Next Article in Journal
Green Chemistry in Organic Synthesis: Recent Update on Green Catalytic Approaches in Synthesis of 1,2,4-Thiadiazoles
Previous Article in Journal
Photo-Induced Holes Initiating Peroxymonosulfate Oxidation for Carbamazepine Degradation via Singlet Oxygen
 
 
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 an Amorphous CoRuP/SiO2 Catalyst

1
College of Urban and Environment Science, Northwest University, Xi’an 710127, China
2
School of Chemical Engineering, Northwest University, Xi’an 710069, China
3
Xi’an Giant Biological Gene Technology Co., Ltd., Xi’an 710077, China
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(11), 1328; https://doi.org/10.3390/catal12111328
Submission received: 11 September 2022 / Revised: 19 October 2022 / Accepted: 27 October 2022 / Published: 29 October 2022
(This article belongs to the Section Catalysis for Sustainable Energy)

Abstract

:
Efficient depolymerization of C–O linkages is essential for converting lignin into fuels and higher value-added chemicals. In this work, CoRuP/SiO2, an amorphous Ru-Co phosphide composite, was fabricated for the efficient hydrogenolysis of ether linkages. The 4–O–5 and αO–4 linkages containing lignin-related compounds, such as diphenyl ether, benzyl phenyl ether, 3-methyl diphenyl ether, and dibenzyl ether, are selected as representatives of linkages in lignin. Under mild conditions, Ru-containing metallic phosphides have high-performance for the catalytic depolymerization of C–O linkages. Compared with other catalysts, CoRuP/SiO2 shows an outstanding selectivity for benzene and excellent efficiency in depolymerizing diphenyl ethers, yielding only a small amount of by-products. Furthermore, the total acidity shows a linear relationship with the hydrogenolysis reactivity in cleaving aromatic ether bonds. The mechanisms for the catalytic hydrogenolysis of 4–O–5 and αO–4 bonds over CoRuP/SiO2 are proposed. Moreover, two-dimensional heteronuclear single quantum coherence nuclear magnetic resonance spectroscopic analysis demonstrates that CoRuP/SiO2 could effectively depolymerize C–O bonds of lignin. These dominant hydrogenolysis products from lignin have excellent potential in the production of high value-added drugs or pharmaceutical intermediates. The hydrogenolysis of lignin can be a highly efficient alternative to the existing method of lignin utilization.

1. Introduction

As a result of the increasing depletion of fossil fuels, the environmental pollution and energy crisis are posing increasingly problems worldwide [1,2]. A promising approach to address these issues is to convert renewable and abundant biomass resources into high grade fuels and value-added chemicals [3,4,5]. Biopolymers rich in natural phenolic substances, i.e., lignin, meet these essential criteria [6,7,8].
Network polymer-like lignin consists of phenylpropane-related structural units connected by bridge bonds [9], and over 80% of the bonds are etheric. Thus, the efficient depolymerization of C–O bonds is essential for a higher value-added utilization of lignin [10].
Several strategies, including pyrolysis, hydrolysis, oxidation, and hydrogenolysis, have been proposed to depolymerize lignin for biodiesel and chemical production. Among these strategies, hydrogenolysis has received increasing interest due to its techno-economic feasibility [11,12].
Depolymerization of the 4–O–5 bond is a highly challenging issue because of its weak reactivity compared to βO–4 and αO–4 bonds [13]. Thus, diphenyl ether is extensively employed as a lignin-related probe compound to explore the depolymerization behaviour of the 4–O–5 bond.
For lignin conversion, transition metal oxides [14,15], phosphides [16,17,18], carbides [19,20], nitrides [21], and sulfides [22,23] are commonly used as catalysts. Specifically, transition metal phosphides could efficiently hydrodeoxygenate bio-oils due to their superior anti-poisoning properties [24]. In addition, the reactivity of monometal phosphides can be significantly improved by doping with precious metals [25,26]. Importantly, metal phosphides with both metal and acidic sites have excellent potential in C–O cleavage [27,28].
In this study, amorphous Ru-Co metallic phosphides supported on gas-phase SiO2 are fabricated and their catalytic activity is evaluated using the 4–O–5 and αO–4 linkages containing lignin-related compounds and lignin. The results show that Ru-Co metallic phosphide has an excellent selectivity to benzene. Most importantly, the total acidity correlates linearly with the hydrogenolysis yield (the sum of the phenol and benzene products). The mechanisms for the CoRuP/SiO2 catalyzed hydrogenolysis of the 4–O–5 and αO–4 bonds are proposed. Moreover, two-dimensional (2D) heteronuclear single quantum coherence (HSQC) nuclear magnetic resonance (NMR) spectroscopic analysis proves that CoRuP/SiO2 could efficiently depolymerize etheric linkages in lignin.

2. Results and Discussion

2.1. Characterization of Catalysts

The broad diffraction hump located at ~22° observed for each catalyst is attributable to SiO2 (Figure 1). However, no apparent diffraction peaks assignable to the metal phosphide were detected for all catalysts, indicating that the metal phosphide was highly dispersed or present in an amorphous form on SiO2.
In accordance with the XRD results, the selected region electron diffraction pattern (Figure 2a) validates CoRuP/SiO2 exists in amorphous structure. Moreover, the elemental distribution showed that Co, Ru, and P were highly dispersed on SiO2 (Figure 2b). Previous studies have proven that the amorphous catalysts possess excellent texture structures, high catalytic activity and better dispersion [29,30].
The pore size distributions and N2 adsorption-desorption isotherms of the gas-phase SiO2 and its supported catalysts are presented in Figure 3a,b, respectively. All catalysts displayed kind IV isotherms with a H1-type hysteresis loop, which suggested the presence of uniform sized and shaped mesopores in the catalysts. Moreover, a type IV isotherms was displayed by the gas-phase SiO2, but with a H3 type hysteresis loop, corresponding to a material formed by aggregated particles. It is well known that the amorphous gas-phase SiO2 is composed of grain-like agglomerates and shows no the especial morphology.
As shown in Table 1, the gas-phase SiO2 has the highest BET area (300.86 m2/g) and pore volume (0.038 cm3/g). Loading metal phosphide, especially ruthenium phosphide, onto SiO2 markedly reduced the specific surface area and pore volume because of partial blocking of the pores by metal phosphide species and the collapse of the pore structure during the fabrication procedures.
In the Co 2p3/2 region (Figure 4a), three significant peaks at 785.7, 782.0, and 778.5 eV are, respectively, assigned to satellite, Co2+, and Coδ+ species. The Coδ+ species are assigned to the Co present in the Co2P and CoP phases with δ ≈ 0, while the Co2+ species are correlated with the cobalt phosphate resulting from the partial passivation on the surface of CoRuP/SiO2 [31,32]. By comparing the binding energy of RuO2 (~463.7 eV), the peak at ~461.8 eV comes from Ru species (Ruδ+) [33]. Previous studies have shown that Ruδ+ species in metal phosphides are effective in absorption and activation of hydrogen [34]. Regardless of whether in metal phosphide or corresponding monometal phosphide, the Co and Ru species exhibit no apparent shift in binding energies. This reveals that there is no notable electron transfer between Co and Ru species.
Table 2 lists the acidic properties of the catalysts according to their NH3-TPD results in Figure 5. With the introduction of Ru, the total acidity preeminently increases (Table 2) and the desorption peaks greatly shift towards higher temperature (Figure 5). The peak at ~250 °C is primarily assigned to Brønsted acid sites arising from PO–H groups [35], while another peak located at ~370 °C might be ascribed to Lewis acid sites originating from the unreduced metal species [24]. This indicates that the Lewis acid sites are stronger than the Brønsted acid sites.

2.2. Catalytic Performance

2.2.1. Screening of Catalysts

As presented in Figure 6, these Ru-containing catalysts with lower BET areas and pore volumes (Table 1) showed high reactivity for C–O cracking, which is consistent with the excellent deoxygenation performance of RuP2/SiO2 [36] and indicates that BET area and pore volume are not determining factors influencing the hydrogenolysis of 4–O–5 linkage. In addition, these catalysts display obviously distinct selectivities in the hydrogenolysis products. For example, NiRuP/SiO2 yielded more phenol, while CoRuP/SiO2 is shown to be an active and selective catalyst for benzene with only a slight amount of by-products. The introduction of ruthenium preeminently improved the hydrodeoxygenation performance of cobalt phosphide. Most importantly, the total acidity exhibits a strong linear correlation with the hydrogenolysis yield of diphenyl ether, as shown in Figure 7. Nevertheless, neither weak nor medium acidity displayed a linear correlation with the hydrogenolysis yield.

2.2.2. Influence of Initial H2 Partial Pressure on Hydrogenolysis Performance

Initial H2 partial pressure significant influenced the distribution of products, as displayed in Figure 8. The hydrogenolysis conversion of diphenyl ether was only 19.2% in the absence of H2, but reached 97.5% at 0.2 MPa, while the benzene yield achieved a maximum value of 59.4%. It is noteworthy that water can act as an important cocatalyst and provide active hydrogen species for hydrogenolysis reaction in the absence of H2 [37]. The hydrogenolysis conversion gradually approached 100% and the benzene yield significantly decreased to 0 with increasing H2 partial pressure to 0.6 MPa. However, the cyclohexane yield climbed from 2.0% to 48.7%. These findings agreed with the previous studies where the hydrogenolysis reaction preferentially occurred with deficient hydrogen species levels [38].

2.2.3. Hydrogenolysis Mechanism

To gain preliminary insights into the hydrogenolysis mechanism of C–O linkages over CoRuP/SiO2, 3-methyl diphenyl ether was further selected as a probe compound, with the results shown in Figure 9. As the reaction proceeded, the content of 3-methyl diphenyl ether dropped dramatically during the initial 60 min and further diminished gradually to 1.7% at 100 min. Concomitantly, the contents of benzene and m-cresol increased equivalently in the first 60 min, following the stoichiometric coefficient of the hydrogenolysis reaction. However, a further elongation of the reaction time resulted in a decrease in the m-cresol content, which finally dropped to 6.0% at 100 min. In contrast, the benzene content reached an almost maximum value of 32.3% at 80 min and remained nearly constant for the remaining time. A small amount of cyclohexane was detected at 100 min, which indicated that the hydrogenation of benzene occurred. The content of toluene increased gradually during the whole reaction process and finally reached 23.1% without hydrogenated derivatives.
Biatomic active hydrogen (H∙∙∙H), originating from the activated H2 adsorption on precious metal Ru, can be heterolytically split into a H and a H+ on metallic phosphides [37,39]. Subsequently, H+ can be trapped by the negatively charged P and PO species to afford Brønsted acidity [24,35,40], while H can transfer towards and deposit on Lewis acid sites, i.e., unreduced Ru and Co metal species [24,32,41]. The higher hydrogenolysis activity of CoRuP/SiO2 might originate from the facile splitting of H2 into H and H+ by its stronger Lewis acidity. Moreover, previous studies found that the H species on Lewis acid sites present much more active than the H+ species for the hydrogenolysis and hydrodeoxygenation reaction on metallic phosphides [24,35]. In contrast, water can act as an important cocatalyst to facilitate direct deoxygenation to afford aromatics [37].
Accordingly, the possible initiation reactions and reaction steps for the catalytic depolymerization of 3-methyl diphenyl ether by CoRuP/SiO2 are proposed as shown in Figure 10 and Scheme 1, respectively. H+ attaching to the electronegative oxygen atom in 3-methyl diphenyl ether does not initiate the C–O cleavage due to the thermodynamic instability of intermediates m11 and m12, which reveals that the proton-assisted initiation reaction, as displayed in Figure 10a, does not easily occur [42]. Actually, the 4–O–5 C–O bond is not depolymerized, even in thermal phosphoric or formic acid [43].
However, even without water and Lewis site assistance (as depicted in Figure 10c,d), the depolymerization of the 4–O–5 linkage initiated by H attacking substituted aromatic carbon atom (Figure 10b) is more thermodynamically favorable due to the stability of the resulting intermediates, m21 and m31 [44]. Additionally, previous research proved that the hydrogenolysis of 3-methyl diphenyl ether, an unsymmetrical diaryl ether, preferentially occurred at the electron-deficient benzene-ring side [45] via intermediate m2, resulting in the formation of benzene and intermediate m21, as shown in Scheme 1. The successive H+ abstraction from the Brønsted acid site or H2O by intermediate m21 leads to the generation of m-cresol, which is a dominant product and can react with absorbed H on metal to finally afford toluene via the thermodynamically favorable intermediate m23.
Furthermore, benzyl phenyl ether and dibenzyl ether were also subjected to hydrogenolysis over CoRuP/SiO2, with their corresponding product distribution shown in Table 3 and possible reaction pathways exhibited in Scheme 2 and Scheme 3, respectively. Similarly, both the principal steps for the depolymerization of dibenzyl ether and benzyl phenyl ether, αO–4 linkage containing compounds, involve H attaching to α–C in the initial step, followed by the subsequent aliphatic C–O bond cleavage. It is generally accepted that an acid possessing a lower pKa value facilitates the formation of the more stable corresponding conjugated base. Therefore, it allows us to deduce that the major pathway for benzyl phenyl ether hydrogenolysis involves the formation of intermediate b21 rather than b31 and b41, while that for the hydrogenolysis of dibenzyl ether prefers b31 to d31 as a dominant intermediate. In the current work, we centered our endeavors on reaction pathways that were in reasonable agreement with experimental results, while not refuting the existence of other complex reaction pathways.

2.2.4. Hydrogenolysis of Lignin over CoRuP/SiO2

To evaluate the changes of etheric bonds in lignin and gain insights into the mechanisms of C–O depolymerization, 2D-HSQC-NMR characterizations of fresh dealkaline lignin and the corresponding residue were performed [46,47,48,49]. The lignin sample in Figure 11 displays the aliphatic ester region. As shown in Figure 11a, obvious cross-signals of Cα–Hα, Cβ–Hβ, and Cγ–Hγ (δC/δH = 67.2/4.5, 76.7/4.4 and 59.0/3.8 ppm, respectively) reveal that A is the predominant structural unit of lignin rather than B, C, and I. Notably, the signals of the A, B, C and I units in residue disappeared completely (Figure 11b). In comparison with our previous work using NiRuP/SiO2 as a catalyst [50], CoRuP/SiO2 can even efficiently catalyze the complete elimination of the methoxy group in lignin under the same reaction conditions, which confirms that CoRuP/SiO2 possesses higher activity in cleaving the C–O bond or deoxygenation.
The organics obtained from the hydrogenolysis of lignin were detected by liquid chromatography-mass spectrometry, and the results are shown in Table 4. The detected organics all contain hydroxyl groups except for 2,4-dimethoxybenzaldehyde and phenyl acetate. As shown in Table 4, the major hydrogenolysis products from dealkaline lignin are acetovanillone, 4-methylcatechol, 4-hydroxy-3-methoxyphenylacetone, and 2-methoxy-4-vinylphenol with yields of 136.1, 683.2, 173.8, and 162.0 μg, respectively. These main products have significant potential in the production of high value-added drugs and pharmaceutical intermediates. It is known that chronic pain and depression-like symptoms can be alleviated using 4-methylcatechol, by which a brain-derived neurotrophic factor is produced [51]. 4-Hydroxy-3-methoxyphenylacetone is a key intermediate for synthesizing carbidopa, which is an important therapeutic drug for Parkinson’s syndrome. Moreover, 4-hydroxy-3-methoxyphenylacetone can also be used as feedstock to produce compounds with biological activities for treating osteoporosis and inhibiting cell mitosis [52]. In addition to anti-inflammatory properties, 2-methoxy-4-vinylphenol is capable of causing cell cycle arrest, which gives it the ability to treat pancreatic cancer [53]. Acetovanillin can be used to treat inflammatory diseases and has been used to treat tracheal asthma [54].

3. Materials and Methods

3.1. Chemicals and Reagents

Ni(NO3)2·6H2O (AR) was obtained from Oukai Chemical Co., Ltd., Xiangyang, China. Co(NO3)2·6H2O (≥99.99%), diphenyl ether (≥99.9%), 3-methyl diphenyl ether (≥98%), benzyl phenyl ether (≥98%), dibenzyl ether (95%), ethyl acetate (99.5% GR), and n-dodecane (>99.0% GC) were purchased from Aladdin Biochemical Technology Co., Ltd., Shanghai, China and other chemicals were obtained from Macklin Biochemical Co., Ltd., Shanghai, China.

3.2. Catalyst Preparation

Co(NO3)2·6H2O, RuCl3·xH2O, citric acid and ammonium hypophosphite were added to deionized water in turn to dissolve. The resulting solution was then impregnated on the gas-phase SiO2. Subsequently, the mixture was aged at ambient temperature and dried at 383 K. The resulting solid was then heated to 927 K under vacuum and held at this temperature for 2 h. The obtained catalyst is labeled as CoRuP/SiO2. The total load of metal is 20 wt.% and the molar ratios of nP/nM (M = Co and Ru) and nCo/nRu were 2/1 and 5/1, respectively. The NiRuP/SiO2 and other catalysts were prepared in the same procedures as CoRuP/SiO2.

3.3. Catalyst Characterization

Ammonia temperature-programmed desorption was conducted with an AutoChem™ II 2920 Chemisorption Analyzer (Micromeritics, Norcross, GA, USA). X-ray diffraction (XRD) patterns were recorded with a SmartLab SE instrument (Rigaku, Tokyo, Japan). X-ray photoelectron spectra were recorded using an ESCALAB instrument (Thermo Fisher, Waltham, MA, USA). Energy-dispersive X-ray spectroscopy and transmission electron microscopy (TEM) images were obtained on a Tecnai™ G2 F20 transmission electron microscope (Thermo Fisher, Waltham, MA, USA). Nitrogen adsorption/desorption examination was employed for calculating the special surface area, pore volume, and its distributions using a Micromeritics ASAP 2460 analyzer (Norcross, GA, USA).

3.4. Hydrogenolysis of Model Compounds

Typically, 5 mL of water, 0.2 mmol of reactant and 0.003 g of the catalyst were put into a reactor, which was then charged with 8 bar N2 and 2 bar H2 after exclusion of air with H2. Reactions were conducted at 250 °C for 1 h 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 products was performed on a GC-FID employing n-dodecane as an internal standard. The hydrogenolysis conversion and product yield were calculated using the following formulas:
Hydrogenolysis   conversion % = moles   of   model   compound   reacted   moles   of   model   compound   supplied × 100
Product   yield   ( x ) % = moles   of   C   atoms   in   product   ( x )   moles   of   C   atoms   in   model   compound × 100

3.5. Hydrogenolysis of Lignin

Due to the less hydrogenolysis reactivity of lignin than model compounds, rather harsh reaction conditions were employed. Typically, 6 mL of water, 0.1 g of dealkaline lignin, and 0.1 g of the catalyst were put into the reactor, which was pressurized in 1.0 MPa H2 after exclusion of air with H2. The temperature was then raised to 280 °C and maintained for 3 h under 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 external standard method. The corresponding residue was characterized by two-dimensional nuclear magnetic resonance heteronuclear single quantum coherence (2D-NMR-HSQC) spectroscopy after drying.

4. Conclusions

In this study, an amorphous Co-Ru phosphide composite catalyst was fabricated and employed to hydrogenolyze lignin and its related model compounds. Under mild reactions, CoRuP/SiO2 can efficiently hydrogenolyze C–O bonds and is highly selective for benzene with a small number of by-products. Importantly, the total acidity is a predominant factor for hydrogenolysis of the 4–O–5 linkage and shows a linear correlation with the hydrogenolysis yield. Regardless of water and Lewis site assistance, the depolymerization of 4–O–5 and αO–4 bonds over CoRuP/SiO2 might be preferentially initiated by H attacking the substituted aromatic carbon atom and the α–carbon atom, respectively.

Author Contributions

Conceptualization, B.C. and Z.-J.D.; methodology, L.-Q.H. and B.C.; validation, L.-Q.H. and Q.-P.D.; formal analysis, K.-Y.D.; investigation, L.-Q.H.; resources, L.-Q.H. and B.C.; data curation, L.-Q.H. and S.-J.Z.; writing—original draft preparation, L.-Q.H.; 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. Patel, M.; Kumar, A. Production of renewable diesel through the hydroprocessing of lignocellulosic biomass-derived bio-oil: A review. Renew. Sustain. Energy Rev. 2016, 58, 1293–1307. [Google Scholar] [CrossRef]
  2. Sheldon, R.A. Green chemistry, catalysis and valorization of waste biomass. J. Mol. Catal. A Chem. 2016, 422, 3–12. [Google Scholar] [CrossRef]
  3. Zhang, Y.; Brown, T.R.; Hu, G.; Brown, R.C. Techno-economic analysis of two bio-oil upgrading pathways. Chem. Eng. J. 2013, 225, 895–904. [Google Scholar] [CrossRef]
  4. Teles, C.A.; de Souza, P.M.; Rabelo-Neto, R.C.; Griffin, M.B.; Mukarakate, C.; Orton, K.A.; Resasco, D.E.; Noronha, F.B. Catalytic upgrading of biomass pyrolysis vapors and model compounds using niobia supported Pd catalyst. Appl. Catal. B Environ. 2018, 238, 38–50. [Google Scholar] [CrossRef]
  5. Sirous-Rezaei, P.; Park, Y.K. Catalytic hydropyrolysis of lignin: Suppression of coke formation in mild hydrodeoxygenation of lignin-derived phenolics. Chem. Eng. J. 2020, 386, 121348. [Google Scholar] [CrossRef]
  6. Gao, F.; Webb, J.D.; Sorek, H.; Wemmer, D.E.; Hartwig, J.F. Fragmentation of lignin samples with commercial Pd/C under ambient pressure of hydrogen. ACS Catal. 2016, 6, 7385–7392. [Google Scholar] [CrossRef]
  7. Zhou, M.; Chen, C.; Liu, P.; Xia, H.; Li, J.; Sharma, B.K.; Jiang, J. Catalytic hydrotreatment of βO–4 ether in lignin: Cleavage of the C-O bond and hydrodeoxygenation of lignin-derived phenols in one pot. ACS Sustain. Chem. Eng. 2020, 8, 14511–14523. [Google Scholar] [CrossRef]
  8. Li, S.; Liu, B.; Truong, J.; Luo, Z.; Ford, P.C.; Abu-Omar, M.M. One-pot hydrodeoxygenation (HDO) of lignin monomers to C9 hydrocarbons co-catalysed by Ru/C and Nb2O5. Green Chem. 2020, 22, 7406–7416. [Google Scholar] [CrossRef]
  9. Mohan, D.; Pittman, C.U., Jr.; Steele, P.H. Pyrolysis of wood/biomass for bio-oil: A critical review. Energy Fuels 2006, 20, 848–889. [Google Scholar] [CrossRef]
  10. Schutyser, W.; Renders, A.T.; Van den Bosch, S.; Koelewijn, S.F.; Beckham, G.T.; Sels, B.F. Chemicals from lignin: An interplay of lignocellulose fractionation, depolymerisation, and upgrading. Chem. Soc. Rev. 2018, 47, 852–908. [Google Scholar] [CrossRef]
  11. 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]
  12. 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]
  13. Parthasarathi, R.; Romero, R.A.; Redondo, A.; Gnanakaran, S. Theoretical study of the remarkably diverse linkages in lignin. J. Phys. Chem. Lett. 2011, 2, 2660–2666. [Google Scholar] [CrossRef]
  14. Rizescu, C.; Sun, C.; Popescu, I.; Urdă, A.; Da Costa, P.; Marcu, I.C. Hydrodeoxygenation of benzyl alcohol on transition-metal-containing mixed oxides derived from layered double hydroxide precursors. Catal. Today 2021, 366, 235–244. [Google Scholar] [CrossRef]
  15. Li, X.; Zhang, B.; Pan, X.; Ji, J.; Ren, Y.; Wang, H.; Ji, N.; Liu, Q.; Li, C. One-pot conversion of lignin into naphthenes catalyzed by a heterogeneous rhenium oxide-modified iridium compound. ChemSusChem 2020, 13, 4409–4419. [Google Scholar] [CrossRef]
  16. Yang, X.; Feng, M.; Choi, J.S.; Meyer, H.M., III; Yang, B. Depolymerization of corn stover lignin with bulk molybdenum carbide catalysts. Fuel 2019, 244, 528–535. [Google Scholar] [CrossRef]
  17. Yan, B.; Lin, X.; Chen, Z.; Cai, Q.; Zhang, S. Selective production of phenolic monomers via high efficient lignin depolymerization with a carbon based nickel-iron-molybdenum carbide catalyst under mild conditions. Bioresour. Technol. 2020, 321, 124503. [Google Scholar] [CrossRef]
  18. Molinari, V.; Clavel, G.; Graglia, M.; Antonietti, M.; Esposito, D. Mild continuous hydrogenolysis of kraft lignin over titanium nitride-nickel catalyst. ACS Catal. 2016, 6, 1663–1670. [Google Scholar] [CrossRef]
  19. Yoosuk, B.; Sanggam, P.; Wiengket, S.; Prasassarakich, P. Hydrodeoxygenation of oleic acid and palmitic acid to hydrocarbon-like biofuel over unsupported Ni-Mo and Co-Mo sulfide catalysts. Renew. Energy 2019, 139, 1391–1399. [Google Scholar] [CrossRef]
  20. Song, W.; Lai, W.; Lian, Y.; Jiang, X.; Yang, W. Sulfated ZrO2 supported CoMo sulfide catalyst by surface exsolution for enhanced hydrodeoxygenation of lignin-derived ethers to aromatics. Fuel 2019, 263, 116705. [Google Scholar] [CrossRef]
  21. Rensel, D.J.; Kim, J.; Jain, V.; Bonita, Y.; Rai, N.; Hicks, J.C. Composition-directed FexMo2-xP bimetallic catalysts for hydrodeoxygenation reactions. Catal. Sci. Technol. 2017, 7, 1857–1867. [Google Scholar] [CrossRef]
  22. Jain, V.; Bonita, Y.; Brown, A.; Taconi, A.; Hicks, J.C.; Rai, N. Mechanistic insights into hydrodeoxygenation of phenol on bimetallic phosphide catalysts. Catal. Sci. Technol. 2018, 8, 4083–4096. [Google Scholar] [CrossRef]
  23. Bonita, Y.; Hicks, J.C. Periodic Trends from metal substitution in bimetallic Mo-based phosphides for hydrodeoxygenation and hydrogenation reactions. J. Phys. Chem. C 2018, 122, 13322–13332. [Google Scholar] [CrossRef]
  24. Li, K.; Wang, R.; Chen, J. Hydrodeoxygenation of anisole over silica-supported Ni2P, MoP, and NiMoP catalysts. Energy Fuel 2011, 25, 854–863. [Google Scholar] [CrossRef]
  25. Li, Y.; Yang, X.; Zhu, L.; Zhang, H.; Chen, B. Hydrodeoxygenation of phenol as a bio-oil model compound over intimate contact noble metal-Ni2P/SiO2 catalysts. RSC Adv. 2015, 5, 80388–80396. [Google Scholar] [CrossRef]
  26. 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, 8, 3650–3658. [Google Scholar] [CrossRef]
  27. 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]
  28. 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]
  29. Nkabinde, S.S.; Ndala, Z.B.; Shumbula, N.P.; Kolokoto, T.; Nchoe, O.; Ngubeni, G.N.; Mubiayi, K.P.; Moloto, N. Delineating the role of crystallinity in the electrocatalytic activity of colloidally synthesized MoP nanocrystals. New J. Chem. 2020, 44, 14041–14049. [Google Scholar] [CrossRef]
  30. Chen, Q.; Cai, C.; Zhang, X.; Zhang, Q.; Chen, L.; Li, Y.; Wang, C.; Ma, L. Amorphous FeNi-ZrO2-catalyzed hydrodeoxygenation of lignin-derived phenolic compounds to naphthenic fuel. ACS Sustain. Chem. Eng. 2020, 8, 9335–9345. [Google Scholar] [CrossRef]
  31. Guo, C.; Rao, K.T.V.; Yuan, Z.; He, S.Q.; Rohani, S.; Xu, C.C. Hydrodeoxygenation of fast pyrolysis oil with novel activated carbon-supported NiP and CoP catalysts. Chem. Eng. Sci. 2017, 178, 248–259. [Google Scholar] [CrossRef]
  32. Rodríguez-Aguado, E.; Infantes-Molina, 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]
  33. Chakroune, N.; Viau, G.; Ammar, S.; Poul, L.; Veautier, D.; Chehimi, M.M.; Mangeney, C.; Villain, F.; Fiévet, F. Acetate-and thiol-capped monodisperse ruthenium nanoparticles: XPS, XAS, and HRTEM studies. Langmuir 2005, 21, 6788–6796. [Google Scholar] [CrossRef] [PubMed]
  34. 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]
  35. Galindo-Ortega, Y.I.; 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]
  36. Bowker, R.H.; Smith, M.C.; Pease, M.L.; Slenkamp, K.M.; Kovarik, L.; Bussell, M.E. Synthesis and hydrodeoxygenation properties of ruthenium phosphide catalysts. ACS Catal. 2011, 1, 917–922. [Google Scholar] [CrossRef]
  37. 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]
  38. Cao, J.P.; Xie, T.; Zhao, X.Y.; Zhu, C.; Jiang, W.; Zhao, M.; Zhao, Y.P.; Wei, X.Y. Selective cleavage of ether C–O bond in lignin-derived compounds over Ru system under different H-sources. Fuel 2021, 284, 119027. [Google Scholar] [CrossRef]
  39. Zeng, G.; Guo, Y.; Li, S. New insights into the molecular mechanism of H2 activation. In Computational Organometallic Chemistry; Wiest, O., Wu, Y., Eds.; Springer: Berlin/Heidelberg, Germany, 2012; pp. 47–60. [Google Scholar]
  40. 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]
  41. 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]
  42. Liu, Z.Q.; Wei, X.Y.; Wu, H.H.; Li, W.T.; Zhang, Y.Y.; Zong, Z.M.; Ma, F.Y.; Liu, J.M. Difunctional nickel/microfiber attapulgite modified with an acidic ionic liquid for catalytic hydroconversion of lignite-related model compounds. Fuel 2017, 204, 236–242. [Google Scholar] [CrossRef]
  43. 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]
  44. Liu, X.X.; Zong, Z.M.; Li, W.T.; Li, X.; Li, Z.K.; Wang, S.K.; Wei, X.Y. A recyclable and highly active magnetic solid superbase for hydrocracking C-O bridged bonds in sawdust. Fuel Process. Technol. 2017, 159, 396–403. [Google Scholar] [CrossRef]
  45. Sergeev, A.G.; Hartwig, J.F. Selective, Nickel-catalyzed hydrogenolysis of aryl ethers. Science 2011, 332, 439–443. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Yuan, T.Q.; Sun, S.N.; Xu, F.; Sun, R.C. Characterization of lignin structures and lignin–carbohydrate complex (LCC) linkages by quantitative 13C and 2D HSQC NMR spectroscopy. J. Agric. Food Chem. 2011, 59, 10604–10614. [Google Scholar] [CrossRef] [PubMed]
  47. Guo, H.; Zhang, B.; Li, C.; Peng, C.; Dai, T.; Xie, H.; Wang, A.; Zhang, T. Tungsten Carbide: A remarkably efficient catalyst for the selective cleavage of lignin C-O bonds. ChemSusChem 2016, 9, 3220–3229. [Google Scholar] [CrossRef] [PubMed]
  48. Chen, W.; McClelland, D.J.; Azarpira, A.; Ralph, J.; Luo, Z.Y.; Huber, G.W. Low temperature hydrogenation of pyrolytic lignin over Ru/TiO2: 2D HSQC and 13C NMR study of reactants and products. Green Chem. 2016, 18, 271–281. [Google Scholar] [CrossRef]
  49. Ren, X.; Wang, P.; Han, X.; Zhang, G.; Gu, J.; Ding, C.; Zheng, X.; Cao, F. Depolymerization of lignin to aromatics by selectively oxidizing cleavage of C–C and C–O bonds using CuCl2/polybenzoxazine catalysts at room temperature. ACS Sustain. Chem. Eng. 2017, 5, 6548–6556. [Google Scholar] [CrossRef]
  50. 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]
  51. Fukuhara, K.; Ishikawa, K.; Yasuda, S.; Kishishita, Y.; Kim, H.K.; Kakeda, T.; Yamamoto, M.; Norii, T.; Ishikawa, T. Intracerebroventricular 4-methylcatechol (4-MC) ameliorates chronic pain associated with depression-like behavior via induction of brain-derived neurotrophic factor (BDNF). Cell. Mol. Neurobiol. 2011, 32, 971–977. [Google Scholar] [CrossRef]
  52. Mann, J.; Wilde, P.D.; Finch, M.W. Synthesis and reactions of 2-aryl-8-oxabicycloc [3.2.1]oct-6-en-3-ones. Tetrahedron 1987, 43, 5431–5441. [Google Scholar] [CrossRef]
  53. Kim, D.H.; Han, S.I.; Go, B.; Oh, U.H.; Kim, C.S.; Jung, Y.H.; Lee, J.; Kim, J.H. 2-methoxy-4-vinylphenol attenuates migration of human pancreatic cancer cells via blockade of fak and akt signaling. Anticancer Res. 2019, 39, 6685–6691. [Google Scholar] [CrossRef] [PubMed]
  54. Stefanska, J.; Sarniak, A.; Wlodarczyk, A.; Sokolowska, M.; Doniec, Z.; Bialasiewicz, P.; Nowak, D.; Pawliczak, R. Hydrogen peroxide and nitrite reduction in exhaled breath condensate of COPD patients. Pulm. Pharmacol. Ther. 2012, 25, 343–348. [Google Scholar] [CrossRef]
Figure 1. XRD spectra of metal phosphides.
Figure 1. XRD spectra of metal phosphides.
Catalysts 12 01328 g001
Figure 2. (a) TEM and (b) elemental-distribution images of CoRuP/SiO2.
Figure 2. (a) TEM and (b) elemental-distribution images of CoRuP/SiO2.
Catalysts 12 01328 g002
Figure 3. (a) Pore size distributions and (b) N2 adsorption/desorption isotherms of samples.
Figure 3. (a) Pore size distributions and (b) N2 adsorption/desorption isotherms of samples.
Catalysts 12 01328 g003
Figure 4. X-ray photoelectron spectrum of CoRuP/SiO2 in (a) Co 2p and (b) Ru 3p regions.
Figure 4. X-ray photoelectron spectrum of CoRuP/SiO2 in (a) Co 2p and (b) Ru 3p regions.
Catalysts 12 01328 g004
Figure 5. NH3-TPD profiles of the catalysts.
Figure 5. NH3-TPD profiles of the catalysts.
Catalysts 12 01328 g005
Figure 6. Depolymerization of diphenyl ether over different metal phosphides.
Figure 6. Depolymerization of diphenyl ether over different metal phosphides.
Catalysts 12 01328 g006
Figure 7. Linear correlation between the total acidity and hydrogenolysis yield.
Figure 7. Linear correlation between the total acidity and hydrogenolysis yield.
Catalysts 12 01328 g007
Figure 8. Influence of initial H2 partial pressure on the distribution of products.
Figure 8. Influence of initial H2 partial pressure on the distribution of products.
Catalysts 12 01328 g008
Figure 9. Time profiles of depolymerization of 3-methyl diphenyl over CoRuP/SiO2.
Figure 9. Time profiles of depolymerization of 3-methyl diphenyl over CoRuP/SiO2.
Catalysts 12 01328 g009
Figure 10. Possible initiation reactions for hydrogenolysis of 3-methyl diphenyl ether over CoRuP/SiO2. (a) H+ attacking oxygen atom; (b) H attacking substituted aromatic carbon atom; (c) with water assistance; (d) with Lewis site assistance.
Figure 10. Possible initiation reactions for hydrogenolysis of 3-methyl diphenyl ether over CoRuP/SiO2. (a) H+ attacking oxygen atom; (b) H attacking substituted aromatic carbon atom; (c) with water assistance; (d) with Lewis site assistance.
Catalysts 12 01328 g010
Scheme 1. Hypothesized pathways for hydrogenolysis of 3-methyl diphenyl ether over CoRuP/SiO2.
Scheme 1. Hypothesized pathways for hydrogenolysis of 3-methyl diphenyl ether over CoRuP/SiO2.
Catalysts 12 01328 sch001
Scheme 2. Hypothesized pathways for hydrogenolysis of benzyl phenyl ether over CoRuP/SiO2.
Scheme 2. Hypothesized pathways for hydrogenolysis of benzyl phenyl ether over CoRuP/SiO2.
Catalysts 12 01328 sch002
Scheme 3. Hypothesized pathways for hydrogenolysis of dibenzyl ether over CoRuP/SiO2.
Scheme 3. Hypothesized pathways for hydrogenolysis of dibenzyl ether over CoRuP/SiO2.
Catalysts 12 01328 sch003
Figure 11. 2D-HSQC-NMR spectroscopy of C–O linkages for (a) fresh dealkaline lignin and (b) the residue; (c) representative structures for ether linkages.
Figure 11. 2D-HSQC-NMR spectroscopy of C–O linkages for (a) fresh dealkaline lignin and (b) the residue; (c) representative structures for ether linkages.
Catalysts 12 01328 g011
Table 1. Structural characteristics of the support and catalysts.
Table 1. Structural characteristics of the support and catalysts.
SampleSurface Area (m2/g)Pore Volume (cm3/g)
SiO2300.860.038
NiP/SiO297.670.036
CoP/SiO294.590.011
RuP/SiO242.470.0004
NiRuP/SiO260.070.0065
CoRuP/SiO265.060.0054
Table 2. Acid sites analysis by NH3-TPD.
Table 2. Acid sites analysis by NH3-TPD.
SampleCoP/SiO2NiP/SiO2RuP/SiO2CoRuP/SiO2NiRuP/SiO2
NH3 acidity/mmol·g−11.11.12.32.43.2
Table 3. Hydrogenolysis of benzyl phenyl ether and dibenzyl ether over RuCoP/SiO2.
Table 3. Hydrogenolysis of benzyl phenyl ether and dibenzyl ether over RuCoP/SiO2.
EntrySubstrateConv./%Yield/%
Catalysts 12 01328 i001Catalysts 12 01328 i002Catalysts 12 01328 i003Catalysts 12 01328 i004
1Catalysts 12 01328 i00598.88.70.621.734.8
2Catalysts 12 01328 i00699.023.50.1-52.8
Table 4. The predominant products from hydrogenolysis of lignin.
Table 4. The predominant products from hydrogenolysis of lignin.
NameYield (μg)
Organic PhaseAqueous phase
2,4-Dimethoxybenzaldehyde1.0-
2,3-Dihydroxytoluene35.317.1
2,6-Dimethoxyphenol29.615.9
2-Methoxy-4-vinylphenol116.245.8
3-(3,4-Dimethoxyphenyl) propionic acid5.13.3
4-Hydroxy-3-methoxyphenylacetone116.757.1
4-Methylcatechol488.5194.7
Acetosyringone2.92.4
Acetovanillone93.642.5
Isoeugenol4.3-
Phenyl acetate22.65.3
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

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. https://doi.org/10.3390/catal12111328

AMA Style

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(11):1328. https://doi.org/10.3390/catal12111328

Chicago/Turabian Style

Huang, Liang-Qiu, Zhi-Jun Diao, Bo Chen, Qing-Pan Du, Kai-Yang Duan, and Si-Jia Zhao. 2022. "Hydrogenolysis of Lignin and C–O Linkages Containing Lignin-Related Compounds over an Amorphous CoRuP/SiO2 Catalyst" Catalysts 12, no. 11: 1328. https://doi.org/10.3390/catal12111328

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