Abstract

Decentralized chemical plants close to circular carbon sources will play an important role in shaping the postfossil society. This scenario calls for carbon technologies which valorize CO2 and CO with renewable H2 and utilize process intensification approaches. The single-reactor tandem reaction approach to convert COx to hydrocarbons via oxygenate intermediates offers clear benefits in terms of improved thermodynamics and energy efficiency. Simultaneously, challenges and complexity in terms of catalyst material and mechanism, reactor, and process gaps have to be addressed. While the separate processes, namely methanol synthesis and methanol to hydrocarbons, are commercialized and extensively discussed, this review focuses on the zeolite/zeotype function in the oxygenate-mediated conversion of COx to hydrocarbons. Use of shape-selective zeolite/zeotype catalysts enables the selective production of fuel components as well as key intermediates for the chemical industry, such as BTX, gasoline, light olefins, and C3+ alkanes. In contrast to the separate processes which use methanol as a platform, this review examines the potential of methanol, dimethyl ether, and ketene as possible oxygenate intermediates in separate chapters. We explore the connection between literature on the individual reactions for converting oxygenates and the tandem reaction, so as to identify transferable knowledge from the individual processes which could drive progress in the intensification of the tandem process. This encompasses a multiscale approach, from molecule (mechanism, oxygenate molecule), to catalyst, to reactor configuration, and finally to process level. Finally, we present our perspectives on related emerging technologies, outstanding challenges, and potential directions for future research.
1. Introduction
While the majority of current C,H-containing fuels and commodities are produced from fossil sources (coal, oil, natural gas), global warming and ocean acidification concerns have triggered a step change in studies of recycled CO and CO2 as carbon sources for the postfossil chemical industry. A recent IEAGHG Technical Report estimated the annual demand for methanol (an intermediate for production of, e.g., chemicals, polyolefins, and heavy duty truck fuel) and middle distillate hydrocarbons (for heavy duty trucks and aviation fuel production) of 825 and 903 megatons (Mt) per year, in the postfossil era.1 Aiming for a cyclic economy with Carbon Capture and Utilization (CCU), the CO2 abatement potential of current processes was roughly estimated to 1537 and 3349 Mt/year, respectively, for the two product groups, presuming use of “green” hydrogen. While these numbers are far from the current, global CO2 emission numbers (36.8 Mt in 2022),2 the potential for CCU is substantial.
Today, fossil sources are first transformed to synthesis gas (syngas, CO/H2) and/or CO2/H2 mixtures, which are subsequently reacted to valuable products. The conversion of CO/CO2/H2 to lower olefins, aromatics, and fuels occurs predominantly via two pathways, either via Fischer–Tropsch Synthesis (FTS) or via oxygenate intermediates (i.e., methanol, dimethyl ether, ketene).3,4
In the FTS process, typical FTS metals (Fe, Co, or Ru) are utilized with promoters to overcome the Anderson–Schulz–Flory distribution in order to attain the desired hydrocarbon distributions.5,6 The combination of sodium and sulfur was proven to suppress the formation of methane,7,8 while manganese oxide and alkali elements such potassium and sodium were effective in promoting chain growth in the hydrocarbon products.5,6,9−12 Alternatively, zeolites were used in combination with the FTS metals, in order to crack/isomerize longer hydrocarbons to the desired hydrocarbon distribution.13,14
The alternative, industrial pathway via oxygenate intermediates is a two-step process, in which synthesis gas is converted to methanol in a first reactor, and methanol is subsequently converted to lower olefins or to gasoline in a second reactor.15,16 Also here, zeolites are key to obtaining high selectivity to desired product groups. As an example, 91% of C2–C4 olefins selectivity was reported at full methanol conversion using a zeotype catalyst with CHA topology.17 In comparison, the highest C2–C4 olefins yield reported for the Fischer–Tropsch to Olefins (FTO) process (without combining the FTO catalysts with a zeolite) is 58%.6 An overview of zeolite structures discussed in this Review is provided in Table 1. The correlation between the product-limiting pore size of some central zeolite structures and the main effluent product groups obtained in the methanol to hydrocarbons (MTH) process are provided in Table 2.
Table 1. An Overview of Zeolite/Zeotype Topologies Used for Processes and Subprocesses Described Hereina.
| Three-letter structure code | Material common name(s) | Dimensionality and type of structure | Effluent product restricting pore(s) (diameter in Å) |
|---|---|---|---|
| AEI | SAPO-18 | 3D cavity-window | 8-ring (3.8 × 3.8) |
| MAPO-18 | |||
| SSZ-39 | |||
| AEL | SAPO-11 | 1D straight channel | 10-ring (4.0 × 6.5) |
| AFI | SAPO-5 | 1D straight channel | 12-ring (7.3 × 7.3) |
| MAPO-5 | |||
| SSZ-24 | |||
| ATS | MAPO-36 | 1D straight channel | 12-ring (6.5 × 7.5) |
| SSZ-55 | |||
| BEA | Beta | 3D disordered channel | 12-ring |
| CHA | SAPO-34 | 3D cavity-window | 8-ring (3.8 × 3.8) |
| MAPO-34 | |||
| SSZ-13 | |||
| SAPO-47 | |||
| DDR | Sigma-1 | 2D cavity-window | 8-ring (3.6 × 4.4) |
| ZSM-58 | |||
| ERI | SAPO-17 | 3D cavity-window | 8-ring (3.6 × 5.1) |
| ETL | EU-12 | 2D sinusoidal channel | 8-ring (n.r.) |
| FAU | Zeolite X | 3D cavity-window | 12-ring (7.4 × 7.4) |
| Zeolite Y | |||
| FER | Ferrierite | 2D channel | 10-ring (4.2 × 5.4) |
| ZSM-35 | 8-ring (3.5 × 4.8) | ||
| GON | GUS-1 | 1D channel | 12-ring (5.4 × 6.8) |
| 8-ring (4.3 × 1.3) | |||
| IRN | ITQ-49 | 1D cavity-window | 8-ring (n.r.) |
| LEV | MAPO-35 | 2D cavity-window | 8-ring (3.6 × 4.8) |
| RUB-1 | |||
| LTA | Linde Type A | 3D cavity-window | 8-ring (4.1 × 4.1) |
| SAPO-42 | |||
| MEL | ZSM-11 | 3D straight channel | 10-ring (5.3 × 5.4) |
| Silicalite-2 | |||
| MFI | ZSM-5 | 3D straight/zigzag channel | 10-ring (5.1 × 5.5) |
| Silicalite-1 | 10-ring (5.3 × 5.6) | ||
| MOR | Mordenite | 1D channel-pocket | 12-ring (6.5 × 7.0) |
| 8-ring (2.6 × 5.7) | |||
| MTT | ZSM-23 | 1D straight channel | 10-ring (4.5 × 5.2) |
| EU-13 | |||
| MTW | ZSM-12 | 1D straight channel | 12-ring (5.6 × 6.0) |
| MWW | MCM-22 | 2D channel | 10-ring (4.0 × 5.5) |
| 10-ring (4.1 × 5.1) | |||
| RTH | RUB-13 | 2D cavity-window | 8-ring (3.8 × 4.1) |
| SSZ-50 | 8-ring (2.5 × 5.6) | ||
| RHO | DNL-6 | 3D cavity-window | 8-ring (3.6 × 3.6) |
| SFH | SSZ-53 | 1D straight channel | 14-ring (6.4 × 8.7) |
| SZR | SUZ-4 | 3D straight channel | 10-ring (4.1 × 5.2) |
| 8-ring (3.2 × 4.8) | |||
| 8-ring (3.0 × 4.8) | |||
| TON | Theta-1 | 1D straight channel | 10-ring (4.6 × 5.7) |
| ZSM-22 | |||
| TUN | TNU-9 | 3D channel | 10-ring (5.6 × 5.5) |
| 10-ring (5.4 × 5.5) | |||
| 10-ring (5.1 × 5.5) |
Structure data from the Database of Zeolite Structures by the International Zeolite Association.45 Abbreviations: n.r. Not reported.
Table 2. Largest and Main Product Groups Eluted from a Few Central Zeolites/Zeotypes during MTH Operationa.
| Three-letter structure code | Largest product group eluted from catalyst during MTH operation | Main product group during MTH operation |
|---|---|---|
| CHA | Linear aliphatics | C2–C4 olefins |
| TON | Monobranched aliphatics | C5+ aliphatics |
| MFI | Tetramethyl benzene | Propene, BTX, Gasolineb |
| BEA | Hexamethyl benzene | Aromatics, C3+ aliphatics |
MFI and CHA catalysts are in commercial use. Data from Teketel et al.46
High temperature, moderate pressure (∼450 °C, 1 atm) favor olefins production. Moderate temperature, high pressure (∼350 °C, 20 atm) favor paraffins/aromatics production.
Moving from the hydrocarbon upgrading processes, significant progress has been achieved in recent years regarding C1 catalysis using bifunctional solid catalysts.3,4,18−20 Of particular interest is the direct pathway from CO/CO2/H2 feeds via an oxygenate (methanol/dimethyl ether/ketene) to form selected hydrocarbon product mixtures in a single reactor. The direct pathway requires bifunctional catalysts, i.e., a function to convert syngas or CO2/H2 to oxygenates and a function to convert oxygenates to hydrocarbons. It is an example of tandem catalysis, which refers to the presence of bi-/multi-catalytic functions operating under different mechanisms in a single reactor for chemical transformations.21,22 It offers opportunities to overcome the thermodynamic limitation and to intensify chemical processes through reducing the number of separation steps and workup of products. However, numerous challenges remain to be resolved, for instance, compatibility of reaction conditions, catalyst activation, and deactivation. Tandem catalysis has a longer heritage in homogeneous and heterogeneous catalytic systems in liquid batch phase, while heterogeneous catalytic systems in gaseous continuous phase are relatively limited.21−25 A classic example of bifunctional solid catalysts operating in continuous gaseous flow conditions is the hydrocracking/hydroisomerization of alkanes, in which the metal sites (e.g., Pt) and acid sites (e.g., zeolites) are responsible for (de)hydrogenation and cracking/isomerization reactions, respectively.26
Returning to the title process, the addition of zeolites/zeotypes such as SAPO-34 to the COx hydrogenation catalysts creates a tandem catalytic system and shifts the thermodynamic equilibrium by applying Le Chatelier’s principle, as reflected in Figure 1. Using CO2 hydrogenation as a case study, Figure 1a shows that the yield of methanol is limited by thermodynamics over a range of reaction temperatures and pressures. The equilibrium yields of methanol are modeled using Aspen Plus according to the RGibbs block (PB-RM property method). Two equilibrium reactions are considered (CO2 to methanol or CO), and the products in the outlet stream are limited to CO2, H2, CO, methanol, and H2O. Experimental trends observed in Figure 1 are in accordance with thermodynamic predictions, but some data are seemingly above the equilibrium limit. The reason may be slightly different reaction conditions than those reported (e.g., higher P, lower T, different partial pressures). Thermodynamics suggest that higher methanol yields could be obtained at lower temperatures and higher pressures; thus, the conversion of CO2 to methanol is typically performed at 225 to 275 °C and 30 to 50 bar. The addition of zeolites/zeotypes to the CO2 hydrogenation catalysts converts the synthesized methanol directly to dimethyl ether or hydrocarbon products, therefore maintaining a relatively low methanol partial pressure in the reactor. The relatively low methanol partial pressure shifts the product selectivity from CO toward dimethyl ether or hydrocarbons, resulting in the theoretical yield of methanol (based on CO selectivity) to be higher than the thermodynamic equilibrium, as illustrated in Figure 1b. However, the accumulation of water with every step of reaction is thermodynamically the major disadvantage of the combined processes. Interestingly, zeolites in the reactor could also act as solid adsorbents to reduce steam partial pressure and further intensify the direct processes.27
Figure 1.
Thermodynamic considerations for CO2 hydrogenation to methanol, dimethyl ether, and hydrocarbons. (a) CO2 hydrogenation to methanol at 200 to 400 °C, 20 to 50 bar, and H2/CO2 = 3. Methanol yields were calculated from the data set in a recent review by Jiang et al.28 (b) CO2 hydrogenation to methanol, dimethyl ether, or hydrocarbons at 200 to 400 °C, 30 bar, and H2/CO2 = 3. Equivalent methanol yields from methanol, dimethyl ether, or hydrocarbon products were calculated from the data sets in recent reviews by Jiang et al., Saravanan et al., and Zhou et al. respectively.3,28,29
Although there is a clear motivation for the combination of zeolites/zeotypes and COx hydrogenation catalytic functions to utilize the tandem catalysis concept, there is an additional complexity to the “usual” operations of the zeolites/zeotypes in individual reactions. For instance, the influence of a relatively high steam partial pressure and the effect of reactants and products from a previous reaction on the kinetics of the next reaction may be significant. Considering again the case of CO2 hydrogenation to methanol and its subsequent conversion to lower olefins in a single pass, CO2 hydrogenation to methanol has a typical reaction window of 225 to 275 °C and 20 to 50 bar, but the methanol to olefins process (MTO) has a typical reaction window above 350 °C and 1 bar. Besides reaction temperatures and pressures, the reactant partial pressures are also different. MTO is commonly carried out with a feed consisting of methanol and an inert gas. However, in the direct conversion of CO2 to lower olefins, the zeolites/zeotypes would be exposed to a gas mixture of CO2, H2, CO, H2O, and oxygenates.
The partial pressure of the reactants and products after CO2 hydrogenation is presented in terms of % outlet pressure in Figure 2, and both absolute partial pressures and relative partial pressures to methanol could further complicate the catalytic function of zeolites/zeotypes in the MTO reaction. From Figure 2, the methanol partial pressure increased with reaction pressure as governed by thermodynamics and accounted for up to 5% of the outlet stream. This is lower than what is typically used for methanol to hydrocarbons (MTH), and a much diluted methanol feed going to the zeolite/zeotype is in principle good for catalyst stability but compromises productivity rates. The partial pressure in the outlet stream followed the order of H2 > CO2 > H2O > CO or methanol. According to Figure 2a, the amount of CO2 in the outlet stream decreased from 25 to 20% as the methanol content increased from 0 to 5%. A feed ratio of H2/CO2 = 3 is required to produce equimolar amounts of methanol and water, so CO2 and H2 made up 25 and 75% of the feed stream, respectively. CO2 conversion is limited by equilibrium conversion and depends on kinetics, so a higher reaction pressure leads to higher conversion, and up to 27% conversion at 50 bar is reported. This results in a range of 20 to 25% CO2 in the outlet stream and corresponds to a CO2 to methanol ratio between 5 and 1000. The same reasoning applies for H2 in the outlet stream (Figure 2b) as H2 in the outlet stream decreased from 75 to 65% with an increase in methanol. The ratio of H2 to methanol was more drastic and ranged from 15 to 4000. On the other hand, the difference in the CO and methanol contents in the outlet stream was smaller, as revealed in Figure 2c. The correlation between CO and methanol was in general positive, with ratios of CO to methanol less than 3 and preferably less than 1. This is understandable since CO and methanol are products of competing reactions. These ratios of CO to methanol are lower than what is typically used for methanol and dimethyl ether (DME) carbonylation (vide infra, Section 4). Last but not least, Figure 2d shows the positive correlation between H2O and methanol, and the ratio of H2O to methanol has to be at least 1. Since the competing reverse water gas shift (RWGS) reaction also produces water, the ratios of H2O to methanol were between 1 and 4. The competitive sorption/reaction of steam and methanol over Brønsted acid sites could potentially affect the overall reaction rates, as well as, for example, relative rates of hydrogen transfer/alkene methylation. Therefore, it is worthwhile to dive deeper into the bifunctional catalysts for COx hydrogenation from the viewpoint of the zeolites/zeotypes catalytic function.
Figure 2.
Partial pressures (in terms of % outlet pressure) of (a) CO2, (b) H2, (c) CO, and (d) H2O as a function of methanol partial pressure at the outlet of a CO2 hydrogenation reactor operating at 200 to 400 °C, 20 to 50 bar, and H2/CO2 = 3. They were calculated from the data set in a recent review by Jiang et al.28
Patent literature reveals early industrial efforts in combining the conversion of syngas to oxygenates with oxygenate conversion to hydrocarbons. While the first efforts (from 1981) focused on a two-reactors-in-series approach, the first single-reactor tandem process was patented in 2001. An overview of patent literature is provided in Table 3. Challenges of this tandem process include the identification of suitable reaction conditions for both the operation of catalytic functions and the inhibition of side reactions. Early investigation on bifunctional catalysts for the direct conversion of synthesis gas via oxygenates concluded with paraffins being the main products,30−32 and the first success stories on light olefins formation were reported in 2016 independently by the groups of Bao and Wang.33,34 Bao and co-workers developed a bifunctional catalyst consisting of Zn-Cr oxide and mesoporous SAPO-34, and proposed that the new process Ox-Zeo operated via ketene as the key intermediate.33 Wang and co-workers reported bifunctional catalysts consisting of Zr-Zn oxides and SAPO-34 as active components for CO activation and C–C coupling steps, respectively, with methanol and dimethyl ether being proposed to be the reaction intermediates.34 Although different oxygenate intermediates were suggested in the two studies, SAPO-34 was used in both cases to convert the oxygenates to lower olefins, understandably due to its commercial success in the methanol to olefins (MTO) process.35−37
Table 3. An Overview of Patent Literature Describing Oxygenate-Mediated Processes for the Formation of Hydrocarbons from COx (x = 1, 2) and H2a.
| Hydrocarbon selectivity (%C) |
||||||||||||
|---|---|---|---|---|---|---|---|---|---|---|---|---|
| Year | Company | Process (Feed) | Methanol synthesis catalyst | MTH cat. | T (°C) | P (bar) | X (% COx) | CH4 | C2–4= | C2–40 | C5+ | ref. |
| 1981 | Metallgesellschaft | 2-stage | Cu/Zn/V | H-ZSM-5 | T1: 250 | P1: 58 | 40 | n.r. | n.r. | n.r. | n.r. | (47) |
| COx, H2 | T2: 375 | P2: 58 | ||||||||||
| 2001 | UOP/Norsk Hydro | 3-stage | Cu/Zn/Cr | H-SAPO-34 | T1: 150–450 | P1: 1–103 | n.r. | <1 | 96 | 1 | 2 | (48) |
| CH4 | H-SAPO-17 | T2: 200–700 | P2: 1–5 | |||||||||
| 2001 | KAIST | H2/CO2 = 3 | Cu/Zn/Zr | H-SAPO-5 | 350 | 28 | 32 | 4 | 89 (43% C4) | 7 | (49) | |
| 2001 | KAIST | H2/CO2 = 3 | Cu/Zn/Zr | Cu-SAPO-5 | 350 | 28 | 28 | 4 | 85 (49% C4) | 12 | (49) | |
| 2001 | KAIST | H2/CO2 = 3 | Cu/Zn/Zr | H-SAPO-34 | 400 | 28 | 35 | 4 | 96 (52% C3) | 1 | (49) | |
| 2001 | KAIST | H2/CO2 = 3 | Cu/Zn/Zr | Cu-SAPO-34 | 400 | 28 | 40 | 2 | 97 (55% C3) | 1 | (49) | |
| 2007 | Japan Gas Synthesize | H2/CO = 2 | Pd/Ca/Si | β-zeolite | 375 | 41 | 74 | 7 | 0 | 84 | 9 | (50) |
| 2007 | Japan Gas Synthesize | H2/CO = 2 | Pd/Ca/Si | β-zeolite | 375 | 51 | 87 | 8 | 0 | 87 | 5 | (50) |
| 2007 | Japan Gas Synthesize | H2/CO = 2 | Pd/Ca/Si | Y-zeolite | 375 | 51 | 88 | 10 | 0 | 84 | 6 | (50) |
| 2011 | Chevron | H2/CO = 2 | Zn/Cr | H-ZSM-5 | 350 | 41 | 30 | 2 | 3 | 26 | 69 | (51) |
| 2011 | Chevron | H2/CO = 2 | Zn/Cr | Ga-ZSM-5 | 350 | 41 | 30 | 1 | 5 | 17 | 77 | (51) |
| 2013 | Chevron | H2/CO/CO2 = 67/25/8 | Zn/Cr | H-ZSM-5 | 380 | 81 | 23 | 1 | 1 | 26 | 72 | (52) |
| 2014 | Sinopec | H2/CO/CO2 = 50/42/8 | Cu/Al/K/C/Mn/Si | H-SAPO-34 | 290 | 10 | 91 | n.r. | 37 | n.r. | n.r. | (53) |
| 2016 | Pioneer Energy | H2/CO = 2 | Cu/Zn | Zn-ZSM-5 | 225–300 | 20 | 87 | 4 | n.r. | 84 | 9 | (54) |
| 2016 | Pioneer Energy | H2/CO = 2 | Cu/Zn/Al | Zn-ZSM-5 | 200–420 | 20 | 67 | 13 | n.r. | 84 | 3 | (54) |
| 2016 | Pioneer Energy | 2-stage | Cu/Zn/Al | Zn-ZSM-5 | T1: 250 | P1: 20 | 61 | 8 | 0 | 54 | 39 | (54) |
| H2/CO = 2 | T2: 400 | P2: 1 | ||||||||||
| 2016 | Pioneer Energy | 2-stage | Cu/Zn/Al | Zn-ZSM-5 | T1: 175–275 | P1: 20 | 65 | 6 | 2 | 33 | 59 | (54) |
| H2/CO = 1 | T2: 450 | P2: 1 | ||||||||||
| 2018 | Dow | H2/CO = 3 | Cu/Zn/Al | H-SAPO-34 | 330 | 20 | 61 | 4 | <3 | 93 | 0 | (55) |
| 2018 | Dow | H2/CO = 3 | Cu/Zn/Al | H-SAPO-34 | 340 | 20 | 69 | 2 | 0 | 98 | 0 | (55) |
| 2018 | Dow | H2/CO = 3 | Cu/Zn/Al | H-SAPO-34 | 350 | 20 | 66 | 2 | 0 | 97 | 0 | (55) |
| 2018 | Dow | H2/CO = 3 | Cu/Zn/Al | H-SAPO-34 | 360 | 20 | 65 | 2 | 0 | 97 | 0 | (55) |
| 2018 | Dow | H2/CO = 3 | Cu/Zn/Al | H-SAPO-34 | 370 | 20 | 59 | 2 | 0 | 97 | 0 | (55) |
| 2018 | Dow | H2/CO = 3 | Cu/Zn/Al | H-SAPO-34 | 380 | 20 | 51 | 2 | 0 | 97 | 0 | (55) |
| 2018 | Dow | H2/CO = 3 | Cu/Zn/Al | H-SAPO-34 | 400 | 20 | 40 | 6 | 0 | 94 | 0 | (55) |
| 2018 | Dow | H2/CO = 3 | Cu/Zn/Al | H-SAPO-34 | 410 | 20 | 34 | 8 | 0 | 91 | 0 | (55) |
| 2018 | Dow | H2/CO = 3 | Cu/Zn/Al | H-SAPO-34 | 430 | 20 | 19 | 22 | 0 | 77 | 0 | (55) |
| 2018 | Dow | H2/CO = 3 | Cu/Zn/Al | H-SAPO-34 | 380 | 5 | 3 | 34 | <2 | 64 | 0 | (55) |
| 2018 | Dow | H2/CO = 3 | Cu/Zn/Al | H-SAPO-34 | 380 | 35 | 74 | 4 | 0 | 96 | 0 | (55) |
| 2018 | Dow | H2/CO = 3 | Cu/Zn/Al | H-SAPO-34 | 380 | 50 | 81 | 5 | <1 | 94 | 0 | (55) |
| 2018 | Dow | H2/CO = 2 | Cu/Zn/Al | H-SAPO-34 | 380 | 50 | 82 | 4 | 0 | 96 | 0 | (55) |
| 2018 | Dow | H2/CO = 4 | Cu/Zn/Al | H-SAPO-34 | 380 | 50 | 80 | 6 | 0 | 94 | 0 | (55) |
| 2018 | Dow | H2/CO = 5 | Cu/Zn/Al | H-SAPO-34 | 380 | 50 | 80 | 7 | <1 | 92 | 0 | (55) |
| 2018 | Dow | H2/CO = 6 | Cu/Zn/Al | H-SAPO-34 | 380 | 50 | 79 | 8 | <1 | 92 | 0 | (55) |
| 2018 | Dow | H2/CO = 3 | Cu/Zn/Al | H-SAPO-34 | 380 | 40 | 72 | 5 | 0 | 94 | 0 | (55) |
| 2018 | Dow | H2/CO/CO2 = 67/23/10 | Cu/Zn/Al | H-SAPO-34 | 380 | 40 | 58 | 3 | <1 | 96 | 0 | (55) |
| 2018 | Dow | H2/CO/CO2 = 60/20/20 | Cu/Zn/Al | H-SAPO-34 | 380 | 40 | 46 | 3 | <1 | 96 | 0 | (55) |
| 2018 | Dow | H2/CO2 = 3 | Cu/Zn/Al | H-SAPO-34 | 400 | 40 | 2b | 23 | 0 | 70 | 0 | (55) |
| 2018 | Dow | H2/CO2 = 3 | Cu/Zn/Al | H-SAPO-34 | 375 | 40 | 5b | 6 | 0 | 85 | 4 | (55) |
| 2018 | Dow | H2/CO2 = 3 | Cu/Zn/Al | H-SAPO-34 | 350 | 40 | 8b | 5 | 0 | 83 | 6 | (55) |
| 2018 | Dow | H2/CO2 = 3 | Cu/Zn/Al | H-SAPO-34 | 325 | 40 | 4b | 6 | 0 | 34 | 0 | (55) |
| 2018 | Dow | H2/CO2 = 3 | Cu/Zn/Al | H-SAPO-34 | 300 | 40 | 3b | 0 | 0 | 1 | 0 | (55) |
| 2018 | Dow | H2/CO2 = 3 | Cu/Zn/Al | H-SAPO-34 | 350 | 28 | 2b | 7 | 0 | 69 | 0 | (55) |
| 2018 | Dow | H2/CO2 = 3 | Cu/Zn/Al | H-SAPO-34 | 350 | 2 | 0b | 4 | 0 | 84 | 4 | (55) |
| 2018 | Dow | H2/CO2 = 1 | Cu/Zn/Al | H-SAPO-34 | 350 | 40 | 2b | 8 | 0 | 84 | 5 | (55) |
| 2018 | Dow | H2/CO2 = 10 | Cu/Zn/Al | H-SAPO-34 | 350 | 40 | 25b | 7 | 0 | 91 | 5 | (55) |
| 2018 | Dow | H2/CO = 3 | Cr/Zn | H-SAPO-34 | 400 | 50 | 67 | 13 | 3 | 82 | 2 | (55) |
| 2018 | Dow | H2/CO = 3 | Cr/Zn | H-SAPO-34 | 400 | 70 | 82 | 14 | 1 | 83 | 2 | (55) |
| 2018 | Dow | H2/CO = 3 | Cu/Zn/Al | H-SAPO-5/H-SAPO-34 | 380 | 50 | 79 | 4 | 0 | 83 | 4 | (55) |
| 2018 | Dow | H2/CO = 3 | Cu/Zn/Al | H-SAPO-18 | 400 | 50 | 69 | 9 | 0 | 89 | 2 | (55) |
| 2018 | Dow | H2/CO = 3 | Cu/Zn/Al | β-zeolite | 385 | 50 | 71 | 13 | 0 | 86 | 2 | (55) |
| 2019 | Dow | H2/CO = 2 | Cr/Zn | H-SAPO-34 | 400 | 20 | 43 | 8 | 51c | 36c | <5 | (56) |
| 2019 | Dow | H2/CO = 1.5 | Cr/Zn | H-SAPO-34 | 400 | 20 | 35 | 11 | 45c | 38c | <5 | (56) |
| 2019 | Dow | H2/CO = 3 | Cr/Zn | H-SAPO-34 | 400 | 20 | 50 | 10 | 20c | 61c | <9 | (56) |
| 2019 | Dow | H2/CO = 3 | Cr/Zn | H-SAPO-34 | 400 | 50 | 76 | 6 | 5c | 76c | <12 | (56) |
| 2019 | Dow | H2/CO = 2 | Cr/Zn | H-SAPO-34 | 400 | 70 | 38 | 16 | 1c | 51c | <3 | (56) |
| 2019 | Dow | H2/CO = 3 | Cr/Zn | H-SAPO-34 | 400 | 70 | 80 | 14 | 2c | 71c | <12 | (56) |
| 2019 | Dow | H2/CO = 2 | Cr/Zn | H-SAPO-34 | 450 | 20 | 32 | 43 | 2c | 51c | <5 | (56) |
| 2020 | DICP | H2/CO = 2 | Cu/Zn/Al | H-SSZ-13 | 365 | 20 | 9b | 2 | 38 | 51 | 10 | (57) |
| 2020 | DICP | H2/CO = 2 | Pd/Zn/Cr | H-SAPO-34 | 400 | 25 | 13b | 2 | 16 | 80 | 3 | (57) |
| 2020 | DICP | H2/CO = 2 | Cu/Zn/Al/Cr/Co | H-SAPO-34 | 370 | 20 | 9b | 3 | 67 | 19 | 11 | (57) |
| 2020 | DICP | H2/CO = 2 | Cu/Zn | H-ZSM-5 | 400 | 25 | 8b | 1 | 50 | 41 | 8 | (57) |
| 2020 | DICP | H2/CO = 2 | Co/Zn/Al | H-SSZ-13 | 400 | 25 | 10b | 8 | 64 | 13 | 17 | (57) |
| 2020 | DICP | H2/CO = 2 | Co/Zn/Ti | H-SAPO-34 | 400 | 20 | 5b | 3 | 57 | 32 | 8 | (57) |
| 2020 | DICP | H2/CO = 2 | Cu/Zn/Mo | H-SSZ-13 | 360 | 8 | 4b | 4 | 50 | 42 | 4 | (57) |
| 2020 | DICP | H2/CO = 2 | Cu/Zn/V | H-SSZ-13 | 360 | 8 | 4b | 4 | 53 | 31 | 12 | (57) |
| 2020 | DICP | H2/CO = 2 | Cu/Zn/Mn | H-SSZ-13 | 360 | 8 | 4b | 4 | 63 | 25 | 8 | (57) |
| 2020 | DICP | H2/CO = 2 | Pd/Zn/Zr | H-SSZ-13 | 380 | 15 | 4b | 4 | 56 | 10 | 30 | (57) |
| 2020 | DICP | H2/CO = 2 | Mo/Zn/Zr/Al | H-SAPO-34 | 400 | 30 | 5b | 7 | 58 | 26 | 9 | (57) |
| 2020 | DICP | H2/CO = 2 | Zn/Ti/Ga | H-SAPO-34 | 400 | 35 | 4b | 3 | 57 | 33 | 7 | (57) |
| 2020 | DICP | H2/CO = 2 | Co/Zn/Ti/Mn | H-SAPO-34 | 400 | 35 | 7b | 5 | 68 | 25 | 2 | (57) |
| 2020 | DICP | H2/CO = 2 | Ca/Mn/Al/Cu/Zn | H-SSZ-13 | 375 | 37 | 16b | 3 | 18 | 74 | 5 | (57) |
| 2020 | DICP | H2/CO = 2 | Ca/Mn/Cr/Co | H-MgSAPO-34 | 415 | 30 | 16b | 3 | 16 | 75 | 6 | (57) |
| 2020 | DICP | H2/CO = 2 | Fe/Zn/Al/Cr/Na | H-SAPO-34 | 400 | 30 | 9b | 6 | 66 | 21 | 7 | (57) |
| 2020 | DICP | H2/CO = 2 | Fe/Zn/Al/Cr/Mg | H-SAPO-34 | 400 | 30 | 10b | 3 | 73 | 17 | 8 | (57) |
| 2020 | DICP | H2/CO = 2 | Fe/Zn/Al/Cr/K | H-SAPO-34 | 400 | 30 | 8b | 9 | 76 | 9 | 6 | (57) |
| 2020 | DICP | H2/CO = 2 | Fe/Zn/Al/Cr/Ce | H-SAPO-34 | 400 | 30 | 8b | 11 | 67 | 18 | 3 | (57) |
| 2020 | DICP | H2/CO = 2 | Fe/Zn/Al/Cr/La | H-SAPO-34 | 400 | 30 | 11b | 2 | 73 | 13 | 12 | (57) |
| 2020 | DICP | H2/CO = 2 | Fe/Zn/Ga/Cr/Mn | H-SAPO-34 | 400 | 30 | 18b | 3 | 75 | 20 | 2 | (57) |
| 2020 | DICP | H2/CO = 2 | Fe/Zn/Ge/Cr/Mn | H-SAPO-34 | 400 | 30 | 16b | 2 | 70 | 21 | 7 | (57) |
| 2020 | DICP | H2/CO = 2 | Fe/Zn/Zr/Cr/Mn | H-SAPO-34 | 400 | 30 | 19b | 4 | 75 | 17 | 4 | (57) |
| 2020 | DICP | H2/CO = 2 | Fe/Zn/In/Cr/Mn | H-SAPO-34 | 400 | 30 | 19b | 6 | 66 | 21 | 7 | (57) |
| 2020 | DICP | H2/CO = 2 | Fe/Mn/Al/Cr/Na | H-SAPO-34 | 400 | 30 | 9b | 6 | 64 | 25 | 6 | (57) |
| 2020 | DICP | H2/CO = 2 | Fe/Mn/Al/Cr/Mg | H-SAPO-34 | 400 | 30 | 10b | 3 | 73 | 17 | 8 | (57) |
| 2020 | DICP | H2/CO = 2 | Fe/Mn/Al/Cr/K | H-SAPO-34 | 400 | 30 | 8b | 9 | 74 | 11 | 6 | (57) |
| 2020 | DICP | H2/CO = 2 | Fe/Mn/Al/Cr/Ce | H-SAPO-34 | 400 | 30 | 10b | 16 | 60 | 15 | 9 | (57) |
| 2020 | DICP | H2/CO = 2 | Fe/Mn/Al/Cr/La | H-SAPO-34 | 400 | 30 | 10b | 12 | 70 | 16 | 2 | (57) |
| 2021 | SXICC | H2/CO = 1.5 | Zn/Mn/In/Ga/Ni/Cr/K | H-SAPO-34 | 420 | 20 | 14b | 7 | 84 | 7 | 2 | (58) |
| 2021 | SXICC | H2/CO = 2.5 | Zn/Mn/In/Ga/Ni/Cr/K | H-SAPO-34 | 400 | 30 | 23b | 7 | 80 | 11 | 2 | (58) |
| 2021 | SXICC | H2/CO = 2.5 | Zn/Mn/In/Ga/Ni/Cr/K | H-SAPO-34 | 380 | 50 | 14b | 4 | 84 | 11 | 2 | (58) |
| 2021 | SXICC | H2/CO = 2.5 | Zn/Mn/In/Ga/Ni/Cr/K | H-SAPO-34 | 390 | 50 | 21b | 6 | 80 | 10 | 4 | (58) |
| 2021 | SXICC | H2/CO = 2.5 | Zn/Mn/In/Ga/Ni/Cr/K | H-SAPO-34 | 400 | 50 | 18b | 9 | 78 | 11 | 2 | (58) |
| 2021 | SXICC | H2/CO = 2.5 | Zn/Mn/In/Ga/Ni/Cr/K | H-SAPO-34 | 390 | 60 | 30b | 8 | 81 | 9 | 2 | (58) |
| 2021 | SXICC | H2/CO = 3.5 | Zn/Mn/In/Ga/Ni/Cr/K | H-SAPO-34 | 400 | 60 | 29b | 10 | 71 | 18 | 1 | (58) |
| 2021 | SXICC | H2/CO = 2.5 | Zn/Mn/In/Ga/Ni/Cr/K | H-SSZ-13 | 390 | 50 | 19b | 4 | 85 | 6 | 4 | (58) |
| 2021 | SXICC | H2/CO = 1.5 | Zn/Mn/In/Ga/Ni/Cr/K | H-ZSM-5 | 380 | 80 | 16b | 6 | 82 | 9 | 3 | (58) |
| 2021 | SXICC | H2/CO = 2.5 | Zn/Mn/In/Ga/Ni/Cr/K | H-ZSM-5 | 410 | 50 | 20b | 6 | 76 | 16 | 3 | (58) |
| 2021 | SXICC | H2/CO = 3.5 | Zn/Mn/In/Ga/Ni/Cr/K | H-ZSM-5 | 380 | 70 | 26b | 8 | 75 | 15 | 2 | (58) |
| 2021 | TYUT | H2/CO = 2 | Zn/Zr | H-SAPO-34/Silicate-1 | 410 | 20 | 9b | 12 | 80 | 6 | 2 | (59) |
| 2021 | DICP | 2-stage | Zn/Al | H-SAPO-34 | T1: 390 | 40 | 28 | n.r. | 82 | n.r. | n.r. | (60) |
| H2/CO = 1 | T2: 390 | |||||||||||
| 2021 | DICP | 2-stage | Zn/Al/Cr | H-SAPO-18 | T1: 390 | 40 | 33 | n.r. | 80 | n.r. | n.r. | (60) |
| H2/CO = 1 | T2: 390 | |||||||||||
| 2021 | DICP | 2-stage | Zn/Al/Zr | H-SSZ-13 | T1: 390 | 40 | 31 | n.r. | 78 | n.r. | n.r. | (60) |
| H2/CO = 1 | T2: 390 | |||||||||||
| 2021 | DICP | 2-stage | Cu/Zn/Al | H-SAPO-34 | T1: 260 | 40 | 58 | n.r. | 76 | n.r. | n.r. | (60) |
| H2/CO = 1 | T2: 390 | |||||||||||
| 2021 | DICP | 2-stage | Cu/Zn/Al + Al2O3 | H-SAPO-34 | T1: 260 | 40 | 46 | n.r. | 77 | n.r. | n.r. | (60) |
| H2/CO = 1 | T2: 390 | |||||||||||
| 2022 | SXICC | H2/CO2 = 3 | Ga/Zr | H-SSZ-13 | 350 | 30 | 23b | 4 | 2 | 94 | 0 | (61) |
| 2022 | SXICC | H2/CO2 = 3 | Cr | H-SAPO-34 | 370 | 5 | 9b | 1 | 96 | 2 | 1 | (62) |
| 2022 | SXICC | H2/CO2 = 3 | Cr/Zn | H-SAPO-34 | 370 | 5 | 5b | 5 | 89 | 3 | 2 | (62) |
| 2022 | SXICC | H2/CO2 = 3 | Cr/In | H-SAPO-34 | 370 | 5 | 7b | 8 | 87 | 3 | 2 | (62) |
| 2022 | SXICC | H2/CO2 = 3 | Cr/Al | H-SAPO-34 | 370 | 5 | 5b | 5 | 85 | 3 | 2 | (62) |
| 2022 | SXICC | H2/CO2 = 3 | Cr/Zr | H-SAPO-34 | 370 | 5 | 5b | 7 | 85 | 5 | 3 | (62) |
| 2022 | Sinopec | H2/CO = 1 | Cr | H-ZSM-5 | 395 | 20 | 18 | n.r. | n.r. | n.r. | 80d | (63) |
| 2022 | Sinopec | H2/CO = 1 | Mn | H-ZSM-5 | 395 | 20 | 19 | n.r. | n.r. | n.r. | 82d | (63) |
| 2022 | Sinopec | H2/CO = 1 | Ce | H-ZSM-5 | 395 | 20 | 16 | n.r. | n.r. | n.r. | 73d | (63) |
| 2022 | Sinopec | H2/CO = 1 | Zr | H-ZSM-5 | 395 | 20 | 14 | n.r. | n.r. | n.r. | 84d | (63) |
| 2022 | Sinopec | H2/CO = 1 | Zn/Cr | H-ZSM-5 | 395 | 20 | 18 | n.r. | n.r. | n.r. | 79d | (63) |
| 2022 | Sinopec | H2/CO = 1 | Zn/Mn | H-ZSM-5 | 395 | 20 | 20 | n.r. | n.r. | n.r. | 78d | (63) |
| 2022 | Sinopec | H2/CO = 1 | Zr/In | H-ZSM-5 | 395 | 20 | 15 | n.r. | n.r. | n.r. | 82d | (63) |
| 2022 | Sinopec | H2/CO = 1 | Cr/Mn | H-ZSM-5 | 395 | 20 | 18 | n.r. | n.r. | n.r. | 82d | (63) |
| 2022 | Sinopec | H2/CO = 1 | Zr/Ce | H-ZSM-5 | 395 | 20 | 15 | n.r. | n.r. | n.r. | 81d | (63) |
| 2022 | Sinopec | H2/CO = 1 | Mn | H-ZSM-11 | 395 | 20 | 20 | n.r. | n.r. | n.r. | 81d | (63) |
| 2022 | Sinopec | H2/CO = 1 | Mn | H-ZSM-5/Silicalite-1 | 395 | 20 | 17 | n.r. | n.r. | n.r. | 84d | (63) |
| 2022 | Sinopec | H2/CO = 1 | Mn | H-ZSM-5/Silicalite-2 | 395 | 20 | 17 | n.r. | n.r. | n.r. | 83d | (63) |
| 2022 | Sinopec | H2/CO = 1 | Mn | H-ZSM-11/Silicalite-2 | 395 | 20 | 17 | n.r. | n.r. | n.r. | 83d | (63) |
| 2022 | Sinopec | H2/CO = 1 | Mn | H-ZSM-11/Silicalite-1 | 395 | 20 | 18 | n.r. | n.r. | n.r. | 83d | (63) |
| 2022 | Sinopec | H2/CO = 1 | Mn | H-ZSM-5 | 350 | 80 | 33 | n.r. | n.r. | n.r. | 72d | (63) |
| 2022 | Sinopec | H2/CO = 0.5 | Mn | H-ZSM-5 | 395 | 80 | 28 | n.r. | n.r. | n.r. | 73d | (63) |
| 2022 | Sinopec | H2/CO = 4 | Mn | H-ZSM-5 | 350 | 50 | 25 | n.r. | n.r. | n.r. | 68d | (63) |
| 2022 | Sinopec | H2/CO = 1 | Mn | H-ZSM-5 | 395 | 50 | 20 | n.r. | n.r. | n.r. | 77d | (63) |
| 2022 | Sinopec | H2/CO = 1 | Mn | H-ZSM-5 | 450 | 50 | 23 | n.r. | n.r. | n.r. | 72d | (63) |
| 2022 | Sinopec | H2/CO = 0.5 | Mn | H-ZSM-5 | 395 | 40 | 20 | n.r. | n.r. | n.r. | 79d | (63) |
| 2022 | DICP | H2/CO = 2 | Zn | H-SAPO-34/18 | 410 | 35 | 24 | 5 | 86 | 5 | 4 | (64) |
| 2022 | DICP | H2/CO = 5.5 | Zn | H-SAPO-34/18 | 400 | 9 | 38 | 5 | 85 | 8 | 2 | (64) |
| 2022 | DICP | H2/CO = 3 | Zn | H-SAPO-34/18 | 380 | 45 | 24 | 3 | 85 | 7 | 5 | (64) |
| 2022 | DICP | H2/CO = 6 | Mn | H-SAPO-34/18 | 370 | 100 | 33 | 7 | 71 | 14 | 8 | (64) |
| 2022 | DICP | H2/CO = 3.5 | Ce | AlPO-34/18 | 470 | 15 | 28 | 4 | 77 | 11 | 8 | (64) |
| 2022 | DICP | H2/CO = 4.5 | Bi | H-SAPO-34/18 | 400 | 70 | 56 | 3 | 83 | 9 | 5 | (64) |
| 2022 | DICP | H2/CO = 6.5 | In | H-SAPO-34/18 | 380 | 25 | 37 | 7 | 76 | 8 | 9 | (64) |
| 2022 | DICP | H2/CO = 8.5 | Ga | AlPO-34/18 | 370 | 5 | 31 | 7 | 69 | 11 | 13 | (64) |
| 2022 | DICP | H2/CO = 1 | Zn/Cr | H-SSZ-13/39 | 370 | 35 | 28 | 5 | 74 | 7 | 14 | (64) |
| 2022 | DICP | H2/CO = 2.5 | Zn/Al | H-SAPO-34/18 | 410 | 50 | 37 | 2 | 88 | 5 | 5 | (64) |
| 2022 | DICP | H2/CO = 2.5 | Zn/Ga | H-GaAPO-34/18 | 430 | 30 | 58 | 5 | 82 | 6 | 7 | (64) |
| 2022 | DICP | H2/CO = 1 | Zn/In | H-GaSAPO-34/18 | 520 | 40 | 19 | 6 | 87 | 6 | 1 | (64) |
| 2022 | DICP | H2/CO = 0.5 | Mn/Cr | H-ZnAPO-34/18 | 480 | 90 | 14 | 7 | 68 | 12 | 13 | (64) |
| 2022 | DICP | H2/CO = 3 | Mn/Al | H-MgAPO-34/18 | 470 | 60 | 38 | 3 | 86 | 6 | 5 | (64) |
| 2022 | DICP | H2/CO = 1.5 | Mn/Zr | H-GaAPO-34/18 | 450 | 50 | 37 | 2 | 86 | 6 | 6 | (64) |
| 2022 | DICP | H2/CO = 2.5 | Mn/In | H-SAPO-34/18 | 450 | 50 | 34 | 3 | 88 | 6 | 3 | (64) |
| 2022 | DICP | H2/CO = 3.5 | Co/Al | H-SAPO-34/18 | 350 | 50 | 27 | 3 | 76 | 10 | 11 | (64) |
| 2022 | DICP | H2/CO = 6 | Fe/Al | H-SAPO-34/18 | 350 | 70 | 20 | 5 | 85 | 7 | 3 | (64) |
| 2022 | DICP | H2/CO = 4 | In/Al/Mn | H-SAPO-34/18 | 400 | 60 | 58 | 5 | 80 | 6 | 9 | (64) |
| 2022 | DICP | H2/CO = 4 | In/Ga/Mn | AlPO-34/18 | 400 | 40 | 33 | 7 | 73 | 7 | 13 | (64) |
Abbreviations: n.r. Not reported.
% Conversion to hydrocarbons, i.e., exclude (R)WGS activity.
C2–C3 selectivity.
Aromatics selectivity.
This Review highlights the opportunities, challenges, prior learnings, and future research in the zeolites/zeotypes catalytic function for the direct conversion of COx to valuable products via oxygenate intermediates. There are recent reviews regarding the advances in C1 catalysis,3,38,39 CO2 to methanol,28,40,41 and methanol to hydrocarbons (MTH),37,42−44 demonstrating the relevance and importance of these research fields. Nonetheless, there has not yet been a critical discussion of the zeolite/zeotype catalytic function in the context of COx valorization using tandem catalysis. Central to zeolite/zeotype functionality in this context is their Brønsted acid properties and, as will be further elaborated in the following sections, their ability to form and maintain a population of methoxy sites (Figure 3).
Figure 3.
Illustration of the direct conversion of COx to hydrocarbons via oxygenate intermediates. Three relevant intermediates, namely, methanol, DME, and ketene (acetyl species), are discussed in this review.
2. Methanol-Mediated COx Conversion to Hydrocarbons
2.1. A Short Introduction to the MTH Process
To date, most studies devoted to oxygenate-mediated conversion of COx to hydrocarbons focused on methanol as the intermediate. As background to the discussion of various process options below, we will first give a brief introduction to the methanol to hydrocarbons (MTH) process. The MTH process (subdivided into olefins (MTO), gasoline (MTG), propene (MTP), and aromatics (MTA)) is a major industrial process for the conversion of C1 feedstock (natural gas, biogas, coal) to light olefins (MTO, MTP) and has been a topic of more than 6000 research papers over the last 10 years (source: Web of Science, May 16, 2022. Search words: Methanol to hydrocarbons, or methanol to gasoline, or methanol to olefins). Its academic interest stems partly from its industrial significance, and partly from fundamental studies published before the major commercialization wave (1977–2010). Those studies enabled the distinction of competing reactions taking place inside the microporous network of the zeolite/zeotype, based on the product distribution in the reactor effluent (see below). As such, fundamental studies of the MTH reaction have contributed greatly to the fundamental understanding of zeolites and zeotypes in general. The early studies are well documented in prior review and perspective articles,35,65−69 and we will only present a few of the early key contributions that are particularly relevant for the methanol-mediated COx to hydrocarbons reaction.
Early studies showed that the MTH reaction is autocatalytic, and that alkenes and methyl benzenes act as cocatalytic species.70−80 They further suggested a sequential reaction scheme, where methanol is first converted to DME, followed by olefin formation and methylation to higher alkenes. Subsequently, alkenes are converted into alkanes and aromatics, and finally to coke.81 Water is a coproduct of all reaction steps. The “hydrocarbon pool” concept next presented by Dahl and Kolboe was based on isotopic labeling studies over H-SAPO-34 and broke with the sequential scheme, suggesting instead that all products are formed from a pool of hydrocarbon species confined inside the zeolite/zeotype.82,83 Subsequent studies combined isotopic labeling with a technique developed by the Guisnet group, in which the zeolite lattice is dissolved after reaction, without altering the hydrocarbons retained inside the zeolite crystals.84 Those studies led to a next groundbreaking discovery, i.e., that propene (and higher alkenes) is formed by alkene cracking as well as polymethyl benzene dealkylation, while ethene is formed predominantly by polymethyl benzene dealkylation, under typical MTH conditions (350–450 °C, 1 bar).85,86 This discovery is the basis of the generally accepted “dual cycle” reaction scheme, which combines sequential and pool-type reaction concepts (Figure 4). The dual cycle concept paved the way for using the C3/C2 ratio in the reactor effluent as a descriptor for the abundance of the alkene versus the arene cycle in various zeolites. The Hydrogen Transfer Index, HTI = [Cn– / (Cn= + Cn–)], where Cn= and Cn– are an alkene and alkane, respectively, with n carbon atoms, is another useful descriptor.87 It is based on the observation of a linear correlation between alkane and arene yields in the effluent of wide-pore zeolites, suggesting that aromatic compounds are mainly formed by hydrogen transfer between alkenes (and (cyclic) polyenes) rather than dehydrogenation reactions. These two descriptors are extremely useful for deriving mechanistic information from conventional screening tests of zeolite and zeotype catalysts.
Figure 4.

Dual cycle mechanism for the conversion of methanol to hydrocarbons, adapted with permission from ref (88). Copyright 2013 Elsevier. Mechanistic details in insets are extracted with permission from refs (89) and (90). Copyright 2017 Elsevier.
Zeolites/zeotypes are molecular sieves, and the kinetic diameter of the smallest pore sets an upper limit to the size of the majority of products that may elute from a material with a given topology (Tables 1–2). Beneath that size, product selectivity relies on a subtle interplay between diffusivity and intrinsic reaction rates, commonly described by the Thiele modulus.91 On the macroscopic level, correlations were found between MTH product selectivity and parameters such as crystal diameter, acid site density, reaction temperature, and methanol pressure. Yet, the plethora of products contained in the zeolite pores during operation, and in some cases a diversity of active site framework positions, complicate the elucidation of structure–function correlations. In that respect, the successful synthesis of Zeolite Socony Mobil–5 (ZSM-5) nanosheets by Ryoo and co-workers represented a turning point of zeolite research, by enabling comparisons of catalyst effectiveness factor and product selectivity between conventional crystals and nanosheets with negligible diffusion length.92,93 For MTH, it was shown that the reaction’s autocatalytic nature leads to an inverse effectiveness factor, but only to minor changes in product distribution between 1 and 3 μm ZSM-5 crystals and 2–4 nm ZSM-5 nanosheets with similar acid site density.94 Later studies showed that larger ZSM-5 crystals yield less arenes and more ethene compared to smaller crystals, indicating promotion of the arene cycle by the longer diffusion length resulting in higher abundance of retained arenes.95
While early studies agree that the MTH reaction occurs via the dual hydrocarbon pool mechanism, no consensus was reached on the formation of the initial C–C bond, and this received much attention in the past decade. Several mechanistic routes pointing to the direct C–C bond formation with key intermediates such as carbene, methane–formaldehyde, methoxymethyl cation, methyleneoxy cation and acetyl have been suggested.42 To highlight the role of CO in the initiation step, the Koch carbonylation mechanism is discussed in more detail here. According to the Koch carbonylation mechanism, methanol is first dehydrogenated to formaldehyde, which is then oxidized to CO and subsequently carbonylated with methoxy species to form acetic acid or methyl acetate (Figure 5). Likewise, methanol dehydrogenation has been found to constitute the first step in the conversion of olefins to aromatic products, and aromatic products to polyaromatic coke precursors (vide infra, Figure 4).89,90,96 As formaldehyde is proposed to be the origin of CO in the MTH reaction, a deeper understanding of the significance of both formaldehyde and CO is relevant for the discussion on the integration of MTH and COx hydrogenation.
Figure 5.
Proposed mechanistic routes to the first direct C–C bond formation and first olefin formation via Koch carbonylation. (a) The surface methoxy group undergoes carbonylation to form a surface acetyl species or ketene, which is the first C–C-containing intermediate.104,105 (b) The surface acetyl group or ketene could then propagate with methanol and decarbonylate to form the first ethene or propene.106,107,109 (c) The surface acetyl group could also react with formaldehyde to form unsaturated carboxylic acid intermediates, which decarboxylate to form the first ethene or propene.102 A propadienone resonance structure might exist for Path (c), analogous to the methyl-substituted ketenes in Path (b).
Formaldehyde is a key intermediate in two of the proposed initiation mechanisms, namely, the methane–formaldehyde and the Koch carbonylation routes. In the first case, formaldehyde is not decomposed to CO but instead reacts directly with methane to form the first C–C bond.97 Methane and formaldehyde could be formed from methanol disproportionation over Brønsted acid sites (BAS), meaning their production rates correlate positively to methanol partial pressure.98 Depending on the coreactant (none, methanol/DME, olefin), H2, CH4, or paraffins are coproduced with the formaldehyde (Figure 4). Co-reaction of methanol with olefins to form paraffins and formaldehyde is in accordance with the hydrogen transfer reactions described previously in this section. A positive correlation has been found between the abundance of hydrogen transfer products and the concentration of Lewis acid sites in the zeolite.96 This is different from the rates of alkene and arene methylation reactions, which have been found to correlate positively with the concentration of Brønsted acid sites.
In addition to its key role in the initial C–C bond formation, formaldehyde is responsible for polyene and aromatics formation, which leads to coking and catalyst deactivation.99,100 Specifically, formaldehyde has been demonstrated through cofeeding and isotopic labeling (13C-formaldehyde) experiments to react with olefins via Prins condensation into dienes and aromatics.101,102 Hence, decomposition of formaldehyde to CO and H2 by a Y2O3 catalyst has been attempted as a strategy to improve catalyst lifetime.100,103 The influence of proximity between the CHA zeolite/zeotype (SAPO-34 or SSZ-13) and Y2O3 was studied at 400 °C, 1 bar, 0.12 bar methanol, CHA/Y2O3 = 3, and the most intimate mixing, i.e., intrapellet at 1 μm proximity, improved the catalyst lifetime by 4 times.103 In the direct conversion of COx to hydrocarbons, the partial pressure of CO would expectedly be much higher than the conventional MTH reaction, and the formaldehyde reactions may be affected.
Moving from the origin of CO to its significance, it is proposed that CO reacts with surface methoxy species to form the first C–C bond via Koch carbonylation. The Lercher group first identified methyl acetate and acetic acid as the first C–C bond containing intermediates from a ZSM-5 catalyst in the early stages of MTH.104 Using an online MS, methyl acetate and acetic acid were detected at the lower temperatures of 200 to 300 °C (1 bar) and were then converted to olefins at higher temperatures. Switching the carrier gas of methanol from He to CO, the olefin formation rate increased, and the existence of acetyl surface groups was confirmed with in situ IR spectroscopy. Such promotional effects of CO were not observed in previous literature, and the authors attributed this to a higher partial pressure of H2O, which inhibited the carbonylation reaction. The acetic acid or methyl acetate is proposed to then react with formaldehyde into unsaturated carboxylate or carboxylic acid, respectively, which decarboxylates into the first olefin (Figure 5).102 Independently, the Weckhuysen group employed advanced UV–vis and solid-state NMR spectroscopy techniques to identify methyl acetate as the first C–C-bond-containing intermediate to be formed in a SAPO-34 catalyst at 400 °C and 1 bar.105,106 Contrary to the mechanism above, their experimental insights supported the theoretical results and mechanism put forward by Plessow and Studt, in which ketene or methyl acetate reacts with a surface methoxy specie followed by decarbonylation to form the first olefin (Figure 5).107−109 In the latter mechanism, CO acts as a cocatalyst and is not consumed (Path (b)). In the former mechanism (Path (c)), CO is indirectly converted to CO2. A hint in favor of Path (b), in which multiple “first C–C bonds” may be formed from a single CO molecule, arises from prior studies in which the coke formed in the MTH initiation zone of a ZSM-23 catalyst contained close to 2 orders of magnitude more C than the CH4 formed in the same zone.110 While the exact mechanism to produce the first olefin still needs to be resolved, the latest work by the Bordiga group observed the presence of free vibrating CO throughout the MTO reaction catalyzed by MAPO-18s and suggested that CO may play a role beyond the first C–C bond initiation.111 Zeolite-catalyzed carbonylation mechanisms are described in more detail in Sections 4.1 and 4.2. For a more extensive review of hydrocarbon pool initiation mechanisms, we refer the reader to the recent review by Yarulina et al.42
Besides CO, water is another molecule of interest in the MTH process, which is relevant for the integrated process from COx to hydrocarbons. Water is well known to compete with methanol and DME for adsorption at the Brønsted acid sites,112 and for forming protonated water clusters at high water contents, thereby slowing down methanol conversion.113−115 Water coverage of the BAS may be reduced by working at high temperature, but may also facilitate hydration of the Si–O–Al bond, leading to extra-framework alumina formation and permanent deactivation of the catalyst.116−118 Hence, the temperature must be chosen with care. On the other hand, water has also been shown to be beneficial for catalytic stability, so it is used as a strategy to improve the lifetime of SAPO-34 catalysts at typical MTO conditions of 400 °C and 1 bar.79,112,119 A probable reason is the hydrolysis of formaldehyde, which is a known coke precursor, thereby reducing the coking rate.120 Another reason may be the reduction of relative reaction vs diffusion rates (Thiele modulus) resulting from unselective quenching of hydrocarbon pool species formation and conversion reactions by water adsorption on BAS, as recently reported for a SAPO-18 catalyst.121
In order to close the process gap between the MTH process and the direct process for COx hydrogenation to hydrocarbons, a “next-generation” MTH catalyst and process has to operate at elevated pressures and lower temperatures in a CO/CO2/H2/H2O environment. Thus, recent breakthroughs in the MTH process have focused on cofeeding these gases at elevated pressures. A main gas component, H2, has been shown to be favorable for zeolite performance at elevated pressures of 16 to 30 bar, by drastically reducing catalyst deactivation by coke formation.123−127 From Figure 6a, the lifetime of four zeolites (SSZ-13, SSZ-39, FER, BEA) improved when 16 bar of H2 was cofed with methanol.122 In particular, H2 was demonstrated to hydrogenate the coke precursors, formaldehyde, and 1,3-butadiene. This effect has a huge impact on process parameters, especially regeneration cycles for 8-ring cavity-window structured catalysts. Interesting to note, regeneration of deactivated ZSM-5 catalysts was attempted with H2 at 20 bar and 480 or 550 °C to yield light hydrocarbons and aromatics.128 Unfortunately, most studies pointed to hydrogenation even of olefins to form paraffinic products, as represented in Figure 6b.
Figure 6.
(a) Methanol conversion profiles versus cumulative turnovers and (b) cumulative selectivity (left ordinate) and the overall olefins-to-paraffins ratio (right ordinate) in the effluent stream observed during methanol feeds with and without H2 cofeeds over HSSZ-13, HSSZ-39, HFER, and HBEA. Reaction conditions: 350 °C, 0.005 bar methanol, 16 bar He or H2 cofeed, 350 [HSSZ-13]; 114 [HSSZ-39]; 61 [HFER]; 89 [HBEA] molC (molH+·ks)−1. The dark- and light-shaded bars in (b) represent the olefinic and paraffinic forms, respectively, of the respective carbon group listed in the dark bars; “MBs” represents methyl-substituted benzenes. Reproduced from ref (122).
Other gases present in the COx hydrogenation/tandem reactor, such as CO and H2O, may then be employed to counteract the hydrogenation of olefins during the MTH reaction due to high pressure H2 cofeeding.126,127 At 450 °C and 40 bar, the cofeeding of methanol, H2, and H2O increased the methanol handling capacity more than 200 times in comparison to typical MTH conditions.126 However, as mentioned earlier about the H2O influences, this lifetime improvement may be yet at the expense of catalyst degradation. Importantly, in another recent contribution, CO was found to have a very strong effect, suppressing olefin hydrogenation while maintaining the favorable turn-over number effect of cofed H2, especially for M(II)-substituted AlPO-18.127 Additional experiments where ethene was cofed with surplus H2/N2 or H2/CO over SAPO-18 confirmed that CO addition suppresses ethene hydrogenation.
The mechanism of product formation from carbonylated intermediates is not yet fully resolved. It could proceed by decarbonylation/decarboxylation (as proposed for the first C–C bond formation, Figure 5) or by dehydration. Dehydration seems plausible in the contribution from Liu and co-workers, where they cofed CO with methanol with CO/methanol ratios in the range of 20–113 at 400 °C, 40–50 bar total pressure, over H-ZSM-5 and observed 13C in methylcyclopentenone products after 13CO/12C-methanol cofeed experiments.129 Notably, very little 13C was found in propene and some more in propane. The thermodynamic drive of aromatic formation may be a reason for the dehydration instead of decarboxylation in longer Cn chains/cyclic compounds. A further interesting point of this contribution is that CO seems to partly suppress the hydrogen transfer route from alkenes to aromatics, leading to much less alkane formation. This result is in line with Xie et al.127 Recently, Shi et al. cofed H2 (0–12 bar), CO (0–8 bar), and/or H2O (0–4.5 bar) with methanol (1 or 10 kPa) over H-SAPO-34 at 400 °C.130 The results were in general agreement with the results described above. Furthermore, they found that the influence of CO on product selectivity vanished at the highest methanol pressure. This observation was ascribed to the higher concentration of organic compounds in the catalyst cavities for higher methanol concentrations, thereby outcompeting the reaction of CO with methoxy species. H2O addition was found to decrease the formation rate and steady state concentration of C2+ compounds in the catalyst, without influencing product selectivity, in line with Valecillos et al.121
2.2. Integration of the MTH Reactions with COx Hydrogenation
The industrial methanol synthesis catalyst (CZA) consists of Cu as the active element, and ZnO and Al2O3 as structural and electronic promoters.131 The operating conditions are 200 to 300 °C, 50 to 100 bar, with a gaseous feed of CO/CO2/H2. The feed composition is determined by the stoichiometry required for methanol synthesis, which is defined as M = (H2 – CO2)/(CO + CO2) ≈ 2. The discovery and historical development of the industrial Cu-based methanol synthesis catalysts is described in the review by Waugh.132 For recent progress on industrial methanol synthesis from CO/CO2/H2, the review by Sehested on catalyst development and the review by Bozzano and Manenti on reactor technologies are recommended.131,133 A current focus is the conversion of CO2 and H2 to methanol in which the Cu-based catalysts face the challenge of selectivity and stability due to oxidation and sintering. At 200 °C, 30 bar, H2/CO2 = 3, and GHSV = 9000 h–1, the space time yield of methanol over a commercial CZA catalyst decreased by 34.5% after 720 h time-on-stream and XPS results of the spent catalyst revealed the oxidation of metallic Cu to Cu2+.134 From XRD and N2O adsorption, Cu particle size did not change, but ZnO dispersion decreased. Hence, a cause for the decrease in methanol time yield was concluded to be oxidation of metallic Cu instead of sintering of Cu nanoparticles. On the other hand, several studies at harsher conditions, e.g., 250 °C and 50 bar, correlated the decrease in methanol space time yield with steam-induced Cu nanoparticle growth.135−137 Currently, oxide-based bulk catalysts are increasingly explored, plausibly inspired by the first-generation ZnCrOx catalysts developed by BASF in 1923.28,40
Methanol-mediated conversion of COx and H2 to hydrocarbons is covered in several recent contributions and review articles.3,18,20 Focusing on the zeolite function, most contributions apply the commercial MTH catalysts, H-ZSM-5 and H-SAPO-34, mixed with metal–support or mixed metal oxide COx hydrogenation catalysts. Reaction temperature is typically 350 °C and higher, in order to thermodynamically favor dealkylation of aromatic intermediates,138,139 as well as cracking of longer-chain alkenes to form lighter analogues.140 Benzene, toluene, and xylenes (BTX) and light alkenes, propene in particular, are the most valuable products. Under conventional MTH conditions (350 to 500 °C, 1 bar, 13 kPa methanol in inert flow), high alkene yields may be obtained over a wide range of zeolites, including H-ZSM-5 and H-SAPO-34, especially at 400 °C. Higher temperatures favor hydrogen transfer reactions and yield more aromatic and paraffinic products.141
The applications of zeolites in C1 conversion have been reviewed in a broad context that includes FTS and methanol-mediated processes,38,142 and that is different from this review, which focuses on the oxygenate-mediated processes. The effects of zeolite topology, crystal size, acidity, and proximity of the two catalytic functions are summarized previously, so only the key developments together with our viewpoints will be highlighted. A main issue related to the methanol-mediated COx hydrogenation process is the extent to which the overall product selectivity is influenced by altered reaction conditions and products formed on the methanol synthesis catalyst, compared to MTO. In qualitative terms, a recent study of the methanol-mediated CO hydrogenation process where ZnAlOx was combined with various SAPO zeotypes showed that the general trend of zeolite shape selectivity observed in prior MTO studies prevail:46,143 8-ring/12-ring window-cavity structures yield C2–C4 olefins and paraffins as main products, while 1D 10-ring structures yield C4–C6 olefins and paraffins as main products. 1D and 3D 12-ring structures yield C3–C5+ as main products, with a higher paraffin-to-olefin ratio than the 8- and 10-ring pore/window structures, due to substantial formation of aromatic products.
A quantitative comparison of the hydrocarbon selectivity attained with four 8-ring SAPOs under MTO or syngas conversion conditions is illustrated in Figure 7. The data sets for the SAPOs tested for MTO were selected based on the synthesis protocols used to prepare the SAPOs tested for the conversion of syngas.143−148 Due to the 8-ring structures, C2–C4 hydrocarbon selectivity was at least 67% in all cases. Furthermore, the distribution of ethene, propene, and butenes produced in syngas conversion fit well to the concept of the “cage-defining ring size” proposed by the Davis group for MTO.148 Specifically, SAPO-35, SAPO-17, and SAPO-34 produced more ethene and propene, while SAPO-18 produced more propene and butenes. In the conversion of syngas, more methane and paraffins were produced than those in the MTO reaction. Methane selectivity in MTO is typically lower than 2%, except during initiation and late deactivation stages.144,149 Both stages are characterized by low concentrations of olefins and aromatic products, which favor methanol–methoxy reactions to form methane and formaldehyde (vide ultra). Although methane has been assigned as a possible product from the methanol synthesis catalysts, it may not be the reason in this case since the same ZnAlOx catalyst was used with the different SAPOs. The methane selectivity also appeared to increase with increasing “cage-defining ring size” of the SAPOs during syngas conversion. Another possible reason for the higher selectivity toward paraffins could be the hydrogenation of the olefins due to the high-pressure H2 environment, as discussed previously in the cofeeding of high-pressure H2 in the MTO process.127 Hydrocarbon formation rates (on a C1 basis) and productivity numbers after 1 and 10 h on stream for MTO and tandem operation of the four catalyst structures are compared in Figure 7b. The initial rates are typically 1–2 orders of magnitude higher for MTO operation but rapidly decline. Instead, conversion rates under tandem operation are slow and rather stable over the test period. While the test duration was limited to 10 h in the examples of Figure 7, the superior time-on-stream (TOS) stability of the tandem process is confirmed by other literature studies, in which the tandem reaction was run for extended periods.3 A recent example is the CO2 conversion over a GaZrOx + SSZ-13 catalyst at 350 °C, 30 bar, H2/CO2 = 6, in which the C-based hydrocarbon production rate was initially 1.3 mmol/gcat·h, and still 0.9 mmol/gcat·h after 500 h on stream.150 The conversion capacity of this system (>550 mmol/gcat) is therefore higher than most of the MTO examples in Figure 7.
Figure 7.
(a) Hydrocarbon product distribution and (b) hydrocarbon productivity over TOS attained using 8-ring zeotypes during MTO or syngas conversion at 400 °C. Data sets for SAPO-35 (MTO) from ref (144), SAPO-17 (MTO) from ref (145), SAPO-34 (MTO) from ref (146), SAPO-18 (MTO) from ref (147), and all SAPOs (syngas) from ref (143).
As outlined by Tian et al.,36 the rapid deactivation of MTO catalysts imposes the use of a circulating fluidized-bed reactor and regenerator configuration similar to a fluid catalytic cracking (FCC) riser reactor concept for rapid and frequent removal of coke to regenerate catalysts. Attrition strength is therefore an important parameter of MTO catalyst design, due to erosion in the fluidized bed, as well as strain and chemical degradation of the zeolite/zeotype lattice imposed by frequent coke formation–regeneration cycles. Efforts targeting improved long-term catalyst cycle stability and carbon yields include regeneration in H2 and alkanes,151 precoking and water cofeed,119 and transformation of coke species to active carbonaceous intermediates via steam cracking.152 While those approaches may be useful even for catalyst regeneration after long cycle times, the fluidized-bed-specific catalyst attrition strength is less of an issue for tandem catalyst operation, for which cycle times are sufficiently long for fixed bed operation. On the other hand, the low product formation rates observed under tandem operation will impact reactor size and recycle ratios.
Another 8-ring zeolite, RUB-13, was recently investigated for the direct conversion of CO2.153 RUB-13 has a slightly wider window size than SAPO-34, and it shifted product selectivity toward C3 and C4 olefins rather than ethene. A total C2–C4 olefin selectivity in hydrocarbons of 65–83% at a CO2 conversion of 10–16%, of which > 90% was C3–C4, was observed over ZnZrOx/RUB-13. ZrCrOx/RUB-13 showed the highest ethene selectivity among the four tandem catalysts. This result was ascribed to the higher strong acid content and proportion of this catalyst, promoting the aromatic-based cycle to produce more ethene.
In addition to the partial pressures of the reactants and products, the extent of olefin hydrogenation has been reported to depend on the acidity of the zeolite/zeotype.154−156 From Figure 8a and 8b, the density of BAS in SSZ-13 was revealed to influence CO conversion and hydrocarbon product selectivity in the conversion of syngas catalyzed by ZnZrOx/SSZ-13, at 400 °C, 30 bar, H2/CO = 2.154 When the BAS density was lower than 0.1 mmol/g, an increase in BAS corresponded to an increase in CO conversion and a decrease in methanol and DME products. The increase in BAS led to a higher conversion and rapid removal of methanol and DME, hence, creating a thermodynamic driving force to increase CO conversion. The C2–C4 olefin/paraffin ratios remained stable between 4 and 5, possibly due to lack of vacant BAS for olefin adsorption. When the BAS density was higher than 0.1 mmol/g, complete conversion of methanol and DME was attained, and the CO conversion did not increase beyond 25%. The availability of BAS led to olefin hydrogenation, so the C2–C4 olefin/paraffin ratios decreased to 0.03 at BAS = 0.23 mmol/g. These effects of the BAS density were also observed for other catalytic systems, such as ZnAlOx/SAPO-34 tested at 400 °C, 30 bar, H2/CO = 2 (Figure 8c) and ZnCrOx/SAPO-18 tested at 400 °C, 40 bar, H2/CO = 2.5 (Figure 8d).155,156 In the study on the role of SAPO-18 acidity, the higher BAS density was also found to increase the C3/C2 ratio, and the authors attributed this to the secondary reactions of ethene, which decreased ethene selectivity.156
Figure 8.
Influence of zeotype acidity on the direct conversion of syngas to lower olefins. (a) CO conversion and (b) hydrocarbon selectivity as a function of BAS density in SSZ-13. Reaction conditions: ZnZrOx as methanol synthesis catalyst, 400 °C, 30 bar, H2/CO = 2. Reproduced from ref (154). (c) Hydrocarbon selectivity as a function of the BAS density in SAPO-34. Reaction conditions: ZnAlOx as methanol synthesis catalyst, 400 °C, 30 bar, H2/CO = 2. Reproduced from ref (155). (d) Hydrocarbon selectivity as a function of the BAS in SAPO-18. Reaction conditions: ZnCrOx as methanol synthesis catalyst, 400 °C, 40 bar, H2/CO = 2.5. Reproduced from ref (156).
The considerations for zeolite/zeotype acidity in the direct COx hydrogenation to hydrocarbons via methanol are unlike in the MTH processes. In MTO, the SAPO-34 zeotype is preferred over its isostructural SSZ-13 zeolite36,37 because the stronger acidic strength of the SSZ-13 often results in faster deactivation and more paraffins.157,158 This preference is not obvious in direct COx hydrogenation, as inferred from Figure 8b and 8c. It should be noted, though, that even MTO conditions may be tuned to favor SSZ-13 over SAPO-34 as MTO catalyst.139 Thus, targeted studies on the influences of zeotype/zeolite acidity, including acidic strength, BAS and LAS, in the oxygenates-mediated COx conversion would be needed to unravel deeper insights.
Considering the kinetics of the tandem reaction from CO2 and H2 to hydrocarbons, a recent study of a mixture of PdZn/ZrO2 and H-SAPO-34 catalysts showed that the reactions to form methanol and CO from CO2 rapidly reached equilibrium over the PdZn/ZrO2 catalyst, while methanol conversion to form hydrocarbons was the rate-limiting step of the reaction (Figure 9).159 In this particular case, the hydrogenating function of the PdZn alloy is so strong that only paraffinic products were formed, with exceptional selectivity toward propane. Overall, this study suggests that the hydrocarbon formation rate remains a challenge in the tandem reaction scheme. The leveling out of CO conversion versus BAS density in Figure 8a points in the same direction. In a follow-up study, Cordero-Lanzac et al. demonstrated an inverse first order correlation between the water content and the rate of methanol conversion to hydrocarbons over the PdZn/ZrO2+H-SAPO-34 catalyst.160
Figure 9.
Experimental data fitting of (a) CO2 to methanol over the PdZn/ZrO2 catalyst at 300 °C and 30 bar and (b) CO2 to propane over the PdZn/ZrO2+SAPO-34 system at 350 °C and 50 bar. (c) Comparison of reaction rates for methanol and CO formation over the PdZn/ZrO2 catalyst at 250 and 350 °C and 30 bar (12,000 mL/gcat/h), and (d) influence of propane formation on these rates over the PdZn/ZrO2+SAPO-34 system at 350 °C and 30 and 50 bar (6000 mL/gcat/h). Reproduced from ref (159).
Beyond effluent selectivity, a general feature of tandem catalysts is their much superior resistance to coking compared to MTO operation conditions (cf. Figure 7b). The process conditions and origin of the improvement in catalytic stability were investigated by Nieskens et al. through a direct comparison of the lifetime and coke species of SAPO-34 in MTO and syngas conversion.124 Notably, different hydrocarbon species were extracted from the spent MTO and syngas catalysts. From the spent SAPO-34 used in converting syngas at 400 °C, 15 bar, and H2/CO = 3, mainly smaller aromatic molecules, such as benzene, toluene, xylenes, and naphthalenes, were found. From the spent SAPO-34 used in converting methanol at 400 °C, 1 bar, and WHSV methanol = 1.3 h–1, the trapped coke species were predominantly anthracenes and pyrenes. The differences were attributed to the presence of H2 in the feed, in particular, the H2/methanol partial pressures. Furthermore, the reaction temperature had to be higher than 300 °C for high-pressure H2 to suppress hard coke formation. These results are in agreement with the high-pressure cofeed studies in MTO (Section 2.1), whereby the H2 content of the feed and effluent gas were observed to hinder the hydrogen transfer reactions that precede aromatics and polyaromatics formation in zeolites/zeotypes.
From the mechanistic perspective, the general understanding is that methanol conversion to hydrocarbons in the direct COx hydrogenation processes occurs via the conventional “dual cycle” reaction mechanism for MTH (Figure 4). However, Hou and co-workers recently proposed a “triple cycle” instead.161 A quasi-in situ, solid-state NMR and online GC/GC-MS analysis methodology was implemented to monitor the direct syngas conversion during start-up and steady state operation. ZnAlOx was mixed with either ZSM-5 (Si/Al = 82) or ZSM-5 (Si/Al = 16) and tested at 300 °C, 25 bar, and H2/CO = 1. Excluding the 45% CO2 selectivity, the tandem catalyst with more BAS showed 80% selectivity toward aromatics, while the other tandem catalyst was more selective toward paraffins (∼60% C2–C4, ∼20% C5+). These hydrocarbon product distributions were different from the MTH control experiments as the aromatics selectivity of both catalysts were only ∼35%. According to the solid state NMR analysis, formate, methanol, and DME were observed after 30 s of reaction, while acetates and carboxylates were detected after 2 min of reaction. The acetates and carboxylates were proposed to be formed from olefins/alkoxyls and CO via Koch carbonylation, and they were presented as evidence for the oxygenate-based cycle.
The new oxygenate-based cycle bears a resemblance to the reaction pathway for the first olefin formation in MTH (Figure 5), with the same intermediate (acetate) being observed in both cases. The three main differences between the MTO and syngas case is first, that the CO/methanol ratio is dramatically different between the two cases. In the syngas case, it is typically > 102 (based on effluent analysis), while in the MTO case, it is likely < 10–2 (when estimated from the CH4 yield). This difference will dramatically shift the relative rates of surface methoxy reactions with either CO to form acetyl/ketene/acetate, or olefins/arenes to form higher olefins/methylated aromatic products. According to Figure 5 Path (b), CO acts as a cocatalyst of methanol conversion in the zeolite/zeotype, just like olefins and monocyclic aromatics–the difference between the cycles being that CO is not an autocatalytic species. Prior studies by the Liu group (vide ultra, Section 2.1) showed that 13C-CO cofed with 12C-methanol over H-ZSM-5 yielded 13C-containing aromatic products, while at the same time suppressing paraffin formation, which is the trademark of aromatics formation by hydrogen transfer reactions in MTH.129 Those results are in line with the “triple-cycle” reaction concept and point to detailed studies of the relative rates of carbonylation versus methylation reactions as a highly interesting topic of further study. The second main difference between the MTO and syngas case is the H2/methanol ratio, which is even more dramatic than the CO/methanol ratio difference. Prior cofeed studies (vide ultra, Section 2.1) showed how hydrogen cofed with methanol suppresses coke formation by hydrogenation of coke precursors, and how H2/CO cofed with methanol forms a well-balanced pool of retained species that produces olefins rather than paraffins, yet with limited deactivation due to coke formation. Again, follow-up studies to unravel individual and combined effects would be highly valuable for future catalyst optimization by design. The third main difference between the syngas and MTO cases is the (de-)hydrogenation function of the methanol-producing catalyst. As pointed out by Hou and co-workers, mixing the two catalyst functions enables acetate/acetic acid (and higher analogues) to diffuse out of the zeolite/zeotype and be hydrogenated to ethene on the methanol-producing catalyst, hence accelerating the final steps of the MTO initiation reactions to first C–C bond formation.161 Together, the recent mechanistic insights in tandem reactions create a novel playground for fundamental as well as more application-oriented studies toward a next-generation tandem process.
3. Dimethyl-Ether-Mediated COx Conversion to Hydrocarbons
With one exception (vide infra, Section 3.3), all contributions cited in this review article report or presume that dimethyl ether is formed via methanol dehydration. As will be seen in Sections 3.2–3.3, the dehydration site may be present in the CO2 hydrogenation catalyst, in the form of a support material or a binder added during catalyst formulation, or in the hydrocarbon-forming zeolite/zeotype. First, however, in Section 3.1, we focus on the conversion of DME to hydrocarbons.
3.1. Dimethyl Ether to Hydrocarbons (DTH)
In the first publication in 1977 on MTH using ZSM-5 at 371 °C, Chang and Silvestri from Mobil proposed that the conversion of methanol to gasoline proceeded via partial methanol dehydration to DME, which resulted in an equilibrated feed mixture.81 The first MTG plant was commercialized by Mobil in New Zealand in 1985, and in the process flowsheets, methanol was dehydrated to DME in the first reactor and the effluent was then converted to gasoline in the second reactor.162 This dual reactor flowsheet removed heat produced during methanol dehydration and allowed better heat control during gasoline production from the methanol/DME/H2O equilibrated feed mixture. Among the subsequent MTH industrial developments, the utilization of DME as an intermediate was included in Topsøe’s TIGAS process, Lurgi’s MTP process, and Dalian Institute of Chemical Physics (DICP)’s DMTO process.35,36,163−165
DME and methanol were often treated as a single reactant feed, because they were considered to be in equilibrium under relevant MTH reaction conditions.65,81,166 This assumption appears to be valid because DTH and MTH were shown to follow the same dual hydrocarbon pool mechanism and to yield similar product distribution.81,167−170 Nonetheless, the mechanistic studies also concluded differences in the reaction rates, and these were highly dependent on zeolite/zeotype properties, such as topology, crystal size, and acidity.
In an early study of SAPO-34 at 425 °C, Holmen et al. used a tapered element oscillating microbalance (TEOM) to determine methanol and DME mass transfer limitations as a function of crystal sizes between 0.25 and 2.5 μm.149 DME diffusion was found to be more hindered than methanol, therefore resulting in slower overall reaction rates.149 In a similar study of SAPO-34 at 350 °C, 0.075 and 0.8 μm crystals were compared for DTH and MTH, and no diffusion limitations were observed.169 There were no significant differences between DTH and MTH, and the smaller crystals resulted in a longer lifetime. In a more recent study of SAPO-34 at 350 to 450 °C, Wang et al. noted a longer initial induction period with corresponding higher propylene selectivity for DTH than MTH.170 DME hydration was found to be unfavorable below 400 °C, which resulted in lower methanol partial pressure in the reactor and slower reaction rates for DTH. Despite the lower reactivity of DTH, it had the advantages of slower catalyst deactivation and higher propylene selectivity during the initial period, and this could potentially be exploited.
Moving on to ZSM-5, there is agreement that methylation rates in DTH are significantly faster than in MTH.90,168,170,171 Svelle et al. performed propylene and toluene methylation experiments using methanol and DME with ZSM-5 at 350 °C, and concluded that 1-butene formation rates were at least two times faster for DME than methanol over 0.02 to 0.08 bar (DME or methanol/propylene = 1 to 4).168 Castano et al. studied DTH and MTH using ZSM-5 at 400 °C and 1.5 bar, and also found faster kinetics of methylation for DTH in comparison to MTH.171 They reasoned that the faster methylation rate and lower water concentration present in DTH led to faster conversion of hydrocarbon pool intermediates. However, this also resulted in the formation of more methylated and condensed aromatics, i.e., coke, hence faster deactivation in DTH. In the direction of arene methylation in ZSM-5, Hibbitts et al. contributed a theoretical investigation on the implications of methanol and DME as methylation agents for all 20 arenes (C6 to C12) at 350 °C and 1 bar.172 They focused on possible confinement effects on the transition state structures and found that reorientations of the transition states could decrease energies up to 45 kJ/mol. This decrease in energy is more significant than the difference between the concerted or sequential methylation mechanism of 20 kJ/mol, implying that both mechanisms could take place. In addition, the energy barriers of methylation by methanol and DME are nearly identical, suggesting that both are equally capable of arene methylation.
To unravel mechanistic differences and origins of DTH and MTH deactivation, Olsbye and co-workers designed a series of experiments which examined DME and methanol influences on individual reaction pathways in the arene cycle. First, the reactivity of DME and methanol with benzene was investigated at 250 to 350 °C, 1 bar, DME or methanol = benzene = 0.06 bar, and < 9% conversion (differential conditions).90 Higher methylation and dealkylation rates were observed with DME than with methanol, regardless of topology (MFI, ZSM-5 and AFI, and SSZ-24) and acidity (SSZ-24 and SAPO-5). Importantly, benzene methylation using DME resulted in the formation of toluene, polymethylbenzenes, and alkenes, while the use of methanol resulted in the formation of diphenylmethanes (DPMs) as well. Additional cofeed and isotopic labeling experiments revealed that methanol was first dehydrogenated to formaldehyde, which then reacted with two benzene molecules to form DPM and ultimately coke. Next, the role of formaldehyde was further investigated by cofeeding formaldehyde with methanol at 350 °C, 1 bar, 0.058 bar methanol, and 0.001 bar formaldehyde.99 The cofeeding of formaldehyde increased aromatics and ethylene formation, thus supporting the conclusions of the prior study. Based on the argument that methanol is a main source of formaldehyde, formaldehyde was demonstrated to be a critical factor for the faster deactivation during MTH than DTH. Finally, the reactivity of DME and methanol with isobutene was investigated at 350 °C, 1 bar, and DME or methanol = isobutene = 0.04 bar.89 Isobutene was used as a probe molecule to distinguish between methylation and H-transfer pathways, and H-transfer is critical for not only the arene cycle but also coke formation. In the case of MTH, methylation and H-transfer rates were equally competitive, whereas in the case of DTH, methylation rates were faster than H-transfer rates so the arene cycle and arene products were less dominating (Figure 10a). From these thorough cofeed studies on methanol vs DME, the origin of faster MTH deactivation was the formation of formaldehyde, which led to DPM, and the competitive rates of methylation and H-transfer. Thus, using DME rather than methanol as feed has clear benefits, and these studies provide mechanistic grounds on the use of a first reactor to dehydrate methanol to DME before a second reactor to produce hydrocarbons in the process flowsheet.
Figure 10.
(a) GC chromatogram comparison of coreaction of 40 mbar isobutene with 40 mbar methanol (blue) and 40 mbar DME (red) over H-ZSM-5 at 350 °C. Reproduced from ref (89). (b) Deactivation profiles based on oxygenate conversion versus time-on-stream at 450 °C over SSZ-24 and AlPO-5/SSZ-24. Reproduced with permission from ref (99). Copyright 2017 Royal Society of Chemistry.
Martinez-Espin et al. further contested the hypothesis of methanol and DME being in equilibrium in the second reactor which produces hydrocarbons in the process.99 Notably, at relevant MTH conditions of 350 °C and 1 bar regardless of methanol, DME, or DME with water feeds, equilibrium was not reached with ZSM-5. This finding is also valid for other topologies with strong BAS, such as SSZ-24 (AFI), and it is due to the similar rates for methanol dehydration to DME and for methylation reactions over the strong BAS. The methanol and DME equilibrium could be established when weak acidic sites such as P–OH were present, as in the case of AlPO-5 addition to the catalyst bed (Figure 10b).
Besides the above studies which compared methanol and DME, it is noteworthy to then highlight relevant insights on DTH. Pérez-Uriarte et al. investigated the effects of zeolite topology using ZSM-5, SAPO-34, and SAPO-18 for DME conversion to lower olefins at 350 and 400 °C, and 1.5 bar pure DME.173 ZSM-5 showed the highest activity and stability, but surprisingly it showed the highest selectivity toward lower olefins (27% ethylene, 50% propylene, and 16% butenes) at low space time (∼10% conversion). As a follow-up, operating conditions were screened to optimize lower olefins yield using the ZSM-5 (Si/Al = 280) catalyst.174 Interestingly, DME was shown to undergo thermal cracking to give CO and methane when the reaction temperature was higher than 400 °C (Figure 11a). As a result, the selectivity and yield of lower olefins decreased at higher temperatures, as presented in Figure 11b. In Figure 11b, the red region of the heat map represented the highest yields of lower olefins, which was between 40 and 50%. The lower steam partial pressure in DTH was suggested to be the cause for higher reactivity but also faster deactivation, presumably due to the high DME partial pressure. Ortega and Kolb also studied the influence of process variables, specifically a temperature range between 325 and 375 °C and space time range between 0.008 and 0.040 h·kgcat/kgDME using a ZSM-5 (Si/Al = 58) catalyst.175 Thorough data and kinetic analysis revealed that product selectivity was dependent on DME conversion and independent of reaction temperatures and space times. They further concluded that equilibrium was not reached between DME and methanol in this study, thereby supporting the earlier finding of Martinez-Espin et al.99
Figure 11.
(a) Thermal cracking behavior of DME and (b) map of lower olefin yield at 1 atm of pure DME for a range of space times and temperatures. Reproduced from ref (174).
The Bhan group conceived several studies on olefins (i.e., ethylene, propylene, 1-butene, trans-butene, cis-butene, and isobutene) and aromatics (i.e., benzene, toluene, para-xylene, and ortho-xylene) methylation reactions using DME. Olefin methylation reactions using DME were performed with various zeolites (FER, MFI, MOR, and BEA) at low temperatures (<127 °C), high DME/olefin ratios (>15), and low conversions (<0.2%) to isolate primary reactions.176,177 Under such conditions, olefin methylation rates were determined to be independent of DME partial pressure and first-order in olefin partial pressure. Higher olefins and higher stabilities of the intermediate carbenium ions resulted in faster methylation rates. Aromatics methylation reactions using DME were carried out with ZSM-5 at similar conditions, and zero-order dependence on DME partial pressure was concluded.178 They further studied DTH over ZSM-5 at more relevant conditions of 275 °C, 1 bar, 0.7 bar DME, and ∼20% conversion.179 Co-feeding of 0.04 bar 13C-propylene with 0.7 bar 12C-DME suggested that C5–C7 olefins were formed mainly from methylation reactions, and these higher olefins subsequently cracked to form propylene. Co-feeding of 0.04 bar of 13C-toluene with 0.7 bar of 12C-DME showed that the higher aromatics originated from C8+ aliphatic cyclization reactions instead of toluene methylation reactions.
In conclusion, DTH has several advantages over MTH: The lower oxygen content of DME leads to a smaller exotherm in a DTH reactor compared to that in an MTH reactor. Furthermore, DME has a higher conversion rate for methylation of alkenes and arenes compared with methanol and is more selective for alkene/arene methylation relative to formaldehyde formation. This selectivity leads to higher turn-over numbers before deactivation by coking for DTH versus MTH. The reaction rate advantage of DTH versus MTH is more pronounced for medium-to-large (10-/12-ring) pore zeolites/zeotypes, since in small-window (8-ring) zeolites/zeotypes (CHA, AEI, ...) DME conversion is slower than methanol conversion due to slower diffusion of DME into the micropores.
3.2. Methanol to Dimethyl Ether (MTD)
Solid acids catalyze the dehydration of methanol to dimethyl ether (MTD), and the two most common catalysts are γ-Al2O3 and ZSM-5.29,180 The advantages of using γ-Al2O3 include low cost, high DME selectivity, and high catalytic and mechanical stability. The weak–moderate acidic sites of γ-Al2O3 catalyze methanol dehydration but are incapable of catalyzing the MTH reaction. Weakly adsorbed surface methoxy species form DME, while the strongly adsorbed surface methoxy species convert to surface formates, which subsequently decompose to CO, H2, and CH4.181 On the other hand, the strong acidic sites of ZSM-5 catalyze both reactions at higher reaction temperatures. ZSM-5 may be preferred over γ-Al2O3 for its superior hydrothermal stability and higher activity at low temperatures, but it has a lower reaction temperature window of 160–250 °C.182
As a connection between γ-Al2O3 and zeolites, Si-modification of Al2O3 was found to increase surface area and acidity, which led to improvement in catalyst lifetime at 300 °C, 1 bar, methanol/N2 = 0.11, and GHSV = 15,600 h–1.183 As another connection between γ-Al2O3 and zeolites, Olsbye and co-workers showed that porous aluminophosphates (i.e., AlPO-5 with AFI topology) could be used for MTD at higher temperatures of 450 °C because they contain weak acidic sites for MTD but not strong acidic sites for MTH.99
Improvements to the catalytic performance of zeolites/zeotypes focus on suppressing hydrocarbon selectivity and coking by eliminating the strong acidic sites, identifying suitable frameworks, and improving catalytic stability by coke studies. The first attempt to remove strong acidic sites in ZSM-5 using Na cations led to 100% DME selectivity over the temperature range of 230 to 340 °C.184 Various cations and zeolites with different Si/Al ratios were later screened for the liquid phase MTD reaction at 250 °C and 30 bar in a slurry reactor.185 Among the zeolites (ZSM-5, Y, MOR, FER, and BEA) with a similar Si/Al ≈ 20, MOR appeared to be the most active and selective. Subsequent ion-exchange of MOR with cations such as Zn, Ni, Al, Zr, Mg, and Na improved selectivity and stability through the removal of strong acidic sites. The catalytic performance of zeolites ZSM-5, MOR, BEA, and FER were also evaluated in typical fixed bed reactors from 180 to 300 °C, and the superior stability and low coke deposition rate of FER was attributed to its two-dimensional structure with smaller pores.186,187 The coke formation mechanism of ZSM-5 during MTD was studied at 200 °C using operando UV–Raman spectroscopy by Li et al.188 Methylbenzenium carbenium ions (MB+), a known coke precursor, was found to transform rapidly into “hard coke” at the top of the catalyst bed, but this transformation was suppressed toward the end of the catalyst bed due to water produced in the reaction.
In addition to catalyst development, mechanistic insight was needed to understand the origin of the influence of reaction parameters in MTD. There are two possible mechanisms, either concerted (associative) or sequential (dissociative) pathways, as shown in Figure 12a.189−192 In the direct associative mechanism, two methanol molecules coadsorb at the BAS to form a protonated methanol dimer. The protonated methanol dimer then rotates to form the transition state, which eliminates DME and water in a kinetically relevant step. In the sequential dissociative mechanism, a methanol molecule first adsorbs on a BAS and dehydrates to form a surface methoxy intermediate in a kinetically relevant step. The surface methoxy intermediate then reacts with another methanol molecule to form DME.
Figure 12.
(a) Reaction coordinate diagram showing DFT-calculated free energies (ΔG; kJ mol–1) of intermediates and transition states involved in methanol dehydration on isolated H+ sites in CHA at 415 K and 1 bar of methanol. Reproduced with permission from ref (193). Copyright 2019 Elsevier. (b) Associative and dissociative rate constants as a function of temperature over ZSM-5. Reproduced with permission from ref (191). Copyright 2014 John Wiley & Sons.
The preference of one mechanism over the other depends on reaction parameters, e.g., temperature, methanol partial pressure, and water partial pressure. Jones and Iglesia used a combination of kinetic, spectroscopic, and theoretical findings on ZSM-5 to support the associative mechanism being dominant at relevant temperatures and pressures for MTD, i.e., below 230 °C and 0.1 bar methanol or below 300 °C and 1 bar methanol (Figure 12b).191 At lower temperatures and higher pressures, the associative mechanism dominates due to the lower enthalpy transition state. At higher temperatures relevant for the title reaction, it is suggested that methanol–DME interconversions become equilibrated, and the dissociative mechanism dominates. These correlations were supported by a more elaborated van der Waals (vdW)-corrected DFT study on ZSM-5 with heterogeneous Al distribution.192 However, Grabow et al. showed that the absolute reaction temperatures and pressures in which the dominating mechanism crossovers from associative to dissociative depend on the location of the BAS, i.e., Al sitting, in ZSM-5. On the other hand, the DFT study on ZSM-22 (TON topology) by Moses and Nørskov concluded that the dissociative mechanism is dominating regardless of reaction temperature.194 Although water lowers the activation energies of key reactions in the associative mechanism, the dissociative mechanism remains more favorable.
The preference of one mechanism over the other also depends on the pore and cavity sizes of the zeolite, since confinement of the transition states would result in stabilization and a lower activation barrier. In this context, BAS descriptors such as acid strength, location, and proximity are also considered. It has been shown that the stronger the acid strength in ZSM-5, the more stable the transition states regardless of the reaction pathways.192,194 Jones and Iglesia surmised that the associative mechanism dominates in catalysts with larger voids, which could fit the larger dimeric transition state of the associative pathway, and such solvation by confinement would be beneficial. However, if the pore and cavity sizes were too small to accommodate an associative transition state, then the dissociative mechanism dominates. They subsequently verified their hypothesis over FAU, SFH, BEA, MOR, MTW, MFI, and MTT zeolites, and illustrated the negative relation between largest free sphere diameter and methanol dehydration rates.195 The confinement effect was exemplified as well in the MTD mechanistic studies using small-pore zeolites, i.e., CHA, AEI, LTA, and LEV, by the group of Gounder.193,196 Interestingly, the confinement effect was not always favorable in these small-pore zeolites, because the turnover rates of MTD became inhibited at high methanol pressures, i.e., > 0.1 bar at 142 °C.193 The origin of such a kinetic inhibition effect was proposed to be the stabilization of methanol clusters greater than three methanol molecules. This was not observed for the medium- or large-pore zeolites.
3.3. Integration of the MTD and DTH Reactions with COx Hydrogenation
DME is proposed to be a more efficient intermediate than methanol for the conversion of COx to hydrocarbons. DME has a lower H/C ratio than methanol, meaning that a broader ratio of syngas feed could be applied with a better C utilization. For instance, a net consumption of CO2 is achievable when the direct DME process starts with dry methane reforming.197 The direct DME process has more favorable thermodynamics, resulting in overall higher energy efficiency and higher conversion per pass.198−200 In addition, DME conversion limits formaldehyde formation and results in higher methylation rates in comparison to methanol, substantially prolonging catalyst lifetime.99 In view of the clear benefits on using DME as an intermediate, a critical assessment of the current status of such processes is needed before considering new strategies in this direction.
Generally there are three types of bifunctional catalysts that are able to convert COx to DME, by integrating methanol synthesis and methanol dehydration catalytic functions. The most obvious is the mixing of the two catalysts, the second type is the synthesis of core–shell catalysts, and the third is the use of oxide catalysts with acidic properties. The first two categories of catalysts have been reviewed, but the latter has scarcely been discussed.29,180,201 We will discuss relevant studies that focus on the MTD (zeolite) catalytic function in the tandem processes for COx conversion, before moving into oxide catalyst developments for CO2 hydrogenation to methanol and DME. We will also highlight improvements from process design, including reactor design and addition of adsorbents. Finally we present our case on the opportunities and risks on using DME as an intermediate, and our proposed solutions to mitigate the risks.
For DME production from syngas, the industrial catalyst consisting of Cu, Zn, and Al (CZA) was almost universally used for the methanol synthesis catalytic function.29,180 ZSM-5 was typically used as reference, and studies include the influences of zeolite framework, ZSM-5 modification and particle size, proximity, and kinetics. The proximity effect was studied with different degree of mixing (slurry, grinding, pelletized) at 260 °C, 40 bar, and 1700 mL/gcat/h, and the catalysts prepared by slurry and grinding methods performed worse due to ion exchange of Cu2+ and Zn2+ to BAS.202 The physical mix of CZA and the ZSM-5 catalyst showed 89% CO conversion with 64% DME selectivity. Acidity of ZSM-5 was modified by MgO addition and evaluated at 260 °C, 40 bar, and 1500 mL/gcat/h (H2/CO/CO2 = 66/30/4). Dispersed MgO (∼5 wt %) introduced basic sites and LAS to ZSM-5 while decreasing BAS, leading to a decrease in CO2 and hydrocarbon selectivities to 31% and 0.4%, respectively (with 64% DME) at 95% CO conversion.203 Referring to Figure 13, crystallite size of ZSM-5 was varied between 65 and 800 nm, and activity (at 260 °C, 20 bar, and 3600 mL/gcat/h) was found to have an inverse relation with particle size due to mass transfer.204 The influence of the crystallite size was suggested to be stronger than the influence of BAS density. Catalyst deactivation was attributed to sintering of Cu nanoparticles on the BAS located at the external surface of ZSM-5. Reaction parameters and kinetics were also considered from 210 to 270 °C and from 10 to 50 bar for the CZA/ZSM-5 catalysts.205 Higher temperature and pressure were beneficial, while water cofeeding decreased methanol synthesis rates. The influence of zeolite topology was studied with ZSM-5, FER, IM-5, TNU-9, MCM-22, and ITQ-2 (Si/Al = 9–14) at 260 °C, 40 bar, and 1700 mL/gcat/h.206 Catalyst deactivation due to Cu nanoparticle sintering was found to be more detrimental than framework-induced coking and was further showed to correlate with BAS located at the external surface of zeolites. Thus, the close intimacy of the two catalytic functions was suggested to be unfavorable.
Figure 13.
(a–f) Representative SEM images of ZSM-5 with varied crystallite sizes and (g) the effect of ZSM-5 crystal sizes on activity and stability for DME formation. Reaction conditions: 260 °C, 20 bar, H2/CO = 2, and 3600 mL/gcat/h. Reproduced with permission from ref (204). Copyright 2016 Elsevier.
DME production from CO2 hydrogenation is more challenging due to unfavorable thermodynamics, which limits CO2 conversion to less than 30% at 200 to 280 °C and 30 bar. Kinetic studies showed that the formate formation step was rate-determining for methanol synthesis, resulting in a much slower methanol formation rate than methanol conversion rate.207,208 CZA was commonly used for methanol synthesis but other metallic catalysts such as Pd were explored recently due to the strong sintering of Cu nanoparticles in the presence of high steam partial pressure.201,209 The high steam partial pressure was detrimental for γ-Al2O3 as discussed earlier, resulting in superior performance of zeolites over γ-Al2O3 at CO2 hydrogenation conditions of 220 to 280 °C, 30 to 40 bar, and H2/CO2 = 3 to 4.210−212 Bonura et al. studied the influences of the mixing of CZA and ZSM-5, and of ZSM-5 acidity by varying Si/Al ratio from 15 to 100 at 240 °C, 30 bar, 10,000 mL/gcat/h, and H2/CO2 = 3.213 The zeolites with varied Si/Al ratios were tested only for MTD reaction, and Si/Al = 25 was selected based on activity and stability during water cofeeding. At 240 °C, 30 bar, 2500 mL/gcat/h, and H2/CO2 = 3, the physical mixture of CZA and ZSM-5 showed 38% DME, 10% methanol, and 52% CO selectivity at 15% CO2 conversion. The catalyst with CZA precipitated on ZSM-5 showed a slight 2% improvement on activity and selectivity. Bonura et al. extended their earlier study with FER zeolite of varied acidity and grain size.214 The catalyst CZA with nanosized FER (200–500 nm) was least active and selective toward DME in comparison to those with FER crystal size of > 1000 nm. The increase in BAS density appeared to cause larger Cu sintering and a faster deactivation rate, which was brought about by higher steam partial pressure.
To attain precise control of the proximity and intimacy of the two catalytic functions for STD, core–shell catalysts were pioneered by the group of Tsubaki in 2010.215 It was challenging to synthesize the zeolite shell over the CZA core because NaOH (commonly used in zeolite synthesis) could not be used to prevent damage to the CZA catalyst and also because the mechanical and hydrothermal stabilities of the CZA catalysts were poor. The core–shell catalyst consisting of a CZA core of 850 to 1700 μm and a zeolite shell of 4 to 5 μm attained the highest DME selectivity of 97% at 6% CO conversion (250 °C, 50 bar, and H2/CO/CO2 = 6/3/0.5). In comparison, the catalyst prepared by mixing attained 41% DME selectivity with 57% CO conversion. The same group extended the core–shell strategy to other zeotypes including SAPO-11 and SAPO-46, which simplified the synthesis protocol toward physical coating in view of potential scale-up and industrial relevance.216,217 Ateka et al. compared the core–shell (100–300 μm) and physically mixed catalysts consisting of CuO-ZnO-ZrO2 and SAPO-11 for syngas and CO2 conversion at 275 °C, 30 bar, and CO2/CO = 1.218 The core–shell and physically mixed catalysts attained 81% DME selectivity at 11% COx conversion and 77% DME selectivity at 9% COx conversion, respectively. Interestingly, the core–shell catalyst showed higher CO2 conversion than the physically mixed catalyst over 24 h TOS (4% vs 2%), and this was attributed to easier H2O diffusion from the methanol synthesis catalytic sites, which shifted the equilibrium. Modeling catalytic performance as a function of ZSM-5 wt % in the core–shell catalyst, Klumpp et al. highlighted the limitations of the core–shell catalysts in mass transfer and varying mass ratios of core and shell catalysts.219 Most literature agrees that the core–shell catalysts hold the advantage in DME selectivity but not activity due to mass transfer limitations.
In addition to the bifunctional catalysts, which consist of 2 distinct catalysts for methanol synthesis and conversion to DME, oxides with acidic sites are an emerging class of catalysts that are capable of producing DME from COx. The industrial methanol synthesis catalyst CZA contains a small fraction of Al2O3, which could in principle catalyze methanol dehydration, but it is typically added as a stabilizer and binder for Cu and Zn, so its catalytic activity has not been considered previously. The Copéret group prepared 3 nm Cu nanoparticles (4 wt % Cu) supported on Al2O3 or SiO2 using the surface organometallic chemistry approach and evaluated their catalytic performance for CO2 conversion to methanol at 230 °C, 25 bar, H2/CO2 = 3, and < 7% CO2 conversion.220 The Cu/Al2O3 catalysts with LAS showed 30%, 15%, and 55% selectivity toward DME, methanol, and CO, respectively. On the other hand, Cu/SiO2 possessed no LAS, so only methanol and CO were formed. The ability of LAS on oxidic supports to catalyze methanol dehydration was also discussed by Prieto et al. in their investigation on LAS on oxidic supports as kinetic descriptors for CO2 conversion to methanol using Cu nanoparticles.221 Negligible DME was produced in their main study due to the low reaction temperature and conversion, which inhibited secondary reactions; however, the capability of LAS of the Ta, Al, Zr, Sc, and Y oxides to dehydrate methanol was verified in control experiments. A recent development from the Copéret group on SiO2-supported PdGa nanoparticles showing DME selectivity further demonstrates the potential of using a single catalyst with two active sites in closest intimacy to convert COx to DME.222 These catalysts could facilitate the rapid conversion of methanol to DME, after which DME would be the major intermediate to hydrocarbons.
Moving from supported catalysts, several studies on mixed oxides mixed with zeolites for conversion of COx to hydrocarbons identified both methanol and DME as reaction intermediates.154,155,223,224 Notably, the Ye Wang group developed several state-of-the-art catalytic systems such as ZnZrOx and ZnAlOx coupled with various zeolites, and the observation of both methanol and DME in the product spectrum provided indirect evidence that the process proceeded via methanol and DME. They further demonstrated the capability of binary oxides with spinel structure to convert both CO and CO2 to an oxygenate mixture of methanol and DME at 400 °C, 30 bar, and H2/CO = 2 or H2/CO2 = 3.155 These spinel oxides were prepared by coprecipitation, and the best-performing catalyst ZnAlOx showed the highest DME selectivity of 50% at 3% CO conversion, with the remaining product spectrum making up of 9% methanol, 6% hydrocarbons, and 35% CO2. Wu et al. reported another binary oxide GaZrOx prepared by evaporation-induced self-assembly (EISA) to be selective toward DME and methanol from CO2 conversion.225 As presented in Figure 14a, ZrO2 showed 10% methanol selectivity but no DME (at < 1% CO2 conversion), while Ga2O3 showed ∼25% methanol and ∼10% DME selectivity (at ∼1.5% CO2 conversion) at 330 °C, 30 bar, 24,000 mL/gcat/h, and H2/CO2 = 3. GaZrOx with 19 to 27 wt % Ga loading showed around 50% methanol and 25% DME selectivity at 6 to 9% CO2 conversion. Detailed characterization performed to understand the synergistic effects of the bimetallic components resulted in the proposed reaction mechanism shown in Figure 14b. Importantly, DME was suggested to be formed directly from hydrolysis and hydrogenation of surface methoxy species, instead of sequentially via methanol dehydration.
Figure 14.
(a) CO2 conversion and selectivity toward methanol and DME over GaZrOx catalysts and (b) the proposed mechanism. Reaction conditions: 330 °C, 30 bar, H2/CO2 = 3, and 24,000 mL/gcat/h. Reproduced from ref (225).
Considering the latest developments in COx conversion and CO2 utilization, there are several research questions and opportunities in the direction of DME-mediated conversion to hydrocarbons. First, new classes of catalysts including supported catalysts and bulk oxides are capable of converting COx to DME on a single catalyst instead of the conventional bifunctional catalysts consisting of copper-based methanol synthesis catalysts and zeolites for methanol dehydration. For the supported catalysts, LAS seems to be a key descriptor for DME production and could be introduced by support doping.222 For the bulk oxides, key descriptors for DME formation are less obvious because there are more variables such as synthesis method, elements and their loading/mixing, structure, and size. For instance, GaZrOx prepared by coprecipitation and tested at 300 °C and 20 bar was not reported to produce DME, while the previous case study showed GaZrOx to produce ∼0.5 mol of DME per mol of methanol.225,226 Thus, catalyst development and identification of key descriptors for the conversion of COx to DME would be useful for further improvements to the DME-mediated tandem process. Second, mechanistic studies thus far focused on active sites for CO2 and H2 activation, and much less is known about the active sites for methanol or DME formation. In the proposed reaction mechanisms, both methanol and DME originated from surface methoxy species, but it is rather ambiguous on when DME would be formed rather than methanol. One mechanism hypothesized that methanol was formed from the surface methoxy species and should dehydrate to form DME.155 The other mechanism inferred that DME could be formed directly from hydrolysis and hydrogenation on surface methoxy species.225 Distinguishing such mechanistic details would be critical for designing active sites for DME formation.
High steam partial pressure remains to be a challenge in the tandem process, and this may be tackled from either catalyst design or process and reactor design. From the catalyst design perspective, the introduction of hydrophobicity into catalysts to suppress water gas shift kinetics has been increasingly explored. This strategy has been proven to be effective for syngas conversion via Fischer–Tropsch Synthesis (FTS),227−230 and it has potential in syngas conversion via oxygenate intermediates as discussed in this Review. Using Cu/ZnO and ZSM-5 as a reference bifunctional catalyst for syngas conversion to DME, the functionalization of the Cu/ZnO nanoparticles using stearic acid to increase hydrophobicity appeared to be effective and resulted in higher DME selectivity.231
From a process intensification perspective, steam partial pressure could be limited by the addition of adsorbents or the use of membrane reactors, and in both cases, zeolites play an essential role as the adsorbent and membrane materials.232,233 Van Kampen et al. authored a recent review on steam separation enhanced reaction processes, in which they discuss the technical aspects and outlook of these technologies.233 Of relevance is the sorption-enhanced DME synthesis from COx conversion, in which the benefits of adding adsorbents to the direct DME synthesis process is demonstrated (Figure 15).27,234 At 275 °C, 25 bar, and H2/CO2 = 3, the addition of an LTA zeolite adsorbent to a typical bifunctional catalyst of CZA and γ-Al2O resulted in > 80% CO2 conversion and > 70% DME selectivity per pass. Notably, sorbent-enhanced processes require frequent sorbent regeneration by pressure or temperature swing, thereby complicating process design and operation. Optional reactor configurations are moving bed and fluidized bed reactors.27 The LTA zeolite is also commonly selected as a membrane for in situ water removal in the relevant processes such as methanol synthesis and MTO.235,236 ZSM-5, MOR, and SIL were theoretically evaluated as membranes for the production of DME from CO2 but were found to be unsuitable.237 These zeolites had low permselectivity toward water, resulting in high loss of methanol. Zeolite membrane reactors are also simulated for CO2 conversion to hydrocarbons via FTS, and it was concluded that the in situ removal of water enhanced hydrocarbon yield by shifting the RWGS equilibrium to a certain extent, and a further increase in water removal led to changes in hydrocarbon product distribution and hot spot formation.238 As concluded by van Kampen et al., these technological advances in process intensification are mostly attained from theoretical thermodynamic calculations, and the feasibility and potential should still be addressed experimentally.
Figure 15.

Thermodynamic (maximally possible) carbon distribution versus experimentally obtained results for sorption-enhanced DME synthesis. A H2O break-through time of approximately 20 min was indicated for the experimental data reported in this figure. Reaction conditions: 275 °C, 40 bar, and stoichiometric H2 to COx feed as shown in figure. Reproduced with permission from ref (233). Copyright 2019 Elsevier.
Last but not least, the tandem process concept of conversion of COx to DME or hydrocarbons is reassessed. Bansode and Urakawa239 reported experimental results for CO2/H2 conversion over Cu/ZnO/Al2O3 alone or mixed with ZSM-5, at an exceptionally high pressure of 360 bar. The CZA catalyst reached > 95% CO2 conversion and > 98% methanol selectivity at 260 °C with a CO2/H2 ratio of 1:10, almost identical to the equilibrium yield under these conditions. A high H2 content in the feed was required to enhance the rate of methanol formation. A mixture of CZA with H-ZSM-5 at the same pressure and CO2/H2 ratio yielded 97% CO2 conversion and 89% DME selectivity at 300 °C. Hydrocarbon production was targeted by a two-stage approach, in which CZA was loaded into the first and H-ZSM-5 into the second of two reactors in series, where the first was operated at 360 bar, 260 °C, and CO2/H2 = 1:10, and the effluent was led into the second reactor which was operated at 375 °C and either 360 or 1 bar. Operation at 360 bar yielded 95% CO2 conversion with 4% selectivity to DME and 85% selectivity to C1–C4 alkanes, while operation at 1 bar yielded 97% CO2 conversion with 10% selectivity to DME, 23% to C1–C4 alkanes, and 42% to C2–C3 alkenes.
The two-stage reactors-in-series concept is also proposed by the team of Liu for the conversion of syngas to gasoline or propane, and the advantage of decoupling reaction conditions, i.e., lower temperatures for DME synthesis and higher temperatures for DME conversion to hydrocarbons, is exemplified.240,241 The same group had earlier demonstrated a dual-bed reactor concept for DME synthesis and conversion to produce olefins from synthesis gas.242 To produce gasoline, the first reactor was loaded with CZA and Al2O3 catalysts to convert syngas to DME at 260 °C, and the second reactor was loaded with a nanosized ZSM-5 (Si/Al = 97) zeolite to convert DME to C5–C11 hydrocarbons at 320 °C.240 At 30 bar and H2/CO = 2, 79% C5–C11 hydrocarbon selectivity (excluding 32% CO2 selectivity) was achieved at 87% CO conversion. Larger ZSM-5 crystals and a lower Si/Al ratio resulted in lower activity and faster catalyst deactivation. To produce propane as illustrated in Figure 16a, CZA and ZSM-5 catalysts were loaded in the first reactor operating at 260 °C, and an SSZ-13 zeolite was loaded in the second reactor operating at 410 °C.241 At 50 bar and H2/CO = 7, propane selectivity reached 77% (excluding 11% CO2 selectivity) at 96% CO conversion. SAPO-34 and SAPO-18 zeotypes were also studied, and the lower acid strength resulted in a poor catalyst lifetime and lower propane selectivity (Figure 16b). Furthermore, the Si/Al ratios in SSZ-13 were varied, and there appeared to be a positive relation between the BAS density and propane selectivity. Three observations could be made regarding the two-stage reactors-in-series concept. First, this concept is more attractive for the production of propane than gasoline/aromatics from CO/CO2/H2, since a higher reaction temperature is needed, which means a larger temperature gap between methanol synthesis and conversion. Second, a higher BAS density is preferred for the selective production of propane but not for gasoline, and this points to different requirements of the zeolite component in such processes. Third, under the conditions of this study, the propane productivity increased from 5 mmol/g/h for the single-reactor test to 7 and 14 mmol/g/h, respectively, for the two reactors in series without and with a methanol dehydration catalyst (ZSM-5) in the first reactor.
Figure 16.
(a) Configuration of the dual-bed catalyst system for syngas-to-propane and (b) effect of different zeolites in the lower bed on the selectivity and yield of propane. Reaction conditions: T (upper bed) = 250 °C, T (lower bed) = 410 °C, 50 bar, H2/CO = 7, and 4000 mL/gcat/h. Reproduced from ref (241).
4. Ketene-Mediated COx Conversion to Hydrocarbons
4.1. Methanol Carbonylation
Methanol carbonylation to acetic acid is an industrial process of more than 15 million tonnes per annum.243 The industrial catalysts are homogeneous Rh or Ir complexes, and the reaction conditions are typically 200 °C and 30 bar. Halogen promoters/cocatalysts are required to facilitate the process. This process was first commercialized in 1960 by BASF using a Co iodide catalyst at 250 °C and 700 bar, and the BASF process was subsequently phased out due to the Monsanto process, which operated at milder conditions.244 The Monsanto process was developed in the late 1960s and is based on homogeneous Rh complex catalysts operated at 150 to 200 °C and 30 to 60 bar. In addition to the challenges associated with homogeneous catalysis, e.g., catalyst costs, stability, and recovery, a drawback of the Rh-based catalysts is the requisite of high water partial pressure to prevent precipitation of the homogeneous catalyst. The high water partial pressure and the WGS activity of the Rh sites result in 10% undesired CO2 selectivity; hence, a breakthrough was made by BP in 1996 with the development of the Cativa process.245,246 The Cativa process utilizes homogeneous Ir-based catalysts and has a similar process window as the Monsanto process, but it operates at low water partial pressure. The low partial pressure coupled with low WGS activity of the Ir-based catalysts imply that > 99% acetic acid selectivity is achieved.
To eliminate the above-mentioned challenges of processes catalyzed by homogeneous catalysts, heterogeneous catalysts based on immobilizing the homogeneous complexes on supports were developed.243,247 Most of such heterogeneous catalysts still require iodide cocatalysts, which would increase cost and complexity due to its high corrosivity. Detailed studies of the homogeneous Rh system suggested that iodide, beyond its role as cocatalyst, enhances the nucleophilicity of the metal complex in the transition state, which leads to methyl group ligand insertion.248 Zeolites are considered as support materials for the homogeneous complexes due to favorable dispersion,249 but discussion on this application of zeolites is beyond the scope here.
Inspired by the acid-catalyzed “Koch” reaction, Fujimoto and colleagues demonstrated in 1984 the feasibility of using zeolites to catalyze methanol carbonylation.250 In this study, zeolite-Y (FAU topology), MOR, and ZSM-5 were tested at 200 to 300 °C, 9.8 bar, and CO/methanol = 1. Although DME was the major product (∼90% selectivity) in all cases, acetic acid and methyl acetate were observed in the product spectrum. Cu incorporation to MOR via ion exchange was found to increase the carbonylation rate but did not change the product distribution. Smith from BP and his collaborators later optimized the Cu-MOR catalyst and process conditions (350 °C, 10 bar, CO/methanol = 10, and GHSV = 3000 h–1) to attain > 70% selectivity of acetic acid and methyl acetate.251 However, selectivity stability remained a challenge as this selectivity was stable for 12 h TOS, and hydrocarbons and DME were the major products during the first 5 h and last 5 h of reaction, respectively.
The Corma group, in collaboration with BP, utilized spectroscopic tools (operando FTIR and in situ magic-angle spinning NMR) to determine reaction intermediates during methanol carbonylation over MOR and Cu-MOR catalysts.252 Cu-MOR was selected as a bifunctional catalyst to contain a function for stabilization of methyl species and a function for CO activation, thereby overcoming the rate-determining CO insertion step. The same intermediate methoxy and acylium cations were observed over both MOR and Cu-MOR catalysts; however, acrylic acid and methyl acetate were the corresponding final products. In the case of MOR, H2O adsorption dominated, resulting in acrylic acid. On the other hand, Cu-MOR possessed an active site composed of two neighboring sites, one bridged hydroxy and a neighboring Cu+. The BAS was responsible for methanol activation, while Cu+ accounted for CO activation and preferential DME adsorption over methanol and H2O, resulting in methyl acetate production.
A relevant tandem catalysis concept was developed by the Hargreaves lab together with BP to convert methanol to acetic acid without CO feed.253 A Pd/CeO2 catalyst was used to decompose methanol to generate CO in situ, and the subsequent methanol carbonylation was catalyzed by Cu-MOR. At 300 °C, 1 bar, methanol/Ar = 0.4, and GHSV = 10,400 h–1, a mixture of CO, DME, and acetic acid was produced in approximate ratios of 300:40:1 in a stacked bed configuration. The stacked bed configuration performed better than the physically mixed bed, because methanol could decompose to formaldehyde and ultimately the needed CO. In a physically mixed bed, formaldehyde would be in close proximity with a BAS and transformed to coke before splitting to CO. This connects with the strategy of Hwang and Bhan to suppress deactivation in the MTO process, when they added Y2O3 to CHA catalysts to decompose formaldehyde into CO and H2 before formaldehyde transformation to polyaromatics (vide ultra, Section 2.1).
The detrimental influence of H2O on methanol carbonylation, stemming from parallel methanol dehydration and/or competitive adsorption, is the main driving force for the shift toward DME carbonylation. A recent breakthrough was achieved by the Liu group with their pyridine-modified MOR catalysts operating at 250 °C, 50 bar, and CO/methanol = 400.254 The higher reaction temperature was used to counter the H2O effects, and the pyridine modification served to block the 12-MR channels to inhibit hydrocarbon selectivity and coke formation. Hence, the pyridine-modified catalyst showed > 90% acetic acid selectivity at full conversion over 145 h TOS.
Beyond zeolites, Qi et al recently reported 54% acetic acid selectivity when feeding methanol and CO over Na-promoted Rh nanoclusters on ZrO2 support at 573 K, with 1.4% CO conversion.255 Even more recently, the same group achieved 96% selectivity to acetic acid over the Rh-ReO4/SiO2 catalyst at 543 K, with 39% CO conversion at a CO:methanol feed ratio of 1:1 (30 mbar each). In both cases, DME was the main byproduct.256 Comparative tests of a series of catalysts revealed that DME is formed on bulk Re2O7 and ReOx crystals, while acetic acid is formed on atomically dispersed ReO4.
4.2. DME Carbonylation
The Iglesia group in cooperation with BP in 2006 introduced the DME carbonylation process with > 99% methyl acetate selectivity over zeolite catalysts at 150 to 190 °C.257 The selectivity decreased at higher temperatures due to hydrocarbon formation. The rate of methyl acetate formation decreased from MOR to FER to ZSM-5 and was not detectable for USY and BEA at 147 to 240 °C, 10 bar, and CO/DME/Ar = 93/2/5. The rate of methyl acetate synthesis was independent of DME partial pressure but increased with CO partial pressure, hence, suggesting reactions of gas-phase or adsorbed CO with DME-derived intermediates to be rate-determining. H2O cofeeding with DME and CO experiments verified that DME carbonylation was more favorable due to the lack of competitive adsorption between H2O and CO at LAS and/or parallel methanol dehydration reactions. The elementary steps of DME carbonylation were proposed to be reversible dissociative adsorption of DME to form surface methyl species, followed by CO insertion to form surface acetyl species, and finally reaction of surface methyl and acetyl species to desorb methyl acetate.258 There were two distinct sites: one site stabilizes acidic protons and methyl/acetyl species, while the other site is responsible for binding CO. In order to locate the selective sites for DME carbonylation, Na+ or Co2+ ion-exchange was performed on MOR to replace selectively H+ sites in 8-MR and 12-MR channels, respectively.259
As demonstrated in Figure 17, the formation of C–C bonds via CO insertion was concluded to occur selectively in 8-MR channels, which matched well with the reactivity in zeolites with 8-MR channels (MOR and FER) and the lack of reactivity in zeolites without 8-MR channels (BEA, FAU, and MFI). Several theories were put forward to explain the 8-MR site requirements, including selective stabilization of cationic transition states, confinement and solvation effects, and lower ion-pair enthalpies together with lower entropies.260−262
Figure 17.
DME carbonylation rates plotted against (a) the number of total H+ sites in MOR (●,■,▼) and other zeolites (▲) and (b) the number of H+ sites per unit mass in 8-MR channels of MOR (▲), FER (◆), and 12-MR channels of MOR (●). Inset shows DME carbonylation rates plotted against the total number of H+ sites in these samples. Reaction conditions: 165 °C, 10 bar, and CO/DME/Ar = 93/2/5. Reproduced from ref (259).
The above studies on MOR were continued by Corma and co-workers with quantum-chemical methods so as to locate the active sites for methanol and DME carbonylation.263 From Figure 18, four nonequivalent tetrahedral sites in the MOR unit cell were considered, namely T1 in the 12-MR channel, T2 and T4 in the intersection between 12-MR and 8-MR, and T3 in the 8-MR channels. At the T1, T2, and T4 positions, the formation of DME and hydrocarbons were kinetically favored over carbonylation due to lack of steric hindrance. The T3-O33 position was the only selective site for carbonylation because at this position, the methoxy group was parallel to the cylinder axis, which allowed the attack of CO to fit perfectly without steric hindrance, as presented in Figure 18. From the calculated energy profile at the T3-O33 position (Figure 18), the transition state for the attack of the CO on the methoxy group possesses the highest activation energy of 23.9 kcal/mol. The acylium cation reaction intermediate is 6.9 kcal/mol more stable than the initial state, and 9.8 kcal/mol less stable than the final state. This theoretical analysis validated prior experimental studies, which proposed CO insertion to the surface methoxy group as the rate-determining step. During methanol carbonylation, H2O or methanol reaction with the acylium cation would produce acetic acid, and the resulting methoxy/water ratio = 1. During DME carbonylation, methyl acetate would be formed instead, and the resulting methoxy/water ratio = 2. Thus, the negative influence of water was doubled during methanol carbonylation, and the formation of water clusters could block the 8-MR channels.264 The negative effect of H2O on carbonylation rates originated from competitive adsorption of CO and H2O, and the displacement of equilibrium toward reactants that decreased the abundance of surface methoxy groups for CO insertion. To disintegrate the water clusters and to compensate for the lower rates, methanol carbonylation required higher reaction temperatures than DME carbonylation, which in turn decreased selectivity toward acetic acid.
Figure 18.
Structure of MOR in the (a) c and (b) b directions, with the respective O labeling. Schematic representation of the relative orientation of the O framework-CH3 bond and the channel axis at (c) the T3-O33 position of MOR and (d) any other position in an 8-MR channel. (e) Calculated energy profile for methanol carbonylation at the T3-O33 position. Reproduced from ref (263).
Several studies devoted to the in situ characterization of the MOR zeolites supported the hypothesis of acetyl intermediates. In situ solid-state NMR studies on three different MOR catalysts (MOR with both 8-MR and 12-MR channels and MOR with accessible 8-MR or 12-MR channels) provided direct evidence of methoxy species and, notably, acetyl species in 8-MR channels.265 In situ DRIFT spectroscopy revealed that the acetyl species were formed only at higher CO partial pressures of 5 to 30 bar.266 In situ solid-state NMR was also used to clarify the nature of the acetyl intermediate, and a covalent acetyl–zeolite complex was identified rather than the usually assumed acylium cation.267
Ketene was first proposed based on DFT calculations by Jensen et al. to be a potential reaction intermediate in MOR-catalyzed DME carbonylation.268 The low energy barrier (8–11 kJ/mol) for the conversion of acetyl carbocation to ketene was in a range similar to that for the direct conversion of acetyl carbocation to acetyl (1–20 kJ/mol), making both pathways feasible. To prove this hypothesis, D2O was introduced in the feed upon reaching a steady state at 165 °C, 10 bar, and CO/DME = 98/2. As the reaction between D2O and ketene was the only possibility to form doubly deuterated acetic acid, the observation of such a molecule confirmed the presence of ketene.
Much attention has been placed on the 8-MR channels due to its reactivity, but the role of the 12-MR channels should not be neglected. Both experimental and theoretical studies point to the 12-MR channels functioning as the main mass transfer routes in MOR.269,270 Due to the size of the 12-MR channels, which accommodates aromatic molecules, coking in the 12-MR channels leading to catalyst deactivation is common. Thus to improve on the catalytic TON of MOR zeolites, synthesis strategies including nanosizing of the crystals,271,272 selective dealumination,273,274 and blocking of 12-MR channels were employed.275
Promoter addition is often used in catalysis to improve performance, and Cu modification to MOR was already attempted in the first publications. Recently, Ma and co-workers performed a study in which the nature and amount of Cu species in H-MOR were systematically varied by applying various insertion and pretreatment methods. The materials were characterized by XRD, TEM, and CO adsorption IR studies and XPS. Then, they were subjected to testing as DME carbonylation catalysts at 15 atm and 200 °C in a flow reactor. A positive correlation was found between the amount of Cu0 in nanoclusters and the carbonylation reaction rate, up to a certain amount of Cu, where Cu migrated to the outer surface of the zeolite and grew into larger particles. On the other hand, the correlation between Cu+ amount and DME carbonylation rate was negative, possibly because Cu+ was covering BAS in the 8-ring pockets of MOR. Further experimental and computational results suggested a synergistic role of Cu0 and the BAS, in which DME adsorbs on the BAS, forming methanol and a methyl group, which adsorbs on Cu0 and reacts with coadsorbed CO and the methanol to finally form methyl acetate.276,277 To maintain Cu stability and dispersion, Zn could be added together with Cu via ion exchange.278 In the screening of Cu, Ni, Co, Zn, and Ag, the cations exchanged in the 8-MR channels, i.e., Cu, Ni, and Co, improved the catalysts.279 In a study focused on Co-MOR, Co2+ ions were exchanged in the 8-MR channels at low loading (Co/Al < 0.1 and Si/Al = 8.5) but were also present in the 12-MR at higher loading Co/Al = 0.1–0.25).280 Co2+ ions at the 8-MR channels facilitated the adsorption of both CO and DME molecules, while those at the 12-MR channels reduced BAS density, leading to less coking. Besides ion exchange, which decreases BAS density, framework heteroatom substitution is an option to introduce promoters. For example, Fe heteroatoms were incorporated in the 12-MR channels of the MOR framework, resulting in lower acidic strength and density which ultimately reduced carbon deposition.281 Ce heteroatom substitution into the framework also resulted in lower acidic strength, but the positive influence on catalytic performance was proposed to be due to an enrichment of Al sites in the 8-MR channels.282
MOR is the clear favorite for DME carbonylation, and in the far second spot is FER, which consists of 8-MR and 10-MR channels. The 10-MR channels of ZSM-35 (FER) are close to the size of a benzene molecule, hinting that diffusion of aromatic molecules and corresponding hard coke formation could be inhibited, arguably increasing the competitiveness of FER from the catalytic stability angle.283 Although the surface methoxy groups were calculated to form preferentially in both 8-MR and 10-MR channels of FER, CO attack on the surface methoxy group took place selectively at the 6-MR zone of the 8-MR channel of FER zeolites.284 Accordingly, the focus has been on positioning BAS sites in the 8-MR channels, and synthesis strategies include the use of various structure-directing agents to direct Al,285 and the seeding and recrystallization approach.286
Other zeolites containing 8-MR channels, including the ETL, SZR, and CHA topologies, are recently explored. The EU-12 (ETL topology) zeolite is two-dimensional with two types of straight 8-MR channels intersected by one type of sinusoidal 8-MR channel.287 At 220 °C, 15 bar, CO/DME/Ar = 93/4/3, and 2400 mL/gcat/h, the highest selectivity toward methyl acetate of 90% was attained at ∼16% DME conversion. Coke was formed mainly on the external surface of ETL due to its small channels, instead of within the channels as in MOR and FER. The SUZ-4 (SZR topology) zeolite is three-dimensional with a 10-MR channel intersected by two arrays of 8-MR channels.288 At 200 °C, 20 bar, CO/DME/Ar/He = 50/5/2.5/42.5, and 1250 mL/gcat/h, 90% methyl acetate selectivity and 10% DME conversion were stable for 100 h of TOS. The abundant 8-MR pore openings on the rod-shaped SUZ-4 facilitated the diffusion of the reactive molecules, leading to superior catalytic stability. Interestingly, zeolite topologies MOR, FER, IRN, ATS, and GON were theoretically screened using Monte Carlo associated with molecular dynamics simulation, and ATS and IRN zeolites were highlighted as potential candidates based on both diffusion dynamics and reaction kinetics.270
Industrial zeolites with MFI and CHA topologies have also been considered. At 165 °C, 1 bar, CO/DME/inert = 95/2/3, 30,000 mL/gcat/h, and Si/Al = 10, SSZ-13 (CHA) was less active than MOR but more active than FER (Figure 19a).289 According to DFT calculations in Figure 19b, the favored sites were located within the plane of the 8MR window, in agreement with the MOR literature. Comparing SSZ-13 and SAPO-34 of similar BAS density (1.7 mmol/g), the weaker acidic strength of SAPO-34 led to lower coverage of methoxy and CO surface species, resulting in lower activity (Figure 19c and d). This positive relation between acidic strength and activity is also reported for ZSM-5 (MFI) in which acidic strength was varied by substituting framework Al with heteroatoms Ga and B.290
Figure 19.
(a) Total methyl acetate (MA) production after 24 h of TOS over each framework. (b) 2 × 2 × 2 supercell of zeolite CHA with Al substituted at the sole T site to form a Brønsted acid site (BAS). Close-up viewpoint of the Brønsted acid site depicts the four crystallographically distinct O sites. Atom colors: Al (purple), Si (yellow), O (red), and H (white). (c) Maximum rates of MA production measured over the SSZ-13s as a function of Si/Al. (d) Maximum rate of MA production as a function of BAS density for SSZ-13s and SAPO-34. Reaction conditions: 165 °C, 1 bar, and CO/DME/He = 95/2/3. Reproduced from ref (289).
4.3. Integration of Methanol and DME Carbonylation with COx Hydrogenation
In the first report on the Ox-Zeo process, Bao and co-workers proposed ketene to be the key intermediate in the conversion of syngas to lower olefins using ZnCrOx/SAPO-34 catalysts, because ketene was detected using vacuum ultraviolet photoionization mass spectrometry.33 The methanol cofeeding experiments leading to a drastic drop in catalytic performance substantiated the importance of ketene instead of methanol. In order to exploit the ketene intermediate for steering hydrocarbon selectivity, the solid acid component was switched from SAPO-34 to MOR zeolites.291 At 360 °C, 20 bar, H2/CO = 1, and 1857 mL/gcat/h, 73% ethylene selectivity (excluding 48% CO2 selectivity) was achieved at 26% CO conversion using a ZnCrOx/MOR catalyst combination in which the MOR had only 8-MR sites accessible to reactivity. Employing solid-state NMR spectroscopy and 129Xe as a probe molecule, the preferential adsorption of ketene in the 8-MR sites and methanol in the 12-MR sites were verified. The subsequent correlations of hydrocarbon selectivity to the number of sites accessible in the 8-MRs were consistent with the proposition of ketene as a key intermediate (Figure 20). To broaden the scope, ZnGa2O4, Ga2O3, and ZnO were mixed with MOR and tested at 400 °C, 40 bar, H2/CO = 2.5, and 1600 mL/gcat/h.292 Arguing that since the three oxides produced similar levels of methanol and DME and an identical MOR was used for the formation of hydrocarbons, the differences in hydrocarbon selectivity pointed to the presence of ketene as an intermediate. Interestingly, the ex situ NMR analysis detected carbonylation products including acetate and propionate. Specifically, these carbonylation products and the initial C–C bond via ketene were proposed to be formed on the oxide surface.
Figure 20.
Hydrocarbon distribution in the conversion of syngas, ketene, and methanol over different sites of the MOR zeolites at 375 °C. (a–c) MOR#2-py with only the 8-MR acid sites accessible, (d–f) MOR#14 with only the 12-MR acid sites accessible, and (g–i) MOR#3 with both the 8-MR and 12-MR acid sites available. (a, d, and g) Syngas over ZnCrOx-MOR, (b, e, and h) ketene conversion over MOR, and (c, f, and i) methanol conversion over MOR. Reproduced with permission from ref (291). Copyright 2018 John Wiley & Sons.
With the intention to determine if the key intermediate is indeed ketene or methanol, theoretical understanding of the ZnCrOx systems is needed.293,294 DFT calculations and microkinetic simulations on the highly reduced ZnCr2O4 (110) surface revealed the propensity of CO to absorb on the O vacancy sites, followed by reaction with H to form a CHO surface species.293 The distinction between ketene and methanol is dependent on what happens next to the CHO surface species. To form ketene, CHO is first dissociated to CH and O, followed by hydrogenation to CH2 and finally CO insertion. To form methanol, CHO goes through consecutive hydrogenation steps. Quantitatively, the pathway to ketene requires less energy, so ketene is more readily formed than methanol. This finding is supported by a parallel microkinetic modeling study on the ZnCr2O4 (111) surface, which is calculated to be the most stable surface under reaction conditions.294 Accordingly, the surface coverage by CH3CO (ketene precursor) is 11 times higher than that by CH3O (methanol precursor).
Although the above theoretical studies supported the hypothesis of ketene as a key intermediate, they were applicable only for ZnCrOx catalysts, and such calculations should be extended to other relevant bulk oxides. With reference to Figure 21, a comparable study on ZnO surfaces found a correlation between surface oxygen vacancies and product selectivity.295 Although ketene is formed on all ZnO surfaces, the methanol production is favored over the more oxidized ZnO surfaces. On a ZnO0.75 surface, the selectivity toward methanol and ketene is 80% and 20%, respectively. With an increase in the number of oxygen vacancies, the major product evolves from methanol to ketene to methane. On a metallic Zn36 cluster surface, the methane and ketene selectivities are 57% and 43%, respectively.
Figure 21.
(a) Complete reaction networks of syngas to methane, ketene, and methanol and (b) a specific reaction pathway for ketene production. Orange arrows represent the evolution cycle of carbon-containing species, and blue arrows represent hydrogenation and dehydrogenation. (c) Collection of ZnO surfaces with a varying concentration of oxygen vacancies. A Zn36 cluster is used to account for the reactivity of small Zn clusters observed in experiments. (d) Microkinetic modeling results of selectivity over ZnO0.75, ZnO0.00, and the Zn cluster. Reproduced from ref (295).
The scope of the ketene-mediated process is expanded with a recent contribution from the Fan group which compared the performance of a set of methanol synthesis catalysts, Cr2O3, InZrOx, and ZnMOx; M = Zr, Al, Ga, and Cr, with SAPO-34, under a given set of conditions (Figure 22).296 In this case, InZrOx and ZnMOx mixed with SAPO-34 yielded very similar product distributions, both with respect to CO versus hydrocarbon selectivity and C2:C3:C4 ratios. Notably, the product distribution obtained with the Cr2O3/SAPO-34 mixture was distinctively different, with the hydrocarbon selectivity to C2 being 66%, compared to 37% over both InZrOx/SAPO-34 and ZnMOx/SAPO-34. Mechanistic investigations showed that acetic acid (as an intermediate product) and ethanol (as a final product) were present both on the surface and in the effluent from Cr2O3, together with methanol. Mixing Cr2O3 with SAPO-34, ethanol was rapidly converted to ethene, thereby yielding the superior ethene selectivity of this tandem catalyst. The reason why ethene persisted in contact with SAPO-34, instead of being incorporated into the hydrocarbon pool, may be linked to the diffusion restrictions of ethene in small-pore zeolites such as SAPO-34, which were recently explored by the van Speybroeck group.297,298 Comparatively, only methanol was detected as the surface and effluent product from ZnZrOx. When mixed with SAPO-34, this catalyst yielded a product distribution more similar to that obtained over SAPO-34 under MTO conditions.139
Figure 22.
(a) CO2 conversion and product distribution and (b) olefins selectivity over various metal oxide/H-SAPO-34 composite catalysts. Reaction conditions: 370 °C, 5 bar, H2/CO2 = 3, and 4000 mL/gcat/h. Reproduced with permission from ref (296). Copyright 2022 Elsevier.
Besides the above studies which highlight ketene as an important intermediate and ethene as a main product, the more classic approach to integrate DME carbonylation with syngas conversion results in acetic acid, methyl acetate, and ethanol as major products. The proof-of-concept was demonstrated by Wang and co-workers, in which syngas was first converted to DME using a mixed bed of CZA and HZM-5 catalysts, and the effluent stream containing DME was passed through MOR to produce methyl acetate and acetic acid.299 Due to the poor CZA catalyst stability at higher temperatures, a ZnAl2O4 catalyst was used to convert syngas to DME at higher temperatures between 325 and 370 °C. At 370 °C, 30 bar, and H2/CO = 1, 87% selectivity toward methyl acetate and acetic acid (excluding 20% CO2 selectivity) was attained at 11% CO conversion. This concept was extended to steer the product selectivity toward ethanol and ethylene by means of catalyst bed configurations (see Figure 23). Recognizing the value of ethanol, a follow-up study was devoted to improving ethanol selectivity via catalyst development, catalyst bed configuration, and reaction parameter optimization.300 The optimal catalyst configuration was three stacked beds of K/ZnO/ZrO2, MOR, and Pt/Sn/SiC catalysts for the formation of methanol, acetic acid, and ethanol, respectively. At 270 °C, 50 bar, and H2/CO = 1, 80% ethanol selectivity (excluding < 10% CO2 selectivity) was achieved at 4% CO conversion. Notably, Al in the 12-MRs of the MOR catalyst was selectively dealuminated to improve the acetic acid selectivity.
Figure 23.

Effect of arrangements of ZnAl2O4 and H-MOR on catalytic conversion of syngas. (a) ZnAl2O4. (b) ZnAl2O4; H-MOR. (c) ZnAl2O4; H-MOR; ZnAl2O4. (d) ZnAl2O4; H-MOR; ZnAl2O4; H-MOR. (e) Mixture of ZnAl2O4 and H-MOR granules with sizes of 250–600 mm. (f) Mixture of ZnAl2O4 particles with sizes of 4–9 nm and H-MOR particles with sizes of 0.3–1 mm. Reaction conditions: H2/CO = 1; P = 3 MPa; F = 25 mL min–1; time on stream = 20 h; total weights of ZnAl2O4 and H-MOR = 0.33 and 0.67 g. Reproduced with permission from ref (299). Copyright 2018 John Wiley & Sons.
Focusing back to the progress in zeolite research, strategies were inspired from prior work on Koch carbonylation, such as improved CO binding, increasing the number of selective sites in the 8-MRs and decreasing the number of unselective sites in 12-MRs in the zeolites. For instance, Cu modification on MOR was investigated and expectedly promoted CO activation and carbonylation.301 However, it was also found that the enhanced electrostatic interaction between Cu+ and the acetyl cation resulted in ethane formation. As this latter observation differs from prior understanding, this could be due to the differences in reaction conditions (e.g., feed composition, temperature, and pressure). Another example is the selective phosphate passivation of the BAS in the 12-MR channels using trimethylphosphite as a phosphate precursor, resulting in over 96% selectivity toward C2 oxygenates.302
Last but not least, the core–shell capsule catalyst approach was investigated by means of a Cu/ZnO core and a micron-sized MOR zeolite shell.303 Importantly, the influence of MOR shell thickness between 2 and 40 μm was studied, and a minimum of 20 μm thickness was required to attain 50% ethanol selectivity at 220 °C, 15 bar, H2/CO = 1, and 10–14% CO conversion. In addition, Cu was introduced via ion exchange to enhance DME carbonylation in the MOR shell.
The successful development of carbonylation catalysts that are active at moderate temperatures hints to an alternative route for C3+ olefins formation, in which acetic acid may be reduced to form ethanol and potentially dehydrolyzed to ethene and subjected to a second carbonylation step. Along this line, Ahlers et al. recently reported the selective formation of 1-propanol from CO2, H2, and ethene over supported Au catalysts.304,305 The reaction was found to proceed by reduction of the CO2 to CO over the Au nanoparticles, followed by selective carbonylation and subsequent partial hydrogenation to form propanol. Small Au nanoparticles on a K-promoted TiO2 support were found to yield close to 100% selectivity to 1-propanol at 20 bar, 200 °C, and CO2/H2/C2H4=1:1:1.
Another moderate-temperature path to C3+ olefins from ethene is the oligomerization reactions. Industrially, ethene oligomerization processes are carried out over metal complex catalysts in solution using organometallic cocatalysts.306 Major efforts have been made in recent years to develop heterogeneous analogues for this process, including zeolite-based catalysts. Those processes are not yet competitive with the homogeneous processes, but some interesting observations have been made and are mentioned here. Please note that a detailed assessment of the oligomerization process is beyond the scope of this study, and we refer the interested reader to two recent, excellent review papers by Finiel et al. and Olivier-Bourbigou et al.307,308 Our brief description focuses on studies carried out in fixed bed reactors with ethene gas flow.
As a first observation, studies of catalysts consisting of Ni ions impregnated or ion-exchanged onto Si-Al-O supports, showed that they are active for ethene oligomerization without a cocatalyst, regardless of the (micro-, meso- or macro-) pore structure of the support material.308 Experimental and theoretical evidence led several groups to conclude that reaction between a sacrificial ethene molecule, an open Ni coordination site, and a neighboring Brønsted acid site leads to the formation of a Ni–H site, active for ethene oligomerization. However, the issue is still under debate.308 Focusing next on zeolite supports, the Martinez group studied Ni-exchanged and -impregnated H-Beta zeolite and reported a linear increase in ethene conversion with increasing Ni content, until the molar concentration of Ni corresponded to the Brønsted acid concentration of the parent sample, where conversion leveled out. They concluded that single Ni ion sites are responsible for ethene oligomerization.309 The conclusion has found support from several independent studies of Ni-zeolite catalysts.308 The Martinez group study was carried out at 120 °C, 26 bar ethene pressure, and WHSV (ethene) = 2.1 h–1. It further showed that the selectivity to odd-numbered alkenes decreased with increasing Ni content for Ni/H < 1, suggesting carbocation-mediated cracking over free BAS.309 Beyond the active site, several studies point to the influence of zeolite pore size and structure. The Hulea group reported a steep increase in ethene conversion level with an increase in pore size of MCM materials, at 150 °C, 35 bar, and 1 h reaction in batch mode.307 They ascribed the difference to rapid coke formation, leading to pore blocking in the microporous materials. Recent studies showed that the microporous structure of the zeolite alters both reaction kinetics and product selectivity compared with amorphous supports. Koch et al. studied the kinetics of ethene oligomerization over 1.8 wt % Ni/SiO2-Al2O3 at 443–503 K, ethene partial pressure 1.5–3.5 bar, and τ = 4.8–14.4 kgcat·s·molC2–1. Microkinetic modeling of the test data revealed a first-order reaction in P(ethene) and a Schulz–Flory product distribution, in accordance with the degenerate polymerization (Cossee–Arlman) mechanism.310 The same product distribution was observed for Na-exchanged Ni/Y zeolite (11.2 Å cavity size) at 100–150 °C, 35 bar, and WHSV (ethene) = 2 h–1. Recently, experimental studies of Ni/SSZ-24 (7.3 Å pore size) at P(ethene) = 4–26 bar and 130–170 °C revealed a second-order reaction rate in ethene with 35 kJ/mol activation energy, and > 98% selectivity to linear butenes at 1–7% conversion. DFT-based static and molecular dynamics simulations pointed to a degenerate polymerization mechanism and suggested that the second-order rate dependence on P(ethene) is due to formation of an [(ethene)2-Ni-ethyl]+ ion which is detached from the framework position and stabilized by the framework in the transition state, leading to a substantially decreased activation energy for butene formation. The high C4 selectivity was allocated to lower rotational freedom of the corresponding [(ethene)2-Ni-butyl]+ ion, leading to retained, higher activation energy for C6 formation.311,312
Overall, lower activity and more rapid deactivation due to pore clogging by coke make Ni/zeolite-based ethene oligomerization catalysts less attractive than their mesoporous counterparts. Future efforts may potentially challenge this conclusion by using lamellar zeolite sheets that retain shape selectivity but offer shorter diffusion paths. Carrying out the process at temperatures and pressures where ethene is in a condensed state, may further promote the heterogeneously catalyzed process by lowering the product desorption energy, as recently demonstrated by Agirrezabal-Telleria and Iglesia for Ni/MCM-41.313 The product solvation effect of condensed ethene leads to a dramatic reduction in catalyst deactivation. However, a recent study revealed that this effect is only observed for mesoporous catalysts, while microporous catalysts (BEA and FAU) suffer from pore clogging by hydrocarbon residues even under capillary condensation conditions. The reason may be that the microporous pores are too small to enable packing of the ethene molecules around the active site containing the alkyl intermediate.314
5. Conclusions and Outlook
CO2 valorization is a key technology for the postfossil society, where it will enable production of consumer goods with properties equal to those obtained by converting fossil carbon, as well liquid fuels for energy storage. The oxygenate-mediated Ox-Zeo tandem process for syngas (CO + H2) conversion to hydrocarbons in a single reactor and its expansion to comprise CO2 hydrogenation in that same reactor have the potential to become a leading technology for such conversion. It offers exceptional selectivity to desired product ranges: Light olefins, BTX, or gasoline range aromatic/paraffin blends. Several decades of zeolite and zeotype catalyst studies of oxygenate (methanol and dimethyl ether) conversion into hydrocarbons yields a platform for further selectivity optimization.
Recent contributions cited in this Review demonstrate that the characteristic shape selectivity offered by zeolites and zeotypes dominates effluent product composition in the tandem process, to a similar extent as in methanol and dimethyl ether to hydrocarbons processes. However, they also demonstrate that the altered gas composition induced by the tandem process, especially the H2, CO, and H2O feed content relative to methanol, influences product formation rate and selectivity. The presence of the external, methanol-producing cocatalyst alters product formation rate and selectivity even further, especially in cases where the zeolite/zeotype pores are large enough to enable diffusion of intermediates and products in and out of the zeolite/zeotype crystals for intermittent interaction with the hydrogenation catalyst.
From a process perspective, a major advantage of the methanol-mediated conversion of COx and H2 to hydrocarbons is the superior TON of this process compared to MTH alone, especially for 8-ring window-cavity structures such as SAPO-34. This feature may enable the use of fixed bed reactors instead of the fluidized bed reactor with continuous regeneration that is currently used for the SAPO-34-catalyzed MTO/DMTO process. On the other hand, the rather low product formation rates induced by the thermodynamically limited, low methanol/water ratio in the tandem reactor may require recycle rates comparable to those used in the methanol-forming reactor of conventional reactor-in-series processes. Rapid conversion of methanol to dimethyl ether could probably mitigate this limitation to some extent due to the higher equilibrium conversion to DME than methanol at a given set of conditions. Recent contributions further suggest that a two-reactor tandem approach may attenuate the conversion limitation, by optimizing the conditions in the first reactor for high DME yields and those in the second reactor for the MTH/DTH conversion. The two-reactor approach maintains the advantage of the tandem process in terms of catalyst TON, due to the ensemble of gas components fed with methanol and DME to the second reactor. On the other hand, the opportunity of direct, synergetic interaction between the two catalyst functions is lost.
Another option to maximize product yield and conversion rates is continuous removal of water in the tandem reactor or between two-stage reactors. However, further studies are needed to elucidate the effect of water removal on coke production.
Focusing beyond effluent product composition and into the zeolite/zeotype, cofeed and tandem process studies combined with operando spectroscopy and (transient) Guisnet-type analysis of retained hydrocarbon and carbohydrate species clearly demonstrate that the hydrocarbon pool composition is strongly affected by the presence of H2, CO, and water. More precisely, hydrogen cofed with methanol leads to hydrogenation of precursors to aromatics within the zeolite, resulting in dramatically higher turn-over numbers (TON) before deactivation by coke formation than for a methanol-only feed. Higher acid strength promotes hydrogenation more than sites with lower acid strength. CO, when cofed with methanol, leads to enhanced carbonylation activity and is incorporated in aromatic products, while simultaneously limiting the conventional hydrogen transfer route to aromatics formation. When cofed with H2 and methanol, CO hinders olefin hydrogenation reactions while maintaining the high TON numbers installed by H2 cofeed. H2O cofed with H2 and methanol has a similar effect on olefins hydrogenation as the CO/H2/CH3OH cofeed but also leads to degradation of the zeolite framework over time. Very recently, the observation of abundant carbonylated species during steady-state operation of tandem catalyst systems by using quasi-in situ solid-state NMR and 12C/13C feedstock led to the proposal of a “triple cycle” mechanism. Here, carbonylation reactions via cyclopentenyl species is suggested as the main route to convert alkenes to arenes in the dual cycle mechanism, hence replacing the traditional methanol dehydrogenation route via formaldehyde. Together, this novel insight emphasizes the complexity and fine balance required between various parameters of the tandem process to ensure stable formation of the desired product mixture over time. It also opens an exciting new playground for fundamental studies of site–structure–conditions–function correlations in this new process.
A special case of the methanol-mediated tandem reaction occurs when C2 products are formed on the COx hydrogenation catalyst alongside methanol. Ketene, acetyl, acetate, acetic acid, and ethene (or its hydrated analogue ethanol) are all constituents of the initiation mechanism of the hydrocarbon pool in traditional MTH chemistry. When those C2 compounds are formed externally to the zeolite/zeotype, the slow step of the autocatalytic MTH reaction, initial C–C bond formation, may become redundant, potentially yielding higher product formation rates. When such acceleration is not observed in tandem studies, it may be due to the high H2O/methanol ratio induced by the thermodynamic limitation of methanol formation under tandem conditions and further enhanced by the competitive adsorption of these two products on the BAS of the zeolite/zeotype. Another reason may be the slow diffusion or low reactivity of C2 products formed externally, especially in small-pore (8-ring) zeolites/zeotypes: the high ethene selectivity observed in the studies referred to above, where ethanol or ketene was produced externally, suggests that they were poorly integrated in the hydrocarbon pool. Nevertheless, the growing number of materials capable of catalyzing C2 compound formation from CO2/H2, some of them referred to in this Review, pave the way for new directions within materials development for the oxygenate-mediated conversion of CO2/H2 to hydrocarbons: Brønsted and Lewis acid sites and their surrounding matrix may be tuned for optimal performance, inside and outside the microporous window/pore/cavity structure of the zeolite/zeotype.
In the Introduction, we promised a comparison between methanol versus DME and ketene as mediators in the oxygenate-mediated conversion of CO2/H2 to hydrocarbons. As an initial statement, recent studies clearly established that both DME and ketene are formed from the methanol feed in the zeolite/zeotype-catalyzed MTH process. Hence, the core of this comparison must be whether it is advantageous to add to the reactor another catalyst component that promotes formation of DME or ketene, beyond the lattice components of the zeolite/zeotype crystals.
Starting with a comparison between methanol and DME, it is clear that in a dual reactor concept, higher yields of DME than methanol are attainable in the first reactor (CO2/H2 to oxygenate) due to thermodynamic limitations. The introduction of a methanol dehydration catalyst in this first reactor will reduce the heat release in the second reactor and enable separation of more H2O from the effluent of the first reactor compared to having methanol as the oxygenate mediator. Potential water separation between the reactors will reduce competitive water sorption on the Brønsted acid sites in the second reactor, thereby enhancing conversion rates and mitigating catalyst degradation. When adding the established role of methanol as a hydrogen transfer agent in aromatics and coke formation during MTH operation (to be reconsidered under CO-rich conditions, vide ultra), the advantages of DME-mediated dual-reactor operation seem overwhelming. The fact that DME production from synthesis gas is already an integral part of several industrial processes for hydrocarbons production (MTP, TIGAS, and DMTO) adds to this picture. However, the influence of H2, CO2, and CO cofeed with DME in the second reactor is yet to be studied in detail and could need further attention.
Proceeding to the single-reactor approach, where CO2 hydrogenation and oxygenate conversion catalysts are mixed, the effect of introducing a methanol dehydration catalyst to establish methanol–DME equilibrium is less straightforward. The slower diffusion of DME versus methanol into 8-ring window-cavity structures might require the dehydration function to be installed inside the cavities, which is where product formation takes place. In medium- to large-pore zeolites and zeotypes, such as ZSM-5, promotion of methanol dehydration has been shown to yield positive effects on catalyst selectivity toward effluent products instead of coke under MTH conditions. Still, detailed studies of H2, CO2, and CO cofeeds with DME (suggested above), as well as the effect of DME/water versus methanol on the dual catalyst functions, are yet to be pursued.
Proceeding next to ketene-mediated routes to hydrocarbons, individual studies of zeolites/zeotypes and metal oxides, respectively, showed that a far higher CO/methanol ratio is needed to carbonylate methanol to ketene in the Brønsted acid-based zeolites/zeotypes compared to dense metal oxides, which have mainly Lewis acid character. Hence, there are clear advantages to external ketene production. As pointed out above, ketene may “kick-start” the MTH reaction in the zeolite/zeotype, provided that the zeolite/zeotype pores are big enough to enable ketene to enter and react. Otherwise, ketene may be a source of a high ethene selectivity. On the opposite side, too high ketene concentrations could be a source of coke formation, so a right balance of conditions leading to a suitable steady-state concentration is important.
Overall, the wealth of literature studies referred to in the individual chapters and subchapters of this Review provide design rules for dense as well as microporous catalysts leading to the formation of either methanol, DME, or ketene in the presence of the molecular precursors to each product. The subchapters devoted to the use of those products as mediators in the tandem reaction reveal further design options for catalyst, process, and reaction conditions. Although some critical issues remain to be solved, we are confident that the current state-of-the-art knowledge represents a stepping-stone for further important discoveries and developments in the years to come, paving the way for a successful transition to the postfossil society.
Among future research opportunities, we highlight:
(1) Catalyst optimization:
A wealth of zeolite/zeotype structures, compositions, and BAS/LAS siting are yet to be explored for the tandem reaction. Inspiration for catalyst design may be found from prior studies cited herein.
Catalyst morphology may be altered to facilitate product diffusion.
Rapid/selective DME or C2 formation before the zeolite/zeotype function may accelerate hydrocarbons formation in the zeolite/zeotype compared to methanol-mediated routes and mitigate catalyst deactivation, and should be further explored.
Long-term catalyst performance, in particular the effect of repeated activation-reaction-regeneration cycles, is underexplored and needs further attention.
Potential element migration within and between catalyst functions during synthesis, activation, testing, and regeneration calls for advanced catalyst characterization studies before, during, and after test cycles.
Organo-catalysis by guest species in zeolite/zeotype pores and cavities (beyond conventional reaction intermediates and products) may promote desired reactions.315
(2) Process optimization:
Continuous water removal by membrane- or sorbent-based reactor design may enhance hydrocarbon formation rates, potentially at the expense of TON. Studies of the effect of CO2, CO, H2, and H2O concentration on process thermodynamics and kinetics is recommended, in particular for the less-explored oxygenate intermediates (DME and C2).
Two temperature zones in a single reactor with a first temperature zone for DME synthesis and a second temperature zone for DME conversion to hydrocarbons is an interesting opportunity. The dual- versus single-reactor tandem options should be compared.
Alternative heating methods may be worthy of consideration.316−318
As a final remark, the purpose of using CO2 as carbon feedstock to the chemical industry is to mitigate the negative impact of human activity on our living environment. In this context, the final selection of catalyst, process, and feedstock for industrial implementation of the oxygenate-mediated conversion of CO2 and H2 to hydrocarbons should not rely only on economic and energy considerations of the individual process but include cradle-to-grave Life Cycle Analysis of all involved elements.
Acknowledgments
This work has been carried out with the financial support of the European Union through the Horizon 2020 research and innovation program under the grant agreement 837733 (COZMOS).
Biographies
Jingxiu Xie attained her BSc from National University of Singapore in 2010 and her MSc in chemical engineering, cum laude from Eindhoven University of Technology in 2012. She received her PhD in heterogeneous catalysis from Utrecht University in 2017. Her PhD thesis was supervised by Prof. Krijn P. de Jong and focused on the Fischer–Trospch to Olefins technology. After working 2 years at BasCat, a UniCat BASF Jointlab in Berlin, and 1 year at University of Oslo, she is appointed as a tenure-track assistant professor at University of Groningen in 2020. Her research group on catalytic processes for gas conversion focuses on process intensification and catalyst development for CO2/CO/H2 conversion to chemicals and fuels.
Unni Olsbye obtained an MSc in Industrial Chemistry at NTNU in 1987 and a PhD in Organic Chemistry at the University of Oslo (UiO) in 1991. She then worked at SINTEF (1991–2000) and Nordox Industrier before returning to UiO as an associate professor in 2001. She was promoted to full professor in 2002. In 2008–15, she was managing director of inGAP, a national center for research-based innovation. She is an elected member of the Norwegian Academy of Technical Sciences and the Norwegian Academy of Science and Letters. She serves as a senior editor of Journal of Catalysis and is a member of several scientific advisory boards. Her research group focuses on catalysis for sustainable valorization of light molecules (C1–C3), using studies of kinetic and mechanistic consequences of single material parameter variation as a guiding tool for catalyst design.
The authors declare no competing financial interest.
References
- IEAGHG . CO2 Utilisation: Hydrogenation Pathways. https://ieaghg.org/ccs-resources/blog/new-ieaghg-report-co2-utilisation-hydrogenation-pathways (accessed June 13, 2023).
- IEA . CO2 Emissions in 2022. https://www.iea.org/reports/global-energy-review-co2-emissions-in-2021-2 (accessed June 13, 2023).
- Zhou W.; Cheng K.; Kang J.; Zhou C.; Subramanian V.; Zhang Q.; Wang Y. New Horizon in C1 Chemistry: Breaking the Selectivity Limitation in Transformation of Syngas and Hydrogenation of CO2 into Hydrocarbon Chemicals and Fuels. Chem. Soc. Rev. 2019, 48, 3193–3228. 10.1039/C8CS00502H. [DOI] [PubMed] [Google Scholar]
- Bao J.; Yang G.; Yoneyama Y.; Tsubaki N. Significant Advances in C1 Catalysis: Highly Efficient Catalysts and Catalytic Reactions. ACS Catal. 2019, 9, 3026–3053. 10.1021/acscatal.8b03924. [DOI] [Google Scholar]
- Khodakov A. Y.; Chu W.; Fongarland P. Advances in the Development of Novel Cobalt Fischer–Tropsch Catalysts for Synthesis of Long-Chain Hydrocarbons and Clean Fuels. Chem. Rev. 2007, 107, 1692–1744. 10.1021/cr050972v. [DOI] [PubMed] [Google Scholar]
- Torres Galvis H. M.; de Jong K. P. Catalysts for Production of Lower Olefins from Synthesis Gas: A Review. ACS Catal. 2013, 3, 2130–2149. 10.1021/cs4003436. [DOI] [Google Scholar]
- Xie J.; Paalanen P. P.; van Deelen T. W.; Weckhuysen B. M.; Louwerse M. J.; de Jong K. P. Promoted Cobalt Metal Catalysts Suitable for the Production of Lower Olefins from Natural Gas. Nat. Commun. 2019, 10, 1–10. 10.1038/s41467-018-08019-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Xie J.; Yang J.; Dugulan A. I.; Holmen A.; Chen D.; de Jong K. P.; Louwerse M. J. Size and Promoter Effects in Supported Iron Fischer–Tropsch Catalysts: Insights from Experiment and Theory. ACS Catal. 2016, 6, 3147–3157. 10.1021/acscatal.6b00131. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhong L.; Yu F.; An Y.; Zhao Y.; Sun Y.; Li Z.; Lin T.; Lin Y.; Qi X.; Dai Y.; Gu L.; Hu J.; Jin S.; Shen Q.; Wang H. Cobalt Carbide Nanoprisms for Direct Production of Lower Olefins from Syngas. Nature 2016, 538, 84–87. 10.1038/nature19786. [DOI] [PubMed] [Google Scholar]
- Zhai P.; Xu C.; Gao R.; Liu X.; Li M.; Li W.; Fu X.; Jia C.; Xie J.; Zhao M.; Wang X.; Li Y. W.; Zhang Q.; Wen X. D.; Ma D. Highly Tunable Selectivity for Syngas-Derived Alkenes over Zinc and Sodium-Modulated Fe5C2 Catalyst. Angew. Chem., Int. Ed. 2016, 55, 9902–9907. 10.1002/anie.201603556. [DOI] [PubMed] [Google Scholar]
- Bezemer G. L.; Radstake P. B.; Falke U.; Oosterbeek H.; Kuipers H. P. C. E.; Van Dillen A. J.; De Jong K. P. Investigation of Promoter Effects of Manganese Oxide on Carbon Nanofiber-Supported Cobalt Catalysts for Fischer–Tropsch Synthesis. J. Catal. 2006, 237, 152–161. 10.1016/j.jcat.2005.10.031. [DOI] [Google Scholar]
- Morales F.; de Smit E.; de Groot F. M. F.; Visser T.; Weckhuysen B. M. Effects of Manganese Oxide Promoter on the CO and H2 Adsorption Properties of Titania-Supported Cobalt Fischer–Tropsch Catalysts. J. Catal. 2007, 246, 91–99. 10.1016/j.jcat.2006.11.014. [DOI] [Google Scholar]
- Sartipi S.; Makkee M.; Kapteijn F.; Gascon J. Catalysis Engineering of Bifunctional Solids for the One-Step Synthesis of Liquid Fuels from Syngas: A Review. Catal. Sci. Technol. 2014, 4, 893–907. 10.1039/C3CY01021J. [DOI] [Google Scholar]
- Duyckaerts N.; Bartsch M.; Trotus I.-T.; Pfander N.; Lorke A.; Schuth F.; Prieto G. Intermediate Product Regulation in Tandem Solid Catalysts with Multimodal Porosity for High-Yield Synthetic Fuel Production Angewandte. Angew. Chem., Int. Ed. 2017, 56, 11480–11484. 10.1002/anie.201705714. [DOI] [PubMed] [Google Scholar]
- de Jong K. P. Surprised by Selectivity. Science 2016, 351, 1030–1032. 10.1126/science.aaf3250. [DOI] [PubMed] [Google Scholar]
- Olsbye U. Single-Pass Catalytic Conversion of Syngas into Olefins via Methanol. Angew. Chem., Int. Ed. 2016, 55, 7294–7295. 10.1002/anie.201603064. [DOI] [PubMed] [Google Scholar]
- Chen J. Q.; Bozzano A.; Glover B.; Fuglerud T.; Kvisle S. Recent Advancements in Ethylene and Propylene Production Using the UOP/Hydro MTO Process. Catal. Today 2005, 106, 103–107. 10.1016/j.cattod.2005.07.178. [DOI] [Google Scholar]
- Ma Z.; Porosoff M. D. Development of Tandem Catalysts for CO2 Hydrogenation to Olefins. ACS Catal. 2019, 9, 2639–2656. 10.1021/acscatal.8b05060. [DOI] [Google Scholar]
- Kasipandi S.; Bae J. W. Recent Advances in Direct Synthesis of Value-Added Aromatic Chemicals from Syngas by Cascade Reactions over Bifunctional Catalysts. Adv. Mater. 2019, 31, 1803390–1803408. 10.1002/adma.201803390. [DOI] [PubMed] [Google Scholar]
- Ye R. P.; Ding J.; Gong W.; Argyle M. D.; Zhong Q.; Wang Y.; Russell C. K.; Xu Z.; Russell A. G.; Li Q.; Fan M.; Yao Y. G. CO2 Hydrogenation to High-Value Products via Heterogeneous Catalysis. Nat. Commun. 2019, 10, 1–10. 10.1038/s41467-019-13638-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lohr T. L.; Marks T. J. Orthogonal Tandem Catalysis. Nat. Chem. 2015, 7, 477–482. 10.1038/nchem.2262. [DOI] [PubMed] [Google Scholar]
- Fogg D. E.; Dos Santos E. N. Tandem Catalysis: A Taxonomy and Illustrative Review. Coord. Chem. Rev. 2004, 248, 2365–2379. 10.1016/j.ccr.2004.05.012. [DOI] [Google Scholar]
- Wasilke J. C.; Obrey S. J.; Baker R. T.; Bazan G. C. Concurrent Tandem Catalysis. Chem. Rev. 2005, 105, 1001–1020. 10.1021/cr020018n. [DOI] [PubMed] [Google Scholar]
- Sancheti S. P.; Urvashi; Shah M. P.; Patil N. T. Ternary Catalysis: A Stepping Stone toward Multicatalysis. ACS Catal. 2020, 10, 3462–3489. 10.1021/acscatal.9b04000. [DOI] [Google Scholar]
- Climent M. J.; Corma A.; Iborra S.; Sabater M. J. Heterogeneous Catalysis for Tandem Reactions. ACS Catal. 2014, 4, 870–891. 10.1021/cs401052k. [DOI] [Google Scholar]
- Zecevic J.; Vanbutsele G.; De Jong K. P.; Martens J. A. Nanoscale Intimacy in Bifunctional Catalysts for Selective Conversion of Hydrocarbons. Nature 2015, 528, 245–254. 10.1038/nature16173. [DOI] [PMC free article] [PubMed] [Google Scholar]
- van Kampen J.; Boon J.; Vente J.; van Sint Annaland M. Sorption Enhanced Dimethyl Ether Synthesis for High Efficiency Carbon Conversion: Modelling and Cycle Design. J. CO2 Util. 2020, 37, 295–308. 10.1016/j.jcou.2019.12.021. [DOI] [Google Scholar]
- Jiang X.; Nie X.; Guo X.; Song C.; Chen J. G. Recent Advances in Carbon Dioxide Hydrogenation to Methanol via Heterogeneous Catalysis. Chem. Rev. 2020, 120, 7984–8034. 10.1021/acs.chemrev.9b00723. [DOI] [PubMed] [Google Scholar]
- Saravanan K.; Ham H.; Tsubaki N.; Bae J. W. Recent Progress for Direct Synthesis of Dimethyl Ether from Syngas on the Heterogeneous Bifunctional Hybrid Catalysts. Appl. Catal. B Environ. 2017, 217, 494–522. 10.1016/j.apcatb.2017.05.085. [DOI] [Google Scholar]
- Fujimoto K.; Saima H.; Tominaga H. Synthesis Gas Conversion Utilizing Mixed Catalyst Composed of CO Reducing Catalyst and Solid Acid. J. Catal. 1985, 94, 16–23. 10.1016/0021-9517(85)90078-8. [DOI] [Google Scholar]
- Inui T.; Kitagawa K.; Takeguchi T.; Hagiwara T.; Makino Y. Hydrogenation of Carbon Dioxide to C1-C7 Hydrocarbons via Methanol on Composite Catalysts. Appl. Catal. A, Gen. 1993, 94, 31–44. 10.1016/0926-860X(93)80043-P. [DOI] [Google Scholar]
- Fujiwara M.; Kieffer R.; Ando H.; Souma Y. Development of Composite Catalysts Made of Cu-Zn-Cr Oxide/Zeolite for the Hydrogenation of Carbon Dioxide. Appl. Catal. A, Gen. 1995, 121, 113–124. 10.1016/0926-860X(95)85014-7. [DOI] [Google Scholar]
- Jiao F.; Li J.; Pan X.; Xiao J.; Li H.; Ma H.; Wei M.; Pan Y.; Zhou Z.; Li M.; Miao S.; Li J.; Zhu Y.; Xiao D.; He T.; Yang J.; Qi F.; Fu Q.; Bao X. Selective Conversion of Syngas to Light Olefins. Science 2016, 351, 1065–1068. 10.1126/science.aaf1835. [DOI] [PubMed] [Google Scholar]
- Cheng K.; Gu B.; Liu X.; Kang J.; Zhang Q.; Wang Y. Direct and Highly Selective Conversion of Synthesis Gas into Lower Olefins: Design of a Bifunctional Catalyst Combining Methanol Synthesis and Carbon-Carbon Coupling. Angew. Chem., Int. Ed. 2016, 128, 4803–4806. 10.1002/ange.201601208. [DOI] [PubMed] [Google Scholar]
- Olsbye U.; Svelle S.; Bjørgen M.; Beato P.; Janssens T. V. W.; Joensen F.; Bordiga S.; Lillerud K. P. Conversion of Methanol to Hydrocarbons: How Zeolite Cavity and Pore Size Controls Product Selectivity. Angew. Chem., Int. Ed. 2012, 51, 5810–5831. 10.1002/anie.201103657. [DOI] [PubMed] [Google Scholar]
- Tian P.; Wei Y.; Ye M.; Liu Z. Methanol to Olefins (MTO): From Fundamentals to Commercialization. ACS Catal. 2015, 5, 1922–1938. 10.1021/acscatal.5b00007. [DOI] [Google Scholar]
- Yang M.; Fan D.; Wei Y.; Tian P.; Liu Z. Recent Progress in Methanol-to-Olefins (MTO) Catalysts. Adv. Mater. 2019, 31, 1902181–1902196. 10.1002/adma.201902181. [DOI] [PubMed] [Google Scholar]
- Zhang Q.; Yu J.; Corma A. Applications of Zeolites to C1 Chemistry: Recent Advances, Challenges, and Opportunities. Adv. Mater. 2020, 32, 2002927–2200958. 10.1002/adma.202002927. [DOI] [PubMed] [Google Scholar]
- Pan X.; Jiao F.; Miao D.; Bao X. Oxide-Zeolite-Based Composite Catalyst Concept That Enables Syngas Chemistry beyond Fischer–Tropsch Synthesis. Chem. Rev. 2021, 121 (121), 6588–6609. 10.1021/acs.chemrev.0c01012. [DOI] [PubMed] [Google Scholar]
- Zhong J.; Yang X.; Wu Z.; Liang B.; Huang Y.; Zhang T. State of the Art and Perspectives in Heterogeneous Catalysis of CO2 Hydrogenation to Methanol. Chem. Soc. Rev. 2020, 49, 1385–1413. 10.1039/C9CS00614A. [DOI] [PubMed] [Google Scholar]
- Sen R.; Goeppert A.; Surya Prakash G. K. Homogeneous Hydrogenation of CO2 and CO to Methanol: The Renaissance of Low-Temperature Catalysis in the Context of the Methanol Economy. Angew. Chem., Int. Ed. 2022, 61, e202207278 10.1002/anie.202207278. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yarulina I.; Chowdhury A. D.; Meirer F.; Weckhuysen B. M.; Gascon J. Recent Trends and Fundamental Insights in the Methanol-to-Hydrocarbons Process. Nat. Catal. 2018, 1, 398–411. 10.1038/s41929-018-0078-5. [DOI] [Google Scholar]
- Wang S.; Qin Z.; Dong M.; Wang J.; Fan W. Recent Progress on MTO Reaction Mechanisms and Regulation of Acid Site Distribution in the Zeolite Framework. Chem Catal. 2022, 2, 1657–1685. 10.1016/j.checat.2022.05.012. [DOI] [Google Scholar]
- Shi Z.; Bhan A. Metrics of Performance Relevant in Methanol-to-Hydrocarbons Catalysis. J. Catal. 2023, 421, 198–209. 10.1016/j.jcat.2023.03.022. [DOI] [Google Scholar]
- IZA . Database of Zeolite Structures. http://www.iza-structure.org/ (accessed June 13, 2023).
- Teketel S.; Erichsen M. W.; Bleken F. L.; Svelle S.; Petter K.; Olsbye U.. Shape Selectivity in Zeolite Catalysis. The Methanol to Hydrocarbons (MTH) Reaction. In Catalysis; Spivey J.; Han Y.-F.; Dooley K. M., Eds.; Royal Society of Chemistry: Cambridge, 2014; pp 179–217. [Google Scholar]
- Möller F.; Marschner F.; Supp E.; Boll W.; Cornelius G.. Process of Producing Gasoline from Synthesis Gas. US4263141A, 1981.
- Marker T. L.; Vora B. V.; Nilsen H. R.. Three Step Process for Producing Light Olefins from Crude Methane and/or Ethane. EP0933345B1, 2001.
- Ihm S.-K.; Park Y.-K.; Jeon J.-K.; Jeong K.-E.. Hybrid Catalyst for Hydrocarbon Synthesis via Hydrogenation of Carbon Dioxide. US6376562B1, 2002.
- Fujimoto K.; Asami F.; Asaoka S.; Li X.; Zhang Q.. Catalyst and Process for LPG Production. US7297825B2, 2007.
- Kibby C. L.; Burton A. W. Jr.; Haas A.; Holz S.. Process for Conversion of Synthesis Gas to Hydrocarbons Using a Zeolite-Methanol Catalyst System. US7943673B2, 2011.
- Kibby C. L.CO2 Injection into Synthesis Feed Gas to Reduce or Eliminate Net CO2 Production during Isosynthesis over ZnO-Cr2O3 plus ZSM-5 Hybrid Catalyst Combinations. US8513315B2, 2013.
- Fu Q.; Zhang X.; Luo Y.; Mu X.; Zong B.. Process for Preparing Methanol, Dimethyl Ether, and Low Carbon Olefins from Syngas. US8552074B2, 2014.
- Kortan A. M.; Kelly M. T.; Rose H. A.; Zubrin R. M.. Synthesis of Drop-in Liquid Fuels and Chemicals from Methanol, Ethanol or Syngas Using Mixed Catalysts. US9296665B2, 2016.
- Chojecki A.; Nieskens D. L. S.; Davidian T.; Groenendijk P. E.; Ruitenbeek M.; Fish B. B.; Tirtowidjojo M. M.; Meima G. R.. Conversion of Carbon Monoxide, Carbon Dioxide, or a Combination Thereof over Hybrid Catalyst. US9919981B2, 2018.
- Nieskens D. L. S.; Sandikci A. C.; Groenendijk P. E.; Malek A.. Process to Convert Synthesis Gas to Olefins over a Bifunctional Chromium Oxide/Zinc Oxide-SAPO-34 Catalyst. US10329209B2, 2019.
- Pan X.; Li J.; Jiao F.; Bao X.. Catalyst and Method of Preparing Light Olefin Directly from Synthesis Gas by One-Step Process. US10532961B2, 2020.
- Tan Y.; Yang G.; Zhang T.; Faen S.; Hongjuan X.; Junxuan P.. Catalyst for Directly Preparing Low-Carbon Olefin from Synthesis Gas, Preparation Method and Application. CN107803219B, 2021.
- Meng F.; Li X.; Li Z.; Zhang P.. High-Stability Catalyst for Direct Low Carbon Olefin Preparation from Synthesis Gas, Preparation Method and Application of High-Stability Catalyst for Direct Low Carbon Olefin Preparation from Synthesis Gas. CN109261199B, 2021.
- Ni Y.; Zhu W.; Liu Z.; Liu Y.. Method for Preparing Lower Olefin from Synthesis Gas. CN111039738B, 2021.
- Wang S.; Fan W.; Dong M.; Qin Z.; Wang J.. Gallium-Zirconium Composite Oxide-Molecular Sieve Catalyst as Well as Preparation Method and Application Thereof. CN113134386B, 2022.
- Wang S.; Fan W.; Qin Z.; Dong M.; Wang J.. Chromium-Based Metal Oxide-Molecular Sieve Catalyst and Preparation Method and Application Thereof. CN112588315B, 2022.
- Liu C.; Liu S.; Zhou J.; Wang Y.; Zhou H.; Su J.; Jiao W.; Zhang L.. Binder-Free Monolithic Catalyst, Preparation Method and Application Thereof. CN111111752B, 2022.
- Pan X.; Jiao F.; Bao X.; Li G.. Supported Catalyst and Method for Preparing Light Olefin Using Direct Conversion of Syngas. US11459279B2, 2022.
- Stöcker M. Methanol-to-Hydrocarbons: Catalytic Materials and Their Behavior. Microporous Mesoporous Mater. 1999, 29, 3–48. 10.1016/S1387-1811(98)00319-9. [DOI] [Google Scholar]
- Chang C. D.The Methanol-to-Hydrocarbons Reaction: A Mechanistic Perspective. In Shape Selective Catalysis: Chemical Synthesis and Hydrocarbon Processing; Song C.; Garces J. M.; Sugi Y., Eds.; American Chemical Society: WA, 2000; pp 96–114. [Google Scholar]
- Barger P.Methanol to Olefins (MTO) and Beyond. In Zeolites from Cleaner Technologies; Guisnet M.; Gilson J.-P., Eds.; Imperial College Press: London, 2002; pp 239–260. [Google Scholar]
- Haw J. F.; Marcus D. M.. Examples of Organic Reactions on Zeolites: Methanol to Hydrocarbon Catalysis. In Handbook of Zeolite Science and Technology; Auerbach S. M.; Carrado K. A.; Dutta P. K., Eds.; Marcel Dekker: New York, 2003; p 833. [Google Scholar]
- Ilias S.; Bhan A. Mechanism of the Catalytic Conversion of Methanol to Hydrocarbons. ACS Catal. 2013, 3, 18–31. 10.1021/cs3006583. [DOI] [Google Scholar]
- Dessau R. M. On the H-ZSM-5 Catalyzed Formation of Ethylene from Methanol or Higher Olefins. J. Catal. 1986, 99, 111–116. 10.1016/0021-9517(86)90204-6. [DOI] [Google Scholar]
- Dessau R. M.; Lapierre R. B. On the Mechanism of Methanol Conversion to Hydrocarbons over HZSM-5. J. Catal. 1982, 78, 136–141. 10.1016/0021-9517(82)90292-5. [DOI] [Google Scholar]
- Haw J. F.; Marcus D. M. Well-Defined (Supra)Molecular Structures in Zeolite Methanol-to-Olefin Catalysis. Top. Catal. 2005, 34, 41–48. 10.1007/s11244-005-3798-0. [DOI] [Google Scholar]
- Mole T.; Whiteside J. A.; Seddon D. Aromatic Co-Catalysis of Methanol Conversion over Zeolite Catalysts. J. Catal. 1983, 82, 261–266. 10.1016/0021-9517(83)90192-6. [DOI] [Google Scholar]
- Mole T.; Bett G.; Seddon D. Conversion of Methanol to Hydrocarbons over ZSM-5 Zeolite: An Examination of the Role of Aromatic Hydrocarbons Using 13Carbon- and Deuterium-Labeled Feeds. J. Catal. 1983, 84, 435–445. 10.1016/0021-9517(83)90014-3. [DOI] [Google Scholar]
- Arstad B.; Kolboe S. Methanol-to-Hydrocarbons Reaction over SAPO-34. Molecules Confined in the Catalyst Cavities at Short Time on Stream. Catal. Lett. 2001, 71, 209–212. 10.1023/A:1009034600533. [DOI] [Google Scholar]
- Arstad B.; Kolboe S. The Reactivity of Molecules Trapped within the SAPO-34 Cavities in the Methanol-to-Hydrocarbons Reaction. J. Am. Chem. Soc. 2001, 123, 8137–8138. 10.1021/ja010668t. [DOI] [PubMed] [Google Scholar]
- Song W.; Haw J. F.; Nicholas J. B.; Heneghan C. S. Methylbenzenes Are the Organic Reaction Centers for Methanol-to-Olefin Catalysis on HSAPO-34. J. Am. Chem. Soc. 2000, 122, 10726–10727. 10.1021/ja002195g. [DOI] [Google Scholar]
- Song W.; Fu H.; Haw J. F. Selective Synthesis of Methylnaphthalenes in HSAPO-34 Cages and Their Function as Reaction Centers in Methanol-to-Olefin Catalysis. J. Phys. Chem. B 2001, 105, 12839–12843. 10.1021/jp012400u. [DOI] [Google Scholar]
- Song W.; Fu H.; Haw J. F. Supramolecular Origins of Product Selectivity for Methanol-to-Olefin Catalysis on HSAPO-34. J. Am. Chem. Soc. 2001, 123, 4749–4754. 10.1021/ja0041167. [DOI] [PubMed] [Google Scholar]
- Fu H.; Song W.; Haw J. F. Polycyclic Aromatics Formation in HSAPO-34 during Methanol-to-Olefin Catalysis : Ex Situ Characterization after Cryogenic Grinding. Catal. Lett. 2001, 76, 89–94. 10.1023/A:1016719924539. [DOI] [Google Scholar]
- Chang C. D.; Silvestri A. J. The Conversion of Methanol and Other O-Compounds to Hydrocarbons over Zeolite Catalysts. J. Catal. 1977, 47, 249–259. 10.1016/0021-9517(77)90172-5. [DOI] [Google Scholar]
- Dahl I. M.; Kolboe S. On the Reaction Mechanism for Propene Formation in the MTO Reaction over SAPO-34. Catal. Lett. 1993, 20, 329–336. 10.1007/BF00769305. [DOI] [Google Scholar]
- Dahl I. M.; Kolboe S. On the Reaction Mechanism for Hydrocarbon Formation from Methanol over SAPO-34: Isotopic Labeling Studies of the Co-Reaction of Ethene and Methanol. J. Catal. 1996, 161, 304–309. 10.1006/jcat.1996.0188. [DOI] [Google Scholar]
- Magnoux P.; Roger P.; Canaff C.; Fouche V.; Gnep N. S.; Guisnet M. New Technique for the Characterization of Carbonaceous Compounds Responsible for Zeolite Deactivation. Stud. Surf. Sci. Catal. 1987, 34, 317–330. 10.1016/S0167-2991(09)60370-0. [DOI] [Google Scholar]
- Svelle S.; Joensen F.; Nerlov J.; Olsbye U.; Lillerud K. P.; Kolboe S.; Bjørgen M. Conversion of Methanol into Hydrocarbons over Zeolite H-ZSM-5: Ethene Formation Is Mechanistically Separated from the Formation of Higher Alkenes. J. Am. Chem. Soc. 2006, 128, 14770–14771. 10.1021/ja065810a. [DOI] [PubMed] [Google Scholar]
- Bjørgen M.; Svelle S.; Joensen F.; Nerlov J.; Kolboe S.; Bonino F.; Palumbo L.; Bordiga S.; Olsbye U. Conversion of Methanol to Hydrocarbons over Zeolite H-ZSM-5: On the Origin of the Olefinic Species. J. Catal. 2007, 249, 195–207. 10.1016/j.jcat.2007.04.006. [DOI] [Google Scholar]
- Tsang C. M.; Dai P. S. E.; Mertens F. P.; Petty R. H. Saturation of olefininc products by hydrogen-transfer during propylene conversion. Abstracts of papers of the American Chemical Society 1994, 208, 54. [Google Scholar]
- Westgård Erichsen M.; Svelle S.; Olsbye U. The Influence of Catalyst Acid Strength on the Methanol to Hydrocarbons (MTH) Reaction. Catal. Today 2013, 215, 216–223. 10.1016/j.cattod.2013.03.017. [DOI] [Google Scholar]
- Martínez-Espín J. S.; De Wispelaere K.; Janssens T. V. W.; Svelle S.; Lillerud K. P.; Beato P.; Van Speybroeck V.; Olsbye U. Hydrogen Transfer versus Methylation: On the Genesis of Aromatics Formation in the Methanol-To-Hydrocarbons Reaction over H-ZSM-5. ACS Catal. 2017, 7, 5773–5780. 10.1021/acscatal.7b01643. [DOI] [Google Scholar]
- Martinez-Espin J. S.; De Wispelaere K.; Westgård Erichsen M.; Svelle S.; Janssens T. V. W.; Van Speybroeck V.; Beato P.; Olsbye U. Benzene Co-Reaction with Methanol and Dimethyl Ether over Zeolite and Zeotype Catalysts: Evidence of Parallel Reaction Paths to Toluene and Diphenylmethane. J. Catal. 2017, 349, 136–148. 10.1016/j.jcat.2017.03.007. [DOI] [Google Scholar]
- Thiele E. W. Relation between Catalytic Activity and Size of Particle. Ind. Eng. Chem. 1939, 31, 916–920. 10.1021/ie50355a027. [DOI] [Google Scholar]
- Choi M.; Na K.; Kim J.; Sakamoto Y.; Terasaki O.; Ryoo R. Stable Single-Unit-Cell Nanosheets of Zeolite MFI as Active and Long-Lived Catalysts. Nature 2009, 461, 246–250. 10.1038/nature08288. [DOI] [PubMed] [Google Scholar]
- Zhang X.; Liu D.; Xu D.; Asahina S.; Cychosz K. A.; Agrawal K. V.; Al Wahedi Y.; Bhan A.; Al Hashimi S.; Terasaki O.; Thommes M.; Tsapatsis M. Synthesis of Self-Pillared Zeolite Nanosheets by Repetitive Branching. Science 2012, 336, 1684–1687. 10.1126/science.1221111. [DOI] [PubMed] [Google Scholar]
- Bleken B. T. L.; Wragg D. S.; Arstad B.; Gunnæs A. E.; Mouzon J.; Helveg S.; Lundegaard L. F.; Beato P.; Bordiga S.; Olsbye U.; Svelle S.; Lillerud K. P. Unit Cell Thick Nanosheets of Zeolite H-ZSM-5: Structure and Activity. Top. Catal. 2013, 56, 558–566. 10.1007/s11244-013-0010-9. [DOI] [Google Scholar]
- Khare R.; Bhan A. Mechanistic Studies of Methanol-to-Hydrocarbons Conversion on Diffusion-Free MFI Samples. J. Catal. 2015, 329, 218–228. 10.1016/j.jcat.2015.05.012. [DOI] [Google Scholar]
- Müller S.; Liu Y.; Kirchberger F. M.; Tonigold M.; Sanchez-Sanchez M.; Lercher J. A. Hydrogen Transfer Pathways during Zeolite Catalyzed Methanol Conversion to Hydrocarbons. J. Am. Chem. Soc. 2016, 138, 15994–16003. 10.1021/jacs.6b09605. [DOI] [PubMed] [Google Scholar]
- Wei Z.; Chen Y. Y.; Li J.; Guo W.; Wang S.; Dong M.; Qin Z.; Wang J.; Jiao H.; Fan W. Stability and Reactivity of Intermediates of Methanol Related Reactions and C-C Bond Formation over H-ZSM-5 Acidic Catalyst: A Computational Analysis. J. Phys. Chem. C 2016, 120, 6075–6087. 10.1021/acs.jpcc.6b00211. [DOI] [Google Scholar]
- Hwang A.; Kumar M.; Rimer J. D.; Bhan A. Implications of Methanol Disproportionation on Catalyst Lifetime for Methanol-to-Olefins Conversion by HSSZ-13. J. Catal. 2017, 346, 154–160. 10.1016/j.jcat.2016.12.003. [DOI] [Google Scholar]
- Martinez-Espin J. S.; Mortén M.; Janssens T. V. W.; Svelle S.; Beato P.; Olsbye U. New Insights into Catalyst Deactivation and Product Distribution of Zeolites in the Methanol-to-Hydrocarbons (MTH) Reaction with Methanol and Dimethyl Ether Feeds. Catal. Sci. Technol. 2017, 7, 2700–2716. 10.1039/C7CY00129K. [DOI] [Google Scholar]
- Wen W.; Yu S.; Zhou C.; Ma H.; Zhou Z.; Cao C.; Yang J.; Xu M.; Qi F.; Zhang G.; Pan Y. Formation and Fate of Formaldehyde in Methanol-to-Hydrocarbon Reaction: In Situ Synchrotron Radiation Photoionization Mass Spectrometry Study. Angew. Chem., Int. Ed. 2020, 59, 4873–4878. 10.1002/anie.201914953. [DOI] [PubMed] [Google Scholar]
- Arora S. S.; Bhan A. The Critical Role of Methanol Pressure in Controlling Its Transfer Dehydrogenation and the Corresponding Effect on Propylene-to-Ethylene Ratio during Methanol-to-Hydrocarbons Catalysis on H-ZSM-5. J. Catal. 2017, 356, 300–306. 10.1016/j.jcat.2017.10.014. [DOI] [Google Scholar]
- Liu Y.; Kirchberger F. M.; Müller S.; Eder M.; Tonigold M.; Sanchez-Sanchez M.; Lercher J. A. Critical Role of Formaldehyde during Methanol Conversion to Hydrocarbons. Nat. Commun. 2019, 10, 1–9. 10.1038/s41467-019-09449-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hwang A.; Bhan A. Bifunctional Strategy Coupling Y2O3-Catalyzed Alkanal Decomposition with Methanol-to-Olefins Catalysis for Enhanced Lifetime. ACS Catal. 2017, 7, 4417–4422. 10.1021/acscatal.7b00894. [DOI] [Google Scholar]
- Liu Y.; Müller S.; Berger D.; Jelic J.; Reuter K.; Tonigold M.; Sanchez-Sanchez M.; Lercher J. A. Formation Mechanism of the First Carbon-Carbon Bond and the First Olefin in the Methanol Conversion into Hydrocarbons. Angew. Chem., Int. Ed. 2016, 128, 5817–5820. 10.1002/ange.201511678. [DOI] [PubMed] [Google Scholar]
- Chowdhury A. D.; Houben K.; Whiting G. T.; Mokhtar M.; Asiri A. M.; Al-thabaiti S. A.; Basahel S. N.; Baldus M.; Weckhuysen B. M. Initial Carbon-Carbon Bond Formation during the Early Stages of the Methanol-to-Olefin Process Proven by Zeolite-Trapped Acetate and Methyl Acetate. Angew. Chem., Int. Ed. 2016, 128, 16072–16077. 10.1002/ange.201608643. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chowdhury A. D.; Paioni A. L.; Houben K.; Whiting G. T.; Baldus M.; Weckhuysen B. M. Bridging the Gap between the Direct and Hydrocarbon Pool Mechanisms of the Methanol-to-Hydrocarbons Process. Angew. Chem., Int. Ed. 2018, 57, 8095–8099. 10.1002/anie.201803279. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Plessow P. N.; Studt F. Unraveling the Mechanism of the Initiation Reaction of the Methanol to Olefins Process Using Ab Initio and DFT Calculations. ACS Catal. 2017, 7, 7987–7994. 10.1021/acscatal.7b03114. [DOI] [Google Scholar]
- Plessow P. N.; Studt F. Theoretical Insights into the Effect of the Framework on the Initiation Mechanism of the MTO Process. Catal. Lett. 2018, 148, 1246–1253. 10.1007/s10562-018-2330-7. [DOI] [Google Scholar]
- Plessow P. N.; Smith A.; Tischer S.; Studt F. Identification of the Reaction Sequence of the MTO Initiation Mechanism Using Ab Initio-Based Kinetics. J. Am. Chem. Soc. 2019, 141, 5908–5915. 10.1021/jacs.9b00585. [DOI] [PubMed] [Google Scholar]
- Rojo-gama D.; Etemadi S.; Kirby E.; Lillerud K. P.; Beato P.; Svelle S.; Olsbye U. Time- and Space-Resolved Study of the Methanol to Hydrocarbons (MTH) Reaction - Influence of Zeolite Topology on Axial Deactivation Patterns. Faraday Discuss. 2017, 197, 421–446. 10.1039/C6FD00187D. [DOI] [PubMed] [Google Scholar]
- Airi A.; Damin A.; Xie J.; Olsbye U.; Bordiga S. Catalyst Sites and Active Species in the Early Stages of MTO Conversion over Cobalt AlPO-18 Followed by IR Spectroscopy. Catal. Sci. Technol. 2022, 12, 2775–2792. 10.1039/D2CY00303A. [DOI] [Google Scholar]
- De Wispelaere K.; Wondergem C. S.; Ensing B.; Hemelsoet K.; Meijer E. J.; Weckhuysen B. M.; Van Speybroeck V.; Ruiz-Martínez J. Insight into the Effect of Water on the Methanol-to-Olefins Conversion in H-SAPO-34 from Molecular Simulations and in Situ Microspectroscopy. ACS Catal. 2016, 6, 1991–2002. 10.1021/acscatal.5b02139. [DOI] [Google Scholar]
- Bordiga S.; Regli L.; Lamberti C.; Zecchina A.; Bjørgen M.; Lillerud K. P. F. FTIR Adsorption Studies of H2O and CH3OH in the Isostructural H-SSZ-13 and H-SAPO-34: Formation of H-Bonded Adducts and Protonated Clusters. J. Phys. Chem. B 2005, 109, 7724–7732. 10.1021/jp044324b. [DOI] [PubMed] [Google Scholar]
- Grifoni E.; Piccini G. M.; Lercher J. A.; Glezakou V. A.; Rousseau R.; Parrinello M. Confinement Effects and Acid Strength in Zeolites. Nat. Commun. 2021, 12, 1–9. 10.1038/s41467-021-22936-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hack J. H.; Dombrowski J. P.; Ma X.; Chen Y.; Lewis N. H. C.; Carpenter W. B.; Li C.; Voth G. A.; Kung H. H.; Tokmakoff A. Structural Characterization of Protonated Water Clusters Confined in HZSM-5 Zeolites. J. Am. Chem. Soc. 2021, 143, 10203–10213. 10.1021/jacs.1c03205. [DOI] [PubMed] [Google Scholar]
- Masuda T.; Fujikata Y.; Mukai S. R.; Hashimoto K. Changes in Catalytic Activity of MFI-Type Zeolites Caused by Dealumination in a Steam Atmosphere. Appl. Catal. A Gen. 1998, 172, 73–83. 10.1016/S0926-860X(98)00120-3. [DOI] [Google Scholar]
- Nielsen M.; Brogaard R. Y.; Falsig H.; Beato P.; Swang O.; Svelle S. Kinetics of Zeolite Dealumination: Insights from H-SSZ-13. ACS Catal. 2015, 5, 7131–7139. 10.1021/acscatal.5b01496. [DOI] [Google Scholar]
- Nielsen M.; Hafreager A.; Brogaard R. Y.; De Wispelaere K.; Falsig H.; Beato P.; Van Speybroeck V.; Svelle S. Collective Action of Water Molecules in Zeolite Dealumination. Catal. Sci. Technol. 2019, 9, 3721–3725. 10.1039/C9CY00624A. [DOI] [Google Scholar]
- Yang L.; Wang C.; Zhang L.; Dai W.; Chu Y.; Xu J.; Wu G.; Gao M.; Liu W.; Xu Z.; Wang P.; Guan N.; Dyballa M.; Ye M.; Deng F.; Fan W.; Li L. Stabilizing the Framework of SAPO-34 Zeolite toward Long-Term Methanol-to-Olefins Conversion. Nat. Commun. 2021, 12, 1–11. 10.1038/s41467-021-24403-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bollini P.; Chen T. T.; Neurock M.; Bhan A. Mechanistic Role of Water in HSSZ-13 Catalyzed Methanol-to-Olefins Conversion. Catal. Sci. Technol. 2019, 9, 4374–4383. 10.1039/C9CY01015G. [DOI] [Google Scholar]
- Valecillos J.; Hita I.; Sastre E.; Aguayo A. T.; Castaño P. Implications of Co-Feeding Water on the Growth Mechanisms of Retained Species on a SAPO-18 Catalyst during the Methanol-to-Olefins Reaction. ChemCatChem 2021, 13, 3140–3154. 10.1002/cctc.202100124. [DOI] [Google Scholar]
- Arora S. S.; Shi Z.; Bhan A. Mechanistic Basis for Effects of High-Pressure H2 Cofeeds on Methanol-to-Hydrocarbons Catalysis over Zeolites. ACS Catal. 2019, 9, 6407–6414. 10.1021/acscatal.9b00969. [DOI] [Google Scholar]
- Arora S. S.; Nieskens D. L. S.; Malek A.; Bhan A. Lifetime Improvement in Methanol-to-Olefins Catalysis over Chabazite Materials by High-Pressure H2 Co-Feeds. Nat. Catal. 2018, 1, 666–672. 10.1038/s41929-018-0125-2. [DOI] [Google Scholar]
- Nieskens D. L. S.; Lunn J. D.; Malek A. Understanding the Enhanced Lifetime of SAPO-34 in a Direct Syngas-to-Hydrocarbons Process. ACS Catal. 2019, 9, 691–700. 10.1021/acscatal.8b03465. [DOI] [Google Scholar]
- Deluca M.; Janes C.; Hibbitts D. Contrasting Arene, Alkene, Diene, and Formaldehyde Hydrogenation in H-ZSM-5, H-SSZ-13, and H-SAPO-34 Frameworks during MTO. ACS Catal. 2020, 10, 4593–4607. 10.1021/acscatal.9b04529. [DOI] [Google Scholar]
- Zhao X.; Li J.; Tian P.; Wang L.; Li X.; Lin S.; Guo X.; Liu Z. Achieving a Superlong Lifetime in the Zeolite-Catalyzed MTO Reaction under High Pressure: Synergistic Effect of Hydrogen and Water. ACS Catal. 2019, 9, 3017–3025. 10.1021/acscatal.8b04402. [DOI] [Google Scholar]
- Xie J.; Firth D. S.; Cordero-lanzac T.; Airi A.; Negri C.; Øien-ødegaard S.; Lillerud K. P.; Bordiga S.; Olsbye U. MAPO-18 Catalysts for the Methanol to Olefins Process: Influence of Catalyst Acidity in a High-Pressure Syngas (CO and H2) Environment. ACS Catal. 2022, 12, 1520–1531. 10.1021/acscatal.1c04694. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Paunovic V.; Sushkevich V.; Rzepka P.; Artiglia L.; Hauert R.; Sik Lee S.; van Bokhoven J. A. Van. Reactivation of Catalysts for Methanol-to-Hydrocarbons Conversion with Hydrogen. J. Catal. 2022, 407, 54–64. 10.1016/j.jcat.2022.01.018. [DOI] [Google Scholar]
- Chen Z.; Ni Y.; Zhi Y.; Wen F.; Zhou Z.; Wei Y.; Zhu W.; Liu Z. Coupling of Methanol and Carbon Monoxide over H-ZSM-5 to Form Aromatics. Angew. Chem., Int. Ed. 2018, 130, 12729–12733. 10.1002/ange.201807814. [DOI] [PubMed] [Google Scholar]
- Shi Z.; Bhan A. Methanol-to-Olefins Catalysis on Window-Cage Type Zeolites/Zeotypes with Syngas Co-Feeds: Understanding Syngas-to-Olefins Chemistry. J. Catal. 2022, 413, 913–922. 10.1016/j.jcat.2022.07.035. [DOI] [Google Scholar]
- Sehested J. Industrial and Scientific Directions of Methanol Catalyst Development. J. Catal. 2019, 371, 368–375. 10.1016/j.jcat.2019.02.002. [DOI] [Google Scholar]
- Waugh K. C. Methanol Synthesis. Catal. Today 1992, 15, 51–75. 10.1016/0920-5861(92)80122-4. [DOI] [Google Scholar]
- Bozzano G.; Manenti F. Efficient Methanol Synthesis : Perspectives, Technologies and Optimization Strategies. Prog. Energy Combust. Sci. 2016, 56, 71–105. 10.1016/j.pecs.2016.06.001. [DOI] [Google Scholar]
- Liang B.; Ma J.; Su X.; Yang C.; Duan H.; Zhou H.; Deng S.; Li L.; Huang Y. Investigation on Deactivation of Cu/ZnO/Al2O3 Catalyst for CO2 Hydrogenation to Methanol. Ind. Eng. Chem. Res. 2019, 58, 9030–9037. 10.1021/acs.iecr.9b01546. [DOI] [Google Scholar]
- 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. 10.1021/ie990078s. [DOI] [Google Scholar]
- 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. 10.1016/S0926-860X(01)00650-0. [DOI] [Google Scholar]
- Prašnikar A.; Pavlišič A.; Ruiz-Zepeda F.; Kovač J.; Likozar B. Mechanisms of Copper-Based Catalyst Deactivation during CO2 Reduction to Methanol. Ind. Eng. Chem. Res. 2019, 58, 13021–13029. 10.1021/acs.iecr.9b01898. [DOI] [Google Scholar]
- Schulz H. “Coking” of Zeolites during Methanol Conversion: Basic Reactions of the MTO-, MTP- and MTG Processes. Catal. Today 2010, 154, 183–194. 10.1016/j.cattod.2010.05.012. [DOI] [Google Scholar]
- Bleken F.; Bjørgen M.; Palumbo L.; Bordiga S.; Svelle S.; Lillerud K. P.; Olsbye U. The Effect of Acid Strength on the Conversion of Methanol to Olefins over Acidic Microporous Catalysts with the CHA Topology. Top. Catal. 2009, 52, 218–228. 10.1007/s11244-008-9158-0. [DOI] [Google Scholar]
- Olsbye U.; Svelle S.; Lillerud K. P.; Wei Z. H.; Chen Y. Y.; Li J. F.; Wang J. G.; Fan W. B. The Formation and Degradation of Active Species during Methanol Conversion over Protonated Zeotype Catalysts. Chem. Soc. Rev. 2015, 44, 7155–7176. 10.1039/C5CS00304K. [DOI] [PubMed] [Google Scholar]
- Borodina E.; Sharbini Harun Kamaluddin H.; Meirer F.; Mokhtar M.; Asiri A. M.; Al-Thabaiti S. A.; Basahel S. N.; Ruiz-Martinez J.; Weckhuysen B. M. Influence of the Reaction Temperature on the Nature of the Active and Deactivating Species during Methanol-to-Olefins Conversion over H-SAPO-34. ACS Catal. 2017, 7, 5268–5281. 10.1021/acscatal.7b01497. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Amoo C. C.; Xing C.; Tsubaki N.; Sun J. Tandem Reactions over Zeolite-Based Catalysts in Syngas Conversion. ACS Cent. Sci. 2022, 8, 1047–1062. 10.1021/acscentsci.2c00434. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang M.; Kang J.; Xiong X.; Zhang F.; Cheng K.; Zhang Q.; Wang Y. Effect of Zeolite Topology on the Hydrocarbon Distribution over Bifunctional ZnAlO/SAPO Catalysts in Syngas Conversion. Catal. Today 2021, 371, 85–92. 10.1016/j.cattod.2020.07.076. [DOI] [Google Scholar]
- Pinilla-Herrero I.; Olsbye U.; Márquez-Álvarez C.; Sastre E. Effect of Framework Topology of SAPO Catalysts on Selectivity and Deactivation Profile in the Methanol-to-Olefins Reaction. J. Catal. 2017, 352, 191–207. 10.1016/j.jcat.2017.05.008. [DOI] [Google Scholar]
- Alshafei F. H.; Park Y.; Zones S. I.; Davis M. E. Methanol-to-Olefins Catalysis on ERI-Type Molecular Sieves: Towards Enhancing Ethylene Selectivity. J. Catal. 2021, 404, 620–633. 10.1016/j.jcat.2021.10.025. [DOI] [Google Scholar]
- Sun Q.; Wang N.; Bai R.; Chen X.; Yu J. Seeding Induced Nano-Sized Hierarchical SAPO-34 Zeolites : Cost-Effective Synthesis and Superior MTO Performance. J. Mater. Chem. A 2016, 4, 14978–14982. 10.1039/C6TA06613E. [DOI] [Google Scholar]
- Martínez-Franco R.; Li Z.; Martínez-Triguero J.; Moliner M.; Corma A. Improving the Catalytic Performance of SAPO-18 for the Methanol-to-Olefins (MTO) Reaction by Controlling the Si Distribution and Crystal Size. Catal. Sci. Technol. 2016, 6, 2796–2806. 10.1039/C5CY02298C. [DOI] [Google Scholar]
- Kang J. H.; Alshafei F. H.; Zones S. I.; Davis M. E. Cage-Defining Ring: A Molecular Sieve Structural Indicator for Light Olefin Product Distribution from the Methanol-to-Olefins Reaction. ACS Catal. 2019, 9, 6012–6019. 10.1021/acscatal.9b00746. [DOI] [Google Scholar]
- 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. 10.1016/S1387-1811(98)00331-X. [DOI] [Google Scholar]
- Wang S.; Zhang L.; Wang P.; Jiao W.; Qin Z.; Dong M.; Wang J.; Olsbye U.; Fan W. Highly Selective Hydrogenation of CO2 to Propane over GaZrOx/H-SSZ-13 Composite. Nat. Catal. 2022, 5, 1038–1050. 10.1038/s41929-022-00871-7. [DOI] [Google Scholar]
- Bauer F.; Ernst H.; Geidel E.; Schödel R. Reactivation of Coked H-ZSM-5 by Treatment with Hydrogen and Alkanes. J. Catal. 1996, 164, 146–151. 10.1006/jcat.1996.0370. [DOI] [Google Scholar]
- 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, 1–11. 10.1038/s41467-020-20193-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guo S.; Wang S.; Zhang W.; Wang H.; Zhang Q.; Qin Z.; Dong M.; Wang J.; Fan W. Catalytic Performance of Various Zinc-Based Binary Metal Oxides/H-RUB-13 for Hydrogenation of CO2. Ind. Eng. Chem. Res. 2022, 61, 10409–10418. 10.1021/acs.iecr.2c00160. [DOI] [Google Scholar]
- Liu X.; Zhou W.; Yang Y.; Cheng K.; Kang J.; Zhang L.; Zhang G.; Min X.; Zhang Q.; Wang Y. Design of Efficient Bifunctional Catalysts for Direct Conversion of Syngas into Lower Olefins via Methanol/Dimethyl Ether Intermediates. Chem. Sci. 2018, 9, 4708–4718. 10.1039/C8SC01597J. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu X.; Wang M.; Yin H.; Hu J.; Cheng K.; Kang J.; Zhang Q.; Wang Y. Tandem Catalysis for Hydrogenation of CO and CO2 to Lower Olefins with Bifunctional Catalysts Composed of Spinel Oxide and SAPO-34. ACS Catal. 2020, 10, 8303–8314. 10.1021/acscatal.0c01579. [DOI] [Google Scholar]
- Li G.; Jiao F.; Pan X.; Li N.; Miao D.; Li L.; Bao X. Role of SAPO-18 Acidity in Direct Syngas Conversion to Light Olefins. ACS Catal. 2020, 10, 12370–12375. 10.1021/acscatal.0c03257. [DOI] [Google Scholar]
- Wu L.; Hensen E. J. M. Comparison of Mesoporous SSZ-13 and SAPO-34 Zeolite Catalysts for the Methanol-to-Olefins Reaction. Catal. Today 2014, 235, 160–168. 10.1016/j.cattod.2014.02.057. [DOI] [Google Scholar]
- Shi Z.; Neurock M.; Bhan A. Methanol-to-Olefins Catalysis on HSSZ-13 and HSAPO-34 and Its Relationship to Acid Strength. ACS Catal. 2021, 11, 1222–1232. 10.1021/acscatal.0c04011. [DOI] [Google Scholar]
- Ramirez A.; Ticali P.; Salusso D.; Cordero-lanzac T.; Ould-chikh S.; Ahoba-sam C.; Bugaev A. L.; Borfecchia E.; Morandi S.; Signorile M.; Bordiga S.; Gascon J.; Olsbye U. Multifunctional Catalyst Combination for the Direct Conversion of CO2 to Propane. JACS Au 2021, 1, 1719–1732. 10.1021/jacsau.1c00302. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cordero-Lanzac T.; Ramirez A.; Cruz-fernandez M.; Zander H.-J.; Joensen F.; Woolass S.; Meiswinkel A.; Styring P.; Gascon J.; Olsbye U. A CO2 Valorization Plant to Produce Light Hydrocarbons : Kinetic Model, Process Design and Life Cycle Assessment. J. CO2 Util. 2023, 67, 102337–102351. 10.1016/j.jcou.2022.102337. [DOI] [Google Scholar]
- Ji Y.; Gao P.; Zhao Z.; Xiao D.; Han Q.; Chen H.; Gong K.; Chen K.; Han X.; Bao X.; Hou G. Oxygenate-Based Routes Regulate Syngas Conversion over Oxide-Zeolite Bifunctional Catalysts. Nat. Catal. 2022, 5, 594–604. 10.1038/s41929-022-00806-2. [DOI] [Google Scholar]
- Chang C. D. The New Zealand Gas-to-Gasoline Plant: An Engineering Tour de Force. Catal. Today 1992, 13, 103–111. 10.1016/0920-5861(92)80190-X. [DOI] [Google Scholar]
- Topp-Jørgensen J. TOPS0E Integrated Gasoline Synthesis- The TIGAS Process. Stud. Surf. Sci. Catal. 1988, 36, 293–305. 10.1016/S0167-2991(09)60523-1. [DOI] [Google Scholar]
- Keil F. J. Methanol-to-Hydrocarbons : Process Technology. Microporous Mesoporous Mater. 1999, 29, 49–66. 10.1016/S1387-1811(98)00320-5. [DOI] [Google Scholar]
- Koempel H.; Liebner W. Lurgi’s Methanol To Propylene (MTP®) Report on a Successful Commercialisation. Stud. Surf. Sci. Catal. 2007, 167, 261–267. 10.1016/S0167-2991(07)80142-X. [DOI] [Google Scholar]
- Hutchings G. J.; Hunter R. Hydrocarbon Formation from Methanol and Dimethyl Ether: A Review of the Experimental Observations Concerning the Mechanism of Formation of the Primary Products. Catal. Today 1990, 6, 279–306. 10.1016/0920-5861(90)85006-A. [DOI] [Google Scholar]
- Song W.; Marcus D. M.; Fu H.; Ehresmann J. O.; Haw J. F. An Oft-Studied Reaction That May Never Have Been: Direct Catalytic Conversion of Methanol or Dimethyl Ether to Hydrocarbons on the Solid Acids HZSM-5 or HSAPO-34. J. Am. Chem. Soc. 2002, 124, 3844–3845. 10.1021/ja016499u. [DOI] [PubMed] [Google Scholar]
- Svelle S.; Kolboe S.; Swang O.; Olsbye U. Methylation of Alkenes and Methylbenzenes by Dimethyl Ether or Methanol on Acidic Zeolites. J. Phys. Chem. B 2005, 109, 12874–12878. 10.1021/jp051125z. [DOI] [PubMed] [Google Scholar]
- Hirota Y.; Murata K.; Miyamoto M.; Egashira Y.; Nishiyama N. Light Olefins Synthesis from Methanol and Dimethylether over SAPO-34 Nanocrystals. Catal. Lett. 2010, 140, 22–26. 10.1007/s10562-010-0421-1. [DOI] [Google Scholar]
- Li Y.; Zhang M.; Wang D.; Wei F.; Wang Y. Differences in the Methanol-to-Olefins Reaction Catalyzed by SAPO-34 with Dimethyl Ether as Reactant. J. Catal. 2014, 311, 281–287. 10.1016/j.jcat.2013.12.004. [DOI] [Google Scholar]
- Valecillos J.; Manzano H.; Aguayo A. T.; Bilbao J.; Castaño P. Kinetic and Deactivation Differences Among Methanol, Dimethyl Ether and Chloromethane as Stock for Hydrocarbons. ChemCatChem 2019, 11, 5444–5456. 10.1002/cctc.201901204. [DOI] [Google Scholar]
- Deluca M.; Kravchenko P.; Hoffman A.; Hibbitts D. Mechanism and Kinetics of Methylating C6-C12 Methylbenzenes with Methanol and Dimethyl Ether in H-MFI Zeolites. ACS Catal. 2019, 9, 6444–6460. 10.1021/acscatal.9b00650. [DOI] [Google Scholar]
- Pérez-Uriarte P.; Ateka A.; Aguayo A. T.; Bilbao J. Comparison of HZSM-5 Zeolite and SAPO (−18 and −34) Based Catalysts for the Production of Light Olefins from DME. Catal. Lett. 2016, 146, 1892–1902. 10.1007/s10562-016-1829-z. [DOI] [Google Scholar]
- Pérez-Uriarte P.; Ateka A.; Gamero M.; Aguayo A. T.; Bilbao J. Effect of the Operating Conditions in the Transformation of DME to Olefins over a HZSM-5 Zeolite Catalyst. Ind. Eng. Chem. Res. 2016, 55, 6569–6578. 10.1021/acs.iecr.6b00627. [DOI] [Google Scholar]
- Ortega C.; Kolb G. DME-to-Hydrocarbon over an MFI Zeolite: Product Selectivity Controlled by Oxygenates under the Kinetic Regime. Ind. Eng. Chem. Res. 2019, 58, 22893–22904. 10.1021/acs.iecr.9b04735. [DOI] [Google Scholar]
- Hill I. M.; Hashimi S. Al; Bhan A. Kinetics and Mechanism of Olefin Methylation Reactions on Zeolites. J. Catal. 2012, 285, 115–123. 10.1016/j.jcat.2011.09.018. [DOI] [Google Scholar]
- Hill I. M.; Ng Y. S.; Bhan A. Kinetics of Butene Isomer Methylation with Dimethyl Ether over Zeolite Catalysts. ACS Catal. 2012, 2, 1742–1748. 10.1021/cs300317p. [DOI] [Google Scholar]
- Hill I.; Malek A.; Bhan A. Kinetics and Mechanism of Benzene, Toluene, and Xylene Methylation over H-MFI. ACS Catal. 2013, 3, 1992–2001. 10.1021/cs400377b. [DOI] [Google Scholar]
- Ilias S.; Bhan A. Tuning the Selectivity of Methanol-to-Hydrocarbons Conversion on H-ZSM-5 by Co-Processing Olefin or Aromatic Compounds. J. Catal. 2012, 290, 186–192. 10.1016/j.jcat.2012.03.016. [DOI] [Google Scholar]
- Sun J.; Yang G.; Yoneyama Y.; Tsubaki N. Catalysis Chemistry of Dimethyl Ether Synthesis. ACS Catal. 2014, 4, 3346–3356. 10.1021/cs500967j. [DOI] [Google Scholar]
- Akarmazyan S. S.; Panagiotopoulou P.; Kambolis A.; Papadopoulou C.; Kondarides D. I. Methanol Dehydration to Dimethylether over Al2O3 Catalysts. Appl. Catal. B Environ. 2014, 145, 136–148. 10.1016/j.apcatb.2012.11.043. [DOI] [Google Scholar]
- Xu M.; Lunsford J. H.; Goodman D. W.; Bhattacharyya A. Synthesis of Dimethyl Ether (DME) from Methanol over Solid-Acid Catalysts. Appl. Catal. A Gen. 1997, 149, 289–301. 10.1016/S0926-860X(96)00275-X. [DOI] [Google Scholar]
- Yaripour F.; Baghaei F.; Schmidt I.; Perregaard J. Catalytic Dehydration of Methanol to Dimethyl Ether (DME) over Solid-Acid Catalysts. Catal. Commun. 2005, 6, 147–152. 10.1016/j.catcom.2004.11.012. [DOI] [Google Scholar]
- Vishwanathan V.; Jun K. W.; Kim J. W.; Roh H. S. Vapour Phase Dehydration of Crude Methanol to Dimethyl Ether over Na-Modified H-ZSM-5 Catalysts. Appl. Catal. A Gen. 2004, 276, 251–255. 10.1016/j.apcata.2004.08.011. [DOI] [Google Scholar]
- Khandan N.; Kazemeini M.; Aghaziarati M. Determining an Optimum Catalyst for Liquid-Phase Dehydration of Methanol to Dimethyl Ether. Appl. Catal. A Gen. 2008, 349, 6–12. 10.1016/j.apcata.2008.07.029. [DOI] [Google Scholar]
- Catizzone E.; Aloise A.; Migliori M.; Giordano G. Dimethyl Ether Synthesis via Methanol Dehydration: Effect of Zeolite Structure. Appl. Catal. A Gen. 2015, 502, 215–220. 10.1016/j.apcata.2015.06.017. [DOI] [Google Scholar]
- Migliori M.; Catizzone E.; Aloise A.; Bonura G.; Gómez-Hortigüela L.; Frusteri L.; Cannilla C.; Frusteri F.; Giordano G. New Insights about Coke Deposition in Methanol-to-DME Reaction over MOR-, MFI- and FER-Type Zeolites. J. Ind. Eng. Chem. 2018, 68, 196–208. 10.1016/j.jiec.2018.07.046. [DOI] [Google Scholar]
- An H.; Zhang F.; Guan Z.; Liu X.; Fan F.; Li C. Investigating the Coke Formation Mechanism of H-ZSM-5 during Methanol Dehydration Using Operando UV-Raman Spectroscopy. ACS Catal. 2018, 8, 9207–9215. 10.1021/acscatal.8b00928. [DOI] [Google Scholar]
- Blaszkowski S. R.; Van Santen R. A. The Mechanism of Dimethyl Ether Formation from Methanol Catalyzed by Zeolitic Protons. J. Am. Chem. Soc. 1996, 118, 5152–5153. 10.1021/ja954323k. [DOI] [Google Scholar]
- Carr R. T.; Neurock M.; Iglesia E. Catalytic Consequences of Acid Strength in the Conversion of Methanol to Dimethyl Ether. J. Catal. 2011, 278, 78–93. 10.1016/j.jcat.2010.11.017. [DOI] [Google Scholar]
- Jones A. J.; Iglesia E. Kinetic, Spectroscopic, and Theoretical Assessment of Associative and Dissociative Methanol Dehydration Routes in Zeolites. Angew. Chem., Int. Ed. 2014, 53, 12177–12181. 10.1002/anie.201406823. [DOI] [PubMed] [Google Scholar]
- Ghorbanpour A.; Rimer J. D.; Grabow L. C. Computational Assessment of the Dominant Factors Governing the Mechanism of Methanol Dehydration over H-ZSM-5 with Heterogeneous Aluminum Distribution. ACS Catal. 2016, 6, 2287–2298. 10.1021/acscatal.5b02367. [DOI] [Google Scholar]
- Di Iorio J. R.; Hoffman A. J.; Nimlos C. T.; Nystrom S.; Hibbitts D.; Gounder R. Mechanistic Origins of the High-Pressure Inhibition of Methanol Dehydration Rates in Small-Pore Acidic Zeolites. J. Catal. 2019, 380, 161–177. 10.1016/j.jcat.2019.10.012. [DOI] [Google Scholar]
- Moses P. G.; Nørskov J. K. Methanol to Dimethyl Ether over ZSM-22: A Periodic Density Functional Theory Study. ACS Catal. 2013, 3, 735–745. 10.1021/cs300722w. [DOI] [Google Scholar]
- Jones A. J.; Zones S. I.; Iglesia E. Implications of Transition State Con Fi Nement within Small Voids for Acid Catalysis. J. Phys. Chem. C 2014, 118, 17787–17800. 10.1021/jp5050095. [DOI] [Google Scholar]
- Di Iorio J. R.; Nimlos C. T.; Gounder R. Introducing Catalytic Diversity into Single-Site Chabazite Zeolites of Fixed Composition via Synthetic Control of Active Site Proximity. ACS Catal. 2017, 7, 6663–6674. 10.1021/acscatal.7b01273. [DOI] [Google Scholar]
- Wittich K.; Krämer M.; Bottke N.; Schunk S. A. Catalytic Dry Reforming of Methane: Insights from Model Systems. ChemCatChem 2020, 12, 2130–2147. 10.1002/cctc.201902142. [DOI] [Google Scholar]
- Shen W. J.; Jun K. W.; Choi H. S.; Lee K. W. Thermodynamic Investigation of Methanol and Dimethyl Ether Synthesis from CO2 Hydrogenation. Korean J. Chem. Eng. 2000, 17, 210–216. 10.1007/BF02707145. [DOI] [Google Scholar]
- Jia G.; Tan Y.; Han Y. A Comparative Study on the Thermodynamics of Dimethyl Ether Synthesis from CO Hydrogenation and CO 2 Hydrogenation. Ind. Eng. Chem. Res. 2006, 45, 1152–1159. 10.1021/ie050499b. [DOI] [Google Scholar]
- Álvarez A.; Bansode A.; Urakawa A.; Bavykina A. V.; Wezendonk T. A.; Makkee M.; Gascon J.; Kapteijn F. Challenges in the Greener Production of Formates/Formic Acid, Methanol, and DME by Heterogeneously Catalyzed CO2 Hydrogenation Processes. Chem. Rev. 2017, 117, 9804–9838. 10.1021/acs.chemrev.6b00816. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Catizzone E.; Bonura G.; Migliori M.; Frusteri F.; Giordano G. CO2 Recycling to Dimethyl Ether: State-of-the-Art and Perspectives. Molecules 2018, 23, 1–28. 10.3390/molecules23010031. [DOI] [PMC free article] [PubMed] [Google Scholar]
- García-Trenco A.; Vidal-Moya A.; Martínez A. Study of the Interaction between Components in Hybrid CuZnAl/HZSM-5 Catalysts and Its Impact in the Syngas-to-DME Reaction. Catal. Today 2012, 179, 43–51. 10.1016/j.cattod.2011.06.034. [DOI] [Google Scholar]
- Mao D.; Yang W.; Xia J.; Zhang B.; Song Q.; Chen Q. Highly Effective Hybrid Catalyst for the Direct Synthesis of Dimethyl Ether from Syngas with Magnesium Oxide-Modified HZSM-5 as a Dehydration Component. J. Catal. 2005, 230, 140–149. 10.1016/j.jcat.2004.12.007. [DOI] [Google Scholar]
- Cai M.; Palčić A.; Subramanian V.; Moldovan S.; Ersen O.; Valtchev V.; Ordomsky V. V.; Khodakov A. Y. Direct Dimethyl Ether Synthesis from Syngas on Copper-Zeolite Hybrid Catalysts with a Wide Range of Zeolite Particle Sizes. J. Catal. 2016, 338, 227–238. 10.1016/j.jcat.2016.02.025. [DOI] [Google Scholar]
- Dadgar F.; Myrstad R.; Pfeifer P.; Holmen A.; Venvik H. J. Direct Dimethyl Ether Synthesis from Synthesis Gas: The Influence of Methanol Dehydration on Methanol Synthesis Reaction. Catal. Today 2016, 270, 76–84. 10.1016/j.cattod.2015.09.024. [DOI] [Google Scholar]
- García-Trenco A.; Valencia S.; Martínez A. The Impact of Zeolite Pore Structure on the Catalytic Behavior of CuZnAl/Zeolite Hybrid Catalysts for the Direct DME Synthesis. Appl. Catal. A Gen. 2013, 468, 102–111. 10.1016/j.apcata.2013.08.038. [DOI] [Google Scholar]
- Aguayo A. T.; Ereña J.; Mier D.; Arandes J. M.; Olazar M.; Bilbao J. Kinetic Modeling of Dimethyl Ether Synthesis in a Single Step on a CuO-ZnO-Al2O3/γ-Al2O3 Catalyst. Ind. Eng. Chem. Res. 2007, 46, 5522–5530. 10.1021/ie070269s. [DOI] [Google Scholar]
- Qin Z.; Su T.; Ji H.; Jiang Y.; Liu R.; Chen J. Experimental and Theoretical Study of the Intrinsic Kinetics for Dimethyl Ether Synthesis from CO2 over Cu-Fe-Zr/HZSM-5. AIChE J. 2015, 61, 1613–1627. 10.1002/aic.14743. [DOI] [Google Scholar]
- Bahruji H.; Armstrong R. D.; Ruiz Esquius J.; Jones W.; Bowker M.; Hutchings G. J. Hydrogenation of CO2 to Dimethyl Ether over Brønsted Acidic PdZn Catalysts. Ind. Eng. Chem. Res. 2018, 57, 6821–6829. 10.1021/acs.iecr.8b00230. [DOI] [Google Scholar]
- Dubois J.-L.; Sayama K.; Arakawa H. Conversion of CO 2 to Dimethylether and Methanol over Hybrid Catalysts. Chem. Lett. 1992, 21, 1115–1118. 10.1246/cl.1992.1115. [DOI] [Google Scholar]
- Jun K.-W.; Jung M.-H.; Rama Rao K.S.; Choi M.-J.; Lee K.-W. Effective Conversion of CO2 to Methanol and Dimethyl Ether over Hybrid Catalysts. Stud. Surf. Sci. Catal. 1998, 114, 447–450. 10.1016/S0167-2991(98)80792-1. [DOI] [Google Scholar]
- Aguayo A. T.; Ereña J.; Sierra I.; Olazar M.; Bilbao J. Deactivation and Regeneration of Hybrid Catalysts in the Single-Step Synthesis of Dimethyl Ether from Syngas and CO2. Catal. Today 2005, 106, 265–270. 10.1016/j.cattod.2005.07.144. [DOI] [Google Scholar]
- Frusteri F.; Bonura G.; Cannilla C.; Drago Ferrante G.; Aloise A.; Catizzone E.; Migliori M.; Giordano G. Stepwise Tuning of Metal-Oxide and Acid Sites of CuZnZr-MFI Hybrid Catalysts for the Direct DME Synthesis by CO2 Hydrogenation. Appl. Catal. B Environ. 2015, 176–177, 522–531. 10.1016/j.apcatb.2015.04.032. [DOI] [Google Scholar]
- Bonura G.; Migliori M.; Frusteri L.; Cannilla C.; Catizzone E.; Giordano G.; Frusteri F. Acidity Control of Zeolite Functionality on Activity and Stability of Hybrid Catalysts during DME Production via CO2 Hydrogenation. J. CO2 Util. 2018, 24, 398–406. 10.1016/j.jcou.2018.01.028. [DOI] [Google Scholar]
- Yang G.; Tsubaki N.; Shamoto J.; Yoneyama Y.; Zhang Y. Confinement Effect and Synergistic Function of H-ZSM-5/Cu-ZnO-Al 2O3 Capsule Catalyst for One-Step Controlled Synthesis. J. Am. Chem. Soc. 2010, 132, 8129–8136. 10.1021/ja101882a. [DOI] [PubMed] [Google Scholar]
- Pinkaew K.; Yang G.; Vitidsant T.; Jin Y.; Zeng C.; Yoneyama Y.; Tsubaki N. A New Core - Shell-like Capsule Catalyst with SAPO-46 Zeolite Shell Encapsulated Cr/ZnO for the Controlled Tandem Synthesis of Dimethyl Ether from Syngas. Fuel 2013, 111, 727–732. 10.1016/j.fuel.2013.03.027. [DOI] [Google Scholar]
- Phienluphon R.; Pinkaew K.; Yang G.; Li J.; Wei Q.; Yoneyama Y.; Vitidsant T.; Tsubaki N. Designing Core (Cu/ZnO/Al 2 O 3)- Shell (SAPO-11) Zeolite Capsule Catalyst with a Facile Physical Way for Dimethyl Ether Direct Synthesis from Syngas. Chem. Eng. J. 2015, 270, 605–611. 10.1016/j.cej.2015.02.071. [DOI] [Google Scholar]
- Sanchez-Contador M.; Ateka A.; Aguayo A.T.; Bilbao J. Direct Synthesis of Dimethyl Ether from CO and CO 2 over a Core-Shell Structured CuO-ZnO-ZrO 2 @ SAPO-11 Catalyst. Fuel Process. Technol. 2018, 179, 258–268. 10.1016/j.fuproc.2018.07.009. [DOI] [Google Scholar]
- Baracchini G.; Machoke A. G. F.; Klumpp M.; Wen R.; Arnold P.; Schwieger W.; Dittmeyer R. Structured Catalysts for the Direct Synthesis of Dimethyl Ether from Synthesis Gas : A Comparison of Core @ Shell versus Hybrid Catalyst Configuration. Catal. Today 2020, 342, 46–58. 10.1016/j.cattod.2019.07.016. [DOI] [Google Scholar]
- Lam E.; Corral-Perez J. J.; Larmier K.; Noh G.; Wolf P.; Comas-Vives A.; Urakawa A.; Coperet C. CO2 Hydrogenation on Cu/Al2O3: Role of the Metal/Support Interface in Driving Activity and Selectivity of a Bifunctional Catalyst. Angew. Chem., Int. Ed. 2019, 58, 13989–13996. 10.1002/anie.201908060. [DOI] [PubMed] [Google Scholar]
- Kim J.; Sarma B. B.; Andrés E.; Pfánder N.; Concepciön P.; Prieto G. Surface Lewis Acidity of Periphery Oxide Species as a General Kinetic Descriptor for CO2 Hydrogenation to Methanol on Supported Copper Nanoparticles. ACS Catal. 2019, 9, 10409–10417. 10.1021/acscatal.9b02412. [DOI] [Google Scholar]
- Docherty S. R.; Phongprueksathat N.; Lam E.; Noh G.; Safonova O. V; Urakawa A.; Copéret C. Silica-Supported PdGa Nanoparticles: Metal Synergy for Highly Active and Selective CO2-to-CH3OH Hydrogenation. J. Am. Chem. Soc. Au 2021, 1, 450–458. 10.1021/jacsau.1c00021. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhou C.; Shi J.; Zhou W.; Cheng K.; Zhang Q.; Kang J.; Wang Y. Highly Active ZnO-ZrO2 Aerogels Integrated with H-ZSM-5 for Aromatics Synthesis from Carbon Dioxide. ACS Catal. 2020, 10, 302–310. 10.1021/acscatal.9b04309. [DOI] [Google Scholar]
- Wang T.; Yang C.; Gao P.; Zhou S.; Li S.; Wang H.; Sun Y. ZnZrOx Integrated with Chain-like Nanocrystal HZSM-5 as Efficient Catalysts for Aromatics Synthesis from CO2 Hydrogenation. Appl. Catal. B Environ. 2021, 286, 119929–119941. 10.1016/j.apcatb.2021.119929. [DOI] [Google Scholar]
- Feng W.-H.; Yu M.-M.; Wang L.-J.; Miao Y.-T.; Shakouri M.; Ran J.; Hu Y.; Li Z.; Huang R.; Lu Y.; Gao D.; Wu J. Insights into Bimetallic Oxide Synergy during Carbon Dioxide Hydrogenation to Methanol and Dimethyl Ether over GaZrOx Oxide Catalysts. ACS Catal. 2021, 11, 4704–4711. 10.1021/acscatal.0c05410. [DOI] [Google Scholar]
- Wang J.; Tang C.; Li G.; Han Z.; Li Z.; Liu H.; Cheng F.; Li C. High-Performance MaZrOx (Ma = Cd, Ga) Solid-Solution Catalysts for CO2 Hydrogenation to Methanol. ACS Catal. 2019, 9, 10253–10259. 10.1021/acscatal.9b03449. [DOI] [Google Scholar]
- Yu X.; Zhang J.; Wang X.; Ma Q.; Gao X.; Xia H.; Lai X.; Fan S.; Zhao T. Fischer–Tropsch Synthesis over Methyl Modified Fe2O3@SiO2 Catalysts with Low CO2 Selectivity. Appl. Catal. B Environ. 2018, 232, 420–428. 10.1016/j.apcatb.2018.03.048. [DOI] [Google Scholar]
- Javed M.; Zhang G.; Gao W.; Cao Y.; Dai P.; Ji X.; Lu C.; Yang R.; Xing C.; Sun J. From Hydrophilic to Hydrophobic: A Promising Approach to Tackle High CO2 Selectivity of Fe-Based Fischer–Tropsch Microcapsule Catalysts. Catal. Today 2019, 330, 39–45. 10.1016/j.cattod.2018.08.010. [DOI] [Google Scholar]
- Xu Y.; Li X.; Gao J.; Wang J.; Ma G.; Wen X.; Yang Y.; Li Y.; Ding M. A Hydrophobic FeMn@Si Catalyst Increases Olefins from Syngas by Suppressing C1 By-Products. Science 2021, 371, 610–613. 10.1126/science.abb3649. [DOI] [PubMed] [Google Scholar]
- Xie J. Designing the Right Protection. Science 2021, 371, 577. 10.1126/science.abg1750. [DOI] [PubMed] [Google Scholar]
- Tan M.; Tian S.; Zhang T.; Wang K.; Xiao L.; Liang J.; Ma Q.; Yang G.; Tsubaki N.; Tan Y. Probing Hydrophobization of a Cu/ZnO Catalyst for Suppression of Water - Gas Shift Reaction in Syngas Conversion. ACS Catal. 2021, 11, 4633–4643. 10.1021/acscatal.0c05585. [DOI] [Google Scholar]
- Diban N.; Aguayo A. T.; Bilbao J.; Urtiaga A.; Ortiz I. Membrane Reactors for in Situ Water Removal: A Review of Applications. Ind. Eng. Chem. Res. 2013, 52, 10342–10354. 10.1021/ie3029625. [DOI] [Google Scholar]
- van Kampen J.; Boon J.; van Berkel F.; Vente J.; van Sint Annaland M. Steam Separation Enhanced Reactions: Review and Outlook. Chem. Eng. J. 2019, 374, 1286–1303. 10.1016/j.cej.2019.06.031. [DOI] [Google Scholar]
- Iliuta I.; Iliuta M. C.; Larachi F. Sorption-Enhanced Dimethyl Ether Synthesis — Multiscale Reactor Modeling. Chem. Eng. Sci. 2011, 66, 2241–2251. 10.1016/j.ces.2011.02.047. [DOI] [Google Scholar]
- Gorbe J.; Lasobras J.; Francés E.; Herguido J.; Menéndez M.; Kumakiri I.; Kita H. Preliminary Study on the Feasibility of Using a Zeolite A Membrane in a Membrane Reactor for Methanol Production. Sep. Purif. Technol. 2018, 200, 164–168. 10.1016/j.seppur.2018.02.036. [DOI] [Google Scholar]
- Rieck genannt Best F.; Mundstock A.; Drager G.; Rusch P.; Bigall N. C.; Richter H.; Caro J. Methanol-to-Olefins in a Membrane Reactor with in Situ Steam Removal - The Decisive Role of Coking. ChemCatChem 2020, 12, 273–280. 10.1002/cctc.201901222. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Diban N.; Urtiaga A. M.; Ortiz I.; Ereña J.; Bilbao J.; Aguayo A. T. Influence of the Membrane Properties on the Catalytic Production of Dimethyl Ether with in Situ Water Removal for the Successful Capture of Co 2. Chem. Eng. J. 2013, 234, 140–148. 10.1016/j.cej.2013.08.062. [DOI] [Google Scholar]
- Najari S.; Gróf G.; Saeidi S. Enhancement of Hydrogenation of CO2 to Hydrocarbons via In-Situ Water Removal. Int. J. Hydrogen Energy 2019, 44, 24759–24781. 10.1016/j.ijhydene.2019.07.159. [DOI] [Google Scholar]
- Bansode A.; Urakawa A. Towards Full One-Pass Conversion of Carbon Dioxide to Methanol and Methanol-Derived Products. J. Catal. 2014, 309, 66–70. 10.1016/j.jcat.2013.09.005. [DOI] [Google Scholar]
- Ni Y.; Wang K.; Zhu W.; Liu Z. Realizing High Conversion of Syngas to Gasoline-Range Liquid Hydrocarbons on a Dual- Bed-Mode Catalyst Realizing High Conversion of Syngas to Gasoline-Range Liquid Hydrocarbons on a Dual-Bed-Mode Catalyst. Chem Catal. 2021, 1, 383–392. 10.1016/j.checat.2021.02.003. [DOI] [Google Scholar]
- Liu Z.; Ni Y.; Gao M.; Wang L.; Fang X.; Liu J.; Chen Z.; Wang N.; Tian P.; Zhu W.; Liu Z. Simultaneously Achieving High Conversion and Selectivity in Syngas-to-Propane Reaction via a Dual-Bed Catalyst System. ACS Catal. 2022, 12, 3985–3994. 10.1021/acscatal.1c05132. [DOI] [Google Scholar]
- Ni Y.; Liu Y.; Chen Z.; Yang M.; Liu H.; He Y.; Fu Y.; Zhu W.; Liu Z. Realizing and Recognizing Syngas-to-Olefins Reaction via a Dual-Bed Catalyst. ACS Catal. 2019, 9, 1026. 10.1021/acscatal.8b04794. [DOI] [Google Scholar]
- Kalck P.; Le Berre C.; Serp P. Recent Advances in the Methanol Carbonylation Reaction into Acetic Acid. Coord. Chem. Rev. 2020, 402, 213078–213088. 10.1016/j.ccr.2019.213078. [DOI] [Google Scholar]
- Le Berre C.; Serp P.; Kalck P.; Torrence G. P.. Acetic Acid. In Ullmann’s Encyclopedia of Industrial Chemistry; Wiley-VCH Verlag GmbH & Co, 2013; pp 1–34. [Google Scholar]
- Jones J. H. The CativaTM Process for the Manufacture of Acetic Acid: Iridium Catalyst Improves Productivity in an Established Industrial Process. Platin. Met. Rev. 2000, 44, 94–105. [Google Scholar]
- Sunley G. J.; Watson D. J. High Productivity Methanol Carbonylation Catalysis Using Iridium. The CativaTM Process for the Manufacture of Acetic Acid. Catal. Today 2000, 58, 293–307. 10.1016/S0920-5861(00)00263-7. [DOI] [Google Scholar]
- Ren Z.; Lyu Y.; Song X.; Ding Y. Review of Heterogeneous Methanol Carbonylation to Acetyl Species. Appl. Catal. A Gen. 2020, 595, 117488–117500. 10.1016/j.apcata.2020.117488. [DOI] [Google Scholar]
- Maitlis P. M.; Haynes A.; Sunley G. J.; Howard M. J. Methanol Carbonylation Revisited: Thirty Years On. J. Chem. Soc. Dalt. Trans. 1996, 2187–2196. 10.1039/dt9960002187. [DOI] [Google Scholar]
- Shang W.; Gao M.; Chai Y.; Wu G.; Guan N.; Li L. Stabilizing Isolated Rhodium Cations by MFI Zeolite for Heterogeneous Methanol Carbonylation. ACS Catal. 2021, 11, 7249–7256. 10.1021/acscatal.1c00950. [DOI] [Google Scholar]
- Fujimoto K.; Shikada T.; Omata K.; Tominaga H. Vapor Phase Carbonylation of Methanol with Solid Acid Catalysts. Chem. Lett. 1984, 13, 2047–2050. 10.1246/cl.1984.2047. [DOI] [Google Scholar]
- Ellis B.; Howard M. J.; Joyner R. W.; Reddy K. N.; Padley M. B.; Smith W. J. Heterogeneous Catalysts for the Direct, Halide-Free Carbonylation of Methanol. Stud. Surf. Sci. Catal. 1996, 101, 771–779. 10.1016/S0167-2991(96)80288-6. [DOI] [Google Scholar]
- Blasco T.; Boronat M.; Concepción P.; Corma A.; Law D.; Vidal-Moya J. A. Carbonylation of Methanol on Metal-Acid Zeolites: Evidence for a Mechanism Involving a Multisite Active Center. Angew. Chem., Int. Ed. 2007, 46, 3938–3941. 10.1002/anie.200700029. [DOI] [PubMed] [Google Scholar]
- Ormsby G.; Hargreaves J. S. J.; Ditzel E. J. A Methanol-Only Route to Acetic Acid. Catal. Commun. 2009, 10, 1292–1295. 10.1016/j.catcom.2009.02.005. [DOI] [Google Scholar]
- Ni Y.; Shi L.; Liu H.; Zhang W.; Liu Y.; Zhu W.; Liu Z. A Green Route for Methanol Carbonylation. Catal. Sci. Technol. 2017, 7, 4818–4822. 10.1039/C7CY01621B. [DOI] [Google Scholar]
- Qi J.; Christopher P. Atomically Dispersed Rh Active Sites on Oxide Supports with Controlled Acidity for Gas-Phase Halide-Free Methanol Carbonylation to Acetic Acid. Ind. Eng. Chem. Res. 2019, 58, 12632–12641. 10.1021/acs.iecr.9b02289. [DOI] [Google Scholar]
- Qi J.; Finzel J.; Robatjazi H.; Xu M.; Hoffman A. S.; Bare S. R.; Pan X.; Christopher P. Selective Methanol Carbonylation to Acetic Acid on Heterogeneous Atomically Dispersed ReO4/SiO2Catalysts. J. Am. Chem. Soc. 2020, 142, 14178–14189. 10.1021/jacs.0c05026. [DOI] [PubMed] [Google Scholar]
- Cheung P.; Bhan A.; Sunley G. J.; Iglesia E. Selective Carbonylation of Dimethyl Ether to Methyl Acetate Catalyzed by Acidic Zeolites. Angew. Chem., Int. Ed. 2006, 45, 1617–1620. 10.1002/anie.200503898. [DOI] [PubMed] [Google Scholar]
- Cheung P.; Bhan A.; Sunley G. J.; Law D. J.; Iglesia E. Site Requirements and Elementary Steps in Dimethyl Ether Carbonylation Catalyzed by Acidic Zeolites. J. Catal. 2007, 245, 110–123. 10.1016/j.jcat.2006.09.020. [DOI] [Google Scholar]
- Bhan A.; Allian A. D.; Sunley G. J.; Law D. J.; Iglesia E. Specificity of Sites within Eight-Membered Ring Zeolite Channels for Carbonylation of Methyls to Acetyls. J. Am. Chem. Soc. 2007, 129, 4919–4924. 10.1021/ja070094d. [DOI] [PubMed] [Google Scholar]
- Bhan A.; Iglesia E. A Link between Reactivity and Local Structure in Acid Catalysis on Zeolites. Acc. Chem. Res. 2008, 41, 559–567. 10.1021/ar700181t. [DOI] [PubMed] [Google Scholar]
- Gounder R.; Iglesia E. The Roles of Entropy and Enthalpy in Stabilizing Ion-Pairs at Transition States in Zeolite Acid Catalysis. Acc. Chem. Res. 2012, 45, 229–238. 10.1021/ar200138n. [DOI] [PubMed] [Google Scholar]
- Gounder R.; Iglesia E. The Catalytic Diversity of Zeolites: Confinement and Solvation Effects within Voids of Molecular Dimensions. Chem. Commun. 2013, 49, 3491–3509. 10.1039/c3cc40731d. [DOI] [PubMed] [Google Scholar]
- Boronat M.; Martínez-Sánchez C.; Law D.; Corma A. Enzyme-like Specificity in Zeolites: A Unique Site Position in Mordenite for Selective Carbonylation of Methanol and Dimethyl Ether with CO. J. Am. Chem. Soc. 2008, 130, 16316–16323. 10.1021/ja805607m. [DOI] [PubMed] [Google Scholar]
- Boronat M.; Martínez C.; Corma A. Mechanistic Differences between Methanol and Dimethyl Ether Carbonylation in Side Pockets and Large Channels of Mordenite. Phys. Chem. Chem. Phys. 2011, 13, 2603–2612. 10.1039/c0cp01996h. [DOI] [PubMed] [Google Scholar]
- Li B.; Xu J.; Han B.; Wang X.; Qi G.; Zhang Z.; Wang C.; Deng F. Insight into Dimethyl Ether Carbonylation Reaction over Mordenite Zeolite from In-Situ Solid-State NMR Spectroscopy. J. Phys. Chem. C 2013, 117, 5840–5847. 10.1021/jp400331m. [DOI] [Google Scholar]
- Zhou H.; Zhu W.; Shi L.; Liu H.; Liu S.; Ni Y.; Liu Y.; He Y.; Xu S.; Li L.; Liu Z. In Situ DRIFT Study of Dimethyl Ether Carbonylation to Methyl Acetate on H-Mordenite. J. Mol. Catal. A Chem. 2016, 417, 1–9. 10.1016/j.molcata.2016.02.032. [DOI] [Google Scholar]
- Lezcano-González I.; Vidal-Moya J. A.; Boronat M.; Blasco T.; Corma A. Identification of Active Surface Species for Friedel-Crafts Acylation and Koch Carbonylation Reactions by Insitu Solid-State NMR Spectroscopy. Angew. Chem., Int. Ed. 2013, 52, 5138–5141. 10.1002/anie.201209907. [DOI] [PubMed] [Google Scholar]
- Rasmussen D. B.; Christensen J. M.; Temel B.; Studt F.; Moses P. G.; Rossmeisl J.; Riisager A.; Jensen A. D. Ketene as a Reaction Intermediate in the Carbonylation of Dimethyl Ether to Methyl Acetate over Mordenite. Angew. Chem., Int. Ed. 2015, 54, 7261–7264. 10.1002/anie.201410974. [DOI] [PubMed] [Google Scholar]
- He T.; Liu X.; Xu S.; Han X.; Pan X.; Hou G.; Bao X. Role of 12-Ring Channels of Mordenite in DME Carbonylation Investigated by Solid-State NMR. J. Phys. Chem. C 2016, 120, 22526–22531. 10.1021/acs.jpcc.6b07958. [DOI] [Google Scholar]
- Liu Z.; Yi X.; Wang G.; Tang X.; Li G.; Huang L.; Zheng A. Roles of 8-Ring and 12-Ring Channels in Mordenite for Carbonylation Reaction: From the Perspective of Molecular Adsorption and Diffusion. J. Catal. 2019, 369, 335–344. 10.1016/j.jcat.2018.11.024. [DOI] [Google Scholar]
- Xue H.; Huang X.; Ditzel E.; Zhan E.; Ma M.; Shen W. Dimethyl Ether Carbonylation to Methyl Acetate over Nanosized Mordenites. Ind. Eng. Chem. Res. 2013, 52, 11510–11515. 10.1021/ie400909u. [DOI] [Google Scholar]
- Wang X.; Li R.; Yu C.; Zhang L.; Xu C.; Zhou H. Dimethyl Ether Carbonylation over Nanosheet-Assembled Hierarchical Mordenite. Microporous Mesoporous Mater. 2019, 274, 227–235. 10.1016/j.micromeso.2018.07.048. [DOI] [Google Scholar]
- Xue H.; Huang X.; Zhan E.; Ma M.; Shen W. Selective Dealumination of Mordenite for Enhancing Its Stability in Dimethyl Ether Carbonylation. Catal. Commun. 2013, 37, 75–79. 10.1016/j.catcom.2013.03.033. [DOI] [Google Scholar]
- Reule A. A. C.; Sawada J. A.; Semagina N. Effect of Selective 4-Membered Ring Dealumination on Mordenite-Catalyzed Dimethyl Ether Carbonylation. J. Catal. 2017, 349, 98–109. 10.1016/j.jcat.2017.03.010. [DOI] [Google Scholar]
- Cao K.; Fan D.; Li L.; Fan B.; Wang L.; Zhu D.; Wang Q.; Tian P.; Liu Z. Insights into the Pyridine-Modified MOR Zeolite Catalysts for DME Carbonylation. ACS Catal. 2020, 10, 3372–3380. 10.1021/acscatal.9b04890. [DOI] [Google Scholar]
- Zhan H.; Huang S.; Li Y.; Lv J.; Wang S.; Ma X. Elucidating the Nature and Role of Cu Species in Enhanced Catalytic Carbonylation of Dimethyl Ether over Cu/H-MOR. Catal. Sci. Technol. 2015, 5, 4378–4389. 10.1039/C5CY00460H. [DOI] [Google Scholar]
- Li Y.; Huang S.; Cheng Z.; Wang S.; Ge Q.; Ma X. Synergy between Cu and BrØnsted Acid Sites in Carbonylation of Dimethyl Ether over Cu/H-MOR. J. Catal. 2018, 365, 440–449. 10.1016/j.jcat.2018.06.026. [DOI] [Google Scholar]
- Reule A. A. C.; Semagina N. Zinc Hinders Deactivation of Copper-Mordenite: Dimethyl Ether Carbonylation. ACS Catal. 2016, 6, 4972–4975. 10.1021/acscatal.6b01464. [DOI] [Google Scholar]
- Wang S.; Guo W.; Zhu L.; Wang H.; Qiu K.; Cen K. Methyl Acetate Synthesis from Dimethyl Ether Carbonylation over Mordenite Modified by Cation Exchange. J. Phys. Chem. C 2015, 119, 524–533. 10.1021/jp511543x. [DOI] [Google Scholar]
- Ma M.; Zhan E.; Huang X.; Ta N.; Xiong Z.; Bai L.; Shen W. Carbonylation of Dimethyl Ether over Co-HMOR. Catal. Sci. Technol. 2018, 8, 2124–2130. 10.1039/C8CY00407B. [DOI] [Google Scholar]
- Zhou H.; Zhu W.; Shi L.; Liu H.; Liu S.; Xu S.; Ni Y.; Liu Y.; Li L.; Liu Z. Promotion Effect of Fe in Mordenite Zeolite on Carbonylation of Dimethyl Ether to Methyl Acetate. Catal. Sci. Technol. 2015, 5, 1961–1968. 10.1039/C4CY01580K. [DOI] [Google Scholar]
- Li Y.; Huang S.; Cheng Z.; Cai K.; Li L.; Milan E.; Lv J.; Wang Y.; Sun Q.; Ma X. Promoting the Activity of Ce-Incorporated MOR in Dimethyl Ether Carbonylation through Tailoring the Distribution of BrØnsted Acids. Appl. Catal. B Environ. 2019, 256, 117777–117789. 10.1016/j.apcatb.2019.117777. [DOI] [Google Scholar]
- Liu J.; Xue H.; Huang X.; Li Y.; Shen W. Dimethyl Ether Carbonylation to Methyl Acetate over HZSM-35. Catal. Lett. 2010, 139, 33–37. 10.1007/s10562-010-0411-3. [DOI] [Google Scholar]
- Feng P.; Zhang G.; Chen X.; Zang K.; Li X.; Xu L. Specific Zone within 8-Membered Ring Channel as Catalytic Center for Carbonylation of Dimethyl Ether and Methanol over FER Zeolite. Appl. Catal. A Gen. 2018, 557, 119–124. 10.1016/j.apcata.2018.03.018. [DOI] [Google Scholar]
- Román-Leshkov Y.; Moliner M.; Davis M. E. Impact of Controlling the Site Distribution of Al Atoms on Catalytic Properties in Ferrierite-Type Zeolites. J. Phys. Chem. C 2011, 115, 1096–1102. 10.1021/jp106247g. [DOI] [Google Scholar]
- Ham H.; Jung H. S.; Kim H. S.; Kim J.; Cho S. J.; Lee W. B.; Park M.-J.; Bae J. W. Gas-Phase Carbonylation of Dimethyl Ether on the Stable Seed-Derived Ferrierite. ACS Catal. 2020, 10, 5135–5146. 10.1021/acscatal.9b05144. [DOI] [Google Scholar]
- Feng X.; Yao J.; Li H.; Fang Y.; Yoneyama Y.; Yang G.; Tsubaki N. A Brand New Zeolite Catalyst for Carbonylation Reaction. Chem. Commun. 2019, 55, 1048–1051. 10.1039/C8CC08411D. [DOI] [PubMed] [Google Scholar]
- Xiong Z.; Zhan E.; Li M.; Shen W. DME Carbonylation over a HSUZ-4 Zeolite. Chem. Commun. 2020, 56, 3401–3404. 10.1039/D0CC00886A. [DOI] [PubMed] [Google Scholar]
- Lusardi M.; Chen T. T.; Kale M.; Kang J. H.; Neurock M.; Davis M. E. Carbonylation of Dimethyl Ether to Methyl Acetate over SSZ-13. ACS Catal. 2020, 10, 842–851. 10.1021/acscatal.9b04307. [DOI] [Google Scholar]
- Wang S.; Li S.; Zhang L.; Qin Z.; Chen Y.; Dong M.; Li J.; Fan W.; Wang J. Mechanistic Insights into the Catalytic Role of Various Acid Sites on ZSM-5 Zeolite in the Carbonylation of Methanol and Dimethyl Ether. Catal. Sci. Technol. 2018, 8, 3193–3204. 10.1039/C8CY00296G. [DOI] [Google Scholar]
- Jiao F.; Pan X.; Gong K.; Chen Y.; Li G.; Bao X. Shape-Selective Zeolites Promote Ethylene Formation from Syngas via a Ketene Intermediate. Angew. Chem., Int. Ed. 2018, 57, 4692–4696. 10.1002/anie.201801397. [DOI] [PubMed] [Google Scholar]
- Chen H.; Liu Z.; Li N.; Jiao F.; Chen Y.; Zhao Z.; Guo M.; Liu X.; Han X.; Pan X.; Gong X.; Hou G.; Bao X. A Mechanistic Study of Syngas Conversion to Light Olefins over OXZEO Bifunctional Catalysts: Insights into the Initial Carbon - Carbon Bond Formation on the Oxide. Catal. Sci. Technol. 2022, 12, 1289–1295. 10.1039/D1CY01807H. [DOI] [Google Scholar]
- Lai Z.; Sun N.; Jin J.; Chen J.; Wang H.; Hu P. Resolving the Intricate Mechanism and Selectivity of Syngas Conversion on Reduced ZnCr2Ox: A Quantitative Study from DFT and Microkinetic Simulations. ACS Catal. 2021, 11, 12977–12988. 10.1021/acscatal.1c03579. [DOI] [Google Scholar]
- Fu X.; Xiao J. Toward Understanding and Simplifying the Reaction Network of Ketene Production on ZnCr2O4 Spinel Catalysts. J. Phys. Chem. C 2021, 125, 24902–24914. 10.1021/acs.jpcc.1c07375. [DOI] [Google Scholar]
- Fu X.; Li J.; Long J.; Guo C.; Xiao J. Understanding the Product Selectivity of Syngas Conversion on ZnO Surfaces with Complex Reaction Network and Structural Evolution. ACS Catal. 2021, 11, 12264–12273. 10.1021/acscatal.1c02111. [DOI] [Google Scholar]
- Wang S.; Zhang L.; Wang P.; Liu X.; Chen Y.; Qin Z.; Dong M.; Wang J.; He L.; Olsbye U.; Fan W. Highly Effective Conversion of CO2 into Light Olefins Abundant in Ethene. Chem 2022, 8, 1376–1394. 10.1016/j.chempr.2022.01.004. [DOI] [Google Scholar]
- Cnudde P.; Demuynck R.; Vandenbrande S.; Waroquier M.; Sastre G.; Speybroeck V. Van. Light Olefin Diffusion during the MTO Process on H-SAPO-34: A Complex Interplay of Molecular Factors. J. Am. Chem. Soc. 2020, 142, 6007–6017. 10.1021/jacs.9b10249. [DOI] [PubMed] [Google Scholar]
- Cnudde P.; Redekop E. A.; Dai W.; Porcaro N. G.; Waroquier M.; Bordiga S.; Hunger M.; Li L.; Olsbye U.; Van Speybroeck V. Experimental and Theoretical Evidence for the Promotional Effect of Acid Sites on the Diffusion of Alkenes through Small-Pore Zeolites. Angew. Chem., Int. Ed. 2021, 60, 10016–10022. 10.1002/anie.202017025. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhou W.; Kang J.; Cheng K.; He S.; Shi J.; Zhou C.; Zhang Q.; Chen J.; Peng L.; Chen M.; Wang Y. Direct Conversion of Syngas into Methyl Acetate, Ethanol, and Ethylene by Relay Catalysis via the Intermediate Dimethyl Ether. Angew. Chem., Int. Ed. 2018, 57, 12012–12016. 10.1002/anie.201807113. [DOI] [PubMed] [Google Scholar]
- Kang J.; He S.; Zhou W.; Shen Z.; Li Y.; Chen M.; Zhang Q.; Wang Y. Single-Pass Transformation of Syngas into Ethanol with High Selectivity by Triple Tandem Catalysis. Nat. Commun. 2020, 11, 1–11. 10.1038/s41467-020-14672-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang Y.; Ding C.; Wang J.; Jia Y.; Xue Y.; Gao Z.; Yu B.; Gao B.; Zhang K.; Liu P. Intermediate Product Regulation over Tandem Catalysts for One-Pass Conversion of Syngas to Ethanol. Catal. Sci. Technol. 2019, 9, 1581–1594. 10.1039/C8CY02593B. [DOI] [Google Scholar]
- Upham D. C.; Orazov M.; Jaramillo T. F. Phosphate-Passivated Mordenite for Tandem-Catalytic Conversion of Syngas to Ethanol or Acetic Acid. 2021, 399, 132–141. 10.1016/j.jcat.2021.04.029. [DOI] [Google Scholar]
- Lu P.; Chen Q.; Yang G.; Tan L.; Feng X.; Yao J.; Yoneyama Y.; Tsubaki N. Space-Confined Self-Regulation Mechanism from a Capsule Catalyst to Realize an Ethanol Direct Synthesis Strategy. ACS Catal. 2020, 10, 1366–1374. 10.1021/acscatal.9b02891. [DOI] [Google Scholar]
- Ahlers S. J.; Bentrup U.; Linke D.; Kondratenko E. V. An Innovative Approach for Highly Selective Direct Conversion of CO2 into Propanol Using C2H4 and H2. ChemSusChem 2014, 7, 2631–2639. 10.1002/cssc.201402212. [DOI] [PubMed] [Google Scholar]
- Mavlyankariev S. A.; Ahlers S. J.; Kondratenko V. A.; Linke D.; Kondratenko E. V. Effect of Support and Promoter on Activity and Selectivity of Gold Nanoparticles in Propanol Synthesis from CO2, C2H4, and H2. ACS Catal. 2016, 6, 3317–3325. 10.1021/acscatal.6b00590. [DOI] [Google Scholar]
- Breuil P.-A. R.; Magna L.; Olivier-Bourbigou H. Role of Homogeneous Catalysis in Oligomerization of Olefins: Focus on Selected Examples Based on Group 4 to Group 10 Transition Metal Complexes. Catal. Lett. 2015, 145, 173–192. 10.1007/s10562-014-1451-x. [DOI] [Google Scholar]
- Finiels A.; Fajula F.; Hulea V. Nickel-Based Solid Catalysts for Ethylene Oligomerization - a Review. Catal. Sci. Technol. 2014, 4, 2412–2426. 10.1039/C4CY00305E. [DOI] [Google Scholar]
- Olivier-Bourbigou H.; Breuil P. A. R.; Magna L.; Michel T.; Espada Pastor M. F.; Delcroix D. Nickel Catalyzed Olefin Oligomerization and Dimerization. Chem. Rev. 2020, 120, 7919–7983. 10.1021/acs.chemrev.0c00076. [DOI] [PubMed] [Google Scholar]
- Martínez A.; Arribas M. A.; Concepción P.; Moussa S. New Bifunctional Ni-H-Beta Catalysts for the Heterogeneous Oligomerization of Ethylene. Appl. Catal. A Gen. 2013, 467, 509–518. 10.1016/j.apcata.2013.08.021. [DOI] [Google Scholar]
- Toch K.; Thybaut J. W.; Marin G. B. Ethene Oligomerization on Ni-SiO2 -Al2O3: Experimental Investigation and Single-Event MicroKinetic Modeling. Appl. Catal. A Gen. 2015, 489, 292–304. 10.1016/j.apcata.2014.10.036. [DOI] [Google Scholar]
- Brogaard R. Y.; Olsbye U. Ethene Oligomerization in Ni-Containing Zeolites : Theoretical Discrimination of Reaction Mechanisms. ACS Catal. 2016, 6, 1205–1214. 10.1021/acscatal.5b01957. [DOI] [Google Scholar]
- Brogaard R. Y.; Kømurcu M.; Dyballa M. M.; Botan A.; Van Speybroeck V.; Olsbye U.; De Wispelaere K. De. Ethene Dimerization on Zeolite-Hosted Ni Ions: Reversible Mobilization of the Active Site. ACS Catal. 2019, 9, 5645–5650. 10.1021/acscatal.9b00721. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Agirrezabal-Telleria I.; Iglesia E. Mechanistic Insights and Consequences of Intrapore Liquids in Ethene, Propene, and Butene Dimerization on Isolated Ni2+ Sites Grafted within Aluminosilicate Mesopores. J. Catal. 2020, 389, 690–705. 10.1016/j.jcat.2020.06.038. [DOI] [Google Scholar]
- Caulkins R.; Joshi R.; Gounder R.; Ribeiro F. H. Effects of Ethene Pressure on the Deactivation of Ni-Zeolites During Ethene Oligomerization at Sub-Ambient Temperatures. ChemCatChem 2022, 14, e202101478 10.1002/cctc.202101478. [DOI] [Google Scholar]
- Yang Z.; Dennis-Smither B. J.; Buda C.; Easey A.; Jackson F.; Price G. A.; Sainty N.; Tan X.; Xu Z.; Sunley G. J. Aromatic Aldehydes as Tuneable and Ppm Level Potent Promoters for Zeolite Catalysed Methanol Dehydration to DME. Catal. Sci. Technol. 2023, 13, 3590. 10.1039/D3CY00105A. [DOI] [Google Scholar]
- Qin Y.; Niu G.; Wang X.; Luo D.; Duan Y. Status of CO2 Conversion Using Microwave Plasma. J. CO2 Util. 2018, 28, 283–291. 10.1016/j.jcou.2018.10.003. [DOI] [Google Scholar]
- Ong M. Y.; Nomanbhay S.; Kusumo F.; Show P. L. Application of Microwave Plasma Technology to Convert Carbon Dioxide (CO2) into High Value Products: A Review. J. Clean. Prod. 2022, 336, 130447–130460. 10.1016/j.jclepro.2022.130447. [DOI] [Google Scholar]
- Ou X.; Tomatis M.; Lan Y.; Jiao Y.; Chen Y.; Guo Z.; Gao X.; Wu T.; Wu C.; Shi K.; Azapagic A.; Fan X. A Novel Microwave-Assisted Methanol-to-Hydrocarbons Process with a Structured ZSM-5/SiC Foam Catalyst: Proof-of-Concept and Environmental Impacts. Chem. Eng. Sci. 2022, 255, 117669–117682. 10.1016/j.ces.2022.117669. [DOI] [Google Scholar]




















