Svoboda | Graniru | BBC Russia | Golosameriki | Facebook
Next Article in Journal
Construction of Cu2O-ZnO/Cellulose Composites for Enhancing the Photocatalytic Performance
Previous Article in Journal
BiVO4-Based Photocatalysts for the Degradation of Antibiotics in Wastewater: Calcination Role after Solvothermal Synthesis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

The Hydrodeoxygenation of Phenol over Ni-P/Hβ and Ni-P/Ce-β: Modifying the Effects in Dispersity and Acidity

1
Yinchuan University of Energy, Yongning Wangtaibu, Yinchuan 750105, China
2
Dalian National Laboratory for Clean Energy, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, Dalian 116023, China
3
PetroChina Dalian Lubricant R&D Center, Dalian 116000, China
*
Author to whom correspondence should be addressed.
Catalysts 2024, 14(8), 475; https://doi.org/10.3390/catal14080475
Submission received: 5 June 2024 / Revised: 15 July 2024 / Accepted: 17 July 2024 / Published: 25 July 2024

Abstract

:
The supported Ni-P catalysts (marked as s-Ni-P/Hβ(3) and s-Ni-P/Ce-β(3)) were prepared by an incipient wetness step-impregnation method, and characterized by XRD, N2 physisorption, TEM, XPS, and NH3-TPD. The catalytic hydrodeoxygenation (HDO) performance was assessed using phenol in water (5.0 wt%) or in decalin (1.0 wt%) as the feed. After the introduction of Ce, the conversion of phenol increased due to the high dispersity of the active site. However, compared to s-Ni-P/Hβ(3), the amount of total and strong acid sites of s-Ni-P/Ce-β(3) decreased, restraining the cycloisomerization of cyclohexane to form methyl-cyclopentane. Moreover, the kinetics of the APHDO and OPHDO of phenol catalyzed by s-Ni-P/Hβ(3) and s-Ni-P/Ce-β(3) were investigated.

1. Introduction

Bio-oils are promising sources for the production of sustainable engine fuels. However, due to the high oxygen content (45~50 wt%), bio-oils are thermolabile and have a low heating value and high viscosity [1]. Hydrodeoxygenation (HDO) is regarded as the most promising route for reducing the oxygen content [2]. Water not only accounts for about 20% of the mass composition of bio-oil, but is also a by-product in HDO. Nowadays, the mixing refining between petroleum and bio-oil attracts more attention. Thus, it is necessary to investigate the HDO performance both in the aqueous phase (APHDO) and the oil phase (OPHDO).
The traditional HDO catalysts were Mo-S or W-S, which were effective in removing oxygen from bio-oils. However, the sulfur was easily replaced by oxygen, leading to the progressive deactivation caused by the destruction of the catalytic phase [3,4,5]. As a result, the non-sulfide catalysts, including metals (precious and base metals) [6,7,8,9,10,11,12,13,14,15,16] and intermetallic compounds (metal carbides, metal nitrides, and metal phosphides) [17,18,19,20,21,22,23,24], have attracted more attention in upgrading bio-oils. Nickel phosphide (including Ni2P, Ni12P5, Ni3P, and so on) exhibited high catalytic activity in the HDO of phenolic compounds [21,23,25,26,27].
The HDO of phenolic compounds to form cycloalkanes was proceeded by the hydrogenation (HYD) pathway, which requires a metal–acid bi-functional catalyst. Hβ was widely used as the catalyst or support and showed high performance in bio-mass conversion [28,29,30]. However, the overly strong acidity tended to deactivate the catalyst by carbon deposition. Ceria was amphoteric between acid and base [31], which could be used to regulate the acidity of zeolite [32,33,34]. Compared with the bulk Ni3P, the addition of Ce significantly influenced the particle size and acidity of the Ni3P catalyst, accelerating both hydrogenation and C-O bond cleavage during the HDO of phenolic compounds [35].
In this context, the aim of the current work was to survey the HDO of phenol in both aqueous and oil phases using nickel phosphide supported on Hβ and Ce-β. The synthesis of the supported catalysts was achieved using phosphate precursors by the step-impregnation method. Approximately 25% of oxygenates in pyrolysis bio-oils were phenolic compounds, which were difficult to remove in HDO treatments [36]. Hence, the catalytic HDO of the prepared catalysts was carried out using phenol as the model substrate. This study could provide important practical information for upgrading pyrolysis oils.

2. Results and Discussion

2.1. Characterization

2.1.1. XRD

Figure 1 shows the XRD patterns of the support and the catalysts. As seen in Figure 1a, after the impregnation of Ce, the crystalline Hβ was not changed, but the peak intensity decreased. The peaks at 28.6, 33.1, 47.5, 56.3, and 76.7° were ascribed to the (111), (200), (220), (311), and (331) planes of CeO2, respectively. Moreover, the XRD patterns of nickel phosphides for s-Ni-P/Hβ(3) and s-Ni-P/Ce-β(3) were Ni-Ni3P-Ni12P5. Compared with s-Ni-P/Hβ(3), the full width at half maxima (FWHM) of the Ni3P crystal phase was wider, indicating that the Ni3P crystal size of the s-Ni-P/Ce-β(3) catalyst was smaller.

2.1.2. N2 Physisorption

After the impregnation of CeO2, the BET surface area (SBET) of the support decreased from 358.3 to 230.4 m2·g−1 (Table 1). Moreover, compared to the support, after the preparation of nickel phosphide the SBET of s-Ni-P/Hβ(3) and s-Ni-P/Ce-β(3) decreased to 295.8 and 153.7 m2·g−1, respectively. The decreasing trend of SBET could be due to the pore blockage by the impregnation components.

2.1.3. TEM

The TEM images of s-Ni-P/Hβ and s-Ni-P/Ce-β are shown in Figure 2. It can be seen that, compared with s-Ni-P/Hβ, the s-Ni-P/Ce-β catalyst had smaller particles (the mean particle size decreased from 17.3 to 11.6 nm), indicating that the incorporation of Ce was beneficial to the formation of smaller particles, increasing the dispersity of active sites.

2.1.4. XPS

Figure 3 shows the XPS spectra of the s-Ni-P/Hβ(3) and s-Ni-P/Ce-β(3) catalysts in the regions of Ni 2p3/2 and P 2p3/2. The electron density of nickel phosphides is shared between the metal and phosphorus atoms because of the characteristics of covalent compounds [37,38]. Three peaks were observed by deconvolution of the broad band of Ni 2p3/2. One peak (852.8–853.1 eV) was usually attributed to the reduced Niδ+ species, while the other two peaks (856.8 eV and 862.6) were due to the oxidized Ni2+ species [39,40]. Similarly, the band of P 2p3/2 could be deconvoluted into two peaks at 129.7–130.0 eV and 133.4–133.5 eV, ascribed to the Pδ− and P5+ species, respectively. Prior to XPS analysis, the samples were inevitably exposed to the air, leading to the formation of oxidized Ni2+ and P5+ species [41]. Compared with the s-Ni-P/Hβ(3) catalyst, the introduction of Ce decreased the Niδ+ binding energy from 853.1 eV to 852.9 eV, but increased the Pδ− binding energy from 129.7 eV to 130.0 eV, indicating the reduced transfer of electron density from Ni to P atoms caused by Ce [42,43].

2.1.5. NH3-TPD

Figure 4 shows the NH3-TPD curves of s-Ni-P/Hβ and s-Ni-P/Ce-β. The s-Ni-P/Hβ catalyst possessed three desorption peaks at 188, 235, and 403 °C, ascribed to weak, medium, and strong acid sites, respectively. However, two desorption peaks at 179 and 399 °C, ascribed to weak and strong acid sites, were observed for the s-Ni-P/Ce-β catalyst. After introducing Ce, the acid strength of both weak and strong acids decreased for the lower desorption peak position, while the medium acid sites disappeared. As seen in Table 1, the total number of acid sites decreased after the introduction of Ce. It has been frequently reported in the literature that the modification of rare earth oxides decreased the acidity, enhanced the oxygen mobility, and improved the metal dispersion of the supported Ni catalysts [44]. After the impregnation of Ce on Hβ, the amount of acid sites decreased by about two thirds [34]. The introduction of CeO2 to Cu/HZSM-5 decreased the amount of both moderate and strong acid sites, while the weak acid sites were slightly affected [33]. A partial exchange of rare earth cations reduced the acid site strength of HY, because the electronegativity of rare earth cations was lower compared to that of protons [32].

2.2. Catalytic Activity

2.2.1. HDO of Phenol in Aqueous Phase (APHDO) and Oil Phase (OPHDO)

Figure 5 shows the APHDO performance of phenol over s-Ni-P/Hβ(3), s-Ni-P/Ce-β(3), and Ni/Hβ catalysts. Over these catalysts, the main products were cyclohexanol, cyclohexane, and methyl-cyclopentane, with a small amount of cyclohexanone and cyclohexene, indicating that the APHDO of phenol was mainly proceeded in the HYD pathway. Compared to the s-Ni-P/Hβ(3) catalyst, the APHDO performance of s-Ni-P/Ce-β(3) increased, especially at low temperatures. At 150 °C, the conversion of phenol increased from 65.8% to 90.6%. As discussed above, the enhanced reactivity of the catalyst in the APHDO of phenol after the impregnation of Ce might be due to the high dispersity and low transfer of electron density from Ni to P. At 350 °C after the introduction of Ce, the selectivity to methyl-cyclopentane decreased from 21.0% to 7.8%, which was consistent with the decrease in strong acid content (Figure 4 and Table 1). However, beneath 350 °C, the selectivity to cyclohexanol and cyclohexane changes slightly, owing to the small decrease in the sum of weak and medium acid content. In addition, the Ni/Hβ catalyst exhibited the worst APHDO performance (Figure 5c), indicating that the nickel phosphide crystal phase was more active than metallic Ni [25]. In our previous study, we found that the activity of the nickel phase decreased as follows: Ni3P > Ni12P5 > Ni2P [40]. Due to the smaller crystal size of Ni3P (Figure 1b), the s-Ni-P/Ce-β(3) catalyst exhibited superior HDO activity than s-Ni-P/Hβ(3).
Figure 6 shows the HDO of phenol in decalin over the s-Ni-P/Hβ(3) and s-Ni-P/Ce-β(3) catalysts. Analogous to s-Ni-P/Hβ(3), over the s-Ni-P/Ce-β(3) catalyst the conversion of phenol was about 100%, with the main products of cyclohexane, methyl-cyclopentane, and cyclohexene under the same reaction conditions. Compared to s-Ni-P/Hβ(3), the decrease in methyl-cyclopentane (the cycloisomerization product) could also be due to the decrease in the amount of strong acid. Compared with the fresh s-Ni-P/Hβ(3) catalyst, the XRD pattern of the spent s-Ni-P/Hβ(3) catalyst remained unchanged (Figure 7), indicating that the crystal phase was stable in the HDO of phenol.

2.2.2. Kinetic Analysis in Aqueous Phase and Oil Phase HDO of Phenol

To study the kinetics of phenol HDO in the aqueous phase, the catalytic performance of s-Ni-P/Hβ(3) and s-Ni-P/Ce-β(3) was investigated at 250–325 °C and 0.09–0.28 g·min·mol−1. Figure 8 shows the APHDO performance of phenol over the s-Ni-P/Hβ(3) and s-Ni-P/Ce-β(3) catalysts. When −ln(1 − x) and weight time were assigned to the vertical coordinate and horizontal coordinate, respectively, two straight lines through the origin were fitted at each temperature for the s-Ni-P/Hβ(3) and s-Ni-P/Ce-β(3) catalysts. It could be concluded that the APHDO of phenol could be treated as a pseudo-first-order kinetics on the variation in phenol concentration. According to the equation: −ln(1 − x) = kτ, the rate constants (k) for phenol APHDO at various temperatures were the slopes of the lines. The calculated k values are summarized in Table 2. Compared to the s-Ni-P/Hβ(3) catalyst, the rate constant of s-Ni-P/Ce-β(3) increased, indicating that the hydrogenation of phenol in the aqueous phase was enhanced after the introduction of CeO2.
Figure 9 shows the OPHDO performance of phenol over the s-Ni-P/Hβ(3) and s-Ni-P/Ce-β(3) catalysts at 150–225 °C and 0.09–0.28 g·min·mol−1. Similarly, two straight lines through the origin were fitted at each temperature for the s-Ni-P/Hβ(3) and s-Ni-P/Ce-β(3) catalysts, suggesting that the reaction order of phenol OPHDO with respect to the phenol concentration was one. The rate constants, calculated by fitting the slope of the lines, are listed in Table 3.
When ln(k) (k represents the rate constant) was plotted against 1/T, a linear relationship was established in both the aqueous-phase and oil-phase HDO of phenol over the s-Ni-P/Hβ(3) and s-Ni-P/Ce-β(3) catalysts (Figure 10). The activation energy (Ea) was calculated according to the Arrhenius equation by fitting the slope of the line and is listed in Table 2 and Table 3. Compared to the s-Ni-P/Hβ(3) catalyst, the Ea of phenol HDO over the s-Ni-P/Ce-β(3) catalyst decreased in the same conditions, both in the aqueous phase and the oil phase. It could be concluded that, after the addition of CeO2, the reaction rate of phenol HDO was elevated.

3. Materials and Methods

3.1. Materials

All reagents were analytical pure grade and purchased from Sinopharm Group CO. LTD., China. They were used as provided: Ni(NO3)2·6H2O, (NH4)2HPO4, Ce(NO3)3·6H2O, Hβ, quartz sand, decalin, dichloromethane, anhydrous MgSO4, phenol.

3.2. Catalyst Preparation

Ce-β was prepared by incipient wetness impregnation, with the amount of Ce loading of 12.5 wt%. Under constant stirring, 1 mL Ce(NO3)3·6H2O solution (0.36 g·mL−1) was dripped on 1.0 g Hβ dropwise, aged (ambient temperature for 12 h), and dried (120 °C overnight) subsequently. Then, the solid was calcined at 450 °C for 3 h to obtain the final product, marked as Ce-β.
The oxidic precursor of supported Ni-P catalysts was prepared by incipient wetness step-impregnation. The initial Ni/P molar ratio was 3, with the amount of Ni loading maintained at 20 wt%. The procedure of impregnation, aging, and drying were similar to the preparation of Ce-β. Then, the obtained product was calcined at 500 °C for 3 h to form the nickel-containing precursor. Afterwards, the phosphorus source ((NH4)2HPO4) was impregnated onto the previous precursor following the analogous procedure to obtain the final oxidic precursor.
The nickel phosphide catalysts were obtained by the in situ temperature-programmed reduction (TPR) of the oxidic precursors in H2. The precursor was heated from 20 °C to 400 °C in 190 min, then to 500 °C in 100 min, and kept at 500 °C for 180 min, with the H2 flow of 150 mL·min−1. The obtained catalysts were named as s-Ni-P/support(3).

3.3. Catalyst Characterization

The reduced samples were easily oxidized. Before exposure to air, the samples were passivated in an O2/Ar (0.5 vol%) flow (20 mL·min−1) at room temperature for 120 min.
The crystal phase of the catalysts was analyzed by X-ray diffraction (XRD) measured with a Rigaku D/Max 2400 diffractometer (Japan) equipped with nickel-filtered Cu-Kα radiation (at 40 kV and 100 mA). A Multilab 2000 X-ray photoelectron spectrometer (XPS, USA) with a Mg-Kα source was used to obtain the XPS spectra. The N2 adsorption isotherms of the catalysts were measured on a Micrometritics Tristar II (3020 model, USA) at −196 °C. The specific surface area was calculated by the BET method.
The TEM images of the catalysts were observed using FEI Tecnai G2 F30 transmission electron microscope equipment (USA, 300 kV). Before measurement, the samples were dispersed in ethanol by ultrasonic treatment. The ethanol suspension was dripped onto a Cu grid coated with carbon, and then dried under an infrared lamp.
The NH3-TPD curves were recorded on a Chembet-3000 (Lab-made) equipped with a TCD. Prior to measurement, the passivated samples needed to be re-reduced (in H2 at 500 °C for 1 h) to remove the passivation layer. Afterwards, the sample was cooled naturally to 40 °C and adsorbed NH3 for 30 min. As the temperature increased (from 40 °C to 500 °C at 10 °C·min−1), the desorption signal was collected. The acid amount was calibrated by integrating the desorption peaks with a standard sample (zeolite HZSM-5) as the reference.

3.4. Catalytic Performance in HDO

The catalytic activity was evaluated in a fixed-bed reactor (Lab-made) (stainless steel, 10.0 mm i.d.) with phenol as the reactant. A 0.2 g catalyst precursor (0.4–0.8 mm) was loaded on the constant temperature zone and was transformed in situ into the catalytic phase (metal phosphide) in a 150 mL·min−1 H2 flow. The reduction procedure was in accordance with the catalyst preparation. Afterwards, the reactor was naturally cooled to the reaction temperature, elevating the total pressure to 4.0 MPa. The HDO reactions were carried out at WHSV of 30 h−1, with phenol (5.0 wt% in water or 1.0 wt% in decalin) as the reactant. The liquid product was collected and analyzed every 2 h. In the APHDO of phenol, dichloromethane was used to extract the organic samples from aqueous products. Prior to gas chromatograph (GC) analysis, the samples were dried using anhydrous MgSO4. The GC (Agilent 6890, USA) was equipped with an HP-INNOWax capillary column (30 m × 320 µm × 30 µm) and a flame ionization detector. In the OPHDO of phenol, the products were analyzed directly, eliminating the extraction and drying procedure. The carbon balance was over 98% in the HDO of phenol over these catalysts.

4. Conclusions

Ce was presented in the form of CeO2 during the preparation of the s-Ni-P/Ce-β(3) catalyst. Compared to s-Ni-P/Hβ(3), s-Ni-P/Ce-β(3) presented higher hydrogenation activity of phenol due to the increased dispersity of Ni3P. After the introduction of CeO2, the amount of total and strong acid sites decreased, suppressing the cycloisomerization of cyclohexane to methyl-cyclopentane. Utilizing, optimizing, and modifying the supports to prepare an excellent HDO catalyst is attractive.

Author Contributions

L.M., Z.Y. and W.W. conceived and designed the experiments; Y.L. and Y.J. performed the experiments; J.Z. and Y.J. analyzed the data; W.W. contributed reagents/materials/analysis tools; Y.L. and Z.Y. wrote the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The data presented in this work are available on request from the corresponding author.

Acknowledgments

The Key Scientific Research Project of Ningxia Higher Education (NYG2022130); the Natural Science Foundation of Ningxia (2023AAC03392); Yinchuan University of Energy university-level scientific research (2022-KY-Z-2).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Corma, A.; Iborra, S.; Velty, A. Chemical routes for the transformation of biomass into chemicals. Chem. Rev. 2007, 107, 2411–2502. [Google Scholar] [CrossRef] [PubMed]
  2. Mortensen, P.M.; Grunwaldt, J.D.; Jensen, P.A.; Knudsen, K.G.; Jensen, A.D. A review of catalytic upgrading of bio-oil to engine fuels. Appl. Catal. A Gen. 2011, 407, 1–19. [Google Scholar] [CrossRef]
  3. Laurent, E.; Delmon, B. Influence of Oxygen-Containing, Nitrogen-Containing, and Sulfur-Containing-Compounds on the Hydrodeoxygenation of Phenols over Sulfided Como/Gamma-Al2O3 and Nimo/Gamma-Al2O3 Catalysts. Ind. Eng. Chem. Res. 1993, 32, 2516–2524. [Google Scholar] [CrossRef]
  4. Elliott, D.C. Historical developments in hydroprocessing bio-oils. Energy Fuels 2007, 21, 1792–1815. [Google Scholar] [CrossRef]
  5. Viljava, T.R.; Komulainen, R.S.; Krause, A.O.I. Effect of H2S on the stability of CoMo/Al2O3 catalysts during hydrodeoxygenation. Catal. Today 2000, 60, 83–92. [Google Scholar] [CrossRef]
  6. Zhao, C.; Kou, Y.; Lemonidou, A.A.; Li, X.; Lercher, J.A. Highly Selective Catalytic Conversion of Phenolic Bio-Oil to Alkanes. Angew. Chem. Int. Ed. 2009, 48, 3987–3990. [Google Scholar] [CrossRef]
  7. Wang, M.; Shi, H.; Camaioni, D.M.; Lercher, J.A. Palladium-Catalyzed Hydrolytic Cleavage of Aromatic C−O Bonds. Angew. Chem. Int. Ed. 2017, 56, 2110–2114. [Google Scholar] [CrossRef] [PubMed]
  8. Xu, X.; Li, Y.; Gong, Y.; Zhang, P.; Li, H.; Wang, Y. Synthesis of Palladium Nanoparticles Supported on Mesoporous N-Doped Carbon and Their Catalytic Ability for Biofuel Upgrade. J. Am. Chem. Soc. 2012, 134, 16987–16990. [Google Scholar] [CrossRef] [PubMed]
  9. Crossley, S.; Faria, J.; Shen, M.; Resasco, D.E. Solid Nanoparticles that Catalyze Biofuel Upgrade Reactions at the Water/Oil Interface. Science 2010, 327, 68–72. [Google Scholar] [CrossRef]
  10. Strapasson, G.B.; Sousa, L.S.; Báfero, G.B.; Leite, D.S.; Moreno, B.D.; Rodella, C.B.; Zanchet, D. Acidity modulation of Pt-supported catalyst enhances C-O bond cleavage over acetone hydrodeoxygenation. Appl. Catal. B Environ. 2023, 335, 122863. [Google Scholar] [CrossRef]
  11. Saleheen, M.; Mamun, O.; Mohan Verma, A.; Sahsah, D.; Heyden, A. Understanding the influence of solvents on the Pt-catalyzed hydrodeoxygenation of guaiacol. J. Catal. 2023, 425, 212–232. [Google Scholar] [CrossRef]
  12. Aqsha, A.; Katta, L.; Mahinpey, N. Catalytic Hydrodeoxygenation of Guaiacol as Lignin Model Component Using Ni-Mo/TiO2 and Ni-V/TiO2 Catalysts. Catal. Lett. 2015, 145, 1351–1363. [Google Scholar] [CrossRef]
  13. Yakovlev, V.A.; Khromova, S.A.; Sherstyuk, O.V.; Dundich, V.O.; Ermakov, D.Y.; Novopashina, V.M.; Lebedev, M.Y.; Bulavchenko, O.; Parmon, V.N. Development of new catalytic systems for upgraded bio-fuels production from bio-crude-oil and biodiesel. Catal. Today 2009, 144, 362–366. [Google Scholar] [CrossRef]
  14. He, J.Y.; Zhao, C.; Lercher, J.A. Ni-Catalyzed Cleavage of Aryl Ethers in the Aqueous Phase. J. Am. Chem. Soc. 2012, 134, 20768–20775. [Google Scholar] [CrossRef] [PubMed]
  15. Abreu Teles, C.; Ciotonea, C.; Le Valant, A.; Canaff, C.; Dhainaut, J.; Clacens, J.-M.; Bellot Noronha, F.; Richard, F.; Royer, S. Optimization of catalyst activity and stability in the m-cresol hydrodeoxygenation through Ni particle size control. Appl. Catal. B Environ. 2023, 338, 123030. [Google Scholar] [CrossRef]
  16. Jiang, C.; Cai, Y.; Xu, T.; Xiao, B.; Hu, Z.; Wang, X. Vapor-phase hydrodeoxygenation of guaiacol for phenol production using bifunctional Ni/Cu-Beta zeolite catalysts. J. Energy Inst. 2023, 109, 101273. [Google Scholar] [CrossRef]
  17. Han, J.; Duan, J.; Chen, P.; Lou, H.; Zheng, X.; Hong, H. Nanostructured molybdenum carbides supported on carbon nanotubes as efficient catalysts for one-step hydrodeoxygenation and isomerization of vegetable oils. Green Chem. 2011, 13, 2561–2568. [Google Scholar] [CrossRef]
  18. Chen, C.-J.; Bhan, A. Mo2C Modification by CO2, H2O, and O2: Effects of Oxygen Content and Oxygen Source on Rates and Selectivity of m-Cresol Hydrodeoxygenation. Acs Catal. 2017, 7, 1113–1122. [Google Scholar] [CrossRef]
  19. Sosa, L.F.; de Souza, P.M.; Rafael, R.A.; Wojcieszak, R.; Briois, V.; Francisco, L.R.; Rabelo-Neto, R.C.; Marceau, E.; Paul, S.; Toniolo, F.S.; et al. Study of the performance of SiO2-supported Mo2C and metal-promoted Mo2C catalysts for the hydrodeoxygenation of m-cresol. Appl. Catal. B Environ. 2023, 331, 122720. [Google Scholar] [CrossRef]
  20. Ghampson, I.T.; Sepúlveda, C.; Garcia, R.; Radovic, L.R.; Fierro, J.L.G.; DeSisto, W.J.; Escalona, N. Hydrodeoxygenation of guaiacol over carbon-supported molybdenum nitride catalysts: Effects of nitriding methods and support properties. Appl. Catal. A Gen. 2012, 439–440, 111–124. [Google Scholar] [CrossRef]
  21. 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]
  22. Ted Oyama, S.; Onkawa, T.; Takagaki, A.; Kikuchi, R.; Hosokai, S.; Suzuki, Y.; Bando, K. Production of Phenol and Cresol from Guaiacol on Nickel Phosphide Catalysts Supported on Acidic Supports. Top. Catal. 2015, 58, 201–210. [Google Scholar] [CrossRef]
  23. Moon, J.-S.; Lee, Y.-K. Support Effects of Ni2P Catalysts on the Hydrodeoxygenation of Guaiacol: In Situ XAFS Studies. Top. Catal. 2015, 58, 211–218. [Google Scholar] [CrossRef]
  24. Wu, S.-K.; Lai, P.-C.; Lin, Y.-C.; Wan, H.-P.; Lee, H.-T.; Chang, Y.-H. Atmospheric Hydrodeoxygenation of Guaiacol over Alumina-, Zirconia-, and Silica-Supported Nickel Phosphide Catalysts. ACS Sustain. Chem. Eng. 2013, 1, 349–358. [Google Scholar] [CrossRef]
  25. Yu, Z.; Wang, Y.; Zhang, G.; Sun, Z.; Liu, Y.-Y.; Shi, C.; Wang, W.; Wang, A. A highly dispersed Ni3P/HZSM-5 catalyst for hydrodeoxygenation of phenolic compounds to cycloalkanes. J. Catal. 2022, 410, 294–306. [Google Scholar] [CrossRef]
  26. Song, H.; Gong, J.; Song, H.; Li, F. A novel surface modification approach for synthesizing supported nickel phosphide catalysts with high activity for hydrodeoxygenation of benzofuran. Appl. Catal. A Gen. 2015, 505, 267–275. [Google Scholar] [CrossRef]
  27. Yu, Z.; Wang, Y.; Sun, Z.; Li, X.; Wang, A.; Camaioni, D.M.; Lercher, J.A. Ni3P as a high-performance catalytic phase for the hydrodeoxygenation of phenolic compounds. Green Chem. 2018, 20, 609–619. [Google Scholar] [CrossRef]
  28. Zhang, Q.; Xiang, S.; Wang, B.; Mayoral, A.; Liu, W.; Wang, Y.; Liu, Y.; Shi, J.; Yang, G.; Luo, J.; et al. Breaking the Si/Al Limit of Nanosized β Zeolites: Promoting Catalytic Production of Lactide. Chem. Mater. 2020, 32, 751–758. [Google Scholar] [CrossRef]
  29. Zhang, Y.; Luo, H.; Kong, L.; Zhao, X.; Miao, G.; Zhu, L.; Li, S.; Sun, Y. Highly efficient production of lactic acid from xylose using Sn-beta catalysts. Green Chem. 2020, 22, 7333–7336. [Google Scholar] [CrossRef]
  30. Tian, Y.; Duan, H.; Zhang, B.; Gong, S.; Lu, Z.; Dai, L.; Qiao, C.; Liu, G.; Zhao, Y. Template Guiding for the Encapsulation of Uniformly Subnanometric Platinum Clusters in Beta-Zeolites Enabling High Catalytic Activity and Stability. Angew. Chem. Int. Ed. 2021, 60, 21713–21717. [Google Scholar] [CrossRef]
  31. An, J.; Wang, Y.; Lu, J.; Zhang, J.; Zhang, Z.; Xu, S.; Liu, X.; Zhang, T.; Gocyla, M.; Heggen, M.; et al. Acid-Promoter-Free Ethylene Methoxycarbonylation over Ru-Clusters/Ceria: The Catalysis of Interfacial Lewis Acid–Base Pair. J. Am. Chem. Soc. 2018, 140, 4172–4181. [Google Scholar] [CrossRef] [PubMed]
  32. Diaz-Mendoza, F.A.; Pernett-Bolaño, L.; Cardona-Martínez, N. Effect of catalyst deactivation on the acid properties of zeolites used for isobutane/butene alkylation. Thermochim. Acta 1998, 312, 47–61. [Google Scholar] [CrossRef]
  33. Liu, X.; Wu, X.; Weng, D.; Shi, L. Modification of Cu/ZSM-5 catalyst with CeO2 for selective catalytic reduction of NOx with ammonia. J. Rare Earths 2016, 34, 1004–1009. [Google Scholar] [CrossRef]
  34. Shi, Y.; Wang, X.; Xia, Y.; Sun, C.; Zhao, C.; Li, S.; Li, W. Promotional effect of CeO2 on the propene poisoning resistance of HBEA zeolite catalyst for NH3-SCR of NOx. Mol. Catal. 2017, 433, 265–273. [Google Scholar] [CrossRef]
  35. Yu, Z.; Wang, Y.; Liu, S.; Yao, Y.; Sun, Z.; Li, X.; Liu, Y.; Wang, W.; Wang, A.; Camaioni, D.M.; et al. Aqueous Phase Hydrodeoxygenation of Phenol over Ni3P-CePO4 Catalysts. Ind. Eng. Chem. Res. 2018, 57, 10216–10225. [Google Scholar] [CrossRef]
  36. Bu, Q.; Lei, H.; Zacher, A.H.; Wang, L.; Ren, S.; Liang, J.; Wei, Y.; Liu, Y.; Tang, J.; Zhang, Q.; et al. A review of catalytic hydrodeoxygenation of lignin-derived phenols from biomass pyrolysis. Bioresour. Technol. 2012, 124, 470–477. [Google Scholar] [CrossRef] [PubMed]
  37. Carenco, S.; Portehault, D.; Boissiere, C.; Mezailles, N.; Sanchez, C. Nanoscaled Metal Borides and Phosphides: Recent Developments and Perspectives. Chem. Rev. 2013, 113, 7981–8065. [Google Scholar] [CrossRef] [PubMed]
  38. Clark, P.; Li, W.; Oyama, S. Synthesis and Activity of a New Catalyst for Hydroprocessing: Tungsten Phosphide. J. Catal. 2001, 200, 140–147. [Google Scholar] [CrossRef]
  39. Blanchard, P.E.R.; Grosvenor, A.P.; Cavell, R.G.; Mar, A. X-ray Photoelectron and Absorption Spectroscopy of Metal-Rich Phosphides M2P and M3P (M = Cr-Ni). Chem. Mater. 2008, 20, 7081–7088. [Google Scholar] [CrossRef]
  40. Yu, Z.; Yao, K.; Wang, Y.; Yao, Y.; Sun, Z.; Liu, Y.; Shi, C.; Wang, W.; Wang, A. Kinetic investigation of phenol hydrodeoxygenation over unsupported nickel phosphides. Catal. Today 2021, 371, 179–188. [Google Scholar] [CrossRef]
  41. Sawhill, S.J.; Layman, K.A.; Van Wyk, D.R.; Engelhard, M.H.; Wang, C.; Bussell, M.E. Thiophene hydrodesulfurization over nickel phosphide catalysts: Effect of the precursor composition and support. J. Catal. 2005, 231, 300–313. [Google Scholar] [CrossRef]
  42. Fuks, D.; Vingurt, D.; Landau, M.V.; Herskowitz, M. Density Functional Theory Study of Sulfur Adsorption at the (001) Surface of Metal-Rich Nickel Phosphides: Effect of the Ni/P Ratio. J. Phys. Chem. C 2010, 114, 13313–13321. [Google Scholar] [CrossRef]
  43. Sun, Z.; Li, X.; Wang, A.; Wang, Y.; Chen, Y. The Effect of CeO2 on the Hydrodenitrogenation Performance of Bulk Ni2P. Top. Catal. 2012, 55, 1010–1021. [Google Scholar] [CrossRef]
  44. Osorio-Vargas, P.; Flores-González, N.A.; Navarro, R.M.; Fierro, J.L.G.; Campos, C.H.; Reyes, P. Improved stability of Ni/Al2O3 catalysts by effect of promoters (La2O3, CeO2) for ethanol steam-reforming reaction. Catal. Today 2016, 259, 27–38. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of (a) Hβ and Ce-β, (b) s-Ni-P/Hβ(3) and s-Ni-P/Ce-β(3).
Figure 1. XRD patterns of (a) Hβ and Ce-β, (b) s-Ni-P/Hβ(3) and s-Ni-P/Ce-β(3).
Catalysts 14 00475 g001
Figure 2. TEM images of (a) s-Ni-P/Hβ and (b) s-Ni-P/Ce-β.
Figure 2. TEM images of (a) s-Ni-P/Hβ and (b) s-Ni-P/Ce-β.
Catalysts 14 00475 g002
Figure 3. XPS spectra of s-Ni-P/Hβ(3) and s-Ni-P/Ce-β(3) in the region of (a) Ni 2p3/2 and (b) P 2p3/2.
Figure 3. XPS spectra of s-Ni-P/Hβ(3) and s-Ni-P/Ce-β(3) in the region of (a) Ni 2p3/2 and (b) P 2p3/2.
Catalysts 14 00475 g003
Figure 4. NH3-TPD curves of s-Ni-P/Hβ(3) and s-Ni-P/Ce-β(3).
Figure 4. NH3-TPD curves of s-Ni-P/Hβ(3) and s-Ni-P/Ce-β(3).
Catalysts 14 00475 g004
Figure 5. Catalytic performance of (a) s-Ni-P/Hβ(3), (b) s-Ni-P/Ce-β(3), and (c) Ni/Hβ in APHDO of phenol.
Figure 5. Catalytic performance of (a) s-Ni-P/Hβ(3), (b) s-Ni-P/Ce-β(3), and (c) Ni/Hβ in APHDO of phenol.
Catalysts 14 00475 g005
Figure 6. Catalytic performance of (a) s-Ni-P/Hβ and (b) s-Ni-P/Ce-β(3) in OPHDO of phenol.
Figure 6. Catalytic performance of (a) s-Ni-P/Hβ and (b) s-Ni-P/Ce-β(3) in OPHDO of phenol.
Catalysts 14 00475 g006
Figure 7. XRD patterns of fresh and spent s-Ni-P/Hβ(3) catalysts.
Figure 7. XRD patterns of fresh and spent s-Ni-P/Hβ(3) catalysts.
Catalysts 14 00475 g007
Figure 8. Kinetic analysis of s-Ni-P/Hβ(3) and s-Ni-P/Ce-β(3) in APHDO of phenol at (a) 250, (b) 275, (c) 300, and (d) 325 °C.
Figure 8. Kinetic analysis of s-Ni-P/Hβ(3) and s-Ni-P/Ce-β(3) in APHDO of phenol at (a) 250, (b) 275, (c) 300, and (d) 325 °C.
Catalysts 14 00475 g008
Figure 9. Kinetic analysis of s-Ni-P/Hβ(3) and s-Ni-P/Ce-β(3) in OPHDO of phenol at (a) 150, (b) 175, (c) 200, and (d) 225 °C.
Figure 9. Kinetic analysis of s-Ni-P/Hβ(3) and s-Ni-P/Ce-β(3) in OPHDO of phenol at (a) 150, (b) 175, (c) 200, and (d) 225 °C.
Catalysts 14 00475 g009
Figure 10. The kinetic study of the (a) APHDO and (b) OPHDO of phenol over s-Ni-P/Hβ(3) and s-Ni-P/Ce-β(3).
Figure 10. The kinetic study of the (a) APHDO and (b) OPHDO of phenol over s-Ni-P/Hβ(3) and s-Ni-P/Ce-β(3).
Catalysts 14 00475 g010
Table 1. The physical and chemical properties of s-Ni-P/Hβ(3) and s-Ni-P/Ce-β(3).
Table 1. The physical and chemical properties of s-Ni-P/Hβ(3) and s-Ni-P/Ce-β(3).
CatalystSBET (m2·g−1)XRD PhaseParticle Size (nm) 1Amount of Acid Sites (mmol·g−1) 2
WeakMediumStrongTotal
358.3--
Ce-β230.4--
s-Ni-P/Hβ(3)295.8Ni-Ni3P-Ni12P517.30.630.770.391.79
s-Ni-P/Ce-β(3)153.7Ni-Ni3P-Ni12P511.61.29-0.151.44
1 Measured from TEM images. 2 Calculated from NH3-TPD curves.
Table 2. Kinetic analysis of APHDO over s-Ni-P/Hβ(3) and s-Ni-P/Ce-β(3).
Table 2. Kinetic analysis of APHDO over s-Ni-P/Hβ(3) and s-Ni-P/Ce-β(3).
CatalystRate Constants (mol·g−1·min−1) 1Apparent Activation
Energy (kJ·mol−1) 2
250275300325
s-Ni-P/Hβ(3)3.04.86.18.936.8
s-Ni-P/Ce-β(3)4.46.57.911.030.7
1 Calculated by pseudo-first order. 2 Calculated by Arrhenius equation.
Table 3. Kinetic analysis of OPHDO over s-Ni-P/Hβ(3) and s-Ni-P/Ce-β(3).
Table 3. Kinetic analysis of OPHDO over s-Ni-P/Hβ(3) and s-Ni-P/Ce-β(3).
CatalystRate Constants (mol·g−1·min−1) Apparent Activation
Energy (kJ·mol−1)
150175200225
s-Ni-P/Hβ(3)5.510.415.123.733.5
s-Ni-P/Ce-β(3)6.912.016.924.929.4
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

Ma, L.; Li, Y.; Yu, Z.; Zou, J.; Jing, Y.; Wang, W. The Hydrodeoxygenation of Phenol over Ni-P/Hβ and Ni-P/Ce-β: Modifying the Effects in Dispersity and Acidity. Catalysts 2024, 14, 475. https://doi.org/10.3390/catal14080475

AMA Style

Ma L, Li Y, Yu Z, Zou J, Jing Y, Wang W. The Hydrodeoxygenation of Phenol over Ni-P/Hβ and Ni-P/Ce-β: Modifying the Effects in Dispersity and Acidity. Catalysts. 2024; 14(8):475. https://doi.org/10.3390/catal14080475

Chicago/Turabian Style

Ma, Lin, Yan Li, Zhiquan Yu, Jie Zou, Yingying Jing, and Wei Wang. 2024. "The Hydrodeoxygenation of Phenol over Ni-P/Hβ and Ni-P/Ce-β: Modifying the Effects in Dispersity and Acidity" Catalysts 14, no. 8: 475. https://doi.org/10.3390/catal14080475

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