Next Article in Journal
Oxidation in Flow Using an Ionic Immobilized TEMPO Catalyst on an Ion Exchange Resin
Previous Article in Journal
Catalytic Ozonation of Sulfachloropyridazine Sodium by Diatomite-Modified Fe2O3: Mechanism and Pathway
Previous Article in Special Issue
Ammonium Phosphotungstate Bonded on Imidazolized Activated Carbon for Selective Catalytic Rearrangement of α-Epoxypinane to Carveol
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Cu/MgO as an Efficient New Catalyst for the Non-Oxidative Dehydrogenation of Ethanol into Acetaldehyde

1
Shanghai Key Laboratory of Molecular Catalysis and Innovative Materials, Department of Chemistry, Fudan University, Shanghai 200438, China
2
SINOPEC Shanghai Research Institute of Petrochemical Technology Co., Ltd., Shanghai 201208, China
*
Author to whom correspondence should be addressed.
Catalysts 2024, 14(8), 541; https://doi.org/10.3390/catal14080541
Submission received: 25 July 2024 / Revised: 14 August 2024 / Accepted: 16 August 2024 / Published: 19 August 2024
(This article belongs to the Special Issue Catalytic Conversion of Renewable Biomass Platform Molecules)

Abstract

:
The non-oxidative dehydrogenation of ethanol into acetaldehyde is one of the efficient solutions for biomass upgrading. In this work, a series of copper catalysts supported on MgO with different Cu loadings ranging from 2.5% to 20% were prepared by an impregnation method. The as-synthesized Cu/MgO catalysts were characterized by N2 adsorption, XRD, TEM, CO2-TPD, XPS and TPR. These catalysts were found to be effective for ethanol dehydrogenation into acetaldehyde. As the Cu loading was increased, the ethanol conversion first increased and then leveled off. At a WHSV of 1.5 h−1 and 250 °C, the 20%Cu/MgO catalyst gave an initial conversion of 81.5%, with 97.7% selectivity toward acetaldehyde. Compared to 20%Cu/SiO2, the 20%Cu/MgO catalyst displayed an equivalent initial acetaldehyde yield, higher acetaldehyde selectivity and longer stability.

Graphical Abstract

1. Introduction

The development of alternative energy sources from renewable biomass and its derivatives contributes to alleviating the dependence on fossil energy. Particularly, ethanol, which can be obtained from crops via fermentation easily and abundantly, has attracted extraordinary attention [1,2]. In the past few decades, in addition to use as a gasoline additive, ethanol was utilized widely as a crucial building block for the production of ethylene, diethyl ether, acetaldehyde, ethyl acetate, propylene, butadiene, butanol, aromatic alcohol and so on [1,2,3,4,5,6,7,8,9,10,11,12,13,14,15]. The non-oxidative of ethanol dehydrogenation into acetaldehyde is regarded as a more promising synthetic route due to its excellent atomic utilization, high selectivity and ease of separation [9].
Copper-based catalysts have been widely recommended for ethanol dehydrogenation into acetaldehyde due to their good activity and relatively low price [9,10,11,12]. The supports can impact the Cu dispersion and valence states, thus adjusting the Cu properties and determining the catalytic performance. The commonly used supports include silica-based materials with relatively weak acidity [16,17,18,19,20,21,22], mixed metal oxides with both acidic and basic properties [23,24,25], and carbon materials with inert surface properties [26,27]. Shen and co-workers reported that copper catalysts supported on basic metal oxides (La2O2CO3 and MgO) could efficiently catalyze the transfer dehydrogenation of primary aliphatic alcohols (i.e., the reaction coupling between primary aliphatic alcohol dehydrogenation and styrene hydrogenation through hydrogen transfer) due to the synergistic effect of basic sites and Cu nanoparticles [28,29]. Consequently, it is envisaged that good performance could be achieved over an MgO-supported Cu catalyst for ethanol dehydrogenation into acetaldehyde.
Inspired by the abovementioned work by Shen’s group [28,29], in this paper, we have developed a new catalyst system, Cu supported on MgO, for the selective dehydrogenation of ethanol into acetaldehyde. Surprisingly, the prepared Cu/MgO catalysts display a superior catalytic performance. The catalytic results are related to the characterization ones and compared with the SiO2-supported Cu catalyst.

2. Results and Discussion

2.1. Structural and Textural Properties

The XRD patterns of a series of Cu/MgO catalysts with different Cu loadings are presented in Figure 1. The diffraction peaks at 2θ = 36.9°, 42.9°, 62.3°, 74.7° and 78.6° are assigned to the (111), (200), (220), (311) and (222) crystal facets of MgO (PDF #45-0946), while those at 2θ = 43.3°, 50.4° and 74.1° correspond to the Cu (111), Cu (200) and Cu (220) facets (PDF #04-0836). The absence of any diffraction peaks of Cu for the Cu/MgO catalysts with a Cu loading up to 15% implies that the Cu metal is highly dispersed on the MgO support. As the Cu loading is increased to 20%, the weak diffraction peak attributed to Cu crystallites was detected, suggesting that bulk Cu is formed at a high Cu loading. At the same Cu loading of 20%, the 20%Cu/SiO2 catalyst shows the obviously stronger intensity of the metallic Cu diffraction peaks than 20%Cu/MgO, indicative of a larger Cu crystallite size for the former catalyst than the latter. According to the Scherrer equation, the size of the Cu crystallites in 20%Cu/MgO and 20%Cu/SiO2 was estimated to be 11.4 and 24.4 nm, respectively. This result is an indication of the better dispersion of Cu on MgO than on SiO2, although the SiO2 support has a markedly higher surface area than MgO (381 vs. 95 m2/g), which signifies that the interaction between Cu and the MgO support is stronger than that between Cu and the SiO2 support.
As the Cu loading is increased from 2.5% to 20%, the BET surface area of the Cu/MgO catalysts drops from 63 to 45 m2/g, and the Cu metal dispersion decreases from 52.6% to 26.1% (Table 1). This result suggests that the size of the Cu crystallites increases with an increase in the Cu loading. Compared to 20%Cu/MgO, the 20%Cu/SiO2 catalyst possesses a much higher surface area (201 vs. 45 m2/g). Much higher Cu dispersion was achieved on 20%Cu/MgO than 20%Cu/SiO2 with the same Cu loading (26.1% vs. 3.6%), which stems from the stronger interaction between Cu species and MgO than between Cu species and SiO2. This result is consistent with the XRD observation.
The typical TEM images and corresponding elemental mapping of the 20%Cu/MgO and 20%Cu/SiO2 catalysts are shown in Figure 2 and Figure 3. Analysis of the TEM images reveals the formation of Cu nanoparticles on the MgO and SiO2 supports, with an average size of ca. 12 nm and 25 nm, respectively, which is consistent with the XRD result calculated using the Scherrer equation. Furthermore, according to the EDS elemental mapping, it can be observed that Cu is distributed uniformly over the MgO and SiO2 supports. A uniform distribution of Cu for the 2.5–15%Cu/MgO catalysts can be rationally expected, because these catalysts have a higher dispersion of surface Cu than the 20%Cu/MgO catalyst.

2.2. Basic Properties

CO2-TPD experiments were conducted to determine the surface basicity of the Cu/MgO catalysts. Figure 4 compares the CO2-TPD curves of the MgO support and the Cu/MgO catalysts with different Cu loadings. All the samples display a broad desorption peak in the temperature range of 80–400 °C. These broad peaks can be deconvoluted into the desorption of CO2 on weak (peak at ca. 140 °C), medium (peak at ca. 200 °C), and strong basic sites (peak at ca. 280 °C) [25,30,31], which correspond to OH groups, Mg2+-O2− pairs, and O2− ions, respectively [28,32]. As presented in Table 1, the Cu/MgO catalysts have more basic sites with medium strength. The number of basic sites present on the Cu/MgO catalysts decreases slightly with the increase in the Cu loading. Compared to MgO, the Cu/MgO catalysts obviously have less basic sites, indicating that the addition of Cu species reduces the number of basic sites on the MgO support.

2.3. XPS and TPR

XPS was used to analyze the chemical state of the Cu species in the 20%CuO/MgO and 20%CuO/SiO2 precursors as well as the reduced 20%Cu/MgO and 20%Cu/SiO2 catalysts. Figure 5 compares the Cu 2p XPS spectra of the samples. For the 20%CuO/MgO and 20%CuO/SiO2 samples, the binding energies of Cu 2p3/2 and 2p1/2 appear at 933.4 and 953.1 eV, respectively, which fit with those of Cu2+ in the CuO species [33,34,35]. Moreover, an accompanying Cu 2p satellite peak centered at around 943 eV can be observed, which is characteristic of the occurrence of Cu2+ species. After the reduction of 20%CuO/MgO and 20%CuO/SiO2 at 300 °C for 1 h under a flow of 10 vol% H2/Ar, the Cu 2p satellite peak disappeared, indicating the absence of Cu2+ species in the reduced 20%Cu/MgO and 20%Cu/SiO2 catalysts, i.e., Cu2+ species has been completely reduced. Moreover, the Cu 2p3/2 peak of the reduced catalysts is shifted to a lower binding energy (BE) of 932.3 eV, which is assigned to Cu0 and/or Cu+ species [35,36,37,38,39]. Considering that Cu0 and Cu+ species present close Cu 2p3/2 BE, in order to reveal the exact nature of the copper, we further performed Cu LMM X-ray-induced Auger electron spectroscopy (XAES) measurements. Figure 6 shows the Cu LMM XAES spectra of the reduced 20%Cu/MgO and 20%Cu/SiO2 catalysts. The signal can be deconvoluted into two symmetric peaks. The peak with a kinetic energy (KE) of 918.3 eV is assigned to metallic Cu, while the one with a KE of 914.0 eV is attributed to Cu+ species [17,18,40,41]. Based on the peak area percentages, the Cu0 and Cu+ species in the 20%Cu/MgO and 20%Cu/SiO2 catalysts were quantified to 81% and 19% as well as 88% and 12%, respectively. This observation indicates that both Cu0 and Cu+ species can be observed in the catalysts, and most Cu species are retained as Cu0. The 20%Cu/MgO catalyst possesses a lower population of Cu0 species than 20%Cu/SiO2 (81% vs. 88%). The presence of Cu+ species can be attributed to the strong interaction between Cu species and the MgO and SiO2 supports via the Cu−O−Mg and Cu−O−Si linkages [19,42,43].
To evaluate the reduction behavior of the Cu species dispersed on the MgO and SiO2 supports, the CuO/MgO and CuO/SiO2 precursors were examined by H2-TPR experiments. As shown in Figure 7, there is only one reduction peak for all the CuO/MgO samples, with the peak temperature ranging from 210 to 237 °C, which is ascribed to the reduction of highly dispersed CuO species similar to those reported for other Cu-based catalysts [17,40,44,45]. Differently, for the 20%CuO/SiO2 sample (Figure 7f), the TPR profile displays two peaks at 217 and 257 °C, which are assigned to the reduction of well-dispersed CuO with a small size and bulk CuO with a large size, respectively [18,46]. The reducibility of Cu species can be influenced by the particle size, support and preparation method. The TPR results suggest the better dispersion of CuO on MgO than on SiO2, thereby leading to the better dispersion of Cu on MgO than on SiO2 in the corresponding reduced Cu catalysts. These findings demonstrate the unique role of MgO in dispersing Cu nanoparticles.

2.4. Catalytic Performance

2.4.1. Effect of Cu Loading and Reaction Temperature

The catalytic performance of a series of Cu/MgO catalysts with different Cu loadings (2.5–20%) was tested in the temperature range of 190 to 290 °C to investigate the influence of the Cu loading and reaction temperature. As shown in Figure 8a, increasing the Cu loading from 2.5% to 15% leads to an increase in the ethanol conversion, which levels off with a further rise from 15% to 20% Cu. A higher Cu loading can provide more Cu sites for the reaction. The enhanced ethanol conversion on the Cu/MgO catalysts with Cu loading can be attributed to an improved number of Cu sites on the surface. As illustrated in Figure 9, the ethanol conversion on the Cu/MgO catalysts achieved at 230 °C is positively correlated with the number of surface Cu atoms.
At 230 °C, the productivity (expressed as the ethanol converted per gram of Cu per hour) drops from 26.8 to 4.8 gethanol/gCu/h as the dispersion of Cu particles on the MgO support is decreased from 52.6% to 26.1% (Figure 10). Small Cu particles (i.e., higher Cu dispersion) possess higher densities of coordinated unsaturated sites, such as corner and kink sites, which are more active than step or terrace sites [47]. Our finding is in accordance with the result for the Cu/SiO2 catalysts reported by Zhang et al. [17].
With the reaction temperature increased from 190 to 290 °C, the ethanol conversion on the 10%Cu/MgO, 15%Cu/MgO and 20%Cu/MgO catalysts improves progressively. However, the 2.5%Cu/MgO and 5%Cu/MgO catalysts show a trend in ethanol conversion with the temperature that first increases and then decreases (Figure 8a). A similar phenomenon was reported by Zhang et al. for ethanol dehydrogenation over Cu/SiO2 catalysts with 0.5% and 1% Cu loadings [17]. As shown in Figure 11, the ethanol conversion on 2.5%Cu/MgO at 270 °C diminishes from 64.8% to 44.6% after 1.5 h on stream. After regenerating the used catalyst by burning off the carbon species deposited on the catalyst in an air stream at 500 °C for 1 h followed by a reduction in the flow of 10 vol% H2/Ar at 300 °C for 1 h, the activity of the 2.5%Cu/MgO catalyst is fully restored. This observation indicates that carbon deposition is the cause of the deactivation. Since the catalytic activity is dependent on the surface Cu sites, the deactivation caused by carbon deposition will be more severe for the samples with low Cu loadings than those with high loadings [48]. The amount of coke (CHx) deposited on the catalyst after the reaction was quantified by TGA [17]. As presented in Table 1, obviously higher values are observed for 2.5%Cu/MgO and 5%Cu/MgO (0.107 and 0.054 g CHx/mmol Cu, respectively) than 10% and 15% Cu/MgO (0.021–0.025 CHx/mmol Cu) after several hours of the reaction, suggesting that the coke formation on Cu/MgO is pronounced at low Cu loadings. Thus, it can be concluded that the decrease in activity at higher temperatures for 2.5%Cu/MgO and 5%Cu/MgO can be attributed to the catalyst deactivation caused by carbon deposition.
The main reaction product is acetaldehyde, with the minor by-products being ethyl acetate, acetic acid, acetone, butyraldehyde and ethylene. The acetaldehyde selectivity for all the Cu/MgO catalysts is >97% at low temperatures, and it declines slightly with the temperature due to the generation of more ethyl acetate (Figure 8b). The selectivity to acetaldehyde can still reach higher than 95% at high temperatures.

2.4.2. Effect of WHSV

The dependence of the activity and selectivity on the space velocity was also investigated at 250 °C over 20%Cu/MgO as a representative catalyst. The results are shown in Figure 12. With the increment of the WHSV from 0.5 h−1 to 6 h−1, the conversion of ethanol obviously declines from 87.5% to 44.4%. The selectivity of acetaldehyde enhances from 94.7% at 0.5 h−1 to 97.7% at 1.5 h−1, followed by a slight increase to 98.2% at 2.5 h−1, and then levels off above 2.5 h−1.

2.4.3. Comparison of Performance for Cu/MgO with Cu/SiO2 in a Prolonged Time

The 20%Cu/MgO catalyst was chosen for comparison because it exhibits higher ethanol conversion than the other catalysts (2.5–15%Cu/MgO) due to its higher number of surface Cu sites. Considering that silica-supported Cu is a conventional catalyst for the non-oxidative dehydrogenation of ethanol, we compared the catalytic performance of the 20%Cu/MgO and 20%Cu/SiO2 catalysts with the same Cu loading at 250 °C over a prolonged reaction time. The results are presented in Figure 13. For the 20%Cu/MgO catalyst, as the reaction of ethanol dehydrogenation goes on, the conversion of ethanol diminishes from 81.5% at 1 h to 72.6% at 10 h and to 61.9% at 24 h, which decreases relatively by ca. 24%. After regeneration of the spent catalyst in flowing air at 500 °C for 1 h, the activity in the initial 10 h can be completely restored, followed by a fast deactivation where the ethanol conversion drops from ca. 73% at 10 h to ca. 52% at 24 h. A relative reduction in the ethanol conversion of ca. 36% is achieved in the second run. This finding suggests that the reason for the catalyst deactivation for 20%Cu/MgO in the first run (24 h) as well as in the initial 10 h of the second run is carbon deposition. The faster deactivation between 10 h and 24 h observed in the second run than in the first run is caused by sintering of the Cu particles. An increase in the intensity of the diffraction peaks of the Cu phase can be found for the used 20%Cu/MgO catalyst compared to the fresh one (Figure 14), signifying the growth of the Cu particles after the second run (from 11.4 to 15.0 nm, as determined by the Scherrer equation).
For the 20%Cu/SiO2 catalyst, the ethanol conversion decreases from 87.7% at 1 h to 52.8% at 24 h, corresponding to a relative reduction of ca. 40%, which is indicative of a faster catalyst deactivation for 20%Cu/SiO2 than 20%Cu/MgO. After pretreatment of the used catalyst in an air flow at 500 °C for 1 h, the ethanol conversion on the regenerated catalyst is 84.7% at 1 h, which is lower than the initial conversion, indicating that the original activity of 20%Cu/SiO2 cannot be fully restored. This result suggests that, in addition to carbon deposition, there is another reason for the catalyst deactivation in the first run of 20%Cu/SiO2, i.e., the growing of the Cu particles. The ethanol conversion on 20%Cu/SiO2 in the second run drops from 84.7% at 1 h to 47.0% at 24 h, which drops relatively by ca. 45%. The intensity of the diffraction peaks of metallic Cu on the spent 20%Cu/SiO2 catalyst is stronger than that on the fresh one (Figure 14), implying the aggregation of the Cu particles after reuse for 2 runs (from 24.4 to 33.0 nm, as determined by the Scherrer equation). Taking into account that less coke is formed on 20%Cu/SiO2 than 20%Cu/MgO (0.008 vs. 0.021 g CHx/mmol Cu), the quicker catalyst deactivation observed for the former catalyst than the latter one is attributed to the more severe Cu sintering for 20%Cu/SiO2 as a consequence of the weaker interfacial interaction between Cu species and the SiO2 support than between Cu species and the MgO support.
As the ethanol dehydrogenation reaction goes on, the acetaldehyde selectivity on 20%Cu/MgO increases slightly from 97.7% to 98.8%, whereas that on 20%Cu/SiO2 improves from 91.6% to 96.4%. Thus, both catalysts exhibit a comparable initial acetaldehyde yield (ca. 80%). More ethylene and ethyl acetate, which are derived from ethanol dehydration and the coupling between ethanol and acetaldehyde, respectively [36,49], are formed on 20%Cu/SiO2, leading to lower selectivity to acetaldehyde for 20%Cu/SiO2 than 20%Cu/MgO. The generation of more ethylene and ethyl acetate on 20%Cu/SiO2 might be related to the weak acid sites present on SiO2, i.e., surface—OH groups [26]. In summary, compared to 20%Cu/SiO2, the 20%Cu/MgO catalyst displays an equivalent initial acetaldehyde yield, higher acetaldehyde selectivity and longer stability.

3. Materials and Methods

3.1. Reagent and Materials

The Mg(NO3)2·6H2O, (NH4)2C2O4·H2O and ethanol were purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). The Cu(NO3)2·3H2O was purchased from Macklin Biochemical Co., Ltd. (Shanghai, China). The SiO2 support was purchased from Qingdao Haiyang Chemical Co., Ltd. (Qingdao, China). All the chemicals were of analytical grade and were used without further purification.

3.2. Catalyst Preparation

The MgO support was synthesized via a precipitation method using Mg(NO3)2·6H2O and (NH4)2C2O4·H2O as a precursor and a precipitating agent, respectively. In a typical synthetic procedure, 50 mL of 0.8 M Mg(NO3)2 aqueous solution was added dropwise to 200 mL of 0.3 M (NH4)2C2O4 aqueous solution under stirring, followed by aging at room temperature for 12 h. The obtained white precipitate was washed with deionized water, dried at 110 °C overnight, and then calcined at 600 °C in an air stream for 6 h. The Cu/MgO catalysts with various Cu loadings (2.5%, 5%, 10%, 15% and 20%, by weight) were prepared by a conventional impregnation method. Different amounts of Cu(NO3)2·3H2O were dissolved in deionized water, and then a certain amount of MgO was added. After drying under an infrared lamp, the samples were dried at 110 °C overnight, and then calcined at 550 °C in an air stream for 6 h to yield the CuO/MgO precursors. The CuO/MgO precursors were reduced at 300 °C for 1 h under a flow of 10 vol% H2/Ar to yield the x%Cu/MgO catalysts, where x% represents the weight percentage of Cu in the catalysts. Accordingly, the corresponding CuO/MgO precursors were labelled as x%CuO/MgO, where x% represents the weight percentage of Cu in the final Cu/MgO catalysts. For comparison, the 20%Cu/SiO2 catalyst was prepared in the same way but employing SiO2 as the support.

3.3. Catalyst Characterization

The X-ray diffraction (XRD) patterns were acquired on a D2 PHASER diffractometer (Brucker, Madison, WI, USA) using Cu Kα radiation at 30 kV and 10 mA. The BET surface areas were determined by N2 adsorption at −196 °C using a Micromeritics Tristar 3020 apparatus (Micromeritics, Atlanta, GA, USA). Before the measurement, the catalyst was pretreated at 300 °C under a vacuum for 5 h. The transmission electron microscopy (TEM) images and high-angle annular dark field scanning transmission electron microscopy (HAADF-STEM), together with the elemental mapping, were obtained on an FEI Tecnai G2 F20 S-TWIN microscope (FEI, Hillsboro, OR, USA). The X-ray photoelectron spectroscopy (XPS) measurements were carried out on a Perkin-Elmer PHI 5000C spectrometer (Perkin-Elmer, Waltham, MA, USA) using an Mg Kα radiation source. All the binding energy values were referenced to the C 1s peak at 284.6 eV. The thermogravimetric analysis (TGA) was conducted in an air stream on a Perkin-Elmer TGA8000 instrument (Perkin-Elmer, Waltham, MA, USA) for measuring the amount of coke deposited on the spent catalyst.
The basicity of the catalysts was determined via the temperature-programed desorption of CO2 (CO2-TPD) on an AutoChem II apparatus (Micromeritics, Atlanta, GA, USA). The catalyst (0.1 g, 40–60 mesh) was pretreated at 500 °C in He flow for 1 h. Afterwards, CO2 was adsorbed at 80 °C for 2 h under 5 vol% CO2/He flow (30 mL/min), followed by purging with He (30 mL/min) for 2 h. At last, the catalyst was heated to 500 °C with a ramp of 10 °C /min in an He flow (30 mL/min), and the desorption curve was recorded by a thermal conductivity detector. The H2 temperature-programed reduction (H2-TPR) characterization was carried out on a Micromeritics AutoChem II apparatus (Micromeritics, Atlanta, GA, USA). Before the TPR measurement, 0.1 g of the CuO/MgO precursor (40–60 mesh) was pretreated at 300 °C in an He flow for 1 h. After cooling to 60 °C in an He flow, the flow was changed to 10 vol% H2/Ar (30 mL/min), and then the temperature was raised to 500 °C with a ramp of 10 °C /min. The dispersion of Cu was determined by N2O oxidation and then by H2 titration [50]. In detail, 0.1 g of the CuO/MgO precursor (40–60 mesh) was first reduced at 300 °C for 1 h under a flow of 10 vol% H2/Ar (30 mL/min). After cooling to 60 °C in an He flow, the flow was switched to an N2O flow (30 mL/min) and maintained at 60 °C for 0.5 h to oxidize the surface Cu atoms to Cu2O, followed by flushing with He (30 mL/min) for 1 h to remove any residual N2O. Finally, another H2-TPR experiment was conducted as above, and the hydrogen consumption was recorded as Y. The hydrogen consumption in the first H2-TPR experiment was denoted as X. The Cu dispersion was calculated using the following equations:
Reduction of all the Cu atoms:
CuO + H2 = Cu + H2O, hydrogen consumption = X
Reduction of the surface Cu atoms only:
Cu2O + H2 = 2Cu + H2O, hydrogen consumption = Y
The dispersion of the surface Cu was calculated by the following formula:
D   ( % ) = 2 Y X × 100 %

3.4. Catalytic Test

Ethanol dehydrogenation was carried out in a flow fixed-bed microreactor under ambient pressure. The CuO/MgO precursor (0.3 g, 40–60 mesh) was in situ reduced at 300 °C for 1 h in a flow of 10 vol% H2/Ar. After the reactor was cooled down to the desired reaction temperature, the flow was changed to the feed gas consisting of 5 vol% ethanol and 95 vol% N2. Ethanol was introduced into the reactor using a bubbler. The products were analyzed online by a gas chromatograph equipped with a flame ionization detector and a FFAP capillary column (30 m × 0.32 mm × 0.25 μm). The conversion of ethanol, selectivity and yield of acetaldehyde were calculated according to the following formulas:
Conversion   ( % ) = [ ethanol ] in [ ethanol ] out [ ethanol ] in × 100 %
Selectivity   ( % ) = [ acetaldehyde ] out [ ethanol ] in   [ ethanol ] out × 100 %
Yield   ( % ) = [ acetaldehyde ] out [ ethanol ] in × 100 %

4. Conclusions

In this work, a series of Cu/MgO catalysts with different Cu loadings (2.5–20%) were prepared by an impregnation method. The as-synthesized Cu/MgO catalysts display a superior catalytic performance for the non-oxidative dehydrogenation of ethanol into acetaldehyde. The ethanol conversion is positively correlated with the number of surface Cu atoms. The productivity (expressed as the ethanol converted per gram of Cu per hour) is positively correlated with the dispersion of Cu particles on the MgO support. The 20%Cu/MgO catalyst affords an initial 81.5% conversion, with 97.7% acetaldehyde selectivity, at a WHSV of 1.5 h−1 and 250 °C. In comparison with 20%Cu/SiO2, the 20%Cu/MgO catalyst exhibits an equivalent initial acetaldehyde yield, higher acetaldehyde selectivity and longer stability. The faster catalyst deactivation observed for the former catalyst than the latter one is attributed to the more severe Cu sintering for 20%Cu/SiO2 as a consequence of the weaker interfacial interaction between Cu species and the SiO2 support than between Cu species and the MgO support.

Author Contributions

Conceptualization, W.H.; methodology, Y.Y. and C.M.; validation, Z.G.; formal analysis, Y.Y., C.M. and Z.G.; investigation, C.T.; data curation, C.T.; writing—original draft preparation, C.T.; writing—review and editing, W.H.; supervision, W.H.; project administration, W.H.; funding acquisition, W.H. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China, grant number 22072027, the Science and Technology Commission of Shanghai Municipality, grant number 19DZ2270100 and the SINOPEC Shanghai Research Institute of Petrochemical Technology Co., Ltd., grant number 33750000-19-ZC0607-0005.

Data Availability Statement

The data presented in this study are available in the article.

Conflicts of Interest

The authors declare that this study received funding from SINOPEC Shanghai Research Institute of Petrochemical Technology Co., Ltd. The funder was not involved in the study design, collection, analysis, interpretation of data, the writing of this article or the decision to submit it for publication.

References

  1. Sun, J.; Wang, Y. Recent advances in catalytic conversion of ethanol to chemicals. ACS Catal. 2014, 4, 1078–1090. [Google Scholar] [CrossRef]
  2. Li, X.; Kant, A.; He, Y.; Thakkar, H.V.; Atanga, M.A.; Rezaei, F.; Ludlow, D.K.; Rownaghi, A.A. Light olefins from renewable resources: Selective catalytic dehydration of bioethanol to propylene over zeolite and transition metal oxide catalysts. Catal. Today 2016, 276, 62–77. [Google Scholar] [CrossRef]
  3. Iwamoto, M. Selective catalytic conversion of bio-ethanol to propene: A review of catalysts and reaction pathways. Catal. Today 2015, 242, 243–248. [Google Scholar] [CrossRef]
  4. Pomalaza, G.; Capron, M.; Ordomsky, V.; Dumeignil, F. Recent breakthroughs in the conversion of ethanol to butadiene. Catalysts 2016, 6, 203. [Google Scholar] [CrossRef]
  5. Wu, X.; Fang, G.; Tong, Y.; Jiang, D.; Liang, Z.; Leng, W.; Liu, L.; Tu, P.; Wang, H.; Ni, J.; et al. Catalytic upgrading of ethanol to n-butanol: Progress in catalyst development. ChemSusChem 2018, 11, 71–85. [Google Scholar] [CrossRef]
  6. Masih, D.; Rohani, S.; Kondo, J.; Tatsumi, T. Catalytic dehydration of ethanol-to-ethylene over Rho zeolite under mild reaction conditions. Microporous Mesoporous Mater. 2019, 282, 91–99. [Google Scholar] [CrossRef]
  7. Wang, Q.; Weng, X.; Zhou, B.; Lv, S.; Miao, S.; Zhang, D.; Han, Y.; Scott, S.L.; Schüth, F.; Lu, A. Direct, selective production of aromatic alcohols from ethanol using a tailored bifunctional cobalt–hydroxyapatite catalyst. ACS Catal. 2019, 9, 7204–7216. [Google Scholar] [CrossRef]
  8. Li, X.; Pang, J.; Wang, C.; Li, L.; Pan, X.; Zheng, M.; Zhang, T. Conversion of ethanol to 1,3-butadiene over high-performance Mg-ZrOx/MFI nanosheet catalysts via the two-step method. Green Chem. 2020, 22, 2852–2861. [Google Scholar] [CrossRef]
  9. Pang, J.; Yin, M.; Wu, P.; Li, X.; Li, H.; Zheng, M.; Zhang, T. Advances in catalytic dehydrogenation of ethanol to acetaldehyde. Green Chem. 2021, 23, 7902–7916. [Google Scholar] [CrossRef]
  10. Huang, Y.; Wang, B.; Yuan, H.; Sun, Y.; Yang, D.; Cui, X.; Shi, F. The catalytic dehydrogenation of ethanol by heterogeneous catalysts. Catal. Sci. Technol. 2021, 11, 1652–1664. [Google Scholar] [CrossRef]
  11. Phung, T.K. Copper-based catalysts for ethanol dehydrogenation and dehydrogenative coupling into hydrogen, acetaldehyde and ethyl acetate. Int. J. Hydrogen Energy 2022, 47, 42234–42249. [Google Scholar] [CrossRef]
  12. He, L.; Zhou, B.; Sun, D.; Li, W.; Lv, W.; Wang, J.; Liang, Y.; Lu, A. Catalytic conversion of ethanol to oxygen-containing value-added chemicals. ACS Catal. 2023, 13, 11291–11304. [Google Scholar] [CrossRef]
  13. Jin, H.; Ding, X.; Tian, C.; Yue, Y.; Hua, W.; Gao, Z. Insights into promoting effect of Sm on catalytic performance of the CeO2/Beta catalyst in direct conversion of bioethanol to propylene. Chem. Res. Chin. Univ. 2022, 38, 1547–1552. [Google Scholar] [CrossRef]
  14. Jin, H.; Miao, C.; Yue, Y.; Tian, C.; Hua, W.; Gao, Z. Sm-CeO2/zeolite bifunctional catalyst for direct and highly selective conversion of bioethanol to propylene. Catalysts 2022, 12, 407. [Google Scholar] [CrossRef]
  15. Jin, H.; Yue, Y.; Miao, C.; Tian, C.; Hua, W.; Gao, Z. Direct and highly selective conversion of bioethanol to propylene over Y-CeO2 and zeolite Beta composite. Catal. Lett. 2023, 153, 230–238. [Google Scholar] [CrossRef]
  16. Sato, A.G.; Volanti, D.P.; de Freitas, I.C.; Longo, E.; Bueno, J.M.C. Site-selective ethanol conversion over supported copper catalysts. Catal. Commun. 2012, 26, 122–126. [Google Scholar] [CrossRef]
  17. Zhang, H.; Tan, H.R.; Jaenicke, S.; Chuah, G.K. Highly efficient and robust Cu catalyst for non-oxidative dehydrogenation of ethanol to acetaldehyde and hydrogen. J. Catal. 2020, 389, 19–28. [Google Scholar] [CrossRef]
  18. Pang, J.; Zheng, M.; Wang, C.; Yang, X.; Liu, H.; Liu, X.; Sun, J.; Wang, Y.; Zhang, T. Hierarchical echinus-like Cu-MFI catalysts for ethanol dehydrogenation. ACS Catal. 2020, 10, 13624–13629. [Google Scholar] [CrossRef]
  19. Pang, J.; Yin, M.; Wu, P.; Song, L.; Li, X.; Zhang, T.; Zheng, M. Redispersion and stabilization of Cu/MFI catalysts by the encapsulation method for ethanol dehydrogenation. ACS Sustain. Chem. Eng. 2023, 11, 3297–3305. [Google Scholar] [CrossRef]
  20. Finger, P.H.; Osmari, T.A.; Cabral, N.M.; Bueno, J.M.C.; Marcel, J.; Gallo, R. Direct synthesis of Cu supported on mesoporous silica: Tailoring the Cu loading and the activity for ethanol dehydrogenation. Catal. Today 2021, 381, 26–33. [Google Scholar] [CrossRef]
  21. Lin, L.; Cao, P.; Pang, J.; Wang, Z.; Jiang, Q.; Su, Y.; Chen, R.; Wu, Z.; Zheng, M.; Luo, W. Zeolite-encapsulated Cu nanoparticles with enhanced performance for ethanol dehydrogenation. J. Catal. 2022, 413, 565–574. [Google Scholar] [CrossRef]
  22. Liu, H.; Chang, Z.; Fu, J.; Hou, Z. A CuZn-BTC derived stable Cu/ZnO@SiO2 catalyst for ethanol dehydrogenation. Appl. Catal. B 2023, 324, 122194. [Google Scholar] [CrossRef]
  23. Campisano, I.S.P.; Rodella, C.B.; Sousa, Z.S.B.; Henriques, C.A.; da Silva, V.T. Influence of thermal treatment conditions on the characteristics of Cu-based metal oxides derived from hydrotalcite-like compounds and their performance in bio-ethanol dehydrogenation to acetaldehyde. Catal. Today 2018, 306, 111–120. [Google Scholar] [CrossRef]
  24. Santos, R.M.M.; Briois, V.; Martins, L.; Santilli, C.V. Insights into the preparation of copper catalysts supported on layered double hydroxide derived mixed oxides for ethanol dehydrogenation. ACS Appl. Mater. Interfaces 2021, 13, 26001–26012. [Google Scholar] [CrossRef] [PubMed]
  25. Liu, H.; Jiang, Y.; Zhou, R.; Chang, Z.; Hou, Z. Co-production of hydrogen and acetaldehyde from ethanol over a highly dispersed Cu catalyst. Fuel 2022, 321, 123980. [Google Scholar] [CrossRef]
  26. Wang, Q.; Shi, L.; Lu, A. Highly selective copper catalyst supported on mesoporous carbon for the dehydrogenation of ethanol to acetaldehyde. ChemCatChem 2015, 7, 2846–2852. [Google Scholar] [CrossRef]
  27. Klinthongchai, Y.; Prichanont, S.; Praserthdam, P.; Jongsomjit, B. Synthesis, characteristics and application of mesocellular foam carbon (MCF-C) as catalyst for dehydrogenation of ethanol to acetaldehyde. J. Environ. Chem. Eng. 2020, 8, 103752. [Google Scholar] [CrossRef]
  28. Shi, R.; Wang, F.; Mu, X.; Ta, N.; Li, Y.; Huang, X.; Shen, W. Transfer dehydrogenation of 1-octanol to 1-octanal over Cu/MgO catalyst: Effect of Cu particle size. Chin. J. Catal. 2010, 31, 626–630. [Google Scholar] [CrossRef]
  29. Wang, F.; Shi, R.; Liu, Z. Highly efficient dehydrogenation of primary aliphatic alcohols catalyzed by Cu nanoparticles dispersed on rod-shaped La2O2CO3. ACS Catal. 2013, 3, 890–894. [Google Scholar] [CrossRef]
  30. Xia, S.; Nie, R.; Lu, X.; Wang, L.; Chen, P.; Hou, Z. Hydrogenolysis of glycerol over Cu0.4/Zn5.6−xMgxAl2O8.6 catalysts: The role of basicity and hydrogen spillover. J. Catal. 2012, 296, 1–11. [Google Scholar] [CrossRef]
  31. Hincapié, G.; López, D.; Moreno, A. Infrared analysis of methanol adsorption on mixed oxides derived from Mg/Al hydrotalcite catalysts for transesterification reactions. Catal. Today 2018, 302, 277–285. [Google Scholar] [CrossRef]
  32. Di Cosimo, J.I.; Díez, V.K.; Xu, M.; Iglesia, E.; Apesteguía, C.R. Structure and surface and catalytic properties of Mg-Al basic oxides. J. Catal. 1998, 178, 499–510. [Google Scholar] [CrossRef]
  33. Gervasini, A.; Manzoli, M.; Martra, G.; Ponti, A.; Ravasio, N.; Sordelli, L.; Zaccheria, F. Dependence of copper species on the nature of the support for dispersed CuO catalysts. J. Phys. Chem. B 2006, 110, 7851–7861. [Google Scholar] [CrossRef] [PubMed]
  34. Gong, J.; Yue, H.; Zhao, Y.; Zhao, S.; Zhao, L.; Lv, J.; Shengping Wang, S.; Ma, X. Synthesis of ethanol via syngas on Cu/SiO2 catalysts with balanced Cu0–Cu+ sites. J. Am. Chem. Soc. 2012, 134, 13922–13925. [Google Scholar] [CrossRef]
  35. Amokrane, S.; Boualouache1, A.; Simon, P.; Capron, M.; Otmanine1, G.; Allam, D.; Hocine, S. Effect of adding transition metals to copper on the dehydrogenation reaction of ethanol. Catal. Lett. 2021, 151, 2864–2883. [Google Scholar] [CrossRef]
  36. Sato, A.G.; Volanti, D.P.; Meira, D.M.; Damyanova, S.; Longo, E.; Bueno, J.M.C. Effect of the ZrO2 phase on the structure and behavior of supported Cu catalysts for ethanol conversion. J. Catal. 2013, 307, 1–17. [Google Scholar] [CrossRef]
  37. Wang, Q.N.; Shi, L.; Li, W.; Li, W.C.; Si, R.; Schüth, F.; Lu, A.H. Cu supported on thin carbon layer-coated porous SiO2 for efficient ethanol dehydrogenation. Catal. Sci. Technol. 2018, 8, 472–479. [Google Scholar] [CrossRef]
  38. Cheng, S.Q.; Weng, X.F.; Wang, Q.N.; Zhou, B.C.; Li, W.C.; Li, M.R.; He, L.; Wang, D.Q.; Lu, A.H. Defect-rich BN-supported Cu with superior dispersion for ethanol conversion to aldehyde and hydrogen. Chin. J. Catal. 2022, 43, 1092–1100. [Google Scholar] [CrossRef]
  39. Hao, Y.; Zhao, D.; Liu, W.; Zhang, M.; Lou, Y.; Wang, Z.; Tang, Q.; Yang, J. Uniformly dispersed Cu nanoparticles over mesoporous silica as a highly selective and recyclable ethanol dehydrogenation catalyst. Catalysts 2022, 12, 1049. [Google Scholar] [CrossRef]
  40. Chen, L.F.; Guo, P.J.; Qiao, M.H.; Yan, S.R.; Li, H.X.; Shen, W.; Xu, H.L.; Fan, K.N. Cu/SiO2 catalysts prepared by the ammonia-evaporation method: Texture, structure, and catalytic performance in hydrogenation of dimethyl oxalate to ethylene glycol. J. Catal. 2008, 257, 172–180. [Google Scholar] [CrossRef]
  41. Yang, H.; Chen, Y.; Cui, X.; Wang, G.; Cen, Y.; Deng, T.; Yan, W.; Gao, J.; Zhu, S.; Olsbye, U.; et al. A highly stable copper-based catalyst for clarifying the catalytic roles of Cu0 and Cu+ species in methanol dehydrogenation. Angew. Chem. Int. Ed. 2018, 57, 1836–1840. [Google Scholar] [CrossRef] [PubMed]
  42. Xu, C.; Chen, G.; Zhao, Y.; Liu, P.; Duan, X.; Gu, L.; Fu, G.; Yuan, Y.; Zheng, N. Interfacing with silica boosts the catalysis of copper. Nat. Commun. 2018, 9, 3367. [Google Scholar] [CrossRef]
  43. Ding, Y.; Tian, J.; Chen, W.; Guan, Y.; Xu, H.; Li, X.; Wu, H.; Wu, P. One-pot synthesized core/shell structured zeolite@copper catalysts for selective hydrogenation of ethylene carbonate to methanol and ethylene glycol. Green Chem. 2019, 21, 5414–5426. [Google Scholar] [CrossRef]
  44. Tu, Y.J.; Chen, Y.W. Effects of alkali metal oxide additives on Cu/SiO2 catalyst in the dehydrogenation of ethanol. Ind. Eng. Chem. Res. 2001, 40, 5889–5893. [Google Scholar] [CrossRef]
  45. Dong, X.; Ma, X.; Xu, H.; Ge, Q. Comparative study of silica-supported copper catalysts prepared by different methods: Formation and transition of copper phyllosilicate. Catal. Sci. Technol. 2016, 6, 4151–4158. [Google Scholar] [CrossRef]
  46. Huang, Z.; Cui, F.; Kang, H.; Chen, J.; Zhang, X.; Xia, C. Highly dispersed silica-supported copper nanoparticles prepared by precipitation−gel method: A simple but efficient and stable catalyst for glycerol hydrogenolysis. Chem. Mater. 2008, 20, 5090–5099. [Google Scholar] [CrossRef]
  47. Shimizu, K.I.; Sugino, K.; Sawabe, K.; Satsuma, A. Oxidant-free dehydrogenation of alcohols heterogeneously catalyzed by cooperation of silver clusters and acid–base sites on alumina. Chem. Eur. J. 2009, 15, 2341–2351. [Google Scholar] [CrossRef]
  48. Han, Y.; Shen, J.; Chen, Y. Microkinetic analysis of isopropanol dehydrogenation over Cu/SiO2 catalyst. Appl. Catal. A 2001, 205, 79–84. [Google Scholar] [CrossRef]
  49. Freitas, I.C.; Damyanova, S.; Oliveira, D.C.; Marques, C.M.P.; Bueno, J.M.C. Effect of Cu content on the surface and catalytic properties of Cu/ZrO2 catalyst for ethanol dehydrogenation. J. Mol. Catal. A 2014, 381, 26–37. [Google Scholar] [CrossRef]
  50. Sato, S.; Takahashi, R.; Sodesawa, T.; Yuma, K.I.; Obata, Y. Distinction between surface and bulk oxidation of Cu through N2O decomposition. J. Catal. 2000, 196, 195–199. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of the catalysts: (a) 2.5%Cu/MgO, (b) 5%Cu/MgO, (c) 10%Cu/MgO, (d) 15%Cu/MgO and (e) 20%Cu/MgO and (f) 20%Cu/SiO2.
Figure 1. XRD patterns of the catalysts: (a) 2.5%Cu/MgO, (b) 5%Cu/MgO, (c) 10%Cu/MgO, (d) 15%Cu/MgO and (e) 20%Cu/MgO and (f) 20%Cu/SiO2.
Catalysts 14 00541 g001
Figure 2. (a) TEM image and (b) HAADF STEM mapping images of the 20%Cu/MgO catalyst.
Figure 2. (a) TEM image and (b) HAADF STEM mapping images of the 20%Cu/MgO catalyst.
Catalysts 14 00541 g002
Figure 3. (a) TEM image and (b) HAADF STEM mapping images of the 20%Cu/SiO2 catalyst.
Figure 3. (a) TEM image and (b) HAADF STEM mapping images of the 20%Cu/SiO2 catalyst.
Catalysts 14 00541 g003
Figure 4. CO2-TPD profiles of the MgO support and Cu/MgO catalysts: (a) MgO, (b) 2.5%Cu/MgO, (c) 5%Cu/MgO, (d) 10%Cu/MgO, (e) 15%Cu/MgO and (f) 20%Cu/MgO.
Figure 4. CO2-TPD profiles of the MgO support and Cu/MgO catalysts: (a) MgO, (b) 2.5%Cu/MgO, (c) 5%Cu/MgO, (d) 10%Cu/MgO, (e) 15%Cu/MgO and (f) 20%Cu/MgO.
Catalysts 14 00541 g004
Figure 5. Cu 2p XPS spectra of the 20%CuO/MgO and 20%CuO/SiO2 precursors as well as the reduced 20%Cu/MgO and 20%Cu/SiO2 catalysts.
Figure 5. Cu 2p XPS spectra of the 20%CuO/MgO and 20%CuO/SiO2 precursors as well as the reduced 20%Cu/MgO and 20%Cu/SiO2 catalysts.
Catalysts 14 00541 g005
Figure 6. Cu LMM XAES spectra of the 20%Cu/MgO and 20%Cu/SiO2 catalysts.
Figure 6. Cu LMM XAES spectra of the 20%Cu/MgO and 20%Cu/SiO2 catalysts.
Catalysts 14 00541 g006
Figure 7. H2-TPR profiles of the CuO/MgO and CuO/SiO2 precursors: (a) 2.5%CuO/MgO, (b) 5%CuO/MgO, (c) 10%CuO/MgO, (d) 15%CuO/MgO, (e) 20%CuO/MgO and (f) 20%CuO/SiO2.
Figure 7. H2-TPR profiles of the CuO/MgO and CuO/SiO2 precursors: (a) 2.5%CuO/MgO, (b) 5%CuO/MgO, (c) 10%CuO/MgO, (d) 15%CuO/MgO, (e) 20%CuO/MgO and (f) 20%CuO/SiO2.
Catalysts 14 00541 g007
Figure 8. (a) Effect of the reaction temperature on the ethanol conversion and (b) selectivity to acetaldehyde over Cu/MgO catalysts with different Cu loadings (2.5–20%). Reaction condition: WHSV = 1.5 h−1, reaction from 190 to 290 °C and holding at each temperature for 1 h. Data were obtained after 1 h of the reaction.
Figure 8. (a) Effect of the reaction temperature on the ethanol conversion and (b) selectivity to acetaldehyde over Cu/MgO catalysts with different Cu loadings (2.5–20%). Reaction condition: WHSV = 1.5 h−1, reaction from 190 to 290 °C and holding at each temperature for 1 h. Data were obtained after 1 h of the reaction.
Catalysts 14 00541 g008
Figure 9. Relationship between the ethanol conversion obtained at 230 °C and the number of surface Cu atoms on the Cu/MgO catalysts. Data were obtained after 1 h of the reaction.
Figure 9. Relationship between the ethanol conversion obtained at 230 °C and the number of surface Cu atoms on the Cu/MgO catalysts. Data were obtained after 1 h of the reaction.
Catalysts 14 00541 g009
Figure 10. Relationship between the productivity obtained at 230 °C and the dispersion of Cu particles on the MgO support. Data were obtained after 1 h of the reaction.
Figure 10. Relationship between the productivity obtained at 230 °C and the dispersion of Cu particles on the MgO support. Data were obtained after 1 h of the reaction.
Catalysts 14 00541 g010
Figure 11. Reuse of the 2.5%Cu/MgO catalyst for 2 runs. Regeneration condition: oxidation at 500 °C for 1 h under an air stream followed by reduction at 300 °C for 1 h under a flow of 10 vol% H2/Ar. Reaction condition: 270 °C, WHSV = 1.5 h−1.
Figure 11. Reuse of the 2.5%Cu/MgO catalyst for 2 runs. Regeneration condition: oxidation at 500 °C for 1 h under an air stream followed by reduction at 300 °C for 1 h under a flow of 10 vol% H2/Ar. Reaction condition: 270 °C, WHSV = 1.5 h−1.
Catalysts 14 00541 g011
Figure 12. Effect of the WHSV on the ethanol conversion and acetaldehyde selectivity over 20%Cu/MgO. Reaction condition: 250 °C, after 1 h of the reaction at each WHSV.
Figure 12. Effect of the WHSV on the ethanol conversion and acetaldehyde selectivity over 20%Cu/MgO. Reaction condition: 250 °C, after 1 h of the reaction at each WHSV.
Catalysts 14 00541 g012
Figure 13. Catalytic performance of the (a) 20%Cu/MgO and (b) 20%Cu/SiO2 catalysts over a prolonged period of time for 2 runs’ reuse. Regeneration condition: oxidation at 500 °C for 1 h under an air stream followed by reduction at 300 °C for 1 h under a flow of 10 vol% H2/Ar. Reaction condition: 250 °C, WHSV = 1.5 h−1.
Figure 13. Catalytic performance of the (a) 20%Cu/MgO and (b) 20%Cu/SiO2 catalysts over a prolonged period of time for 2 runs’ reuse. Regeneration condition: oxidation at 500 °C for 1 h under an air stream followed by reduction at 300 °C for 1 h under a flow of 10 vol% H2/Ar. Reaction condition: 250 °C, WHSV = 1.5 h−1.
Catalysts 14 00541 g013
Figure 14. XRD patterns of the fresh and used 20%Cu/MgO and 20%Cu/SiO2 catalysts after reuse for 2 runs, as depicted in Figure 13.
Figure 14. XRD patterns of the fresh and used 20%Cu/MgO and 20%Cu/SiO2 catalysts after reuse for 2 runs, as depicted in Figure 13.
Catalysts 14 00541 g014
Table 1. Physicochemical properties of the different catalysts.
Table 1. Physicochemical properties of the different catalysts.
CatalystSBETBasicity (mmol/g)DCu bCoke c
(m2/g)WeakMediumStrongTotal(%)(g CHx/mmol Cu)
MgO950.0280.0980.0270.153
2.5%Cu/MgO630.0260.0780.0130.11752.60.107
5%Cu/MgO610.0260.0700.0140.11044.80.054
10%Cu/MgO590.0280.0680.0130.10934.40.025
15%Cu/MgO550.0280.0640.0130.10529.20.021
20%Cu/MgO450.0270.0630.0130.10326.10.021
20%Cu/SiO2201 a0.0130.0170.0060.0363.60.008
a The BET surface area of the SiO2 support is 381 m2/g; b Cu dispersion; c for 2.5%, 5%, 10% and 15% Cu/MgO, reaction from 190 to 290 °C and holding at each temperature for 1 h. For 20%Cu/MgO and 20%Cu/SiO2, reaction was performed for 48 h at 250 °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

Tian, C.; Yue, Y.; Miao, C.; Hua, W.; Gao, Z. Cu/MgO as an Efficient New Catalyst for the Non-Oxidative Dehydrogenation of Ethanol into Acetaldehyde. Catalysts 2024, 14, 541. https://doi.org/10.3390/catal14080541

AMA Style

Tian C, Yue Y, Miao C, Hua W, Gao Z. Cu/MgO as an Efficient New Catalyst for the Non-Oxidative Dehydrogenation of Ethanol into Acetaldehyde. Catalysts. 2024; 14(8):541. https://doi.org/10.3390/catal14080541

Chicago/Turabian Style

Tian, Chao, Yinghong Yue, Changxi Miao, Weiming Hua, and Zi Gao. 2024. "Cu/MgO as an Efficient New Catalyst for the Non-Oxidative Dehydrogenation of Ethanol into Acetaldehyde" Catalysts 14, no. 8: 541. https://doi.org/10.3390/catal14080541

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