Next Article in Journal
Maximizing Anhydrosugar Production from Fast Pyrolysis of Eucalyptus Using Sulfuric Acid as an Ash Catalyst Inhibitor
Previous Article in Journal
Thermal Stability and Potential Cycling Durability of Nitrogen-Doped Graphene Modified by Metal-Organic Framework for Oxygen Reduction Reactions
Previous Article in Special Issue
Versatile Synthesis of Pd and Cu Co-Doped Porous Carbon Nitride Nanowires for Catalytic CO Oxidation Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Improving Fe/Al2O3 Catalysts for the Reverse Water-Gas Shift Reaction: On the Effect of Cs as Activity/Selectivity Promoter

by
Laura Pastor-Pérez
1,2,*,
Mihir Shah
1,
Estelle Le Saché
1 and
Tomas Ramirez Reina
1,*
1
Department of Chemical and Process Engineering, University of Surrey, Guildford GU2 7XH, UK
2
Laboratorio de Materiales Avanzados, Departamento de Química Inorgánica Instituto Universitario de Materiales de Alicante Universidad de Alicante, Apartado 99, E03080 Alicante, Spain
*
Authors to whom correspondence should be addressed.
Catalysts 2018, 8(12), 608; https://doi.org/10.3390/catal8120608
Submission received: 14 November 2018 / Revised: 21 November 2018 / Accepted: 26 November 2018 / Published: 3 December 2018
(This article belongs to the Special Issue Preferential Oxidation of Carbon Monoxide)

Abstract

:
The conversion of CO2 into CO via the Reverse Water–Gas Shift (RWGS) reaction is a suitable route for CO2 valorisation. Fe-based catalysts are highly active for this reaction, but their activity and selectivity can be substantially boosted by adding Cs as a promoter. In this work we demonstrate that Cs modifies the redox behaviour and the surface chemistry of the iron-based materials. The metallic dispersion and the amount of metallic Fe centres available for the reaction depends on Cs loading. 5 wt. % of Cs is an optimum amount of dopant to achieve a fair activity/selective balance. Nevertheless, depending on the RWGS reactor operational temperature, lower concentrations of Cs also lead to acceptable catalytic performance. Along with the excellent activity of the prepared materials this work showcases their robustness for long-term runs and the strong impact of H2/CO ratio in the overall catalytic performance.

1. Introduction

Some of the most severe issues facing society today include anthropogenic induced climate change and ocean acidification. Atmospheric CO2 is rising year on year and global average temperatures have risen in parallel. In 2018 the concentrations of CO2 in the atmosphere passed 410 parts per million for the first time in several million years [1], and became the hottest year since records began. One method for reducing CO2 emissions and mitigating the associated climate change is to capture anthropogenic emissions and store it in geological formations, such as depleted gas reservoirs [2]. Alternatively, CO2 can be converted into value-added fuels or chemicals such as methanol.
Currently however, global demand for chemicals derived from sequestered CO2 fails to contribute significantly to help meet global carbon reduction targets; CO2 consumption by catalytic conversion is equivalent to only 0.5% of the total global CO2 emissions [3]. The limited adoption of carbon capture and utilisation in industry is partially due to the thermodynamic stability of CO2. Because of its stability, it is relatively unreactive and therefore its conversion is energy intensive. Furthermore, many of the methods for CO2 valorisation require hydrogen as a feedstock and finding a carbon neutral and cost effective hydrogen source is not always straightforward [4]. In any case, the development of an economically viable CO2 conversion process is key to achieving a circular economy pathway.
At present, the Fischer Tropsch (FT) process, which produces synthetic liquid hydrocarbon fuels, as well as the CAMERE process, which produces methanol, have both been examined extensively due to their viability for CO2 conversion on a large scale [5,6,7]. Both reactions can be carried out in conjunction with the Reverse Water–Gas Shift (RWGS) reaction, which converts CO2 and H2 into CO and H2O.
Because of its endothermic nature, RWGS is thermodynamically favourable at high temperatures as shown in Equation (1). At lower temperatures, the equilibrium will increasingly favour the water-gas shift reaction—WGS (reverse of Equation (1)) and methanation (Equation (2)) reactions, given their exothermic nature. Due to the stoichiometry of the reaction, altering pressure has no significant effect on the reaction activity or the position of the equilibrium [8,9].
H 2 + CO 2 CO + H 2 O   Δ H ° 298 k = 41.2   kJ · mol 1 ,
CO 2 + 4 H 2 CH 4 + 2 H 2 O   Δ H ° 298 k = 165   kJ · mol 1 ,
In recent years, numerous studies have been dedicated to improving the catalytic performance of the RWGS reaction, with considerable attention on monometallic catalysts such as Pt [10] and Cu [11], bimetallic catalysts such as Cu–Ni [12], and transition metal carbide catalysts, such as Mo2C [13]. Among the wide variety of catalysts reviewed in literature, iron-based catalysts have shown the greatest potential, due to their thermal stability and high oxygen mobility [14,15], while remaining a credible option in terms of costs of manufacture [16]. Transition metal oxides, such as TiO2 [17], CeO2 [18], and Al2O3 [19], have been investigated as supports for catalysts used in the RWGS reaction. The transition metal oxides support helps with dispersion of the active phase due to their large surface area. Additionally, reducible oxides boost oxygen mobility which in turns prevents coking [20]. In addition to the support, the active phase can also be promoted by adding an alkali metal [21]. Indeed the addition of alkali metals has repeatedly shown to improve CO2 adsorption, as the presence of these species significantly modifies the electronic density of the catalyst’s surface as well as improves the dispersion of the active phase [22,23]. Previous reports in literature pointed out that the valence s-orbital of the alkali metal hybridises with the valence band of the active metal, allowing the alkali metal to easily donate electrons, consequently enhancing the catalyst basicity [24].
Among the alkaline metals, potassium has been the most extensively studied as a catalytic promoter, often in conjunction with iron [25,26]. Caesium, on the other hand, has been subject to far fewer studies in the field. A recent work of our team revealed that Cs promoters can improve conversion and selectivity compared to un-promoted Fe/Cu-based catalysts [27]. Furthermore, the study also highlighted that Cs is more prone to donate electrons due to its larger ionic radius compared to potassium, implying that a Cs-promoted catalyst could be better at adsorbing CO2 compared to a K-promoted catalyst.
While it is understood why the addition of an alkali metal promoter would improve CO2 adsorption, a study by Yang et al., which tested K-doped Pt/zeolites for the RWGS reaction, has shown that while increasing the K loading initially improves CO2 conversion in the RWGS, after the K/Pt atomic ratio exceeded 80, the CO2 conversion decreased [28]. A study on CO2 conversion to propanol by Ahlers et al. which tested Cs-doped Au/TiO2, showed that there is a maximum effect on propanol selectivity between 4–7% of Cs and such a positive effect disappears for the catalysts with Cs loading above 7 wt. % [29]. A similar question arises for the RWGS—what is the effect of Cs loading as a promoter, and is there a threshold in terms of activity/selectivity improvement.
Under these premises, this study investigates the effect of Cs loading on a Fe2O3/Al2O3 catalyst for CO2 valorisation via reverse water-gas shift. The different weight loadings (1, 2.5, and 5 wt. %) help elucidate the impact of Cs on the catalytic performance and the characterisation study by means of XPS, XRD, BET, and H2-TPR is useful to correlate the catalytic trends with the structural and electronic properties of the different catalysts.

2. Results and Discussion

2.1. Textural Properties

The N2 adsorption-desorption isotherms of the calcined samples are presented in Figure 1. All samples are mesoporous materials presenting a type IV isotherm according to IUPAC classifications, also indicating multilayer adsorption. The textural properties of the catalysts are governed by the primary support (γ-Al2O3).
Table 1 lists the surface area, pore volume, and average pore diameter of each catalyst after calcination. The addition of a caesium promoter has a noticeable effect on the specific surface area and pore volume. With each increase in caesium loading, the surface area decreases, with the 5% CsFe/Al2O3 sample exhibiting a value of 149 m2/g compared to Fe/Al2O3, which has a surface area of 174 m2/g. The reduction of surface area and total pore volume are in good agreement with previous observations in literature [29], and indicates that Cs species could be partially covering the mesopores of the alumina support.

2.2. XRD

The XRD profiles in Figure 2 reveal the crystalline structure of the calcined catalysts. The samples are composed of small peaks of hematite α-Fe2O3 structure (JCPDS#-24-0072) at 2θ = 33.1°, 35.6°, 49.4°, 54.0°, 62.4°, 64.0°, and face-centred cubic crystal structure of ferrite Fe3O4 (JCPDS#-19-0629) at 2θ = 30.0°, 35.4°, 56.9°, 62.5°. This reveals that iron oxide has been through different oxidation states. Overall the peaks ascribed to iron crystalline phases are very small and hard to detect indicating a good dispersion of the iron oxide phase over the commercial alumina support. Some typical reflexions attributed to the primary support γ-Al2O3 (JCPDS# 00-048-0367) are identified by assigning peaks at 2θ = 37.6°, 45.8°, 67°. No peaks relating to Cs2O are observed, suggesting that the Cs-particles are small and dispersed over the support corroborating the successful catalysts preparation.

2.3. H2-TPR

Assessing the redox properties of the catalysts, as well as interactions between the support and metallic species, is of great importance for hydrogenation reactions like RWGS. The reducibility of the catalysts was assessed by H2-TPR studies, the profiles of which are shown in Figure 3.
All catalyst profiles present two reduction zones, which can be attributed to the reduction of the iron oxides, which occur over two separate stages. The first and larger zone, which has a range of 290–520 °C, is the first stage of the overall reduction, whereupon Fe2O3 was reduced to Fe3O4. The second zone, with a range of 580–800 °C, corresponds to the reduction of Fe3O4 to Fe through FeO as previously reported elsewhere [30].
H2-TPR can also help to elucidate the influence of the promoter on the reducibility of a specific compound in a catalyst. A remarkable shift of both peaks towards lower temperatures was observed with the addition of 1% of caesium, indicating that Cs improves the overall reducibility. Cs basic character eases electron transfer from Cs to FeOx, facilitating iron phases reduction at lower temperatures. Further increments on Cs loading (2.5 and 5%) do not have a remarkable effect on the overall reducibility of the catalysts and although the reduction zones are shifted to lower temperatures upon increasing Cs loading, it seems that 1% is enough to alter significantly the redox behaviour.

2.4. XPS

The Fe 2p3/2 and Cs 3d5/2 spectra of the reduced samples are represented in Figure 4 with the main peaks summarised in Table 2. It can be observed on the Fe 2p3/2 spectra that after the reduction at 800 °C, only a small fraction of the iron species was in the metallic state and that different iron species, such as Fe2O3 and Fe3O4, co-exist in the surface of the reduced catalysts. The Fe 2p3/2 binding energies (BE) peaks around ~706 eV are characteristic of metallic Fe, while the bands around 708.6–709.8 and 710.9–711.9 correspond to Fe2+ and Fe3+, respectively [31]. From Table 2, it is clear that increasing the Cs loading increases the Fe metallic content in the surface, with the 5% CsFe/Al2O3 sample having the highest amount of metallic Fe. An interesting effect is observed in terms of Fe3+/Fe2+/Fe proportions. For low Cs loadings, the reduction of Fe3+ to Fe2+ seems to be favoured and the Fe0 remains constant. In other words, low Cs loading seems to stabilise FeO species in the surface. However, higher Cs concentrations (5%) seem to favour the complete reduction of Fe3+ to Fe0 since this sample presents the greater proportion of metallic iron in the surface, avoiding the stabilisation of Fe2+ species. This effect is also corroborated by the Fe/Al ratios depicted in Table 2. For low Cs loadings there is less Fe exposed in the surface, probably due to the partial coverage of these species by Cs, while for the 5% sample the amount of iron exposed in the surface increases, likely due to the presence of exposed metallic iron clusters re-dispersed upon intimate interaction with Cs.
It is important to note that the reducibility of FeOx increases heavily upon addition of Cs, in good agreement with the TPR data. In fact, the total metallic iron on the surface goes from 4% in the non-promoted material to 20% in the 5% Cs doped catalysts. It is the clear that Fe in the Cs-doped catalyst is electronically richer compared to the reference system. This is a very relevant result and reflects how Cs tunes the surface chemistry of the Fe based catalysts. The electronic enrichment also implies that the Cs-doped Fe/Al2O3, especially the 5% FeCs/Al2O3 sample, should be more prone to interact and activate CO2.
As for the chemical nature of caesium species on the surface, Figure 4b depicts the XPS spectra. Overall for Cs-based samples the peak identification is relatively difficult and there is some debate in literature. For the samples with 1 and 2.5% of Cs the binding energy of the Cs 3d5/2 photoelectron around 724.5–725.6 can be ascribed to Cs in contact with FeOx oxides or CsxOy species. The shift to higher B.E. for the 2.5% Cs-doped sample could be due to the surface enrichment of Fe2+ and metallic Fe and the interaction of these species with Cs. As reported in NIST database, the close contact between Cs and metallic Fe shifts the binding energy towards higher values [32]. Interestingly, the deconvolution of Cs 3d5/2 spectra of 5% caesium-doped catalyst showed two peaks at 723.7 and 726.5 eV. Herein, the shifts to higher values are more notorious (726.5 eV) and it could be related with the higher amount of Fe metallics in this sample. Furthermore, the presence of a secondary peak at 723.7 eV can be assigned to Cs+, a more electropositive species [32].

2.5. Catalytic Behaviour

2.5.1. Effect of the Promoter on Catalytic Performance

Once the structural and electronic properties of the catalysts had been analysed, the synthesised catalysts were tested in the RWGS reaction. The catalytic activities in terms of CO2 conversion of the prepared samples are shown in Figure 5.
There is a clear effect of the temperature of the RWGS—conversion steadily increases with temperature reflecting the endothermic nature of the reaction. Interestingly, at lower temperatures (400 °C–600 °C), 2.5% CsFe/Al2O3 displayed the highest level of conversion, whilst at higher temperatures (600 °C–800 °C), 5% CsFe/Al2O3 exhibited the best conversion levels. In all the studied temperature ranges, the reference Fe/Al2O3 showed the lowest CO2 conversion, with 1% CsFe/Al2O3 having slightly higher CO2 conversion.
It could be anticipated that the 5% CsFe/Al2O3 catalyst would have the highest conversion throughout; the catalyst presented the best reducibility (as discussed in the TPR section) and a higher amount of metallic Fe, favouring the CO2 adsorption. This catalyst also presents a better Fe dispersion on the surface compared to that of the 2.5% CsFe/Al2O3, which explains its superior activity in the medium–high temperature range. The study by Yang et al. found that when loading of the alkali metal promoter becomes too great, the promoter can fail to disperse uniformly, leading to an uneven distribution of the active phase. However, the XRD profile (Figure 2) and XPS Fe/Al ratio (Table 2) show that caesium and iron were evenly distributed for all catalysts, even for relatively high Cs loadings (5% CsFe/Al2O3), providing suitable availability of active phases.
Although the 5% Cs loading seems to be the best performing catalyst, if the RWGS reaction is integrated with Fischer-Tropsch process, as proposed by Pastor-Pérez et al. [27], the 2.5% CsFe/Al2O3 catalyst would be the most logical choice, based solely on CO2 conversion. Indeed, as the Fischer-Tropsch process takes place at low temperatures, running the RWGS reactor at temperatures below 600 °C could facilitate heat and energy integration. However, if the RWGS reaction is integrated with methanol production via the CAMERE process, the 5% CsFe/Al2O3 catalyst would be most logical choice, as the CAMERE process is conducted at higher temperatures [33].
Along with conversion, selectivity is a key factor when assessing catalytic performance for RWGS, especially at low temperature, due to the strong competition with CO2 methanation which consumes a significantly larger amount of hydrogen [34]. For all catalysts, the selectivity of CO increases with temperature, while CH4 selectivity shows the opposite trend. In terms of CO selectivity, the catalyst performances were ranked as: 5% CsFe/Al2O3 > Fe/Al2O3 > 2.5% CsFe/Al2O3 > 1% CsFe/Al2O3. The excellent selectivity of the 5% CsFe/Al2O3 sample could be attributed to the boosted electronic effects of Cs on Fe. As revealed in the XPS study, the amount of metallic Fe is higher for this sample, indicating a higher charge transfer from Cs to Fe. Lower CO selectivities were obtained for the 1% and 2.5% CsFe/Al2O3 samples compared to the non-doped sample, which could be attributed to the lower Fe dispersion obtained with these samples (see Fe/Al ratios on Table 2), despite their larger content of metallic iron on the surface.
In terms of CH4 selectivity, the catalyst performances were ranked as: Fe/Al2O3 > 1% FeCs/Al2O3 > 2.5% FeCs/Al2O3 > 5% FeCs/Al2O3. The results from Figure 6 indicate that the caesium promoter inhibits the methanation reaction at lower temperatures, and that the greater amount of caesium the stronger the inhibition of CH4 production is achieved. The result of an alkali metal promoter suppressing methanation is in good agreement with previous observations in literature [10,35], highlighting that the alkali metal potassium weakens the CO adsorption catalyst’s surface. The weakened adsorption hinders further hydrogenation of CO into CH4 via C–O bond dissociation.
In summary, we identified 5% loading of Cs as the best choice in this study. In fact there are no significant benefits in terms of CO2 conversion when 5% Cs is compared to 2.5%, but there is a remarkable gain in selectivity. Indeed, the 5% Cs loaded catalyst barely forms methane even at the low temperature range. Therefore, the 5% Cs loaded sample was selected for further catalytic studies, such as the effect of H2/CO2 ratio and the stability test.

2.5.2. Effect of H2:CO2 Ratio on Catalytic Performance

A major concern of adapting RWGS for industrial applications is obtaining a sustainable and cheap source of hydrogen. This so-called renewable hydrogen is typically produced by electrolysis, however in recent times, alternative hydrogen sources have been investigated, such as using cyanobacteria, biomass, microbial electrolysis, and solar energy [4]. Despite investigation into new sources, hydrogen remains expensive and poses a significant barrier to the implementation of the RWGS for practical applications.
Given its high cost, the catalytic performance tests were repeated with H2/CO2 ratio being altered to observe the impact that decreasing the ratio would have. While it was expected that reducing the H2/CO2 ratio would decrease CO2 conversion [25,36], if the decrease in conversion is relatively small and the profitability of the process remained the same or increased, then it could be recommended to use a lower H2/CO2 ratio for the RWGS reaction.
For the test to measure the effect of H2/CO2 ratio on the catalytic performance, the H2/CO2 ratio was varied from 4:1 to 1:1 using the 5% FeCs/Al2O3 sample (Figure 7). Between 400–450 °C, CO2 conversion is similar between the H2/CO2 ratios, with conversion around 35–40% for both 4:1 and 1:1 ratio at 400 °C. However, at 500 °C a difference in CO2 conversion between the ratios begins to arise, with a difference of 11% between the 4:1 and 1:1 H2/CO2 ratios. At 800 °C, the difference in conversion widened, with a difference of 21% between the different H2/CO2 ratios. While the difference in conversion increases with temperature, the opposite is true with CO selectivity. At 450 °C, the difference in CO selectivity between the ratios was 38%, while at 800 °C, the difference in selectivity had fallen to only 9%.
If the RWGS is to be integrated with the Fischer-Tropsch process, then the 1:1 H2/CO2 ratio would be suitable, based solely on CO2 conversion results. However, the poor CO selectivity at low temperatures limits the appeal of using the 1:1 ratio, and it would be necessary to use a hydrogen rich stream with H2/CO2 ratio around 4 to maximise both activity and selectivity.

2.5.3. Stability Test

CO2 valorisation technology may require continuous operation, for instance if the CO2 conversion unit is meant to deal with continuous emissions of flue gases in a heavy carbon industry. For example, at a power plant, converting CO2 exhausted in flue gas into CO would be an ongoing process, and therefore the long-term stability of the catalyst would be vital. For the stability test, the reaction conditions were sufficiently far from equilibrium, with temperature held at 550 °C, and tested for 40 h (WHSV of 12,000 mL/gcath with a H2/CO2 ratio of 4:1). From Figure 8, the catalyst maintained stability after 40 h, with CO2 conversion remaining approximately constant at 57%. It is interesting to note that there is a slight induction period at the initial stages of the stability test with smooth increase of the conversion which may indicate an “in-situ” activation of catalysts, perhaps due to active phase re-dispersion. Overall, the promising performance of the 5% CsFe/Al2O3 catalyst in terms of stability can be linked to its robustness to sintering and carbon deposition resistance provided by Cs as a dopant, as previously demonstrated by our group [27].

3. Materials and Methods

3.1. Catalyst Synthesis

The catalysts were synthesised by sequential wet impregnation. The necessary amount of Fe (NO3)2·9H2O (Sigma-Aldrich, St. Louis, MO, USA, 99.95%) to obtain 15 wt.% Fe2O3 was dissolved in acetone. The Puralox SCFa-230 γ-alumina powder (Sasol, Johannesburg, South Africa, ≥99%) was then impregnated with the solution containing the metallic precursor for 4 h in a rotavapor. The solvent was removed by evaporation at reduced pressure in the rotavapor and the resultant slurry was dried at 100 °C overnight and calcined at 750 °C for 4 h.
The resultant Fe/Al2O3 (oxygen is Fe formulation is omitted for sake of simplicity) was divided equally, with each batch designated different loadings of Cs from 0% to 5% wt.%. For each batch, the mass of Cs2CO3 precursor (Sigma-Aldrich, ≥99%) was determined and dissolved in ethanol. Fe/Al2O3 was then impregnated for 4 h in the rotavapor with the solution containing the Cs2CO3 precursor. The solvent was removed by evaporation at reduced pressure and the resultant slurry was dried at 100 °C overnight and calcined at 750 °C for 4 h. The resulting catalyst produced were Fe/Al2O3, 1% FeCs/Al2O3, 2.5% FeCs/Al2O3, and 5% FeCs/Al2O3, where the percentage denotes the weight loading of Cs in the catalyst.

3.2. Catalyst Characterisation

X-ray diffraction (XRD) analysis was conducted on fresh and spent catalysts using an X’Pert Powder diffractometer from PANalytical (Royston, United Kingdom) at room temperature, using the powder method. Diffraction patterns were recorded at 45 kV and 40 mA, using Cu Kα radiation (λ = 0.154 nm). The 2θ angle was increased by a step size of 0.05° every 450 s over a scanning range of 10–90°.
The elemental composition was determined via means of X-ray fluorescence (XRF), which was conducted on an EDAX Eagle III spectrophotometer (Mahwah, NJ, USA) with a rhodium source of radiation (at 40 kV).
The textural properties of catalysts were determined from N2 adsorption–desorption isotherms recorded on a Micrometrics TriStar II 3020 apparatus (Norcross, GA, USA) at the boiling point of nitrogen (77 K). Prior to the adsorption–desorption measurements, the samples were degassed at 250 °C for 2 h in a vacuum. The specific surface area was determined using the Brunauer–Emmett–Teller (BET) method, whilst average pore size and pore volume were obtained by the Barret–Joyner–Halenda (BJH) method.
Temperature programmed reduction with H2 (H2-TPR) was conducted in a conventional U-shaped quartz reactor, using a gaseous mixture of 5% H2/He and flow rate of 50 mL/min, with a heating rate of 10 °C/min. Prior to the TPR runs, the catalyst was pre-treated with He (50 mL/min) at 150 °C for 1 h.
XPS measurements were performed with a K-ALPHA spectrometer (Thermo Fisher Scientific, Waltham, MA, USA) operated in the constant energy mode with survey scan pass energies of 200 eV and narrow scan energies of 50 eV, to measure the whole energy band as well as selectively measure particular elements. All XPS spectra were acquired using Al-Kα radiation (1486.6 eV) with a twin crystal monochromator, yielding a focused X-ray spot (elliptical in shape with a major axis length of 400 μm) at 3 mA × 12 kV. Charge compensation was attained with the system flood gun, which provides low energy electrons and low energy argon ions from a single source. For the reference binding energy, the C1s core level was used, located at 284.6 eV. All samples were reduced ex-situ at 800 °C, and before recording the spectrum, the samples were maintained in the analysis chamber until a residual pressure of ca. 5 × 10−7 N/m2 was reached.

3.3. Catalytic Tests

The behaviour of the studied catalysts was evaluated in a vertical fixed-bed quartz reactor (10 mm ID), at atmospheric pressure, with the product stream monitored by an on-line gas analyser (ABB AO2020 Advanced Optima Process Gas Analyser, ABB, Mannheim, Germany). 250 mg of catalyst was used for each test, supported by a bed of quartz wool centrally positioned in the reactor. Prior to each reaction, the catalysts were reduced for 1 h in situ under a total gas flow of 100 mL/min, composed of 10% H2 carried by N2 at 800 °C.
All tests were conducted at a constant weight hourly space velocity (WHSV) of 12000 mL/gcath with H2/CO2 ratio of 4:1 or 1:1 balanced in N2. The temperature range of the tests was 400 to 800 °C, increased in 50 °C increments, with a heating rate of 10 °C/min. The effluent compositions were documented for each temperature interval once the reaction had reached steady-state. The stability test was conducted for 36 h at 550 °C with a H2/CO2 ratio of 4:1. The conversions and selectivities were calculated as follows:
CO 2   conversion   ( % ) = [ CO 2 ] In [ CO 2 ] Out [ CO 2 ] In × 100 ,
CO   selectivity   ( % ) = ( [ CO ] Out ) / ( [ CO 2 ] In [ CO 2 ] Out ) × 100 ,
CH 4   selectivity   ( % ) = ( [ CH 4 ] In ) / ( [ CO 2 ] In [ CO 2 ] Out ) × 100 ,
where [CO2]In and [CO2]Out are CO2 inlet and outlet mole concentrations respectively, [CH4]Out is the CO outlet mole concentration, and [CH4]Out is the methane outlet mole concentration.

4. Conclusions

Highly effective multicomponent iron-based catalysts for CO2 valorisation via RWGS have been developed in this study. All of the catalysts tested were highly active, with high conversion and CO selectivity. However, this study reveals that the reference catalyst Fe/Al2O3 can be remarkably promoted by the addition of Cs, with the 5% FeCs/Al2O3 exhibiting the best CO selectivity and lowest CH4 selectivity. The 2.5% FeCs/Al2O3 catalyst had the best CO2 conversion at low temperatures, while the 5% FeCs/Al2O3 catalyst had the best CO2 conversion at high temperatures.
The excellent CO2 conversion of the 5% FeCs/Al2O3 at medium-high temperatures can be attributed to the fact that increasing the Cs loading improves the overall reducibility of the catalyst. The presence of Cs improves CO2 activation by facilitating the CO2 adsorption by easing electron transfer from the catalyst to the reactive species. In addition, the high CO selectivity of 5% FeCs/Al2O3 can be ascribed by Cs suppressing the formation of CH4, very likely by weakening the strength of the CO adsorption on the catalyst surface and avoiding its hydrogenation to methane. However, the better CO2 conversion of the 2.5% FeCs/Al2O3 at low temperatures indicates that too much Cs can have a counterproductive effect. Therefore, the amount of dopant and overall catalyst choice (i.e., 2.5% of Cs vs. 5%) will strongly depend on the operation conditions of the shift reactor.
The strong effect of the H2/CO2 ratio highlighted the need of a H2-rich stream to achieve high levels of activity and selectivity in the RWGS reaction. Finally, our Cs-doped catalysts presented excellent stability for long-term runs, with remarkable selectivity towards CO production while operating at relatively high space velocities indicating their suitability to be further explored for their utilisation in compact CO2 conversion reactors.

Author Contributions

Methodology, L.P.-P. and E.l.S.; formal analysis, L.P.-P.; investigation, M.S. and E.l.S.; writing—original draft preparation, M.S.; writing—review and editing, L.P.-P.; supervision, T.R.R.; funding acquisition, T.R.R.

Funding

This research was funded by the Department of Chemical and Process Engineering at the University of Surrey and by the EPSRC, grant EP/R512904/1, as well as the Royal Society, Research Grant RSGR1180353. LPP acknowledge Comunitat Valenciana for her APOSTD2017 fellowship. This work was also partially sponsored by the CO2 Chem through the EPSRC grant EP/P026435/1.

Acknowledgments

The authors acknowledge Sasol for supplying the alumina.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Nooa Research. Available online: https://research.noaa.gov/article/ArtMID/587/ArticleID/2362/Another-climate-milestone-falls-at-NOAA%E2%80%99s-Mauna-Loa-observatory (accessed on 7 November 2018).
  2. Cuéllar-Franca, R.M.; Azapagic, A. Carbon capture, storage and utilisation technologies: A critical analysis and comparison of their life cycle environmental impacts. J. CO2 Util. 2015, 9, 82–102. [Google Scholar] [CrossRef]
  3. Olivier, J.G.J.; Schure, K.M.; Peters, J.A.H.W. Trends in Global CO2 and Total Greenhouse Gas Emissions: 2017 Report; PBL Netherlands Environmental Assessment Agency: The Hague, The Netherlands, 2017. [Google Scholar]
  4. Centi, G.; Quadrelli, E.A.; Perathoner, S. Catalysis for CO2 conversion: A key technology for rapid introduction of renewable energy in the value chain of chemical industries. Energy Environ. Sci. 2013, 6, 1711–1731. [Google Scholar] [CrossRef]
  5. Jahangiri, H.; Bennett, J.; Mahjoubi, P.; Wilson, K.; Gu, S. A review of advanced catalyst development for fischer-tropsch synthesis of hydrocarbons from biomass derived syn-gas. Catal. Sci. Technol. 2014, 4, 2210–2229. [Google Scholar] [CrossRef]
  6. Joo, O.-S.; Jung, K.-D.; Yonsoo, J. Camere process for methanol synthesis from CO2 hydrogenation. In Studies in Surface Science and Catalysis; Park, S.-E., Chang, J.-S., Lee, K.-W., Eds.; Elsevier: Amsterdam, The Netherlands, 2004; Volume 153, pp. 67–72. [Google Scholar]
  7. Samimi, F.; Rahimpour, M.R.; Shariati, A. Development of an Efficient Methanol Production Process for Direct CO2 Hydrogenation over a Cu/ZnO/Al2O3 Catalyst. Catalysts 2017, 7, 332. [Google Scholar] [CrossRef]
  8. Daza, Y.A.; Kuhn, J.N. CO2 conversion by reverse water gas shift catalysis: Comparison of catalysts, mechanisms and their consequences for CO2 conversion to liquid fuels. RSC Adv. 2016, 6, 49675–49691. [Google Scholar] [CrossRef]
  9. Zhang, L.; Chen, L.; Xia, S.; Wang, C.; Sun, F. Entropy generation minimization for the reverse water gas shift (RWGS) reactors. Entropy 2018, 20, 415. [Google Scholar] [CrossRef]
  10. Liang, B.; Duan, H.; Su, X.; Chen, X.; Huang, Y.; Chen, X.; Delgado, J.J.; Zhang, T. Promoting role of potassium in the reverse water gas shift reaction on Pt/mullite catalyst. Catal. Today 2017, 281, 319–326. [Google Scholar] [CrossRef]
  11. Álvarez Galván, C.; Schumann, J.; Behrens, M.; Fierro, J.L.G.; Schlögl, R.; Frei, E. Reverse water-gas shift reaction at the Cu/ZnO interface: Influence of the cu/zn ratio on structure-activity correlations. Appl. Catal. B 2016, 195, 104–111. [Google Scholar] [CrossRef]
  12. 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]
  13. Nagai, M.; Kurakami, T. Reverse water gas shift reaction over molybdenum carbide. J. Chem. Eng. Jpn. 2005, 38, 807–812. [Google Scholar] [CrossRef]
  14. Bligaard, T.; Nørskov, J.K.; Dahl, S.; Matthiesen, J.; Christensen, C.H.; Sehested, J. The brønsted–evans–polanyi relation and the volcano curve in heterogeneous catalysis. J. Catal. 2004, 224, 206–217. [Google Scholar] [CrossRef]
  15. Peña, D.; Cognigni, A.; Neumayer, T.; van Beek, W.; Jones, D.S.; Quijada, M.; Rønning, M. Identification of carbon species on iron-based catalysts during fischer-tropsch synthesis. Appl. Catal. A 2018, 554, 10–23. [Google Scholar] [CrossRef]
  16. Wenzel, M.; Aditya Dharanipragada, N.V.R.; Galvita, V.V.; Poelman, H.; Marin, G.B.; Rihko-Struckmann, L.; Sundmacher, K. Co production from CO2 via reverse water–gas shift reaction performed in a chemical looping mode: Kinetics on modified iron oxide. J. CO2 Util. 2017, 17, 60–68. [Google Scholar] [CrossRef]
  17. Chen, X.; Su, X.; Duan, H.; Liang, B.; Huang, Y.; Zhang, T. Catalytic performance of the Pt/TiO2 catalysts in reverse water gas shift reaction: Controlled product selectivity and a mechanism study. Catal. Today 2017, 281, 312–318. [Google Scholar] [CrossRef]
  18. Wang, L.; Zhang, S.; Liu, Y. Reverse water gas shift reaction over co-precipitated Ni-CeO2 catalysts. J. Rare Earths 2008, 26, 66–70. [Google Scholar] [CrossRef]
  19. Kharaji, A.G.; Shariati, A.; Takassi, M.A. A novel γ-Alumina supported Fe-Mo bimetallic catalyst for reverse water gas shift reaction. Chin. J. Chem. Eng. 2013, 21, 1007–1014. [Google Scholar] [CrossRef]
  20. Jurković, D.L.; Pohar, A.; Dasireddy, V.D.B.C.; Likozar, B. Effect of copper-based catalyst support on reverse water-gas shift reaction (RWGS) activity for CO2 reduction. Chem. Eng. Technol. 2017, 40, 973–980. [Google Scholar] [CrossRef]
  21. Santos, J.L.; Bobadilla, L.F.; Centeno, M.A.; Odriozola, J.A. Operando DRIFTS-MS Study of WGS and rWGS Reaction on Biochar-Based Pt Catalysts: The Promotional Effect of Na. C 2018, 4, 47. [Google Scholar] [CrossRef]
  22. Sai Prasad, P.S.; Bae, J.W.; Jun, K.-W.; Lee, K.-W. Fischer–tropsch synthesis by carbon dioxide hydrogenation on fe-based catalysts. Catal. Surv. Asia 2008, 12, 170–183. [Google Scholar] [CrossRef]
  23. 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 2017, 200, 530–542. [Google Scholar] [CrossRef]
  24. Park, J.C.; Yeo, S.C.; Chun, D.H.; Lim, J.T.; Yang, J.-I.; Lee, H.-T.; Hong, S.; Lee, H.M.; Kim, C.S.; Jung, H. Highly activated K-doped iron carbide nanocatalysts designed by computational simulation for fischer–tropsch synthesis. J. Mater. Chem. A 2014, 2, 14371–14379. [Google Scholar] [CrossRef]
  25. 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]
  26. Bobadilla, L.F.; Riesco-García, J.M.; Penelás-Pérez, G.; Urakawa, A. Enabling continuous capture and catalytic conversion of flue gas CO2 to syngas in one process. J. CO2 Util. 2016, 14, 106–111. [Google Scholar] [CrossRef]
  27. Pastor-Pérez, L.; Baibars, F.; le Saché, E.; Arellano-García, H.; Gu, S.; Reina, T.R. CO2 valorisation via reverse water-gas shift reaction using advanced Cs doped Fe-Cu/Al2O3 catalysts. J. CO2 Util. 2017, 21, 423–428. [Google Scholar] [CrossRef]
  28. Yang, X.; Su, X.; Chen, X.; Duan, H.; Liang, B.; Liu, Q.; Liu, X.; Ren, Y.; Huang, Y.; Zhang, T. Promotion effects of potassium on the activity and selectivity of Pt/zeolite catalysts for reverse water gas shift reaction. Appl. Catal. B 2017, 216, 95–105. [Google Scholar] [CrossRef]
  29. Ahlers, S.J.; Pohl, M.-M.; Radnik, J.; Linke, D.; Kondratenko, E.V. Catalytic role and location of cs promoter in Cs–Au/TiO2 catalysts for propanol synthesis from CO2, C2H4 and H2. Appl. Catal. B 2015, 176–177, 570–577. [Google Scholar] [CrossRef]
  30. Ali, S.; Mohd Zabidi, N.A.; Subbarao, D. Correlation between fischer-tropsch catalytic activity and composition of catalysts. Chem. Cent. J. 2011, 5, 68. [Google Scholar] [CrossRef]
  31. Yang, L.; Pastor-Pérez, L.; Gu, S.; Sepúlveda-Escribano, A.; Reina, T.R. Highly efficient Ni/CeO2-Al2O3 catalysts for CO2 upgrading via reverse water-gas shift: Effect of selected transition metal promoters. Appl. Catal. B 2018, 232, 464–471. [Google Scholar] [CrossRef]
  32. Nist X-ray Photoelectron Spectroscopy Database. Available online: https://srdata.nist.gov/xps/Default.aspx (accessed on 8 November 2018).
  33. Park, S.-W.; Joo, O.-S.; Jung, K.-D.; Kim, H.; Han, S.-H. Development of ZnO/Al2O3 catalyst for reverse-water-gas-shift reaction of camere (carbon dioxide hydrogenation to form methanol via a reverse-water-gas-shift reaction) process. Appl. Catal. A 2001, 211, 81–90. [Google Scholar] [CrossRef]
  34. Aljishi, A.; Veilleux, G.; Lalinde, J.A.H.; Kopyscinski, J. The effect of synthesis parameters on ordered mesoporous nickel alumina catalyst for CO2 methanation. Appl. Catal. A 2018, 549, 263–272. [Google Scholar] [CrossRef]
  35. Zhang, C.; Zhao, G.; Liu, K.; Yang, Y.; Xiang, H.; Li, Y. Adsorption and reaction of CO and hydrogen on iron-based fischer–tropsch synthesis catalysts. J. Mol. Catal. A Chem. 2010, 328, 35–43. [Google Scholar] [CrossRef]
  36. Ghodoosi, F.; Khosravi-Nikou, M.R.; Shariati, A. Mathematical modeling of reverse water-gas shift reaction in a fixed-bed reactor. Chem. Eng. Technol. 2017, 40, 598–607. [Google Scholar] [CrossRef]
Figure 1. N2 adsorption-desorption isotherms of the calcined samples.
Figure 1. N2 adsorption-desorption isotherms of the calcined samples.
Catalysts 08 00608 g001
Figure 2. X-ray diffraction profiles of the calcined samples.
Figure 2. X-ray diffraction profiles of the calcined samples.
Catalysts 08 00608 g002
Figure 3. TPR profiles of the calcined samples.
Figure 3. TPR profiles of the calcined samples.
Catalysts 08 00608 g003
Figure 4. XPS core level of (a) Fe 2p and (b) Cs3d spectra.
Figure 4. XPS core level of (a) Fe 2p and (b) Cs3d spectra.
Catalysts 08 00608 g004
Figure 5. CO2 Conversion at 400–800 °C, WHSV of 12,000 mL/gcath, and H2/CO2 ratio of 4:1.
Figure 5. CO2 Conversion at 400–800 °C, WHSV of 12,000 mL/gcath, and H2/CO2 ratio of 4:1.
Catalysts 08 00608 g005
Figure 6. CO and CH4 Selectivity at 400–800 °C, WHSV of 12,000 mL/gcath, and H2/CO2 ratio of 4:1.
Figure 6. CO and CH4 Selectivity at 400–800 °C, WHSV of 12,000 mL/gcath, and H2/CO2 ratio of 4:1.
Catalysts 08 00608 g006
Figure 7. (a) CO2 Conversion and (b) CO and CH4 selectivity at 400–800 °C, WHSV of 12,000 mL/gcath, and various H2/CO2 ratios using 5% CsFe/Al2O3.
Figure 7. (a) CO2 Conversion and (b) CO and CH4 selectivity at 400–800 °C, WHSV of 12,000 mL/gcath, and various H2/CO2 ratios using 5% CsFe/Al2O3.
Catalysts 08 00608 g007
Figure 8. Stability test for CO2 Conversion at 550 °C, WHSV of 12,000 mL/gcath, and H2/CO2 ratio of 4:1.
Figure 8. Stability test for CO2 Conversion at 550 °C, WHSV of 12,000 mL/gcath, and H2/CO2 ratio of 4:1.
Catalysts 08 00608 g008
Table 1. Textural Properties of the prepared catalysts.
Table 1. Textural Properties of the prepared catalysts.
CatalystSurface Area (m2/g)Pore Volume (cm3/g)Pore Diameter (nm)
Fe/Al2O31740.4708.73
1% CsFe/Al2O31680.4678.83
2.5% CsFe/Al2O31530.4218.75
5% CsFe/Al2O31490.4018.83
Table 2. Binding energies of the Fe 2p3/2 and Ce 3d5/2 levels for the pre-reduced catalysts and Fe/Al atomic ratios.
Table 2. Binding energies of the Fe 2p3/2 and Ce 3d5/2 levels for the pre-reduced catalysts and Fe/Al atomic ratios.
CatalystsFe 2p3/2 (eV)Cs 3d5/2 (eV)Fetotal/Al (at/at)Cs/Fe
Fe3+Fe2+FeCs1+
5% CsFe/Al2O3710.9 (62%)708.6 (18%)706.7 (20%)723.7–725.90.0770.708
2.5% CsFe/Al2O3711.5 (25%)709.5 (63%)706.7 (14%)725.50.0480.460
1% CsFe/Al2O3711.9 (34%)709.8 (52%)706.5 (13%)724.50.0590.202
Fe-Al2O3712.8 (43%)710.3 (53%)706.5 (4%)-0.077-

Share and Cite

MDPI and ACS Style

Pastor-Pérez, L.; Shah, M.; Le Saché, E.; Ramirez Reina, T. Improving Fe/Al2O3 Catalysts for the Reverse Water-Gas Shift Reaction: On the Effect of Cs as Activity/Selectivity Promoter. Catalysts 2018, 8, 608. https://doi.org/10.3390/catal8120608

AMA Style

Pastor-Pérez L, Shah M, Le Saché E, Ramirez Reina T. Improving Fe/Al2O3 Catalysts for the Reverse Water-Gas Shift Reaction: On the Effect of Cs as Activity/Selectivity Promoter. Catalysts. 2018; 8(12):608. https://doi.org/10.3390/catal8120608

Chicago/Turabian Style

Pastor-Pérez, Laura, Mihir Shah, Estelle Le Saché, and Tomas Ramirez Reina. 2018. "Improving Fe/Al2O3 Catalysts for the Reverse Water-Gas Shift Reaction: On the Effect of Cs as Activity/Selectivity Promoter" Catalysts 8, no. 12: 608. https://doi.org/10.3390/catal8120608

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