Next Article in Journal
Spent Nuclear Fuel—Waste or Resource? The Potential of Strategic Materials Recovery during Recycle for Sustainability and Advanced Waste Management
Next Article in Special Issue
Computational Modelling on Gasification Processes of Municipal Solid Wastes Including Molten Slag
Previous Article in Journal
A Detailed Database of the Chemical Properties and Methane Potential of Biomasses Covering a Large Range of Common Agricultural Biogas Plant Feedstocks
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Concise Review of Catalytic Synthesis of Methanol from Synthesis Gas

1
Agricultural & Biological Engineering Department, Packaging Engineering Program, University of Florida, 1741 Museum Rd, Gainesville, FL 32611, USA
2
Chemical Engineering Department, University of Florida, Gainesville, FL 32611, USA
*
Author to whom correspondence should be addressed.
Waste 2023, 1(1), 228-248; https://doi.org/10.3390/waste1010015
Submission received: 21 October 2022 / Revised: 8 December 2022 / Accepted: 3 January 2023 / Published: 11 January 2023

Abstract

:
Regenerative Robust Gasification promises to convert unsorted organic waste, including all plastic waste, into the fungible, primary feedstock chemical methanol. As the backbone of the C1 chemical industry, methanol has broad application in circular economy chemical synthesis. This paper summarizes traditional and newer approaches for producing methanol from synthesis gas. Approaches, methods, reaction mechanisms, catalyst systems, catalyst synthesis methods, reactor types, and many other aspects are summarized.

1. Research Necessity

Methanol is a primary feedstock chemical for the subsequent synthesis of chemicals and products [1]. Some examples include formaldehyde, acetic acid, acetic anhydride, dimethyl ether (via Methanol-to-Dimethyl Ether, MTD) [2], olefins (via Methanol-to-Olefins, MTO) [3], gasoline (via Methanol-to-Gasoline, MTG) [4], aromatics (via Methanol-to-Aromatics, MTA) [5], and dimethyl carbonate (via Methanol-to-dimethyl carbonate, MTC) [6].
Methanol is a traditional value-added chemical product. Compared with traditional petroleum-based fuels, alcohol can burn more fully due to the presence of oxygen. Methanol is the simplest alcohol with a small molecular weight. Second, methanol has a high-octane value and good anti-knock properties and can be used as a clean gasoline additive. As a fuel, the effective thermal efficiency of pure methanol is nearly 30% higher than that of gasoline. Therefore, methanol is a clean and efficient liquid vehicle fuel and a high-power raw material for fuel cells [7].
In general, methanol is widely used in chemical, agricultural, medical, and other fields. Research on methanol synthesis is, therefore, of great significance to the development of many fields.
Recently, the Consortium for Waste Circularity has advocated the use of “Regenerative Robust Gasification” to simplify waste collection and boost recycling rates. Robust Gasification refers to a type of gasifier that can accept everything in municipal solid waste (MSW) without the need for sorting. While Robust Gasifiers melt down and recover inorganics (e.g., metals and glass), these materials do not add benefit to the process, so inorganics that can be removed from the waste stream would likely be removed and recycled traditionally. To the extent that regional markets demand specific organic materials, such as biomass and particular plastics, these materials may also be selectively removed to serve the demand of other material-specific recycling processes. However, all other materials, including multi-layer plastics, metallized flexible films, and many materials that are not conducive to traditional recycling can be gasified into syngas. “Regenerative robust gasification” refers to the use of syngas for the subsequent manufacture of new products, plastics, and packaging, as opposed to use for fuel and/or power. Since syngas is not easily transported, the conversion of syngas into a fungible, widely used chemical feedstock, such as methanol, is proposed. Therefore, a small-scale, distributed manufacture of methanol from waste-derived syngas is required. This work provides a concise review of current state-of-the-art methanol synthesis from syngas.

2. Research Background and Current Situation

The methanol synthesis process includes gas and liquid phase synthesis methods. Due to the differences between these methods, each one will be reviewed separately below.

2.1. Gas Phase Processes

2.1.1. Synthesis Methods and Catalysts

At present, commercial methanol synthesis is primarily a gas-phase process. According to different operating pressures, the gas-phase process can be characterized as high-pressure, medium-pressure, and low-pressure [8].

High-Pressure Gas-Phase Methanol Synthesis

In 1923, Mittash and Schneider of BASF (Germany) first synthesized methanol through a zinc oxide/chromium-based (ZnO/Cr2O3) catalyst with carbon monoxide (CO) and hydrogen (H2) at 300–400 °C and 30 MPa. Then, they established the world’s first commercial methanol plant capable of producing 300 tons of methanol annually [9].
Before the mid−1960s, almost all methanol synthesis factories used the high-pressure method with the ZnO/Cr2O3 catalyst. However, the high temperatures and pressures resulted in high costs. Additionally, the activity of the catalyst was poor, the purity of methanol was low, and large quantities of chromium posed pollution risks.

Low-Pressure Gas-Phase Methanol Synthesis

In 1966, ICI (Britain) successfully developed a copper/zinc oxide/alumina (Cu/ZnO/Al2O3) catalyst. Copper-based catalysts have been widely used in the industry because of their good activity and selectivity. The ICI method requires lower pressures and temperatures (5.0–8.0 MPa, 240–270 °C), which reduces the cost and consumption of raw materials. Additionally, side reactions, such as alkylation, are reduced, and the purity of crude methanol increases [10].
Five years later, Lurgi (Germany) introduced another low-pressure methanol synthesis process using a more active copper/zinc oxide/alumina/vanadium oxide (Cu/ZnO/Al2O3/V2O5) catalyst system, which has the same excellent performance. This led to an upsurge in research on low-pressure methanol synthesis processes using copper-based catalysts [11].

Medium-Pressure Gas-Phase Methanol Synthesis

While the low-pressure method offered considerable advantages, one disadvantage is the requirement for large-volume reaction vessels. Therefore, the medium-pressure method was developed based on the low-pressure method.
The medium-pressure method has the same catalytic system as the low-pressure method, using a copper-based catalyst and the same reaction temperature, but the reaction pressure is increased to 10–20 MPa. Therefore, this method not only has similar advantages to the low-pressure method but also solves the shortcomings of the excessively large footprint required in the production of the low-pressure method.

Non-Copper Catalyst

It has been discovered that noble metals, such as Pd, Pt, Zr, and Rh, can also be active catalysts for methanol synthesis [12,13,14,15]. Although the mechanism of methanol formation on these catalysts has not yet been elucidated, it has been recognized that these metals offer good activity for methanol synthesis due to their strong hydrogenation properties, accompanied by the ability to activate adsorbed CO. Studies have shown that the activity and selectivity of Pd and Rh catalysts are largely affected by the composition of the support and additives. It has been found that Pd supported on some types of materials, such as silica or cerium dioxide, has high activity and selectivity for methanol synthesis. Song et al. [16] used hollow silica as a carrier to prepare a new eggshell-type Pd/SiO2 catalyst and Ca-Pd/SiO2 catalyst with Ca as an additive by the “incipient wetness impregnation” (IWI) method and investigated methanol synthesis performance of two types of catalysts. It was found that the activities of Ca-Pd/SiO2 and Pd/SiO2 were 36.50 and 2.54 mmol CO/mol Pd/s at 250 °C and 2.5 MPa, respectively. The selectivity to methanol was 95.2% for both catalysts.

2.1.2. Types of Reactors

The synthesis of methanol generally involves synthesis gas (syngas) production, conversion, purification, crude methanol synthesis, and methanol refining (Figure 1). The reactor is central to the entire methanol synthesis process.
Continuing research on methanol production focuses on the reduction of costs and upsizing scale. The high-pressure method cannot compete with low- and medium-pressure methods, so it is being phased out. Since the mid−1970s, almost all new or expanded methanol factories in the world have adopted low- or medium-pressure methods. The following are the representative methods among them:

ICI Method

The ICI process is most widely used in cold shock synthesis towers [17,18]. The tower divides the reaction bed into several adiabatic sections. Between the two sections, cold syngas is introduced through a lozenge sparger. The temperature of each section is maintained at distinct values. Advantages of the ICI cold shock synthesis tower are: the reactor structure is simple and can be operated in a single tower; the temperature is easy to control, and the operation is stable; the production capacity is high (up to 3000 tons per day); catalysts are not layered, simplifying loading and unloading; utilization of reactants is high; and the investment is relatively low. The disadvantages are as follows: catalyst loading is relatively high, and the recovered heat of the reaction is relatively low. Compared to the Lurgi process, the ICI process control system is more complex and more important for better controlling the temperature in each section.

Lurgi Method

The Lurgi method mainly uses shell-and-tube isothermal reactors [17]. Reactors are generally of two types: catalysts are packed in tubes, with water flowing between tubes to remove the heat of the reaction; catalysts are packed between the tubes, and water flows in the tubes. The reaction gas flow is axial in both types of reactors. Its advantages are: temperature distributions in the catalyst bed are uniform, temperature changes are small, few side reactions, and product quality is high; reaction temperatures can be accurately controlled by adjusting the pressure of the by-product steam, and the catalyst has a long service life; yield is high; heat released during the methanol synthesis reaction can be recovered and utilized. Saturated medium-pressure steam, as a by-product, can also be used as a heat source of the distillation tower reboiler during the subsequent refining of the crude methanol product. The process becomes less sensitive to fluctuations as production capacity increases, and crude methanol is relatively pure. The disadvantage is that the reactor structure is complex, and cost is relatively high, thus limiting its production scale.

Other

In addition to the above two reactor types, there are other reactors that also perform well. For example, the Linde reactor is an isothermal coil heat exchanger [10]. Catalysts are placed outside the coil, and water inside the coil is used to remove heat, creating high-pressure steam as a by-product. Its principles and characteristics are very similar to those of the Lurgi method. The Linde process offers good performance but with complexity and cost. The Linde method may not scale as well for high production requirements.
The Topsoe reactor is similar to the ICI method [19]. The reactor is a multi-stage adiabatic synthesis reactor with a heat exchanger placed between catalysts to remove heat. Its advantages are: the pressure drop between beds is small, and the methanol synthesis reaction rate can be accelerated by increasing the gas velocity, also increasing methanol yield; the reactor is allowed to be filled with smaller catalyst particles, which improves reaction rate; and the production scale can be expanded by increasing the height of the reactor with the diameter unchanged. The disadvantages are as follows: the reaction gas flows axially, resulting in large pressure drops within the bed, which increases the difficulty of equipment manufacturing and investment costs; the gas distribution in the recycle loop is relatively uneven; and the recycled gas volume is large, so catalysts placed in the reactor cannot be fully utilized.

2.1.3. Reaction Mechanism

The primary methanol synthesis reaction stoichiometries are as follows:
CO + 2 H 2   CH 3 OH , Δ H = 90.64   kJ / mol
CO 2 + 3 H 2   CH 3 OH + H 2 O , Δ H = 49.45   kJ / mol  
The production of methanol from syngas is exothermic. Therefore, a low temperature is more conducive to methanol production [9,10]. However, traditional industrial methanol synthesis is carried out under relatively high temperatures and high pressures, so the theoretical conversion rate of CO is equilibrium limited. There have been many reports on traditional gas-phase methanol synthesis, and much effort has been made to avoid the accumulation of reaction heat. Usually, the single-pass conversion rate of CO for methanol synthesis is still less than 20%.
This fundamental thermodynamic equilibrium limitation is difficult to overcome. At the same time, various side reactions and water generated during the hydrogenation of CO2 to produce methanol reduce the purity of crude methanol. The methanol concentration in the output is very limited. Therefore, work is needed to refine crude methanol, such as recycling and distillation, which increases capital and operating costs. In addition, the most popular copper-based catalysts in gas phase methanol production tend to be thermally unstable. Furthermore, they are easily poisoned by S, Cl, and other elements, requiring multi-step purification of the syngas feed. If the reaction temperature exceeds the catalyst tolerance, thermal sintering deactivates the catalyst. To control the temperature of the reaction bed within an acceptable range, the single-pass conversion rate is maintained at a low level.

2.1.4. The Role of CO2 in Methanol Synthesis

In the industrial synthesis of methanol, a certain amount of CO2 is usually required for raw synthesis gas. The conversion of syngas to methanol may involve several undesirable side reactions. Additionally, Cu-based catalysts also have activity for some of these side reactions, so the specific reaction mechanism is still undetermined. A widely accepted route is the stepwise hydrogenation of chemically adsorbed CO to produce methanol without the dissociation of CO. It is generally believed that the presence of a certain amount of CO2 promotes the production of methanol from syngas, but the role of CO2 is still debated, as highlighted below.
Klier [20] believed that CO2 does not participate in methanol synthesis, and its role was to regulate the surface structure, valence, and dispersion of copper in copper-based methanol synthesis catalysts.
Denis et al. [21] believed that CO2 could alter CO hydrogenation by forming formate and then serving as a catalyst to increase the reaction rate of methanol synthesis.
Chinchen et al. [22] used tracer atom technology to show that when methanol is synthesized from CO/CO2/H2 on Cu/ZnO/Al2O3 catalyst, methanol is directly generated from CO2 instead of first generating CO through reverse water gas shift and then producing methanol. CO2 is the direct precursor of methanol synthesis, and CO can regenerate the active Cu surface by reducing CuO to Cu.
Zhang et al. [23] used three different gas compositions CO + H2, CO2 + H2, and CO + CO2 + H2 to synthesize methanol and studied the role of CO2 in methanol synthesis. It was shown that the conversion and efficiency of the three types of gas compositions to methanol follow this order: CO + CO2 + H2 > CO2 + H2 > CO + H2. CO2 significantly promoted the methanol synthesis reaction, and CO2 was also the carbon source for methanol synthesis.
Daniel et al. [24] indicated that CO2 content in syngas facilitates methanol synthesis by controlling the oxidation state of the interface sites. In the absence of CO2, interface sites are obviously less active and more metallic.

2.1.5. Research on the Active Center of Copper-Based Catalysts

At present, there is no consensus on the active center of copper-based methanol synthesis catalysts, and sometimes even conflicting results are obtained [25]. There are mainly the following three points of view:

Cu0 Is the Active Center

Represented by ICI [26], it is believed that Cu0 is the only effective component in low-temperature and low-pressure methanol synthesis catalysts. Chinchen et al. [27] used copper-based catalysts with different supports to determine the relationship between the reaction activity and the surface area of copper metal. They found that the two were directly proportional. Robbins et al. [28] studied the relationship between reaction activity and the surface area of copper metal on different Cu/SiO2 structures and reached the same conclusions as Chinchen. Therefore, they believe that the surface of copper metal is the active center for methanol synthesis. Since the activity per unit surface area of copper is not influenced by the carrier, they believe that the carrier only functions as a structural agent to maintain dispersion, and they did not find a strong interaction between Cu-ZnO. Another experimental result that supports the Cu0 active center theory is that no Cu+ and Cu2+ are found through in-situ detection of the catalyst in the reducing or reaction atmosphere. In-situ surface analysis by Fleisch et al. [29] showed that only Cu0 existed in the industrial Cu/ZnO/Al2O3 catalyst under synthesis conditions.

Cu+ Is the Active Center

As early as 1955, Nata believed that it was the Cu+ oxidation state copper, instead of metallic copper, that really played the role of an active center for methanol synthesis. Herman et al. [30] proposed that the active center of Cu-based catalysts should be Cu+ dissolved in ZnO. They found that the activity of the catalyst with more Cu+ sites was better, and the increase in activity came from the Cu+/ZnO solid solution structure. Duprez et al. [31] found Cu+ by XPS during the reaction and found a good correlation between the Cu+ content and methanol synthesis activity. Through NMR research, Chu et al. [32] found that the activity of synthesizing methanol on copper catalysts containing different alkali metal additives is only related to the amount of Cu+.

Cu0-Cu+ Is the Active Center

Duprez‘s [31] research results show that in low CuO content catalysts (<30%), the Cu2+ ions dissolved in the ZnO lattice are reduced to Cu0-Cu+. Due to the easy oxidation-reduction properties of CuO and the high stability of Cu+, they believe that the two-dimensional Cu0-Cu+ species are the active species in Cu/ZnO catalyst. Denise et al. [33] studied the relationship between copper surface area and the activity of Cu/ZnO/Al2O3 catalysts under normal pressure. The yield of methanol was shown to be proportional to the difference between the copper surface area before and after the reaction, and this difference is equal to the amount of copper oxidized during the reaction.

2.1.6. Summary

In summary, the Zn-Cr-based catalyst that was first invented for the high-pressure process has now been eliminated due to harsh operating conditions, low activity, and environmental pollution caused by heavy metals in the catalyst. Pd-based catalysts belong to noble metal catalysts, and the industrialization cost is relatively high. The low-temperature methanol synthesis technology on Cu-based catalysts is relatively mature, and its good activity and high methanol selectivity make it a representative catalyst for the middle- and low-pressure methanol synthesis processes. However, the commonly used copper-based catalysts are thermally unstable, and are sensitive to poisoning by S, Cl, and other elements. Moreover, the fundamental limitations of thermodynamics are difficult to overcome for the traditional gas-phase methanol synthesis process, and the reaction conditions are relatively harsh, resulting in a series of shortcomings, such as low synthesis efficiency, high energy consumption, and high cost. Under this circumstance, research has been devoted to low-temperature liquid-phase processes that can reach higher conversion.

2.2. Liquid Phase Processes

Danish scientist J. Christiansen first proposed the liquid phase methanol synthesis process in 1919, but it was not until 1975 that it finally attracted attention when it was proposed again by Chemical Systems Inc. (Fort Smith, AR, USA) [34]. Different from the gas-solid two-phase reaction of the industrial synthesis process, the liquid-phase synthesis method is a gas-liquid-solid three-phase system. Briefly, a liquid phase medium with a high heat capacity and high thermal conductivity is introduced as a solvent, which removes reaction heat rapidly, thereby keeping the reaction system at a lower temperature. At the same time, the catalyst is dispersed in the liquid phase medium, which increases the reaction surface area and accelerates the reaction process. Therefore, methanol synthesis can be carried out at a lower reaction temperature and pressure. According to the thermodynamic equilibrium, a lower temperature corresponds to a higher equilibrium conversion rate. Generally, the research on the low-temperature liquid-phase methanol synthesis process mainly uses CO/H2 as the raw material and synthesizes methanol at 100∼150 °C and 1∼3 MPa. The single-pass conversion rate of CO can reach 90%. This process solves the problems of the traditional gas phase method and has the advantages of high single-pass conversion, low water content in product methanol, and low energy consumption, so it should offer commercial benefits.

2.2.1. Reaction Mechanism

CO/H2 Reaction System

Unlike the one-step reaction of direct hydrogenation of CO to produce methanol in the gas phase method, current researchers generally believe that the mechanism of low-temperature liquid-phase methanol synthesis that uses CO/H2 as a raw material follows a two-step reaction mechanism. Methanol already produced facilitates the reaction and is first carbonylated to produce methyl formate, and then one molecule of methyl formate is hydrogenolyzed to produce two molecules of methanol. The net reaction result is the formation of one molecule of methanol.
CH 3 OH + CO   HCOOCH 3
HCOOCH 3 + 2 H 2   2 CH 3 OH
The first carbonylation reaction step is the rate-controlling step [35].

CO/CO2/H2 Reaction System

Zeng et al. [36] used CO/CO2/H2 as the raw material to synthesize methanol in the low-temperature liquid phase, and believe that the reaction mechanism is mainly composed of three steps: CO in the raw material undergo a water-gas shift to generate additional CO2, and then the CO2 reacts with H2 to produce formate; the formate reacts with the alcohol in the catalytic system to generate an ester; the ester is hydrogenated to generate methanol. Figure 2 shows the difference between the mechanisms of the traditional gas phase methanol synthesis (A) and the low-temperature methanol liquid phase synthesis of the alcohol system (B).
Yang et al. [37,38,39] used Cu/ZnO and Cu/MgO as catalysts, respectively, and observed changes in the functional groups of adsorbed intermediates on the catalyst surface during low-temperature methanol synthesis by DRIFTS (Diffuse Reflectance infrared Fourier Transform Spectroscopy). Then, they inferred the corresponding reaction mechanism according to changes. For these two catalysts, the conclusions were roughly the same, except for the differences in the adsorption sites on the catalysts. The mechanism of methanol liquid phase synthesis in the ethanol system is shown in Figure 2B. In the process of mechanism research, ethanol is introduced into the reaction system in the form of ethanol vapor under atmospheric pressure, so the products measured in the experiment exist in gaseous form. The research results show that adsorbed formate is easily formed when the catalyst is exposed to synthesis gas. Formate can react with ethanol at low temperatures to produce adsorbed or gaseous ethyl formate. In the absence of ethanol, formate cannot be hydrogenated. As an intermediate product of the reaction, ethyl formate is easily reduced by hydrogen atoms on Cu to obtain gaseous methanol and ethanol. Moreover, the study found that the reaction of adsorbed formate species with ethanol in the gas phase to form ethyl formate is the key step in low-temperature methanol synthesis. This step changes the reaction process such that formate is the intermediate rather than the methoxy group as in the traditional gas phase of methanol synthesis. Since the generation of methanol no longer goes through the formation of methoxy groups, the methanol synthesis can be achieved at low temperatures. Ethanol significantly reduces the temperature required for the reaction and can be recycled. It not only plays the role of heat carrier but also acts as a catalyst. At the same time, the carbonylation reaction was found to be a Rideal-type process, which was affected by electronic effects and steric hindrance [40].
These studies suggest that in the alcohol solvent system, the mechanism of methanol synthesis is changed by the solvent. Whether it is a single copper-based catalyst or a composite catalyst, the reaction produces the corresponding intermediate formic acid ester, and then further undergoes hydrogenolysis to obtain methanol. This is a common feature of the reaction mechanism and is the most obvious difference from the traditional gas-phase methanol synthesis mechanism.

2.2.2. Main Catalyst

At present, there have been many studies on the low-temperature liquid phase methanol synthesis process, and there are various types of catalysts. Generally, the catalyst for the liquid phase methanol process is an inseparable composite system composed of a carbonylation catalyst, hydrogenolysis catalyst, solvent, and additives. It can also be divided into homogeneous and heterogeneous catalytic systems. The homogeneous catalysis system uses alcohol as the solvent and a strong alkaline substance as the carbonylation catalyst. In heterogeneous catalytic systems, Cu catalysts are commonly used as the main active catalysts and alcohol solvents, or inert solvents are used as dispersion media.

Carbonylation Catalyst

Alkali metal methanolates are the most effective carbonylation catalysts at present. Considering factors such as activity and price, the most used are sodium methoxide or potassium methoxide. The outstanding advantages of sodium methoxide catalysts are high activity and good selectivity. However, sodium methoxide catalysts have some disadvantages, such as rapid deactivation and the production of unwanted products (HCOONa, Na2CO3), which may be problematic. Therefore, attempts have been made to find a carbonylation catalyst that can replace sodium methoxide.
The catalytic activity of carbonylation of alkali metal formate is slightly lower than that of sodium methoxide, but it is more stable. Sodium formate has the best catalytic activity [41]. Under reaction conditions of 170 °C, 5 MPa, with a H2 to CO ratio of 2, the single-pass conversion rate of CO can be about 84%. After sodium formate is deactivated, it can be regenerated by the hydrogenation catalyst, so sodium formate is a promising catalyst for carbonylation.
Organic nitrogen-containing heterocyclic olefins are also catalysts for methanol carbonylation (Green and Kiyoshi) [42,43,44,45,46]. Under the reaction conditions of T = 40–80 °C and DBN (1,5-Diazabicyclo [4.3.0]non-5-ene) as a catalyst, the molar yield of methyl formate reached 50% to 80%. Due to the high synthesis cost of DBN, it has not been commercialized, which limits its development and application.
The commonly used carbonylation catalysts mentioned above (CH3ONa or HCOOK) are sensitive to CO2 and H2O in feed gas. Trace amounts of either can quickly deactivate the catalysts. Reactions of sodium methoxide inactivation include:
C H 3 O N a + C O 2   C H 3 O C O O N a
C H 3 O N a + H 2 O   C H 3 O H + N a O H
However, industrial synthesis gas usually contains CO2 and H2O, and the purification cost is high. Therefore, the industrial application of the liquid-phase methanol process has been limited [47].

Hydrogenolysis Catalyst

Highly active hydrogenolysis catalysts can improve the hydrogenolysis rate of the intermediate product ester RCOOCH3, reduce the RCOOCH3 concentration in the product, and increase the selectivity to methanol. Most researched among them are nickel-based hydrogenolysis catalysts and copper-based hydrogenolysis catalysts.

Nickel-Based Catalyst

In 1986, Brookhaven National Laboratory (BNL) [48] first reported the use of Ni-based catalysts to achieve 90% single-pass conversion of CO under reaction conditions of 373 K and 1–5 MPa. The catalyst composition is NaH-RONa-M(OAc)2 (M = Ni, Pd, and Co), in which sodium alkoxide has CO carbonylation activity and acts on the esterification reaction, while the Ni-based catalyst has good hydrogenolysis reaction activity. Mahajan et al. [49,50] synthesized methanol using Ni(CO)4/KOCH3/methanol/glyme catalyst system at 150 °C, 1–3 MPa. The selectivity to methanol can be greater than 95%, and the single-pass conversion rate of syngas exceeds 90%. An inert coolant that is not miscible with methanol was added into the liquid phase at 0–70 °C to overcome the problem of indirect heat exchange on a large number of surfaces required for the reaction mixing process. Ohyama [51,52,53] developed two catalysts, NaH/2-methyl-2-butanol/Ni(CH3COO)2 or Ni(CO)4. Under conditions of 423 K and 5 MPa, the single-pass conversion rate of CO over these catalysts reached 90%, with a methanol selectivity of 99%.
Although the nickel-based catalyst has high activity and selectivity, toxic, volatile, and flammable intermediates, such as Ni(CO)4 and [HNi2(CO)6], are easily produced during the reaction, which reduces their research significance and practical value. Organic nickel/CH3OK catalyst has the same activity and selectivity as the first two catalysts. No toxic intermediates are generated during the reaction [54]. However, removal of residual water in the preparation process of organic nickel is complicated, which increases the catalyst preparation cost. At present, only Amoco and Brookhaven National Laboratory are still working on the exploration and development of nickel-based catalyst systems.

Copper-Based Catalyst

Cu-based catalysts are the most used catalysts in gas phase medium and low-pressure processes, and they are also the key point of methanol synthesis research. This type of catalyst has good activity, high methanol selectivity, low operating pressure, and temperature, so it is an excellent choice for hydrogenation catalysts in liquid-phase, low-temperature methanol processes.
Machionna et al. developed a copper-based catalyst with activity and selectivity comparable to those of nickel-based catalysts [55]. Kokubu et al. used Raney Cu, CH3OK and solvent to synthesize methanol in a liquid phase at 120 °C and 5 MPa [56]. Researchers at the University of Pittsburgh used a slurry-bed reactor to synthesize methanol with Cu-Cr2O3/KOCH3 catalyst under 373–453 K and 5 MPa conditions [57]. All of them obtained high synthesis gas conversion and selectivity.

2.2.3. Solvent

Solvents are also very important in liquid methanol synthesis systems. According to their characteristics, they can be roughly divided into three categories. (1) Inert solvents that do not participate in the reaction process and are purely used as heat carriers, such as paraffin and mineral oil. (2) A mixed solvent system that uses a heat carrier solution together with a solution participating in the catalytic process, such as the p-xylene (heat carrier) and sodium methoxide (carbonylation agent) systems. (3) An alcohol solvent that acts as a dual heat carrier and catalyst. Studies suggest that this type of system can be further divided into two sub-categories, depending on the additional catalyst used. One is early developed; strong alkaline substances, such as sodium methoxide or potassium methoxide, were added to methanol solution as homogeneous catalysts for methanol carbonylation. Methyl formate was hydrogenolyzed by a copper-based catalyst; the second was developed later, without adding strong alkaline substances, such as sodium methoxide, and using only a copper-based catalyst in an alcohol solution (usually ethanol solution) to achieve low-temperature methanol liquid phase synthesis. These three types of solution systems are discussed separately as follows.

Inert Solvent

The principle of this reaction system is the same as that of the gas-phase methanol process. The essence is to disperse the catalyst in an inert solvent that does not participate in the reaction, so no carbonylation catalyst is needed. Paraffin oil, as an inert solvent heat carrier, is one of the more commonly used solutions for studying liquid-phase methanol synthesis. The use of paraffin oil is that, on the one hand, it can transfer the heat released by the reaction over time and control the temperature of the reaction bed within the working range of the catalyst; on the other hand, it facilitates uniform dispersion of the catalyst and reduces the accumulation of reaction heat on catalyst particles, which prevents sintering and promotes uniform mass transfer.
Aiming at the liquid phase methanol synthesis of a paraffin system, Xiaobing et al. studied the effect of co-precipitation conditions on catalytic activity [58]. They found that precipitation and aging processes are the main factors affecting the microstructure of the catalyst. With increased precipitation and aging, the crystallinity of each species in the precursor increases. With increased calcination temperature or time, the average particle size of the microcrystals in the catalyst increases. Thus, the activity and stability of the catalysts are affected.
Wei et al. [59] prepared Cu/ZnO/Al2O3 catalyst in situ in paraffin by a direct liquid phase method and tested its catalytic activity on syngas to lower alcohols at low temperature. The catalyst preparation and reaction were coupled to make the catalyst uniformly dispersed. The effect of carbon nanotubes and carbon microspheres on catalytic activity was also studied. The results showed that the addition of carbon nanospheres significantly reduced the Cu0 crystal size of Cu/ZnO/Al2O3 catalyst and increased the dispersion of Cu components due to the localized effect of the pores, thereby increasing methanol selectivity. Because the large Cu0 crystal size is beneficial to CO insertion and carbon chain growth, the smaller the grain size, the higher the methanol selectivity, and the larger the Cu0 grain size, the higher the ethanol selectivity.
Yuan et al. [60] developed Cu/MgO/ZnO catalyst for low-temperature methanol synthesis in paraffin medium. No carbonylation agent is required, and a CO conversion rate of 63% and methanol selectivity of 99% can be achieved at low temperatures. Studies have shown that the introduction of MgO can promote the formation of Cu+, which is conducive to the dispersion of Cu0 and Cu+ in the carrier and increases the catalytic activity. They also analyzed the deactivation of the catalyst and verified again that low temperatures are desired for methanol synthesis as the main reason for the catalyst deactivation was copper sintering.

Mixed Solvent

In this type of system, one of the solvents will participate in the reaction, so the principle follows the liquid phase methanol process. The representative solvent mixture of this approach is the p-xylene, sodium methoxide, and alcohol mixture system. The carbonylation reaction of methanol is an ionic reaction. When reacting in a non-polar solvent, the reaction rate will increase as the dielectric constant of the solution decreases. Therefore, adding a non-polar solvent with a lower dielectric constant than methanol may increase catalytic activity. It was found that when p-xylene was used as the solvent, the yield and selectivity of methanol were the highest among common solvents.
Wei et al. researched the p-xylene system [61,62]. Aiming at the low activity and poor stability of Cu/Cr2O3 catalyst, Cu/Cr2O3/SiO2 catalyst prepared by the co-precipitation method and Cu/Cr2O3/Al2O3 catalyst prepared by a sol state co-precipitation were developed. For the Cu/Cr2O3/SiO2 catalyst, it was found that the addition of silicon makes the catalyst structure looser, the specific surface area larger, and the Cu reduction temperature lower, thereby improving the reactivity, selectivity, and stability of the Cu/Cr2O3/SiO2 catalyst. The researchers believed that, on one hand, the addition of silica support may increase the hydrogenolysis activity of the catalyst, reduce the concentration of methyl formate in the catalyst system, and inhibit the deactivation of sodium methoxide; on the other hand, the silica support stabilizes the catalyst structure, enhancing the stability of the hydrogenolysis catalyst itself.
When investigating the effect of zirconium additives on the performance of Cu/Cr2O3/SiO2 catalysts, the results showed that zirconium dioxide, as a structural and electronic promoter, had an effect on low-temperature liquid phase methanol synthesis [62]. This is because zirconium additives can increase the specific surface area of the catalyst and promote dispersion of the copper and chromium components in the catalyst. In addition, ZrO2 enables Cu+, Zr4+, and Cr3+ generated on the catalytic surface to form a composite center, which can stabilize Cu+. Additionally, Cu+ plays an important role in the reaction steps, such as H2 activation and CO bond breakage, to improve catalyst activity.
A study on the Cu/Cr2O3/Al2O3 catalyst prepared by the sol state co-precipitation method showed that there was an optimal value for the aluminum content [61]. The addition of aluminum also improved the catalyst activity and methanol selectivity and made the active components in the catalyst more dispersed, more reactive sites on the catalyst surface, increased specific surface area, and enhanced adsorption for raw gas, thereby improving the activity and stability of the catalyst.

Alcohol Solvent

The alcohol solution system for liquid phase methanol synthesis can be subdivided into two types: homogeneous and heterogeneous.
Wender et al. [63] proposed a methanol solution system using sodium methoxide. A mixed catalyst composed of alkali metal alkoxide and copper-chromium catalysts was used to catalyze methanol carbonylation and methyl formate hydrogenolysis, respectively. It was found that when the concentration of alkali metal alkoxide was low, the amount of methanol produced increased with increasing concentrations of methyl formate, but when the concentration of alkali metal alkoxide was high, the rate of methanol synthesis decreased. The researchers believed that a high concentration of alkali metal alkoxides covers the hydrogenolysis active sites on the surface of the copper-chromium catalyst.
Tiansheng et al. [64,65] found that Cu/MgO catalyst could effectively catalyze the hydrogenolysis step. When combined with the esterification catalyst potassium formate, it showed high catalytic activity, reaching a total carbon conversion rate of 91% and 99% methanol selectivity under 150 °C and 5 MPa. The disadvantage is that the composition of the catalyst is complex. Hu et al. [41,66] found that Cu/MgO catalyst modified by alkali metal had better performance. Cu/MgO-Na and sodium formate worked together to increase the carbonylation activity of the catalyst. When the atomic ratio of Cu:Mg:Na was 1:2:1, the catalyst had the highest activity for methanol synthesis, and the single-pass conversion of CO reached 90%. However, CO2 was formed during the reaction, and the stability needed to be improved.
As mentioned above, although sodium methoxide has relatively high catalytic activity for methanol carbonylation, it is easily deactivated due to the presence of trace amounts of CO2 and H2O in the synthesis gas. Tsubaki et al. [67] discovered and proposed that liquid phase methanol production from synthesis gas can be realized with merely copper-based catalysts in an alcohol solution, without strong alkaline substances such as sodium methoxide, as a catalyst for carbonylation. They prepared a Cu/ZnO catalyst at 150 °C and 5 MPa by the co-precipitation method in a slurry-bed reactor with C1–C4 alcohol as the liquid medium, without adding strong alkaline carbonylation reagents. It achieved a CO conversion over 50% and methanol selectivity over 90%. The alcohol solvent not only served as the liquid medium but also participated in the reaction process, but it was not consumed (just like the catalyst). It should be noted that, similar to the ICI process, CO2 (φ = 2–5%) must be used as an inducer in this process. If there is no CO2 in the raw gas, the product methanol cannot be obtained [68]. The inferred reaction process is as follows:
C O + H 2 O   C O 2 + H 2
C O 2 + 1 / 2 H 2 + Cu   H C O O C u
H C O O C u + ROH HCOOR + C u O H
H C O O R + 2 H 2   ROH + C H 3 O H
C u O H + 1 / 2 H 2   Cu + H 2 O
The overall main reaction of steps (7)–(11) is methanol synthesis from carbon monoxide, i.e., Equation (1).
After Tsubaki et al. proposed a pure alcohol catalyst system, Cu-ZnO catalysts have been researched more thoroughly. Reubroycharoen et al. [69,70] found that when the atomic ratio of Cu to Zn is one, the catalyst showed the highest activity. The single-pass conversion of total carbon reached 47%. Bao et al. [71] found that the addition of Zn could improve the pore size distribution of CuO, prevent Cu aggregation caused by reduction, and obtain highly dispersed Cu and Cu-ZnOx particles.
The type of alcohol solvent also has an impact on catalytic activity [72]. Isopropanol has a strong nucleophilic attack ability, which can capture the adsorbed formic acid group and esterify it, and then generate two molecules of alcohol. Zeng et al. [68] prepared Cu/ZnO catalyst by the co-precipitation method and studied 13 kinds of common alcohols as liquid phase methanol synthesis solutions. The research results showed that the influence of the alcohol solution was mainly reflected in the formation rate of the intermediate ester, which also confirmed the reaction process inferred by Tsubaki et al., as mentioned above (solvent alcohol participates in the carbonylation reaction). In contrast to the hydrogenolysis of esters, the carbonylation reaction was considered to be the rate control step. Carbonylation reaction is a nucleophilic reaction [73], so high hydroxyl electronegativity and low steric hindrance facilitate the progress of the reaction. For example, 1-butanol has a lower steric hindrance, but the oxygen atom on the hydroxyl group has a lower electronegativity, making the activity not as high as 2-butanol. The oxygen atom on the hydroxyl group of isobutanol is highly electronegative, but the steric hindrance is large, which makes it less active than 2-butanol. Therefore, the oxygen of 2-butanol has good electronegativity and suitable steric hindrance, and thus exhibits the highest catalytic activity. According to the research of Reubroycharoen et al. [70], it can be found that, as a result of the balance between the electronic effect and the steric effect, when the carbon number of the alcohol is the same, secondary alcohol has the highest activity. For the six primary alcohols, from ethanol to hexanol to benzyl alcohol, the yield of methanol decreased as the carbon number of the alcohol increased.
In the pure alcohol system, since sodium methoxide is not used as a carbonylation catalyst, it can tolerate a certain amount of CO2 and H2O in syngas, which reduces feedstock purification costs. From this point of view, when catalyst activity can be effectively improved, the pure alcohol system has more industrial application prospects.

2.2.4. Additives and Supports

The performance of catalysts largely depends on their composition. In the copper-based methanol synthesis catalyst, copper is considered the main active component, but the catalytic activity of copper alone is relatively poor and thermally unstable. The appropriate addition of other components can greatly improve the performance of the catalyst. Many additives and supports have been mentioned, such as ZnO, Al2O3, ZrO2, MgO, SiO2, and carbon nanotubes. Most additives loosen structure, improve dispersion of Cu components, provide stability, and mitigate thermal sintering. At present, rare-earth and alkaline earth metals and their oxides are commonly used as additives for catalysts as attempts to solve problems by adjusting structure, surface acidity and alkalinity, electronic properties, etc.
Qi et al. [74] found that when TiO2 was added as an additive, the Cu crystal size was reduced, the surface area increased, and catalytic activity was significantly improved. It was also found that TiO2 moved the reduction peak to a lower temperature and promoted the reduction of copper as an electronic promoter.
Cabrera et al. [75] used palladium as an additive of a copper-based catalyst. They found that Pd could significantly improve the reduction properties of Cu and increase the rate of the methanol synthesis reaction due to its excellent ability to dissociate and overflow hydrogen.
Chen et al. [76] believed that Mn additives could improve the activity of CuO/ZnO/Al2O3 catalysts and adjust the ratio of Cu+/Cu0, therefore increasing the number of active centers and making Cu2+ reduction harder. The addition of Mn additives could also increase the dispersion of Cu on catalyst surfaces and stabilize the active components, thereby improving the resistance to sintering.
Alkali metal salts can catalyze the carbonylation reaction of methanol and increase the conversion rate of feed gas. For example, the catalytic activity of Cu-MnOx alone is low, but the addition of K2CO3 can improve the activity of the entire reaction system [64]. The conversion of total carbon reached 90.2%, and the methanol selectivity was about 99.1%.
Chen et al. [77] believed that Al, Sc, and Cr could cause cation defects on the ZnO surface and increase the concentration and stability of Cu+ during catalyst activation and the reaction process. Especially after adding Sc, the Cu/ZnO catalyst had a larger specific surface area, which was conducive to the production of methanol.
Bell et al. [78] believed that adding Ce to Cu/ZrO2 catalyst could improve the performance of the catalyst. At 473–523 K and 3 MPa, the methanol synthesis activity of Cu/Ce0.3Zr0.7O2 is 2.7 times higher than that of Cu/m-ZrO2.

2.2.5. Preparation Method for Catalyst

The preparation method of the catalyst can directly affect the physical and chemical properties of the catalysts, such as the crystal size, dispersion degree, and pore structure. The preparation of Cu-based catalysts for methanol synthesis includes precipitation, impregnation, ball milling, and sol-gel methods.

Co-Precipitation Method

The co-precipitation method is the most used method for preparing copper-based methanol synthesis catalysts. A common process of Cu/ZnO catalyst preparation by co-precipitation is shown in Figure 3. By adjusting the precipitation temperature and pH value, uniformly dispersed metal cations in solution can be converted into organometal/oxide precursors through rapid solidification. After calcination and reduction, highly dispersed and well-mixed solid metal/oxide catalyst are formed. The preparation process of the precipitation method includes precipitation, aging, filtration and washing, as well as drying, calcining, granulation, and reduction. The adjustment of each step affects the physical structure and chemical properties of the product catalyst.
In the process of co-precipitation, the control of pH and temperature is extremely important. Fujita et al. [79] believed that due to the exothermic reduction reaction of CuO in H2, the local temperature of the CuO particles would increase during the reduction process, leading to the aggregation and growth of CuO crystal particles. Therefore, keeping the reduction temperature lower can control the Cu crystal sizes to be smaller. Through sodium carbonate titration, Behrens et al. [80] found that when the pH value was low, Zn2+ would stay in the liquor and could not be precipitated. The optimal precipitation pH was the lowest pH for the complete precipitation of Zn2+ (pH = 7). When the precipitation temperature was 60 °C and pH = 7, after calcining and reduction, the catalyst components obtained were highly dispersed, which is conducive to high Cu-Zn catalyst activity. Jeong et al. [81] used the co-precipitation method to synthesize a series of Cu/ZnO catalysts for low-temperature methanol synthesis with a pH value of 6–10. They found that the catalyst produced at pH = 8 had the smallest copper crystals, the largest number of strong acid sites on the surface, and the highest catalyst activity. The catalytic activity was positively correlated with the number of strong acid sites on the catalyst surface. The strong acid sites on the surface helped in the adsorption and activation of CO and increased the conversion rate of CO at low temperatures.
Paul et al. [82] prepared zincian georgeite (an amorphous copper–zinc hydroxycarbonate), the precursor of their Cu/ZnO catalyst, by co-precipitation using acetate salts and ammonium carbonate. They attributed the superior performance of their catalysts to the exclusion of alkali metals from the synthesis procedure, which are known to act as catalyst poisons.
Precipitation methods include direct precipitation, co-precipitation, and complex precipitation [62,79]. Catalysts prepared by this method have high dispersion, small crystal grains, and strong interaction between the components, but the processes are complicated, and sometimes the strong interaction is not conducive to the reduction of the catalyst.

Sol-Gel Method

The sol-gel method is a common method for preparing nanocatalysts. Metal ions and organic ligands are uniformly dispersed in a solvent, and then hydrolyzed and condensed to form a sol. The colloidal particles in the sol polymerize during the aging process to form a gel with a three-dimensional network structure. After calcination, reduction, and passivation of the gel, a product catalyst can be obtained.
Wei et al. [61] used the impregnation method, co-precipitation method, and sol-gel method to prepare Cu/Cr2O3/SiO2 catalysts and to evaluate their activity. The results showed that the catalyst prepared by the sol-gel method had the best catalytic activity and methanol selectivity. Ali et al. [83] added the surfactant sodium dodecyl sulfonate (SDS) to the gel-forming process to obtain MgO particles with an average particle size of 13.2 nm. However, the addition of the surfactant introduced sulfur, which was toxic to the catalyst.
Citric acid (C6H8O7∙H2O) is a commonly used chelating agent, which contains three (-COOH) groups available for complexation. Citric acid can chelate with Cu2+ ions to form a sol, which can be transformed into a gel after heating, evaporation, dehydration, and condensation. Chelated citric acid in gel undergoes self-combustion at about 200 °C. Combustion consumes C, O, and H elements to generate a large amount of gas. Therefore, the prepared catalyst has a larger surface area and pore volume. Lei et al. [84] prepared Cu/ZnO catalyst by the citric acid sol-gel combustion method and found that the gel of citric acid during calcination in Ar atmosphere could generate a large amount of gas. The reducing gas, such as CH4 and CO, could, in situ, reduce the calcination product CuO to Cu, so the subsequent additional reduction treatment was no longer needed. The activity was highest when the stoichiometric ratio of citric acid to Cu was 0.8.
Because the gel has a three-dimensional network structure, it can maintain a certain structure and morphology after calcination, so the catalyst prepared by the sol-gel method has a high dispersion of active metals and a smaller particle size. However, the disadvantage is that the organic ligands are expensive and the preparation conditions are harsh.

Other Synthesis Methods

There are reports of other catalyst preparation methods, but the reports are also very limited, and each presents problems.
The physical pulverization method prepares nanoparticles by mechanical pulverization, electric sparks, or other methods. Ball mill, colloid mill, and other methods of size reduction are commonly used. Huang et al. [85] used a ball mill method to grind and mix the powders of copper and zinc oxide in air or vacuum. They found that mixing in air was better than mixing in a vacuum. Because the partial oxidation of copper in the air enhanced the interaction with zinc oxide, the resulting catalyst had a higher copper surface area. The preparation process of this method was simple and the cost was low, but high-strength mechanical mixing can destroy internal catalyst structures, introduce impurities, reduce the purity of the catalyst, and controlling the particle size distribution is difficult.
Impregnation involves the immersion of the solid carrier into a salt solution containing active components, followed by separation of the residual liquid after a certain period. In this way, active components are attached to the solid in the form of ions or compounds, and when the impregnation is balanced, the remaining liquid is removed, followed by similar drying, calcination, and other post-treatments. The active component salts can be left on the inner surface of the carrier. The salts of these metals or metal oxides are evenly distributed in the pores of the carrier. After being decomposed and activated by heating, a highly dispersed catalyst is obtained. Chappell et al. [86] prepared Cu/SiO2 catalyst by the impregnation method using SiO2 as a carrier and added Al, Zr, Ga, Zn, and other metal additives to improve its performance. The impregnation method has the advantages of simple operation, low preparation cost, and a high single batch yield. The prepared catalyst also has good dispersity, moderate inner interaction, and good catalytic performance, but it has high requirements for the specific surface, thermal stability, and electronic effects of the support.
The flame combustion synthesis method is a general method for the preparation of metal oxide catalysts with a high specific surface area. Burner design and combustion conditions determine the specific surface area, structure, and phase composition of the catalyst. Jensen et al. [87] mixed acetyl pyruvate of volatile Cu, Zn, Al precursors with fuel (H2 and CH4), and air for high temperature combustion and gasification, resulting in decomposition, condensation, rearrangement, and rapid cooling to obtain catalysts. Moreover, the metal particles could be more fully mixed in the gas phase, which increased the interaction between the components. The specific surface, morphology, and crystal phase structure of the particles can be changed by adjusting the flame temperature, the residence time in the high-temperature zone, and the solution concentration to improve the activity, selectivity, and stability of the catalyst. Methanol synthesis is a structure-sensitive reaction over copper-based material, and in the traditional precipitation method it is difficult to control the structure during the calcination process, which leads to losses of the active surface area. The flame combustion synthesis method offers better control of the structure and thus limits surface area losses.

2.2.6. Reactor Types

Unlike the gas-solid two-phase reaction of the gas-phase synthesis process, the liquid-phase synthesis method is a gas-liquid-solid three-phase reaction, so the reactor must be adjusted accordingly. At present, the most used reactors in liquid phase methanol synthesis are slurry beds and trickle beds.

Slurry Bed Process

The three-phase slurry bed reactor is like a bubbler, with a heat exchanger placed in the bed. In the reactor, a paraffin-like long-chain hydrocarbon compound with a high heat capacity and large thermal conductivity is added as an inert medium, and the fine powder catalyst is dispersed in a liquid phase medium to form a slurry and placed in the reactor. At the beginning of the reaction, syngas must be dissolved and dispersed in an inert oil medium to reach the surface of the catalyst, and the product after the reaction must undergo a similar process before it can be collected. The reactor can be batch or continuous, and a single reactor or multiple reactors can be connected in series. The general process is that the purified and compressed synthesis gas flows through the reactor from bottom to top and forms a three-phase bubbling bed with the slurry, where mass transfer occurs between the phases and reacts on the surface of the catalyst. Output gas is discharged from the upper part of the reactor, where reaction heat is absorbed by circulating inert hot liquid, which is exchanged via heat exchanger, so as to maintain the bed temperature, while producing steam as a by-product.
Since the liquid inside the reactor is in a fully mixed flow state, the reaction temperature is uniform, and temperature control is simple and effective. At the same time, the average temperature of the slurry-bed reactor is much higher than that of the tubular fixed-bed reactor, so it has a higher reaction rate and product selectivity. The reaction pressure is reduced, which can reduce gas compression costs. Additionally, replacement and addition of catalysts during operation is possible, greatly improving the utilization of the equipment, which is a significant improvement over the tubular fixed-bed reactor, which must be shut down often to replace the catalyst. At the same time, it also has the advantages of simple operation and easy separation of solid and liquid.
The slurry-bed reactor has a simple structure, uniform dispersion, good heat transfer performance, and convenient operation. However, catalyst sedimentation and aggregation in the reactor may occur. Therefore, catalyst utilization efficiency may be low, and overall efficiency may not be high.

Trickle Bed Process

In 1990, Pass et al. proposed a trickle bed method to synthesize methanol [88]. The structure of the trickle bed reactor is similar to that of the traditional gas phase fixed bed reactor. The fixed layer is composed of catalysts with larger particles. Liquid flows from top to bottom in the form of droplets, and gas also flows from top to bottom. The gas and liquid are distributed among the catalyst particles.
The trickle bed has the advantages of both a slurry bed and a fixed bed. Similar to a fixed bed, it has a large catalyst loading and no abrasion. The reactor is equipped with heat exchangers, making material flow in the catalyst bed close to plug flow without back mixing. The trickle bed has a high conversion and isothermal reaction similar to the slurry bed and is more suitable for synthesis gas with a low hydrogen to carbon monoxide ratio. From an industrial point of view, the liquid phase fluid in the trickle bed contains very little catalyst powder, and the conveying equipment is easy to seal and has little abrasion. Therefore, it may be more reliable for long-term operations.

2.3. CO2 and H2 Methanol Synthesis

In addition to the mature industrialized gas phase method and the hot liquid phase method currently being studied, there is also a CO2/H2 synthesis process for methanol production. However, compared to the CO-synthesis process, CO2 process is less reactive, requires a larger reactor, and is not yet mature, so only a brief introduction is given here [89].
In the early 1980s, Topsoe used CO2 and H2 contained in the exhaust gas of a refinery to obtain methanol successfully through the adiabatic reaction of the catalyst. Then, the Tokyo Gas Company synthesized methanol with CO2 and H2 using the Cu/ZnO/Al2O3 catalyst. Lurgi and Southern Chemical Corporation used CO2 and H2 to synthesize methanol with a low-pressure catalytic system. Both showed that although CO2/H2 methanol synthesis process reduced the circulation of the system, the methanol synthesis rate was low. Although various highly active metal catalysts have been reported for CO2 hydrogenation to methanol, the methanol selectivity of most of them is still below 60%. However, the CO byproduct can still be reused, so it still has industrial application prospects. In addition, the reaction mechanism of catalytic CO2 hydrogenation to methanol is still controversial and needs to be studied further [90,91,92].
It has been reported that a new type of Ni-Ga bimetallic catalyst has good performance of CO2 hydrogenation to methanol [93]. The catalyst can complete the CO2 hydrogenation process under atmospheric pressure and has good stability. In2O3-based catalysts are also promising in the CO2 hydrogenation to methanol and are currently receiving a lot of attention [94]. These catalysts are still in the optimization stage, and the catalytic technology is not yet mature. However, this process has great potential in terms of operating conditions and methanol synthesis performance. Therefore, the development of a suitable catalyst is the key to the industrialization of methanol synthesis from direct CO2 hydrogenation in the future.

3. Summary and Prospect

The gas phase method for synthesizing methanol is commercially mature, and the high-pressure method has been eliminated due to high energy consumption, large investment in equipment, and poor product quality. Medium- and low-pressure methods are now widely used. However, the traditional gas phase methanol synthesis process requires a high reaction temperature and is limited by thermodynamics, resulting in a series of disadvantages such as low synthesis efficiency, high energy consumption, and high cost. In view of the many shortcomings of the gas-phase process, a low-temperature and high-conversion liquid-phase methanol synthesis process has emerged. In the low-temperature liquid phase methanol synthesis process, the catalyst is a composite system in which the main catalyst, solvent, carrier, and additive are all important. Currently, however, each has limitations. Therefore, finding a catalytic system that has high catalytic performance and high selectivity but is also stable and does not easily deactivate, and can be recycled and regenerated, is the focus of ongoing research. The synthesis method of the catalyst and the type of reactor will also affect methanol synthesis. Therefore, their optimization and development are also future research directions. The development of the liquid phase methanol synthesis process into commercial applications, thus, requires significant further research.

Author Contributions

Conceptualization, B.W. and G.L.; investigation, G.L.; resources, B.W.; writing—original draft preparation, G.L.; writing—review and editing, B.W. and H.H.-W.; visualization, G.L. and H.H.-W.; supervision, B.W. and H.H.-W.; project administration, B.W.; funding acquisition, B.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded in part by the Florida Agricultural Experiment Station and the Consortium for Waste Circularity.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were collected.

Acknowledgments

The authors are grateful for support from the Consortium for Waste Circularity (wastecircularity.org) and the Florida Agricultural Experiment Station.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Roode-Gutzmer, Q.I.; Kaiser, D.; Bertau, M. Renewable Methanol Synthesis. ChemBioEng Rev. 2019, 6, 209–236. [Google Scholar] [CrossRef]
  2. Brunetti, A.; Migliori, M.; Cozza, D.; Catizzone, E.; Giordano, G.; Barbieri, G. Methanol Conversion to Dimethyl Ether in Catalytic Zeolite Membrane Reactors. ACS Sustain. Chem. Eng. 2020, 8, 10471–10479. [Google Scholar] [CrossRef]
  3. Tian, P.; Wei, Y.; Ye, M.; Liu, Z. Methanol to Olefins (MTO): From Fundamentals to Commercialization. ACS Catal. 2015, 5, 1922–1938. [Google Scholar] [CrossRef]
  4. Keil, F.J. Methanol-to-hydrocarbons: Process technology. Microporous Mesoporous Mater. 1999, 29, 49–66. [Google Scholar] [CrossRef]
  5. Wang, T.; Tang, X.; Huang, X.; Qian, W.; Cui, Y.; Hui, X.; Yang, W.; Wei, F. Conversion of methanol to aromatics in fluidized bed reactor. Catal. Today 2014, 233, 8–13. [Google Scholar] [CrossRef]
  6. Kohli, K.; Sharma, B.K.; Panchal, C.B. Dimethyl Carbonate: Review of Synthesis Routes and Catalysts Used. Energies 2022, 15, 5133. [Google Scholar] [CrossRef]
  7. Din, I.U.; Shaharun, M.S.; Alotaibi, M.A.; Alharthi, A.I.; Naeem, A. Recent developments on heterogeneous catalytic CO2 reduction to methanol. J. CO2 Util. 2019, 34, 20–33. [Google Scholar] [CrossRef]
  8. Khadzhiev, S.N.; Kolesnichenko, N.V.; Ezhova, N.N. Slurry technology in methanol synthesis (Review). Pet. Chem. 2016, 56, 77–95. [Google Scholar] [CrossRef]
  9. Marie-Rose, S.; Lemieux, A.; Lavoie, J.-M. Conversion of Non-Homogeneous Biomass to Ultraclean Syngas and Catalytic Conversion to Ethanol. In Biofuel’s Engineering Process Technology; IntechOpen: London, UK, 2011. [Google Scholar] [CrossRef] [Green Version]
  10. Tijm, P.J.A.; Waller, F.J.; Brown, D.M. Methanol technology developments for the new millennium. Appl. Catal. Gen. 2001, 221, 275–282. [Google Scholar] [CrossRef]
  11. Bozzano, G.; Manenti, F. Efficient methanol synthesis: Perspectives, technologies and optimization strategies. Prog. Energy Combust. Sci. 2016, 56, 71–105. [Google Scholar] [CrossRef]
  12. Kim, W.-Y.; Hayashi, H.; Kishida, M.; Nagata, H.; Wakabayashi, K. Methanol synthesis from syngas over supported palladium catalysts prepared using water-in-oil microemulsion. Appl. Catal. A Gen. 1998, 169, 157–164. [Google Scholar] [CrossRef]
  13. Mériaudeau, P.; Albano, K.; Naccache, C. Promotion of Platinum-based Catalysts for Methanol Synthesis from Syngas. J. Chem. Soc. Faraday Trans. I 1987, 83, 2113–2118. [Google Scholar] [CrossRef]
  14. Yang, C.; Ma, Z.; Zhao, N.; Wei, W.; Hu, T.; Sun, Y. Methanol synthesis from CO2-rich syngas over a ZrO2 doped CuZnO catalyst. Catal. Today 2006, 115, 222–227. [Google Scholar] [CrossRef]
  15. Liu, L.; Fan, F.; Bai, M.; Xue, F.; Ma, X.; Jiang, Z.; Fang, T. Mechanistic study of methanol synthesis from CO2 hydrogenation on Rh-doped Cu(111) surfaces. Mol. Catal. 2019, 466, 26–36. [Google Scholar] [CrossRef]
  16. Song, J.-R.; Wen, L.-X.; Shao, L.; Chen, J.-F. Preparation and characterization of novel Pd/SiO2 and Ca–Pd/SiO2 egg-shell catalysts with porous hollow silica. Appl. Surf. Sci. 2006, 253, 2678–2684. [Google Scholar] [CrossRef]
  17. Dry, R.J. Possibilities for the Development of Large-Capacity Methanol Synthesis. Ind. Eng. Chem. Res. 1988, 27, 616–624. [Google Scholar] [CrossRef]
  18. Joo, O.S.; Jung, K.D.; Moon, I.; Rozovskii, A.Y.; Lin, G.I.; Han, S.H.; Uhm, S.J. Carbon Dioxide Hydrogenation To Form Methanol via a Reverse-Water-Gas-Shift Reaction (the CAMERE Process). Ind. Eng. Chem. Res. 1999, 38, 1808–1812. [Google Scholar] [CrossRef]
  19. Dybkjaer, I. Topsoe Methanol Technology. Chem. Econ. Eng. Rev. 1981, 6, 17–25. [Google Scholar]
  20. Klier, K. Methanol Synthesis. Adv. Catal. 1982, 31, 243–313. [Google Scholar]
  21. Denise, B.; Sneeden, R.P.A.; Hamon, C. Hamon Hydrocondensation of carbon dioxide: IV. J. Mol. Catal. 1982, 17, 359–366. [Google Scholar] [CrossRef]
  22. Chinchen, G.C.; Denny, P.J.; Parker, D.G.; Spencer, M.S.; Whan, D.A. Mechanism of methanol synthesis from CO2/CO/H2 mixtures over copper/zinc oxide/alumina catalysts: Use of14C-labelled reactants. Appl. Catal. 1987, 30, 333–338. [Google Scholar] [CrossRef]
  23. Zhang, Y.; Sun, Q.; Deng, J.; Wu, D.; Chen, S. A high activity Cu/ZnO/A1203 catalyst for methanol synthesis: Preparation and catalytic properties. Appl. Catal. A Gen. 1997, 158, 105–120. [Google Scholar] [CrossRef]
  24. Laudenschleger, D.; Ruland, H.; Muhler, M. Identifying the nature of the active sites in methanol synthesis over Cu/ZnO/Al2O3 catalysts. Nat. Commun. 2020, 11, 3898. [Google Scholar] [CrossRef]
  25. Sehested, J. Industrial and scientific directions of methanol catalyst development. J. Catal. 2019, 371, 368–375. [Google Scholar] [CrossRef]
  26. Waugh, K.C. Methanol Synthesis. Catal. Today 1992, 15, 51–75. [Google Scholar] [CrossRef]
  27. Chinchen, G.C.; Waugh, K.C.; Whan, D.A. Whan The activity and state of the copper surface in methanol synthesis catalysts. Appl. Catal. 1986, 25, 101–107. [Google Scholar] [CrossRef]
  28. Robbins, J.L. Methanol synthesis over Cu/SiO2 catalysts. Catal. Lett. 1991, 10, 1–10. [Google Scholar] [CrossRef]
  29. Fleisch, T.H.; Fleisch, R.L.M. Studies on the chemical state of Cu during methanol synthesis. J. Catal. 1984, 90, 165–172. [Google Scholar] [CrossRef]
  30. Herman, R.G.; Herman, K.K.; Simmons, G.; Finn, B. Catalytic synthesis of methanol from COH2: I. Phase composition, electronic properties, and activities of the Cu/ZnO/M2O3 catalysts. J. Catal. 1979, 56, 407–429. [Google Scholar] [CrossRef]
  31. Duprez, D.; Ferhat-Hamida, Z.; Bettahar, M.M. Surface mobility and reactivity of oxygen species on a copper-zinc catalyst in methanol synthesis. J. Catal. 1990, 124, 1–11. [Google Scholar] [CrossRef]
  32. Chu, P.J.; Gerstein, B.C.; Sheffer, G.R.; King, T.S. NMR studies of 65Cu and 133Cs in alkali-metal-promoted copper catalysts. J. Catal. 1989, 115, 194–204. [Google Scholar] [CrossRef]
  33. Denise, B.; Sneeden, R.P.A.; Beguin, B.; Cherifi, O. Supported copper catalysts in the synthesis of methanol: N2O-titrations. Appl. Catal. 1987, 30, 353–363. [Google Scholar] [CrossRef]
  34. Cybulski, A.; Edvinsson, R.; Irandoust, S.; Andersson, B. Liquid-phase methanol synthesis: Modelling of a monolithic reactor. Chem. Eng. Sci. 1993, 48, 3463–3478. [Google Scholar] [CrossRef]
  35. Hu, B.; Fujimoto, K. Low Temperature Methanol Synthesis in Slurry Phase With a Hybrid Copper-Formate System. Catal. Lett. 2009, 129, 416–421. [Google Scholar] [CrossRef]
  36. Zeng, J.-Q.; Tsubaki, N.; Fujimoto, K. The promoting effect of alcohols in a new process of low-temperature synthesis of methanol from CO/CO2/H2. Fuel 2002, 81, 125–127. [Google Scholar] [CrossRef]
  37. Yang, R.; Fu, Y.; Zhang, Y.; Tsubaki, N. In situ DRIFT study of low-temperature methanol synthesis mechanism on Cu/ZnO catalysts from CO2-containing syngas using ethanol promoter. J. Catal. 2004, 228, 23–35. [Google Scholar] [CrossRef]
  38. Yang, R.; Zhang, Y.; Iwama, Y.; Tsubaki, N. Mechanistic study of a new low-temperature methanol synthesis on Cu/MgO catalysts. Appl. Catal. A Gen. 2005, 288, 126–133. [Google Scholar] [CrossRef]
  39. Zhang, Y.; Yang, R.; Tsubaki, N. A new low-temperature methanol synthesis method: Mechanistic and kinetics study of catalytic process. Catal. Today 2008, 132, 93–100. [Google Scholar] [CrossRef]
  40. Yang, R.; Zhang, Y.; Tsubaki, N. Rideal-type reaction of formate species with alcohol: A key step in new low-temperature methanol synthesis method. Catal. Commun. 2007, 8, 1829–1833. [Google Scholar] [CrossRef]
  41. Hu, B.; Fujimoto, K. High-performance Cu/MgO–Na catalyst for methanol synthesis via ethyl formate. Appl. Catal. A Gen. 2008, 346, 174–178. [Google Scholar] [CrossRef]
  42. EP 0104875; Green, M.J. Process for the Production of Formates. European Patent Office: Munich, Germany, 1984.
  43. EP 0115387; Green, M.J. Decarbonylation Process. European Patent Office: Munich, Germany, 1984.
  44. EP 0107441; Green, M.J. Process for the Preparation of Formamides. European Patent Office: Munich, Germany, 1984.
  45. EP 0168167; Green, M.J. Catalysis Using Supported Strong Base Catalysts. European Patent Office: Munich, Germany, 1986.
  46. Kudo, K. Kinetic Study of Methyl Formate Synthesis from Methanol and Carbon Monoxide Catalyzed by 1, 8-Diazabicyclo [5. 4. 0] undec-7-ene. Nippon Kagaku Kaishi 1976, 4, 457–465. [Google Scholar]
  47. He, L.; Liu, H.; Xiao, C.; Kou, Y. Liquid-phase synthesis of methyl formate via heterogeneous carbonylation of methanol over a soluble copper nanocluster catalyst. Green Chem. 2008, 10, 619–622. [Google Scholar] [CrossRef]
  48. US 4619946; Sapienza, R.S.; Slegeir, W.A.; O’Hare, T.E.; Mahajan, D. Low Temperature Catalysts for Methanol Production. U.S. Patent and Trademark Office: Alexandria, Egypt, 1986.
  49. Mahajan, D. Atom-economical reduction of carbon monoxide to methanol catalyzed by soluble transition metal complexes at low temperatures. Top. Catal. 2005, 32, 209–214. [Google Scholar] [CrossRef]
  50. US 4992480; Mahajan, D.; Sapienza, R.S.; Slegeir, W.A.; O’Hare, T.E. Homogeneous Catalyst Formulations for Methanol Production. U.S. Patent and Trademark Office: Alexandria, Egypt, 1990.
  51. Ohyama, S. Low-temperature methanol synthesis in catalytic systems composed of nickel compounds and alkali alkoxides in liquid phases. Appl. Catal. A Gen. 1999, 180, 217–225. [Google Scholar] [CrossRef]
  52. Ohyama, S. In situ FTIR study on reaction pathways in Ni(CO)4/CH3OK catalytic system for low-temperature methanol synthesis in a liquid medium. Appl. Catal. A Gen. 2001, 220, 235–242. [Google Scholar] [CrossRef]
  53. Ohyama, S. Transformation of the nickel precursor in catalytic systems for low-temperature methanol synthesis in liquid phase. Appl. Catal. A Gen. 1999, 181, 87–93. [Google Scholar] [CrossRef]
  54. Lee, E.S.; Aika, K.I. Low-temperature methanol synthesis in liquid-phase with a Raney Nickel–alkoxide system: Effect of Raney Nickel pretreatment and reaction conditions. J. Mol. Catal. A Chem. 1999, 141, 241–248. [Google Scholar] [CrossRef]
  55. EP 0504981; Marchionna, M.; Lami, M. Catalyst System and Process for the Liquid-Phase Production of Methanol from Synthesis Gas. European Patent Office: Munich, Germany, 1992.
  56. US 6028119; Kokubu, Y. Process for Methanol Production. U.S. Patent and Trademark Office: Alexandria, Egypt, 2000.
  57. Vishwesh, M.; Palekar, H.J.; John, W. Tiemey and Irving Wender Slurry phase synthesis of methanol with a potassium methoxide/copper chromite catalytic system. Appl. Catal. A Gen. 1993, 102, 13–34. [Google Scholar]
  58. Zhang, X.; Zhong, L.; Guo, Q.; Fan, H.; Zheng, H.; Xie, K. Influence of the calcination on the activity and stability of the Cu/ZnO/Al2O3 catalyst in liquid phase methanol synthesis. Fuel 2010, 89, 1348–1352. [Google Scholar] [CrossRef]
  59. Huang, W.; Li, W.; Sun, J.; Yin, L. Effect of Carbon Promoters on the Performance of Cu-Zn-Al Catalyst Prepared by Complete Liquid-Phase Technology. Chin. J. Catal. (Chin. Version) 2011, 31, 1393–1398. [Google Scholar] [CrossRef]
  60. Li, Y. Study on Cu/MgO/ZnO Catalyst for Low-temperature Synthesis of Methanol. Chin. J. Process Eng. 2010, 10, 781–787. [Google Scholar]
  61. Chu, W.; Zhang, T.; He, C.; Wu, Y. Low-temperature methanol synthesis (LTMS) in liquid phase on novel copper-based catalysts. Catal. Lett. 2002, 79, 129–132. [Google Scholar] [CrossRef]
  62. Huang, L.; Chu, W.; Long, Y.; Ci, Z.; Luo, S. Influence of Zirconia Promoter on Catalytic Properties of Cu–Cr–Si Catalysts for Methanol Synthesis at High CO Conversion in Slurry Phase. Catal. Lett. 2006, 108, 113–118. [Google Scholar] [CrossRef]
  63. Palekar, V.M.; Tierney, J.W.; Wender, I. Alkali compounds and copper chromite as low-temperature slurry phase methanol catalysts. Appl. Catal. A Gen. 1993, 103, 105–122. [Google Scholar] [CrossRef]
  64. Zhao, T.-S.; Yoneyama, Y.; Fujimoto, K.; Yamane, N.; Fujimoto, K.; Tsubaki, N. Promotional Effect of Potassium Salt in Low-temperature Formate and Methanol Synthesis from CO/CO2/H2 on Copper Catalyst. Chem. Lett. 2007, 36, 734–735. [Google Scholar] [CrossRef]
  65. Zhao, T.-S.; Zhang, K.; Chen, X.; Ma, Q.; Tsubaki, N. A novel low-temperature methanol synthesis method from CO/H2/CO2 based on the synergistic effect between solid catalyst and homogeneous catalyst. Catal. Today 2010, 149, 98–104. [Google Scholar] [CrossRef]
  66. Hu, B.; Fujimoto, K. Promoting behaviors of alkali compounds in low temperature methanol synthesis over copper-based catalyst. Appl. Catal. B Environ. 2010, 95, 208–216. [Google Scholar] [CrossRef]
  67. Tsubaki, N.; Ito, M.; Fujimoto, K. A New Method of Low-Temperature Methanol Synthesis. J. Catal. 2001, 197, 224–227. [Google Scholar] [CrossRef]
  68. Zeng, J.; Fujimoto, K.; Tsubaki, N. A New Low-Temperature Synthesis Route of Methanol:  Catalytic Effect of the Alcoholic Solvent. Energy Fuels 2002, 16, 83–86. [Google Scholar] [CrossRef]
  69. Reubroycharoen, P.; Yamagami, T.; Vitidsant, T.; Yoneyama, Y.; Ito, M.; Tsubaki, N. Continuous Low-Temperature Methanol Synthesis from Syngas Using Alcohol Promoters. Energy Fuels 2003, 17, 817–821. [Google Scholar] [CrossRef]
  70. Reubroycharoen, P.; Vitidsant, T.; Yoneyama, Y.; Tsubaki, N. Development of a new low-temperature methanol synthesis process. Catal. Today 2004, 89, 447–454. [Google Scholar] [CrossRef]
  71. Bao, J.; Liu, Z.; Zhang, Y.; Tsubaki, N. Preparation of mesoporous Cu/ZnO catalyst and its application in low-temperature methanol synthesis. Catal. Commun. 2008, 9, 913–918. [Google Scholar] [CrossRef]
  72. Yang, R.; Yu, X.; Zhang, Y.; Li, W.; Tsubaki, N. A new method of low-temperature methanol synthesis on Cu/ZnO/Al2O3 catalysts from CO/CO2/H2. Fuel 2008, 87, 443–450. [Google Scholar] [CrossRef]
  73. Xu, B.; Yang, R.; Meng, F.; Reubroycharoen, P.; Vitidsant, T.; Zhang, Y.; Yoneyama, Y.; Tsubaki, N. A New Method of Low Temperature Methanol Synthesis. Catal. Surv. Asia 2009, 13, 147–163. [Google Scholar] [CrossRef]
  74. Qi, G.X.; Zheng, X.M.; Fei, J.H.; Hou, Z.Y. Low-temperature methanol synthesis catalyzed over Cu/γ-Al2O3–TiO2 for CO2 hydrogenation. Catal. Lett. 2001, 72, 191–196. [Google Scholar] [CrossRef]
  75. Melián-Cabrera, I.; López Granados, M.; Fierro, J.L.G. Reverse Topotactic Transformation of a Cu–Zn–Al Catalyst during Wet Pd Impregnation: Relevance for the Performance in Methanol Synthesis from CO2/H2 Mixtures. J. Catal. 2002, 210, 273–284. [Google Scholar] [CrossRef]
  76. Chen, H.Y.; Lin, J.; Tan, K.L.; Li, J. Comparative studies of manganese-doped copper-based catalysts: The promoter effect of Mn on methanol synthesis. Appl. Surf. Sci. 1998, 126, 323–331. [Google Scholar] [CrossRef]
  77. Chen, H.-B.; Liao, D.-W.; Yu, L.-J.; Lin, Y.-J.; Yi, J.; Zhang, H.-B.; Tsai, K.-R. Influence of trivalent metal ions on the surface structure of a copper-based catalyst for methanol synthesis. Appl. Surf. Sci. 1999, 147, 85–93. [Google Scholar] [CrossRef]
  78. Pokrovski, K.A.; Rhodes, M.D.; Bell, A.T. Effects of cerium incorporation into zirconia on the activity of Cu/ZrO2 for methanol synthesis via CO hydrogenation. J. Catal. 2005, 235, 368–377. [Google Scholar] [CrossRef] [Green Version]
  79. Fujita, S.-i.; Moribe, S.; Kanamori, Y.; Kakudate, M.; Takezawa, N. Preparation of a coprecipitated Cu/ZnO catalyst for the methanol synthesis from CO2—effects of the calcination and reduction conditions on the catalytic performance. Appl. Catal. A Gen. 2001, 207, 121–128. [Google Scholar] [CrossRef]
  80. Behrens, M.; Brennecke, D.; Girgsdies, F.; Kißner, S.; Trunschke, A.; Nasrudin, N.; Zakaria, S.; Idris, N.F.; Hamid, S.B.A.; Kniep, B. Understanding the complexity of a catalyst synthesis: Co-precipitation of mixed Cu,Zn,Al hydroxycarbonate precursors for Cu/ZnO/Al2O3 catalysts investigated by titration experiments. Appl. Catal. A Gen. 2011, 392, 93–102. [Google Scholar] [CrossRef] [Green Version]
  81. Jeong, Y.; Kim, I.; Kang, J.Y.; Jeong, H.; Park, J.K.; Park, J.H.; Jung, J.C. Alcohol-assisted low temperature methanol synthesis from syngas over Cu/ZnO catalysts: Effect of pH value in the co-precipitation step. J. Mol. Catal. A Chem. 2015, 400, 132–138. [Google Scholar] [CrossRef]
  82. Smith, P.J.; Kondrat, S.A.; Chater, P.A.; Yeo, B.R.; Shaw, G.M.; Lu, L.; Bartley, J.K.; Taylor, S.H.; Spencer, M.S.; Kiely, C.J. A new class of Cu/ZnO catalysts derived from zincian georgeite precursors prepared by co-precipitation. Chem. Sci. 2017, 8, 2436–2447. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Mashayekh-Salehi, A.; Moussavi, G.; Yaghmaeian, K. Preparation, characterization and catalytic activity of a novel mesoporous nanocrystalline MgO nanoparticle for ozonation of acetaminophen as an emerging water contaminant. Chem. Eng. J. 2017, 310, 157–169. [Google Scholar] [CrossRef]
  84. Shi, L.; Tao, K.; Yang, R.; Meng, F.; Xing, C.; Tsubaki, N. Study on the preparation of Cu/ZnO catalyst by sol–gel auto-combustion method and its application for low-temperature methanol synthesis. Appl. Catal. A Gen. 2011, 401, 46–55. [Google Scholar] [CrossRef]
  85. Huang, L.; Kramer, G.J.; Wieldraaijer, W.; Brands, D.S.; Poels, E.K.; Castricum, H.L.; Bakker, H. Methanol synthesis over Cu/ZnO catalysts prepared by ball milling. Catal. Lett. 1997, 48, 55–59. [Google Scholar] [CrossRef] [Green Version]
  86. Burch, R.; Chappell, R.J. Support and additive effects in the synthesis of methanol over copper catalysts. Appl. Catal. 1988, 45, 131–150. [Google Scholar] [CrossRef]
  87. Jensen, J.R.; Johannessen, T.; Wedel, S.; Livbjerg, H. A study of Cu/ZnO/Al2O3 methanol catalysts prepared by flame combustion synthesis. J. Catal. 2003, 218, 67–77. [Google Scholar] [CrossRef]
  88. Pass, G.; Holzhauser, C.; Akgerman, A.; Anthony, R.G. Methanol synthesis in a trickle-bed reactor. AIChE J. 1990, 36, 1054–1060. [Google Scholar] [CrossRef]
  89. Biswal, T.; Shadangi, K.P.; Sarangi, P.K.; Srivastava, R.K. Conversion of carbon dioxide to methanol: A comprehensive review. Chemosphere 2022, 298, 134299. [Google Scholar] [CrossRef]
  90. Bowker, M. Methanol Synthesis from CO2 Hydrogenation. ChemCatChem 2019, 11, 4238–4246. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Guil-López, R.; Mota, N.; Llorente, J.; Millán, E.; Pawelec, B.; Fierro, J.L.G.; Navarro, R.M. Methanol Synthesis from CO2: A Review of the Latest Developments in Heterogeneous Catalysis. Materials 2019, 12, 3902. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Dang, S.; Yang, H.; Gao, P.; Wang, H.; Li, X.; Wei, W.; Sun, Y. A review of research progress on heterogeneous catalysts for methanol synthesis from carbon dioxide hydrogenation. Catal. Today 2019, 330, 61–75. [Google Scholar] [CrossRef]
  93. Studt, F.; Sharafutdinov, I.; Abild-Pedersen, F.; Elkjaer, C.F.; Hummelshoj, J.S.; Dahl, S.; Chorkendorff, I.; Norskov, J.K. Discovery of a Ni-Ga catalyst for carbon dioxide reduction to methanol. Nat. Chem. 2014, 6, 320–324. [Google Scholar] [CrossRef]
  94. Wang, J.; Zhang, G.; Zhu, J.; Zhang, X.; Ding, F.; Zhang, A.; Guo, X.; Song, C. CO2 Hydrogenation to Methanol over In2O3-Based Catalysts: From Mechanism to Catalyst Development. ACS Catal. 2021, 11, 1406–1423. [Google Scholar] [CrossRef]
Figure 1. Process of synthesizing methanol [9].
Figure 1. Process of synthesizing methanol [9].
Waste 01 00015 g001
Figure 2. Illustration of two commonly accepted pathways for methanol synthesis. (A) The typical gas phase mechanism proceeds via a formate intermediate, which is hydrogenated into a methoxy group before methanol is released. (B) In the liquid phase mechanism, the formate reacts with an alcohol to form an ester before methanol is released [37].
Figure 2. Illustration of two commonly accepted pathways for methanol synthesis. (A) The typical gas phase mechanism proceeds via a formate intermediate, which is hydrogenated into a methoxy group before methanol is released. (B) In the liquid phase mechanism, the formate reacts with an alcohol to form an ester before methanol is released [37].
Waste 01 00015 g002
Figure 3. Flow Jdiagram for the precipitation preparation of Cu/ZnO catalysts [79].
Figure 3. Flow Jdiagram for the precipitation preparation of Cu/ZnO catalysts [79].
Waste 01 00015 g003
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Liu, G.; Hagelin-Weaver, H.; Welt, B. A Concise Review of Catalytic Synthesis of Methanol from Synthesis Gas. Waste 2023, 1, 228-248. https://doi.org/10.3390/waste1010015

AMA Style

Liu G, Hagelin-Weaver H, Welt B. A Concise Review of Catalytic Synthesis of Methanol from Synthesis Gas. Waste. 2023; 1(1):228-248. https://doi.org/10.3390/waste1010015

Chicago/Turabian Style

Liu, Guanfu, Helena Hagelin-Weaver, and Bruce Welt. 2023. "A Concise Review of Catalytic Synthesis of Methanol from Synthesis Gas" Waste 1, no. 1: 228-248. https://doi.org/10.3390/waste1010015

Article Metrics

Back to TopTop