Next Article in Journal
Intercalated-Laurate-Enhanced Photocatalytic Activities of Ni/Cr-Layered Double Hydroxides
Previous Article in Journal
Controlled Synthesis of Ag-SnO2/α-Fe2O3 Nanocomposites for Improving Visible-Light Catalytic Activities of Pollutant Degradation and CO2 Reduction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Some Insights into the Use of Heterogeneous Copper Catalysts in the Hydroprocessing of Levulinic Acid

1
Consiglio Nazionale delle Ricerche, Istituto di Scienze e Tecnologie Chimiche “G. Natta”, Via Golgi 19, 20133 Milano, Italy
2
Dipartimento di Chimica, Università degli Studi di Milano, Via Golgi 19, 20133 Milano, Italy
*
Author to whom correspondence should be addressed.
Catalysts 2023, 13(4), 697; https://doi.org/10.3390/catal13040697
Submission received: 28 February 2023 / Revised: 29 March 2023 / Accepted: 29 March 2023 / Published: 4 April 2023
(This article belongs to the Special Issue Advances on Catalysts Based on Copper, 2nd Edition)

Abstract

:
Levulinic acid and its esters are close to being extensively produced through consolidated industrial processes, thus playing a central role in biobased industries producing commodities within the principles of the circular economy. One of the main pathways of levulinic acid and ester valorization is their transformation with hydrogen to obtain γ-valerolactone, valeric esters, 1,4-pentanediol and 2-methyl tetrahydrofuran. These reactions are catalyzed by noble and non-noble metal-based heterogeneous catalysts. The use of an abundant and non-toxic element, such as copper, is advantageous with respect to expensive or harmful metals, such as Rh, Ru, Pt or Ni. In this critical review, we wish to give a deeper insight into research advancements in the last ten years regarding the processing of levulinic acid and its esters with hydrogen using heterogeneous copper catalysts.

1. Introduction

1.1. Importance of LA/LEs and Scope of the Review

Levulinic acid (LA) is likely the most attractive platform molecule derived from lignocellulose, as it can be produced in relatively high yields from this renewable source. Indeed, it has been identified by the United States Department of Energy as one of the top 12 promising bio-based building blocks [1,2]. LA is typically produced using mineral acids (i.e., HCl or H2SO4) at around 200 °C, in a process in which C-6 sugars are converted into LA and formic acid, while C-5 sugars yield furfural. Biofine (by Biofine Renewables) is one of the most celebrated technologies in the production of this chemical, even if the first attempt at a conversion plant based on this process, which was developed in Caserta (Italy) by GFBiochemicals, has not gone into operation thus far. An improved process was recently designed, and a joint venture between GFBiochemicals and Towell Engineering group (named NXTLEVVEL) is currently working on a new production plant [1,2,3]. Levulinic esters (levulinates, LEs), especially methyl and ethyl levulinates, are also convenient platform chemicals, instead of LA, because of their specific physicochemical properties. LEs can be directly obtained from lignocellulose in the presence of an alcohol, in particular from the cellulose stream through the transformation of glucose [4,5,6], or from the hemicellulose stream, via a domino reaction from furfural [7,8,9,10]. Compared to LA, LEs show easier product separation [4] and they limit the metal leaching when a reaction is carried out in the presence of a metal-based heterogeneous catalyst [5,11].
As already pointed out, the synthetic routes of LA and LEs commonly involve the use of mineral acids that result in equipment corrosion, difficult separation and purification and potential environmental pollution, thus limiting the sustainability of the process. For these reasons, research is also oriented towards heterogeneous catalysts (i.e., Brønsted acid-based ones, such as zeolites, or ion exchange resins), but they result in lower LA/LE yields and easy deactivation [2,6,12].
LA and LEs are near to being extensively produced through consolidated and sustainable processes, and therefore they could play a central role in biobased industries producing commodities within the principles of the circular economy. Indeed, they can be transformed into different chemicals, such as γ-valerolactone (GVL), succinic acid, diphenolic acid, 1,4-pentanediol (1,4-PDO), valeric esters, 2-methyl tetrahydrofuran (2-MTHF), aminolevulinic acid, ketals and others [1,13,14,15], by different catalyzed processes (e.g., hydrogenation, oxidation, amination, condensation). In this critical review, we wish to give a deeper insight into recent research advancements (over the last ten years) related to the transformation of LA/LEs with hydrogen using heterogeneous copper-based catalysts. Thus, we will focus our attention on the following products: GVL, valeric esters, 1,4-PDO, and 2-MTHF (Scheme 1). The processes of catalytic transfer hydrogenation (CTH), as an alternative to the use of molecular hydrogen, will also be examined, although copper-based materials find a limited diffusion compared to others (e.g., Zr-catalysts).

1.2. General Considerations about Copper Catalysts

Heterogeneous Cu-based catalysts are materials used in a wide range of chemical transformations, especially those linked to the methanol economy, but also in organic chemistry (i.e., C=O and C=C hydrogenation, coupling reactions and click chemistry, dehydrogenations, deoxygenation) [16], environmental remediation and deNOx activity [17,18,19,20,21]. Copper is a non-noble metal, with a very low toxicity, compared to other metals such as Ni [22,23].
Moving to the objective of this review, several copper-based catalysts have been designed for the hydrogenation of LA/LEs in the last ten years. The dominant controlling factors are Cu-support interactions, morphology and dispersion of the Cu phase and the use of a second metal (e.g., Ni, Ag). The preparation method is the key, but the combination and the rationalization of all these parameters to obtain a highly performant catalyst is not an easy task and, at the same time, not always linear. One of the main issues preventing their large-scale application is deactivation, which mainly occurs through copper leaching, metal sintering or carbon deposition [24,25,26,27,28]. Catalyst stability and durability represent priorities in the scaling-up of the process at an industrial level. It is worth noting that sintering phenomena are in strong correlation with temperature. In particular, copper Hüttig temperature (that is, the semi-empirical value at which atoms at the surface start to exhibit mobility) is only 134 °C, while the reactions of LA/LE hydrogenation are normally carried out over 200 °C [29,30], thus limiting the lifetime (and the applicability) of copper catalysts. Traditionally, Adkins-type catalysts, namely copper chromites, have been used for hydrogenation of carboxylic acids and esters to alcohols due to their effectiveness and inexpensiveness [31,32]. Because of these reasons, since the 1950s, they have been applied in the production of biomass-derived chemicals, in particular in the hydrogenation of LA and LEs to GVL [33]; some examples can also be found in the recent literature [34,35]. However, despite copper chromites being simple and cheap catalysts, they suffer from toxicity and recyclability problems and could rapidly undergo deactivation, especially when free carboxylic acids are used. In particular, the presence of free acids or short-chain alcohols slows down the hydrogenation reaction rate [31,36].

2. LA/LE Hydrogenation to GVL and Valerates

There is a growing interest in the production of GVL under environmentally friendly conditions because it is a safe, non-toxic, biodegradable and renewable-based added-value chemical, with many applications [37,38,39]. GVL can be (i) used as a biofuel additive, with combustion properties comparable to ethanol when blended with gasoline [40,41,42,43], (ii) converted to alkyl valerates, which are used as fuel components for gasoline and diesel engines [24,39,44], (iii) an environmentally safe solvent [25,40] hydrogenated to form 1,4-PDO, which undergoes dehydration to form 2-MTHF [45,46], and (iv) employed as a renewable monomer to produce polyesters [47,48]. In addition, it can be used as cutting oil, brake fluid and as a coupling agent in dye baths [26], converted to methyl pentenoate, which can be further transformed into nylon intermediates, such as adipic acid, dimethyl adipate or caprolactam [27], or even used as a food additive [49]. Several noble metals, such as supported Ru, Rh, Pt and Ir, have also been recently employed [41,50,51,52,53,54,55,56,57,58,59] for the heterogeneous catalyzed selective hydrocyclization of LA/LEs to GVL, in high yield. However, noble metals are expensive, and their supply is a limitation for large-scale plant applications. On the contrary, the use of abundant elements, such as Cu and Zr, offers more environmentally friendly routes because of their higher abundance and lower price.

2.1. Hydrogenation of LA/LEs to GVL with Molecular Hydrogen

In Table 1, the main results obtained in the last ten years in the hydrogenation of LA/LEs to GVL with copper-based heterogeneous catalysts are summarized. The reactions are performed by using water, alcohols (primary or secondary) or, rarely, organic solvents such as THF [11], dioxane [60] or even n-hexane [61]. However, some of these compounds are toxic or harmful, thus their replacement is a priority to improve the sustainability of the process. This is the case with n-hexane, which is neurotoxic [62,63], and of THF, which is suspected to be carcinogenic [64]. 2-MTHF, which is less toxic and which can be derived from LA/LEs via GVL (vide infra), can be proposed instead of THF [46,65,66]. The solvent has a strong impact on the reaction, and from this point of view, water is considered a green solvent and LA is highly soluble in this medium. However, easier copper leaching is observed in water, thus reducing the lifetime of the catalyst [5,67]. Therefore, the use of organic solvents and/or LEs can have a positive effect on the phenomenon. It is worth noting that a reaction under no solvent conditions is reported [68], following the principles that ‘‘the best solvent is no solvent’’ [69]. The vapor phase hydrogenation of LA was indeed performed in a fixed-bed down-flow gas reactor, over a bimetallic Cu-Ni/SiO2 catalyst, and provided a conversion (C) and GVL selectivity (S) of 98% and 97%, respectively [68].
The temperature range is generally between 180 and 250 °C. Remarkably, in some cases, it has been possible to reduce the temperature to 110, 120 or 140 °C [70,71,72]. This was achieved using, respectively, Cu–Ni bimetallic NPs supported on functionalized SiO2 nanosphere (C = 99% and S = 97% in 13 h), CuMgAl mixed oxides derived from calcined hydrotalcites (C = 100% and S = 95%) or Cu/Al2O3-SiO2 (C = 79% and S = 96%) in continuous plants.
Usually, the reactions are performed using high-pressure batch reactors at a H2 pressure >25 atm. On the other hand, the literature on continuous-flow systems is still scarce [41,42,43,72], despite their role as a step ahead towards industrial scale-up [30,68,73,74,75].
As summarized in Scheme 2, GVL can be synthesized from LA/LEs via two reaction pathways, both containing hydrogenation and dehydration steps. The first one involves the hydrogenation of the carbonyl functionality of LA, thereby forming the unstable intermediate 4-hydroxypentanoic acid (4-HPA), and the subsequent dehydration followed by an intramolecular esterification (ring closure) to yield GVL. The other one is promoted by an acid catalyzed dehydration of LA to form angelica lactone (AL), followed by its hydrogenation to GVL [24,39,44]. Starting from LEs, the reaction was found to preferentially proceed via the first pathway, in which the levulinate hydrogenation forms the 4-hydroxy pentanoic ester (HPE), which undergoes a rapid lactonization to yield GVL [25,45].
As already stated, the use of LEs instead of LA has a positive effect on catalyst stability, since it can limit copper leaching [5,11]. Regardless of the reaction pathway, a well-balanced amount of metal sites and of acidic ones for the hydrogenation and the dehydration, respectively, is a prerequisite to obtain an efficient catalyst [44,45,46].
Metallic copper is indeed required for the hydrocyclization of LA/LEs to give GVL, as well highlighted in the literature [76,77,78]. Interestingly, some publications have pointed out that a balanced Cu(0) and Cu(I) surface distribution can promote the hydrogenation of ester, acyl and ketone groups [30,79,80,81]. In particular, Cu(0) facilitates the H2 adsorption and dissociation, while Cu(I) can activate LA/LEs C=O groups [30]. Small Cu nanoparticles (NPs) and a highly dispersed copper phase are reported to be crucial factors to achieve good catalytic performances [76,82] on a high surface area support [83], although they would easily sinter and deactivate during the reaction (we already pointed out the low Cu Hüttig temperature, namely 134 °C) [30]. Therefore, the catalyst preparation method and support are precious tools helping to overcome this issue.
Copper-based catalysts for the hydrogenation of LA/LEs to GVL were prepared over different oxides or mixed oxides such as SiO2, TiO2, ZrO2, Al2O3, Cr2O3, BaO, SiO2-ZrO2, Al2O3-ZrO2, WO3–ZrO2 and others [76,77,82]. The ZrO2-based ones are the most widespread [5,30,61,73,76,77,82,84]. Indeed, the fruitful combination between copper and zirconia has been demonstrated to be a winning strategy in the design of active catalysts with tunable properties for many applications (e.g., methanol economy) [85]. High surface area [30,76] and strong Cu-support interactions [30,61,75,77,82] enhance the dispersion of the metal phase, especially at high Cu loading, and reduce the sintering. Despite this, many papers (among these the majority of those related to Zr-based materials) report catalysts with very low surface area (i.e., <100 m2/g, but also <50 m2/g) [5,61,71,73,77,82,84] and this usually results in either reduced copper dispersion or metal loading, thus leaving space for further improvements. Some oxides (e.g., Al2O3) [11,74,86,87] more commonly lead to high surface area materials, while for other systems, the use of an amorphous phase [88], instead of crystalline ones, or their dispersion into a high surface area matrix (e.g., SiO2) would be preferable [89,90]. This is the case, for example, with ZrO2.
The support not only affects the metal dispersion, but it also plays a direct role in the activation of the reactants. The presence of acid sites, both Lewis and Brønsted ones, for instance, favors C=O polarization or the dehydration reaction [5,71,74]. It is curious that this aspect is often neglected in many papers that focus their attention mainly on hydrogenation activity. Balla et al. [73] compared a series of Cu-supported catalysts prepared on different oxides (ZrO2, Al2O3, SiO2, TiO2). They found that the most acidic material, namely Cu/Al2O3, was the most active one, resulting in conversion and selectivity even higher than the Cu/ZrO2 (C = 98% and S = 87%, respectively, vs. C = 81% and S = 83%). Previously, Hengne et al. also showed that supports not exhibiting acid sites (e.g., BaO) have a considerably low activity when compared to ZrO2- and Al2O3-based catalysts, which both provided C = 100% and S > 90% [5]. Unfortunately, these papers do not discuss in depth the role of the acid sites and, in particular, their nature (Lewis vs. Brønsted) and strength. In 2019, a more in-depth analysis of the acid–base properties of the catalyst was performed, through pyridine-adsorbed FT-IR, and NH3 and CO2 TPD measurements on a series of CuMgAl catalysts with different Mg/Al ratios [71]. In this case, moderately strong Brønsted acidity appears to be the most suitable feature to obtain GVL from LA, while the introduction of basic sites promotes the formation of 1,4-PDO through ring opening. In this paper, the authors present the Brønsted acid sites’ role in suppressing the ring-opening reaction. This contrasts with what was reported one year earlier in another work, where this side reaction was ascribed to the strong Brønsted acidity [70]. In this view, we agree that the control of the acid strength is critical to promote/inhibit the ring-opening reaction, analogous to what we reported on furfural and HMF [91].
The choice of the support is not the only way to tune the acid–base properties, but also the metal phase can have an important role. Ravasio and co-workers [92] showed the possibility of preparing very small copper nanoparticles with a Lewis acid character (Figure 1) with a peculiar, but easy, preparation method (namely, chemisorption–hydrolysis—CH). A similar effect can be obtained by stabilizing Cu+ species on the catalyst surface using a support such as Al2O3 or ZrO2. Indeed, as discussed above, it has been shown that Cu+ possesses an acidic character [61,93]. Alternatively, the acidity of the catalyst can be tuned by introducing a second metal (e.g., Ni) [86] or by catalyst doping strategies [30]. This is the case of Li et al., who used boron oxides to modulate the distribution of Cu0 and Cu+ on a Cu/ZrO2. This resulted in a double effect: on one side, the co-presence of Cu0 and Cu+ species facilitates the H2 adsorption and the activation of LA C=O groups; on the other side, the presence of boron oxide retarded the Cu sintering. This led to a total conversion and a selectivity above 90% in a long-term experiment (up to 100 h) in a fixed-bed reactor at a low temperature of 140 °C [30].
One of the most common strategies to obtain more stable materials is the introduction of another metal (e.g., Ni or Ag) to obtain bimetallic catalysts. Monometallic Ni-based supported materials show high activity, but low selectivity due to the formation of side products and a significant amount of deposited carbon [87]. On the other hand, the intrinsic synergy between Cu and Ni particles in supported bimetallic Cu-Ni catalysts was proposed to be responsible for the higher activity, stability and recyclability. Indeed, the presence of Cu in bimetallic catalysts enhances the reducibility of NiO, behaving as an activation site for H2 molecules [86]. In the meantime, the reduced Ni species act as a co-active site, able to accelerate the H2 adsorption and activation, thus lowering the reduction temperature of bimetallic samples [70]. Compared to monometallic systems, the promotion effects of Ni allowed the achieving of full LA conversion after 2 h, while Cu sites were responsible for selectivity improvements (S = 96%) [87]. Moreover, Ni was found to prevent copper leaching [60] and sintering [94], while Cu was capable of suppressing coke formation [87]. Nevertheless, nickel has been classified as a toxic metal and its use should be avoided for a more environmentally friendly process [95,96].
A remarkable attempt to improve Cu catalysts using a different doping metal, namely Ag, was carried out by Zhang et al. [11]. The authors proposed that Ag promotes the CuO in situ reduction during the reaction with less severe sintering. In addition, Ag prevents Cu leaching by maintaining it at metallic state. This improved the catalyst stability, allowing nine catalyst recycles under mild reaction conditions, and providing complete conversion and selectivity in only 1 h at 180 °C.
In Table 1, we summarized the catalytic behavior of the supported Cu or Cu-Ni catalysts for the hydrocyclization of LA/LEs to GVL in the presence of molecular H2.
Table 1. Catalysts for the selective hydrogenation of LA/LE to GVL.
Table 1. Catalysts for the selective hydrogenation of LA/LE to GVL.
Batch
CatalystPreparation ProcedureReaction ConditionsCatalytic Performances (%)Ref.
Cu/ZrO2CPLA/LEs, H2O or
MeOH, molLA/gcat = 0.086 mol g−1 in H2O,
molLE/gcat = 0.069 mol g−1 in MeOH,
molLA/gcat = 0.086 mol g−1 in MeOH, 200 °C, 5 h,
P(H2) = 35 bar, batch
C = 100
S > 90
[5]
Cu-Ni/γAl2O3WI
SG
LA, H2O, molLA/gcat = 0.086 mol g−1, 250 °C, 6 h,
P(H2) = 65 bar, batch
C = 100
S = 96
[87]
(1) Cu-WO3/ZrO2
(2) Cu/ZrO2
CP
OG
LA, (1) EtOH or (2) H2O,
molLA/gcat = 0.022 mol g−1, 200 °C, 6 h,
P(H2) = 50 bar, batch
C = 100
(1) S = 94
(2) S = 99
[84]
Cu/ZrO2CPLA, H2O, molLA/gcat = 0.086 mol g−1, 200 °C, 2 h,
P(H2) = 35 bar, batch.
C = 100
S = 80
[76]
Cu/ZrO2OG
DP
MT
LA, H2O, molLA/gcat = 0.086 mol g−1, 200 °C, 1 h,
P(H2) = 35 bar, batch
C = 90
S = 100
[77]
Cu-Ag/Al2O3WILA, THF, molLA/gcat = 0.017 mol g−1, 180 °C, 1 h,
P(H2) = 14 bar, batch
C = 100
S = 100
[11]
Cu-Ni
organosilica nanospheres
ST
WI
CP
LA, 2-PrOH, molLA/gcat = 0.031 mol g−1, 120 °C,
13 h, P(H2) = 40 bar, batch
C = 99
S = 97
[70]
Cu-Al2O3CPLA, EtOH, molLA/gcat = 0.0086 mol g−1, 110 °C, 2 h,
P(H2) = 30 bar, batch
C = 100
S = 95
[71]
Al2O3-Cu/ZrO2CPLA, H2O, molLA/gcat = 0.086 mol g−1, 200 °C, 2 h,
P(H2) = 30 bar, batch
C = 100
S = 100
[82]
Cu-Ni/γ-Al2O3WILA, solvent-free, molLA/gcat = 0.196 mol g−1, 220 °C, 6 h,
P(H2) = 30 bar, batch
C = 100
S = 99
[86]
Cu-Ni/Al2O3CILEs, n-HEX, molLE/gcat = 0.069 mol g−1, 180 °C, 6 h,
P(H2) = 25 bar, batch
C = 100
S = 98
[61]
B2O3-Cu/ZrO2CP for Cu/ZrO2, then boric acidLA, H2O, molLA/gcat = 0.069 mol g−1, 150 °C, 5 h,
P(H2) = 30 bar, batch
C = 100
S = 100
[30]
Flow
CatalystPreparation ProcedureReaction ConditionsCatalytic Performances (%)Ref.
Cu/γAl2O3WILA, H2O, 265 °C, P(H2) = 1 bar, H2 = 30 mL min−1,
WHSV = 0.169 h−1, H2/LA molar ratio = 201, flow
C = 98
S = 87
[74]
Cu/ZrO2 WILA, H2O, 265 °C, P(H2) = 1 bar, H2 = 30 mL min−1,
WHSV = 0.169 h−1, H2/LA molar ratio = 201, flow
C = 81
S = 83
[73]
Cu/Al2O3-SiO2CPLEs, EtOH, 140 °C,
1000 h, P(H2) = 30 bar, WHSV = 0.6 h−1,
H2/LE molar ratio = 50, flow
C = 79
S = 96
[72]
Cu-Ni/SiO2 WI with
citric acid
LA, solvent-free, 250 °C, P(H2) = 1 bar, H2 = 30 mL min−1, WHSV = 13.2 h−1, H2/LA molar ratio = 6, flowC = 98
S = 98
[68]
Cu-Ni/KIT6/
ZSM-5
WI LA, 250 °C, P(H2) = 1 bar, H2 = 30 mL min−1, time on stream = 24 h, H2/LA molar ratio = 0.0024, flowC = 100
S = 80
[75]
B2O3-Cu/ZrO2 CP for Cu/ZrO2, then boric acidLA, H2O, 200 °C, P(H2) = 30 bar, H2 = 40 mL min−1, H2/LA molar ratio = 2711, flowC = 100
S = 100
[30]

2.2. Catalytic Transfer Hydrogenation (CTH) of LA/LEs to GVL

The use of molecular hydrogen in high-pressure reactors poses several concerns in terms of sustainability, safety and applicability of the process. An alternative strategy is the catalytic transfer hydrogenation (CTH) that exploits an alcohol, both primary or secondary, or formic acid (FA) as hydrogen donor [78,97,98,99,100,101,102], reducing most of the problems related to the use of high-pressure molecular hydrogen (purchase, transport, safety hazard, expensive high-pressurized industrial plants) [103,104]. Moreover, several hydrogen-donor molecules can be obtained from renewable feedstock, such as lignocellulosic biomass, as in the case of bio-alcohols (obtained from biomass fermentation), and FA (that is a side product in LA/LE production). CTH was performed over Ni-, Zr- and Cu-based systems without the use of precious metal catalysts [105,106]. One of the drawbacks of the CTH process with alcohols is that the primary or secondary alcohol is transformed into the corresponding aldehyde or ketone, respectively, as a side product. Indeed, this may significantly affect the selectivity of the process, opening the path to aldol condensation side reactions [107,108,109].
The CTH reaction of LA/LEs with alcohols is reported in Scheme 3. First, the alcohol donates the hydrogen to LA/LEs, thus forming 4-hydroxypentanoic acid or its ester, and then GVL is formed by cyclization and ROH elimination.
Two mechanisms have been proposed for this reaction in the literature, depending on the catalyst used. When non-noble metals (e.g., Cu, Ni) are used, the reaction proceeds via a hydrogen-borrowing mechanism [94,110,111]. On the other hand, when ZrO2 is used, the reaction occurs without the evolution of molecular hydrogen through a concerted Meerwein–Ponndorf–Verley (MPV) mechanism, in which both the reducing alcohol and the carbonyl group are coordinated to the same metal center [91,98,112,113]. In the CTH process of LA/LEs in the presence of alcohols, the transesterification reaction is usually a competitive process. However, the resulting product can, in turn, undergo CTH, thus not usually hindering the GVL formation (Scheme 3).
Several heterogeneous Cu catalysts prepared on different supports, such as ZrO2, SiO2, TiO2, ZnO and γ-Al2O3, were proposed for this reaction. Monometallic and bimetallic (i.e., Cu in combination with Ni) materials have been designed by using different preparation methods, such as wet impregnation (WI), deposition–precipitation (DP) or oxalate gel coprecipitation method (OG), and tested, providing high conversion and selectivity to GVL under mild reaction conditions [78,110]. Among these, the best results were obtained when ZrO2 was employed, thus confirming the importance of Lewis acid-base pairs on the catalyst surface for the CTH reaction. An attempt to improve the catalytic system using bifunctional catalysts with aluminum oxide and hydrogenation sites (Cu-Ni alloy) supported on active carbon (AC) was made by Yu et al. [114]. An adequate Al/Cu-Ni ratio of 1, resulting in the best acidity of the catalyst, was found to promote the esterification reaction step, and thus increase the GVL yield (C = 100% and S = 97%, respectively).
The use of a high surface area support, such as SBA-15, provided high metallic dispersion and a limited diffusion of the metal NPs in the SBA channels, and repressed agglomeration as well as leaching [83]. The introduction of Ni provided benefits also for the CTH reaction, by remarkably enhancing the catalytic activity and stability to poisoning and sintering, thus allowing a better catalyst recyclability [94].
The literature reports the notable work of Yuan et al. [78], in which FA, which is co-produced in an equimolar amount along with LA during the biomass dehydration process [115,116], was employed as an alternative in situ hydrogen source via its selective decomposition. The CTH reaction proceeded in stream and in water on Cu/ZrO2, providing 100% of C and S, and is an interesting example of an integrated process. In Table 2, the catalytic behavior of the supported Cu or Cu-Ni catalysts for the hydrocyclization of LA/LEs to GVL under CTH conditions is summarized.
It is worth noting that bare ZrO2 is a very active and selective material for CTH processes, including the transformation of LA/LEs to GVL [91,97]. It is our opinion that although the addition of Cu or other metals deserves to be explored to optimize the catalytic performance, the bare ZrO2 behaves as an effective system by itself [10,67,104,106,117,118,119,120,121].
Table 3 highlights the main milestones in the comprehension of supported Cu and Cu-bimetallic catalysts for the hydrocyclization of LA/LEs to GVL under molecular H2 or under CTH conditions. Metallic Cu was found to be the active phase, and highly dispersed Cu NPs are crucial to provide an active catalytic surface. The preparation method, together with high surface area and/or 2D–3D supports, are key factors guaranteeing a highly dispersed metal phase. Despite supported Cu catalysts being highly efficient and selective, both for conventional hydrogenation using molecular H2 and for CTH, they suffer some drawbacks, such as Cu leaching, especially when LA instead of LEs is used, or when the reaction occurs in water. As many reactions work at temperatures higher than the Cu Hüttig one, sintering occurs and catalysts deactivate. To overcome this, the use of an additive, such as B2O3 or Al2O3 during catalyst preparation, helps prevent sintering. Moreover, the addition of a second metal, namely Ni or Ag, improves the catalytic lifetime because of synergic effects between the two metals: Ni or Ag prevent metal leaching, while Cu prevents coking and facilitates Ni or Ag reduction. Lewis or Brønsted acid-base pairs are crucial to the promotion of the dehydration step and thus the cyclization to GVL.

2.3. Transformation of LA/LEs and GVL to VE

Valeric esters (VE) have been identified as a new class of cellulosic-derived potential biofuels, suitable both for gasoline and diesel engines, as they provide fully compatible components with transportation fuels [39,44]. VE, depending on the alcoholic residue, have shown better fuel blending properties than GVL because of energy density, appropriate polarity and volatility–ignition properties [24,53,122].
The reaction occurs under H2 either from LA/LEs or GVL through nucleophilic addition of the alcohol to the carboxylic group, followed by dehydration to pentenoate and hydrogenation to valeric ester (Scheme 4).
Lange et al. [24] were pioneers in developing a process to convert GVL into valerates, in continuous high-pressure tubular reactor units, using a Pt/ZSM-5 bound with a SiO2 catalyst with balanced acidic and hydrogenation properties, or Pt supported on ZrO2 or TiO2. Although Pt-based catalysts tend to convert GVL, providing a high PV/VA ratio, they produce a large amount of undesired light hydrocarbons and are expensive, as already discussed above. Some investigations [123] were reported of GVL conversion or one-pot conversion of LA over Ni- [124], Ru- [125,126], Re-Ru- [127], Co- [128], Pd- [129], Rh- [130] or Pt-based catalysts [131,132], whereas Cu-supported catalysts, although advantageous, are still scarcely studied.
In an attempt to switch from noble to non-noble metal-based catalysts, and to develop a bifunctional system for a one-pot cascade process, some of us [122] successfully combined the acidic properties of several solid supports with the hydrogenation activity of pre-reduced CuO catalysts prepared by CH for the one-pot ring opening and hydrogenation of GVL to valerate esters in a batch reactor. The best performances were obtained with Cu/SiO2-ZrO2 (4.7 wt.% ZrO2) in the presence of ethanol or 1-pentanol to form ethyl valerate (C = 69%; S = 59%) or pentyl pentanoate (C = 90%; S = 83%). The significant hydrogenation activity of finely dispersed supported Cu, together with relevant Lewis acidity exhibited by very small Cu particles, allowed us to design a new Cu/SiO2 bifunctional catalyst [92], without the need of an acidic support, converting 91% of GVL and improving the selectivity to 1-pentyl valerate up to 92%.
Liu et al. [93] claimed a novel separate nucleation and aging steps assistant reduction–oxidation strategy, developed to synthesize CuO/ZrO2 with homogeneously distributed Cu and Zr components, which were reduced to obtain highly dispersed Cu/ZrO2 catalysts and used for the transformation of GVL to valerate esters under H2 pressure, exhibiting excellent catalytic performances with several alcohols (e.g., with 1-pentanol C = 85%, S = 98%).

3. Hydrogenation of LA/LEs and GVL to Give 1,4-PDO

As shown in the previous section, LA/LEs can be converted into GVL through a combination of hydrogenation and dehydration reactions. However, GVL can be further hydrogenated to produce other interesting and useful compounds, such as 1,4-PDO and 2-MTHF (Scheme 5). Biomass-derived diols can be utilized not only as bio-monomers to synthetize bioderived polyesters [133,134] and polyurethanes [135,136] but also as solvents [137,138]. The reaction pathways of LA/LEs to give GVL and 1,4-PDO diverts depending on dehydration or hydrogenation sequences, as already reported. In particular, it involves the formation of the 4-HPA intermediate if the hydrogenation occurs first; on the other hand, if dehydration is the first step, angelica lactone (AL) is formed instead. GVL can be subsequently hydrogenated to 1,4-PDO, which can be further dehydrated to give 2-MTHF [139].
The Resasco group has given a possible reaction pathway and intermediate formation thanks to a combination of experimental and theoretical studies, considering, in particular, the hydrogenation of GVL in aqueous phase on 5% Ru/C catalysts [140]. The reaction may start with a ring-opening step breaking the bond between the carbonylic carbon and the oxygen, forming the surface intermediate CH3CH(O)–(CH2)2–CO. This species can undergo three reversible parallel pathways: (i) decarbonylation to 2-butanol, (ii) hydrogenation to 1,4-PDO and (iii) C-O hydrogenolysis to 2-pentanol [140].
Successively, the same research group had further investigated the reaction profile to shed light on the rate-limiting step. In particular, DFT studies had suggested that the rate-limiting step for the formation of the 1,4-PDO is the hydrogenation one, underlining the need to design hydrogenation-promoting catalysts [141]. Actually, GVL hydrogenation can also result in the formation of valeric acid (VA) or its esters [142].
The fine-tuning of the acid/base and hydrogenation properties of the catalysts allows one to move toward the desired product through a different reaction pathway.
Scheme 5. LA hydrogenation pathways to give GVL and/or 1,4-PDO. Adapted from ref. [65,143].
Scheme 5. LA hydrogenation pathways to give GVL and/or 1,4-PDO. Adapted from ref. [65,143].
Catalysts 13 00697 sch005
There are several papers in the literature relying on noble metal catalysts to produce 1,4-PDO [47,48,144]. However, some examples of supported Cu catalysts that are effective in the synthesis of 1,4-PDO from LA/LEs could be found. In these cases, the fine-tuning of acidic and basic sites becomes of crucial importance to modulating the selectivity, having a role in the GVL ring-opening process, has already been highlighted. In particular, basic sites could boost the selectivity towards the diol, enabling the GVL ring opening. The introduction of basic Mg species pushed the selective conversion of GVL to 1,4-PDO, with a yield up to 98% at 140 °C and 60 bar of hydrogen [71]. On the contrary, a high Brønsted acidity, similar to the one found over the CuAl catalyst, suppresses the ring-opening reaction, inhibiting 1,4-PDO formation and thus improving the GVL yield [71]. These results were further confirmed by the same group in a study in which the hydrogenation of ethyl levulinate to give 1,4-PDO with a 99% yield was favored on a CuMg catalyst with abundant basic sites [145].
Due to the complexity of the whole process, the production of 1,4-PDO can be advantageously separated into two different steps: the first one is the synthesis of GVL from LA/LEs, and the second is the transformation of GVL into the diol. In particular, GVL is more prone to be hydrogenated using non-noble metal catalysts, as shown below in this paragraph (see Table 4). Moreover, GVL can be easily obtained from LA/LEs using non-noble metal catalysts, as shown in the previous section.
Most recently, supported copper-based catalysts were shown to be effective in the selective hydrogenation of GVL to give 1,4-PDO. Copper can be supported on a wide fan of metal oxide supports with different properties. A Cu/MgO prepared with a metal–organic chemical vapor deposition method achieved a C = 91% and S = 94% at 100 bar of H2 for 10 h [146]. The loading of copper was found to be crucial for the catalytic activity and selectivity towards 1,4-PDO. Moving from 6 to 18 copper wt.% loading, conversion increased from 71% to 80%, reaching the maximum selectivity of 86%. On the other hand, when the copper loading was further increased, a rise in selectivity for 2-MTHF and 1-pentanol was observed [146].
As already cited, it is important to underline the major role of acid/basic sites when considering this hydrogenation reaction. In fact, the major side reaction in the hydrogenation of GVL to 1,4-PDO is the dehydration of the diol to give 2-MTHF, which is a typical acid catalyzed reaction. Thus, it is likely to work with bifunctional basic catalysts or to partially suppress their acidity. For instance, the basic Cu1.5/Mg1.5AlO catalyst derived from layered double hydroxides precursors was found to be particularly efficient, achieving a GVL conversion of 93% and a 1,4-PDO selectivity >99% [147]. These results were mainly ascribed to the cooperative effect between its well-dispersed active Cu nanoparticles and the proper surface basic sites nearby. Notably, when the acidic Cu/HZSM-5 is used, larger amounts of 2-MTHF were obtained due to the acid-assisted diol dehydration. This was observed also when 1,4-PDO was used as the substrate: a C = 69% was found over the Cu/HZSM-5 catalyst when, in contrast, no 1,4-PDO conversion resulted using the basic Cu1.5/Mg1.5AlO catalyst [147]. On the other hand, the addition of Zn on the Cu/Al2O3 catalyst significantly improved the selectivity towards the diol, passing from 46% obtained with the bare Cu/Al2O3 to 98% in the case of the ZnCu/Al2O3 catalyst, with a Zn/Cu molar ratio of 1. In fact, Zn reduced the quantity of surface Lewis acid sites and promoted the formation of 1,4-PDO with respect to 2-MTHF [148].
However, it was demonstrated that high yields of 1,4-PDO can be obtained even using mild acidic supports such as SiO2, both under flow conditions [150] and in batch [23]. Some of us were able to achieve a GVL conversion of 78% with a 1,4-PDO selectivity of 98% using Cu/SiO2 prepared with the CH on a pyrogenic silica in the green solvent cyclopenthyl methyl ether (T = 160 °C, P(H2) = 50 bar, t = 22 h) [23]. It was found that the right acid strength is crucial in the activation of the carbonyl group of the lactone, while the formation of 2-MTHF can be avoided, limiting the surface hydrophilicity of the catalyst. In fact, the proper combination of solvent and catalyst hydrophobicity/hydrophilicity allows one to improve the 1,4-PDO yield. In particular, when cyclopentylmethylether and a less hydrophilic SiO2 are used, the yield in the diol rises from 46% to 78%. It was proposed that a less hydrophilic surface facilitates the desorption of the diol, thus preventing side reactions. On the other hand, the solvent affects the acidic strength [23]. The catalytic system was further optimized by decreasing the hydrophilicity of a Cu/SiO2 prepared with a silica gel carefully functionalizing the surface with trietoxyoctylsilane, achieving 99% of selectivity of 1,4-PDO, even at 1% loading of the silane under the same conditions as the previous work [149]. The best yield of 80% was achieved with a silane loading of 10%. However, a higher silane loading hinders the GVL from reaching the catalytic active Cu sites on the surface, decreasing the conversion despite keeping the 1,4-PDO selectivity at 99%. This phenomenon was caused by the tendency of the organosilane to reticulate on the surface at higher concentration [149].
It is worth noting that there are also few examples relying on heterogeneous copper catalysts for the production of 1,4-PDO from GVL through flow processes. This is the case with a Cu/ZnO prepared with a co-precipitation method using different copper-loading and calcination temperatures. The 40% Cu/ZnO calcined at 500 °C was the best catalyst and gave a C = 82% and S = 99%, at a GVL feed rate of 0.4 g h−1, 140 °C, H2 pressure of 15 bar and H2 flow rate of 90 mL min−1 [143]. In addition, in this case, the selectivity towards 1,4-PDO increased with the H2 pressure, while the increase in temperature favored the formation of 2-MTHF [143]. These trends were observed also in the work of Simakova et al., where Cu/SiO2 was used [150], obtaining a C = 32% and S = 67% in a continuous flow reactor using 1-butanol as the solvent (10% GVL in 1-butanol, GVL flow rate 2.1 g h−1 cat = 0.455 g, T = 130 °C, P(H2) = 13 bar, H2 flow rate 167 mL min−1, 4 h). Once again, the lower selectivity towards 1,4-PDO compared to that achieved with Cu/ZnO was ascribed to the acidity of the silanol groups, which promotes the dehydration reaction to form 2-MTHF [150].
The analysis of the literature confirms the necessity of having a high hydrogenating system to reach high 1,4-PDO yields, by choosing an appropriate catalyst and/or by increasing the hydrogen pressure. Moreover, strategies to prevent 1,4-PDO dehydration to 2-MTHF should be pursued through an accurate catalyst design. This can be achieved by properly working on the catalyst surface, with the introduction of basic sites and/or with the suppression of the acidic ones, or by playing with the catalyst wettability to promote the diol desorption, thus preventing side reactions. It is worth noting that the wettability is often an underrated parameter in catalyst design [151,152,153].

4. Hydrogenation of LA/LEs and GVL to Give 2-MTHF

2-MTHF is a very important product of the LA stream due to its role in bioderived gasoline blends. In particular, it represents up to 20% of the “P-series fuels”, a family of renewable, non-petroleum liquid fuels, containing about 35% liquid by-products, known as “C5+” or “pentanes-plus”, and 45% ethanol, fermented from corn [154]. Moreover, it is considered a sustainable solvent, with a promising environmental footprint (not only biobased, but also easy to degrade) and low toxicity [66,155].
As already pointed out, to produce 2-MTHF from LA/LEs, GVL is initially formed (Scheme 5) and therefore hydrogenated to form 2-hydroxy-5-methyltetrahydrofuran as intermediate. From here, two main routes can take place. The first one involves the formation of 1,4-PDO (via a ring-opening and hydrogenation reaction), followed by intramolecular etherification, as already explained. The second pathway proceeds by dehydration to form 2,3-dihydro-2-methylfuran that, in turn, is hydrogenated to 2-MTHF [65,143].
The literature suggests that the reaction preferentially occurs via the ring opening [84,156,157]. It is worth noting that noble metal-based catalysts also favor the C-O cleavage at the methyl-group side with the formation of valeric acid that can be converted into 1-pentanol, as highlighted by Sun et al. [65,143]. Based on these considerations, the use of non-noble metals, such as copper, for the preparation of 2-MTHF is even more advantageous because they combine low cost and higher availability with superior selectivity, compared to noble-ones, as they exclude the formation of 1-pentanol via valeric acid.
Activity and selectivity are strongly related to both reaction conditions and catalyst features. In particular, H2 pressure and its availability are critical parameters to obtaining good yields in 2-MTHF, while acid/base properties and stability are the main points to look at for catalyst optimization. Based on the previous considerations, we know that basicity favors the selective formation of 1,4-PDO while acidity pushes towards 2-MTHF.
In Table 5, the main results of 2-MTHF synthesis from LA/LEs and GVL obtained with copper catalysts are reported.
The conversion of LA/LEs into 2-MTHF is usually obtained with high-loading copper catalysts. A Cu/SiO2 system with 80 wt.% of Cu was found to be selective, allowing one to obtain a 64% yield under 25 bar of H2 at 265 °C [161]. Interestingly, the results reported show that the selectivity of the reaction can be tuned by changing the Cu loading: low loading favors the selective hydrocyclization of LA to GVL, which, in turn, undergoes a further hydrogenation to 2-MTHF and 1-pentanol in the presence of high-loading catalysts. Other authors proposed the use of a Cu/SiO2 system doped with Al2O3, obtaining a 65% yield in 2-MTHF starting from LEs [72]. The authors underline the important role of alumina in decreasing the particle size of CuO clusters in the co-precipitation process. Smaller Cu particles after reduction and hydrogen spillover from Cu to Lewis acidic sites may account for the much-improved activity of the Cu/Al2O3-SiO2 catalyst. The CuO particles distribute around 20–60 nm in the spent and re-calcined catalyst due to sintering inhibition by anchoring of Cu particles on the Lewis acidic sites and much lower reaction temperature. The stronger Lewis acidic sites also accelerate the dehydration reaction to form GVL as the main product, suppressing the formation of 1,4-PDO and 2-pentanol. The great advantage obtained in the use of a bimetallic catalyst starting from LA/LEs is also claimed in some papers relying on the hydrogenation promoted by Cu-Ni systems. A Cu-Ni/Al2O3 obtained 98% of 2-MTHF at 180 °C and 40 bar of H2 [159]. The excellent catalytic activity observed was ascribed by the authors to both Ni and Cu compositions, where Ni was more readily able to activate the H2 and hydrogenate EL into GVL, while Cu catalyzed the ring opening of GVL to yield the target molecule 2-MTHF. Moreover, the use of a non-polar solvent, such as n-hexane, significantly reduced the apparent activation energy for the ring opening of GVL. The synergistic effect of Ni and Cu has also been highlighted in another catalytic system, namely a Cu-Ni/ Al2O3-ZrO2 [158], which provided a 100% yield in 2-MTHF starting from LA at 210 °C and 35 bar of H2, mainly due to the tunable acidic properties of the support.
Some important differences arise when starting from GVL. Thus, while the effect of increased acidity still remains a strong point in selectively obtaining 2-MTHF, the particle size parameter shows a different effect.
Sun et al. compared Co, Ni and Cu catalysts in the vapor phase hydrogenation of GVL, observing a marked higher activity of Cu with respect to Ni and Co, particularly in terms of selectivity. Thus, a commercial Cu/Al2O3 obtained more than 93% selectivity in 2-MTHF vs. the 75 and 44%, respectively, obtained with Co/Al2O3 and Ni/Al2O3 [65]. The optimization of the catalyst and of the reaction protocol led to excellent catalytic activity, which means a 2-MTHF selectivity of 99% and a GVL conversion of 98%. Besides acidity, the authors highlighted the importance of particle size as one of the main parameters to work on to maximize 2-MTHF yield. In particular, by increasing the calcination temperature up to 700 °C, the particle size increases with a coherent higher selectivity in 2-MTHF with respect to 1,4-PDO. A further increase in the calcination temperature leads, on the contrary, to an inversion in product distribution due to the formation of a CuAl2O4 phase, which reduces the amount of acid sites. This, once again, underlines the effect of acidity/basicity in addressing the formation of 1,4-PDO or 2-MTHF. The optimization of the reaction conditions and of the catalyst textural properties also allowed the obtaining of up to 81% of 2-MTHF at 300 °C under flow conditions with a Cu/SiO2-impregnated system [157].
Different catalyst properties and reaction conditions for each step would be necessary for these three steps, which would increase the difficulty of the direct production of 2-MTHF from LA. In contrast to the production of 1,4-PDO from LA, which needs precious metal catalysts and high H2 pressures, the production of 1,4-PDO from GVL can be efficiently catalyzed by Cu-based catalysts under mild reaction conditions [26].
Also, when starting from GVL, some bimetallic systems have been proposed, once again Ni-Cu catalysts. Eighty-eight percent selectivity in 2-MTHF resulted from the use of a Ni2Cu1/Al2O3 derived from layered double hydroxide (LDH) precursors [160], while the strong beneficial effect of Ni in limiting sintering and leaching of copper was underlined in using a Ni-Cu/SiO2 system [161].

5. Conclusions and Future Directions

5.1. General Strategies to Selectively Obtain the Desired Product

Copper-based materials are very promising catalysts for the cascade transformation of LA/LEs in the presence of hydrogen. By tuning the experimental conditions (i.e., H2 pressure, temperature) and the catalyst properties, it is possible to move from one product to another (Figure 2). All these products (GVL, valerates, 1,4-PDO, and 2-MTHF) are obtained through consecutive reaction steps, starting from LA/LEs. A good catalyst is a material with high hydrogenation activity and, depending on the target product, with different acid-base properties. From the analysis of the literature, to draw a clear picture is not an easy task, but some general considerations can be summarized:
-
GVL and valerates are preferentially obtained when a catalyst with a pronounced acidic character is used. Clearly, the presence of an alcoholic medium is required for the preparation of valerates.
-
By tuning the acid/base properties, the reaction is more prone to proceed through the ring opening, leading to the formation of 1,4-PDO [71]. 1,4-PDO dehydration can be avoided by limiting catalyst acidity or polarity. High H2 pressure is usually needed to reach high yields.
-
The yield in 2-MTHF can be optimized by increasing the acidity of the catalyst and the copper particle size.

5.2. Catalyst Deactivation and Stability Considerations

The use of a non-noble, non-toxic metal, such as copper, is a clear advantage for the valorization of LA/LEs, and the literature shows some nice examples in which high conversions and high selectivity are reported. Despite this, catalyst deactivation represents one of the major obstacles to industrial scale-up (Figure 3). Different approaches have been enacted to improve stability, but there is still much to do. Leaching can be limited by using LEs instead of LA and by selecting the most suitable solvent. Water, which is considered a green solvent, favors this phenomenon, but its replacement with an organic solvent could entail sustainability problems. At the same time, leaching can decrease by controlling copper–support or copper–metal (in the case of a bimetallic system) interactions. These strategies are also effective in contrasting sintering, along with the use of high surface area supports, or of materials with a particular 2D/3D structure. Despite its toxicity, the addition of nickel appears to be a good strategy to facilitate H2 activation and to limit sintering and leaching of copper. It is important to note that the reactions of hydrogenation of LA/LEs usually require temperatures over 200 °C, while copper Hüttig temperature is only 135 °C. It follows that the research should also focus on developing low-temperature processes to improve the catalyst lifetime. Unlike other reactions, the deactivation of copper-based systems due to carbon deposition in LA/LE hydrogenation seems to be less relevant.

5.3. Future Directions

It is our opinion that future papers should put more stress on aspects related to catalyst stability and recyclability, as well as implementation of continuous flow processes. Moreover, there is still a gap in research concerning one-pot, cascade processes that produce 1,4-PDO or 2-MTHF starting from LA/LEs using copper-based catalysts.

Author Contributions

Conceptualization, N.S.; writing—original draft preparation, F.Z., D.C., L.A. and N.S.; writing—review and editing, N.S., F.Z. and N.R. All authors have read and agreed to the published version of the manuscript.

Funding

The authors are grateful to Italian Ministero dell’Università e della Ricerca for the financial support provided through the PRIN 2020 LEVANTE project “LEvulinic acid Valorization through Advanced Novel Technologies” (Progetti di Ricerca di Rilevante Interesse Nazionale-Bando 2020, 2020CZCJN7).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

List of Abbreviations

LA = levulinic acid; LEs = levulinic esters; GVL = γ-valerolactone; FA = formic acid; VA = valeric acid; VE = valeric esters; 1,4-PDO = 1,4 pentanediol; 2-MTHF = 2-methyl tetrahydrofuran; OG = oxalate gel; CP = co-precipitation; WI = wet impregnation; CI = co-impregnation; SG = sol-gel; DP = deposition–precipitation; MT = methanothermal; ST = solvothermal method; CVD = chemical vapor deposition; CH = chemisorption hydrolysis; AC = activated carbon.

References

  1. Pileidis, F.D.; Titirici, M.-M. Levulinic Acid Biorefineries: New Challenges for Efficient Utilization of Biomass. ChemSusChem 2016, 9, 562–582. [Google Scholar] [CrossRef]
  2. Morone, A.; Apte, M.; Pandey, R.A. Levulinic Acid Production from Renewable Waste Resources: Bottlenecks, Potential Remedies, Advancements and Applications. Renew. Sustain. Energy Rev. 2015, 51, 548–565. [Google Scholar] [CrossRef]
  3. De Vries, J.G. Industrial Implementation of Chemical Biomass Conversion. Curr. Opin. Green Sustain. Chem. 2023, 39, 100715. [Google Scholar] [CrossRef]
  4. Yang, Z.; Huang, Y.-B.; Guo, Q.-X.; Fu, Y. RANEY® Ni Catalyzed Transfer Hydrogenation of Levulinate Esters to γ-Valerolactone at Room Temperature. Chem. Commun. 2013, 49, 5328–5330. [Google Scholar] [CrossRef]
  5. Hengne, A.M.; Rode, C.V. Cu-ZrO2 Nanocomposite Catalyst for Selective Hydrogenation of Levulinic Acid and Its Ester to γ-Valerolactone. Green Chem. 2012, 14, 1064–1072. [Google Scholar] [CrossRef]
  6. Di Menno Di Bucchianico, D.; Wang, Y.; Buvat, J.-C.; Pan, Y.; Casson Moreno, V.; Leveneur, S. Production of Levulinic Acid and Alkyl Levulinates: A Process Insight. Green Chem. 2022, 24, 614–646. [Google Scholar] [CrossRef]
  7. Ahmad, E.; Alam, M.D.I.; Pant, K.K.; Haider, M.A. Catalytic and Mechanistic Insights into the Production of Ethyl Levulinate from Biorenewable Feedstocks. Green Chem. 2016, 18, 4804–4823. [Google Scholar] [CrossRef]
  8. Lange, J.-P.; van de Graaf, W.D.; Haan, R.J. Conversion of Furfuryl Alcohol into Ethyl Levulinate Using Solid Acid Catalysts. ChemSusChem 2009, 2, 437–441. [Google Scholar] [CrossRef]
  9. Tiwari, M.S.; Gawade, A.B.; Yadav, G.D. Magnetically Separable Sulfated Zirconia as Highly Active Acidic Catalysts for Selective Synthesis of Ethyl Levulinate from Furfuryl Alcohol. Green Chem. 2017, 19, 963–976. [Google Scholar] [CrossRef]
  10. Bui, L.; Luo, H.; Gunther, W.R.; Román-Leshkov, Y. Domino Reaction Catalyzed by Zeolites with Brønsted and Lewis Acid Sites for the Production of γ-Valerolactone from Furfural. Angew. Chem. Int. Ed. 2013, 52, 8022–8025. [Google Scholar] [CrossRef]
  11. Zhang, L.; Mao, J.; Li, S.; Yin, J.; Sun, X.; Guo, X.; Song, C.; Zhou, J. Hydrogenation of Levulinic Acid into Gamma-Valerolactone over in Situ Reduced CuAg Bimetallic Catalyst: Strategy and Mechanism of Preventing Cu Leaching. Appl. Catal. B 2018, 232, 1–10. [Google Scholar] [CrossRef]
  12. Démolis, A.; Essayem, N.; Rataboul, F. Synthesis and Applications of Alkyl Levulinates. ACS Sustain. Chem. Eng. 2014, 2, 1338–1352. [Google Scholar] [CrossRef]
  13. Xue, Z.; Liu, Q.; Wang, J.; Mu, T. Valorization of Levulinic Acid over Non-Noble Metal Catalysts: Challenges and Opportunities. Green Chem. 2018, 20, 4391–4408. [Google Scholar] [CrossRef]
  14. Antonetti, C.; Licursi, D.; Fulignati, S.; Valentini, G.; Raspolli Galletti, A.M. New Frontiers in the Catalytic Synthesis of Levulinic Acid: From Sugars to Raw and Waste Biomass as Starting Feedstock. Catalysts 2016, 6, 196. [Google Scholar] [CrossRef]
  15. Melchiorre, M.; Lentini, D.; Cucciolito, M.E.; Taddeo, F.; Hmoudah, M.; di Serio, M.; Ruffo, F.; Russo, V.; Esposito, R. Sustainable Ketalization of Glycerol with Ethyl Levulinate Catalyzed by the Iron (III)-Based Metal-Organic Framework MIL-88A. Molecules 2022, 27, 7229. [Google Scholar] [CrossRef]
  16. Zaccheria, F.; Ravasio, N.; Ercoli, M.; Allegrini, P. Heterogeneous Cu-Catalysts for the Reductive Deoxygenation of Aromatic Ketones without Additives. Tetrahedron Lett. 2005, 46, 7743–7745. [Google Scholar] [CrossRef]
  17. Mondal, U.; Yadav, G.D. Methanol Economy and Net Zero Emissions: Critical Analysis of Catalytic Processes, Reactors and Technologies. Green Chem. 2021, 23, 8361–8405. [Google Scholar] [CrossRef]
  18. Zhang, R.; Chen, Y.; Ding, M.; Zhao, J. Heterogeneous Cu Catalyst in Organic Transformations. Nano Res. 2022, 15, 2810–2833. [Google Scholar] [CrossRef]
  19. Fang, Y.; Guo, Y. Copper-Based Non-Precious Metal Heterogeneous Catalysts for Environmental Remediation. Chin. J. Catal. 2018, 39, 566–582. [Google Scholar] [CrossRef]
  20. Zaccheria, F.; Ravasio, N.; Chan-Thaw, C.E.; Scotti, N.; Bondioli, P. A Bifunctional Copper Catalyst for the One Pot-One Step Esterification + Hydrogenation of Tall Oil Fatty Acids. Top. Catal. 2012, 55, 631–636. [Google Scholar] [CrossRef]
  21. Zaccheria, F.; Shaikh, N.I.; Scotti, N.; Psaro, R.; Ravasio, N. New Concepts in Solid Acid Catalysis: Some Opportunities Offered by Dispersed Copper Oxide. Top. Catal. 2014, 57, 1085–1093. [Google Scholar] [CrossRef]
  22. Zaccheria, F.; Mariani, M.; Scotti, N.; Psaro, R.; Ravasio, N. Catalytic Upgrading of Lactose: A Rest Raw Material from the Dairy Industry. Green Chem. 2017, 19, 1904–1910. [Google Scholar] [CrossRef]
  23. Cavuoto, D.; Ravasio, N.; Scotti, N.; Gervasini, A.; Campisi, S.; Marelli, M.; Cappelletti, G.; Zaccheria, F. A Green Solvent Diverts the Hydrogenation of γ–Valerolactone to 1,4—Pentandiol over Cu/SiO2. Mol. Catal. 2021, 516, 111936. [Google Scholar] [CrossRef]
  24. Lange, J.-P.; Price, R.; Ayoub, P.M.; Louis, J.; Petrus, L.; Clarke, L.; Gosselink, H. Valeric Biofuels: A Platform of Cellulosic Transportation Fuels. Angew. Chem. Int. Ed. 2010, 49, 4479–4483. [Google Scholar] [CrossRef] [PubMed]
  25. Qi, L.; Mui, Y.F.; Lo, S.W.; Lui, M.Y.; Akien, G.R.; Horváth, I.T. Catalytic Conversion of Fructose, Glucose, and Sucrose to 5-(Hydroxymethyl)Furfural and Levulinic and Formic Acids in γ-Valerolactone as a Green Solvent. ACS Catal. 2014, 4, 1470–1477. [Google Scholar] [CrossRef]
  26. Du, X.-L.; Bi, Q.-Y.; Liu, Y.-M.; Cao, Y.; He, H.-Y.; Fan, K.-N. Tunable Copper-Catalyzed Chemoselective Hydrogenolysis of Biomass-Derived γ-Valerolactone into 1,4-Pentanediol or 2-Methyltetrahydrofuran. Green Chem. 2012, 14, 935–939. [Google Scholar] [CrossRef]
  27. Lange, J.-P.; Vestering, J.Z.; Haan, R.J. Towards ‘Bio-Based’ Nylon: Conversion of γ-Valerolactone to Methyl Pentenoate under Catalytic Distillation Conditions. Chem. Commun. 2007, 33, 3488–3490. [Google Scholar] [CrossRef] [PubMed]
  28. Weng, R.; Yu, Z.; Xiong, J.; Lu, X. Effects of water in the heterogeneous catalytic valorization of levulinic acid into γ-valerolactone and its derivatives. Green Chem. 2020, 22, 3013–3027. [Google Scholar] [CrossRef]
  29. Dai, Y.; Lu, P.; Cao, Z.; Campbell, C.T.; Xia, Y. The Physical Chemistry and Materials Science behind Sinter-Resistant Catalysts. Chem. Soc. Rev. 2018, 47, 4314–4331. [Google Scholar] [CrossRef]
  30. Li, J.-F.; Zhao, L.; Li, J.; Li, M.; Liu, C.-L.; Yang, R.-Z.; Dong, W.-S. Highly Selective Synthesis of γ-Valerolactone from Levulinic Acid at Mild Conditions Catalyzed by Boron Oxide Doped Cu/ZrO2 Catalysts. Appl. Catal. A Gen. 2019, 587, 117244. [Google Scholar] [CrossRef]
  31. Adkins, H.; Folkers, K. The Catalytic Hydrogenation of Esters to Alcohols. J. Am. Chem. Soc. 1931, 53, 1095–1097. [Google Scholar] [CrossRef]
  32. Folkers, K.; Adkins, H. The Catalytic Hydrogenation of Esters to Alcohols. II. J. Am. Chem. Soc. 1932, 54, 1145–1154. [Google Scholar] [CrossRef]
  33. Christian, R.V., Jr.; Brown, H.D.; Hixon, R.M. Derivatives of γ-Valerolactone, 1,4-Pentanediol and 1,4-Di-(β-Cyanoethoxy)-Pentane1. J. Am. Chem. Soc. 1947, 69, 1961–1963. [Google Scholar] [CrossRef]
  34. Bonrath, W.; Castelijns, A.M.C.F.; de Vries, J.G.; Guit, R.P.M.; Schütz, J.; Sereinig, N.; Vaessen, H.W.L.M. Gas Phase Hydrogenation of Levulinic Acid to γ-Valerolactone. Catal. Lett. 2016, 146, 28–34. [Google Scholar] [CrossRef]
  35. Li, Z.; Zuo, M.; Jiang, Y.; Tang, X.; Zeng, X.; Sun, Y.; Lei, T.; Lin, L. Stable and Efficient CuCr Catalyst for the Solvent-Free Hydrogenation of Biomass Derived Ethyl Levulinate to γ-Valerolactone as Potential Biofuel Candidate. Fuel 2016, 175, 232–239. [Google Scholar] [CrossRef]
  36. Karam, L.; Neumann, C.N. Heterogeneously Catalyzed Carboxylic Acid Hydrogenation to Alcohols. ChemCatChem 2022, 14, e202200953. [Google Scholar] [CrossRef]
  37. Bond, J.Q.; Martin Alonso, D.; West, R.M.; Dumesic, J.A. γ-Valerolactone Ring-Opening and Decarboxylation over SiO2/Al2O3 in the Presence of Water. Langmuir 2010, 26, 16291–16298. [Google Scholar] [CrossRef] [PubMed]
  38. Serrano-Ruiz, J.C.; Dumesic, J.A. Catalytic Routes for the Conversion of Biomass into Liquid Hydrocarbon Transportation Fuels. Energy Environ. Sci. 2011, 4, 83–99. [Google Scholar] [CrossRef]
  39. Yan, K.; Yang, Y.; Chai, J.; Lu, Y. Catalytic Reactions of Gamma-Valerolactone: A Platform to Fuels and Value-Added Chemicals. Appl. Catal. B 2015, 179, 292–304. [Google Scholar] [CrossRef]
  40. Horváth, I.T. Solvents from Nature. Green Chem. 2008, 10, 1024–1028. [Google Scholar] [CrossRef]
  41. Yan, Z.; Lin, L.; Liu, S. Synthesis of γ-Valerolactone by Hydrogenation of Biomass-Derived Levulinic Acid over Ru/C Catalyst. Energy Fuels 2009, 23, 3853–3858. [Google Scholar] [CrossRef]
  42. Bond, J.Q.; Alonso, D.M.; Wang, D.; West, R.M.; Dumesic, J.A. Integrated Catalytic Conversion of γ-Valerolactone to Liquid Alkenes for Transportation Fuels. Science (1979) 2010, 327, 1110–1114. [Google Scholar] [CrossRef]
  43. Fábos, V.; Lui, M.Y.; Mui, Y.F.; Wong, Y.Y.; Mika, L.T.; Qi, L.; Cséfalvay, E.; Kovács, V.; Szűcs, T.; Horváth, I.T. Use of Gamma-Valerolactone as an Illuminating Liquid and Lighter Fluid. ACS Sustain. Chem. Eng. 2015, 3, 1899–1904. [Google Scholar] [CrossRef] [Green Version]
  44. Kon, K.; Onodera, W.; Shimizu, K. Selective Hydrogenation of Levulinic Acid to Valeric Acid and Valeric Biofuels by a Pt/HMFI Catalyst. Catal. Sci. Technol. 2014, 4, 3227–3234. [Google Scholar] [CrossRef]
  45. Horváth, I.T.; Mehdi, H.; Fábos, V.; Boda, L.; Mika, L.T. γ-Valerolactone—A Sustainable Liquid for Energy and Carbon-Based Chemicals. Green Chem. 2008, 10, 238–242. [Google Scholar] [CrossRef]
  46. Al-Shaal, M.G.; Dzierbinski, A.; Palkovits, R. Solvent-Free γ-Valerolactone Hydrogenation to 2-Methyltetrahydrofuran Catalysed by Ru/C: A Reaction Network Analysis. Green Chem. 2014, 16, 1358–1364. [Google Scholar] [CrossRef]
  47. Mizugaki, T.; Nagatsu, Y.; Togo, K.; Maeno, Z.; Mitsudome, T.; Jitsukawa, K.; Kaneda, K. Selective Hydrogenation of Levulinic Acid to 1,4-Pentanediol in Water Using a Hydroxyapatite-Supported Pt–Mo Bimetallic Catalyst. Green Chem. 2015, 17, 5136–5139. [Google Scholar] [CrossRef]
  48. Li, M.; Li, G.; Li, N.; Wang, A.; Dong, W.; Wang, X.; Cong, Y. Aqueous Phase Hydrogenation of Levulinic Acid to 1,4-Pentanediol. Chem. Commun. 2014, 50, 1414–1416. [Google Scholar] [CrossRef] [PubMed]
  49. Zhang, Z. Synthesis of γ-Valerolactone from Carbohydrates and Its Applications. ChemSusChem 2016, 9, 156–171. [Google Scholar] [CrossRef] [PubMed]
  50. Xiao, C.; Goh, T.-W.; Qi, Z.; Goes, S.; Brashler, K.; Perez, C.; Huang, W. Conversion of Levulinic Acid to γ-Valerolactone over Few-Layer Graphene-Supported Ruthenium Catalysts. ACS Catal. 2016, 6, 593–599. [Google Scholar] [CrossRef] [Green Version]
  51. Yan, K.; Lafleur, T.; Jarvis, C.; Wu, G. Clean and Selective Production of γ-Valerolactone from Biomass-Derived Levulinic Acid Catalyzed by Recyclable Pd Nanoparticle Catalyst. J. Clean Prod. 2014, 72, 230–232. [Google Scholar] [CrossRef]
  52. Du, X.; Liu, Y.; Wang, J.; Cao, Y.; Fan, K. Catalytic Conversion of Biomass-Derived Levulinic Acid into γ-Valerolactone Using Iridium Nanoparticles Supported on Carbon Nanotubes. Chin. J. Catal. 2013, 34, 993–1001. [Google Scholar] [CrossRef]
  53. Sun, P.; Gao, G.; Zhao, Z.; Xia, C.; Li, F. Stabilization of Cobalt Catalysts by Embedment for Efficient Production of Valeric Biofuel. ACS Catal. 2014, 4, 4136–4142. [Google Scholar] [CrossRef]
  54. Manzer, L.E. Catalytic Synthesis of α-Methylene-γ-Valerolactone: A Biomass-Derived Acrylic Monomer. Appl. Catal. A Gen. 2004, 272, 249–256. [Google Scholar] [CrossRef]
  55. Primo, A.; Concepción, P.; Corma, A. Synergy between the Metal Nanoparticles and the Support for the Hydrogenation of Functionalized Carboxylic Acids to Diols on Ru/TiO2. Chem. Commun. 2011, 47, 3613–3615. [Google Scholar] [CrossRef]
  56. Tan, J.; Cui, J.; Ding, G.; Deng, T.; Zhu, Y.; Li, Y. Efficient Aqueous Hydrogenation of Levulinic Acid to γ-Valerolactone over a Highly Active and Stable Ruthenium Catalyst. Catal. Sci. Technol. 2016, 6, 1469–1475. [Google Scholar] [CrossRef]
  57. Wang, R.; Chen, L.; Zhang, X.; Zhang, Q.; Li, Y.; Wang, C.; Ma, L. Conversion of Levulinic Acid to γ-Valerolactone over Ru/Al2O3–TiO2 Catalyst under Mild Conditions. RSC Adv. 2018, 8, 40989–40995. [Google Scholar] [CrossRef] [Green Version]
  58. Almeida, L.D.; Rocha, A.L.A.; Rodrigues, T.S.; Robles-Azocar, P.A. Highly Selective Hydrogenation of Levulinic Acid Catalyzed by Ru on TiO2-SiO2 Hybrid Support. Catal. Today 2020, 344, 158–165. [Google Scholar] [CrossRef]
  59. Ruiz-Bernal, Z.; Lillo-Ródenas, M.Á.; Román-Martínez, M.d.C. Ru Catalysts Supported on Commercial and Biomass-Derived Activated Carbons for the Transformation of Levulinic Acid into γ-Valerolactone under Mild Conditions. Catalysts 2021, 11, 559. [Google Scholar] [CrossRef]
  60. Gupta, S.S.R.; Kantam, M.L. Selective Hydrogenation of Levulinic Acid into γ-Valerolactone over Cu/Ni Hydrotalcite-Derived Catalyst. Catal. Today 2018, 309, 189–194. [Google Scholar] [CrossRef]
  61. Li, Y.; Lan, X.; Liu, B.; Wang, T. Synthesis of γ-Valerolactone from Ethyl Levulinate Hydrogenation and Ethyl 4-Hydroxypentanoate Lactonization over Supported Cu-Ni Bimetallic, Bifunctional Catalysts. J. Ind. Eng. Chem. 2022, 107, 215–223. [Google Scholar] [CrossRef]
  62. Clough, S.R. Hexane. In Encyclopedia of Toxicology, 3rd ed.; Wexler, P., Ed.; Academic Press: Oxford, UK, 2014; pp. 900–904. ISBN 978-0-12-386455-0. [Google Scholar]
  63. Chang, Y.C. Neurotoxic Effects of N-Hexane on the Human Central Nervous System: Evoked Potential Abnormalities in n-Hexane Polyneuropathy. J. Neurol. Neurosurg. Psychiatry 1987, 50, 269–274. [Google Scholar] [CrossRef] [Green Version]
  64. Chhabra, R.S.; Herbert, R.A.; Roycroft, J.H.; Chou, B.; Miller, R.A.; Renne, R.A. Carcinogenesis Studies of Tetrahydrofuran Vapors in Rats and Mice. Toxicol. Sci. 1998, 41, 183–188. [Google Scholar] [CrossRef] [PubMed]
  65. Sun, D.; Saito, T.; Otsuka, S.; Ozawa, T.; Yamada, Y.; Sato, S. Selective Hydrogenation of γ-Valerolactone to 2-Methyltetrahydrofuran over Cu/Al2O3 Catalyst. Appl. Catal. A Gen. 2020, 590, 117309. [Google Scholar] [CrossRef]
  66. Bijoy, R.; Agarwala, P.; Roy, L.; Thorat, B.N. Unconventional Ethereal Solvents in Organic Chemistry: A Perspective on Applications of 2-Methyltetrahydrofuran, Cyclopentyl Methyl Ether, and 4-Methyltetrahydropyran. Org. Process Res. Dev. 2022, 26, 480–492. [Google Scholar] [CrossRef]
  67. Dutta, S.; Yu, I.K.M.; Tsang, D.C.W.; Ng, Y.H.; Ok, Y.S.; Sherwood, J.; Clark, J.H. Green Synthesis of Gamma-Valerolactone (GVL) through Hydrogenation of Biomass-Derived Levulinic Acid Using Non-Noble Metal Catalysts: A Critical Review. Chem. Eng. J. 2019, 372, 992–1006. [Google Scholar] [CrossRef]
  68. Yoshida, R.; Sun, D.; Yamada, Y.; Sato, S. Stable Cu-Ni/SiO2 Catalysts Prepared by Using Citric Acid-Assisted Impregnation for Vapor-Phase Hydrogenation of Levulinic Acid. Mol. Catal. 2018, 454, 70–76. [Google Scholar] [CrossRef]
  69. Sheldon, R.A. Green Solvents for Sustainable Organic Synthesis: State of the Art. Green Chem. 2005, 7, 267–278. [Google Scholar] [CrossRef]
  70. Pendem, S.; Mondal, I.; Shrotri, A.; Rao, B.S.; Lingaiah, N.; Mondal, J. Unraveling the Structural Properties and Reactivity Trends of Cu–Ni Bimetallic Nanoalloy Catalysts for Biomass-Derived Levulinic Acid Hydrogenation. Sustain. Energy Fuels 2018, 2, 1516–1529. [Google Scholar] [CrossRef]
  71. Shao, Y.; Sun, K.; Li, Q.; Liu, Q.; Zhang, S.; Liu, Q.; Hu, G.; Hu, X. Copper-Based Catalysts with Tunable Acidic and Basic Sites for the Selective Conversion of Levulinic Acid/Ester to γ-Valerolactone or 1,4-Pentanediol. Green Chem. 2019, 21, 4499–4511. [Google Scholar] [CrossRef]
  72. Zheng, J.; Zhu, J.; Xu, X.; Wang, W.; Li, J.; Zhao, Y.; Tang, K.; Song, Q.; Qi, X.; Kong, D.; et al. Continuous Hydrogenation of Ethyl Levulinate to γ-Valerolactone and 2-Methyl Tetrahydrofuran over Alumina Doped Cu/SiO2 Catalyst: The Potential of Commercialization. Sci. Rep. 2016, 6, 28898. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Balla, P.; Perupogu, V.; Vanama, P.K.; Komandur, V.R.C. Hydrogenation of Biomass-Derived Levulinic Acid to γ-Valerolactone over Copper Catalysts Supported on ZrO2. J. Chem. Technol. Biotechnol. 2016, 91, 769–776. [Google Scholar] [CrossRef]
  74. Putrakumar, B.; Nagaraju, N.; Kumar, V.P.; Chary, K.V.R. Hydrogenation of Levulinic Acid to γ-Valerolactone over Copper Catalysts Supported on γ-Al2O3. Catal. Today 2015, 250, 209–217. [Google Scholar] [CrossRef]
  75. Popova, M.; Trendafilova, I.; Oykova, M.; Mitrev, Y.; Shestakova, P.; Mihályi, M.R.; Szegedi, Á. Hydrodeoxygenation of Levulinic Acid to &gamma;-Valerolactone over Mesoporous Silica-Supported Cu-Ni Composite Catalysts. Molecules 2022, 27, 5383. [Google Scholar] [CrossRef] [PubMed]
  76. Jones, D.R.; Iqbal, S.; Ishikawa, S.; Reece, C.; Thomas, L.M.; Miedziak, P.J.; Morgan, D.J.; Edwards, J.K.; Bartley, J.K.; Willock, D.J.; et al. The Conversion of Levulinic Acid into γ-Valerolactone Using Cu–ZrO2 Catalysts. Catal. Sci. Technol. 2016, 6, 6022–6030. [Google Scholar] [CrossRef]
  77. Ishikawa, S.; Jones, D.R.; Iqbal, S.; Reece, C.; Morgan, D.J.; Willock, D.J.; Miedziak, P.J.; Bartley, J.K.; Edwards, J.K.; Murayama, T.; et al. Identification of the Catalytically Active Component of Cu–Zr–O Catalyst for the Hydrogenation of Levulinic Acid to γ-Valerolactone. Green Chem. 2017, 19, 225–236. [Google Scholar] [CrossRef]
  78. Yuan, J.; Li, S.-S.; Yu, L.; Liu, Y.-M.; Cao, Y.; He, H.-Y.; Fan, K.-N. Copper-Based Catalysts for the Efficient Conversion of Carbohydrate Biomass into γ-Valerolactone in the Absence of Externally Added Hydrogen. Energy Environ. Sci. 2013, 6, 3308–3313. [Google Scholar] [CrossRef]
  79. 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]
  80. Zhu, Y.; Kong, X.; Zhu, S.; Dong, F.; Zheng, H.; Zhu, Y.; Li, Y.-W. Construction of Cu/ZrO2/Al2O3 Composites for Ethanol Synthesis: Synergies of Ternary Sites for Cascade Reaction. Appl. Catal. B 2015, 166–167, 551–559. [Google Scholar] [CrossRef]
  81. He, Z.; Lin, H.; He, P.; Yuan, Y. Effect of Boric Oxide Doping on the Stability and Activity of a Cu–SiO2 Catalyst for Vapor-Phase Hydrogenation of Dimethyl Oxalate to Ethylene Glycol. J. Catal. 2011, 277, 54–63. [Google Scholar] [CrossRef]
  82. He, D.; He, Q.; Jiang, P.; Zhou, G.; Hu, R.; Fu, W. Novel Cu/Al2O3-ZrO2 Composite for Selective Hydrogenation of Levulinic Acid to γ-Valerolactone. Catal. Commun. 2019, 125, 82–86. [Google Scholar] [CrossRef]
  83. Fang, C.; Kuboon, S.; Khemthong, P.; Butburee, T.; Chakthranont, P.; Itthibenchapong, V.; Kasamechonchung, P.; Witoon, T.; Faungnawakij, K. Highly Dispersed NiCu Nanoparticles on SBA-15 for Selective Hydrogenation of Methyl Levulinate to γ-Valerolactone. Int. J. Hydrogen Energy 2020, 45, 24054–24065. [Google Scholar] [CrossRef]
  84. Xu, Q.; Li, X.; Pan, T.; Yu, C.; Deng, J.; Guo, Q.; Fu, Y. Supported Copper Catalysts for Highly Efficient Hydrogenation of Biomass-Derived Levulinic Acid and γ-Valerolactone. Green Chem. 2016, 18, 1287–1294. [Google Scholar] [CrossRef]
  85. Scotti, N.; Bossola, F.; Zaccheria, F.; Ravasio, N. Copper–Zirconia Catalysts: Powerful Multifunctional Catalytic Tools to Approach Sustainable Processes. Catalysts 2020, 10, 168. [Google Scholar] [CrossRef] [Green Version]
  86. Gebresillase, M.N.; Raguindin, R.Q.; Kim, H.; Seo, J.G. Supported Bimetallic Catalysts for the Solvent-Free Hydrogenation of Levulinic Acid to γ-Valerolactone: Effect of Metal Combination (Ni-Cu, Ni-Co, Cu-Co). Catalysts 2020, 10, 1354. [Google Scholar] [CrossRef]
  87. Obregón, I.; Corro, E.; Izquierdo, U.; Requies, J.; Arias, P.L. Levulinic Acid Hydrogenolysis on Al2O3-Based Ni-Cu Bimetallic Catalysts. Chin. J. Catal. 2014, 35, 656–662. [Google Scholar] [CrossRef]
  88. Scotti, N.; Zaccheria, F.; Evangelisti, C.; Psaro, R.; Ravasio, N. Dehydrogenative Coupling Promoted by Copper Catalysts: A Way to Optimise and Upgrade Bio-Alcohols. Catal. Sci. Technol. 2017, 7, 1386–1393. [Google Scholar] [CrossRef]
  89. Armelao, L.; Bertagnolli, H.; Bleiner, D.; Groenewolt, M.; Gross, S.; Krishnan, V.; Sada, C.; Schubert, U.; Tondello, E.; Zattin, A. Highly Dispersed Mixed Zirconia and Hafnia Nanoparticles in a Silica Matrix: First Example of a ZrO2–HfO2–SiO2 Ternary Oxide System. Adv. Funct. Mater. 2007, 17, 1671–1681. [Google Scholar] [CrossRef] [Green Version]
  90. Sulym, I.; Sternik, D.; Oleksenko, L.; Lutsenko, L.; Borysenko, M.; Derylo-Marczewska, A. Highly Dispersed Silica-Supported Ceria–Zirconia Nanocomposites: Preparation and Characterization. Surf. Interfaces 2016, 5, 8–14. [Google Scholar] [CrossRef]
  91. Zaccheria, F.; Bossola, F.; Scotti, N.; Evangelisti, C.; Dal Santo, V.; Ravasio, N. On Demand Production of Ethers or Alcohols from Furfural and HMF by Selecting the Composition of a Zr/Si Catalyst. Catal. Sci. Technol. 2020, 10, 7502–7511. [Google Scholar] [CrossRef]
  92. Scotti, N.; Dangate, M.; Gervasini, A.; Evangelisti, C.; Ravasio, N.; Zaccheria, F. Unraveling the Role of Low Coordination Sites in a Cu Metal Nanoparticle: A Step toward the Selective Synthesis of Second Generation Biofuels. ACS Catal. 2014, 4, 2818–2826. [Google Scholar] [CrossRef]
  93. Liu, S.; Fan, G.; Yang, L.; Li, F. Highly Efficient Transformation of γ-Valerolactone to Valerate Esters over Structure-Controlled Copper/Zirconia Catalysts Prepared via a Reduction-Oxidation Route. Appl. Catal. A Gen. 2017, 543, 180–188. [Google Scholar] [CrossRef]
  94. Cai, B.; Zhou, X.-C.; Miao, Y.-C.; Luo, J.-Y.; Pan, H.; Huang, Y.-B. Enhanced Catalytic Transfer Hydrogenation of Ethyl Levulinate to γ-Valerolactone over a Robust Cu–Ni Bimetallic Catalyst. ACS Sustain Chem. Eng. 2017, 5, 1322–1331. [Google Scholar] [CrossRef]
  95. Genchi, G.; Carocci, A.; Lauria, G.; Sinicropi, M.S.; Catalano, A. Nickel: Human Health and Environmental Toxicology. Int. J. Environ. Res. Public Health 2020, 17, 679. [Google Scholar] [CrossRef] [Green Version]
  96. Das, K.K.; Reddy, R.C.; Bagoji, I.B.; Das, S.; Bagali, S.; Mullur, L.; Khodnapur, J.P.; Biradar, M.S. Primary Concept of Nickel Toxicity—An Overview. J. Basic Clin. Physiol. Pharmacol. 2019, 30, 141–152. [Google Scholar] [CrossRef] [Green Version]
  97. Gilkey, M.J.; Xu, B. Heterogeneous Catalytic Transfer Hydrogenation as an Effective Pathway in Biomass Upgrading. ACS Catal. 2016, 6, 1420–1436. [Google Scholar] [CrossRef]
  98. Wang, D.; Astruc, D. The Golden Age of Transfer Hydrogenation. Chem. Rev. 2015, 115, 6621–6686. [Google Scholar] [CrossRef] [PubMed]
  99. Espro, C.; Gumina, B.; Szumelda, T.; Paone, E.; Mauriello, F. Catalytic Transfer Hydrogenolysis as an Effective Tool for the Reductive Upgrading of Cellulose, Hemicellulose, Lignin, and Their Derived Molecules. Catalysts 2018, 8, 313. [Google Scholar] [CrossRef] [Green Version]
  100. He, J.; Li, H.; Lu, Y.-M.; Liu, Y.-X.; Wu, Z.-B.; Hu, D.-Y.; Yang, S. Cascade Catalytic Transfer Hydrogenation–Cyclization of Ethyl Levulinate to γ-Valerolactone with Al–Zr Mixed Oxides. Appl. Catal. A Gen. 2016, 510, 11–19. [Google Scholar] [CrossRef]
  101. Tang, X.; Chen, H.; Hu, L.; Hao, W.; Sun, Y.; Zeng, X.; Lin, L.; Liu, S. Conversion of Biomass to γ-Valerolactone by Catalytic Transfer Hydrogenation of Ethyl Levulinate over Metal Hydroxides. Appl. Catal. B 2014, 147, 827–834. [Google Scholar] [CrossRef]
  102. Du, X.-L.; He, L.; Zhao, S.; Liu, Y.-M.; Cao, Y.; He, H.-Y.; Fan, K.-N. Hydrogen-Independent Reductive Transformation of Carbohydrate Biomass into γ-Valerolactone and Pyrrolidone Derivatives with Supported Gold Catalysts. Angew. Chem. Int. Ed. 2011, 50, 7815–7819. [Google Scholar] [CrossRef]
  103. Tabanelli, T.; Paone, E.; Blair Vásquez, P.; Pietropaolo, R.; Cavani, F.; Mauriello, F. Transfer Hydrogenation of Methyl and Ethyl Levulinate Promoted by a ZrO2 Catalyst: Comparison of Batch vs Continuous Gas-Flow Conditions. ACS Sustain. Chem. Eng. 2019, 7, 9937–9947. [Google Scholar] [CrossRef]
  104. Tang, X.; Zeng, X.; Li, Z.; Hu, L.; Sun, Y.; Liu, S.; Lei, T.; Lin, L. Production of γ-Valerolactone from Lignocellulosic Biomass for Sustainable Fuels and Chemicals Supply. Renew. Sustain. Energy Rev. 2014, 40, 608–620. [Google Scholar] [CrossRef]
  105. Komanoya, T.; Nakajima, K.; Kitano, M.; Hara, M. Synergistic Catalysis by Lewis Acid and Base Sites on ZrO2 for Meerwein–Ponndorf–Verley Reduction. J. Phys. Chem. C 2015, 119, 26540–26546. [Google Scholar] [CrossRef]
  106. Kuwahara, Y.; Kaburagi, W.; Osada, Y.; Fujitani, T.; Yamashita, H. Catalytic Transfer Hydrogenation of Biomass-Derived Levulinic Acid and Its Esters to γ-Valerolactone over ZrO2 Catalyst Supported on SBA-15 Silica. Catal. Today 2017, 281, 418–428. [Google Scholar] [CrossRef]
  107. Campbell, E.J.; Zhou, H.; Nguyen, S.T. Catalytic Meerwein−Pondorf−Verley Reduction by Simple Aluminum Complexes. Org. Lett. 2001, 3, 2391–2393. [Google Scholar] [CrossRef]
  108. Assary, R.S.; Curtiss, L.A.; Dumesic, J.A. Exploring Meerwein–Ponndorf–Verley Reduction Chemistry for Biomass Catalysis Using a First-Principles Approach. ACS Catal. 2013, 3, 2694–2704. [Google Scholar] [CrossRef]
  109. Zhang, M.; Li, R.; Wu, Y.; Yu, Y. Mechanistic Insights into the Meerwein–Ponndorf–Verley Reaction and Relative Side Reactions over MgO in the Process of Ethanol to 1,3-Butadiene: A DFT Study. Ind. Eng. Chem. Res. 2021, 60, 2871–2880. [Google Scholar] [CrossRef]
  110. Douthwaite, M.; Zhang, B.; Iqbal, S.; Miedziak, P.J.; Bartley, J.K.; Willock, D.J.; Hutchings, G.J. Transfer Hydrogenation of Methyl Levulinate with Methanol to Gamma Valerolactone over Cu-ZrO2: A Sustainable Approach to Liquid Fuels. Catal. Commun. 2022, 164, 106430. [Google Scholar] [CrossRef]
  111. Shimizu, K. Heterogeneous Catalysis for the Direct Synthesis of Chemicals by Borrowing Hydrogen Methodology. Catal. Sci. Technol. 2015, 5, 1412–1427. [Google Scholar] [CrossRef]
  112. de Graauw, J.A.; van Bekkum, H.; Huskens, J.C.F.P. Meerwein-Ponndorf-Verley Reductions and Oppenauer Oxidations: An Integrated Approach. Synth. (Stuttg.) 1994, 1994, 1007–1017. [Google Scholar] [CrossRef]
  113. Wei, J.; Cao, X.; Wang, T.; Liu, H.; Tang, X.; Zeng, X.; Sun, Y.; Lei, T.; Liu, S.; Lin, L. Catalytic Transfer Hydrogenation of Biomass-Derived 5-Hydroxymethylfurfural into 2,5-Bis(Hydroxymethyl)Furan over Tunable Zr-Based Bimetallic Catalysts. Catal. Sci. Technol. 2018, 8, 4474–4484. [Google Scholar] [CrossRef]
  114. Yu, N.; Lu, H.; Yang, W.; Zheng, Y.; Hu, Q.; Liu, Y.; Wu, K.; Liang, B. Transfer Hydrogenation of Levulinic Acid to γ-Valerolactone over Acid Site-Modified CuNi Alloy. Biomass Convers. Biorefin. 2022. [Google Scholar] [CrossRef]
  115. Deng, L.; Li, J.; Lai, D.-M.; Fu, Y.; Guo, Q.-X. Catalytic Conversion of Biomass-Derived Carbohydrates into γ-Valerolactone without Using an External H2 Supply. Angew. Chem. Int. Ed. 2009, 48, 6529–6532. [Google Scholar] [CrossRef]
  116. Deng, L.; Zhao, Y.; Li, J.; Fu, Y.; Liao, B.; Guo, Q.-X. Conversion of Levulinic Acid and Formic Acid into γ-Valerolactone over Heterogeneous Catalysts. ChemSusChem 2010, 3, 1172–1175. [Google Scholar] [CrossRef]
  117. Scotti, N.; Zaccheria, F.; Bisio, C.; Vittoni, C.; Ravasio, N. Switching Selectivity in the Hydrogen Transfer Reduction of Furfural. ChemistrySelect 2018, 3, 8344–8348. [Google Scholar] [CrossRef]
  118. Wang, J.; Jaenicke, S.; Chuah, G.-K. Zirconium–Beta Zeolite as a Robust Catalyst for the Transformation of Levulinic Acid to γ-Valerolactone via Meerwein–Ponndorf–Verley Reduction. RSC Adv. 2014, 4, 13481–13489. [Google Scholar] [CrossRef]
  119. Hengne, A.M.; Kadu, B.S.; Biradar, N.S.; Chikate, R.C.; Rode, C.V. Transfer Hydrogenation of Biomass-Derived Levulinic Acid to γ-Valerolactone over Supported Ni Catalysts. RSC Adv. 2016, 6, 59753–59761. [Google Scholar] [CrossRef]
  120. Tang, X.; Zeng, X.; Li, Z.; Li, W.; Jiang, Y.; Hu, L.; Liu, S.; Sun, Y.; Lin, L. In Situ Generated Catalyst System to Convert Biomass-Derived Levulinic Acid to γ-Valerolactone. ChemCatChem 2015, 7, 1372–1379. [Google Scholar] [CrossRef]
  121. Tabanelli, T.; Vásquez, P.B.; Paone, E.; Pietropaolo, R.; Dimitratos, N.; Cavani, F.; Mauriello, F. Improved Catalytic Transfer Hydrogenation of Levulinate Esters with Alcohols over ZrO2 Catalyst. Chem. Proc. 2020, 2, 28. [Google Scholar] [CrossRef]
  122. Chan-Thaw, C.E.; Marelli, M.; Psaro, R.; Ravasio, N.; Zaccheria, F. New Generation Biofuels: γ-Valerolactone into Valeric Esters in One Pot. RSC Adv. 2013, 3, 1302–1306. [Google Scholar] [CrossRef]
  123. Pothu, R.; Gundeboyina, R.; Boddula, R.; Perugopu, V.; Ma, J. Recent Advances in Biomass-Derived Platform Chemicals to Valeric Acid Synthesis. New J. Chem. 2022, 46, 5907–5921. [Google Scholar] [CrossRef]
  124. Velisoju, V.K.; Jampaiah, D.; Gutta, N.; Bentrup, U.; Brückner, A.; Bhargava, S.K.; Akula, V. Conversion of γ-Valerolactone to Ethyl Valerate over Metal Promoted Ni/ZSM-5 Catalysts: Influence of Ni0/Ni2+ Heterojunctions on Activity and Product Selectivity. ChemCatChem 2020, 12, 1341–1349. [Google Scholar] [CrossRef] [Green Version]
  125. Yi, Z.; Hu, D.; Xu, H.; Wu, Z.; Zhang, M.; Yan, K. Metal Regulating the Highly Selective Synthesis of Gamma-Valerolactone and Valeric Biofuels from Biomass-Derived Levulinic Acid. Fuel 2020, 259, 116208. [Google Scholar] [CrossRef]
  126. Pan, T.; Deng, J.; Xu, Q.; Xu, Y.; Guo, Q.-X.; Fu, Y. Catalytic Conversion of Biomass-Derived Levulinic Acid to Valerate Esters as Oxygenated Fuels Using Supported Ruthenium Catalysts. Green Chem. 2013, 15, 2967–2974. [Google Scholar] [CrossRef]
  127. Bacchiocchi, R.; De Maron, J.; Tabanelli, T.; Bianchi, D.; Cavani, F. Supported Rhenium Catalysts for the Hydrogenation of Levulinic Acid Derivatives: Limits and Potential. Sustain. Energy Fuels 2023, 7, 671–681. [Google Scholar] [CrossRef]
  128. Muñoz-Olasagasti, M.; López Granados, M.; Jiménez-Gómez, C.P.; Cecilia, J.A.; Maireles-Torres, P.; Dumesic, J.A.; Mariscal, R. The Relevance of Lewis Acid Sites on the Gas Phase Reaction of Levulinic Acid into Ethyl Valerate Using CoSBA-XAl Bifunctional Catalysts. Catal. Sci. Technol. 2021, 11, 4280–4293. [Google Scholar] [CrossRef]
  129. Muñoz-Olasagasti, M.; Martínez-Salazar, I.; Granados, M.L.; López-Aguado, C.; Iglesias, J.; Morales, G.; Mariscal, R. Elucidating the Roles of Acid Site Nature and Strength in the Direct Conversion of Levulinic Acid into Ethyl Valerate: The Case of Zr-Modified Beta Zeolite-Supported Pd Catalysts. Sustain. Energy Fuels 2022, 6, 1164–1174. [Google Scholar] [CrossRef]
  130. Martínez Figueredo, K.G.; Virgilio, E.M.; Segobia, D.J.; Bertero, N.M. Production of Pentyl Valerate from γ-Valerolactone, Pentanol and H2 Using Pd and Rh-Based Bifunctional Catalysts. React. Chem. Eng. 2022, 7, 1997–2008. [Google Scholar] [CrossRef]
  131. Yu, Z.; Lu, X.; Xiong, J.; Ji, N. Transformation of Levulinic Acid to Valeric Biofuels: A Review on Heterogeneous Bifunctional Catalytic Systems. ChemSusChem 2019, 12, 3915–3930. [Google Scholar] [CrossRef]
  132. Al-Naji, M.; Van Aelst, J.; Liao, Y.; d’Hullian, M.; Tian, Z.; Wang, C.; Gläser, R.; Sels, B.F. Pentanoic Acid from γ-Valerolactone and Formic Acid Using Bifunctional Catalysis. Green Chem. 2020, 22, 1171–1181. [Google Scholar] [CrossRef] [Green Version]
  133. Arnaud, S.P.; Wu, L.; Wong Chang, M.-A.; Comerford, J.W.; Farmer, T.J.; Schmid, M.; Chang, F.; Li, Z.; Mascal, M. New Bio-Based Monomers: Tuneable Polyester Properties Using Branched Diols from Biomass. Faraday Discuss. 2017, 202, 61–77. [Google Scholar] [CrossRef] [Green Version]
  134. Stadler, B.M.; Brandt, A.; Kux, A.; Beck, H.; de Vries, J.G. Properties of Novel Polyesters Made from Renewable 1,4-Pentanediol. ChemSusChem 2020, 13, 556–563. [Google Scholar] [CrossRef] [PubMed]
  135. Tang, D.; Macosko, C.W.; Hillmyer, M.A. Thermoplastic Polyurethane Elastomers from Bio-Based Poly(δ-Decalactone) Diols. Polym. Chem. 2014, 5, 3231–3237. [Google Scholar] [CrossRef]
  136. Shen, R.; Long, M.; Lei, C.; Dong, L.; Yu, G.; Tang, J. Anticorrosive Waterborne Polyurethane Coatings Derived from Castor Oil and Renewable Diols. Chem. Eng. J. 2022, 433, 134470. [Google Scholar] [CrossRef]
  137. Zhang, Y.; Wang, H.; Zhao, H.; Liu, Z.; Huang, J.; Yang, Y.; Chen, Y. Characterization of Liquefied Products from Corn Stalk in the Presence of Polyhydric Alcohols with Acid Catalysis. Bioresources 2022, 17, 4262–4279. [Google Scholar] [CrossRef]
  138. Samoilov, V.; Ni, D.; Goncharova, A.; Zarezin, D.; Kniazeva, M.; Ladesov, A.; Kosyakov, D.; Bermeshev, M.; Maximov, A. Bio-Based Solvents and Gasoline Components from Renewable 2,3-Butanediol and 1,2-Propanediol: Synthesis and Characterization. Molecules 2020, 25, 1723. [Google Scholar] [CrossRef]
  139. Piskun, A.S.; van de Bovenkamp, H.H.; Rasrendra, C.B.; Winkelman, J.G.M.; Heeres, H.J. Kinetic Modeling of Levulinic Acid Hydrogenation to γ-Valerolactone in Water Using a Carbon Supported Ru Catalyst. Appl. Catal. A Gen. 2016, 525, 158–167. [Google Scholar] [CrossRef]
  140. Rozenblit, A.; Avoian, A.J.; Tan, Q.; Sooknoi, T.; Resasco, D.E. Reaction Mechanism of Aqueous-Phase Conversion of γ-Valerolactone (GVL) over a Ru/C Catalyst. J. Energy Chem. 2016, 25, 1008–1014. [Google Scholar] [CrossRef] [Green Version]
  141. Bababrik, R.M.; Wang, B.; Resasco, D.E. Reaction Mechanism for the Conversion of γ-Valerolactone (GVL) over a Ru Catalyst: A First-Principles Study. Ind. Eng. Chem. Res. 2017, 56, 3217–3222. [Google Scholar] [CrossRef]
  142. Zhang, G.; Li, W.; Fan, G.; Yang, L.; Li, F. Controlling Product Selectivity by Surface Defects over MoOx-Decorated Ni-Based Nanocatalysts for γ-Valerolactone Hydrogenolysis. J. Catal. 2019, 379, 100–111. [Google Scholar] [CrossRef]
  143. Sun, D.; Saito, T.; Yamada, Y.; Chen, X.; Sato, S. Hydrogenation of γ-Valerolactone to 1,4-Pentanediol in a Continuous Flow Reactor. Appl. Catal. A Gen. 2017, 542, 289–295. [Google Scholar] [CrossRef]
  144. Bucciol, F.; Tabasso, S.; Grillo, G.; Menegazzo, F.; Signoretto, M.; Manzoli, M.; Cravotto, G. Boosting Levulinic Acid Hydrogenation to Value-Added 1,4-Pentanediol Using Microwave-Assisted Gold Catalysis. J. Catal. 2019, 380, 267–277. [Google Scholar] [CrossRef]
  145. Fan, M.; Shao, Y.; Sun, K.; Li, Q.; Zhang, S.; Wang, Y.; Xiang, J.; Hu, S.; Wang, S.; Hu, X. Switching Production of γ-Valerolactone and 1,4-Pentanediol from Ethyl Levulinate via Tailoring Alkaline Sites of CuMg Catalyst and Hydrogen Solubility in Reaction Medium. Mol. Catal. 2021, 510, 111680. [Google Scholar] [CrossRef]
  146. Zhai, X.; Li, C.; Di, X.; Yin, D.; Liang, C. Preparation of Cu/MgO Catalysts for γ-Valerolactone Hydrogenation to 1,4-Pentanediol by MOCVD. J. Fuel Chem. Technol. 2017, 45, 537–546. [Google Scholar] [CrossRef]
  147. Wu, J.; Gao, G.; Li, Y.; Sun, P.; Wang, J.; Li, F. Highly Chemoselective Hydrogenation of Lactone to Diol over Efficient Copper-Based Bifunctional Nanocatalysts. Appl. Catal. B 2019, 245, 251–261. [Google Scholar] [CrossRef]
  148. Liu, Q.; Zhao, Z.; Arai, M.; Zhang, C.; Liu, K.; Shi, R.; Wu, P.; Wang, Z.; Lin, W.; Cheng, H.; et al. Transformation of γ-Valerolactone into 1,4-Pentanediol and 2-Methyltetrahydrofuran over Zn-Promoted Cu/Al2O3 Catalysts. Catal. Sci. Technol. 2020, 10, 4412–4423. [Google Scholar] [CrossRef]
  149. Cavuoto, D.; Ravasio, N.; Zaccheria, F.; Marelli, M.; Cappelletti, G.; Campisi, S.; Gervasini, A. Tuning the Cu/SiO2 Wettability Features for Bio-Derived Platform Molecules Valorization. Mol. Catal. 2022, 528, 112462. [Google Scholar] [CrossRef]
  150. Simakova, I.; Demidova, Y.; Simonov, M.; Prikhod’ko, S.; Niphadkar, P.; Bokade, V.; Dhepe, P.; Murzin, D.Y. Heterogeneously Catalyzed γ-Valerolactone Hydrogenation into 1,4-Pentanediol in Milder Reaction Conditions. Reactions 2020, 1, 54–71. [Google Scholar] [CrossRef]
  151. Wang, L.; Xiao, F.-S. The Importance of Catalyst Wettability. ChemCatChem 2014, 6, 3048–3052. [Google Scholar] [CrossRef]
  152. Liu, F.; Huang, K.; Zheng, A.; Xiao, F.-S.; Dai, S. Hydrophobic Solid Acids and Their Catalytic Applications in Green and Sustainable Chemistry. ACS Catal. 2018, 8, 372–391. [Google Scholar] [CrossRef]
  153. Cavuoto, D.; Zaccheria, F.; Ravasio, N. Some Critical Insights into the Synthesis and Applications of Hydrophobic Solid Catalysts. Catalysts 2020, 10, 1337. [Google Scholar] [CrossRef]
  154. Bozell, J.J.; Moens, L.; Elliott, D.C.; Wang, Y.; Neuenscwander, G.G.; Fitzpatrick, S.W.; Bilski, R.J.; Jarnefeld, J.L. Production of Levulinic Acid and Use as a Platform Chemical for Derived Products. Resour. Conserv. Recycl. 2000, 28, 227–239. [Google Scholar] [CrossRef]
  155. Pace, V.; Hoyos, P.; Castoldi, L.; Domínguez de María, P.; Alcántara, A.R. 2-Methyltetrahydrofuran (2-MeTHF): A Biomass-Derived Solvent with Broad Application in Organic Chemistry. ChemSusChem 2012, 5, 1369–1379. [Google Scholar] [CrossRef] [PubMed]
  156. Huang, X.; Kudo, S.; Ashik, U.P.M.; Einaga, H.; Hayashi, J. Selective Hydrodeoxygenation of γ-Valerolactone over Silica-Supported Rh-Based Bimetallic Catalysts. Energy Fuels 2020, 34, 7190–7197. [Google Scholar] [CrossRef]
  157. Pothu, R.; Challa, P.; Rajesh, R.; Boddula, R.; Balaga, R.; Balla, P.; Perugopu, V.; Radwan, A.B.; Abdullah, A.M.; Al-Qahtani, N. Vapour-Phase Selective Hydrogenation of γ-Valerolactone to 2-Methyltetrahydrofuran Biofuel over Silica-Supported Copper Catalysts. Nanomaterials 2022, 12, 3414. [Google Scholar] [CrossRef] [PubMed]
  158. Xie, Z.; Chen, B.; Wu, H.; Liu, M.; Liu, H.; Zhang, J.; Yang, G.; Han, B. Highly Efficient Hydrogenation of Levulinic Acid into 2-Methyltetrahydrofuran over Ni-Cu/Al2O3-ZrO2 Bifunctional Catalysts. Green Chem. 2019, 21, 606–613. [Google Scholar] [CrossRef]
  159. Huang, Y.-B.; Liu, A.-F.; Zhang, Q.; Li, K.-M.; Porterfield, W.B.; Li, L.-C.; Wang, F. Mechanistic Insights into the Solvent-Driven Adsorptive Hydrodeoxygenation of Biomass Derived Levulinate Acid/Ester to 2-Methyltetrahydrofuran over Bimetallic Cu–Ni Catalysts. ACS Sustain. Chem. Eng. 2020, 8, 11477–11490. [Google Scholar] [CrossRef]
  160. Wang, X.; Yu, Z.; Ye, L.; Zhang, M.; Xiong, J.; Zhang, R.; Li, X.; Ji, N.; Lu, X. Layered Double Hydroxide-Derived Bimetallic Ni−Cu Catalysts Prompted the Efficient Conversion of γ-Valerolactone to 2-Methyltetrahydrofuran. ChemCatChem 2022, 14, e202101441. [Google Scholar] [CrossRef]
  161. Upare, P.P.; Lee, J.-M.; Hwang, Y.K.; Hwang, D.W.; Lee, J.-H.; Halligudi, S.B.; Hwang, J.-S.; Chang, J.-S. Direct Hydrocyclization of Biomass-Derived Levulinic Acid to 2-Methyltetrahydrofuran over Nanocomposite Copper/Silica Catalysts. ChemSusChem 2011, 4, 1749–1752. [Google Scholar] [CrossRef]
Scheme 1. Main hydrogenation products derived from LA/LEs.
Scheme 1. Main hydrogenation products derived from LA/LEs.
Catalysts 13 00697 sch001
Scheme 2. Two possible reaction pathways for the hydrocyclization of LA/LEs to GVL.
Scheme 2. Two possible reaction pathways for the hydrocyclization of LA/LEs to GVL.
Catalysts 13 00697 sch002
Figure 1. C=O activation over a copper catalyst prepared by CH.
Figure 1. C=O activation over a copper catalyst prepared by CH.
Catalysts 13 00697 g001
Scheme 3. Reaction pathway for the CTH of LA to GVL, adapted from ref. [106].
Scheme 3. Reaction pathway for the CTH of LA to GVL, adapted from ref. [106].
Catalysts 13 00697 sch003
Scheme 4. Reaction mechanism for the conversion of EL into valeric esters, adapted from refs. [92,122].
Scheme 4. Reaction mechanism for the conversion of EL into valeric esters, adapted from refs. [92,122].
Catalysts 13 00697 sch004
Figure 2. Products obtainable from LA/LEs by varying catalyst properties and reaction conditions.
Figure 2. Products obtainable from LA/LEs by varying catalyst properties and reaction conditions.
Catalysts 13 00697 g002
Figure 3. Main factors limiting the catalyst scale-up and the design strategies aimed at optimizing the performance.
Figure 3. Main factors limiting the catalyst scale-up and the design strategies aimed at optimizing the performance.
Catalysts 13 00697 g003
Table 2. Catalysts for the CTH of LA/LEs to GVL.
Table 2. Catalysts for the CTH of LA/LEs to GVL.
CatalystPreparation ProcedureReaction ConditionsCatalytic Performances (%)Ref.
Cu/ZrO2WI
DP
OG
LA, FA, H2O, LA/FA molar ratio = 1,
200 °C, 5 h, flow.
C = 100
S = 100
[78]
Cu-Ni/Al2O3WILEs, 2-BuOH, molLE/gcat = 0.01mol g−1, 150 °C,
12 h, batch.
C = 100
S = 97
[94]
Cu-Ni/SBA15glycol WILEs, 2-PrOH, molLE/gcat = 0.0048
mol g−1, 140 °C, 3 h, batch.
C = 91
S = 90
[83]
Cu/ZrO2OGLEs, MeOH, molLE/gcat = 0.019 mol g−1, 220 °C, 1 h, batch.C = 99
S = 88
[110]
Cu-Ni-Al2O3/ACWILA, 2-PrOH, molLA/gcat = 0.086 mol g−1, 220 °C, 2 h, batch.C = 100
S = 97
[114]
Table 3. Selected key catalytic aspects of the comprehension of Cu and bimetallic catalysts.
Table 3. Selected key catalytic aspects of the comprehension of Cu and bimetallic catalysts.
FocusKey Catalytic Aspects[Ref.]
Bimetallic Cu-Ni
  • Balanced amount of acid and metallic sites
  • NP synergistic effects
  • Cu inhibits coke, Ni enhances activity and prevents sintering
[87]
Cu-γAl2O3
  • Importance of catalytic acidity for the C-O bond cleavage
[74]
Cu-MOx
  • Effect of the support on acidity
  • Catalytic performances: Cu-γAl2O3 > Cu-ZrO2 > Cu/SiO2 > Cu/TiO2
  • Effect of Cu loading on acid site concentration
[73]
Cu/SiO2 vs. Cu/Al2O3-SiO2
  • Importance of the support acidity
  • Tuning product selectivity with activation temperature
  • Particle size matters
[72]
Cu/t-ZrO2
  • Identification of Cu0 as the catalytically active site
  • Importance of Cu NP dispersion
[77]
Bimetallic
Cu-Ag/Al2O3
Effect of Ag:
  • limitation of Cu leaching
  • facilitated in situ CuO reduction
  • maintenance of Cu in metallic state
[11]
Cu-Ni organosilica nanospheres
  • Organosilica as innovative support allows lowering of the reaction T
  • Synergic effect of Cu-Ni is responsible for high reactivity and stability
[70]
Cu-Ni on KIT-6 ZSM-5 doped
  • Confinement effect of the support on metal nanoparticles: high dispersed Cu-Ni and limited sintering
  • Hydrogenation at ambient P due to highly dispersed and active particles
  • Importance of Lewis–Brønsted acid sites for the ring opening
[75]
B2O3-Cu/ZrO2Effect of B:
  • favored Cu dispersion
  • retarded Cu sintering
  • inhibited ZrO2 phase change
  • tuning of Cu0/Cu+ surface distribution
  • improved surface acidity
[30]
Al2O3-Cu/ZrO2
  • enhanced electronic interaction between Cu and ZrO2 due to the dilution of Cu by Al2O3
[114]
Cu-Ni on mesoporous high surface SBA-15
  • Low reaction temperature (140 °C), stable Cu-Ni into SBA-15 mesopores
[83]
Table 4. Most relevant copper catalysts used for hydrogenation of GVL to give 1,4-PDO from the literature.
Table 4. Most relevant copper catalysts used for hydrogenation of GVL to give 1,4-PDO from the literature.
Batch
CatalystPreparation MethodReaction ConditionsCatalytic Performances (%)Ref.
Cu/MgOCVD1,4-dioxane, 240 °C, 10 h, P(H2) = 100 bar,
molGVL/g cat = 0.0125 mol g−1, batch
C = 91
S = 94
[146]
Cu1.5/Mg1.5AlOCP1,4-dioxane, 160 °C, 12 h, P(H2) = 50 bar, molGVL/g cat = 0.0625 mol g−1, batchC = 93
S = 99
[147]
Zn1.5Cu/Al2O3WI1,4-dioxane, 200 °C, 2 h, P(H2) = 40 bar,
molGVL/g cat = 0.02 mol g−1, batch
C = 91
S = 97
[148]
Cu/SiO2CHCyclopentyl-methylether, 160 °C, 22 h, P(H2) = 50 bar,
molGVL/g cat = 0.05 mol g−1, batch
C = 78
S = 98
[23]
Cu/SiO2 –10% TEOCSCHCyclopentyl-methylether, 160 °C, 22 h, P(H2) = 50 bar,
molGVL/g cat= 0.05 mol g−1, batch
C = 80
S = 99
[149]
Flow
CatalystPreparation MethodReaction ConditionsCatalytic Performances (%)Ref.
Cu/ZnOWI1,4-dioxane, 140 °C, WHSV = 0.4 h−1,
H2/GVL molar ratio = 810, flow
C = 82
S = 99
[143]
Cu/SiO2WI1-butanol, 130 °C, WHSV = 4.61 h−1,
H2/GVL molar ratio = 2485, flow
C = 32
S = 67
[150]
Table 5. Most relevant copper catalysts used for hydrogenation of LA/LEs or GVL to give 2-MTHF from the literature.
Table 5. Most relevant copper catalysts used for hydrogenation of LA/LEs or GVL to give 2-MTHF from the literature.
Batch
CatalystPreparation
Method
Reaction ConditionsCatalytic Performances (%)Ref.
Cu-Ni/Al2O3–ZrO2SG
WI
LA, 2-BuOH, 210 °C, 10 h,
P (H2) = 35 bar, molLA/gcat = 0.043 mol g−1, batch
C = 100
S = 100
[158]
Cu-Ni/Al2O3WILEs, n-HEX, 180 °C, 4 h,
P (H2) = 40 bar, molLE/g cat = 0.0083 mol g−1, batch
C = 100
S = 98
[159]
Ni2Cu1/Al2O3CPGVL, 2-PrOH, 200 °C, 5 h,
P (H2) = 50 bar, molGVL/g cat = 0.04 mol g−1, batch
C = 100
S = 88
[160]
Flow
CatalystPreparation
Method
Reaction ConditionsCatalytic Performances (%)Ref.
Cu/Al2O3-SiO2CPLEs, EtOH, 250 °C, 1000 h,
P (H2) = 30 bar, WHSV = 0.6 h−1,
H2/LE molar ratio = 50, flow
C = 100
S = 65
[72]
Ni–Cu/SiO2CPGVL, 1,4 dioxane, 265 °C,
P (H2) = 25 bar, WHSV = 0.5 h−1
H2/LE molar ratio = 80, flow
C = 100
S = 89
[161]
Cu/SiO2DPGVL, H2O, 300 °C,
P (H2) = 1 bar, H2 = 50 mL min−1, WHSV = 1.0 h−1, H2/GVL molar ratio = 24, flow
C = 97
S = 84
[157]
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

Cavuoto, D.; Ardemani, L.; Ravasio, N.; Zaccheria, F.; Scotti, N. Some Insights into the Use of Heterogeneous Copper Catalysts in the Hydroprocessing of Levulinic Acid. Catalysts 2023, 13, 697. https://doi.org/10.3390/catal13040697

AMA Style

Cavuoto D, Ardemani L, Ravasio N, Zaccheria F, Scotti N. Some Insights into the Use of Heterogeneous Copper Catalysts in the Hydroprocessing of Levulinic Acid. Catalysts. 2023; 13(4):697. https://doi.org/10.3390/catal13040697

Chicago/Turabian Style

Cavuoto, Denise, Leandro Ardemani, Nicoletta Ravasio, Federica Zaccheria, and Nicola Scotti. 2023. "Some Insights into the Use of Heterogeneous Copper Catalysts in the Hydroprocessing of Levulinic Acid" Catalysts 13, no. 4: 697. https://doi.org/10.3390/catal13040697

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop