Next Article in Journal
Improved Microwave-Assisted Ethyl Levulinate Production Using Rice Husk-Derived Biobased Mesoporous Silica as Catalyst
Previous Article in Journal
Catalytic Cascade for Biomass Valorisation: Coupled Hydrogen Transfer Initiated Dehydration and Self-Aldol Condensation for the Synthesis of 2-methyl-pent-2-enal from 1,3-propanediol
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Influence of the ZrO2 Crystal Phase on Cu/ZrO2-Al2O3 Catalysts in Methanol Steam Reforming

by
Mouxiao Song
,
Li Li
,
Xueshuang Wu
,
Haiqing Cai
,
Guiying Li
* and
Changwei Hu
Key Laboratory of Green Chemistry and Technology, Ministry of Education, College of Chemistry, Sichuan University, Chengdu 610064, China
*
Author to whom correspondence should be addressed.
Catalysts 2024, 14(8), 480; https://doi.org/10.3390/catal14080480
Submission received: 27 June 2024 / Revised: 21 July 2024 / Accepted: 24 July 2024 / Published: 27 July 2024
(This article belongs to the Section Catalysis for Sustainable Energy)

Abstract

:
Copper-based catalysts are widely used in methanol steam reforming to produce hydrogen. In this paper, the supportive effect of the crystal phase of ZrO2 on Cu-based catalysts in methanol steam reforming is discussed. Monoclinic(m-), Tetragonal(t-) and mixed ZrO2 phases were prepared, and structure–activity relationships were investigated with XRD, H2-TPR, BET, HR-TEM and XPS. It was found that the catalyst with a 81.4% monoclinic ZrO2 crystal phase exhibited the highest methanol conversion (88.5%) and the highest hydrogen production rate (104.6 μmol/gcat·s) at 275 °C as it displayed the best reducing properties and more oxygen vacancies on the catalyst surface. Oxygen vacancies can produce more Cu1+ + Cu0, which is the active species for methanol steam reforming on the catalyst surface, and therefore affect catalytic activity.

1. Introduction

With the rapid growth of population and industrialization, environmental pollution is becoming more and more serious. The overuse of traditional fossil fuels is one of the reasons for this [1,2,3]. As one kind of clean and green secondary energy, hydrogen can be a good alternative to traditional fossil fuels. It comes from a wide range of sources, and is more efficient than gasoline in energy conversion [4]. However, the efficient utilization of hydrogen energy is still a difficult but urgent problem. As we all know, hydrogen is an inflammable and explosive gas, the explosion limit of which is 4–75.6% (volume concentration). Therefore, the safe storage and transportation of hydrogen is a huge problem. What is more, the establishment of hydrogen stations is also a problem, since it is very expensive and complicated with current technology [5].
Recent studies have shown that methanol steam reforming (MSR) is a promising method for hydrogen production; the stoichiometric equations of MSR are shown in Equations (1) and (2) [6]. Methanol has many advantages. Firstly, methanol has a high H/C ratio, high volume energy density, biodegradability and low sulfur content [7,8,9,10]. Secondly, methanol is liquid at room temperature and pressure, which is more conducive to storage and transportation. Thirdly, methanol has a low boiling point (65 °C), which means its vaporization amplitude is roughly the same as that of water, and it is soluble in water and easily handled [7]. In addition, methanol lacks a C-C bond, resulting in a temperature requirement of 150–300 °C for its conversion to hydrogen, significantly lower than that of other alcohols. For example, the steam reforming temperature of ethanol is about 400 °C [11].
CH3OH + H2O → 3H2 + CO2 (ΔHθ = +50 kJ/mol)
CO2 + H2 → CO + H2O (ΔHθ = +41 kJ/mol)
In recent years, the development of fuel cells has been much more rapid, the technological advancement of which also urgently requires the efficient utilization of hydrogen energy [12]. Methanol steam reforming is widely used in polymer electrolyte membrane fuel cells, which is a power generation component used in small transport vehicles such as new energy cars [13,14,15,16]. In general, hydrogen production from methanol reforming is very promising since it can not only produce hydrogen efficiently, but also solve the hydrogen transportation problem.
High concentrations of carbon monoxide cause irreversible damage to fuel cells [17]. Therefore, they require both low carbon monoxide selectivity and high hydrogen selectivity for methanol steam reforming with high methanol conversion at low temperatures. Catalysts for MSR are mainly divided into noble metal catalysts [18,19,20], and copper-based catalysts [21,22,23]. Although noble metal catalysts have the advantages of high stability and good selectivity and activity, its cost is too high, which makes its commercial application at a large scale unrealistic. For this reason, copper-based catalysts have received more focus from researchers. Initially, Zn and Cu were used to synthesize catalysts for methanol steam reforming. It was found that the addition of Zn significantly improved the activity of the catalyst [24]. In order to overcome the shortcomings of copper-based catalysts such as easy sintering and easy deactivation, and improve the activity of the catalyst, stabilizers and promoters are often added. Alejo et al. [25] investigated the effect of the Al2O3 component on Cu/ZnO/Al2O3 catalysts for MSR. It was found that the catalyst without Al2O3 was completely deactivated after a reaction at 230 °C for 20 h, while for the catalyst with Al2O3, at the same temperature for 110 h, the catalyst exhibited observable activity, and the selectivity of hydrogen and carbon dioxide were improved compared to that without Al2O3. Meanwhile, Al2O3 is also used as catalyst support due to its excellent properties [26,27]. In addition to the Al component, ZrO2 is also used as promoter for methanol steam reforming reactions. Wu et al. [22] found that adding Zr components to Cu/ZnO catalysts could improve the stability and performance of copper-based catalysts by preventing Cu particles from aggregating, therefore stabilizing Cu crystals. Jeong et al. [28] also reported that the addition of ZrO2 to Cu/ZnO/Al2O3 catalysts promoted the dispersion of Cu on the catalyst surface and the formation of smaller copper particles on the catalyst surface.
Numerous studies have explored the impact of the ZrO2 crystal phase on catalytic activity for methanol synthesis from CO2. Scholars have demonstrated that copper-based catalysts, when supported by ZrO2 in various crystalline forms, including monoclinic (m-ZrO2), tetragonal (t-ZrO2) and amorphous (a-ZrO2), exhibits distinctly different catalytic performances in the hydrogenation of carbon dioxide [29,30,31]. Witoon et al. [29] found that Cu/a-ZrO2 catalysts had bigger copper surface area than Cu/m-ZrO2 and Cu/t-ZrO2. The surface area of copper is the main factor controlling the yield of methanol and CO. Therefore, Cu/a-ZrO2 catalyst shows higher catalytic activity. Samson et al. [31] discussed the catalytic activity of Cu-based catalysts for methanol synthesis from CO2 with different ZrO2 crystal phase contents. They discovered that the catalytic activity escalates as the proportion of t-ZrO2 to m-ZrO2 increases. Therefore, enhancing the activity and stability of catalysts by altering the ZrO2 crystal phase is scientifically grounded. While these insights offer hints for MSR, the literature on the specific influences of ZrO2 crystal phases in relation to MSR remains scarce.
MSR is the reverse reaction of carbon dioxide hydrogenation to methanol, and many catalysts for carbon dioxide hydrogenation to methanol might be also suitable for MSR. Therefore, it is interesting to see whether changes in the ZrO2 crystal phase affect the activity of MSR. In our previous work [32], a Cu/ZrO2-Al2O3 catalyst was prepared via mechanical mixing and the impregnation method for MSR. It was found that the catalyst with a 1:8 molar ratio of ZrO2 and Al2O3 showed the best activity, and the catalyst obtained by mechanical mixing was not just a simple mixture of Cu/ZrO2 and Cu/Al2O3. Several publications have reported that ZrO2 of varying crystal phases can be obtained through different preparation methods [29,30]. Therefore, based on our previous work, different ZrO2 crystal phases have been prepared in this study by altering the preparation techniques, obtaining various catalysts through mechanical mixing and impregnation methods, where all catalysts maintained the same Zr/Al molar ratio (1:8). Subsequently, the activity differences among these catalysts were investigated, and several characterizations were conducted to explore the relationship between catalyst structure and its activity.

2. Results

2.1. Catalyst Activity for MSR

The activity of different catalysts for MSR is shown in Figure 1. It can be seen that the methanol conversion and hydrogen production rate of all catalysts except Cu/Zr-Al-2 increased with the increase of temperature, and the Cu/Zr-Al-3 catalyst shows the best performance. For Cu/Zr-Al-2, it seems to be less active below 275 °C and the methanol conversion only increases slightly at 300 °C. As shown in Figure 1c, the CO selectivity of all catalysts increase with increasing temperature. It can be speculated that high temperature is conducive to the production of CO byproducts on all the catalysts used in the present work. Too high of a temperature will result in high CO selectivity, and too of a low temperature will result in low methanol conversion. The aim of the present work was to produce hydrogen from MSR at a low temperature; therefore, 275 °C was chosen as a suitable MSR reaction temperature for the Cu/Zr-Al catalyst. The activity data of different catalysts at 275 °C are shown in Figure 2. It is clear that the Cu/Zr-Al-3 catalyst shows the highest methanol conversion (88.5%) and hydrogen production rate (104.6 μmol/gcat·s). At the same time, the Cu/Zr-Al-3 catalyst shows low CO selectivity (1.9%). However, the Cu/Zr-Al-2 catalyst shows the lowest methanol conversion (10.2%), hydrogen production rate (9.5 μmol/gcat·s) and the highest CO selectivity (3.5%). Despite the small difference in CO selectivity, there is significant variation in the catalytic activity results. In addition, a stability experiment of the Cu/Zr-Al-3 catalyst was also performed, and the result is shown in Figure S1. It can be seen that this catalyst exhibits good stability. Therefore, Cu/Zr-Al-3 is relatively more suitable for MSR.

2.2. XRD Analysis

XRD patterns of the catalysts are shown in Figure 3. Figure 3a shows the XRD pattern of calcined catalysts. The characteristic peaks of CuO (PDF:89-2529), m-ZrO2 (PDF: 83-0938) and t-ZrO2 (PDF: 79-1764) can be observed [33,34,35]. The diffraction peaks of CuO on all catalysts are not obvious, which means that highly dispersed CuO species have been formed on the catalysts. Unusually, Cu/Zr-Al-2 contains only tetragonal crystal phase ZrO2 and Cu/Zr-Al-4 contains only monoclinic crystal phase ZrO2. However, other catalysts contain the mixed monoclinic and tetragonal crystal phase ZrO2. Figure 3b shows the XRD pattern of reduced catalysts. It can be seen that there is no clear change in the characteristic peaks of ZrO2, because it is hard to reduce bulk ZrO2 below 500 °C [36]. It can be also seen from the figure that the reduced catalyst has few characteristic peaks of CuO and the characteristic peaks of Cu (PDF: 85-1326) appears, which indicates that most CuO has been reduced to Cu under this reduction condition, while the reduction process might cause some aggregation of metal Cu species.
In order to further investigate the influence of the ZrO2 crystal phase on activity, the proportions of different ZrO2 crystal phases in all catalysts were semi-quantitatively calculated with the normalized RIR method [37], and the results are shown in Figure 4 and Table S1. It can be seen that the proportion of monoclinic crystal phase ZrO2 in mixed crystal phase catalysts is Cu/Zr-Al-3 > Cu/Zr-Al-1 > Cu/Zr-Al-5. Cu/Zr-Al-2 has pure tetragonal crystal phase ZrO2 and Cu/Zr-Al-4 has pure monoclinic crystal phase ZrO2. It can be also found that the proportion of monoclinic crystal phase ZrO2 of the reduced catalyst is slightly higher than the calcined catalyst. According to the results of catalytic activity, which is shown in Figure S2, the catalyst with a higher m-ZrO2 proportion is more active, and the catalytic activity of the catalysts containing pure monoclinic ZrO2 is much higher than that of the catalysts containing pure tetragonal ZrO2.

2.3. H2-TPR Analysis

Figure 5 shows the reducibility of the catalysts. There are four reduction peaks which are labeled a, b, c, d from low to high temperature. All the reduction peaks originate from the CuO species, because the reduction peaks of ZrO2 and Al2O3 are higher than 500 °C [36]. The reduction peaks of a and b originate from the stepwise reduction of highly dispersed Cu2+ to Cu+ and Cu0 [38]. It can be seen that the a and b reduction peaks of Cu/Zr-Al-3 move to a lower temperature region. The results show that the Cu/Zr-Al-3 has better reducing capacity, and the stepwise reduction of CuO species is promoted. The peak of c originates from the reduction of highly dispersed CuO with interactions between ZrO2 species [38,39]. Cu/Zr-Al-2 and Cu/Zr-Al-4 have the obvious reduction peak of c, but the c reduction peak of Cu/Zr-Al-1 and Cu/Zr-Al-5 is relatively weak and Cu/Zr-Al-3 has almost no c reduction peak. The reduction peak of d originates from the bulk copper oxide crystalline [40,41,42]. It can be seen that this peak exists only in the Cu/Zr-Al-2 catalyst. Because the reduction temperature of d reaches 300.8 °C, the reduction capacity of Cu/Zr-Al-2 is poor. This may be the reason why this catalyst has the worst performance in addition to the presence of pure tetragonal ZrO2.

2.4. BET Surface Area and ICP-AES Analysis

Figure 6 shows the nitrogen adsorption and desorption isotherms and pore diameter distributions of the catalysts. The hysteresis loops at P/P0 = 0.6–1.0 mean that the catalyst has a mesoporous character, and all isotherms are type IV [43]. The pore diameter distribution of the catalysts is mostly between 10–30 nm. Table 1 shows the actual Cu contents, the texture and chemical properties of the catalysts and supports. Cu/Zr-Al-2 with pure tetragonal crystal phase ZrO2 has the smallest specific surface area, pore volume, pore diameter and particle size of ZrO2. This may be related to the step-wise impregnation of the catalyst in preparation. Unusually, Cu/Zr-Al-5 has the biggest specific surface area, the biggest particle size of ZrO2 and the smallest particle size of Cu. For other catalysts, there is no obvious change in specific surface area, pore volume and pore diameter. In addition, Cu dispersion and surface area were calculated and the results are shown in Table 1. It can be seen that the Cu dispersion and surface area follow the order Cu/Zr-Al-3 > Cu/Zr-Al-1 > Cu/Zr-Al-4 > Cu/Zr-Al-5 > Cu/Zr-Al-2. Higher metal surface area illustrates better dispersion of the metal, which means more active metal sites, ultimately leading to higher catalytic activity. The results show that the catalyst with higher Cu dispersion and Cu surface area has higher activity in this work.

2.5. XPS and AES Analysis

Figure 7 shows the surface elemental chemical valence states of reduced catalysts by X-ray photoelectron spectroscopy characterization. All the XPS data were analyzed by XPS PEAK41 software. The FWHM data of all fitting peaks are less than 2.7 eV, and the type of baseline used was the Shirley type. As shown in Figure 7a, there are two peaks at about 932.5 eV and 951.2 eV originating from Cu 2p3/2 and Cu 2p1/2, respectively [44]. It is reported that the peaks at 953.8 eV and 934 eV are related to Cu2+, and the peaks at 952.1 eV and 932.3 eV are related to Cu0 + Cu1+ [45,46,47,48]. The reason for the existence of Cu2+ in the reduced catalyst is that the reduced catalyst was passivated for 12 h under an atmosphere of 5% O2/Ar, so there will be Cu2+ on the surface of the catalyst. Cu0 and Cu1+ are too difficult to be distinguished since their binding energies are located too close to one another (Cu+ at 932.18 eV and Cu0 at 932.63 eV) [49]. Cu Auger spectra were used to distinguish Cu0 and Cu1+ species and the result is shown in Figure S3. It can be seen that the characteristic peaks of Cu0 and Cu1+ are also not obvious. There is also one peak at 942.7 eV, which originates from the satellite peak of Cu2+, that confirms the existence of Cu2+ [50]. In Figure 7b, there are two peaks about 530.0 eV and 531.2 eV originating from the lattice oxygen species and the surface oxygen species, respectively [51,52]. In Figure 7c, there are two peaks at about 180.7 eV and 183.1 eV originating from Zr 3d3/2 and Zr 3d5/2, respectively [53,54]. It is reported that the peaks at 181.6 eV and 183.8 eV are related to Zr4+ and the peaks at 180.7 eV and 183 eV are related to Zr3+ [34].
Table 2 shows the relative content of elements on the reduced catalysts’ surface. The Cu/Zr-Al-3 catalyst has the highest proportion of (Cu1+ + Cu0) on its surface, which may be the active site of MSR [55]. A high proportion of (Cu1+ + Cu0) is beneficial to MSR processing and this is why the Cu/Zr-Al-3 catalyst has the best performance. Zr4+ is derived from the lattice ions of ZrO2 species, while Zr3+ is derived from the partially reduced oxygen-deficient ZrO2 [52,54]. Therefore, the ratio of Zr3+ can be used to reflect the concentration of oxygen vacancy on the ZrO2 support [52,54]. The Cu/Zr-Al-3 catalyst has the highest surface proportion of Zr3+ (89.1%) and the Cu/Zr-Al-2 catalyst has the lowest proportion of Zr3+ (79.8%). This means that the Cu/Zr-Al-3 catalyst has the highest concentration of oxygen vacancy and the Cu/Zr-Al-2 catalyst has the lowest concentration of oxygen vacancy. It is reported that oxygen vacancy concentration can be evaluated using the formula Osurface/(Olattice + Osurface) [35,52,54]. The highest proportion of Osurface is observed in Cu/Zr-Al-3 and the lowest proportion of Osurface is observed in Cu/Zr-Al-2. This is consistent with the results for the Zr 3d spectra. For other catalysts, it can be seen that the concentration of oxygen vacancy is positively correlated with the activity of catalysts. In addition, the O2-TPD characterization experiment was used to detect the concentration of oxygen vacancy and the results are shown in Figure 7d. The α region oxygen desorption peak (70–200 °C) is attributed to the desorption of weakly chemically oxygen [56], and the β region oxygen desorption peak (200–450 °C) is attributed to the desorption of strongly chemically oxygen which originates from the oxygen vacancy [57]. The γ region oxygen desorption peak (450–650 °C) is attributed to the desorption of O [58]. It can be seen that the strongest β region oxygen desorption peak appears on the Cu/Zr-Al-3 catalyst, which proves that this catalyst exhibits the most oxygen vacancies. This result is consistent with the results from O 1s and Zr 3d spectra. In general, the Cu/Zr-Al-3 catalyst has the highest concentration of oxygen vacancy, which may affect the valence state of Cu and further promote the occurrence of MSR.

2.6. TEM and HRTEM Analysis

Figure 8 shows TEM and HRTEM images of the reduced catalysts. The interplanar spacings observed in the image of Cu/Zr-Al-1 and Cu/Zr-Al-5 are 0.29, 0.31 and 0.21 nm, respectively, indicating the exposure of t-ZrO2 (101), m-ZrO2 (−111) and Cu (111) lattice planes. For Cu/Zr-Al-2, the lattice fringes in the spacings of 0.21, 0.26 and 0.29 nm are attributed to Cu (111), t-ZrO2 (002) and t-ZrO2 (101) lattice planes. For Cu/Zr-Al-4, the lattice fringes in the spacings of 0.21, 0.28 and 0.31 nm are attributed to Cu (111), m-ZrO2 (111) and m-ZrO2 (−111) lattice planes. For Cu/Zr-Al-3, the lattice fringes in the spacings of 0.26 and 0.28 nm are attributed to t-ZrO2 (002) and m-ZrO2 (111) lattice planes, which is not present on other catalysts with mixed crystal phase ZrO2. Significantly, the lattice planes of t-ZrO2 (101) and m-ZrO2 (−111) in Cu/Zr-Al-1 and Cu/Zr-Al-5 are also not present on the Cu/Zr-Al-3 catalyst. Based on the results of XPS and activity, it can be deduced that the co-existence of t-ZrO2 (101) and m-ZrO2 (−111) lattice planes leads to the appearance of more defective ZrO2 species on the Cu/Zr-Al-3 catalyst, resulting in a higher concentration of oxygen vacancy, which further affects the valence change of the active species Cu. Eventually, the Cu/Zr-Al-3 catalyst has the best catalytic activity.

3. Discussion

Cu/Zr-Al catalysts were synthesized by different methods to obtain different crystal phase ZrO2. It was found that the Cu/Zr-Al-3 catalyst had the best performance with the highest methanol conversion (88.5%), hydrogen production rate (104.6 μmol/gcat·s) and low CO selectivity (1.9%) at 275 °C. From the characterization of the catalysts, it can be seen that the mixed crystal phase catalyst with a higher proportion of m-ZrO2 exhibits higher catalytic activity, and the catalytic activity of the catalysts containing pure m-ZrO2 is much higher than that of the catalysts containing pure t-ZrO2. Meanwhile, the highly dispersed Cu2+ stepwise reduction peaks move to a low temperature region in the Cu/Zr-Al-3 catalyst. This means that this catalyst has better reducing capacity, and the stepwise reduction of Cu2+ species is promoted. The Cu/Zr-Al-3 catalyst has the characteristic lattice planes of t-ZrO2 (002) and m-ZrO2 (111), which is absent on other catalysts with mixed crystal phase ZrO2. It is possible that the co-existence of these two characteristic lattice planes influences the growth of copper species. Catalysts with higher m-ZrO2 ratios have the highest concentrations of oxygen vacancy, which further affects the valence state of Cu over the catalyst. Therefore, the Cu/Zr-Al-3 catalyst contains the highest proportion of Cu1++Cu0, which promotes the occurrence of MSR. In summary, a series of catalysts are obtained by changing the preparation method of ZrO2. These catalysts have different structural characteristics, such as the ratio of monocline and the tetragonal ZrO2 crystal phase, the exposed lattice planes of ZrO2, the content of oxygen vacancy, and the content of active species Cu1++Cu0. It is the structural changes of these catalysts that lead to the differences in catalytic activity. Meanwhile, in order to better compare the performance of the catalysts, the performance of some reported catalysts for MSR are listed in Table 3. Compared to other reported catalysts, the Cu/Zr-Al-3 catalyst exhibits good activity at lower temperatures with higher WHSVs. Compared to our previous work [32], the Cu/Zr-Al-3 catalyst also exhibits higher methanol conversion and higher hydrogen production rate at a lower temperature than the Cu/ZrO2-8Al2O3 catalyst in our previous work, which indicates that Cu/Zr-Al-3 performs well.

4. Experimental Section

4.1. Materials

NH3·H2O, Cu(NO3)2·3H2O, ZrOCl2·8H2O, ZrO(NO3)2·xH2O, Zr(NO3)4·5H2O, acetone and polyethylene oxide-polypropylene oxide-polyethylene oxide (P123) were purchased from Chron Chemical in China, and Al2O3 was purchased from Adamas-beta in China. All chemical reagents are analytically pure.

4.2. Synthesis of ZrO2-Al2O3

Zr-Al-1: ZrO2 was obtained by the calcination of ZrO(NO3)2·xH2O at 600 °C for 4 h. Then, Zr-Al-1 was obtained by mechanically mixing ZrO2 and Al2O3 with a molar ratio of 1:8 via placing the ZrO2 and the Al2O3 in a mortar to grind evenly.
Zr-Al-2: Zr-Al-2 was obtained by the impregnation method to obtain the designed phase [33,66]. Firstly, Al2O3 was impregnated in the ZrOCl2·8H2O solution which was prepared by adding 30 grams of solid ZrOCl2·8H2O to 100 mL of deionized water (the Zr/Al molar ratio is 1:8). Then, the solid component was obtained by oscillating, oil bath drying and drying in oven at 80 °C for 12 h. Finally, the solid component was calcined at 600 °C for 4 h to obtain Zr-Al-2.
Zr-Al-3: Zr-Al-3 was obtained by our previous method [53]. Zr(NO3)4·5H2O (30 g) was dissolved in 100 mL ultra-pure water and stirred until completely dissolved. Then, NH3·H2O was added with stirring to pH=9.5 to produce zirconium hydroxide precipitation. After stirring for another hour, the mixture was transferred to a hydrothermal reactor (200 mL) where it was kept at 140 °C for 24 h. The precipitate was then filtered and the resulting solid portion as rinsed with deionized water to pH=7. The solid was then dried at 80 °C for 12 h and calcined at 600 °C for 4 h to obtain ZrO2. Finally, Zr-Al-3 as obtained by the mechanical mixing of ZrO2 and Al2O3 with a molar ratio of 1:8.
Zr-Al-4: Zr-Al-4 was obtained by the same method as Zr-Al-3. The only difference is that the precursor was ZrOCl2·8H2O instead of Zr(NO3)4·5H2O.
Zr-Al-5: 9.25 g ZrO(NO3)2·xH2O and 6.96 g P123 were dissolved in ultrapure water. Subsequently, ammonia was added to adjust the pH of the solution to 11, and the mixture was stirred at 88 °C for 24 h. Then, the mixture was washed three times with acetone and three times with water. Finally, ZrO2 was dried at 80 °C for 12 h and calcined at 600 °C for 4 h. Zr-Al-5 was obtained by mechanically mixing ZrO2 and Al2O3 with a molar ratio of 1:8.

4.3. Synthesis of Cu/ZrO2-Al2O3

The Cu/ZrO2-Al2O3 catalyst was prepared by the impregnation method. Copper (II) nitrate trihydrate (0.4 g) was dissolved in 10 mL ultrapure water by ultrasound for 20 min. Then, the designed amount of prepared ZrO2-Al2O3 was added into the solution. Subsequently, the slurry was maintained at 25 °C for 24 h under stirring, and then dried at 80 °C to remove the remaining water and desiccated at 110 °C in a drying oven overnight. Lastly, the prepared solid was calcined in air at 350 °C (2 °C/min) for 4 h. The prepared catalysts were named as 5 wt% Cu/Zr-Al-x, where x = 1, 2, 3, 4, 5.

4.4. Catalyst Activity Study

The catalyst activity test was performed in a stainless-steel fixed bed reactor of length 60 cm and 1.4 cm inside diameter at atmospheric pressure, with 0.5 g prepared catalyst (20–40 mesh) used. The temperature was increased to 300 °C at 10 °C/min in an Ar atmosphere at 30 mL/min. Then, the catalyst was reduced at 300 °C for 2 h under a hydrogen and argon atmosphere at a total flow rate of 60 mL/min (VH2:VAr = 1:1). After cooling to 200 °C, the feed flow mixture (nMeOH:nH2O = 1:1) was introduced through the micro pump at 4 mL/h. Then, the mixture was vaporized by the carburetor and preheater at 135 °C. The performance of the Cu/Zr-Al-x catalysts over the temperature range of 200 to 300 °C was investigated. Finally, product gas flow was detected by a mass flowmeter and all the gaseous outcomes were detected by on-line gas chromatography (GC Plot-C 2000 Capillary column). The gas chromatography model used is PANNA-A91 and the specific parameters were as follows: column temperature: 160 °C, detector temperature: 200 °C, injector temperature: 120 °C and carrier (Ar) gas flow: 20 mL/min.
According to the literature, methanol conversion ( X M e O H ), CO selectivity ( S C O ) and H2 production rate ( P H 2 ) are calculated using the following Equations [36,63]:
X M e O H % = n C O + n C O 2 + n C H 4 n M e O H × 100
S C O % = n C O n C O + n C O 2 + n C H 4 × 100
P H 2 [ μ m o l / ( g c a t · s ) ] = n H 2 × 10 6 m c a t a l y s t × 60
where n C O , n C H 4 , n C O 2 and n H 2 are the product gas molar flow rates of CO, CH4, CO2 and H2 (mol/min) and n M e O H is the feed of molar rates of methanol (mol/min).

4.5. Catalyst Characterizations

The calcined catalysts were reduced at 300 °C for 2 h under a hydrogen and argon atmosphere at a total flow rate of 60 mL/min (VH2:VAr = 1:1). Then, the reduced catalysts were passivated for 12 h under an atmosphere of 5% O2/Ar to prevent catalyst oxidation according to the literature [67,68].
The actual loading of Cu on all calcined catalysts were analyzed by inductively coupled plasma atomic emission spectroscopy (ICP-AES) and was carried out on an Optima 8000 from Perkin Elmer device in Waltham, MA, USA. The prepared catalysts were absolutely dissolved in mixed acid (3 mL HCl, 1 mL HNO3, three or four drops of HF) before testing.
The texture and chemical properties of the reduced catalysts were analyzed by nitrogen adsorption on a Tristar II 3020 from Micromeritics company in Norcross, GA, USA. Approximately 120 mg of catalysts were pretreated at 120 °C for 2 h and 300 °C for 2 h under vacuum.
X-ray diffraction (XRD) of the catalysts was tested on a Shimazu XRD-6100 diffractometer from PANalytical B.V company in Shanghai, China with Cu-Kα radiation (λ = 1.5406 Å) at 40 eVkV and 30 mA, over a 2θ range of 20–80°. The crystal size (d) was calculated by the Debye–Scherrer Equation (6).
d = K λ / ( β cos θ )
where K is the Scherrer constant, λ is the wavelength of the incident X-ray, β is the FWHM of the diffraction peak and θ is the Bragg diffraction angle.
The reduction behavior of the calcined catalysts was measured by hydrogen temperature programmed reduction (H2-TPR) on an Autochem II 2920 from Micromeritics company in USA. Approximately 100 mg of catalysts were treated in ab argon atmosphere at 140 °C for 30 min and then cooled to 50 °C. After that, the gas flow was changed to 5% H2 + N2 (totally 50 mL/min) and the TPR test was operated with a temperature increasing rate of 5 °C/min from 50 °C to 800 °C.
The H2 pulse chemisorption of the catalysts was performed at 40 °C by injecting loop gas (5% H2-He totally 50 mL/min) repeatedly until the adsorption became saturated. Metal surface area (S) and dispersion (D) were calculated by the following (Equations (7) and (8)):
S = V m × A x × S F × N A
D = V m × S F × M / ( 100 L )
where Vm is the adsorbed hydrogen volume, Ax is the cross-sectional area of metal atoms, SF is the chemisorption measurement number, NA is the Avogadro constant, M is the metal molar mass and L is the metal load.
O2-TPD was carried out to test the types of adsorption oxygen on catalysts’ surfaces. This was conducted from 40 °C to 900 °C at 10 °C/min under a helium atmosphere (50 mL/min) after the adsorption of O2 was saturated.
The surface elemental chemical valence states of reduced catalysts were analyzed by X-ray photoelectron spectroscopy (XPS) and Auger electron spectroscopy (AES) characterization, using an AXIS Ultra DLD spectrometer from Krato company in Manchester, UK. All the binding energies were corrected by the peak of C1s at 284.6 eV.
The catalyst samples were subjected to transmission electron microscopy (TEM) and high-resolution transmission electron microscopy (HRTEM) characterization using an FEI Talos F200S apparatus (from Thermo Fisher Scientific company, Waltham, MA, USA), operated with 0.25 nm resolution and 10–10,000 magnification.

5. Conclusions

In this paper, the catalyst with 81.4% monoclinic ZrO2 crystal phases (Cu/Zr-Al-3) exhibited the highest methanol conversion (88.5%) and the highest hydrogen production rate (104.6 μmol/gcat·s) at 275 °C. This catalyst showed the best reducing properties and many more oxygen vacancies in its surface. The amount of oxygen vacancies can affect the chemical valence state of the active center, and accordingly affect the catalytic activity. In addition, the presence of characteristic t-ZrO2 (002) and m-ZrO2 (111) lattice planes on the Cu/Zr-Al-3 catalyst may influence the growth of copper species and promote catalytic activity. This study not only successfully prepared MSR catalysts with high activity and stability but also confirmed the feasibility of enhancing the catalytic activity of the MSR reaction by altering the ZrO2 crystal phase. This provides guidance for designing and developing efficient MSR reaction catalysts in mixed-phase systems.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal14080480/s1, Figure S1: 9h stability of Cu/Zr-Al-3 catalyst; Figure S2: Effect of m-ZrO2 ratio and oxygen vacancy amount on methanol conversion; Figure S3: Cu Auger spectra of the reduced catalysts; Table S1: The proportion of different crystalline ZrO2 in catalysts after calcination and reduction.

Author Contributions

M.S.: investigation, methodology, data curation and writing—original draft preparation; L.L.: investigation and methodology; X.W.: investigation and methodology; H.C.: investigation and methodology; C.H.: project administration, funding acquisition, supervision, resources and writing—reviewing and editing; G.L.: supervision, methodology, formal analysis, validation and writing—reviewing and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was founded by the Application Foundation Program of Sichuan Province (No. 2023YFG0106), the 111 Center (B17030) and the Fundamental Research Funds for the Central Universities. The APC was funded by the Fundamental Research Funds for the Central Universities.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors sincerely thank the Analysis and Testing Center of Sichuan University for the characterization.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Kwon, B.W.; Kim, G.S.; Oh, J.H.; Jang, S.C.; Yoon, S.P.; Han, J.; Nam, S.W.; Ham, H.C. H2 Production from Reforming Biogas (CH4 + CO2) in the Ni-doped Sr0.92 Y0.08 TiO3 Perovskite Catalyst. Energy Procedia 2017, 105, 1942–1947. [Google Scholar] [CrossRef]
  2. Liu, H.; Hadjltaief, H.B.; Benzina, M.; Gálvez, M.E.; Da Costa, P. Natural clay based nickel catalysts for dry reforming of methane: On the effect of support promotion (La, Al, Mn). Int. J. Hydrogen Energy 2019, 44, 246–255. [Google Scholar] [CrossRef]
  3. Sutthiumporn, K.; Maneerung, T.; Kathiraser, Y.; Kawi, S. CO2 dry-reforming of methane over La0.8Sr0.2Ni0.8M0.2O3 perovskite (M = Bi, Co, Cr, Cu, Fe): Roles of lattice oxygen on C–H activation and carbon suppression. Int. J. Hydrogen Energy 2012, 37, 11195–11207. [Google Scholar] [CrossRef]
  4. Nunes, P.; Oliveira, F.; Hamacher, S.; Almansoori, A. Design of a hydrogen supply chain with uncertainty. Int. J. Hydrogen Energy 2015, 40, 16408–16418. [Google Scholar] [CrossRef]
  5. Balcombe, P.; Speirs, J.; Johnson, E.; Martin, J.; Brandon, N.; Hawkes, A. The carbon credentials of hydrogen gas networks and supply chains. Renew. Sustain. Energy Rev. 2018, 91, 1077–1088. [Google Scholar] [CrossRef]
  6. Palo, D.R.; Dagle, R.A.; Holladay, J.D. Methanol Steam Reforming for Hydrogen Production. Chem. Rev. 2007, 107, 3992–4021. [Google Scholar] [CrossRef]
  7. Chen, W.H.; Lin, B.J. Effect of microwave double absorption on hydrogen generation from methanol steam reforming. Int. J. Hydrogen Energy 2010, 35, 1987–1997. [Google Scholar] [CrossRef]
  8. Silva, H.; Mateos-Pedrero, C.; Ribeirinha, P.; Boaventura, M.; Mendes, A. Low-temperature methanol steam reforming kinetics over a novel CuZrDyAl catalyst. React. Kinet. Mech. Catal. 2015, 115, 321–339. [Google Scholar] [CrossRef]
  9. Ribeirinha, P.; Mateos-Pedrero, C.; Boaventura, M.; Sousa, J.; Mendes, A. CuO/ZnO/Ga2O3 catalyst for low temperature MSR reaction: Synthesis, characterization and kinetic model. Appl. Catal. B Environ. 2018, 221, 371–379. [Google Scholar] [CrossRef]
  10. 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]
  11. Basile, A.; Parmaliana, A.; Tosti, S.; Iulianelli, A.; Gallucci, F.; Espro, C.; Spooren, J. Hydrogen production by methanol steam reforming carried out in membrane reactor on Cu/Zn/Mg-based catalyst. Catal. Today 2008, 137, 17–22. [Google Scholar] [CrossRef]
  12. Davda, R.R.; Shabaker, J.W.; Huber, G.W.; Cortright, R.D.; Dumesic, J.A. A review of catalytic issues and process conditions for renewable hydrogen and alkanes by aqueous-phase reforming of oxygenated hydrocarbons over supported metal catalysts. Appl. Catal. B Environ. 2005, 56, 171–186. [Google Scholar] [CrossRef]
  13. Cacciola, G.; Antonucci, V.; Freni, S. Technology up date and new strategies on fuel cells. J. Power Sources 2001, 100, 67–79. [Google Scholar] [CrossRef]
  14. King, J.M.; O’Day, M.J. Applying fuel cell experience to sustainable power products. J. Power Sources 2000, 86, 16–22. [Google Scholar] [CrossRef]
  15. Ramirez, D.; Beites, L.; Blazquez, F.; Ballesteros, J. Distributed generation system with PEM fuel cell for electrical power quality improvement. Int. J. Hydrogen Energy 2008, 33, 4433–4443. [Google Scholar] [CrossRef]
  16. Stone, C.; Morrison, A.E. From curiosity to “power to change the world”. Solid State Ion. 2002, 152-153, 1–13. [Google Scholar] [CrossRef]
  17. Dhar, H.P.; Christner, L.G.; Kush, A.K. Nature of CO Adsorption during H2 Oxidation in Relation to Modeling for CO Poisoning of a Fuel Cell Anode. J. Electrochem. Soc. 1987, 134, 3021–3026. [Google Scholar] [CrossRef]
  18. Ranganathan, E.S.; Bej, S.K.; Thompson, L.T. Methanol steam reforming over Pd/ZnO and Pd/CeO2 catalysts. Appl. Catal. A Gen. 2005, 289, 153–162. [Google Scholar] [CrossRef]
  19. Suwa, Y.; Ito, S.-i.; Kameoka, S.; Tomishige, K.; Kunimori, K. Comparative study between Zn–Pd/C and Pd/ZnO catalysts for steam reforming of methanol. Appl. Catal. A Gen. 2004, 267, 9–16. [Google Scholar] [CrossRef]
  20. 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]
  21. Papavasiliou, J.; Avgouropoulos, G.; Ioannides, T. Steady-state isotopic transient kinetic analysis of steam reforming of methanol over Cu-based catalysts. Appl. Catal. B Environ. 2009, 88, 490–496. [Google Scholar] [CrossRef]
  22. Wu, G.-S.; Mao, D.-S.; Lu, G.-Z.; Cao, Y.; Fan, K.-N. The Role of the Promoters in Cu Based Catalysts for Methanol Steam Reforming. Catal. Lett. 2009, 130, 177–184. [Google Scholar] [CrossRef]
  23. 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]
  24. 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]
  25. Alejo, L.; Lago, R.; Pena, 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]
  26. Paglia, G.; Buckley, C.E.; Udovic, T.J.; Rohl, A.L.; Jones, F.; Maitland, C.F.; Connolly, J. Boehmite-Derived γ-Alumina System. 2. Consideration of Hydrogen and Surface. Effects. Chem. Mater. 2004, 16, 1914–1923. [Google Scholar] [CrossRef]
  27. Byun, M.Y.; Kim, J.S.; Park, D.W.; Lee, M.S. Influence of calcination temperature on the structure and properties of Al2O3 as support for Pd catalyst. Korean, J. Chem. Eng. 2018, 35, 1083–1088. [Google Scholar] [CrossRef]
  28. 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]
  29. Witoon, T.; Chalorngtham, J.; Dumrongbunditkul, P.; Chareonpanich, M.; Limtrakul, J. CO2 hydrogenation to methanol over Cu/ZrO2 catalysts: Effects of zirconia phases. Chem. Eng. J. 2016, 293, 327–336. [Google Scholar] [CrossRef]
  30. Tada, S.; Katagiri, A.; Kiyota, K.; Honma, T.; Kamei, H.; Nariyuki, A.; Uchida, S.; Satokawa, S. Cu Species Incorporated into Amorphous ZrO2 with High Activity and Selectivity in CO2-to-Methanol Hydrogenation. J. Phys. Chem. C 2018, 122, 5430–5442. [Google Scholar] [CrossRef]
  31. Samson, K.; Śliwa, M.; Socha, R.P.; Góra-Marek, K.; Mucha, D.; Rutkowska-Zbik, D.; Paul, J.F.; Ruggiero-Mikołajczyk, M.; Grabowski, R.; Słoczyński, J. Influence of ZrO2 Structure and Copper Electronic State on Activity of Cu/ZrO2 Catalysts in Methanol Synthesis from CO2. ACS Catal. 2014, 4, 3730–3741. [Google Scholar] [CrossRef]
  32. Song, M.; Zeng, W.; Li, L.; Wu, X.; Li, G.; Hu, C. Effect of the Zr/Al molar ratio on the performance of Cu/ZrO2–Al2O3 catalysts for methanol steam reforming. Ind. Eng. Chem. Res. 2023, 62, 3898–3908. [Google Scholar] [CrossRef]
  33. Lv, X.; Cheng, D.H.; Shou, X.K.; Liu, J.C.; Xu, H.T. Bifunctional catalyst Ga2O3-mZrO2/SAPO-34 for CO2 hydrogenation: Ga2O3-mZrO2 improving light olefins. Catal. Sci. Technol. 2024, 14, 3652–3659. [Google Scholar] [CrossRef]
  34. Zhou, Y.; Liu, L.; Li, G.; Hu, C. Insights into the Influence of ZrO2 Crystal Structures on Methyl Laurate Hydrogenation over Co/ZrO2 Catalysts. ACS Catal. 2021, 11, 7099–7113. [Google Scholar] [CrossRef]
  35. Mateos-Pedrero, C.; Azenha, C.; Pacheco Tanaka, D.A.; Sousa, J.M.; Mendes, A. The influence of the support composition on the physicochemical and catalytic properties of Cu catalysts supported on Zirconia-Alumina for methanol steam reforming. Appl. Catal. B Environ. 2020, 277, 119243. [Google Scholar] [CrossRef]
  36. Chen, M.; Sun, G.; Wang, Y.; Liang, D.; Li, C.; Wang, J.; Liu, Q. Steam reforming of methanol for hydrogen production over attapulgite-based zeolite-supported Cu-Zr catalyst. Fuel 2022, 314, 122733. [Google Scholar] [CrossRef]
  37. Snyder, R.L. The Use of Reference Intensity Ratios in X-Ray Quantitative Analysis. Powder Diffr. 1992, 7, 186–193. [Google Scholar] [CrossRef]
  38. Yu, Q.; Yao, X.; Zhang, H.; Gao, F.; Dong, L. Effect of ZrO2 addition method on the activity of Al2O3-supported CuO for NO reduction with CO: Impregnation vs. coprecipitation. Appl. Catal. A Gen. 2012, 423, 42–51. [Google Scholar] [CrossRef]
  39. Mohtashami, Y.; Taghizadeh, M. Performance of the ZrO2 promoted Cu-ZnO catalyst supported on acetic acid-treated MCM-41 in methanol steam reforming. Int. J. Hydrogen Energy 2019, 44, 5725–5738. [Google Scholar] [CrossRef]
  40. Luo, M.-F.; Fang, P.; He, M.; Xie, Y.-L. In situ XRD, Raman, and TPR studies of CuO/Al2O3 catalysts for CO oxidation. J. Mol. Catal. A Chem. 2005, 239, 243–248. [Google Scholar] [CrossRef]
  41. Liang, C.; Li, X.; Qu, Z.; Tade, M.; Liu, S. The role of copper species on Cu/γ-Al2O3 catalysts for NH3–SCO reaction. Appl. Surf. Sci. 2012, 258, 3738–3743. [Google Scholar] [CrossRef]
  42. Li, S.; Wang, Y.; Yang, B.; Guo, L. A highly active and selective mesostructured Cu/AlCeO catalyst for CO2 hydrogenation to methanol. Appl. Catal. A Gen. 2019, 571, 51–60. [Google Scholar] [CrossRef]
  43. Liu, R.; Wang, C.A. Synthesis of aluminum-doped mesoporous zirconia with improved thermal stability. Microporous Mesoporous Mater. 2014, 186, 1–6. [Google Scholar] [CrossRef]
  44. Bin, F.; Wei, X.; Li, B.; Hui, K.S. Self-sustained combustion of carbon monoxide promoted by the Cu-Ce/ZSM-5 catalyst in CO/O2/N2 atmosphere. Appl. Catal. B Environ. 2015, 162, 282–288. [Google Scholar] [CrossRef]
  45. Liu, P.; Hensen, E.J. Highly efficient and robust Au/MgCuCr2O4 catalyst for gas-phase oxidation of ethanol to acetaldehyde. J. Am. Chem. Soc. 2013, 135, 14032–14035. [Google Scholar] [CrossRef] [PubMed]
  46. Oliveira, M.L.M.d.; Silva, C.M.; Moreno-Tost, R.; Farias, T.L.; Jiménez-López, A.; Rodríguez-Castellón, E. A study of copper-exchanged mordenite natural and ZSM-5 zeolites as SCR–NOx catalysts for diesel road vehicles: Simulation by neural networks approach. Appl. Catal. B Environ. 2009, 88, 420–429. [Google Scholar] [CrossRef]
  47. Azenha, C.; Lagarteira, T.; Mateos-Pedrero, C.; Mendes, A. Production of hydrogen from methanol steam reforming using CuPd/ZrO2 catalysts–Influence of the catalytic surface on methanol conversion and CO selectivity. Int. J. Hydrogen Energy 2021, 46, 17490–17499. [Google Scholar] [CrossRef]
  48. Li, H.; Tian, H.; Chen, S.; Sun, Z.; Liu, T.; Liu, R.; Assabumrungrat, S.; Saupsor, J.; Mu, R.; Pei, C.; et al. Sorption enhanced steam reforming of methanol for high-purity hydrogen production over Cu-MgO/Al2O3 bifunctional catalysts. Appl. Catal. B Environ. 2020, 276, 119052. [Google Scholar] [CrossRef]
  49. Biesinger, M.C. Advanced analysis of copper X-ray photoelectron spectra. Surf. Interface Anal. 2017, 49, 1325–1334. [Google Scholar] [CrossRef]
  50. Pakharukova, V.P.; Moroz, E.M.; Kriventsov, V.V.; Larina, T.V.; Boronin, A.I.; Dolgikh, L.Y.; Strizhak, P.E. Structure and state of copper oxide species supported on Yttria-stabilized zirconia. J. Phys. Chem. C 2009, 113, 21368–21375. [Google Scholar] [CrossRef]
  51. Li, W.; Liu, Y.; Mu, M.; Ding, F.; Liu, Z.; Guo, X.; Song, C. Organic acid-assisted preparation of highly dispersed Co/ZrO2 catalysts with superior activity for CO2 methanation. Appl. Catal. B Environ. 2019, 254, 531–540. [Google Scholar] [CrossRef]
  52. Jiang, F.; Wang, S.; Liu, B.; Liu, J.; Wang, L.; Xiao, Y.; Xu, Y.; Liu, X. Insights into the Influence of CeO2 Crystal Facet on CO2 Hydrogenation to Methanol over Pd/CeO2 Catalysts. ACS Catal. 2020, 10, 11493–11509. [Google Scholar] [CrossRef]
  53. Zhou, Y.; Liu, X.; Yu, P.; Hu, C. Temperature-tuned selectivity to alkanes or alcohol from ethyl palmitate deoxygenation over zirconia-supported cobalt catalyst. Fuel 2020, 278, 118295. [Google Scholar] [CrossRef]
  54. Wang, Y.; Zhao, Q.; Li, L.; Hu, C.; Da Costa, P. Dry reforming of methane over Ni–ZrOx catalysts doped by manganese: On the effect of the stability of the structure during time on stream. Appl. Catal. A Gen. 2021, 617, 118120. [Google Scholar] [CrossRef]
  55. Cheng, Z.; Jiang, C.; Sun, X.; Lan, G.; Wang, X.; He, L.; Li, Y.; Tang, H.; Li, Y. Insights into the inducing effect of aluminum on Cu–ZnO synergy for methanol steam reforming. Ind. Eng. Chem. Res. 2022, 61, 11699–11707. [Google Scholar] [CrossRef]
  56. He, C.; Yu, Y.; Chen, C.; Yue, L.; Qiao, N.; Shen, Q.; Chen, J.; Hao, Z. Facile preparation of 3D ordered mesoporous CuOx–CeO2 with notably enhanced efficiency for the low temperature oxidation of heteroatom-containing volatile organic compounds. RSC Adv. 2013, 3, 19639–19656. [Google Scholar] [CrossRef]
  57. Yang, H.; Zeng, Y.; Zhou, Y.; Du, X.; Li, D.; Hu, C. One-step synthesis of highly active and stable Ni-ZrO2 catalysts for the conversion of methyl laurate to alkanes. J. Catal. 2022, 413, 297–310. [Google Scholar] [CrossRef]
  58. Fu, M.; Yue, X.; Ye, D.; Ouyang, J.; Huang, B.; Wu, J.; Liang, H. Soot oxidation via CuO doped CeO2 catalysts prepared using coprecipitation and citrate acid complex-combustion synthesis. Catal. Today 2010, 153, 125–132. [Google Scholar] [CrossRef]
  59. Taghizadeh, M.; Abbandanak, M.H. Production of hydrogen via methanol steam reforming over mesoporous CeO2–Cu/KIT-6 nanocatalyst: Effects of polar aprotic tetrahydrofuran solvent and ZrO2 promoter on catalytic performance. Int. J. Hydrogen Energy 2022, 47, 16362–16374. [Google Scholar] [CrossRef]
  60. Li, D.; Xu, F.; Tang, X.; Dai, S.; Pu, T.; Liu, X.; Tian, P.; Xuan, F.; Xu, Z.; Wachs, I.E.; et al. Induced activation of the commercial Cu/ZnO/Al2O3 catalyst for the steam reforming of methanol. Nat. Catal. 2022, 5, 99–108. [Google Scholar] [CrossRef]
  61. Khanchi, M.; Mousavian, S.M.A.; Soltanali, S. Methanol steam reforming over Cu supported on SiO2, amorphous SiO2-Al2O3, and Al2O3 catalysts: Influence of support nature. Int. J. Energy Res. 2021, 46, 3057–3071. [Google Scholar] [CrossRef]
  62. Shahmohammadi, N.; Rezaei, M.; Alavi, S.M.; Akbari, E. Methanol steam reforming over highly efficient CuO–Al2O3 nanocatalysts synthesized by the solid-state mechanochemical method. Int. J. Hydrogen Energy 2023, 48, 13139–13150. [Google Scholar] [CrossRef]
  63. Cheng, Z.; Zhou, W.; Lan, G.; Sun, X.; Wang, X.; Jiang, C.; Li, Y. High-performance Cu/ZnO/Al2O3 catalysts for methanol steam reforming with enhanced Cu-ZnO synergy effect via magnesium assisted strategy. J. Energy Chem. 2021, 63, 550–557. [Google Scholar] [CrossRef]
  64. Li, L.; Chen, J.; Zeng, C.; Liu, Q.; Hu, H.; Huang, Q.; Chen, X. Preparation of CuZnZrAl catalysts by coprecipitation-ammonia method for methanol steam reforming and the effect of promoters Y and Ce. Mol. Catal. 2023, 537, 112887. [Google Scholar] [CrossRef]
  65. Chen, Y.; Huang, Y. Influence of ceria existence form on deactivation behavior of Cu-Ce/SBA-15 catalysts for methanol steam reforming. Int. J. Hydrogen Energy 2023, 48, 1323–1336. [Google Scholar] [CrossRef]
  66. Abate, S.; Mebrahtu, C.; Giglio, E.; Deorsola, F.; Bensaid, S.; Perathoner, S.; Pirone, R.; Centi, G. Catalytic Performance of γ-Al2O3–ZrO2–TiO2–CeO2 Composite Oxide Supported Ni-Based Catalysts for CO2 Methanation. Int. J. Hydrogen Energy 2016, 55, 4451–4460. [Google Scholar] [CrossRef]
  67. Zeng, W.; Li, L.; Song, M.; Wu, X.; Li, G.; Hu, C. The effect of different atmosphere treatments on the performance of Ni/Nb–Al2O3 catalysts for methane steam reforming. Int. J. Hydrogen Energy 2023, 48, 6358–6369. [Google Scholar] [CrossRef]
  68. Cui, C.; Wang, Y.; Du, X.; Li, L.; Song, M.; Zeng, W.; Li, G.; Hu, C. Carbon resistance of xNi/HTASAO5 catalyst for the production of H2 via CO2 reforming of methane. Int. J. Hydrogen Energy 2021, 46, 20835–20847. [Google Scholar] [CrossRef]
Figure 1. (a) Methanol conversion, (b) hydrogen production rate and (c) CO selectivity. Reaction conditions: nCH3OH:nH2O = 1:1, mcat = 0.5 g, feed flow rate = 4 mL/h and 1 atm, WHSV = 9.9 h−1.
Figure 1. (a) Methanol conversion, (b) hydrogen production rate and (c) CO selectivity. Reaction conditions: nCH3OH:nH2O = 1:1, mcat = 0.5 g, feed flow rate = 4 mL/h and 1 atm, WHSV = 9.9 h−1.
Catalysts 14 00480 g001
Figure 2. Comparison of the activities of different catalysts at 275 °C.
Figure 2. Comparison of the activities of different catalysts at 275 °C.
Catalysts 14 00480 g002
Figure 3. (a) XRD pattern of calcined catalysts. (b) XRD pattern of reduced catalysts.
Figure 3. (a) XRD pattern of calcined catalysts. (b) XRD pattern of reduced catalysts.
Catalysts 14 00480 g003
Figure 4. The proportion of different crystal phase ZrO2 in calcined (left column) and reduced (right column) catalysts.
Figure 4. The proportion of different crystal phase ZrO2 in calcined (left column) and reduced (right column) catalysts.
Catalysts 14 00480 g004
Figure 5. H2-TPR profiles of the calcined catalysts.
Figure 5. H2-TPR profiles of the calcined catalysts.
Catalysts 14 00480 g005
Figure 6. (a) Nitrogen adsorption and desorption isotherms and (b) pore diameter distribution of reduced catalysts.
Figure 6. (a) Nitrogen adsorption and desorption isotherms and (b) pore diameter distribution of reduced catalysts.
Catalysts 14 00480 g006
Figure 7. XPS curves of (a) Cu 2p, (b) O 1s and (c) Zr 3d for reduced catalysts; (d) O2-TPD of different catalysts.
Figure 7. XPS curves of (a) Cu 2p, (b) O 1s and (c) Zr 3d for reduced catalysts; (d) O2-TPD of different catalysts.
Catalysts 14 00480 g007
Figure 8. TEM and HRTEM images of the reduced catalysts.
Figure 8. TEM and HRTEM images of the reduced catalysts.
Catalysts 14 00480 g008
Table 1. Texture and properties of the catalysts and supports.
Table 1. Texture and properties of the catalysts and supports.
CatalystCu Content a (wt%)SBET b (m2·g−1)VBJH c (cm3·g−1)Dp d (nm)Cu Dispersion e
(%)
Cu
Surface e Area (m2/gcat)
Cu Particle Size f
(nm)
ZrO2 Particle Size f
(nm)
Cu/Zr-Al-15.1116.40.3812.74.635.4035.816.0
Cu/Zr-Al-24.586.40.249.71.982.2435.513.8
Cu/Zr-Al-35.1106.70.3913.67.428.4831.016.6
Cu/Zr-Al-45.1113.90.3913.32.192.5532.316.3
Cu/Zr-Al-54.7117.60.4013.02.012.3526.917.2
a: measured by ICP-AES; b: calculated by the BET equation; c: BJH adsorption pore volume; d: BJH adsorption average pore diameter; e: calculated by H2 pulse chemisorption; f: calculated by the Debye–Scherrer equation from the diffraction peaks of Cu and ZrO2 in Figure 3.
Table 2. The relative content of elements on the reduced catalyst surface.
Table 2. The relative content of elements on the reduced catalyst surface.
CatalystsCu2+
(%)
Cu1+
and
Cu0
(%)
Zr3+
(%)
Zr4+
(%)
OL a
(%)
OS b
(%)
Ov cOv d
(mmol/g)
Cu/Zr-Al-121.079.087.712.370.429.60.2960.102
Cu/Zr-Al-231.168.979.820.279.520.50.2050.056
Cu/Zr-Al-310.789.389.110.965.834.20.3420.136
Cu/Zr-Al-423.976.185.614.471.728.30.2830.075
Cu/Zr-Al-525.574.583.216.874.625.40.2540.063
a: OL: lattice oxygen species. b: OS: surface oxygen species. c: Ov: amount of surface oxygen vacancy. Ov = OS/(OL + OS). d: Ov: measurement by O2-TPD.
Table 3. Performance of different copper-based catalysts for MSR.
Table 3. Performance of different copper-based catalysts for MSR.
CatalystTemperature
(°C)
S/C aSpace-Time Ratio (h−1)Conversion
(%)
H2
Production Rate (μmol/gcat·s)
CO Selectivity (%)References
ZrO2-CeO2-Cu/KIT-630022 (WHSV)96SH2 = 99.8%0.7[59]
Cu/ZnO-3Al250115.3 (WHSV)57.3-1.3[55]
Cu/ZnAl-R102251.36 (WHSV)67--[60]
1Cu/1Zr/AZ4001.210.8 (WHSV)90.6182.03.2[36]
Cu/ASA13000.51.8 (WHSV)91.0SH2 = 96.8%9.6[61]
CuO-Al2O3350536,000(GHSV)72.2SH2 = 87.5%0.5[62]
CuZnAl-5Mg20013.84 (WHSV)7047.21.0[63]
Cu/ZnCeZrAl2501.22.6 (WHSV)96.1-0.37[64]
Cu-Ce/SBA-152701.23 (WHSV)90.1-2.1[65]
Cu/ZrO2-8Al20330019.9 (WHSV)84.598.71.5[32]
Cu/Zr-Al-327519.9 (WHSV)88.5104.61.9This work
a: molar ratio of steam to methanol.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Song, M.; Li, L.; Wu, X.; Cai, H.; Li, G.; Hu, C. The Influence of the ZrO2 Crystal Phase on Cu/ZrO2-Al2O3 Catalysts in Methanol Steam Reforming. Catalysts 2024, 14, 480. https://doi.org/10.3390/catal14080480

AMA Style

Song M, Li L, Wu X, Cai H, Li G, Hu C. The Influence of the ZrO2 Crystal Phase on Cu/ZrO2-Al2O3 Catalysts in Methanol Steam Reforming. Catalysts. 2024; 14(8):480. https://doi.org/10.3390/catal14080480

Chicago/Turabian Style

Song, Mouxiao, Li Li, Xueshuang Wu, Haiqing Cai, Guiying Li, and Changwei Hu. 2024. "The Influence of the ZrO2 Crystal Phase on Cu/ZrO2-Al2O3 Catalysts in Methanol Steam Reforming" Catalysts 14, no. 8: 480. https://doi.org/10.3390/catal14080480

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