Next Article in Journal
Simulation of CO2 Catalytic Absorption Process Using Amine Solutions Based on the Lattice Boltzmann Method
Previous Article in Journal
Synergistic Effect of Physicochemical Properties of Ni Nanofibrous Catalysts on Catalytic Performance for Methane Partial Oxidation
Previous Article in Special Issue
Study on the Effect of Precious Metal Loading and Pt/Pd Ratio on Gaseous Pollutant Emissions from Diesel Engines
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Mn/Cu Ratio on the Structure–Performance Relationship of Spinel-Type Mn–Cu/Al2Ox Catalysts for Methanol Steam Reforming

Guangxi Key Laboratory for High-Value Utilization of Manganese Resources, College of Chemistry & Materials, Guangxi Minzu Normal University, Chongzuo 532200, China
*
Author to whom correspondence should be addressed.
Catalysts 2025, 15(11), 1091; https://doi.org/10.3390/catal15111091
Submission received: 7 October 2025 / Revised: 11 November 2025 / Accepted: 14 November 2025 / Published: 20 November 2025

Abstract

The development of highly active, thermally stable, and low-CO-selective catalysts is critical for practical methanol steam reforming (MSR) to produce high-purity hydrogen for fuel cell applications. In this work, a series of Mn–Cu/Al2Ox catalysts with varying Mn/Cu/Al molar ratios were synthesized via co-precipitation and systematically investigated to establish the relationship between composition, structure, and catalytic performance. XRD analysis revealed the formation of spinel-type CuAl2O4 and MnAl2O4 phases, with Mn preferentially occupying octahedral B-sites to form MnAl2O4, thereby inducing lattice distortion and inhibiting grain growth. SEM and TEM–EDS mapping confirmed uniform elemental distribution and a porous nanoscale morphology, while H2-TPR results suggested that increasing the Mn/Cu ratio strengthens Mn–Cu interactions, shifts Cu2+ reduction to higher temperatures, and enhances Cu dispersion (up to 26.11 m2/g). XPS analysis indicated that Mn doping enriches Mn3+ species and facilitates oxygen vacancy formation, which promotes water–gas shift (WGS) activity and suppresses CO formation. Catalytic testing (240–300 °C) showed that Mn2Cu2Al4Ox achieved the highest methanol conversion while maintaining low CO selectivity; in contrast, reducing the Mn/Cu ratio increased CO selectivity, detrimental to hydrogen purification. Stability tests under continuous steam exposure for 24 h demonstrated minimal activity loss (~2%) and negligible increase in CO selectivity (<1%), confirming excellent hydrothermal stability. The results indicate that tailoring the Mn/Cu ratio optimizes the balance between redox properties and metallic Cu dispersion, offering a promising route to design low-CO, durable catalysts for on-site hydrogen generation via MSR.

1. Introduction

The growing shift toward low-carbon energy systems has heightened interest in hydrogen as a clean energy carrier. Among the various production routes, methanol steam reforming (SRM: CH3OH + H2O → CO2 + 3H2) is considered highly promising for on-site hydrogen generation due to methanol’s high hydrogen content, easy handling, and mild operation temperatures (200–300 °C) [1]. The core advantages of SRM-based hydrogen production focus on three aspects: (1) the raw material methanol has high hydrogen content and can be efficiently converted into hydrogen, (2) methanol exhibits stable physical properties, making it easy to store and transport, and (3) the reaction requires a mild operating temperature, only 200–300 °C, which reduces equipment and energy consumption costs [2]. However, the commercialization process of SRM technology is still limited by catalyst performance. The current core challenge lies in the lack of catalysts that simultaneously possess high activity, long-term thermal stability, and ultra-low CO selectivity (<1%). These three properties are exactly the key prerequisites for ensuring the stable operation of fuel cells and preventing CO from poisoning the electrodes, and they directly hinder the practical implementation of SRM technology [3,4].
Copper-based catalysts, particularly Cu/ZnO/Al2O3, remain the industrial benchmark for MSR due to their excellent C–H and O–H bond activation capabilities [5,6]. Yet, they are prone to Cu nanoparticle sintering at higher temperatures (>250 °C), structural degradation, and deactivation mechanisms, including carbon deposition [7,8]. Additionally, unwanted CO production via the reverse water–gas shift (RWGS) (CO2 + H2 ⇌ CO + H2O) reaction compromises fuel-cell compatibility [9,10]. D. Hammoud et al. found that 10% Cu/Al-400-500 exhibits the best catalytic activity, with about 75.44% of H2 yield and 51.87% of methanol conversion at 250 °C [11]. These shortcomings have steered research toward the development of improved supports and doping strategies to enhance performance [12]. Alumina (Al2O3) is a commonly employed support, valued for its thermal stability, high surface area, and its ability to form CuAl2O4 spinel phases that anchor Cu species, mitigate sintering, and extend catalyst life [13,14]. Spinel structures serve as reservoirs of Cu, facilitating the gradual migration of active species under reducing conditions, effectively stabilizing the catalytic activity [15]. Recent studies highlight that Mn incorporation into Al2O3 forms MnAl2O4 spinels, fostering stronger Cu–Mn interactions and further inhibiting sintering while enhancing redox behavior [16,17]. For instance, Mn-doped Cu/Al2O3 catalysts have shown promise in suppressing CO formation by promoting water–gas shift (WGS) activity through oxygen vacancy generation and improved redox cycling. Mao et al. found that the Cu–Al2O3 interface serves as the main active site in methanol steam reforming, guiding the construction of high-performance catalysts [18].
Recently, manganese (Mn) has gained attention as a promoter in Cu-based catalysts. Mn’s multi-valent nature (Mn2+/Mn3+/Mn4+) enables it to modulate redox behavior, enhance oxygen mobility, and create oxygen vacancies [19,20]. Incorporating Mn into Al2O3 can also form MnAl2O4 spinels, fostering stronger interaction with Cu and further inhibiting sintering while enhancing redox cycling [21,22]. Experimental studies reveal that Mn doping in Cu/Al2Ox systems enhances reducibility, modifies surface acidity-basicity, and suppresses CO formation by favoring water–gas shift (WGS) activity [23,24]. Component optimization, support effects, and surface modification strategies synergistically enhance the exposure of active sites, improve the anti-sintering capability, and boost the resistance to carbon deposition in copper-based catalysts for methanol steam reforming [7]. Despite these promising findings, systematic investigations of the effect of different Mn/Cu ratios on catalyst structure and SRM performance remain lacking.
Herein, we prepare Mn-doped Cu/Al2Ox catalysts via co-precipitation, varying the Mn/Cu ratio. We employ comprehensive characterizations (XRD, BET, TEM-EDS, H2-TPR, XPS) and assess SRM performance (240–300 °C). The goals are to understand how Mn influences CuAl2O4 phase formation, Cu dispersion, redox behavior, and CO suppression—providing rational design strategies for Mn–Cu/Al2Ox catalysts with enhanced SRM performance.

2. Results and Discussion

2.1. XRD Analysis

The XRD patterns of catalyst samples with different molar ratios are presented in Figure 1.
As shown in Figure 1, characteristic diffraction peaks of the CuAl2O4 spinel are observed at 31.3° (220 plane), 36.9° (311 plane), and 65.2° (440 plane) in the sample without manganese. With increasing Mn content in the catalysts, the diffraction peak intensity of the Mn2Al2O4 phase (36.1° (311)) significantly enhances [25]. This indicates that Mn preferentially occupies the B-sites of the spinel structure to form MnAl2O4. Concurrently, peak broadening occurs, which may be attributed to lattice distortion induced by Mn2+ doping, thereby suppressing crystallite growth. At excessive Mn concentrations, the formation of amorphous MnOx structures becomes dominant, leading to the weakening of spinel phase peaks. Consistent with findings by Dasireddy et al., Mn doping enhances sintering resistance [19].
When Cu content increases, the CuAl2O4 phase becomes predominant, evidenced by the highest intensity of the 36.9° (311) peak and improved crystallinity. Controlled variation in the Mn/Cu ratio reveals that high Mn content promotes MnAl2O4 formation, while high Cu content favors CuAl2O4 crystallization. At a Mn/Cu/Al molar ratio of 2:2:4, a composite spinel phase of Mn2Cu1Al4 forms, where the Mn-Cu synergistic effect balances redox properties and structural stability, potentially yielding superior performance in steam reforming of methanol (SRM). Subsequent characterizations via H2-TPR and XPS will be conducted to validate the valence distributions of Mn, Cu, and oxygen species. As Cu serves as the catalytic core for hydrogen production via methanol steam reforming, an optimal Mn/Cu/Al ratio is critical to simultaneously achieve high dispersion, structural stability, and catalytic activity—key factors for realizing efficient hydrogen generation from methanol [26].

2.2. SEM Analysis

The morphologies of MnCuAl catalysts with different molar ratios are depicted in Figure 2. The overall morphology exhibits grain aggregates composed of nanoscale particles, with mesoporous gaps existing between the particles. The molar ratio has an impact on the morphology; specifically, samples with higher Cu content, namely Mn2Cu2Al4 and Mn1Cu2Al5, exhibit more densely packed blocky structures with smoother surfaces and fewer pores [14]. Samples with higher Mn/Al ratios or elevated Al (Mn2Cu1Al5Ox) proportions exhibit a more loosely packed structure with abundant porosity, facilitating mass diffusion. This observation aligns with the established mechanism that Cu species are more prone to migration and neck growth during thermal treatment, whereas high-valence Mn ions effectively suppress sintering by stabilizing the structural framework. A significant improvement in particle dispersion is observed, which corroborates the XRD-derived conclusion that Mn doping effectively inhibits sintering by stabilizing the crystal lattice when Mn content is ≥2 (in atomic/molar ratio) [27]. An appropriate Mn/Cu ratio can enhance the specific surface area and optimize pore connectivity of the catalyst, as Mn2+ acts as a “structural directing agent” to suppress self-agglomeration of active components [28].

2.3. H2-TPR Analysis

The H2-TPR profiles of different catalysts are presented in Figure 3. As shown in Figure 3 (left), the Cu3Al5Ox catalyst exhibits the most intense reduction peak in the 220–300 °C range, consistent with its higher Cu content. The Mn2Cu2Al4Ox catalyst also displays a pronounced reduction peak in this temperature region, albeit with lower intensity compared to Cu3Al5Ox. These peaks are primarily attributed to the reduction of surface CuO species (Cu2+ → Cu0) [29]. With increasing Mn/Cu ratios, the reduction peaks shift toward higher temperatures, indicating enhanced dispersion of copper species, reduced reducibility, and improved thermal stability. This observation suggests a synergistic interaction between Mn and Cu, likely forming Mn-Cu composite oxides, which aligns with XRD analysis results. Such interactions contribute to superior high-temperature stability and sintering resistance during catalytic reactions [28].
Post-N2O oxidation and subsequent H2-TPR analysis (Figure 3, right) reveal that high-Cu catalysts (Mn:Cu < 1), such as Cu3Al5Ox, exhibit reduction peaks predominantly in the low-temperature region (~150 °C) with large peak areas, indicating facile reduction of CuO species and weak Cu-O bonding. Mn doping shifts the reduction temperature to higher values and increases reduction difficulty, thereby enhancing the sintering resistance of Cu-based catalysts [30].
The key structural properties derived from H2-TPR and BET analyses are summarized in Table 1. Notably, the Mn2Cu2Al4Ox sample exhibits the highest copper-specific surface area (S_Cu) of 26.11 m2/g. This finding, combined with its high copper dispersion (D_Cu), confirms that an optimal Mn/Cu ratio is crucial for achieving superior Cu dispersion and exposing a larger active Cu surface area, which is beneficial for the MSR reaction [31].

2.4. BET Analysis

The N2 physical adsorption/desorption isotherms and key textural properties of the catalysts are presented in Figure 4 and Table 1, respectively. All six catalysts exhibit Type IV isotherms with H3-type hysteresis loops, indicative of mesoporous structures formed by the stacking of plate-like particles [32]. As summarized in Table 1, the Mn doping significantly influences the specific surface area. The non-Mn-doped Cu3Al5Ox catalyst shows the smallest specific surface area (S_BET = 51.43 m2/g), whereas the Mn2Cu2Al4Ox sample achieves the highest S_BET of 76.44 m2/g. This enhancement is attributed to the optimal Mn/Cu ratio promoting better dispersion of active components and inhibiting particle aggregation during calcination. Furthermore, the copper dispersion (D_Cu) data in Table 1 reveals a clear trend. The Cu3Al5Ox and Mn1Cu2Al5Ox samples exhibit relatively low Cu dispersion values of 11.71% and 13.47%, respectively. In stark contrast, the Mn2Cu2Al4Ox catalyst demonstrates the highest Cu dispersion of 24.6%, unequivocally confirming that appropriate Mn doping is highly effective in improving Cu dispersion and increasing the number of exposed active sites in Cu-based catalysts [33].

2.5. XPS Analysis

To investigate the surface composition and chemical states of spinel catalysts, XPS analysis was performed on the Cu 2p, Mn 2p, and O 1s regions for Mn2Cu2Al4Ox; samples before and after reduction [34], with results presented in Figure 5. As shown in Figure 5a, the pre-reduction Cu 2p3/2 spectrum exhibits characteristic peaks of Cu2+, attributable to both non-spinel and spinel phases, with distinct satellite peaks observed in the 943–945 eV range [35]. In comparison with the Cu 2p spectrum of CuAl2O4 reported in the literature, Mn doping induces a negative binding energy shift in the Mn2Cu2Al4Ox catalyst [36]. Following H2 reduction, the main Cu 2p peak of Mn2Cu2Al4Ox shifts to lower binding energies (~932–933 eV), accompanied by a significant decrease in satellite peak intensity. These spectral changes indicate partial reduction of Cu2+ to Cu0/Cu+ species [37].
As shown in Figure 5b, the Mn 2p3/2 spectrum can be deconvoluted into multiple components, indicating the coexistence of multiple Mn oxidation states (Mn2+, Mn3+, Mn4+) [38]. Notably, the as-prepared Mn2Cu2Al4Ox catalyst exhibits a high proportion of Mn3+ (binding energy ~641 eV) [39]. After reduction, the Mn4+/Mn3+ ratio decreases while the Mn2+ content increases. In methanol steam reforming for hydrogen production, the Mn4+/Mn3+ redox couple facilitates CO oxidation to CO2, thereby reducing CO selectivity [40].
As depicted in Figure 5c, the O 1s spectrum can be deconvoluted into two distinct components: lattice oxygen (~529.6 eV) and surface-adsorbed oxygen (~531.3 eV) [41]. Notably, the Mn 2p3/2 peak also shifts to lower binding energies after reduction. Concurrently, the lattice oxygen peak intensity decreases while the surface-adsorbed oxygen peak intensity increases with a high-energy shift in its position. These spectral changes indicate the formation of oxygen vacancies and an increase in surface hydroxyl groups during reduction. The enhanced presence of chemically adsorbed oxygen species facilitates CO oxidation to CO2, thereby reducing CO selectivity and improving the catalyst’s CO2 selectivity [42]. It is noteworthy that surface-adsorbed oxygen is considered one of the key factors influencing the reaction rate of methanol steam reforming (MSR), as it can regulate the redox reaction rate and facilitate the dissociative adsorption of methanol [43].

2.6. TEM Characterization

TEM image of fresh and reduced Mn2Cu2Al4Ox sample was shown in Figure 6. As shown in Figure 6A, the fresh Mn2Cu2Al4Ox catalyst is composed of highly dispersed nanoparticles with an average particle size of 15.53 nm. After reduction with hydrogen, the active metal particles undergo significant sintering, resulting in a notable increase in the average particle size to 22.58 nm.
HRTEM and STEM of fresh and reduced Mn2Cu2Al4Ox sample was shown in Figure 7. As displayed in Figure 7A, the fresh Mn2Cu2Al4Ox catalyst exhibits two distinct lattice spacings of 0.14 nm and 0.16 nm. The 0.14 nm spacing can be assigned to the (440) crystal plane of the spinel-type structure, whereas the 0.16 nm spacing likely corresponds to the (111) plane of CuO, suggesting the formation of a composite oxide with well-intermixed phases in the as-prepared catalyst [44]. After H2 reduction, the observed lattice spacings become 0.14 nm and 0.18 nm. The persistence of the 0.14 nm spacing indicates that the spinel framework remains structurally stable under reductive conditions. In contrast, the expansion from 0.16 nm to 0.18 nm suggests lattice relaxation, which can be attributed to the creation of oxygen vacancies and the reduction of metal cations, most likely arising from the transformations of Cu2+ to Cu+/Cu0 and Mn4+ to Mn3+/Mn2+ [45]. Similar lattice expansions upon reduction have been reported for transition metal oxides, where removal of lattice oxygen and partial reduction of metal ions lead to weakened metal–oxygen bonds, increased inter-planar distances, and defect-mediated structural relaxation. The formation of oxygen vacancies not only modifies the local coordination environment but also enhances the mobility of surface oxygen species, as corroborated by the O 1s XPS results (Figure 5c), thereby potentially contributing to the improved CO2 selectivity observed during methanol steam reforming [19].
Elemental mapping of the fresh Mn2Cu2Al4Ox sample shows that Mn (yellow), Cu (cyan), Al (orange), and O (blue) are homogeneously distributed, indicating that the co-precipitation method effectively achieved uniform mixing of the constituent elements without noticeable elemental segregation [19]. This uniform elemental distribution is consistent with the particle morphology observed in the TEM images, suggesting the formation of a homogeneous multi-component composite oxide. After H2 reduction, the overall elemental distribution remains relatively uniform; however, slight aggregation of Cu is observed, implying element migration during the reduction process. This phenomenon is likely due to the reduction of CuO to metallic Cu. The Al distribution remains highly stable, indicating that the spinel framework is preserved under the reductive conditions [46].
The elemental content of Mn2Cu2Al4Ox samples before and after reduction was shown in Table 2. As shown in Table 2, after H2 reduction, the Cu content in the Mn2Cu2Al4Ox sample increases by approximately 31%, while the Mn content increases by about 40%, accompanied by a decrease of around 15% in oxygen content. The Al content remains nearly unchanged, indicating that the spinel framework is highly stable during the reduction process.
TEM–mapping analysis further reveals that the Mn2Cu2Al4Ox catalyst prepared by the co-precipitation method undergoes notable structural and surface compositional changes upon reduction. The observed lattice expansion—in particular, the increase in the interplanar spacing from 0.16 nm to 0.18 nm—suggests the formation of oxygen vacancies. Meanwhile, elemental distribution analysis shows that, despite the structural changes, the overall distribution of elements remains relatively homogeneous after reduction. However, the surface concentration of metallic species increases, which is favorable for the catalytic reaction.

2.7. Assessment of Catalytic Activity and Selectivity

Figure 8 compares the methanol conversion (solid lines) and CO selectivity (dashed lines) of the catalysts with different Mn/Cu/Al molar ratios in the temperature range of 240–300 °C. For all samples, methanol conversion increases markedly with temperature, reflecting the endothermic nature of the methanol steam reforming (MSR) reaction. The variation in the Mn/Cu/Al ratio has a pronounced influence on catalytic activity. Notably, the Mn2Cu2Al4Ox catalyst consistently demonstrates the highest methanol conversion across the entire temperature range, reaching over 95% at 300 °C, which correlates well with its high specific surface area and optimal Cu dispersion as shown in Table 1.
Regarding CO selectivity, the differences among the catalysts are relatively small and largely within the margin of experimental error at lower temperatures (240 and 260 °C). However, as the temperature increases to 280 °C and 300 °C, where side reactions such as methanol decomposition and reverse water-gas shift (RWGS) become more significant [47], clear and statistically meaningful trends emerge.
Specifically, at 300 °C, catalysts with a lower Mn/Cu ratio, such as Mn1Cu3Al4Ox, exhibit a significantly higher CO selectivity of approximately 9–10%. In contrast, the catalysts with a balanced or high Mn/Cu ratio, particularly our optimal Mn2Cu2Al4Ox, effectively maintain a lower CO selectivity (around 6%). This trend strongly suggests that a sufficient amount of manganese plays a crucial role in suppressing CO formation under more demanding reaction conditions. This is consistent with our XPS analysis, which indicates that Mn doping enriches Mn3+ species and facilitates oxygen vacancy formation, thereby promoting the water-gas shift (WGS) reaction that consumes CO [48]. While this effect is less discernible at lower temperatures, the data at higher temperatures clearly validates that catalysts with an insufficient Mn content are more prone to undesirable CO production. Therefore, catalysts with balanced Mn and Cu contents, like Mn2Cu2Al4Ox, offer the best compromise between achieving high conversion and maintaining low CO selectivity.
Figure 9 illustrates the temporal evolution of methanol conversion and CO selectivity for the optimal Mn2Cu2Al4Ox catalyst during a 24 h stability test at 240 °C. The catalyst shows an initial steady-state methanol conversion of approximately 90%. It is noted that this conversion is higher than the ~76% reported in the activity test (Figure 8) at the same temperature. This is attributed to a catalyst ‘break-in’ period, where the catalyst reaches a more stable and highly active state after being held on stream at a constant temperature. Over the 24 h period, the conversion only experiences a slight decline, demonstrating the catalyst’s high structural and catalytic stability under prolonged hydrothermal conditions. Simultaneously, the CO selectivity remains consistently low (~1.1%) with only a marginal change (<0.3% absolute), confirming its excellent resistance to side reactions like methanol decomposition.
Varying the Mn:Cu:Al molar ratio tunes the balance between redox properties and metallic copper dispersion, thereby controlling both methanol conversion efficiency and CO selectivity. Mn-rich spinel-type catalysts (for instance Mn2Cu2Al4Ox) offer higher conversions with no apparent penalty in CO selectivity, making them promising candidates for low-CO hydrogen production via MSR.
To further contextualize the performance of our optimized catalyst, Table 3 provides a comparison of the catalytic activity of the Mn2Cu2Al4Ox catalyst with other recently reported Cu-based catalysts for methanol steam reforming.
As shown in Table 3, our optimal Mn2Cu2Al4Ox catalyst demonstrates outstanding performance. Compared to the simple Cu/Al2O3 catalyst reported by Hammoud et al. [11], our catalyst achieves significantly higher methanol conversion at similar temperatures. While the commercial Cu/ZnO/Al2O3 catalyst shows high initial activity, it is known to suffer from poor thermal stability, an issue our spinel-based catalyst effectively overcomes, as evidenced by the 24 h stability test.
More importantly, when benchmarked against other Mn-promoted systems, the Mn2Cu2Al4Ox catalyst remains highly competitive. For instance, compared to a reported Mn-doped Cu/Al2O3 catalyst, our system achieves higher conversion and lower CO selectivity at 300 °C. Although catalysts derived from LDH precursors show excellent results [23], our spinel-based catalyst prepared by a straightforward co-precipitation method offers a comparable high-performance profile. The superior activity and significantly suppressed CO formation compared to the pure CuAl2O4 spinel [14] unequivocally highlight the critical synergistic effect of incorporating an optimal amount of manganese.
In summary, this comparison validates that tailoring the Mn/Cu ratio to form a Mn2Cu2Al4Ox spinel structure is a highly effective strategy, yielding a catalyst that not only surpasses many existing systems in activity but also provides the crucial durability and low-CO selectivity required for practical hydrogen production applications.

3. Experimental Section

3.1. Materials

The chemical reagents and experimental water used in this experiment are as follows: Copper nitrate trihydrate (Cu(NO3)2·3H2O), manganese nitrate tetrahydrate (Mn(NO3)2·4H2O), aluminum nitrate nonahydrate (Al(NO3)3·9H2O), sodium hydroxide (NaOH), and Graphite: All are of Analytical Reagent (AR) grade and purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). Methanol (CH3OH): Of Guaranteed Reagent (GR) grade, purchased from Shanghai Macklin Biochemical Technology Co. (Shanghai, China), Ltd. Ammonia solution: With a mass fraction of 5%, purchased from Sinopharm Chemical Reagent Co., Ltd. Sodium hydroxide (NaOH): Of Analytical Reagent (AR) grade, provided by Aladdin Reagents (Shanghai) Co., Ltd. (Shanghai, China). Graphite (C60): Provided by Aladdin Reagents Co., Ltd. (Shanghai, China).

3.2. Instruments

The catalytic performance in the methanol steam reforming (MSR) reaction was evaluated using a gas–solid fixed-bed reactor (Tianjin Honghua Technology Co., Ltd., Tianjin, China). The crystalline phase structures were characterized by an X-ray powder diffractometer (Bruker Corporation, Karlsruhe, Germany). The product compositions were analyzed using a gas chromatograph (Shimadzu Corporation, Kyoto, Japan). Surface properties were investigated using a dynamic chemical adsorption analyzer (Beijing JWGB Sci. & Tech. Co., Ltd., Beijing, China), while morphological observations were conducted via a scanning electron microscope (Carl Zeiss AG, Oberkochen, Germany). The elemental composition was determined by inductively coupled plasma optical emission spectrometry (ICP-OES, Thermo Fisher Scientific Inc., Waltham, MA, USA). The calcination process was carried out in a muffle furnace (Qixin, Shanghai, China). Textural properties were measured using a BSDPM2 surface area and micropore analyzer (Beishide Instrument Technology (Beijing) Co., Ltd., Beijing, China). X-ray photoelectron spectroscopy (XPS) analyses were performed using a Nexsa system (Thermo Fisher Scientific Inc., Waltham, MA, USA), and transmission electron microscopy (TEM) tests were conducted with a Talos F200X G2 instrument (Thermo Fisher Scientific Inc., Waltham, MA, USA).

3.3. Catalyst Preparation

Mn–Cu/Al2O4 catalysts were synthesized via a co-precipitation method using copper(II) nitrate trihydrate, manganese nitrate, and aluminum nitrate as the respective precursors of Cu, Mn, and Al. Catalysts with varying Mn:Cu:Al molar ratios (0:2:5, 1:2:5, 2:2:4, 1:3:4, 2:1:5, and 3:1:4) were prepared. The calculated amounts of metal salts were dissolved in 80 mL of deionized water, followed by the addition of five drops of 0.1 M nitric acid to obtain solution A. The mixture was maintained at 30 °C in an oil bath under constant stirring for 30 min. A 1 wt% NH3·H2O solution was then introduced dropwise at a rate of 1 mL/min under magnetic stirring (200 rpm). After complete addition, the pH of the suspension was adjusted to 9.0, and the slurry was aged for 2 h with continuous stirring. The resulting precipitate was filtered, washed three times with deionized water and ethanol, and subsequently dried in a vacuum oven at 80 °C (0.08 MPa) for 12 h. The dried solid was ground into fine precursor powders, which were finally calcined in air at 400 °C for 4 h in a tubular furnace (heating rate: 5 °C/min). The obtained catalysts with different molar ratios were denoted as MnCuAlOx. The sample preparation process is shown in Figure 10.

3.4. Characterization of Catalysts

3.4.1. X-Ray Diffraction (XRD)

The crystalline phases of the catalysts were analyzed using a Bruker D8 Advance X-ray diffractometer (Bruker Corporation, Karlsruhe, Germany) equipped with Cu Kα radiation (λ = 0.154 nm). The instrument was operated at a tube voltage of 40 kV and a tube current of 40 mA. Diffraction patterns were collected over a 2θ range of 10–90° with a step size of 0.02° and a scan rate of 2°/min. For sample preparation, catalyst powder was uniformly packed into a standard sample holder and flattened with a glass slide to ensure a planar surface coplanar with the holder edge. Phase identification and crystallite size analysis were performed using MDI Jade software (version 6.5) by matching diffraction peaks to reference patterns in the ICDD PDF-4+ database.

3.4.2. Scanning Electron Microscopy (SEM)

Surface morphology and elemental composition were examined using a Zeiss Evo 18 scanning electron microscope (Carl Zeiss AG, Oberkochen, Germany) equipped with an energy-dispersive X-ray spectroscopy (EDS) detector. Samples were prepared by affixing double-sided conductive carbon tape to an aluminum stub. Catalyst powder was dispersed onto the tape using a capillary tube, and loose particles were removed with compressed air. To enhance conductivity, samples were sputter-coated with a 5 nm gold layer using a Quorum Q150R ES sputter coater (Quorum Technologies Ltd., Ramsgate, Kent, UK). Imaging was performed under high-vacuum conditions at accelerating voltages of 0.2–30 kV, with a working distance of 8–10 mm. Secondary electron images were acquired at various magnifications (500–50,000×), while EDS analysis was conducted at 20 kV to quantify elemental distribution (Mn, Cu, Al, O) with an acquisition time of 60 s per spectrum.

3.4.3. H2-TPR Characterization

The dispersion and specific surface area of Cu in the catalysts were determined using a JW-DTP1009 chemisorption analyzer (Beijing JWGB Sci. & Tech. Co., Ltd., Beijing, China). The reduction of the catalyst samples with hydrogen, followed by dynamic adsorption, was employed to reveal the distribution characteristics and variation trends of copper under different reaction conditions.
Approximately 30 mg of the prepared catalyst sample was placed into a U-shaped quartz tube packed with asbestos, with an additional asbestos layer used for fixation. The tube was then mounted on the chemisorption analyzer. The reducibility of the catalysts was investigated under a 10% H2/Ar mixed atmosphere in the temperature range of 30–350 °C, referred to as TPR1. After reduction, the sample was titrated with N2O, followed by a second reduction in a 10% H2/Ar mixture at 50–300 °C, designated as TPR2. The copper dispersion (D_Cu) and S-Cu data of Mn/Cu/Al catalyst samples were calculated by Equations (1) and (2) as following:
D _ C u = 2 n 2 / n 1 × 100 %
S _ C u = 2 n 2 N A / n 1 M C u N C u
where n1 is the amount of H2 consumed during TPR1 (mol), n2 is the amount of H2 consumed during the reaction in Equation (2) (mol), NA is Avogadro’s constant (6.02 × 1023 mol−1), MCu is the atomic weight of Cu (63.55 g/mol), ρCu is the density of Cu (8.92 g/cm3), and NCu is the number of copper atoms per unit surface area (1.4 × 1019 atoms/m2).

3.4.4. Determination of Elemental Content by ICP

The Perkin-Elmer Optima 8000 Inductively Coupled Plasma Optical Emission Spectrometer (ICP-OES, Thermo Fisher Scientific Inc., Waltham, MA, USA) was used to analyze the content of various metal elements in the catalyst. Prior to sample loading and analysis, the samples were digested using nitric acid and hydrofluoric acid.

3.4.5. N2 Physisorption Analysis

Surface area and porosity measurements were conducted on a BSD-PM2 automated analyzer (Beijing JWGB Sci. & Tech. Co., Ltd., Beijing, China) equipped with a high-precision pressure transducer (±0.15% reading accuracy). Specific surface areas were calculated via multi-point BET analysis (P/P0 = 0.05–0.30), and pore volume distributions were determined using the BJH model applied to the desorption isotherms.

3.4.6. XPS and TEM Analysis

X-ray photoelectron spectroscopy (XPS) was performed on a Nexsa system (Thermo Fisher Scientific Inc., Waltham, MA, USA), and transmission electron microscopy (TEM) was carried out using a Talos F200X G2 microscope (Thermo Fisher Scientific Inc., Waltham, MA, USA).

3.5. Catalyst Evaluation

The catalytic performance for methanol steam reforming was evaluated in a conventional gas–solid fixed-bed reactor [48]. Methanol conversion and carbon monoxide selectivity were used as the primary indicators.
A total of 2.0 g of catalyst mixed with graphite at a 93:7 weight ratio (catalyst:graphite) to facilitate pelletization was loaded into the reactor for testing. A mixed gas of H2 (15 mL/min) and N2 (75 mL/min) was introduced, and the sample was pretreated at 30 °C for 30 min. The temperature was then increased to 260 °C at a heating rate of 3 °C/min, and the catalyst was reduced for 3 h. After reduction, the reactor was cooled directly to 240 °C.
Subsequently, a water–methanol mixture (mass ratio 1.2:1) was fed into the reactor at a flow rate of 0.2 mL/min using a micro-liquid pump. The catalyst bed temperature was controlled in the range of 240–300 °C, with a heating rate of 1 °C/min. Catalyst stability tests were carried out at 240 °C for 24 h under identical reaction conditions.
The gas-phase products were analyzed using a Shimadzu GC-2014PLUS gas chromatograph (Shimadzu Corporation, Kyoto, Japan). Methanol conversion was determined by collecting and analyzing the condensed liquid. The methanol concentration in the condensate was calculated by the area normalization method, from which the methanol conversion was obtained according to the following Equation:
α Methanol   conversion = 1 V effluent × ρ effluent × ω m e t n a n o l   c o n t e n t   o f   effluent 0.2 × t × ρ m e t n a n o l   c o n t e n t   o f   feed × 0 . 4
S CO = F C O F C O + F C O 2 + F C H 4 × 100 %
where Fi represents the flow rate of CO, CO2, CH4. The molar flow rates were obtained from the concentrations measured by GC and the total volumetric flow rate of the dry gas effluent.
The carbon balance was calculated for each data point by comparing the total moles of carbon in the outlet gas stream (CO, CO2, and unreacted CH3OH in the condensate) to the moles of carbon fed into the reactor. For all reported catalytic tests, the carbon balance was maintained in the range of 97–102%, confirming the reliability of the collected data and indicating that the formation of other carbonaceous byproducts, such as coke, was negligible under these reaction conditions.

4. Conclusions

This study demonstrates that precise tuning of the Mn/Cu ratio in spinel-type Mn–Cu/Al2Ox catalysts is an effective strategy to optimize structural stability, active phase dispersion, and CO suppression in methanol steam reforming. Higher Mn content promotes MnAl2O4 spinel formation, induces lattice distortion, and inhibits Cu crystallite growth, while balanced Mn–Cu compositions yield dual-spinels that combine Mn–O redox cycles with well-dispersed metallic Cu sites. H2-TPR, BET, and XPS analyses confirm that Mn doping enhances Cu dispersion (up to 24.6%), enriches Mn3+ species, and generates abundant oxygen vacancies, which collectively facilitate CH3OH activation and water–gas shift coupling while suppressing methanol decomposition routes to CO. Catalytic tests reveal that Mn2Cu2Al4Ox achieves near-complete methanol conversion at 300 °C with CO selectivity below 10% and maintains >88% conversion with negligible CO increase during 24 h hydrothermal stability trials. These findings establish Mn2Cu2Al4Ox as a promising, durable, and low-CO catalyst for on-site hydrogen production in fuel-cell-powered energy systems.

Author Contributions

Conceptualization, Q.Z. and Y.Z.; methodology, S.Q.; software, Q.Z. and S.Q.; validation, Q.Z., S.Q., Y.Z. and Y.H.; formal analysis, Q.Z.; resources, S.Q.; data curation, Q.Z. and S.Q.; writing—original draft preparation, Q.Z.; writing—review and editing, Y.H.; visualization, Y.Z. and Y.H.; supervision, S.Q. and Y.H.; project administration, Q.Z. and S.Q.; funding acquisition, Q.Z. and S.Q. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by Chongzuo Science and Technology Bureau, Project No. Chongke2024088520, Chognzuo Youth Science and Technology Talent Training Project, Project No. Chongke 2022QN013010, High level Talent Introduction Project of Guangxi Minzu Normal University, Project No. 2024SBNGCC05 and 2021BS003, Chongzuo “Sharp Edge” Science and Technology Special Project (No. Chongke 20241205). The funding agency played no role in the study design, data collection, analysis, interpretation of data, writing of the report, or the decision to submit the article for publication. The authors are solely responsible for the content and the writing of the manuscript.

Data Availability Statement

The data presented in this study are available on request from the corresponding authors.

Acknowledgments

The authors are grateful to the Fundamental Research Program (Free Exploration) of Shanxi Province of China (20210302124338) and the Postdoctoral program of Administrative Committee of Taiyuan Economic Development Zone.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Tiba, S.; Omri, A. Literature survey on the relationships between energy, environment and economic growth. Renew. Sustain. Energy Rev. 2017, 69, 1129–1146. [Google Scholar] [CrossRef]
  2. Ramasubramanian, B.; Sundarrajan, S.; Rao, R.P.; Reddy, M.; Chellappan, V.; Ramakrishna, S. Novel low-carbon energy solutions for powering emerging wearables, smart textiles, and medical devices. Energy Environ. Sci. 2022, 15, 4928–4981. [Google Scholar] [CrossRef]
  3. Ranjekar, A.M.; Yadav, G.D. Steam reforming of methanol for hydrogen production: A critical analysis of catalysis, processes, and scope. Ind. Eng. Chem. Res. 2021, 60, 89–113. [Google Scholar] [CrossRef]
  4. Rostami, M.; Farajollahi, A.H.; Amirkhani, R.; Farshchi, M.E. A review study on methanol steam reforming catalysts: Evaluation of the catalytic performance, characterizations, and operational parameters. AIP Adv. 2023, 13, 030701. [Google Scholar] [CrossRef]
  5. Jiang, L.; Yuan, S.; Ma, J.; Deng, S.; Fang, X.; Xu, X.; Meng, H.; Wang, X. Enhancing the Reactivity of Cu/Al2O3 for Methanol Steam Reforming through adding CrOx: Unraveling Reaction Pathways and the Mechanism for Improvement. ACS Catal. 2025, 15, 7138–7152. [Google Scholar] [CrossRef]
  6. 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]
  7. Pang, S.; Dou, X.; Zhao, W.; Bai, S.; Wan, B.; Wang, T.; Yang, J.H. A Review on the Design Strategies of Copper-Based Catalysts for Enhanced Activity and Stability in Methanol Reforming to Hydrogen. Nanomaterials 2025, 15, 1118. [Google Scholar] [CrossRef]
  8. Kamsuwan, T.; Guntida, A.; Praserthdam, P.; Jongsomjit, B. Differences in Deterioration Behaviors of Cu/ZnO/Al2O3 Catalysts with Different Cu Contents toward Hydrogenation of CO and CO2. ACS Omega 2022, 7, 25783–25797. [Google Scholar] [CrossRef]
  9. Zamel, N.; Groos, U. Hydrogen Energy Recovery–H2-Based Fuel Cells. Adv. Energy Storage Latest Dev. RD Mark. 2022, 559–588. [Google Scholar]
  10. Jambur, S.R. Techno-Economic Assessment of High-Temperature H2O/CO2: Co-Electrolysis in Solid Oxide Electrolysers for Syngas Production; Technical University of Denmark/DTU: Kongens Lyngby, Denmark, 2022. [Google Scholar]
  11. Hammoud, D.; Gennequin, C.; Aboukaïs, A.; Aad, E.A. Steam reforming of methanol over x% Cu/Zn–Al 400 500 based catalysts for production of hydrogen: Preparation by adopting memory effect of hydrotalcite and behavior evaluation. Int. J. Hydrogen Energy 2015, 40, 1283–1297. [Google Scholar] [CrossRef]
  12. Shan, J.; Zheng, Y.; Shi, B.; Davey, K.; Qiao, S.-Z. Regulating Electrocatalysts via Surface and Interface Engineering for Acidic Water Electrooxidation. ACS Energy Lett. 2019, 4, 2719–2730. [Google Scholar] [CrossRef]
  13. Liao, M.; Duan, Q.; Xiang, R.; Tan, X.; Dai, Z.; Qin, H.; Xiao, H. Compounds: Rare earth-doped CuAl2O4 catalysts with superior catalytic performance for methanol steam reforming. J. Alloys Compd. 2025, 1024, 180229. [Google Scholar] [CrossRef]
  14. Frank, B. Steam Reforming of Methanol over Cu/ZnO/Al2O3 Modified with Hydrotalcites. Catal. Commun. 2007, 8, 1684–1690. [Google Scholar] [CrossRef]
  15. Tan, Z.; Fan, G.; Zheng, L.; Li, F. Promoted Surface-Interface Catalysis over Mn–Cr Incorporated Cu-Based Catalysts for Efficient Hydrogen Production from Methanol Decomposition. ACS Catal. 2024, 14, 11218–11230. [Google Scholar] [CrossRef]
  16. Ye, R.; Xiao, S.; Lai, Q.; Wang, D.; Huang, Y.; Feng, G.; Zhang, R.; Wang, T. Advances in enhancing the stability of Cu-based catalysts for methanol reforming. Catalysts 2022, 12, 747. [Google Scholar] [CrossRef]
  17. Zhang, M.; Liu, D.; Wang, Y.; Zhao, L.; Xu, G.; Yu, Y.; He, H. Recent advances in methanol steam reforming catalysts for hydrogen production. Catalysts 2025, 15, 36. [Google Scholar] [CrossRef]
  18. Mao, Q.; Gao, Z.; Liu, X.; Guo, Y.; Wang, Y.; Ma, D. The Cu–Al2O3 interface: An unignorable active site for methanol steam reforming hydrogen production. Catal. Sci. Technol. 2024, 14, 3448–3458. [Google Scholar] [CrossRef]
  19. Dasireddy, V.D.; Likozar, B. Cu–Mn–O nano-particle/nano-sheet spinel-type materials as catalysts in methanol steam reforming (MSR) and preferential oxidation (PROX) reaction for purified hydrogen production. Renew. Energy 2022, 182, 713–724. [Google Scholar] [CrossRef]
  20. Kang, N.; Zhao, Z.; Yu, X.; Lin, J.; Lu, S. Tuning surface properties of CuCeOx catalysts with manganese and cobalt co-doping for CO oxidation. J. Rare Earths 2025, 43, 1643–1651. [Google Scholar] [CrossRef]
  21. Rani, B.J.; Sivanantham, A.; Cho, I.S. Nanostructured spinel manganates and their composites for electrochemical energy conversion and storage. Adv. Funct. Mater. 2023, 33, 2303002. [Google Scholar] [CrossRef]
  22. Liao, M.; Xiang, R.; Zhou, X.; Dai, Z.; Wang, L.; Qin, H.; Xiao, H. Enhancing effect of Mn2+ substitution in CuAl2O4 spinel for methanol steam reforming in a microreactor. Renew. Energy 2024, 230, 120815. [Google Scholar] [CrossRef]
  23. Meng, H.; Yang, Y.; Shen, T.; Yin, Z.; Wang, L.; Liu, W.; Yin, P.; Ren, Z.; Zheng, L.; Zhang, J.; et al. Designing Cu0−Cu+ dual sites for improved C−H bond fracture towards methanol steam reforming. Nat. Commun. 2023, 14, 7980. [Google Scholar] [CrossRef]
  24. Yan, Q.; Chen, S.; Qiu, L.; Gao, Y.; O’Hare, D.; Wang, Q. The synthesis of CuyMnzAl1−zOx mixed oxide as a low-temperature NH3-SCR catalyst with enhanced catalytic performance. Dalton Trans. 2018, 47, 2992–3004. [Google Scholar] [CrossRef]
  25. Herdem, M.S.; Sinaki, M.Y.; Farhad, S.; Hamdullahpur, F. An overview of the methanol reforming process: Comparison of fuels, catalysts, reformers, and systems. Int. J. Energy Res. 2019, 43, 5076–5105. [Google Scholar] [CrossRef]
  26. Li, Y.; Zhang, Y.; Qian, K.; Huang, W. Metal–Support Interactions in Metal/Oxide Catalysts and Oxide–Metal Interactions in Oxide/Metal Inverse Catalysts. ACS Catal. 2022, 12, 1268–1287. [Google Scholar] [CrossRef]
  27. Pang, J.; Li, Q.; Su, G.; Sun, B.; Song, M.; Hua, Y.; Zhang, W.; Meng, J.; Shi, B. Tailoring Dual High-Valence Cu–O–Mn Active Sites to Enhance VOC Catalytic Oxidation. Environ. Sci. Technol. 2025, 59, 9812–9826. [Google Scholar] [CrossRef] [PubMed]
  28. Ding, J.; Zhao, L.; Yang, Y.; Liu, J. Unraveling the role of a Cu dopant in formaldehyde catalytic oxidation over a La0.8Sr0.2Mn1−xCuxO3 perovskite: An experimental and theoretical study. J. Mater. Chem. A 2025, 13, 7371–7380. [Google Scholar] [CrossRef]
  29. Lu, T.; Su, F.; Zhao, Q.; Li, J.; Zhang, C.; Zhang, R.; Liu, P. Catalytic oxidation of volatile organic compounds over manganese-based oxide catalysts: Performance, deactivation and future opportunities. Sep. Purif. Technol. 2022, 296, 121436. [Google Scholar] [CrossRef]
  30. Saw, S.; Datta, S.; Chavan, P.; Gupta, P.; Kumari, S.; Sahu, G.; Chauhan, V. Significance and influence of various promoters on Cu-based catalyst for synthesizing methanol from syngas: A critical review. J. Chem. Technol. Biotechnol. 2023, 98, 1083–1102. [Google Scholar] [CrossRef]
  31. Fajin, J.L.; Cordeiro, M.N.D. Insights into the mechanism of methanol steam reforming for hydrogen production over Ni–Cu-based catalysts. ACS Catal. 2021, 12, 512–526. [Google Scholar] [CrossRef]
  32. Lin, J.; Guo, Y.; Chen, X.; Li, C.; Lu, S.; Liew, K.M. CO Oxidation over Nanostructured Ceria Supported Bimetallic Cu–Mn Oxides Catalysts: Effect of Cu/Mn Ratio and Calcination Temperature. Catal. Lett. 2018, 148, 181–193. [Google Scholar] [CrossRef]
  33. Kang, R.; Wei, X.; Bin, F.; Wang, Z.; Hao, Q.; Dou, B. Reaction mechanism and kinetics of CO oxidation over a CuO/Ce0.75Zr0.25O2-δ catalyst. Appl. Catal. A Gen. 2018, 565, 46–58. [Google Scholar] [CrossRef]
  34. Zedan, A.F.; Polychronopoulou, K.; Asif, A.; AlQaradawi, S.Y.; AlJaber, A.S. Cu-Ce-O catalyst revisited for exceptional activity at low temperature CO oxidation reaction. Surf. Coat. Technol. 2018, 354, 313–323. [Google Scholar] [CrossRef]
  35. Abed, S.H.; Reshak, A.H. Optimization of CuAl2O4 Nanoparticles for Enhanced Photocatalytic Degradation of Organic Dyes Under UV and Visible Light. J. Fluoresc. 2025, 35, 6797–6812. [Google Scholar] [CrossRef]
  36. Torres-Ochoa, J.A.; Cabrera-German, D.; Cortazar-Martinez, O.; Bravo-Sanchez, M.; Gomez-Sosa, G.; Herrera-Gomez, A. Peak-fitting of Cu 2p photoemission spectra in Cu0, Cu1+, and Cu2+ oxides: A method for discriminating Cu0 from Cu1+. Appl. Surf. Sci. 2023, 622, 156960. [Google Scholar] [CrossRef]
  37. Napruszewska, B.; Walczyk, A.; Duraczyńska, D.; Kryściak-Czerwenka, J.; Michalik, A.; Karcz, R.; Śliwa, M.; Serwicka, E. The Synthesis of Cu–Mn–Al Mixed-Oxide Combustion Catalysts by Co-Precipitation in the Presence of Starch: A Comparison of NaOH with Organic Precipitants. Catalysts 2022, 12, 1159. [Google Scholar] [CrossRef]
  38. Liu, P.; Wei, G.; Liang, X.; Chen, D.; He, H.; Chen, T.; Xi, Y.; Chen, H.; Han, D. Synergetic effect of Cu and Mn oxides supported on palygorskite for the catalytic oxidation of formaldehyde: Dispersion, microstructure, and catalytic performance. Appl. Clay Sci. 2018, 161, 265–273. [Google Scholar] [CrossRef]
  39. Liu, T.; Yao, Y.; Wei, L.; Shi, Z.; Han, L.; Yuan, H.; Li, B.; Dong, L.; Wang, F.; Sun, C.-Z. Preparation and Evaluation of Cu-Mn-Oxide as High Efficiency Catalyst for CO Oxidation and NO Reduction by CO. J. Phys. Chem. C 2017, 121, 12757–12770. [Google Scholar] [CrossRef]
  40. Xu, X.; Liu, X.; Ma, L.; Liang, N.; Yang, S.; Liu, H.; Sun, J.; Huang, F.; Sun, C.; Dong, L. Construction of Surface Synergetic Oxygen Vacancies on CuMn2O4 Spinel for Enhancing NO Reduction with CO. ACS Catal. 2024, 14, 3028–3040. [Google Scholar] [CrossRef]
  41. Venkataswamy, P.; Rao, K.N.; Jampaiah, D.; Reddy, B.M. Nanostructured manganese doped ceria solid solutions for CO oxidation at lower temperatures. Appl. Catal. B Environ. 2015, 162, 122–132. [Google Scholar] [CrossRef]
  42. Yang, S.; Zhou, F.; Liu, Y.; Zhang, L.; Chen, Y.; Wang, H.; Tian, Y.; Zhang, C.; Liu, D. Morphology effect of ceria on the performance of CuO/CeO2 catalysts for hydrogen production by methanol steam reforming. Int. J. Hydrogen Energy 2019, 44, 7252–7261. [Google Scholar] [CrossRef]
  43. Gawande, M.B.; Goswami, A.; Felpin, F.-X.; Asefa, T.; Huang, X.; Silva, R.; Zou, X.; Zboril, R.; Varma, R.S. Cu and Cu-Based Nanoparticles: Synthesis and Applications in Catalysis. Chem. Rev. 2016, 116, 3722–3811. [Google Scholar] [CrossRef] [PubMed]
  44. Lin, X.; Li, S.; He, H.; Wu, Z.; Wu, J.; Chen, L.; Ye, D.; Fu, M. Evolution of oxygen vacancies in MnOx-CeO2 mixed oxides for soot oxidation. Appl. Catal. B Environ. 2018, 223, 91–102. [Google Scholar] [CrossRef]
  45. Castaño, M.H.; Molina, R.; Moreno, S. Cooperative effect of the Co–Mn mixed oxides for the catalytic oxidation of VOCs: Influence of the synthesis method. Appl. Catal. A Gen. 2015, 492, 48–59. [Google Scholar] [CrossRef]
  46. Daza, Y.A.; Kuhn, J.N. CO2 conversion by reverse water gas shift catalysis: Comparison of catalysts, mechanisms and their consequences for CO2 conversion to liquid fuels. RSC Adv. 2016, 6, 49675–49691. [Google Scholar] [CrossRef]
  47. Nishimoto, M.; Tokura, R.; Nguyen, M.T.; Yonezawa, T. Copper materials for low temperature sintering. Mater. Trans. 2022, 63, 663–675. [Google Scholar] [CrossRef]
  48. Sanches, S.; Flores, J.H.; De Avillez, R.; Da Silva, M.P. Influence of preparation methods and Zr and Y promoters on Cu/ZnO catalysts used for methanol steam reforming. Int. J. Hydrogen Energy 2012, 37, 6572–6579. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of the Mn/Cu/Al catalysts with different mole ratio.
Figure 1. XRD patterns of the Mn/Cu/Al catalysts with different mole ratio.
Catalysts 15 01091 g001
Figure 2. SEM scan of MnCuAl catalyst.
Figure 2. SEM scan of MnCuAl catalyst.
Catalysts 15 01091 g002
Figure 3. TPR1 (left) and TPR2 (right) spectra of different MnCuAl catalysts.
Figure 3. TPR1 (left) and TPR2 (right) spectra of different MnCuAl catalysts.
Catalysts 15 01091 g003
Figure 4. N2 physical adsorption/desorption isotherms of different catalysts.
Figure 4. N2 physical adsorption/desorption isotherms of different catalysts.
Catalysts 15 01091 g004
Figure 5. Cu 2p XPS spectra (a), Mn 2p XPS spectra (b), and O 1s XPS spectra (c) of the Mn2Cu2Al4Ox catalyst before and after reduction.
Figure 5. Cu 2p XPS spectra (a), Mn 2p XPS spectra (b), and O 1s XPS spectra (c) of the Mn2Cu2Al4Ox catalyst before and after reduction.
Catalysts 15 01091 g005
Figure 6. TEM images of the spinels: images of Mn2Cu2Al4Ox: (A) Fresh catalyst; (B) The catalyst after reduction.
Figure 6. TEM images of the spinels: images of Mn2Cu2Al4Ox: (A) Fresh catalyst; (B) The catalyst after reduction.
Catalysts 15 01091 g006
Figure 7. HRTEM and STEM mapping images of Mn2Cu2Al4Ox: (A) Fresh catalyst; (B) The catalyst after reduction.
Figure 7. HRTEM and STEM mapping images of Mn2Cu2Al4Ox: (A) Fresh catalyst; (B) The catalyst after reduction.
Catalysts 15 01091 g007
Figure 8. Effect of Mn/Cu/Al molar ratio on methanol conversion and CO selectivity. (The solid lines represent methanol conversion, and the dashed lines denote CO selectivity. Error bars represent the standard deviation from three independent experiments.).
Figure 8. Effect of Mn/Cu/Al molar ratio on methanol conversion and CO selectivity. (The solid lines represent methanol conversion, and the dashed lines denote CO selectivity. Error bars represent the standard deviation from three independent experiments.).
Catalysts 15 01091 g008
Figure 9. Operational Stability of Mn2Cu2Al4Ox Catalyst.
Figure 9. Operational Stability of Mn2Cu2Al4Ox Catalyst.
Catalysts 15 01091 g009
Figure 10. Sample Preparation Flowchart.
Figure 10. Sample Preparation Flowchart.
Catalysts 15 01091 g010
Table 1. Structural and surface properties of the prepared Mn-Cu-Al catalysts.
Table 1. Structural and surface properties of the prepared Mn-Cu-Al catalysts.
SampleS_BET (m2/g)C_Cu (%)D_Cu (%)S_Cu (m2/g)TPR1 (°C)TPR2 (°C)H2 Consumption of TPR1 (mmol/g)H2 Consumption of TPR2 (mmol/g)
Cu3Al5Ox51.430.3411.7115.91287.17171.2212.370.99
Mn1Cu3Al4Ox55.870.3118.4021.65244.01171.6311.351.05
Mn1Cu2Al5Ox57.450.2613.4717.56251.51162.458.571.26
Mn2Cu2Al4Ox76.440.2524.6026.11257.86144.906.340.88
Mn2Cu1Al5Ox71.620.1627.4219.88215.23170.8314.451.97
Mn3Cu1Al4Ox65.320.1531.2521.65277.38151.6611.291.77
Table 2. Elemental Composition Changes in Mn2Cu2Al4Ox Samples Before and After Reduction.
Table 2. Elemental Composition Changes in Mn2Cu2Al4Ox Samples Before and After Reduction.
ElementComposition Before Reduction (%)Composition After Reduction (%)Numerical Change (%)
Mn10.2114.32+40.3
Cu11.8615.56+31.2
Al18.2019.37+6.4
O59.7350.75−15.0
Table 3. Comparison of catalytic performance for methanol steam reforming over the Mn2Cu2Al4Ox catalyst and other reported Cu-based catalysts.
Table 3. Comparison of catalytic performance for methanol steam reforming over the Mn2Cu2Al4Ox catalyst and other reported Cu-based catalysts.
Catalyst CompositionTemperature (°C)MeOH Conv. (%)CO Sel. (%)Stability/NotesReference
Mn2Cu2Al4Ox300>956Excellent stability (>88% conv. after 24h)This work
Mn2Cu2Al4Ox260873 This work
10% Cu/Al-400-50025051.9N/A [11]
Cu/ZnO/Al2O3250>90~3–5Prone to sintering and deactivation[5,6]
5%Mn-15%Cu/Al2O3300~9012.0 [18]
4.25Cu/Cu(Al)Ox270~92.3N/AGood performance, complex structure[23]
Cu/ZnO/Al2O3280~75~15Lower activity, higher CO selectivity[14]
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

Zhang, Q.; Qiu, S.; Zheng, Y.; Huang, Y. Effect of Mn/Cu Ratio on the Structure–Performance Relationship of Spinel-Type Mn–Cu/Al2Ox Catalysts for Methanol Steam Reforming. Catalysts 2025, 15, 1091. https://doi.org/10.3390/catal15111091

AMA Style

Zhang Q, Qiu S, Zheng Y, Huang Y. Effect of Mn/Cu Ratio on the Structure–Performance Relationship of Spinel-Type Mn–Cu/Al2Ox Catalysts for Methanol Steam Reforming. Catalysts. 2025; 15(11):1091. https://doi.org/10.3390/catal15111091

Chicago/Turabian Style

Zhang, Qiang, Shiming Qiu, Yanfei Zheng, and Yingying Huang. 2025. "Effect of Mn/Cu Ratio on the Structure–Performance Relationship of Spinel-Type Mn–Cu/Al2Ox Catalysts for Methanol Steam Reforming" Catalysts 15, no. 11: 1091. https://doi.org/10.3390/catal15111091

APA Style

Zhang, Q., Qiu, S., Zheng, Y., & Huang, Y. (2025). Effect of Mn/Cu Ratio on the Structure–Performance Relationship of Spinel-Type Mn–Cu/Al2Ox Catalysts for Methanol Steam Reforming. Catalysts, 15(11), 1091. https://doi.org/10.3390/catal15111091

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