Next Article in Journal
Comparison of Long-Term Strength Development of Steel Fiber Shotcrete with Cast Concrete Based on Accelerator Type
Previous Article in Journal
Recovery and Characterization Studies of Post-Production Alloy Waste from the Automotive Industry
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Catalysts for Hydrogen Generation via Oxy–Steam Reforming of Methanol Process

by
Magdalena Mosińska
,
Małgorzata I. Szynkowska-Jóźwik
and
Paweł Mierczyński
*
Institute of General and Ecological Chemistry, Lodz University of Technology, Zeromskiego 116, 90 – 924 Lodz, Poland
*
Author to whom correspondence should be addressed.
Submission received: 23 October 2020 / Revised: 24 November 2020 / Accepted: 1 December 2020 / Published: 8 December 2020
(This article belongs to the Section Catalytic Materials)

Abstract

:
The production of pure hydrogen is one of the most important problems of the modern chemical industry. While high volume production of hydrogen is well under control, finding a cheap method of hydrogen production for small, mobile, or his receivers, such as fuel cells or hybrid cars, is still a problem. Potentially, a promising method for the generation of hydrogen can be oxy–steam-reforming of methanol process. It is a process that takes place at relatively low temperature and atmospheric pressure, which makes it possible to generate hydrogen directly where it is needed. It is a process that takes place at relatively low temperature and atmospheric pressure, which makes it possible to generate hydrogen directly where it is needed. This paper summarizes the current state of knowledge on the catalysts used for the production of hydrogen in the process of the oxy–steam-reforming of methanol (OSRM). The development of innovative energy generation technologies has intensified research related to the design of new catalysts that can be used in methanol-reforming reactions. This review shows the different pathways of the methanol-reforming reaction. The paper presents a comparison of commonly used copper-based catalysts with other catalytic systems for the production of H2 via OSRM reaction. The surface mechanism of the oxy–steam-reforming of methanol and the kinetic model of the OSRM process are discussed.

1. Introduction

The economic and civilization development of the world has caused the population and the total energy consumption to have increased significantly recently. The current state of the natural environment and the shrinking reserves of crude oil, resulting from excessive consumption of fossil fuels, have intensified research related to the use of alternative fuels and energy generation technologies [1]. The use of fossil fuels has a negative impact on the environment and causes the emission of harmful oxides into the atmosphere. The emission of an excessive amount of harmful gases into the atmosphere is responsible for the formation of smog and the greenhouse effect. This is a serious problem for animals, plants and human health [2]. In addition, it should be emphasized that fossil fuel sources are not renewable. Their continuous exploitation may finally lead to the exhaustion of their sources. The greatest challenge nowadays is the production of energy from renewable sources. One of the promising alternatives to fossil fuels is hydrogen [3]. Hydrogen is a pro-environmental energy carrier. In the burning flame arise only water vapor and nitrogen oxides, and when the oxidation process is carried out in fuel cells, the only byproduct is water vapor. Hydrogen also has a low ignition energy of initiation, which improves combustion. It is worth noting that the sources of hydrogen are practically inexhaustible. The increased interest in hydrogen is also associated with its usage to power fuel cells as one of the possible alternatives to replace fossil fuels [4,5]. Current research proves that the use of hydrogen to power fuel cells is one of the most effective and environmentally friendly technologies of transforming chemical energy into electricity, which does not generate additional pollution [5,6]. Hydrogen and fuel cell are a great opportunity to reduce emissions of compounds such as nitrogen oxides, carbon oxides, various hydrocarbons, which cause acid rain and possibly the greenhouse effect. The utilization of hydrogen and fuel cells can eliminate the formation of smog in highly urbanized cities. In addition, due to the fact that hydrogen used to power fuel cells should be highly pure, because small amounts of carbon monoxide (above 10 ppm) in the gas supplying the fuel cell cause its irreversible adsorption and poisoning of platinum electrodes used in fuel cells [7]. Therefore, the key task seems to be the development of a method or technology for the production of high purity hydrogen. The production of pure hydrogen is an important problem in the modern chemical industry. However, the industrial production of hydrogen-based on natural gas reforming is well known; it is still a problem to find low-cost methods for obtaining hydrogen intended for supplying small or mobile devices, such as fuel cells. High efficiency, durability and reliability of fuel cell technology represents a major potential, pointing to the possibility of the rapid development of industries that use fuel cells. The possible ways of hydrogen generation and application are presented in Figure 1. There are many sources of hydrogen that can include hydrocarbons, alcohols, ammonia and others. However, methanol is one of the most promising sources of hydrogen because it is the simplest alcohol without a C–C bond in the molecule and provides a high H:C ratio in a molecule. Its properties indicate that methanol can be easily reformable in the low-temperature range and under atmospheric pressure. Considering the relatively mild conditions of the methanol-reforming process (p = 1 atm, T = 160–350 °C), it is possible to generate hydrogen practically everywhere. Production of hydrogen through mixed oxy–steam-reforming of methanol enables to be carried out the process autothermally, which is extremely beneficial from the economics point of view.
Methanol fuel cell DMFC (direct methanol fuel cell) is particularly promising. The choice of methanol as a fuel is dictated by the high energy density of methanol and relatively easy oxidation of methanol at a relatively low temperature of 40–80 °C. The principle of operation of a methanol-powered fuel cell is shown in Figure 2. One of the most important applications of fuel cell technology is hybrid systems, which are widely used in the first commercial portable power generators and prototype electric vehicles [8]. Portable power generators are based on a methanol fuel cell (direct or generators equipped with a micro-reformer) [9] and lithium–ion battery. Prototype mobile vehicles use nickel-hydride battery used for starting purposes or for maintaining stable engine operation. These vehicles achieve an efficiency of about 50%, which is a value higher than in hybrid cars equipped with internal combustion engines. The disadvantage of these systems is their present cost, which is difficult to reduce [9]. The basic condition that hydrogen could replace current energy sources is to develop a cheap, efficient and rapid method for hydrogen production.
The importance of hydrogen as a fuel in the twenty-first century still increases with the further development of fuel cells, which are used in automotive, power generation, power supplies, cell phones and laptops. Due to the above reasons, one of the most important areas of the research in the field of hydrogen production, which are undertaken by scientific groups around the world, is the design of new catalytic systems, which must be characterized by high activity and selectivity in the reforming of methanol reaction. An important aspect of the developed catalysts is their high stability, which they should have in order to be used to power fuel cells applied in mobile vehicles or stationary systems constituting an emergency energy source in public places (including military bases, offices, hospitals, administrative buildings, etc.). Their applicability increases due to their quiet operation, the quality of the supplied energy, the possibility to transport and application practically everywhere.

2. Methanol Reforming Reactions

There are many possible sources of hydrogen that can include hydrocarbons, alcohols, ammonia and others [5,10,11,12,13,14,15,16]. An attractive alternative solution to problems associated with storing molecular hydrogen involves on board catalytic production of hydrogen from a high energy liquid fuel such as methanol. Methanol is one of the most promising sources of hydrogen because it is the simplest alcohol without a C–C bond in the molecule and provides a high H: C ratio in a molecule. Its properties indicate that methanol can be easily decomposed to a hydrogen-rich mixture practically everywhere. The hydrogen can be synthesized from the catalytic reforming of methanol, decomposition of methanol, partial oxidation of methanol, or a combination of two processes into one reaction, which is named the oxy–steam-reforming of methanol [4,17,18,19,20,21]. Thermodynamics of the processes of partial oxidation and steam-reforming of methanol suggests that the most energetically favorable solution is a combination of these two reactions in one process. Depending on the stoichiometry of the reaction between methanol, water vapor and oxygen, the methanol processing reactions run according to the following equations:
Steam reforming of methanol (SRM)
CH3OH + H2O → CO2 + 3 H2 ΔH° = 49.7 kJ/mol
Decomposition of methanol (MD)
CH3OH → CO + 2 H2 ΔH° = 128.5 kJ/mol
Partial oxidation of methanol (POM)
CH3OH + 0.5 O2 → CO2 + 2 H2 ΔH° = −192.2 kJ/mol
Oxy–Steam Reforming of Methanol (OSRM—a combination of SRM and POM)
CH3OH + 0.5 H2O + 0.25 O2 → CO2 + 2.5 H2 ΔH° = −71.4 kJ/mol
It is worth emphasizing that the most energetically beneficial process is oxy–steam-reforming of methanol being a combination of steam and partial oxidation of methanol in one process. In addition, this process can be carried out at low temperature (160–350 °C) at atmospheric pressure without the formation of carbon deposits [18,22,23]. This means that hydrogen can be directly produced where it is needed. The production of hydrogen by the oxy–steam-reforming methanol method allows the reaction to be carried out in an auto-thermal way, which is very beneficial from an economic point of view. The supported catalysts used in this reaction undergo deactivation process due to carbon deposition or as a result of overheating. The problem is obtaining pure hydrogen without carbon monoxide. CO is a strong poison of platinum catalyst used directly in fuel cells.
Geissler et al. [24] developed the kinetic model for the autothermal reforming of methanol reaction over CuO/ZnO/Al2O3 catalyst. They reported that the eight possible reactions take place during the autothermal reforming of the methanol process, and seven components were present in the feed or product gases of the reaction (see Figure 3). The reaction scheme showed that the reactions strongly depend on each other because most of the components occur in more than one reaction. In addition, the products of one reaction can be reactants in other reactions. All reactions which run in the presence of oxygen are very fast and strongly exothermic compared to the reverse water gas shift (RWGS) process. This result proves that the major source of hydrogen is the steam-reforming of methanol reaction. Moreover, the dimethyl ether (DME) formation can be considered independently from the other reactions, which runs parallel with another autothermal reforming of methanol reactions. The reactor to autothermal reforming of methanol process is divided into two parts. Strongly exothermic reactions with oxygen take place at the entrance to the reactor. In the lower parts of the catalyst bed, the endothermic reactions, including mainly steam-reforming of methanol, are carried out. Furthermore, two models of autothermal reforming reactions (ATRM) can be considered, named partial oxidation–steam-reforming of methanol (POM-SRM) and total oxidation–steam-reforming of methanol (TOM-SRM). The coefficients describing the number of reactants, which must be taken for the mentioned processes of the methanol processing, are shown in Table 1. The experimental analysis clearly showed that in the presence of oxygen, the methanol conversion occurs mainly via the total oxidation (TOM) reaction. However, the hydrogen production is subsequent to the steam-reforming (SRM) reaction.

3. Surface Reaction Mechanism and Kinetic Models of the Oxy–Steam Reforming of Methanol Process on Copper-Based System

The literature review concerning the catalysts used in the steam, partial oxidation and oxy–steam-reforming of methanol reactions shows that the typical catalysts of the discussed processes are copper catalysts supported on both monoxides: Al2O3, ZnO, CeO2, MgO, La2O3, SiO2, and bi-oxides: ZnO–Al2O3, CeO2–ZrO2, SiO2–SnO2, Al2O3–CeO2 [18,25,26,27,28,29,30,31,32,33,34,35,36,37,38]. Metallic copper surface determined by the chemisorption method, high values of copper dispersion is the targets to attain for the achievement of highly active catalysts. In the literature data, there are many works investigating the addition of promoters [30,32,39,40] and the influence of the preparation method [41,42,43] on catalytic properties of copper supported catalysts. Despite the high activity and selectivity of copper catalysts in the methanol conversion reaction, research is still ongoing to improve their catalytic properties in the investigated process [6]. Therefore, many research centers are trying to understand the mechanism and kinetics of individual methanol processing reactions [44,45,46,47,48,49,50,51,52]. The mechanism of methanol processing reactions has been studied in the literature data by various scientists, but H2 production from methanol is an issue still raised in many works. Various mechanisms of methanol reforming processes have been proposed in the literature, depending on the role of various active centers present on the copper catalyst surface, such as Cu0 [53], Cu+ [54] or Cu0–Cu+ [49] couples. The methanol adsorption and decomposition measurements performed on Cu100 and Cu110 surfaces showed that methanol dissociates from methoxy species (CH3O). The recent studies proved that the presence of absorbed O on the copper surface enhances the formation of methoxy species [49]. Other authors also reported that methanol interacted very weakly on clean Cu(100) [55], Cu(110) [56], and Cu(111) [57] surface, what confirmed the need for activation of the catalysts by partial exposure of the catalyst surface by O2. Wachs et al. [56] reported that the dissociative chemisorption of CH3O-H on Cu increased by the specific interaction, which takes place between the hydroxyl end of the methanol molecule and surface oxygen atoms (see Table 3). Moreover, the Cuδ+Oδ- site facilitates the breakdown of the O-H bond present in the methanol molecule. The surface oxygen comes from the incomplete reduction of the catalyst or reaction atmosphere formed near the space of the catalyst bed during the reaction. In addition, Fisher et al. [56] reported about the dissociation of water to H2 and O even at 73 °C on reduced polycrystalline Cu [58]. The reaction scheme of POM, SRM and OSRM reactions is complicated by a few secondary reactions which run in parallel to the methanol dehydrogenation process. These reactions are listed below:
Total oxidation of methanol (TOM)
CH3OH + 3/2O2 → 2CO2 + 2H2
Water-gas shift reaction (WGS)
CO   +   H 2 O   H 2 + CO 2
PROX reaction
CO   +   1 / 2 O 2   CO 2
Oxidation of H2
H 2   +   1 / 2 O 2   H 2 O
Taking into consideration the above processes, it could be emphasized that CO could be produced or reacted within all presented above reactions. That is why the concentration of carbon monoxide should be carefully controlled, taking into account the further use of reforming of methanol reactions for the production of hydrogen for fuel cells. It should be remembered that CO is a serious poison of platinum electrodes used in fuel cell technology. It is well known that even trace amounts of carbon monoxide can chemisorb irreversibly on the surface of the platinum electrodes, stopping the operation of the fuel cell. Methanol decomposition has been regarded as the easiest process of H2 production from methanol, and it runs during other methanol-reforming processes, including SRM, OSRM and POM. Up to date, there are four reaction possible schemes proposed for SRM reaction. Namely, MD-WGSR scheme, 1-step SRM scheme, SRM-MD-reverse-WGSR scheme and methyl formate scheme [59] (see Table 3). According to the MD-WGSR reaction scheme, CO is the primary product of the reaction, which is produced from the methanol dehydrogenation process. Then in the next step, CO subsequently is converted into CO2 via water gas shift (WGS) reaction [60]. Within the 1-step of the SRM reaction, CO2 and H2 are formed directly from the methanol dehydrogenation process [24]. This 1-step SRM scheme may be complicated in the case of the reaction carried out with high methanol conversion and contact time where CO started to be formed. Breen et al. [61] investigated the Cu/ZnO/Al2O3 catalysts promoted by ZrO2 and reported that the conversion of methanol was completed at about 345 °C and CO has formed starts from 300 °C, even though its formation is thermodynamically permitted at a lower temperature. In addition, the concentration of CO increase with increase of the reaction temperature and is not formed at low contact time. The reaction conditions of the process are presented in Table 3. Agrell et al. [62] reported that the low CO concentration is connected with the fact that CO is formed during the reverse-WGSR process, which runs during the reforming process. The authors also have experimentally confirmed that the quantity of CO formed during the process was decreasing with reducing contact time. This result indicates that a short contact time decreases the quantity of CO produced from the reverse-WGSR reaction. Peppley et al. [45] also investigated the SRM reaction and claimed that MD-WGSR and 1-step SRM schemes did not describe the mechanism of this reaction. Based on their results, they reported that all the MD, WGSR and SRM steps must be included in the mechanism model of the investigated reaction. The authors claimed that during SRM, MD process run parallel as a side reaction. The products of steam-reforming of methanol process are consumed via reverse-WGSR process (see Table 3). The kinetic model suggested by the authors [63,64,65] confirmed that two different centers present on the catalyst surface were required. One kind of the active centers are required for both the SRM and reverse WGSR processes (see Table 3). The second kinds of the active centers are required for the methanol decomposition reaction. On the other hand, a methyl formate scheme of steam-reforming of methanol has been also proposed by several scientists [66,67,68] (see Table 3). According to this scheme at first step methanol is dehydrogenate to methyl formate and then, the methyl formate is hydrolyzed into formic acid. In the next stage of the process the formic acid decompose to CO2 and H2 which were the primary products of the system. All of the steam-reforming of methanol reaction stages remaining above are presented in Figure 4.
Authors based on their results reported that methanol dehydrogenation reaction is the rate-determining step of the SRM process. However, the MD-WGSR scheme could be ruled out because the concentration of CO in the product mixture was below the predicted value resulting from the equilibrium of the SRM process. The produced CO via SRM reaction is a primary product, which was produced by the decomposition of methyl formate according to the reaction below:
HCOOCH3 → CH3OH + CO
Partial oxidation of methanol reaction carried out using molar ratios of O2/CH3OH equal 0.3 and 0.4 were tested by Murcia-Mascaros et al. [70] (see Table 3). The catalytic activity results obtained by the authors indicated that both methanol conversion and concentrations of products strongly depend on the O2 concentration in the reaction mixture. During the partial oxidation of methanol, a significant amount of water and CO2 together with a small amount of H2 was observed at low methanol conversion and in the case of the reaction during which O2 was not completely consumed. Therefore, TOM and MD processes were proposed to be the dominating reactions at the initial stage of the partial oxidation of methanol. In contrast, when oxygen in the reaction mixture is almost consumed, the conversion of methanol increased to a value higher than that resulting from the reaction stoichiometry. At the same time, selectivity towards H2 production increases and towards CO formation decreases. Based on the obtained results, the authors reported that during the TOM process, a parallel SRM reaction takes place. Agrell et al. [11] also proposed a scheme of TOM-SRM consecutive pathway reactions for the partial oxidation of methanol reaction. Rabe and Vogel [71] also investigated POM reaction on Cu/ZnO/Al2O3 catalyst using thermogravimetric analysis coupled with Fourier-transform infrared (FTIR) spectroscopy (see Table 3). The authors reported that the POM reaction from an oxygen-poor mixture leads to the formation of formaldehyde and water. The results also confirmed that CO2 was produced as a primary product of the POM reaction. According to the literature [11], the OSRM reaction is very similar to the partial oxidation of the methanol reaction scheme, in which TOM runs successively through the SRM. The reaction scheme ends with a partial CO oxidation or RWGS reaction. The hydrogen and water production via the OSRM process is related to the relation between copper species and the TOM-SRM consecutive reaction scheme. During the TOM reaction, the metallic copper surface of the pre-reduced Cu/ZnO/Al2O3 catalyst was completely oxidized into Cu2+ by the oxygen present in the reaction mixture. At the point at which oxygen is consumed in the subsequent reaction SRM, the reaction mixture may again become a reduction. As a result, Cu2+ can be converted back into metallic copper. In addition, the copper species play different roles during OSRM depending on the degree of oxidation of Cu. Cu2+ species show negligible activity in H2 production. These centers are active in the formation of water and carbon dioxide. Whereas, metallic copper are very active for the H2 production. The above proposed mechanism is compatible with the results reported by Reitz et al. [53]. They studied the OSRM reaction using Cu/ZnO catalyst by X-ray absorption near edge structure (XANES) technique. The results showed that at low methanol conversion value, Cu2+ was the dominant Cu species, and combustion was the main reaction. In the case of the total conversion of O2, Cu2+ species were reduced to Cu0, and as a consequence, H2 is produced via the SRM process. Agrell et al. [11] reported that the gaseous products formed in POM and OSRM reactions and methanol conversions value were very similar. At the same time, during the SRM reaction, both the methanol conversion and H2 production were initiated at a lower reaction temperature on the metallic copper surface. Whereas, during POM and OSRM reactions the methanol conversion and H2 production were slower at a higher reaction temperature. The lower catalytic activity of both POM and OSRM reactions should be related to TOM process occurred on the Cu2+ species. Whereas, the higher catalytic efficiency is attributed to the SRM reaction that takes place on the metallic copper surface. On the other hand, Patel and Pant [72] proposed a different reaction scheme for OSRM reaction. The conditions of the OSRM process are given in Table 3. Authors reported that during OSRM reaction, partial oxidation of methanol, steam-reforming of methanol and RWGS processes run parallel. A kinetic model for the OSRM system over Cu/ZnO/CeO2/Al2O3 catalyst was developed by using Langmuir–Hinshelwood mechanism. The authors based on their results proposed that TOM was impossible due to the low molar ratio of O2 to CH3OH in a reaction mixture together with the excess of steam introduced in a reaction mixture used during the oxy–steam of methanol reaction. The scientists claimed that there are two different types of active sites on the catalyst surface. First kind of the active centers are used for the adsorption of C- and O-containing species. While, the second types of the active centers present on the catalyst surface are designed for the adsorption of H. According to the previous work describing in the literature data [73,74,75] (see Table 3), CO is formed as a secondary product from the consecutive RWGS process. The kinetic model proposed by the authors is based on the assumption that formate is formed from oxymethylene in the process of partial oxidation of methanol, which determines the reaction rate determining step (RDS). In this model steam-reforming of methanol run through the dissociation of formic acid, from which adsorbed carbon monoxide and hydroxyl groups are formed. The predicted mechanism was also proved by the experimental data obtained by the authors in the OSRM process. Turco et al. [49] also determined the kinetic parameters based on the methanol conversion values in order to understand the differences in the activity of the investigated catalytic systems (see Table 3). They calculated the kinetic constant k (s−1 cm3 gcat−1) from the equation presented below and assumed for OSRM reaction that fractional expansion ɛ = 0.32.
k = 1 τ ε x + 1 + ε l n 1 x
where, τ —contact time (s∙cm−3 ∙gcat) and x—fractional conversion;
Kinetic studies for the OSRM process based on copper catalysts have been conducted. Moreover, the Arrhenius plots presented by the authors show a satisfactory linear regression. Table 2 presents the activation energy values (Ea) obtained for the investigated catalysts. It was also observed that the k values for the OSRM process are not correlated with the copper area. This result is a consequence of the complexity of the OSRM reaction. During the OSRM process, the POM reaction takes place, which is catalyzed by copper oxide in the first zone of the catalyst bed with access to high oxygen concentration. They reported that the interaction with other oxides could influence the activity of such Cu oxide, which may have an effect on the kinetics of POM reaction.
Turco et al. [48] also investigated the OSRM mechanism on Cu/ZnO/Al2O3 by FTIR technique, and they reported in their work that methanol adsorbs dissociatively on the catalyst surface (see Table 3). Methoxy groups are adsorbed both on the active sites and on inert regions of the catalyst surface, mainly associated with alumina. Methoxy groups in the vicinity of copper or adsorbed directly on it are easily transformed into surface forms of the formate. Based on their research, the authors discovered traces of formaldehyde also adsorbed on the catalyst surface. In the next step, the formate groups decompose to CO at or above 300 °C. The FTIR measurements also confirmed the formation of dimethyl ether during the OSRM process. Base on the above discussion, the authors proposed the following reaction scheme of the OSRM process carried out on the copper catalyst supported on ZnO·Al2O3 carrier (see Figure 5).
The authors reported in their work that the mechanisms of MD (reaction B), POM (reaction D), and SRM (reaction A) reactions are closely related and that the surface phenomena are similar. They also observed that the conversion of methanol in the presence of both water (SRM) and oxygen (POM) is higher compared to the methanol decomposition reaction what is directly related to the oxidation state of Cu present on the catalyst surface. It should also be emphasized that a small amount of carbon monoxide is formed in the presence of water (SRM), and CO2 is the dominant product. The FTIR results clearly showed that CO and CO2 are formed in appreciable amounts in the POM reaction. It is also known that CO is adsorbed very weakly on metallic copper particles [52], while it is adsorbed very strongly on Cu(I) centers. CO is easily oxidized to CO2 on Cu(II) centers. During the methanol decomposition reaction, the catalyst is completely reduced, and in this case, CO is produced by the decomposition of the formate groups. In contrast, the catalyst is partially oxidized during the POM reaction. Therefore, the Cu(I) sites present on the catalyst surface may strongly adsorb CO, while Cu(II) allows its oxidation to CO2 before its desorption. Based on this assumption, CO2 is the main product of the POM process. During the POM reaction, when O2 is consumed, the excess methanol partially decomposes to CO. However, in the case of the SRM reaction, CO is formed in a limited amount. It can also be assumed that the water vapor present during the SRM process oxidizes the catalyst surface, leading to the formation of Cu (II) centers on which CO2 is formed. The authors also report that the scheme of the methanol decomposition mechanism was reasonably proposed by Riva et al. [76] (see Figure 6).

4. Catalysts Configuration Systems Applied for Hydrogen Production in the Oxy–Steam Reforming of Methanol Process

A literature review concerning catalysts systems applying in reforming of methanol processes indicates that the typical catalysts are Cu, Ni, Co, Fe, Pd, Pt, Ru, Au, Ir, Ag supported on mono-Al2O3, ZnO, CeO2, MgO, La2O3, SiO2 [21,73,78,79], and binary oxides: ZnO-Al2O3, CeO2-ZrO2. SiO2-SnO2, A12O3-CeO2 [32,73,79,80]. Those catalytic systems are prepared using different methods including: microemulsion [81], aerogel [82], co-precipitation [43], sol–gel [83], impregnation [46], combustion synthesis [84], and others. However, still, the most common catalysts used for oxy–steam reforming of methanol reaction are the copper-based systems. Their metal surface area is determined by the chemisorption method; high values of copper dispersion are the targets to attain for the achievement of highly active catalysts. In the literature data, there are many works investigating the addition of promoters [30,32,39,40] and the influence of the preparation method [41,42,43] on catalytic properties of copper supported catalysts. Despite the high activity and selectivity of copper catalysts in the methanol conversion reaction, research is still ongoing to improve their catalytic properties in the investigated process [6].

4.1. The Influence of the Preparation Method on the Catalytic Properties of the Tested Catalytic Systems in the Oxy–Steam Reforming of Methanol Process

The effect of the preparation method of copper catalysts on their catalytic properties in the oxidative steam-reforming of methanol was studied by Shen and Song [42]. They have compared the physicochemical properties of copper catalysts synthesized by impregnation, co-precipitation and hydrothermal synthesis methods, and they claimed that the systems prepared using the co-precipitation method exhibited higher surface area (46% higher than wet impregnation) and methanol conversion in the oxidative steam-reforming of methanol reaction. The activation of the catalyst through the reduction process performed before each catalyst test at lower temperature results in increasing the activity towards hydrogen production. The CuO/ZnO/Al2O3 catalyst prepared by the co-impregnation method was also examined in the oxidative steam-reforming of the methanol process. The lowest CO concentration close to zero was observed when H2O/methanol and O2/methanol ratio equal to 1.43 and 0.47, respectively. Papavasiliou et al. [39] investigated Cu-Mn spinel oxide catalysts as an alternative to a commercial CuO/ZnO/Al2O3 catalyst. The physicochemical properties and catalytic characterization of the investigated catalysts applied in OSR of methanol reaction are given in Table 4. The Cux-Mny catalytic systems were synthesized by the urea nitrate combustion method. The activity results showed that the investigated catalysts exhibited high activity to H2 production in combined (oxy–steam) reforming of methanol reaction. The most active catalysts which exhibited almost identical activity and selectivity results were Cu0.30-Mn0.70 and Cu0.40-Mn0.60 catalysts. Their high activity in the studied reaction is explained by their almost identical physicochemical properties which were confirmed by XRD and XPS analysis. The authors also investigated the stability of Cu0.30-Mn0.70 and commercial CuO/ZnO/Al2O3 catalysts in combined (oxy–steam) reforming of methanol reaction at 300 °C within 8 h of operation. The only observed difference between the tested systems was the value of the hydrogen selectivity, which was higher for the Cu0.30-Mn0.70 catalyst. Catalysts operated stably under the reaction conditions, demonstrating constant selectivity towards H2 formation and a slight decrease in methanol conversion from 99 to 96% after 8 h of catalyst operation. Other catalysts prepared by Papavasiliou et al. [79] were CuO–CeO2 systems prepared by the urea–nitrate combustion method. The activity tests were performed in combined (oxy–steam (CRM) reforming of methanol) reaction. The result presented in Table 4 shows that the catalyst surface area and crystallites size could be controlled and optimized by the preparation method. The authors reported that the optimal fuel to oxidant ratio (urea/nitrates) was 4.17 and the optimal Cu/(Cu + Ce) atomic ratio was 0.15, with the surface area increasing by about 4.3 times, while the methanol conversion increased from 52% to 100%, and the activity increased by about 1.8 times. The comparison of the activity in SRM and CRM showed that under autothermal conditions, the activity of the catalyst was improved. The higher methanol conversion obtained during the CRM process is assigned to better efficiency of the heat transferring in the catalyst bed. In CRM reaction, the methanol conversion obtained at 300 °C was equal to 100% with more than 97% selectivity toward hydrogen formation. However, the CO concentration increased in all cases by increasing the reaction temperature. Liu et al. [85] studied Pd/ZnO catalysts prepared by impregnation and co-precipitation methods and tested their catalytic activity in the oxidative steam-reforming of the methanol process (see Table 4). In addition, the authors also investigated the influence of Pd loading on the activity results in the tested reaction. The Pd/ZnO catalysts with Pd loading below 5% prepared by impregnation method (IP method) showed better activity in the oxidative reforming of methanol process compared to samples prepared by co-precipitation (CP) method. This is due to the higher concentration of the PdZn alloy on the surface of the catalyst prepared by the IP method. Nevertheless, for the higher Pd loading, the Pd/ZnO catalysts prepared by CP method were more active than system prepared by IP method. Based on the obtained results, the authors concluded that the activity of the catalyst strongly depends on the crystal size and the dispersion of the PdZn alloy on the ZnO support. Increasing the size of the Pd crystals on the ZnO support improves the catalytic activity and selectivity of the palladium catalysts. Moreover, the amount of CO generated during the tested process is effectively reduced, along with an increase in the Pd content in the catalytic material.

4.2. Effect of the Type of Carrier on the Catalytic Reactivity of the Catalytic Systems Applied in the Oxy–Steam Reforming of Methanol

The type of support and its nature has a great influence on the catalyst material activity and stability [86]. The promotion of copper catalyst by CeO2 leads to an increase of the thermal stability in steam-reforming of methanol reaction [78]. CeO2 itself can provide mobile oxygen, which has a direct influence on the catalytic activity and may affect the oxidation state of the metal present on the support surface under a reducing environment. Recently, an increase in interest has in the case of catalysts supported on CeO2 has been observed. The crystal structure of CeO2 is a cubic fluorite network for which it is possible to introduce other cations, such as Si4+, Th4+, Zr4+, Y3+, La3+, Sc3+, Mg2+, Ca2+ or Cu2+ to improve catalytic properties of CeO2 [87]. High mobility oxygen present in the CeO2 containing systems [88], the strong interaction of CeO2 with supported metal—strong metal-support interaction (SMSI) [89], all these features make these systems as promising catalytic materials. The role of promoters of copper supported catalysts is one of the main factors which was brought up in many publications. Promoters have been used to influence the status of copper and improve the activity of the catalyst. Promotion of copper catalyst by CeO2 improves activity and stability of the copper-supported catalyst and leads also to the improvement of copper dispersion on catalyst surface and in the same time protects copper crystallites against poisoning and has influence on greatest crystallites formation [90]. Mierczynski et al. [91] studied the influence of copper content in monometallic xCu (where, x = 5, 20, 40 and 60 wt.%) catalysts supported on binary oxide CeO2·Al2O3 on the activity of these systems in the oxy–steam reforming of methanol reaction and the results are presented in Table 4. The authors confirmed that the reactivity of the investigated systems depends on the copper content and its dispersion on the catalyst surface. These studies confirmed that the 20 wt.% of Cu content is the optimum content of copper in order to obtain the highest methanol conversion and reaction rate value compared to other investigated catalytic systems. In addition, the activation energy in the OSRM process for the 20% Cu/CeO2·Al2O3 catalyst was the lowest and equal 66.56 kJ/mol. In other work [23], authors studied both mono-Ni and bimetallic Pd-Ni catalysts supported on CeO2, Al2O3 and CeO2·Al2O3 carriers and they reported about the highest activity of bimetallic 2% Pd-40% Ni/CeO2·Al2O3 system compared to the other investigated catalysts. This catalytic system showed the highest stability and selectivity to H2 production in the oxy–steam-reforming of methanol process compared to the systems supported on monoxide. In addition, they investigated the influence of the Ni loading on the catalytic activity of the prepared monometallic systems in the OSRM process. The results showed that system containing 40 wt.% of Ni exhibited the highest methanol conversion value and the highest selectivity to hydrogen formation compared to the rest of the monometallic catalysts. Mierczynski et al. [18] extensively investigated the influence of support composition on the catalytic activity of copper catalysts in order to obtain optimal catalyst composition for the OSRM reaction (see the results presented in Table 4). They prepared various copper catalysts supported on binary oxides systems (ZrO2·Al2O3, where Zr:Al = 2:1, 1:1 and 1:2). The reactivity results performed in the oxy–steam-reforming of methanol confirmed that the highest active system was 20% Cu/ZrO2·Al2O3 (1:2) catalyst. In further studies, the authors compared the physicochemical and catalytic properties of Cu catalyst with Ni system supported on ZrO2·Al2O3 (1:2) carrier. The reactivity measurements showed that the supported copper catalyst was more active than the nickel catalyst. They have found that the catalytic activity of the investigated catalysts strongly dependent on their acidity and sorption properties in relation to methanol. The catalytic tests confirmed the highest activity of copper catalysts supported on ZrO2-Al2O3 (Zr:Al = 0.5) binary oxide promoted by noble metals such as Pd or Rh. The most active systems exhibited the highest specific surface area and the highest number of acidic centers on their surfaces. The XPS measurements showed that the catalyst with the lowest ratio between Cu0 and Cu+ species present on the catalyst surface exhibited the highest activity in the OSRM process, and the relationship between these species is a critical parameter to achieve highly active systems in the OSRM reaction. Furthermore, they have found that the pre-treatment process of Ni catalysts carried out before the activity tests play a significant role in terms of the catalytic activity of the tested systems in the OSRM process. In other work [22], authors studied the influence of the binary oxide composition of ZnO·Al2O3 (Zn:Al = 2:1, 1:1, 1:2 and 1:4) on the physicochemical and catalytic properties of nickel supported catalysts. The reactivity tests performed in the oxy–steam-reforming of methanol reaction showed that 20% Ni/ZnO·Al2O3 (1:1) system was the most active catalysts in the investigated reaction at 300 °C. This result was explained by the easiest reducibility, and the highest acidity of this catalyst compare to other investigated catalysts. Iwasa et al. investigated various metal catalysts Me/ZnO (where Me = Fe, Co, Ni, Pd, Pt, Ir or Ru) catalysts in transformation of methanol in the presence of steam and oxygen [92]. He proved that the best promising system was Pd/ZnO monometallic catalyst. Their high activity was explained by intermetallic Pd-Zn formation. Similar to the Pd/ZnO catalyst, the low selectivity of Pt/ZnO towards CO was assigned to PtZn alloy formation. Whereas, in the case of Ni, Ru, Co and Ir catalyst, no alloy phase was formed after reducing the catalysts at 500 °C. These catalysts showed low methanol conversion and produce a higher concentration of carbon monoxide [93,94]. Palladium is active component for methanol-reforming reactions. It is an effective decomposition catalyst, selectively forming H2 and CO when it is supported on metal oxide [95,96,97,98]. Udani et al. [99] also tested Cu-CeO2 catalysts in the oxy–steam-reforming of methanol reaction. They investigated copper catalysts with various content of copper. They reported that catalyst with 70 at.% Cu showed the highest catalytic activity in the studied process and was slightly better than commercial SRM catalyst (Synetix 33–5). In addition, the catalytic measurements performed in the OSRM process confirmed the stable operation of the catalytic system at 300 °C. In the literature, data can be found only a few works concerning the use of carbon nanotubes as a carrier of catalytic system used in oxy–steam-reforming of methanol processes. The use of carbon nanotubes (CNTs) or multiwalled carbon nanotubes (MWCNTs) as a carrier for methanol-reforming catalysts is associated with the attractive properties of this material [100]. Carbon nanotubes were applied as a carrier of catalytic materials due to their specific structural morphology and their specific physicochemical properties. In addition, CNTs properties can be modified by functions groups introduced onto their surface. This modification possibilities gives capabilities to obtain specific material. Furthermore, catalysts supported on CNTs can be prepared by different methods such as: impregnation, precipitation, colloidal, electroless plating, and hydrothermal method [100]. The main advantage of carbon nanotubes used as a catalyst carrier is their high purity, high thermal and mechanical stability, the presence of specific interactions on the boundary metal-support, the possibility of adsorption of catalytically active nanoparticles inside or on the external wall of CNTs and their specific electron structure. Good conductivity of carbon nanotubes promotes the “spillover” effect in a place of interfacial boundary created by the active center [101]. Nanoparticles of an active phase dispersed on the functional surface of the support (CNTs) make that created active centers are easily accessible to the reactants. All remaining above properties of CNTs can directly affect the catalytic activity and selectivity [102,103,104,105,106]. All mentioned properties of CNTs material have an important influence on the catalytic activity and selectivity of the synthesized catalysts. Mierczynski et al. [19] reported the potential usages of MWCNTs in the OSRM reaction to hydrogen generation. They have confirmed that magnetite, metallic iron and defects formed by encapsulated or removal of metal particles play an important role in the oxy–steam-reforming of methanol. Their results indicate that MWCNTs are potential carriers for various metallic catalysts which can be applied in methanol processing reactions.

4.3. Role of Promotors Addition on the Catalytic and Physicochemical Properties of Catalytic Materials Tested in the Oxy–Steam Reforming of Methanol Reaction

The role of promoters of copper supported catalysts is one of the main factors which was brought up in many publications. Promoters have been used to influence the status of copper and improve the activity of the catalyst. Modification of copper catalyst by CeO2 improves activity, the addition of ZnO and ZrO2 cause increase of catalyst surface, and also stabilize crystallites size of copper, and in the same time protects crystallites against their aggregations. Additionally, ZrO2 stabilizes the copper Cu+ ions on the catalyst surface [39]. Agrell et al. [107] determined the influence of ZrO2 and Al2O3 promotion on physicochemical and catalytic properties of Cu/ZnO catalyst (see the results presented in Table 4). The authors observed an increase in copper dispersion after adding ZrO2 to Cu/ZnO catalyst. While the structural promoter Al2O3 provides a larger surface into catalytic systems, which directly leads to high dispersion of copper [30,107]. The authors observed an increase in both the total surface area and the degree of copper dispersion after the introduction of Al2O3 into the catalytic system. They also performed the activity tests of catalytic systems in the steam, partial oxidation and oxy–steam reforming of methanol processes. They reported that ZrO2 containing catalysts were more active compared to other investigated catalysts. Cu/ZnO/ZrO2 and Cu/ZnO/ZrO2/A12O3 catalysts exhibited high activity in oxy–steam-reforming of methanol. It should be emphasized that the Cu/ZnO catalyst was more efficient at low CH3OH conversion values. In the case of this system, the authors observed a lower amount of CO formed during the studied process compared to the amount of carbon monoxide generated during the steam-reforming reaction of methanol (SRM). As part of the work, the authors conducted stability studies in the process of oxy–steam reforming of methanol on Cu/ZnO and Cu/ZnO/ZrO2/Al2O3 catalysts at a temperature of 260 °C within 20 h of operation. The results showed that ZrO2 addition improves the lifetime of the studied catalyst. A similar promotional effect attributed to Cr2O3 was observed in other works [40,108]. The addition of this promotor acts as a stabilizer of the copper structure, protecting it against sintering. The promotion effect of palladium on the catalytic activity and selectivity towards hydrogen production of nickel-supported catalysts in the OSRM was studied by Mierczynski et al. [23]. The obtained results showed that the palladium addition facilitates the reducibility of the nickel supported catalysts. In addition, they confirmed the highest activity of Pd-Ni catalyst in the studied process, which exhibited the highest selectivity towards hydrogen production and towards carbon monoxide formation in the investigated reaction. Liu et al. [109] investigated the effects of modifier (Zr, Fe, Mg, Al and Cr) of a Pd/ZnO catalyst prepared by co-precipitation method on the catalytic activity in the oxidative reforming of methanol reaction. The addition of modifier to Pd/ZnO catalyst in the form of metal significantly modifies its physicochemical and catalytic properties in an oxidative reforming of methanol process. The ZnO crystal size decreases as a result of modifier addition. During high-temperature reduction treatment, the smaller PdZn alloy particles are formed. The introduced modifier can interact with ZnO, which leads to the formation of binary oxides systems such as ZnAl2O4, ZnCr2O4 and ZnFe2O4. The promotion effect of a small amount of Cr or Fe into Pd/ZnO catalyst results in decreases in the amount of CO formed during the process and increased the selectivity toward H2 production. However, the CO selectivity of the Pd/ZnO catalyst increases as a result of Mg, Zr or Al addition. Mierczynski et al. [18,22,23] also investigated the influence of Pd addition into Ni catalysts on their activity results in the OSRM process. The authors confirmed that the palladium addition has a promotion effect on the performance of Ni supported catalyst in the OSRM reaction. They proved that Pd facilitates the NiO reduction by the spillover effect occurring between Pd and NiO species. This phenomenon is related to the generation of new adsorption centers on the catalyst surface, which may be responsible for increasing the activity and selectivity of the bimetallic catalyst towards H2 formation. They also reported that the activity in the OSRM reaction strongly depends on the Ni content in the investigated system and their reduced properties. The acidity measurement also confirmed that acid sites play a crucial role in the oxy–steam-reforming of the methanol process. In addition, the presence of acid centers located close to high dispersed metal centers may have a significant influence on the reactivity of the tested catalyst systems. Catalytic activity tests performed for 48 h confirmed the stable operation of the Pd-Ni bimetallic catalyst. (see Figure 7). In another work [18], the authors also reported the high activity of bimetallic Rh-Cu, Rh-Ni and Pd-Cu, Pd-Ni catalysts tested in the OSRM process. They confirmed that the most active system in the studied process was 0.5% Rh-20% Cu/ZrO2·Al2O3 (1:2) catalyst (see Figure 8). They also claimed that the occurrence of Cu0 and Cu+ species and their ratio is a critical parameter to achieve highly active systems in the OSRM reaction.
In other work [92], the authors reported the potential capability to use palladium as a promoter for methanol-reforming catalysts. Despite the relatively high price of gold, even a small addition of gold can significantly improves the activity of bimetallic catalysts through the formation of intermetallic compound Au–Cu, Au–Ni, which catalyze the OSRM process. Since Haruta’s discovery that catalysts containing nanosized gold particles have extraordinary activities for reactions, including CO oxidation [110], there has been substantial interest in their use and the origin of their exceptional catalytic properties. The use of gold as a promoter of the active phase (Cu, Ni) results in lowering the reaction temperature and improves selectivity toward hydrogen and carbon dioxide formation. The selectivity improvement is associated with the surface modification of copper and nickel by gold atoms. These new atoms can generate new adsorption centers, which may be involved in the dissociative adsorption of methanol or adsorption of atomic oxygen as a result of dissociation of previously adsorbed methanol. Adsorbed oxygen atoms can leave the surface of the catalyst as a product of CO2, thereby reducing the formation of CO. Mierczynski et al. [17,111,112] for the first time investigated the physicochemical and catalytic activity of the mono-(Cu, Ni) and bimetallic (Au–Cu, Au–Ni) catalysts supported on multiwalled carbon nanotubes (MWCNTs) in oxy–steam-reforming of methanol process (see Figure 9). They proved that the AuCu and AuNi alloys are formed in the case of bimetallic Au–Cu/MWCNTs and Au–Ni/MWCNTs catalyst after their reduction. These results were confirmed by XRD, XPS and SEM-EDS measurements. The spillover effect between metallic gold and nickel (copper) oxide was confirmed by TPR-H2 studies. The bimetallic Au–Cu catalyst exhibited a significant improvement of the activity and selectivity towards hydrogen formation compared to the monometallic systems tested in the OSRM process at higher temperatures (300 °C). The higher activity and selectivity towards H2 production is explained by an alloy AuCu formation. The Au–Ni/MWCNTs catalyst showed the lowest selectivity towards CO formation at 200 and 300 °C, which is very important from the application point of view of these catalytic systems in fuel cell technology.

4.4. Role of Active Phase Composition on Reactivity Properties of Catalytic Materials Applied in Oxy–Steam Reforming of Methanol

The type of an active phase has an important influence on the reactivity properties of synthesized catalysts in methanol processing reactions. It is well known that copper and nickel catalysts supported on metal oxide support are highly active in methanol processing reactions, as evidenced by a number of papers published in reputable journals concerning copper and nickel catalysts using in methanol reforming processes. In addition, transition metal catalysts and noble metals are also used in oxy–steam reforming of methanol reaction. The activity results obtained in steam-reforming of methanol reaction on Pd/ZnO catalyst showed that palladium is an effective catalyst in the studied process. Pd/ZnO system showed high selectivity for CO2 production in the SRM process. The high selectivity towards CO2 formation is explained by the Pd-Zn alloys formed on the catalyst surface [113]. These palladium alloys are formed at moderate temperatures under reducing conditions [114,115]. Iwasa et al. [92] also compared the catalytic properties of Pd/ZnO catalyst with copper-containing systems (Cu/ZnO, Cu/ZrO2 and Cu/SiO2) in steam reforming of methanol reaction. The methanol conversion and hydrogen concentration obtained for Pd/ZnO were higher compared to Cu/ZrO2 and Cu/SiO2 catalysts and slightly lower than for Cu/ZnO system. The influence of Pd loading on Pd/ZnO catalyst was examined, and obtained results showed the CH3OH conversion increased with increasing Pd loading. In the presence of oxygen, the distribution of reaction products depends on Pd loading. The increase od Pd concentration in Pd/ZnO catalyst results in increasing of H2 and decreasing of CO level in an obtained reaction product. The 1% Pd/ZnO catalyst exhibited the highest concentration of H2O in the product in contrast to the 10% Pd/ZnO catalyst for which water was not a product of the reaction. Xu et al. [116] have been investigated Pt-based alloys, including Pt–Cr, Pt–Fe, Pt–Co, Pt–Ni and Pt–Au, as methanol-tolerant cathode catalysts. They reported that the Pt–Au catalyst had been the most promising one in terms of both the catalytic activity and stability. Studies indicate that the methanol-tolerant mechanism of the Pt-based alloys can be attributed to the diluted Pt sites for methanol dehydrogenation as compared with the pure Pt catalyst.
Manzoli et al. [117] investigated CuO/ZnO, Au/ZnO, Cu/TiO2 and Au/TiO2 catalysts in the decomposition and oxy–steam reforming of methanol process. The catalyst systems supported on ZnO were prepared by the co-precipitation method, while catalysts containing TiO2 oxide were prepared by deposition–precipitation method. The catalytic activity of the investigated systems expressed as H2/CO ratio can be described by the following row: Cu/ZnO > Cu/TiO2 > Au/ZnO > Au/TiO2, respectively. In contrast, the CO2/CO ratios formed during the investigated process for Au containing systems were higher than for Cu catalysts. Authors indicated that copper catalysts used in the OSRM process (reaction mixture—CH3OH-H2O-O2 in a molar ratio equal 1:1:0.2) at 200 °C exhibited a higher molar ratio of H2/CO compared to the process realized over gold-containing systems. However, the produced CO2/CO ratio in the final product obtained in the OSRM process over gold catalysts was higher than in the case of copper catalysts, which means that the use of a gold catalyst reduces the amount of CO generated in the reaction. In addition, the authors reported about lower activity of the TiO2 supported catalysts in the OSRM process. These results are related to the high selectivity of TiO2-containing systems towards methane formation, which reduces the amount of produced hydrogen. Figure 10 present the surface and gaseous species which are produced during the OSRM process. They have reported about undefined C-containing species (possibly polyoxymethylene or bidentate carbonate) formed during contact of the reaction mixture with the catalyst surface.
Literature data show that bimetallic catalysts Ni–Cu have also been extensively studied in the oxy–steam reforming of both ethanol [118] and methanol reactions [119]. Catalytic tests of NixCuy-Al catalysts with different Ni to Cu contents [120] carried out in the studied processes showed that the Ni-Cu alloy containing catalyst had better performance in reforming of ethanol and methanol process compared to monometallic copper catalysts. It has also been shown that the introduction of Cu into nickel-based catalysts prevents carbon deposition and sintering of the active phase of catalysts used in the methanol-reforming process. The addition of Cu to the nickel catalyst also prevents the formation of methane and increases the stability of the Ni catalyst during the methanol-reforming process [119]. While, the addition of nickel into copper catalysts also improves Cu dispersion compared to the dispersion of copper species observed in the case of the Cu/ZnO/Al2O3 catalyst. The authors reported that a bimetallic 5% Ni-5% Cu/Al2O3 catalyst was very active in both methane and methanol steam reforming reactions compared to commercial catalysts. Perez-Hernandez et al. [121] also studied Cu/ZrO2, Ni/ZrO2 and Cu–Ni/ZrO2 catalysts in oxidative steam reforming of methanol reaction in order to produce H2-rich gas at relatively low temperature (see the results presented in Table 4). The activity results showed that the monometallic Ni/ZrO2 catalyst was more active than the Cu/ZrO2 system in the OSRM reaction at higher temperatures. However, the bimetallic Cu–Ni/ZrO2 catalyst showed the best catalytic performance at low reaction temperatures compared to the monometallic catalysts. This activity result was attributed to the bimetallic nanoparticles present on the catalyst surface with different Cu/Ni weight ratios. The investigated catalysts showed a similar selectivity toward H2 production equal about 60–70% at higher reaction temperatures. The Cu–Ni/ZrO2 system exhibited high selectivity toward CO formation, which is related to the presence of bimetallic nanoparticles on the catalyst surface. Mosinska et al. [4] also studied the bimetallic x Cu-y Ni (where x(y) = 10, 20 and 30 wt.%) catalysts supported on binary oxides (ZnO·Al2O3, CeO2·Al2O3, ZrO2·Al2O3), and the results are given in Table 4. Authors reported that the hydrogen can be effectively produce in the OSRM process over an investigated Cu–Ni catalyst systems. The 30% Cu–10% Ni/ZrO2·Al2O3 system was the most active system at 160 °C. The high reactivity of this catalyst is related with the Cu0.8Ni0.2 alloy formation which was confirmed by XRD, ToF-SIMS and XPS techniques (see Figure 11).
In addition, the most active system in the studied process at 200 °C was 30% Cu–10% Ni/CeO2·Al2O3 catalyst. This catalytic system was easily reducible and showed the presence of the Cu0.8Ni0.2 alloy on its surface. The authors also confirmed that the alloy composition is an important parameter influencing the reactivity of the bimetallic Cu–Ni catalysts in the OSRM process. At the same time, the bimetallic 30% Cu–10% Ni/ZnO·Al2O3 catalyst containing the equimolar Cu–Ni alloy and irreducible support showed the lowest activity in the studied oxy–steam-reforming of methanol reaction. The authors of the work [122] investigated the catalytic activity of the Au/CeO2-0.135 system in the OSRM process. The results showed that it has the highest methanol conversion and the H2 production rate for oxidative steam reforming of methanol (OSRM). The reducibility and gold particle size played a crucial role in determining the H2/CO2 ratio at the temperature of 200 °C. The authors reported that the reducibility of the catalytic material, small gold particle size and cationing gold centers as active sites play a crucial role in the catalytic activity of prepared catalysts in the OSRM process. Wang et al. [123] reported the high activity of the ZnO–Cr2O3/CeO2–ZrO2 catalyst in oxidative steam reforming of methanol at higher temperatures, which exhibited fast reaction rates compared to conventional Cu-based catalyst. The investigated catalysts produced much lower CO compared to the Cu catalyst and can be used to determine detailed kinetic modeling and design of optimal parameters of the OSRM reactors.

5. Conclusions

This work provides valuable information about catalyst systems used in reforming methanol processes. This paper sheds light on the role of the catalyst in oxy–steam reforming of methanol reaction and presents possible methods of modification of catalytic systems in order to achieve high active, stable and selective catalysts of this process. It was shown that the selection of the suitable preparation method, type of support, the addition of promoters to copper, nickel or transition metal catalysts have a great influence on the catalyst performance in the OSRM reaction. The mechanism of the OSRM process on the copper-based catalyst surface was also discussed. However, it is still a controversial issue to show the clear approach of the OSRM mechanism. In this review, we presented the latest data concerning the catalytic materials and mechanism approach applied in the oxy–steam reforming of methanol process. The presented data may become the basis for the development of the industrial catalyst used in the methanol processing reaction and may also contribute to the development of new technologies based on fuel cells.

Author Contributions

The work was designed and presented by M.M., M.I.S.-J. and P.M. All authors have read and agreed to the published version of the manuscript.

Funding

The work was funded by the National Science Center within the “OPUS” Program, Poland (Grant no. 2018/29/B/ST8/01317).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

ATRMAutothermal reforming of methanol
CIPCo-impregnation method
CNTsCarbon nanotubes
CPCo-precipitation method
CRMCombined (oxy–steam) reforming of methanol
DMEDimethyl ether
DMFCDirect methanol fuel cell
DPDeposition–precipitation method
FTIRFourier-transform infrared spectroscopy
GHSVGas hourly space velocity
IPImpregnation method
MDDecomposition of methanol
MWCNTsMultiwalled carbon nanotubes
OSROxy–steam-reforming of methanol
POMPartial oxidation of methanol
RDSRate-determining step
RWGSReverse water gas shift
SBETSpecific surface area
SIPSubsequent impregnation method
SMSIStrong metal-support interaction
SRMSteam reforming of methanol
TOMTotal oxidation of methanol
UNCUrea nitrate combustion method
W/FCatalyst weight/volume flow rate ratio
WGSWater-gas shift
XANESX-ray absorption near edge structure
XPSX-ray photoelectron spectroscopy
XRDX-ray diffraction spectroscopy

References

  1. Salvi, B.L.; Subramanian, K.A.; Panwar, N.L. Alternative fuels for transportation vehicles: A technical review. Renew. Sustain. Energy Rev. 2013, 25, 404–419. [Google Scholar] [CrossRef]
  2. He, L.-Y.; Qiu, L.-Y. Transport demand, harmful emissions, environment and health co-benefits in China. Energy Policy 2016, 97, 267–275. [Google Scholar] [CrossRef]
  3. Mierczynski, P.; Vasilev, K.; Mierczynska, A.; Maniukiewicz, W.; Maniecki, T.P. Highly selective Pd–Cu/ZnAl2O4 catalyst for hydrogen production. Appl. Catal. A Gen. 2014, 479, 26–34. [Google Scholar] [CrossRef]
  4. Mosinska, M.; Stępińska, N.; Maniukiewicz, W.; Rogowski, J.; Mierczynska-Vasilev, A.; Vasilev, K.; Szynkowska, M.I.; Mierczynski, P. Hydrogen Production on Cu-Ni Catalysts via the Oxy-Steam Reforming of Methanol. Catalysts 2020, 10, 273. [Google Scholar] [CrossRef] [Green Version]
  5. Alves, H.J.; Bley Junior, C.; Niklevicz, R.R.; Frigo, E.P.; Frigo, M.S.; Coimbra-Araújo, C.H. Overview of hydrogen production technologies from biogas and the applications in fuel cells. Int. J. Hydrog. Energy 2013, 38, 5215–5225. [Google Scholar] [CrossRef]
  6. Mierczynski, P. Comparative Studies of Bimetallic Ru–Cu, Rh–Cu, Ag–Cu, Ir–Cu Catalysts Supported on ZnO–Al2O3, ZrO2–Al2O3 Systems. Catal. Lett. 2016, 146, 1825–1837. [Google Scholar] [CrossRef]
  7. Echigo, M.; Tabata, T. A study of CO removal on an activated Ru catalyst for polymer electrolyte fuel cell applications. Appl. Catal. A Gen. 2003, 251, 157–166. [Google Scholar] [CrossRef]
  8. Grochala, W.; Edwards, P.P. Thermal Decomposition of the Non-Interstitial Hydrides for the Storage and Production of Hydrogen. Chem. Rev. 2004, 104, 1283–1316. [Google Scholar] [CrossRef]
  9. Samimi, F.; Rahimpour, M.R. Chapter 14—Direct Methanol Fuel Cell. In Methanol; Basile, A., Dalena, F., Eds.; Elsevier: Amsterdam, The Netherlands, 2018; pp. 381–397. [Google Scholar]
  10. Adhikari, S.; Fernando, S.D.; Haryanto, A. Hydrogen production from glycerol: An update. Energy Convers. Manag. 2009, 50, 2600–2604. [Google Scholar] [CrossRef]
  11. Agrell, J.; Boutonnet, M.; Melián-Cabrera, I.; Fierro, J. Production of hydrogen from methanol over binary Cu/ZnO catalysts. Appl. Catal. A Gen. 2003, 253, 201–211. [Google Scholar] [CrossRef]
  12. Alejo, L.; Lago, R.; Peña, M.A.; Fierro, J.L.G. Partial oxidation of methanol to produce hydrogen over CuZn-based catalysts. Appl. Catal. A Gen. 1997, 162, 281–297. [Google Scholar] [CrossRef]
  13. Azadi, P.; Otomo, J.; Hatano, H.; Oshima, Y.; Farnood, R. Hydrogen production by catalytic near-critical water gasification and steam reforming of glucose. Int. J. Hydrog. Energy 2010, 35, 3406–3414. [Google Scholar] [CrossRef]
  14. Bae, J.; Lee, S.; Kim, S.; Oh, J.; Choi, S.; Bae, M.; Kang, I.; Katikaneni, S.P. Liquid fuel processing for hydrogen production: A review. Int. J. Hydrog. Energy 2016, 41, 19990–20022. [Google Scholar] [CrossRef]
  15. Zhao, X.-l.; LÜ, Y.-a.; Liao, W.-p.; Jin, M.-s.; Suo, Z.-h. Hydrogen production from steam reforming of ethylene glycol over supported nickel catalysts. J. Fuel Chem. Technol. 2015, 43, 581–588. [Google Scholar] [CrossRef]
  16. Profeti, L.P.R.; Ticianelli, E.A.; Assaf, E.M. Production of hydrogen via steam reforming of biofuels on Ni/CeO2–Al2O3 catalysts promoted by noble metals. Int. J. Hydrog. Energy 2009, 34, 5049–5060. [Google Scholar] [CrossRef]
  17. Mierczynski, P.; Vasilev, K.; Mierczynska, A.; Maniukiewicz, W.; Szynkowska, M.I.; Maniecki, T.P. Bimetallic Au–Cu, Au–Ni catalysts supported on MWCNTs for oxy-steam reforming of methanol. Appl. Catal. B Environ. 2016, 185, 281–294. [Google Scholar] [CrossRef]
  18. Mierczynski, P.; Mosinska, M.; Maniukiewicz, W.; Vasilev, K.; Szynkowska, M.I. Novel Rh(Pd)-Cu(Ni) supported catalysts for oxy-steam reforming of methanol. Arab. J. Chem. 2020, 13, 3183–3195. [Google Scholar] [CrossRef]
  19. Mierczynski, P.; Mierczynska, A.; Maniukiewicz, W.; Maniecki, T.P.; Vasilev, K. MWCNTs as a catalyst in oxy-steam reforming of methanol. RSC Adv. 2016, 6, 81408–81413. [Google Scholar] [CrossRef]
  20. Barthos, R.; Solymosi, F. Hydrogen production in the decomposition and steam reforming of methanol on Mo2C/carbon catalysts. J. Catal. 2007, 249, 289–299. [Google Scholar] [CrossRef]
  21. Wang, J.B.; Li, C.-H.; Huang, T.-J. Study of Partial Oxidative Steam Reforming of Methanol over Cu–ZnO/samaria-doped Ceria Catalyst. Catal. Lett. 2005, 103, 239–247. [Google Scholar] [CrossRef]
  22. Mierczynski, P.; Mosinska, M.; Zakrzewski, M.; Dawid, B.; Ciesielski, R.; Maniukiewicz, W.; Maniecki, T. Influence of the Zn–Al binary oxide composition on the physicochemical and catalytic properties of Ni catalysts in the oxy-steam reforming of methanol. React. Kinet. Mech. Catal. 2017, 121, 453–472. [Google Scholar] [CrossRef] [Green Version]
  23. Mierczynski, P.; Mierczynska, A.; Ciesielski, R.; Mosinska, M.; Nowosielska, M.; Czylkowska, A.; Maniukiewicz, W.; Szynkowska, I.M.; Vasilev, K. High Active and Selective Ni/CeO2–Al2O3 and Pd–Ni/CeO2–Al2O3 Catalysts for Oxy-Steam Reforming of Methanol. Catalysts 2018, 8. [Google Scholar] [CrossRef] [Green Version]
  24. Geissler, K.; Newson, E.; Vogel, F.; Truong, T.-B.; Hottinger, P.; Wokaun, A. Autothermal methanol reforming for hydrogen production in fuel cell applications. Phys. Chem. Chem. Phys. 2001, 3, 289–293. [Google Scholar] [CrossRef]
  25. Abrokwah, R.Y.; Deshmane, V.G.; Kuila, D. Comparative performance of M-MCM-41 (M: Cu, Co, Ni, Pd, Zn and Sn) catalysts for steam reforming of methanol. J. Mol. Catal. A Chem. 2016, 425, 10–20. [Google Scholar] [CrossRef] [Green Version]
  26. Ahn, S.H.; Kwon, O.J.; Choi, I.; Kim, J.J. Synergetic effect of combined use of Cu–ZnO–Al2O3 and Pt–Al2O3 for the steam reforming of methanol. Catal. Commun. 2009, 10, 2018–2022. [Google Scholar] [CrossRef]
  27. Chang, C.-C.; Chang, C.-T.; Chiang, S.-J.; Liaw, B.-J.; Chen, Y.-Z. Oxidative steam reforming of methanol over CuO/ZnO/CeO2/ZrO2/Al2O3 catalysts. Int. J. Hydrog. Energy 2010, 35, 7675–7683. [Google Scholar] [CrossRef]
  28. Chang, C.-C.; Hsu, C.-C.; Chang, C.-T.; Chen, Y.-P.; Liaw, B.-J.; Chen, Y.-Z. Effect of noble metal on oxidative steam reforming of methanol over CuO/ZnO/Al2O3 catalysts. Int. J. Hydrog. Energy 2012, 37, 11176–11184. [Google Scholar] [CrossRef]
  29. Clancy, P.; Breen, J.P.; Ross, J.R.H. The preparation and properties of coprecipitated Cu–Zr–Y and Cu–Zr–La catalysts used for the steam reforming of methanol. Catal. Today 2007, 127, 291–294. [Google Scholar] [CrossRef]
  30. Jeong, H.; Kim, K.I.; Kim, T.H.; Ko, C.H.; Park, H.C.; Song, I.K. Hydrogen production by steam reforming of methanol in a micro-channel reactor coated with Cu/ZnO/ZrO2/Al2O3 catalyst. J. Power Sources 2006, 159, 1296–1299. [Google Scholar] [CrossRef]
  31. Das, D.; Llorca, J.; Dominguez, M.; Colussi, S.; Trovarelli, A.; Gayen, A. Methanol steam reforming behavior of copper impregnated over CeO2–ZrO2 derived from a surfactant assisted coprecipitation route. Int. J. Hydrog. Energy 2015, 40, 10463–10479. [Google Scholar] [CrossRef] [Green Version]
  32. Yong-Feng, L.; Xin-Fa, D.; Wei-Ming, L. Effects of ZrO2-promoter on catalytic performance of CuZnAlO catalysts for production of hydrogen by steam reforming of methanol. Int. J. Hydrog. Energy 2004, 29, 1617–1621. [Google Scholar] [CrossRef]
  33. Lindström, B.; Pettersson, L.J.; Govind Menon, P. Activity and characterization of Cu/Zn, Cu/Cr and Cu/Zr on γ-alumina for methanol reforming for fuel cell vehicles. Appl. Catal. A Gen. 2002, 234, 111–125. [Google Scholar] [CrossRef]
  34. Ilinich, O.M.; Liu, Y.; Waterman, E.M.; Farrauto, R.J. Kinetics of Methanol Steam Reforming with a Pd–Zn–Y/CeO2 Catalyst under Realistic Operating Conditions of a Portable Reformer in Fuel Cell Applications. Ind. Eng. Chem. Res. 2013, 52, 638–644. [Google Scholar] [CrossRef]
  35. Sá, S.; Silva, H.; Brandão, L.; Sousa, J.M.; Mendes, A. Catalysts for methanol steam reforming—A review. Appl. Catal. B Environ. 2010, 99, 43–57. [Google Scholar] [CrossRef]
  36. Talkhoncheh, S.K.; Haghighi, M.; Minaei, S.; Ajamein, H.; Abdollahifar, M. Synthesis of CuO/ZnO/Al2O3/ZrO2/CeO2 nanocatalysts via homogeneous precipitation and combustion methods used in methanol steam reforming for fuel cell grade hydrogen production. RSC Adv. 2016, 6, 57199–57209. [Google Scholar] [CrossRef]
  37. Luengnaruemitchai, A.; Pojanavaraphan, C.; Kumyam, A.; Thunyaratchatanon, C.; Gulari, E. Hydrogen production from the oxidative steam reforming of methanol over AuCu nanoparticles supported on Ce1-xZrxO2 in a fixed-bed reactor. Int. J. Hydrog. Energy 2019, 44, 1686–1700. [Google Scholar] [CrossRef]
  38. Lu, J.; Li, X.; He, S.; Han, C.; Wan, G.; Lei, Y.; Chen, R.; Liu, P.; Chen, K.; Zhang, L.; et al. Hydrogen production via methanol steam reforming over Ni-based catalysts: Influences of Lanthanum (La) addition and supports. Int. J. Hydrog. Energy 2017, 42, 3647–3657. [Google Scholar] [CrossRef]
  39. Papavasiliou, J.; Avgouropoulos, G.; Ioannides, T. Combined steam reforming of methanol over Cu–Mn spinel oxide catalysts. J. Catal. 2007, 251, 7–20. [Google Scholar] [CrossRef]
  40. Ma, L.; Gong, B.; Tran, T.; Wainwright, M.S. Cr2O3 promoted skeletal Cu catalysts for the reactions of methanol steam reforming and water gas shift. Catal. Today 2000, 63, 499–505. [Google Scholar] [CrossRef]
  41. Yao, C.-Z.; Wang, L.-C.; Liu, Y.-M.; Wu, G.-S.; Cao, Y.; Dai, W.-L.; He, H.-Y.; Fan, K.-N. Effect of preparation method on the hydrogen production from methanol steam reforming over binary Cu/ZrO2 catalysts. Appl. Catal. A Gen. 2006, 297, 151–158. [Google Scholar] [CrossRef]
  42. Shen, J.-P.; Song, C. Influence of preparation method on performance of Cu/Zn-based catalysts for low-temperature steam reforming and oxidative steam reforming of methanol for H2 production for fuel cells. Catal. Today 2002, 77, 89–98. [Google Scholar] [CrossRef]
  43. Shishido, T.; Yamamoto, Y.; Morioka, H.; Takaki, K.; Takehira, K. Active Cu/ZnO and Cu/ZnO/Al2O3 catalysts prepared by homogeneous precipitation method in steam reforming of methanol. Appl. Catal. A Gen. 2004, 263, 249–253. [Google Scholar] [CrossRef]
  44. Larrubia Vargas, M.A.; Busca, G.; Costantino, U.; Marmottini, F.; Montanari, T.; Patrono, P.; Pinzari, F.; Ramis, G. An IR study of methanol steam reforming over ex-hydrotalcite Cu–Zn–Al catalysts. J. Mol. Catal. A Chem. 2007, 266, 188–197. [Google Scholar] [CrossRef]
  45. Peppley, B.A.; Amphlett, J.C.; Kearns, L.M.; Mann, R.F. Methanol–steam reforming on Cu/ZnO/Al2O3 catalysts. Part 2. A comprehensive kinetic model. Appl. Catal. A Gen. 1999, 179, 31–49. [Google Scholar] [CrossRef]
  46. Liu, S.; Takahashi, K.; Uematsu, K.; Ayabe, M. Hydrogen production by oxidative methanol reforming on Pd/ZnO. Appl. Catal. A Gen. 2005, 283, 125–135. [Google Scholar] [CrossRef]
  47. Takahashi, K.; Takezawa, N.; Kobayashi, H. The mechanism of steam reforming of methanol over a copper-silica catalyst. Appl. Catal. 1982, 2, 363–366. [Google Scholar] [CrossRef]
  48. Turco, M.; Bagnasco, G.; Costantino, U.; Marmottini, F.; Montanari, T.; Ramis, G.; Busca, G. Production of hydrogen from oxidative steam reforming of methanol: II. Catalytic activity and reaction mechanism on Cu/ZnO/Al2O3 hydrotalcite-derived catalysts. J. Catal. 2004, 228, 56–65. [Google Scholar] [CrossRef]
  49. Turco, M.; Bagnasco, G.; Cammarano, C.; Senese, P.; Costantino, U.; Sisani, M. Cu/ZnO/Al2O3 catalysts for oxidative steam reforming of methanol: The role of Cu and the dispersing oxide matrix. Appl. Catal. B Environ. 2007, 77, 46–57. [Google Scholar] [CrossRef]
  50. Yong, S.T.; Ooi, C.W.; Chai, S.P.; Wu, X.S. Review of methanol reforming-Cu-based catalysts, surface reaction mechanisms, and reaction schemes. Int. J. Hydrog. Energy 2013, 38, 9541–9552. [Google Scholar] [CrossRef]
  51. Turco, G.; Donati, I.; Grassi, M.; Marchioli, G.; Lapasin, R.; Paoletti, S. Mechanical Spectroscopy and Relaxometry on Alginate Hydrogels: A Comparative Analysis for Structural Characterization and Network Mesh Size Determination. Biomacromolecules 2011, 12, 1272–1282. [Google Scholar] [CrossRef]
  52. Turco, M.; Bagnasco, G.; Costantino, U.; Marmottini, F.; Montanari, T.; Ramis, G.; Busca, G. Production of hydrogen from oxidative steam reforming of methanol: I. Preparation and characterization of Cu/ZnO/Al2O3 catalysts from a hydrotalcite-like LDH precursor. J. Catal. 2004; 228, 43–55. [Google Scholar] [CrossRef]
  53. Reitz, T.L.; Lee, P.L.; Czaplewski, K.F.; Lang, J.C.; Popp, K.E.; Kung, H.H. Time-Resolved XANES Investigation of CuO/ZnO in the Oxidative Methanol Reforming Reaction. J. Catal. 2001, 199, 193–201. [Google Scholar] [CrossRef] [Green Version]
  54. Mrad, M.; Gennequin, C.; Aboukaïs, A.; Abi-Aad, E. Cu/Zn-based catalysts for H2 production via steam reforming of methanol. Catal. Today 2011, 176, 88–92. [Google Scholar] [CrossRef]
  55. Sexton, B.A. Surface vibrations of adsorbed intermediates in the reaction of alcohols with Cu(100). Surf. Sci. 1979, 88, 299–318. [Google Scholar] [CrossRef]
  56. Wachs, I.E.; Madix, R.J. The selective oxidation of CH3OH to H2CO on a copper(110) catalyst. J. Catal. 1978, 53, 208–227. [Google Scholar] [CrossRef]
  57. Russell, J.N.; Gates, S.M.; Yates, J.T. Reaction of methanol with Cu(111) and Cu(111) + O(ads). Surf. Sci. 1985, 163, 516–540. [Google Scholar] [CrossRef]
  58. Chinchen, G.C.; Spencer, M.S.; Waugh, K.C.; Whan, D.A. Promotion of methanol synthesis and the water-gas shift reactions by adsorbed oxygen on supported copper catalysts. J. Chem. Soc. Faraday Trans. 1 Phys. Chem. Condens. Phases 1987, 83, 2193–2212. [Google Scholar] [CrossRef]
  59. Lwin, Y.; Daud, W.R.W.; Mohamad, A.B.; Yaakob, Z. Hydrogen production from steam–methanol reforming: Thermodynamic analysis. Int. J. Hydrog. Energy 2000, 25, 47–53. [Google Scholar] [CrossRef]
  60. Santacesaria, E.; Carrá, S. Kinetics of catalytic steam reforming of methanol in a cstr reactor. Appl. Catal. 1983, 5, 345–358. [Google Scholar] [CrossRef]
  61. Breen, J.P.; Meunier, F.C.; Ross, J.R.H. Mechanistic aspects of the steam reforming of methanol over a CuO/ZnO/ZrO2/Al2O3 catalyst. Chem. Commun. 1999, 2247–2248. [Google Scholar] [CrossRef]
  62. Agrell, J.; Birgersson, H.; Boutonnet, M. Steam reforming of methanol over a Cu/ZnO/Al2O3 catalyst: a kinetic analysis and strategies for suppression of CO formation. J. Power Sources 2002, 106, 249–257. [Google Scholar] [CrossRef]
  63. Mastalir, A.; Frank, B.; Szizybalski, A.; Soerijanto, H.; Deshpande, A.; Niederberger, M.; Schomäcker, R.; Schlögl, R.; Ressler, T. Steam reforming of methanol over Cu/ZrO2/CeO2 catalysts: a kinetic study. J. Catal. 2005, 230, 464–475. [Google Scholar] [CrossRef] [Green Version]
  64. Gallucci, F.; Paturzo, L.; Basile, A. Hydrogen Recovery from Methanol Steam Reforming in a Dense Membrane Reactor:  Simulation Study. Ind. Eng. Chem. Res. 2004, 43, 2420–2432. [Google Scholar] [CrossRef]
  65. Morillo, A.; Freund, A.; Merten, C. Concept and Design of a Novel Compact Reactor for Autothermal Steam Reforming with Integrated Evaporation and CO Cleanup. Ind. Eng. Chem. Res. 2004, 43, 4624–4634. [Google Scholar] [CrossRef]
  66. Jiang, C.J.; Trimm, D.L.; Wainwright, M.S.; Cant, N.W. Kinetic study of steam reforming of methanol over copper-based catalysts. Appl. Catal. A Gen. 1993, 93, 245–255. [Google Scholar] [CrossRef]
  67. Breen, J.P.; Ross, J.R.H. Methanol reforming for fuel-cell applications: development of zirconia-containing Cu–Zn–Al catalysts. Catal. Today 1999, 51, 521–533. [Google Scholar] [CrossRef]
  68. Iwasa, N.; Kudo, S.; Takahashi, H.; Masuda, S.; Takezawa, N. Highly selective supported Pd catalysts for steam reforming of methanol. Catal. Lett. 1993, 19, 211–216. [Google Scholar] [CrossRef]
  69. Amiri, T.Y.; Moghaddas, J. Reaction parameters influence on the catalytic performance of copper-silica aerogel in the methanol steam reforming. J. Fuel Chem. Technol. 2016, 44, 84–90. [Google Scholar] [CrossRef]
  70. Murcia-Mascarós, S.; Navarro, R.M.; Gómez-Sainero, L.; Costantino, U.; Nocchetti, M.; Fierro, J.L.G. Oxidative Methanol Reforming Reactions on CuZnAl Catalysts Derived from Hydrotalcite-like Precursors. J. Catal. 2001, 198, 338–347. [Google Scholar] [CrossRef]
  71. Rabe, S.; Vogel, F. A thermogravimetric study of the partial oxidation of methanol for hydrogen production over a Cu/ZnO/Al2O3 catalyst. Appl. Catal. B Environ. 2008, 84, 827–834. [Google Scholar] [CrossRef]
  72. Patel, S.; Pant, K.K. Kinetic modeling of oxidative steam reforming of methanol over Cu/ZnO/CeO2/Al2O3 catalyst. Appl. Catal. A Gen. 2009, 356, 189–200. [Google Scholar] [CrossRef]
  73. Patel, S.; Pant, K.K. Hydrogen production by oxidative steam reforming of methanol using ceria promoted copper–alumina catalysts. Fuel Process. Technol. 2007, 88, 825–832. [Google Scholar] [CrossRef]
  74. Horny, C.; Renken, A.; Kiwi-Minsker, L. Compact string reactor for autothermal hydrogen production. Catal. Today 2007, 120, 45–53. [Google Scholar] [CrossRef] [Green Version]
  75. Horng, R.-F.; Chou, H.-M.; Lee, C.-H.; Tsai, H.-T. Characteristics of hydrogen produced by partial oxidation and auto-thermal reforming in a small methanol reformer. J. Power Sources 2006, 161, 1225–1233. [Google Scholar] [CrossRef]
  76. Riva, A.; Trifirò, F.; Vaccari, A.; Mintchev, L.; Busca, G. Structure and reactivity of zinc–chromium mixed oxides. Part 2.—Study of the surface reactivity by temperature-programmed desorption of methanol. J. Chem. Soc. Faraday Trans. 1 Phys. Chem. Condens. Phases 1988, 84, 1423–1435. [Google Scholar] [CrossRef]
  77. Peppley, B.A.; Amphlett, J.C.; Kearns, L.M.; Mann, R.F. Methanol–steam reforming on Cu/ZnO/Al2O3. Part 1: the reaction network. Appl. Catal. A Gen. 1999, 179, 21–29. [Google Scholar] [CrossRef]
  78. Shan, W.; Feng, Z.; Li, Z.; Zhang, J.; Shen, W.; Li, C. Oxidative steam reforming of methanol on Ce0.9Cu0.1OY catalysts prepared by deposition–precipitation, coprecipitation, and complexation–combustion methods. J. Catal. 2004, 228, 206–217. [Google Scholar] [CrossRef]
  79. Papavasiliou, J.; Avgouropoulos, G.; Ioannides, T. Production of hydrogen via combined steam reforming of methanol over CuO–CeO2 catalysts. Catal. Commun. 2004, 5, 231–235. [Google Scholar] [CrossRef]
  80. Shishido, T.; Yamamoto, Y.; Morioka, H.; Takehira, K. Production of hydrogen from methanol over Cu/ZnO and Cu/ZnO/Al2O3 catalysts prepared by homogeneous precipitation: Steam reforming and oxidative steam reforming. J. Mol. Catal. A Chem. 2007, 268, 185–194. [Google Scholar] [CrossRef] [Green Version]
  81. Ritzkopf, I.; Vukojević, S.; Weidenthaler, C.; Grunwaldt, J.-D.; Schüth, F. Decreased CO production in methanol steam reforming over Cu/ZrO2 catalysts prepared by the microemulsion technique. Appl. Catal. A Gen. 2006, 302, 215–223. [Google Scholar] [CrossRef]
  82. Wang, C.-T.; Willey, R.J. Fine particle iron oxide based aerogels for the partial oxidation of methanol. Catal. Today 1999, 52, 83–89. [Google Scholar] [CrossRef]
  83. Matsumura, Y.; Tanaka, K.; Tode, N.; Yazawa, T.; Haruta, M. Catalytic methanol decomposition to carbon monoxide and hydrogen over nickel supported on silica. J. Mol. Catal. A Chem. 2000, 152, 157–165. [Google Scholar] [CrossRef]
  84. Dinka, P.; Mukasyan, A. Solution combustion synthesis of nano materials. In Proceedings of the 2006 NSTI Nanotechnology Conference and Trade Show-NSTI Nanotech 2006 Technical Proceedings, Boston, MA, USA, 7–11 May 2006; pp. 456–459. [Google Scholar]
  85. Liu, S.; Takahashi, K.; Ayabe, M. Hydrogen production by oxidative methanol reforming on Pd/ZnO catalyst: effects of Pd loading. Catal. Today 2003, 87, 247–253. [Google Scholar] [CrossRef]
  86. Haynes, D.J.; Shekhawat, D.; Spivey, J.J.; Berry, D.A. Chapter 6—Oxidative Steam Reforming. In Fuel Cells; Elsevier: Amsterdam, The Netherlands, 2011; pp. 129–190. [Google Scholar]
  87. Pijolat, M.; Prin, M.; Soustelle, M.; Touret, O.; Nortier, P. Thermal stability of doped ceria: experiment and modelling. J. Chem. Soc. Faraday Trans. 1995, 91, 3941–3948. [Google Scholar] [CrossRef]
  88. Fornasiero, P.; Balducci, G.; di Monte, R.; Kašpar, J.; Sergo, V.; Gubitosa, G.; Ferrero, A.; Graziani, M. Modification of the Redox Behaviour of CeO2Induced by Structural Doping with ZrO2. J. Catal. 1996, 164, 173–183. [Google Scholar] [CrossRef]
  89. Fan, L.; Fujimoto, K. Reaction Mechanism of Methanol Synthesis from Carbon Dioxide and Hydrogen on Ceria-Supported Palladium Catalysts with SMSI Effect. J. Catal. 1997, 172, 238–242. [Google Scholar] [CrossRef]
  90. Zhang, X.; Shi, P. Production of hydrogen by steam reforming of methanol on CeO2 promoted Cu/Al2O3 catalysts. J. Mol. Catal. A Chem. 2003, 194, 99–105. [Google Scholar] [CrossRef]
  91. Mierczynski, P.; Mosinska, M.; Maniukiewicz, W.; Nowosielska, M.; Czylkowska, A.; Szynkowska, M.I. Oxy-steam reforming of methanol on copper catalysts. React. Kinet. Mech. Catal. 2019, 127, 857–874. [Google Scholar] [CrossRef] [Green Version]
  92. Iwasa, N.; Yoshikawa, M.; Nomura, W.; Arai, M. Transformation of methanol in the presence of steam and oxygen over ZnO-supported transition metal catalysts under stream reforming conditions. Appl. Catal. A Gen. 2005, 292, 215–222. [Google Scholar] [CrossRef]
  93. Iwasa, N.; Masuda, S.; Ogawa, N.; Takezawa, N. Steam reforming of methanol over Pd/ZnO: Effect of the formation of PdZn alloys upon the reaction. Appl. Catal. A Gen. 1995, 125, 145–157. [Google Scholar] [CrossRef]
  94. Iwasa, N.; Takezawa, N. New Supported Pd and Pt Alloy Catalysts for Steam Reforming and Dehydrogenation of Methanol. Top. Catal. 2003, 22, 215–224. [Google Scholar] [CrossRef]
  95. Kapoor, M.P.; Ichihashi, Y.; Kuraoka, K.; Matsumura, Y. Catalytic methanol decomposition over palladium deposited on thermally stable mesoporous titanium oxide. J. Mol. Catal. A Chem. 2003, 198, 303–308. [Google Scholar] [CrossRef]
  96. Kapoor, M.P.; Ichihashi, Y.; Kuraoka, K.; Shen, W.-J.; Matsumura, Y. Catalytic Methanol Decomposition Over Palladium Deposited on Mesoporous Cerium Oxide. Catal. Lett. 2003, 88, 83–87. [Google Scholar] [CrossRef]
  97. Lenarda, M.; Moretti, E.; Storaro, L.; Patrono, P.; Pinzari, F.; Rodríguez-Castellón, E.; Jiménez-López, A.; Busca, G.; Finocchio, E.; Montanari, T.; et al. Finely dispersed Pd-Zn catalyst supported on an organized mesoporous alumina for hydrogen production by methanol steam reforming. Appl. Catal. A Gen. 2006, 312, 220–228. [Google Scholar] [CrossRef]
  98. Usami, Y.; Kagawa, K.; Kawazoe, M.; Yasuyuki, M.; Sakurai, H.; Haruta, M. Catalytic methanol decomposition at low temperatures over palladium supported on metal oxides. Appl. Catal. A Gen. 1998, 171, 123–130. [Google Scholar] [CrossRef]
  99. Udani, P.P.C.; Gunawardana, P.V.D.S.; Lee, H.C.; Kim, D.H. Steam reforming and oxidative steam reforming of methanol over CuO–CeO2 catalysts. Int. J. Hydrog. Energy 2009, 34, 7648–7655. [Google Scholar] [CrossRef]
  100. Serp, P.; Corrias, M.; Kalck, P. Carbon nanotubes and nanofibers in catalysis. Appl. Catal. A Gen. 2003, 253, 337–358. [Google Scholar] [CrossRef]
  101. Seelam, P.K.; Huuhtanen, M.; Sápi, A.; Szabó, M.; Kordás, K.; Turpeinen, E.; Tóth, G.; Keiski, R.L. CNT-based catalysts for H2 production by ethanol reforming. Int. J. Hydrog. Energy 2010, 35, 12588–12595. [Google Scholar] [CrossRef]
  102. Pham-Huu, C.; Keller, N.; Ehret, G.; Charbonniere, L.c.J.; Ziessel, R.; Ledoux, M.J. Carbon nanofiber supported palladium catalyst for liquid-phase reactions: An active and selective catalyst for hydrogenation of cinnamaldehyde into hydrocinnamaldehyde. J. Mol. Catal. A Chem. 2001, 170, 155–163. [Google Scholar] [CrossRef]
  103. Hou, T.; Yuan, L.; Ye, T.; Gong, L.; Tu, J.; Yamamoto, M.; Torimoto, Y.; Li, Q. Hydrogen production by low-temperature reforming of organic compounds in bio-oil over a CNT-promoting Ni catalyst. Int. J. Hydrog. Energy 2009, 34, 9095–9107. [Google Scholar] [CrossRef]
  104. Solhy, A.; Machado, B.F.; Beausoleil, J.; Kihn, Y.; Gonçalves, F.; Pereira, M.F.R.; Órfão, J.J.M.; Figueiredo, J.L.; Faria, J.L.; Serp, P. MWCNT activation and its influence on the catalytic performance of Pt/MWCNT catalysts for selective hydrogenation. Carbon 2008, 46, 1194–1207. [Google Scholar] [CrossRef]
  105. Tang, J.M.; Jensen, K.; Waje, M.; Li, W.; Larsen, P.; Pauley, K.; Chen, Z.; Ramesh, P.; Itkis, M.E.; Yan, Y.; et al. High Performance Hydrogen Fuel Cells with Ultralow Pt Loading Carbon Nanotube Thin Film Catalysts. J. Phys. Chem. C 2007, 111, 17901–17904. [Google Scholar] [CrossRef]
  106. Liu, Z.-T.; Wang, C.-X.; Liu, Z.-W.; Lu, J. Selective hydrogenation of cinnamaldehyde over Pt-supported multi-walled carbon nanotubes: Insights into the tube-size effects. Appl. Catal. A Gen. 2008, 344, 114–123. [Google Scholar] [CrossRef]
  107. Agrell, J.; Birgersson, H.; Boutonnet, M.; Melián-Cabrera, I.; Navarro, R.M.; Fierro, J.L.G. Production of hydrogen from methanol over Cu/ZnO catalysts promoted by ZrO2 and Al2O3. J. Catal. 2003, 219, 389–403. [Google Scholar] [CrossRef]
  108. Huang, X.; Ma, L.; Wainwright, M.S. The influence of Cr, Zn and Co additives on the performance of skeletal copper catalysts for methanol synthesis and related reactions. Appl. Catal. A Gen. 2004, 257, 235–243. [Google Scholar] [CrossRef]
  109. Liu, S.; Takahashi, K.; Uematsu, K.; Ayabe, M. Hydrogen production by oxidative methanol reforming on Pd/ZnO catalyst: effects of the addition of a third metal component. Appl. Catal. A Gen. 2004, 277, 265–270. [Google Scholar] [CrossRef]
  110. Haruta, M.; Tsubota, S.; Kobayashi, T.; Kageyama, H.; Genet, M.J.; Delmon, B. Low-Temperature Oxidation of CO over Gold Supported on TiO2, α-Fe2O3, and Co3O4. J. Catal. 1993, 144, 175–192. [Google Scholar] [CrossRef]
  111. Mierczynski, P.; Vasilev, K.; Mierczynska, A.; Maniukiewicz, W.; Ciesielski, R.; Rogowski, J.; Szynkowska, I.M.; Trifonov, A.Y.; Dubkov, S.V.; Gromov, D.G.; et al. The effect of gold on modern bimetallic Au–Cu/MWCNT catalysts for the oxy-steam reforming of methanol. Catal. Sci. Technol. 2016, 6, 4168–4183. [Google Scholar] [CrossRef] [Green Version]
  112. Mierczynski, P.; Ciesielski, R.; Kedziora, A.; Nowosielska, M.; Kubicki, J.; Maniukiewicz, W.; Czylkowska, A.; Maniecki, T.P. Monometallic copper catalysts supported on multi-walled carbon nanotubes for the oxy-steam reforming of methanol. React. Kinet. Mech. Catal. 2016, 117, 675–691. [Google Scholar] [CrossRef]
  113. Chin, Y.-H.; Dagle, R.; Hu, J.; Dohnalkova, A.C.; Wang, Y. Steam reforming of methanol over highly active Pd/ZnO catalyst. Catal. Today 2002, 77, 79–88. [Google Scholar] [CrossRef]
  114. Cubeiro, M.L.; Fierro, J.L.G. Partial oxidation of methanol over supported palladium catalysts. Appl. Catal. A Gen. 1998, 168, 307–322. [Google Scholar] [CrossRef]
  115. Cubeiro, M.L.; Fierro, J.L.G. Selective Production of Hydrogen by Partial Oxidation of Methanol over ZnO-Supported Palladium Catalysts. J. Catal. 1998, 179, 150–162. [Google Scholar] [CrossRef]
  116. Xu, J.B.; Zhao, T.S.; Yang, W.W.; Shen, S.Y. Effect of surface composition of Pt-Au alloy cathode catalyst on the performance of direct methanol fuel cells. Int. J. Hydrog. Energy 2010, 35, 8699–8706. [Google Scholar] [CrossRef]
  117. Manzoli, M.; Chiorino, A.; Boccuzzi, F. Decomposition and combined reforming of methanol to hydrogen: A FTIR and QMS study on Cu and Au catalysts supported on ZnO and TiO2. Appl. Catal. B Environ. 2005, 57, 201–209. [Google Scholar] [CrossRef]
  118. Vizcaíno, A.; Carrero, A.; Calles, J. Hydrogen production by ethanol steam reforming over Cu-Ni supported catalysts. Int. J. Hydrog. Energy 2007, 32, 1450–1461. [Google Scholar] [CrossRef]
  119. Khzouz, M.; Wood, J.; Pollet, B.; Bujalski, W. Characterization and activity test of commercial Ni/Al2O3, Cu/ZnO/Al2O3 and prepared Ni–Cu/Al2O3 catalysts for hydrogen production from methane and methanol fuels. Int. J. Hydrog. Energy 2013, 38, 1664–1675. [Google Scholar] [CrossRef] [Green Version]
  120. De Rogatis, L.; Montini, T.; Lorenzut, B.; Fornasiero, P. NixCuy/Al2O3 based catalysts for hydrogen production. Energy Environ. Sci. 2008, 1, 501–509. [Google Scholar] [CrossRef]
  121. Pérez-Hernández, R.; Mondragón Galicia, G.; Mendoza Anaya, D.; Palacios, J.; Angeles-Chavez, C.; Arenas-Alatorre, J. Synthesis and characterization of bimetallic Cu–Ni/ZrO2 nanocatalysts: H2 production by oxidative steam reforming of methanol. Int. J. Hydrog. Energy 2008, 33, 4569–4576. [Google Scholar] [CrossRef]
  122. Eaimsumang, S.; Chollacoop, N.; Luengnaruemitchai, A.; Taylor, S.H. Ceria nanorod supported gold nanoparticles as structured catalysts for the oxidative steam reforming of methanol: Effect of CTAB concentration on physiochemical properties and catalyst performance. J. Catal. 2020, 392, 254–265. [Google Scholar] [CrossRef]
  123. Wang, C.; Liu, N.; Pan, L.; Wang, S.; Yuan, Z.; Wang, S. Measurement of concentration profiles over ZnO–Cr2O3/CeO2–ZrO2 monolithic catalyst in oxidative steam reforming of methanol. Fuel Process. Technol. 2007, 88, 65–71. [Google Scholar] [CrossRef]
Figure 1. Hydrogen as an energy carrier.
Figure 1. Hydrogen as an energy carrier.
Materials 13 05601 g001
Figure 2. Direct methanol fuel cells (DMFC).
Figure 2. Direct methanol fuel cells (DMFC).
Materials 13 05601 g002
Figure 3. The possible reactions for the autothermal reforming of methanol process [24].
Figure 3. The possible reactions for the autothermal reforming of methanol process [24].
Materials 13 05601 g003
Figure 4. The reaction steps taking place during the steam-reforming of methanol process [69].
Figure 4. The reaction steps taking place during the steam-reforming of methanol process [69].
Materials 13 05601 g004
Figure 5. The reaction network taking place during oxidative steam-reforming of methanol (OSRM) process [48].
Figure 5. The reaction network taking place during oxidative steam-reforming of methanol (OSRM) process [48].
Materials 13 05601 g005
Figure 6. The methanol decomposition (MD) mechanism [60].
Figure 6. The methanol decomposition (MD) mechanism [60].
Materials 13 05601 g006
Figure 7. XPS and stability results during 48 h of the OSRM reaction performed at 250 °C after reduction 1 h at 300 °C [23].
Figure 7. XPS and stability results during 48 h of the OSRM reaction performed at 250 °C after reduction 1 h at 300 °C [23].
Materials 13 05601 g007
Figure 8. The activity results of Rh(Pd)-Cu(Ni)/ZrO2·Al2O3 (1:2) catalysts in the OSRM process [18].
Figure 8. The activity results of Rh(Pd)-Cu(Ni)/ZrO2·Al2O3 (1:2) catalysts in the OSRM process [18].
Materials 13 05601 g008
Figure 9. Phase composition studies of spent catalysts and reactivity results of bimetallic Au–Cu and Au–Ni system in the OSRM process [17].
Figure 9. Phase composition studies of spent catalysts and reactivity results of bimetallic Au–Cu and Au–Ni system in the OSRM process [17].
Materials 13 05601 g009
Figure 10. The species formed during the combined reforming of methanol process [117].
Figure 10. The species formed during the combined reforming of methanol process [117].
Materials 13 05601 g010
Figure 11. XRD patterns of spent bimetallic Cu–Ni systems [4].
Figure 11. XRD patterns of spent bimetallic Cu–Ni systems [4].
Materials 13 05601 g011
Table 1. The coefficients of combination for two models of autothermal reforming of methanol process (ATRM) [24].
Table 1. The coefficients of combination for two models of autothermal reforming of methanol process (ATRM) [24].
POM-SRMTOM-SRM
SRMCH3OH + H2O CO2 + 3H23 3 2 3
POMCH3OH + 1 2 O2 CO2 + 2H21-
TOMCH3OH + 3 2 O2 CO2 + 2H2O- 1 3
A T R M 4CH3OH + 3H2O + 1 2 O2 + 2N2 4CO2 + 11H2 + 2N2
Table 2. Activation energy for the oxy–steam-reforming of methanol process [49].
Table 2. Activation energy for the oxy–steam-reforming of methanol process [49].
CatalystEa (cal mol−1)
Cu(5)Zn(50)Al(45)24
Cu(15)Zn(48)Al(37)27
Cu(18)Zn(33)Al(49)16
Cu(45)Zn(31)Al(24)22
Cu(75)Zn(25)Al(0)-
Table 3. Reaction conditions and catalytic materials applied in the studied process.
Table 3. Reaction conditions and catalytic materials applied in the studied process.
ReferencesInvestigated MechanismEmployed
Catalyst
Operating Conditions
Total Flow
[cm3/min]
Temperature
[K]
Pressure
[atm]
H2O/CH3OHO2/CH3OH
[56]Selective oxidation of CH3OH to H2COCopper (110)-295---
[59]SRMCopper containing catalyst-360–57311.5-
[60]SRMCu/ZnO/Al2O3-433–4731--
[61]SRMCuO/ZnO/ZrO2/Al2O3-473 and 57311.3-
[62]SRMCu/ZnO/Al2O3230448–62311.3-
[77]SRMCu/ZnO/Al2O350433–5331–350–1.2-
[45]SRMCu/ZnO/Al2O350433–5331–350–1.2-
[63]SRMCu/ZrO2/CeO2-52311.0-
[66]SRMCu/ZnO/Al2O3-443–5331--
[67]SRMCu/Zn/Zr/Al38.6413–61811.3-
[47]SRMcopper–silica-433 and 4531--
[69]SRMcopper–silica aerogel-423–67312.0-
[70]OSRMCu/ZnO/Al2O3120473–63311.10.3
[71]POMCu/ZnO/Al2O3-453 and 493--0.1–0.5
[72]OSRMCu/ZrO2/CeO2/Al2O3-473–57311.50.1–0.2
[73]OSRMCu/CeO2/Al2O3-473–57311.50–0.5
[74]OSRMCu/ZnO/Al2O3
Cu/ZnO/Al2O3/Cr2O3
100538–5481.2810.9
[48]OSRMCu/ZnO/Al2O3-473–673-1.10.12
Table 4. The physicochemical and catalytic characterization of the various catalysts applied in OSR of methanol reaction.
Table 4. The physicochemical and catalytic characterization of the various catalysts applied in OSR of methanol reaction.
CatalystPreparation MethodSBET
(m2/g)
Metal Dispersion
(%)
Reduction
Temp.
(°C)
Reduction
Time
(h)
H 2 O methanol O 2 methanol W/F
(gscm−3)
GHSV
(h−1)
TOSR
(°C)
Catalyst
Weight
(g)
CH3OH
Conv.
(%)
H2COCO2
CuO/ZnO/Al2O3 [42]CP94-350–45021.430.47--2300.3100.071.0 y0.1 y28.9 y
Cu0.30Mn0.70 [39]UNC8-32021.260.100.257-2400.3100.097.0 s3.0 s-
CuO–CeO2 (mol.rat. = 0.15, u/n rat. = 1.2) [79]UNC10---1.500.100.257-3000.356.090.3 s2.1 s-
CuO–CeO2 (mol.rat. = 0.15, u/n rat. = 1.2) [79]UNC10---1.500.100.257-3400.35.775.0 s0.8 s-
CuO–CeO2 (mol.rat. = 0.15, u/n rat. = 2.38) [79]UNC5---1.500.100.257-3000.345.084.3 s4.8 s-
CuO–CeO2 (mol.rat. = 0.15, u/n rat. = 2.38) [79]UNC5---1.500.100.257-2400.312.069.0 s3.1 s-
CuO–CeO2 (mol.rat. = 0.15, u/n rat. = 3.30) [79]UNC20---1.500.100.257-3000.395.096.6 s3.4 s-
CuO–CeO2 (mol.rat. = 0.15, u/n rat. = 3.30) [79]UNC20---1.500.100.257-2400.332.492.6 s1.0 s-
CuO–CeO2 (mol.rat. = 0.15, u/n rat. = 4.17) [79]UNC43---1.500.100.257-3000.310095.9 s4.0 s-
CuO–CeO2 (mol.rat. = 0.15, u/n rat. = 4.17) [79]UNC43---1.500.100.257-2400.336.593.7 s0.8 s-
Cu(5)/CeO2·Al2O3 [91]IP1262.14 *3001---26,7002000.297.769.1 s0 s30.6 s
Cu(20)/CeO2·Al2O3 [91]IP1010.16 *3001---26,7002000.299.668.2 s0 s31.7 s
Cu(40)/CeO2·Al2O3 [91]IP900.13 *3001---26,7002000.287.965.7 s0 s33.8 s
Cu(60)/CeO2·Al2O3 [91]IP30-3001---26,7002000.29.628.5 s0 s71.5 s
Ni(5)/CeO2·Al2O3 [23]IP1321.06 *3001---26,7002500.2230.5 s-25 s
Ni(20)/CeO2·Al2O3 [23]IP1281.16 *3001---26,7002500.23133 s-55.6 s
Ni(40)/CeO2·Al2O3 [23]IP780.56 *3001---26,7002500.29868.9 s13.3 s14.3 s
Ni(60)/CeO2·Al2O3 [23]IP1330.69 *3001---26,7002500.27758.2 s27.7 s14.1 s
Ni(40)/CeO2 [23]IP340.63 *3001---26,7002500.28568.6 s22.6 s0.5 s
Ni(40)/Al2O3 [23]IP580.28 *3001---26,7002500.220.553.4 s-25.2 s
Pd(2)–Ni(40)/CeO2·Al2O3 [23]SIP42-3001---26,7002500.299.971.5 s14.9 s9.1 s
Pd/ZnO [85]IP--40021.500.10-110,0002500.3----
Pd/ZnO [85]CP--40021.500.10-110,0002500.3----
Cu/ZnO [107]CP499.6 *250–30011.300.20--3000.590.050.0 m0.07 m20.0 m
Cu/ZnO/Al2O3 [107]CP9211.3 *250–30011.300.20--3250.590.0-0.13 m-
Cu/ZnO/ZrO2 [107]CP8213.2 *250–30011.300.20--2950.590.0-0.04 m-
Cu/ZnO/ZrO2/Al2O3 [107]CP11623.2 *250–30011.300.20--2950.590.0-0.05 m-
Cu(20)/ZrO2·Al2O3 (2:1) [18]IP143-3001---26,7002000.22241 s0s 59 s
Cu(20)/ZrO2·Al2O3 (1:1) [18]IP138-3001---26,7002000.24670 s0 s29 s
Cu(20)/ZrO2·Al2O3 (1:2) [18] IP167-3001---26,7002000.25868 s0 s31 s
Ni(20)/ZrO2·Al2O3 (1:2) [18]IP116-3001---26,7003000.29470 s25 s5 s
Ni(20)/ZrO2·Al2O3 (1:2) [18]IP116-5001---26,7003000.26165 s0 s22 s
Pd(1)–Cu(20)/
ZrO2·Al2O3 (1:2) [18]
SIP171-3001---26,7002000.26066 s0 s33 s
Rh(0.5)–Cu(20)/
ZrO2·Al2O3 (1:2) [18]
SIP--3001---26,7002000.28668 s14 s18s
Rh(1)–Cu(20)/
ZrO2·Al2O3 (1:2) [18]
SIP164-3001---26,7002000.27471 s4 s25 s
Rh(2)–Cu(20)/
ZrO2·Al2O3 (1:2) [18]
SIP--3001---26,7002000.26160 s18 s22 s
Pd(1)–Ni(20)/
ZrO2·Al2O3 (1:2) [18]
SIP120-3001---26,7002000.25863 s19s18 s
Rh(1)–Ni(20)/
ZrO2·Al2O3 (1:2) [18]
SIP123-3001---26,7002000.26664 s18s18 s
Ni(20)/ZnO·Al2O3 (2:1) [22]IP108-3001---26,7003000.27876 s0 s24 s
Ni(20)/ZnO·Al2O3 (1:1) [22]IP123-3001---26,7003000.29976 s0 s24 s
Ni(20)/ZnO·Al2O3 (1:2) [22]IP231-3001---26,7003000.27340 s10 s24 s
Ni(20)/ZnO·Al2O3 (1:4) [22]IP246-3001---26,7003000.28365 s0 s21 s
Pd(0.5)–Ni(20)/
ZnO·Al2O3 (1:1) [22]
SIP106-3001---26,7003000.29973 s10 s17 s
Pd(2)–Ni(20)/
ZnO·Al2O3 (1:1) [22]
SIP104-3001---26,7003000.29972 s8 s20 s
Pd(6.5)/ZnO [109]CP--400/50021.500.10--2500.3----
Pd(6.5)/ZnO–ZrO2 [109]CP--400/50021.500.10--2500.3----
Pd(6.5)/ZnO–Fe3O4 [109]CP--400/50021.500.10--2500.3----
Pd(6.5)/ZnO–MgO [109]CP--400/50021.500.10--2500.3----
Pd/(6.5)ZnO–Cr2O3 [109]CP--400/50021.500.10--2500.3----
Pd/(6.5)ZnO–Al2O3 [109]CP--400/50021.500.10--2500.3----
Pd(10)/ZnO [92]CP--0–5001----3000.110067 m6 m27 m
Pt(10)/ZnO [92]CP--0–5001----3000.110070 m2 m28 m
Co(10)/ZnO [92]CP--0–5001----3000.15343 m5 m30 m
Ni(10)/ZnO [92]CP--0–5001----3000.19654 m23 m13 m
Ir(10)/ZnO [92]IP--0–5001----3000.15949 m3 m31 m
Ru(10)/ZnO [92]CP--0–5001----3000.18848 m25 m11 m
Pd(10)/ZnO [92]CP--0–5001----2200.18960 m3 m27 m
Pd(10)/SiO2 [92]CP--0–5001----2200.1222 m10 m21 m
Pd(1)/CeO2 [92]CP--0–5001----2200.13911 m14 m21 m
Pd(1)/ZnO [92]CP--0–5001----2200.15535 m14 m24 m
Pd(5)/ZnO [92]CP--0–5001----2200.18060 m10 m25 m
Cu(25)/ZnO [92]CP--0–5001----2200.19920 c--
Cu(25)/ZrO2 [92]IP--0–5001----2200.17513 c--
Cu(25)/SiO2 [92]IP--0–5001----2200.150.1 c--
Cu/ZrO2 [121]DP33-25–3001---30,0003100.14068 m2 m98 m
Cu/ZrO2 [121]DP33-25–3001---30,0003500.15070 m10 m90 m
Ni/ZrO2 [121]DP34-25–3001---30,0003100.13060 m19 m80 m
Ni/ZrO2 [121]DP34-25–3001---30,0003500.110062 m80 m15 m
Cu–Ni/ZrO2 [121]DP35-25–3001---30,0003100.19072 m87 m13 m
Cu–Ni/ZrO2 [121]DP35-25–3001---30,0003500.19963 m80 m20 m
Cu(10)–Ni(30)/
ZrO2·Al2O3 [4]
CIP120-3001---26,7001600.2223.0 y0 s100 s
Cu(10)–Ni(30)/
ZrO2·Al2O3 [4]
CIP120-3001---26,7002000.2852.2 y48 s52 s
Cu(20)–Ni(20)/
ZrO2·Al2O3 [4]
CIP142-3001---26,7001600.2353.0 y0 s100 s
Cu(20)–Ni(20)/
ZrO2·Al2O3 [4]
CIP142-3001---26,7002000.2862.0 y48 s52 s
Cu(30)–Ni(10)/
ZrO2·Al2O3 [4]
CIP119-3001---26,7001600.2793.0 y0 s100 s
Cu(30)–Ni(10)/
ZrO2·Al2O3 [4]
CIP119-3001---26,7002000.2912.3 y39 s61 s
Cu(30)–Ni(10)/
CeO2·Al2O3 [4]
CIP120-3001---26,7001600.2263.0 y0 s100 s
Cu(30)–Ni(10)/
CeO2·Al2O3 [4]
CIP120-3001---26,7002000.2962.1 y30 s70 s
Cu(30)–Ni(10)/
ZnO·Al2O3 [4]
CIP150-3001---26,7001600.2193.0 y0 s100 s
Cu(30)–Ni(10)/
ZnO·Al2O3 [4]
CIP150-3001---26,7002000.2872.3 y23 s77 s
Cu(20)/MWCNTs [112]IP2900.35 *3001---26,7002000.11133 s0 s62 s
Cu(20)/MWCNTs [112]IP2900.35 *3001---26,7003000.17563 s8.5 s28.5 s
Ni(20)/MWCNTs [17]IP271-3001---26,7002000.17.578.5 s0 s21.5 s
Ni(20)/MWCNTs[17]IP271-3001---26,7003000.199.767.6 s16.5 s15.9 s
Au(1)–Cu(20)/MWCNTs [111]DP272-3001---26,7002000.11429.8 s0 s70.2 s
Au(1)–Cu(20)/MWCNTs [111]DP272-3001---26,7003000.18373 s8.6 s18.4 s
Au(1)–Ni(20)/MWCNTs [17]DP311-3001---26,7002000.1863.2 s0 s36.8 s
Au(1)–Ni(20)/MWCNTs [17]DP311-3001---26,7003000.199.870.4 s0 s29.6 s
*—based on chemisorption measurement; IP—impregnation method, SIP—subsequent impregnation method, CIP—co-impregnation method, CP—co-precipitation method, DP—deposition-precipitation method, UNC—urea nitrate combustion method, SBET—specific surface area, y—yield of the product (%), s—selectivity of the product (%), c—concentration of the product (%), m—level in product gas (mol%).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mosińska, M.; Szynkowska-Jóźwik, M.I.; Mierczyński, P. Catalysts for Hydrogen Generation via Oxy–Steam Reforming of Methanol Process. Materials 2020, 13, 5601. https://0-doi-org.brum.beds.ac.uk/10.3390/ma13245601

AMA Style

Mosińska M, Szynkowska-Jóźwik MI, Mierczyński P. Catalysts for Hydrogen Generation via Oxy–Steam Reforming of Methanol Process. Materials. 2020; 13(24):5601. https://0-doi-org.brum.beds.ac.uk/10.3390/ma13245601

Chicago/Turabian Style

Mosińska, Magdalena, Małgorzata I. Szynkowska-Jóźwik, and Paweł Mierczyński. 2020. "Catalysts for Hydrogen Generation via Oxy–Steam Reforming of Methanol Process" Materials 13, no. 24: 5601. https://0-doi-org.brum.beds.ac.uk/10.3390/ma13245601

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop