Next Article in Journal
Photocatalytic Inactivation of Viruses Using Graphitic Carbon Nitride-Based Photocatalysts: Virucidal Performance and Mechanism
Next Article in Special Issue
CO2 Reduction to Valuable Chemicals on TiO2-Carbon Photocatalysts Deposited on Silica Cloth
Previous Article in Journal
Atomic Structure of Pd-, Pt-, and PdPt-Based Catalysts of Total Oxidation of Methane: In Situ EXAFS Study
Previous Article in Special Issue
The Effect of Si on CO2 Methanation over Ni-xSi/ZrO2 Catalysts at Low Temperature
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in the Mitigation of the Catalyst Deactivation of CO2 Hydrogenation to Light Olefins

1
Chemistry Department, Long Island University (Post), Brookville, NY 11548, USA
2
College of Natural Science, The University of Texas at Austin, Austin, TX 78712, USA
3
College of Engineering, Cornell University, Ithaca, NY 14850, USA
4
John Jay College of Criminal Justice, New York, NY 10019, USA
*
Authors to whom correspondence should be addressed.
These authors contributed equally.
Catalysts 2021, 11(12), 1447; https://doi.org/10.3390/catal11121447
Submission received: 10 November 2021 / Revised: 25 November 2021 / Accepted: 25 November 2021 / Published: 28 November 2021
(This article belongs to the Special Issue New Trends in Catalysis for Sustainable CO2 Conversion)

Abstract

:
The catalytic conversion of CO2 to value-added chemicals and fuels has been long regarded as a promising approach to the mitigation of CO2 emissions if green hydrogen is used. Light olefins, particularly ethylene and propylene, as building blocks for polymers and plastics, are currently produced primarily from CO2-generating fossil resources. The identification of highly efficient catalysts with selective pathways for light olefin production from CO2 is a high-reward goal, but it has serious technical challenges, such as low selectivity and catalyst deactivation. In this review, we first provide a brief summary of the two dominant reaction pathways (CO2-Fischer-Tropsch and MeOH-mediated pathways), mechanistic insights, and catalytic materials for CO2 hydrogenation to light olefins. Then, we list the main deactivation mechanisms caused by carbon deposition, water formation, phase transformation and metal sintering/agglomeration. Finally, we detail the recent progress on catalyst development for enhanced olefin yields and catalyst stability by the following catalyst functionalities: (1) the promoter effect, (2) the support effect, (3) the bifunctional composite catalyst effect, and (4) the structure effect. The main focus of this review is to provide a useful resource for researchers to correlate catalyst deactivation and the recent research effort on catalyst development for enhanced olefin yields and catalyst stability.

1. Introduction

1.1. General Aspects

While carbon-rich fossil fuels like coal, oil, and natural gas have powered human civilization, the massive emission of CO2 as a greenhouse gas has caused severe and harmful effects on the ecological environment [1]. For example, the rise of sea levels is accelerating, the number of large hurricanes and wildfires is growing, and dangerous heat waves and more severe droughts are occurring in many areas. The CO2 concentration in the atmosphere had climbed to 415 ppm by 2020 (Figure 1), an increase of more than 40% relative to the pre-industrial era [2]. The atmospheric CO2 concentration will continue to rise to ~570 ppm by the end of the 21st century if no alleviation measures are taken [3]. Therefore, there is an urgent need to control CO2 emissions in order to mitigate their negative impact on the environment. In recent years, capture and storage technologies for the CO2 released from the burning of fossil fuels have emerged and developed in potential commercial scale applications [4,5,6,7]. In order to close the carbon gap, transforming the captured gas into value-added fuels and chemicals has become an urgent task for CO2 remediation [8,9].
The catalytic conversion of CO2 is a favorable approach to the mitigation of CO2 emissions by producing chemicals and fuels [8,10,11,12,13,14,15,16,17]. Light olefins such as ethylene, propylene and butylene (C2=−C4=), which are currently among the top petrochemicals, are the building blocks for the production of a wide variety of polymers, plastics, solvents, and cosmetics [8,13,18,19,20,21]. Moreover, light olefins can be oligomerized into long-chain hydrocarbons which can be used as fuels, making them a desirable product with high potential for the utilization—and therefore elimination—of up to 23% of CO2 emissions [8]. A highly promising route is selective CO2 hydrogenation to produce light olefins [10]. The huge market demand for the lower olefins offers a great opportunity for the target technology to profoundly impact the scale of CO2 utilization once it is developed with renewable hydrogen. The current chemical industry relies heavily on petroleum (the steam cracking of naphtha) for the production of light olefins [22]. The depletion or movement away from the refining of petroleum and the gap between the supply and demand of light olefins call for a new strategy to synthesize light olefins from alternative carbon sources [18,19,20,21,23,24]. A one-step process for the conversion of CO2 to light olefins is a highly desirable tactic to address the “3Rs” (reduce, reuse, and recycle) associated with ever-increasing CO2 levels, and to solve the paradox between the supply and demand of light olefins [25].
Currently, there are two primary pathways, as shown in Scheme 1, to produce light olefins from CO2 reduction by hydrogen (H2) in a one-step process: (1) the CO2 Fischer-Tropsch synthesis (CO2–FTS) route consists of two consecutive processes, the reverse water–gas shift (RWGS) reaction (Equation (1)) and subsequent Fischer–Tropsch synthesis (FTS) (Equation (2)); (2) the methanol (MeOH) mediated route consists of two consecutive processes, i.e., CO2-to-MeOH (Equation (3)) and a subsequent MeOH-to-olefins process (MTO) (Equation (4)). The complex reaction network in Scheme 2 indicates the competing reactions (i.e., Equation (5)) with the formation of light olefins. The control of the selectivity of the CO2 hydrogenation to the desired olefin product requires the design of catalysts for reaction pathways that are compatible with favorable thermodynamics and a good understanding of the reaction kinetics [26]. The thermodynamic values in the equations (Equations (1)–(5)) indicate that lower temperatures favor FTS (Equation (2)), methanol (Equation (3)), and methane synthesis (Equation (5)), while higher temperatures are needed to activate CO2 (Equation (1)) for rapid reaction rates [27]. The complex reaction network in Scheme 2 and thermodynamics suggest that the design and synthesis of catalysts for a one-step process to selectively produce olefins are challenging.
CO2–FTS reaction pathway:
Reverse water-gas shift reaction (RWGS):
CO2 + H2 → CO + H2O    △H0298 = 41.1 kJ mol−1
Fischer-Tropsch synthesis to olefins (FTS):
nCO + 2nH2 → (CH2)n + nH2O   △H0298 = −210.2 kJ mol−1 (n = 2)
Methanol mediated reaction pathway:
Methanol synthesis:
CO2 + 3H2 → CH3OH+ H2O   △H0298 = −49.3 kJ mol−1
Methanol to olefins (MTO):
nCH3OH → (CH2)n + H2O   △H0298 = −29.3 kJ mol−1 (n = 2)
CO2 methanation:
CO2 + 4H2 → CH4 + 2H2O   △H0298 = −165.0 kJ mol−1

1.2. Mechanistic Insights for CO2 Conversion to Light Olefins

In reviewing the mechanistic details of light olefin formation, it is clear that controlling the active H to C ratio is of primary importance. The presence of too much H* on the surface will result in excessive hydrogenation, and therefore methanation, while too little H* on the surface will restrict the hydrogenation ability of the catalyst and therefore reduce the CO2 conversion activity. At its most fundamental, the pivotal steps of CO2 conversion to light olefins are the cleavage of the C–O bonds and the formation of C–C bonds [25].
Iron-based catalysts have been extensively studied for use in the CO2–FTS route due to their relatively high utility and activity for both the RWGS and FTS component reactions. When using Fe-based catalysts for CO2–FTS, the initial Fe2O3 phase is reduced by hydrogen to Fe3O4 or a mixture of Fe3O4 and FeO. The resulting Fe3O4 is the active component for the RWGS reaction, and can be further reduced to form metallic Fe [27]. The reaction mechanism for the CO2–FTS pathway is suggested as shown in Scheme 3a. CO2 is first adsorbed and activated on the RWGS active phases (e.g., Fe3O4) to form a carboxylate (*CO2, * representing the adsorption state). The *CO2 can then be hydrogenated by adsorbed H to form an *HOCO intermediate. The intermediate then dissociates into *OH and *CO. The *OH is then hydrogenated into *H2O. Then, *CO either desorbs as CO gas or reacts further via successive FTS. In order to form hydrocarbons, the *CO is first partially hydrogenated into *HCO and then undergoes complete hydrogenation, dissociation, and finally dehydration to form *CHx species. The *CHx species are precursors for the formation of olefins. In an alternative mechanism, *CO can dissociate into *C and *O. Some *C can diffuse into the Fe-metal lattice to form metal carbides as χ-Fe5C2, the active component for the FTS reaction [27]. The C* on the χ-Fe5C2 surface can then be hydrogenated to CHx* species. C* + CHx* and CHx* + CHx* were the most likely coupling pathways [25].
As indicated above, the *C from the dissociation of *CO during the FTS reaction may diffuse into the α-Fe metal lattice, resulting in the formation of Fe7C3, χ-Fe5C2, θ-Fe3C, ε′-Fe2.2C, and ε-Fe2C phases, depending on reaction conditions [27]. Iron carbides play an essential role in CO hydrogenation/dissociation and C–C coupling. Some researchers have proposed that χ-Fe5C2 is the active phase, while θ-Fe3C is less active and can cause catalyst deactivation due to production of graphite, which has increased stability under typical FTS reaction conditions and may block the production of other active phases [27,28].
Alternatively, the reaction mechanism for the MeOH pathway is suggested as shown in Scheme 3b,c. The synthesis of MeOH can proceed via two pathways: (1) CO-mediated, in which the *CO intermediate, which was produced from the RWGS reaction via the dissociation of the carboxyl (*HOCO) species, is hydrogenated to methanol via *HCO and *COH, and (2) formate-mediated, in which the formate (*HCOO) species results from the hydrogenation of the carboxylate intermediate (*CO2), which is then reacted further to *H2COOH, *H2CO, *H2COH, and *H3CO. Through dehydration coupling, the methanol forms *CH2CH, and then forms olefins via subsequent hydrogenation [27].
The factors that may affect the CO2 conversion and light olefin selectivity are the catalyst composition (metals, supports, promotors, etc.), functionality (i.e., metal/zeolite bifunctionality), structure (i.e., layered metal oxide, core–shell, etc.), preparation methods (e.g., impregnation, hydrothermal, sol-gel, etc.) and testing conditions (e.g., temperature, pressure, CO2/H2 molar ratio, gas hourly space velocity, etc.). The focus of this review will be on the catalyst composition, functionality and structure. Other factors of catalyst preparation methods and testing conditions for CO2 conversion to light olefins can be found elsewhere [15,20,27,29,30].

1.3. Catalysts for CO2 Conversion to Light Olefins

As can be seen above, because each route has its own unique pathway of species and intermediates, different catalysts must be employed for the hydrogenation of CO2 to olefins depending on the chosen route. In the CO2–FTS path, Fe is one of the most widely used components in the catalysts, as catalysts containing Fe offer less methanation activity under higher reaction temperatures. As described above, it has been reported that Fe3O4 was the active phase responsible for RWGS; the metallic Fe and iron carbides could activate CO and produce hydrocarbons [31,32]. When incorporating alkali promoters, Fe-based catalysts showed greater olefin selectivity. The alkali metals, acting as electron donors to the Fe metal, facilitate the adsorption of CO2 while lowering the affinity for H2. The net result is a higher olefin yield [33,34,35]. There is also some indication that doping the catalyst with an additional metal may promote even higher olefin yields by forming a highly active interface. The second metal components allow for even greater adjustment of the CO2 and H2 adsorption and activation, shifting the distribution of the product more towards the desired hydrocarbons. By supporting the Fe-based catalysts on supports such as silica (SiO2), alumina (Al2O3), titania (TiO2), zirconia (ZrO2) and carbon materials (i.e., carbon nanotubes (CNTs), carbon nanospheres (CNSs), graphene oxide (GO)), the catalytic performance may be further enhanced by improving the active metal dispersion and slowing down the sintering of the active particles [36,37]. Controlling hydrocarbon chain growth to achieve a desired carbon range (i.e., C2–C4) remains a challenge for CO2 conversion due to the product selectivity limit governed by the Anderson–Schulz–Flory (ASF) distribution with a maximum achievable C2–C4 hydrocarbons selectivity of less than 60%, as shown in Figure 2 [30,38].
Alternatively, in the MeOH path, light olefins can be synthesized with selectivity as high as 80–90% among hydrocarbons, exceeding the ASF product distribution limit for FTS reactions [18,39,40,41]. Some plausible reasons for the reported ASF distribution deviation are the blockage of surface polymerization by intermediates, (e.g., ketene (CH2CO)), space confinement, or the use of catalysts with two types of active sites (i.e., bifunctional catalysts) [27]. Regardless of the reasons, the observed deviation from the ASF distribution offers opportunities to increase the selectivity to olefins [27]. Several recent studies have reported the results for the combination of MeOH synthesis catalysts (i.e., In2O3, In-Zr, ZnGa2O4, MgGa2O4, ZnAl2O4, MgAl2O4, ZnZrO and In2O3-ZnZrO2) with an MTO catalyst (i.e., SAPO-34, SSZ-13 and ZSM-5), and their ability to produce light olefins with enhanced selectivity for CO2 hydrogenation [13,42,43,44]. It has been proposed that the secondary functionality of acid–base sites on the catalytic support significantly impacts the light olefin selectivity. For example, by passivating the Brønsted acid sites of In2O3-ZnZrOx/SAPO-34, the secondary hydrogenation reaction is inhibited, thereby improving the olefin selectivity [27,30].

1.4. The Main Focus of This Review

Even though significant efforts have been made, considerable challenges remain in the development of highly efficient catalysts with selective pathways to light olefins due to the thermodynamically stable nature of the CO2 molecule, the complexity of the reaction networks, and catalyst deactivation [8,45,46]. Several recent reviews have summarized CO2 hydrogenation to value-added products, including light olefins [25,27,30,38]. However, it is necessary to present a review focused on the recent advances in the mitigation of the catalyst deactivation of CO2 hydrogenation to light olefins, as catalyst deactivation has been a big challenge that provides economic hurdles to the adoption of the new technologies.
Because catalysts and mechanisms have been extensively reviewed in numerous review papers [25,27,30,38], the focuses of the current article are to identify possible causes that trigger catalyst deactivation and summarize recent advances on catalyst development with enhanced catalyst stability and light olefin selectivity for CO2 hydrogenation. In this review, we first provide a brief summary of the two dominant reaction pathways (CO2–FTS and MeOH-mediated), mechanistic insights and catalytic materials for CO2 hydrogenation to light olefins. We then list the deactivation mechanism caused by carbon deposition, water formation, phase transformation and metal sintering/agglomeration. Finally, we summarize the recent progress published within five years on catalyst development that improves catalyst deactivation by the following catalyst functionalities: (1) the promoter effect, (2) the support effect, (3) the hybrid functional effect, and (4) the structure effect.
Each one of these aspects is accompanied by a suitable table in which the most significant literature findings are comparatively presented. To the best of our knowledge, no review has ever directly correlated the causes of catalyst deactivation and catalyst mitigation for CO2 hydrogenation to light olefins. Herein, we attempt to provide a useful resource for researchers to correlate the catalyst deactivation and the recent research effort on catalyst development for enhanced olefin yield and catalyst stability.

2. Causes of Catalyst Deactivation

During CO2 hydrogenation, catalyst deactivation can occur via several mechanisms, resulting in decreased activity and selectivity toward the desired olefins. The determination of the mechanism of deactivation is an important step toward mitigation. The primary causes of catalyst deactivation are the sintering (or agglomeration) of metal particles, phase transformation at the catalyst’s surface, and catalyst poisoning by water or carbonaceous deposits (i.e., coke). An understanding of the deactivation causes is necessary to develop a mitigation strategy and sustain high selectivity toward the desired olefins during CO2 hydrogenation. For context, we present brief descriptions of each of these causes with a few representative examples from the literature that demonstrate the necessity of robust and novel mitigation studies. More thorough reviews of the deactivation causes and their mechanisms can be found elsewhere [47,48,49].

2.1. Sintering

Catalyst sintering can occur through either Ostwald ripening or particle migration and coalescence, as shown in Figure 3 [50]. Through sintering, the agglomeration of smaller catalyst crystals into larger ones will bring about the loss of the pore structure, which lowers the internal surface area of the catalyst, leading to the deactivation. In the area of FT by cobalt catalysts, several groups have determined that the particle growth of cobalt is the largest factor causing deactivation [51,52].
Sun et al. [53] examined sintering in zinc- and alumina-supported copper catalysts (Cu/ZnO/Al2O3). It was found that the presence of CO in the process employed for CH3OH synthesis strongly contributed to the deactivation of the catalysis over 0 to 50 h. Taken with corroborative evidence from the Cu surface area determination, the deactivation was likely attributed to the sintering of the Cu metal.
As mentioned above, sintering negatively affects the catalytic performance due to many reasons: for example, the overall catalytically active surface area is reduced due to the collapse of the structure and the chemical alteration of the catalytically active phases to non-active phases [50,54,55]. As this form of deactivation involves the coalescence of larger particles from smaller, it is extremely difficult to reverse. Sintering, therefore, is easier to prevent through careful catalyst design [50,56]. For example, Li et al. observed remarkable metal sintering on supported FeCo/ZrO2 catalysts [56]. As shown in Figure 4A(a), for the 13Fe2Co/ZrO2 supported catalyst precursor prepared using the conventional impregnation method, the Co and Fe are distributed into separate oxide particles, which increased the possibility of sintering. As confirmed in Figure 4A(b,c), aggregates composed of Fe and Co oxide nanoparticles were observed on the ZrO2 fibers, with an average diameter of ca. 15 nm before the reaction. The particle size increased to 48 nm after the reaction, which was responsible for the rapid deactivation of activity (Figure 4A(d–f)). By comparison, Fe-Co-Zr polymetallic fibers obtained via a one-step electrospinning technique showed that Fe and Co were dispersed in proximity to ZrO2, as shown in Figure 4C(a), but separately from each other. In order to reduce the possibility of sintering, as demonstrated in Figure 4B(a–f), the Fe and Co oxides nanoparticles successfully dispersed with the ZrO2 particles for the polymetallic oxide fibers, with an average size of roughly 1–2 nm before the reaction, and after the reaction, the particle size barely changed, which contributed to the stable catalytic activity after 500 mins on stream (Figure 4C(a,b)).

2.2. Phase Transformations

Phase transformations are processes of deactivation involving the conversion of an active crystalline phase of the catalyst (or one of its components) into a different inactive one. These transformations can involve both metal-supported and metal-oxide catalysts. In the former type of catalyst, atoms from the catalyst’s support will diffuse into the catalyst’s surface. A reaction at the surface can then result in an inactive phase, deactivating the catalyst.
Riedel et al. was able to demonstrate that the steady states of the synthesis of hydrocarbons using iron oxides could be separated into five episodes of distinct kinetic regimes. In episode I, the adsorption of the reactants takes place on the catalyst surface and carbonization occurs. During episodes II and III, products from the RWGS reaction dominate during ongoing carbon deposition. In episode IV, the rate of FT activity increases up to the steady state, and the maintenance of the steady state occurs in episode V. Prior to the reaction, the iron phases of the reduced catalyst are mainly α-Fe and Fe3O4, along with a small amount of Fe2O3. As the process proceeds, the Fe3O4 and Fe2O3 phases are consumed and a new oxidic iron amorphous phase is formed, which appears to be active for the RWGS reaction. Through a reaction of iron with carbon from the CO dissociation, FTS activity commences with the formation of iron carbide (Fe5C2). Upon the formation of the stable but inactive carbide (Fe3C), which is the result of Fe5C2 carburization, the catalyst begins deactivating [57,58,59]. Lee et al. studied the causes of the deactivation of Fe–K/γ-Al2O3 for CO2 hydrogenation to hydrocarbons, and found the causes for deactivation varied based on positioning inside the reactor. Over time, the Fe2O3 was reduced to active phase χ-Fe5C3, and then the χ-Fe5C3 was transformed to θ-FeC3, a form which is not active for CO2 hydrogenation. The primary reason for deactivation was the phase transformation at the top of the reactor. Conversely, at the bottom of the reactor, deactivation was largely the result of deposited coke generated by secondary reactions [57].
Zhang et al. reported the structure evolution of the iron catalyst during its full catalytic life cycle of CO2 to olefins (CTO), including the catalyst activation, reaction/deactivation (120 h) and regeneration. The phase transition during the CO activation was observed to follow the sequence of Fe2O3 → Fe3O4 → Fe → Fe5C2. The primary deactivation mechanism during CTO was identified as the irreversible transition of iron phases under reaction conditions. Two possible pathways of the phase transition of the iron catalyst under CTO conditions have been identified, i.e., Fe5C2 → Fe3O4 and Fe5C2 → Fe3C → Fe3O4. Moreover, carbon deposition and the agglomeration of the catalyst particle proves to have relatively minor impacts on the catalytic activity compared with phase transition during the 120 h of reaction [60].
It appears that transformation to iron oxides will destroy catalyst activity. There is some question as to whether the cementite phase itself is problematic to activity, as higher-surface-area cementite phases have been reported to perform CO2 hydrogenation quite effectively [61].

2.3. Poisoning

Catalytic poisoning is a result of the strong chemisorption of reactants, products or impurities on sites that would otherwise be capable of catalysis. In essence, the poisoning ability of a particular species is related to the strength of its chemisorption to the catalysis relative to the other reactants that are competing for the catalytic active sites. The poisoning has two deactivating effects; a poison physically blocks the active sites from receiving additional reactants, and a poison can alter the electronic or structural properties of the catalytic surface, rendering it partially or completely ineffective toward catalysis [62,63].
While there are several different poisons which have been reported in the literature that have shown to deactivate CO2 hydrogenation catalysts [64,65], we will focus on the two which are the most pervasive, namely water and carbonaceous deposits or coke.

2.3.1. Water Poisoning

As seen in the above CO2 hydrogenation reactions, the dissociation of CO2 produces oxygen atoms, which in turn results in the formation of water. This byproduct is necessary for the thermodynamic favorability of the entire process, but can be unfortunately detrimental to catalytic performance. It is because of this unavoidable mechanistic absolute that the mitigation of water poisoning must be part of all catalytic investigations [66].
Wu et al. [67,68] examined the effect of the produced water on the stability of Cu/ZnO-based catalysts in methanol synthesis from the high temperature hydrogenation of CO2. Specifically, Cu/ZnO/ZrO2/Al2O3 (40/30/25/5) was subjected to a CO2-rich feed, which produces water, and a CO-rich feed, which does not produce water. The examination of the catalysts by X-ray powder diffraction (XRD) after 1 h and 500 h time-on-stream of a CO2-rich feed containing steam showed that the Cu and ZnO crystallized more rapidly when compared to identical catalysts exposed to a CO-rich feed not containing steam. In particular, the Cu particle size in the catalyst used with the CO2-rich feed containing steam grew from 94 Å to 166 Å from 1 h to 500 h. The particle size growth under steam might be the key reason causing catalyst deactivation.
Huber et al. observed the rapid deactivation of Co/SiO2 during an FTS reaction at high water partial pressure, and the loss of activity was attributed to the support breakdown byproduct water accompanied by the formation of stable, inactive cobalt-silicates and the loss of the BET surface area [69]. van Steen et al. stated that metallic cobalt crystallites with a diameter less than 4.4 nm are more susceptible to oxidation by water to form Co(II)O [70]. This is in agreement with Iglesia’s work showing that small Co metal crystallites (<5–6 nm diameter) appear to re-oxidize and deactivate rapidly in the presence of a water reaction product in typical FTS conditions [71].
Water poisoning has the most dramatic effect on zeolite-based CO2 hydrogenation catalysts for which the acidic sites of the zeolite are essential for catalysis. Recently, Zhang et al. investigated the water effect over zeolite-based catalysts at high temperatures, and found that water caused the loss of crystallinity and modified acid sites, thereby deactivating the catalyst [72]. Their studies show, by functionalization with organosilanes, that the tolerance of defective zeolites to hot liquid water can be greatly enhanced. This method renders the zeolite hydrophobic, which prevents the wetting of the surface. At the same time, the organosilanes act as a capping agent of Si−OH species, reducing their reactivity. Both aspects are important for the prevention of water attack [72].
It appears that there are several analogies of Fe catalysts for CO2 hydrogenation and CO catalysts for conventional FT synthesis. Kliewer et al., for example, showed that for a supported CO catalyst, water can oxidize the surface of the CO to an inactive oxide phase, and it also plays a large role in sintering. With a high water partial pressure in the Fe system, it appears that this can also oxidize iron carbides to inactive surface oxide phases and also promotes particle growth sintering [51].

2.3.2. Carbonaceous Deposits (Coke)

Coke is produced by the decomposition or condensation of hydrocarbons on the surfaces of catalysts, and is primarily is comprised of polymerized hydrocarbons. There have been several books and reviews that describe the formation of coke on catalysts, and the resulting deactivation [73,74,75,76,77,78].
These deposits are most problematic for catalysis involving zeolites, because the active sites of the zeolites become blocked or fouled by the coke deposits. The deactivation of MTO reactions over zeolites due to coke deposition results in a reduction in both the catalyst activity and product selectivity [79,80,81].
Nishiyama et al. [82] studied the effect of the SAPO-34 crystal size on the catalyst lifetime, and found that the amount of coke deposited on the deactivated SAPO-34 catalyst increased with the decreasing crystal size, indicating that for larger crystals, the reactants were unable to penetrate further into the larger crystals to reach other acidic sites. Because MTO reactions and coke formation take place simultaneously in the same pores, it seems likely that the effectiveness of the catalyst increased with the decreasing crystal size. Their studies demonstrated that the coke formation was inhibited in small-crystal SAPO-34 due to reduced diffusive resistance.
The work of Wei et al. on CO2 hydrogenation found that the deactivation of the zeolites HMCM-22 and HBeta was the result of coke formation, which deposited in the zeolites’ cavities and channels. The deposition blocked the reactants’ access to the zeolites’ acid sites, leading to the deactivation [11]. Muller et al. investigated the MTO process on H-ZSM-5 catalysts in plug-flow (PFR) and fully back-mixed reactors (CSTR). They found that the catalysts deactivated under the homogeneous gas phase in the CSTR. It was shown unequivocally that, in the early stages of the reaction, the zeolite deactivates via Brønsted acid site blocking, and not by coke-induced deposition restricting the pore access. The deactivation of H-ZSM-5 in the CSTR occurred at first rapidly, and then at a much slower rate (Figure 5). The rapid deactivation was observed in a PFR due to the formation of a larger fraction of the oxygen-containing carbon species. The larger fraction of oxygen-containing carbon species increases the reaction with the desired olefins, which results in a strongly adsorbed aromatic molecule. The formation of aromatic coke proceeds mostly by hydride transfer between olefins and carbon growth via multiple methylations of such aromatic species [83].
Zeolite-based catalysts that show promise for high olefin selectivity are unfortunately typically limited by mass transfer, suffering from rapid deactivation due to carbon deposition and water poisoning [83]. The issues with coke deactivation on the zeolite catalysts involved in MTO reactions are seen in classical MTO chemistry. The directed transformation of coke into active intermediates in a methanol-to-olefins catalyst was reported to boost the light olefin selectivity [84]. Another strategy to mitigate the deactivation was to synthesize nanozeolites, which have shortened diffusion paths, or mesoporous hierarchical zeolites, which exhibit longer catalyst lifetimes because of their larger pores and improved mass transfer [85,86,87].
With an Fe catalyst, the deactivation by coke is not related to the constriction of narrow pores, but several authors have reported the formation of carbonaceous residues on the active sites. Lee et al. investigated the deactivation behavior of an Fe–K/-Al2O3 catalyst, and found that the deactivation pathway was different according to the reaction position and reaction time. The main deactivation reason was the phase transformation at the top of the reactor. Conversely, the main factor at the bottom of the reactor was the deposited coke generated by secondary reactions. In particular, the produced olefins may have been adsorbed on acidic sites, and thus the olefins served as major precursors to coke. The SEM micrographs of the used catalysts clearly showed that most of the surface was covered by deposited graphite and graphite clusters protruding on the surface, mixed with some fine filamentous carbon (Figure 6) [57].
With these multiple deactivation pathways having been identified, it now becomes a critical issue to find ways of modifying the catalyst to become more stable. In the section directly below, we will describe the approaches that multiple researchers have examined in an attempt to mitigate deactivation.

3. Recent Progress on the Mitigation of the Catalyst Deactivation

We will discuss the effect of promotors, metal oxide support, bifunctional composition and structure on the catalyst design to minimize the catalyst deactivation. We will summarize reports published within the last five years showing that promoters, supports, and novel morphology designs have mitigated the deactivation effects.

3.1. Promoter Effect

Fe-based catalysts have been widely studied in CO2 hydrogenation, and usually show unsatisfactory selectivity toward lower olefins. The addition of suitable promotors to increase the yield of light olefins and the stability of the catalysts by controlling the electronic and structural properties have been extensively studied. Alkali metals such as K and Na have been broadly used as promotors to control the electronic properties. Mn, Ce, Ca metals have been used as structural promotors. Transition metals such as Zn, Co, Cu, V, Zr, etc., have been used as both electronic and structural promotors. Some representative catalysts on the promoter effect for CO2 hydrogenation to light olefins with improved catalyst stability are presented in Table 1.
Adding alkali metals (i.e., Na, K) could increase the selectivity towards light olefins due to the enhanced CO2 adsorption on the more electron-rich Fe phases and suppressed H2 chemisorption, which inhibits olefin re-adsorption. Numpilai et al. reported on the effect of varying the content of the K promoter on the Fe-Co/K-Al2O3 catalysts via the CO2–FTS reaction pathway. Unpromoted catalysts evidenced low-light olefin yields when compared to K-promoted ones with an ascending K/Fe ratio from 0 to 2.5. The maximum light olefin (C2=−C4=) distribution of 46.7% and O/P ratio of 7.6 were achieved over the catalyst promoted with a K/Fe atomic ratio of 2.5. The positive effect of K’s addition is attributed to the strong interaction of H adsorbed with the catalyst surface caused by the electron donor from K to Fe species. This notion is also rationalized by the fact that the K promoter enhances the bond strength of absorbed CO2 and H2, retarding the hydrogenation of olefins to paraffins. In the same operating conditions, the catalyst promoted with a K/Fe atomic ratio of 0.5 provides the maximum light olefin (C2=−C4=) yield of 16.4%, which is significantly higher than that of 2.5 KFe catalysts (13.4%). This is explained by the K enriched surface of 2.5 KFe catalysts significantly reducing the BET surface area and generating a hydrogen-lean environment, ultimately lessening the catalytic activity [101].
A different promoter source plays an important role to affect catalytic CO2 hydrogenation. Han et al. demonstrated that as the series of K-promoters changes from K2CO3, CH3COOK, KHCO3, and KOH, the electron transfer from potassium to iron species is facilitated, which forms a more active and distinct χ-Fe5C2-K2CO3 interface during CO2 hydrogenation. This results in a higher selectivity to light olefins (75%) and a higher CO2 conversion (32%). In contrast, the non-carbonaceous K-promoters do not facilitate iron species to form iron carbides, which causes an undesirable catalytic performance (Figure 7a). Additionally, the close proximity between carbonaceous K-promoters and Fe/C catalyst components produced high olefin yields and catalytic stability (Figure 7b) [94]. Guo et al. reported that K derived from biological rather than inorganic precursors showed a stronger migration ability during the CO2 hydrogenation to light olefins. These surface-enriched K ions extracted from corncobs could promote the carburization of iron species to form more Fe5C2, promoting both the reverse water–gas shift reaction and subsequent C–C coupling [97].
Metal organic frameworks as precursors for the preparation of heterogeneous catalysts have been used recently [99,106,107]. Ramirez et al. used a metal organic framework as a catalyst precursor to synthesize a highly active, selective, and stable catalyst, as shown in Figure 8a–f for the hydrogenation of CO2 to light olefins. Comparing the addition of Cu, Mo, Li, Na, K, Mg, Ca, Zn, Ni, Co, Mn, Fe, Pt, and Rh to an Fe/C composite, only K is able to enhance olefin selectivity, as shown in Figure 8c. The presence of K promoted the formation of Fe5C2 and Fe7C3 carbides, as confirmed by XRD (Figure 8e). K helped keep a good balance between the iron oxide for RWGS and iron carbide for FTS. The results presented in Figure 8f indicated a trend in which methane formation decreased and olefin selectivity increased as the K loading increased. The catalyst Fe/C+K(0.75) exhibited good stability (Figure 8d) and outstanding C2−C4 olefin space time yields of 33.6 mmol·gcat−1·h−1 at XCO2 = 40%, 593 K, 30 bar, H2/CO2 = 3, and 24,000 mL·g−1·h−1 [99].
Some work may shed light on the ways in which the alkali promoters affect the behavior of iron catalysts. By the precisely controlled addition of promoters to fine tune the catalytic performance for the hydrogenation of CO2 to olefins, Yang et al. investigated how a zinc ferrite catalyst system could be affected by the addition of sodium and potassium promoters, specifically on the conversion of CO and CO2 to olefins. It was found that the catalyst’s composition of iron oxides and iron carbides was altered in the presence of the promoters, which affected the CO and CO2 conversion. The production of C2+ olefins was greatly facilitated by the Na- and K-promoted catalysts. The Na/Fe-Zn catalyst was found to possess the optimal olefin productivity, and inhibited the competitive methanation reaction. It showed a total carbon conversion of 34.0%, which decreased by only 12.2% over 200 h [96]. Similarly, Wei et al. unraveled the effect of the Na promoter on the evolution of iron and carbon species, as well as the consequent tuning effect on the hydrogenation of CO2 to olefins. With the contents of the Na promoter increasing from 0 wt% to 0.5 wt%, the ratio of olefins to paraffins (C2+) rose markedly, from 0.70 to 5.67. The in situ XRD and temperature programed surface reaction (TPSR) confirmed that the introduction of the Na promoter decreased the particle size of Fe5C2 and regulated the distribution of surface carbon species. Furthermore, the in situ XRD and Raman demonstrated that the interaction between the Na promoter and the catalysts inhibited the hydrogenation of Fe5C2 and surface graphitic carbon species, consequently improving the stability of the Fe5C2 and enhancing the formation of olefins by inhibiting the hydrogenation of the intermediate carbon species [92]. Using a similar approach, Liang et al. modified the xNa/Fe-based catalysts with tunable amounts of sodium promoter for CO2 hydrogenation to alkenes, with CO2 conversion at 36.8% and a light olefin selectivity of 64.3%. It was found that the addition of the Na promoter into Fe-based catalysts boosted the adsorption of CO2, facilitated the formation and stability of the active Fe5C2 phase, and inhibited the secondary hydrogenation of alkenes under the CO2 hydrogenation reaction conditions (Figure 9a–c). The content of Fe5C2 correlated with the amount of Na is shown in Figure 9d [98].
Wei et al. synthesized a series of Fe3O4-based nanocatalysts with varying sodium contents. The residual sodium markedly influenced the textural properties of the Fe3O4-based catalysts, and faintly hampered the reduction of the catalysts. However, it discernibly promoted the surface basicity and prominently improved the carburization degrees of the iron catalysts, which is favored for olefin production. Compared with the sodium-free Fe3O4 catalysts, the sodium-promoted Fe3O4 catalysts displayed higher activity and selectivity for C2–C4 olefins. The FeNa catalyst (1.18) (Na/Fe weight ratio of 1.18/100) exhibited a high degree of catalytic activity with a high olefin/paraffin ratio (6.2) and selectivity to C2–C4 olefins (46.6%), and fairly low CO and CH4 production at a CO2 conversion of 40.5%. This catalyst also exhibited superb stability during the 100 h test at 593K. Comparing the scanning electron microscopy (SEM) image after reduction, there was no apparent indication of particle size growth after catalytic reaction for 100 h, further revealing the improved reaction stability of these iron nanoparticles [31]. Zhang et al. fabricated a Na- and Zn-promoted iron catalyst by a sol-gel method, and demonstrated its high activity, selectivity and stability towards the formation of C2+ olefins in the hydrogenation of CO2 into C2+ olefins. The selectivity of the C2+ olefins reached 78%, and the space–time yield of olefins was as high as 3.4 g gcat−1 h−1. The catalyst was composed of ZnO and χ-Fe5C2 phases with Na+ dispersed on both ZnO and χ-Fe5C2. Zhang et al. found that ZnO functions for the RWGS reaction of CO2 to CO, while χ-Fe5C2 is responsible for CO hydrogenation to olefins. The presence of Na+ enhanced the selectivity of C2+ olefins by regulating the hydrogenation ability and facilitating the desorption of olefins (Figure 10a). The presence of ZnO not only efficiently catalyzes the RWGS reaction but also improves the activity and stability of CO2 hydrogenation by controlling the size of χ-Fe5C2 (Figure 10c,d). It was further discovered that the close proximity between ZnO and χ-Fe5C2 is beneficial for the conversion of CO2 to olefins (Figure 10b). The larger interface could facilitate the diffusion and transfer of intermediate CO from ZnO to χ-Fe5C2, favoring CO2 adsorption and subsequent CO hydrogenation to C2+ olefins [90]. Malhi et al. also investigated the effect of Na and Zn on iron-based catalysts, and found that the modified Fe-based catalyst exhibited a good performance for CO2 hydrogenation to olefins, with a CO2 conversion of 43%, a selectivity of 54.1% to C2+= olefins, and a high olefins-to-paraffins ratio of 7.3 [93].
Chaipraditgul et al. investigated the effect of transition metals (Cu, Co, Zn, Mn or V) on the Fe/K-Al2O3 catalyst and found that the inclusion of the transition metal remarkably affected the interaction between the catalysts’ surface and the adsorptive CO2 and H2. The Fe/K-Al2O3 promoted with Cu, Co or Zn showed a lower the olefin to paraffin ratio, owning to a markedly increased number of weakly adsorbed H atoms resulting from the enhanced hydrogenation ability of the promoted catalysts. On the contrary, the addition of a Mn promoter to Fe/K-Al2O3 reduced the number of weakly adsorbed H atoms, lowering the hydrogenation ability to result in a high olefin to paraffin ratio of 7.4. The presence of either Mn or V inhibited the CO hydrogenation to hydrocarbon, leading to the low CO2 conversion, while the CO2 conversion was enhanced by incorporating either Co or Cu onto the Fe/K-Al2O3 catalyst [33]. Gong et al. investigated the promoting effect that Cu had on Fe-Mn-based catalysts in the production of light olefins via the CO2–FTS process. The Cu promoter was found to facilitate the reduction process and enhance CO dissociative adsorption by altering the interactions between Fe, Mn and the SiO2 binder, which led to increased activity. The addition of Cu weakened the surface basicity, which in turn decreased the chain growth probability and yielded a higher selectivity of light olefins [108].
Jiang et al. reported the synthesis of Mn-modified Fe3O4 microsphere catalysts. These catalysts demonstrated excellent catalytic performance, with a 44.7% CO2 conversion, 46.2% light olefin selectivity, and 18.7% light olefin yield over the 10 Mn−Fe3O4 catalyst. The O/P ratio increased from 3.7 for the unpromoted Fe3O4 catalyst to 6.5 for the Mn-promoted catalyst. An even distribution of manganese was found over the surface of the Fe3O4 microsphere. Such homogeneous dispersion allows for an increase in the basicity of the catalyst, which prevents the further hydrogenation of olefins into paraffins. It was noted that the synergistic effects between Fe and Mn improve the dissociation and conversion of CO2 to hydrocarbons. The addition of Mn was found to promote the production of Fe carbides and enhance the active phases of CO2 hydrogenation and the FTS reaction, as well as preventing the hydrogenation of light olefins into paraffins and chain growth into longer hydrocarbons [88]. A similar effect of the addition of Mn to Na/Fe catalysts was also observed by Liang et al. [95].
Zhang et al. synthesized uniform microspheres of Fe-Zr-Ce-K catalysts by microwave-assisted homogeneous precipitation, and found that the reducibility, surface basicity and surface atom composition of the catalysts were greatly affected by varying the Ce content. CeO2, as the structural promoter, restrained the growth of Fe2O3 crystallite, weakening the interaction between Fe species and zirconia, and enabling the easier reduction of Fe2O3. The best performance was obtained on a 35Fe-7Zr-1Ce-K catalyst at 593 K and 2 MPa, with a CO2 conversion of 57.34%, a C2–C4 olefin selectivity of 55.67%, and a ratio of olefin/paraffin of 7 [100].
Extensive research efforts have been exerted on the development of bi-metallic catalysts for the conversion of CO2 to light olefins. Yuan et al. demonstrated the influence of Na, Co and intimacy between Fe and Co on the catalytic performance of Fe-Co bimetallic catalysts for CO2 hydrogenation that offers an olefin to paraffin ratio of 6 at a CO2 conversion rate of 41%. With the introduction of Co into the Fe catalyst, the CO2 conversion is significantly enhanced. The intimate contact between the Fe and Co sites favored the production of C2–C4=. When Na was added to the system, the surface of the catalyst became carbon-rich and hydrogen-poor, allowing C–C coupling to form light olefins and suppress the methane formation. Moreover, the addition of a Na promoter facilitated the generation of χ-(Fe1-xCox)5C2 under the CO2 hydrogenation reaction conditions, and thus further improved the catalytic performances. A superb stability over 100 h was observed (Figure 11) [91]. Witoon et al. investigated the effect of Zn addition to Fe-Co/K-Al2O3 catalysts. The addition of Zn resulted in the improved dispersion and reducibility of iron oxides. For example, the 0.58 wt% Zn-promoted Fe-Co/K-Al2O3 catalyst afforded a large number of active sites for the adsorption of CO and H2 due to higher dispersion and an eased reducibility (Figure 12a). The catalyst exhibited superior activity for light olefin formation with yield of 19.9% under the optimum testing conditions of 613 K, 25 bar, 9000 mL gcat−1 h−1 and a H2/CO2 ratio of 4. Figure 12b also shows a gradual decrease in the olefin to paraffin ratio, with an almost constant CO2 conversion as a function of the time-on-stream (TOS). The X-ray photoelectron spectroscopy (XPS) analysis of the spent catalyst showed the continuous growth of iron carbide with the time-on-stream, indicating that iron carbide may be the active component resulting in paraffin production (Figure 12c). XRD confirmed the formation of Fe-C phases over the spent 0.58 wt% Zn-promoted Fe-Co/K-Al2O3 catalyst at the time-on-stream (Figure 12d) [89].
Wang et al. reported the synthesis of γ-alumina supported Fe-Cu bimetallic catalysts, and found a strong bimetallic promotion for selective CO2 conversion to olefin-rich C2+ hydrocarbons resulting from the combination of Fe and Cu at a specific composition. The suppression of the undesired CH4 formation was achieved by the addition of Cu to Fe while simultaneously enhancing the C–C coupling for C2+ hydrocarbon formation. The formation of the Fe-Cu alloy in the Fe-Cu(0.17)/Al2O3 catalyst is suggested by the XRD results. Furthermore, the addition of K into the Fe-Cu considerably enhanced the production of C2=–C4= light olefins and the O/P ratio over Fe-Cu bimetallic catalysts. The Fe-Cu/K catalysts exhibited the superior selectivity of C2+ hydrocarbons compared to Fe-Co/K catalysts under the same reaction conditions [102]. Kim et al. synthesized monodisperse nanoparticles (NPs) of CoFe2O4 by the thermal decomposition of metal−oleate complexes, as shown in Scheme 4. The prepared NPs were supported on carbon nanotubes (CNTs), and Na was added to investigate the promoter and support effects on the catalyst for CO2 hydrogenation to light olefins. The resulting Na-CoFe2O4/CNT exhibited a superior CO2 conversion of 34% and a light olefin selectivity of 39% compared to other reported Fe-based catalysts under similar reaction conditions. The superb performance of Na-CoFe2O4/CNT was attributed to the formation of a bimetallic alloy carbide, (Fe1−xCox)5C2. Higher CO2 conversion and better light olefin selectivity were found in comparison with conventional Fe-only catalysts which possess χ- Fe5C2 active sites and drastically improved the C2+ hydrocarbon formation in comparison with Co-only catalysts which contain Co2C sites [103].
Song et al. investigated titania-supported monometallic and bimetallic Fe-based catalysts for CO2 conversion, and found that the mono-metallic catalyst (Fe-, Co-, Cu-) performed poorly for C–C coupling reactions. However, adding a small amount of a second metal (Co and Cu) to Fe revealed the synergetic promotion on the CO2 conversion and the space–time yields (STY) of hydrocarbon products. The inclusion of K and La as promoters further improved the activity, giving a higher hydrocarbon selectivity and O/P ratio, indicating that the promotor facilitated the CO2 activation and C–C couplings over bi-metallic catalysts [106]. Zhang et al. investigated Fe-Zn bimetallic catalysts for CO2 hydrogenation to C2+ olefins. A high C2+ olefin selectivity of 57.8% after 200 h of time-on-stream at a CO2 conversion of 35.0% was obtained over an Fe2Zn1 catalyst. In bimetallic Fe5C2-ZnO catalysts, the ZnO plays a crucial role in improving the performance by altering the structure of the Fe components. Without ZnO, the chief deactivation mechanism was attributed to a phase transition from FeCx to FeOx over Fe2O3. However, with the addition of Zn to Fe2O3, the phase transformation and the carbon deposits over Fe2Zn1 were greatly diminished. Furthermore, the addition of Na inhibited the oxidation of χ-Fe5C2 active species for Fe-Zn bimetallic catalysts. During activation, both Zn and Na were shown to migrate onto the catalysts’ surfaces. The oxidation of FeCx by H2O and CO2 was shown to be diminished by the interaction between Zn and Na [28].
Xu et al. investigated the roles of Fe-Co interactions over ternary spinel-type ZnCoxFe2-xO4 catalysts for CO2 hydrogenation to produce light olefins. As shown in Figure 13, a high light olefin selectivity of 36.1%, a low CO selectivity of 5.8% at a high CO2 conversion of 49.6%, and an excellent catalyst stability were obtained over the ZnCo0.5Fe1.5O4 via the RWGS–FTS reaction pathway. It was shown that during the CO2 hydrogenation over ternary ZnCo0.5Fe1.5O4 catalysts, the formation of electron-rich Fe0 atoms in the CoFe alloy phase significantly boosted the generation of the active χ-Fe5C2, Co2C, and θ-Fe3C phases, in which the χ-Fe5C2 phase facilitated the C–C coupling, the Co2C species suppressed the formation of CH4, and the formation of the θ-Fe3C phase with lower hydrogenation activity inhibited the second hydrogenation of light olefins [105].
In summary, the use of the appropriate K or Na promoter, the inclusion of Cu, Co, Zn, Mn or Ce in the Fe phase, and the bi-metallic formation played important roles for enhanced catalytic performance and stability.

3.2. Support Effect

Catalyst support plays an important role in the overall activity and selectivity due to the interactions between the active metal components and the support during CO2–FTS. Some representative catalysts of the support effect for CO2 hydrogenation to light olefins with improved catalyst stability are presented in Table 2.
Owen et al. investigated the effect of Co-Na-Mo on various supports (SiO2, CeO2, ZrO2, γ-Al2O3, TiO2, ZSM-5 (NH4+) and MgO) for CO2 hydrogenation. It was found that the surface area of the support and the metal–support interaction played a key role in the determination of the cobalt crystallite size, which strongly affected the catalytic activity. Cobalt particles with sizes < 2 nm supported on MgO showed low RWGS conversion with negligible FT activity, which is in agreement with the work of de Jong et al. [51]. When the cobalt particle size increased to 15 nm supported on SiO2 and ZSM-5, both the CO2 conversion and C2+ hydrocarbon selectivity increased markedly. When the cobalt particle size further increased to 25–30 nm, a lower CO2 conversion but higher C2+ light olefin selectivity was obtained. The authors reported that the higher the metal–support interaction, the higher the growth chain probability of the hydrocarbons. By altering the TiO2/SiO2 ratio in the support, the CO2 conversion and C2+ light olefin selectivity could be tuned [115]. Li et al. evaluated cobalt catalysts supported on TiO2 with different crystal forms of anatase (a-TiO2) and rutile (r-TiO2), and it was found that the addition of Zr, K, and Cs improved the CO, CO2, and H2 adsorption in both the capacity and strength over a-TiO2- and r-TiO2-supported catalysts. The surface C/H ratio increased drastically in the presence of promoters, leading to a high C2+ selectivity of 17% with 70% CO2 conversion over a K-Zr-Co/a-TiO2 catalyst. As a result, the product distribution could be tuned by adjusting the metal–support interaction and surface C/H ratio through Zr, K, and Cs modification over Co-based catalysts for CO2 hydrogenation, as shown in Scheme 5 [10].
Da Silva et al. found the Fe-Cr catalyst, promoted with K and supported on niobium oxide, was more active (CO2 conversion = 20%) and selective to light olefins (25%) compared to the same composition supported on silica (CO2 conversion = 11%, light olefin selectivity = 18%) under the same testing conditions. Alkali metal promotion increased the selectivity of olefins, probably due to electron-donor effects and the basicity of niobium oxide. A niobium oxide-supported Fe-Cr catalyst presented higher activity and selectivity to olefins, which is probably due to strong metal–support interactions when compared with traditional SiO2 [4]. Very recently, Huang et al. revealed the dynamic evolution of the active Fe and carbon species over different phases of zirconia (m-ZrO2 and t-ZrO2) on CO2 hydrogenation to light olefins, as shown in Scheme 6. Fe-K/m-ZrO2 catalysts performed better than the corresponding Fe-K/t-ZrO2 catalysts under the optimal reaction conditions. Among them, the 15Fe-K/m-ZrO2 catalyst showed remarkable catalytic activity, with a CO2 conversion of 38.8% and a C2–C4= selectivity of 42.8%. More active species (Fe3O4 and χ-Fe5C2) with smaller particle sizes were obtained for the Fe-K/m-ZrO2 catalysts. The larger specific surface area facilitated the highly dispersed Fe species on the surface of the m-ZrO2 support when compared to the t-ZrO2 support. In addition, the monoclinic phase m-ZrO2 support provided more strong basic sites, effectively decreasing the deposited carbon species and coke generation. Moreover, the electron-donating ability of iron elements and more oxygen vacancies (Ov) improved the charge transfer between ZrO2 and Fe. The synergy effect between K2O and ZrO2 fostered the generation of active carbide species. The formation of more χ-Fe5C2 species contributed to the high yield of light olefins [112]. Similarly, Gu et al. investigated Fe-K supported on ZrO2 with different crystal phases, revealing 40.5% CO2 conversion, 15.0% light olefin selectivity, and excellent stability (Figure 14) over 10Fe-1K/m-ZrO2 (10 wt% Fe and 1 wt% K) at 2.0 MPa and 613 K. The CO2 conversion was almost 200% higher than that of 10Fe-1K/t-ZrO2 [114].
Torrente-Murciano demonstrated that iron-based catalysts could be improved not only through the inclusion of promoters but also by the judicious control of the morphology of the ceria support (nanoparticle, nanorods, nanocubes) for CO2 hydrogenation to light olefins. For example, 20 wt% Fe/CeO2 cubes provided better catalytic performance (CO2 conversion = 15.2%, C2–C4= selectivity = 20.2%) when compared with nanorods and their nanoparticle counterparts. TPR showed that the ceria reducibility decreased in the order of rods > particles > cubes, suggesting that the catalytic effect had a direct dependence on the reducibility of the different nanostructured ceria supports and their interaction with the iron particles [113]. By the physical mixing of Fe5C2 and K-modified Al2O3, Liu et al. discovered that Fe5C2-10K/a-Al2O3 exhibited a CO2 conversion of 40.9% and C2+ selectivity of 73.5%, containing 37.3% C2–C4= and 31.1% C5+ (Figure 15). The superior catalytic performance was due to the potassium which migrated into the Fe5C2 during the reaction, and the intimate contact between the Fe5C2 and K/a-Al2O3. Among the various supports tested, as shown in Figure 15, alkaline Al2O3 is the best support for the high selectivity of value-added hydrocarbons [15].
Dai et al. synthesized hierarchical porous carbon monoliths (HPCMs) by an adaptable strategy employing a one-step desilication process for a coke-deposited spent zeolite catalyst. This hierarchical porous carbon was shown to be a better support for the reduction of the nanoparticle size and heightening the synergism of the Fe–K catalyst for CO2 hydrogenation, with a CO2 conversion of 33.4% and a C2=–C4= selectivity of 18.0% [109].
Metal organic frameworks (MOFs) as novel porous materials had a considerable effect on the activity and selectivity of Fe-based catalysts. Hu et al. synthesized a type of hydrothermally stable MOF, zeolitic imidazolate frameworks (ZIF-8) with different sizes and morphologies, which were used as supports for CO2 hydrogenation. The acidity, internal diffusion process and crystal size enabled the ZIF-8 supports to show different levels of substantial light olefin selectivity [110]. Raghav et al. developed a simple method for the synthesis of hierarchical molybdenum carbide (β-Mo2C). The β-Mo2C phase exhibited the strongest metallic and some ionic character, and it behaved as both a support and co-catalyst for CO2 hydrogenation to light olefins. The Fe(0.5)-Mo2C catalyst exhibited a conversion of CO2 of 7.3% and a C2= olefin selectivity of 79.4% at 300 °C and 4.0 mPa. The XRD patterns of the fresh and used Fe(0.5)-Mo2C catalyst did not show a noticeable difference, indicating the stability of the catalysts to achieve high olefin selectivity [111].
In summary, various supports (SiO2, CeO2, m-ZrO2, γ-Al2O3, TiO2, ZSM-5, MgO, NbO HPCMs, MOFs, β-Mo2C) have been used for the dispersal of active species. The surface area, basicity, reducibility, oxygen vacancies, and morphology of the support played important roles, in most cases with the presence of promoters (K, Zr, Cs), in affecting the amount and particle size of the active carbide species; the synergy effect; the metal–support interaction; the strength and capacity of the CO, CO2, and H2 adsorption on support; and the surface C/H ratio for CO2 hydrogenation. By tuning the above-mentioned characteristics properly, the physically deposited carbon species, coke generation and metal sintering could be mitigated as reported.

3.3. Bifunctional Composite Catalyst Effect

The zeolite–methanol composite catalyst can also be improved by compositional modifications. The composite catalyst is composed of two functional components: one is the target for methanol synthesis, mainly Cu, Zn, and In metal oxide catalysts; the other one is for the MTO process, mainly zeolite catalysts. Here, in this section, the recent progress on composite catalysts for improved catalytic performance and stability are described accordingly. Some representative catalysts for the bifunctional composite catalyst effect for CO2 hydrogenation to light olefins with improved catalyst stability are presented in Table 3.
Wang et al. prepared kaolin-supported CuO-ZnO/SAPO-34 catalysts using kaolin as the support and raw material to prepare SAPO-34 molecular sieves. It was found that the resultant SAPO-34 molecular sieves showed a lamellar structure, relatively high crystallinity, and a larger specific surface area, which enabled the good dispersion of CuO-ZnO on the surface of the kaolin, and exposed more active sites for CO2 conversion. The confinement effect of (CuO-ZnO)-kaolin/SAPO-34 catalysts could prevent methanol dissipation, and provided an increased driving force for the conversion of CO2. Furthermore, the lamellar structure of SAPO-34 molecular sieves shortened the diffusion path of the intermediate product, and therefore enhanced the catalytic lifetime [116].
Gao et al. shown a selective hydrogenation process to directly convert CO2 to light olefins via a bifunctional catalyst composed of a methanol synthesis catalyst (In2O3-ZrO2) and a MTO catalyst (SAPO-34) by simple physical mixing. This bifunctional process exhibited an outstanding light olefin (C2–C3=) selectivity of 80–90% with a CO2 conversion of ~20% and superior catalyst stability, running 50 h without obvious deactivation. The excellent catalytic performance was ascribed to the hybrid catalyst that suppressed the usually uncontrollable surface polymerization of CHx in conventional CO2–FTS. This was the highest selectivity reported to date, which dramatically surpassed the value obtained from traditional Fe or Co CO2–FTS catalysts (typically less than 50%) [43].
Similarly, Tan et al. evaluated CO2 conversion to light olefins over an In2O3-ZrO2/SAPO-34 hybrid catalyst. This hybrid catalyst combined a In2O3-ZrO2 component, which would provide the benefit of oxygen vacancy to foster CO2 activation for hydrogenation into methanol, and a SAPO-34 component, to provide sites for the dehydration of the formed methanol into light olefins (Figure 16a). The light olefin selectivity reached 77.6% with less than 5% CO formation, which was ascribed to the strong adsorption of CO2 to defects in the In2O3 and ZrO2 components, creating a large energy barrier that suppressed CO2 dissociation into CO. The weaker acidity from In2O3-ZrO2 suppressed the further hydrogenation of the generated light olefins to paraffins. The catalyst displayed excellent stability, running for 100 h without obvious deactivation (Figure 16b) [44].
Furthermore, Gao et al. discovered that a bifunctional catalyst with an appropriate proximity containing In−Zr oxide, which was responsible for the CO2 activation, and SAPO-34, which was responsible for the selective C−C coupling, could greatly improve the CO2 hydrogenation to lower olefins with excellent selectivity (80%) and high activity (35% CO2 conversion) (Figure 17a). They showed that the incorporation of zirconium significantly improved the catalytic stability by preventing the sintering of the oxide nanoparticles caused by the increase in surface oxygen vacancies. No obvious deactivation was observed over 150 h (Figure 17b) [19].
Wang et al. developed a new catalyst system composed of a Zn0.5Ce0.2Zr1.8O4 solid solution and H-RUB-13 zeolite. This composite exhibited a remarkable C2=–C4= yield as high as 16.1%, with a CO selectivity of only 26.5% due to the hindering of the RWGS reaction. It was demonstrated that methanol was first generated on the Zn0.5Ce0.2Zr1.8O4 solid solution via the formate–methoxyl intermediate mechanism, and was then converted into light olefins on H-RUB-13. By adjusting the H-RUB-13 acidity, the light olefin distribution can be effectively regulated, with propene and butene accounting for 90% of the light olefins [117].
Li et al. proposed a new synthetic strategy to prepare the bifunctional catalysts ZnZrOx/bio-ZSM-5. Hierarchically porous structured bio-ZSM-5 was prepared by using a natural rice husk as a template, which was then integrated with the ZnZrOx solid solution nanoparticles by physical mixing. The derived bifunctional catalysts ZnZrOx and bio-ZSM-5 exhibited superior light olefin selectivity and stability due to their unique pore structure, which was advantageous for mass transport and coke formation inhibition. *CHxO was identified to be the key intermediate formed on the ZnZrOx surface, and was transferred to the Brønsted acid sites in the bio-ZSM-5 for the subsequent conversion to light olefins. The addition of a Si promoter to the ZnZrOx/bio-ZSM-5 catalyst prominently enhanced the light olefin selectivity. The ZnZrOx/bio-ZSM-5−Si catalyst exhibited an outstanding light olefin selectivity of 64.4%, with a CO2 conversion of 10% and an excellent stability without noticeable deactivation during 60 h on stream (Figure 18a). In addition, the proximity of the catalyst components plays a key role in light olefin selectivity. As seen in Figure 18b, increasing the proximity resulted in a greater olefin selectivity [118]. By incorporating proper amounts of Ce or Cr ions into indium oxides, the methanol selectivity is increased, along with a reduction in the CH4 amount, as shown in Figure 19. Upon complexing with SAPO-34, a CO2 conversion of 33.6% and a C2=–C4= selectivity of 75.0% were achieved over InCrOx(0.13)/SAPO-34, which was about 1.5–2.0 times those obtained on In2O3/SAPO-34 and In–Zr/SAPO-34. This is because the incorporation of Ce or Cr ions into In2O3 lattice sites promoted the generation of more surface oxygen vacancies, as shown in Figure 19a, and enhanced the electronic interaction of HCOO* with InCeOx(0.13) and InCrOx(0.13) surfaces, which decreased the free energy barrier and enthalpy barrier for the formation of HCOO* and CH3OH. The composite catalysts also displayed excellent stability after 120 h on stream (Figure 19b) [119].
Similarly, Li et al. developed a bifunctional composite catalyst ZnZrO/SAPO-34 containing a ZnOZrO2 component to activate CO2 and H2 to form methanol, and a SAPO-34 component to perform C–C bond formation for the conversion of the produced methanol to light olefins. The derived dual function tandem catalyst exhibited an outstanding light olefin selectivity of 80% with good stability, and a CO2 conversion of 12.6% (Figure 20a,b). The kinetic and thermodynamic coupling between the tandem reactions enabled the highly efficient conversion of CO2 to lower olefins through the transfer and migration of CHxO intermediate species [13].
Dang et al. advanced a series of dual function tandem catalysts containing In2O3-ZnZrOx oxides and various SAPO-34 zeolites with varying crystal sizes (0.4–1.5 mm) and pore structures. It was found that decreasing the crystal size of SAPO-34 could shorten the diffusion path from the surface to the acid sites inside the zeolite pores, thus favoring the mass transfer of intermediate species for efficient C–C coupling to produce lower olefins and enhance the selectivity of C2=–C4=. Interestingly, further HNO3 post-treatment caused the formation of the SAPO-34 zeolites with a hierarchical structure comprised of micro-/meso-/macropores, and reduced the amount of the Brønsted acid sites, both of which led to a significant increase in the catalytic performance, with the C2=–C4= selectivity reaching as high as 85% among all of the hydrocarbons (Figure 21a), a very low CH4 selectivity of only 1%, and an O/P ratio of 7.7 at a CO2 conversion of 17%. The C2=–C4= selectivity is much higher than the maximum predicted by the Anderson–Schulz–Flory distribution over modified FTS catalysts. The composite catalysts also exhibited excellent stability after 90 h on stream (Figure 21b) [42].
Liu et al. synthesized bifunctional composite catalysts composed of a spinel binary metal oxide ZnAl2O4/ZnGa2O4 and SAPO-34, with the selectivity of C2–C4 olefins reaching 87% at CO2 conversions of 15%. This study revealed that the oxygen vacancy site on metal oxides played a crucial role in the adsorption and activation of CO2, while the -Zn-O- domain accounted for H2 activation. It was demonstrated that the methanol reaction intermediates formed on the metal oxide, then converted to lower olefins at the Brønsted acid sites in SAPO-34 zeolite [121]. Tong et al. developed a dual-function composite catalyst, 13%ZnO-ZrO2/Mn0.1SAPO-34, and attained a high CO2 conversion of 21.3% with a light olefin selectivity of 61.7%, and suppressed the selectivity of CO below 43% and the CH4 selectivity below 4%. The fine-tuned acidity of zeolite by the addition of Mn and the granule stacking arrangement contributed to the excellent catalytic performance. Mn was embedded into the zeolite ionic structure to tune the acidity of the molecular sieve and limit secondary hydrogenation reactions. The granule stacking arrangement facilitated the tandem catalysis [122]. Dang et al. presented a series of bifunctional catalysts containing In-Zr composite oxides with different In/Zr atomic ratios and SAPO-34 zeolite for CO2 conversion to light olefins. It was demonstrated that the inclusion of a certain amount of ZrO2 could provide more oxygen vacancy sites (Figure 22a), stabilize the intermediates in the CO2 hydrogenation, and prevent the sintering of the active nanoparticles. This, in turn, would lead to significantly enhanced catalytic activity, selectivity of hydrocarbons and stability for direct CO2 hydrogenation to lower olefins at the relatively high reaction temperature of 653K. A light olefin selectivity as high as 80% at a CO2 conversion rate of 27% and less than 2.5% methane selectivity was obtained over the optimized indium-zirconium/SAPO-34 bifunctional catalyst. The catalyst exhibited excellent stability for over 140 h without showing obvious deactivation (Figure 22b) [18].
In summary, the majority of the catalysts tested for CO2 hydrogenation to light olefins via the MeOH-mediated route involve two active components (metal oxides and zeolite), which are so-called bifunctional composite catalysts. In this section, multiple variations (acidity, particle size, proximity, oxygen vacancy) in the combination of methanol synthesis catalysts (Cu, Zn, In, Ce, Zr, etc. metal oxides) with various zeolites (SAPO-34 and ZSM5) have been reported to give improved olefin selectivity and catalyst stability by mitigating coke formation, reducing the particle size growth of active carbide species, and inhibiting inactive species formation for CO2 hydrogenation to light olefins.

3.4. Structure Effect

The structure of the catalysts plays an important role in converting CO2 to light olefins. In this section, we will report the recent progress on the ways in which morphology changes in both the Fe-based and methanol zeolite composite catalysts can improve the catalytic performance. Some representative catalysts on the structure effect for CO2 hydrogenation to light olefins with improved catalyst stability are presented in Table 4.
Wang et al. developed a layered metal oxides (LMO) structure, K-Fe-Ti, that displayed high catalytic activity, olefin selectivity and decent stability toward CO2–FTS. The light olefin selectivity achieved approximately 60% with an olefin/paraffin ratio of 7.3 over the catalyst 0.8K-2.4Fe-1.3Ti (Figure 23). The LMO structure exfoliated through the acid treatment was found to weaken the interaction between Fe and Ti, which made it easier for the reduction and activation of iron oxides to form active iron carbide species that favored a shift from the RWGS to the FTS reaction. Meantime, C2H4 adsorption was hindered due to the low surface area of the LMO structure, contributing to higher olefin selectivity by inhibiting the secondary hydrogenation of primary olefins. The acid treatment played a key role in the formation of a slice structure that favored CO2 conversion to light olefins with lower CO selectivity [123]. Fujiwara et al. found the composite catalysts obtained from the simple mixing of Cu–Zn–Al oxide together with HB zeolite, which was modified with 1,4-bis(hydroxydimethylsilyl) benzene, to be very effective for CO2 hydrogenation to C2+ hydrocarbons. The modification of zeolite with the disilane compound made the catalysts’ surface hydrophobic, a characteristic which was effective in preventing catalyst deactivation by the formation of water during CO2 hydrogenation. The highest yield of C2+ hydrocarbons over the modified composite catalysts reached about 12.6 C-mol% at 573 K under a pressure of 0.98 Mpa. The diminishing of the deactivation of the strong acid sites of HB zeolite with the hydrophobic surface is the source of the enhanced catalytic activity [129].
Liu et al. synthesized a unique structure with ZnO and nitrogen-doped carbon (NC)-overcoated Fe-based catalysts (Fe@NC) (Figure 24), and found that the reaction rate increased by ~25%, while the O/P ratio increased from 0.07 to 1.68 when compared with the benchmark Fe3O4 catalyst. The inactive θ-Fe3C phase disappeared, and the active phases (Fe3O4 and Fe5C2) formed for CO2 hydrogenation. The introduction of NC to the surface of the Fe catalysts significantly boosted the catalyst activity, the selectivity toward light olefins, and the stability due to the enhanced metal–support-reactant interaction and interfacial charge transfer [36].
Numpilai et al. studied the hydrogenation of CO2 to light olefins over Fe-Co/K-Al2O3 catalysts, and discovered that the pore sizes of the Al2O3 support had profound effects on the Fe2O3 crystallite size, the reducibility, the adsorption–desorption of CO2 and H2, and the catalytic performances. The highest olefins to paraffins ratio of 6.82 was obtained from the largest pore catalyst (CL-Al2O3) due to the suppression of the hydrogenation of olefins to paraffins by increasing the pore sizes of Al2O3 to eliminate diffusion limitation. The maximum light olefin yield of 14.38% was obtained over the catalyst with an appropriated Al2O3 pore size (49.7 nm) owing to the suppression of the olefins’ hydrogenation and chain growth reaction [37].
The electrospun ceramic K/Fe-Al-O nanobelt catalysts synthesized by Elishav et al. showed a much higher CO2 conversion of 48%, a C2-C5 olefin selectivity of 52%, and a high olefin/paraffin ratio of 10.4, while the K/Fe-Al-O spinel powder catalyst produced mainly C6+ hydrocarbons. The enhanced olefin selectivity of the electrospun materials is related to a high degree of reduction of the surface Fe atoms due to the more efficient interaction with the K promoter [124].
A defect-rich MgH2/CuxO hydrogen storage composite might inspire the catalysts’ design for the hydrogenation of CO2 to lower olefins. Chen et al. presented a defect-rich MgH2/CuxO composite catalyst that achieved a C2=–C4= selectivity of 54.8% and a CO2 conversion of 20.7% at 623 K under a low H2/CO2 ratio of 1:5. It is the defective structure of MgH2/CuxO that promotes CO2 molecule adsorption and activation, while the electronic structure of MgH2 was more conducive to the provision of lattice H for the hydrogenation of the CO2 molecule. The lattice H could combine with the C site of the CO2 molecule to promote the formation of Mg formate, which was further hydrogenated to lower olefins under a low H concentration [125]. The same group reported carbon-confined MgH2 nano-lamellae which stored solid hydrogen for the hydrogenation of CO2 to lower olefins and demonstrated a high selectivity under low H2/CO2 ratios. The high selectivity of lower olefins was attributed to the low concentration of solid hydrogen under low H2/CO2 ratios that suppressed the further hydrogenation of light olefins from Mg formate [128].
SAPO-34 molecular sieves were considered to be the best catalysts due to their excellent structure selectivity, suitable acidity, favorable thermal stability, and hydrothermal stability, as well as their high selectivity for light olefins. Tian et al. used Palygorskite as a silicon and partial aluminum source, and DEA, TEA, MOR and TEAOH as template agents to prepare SAPO-34 molecular sieves with higher purity. Composite catalysts of CuO-ZnO-Al2O3/SAPO-34 were prepared by mechanically mixing SAPO-34 molecular sieves with CuO-ZnO-Al2O3 (CZA), and a superb CO2 conversion of 53.5%, a light olefin selectivity of 62.1% and a yield of 33.2% were obtained over the CZA/SAPO-34(TEAOH)HCl composite catalyst [126]. CO2 conversion and product distribution are strongly dependent on the oxide composition and structure. Li et al. developed a bifunctional catalyst composed of ZnO-Y2O3 oxide and SAPO-34 zeolite that offered a CO2 conversion of 27.6% and a light olefin selectivity of 83.6% [40].
Some Fe-containing catalysts can also be improved by creating unique architectures. Wei et al. created Fe-based catalysts with honeycomb-structured graphene (HSG) as the catalyst support and K as the promoter, and achieved the 59% selectivity of light olefins over a FeK1.5/HSG catalyst. No obvious deactivation was observed within 120 h on stream (Figure 25). The excellent catalytic performance was ascribed to the confinement effect of HSG and the K promotion effect on the activation of inert CO2 and the formation of iron carbide. The complex three-dimensional (3D) architecture of the porous HSG effectively impeded the sintering of the active sites’ iron carbide nanoparticles (NPs). Meanwhile, CO2 and H2 could more easily permeate the mesoporous–macroporous framework of HSG and access the catalysts’ active sites. Similarly, the generated light olefins could more easily emerge from the catalyst so as to avoid further unwanted hydrogenation [127].
Consequently, multiple reports indicate that the modification of the morphology of zeolite–methanol synthesis composites by creating core–shell configurations can have a beneficial effect [16,120,130]. For example, dual-function composite catalysts containing CuZnZr (CZZ) and SAPO-34 were synthesized by Chen et al. for the tandem reactions of CO2 to methanol and methanol to olefins. The assembled core–shell CZZ@SAPO-34 catalyst, as shown in Figure 26, exhibited an enhanced light olefin selectivity of 72% and inhibited CH4 formation due to reduced contact interface between CZZ and SAPO-34 and weakened hydrogenation ability at the metal sites. Furthermore, the addition of Zn reduced the acidity of SAPO-34; as a result, the secondary reactions of the primary olefins were significantly diminished (Figure 26) [120].
In summary of this section, the structure and the properties associated with the structure of the catalysts are pivotal for CO2 hydrogenation to light olefins. The low surface area of the LMO structure could hinder the C2H4 secondary reaction, contributing to higher olefin selectivity. The surface modification of zeolite from hydrophilic to hydrophobic could prevent the catalyst deactivation caused by the formation of water. The unique structure of Fe@NC enables phase transformation from the inactive (θ-Fe3C) phase to active species (Fe3O4 and Fe5C2). Increasing the pore sizes of Al2O3 could eliminate the diffusion limitation for CO2 and H2. The electrospun ceramic K/Fe-Al-O nanobelt catalysts led to a high degree of the reduction of surface iron atoms. The defective structure of MgH2/CuxO and carbon-confined MgH2/C nano-lamellae could promote CO2 adsorption and activation, with the electronic structure of MgH2 offering lattice H for CO2 hydrogenation. The 3D architecture of the porous HSG could impede the sintering of the active sites’ iron carbide NPs. The confinement of core–shell CZZ@SAPO-34 structure could increase the access frequency of the methanol intermediate to the active zeolite sites, consequently improving the light olefine selectivity.

4. Conclusions

There is an urgent need to control CO2 emissions in order to mitigate their negative impact on the environment. The catalytic conversion of CO2 is an encouraging approach to mitigate CO2 emissions by producing chemicals and fuels. A highly promising route is selective CO2 hydrogenation to produce light olefins. The huge market demand for the lower olefins offers a great opportunity for the target technology to profoundly impact the scale of CO2 utilization once it is developed with renewable hydrogen. Currently, there are two primary pathways (the CO2−FTS route and the MeOH-mediated route) to produce light olefins from CO2 hydrogenation in a one-step process. In the CO2–FTS path, Fe is one of the most widely used components, while in the MeOH path, Cu/zeolite has been used the most. Even though significant efforts have been made, considerable challenges remain in the development of highly efficient catalysts with selective pathways to light olefins due to the thermodynamically stable nature of the CO2 molecule, the complexity of the reaction networks, and catalyst deactivation. During CO2 hydrogenation, the primary causes for catalyst deactivation are the sintering (or agglomeration) of metal particles, phase transformation at the catalyst’s surface, and catalyst poisoning by water or carbonaceous deposits (i.e., coke). A firm grasp of the causes for deactivation is essential in order to develop a mitigation strategy and sustain a high selectivity toward the desired olefins during CO2 hydrogenation. In this review, we summarized the reports published within five years on the effect of the promotors, metal oxide support, bifunctional composites and structure on the catalyst design in order to minimize catalyst deactivation.
Promoter effect: Alkali metals such as K and Na have been broadly used as promotors to control the electronic properties. Mn, Ce, and Ca metals have been used as structural promotors. Transition metals such as Zn, Co, Cu, V, Zr, etc., have been used as both electronic and structural promotors. With the inclusion of alkali promoters, Fe-based catalysts can possess higher olefin selectivity. The alkali metals act as electron donors to Fe metal centers, fostering CO2 adsorption while decreasing their affinity with H2, and consequently leading to a higher olefin yield. Some studies show that doping the catalyst with a second metal improves the olefin yield by forming a highly active interface. The second metal promoters may provide a way to tune the CO2 and H2 adsorption and activation, shifting the product distribution towards the desired hydrocarbons.
Support effect: Supporting the Fe-based species on supports such as SiO2, CeO2, m-ZrO2, γ-Al2O3, TiO2, ZSM-5, MgO, NbO HPCMs, MOFs, and β-Mo2C may enhance the catalytic performance by improving the active metal dispersion and retarding the sintering of the active particles. The surface area, basicity, reducibility, oxygen vacancies, and morphology of the support played important roles—in most cases, with the presence of promoters (K, Zr, Cs)—in affecting the amount and particle size of the active carbide species; the synergy effect; the metal–support interaction; the strength and capacity of CO, CO2, and H2 adsorption on support; and the surface C/H ratio for CO2 hydrogenation. By tuning the above-mentioned characteristics properly, the physically deposited carbon species, coke generation and metal sintering could be mitigated.
Bifunctional composite catalyst effect: The catalysts tested for CO2 hydrogenation to light olefins via the MeOH-mediated route mainly involve two active components (metal oxides and zeolite), and so are called bifunctional composite catalysts. In this review, multiple variations (acidity, particle size, proximity, oxygen vacancy) of the combination of methanol synthesis catalysts (Cu, Zn, In, Ce, Zr, etc. metal oxides) with various zeolites (SAPO-34 and ZSM5) have been reported for enhanced olefin selectivity and catalyst stability by mitigating coke formation, reducing the particle size growth of active carbide species, and inhibiting inactive species formation for CO2 hydrogenation to light olefins.
Structure effect: The structure of the catalysts plays a pivotal role in CO2 hydrogenation to light olefins. The structures and properties (for example, LMO, the surface modification of zeolite from hydrophilic to hydrophobic, Fe@NC, the pore sizes of Al2O3, the defective structure of MgH2/CuxO and carbon-confined MgH2/C nano-lamellae, the 3D architecture of the porous HSG, and core–shell CZZ@SAPO-34) could be tuned to mitigate catalyst deactivation by retarding the sintering of active species and coke deposition, tolerating water formation and enabling favorable phase transformation for an enhanced light olefin yield and catalyst stability.
Despite the many advances made in catalytic development, especially with light olefin yield and stability, a novel catalytic system that is both economically viable and resistant to deactivation has not yet been achieved. Most research efforts have focused on the development of catalytic materials and the adjustment of properties and metal interactions for the desired catalyst activity and long-term stability. Future research directions for CO2 hydrogenation should consider: (1) the further modification of the catalytic surface H/C molar ratio and the fostering of C-C coupling; (2) tuning the basicity and oxygen vacancies of the catalyst support to facilitate the CO2 adsorption and activation; (3) examining more novel catalytic materials/structures to boost the catalyst stability; and (4) exploring more energy-saving catalysts for CO2 hydrogenation to light olefins. In addition, in situ measurements using synchrotron-based techniques, such as X-ray adsorption spectroscopy (XAS), should be performed in order to understand the ways in which the local environment of the catalysts affects their activity, stability and efficient mitigation.

Author Contributions

Conceptualization, C.Z.; writing—original draft, C.Z., N.J.R., T.H., D.W., M.W., C.M., A.Z. and N.F.; writing—review and editing, C.Z. and N.J.R.; funding acquisition, C.Z.; resources, C.Z.; supervision, C.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Science Foundation under Grant No. 1955521 (C.Z.).

Acknowledgments

The authors are grateful for the U.S. Department of Energy, the Office of Science, and the Office of Workforce Development for Teachers and Scientists under the Science Undergraduate Laboratory Internships Program (T.H. and A.Z.) and Visiting Faculty Program (C.Z.).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

FTSFisch–Tropsch Synthesis
MeOHmethanol
RWGSreverse water–gas shift
MTOMeOH-to-olefins
ASFAnderson–Schulz–Flory
CTOCO2 to olefins
XRDX-ray powder diffraction
PFRplug-flow
CSTRfully back-mixed reactors
TPSRTemperature-programed surface reaction
XPSX-ray photoelectron spectroscopy
SEMscanning electron microscopy
TEMtransmission electron microscopy
HAADF−STEMhigh-angle annular dark-field-scanning transmission electron microscopy
O/P ratioolefins/paraffin ratio
NPsnanoparticles
CNTcarbon nanotubes
FTYFe time yield
STYspace–time yields
HPCMshierarchical porous carbon monoliths
LMOlayered metal oxides
MOFmetal organic framework

References

  1. Zhang, X.; Zhang, A.; Jiang, X.; Zhu, J.; Liu, J.; Li, J.; Zhang, G.; Song, C.; Guo, X. Utilization of CO2 for aro-matics production over ZnO/ZrO2-ZSM-5 tandem catalyst. J. CO2 Util. 2019, 29, 140–145. [Google Scholar] [CrossRef]
  2. Atmospheric CO2 Levels Defy the Pandemic to Hit Record High. Available online: https://newatlas.com/environment/atmospheric-co2-pandemic-record-concentrations/ (accessed on 15 October 2021).
  3. Monastersky, R. Global carbon dioxide levels near worrisome milestone. Nature 2013, 497, 13–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Da Silva, I.A.; Mota, C. Conversion of CO2 to light olefins over iron-based catalysts supported on niobium oxide. Front. Energy Res. 2019, 7, 49. [Google Scholar] [CrossRef] [Green Version]
  5. Keith, D.; Holmes, G.; Angelo, D.S.; Heidel, K. A process for capturing CO2 from the atmosphere. Joule 2018, 2, 1573–1594. [Google Scholar] [CrossRef] [Green Version]
  6. Centi, G.; Quadrelli, E.A.; Perathoner, S. Catalysis for CO2 conversion: A key technology for rapid introduc-tion of renewable energy in the value chain of chemical industries. Energy Environ. Sci. 2013, 6, 1711. [Google Scholar] [CrossRef]
  7. Dutta, A.; Farooq, S.; Karimi, I.A.; Khan, S.A. Assessing the potential of CO2 utilization with an integrated framework for producing power and chemicals. J. CO2 Util. 2017, 19, 49–57. [Google Scholar] [CrossRef]
  8. Ma, Z.; Porosoff, M. Development of tandem catalysts for CO2 hydrogenation to olefins. ACS Catal. 2019, 9, 2639–2656. [Google Scholar] [CrossRef]
  9. Science Daily. Available online: https://www.sciencedaily.com/releases/2018/11/181108130533.htm (accessed on 15 August 2021).
  10. Li, W.; Zhang, G.; Jiang, X.; Liu, Y.; Zhu, J.; Ding, F.; Liu, Z.; Guo, X.; Song, C. CO2 hydrogenation on un-promoted and M-promoted Co/TiO2 catalysts (M = Zr, K, Cs): Effects of crystal phase of supports and met-al−support interaction on tuning product distribution. ACS Catal. 2019, 9, 2739–2751. [Google Scholar] [CrossRef]
  11. Wei, J.; Yao, R.W.; Ge, Q.J.; Wen, Z.Y.; Ji, X.W.; Fang, C.Y.; Zhang, J.X.; Xu, H.Y.; Sun, J. Catalytic hydrogena-tion of CO2 to isoparaffins over Fe-based multifunctional catalysts. ACS Catal. 2018, 8, 9958–9967. [Google Scholar] [CrossRef]
  12. Wei, J.; Ge, Q.; Yao, R.; Wen, Z.; Fang, C.; Guo, L.; Xu, H.; Sun, J. Directly converting CO2 into a gaso-line fuel. Nat. Commun. 2018, 8, 15174. [Google Scholar] [CrossRef] [Green Version]
  13. Li, Z.; Wang, J.; Qu, Y.; Liu, H.; Tang, C.; Miao, S.; Feng, Z.; An, H.; Li, C. Highly selective con-version of carbon dioxide to lower olefins. ACS Catal. 2017, 7, 8544–8548. [Google Scholar] [CrossRef]
  14. Li, Z.L.; Qu, Y.Z.; Wang, J.J.; Liu, H.L.; Li, M.R.; Miao, S.; Li, C. Highly selective conversion of carbon dioxide to aromatics over tandem catalysts. Joule 2019, 3, 570–583. [Google Scholar] [CrossRef] [Green Version]
  15. Liu, J.H.; Zhang, A.F.; Jiang, X.; Liu, M.; Zhu, J.; Song, C.S.; Guo, X.W. Direct transformation of carbon dioxide to value-added hydrocarbons by physical mixtures of Fe5C2 and K-modified Al2O3. Ind. Eng. Chem. Res. 2018, 57, 9120–9126. [Google Scholar] [CrossRef]
  16. Xie, C.L.; Chen, C.; Yu, Y.; Su, J.; Li, Y.F.; Somorjai, G.A.; Yang, P.D. Tandem catalysis for CO2 hydrogenation to C2−C4 hydrocarbons. Nano Lett. 2017, 17, 3798−3802. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Liu, M.; Yi, Y.H.; Wang, L.; Guo, H.C.; Bogaerts, A. Hydrogenation of carbon dioxide to value-added chemi-cals by heterogeneous catalysis and plasma catalysis. Catalysts 2019, 9, 275. [Google Scholar] [CrossRef] [Green Version]
  18. Dang, S.S.; Gao, P.; Liu, Z.Y.; Chen, X.Q.; Yang, C.G.; Wang, H.; Zhong, L.S.; Li, S.G.; Sun, Y.H. Role of zirconium in direct CO2 hydrogenation to lower olefins on oxide/zeolite bifunctional catalysts. J. Catal. 2018, 364, 382–393. [Google Scholar] [CrossRef]
  19. Gao, P.; Dang, S.S.; Li, S.G.; Bu, X.N.; Liu, Z.Y.; Qiu, M.H.; Yang, C.G.; Wang, H.; Zhong, L.S.; Han, Y.; et al. Direct production of lower olefins from CO2 conversion via bifunctional catalysis. ACS Catal. 2018, 8, 571–578. [Google Scholar] [CrossRef]
  20. Saeidi, S.; Najari, S.; Hessel, V.; Wilson, K.; Keil, F.J.; Concepción, P.; Suib, S.L.; Rodrigues, A.E. Recent ad-vances in CO2 hydrogenation to value-added products-Current challenges and future directions. Prog. Energy Combust. Sci. 2021, 85, 100905. [Google Scholar] [CrossRef]
  21. Gnanamani, M.K.; Jacobs, G.; Hamdeh, H.H.; Shafer, W.D.; Liu, F.; Hopps, S.D.; Thomas, G.A.; Davis, B.H. Hydrogenation of carbon dioxide over Co−Fe bimetallic catalysts. ACS Catal. 2016, 6, 913–927. [Google Scholar] [CrossRef]
  22. Ren, T.; Patel, M.; Blok, K. Olefins from conventional and heavy feedstocks: Energy use in steam cracking and alternative processes. Energy 2006, 31, 425–451. [Google Scholar] [CrossRef] [Green Version]
  23. Amghizar, I.; Vandewalle, L.A.; Geem, K.M.V.; Marin, G.B. New trends in olefin production. Engineering 2017, 3, 171–178. [Google Scholar] [CrossRef]
  24. One, O.; Niziolek, A.M.; Floudas, C.A. Optimal production of light olefins from natural gas via the methanol intermediate. Ind. Eng. Chem. Res. 2016, 5511, 3043–3063. [Google Scholar]
  25. Li, W.H.; Wang, H.Z.; Jiang, X.; Zhu, J.; Liu, Z.M.; Guo, X.W.; Song, C.S. A short review of recent advances in CO2 hydrogenation to hydrocarbons over heterogeneous catalysts. RCS Adv. 2018, 8, 7651–7669. [Google Scholar] [CrossRef] [Green Version]
  26. Porosoff, M.; Yan, B.; Chen, J. Catalytic reduction of CO2 by H2 for synthesis of CO, methanol and hydrocarbons: Challenges and opportunities. Energy Environ. Sci. 2016, 9, 62–73. [Google Scholar] [CrossRef]
  27. Wang, D.; Xie, Z.; Porosoff, M.D.; Chen, J. Recent advances in carbon dioxide hydrogenation to produce ole-fins and aromatics. Chem 2021, 7, 1–35. [Google Scholar] [CrossRef]
  28. Zhang, C.; Cao, C.; Zhang, Y.; Liu, X.; Xu, J.; Zhu, M.; Tu, W.; Han, Y. Unraveling the role of zinc on bimetal-lic Fe5C2−ZnO catalysts for highly selective carbon dioxide hydrogenation to high carbon α-olefins. ACS Catal. 2021, 11, 2121–2133. [Google Scholar] [CrossRef]
  29. Wang, X.; Zhang, J.; Chen, J.; Ma, Q.; Fan, S.; Zhao, T.S. Effect of preparation methods on the structure and catalytic performance of Fe–Zn/K catalysts for CO2 hydrogenation to light olefins. Chin. J. Chem. Eng. 2018, 26, 761–767. [Google Scholar] [CrossRef]
  30. Numpilai, T.; Cheng, C.K.; Limtrakul, J.; Witoon, T. Recent advances in light olefins production from cata-lytic hydrogenation of carbon dioxide. Proc. Saf. Environ. Prot. 2021, 151, 401–427. [Google Scholar] [CrossRef]
  31. Wei, J.; Sun, J.; Wen, Z.; Fang, C.; Ge, Q.; Xu, H. New insights into the effect of sodium on Fe3O4-based nano-catalysts for CO2 hydrogenation to light olefins. Catal. Sci. Technol. 2016, 6, 4786. [Google Scholar] [CrossRef]
  32. Wezendonk, T.A.; Sun, X.; Duglan, A.I.; van Hoof, A.J.F.; Hensen, E.J.M.; Kapteijn, F.; Gascon, J. Controlled formation of iron carbides and their performance in Fischer-Tropsch synthesis. J. Catal. 2018, 362, 106–117. [Google Scholar] [CrossRef]
  33. Chaipraditgul, N.; Numpilai, T.; Cheng, C.K.; 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]
  34. Gnanamani, M.K.; Jacobs, G.; Hamdeh, H.H.; Shafer, W.D.; Liu, F.; Hopps, S.D.; Thomas, G.A.; Davis, B.H. Hydrogenation of carbon dioxide over iron carbide prepared from alkali metal promoted iron oxalate. Appl. Catal. A Gen. 2018, 564, 243–249. [Google Scholar] [CrossRef]
  35. Guo, L.; Sun, J.; Ge, Q.; Tsubaki, N. Recent advances in direct catalytic hydro-genation of carbon dioxide to valuable C2+ hydrocarbons. J. Mater. Chem. A 2018, 6, 23244–23262. [Google Scholar] [CrossRef]
  36. Liu, J.; Zhang, A.; Jiang, X.; Zhang, G.; Sun, Y.; Liu, M.; Ding, F.; Song, C.; Guo, X. Overcoating the surface of Fe-based catalyst with ZnO and nitrogen-doped carbon toward high selectivity of light olefins in CO2 Hy-drogenation. Ind. Eng. Chem. Res. 2019, 58, 4017–4023. [Google Scholar] [CrossRef]
  37. Numpilai, T.; Chanel, N.; Poo-Arporn, Y.; Cheng, C.; Siri-Nguan, N.; Sornchamni, T.; Chareonpanich, M.; Kongkachuichay, P.; Yigit, N.; Rupprechter, G.; et al. Pore size effects on physicochemical properties of Fe-Co/K-Al2O3 catalysts and their catalytic activity in CO2 hydrogenation to light olefins. Appl. Surf. Sci. 2019, 483, 581–592. [Google Scholar] [CrossRef]
  38. Yang, H.; Zhang, C.; Gao, P.; Wang, H.; Li, X.; Zhong, L.; Wei, W.; Sun, Y. A review of the catalytic hydro-genation of carbon dioxide into value-added hydrocarbons. Catal. Sci. Technol. 2017, 7, 4580. [Google Scholar] [CrossRef]
  39. Jiao, F.; Li, J.; Pan, X.; Xiao, J.; Li, H.; Ma, H.; Wei, M.; Pan, Y.; Zhou, Z.; Li, M.; et al. Selective conversion of syngas to light olefins. Science 2016, 351, 1065–1068. [Google Scholar] [CrossRef] [PubMed]
  40. Li, J.; Yu, T.; Miao, D.; Pan, X.; Bao, X. Carbon dioxide hydrogenation to light olefins over ZnO/Y2O3 and SAPO-34 bifunctional catalysts. Catal. Commun. 2019, 129, 105711. [Google Scholar] [CrossRef]
  41. Liu, X.; Wang, M.; Zhou, C.; Zhou, W.; Cheng, K.; Kang, J.; Zhang, Q.; Deng, W.; Wang, Y. Selective transfor-mation of carbon dioxide into lower olefins with a bifunctional catalyst composed of ZnGa2O4 and SAPO-34. Chem. Commun. 2018, 54, 140–143. [Google Scholar] [CrossRef]
  42. Dang, S.; Li, S.; Yang, C.; Chen, X.; Li, X.; Zhong, L.; Gao, P.; Sun, Y. Selective transformation of CO2 and H2 into lower olefins over In2O3-ZnZrOx/SAPO-34 bifunctional catalysts. ChemSusChem 2019, 12, 3582–3591. [Google Scholar] [CrossRef]
  43. Gao, J.; Jia, C.; Liu, B. Direct and selective hydrogenation of CO2 to ethylene and propene by bifunctional catalysts. Catal. Sci. Technol. 2017, 7, 5602–5607. [Google Scholar] [CrossRef]
  44. Tan, L.; Zhang, P.; Cui, Y.; Suzuki, Y.; Li, H.; Guo, L.; Yang, G.; Tsubaki, N. Direct CO2 hydrogenation to light olefins by suppressing CO by-product formation. Fuel Process. Technol. 2019, 196, 106174–106178. [Google Scholar] [CrossRef]
  45. Porosoff, M.D.; Kattel, S.; Li, W.; Liu, P.; Chen, J.G. Identifying trends and descriptors for selective CO2 conversion to CO over transition metal carbides. Chem. Commun. 2015, 51, 6988–6991. [Google Scholar] [CrossRef] [PubMed]
  46. Porosoff, M.D.; Chen, J.G. Trends in the catalytic reduction of CO2 by hydrogen over supported monometal-lic and bimetallic catalysts. J. Catal. 2013, 301, 30–37. [Google Scholar] [CrossRef]
  47. Butt, J.B.; Petersen, E. Activation, Deactivation and Poisoning of Catalysts; Academic Press: Cambridge, MA, USA, 1988. [Google Scholar]
  48. Bartholomew, C. Mechanisms of catalyst deactivation. Appl. Catal. A Gen. 2001, 212, 17–60. [Google Scholar] [CrossRef]
  49. Moulijn, J.A.; van Diepen, A.E.; Kapteijn, F. Catalyst deactivation: Is it predictable? What to do? Appl. Catal. A Gen. 2001, 212, 3–16. [Google Scholar] [CrossRef]
  50. Price, C.A.H.; Reina, T.R.; Liu, J. Engineering heterogenous catalysts for chemical CO2 utilization: Lessons from thermal catalysis and advantages of yolk@shell structured nanoreactors. J. Energy Chem. 2021, 57, 304–324. [Google Scholar] [CrossRef]
  51. Kliewer, C.E.; Soled, S.L.; Kiss, G. Morphological transformations during Fischer-Tropsch synthesis on a titania-supported cobalt catalyst. Catal. Today 2019, 323, 233–256. [Google Scholar] [CrossRef]
  52. Eschemann, T.O.; de Jong, K.P. Deactivation behavior of Co/TiO2 catalysts during Fischer–Tropsch synthesis. ACS Catal. 2015, 5, 3181–3188. [Google Scholar] [CrossRef]
  53. Sun, J.T.; Metcalfe, I.S.; Sahibzada, M. Deactivation of Cu/ZnO/Al2O3 methanol synthesis catalyst by sintering. Ind. Eng. Chem. Res. 1999, 38, 3868–3872. [Google Scholar] [CrossRef]
  54. Argyle, M.; Bartholomew, C. Heterogeneous catalyst deactivation and regeneration: A review. Catalysts 2015, 5, 145–269. [Google Scholar] [CrossRef] [Green Version]
  55. Hansen, T.W.; Delariva, A.T.; Challa, S.R.; Datye, A.K. Sintering of catalytic nanoparticles: Particle migration or Ostwald ripening. Acc. Chem. Res. 2013, 46, 1720–1730. [Google Scholar] [CrossRef] [PubMed]
  56. Li, W.; Zhang, A.; Jiang, X.; Janik, M.J.; Qiu, J.; Liu, Z.; Guo, X.; Song, C. The anti-sintering catalysts: Fe–Co–Zr polymetallic fibers for CO2 hydrogenation to C2 =–C4 =—Rich hydrocarbons. J. CO2 Util. 2018, 23, 219–225. [Google Scholar] [CrossRef]
  57. Lee, S.-C.; Kim, J.-S.; Shin, W.C.; Choi, M.-J.; Choung, S.-J. Catalyst deactivation during hydrogenation of carbon dioxide: Effect of catalyst position in the packed bed reactor. J. Mol. Catal. A Chem. 2009, 301, 98–105. [Google Scholar] [CrossRef]
  58. Riedel, T.; Schulz, H.; Schaub, G.; Jun, K.W.; Hwang, J.S.; Lee, K.W. Fischer–Tropsch on iron with H2-CO and H2-CO2 as synthesis gases: The episodes of formation of the Fischer–Tropsch regime and construction of the catalyst. Top. Catal. 2003, 26, 41–54. [Google Scholar] [CrossRef]
  59. Wang, W.; Wang, S.P.; Ma, X.B.; Gong, J.L. Recent advances in catalytic hydrogenation of carbon dioxide. Chem. Soc. Rev. 2011, 40, 3703–3727. [Google Scholar] [CrossRef] [Green Version]
  60. Zhang, Y.; Cao, C.; Zhang, C.; Zhang, Z.; Liu, X.; Yng, Z.; Zhu, M.; Meng, B.; Jing, X.; Han, Y.-F. The study of structure-performance relationship of iron catalyst during a full life cycle for CO2 hydrogenation. J. Catal. 2019, 378, 51–62. [Google Scholar] [CrossRef]
  61. Fiato, R.A.; Rice, G.W.; Miseo, S.; Soled, S.L. Laser Produced Iron Carbide-Based Catalysts. U.S. Patent 4,687,753, 18 August 1987. [Google Scholar]
  62. Hegedus, L.L.; McCabe, R.W. Catalyst Poisoning. In Catalyst Deactivation 1980 (Studies in Surface Science and Catalysis); Delmon, B., Froment, G.F., Eds.; Elsevier: Amsterdam, The Netherlands, 1980; Volume 6, pp. 471–505. [Google Scholar]
  63. Bartholomew, C.H. Mechanisms of nickel catalyst poisoning. In Catalyst Deactivation 1987 (Studies in Surface Science and Catalysis); Delmon, B., Froment, G.F., Eds.; Elsevier: Amsterdam, The Netherlands, 1987; Volume 34, pp. 81–104. [Google Scholar]
  64. Schühle, P.; Schmidt, M.; Schill, L.; Rilsager, A.; Wasserscheid, P.; Albert, J. Influence of gas impurities on the hydrogenation of CO2 to methanol using indium-based catalysts. Catal. Sci. Technol. 2020, 10, 7309–7322. [Google Scholar] [CrossRef]
  65. Szailer, T.; Novák, É.; Oszkó, A.; Erdőhelyi, A. Effect of H2S on the hydrogenation of carbon dioxide over supported Rh catalysts. Top. Catal. 2007, 46, 79–86. [Google Scholar] [CrossRef]
  66. Rytter, E.; Holmen, A. Perspectives on the effect of water in cobalt Fischer–Tropsch synthesis. ACS Catal. 2017, 7, 5321–5328. [Google Scholar] [CrossRef]
  67. Wu, J.; Luo, S.; Toyir, J.; Saito, M.; Takeuchi, M.; Watanabe, T. Optimization of preparation conditions and improvement of stability of Cu/ZnO-based multicomponent catalysts for methanol synthesis from CO2 and H2. Catal. Today 1998, 45, 215–220. [Google Scholar] [CrossRef]
  68. Wu, J.; Saito, M.; Takeuchi, M.; Watanabe, T. The stability of Cu/ZnO-based catalysts in methanol synthesis from a CO2-rich feed and from a CO-rich feed. Appl. Catal. A Gen. 2001, 218, 235–240. [Google Scholar] [CrossRef]
  69. Huber, G.W.; Guymon, C.G.; Conrad, T.L.; Stephenson, B.C.; Bartholomew, C.H. Hydrothermal stability of Co/SiO2 Fischer-Tropsch Synthesis Catalysts. Stud. Surf. Sci. Catal. 2001, 139, 423–429. [Google Scholar]
  70. Van Steen, E.; Claeys, M.; Dry, M.E.; van de Loosdrecht, J.; Viljoen, E.L.; Visagie, J.L. Stability of nanocrystals: Thermodynamic analysis of oxidation and re-reduction of cobalt in water/hydrogen mixtures. J. Phys. Chem. B 2005, 109, 3575–3577. [Google Scholar] [CrossRef] [PubMed]
  71. Iglesia, E. Design, synthesis, and use of cobalt-based Fischer-Tropsch synthesis catalysts. Appl. Catal. A Gen. 1997, 161, 59–78. [Google Scholar] [CrossRef]
  72. Zhang, L.; Chen, K.; Chen, B.; White, J.L.; Resasco, D.E. Factors that determine zeolite stability in hot liquid water. J. Am. Chem. Soc. 2015, 137, 11810–11819. [Google Scholar] [CrossRef]
  73. Rostrup-Neilson, J.; Trimm, D.L. Mechanisms of carbon formation on nickel-containing catalysts. J. Catal. 1977, 48, 155–165. [Google Scholar] [CrossRef]
  74. Trimm, D.L. The formation and removal of coke from nickel catalyst. Catal. Rev. Sci. Eng. 1977, 16, 155–189. [Google Scholar] [CrossRef]
  75. Trimm, D.L. Catalyst design for reduced coking (review). Appl. Catal. 1983, 5, 263–290. [Google Scholar] [CrossRef]
  76. Bartholomew, C.H. Carbon deposition in steam reforming and methanation. Catal. Rev. Sci. Eng. 1982, 24, 67–112. [Google Scholar] [CrossRef]
  77. Albright, L.F.; Baker, R.T.K. (Eds.) Coke Formation on Metal Surfaces; ACS Symposium Series; American Chemical Society: Washington, DC, USA, 1982; Volume 202, pp. 1–200. [Google Scholar]
  78. Menon, P.G. Coke on catalysts-harmful, harmless, invisible and beneficial types. J. Mol. Catal. 1990, 59, 207–220. [Google Scholar] [CrossRef]
  79. Chen, D.; Moljord, K.; Holmen, A. Methanol to Olefins: Coke Formation and Deactivation. In Deactivation and Regeneration of Zeolite Catalysts; Guisnet, M., Ribeiro, F., Eds.; Imperial College Press: London, UK, 2011; Volume 9, pp. 269–292. [Google Scholar]
  80. Chen, D.; Rebo, H.P.; Grønvold, A.; Moljord, K.; Holmen, A. Methanol conversion to light olefins over SAPO-34: Kinetic modeling of coke formation. Microporous Mesoporous Mater. 2000, 35−36, 121–135. [Google Scholar] [CrossRef]
  81. Chen, D.; Moljord, K.; Fuglerud, T.; Holmen, A. The effect of crystal size of SAPO-34 on the selectivity and deactivation of the MTO reaction. Microporous Mesoporous Mater. 1999, 29, 191–203. [Google Scholar] [CrossRef]
  82. Nishiyama, N.; Kawaguchi, M.; Hirota, Y.; Van Vu, D.; Egashira, Y.; Ueyama, K. Size control of SAPO-34 crystals and their catalyst lifetime in the methanol-to-olefin reaction. Appl. Catal. A Gen. 2009, 362, 193–199. [Google Scholar] [CrossRef]
  83. Müller, S.; Liu, Y.; Vishnuvarthan, M.; Sun, X.; van Veen, A.C.; Haller, G.L.; Sanchez-Sanchez, M.; Lercher, J.A. Coke formation and deactivation pathways on H-ZSM-5 in the conversion of methanol to olefins. J. Catal. 2015, 325, 48–59. [Google Scholar] [CrossRef]
  84. Zhou, J.; Gao, M.; Zhang, J.; Liu, W.; Zhang, T.; Li, H.; Xu, Z.; Ye, M.; Liu, Z. Directed transforming of coke to active intermediates in methanol-to-olefins catalyst to boost light olefins selectivity. Nat. Commun. 2021, 12, 17. [Google Scholar] [CrossRef]
  85. Yang, M.; Tian, P.; Wang, C.; Yuan, Y.; Yang, Y.; Xu, S.; He, Y.; Liu, Z. A top-down approach to prepare sili-coaluminophosphate molecular sieve nanocrystals with improved catalytic activity. Chem. Commun. 2014, 50, 1845–1847. [Google Scholar] [CrossRef]
  86. Xi, D.; Sun, Q.; Chen, X.; Wang, N.; Yu, J. The recyclable synthesis of hierarchical zeolite SAPO-34 with ex-cellent MTO catalytic performance. Chem. Commun. 2015, 51, 11987–11989. [Google Scholar] [CrossRef]
  87. Guo, G.; Sun, Q.; Wang, N.; Bai, R.; Yu, J. Cost-effective synthesis of hierarchical SAPO-34 zeolites with abundant intracrystalline mesopores and excellent MTO performance. Chem. Commun. 2018, 54, 3697–3700. [Google Scholar] [CrossRef]
  88. Jiang, J.; Wen, C.; Tian, Z.; Wang, Y.; Zhai, Y.; Chen, L.; Li, Y.; Liu, Q.; Wang, C.; Ma, L. Manga-nese-promoted Fe3O4 microsphere for efficient conversion of CO2 to light olefins. Ind. Eng. Chem. Res. 2020, 59, 2155–2162. [Google Scholar] [CrossRef]
  89. Witoon, T.; Chaipraditgul, N.; Numpilai, T.; Lapkeatseree, V.; Ayodele, B.; Cheng, C.; Siri-Nguan, N.; Sorn-chamni, T.; Limtrakul, J. Highly active Fe-Co-Zn/K-Al2O3 catalysts for CO2 hydrogenation to light olefins. Chem. Eng. Sci. 2021, 233, 116428. [Google Scholar] [CrossRef]
  90. Zhang, Z.; Yin, H.; Yu, G.; He, S.; Kang, J.; Liu, Z.; Cheng, K.; Zhang, Q.; Wang, Y. Selective hydrogenation of CO2 and CO into olefins over sodium- and zinc-promoted iron carbide catalysts. J. Catal. 2021, 395, 350–361. [Google Scholar] [CrossRef]
  91. Yuan, F.; Zhang, G.; Zhu, J.; Ding, F.; Zhang, A.; Song, C.; Guo, X. Boosting light olefin selectivity in CO2 hy-drogenation by adding Co to Fe catalysts within close proximity. Catal. Today 2021, 371, 142–149. [Google Scholar] [CrossRef]
  92. Wei, C.; Tu, W.; Jia, L.; Liu, Y.; Lian, H.; Wang, P.; Zhang, Z. The evolutions of carbon and iron species modi-fied by Na and their tuning effect on the hydrogenation of CO2 to olefins. Appl. Surf. Sci. 2020, 525, 146622. [Google Scholar] [CrossRef]
  93. Malhi, H.S.; Sun, C.; Zhang, Z.; Liu, Y.; Liu, W.; Ren, P.; Tu, W.; Han, Y.-F. Catalytic consequences of the dec-oration of sodium and zinc atoms during CO2 hydrogenation to olefins over iron-based catalyst. Catal. Today 2021, 4. [Google Scholar] [CrossRef]
  94. 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]
  95. Liang, B.; Sun, T.; Ma, J.; Duan, H.; Li, L.; Yang, X.; Zhang, Y.; Su, X.; Huang, Y.; Zhang, T. Mn decorated Na/Fe catalysts for CO2 hydrogenation to light olefins. Catal. Sci. Technol. 2019, 9, 456–464. [Google Scholar] [CrossRef]
  96. Yang, S.; Chun, H.; Lee, S.; Han, S.J.; Lee, K.; Kim, Y.T. Comparative study of olefin production from CO and CO2 Using Na-and K-promoted zinc ferrite. ACS Catal. 2020, 10, 10742–10759. [Google Scholar] [CrossRef]
  97. Guo, L.; Sun, J.; Ji, X.; Wei, J.; Wen, Z.; Yao, R.; Xu, H.; Ge, Q. Directly converting carbon dioxide to linear α-olefins on bio-promoted catalysts. Commun. Chem. 2018, 1, 11. [Google Scholar] [CrossRef]
  98. Liang, B.; Duan, H.; Sun, T.; Ma, T.; Liu, X.; Xu, J.; Su, X.; Huang, Y.; Zhang, T. Effect of Na promoter on Fe-based catalyst for CO2 hydrogenation to alkenes. ACS Sustain. Chem. Eng. 2019, 7, 925–932. [Google Scholar] [CrossRef]
  99. Ramirez, A.; Gevers, L.; Bavykina, A.; Ould-Chikh, S.; Gascon, J. Metal organic framework-derived iron cat-alysts for the direct hydrogenation of CO2 to short chain olefins. ACS Catal. 2018, 8, 9174–9182. [Google Scholar] [CrossRef]
  100. Zhang, J.; Su, X.; Wang, X.; Ma, Q.; Fan, S.; Zhao, T. Promotion effects of Ce added Fe–Zr–K on CO2 hydro-genation to light olefins. React. Kinet. Mech. Catal. 2018, 124, 575–585. [Google Scholar] [CrossRef]
  101. Numpilai, T.; Chanlek, N.; Poo-Arporn, Y.; Cheng, C.K.; Siri-Nguan, N.; Sornchamni, T.; Chareonpanich, M.; Kongkachuichay, P.; Yigit, N.; Rupprechter, G.; et al. Tuning interactions of sur-face-adsorbed species over Fe-Co/K-Al2O3 catalyst by different K content: Selective CO2 hydrogenation to light olefins. ChemCatChem 2020, 12, 3306–3320. [Google Scholar] [CrossRef]
  102. Wang, W.; Jiang, X.; Wang, X.; Song, C. Fe−Cu bimetallic catalysts for selective CO2 hydrogenation to ole-fin-rich C2+ hydrocarbons. Ind. Eng. Chem. Res. 2018, 57, 4535–4542. [Google Scholar] [CrossRef]
  103. 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]
  104. Boreriboon, N.; Jiang, X.; Song, C.; Prasassarakich, P. Fe-based bimetallic catalysts supported on TiO2 for se-lective CO2 hydrogenation to hydrocarbons. J. CO2 Util. 2018, 25, 330–337. [Google Scholar] [CrossRef]
  105. Xu, Q.; Xu, X.; Fan, G.; Yang, L.; Li, F. Unveiling the roles of Fe-Co interactions over ternary spinel-type ZnCoxFe2-xO4 catalysts for highly efficient CO2 hydrogenation to produce light olefins. J. Catal. 2021, 400, 355–366. [Google Scholar] [CrossRef]
  106. Numpilai, T.; Witoon, T.; Chanlek, N.; Limphirat, W.; Bonura, G.; Chareonpanich, M.; Limtrakul, J. Struc-ture–activity relationships of Fe-Co/K-Al2O3 catalysts calcined at different temperatures for CO2 hydro-genation to light olefins. Appl. Catal. A Gen. 2017, 547, 219–229. [Google Scholar] [CrossRef]
  107. Rodemerck, U.; Holeňa, M.; Wagner, E.; Smejkal, Q.; Barkschat, A.; Baerns, M. Catalyst development for CO2 hydrogenation to fuels. ChemCatChem 2013, 5, 1948–1955. [Google Scholar] [CrossRef]
  108. Gong, W.; Ye, R.-P.; Ding, J.; Wang, T.; Shi, X.; Russell, C.K.; Tang, J.; Eddings, E.G.; Zhang, Y.; Fan, M. Effect of copper on highly effective Fe-Mn based catalysts during production of light olefins via Fischer-Tropsch process with low CO2 emission. Appl. Catal. 2020, 278, 119302. [Google Scholar] [CrossRef]
  109. Dai, C.; Zhang, A.; Liu, M.; Song, F.; Song, C.; Guo, X. Facile one-step synthesis of hierarchical porous carbon monoliths as superior supports of Fe-based catalysts for CO2 hydrogenation. RSC Adv. 2016, 6, 10831–10836. [Google Scholar] [CrossRef]
  110. Hu, S.; Liu, M.; Ding, F.; Song, C.; Zhang, G.; Guo, X. Hydrothermally stable MOFs for CO2 hydrogenation over iron-based catalyst to light olefins. J. CO2 Util. 2016, 15, 89–95. [Google Scholar] [CrossRef]
  111. Raghav, H.; Siva Kumar Konathala, L.N.; Mishra, N.; Joshi, B.; Goyal, R.; Agrawal, A.; Sarkar, B. Fe-decorated hierarchical molybdenum carbide for direct conversion of CO2 into ethylene: Tailoring activity and stability. J. CO2 Util. 2021, 50, 101607. [Google Scholar] [CrossRef]
  112. Huang, J.; Jiang, S.; Wang, M.; Wang, X.; Gao, J.; Song, C. Dynamic evolution of Fe and carbon species over different ZrO2 supports during CO prereduction and their effects on CO2 hydrogenation to light olefins. ACS Sustain. Chem. Eng. 2021, 9, 7891–7903. [Google Scholar] [CrossRef]
  113. Torrente-Murciano, L.; Chapman, R.S.L.; Narvaez-Dinamarca, A.; Mattia, D.; Jones, M.D. Effect of nanostructured ceria as support for the iron catalysed hydrogenation of CO2 into hydrocarbons. Phys. Chem. Chem. Phys. 2016, 18, 15496–15500. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Gu, H.; Ding, J.; Zhong, Q.; Zeng, Y.; Song, F. Promotion of surface oxygen vacancies on the light olefins synthesis from catalytic CO2 hydrogenation over Fe–K/ZrO2 catalysts. Int. J. Hydrogen Energy 2019, 44, 11808–11816. [Google Scholar] [CrossRef]
  115. Owen, R.E.; Pluckinski, P.; Mattia, D.; Torrente-Murciano, L.; Ting, V.; Jones, M.D. Effect of support of Co-Na-Mo catalysts on the direct conversion of CO2 to hydrocarbons. J. CO2 Util. 2016, 16, 97–103. [Google Scholar] [CrossRef] [Green Version]
  116. Wang, P.; Zha, F.; Yao, L.; Chang, Y. Synthesis of light olefins from CO2 hydrogenation over (CuO-ZnO)-kaolin/SAPO-34 molecular sieves. Appl. Clay Sci. 2018, 163, 249–256. [Google Scholar] [CrossRef]
  117. Wang, S.; Zhang, L.; Zhang, W.; Wang, P.; Qin, Z.; Yan, W.; Dong, M.; Li, J.; Wang, J.; He, L.; et al. Selective conversion of CO2 into propene and butene. Chem 2020, 6, 3344–3363. [Google Scholar] [CrossRef]
  118. Li, W.; Wang, K.; Zhan, G.; Huang, J.; Li, Q. Design and synthesis of bioinspired ZnZrOx & bio-ZSM-5 inte-grated nanocatalysts to boost CO2 hydrogenation to light olefins. ACS Sustain. Chem. Eng. 2021, 9, 6446–6458. [Google Scholar]
  119. Wang, S.; Wang, P.; Qin, Z.; Yan, W.; Dong, M.; Li, J.; Wang, J.; Fan, W. Enhancement of light olefin produc-tion in CO2 hydrogenation over In2O3-based oxide and SAPO-34 composite. J. Catal. 2020, 391, 459–470. [Google Scholar] [CrossRef]
  120. Chen, J.; Wang, X.; Wu, D.; Zhang, J.; Ma, Q.; Gao, X.; Lai, X.; Xia, H.; Fan, S.; Zhao, T. Hydrogenation of CO2 to light olefins on CuZnZr@(Zn-)SAPO-34 catalysts: Strategy for product distribution. Fuel 2019, 239, 44–52. [Google Scholar] [CrossRef]
  121. Liu, X.; Wang, M.; Yin, H.; Hu, J.; Cheng, K.; Kang, J.; Zhang, Q.; Wang, Y. Tandem catalysis for hydrogena-tion of CO and CO2 to lower olefins with bifunctional catalysts composed of spinel oxide and SAPO-34. ACS Catal. 2020, 10, 8303–8314. [Google Scholar] [CrossRef]
  122. Tong, M.; Chizema, L.G.; Chang, X.; Hondo, E.; Dai, L.; Zeng, Y.; Zeng, C.; Ahmad, H.; Yang, R.; Lu, P. Tan-dem catalysis over tailored ZnO-ZrO2/MnSAPO-34 composite catalyst for enhanced light olefins selectivity in CO2 hydrogenation. Microporous Mesoporous Mater. 2021, 320, 111105. [Google Scholar] [CrossRef]
  123. Wang, X.; Wu, D.; Zhang, J.; Gao, X.; Ma, Q.; Fan, S.; Zhao, T. Highly selective conversion of CO2 to light ole-fins via Fischer-Tropsch synthesis over stable layered K–Fe–Ti catalysts. Appl. Catal. A Gen. 2019, 573, 32–40. [Google Scholar] [CrossRef]
  124. Elishav, O.; Shener, Y.; Beilin, V.; Langau, M.V.; Herskowitz, M.; Shter, G.E.; Grader, G.S. Electrospun Fe−Al−O nanobelts for selective CO2 hydrogenation to light olefins. ACS Appl. Mater. Interfaces 2020, 12, 24855–24867. [Google Scholar] [CrossRef] [PubMed]
  125. Chen, H.; Liu, P.; Li, J.; Wang, Y.; She, C.; Liu, J.; Zhang, L.; Yang, Q.; Zhou, S.; Feng, X. MgH2/CuxO hydrogen storage composite with defect-rich surfaces for carbon dioxide hydrogenation. ACS Appl. Mater. Interfaces 2019, 11, 31009–31017. [Google Scholar] [CrossRef] [PubMed]
  126. Tian, H.; Yao, J.; Zha, F.; Yao, L.; Chang, Y. Catalytic activity of SAPO-34 molecular sieves prepared by using palygorskite in the synthesis of light olefins via CO2 hydrogenation. Appl. Clay Sci. 2020, 184, 105392. [Google Scholar] [CrossRef]
  127. Wu, T.; Lin, J.; Cheng, Y.; Tian, J.; Wang, S.; Xie, S.; Pei, Y.; Yan, S.; Qiao, M.; Xu, H.; et al. Porous gra-phene-confined Fe−K as highly efficient catalyst for CO2 direct hydrogenation to light olefins. ACS Appl. Mater. Interfaces 2018, 10, 23439–23443. [Google Scholar] [CrossRef]
  128. Chen, H.; Liu, J.; Liu, P.; Wang, Y.; Xiao, H.; Yang, Q.; Feng, X.; Zhou, S. Carbon-confined magnesium hydride nano-lamellae for catalytic hydrogenation of carbon dioxide to lower olefins. J. Catal. 2019, 379, 121–128. [Google Scholar] [CrossRef]
  129. Fujiwara, M.; Satake, T.; Shiokawa, K.; Sakurai, H. CO2 hydrogenation for C2+ hydrocarbon synthesis over composite catalyst using surface modified HB zeolite. Appl. Catal. B Environ. 2015, 179, 37–43. [Google Scholar] [CrossRef]
  130. Tan, L.; Wang, F.; Zhang, P.P.; Suzuki, Y.; Wu, Y.; Chen, J.; Yang, G.; Tsubaki, N. Design of a core–shell catalyst: An effective strategy for suppressing side reactions in syngas for direct selective conversion to light olefins. Chem. Sci. 2020, 11, 4097–4105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Trends in the atmospheric CO2 concentration (ppm) [2].
Figure 1. Trends in the atmospheric CO2 concentration (ppm) [2].
Catalysts 11 01447 g001
Scheme 1. Reaction route for CO2 hydrogenation to light olefins.
Scheme 1. Reaction route for CO2 hydrogenation to light olefins.
Catalysts 11 01447 sch001
Scheme 2. Complex reaction network for CO2 conversion to chemicals through hydrogenation.
Scheme 2. Complex reaction network for CO2 conversion to chemicals through hydrogenation.
Catalysts 11 01447 sch002
Scheme 3. (ac) Reaction mechanism for CO2 hydrogenation to light olefins (modified and adapted with permission from ref. [27]. Copyright 2021 Elsevier).
Scheme 3. (ac) Reaction mechanism for CO2 hydrogenation to light olefins (modified and adapted with permission from ref. [27]. Copyright 2021 Elsevier).
Catalysts 11 01447 sch003
Figure 2. Product distribution as predicted by the Anderson–Schulz–Flory (ASF) model. Adapted with permission from ref. [30]. Copyright 2021 Elsevier.
Figure 2. Product distribution as predicted by the Anderson–Schulz–Flory (ASF) model. Adapted with permission from ref. [30]. Copyright 2021 Elsevier.
Catalysts 11 01447 g002
Figure 3. Diagram of active phase sintering occurring over a support material: the blue ring represents atomic migration to form larger crystallites; the red ring represents the coalescence of crystallites. Adapted with permission from ref. [50]. Copyright 2021 Elsevier.
Figure 3. Diagram of active phase sintering occurring over a support material: the blue ring represents atomic migration to form larger crystallites; the red ring represents the coalescence of crystallites. Adapted with permission from ref. [50]. Copyright 2021 Elsevier.
Catalysts 11 01447 g003
Figure 4. (A) (af) Schematic illustration of the metal distribution and TEM images of the 13Fe2Co/ZrO2-supported catalyst precursor (ac) and the spent catalyst (df). (B) (af) Schematic illustration of the metal distribution and TEM images of the 13Fe2Co100Zr polymetallic oxide fiber catalyst (ac) and the spent catalyst (df). (C) CO2 conversion (a) and the C2+/C2= –C4= selectivity and C2= –C4= yield (b) over different catalysts after 8 h TOS (testing conditions: H2/CO2 molar ratio = 3/1, GHSV = 7200 mL g−1 h−1, P = 3 MPa, T = 673 K). Adapted with permission from ref. [56]. Copyright 2019 Elsevier.
Figure 4. (A) (af) Schematic illustration of the metal distribution and TEM images of the 13Fe2Co/ZrO2-supported catalyst precursor (ac) and the spent catalyst (df). (B) (af) Schematic illustration of the metal distribution and TEM images of the 13Fe2Co100Zr polymetallic oxide fiber catalyst (ac) and the spent catalyst (df). (C) CO2 conversion (a) and the C2+/C2= –C4= selectivity and C2= –C4= yield (b) over different catalysts after 8 h TOS (testing conditions: H2/CO2 molar ratio = 3/1, GHSV = 7200 mL g−1 h−1, P = 3 MPa, T = 673 K). Adapted with permission from ref. [56]. Copyright 2019 Elsevier.
Catalysts 11 01447 g004
Figure 5. MTO reaction over the H-ZSM-5-S catalyst in the PFR at ambient pressure and the H-ZSM-5-S catalyst in the CSTR at 6.5 bar, T = 723 K and pMeOH = 178 mbar. Adapted with permission from ref. [83]. Copyright 2015 Elsevier.
Figure 5. MTO reaction over the H-ZSM-5-S catalyst in the PFR at ambient pressure and the H-ZSM-5-S catalyst in the CSTR at 6.5 bar, T = 723 K and pMeOH = 178 mbar. Adapted with permission from ref. [83]. Copyright 2015 Elsevier.
Catalysts 11 01447 g005
Figure 6. SEM image of the Fe–K/γ-Al2O3 catalysts after the CO2 hydrogenation reaction: (a) 100 h, (b) 300 h, and (c) 500 h. Adapted with permission from ref. [57]. Copyright 2009 Elsevier.
Figure 6. SEM image of the Fe–K/γ-Al2O3 catalysts after the CO2 hydrogenation reaction: (a) 100 h, (b) 300 h, and (c) 500 h. Adapted with permission from ref. [57]. Copyright 2009 Elsevier.
Catalysts 11 01447 g006
Figure 7. (a) Distribution of iron species content over different spent catalysts. (b) CO2 hydrogenation over Fe/C−K2CO3 catalysts with varying proximity. Adapted with permission from ref. [94]. Copyright 2020 American Chemical Society.
Figure 7. (a) Distribution of iron species content over different spent catalysts. (b) CO2 hydrogenation over Fe/C−K2CO3 catalysts with varying proximity. Adapted with permission from ref. [94]. Copyright 2020 American Chemical Society.
Catalysts 11 01447 g007
Figure 8. (a) Illustrated synthesis of Fe/C catalysts. (b) TEM image of Fe/C catalysts. (c) Catalytic performance over promoted and unpromoted Fe catalysts. (d) CO2 conversion after 50 h of TOS. (e) XRD of promoted and unpromoted Fe catalysts. (f) Effect of K loading on the selectivity and CO2 conversion after 50 h of TOS. Testing conditions: 593 K, 30 bar, H2/CO2 molar ratio = 3, and GHSV = 24,000 mL·g−1·h−1. Adapted with permission from ref. [99]. Copyright 2018 American Chemical Society.
Figure 8. (a) Illustrated synthesis of Fe/C catalysts. (b) TEM image of Fe/C catalysts. (c) Catalytic performance over promoted and unpromoted Fe catalysts. (d) CO2 conversion after 50 h of TOS. (e) XRD of promoted and unpromoted Fe catalysts. (f) Effect of K loading on the selectivity and CO2 conversion after 50 h of TOS. Testing conditions: 593 K, 30 bar, H2/CO2 molar ratio = 3, and GHSV = 24,000 mL·g−1·h−1. Adapted with permission from ref. [99]. Copyright 2018 American Chemical Society.
Catalysts 11 01447 g008
Figure 9. (a) Illustrated scheme of CO2 hydrogenation. (b) Conversion and selectivity of CO2 hydrogenation over an xNa/Fe catalyst (testing conditions: H2/CO2 molar ratio = 3; P = 3 MPa; T = 593 K; GHSV = 2040 mL h−1 gcat−1; TOS = 10 h). (c) Mössbauer spectra of the spent Na-free/Fe and 1Na/Fe catalysts. (d) Fe5C2 content of the spent xNa/Fe catalyst vs. the Na content. Adapted with permission from ref. [98]. Copyright 2019 American Chemical Society.
Figure 9. (a) Illustrated scheme of CO2 hydrogenation. (b) Conversion and selectivity of CO2 hydrogenation over an xNa/Fe catalyst (testing conditions: H2/CO2 molar ratio = 3; P = 3 MPa; T = 593 K; GHSV = 2040 mL h−1 gcat−1; TOS = 10 h). (c) Mössbauer spectra of the spent Na-free/Fe and 1Na/Fe catalysts. (d) Fe5C2 content of the spent xNa/Fe catalyst vs. the Na content. Adapted with permission from ref. [98]. Copyright 2019 American Chemical Society.
Catalysts 11 01447 g009
Figure 10. (a) Illustrated reaction mechanism for CO2 hydrogenation over the catalyst Na-Zn-Fe. (b) Effect of the proximity between ZnO and Na+Fe5C2 on catalytic behaviors for CO2 hydrogenation. Catalyst stability in CO2 hydrogenation over (c) an Na+Fe5C2 catalyst and (d) a Na-Zn-Fe catalyst. Testing conditions: H2/CO2 molar ratio = 3, P = 2.5 MPa, W = 0.10 g, F = 25 mL min−1, T = 613 K. Adapted with permission from ref. [90]. Copyright 2021 Elsevier.
Figure 10. (a) Illustrated reaction mechanism for CO2 hydrogenation over the catalyst Na-Zn-Fe. (b) Effect of the proximity between ZnO and Na+Fe5C2 on catalytic behaviors for CO2 hydrogenation. Catalyst stability in CO2 hydrogenation over (c) an Na+Fe5C2 catalyst and (d) a Na-Zn-Fe catalyst. Testing conditions: H2/CO2 molar ratio = 3, P = 2.5 MPa, W = 0.10 g, F = 25 mL min−1, T = 613 K. Adapted with permission from ref. [90]. Copyright 2021 Elsevier.
Catalysts 11 01447 g010
Figure 11. Catalytic performance of the CO2 hydrogenation over the Na-CoFe2O4 catalyst at TOS (reaction conditions: H2/CO2 molar ratio = 3, T= 593 K, P = 3 MPa, GHSV = 7200 mL h−1 gcat−1, TOS = 100 h). Adapted with permission from ref. [91]. Copyright 2021 Elsevier.
Figure 11. Catalytic performance of the CO2 hydrogenation over the Na-CoFe2O4 catalyst at TOS (reaction conditions: H2/CO2 molar ratio = 3, T= 593 K, P = 3 MPa, GHSV = 7200 mL h−1 gcat−1, TOS = 100 h). Adapted with permission from ref. [91]. Copyright 2021 Elsevier.
Catalysts 11 01447 g011
Figure 12. (a) Illustrated reaction mechanism. (b) Catalytic performance of the CO2 hydrogenation over a Zn-promoted Fe-Co/K-Al2O3 catalyst. (c) XPS spectra (Fe 2p region) of the 0.58 wt% Zn-promoted Fe-Co/K-Al2O3 catalysts. (d) XRD pattern of the 0.58 wt% Zn-promoted Fe-Co/K-Al2O3 catalyst at varying TOS. Testing conditions: T = 613 K, P = 25 bar, GHSV = 9000 mL gcat−1 h−1 and H2/CO2 molar ratio = 4. Adapted with permission from ref. [89]. Copyright 2021 Elsevier.
Figure 12. (a) Illustrated reaction mechanism. (b) Catalytic performance of the CO2 hydrogenation over a Zn-promoted Fe-Co/K-Al2O3 catalyst. (c) XPS spectra (Fe 2p region) of the 0.58 wt% Zn-promoted Fe-Co/K-Al2O3 catalysts. (d) XRD pattern of the 0.58 wt% Zn-promoted Fe-Co/K-Al2O3 catalyst at varying TOS. Testing conditions: T = 613 K, P = 25 bar, GHSV = 9000 mL gcat−1 h−1 and H2/CO2 molar ratio = 4. Adapted with permission from ref. [89]. Copyright 2021 Elsevier.
Catalysts 11 01447 g012
Scheme 4. Schematic demonstration of CO2 hydrogenation over CNT supported bi-metallic catalyst CoFe2O4. Adapted with permission from ref. [103]. Copyright 2020 American Chemical Society.
Scheme 4. Schematic demonstration of CO2 hydrogenation over CNT supported bi-metallic catalyst CoFe2O4. Adapted with permission from ref. [103]. Copyright 2020 American Chemical Society.
Catalysts 11 01447 sch004
Figure 13. (a) Schematic illustration of the structural transformations of as-formed ZnCoxFe2-xO4 catalysts during the reduction and reaction steps. (b) CO2 conversion and product distributions over K-containing ZnCoxFe2-xO4 catalysts with various Fe/Co molar ratios. (c) The stability of the K-containing ZnCo0.5Fe1.5O4 catalyst in CO2 hydrogenation (testing conditions: T = 583 K, P = 2.5 MPa, GHSV = 4800 mLh−1gcat−1, CO2/H2 molar ratio = 1:3). Adapted with permission from ref. [105]. Copyright 2021 Elsevier.
Figure 13. (a) Schematic illustration of the structural transformations of as-formed ZnCoxFe2-xO4 catalysts during the reduction and reaction steps. (b) CO2 conversion and product distributions over K-containing ZnCoxFe2-xO4 catalysts with various Fe/Co molar ratios. (c) The stability of the K-containing ZnCo0.5Fe1.5O4 catalyst in CO2 hydrogenation (testing conditions: T = 583 K, P = 2.5 MPa, GHSV = 4800 mLh−1gcat−1, CO2/H2 molar ratio = 1:3). Adapted with permission from ref. [105]. Copyright 2021 Elsevier.
Catalysts 11 01447 g013
Scheme 5. Schematic illustration of CO2 hydrogenation over unpromoted and Zr- and K-promoted cobalt catalysts supported on a-TiO2 and r-TiO2. Adapted with permission from ref. [10]. Copyright 2013 American Chemical Society.
Scheme 5. Schematic illustration of CO2 hydrogenation over unpromoted and Zr- and K-promoted cobalt catalysts supported on a-TiO2 and r-TiO2. Adapted with permission from ref. [10]. Copyright 2013 American Chemical Society.
Catalysts 11 01447 sch005
Scheme 6. CO pre-reduction and CO2 hydrogenation process on (a) m-ZrO2- and (b) t-ZrO2-supported Fe-Zr catalysts. Adapted with permission from ref. [112]. Copyright 2021 American Chemical Society.
Scheme 6. CO pre-reduction and CO2 hydrogenation process on (a) m-ZrO2- and (b) t-ZrO2-supported Fe-Zr catalysts. Adapted with permission from ref. [112]. Copyright 2021 American Chemical Society.
Catalysts 11 01447 sch006
Figure 14. The stability of 10Fe1K/m-ZrO2 and 10Fe1K/t-ZrO2 for CO2 conversion, and the light olefin selectivity at a TOS of 100 h at 613K. Adapted with permission from ref. [114]. Copyright 2019 Elsevier.
Figure 14. The stability of 10Fe1K/m-ZrO2 and 10Fe1K/t-ZrO2 for CO2 conversion, and the light olefin selectivity at a TOS of 100 h at 613K. Adapted with permission from ref. [114]. Copyright 2019 Elsevier.
Catalysts 11 01447 g014
Figure 15. (a) Catalytic performance of CO2 hydrogenation over Fe5C2-based catalysts on various supports. (b) Catalytic performance and stability over an Fe5C2-10K/a-Al2O3 catalyst (testing conditions: T = 593 K, P = 3.0 MPa, GHSV = 3600 mLg−1h−1, H2/CO2 molar ratio = 3). Adapted with permission from ref. [15]. Copyright 2018 American Chemical Society.
Figure 15. (a) Catalytic performance of CO2 hydrogenation over Fe5C2-based catalysts on various supports. (b) Catalytic performance and stability over an Fe5C2-10K/a-Al2O3 catalyst (testing conditions: T = 593 K, P = 3.0 MPa, GHSV = 3600 mLg−1h−1, H2/CO2 molar ratio = 3). Adapted with permission from ref. [15]. Copyright 2018 American Chemical Society.
Catalysts 11 01447 g015
Figure 16. (a) Illustrated reaction mechanism over the bifunctional composite catalyst In2O3-ZrO2/SAPO-34, and (b) the stability of the In2O3-ZrO2/SAPO-34 composite catalyst for CO2 hydrogenation to light olefins (testing conditions: P = 2.0 MPa, T = 573 K, GHSV = 2160 cm3h−1gcat−1). Adapted with permission from ref. [44]. Copyright 2019 Elsevier.
Figure 16. (a) Illustrated reaction mechanism over the bifunctional composite catalyst In2O3-ZrO2/SAPO-34, and (b) the stability of the In2O3-ZrO2/SAPO-34 composite catalyst for CO2 hydrogenation to light olefins (testing conditions: P = 2.0 MPa, T = 573 K, GHSV = 2160 cm3h−1gcat−1). Adapted with permission from ref. [44]. Copyright 2019 Elsevier.
Catalysts 11 01447 g016
Figure 17. (a) Effect of the proximity of the active components on the CO2 conversion and product selectivity, and (b) the catalytic stability of the composite catalyst In-Zr/SAPO-34 (testing conditions: T = 673 K, P = 3.0 MPa, GHSV = 9000 mL gcat−1 h−1, molar ratio of H2/CO2/N2 = 73/24/3, and mass ratio of oxide/zeolite = 2). Adapted with permission from ref. [19]. Copyright 2018 American Chemical Society.
Figure 17. (a) Effect of the proximity of the active components on the CO2 conversion and product selectivity, and (b) the catalytic stability of the composite catalyst In-Zr/SAPO-34 (testing conditions: T = 673 K, P = 3.0 MPa, GHSV = 9000 mL gcat−1 h−1, molar ratio of H2/CO2/N2 = 73/24/3, and mass ratio of oxide/zeolite = 2). Adapted with permission from ref. [19]. Copyright 2018 American Chemical Society.
Catalysts 11 01447 g017
Figure 18. (a) Catalytic performance over the bifunctional composite catalysts ZnZrOx/bio-ZSM-5−Si at TOS (testing conditions: mass of catalyst = 0.6 g, T = 653 K, P = 3 MPa, gas flow rate = 20 mL min−1). (b) Effect of the proximity of the active components of ZnZrOx/bio-ZSM-5-Si on the catalytic performance. Adapted with permission from ref. [118]. Copyright 2021 American Chemical Society.
Figure 18. (a) Catalytic performance over the bifunctional composite catalysts ZnZrOx/bio-ZSM-5−Si at TOS (testing conditions: mass of catalyst = 0.6 g, T = 653 K, P = 3 MPa, gas flow rate = 20 mL min−1). (b) Effect of the proximity of the active components of ZnZrOx/bio-ZSM-5-Si on the catalytic performance. Adapted with permission from ref. [118]. Copyright 2021 American Chemical Society.
Catalysts 11 01447 g018
Figure 19. (a) The content of the surface oxygen vacancies (Ov) from O (1s) XPS spectra for the catalysts In2O3, InCeOx(0.13), and InCrOx(0.13). (b) The catalytic stability of the bifunctional composite catalysts InCrOx(0.13) and SAPO-34 for CO2 hydrogenation (testing conditions: H2/CO2 molar ratio = 3/1, T = 623 K, P = 3.5 MPa, and GHSV = 1140 mLgcat−1h−1). Adapted with permission from ref. [119]. Copyright 2020 Elsevier.
Figure 19. (a) The content of the surface oxygen vacancies (Ov) from O (1s) XPS spectra for the catalysts In2O3, InCeOx(0.13), and InCrOx(0.13). (b) The catalytic stability of the bifunctional composite catalysts InCrOx(0.13) and SAPO-34 for CO2 hydrogenation (testing conditions: H2/CO2 molar ratio = 3/1, T = 623 K, P = 3.5 MPa, and GHSV = 1140 mLgcat−1h−1). Adapted with permission from ref. [119]. Copyright 2020 Elsevier.
Catalysts 11 01447 g019
Figure 20. (a) CO2 hydrogenation over the bifunctional composite catalyst ZnZrO/SAPO-34, with the effect of the proximity of the active components of ZnZrO and SAPO-34 on the catalytic performance. (b) The catalytic stability of the catalyst ZnZrO/SAPO-34 (testing conditions: T = 653K, P = 2 MPa, and GHSV = 3600 mL gcat−1 h−1). Adapted with permission from ref. [13]. Copyright 2017 American Chemical Society.
Figure 20. (a) CO2 hydrogenation over the bifunctional composite catalyst ZnZrO/SAPO-34, with the effect of the proximity of the active components of ZnZrO and SAPO-34 on the catalytic performance. (b) The catalytic stability of the catalyst ZnZrO/SAPO-34 (testing conditions: T = 653K, P = 2 MPa, and GHSV = 3600 mL gcat−1 h−1). Adapted with permission from ref. [13]. Copyright 2017 American Chemical Society.
Catalysts 11 01447 g020
Figure 21. (a) Catalytic performance for CO2 hydrogenation over a In2O3-ZnZrOx catalyst with different types of SAPO-34. (b) The stability of the bifunctional composite catalysts In2O3-ZnZrOx/SAPO-34-H-a (testing conditions: T = 653 K, P = 3.0 MPa, GHSV = 9000 mLgcat−1h−1, molar ratio of H2/CO2/N2 = 73:24:3, mass ratio of oxide/zeolite = 0.5). Adapted with permission from ref. [42]. Copyright 2019 Wiley-VCH.
Figure 21. (a) Catalytic performance for CO2 hydrogenation over a In2O3-ZnZrOx catalyst with different types of SAPO-34. (b) The stability of the bifunctional composite catalysts In2O3-ZnZrOx/SAPO-34-H-a (testing conditions: T = 653 K, P = 3.0 MPa, GHSV = 9000 mLgcat−1h−1, molar ratio of H2/CO2/N2 = 73:24:3, mass ratio of oxide/zeolite = 0.5). Adapted with permission from ref. [42]. Copyright 2019 Wiley-VCH.
Catalysts 11 01447 g021
Figure 22. (a) XPS spectra (O1s) of various oxides and the content of surface oxygen vacancies (Ov). (b) The stability tests of the bifunctional composite catalysts In-Zr(4:1)/SAPO-34 (testing conditions: T = 653 K, P = 3.0 MPa, GHSV = 9000 mL gcat−1 h−1, molar ratio of H2/CO2/N2 = 73/24/3, and mass ratio of oxide/zeolite = 0.5). Adapted with permission from ref. [18]. Copyright 2018 Elsevier.
Figure 22. (a) XPS spectra (O1s) of various oxides and the content of surface oxygen vacancies (Ov). (b) The stability tests of the bifunctional composite catalysts In-Zr(4:1)/SAPO-34 (testing conditions: T = 653 K, P = 3.0 MPa, GHSV = 9000 mL gcat−1 h−1, molar ratio of H2/CO2/N2 = 73/24/3, and mass ratio of oxide/zeolite = 0.5). Adapted with permission from ref. [18]. Copyright 2018 Elsevier.
Catalysts 11 01447 g022
Figure 23. (a) Catalytic performance over different catalysts. (b) The catalytic stability of 0.8K-2.4Fe-1.3Ti at TOS (testing conditions: H2/CO2 molar ratio = 3/1, T = 593 K, P = 2.0 MPa and GHSV = 10,000 mL gcat−1h−1). Adapted with permission from ref. [123]. Copyright 2019 Elsevier.
Figure 23. (a) Catalytic performance over different catalysts. (b) The catalytic stability of 0.8K-2.4Fe-1.3Ti at TOS (testing conditions: H2/CO2 molar ratio = 3/1, T = 593 K, P = 2.0 MPa and GHSV = 10,000 mL gcat−1h−1). Adapted with permission from ref. [123]. Copyright 2019 Elsevier.
Catalysts 11 01447 g023
Figure 24. Schematic illustration of the formation of Fe@NC catalysts and the reaction for CO2 hydrogenation. Adapted with permission from ref. [36]. Copyright 2019 American Chemical Society.
Figure 24. Schematic illustration of the formation of Fe@NC catalysts and the reaction for CO2 hydrogenation. Adapted with permission from ref. [36]. Copyright 2019 American Chemical Society.
Catalysts 11 01447 g024
Figure 25. (a) N2 physisorption isotherms, (b) SEM image, (c) HAADF−STEM image, and (d) TEM image and particle distribution of the FeK1.5/HSG catalyst. (e) CO2 hydrogenation over the catalyst FeK1.5/HSG during a TOS of 120 h (testing conditions: mass of catalyst = 0.15 g, T = 613 K, P = 20 bar, H2/CO2 molar ratio = 3, and GHSV = 26 L h−1g−1). Adapted with permission from ref. [127]. Copyright 2018 American Chemical Society.
Figure 25. (a) N2 physisorption isotherms, (b) SEM image, (c) HAADF−STEM image, and (d) TEM image and particle distribution of the FeK1.5/HSG catalyst. (e) CO2 hydrogenation over the catalyst FeK1.5/HSG during a TOS of 120 h (testing conditions: mass of catalyst = 0.15 g, T = 613 K, P = 20 bar, H2/CO2 molar ratio = 3, and GHSV = 26 L h−1g−1). Adapted with permission from ref. [127]. Copyright 2018 American Chemical Society.
Catalysts 11 01447 g025
Figure 26. (a) Schematic illustration of the interface between CZZ and SAPO-34. (b) Core–shell interface of CZZ and SAPO-34. (c) Stability of the composite catalyst CZZ@Zn-SAPO-34 at TOS (testing conditions: H2/CO2 molar ratio = 3, T = 673 K). Adapted with permission from ref. [120]. Copyright 2019 Elsevier.
Figure 26. (a) Schematic illustration of the interface between CZZ and SAPO-34. (b) Core–shell interface of CZZ and SAPO-34. (c) Stability of the composite catalyst CZZ@Zn-SAPO-34 at TOS (testing conditions: H2/CO2 molar ratio = 3, T = 673 K). Adapted with permission from ref. [120]. Copyright 2019 Elsevier.
Catalysts 11 01447 g026
Table 1. Some representative catalysts on promoter effect for CO2 hydrogenation to light olefins.
Table 1. Some representative catalysts on promoter effect for CO2 hydrogenation to light olefins.
CatalystCO2 Conv., %Selectivity, %Yield, %
C2–C4=
O/P RatioStabilityRef.
COCH4.C2–C4C2–C40
10Mn-Fe3O444.79.422.046.27.118.76.524 h[88]
0.58% Zn-Fe-Co/K-Al2O357.88.86.263.221.819.92.950 h[89]
Na-Zn-Fe38.015.013.042.04.916.08.5100 h[90]
Na-CoFe2O441.810.0~18.037.2~7.015.5~5.3100 h[91]
Fe-Co/K-Al2O340.012.224.846.17.916.25.96 h[33]
0.5%Na-Fe5C235.313.231.857.010.120.15.710 h[92]
Fe-Zn-2Na43.015.722.854.17.423.27.310 h[93]
Fe/C-KHCO333.020.812.759.8 a27.39.02.2100 h[94]
5Mn-Na/Fe38.611.711.830.24.011.711.010 h[95]
FeNa(1.18)40.513.515.846.67.515.76.260 h[31]
Na/Fe-Zn30.6n/a13.026.83.98.46.9200 h[96]
Fe/C-Bio31.023.211.821.724.46.70.96 h[97]
5%Na/Fe3O436.8~11.0~5.064.3~13.023.7~4.910 h[98]
Fe/C+K(0.75)40.0~16.0~22.0~39.0~12.0~15.6~3.350 h[99]
35Fe-7Zr-1Ce-K57.33.0520.655.67.931.87.184 h[100]
Fe-Mn/K-Al2O329.420.218.748.76.514.37.4>6 h[101]
Fe-Cu(0.17)/K(1.0)29.317.07.063.812.219.15.250 h[102]
Na-CoFe2O4/CNT34.419.0~5.038.818.013.312.9>24 h[103]
Fe-Co-K(0.3)/TiO221.254.09.037.0 bn/an/a4.118 h[104]
Fe2Zn135.0~15.0~20.057.8 c7.220.28.0200 h[28]
ZnCo0.5Fe1.5O449.6~7.5~17.536.1~10.017.9~3.680 h[105]
a refers to high valued olefins (HVO); b includes C2-C4 and C5+ hydrocarbons; c refer to C2–C7= olefins.
Table 2. Some representative catalysts of the support effect for CO2 hydrogenation to light olefins.
Table 2. Some representative catalysts of the support effect for CO2 hydrogenation to light olefins.
CatalystCO2 Conv., %Selectivity, %Yield, %
C2–C4=
O/P RatioStabilityRef.
COCH4C2–C4=C2–C40
Fe-K/HPCMs-133.438.913.518.011.56.01.635 h[109]
ZIF-8(a)/Fe2O3~24.0~24.0~21.0~20.0~24.0~4.80.83n/a[110]
Fe(0.5)-Mo2Cc9.80.52.192.03.59.026.32 h[111]
K-Zr-Co/aTiO270.0n/an/a17.0n/a11.9n/a8h[10]
Fe-Cr-K/Nb2O531.057.032.010.01.03.13.1n/a[4]
15Fe-K/m-ZrO238.819.930.142.812.816.63.312 h[112]
20%Fe/CeO2-NC18.973.575.518.24.03.44.1n/a[113]
10Fe-1K/m-ZrO240.5n/an/a15.0n/a6.1n/a100 h[114]
Fe5C2-10K/a-Al2O340.9n/an/a73.5n/a30.1n/a100 h[15]
Co-Na-Mo/CeO215.170.222.110.736.01.60.03n/a[115]
Table 3. Some representative catalysts for the bifunctional composite catalyst effect for CO2 hydrogenation to light olefins.
Table 3. Some representative catalysts for the bifunctional composite catalyst effect for CO2 hydrogenation to light olefins.
CatalystCO2 Conv., %Selectivity, %Yield, %
C2–C4=
O/P RatioStabilityRef.
COCH4C2–C4=C2–C40
CuO-ZnO & SAPO-3441.39.311.863.415.526.24.113 h[116]
(CuO-ZnO)-kaolin & SAPO-3457.69.611.463.815.236.74.220 h[116]
In2O3/ZrO2 & SAPO19.087.0~17.090.0 an/a17.1n/a>50 h[43]
In2O3/ZrO2 & SAPO~14.0<5.0<5.070.0n/a9.8n/a>100 h[44]
In−Zr/SAPO-3426.7n/a4.376.4 a~14.0 a20.45.5>150 h[19]
Zn0.5Ce0.2Zr1.8O4 & H-RUB-13 (200)10.728.32.983.45.48.915.4>30 h[117]
ZnZrOx & bio-ZSM-Si10.0~80.05.564.430.16.42.160 h[118]
InCrOx(0.13) & SAPO33.655.035.075.0 a20.0 a11.33.8>120 h[119]
ZnZrO & SAPO-3412.647.03.080.0 a14.0 a10.15.7>100 h[14]
CZZ@Zn & SAPO-34~7.0n/a~18.072.08.01.38.6>120 h[120]
In2O3-ZnZrOx & SAPO-34-S-a17.055.81.685.0 a11.1 a14.57.7>90 h[42]
In2O3-ZnZrOx & SAPO-34-H-a17.053.41.284.5 a11.0 a14.47.7>90 h[42]
ZnAl2O4 & SAPO-3415.049.00.787.0 a10.0 a13.18.710 h[121]
ZnGa2O4 & SAPO-3413.046.01.086.0 a11.0 a11.27.810 h[121]
ZnO-ZrO2 & Mn0.1SAPO-3424.442.23.761.733.615.11.810 h[122]
In-Zr (4:1) & SAPO-3426.263.92.074.5 a21.5 a19.53.5>140 h[18]
a CO is not considered when calculating selectivity.
Table 4. Some representative catalysts on the structure effect for CO2 hydrogenation to light olefins.
Table 4. Some representative catalysts on the structure effect for CO2 hydrogenation to light olefins.
CatalystCO2 Conv., %Selectivity, %Yield, %
C2–C4=
O/P RatioStabilityRef.
COCH4C2–C4=C2–C40
0.8K-2.4Fe-1.3Ti35.036.322.060.08.021.07.5200 h[123]
Fe@NC-40029.017.527.021.012.06.11.7>15 h[36]
K/Fe-Al-O Spinel E.1 nanobelts48.016.010.052.05.024.03.1120 h[124]
MgH2/CuxO20.7n/a40.054.87.011.37.8210 h[125]
CZA/SAPO-3450.03.010.062.025.033.02.512 h[126]
FeK1.5/HSG503931569.9285.7>120 h[127]
Carbon-confined MgH2 nano-lamellae10.527.617.550.94.05.312.7>2 h[128]
Fe-Co/K-(CM-Al2O3)41.012.433.741.16.414.46.450 h[37]
ZnO-Y2O3 & SAPO-3427.685.01.883.9 a12.9 a23.26.5n/a[40]
Cu-Zn-Al (6:3:1) oxide & HB zeolite27.653.40.745.5 bn/a12.6 bn/a9 h[129]
a Refers to hydrocarbon distribution %; b refers to C2-C5+ hydrocarbons.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Weber, D.; He, T.; Wong, M.; Moon, C.; Zhang, A.; Foley, N.; Ramer, N.J.; Zhang, C. Recent Advances in the Mitigation of the Catalyst Deactivation of CO2 Hydrogenation to Light Olefins. Catalysts 2021, 11, 1447. https://doi.org/10.3390/catal11121447

AMA Style

Weber D, He T, Wong M, Moon C, Zhang A, Foley N, Ramer NJ, Zhang C. Recent Advances in the Mitigation of the Catalyst Deactivation of CO2 Hydrogenation to Light Olefins. Catalysts. 2021; 11(12):1447. https://doi.org/10.3390/catal11121447

Chicago/Turabian Style

Weber, Daniel, Tina He, Matthew Wong, Christian Moon, Axel Zhang, Nicole Foley, Nicholas J. Ramer, and Cheng Zhang. 2021. "Recent Advances in the Mitigation of the Catalyst Deactivation of CO2 Hydrogenation to Light Olefins" Catalysts 11, no. 12: 1447. https://doi.org/10.3390/catal11121447

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