Next Article in Journal
Research Progress of Bioactive Components in Sanghuangporus spp.
Previous Article in Journal
Advanced Photodegradation of Azo Dye Methyl Orange Using H2O2-Activated Fe3O4@SiO2@ZnO Composite under UV Treatment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Brief Review of Recent Theoretical Advances in Fe-Based Catalysts for CO2 Hydrogenation

by
Haoxiang Tang
,
Tongyue Qiu
,
Xuerui Wang
,
Chundong Zhang
* and
Zunmin Zhang
*
State Key Laboratory of Materials-Oriented Chemical Engineering, College of Chemical Engineering, Nanjing Tech University, Nanjing 211816, China
*
Authors to whom correspondence should be addressed.
Molecules 2024, 29(6), 1194; https://doi.org/10.3390/molecules29061194
Submission received: 9 February 2024 / Revised: 4 March 2024 / Accepted: 5 March 2024 / Published: 7 March 2024
(This article belongs to the Section Green Chemistry)

Abstract

:
Catalytic hydrogenation presents a promising approach for converting CO2 into valuable chemicals and fuels, crucial for climate change mitigation. Iron-based catalysts have emerged as key contributors, particularly in driving the reverse water–gas shift and Fischer–Tropsch synthesis reactions. Recent research has focused on enhancing the efficiency and selectivity of these catalysts by incorporating alkali metal promoters or transition metal dopants, enabling precise adjustments to their composition and properties. This review synthesizes recent theoretical advancements in CO2 hydrogenation with iron-based catalysts, employing density functional theory and microkinetic modeling. By elucidating the underlying mechanisms involving metallic iron, iron oxides, and iron carbides, we address current challenges and provide insights for future sustainable CO2 hydrogenation developments.

1. Introduction

Fossil fuels, including coal, oil, and natural gas, have historically served as the primary sources of energy to meet increasing global demands. However, their combustion releases substantial amounts of greenhouse gases, notably carbon dioxide (CO2), significantly contributing to climate change [1,2,3]. Efforts to develop efficient carbon dioxide capture and utilization technologies arise not only from the need to mitigate atmospheric CO2 levels but also from the opportunity to convert captured CO2 into valuable products [4,5]. Among these technologies, catalytic CO2 hydrogenation emerges as a highly promising approach for CO2 recycling, particularly when coupled with renewable hydrogen sources like solar, wind, or hydropower [6,7]. This approach’s attractiveness lies in its ability to utilize CO2 as a resource, transforming it into valuable chemicals such as methanol, olefins, and paraffin [8], thereby fostering sustainability and reducing reliance on fossil fuels [9]. As renewable energy technologies advance, catalytic CO2 hydrogenation is expected to play a pivotal role in carbon cycling and carbon neutrality initiatives [10]. Despite the substantial promise of catalytic CO2 hydrogenation, its broad adoption remains constrained by certain barriers [11]. Researchers are currently involved in actively optimizing catalyst design to improve reaction rates, selectivity, and stability in the hydrogenation process, utilizing both experimental and theoretical methods [12].
Fe-based catalysts are distinguished among metal catalysts by their demonstrated cost-effectiveness, abundant product yields, and exceptional catalytic performance in CO2 hydrogenation, encompassing both the endothermic Reverse Water–Gas Shift (RWGS) and exothermic Fischer–Tropsch Synthesis (FTS) processes [13,14,15]. Their proven activity in RWGS and FTS reactions has attracted attention for direct CO2 hydrogenation [16]. For example, iron oxide-based catalysts have exhibited notable efficiency in high-temperature water–gas shift reactions [17,18], while iron carbides (FeCx), particularly Hägg-carbide (χ-Fe5C2), are usually considered as the catalytically active phases in CO2-FTS [19,20,21]. Compared to Co-based and Ru-based catalysts, iron-based catalysts exhibit higher affinity for olefins and long-chain hydrocarbons [22]. Recent research has focused on alkali metal-modified Fe-based catalysts, such as sodium (Na), potassium (K), and rubidium (Rb) [23]. These modifiers have demonstrated the capacity to reshape the electronic structure of catalyst interfaces, leading to enhanced CO2 adsorption, the formation of Fe5C2-carbonates, and reduced energy barriers for olefin desorption [24,25,26]. Additionally, doping with transition metals like zinc (Zn), copper (Cu), manganese (Mn), and cobalt (Co) has refined catalyst evolution and reliability [24,27,28,29]. The interplay between different metals and oxygen adsorbed on iron species significantly influences dissociation activation energy and CO2 conversion pathways. In addition, zeolite-confined Fe-based catalysts have also garnered significant interest for CO2 hydrogenation, owing to their exceptional efficiency, versatile framework, and customizable composition. The role of zeolite properties, including topology, the nature, density, and strength of acid sites, and metal–zeolite interactions, has been comprehensively reviewed in the literature for its influence on catalytic activity and product selectivity [30,31,32].
In the field of catalysis research, quantum mechanics methods, particularly density functional theory (DFT), have become indispensable in recent decades [33,34,35]. These computational tools offer invaluable insights into the intricate atomic-scale mechanisms underlying CO2 hydrogenation. DFT has exhibited remarkable efficacy in specific aspects of CO2 hydrogenation, enabling the exploration of novel intermediates, the scrutiny of potential catalysts, and the elucidation of reaction pathways [36,37]. Moreover, the integration of experimental data with theoretical findings establishes crucial connections between catalytic activity and characteristics in RWGS, facilitating the advancement of cutting-edge catalytic materials [38,39]. Notably, DFT calculations of CO2 behavior on Fe-based catalyst surfaces provide crucial insights into adsorbed species evolution, offer information about rate-determining steps, and reveal overall reaction mechanisms [40,41]. This computational paradigm not only accelerates catalyst development for CO2 hydrogenation but also improves our understanding of chemical reactivity and the simulation of intricate catalytic reaction pathways [42,43].
While some previous studies have offered general or specific summaries of CO2 hydrogenation with iron-based catalysts [12,22,23,42,44], this review exclusively focuses on recent theoretical achievements in this field, predominantly employing density functional theory and microkinetic modeling. The principal objective is to elucidate the underlying mechanisms involving metallic iron, iron oxides, iron carbides, and the influence of alkali or transition metals on CO2 redox and association processes. This work aims to provide a comprehensive summary of the oxidation-reduction mechanism, as well as the associative pathways of CO2. Additionally, we anticipate the potential contribution of machine learning and artificial intelligence in driving novel advancements in catalyst design for CO2 hydrogenation.

2. Related Reactions and Computational Methods

In the RWGS reaction, renewable hydrogen (H2) and CO2 undergo a redox reaction to generate carbon monoxide (CO) and water (H2O) (Equation (1)). Potential pathways include the formation of methane (CH4) (Equation (2)) and methanol (CH3OH) (Equation (3)) [45]. Governed by Le Chatelier’s principle, the endothermic nature of the reaction thermodynamically favors higher temperatures. Consequently, as temperatures decrease, the equilibrium gradually shifts to promote the exothermic reverse of Equation (1) and methanation reactions, which predominantly occur as side reactions [46]. Effective catalyst surfaces should facilitate CO desorption while impeding further hydrogenation into hydrocarbons, necessitating the precise design of catalysts and careful regulation of reaction conditions to optimize CO2 valorization. In contrast to the indirect route, where CO2 transforms into more active intermediates like CO or MeOH before conversion into hydrocarbons, the direct RWGS reaction is often coupled with FTS for the production of fuel range hydrocarbons from syngas (Figure 1a). In this process, Fe3O4 serves as a RWGS catalyst, while χ-Fe5C2 is employed for FTS [47,48]. This tandem CO2-FTS integrates RWGS for CO2 activation with C-C coupling through FTS, thus facilitating the generation of extended hydrocarbon chains (Figure 1b) [49,50,51].
CO2 + H2 ⇆ CO + H2O ΔH = 42.1 kJ/mol
CO2 + 4H2 → CH4 + 2H2O ΔH = −165 kJ/mol
CO2 + 3H2 → CH3OH + H2O ΔH = −49.5 kJ/mol
The reaction kinetics data indicate that the main pathways for the RWGS reaction are the redox and association mechanisms [44,52,53], as presented in Table 1. In the redox mechanism, the catalyst oxidizes CO2 to form CO (CO2 → CO2* → CO* → CO). In the association mechanism, CO2 typically adsorbs onto the catalyst surface, leading to the formation of intermediate species such as COOH and HCOO during the reaction, which subsequently decompose into CO and H2O. COOH and HCOO pathways depend on CO2* and H* (CO2 → CO2* → COOH* → CO* → CO, CO2 → CO2* → HCOO* → HCO* → CO* → CO). The final products and the related multiphase catalytic processed involving CO2 intermediates and reaction surfaces are mainly affected by the reaction operating conditions and the composition of the Fe-based catalyst.
While catalyst development traditionally relies on trial-and-error experimental methods, the emergence of multiscale models for designing and refining catalysts has become indispensable [54]. Typically, scientists employ DFT to validate experimental findings, including catalytic conversion, selectivity, and characterizations [55]. The generalized gradient approximation (GGA) with the Perdew–Burke–Ernzerhof (PBE) functional for electron exchange–correlation energy approximation is commonly used for the structural optimization of Fe-based catalysts [56]. For transition metal oxides, conventional treatments often underestimate equilibrium lattice constants and overestimate binding energies, necessitating the incorporation of an effective Hubbard U value for accurate predictions [57]. Furthermore, conventional semi-local DFT-GGA functionals lack inherent inclusion of the long-range electronic correlations responsible for van der Waals (vdW) interactions. Consequently, there is a tendency to underestimate the binding affinity of sizable, closed-shell adsorbates interacting predominantly with catalytic surfaces via vdW interactions, such as aromatic compounds and extended hydrocarbon chains [58]. In heterogeneous catalysis, empirical vdW corrections are commonly used in DFT studies to address vdW interactions, rather than relying solely on vdW functionals [59]. Electronic structure calculations entail constructing an atomic-scale structural catalyst model and scrutinizing energy variations associated with fundamental steps in the reaction mechanism [60].
Despite the complexity of real catalyst materials, first-principle methods are employed to model chemical reactions at the atomic level, enabling the assessment of reaction properties and catalyst composition [61,62]. Microkinetic models (MKMs) are commonly utilized to unveil the fundamental reaction mechanisms of heterogeneous catalytic reactions, leveraging energy insights obtained from DFT calculations. Unlike electronic structure methods that elucidate how intermediates move on catalytic surfaces, MKMs bridge the gap between atomic-scale phenomena and macroscopic properties, facilitating the prediction of macroscopic reaction kinetics data. To precisely characterize the interactions between adsorbates and reaction energetics, Bhandari et al. [63] proposed an algorithmic framework that integrates first-principles calculations, microkinetic modeling, and reaction kinetics experiments (Figure 2). While the multiscale methodology remains the preferred technique for modeling heterogeneous catalysis, theoretical progress has enhanced the faithfulness and precision of each phase of the multiscale procedure [64]. The remarkable power of computational techniques to investigate the atomic-level intricacies of catalytic systems offers tremendous opportunities for the fundamental, bottom-up design of new heterogeneous catalysts [65].

3. Modelling Fe-Based Catalyst Surfaces

The heterogeneous catalytic conversion of CO2 and H2 occurs across a diverse range of iron-based catalysts, highlighting the need for a comprehensive understanding of CO2 activation mechanisms at their active sites. This demands thorough analysis, bridging diverse mechanisms with atomic-scale properties [66,67]. Numerous DFT studies have investigated CO2 adsorption on specific metal surfaces and H2 dissociation to elucidate conversion mechanisms [68,69]. However, a cohesive and accessible synthesis of these findings is still lacking [70]. Thus, this section aims to systematically summarize recent theoretical research on iron-based catalytic surfaces for CO2 hydrogenation and subsequent conversion into value-added products.

3.1. Iron-Based Catalysts

In FTS, the active phase of Co-based catalysts primarily exists in the metallic form, whereas Fe-based catalysts typically comprise complex species, including iron carbides, oxides, or metallic iron [71,72]. Iron species demonstrate the ability to produce both long-chain and short-chain hydrocarbons, with variations in product selectivity potentially attributed to iron’s comparatively weaker hydrogenation capability [73]. Nonetheless, the intricate interplay among these species considerably complicates the elucidation of their individual contributions to the overall reaction.

3.1.1. Metallic Irons

Significant scientific attention has been devoted to studying the intrinsic catalytic properties of CO2 adsorption and activation on Fe (100) and (110) surfaces. Liu et al. [74] employed DFT to explore the CO2 reduction process to CO on Fe, Co, Ni, and Cu surfaces. Their investigation highlighted the intrinsic chemical adsorption of CO2 on these surfaces (Figure 3a), driven by thermodynamic attributes. Notably, Fe (100) exhibited superior propensity for CO2 adsorption compared to the other metals studied. Valence electron density calculations provide insight into the CO2 activation, revealing charge transfer from the metal surface to the CO2 component. Overall, the total reaction energy barriers aligned with the trend of reaction energy: Fe < Co < Ni < Cu.
To evaluate the potential of metallic iron as a catalyst for the water–gas shift reaction, Liu et al. [75] meticulously simulated the interactions of CO and H2O on the Fe (110) surface at varying ratios (Figure 3b). They also investigated the effects of surface precoverage with 0.25 monolayer O, OH, and H. Their findings indicated that Fe (110) favored CO2 dissociation in terms of both kinetics and thermodynamics. Chen et al. [78] explored H2 activation and CO2 pre-hydrogenation reactions on four transition metal surfaces: Fe (111), Ni (111), Ru (111), and Pt (111). They discovered that the active hydrogen generated from H2 dissociation can influence the hydrogenation pathway of CO2. The metal electronegativity affects the selective hydrogenation of CO2. Specifically, on Pt (111) surface, active hydrogen can provide electrons through the Pt dxy orbital, generating positively charged H+ species and favoring O binding of CO2. Conversely, on Fe (111) surface, active hydrogen accepts electrons via electron back-donation from the Fe dxy orbital, leading to the formation of negatively charged H species, thereby facilitating the hydrogenation of the C-terminal of CO2. In another study, Liu et al. [76] systematically investigated the adsorption and dissociation of CO2 on various low-index transition metal surfaces (Figure 3c). They observed CO2 chemisorbing with a bent configuration on Fe (110). Moreover, they noted that CO2 activation is influenced by the type of transition metal and the surface geometry, with changes in charge transfer correlating with variations in adsorption energy.
Additionally, detailed investigations into the reaction mechanisms of CO2 and H2 on various iron surfaces, including changes in reaction energetics for intermediate formation, have been reported. Wang et al. [77,79] conducted comprehensive research exploring the adsorption, dissociation, and hydrogenation mechanisms of CO2 on Fe (100), (110), (111), and (211) surfaces (Figure 3d). They identified optimal configurations for CO2 and H2 on the Fe (100) surface, achieving enhanced adsorption and reactant activation through appropriate adjustment of the CO2/H2 ratio. Their research revealed that CO2 adsorption was more favorable on Fe (111) and (211) surfaces, while it was weakest on the (110) surface. Due to lower dissociation barriers, H2 on the Fe surface undergoes rapid activation and dissociation, emphasizing the significance of maintaining co-adsorption equilibrium of H2-CO2 for effective reactant activation. On the (100) and (110) surfaces, CO2 selectively undergoes direct dissociation to form CO*, while Fe (111) is more favorable for HCOO* formation owing to a lower kinetic barrier. However, the (211) surface shows a stronger competition between CO* and HCOO* formation. Notably, none of the Fe surfaces are conducive to COOH* intermediate formation. It is speculated that Fe (111) demonstrates excellent activation ability in CO2 conversion due to its lower initial hydrogenation barrier.

3.1.2. Iron Oxides

Fe catalysts characterized in situ and after the reaction have been identified in their oxidized states [12,52,80,81,82]. Parkinson [83] conducted a systematic review of various iron oxides’ surfaces, including magnetite (Fe3O4), maghemite (γ-Fe2O3), hematite (α-Fe2O3), and wustite (Fe1−xO) (Figure 4a). The discussion explored the correlation between the adsorption of molecules on the surface and the reactions of iron oxide catalysts, with a particular emphasis on the extraordinary thermal stability mechanism of isolated metal adatoms on the Fe3O4 surface. Chou et al. [84] demonstrated the promising activity of an unsupported Fe3O4 catalyst for the oxidative-reductive conversion of CO2 to form CO. Fishman et al. [85] focused on hematite nanosheets, nanowires, and nanoparticles to explore the nano-dimensionality of iron oxide based catalysis on the RWGS reaction. They investigated the mechanism by which iron oxide nanomaterials convert CO2 to CO through H2-TPR and RWGS and compared their findings with DFT-based H2 binding energy results. Band gap and reaction activity exhibit a certain correlation, with nanowires and nanosheets having lower band gaps being more active at low temperatures compared to nanoparticles with higher band gaps. The results showed that nanosheets exhibited a higher CO2 conversion rate of 28% at 510 °C, while nanowires achieved a remarkable conversion rate of 50% at 750 °C under atmospheric pressure.
In theoretical terms, Fe3O4 (111) emerges as the energetically favored surface [86]. Yang et al. [87] reported a systematic DFT study on the hydrogen adsorption behavior on the Fe3O4 (111) surface with two distinct terminal configurations, Fetet1- and Feoct2-termination. The findings revealed that on the Fetet1-termination surface, H tends to adsorb and dissociate more readily, which is contrary to the adsorption behavior of CO on the same surface. Yu et al. [88] employed the GGA+U method to compute the properties of different Fe3O4 (111), (110), and (001) surfaces (Figure 4b). They utilized the GGA+U method to accurately capture the strong correlation effects of iron’s 3d orbitals and investigated the electronic structure, stability, and magnetism of Fe3O4. The computational results closely aligned with experimental lattice parameters and magnetic moments. Their study revealed that Fetet1 and Feoct2 are the most stable among the six Fe3O4 (111) terminations (Ooct2, Ooct1, Feoct1, and Fetet2), exhibiting similar surface energies. David et al. [89] performed the GGA+U approximation of DFT to model three different orientations of the Fe3O4 bulk. They investigated the stability of non-dipolar stoichiometric surface terminations and explored the redox properties of the surfaces. By introducing oxygen vacancies or adding oxygen atoms to the most stable non-dipolar stoichiometric surface terminations, they altered the redox conditions. Consistent with previous experimental STM images, the study concluded that (001) and (111) are the most stable surfaces of Fe3O4. While both (001) and (111) surfaces undergo oxidation under ambient conditions, they both experience a gradual reduction process. On the (001) surface, the reduction process initiates at lower chemical potentials and encompasses the stoichiometric plane. Su et al. [90] conducted first-principles calculations on the chemical adsorption of CO2 on Feoct2-tet1- or Fetet1-terminated Fe3O4 (111) surfaces (Figure 4c). The study revealed that the Feoct2-tet1-terminated surface is more active compared to the Fetet1-terminated surface. On the Feoct2-tet1-terminated surface, CO2 acts as an electron acceptor from the surface, leading to the formation of covalent bonds between the C and O atoms of CO2 and the surface. This observation is supported by the analysis of the electron localization function, the difference electron density, and the partial density of states (PDOS). Han et al. [52] combined DFT and microkinetic analysis, utilizing CO2 and OH binding energies as descriptors, to investigate the CO2 hydrogenation reactions on typical iron-based surfaces (Fe3C, Fe5C2, Fe2O3, Fe3O4). In this theoretical framework, Fe3O4 is considered the predominant active phase for the RWGS reaction, and the direct CO2 dissociation mechanism governs the RWGS process. A degree of rate control analysis reveals that OH* removal is the rate-determining step for most iron-based surfaces [91,92,93]. By incorporating Cu and Zn into the Fe3O4 surface structure, the reaction can be effectively promoted. The authors believe that identifying the active phases before screening promoters is beneficial for analyzing multiphase catalysts with various surfaces.
Figure 4. (a) The iron oxides are based on a close packed O2− anion lattice with metal cations in octahedral and tetrahedral coordinated interstitial sites. Reprinted with permission from Parkinson [83]. Copyright 2016 Elsevier. (b) Fe3O4 (111), (110), and (001) surface and its different terminations (red for O; blue for Fe). Reprinted with permission from Yu et al. [88]. Copyright 2012 Elsevier. (c) Side view and top view for the Fetet1-terminated and Feoct2-tet1-terminated Fe3O4 (111) surfaces (blue, Fe atom; red, oxygen atom). Reprinted with permission from Su et al. [90]. Copyright 2016 Elsevier.
Figure 4. (a) The iron oxides are based on a close packed O2− anion lattice with metal cations in octahedral and tetrahedral coordinated interstitial sites. Reprinted with permission from Parkinson [83]. Copyright 2016 Elsevier. (b) Fe3O4 (111), (110), and (001) surface and its different terminations (red for O; blue for Fe). Reprinted with permission from Yu et al. [88]. Copyright 2012 Elsevier. (c) Side view and top view for the Fetet1-terminated and Feoct2-tet1-terminated Fe3O4 (111) surfaces (blue, Fe atom; red, oxygen atom). Reprinted with permission from Su et al. [90]. Copyright 2016 Elsevier.
Molecules 29 01194 g004

3.1.3. Iron Carbides

In the study of iron carbides, researchers have focused on identifying the most stable active phases. Smit et al. [94] conducted a thorough analysis, integrating experimental and theoretical approaches to compare the reactivity of different ε-χ-θ carbide phases under varying conditions. The carbon chemical potential (μC) was used as an indicator to describe thermodynamically induced phase transitions in the catalyst system. At low temperatures and high μC regimes, ε-carbide is primarily formed, whereas at lower μC conditions, the emergence of θ-Fe3C is observed. Under high-pressure FTS conditions, a gradual evolution from θ-Fe3C to χ-Fe5C2 was found. Additionally, the presence of an amorphous carbon/carbide layer commonly found on Fe-based catalysts tends to catalyze the transformation of carbides during FTS. Lyu et al. [95] proposed a novel catalytic approach by confining pure ε-Fe2C nanocrystals within a graphene matrix, resulting in enhanced activity and stability in contrast to traditional carbon-loaded Fe catalysts (Figure 5a). Additionally, DFT analysis further elucidates the interfacial interactions within the ε-Fe2C@graphene composite, demonstrating the efficacious suppression of amorphous carbon layer formation. Zhao et al. [96] synthesized a series of iron carbide (Fe3C, Fe7C3, and Fe5C2) nanoparticles, investigating their catalytic activity in FTS. They discovered that Fe5C2 exhibited superior catalytic activity and stability compared to other carbide phases (Figure 5b). DFT calculations further revealed that due to strong interactions with the dissociated atomic carbon on iron, iron carbides are inherently more active than elemental iron in terms of C-C coupling and methane formation. Moreover, the study demonstrated that while methane formation occurs on iron carbides, C-C coupling reactions are notably more facile, thus underscoring the exceptional activity of iron carbides in FTS and their remarkable selectivity towards olefins.
Furthermore, a comparison of CO2 conversion pathways on the two carbides have been conducted. Liu et al. [97] investigated the mechanisms of CO2 adsorption and activation on thermodynamically stable χ-Fe5C2 (510) and θ-Fe3C (031) surfaces (Figure 6). They explored four primary pathways for CO2 activation, including direct dissociation and intermediates involving H, such as COOH*, HCOO*, and CO* + OH* (Figure 5c). The findings suggested that both surfaces exhibited activity for the direct dissociation pathways and were unfavorable for the formation of COOH* [19,77,98]. On χ-Fe5C2 (510), CO* + OH* could be formed in a single step, while θ-Fe3C (031) favored the formation of HCOO* and CO* + OH* intermediates. Additionally, on Fe5C2 (510), hydrogen displayed distinct behavior across four different adsorption sites (four-fold site, three-fold site, bridge site, and top site). Zhang et al. [99] observed that hydrogen at the three-fold site and four-fold site did not participate in hydrogenation reactions but tended to migrate to another three-fold site or four-fold site, resulting in increased hydrogen coverage on Fe5C2 (510). Meanwhile, hydrogen at the top site and bridge site engaged in hydrogenation reactions on Fe5C2 (510). Wang et al. [100] utilized DFT and microkinetic modeling to investigate the mechanism of CO2 hydrogenation to hydrocarbons over iron carbide catalysts. The study explored three possible initial conversion pathways of CO2 on the χ-Fe5C2 (510) surface, including CO2 dissociation to CO, CO2 hydrogenation to COOH*, and HCOO* intermediates (Figure 7). The formation of COOH* is energetically unfavorable, and even if it can be formed on the χ-Fe5C2 (510) surface, it will dissociate into CO* and OH* species with a potential barrier of 0.72 eV. On the other hand, the formation of the HCOO* intermediate occurs when the surface H* attacks the C atom of CO2 to form a C-H bond, and this process is kinetically more favorable than the formation of COOH* via an O-H bond. The energy barrier for CO2 dissociation into CO* and O* is 0.64 eV, making it kinetically more favorable compared to COOH* and HCOO*. The calculated results suggest that direct dissociation to CO* and hydrogenation to HCOO* are the two main pathways on the χ-Fe5C2 (510) surface.
An in-depth understanding of the adsorption and dissociation processes of H2O is essential for comprehension of the impact of iron carbide oxidation by H2O on the FTS process. Gao et al. [101] employed the GGA-PBE method to investigate the adsorption and dissociation of H2O on the Fe5C2 (010) surface. Their investigation revealed distinct iron-rich and carbon-rich regions on the Fe5C2 (010) surface, where the adsorption energy results suggested that the iron-rich region is active for H2O adsorption and dissociation, while the carbon-rich region remains inert. Notably, H2O tends to preferentially adsorb on the top position of surface iron atoms within the iron-rich region. Both thermodynamic and kinetic analyses indicated that H2O dissociation in the iron-rich region is energetically favored. Furthermore, thermodynamic analysis elucidated that the catalyst surface continuously adsorbs oxygen atoms in a water environment, with the quantity of oxygen atoms in the iron-rich region contingent on temperature and water content. High temperatures and low H2O partial pressures were found to promote catalyst stability. By comparing relevant findings, it can be inferred that the adsorption and dissociation of H2O on Fe5C2 (010) at low coverage closely resemble those observed on Fe (100).

3.2. Promoters

3.2.1. Alkali Metals

Alkali metal salts (or oxides) are commonly utilized as electronic promoters in iron-based catalysts, facilitating electron donation and adjustment of the electronic structure on catalyst surfaces. This enhancement augments the adsorption strength of CO2 and reduces the energy barrier for olefin desorption. Nevertheless, their specific impact on catalytic properties and mechanisms remains a topic of ongoing debate [102].
The electron interaction between potassium (K) atoms and CO2 molecules enhances the adsorption strength of CO2, while K2O promotes the formation of highly active surfaces. Nie et al. [19] studied the effects of potassium on the adsorption, activation, and dissociation of CO2 on the surfaces of Fe (100), Fe5C2 (510), and Fe3O4 (111). The research demonstrated that the electronic interactions between CO2 molecules and the K-covered surfaces significantly enhance the adsorption and activation of CO2 by potassium. The order of adsorption strength follows the order oct-Fe3O4 (111) > Fe (100) > Fe5C2 (510). In the presence of potassium, the dissociation barrier of CO2 molecules is reduced, favoring CO2 dissociation on the Fe (100) and Fe5C2 (510) surfaces. However, the influence of potassium on the dissociation of CO2 on the Fe3O4 (111) surface appears limited, as a relatively high dissociation barrier persists. Ma et al. [103] examined the adsorption and activation of CO2 on clean and K-precovered transition metals. CO2 activation on K is primarily governed by electronic and geometric factors. Introduction of potassium significantly enhances the binding strength of CO2 and generally reduces activation energy (Figure 8a). This enhancement is mainly attributed to direct electron transfer and dipole–dipole interactions. Huo et al. [104] utilized K2O/Fe as a model catalyst and investigated its morphology control effect through both theoretical and experimental approaches. Their findings indicate that the K promoter can modify crystal orientations, favoring the formation of crystals with abundant highly active surfaces. This stabilization effect alters the relative growth rates of crystals in different directions, providing insights for designing catalysts with controllable surface structures. Chen et al. [105] explored the activation of CO2 and H2 molecules on the surfaces of Fe5C2 (110) and K2O/Fe5C2 (110), revealing the promoting effect of K2O on the hydrogenation of CO2 to olefins (Figure 8b). Their research suggests that the adsorption of K2O facilitates the direct activation of CO2 into CO*, thereby promoting the C-C coupling reaction between CO* and the surface carbon of Fe5C2. Regarding the activation of H2, the K2O-adhered Fe5C2 (110) surface forms Fe-H bonds, which not only lowers the reactivity of H species but also reduces CH4 species, preventing the excessive hydrogenation of olefins to saturated alkanes.
Sodium (Na) not only engages in electrostatic interactions similar to potassium (K) but also, Na2O significantly alters the surface electronic structure, thereby enhancing olefin selectivity. Mahyuddin et al. [108] investigated the impact of Na and K on the adsorption properties of CO and H2S on the Fe (100) surface. Through local density of states and Bader charge analysis, they found that the alkali metal adsorbates reinforce CO adsorption due to the electrostatic interactions between the adsorbates and the molecules. However, they noted a poisoning effect hindering H2S adsorption. Liu et al. [106] combined experiments and DFT calculations to systematically study the promoting effect of sodium on CO2 hydrogenation over Fe5C2 (Figure 8c). Through a comparative analysis between Na2O/Fe5C2 (510) and Fe5C2 (510) models, they elucidated that Na could significantly modify the electronic structure of the Fe5C2 (510) surface, leading to a reduction in the CO2 dissociation barrier from 0.45 eV to 0.08 eV. Furthermore, Na impedes further hydrogenation of CH2, enhances C-C coupling, promotes CH2 chain growth, lowers methane selectivity, and enhances olefin selectivity. Wang et al. [107] successfully designed a novel multifunctional catalyst integrating Na-doped Fe-based catalyst (Na-Fe@C) and K-doped CuZnAl (K-CuZnAl) for promoting the direct synthesis of ethanol and olefins through CO2 hydrogenation. Based on experimental results, they established two surface models, Fe5C2 (510) and Na2O/Fe5C2 (510), to estimate the C-C coupling barriers (Figure 8d). The computational results are consistent with the changes in ethanol selectivity obtained from multifunctional catalysts with different levels of Na doping. According to Partial Density of States (PDOS) analysis, the low adsorption energy of high-occupied intermediate species is attributable to anti-bonding states on the Na-doped Fe5C2 (510) surface.
Na binds to the metal oxide MgO, which acts as a catalyst for electron transfer and structural stability. Ahmed et al. [109] designed a dual-functional Na-FeMgOx catalyst and explored the promoting effect of MgO in Fe-based catalysts. Their study revealed that MgO facilitates electron transfer to the Fe-based phase, promoting the reduction of Fe oxides and the creation of oxygen vacancies. Furthermore, MgO enhances the adsorption of CO2 and H2, thereby facilitating the formation of long-chain hydrocarbons. However, as the reaction progresses, MgO may convert to inactive MgCO3, leading to catalyst deactivation and decreased selectivity towards C5+ products.

3.2.2. Transition Metals

Transition metals such as zinc (Zn), nickel (Ni), copper (Cu), and manganese (Mn) are commonly incorporated into iron-based catalysts to fine-tune the content and composition of active phases. This adjustment can lead to improved carbon chain elongation and the overall durability of iron-based catalysts.
Zn prevents the re-oxidation of iron carbide and carbon deposition to control surface coverage, ensuring that the phase maintains sufficient dispersion and stability. Xu et al. [110] reported a ternary iron-based catalyst (Fe-Zn-Al) in which Zn was doped into a binary Fe-Al spinel for the direct conversion of CO2 into linear α-olefins (Figure 9a). A combination of various in situ characterization methods and DFT calculations indicated that the introduction of Al led to the wrapping of active Fe5C2 nanoparticles within the Fe-Al spinel, facilitating hydrogenation reactions and inhibiting C-C coupling on the catalyst. The additional introduction of Zn allowed for the redistribution of Al, weakening the strong interactions between Fe5C2 and the spinel phase and resulting in higher selectivity towards olefins. In a subsequent study, Xu et al. [111] synthesized a series of FeAl catalysts (FeAl350, FeAl450, FeAl600, FeAl750, and FeAl900) at various calcination temperatures for CO2 hydrogenation. They analyzed the activity and selectivity of different iron species during the hydrogenation process. Combining experimental and DFT computational approaches, it was found that strong Fe-Al interactions hindered the reduction and carburization of Fe species, resulting in a higher FeOx content. The FeAl350 catalyst, with the highest FeCx content, exhibited the highest CO2 conversion rate (48.2%). Lower FeCx content led to decreased activity in CO2 hydrogenation to olefins, as observed in the FeAl900 catalyst (36.6%). Therefore, adjusting the composition of active phases offers a promising strategy for the rational design of catalysts for CO2 hydrogenation to olefins. The study by Liu et al. [21] investigates the role of Zn-promoted iron-based catalysts in the direct hydrogenation of CO2 to produce olefins. The research reveals that in the presence of Zn, the active phase Fe5C2 (510) can prevent further oxidation under reaction conditions and the deposition of carbon on the catalyst surface (Figure 9b), leading to enhanced long-term stability. Zn facilitates the adsorption of surface oxygen atoms and promotes the desorption of H2O and H2O during hydrogenation, reducing the likelihood of surface carbides being oxidized. The study speculates that in terms of stabilizing iron carbides, the effect of ZnO is more significant compared to standalone Zn.
The introduction of Ni impedes the adsorption of CO2 on Fe (111). Belelli et al. [112] examined the adsorption and dissociation of CO2 on the Fe (111) surface and three different Ni-Fe (111) alloy surfaces, focusing primarily on the most stable (MS) and most active (MA) modes (Figure 9c). Overall, the presence of nickel inhibits the adsorption of CO2 compared to the Fe (111) surface. The study proposed two different mechanisms for CO2 dissociation: one starting from the MS mode, directly producing CO and O through dissociation; the other involving a two-step reaction, first migrating from the MS mode to the MA mode, followed by dissociation in the MA mode. From a kinetic perspective, the pure Fe (111) surface is highly favorable for hydrogenation reactions, while the presence of nickel increases the hydrogenation barrier in all cases, leading to the formation of the HCOO* intermediate.
Cu modifiers demonstrate a comparable effect to Zn in promoting the reduction and carburation of iron-based catalysts, thereby favoring the formation of Fe3O4 and Fe5C2 phases. However, strong interactions between Cu and iron species impede CO adsorption, unfavorably influencing the formation of CO intermediates. Yang et al. [113] conducted a combination of various characterization techniques and DFT calculations to investigate the promoting effects of Zn, Cu, and Mn on Fe-based catalysts for the catalytic hydrogenation of CO2 to produce olefins. FeZn-Na exhibited excellent activity, displaying the highest overall selectivity towards olefins. Based on experimental characterization, models of the Fe5C2 (111), Cu/Fe5C2 (111), and ZnO/Fe5C2 (111) interfaces were constructed. In the transition from C3H6 through two hydrogenation steps to C3H8, a notable increase in the energy barrier in the second step was observed, suggesting that the Cu/Fe5C2 (111) (Figure 9d). However, calculations indicated that further adsorption and hydrogenation of olefins on ZnO/Fe5C2 (111) are unfavorable.

3.2.3. Synergistic Effect

Exploiting the inherent advantages of incorporating alkali metals to catalyze active phase formation and enhancing the structural robustness of iron-based catalysts through the addition of transition metals, the concurrent integration of alkali and transition metals represents a strategic approach to amplify catalytic efficacy in CO2 hydrogenation. Hwang et al. [114] developed a Fe-Cu-K catalyst for investigating the hydrogenation of CO2 (Figure 10). Experimental findings demonstrated that the yield of C5+ products was 18.1%, which is 1.4 times higher than that of the FeK catalyst (12.8%) and 7.8 times higher than that of the Fe-Cu catalyst (2.3%). Experimental characterization confirmed the incorporation of copper into the lattice of iron in the presence of potassium. The addition of K ensured the activity of the FeCu alloy phase during the reaction. Electron structure calculations revealed that the Fe-Cu surface promotes CO2 hydrogenation by lowering the energy required for deoxygenation. Regardless of the reaction pathway, the addition of K increased the binding energy of carbon and decreased the binding energy of hydrogen. The study speculated that surface modification might increase the surface carbon density, explaining the high selectivity of the K-promoted catalyst for the formation of long-chain hydrocarbons. Subsequently, Hwang et al. [115] synthesized a novel Fe-Co alloy catalyst derived from N-coordinated Co single-atom carbon (FeK/Co-NC) and applied this catalyst system to CO2 hydrogenation. Due to the Co-NC support, Co atoms were efficiently supplied to Fe nanoparticles, inducing the effective formation of Fe-Co alloy. DFT calculations indicated that Fe-Co mixed oxides accelerate oxygen removal in the RWGS reaction, while Fe-Co mixed carbides promote chain growth and suppress methane formation in FTS. This innovative catalyst design not only showcases the potential of Fe-Co alloys for catalytic applications but also highlights the synergy between different components in achieving enhanced catalytic performance in CO2 hydrogenation.

4. Conclusions and Outlook

In conclusion, this review provides a concise overview of recent theoretical advancements in iron-based catalysts for CO2 hydrogenation, employing DFT simulations and microkinetic models. Various types of iron-based catalysts, including metallic iron, iron oxides, iron carbides, and other derivatives, have systematically been synthesized to understand the mechanisms underlying CO2 hydrogenation. Notably, alkali metals have been shown to significantly enhance CO2 adsorption and stabilize active phases within iron-based catalysts. Additionally, the incorporation of doped transition metals contributes to the development of versatile catalysts with multifunctional capabilities. The catalytic functionalities of each metal species enable distinct transformations of target molecules. The synergistic integration of alkali metals and transition metals allows precise adjustments of key attributes governing the CO2 hydrogenation process. These findings highlight the efficacy of the bottom-up catalyst design strategy, facilitating the production of high-performance iron-based catalysts for CO2 hydrogenation.
Despite the remarkable achievements of computational methods in identifying active sites and elucidating structure–activity relationships in Fe-based catalysts for CO2 hydrogenation, their practical applicability has been confined to specific spatial and temporal scales for which they were developed. Artificial intelligence (AI) and machine learning (ML) present promising avenues for enhancing theoretical descriptions across various time and length scales. Integrating density functional theory (DFT)-based materials databases with ML approaches can facilitate the transformation of data into valuable insights, thereby accelerating the discovery of catalyst candidates and enriching our understanding of underlying catalytic mechanisms. Although in its early stages with many underexplored research opportunities, ML techniques have demonstrated successful applications in various solid heterogeneous catalysis in recent years [116,117,118]. With the ongoing increase in applied research, we anticipate that machine learning methods will emerge as pivotal computational tools for advancing the development of high-performance catalysts for CO2 hydrogenation in the near future.

Author Contributions

Conceptualization, Z.Z., C.Z. and X.W.; investigation, H.T.; data curation, H.T. and T.Q.; writing—original draft preparation, H.T. and T.Q.; writing—review and editing, Z.Z., C.Z. and X.W.; visualization, H.T.; supervision, Z.Z., C.Z. and X.W.; project administration, Z.Z.; funding acquisition, C.Z. and X.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (22178164 and 22378187), the Natural Science Foundation of Jiangsu Province (BZ2023051, BK20200694, 20KJB530002, and 21KJB480014), and the Jiangsu Specially-Appointed Professors Program.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in this article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Goeppert, A.; Czaun, M.; Jones, J.-P.; Surya Prakash, G.K.; Olah, G.A. Recycling of carbon dioxide to methanol and derived products–closing the loop. Chem. Soc. Rev. 2014, 43, 7995–8048. [Google Scholar] [CrossRef]
  2. Snoeckx, R.; Bogaerts, A. Plasma technology—A novel solution for CO2 conversion? Chem. Soc. Rev. 2017, 46, 5805–5863. [Google Scholar] [CrossRef]
  3. Nazir, G.; Rehman, A.; Hussain, S.; Mahmood, Q.; Fteiti, M.; Heo, K.; Ikram, M.; Aizaz Ud Din, M. Towards a sustainable conversion of biomass/biowaste to porous carbons for CO2 adsorption: Recent advances, current challenges, and future directions. Green Chem. 2023, 25, 4941–4980. [Google Scholar] [CrossRef]
  4. Kätelhön, A.; Meys, R.; Deutz, S.; Suh, S.; Bardow, A. Climate change mitigation potential of carbon capture and utilization in the chemical industry. Proc. Natl. Acad. Sci. USA 2019, 116, 11187–11194. [Google Scholar] [CrossRef]
  5. de Kleijne, K.; Hanssen, S.V.; van Dinteren, L.; Huijbregts, M.A.J.; van Zelm, R.; de Coninck, H. Limits to Paris compatibility of CO2 capture and utilization. One Earth 2022, 5, 168–185. [Google Scholar] [CrossRef]
  6. Choi, Y.H.; Jang, Y.J.; Park, H.; Kim, W.Y.; Lee, Y.H.; Choi, S.H.; Lee, J.S. Carbon dioxide Fischer-Tropsch synthesis: A new path to carbon-neutral fuels. Appl. Catal. B 2017, 202, 605–610. [Google Scholar] [CrossRef]
  7. Ra, E.C.; Kim, K.Y.; Kim, E.H.; Lee, H.; An, K.; Lee, J.S. Recycling Carbon Dioxide through Catalytic Hydrogenation: Recent Key Developments and Perspectives. ACS Catal. 2020, 10, 11318–11345. [Google Scholar] [CrossRef]
  8. Atsbha, T.A.; Yoon, T.; Seongho, P.; Lee, C.-J. A review on the catalytic conversion of CO2 using H2 for synthesis of CO, methanol, and hydrocarbons. J. CO2 Util. 2021, 44, 101413. [Google Scholar] [CrossRef]
  9. Hepburn, C.; Adlen, E.; Beddington, J.; Carter, E.A.; Fuss, S.; Mac Dowell, N.; Minx, J.C.; Smith, P.; Williams, C.K. The technological and economic prospects for CO2 utilization and removal. Nature 2019, 575, 87–97. [Google Scholar] [CrossRef] [PubMed]
  10. Krevor, S.; de Coninck, H.; Gasda, S.E.; Ghaleigh, N.S.; de Gooyert, V.; Hajibeygi, H.; Juanes, R.; Neufeld, J.; Roberts, J.J.; Swennenhuis, F. Subsurface carbon dioxide and hydrogen storage for a sustainable energy future. Nat. Rev. Earth Environ. 2023, 4, 102–118. [Google Scholar] [CrossRef]
  11. Kamkeng, A.D.N.; Wang, M.; Hu, J.; Du, W.; Qian, F. Transformation technologies for CO2 utilisation: Current status, challenges and future prospects. Chem. Eng. J. 2021, 409, 128138. [Google Scholar] [CrossRef]
  12. Pahija, E.; Panaritis, C.; Gusarov, S.; Shadbahr, J.; Bensebaa, F.; Patience, G.; Boffito, D.C. Experimental and Computational Synergistic Design of Cu and Fe Catalysts for the Reverse Water–Gas Shift: A Review. ACS Catal. 2022, 12, 6887–6905. [Google Scholar] [CrossRef]
  13. Amoyal, M.; Vidruk-Nehemya, R.; Landau, M.V.; Herskowitz, M. Effect of potassium on the active phases of Fe catalysts for carbon dioxide conversion to liquid fuels through hydrogenation. J. Catal. 2017, 348, 29–39. [Google Scholar] [CrossRef]
  14. Wang, X.; Yang, G.; Zhang, J.; Chen, S.; Wu, Y.; Zhang, Q.; Wang, J.; Han, Y.; Tan, Y. Synthesis of isoalkanes over a core (Fe–Zn–Zr)–shell (zeolite) catalyst by CO2 hydrogenation. Chem. Comm. 2016, 52, 7352–7355. [Google Scholar] [CrossRef] [PubMed]
  15. Lee, S.; Seo, J.-C.; Chun, H.-J.; Yang, S.; Sim, E.-h.; Lee, J.; Kim, Y.T. Selective olefin production on silica based iron catalysts in Fischer–Tropsch synthesis. Catal. Sci. Technol. 2022, 12, 5814–5828. [Google Scholar] [CrossRef]
  16. Samanta, A.; Landau, M.V.; Vidruk-Nehemya, R.; Herskowitz, M. CO2 hydrogenation to higher hydrocarbons on K/Fe–Al–O spinel catalysts promoted with Si, Ti, Zr, Hf, Mn and Ce. Catal. Sci. Technol. 2017, 7, 4048–4063. [Google Scholar] [CrossRef]
  17. 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]
  18. Theofanidis, S.A.; Kasun Kalhara Gunasooriya, G.T.; Itskou, I.; Tasioula, M.; Lemonidou, A.A. On-purpose Ethylene Production via CO2-assisted Ethane Oxidative Dehydrogenation: Selectivity Control of Iron Oxide Catalysts. ChemCatChem 2022, 14, e202200032. [Google Scholar] [CrossRef]
  19. Nie, X.; Meng, L.; Wang, H.; Chen, Y.; Guo, X.; Song, C. DFT insight into the effect of potassium on the adsorption, activation and dissociation of CO2 over Fe-based catalysts. Phys. Chem. Chem. Phys. 2018, 20, 14694–14707. [Google Scholar] [CrossRef]
  20. Claeys, M.; van Steen, E.; Botha, T.; Crous, R.; Ferreira, A.; Harilal, A.; Moodley, D.J.; Moodley, P.; du Plessis, E.; Visagie, J.L. Oxidation of Hägg Carbide during High-Temperature Fischer–Tropsch Synthesis: Size-Dependent Thermodynamics and In Situ Observations. ACS Catal. 2021, 11, 13866–13879. [Google Scholar] [CrossRef]
  21. Liu, X.; Xu, M.; Cao, C.; Yang, Z.; Xu, J. Effects of zinc on χ-Fe5C2 for carbon dioxide hydrogenation to olefins: Insights from experimental and density function theory calculations. Chin. J. Chem. Eng. 2023, 54, 206–214. [Google Scholar] [CrossRef]
  22. Saeidi, S.; Najari, S.; Hessel, V.; Wilson, K.; Keil, F.J.; Concepción, P.; Suib, S.L.; Rodrigues, A.E. Recent advances in CO2 hydrogenation to value-added products—Current challenges and future directions. Prog. Energy Combust. Sci. 2021, 85, 100905. [Google Scholar] [CrossRef]
  23. Liu, W.; Cheng, S.; Malhi, H.S.; Gao, X.; Zhang, Z.; Tu, W. Hydrogenation of CO2 to Olefins over Iron-Based Catalysts: A Review. Catalysts 2022, 12, 1432. [Google Scholar] [CrossRef]
  24. Xu, Y.; Zhai, P.; Deng, Y.; Xie, J.; Liu, X.; Wang, S.; Ma, D. Highly Selective Olefin Production from CO2 Hydrogenation on Iron Catalysts: A Subtle Synergy between Manganese and Sodium Additives. Angew. Chem. Int. Ed. 2020, 59, 21736–21744. [Google Scholar] [CrossRef]
  25. Han, Y.; Fang, C.; Ji, X.; Wei, J.; Ge, Q.; Sun, J. Interfacing with Carbonaceous Potassium Promoters Boosts Catalytic CO2 Hydrogenation of Iron. ACS Catal. 2020, 10, 12098–12108. [Google Scholar] [CrossRef]
  26. Gnanamani, M.K.; Hamdeh, H.H.; Shafer, W.D.; Hopps, S.D.; Davis, B.H. Hydrogenation of carbon dioxide over iron carbide prepared from alkali metal promoted iron oxalate. Appl. Catal. A 2018, 564, 243–249. [Google Scholar] [CrossRef]
  27. Chaipraditgul, N.; Numpilai, T.; Kui Cheng, C.; Siri-Nguan, N.; Sornchamni, T.; Wattanakit, C.; Limtrakul, J.; Witoon, T. Tuning interaction of surface-adsorbed species over Fe/K-Al2O3 modified with transition metals (Cu, Mn, V, Zn or Co) on light olefins production from CO2 hydrogenation. Fuel 2021, 283, 119248. [Google Scholar] [CrossRef]
  28. Kim, K.Y.; Lee, H.; Noh, W.Y.; Shin, J.; Han, S.J.; Kim, S.K.; An, K.; Lee, J.S. Cobalt Ferrite Nanoparticles to Form a Catalytic Co–Fe Alloy Carbide Phase for Selective CO2 Hydrogenation to Light Olefins. ACS Catal. 2020, 10, 8660–8671. [Google Scholar] [CrossRef]
  29. Barrios, A.J.; Peron, D.V.; Chakkingal, A.; Dugulan, A.I.; Moldovan, S.; Nakouri, K.; Thuriot-Roukos, J.; Wojcieszak, R.; Thybaut, J.W.; Virginie, M.; et al. Efficient Promoters and Reaction Paths in the CO2 Hydrogenation to Light Olefins over Zirconia-Supported Iron Catalysts. ACS Catal. 2022, 12, 3211–3225. [Google Scholar] [CrossRef]
  30. Karre, A.V.; Dadyburjor, D.B. Review of iron-based catalysts with and without zeolite supports used in fischer-tropsch processes. Chem. Eng. Commun. 2022, 209, 967–987. [Google Scholar] [CrossRef]
  31. Yan, P.; Peng, H.; Vogrin, J.; Rabiee, H.; Zhu, Z. Selective CO2 hydrogenation over zeolite-based catalysts for targeted high-value products. J. Mater. Chem. A 2023, 11, 17938–17960. [Google Scholar] [CrossRef]
  32. Han, X.; Xia, H.; Tu, W.; Wei, Y.; Xue, D.; Li, M.; Yan, W.; Zhang, J.-N.; Han, Y.-F. Zeolite-confined Fe-site Catalysts for the Hydrogenation of CO2 to Produce High-value Chemicals. Chem. Res. Chin. Univ. 2024, 40, 78–95. [Google Scholar] [CrossRef]
  33. Hoffmann, R. Small but strong lessons from chemistry for nanoscience. Angew. Chem. Int. Ed. 2013, 52, 93–103. [Google Scholar] [CrossRef] [PubMed]
  34. Horton, M.K.; Montoya, J.H.; Liu, M.; Persson, K.A. High-throughput prediction of the ground-state collinear magnetic order of inorganic materials using Density Functional Theory. npj Comput. Mater. 2019, 5, 64. [Google Scholar] [CrossRef]
  35. Xu, L.; Papanikolaou, K.G.; Lechner, B.A.J.; Je, L.; Somorjai, G.A.; Salmeron, M.; Mavrikakis, M. Formation of active sites on transition metals through reaction-driven migration of surface atoms. Science 2023, 380, 70–76. [Google Scholar] [CrossRef]
  36. Cheng, D.; Negreiros, F.R.; Aprà, E.; Fortunelli, A. Computational Approaches to the Chemical Conversion of Carbon Dioxide. ChemSusChem 2013, 6, 944–965. [Google Scholar] [CrossRef]
  37. Nezam, I.; Zhou, W.; Gusmão, G.S.; Realff, M.J.; Wang, Y.; Medford, A.J.; Jones, C.W. Direct aromatization of CO2 via combined CO2 hydrogenation and zeolite-based acid catalysis. J. CO2 Util. 2021, 45, 101405. [Google Scholar] [CrossRef]
  38. Choi, S.; Sang, B.-I.; Hong, J.; Yoon, K.J.; Son, J.-W.; Lee, J.-H.; Kim, B.-K.; Kim, H. Catalytic behavior of metal catalysts in high-temperature RWGS reaction: In-situ FT-IR experiments and first-principles calculations. Sci. Rep. 2017, 7, 41207. [Google Scholar] [CrossRef]
  39. Yang, H.; Zhang, C.; Gao, P.; Wang, H.; Li, X.; Zhong, L.; Wei, W.; Sun, Y. A review of the catalytic hydrogenation of carbon dioxide into value-added hydrocarbons. Catal. Sci. Technol. 2017, 7, 4580–4598. [Google Scholar] [CrossRef]
  40. Weijing, D.; Weihong, Z.; Xiaodong, Z.; Baofeng, Z.; Lei, C.; Laizhi, S.; Shuangxia, Y.; Haibin, G.; Guanyi, C.; Liang, Z.; et al. The application of DFT in catalysis and adsorption reaction system. Energy Procedia 2018, 152, 997–1002. [Google Scholar] [CrossRef]
  41. Cui, W.-G.; Zhang, G.-Y.; Hu, T.-L.; Bu, X.-H. Metal-organic framework-based heterogeneous catalysts for the conversion of C1 chemistry: CO, CO2 and CH4. Coord. Chem. Rev. 2019, 387, 79–120. [Google Scholar] [CrossRef]
  42. Podrojková, N.; Sans, V.; Oriňak, A.; Oriňaková, R. Recent Developments in the Modelling of Heterogeneous Catalysts for CO2 Conversion to Chemicals. ChemCatChem 2020, 12, 1802–1825. [Google Scholar] [CrossRef]
  43. Zheng, Z.; Zhang, O.; Borgs, C.; Chayes, J.T.; Yaghi, O.M. ChatGPT Chemistry Assistant for Text Mining and the Prediction of MOF Synthesis. J. Am. Chem. Soc. 2023, 145, 18048–18062. [Google Scholar] [CrossRef]
  44. Rommens, K.T.; Saeys, M. Molecular Views on Fischer–Tropsch Synthesis. Chem. Rev. 2023, 123, 5798–5858. [Google Scholar] [CrossRef]
  45. González-Castaño, M.; Dorneanu, B.; Arellano-García, H. The reverse water gas shift reaction: A process systems engineering perspective. React. Chem. Eng. 2021, 6, 954–976. [Google Scholar] [CrossRef]
  46. Wei, J.; Yao, R.; Han, Y.; Ge, Q.; Sun, J. Towards the development of the emerging process of CO2 heterogenous hydrogenation into high-value unsaturated heavy hydrocarbons. Chem. Soc. Rev. 2021, 50, 10764–10805. [Google Scholar] [CrossRef] [PubMed]
  47. 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]
  48. Yao, B.; Xiao, T.; Makgae, O.A.; Jie, X.; Gonzalez-Cortes, S.; Guan, S.; Kirkland, A.I.; Dilworth, J.R.; Al-Megren, H.A.; Alshihri, S.M.; et al. Transforming carbon dioxide into jet fuel using an organic combustion-synthesized Fe-Mn-K catalyst. Nat. Commun. 2020, 11, 6395. [Google Scholar] [CrossRef]
  49. Riedel, T.; Claeys, M.; Schulz, H.; Schaub, G.; Nam, S.-S.; Jun, K.-W.; Choi, M.-J.; Kishan, G.; Lee, K.-W. Comparative study of Fischer–Tropsch synthesis with H2/CO and H2/CO2 syngas using Fe- and Co-based catalysts. Appl. Catal. A 1999, 186, 201–213. [Google Scholar] [CrossRef]
  50. He, Z.; Cui, M.; Qian, Q.; Zhang, J.; Liu, H.; Han, B. Synthesis of liquid fuel via direct hydrogenation of CO2. Proc. Natl. Acad. Sci. USA 2019, 116, 12654–12659. [Google Scholar] [CrossRef]
  51. Ye, R.-P.; Ding, J.; Gong, W.; Argyle, M.D.; Zhong, Q.; Wang, Y.; Russell, C.K.; Xu, Z.; Russell, A.G.; Li, Q.; et al. CO2 hydrogenation to high-value products via heterogeneous catalysis. Nat. Commun. 2019, 10, 5698. [Google Scholar] [CrossRef]
  52. Han, S.J.; Hwang, S.-M.; Park, H.-G.; Zhang, C.; Jun, K.-W.; Kim, S.K. Identification of active sites for CO2 hydrogenation in Fe catalysts by first-principles microkinetic modelling. J. Mater. Chem. A 2020, 8, 13014–13023. [Google Scholar] [CrossRef]
  53. Zhang, Z.; Chen, B.; Jia, L.; Liu, W.; Gao, X.; Gao, J.; Meng, B.; Tan, Y.; He, Y.; Tu, W.; et al. Unraveling the role of Fe5C2 in CH4 formation during CO2 hydrogenation over hydrophobic iron catalysts. Appl. Catal. B 2023, 327, 122449. [Google Scholar] [CrossRef]
  54. Huš, M.; Grilc, M.; Pavlišič, A.; Likozar, B.; Hellman, A. Multiscale modelling from quantum level to reactor scale: An example of ethylene epoxidation on silver catalysts. Catal. Today 2019, 338, 128–140. [Google Scholar] [CrossRef]
  55. Li, B.; Shao, Z.-G.; Feng, Y.-T. First-principles investigation of CO and CO2 adsorption on pristine and Fe-doped planar carbon allotrope net-Y. Phys. Chem. Chem. Phys. 2021, 23, 12771–12779. [Google Scholar] [CrossRef]
  56. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed]
  57. Dudarev, S.L.; Botton, G.A.; Savrasov, S.Y.; Humphreys, C.J.; Sutton, A.P. Electron-energy-loss spectra and the structural stability of nickel oxide: An LSDA+U study. Phys. Rev. B 1998, 57, 1505–1509. [Google Scholar] [CrossRef]
  58. Liu, W.; Carrasco, J.; Santra, B.; Michaelides, A.; Scheffler, M.; Tkatchenko, A. Benzene adsorbed on metals: Concerted effect of covalency and van der Waals bonding. Phys. Rev. B 2012, 86, 245405. [Google Scholar] [CrossRef]
  59. Grimme, S. Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 2006, 27, 1787–1799. [Google Scholar] [CrossRef]
  60. Suvarna, M.; Araújo, T.P.; Pérez-Ramírez, J. A generalized machine learning framework to predict the space-time yield of methanol from thermocatalytic CO2 hydrogenation. Appl. Catal. B 2022, 315, 121530. [Google Scholar] [CrossRef]
  61. Al-Mahayni, H.; Wang, X.; Harvey, J.-P.; Patience, G.S.; Seifitokaldani, A. Experimental methods in chemical engineering: Density functional theory. Can. J. Chem. Eng. 2021, 99, 1885–1911. [Google Scholar] [CrossRef]
  62. Naimatullah; Li, D.; Gahungu, G.; Li, W.; Zhang, J. First-principles calculations on CO2 hydrogenation to formic acid over a metal-doped boron phosphide. Mol. Catal. 2022, 527, 112412. [Google Scholar] [CrossRef]
  63. Bhandari, S.; Rangarajan, S.; Mavrikakis, M. Combining Computational Modeling with Reaction Kinetics Experiments for Elucidating the In Situ Nature of the Active Site in Catalysis. Acc. Chem. Res. 2020, 53, 1893–1904. [Google Scholar] [CrossRef]
  64. Chen, B.W.J.; Xu, L.; Mavrikakis, M. Computational Methods in Heterogeneous Catalysis. Chem. Rev. 2021, 121, 1007–1048. [Google Scholar] [CrossRef]
  65. Wang, Y.; Su, Y.-Q.; Hensen, E.J.M.; Vlachos, D.G. Insights into Supported Subnanometer Catalysts Exposed to CO via Machine-Learning-Enabled Multiscale Modeling. Chem. Mater. 2022, 34, 1611–1619. [Google Scholar] [CrossRef]
  66. Clay, J.P.; Greeley, J.P.; Ribeiro, F.H.; Nicholas Delgass, W.; Schneider, W.F. DFT comparison of intrinsic WGS kinetics over Pd and Pt. J. Catal. 2014, 320, 106–117. [Google Scholar] [CrossRef]
  67. Dietz, L.; Piccinin, S.; Maestri, M. Mechanistic Insights into CO2 Activation via Reverse Water–Gas Shift on Metal Surfaces. J. Phys. Chem. C 2015, 119, 4959–4966. [Google Scholar] [CrossRef]
  68. Huynh, H.L.; Zhu, J.; Zhang, G.; Shen, Y.; Tucho, W.M.; Ding, Y.; Yu, Z. Promoting effect of Fe on supported Ni catalysts in CO2 methanation by in situ DRIFTS and DFT study. J. Catal. 2020, 392, 266–277. [Google Scholar] [CrossRef]
  69. Serrer, M.-A.; Gaur, A.; Jelic, J.; Weber, S.; Fritsch, C.; Clark, A.H.; Saraçi, E.; Studt, F.; Grunwaldt, J.-D. Structural dynamics in Ni–Fe catalysts during CO2 methanation—Role of iron oxide clusters. Catal. Sci. Technol. 2020, 10, 7542–7554. [Google Scholar] [CrossRef]
  70. Krishnan, R.; Yang, K.; Hashem, K.; Jiang, J. Metallated porphyrinic metal–organic frameworks for CO2 conversion to HCOOH: A computational screening and mechanistic study. Mol. Catal. 2022, 527, 112407. [Google Scholar] [CrossRef]
  71. Khodakov, A.Y.; Chu, W.; Fongarland, P. Advances in the Development of Novel Cobalt Fischer–Tropsch Catalysts for Synthesis of Long-Chain Hydrocarbons and Clean Fuels. Chem. Rev. 2007, 107, 1692–1744. [Google Scholar] [CrossRef]
  72. de Smit, E.; Weckhuysen, B.M. The renaissance of iron-based Fischer–Tropsch synthesis: On the multifaceted catalyst deactivation behaviour. Chem. Soc. Rev. 2008, 37, 2758–2781. [Google Scholar] [CrossRef]
  73. Li, X.; Yang, Z.; Zhang, L.; He, Z.; Fang, R.; Wang, Z.; Yan, Y.; Ran, J. Effect of Pd doping in (Fe/Ni)/CeO2 catalyst for the reaction path in CO2 oxidative ethane dehydrogenation/reforming. Energy 2021, 234, 121261. [Google Scholar] [CrossRef]
  74. Liu, C.; Cundari, T.R.; Wilson, A.K. CO2 Reduction on Transition Metal (Fe, Co, Ni, and Cu) Surfaces: In Comparison with Homogeneous Catalysis. J. Phys. Chem. C 2012, 116, 5681–5688. [Google Scholar] [CrossRef]
  75. Liu, S.; Li, Y.-W.; Wang, J.; Jiao, H. Reactions of CO, H2O, CO2, and H2 on the Clean and Precovered Fe(110) Surfaces—A DFT Investigation. J. Phys. Chem. C 2015, 119, 28377–28388. [Google Scholar] [CrossRef]
  76. Liu, X.; Sun, L.; Deng, W.-Q. Theoretical Investigation of CO2 Adsorption and Dissociation on Low Index Surfaces of Transition Metals. J. Phys. Chem. C 2018, 122, 8306–8314. [Google Scholar] [CrossRef]
  77. Wang, H.; Nie, X.; Chen, Y.; Guo, X.; Song, C. Facet effect on CO2 adsorption, dissociation and hydrogenation over Fe catalysts: Insight from DFT. J. CO2 Util. 2018, 26, 160–170. [Google Scholar] [CrossRef]
  78. Chen, H.; Yang, M.; Liu, J.; Lu, G.; Feng, X. Insight into the effects of electronegativity on the H2 catalytic activation for CO2 hydrogenation: Four transition metal cases from a DFT study. Catal. Sci. Technol. 2020, 10, 5641–5647. [Google Scholar] [CrossRef]
  79. Wang, H.; Nie, X.; Guo, X.; Song, C. A computational study of adsorption and activation of CO2 and H2 over Fe(100) surface. J. CO2 Util. 2016, 15, 107–114. [Google Scholar] [CrossRef]
  80. Yang, L.; Pastor-Pérez, L.; Villora-Pico, J.J.; Gu, S.; Sepúlveda-Escribano, A.; Reina, T.R. CO2 valorisation via reverse water-gas shift reaction using promoted Fe/CeO2-Al2O3 catalysts: Showcasing the potential of advanced catalysts to explore new processes design. Appl. Catal. A 2020, 593, 117442. [Google Scholar] [CrossRef]
  81. 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. 2015, 23, 67–71. [Google Scholar] [CrossRef]
  82. Mirabella, F.; Zaki, E.; Ivars-Barcelo, F.; Schauermann, S.; Shaikhutdinov, S.; Freund, H.J. CO2 Adsorption on Magnetite Fe3O4(111). J. Phys. Chem. C 2018, 122, 27433–27441. [Google Scholar] [CrossRef]
  83. Parkinson, G.S. Iron oxide surfaces. Surf. Sci. Rep. 2016, 71, 272–365. [Google Scholar] [CrossRef]
  84. Chou, C.-Y.; Loiland, J.A.; Lobo, R.F. Reverse Water-Gas Shift Iron Catalyst Derived from Magnetite. Catalysts 2019, 9, 773. [Google Scholar] [CrossRef]
  85. 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]
  86. Li, X.; Paier, J. Vibrational properties of CO2 adsorbed on the Fe3O4 (111) surface: Insights gained from DFT. J. Chem. Phys. 2020, 152, 104702. [Google Scholar] [CrossRef]
  87. Yang, T.; Wen, X.-D.; Huo, C.-F.; Li, Y.-W.; Wang, J.; Jiao, H. Structure and energetics of hydrogen adsorption on Fe3O4(111). J. Mol. Catal. A Chem. 2009, 302, 129–136. [Google Scholar] [CrossRef]
  88. Yu, X.; Huo, C.-F.; Li, Y.-W.; Wang, J.; Jiao, H. Fe3O4 surface electronic structures and stability from GGA+U. Surf. Sci. 2012, 606, 872–879. [Google Scholar] [CrossRef]
  89. Santos-Carballal, D.; Roldan, A.; Grau-Crespo, R.; de Leeuw, N.H. A DFT study of the structures, stabilities and redox behaviour of the major surfaces of magnetite Fe3O4. Phys. Chem. Chem. Phys. 2014, 16, 21082–21097. [Google Scholar] [CrossRef]
  90. Su, T.; Qin, Z.; Huang, G.; Ji, H.; Jiang, Y.; Chen, J. Density functional theory study on the interaction of CO2 with Fe3O4(111) surface. Appl. Surf. Sci. 2016, 378, 270–276. [Google Scholar] [CrossRef]
  91. Stegelmann, C.; Andreasen, A.; Campbell, C.T. Degree of Rate Control: How Much the Energies of Intermediates and Transition States Control Rates. J. Am. Chem. Soc. 2009, 131, 8077–8082. [Google Scholar] [CrossRef]
  92. Campbell, C.T. The Degree of Rate Control: A Powerful Tool for Catalysis Research. ACS Catal. 2017, 7, 2770–2779. [Google Scholar] [CrossRef]
  93. Campbell, C.T.; Mao, Z. Analysis and prediction of reaction kinetics using the degree of rate control. J. Catal. 2021, 404, 647–660. [Google Scholar] [CrossRef]
  94. de Smit, E.; Cinquini, F.; Beale, A.M.; Safonova, O.V.; van Beek, W.; Sautet, P.; Weckhuysen, B.M. Stability and Reactivity of ϵ–χ–θ Iron Carbide Catalyst Phases in Fischer–Tropsch Synthesis: Controlling μC. J. Am. Chem. Soc. 2010, 132, 14928–14941. [Google Scholar] [CrossRef]
  95. Lyu, S.; Wang, L.; Li, Z.; Yin, S.; Chen, J.; Zhang, Y.; Li, J.; Wang, Y. Stabilization of ε-iron carbide as high-temperature catalyst under realistic Fischer–Tropsch synthesis conditions. Nat. Commun. 2020, 11, 6219. [Google Scholar] [CrossRef]
  96. Zhao, H.; Liu, J.-X.; Yang, C.; Yao, S.; Su, H.-Y.; Gao, Z.; Dong, M.; Wang, J.; Rykov Alexandre, I.; Wang, J.; et al. Synthesis of Iron-Carbide Nanoparticles: Identification of the Active Phase and Mechanism of Fe-Based Fischer–Tropsch Synthesis. CCS Chem. 2020, 3, 2712–2724. [Google Scholar] [CrossRef]
  97. Liu, X.; Cao, C.; Tian, P.; Zhu, M.; Zhang, Y.; Xu, J.; Tian, Y.; Han, Y.-F. Resolving CO2 activation and hydrogenation pathways over iron carbides from DFT investigation. J. CO2 Util. 2020, 38, 10–15. [Google Scholar] [CrossRef]
  98. Huang, L.; Han, B.; Zhang, Q.; Fan, M.; Cheng, H. Mechanistic Study on Water Gas Shift Reaction on the Fe3O4 (111) Reconstructed Surface. J. Phys. Chem. C 2015, 119, 28934–28945. [Google Scholar] [CrossRef]
  99. Zhang, M.; Ren, J.; Yu, Y. Insights into the Hydrogen Coverage Effect and the Mechanism of Fischer–Tropsch to Olefins Process on Fe5C2 (510). ACS Catal. 2020, 10, 689–701. [Google Scholar] [CrossRef]
  100. Wang, H.; Nie, X.; Liu, Y.; Janik, M.J.; Han, X.; Deng, Y.; Hu, W.; Song, C.; Guo, X. Mechanistic Insight into Hydrocarbon Synthesis via CO2 Hydrogenation on χ-Fe5C2 Catalysts. ACS Appl. Mater. Interfaces 2022, 14, 37637–37651. [Google Scholar] [CrossRef] [PubMed]
  101. Gao, R.; Cao, D.-B.; Liu, S.; Yang, Y.; Li, Y.-W.; Wang, J.; Jiao, H. Density functional theory study into H2O dissociative adsorption on the Fe5C2(010) surface. Appl. Catal. A 2013, 468, 370–383. [Google Scholar] [CrossRef]
  102. Raje, A.P.; O’Brien, R.J.; Davis, B.H. Effect of Potassium Promotion on Iron-Based Catalysts for Fischer–Tropsch Synthesis. J. Catal. 1998, 180, 36–43. [Google Scholar] [CrossRef]
  103. Ma, Y.-P.; Wang, G.-C. Comparative theoretical study of CO2 activation on clean and potassium-preadsorbed low index surfaces of transition metals. J. Mol. Model. 2023, 29, 375. [Google Scholar] [CrossRef] [PubMed]
  104. Huo, C.-F.; Wu, B.-S.; Gao, P.; Yang, Y.; Li, Y.-W.; Jiao, H. The Mechanism of Potassium Promoter: Enhancing the Stability of Active Surfaces. Angew. Chem. Int. Ed. 2011, 50, 7403–7406. [Google Scholar] [CrossRef] [PubMed]
  105. Chen, H.; Ma, N.; Wang, C.; Liu, C.; Shen, J.; Wang, Y.; Xu, G.; Yang, Q.; Feng, X. Insight into the activation of CO2 and H2 on K2O-adsorbed Fe5C2(110) for olefins production: A density functional theory study. Mol. Catal. 2022, 524, 112323. [Google Scholar] [CrossRef]
  106. Liu, X.; Zhang, C.; Tian, P.; Xu, M.; Cao, C.; Yang, Z.; Zhu, M.; Xu, J. Revealing the Effect of Sodium on Iron-Based Catalysts for CO2 Hydrogenation: Insights from Calculation and Experiment. J. Phys. Chem. C 2021, 125, 7637–7646. [Google Scholar] [CrossRef]
  107. Wang, Y.; Wang, K.; Zhang, B.; Peng, X.; Gao, X.; Yang, G.; Hu, H.; Wu, M.; Tsubaki, N. Direct Conversion of CO2 to Ethanol Boosted by Intimacy-Sensitive Multifunctional Catalysts. ACS Catal. 2021, 11, 11742–11753. [Google Scholar] [CrossRef]
  108. Mahyuddin, M.H.; Belosludov, R.V.; Khazaei, M.; Mizuseki, H.; Kawazoe, Y. Effects of Alkali Adatoms on CO and H2S Adsorptions on the Fe(100) Surface: A Density Functional Theory Study. J. Phys. Chem. C 2011, 115, 23893–23901. [Google Scholar] [CrossRef]
  109. Ahmed, S.; Irshad, M.; Yoon, W.; Karanwal, N.; Sugiarto, J.R.; Khan, M.K.; Kim, S.K.; Kim, J. Evaluation of MgO as a promoter for the hydrogenation of CO2 to long-chain hydrocarbons over Fe-based catalysts. Appl. Catal. B 2023, 338, 123052. [Google Scholar] [CrossRef]
  110. Xu, M.; Liu, X.; Cao, C.; Sun, Y.; Zhang, C.; Yang, Z.; Zhu, M.; Ding, X.; Liu, Y.; Tong, Z.; et al. Ternary Fe–Zn–Al Spinel Catalyst for CO2 Hydrogenation to Linear α-Olefins: Synergy Effects between Al and Zn. ACS Sustain. Chem. Eng. 2021, 9, 13818–13830. [Google Scholar] [CrossRef]
  111. Xu, M.; Liu, X.; Song, G.; Cai, Y.; Shi, B.; Liu, Y.; Ding, X.; Yang, Z.; Tian, P.; Cao, C.; et al. Regulating iron species compositions by Fe-Al interaction in CO2 hydrogenation. J. Catal. 2022, 413, 331–341. [Google Scholar] [CrossRef]
  112. Belelli, P.G.; Rossi-Fernández, A.C.; Ferullo, R.M. CO2 dissociation and hydrogenation on pure and Ni-doped Fe(111). A DFT theoretical approach. Appl. Surf. Sci. 2023, 617, 156569. [Google Scholar] [CrossRef]
  113. Yang, H.; Dang, Y.; Cui, X.; Bu, X.; Li, J.; Li, S.; Sun, Y.; Gao, P. Selective synthesis of olefins via CO2 hydrogenation over transition-metal-doped iron-based catalysts. Appl. Catal. B 2023, 321, 122050. [Google Scholar] [CrossRef]
  114. Hwang, S.-M.; Han, S.J.; Min, J.E.; Park, H.-G.; Jun, K.-W.; Kim, S.K. Mechanistic insights into Cu and K promoted Fe-catalyzed production of liquid hydrocarbons via CO2 hydrogenation. J. CO2 Util. 2019, 34, 522–532. [Google Scholar] [CrossRef]
  115. Hwang, S.-M.; Han, S.J.; Park, H.-G.; Lee, H.; An, K.; Jun, K.-W.; Kim, S.K. Atomically Alloyed Fe–Co Catalyst Derived from a N-Coordinated Co Single-Atom Structure for CO2 Hydrogenation. ACS Catal. 2021, 11, 2267–2278. [Google Scholar] [CrossRef]
  116. Schlexer Lamoureux, P.; Winther, K.T.; Garrido Torres, J.A.; Streibel, V.; Zhao, M.; Bajdich, M.; Abild-Pedersen, F.; Bligaard, T. Machine Learning for Computational Heterogeneous Catalysis. ChemCatChem 2019, 11, 3581–3601. [Google Scholar] [CrossRef]
  117. Yang, W.; Fidelis, T.T.; Sun, W.-H. Machine Learning in Catalysis, From Proposal to Practicing. ACS Omega 2020, 5, 83–88. [Google Scholar] [CrossRef] [PubMed]
  118. Guan, Y.; Chaffart, D.; Liu, G.; Tan, Z.; Zhang, D.; Wang, Y.; Li, J.; Ricardez-Sandoval, L. Machine learning in solid heterogeneous catalysis: Recent developments, challenges and perspectives. Chem. Eng. Sci. 2022, 248, 117224. [Google Scholar] [CrossRef]
Figure 1. (a) Reaction scheme for CO2 hydrogenation to jet fuel range hydrocarbons. The CO2 hydrogenation to jet fuel range hydrocarbons process through a Tandem Mechanism in which the reverse water–gas shift reaction (RWGS) and Fischer–Tropsch synthesis (FTS) reaction are catalysed by Fe3O4 and χ-Fe5C2, respectively. Reprinted with permission from Yao et al. [48]. Copyright 2020 Springer Nature. (b) Scheme of CO2 modified FTS-based catalytic mechanism. Reprinted with permission from Ye et al. [51]. Copyright 2019 Springer Nature.
Figure 1. (a) Reaction scheme for CO2 hydrogenation to jet fuel range hydrocarbons. The CO2 hydrogenation to jet fuel range hydrocarbons process through a Tandem Mechanism in which the reverse water–gas shift reaction (RWGS) and Fischer–Tropsch synthesis (FTS) reaction are catalysed by Fe3O4 and χ-Fe5C2, respectively. Reprinted with permission from Yao et al. [48]. Copyright 2020 Springer Nature. (b) Scheme of CO2 modified FTS-based catalytic mechanism. Reprinted with permission from Ye et al. [51]. Copyright 2019 Springer Nature.
Molecules 29 01194 g001
Figure 2. Proposed algorithmic scheme for elucidating the nature of the catalytic active site and the reaction mechanism, using a combination of DFT, reaction kinetics experiments, and microkinetic modeling. Reprinted with permission from Bhandari et al. [63]. Copyright 2020 American Chemical Society.
Figure 2. Proposed algorithmic scheme for elucidating the nature of the catalytic active site and the reaction mechanism, using a combination of DFT, reaction kinetics experiments, and microkinetic modeling. Reprinted with permission from Bhandari et al. [63]. Copyright 2020 American Chemical Society.
Molecules 29 01194 g002
Figure 3. (a) The most stable calculated structures of (CO + O)/M(100) (M = Fe, Co, Ni, Cu). Reprinted with permission from Liu et al. [74]. Copyright 2012 American Chemical Society. (b) Adsorption of CO, CO2, HCO, COH, COOH, and HCOO on Fe (110). Adapted with permission from Liu et al. [75]. Copyright 2015 American Chemical Society. (c) CO2 adsorption and dissociation on low index surfaces of different transition metals. Reprinted with permission from Liu et al. [76]. Copyright 2018 American Chemical Society. (d) CO2 adsorption energy, dissociation and hydrogenation barriers on the various Fe surfaces. Reprinted with permission from Wang et al. [77]. Copyright 2018 Elsevier.
Figure 3. (a) The most stable calculated structures of (CO + O)/M(100) (M = Fe, Co, Ni, Cu). Reprinted with permission from Liu et al. [74]. Copyright 2012 American Chemical Society. (b) Adsorption of CO, CO2, HCO, COH, COOH, and HCOO on Fe (110). Adapted with permission from Liu et al. [75]. Copyright 2015 American Chemical Society. (c) CO2 adsorption and dissociation on low index surfaces of different transition metals. Reprinted with permission from Liu et al. [76]. Copyright 2018 American Chemical Society. (d) CO2 adsorption energy, dissociation and hydrogenation barriers on the various Fe surfaces. Reprinted with permission from Wang et al. [77]. Copyright 2018 Elsevier.
Molecules 29 01194 g003
Figure 5. (a) Schematic models of iron-based catalysts for Fischer–Tropsch synthesis. Conventional catalysts with unconfined iron carbide (FexC) particles as the active phase. Graphene layer-confined ε-Fe2C. Reprinted with permission from Lyu et al. [95]. Copyright 2020 Springer Nature. (b) The potential energy diagrams for CO dissociation on Fe5C2 (100) (red) surface. The solid and dashed lines present direct and H-assisted CO activation pathways, respectively. The apparent activation barriers (in eV) are indicated. Reprinted with permission from Zhao et al. [96]. Copyright 2020 Chinese Chemical Society. (c) CO2 adsorption and activation energies on metallic iron, magnetite and iron carbides surfaces. Reprinted with permission from Liu et al. [97]. Copyright 2020 Elsevier.
Figure 5. (a) Schematic models of iron-based catalysts for Fischer–Tropsch synthesis. Conventional catalysts with unconfined iron carbide (FexC) particles as the active phase. Graphene layer-confined ε-Fe2C. Reprinted with permission from Lyu et al. [95]. Copyright 2020 Springer Nature. (b) The potential energy diagrams for CO dissociation on Fe5C2 (100) (red) surface. The solid and dashed lines present direct and H-assisted CO activation pathways, respectively. The apparent activation barriers (in eV) are indicated. Reprinted with permission from Zhao et al. [96]. Copyright 2020 Chinese Chemical Society. (c) CO2 adsorption and activation energies on metallic iron, magnetite and iron carbides surfaces. Reprinted with permission from Liu et al. [97]. Copyright 2020 Elsevier.
Molecules 29 01194 g005
Figure 6. (a) Geometric structures of χ-Fe5C2 and θ-Fe3C unit cells. (b) Side and top views of χ-Fe5C2 (510) and θ-Fe3C (031) surfaces. (Fe atoms in purple, C atoms in gray). (c) Projected density of states (PDOS) of the D-orbitals of the surface Fe atoms for χ-Fe5C2 (510) and θ-Fe3C (031) surfaces. The vertical black dashed lines represent the Fermi level and the vertical blue lines indicate the D-band center. Reprinted with permission from Liu et al. [97]. Copyright 2020 Elsevier. (Fe atoms in purple, C atoms in gray).
Figure 6. (a) Geometric structures of χ-Fe5C2 and θ-Fe3C unit cells. (b) Side and top views of χ-Fe5C2 (510) and θ-Fe3C (031) surfaces. (Fe atoms in purple, C atoms in gray). (c) Projected density of states (PDOS) of the D-orbitals of the surface Fe atoms for χ-Fe5C2 (510) and θ-Fe3C (031) surfaces. The vertical black dashed lines represent the Fermi level and the vertical blue lines indicate the D-band center. Reprinted with permission from Liu et al. [97]. Copyright 2020 Elsevier. (Fe atoms in purple, C atoms in gray).
Molecules 29 01194 g006
Figure 7. Energy profiles of the three possible pathways for CO2 initial conversions over the χ-Fe5C2(510) surface. (Purple: Fe; black: C of adsorbates; red: O; white: H; and gray: C of χ-Fe5C2 catalyst). Reprinted with permission from Wang et al. [100]. Copyright 2022 American Chemical Society.
Figure 7. Energy profiles of the three possible pathways for CO2 initial conversions over the χ-Fe5C2(510) surface. (Purple: Fe; black: C of adsorbates; red: O; white: H; and gray: C of χ-Fe5C2 catalyst). Reprinted with permission from Wang et al. [100]. Copyright 2022 American Chemical Society.
Molecules 29 01194 g007
Figure 8. (a) CO2 adsorption energies on clean and K-preadsorbed transition metals. Reprinted with permission Ma et al. [103]. Copyright 2023 Springer Nature. (b) Supercells of clean and K2O-adsorbed Fe5C2 (110). Reprinted with permission Chen et al. [105]. Copyright 2022 Elsevier. (c) Electron distribution diagrams of surface atoms on Fe5C2 (510) and Na2O/Fe5C2 (510) (Fe atoms in purple, C atoms in gray, O atoms in red, and Na atoms in amaranth). Reprinted with permission Liu et al. [106]. Copyright 2021 American Chemical Society. (d) Energy profiles of the C–C coupling between CH2* and CO* species (CH2* +CO* → * + CH2CO*) on Fe5C2 (510) and Na2O/Fe5C2 (510). Na, Fe, C, O, and H are shown in purple, blue, gray, red, and white, respectively. Reprinted with permission Wang et al. [107]. Copyright 2021 American Chemical Society.
Figure 8. (a) CO2 adsorption energies on clean and K-preadsorbed transition metals. Reprinted with permission Ma et al. [103]. Copyright 2023 Springer Nature. (b) Supercells of clean and K2O-adsorbed Fe5C2 (110). Reprinted with permission Chen et al. [105]. Copyright 2022 Elsevier. (c) Electron distribution diagrams of surface atoms on Fe5C2 (510) and Na2O/Fe5C2 (510) (Fe atoms in purple, C atoms in gray, O atoms in red, and Na atoms in amaranth). Reprinted with permission Liu et al. [106]. Copyright 2021 American Chemical Society. (d) Energy profiles of the C–C coupling between CH2* and CO* species (CH2* +CO* → * + CH2CO*) on Fe5C2 (510) and Na2O/Fe5C2 (510). Na, Fe, C, O, and H are shown in purple, blue, gray, red, and white, respectively. Reprinted with permission Wang et al. [107]. Copyright 2021 American Chemical Society.
Molecules 29 01194 g008
Figure 9. (a) Mechanistic diagram of binary Fe–Al and ternary Fe–Zn–Al catalysts for selective synthesis of α-Olefins from CO2 hydrogenation. Reprinted with permission from Xu et al. [110]. Copyright 2021 American Chemical Society. (b) The average binding energy (Eb) of isolated carbon atoms on χ-Fe5C2 (510), Zn/χ-Fe5C2 (510), and ZnO/χ-Fe5C2 (510) surfaces at different coverages. Reprinted with permission from Liu et al. [21]. Copyright 2023 Elsevier. (c) Energy profiles for HCOO* formation from the MS configurations of CO2 on all the surfaces evaluated. Reprinted with permission from Belelli et al. [112]. Copyright 2023 Elsevier. (d) Reaction pathways for propene hydrogenation into propane on the Fe5C2 (111) (black line) and Cu/Fe5C2 (111) (red line) surfaces. Reprinted with permission from Yang et al. [113]. Copyright 2023 Elsevier.
Figure 9. (a) Mechanistic diagram of binary Fe–Al and ternary Fe–Zn–Al catalysts for selective synthesis of α-Olefins from CO2 hydrogenation. Reprinted with permission from Xu et al. [110]. Copyright 2021 American Chemical Society. (b) The average binding energy (Eb) of isolated carbon atoms on χ-Fe5C2 (510), Zn/χ-Fe5C2 (510), and ZnO/χ-Fe5C2 (510) surfaces at different coverages. Reprinted with permission from Liu et al. [21]. Copyright 2023 Elsevier. (c) Energy profiles for HCOO* formation from the MS configurations of CO2 on all the surfaces evaluated. Reprinted with permission from Belelli et al. [112]. Copyright 2023 Elsevier. (d) Reaction pathways for propene hydrogenation into propane on the Fe5C2 (111) (black line) and Cu/Fe5C2 (111) (red line) surfaces. Reprinted with permission from Yang et al. [113]. Copyright 2023 Elsevier.
Molecules 29 01194 g009
Figure 10. Relative energy diagrams for (a) the RWGS mechanism over Cu and CuK surfaces; (b) the CH* formation mechanism over Fe, FeK, Fe-Cu, and Fe-Cu-K surfaces. The effect of K on the chain growth mechanism from (c) CH* and (d) CH2* monomer. T = 300 °C and P = 2.5 MPa. Reprinted with permission from Hwang et al. [114]. Copyright 2019 Elsevier.
Figure 10. Relative energy diagrams for (a) the RWGS mechanism over Cu and CuK surfaces; (b) the CH* formation mechanism over Fe, FeK, Fe-Cu, and Fe-Cu-K surfaces. The effect of K on the chain growth mechanism from (c) CH* and (d) CH2* monomer. T = 300 °C and P = 2.5 MPa. Reprinted with permission from Hwang et al. [114]. Copyright 2019 Elsevier.
Molecules 29 01194 g010
Table 1. Microkinetic analysis of the redox and association reaction steps in the reverse water–gas shift reaction. Reprinted with permission from Pahija et al. [12]. Copyright 2022 American Chemical Society.
Table 1. Microkinetic analysis of the redox and association reaction steps in the reverse water–gas shift reaction. Reprinted with permission from Pahija et al. [12]. Copyright 2022 American Chemical Society.
Reaction StepsRedox PathwayCOOH PathwayHCOO Pathway
CO + * ⇆ CO*CO2 + H2CO2 + H2CO2 + H2
H2 + 2* ⇆ 2H*CO2* + 2H*COOHtrans* + H*HCOO* + H*
H2O + * ⇆ H2O*CO* + O* + 2H*COOHcis* + H*HCO* + O* + H*
H2O + * ⇆ H* + OH*CO* + OH* + H*CO* + OH* + H*CO* + O* + 2H*
CO2* ⇆ CO2 + *CO* + H2O*CO* + H2O*CO* + OH* + H*
OH + * ⇆ O* + H*CO* + H2OCO* + H2OCO* + H2O*
OH* + OH* ⇆ H2O* + O*CO + H2OCO + H2OCO* + H2O
CO* + O* ⇆ CO2* + * CO + H2O
CO* + OH* ⇆ COOHcis* + *
COOHtrans* + * ⇆ CO2* + H*
COOHtrans* + OH* ⇆ CO2* + H2O
CO* + OH* ⇆ HCOO* + *
CO2* + H2O* ⇆ HCOO* + OH*
OH* + OH* ⇆ H2O* + O*
COOH* + * ⇆ HCOOH**
COOHcis* ⇆ COOHtrans*
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Tang, H.; Qiu, T.; Wang, X.; Zhang, C.; Zhang, Z. A Brief Review of Recent Theoretical Advances in Fe-Based Catalysts for CO2 Hydrogenation. Molecules 2024, 29, 1194. https://doi.org/10.3390/molecules29061194

AMA Style

Tang H, Qiu T, Wang X, Zhang C, Zhang Z. A Brief Review of Recent Theoretical Advances in Fe-Based Catalysts for CO2 Hydrogenation. Molecules. 2024; 29(6):1194. https://doi.org/10.3390/molecules29061194

Chicago/Turabian Style

Tang, Haoxiang, Tongyue Qiu, Xuerui Wang, Chundong Zhang, and Zunmin Zhang. 2024. "A Brief Review of Recent Theoretical Advances in Fe-Based Catalysts for CO2 Hydrogenation" Molecules 29, no. 6: 1194. https://doi.org/10.3390/molecules29061194

Article Metrics

Back to TopTop