Next Article in Journal
Special Issue “10th Anniversary of Catalysts: Biocatalysis in Analysis and Synthesis—Past, Present and Future”
Previous Article in Journal
Synthesis, Characterization and Dye Removal Capability of Conducting Polypyrrole/Mn0.8Zn0.2Fe2O4/Graphite Oxide Ternary Composites
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hydrogenolysis of Lignin and C–O Linkages Containing Lignin-Related Compounds over a Macroporous Silicalite-1 Array-Supported Ru-Ni Phosphide Composite

1
School of Chemical Engineering, Northwest University, Xi’an 710069, China
2
Xi’an Giant Biological Gene Technology Co., Ltd., Xi’an 710077, China
3
College of Urban and Environment Science, Northwest University, Xi’an 710127, China
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(12), 1625; https://doi.org/10.3390/catal12121625
Submission received: 9 November 2022 / Revised: 5 December 2022 / Accepted: 9 December 2022 / Published: 11 December 2022
(This article belongs to the Section Biomass Catalysis)

Abstract

:
Hydrogenolysis via targeted depolymerization of C–O linkages is a techno-economic beneficial process for converting lignin into highly valuable chemicals and clean fuels. In this work, a macroporous silicalite-1 (S-1) array-supported Ru-Ni metallic phosphide composite (Ru-Ni12P5/S-15) was prepared as a catalyst and hydrogenolysis activity under relative mild conditions was investigated using a series of compounds containing ether linkages as lignin-related model compounds. The Lewis acid sites originating from the unreduced Ru species and the macroporous geometry of S-1 significantly influenced hydrogenolysis activity and product selectivity. Analysis of the mechanism demonstrated that both the aryl ether and aliphatic ether linkages were directly hydrogenated over Ru-Ni12P5/S-15. 2D-HSQC-NMR spectroscopy demonstrated that the ether linkages of lignin were efficiently cleaved by Ru-Ni12P5/S-15. Furthermore, the obtained liquid hydrogenolysis products are high value-added chemicals used for pharmaceutical production and can be facilely tuned via the reaction conditions.

1. Introduction

The depletion of oil reserves and environmental issues relevant to the use of traditional fuels has prompted the manufacture of chemicals and fuels from renewable sources [1]. The conversion of lignocellulosic biomass into chemical raw materials is particularly significant to diminish dependence on traditional fuels [2].
Lignin, a conceivable lignocellulosic biomass for the production of aromatic chemicals and fuels [3,4], consists of phenylpropane-related structural units linked by bridge linkages [5]. However, more than 80% of bridge linkages are etheric, implying that the depolymerization of ether linkages is a critical step for a higher value-added exploitation of lignin. Notably, 4–O–5 bonds are highly stable due to the higher bond dissociation energy as compared to αO–4 and βO–4 linkages. Thus, the efficient cleavage of 4–O–5 bonds remains a significant challenge to the biorefinery industry. Usually, diphenyl ether is selected as a lignin-related probe compound to investigate potential catalysts for cleaving 4–O–5 linkages.
Hydrogenolysis via targeted depolymerization of C–O linkages is a techno-economic beneficial process for the conversion of lignin to highly valuable chemicals and clean fuels as compared to pyrolysis, oxidation degradation, and gasification [6,7].
Various hierarchically macroporous materials have been investigated as potential catalysts to enhance the diffusion of reactants and products in liquid-phase reactions and also reduce the distance among molecules to improve catalytic activity and selectivity [8,9]. The improved reactant-catalyst contact efficiency of three-dimensional ordered macroporous materials has enhanced the performance of gas phase-related reactions [10,11].
However, the use of macroporous materials as catalysts for macromolecular reactants, such as lignin and thermal extractions of coals, in liquid-phase reactions has been relatively limited. Recently, Luo et al. found that a hierarchical macro/mesoporous Ni/ASA-supported catalyst facilitated the transport and hydrodepolymerization of lignin, resulting in a yield of ca. 43 wt% for liquid hydrocarbons [12].
Transition metal phosphides are widely applied as catalysts for the hydrotreatment of bio-oils [13], whereas monometallic phosphides are typically used as catalysts under harsh reaction conditions, such as high temperatures and H2 partial pressure [14,15]. The incorporation of noble-metal like Ru, one of the cheapest noble metals, can efficiently enhance the performance of transition metal phosphides and simultaneously moderate the reaction conditions [16,17,18]. However, the dominant disadvantage of metal-based catalysts is the inevitable deactivation due to carbon deposition and metal particles sintering [19,20,21,22]. It is reasonable to believe that catalysts with a macroporous structure should diminish transport limitations and possess better coke-resistant ability, and thus, could extend the catalysts lifetime [23].
In the present work, a macroporous silicalite-1 (S-1) array-supported Ru-Ni metallic phosphide composite (Ru-Ni12P5/S-15) was prepared and employed for the depolymerization of lignin and C–O linkages containing lignin-related compounds. The results demonstrated that the electron-enriched Ruδ+ species and the macroporous array-confined effect promoted the hydrogenolysis of C–O linkages and hydrodeoxygenation even under relative mild conditions, indicating that the geometry of the porous structure plays a pivotal role in macromolecule-related reactions. Macromolecular reactants can be efficiently captured using a fishing net-like macroporous array, and preferentially undergo hydrodeoxygenation/hydrogenation due to the higher concentration of active hydrogen species in the interior of the microporous. Moreover, the liquid hydrogenolysis products from lignin are important pharmaceutical precursors and can be obtained by facilely tuning the reaction conditions. Our present work indicates a feasible strategy for the production of high value-added chemicals from the lignin and thermal extractions of coals.

2. Results and Discussion

2.1. Characterization of the Catalysts

The particle size of the solid S-1 nanocubes was about 300 nm, as shown in Figure 1a. After three days of recrystallization, the shells of the nanocubes were punctuated with macropores of about 200 nm in diameter (Figure 1b). As the recrystallization time was extended to four days, the macroporous nanocubes formed short-range ordered aggregates (Figure 1c), which was attributed to continuous desilication and recrystallization during the alkali treatment process [24]. Moreover, as displayed in Figure 1d,e, a highly ordered macroporous array-like structure formed when the recrystallization time was prolonged to five days. TEM (Figure 1f) further confirmed that the macropores were hollow and interconnected. As compared to S-1, the macroporous array structure of S-15 allowed for macromolecule-related reactions in the liquid phase.
As shown in Figure 2a and Figure 3a, the Ni phosphide was successfully loaded into the interior cavity of the S-15 structure, as demonstrated by characteristic peaks of the Ni12P5 phase at ~49.0°, 47.0°, and 38.4° (Figure 2b) [25]. However, Ru phosphide might exist in an amorphous form [26] due to the lack of signals of any feature peaks, and thus, was denoted as RuPx. In addition, the X-ray diffraction patterns of Ru-Ni bimetallic phosphide (Ru-Ni12P5) closely resembles that of Ni12P5, indicating that Ru-Ni12P5 is likely a composite of RuPx and Ni12P5. Due to the limitations of the incipient wetness impregnation method, P, Ni, and Ru were non-homogeneously distributed on the interior surface of the macropores of the S-15 structure, as displayed in Figure 3a–f. Moreover, the distinct distributions of Ni and Ru further revealed that Ru-Ni12P5 is a composite rather than solid solution.
The surface chemical states of Ru, Ni, and P in Ru-Ni12P5/S-15 are presented in Figure 4. In the Ni12P5 phase, four apparent Ni (2p3/2) peaks centered at 862.3, 858.5, 856.5 and 853.1 eV were assigned to satellite, Ni3+, Ni2+, and Niδ+ species [27,28], respectively, as displayed in Figure 4a. Similar to amorphous RuPx, two Ru (3p3/2) peaks with binding energies of 466.1 and 462.6 eV, respectively, were attributed to Ru4+ and Ruδ+ species [29] (Figure 4b). Previous studies found that these metal species with a very small charge (δ ≈ 0) in metallic phosphides, especially the Ruδ+ species, facilitates hydrogen activation [26,30]. Notably, no apparent electron transfer between Ru and Ni species was detected in Ru-Ni metal phosphide, confirming that Ru-Ni12P5 is a composite.
According to the NH3-TPD results (shown in Figure 5), the acid properties of the catalysts are listed in Table 1. As displayed in Figure 5, only RuPx/S-15 exhibits an apparent NH3 desorption peak at ~450 °C which might be attributed to Lewis acid sites originating from the unreduced Ru species [31]. Due to the low content of Ru, the NH3-TPD curve of Ru-Ni12P5/S-15 is closely similar to that of Ni12P5/S-15. However, the total acidity of Ru-Ni12P5/S-15 as a composite is much higher than those of monometallic phosphides, i.e., Ni12P5/S-15 and RuPx/S-15.

2.2. Hydrogenolysis of Diphenyl Ether

2.2.1. Effect of a Porous Structure on Catalytic Performance

The activity of the catalyst Ni12P5/S-1 for depolymerization of diphenyl ether with hydrogenolysis conversion was negligible at only 0.5% (Table 2, entry 1). As compared to Ni12P5/S-1, the catalytic reactivity of RuPx/S-1 for hydrogenolysis of diphenyl ether was much higher with a conversion rate of 66.5% (Table 2, entry 2). However, when employing Ru-Ni12P5/S-1 as a catalyst, the hydrogenolysis conversion rate and benzene yield were 94.7% and 40.5%, respectively (Table 2, entry 3). These results indicate that Lewis acid sites, i.e., unreducted Ru species (as illustrated by NH3-TPD results in Figure 5), play a pivotal role in the hydrogenolysis of C−O linkages.
Importantly, the macroporous structure of a support also has a significant influence on hydrogenolysis (Table 2, entries 3–6). A previous study illustrated that catalysts with larger pores favor the hydrodepolymerization of kraft lignin in bio-oil production [32]. The use of the highly ordered macroporous array S-15 as a support afforded the maximum hydrogenolysis conversion rate and benzene yield of 99.4% and 57.2%, respectively (Table 2, entry 6) and resulted in a minimum yield of oxygen-containing products, including phenol, cyclohexanol, and cyclohexanone. The interconnected macroporous array with pore diameters of ~200 nm promotes mass transport [33,34,35] and acts as a “fishing net” that captures lignin macromolecules in the liquid phase. Subsequently, the captured lignin macromolecules moved into the interior of the macropores and were hydrogenolyzed into smaller molecular products. Notably, even though S-1 has a larger total surface area than S-15, the corresponding supported Ru-Ni12P5 exhibited distinct reactivity for the catalytic hydrogenolysis of diphenyl ether (Table 2, entries 3 and 6). This phenomenon is supposed to be related to pore confined-like effect, as illustrated in Figure 6. It can be seen that the concentration of active hydrogen species in the interior of the macroporous array S-15 was greater than that on the surface of S-1, which might account for the preeminent hydrodeoxygenation selectivity of Ru-Ni12P5/S-15. Moreover, the highly ordered macroporous array S-15 also facilitated hydrogenation with a maximum cyclohexane yield of 13.8%.

2.2.2. Effect of H2 Partial Pressure on Catalytic Performance

The initial H2 partial pressure had a significant effect on the hydrogenolysis of performance (Figure 7). The hydrogenolysis conversion of diphenyl ether was merely 16.9% under an initial H2 partial pressure of 0.2 MPa, but increased to 99.2% at 0.6 MPa, while the benzene yield achieved a maximum value of 57.2%. However, the phenol and cyclohexanol yields sharply decreased to 1.3%, while the cyclohexane yield increased to 13.8%. As the initial partial pressure was continuously increased to 1.0 MPa, the hydrogenolysis conversion rate and cyclohexane yield remained nearly unchanged, whereas the benzene yield dropped dramatically to 34.5%, likely due to subsequent hydrodecyclization of benzene-derived hydrogenation products to afford paraffin (C1-C6) [36], which is less soluble in the aqueous phase. These results demonstrate that increasing the H2 partial pressure did not accelerate the hydrogenolysis reaction [26].

2.3. Reaction Mechanism

To obtain further insights into the hydrogenolysis mechanism, 3-phenoxytoluene was selected as a probe compound. As shown in Figure 8, metal phosphides, as a bifunctional catalyst, have excellent potential for cleaving C–O linkages [37,38]. Over metal phosphides, H2 can be heterolytically split into H and H+ species [39], which become subsequently trapped by Lewis acid sites (e.g., Ruδ+ and Niδ+) [40,41], while the negatively charged P and PO species form Brønsted sites (PO–H) [40,42,43] are notably much less active than Lewis acid sites in hydrogenolysis and hydrodeoxygenation [40,44], indicating that H species are essential for C–O cleavage. Essentially, the 4–O–5 linkage is not easily cleaved, even in thermal strong acid solution [45]. However, due to the complexity of the chemical reaction, there is presently no clear atomistic description for the hydrogenolysis of C–O linkages on metal phosphides. Even so, we centered our endeavors on the reaction pathways that were consistent with the catalytic experiments, while not denying the existence of other complex reaction pathways.
Accordingly, possible reaction pathways for the catalytic hydrogenolysis of 3-phenoxytoluene over Ru-Ni12P5/S-15 are proposed in Scheme 1. As demonstrated by the time profile of the product evolution presented in Figure 8, during hydrogenolysis, route 2 is thermodynamically favorable as compared to route 1. In route 2, 3-phenoxytoluene, an unsymmetrical diaryl ether, is preferentially initially hydrogenolyzed by H attacking substituted aromatic carbon atom at the electron-deficient benzene-ring side with Lewis sites (e.g., unreducted Ru species) assistance [46], resulting in the generation of intermediate im22 and benzene via intermediate im2. Subsequently, intermediate im22 abstracts H+ from Brønsted sites or other H-donor, such as H2O, leading to the formation of m-cresol. As a dominant intermediate product, m-cresol was sequentially dehydroxylated in the formation of toluene via the thermodynamically favorable intermediate im23, which resulted from H attacking hydroxyl-substituted aromatic carbon atom. Concomitantly, a slight amount of m-cresol was hydrogenated to 3-methylcyclohexanol as a by-product.

2.4. Catalytic Hydrogenolysis of Other Model Compounds

Encouraged by the efficient depolymerization of diphenyl ether and 3-phenoxytoluene, hydrogenolysis of other C–O linkages containing lignin-related compounds over Ru-Ni12P5/S-15 was evaluated. As presented in Table 3, similar to 3-phenoxytoluene, 4,4’-oxybis(methylbenzene) was depolymerized via the direct hydrogenolysis of aromatic ether linkages, resulting in the formation of toluene and p-cresol without detectable hydrogenation products (Table 3, entry 1). Phenol, a vital intermediate resulting from hydrogenolysis of aryl ether linkages, was efficiently dehydroxylated and hydrogenolyzed despite the bond dissociation energy of the C–O bond as high as 463.6 ± 4.2 kJ/mol [47] (Table 3, entry 2). As shown in entries three–five in Table 3, all of the probe compounds were directly hydrogenolyzed via selective cleavage of α–C–O linkages, which all resulted in toluene as the most dominant product. In addition to toluene, phenylmethanol and phenol were suspected products from the direct hydrogenolysis of dibenzyl ether and benzyl phenyl ether, respectively. However, successive dehydroxylation from phenol led to the formation of benzene with a yield of 20.3%, while that from benzyl alcohol afforded more toluene, reaching a maximum yield of 88.0%.

2.5. Catalytic Hydrogenolysis of Lignin by Ru-Ni12P5/S-15

2D-HSQC-NMR spectroscopy was employed to investigate changes to the C–O linkages of lignin and any corresponding residue [48,49,50]. As displayed in Figure 9a, the strong signals of –OCH3CH = (3.7–4.0)/(53.2–55.8) ppm) illustrate that lignin contains large amounts of methoxyaryl moieties. Moreover, obvious correlations among Cα−Hα, Cβ−Hβ, and Cγ−HγCH = 61.3/3.5, (77.1–79.2)/(4.4–4.5), and (58.0–60.0)/3.9 ppm, respectively) demonstrate that A is the predominant structural unit of lignin rather than B and C. After three hours of hydrogenolysis, the signals of A, B, and C had disappeared and that of the methoxyl group became weaker, as shown in Figure 9b. After six hours of hydrogenolysis, the signals of the methoxyl groups finally vanished, indicating that the C–O linkages of lignin had been completely cleaved (Figure 9c).
The hydrogenolysis products of lignin, as determined by HPLC-TQMS, are listed in Table 4. Notably, the yields of almost all products had increased over an initial six hours, revealing that cleavage of the ether linkages of lignin was incomplete for the initial three hours over Ru-Ni12P5/S-15, in accordance with the results exhibited in Figure 9. When further prolonging the hydrogenolysis time to 12 h, the yields of all methoxyl-containing products remarkably decreased to zero, whereas those of 4-methylcatechol and 3-methylcatechol markedly increased to 103.6 and 25.3 μg, respectively. These results suggest that polyphenolics might be produced by hydrogenolysis of the methoxyl-containing moieties of lignin. Moreover, these findings demonstrate that different higher value-added compounds, such as vanillin and 4-methylcatechol, can be obtained by facilely tuning the reaction conditions. Vanillin is an important pharmaceutical intermediate with antiepileptic and antibacterial activities that is used for the synthesis of drugs for treatment of hypertension and heart disease [51]. Acting as a hapten and an antimicrobial as well as an antioxidant, 4-methylcatechol has a strong antiplatelet effect that is useful for the treatment of diabetic neuropathy and tumors [52,53].

3. Materials and Methods

3.1. Chemicals and Reagents

Tetrapropylammonium hydroxide (TPAOH; 25 wt% in water), RuCl3·xH2O (Ru, 36.0–40.0%), and dealkaline lignin were obtained from Macklin Biochemical Co., Ltd.. Sodium citrate dihydrate (99.0%), tetraethyl orthosilicate (>99%), bibenzyl ether (95.0%), phenol (99.0%), anisole (99.0%), diphenyl ether (≥99.9%), benzyl methyl ether (99.0%), ethyl acetate (99.5%), 3-phenoxytoluene (98.0%), and dodecane (>99.0%) were got from Aladdin Biochemical Technology Co., Ltd. Ni(NO3)2·6H2O was obtained from Oukai Chemical Co., Ltd., benzyl phenyl ether (98.0%) from Energy Chemicals Co. Ltd., and di-p-tolyl ether (99.0%) from Alfa Chemical Co., Ltd.

3.2. Synthesis of S-1 and Macroporous S-1

S-1 and macroporous S-1 were prepared using a previously reported procedure with slight modifications [24]. In brief, tetraethyl orthosilicate was vigorously mixed with TPAOH (with a molar composition of 1 SiO2:0.2 TPAOH:37 H2O) until the formation of a clear solution. Afterward, the solution was heated in a 100 mL polytetrafluoroethylene-lined autoclave at 453 K for 3 days. After cooling to ~300 K, the solid precipitate was centrifuged and washed, then dried at 383 K, followed by calcination at 823 K to remove the templates. Finally, a white powdered product (S-1) was obtained.
Then, S-1 (0.4 g) was added to 4 mL of alkaline solution (TPAOH + NaCl) and heated in a polytetrafluoroethylene-lined steel autoclave at 453 K while stirring for 3–5 days. The solid was obtained by filtration, thoroughly washed, and then dried at 383 K. Finally, the resulting product was calcined at 823 K for 6 h to obtain macroporous S-1x, where x denotes recrystallization time. For example, “S-15” indicates that the macroporous S-1 was synthesized by recrystallization of solid S-1 at 453 K for 5 days.

3.3. Preparation of Catalysts

In brief, S-1 zeolite or macroporous S-1 was impregnated with a solution of NaH2PO2, RuCl3·xH2O, NiCl2·6H2O, and sodium citrate, then pyrolysized at 773 K under a vacuum after drying at 383 K. Finally, the catalysts were passivated in N2 at room temperature. For each catalyst, the molar ratios of nP/nM (M = Ru and/or Ni) and nNi/nRu were 6/1 and the total load was 10 wt%.

3.4. Characterization of Catalysts and Supports

X-ray diffraction patterns were obtained with a SmartLab SE instrument. X-ray photoelectron spectra were obtained using an ESCALAB instrument. Energy-dispersive X-ray spectroscopy and transmission electron microscopy (TEM) images were obtained using a Tecnai™ G2 F20 transmission electron microscope.

3.5. Hydrogenolysis of Model Compounds

In a typical hydrogenolysis reaction, fresh catalyst (0.003 g), a model compound (0.2 mmol), and water (5 mL) were put into a reactor, which was then charged with 0.4 MPa N2 and 0.6 MPa H2 after exclusion of air. Reactions were conducted at 250 °C for a predetermined time while vigorously stirring. The reaction was terminated by cooling the reactor to ambient temperature using ice water. Ethyl acetate was used to retrieve the organic products from the reaction mixture. Quantitative analysis of the organic products was performed on a GC-FID employing n-dodecane as an internal standard. The hydrogenolysis conversion rate and product yield were calculated using the following equations.
Conversion   of   reactant % = moles   of   reactant   reacted   moles   of   reactant   supplied × 100
Yield   of   product   x % = moles   of   C   atoms   in   product   x moles   of   C   atoms   in   reactant × 100

3.6. Hydrogenolysis of Lignin

Rather severe reaction conditions were employed due to the less depolymerization reactivity of lignin than model compounds. Typically, the catalyst (0.1 g), dealkaline lignin (0.1 g), and water (5 mL) were put into a reactor, which was sequentially charged with 10 bar H2 after exclusion of air with H2. Thereafter, the reaction was conducted at 280 °C for 3–12 h with vigorous stirring. After termination of the reaction, the hydrogenolysis products were retrieved with ethyl acetate. The liquid products in the aqueous and organic phases were analyzed quantitatively by HPLC-TQMS with the external standard method. The corresponding residues were characterized by two-dimensional nuclear magnetic resonance heteronuclear single quantum coherence (2D-NMR-HSQC) spectroscopy.

4. Conclusions

In this study, a macroporous S-1 array-supported Ru-Ni metallic phosphide composite was fabricated and introduced for the hydrogenolysis of lignin and C–O linkages containing lignin-related compounds. The unreduced Ru species-originated Lewis acid sites and the geometrical structure of the macroporous support significantly influenced the hydrogenolysis conversion rate and product selectivity. Notably, both the aryl ether and aliphatic ether linkages favor direct hydrogenation in the initiation reaction over Ru-Ni12P5/S-15. Furthermore, 2D-HSQC-NMR spectroscopy demonstrated that Ru-Ni12P5/S-15 efficiently depolymerized the C–O linkages of lignin. Different higher value-added products suitable for pharmaceutical production can be obtained by facilely tuning the reaction conditions.

Author Contributions

Conceptualization, B.C. and Z.-J.D.; methodology, Z.-Z.C. and B.C.; validation, Z.-Z.C. and L.-Q.H.; formal analysis, H.Y. and J.-M.H.; investigation, Z.-Z.C.; resources, Z.-Z.C. and B.C.; data curation, Z.-Z.C. and S.-J.Z.; writing—original draft preparation, Z.-Z.C.; writing—review and editing, B.C. and Z.-J.D.; supervision, B.C. and Z.-J.D.; project administration, B.C. and Z.-J.D.; funding acquisition, B.C. and Z.-J.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research work was supported by the Fund from National Natural Science Foundation of China (Grant 21875186), Natural Science Basic Research Plan in Shaanxi Province of China (Grant 2019JM-259), Foundation of State Key Laboratory of High-efficiency Utilization of Coal and Green Chemical Engineering (Grant 2019-KF-17), China Postdoctoral Science Foundation (Grant 2017M623205) and Special Research Foundation of Education Bureau of Shaanxi Province (Grant 15JK1692).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Alonso, D.M.; Wettstein, S.G.; Dumesic, J.A. Bimetallic catalysts for upgrading of biomass to fuels and chemicals. Chem. Soc. Rev. 2012, 41, 8075–8898. [Google Scholar] [CrossRef] [PubMed]
  2. Sahoo, A.; Saini, K.; Jindal, M.; Bhaskar, T.; Pant, K.K. Co-hydrothermal liquefaction of algal and lignocellulosic biomass: Status and perspectives. Bioresour. Technol. 2021, 342, 125948. [Google Scholar] [CrossRef] [PubMed]
  3. Mohan, D.; Pittman, C.U.; Steele, P.H. Pyrolysis of wood/biomass for bio-oil: A critical review. Energy Fuels 2006, 20, 848–889. [Google Scholar] [CrossRef]
  4. Wang, H.; Male, J.; Wang, Y. Recent advances in hydrotreating of pyrolysis bio-oil and its oxygen-containing model compounds. ACS Catal. 2013, 3, 1047–1070. [Google Scholar] [CrossRef]
  5. Tuck, C.O.; Pérez, E.; Horváth, I.; Sheldon, R.; Poliakoff, M. Valorization of biomass: Deriving more value from waste. Science 2012, 337, 695–699. [Google Scholar] [CrossRef]
  6. Zhao, M.X.; Wei, X.Y.; Qu, M.; Kong, J.; Li, Z.K.; Liu, J.; Zong, Z.M. Complete hydrocracking of dibenzyl ether over a solid acid under mild conditions. Fuel 2016, 183, 531–536. [Google Scholar] [CrossRef]
  7. Luo, H.; Wang, L.; Li, G.; Shang, S.; Lv, Y.; Niu, J.; Gao, S. Nitrogen-doped carbon-modified cobalt-nanoparticle-catalyzed oxidative cleavage of lignin β-O-4 model compounds under mild conditions. ACS Sustain. Chem. Eng. 2018, 6, 14188–14196. [Google Scholar] [CrossRef]
  8. Toor, S.S.; Reddy, H.; Deng, S.; Hoffmann, J.; Spangsmark, D.; Madsen, L.B.; Holm-Nielsen, J.B.; Rosendahl, L.A. Hydrothermal liquefaction of Spirulina and nannochloropsis salina under subcritical and supercritical water conditions. Bioresour. Technol. 2013, 131, 413–419. [Google Scholar] [CrossRef]
  9. Liu, L.; Zhao, C.; Zheng, F.; Deng, D.; Anderson, M.A. Three-dimensional electrode design with conductive fibers and ordered macropores for enhanced capacitive deionization performance. Desalination 2021, 498, 114794. [Google Scholar] [CrossRef]
  10. Mei, X.; Xiong, J.; Wei, Y.; Zhang, Y.; Zhang, P.; Yu, Q.; Zhao, Z.; Liu, J. High-efficient non-noble metal catalysts of 3D ordered macroporous perovskite-type La2NiB’O6 for soot combustion: Insight into the synergistic effect of binary Ni and B’ sites. Appl. Catal. B Environ. 2020, 275, 119108. [Google Scholar] [CrossRef]
  11. Zhang, P.; Mei, X.; Zhao, X.; Xiong, J.; Li, Y.; Zhao, Z.; Wei, Y. Boosting catalytic purification of soot particles over double Perovskite-Type La2−xKxNiCoO6 catalysts with an ordered macroporous structure. Environ. Sci. Technol. 2021, 55, 11245–11254. [Google Scholar] [CrossRef] [PubMed]
  12. Luo, Z.; Kong, J.; Ma, B.; Wang, Z.; Huang, J.; Zhao, C. Liquefaction and hydrodeoxygenation of polymeric lignin using a hierarchical Ni microreactor catalyst. ACS Sustain. Chem. Eng. 2020, 8, 2158–2166. [Google Scholar] [CrossRef]
  13. Golubeva, M.A.; Zakharyan, E.M.; Maximov, A.L. Transition metal phosphides (Ni, Co, Mo, W) for hydrodeoxygenation of biorefinery products (a review). Pet. Chem. 2020, 60, 1109–1128. [Google Scholar] [CrossRef]
  14. Li, H.; Li, G. Enhancing activity of Ni2P-based catalysts by a Yolk–Shell structure and transition metal-doping for catalytic transfer hydrogenation of vanillin. Energy Fuels 2021, 35, 4158–4168. [Google Scholar] [CrossRef]
  15. Zhao, H.Y.; Li, D.; Bui, P.; Oyama, S.T. Hydrodeoxygenation of guaiacol as model compound for pyrolysis oil on transition metal phosphide hydroprocessing catalysts. Appl. Catal. A Gen. 2011, 391, 305–310. [Google Scholar] [CrossRef]
  16. García-Pérez, D.; Alvarez-Galvan, M.C.; Capel-Sanchez, M.C.; Blanco-Brieva, G.; Morales-delaRosa, S.; Campos-Martin, J.M.; Fierro, J.L.G. Influence of bimetallic characteristics on the performance of MoCoP and MoFeP catalysts for methyl laurate hydrodeoxygenation. Catal. Today 2021, 367, 43–50. [Google Scholar] [CrossRef]
  17. Bonita, Y.; O’Connell, T.P.; Miller, H.E.; Hicks, J.C. Revealing the hydrogenation performance of RuMo phosphide for chemoselective reduction of functionalized aromatic hydrocarbons. Ind. Eng. Chem. Res. 2019, 58, 3650–3658. [Google Scholar] [CrossRef]
  18. Bonita, Y.; Jain, V.; Geng, F.; O’Connell, T.P.; Wilson, W.N.; Rai, N.; Hicks, J.C. Direct synthesis of furfuryl alcohol from furfural: Catalytic performance of monometallic and bimetallic Mo and Ru phosphides. Catal. Sci. Technol. 2019, 9, 3656–3668. [Google Scholar] [CrossRef]
  19. Charisiou, N.D.; Siakavelas, G.I.; Papageridis, K.N.; Motta, D.; Dimitratos, N.; Sebastian, V.; Polychronopoulou, K.; Goula, M.A. The effect of noble metal (M: Ir, Pt, Pd) on M/Ce2O3-γ-Al2O3 catalysts for hydrogen production via the steam reforming of glycerol. Catalysts 2020, 10, 790. [Google Scholar] [CrossRef]
  20. Gao, K.; Sahraei, O.A.; Iliuta, M.C. Development of residue coal fly ash supported nickel catalyst for H2 production via glycerol steam reforming. Appl. Catal. B-Environ. 2021, 298, 119958. [Google Scholar] [CrossRef]
  21. Ma, Y.X.; Ma, Y.Y.; Li, J.J.; Ye, Z.M.; Hu, X.; Dong, D.H. Electrospun nanofibrous Ni/LaAlO3 catalysts for syngas production by high temperature methane partial oxidation. Int. J. Hydrogen Energy 2022, 47, 3867–3875. [Google Scholar] [CrossRef]
  22. Charisiou, N.D.; Siakavelas, G.; Tzounis, L.; Sebastian, V.; Monzon, A.; Baker, M.A.; Hinder, S.J.; Polychronopoulou, K.; Yentekakis, I.V.; Goula, M.A. An in depth investigation of deactivation through carbon formation during the biogas dry reforming reaction for Ni supported on modified with CeO2 and La2O3 zirconia catalysts. Int. J. Hydrogen Energy 2018, 43, 18955–18976. [Google Scholar] [CrossRef]
  23. Yang, X.Y.; Tian, G.; Chen, L.H.; Li, Y.; Rooke, J.C.; Wei, Y.X.; Liu, Z.M.; Deng, Z.; Tendeloo, G.V.; Su, B.L. Well-organized zeolite nanocrystal aggregates with interconnected hierarchically micro–meso–macropore systems showing enhanced catalytic performance. Chem. A Eur. J. 2011, 17, 14987–14995. [Google Scholar] [CrossRef]
  24. Dai, C.; Zhang, A.; Li, L.; Hou, H.; Ding, F.; Li, J.; Mu, D.; Song, C.; Liu, M.; Guo, X. Synthesis of hollow nanocubes and macroporous monoliths of silicalite-1 by alkaline treatment. Chem. Mater. 2013, 25, 4197–4205. [Google Scholar] [CrossRef]
  25. Larsson, E. An x-ray investigation of the Ni-P system and the crystal structures of NiP and NiP2. Arkiv Kemi 1965, 23, 335–365. [Google Scholar]
  26. Diao, Z.J.; Huang, L.Q.; Chen, B.; Gao, T.; Cao, Z.Z.; Ren, X.D.; Zhao, S.J.; Li, S. Amorphous Ni-Ru bimetallic phosphide composites as efficient catalysts for the hydrogenolysis of diphenyl ether and lignin. Fuel 2022, 324, 124489. [Google Scholar] [CrossRef]
  27. Wang, S.; Hu, J.; Gui, X.; Lin, S.; Tu, Y. RuO2 active particles supported on Ni12P5 as excellent electrocatalysts for LiO2 batteries. Solid State Ionics 2021, 372, 115773. [Google Scholar] [CrossRef]
  28. Wang, Y.; Qiao, M.; Li, Y.; Wang, S. Tuning surface electronic configuration of NiFe LDHs nanosheets by introducing cation vacancies (Fe or Ni) as highly efficient electrocatalysts for oxygen evolution reaction. Small 2018, 14, 1800136. [Google Scholar] [CrossRef]
  29. Morgan, D.J. Resolving ruthenium: XPS studies of common ruthenium materials. Surf. Interface Anal. 2015, 47, 1072–1079. [Google Scholar] [CrossRef]
  30. Topalian, P.J.; Carrillo, B.A.; Cochran, P.M.; Takemura, M.F.; Bussell, M.E. Synthesis and hydrodesulfurization properties of silica-supported nickel-ruthenium phosphide catalysts. J. Catal. 2021, 403, 173–180. [Google Scholar] [CrossRef]
  31. Huang, L.-Q.; Diao, Z.-J.; Chen, B.; Du, Q.-P.; Duan, K.-Y.; Zhao, S.-J. Hydrogenolysis of lignin and C–O linkages containing lignin-related compounds over an amorphous CoRuP/SiO2 catalyst. Catalysts 2022, 12, 1328. [Google Scholar] [CrossRef]
  32. Wang, Y.; Chen, M.; Shi, J.; Zhang, J.; Li, C.; Wang, J. Catalytic depolymerization of kraft lignin for liquid fuels and phenolic monomers over molybdenum-based catalysts: The effect of supports. J. Fuel Chem. Technol. 2021, 49, 1922–1934. [Google Scholar] [CrossRef]
  33. Rolison, D.R. Catalytic nanoarchitectures—The importance of nothing and the unimportance of periodicity. Science 2003, 299, 1698–1701. [Google Scholar] [CrossRef]
  34. Wang, A.; Shi, Y.; Yang, L.; Fan, G.; Li, F. Ordered macroporous Co3O4-supported Ru nanoparticles: A robust catalyst for efficient hydrodeoxygenation of anisole. Catal. Commun. 2021, 153, 106302. [Google Scholar] [CrossRef]
  35. Yang, P.; Zhou, S.; Du, Y.; Li, J.; Lei, J. Synthesis of ordered meso/macroporous H3PW12O40/SiO2 and its catalytic performance in oxidative desulfurization. RSC Adv. 2016, 6, 53860–53866. [Google Scholar] [CrossRef]
  36. Molinari, V.; Giordano, C.; Antonietti, M.; Esposito, D. Titanium nitride-nickel nanocomposite as heterogeneous catalyst for the hydrogenolysis of aryl ethers. J. Am. Chem. Soc. 2014, 136, 1758–1761. [Google Scholar] [CrossRef]
  37. Yu, Z.; Wang, A.; Liu, S.; Yao, Y.; Sun, Z.; Li, X.; Liu, Y.; Wang, Y.; Camaioni, D.M.; Lercher, J.A. Hydrodeoxygenation of phenolic compounds to cycloalkanes over supported nickel phosphides. Catal. Today 2019, 319, 48–56. [Google Scholar] [CrossRef]
  38. Gutiérrez-Rubio, S.; Berenguer, A.; Přech, J.; Opanasenko, M.; Ochoa-Hernández, C.; Pizarro, P.; Čejka, J.; Serrano, D.P.; Coronado, J.M.; Moreno, I. Guaiacol hydrodeoxygenation over Ni2P supported on 2D-zeolites. Catal. Today 2020, 345, 48–58. [Google Scholar] [CrossRef]
  39. Nelson, R.C.; Baek, B.; Ruiz, P.; Goundie, B.; Brooks, A.; Wheeler, M.C.; Frederick, B.G.; Grabow, L.C.; Austin, R.N. Experimental and theoretical insights into the hydrogen-efficient direct hydrodeoxygenation mechanism of phenol over Ru/TiO2. ACS Catal. 2015, 5, 6509–6523. [Google Scholar] [CrossRef]
  40. Li, K.; Wang, R.; Chen, J. Hydrodeoxygenation of anisole over silica-supported Ni2P, MoP, and NiMoP catalysts. Energy Fuels 2011, 25, 854–863. [Google Scholar] [CrossRef]
  41. Shi, Y.; Zhang, B. Recent advances in transition metal phosphide nanomaterials: Synthesis and applications in hydrogen evolution reaction. Chem. Soc. Rev. 2016, 45, 1529–1541. [Google Scholar] [CrossRef] [PubMed]
  42. Rodríguez-Aguado, E.; Rodríguez-Aguado, A.; Cecilia, J.A.; Ballesteros-Plata, D.; López-Olmo, R.; Rodríguez-Castellón, E. CoxPy catalysts in HDO of phenol and dibenzofuran: Effect of P content. Top. Catal. 2017, 60, 1094–1107. [Google Scholar] [CrossRef]
  43. Cecilia, J.A.; Infantes-Molina, A.; Sanmartín-Donoso, J.; Rodríguez-Aguado, E.; Ballesteros-Plata, D.; Rodríguez-Castellón, E. Enhanced HDO activity of Ni2P promoted with noble metals. Catal. Sci. Technol. 2016, 6, 7323–7333. [Google Scholar] [CrossRef]
  44. Galindo-Ortega, Y.L.; Infantes-Molina, A.; Huirache-Acuña, R.; Barroso-Martín, I.; Rodríguez-Castellón, E.; Fuentes, S.; Alonso-Nuñez, G.; Zepeda, T.A. Active ruthenium phosphide as selective sulfur removal catalyst of gasoline model compounds. Fuel Process. Technol. 2020, 208, 106507. [Google Scholar] [CrossRef]
  45. Zhao, C.; Lercher, J.A. Upgrading pyrolysis oil over Ni/HZSM-5 by cascade reactions. Angew. Chem. Int. Ed. 2012, 51, 5935–5940. [Google Scholar] [CrossRef]
  46. Sergeev, A.G.; Hartwig, J.F. Selective, nickel-catalyzed hydrogenolysis of aryl ethers. Science 2011, 332, 439–443. [Google Scholar] [CrossRef] [Green Version]
  47. Luo, Y.R. Handbook of Bond Dissociation Energies in Organic Compounds; CRC: New York, NY, USA, 2003. [Google Scholar]
  48. Si, X.; Chen, J.; Lu, F.; Liu, X.; Ren, Y.; Lu, R.; Jiang, H.; Liu, H.; Miao, S.; Zhu, Y.; et al. Immobilized Ni clusters in mesoporous aluminum silica nanospheres for catalytic hydrogenolysis of lignin. ACS Sustain. Chem. Eng. 2019, 7, 19034–19041. [Google Scholar] [CrossRef]
  49. Kong, J.; Li, L.; Zeng, Q.; Long, J.; He, H.; Wang, Y.; Liu, S.; Li, X. Production of 4-Ethylphenol from lignin depolymerization in a novel Surfactant-Free Microemulsion reactor. Ind. Eng. Chem. Res. 2021, 60, 17897–17906. [Google Scholar] [CrossRef]
  50. Amiri, M.T.; Bertella, S.; Questell-Santiago, Y.M.; Luterbacher, J.S. Establishing lignin structure-upgradeability relationships using quantitative 1H–13C heteronuclear single quantum coherence nuclear magnetic resonance (HSQC-NMR) spectroscopy. Chem. Sci. 2019, 10, 8135–8142. [Google Scholar] [CrossRef] [Green Version]
  51. Wang, X.B.; Du, L.D.; Wang, S.M.; Du, G.H. Vanillin. In Natural Small Molecule Drugs from Plants; Springer: Singapore, 2018; pp. 343–346. [Google Scholar]
  52. Applová, L.; Karlíčková, J.; Warncke, P.; Macáková, K.; Hrubša, M.; Macháček, M.; Tvrdý, V.; Fischer, D.; Mladěnka, P. 4-Methylcatechol, a flavonoid metabolite with potent antiplatelet effects. Mol. Nutr. Food Res. 2019, 63, 1900261. [Google Scholar] [CrossRef]
  53. Morita, K.; Arimochi, H.; Ohnishi, Y. In vitro cytotoxicity of 4-Methylcatechol in murine tumor cells: Induction of apoptotic cell death by extracellular pro-oxidant action. J. Pharmacol. Exp. Ther. 2003, 306, 317–323. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Scanning electron microscopy images of (a) S-1, (b) S-13, (c) S-14 and (d,e) S-15; and (f) TEM image of S-15.
Figure 1. Scanning electron microscopy images of (a) S-1, (b) S-13, (c) S-14 and (d,e) S-15; and (f) TEM image of S-15.
Catalysts 12 01625 g001
Figure 2. (a) TEM-EDS results of Ru-Ni12P5/S-15 and (b) XPD patterns of catalysts.
Figure 2. (a) TEM-EDS results of Ru-Ni12P5/S-15 and (b) XPD patterns of catalysts.
Catalysts 12 01625 g002
Figure 3. (af) TEM mapping of Ru-Ni12P5/S-15.
Figure 3. (af) TEM mapping of Ru-Ni12P5/S-15.
Catalysts 12 01625 g003
Figure 4. X-ray photoelectron spectrum of Ru-Ni12P5/S-15 in (a) Ni 2p and (b) Ru 3p regions.
Figure 4. X-ray photoelectron spectrum of Ru-Ni12P5/S-15 in (a) Ni 2p and (b) Ru 3p regions.
Catalysts 12 01625 g004
Figure 5. NH3-TPD profiles of the catalysts.
Figure 5. NH3-TPD profiles of the catalysts.
Catalysts 12 01625 g005
Figure 6. Schematic illustration of hydrogenolysis reactivity under a pore confined-like effect.
Figure 6. Schematic illustration of hydrogenolysis reactivity under a pore confined-like effect.
Catalysts 12 01625 g006
Figure 7. Effects of initial H2 partial pressure on the hydrogenolysis of diphenyl ether over Ru-Ni12P5/S-15.
Figure 7. Effects of initial H2 partial pressure on the hydrogenolysis of diphenyl ether over Ru-Ni12P5/S-15.
Catalysts 12 01625 g007
Figure 8. Time profiles of hydrogenolysis of 3-phenoxytoluene over Ru-Ni12P5/S-15. CHX-derivs: Cyclohexene-derivatives, including cyclohexane, cyclohexanol, and cyclohexanone; MCH-derivs: methylcyclohexane-derivatives, including methylcyclohexane, 3-methylcyclohexanol and 3-methylcyclohexanone.
Figure 8. Time profiles of hydrogenolysis of 3-phenoxytoluene over Ru-Ni12P5/S-15. CHX-derivs: Cyclohexene-derivatives, including cyclohexane, cyclohexanol, and cyclohexanone; MCH-derivs: methylcyclohexane-derivatives, including methylcyclohexane, 3-methylcyclohexanol and 3-methylcyclohexanone.
Catalysts 12 01625 g008
Scheme 1. Proposed pathways for hydrogenolysis of 3-methyl diphenyl ether over Ru-Ni12P5/S-15. L denotes Lewis acid sites.
Scheme 1. Proposed pathways for hydrogenolysis of 3-methyl diphenyl ether over Ru-Ni12P5/S-15. L denotes Lewis acid sites.
Catalysts 12 01625 sch001
Figure 9. 2D-HSQC-NMR spectroscopy of C–O linkages for (a) fresh dealkaline lignin and (b) after three hours and (c) six hours of hydrogenolysis; (d) representative structures of ether linkages for lignin.
Figure 9. 2D-HSQC-NMR spectroscopy of C–O linkages for (a) fresh dealkaline lignin and (b) after three hours and (c) six hours of hydrogenolysis; (d) representative structures of ether linkages for lignin.
Catalysts 12 01625 g009
Table 1. Acid sites analysis by NH3-TPD.
Table 1. Acid sites analysis by NH3-TPD.
SampleNi12P5/S-15RuPx/S-15Ru-Ni12P5/S-15
NH3 acidity/mmol·g−10.190.130.25
Table 2. Hydrogenolysis of diphenyl ether over different catalysts.
Table 2. Hydrogenolysis of diphenyl ether over different catalysts.
EntryCatalystConv. (%)Yield (%)
Catalysts 12 01625 i001Catalysts 12 01625 i002Catalysts 12 01625 i003Catalysts 12 01625 i004Catalysts 12 01625 i005
1Ni12P5/S-10.50.2----
2RuPx/S-166.520.41.317.40.61.2
3Ru-Ni12P5/S-1 a94.740.53.621.83.12.7
4Ru-Ni12P5/S-1397.445.37.012.62.61.7
5Ru-Ni12P5/S-1497.350.28.03.13.11.9
6Ru-Ni12P5/S-15 b99.457.213.81.30.30.2
a SBET(S-1) = 482 m2·g−1 [24]; b SBET(S-15) = 389 m2·g−1 [24].
Table 3. Hydrogenolysis of different model compounds over Ru-Ni12P5/S-15.
Table 3. Hydrogenolysis of different model compounds over Ru-Ni12P5/S-15.
EntrySubstrateConv. (%)Yield (%)
Catalysts 12 01625 i006Catalysts 12 01625 i007Catalysts 12 01625 i008Catalysts 12 01625 i009Catalysts 12 01625 i010Catalysts 12 01625 i011Catalysts 12 01625 i012
1Catalysts 12 01625 i01392.340.812.6-----
2Catalysts 12 01625 i01498.9---35.837.65.40.8
3Catalysts 12 01625 i015100.067.8--3.0---
4Catalysts 12 01625 i016100.042.3-1.620.32.31.51.3
5Catalysts 12 01625 i01799.688.0--9.8---
Table 4. The predominant products from the hydrogenolysis of lignin.
Table 4. The predominant products from the hydrogenolysis of lignin.
ProductsStructureYield (μg)
3 h6 h12 h
VanillinCatalysts 12 01625 i01894.9103.20.0
IsoeugenolCatalysts 12 01625 i01919.332.10.0
AcetovanilloneCatalysts 12 01625 i02019.527.02.9
AcetosyringoneCatalysts 12 01625 i0210.20.10.0
4-VinylguaiacolCatalysts 12 01625 i02212.19.48.2
4-MethylcatecholCatalysts 12 01625 i02312.013.1103.6
4-AcetonylguaiacolCatalysts 12 01625 i0245.87.24.9
3-MethylcatecholCatalysts 12 01625 i0252.29.125.3
2-AcetoxyacetophenoneCatalysts 12 01625 i0260.30.90.0
2,6-DimethoxyphenolCatalysts 12 01625 i0274.78.00.5
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chen, B.; Cao, Z.-Z.; Diao, Z.-J.; Huang, L.-Q.; Zhao, S.-J.; Yuan, H.; He, J.-M. Hydrogenolysis of Lignin and C–O Linkages Containing Lignin-Related Compounds over a Macroporous Silicalite-1 Array-Supported Ru-Ni Phosphide Composite. Catalysts 2022, 12, 1625. https://doi.org/10.3390/catal12121625

AMA Style

Chen B, Cao Z-Z, Diao Z-J, Huang L-Q, Zhao S-J, Yuan H, He J-M. Hydrogenolysis of Lignin and C–O Linkages Containing Lignin-Related Compounds over a Macroporous Silicalite-1 Array-Supported Ru-Ni Phosphide Composite. Catalysts. 2022; 12(12):1625. https://doi.org/10.3390/catal12121625

Chicago/Turabian Style

Chen, Bo, Zhi-Ze Cao, Zhi-Jun Diao, Liang-Qiu Huang, Si-Jia Zhao, Hong Yuan, and Jia-Meng He. 2022. "Hydrogenolysis of Lignin and C–O Linkages Containing Lignin-Related Compounds over a Macroporous Silicalite-1 Array-Supported Ru-Ni Phosphide Composite" Catalysts 12, no. 12: 1625. https://doi.org/10.3390/catal12121625

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