Next Article in Journal
A Simplified Method for Predicting Bond–Slip Behaviour of Ribbed Bars and Threaded Rods Glued in Glulam along the Grain
Previous Article in Journal
Argon Plasma Treatment Effects on the Micro-Shear Bond Strength of Lithium Disilicate with Dental Resin Cements
Previous Article in Special Issue
Synergistic Effect of P Doping and Mo-Ni-Based Heterostructure Electrocatalyst for Overall Water Splitting
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Potassium Doping on the Structural and Catalytic Properties of Co/MnOx Catalyst in the Steam Reforming of Ethanol

by
Magdalena Greluk
1,*,
Marek Rotko
1,
Grzegorz Słowik
1,
Sylwia Turczyniak-Surdacka
2,
Gabriela Grzybek
3 and
Katarzyna Tyszczuk-Rotko
4
1
Department of Chemical Technology, Faculty of Chemistry, Maria Curie-Sklodowska University in Lublin, Maria Curie-Sklodowska Sq. 3, 20-031 Lublin, Poland
2
Biological and Chemical Research Centre, University of Warsaw, 101 Żwirki i Wigury Street, 20-089 Warsaw, Poland
3
Faculty of Chemistry, Jagiellonian University, Gronostajowa 2, 30-387 Krakow, Poland
4
Department of Analytical Chemistry, Faculty of Chemistry, Maria Curie-Sklodowska University in Lublin, Maria Curie-Sklodowska Sq. 3, 20-031 Lublin, Poland
*
Author to whom correspondence should be addressed.
Materials 2023, 16(15), 5377; https://doi.org/10.3390/ma16155377
Submission received: 19 April 2023 / Revised: 19 July 2023 / Accepted: 20 July 2023 / Published: 31 July 2023

Abstract

:
The promotional effect of potassium (~1.25 wt%) on a Co/MnOx catalyst was studied for samples prepared by the impregnation method in the steam reforming of ethanol (SRE) process at 420 °C for a H2O/EtOH molar ratio of 12/1. The catalysts were characterized using physicochemical methods to study their textural, structural, and redox properties. The XRD studies revealed that, during the treatment of both cobalt-based catalysts under a hydrogen atmosphere at 500 °C, Co0 and MnO phases were formed by the reduction in Co3O4 and Mn2O3/Mn3O4 phases, respectively. Potassium doping significantly improved stability and ability for the C–C bond cleavage of the Co/MnOx catalyst. The enhancement of activity (at ~25%) and selectivity to hydrogen (at ca. 10%) and the C1 product, mainly carbon dioxide (at ~20%), of the Co/MnOx catalyst upon potassium doping was clarified by the alkali promoter’s impact on the reducibility of the cobalt and manganese oxides. The microscopic observations revealed that fibrous carbon deposits are present on the surface of Co/MnOx and KCo/MnOx catalysts after the SRE reaction and their formation is the main reason these catalysts deactivate under SRE conditions. However, carbon accumulation on the surface of the potassium-promoted catalyst was ca. 12% lower after 18 h of SRE reaction compared to the unpromoted sample.

Graphical Abstract

1. Introduction

Hydrogen is considered an ideal energy carrier in fuel cell technology because of its high efficiency. Hydrogen is mainly produced from the steam reforming of fossil fuels, which emit harmful air pollutants into the atmosphere. The steam reforming of ethanol (SRE), C2H5OH + 3H2O → 6H2 + 2CO2, is of interest since it provides a promising method of hydrogen production from renewable resources such as bioethanol, which are easily obtained from biomass. The hydrogen production efficiency first depends on the nature of the catalyst used in the SRE reaction. The preferred catalysts should not be poisoned by carbon monoxide at the reaction temperature, should allow for a fast oxygen transfer and the formation of oxygen vacancy defects, and should contain metals that remain highly active after numerous oxidation-reduction cycles. Cobalt-based catalysts are proposed as more cost-efficient substitutes with a similar C–C bond scission ability in the SRE process to noble-metal-based catalysts. However, their drawback is significant deactivation, which occurs under SRE conditions mainly due to oxidation and sintering of the metallic active phase, the migration of cobalt ions into the support to form nonreducible mixed-metal cobalt oxides such as aluminates or silicates, and the accumulation of surface carbon species that poison the catalyst by blocking the active surface cobalt sites.
The support materials may strongly influence the catalytic performance of cobalt-based catalysts in the SRE reaction when allowing for the suppression of both cobalt sintering and coke formation [1,2,3,4,5,6]. Moreover, reducible oxides such as ZnO and CeO2 provide a spillover of oxygen species to the metal particles dispersed on their surfaces, thereby facilitating the oxidation of the carbon deposit [2,3,5,7,8,9,10,11]. However, although manganese oxides possess plenty of active lattice oxygen species that play a role in the efficient elimination of carbon deposition, there are not many reports [1,12,13,14,15,16] on their use as a support for the catalysts of SRE reaction. Both types of manganese oxides with layered and tunnel structures were the subject of this research [12,14,15]. Conventional methods were used to prepare these materials, such as redox precipitation to obtain tunneled structured potassium containing cryptomelane-type manganese dioxide [15], comproportionation reactions of Mn2+ and MnO4 under alkaline conditions to obtain layer-structure birnessite-type manganese oxide [12,14], or the hydrothermal treatment of layered manganese oxide precursor (birnessite) to obtain tunneled-structure todorokite-type manganese oxide [14,16]. Unfortunately, all these preparative processes of layer- or tunnel-structure manganese oxide have multiple steps and/or are time-consuming. Moreover, all these complex layer- and tunnel-structures of manganese oxide, during hydrogen pretreatment at elevated temperatures, are reduced to lower valence states, an effect that generally gives rise to strong structural and surface changes. The decomposition of the manganese oxides phase into a stable lower valent during the reduction can lead to the collapse of the tunnel structure. For example, Gac et al. [15] indicated that a reduction in the cryptomelane-type manganese oxide led to the development of large, cube-like MnO particles and the disappearance of its tunnel structure. As the activation of the SRE catalysts with hydrogen at elevated temperatures leads to a reduction in the manganese oxides to lower oxidation states with the concomitant evolution of oxygen and complete structural rearrangement, Greluk et al. [16] proposed a simple precipitation method for the preparation of a mixture of chemically simple and inexpensive manganese oxides, Mn2O3 and Mn3O4, as a support for the SRE catalyst.
A limitation of the coke deposition might be also achieved by the addition of alkali promoters [16,17,18,19], which can significantly change both the structure and surface of the catalyst. The key role of alkali metal additives is dependent on the nature of the catalytic material, e.g., the type and crystallite size of the active phase, the phase composition and the morphology of the support. Moreover, because of the high reactivity and volatility of alkalis, the high-temperature pretreatment of SRE catalysts and the interaction of numerous reagents of the SRE process can result in undesired processes such as alkali surface redistribution, a solid-state reaction with the active phase/support, their agglomeration, or even alkali promoters’ desorption, affecting the resulting performance of the catalyst. Therefore, the promotional effect of alkali must be elucidated individually for each catalytic system [17,20].
To our knowledge, no studies were conducted to investigate the impact of the potassium promoter on manganese-oxide-supported, cobalt-based catalysts in the SRE reaction. Therefore, this work aimed to determine the role that potassium promotion plays in the performance of the Co/MnOx catalyst during the SRE process. The chemical and physical properties of catalysts were characterized by BET, XRF, XRD, TPR, SEM, TEM and XPS methods.

2. Materials and Methods

2.1. Materials

The materials used in the preparation of samples: manganese(II) acetate tetrahydrate (≥97.0%, Sigma-Aldrich, Darmstadt, Germany), cobalt(II) nitrate hexahydrate (≥98.0%, Sigma-Aldrich), potassium nitrate (≥99.0%, Merck, Darmstadt, Germany) and ammonium carbonate (100%, Avantor, Gliwice, Poland) were used without further purification.

2.2. Preparation of Catalysts

The MnOx support was prepared using the precipitation method with 1 mol/L ammonium carbonate solution as a precipitant from the precursor of manganese acetate. Ammonium carbonate solution was added dropwise to manganese acetate solution with vigorous stirring until a pH of 8 was maintained at 40 °C. The resulting precipitate was aged at 60 °C for 2 h. The suspension was filtered and then washed several times with deionized water until reaching a pH of 7. The precipitate was then dried at 110 °C and calcined at 500 °C in air.
The Co/MnOx catalysts were prepared using the impregnation method by mixing cobalt nitrate solution and MnOx support to achieve 10 wt% cobalt loading. The sample was dried at 110 °C and calcined at 500 °C in air.
The KCo/MnOx catalysts were prepared using the impregnation method by mixing potassium nitrate solution and Co/MnOx sample to achieve 2 wt% potassium loading. The sample was dried at 110 °C and calcined at 500 °C in air.
A schematic drawing of the preparations of MnOx support and Co/MnOx and KCo/MnOx catalysts is presented in Figure 1.

2.3. Catalysts Characterization

Details about the physicochemical characterization of materials can be found in our previous work [16]. The metal loading of catalysts was determined by the X-ray fluorescence method, using an Axios mAX (PANalytical, Malvern, UK) fluorescence spectrometer. The BET surface area, pore volume, and pore diameter of the samples were measured by ASAP 2405N (Micromeritics, Norcross, GA, USA) instrument, using nitrogen adsorption/desorption isotherms collected at −196 °C. The X-ray powder diffraction (XRD) patterns were obtained with an Empyrean X-ray (PANalytical, Malvern, UK) diffractometer. Temperature-programmed reduction (TPR) studies were carried out with an AutoChem II 2920 (Micromeritics, Norcross, GA, USA) analyzer. The high-resolution transmission electron microscopy (HRTEM) images of catalysts were obtained with an electron transmission microscope Titan G2 60–300 kV (FEI Company, Hillsboro, OR, USA). X-ray photoelectron spectroscopy (XPS) studies were performed in a Kratos Axis Supra Spectrometer (Kratos Analytical, Manchester, UK) equipped with a monochromatized Al source.

2.4. SRE Experimental Tests

The catalyst activity test was carried out in a quartz tube reactor (inner diameter: 10 mm) using a Microactivity Reference unit (PID Eng & Tech., Alcobendas, Spain). The reactor was filled with a catalyst sample diluted with quartz. A total of 100 mg of catalyst sample was reduced with hydrogen at 500 °C for 1 h. After this reduction, the performance was evaluated at atmospheric pressure, a temperature of 420 °C and a H2O/EtOH molar ratio of 12/1, with a total flow rate equal to 100 mL min−1. The reaction substrates and products were analyzed by online gas chromatographs (Bruker 430-GC and Bruker 450-GC, (Bruker, Billerica, MA, USA) equipped with thermal conductivity detectors (TCD).
The conversion of ethanol, conversions of ethanol into individual carbon-containing products and the selectivity of hydrogen formation were calculated based on the previously reported equations [16].

3. Results and Discussion

3.1. The Effect of Potassium Doping on the Catalysts’ Physicochemical Properties

The results of BET analysis, quantitative analysis (analyzed by XRF method), and values of average crystallite size (analyzed by XRD and TEM methods) of Co/MnOx and KCo/MnOx catalysts are summarized in Table 1. The desired loading of cobalt (10 wt%) was obtained with good accuracy. However, the potassium promoter content is less than the intended loading of 2 wt%. The Co/MnOx catalyst exhibits a low surface area, which decreases by more than 50% following re-calcination after the addition of a potassium promoter. Because the low surface area of support does not facilitate the dispersion of metal particles, both Co/MnOx and KCo/MnOx catalysts exhibit a large Co0 particle size (see also Supplementary Materials, Figure S1). The slightly larger size of Co0 particles for KCo/MnOx catalysts results from the agglomeration of the initially formed cobalt species crystallites of the Co/MnOx catalyst during its second calcination, after impregnation with potassium salt [18]. STEM-EDS maps of samples, after their reduction with hydrogen at 500° (Figure 2 and Figure 3), reveal a comparatively homogenous cobalt dispersion on the MnOx support of both catalysts, as well as potassium accumulation on both MnOx and Co0 particles for the K-doped sample.
Figure 4A,B shows the XRD patterns of the calcined Co/MnOx and KCo/MnOx catalysts, which reveal that both Co/MnOx and KCo/MnOx samples contain a mixture of tetragonal Mn3O4 (JCPDS file no. 04-005-9818) and cubic Mn2O3 phases (JCPDS file no. 00-001-1127). However, in the case of both catalysts, the major species is Mn2O3. Diffraction peaks observed at diffraction angles (2θ) of 24.0°, 33.1°, 38.6°, 45.3°, 49.5°, 55.2°, 60.9° and 66.0° can be assigned to the (211), (222), (400), (332), (431), (400), (611) and (622) crystalline planes of the cubic Mn2O3 crystal. Similarly, prominent diffraction peaks appeared at 2θ of 28.0°, 31.4°, 32.2°, 36.1°, 44.8°, 58.8°, 60.0°, which can be assigned to the (112), (200), (103), (211), (220), (321) and (224) crystalline planes of the tetragonal Mn3O4 phase. Moreover, because the main reflection for Mn3O4 at 2θ = 33.1° is nearly identical to the main reflection of Co3O4 (JCPDS file no. 04-002-0644) at 2θ = 36.1°, the possibility that these both reflections overlap should not be excluded. However, the XRD pattern of both catalysts mainly diffracts at angles that can attributed to the phases of support due to its much higher content compared to the Co3O4 active phase. Upon the reduction in both cobalt-based catalysts for 2 h at 500 °C, the XRD peaks corresponding to the cubic-phase Mn2O3 and tetragonal-phase Mn3O4 disappeared completely, and new reflections that appeared at 34.9°, 40.4°, 58.5°, 70.0 and 73.6° are mainly present, which can be assigned to the (111), (200), (220), (311) and (222) planes of the cubic structure of MnO (JCPDS file no. 00-075-1090) (Figure 4C,D). In addition to the predominant diffraction peaks derived from the MnO support, weaker diffraction lines at 2θ = 43.7° and 51.2°, corresponding to the reflection planes of the Co0 phase, are visible [16,21,22,23,24,25,26,27,28]. Moreover, the absence of any potassium phase for the KCo/MnOx catalyst in the XRD patterns may result from several factors, namely its considerably low amount and/or amorphous form and/or high dispersion over the catalyst surface [16,29]. Unfortunately, due to the low potassium-promoter content, there is a problem with the understanding of its role, since it is difficult to characterize this on the catalytic surface [29].
Detailed HRTEM measurements (Figure 5 and Figure 6) were taken for both cobalt-based catalysts after their reduction at 500 °C, and the results are in good agreement with the XRD findings. Well-crystallized Co0 and MnO phases are identified for both Co/MnOx (Figure 5) and KCo/MnOx (Figure 6) catalysts. For both catalysts, HRTEM images reveal lattice fringes with an interplanar spacing of 0.205, 0.222 and 0.225 nm, which may be ascribed to the (111) plane of Co0, (200) and (111) planes of MnO, respectively. Moreover, for the Co/MnOx sample, lattice fringes with an interplanar spacing of 0.157 nm, which can be ascribed to the (220) planes of MnO, are also observed. The planes of both samples were also identified by the spots on the corresponding FFT pattern.
The reducibility of the MnOx support, Co/MnOx and KCo/MnOx catalysts, was also studied by the H2-TPR method, and the results are displayed in Figure 7. According to the literature [16,30,31,32,33], bare MnOx exhibits a low-temperature peak at 220–330 °C, which can be assigned to the reduction from Mn2O3 to Mn3O4 and high temperature at ca. 330–520 °C, corresponding to the transformation of Mn3O4 to MnO. Based on the study of the literature [34,35], due to two reduction steps, the H2-TPR profile of Co3O4 has a low-temperature peak below 300 °C, assigned to the reduction from Co3O4 to CoO, and a high-temperature peak between 300 and 700 °C, corresponding to the further reduction from CoO to Co0. This means that the reduction peaks in manganese oxides overlapped with those of cobalt oxides and could not be distinguished. The H2-TPR profile of the Co/MnOx catalyst has two regions of hydrogen consumption: a low-temperature region at 220–310 °C and a high-temperature region at 310–750 °C, which could correspond to the reduction from Mn2O3 to Mn3O4 and Co3O4 to CoO, and the reduction from Mn3O4 to MnO and CoO to Co0, respectively. Moreover, the broad peak observed in the high-temperature region (above 500 °C) could be a result of the synergetic interactions between Mn and Co species [36]. A reduction in the K-containing sample leads to a lower temperature compared to a reduction in Co/MnOx material. In addition, two reduction peaks were observed that could be ascribed to the stepwise reduction in manganese and cobalt oxides. Moreover, a small shoulder below 260 °C is related to the presence of surface-active species with different Mn–O bond strengths [16,37]. The KCo/MnOx sample (Table 1) exhibits high hydrogen consumption, which may result from the reduction in nitrates stored on the catalyst’s surface, facilitated by the presence of potassium [16]. It can be supposed that the nitrate reducibility is increased by the presence of cobalt and hydrogen, which is dissociated upon the Co0 particles’ spilling onto the nitrate, reducing their quantity [38].

3.2. The Effect of Potassium Doping on the Performance of the Catalysts in the Ethanol Steam Reforming Reaction

The effect of potassium doping on the cobalt-based catalyst in the SRE process was determined by ethanol conversion (Figure 8) and selectivity to products (Figure 9) at the temperature of 420 °C for a H2O/EtOH molar ratio of 12/1. The initial complete conversion of ethanol decreases after 18 h of the SRE process to ca. 40 and 70% over Co/MnOx and KCo/MnOx catalysts, respectively. At the beginning of the SRE reaction, the KCo/MnOx catalyst exhibits high selectivity to hydrogen and carbon dioxide. In contrast, hydrogen and carbon monoxide are mainly produced in the presence of the Co/MnOx sample, which suggests an increase in activity in water gas shift (WGS, Reaction (1)) reactions in the presence of a potassium-doped catalyst:
CO + H2O ↔ CO2 + H2
Furthermore, the C2 and C3 products, i.e., acetaldehyde, ethylene and acetone, are present among the by-products of a Co/MnOx sample from the beginning of the SRE reaction, which indicates the poor ability of this catalyst in terms of the C–C bond cleavage. The addition of potassium allows for restrictions to the production of these products and, initially, only small amounts of acetaldehyde are formed over the KCo/MnOx sample. However, the amount of this product drastically increases with the increase in time-on-stream over both cobalt-based catalysts. After 18 h of the SRE process, selectivity to acetaldehyde is ca. 70 and 55% for Co/MnOx and KCo/MnOx catalysts, respectively. Next to hydrogen, it is the second product that is mainly produced in the presence of these catalysts during the SRE process. This indicates that the ability of both samples in the C–C bond scission decreases with the increase in time-on-stream because of the deactivation of cobalt active sites on the catalysts’ surface under SRE conditions. However, the lower selectivity to the C2 and C3 products and higher selectivity to the C1 products suggests that more cobalt-active sites remained available on the surface of the KCo/MnOx catalyst to break the C–C bond.

3.3. The Effect of Potassium Doping on Prevention of the Cobalt-Based Catalyst Deactivation under SRE Conditions

The microscopic (Figure 10 and Figure 11) observations revealed that fibrous carbon deposits are present on the surface of Co/MnOx and Kco/MnOx catalysts. The cobalt nanocrystals are almost completely detached from the MnOx support by carbon deposits, and they are either confined to the fibres or located at the tip of the carbon filaments, being totally or partially surrounded by a carbon shell.
Based on the HRTEM images and the corresponding phase identification obtained using the FFT method (Figure 11), two forms of cobalt-active phase were identified in the case of the spent Co/MnOx catalyst, namely Co0 and CoO. However, the Co0 form of the cobalt-active phase was detected for the spent KCo/MnOx catalyst. Moreover, support for the Co/MnOx sample was identified as including two forms, MnO and Mn2O3, where only the MnO phase was detected for the support of the spent KCo/MnOx catalyst.
Figure 12 shows the amount of carbon deposits that were formed on the surface of the Co/MnOx and KCo/MnOx catalysts as a function of time. The rate of carbon formation significantly decreased in the presence of a potassium promoter on the surface of a cobalt-based catalyst, indicating that its addition can inhibit carbon deposition and/or promote gasification [16]. Potassium species probably migrated and covered the surface of the Co0 particles, which favoured the adsorption-dissociation of the H2O molecule, and therefore increased the direct oxidation of the carbonaceous species [39].
The survey spectra of spent cobalt-based catalysts (H2O/ethanol = 12/1, 420 °C) show the major signals originating from C, O and Mn elements, and minor signals originating from Co. The atomic surface coverage by carbon deposits was estimated at around 81.5 at.% for Co/MnOx and 91 at.% for KCo/MnOx catalysts. Most of the carbon exists in graphitic-like compounds, as confirmed by the asymmetric C 1s peak shape (284.5 eV, not shown here). The high carbon coverage of the surface leads us to expect that most crystallites of the active phase and support were occluded in carbon; therefore, their oxidation state remained unchanged even upon exposure to air (Figure 13). The Co 2p3/2 spectrum can be resolved into components, which are attributed to the Co–Co (778.4 eV) and Co–O bonding (780.4 eV), respectively [40,41]. A higher concentration of metallic cobalt species can be observed for the potassium-promoted catalyst, suggesting that this promotor plays a very important role in the reduction in the active phase. For the manganese oxidation state, the literature data [41] suggest that MnO and Mn2O3 phases are present. The relative concentration of Mn(II)/(Mn(II) + Mn(III)) in both samples is very similar, in the range of 51%. The interpretation of the results obtained for bare manganese support requires a comparison of both the spectrum recorded for the as-calcined sample and after ESR reaction. This is important, as bare support was rather inactive in the ESR [18], whereas the survey spectrum suggests almost 55 at.% coverage of the surface by carbon-containing species. This result seems less surprising when one notes that the atomic contribution of adventitious carbon species on the surface of the calcined sample is equal to 39 at.%. The calculation of the relative concentration of Mn(II) in the overall Mn 2p3/2 spectrum suggests a lower concentration of these species (43%) as compared to the results obtained for cobalt-based catalysts.

4. Conclusions

The doping effect of a potassium-promoter on the novel Co/MnOx catalytic material’s performance in the SRE reaction was studied, focusing on the catalyst’s ability to resist carbon growth. It is beneficial to consider the significant increase in catalytic stability achieved by the Kco/MnOx catalyst in comparison with the Co/MnOx sample, and the distribution of the products over these catalysts under SRE conditions. Compared with the unpromoted Co/MnOx catalyst, the Kco/MnOx sample indicated an increase in the selectivity to two main products of the SRE reaction, i.e., hydrogen (Co/MnOx—ca. 55% and Kco/MnOx—ca. 70%) and carbon dioxide (Co/MnOx—ca. 20% and Kco/MnOx—ca. 40%). In turn, selectivity to C2 products was ca. 20% higher in the presence of an unpromoted Co/MnOx sample, indicating that more cobalt-active sites remained available on the surface of the KCo/MnOx catalyst to cleave the C–C bond. Although fibrous carbon was present on the surface of both Co/MnOx and KCo/MnOx samples, the carbon accumulation was ca. 12% higher on the surface of unpromoted catalysts after 18 h of SRE reaction. This indicates that potassium species probably migrated and covered the surface of the Co0 particles, which favoured the adsorption–dissociation of the H2O molecule, and therefore increased the direct oxidation of the carbonaceous species. It could be concluded that the crucial effect of potassium dopant is the decrease in carbon accumulation on the catalyst surface and the increase in the number of sites for water adsorption, which improve the catalyst stability. Furthermore, the alkali metal could change the electronic properties of the cobalt species, and hence facilitate the redox ability of active sites both during activation with hydrogen and under SRE conditions. The maintenance of a reduced oxidation state for the catalyst surface during the SRE process is important to ensure an effective C—C bond; thus, the catalytic stability of a potassium cobalt-based catalyst was improved. According to the results of the characterization, the addition of the alkali metal did not influence the cobalt active-phase dispersion, but was not a key parameter improving Co/MnOx catalysts’ activity, selectivity and resistance to deactivation.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma16155377/s1, Figure S1: Particle size distribution histograms for (a) Co/MnOx and (b) KCo/MnOx catalysts reduced at 500 °C.

Author Contributions

Conceptualization, M.G.; formal analysis, M.G. and G.G.; funding acquisition, G.G.; investigation, M.G., M.R., G.S., S.T.-S., G.G. and K.T.-R.; methodology, M.G., M.R., G.S., S.T.-S., G.G. and K.T.-R.; supervision, M.G.; writing—original draft, M.G.; visualization, M.G.; writing—review and editing, M.G. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Science Centre, Poland within the SONATA program under grant no. 2021/43/D/ST4/03016.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sohrabi, S.; Irankhah, A. Synthesis, characterization, and catalytic activity of Ni/CeMnO2 catalysts promoted by copper, cobalt, potassium and iron for ethanol steam reforming. Int. J. Hydrogen Energy 2021, 46, 12846–12856. [Google Scholar] [CrossRef]
  2. Yu, S.-W.; Huang, H.-H.; Tang, C.-W.; Wang, C.-B. The effect of accessible oxygen over Co3O4-CeO2 catalysts on the steam reforming of ethanol. Int. J. Hydrogen Energy 2014, 39, 20700–20711. [Google Scholar] [CrossRef]
  3. Liu, Z.; Duchoň, T.; Wang, H.; Peterson, E.W.; Zhou, Y.; Luo, S.; Zhou, Y.; Matolín, V.; Stacchiola, D.J.; Rodriguez, J.A.; et al. Mechanistic insight of ethanol steam reforming over Ni-CeOx(111): The importance of hydroxyl groups for suppressing coke formation. J. Phys. Chem C 2015, 119, 18248–18256. [Google Scholar] [CrossRef]
  4. Zhu, S.; Liu, L. A comparative DFT study of ethanol steam reforming over Co(1 0 0) and CoO(1 0 0) surfaces: Molecular reaction mechanism. Comput. Theor. Chem. 2022, 1217, 113926. [Google Scholar] [CrossRef]
  5. Idriss, H. Ethanol reactions over the surfaces of noble metal/cerium oxide catalysts. Platin. Met. Rev. 2004, 48, 105–115. [Google Scholar] [CrossRef]
  6. Passos, A.R.; Martins, L.; Pulcinelli, S.H.; Santilli, C.V.; Briois, V. Effect of the balance between Co(II) and Co(0) oxidation states on the catalytic activity of cobalt catalysts for Ethanol Steam Reforming. Catal. Today 2014, 229, 88–94. [Google Scholar] [CrossRef]
  7. Liu, X.; Murata, K.; Inaba, M. Steam Reforming of Bio-Ethanol to Produce Hydrogen over Co/CeO2 Catalysts Derived from Ce1−xCoxO2−y Precursors. Catalysts 2016, 6, 26. [Google Scholar] [CrossRef]
  8. You, W.; Zhang, Q.; Jia, H.; Ta, N.; Sheng, X.; Yang, X.; Wang, Y.; Shen, W.; Goldbach, A. Insights into the state of ceria during ethanol steam reforming over Ir/CeO2. Catal. Sci. Technol. 2023, 13, 558–572. [Google Scholar] [CrossRef]
  9. Palma, V.; Castaldo, F.; Ruocco, C.; Ciambelli, P.; Iaquaniello, G. Low temperature-ethanol steam reforming over Ni-based catalysts supported on CeO2. J. Power Technol. 2015, 95, 54–66. [Google Scholar]
  10. Chiou, J.Y.Z.; Wang, W.-Y.; Yang, S.-Y.; Lai, C.-L.; Huang, H.-H.; Wang, C.-B. Ethanol Steam Reforming to Produce Hydrogen Over Co/ZnO and PtCo/ZnO Catalysts. Catal. Lett. 2013, 143, 501–507. [Google Scholar] [CrossRef]
  11. Yu, N.; Zhang, H.; Davidson, S.D.; Sun, J.; Wang, Y. Effect of ZnO facet on ethanol steam reforming over Co/ZnO. Catal. Commun. 2016, 73, 93–97. [Google Scholar] [CrossRef]
  12. Da Costa-Serra, J.F.; Chica, A. Catalysts based on Co-Birnessite and Co-Todorokite for the efficient production of hydrogen by ethanol steam reforming. Int. J. Hydrogen Energy 2018, 43, 16859–16865. [Google Scholar] [CrossRef]
  13. Kwak, B.S.; Lee, G.; Park, S.-M.; Kang, M. Effect of MnOx in the catalytic stabilization of Co2MnO4 spinel during the ethanol steam reforming reaction. Appl. Catal. A Gen. 2015, 503, 165–175. [Google Scholar] [CrossRef]
  14. Fuertes, A.; Da Costa-Serra, J.F.; Chica, A. New catalysts based on Ni-Birnessite and Ni-Todorokite for the efficient production of hydrogen by bioethanol steam reforming. Energy Procedia 2012, 29, 181–191. [Google Scholar] [CrossRef]
  15. Gac, W.; Greluk, M.; Słowik, G.; Turczyniak-Surdacka, S. Structural and surface changes of cobalt modified manganese oxide during activation and ethanol steam reforming reaction. Appl. Surf. Sci. 2018, 440, 1047–1062. [Google Scholar] [CrossRef]
  16. Greluk, M.; Rotko, M.; Słowik, M.; Turczyniak-Surdacka, S.; Grzybek, G.; Góra-Marek, K.; Kotarba, A. Effect of Potassium Promoter on the Performance of Nickel-Based Catalysts Supported on MnOx in Steam Reforming of Ethanol. Catalysts 2022, 12, 600. [Google Scholar] [CrossRef]
  17. Grzybek, G.; Góra-Marek, K.; Tarach, K.; Pyra, K.; Patulski, P.; Greluk, M.; Słowik, G.; Kotarba, A. Tuning the properties of the cobalt-zeolite nanocomposite catalyst by potassium: Switching between dehydration and dehydrogenation of ethanol. J. Catal. 2022, 407, 364–380. [Google Scholar] [CrossRef]
  18. Grzybek, G.; Greluk, M.; Tarach, K.; Pyra, K.; Słowik, G.; Rotko, M.; Góra-Marek, K. Bioethanol Steam Reforming over Cobalt-Containing USY and ZSM-5 Commercial Zeolite Catalysts. Front. Mater. 2020, 7, 597528. [Google Scholar] [CrossRef]
  19. Yoo, S.; Park, S.; Song, H.; Kim, D.H. Hydrogen production by the steam reforming of ethanol over K-promoted Co/Al2O3–CaO xerogel catalysts. Mol. Catal. 2020, 491, 110980. [Google Scholar] [CrossRef]
  20. Grzybek, G.; Wójcik, S.; Legutko, P.; Gryboś, J.; Indyka, P.; Leszczyński, B.; Kotarba, A.; Sojka, Z. Thermal stability and repartition of potassium promoter between the support and active phase in the K-Co2.6Zn0.4O4|α-Al2O3 catalyst for N2O decomposition: Crucial role of activation temperature on catalytic performance. Appl. Catal. B Environ. 2017, 205, 597–604. [Google Scholar] [CrossRef]
  21. Sengottaiyan, C.; Jayavel, R.; Shrestha, R.G.; Hill, J.P.; Ariga, K.; Shrestha, L.K. Electrochemical Supercapacitance Properties of Reduced Graphene Oxide/Mn2O3:Co3O4 Nanocomposite. J. Inorg. Organomet. Polym. 2017, 27, 576–585. [Google Scholar] [CrossRef]
  22. Sharrouf, M.; Awad, R.; Roumié, M.; Marhaba, S. Structural, Optical and Room Temperature Magnetic Study of Mn2O3 Nanoparticles. Mater. Sci. Appl. 2015, 6, 850–859. [Google Scholar]
  23. Gnana Sundara Raj, B.; Asiri, A.M.; Wu, J.J.; Anandan, S. Synthesis of Mn3O4 nanoparticles via chemical precipitation approach for supercapacitor application. J. Alloys Compd. 2015, 636, 234–240. [Google Scholar] [CrossRef]
  24. Khalaji, A.D.; Soleymanifard, M.; Jarosova, M.; Machek, P. Facile synthesis and characterization of Mn3O4, Co3O4, and NiO. Acta Phys. Pol. 2002, 137, 1043–1045. [Google Scholar] [CrossRef]
  25. Ciura, K.; Grzybek, G.; Wójcik, S.; Indyka, P.; Kotarba, A.; Sojka, S. Optimization of cesium and potassium promoter loading in alkali-doped Zn0.4Co2.6O4|Al2O3 catalysts for N2O abatement. React. Kinet. Mech. Catal. 2017, 121, 645–655. [Google Scholar] [CrossRef]
  26. Hou, H.; Li, J.; Lan, J.; Chen, F.; Huang, B. Porous walnut-like Mn2O3 anode derived from the waste graphene production effluent. J. Porous Mater. 2022, 29, 837–847. [Google Scholar] [CrossRef]
  27. Atique Ullah, A.K.M.; Fazle Kibria, A.K.M.; Akter, M.; Khan, M.N.I.; Tareq, A.R.M.; Firoz, S.H. Oxidative Degradation of Methylene Blue Using Mn3O4 Nanoparticles. Water Conserv. Sci. Eng. 2017, 1, 249–256. [Google Scholar] [CrossRef]
  28. Martinez de la Torre, C.; Bennewitz, M.F. Manganese oxide nanoparticle synthesis by thermal decomposition of manganese(II) acetylacetonate. J. Vis. Exp. 2020, 160, e61572. [Google Scholar]
  29. Grzybek, G.; Greluk, M.; Indyka, P.; Góra-Marek, K.; Legutko, P.; Słowik, G.; Turczyniak-Surdacka, S.; Rotko, M.; Sojka, Z.; Kotarba, A. Cobalt catalyst for steam reforming of ethanol–Insights into the promotional role of potassium. Int. J. Hydrogen Energy 2020, 45, 22658–22673. [Google Scholar] [CrossRef]
  30. Li, J.; Li, L.; Cheng, W.; Wu, F.; Lu, X.; Li, Z. Controlled synthesis of diverse manganese based catalysts for complete oxidation of toluene and carbon monoxide. Chem. Eng. J. 2014, 22, 59–67. [Google Scholar] [CrossRef]
  31. Shu, S.; Guo, J.; Li, J.; Fang, N.; Li, J.; Yuan, S. Effect of post-treatment on the selective catalytic reduction of NO with NH3 over Mn3O4. Mater. Chem. Phys. 2019, 237, 121845. [Google Scholar] [CrossRef]
  32. Gil, A.; Gandía, L.M.; Korili, S.A. Effect of the temperature of calcination on the catalytic performance of manganese- and samarium-manganese-based oxides in the complete oxidation of acetone. Appl. Catal. A Gen. 2004, 274, 229–235. [Google Scholar] [CrossRef]
  33. Chen, Z.; Yang, Q.; Li, H.; Li, X.; Wang, L.; Tsang, S.C. Cr–MnOx mixed-oxide catalysts for selective catalytic reduction of NOx with NH3 at low temperature. J. Catal. 2010, 276, 56–65. [Google Scholar] [CrossRef]
  34. Lin, H.-Y.; Chen, Y.-W. The mechanism of reduction of cobalt by hydrogen. Mater. Chem. Phys. 2004, 85, 171–175. [Google Scholar] [CrossRef]
  35. Darda, S.; Pachatouridou, E.; Lappas, A.; Iliopoulou, E. Effect of preparation method of Co-Ce catalysts on CH4 combustion. Catalysts 2019, 9, 219. [Google Scholar] [CrossRef]
  36. Zhong, L.; Barreau, M.; Chen, D.; Caps, V.; Haevecker, M.; Teschnerb, D.; Simonne, D.H.; Borfecchia, E.; Baaziz, W.; Šmí, B.; et al. Effect of manganese promotion on the activity and selectivity of cobalt catalysts for CO preferential oxidation. Appl. Catal. B Environ. 2021, 297, 120397. [Google Scholar] [CrossRef]
  37. Zhang, X.; Yang, Y.; Zhu, Q.; Ma, M.; Jiang, Z.; Liao, X.; He, C. Unraveling the effects of potassium incorporation routes and positions on toluene oxidation over α-MnO2 nanorods: Based on experimental and density functional theory (DFT) studies. J. Colloid Interface Sci. 2021, 598, 324–338. [Google Scholar] [CrossRef]
  38. Grzona, C.B.; Lick, I.D.; Rodriguez Castellón, E.; Ponzi, M.I.; Ponzi, E.N. Cobalt and KNO3 supported on alumina catalysts for diesel soot combustion. Mater. Chem. Phys. 2010, 123, 557–562. [Google Scholar] [CrossRef]
  39. Claude, V.; Mahy, J.G.; Lohay, T.; Tilkin, R.G.; Micheli, F.; Lambert, S.D. Sol–gel synthesis of Ni/Al2O3 catalysts for toluene reforming: Support modification with alkali, alkaline earth or rare-earth dopant (Ca, K, Mg or Ce). Surf. Interfaces 2020, 20, 100511. [Google Scholar] [CrossRef]
  40. Yildirim, O.A.; Arslan, H.; Sönmezoglu, S. Facile synthesis of cobalt-doped zinc oxide thin films for highly efficient visible light photocatalysts. Appl. Surf. Sci. 2016, 390, 111–121. [Google Scholar] [CrossRef]
  41. Biesinger, M.C.; Payne, B.P.; Grosvenor, A.P.; Lau, L.V.M.; Gerson, A.R.; Smart, R.S.C. Resolving surface chemical states in XPS analysis of first row transition metals, oxides and hydroxides: Cr, Mn, Fe, Co and Ni. Appl. Surf. Sci. 2011, 257, 2717–2730. [Google Scholar] [CrossRef]
Figure 1. A schematic drawing of the preparations of MnOx support and Co/MnOx and KCo/MnOx catalysts.
Figure 1. A schematic drawing of the preparations of MnOx support and Co/MnOx and KCo/MnOx catalysts.
Materials 16 05377 g001
Figure 2. STEM-EDS analysis of Co/MnOx catalyst after their reduction with hydrogen at 500 °C.
Figure 2. STEM-EDS analysis of Co/MnOx catalyst after their reduction with hydrogen at 500 °C.
Materials 16 05377 g002
Figure 3. STEM-EDS analysis of KCo/MnOx catalyst after their reduction with hydrogen at 500 °C.
Figure 3. STEM-EDS analysis of KCo/MnOx catalyst after their reduction with hydrogen at 500 °C.
Materials 16 05377 g003
Figure 4. XRD patterns of Co/MnOx and KCo/MnOx catalysts after their calcination in-air at 500 °C (A,B) and after their reduction with hydrogen at 500 °C (C,D).
Figure 4. XRD patterns of Co/MnOx and KCo/MnOx catalysts after their calcination in-air at 500 °C (A,B) and after their reduction with hydrogen at 500 °C (C,D).
Materials 16 05377 g004
Figure 5. HRTEM images from different points (A1,B1) of Co/MnOx catalyst after reduction with hydrogen at 500 °C with the complementary phase identification ((A2,B2), Co0—yellow; MnO—turquoise) and the corresponding fast Fourier transform (FFT) patterns in the turquoise squares as the inset of (A1,B1). Arrows represent distance between the two the two lattice planes. Magnified view of the selected areas was enclosed in the dashed square to show diffraction fringes and their respective interplanar distances.
Figure 5. HRTEM images from different points (A1,B1) of Co/MnOx catalyst after reduction with hydrogen at 500 °C with the complementary phase identification ((A2,B2), Co0—yellow; MnO—turquoise) and the corresponding fast Fourier transform (FFT) patterns in the turquoise squares as the inset of (A1,B1). Arrows represent distance between the two the two lattice planes. Magnified view of the selected areas was enclosed in the dashed square to show diffraction fringes and their respective interplanar distances.
Materials 16 05377 g005
Figure 6. HRTEM images from different points (A1,B1) of KCo/MnOx catalyst after its reduction with hydrogen at 500 °C with the complementary phase identification ((A2,B2), Co0—yellow; MnO—turquoise) and the corresponding fast Fourier transform (FFT) patterns in the turquoise squares as the inset of (A1,B1). Arrows represent distance between the two the two lattice planes. Magnified view of the selected areas was enclosed in the dashed square to show diffraction fringes and their respective interplanar distances. Obtained from the HRTEM image in (A).
Figure 6. HRTEM images from different points (A1,B1) of KCo/MnOx catalyst after its reduction with hydrogen at 500 °C with the complementary phase identification ((A2,B2), Co0—yellow; MnO—turquoise) and the corresponding fast Fourier transform (FFT) patterns in the turquoise squares as the inset of (A1,B1). Arrows represent distance between the two the two lattice planes. Magnified view of the selected areas was enclosed in the dashed square to show diffraction fringes and their respective interplanar distances. Obtained from the HRTEM image in (A).
Materials 16 05377 g006
Figure 7. H2-TPR profiles of MnOx support and Co/MnOx and KCo/MnOx catalysts.
Figure 7. H2-TPR profiles of MnOx support and Co/MnOx and KCo/MnOx catalysts.
Materials 16 05377 g007
Figure 8. The ethanol conversion over Co/MnOx and KCo/MnOx catalysts after 18 h of the SRE process at 420 °C for an H2O/EtOH molar ratio of 12/1.
Figure 8. The ethanol conversion over Co/MnOx and KCo/MnOx catalysts after 18 h of the SRE process at 420 °C for an H2O/EtOH molar ratio of 12/1.
Materials 16 05377 g008
Figure 9. The selectivity to products at the beginning (A) and after 18 h (B) of the SRE process at 420 °C for a H2O/EtOH molar ratio of 12/1 over Co/MnOx and Kco/MnOx catalysts.
Figure 9. The selectivity to products at the beginning (A) and after 18 h (B) of the SRE process at 420 °C for a H2O/EtOH molar ratio of 12/1 over Co/MnOx and Kco/MnOx catalysts.
Materials 16 05377 g009
Figure 10. SEM images of Co/MnOx (A) and KCo/MnOx (B) catalyst after 18 h of SRE reaction at 420 °C for H2O/EtOH molar ratio of 12/1.
Figure 10. SEM images of Co/MnOx (A) and KCo/MnOx (B) catalyst after 18 h of SRE reaction at 420 °C for H2O/EtOH molar ratio of 12/1.
Materials 16 05377 g010
Figure 11. Microscopic analysis of Co/MnOx (A) and KCo/MnOx (B) catalyst after 18 h of SRE reaction at 420 °C for H2O/EtOH molar ratio of 12/1. TEM images (A1,B1), HRTEM images (A2,B2) and the corresponding phase identification (A3,B3), Co0—yellow; CoO—green, MnO—turquoise, C—blue).
Figure 11. Microscopic analysis of Co/MnOx (A) and KCo/MnOx (B) catalyst after 18 h of SRE reaction at 420 °C for H2O/EtOH molar ratio of 12/1. TEM images (A1,B1), HRTEM images (A2,B2) and the corresponding phase identification (A3,B3), Co0—yellow; CoO—green, MnO—turquoise, C—blue).
Materials 16 05377 g011
Figure 12. Changes in Co/MnOx and KCo/MnOx catalysts’ weight during 18 h of SRE reaction at 420 °C for H2O/EtOH molar ratio of 12/1.
Figure 12. Changes in Co/MnOx and KCo/MnOx catalysts’ weight during 18 h of SRE reaction at 420 °C for H2O/EtOH molar ratio of 12/1.
Materials 16 05377 g012
Figure 13. High-resolution XPS spectra of Mn 2p and Co 2p regions collected from the surface of Co/MnOx and KCo/MnOx catalysts after 18 h of SRE reaction at 420 °C for H2O/EtOH molar ratio of 12/1.
Figure 13. High-resolution XPS spectra of Mn 2p and Co 2p regions collected from the surface of Co/MnOx and KCo/MnOx catalysts after 18 h of SRE reaction at 420 °C for H2O/EtOH molar ratio of 12/1.
Materials 16 05377 g013
Table 1. Physicochemical parameters for Co/MnOx and KCo/MnOx catalysts.
Table 1. Physicochemical parameters for Co/MnOx and KCo/MnOx catalysts.
Parameter Co/MnOxKCo/MnOx
Metal content (wt%) 9.79.7
Potassium content (wt%) -1.25
BET surface area (m2/g) 11.25.8
Co0 crystallite size (nm) 1By XRD1423
By TEM2126
H2 consumption Theoretical (mmol/gCo)2.182.19
Experimental (mmol/g)5.217.23
1 Co0 denoted Co particle size for a catalyst reduced at 500 °C.
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

Greluk, M.; Rotko, M.; Słowik, G.; Turczyniak-Surdacka, S.; Grzybek, G.; Tyszczuk-Rotko, K. Effect of Potassium Doping on the Structural and Catalytic Properties of Co/MnOx Catalyst in the Steam Reforming of Ethanol. Materials 2023, 16, 5377. https://doi.org/10.3390/ma16155377

AMA Style

Greluk M, Rotko M, Słowik G, Turczyniak-Surdacka S, Grzybek G, Tyszczuk-Rotko K. Effect of Potassium Doping on the Structural and Catalytic Properties of Co/MnOx Catalyst in the Steam Reforming of Ethanol. Materials. 2023; 16(15):5377. https://doi.org/10.3390/ma16155377

Chicago/Turabian Style

Greluk, Magdalena, Marek Rotko, Grzegorz Słowik, Sylwia Turczyniak-Surdacka, Gabriela Grzybek, and Katarzyna Tyszczuk-Rotko. 2023. "Effect of Potassium Doping on the Structural and Catalytic Properties of Co/MnOx Catalyst in the Steam Reforming of Ethanol" Materials 16, no. 15: 5377. https://doi.org/10.3390/ma16155377

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