Next Article in Journal
A Novel Choline Alkali–Thiourea Pretreatment for Enhanced Enzymatic Hydrolysis of Reed Straw
Previous Article in Journal
Selective Ammoxidation of Methanol to Hydrogen Cyanide over Silica-Supported FeMo Oxide Catalysts: Experiments and Kinetic Modeling
Previous Article in Special Issue
Evaluation of the Environmental and Operating Parameters of a Modern Compression-Ignition Engine Running on Vegetable Fuels with a Catalytic Additive
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hydrogenolysis of Biomass-Based Furfuryl Alcohol into 1,2-Pentanediol over Magnesium Oxide-Supported Pt-Y Bimetallic Catalysts

1
College of Chemistry and Chemical Engineering, Northeast Petroleum University, Daqing 163318, China
2
Department of Chemistry and Chemical Engineering, College of Ecology, Lishui University, Lishui 323000, China
*
Authors to whom correspondence should be addressed.
Catalysts 2025, 15(11), 1005; https://doi.org/10.3390/catal15111005
Submission received: 11 September 2025 / Revised: 15 October 2025 / Accepted: 21 October 2025 / Published: 22 October 2025

Abstract

The catalytic synthesis of 1,2-pentanediol from biomass-derived feedstocks is of remarkable significance for addressing current environmental challenges and energy crises. In this paper, a series of Pt-based catalysts were prepared and evaluated in the hydrogenolysis of furfuryl alcohol. The 5Pt0.5Y/MgO provided a 1,2-pentanediol yield of 68.9% and a tetrahydrofurfuryl alcohol yield of 19.8% with 98.1% conversion of furfuryl alcohol, at 200 °C and 2 MPa H2 for 10 h. The promotional effect of yttrium on the catalytic performance was investigated through catalytic reaction and comprehensive characterization. It was found that the reducibility of Pt species was suppressed by the introduction of Y species, resulting in reduced activity compared to the 5Pt/MgO catalyst. However, the addition of Y notably shifted the reaction pathway towards 1,2-pentanediol formation at the expense of tetrahydrofurfuryl alcohol selectivity. This increase in 1,2-pentanediol selectivity was attributed to a higher concentration of medium-strength basic sites on the Y-modified Pt catalyst. Furthermore, the strong interaction between Y2O3, Pt particles, and the MgO support led to high Pt dispersion and stability on the MgO surface, consequently yielding satisfactory recyclability.

1. Introduction

With the escalating depletion of fossil resources and the rising global energy consumption, exploring sustainable and green alternatives has become increasingly imperative. As a renewable and abundantly available carbon resource, biomass is widely recognized as a next-generation chemical feedstock. The utilization of biomass-derived platform compounds for producing high-grade biofuels or high-value-added chemicals has become a prominent research focus in both scientific and industrial communities [1,2]. Among various bioderived platform molecules, furfuryl alcohol is a promising C5 biomass-derived molecule derived from the hydrolysis and dehydration of hemicellulose, and has been commercially produced from furfural [3]. The selective hydrogenation and hydrogenolysis of furfural/furfuryl alcohol has been considered as the most effective route to prepare valuable chemicals [4,5,6,7,8]. In particular, 1,2-pentanediol (1,2-PeD) has attracted significant attention due to its high demand as a crucial raw material for manufacturing of fungicides, printing inks, polyester, and cosmetics [9,10]. Currently, a substantial portion of 1,2-pentanediol production relies on petroleum-derived sources such as n-pentanol, n-valeric acid, and n-pentene [11]. Considering the environmental pollution associated with the petrochemical industry and the limitations of petroleum resources, the hydrogenolysis of furfuryl alcohol provides a green and sustainable route for producing bio-based 1,2-pentanediol, offering considerable economic and environmental benefits [12,13].
It is generally accepted that the formation of 1,2-pentanediol is more challenging than that of 1,5-pentanediol in the hydrogenolysis of furfuryl alcohol [14,15]. As illustrated in Scheme 1, the production of pentanediols from furfural/furfuryl alcohol typically proceeds via two distinct pathways. The first pathway involves the hydrogenolysis of either the C5-O or C2-O bond of furfuryl alcohol, yielding 1,2-pentanediol or 1,5-pentanediol, respectively. Alternatively, furfuryl alcohol can hydrogenated to form tetrahydrofurfuryl alcohol. Due to the higher stability of the secondary carbenium ion structure compared to the primary carbenium ion, the hydrogenolysis of tetrahydrofurfuryl alcohol usually produces 1,5-pentanediol as the dominant product [16]. Consequently, the selective formation of 1,2-pentanediol is primarily achieved through the direct conversion of furfural or furfural alcohol. That is to say, the presence of an unsaturated C=C bond is essential for ring-opening, as the previous reports indicated tetrahydrofurfuryl alcohol cannot undergo the ring-opening reaction under the same conditions used for furfuryl alcohol [15,17].
Considering the reaction pathway mentioned above, preventing the formation of tetrahydrofurfuryl alcohol while ensuring the highly selective cleavage of the C5-O bond is indispensable for the selective synthesis of 1,2-pentanediol. Density functional theory (DFT) calculations and surface science measurements have revealed that the binding energy of the furan ring on Cu (111) is lower than on other metals, indicating reduced furan ring hydrogenation capability of Cu-based catalysts [17]. This makes Cu a preferred active component for producing 1,2-pentanediol from furfuryl alcohol (Table S1). However, copper-catalyzed reaction typically requires high hydrogen pressure (>6 MPa) and produce large amounts of 1,5-pentanediol [10,18,19,20]. Pt- and Ru-based catalysts have also been employed in the production of 1,2-pentanediol, as the interaction between the furan ring and Pt/Ru is significantly weaker than that of Pd or Ni (Table S1) [15,16,21,22]. However, the selective scission of C5-O bonds remains limited on these catalysts due to the competitive parallel reaction of furan ring hydrogenation during the hydrogenolysis of furfuryl alcohol compared to Cu [23]. Overall, it remains a challenging task to achieve higher selectivity to 1,2-pentanediol.
It has been recommended that the association of basic sites along with metal active sites for hydrogenation could be the optimal strategy for the selective production of 1,2-pentanediol [14]. Indeed, basic supports such as MgO, CeO2, ZnO, MnOx, and MgAl-LDHs are commonly employed in the reported catalysts. For instance, Shao et al. [18] found that the strong chemical adsorption of the furan ring in furfuryl alcohol on the basic sites of the CuMgAl catalyst inhibited further hydrogenation to tetrahydrofurfuryl alcohol. Critically, tetrahydrofurfuryl alcohol could not be converted to diols over Cu-based catalysts via hydrogenolysis. More importantly, the basic sites facilitate the opening of the furan ring in furfuryl alcohol, which is the rate-determining step for diols formation. Zhu et al. [10] also emphasized the importance of abundant basic sites for the structure-sensitive ring-opening hydrogenolysis of furfuryl alcohol. They demonstrated that furfuryl alcohol was adsorbed on the basic sites of the support through the O atoms of the furan ring and –CH2OH, inducing ring-opening to form enol intermediates. These intermediates are subsequently hydrogenated to pentanediol over metal nanoparticles. Numerous studies indicated that modifying catalysts with a second metal is an effective strategy for regulating reaction selectivity. Bimetallic catalysts offer distinct structural and electronic environments compared to monometallic catalysts, providing additional level of flexibility for tuning catalytic activity, selectivity and stability [9,14]. Particularly in the hydrogenation or hydrogenolysis of furfural/furfuryl alcohol, bimetallic catalysts can modulate the adsorption modes of substrate molecules, thereby suppressing hydrogenation saturation of the furan ring. Inspired by these findings, this work focused on developing a bimetallic Pt-based catalyst by introducing a second metal to modify the electronic and geometric properties of the active Pt sites. This strategy provides an alternative approach to increase the utilization efficiency of biomass resources.

2. Results

2.1. Catalyst Screening

Basic supports are beneficial to the transformation of furfural or furfuryl alcohol to 1,2-pentanediol by partially suppressing the dehydration of furfuryl alcohol to 2-methylfuran and/or stabilizing the hydrogenolysis intermediates of furfuryl alcohol [14,24,25]. Moreover, basic supports may also prevent the polymerization of furfuryl alcohol, which typically occurs with Lewis acid catalysts [23]. Firstly, we prepared and tested several platinum-based catalysts supported on commercially available metal oxides with basic sites, such as MgO, CeO2, and Al2O3, as summarized in Table 1 (entries 1–3). The amphoteric oxide Al2O3 supported 5 wt% Pt catalyst exhibited the highest activity, achieving complete conversion of furfuryl alcohol with a low 1,2-pentanediol yield of 9.8%. In contrast, the basic oxides MgO and CeO2 supported 5 wt% Pt catalysts showed moderate activity, yielding 20.4% and 15.6% 1,2-pentanediol, respectively. Despite the lower activity, these catalysts demonstrated higher selectivity to 1,2-pentanediol, indicating that basic supports were more favorable for 1,2-pentanediol production via hydrogenolysis of furfuryl alcohol.
Given the highest yield of 1,2-pentanediol obtained over magnesium oxide supported Pt, a series of Pt/MgO catalysts with different Pt loadings (0.5, 1, 3, and 5 wt%) were prepared and tested to investigate the effect of Pt loading amount. As shown in Table 1 (entries 3–6), both the conversion of furfuryl alcohol and the yield of 1,2-pentanediol significantly increased with the Pt loading varied from 0.5 wt% to 5 wt%, while the 1,2-pentanediol selectivity decreased. Typically, higher metal loading provides a higher density of active sites per unit mass of catalyst for the reaction, which accounts for the highest conversion and yield obtained with 5 wt% Pt loading [23]. However, increased loading amount may also lead to the formation of larger metal particles due to the growth of metal crystalline grains. These larger Pt particles are more effective at activating the C=C bond in the furan ring because they tend to facilitate a parallel adsorption mode for furan rings, thereby promoting the hydrogenation of C=C bonds [3,26]. Consequently, a lower 1,2-pentanediol selectivity was obtained over catalyst with 5 wt% Pt loading.
To obtain a higher 1,2-pentanediol yield over MgO supported Pt catalyst, a series of bimetallic catalysts denoted as 5Pt1M/MgO (M = Re, Fe, Cu, Ru, Ni, K, La, and Y; 5 and 1 represent the mass loadings of Pt and M, respectively) were subsequently synthesized and evaluated, as presented in Table 1 (entries 7–14). Based on reducibility, these bimetallic catalysts can be classified into two groups. Readily reducible second metals, such as Re, Fe, Cu, Ru, Ni, are commonly employed in hydrogenation [2,27]. Almost complete conversion of furfuryl alcohol was observed after modification with these reducible metals except for Fe and Cu (Table 1, entries 7–11). However, a large amount of tetrahydrofurfuryl alcohol emerged as the predominant product over these catalysts. Modification with Fe or Cu resulted in lower activity and 1,2-pentanediol yield. These results indicated lack of positive synergy between the bimetals for the desired cleavage of furan C−O bond to form 1,2-pentanediol. In contrast, second metals such as K, La, and Y are difficult to reduce and typically used as basic promoters [28,29]. Although the above metals did not significantly enhance the activity of the Pt/MgO catalyst, higher selectivity to 1,2-pentanediol was obtained compared to the reducible metals (Table 1, entries 12–14). Particularly, the 5Pt1Y/MgO bimetallic catalysts achieved only 36.2% conversion of furfuryl alcohol but a 1,2-pentanediol yield of 17.1%, which was comparable to that of 5Pt/MgO. This demonstrated the pronounced positive effect of yttrium addition in promoting the cleavage of the furan C−O bond to yield 1,2-pentanediol.

2.2. Hydrogenolysis of Furfuryl Alcohol over PtY/MgO

The influence of yttrium on furfuryl alcohol hydrogenolysis was further investigated by varying the Y loading, while the Pt loading was fixed at 5 wt%. To enable a more distinct comparison of the differences among catalysts, the reactions were carried out at an elevated temperature of 200 °C. As shown in Figure 1a, the 5Pt/MgO catalyst without Y additive achieved a conversion of 97.5% under the current conditions, with 1,2-pentanediol yield and selectivity as 46.8% and 48.0%, respectively. With increasing Y loading, the furfuryl alcohol conversion gradually decreased, while the yield and selectivity of 1,2-pentanediol exhibited an initial increase followed by a subsequent decline. Notably, the 5Pt0.5Y/MgO catalyst demonstrated the highest 1,2-pentanediol selectivity of 64.4%. Although there was no significant improvement in 1,2-pentanediol yield compared to the 5Pt/MgO catalyst under identical conditions, the substantial enhancement in selectivity suggested that further optimization of reaction conditions could lead to higher 1,2-pentanediol yield.
To gain further insight into the reaction catalyzed by the Y promoted Pt catalyst, the effects of reaction temperature and hydrogen pressure were also investigated. The effect of reaction temperature was tested in the range of 160 to 240 °C, with other reaction conditions holding constant. As shown in Figure 1b, the conversion of furfuryl alcohol gradually increased from 39.5% to 97.4% as the reaction temperature rose from 160 °C to 240 °C, accompanied by a corresponding enhancement of 1,2-pentanediol yield from 20.3% to 55.0%. However, the yield of tetrahydrofurfuryl alcohol did not increase obviously, and a gradual decline in its selectivity with increasing temperature was observed. Based on the above results, it can be inferred that the hydrogenolysis of furfuryl alcohol (cleavage of furan C−O) could be better activated with higher temperatures, while elevated temperature did not promote C−C hydrogenation of the furan ring [23,30]. Notably, the selectivity for 1,2-pentanediol exhibited an initial increase before a decrease, reaching a maximum of 64.4% at 200 °C. This phenomenon could be attributed to the fact that the excessive temperature of 240 °C also promoted the formation of byproducts such as 2-methylfuran, 2-methyltetrahydrofuran, n-pentanol, and 1,5-pentanediol, which was collectively contributed to the decrease of 1,2-pentanediol selectivity.
The effect of hydrogen pressure on the hydrogenolysis of furfuryl alcohol over 5Pt0.5Y/MgO was investigated at 200 °C. As shown in Figure 1c, the conversion of furfuryl alcohol increased significantly from 23.6% to 78.4% as the reaction pressurized from 1.0 to 3.0 MPa. Both the yield and selectivity of tetrahydrofurfuryl alcohol exhibited remarkable enhancement throughout this pressure range. Although the yield of 1,2-pentanediol increased due to the improved conversion, its selectivity showed a gradual decline. Notably, when the pressure reached 3.0 MPa, the influence of hydrogen pressure on conversion became less significant, as both tetrahydrofurfuryl alcohol and 1,2-pentanediol yields no longer changed appreciably. Based on the above results, it was concluded that the optimal reaction pressure for the production of 1,2-pentanediol was determined to be 2.0 MPa.
The product evolution with respect to reaction time over 5Pt0.5Y/MgO was traced at the optimal reaction conditions (200 °C, 2 MPa). As seen in Figure 1d, the conversion of furfuryl alcohol increased rapidly within the first 6 h, accompanied by a remarkable increase in the yields of tetrahydrofurfuryl alcohol and 1,2-pentanediol. After reaction for 10 h, a maximum 1,2-pentanediol yield of 68.9% was obtained with nearly complete conversion of furfuryl alcohol (98.1%). The reaction order was determined from the temporal concentration profile of furfuryl alcohol hydrogenolysis over 5Pt0.5Y/MgO (Figure S1). It showed that ln(CFA) was proportional to the reaction time, proving it is a first-order reaction. The calculated kinetic constant was 3.77 × 10−2, which is more than three times the value reported by Zhu et al. [10] for a Cu@MgO-La2O3 catalyst. This indicated that the hydrogenolysis reaction proceeded much faster on our Pt-based catalyst than on the reported Cu-based system. Further prolonging the reaction time to 12 h resulted in no substantial variation in the yields of tetrahydrofurfuryl alcohol and 1,2-pentanediol, indicating that tetrahydrofurfuryl alcohol could hardly be further converted into 1,2-pentanediol under the current catalytic system and reaction conditions [17,23]. Furthermore, it should be noted that extended reaction durations resulted in slight increases in the yields of 2-methyltetrahydrofuran, n-pentanol, and 1,5-pentanediol (1,5-PeD), along with a gradual accumulation of undetectable by-products. Considering both conversion efficiency and product selectivity, we concluded that 10 h represented the optimal reaction duration.

2.3. Catalyst Recyclability

The recyclability of 5Pt0.5Y/MgO for the hydrogenolysis of furfuryl alcohol was tested under 200 °C and 2 MPa for 4 h.
As shown in Figure 2, although a gradual decrease in furfuryl alcohol conversion and 1,2-pentanediol yield was observed over repeated cycles, the 5Pt0.5Y/MgO catalyst was considered relatively stable with taking account of inevitable loss of catalyst amount during recovery and factors such as oxidation.

2.4. Characterization Results

The textural characteristics of the MgO support and catalysts were examined by N2 adsorption–desorption analysis, and the results are summarized in Table 2. The BET surface area, total pore volume, and average pore size of the MgO support were 29.26 m2·g−1, 0.309 cm3·g−1, and 17.68 nm, respectively. Upon loading with 5% Pt nanoparticles, the surface area decreased to 21.32 m2·g−1, and the total pore volume and pore size decreased to 0.155 cm3·g−1 and 3.79 nm, respectively. The dramatic decrease could be attributed to structure change of MgO or pore blocking by Pt species [31,32]. However, the surface area and total volume of 5Pt0.5Y/MgO slightly increased to 28.48 m2·g−1, and 3.82 nm, respectively. This rebound in surface area and pore volume might be due to the well-dispersed metal species [33]. Compared with the fresh 5Pt0.5Y/MgO catalyst, only negligible changes in the textural characteristics of the used catalyst were observed, indicating the structural stability of the catalyst under the reaction conditions.
The X-ray diffraction (XRD) patterns for MgO, 5Pt/MgO and 5Pt0.5Y/MgO are shown in Figure 3a. It can be seen that all samples showed diffraction peaks at 2θ = 36.9, 42.9, 62.2, 74.6, and 78.5°, corresponding to the (111), (200), (220), (311), and (222) crystal planes of MgO (PDF # 45-0946), respectively [22,34]. Additionally, diffraction peaks at 2θ = 18.6, 38.0, 50.9, 58.6, and 68.3° were also observed in the XRD patterns of 5Pt/MgO and 5Pt0.5Y/MgO, corresponding to the (001), (101), (102), (110), and (103) crystal planes of Mg(OH)2, respectively, indicating the formation of Mg(OH)2 phase (PDF # 07-0239). It is well known that the conversion of MgO to Mg(OH)2 in wet environment is thermodynamically favorable and this conversion can be accelerated by elevated temperatures [35,36]. Estrada et al. [31] have reported the complete transformation of MgO into Mg(OH)2 via hydration of MgO in aqueous solution during the Au/MgO catalyst preparation. They suggested that the process seemed to be accelerated by the presence of the strong HAuCl4 acid at the beginning of the catalyst synthesis procedure. Consequently, the formation of Mg(OH)2 was inevitable during the impregnation and drying steps in the preparation of the 5Pt/MgO and 5Pt0.5Y/MgO catalysts. As also noted by Estrada et al., the phase transformation resulted in a reduction of the specific surface area, pore volume and average pore size. Such textural deterioration can be attributed to multiple factors like pore collapse, crystal structure transition and volume expansion [37,38].
Notably, no apparent peaks associated with Pt species were observed, which could be ascribed to the low loadings or high dispersion of the Pt metal [5,33]. Figure 3b shows the XRD patterns of the used 5Pt/MgO and 5Pt0.5Y/MgO. Both samples exhibited similar diffraction patterns to the fresh counterparts, with no distinct diffraction peaks of Pt species detected, indicating that Pt species remained stable on the surface of MgO support.
To investigate potential carbon deposition on the catalyst surface, thermogravimetry-derivative thermogravimetry (TG-DTG) analysis was performed on the fresh and used catalysts. The analysis involved heating the samples from room temperature to 800 °C in an air atmosphere. As shown in Figure 4a, the TG profile of fresh 5Pt0.5Y/MgO exhibited a weight loss at approximately 92.0 °C, which was attributed to the removal of physically adsorbed moisture [39]. Another weight loss stage were observed at approximately 355.4 °C, assigned to the thermal decomposition of Mg(OH)2 into MgO and water [40]. The TG profile of used 5Pt0.5Y/MgO in Figure 4b showed a similar pattern to the fresh catalyst. The only difference was the appearance of an additional weight loss event at approximately 295.0 °C, alongside the weight loss at approximately 100.8 °C (removal of adsorbed moisture) and 375.3 °C (thermal decomposition of Mg(OH)2). We ascribed the weight loss at approximately 295.0 °C to the removal of adsorbed small organic molecules [40]. Nevertheless, the total weight loss for both catalysts was nearly identical, suggesting negligible carbon deposition on the used catalysts. Zhang et al. [23] reported that the basic support suppressed the polymerization of furfuryl alcohol in the hydrogenolysis process, which also contributed to the enhanced selectivity to 1,2-pentanediol. It was reasonably assumed that the MgO support with extremely weak acidity is unlikely to induce polymerization reactions. Moreover, the above-mentioned results verified that the catalyst did not decompose under the reaction conditions, demonstrating the good thermal stability of the 5Pt0.5Y/MgO catalyst.
To investigate the distribution of Pt species on the MgO support, the morphological characteristics of 5Pt/MgO and 5Pt0.5Y/MgO were investigated by transmission electron microscopy (TEM) and high-angle annular dark field scanning transmission electron microscopy (HAADF-STEM). For the 5Pt/MgO and 5Pt0.5Y/MgO catalysts, the prominent dark regions in the bright field TEM images (Figure 5a,b and Figure S2a–d) corresponding to Pt particles (verified by the lattice stripe of Pt (111) with a spacing of around 0.227 nm in Figure 5d,e) seemed to be evenly distributed on the support [33,41]. The average particle sizes of the 5Pt/MgO and 5Pt0.5Y/MgO catalysts were determined to be 2.56 and 2.02 nm, respectively, indicating that the introduction of Y inhibited the aggregation of Pt particles, accordingly resulting in improved dispersion and smaller particles sizes. This finding explained the absence of Pt diffraction peaks in the XRD patterns of both 5Pt/MgO and 5Pt0.5Y/MgO catalysts. For the used 5Pt0.5Y/MgO, Pt particles remained well dispersed on the surface of the support, even though the average particle size increased slightly to 2.27 nm (Figure 5c,f). However, it is worth noting that clear lattice fringes corresponding to Y species were not identified in any of the samples. To verify the distribution of Y species, the 5Pt0.5Y/MgO catalyst was also characterized by HAADF-STEM and STEM-EDS mapping. As seen from Figure 5g, no obvious Y-containing aggregates were observed, confirming the homogeneous distribution of Y species [42,43].
In situ XPS was employed to investigate the surface elemental composition of the reduced 5Pt/MgO and 5Pt0.5Y/MgO. As shown in Figure 6a, Mg 1s, Pt 4f, and O 1s peaks were observed in both survey spectra of the two catalysts. However, the Y 3d peak was barely visible for the 5Pt0.5Y/MgO catalyst, likely owing to the significantly lower Y content compared to other elements. Figure 6b showed the high-resolution Pt 4f spectra for the two samples. Each spectrum exhibited the typical Pt 4f7/2 and Pt 4f5/2 spin–orbit components separated by 3.3 eV. The Pt 4f7/2 peak for the 5Pt/MgO was deconvoluted into two distinct components at binding energies of 70.6 eV and 71.8 eV, indicating the coexistence of Pt0 and Pt2+ species [44,45]. Similarly, the Pt 4f5/2 peaks could be deconvoluted into same patterns, further confirming the presence of Pt0 and Pt2+ species. In contrast, the Pt 4f7/2 peaks for the 5Pt0.5Y/MgO catalysts could be decomposed into two components at binding energies of 70.8 eV and 72.0 eV, displaying a shift to the higher binding energy. This change in the Pt electronic state suggested that the Pt atoms in this catalyst were partially positively charged [45], which provided critical evidence for the existence of Pt-Y synergy [41]. Notably, the Pt0/Pt2+ ratio in the 5Pt0.5Y/MgO catalyst was significantly lower than that in the 5Pt/MgO catalyst, indicating a negative effect of Y addition on the reduction of Pt species [21]. Therefore, the relatively lower catalytic activity of the Y-modified catalyst was likely attributed to the declined Pt reduction extent. The Y 3d spectrum of the 5Pt0.5Y/MgO catalyst in Figure 6c displayed two characteristic peaks at binding energies of 157.8 eV and 159.8 eV, corresponding to the Y 3d5/2 and Y 3d3/2 orbitals of the Y2O3 phase, respectively [46,47]. This confirmed that yttrium existed as Y2O3 within the 5Pt0.5Y/MgO catalyst and remained in an unreduced state during the reduction process.
The basic properties of MgO, 5Pt/MgO, and 5Pt0.5Y/MgO were probed by CO2-TPD, as shown in Figure 7. The MgO support exhibited two desorption peaks centered between 50~300 °C and 500~800 °C. The desorption peak between 50~300 °C could be ascribed to the weak basic sites or physiosorbed CO2 (desorption peak below 100 °C) [48]. The desorption peaks located in the temperature higher than 500 °C could be closely related to the strong basic sites [32]. The intense CO2 desorption peak in the high-temperature region reflected the abundance of strong basic sites on MgO. After loading Pt, the strong basic sites were significantly suppressed on both 5Pt/MgO and 5Pt0.5Y/MgO, as the main desorption peak shifted to lower temperatures region (between 300~700 °C). This indicated the formation of abundant moderate-strength basic sites. The decrease of strong basic sites on MgO surface and the concomitant increase of moderate-strength basic sites could be ascribed to the hydration reaction of MgO forming Mg(OH)2 [40]. This was also confirmed by the CO2-TPD profile of Mg(OH)2 (Figure S10), as the desorption peaks in the high-temperature region were also observed between 300 and 700 °C. Notably, upon the introduction of Y species, the desorption peaks for 5Pt0.5Y/MgO were broader and exhibited a larger integrated area compared to 5Pt/MgO. These features indicated that the addition of Y enhanced the surface basicity of the catalyst in terms of both increased strength and higher density of basic sites [40]. Consequently, the 5Pt0.5Y/MgO catalyst exhibited superior selectivity toward 1,2-pentanediol compared to 5Pt/MgO, which owed to this enhanced basicity.
H2-TPR analysis was performed to investigate the interaction between the active metal and the support. As shown in Figure 8, the 5Pt/MgO and 5Pt0.5Y/MgO catalysts displayed two hydrogen consumption peaks within the temperature ranges of 150~350 °C and 400~600 °C. These peaks can be assigned to the reduction of Pt species, exhibiting weak and strong interactions with the support, respectively [21,49]. With the introduction of Y, the hydrogen consumption peak in the 150~350 °C region shifted from 220 °C to a higher temperature of 257 °C. It is noteworthy that the 5Pt0.5Y/MgO samples still exhibited a minor shoulder at 220 °C, which was similar to the TPR results of La-modified Co/SBA-15 catalysts reported by Mao et al. [50]. Given the low yttrium loading in 5Pt0.5Y/MgO, we attribute this feature to interaction of Y2O3 with partial Pt species, which was further supported by the incomplete spatial overlap of Pt and Y species observed in STEM-EDS mapping. Shi et al. [51] reported that the reduction peak of PdY/Al2O3 is much higher than that of Pd/Al2O3, indicating a strong interaction between Pd- and Y-modified support. Similarly, Liu et al. [52] also reported that interaction between iron species and Y2O3 in Fe2O3–Y2O3/Al2O3 catalysts shifted the reduction temperature of iron species to higher values and improved the catalyst stability. Therefore, our H2-TPR results suggested that the introduction of yttrium species enhanced Pt-support interaction in the 5Pt0.5Y/MgO catalyst. However, the increased reduction temperature also indicated that the reducibility of Pt species was suppressed by the Y species. This finding was consistent with the XPS characterization, which showed a lower metallic Pt0 content in the 5Pt0.5Y/MgO catalyst.

3. Discussion

A series of MgO supported Pt catalysts were prepared and evaluated in the hydrogenolysis of furfuryl alcohol to 1,2-pentanediol. The bimetallic 5Pt0.5Y/MgO catalyst exhibited higher selectivity towards 1,2-pentanediol compared to 5Pt/MgO, despite its lower activity, making it the more efficient catalyst. Contrary to some previous reports, adding yttrium decreased the activity of the Pt/MgO catalyst. BET and TEM analyses revealed that Y species enhanced the dispersion of Pt nanoparticles on the MgO support, leading to increases in both the specific surface area and metal dispersion. However, the reduction ability of the catalyst was influenced by the dispersion of Pt and the interaction between the support material and the Pt component. The higher the dispersion of Pt, the stronger its interaction with MgO, resulting in a higher reduction temperature [50,51]. Additionally, the electronic interactions between Y2O3 and Pt species also suppressed the reduction of the Pt species [53,54]. Consequently, the catalytic activity of Pt/MgO was diminished with the introduction of Y species. This conclusion was confirmed by XPS and H2-TPR analyses. These findings are in accordance with previous reports, in which alkaline promoters commonly reduced the reducibility of active metals, thereby lowering the hydrogenation activity of catalysts [55,56].
Based on the product distribution, the reaction selectivity is determined by parallel competitive pathways: furan ring hydrogenation versus ring-opening. The critical role of base sites in the conversion of furfuryl alcohol to pentanediols has been widely verified [57]. Mizugaki et al. [24] proposed that surface basic sites strongly interacted with the hydroxymethyl moiety of furfuryl alcohol, giving an alkoxide species, while the furan ring was adsorbed on Pt nanoparticles in a vertical orientation. The cooperative behavior between metal particles and basic sites facilitated an adsorption configuration which promoted the selective scission of C5-O bond. In situ FTIR experiments also demonstrated that furfuryl alcohol was adsorbed on the alkaline sites of the carrier via the O atoms of the furan ring and CH2OH group, inducing ring-opening to generate enol intermediates [10]. These intermediates are subsequently hydrogenated to 1,2-pentanediol over the metal nanoparticles. In summary, basic sites could not only inhibit the furan ring saturation to form tetrahydrofurfuryl alcohol, but also play a critical role in promoting selective cleavage of C5-O bond to produce 1,2-pentanediol. In this regard, the introduction of Y provides an effective strategy to enhance selectivity towards 1,2-pentanediol over tetrahydrofurfuryl alcohol. This enhancement is attributed to the increased basicity of 5Pt0.5Y/MgO compared to 5Pt/MgO, as confirmed by CO2-TPD results.
In regard to the stability, H2-TPR analysis revealed a strong interaction between Y2O3 and both Pt particles and MgO. The addition of Y enhanced the stability of the metal particles on the support, consistent with the results of XRD and TEM. Furthermore, TG-DTG analysis confirmed negligible carbon deposition on the catalyst during the reaction, further improving the catalyst stability.

4. Materials and Methods

4.1. Materials

Chemicals and solvents, including furfuryl alcohol (FA, 98%), tetrahydrofurfuryl alcohol (THFA, 98%), 2-methylfuran (2-MF, 98%), 2-methyltetrahydrofuran (2-MTHF, 99%), 1,2-pentanediol (1,2-PeD, 98%), 1,5-pentanediol (1,5-PeD, 97%), and isopropanol (IPA, 99.5) were purchased from Shanghai Macklin Biochemical Co., Ltd. (Shanghai, China). Metal precursors, such as H2PtCl6·6H2O (Pt ≥ 37.5%), NH4ReO4 (≥99%), Cu(NO3)2·3H2O (≥99.5%), RuCl3·3H2O (≥99.9%), Ni(NO3)2·6H2O (≥99%) were also purchased from Shanghai Macklin Biochemical Co., Ltd. Y(NO3)3·6H2O (99.99% metal basis), La(NO3)3·6H2O (99.99% metal basis), and KNO3 (≥99.997% metals basis) were purchased from Aladdin Reagent Co., Ltd. (Shanghai, China). MgO, CeO2, γ-Al2O3 were also sourced from Shanghai Macklin Biochemical Co., Ltd. Hydrogen and nitrogen gases were provided by Lishui Li՛an Special Gas Co. Ltd. (Lishui, China). with purity higher than 99.99%. All the materials and chemicals were used as received without any further treatment.

4.2. Catalyst Preparation and Characterization

The catalysts were prepared by an incipient wetness co-impregnation method. Typically, MgO (1.0 g) was first impregnated with an aqueous solution that contained H2PtCl6·6H2O (0.133 g) and Y(NO3)3·6H2O (0.022 g). The mixture was impregnated for 24 h and dried at 110 °C overnight. The catalyst was reduced in a flow of 10% H2/Ar from RT to 300 °C at a rate of 10 °C·min−1 then kept for another 2 h. This catalyst was denoted as 5Pt1Y/MgO (5 wt% Pt−1 wt% Y/MgO). The other catalysts were prepared in the same procedure. Characterization details were included in the Supplementary Materials for duplicate-checking reasons.

4.3. Typical Procedures of Catalytic Reactions

Typically, 20 mL of isopropanol, 0.5 g of furfural and 0.1 g of catalyst were introduced into a 50 mL stainless steel autoclave. The reactor was sealed and purged with N2 (five times) to remove air, and hydrogen with designated pressure was introduced. Afterward, the reactor was heated to 200 °C and maintained for 4 h with continuous stirring at 800 rpm. At the end of the reaction, the reactor was cooled to RT, and the reaction bulk was sampled and subjected to product analysis. For the recycling of the catalyst, the catalyst was centrifuged after the reaction, washed with isopropanol and returned to the reactor for the next run.

4.4. Analytical Method

The products were analyzed with a Shimadzu GC-2010 (Shimadzu Corporation, Kyoto, Japan) gas chromatograph equipped with a WondaCap 5 capillary column (30.0 m × 0.25 mm × 0.25 µm) (Shimadzu Corporation, Kyoto, Japan) and a flame ionization detector (FID) (Shimadzu Corporation, Kyoto, Japan). The initial column temperature was 50 °C, and then the temperature was ramped to 250 °C at a rate of 10 °C·min−1 and held for 5 min. Both the injector temperature and the detector temperature were set at 260 °C. The products were quantitatively analyzed by standard curves of each compound. Structure of the reactants was further identified by Shimadzu QP 2010 Plus gas chromatography–mass spectrometry (GC-MS) as well as standard retention times of known compounds. The conversion was calculated as the moles of furfuryl alcohol of consumption divided by the moles of furfuryl alcohol before reaction. The selectivity was defined by moles of the products divided by the moles of converted furfuryl alcohol.
Conversion = n FA 0 n FA 1 n FA 0 × 100 %
Selectivity = n i n FA 0 n FA 1 × 100 %
Yield = n i n FA 0 × 100 %
Here, n FA 0 is the initial moles of furfuryl alcohol in the solution, ni and n FA 1 refer to the moles of the products and furfuryl alcohol after the reaction, respectively.

5. Conclusions

In conclusion, we developed supported Pt-Y bimetallic catalysts for the hydrogenolysis of furfuryl alcohol into 1,2-pentanediol. An optimal Pt:Y weight ratio of 5:0.5 showed the best performance, providing the highest 1,2-pentanediol yield of 68.9% and a tetrahydrofurfuryl alcohol yield of 19.8% with 98.1% conversion of furfuryl alcohol, at 200 °C and 2 MPa H2 for 10 h. Catalyst evaluation revealed that the MgO supported Pt was more favorable for 1,2-pentanediol production via hydrogenolysis of furfuryl alcohol, while the introduction of Y species enhanced the selectivity to 1,2-pentanediol. Catalyst characterization revealed that Y addition suppressed Pt reduction, thereby diminishing catalytic activity, but concurrently increased the density of medium-strength basic sites, promoting 1,2-pentanediol selectivity. Furthermore, the interaction between Y, the MgO support, and Pt species improved the dispersion of platinum and surface stability, resulting in satisfactory catalyst stability during reuse.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal15111005/s1, Characterization Details; Figure S1: Data plot for ln(CFA) versus reaction time for furfuryl hydrogenolysis over 5Pt0.5Y/MgO catalysts; Figure S2: Representative TEM images of 5Pt/MgO (a) (b) and 5Pt0.5Y/MgO (c) (d).; Figure S3: Mass spectra of the main components in the hydrogenolysis of furfuryl alcohol; Figure S4: Photographs of the reaction mixture and catalyst before and after the reaction; Figure S5: Gas chromatogram of the crude reaction mixture concentrated reaction mixture; Figures S6–S9: NMR spectrums of crude and concentrated mixture; Figure S10: CO2-TPD profiles of Mg(OH)2. Table S1: Reported hydrogenolysis of furfuryl alcohol into 1,2-pentanediol. References [3,9,10,11,15,16,18,19,20,21,22,23,30,58,59,60,61,62] are cited in the Supplementary Materials.

Author Contributions

K.Z.: Conceptualization, Formal analysis, Investigation, Writing—original draft, Visualization, Funding acquisition. J.X.: data curation, Validation, Writing—review and editing. S.G.: Supervision, Project administration, Funding acquisition. H.W.: Supervision, Project administration. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by National Natural Science Foundation of China (22408142), the funding of Guangdong Basic and Applied Basic Research Foundation (2020A1515110127), and Project of Lishui Science and Technology Bureau (2023GYX08, 2023KJTP07).

Data Availability Statement

The original contributions presented in this study are included in the article/Supplementary Materials. Further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Mika, L.T.; Csefalvay, E.; Nemeth, A. Catalytic Conversion of Carbohydrates to Initial Platform Chemicals: Chemistry and Sustainability. Chem. Rev. 2018, 118, 505–613. [Google Scholar] [CrossRef]
  2. 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]
  3. Zhu, Y.; Zhao, W.; Zhang, J.; An, Z.; Ma, X.; Zhang, Z.; Jiang, Y.; Zheng, L.; Shu, X.; Song, H.; et al. Selective Activation of C–OH, C–O–C, or C=C in Furfuryl Alcohol by Engineered Pt Sites Supported on Layered Double Oxides. ACS Catal. 2020, 10, 8032–8041. [Google Scholar] [CrossRef]
  4. Chen, L.; Ye, J.; Yang, Y.; Yin, P.; Feng, H.; Chen, C.; Zhang, X.; Wei, M.; Truhlar, D.G. Catalytic Conversion Furfuryl Alcohol to Tetrahydrofurfuryl Alcohol and 2-Methylfuran at Terrace, Step, and Corner Sites on Ni. ACS Catal. 2020, 10, 7240–7249. [Google Scholar] [CrossRef]
  5. Zhou, K.; Chen, J.; Cheng, Y.; Chen, Z.; Kang, S.; Cai, Z.; Xu, Y.; Wei, J. Enhanced Catalytic Transfer Hydrogenation of Biomass-Based Furfural into 2-Methylfuran over Multifunctional Cu–Re Bimetallic Catalysts. ACS Sustain. Chem. Eng. 2020, 8, 16624–16636. [Google Scholar] [CrossRef]
  6. Kalong, M.; Praikaew, W.; Ratchahat, S.; Chaiwat, W.; Koo-amornpattana, W.; Klysubun, W.; Limphirat, W.; Assabumrungrat, S.; Srifa, A. Continuous Furfural Hydrogenolysis into 2-Methylfuran and 2-Methyltetrahydrofuran over Cu/γ–Al2O3 with ReOx and WOx as Catalyst Boosters. Energy Fuels 2024, 38, 9836–9848. [Google Scholar] [CrossRef]
  7. Cao, C.; Guan, W.; Liu, Q.; Li, L.; Su, Y.; Liu, F.; Wang, A.; Zhang, T. Selective hydrogenolysis of furfural to 1,2-pentanediol over a Pt–Fe/MT catalyst under mild conditions. Green Chem. 2024, 26, 6511–6519. [Google Scholar] [CrossRef]
  8. 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 Environ. 2023, 320, 121971. [Google Scholar] [CrossRef]
  9. Nimbalkar, A.S.; Oh, K.R.; Hong, D.Y.; Park, B.G.; Lee, M.; Hwang, D.W.; Awad, A.; Upare, P.P.; Han, S.J.; Hwang, Y.K. Continuous production of 1,2-pentanediol from furfuryl alcohol over highly stable bimetallic Ni–Sn alloy catalysts. Green Chem. 2024, 26, 11164–11176. [Google Scholar] [CrossRef]
  10. Zhu, Y.; Li, B.; Zhao, C. Cu nanoparticles supported on core–shell MgO-La2O3 catalyzed hydrogenolysis of furfuryl alcohol to pentanediol. J. Catal. 2022, 410, 42–53. [Google Scholar] [CrossRef]
  11. Weng, Y.; Meng, S.; Zhu, W.; Zhang, M.; Sun, Q.; Zhang, Y. Synergistic hydrogenolysis of biomass furfuryl alcohol over Ru/solid base catalysts in hydrothermal reaction environment. J. Fuel Chem. Technol. 2021, 49, 1867–1874. [Google Scholar] [CrossRef]
  12. Sun, X.; Wen, B.; Wang, F.; Zhang, W.; Zhao, K.; Liu, X. Research advances on the catalytic conversion of biomass-derived furfural into pentanediols. Catal. Commun. 2024, 187, 106864. [Google Scholar] [CrossRef]
  13. Khemthong, P.; Yimsukanan, C.; Narkkun, T.; Srifa, A.; Witoon, T.; Pongchaiphol, S.; Kiatphuengporn, S.; Faungnawakij, K. Advances in catalytic production of value-added biochemicals and biofuels via furfural platform derived lignocellulosic biomass. Biomass Bioenergy 2021, 148, 106033. [Google Scholar] [CrossRef]
  14. Upare, P.P.; Kim, Y.; Oh, K.R.; Han, S.J.; Kim, S.K.; Hong, D.Y.; Lee, M.; Manjunathan, P.; Hwang, D.W.; Hwang, Y.K. A Bimetallic Ru3Sn7 Nanoalloy on ZnO Catalyst for Selective Conversion of Biomass-Derived Furfural into 1,2-Pentanediol. ACS Sustain. Chem. Eng. 2021, 9, 17242–17253. [Google Scholar] [CrossRef]
  15. Wang, J.; Zhu, S.; He, Y.; Fan, G.; Li, X.; Jia, X.; Dong, M.; Fan, W. Pt/MgAlO bifunctional catalysts with various Mg/Al ratios for selective hydrogenation of furfural alcohol to 1,2-pentanediol. Catal. Today 2024, 433, 114647. [Google Scholar] [CrossRef]
  16. Ma, R.; Wu, X.P.; Tong, T.; Shao, Z.J.; Wang, Y.; Liu, X.; Xia, Q.; Gong, X.Q. The Critical Role of Water in the Ring Opening of Furfural Alcohol to 1,2-Pentanediol. ACS Catal. 2016, 7, 333–337. [Google Scholar] [CrossRef]
  17. Shao, Y.; Guo, M.; Wang, J.; Sun, K.; Zhang, L.; Zhang, S.; Hu, G.; Xu, L.; Yuan, X.; Hu, X. Selective Conversion of Furfural into Diols over Co-Based Catalysts: Importance of the Coordination of Hydrogenation Sites and Basic Sites. Ind. Eng. Chem. Res. 2021, 60, 10393–10406. [Google Scholar] [CrossRef]
  18. Shao, Y.; Wang, J.; Du, H.; Sun, K.; Zhang, Z.; Zhang, L.; Li, Q.; Zhang, S.; Liu, Q.; Hu, X. Importance of Magnesium in Cu-Based Catalysts for Selective Conversion of Biomass-Derived Furan Compounds to Diols. ACS Sustain. Chem. Eng. 2020, 8, 5217–5228. [Google Scholar] [CrossRef]
  19. Liu, H.; Huang, Z.; Zhao, F.; Cui, F.; Li, X.; Xia, C.; Chen, J. Efficient hydrogenolysis of biomass-derived furfuryl alcohol to 1,2- and 1,5-pentanediols over a non-precious Cu–Mg3AlO4.5 bifunctional catalyst. Catal. Sci. Technol. 2016, 6, 668–671. [Google Scholar] [CrossRef]
  20. Liu, H.; Huang, Z.; Kang, H.; Xia, C.; Chen, J. Selective hydrogenolysis of biomass-derived furfuryl alcohol into 1,2- and 1,5-pentanediol over highly dispersed Cu-Al2O3 catalysts. Chin. J. Catal. 2016, 37, 700–710. [Google Scholar] [CrossRef]
  21. Li, X.; Pang, J.; Zhang, J.; Li, X.; Jiang, Y.; Su, Y.; Li, W.; Zheng, M. Tuning the Reaction Selectivity over MgAl Spinel-Supported Pt Catalyst in Furfuryl Alcohol Conversion to Pentanediols. Catalysts 2021, 11, 415. [Google Scholar] [CrossRef]
  22. Yang, Y.; Liu, Q.; Liu, Z. Selective Hydrogenolysis of Furfuryl Alcohol to Pentanediol over Pt Supported on MgO. Catalysts 2024, 14, 233. [Google Scholar] [CrossRef]
  23. Zhang, B.; Zhu, Y.; Ding, G.; Zheng, H.; Li, Y. Selective conversion of furfuryl alcohol to 1,2-pentanediol over a Ru/MnOx catalyst in aqueous phase. Green Chem. 2012, 14, 3402–3409. [Google Scholar] [CrossRef]
  24. Mizugaki, T.; Yamakawa, T.; Nagatsu, Y.; Maeno, Z.; Mitsudome, T.; Jitsukawa, K.; Kaneda, K. Direct Transformation of Furfural to 1,2-Pentanediol Using a Hydrotalcite-Supported Platinum Nanoparticle Catalyst. ACS Sustain. Chem. Eng. 2014, 2, 2243–2247. [Google Scholar] [CrossRef]
  25. Liu, Y.; Zhou, K.; Lu, M.; Wang, L.; Wei, Z.; Li, X. Acidic/Basic Oxides-Supported Cobalt Catalysts for One-Pot Synthesis of Isophorone Diamine from Hydroamination of Isophorone Nitrile. Ind. Eng. Chem. Res. 2015, 54, 9124–9132. [Google Scholar] [CrossRef]
  26. Kliewer, C.J.; Aliaga, C.; Bieri, M.; Huang, W.; Tsung, C.-K.; Wood, J.B.; Komvopoulos, K.; Somorjai, G.A. Furan Hydrogenation over Pt(111) and Pt(100) Single-Crystal Surfaces and Pt Nanoparticles from 1 to 7 nm: A Kinetic and Sum Frequency Generation Vibrational Spectroscopy Study. J. Am. Chem. Soc. 2010, 132, 13088–13095. [Google Scholar] [CrossRef]
  27. Luo, J.; Liang, C. Rhenium in Heterogeneous Catalysis: A Rising Star for Hydrogenation Reactions. ACS Catal. 2024, 14, 7032–7049. [Google Scholar] [CrossRef]
  28. Rukini, A.; Rhamdhani, M.A.; Brooks, G.A.; Van den Bulck, A. Metals Production and Metal Oxides Reduction Using Hydrogen: A Review. J. Sustain. Metall. 2022, 8, 1–24. [Google Scholar] [CrossRef]
  29. Sun, C.; Świrk Da Costa, K.; Wang, Y.; Li, L.; Fabbiani, M.; Hulea, V.; Rønning, M.; Hu, C.; Da Costa, P. Unraveling catalytic properties by yttrium promotion on mesoporous SBA-16 supported nickel catalysts towards CO2 methanation. Fuel 2022, 317, 122829. [Google Scholar] [CrossRef]
  30. Wang, X.; Weng, Y.; Zhao, X.; Xue, X.; Meng, S.; Wang, Z.; Zhang, W.; Duan, P.; Sun, Q.; Zhang, Y. Selective Hydrogenolysis and Hydrogenation of Furfuryl Alcohol in the Aqueous Phase Using Ru–Mn-Based Catalysts. Ind. Eng. Chem. Res. 2020, 59, 17210–17217. [Google Scholar] [CrossRef]
  31. Estrada, M.; Costa, V.V.; Beloshapkin, S.; Fuentes, S.; Stoyanov, E.; Gusevskaya, E.V.; Simakov, A. Aerobic oxidation of benzyl alcohol in methanol solutions over Au nanoparticles: Mg(OH)2 vs MgO as the support. Appl. Catal. A Gen. 2014, 473, 96–103. [Google Scholar] [CrossRef]
  32. Ding, J.; Li, Y.; Zhang, X.; Zhang, Z.; Zhang, Y.; Wang, Y.; Yu, G.; Wang, Z.; Guo, X.; Wang, Y.; et al. Study on the homologous effects on the performance of alkali metal-doping MgO-based adsorbents for CO2 capture. Sep. Purif. Technol. 2025, 355, 129673. [Google Scholar] [CrossRef]
  33. Zhang, J.; Zhao, C.; Li, C.; Li, S.; Tsang, C.; Liang, C. The role of oxophilic Mo species in Pt/MgO catalysts as extremely active sites for enhanced hydrodeoxygenation of dibenzofuran. Catal. Sci. Technol. 2020, 10, 2948–2960. [Google Scholar] [CrossRef]
  34. Wang, J.; Zhao, J.; Fu, J.; Miao, C.; Jia, S.; Yan, P.; Huang, J. Highly selective hydrogenation of 5-hydroxymethylfurfural to 2,5-bis(hydroxymethyl)furan over metal-oxide supported Pt catalysts: The role of basic sites. Appl. Catal. A Gen. 2022, 643, 118762. [Google Scholar] [CrossRef]
  35. Newberg, J.T. Surface Thermodynamics and Kinetics of MgO(100) Terrace Site Hydroxylation. J. Phys. Chem. C 2014, 118, 29187–29195. [Google Scholar] [CrossRef]
  36. Bracco, J.N.; Camacho Meneses, G.; Colon, O.; Yuan, K.; Stubbs, J.E.; Eng, P.J.; Wanhala, A.K.; Einkauf, J.D.; Boebinger, M.G.; Stack, A.G.; et al. Reaction Layer Formation on MgO in the Presence of Humidity. ACS Appl. Mater. Interfaces 2024, 16, 712–722. [Google Scholar] [CrossRef]
  37. Huang, L.; Yang, Z.; Wang, S. Influence of calcination temperature on the structure and hydration of MgO. Constr. Build. Mater. 2020, 262, 120776. [Google Scholar] [CrossRef]
  38. Puriwat, J.; Chaitree, W.; Suriye, K.; Dokjampa, S.; Praserthdam, P.; Panpranot, J. Elucidation of the basicity dependence of 1-butene isomerization on MgO/Mg(OH)2 catalysts. Catal. Commun. 2010, 12, 80–85. [Google Scholar] [CrossRef]
  39. Khamkongkaeo, A.; Mothaneeyachart, N.; Sriwattana, P.; Boonchuduang, T.; Phetrattanarangsi, T.; Thongchai, C.; Sakkomolsri, B.; Pimsawat, A.; Daengsakul, S.; Phumying, S.; et al. Ferromagnetism and diamagnetism behaviors of MgO synthesized via thermal decomposition method. J. Alloys Compd. 2017, 705, 668–674. [Google Scholar] [CrossRef]
  40. Liu, J.; Ruan, L.; Liao, J.; Pei, A.; Yang, K.; Zhu, L.; Chen, B.H. Magnesium hydroxide–supported ruthenium as an efficient and stable catalyst for glycerol-selective hydrogenolysis without addition of base and acid additives. New J. Chem. 2020, 44, 16054–16061. [Google Scholar] [CrossRef]
  41. Liu, J.; Li, C.; Niu, H.; Liang, C. Role of metal (Pt)–support (MgO) interactions in base-free glucose dehydrogenation. Catal. Sci. Technol. 2022, 12, 6849–6855. [Google Scholar] [CrossRef]
  42. Ye, J.; Chen, C.; Zheng, Y.; Zhou, D.; Liu, Y.; Chen, D.; Ni, L.; Xu, G.; Wang, F. Efficient conversion of cellulose to lactic acid over yttrium modified siliceous Beta zeolites. Appl. Catal. A Gen. 2021, 619, 118133. [Google Scholar] [CrossRef]
  43. Sandström, R.; Gracia-Espino, E.; Hu, G.; Shchukarev, A.; Ma, J.; Wågberg, T. Yttria stabilized and surface activated platinum (PtxYOy) nanoparticles through rapid microwave assisted synthesis for oxygen reduction reaction. Nano Energy 2018, 46, 141–149. [Google Scholar] [CrossRef]
  44. Wang, Q.; Shao, Y.; Zhang, T.; Liu, C.L.; Dong, W.S. One-pot conversion of xylose to 1,2-pentanediol catalyzed by an organic acid-assisted Pt/NC in aqueous phase. ChemSusChem 2024, 18, e202401109. [Google Scholar] [CrossRef] [PubMed]
  45. Li, X.; Pang, J.; Luo, W.; Zhao, Y.; Pan, X.; Zheng, M. Catalytic Conversion of Tetrahydrofurfuryl Alcohol over Stable Pt/MoS2 Catalysts. Catal. Lett. 2021, 151, 2734–2747. [Google Scholar] [CrossRef]
  46. Mansouri, A.; Semagina, N. Enhancement of palladium-catalyzed direct desulfurization by yttrium addition. Appl. Catal. A Gen. 2017, 543, 43–50. [Google Scholar] [CrossRef]
  47. Ilieva, L.; Venezia, A.; Petrova, P.; Pantaleo, G.; Liotta, L.; Zanella, R.; Kaszkur, Z.; Tabakova, T. Effect of Y Modified Ceria Support in Mono and Bimetallic Pd–Au Catalysts for Complete Benzene Oxidation. Catalysts 2018, 8, 283. [Google Scholar] [CrossRef]
  48. Shen, H.; Quan, Y.; Ding, X.; Ren, J. CuLaAl ternary catalysts for the selective hydrogenolysis of furfuryl alcohol to 1,2-pentanediol prepared by cation-anion double hydrolysis method: New insights into La promoter. Int. J. Hydrogen Energy 2025, 140, 392–404. [Google Scholar] [CrossRef]
  49. Aramendía, M. Synthesis and characterization of Pt/MgO catalysts and their use in n-hexane conversion. Colloids Surf. A Physicochem. Eng. Asp. 2003, 225, 137–143. [Google Scholar] [CrossRef]
  50. Mao, B.; Xie, Y.; Qiu, X.; He, F.; Zhao, K.; Xu, E.; Lu, J. Impact of various metal promoters on the propane dehydrogenation performance of cobalt-based catalysts. Mol. Catal. 2024, 567, 114457. [Google Scholar] [CrossRef]
  51. Shi, C.; Zhang, P. Effect of a second metal (Y, K, Ca, Mn or Cu) addition on the carbon dioxide reforming of methane over nanostructured palladium catalysts. Appl. Catal. B Environ. 2012, 115–116, 190–200. [Google Scholar] [CrossRef]
  52. Liu, B.; Wang, H.; Liu, H.; You, C. Aluminum-yttrium oxides supported iron for effective catalytic reforming of toluene: Promotion roles of yttrium. Fuel Process. Technol. 2022, 235, 107344. [Google Scholar] [CrossRef]
  53. Wang, Y.; Chen, M.; Zhang, K.; Wu, H.; Wang, J.; Cheng, Y.; Liu, Y.; Wei, Z. CoNi Alloy Nanoparticles Confined in N-Doped Porous Carbon as an Efficient and Versatile Catalyst for Reductive Amination of Levulinic Acid/Esters to N-Substituted Pyrrolidones. ACS Catal. 2023, 13, 12601–12616. [Google Scholar] [CrossRef]
  54. Liu, D.; Fu, J.; Wang, J.; Zhu, X.; Xu, J.; Zhao, Y.; Huang, J. Interfacial synergy within bimetallic oxide promotes selective hydrogenolysis of furfuryl alcohol to 1,5-pentanediol. Appl. Surf. Sci. 2024, 642, 158571. [Google Scholar] [CrossRef]
  55. Gholami, Z.; Tišler, Z.; Rubáš, V. Recent advances in Fischer-Tropsch synthesis using cobalt-based catalysts: A review on supports, promoters, and reactors. Catal. Rev. 2020, 63, 512–595. [Google Scholar] [CrossRef]
  56. Dai, Y.; Yu, F.; Li, Z.; An, Y.; Lin, T.; Yang, Y.; Zhong, L.; Wang, H.; Sun, Y. Effect of Sodium on the Structure-Performance Relationship of Co/SiO2 for Fischer-Tropsch Synthesis. Chin. J. Chem. 2017, 35, 918–926. [Google Scholar] [CrossRef]
  57. Li, H.; Nie, X.; Du, H.; Zhao, Y.; Mu, J.; Zhang, Z.C. Understanding the Role of Base Species on Reversed Cu Catalyst in Ring Opening of Furan Compounds to 1,2-Pentanediol. ChemSusChem 2024, 17, e202300880. [Google Scholar] [CrossRef] [PubMed]
  58. Tong, T.; Liu, X.; Guo, Y.; Norouzi Banis, M.; Hu, Y.; Wang, Y. The critical role of CeO2 crystal-plane in controlling Pt chemical states on the hydrogenolysis of furfuryl alcohol to 1,2-pentanediol. J. Catal. 2018, 365, 420–428. [Google Scholar] [CrossRef]
  59. Yamaguchi, A.; Murakami, Y.; Imura, T.; Wakita, K. Hydrogenolysis of Furfuryl Alcohol to 1,2-Pentanediol Over Supported Ruthenium Catalysts. ChemistryOpen 2021, 10, 731–736. [Google Scholar] [CrossRef]
  60. Götz, D.; Lucas, M.; Claus, P. Aqueous Phase Hydrogenolysis of Bio-Derivable Furfuryl Alcohol to Pentanediols Using Copper Catalysts. Catalysts 2017, 7, 50. [Google Scholar] [CrossRef]
  61. Götz, D.; Lucas, M.; Claus, P. C–O bond hydrogenolysis vs. C=C group hydrogenation of furfuryl alcohol: Towards sustainable synthesis of 1,2-pentanediol. React. Chem. Eng. 2016, 1, 161–164. [Google Scholar] [CrossRef]
  62. Lee, J.; Burt, S.P.; Carrero, C.A.; Alba-Rubio, A.C.; Ro, I.; O’Neill, B.J.; Kim, H.J.; Jackson, D.H.K.; Kuech, T.F.; Hermans, I.; et al. Stabilizing cobalt catalysts for aqueous-phase reactions by strong metal-support interaction. J. Catal. 2015, 330, 19–27. [Google Scholar] [CrossRef]
Scheme 1. Conversion of furfural/furfuryl alcohol into pentanediols.
Scheme 1. Conversion of furfural/furfuryl alcohol into pentanediols.
Catalysts 15 01005 sch001
Figure 1. Hydrogenolysis of furfuryl alcohol over PtY/MgO at varying (a) loading amount of Y, (b) reaction temperature, (c) effect of H2 pressure, and (d) reaction time. Conventional reaction conditions: furfuryl alcohol, 0.5 g; IPA, 20 mL; catalyst, 0.1 g; stirring speed, 800 rpm. Specified reaction conditions: (a) reaction temperature, 200 °C; H2 pressure, 3.0 MPa; reaction time, 4 h; (b) 5Pt0.5Y/MgO, 0.1 g, H2 pressure, 3.0 MPa; reaction time, 4 h; (c) 5Pt0.5Y/MgO, 0.1 g, reaction temperature, 200 °C; reaction time, 4 h; (d) 5Pt0.5Y/MgO, 0.1 g, reaction temperature, 200 °C; H2 pressure, 2.0 MPa.
Figure 1. Hydrogenolysis of furfuryl alcohol over PtY/MgO at varying (a) loading amount of Y, (b) reaction temperature, (c) effect of H2 pressure, and (d) reaction time. Conventional reaction conditions: furfuryl alcohol, 0.5 g; IPA, 20 mL; catalyst, 0.1 g; stirring speed, 800 rpm. Specified reaction conditions: (a) reaction temperature, 200 °C; H2 pressure, 3.0 MPa; reaction time, 4 h; (b) 5Pt0.5Y/MgO, 0.1 g, H2 pressure, 3.0 MPa; reaction time, 4 h; (c) 5Pt0.5Y/MgO, 0.1 g, reaction temperature, 200 °C; reaction time, 4 h; (d) 5Pt0.5Y/MgO, 0.1 g, reaction temperature, 200 °C; H2 pressure, 2.0 MPa.
Catalysts 15 01005 g001
Figure 2. Recyclability of the 5Pt0.5Y/MgO catalyst for the hydrogenolysis of furfuryl alcohol. Reaction conditions: furfuryl alcohol, 0.5 g; IPA, 20 mL; catalyst, 0.1 g; reaction temperature, 200 °C; H2 pressure, 2.0 MPa; reaction time, 4 h; stirring speed, 800 rpm.
Figure 2. Recyclability of the 5Pt0.5Y/MgO catalyst for the hydrogenolysis of furfuryl alcohol. Reaction conditions: furfuryl alcohol, 0.5 g; IPA, 20 mL; catalyst, 0.1 g; reaction temperature, 200 °C; H2 pressure, 2.0 MPa; reaction time, 4 h; stirring speed, 800 rpm.
Catalysts 15 01005 g002
Figure 3. (a) X-ray diffraction patterns of MgO, 5Pt/MgO, and 5Pt0.5Y/MgO; (b) X-ray diffraction patterns of used 5Pt/MgO and 5Pt0.5Y/MgO.
Figure 3. (a) X-ray diffraction patterns of MgO, 5Pt/MgO, and 5Pt0.5Y/MgO; (b) X-ray diffraction patterns of used 5Pt/MgO and 5Pt0.5Y/MgO.
Catalysts 15 01005 g003
Figure 4. TG-DTG profiles for the (a) fresh and (b) spent 5Pt0.5Y/MgO.
Figure 4. TG-DTG profiles for the (a) fresh and (b) spent 5Pt0.5Y/MgO.
Catalysts 15 01005 g004
Figure 5. TEM images of fresh 5Pt/MgO (a,d), 5Pt0.5Y/MgO (b,e), and used 5Pt0.5Y/MgO (c,f). (g) STEM with the elemental mapping images for 5Pt0.5Y/MgO.
Figure 5. TEM images of fresh 5Pt/MgO (a,d), 5Pt0.5Y/MgO (b,e), and used 5Pt0.5Y/MgO (c,f). (g) STEM with the elemental mapping images for 5Pt0.5Y/MgO.
Catalysts 15 01005 g005
Figure 6. XPS spectra of 5Pt/MgO and 5Pt0.5Y/MgO: (a) full spectra, (b) Pt 4f, and (c) Y 3d.
Figure 6. XPS spectra of 5Pt/MgO and 5Pt0.5Y/MgO: (a) full spectra, (b) Pt 4f, and (c) Y 3d.
Catalysts 15 01005 g006
Figure 7. CO2-TPD profiles of the MgO, 5Pt/MgO, and 5Pt0.5Y/MgO.
Figure 7. CO2-TPD profiles of the MgO, 5Pt/MgO, and 5Pt0.5Y/MgO.
Catalysts 15 01005 g007
Figure 8. H2-TPR results of Pt/MgO and 5Pt0.5Y/MgO.
Figure 8. H2-TPR results of Pt/MgO and 5Pt0.5Y/MgO.
Catalysts 15 01005 g008
Table 1. Hydrogenolysis of furfuryl alcohol over different catalysts.
Table 1. Hydrogenolysis of furfuryl alcohol over different catalysts.
EntryCatalystConv.
(%)
Yield (Selectivity) (%)
Catalysts 15 01005 i001Catalysts 15 01005 i002Catalysts 15 01005 i003Catalysts 15 01005 i004Catalysts 15 01005 i005Catalysts 15 01005 i006Catalysts 15 01005 i007
15Pt/Al2O31000.36.720.928.031.09.8 (9.8)2.3
25Pt/CeO268.80.40.31.71.244.715.6 (22.7)4.9
35Pt/MgO61.50.6---32.620.4 (33.2)7.9
43Pt/MgO42.50.5---20.016.9 (39.8)5.1
51Pt/MgO14.90.4---5.46.9 (46.3)2.2
60.5Pt/MgO1.5----0.40.8 (53.3)0.3
75Pt1Re/MgO1000.3---74.020.5 (20.5)5.2
85Pt1Fe/MgO49.50.827.21.41.03.47.8 (15.8)2.7
95Pt1Cu/MgO6.80.3---5.31.2 (17.9)-
105Pt1Ru/MgO1000.2-0.4-89.83.9 (3.9)5.7
115Pt1Ni/MgO1000.3-0.3-95.31.3 (1.3)2.8
125Pt1K/MgO17.80.5---11.04.7 (26.4)1.6
135Pt1La/MgO75.90.4---44.122.1 (29.4)8.8
145Pt1Y/MgO36.30.4---16.217.1 (47.2)2.6
Reaction conditions: furfuryl alcohol, 0.5 g; IPA, 20 mL; catalyst, 0.1 g; reaction temperature, 160 °C; H2 pressure, 3.0 MPa; reaction time, 4 h; stirring speed, 800 rpm.
Table 2. Main physiochemical properties of the MgO support and catalysts.
Table 2. Main physiochemical properties of the MgO support and catalysts.
SampleSBET [a]
[m2·g−1]
Vp [b]
[cm3·g−1]
Dp [c]
[nm]
Base Sites [d] [mmol·g−1]
50~300 °C300~700 °C500~800 °C
MgO29.260.30917.680.029-0.177
5Pt/MgO21.320.1553.790.0040.095-
5Pt0.5Y/MgO28.480.1523.820.0100.175-
Used 5Pt0.5Y/MgO26.240.1463.81---
[a] BET surface area. [b] BJH desorption cumulative volume of pores. [c] BJH desorption average pore width. [d] Amounts of surface basic site were determined by CO2-TPD.
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

Zhou, K.; Xu, J.; Guo, S.; Wu, H. Hydrogenolysis of Biomass-Based Furfuryl Alcohol into 1,2-Pentanediol over Magnesium Oxide-Supported Pt-Y Bimetallic Catalysts. Catalysts 2025, 15, 1005. https://doi.org/10.3390/catal15111005

AMA Style

Zhou K, Xu J, Guo S, Wu H. Hydrogenolysis of Biomass-Based Furfuryl Alcohol into 1,2-Pentanediol over Magnesium Oxide-Supported Pt-Y Bimetallic Catalysts. Catalysts. 2025; 15(11):1005. https://doi.org/10.3390/catal15111005

Chicago/Turabian Style

Zhou, Kuo, Jialin Xu, Shengrong Guo, and Hongjun Wu. 2025. "Hydrogenolysis of Biomass-Based Furfuryl Alcohol into 1,2-Pentanediol over Magnesium Oxide-Supported Pt-Y Bimetallic Catalysts" Catalysts 15, no. 11: 1005. https://doi.org/10.3390/catal15111005

APA Style

Zhou, K., Xu, J., Guo, S., & Wu, H. (2025). Hydrogenolysis of Biomass-Based Furfuryl Alcohol into 1,2-Pentanediol over Magnesium Oxide-Supported Pt-Y Bimetallic Catalysts. Catalysts, 15(11), 1005. https://doi.org/10.3390/catal15111005

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