Next Article in Journal
Highly Active and Water-Resistant Mn-Loaded MgAlOx Catalysts for NH3-SCR at Low Temperature
Previous Article in Journal
Theoretical Study on the Electronic Structure of Fe(0)–, Pd(0)–, and Pt(0)–Phosphine–Carbon Dioxide Complexes
Previous Article in Special Issue
Electrochemical Hydrogenation of Furfural Enhancing Furfuryl Alcohol Selectivity over Flower-like Zn-Based MBON-2 in Alkaline Medium
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Brief Review of Cu-Based Catalysts for the Selective Liquid-Phase Hydrogenation of Furfural to Furfuryl Alcohol

1
College of Chemistry, Zhengzhou University, Zhengzhou 450001, China
2
State Key Laboratory of Coal Conversion, Institute of Coal Chemistry, Chinese Academy of Sciences, Taiyuan 030001, China
3
State Key Laboratory of Coking Coal Resources Green Exploitation, Zhengzhou University, Zhengzhou 450001, China
*
Authors to whom correspondence should be addressed.
Chemistry 2025, 7(5), 153; https://doi.org/10.3390/chemistry7050153
Submission received: 4 August 2025 / Revised: 17 September 2025 / Accepted: 17 September 2025 / Published: 22 September 2025
(This article belongs to the Special Issue Catalytic Conversion of Biomass and Its Derivatives)

Abstract

With the rapid industrialization, excessive reliance on fossil fuels has resulted in energy crises and environmental pollution, driving the search for sustainable alternatives. Biomass-derived resources have emerged as promising candidates to replace fossil-based feedstocks. Among these, furfural (FF) serves as a key platform molecule that can be catalytically hydrogenated to various high-value chemicals, with furfuryl alcohol (FA) representing one of the most valuable products. Currently, Cr-based catalysts remain predominant for the selective hydrogenation of FF to FA. However, the severe environmental toxicity of Cr necessitates urgent development of alternative Cr-free catalytic systems. This study systematically reviews recent advances in FF hydrogenation to FA, providing an in-depth discussion of reaction mechanisms, including adsorption configurations, solvent effects, and side reactions, as well as a comprehensive analysis of structure–activity relationships, involving active metal, support, promoter, and preparation methods. Furthermore, we evaluate the application of the advanced characterization techniques for monitoring the reaction processes. Finally, we propose the future research directions: (1) designing efficient and stable non-noble metal catalysts and (2) elucidating reaction mechanisms via the combined in situ characterization and theoretical calculations. These efforts would facilitate the academic understanding and industrial implementation of the FF-to-FA conversion process.

Graphical Abstract

1. Introduction

Since the Industrial Revolution, the utilization of fossil fuels has significantly contributed to societal development and improvements in material living standards. However, the excessive exploitation has led to severe energy and resource supply crises, as well as environmental pollution, which have caused global attention and concern [1,2,3]. To address these challenges, researchers must identify an energy alternative with cost-effective, clean, readily available, and renewable advantages. Fortunately, biomass satisfies these criteria and holds potential as a substitute for fossil fuels [4,5,6]. It encompasses a diverse range of organic materials produced through photosynthesis, with numerous types naturally occurring in the environment. Nevertheless, only a limited number of biomass resources can simultaneously fulfill all the aforementioned requirements for energy substitution. In April 2004, the U.S. Department of Energy identified 12 representative biomass-derived compounds known as platform molecules [7]. Among them, furfural (FF) is regarded as one of the most significant biomass-based chemicals and is considered a highly promising platform molecule for the sustainable production of fuels and chemicals in the 21st century [8,9]. FF contains several functional groups, such as furan cyclic, aldehyde, and diene ethers, which render it highly active. Leveraging these chemical properties, a range of high-value-added chemicals can be synthesized using FF as a starting material [10,11,12,13,14,15,16,17], as illustrated in Scheme 1.
Although there are numerous routes and products for FF conversion, the selective hydrogenation to produce furfuryl alcohol (FA) warrants priority discussion. As one of the most significant derivatives of FF, FA accounts for approximately 65% of FF’s annual consumption. As early as 1934, Miller explored the potential transformation of FF to FA and measured the activation energy of the reaction. Subsequently, the catalytic conversion of FF to FA was successfully demonstrated in the literature in 1947. Subsequent studies employed a reduced CuCr catalyst, achieving a yield greater than 90% under high-temperature and high-pressure conditions. Furthermore, after the addition of CaO, the yield of FA approached 98% at lower pressure [18]. However, the high toxicity of chromium renders it unsuitable for use in industrial catalytic applications. In light of the growing demand for FA, there is an urgent need to develop environmentally benign alternatives [19,20].
In addition to the composition and structure, the design of the novel catalysts should also take into account the operational conditions of hydrogenation reactions. FF hydrogenation can generally be categorized into gas-phase and liquid-phase processes [21,22,23]. The gas-phase hydrogenation is widely preferred in industrial settings due to the more straightforward separation of products and catalysts, particularly in fixed-bed reactor configurations. Compared to the gas-phase hydrogenation, the liquid-phase hydrogenation offers higher reaction efficiency, improved mass transfer rates, as well as higher selectivity and product yield. As a result, the current research efforts are predominantly focused on the liquid-phase hydrogenation process [24,25,26].
Therefore, to bridge the existing gap between the development of environmentally friendly and highly efficient catalysts and their industrial catalytic liquid-phase hydrogenation application, it is necessary to review the catalytic reaction mechanisms of FF hydrogenation to FA, as well as the composition of the catalyst, and other related issues. This will provide theoretical guidance for the design and preparation of novel and highly efficient catalysts.

2. Reaction Mechanisms

Theoretical calculations and surface science studies have demonstrated that the carbonyl group of FF can be adsorbed on the surface of the active metal in either a perpendicular or parallel manner [27]. The adsorption mode of FF on the surface of metals plays a critical role in determining the pathway of the hydroreduction reaction, and this mode is highly dependent on the properties of the metal. Combining detailed experimental and theoretical studies, two different reaction mechanisms based on Cu-based catalysts [28] and other transition metals-based catalysts (e.g., Fe, Co, Ni, Pd, Pt) [29,30,31] have been proposed.

2.1. η1-(O)-aldehyde Configuration

As shown in Scheme 2 [27], for Cu-based catalysts, FF is adsorbed on the metal surface in a perpendicular configuration, bound in an η1-(O) configuration by lone electron pairs of carbonyl O atoms [28]. Due to the partial overlap between the antibonding orbital of the furan ring and the 3d orbital of the Cu atom, there is a strong repulsion between the furan ring and the Cu(111) crystal plane [28,32]. This repulsion prevents the stabilization of a parallel adsorption configuration of FF on the Cu surface. Consequently, the H atoms generated from H2 dissociation preferentially react with the adsorbed carbonyl group, thus limiting the hydroconversion of the furan ring and resulting in high selectivity for FA. A similar adsorption configuration is observed for FF on the Ag(110) crystal plane. The subsequent hydrogenation step follows two reaction mechanisms: (1) one H atom attacks the carbon atom of the carbonyl group to form an alcohol oxide intermediate, which is then attacked by a second H atom to produce FA; (2) the other H atom first attacks the carbonyl O atom to form a hydroxylalkyl intermediate, which subsequently combines with a second H atom to form FA. When the H atom initially attacks the carbonyl O, the electric charge in the C atom of the intermediate can be delocalized by the furan ring. This makes the intermediate more stable, reducing the activation energy of the reaction process, and more conducive to the formation of FA, so route (2) is more conducive to the occurrence of the reaction [28]. In addition, two configurations of the η1-(O) binding mode exist: vertical and tilted. Among these, the tilted configuration is more stable than the vertical one, favoring the interaction between the aldehyde group and adsorbed H atoms, thereby restricting the hydrogenation of the furan ring [33]. As shown in Scheme 3a. In recent years, studies have also indicated that for moderately or weakly binding metals such as Cu and Ag, stepped (211) surfaces may exhibit higher activity than terrace (111) surfaces [34]. It can be noticed that the adsorbed furfural on the surface of Cu sites exhibits a unique mode, which is different from that on other metals (Scheme 3b), which will be discussed in detail in Section 2.2.

2.2. η2-(C,O)-aldehyde Configuration

Due to the strong interaction between the transition metals other than Cu (e.g., Fe, Co, Ni, Pd, Pt) and the π bonds in FF molecules, there is no repulsion between the furan ring and the metal surface. As a result, FF can be adsorbed on the metal surface in a parallel manner [30,35]. On these metal surfaces, FF adopts an η2-(C,O)-aldehyde adsorption mode, where both the O and C atoms of the carbonyl group, as well as the C atom of the furan ring, are all co-bound to the metal surface [31], as shown in Scheme 4 [27]. The hydrogenation of FF to produce FA occurs in two steps: (1) carbonyl oxyhydrogenation to form hydroxyalkyl intermediates; (2) hydrogenation of the C linked to the hydroxyl group to form FA (Route 3). These two steps might also be carried out simultaneously, utilizing the active H species adsorbed on the surface [36]. At low temperatures, the second step of the hydrogenation reaction mechanism occurs preferentially. However, at elevated temperatures, the adsorption configuration of FF will change from η2-(C,O)-aldehyde adsorption to η1-(C)-acyl adsorption (Route 2). In this configuration, FF tends to undergo a decarbonylation reaction, leading to the formation of a byproduct, furan, resulting in a reduction in the selectivity of the product FA. In addition, when the adsorption active site is located on the oxygen atom (Route 1), the reaction follows a mechanism analogous to that of η1-(O)-furfural, leading to the formation of adsorbed FA intermediates. However, owing to the strong adsorption affinity of metals for oxygen, the FA intermediate undergoes subsequent hydrogenation and deoxygenation reactions, ultimately resulting in its conversion to 2-methylfuran [37]. Both the furan ring and the aldehyde group interact with the metal, enabling hydrogenation reactions. The furan ring can also completely hydrogenate to tetrahydrofuran, and even the ring-opening reaction will occur, further affecting the product distribution and selectivity of FF hydrogenation.
An accurate understanding of the reaction mechanisms contributes to the rational design of the novel catalysts. Although numerous studies have been devoted to exploring the active sites of FF hydrogenation catalysts, significant controversy persists at present.

2.3. Solvent Effects

The aforementioned adsorption model focuses on the direct adsorption of H2 on the catalyst surface. Although studies demonstrate that FF liquid-phase hydrogenation can proceed in solvent-free systems [38], the influence of solvents on catalytic performance remains significant. Wang et al. [39] revealed that solvents markedly affect FA selectivity. Specifically, FF is first hydrogenated to form FA; however, the incomplete or delayed desorption of FA may lead to its further hydrogenation to tetrahydrofurfuryl alcohol (THFA). When the alcoholic solvents are employed, the competitive adsorption between FA and the solvent molecules suppresses the overhydrogenation, thereby enhancing FA selectivity
In addition to the adsorption and activation of the substrate, the solvent also affects the activation of H2. Current studies have employed solvents such as isopropanol, methanol, and ethanol for FF hydrogenation to FA using Cu-based catalysts, as summarized in Table 1. Regarding the role of solvents, Bordoloi et al. [40] systematically investigated FF dissolution using a Cu/Zn/Al2O3 (CZAl) catalyst (Table 1). The study revealed that the high hydrogen bond donor (HBD) capacity of H2O enhances the reaction performance. Notably, H2 solubility in solvents did not play a decisive role; instead, hydrogen bond acceptance (HBA) and HBD capacities were identified as key factors. FF conversion exhibited a positive correlation with HBA capacity, generally following the solvent trend: polar protic > polar aprotic > non-polar. Consequently, 100% FF conversion was achieved with H2O as the solvent. While FA selectivity might be correlated with H2 solubility and HBD capacity, further research was required to elucidate these relationships.

2.4. Side Reactions

The hydrogenation of FF to FA is highly susceptible to excessive hydrogenation, leading to various side reactions. These include hydrodeoxygenation to 2-methylfuran (MF), ring-opening to form 1,5-pentanediol, and ring-hydrogenation to tetrahydrofurfuryl alcohol (THFA), as shown in Scheme 5 [62]. As illustrated, the primary by-products of FF hydrogenation are THFA and furan derivatives, while other by-products originate from the subsequent reactions of FA, THFA, or furans. To suppress the side reactions, it is critical to (1) selectively hydrogenate the carbonyl group to produce FA and (2) prevent further hydrogenation of FA. Compared to other transition metal catalysts, Cu-based catalysts exhibit superior selectivity by preferentially adopting the η1-(O)-aldehyde adsorption configuration, which avoids the interaction with the furan ring and effectively inhibits THFA formation. The formation of furan derivatives requires the decarboxylation of FF or FA, a process typically facilitated by precious metals (e.g., Pt, Pd) or Ni due to their strong CO adsorption capability [63]. Consequently, employing Cu-based catalysts and avoiding the incorporation of such active metal components can effectively suppress the generation of these by-products. However, the potential for further conversion of FA (e.g., to THFA or via ring-opening reactions) remains a challenge. In addition, there is a possibility of a humin reaction, which leads to catalyst inactivation [64].
Under the liquid-phase hydrogenation conditions, the formation of 2-methylfuran (2-MF) necessitates the hydrodeoxygenation of FA, a process requiring strong surface acidity [65]. Furthermore, the introduction of additional active metal components (e.g., Pt, Pd, or Ni) may promote this pathway, leading to increased 2-MF selectivity [66].
In conclusion, Cu-based catalysts effectively suppress the formation of major by-products during the liquid-phase hydrogenation of FF to FA. Furthermore, the prevention of FA overconversion can be achieved by selecting appropriate supports or additives that avoid introducing strong acidic sites, thereby minimizing undesired side reactions.

3. Catalyst

3.1. Active Metal

3.1.1. Noble Metal

In order to avoid the environmental damage caused by the use of Cr in the preparation and generation of Cr-containing wastes in the catalyst preparation and reaction, an increasing number of researchers are developing Cr-free catalysts with superior performance, including noble metals such as Pd, Pt, and Ru.
Noble metal Pd has very high activity, but its selectivity is usually low when directly used to catalyze the production of FA from FF. The selectivity of the hydrogenation product FA can be improved by reducing the Pd loading, reaction temperature, and H2 pressure [67]. Wang et al. [68] prepared Pd/TiH2 catalysts with different mass fractions, achieving a maximum conversion rate of 97.8% with 1 h of reaction time, which indicates that Pd has high activity in the liquid-phase hydrogenation of FF to prepare FA. However, its conversion is only 45.9%.
Noble metal Pt is usually used to catalyze α, hydrogenation of β-unsaturated aldehydes, obtaining a variety of chemicals and fuels with high energy density when applied to catalyze the FF hydrogenation reaction [69]. Papadogianakis et al. [70] employed the water-soluble platinum catalyst PtCl2/TPPTS (where TPPTS denotes tris(3-sulfophenyl)phosphine trisodium salt, a sulfonated analog of triphenylphosphine widely used in industrial aqueous-phase catalysis) for FA hydrogenation. Under mild and neutral conditions (130 °C, 4 MPa H2) in a green aqueous solvent, the system achieved a remarkable turnover frequency (TOF) exceeding 20,000 h−1 and a furfuryl alcohol selectivity of up to 99%. Yang et al. [71] synthesized Pt/CeO2-270 catalyst that achieved nearly 100% conversion within 1 h of reaction time, with 97.3% selectivity. These results demonstrate the high reactivity of Pt in this catalytic system. Although the selectivity is markedly superior to that of conventional precious metal catalysts, this enhancement is contingent on the synergistic role of Ce species.
Ru-based catalysts exhibit broad applicability, not only in the reduction of aromatic hydrocarbons [72], nitrile [73], CO methanation [74], α, β-unsaturated aldehydes [75] and hydrolysis of polyols to alkanes [76], but also in the reduction of FF under mild conditions. Yang et al. [77] prepared Al-MIL-53 and employed it as a support for Ru. The results indicate that the Ru/Al-MIL-53 catalyst with aromatic linker demonstrated superior performance, achieving a 100% yield of furfuryl alcohol at 20 °C and 0.5 MPa.
All these noble metal-based catalysts showed high activity and selectivity in the preparation of FA through liquid-phase hydrogenation of FF. Moreover, the use of noble metal-based catalysts effectively eliminates the need for Cr. However, the high cost of the noble metal catalysts does not make them suitable for large-scale industrial applications, and they are difficult to recycle. Therefore, the cheap transition metals have more potential for experimental research and industrial applications.

3.1.2. Fe, Co, and Ni

Catalysts with the active components of Fe, Co, and Ni have been increasingly developed in recent years for the hydrogenation of FF to produce FA [18,78,79,80]. Compared with noble metals, these transition metals can also exhibit high activity and stability in FF hydrogenation reactions but require more detailed design.
Fe is mainly used as an oxygenophilic metal promoter to enhance the hydrogen uptake activity of metals such as Pd, Cu, Ni, and Co [81]. Li et al. [82] synthesized a Ni3Fe1 bimetallic catalyst via a solvothermal method, which achieved over 98% furfural conversion and furfuryl alcohol selectivity under reaction conditions of 130 °C and 1 MPa H2 for 1 h.
For the supported Co catalysts, Audema et al. [83] reported a monometallic Co/SBA-15 catalyst in which Co nanoparticles were encapsulated within the mesoporous structure of SBA-15, and the optimal Co nanoparticle size suppressed the generation of side reactions.
The metal Ni is capable of catalyzing the conversion of FF to a variety of products, including FA, cyclopentanone, tetrahydroFA, furan, tetrahydrofuran, and many other hydrogenation products. Meng et al. [84] reported two heterogeneous Ni catalysts loaded on mixed metal oxides prepared by the structural topological transformation of NiAl hydrotalcite precursor interlayer containing nitrate or carbonate to tune the hydrogenation product species, denoted as Ni/MMO-CO3 and Ni/MMO-NO3, respectively.
Fe, Co, and Ni are widely distributed in nature, which are cheap and easy to obtain, and they have an excellent ability to activate H2 [85,86,87]. However, due to their distinct catalytic mechanisms compared to Cu-based systems (as discussed in Section 2.2), these metals might result in low selectivity toward FA [88]. Therefore, using these metals as the active sites still requires more careful design and precise control.

3.1.3. Cu

In recent years, studies have shown that Cu0 can facilitate the transfer of H species during hydrogenation reactions, such as CO2 and methanol hydrogenation, while Cu+ is generally recognized as the primary active site for hydrogenation processes [89,90]. This dual role has also been observed in FF hydrogenation to FA [91]. Therefore, further investigation into the mechanism and structure–activity relationships of Cu-based catalysts in FF hydrogenation toward FA is essential. Among all catalysts with non-noble metals as active components, Cu-based catalysts have garnered extensive attention because of their unique activity toward C=O bonds and exceptional selectivity for FAs [92,93,94].
Zhang et al. [41] prepared a series of Cu/MgO catalysts reduced at 300, 350, 400, 450, and 500 °C, respectively (Table 1). Characterization results revealed that a reduction temperature of 350 °C facilitated the optimal dispersion and reduction in Cu on the basic interface of MgO. With a higher active surface area, the weak adsorption of FA on the catalyst surface area avoids its excessive hydrogenation, while simultaneously exposing more active sites. The synergistic interaction between the basic interface of MgO and active Cu species contributed to the excellent performance of Cu/MgO-350.
Mohammed et al. [42] prepared monometallic Cu/Al2O3 catalysts with loadings of 1.0 wt.% and 5.0 wt.%, respectively, using copper nitrate, copper acetate, and copper sulfate as precursors, respectively (Table 1). At a 5.0 wt.% loading, Cu formed well-dispersed nanoparticles on the surface of Al2O3. Whereas at 1.0 wt%, Cu mainly existed in the form of single atoms and dimers. Characterization results indicated that the Cu/Al2O3 derived from the three precursors were sensitive to the structure of FF hydrogenation. At low Cu loadings, the decarbonylation reaction dominated, while at higher loadings, the hydrogenation reaction became the primary reaction. The sulfur-containing Cu/Al2O3 catalyst derived from copper sulfate precursor exhibited substantial side reactions, resulting in a FA selectivity of merely 0.8%. In contrast, the catalysts prepared with copper acetate as precursors provide the best performance under all conditions, demonstrating that metal loading and metal precursor selection are key to achieving optimal catalytic activity.
Li et al. [43] prepared a series of CuxMg3AlOy (x = 0.3, 0.5, 0.7, 0.9, and 1.0) composite metal oxide catalysts using CuMgAl-LDHs as precursors. These catalysts were directly employed for liquid-phase furfuryl hydrogenation without H2 pre-reduction. As shown in Figure 1a. Characterization results showed that Cu0.9Mg3AlOy has the highest catalytic activity, as shown in Table 1. Notably, the catalyst maintained its activity without significant degradation after five reuse cycles. Further analysis revealed that Cu2+ on the surface of Cu0.9Mg3AlOy was reduced in situ to Cu0 and Cu+ species during liquid-phase hydrogenation. The study of the structure–activity relationship demonstrated that Cu0 mainly facilitated activation and dissociation of H2, while the coexistence of Cu0 and Cu+ enabled co-adsorption of carbonyl groups, thereby enhancing the selectivity of FA (Figure 1b,c). Using an LDH precursor, Liu et al. [44] successfully anchored Cu species onto the Mg-Al LDH surface and subsequently prepared the CuMg3Al-R catalyst via high-temperature reduction (shown in Figure 2a). Characterization results revealed that the catalyst surface contained well-dispersed Cu0/Cu+ nanodots, which exhibited exceptional catalytic performance for FF hydrogenation to FA, achieving both conversion and selectivity exceeding 99%, as shown in Table 1. These findings demonstrate the following: (1) Cu species serve as highly active catalytic centers for this transformation, exhibiting remarkable intrinsic activity; (2) the catalyst’s structure and composition offer significant tunability through synthetic modifications. The study highlights the potential for optimizing catalytic performance through the rational design of Cu-based nanomaterials (Figure 1b,c).
In summary, Cu should maintain the highest possible dispersion, while having a high Cu+ content and an appropriate Cu0 content; its interaction mechanism requires further investigation. Cu is abundant in nature, inexpensive, easily available, highly selective, mechanistically well-defined, and can achieve high catalytic activity under mild reaction conditions, and has great potential for industrial application, and therefore is the best choice of the main active component for FF catalysts for furfuryl aldehyde selective direct hydrogenation to prepare FA.

3.2. Support

During the preparation and reduction of the catalysts, the active metal typically interacts with the support, which would further affect the activity and product selectivity of the catalyst [95,96]. Therefore, the selection of catalyst supports should be considered from multiple perspectives, such as specific surface area, pore structure, mechanical strength, and surface properties. In recent years, researchers have focused on optimizing the topography and surface characteristics of the support to enhance the performance of catalysts for the high-performance FF hydrogenation to produce FA. However, different supports exhibit unique characteristics; therefore, their selection and modification should be based on their structural and physicochemical properties in catalyst design.

3.2.1. Metal Oxide

In the design of catalysts for the hydrogenation of FF to FA, metal oxide supports play a crucial role and are extensively employed [97,98,99]. They not only support active metals but also actively participate in and optimize the entire catalytic process. Appropriate metal oxide supports can significantly enhance the activity, selectivity, and stability of the catalysts.
Liu et al. [100] using MgO-La2O3 mixed support to load Cu, which could promote the selective hydrogenation of FF to FA. The catalytic performance and characterization confirm that MgO effectively disperses Cu nanoparticles; meanwhile, La2O3 enhances the adsorption of the catalyst to FF. In addition, compared with single MgO, MgO-La2O3 support can effectively promote the reduction of Cu. Under optimal reaction conditions, the conversion rate of FF and the selectivity of FA exceeded 99.9%. The activity of the catalyst remains at about 97.6% after five cycles. This result demonstrates that suitable mixed metal oxide supports generate a synergistic effect, where the combined performance exceeds that of individual components.
Structural and morphological changes in metal oxides may lead to altered surface crystal exposure and increased oxygen vacancy formation, both of which are conducive to enhanced catalytic performance. Tan et al. [45] prepared three Cu-based catalysts supported on CeO2 supports with distinct morphologies—nanorods, nanocubes, and nanopolyhedra—for the catalytic hydrogenation of FF. The results revealed that compared to the other two catalysts, Cu/CeO2 catalysts with a nanorod (CeO2-R) morphology exhibited superior performance (Table 1). Characterization results indicated that the main exposed surface of the rod-shaped CeO2 support was the (111) crystal surface with more oxygen vacancies and higher Cu0 content of the reduced Cu/CeO2-R catalyst. The cooperative interaction between Cu0 and oxygen vacancies led to the excellent performance of Cu/CeO2-R.
The crystal phase of metal oxide supports also plays a critical role in determining catalytic performance. Liu et al. [46] employed Al2O3 with different crystalline phases to support Cu and catalyze FF hydrogenation for the production of FA. The results revealed that η-Al2O3 was more effective in dispersing Cu nanoparticles and promoting the highest surface concentrations of Cu+ and Cu0 species. This enhanced performance is attributed to the stronger FF adsorption and FA desorption capabilities of η-Al2O3 compared to other crystalline forms, which helps prevent hydrogenation of the furan ring and thereby achieves a high FF conversion and FA selectivity (Table 1).
In summary, metal oxides serve as promising catalyst supports and can be effectively utilized in the hydrogenation of FF to yield FA. However, metal oxide supports exhibit certain limitations, such as an undeveloped porous structure. Consequently, the choice of support material should consider appropriate modification strategies to mitigate these drawbacks.

3.2.2. SiO2

Compared to metal oxides, non-metallic SiO2 is also widely employed as a catalyst support [101,102,103]. It offers broader availability at lower cost, exhibits higher chemical inertness, and facilitates the formation of well-defined porous structures. Zhang et al. [104] modified granular SiO2 with ethanolamine to adjust the surface functional groups and hydrophilic properties while keeping the skeleton structure unchanged, followed by impregnation with Cu. Further studies have shown that this support can effectively improve the dispersion of Cu and form an appropriate Cu+/Cu0 ratio, thereby improving the activity and stability of the catalyst at the same time. However, the experimental results also show that this kind of SiO2 support cannot completely inhibit the sintering of carbon and reactive metals in the surface area of the catalyst.
Zhu et al. [105] selected nano-SiO2 balls as the support to prepare PtCuCo/SiO2 catalysts, and the reaction performance of FF hydrogenation to FA could reach 97.9% conversion and 95.4% selectivity. This is because the SiO2 sphere with a regular structure and well-developed pores can not only effectively reduce mass transfer resistance but also promote the formation of alloys while dispersing a variety of active components on the surface. This suggests that some of the shortcomings of the SiO2 supports can be compensated for by adjusting their structure.
Marco et al. [47] synthesized Cu/SiO2, SiO2@Cu, and nNPCu/SiO2 catalysts, and found that the NPCu@SiO2 and SiO2@Cu of the core–shell structure can promote different degrees of bonding between copper and silica, thereby adjusting the metal-support interaction between copper and silica. At the same time, the study fully demonstrates that catalysts with weak metal–support interactions do not have a significant effect on the distribution of products even in anhydrous reaction systems.
To sum up, the tunable properties of SiO2 render it a promising candidate as a support material for catalysts in the hydrogenation of FF to FA. Moreover, SiO2 supports featuring regular structures and well-developed porosity, offering distinct advantages and warranting further investigation into compositionally analogous materials such as zeolites.

3.2.3. Molecular Sieve

As mentioned above, SiO2 often suffers from underdeveloped porosity, a limitation that can be effectively addressed by molecular sieves [106]. Based on their pore structure, molecular sieves are generally classified into zeolitic and mesoporous types. Zeolite molecular sieves possess a well-defined topological framework with uniform micropores, whereas mesoporous molecular sieves feature ordered mesopores formed using templating methods [107,108].
Li et al. [48] synthesized PtCu@S-1 (S-1, silicalite-1) catalysts by the zeolite encapsulation method. As shown in Figure 3a, the rigid framework of S-1 effectively prevents the sintering of active metal particles, while its well-developed pore structure minimizes mass transfer limitations. PtCu@S-1 showed the remarkable activity in the hydrogenation of FF to FA (Table 1). The high performance of PtCu@S-1 is related to the high dispersion of active metals, while the developed pore structure and large specific surface area facilitate the dispersion of metals (Figure 3b).
Fu et al. [49] synthesized a Cu@MFI catalyst via a high-temperature rapid mixing hydrothermal method. Characterization results indicate that the MFI framework can coordinate with Cu to form a [Cu2(μ-O)x]2+ structure, enabling Cu to stably exist as single atoms within the 10-membered ring of the MFI zeolite. Compared to conventional impregnation methods, this Cu@MFI catalyst demonstrates excellent catalytic performance in aqueous-phase FF hydrogenation (Table 1). However, zeolites affect the FF hydrogenation reaction to produce by-products due to their acidic sites, and zeolites also have problems, such as a small pore size and low quantity. Therefore, special strategies are often required to mitigate the adverse effects of side reactions. It is worth noting that although the above results have fully demonstrated the feasibility of zeolites as catalyst supports in this reaction, some inherent problems remain. It is noteworthy that molecular sieves, characterized by well-developed pore structures and high specific surface areas, exhibit valence state distributions of Cu species that are highly sensitive to H2 reduction [109]. Given that Cu+ species are more favorable for hydrogenation reactions, excessive reduction temperatures should be avoided. Furthermore, the strong Brønsted acid sites inherent to molecular sieves may promote side reactions, such as the formation of cyclopentanone [110]. Consequently, further research and a more precise design are essential for optimizing Cu-loaded molecular sieve catalysts in the liquid-phase hydrogenation of FF to FA.
Compared with zeolites, mesoporous molecular sieves have weaker acidity and larger pore channels. Parikh et al. [50] investigated a Co-Cu/SBA-15 catalyst. Its high surface-to-volume ratio reduces the diffusion barrier of reactants and products. It was found that Co-Cu/SBA-15 had higher performance than the amorphous silica catalyst (Table 1). However, mesoporous molecular sieves may exhibit structural instability at elevated temperatures and typically possess fewer surface functional groups, which hinders the effective anchoring of active sites and may result in compromised catalyst stability.
In addition to SBA-15, another mesoporous molecular sieve material, MCM-41, has also been utilized for the liquid-phase hydrogenation of FF to FA. Fang et al. [51] employed Cu-loaded MCM-41 for FF hydrogenation, achieving nearly 100% FF conversion and FA selectivity at only 1 bar H2 pressure (Table 1). This exceptional performance is attributed to the unique spatial structure and high specific surface area of MCM-41, which facilitates FF diffusion and adsorption while enabling the stable and uniform dispersion of Cu0 and Cu+ sites on both internal and external surfaces.
In summary, molecular sieves offer advantages such as tunable structure and surface properties, as well as chemical inertness. However, zeolite molecular sieves suffer from strong surface acidity and limited pore size, whereas mesoporous molecular sieves often exhibit poor thermal stability and challenges in anchoring active sites. Therefore, effective modification of molecular sieve supports is essential to address these inherent limitations and enhance their performance as catalysts.

3.2.4. Carbon-Containing Material

In addition to the aforementioned supports, carbon-based materials offer advantages such as structural diversity and facile functionalization, making them widely employed in the synthesis of catalysts [111,112]. A carbon-coated Cu catalyst, CuOx@NC-150, was prepared by Qiu et al. using a thermal decomposition strategy of Cu-EDTA complexes [113]. Cu nanoparticles with complex valence states are prone to agglomerate during the hydrogenation reaction, which leads to catalyst deactivation. The NC formed after EDTA decomposition effectively coats Cu nanoparticles, preventing their sintering. Meanwhile, the good conductivity of NC helps maintain its catalytic reactivity.
The advantage of activated carbon lies in its large specific surface area and abundant surface functional groups, which facilitate the effective dispersion of Cu nanoparticles and enable efficient surface modification. Gong et al. [52] developed a sulfonic acid-based grafted activated carbon-supported Cu-based catalyst denoted as Cu/AC-SO3H, which had better catalytic performance than the unmodified Cu/AC (Table 1). That is, grafting -SO3H onto the surface of activated carbon enhances the dispersion of Cu nanoparticles, promotes a higher reduction degree of Cu, and strengthens the adsorption capacity of FF. These improvements collectively contributed to the enhanced hydrogenation performance of the catalyst.
At present, research on carbon-containing material supports remains relatively limited; however, it is undeniable that such materials hold significant potential for Cu-based catalyst support.

3.2.5. MOF

Metal–organic framework (MOF) is a porous crystalline material with a periodic network structure formed by the self-assembly of metal ions or metal clusters (as nodes) and organic ligands (as connecting rods) via ligand bonding. Taking MIL-100 and MIL-101 as examples, the composition and structure of the MOFs are illustrated in Scheme 6 [114]. MOF has an ultra-high specific surface area and porosity, which provides a large number of anchoring sites for activated metal nanoparticles (NPs) to promote a high degree of dispersion of the metal NPs and to prevent agglomeration efficiently. So, MOF can be used as a support. Moyo et al. [53] prepared a novel Cu-MOF, which showed a high conversion rate for FF and a high selectivity for FA, as listed in Table 1. However, the organic ligands of MOFs are prone to decomposition when exposed to high temperatures for a long time, which directly leads to the collapse of the MOF’s skeletal structure, making large-scale usage of MOFs difficult. This can also be used to synthesize novel coated catalysts, and the relevant description will be elaborated in the following text.
The advantages and disadvantages of the aforementioned supports are listed in Table 2. In summary, suitable supports can effectively enhance the dispersibility of active components, improving both the activity and stability of the catalysts. An ideal support should possess a well-developed mesoporous structure to facilitate reactant mass transfer, a large specific surface area to promote Cu dispersion, appropriate metal–support interactions to favor the Cu+ ratio, as well as good FF adsorption and FA desorption capabilities, along with the ability to promote H2 adsorption, activation, and hydrogen species overflow. However, most supports fail to satisfy all these requirements simultaneously, making it necessary to introduce promoters or modify the preparation method to compensate for their limitations.

3.3. Promoter

The incorporation of a metal promoter can modify the electronic structure and surface adsorption capacity of the active phases, thereby enhancing the performance of the catalyst. Promoters include metal, metal oxide, and other promoters. According to the literature, oxygen-friendly metals such as Fe [54], Zn [115], Sn [116], Ir [117], Mn [118], Co [119] and Mg [43] et al. can promote the adsorption of carbonyl groups on metal surfaces. In addition, Other types of promoters have also been found to similarly enhance the activity of the catalysts [120]. Some promoters have anti-carbon deposition properties. So even if the promoter does not have catalytic activity or has weak catalytic activity, its modification effect on the active component or support can effectively compensate for the defects of the catalyst, improve the activity or stability of the catalyst, and is of great research value.

3.3.1. Metal Promoter

For Cu-based catalysts, metal promoters can combine with Cu to form mixtures with phase interfaces or even directly form solid solutions. This effect will regulate the dispersion and electron state of Cu, thereby improving its adsorption and activation ability for FF, while preventing Cu sintering. Zhao et al. [55] prepared a series of layered bimetallic hydroxide-derived CunCo/MgAlOx (n = 1, 2, 3, 4) catalysts with varying Cu/Co ratios to investigate their catalytic reactivity in FF hydrogenation. The results demonstrated that the synergistic effect between Cu and Co facilitated the formation of Cu2O, which acted as a Lewis acid site, preferentially adsorbing C=O bonds and polarizing them to promote the formation of FA [50,121,122]. Among these catalysts, Cu3Co1/MgOx exhibited the strongest synergy and achieved the highest activity (Table 1). Additionally, Zhang et al. [56] explored the impact of Zn doping on the activity of Cu/SiO2 catalysts. They found that Zn not only promoted the reduction in CuO and improved the dispersion of metallic Cu but also effectively inhibited the decarbonylation reaction, enhancing the selectivity for FA. The Cu2Zn/SiO2 catalyst displayed the best catalytic performance (Table 1).
In summary, loading Cu onto different supports might lead to either overly strong or insufficient metal–support interactions, resulting in low Cu+ ratios or poor Cu dispersion. The use of various metal or metal oxide promoters can help modulate these interactions. Additionally, excessive acidity of the support may negatively affect the catalyst’s selectivity toward FA, which can be mitigated by incorporating basic promoters. Therefore, the rational selection of promoters based on the characteristics of the support is key to further enhancing catalytic performance.

3.3.2. Metal Oxide Promoter

Similarly to metal elements, metal oxide promoters are able to affect the activity of catalysts by adjusting the chemical environment of Cu, modulating some properties of the support. Zhu et al. [123] explored the effect of Al2O3 introduction during the synthesis of Cu/ZnO catalysts using the co-precipitation method. As shown in Figure 4a, XRD analysis revealed that these catalysts contained both ZnO and Cu phases, though with distinct crystallinity. H2-TPR further demonstrated significant differences in the reduction temperature of Cu species. These findings indicate that the Zn–Al interaction simultaneously modulates the metal-support interaction and Cu dispersion (as illustrated in Figure 4b,c), ultimately changing the catalytic performance.
Li et al. [54] prepared Fe3O4/Cu@C by N2 atmosphere roasting using Fe(NO3)3-loaded Cu-MOF as precursor. These Fe3O4 particles embedded in the middle of the C support can serve to reduce the electron cloud density at the neighboring Cu active sites and promote hydrogen overflow, thus improving the intrinsic activity and selectivity of the catalyst (Table 1). In contrast, the Fe/Cu@C catalysts generated after H2 roasting showed lower activity because Fe(0) is more likely to promote deep hydrogenation and carbon buildup.
In summary, metal oxide promoters can simultaneously modulate the active components and supports, and there are many related studies. Therefore, they hold significant potential for further research and large-scale application.

3.3.3. Other Promoter

In addition to metal elements and metal oxides, there are some basic substances that can also serve as promoters, primarily for regulating the surface properties of supports. Researchers explored the use of inorganic salts such as K2CO3 as a promoter to improve the catalytic activity. Zhang et al. [124] reported an effective strategy involving K2CO3 assisted FF hydrogenation using a CuO#TiO2 catalyst, in which the FA production rate of CuO#TiO2 catalyst was as high as 24.2 molFOL/(molCu·h) at 100 °C and 1.4 MPa H2 pressure, which was 4 times that of CuO#TiO2 catalyst without K2CO3 accelerator. The results showed that K2CO3 assistance has outstanding dual functions: (1) promoting the dissolution of gaseous H2 molecules in solvents; (2) promoting the formation of Cu active sites on the surface.
Wu et al. [57] employed a Na ion-exchange process to modulate the zeolite microenvironment, obtaining a Na–Cu@TS-1 catalyst with enhanced activity and selectivity for the selective hydrogenation of FF to FA. The introduced Na species modulate the zeolite’s acid-base properties to suppress side reactions and promote an electron-rich state in the Cu species, thereby enhancing the FF hydrogenation. Catalyst deactivation primarily results from Na leaching into the liquid phase, because activity is nearly fully restored through a Na re-addition process.
In summary, loading Cu onto different supports may lead to either overly strong or insufficient metal–support interactions, resulting in low Cu+ ratios or poor Cu dispersion. The use of various metal or metal oxide promoters can help modulate these interactions. Additionally, excessive acidity of the support may negatively affect the catalyst’s selectivity toward FA, an issue that can be mitigated by incorporating basic promoters. Therefore, the rational selection of promoters based on the characteristics of the support is key to further enhancing catalytic performance.

3.4. Preparation Method

3.4.1. Supported Cu Catalysts

One of the most commonly used preparation methods for catalysts involves loading active components onto a support. This method has a wide range of applicability and is easy to operate. However, this method can easily cause changes in catalyst performance through changes in supports. Liu et al. [58] compared Cu/SiO2 catalysts prepared via ammonia evaporation (AE) and impregnation methods. The results showed that the AE-synthesized sample formed a layered copper silicate species, which facilitated the generation of a higher concentration of Cu+ ions. Consequently, under reaction conditions of 90 °C and 1 MPa H2, the AE catalyst achieved a turnover frequency (TOF) of 36.0 h−1 for the hydrogenation of FF to FA.
The morphology, dispersion, and chemical state of the active components, as well as the interaction between the active metal and the support, will differ for supported catalysts obtained through different preparation methods, which will further affect the performance of the catalysts. Furthermore, this preparation method offers advantages such as tunable catalyst composition and straightforward fabrication, making it highly suitable for large-scale catalyst production at present. However, since the active metal sites are directly exposed on the support surface, challenges including metal sintering, surface carbon deposition, and leaching of soluble components might arise. Notably, in liquid-phase hydrogenation, the solvent’s influence on the catalyst cannot be overlooked. Thus, future efforts should not only focus on improving catalyst stability but may also require strategies for efficient catalyst regeneration.

3.4.2. Coated Cu Catalysts

Using support to encapsulate the active components will effectively avoid sintering of Cu, thereby improving the stability of the catalyst [125,126]. The core–shell structure catalyst is a special nanocomposite material consisting of an active component as the core and an outer shell as a support. In addition to regulating the electronic and geometrical structure of the catalyst, this structure can also effectively avoid the sintering of the active components and improve the life of the catalyst. Tu, R et al. [59] synthesized the novel ultra-dilute high-entropy alloys (HEAs) with multiple metal sites and core–shell structure for boosting the selectivity of FF hydrogenation to FA, as shown in Figure 5a. The results showed that the FF was converted to FA with a conversion of 87.32% and selectivity of 100% at 90 °C over ultra-dilute HEA (NiCoCuZnFe/C-800) (Table 1). This is because different metal sites play different roles in hydrogenation reactions, and the core–shell structure effectively stabilizes the structure of ultra-dilute HEAs (Figure 5b,c). The core–shell structure enhanced the stability of the catalysts during the hydrogenation reaction, and the porous carbon shell improved the transfer of FF and electrons. However, the core–shell structure is complicated to prepare, the process is demanding and costly, there are interface defects, and the shell may rupture during prolonged use, among other disadvantages.
Although core–shell catalysts can effectively protect and confine copper species, thereby mitigating active metal sintering and enhancing suitability for large-scale applications, their performance is inherently dependent on the stability of the shell structure. Furthermore, the shell must exhibit well-developed porosity to facilitate efficient mass transfer of reactants and products. Consequently, the design and synthesis of such catalysts entail greater complexity compared to traditional supported catalysts. Additionally, in liquid-phase hydrogenation processes, fluid-induced shear forces may compromise the structural integrity of the shell, necessitating further investigation into coating methodologies to ensure mechanical robustness under reaction conditions.

3.4.3. Integrated Cu and Support

In addition to the two methods of loading active metals onto supports and coating supports onto active metals, researchers have developed catalyst synthesis strategies in recent years to simultaneously prepare active metal particles and supports. MOF is a porous crystalline material formed by the self-assembly of metal ions or metal clusters with organic ligands through coordination bonds. This precursor can be thermally decomposed to form C-coated Cu. Yang et al. [60] reported a CuO#TiO2 catalyst prepared by using Cu3(BTC)2 as the backbone, first loading TiO2 support inversely into the interior of the MOF’s backbone, and then decomposing Cu3(BTC)2 by roasting. The particle size of the catalyst was smaller than that of the conventional loaded catalyst, and the catalyst had a superior activity (Table 1), and an FF conversion rate of 20.8 molFF/molCu·h. While MOF as a precursor can effectively enhance the dispersion of the active components, the organic ligand is completely decomposed during the roasting process, thus requiring the addition of TiO2 as an extra support. In fact, organic ligands have the potential to be converted into polymer supports. Qiao et al. [61] prepared a MOF (ZJU-199) with a ligand having an acrylate-like structure, which was subsequently pyrolyzed at a controlled roasting temperature to prepare the catalyst. As shown in Figure 6a. Under appropriate calcination temperature, the acrylate substituent on the organic ligand undergoes decarboxylation to form a C=C bond, which subsequently undergoes polymerization, leading to the formation of a cross-linked polystyrene network by the ligand; Cu2+ undergoes sintering and partial reduction during this process, ultimately forming a Cu/Cu2O heterojunction (Figure 6b). The crosslinked polystyrene effectively wrapped and isolated the heterojunction, preventing it from further sintering, thus ensuring the activity and stability of the catalyst at the same time. The catalyst was prepared with a single MOF precursor without the need for additional supports.
The preparation method of integrated Cu and support may have potential advantages in terms of activity and stability. However, considering that the synthesis of MOFs may require some expensive organic ligands, it may not be suitable for large-scale production and industrial applications. Nevertheless, this method can be utilized for the directed synthesis of Cu-based catalysts with different valence state distributions and particle sizes, which is beneficial for further understanding the mechanisms.
In summary, for specific active components and supports, the catalyst structure and metal-support interaction will fully affect properties such as metal particle size and metal valence distribution, which are highly correlated with the performance of the catalyst. An ideal support should exhibit high (FF adsorption capacity and strong FA desorption capability, facilitate H2 adsorption and activation, maximize Cu dispersion and the Cu+/Cu ratio, and maintain structural stability under reaction conditions. Therefore, in the design of Cu-based catalysts for the preparation of FA by FF hydrogenation, it is necessary to fully consider the synthesis mode of the catalyst based on the selected support.

4. Characterization Method

The development of efficient catalysts for the hydrogenation of FF to FA is critical, and the characterization methods are fundamental to understanding and optimizing catalyst performance. Traditional characterization includes XRD [40,127], N2-physisorption [128,129], SEM [130,131] and TEM [53,132] only discloses = readily accessible structure and morphology information (Figure 7). The spectrum techniques, such as FT-IR [61,133], Raman [134,135], XPS [136,137], EPR [134], and the various temperature programmed (TP) techniques [131,138,139] could also provide further insights into the intrinsic surface chemical properties of the catalyst (Figure 8). For example, functional groups, elemental valence, surface acidity and alkalinity, and redox properties. It is particularly important to note that XPS measurements require careful consideration of testing conditions. Since electronic states under operational conditions (e.g., in H2 atmosphere) differ significantly from those of air-exposed samples, XPS results may not fully represent the catalyst’s working state. Furthermore, XPS alone often cannot reliably distinguish between Cu0 and Cu+ species due to their similar binding energies. Therefore, Auger electron spectroscopy is frequently employed in conjunction with XPS to achieve a more accurate identification of copper valence states. Extensive literature has demonstrated that these analytical methods can effectively elucidate the structural and compositional characteristics of catalysts and establish the superficial structure–activity relationships. Furthermore, the development and application of more advanced characterization techniques may facilitate the acquisition of certain critical data more efficiently.
Defining the metal dispersion is crucial for unraveling the intrinsic activity and kinetic experiments. Conventionally, Cu dispersion is determined via N2O pulse chemisorption [43], while Ni dispersion is generally calculated via H2-TPD or chemisorption data [140,141]. CO pulse adsorption is a useful technique employed to determine metal dispersion and quantify the active sites on catalyst surfaces [142]. CO molecules exhibit strong chemisorption on the surfaces of certain transition metals (e.g., Ni, Fe), but this interaction is generally weak on Cu surfaces. However, at lower temperatures (e.g., 193 K), CO can undergo reversible chemisorption on Cu as well. In fact, even Cu+ species can effectively coordinate with CO to form Cu+(CO)2 complexes [143]. Based on this principle, the total dispersion of an alloy can be calculated by combining data from CO pulse chemisorption with elemental analysis. Although this method cannot separately determine the dispersion of individual metals, it may still hold practical value, considering that not all exposed metal atoms are capable of chemisorption or catalytic activity. Liu et al. [144] synthesized a series of SiO2-supported Cu-Ni bimetallic catalysts and systematically investigated the influence of the Cu/Ni ratio on FF hydrogenation to FA. Obviously, the larger the specific surface area of active metal particles and the higher the metal dispersion, the more favorable it is for the catalytic reaction to proceed. The metal dispersion and surface area of Cu3Ni1/SiO2 were the largest, at 55.7% and 20.8 m2/g, respectively. Under the optimum conditions (333 K, 2.0 MPa, 3 h), Cu3Ni1/SiO2 gives 99.9 % yield to FOL at 333 K in isopropanol. Compared to Ni/SiO2, it has a yield of about 60% higher. CO pulse adsorption enables simultaneous measurement of both metals’ dispersion in Cu–Ni alloys, offering a convenient approach for establishing catalyst structure–activity relationships. Notably, the Cu3Ni/SiO2 catalyst exhibited the highest catalytic activity, correlating with its superior metal dispersion.
As a novel spectroscopic technique that provides deeper insight into the intrinsic properties of materials, X-ray absorption spectroscopy (XAS) consists of X-ray absorption near-edge structure (XANES) and extended X-ray absorption fine structure (EXAFS). Specifically, XANES can be used to study valence and electronic structure of the metal elements in the catalysts, while EXAFS can provide information about the coordination environments of the metal atoms and the particle size information [145,146].
Dalai et al. [147] prepared a series of Cu–Zn–Cr–Zr catalysts. Owing to the complex composition of metallic elements in the catalyst, identifying the presence of active Cu species on the catalyst surface after fresh and reduced states poses significant analytical challenges. However, as demonstrated in Figure 9a, XANES analysis clearly reveals the presence of CuO in the fresh catalyst and Cu0 in the reduced catalyst, indicating that catalyst deactivation is caused by Cu oxidation under reaction conditions. These findings corroborate the experimental results obtained through the conventional characterization techniques (Figure 9b).
It is well known that at equivalent metal loading, higher metal dispersion corresponds to a greater number of catalytic active sites. Single-atom catalysts (SACs) not only achieve the highest possible dispersion, but the isolated metal sites can also engage in strong metal–support interactions (SMSI) with the support, effectively modulating the electronic states of the metal atoms and enhancing their intrinsic activity [148]. SACs can effectively enhance the utilization of active metal components. However, due to the extremely small size of SACs, the conventional characterization methods struggle to directly observe the presence of metal single atoms. Mesa et al. [149] prepared a PdCu single-atom alloy supported on Al2O3 catalyst. They used the EXAFS technique to demonstrate that palladium atoms successfully replaced copper atoms on the nanoparticles while maintaining atomic dispersion, which confirms the formation of single-atom Pd, as shown in Figure 10a,b. TOF analysis clearly demonstrated that the optimal Pd1Cu216 SAA catalyst exhibited a copper site activity of 7.21 h−1, representing an 85% increase compared to the pure Cu100 reference. The palladium site activity also increased significantly to 813 ± 81 h−1, surpassing that of the monometallic Pd100. This enhancement is attributed to the mutual reactivity promoted by the interaction between atomically dispersed Pd sites and the Cu surface. In fact, STEM and other advanced electron microscope techniques can offer a more direct visualization of single-atom sites; however, due to the high magnification involved, they cannot conclusively demonstrate that all active sites are uniformly dispersed as single atoms. The above results show that the rational use of XAS technique associated with electron microscope can help us to fully understand the structure of catalysts at the atomic level and thus associate the conformational relationships (Figure 10c,d).
The aforementioned characterization techniques can effectively provide static structural and compositional information about catalysts, but they are less capable of revealing dynamic changes that occur during actual preparation or reaction processes—information critical for tracing reaction pathways and investigating deactivation mechanisms. In situ experiments using conventional characterization methods can effectively address this limitation. For example, in situ XRD allows direct observation of phase transitions under specific conditions. Liu et al. [150] employed in situ XRD to monitor the crystalline evolution of Cu/Al2O3 during programmed-temperature calcination in different atmospheres. The results showed that calcination in N2 effectively suppressed sintering of Cu species compared to that in air, leading to a more facile reduction in Cu active components to lower oxidation states in subsequent hydrogenation steps (Figure 11). The Cu/Al2O3-N2-R with the smallest Cu particles exhibited excellent performance, achieving a FA yield of up to 99.9% after 2 h at 120 °C and 1 MPa of H2 pressure. To investigate the valence state changes in Cu species, in situ XPS is a suitable technique. Zhang et al. [151] used in situ XPS to study the redox behavior of Cu in Cu/SiO2 catalysts prepared using different loading methods during H2 reduction. Their results indicated that the Cu loading method influenced the formation of lamellar copper phyllosilicate phases after calcination (Figure 12a), which in turn affected the valence distribution of Cu during the subsequent reduction process (Figure 12b). The Cu/SiO2-EA and Cu/SiO2-HDP catalysts exhibited significantly enhanced Cu dispersion after reduction, attributed to the formation of a lamellar copper phyllosilicate structure during synthesis. This structural feature resulted in the superior activity and stability compared to other catalysts.
While TP techniques are valuable for catalyst characterization, their application is limited for FF and FA due to the high boiling points of these compounds. In such cases, thermogravimetric analysis (TG-DTG) offers an effective alternative for qualitative and kinetic studies by monitoring mass changes and amplifying differential details. Furthermore, thermal gravimetric analysis of catalysts after FF adsorption can provide insights into adsorption behavior. For example, Liu et al. [46] synthesized a series of Cu-supported Al2O3 catalysts with different crystal phases and employed TG analysis to investigate FA adsorption/desorption properties. The results demonstrated that the Al2O3 crystal phase significantly influences FA adsorption strength. Notably, FA desorbed more readily from Cu/η-Al2O3 compared to other catalysts, which may account for its superior catalytic activity (FAL conversion: 99.7 %, FA selectivity: 94.0 %), highest TOF value (29.3 h−1), and lowest apparent activation energy (66.3 kJ/mol).
Liu et al. [152] prepared a series of MgAlOx-supported Cu or CuCo catalysts, and characterized them using in situ DRIFTS. Unsupported Cu2O was employed as the reference (Figure 13a,b). In situ DRIFTS results show that on the Cu3Co1/MgAlOx catalyst, the intensity of the characteristic spectral band at 1713 cm−1 decreases rapidly, accompanied by an increase in the intensity of the characteristic spectral band at 1089 cm−1 attributed to the C-O bond of alcohols, indicating that C=O is rapidly reduced. Based on the free energy of the FF hydrogenation process to form FA on different crystal faces (Figure 13c), the Cu3Co1 (111) crystal face is favorable for H* attacking C=O to form the -CH2O+ intermediate, and also favorable for the subsequent hydrogenation steps of this intermediate. Another possible pathway, where H* attacks O to form C+, has relatively high activation energy (Figure 13d), resulting in a lower probability of this reaction pathway occurring. However, the activation energy for FF hydrogenation on the Cu3Co1 (111) crystal face still has a relative advantage.
In summary, the novel characterization method facilitates researchers to determine reaction paths more quickly and accurately, and analyze the interactions between different components more precisely and in detail. Novel in situ characterization methods enable real-time tracking of reaction paths and guide efficient interface design. The future trend is to focus on high temporal and spatial resolution in situ platforms (e.g., XAS-DRIFTS-MS [152]) and machine learning-assisted data parsing to promote FF hydrogenation catalysts towards the industrialization goal of “precise design, efficient conversion, and long-term stability”.

5. Conclusions and Perspectives

5.1. Summary and Significance

Cr-based catalysts have been widely employed in the industrial hydrogenation of FF to produce FA. However, concerns over the toxicity of Cr have driven efforts to develop non-toxic alternatives. Although extensive research has highlighted the potential of Cu-based catalysts as substitutes, their activity and stability remain inferior to those of Cr-based systems. While current studies have provided insights into the adsorption and activation mechanisms of FF on Cu-based surfaces, a detailed understanding of the overall reaction mechanism remains lacking. Furfural vapor-phase hydrogenation has advantages in terms of reaction efficiency and continuous production, but the gas-phase reaction is usually carried out at higher temperatures, which requires strict thermal stability and mechanical strength of the catalyst. The liquid-phase reaction conditions are relatively mild, which is conducive to the development and use of more diverse catalyst systems. In addition, since the liquid-phase method is easier to perform at the laboratory level, the reaction mechanism and the structure–activity relationship of the catalyst can be efficiently studied with the help of the liquid-phase method. Although liquid-phase direct hydrogenation of FF lacks the inherent capacity for continuous production characteristic of gas-phase processes, it retains significant potential for industrial application. This potential can be realized if catalysts achieve higher turnover TOF and superior stability under milder reaction conditions, thereby demonstrating a decisive performance advantage over gas-phase hydrogenation systems.

5.2. Composition and Preparation Method for Better Catalyst Design

Compared to noble metals or active transition metals such as Fe, Co, and Ni, Cu offers both low cost and high selectivity, making Cu-based catalysts promising alternatives to Cr-based systems. Current research suggests that maximizing the dispersion of Cu is essential for enhancing catalytic performance, with single-atom-dispersed Cu being particularly beneficial in this regard. Cu+ is known as a highly active site, while Cu0 may favor the adsorption of FF, implying that an optimal Cu+/Cu0 ratio could significantly improve catalytic activity. A support with a well-developed mesoporous structure and a large specific surface area facilitates better dispersion of Cu species. Additionally, supports that provide adsorption sites for FF or H2 can further enhance catalytic performance. However, strong acidity on the support surface may promote undesirable side reactions. This drawback can be mitigated by incorporating promoters, which help regulate the oxidation state of Cu, improve its dispersion, and promote reduction through modulation of the metal–support interaction. Moreover, for the same support, variations in pretreatment methods and Cu loading techniques can influence the metal–support interaction, potentially suppressing or even preventing Cu sintering. Therefore, when designing Cu-based catalysts, it is crucial to carefully select suitable promoters and synthesis strategies that are tailored to the nature of the support.

5.3. Using Novel Characterization Technique

Conventional characterization methods are essential for understanding the structure–activity relationships of catalysts; however, advanced techniques are still required to probe their fine structural features and dynamic behavior under real reaction conditions. CO pulse adsorption is an effective tool for evaluating the dispersion of multi-metal active components. When combined with XAS and advanced electron microscopy, it enables a more intuitive analysis of Cu distribution, particularly in SACs. In situ characterization techniques offer significant advantages in monitoring structural and compositional changes during catalyst preparation, reduction, and reaction processes, as well as the dynamic evolution of reactants on the catalyst surface. TG can serve as an alternative to traditional TP experiments, providing more accurate insights into surface adsorption and desorption behaviors. Looking ahead, the development and application of novel characterization methods will further advance the understanding of structure–activity relationships and support the rational design of more efficient Cu-based catalysts.

5.4. Perspectives

In summary, well-designed Cu-based catalysts enable high activity and stability for the liquid-phase hydrogenation of FF to FA under relatively mild conditions, demonstrating potential for industrial implementation. However, unlike gas-phase processes, liquid-phase systems lack inherent advantages for continuous operation. Consequently, future research should prioritize achieving milder reaction conditions, higher TOF, and enhanced catalyst reusability. This direction implies that conventionally supported catalysts with simple preparation methods might be less competitive, while encapsulated catalyst architectures could offer greater promise for advancing FF hydrogenation research and applications.

Author Contributions

Conceptualization, M.Z.; methodology, M.Z. and C.L.; resources, Z.L.; writing—original draft preparation, T.L., Y.G. and C.L.; writing—review and editing, M.Z.; supervision, Z.L.; project administration, M.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Nijnens, J.; Behrens, P.; Kraan, O.; Sprecher, B.; Kleijn, R. Energy transition will require substantially less mining than the current fossil system. Joule 2023, 7, 2408–2413. [Google Scholar] [CrossRef]
  2. Niu, Y.; Yan, Y.; Dong, J.; Yue, K.; Duan, X.; Hu, D.; Li, J.; Peng, L. Evidence for sustainably reducing secondary pollutants in a typical industrial city in China: Co-benefit from controlling sources with high reduction potential beyond industrial process. J. Hazard. Mater. 2024, 478, 135556. [Google Scholar] [CrossRef]
  3. Ali, H.; Ali, I.; Baz, K. Do industrialization and nonrenewable energy affect environmental quality? Evidence from top fossil fuel–consuming countries. Environ. Sci. Pollut. Res. 2023, 30, 109800–109809. [Google Scholar] [CrossRef]
  4. Vogt, E.T.C.; Weckhuysen, B.M. The refinery of the future. Nature 2024, 629, 295–306. [Google Scholar] [CrossRef]
  5. Phang, F.J.F.; Chew, J.J.; Chakraborty, S.; Sunarso, J. Synthesis of hydrochars via wet torrefaction of biomass for sustainable energy production: A life cycle assessment study. ACS Sustain. Chem. Eng. 2025, 13, 2884–2892. [Google Scholar] [CrossRef]
  6. Ge, H.; Kuwahara, Y.; Okada, M.; Yamashita, H. Mild hydrodeoxygenation of aromatic ketones by Pd/HxWO3–y with plasmonic features assisted by visible-NIR light irradiation. ACS Sustain. Chem. Eng. 2024, 12, 2162–2171. [Google Scholar] [CrossRef]
  7. Gunukula, S.; Pendse, H.P.; DeSisto, W.J.; Wheeler, M.C. Heuristics to guide the development of sustainable, biomass-derived, platform chemical derivatives. ACS Sustain. Chem. Eng. 2018, 6, 5533–5539. [Google Scholar] [CrossRef]
  8. Liao, Y.; Koelewijn, S.F.; Van den Bossche, G.; Van Aelst, J.; Van den Bosch, S.; Renders, T.; Navare, K.; Nicolaï, T.; Van Aelst, K.; Maesen, M.; et al. A sustainable wood biorefinery for low–carbon footprint chemicals production. Science 2020, 367, 1385–1390. [Google Scholar] [CrossRef]
  9. 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]
  10. 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]
  11. Li, Y.; Zhang, J.; Zhang, W.; Zhou, C.; Huang, J.; Su, Z.; Qiao, Z.; Qin, X.; Xiong, P.; Xiao, F.-S. Carbothermal shock synthesis of a PtCoCe ternary oxide catalyst for selective hydrogenolysis of furfural to 1,5-Pentanediol. ACS Catal. 2025, 15, 7731–7740. [Google Scholar] [CrossRef]
  12. Li, H.; Luo, M.; Zhu, T.; Wang, C.; Liao, Y.; Wang, H.; Ma, L. Valorization of furfural to cyclopentanone/cyclopentanol over platinum cluster catalysts: The ensemble effect. Appl. Catal. B Environ. Energy 2025, 366, 125052. [Google Scholar] [CrossRef]
  13. Di, J.; Li, Q.; Ma, C.; He, Y.-C. An efficient and sustainable furfurylamine production from biomass-derived furfural by a robust mutant ω-transaminase biocatalyst. Bioresour. Technol. 2023, 369, 128425. [Google Scholar] [CrossRef]
  14. Zhang, L.; Meng, G.; Zhang, W.; Li, X.; Zhang, Z.; Yang, M.; Wu, Y.; Wang, D.; Li, Y. Oriented conversion of a LA/HMF mixture to GVL and FDCA in a biphasic solvent over a Ru single-atom/nanoparticle dual-site catalyst. ACS Catal. 2023, 13, 2268–2276. [Google Scholar] [CrossRef]
  15. Ramos, N.C.; Manyé Ibáñez, M.; Mittal, R.; Janik, M.J.; Holewinski, A. Combining renewable electricity and renewable carbon: Understanding reaction mechanisms of biomass-derived furanic compounds for design of catalytic nanomaterials. Acc. Chem. Res. 2023, 56, 2631–2641. [Google Scholar] [CrossRef]
  16. Ren, Z.; Yu, L.; Song, X.; Wang, S.; Li, H.; Wang, L.; Zhang, X.; Li, X.; Yang, Y.; Wang, Z.; et al. Atomically dispersed Sn catalysts toward selective oxidation of furfural to biodegradable polymer monomers. ACS Catal. 2025, 15, 3799–3809. [Google Scholar] [CrossRef]
  17. Liu, B.; Chen, X.; Xu, Y.; Qiao, C.; Lu, Z.; Tian, Y. A combo Zr–zeolite and Zr(OH)4 mixture composition for one–pot production of γ–valerolactone from furfural. Renew. Energy 2024, 229, 120751. [Google Scholar] [CrossRef]
  18. 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]
  19. Li, M.-Y.; Shi, Y.-C.; Xu, W.-X.; Zhao, L.; Zhang, A.-Z. Exploring Cr(VI)-induced blood-brain barrier injury and neurotoxicity in zebrafish and snakehead fish, and inhibiting toxic effects of astaxanthin. Environ. Pollut. 2024, 355, 124280. [Google Scholar] [CrossRef]
  20. Ma, L.; Liu, T.; Li, J.; Yang, Q. Interaction characteristics and mechanism of Cr(VI)/Cr(III) with microplastics: Influence factor experiment and DFT calculation. J. Hazard. Mater. 2024, 476, 134957. [Google Scholar] [CrossRef] [PubMed]
  21. Jorqueira, D.S.S.; de Lima, L.F.; Moya, S.F.; Vilcocq, L.; Richard, D.; Fraga, M.A.; Suppino, R.S. Critical review of furfural and furfuryl alcohol production: Past, present, and future on heterogeneous catalysis. Appl. Catal. A Gen. 2023, 665, 119360. [Google Scholar] [CrossRef]
  22. Racha, A.; Samanta, C.; Sreekantan, S.; Marimuthu, B. Review on catalytic hydrogenation of biomass-derived furfural to furfuryl alcohol: Recent advances and future trends. Energy Fuels 2023, 37, 11475–11496. [Google Scholar] [CrossRef]
  23. Shi, D.; Zhang, A.; Bao, W.; Zhang, L. Recent advances in catalytic hydrogenation of furfural over non-noble metal catalysts. Ind. Eng. Chem. Res. 2025, 64, 13966–13983. [Google Scholar] [CrossRef]
  24. Wang, J.; Quan, Y.; Ren, J. Effect of Cu loading on the structure and catalytic properties of Cu/SBA-15 for the hydrogenation of furfural to furfuryl alcohol. Sep. Purif. Technol. 2025, 362, 131568. [Google Scholar] [CrossRef]
  25. Mensah, J.; Jampaiah, D.; Ahmed, M.H.M.; Konarova, M.; Durndell, L.J.; Bhargava, S.K.; Lee, A.F.; Wilson, K. Liquid-phase furfural hydrogenation over Ni/Alumina catalysts. ACS Sustain. Chem. Eng. 2025, 13, 2893–2905. [Google Scholar] [CrossRef]
  26. Kuchkina, N.; Sorokina, S.; Lisichkin, D.; Bykov, A.; Bukalov, S.; Matveeva, V.; Shifrina, Z. Simultaneous synthesis of a microporous polymer framework with embedded Nickel nanoparticles for enhanced selective hydrogenation of furfural to furfuryl alcohol. ACS Appl. Polym. Mater. 2025, 7, 2017–2028. [Google Scholar] [CrossRef]
  27. 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]
  28. Sitthisa, S.; Sooknoi, T.; Ma, Y.; Balbuena, P.B.; Resasco, D.E. Kinetics and mechanism of hydrogenation of furfural on Cu/SiO2 catalysts. J. Catal. 2011, 277, 1–13. [Google Scholar] [CrossRef]
  29. Sitthisa, S.; An, W.; Resasco, D.E. Selective conversion of furfural to methylfuran over silica-supported NiFe bimetallic catalysts. J. Catal. 2011, 284, 90–101. [Google Scholar] [CrossRef]
  30. Sitthisa, S.; Pham, T.; Prasomsri, T.; Sooknoi, T.; Mallinson, R.G.; Resasco, D.E. Conversion of furfural and 2-methylpentanal on Pd/SiO2 and Pd–Cu/SiO2 catalysts. J. Catal. 2011, 280, 17–27. [Google Scholar] [CrossRef]
  31. Medlin, J.W. Understanding and controlling reactivity of unsaturated oxygenates and polyols on metal catalysts. ACS Catal. 2011, 1, 1284–1297. [Google Scholar] [CrossRef]
  32. Shi, Y.; Zhu, Y.; Yang, Y.; Li, Y.-W.; Jiao, H. Exploring furfural catalytic conversion on Cu(111) from computation. ACS Catal. 2015, 5, 4020–4032. [Google Scholar] [CrossRef]
  33. 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]
  34. Liu, S.; Govindarajan, N.; Chan, K. Understanding activity trends in furfural hydrogenation on transition metal surfaces. ACS Catal. 2022, 12, 12902–12910. [Google Scholar] [CrossRef]
  35. Vorotnikov, V.; Mpourmpakis, G.; Vlachos, D.G. DFT study of furfural conversion to furan, furfuryl alcohol, and 2-Methylfuran on Pd(111). ACS Catal. 2012, 2, 2496–2504. [Google Scholar] [CrossRef]
  36. Nakagawa, Y.; Nakazawa, H.; Watanabe, H.; Tomishige, K. Total Hydrogenation of Furfural over a Silica-Supported Nickel Catalyst Prepared by the Reduction of a Nickel Nitrate Precursor. ChemCatChem 2012, 4, 1791–1797. [Google Scholar] [CrossRef]
  37. Saha, B.; Racha, A.; Chaudhary, P.K.; Singh, B.K.; Samanta, C.; Newalkar, B.L. Enhanced production and techno-economic analysis of sustainable biofuel production via continuous hydrogenation of furfural using the Cu–ZnO–Al2O3 catalyst. ACS Sustain. Chem. Eng. 2025, 13, 3183–3199. [Google Scholar] [CrossRef]
  38. 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]
  39. Jia, P.; Lan, X.; Li, X.; Wang, T. Highly active and selective NiFe/SiO2 bimetallic catalyst with optimized solvent effect for the liquid-phase hydrogenation of furfural to furfuryl alcohol. ACS Sustain. Chem. Eng. 2018, 6, 13287–13295. [Google Scholar] [CrossRef]
  40. Singh, G.; Singh, L.; Gahtori, J.; Gupta, R.K.; Samanta, C.; Bal, R.; Bordoloi, A. Catalytic hydrogenation of furfural to furfuryl alcohol over chromium-free catalyst: Enhanced selectivity in the presence of solvent. Mol. Catal. 2021, 500, 111339. [Google Scholar] [CrossRef]
  41. 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, 143755. [Google Scholar] [CrossRef]
  42. 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 Environ. 2020, 273, 119062. [Google Scholar] [CrossRef]
  43. 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]
  44. Yi, W.-J.; Yang, J.; Gao, Y.; Zhou, X.; Zhang, W.; Zhang, M.; Liu, Z. Well-dispersed Cu0/Cu+ nanodots confined in hydrotalcite-derived oxide for highly efficient and selective furfural hydrogenation to furfuryl alcohol. Chem. Eng. J. 2025, 515, 163742. [Google Scholar] [CrossRef]
  45. 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]
  46. Gao, Y.; Hu, J.; Yi, W.-J.; Yang, J.; Jiang, K.; Yang, T.; Li, Z.; Zhang, M.; Liu, Z.; Wu, B. Unveiling the effects of support crystal phase on Cu/Al2O3 catalysts for furfural selective hydrogenation to furfuryl alcohol. Chem. Eng. J. 2024, 498, 155620. [Google Scholar] [CrossRef]
  47. Soares, W.L.S.; Feitosa, L.F.; Moreira, C.R.; Bertella, F.; Lopes, C.W.; de Farias, A.M.D.; Fraga, M.A. Tailoring Cu–SiO2 interaction through nanocatalyst architecture to assemble surface sites for furfural aqueous-phase hydrogenation to cycloketones. ACS Appl. Mater. Interfaces 2025, 17, 13146–13161. [Google Scholar] [CrossRef] [PubMed]
  48. Luo, W.; Song, S.; Zhao, W.; Ding, T.; Tian, Y.; Li, X. Cis-trans adsorption configuration influenced high-efficient hydrogenation of furfural over zeolite-encapsulated PtCu bimetallic catalyst. Chem. Eng. J. 2024, 489, 151471. [Google Scholar] [CrossRef]
  49. Liang, W.; Xu, G.; Fu, Y. Accurate restricted transition-state shape selective hydrogenation of furfural over zeolite confined Cu catalyst. Chin. J. Catal. 2025, 74, 71–81. [Google Scholar] [CrossRef]
  50. Srivastava, S.; Mohanty, P.; Parikh, J.K.; Dalai, A.K.; Amritphale, S.S.; Khare, A.K. Cr-free Co–Cu/SBA-15 catalysts for hydrogenation of biomass-derived α-, β-unsaturated aldehyde to alcohol. Chin. J. Catal. 2015, 36, 933–942. [Google Scholar] [CrossRef]
  51. Wang, Y.; Wang, Y.; Tang, Q.; Jing, F.; Cao, Q.; Fang, W. Efficient activation of H2 on copper species immobilized by MCM-41 for selective hydrogenation of furfural at ambient pressure. Mol. Catal. 2021, 515, 111921. [Google Scholar] [CrossRef]
  52. 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]
  53. Moyo, P.S.; Mehlana, G.; Matsinha, L.C.; Makhubela, B.C.E. Copper-based metal–organic framework: Synthesis, characterization and evaluation for the hydrogenation of furfural to furfuryl alcohol. J. Inorg. Organomet. Polym. Mater. 2025, 35, 2257–2273. [Google Scholar] [CrossRef]
  54. Li, P.; Ma, F.; Fu, M.; Lu, S.; Xia, X.; Li, C.; Gao, Y.; Li, F. Hydrogenation of furfural to furfuryl alcohol over MOF-derived Fe/Cu@C and Fe3O4/Cu@C catalysts. React. Chem. Eng. 2022, 7, 994–1004. [Google Scholar] [CrossRef]
  55. Zhao, H.; Liao, X.Q.; Cui, H.S.; Zhu, M.C.; Hao, F.; Xiong, W.; Luo, H.; Lv, Y.; Liu, P.L. Efficient Cu-Co bimetallic catalysts for the selective hydrogenation of furfural to furfuryl alcohol. Fuel 2023, 351, 128887. [Google Scholar] [CrossRef]
  56. 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]
  57. 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]
  58. Wang, J.; Yang, J.; Bing, Z.; Gao, Y.; Yang, T.; Liu, Q.; Zhang, M.; Liu, Z. Refining surface copper species on Cu/SiO2 catalysts to boost furfural hydrogenation to furfuryl alcohol. Molecules 2025, 30, 225. [Google Scholar] [CrossRef] [PubMed]
  59. Tu, R.; Liang, K.; Sun, Y.; Wu, Y.; Lv, W.; Jia, C.Q.; Jiang, E.; Wu, Y.; Fan, X.; Zhang, B.; et al. Ultra-Dilute high-entropy alloy catalyst with core-shell structure for high-active hydrogenation of furfural to furfuryl alcohol at mild temperature. Chem. Eng. J. 2023, 452, 139526. [Google Scholar] [CrossRef]
  60. Yang, X.; Zhang, Z.; Liu, W.; Liang, T.; Dang, D.; Tian, X. In Situ hybridizing Cu3(BTC)2 and Titania to attain a high-performance Copper catalyst: Dual-functional role of metal-support interaction on the activity and selectivity. ChemCatChem 2021, 13, 3846–3856. [Google Scholar] [CrossRef]
  61. Chen, K.; Ling, J.-L.; Wu, C.-D. In Situ generation and stabilization of accessible Cu/Cu2O heterojunctions inside organic frameworks for highly efficient catalysis. Angew. Chem. Int. Ed. 2020, 59, 1925–1931. [Google Scholar] [CrossRef] [PubMed]
  62. Yan, K.; Ling, J.; Wu, 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]
  63. Zheng, H.-Y.; Zhu, Y.-L.; Teng, B.-T.; Bai, Z.-Q.; Zhang, C.-H.; Xiang, H.-W.; Li, Y.-W. Towards understanding the reaction pathway in vapour phase hydrogenation of furfural to 2-methylfuran. J. Mol. Catal. A Chem. 2006, 246, 18–23. [Google Scholar] [CrossRef]
  64. Sajid, M.; Rizwan Dilshad, M.; Saif Ur Rehman, M.; Liu, D.; Zhao, X. Catalytic conversion of xylose to furfural by p-Toluenesulfonic acid (pTSA) and chlorides: Process optimization and kinetic modeling. Molecules 2021, 26, 2208. [Google Scholar] [CrossRef]
  65. Yang, X.; Xiang, X.; Chen, H.; Zheng, H.; Li, Y.-W.; Zhu, Y. Efficient Synthesis of Furfuryl Alcohol and 2-Methylfuran from Furfural over Mineral-Derived Cu/ZnO Catalysts. ChemCatChem 2017, 9, 3023–3030. [Google Scholar] [CrossRef]
  66. Zhang, Z.; Pei, Z.; Chen, H.; Chen, K.; Hou, Z.; Lu, X.; Ouyang, P.; Fu, J. Catalytic in-situ hydrogenation of furfural over bimetallic Cu–Ni alloy catalysts in isopropanol. Ind. Eng. Chem. Res. 2018, 57, 4225–4230. [Google Scholar] [CrossRef]
  67. Nehra, P.; Betsy, K.; Vinod, C. Selective Hydrogenation of Furfural to Furfuryl Alcohol over Pd Supported on Ternary Oxide in Aqueous Medium Under Mild Conditions. ChemCatChem 2025, 17, e02092. [Google Scholar] [CrossRef]
  68. 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]
  69. Kijeński, J.; Winiarek, P. Selective hydrogenation of α,β-unsaturated aldehydes over Pt catalysts deposited on monolayer supports. Appl. Catal. A Gen. 2000, 193, L1–L4. [Google Scholar] [CrossRef]
  70. Seretis, A.; Mertika, I.; Gabrielatou, E.; Patatsi, E.; Thanou, I.; Diamantopoulou, P.; Tzevelekidis, P.; Fakas, C.; Lilas, P.; Kanellopoulos, P.G.; et al. Novel hydrogenation reaction of renewable furfural into furfuryl alcohol using highly efficient and selective water-soluble platinum catalysts modified with phosphines and nitrogen-containing ligands in green aqueous media. Catal. Today 2025, 444, 115019. [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 Environ. 2023, 328, 122458. [Google Scholar] [CrossRef]
  72. Cui, X.; Surkus, A.E.; Junge, K.; Topf, C.; Radnik, J.; Kreyenschulte, C.; Beller, M. Highly selective hydrogenation of arenes using nanostructured ruthenium catalysts modified with a carbon-nitrogen matrix. Nat. Commun. 2016, 7, 11326. [Google Scholar] [CrossRef] [PubMed]
  73. Huang, Y.; Adeeva, V.; Sachtler, W.M.H. Stability of supported transition metal catalysts in the hydrogenation of nitriles. Appl. Catal. A Gen. 2000, 196, 73–85. [Google Scholar] [CrossRef]
  74. Abdel-Mageed, A.M.; Widmann, D.; Olesen, S.E.; Chorkendorff, I.; Biskupek, J.; Behm, R.J. Selective CO methanation on Ru/TiO2 catalysts: Role and influence of metal–support interactions. ACS Catal. 2015, 5, 6753–6763. [Google Scholar] [CrossRef]
  75. Hájek, J.; Kumar, N.; Mäki-Arvela, P.; Salmi, T.; Murzin, D.Y. Selective hydrogenation of cinnamaldehyde over Ru/Y zeolite. J. Mol. Catal. A Chem. 2004, 217, 145–154. [Google Scholar] [CrossRef]
  76. Miyazawa, T.; Koso, S.; Kunimori, K.; Tomishige, K. Development of a Ru/C catalyst for glycerol hydrogenolysis in combination with an ion-exchange resin. Appl. Catal. A Gen. 2007, 318, 244–251. [Google Scholar] [CrossRef]
  77. 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]
  78. Merlo, A.B.; Bengoa, J.F.; Bianco, M.V.; Marchetti, S.G.; Vetere, V. Characterization of NiFe alloy phases in nanoparticles used as selective catalysts for the production of furfuryl alcohol from furfural hydrogenation. Mol. Catal. 2024, 569, 114614. [Google Scholar] [CrossRef]
  79. Lara, K.; Arias, D.; Sepulveda, C.; Buljan, A.; Lizana, I.; Delgado, E.J.; Pecchi, G.; Herrera, C. Furfural hydrogenation over reduced pure (LaBO3) and substituted (LaB0.5B’0.5O3) (B, B’: Fe, Co, Ni) perovskites. ChemCatChem 2024, 16, e202301507. [Google Scholar] [CrossRef]
  80. Ren, W.; Tian, J.; Wang, Z.; Zhang, M. Which is better for converting furfural into furfuryl alcohol: The strong metal-support interaction effect or the formation of alloy? Appl. Catal. A Gen. 2024, 673, 119609. [Google Scholar] [CrossRef]
  81. Manikandan, M.; Venugopal, A.K.; Nagpure, A.S.; Chilukuri, S.; Raja, T. Promotional effect of Fe on the performance of supported Cu catalyst for ambient pressure hydrogenation of furfural. RSC Adv. 2016, 6, 3888–3898. [Google Scholar] [CrossRef]
  82. Li, Y.; Zhu, L.; Zhao, J.; Qiu, M.; Liu, J.; He, J.; Wang, L.; Lei, J.; Tian, W.; Rong, L. Facile synthesis of a high-efficiency NiFe bimetallic catalyst without pre-reduction for the selective hydrogenation reaction of furfural. Catal. Sci. Technol. 2023, 13, 457–467. [Google Scholar] [CrossRef]
  83. Audemar, M.; Ciotonea, C.; De Oliveira Vigier, K.; Royer, S.; Ungureanu, A.; Dragoi, B.; Dumitriu, E.; Jérôme, F. Selective hydrogenation of furfural to furfuryl alcohol in the presence of a recyclable Cobalt/SBA-15 catalyst. ChemSusChem 2015, 8, 1885–1891. [Google Scholar] [CrossRef] [PubMed]
  84. 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]
  85. Lin, W.; Chen, Y.; Zhang, Y.; Zhang, Y.; Wang, J.; Wang, L.; Xu, C.C.; Nie, R. Surface synergetic effects of Ni–ReOx for promoting the mild hydrogenation of furfural to tetrahydrofurfuryl alcohol. ACS Catal. 2023, 13, 11256–11267. [Google Scholar] [CrossRef]
  86. Chen, J.; Jia, W.; Yu, X.; Sun, Y.; Zhang, J.; Yang, S.; Tang, X.; Peng, L.; Zeng, X.; Lin, L. Insight into the Co2+/Co3+ sites for the selective reduction of furfural to furfuryl alcohol. Fuel 2023, 332, 126137. [Google Scholar] [CrossRef]
  87. Shao, Y.; Wu, J.; Zheng, Z.; Fan, M.; Sun, K.; Bkangmo Kontchouo, F.M.; Zhang, L.; Zhang, S.; Hu, G.; Hu, X. Alloying cobalt in Co–Fe–Al catalyst for achieving the selective conversion of furfural to cyclopentanone. Renew. Energy 2022, 195, 957–971. [Google Scholar] [CrossRef]
  88. Ratthiwal, J.; Lazaro, N.; Pineda, A.; Esposito, R.; Alothman, Z.A.; Reubroycharoen, P.; Luque, R. Furfural conversion over calcined Ti and Fe metal-organic frameworks under continuous flow conditions. Catal. Commun. 2023, 177, 106649. [Google Scholar] [CrossRef]
  89. Gong, J.; Yue, H.; Zhao, Y.; Zhao, S.; Zhao, L.; Lv, J.; Wang, S.; Ma, X. Synthesis of ethanol via syngas on Cu/SiO2 catalysts with balanced Cu0–Cu+ sites. J. Am. Chem. Soc. 2012, 134, 13922–13925. [Google Scholar] [CrossRef]
  90. Paris, C.; Karelovic, A.; Manrique, R.; Le Bras, S.; Devred, F.; Vykoukal, V.; Styskalik, A.; Eloy, P.; Debecker, D.P. CO2 hydrogenation to methanol with Ga- and Zn-doped mesoporous Cu/SiO2 catalysts prepared by the aerosol-assisted Sol-Gel process. ChemSusChem 2020, 13, 6409–6417. [Google Scholar] [CrossRef]
  91. 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 Environ. 2021, 282, 119576. [Google Scholar] [CrossRef]
  92. 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]
  93. Rao, R.; Dandekar, A.; Baker, R.T.K.; Vannice, M.A. Properties of copper chromite catalysts in hydrogenation reactions. J. Catal. 1997, 171, 406–419. [Google Scholar] [CrossRef]
  94. Shoosri, T.; Chotiwilaiwan, P.; Rattanapornchaiwat, T.; Teerawatananond, T.; Miyake, T.; Panpranot, J.; Weerachawanasak, P. Bimetallic copper-and nickel-rich Cu–Ni phyllosilicate catalysts for the liquid phase selective hydrogenation of furfural to furfuryl alcohol. RSC Adv. 2024, 14, 38232–38244. [Google Scholar] [CrossRef] [PubMed]
  95. Hu, S.; Li, W.-X. Sabatier principle of metal-support interaction for design of ultrastable metal nanocatalysts. Science 2021, 374, 1360–1365. [Google Scholar] [CrossRef]
  96. Wang, T.; Hu, J.; Ouyang, R.; Wang, Y.; Huang, Y.; Hu, S.; Li, W.-X. Nature of metal-support interaction for metal catalysts on oxide supports. Science 2024, 386, 915–920. [Google Scholar] [CrossRef]
  97. Ghashghaee, M.; Sadjadi, S.; Shirvani, S.; Farzaneh, V. A novel consecutive approach for the preparation of Cu–MgO catalysts with high activity for hydrogenation of furfural to furfuryl alcohol. Catal. Lett. 2017, 147, 318–327. [Google Scholar] [CrossRef]
  98. Li, F.; Cao, B.; Ma, R.; Liang, J.; Song, H.; Song, H. Performance of Cu/TiO2-SiO2 catalysts in hydrogenation of furfural to furfuryl alcohol. Can. J. Chem. Eng. 2016, 94, 1368–1374. [Google Scholar] [CrossRef]
  99. Jackson, M.A.; White, M.G.; Haasch, R.T.; Peterson, S.C.; Blackburn, J.A. Hydrogenation of furfural at the dynamic Cu surface of CuOCeO2/Al2O3 in a vapor phase packed bed reactor. Mol. Catal. 2018, 445, 124–132. [Google Scholar] [CrossRef]
  100. Zhang, J.; Yang, X.; Jia, Z.; Han, Q.; Yu, S.; Liu, S.; Li, L.; Wu, Q.; Yu, H.; Liu, Y.; et al. Bi-functional active sites Cu/MgO-La2O3 catalysts the selective hydrogenation of furfural to furfuryl alcohol. Chem. Eng. J. 2024, 500, 157400. [Google Scholar] [CrossRef]
  101. Chen, B.; Hu, J.; Yu, Z.; Zhu, L.; Wu, C.; Yu, H. Improving the activity and selectivity of Ir/SiO2 for the selective hydrogenation of unsaturated carbonyl compounds by SnO2 promotion. Ind. Eng. Chem. Res. 2025, 64, 1181–1188. [Google Scholar] [CrossRef]
  102. Li, C.; Liu, Q.; Huang, D.; Wang, J.; Xi, Y.; Huang, Z.; Li, F. Catalytic tandem CO2 hydrogenation and hydroformylation for high-yield synthesis of C2+ alcohols. ACS Catal. 2025, 15, 11440–11451. [Google Scholar] [CrossRef]
  103. Coeck, R.; Claes, N.; Cuypers, T.; Bals, S.; De Vos, D.E. The sustainable and catalytic synthesis of N,N-alkylated fatty amines from fatty acids and esters. Green Chem. 2025, 27, 1410–1422. [Google Scholar] [CrossRef]
  104. Du, H.; Ma, X.; Jiang, M.; Yan, P.; Conrad Zhang, Z. Highly efficient Cu/SiO2 catalyst derived from ethanolamine modification for furfural hydrogenation. Appl. Catal. A Gen. 2020, 598, 117598. [Google Scholar] [CrossRef]
  105. Ruan, L.; Zhu, L.; Zhang, X.; Guo, G.; Shang, C.; 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]
  106. Kerstens, D.; Smeyers, B.; Van Waeyenberg, J.; Zhang, Q.; Yu, J.; Sels, B.F. State of the art and perspectives of hierarchical zeolites: Practical overview of synthesis methods and use in catalysis. Adv. Mater. 2020, 32, 2004690. [Google Scholar] [CrossRef]
  107. Li, Y.; Yu, J. New stories of zeolite structures: Their descriptions, determinations, predictions, and evaluations. Chem. Rev. 2014, 114, 7268–7316. [Google Scholar] [CrossRef] [PubMed]
  108. Li, W.; Yue, Q.; Deng, Y.; Zhao, D. Ordered mesoporous materials based on interfacial assembly and engineering. Adv. Mater. 2013, 25, 5129–5152. [Google Scholar] [CrossRef] [PubMed]
  109. Santa Cruz-Navarro, D.; Torres-Rodríguez, M.; Gutiérrez-Arzaluz, M.; Mugica-Álvarez, V.; Pergher, S.B. Comparative study of Cu/ZSM-5 catalysts synthesized by two Ion-exchange methods. Crystals 2022, 12, 545. [Google Scholar] [CrossRef]
  110. Yu, Z.; Tian, H.; Sun, K.; Shao, Y.; Zhang, L.; Zhang, S.; Duan, P.; Liu, Q.; Niu, S.; Dong, D.; et al. Impacts of externally added Brønsted and Lewis acid on conversion of furfural to cyclopentanone over Ni/SiC catalyst. Mol. Catal. 2020, 496, 111187. [Google Scholar] [CrossRef]
  111. Liu, X.; Huang, L.; Ma, Y.; She, G.; Zhou, P.; Zhu, L.; Zhang, Z. Enable biomass-derived alcohols mediated alkylation and transfer hydrogenation. Nat. Commun. 2024, 15, 7012. [Google Scholar] [CrossRef]
  112. Yao, Y.; Yin, C.; Ma, C.; Li, Y.; Wang, Y.; Jiang, R.; He, W.; Xiang, Z.; Liu, Y.; Li, X.; et al. Aromatic ethers induced electronic structure reconstruction on encapsulated Nickel catalysts for high-performance catalytic hydrogenation. ACS Appl. Mater. Interfaces 2024, 16, 14680–14693. [Google Scholar] [CrossRef] [PubMed]
  113. Xi, N.; Chen, S.; Bao, R.; Wang, Q.; Lin, Y.; Yue, J.; Wang, R.; Yang, C.; Yin, W.; Qiu, T. Layered carbon encapsulated CuOx nanopaticles for selective hydrogenation of furfural to furfuryl alcohol. Mol. Catal. 2024, 565, 114364. [Google Scholar] [CrossRef]
  114. Kholdeeva, O.A.; Skobelev, I.Y.; Ivanchikova, I.D.; Kovalenko, K.A.; Fedin, V.P.; Sorokin, A.B. Hydrocarbon oxidation over Fe- and Cr-containing metal-organic frameworks MIL-100 and MIL-101–a comparative study. Catal. Today 2014, 238, 54–61. [Google Scholar] [CrossRef]
  115. Yan, X.; Zhang, G.; Zhu, Q.; Kong, X. CuZn@N-doped graphene layer for upgrading of furfural to furfuryl alcohol. Mol. Catal. 2022, 517, 112066. [Google Scholar] [CrossRef]
  116. Rodiansono; Astuti, M.D.; Santoso, U.T.; Shimazu, S. Hydrogenation of biomass-derived furfural over highly dispersed-Aluminium hydroxide supported Ni-Sn(3.0) alloy catalysts. Procedia Chem. 2015, 16, 531–539. [Google Scholar] [CrossRef]
  117. Wang, Y.; Miao, Y.; Li, S.; Gao, L.; Xiao, G. Metal-organic frameworks derived bimetallic Cu-Co catalyst for efficient and selective hydrogenation of biomass-derived furfural to furfuryl alcohol. Mol. Catal. 2017, 436, 128–137. [Google Scholar] [CrossRef]
  118. Alba-Rubio, A.C.; Cecilia, J.A.; Jiménez-Gómez, C.P.; García-Sancho, C.; Cassidy, A.; Moreno-Tost, R.; Maireles-Torres, P. The role of MnOx in Cu-MnOx/SiO2 catalysts for the gas-phase hydrogenation of furfural. Mol. Catal. 2023, 546, 113224. [Google Scholar] [CrossRef]
  119. Sulmonetti, T.P.; Pang, S.H.; Claure, M.T.; Lee, S.; Cullen, D.A.; Agrawal, P.K.; Jones, C.W. Vapor phase hydrogenation of furfural over nickel mixed metal oxide catalysts derived from layered double hydroxides. Appl. Catal. A Gen. 2016, 517, 187–195. [Google Scholar] [CrossRef]
  120. Song, T.; Qi, Y.; Zhao, C.; Wu, P.; Li, X. Sorbitol-derived carbon overlayers encapsulated Cu nanoparticles on SiO2: Stable and efficient for the continuous hydrogenation of ethylene carbonate. iScience 2022, 25, 105239. [Google Scholar] [CrossRef]
  121. Yao, Y.; Wu, X.; Gutiérrez, O.Y.; Ji, J.; Jin, P.; Wang, S.; Xu, Y.; Zhao, Y.; Wang, S.; Ma, X.; et al. Roles of Cu+ and Cu0 sites in liquid-phase hydrogenation of esters on core-shell CuZnx@C catalysts. Appl. Catal. B Environ. 2020, 267, 118698. [Google Scholar] [CrossRef]
  122. Li, Y.; Guo, X.; Liu, D.; Mu, X.; Chen, X.; Shi, Y. Selective Conversion of Furfural to Cyclopentanone or Cyclopentanol Using Co-Ni Catalyst in Water. Catalysts 2018, 8, 193. [Google Scholar] [CrossRef]
  123. Yang, X.; Meng, Q.; Ding, G.; Wang, Y.; Chen, H.; Zhu, Y.l.; Li, Y.W. Construction of novel Cu/ZnO-Al2O3 composites for furfural hydrogenation: The role of Al components. Appl. Catal. A Gen. 2018, 561, 78–86. [Google Scholar] [CrossRef]
  124. 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]
  125. Zhu, T.; Chen, X.; Zhu, Z.; Tan, L.; Yuan, H.; Wu, J.; Zhu, C.; Xu, J. In Situ synthesis of Nitrogen-doped carbon coated copper: Boosting superhydrophobicity, conductivity, and oxidation resistance. ACS Appl. Mater. Interfaces 2024, 16, 68703–68711. [Google Scholar] [CrossRef]
  126. Wang, J.; Liu, S.; Song, P.; Zhang, X.; Song, Y.; Guo, W.; Yao, Z. Highly efficient activation of molecular oxygen by Cu@N doped carbon for antibiotic degradation. Chem. Eng. J. 2025, 507, 160769. [Google Scholar] [CrossRef]
  127. Guo, X.; Yu, B.; Wang, Z.; Li, S.; Chen, X.; Yang, Y. Selective hydrogenation of furfural to furfuryl alcohol over Cu/CeCoOx in aqueous phase. Mol. Catal. 2022, 529, 112553. [Google Scholar] [CrossRef]
  128. Wang, X.; Liang, T.; Zheng, Z.; Guo, R.; Zhang, Z.; Miao, L.; Yang, X. MOF-silica hybrid derived high performance K-Cu#SiO2 catalyst for furfural valorization: The functional role of potassium acetate (KAc) in hybridization and copper electronic state. Appl. Catal. A Gen. 2022, 640, 118603. [Google Scholar] [CrossRef]
  129. Xu, J.; Han, X.; Zhu, L.; Wang, W.; Ruan, L.; Yang, Z.; Ye, H.; Chen, B.H. Revealing the intrinsic relationship between nano/electronic structure of CuCo/NC (NC drived from ZIF-67) and their catalytic performance for furfural selective hydrogenation. J. Catal. 2025, 447, 116140. [Google Scholar] [CrossRef]
  130. Yu, X.; Zhang, J.; Yang, H.; Chong, S.; Liu, G.; Zhang, Y.; Wang, K. Effect of the metal-support interaction in the Cu/ZnO catalyst on its performance in the hydrogenation of furfural to furfuryl alcohol. J. Fuel Chem. Technol. 2024, 52, 1045–1056. [Google Scholar] [CrossRef]
  131. Kaim, J.; Śliwa, M.; Kuterasiński, Ł.; Samson, K.; Ruggiero-Mikołajczyk, M.; Zimowska, M.; Mordarski, G.; Karcz, R.; Podobiński, J.; Datka, J.; et al. Gas-phase hydrogenation and decarbonylation of furfuryl aldehyde over catalysts containing Cu and Ni. Catal. Today 2024, 441, 114897. [Google Scholar] [CrossRef]
  132. Kumar, A.; Shivhare, A.; Bal, R.; Srivastava, R. Metal and solvent-dependent activity of spinel-based catalysts for the selective hydrogenation and rearrangement of furfural. Sustain. Energy Fuels 2021, 5, 3191–3204. [Google Scholar] [CrossRef]
  133. Yang, X.; Chen, H.; Meng, Q.; Zheng, H.; Zhu, Y.; Li, Y.W. Insights into influence of nanoparticle size and metal–support interactions of Cu/ZnO catalysts on activity for furfural hydrogenation. Catal. Sci. Technol. 2017, 7, 5625–5634. [Google Scholar] [CrossRef]
  134. Yang, X.; Chen, Z.; Tan, J.; Zhang, Y.; Cui, J.; Wang, C.; Fang, L.; Zhu, Y.; Huang, L.; Shi, H.; et al. Understanding the synergistic catalysis of balanced Cu0-Cu+ sites and oxygen vacancies in Cu/ZrO2 catalysts for the efficient hydrogenation of furfural. Chem. Eng. J. 2024, 501, 157796. [Google Scholar] [CrossRef]
  135. Yu, D.; Yu, X.; Zhang, Y.; Wang, K. Deactivation mechanism of Cu/SiO2 catalyst in gas phase hydrogenation of furfural to furfuryl alcohol. J. Fuel Chem. Technol. 2023, 51, 1751–1760. [Google Scholar] [CrossRef]
  136. Costa, M.M.; Coelho, E.C.D.A.; Moya, S.F.; Suppino, R.S. Cascade reactions of furfural conversion with different bifunctional niobia-supported metal catalysts. React. Kinet. Mech. Catal. 2025, 1–23. [Google Scholar] [CrossRef]
  137. Zhang, C.; Xing, Z.; Li, Y.; Xu, T.; Sun, Y.; Bai, J. Alloying palladium with copper on zirconium dioxide for accelerating the selective catalytic transfer hydrogenation of furfural to furfuryl alcohol. New J. Chem. 2025, 49, 6684–6693. [Google Scholar] [CrossRef]
  138. 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 Gen. 2022, 633, 118527. [Google Scholar] [CrossRef]
  139. Zhao, S.; Zhang, Y.; Rao, Z.; Liu, H.; Zhang, R.; Jia, W.; Zhang, J.; Sun, Y.; Peng, L. High-performance CuNi-alloy catalysts for efficient solvent-free hydrogenation furfural to furfuryl alcohol. Appl. Catal. B Environ. Energy 2025, 371, 125228. [Google Scholar] [CrossRef]
  140. Pholjaroen, B.; Li, N.; Huang, Y.; Li, L.; Wang, A.; Zhang, T. Selective hydrogenolysis of tetrahydrofurfuryl alcohol to 1,5-pentanediol over vanadium modified Ir/SiO2 catalyst. Catal. Today 2015, 245, 93–99. [Google Scholar] [CrossRef]
  141. Tang, Q.; Sun, X.; Duan, Z.; Xiao, Z.; Liu, X. Efficient hydrogenation of biomass-derived furfural to tetrahydrofurfuryl alcohol over a non-noble Ni/CeO2 catalyst under mild conditions. Mol. Catal. 2025, 584, 115261. [Google Scholar] [CrossRef]
  142. Sahraie, N.R.; Kramm, U.I.; Steinberg, J.; Zhang, Y.; Thomas, A.; Reier, T.; Paraknowitsch, J.-P.; Strasser, P. Quantifying the density and utilization of active sites in non-precious metal oxygen electroreduction catalysts. Nat. Commun. 2015, 6, 8618. [Google Scholar] [CrossRef]
  143. Hadjiivanov, K.; Tsoncheva, T.; Dimitrov, M.; Minchev, C.; Knözinger, H. Characterization of Cu/MCM-41 and Cu/MCM-48 mesoporous catalysts by FTIR spectroscopy of adsorbed CO. Appl. Catal. A Gen. 2003, 241, 331–340. [Google Scholar] [CrossRef]
  144. Liao, X.; Zhao, H.; Liu, R.; Luo, H.; Lv, Y.; Liu, P. Highly efficient and selective hydrogenation of furfural to furfuryl alcohol and cyclopentanone over Cu-Ni bimetallic Catalysts: The crucial role of CuNi alloys and Cu+ species. J. Catal. 2024, 436, 115603. [Google Scholar] [CrossRef]
  145. Rehr, J.J.; Albers, R.C. Theoretical approaches to x-ray absorption fine structure. Rev. Mod. Phys. 2000, 72, 621–654. [Google Scholar] [CrossRef]
  146. Wang, M.; Árnadóttir, L.; Xu, Z.J.; Feng, Z. In Situ X-ray absorption spectroscopy studies of nanoscale electrocatalysts. Nano-Micro Lett. 2019, 11, 47. [Google Scholar] [CrossRef]
  147. Sharma, R.V.; Das, U.; Sammynaiken, R.; Dalai, A.K. Liquid phase chemo-selective catalytic hydrogenation of furfural to furfuryl alcohol. Appl. Catal. A Gen. 2013, 454, 127–136. [Google Scholar] [CrossRef]
  148. Kment, Š.; Bakandritsos, A.; Tantis, I.; Kmentová, H.; Zuo, Y.; Henrotte, O.; Naldoni, A.; Otyepka, M.; Varma, R.S.; Zbořil, R. Single atom catalysts based on earth-abundant metals for energy-related applications. Chem. Rev. 2024, 124, 11767–11847. [Google Scholar] [CrossRef]
  149. 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 Environ. 2021, 299, 120652. [Google Scholar] [CrossRef]
  150. 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. [Google Scholar] [CrossRef]
  151. 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]
  152. Meira, D.M.; Monte, M.; Fernández-García, M.; Meunier, F.; Mathon, O.; Pascarelli, S.; Agostini, G. A flexible cell for in situ combined XAS-DRIFTS-MS experiments. J. Synchrotron Radiat. 2019, 26, 801–810. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Diagram of multiple reaction pathways for FF conversion.
Scheme 1. Diagram of multiple reaction pathways for FF conversion.
Chemistry 07 00153 sch001
Scheme 2. Reaction mechanism diagram of FA preparation from FF hydrogenation catalyzed by Cu-based catalysts [27]. Reprinted with permission from Ref. [27]. Copyright 2022, Royal Society of Chemistry.
Scheme 2. Reaction mechanism diagram of FA preparation from FF hydrogenation catalyzed by Cu-based catalysts [27]. Reprinted with permission from Ref. [27]. Copyright 2022, Royal Society of Chemistry.
Chemistry 07 00153 sch002
Scheme 3. Preferential adsorption configurations of furfural over different surfaces, based on DFT calculations [33]. Reprinted with permission from Ref. [33].
Scheme 3. Preferential adsorption configurations of furfural over different surfaces, based on DFT calculations [33]. Reprinted with permission from Ref. [33].
Chemistry 07 00153 sch003
Scheme 4. Reaction mechanism diagram for the preparation of FA by catalytic hydrogenation of FF by the transition metals without Cu [27]. Reprinted with permission from Ref. [27]. Copyright 2022, Royal Society of Chemistry.
Scheme 4. Reaction mechanism diagram for the preparation of FA by catalytic hydrogenation of FF by the transition metals without Cu [27]. Reprinted with permission from Ref. [27]. Copyright 2022, Royal Society of Chemistry.
Chemistry 07 00153 sch004
Scheme 5. Possible reaction pathways during FF hydrogenation [62]. Adapted with permission from Ref. [62]. Copyright 2014, Elsevier.
Scheme 5. Possible reaction pathways during FF hydrogenation [62]. Adapted with permission from Ref. [62]. Copyright 2014, Elsevier.
Chemistry 07 00153 sch005
Scheme 6. Schematic representation of the MIL-101 and MIL-100 structures: M3O-carboxylate trimer—primary building unit, A and D—supertetrahedra, secondary building units, B and E—small cages, C and F—large cages of MIL-101 and MIL-100, respectively [114]. Reprinted with permission from Ref. [114]. Copyright 2014, Elsevier.
Scheme 6. Schematic representation of the MIL-101 and MIL-100 structures: M3O-carboxylate trimer—primary building unit, A and D—supertetrahedra, secondary building units, B and E—small cages, C and F—large cages of MIL-101 and MIL-100, respectively [114]. Reprinted with permission from Ref. [114]. Copyright 2014, Elsevier.
Chemistry 07 00153 sch006
Figure 1. (a) Cu LDH used as a precursor to prepare highly dispersed Cu/Cu0 active sites, (b) Cu 2p XPS, and (c) Cu LMM AES spectra of CuxMg3AlOy-1 catalysts [43]. Reprinted with permission from Ref. [43]. Copyright 2023, Elsevier.
Figure 1. (a) Cu LDH used as a precursor to prepare highly dispersed Cu/Cu0 active sites, (b) Cu 2p XPS, and (c) Cu LMM AES spectra of CuxMg3AlOy-1 catalysts [43]. Reprinted with permission from Ref. [43]. Copyright 2023, Elsevier.
Chemistry 07 00153 g001
Figure 2. (a) LDH-supported Cu as precursor to prepare well-dispersed Cu0/Cu+ nanodots, (b) Cu 2p XPS spectra, and (c) the corresponding AES spectra of Cu/Al2O3-R, Cu/MgO-R, Cu/Mg3Al-R, and CuMg3Al-R [44]. Reprinted with permission from Ref. [44]. Copyright 2025, Elsevier.
Figure 2. (a) LDH-supported Cu as precursor to prepare well-dispersed Cu0/Cu+ nanodots, (b) Cu 2p XPS spectra, and (c) the corresponding AES spectra of Cu/Al2O3-R, Cu/MgO-R, Cu/Mg3Al-R, and CuMg3Al-R [44]. Reprinted with permission from Ref. [44]. Copyright 2025, Elsevier.
Chemistry 07 00153 g002
Figure 3. (a) Synthesis flowchart of PtCu@S-1; (b) N2 adsorption–desorption isotherms of the as-synthesized samples [48]. Reprinted with permission from Ref. [48]. Copyright 2024, Elsevier.
Figure 3. (a) Synthesis flowchart of PtCu@S-1; (b) N2 adsorption–desorption isotherms of the as-synthesized samples [48]. Reprinted with permission from Ref. [48]. Copyright 2024, Elsevier.
Chemistry 07 00153 g003
Figure 4. (a) Schematic diagram of the adjusted metal-support interaction by Al promoter, (b) XRD patterns of the reduced samples, (c) H2-TPR patterns of the samples [123]. Reprinted with permission from Ref. [123]. Copyright 2018, Elsevier.
Figure 4. (a) Schematic diagram of the adjusted metal-support interaction by Al promoter, (b) XRD patterns of the reduced samples, (c) H2-TPR patterns of the samples [123]. Reprinted with permission from Ref. [123]. Copyright 2018, Elsevier.
Chemistry 07 00153 g004
Figure 5. (a) Schematic diagram of HEA’s structure with multiple core–shell structures, (b,c) the TEM and IFFT images, and SEM images of ultra-dilute HEAs with core–shell structure [59]. Reprinted with permission from Ref. [59]. Copyright 2023, Elsevier.
Figure 5. (a) Schematic diagram of HEA’s structure with multiple core–shell structures, (b,c) the TEM and IFFT images, and SEM images of ultra-dilute HEAs with core–shell structure [59]. Reprinted with permission from Ref. [59]. Copyright 2023, Elsevier.
Chemistry 07 00153 g005
Figure 6. (a) The synthesis strategy of ZJU-199-350; (b) HRTEM images of ZJU-199-350 (insets show representations of Cu/Cu2O heterostructures; Cu green, Cu2O red) [61]. Reprinted with permission from Ref. [61]. Copyright 2020, John Wiley and Sons.
Figure 6. (a) The synthesis strategy of ZJU-199-350; (b) HRTEM images of ZJU-199-350 (insets show representations of Cu/Cu2O heterostructures; Cu green, Cu2O red) [61]. Reprinted with permission from Ref. [61]. Copyright 2020, John Wiley and Sons.
Chemistry 07 00153 g006
Figure 7. (a) XRD patterns of Cu/CeCoOx catalysts: (a) Cu/CoOx; (b) Cu(5)/Ce(2)Co(18)Ox; (c) Cu(5)/Ce(6)Co(14)Ox; (d) Cu(5)/Ce(10)Co(10)Ox; (e) Cu(5)/Ce(14)Co(6)Ox; (f) Cu(5)/Ce(18)Co(2)Ox; (g) Cu/CeO2 [127]; (b) nitrogen sorption plots for the Cu#SiO2 and K-Cu#SiO2 [128]; (c) SEM images of the catalysts Cu5.0CeO2 [131]; (d) HR-TEM image for Cu-MOF before catalysis at 50 nm scale [53]. 7 (a) was reprinted with permission from Ref. [127]. Copyright 2022, Elsevier. 7 (b) was reprinted with permission from Ref. [128]. Copyright 2022, Elsevier. 7 (c) was reprinted with permission from Ref. [131]. Copyright 2024, Elsevier. 7 (d) was reprinted with permission from Ref. [53].
Figure 7. (a) XRD patterns of Cu/CeCoOx catalysts: (a) Cu/CoOx; (b) Cu(5)/Ce(2)Co(18)Ox; (c) Cu(5)/Ce(6)Co(14)Ox; (d) Cu(5)/Ce(10)Co(10)Ox; (e) Cu(5)/Ce(14)Co(6)Ox; (f) Cu(5)/Ce(18)Co(2)Ox; (g) Cu/CeO2 [127]; (b) nitrogen sorption plots for the Cu#SiO2 and K-Cu#SiO2 [128]; (c) SEM images of the catalysts Cu5.0CeO2 [131]; (d) HR-TEM image for Cu-MOF before catalysis at 50 nm scale [53]. 7 (a) was reprinted with permission from Ref. [127]. Copyright 2022, Elsevier. 7 (b) was reprinted with permission from Ref. [128]. Copyright 2022, Elsevier. 7 (c) was reprinted with permission from Ref. [131]. Copyright 2024, Elsevier. 7 (d) was reprinted with permission from Ref. [53].
Chemistry 07 00153 g007
Figure 8. (a) FTIR spectra of annealed samples of ZJU-199 at different temperatures [61]; (b) Raman spectra of the catalysts. a: CZ-300, b: CZ-400, c: CZ-500, d: CZ-600 [134], (c) XPS spectra of Pd–ZrO2 catalyst Cu 2p [137]; (d) EPR spectra of the catalysts. a: CZ-300, b: CZ-400, c: CZ-500, d: CZ-600 [134], (e) H2-TPR curves of Cu-SBA-GLY and 10Cu-SBA-DP samples [138]; (f) NH3-TPD profiles for: SiO2 support (dashed line), Cu5.0SiO2 (solid line) and Ni5.0SiO2 (dotted line) [131]; (g) CO2-TPD of CuaNibOx/C catalysts [139]. 8 (a) was reprinted with permission from Ref. [61]. Copyright 2020, John Wiley and Sons. 8 (b) was reprinted with permission from Ref. [134]. Copyright 2024, Elsevier. 8 (c) was reprinted with permission from Ref. [137]. Copyright 2025, Royal Society of Chemistry. 8 (d) was reprinted with permission from Ref. [134]. Copyright 2024, Elsevier. 8 (e) was reprinted with permission from Ref. [138]. Copyright 2022, Elsevier. 8 (f) was reprinted with permission from Ref. [131]. Copyright 2024, Elsevier. 8 (g) was reprinted with permission from Ref. [139]. Copyright 2025, Elsevier.
Figure 8. (a) FTIR spectra of annealed samples of ZJU-199 at different temperatures [61]; (b) Raman spectra of the catalysts. a: CZ-300, b: CZ-400, c: CZ-500, d: CZ-600 [134], (c) XPS spectra of Pd–ZrO2 catalyst Cu 2p [137]; (d) EPR spectra of the catalysts. a: CZ-300, b: CZ-400, c: CZ-500, d: CZ-600 [134], (e) H2-TPR curves of Cu-SBA-GLY and 10Cu-SBA-DP samples [138]; (f) NH3-TPD profiles for: SiO2 support (dashed line), Cu5.0SiO2 (solid line) and Ni5.0SiO2 (dotted line) [131]; (g) CO2-TPD of CuaNibOx/C catalysts [139]. 8 (a) was reprinted with permission from Ref. [61]. Copyright 2020, John Wiley and Sons. 8 (b) was reprinted with permission from Ref. [134]. Copyright 2024, Elsevier. 8 (c) was reprinted with permission from Ref. [137]. Copyright 2025, Royal Society of Chemistry. 8 (d) was reprinted with permission from Ref. [134]. Copyright 2024, Elsevier. 8 (e) was reprinted with permission from Ref. [138]. Copyright 2022, Elsevier. 8 (f) was reprinted with permission from Ref. [131]. Copyright 2024, Elsevier. 8 (g) was reprinted with permission from Ref. [139]. Copyright 2025, Elsevier.
Chemistry 07 00153 g008
Figure 9. (a) XANES spectra: (A) Cu foil, (B) Cat-C (freshly reduced), and (C) Cat-C (fresh), (b) TPR profiles of Cat-C (fresh), Cat-C (regenerated), Cat-C (used), and Cat-C (freshly reduced) [147]. Reprinted with permission from Ref. [147]. Copyright 2013, Elsevier.
Figure 9. (a) XANES spectra: (A) Cu foil, (B) Cat-C (freshly reduced), and (C) Cat-C (fresh), (b) TPR profiles of Cat-C (fresh), Cat-C (regenerated), Cat-C (used), and Cat-C (freshly reduced) [147]. Reprinted with permission from Ref. [147]. Copyright 2013, Elsevier.
Chemistry 07 00153 g009
Figure 10. (a) EXAFS spectra in k-space (k-weight = 3) and (b) R-space (k-weight = 3) for PdCu/Al2O3 catalysts (reduced ex situ) along with Pd and Cu reference foils. Dashed-lined rectangles indicate k ranges over which the data were then Fourier transformed and analyzed, (c) STEM/TEM images of Pd1Cu216/Al2O3, and (d) lognormal STEM/TEM size distributions for the catalysts reduced ex situ at 300 °C for 0.5 h under flowing H2 [149]. Reprinted with permission from Ref. [149]. Copyright 2021, Elsevier.
Figure 10. (a) EXAFS spectra in k-space (k-weight = 3) and (b) R-space (k-weight = 3) for PdCu/Al2O3 catalysts (reduced ex situ) along with Pd and Cu reference foils. Dashed-lined rectangles indicate k ranges over which the data were then Fourier transformed and analyzed, (c) STEM/TEM images of Pd1Cu216/Al2O3, and (d) lognormal STEM/TEM size distributions for the catalysts reduced ex situ at 300 °C for 0.5 h under flowing H2 [149]. Reprinted with permission from Ref. [149]. Copyright 2021, Elsevier.
Chemistry 07 00153 g010
Figure 11. 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 [150]. Reprinted with permission from Ref. [150].
Figure 11. 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 [150]. Reprinted with permission from Ref. [150].
Chemistry 07 00153 g011
Figure 12. Cu 2p in situ XPS spectra of (a) calcined samples and (b) fresh catalysts: a Cu/SiO2-IM, b Cu/SiO2-DP, c Cu/SiO2-IE, d Cu/SiO2-HDP, and e Cu/SiO2-EA [151]. Reprinted with permission from Ref. [151]. Copyright 2019, Elsevier.
Figure 12. Cu 2p in situ XPS spectra of (a) calcined samples and (b) fresh catalysts: a Cu/SiO2-IM, b Cu/SiO2-DP, c Cu/SiO2-IE, d Cu/SiO2-HDP, and e Cu/SiO2-EA [151]. Reprinted with permission from Ref. [151]. Copyright 2019, Elsevier.
Chemistry 07 00153 g012
Figure 13. Panels (a,b) are in situ DRIFTS spectra of the FF hydrogenation on the Cu/MgAlOx, while panels (c,d) are calculated energy profiles for the potential comparative free energy of FF to FA on the Co (111), Cu (111), and CuxCo1 (111) with different Cu/Co ratios crystal surfaces [55]. Reprinted with permission from Ref. [55]. Copyright 2023, Elsevier.
Figure 13. Panels (a,b) are in situ DRIFTS spectra of the FF hydrogenation on the Cu/MgAlOx, while panels (c,d) are calculated energy profiles for the potential comparative free energy of FF to FA on the Co (111), Cu (111), and CuxCo1 (111) with different Cu/Co ratios crystal surfaces [55]. Reprinted with permission from Ref. [55]. Copyright 2023, Elsevier.
Chemistry 07 00153 g013
Table 1. Reaction conditions and results of FF liquid-phase hydrogenation to FA over the Cu-based catalysts.
Table 1. Reaction conditions and results of FF liquid-phase hydrogenation to FA over the Cu-based catalysts.
CatalystMetal Loading (wt%)SolventTemperature (°C)Time (h)H2 Pressure (bar)FF Conversion (%)FA Selectivity (%)FA Yield (%)Ref.
Cu/Zn/Al2O3 (CZAl)10H2O100410100~9999[40]
Cu/Zn/Al2O3 (CZAl)10Ethanol10041099~9998.01
Cu/Zn/Al2O3 (CZAl)10Methanol10041099~9594.05
Cu/Zn/Al2O3 (CZAl)102-Propanol10041019~9017.1
Cu/Zn/Al2O3 (CZAl)10DMF10041098~9997.02
Cu/Zn/Al2O3 (CZAl)10Acetone10041098~8078.4
Cu/Zn/Al2O3 (CZAl)10Toluene10041077~9976.23
Cu/Zn/Al2O3 (CZAl)10Cyclohexane1004104~994.0
Cu/MgO5Isopropyl alcohol109.852.32099.999.999.98[41]
Cu/Al2O3 (Acetate)1MeOH5071.524.2 ± 1.296.0 ± 4.823.23[42]
Cu/Al2O3 (Acetate)5MeOH5071.547.7 ± 2.497.6 ± 4.946.56
Cu/Al2O3 (Sulfate)1MeOH5071.52.2 ± 0.15.1 ± 0.30.11
Cu/Al2O3 (Sulfate)5MeOH5071.57.8 ± 0.40.8 ± 0.10.062
Cu0.9Mg3AlOy-Isopropanol1201.5169998.497.42[43]
CuMg3Al-R20i-PrOH13032010099.399.3[44]
Cu/CeO2-R (nanorod)5γ-butyrolactone10042097.596.293.8[45]
Cu/η-Al2O35Isopropanol15021099.79493.7[46]
NPCu@SiO2-H2O150430623119.2[47]
SiO2@Cu7.6H2O150430841714.3
PtCu@S-1 (S-1, silicalite-1)0.8THF16062099.910099.9[48]
Cu@MFI-H2O70540100100100[49]
Co-Cu/SBA-1510Isopropanol170420998079.2[50]
Cu/MCM-415i-PrOH12041100100100[51]
Cu/AC-SO3H162-propanol10524100100100[52]
[Cu2(L1)2·5DMF·4H2O]n-Methanol14024507610076[53]
Fe3O4/Cu@C27.4n-butanol18041098.589.488.1[54]
Fe/Cu@C30.4i-propanol300410~60%~9054[54]
Cu3Co1/MgOx20Isopropanol1102201009999[55]
Cu2Zn/SiO213Deionized water12042581.994.877.6[56]
Na–Cu@TS-12.1Isopropanol1102109398.191.2[57]
Cu/SiO2-AE5Isopropanol9021055.299.955.1[58]
NiCoCuZnFe/C-800-Isopropanol9093087.3210087.32[59]
CuO#TiO212Ethanol140120999999[60]
ZJU-199-350-Isopropanol130310979996[61]
Table 2. Comparison of the advantages and disadvantages of different support.
Table 2. Comparison of the advantages and disadvantages of different support.
SupportAdvantageDisadvantage
Metal oxideLow cost, easy to regulate, well-studied for applicationRelatively undeveloped porous structure
SiO2Low cost, high chemical inertnessIrregular structure, poor porosity
Microporous molecular sievesWell-defined topological framework with uniform micropores, high hydrothermal stabilityStrong surface acidity, limited pore size
Mesoporous molecular sievesOrdered mesoporous structure, relatively facile preparationStructural hydrothermal instability
Carbon-containing materialStructural diversity, facile functionalizationSevere side reactions
MOFDeveloped pore structure, high surface area, and abundant metal anchoring sitesExpensive organic ligands, easy decomposability
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

Lin, T.; Gao, Y.; Li, C.; Zhang, M.; Liu, Z. A Brief Review of Cu-Based Catalysts for the Selective Liquid-Phase Hydrogenation of Furfural to Furfuryl Alcohol. Chemistry 2025, 7, 153. https://doi.org/10.3390/chemistry7050153

AMA Style

Lin T, Gao Y, Li C, Zhang M, Liu Z. A Brief Review of Cu-Based Catalysts for the Selective Liquid-Phase Hydrogenation of Furfural to Furfuryl Alcohol. Chemistry. 2025; 7(5):153. https://doi.org/10.3390/chemistry7050153

Chicago/Turabian Style

Lin, Tiantian, Yongzhen Gao, Chao Li, Meng Zhang, and Zhongyi Liu. 2025. "A Brief Review of Cu-Based Catalysts for the Selective Liquid-Phase Hydrogenation of Furfural to Furfuryl Alcohol" Chemistry 7, no. 5: 153. https://doi.org/10.3390/chemistry7050153

APA Style

Lin, T., Gao, Y., Li, C., Zhang, M., & Liu, Z. (2025). A Brief Review of Cu-Based Catalysts for the Selective Liquid-Phase Hydrogenation of Furfural to Furfuryl Alcohol. Chemistry, 7(5), 153. https://doi.org/10.3390/chemistry7050153

Article Metrics

Back to TopTop