Next Article in Journal
The Influence of Clustered DNA Damage Containing Iz/Oz and OXOdG on the Charge Transfer through the Double Helix: A Theoretical Study
Previous Article in Journal
Efficient Preparation of Biodiesel Using Sulfonated Camellia oleifera Shell Biochar as a Catalyst
Previous Article in Special Issue
Dry Reforming of Methane over Pyrochlore-Type La2Ce2O7-Supported Ni Catalyst: Effect of Particle Size of Support
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Calcination Atmosphere on the Performance of Cu/Al2O3 Catalyst for the Selective Hydrogenation of Furfural to Furfuryl Alcohol

by
Yongzhen Gao
1,
Wenjing Yi
2,
Jingyi Yang
2,
Kai Jiang
1,
Tao Yang
1,
Zhihan Li
1,
Meng Zhang
1,*,
Zhongyi Liu
1,3,* and
Benlai Wu
1,*
1
College of Chemistry, Zhengzhou University, Zhengzhou 450001, China
2
School of Chemical Engineering, Zhengzhou University, Zhengzhou 450001, China
3
State Key Laboratory of Coking Coal Resources Green Exploitation, Zhengzhou University, Zhengzhou 450001, China
*
Authors to whom correspondence should be addressed.
Molecules 2024, 29(12), 2753; https://doi.org/10.3390/molecules29122753
Submission received: 17 May 2024 / Revised: 5 June 2024 / Accepted: 7 June 2024 / Published: 9 June 2024
(This article belongs to the Special Issue New Strategies for Metal Catalysis in Heterogeneous System)

Abstract

:
The selective hydrogenation of the biomass platform molecule furfural (FAL) to produce furfuryl alcohol (FA) is of great significance to alleviate the energy crisis. Cu-based catalysts are the most commonly used catalysts, and their catalytic performance can be optimized by changing the preparation method. This paper emphasized the effect of calcination atmosphere on the performance of a Cu/Al2O3 catalyst for the selective hydrogenation of FAL. The precursor of the Cu/Al2O3 catalyst prepared by the ammonia evaporation method was treated with different calcination atmospheres (N2 and air). On the basis of the combined results from the characterizations using in situ XRD, TEM, N2O titration, H2-TPR and XPS, the Cu/Al2O3 catalyst calcined in the N2 atmosphere was more favorable for the dispersion and reduction of Cu species and the reduction process could produce more Cu+ and Cu0 species, which facilitated the selective hydrogenation of FAL to FA. The experimental results showed that the N2 calcination atmosphere improved the FAL conversion and FA selectivity, and the FAL conversion was further increased after reduction. Cu/Al2O3-N2-R exhibited the outstanding performance, with a high yield of 99.9% of FA after 2 h at 120 °C and an H2 pressure of 1 MPa. This work provides a simple, efficient and economic method to improve the C=O hydrogenation performance of Cu-based catalysts.

Graphical Abstract

1. Introduction

The energy supply crisis and environmental pollution caused by the excessive consumption of the fossil energy have aroused worldwide concerns [1]. In order to achieve the sustainable development, researchers have spared no efforts to explore and develop new renewable energy as the alternative, which has included solar energy, hydroelectric energy, wind energy, biomass energy, geothermal energy, tidal energy and so on [2,3]. Among them, biomass energy is the only carbon-containing energy and is the fourth largest category after the traditional fossil energy of coal, oil and natural gas [4]. It occupies a very important position in the overall energy system. Compared with other energy products, biomass energy has the advantages of abundant reserves, strong renewable ability and wide distribution and is an ideal substitute for fossil energy [2,5].
Furfural (FAL) is an important biomass platform compound and considered to be one of the most promising platform molecules for sustainable production of fuels and chemicals in the 21st century [6,7]. FAL has a variety of functional groups such as furan ring, aldehyde group and diene ether, and it can generate various derivatives through selective hydrogenation reactions [8,9], as shown in Figure 1. Among these products, furfuryl alcohol (FA), as one of the important value-added intermediates, accounts for about 65% of the total output of FAL derivatives [10]. FA is mainly used in the production of resins [8], plasticizers [11], wood preservatives [12] and as a platform molecule for the synthesis of other chemicals [13]. Due to the complexity of the FAL reaction pathway, obtaining high FAL conversion and high FA selectivity simultaneously remains a great challenge [11].
The CuCr oxide catalyst is mainly used in industry to catalyze FAL hydrogenation to produce FA [14]. However, the high conversion and remarkable yield require high temperature and high pressure, and the employed catalysts also produce the toxic Cr-containing waste for the environment [15]. In recent years, developing the Cr-free catalysts with superior performance as the alternatives has been an interesting topic [16]. In the beginning, most of the catalysts used to catalyze the FAL hydrogenation conversion were noble metals, including Pd [17], Pt [18] and Ru [19]. Although the noble metal catalysts have the high intrinsic hydrogenation activity, the low selectivity for FA and high price limit their large-scale industrial application. Recently, non-noble metals, including Fe [20], Co [21], Ni [22] and Cu [23], have attracted wide interest because of their abundant reserves, low price and simple availability. As long as the reactivity, selectivity and stability can be effectively improved, they can be the promising alternatives for the industrial applications [14]. Among the non-noble metals, Cu-based catalysts have received much attention due to their specific hydrogenation activity towards C=O [24]. Theoretical calculations and experimental results have well demonstrated that FAL is linear chemical adsorbed in a η1(O) configuration on the active Cu species [14]. This avoids the hydrogenation of the furan ring, resulting in high FA selectivity. However, the low reactivity of Cu-based catalysts urgently needs to be improved [25].
Many strategies, including but not limited to improving the active components [26,27], modulating the supports [28,29], doping the promotors [30,31] and tuning the preparation methods [32,33], have been used to develop the novel Cu-based catalysts. Zhang et al. [26] obtained the highly distributed Cu/MgO catalysts, where the Cu0/Cu+ molar ratio increased gradually with the upward reduction temperature. The experimental results showed that the Cu/MgO-350 exhibited excellent stability and reusability without significant activity loss. The reduction temperature could modulate the nature of the interface between the Cu0/Cu+ active species and the metal oxide interface in Cu/MgO catalysts. Gong et al. [34] reported a copper-based catalyst supported by sulfonate group (-SO3H)-grafted active carbon (AC). The modified Cu/AC-SO3H catalyst exhibited an enhanced catalytic performance for selective hydrogenation of FAL to FA in the liquid phase. Through grafting the sulfonate group on the support, better metal dispersion, more Cu0/Cu+ species and stronger FAL adsorption capacity were attained to contribute the high hydrogenation performance. Zhang et al. [30] synthesized a Pd-CuOx nanocomposite catalyst with outstanding performance for the selective hydrogenation of FAL to FA. The valence state of Cu and Pd-Cu interactions played the critical roles in determining the intrinsic activity of the prepared Pd-Cu catalysts. Various characterizations combined with the kinetic experiments and in situ chemisorption clearly unraveled the adsorption and activation processes of the C=O bond and H2 molecule on Pd0, Cu0 and Cu+ sites. Zhang et al. [35] provided an efficient strategy of K2CO3 assisting the CuO#TiO2 catalyst for the liquid-phase hydrogenation of FAL. The reason for the increased yield was ascribed to the boosted generation of the surface Cu+/Cu0 active sites and the promoted gaseous hydrogen dissolution in ethanol media by K2CO3. Cheng et al. [36] developed a new catalytic transfer hydrogenation (CTH) strategy for the continuous conversion of FAL into FA with Cu/ZnO/Al2O3 as the catalyst under mild conditions. The highly dispersive Cu species (Cu0, Cu+), with few Lewis acidic sites and η1(O)-type adsorption, ensured high selectivity and mild conversion. Du et al. [32] investigated the effect of preparation methods on the structure and performance of Cu/SiO2 catalysts. The results showed that the excellent performance was associated with the highest Cu0 surface area, the smallest Cu particle size and the suitable Cu+/(Cu+ + Cu0) ratio. The above studies have confirmed that the efficient Cu-based catalysts are closely related to the abundance of Cu+ and Cu0 species. Cu0 and Cu+ active species synergistically catalyze C=O bond hydrogenation, where Cu0 promotes the dissociation of H2, while Cu+ adsorbs and activates C=O bonds as the Lewis acid sites [37].
In addition to the above methods, the calcination atmosphere also has great influences on the size of the active metal [38], the metal–support interaction [39] and the surface active species [40], which could also affect the catalytic performance [41]. However, the effect of the calcination atmosphere on the activity of Cu-based catalysts is rarely reported. Moreover, compared with other modification methods, tuning the calcination atmosphere to pretreat catalyst precursor is easier to operate and more economic, which has the potential application prospects.
Herein, Cu/Al2O3 catalyst precursor prepared by the ammonia evaporation method was calcined in the different atmospheres and employed for FAL hydrogenation. The physical and chemical properties of Cu/Al2O3 samples were systematically characterized by in situ XRD, TEM, N2O titration, H2-TPR, and XPS, and the correlation between catalytic activity and catalyst properties was discussed to unravel the effects of calcination atmosphere on the hydrogenation performance.

2. Experimental

2.1. Materials

FAL (99.8%), FA (99.8%), Al2O3 and Cu(NO3)2·3H2O were purchased from McLean Chemical Reagent Co., Ltd. (Shanghai, China). NH3·H2O (25 wt%) was purchased from Aladdin Reagent Shanghai Co., Ltd. (Shanghai, China). Isopropanol was purchased from Fengchuan Chemical Reagent Technology Co., Ltd. (Tianjin, China). The deionized water was self-prepared. All reagents were used without further purification.

2.2. Catalyst Preparation

The Cu/Al2O3 catalyst precursor was prepared by the ammonia evaporation method [42]. Firstly, 2.4030 g of Cu(NO3)2·3H2O was dissolved in 140 mL of deionized water. After stirring for 0.5 h, 25 wt% ammonia solution was slowly added to the above solution under stirring to adjust the pH to 12. Then, Al2O3 (12 g) was added to the above solution, and the suspension was stirred for another 4 h. Thereafter, the mixture was heated to 90 °C to evaporate ammonia until the pH decreased to 7. Finally, the resultant solid was filtered, washed thoroughly with the deionized water and dried at 120 °C for 12 h to obtain the Cu/Al2O3 catalyst precursor.
The Cu/Al2O3 catalyst precursor was calcined at 450 °C for 4 h in the different atmospheres, including N2 and air in the tube furnace. Before each test, Cu/Al2O3 catalyst was reduced in 10% vol. H2/Ar at 350 °C for 4 h. The calcined and reduced catalysts were denoted as Cu/Al2O3-N2, Cu/Al2O3-air, Cu/Al2O3-N2-R and Cu/Al2O3-air-R, respectively.

2.3. Catalyst Characterization

In situ X-ray diffraction (XRD) measurement was carried out on a SmartLab SE diffractometer (Rigaku Corporation), and the reaction cell was controlled by PTC-EVO. The patterns were recorded with a 2 Theta range of 20–80° and scan rate of 10 °/min. Transmission electron microscopy (TEM) and high-resolution TEM (HRTEM) were taken on a Tecnai G2 S-Twin F20 TEM microscope (FEI Company) at 200 kV. The textural properties were measured on an ASAP 2460 sorptometer (Micromeritics) at −196 °C. The sample was degassed at 150 °C for 3 h before each test. The specific surface area (SBET) was calculated by the Brunauer–Emmett–Teller (BET) method, and the pore volume, as well as the pore size, was calculated by the Barrett–Joyner–Halenda (BJH) model. The actual Cu loading was determined by an inductively coupled plasma atomic mission spectrometer (ICP-AES) on Shimadzu ICPE-9820. X-ray photoelectron spectra (XPS) and X-ray excited Auger electron spectra (AES) were acquired on a Thermo VG Scientific ESCALAB 250 (UK) equipped with Al Kα X-ray excitation. All binding energies were referenced to C 1s (284.8 eV).
H2 temperature-programmed reduction (H2-TPR) and Cu dispersion measurement were performed on AutoChem Ⅱ 2920 chemisorption analyzer (Micromeritics) equipped with an online thermal conductivity detector (TCD). For H2-TPR, the catalyst (~100 mg) was loaded into the U-shaped quartz tube, and pretreatment occurred under Ar flow at 120 °C for 60 min. The signal of H2 consumption was subsequently monitored under 10 vol.% H2/Ar (50 mL/min) at a heating rate of 10 °C/min from room temperature to 800 °C. Cu dispersion was measured by the technique of N2O titration [43,44]. Firstly, the fresh catalyst (~100 mg) was pretreated at 120 °C for 60 min under Ar flow and cooled to 50 °C. Afterward, the sample was heated to 350 °C under 10 vol.% H2/Ar gas (50 mL/min) with a rate of 10 °C/min. The area of H2 consumption was recorded as X. Then, the tube was cooled down to 50 °C and re-oxidized by 10 vol.% N2O/He for 1 h. Finally, the sample was heated to 350 °C under 10 vol.% H2/Ar (50 mL/min) with a rate of 10 °C/min. The area of H2 consumption was recorded as Y. Cu dispersion and the surface area of surface Cu species per gram catalyst (S(m2∙gcat−1)) were calculated according to the following equations:
Cu   dispersion = [ 2 Y   /   X ]   × 100 %
S = ( 2   ×   N A   ×   Y )   /   ( X   ×   1.4   ×   10 19 ×   M Cu   ×   Wt Cu % ) = ( 1353   ×   Y )   /   ( X   ×   Wt Cu % )
where, MCu, WtCu% and NA are the molecular weight of Cu, actual Cu loading and Avogadro’s constant, respectively. In addition, 1.4 × 1019 is the number of copper atoms per square meter.
In situ diffuse reflectance infrared Fourier transform (DRIFT) spectra of FAL adsorption and hydrogenation were conducted on INVENIO S FTIR (Bruker) equipped with a diffuse reflectance cell and mercury-cadmium telluride (MCT) detector. Before each test, the sample was reduced under 10 vol.% H2/Ar at 350 °C for 1 h. After cooling down to 30 °C under N2, the background signal was collected. Then, FAL was introduced into the cell with a bubbler using N2 as the carrier. Thereafter, the temperature was heated to 120 °C to desorb the physically adsorbed FAL. When the desorption was completed, 10 vol.% H2/Ar was introduced for in situ hydrogenation reaction, and the spectra were finally collected every 10 min.

2.4. Catalyst Evaluation

The activity of the catalysts was evaluated in a 100 mL stainless-steel batch reactor, which was equipped with mechanical stirrer and temperature controller. Typically, catalyst (0.1 g), FAL (0.5 g) and solvent (30 mL) were added into the reactor. After sealing, H2 was charged to replace the internal air for 5 times. Then, the reactor was filled with the desired H2 pressure and heated to the predetermined temperature under the continuous mechanical stirring (800 rpm). After the required time, the reactor was naturally cooled down to room temperature, and the mixture was separated by filtration and collected. The products were analyzed by a gas chromatographer (GC9600) equipped with flame ionization detector (FID), and the carrier gas was ultra-high-purity N2. The standard solutions of FAL and FA with different concentrations were precisely prepared. External standard method was used to quantify FAL conversion and product selectivity as follows [45,46]:
Conversion   ( % ) = Mole   of   FAL   reacted Mole   of   initial   FAL   ×   100 %
Selectivity   ( % ) = Mole   of   FA   Mole   of   FAL   converted   ×   100 %
Yield   ( % ) =   Conversion   ×   selectivity

3. Results and Discussion

3.1. Structure and Morphology

Figure 2a–f show the in situ XRD patterns of the Cu/Al2O3 catalyst precursor during calcination in N2 and air and the following reduction in H2/Ar. As shown in Figure 2a, with the increased temperature from 50 to 450 °C in the N2 atmosphere, only the CuO species with the characteristic diffraction peaks at 35.6° and 38.7° were observed [43,47]. Here, the CuO species might be derived from the thermal decomposition of Cu(OH)2 and [Cu(NH3)x]2+ in the precursor. Figure 2b displays the further thermal treatment in N2 atmosphere at 450 °C. The peaks at 43.3° and 50.5° corresponded to the (111) and (200) crystalline planes of Cu0, respectively [48]. Notably, the intensity of the diffraction peaks ascribed to the CuO species decreased gradually with the extended time (after 2 h, the peaks disappear), while the diffraction peaks indexed to the Cu0 species appeared after 1 h and became sharper with the prolonging time. Figure 2c displays the reduction process of the Cu/Al2O3-N2 catalyst in H2/Ar atmosphere with the increased temperature from 50 to 350 °C. Observably, the intensity of the diffraction peaks assigned to Cu0 gradually enhanced with increasing the reduction temperature, implying a further reduction of the CuO species. Figure 2d shows the patterns of the Cu/Al2O3-N2 catalyst reduced for another 4 h at 350 °C. It was noted that the diffraction peaks of Cu0 were almost unchanged, suggesting the high anti-sintering ability of the Cu/Al2O3 catalyst [49].
From in situ XRD patterns in Figure 2e,f during the calcination in air, one can note that only the diffraction peaks of CuO species were present, and no diffraction peaks of Cu0 appeared. When the Cu/Al2O3-air catalyst was reduced by H2/Ar (Figure 2g), the intensity of the diffraction peaks attributed to CuO decreased with increasing the reduction temperature, and the diffraction peaks of Cu0 species came to exist at 200 °C. When the reduction temperature was increased to 300 °C, the diffraction peaks of CuO completely disappeared. The intensity of the diffraction peaks corresponding to Cu0 continuously increased when the reduction temperature rose to 350 °C. As shown in Figure 2h, the diffraction peaks intensity of Cu0 remained stable after 3 h reduction at 350 °C. The aforementioned analysis indicates that the presence of Cu0 species during the calcination in N2 may have been to the generation of Cu(NH3)4(NO3)2 in the catalyst precursor prepared by ammonia evaporation method. The thermal decomposition of Cu(NH3)4(NO3)2 at a high temperature can produce reduction by NH3, which could reduce CuO to Cu0. Therefore, the Cu/Al2O3 precursor was easier to reduce to the lower valence state in the N2 atmosphere.
To study the textural properties of the Cu/Al2O3 precursor and the employed catalysts, N2 physical adsorption-desorption experiments were carried out (Table 1 and Figure 3). According to the IUPAC classification [50], all the isotherms exhibited type IV with H3 hysteresis loop (Figure 3a), indicating the typical mesoporous structure [48,51]. As shown in Figure 3b and Table 1, the specific surface areas, pore volumes and average pore sizes of the precursor and the catalysts, including the calcined and reduced samples, were very similar. It illustrates that the calcination atmosphere and further reduction did not affect the textural properties of the Cu/Al2O3 catalyst [52].
TEM and HRTEM were employed to investigate the size of Cu nanoparticles and the morphology of the reduced Cu/Al2O3 catalysts. As shown in Figure 4a,b, the average size of Cu nanoparticles was 4.9 and 6.6 nm for Cu/Al2O3-N2-R and Cu/Al2O3-air-R, respectively. That is, the average particle size of the Cu species on Cu/Al2O3-N2-R was smaller and more uniformly dispersed than that on Cu/Al2O3-air-R. The lattice fringe spacing was measured to be 0.21 and 0.18 nm in the HRTEM images (Figure 4c,d), corresponding to the (111) and (200) crystalline plane of Cu, respectively [48]. The assignment of the above crystalline planes is consistent with the XRD results. Moreover, the high-angle annular darkfield scanning transmission electron microscopy (HAADF-STEM) and the corresponding elemental mapping images (Figure 4f–h,j–l) confirm that Cu, O and Al were homogeneously distributed on the Cu/Al2O3 catalysts [53].
The actual Cu loadings on the catalysts determined by ICP-AES are listed in Table 1. The actual loading of 4.3 wt.% was lower than the theoretical loading of 5 wt.%, which may have been caused by the incomplete precipitation and washing during the preparation [42]. The actual loadings on the different catalysts were the same, which suggests that the calcination atmosphere and reduction treatment could not have resulted in the metal loss and thus did not affect the experimental results and conclusions. Moreover, the dispersion (DCu) and specific surface area of Cu (SCu) on the catalysts were calculated (Table 1). The results show that the dispersion on Cu/Al2O3-N2-R catalyst was 48.3%, which was much higher than that on Cu/Al2O3-air-R catalyst, and the specific surface area of Cu was 14.1 m2Cu/gcat. The Cu dispersion and specific surface area on Cu/Al2O3-air-R catalyst were only 21.5% and 6.3 m2Cu/gcat, respectively. The data are consistent with the TEM results. In conclusion, the catalyst calcined in the N2 atmosphere was more favorable for the dispersion and reduction of Cu species.

3.2. Surface Chemical State

H2-TPR was performed to study the reduction behavior of the calcined Cu/Al2O3 catalysts, and the profiles are given in Figure 5a. Apparently, the calcination atmosphere significantly affected the reduction behaviors of the CuO species. The reduction peaks centered at 208 °C/218 °C (a) and 245 °C/290 °C (b) were observed for the Cu/Al2O3 catalysts [28]. The peak a was attributed to the reduction of the highly dispersed CuO species, and the peak b was related to the reduction of the small-sized CuO species [54,55]. The temperature of the peak a for Cu/Al2O3-N2 was lower than that for Cu/Al2O3-air. This indicates that the former was more easily reduced, which may have been due to its higher Cu dispersion. Peak fitting and integration were performed to better quantify the percentage of different CuO species, and the quantitative results are plotted in Figure 5b. Cu/Al2O3-N2 had a higher proportion of highly dispersed CuO, accounting for 95.1% of all Cu species, while Cu/Al2O3-air only contained 45.4% of highly dispersed CuO. This again verifies that for the Cu/Al2O3 precursor, the calcination in N2 was more favorable for the dispersion of Cu species, which are easier to reduce. The analysis is in good agreement with the TEM and N2O titration results.
To further examine the chemical state of Cu species, XPS was also conducted, and the spectra are displayed in Figure 6a. The spectra at the region of 930.0–937.9 eV were fitted to two characteristic peaks, where the feature at 932.5 eV corresponded to the low oxidation state species (Cu+/Cu0), and the one at 934.4 eV was assigned to the Cu2+ species [54]. According to the fitting results (Table 2), the ratio of (Cu+/Cu0)/Cu on the surface of Cu/Al2O3-air was only 21.8%, but the Cu/Al2O3-N2 reached 37.1%. After reduction in H2/Ar, the ratio of (Cu+/Cu0)/Cu increased to 48.8% and 60.8%, respectively. This suggests that the Cu/Al2O3 catalyst calcined in the N2 atmosphere was more easily reduced to the low oxidation state, which is consistent with the H2-TPR results.
It is difficult to make a clear distinction between Cu+ and Cu0 on the basis of XPS analysis [56]. Thus, the Cu AES spectra were also recorded (Figure 6b). Three peaks were extracted by peak fitting in the Kinetic energy (KE) range of 903-922 eV, including the characteristic of Cu0 (KE = 918.3 eV), Cu2+ (KE = 916.3 eV) and Cu+ (KE = 914.3 eV) species [57]. Further quantitative analysis of the fitted peaks is listed in Table 2. The Cu+ and Cu0 ratio of Cu/Al2O3 catalyst calcined in air atmosphere were 13.8% and 8.0%, respectively. However, the Cu+ and Cu0 content of Cu/Al2O3-N2 reached higher levels of 27.4% and 9.7%. It further indicates that the calcination in the N2 atmosphere was more favorable for the generation of low valence Cu species, which is in accordance with the results of in situ XRD and TPR. After reduction by H2/Ar atmosphere, the Cu+ and Cu0 content on both Cu/Al2O3-air-R and Cu/Al2O3-N2-R further increased, where on the former, an increase to 38.6% and 10.2% was observed, respectively, and on the latter, the Cu+ and Cu0 ratio had the highest values of 45.3% and 15.5%, respectively. According to previous reports, high Cu+ and Cu0 ratio facilitates the selective hydrogenation of FAL [24,48,54].
Full-spectrum analysis of XPS spectra was performed. The surface Cu/Al atomic ratios of Cu/Al2O3-N2-R and Cu/Al2O3-air-R were 0.056 and 0.092, respectively. The surface Cu/Al atomic ratio of Cu/Al2O3-N2-R was lower, indicating that Cu nanoparticles were highly dispersed on the Al2O3 support. The small Cu nanoparticles easily entered the support pores, leading to a decrease in the surface Cu/Al atomic ratio. However, the Cu/Al2O3-air-R dispersion was poor. A large number of Cu nanoparticles aggregated on the support surface, resulting in a high surface Cu/Al atomic ratio. The above results are in good agreement with the N2O titration and TEM results, which prove that Cu/Al2O3-N2-R had higher dispersion.

3.3. Catalytic Performance

The selective hydrogenation of FAL over the employed Cu/Al2O3 catalysts was examined in the stainless-steel batch reactor, and the obtained data are given in Table 3. FA yield over the Cu/Al2O3 catalysts showed the following trend: Cu/Al2O3-N2-R > Cu/Al2O3-N2 > Cu/Al2O3-air-R > Cu/Al2O3-air. Cu/Al2O3 precursor calcined in air (Cu/Al2O3-air) had very low FAL conversion and FA selectivity. Even after reduction (Cu/Al2O3-air-R), the improvement of the catalytic activity was still very limited, where FAL conversion and FA selectivity were only 26.7% and 85.8%, respectively. Interestingly, the precursor calcined in N2 had the relatively superior activity. FAL conversion reached 42.7%, and FA selectivity was even 99.9% over Cu/Al2O3-N2. After reduction, the activity was greatly enhanced. Specifically, FAL conversion and FA selectivity reached 99.9% over Cu/Al2O3-N2-R.
The evaluation data indicate that the calcination atmosphere significantly affected the performance of Cu/Al2O3 catalyst for FAL hydrogenation. Compared with the Cu/Al2O3-air-R, the FAL conversion and FA selectivity of Cu/Al2O3-N2-R were greatly improved. In general, among the Cu0 and Cu+ active hydrogenation species on the catalyst surface, the Cu0 species contributed the dissociation and activation of H2 [27]. The above in situ XRD, TEM, N2O titration and H2-TPR characterization results demonstrate that Cu/Al2O3 catalyst calcined in the N2 atmosphere was more favorable for the dispersion and reduction of Cu species. This facilitated the reduction of CuO to the lowest valence state of the Cu0 species. The XPS results were the same as the above conclusion. The Cu0 content over Cu/Al2O3-N2-R was the highest, accounting for 15.5%. Then, Cu+ species could act as Lewis acid sites and polarize the carbonyl group (C=O), thus promoting the formation of FA [24]. The Cu AES spectra fitting results show that Cu/Al2O3-N2-R also contained the highest content of Cu+ species, accounting for 45.3%. Compared to Cu/Al2O3-air, the Cu/Al2O3 catalyst calcined in the N2 atmosphere had higher Cu dispersion and higher content of Cu+/Cu0, resulting in the improved FA conversion and FAL selectivity. After reduction, Cu/Al2O3-N2-R obtained the highest content of Cu+ and Cu0 active hydrogenated species, and the FAL conversion was further improved. Cu0 and Cu+ synergistically catalyzed the selective hydrogenation of FAL to prepare FA. Therefore, Cu/Al2O3-N2-R had the highest catalytic activity.
The reusability of the Cu/Al2O3-N2-R catalyst for FAL hydrogenation was also tested. After every cycle, the catalyst was washed with isopropanol and water to remove the adsorbed organics on the surface and dried overnight in a vacuum oven at 60 °C. As displayed in Figure 7, FAL conversion and FA selectivity was not obviously declined after four recycles. These results imply that the catalyst possessed a good reusability performance.
Table 4 gives a brief comparison of the performance over the representative Cu-based catalysts for FAL hydrogenation to FA using the batch reactor. Notably, the present work provides an efficient FAL hydrogenation catalyst under the relatively mild conditions. Therefore, the strategy in this work has certain advantages for the selective hydrogenation of FAL catalyzed by non-noble metal catalysts.

3.4. Reaction Mechanisms

Figure 8a shows the in situ DRIFT spectra of FAL adsorption at 120 °C. The observed bands located at 1643 cm−1 were attributed to the C=O stretching vibration [71], and the bands at 1592 cm−1 and 1452 cm−1 corresponded to the furan ring breath and C=C characteristic vibrations, respectively [51,71]. Compared with the v(C=O) for gaseous FAL, the significant red shift in v(C=O) from 1670 cm−1 to 1643 cm−1 implies that FAL was linearly chemical adsorbed in the η1(O) configuration on the catalysts [52]. This facilitates FAL hydrogenation towards the production of FA. Cu/Al2O3-N2-R had larger C=O peak intensity, which indicates that Cu/Al2O3-N2-R was more favorable for C=O chemical adsorption. Moreover, in situ DRIFT spectra of FAL hydrogenation are also provided in Figure 8b to further understand the evolution of C=O hydrogenation. Apparently, the strong characteristic bands assigned to C=O declined faster on Cu/Al2O3-N2-R with the continuous inflow of H2 than the Cu/Al2O3-air-R catalyst, indicating the higher catalytic activity of Cu/Al2O3-N2-R. This is consistent with experimental data on FAL hydrogenation.

4. Conclusions

In summary, the Cu/Al2O3 catalysts pretreated by calcining in the N2 and air atmosphere were prepared and applied for FAL liquid-phase hydrogenation. The activity of the catalyst calcined in air was lower than that calcined in N2, which was related to the generated lower valence state during the calcination process on the latter. After reduction, the performance of FAL selective hydrogenation to FA was proliferated. Specially, the Cu/Al2O3-N2-R catalyst exhibited the most excellent performance. At 120 °C, 1 MPa H2, 2 h, both FAL conversion and FA selectivity over Cu/Al2O3-N2-R reached 99.9%, which was comparable to the reported Cu-based catalysts. The comprehensive characterization results based on in situ XRD, TEM, N2O titration, H2-TPR and XPS demonstrate that the Cu/Al2O3 catalyst calcined in the N2 atmosphere was more favorable for the dispersion and reduction of Cu species and the reduction process could produce more Cu+ and Cu0 species. Thus, Cu/Al2O3-N2-R had the highest content of Cu+/Cu0 active hydrogenation species among the employed catalysts, which provides a reasonable explanation for its efficient activity under the relatively mild conditions. This work designs a simple strategy for optimizing the reactivity of Cu-based catalysts and is conducive to efficiently utilizing biomass and its derivatives to develop a renewable energy system.

Author Contributions

Conceptualization, M.Z.; methodology, W.Y., M.Z. and B.W.; software, J.Y. and T.Y.; validation, K.J., T.Y. and Z.L. (Zhihan Li); formal analysis, Y.G., J.Y. and K.J.; investigation, Y.G., W.Y. and J.Y.; resources, Z.L. (Zhongyi Liu); data curation, K.J. and T.Y.; writing—original draft preparation, Y.G. and T.Y.; writing—review and editing, Z.L. (Zhihan Li), M.Z. and Z.L. (Zhongyi Liu); supervision, Z.L. (Zhongyi Liu) and B.W.; project administration, M.Z. and B.W.; funding acquisition, Z.L. (Zhongyi Liu). All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data will be made available on request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. An, Z.; Li, J. Recent advances in the catalytic transfer hydrogenation of furfural to furfuryl alcohol over heterogeneous catalysts. Green Chem. 2022, 24, 1780–1808. [Google Scholar] [CrossRef]
  2. Climent, M.J.; Corma, A.; Iborra, S. Conversion of biomass platform molecules into fuel additives and liquid hydrocarbon fuels. Green Chem. 2014, 16, 516–547. [Google Scholar] [CrossRef]
  3. He, D.; Li, T.; Dai, X.; Liu, S.; Cui, X.; Shi, F. Construction of highly active and selective molecular imprinting catalyst for hydrogenation. J. Am. Chem. Soc. 2023, 145, 20813–20824. [Google Scholar] [CrossRef]
  4. Jaswal, A.; Singh, P.P.; Mondal, T. Furfural-a versatile, biomass-derived platform chemical for the production of renewable chemicals. Green Chem. 2022, 24, 510–551. [Google Scholar] [CrossRef]
  5. Zhang, X.; Xu, S.; Li, Q.; Zhou, G.; Xia, H. Recent advances in the conversion of furfural into bio-chemicals through chemo- and bio-catalysis. RSC Adv. 2021, 11, 27042–27058. [Google Scholar] [CrossRef] [PubMed]
  6. Shi, N.; Liu, Q.; Zhang, Q.; Wang, T.; Ma, L. High yield production of 5-hydroxymethylfurfural from cellulose by high concentration of sulfates in biphasic system. Green Chem. 2013, 15, 1967–1974. [Google Scholar] [CrossRef]
  7. Chen, H.; Ruan, H.; Lu, X.; Fu, J.; Langrish, T.; Lu, X. Efficient catalytic transfer hydrogenation of furfural to furfuryl alcohol in near-critical isopropanol over Cu/MgO-Al2O3 catalyst. Mol. Catal. 2018, 445, 94–101. [Google Scholar] [CrossRef]
  8. Mariscal, R.; Maireles-Torres, P.; Ojeda, M.; Sádaba, I.; López Granados, M. Furfural: A renewable and versatile platform molecule for the synthesis of chemicals and fuels. Energy Environ. Sci. 2016, 9, 1144–1189. [Google Scholar] [CrossRef]
  9. Ranaware, V.; Kurniawan, R.G.; Verma, D.; Kwak, S.K.; Ryu, B.C.; Kang, J.W.; Kim, J. Solvent-mediated selectivity control of furfural hydrogenation over a N-doped carbon-nanotube-supported Co/CoOx catalyst. Appl. Catal., B 2022, 318, 121838. [Google Scholar] [CrossRef]
  10. Xu, C.; Paone, E.; Rodriguez-Padron, D.; Luque, R.; Mauriello, F. Recent catalytic routes for the preparation and the upgrading of biomass derived furfural and 5-hydroxymethylfurfural. Chem. Soc. Rev. 2020, 49, 4273–4306. [Google Scholar] [CrossRef]
  11. Yan, K.; Wu, G.; Lafleur, T.; Jarvis, C. Production, properties and catalytic hydrogenation of furfural to fuel additives and value-added chemicals. Renew. Sustain. Energy Rev. 2014, 38, 663–676. [Google Scholar] [CrossRef]
  12. Chen, H.; Ding, J.; Liang, H.; Yu, H. Synthesis and application of sustainable furfuryl alcohol-based plasticizer. ChemistrySelect 2020, 5, 4085–4090. [Google Scholar] [CrossRef]
  13. Merat, N.; Godawa, C.; Gaset, A. High selective production of tetrahydrofurfuryl alcohol: Catalytic hydrogenation of furfural and furfuryl alcohol. J. Chem. Technol. Biotechnol. 1990, 48, 145–159. [Google Scholar] [CrossRef]
  14. Chen, S.; Wojcieszak, R.; Dumeignil, F.; Marceau, E.; Royer, S. How catalysts and experimental conditions determine the selective hydroconversion of furfural and 5-hydroxymethylfurfural. Chem. Rev. 2018, 118, 11023–11117. [Google Scholar] [CrossRef] [PubMed]
  15. Jiménez-Gómez, C.P.; Cecilia, J.A.; Durán-Martín, D.; Moreno-Tost, R.; Santamaría-González, J.; Mérida-Robles, J.; Mariscal, R.; Maireles-Torres, P. Gas-phase hydrogenation of furfural to furfuryl alcohol over Cu/ZnO catalysts. J. Catal. 2016, 336, 107–115. [Google Scholar] [CrossRef]
  16. Arundhathi, R.; Reddy, P.L.; Samanta, C.; Newalkar, B.L. Chromium-free Cu@Mg/γ-Al2O3-an active catalyst for selective hydrogenation of furfural to furfuryl alcohol. RSC Adv. 2020, 10, 41120–41126. [Google Scholar] [CrossRef] [PubMed]
  17. Wang, Z.; Wang, X.; Zhang, C.; Arai, M.; Zhou, L.; Zhao, F. Selective hydrogenation of furfural to furfuryl alcohol over Pd/TiH2 catalyst. Mol. Catal. 2021, 508, 111599. [Google Scholar] [CrossRef]
  18. Ruan, L.; Zhu, L.; Zhang, X.; Guo, G.; Shang, C.; Hui Chen, B.; Guo, Z. Porous SiO2 nanosphere-supported PtCuCo trimetallic nanoparticles for highly efficient and selective furfural hydrogenation. Fuel 2023, 335, 126935. [Google Scholar] [CrossRef]
  19. Yang, J.; Ma, J.; Yuan, Q.; Zhang, P.; Guan, Y. Selective hydrogenation of furfural on Ru/Al-MIL-53: A comparative study on the effect of aromatic and aliphatic organic linkers. RSC Adv. 2016, 6, 92299–92304. [Google Scholar] [CrossRef]
  20. Li, J.; Liu, J.-L.; Zhou, H.-j.; Fu, Y. Catalytic transfer hydrogenation of furfural to furfuryl alcohol over nitrogen-doped carbon-supported iron catalysts. ChemSusChem 2016, 9, 1339–1347. [Google Scholar] [CrossRef]
  21. Ruan, L.; Pei, A.; Liao, J.; Zeng, L.; Guo, G.; Yang, K.; Zhou, Q.; Zhao, N.; Zhu, L.; Chen, B.H. Nitrogen-doped carbon nanotubes-supported PdNiCo nanoparticles as a highly efficient catalyst for selective hydrogenation of furfural. Fuel 2021, 284, 119015. [Google Scholar] [CrossRef]
  22. Sheng, Y.; Tian, F.; Wang, X.; Jiang, N.; Zhang, X.; Chen, X.; Liang, C.; Wang, A. Carbon-encapsulated Ni catalysts derived from citrate complexes for highly efficient hydrogenation of furfural to tetrahydrofurfuryl alcohol. Energy 2024, 292, 130360. [Google Scholar] [CrossRef]
  23. Islam, M.J.; Granollers Mesa, M.; Osatiashtiani, A.; Taylor, M.J.; Manayil, J.C.; Parlett, C.M.A.; Isaacs, M.A.; Kyriakou, G. The effect of metal precursor on copper phase dispersion and nanoparticle formation for the catalytic transformations of furfural. Appl. Catal. B 2020, 273, 119062. [Google Scholar] [CrossRef]
  24. Li, J.; Niu, X.; Zhu, Y. Synergistic effect of surface Cu0 and Cu+ species on improved selective hydrogenation of furfural to furfuryl alcohol over hydrotalcite-derived CuxMg3Al oxides. Appl. Surf. Sci. 2023, 635, 157774. [Google Scholar] [CrossRef]
  25. Li, X.; Jia, P.; Wang, T. Furfural: A promising platform compound for sustainable production of C4 and C5 chemicals. ACS Catal. 2016, 6, 7621–7640. [Google Scholar] [CrossRef]
  26. Zhang, J.; Jia, Z.; Yu, S.; Liu, S.; Li, L.; Xie, C.; Wu, Q.; Zhang, Y.; Yu, H.; Liu, Y.; et al. Regulating the Cu0-Cu+ ratio to enhance metal-support interaction for selective hydrogenation of furfural under mild conditions. Chem. Eng. J. 2023, 468, 143955. [Google Scholar] [CrossRef]
  27. Yi, W.-J.; Gao, Y.; Yang, J.; Zhou, X.; Liu, Z.; Zhang, M. Synergistic effect of surface Cu0 and Cu+ species over hydrotalcite-derived CuxCo3-xAlOy mixed-metal oxides toward efficient hydrogenation of furfural to furfuryl alcohol. Appl. Surf. Sci. 2023, 641, 158559. [Google Scholar] [CrossRef]
  28. Tan, J.; He, J.; Gao, K.; Zhu, S.; Cui, J.; Huang, L.; Zhu, Y.; Zhao, Y. Catalytic hydrogenation of furfural over Cu/CeO2 catalyst: The effect of support morphology and exposed facet. Appl. Surf. Sci. 2022, 604, 154472. [Google Scholar] [CrossRef]
  29. Jiménez-Gómez, C.P.; Cecilia, J.A.; Alba-Rubio, A.C.; Cassidy, A.; Moreno-Tost, R.; García-Sancho, C.; Maireles-Torres, P. Tailoring the selectivity of Cu-based catalysts in the furfural hydrogenation reaction: Influence of the morphology of the silica support. Fuel 2022, 319, 123827. [Google Scholar] [CrossRef]
  30. Zhang, W.; Wang, Y.; Gu, B.; Tang, Q.; Cao, Q.-E.; Fang, W. Regulating the interaction within Pd-Cu dual metal sites for selective hydrogenation of furfural using ambient H2 pressure. ACS Sustain Chem. Eng. 2023, 11, 12798–12808. [Google Scholar] [CrossRef]
  31. Kumar, A.; Bal, R.; Srivastava, R. Modulation of Ru and Cu nanoparticle contents over CuAlPO-5 for synergistic enhancement in the selective reduction and oxidation of biomass-derived furan based alcohols and carbonyls. Catal. Sci. Technol. 2021, 11, 4133–4148. [Google Scholar] [CrossRef]
  32. Du, H.; Ma, X.; Yan, P.; Jiang, M.; Zhao, Z.; Zhang, Z.C. Catalytic furfural hydrogenation to furfuryl alcohol over Cu/SiO2 catalysts: A comparative study of the preparation methods. Fuel Process. Technol. 2019, 193, 221–231. [Google Scholar] [CrossRef]
  33. Wang, S.; Lv, Y.; Ren, J.; Xu, Z.; Yang, Q.; Zhao, H.; Gao, D.; Chen, G. Ultrahigh selective hydrogenation of furfural enabled by modularizing hydrogen dissociation and substrate activation. ACS Catal. 2023, 13, 8720–8730. [Google Scholar] [CrossRef]
  34. Gong, W.; Chen, C.; Zhang, Y.; Zhou, H.; Wang, H.; Zhang, H.; Zhang, Y.; Wang, G.; Zhao, H. Efficient synthesis of furfuryl alcohol from H2-hydrogenation/transfer hydrogenation of furfural using sulfonate group modified Cu catalyst. ACS Sustain. Chem. Eng. 2017, 5, 2172–2180. [Google Scholar] [CrossRef]
  35. Zhang, Z.; Guo, R.; Yang, X.; Fang, Y.-X. Potassium carbonate (K2CO3)-assisted copper-catalyzed liquid-phase hydrogenation of furfural: Striking promotion synergy enables a superior high furfuryl alcohol yield at mild reaction conditions. Ind. Eng. Chem. Res. 2022, 61, 16643–16652. [Google Scholar] [CrossRef]
  36. Cheng, S.; Lei, Q.; Deng, C.; Liang, L.; Chen, Y.; Meng, H.; Lei, W.; Chen, H. Mild and selective transfer hydrogenation of biomass-derived furfural to furfuryl alcohol over Cu/ZnO/Al2O3 with methanediol as the hydrogen donor. Green Chem. 2024. Advance Article. [Google Scholar] [CrossRef]
  37. Lee, J.; Seo, J.H.; Nguyen-Huy, C.; Yang, E.; Lee, J.G.; Lee, H.; Jang, E.J.; Kwak, J.H.; Lee, J.H.; Lee, H.; et al. Cu2O(100) surface as an active site for catalytic furfural hydrogenation. Appl. Catal. B 2021, 282, 119576. [Google Scholar] [CrossRef]
  38. Liu, J.; Wang, H.; Chen, Y.; Yang, M.; Wu, Y. Effects of pretreatment atmospheres on the catalytic performance of Pd/γ-Al2O3 catalyst in benzene degradation. Catal. Commun. 2014, 46, 11–16. [Google Scholar] [CrossRef]
  39. Gao, X.; Ashok, J.; Kawi, S. A review on roles of pretreatment atmospheres for the preparation of efficient Ni-based catalysts. Catal. Today 2022, 397–399, 581–591. [Google Scholar] [CrossRef]
  40. Zhang, M.; Zhang, J.; Wu, Y.; Pan, J.; Zhang, Q.; Tan, Y.; Han, Y. Insight into the effects of the oxygen species over Ni/ZrO2 catalyst surface on methane reforming with carbon dioxide. Appl. Catal. B 2019, 244, 427–437. [Google Scholar] [CrossRef]
  41. Liu, D.; Wang, Y.; Shi, D.; Jia, X.; Wang, X.; Borgna, A.; Lau, R.; Yang, Y. Methane reforming with carbon dioxide over a Ni/ZiO2-SiO2 catalyst: Influence of pretreatment gas atmospheres. Int. J. Hydrogen Energy 2012, 37, 10135–10144. [Google Scholar] [CrossRef]
  42. Chen, C.-C.; Lin, L.; Ye, R.-P.; Sun, M.-L.; Yang, J.-X.; Li, F.; Yao, Y.-G. Mannitol as a novel dopant for Cu/SiO2: A low-cost, environmental and highly stable catalyst for dimethyl oxalate hydrogenation without hydrogen prereduction. J. Catal. 2020, 389, 421–431. [Google Scholar] [CrossRef]
  43. Cui, G.; Zhang, X.; Wang, H.; Li, Z.; Wang, W.; Yu, Q.; Zheng, L.; Wang, Y.; Zhu, J.; Wei, M. ZrO2−x modified Cu nanocatalysts with synergistic catalysis towards carbon-oxygen bond hydrogenation. Appl. Catal. B 2021, 280, 119406. [Google Scholar] [CrossRef]
  44. Cui, G.; Meng, X.; Zhang, X.; Wang, W.; Xu, S.; Ye, Y.; Tang, K.; Wang, W.; Zhu, J.; Wei, M.; et al. Low-temperature hydrogenation of dimethyl oxalate to ethylene glycol via ternary synergistic catalysis of Cu and acid-base sites. Appl. Catal. B 2019, 248, 394–404. [Google Scholar] [CrossRef]
  45. Kurniawan, R.G.; Karanwal, N.; Park, J.; Verma, D.; Kwak, S.K.; Kim, S.K.; Kim, J. Direct conversion of furfural to 1,5-pentanediol over a nickel-cobalt oxide-alumina trimetallic catalyst. Appl. Catal., B 2023, 320, 121971. [Google Scholar] [CrossRef]
  46. Gao, G.; Remón, J.; Jiang, Z.; Yao, L.; Hu, C. Selective hydrogenation of furfural to furfuryl alcohol in water under mild conditions over a hydrotalcite-derived Pt-based catalyst. Appl. Catal. B 2022, 309, 121260. [Google Scholar] [CrossRef]
  47. Thongratkaew, S.; Luadthong, C.; Kiatphuengporn, S.; Khemthong, P.; Hirunsit, P.; Faungnawakij, K. Cu-Al spinel-oxide catalysts for selective hydrogenation of furfural to furfuryl alcohol. Catal. Today 2021, 367, 177–188. [Google Scholar] [CrossRef]
  48. Ren, Y.; Yang, Y.; Chen, L.; Wang, L.; Shi, Y.; Yin, P.; Wang, W.; Shao, M.; Zhang, X.; Wei, M. Synergetic effect of Cu0-Cu+ derived from layered double hydroxides toward catalytic transfer hydrogenation reaction. Appl. Catal. B 2022, 314, 121515. [Google Scholar] [CrossRef]
  49. Zhang, Z.; Sun, S.; Chen, Y.; Wei, Y.; Zhang, M.; Li, C.; Sun, Y.; Zhang, S.; Jiang, Y. Epitaxial growth of Cu2−xSe on Cu (220) crystal plane as high property anode for sodium storage. Chin. Chem. Lett. 2023, 35, 108922. [Google Scholar] [CrossRef]
  50. Islam, M.J.; Granollers Mesa, M.; Osatiashtiani, A.; Manayil, J.C.; Isaacs, M.A.; Taylor, M.J.; Tsatsos, S.; Kyriakou, G. PdCu single atom alloys supported on alumina for the selective hydrogenation of furfural. Appl. Catal. B 2021, 299, 120652. [Google Scholar] [CrossRef]
  51. Meng, X.; Yang, Y.; Chen, L.; Xu, M.; Zhang, X.; Wei, M. A control over hydrogenation selectivity of furfural via tuning exposed facet of Ni catalysts. ACS Catal. 2019, 9, 4226–4235. [Google Scholar] [CrossRef]
  52. Wang, Q.; Feng, J.; Zheng, L.; Wang, B.; Bi, R.; He, Y.; Liu, H.; Li, D. Interfacial structure-determined reaction pathway and selectivity for 5-(hydroxymethyl)furfural hydrogenation over Cu-based catalysts. ACS Catal. 2019, 10, 1353–1365. [Google Scholar] [CrossRef]
  53. Shao, Y.; Li, Q.; Dong, X.; Wang, J.; Sun, K.; Zhang, L.; Zhang, S.; Xu, L.; Yuan, X.; Hu, X. Cooperation between hydrogenation and acidic sites in Cu-based catalyst for selective conversion of furfural to γ-valerolactone. Fuel 2021, 293, 120457. [Google Scholar] [CrossRef]
  54. Zhao, H.; Liao, X.; Cui, H.; Zhu, M.; Hao, F.; Xiong, W.; Luo, H.; Lv, Y.; Liu, P. Efficient Cu-Co bimetallic catalysts for the selective hydrogenation of furfural to furfuryl alcohol. Fuel 2023, 351, 128887. [Google Scholar] [CrossRef]
  55. Park, S.; Kannapu, H.P.R.; Jeong, C.; Kim, J.; Suh, Y.-W. Highly active mesoporous Cu-Al2O3 catalyst for the hydrodeoxygenation of furfural to 2-methylfuran. ChemCatChem 2020, 12, 105–111. [Google Scholar] [CrossRef]
  56. Wang, S.; Song, L.; Qu, Z. Cu/ZnAl2O4 catalysts prepared by ammonia evaporation method: Improving methanol selectivity in CO2 hydrogenation via regulation of metal-support interaction. Chem. Eng. J. 2023, 469, 144008. [Google Scholar] [CrossRef]
  57. Fang, W.; Liu, S.; Steffensen, A.K.; Schill, L.; Kastlunger, G.; Riisager, A. On the role of Cu+ and CuNi alloy phases in mesoporous CuNi catalyst for furfural hydrogenation. ACS Catal. 2023, 13, 8437–8444. [Google Scholar] [CrossRef]
  58. Luo, L.; Yuan, F.; Zaera, F.; Zhu, Y. Catalytic hydrogenation of furfural to furfuryl alcohol on hydrotalcite-derived CuxNi3−xAlOy mixed-metal oxides. J. Catal. 2021, 404, 420–429. [Google Scholar] [CrossRef]
  59. Chen, S.; de Souza, P.M.; Ciotonea, C.; Marinova, M.; Dumeignil, F.; Royer, S.; Wojcieszak, R. Micro-/mesopores confined ultrasmall Cu nanoparticles in SBA-15 as a highly efficient and robust catalyst for furfural hydrogenation to furfuryl alcohol. Appl. Catal. A 2022, 633, 118527. [Google Scholar] [CrossRef]
  60. Villaverde, M.M.; Bertero, N.M.; Garetto, T.F.; Marchi, A.J. Selective liquid-phase hydrogenation of furfural to furfuryl alcohol over Cu-based catalysts. Catal. Today 2013, 213, 87–92. [Google Scholar] [CrossRef]
  61. Wang, C.; Liu, Y.; Cui, Z.; Yu, X.; Zhang, X.; Li, Y.; Zhang, Q.; Chen, L.; Ma, L. In situ synthesis of Cu nanoparticles on carbon for highly selective hydrogenation of furfural to furfuryl alcohol by using pomelo peel as the carbon source. ACS Sustain. Chem. Eng. 2020, 8, 12944–12955. [Google Scholar] [CrossRef]
  62. Ghashghaee, M.; Shirvani, S.; Ghambarian, M. Kinetic models for hydroconversion of furfural over the ecofriendly Cu-MgO catalyst: An experimental and theoretical study. Appl. Catal. A 2017, 545, 134–147. [Google Scholar] [CrossRef]
  63. Zhang, J.; Wu, D. Aqueous phase catalytic hydrogenation of furfural to furfuryl alcohol over in-situ synthesized Cu-Zn/SiO2 catalysts. Mater. Chem. Phys. 2021, 260, 124152. [Google Scholar] [CrossRef]
  64. Cao, P.; Lin, L.; Qi, H.; Chen, R.; Wu, Z.; Li, N.; Zhang, T.; Luo, W. Zeolite-encapsulated Cu nanoparticles for the selective hydrogenation of furfural to furfuryl alcohol. ACS Catal. 2021, 11, 10246–10256. [Google Scholar] [CrossRef]
  65. Li, X.-L.; Deng, J.; Shi, J.; Pan, T.; Yu, C.-G.; Xu, H.-J.; Fu, Y. Selective conversion of furfural to cyclopentanone or cyclopentanol using different preparation methods of Cu-Co catalysts. Green Chem. 2015, 17, 1038–1046. [Google Scholar] [CrossRef]
  66. Tang, F.; Wang, L.; Dessie Walle, M.; Mustapha, A.; Liu, Y.-N. An alloy chemistry strategy to tailoring the d-band center of Ni by Cu for efficient and selective catalytic hydrogenation of furfural. J. Catal. 2020, 383, 172–180. [Google Scholar] [CrossRef]
  67. Şebin, M.E.; Akmaz, S.; Koc, S.N. Hydrogenation of furfural to furfuryl alcohol over efficient sol-gel nickel-copper/zirconia catalyst. J. Chem. Sci. 2020, 132, 157. [Google Scholar] [CrossRef]
  68. Khromova, S.A.; Bykova, M.V.; Bulavchenko, O.A.; Ermakov, D.Y.; Saraev, A.A.; Kaichev, V.V.; Venderbosch, R.H.; Yakovlev, V.A. Furfural hydrogenation to furfuryl alcohol over bimetallic Ni-Cu sol-gel catalyst: A model reaction for conversion of oxygenates in pyrolysis liquids. Top. Catal. 2016, 59, 1413–1423. [Google Scholar] [CrossRef]
  69. O’Neill, B.J.; Sener, C.; Jackson DH, K.; Kuech, T.F.; Dumesic, J.A. Control of thickness and chemical properties of atomic layer deposition overcoats for stabilizing Cu/γ-Al2O3 catalysts. ChemSusChem 2014, 7, 3247–3251. [Google Scholar] [CrossRef]
  70. Jiménez-Gómez, C.P.; Cecilia, J.A.; Márquez-Rodríguez, I.; Moreno-Tost, R.; Santamaría-González, J.; Mérida-Robles, J.; Maireles-Torres, P. Gas-phase hydrogenation of furfural over Cu/CeO2 catalysts. Catal. Today 2017, 279, 327–338. [Google Scholar] [CrossRef]
  71. Yang, Q.; Gao, D.; Li, C.; Wang, S.; Hu, X.; Zheng, G.; Chen, G. Highly dispersed Pt on partial deligandation of Ce-MOFs for furfural selective hydrogenation. Appl. Catal. B 2023, 328, 122458. [Google Scholar] [CrossRef]
Figure 1. Reaction pathways of FAL hydrogenation to FA and other chemicals.
Figure 1. Reaction pathways of FAL hydrogenation to FA and other chemicals.
Molecules 29 02753 g001
Figure 2. In situ XRD patterns of the Cu/Al2O3 precursor calcinated in N2 (a,b) and further reduced in H2/Ar (c,d): (a) 50–450 °C with different temperatures, (b) 450 °C with different time, (c) 50–350 °C with different temperatures, (d) 350 °C with different time; In situ XRD patterns of the Cu/Al2O3 precursor calcined in air (e,f) and further reduced in H2/Ar (g,h): (e) 50–450 °C with different temperatures, (f) 450 °C with different time, (g) 50–350 °C with different temperatures, (h) 350 °C with different time.
Figure 2. In situ XRD patterns of the Cu/Al2O3 precursor calcinated in N2 (a,b) and further reduced in H2/Ar (c,d): (a) 50–450 °C with different temperatures, (b) 450 °C with different time, (c) 50–350 °C with different temperatures, (d) 350 °C with different time; In situ XRD patterns of the Cu/Al2O3 precursor calcined in air (e,f) and further reduced in H2/Ar (g,h): (e) 50–450 °C with different temperatures, (f) 450 °C with different time, (g) 50–350 °C with different temperatures, (h) 350 °C with different time.
Molecules 29 02753 g002
Figure 3. (a) N2 adsorption-desorption isotherms and (b) pore size distributions of the Cu/Al2O3 precursor and catalysts.
Figure 3. (a) N2 adsorption-desorption isotherms and (b) pore size distributions of the Cu/Al2O3 precursor and catalysts.
Molecules 29 02753 g003
Figure 4. TEM images: (a) Cu/Al2O3-N2-R, (b) Cu/Al2O3-air-R. HRTEM images: (c) Cu/Al2O3-N2-R, (d) Cu/Al2O3-air-R. HAADF-STEM images: (e) Cu/Al2O3-N2-R, (i) Cu/Al2O3-air-R. EDS mapping images: (fh) Cu/Al2O3-N2-R, (jl) Cu/Al2O3-air-R.
Figure 4. TEM images: (a) Cu/Al2O3-N2-R, (b) Cu/Al2O3-air-R. HRTEM images: (c) Cu/Al2O3-N2-R, (d) Cu/Al2O3-air-R. HAADF-STEM images: (e) Cu/Al2O3-N2-R, (i) Cu/Al2O3-air-R. EDS mapping images: (fh) Cu/Al2O3-N2-R, (jl) Cu/Al2O3-air-R.
Molecules 29 02753 g004
Figure 5. (a) H2-TPR profiles and (b) reduction peaks distribution of the Cu/Al2O3-N2 and Cu/Al2O3-air catalyst.
Figure 5. (a) H2-TPR profiles and (b) reduction peaks distribution of the Cu/Al2O3-N2 and Cu/Al2O3-air catalyst.
Molecules 29 02753 g005
Figure 6. Distribution of Cu species over the employed Cu/Al2O3 catalysts: (a) Cu 2p3/2 XPS spectra; (b) Cu AES spectra.
Figure 6. Distribution of Cu species over the employed Cu/Al2O3 catalysts: (a) Cu 2p3/2 XPS spectra; (b) Cu AES spectra.
Molecules 29 02753 g006
Figure 7. Reusability performance of Cu/Al2O3-N2-R. Reaction conditions: Cu/Al2O3-N2-R (0.1 g), FAL (0.5 g), isopropanol (30 mL), 120 °C, H2 pressure (1 MPa), time (2 h).
Figure 7. Reusability performance of Cu/Al2O3-N2-R. Reaction conditions: Cu/Al2O3-N2-R (0.1 g), FAL (0.5 g), isopropanol (30 mL), 120 °C, H2 pressure (1 MPa), time (2 h).
Molecules 29 02753 g007
Figure 8. In situ DRIFT spectra of FAL adsorption and hydrogenation at 120 °C.
Figure 8. In situ DRIFT spectra of FAL adsorption and hydrogenation at 120 °C.
Molecules 29 02753 g008
Table 1. Physicochemical properties of the Cu/Al2O3 precursor and the employed catalysts.
Table 1. Physicochemical properties of the Cu/Al2O3 precursor and the employed catalysts.
SampleSBET (m2/g) aVpore (cm3/g) aDpore (nm) aCu Loading (wt.%) bDCu
(%) c
SCu
(m2Cu/gcat) c
Cu/Al2O3 precursor 131.50.8020.5---
Cu/Al2O3-N2125.10.8020.74.3--
Cu/Al2O3-N2-R128.50.7619.94.348.314.1
Cu/Al2O3-air125.00.7821.14.3--
Cu/Al2O3-air-R127.20.7720.84.321.56.3
a: Measured by N2 adsorption; b: Determined by ICP-AES; c: Calculated by N2O titration.
Table 2. Surface element compositions on the Cu/Al2O3 catalysts.
Table 2. Surface element compositions on the Cu/Al2O3 catalysts.
Catalyst(Cu0 + Cu+)/Cu(%) aCu+/Cu(%) bCu0/Cu(%) b
Cu/Al2O3-N237.127.49.7
Cu/Al2O3-N2-R60.845.315.5
Cu/Al2O3-air21.813.88.0
Cu/Al2O3-air-R48.838.610.2
a: calculated from the XPS data. b: calculated from the AES data.
Table 3. Hydrogenation performance of FAL to FA over the employed Cu/Al2O3 catalysts.
Table 3. Hydrogenation performance of FAL to FA over the employed Cu/Al2O3 catalysts.
EntryCatalystFAL Conversion (%)FA Selectivity (%)FA Yield (%)
1Cu/Al2O3-N242.799.942.7
2Cu/Al2O3-N2-R99.999.999.9
3Cu/Al2O3-air14.362.28.9
4Cu/Al2O3-air-R26.785.822.9
Reaction conditions: (0.1 g), FAL (0.5 g), isopropanol (30 mL), temperature (120 °C), time (2 h), H2 pressure (1 MPa).
Table 4. A brief comparison of the performance over the representative Cu-based catalysts for FAL hydrogenation to FA using the batch reactor.
Table 4. A brief comparison of the performance over the representative Cu-based catalysts for FAL hydrogenation to FA using the batch reactor.
CatalystReaction ConditionsCon. (%)Sel. (%)Yield (%)Ref.
Cu2Ni1AlOy120 °C, 1.6 MPa H2, 1.5 h98.099.097.0[58]
10Cu-MCM-41-GLY150 °C, 2 MPa H2, 2 h98.180.078.0[59]
Cu/SiO2110 °C, 1 MPa H2, 4 h66.0100.066.0[60]
Cu/C-400170 °C, 2 MPa H2, 3 h99.699.398.9[61]
Cu/MgO180 °C, 0.9 MPa95.094.089.3[62]
CuAlOx220 °C, 3.5 MPa H2, 5 h91.098.489.5[63]
Na-Cu@TS-1110 °C, 1 MPa H2, 2 h93.098.191.2[64]
Cu/Co3O4170 °C, 2 MPa H2, 1 h100.067.067.0[65]
NiCu0.33/C120 °C, 1.5 MPa H2, 12 h96.793.889.6[66]
7NiCu/ZrO2200 °C, 1.5 MPa H2, 4 h99.593.093.0[67]
Ni-Cu110 °C, 6 MPa H2, 2.3 h50.0100.050.0[68]
Cu/γ-Al2O3130 °C, 4 MPa H2, 4 h64.272.346.4[69]
Cu/CeO2210 °C, 0.1 MPa H2836755.6[70]
Cu/Al2O3-N2-R120 °C, 1 MPa H2, 2 h99.999.999.9This work
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

Gao, Y.; Yi, W.; Yang, J.; Jiang, K.; Yang, T.; Li, Z.; Zhang, M.; Liu, Z.; Wu, B. Effect of Calcination Atmosphere on the Performance of Cu/Al2O3 Catalyst for the Selective Hydrogenation of Furfural to Furfuryl Alcohol. Molecules 2024, 29, 2753. https://doi.org/10.3390/molecules29122753

AMA Style

Gao Y, Yi W, Yang J, Jiang K, Yang T, Li Z, Zhang M, Liu Z, Wu B. Effect of Calcination Atmosphere on the Performance of Cu/Al2O3 Catalyst for the Selective Hydrogenation of Furfural to Furfuryl Alcohol. Molecules. 2024; 29(12):2753. https://doi.org/10.3390/molecules29122753

Chicago/Turabian Style

Gao, Yongzhen, Wenjing Yi, Jingyi Yang, Kai Jiang, Tao Yang, Zhihan Li, Meng Zhang, Zhongyi Liu, and Benlai Wu. 2024. "Effect of Calcination Atmosphere on the Performance of Cu/Al2O3 Catalyst for the Selective Hydrogenation of Furfural to Furfuryl Alcohol" Molecules 29, no. 12: 2753. https://doi.org/10.3390/molecules29122753

Article Metrics

Back to TopTop