Louis, MO, USA), and D2 ( 99.6% gas purity, 99.8% isotope purity, Cambridge Isotopes, Tewksbury, MA, USA).

3.2. Reactor Setup for Kinetics and Gas-Switching Experiments

The reaction rates and other kinetic parameters were measured using a packed-bed microreactor operated in a down-flow mode at atmospheric pressure. The catalyst particles were supported on a plug of quartz wool within a 7 mm I.D. quartz tube reactor. The quartz tube was positioned inside an Omega CRFC-26/120-A ceramic radiant full cylinder heater. The temperature was controlled by an Omega CN/74000 temperature controller using the input from a K-type, 1/16 in. diameter thermocouple (Omega Engineering, Inc., Norwalk, CT, USA) placed around the outside of the quartz tube at the center of the catalyst bed. Gas flows were controlled by mass flow controllers (Brooks Instruments) through the reactor or the other instruments.
Gas transfer lines were heated to a temperature above 373 K at all times to avoid water condensation. The composition of the effluent stream was analyzed online by a gas chromatograph (GC, 7890A, Agilent, Santa Clara, CA, USA) during continuous flow experiments or by a mass spectrometer (MS, GSD320, Pfeiffer Vacuum Technology AG, Aßlar, Germany) during gas-switching experiments or isotopic experiments. The GC was equipped with both a thermal conductivity detector (TCD) and a flame-ionization detector (FID). The TCD was used to quantify CO2, CO and H2 concentrations, while the FID was used to quantify hydrocarbon concentrations. An Agilent 2 mm ID × 12 ft Hayesep Q column was used in the GC to separate products quantified with the TCD, and an Agilent 0.32 mm ID × 30 m HP-Plot Q column was used to separate products quantified with the FID.

3.3. Catalyst Characterization

Temperature-programmed reduction (TPR) was performed by using a MS including the reduction period (773 K, Ptot = 1 bar, PH2 = 15 kPa in He for 2 h), and then the temperature was ramped up to 1073 K. The ramping rate of each step was 5 K min−1. Another TPR experiment was performed after the reverse water-gas shift (RWGS) reaction (553K, Ptot = 1 bar, GHSV = 4.5 × 104 cm3 h−1 gcat.−1, PCO2 = PH2 = 15 kPa with He as an inert balance gas). The post-reaction sample was cooled down to room temperature in He flow inside the microreactor. The TPR was then continued in 15 kPa H2 and 86.3 kPa He from room temperature to 1073 K at the rate of 5 K min−1. The H2 consumption was converted from the ion current (m/z = 2) after the calibration of the H2 signal each time before the experiment using a mass spectrometer.
X-Ray Diffraction (XRD) patterns of catalyst powders were collected at room temperature on a Bruker diffractometer using Cu Kα radiation (λ = 1.5418 Å). Measurements were taken over the range of 5° < 2θ < 70°, with a step size of 0.02° before and after the RWGS reaction. X-ray photoelectron spectroscopy (XPS) measurements were performed on a K-alpha Thermo Fisher Scientific spectrometer using a monochromated Al Kα X-ray source. The measurements of iron oxide samples (pre- and post-reaction) were done with a spot size of 50 μm at ambient temperature and a chamber pressure of ~10−7 mbar. A flood gun was used for charge compensation. All the spectra measured were calibrated by setting the reference binding energy of carbon 1s at 284.8 eV. The spectra were analyzed by Thermo Fisher Scientific (Waltham, MA, USA) Avantage® commercial software (v5.986). For the fitting, each component consists of a linear combination of Gaussian and Lorentzian product functions, and the full width at half maximum (FWHM) and differences in binding energy of the same species between the Fe2p3/2 and Fe2p1/2 scan were kept constant. SMART background in Avantage® was used over the region to define the peaks.

3.4. Measurement of Product Formation Rates and Reaction Rates with Gas-Switching or Isotopic Experiments

Most of the procedures in the measurement of the reaction rates, gas-switching, or isotopic experiments are identical to the ones described in our previous report [14]. Fe3O4 samples were pretreated before all experiments by increasing the reactor temperature at a rate of 5 K min−1 to 773 K under a gas flow of 15 kPa H2. After the pretreatment at 773 K for 2 h, the temperature was lowered to the initial reaction temperature of 753 K. During the measurement of the reaction rates, the partial pressures of the reactants were PCO2 = PH2 = 15 kPa. A constant total flow rate of 75 cm3 min−1 (sccm) was maintained with He as an inert balance gas.
Gas hourly space velocity (GHSV) in cm3 h−1 gcat.−1 was calculated under STP condition (273 K and 1 atm) according to the equation below:
G H S V = F t o t m c a t .
where m c a t . is the mass of the catalyst and Ftot is the total flow rate.
CO2 and H2 reaction orders were measured by independently varying the inlet CO2 and H2 partial pressures. The total pressure remained at 1 bar. The activation energy was estimated by using the Arrhenius plot with the temperature varied between 723 K and 773 K while monitoring the CO formation rate.
Rates of CO formation were calculated assuming differential reactor operation according to Equation (8):
r C O = V ˙ Δ C C O m c a t .
where V ˙ is the total volumetric flow rate, Δ C C O is the change in CO concentration. Measured reaction rates are the net rate of the forward and reverse reactions; therefore, the observed rate must be transformed into the reaction rate for the forward reaction by using Equations (9)–(11). The equilibrium constant (KC) is low (<1) for the RWGS at the temperatures investigated, although the reverse reaction had a negligible contribution to the observed rates because of the low conversion (<12%) under conditions at which the reactor was operated. Note that Co (Equation (11)) represents the standard state (1 mol L−1) and equals 1, since the reaction is equimolar.
r o b s . = r + r = r + ( 1 η )
η = [ C O ] [ H 2 O ] K C [ C O 2 ] [ H 2 ]
K C = ( i C i e q . γ ) 1 C o
where r o b s . is the observed rate, r + and r are the rate of forward and reverse reactions, η is the ratio of the rate of the reverse and forward reactions and K C is the equilibrium constant.
Experiments were also conducted to (i) determine reaction rates in excess (i.e. non-equimolar) CO2 or H2, and (ii) to determine apparent kinetic parameters. In the first case, CO2 and H2 were fed with the catalyst—Fe3O4 (100 mg)—held at a temperature of 753 K. The initial partial pressure of both CO2 and H2 was 15 kPa. After a period of 16 h, the partial pressure of CO2 was increased to 60 kPa, while the partial pressure of H2 was held at 15 kPa. After another period of 2 h, the partial pressure of CO2 was decreased to 15 kPa and the partial pressure of H2 was increased to 60 kPa. Finally, both partial pressures were returned to 15 kPa. CO2 conversion was quantified under the same conditions.
For the second case, apparent kinetic parameters (activation energy and reaction orders) were determined with near equimolar concentrations of CO2 and H2 on Fe3O4, and under large H2 excess on Fe3O4 as well. With near equimolar concentrations of CO2 and H2, the reaction was first performed for 15–16 h at a temperature of 753 K with reactant partial pressures of 15 kPa. The temperature was then lowered in 10 K increments to 723 K, with 5–6 GC injections (a period of about 60 min) taken at each temperature. After the period at 723 K, the CO2 partial pressure was reduced to 10 kPa and increased in 2.5 kPa increments to a final partial pressure of 20 kPa. Finally, the CO2 partial pressure was returned to 15 kPa and the H2 partial pressure was lowered to 10 kPa and increased in 2.5 kPa increments. The basic outline of experiments conducted with excess H2 was the same as that used for near equimolar reactant concentrations (see also our previous report) [14]. Reactant partial pressures during the initial period were 90 kPa H2 and 10 kPa CO2. During the variable CO2 partial pressure period, the H2 partial pressure was 85 kPa, and the CO2 partial pressure was varied between 5 and 12.5 kPa in 2.5 kPa increments. To investigate the effect of H2 partial pressure, the CO2 partial pressure was kept at 10 kPa and the H2 partial pressure was varied between 70–90 kPa in 5 kPa increments.
The measurement of the CO formation rate over Fe3C has been performed in two different manners. First, the temperature was ramped under a flow of H2 (15 kPa) and He at the rate of 5 K min−1 up to 753 K. 15 kPa CO2 was then added into the flow while monitoring the CO formation rate. The second part of the Fe3C activity test was done by ramping the temperature to 573 K, 623 K, 673 K, 723 K and then cooling back to 673 K again at the rate of 5 K min−1 under the flow of H2 (15 kPa), CO2 (15 kPa), and He (remainder). Each temperature was held for 80 min respectively for GC injection.
Gas-switching experiments were done by measuring CO formation rates while alternating between CO2 and H2 gas flows. Catalysts were pretreated as described above, and after pretreatment, CO2 was added into the reactor to allow the RWGS reaction to proceed for 2 h.
After the reaction, the gas flow rates were changed to 36 sccm helium and 4 sccm H2. After 20 min, H2 flow was stopped and was replaced by 4 sccm of CO2. After 20 min, CO2 in the gas stream was replaced by H2. This CO2→H2 sequence was repeated three times. The reactor was then purged with helium for 20 min before CO2 was readmitted into the gas stream. After 20 min, the reactor was again purged with helium before H2 was readmitted to the gas stream. All sequences with a given gas composition lasted for 20 min, and the temperature of the reactor was 773 K throughout the duration of the gas switching portion of the experiment.
Additional gas-switching experiments involving purge times of varying length with an inert gas were carried out at 753 K. Following the same pretreatment and the reaction in 15 kPa H2 and 15 kPa for 2 h at 773 K, 15 kPa H2 was admitted to the reactor. After 15 min, H2 was replaced by 15 kPa CO2 for 15 min, and CO2 was then replaced by H2 for another 15 min. Then the reactor was purged with helium for 5 min. This sequence (CO2 → H2 → He) was repeated several times, but each time the length of the inert purge was increased by 5 min.
An isotopic experiment for the CO formation rate was monitored by MS on Fe3O4 (100 mg) while alternating between CO2 (4 sccm) and C18O2 (4 sccm) after the RWGS reaction for 2 h. The temperature was kept constant at 753 K. The total flow rate was 40 sccm with H2 maintained at 4 sccm. The kinetic isotope effect (KIE) of H2/D2 was investigated on Fe3O4 (100 mg) for various H2:CO2 ratios. After pretreatment, the reaction began at a temperature of 753 K with CO2 and H2 partial pressures of 15 kPa. After 16 h, the temperature was lowered to 723 K, and after 1.5 h, H2 in the feed was replaced by D2.

4. Conclusions

An unsupported Fe3O4-derived catalyst showed very promising activity toward CO formation via CO2 hydrogenation. The high selectivity (~100% under H2:CO2 = 1:1) and great stability make the catalyst feasible to consider in extensive use. The catalyst exhibited only slight deactivation under conditions of excess CO2, but it can be quickly regenerated under excess H2. Reaction rates depended more strongly on H2 (0.8 in reaction order) compared to CO2 (0.33 in reaction order) under a near equimolar gas-phase composition. The post-reaction analyses of the catalyst indicated that the catalyst was reduced to metallic iron first in the pretreatment of H2, but the working catalyst remained partially oxidized with the composition of Fe2+, Fe3+, Fe0 and Fe3C. The main active sites are believed to be the combination of the above species, except that Fe3C is unlikely to directly contribute to the very steady CO formation at our reaction conditions (1 atm, 723–773 K). Gas-switching experiments revealed that CO was formed only when switching from H2 to CO2, and H2O was formed when switching from CO2 to H2, but not when switching from H2 to CO2 if purging of helium was in between with the gas admission. The redox mechanism is identified as the dominant reaction pathway for the unsupported iron catalyst.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/9/9/773/s1, Figure S1. XRD pattern of Fe3O4 after the reduction in H2, Figure S2. Ion current at m/z = 18 (H2O), 28 (CO), and 44 (CO2) during H2/CO2 switching experiments on Fe3O4. Arrows with a label indicate a change in gas composition to the indicated gas. Reaction conditions: T = 773 K, FHe = 36 sccm, FH2 or FCO2 = 4 sccm. The figure is a modification of Figure 6, Figure S3. Ion current at m/z = 2 (H2), 3 (HD), 4 (D2), and 28 (CO) during flow of 7.5 kPa H2 + 7.5 kPa D2 +15 kPa CO2 and 7.5 kPa H2 +15 kPa CO2 on Fe3O4. Reaction conditions: T = 753 K, Ftot = 75 sccm, Figure S4. (a) Ion current at m/z = 28 (CO) and m/z = 18 (H2O) during H2/CO2 switching experiments on Fe3O4. Arrows with a label indicate a change in gas composition to the indicated gas. The catalysts were in flowing H2 for 2 h followed by the reaction in CO2+H2 for 2 h before the first admission of H2 (relative time: 31 min) and CO2 (relative time: 46 min). Reaction conditions: T = 753 K, Ftot = 75 sccm, PH2 or PCO2 = 15 kPa. (b) is the modification of (a), Figure S5. XPS O1s spectra of Fe3O4 and post-reaction Fe3O4. The curves under the fitted envelope and above the background are contributions of estimated components from peak fitting. The peak deconvolution and their atomic % are listed on the right of the spectra for each sample, Table S1 Fitted initial slopes and area of H2O in H2/CO2 switching experiment with different He purging time in Figure S4.

Author Contributions

Conceptualization, C.-Y.C. and R.F.L.; methodology, C.-Y.C., J.A.L., and R.F.L.; investigation, C.-Y.C. and R.F.L.; resources, R.F.L.; data curation, C.-Y.C.; writing—original draft preparation, C.-Y.C.; writing—review and editing, C.-Y.C., J.A.L., and R.F.L.; supervision, R.F.L.; project administration, R.F.L.; funding acquisition, R.F.L.”

Funding

This research was funded by U.S. Army, grant number GTS-S-17-013.

Acknowledgments

The authors would also like to acknowledge Terry Dubois for the help and suggestions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. National Academies of Sciences, Engineering, and Medicine. Gaseous Carbon Waste Streams Utilization: Status and Research Needs; The National Academies Press: Washington, DC, USA, 2019. [Google Scholar]
  2. Joo, O.S.; Jung, K.D.; Moon, I.; Rozovskii, A.Y.; Lin, G.I.; Han, S.H.; Uhm, S.J. Carbon dioxide hydrogenation to form methanol via a reverse-water-gas-shift reaction (the CAMERE process). Ind. Eng. Chem. Res. 1999, 38, 1808–1812. [Google Scholar] [CrossRef]
  3. Kim, S.S.; Lee, H.H.; Hong, S.C. A study on the effect of support’s reducibility on the reverse water-gas shift reaction over Pt catalysts. Appl. Catal. A Gen. 2012, 423–424, 100–107. [Google Scholar] [CrossRef]
  4. Kim, S.S.; Lee, H.H.; Hong, S.C. The effect of the morphological characteristics of TiO2 supports on the reverse water-gas shift reaction over Pt/TiO2 catalysts. Appl. Catal. B Environ. 2012, 119–120, 100–108. [Google Scholar] [CrossRef]
  5. Goguet, A.; Shekhtman, S.O.; Burch, R.; Hardacre, C.; Meunier, F.C.; Yablonsky, G.S. Pulse-response TAP studies of the reverse water-gas shift reaction over a Pt/CeO2 catalyst. J. Catal. 2006, 237, 102–110. [Google Scholar] [CrossRef]
  6. Pettigrew, D.J.; Cant, N.W. The effects of rare earth oxides on the reverse water-gas shift reaction on palladium/alumina. Catal. Lett. 1994, 28, 313–319. [Google Scholar] [CrossRef]
  7. Park, J.N.; McFarland, E.W. A highly dispersed Pd-Mg/SiO2 catalyst active for methanation of CO2. J. Catal. 2009, 266, 92–97. [Google Scholar] [CrossRef]
  8. Wang, G.C.; Nakamura, J. Structure sensitivity for forward and reverse water-gas shift reactions on copper surfaces: A DFT study. J. Phys. Chem. Lett. 2010, 1, 3053–3057. [Google Scholar] [CrossRef]
  9. Wang, G.-C.; Jiang, L.; Pang, X.-Y.; Cai, Z.-S.; Pan, Y.-M.; Zhao, X.-Z.; Morikawa, Y.; Nakamura, J. A theoretical study of surface-structural sensitivity of the reverse water-gas shift reaction over Cu(hkl) surfaces. Surf. Sci. 2003, 543, 118–130. [Google Scholar] [CrossRef]
  10. Chen, C.S.; Cheng, W.H.; Lin, S.S. Enhanced activity and stability of a Cu/SiO2 catalyst for the reverse water gas shift reaction by an iron promoter. Chem. Commun. 2001, 1, 1770–1771. [Google Scholar] [CrossRef]
  11. Fujitani, T.; Nakamura, J. The effect of ZnO in methanol synthesis catalysts on Cu dispersion and the specific activity. Catal. Lett. 1998, 56, 119–124. [Google Scholar] [CrossRef]
  12. Sun, F.M.; Yan, C.F.; Wang, Z.D.; Guo, C.Q.; Huang, S.L. Ni/Ce-Zr-O catalyst for high CO2 conversion during reverse water gas shift reaction (RWGS). Int. J. Hydrogen Energy 2015, 40, 15985–15993. [Google Scholar] [CrossRef]
  13. Wolf, A.; Jess, A.; Kern, C. Syngas Production via Reverse Water-Gas Shift Reaction over a Ni-Al2O3 Catalyst: Catalyst Stability, Reaction Kinetics, and Modeling. Chem. Eng. Technol. 2016, 39, 1040–1048. [Google Scholar] [CrossRef]
  14. Loiland, J.A.; Wulfers, M.J.; Marinkovic, N.S.; Lobo, R.F. Fe/γ-Al2O3 and Fe–K/γ-Al2O3 as reverse water-gas shift catalysts. Catal. Sci. Technol. 2016, 6, 5267–5279. [Google Scholar] [CrossRef]
  15. Liu, Y.; Liu, D. Study of bimetallic Cu–Ni/γ-Al2O3 catalysts for carbon dioxide hydrogenation. Int. J. Hydrogen Energy 1999, 24, 351–354. [Google Scholar] [CrossRef]
  16. Stone, F.S.; Waller, D. Cu–ZnO and Cu–ZnO/Al2O3 catalysts for the reverse water-gas shift reaction. The effect of the Cu/Zn ratio on precursor characteristics and on the activity of the derived catalysts. Top. Catal. 2003, 22, 305–318. [Google Scholar] [CrossRef]
  17. Chen, C.S.; Cheng, W.H.; Lin, S.S. Study of reverse water gas shift reaction by TPD, TPR and CO2 hydrogenation over potassium-promoted Cu/SiO2 catalyst. Appl. Catal. A Gen. 2002, 238, 55–67. [Google Scholar] [CrossRef]
  18. Schulz, H. Short history and present trends of Fischer-Tropsch synthesis. Appl. Catal. A Gen. 1999, 186, 3–12. [Google Scholar] [CrossRef]
  19. Van Der Laan, G.P.; Beenackers, A.A.C.M. Kinetics and Selectivity of the Fischer-Tropsch Synthesis: A Literature Review. Catal. Rev. Sci. Eng. 1999, 41, 255–318. [Google Scholar] [CrossRef]
  20. Zhu, M.; Wachs, I.E. Iron-Based Catalysts for the High-Temperature Water–Gas Shift (HT-WGS) Reaction: A Review. ACS Catal. 2016, 6, 722–732. [Google Scholar] [CrossRef]
  21. Newsome, D.S. The Water-Gas Shift Reaction. Catal. Rev. 1980, 21, 275–318. [Google Scholar] [CrossRef]
  22. Reddy, G.K.; Gunasekara, K.; Boolchand, P.; Smirniotis, P.G. Cr- and Ce-doped ferrite catalysts for the high temperature water-gas shift reaction: TPR and mossbauer spectroscopic study. J. Phys. Chem. C 2011, 115, 920–930. [Google Scholar] [CrossRef]
  23. Schulz, H.; Riedel, T.; Schaub, G. Fischer-Tropsch principles of co-hydrogenation on iron catalysts. Top. Catal. 2005, 32, 117–124. [Google Scholar] [CrossRef]
  24. Riedel, T.; Schulz, H.; Schaub, G.; Jun, K.W.; Hwang, J.S.; Lee, K.W. Fischer-Tropsch on iron with H2/CO and H2/CO2 as synthesis gases: The episodes of formation of the Fischer-Tropsch regime and construction of the catalyst. Top. Catal. 2003, 26, 41–54. [Google Scholar] [CrossRef]
  25. Ratnasamy, C.; Wagner, J.P. Water Gas Shift Catalysis. Catal. Rev. 2009, 51, 325–440. [Google Scholar] [CrossRef]
  26. Visconti, C.G.; Martinelli, M.; Falbo, L.; Infantes-Molina, A.; Lietti, L.; Forzatti, P.; Iaquaniello, G.; Palo, E.; Picutti, B.; Brignoli, F. CO2 hydrogenation to lower olefins on a high surface area K-promoted bulk Fe-catalyst. Appl. Catal. B Environ. 2017, 200, 530–542. [Google Scholar] [CrossRef]
  27. Lee, D.W.; Lee, M.S.; Lee, J.Y.; Kim, S.; Eom, H.J.; Moon, D.J.; Lee, K.Y. The review of Cr-free Fe-based catalysts for high-temperature water-gas shift reactions. Catal. Today 2013, 210, 2–9. [Google Scholar] [CrossRef]
  28. Fishman, Z.S.; He, Y.; Yang, K.R.; Lounsbury, A.W.; Zhu, J.; Tran, T.M.; Zimmerman, J.B.; Batista, V.S.; Pfefferle, L.D. Hard templating ultrathin polycrystalline hematite nanosheets: Effect of nano-dimension on CO2 to CO conversion: Via the reverse water-gas shift reaction. Nanoscale 2017, 9, 12984–12995. [Google Scholar] [CrossRef] [PubMed]
  29. Kim, D.H.; Han, S.W.; Yoon, H.S.; Kim, Y.D. Reverse water gas shift reaction catalyzed by Fe nanoparticles with high catalytic activity and stability. J. Ind. Eng. Chem. 2014, 23, 67–71. [Google Scholar] [CrossRef]
  30. Kunkes, E.L.; Studt, F.; Abild-Pedersen, F.; Schlögl, R.; Behrens, M. Hydrogenation of CO2 to methanol and CO on Cu/ZnO/Al2O3: Is there a common intermediate or not? J. Catal. 2015, 328, 43–48. [Google Scholar] [CrossRef]
  31. Kalamaras, C.M.; Americanou, S.; Efstathiou, A.M. “Redox” vs “associative formate with -OH group regeneration” WGS reaction mechanism on Pt/CeO2: Effect of platinum particle size. J. Catal. 2011, 279, 287–300. [Google Scholar] [CrossRef]
  32. Temkin, M.I. The Kinetics of Some Industrial Heterogeneous Catalytic Reactions. Adv. Catal. 1979, 28, 173–291. [Google Scholar]
  33. Rhodes, C.; Hutchings, G.J.; Ward, A.M. Water-gas shift reaction: Finding the mechanistic boundary. Catal. Today 1995, 23, 43–58. [Google Scholar] [CrossRef]
  34. Ginés, M.J.L.; Marchi, A.J.; Apesteguía, C.R. Kinetic study of the reverse water-gas shift reaction over CuO/ZnO/Al2O3 catalysts. Appl. Catal. A Gen. 1997, 154, 155–171. [Google Scholar] [CrossRef]
  35. Jozwiak, W.K.; Kaczmarek, E.; Maniecki, T.P.; Ignaczak, W.; Maniukiewicz, W. Reduction behavior of iron oxides in hydrogen and carbon monoxide atmospheres. Appl. Catal. A Gen. 2007, 326, 17–27. [Google Scholar] [CrossRef]
  36. Pineau, A.; Kanari, N.; Gaballah, I. Kinetics of reduction of iron oxides by H2. Part II. Low temperature reduction of magnetite. Thermochim. Acta 2007, 456, 75–88. [Google Scholar] [CrossRef]
  37. Zhang, C.L.; Li, S.; Wang, L.J.; Wu, T.H.; Peng, S.Y. Studies on the decomposing carbon dioxide into carbon with oxygen-deficient magnetite. II. The effects of properties of magnetite on activity of decomposition CO2 and mechanism of the reaction. Mater. Chem. Phys. 2000, 62, 52–61. [Google Scholar] [CrossRef]
  38. Bonnet, F.; Ropital, F.; Lecour, P.; Espinat, D.; Huiban, Y.; Gengembre, L.; Berthier, Y.; Marcus, P. Study of the oxide/carbide transition on iron surfaces during catalytic coke formation. Surf. Interface Anal. 2002, 34, 418–422. [Google Scholar] [CrossRef]
  39. Mondal, K.; Lorethova, H.; Hippo, E.; Wiltowski, T.; Lalvani, S.B. Reduction of iron oxide in carbon monoxide atmosphere—Reaction controlled kinetics. Fuel Process. Technol. 2004, 86, 33–47. [Google Scholar] [CrossRef]
  40. Yamashita, T.; Hayes, P. Analysis of XPS spectra of Fe2+ and Fe3+ ions in oxide materials. Appl. Surf. Sci. 2008, 254, 2441–2449. [Google Scholar] [CrossRef]
  41. Biesinger, M.C.; Payne, B.P.; Grosvenor, A.P.; Lau, L.W.M.; Gerson, A.R.; Smart, R.S.C. Resolving surface chemical states in XPS analysis of first row transition metals, oxides and hydroxides: Cr, Mn, Fe, Co and Ni. Appl. Surf. Sci. 2011, 257, 2717–2730. [Google Scholar] [CrossRef]
  42. McIntyre, N.S.; Zetaruk, D.G. X-ray photoelectron spectroscopic studies of iron oxides—Analytical Chemistry (ACS Publications). Anal. Chem. 1977, 49, 1521–1529. [Google Scholar] [CrossRef]
  43. Shroff, M.D.; Kalakkad, D.S.; Coulter, K.E.; Kohler, S.D.; Harrington, M.S.; Jackson, N.B.; Sault, A.G.; Datye, A.K. Activation of Precipitated Iron Fischer-Tropsch Synthesis Catalysts. J. Catal. 1995, 156, 185–207. [Google Scholar] [CrossRef]
  44. De Smit, E.; Cinquini, F.; Beale, A.M.; Safonova, O.V.; Van Beek, W.; Sautet, P.; Weckhuysen, B.M.; Supe, N.; Horowitz, R.J.; Cedex, F.-G. Stability and Reactivity of E-X-θ Iron Carbide Catalyst Phases in Fischer—Tropsch Synthesis: Controlling µC. J. Am. Chem. Soc. 2010, 132, 14928–14941. [Google Scholar] [CrossRef] [PubMed]
  45. Davis, B.H. Fischer-Tropsch Synthesis: Reaction mechanisms for iron catalysts. Catal. Today 2009, 141, 25–33. [Google Scholar] [CrossRef]
  46. Fujita, S.I.; Usui, M.; Takezawa, N. Mechanism of the reverse water gas shift reaction over Cu/ZnO catalyst. J. Catal. 1992, 134, 220–225. [Google Scholar] [CrossRef]
Figure 1. CO formation rates and their turnover frequencies (T.O.F.) on Fe3O4 at partial pressures of CO2 and H2 indicated in the legend. Other reaction conditions: Ptot = 1 bar, T = 753 K, Ftot. = 75 sccm, GHSV = 4.5 × 104 cm3 h−1 gcat.−1.
Figure 1. CO formation rates and their turnover frequencies (T.O.F.) on Fe3O4 at partial pressures of CO2 and H2 indicated in the legend. Other reaction conditions: Ptot = 1 bar, T = 753 K, Ftot. = 75 sccm, GHSV = 4.5 × 104 cm3 h−1 gcat.−1.
Catalysts 09 00773 g001
Figure 2. (a) H2 consumption (%) of the reduction period during the pretreatment following the temperature-programmed reduction (TPR) of 100 mg Fe3O4. (b) TPR curve of the post-reaction Fe3O4 sample.
Figure 2. (a) H2 consumption (%) of the reduction period during the pretreatment following the temperature-programmed reduction (TPR) of 100 mg Fe3O4. (b) TPR curve of the post-reaction Fe3O4 sample.
Catalysts 09 00773 g002
Figure 3. XRD patterns of fresh Fe3O4 (down) and post-reaction Fe3O4 (up); inset is the magnification of post-reaction Fe3O4 from 44° to 45.5°.
Figure 3. XRD patterns of fresh Fe3O4 (down) and post-reaction Fe3O4 (up); inset is the magnification of post-reaction Fe3O4 from 44° to 45.5°.
Catalysts 09 00773 g003
Figure 4. XPS Fe 2p3/2 spectra of Fe3O4 and post-reaction Fe3O4. The curves under the fitted envelope and above the background are contributions of estimated components from peak fitting.
Figure 4. XPS Fe 2p3/2 spectra of Fe3O4 and post-reaction Fe3O4. The curves under the fitted envelope and above the background are contributions of estimated components from peak fitting.
Catalysts 09 00773 g004
Figure 5. (a) CO formation rate on 100 mg Fe3C. Reaction conditions: Ftot = 75 sccm; the reactor was ramped to 753 K with PH2 = 15 kPa and He as remainder. No further reduction was applied after the ramping. During the reaction, T = 753 K, Ptot = 1 bar, PH2 = PCO2 = 15 kPa. (b) CO formation rate on 100 mg Fe3C. The temperature was directly ramped from room temperature to 773 K as shown in the figure with Ptot = 1 bar, PH2 = PCO2 = 15 kPa, and the remainder is He.
Figure 5. (a) CO formation rate on 100 mg Fe3C. Reaction conditions: Ftot = 75 sccm; the reactor was ramped to 753 K with PH2 = 15 kPa and He as remainder. No further reduction was applied after the ramping. During the reaction, T = 753 K, Ptot = 1 bar, PH2 = PCO2 = 15 kPa. (b) CO formation rate on 100 mg Fe3C. The temperature was directly ramped from room temperature to 773 K as shown in the figure with Ptot = 1 bar, PH2 = PCO2 = 15 kPa, and the remainder is He.
Catalysts 09 00773 g005
Figure 6. Ion current at m/z = 18 (H2O) and 28 (CO) during H2/CO2 switching experiments on Fe3O4. Arrows with a label indicate a change in gas composition to the indicated gas. The catalyst first was reduced in flowing H2 for 2 h, following the reaction in H2 and CO2 for 2 h, and H2/He in 20 min before the first admission of CO2. Reaction conditions: T = 773 K, FHe = 36 sccm, FH2 or FCO2 = 4 sccm.
Figure 6. Ion current at m/z = 18 (H2O) and 28 (CO) during H2/CO2 switching experiments on Fe3O4. Arrows with a label indicate a change in gas composition to the indicated gas. The catalyst first was reduced in flowing H2 for 2 h, following the reaction in H2 and CO2 for 2 h, and H2/He in 20 min before the first admission of CO2. Reaction conditions: T = 773 K, FHe = 36 sccm, FH2 or FCO2 = 4 sccm.
Catalysts 09 00773 g006
Figure 7. Ion current at m/z = 28 (CO) and 30 (C18O) during CO2/C18O2 switching experiments on Fe3O4 (100 mg). Labels in each region indicate a change in gas composition. A standard pretreatment (reduced in flowing H2 for 2 h) and RWGS reaction (>2 h) had been done before the switching experiment. Switching experiment conditions: T = 753 K, Ftot = 40 sccm, FH2 = 4 sccm, FCO2 or FC18O2 = 4 sccm.
Figure 7. Ion current at m/z = 28 (CO) and 30 (C18O) during CO2/C18O2 switching experiments on Fe3O4 (100 mg). Labels in each region indicate a change in gas composition. A standard pretreatment (reduced in flowing H2 for 2 h) and RWGS reaction (>2 h) had been done before the switching experiment. Switching experiment conditions: T = 753 K, Ftot = 40 sccm, FH2 = 4 sccm, FCO2 or FC18O2 = 4 sccm.
Catalysts 09 00773 g007
Scheme 1. Redox reaction pathway for CO formation.
Scheme 1. Redox reaction pathway for CO formation.
Catalysts 09 00773 sch001
Table 1. Measured reaction orders with respect to CO2 and H2, and measured activation energies (Emeas) over Fe3O4. Reaction conditions: 100 mg Fe3O4, Ftot = 75 sccm, T = 723 K.
Table 1. Measured reaction orders with respect to CO2 and H2, and measured activation energies (Emeas) over Fe3O4. Reaction conditions: 100 mg Fe3O4, Ftot = 75 sccm, T = 723 K.
PH2 (kPa)PCO2 (kPa)PH2: PCO2Reaction Order in CO2Reaction Order in H2Emeas (kJ/mol)
1510–20~1:10.33-28.9 ± 0.9
10–2015-0.79
3010–20~2:10.39-27.1 ± 0.5
25–3515-0.72
405–12.5~4:10.43-34.2 ± 1.9
35–4510-0.31
855–12.5~9:11.30-39.0 ± 3.4
70–9010-0
Table 2. XPS peak fittings for Fe 2p3/2 spectra for Fe3O4 and post-reaction Fe3O4.
Table 2. XPS peak fittings for Fe 2p3/2 spectra for Fe3O4 and post-reaction Fe3O4.
SamplePeak DeconvolutionBinding Energy (eV)FWHM (eV)Atomic %
Fe3O4Fe3+711.33.0865.2
Fe2+709.81.9534.8
post-reaction Fe3O4Fe3+711.22.5843.9
Fe2+709.81.5127.5
Fe3C707.91.7816.4
Metallic Fe706.70.9412.2
Table 3. Estimated initial rates of CO production after gas switches from H2 to CO2 and from CO2 to H2 during gas-switching experiment on Fe3O4 in Figure 6.
Table 3. Estimated initial rates of CO production after gas switches from H2 to CO2 and from CO2 to H2 during gas-switching experiment on Fe3O4 in Figure 6.
PeriodRate after H2 to CO2 Gas Switch
(µmol L−1 s−1 gcat.−1)
Rate after CO2 to H2 Gas Switch
(µmol L−1 s−1 gcat.−1)
(H2 to CO2 Rate)/(CO2 to H2 Rate) Ratio
1st CO22.781.641.70
2nd CO22.981.002.98
3rd CO22.740.913.01
4th CO2 (after He purge)1.940-

Share and Cite

MDPI and ACS Style

Chou, C.-Y.; Loiland, J.A.; Lobo, R.F. Reverse Water-Gas Shift Iron Catalyst Derived from Magnetite. Catalysts 2019, 9, 773. https://doi.org/10.3390/catal9090773

AMA Style

Chou C-Y, Loiland JA, Lobo RF. Reverse Water-Gas Shift Iron Catalyst Derived from Magnetite. Catalysts. 2019; 9(9):773. https://doi.org/10.3390/catal9090773

Chicago/Turabian Style

Chou, Chen-Yu, Jason A. Loiland, and Raul F. Lobo. 2019. "Reverse Water-Gas Shift Iron Catalyst Derived from Magnetite" Catalysts 9, no. 9: 773. https://doi.org/10.3390/catal9090773

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