Next Article in Journal
Aldolase: A Desirable Biocatalytic Candidate for Biotechnological Applications
Previous Article in Journal
Heterogeneous Photocatalytic Oxidation and Detoxification of Simulated Agricultural Wastewater Contaminated with Boscalid Fungicide Using g-C3N4 Catalyst
Previous Article in Special Issue
Exclusive Papers on Environmentally Friendly Catalysis in China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ion-Exchange Synthesis of Surrounded CoNi@Al2O3 Catalyst for Levulinic Acid Hydrogenation to γ-Valerolactone under Mild Conditions

1
Key Laboratory of Mesoscopic Chemistry MOE, School of Chemistry and Chemical Engineering, Nanjing University, Nanjing 210023, China
2
Jiangsu Key Laboratory of Vehicle Emissions Control, Nanjing University, Nanjing 210023, China
3
Nanjing University-Yangzhou Institute of Chemistry and Chemical Engineering, Yangzhou 211900, China
*
Author to whom correspondence should be addressed.
Catalysts 2024, 14(2), 113; https://doi.org/10.3390/catal14020113
Submission received: 7 January 2024 / Revised: 27 January 2024 / Accepted: 29 January 2024 / Published: 31 January 2024

Abstract

:
The use of eco-friendly biomass as a resource is an efficient way to address the problems of fossil fuel depletion and climate change. In biomass conversion, versatile γ-valerolactone (GVL) is generally obtained from levulinic acid (LA) hydrogenation via a multimetallic catalyst system. Despite conversion efficiency being enhanced in mild conditions due to metal interactions, maintaining high catalyst stability is still a challenge. In this study, we synthesized a surrounded Co0.52Ni0.48@Al2O3-IE catalyst that exhibited excellent alloying and synergistic interaction between the metal constituents. Under relatively mild reaction conditions, the GVL yield over the catalyst exceeded 99% in LA hydrogenation. The catalyst showed no deactivation in a test of five cycles, displaying superiority in stability, possibly due to reasons of the physical isolation of the shell and the alumina retention on the Co-Ni alloys surface caused by the reversibility of exchange equilibrium. The present work demonstrated that a surrounded structured catalyst fabricated by ion exchange (IE) with active metals physically enclosed can lead to high catalytic activity and superior stability.

Graphical Abstract

1. Introduction

In light of the depletion of fossil fuels and escalating concerns about global warming, the scientific communities have redoubled their efforts to explore sustainable and renewable resources. Biomass, characterized by its renewability, environmental friendliness, and sustainability, has emerged as a promising solution [1,2]. However, an uneconomic aspect lies in the fact that a substantial portion of biomass undergoes natural decomposition by microorganisms, resulting in the release of carbon dioxide, water, and heat, and this natural breakdown process fails to fully harness the energy and organic chemicals inherent in biomass. Therefore, further exploration and optimization of strategies for comprehensive biomass utilization are indispensable. LA is recognized as one of the “Top Ten” biomass-derived platform molecules by the U.S. Department of Energy due to its tremendous commercial attributes, which can be obtained through acid-catalyzed hydrolysis of cellulose [3,4,5]. The selective hydrogenation of LA or its esters holds the potential for producing a variety of high-value derivatives, including GVL, 1,4-butanediol (PDO), 2-methyltetrahydrofuran (MTHF), valeric acid, etc. [6]. Among these biomass-derived compounds, GVL has been found in widespread applications, such as food components, fuel additives, green solvent, advanced renewable fuel, pharmaceutical intermediates, and so on [7,8].
In the realm of catalytic processes from LA to GVL, both noble metals (e.g., Ru, Rh, Pt, Pd, and Au) [9,10,11,12,13,14], and non-precious metals (e.g., Ni, Cu, an Co) are used in the hydrogenation of LA to GVL [5,7,15,16,17,18,19,20]. Noble metal catalysts generally operate under mild reaction conditions and demonstrate notable catalytic efficiency. Nevertheless, it is true that their high costs pose challenges for large-scale industrial applications [21]. Moreover, the extensively employed Ru/C catalyst is plagued by relatively weak metal–support interactions, resulting in the facile loss of active components and consequently diminishing catalyst stability [11,14,18]. Non-precious metals have received widespread attention due to their cheapness and easy availability. Nevertheless, non-noble metals also manifest typical drawbacks, such as suboptimal catalytic activity and stringent reaction conditions, limiting their developments [22]. Over the past several decades, there has been a large amount of studies dedicated to the augmentation of catalytic efficiency in non-noble metals through the synergistic effects of multimetallic systems. These endeavors aim to harness the combined impact of various metal species to modify electronic configurations and improve adsorption properties, thereby delving into the intricate interplay of factors that influence catalytic performance. This multidimensional approach seeks to unravel the intricate mechanisms underlying the cooperative behavior of multiple metals in the catalytic system, offering insights that contribute to the refinement and optimization of non-noble metal catalysts for diverse applications [23,24,25,26,27,28]. For example, Huang et al. reported the effectiveness of the bimetallic Fe-Re catalyst supported by TiO2, which contained metallic Re nanoparticles coated by FeOx species and small quantities of Fe-Re alloys, giving a GVL yield of 95% at 180 °C under 40 bar of H2 in water [6]. Tang et al. successfully synthesized the bimetallic Ni1-Zn1@OMC catalyst, achieving a 93% GVL yield at 180 °C within a 90 min reaction time. The incorporation of Zn metals led to the formation of Ni-Zn alloys, playing a pivotal role in activating the carbonyl group in LA. Simultaneously, the active sites of Niδ–Znδ+ facilitated the weakening and cleavage of C–O bonds in 4-hydroxyvaleric acid (4-HA) [29]. Wang et al. designed a bimetallic Ni3Fe NPs@C catalyst by a simple one-step carbothermal reduction method, which was capable of showing high activity for selective conversion of LA to GVL by both direct and transfer hydrogenation mechanisms. Staggeringly, the bimetallic Ni3Fe NPs@C catalyst was 6 and 40 times more reactive than both monometallic Ni NPs@C and Fe NPs@C in direct hydrogenation, respectively [30]. Notwithstanding the considerable body of researches on bimetallic catalysts, normally, elevating temperatures and pressures are necessary for achieving the optimal reaction conditions of the hydrogenation of levulinic acid (LA) to gamma-valerolactone (GVL). As such, the quest for bimetallic catalysts capable of facilitating LA hydrogenation under mild conditions, coupled with outstanding cyclic stability, still represents a formidable challenge [2,21,31,32,33].
In this study, a novel bimetallic catalyst for the hydrogenation of LA to GVL was developed by ion exchange directly. Notably, the ion exchanging reactions are more inclined to proceed in the direction of a lower dissolution equilibrium constant (A(OH)x + By+  B(OH)y + Ax+, KspB(OH)x < KspA(OH)y). CoNi-MMH (mixed-metal hydroxide) precursors were synthesized according to methodologies outlined in the prior literature, exhibiting a well-defined nanosheet structure (Scheme 1) [34,35,36]. Consequently, the bimetallic hydroxide served as the primary precursor, and then, the Co2+ and Ni2+ were partially exchanged by Al3+ due to the solubility constants. Lastly, alumina-encapsulated Co-Ni alloys could be obtained after H2 reduction (Scheme 1). The catalyst exhibited a high level of alloying effect and the synergistic interaction between metals, resulting in improved LA hydrogenation activity. The alumina encapsulation and the residual traces of alumina after ion exchange (IE) on the alloy surface increased the stability of the catalytic system. Our findings served to underscore the high efficacy and stability of the bimetallic catalysts synthesized through the ion-exchange method in the context of LA hydrogenation for GVL production.

2. Results and Discussion

Previous studies by Guo and coworkers [37] demonstrated that the spontaneous replacement of Ni2+ by Al3+ can be facilitated by utilizing the discrepancies in solubility constants. In this work, Co2+ and Ni2+ were directly exchanged by Al3+ because they have similar exchange rates due to the comparable solubility constants (KspAl(OH)3 = 1.3 × 10−33 < KspNi(OH)2 = 2 × 10−15 ≈ KspCo(OH)2 = 1.6 × 10−15) of Ni(OH)2 and Co(OH)2, resulting in the formation of CoxNiy(OH)2@Al(OH)3 which exhibited diffraction peaks of hydrotalcite-like structure (Figure 1a) [35,38]. Following a controlled calcination process, the XRD patterns of the sample revealed distinctive peaks at 2θ of 31.4°, 36.9°, 45.0°, 59.5°, and 65.3°. These discernible peaks can be confidently attributed to Co3O4 (PDF#43-1003). Additionally, other notable peaks detected at 2θ of 37.3°, 43.4°, and 62.9° can be ascribed to NiO (PDF#44-1159) [3,39,40]. Compared to the catalyst prepared by the impregnation method [37], the diffraction peaks of Co3O4 and NiO shifted to higher angles, plausibly due to the entry of Al3+ into the Co3O4 and NiO lattice (Figure 1b). After H2 reduction, the oxides were transformed into the corresponding metal components. Due to the similarity in structural features and closeness in peak position, it is difficult to distinguish the Co, Ni, and alloy phases by XRD (Figure 1c,d). Within a narrow range, the Co-Ni alloys peaks were located between the metal peaks of Co and Ni, and negatively shifted with increasing Co/Ni ratio, indicating enhanced formation of Co-Ni alloys. The shift is primarily attributed to the fact that Ni and Co atoms are alike in crystal structure, atomic radius, and valence state [41,42]. Furthermore, the peak intensity increased with the rise i Ni content, signifying improved crystallinity and enlarged particle size.
The EDX maps of Co0.52Ni0.48@Al2O3-IE showed that the cobalt and nickel signals highly overlap at localized positions, while those of aluminum are highly dispersed, suggesting a surrounded structure of the alumina-coated Co-Ni alloy [39]. The signal of Mg was extremely weak which meant that there was almost no retention of Mg (Figure 2a). Further evidence of the generation of the surrounded structure, as well as the partial residual of Al2O3 on the metal surface, could be found on the basis of TEM (Figure 2b). Upon reduction, the lattice spacing converged between that characteristic of metallic cobalt (111) and metallic nickel (111), offering further substantiation for the emergence of an alloy phase. Conversely, the catalyst synthesized through impregnation exhibited a particle size of 14.3 nm, and a notable increase compared to the Co0.52Ni0.48@Al2O3-IE catalyst (10.2 nm) (Figure 2b). This observation served as confirmation that the encapsulated structure fabricated through ion exchange adeptly restrained the growth of metal particles during the processes of calcination and high-temperature reduction. This encapsulation mechanism, arising from ion exchange, played a pivotal role in controlling the particle size distribution and preserving the nanostructure integrity of the Co0.52Ni0.48@Al2O3-IE catalyst, contributing to its enhanced stability and catalytic performance, eventually.
Moreover, we investigated the reduction behavior of the catalysts by H2-TPR. Ni@Al2O3-IE had two hydrogen-consuming peaks at 485 °C and 588 °C attributable to β-NiO which had strong interaction with alumina [37,43]. The TPR curves of Co@Al2O3-IE showed peaks at 327 °C and 422 °C corresponding to the reduction of Co3O4 and CoO, respectively [16,44]. The peak at 678 °C was due to CoOx-Al2O3 or CoAlxOy. Owing to the similar crystal structures of Co3O4 and γ-Al2O3, it was likely to have ion migration from cobalt oxide into the Al2O3 support [45,46] (Figure 3a). For bimetallic catalysts with different Co/Ni ratios, the reduction peaks gradually shifted to lower temperatures with increasing Ni content, suggesting interaction between cobalt–nickel oxides. Meanwhile the rise in peak area with an escalating Co/Ni ratio was attributed to the higher molar consumption of H2 by cobalt compared to nickel, partly due to the presence of trivalent Co. In contrast to the catalysts synthesized through impregnation, the Co0.52Ni0.48@Al2O3-IE catalyst exhibited a comparatively elevated reduction temperature. This observation implied the more robust interactions with the support, a factor known to enhance catalyst stability, as illustrated in Figure 3b. The higher reduction temperature signified enhanced metal–support interactions, leading to a more secure anchoring of the active metal species on the Al2O3 support. This strengthened interaction was instrumental in fortifying the catalyst against sintering and agglomeration during the reduction process, thereby contributing to its sustained catalytic activity over prolonged reaction cycles [31,47].
The BET surface area, total pore volume, and pore size of the catalysts are listed in Table 1. The Co0.52Ni0.48@Al2O3-IE catalyst exhibited a large specific surface area of 246.9 m2/g and a relatively homogeneous mesoporous structure, whereas the specific surface area decreased with the variation in the Co/Ni ratio, and the slight deviations may be due to the slight difference in Ksp between Co and Ni, which caused inhomogeneity in the carrier layer after exchange and may have resulted in the accumulation of pore structures. Relatively, the Co0.50Ni0.50/Al2O3-IM catalyst prepared by the impregnation method, of which the specific surface area was only 132.5 m2/g, was significantly lower than that of the catalyst prepared by the ion-exchange method. Figure 4 displays the isothermal adsorption–desorption and pore size distribution curves of the catalysts. The Co0.52Ni0.48@Al2O3-IE catalyst exhibited a characteristic type IV isotherm profile with an H2-type hysteresis loop. Moreover, the pore size distributions of the ion exchange (IE) series were noticeably more uniform than the catalyst of the impregnation method (IM). In comparison to other catalysts, Co0.52Ni0.48@Al2O3-IE possessed substantial specific surface area and uniform pore structure, favoring the promotion of reactant diffusion and molecular transfer, and contributing to the enhancement in catalyst activity.
The investigation of catalyst chemical compositions and electronic states was performed through X-ray photoelectron spectroscopy (XPS) characterization. This analytical technique enabled a detailed examination of the surface elemental composition and oxidation states of the catalyst, providing valuable insights into its chemical structure and electronic configuration. After the reduction treatment of the calcined metal oxide catalysts, it was clearly observed that the catalysts were reduced to the metallic states. The presence of some oxide signals in the peaks may be attributed to oxidation during the testing process. Notably, the Ni 2p region (Figure 5a) showed two predominant peaks at 852.93 eV and 855.74 eV which could be ascribed to Ni0 and Ni2+, respectively [38]. Figure 5b shows that the Co 2p spectra had three main peaks including Co0 (~777.87 eV), Co3+ (~780.13 eV), and Co2+ (~781.86 eV) [39,40]. The O 1s fine spectrum revealed three distinct signal peaks, each associated with specific oxygen species (Figure 5c). These included lattice oxygen (OI), surface defect-related oxygen (e.g., oxygen vacancies), and surface hydroxyl groups (OII) encompassing O22− and O, as well as adsorbed CO2 (OIII). Notably, the higher OII/OI suggested that the bimetallic catalysts possessed a higher density of oxygen vacancies. This increased abundance of oxygen vacancies enhanced the adsorption capacity of carbonyl groups present in the reactant molecules [48]. In contrast to the monometallic catalysts, the Co peaks of the bimetallic catalyst exhibited a discernible shift toward lower binding energy, whereas there was a noticeable increase in the binding energy of the Ni 2p peaks of Ni0 and Ni2+. These shifts served as a compelling indicator of intermetallic electron interactions between Ni and Co, indicating the generation of CoNi alloys [49]. As a consequence of this charge transfer mechanism, there was the generation of electron-enriched Co0 centers on the catalyst surface [50,51].
The catalytic efficiency of catalysts was assessed in the context of LA hydrogenation to GVL. Among the bimetallic catalysts with different Co/Ni ratios, Co0.52Ni0.48@Al2O3-IE showed the highest GVL yield, and thus we optimized the reaction conditions using this catalyst. Interestingly, the type of alcohol solvent had a significant effect on GVL selectivity, with isopropanol (IPA) leading to higher selectivity and yield compared to the primary alcohol. This finding was consistent with those of previous reports [5,30,52] (Figure 6a).
Regarding H2 pressure, a pivotal reducing agent in the hydrogenation of LA to GVL, its significance is demonstrated in Figure 6b. Notably, the GVL yield reached nearly 94% when the H2 pressure was maintained at 2 MPa, underscoring the facilitating role of moderate H2 pressure in the reaction. Furthermore, elevating the H2 pressure from 1 MPa to 3 MPa led to LA conversion and GVL yield exceeding 99.5%. Nevertheless, raising the H2 pressure beyond 2 MPa failed to yield further enhancement in GVL yield, signifying that a pressure of 2 MPa was adequate for the hydrogenation of LA to GVL. Figure 6c illustrates the outcomes of temperature adjustments spanning from 110 °C to 150 °C. Notably, no significant alterations were observed beyond this temperature range (130 °C to 150 °C). The reaction time played a pivotal role in the hydrogenation of LA to GVL. We meticulously optimized the reaction time, and the results are depicted in Figure 6d. The optimal conditions were reached with a 4 h reaction time, where a remarkable 97.6% GVL yield and complete conversion of LA were achieved. Extending the reaction time beyond 4 h did not result in a significant improvement in GVL yield, indicating that the GVL product is stable in the current system. Overall, we determined that the optimal reaction conditions are as follows: 4 h of reaction time, 130 °C, and 2 MPa H2. These conditions were notably milder than those of typical ones reported in the literature, and we could reach a remarkable GVL yield exceeding 99% (Table 2).
When comparing bimetallic catalysts to their monometallic counterparts (Figure 7), it was evident that the bimetallic catalyst exhibits significantly higher activity. The enhanced activity could be attributed to the synergistic interaction between Ni and Co, and the prominent specific surface area facilitating the reactant diffusion and molecular transfer [50]. Ultimately, we speculate that the reaction mechanism is as follows: In comparison to their monometallic counterparts, the bimetallic CoNi nanoparticles (NPs) demonstrated a heightened abundance of surface oxygen vacancies. These vacancies readily coordinated with oxygen atoms present on the carbonyl groups of LA, resulting in the elongation and weakening of the C=O double bond. Apart from the generation of active hydrogen through the dissociation of molecular hydrogen, the bimetallic CoNi NPs exhibited an additional capability: they facilitated the dehydrogenation of isopropanol (IPA), leading to the production of active hydrogen. This dual activation of hydrogen proved to be effective in attacking the less robust C=O double bond, thereby expediting a swift hydrogenation reaction to yield the 4-HA intermediate [48,49]. Both the partial acid sites provided by acidic support Al2O3 and a relatively high temperature facilitated the dehydration cyclization of 4-HA into GVL [3,57,58].
We assessed the recyclability of the catalyst (Figure 8). The capacity of a catalyst to consistently sustain its catalytic activity and selectivity across successive reaction cycles holds paramount significance for its pragmatic utilization and economic feasibility in industrial applications. A catalyst characterized by superior cyclic stability not only ensures prolonged effectiveness over an extended operational duration but also plays a crucial role in diminishing the overall cost associated with the catalytic process. This dual advantage enhances the environmental sustainability and economic viability of the process. When comparing the recyclability performance between the ion-exchange (IE) and impregnation (IM) catalysts, a discernible difference emerged. The former demonstrated consistent and stable activity over five consecutive test runs, whereas the latter experienced deactivation after only two runs. This observation underscores the pivotal role played by the encapsulation of alloy particles and the potential influence of residual Al2O3 on the surface of the cobalt–nickel alloy, both possibly contributing to the stabilization of Co0.52Ni0.48@Al2O3-IE.

3. Materials and Methods

3.1. Materials

Magnesium carbonate basic (AR, 40–45% MgO basis) was procured from Rhawn (Shanghai, China). Nickel nitrate hexahydrate (Ni(NO3)2·6H2O, AR, ≥98.0%) and cobalt nitrate hexahydrate (Co(NO3)2·6H2O, AR, ≥98.5%) were acquired from Sinopharm (Shanghai, China). Levulinic acid (LA, AR, 99.0%) was sourced from Shanghai Macklin Biochemical Co., Ltd. (Shanghai, China). γ-Valerolactone (GVL) and γ-Alumina (Al2O3) were obtained from Aladdin (Shanghai, China). Aluminum nitrate (Al(NO3)3·9H2O, AR, 99.5%) was procured from Nanjing Reagent (Nanjing, China). All of the aforementioned chemicals were utilized as received without further purification, ensuring their chemical integrity and suitability for the intended experimental procedures.

3.2. Catalyst Preparation

3.2.1. Synthesis of CoNi-MMH Precursors

Magnesium carbonate basic underwent calcination at 750 °C for 2 h in a static air atmosphere, resulting in the formation of magnesium oxide (MgO). In a standard synthesis protocol, the MgO obtained (1.21 g, 0.03 mol) was immersed in a solution containing cobalt and nickel ions with varying Co/Ni ratios, maintaining a total “Co + Ni” concentration equivalent to that of MgO, stirring at 500 revolutions per minute for 24 h at room temperature. Following this, the resulting dark green CoNi-MMH products were filtered and washed seven times with deionized water (the green color deepened as the proportion of Co increased). Ultimately, the products were dried at 60 °C overnight. This meticulously executed procedure ensured the complete exchange of cobalt and nickel ions with magnesium ions of the template, resulting in the desired CoNi-MMH for further study.

3.2.2. Synthesis of CoxNiy(OH)2@Al(OH)3

CoNi-MMH and aluminium nitrate were dissolved in 60 mL of deionized water solution at a molar ratio of 2/1, followed by rapid stirring for 30 min, otherwise a gel may form. Subsequently, this mixture was transferred into a 100 mL Teflon-lined stainless steel autoclave, then hermetically sealed, and maintained at a temperature of 120 °C for a duration of 12 h. Following the completion of the reaction, the resulting precipitate, denoted as CoxNiy(OH)2@Al(OH)3, was isolated through centrifugation, followed by thorough rinsing with deionized water. Lastly, the obtained product underwent a drying process at 80 °C overnight. This meticulous procedure ensured the successful formation of CoxNiy(OH)2@Al(OH)3, a precursor material for subsequent investigations.

3.2.3. Synthesis of CoxNiy@Al2O3-IE

Following calcination of CoxNiy(OH)2@Al(OH)3 under a static air atmosphere at 400 °C for 2 h, a gray powder was obtained. Subsequently, the powder underwent a reduction in a 5% H2/Ar gas flow at 650 °C (with a heating rate of 5 °C min−1) for a duration of 2 h, resulting in the formation of CoxNiy@Al2O3-IE. Various samples with different metal loadings can be prepared by carefully adjusting the Al3+/bimetallic hydroxide molar ratio. In CoxNiy@Al2O3-IE, the variables x and y represent the Ni and Co mass percentages determined by ICP-OES, respectively. The cumulative metal loading was meticulously controlled to be 25%. This systematic procedure ensured the reproducibility and accuracy of the metal loadings in CoxNiy@Al2O3-IE, facilitating precise analysis and interpretation of its catalytic properties.

3.2.4. Synthesis of Comparative Catalysts

The catalyst for comparison was prepared using a conventional impregnation method, noted as Co0.50Ni0.50/Al2O3-IM. The preparation process for surrounded single (Co or Ni) catalysts was like that of bimetallic catalysts described earlier: Through ion exchange, single-metal nanosheets were prepared. Initially, a concentrated 1 M solution of nickel nitrate or cobalt nitrate was prepared. Subsequently, 6.40 g of freshly prepared magnesium oxide was added to the above solution. The mixture was stirred at room temperature for 48 h. The excess nickel nitrate or cobalt nitrate and an extended reaction time facilitated complete ion exchange between metal ions and Mg(OH)2 templates. Following filtration, the resulting product was washed multiple times with water and ethanol, filtered, dried, and then ground to obtain light green Ni(OH)2 or dark green Co(OH)2 nanosheets. The ensuing preparation process was similar to that of the previous bimetallic catalysts.

3.3. Catalyst Characterization

The chemical composition was analyzed using ICP-OES (Avio500, PerkinElmer, Waltham, MA, USA). XRD patterns were obtained using a Philips X′Pert XRD (PANalytical, Almelo, the Netherlands) with Co Kα radiation (35 kV, 40 mA, λ = 1.7902 Å). Scanning electron microscopy (Hitachi S-4800 of JPN) was employed to characterize the morphology of the products. HRTEM images were captured on a JEM-2100F JEOL electron microscope (Tokyo, Japan) operating at 200 kV. Element mapping results were acquired with an X-maxN80T (Oxford Instruments, Oxford, UK) equipped with an energy dispersive spectrometer (EDX). The Micromeritics ASAP 3020 (Norcross, GA, USA) was employed for the analysis of the specific surface area in the samples. The BET method within the relative pressure (P/Po) range of 0.04–0.3 was used to determine the specific surface area, and the pore size distribution curve was calculated using the BJH method. XPS analysis was performed on an Escalab-250Xi system using a monochromatic Al Kα X-ray source, and all binding energies were referenced to the C 1s line at 284.8 eV. H2-TPR experiments were conducted on a FINESORB-3010 (Finetec Instruments, Zhejiang, China) instrument with a thermal conductivity detector (TCD). Prior to TPR measurements, a standard pretreatment procedure involved subjecting the sample (40 mg in all runs) to a 2 h pretreatment at 300 °C under Ar flow to eliminate any absorbed gases. Following pretreatment, the sample was gradually cooled to room temperature and subsequently heated to 800 °C at a rate of 10 °C min−1 in a 5% H2/Ar gas flow (25 mL min−1).

3.4. Catalytic Test

The performance testing of the catalyst for LA hydrogenation was conducted in a high-pressure liquid-phase reaction vessel. Initially, 10 mL of isopropanol solvent, 102.4 μL of LA, 36 μL of internal standard 1,4-dioxane, and 35 mg of freshly reduced catalyst were placed in a Teflon liner. Following this, the reaction vessel was then sealed, with subsequent purging of 0.1 MPa H2 to displace air from the reaction vessel multiple times. Finally, after introducing 2 MPa H2, the reaction proceeded at 130 °C for 4 h. The results were then analyzed for composition by GC (GC-9860) with a capillary column and a flame ionization detector (FID). The chromatographic column was FFAP (length: 30 m, inner diameter: 0.32, and film thickness: 0.50 μm). The catalyst’s performance was evaluated based on several reference values, including the following:
The conversion of LA was calculated as follows:
Conv . ( % ) = ( m L A , 0 m L A , T m L A , 0 ) × 100
GVL selectivity was calculated as follows:
S e l . ( % ) = ( m G V L m L A , 0 m L A , T ) × 100
The yield of GVL was calculated as follows:
Yield ( % ) = Conv . × S e l . 100

4. Conclusions

In this work, we successfully generated an encapsulated catalyst with good alloying, which can achieve more than 99% GVL yield in LA hydrogenation under relatively mild reaction conditions thanks to the synergistic interaction between metals. Both the active Co-Ni alloy particles physically encapsulated by alumina, because of the surrounded structure resulting from ion exchange, and some alumina on the surface of the alloy particles, due to the reversibility of exchange equilibrium, may have contributed to the stabilization of the catalyst. Compared to the catalyst prepared by impregnation where the activity started to decrease after two cycles, the encapsulated Co0.52Ni0.48@Al2O3-IE counterpart showed no deactivation in a test of five cycles, displaying superiority in stability. The results show that the bimetallic catalyst synthesized by the ion-exchange method has high efficiency and stability in the production of GVL for the LA hydrogenation reaction. Moreover, the inverse loading synthesis by the ion-exchange method can be extended to many other valuable catalytic systems based on the differences in solubility product constants.

Author Contributions

H.D., X.G. and C.Y. conceived and designed the experiments; H.D., C.Y. and C.J. performed the experiments; W.L., Q.W. and H.D. ananlyzed the data; H.D. and Q.W. contributed reagents/materials/analysis tools; H.D. and X.G. wrote the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Key R&D Program of China (No. 2021YFA1502804), the National Natural Science Foundation of China (22172073, 52003119, 21773112, 21173119, 21273109, 11874199, 21902027, and U19B2003), the Jiangsu Provincial Key Research and Development Program (BE2022611), and the Natural Science Foundation of Jiangsu Province (BK20221286).

Data Availability Statement

The data presented in this study are available.

Acknowledgments

The authors thank the financial supports from the funding mentioned above.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. 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]
  2. 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 Environ. 2018, 232, 1–10. [Google Scholar] [CrossRef]
  3. Bai, J.; Cheng, C.; Liu, Y.; Wang, C.; Liao, Y.; Chen, L.; Ma, L. Selective hydrogenation of levulinic acid to γ-valerolactone on Ni-based catalysts. Mol. Catal. 2021, 516, 112000. [Google Scholar] [CrossRef]
  4. 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. 2016, 5, 1322–1331. [Google Scholar] [CrossRef]
  5. Sakakibara, K.; Endo, K.; Osawa, T. Facile synthesis of γ-valerolactone by transfer hydrogenation of methyl levulinate and levulinic acid over Ni/ZrO2. Catal. Commun. 2019, 125, 52–55. [Google Scholar] [CrossRef]
  6. Huang, X.; Liu, K.; Vrijburg, W.L.; Ouyang, X.; Iulian Dugulan, A.; Liu, Y.; Tiny Verhoeven, M.W.G.M.; Kosinov, N.A.; Pidko, E.A.; Hensen, E.J.M. Hydrogenation of levulinic acid to γ-valerolactone over Fe-Re/TiO2 catalysts. Appl. Catal. B Environ. 2020, 278, 119314. [Google Scholar] [CrossRef]
  7. Raguindin, R.Q.; Desalegn, B.Z.; Vishwanath, H.; Gebresillase, M.N.; Seo, J.G. Enhanced Hydrogenation of Levulinic Acid over Ordered Mesoporous Alumina-Supported Catalysts: Elucidating the Effect of Fabrication Strategy. ChemSusChem 2022, 15, e202102662. [Google Scholar] [CrossRef]
  8. 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]
  9. Zhang, Y.; Chen, C.; Gong, W.; Song, J.; Zhang, H.; Zhang, Y.; Wang, G.; Zhao, H. Self-assembled Pd/CeO2 catalysts by a facile redox approach for high-efficiency hydrogenation of levulinic acid into gamma-valerolactone. Catal. Commun. 2017, 93, 10–14. [Google Scholar] [CrossRef]
  10. Luo, W.; Deka, U.; Beale, A.M.; van Eck, E.R.H.; Bruijnincx, P.C.A.; Weckhuysen, B.M. Ruthenium-catalyzed hydrogenation of levulinic acid: Influence of the support and solvent on catalyst selectivity and stability. J. Catal. 2013, 301, 175–186. [Google Scholar] [CrossRef]
  11. Cao, W.; Lin, L.; Qi, H.; He, Q.; Wu, Z.; Wang, A.; Luo, W.; Zhang, T. In-situ synthesis of single-atom Ir by utilizing metal-organic frameworks: An acid-resistant catalyst for hydrogenation of levulinic acid to γ-valerolactone. J. Catal. 2019, 373, 161–172. [Google Scholar] [CrossRef]
  12. Yan, K.; Lafleur, T.; Wu, G.; Liao, J.; Ceng, C.; Xie, X. Highly selective production of value-added γ-valerolactone from biomass-derived levulinic acid using the robust Pd nanoparticles. Appl. Catal. A Gen. 2013, 468, 52–58. [Google Scholar] [CrossRef]
  13. Li, S.; Wang, Y.; Yang, Y.; Chen, B.; Tai, J.; Liu, H.; Han, B. Conversion of levulinic acid to γ-valerolactone over ultra-thin TiO2 nanosheets decorated with ultrasmall Ru nanoparticle catalysts under mild conditions. Green Chem. 2019, 21, 770–774. [Google Scholar] [CrossRef]
  14. Wang, S.; Zhuang, Z.; Chen, X.; Wang, Y.; Li, X.; Yang, M.; Wu, Y.; Peng, Q.; Chen, C.; Li, Y. 3D Oxide-Derived Ru Catalyst for Ultra-Efficient Hydrogenation of Levulinic Acid to γ-Valerolactone. Small 2023. early view. [Google Scholar] [CrossRef]
  15. Sun, D.; Ohkubo, A.; Asami, K.; Katori, T.; Yamada, Y.; Sato, S. Vapor-phase hydrogenation of levulinic acid and methyl levulinate to γ-valerolactone over non-noble metal-based catalysts. Mol. Catal. 2017, 437, 105–113. [Google Scholar] [CrossRef]
  16. Shao, Y.; Ba, S.; Sun, K.; Gao, G.; Fan, M.; Wang, J.; Fan, H.; Zhang, L.; Hu, X. Selective production of γ-valerolactone or 1,4-pentanediol from levulinic acid/esters over Co-based catalyst: Importance of the synergy of hydrogenation sites and acidic sites. Chem. Eng. J. 2022, 429, 132433. [Google Scholar] [CrossRef]
  17. Ren, D.; Zhao, C.; Zhang, N.; Norinaga, K.; Zeng, X.; Huo, Z. Catalytic transfer hydrogenation of ethyl levulinate into γ-valerolactone over air-stable skeletal cobalt catalyst. J. Environ. Chem. Eng. 2022, 10, 107188. [Google Scholar] [CrossRef]
  18. Lan, F.; Zhang, H.; Zhao, C.; Shu, Y.; Guan, Q.; Li, W. Copper Clusters Encapsulated in Carbonaceous Mesoporous Silica Nanospheres for the Valorization of Biomass-Derived Molecules. ACS Catal. 2022, 12, 5711–5725. [Google Scholar] [CrossRef]
  19. Zhang, C.; Huo, Z.; Ren, D.; Song, Z.; Liu, Y.; Jin, F.; Zhou, W. Catalytic transfer hydrogenation of levulinate ester into γ-valerolactone over ternary Cu/ZnO/Al2O3 catalyst. J. Energy Chem. 2019, 32, 189–197. [Google Scholar] [CrossRef]
  20. 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]
  21. Xu, H.; Hu, D.; Zhang, M.; Wang, Y.; Zhao, Z.; Jiang, Z.; Garces, H.F.; Yan, K. Bimetallic NiCu Alloy Catalysts for Hydrogenation of Levulinic Acid. ACS Appl. Nano Mater. 2021, 4, 3989–3997. [Google Scholar] [CrossRef]
  22. Kadu, B.S.; Hengne, A.M.; Biradar, N.S.; Rode, C.V.; Chikate, R.C. Reductive Cyclization of Levulinic Acid to γ-Valerolactone over Non-Noble Bimetallic Nanocomposite. Ind. Eng. Chem. Res. 2016, 55, 13032–13039. [Google Scholar] [CrossRef]
  23. Kim, J.; Choi, H.; Kim, D.; Park, J.Y. Operando Surface Studies on Metal-Oxide Interfaces of Bimetal and Mixed Catalysts. ACS Catal. 2021, 11, 8645–8677. [Google Scholar] [CrossRef]
  24. Nakaya, Y.; Furukawa, S. Catalysis of Alloys: Classification, Principles, and Design for a Variety of Materials and Reactions. Chem. Rev. 2022, 123, 5859–5947. [Google Scholar] [CrossRef] [PubMed]
  25. Garkul’, I.A.; Zadesenets, A.V.; Plyusnin, P.E.; Filatov, E.Y.; Asanova, T.I.; Kozlov, D.V.; Korenev, S.V. Zinc(II) and Manganese(II) Oxalatopalladates as Precursors of Bimetallic Nanomaterials. Russ. J. Inorg. Chem. 2020, 65, 1571–1576. [Google Scholar] [CrossRef]
  26. Yudanova, L.I.; Logvinenko, V.A.; Ishchenko, A.V.; Rudina, N.A. Quantum Size Effect in Bimetallic Nanoparticles Obtained via Thermolysis of Solid Solutions of Co(II), Ni(II), Zn(II) Salts of Maleic Acid. Russ. J. Phys. Chem. A 2020, 94, 2108–2114. [Google Scholar] [CrossRef]
  27. Yang, Y.; Sun, Y.; Luo, X. The Relationship between Structure and Catalytic Activity-Stability of Non-Precious Metal-Based Catalysts towards Levulinic Acid Hydrogenation to γ-Valerolactone: A Review. Energies 2022, 15, 8093. [Google Scholar] [CrossRef]
  28. Minamihara, H.; Kusada, K.; Wu, D.; Yamamoto, T.; Toriyama, T.; Matsumura, S.; Kumara, L.S.R.; Ohara, K.; Sakata, O.; Kawaguchi, S.; et al. Continuous-Flow Reactor Synthesis for Homogeneous 1 nm-Sized Extremely Small High-Entropy Alloy Nanoparticles. J. Am. Chem. Soc. 2022, 144, 11525–11529. [Google Scholar] [CrossRef]
  29. Tang, Y.; Fu, J.; Wang, Y.; Guo, H.; Qi, X. Bimetallic Ni-Zn@OMC catalyst for selective hydrogenation of levulinic acid to γ-valerolactone in water. Fuel Process. Technol. 2023, 240, 107559. [Google Scholar] [CrossRef]
  30. Wang, H.; Chen, C.; Zhang, H.; Wang, G.; Zhao, H. An efficient and reusable bimetallic Ni3Fe NPs@C catalyst for selective hydrogenation of biomass-derived levulinic acid to γ-valerolactone. Chin. J. Catal. 2018, 39, 1599–1607. [Google Scholar] [CrossRef]
  31. Huo, J.; Tessonnier, J.-P.; Shanks, B.H. Improving Hydrothermal Stability of Supported Metal Catalysts for Biomass Conversions: A Review. ACS Catal. 2021, 11, 5248–5270. [Google Scholar] [CrossRef]
  32. Jiang, L.; Xu, G.; Fu, Y. A nitrogen-doped carbon modified nickel catalyst for the hydrogenation of levulinic acid under mild conditions. Green Chem. 2021, 23, 7065–7073. [Google Scholar] [CrossRef]
  33. 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]
  34. Zheng, Z.; Lin, L.; Mo, S.; Ou, D.; Tao, J.; Qin, R.; Fang, X.; Zheng, N. Economizing Production of Diverse 2D Layered Metal Hydroxides for Efficient Overall Water Splitting. Small 2018, 14, e1800759. [Google Scholar] [CrossRef]
  35. Wang, X.; Song, H.; Ma, S.; Li, M.; He, G.; Xie, M.; Guo, X. Template ion-exchange synthesis of Co-Ni composite hydroxides nanosheets for supercapacitor with unprecedented rate capability. Chem. Eng. J. 2022, 432, 134319. [Google Scholar] [CrossRef]
  36. Wang, X.; Ding, H.; Luo, W.; Yu, Y.; Chen, Q.; Luo, B.; Xie, M.; Guo, X. Morphology evolution of CoNi-LDHs synergistically engineered by precipitant and variable cobalt for asymmetric supercapacitor with superior cycling stability. EcoEnergy 2023, 1, 448–459. [Google Scholar] [CrossRef]
  37. Hao, P.; Xie, M.; Chen, S.; Li, M.; Bi, F.; Zhang, Y.; Lin, M.; Guo, X.; Ding, W.; Guo, X. Surrounded catalysts prepared by ion-exchange inverse loading. Sci. Adv. 2020, 6, eaay7031. [Google Scholar] [CrossRef]
  38. Wei, D.; Cao, Y.; Yan, L.; Gang, H.; Wu, B.; Ouyang, B.; Chen, P.; Jiang, Y.; Wang, H. Enhanced Pseudo-Capacitance Process in Nanoarchitectural Layered Double Hydroxide Nanoarrays Hollow Nanocages for Improved Capacitive Deionization Performance. ACS Appl. Mater. Interfaces 2023, 15, 24427–24436. [Google Scholar] [CrossRef] [PubMed]
  39. Cheng, S.; Liu, Y.; Zhao, Y.; Zhao, X.; Lang, Z.; Tan, H.; Qiu, T.; Wang, Y. Superfine CoNi alloy embedded in Al2O3 nanosheets for efficient tandem catalytic reduction of nitroaromatic compounds by ammonia borane. Dalton Trans. 2019, 48, 17499–17506. [Google Scholar] [CrossRef] [PubMed]
  40. 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]
  41. He, Y.; Abedi, Z.; Ni, C.; Milliken, S.; O’Connor, K.M.; Ivey, D.G.; Veinot, J.G.C. CoNi Nanoparticle-Decorated ZIF-67-Derived Hollow Carbon Cubes as a Bifunctional Electrocatalyst for Zn–Air Batteries. ACS Appl. Nano Mater. 2022, 5, 12496–12505. [Google Scholar] [CrossRef]
  42. Wu, J.; Yan, X.; Wang, W.; Jin, M.; Xie, Y.; Wang, C. Highly Dispersed CoNi Alloy Embedded in N-doped Graphitic Carbon for Catalytic Transfer Hydrogenation of Biomass-derived Furfural. Chem.–Asian J. 2021, 16, 3194–3201. [Google Scholar] [CrossRef] [PubMed]
  43. Ma, R.; Gao, J.; Kou, J.; Dean, D.P.; Breckner, C.J.; Liang, K.; Zhou, B.; Miller, J.T.; Zou, G. Insights into the Nature of Selective Nickel Sites on Ni/Al2O3 Catalysts for Propane Dehydrogenation. ACS Catal. 2022, 12, 12607–12616. [Google Scholar] [CrossRef]
  44. Son, I.H.; Lee, S.J.; Roh, H.-S. Hydrogen production from carbon dioxide reforming of methane over highly active and stable MgO promoted Co–Ni/γ-Al2O3 catalyst. Int. J. Hydrogen Energy 2014, 39, 3762–3770. [Google Scholar] [CrossRef]
  45. Pardo-Tarifa, F.; Cabrera, S.; Sanchez-Dominguez, M.; Boutonnet, M. Ce-promoted Co/Al2O3 catalysts for Fischer–Tropsch synthesis. Int. J. Hydrogen Energy 2017, 42, 9754–9765. [Google Scholar] [CrossRef]
  46. Yang, X.; Yang, J.; Zhao, T.; Qian, W.; Wang, Y.; Holmen, A.; Jiang, W.; Chen, D.; Ben, H. Kinetic insights into the effect of promoters on Co/Al2O3 for Fischer-Tropsch synthesis. Chem. Eng. J. 2022, 445, 136655. [Google Scholar] [CrossRef]
  47. Li, Y.; Zhang, Y.; Qian, K.; Huang, W. Metal–Support Interactions in Metal/Oxide Catalysts and Oxide–Metal Interactions in Oxide/Metal Inverse Catalysts. ACS Catal. 2022, 12, 1268–1287. [Google Scholar] [CrossRef]
  48. Liu, M.; Zhang, J.; Zheng, L.; Fan, G.; Yang, L.; Li, F. Significant Promotion of Surface Oxygen Vacancies on Bimetallic CoNi Nanocatalysts for Hydrodeoxygenation of Biomass-derived Vanillin to Produce Methylcyclohexanol. ACS Sustain. Chem. Eng. 2020, 8, 6075–6089. [Google Scholar] [CrossRef]
  49. Wang, Y.; Chen, M.; Zhang, K.; Wu, H.; Wang, J.; Cheng, Y.; Liu, Y.; Wei, Z. CoNi Alloy Nanoparticles Confined in N-Doped Porous Carbon as an Efficient and Versatile Catalyst for Reductive Amination of Levulinic Acid/Esters to N-Substituted Pyrrolidones. ACS Catal. 2023, 13, 12601–12616. [Google Scholar] [CrossRef]
  50. Yan, H.; Yao, S.; Zhang, T.; Li, D.; Tang, X.; Chen, M.; Zhou, Y.; Zhang, M.; Liu, Y.; Zhou, X.; et al. Promoting catalytic transfer hydrodecarbonylation of methyl stearate over bimetallic CoNi/HAP catalysts with strong electronic coupling effect. Appl. Catal. B Environ. 2022, 306, 121138. [Google Scholar] [CrossRef]
  51. Hao, X.; Liu, Y.; Deng, J.; Jing, L.; Wang, J.; Pei, W.; Dai, H. Bimetallic CoNi single atoms supported on three-dimensionally ordered mesoporous chromia: Highly active catalysts for n-hexane combustion. Green Energy Environ. 2022, in press. [CrossRef]
  52. Wu, Y.; Wang, H.; Peng, J.; Zhang, J.; Ding, M. Single-atom Cu catalyst in a zirconium-based metal–organic framework for biomass conversion. Chem. Eng. J. 2023, 454, 140156. [Google Scholar] [CrossRef]
  53. Shimizu, K.-i.; Kanno, S.; Kon, K. Hydrogenation of levulinic acid to γ-valerolactone by Ni and MoOx co-loaded carbon catalysts. Green Chem. 2014, 16, 3899–3903. [Google Scholar] [CrossRef]
  54. Rodiansono, R.; Astuti, M.D.; Hara, T.; Ichikuni, N.; Shimazu, S. Efficient hydrogenation of levulinic acid in water using a supported Ni–Sn alloy on aluminium hydroxide catalysts. Catal. Sci. Technol. 2016, 6, 2955–2961. [Google Scholar] [CrossRef]
  55. Gundekari, S.; Srinivasan, K. In situ generated Ni(0)@boehmite from NiAl-LDH: An efficient catalyst for selective hydrogenation of biomass derived levulinic acid to γ-valerolactone. Catal. Commun. 2017, 102, 40–43. [Google Scholar] [CrossRef]
  56. 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]
  57. Shao, S.; Ding, Z.; Shang, C.; Zhang, S.; Ke, Y.; Zhu, G.; Yang, Y. Yolk-shell Co catalysts with controlled nanoparticle/single-atom ratio for aqueous levulinic acid hydrogenation to γ-valerolactone. Chem. Eng. J. 2022, 450, 138153. [Google Scholar] [CrossRef]
  58. Meng, F.; Yang, X.; Zhao, S.; Li, Z.; Zhang, G.; Qi, Y.; Chu, S.; Wang, G.; Zhang, J.; Qin, Y.; et al. Shifting reaction path for levulinic acid aqueous-phase hydrogenation by Pt-TiO2 metal-support interaction. Appl. Catal. B Environ. 2023, 324, 122236. [Google Scholar] [CrossRef]
Scheme 1. Flow chart of catalyst preparation.
Scheme 1. Flow chart of catalyst preparation.
Catalysts 14 00113 sch001
Figure 1. XRD patterns of (a) CoxNiy(OH)2@Al(OH)3 precursors; (b) CoNiOx@Al2O3 with different Co/Ni ratios; (c) CoxNiy@Al2O3 after reduction; (d) Figure (c) with scale enlarged at 2θ = 35~55°.
Figure 1. XRD patterns of (a) CoxNiy(OH)2@Al(OH)3 precursors; (b) CoNiOx@Al2O3 with different Co/Ni ratios; (c) CoxNiy@Al2O3 after reduction; (d) Figure (c) with scale enlarged at 2θ = 35~55°.
Catalysts 14 00113 g001
Figure 2. (a) Nanoscale elemental EDX maps of Co0.52Ni0.48@Al2O3-IE; (b) TEM results of Co0.52Ni0.48@Al2O3-IE and Co0.50Ni0.50/Al2O3-IM.
Figure 2. (a) Nanoscale elemental EDX maps of Co0.52Ni0.48@Al2O3-IE; (b) TEM results of Co0.52Ni0.48@Al2O3-IE and Co0.50Ni0.50/Al2O3-IM.
Catalysts 14 00113 g002
Figure 3. H2-TPR profile of (a) CoxNiy@Al2O3-IE with different Co/Ni ratios; (b) Co0.52Ni0.48@Al2O3-IE prepared by ion exchange and Co0.50Ni0.50/Al2O3-IM prepared by impregnation.
Figure 3. H2-TPR profile of (a) CoxNiy@Al2O3-IE with different Co/Ni ratios; (b) Co0.52Ni0.48@Al2O3-IE prepared by ion exchange and Co0.50Ni0.50/Al2O3-IM prepared by impregnation.
Catalysts 14 00113 g003
Figure 4. The specific surface area of: (a) ion exchanging catalysts with different Co/Ni ratios and (b) catalysts prepared by different methods; The pore size distribution of: (c) ion exchanging catalysts with different Co/Ni ratios and (d) catalysts prepared by different methods.
Figure 4. The specific surface area of: (a) ion exchanging catalysts with different Co/Ni ratios and (b) catalysts prepared by different methods; The pore size distribution of: (c) ion exchanging catalysts with different Co/Ni ratios and (d) catalysts prepared by different methods.
Catalysts 14 00113 g004
Figure 5. XPS spectra of catalysts prepared by ion exchange: (a) Ni 2p; (b) Co 2p; (c) O 1s.
Figure 5. XPS spectra of catalysts prepared by ion exchange: (a) Ni 2p; (b) Co 2p; (c) O 1s.
Catalysts 14 00113 g005
Figure 6. LA conversion and GVL yield as a function of (a) solvents (120 °C, 2 MPa H2, and 30 mg catalysts); (b) H2 pressure (130 °C,10 mL IPA, 4 h, and 30 mg catalysts); (c) temperature (2 MPa H2, 4 h, 10 mL IPA, and 30 mg catalysts); (d) time (130 °C, 2 MPa H2, 10 mL IPA, and 30 mg catalysts).
Figure 6. LA conversion and GVL yield as a function of (a) solvents (120 °C, 2 MPa H2, and 30 mg catalysts); (b) H2 pressure (130 °C,10 mL IPA, 4 h, and 30 mg catalysts); (c) temperature (2 MPa H2, 4 h, 10 mL IPA, and 30 mg catalysts); (d) time (130 °C, 2 MPa H2, 10 mL IPA, and 30 mg catalysts).
Catalysts 14 00113 g006
Figure 7. Comparison with monometallic catalysts.
Figure 7. Comparison with monometallic catalysts.
Catalysts 14 00113 g007
Figure 8. Recycling performance (conditions: 30 mg catalysts, 130 °C, 2 MPa H2, and 4 h).
Figure 8. Recycling performance (conditions: 30 mg catalysts, 130 °C, 2 MPa H2, and 4 h).
Catalysts 14 00113 g008
Table 1. N2 adsorption–desorption isotherms, pore size distributions, and pore volume.
Table 1. N2 adsorption–desorption isotherms, pore size distributions, and pore volume.
SBET (m2/g)Vtotal (cm3/g)Pore Size (nm)
Co0.52Ni0.48@Al2O3-IE246.90.396.3
Co0.27Ni0.73@Al2O3-IE199.80.408.1
Co0.71Ni0.29@Al2O3-IE176.10.419.3
Co0.50Ni0.50/Al2O3-IM132.50.6721.8
γ-Al2O3141.00.8624.6
Table 2. Hydrogenation of LA to GVL over various catalysts under optimal reaction conditions.
Table 2. Hydrogenation of LA to GVL over various catalysts under optimal reaction conditions.
EntryCatalystSolventT/°CH2/MPat/hY/%CyclesRef.
1Co0.52Ni0.48@Al2O3IPA13024>995this work
2Ni–MoOx/C250524>99[53]
3Ni–Sn(1.4)/AlOHH2O12042>99[54]
425%Fe-25%Ni/MMTIPA2003.51>996[22]
5Ni1-Zn1@OMCH2O18021.5>932[29]
6Ni(0)@boehmiteH2O20036>992[55]
7Fe-Re/TiO2H2O1804495[6]
8RuHZSM-51, 4-dioxane22031093.13[56]
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

Ding, H.; Yang, C.; Jiang, C.; Luo, W.; Wang, Q.; Guo, X. Ion-Exchange Synthesis of Surrounded CoNi@Al2O3 Catalyst for Levulinic Acid Hydrogenation to γ-Valerolactone under Mild Conditions. Catalysts 2024, 14, 113. https://doi.org/10.3390/catal14020113

AMA Style

Ding H, Yang C, Jiang C, Luo W, Wang Q, Guo X. Ion-Exchange Synthesis of Surrounded CoNi@Al2O3 Catalyst for Levulinic Acid Hydrogenation to γ-Valerolactone under Mild Conditions. Catalysts. 2024; 14(2):113. https://doi.org/10.3390/catal14020113

Chicago/Turabian Style

Ding, Hongzhi, Chenyu Yang, Congyan Jiang, Wei Luo, Qiuyue Wang, and Xuefeng Guo. 2024. "Ion-Exchange Synthesis of Surrounded CoNi@Al2O3 Catalyst for Levulinic Acid Hydrogenation to γ-Valerolactone under Mild Conditions" Catalysts 14, no. 2: 113. https://doi.org/10.3390/catal14020113

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