Next Article in Journal
Catalytic Acetalization: An Efficient Strategy for High-Value Utilization of Biodiesel-Derived Glycerol
Next Article in Special Issue
DBD Plasma Assisted CO2 Decomposition: Influence of Diluent Gases
Previous Article in Journal
Insights into the Metal Salt Catalyzed 5-Ethoxymethylfurfural Synthesis from Carbohydrates
Previous Article in Special Issue
Effect of Citric Acid on MoO3/Al2O3 Catalysts for Sulfur-Resistant Methanation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Review on Copper and Palladium Based Catalysts for Methanol Steam Reforming to Produce Hydrogen

1
School of Mechanical Engineering and Automation, Harbin Institute of Technology, Shenzhen Graduate School, Shenzhen 518055, China
2
Department of Mechanical Engineering, University of Texas Rio Grande Valley, Edinburg, TX 78539, USA
*
Author to whom correspondence should be addressed.
Catalysts 2017, 7(6), 183; https://doi.org/10.3390/catal7060183
Submission received: 4 May 2017 / Revised: 26 May 2017 / Accepted: 26 May 2017 / Published: 8 June 2017

Abstract

:
Methanol steam reforming is a promising technology for producing hydrogen for onboard fuel cell applications. The methanol conversion rate and the contents of hydrogen, carbon monoxide and carbon dioxide in the reformate, significantly depend on the reforming catalyst. Copper-based catalysts and palladium-based catalysts can effectively convert methanol into hydrogen and carbon dioxide. Copper and palladium-based catalysts with different formulations and compositions have been thoroughly investigated in the literature. This work summarized the development of the two groups of catalysts for methanol steam reforming. Interactions between the activity components and the supports as well as the effects of different promoters were discussed. Compositional and morphological characteristics, along with the methanol steam reforming performances of different Cu/ZnO and Pd/ZnO catalysts promoted by Al2O3, CeO2, ZrO2 or other metal oxides, were reviewed and compared. Moreover, the reaction mechanism of methanol steam reforming over the copper based and palladium based catalysts were discussed.

1. Introduction

Hydrogen is a promising alternative fuel because of its high energy density and low greenhouse gas emission rate [1]. As an energy carrier, hydrogen can be used to produce electrical power in a fuel cell. The use of fuel cells is attractive for the vehicle industry due to their high energy conversion efficiency and low emissions [2]. However, refueling of hydrogen is a major obstacle for the deployment of fuel cell vehicles owing to the lack of infrastructure [3]. Onboard hydrogen production by a processing system which converts liquid hydrocarbon fuels into hydrogen is a practical solution [4,5]. Among various fuels, methanol receives attention for onboard processing applications because: (a) It is easy to handle and store; (b) It can be extracted from renewable sources and it is biodegradable; (c) Its reforming temperature is relatively low compared to the other fuels; and (d) Its tendency to form coke during reforming is low due to the high H/C ratio [6,7].
Three main onboard fuel processing mechanisms to produce hydrogen are steam reforming, partial oxidation and auto thermal reforming [8]. Steam reforming is the most industrial kind of mature technology and it is also the most commonly employed process to produce hydrogen from methanol reforming [9]. Among all the three mechanisms, steam reforming has the highest ratio of hydrogen and carbon monoxide in the reformate.
The performance of a methanol steam reforming process significantly depends on the catalyst. Ni-based catalysts are mostly used in large scale industrial processes. However, only around 20% of the catalysts can be effectively utilized in industrial processes [10]. Thereafter, excess amounts of catalysts need to be used to counterbalance the effect of the low catalyst-utilization rate. For onboard methanol steam reforming, both the mass and size of the catalyst are restricted. Thus catalysts with higher activity and a better utilization rate are investigated. Interactions between the metal components and the support materials are critical for catalyst performance [11]. Promoters such as alkaline metals or acidic oxides are also important because they could promote the reactions following expected paths and help remove undesired matters that are formed during the reactions such as the carbonate species. Deactivation mechanisms including sintering and coke formation are also important concerns [12]. Overall, the objective for the development of an effective catalyst is to achieve the highest activity and hydrogen selectivity while minimizing the activity for side reactions such as methanol decomposition and maintaining the longest possible durability [9]. For different catalyst formulations, the influence of interactions between the catalyst components and the support as well as the addition of promoters on the structural, morphological and redox properties of the active phase needs to be examined.
The present work summarized the development of copper-based catalysts and palladium-based catalysts for methanol steam reforming. Interactions between the activity components and the supports as well as the effects of different promoters were discussed in detail. Methanol steam reforming performances based on different catalysts and the reaction mechanism were also presented. Besides copper-based and palladium-based catalysts, other metals such as nickel and cobalt are also used in catalysts for methanol steam reforming. Ni-based catalysts supported on alumina, zirconia, ceria and those containing Zn-Al or Mg-Al were intensively investigated for ethanol steam reforming instead of methanol reforming. The control of metallic Ni particle size and shape, as well as the redox properties of the catalyst system, are key factors in the preparation of Ni-based catalysts. Co and NiCo catalysts supported on cerium, lanthanum, zinc oxides, alumina and lantana-alumina were also studied for ethanol steam reforming. The key factors of Co-based catalysts are the control of metal morphology properties and the redox state of the system. A recent review summarized the progress of Ni-based and Co-based catalysts for methanol steam reforming [9]. Therefore they are not discussed in the present study.

2. Catalysts

2.1. Copper-Based Catalysts

Cu-based catalysts are commonly employed in methanol steam reforming reactions for onboard applications. Cu/ZnO catalysts are originally designed for methanol synthesis, and are also often used for methanol steam reforming because of the high activity at low operating temperature in the range of 250 °C and 300 °C. However, they need pre-treatment in hydrogen to obtain full activity, which is not convenient in small scale onboard systems [13]. Other disadvantages of the Cu/ZnO catalysts include the possibility of pyrophoric behavior occurring when exposed to air and the high sensitivity to a few ppm of sulfur [14].
Although it is confirmed that zinc leads to an improvement in the desired properties of the Cu/ZnO catalysts, the mechanism of the promotion effect of zinc is not completely understood [15]. Empirical experience indicates a particularly large Cu-ZnO contact is desired in order to improve the catalyst performance, whereas the casualty of the contact on the catalyst is still ambiguous [16]. Several different explanations of the mechanism can be found in the literature such as the spillover model, the metallic copper model and the Cu-Zn alloy model. The two-way transfer of hydrogen between Cu and ZnO is supposed to be crucial in the ZnO-promoted Cu system as explained in the spillover model [17]. Because ZnO has a large capacity for adsorbed hydrogen [18], it may trap H atoms originally produced on the Cu surface and serve as a reservoir to facilitate hydrogen spillover in the catalyst system. However, the hydrogen spillover from Cu to ZnO was found to be rapid from a partially oxidized Cu surface but limited from a fully reduced Cu surface [17]. In the metallic copper model, Cu particles supported on ZnO are confirmed in the metallic state by extended X-ray absorption fine-structure spectroscopy (EXAFS) [19]. However, the morphology of the Cu particle may reversibly change under different operating gaseous conditions, while no detectable change of the ZnO phase can be observed. An increase of the oxidation potential in the operating gas can lead to more spherical Cu particles with larger particle size. In contrast, flat dish-like Cu particles can be obtained when the operating gaseous condition is reduced as shown in Figure 1a [20,21]. The metallic copper model also argues that Cu-Zn alloy may play a role in the Cu/ZnO system, but Cu-Zn alloy is only formed at extreme reducing operating conditions. However, the Cu-Zn alloy model claims that the role of ZnO in the Cu/ZnO system is not to support Cu but to separate Cu particles from sintering as a spacer. Thus the Cu/ZnO catalyst is a mixed system of Cu and ZnO particles as shown in Figure 1b. The Cu-Zn or Cu-O-Zn surface alloy is supposed to be responsible for the promotion effect of ZnO depending on the reaction conditions and the amount of Zn [22]. Activity of the Cu/ZnO catalyst was proposed to be related to the copper metal surface area. Alejo et al. [23] found that the catalyst Cu40Zn60 (containing Cu 40 wt.% and Zn 60 wt.%) showed the highest copper area in the series of Cu-Zn contents with copper relative content of 20–70 wt.%. Correspondingly, Cu40Zn60 has the highest activity for methanol reforming to produce hydrogen. Therefore it can be confirmed that the Cu-ZnO interaction also depends on the catalyst compositions. Both the turnover frequencies (TOF) and apparent activation energy change with the catalyst compositions.
Lunkenbein et al. [24] were the first to report clear evidence for the formation of metastable graphite-like (GL) ZnO layers in industrial Cu/ZnO/Al2O3 catalysts induced by strong metal-support interactions. High-angle annular dark field scanning transmission electron microscopy (HAADF-STEM) was employed to analyze the microstructure of the fresh catalyst in the reduced state. The results show that the Cu nanoparticles (in pink color) are homogeneously dispersed within a porous ZnO (in yellow color) rod-like matrix as shown in Figure 2. High resolution TEM results also indicate that the ZnO layer is 1 to 2 nm thick comprising four to six sheets with interlayer distance between 1.8 and 2.8 Å. The proposed scheme of Cu and ZnO interaction is shown in Figure 3. Although the TEM imaging shows almost complete surface coverage of Cu particles, the accessibility of the Cu surface is still unclear due to the fact that only 2D projections are indicated in TEM imaging. Fichtl et al. [25] concluded that about 30% of the Cu surface area is accessible based on H2-temperature programed reduction (H2-TPD) results. Lunkenbein et al. [24] argued that these results discredit the direct correlation between catalyst activity and surface area.
Stabilizers and promoters are often employed in Cu/ZnO catalysts. Alumina can be added as support of Cu-ZnO to increase the catalyst durability and selectivity of H2 in the reformate [9]. Alejo et al. [23] found that the presence of alumina in Cu40Zn55Al5 (containing Cu 40 wt.%, Zn 55 wt.% and Al 5 wt.%) had a slightly inhibiting effect on methanol conversion, whereas it significantly improved the catalyst stability and selectivity for H2 and CO2 with only traces of CO existing in the reformate. Without the addition of alumina, the Cu40Zn60 catalyst started to deactivate after 20 h at 503 K. In comparison, the durability of Cu40Zn55Al5 was increased to more than 110 h at the same operating temperature [23]. A recent work reported that a Cu-Al spinel oxide catalyst can improve the activity for methanol steam reforming compared to the commercial Cu/ZnO/Al2O3 catalyst [26]. The Cu-Al spinel oxide catalysts can be used for methanol reforming without pre-reduction by H2, which not only avoids the copper sintering prior to the reaction, but also slows down the sintering rate during reaction. The small amount of CuO phase in the catalyst can initiate the reforming reaction at a low temperature, and Cu can be released gradually from CuAl2O4 to sustain the reaction after initiation. Moreover, the activity of the Cu-Al spinel catalyst can be further improved by adding excess amount of Al2O3, which can increase the surface area and CuAl2O4 crystal dispersion [26].
Other alkaline oxides such as CeO2 and ZrO2 can also be added into the Cu-based catalysts as promoters. The addition of ceria oxide is beneficial for maintaining copper dispersion and coke suppression due to its oxygen storage property [27]. Cu/CeO2 catalysts have been proven to have superior activities for steam methanol reforming compared to industrial Cu/ZnO catalysts [28]. The addition of cerium to Cu/ZnO/Al2O3 catalysts can effectively improve the activity and hydrogen selectivity. In any case, the content of CO in the reformate can be kept low due to the suppression of methanol decomposition and reverse water-gas shift reactions [29]. Men et al. [30] reported that the activities of Cu/CeO2/Al2O3 catalysts were dependent on the loading of Cu. They found that while the molar ratio of Cu and Ce in the catalyst increases from 0.1 to 0.9, the catalytic activity monotonously decreases. They attributed this to the fact that copper disperses preferentially on the ceria surface at low copper loading and the aggregation of copper particles occurs at high copper loading. Thermal desorption spectroscopy (TDS) results indicate that the decrease of copper loading and increase of ceria loading can enhance the basicity of the catalyst [31]. The methanol steam reforming reaction was believed to occur at the Cu/CeO2 interface and the proposed reaction mechanism is shown in Figure 4 [30,32]. It includes four steps to fulfill a catalysis cycle, which are the adsorption of methanol and water at the Cu/CeO2 interface, followed by the surface reaction and the oxygen reverse spillover (the migration of surface oxygen from CeO2 to Cu), and, finally, the regeneration of partially oxidized copper and oxygen vacancies. ZrO2 added to Cu/ZnO catalysts can also enhance the copper dispersion on the catalyst surface and thus lead to higher activity [33,34]. Moreover, the introduction of ZrO2 into Cu/ZnO improves the reducibility and stability of the catalysts [35].

2.2. Performances of Copper-Based Catalysts

Pan and Wang [36] experimentally tested a commercial Cu/ZnO/Al2O3 catalyst in a compact plate-fin reformer. The catalyst was grounded and sieved to the particle size of 0.442–0.681 mm before loading. The feeding reactants (CH3OH and H2O) flow rate was kept at a constant of 5 mL/min. Different molar ratios of H2O and CH3OH between 1.2 and 1.6 were examined. The results showed that the content of H2 in the dry reformate was 74.5% regardless of the ratio of H2O/CH3OH, whereas the content of CO decreased from 1.0% to less than 0.5% as the ratio of H2O/CH3OH increased from 1.2 to 1.6. Thus, complete methanol conversion can be achieved in all the tests.
Zhang et al. [37,38] investigated the methanol steam reforming performances of the commercial Cu/ZnO, Cu/ZnO/Al2O3 (CB-7), Cu/ZnO/CeO2, Cu/ZnO/ZrO2, and Cu/ZnO/CeO2/ZrO2 catalysts. The experiments were conducted in a fixed-bed quartz tube reactor with the gas hourly space velocity (GHSV) of 1200 h−1. The ratio of H2O and CH3OH in the reactants was kept at 1.2. At the reforming temperature of 250 °C, the methanol conversion is 97%, 97%, 80%, 75%, and 68% over Cu/ZnO/ZrO2, Cu/ZnO/CeO2/ZrO2, Cu/ZnO/CeO2, Cu/ZnO/Al2O3, and Cu/ZnO catalysts, respectively. The corresponding CO concentration in the dry reformate is 0.8%, 0.55%, 0.25%, 1.8%, and 0.3%, respectively. The results also show that no noticeable deactivation was observed during a 360 h continuous operation over the Cu/ZnO/CeO2/ZrO2 catalyst.
Baneshi et al. [39] compared the methanol steam reforming performances over Cu/CeO2/Al2O3, Cu/ZrO2/Al2O3, and Cu/CeO2/ZrO2/Al2O3 catalysts. The catalysts were evaluated in a fixed bed reactor with a constant GHSV of 10,000 cm3 h−1/g-catalyst. The ratio of H2O and CH3OH in the reactant flow was kept at 1.5. At the reforming temperature of 240 °C, the methanol conversion is 100%, 95% and 94% over the Cu/CeO2/Al2O3, Cu/CeO2/ZrO2/Al2O3, and Cu/ZrO2/Al2O3 catalysts, respectively. For Cu/CeO2/Al2O3, the H2 content in the dry reformate is around 65% and only a trace amount of CO is detectable. In a long run test, the activity of Cu/CeO2/Al2O3 remains at 100% after 110 h of continuous operation.
Zeng et al. [40] examined the Cu/ZnO/ZrO2/Al2O3 catalyst in cube-post micro reactors. The catalyst was prepared by co-precipitation method and coated in the plate by the wash coating method. The fed reactant mixture had a flow rate varying from 2 to 6 mL/h. The ratio of H2O and CH3OH was 1.3. The results indicate that the highest methanol conversion of 70.27% was achieved at 280 °C. The H2 concentration in the reformate were around 74.4%, and the CO concentration was about 1%. The low methanol conversion in this study resulted from the microreactor structure. With optimized microchannel structure and dimensions, the methanol conversion can be improved to around 95% [41].
Jones et al. [42] studied the performance of a Cu/ZnO/ZrO2/Al2O3 catalyst containing ZrO2 from a monoclinic nanoparticle ZrO2 precursor instead of a Zr(NO3)2 precursor. The catalyst was prepared by the co-impregnation method and its compositions were 15 wt.% CuO-15 wt.% ZnO-10 wt.% ZrO2-60 wt.% Al2O3. The total reactants fed rate was 0.8 mL/h and the ratio of H2O and CH3OH was 3. At 305 °C, the methanol conversation rate can achieve 80%, and the corresponding H2 content in the dry reformate is almost 80%. Only a trace amount of CO exists in the reformate.
Chang et al. [43] compared the oxidative steam reforming of methanol performances over the Cu/ZnO/ZrO2/Al2O3 catalyst and the commercial Cu/ZnO/Al2O3 catalyst. The Cu/ZnO/ZrO2/Al2O3 catalyst was prepared by the co-precipitation method. Under the reaction conditions of fed rate = 3.5 mL/h, H2O/CH3OH = 1.1, O2/CH3OH = 0.1, temperature = 250 °C, Cu/ZnO/ZrO2/Al2O3 (20/50/20/10 wt.%) shows higher activity than the commercial Cu/ZnO/Al2O3 (30/60/10 wt.%) catalyst. Methanol conversion rates are 75.3% and 65.5% over the two catalysts, respectively.
Ahmadi et al. [44] investigated the nanostructured CuO/ZnO/ZrO2/Al2O3 catalyst synthesized by the sonochemically-assisted co-precipitation method using different ultrasonic power. BET results show that higher ultrasonic power results in higher surface area and better activity. When the catalyst is synthesized and assisted by 90 W ultrasonic power, it can achieve 100% methanol conversion at a relatively low temperature of 200 °C. The H2 and CO concentrations in the dry reformate is above 75% and below 1%, respectively. No deactivation is observed after continuously running for 24 h.

2.3. Palladium Based Catalysts

Iwasa et al. [45] first experimentally tested Pd/ZnO, Pd/Al2O3, and Pd/ZrO2 catalysts for methanol steam reforming. The results show that the catalyst characteristics were greatly affected by the support. For catalysts with 1.0 wt.% Pd loading, Pd/ZrO2 has higher dispersion (28.9%) than Pd/ZnO (10.7%) and Pd/Al2O3 (13.1%). However, the TOF is the higher for Pd/ZnO (0.829 s−1) compared to Pd/ZrO2 (0.148 s−1) and Pd/Al2O3 (0.152 s−1). Pd/ZnO also has the highest selectivity for the steam reforming reaction of methanol. Only Pd/Al2O3 gave dimethyl ether in addition to hydrogen, carbon dioxide and carbon monoxide in the reformate.
Further investigation concludes that the catalytic performance of Pd/ZnO for methanol steam reforming can be greatly improved by reducing the catalyst prior to the reaction. Iwasa et al. [46] proposed that the Pd-Zn alloy formed in hydrogen treatment has positive effects on the selectivity of methanol steam reforming. Formaldehyde species formed in the reaction were able to be effectively transformed into H2 and CO2 over the catalyst containing Pd-Zn alloys. In contrast, the formaldehyde species selectively decomposed to H2 and CO over metallic-palladium-dominated catalysts without reducing treatment.
Chin et al. [47] confirmed the formation of the Pd-Zn alloy in the reduced Pd/ZnO catalyst by XRD, TPR and TEM characterization techniques. Figure 5 shows the high resolution TEM graph of a spent 16.7 wt.% Pd/ZnO catalyst. The presence of the Pd-Zn alloy was verified by the distinct d-lines of the small crystallites. It can also be concluded that Pd and Zn formed a highly ordered crystallites structure because the d-lines are continuous over an extended area as shown in Figure 5.
Chin et al. [48] also studied the effects of the preparation and pretreatment procedures on the Pd/ZnO catalyst. Using highly acidic Pd nitrate aqueous precursors such as the Pd(NO3)2 solution can alter the textural properties of the ZnO support including porosities and crystalline structures. The dissolution of ZnO into Zn2+ influenced the extent of mixing between Zn2+ and Pd2+ in pretreatment. As mentioned earlier, the formation of the Pd-Zn alloy was verified in the reducing pretreatment. Karim et al. [49] reported that the high temperature pretreatment (>300 °C) not only completely transformed Pd to Pd-Zn alloy, but also lead to particle growth. It is necessary to eliminate the small particles (<2 nm) because they lower the CO2 selectivity of the Pd/ZnO catalyst.
XRD characterization of the spent catalyst indicates that Pd-Zn crystallites are strongly dependent on Pd loading, as shown by the catalytic performance [50]. Surface concentration of the Pd-Zn alloy increases as the Pd loading increases, but larger Pd-Zn crystallites resulting from higher Pd loading also inhibit the further increase of the surface concentration of the Pd-Zn alloy. Excess amounts of Pd loading could decrease the catalyst activity and increase the CO content in the reformate [51]. Liu et al. [52] suggested that 5 wt.% Pd loading is appropriate with regard to the catalytic activity. However, Pfeifer et al. [53] reported that the catalyst with 10 wt.% Pd loading showed the highest activity. Liu et al. [54] also studied the deactivation mechanisms of the Pd/ZnO catalyst. Two major routes that caused deactivation are carbon deposited on the catalyst surface and surface oxidation of the Pd-Zn alloy.
Pd/ZnO supported on alumina was also investigated for methanol steam reforming [55]. The activity of Pd/ZnO catalysts can be greatly improved by supporting them on alumina, and the Pd/ZnO/Al2O3 catalysts show better stability than the commercial Cu/ZnO/Al2O3 catalysts [56]. Zn-modified Pd catalysts supported on CeO2 also show high selectivity for methanol steam reforming. XRD results indicate the formation of the Pd-Zn alloy, as well as the complete disappearance of the metallic Pd after reduction [57]. Methanol steam reforming performances over Pd/ZnO and Pd/CeO2 catalysts were compared. The results show that Pd/ZnO has a lower methanol conversion rate and higher CO2 selectivity than the Pd/CeO2 catalyst. The average TOF at 230 °C for Pd/ZnO and Pd/CeO2 was 0.8 ± 0.3 s−1 and 0.4 ± 0.2 s−1, respectively. Pd/ZnO has a higher density of acidic sites while Pd/CeO2 has a higher density of basic sites [58].
Except for Zn, the catalytic functions of the Pd/Al2O3 and Pd/CeO2 catalysts can be greatly modified by the addition of some other metals such as In and Ga due to the formation of Pd-In and Pd-Ga alloys [59,60,61]. Similarly to the Pd/ZnO catalysts, the alloys are responsible for the high CO2 selectivity and the metallic Pd is responsible for CO selectivity. The ratios of Pd and the second metal as well as the loadings on the support have significant effects on the catalyst activity [62,63].
Pd/ZrO2 catalysts were found to have high CO selectivity of approximately 50% at the reforming temperature of 260 °C [64,65]. Azenha et al. [64] first investigated the effect of Cu as a promoter of Pd/ZrO2 catalysts. The results show that the addition of Cu into Pd supported on ZrO2 with monoclinic crystalline structure leads to a remarkable decrease of the CO selectivity from 50% to 5%. SEM images shown in Figure 6 indicate the Cu and Pd particles are in intimate contact and highly dispersed in the support. The superior methanol steam reforming performance of the Cu-Pd/ZrO2 catalyst is due to the strong interaction between Pd and Cu as well as the enhanced dispersion of the metal phase on the support.

2.4. Performances of Palladium Based Catalysts

Echave et al. [66] experimentally examined the methanol steam reforming performance of the 2.5 wt.% Pd/ZnO catalyst. The catalyst was prepared by incipient wetness impregnation of pre-synthesized ZnO with a palladium nitrate solution. The catalyst was evaluated at the reforming temperature of 350 °C with a reactant flow rate of 0.064 mL/min and a steam to carbon ratio of 1.5. No obvious deactivation was observed during a 4000 min continuous operation. Methanol conversion higher than 90% can be reached. H2 and CO contents in the dry reformate are 70% and less than 1%, respectively. Sanz et al. [67] washcoated 2.5 wt.% Pd/ZnO catalyst in microchannel reformers, and similar results were reported.
Xia et al. [68] investigated the methanol steam reforming performance over a series of Pd/ZnO/Al2O3 catalysts. Under the reforming conditions of 220 °C, GHSV = 14,400 h−1, and H2O/CH3OH = 1.78, the methanol conversion is 14.3% and 46.5% over Pd/ZnO (Pd loading 8.6 wt.%) and Pd/ZnO/Al2O3 (Pd loading 8.9 wt.%), respectively. Both catalysts achieve CO2 selectivity higher than 99%. While elevating the reforming temperature to 270 °C, 100% methanol conversion is reached over the Pd/ZnO/Al2O3 catalyst. Pd loading and the Pd/Zn ratio were also varied and the performances were compared with respect to the methanol conversion and CO2 selectivity. The results showed that the catalyst with a Pd loading of 8.9 wt.% and a Pd/Zn ratio of 0.38 had the highest methanol conversion and CO2 selectivity. At the reforming temperature of 250 °C, 80% methanol conversion can be obtained over this catalyst.
Barrios et al. [69] tested the performances of Pd/CeO2 and Pd/ZnO/CeO2 catalysts for methanol steam reforming. Pd/CeO2 can only decompose methanol to CO at the reforming temperature of 250 °C and the steam to carbon ratio of 1. The methanol steam reforming reaction only occurs at the presence of zinc oxide. At 300 °C, the methanol conversion rate is 65% and 95% over Pd/ZnO and Pd/ZnO/CeO2 catalysts, respectively. Although the Pd/ZnO/CeO2 catalyst significantly improves the methanol conversion compared to Pd/ZnO, its CO2 selectivity (80%) is lower than that of Pd/ZnO (90%).
Men et al. [62] studied the methanol steam reforming performance of Pd-In/Al2O3 catalysts with different Pd loadings and Pd/In ratios. The results show that the catalyst with 7.5 wt.% Pd and 10 wt.% In2O3 has the overall best performance considering both the methanol conversion rate and the CO content in the reformate. About 90% methanol conversion and 0.6% CO concentration are achieved at the reforming temperature of 375 °C, with total flow rate of 30 mL/min and a steam to carbon ratio of 1.5.
Kuc et al. [70] experimentally investigated methanol steam reforming over perovskite-type oxides catalysts of LaCo0.87Pd0.13O3±δ (denoted as LCPO) and LaCo0.85Pd0.075Zn0.075O3±δ (denoted as LCPZO). At the reforming temperature of 250 °C, with a total flow rate of 100 mL/min and a H2O to CH3OH ratio of 1.3, the methanol conversion rate is about 90% and 30% for LCPO and LCPZO, respectively. The corresponding CO2 selectivity is 10% and 60%, respectively.

2.5. Reaction Mechanism of Methanol Steam Reforming

Regarding the reaction mechanism of methanol steam reforming over copper-based catalysts, four schemes have been proposed [7] including a methanol decomposition-water gas shift reaction scheme [71], a 1-step methanol steam reforming scheme [72,73], a methanol steam reforming-methanol decomposition-reverse water gas shift reaction scheme [74,75] and a methyl formate scheme [76,77,78]. The methanol steam reforming-methanol decomposition-reverse water gas shift reaction scheme is also adopted for methanol steam reforming over palladium catalysts [79,80].
The reaction scheme proposed by Peppley et al. [74,75] claimed the three reactions of methanol decomposition, water gas shift reaction, and methanol steam reforming all occur in parallel. This scheme has been supported and experimentally validated by different researchers [81,82,83,84,85,86,87]. The three parallel reactions are shown in Equations (1)–(3).
CH 3 OH + H 2 O k 1 k 1 3 H 2 + CO 2
CO 2 + H 2 k 2 k 2 H 2 O + CO
CH 3 OH k 3 2 H 2 + CO
As indicated above, both the methanol steam reforming and reverse water gas shift reactions are reversible, and the methanol decomposition reaction is irreversible. k1, k2 and k3 are the forward rate constants, while k−1 and k−2 are the backward rate constants.
Mastalir et al. [82] evaluated the reaction rate coefficients of the above three reactions. The Arrhenius equations are shown in Equations (4)–(6).
r 1 = k 1 C CH 3 OH 0.6 C H 2 O 0.4 exp ( E a 1 R T ) k 1 C CO 2 C H 2 exp ( E a 1 R T )
r 2 = k 2 C CO 2 C H 2 exp ( E a 2 R T ) k 2 C CO C H 2 O exp ( E a 2 R T )
r 3 = k 3 C CH 3 OH 1.3 exp ( E a 3 R T )
Chein et al. [88] proposed a different model including the methanol steam reforming reaction, methanol decomposition reaction and water gas shift reaction as shown in Equations (7)–(9). All the reactions are assumed to be irreversible.
CH 3 OH + H 2 O 3 H 2 + CO 2
CH 3 OH 2 H 2 + CO
H 2 O + CO CO 2 + H 2
Semi-empirical kinetic models proposed by Suh et al. [89] were used for the methanol steam reforming and decomposition reactions, which are,
r R = ( 1 ε ) ρ s k R C CH 3 OH
r D = ( 1 ε ) ρ s k D
where C is the molar concentration, ε is the catalyst layer porosity, and ρs is the catalyst density. The rate constants kR and kD can be expressed as,
k R = 5.5 ( 1.15 × 10 6 + 9.41 × 10 5 ln φ ) e E R / R T
k D = 5.5 × 7.09 × 10 7 e E D / R T
where ϕ is the molar ratio of water to methanol, and ER and ED are the activation energies of the reforming and decomposition reactions, respectively.
Studies of Purnama et al. [90] and Chen et al. [91] were used for the water gas shift reaction. The reaction rate is expressed as,
r W G S = 11.2 k W G S ( p CO p H 2 O p CO 2 p H 2 / K eq )
k W G S = 1.78 × 10 22 ( 1 0.154 δ + 0.008 δ 2 ) T 8 e E W G S / R T
K eq = exp ( 4577.8 T 4.33 )
where kWGS and Keq are the rate constant and the equilibrium constant of the water gas shift reaction, respectively. p is the partial pressure of a species, and δ is the molar ratio of water to carbon monoxide.

3. Conclusions

Methanol steam reforming is a promising technology to produce hydrogen-rich syngas with only a trace amount of carbon monoxide for onboard fuel cell applications. The performance of methanol steam reforming is significantly dependent on the reforming catalyst. The present work reviewed the commonly used copper-based and palladium-based catalysts for methanol steam reforming. Compositional and morphological characteristics, along with the methanol steam reforming performances of the two groups of catalysts were summarized.
Concerning the copper-based catalysts, it is confirmed that the Cu/ZnO catalysts have high activity towards methanol steam reforming reactions. However, the mechanism of the promotion effect of zinc oxide is still under discussion. One widely accepted theory is that Cu-Zn alloys formed in the Cu/ZnO catalysts are responsible for the promotion effect of ZnO. Alumina can be added to the Cu/ZnO catalysts as a stabilizer to increase the durability and selectivity of hydrogen, although it slightly reduces the methanol conversion rate. Other alkaline oxides such as CeO2 and ZrO2 can also be added as promoters. The addition of ceria oxide is beneficial for maintaining copper dispersion and coke suppression. The catalytic activity and hydrogen selectivity of the Cu/ZnO catalysts are also enhanced by the addition of ceria oxide. The introduction of ZrO2 into Cu/ZnO improves the reducibility and stability of the catalysts.
As for the palladium-based catalysts, the Pd/ZnO catalysts were found to have similar activities with the copper-based catalysts for methanol steam reforming. Moreover, they have higher carbon dioxide selectivity due to the formation of Pd-Zn alloys. The Pd-Zn crystallites and the catalytic performance of the catalysts are both strongly affected by the loading of Pd. An optimum range of Pd loading between 5% and 10% exists as reported in the literature. The activity of Pd/ZnO catalysts can be greatly improved by supporting them on alumina, and the Pd/ZnO/Al2O3 catalysts show better stability than the commercial Cu/ZnO/Al2O3 catalysts.
Several schemes have been proposed regarding the reaction mechanism of methanol steam reforming. The methanol steam reforming-methanol decomposition-reverse water gas shift reaction scheme is generally accepted for methanol steam reforming reactions occurring over the copper-based or palladium-based catalysts. In this scheme, the three reactions of the methanol decomposition, water gas shift reaction and methanol steam reforming are assumed to occur in parallel. Two different models based on this scheme were also presented in this work.

Acknowledgments

Support from Shenzhen Science and Technology Research and Development Funds (JCYJ20160318094917454) and Harbin Institute of Technology is gratefully acknowledged.

Author Contributions

All authors contributed to write the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, S.; Wang, X.; Xu, X.; Li, P. Hydrogen production via catalytic autothermal reforming of desulfurized Jet—A fuel. Int. J. Hydrog. Energy 2017, 42, 1932–1941. [Google Scholar] [CrossRef]
  2. Aicher, T.; Lenz, B.; Gschnell, F.; Groos, U.; Federici, F.; Caprile, L.; Parodi, L. Fuel processors for fuel cell APU applications. J. Power Sources 2006, 154, 503–508. [Google Scholar] [CrossRef]
  3. Xu, X.; Zhang, S.; Li, P. Autothermal reforming of n-dodecane and desulfurized Jet-A fuel for producing hydrogen-rich syngas. Int. J. Hydrog. Energy 2014, 39, 19593–19602. [Google Scholar] [CrossRef]
  4. Choudhury, A.; Chandra, H.; Arora, A. Application of solid oxide fuel cell technology for power generation—A review. Renew. Sustain. Energy Rev. 2013, 20, 430–442. [Google Scholar] [CrossRef]
  5. Xu, X.; Zhang, S.; Wang, X.; Li, P. Fuel adaptability study of a lab-scale 2.5 kWth autothermal reformer. Int. J. Hydrog. Energy 2015, 40, 6798–6808. [Google Scholar] [CrossRef]
  6. Iulianelli, A.; Ribeirinha, P.; Mendes, A.; Basile, A. Methanol steam reforming for hydrogen generation via conventional and membrane reactors: A review. Renew. Sustain. Energy Rev. 2014, 29, 355–368. [Google Scholar] [CrossRef]
  7. Yong, C.; Ooi, C.; Chai, S.; Wu, X. Review of methanol reforming-Cu-based catalysts, surface reaction mechanisms, and reaction schemes. Int. J. Hydrog. Energy 2013, 38, 9541–9552. [Google Scholar] [CrossRef]
  8. Xu, X.; Liu, X.; Xu, B. A survey of nickel-based catalysts and monolithic reformers of the onboard fuel reforming system for fuel cell APU applications. Int. J. Energy Res. 2016, 40, 1157–1177. [Google Scholar] [CrossRef]
  9. Kubacka, A.; Fernandez-Garcia, M.; Martinez-Arias, A. Catalytic hydrogen production through WGS or steam reforming of alcohols over Cu, Ni and Co catalysts. Appl. Catal. A 2016, 518, 2–17. [Google Scholar] [CrossRef]
  10. Christensen, T. Adiabatic prereforming of hydrocarbons—An important step in syngas production. Appl. Catal. A 1996, 138, 285–309. [Google Scholar] [CrossRef]
  11. Kubacka, A.; Martines-Arias, A.; Fernandez-Garcia, M. Role of the interface in base-metal ceria-based catalysts for hydrogen purification and production processes. ChemCatChem 2015, 7, 3614–3624. [Google Scholar] [CrossRef]
  12. Alenazey, F.; Cooper, C.; Dave, C.; Elnashaie, S.; Susu, A.; Adesina, A. Coke removal from deactivated Co–Ni steam reforming catalyst using different gasifying agents: An analysis of the gas–solid reaction kinetics. Catal. Commun. 2009, 10, 406–411. [Google Scholar] [CrossRef]
  13. Kolb, G. Review: Microstructured reactors for distributed and renewable production of fuels and electrical energy. Chem. Eng. Process. 2013, 65, 1–44. [Google Scholar] [CrossRef]
  14. Lindstrom, B.; Petterson, L. Deactivation of copper-based catalysts for fuel cell applications. Catal. Lett. 2001, 74, 27–30. [Google Scholar] [CrossRef]
  15. Rameshan, C.; Stadlmayr, W.; Penner, S.; Lorenz, H.; Memmel, N.; Havecker, M.; Blume, R.; Teschner, D.; Rocha, T.; Zemlyanov, D.; et al. Hydrogen production by methanol steam reforming on copper boosted by zinc-assisted water activation. Angew. Chem. Int. Ed. 2012, 51, 3002–3006. [Google Scholar] [CrossRef] [PubMed]
  16. Kasatkin, I.; Kurr, P.; Kniep, B.; Trunschke, A.; Schlogl, R. Role of lattice strain and defects in copper particles on the activity of Cu/ZnO/Al2O3 catalysts for methanol synthesis. Angew. Chem. Int. Ed. 2007, 119, 7465–7468. [Google Scholar] [CrossRef]
  17. Burch, R.; Golunski, E.; Spencer, M. The role of copper and zinc oxide in methanol synthesis catalysts. J. Chem. Soc. Faraday Trans. 1990, 86, 2683–2691. [Google Scholar] [CrossRef]
  18. Bowker, M.; Hadden, R.; Houghton, H.; Hyland, J.; Waugh, K. The mechanism of methanol synthesis on copper/zinc oxide/alumina catalysts. J. Catal. 1988, 109, 263–273. [Google Scholar] [CrossRef]
  19. Clausen, B.; Schiøtz, J.; Gråbæk, L.; Ovesen, C.; Jacobsen, K.; Nørskov, J.; Topsøe, H. Wetting/non-wetting phenomena during catalysis: Evidence from in situ on-line EXAFS studies of Cu-based catalysts. Top. Catal. 1994, 1, 367–376. [Google Scholar] [CrossRef]
  20. Grunwaldt, J.D.; Molenbroek, A.; Topsoe, N.Y.; Topsoe, H.; Clausen, B. In Situ investigations of structural changes in Cu/ZnO catalysts. J. Catal. 2000, 194, 452–460. [Google Scholar] [CrossRef]
  21. Yurieva, T.; Plyasova, L.; Makarova, O.; Krieger, T. Mechanisms for hydrogenation of acetone to isopropanol and of carbon oxides to methanol over copper-containing oxide catalysts. J. Mol. Catal. 1996, 113, 455–468. [Google Scholar] [CrossRef]
  22. Nakamura, J.; Choi, Y.; Fujitani, T. On the issue of the active site and the role of ZnO in Cu/ZnO methanol synthesis catalysts. Top. Catal. 2003, 22, 277–285. [Google Scholar] [CrossRef]
  23. Alejo, L.; Lago, R.; Pena, M.; Fierro, J. Partial oxidation of methanol to produce hydrogen over Cu-Zn-based catalysts. Appl. Catal. A 1997, 162, 281–297. [Google Scholar] [CrossRef]
  24. Lunkenbein, T.; Schumann, J.; Behrens, M.; Schlogl, R.; Willinger, M. Formation of a ZnO overlayers in industrial Cu/ZnO/Al2O3 catalysts induced by strong metal-support interactions. Angew. Chem. Int. Ed. 2015, 54, 4544–4548. [Google Scholar] [CrossRef] [PubMed]
  25. Fichtl, M.; Schumann, J.; Kasatkin, I.; Jacobsen, N.; Behrens, M.; Schlogl, R.; Muhler, M.; Hinrichsen, O. Counting of oxygen defects versus metal surface sites in methanol synthesis catalysts by different probe molecules. Angew. Chem. Int. Ed. 2014, 53, 7043–7047. [Google Scholar] [CrossRef] [PubMed]
  26. Xi, H.; Hou, X.; Liu, Y.; Qing, S.; Gao, Z. Cu-Al spinel oxide as an efficient catalyst for methanol steam reforming. Angew. Chem. Int. Ed. 2014, 53, 11886–11889. [Google Scholar] [CrossRef] [PubMed]
  27. Gao, J.; Guo, J.; Liang, D.; Hou, Z.; Fei, J.; Zheng, X. Production of syngas via autothermal reforming of methane in a fluidized-bed reactor over the combined CeO2-ZrO2/SiO2 supported Ni catalysts. Int. J. Hydrog. Energy 2008, 33, 5493–5500. [Google Scholar] [CrossRef]
  28. Liu, Y.; Hayakawa, T.; Suzuki, K.; Hamakawa, S. Production of hydrogen by steam reforming of methanol over Cu/CeO2 catalysts derived from Ce1−xCuxO2−x precursors. Catal. Commun. 2001, 2, 195–200. [Google Scholar] [CrossRef]
  29. Patel, S.; Pant, K. Activity and stability enhancement of copper-alumina catalysts using cerium and zinc promoters for the selective production of hydrogen via steam reforming of methanol. J. Power Sources 2006, 159, 139–143. [Google Scholar] [CrossRef]
  30. Men, Y.; Gnaser, H.; Zapf, R.; Hessel, V.; Ziegler, C. Parallel screening of Cu/CeO2/γ-Al2O3 catalysts for steam reforming of methanol in a 10-channel micro-structured reactor. Catal. Commun. 2004, 5, 671–675. [Google Scholar] [CrossRef]
  31. Men, Y.; Gnaser, H.; Ziegler, C.; Zapf, R.; Hessel, V.; Kolb, G. Characterization of Cu/CeO2/γ-Al2O3 thin film catalysts by thermal desorption spectroscopy. Catal. Lett. 2005, 105, 35–40. [Google Scholar] [CrossRef]
  32. Men, Y.; Gnaser, H.; Zapf, R.; Hessel, V.; Ziegler, C.; Kolb, G. Steam reforming of methanol over Cu/CeO2/γ-Al2O3 catalysts in a microchannel reactor. Appl. Catal. A 2004, 277, 83–90. [Google Scholar] [CrossRef]
  33. Wu, G.; Mao, D.; Lu, G.; Cao, Y.; Fan, K. The role of the promoters in Cu based catalysts for methanol steam reforming. Catal. Lett. 2009, 130, 177–184. [Google Scholar] [CrossRef]
  34. Jeong, H.; Kim, K.; Kim, T.; Ko, C.; Park, H.; Song, I. 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]
  35. Huang, G.; Liaw, B.; Jhang, C.; Chen, Y. Steam reforming of methanol over CuO/ZnO/CeO2/ZrO2/Al2O3 catalysts. Appl. Catal. A 2009, 358, 7–12. [Google Scholar] [CrossRef]
  36. Pan, L.; Wang, S. Modeling of a compact plate-fin reformer for methanol steam reforming in fuel cell systems. Chem. Eng. J. 2005, 108, 51–58. [Google Scholar] [CrossRef]
  37. Zhang, L.; Pan, L.; Ni, C.; Sun, T.; Zhao, S.; Wang, S.; Wang, A.; Hu, Y. CeO2-ZrO2-promoted CuO/ZnO catalyst for methanol steam reforming. Int. J. Hydrog. Energy 2013, 38, 4397–4406. [Google Scholar] [CrossRef]
  38. Zhang, L.; Pan, L.; Ni, C.; Sun, T.; Wang, S.; Hu, Y.; Wang, A.; Zhao, S. Effects of precipitation aging time on the performance of CuO/ZnO/CeO2-ZrO2 for methanol steam reforming. J. Fuel Chem. Technol. 2013, 41, 883–888. [Google Scholar] [CrossRef]
  39. Baneshi, J.; Haghighi, M.; Jodeiri, N.; Abdollahifar, M.; Ajamein, H. Homogeneous precipitation synthesis of CuO-ZrO2-CeO2-Al2O3 nanocatalyst used in hydrogen production via methanol steam reforming for fuel cell applications. Energy Convers. Manag. 2014, 87, 928–937. [Google Scholar] [CrossRef]
  40. Zeng, D.; Pan, M.; Wang, L.; Tang, Y. Fabrication and characteristics of cube-post microreactors for methanol steam reforming. Appl. Energy 2012, 91, 208–213. [Google Scholar] [CrossRef]
  41. Pan, M.; Wu, Q.; Jiang, L.; Zeng, D. Effect of microchannel structure on the reaction performance of methanol steam reforming. Appl. Energy 2015, 154, 416–427. [Google Scholar] [CrossRef]
  42. Jones, S.; Neal, L.; Everett, M.; Hoflund, G.; Hagelin-Weaver, H. Characterization of ZrO2-promoted Cu/ZnO/nano-Al2O3 methanol steam reforming catalysts. Appl. Surf. Sci. 2010, 256, 7345–7353. [Google Scholar] [CrossRef]
  43. Chang, C.; Chang, C.; Chiang, S.; Liaw, B.; Chen, Y. Oxidative steam reforming of methanol over CuO/ZnO/CeO2/ZrO2/Al2O3 catalysts. Int. J. Hydrog. Energy 2010, 35, 7675–7683. [Google Scholar] [CrossRef]
  44. Ahmadi, F.; Haghighi, M.; Ajamein, H. Sonochemically coprecipitation synthesis of CuO/ZnO/ZrO2/Al2O3 nanocatalyst for fuel cell grade hydrogen production via steam methanol reforming. J. Mol. Catal. A 2016, 421, 196–208. [Google Scholar] [CrossRef]
  45. 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]
  46. 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 1995, 125, 145–157. [Google Scholar] [CrossRef]
  47. Chin, Y.; Dagle, R.; Hu, J.; Dohnalkova, A.; Wang, Y. Steam reforming of methanol over highly active Pd/ZnO catalyst. Catal. Today 2002, 77, 79–88. [Google Scholar] [CrossRef]
  48. Chin, Y.; Wang, Y.; Dagle, R.; Li, X. Methanol steam reforming over Pd/ZnO: Catalyst preparation and pretreatment studies. Fuel Process Technol. 2003, 83, 193–201. [Google Scholar] [CrossRef]
  49. Karim, A.; Conant, T.; Datye, A. The role of PdZn alloy formation and particle size on the selectivity for steam reforming of methanol. J. Catal. 2006, 243, 420–427. [Google Scholar] [CrossRef]
  50. Zhang, H.; Sun, J.; Dagle, V.; Halevi, B.; Datye, A.; Wang, Y. Influence of ZnO facets on Pd/ZnO catalysts for methanol steam reforming. ACS Catal. 2014, 4, 2379–2386. [Google Scholar] [CrossRef]
  51. Liu, S.; Takahashi, K.; Uematsu, K.; Ayabe, M. Hydrogen production by oxidative methanol reforming on Pd/ZnO. Appl. Catal. A 2005, 283, 125–135. [Google Scholar] [CrossRef]
  52. 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]
  53. Pfeifer, P.; Schubert, K.; Liauw, M.; Emig, G. PdZn catalysts prepared by washcoating microstructured reactors. Appl. Catal. A 2004, 270, 165–175. [Google Scholar] [CrossRef]
  54. Liu, S.; Takahashi, K.; Fuchigami, K.; Uematsu, K. Hydrogen production by oxidative methanol reforming on Pd/ZnO: Catalyst deactivation. Appl. Catal. A 2006, 299, 58–65. [Google Scholar] [CrossRef]
  55. Ota, A.; Kunkes, E.; Kasatkin, I.; Groppo, E.; Ferri, D.; Poceiro, B.; Yerga, R.; Behrens, M. Comparative study of hydrotalcite-derived supported Pd2Ga and PdZn intermetallic nanoparticles as methanol synthesis and methanol steam reforming catalysts. J. Catal. 2012, 293, 27–38. [Google Scholar] [CrossRef]
  56. Conant, T.; Karim, A.; Lebarbier, V.; Wang, Y.; Girgsdies, F.; Schlogl, R.; Datye, A. Stability of bimetallic Pd-Zn catalysts for the steam reforming of methanol. J. Catal. 2008, 257, 64–70. [Google Scholar] [CrossRef]
  57. Iwasa, N.; Mayanagi, T.; Nomura, W.; Arai, M.; Takezawa, N. Effect of Zn addition to supported Pd catalysts in the steam reforming of methanol. Appl. Catal. A 2003, 248, 153–160. [Google Scholar] [CrossRef]
  58. Ranganathan, E.; Bej, S.; Thompson, L. Methanol steam reforming over Pd/ZnO and Pd/CeO2 catalysts. Appl. Catal. A 2005, 289, 153–162. [Google Scholar] [CrossRef]
  59. 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]
  60. Collins, S.; Delgado, J.; Mira, C.; Calvino, J.; Bernal, S.; Chiavassa, D.; Baltanas, M.; Bonivardi, A. The role of Pd-Ga bimetallic particles in the bifunctional mechanism of selective methanol synthesis via CO2 hydrogenation on a Pd/Ga2O3 catalyst. J. Catal. 2012, 292, 90–98. [Google Scholar] [CrossRef]
  61. Mayr, L.; Lorenz, H.; Armbruster, M.; Villaseca, S.; Luo, Y.; Cardoso, R.; Burkhardt, U.; Zemlyanov, D.; Haevecker, M.; Blume, R.; et al. The catalytic properties of thin film Pd-rich GaPd2 in methanol steam reforming. J. Catal. 2014, 309, 231–240. [Google Scholar] [CrossRef]
  62. Men, Y.; Kolb, G.; Zapf, R.; O’Connell, M.; Ziogas, A. Methanol steam reforming over bimetallic Pd-In/Al2O3 catalysts in a microstructured reactor. Appl. Catal. A 2010, 380, 15–20. [Google Scholar] [CrossRef]
  63. Föttinger, K. The effect of CO on intermetallic PdZn/ZnO and Pd2Ga/Ga2O3 methanol steam reforming catalysts: A comparative study. Catal. Today 2013, 208, 106–112. [Google Scholar] [CrossRef]
  64. Azenha, C.; Mateos-Pedrero, C.; Queiros, S.; Concepcion, P.; Mendes, A. Innovative ZrO2-supported CuPd catalysts for the selective production of hydrogen from methanol steam reforming. Appl. Catal. B 2017, 203, 400–407. [Google Scholar] [CrossRef]
  65. Kopfle, N.; Mayr, L.; Schmidmair, D.; Bernardi, J.; Knop-Gericke, A.; Havecker, M.; Klotzer, B.; Penner, S. A Comparative Discussion of the Catalytic Activity and CO2-Selectivity of Cu-Zr and Pd-Zr (Intermetallic) Compounds in Methanol Steam Reforming. Catalysts 2017, 7, 53. [Google Scholar] [CrossRef]
  66. Echave, F.; Sanz, O.; Montes, M. Washcoating of micro-channel reactors with PdZnO catalyst for methanol steam reforming. Appl. Catal. A 2014, 474, 159–167. [Google Scholar] [CrossRef]
  67. Sanz, O.; Velasco, I.; Reyero, I.; Legorburu, I.; Arzamendi, G.; Gandia, L.; Montes, M. Effect of the thermal conductivity of metallic monoliths on methanol steam reforming. Catal. Today 2016, 273, 131–139. [Google Scholar] [CrossRef]
  68. Xia, G.; Holladay, J.; Dagle, R.; Jones, E.; Wang, Y. Development of highly active Pd-ZnO/Al2O3 catalysts for microscale fuel processor applications. Chem. Eng. Technol. 2005, 28, 515–519. [Google Scholar] [CrossRef]
  69. Barrios, C.; Bosco, M.; Baltanas, M.; Bonivardi, A. Hydrogen production by methanol steam reforming: Catalytic performance of supported-Pd on zinc-cerium oxides’ nanocomposites. Appl. Catal. B 2015, 179, 262–275. [Google Scholar] [CrossRef]
  70. Kuc, J.; Weidenkaff, A.; Matam, S. Methanol steam reforming on perovskite-type oxides LaCo1−x−yPdxZnyO3±δ: Effect of Pd/Zn on CO2 Selectivity. Top. Catal. 2015, 58, 905–909. [Google Scholar] [CrossRef]
  71. Santacesaria, E.; Carra, S. Kinetics of a catalytic steam reforming of methanol in a CSTR reactor. Appl. Catal. 1983, 5, 345–358. [Google Scholar] [CrossRef]
  72. Breen, J.; Meunier, F.; Ross, J. Mechanistic aspects of the steam reforming of methanol over a CuO/ZnO/ZrO2/Al2O3 catalyst. Chem. Commun. 1999, 22, 2247–2248. [Google Scholar] [CrossRef]
  73. Agrell, J.; Bigersson, H.; Boutonnet, M. Steam reforming of methanol over Cu/ZnO/Al2O3 catalysts: A kinetic analysis and trategies for suppression of CO formation. J. Power Sources 2002, 106, 249–257. [Google Scholar] [CrossRef]
  74. Peppley, B.; Amphlett, J.; Kearns, L.; Mann, R. Methanol-steam reforming on Cu/ZnO/Al2O3. Part 1. The reaction network. Appl. Catal. A 1999, 179, 21–30. [Google Scholar] [CrossRef]
  75. Peppley, B.; Amphlett, J.; Kearns, L.; Mann, R. Methanol-steam reforming on Cu/ZnO/Al2O3. Part 2. A comprehensive kinetic model. Appl. Catal. A 1999, 179, 31–49. [Google Scholar] [CrossRef]
  76. Jiang, C.; Trimm, D.; Wainwright, M. Kinetic mechanism for the reaction between methanol and water over a Cu-ZnO-Al2O3 catalyst. Appl. Catal. A 1993, 97, 145–158. [Google Scholar] [CrossRef]
  77. Jiang, C.; Trimm, D.; Wainwright, M. Kinetic study of stream reforming of methanol over copper-based catalysts. Appl. Catal. A 1993, 93, 245–255. [Google Scholar] [CrossRef]
  78. Breen, J.; Ross, J. Methanol reforming for fuel-cell applications: Development of zirconia-containing Cu-Zn-Al catalysts. Catal. Today 1999, 51, 521–533. [Google Scholar] [CrossRef]
  79. Cao, C.; Xia, G.; Holladay, J.; Jones, E.; Wang, Y. Kinetic studies of methanol steam reforming over Pd/ZnO catalyst using a microchannel reactor. Appl. Catal. A 2004, 262, 19–29. [Google Scholar] [CrossRef]
  80. Uriz, I.; Arzamendi, G.; Dieguez, P.; Echave, F.; Sanz, O.; Montes, M.; Gandia, L. CFD analysis of the effects of the flow distribution and heat losses on the steam reforming of methanol in catalytic (Pd/ZnO) microreactors. Chem. Eng. J. 2014, 238, 37–44. [Google Scholar] [CrossRef]
  81. Frank, B.; Jentoft, F.; Soerijanto, H.; Krohnert, J.; Schlogl, R.; Schomacker, R. Steam reforming of methanol over copper-containing catalysts: Influence of support material on microkinetics. J. Catal. 2007, 246, 177–192. [Google Scholar] [CrossRef]
  82. Mastalir, A.; Frank, B.; Szizybalski, A.; Soerijanto, H.; Deshpande, A.; Niederberger, M.; Schomacker, R.; Schlogl, R.; Ressler, T. Steam reforming of methanol over Cu/ZrO2/CeO2 catalysts: A kinetic study. J. Catal. 2005, 230, 464–475. [Google Scholar] [CrossRef]
  83. Harold, M.; Nair, B.; Kolios, G. Hydrogen generation in a Pd membrane fuel processor: Assessment of methanol-based reaction systems. Chem. Eng. Sci. 2003, 58, 2551–2571. [Google Scholar] [CrossRef]
  84. 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]
  85. 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]
  86. Hsueh, C.; Chu, H.; Yan, W.; Leu, G.; Tsai, J. Three-dimensional analysis of a plate methanol steam micro-reformer and a methanol catalytic combustor with different flow channel designs. Int. J. Hydrog. Energy 2011, 36, 13575–13586. [Google Scholar] [CrossRef]
  87. Tadbir, M.; Akbari, M. Methanol steam reforming in a planar wash coated microreactor integrated with a micro-combustor. Int. J. Hydrog. Energy 2011, 36, 12822–12832. [Google Scholar] [CrossRef]
  88. Chein, R.; Chen, Y.; Chung, J. Thermal resistance effect on methanol-steam reforming performance in micro-scale reformers. Int. J. Hydrog. Energy 2012, 37, 250–262. [Google Scholar] [CrossRef]
  89. Suh, J.; Lee, M.; Greif, R.; Grigoropoulos, C. A study of steam methanol reformer in a micro-reactor. J. Power Sources 2007, 173, 458–466. [Google Scholar] [CrossRef]
  90. Purnama, H.; Ressler, T.; Jentoft, R.; Soerijanto, H.; Schlogl, R.; Schomacker, R. CO formation/selectivity for steam reforming of methanol with a commercial CuO/ZnO/Al2O3 catalysts. Appl. Catal. A 2004, 259, 83–94. [Google Scholar] [CrossRef]
  91. Chen, W.; Lin, M.; Jiang, T.; Chen, M. Modeling and simulation of hydrogen generation from high temperature and low temperature water gas shift reactions. Int. J. Hydrog. Energy 2008, 33, 6644–6656. [Google Scholar] [CrossRef]
Figure 1. (a) The metallic copper model; and (b) The Cu-Zn alloy model (Adapted from [22], Copyright 2003, Springer).
Figure 1. (a) The metallic copper model; and (b) The Cu-Zn alloy model (Adapted from [22], Copyright 2003, Springer).
Catalysts 07 00183 g001
Figure 2. High-angle annular dark field scanning transmission electron microscopy (HAADF-STEM) image of Cu/ZnO/Al2O3 fresh catalyst (Adapted from [24], Copyright 2015, Wiley-VCH).
Figure 2. High-angle annular dark field scanning transmission electron microscopy (HAADF-STEM) image of Cu/ZnO/Al2O3 fresh catalyst (Adapted from [24], Copyright 2015, Wiley-VCH).
Catalysts 07 00183 g002
Figure 3. Illustration of Cu/ZnO/Al2O3 fresh catalyst after reduction (Adapted from [24], Copyright 2015, Wiley-VCH).
Figure 3. Illustration of Cu/ZnO/Al2O3 fresh catalyst after reduction (Adapted from [24], Copyright 2015, Wiley-VCH).
Catalysts 07 00183 g003
Figure 4. Proposed reaction mechanism of methanol steam reforming on Cu/CeO2/Al2O3 catalyst (Adapted from [30], Copyright 2004, Elsevier).
Figure 4. Proposed reaction mechanism of methanol steam reforming on Cu/CeO2/Al2O3 catalyst (Adapted from [30], Copyright 2004, Elsevier).
Catalysts 07 00183 g004
Figure 5. High resolution TEM graph of a spent 16.7 wt.% Pd/ZnO catalyst (Adapted from [47], Copyright 2002, Elsevier).
Figure 5. High resolution TEM graph of a spent 16.7 wt.% Pd/ZnO catalyst (Adapted from [47], Copyright 2002, Elsevier).
Catalysts 07 00183 g005
Figure 6. SEM images of the CuPd/ZrO2 catalyst with EDX quantification (Adapted from [64], Copyright 2017, Elsevier).
Figure 6. SEM images of the CuPd/ZrO2 catalyst with EDX quantification (Adapted from [64], Copyright 2017, Elsevier).
Catalysts 07 00183 g006

Share and Cite

MDPI and ACS Style

Xu, X.; Shuai, K.; Xu, B. Review on Copper and Palladium Based Catalysts for Methanol Steam Reforming to Produce Hydrogen. Catalysts 2017, 7, 183. https://doi.org/10.3390/catal7060183

AMA Style

Xu X, Shuai K, Xu B. Review on Copper and Palladium Based Catalysts for Methanol Steam Reforming to Produce Hydrogen. Catalysts. 2017; 7(6):183. https://doi.org/10.3390/catal7060183

Chicago/Turabian Style

Xu, Xinhai, Kaipeng Shuai, and Ben Xu. 2017. "Review on Copper and Palladium Based Catalysts for Methanol Steam Reforming to Produce Hydrogen" Catalysts 7, no. 6: 183. https://doi.org/10.3390/catal7060183

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