The hydrogenation of aromatic-naphthalene with Ni2P/CNTs

Minzhi Ruana, Jun Guana, Demin Hea, Tao Mengc and Qiumin Zhang*ab
aSchool of Chemical Engineering, Dalian University of Technology (DUT), No. 2 Linggong Road, Dalian 116024, P. R. China. E-mail: zhangqm@dlut.edu.cn; Fax: +86 0411 84986150; Tel: +86 0411 84986150
bState Key Laboratory of Fine Chemicals, Dalian University of Technology (DUT), China
cBeijing Guodian Longyuan Environmental Engineering Co., LTD, Building No. 1, Yard No. 16, Xisihuan Mid Road, Haidian District, Beijing 100039, PR China

Received 26th March 2015 , Accepted 16th June 2015

First published on 16th June 2015


Abstract

Ni2P/CNTs was synthesized using an impregnation method. XPS revealed that CNTs could affect the electronic properties of bulk Ni2P. The catalyst shows superior activity for HYD of naphthalene with a conversion of 99%, and demonstrates superior tolerance towards potential catalyst poisons, which is higher than Ni/CNTs with a conversion of 89%.


Due to the increasingly low quality of petroleum feedstocks and stringent environmental regulations, it is of paramount importance to develop new environmentally friendly catalysts for hydrotreatment. The hydrogenation (HYD) and hydrodesulfurization (HDS) reactions, which transfer polycyclic aromatic hydrocarbons with sulfur into high RON oil, underpin many clean-energy technologies. Although Pt or Pd have been used in hydrotreatment catalysts, they have poor performances when in contact with sulfur.1 Furthermore Pt and Pd are luxurious and relatively infrequent in the Earth’s shell limiting the use of them in energy systems deployed on a global scale. More recently, nanoparticulate films of Ni2P composed of inexpensive materials to show high HDS activity. As the nickel phosphorus catalyst was firstly introduced by Oyama, it has aroused lots of interest all around the world as it has been used in energy applications with high performances of HDS and hydrodenitrogenation (HDN) activity and even has outstanding performances in fuel cells. This indicated that the sequence of HYD activity over transition metal phosphides is as follows: Fe2P < CoP < MoP < WP < Ni2P, where the Ni2P catalyst was found to be the most active.2–4 Among such classes of catalysts, various carbon nanostructures, including carbon, carbon nanotubes, and graphene and so on, have been used for catalyst carriers which have excellent performances.5,6 However, carbon nanotubes (CNTs) with unique properties have experienced limited use as a support material for Ni2P, and the potential industrial application value of Ni2P/CNTs in aromatic HYD has not been sufficiently explored. Herein, we have thus applied Ni2P catalysts supported on carbon nanotubes to examine the effect of HYD. CNTs was reported to have huge specific areas, therefore they have a better hydrogen storage function and strictly mesoporous nature, differing greatly from the classical activated carbon support.7–13 The CNTs was functionalized using an acid mixture which led to the unique CNT properties, and the Ni2P/CNTs particles were synthesized using an impregnation method.14

X-ray diffraction patterns of reduced Ni2P/CNTs and Ni2P samples are presented in Fig. 1a. The Ni2P loadings were varied from 5 to 40 wt%, and the Ni/P mole ratios were varied from 0.5 to 2.0. With the increase of Ni2P loadings and Ni/P mole ratios, the sharp and symmetric peaks evidently indicate that the samples were well crystallized. The patterns all display a broad feature at 20–30° and 48.8° due to the carbon nanotube supports. For all of the catalysts, the peaks at 2θ = 40.6°, 44.5°, 47.1°, 54.1°, 54.8° (PDF: 74-1385) can be ascribed to Ni2P, and no additional phase of Ni and P is observed, indicating that the active phase formed is mainly Ni2P. The 30 wt% Ni2P/CNTs sample (Fig. 1b) was substantiated by energy dispersive spectroscopy (EDS) in Fig. 1c.


image file: c5ra05364a-f1.tif
Fig. 1 (a) XRD patterns of catalyst samples with different wt% of Ni2P (Ni/P = 1.25, atomic) and Ni/P proportion (30 wt% Ni2P); (b) SEM image of 30 wt% Ni2P/CNTs; (c) EDS of 30 wt% Ni2P/CNTs; (d) FTIR of catalyst samples.

The IR spectrum of the acid-treated CNTs and Ni2P/CNTs catalysts is shown in Fig. 1d. The bands appearing at 1000 cm−1 can be assigned to the Ni2P, and the bands appearing at 1635 and 3450 cm−1 can be assigned to the C[double bond, length as m-dash]O and O–H stretching vibration bands, which confirm the presence of functional groups in the acid-treated CNTs, and could help adsorb active metal on the surface.15

Table 1 shows the physical properties of the Ni2P/CNTs catalysts, including the Ni2P loading content, average Ni2P size and BET surface area. The results showed that the BET surface area of the catalysts declined with the increase in the amount of Ni2P loading. This indicated that the Ni2P particles occupied the surface holes of the CNTs. TEM observations indicate that the size and the morphology of the Ni2P/CNTs composites differ from those of the CNTs. Most of the Ni2P particles are displayed outside of the CNTs, the loading content could affect the Ni particle size and its distribution.15 In addition, as the histograms in Fig. 2 show, the mean particle size increases from 6.7 nm to 20.0 nm with the increasing loading content. In the following section, it was found that a smaller Ni2P particle size with a higher dispersion contributes to the inhibition of the agglomeration of Ni2P particles, which should favour the hydrogenation reaction.

Table 1 BET results and pore structure parameters of Ni2P/CNTs catalysts
Samples Surface areaa (m2 g−1) Pore volumea (cm3 g−1) dXRDb (nm) dTEMc (nm)
a Evaluated from N2 adsorption–desorption isotherms.b Calculated using the Scherrer equation from the XRD.c Measured using TEM.
10 wt% 171 0.33 13.6 6.7
20 wt% 151 0.31 14.6 11.0
30 wt% 143 0.26 21.9 13.1
40 wt% 94 0.23 28.0 20.0



image file: c5ra05364a-f2.tif
Fig. 2 TEM images and size distribution histograms: (a) 10 wt% Ni2P/CNTs, (b) 20 wt% Ni2P/CNTs, (c) 30 wt% Ni2P/CNTs and (d) 40 wt% Ni2P/CNTs.

The XPS Ni 2p and P 2p spectra of 30 wt% Ni2P/CNTs catalysts are presented in Fig. 3a and b. For samples (1)–(3), the Ni 2p core level spectrum includes two contributions (Fig. 3a and b), which correspond to Ni 2p3/2 and Ni 2p1/2, respectively. The former one, is assigned to Niδ+ in the Ni 2p phase and is centered at 853.31 eV, and the later one at 857.9 eV is assigned to oxidized Ni. With regards to the P 2p3/2 binding energy, the peaks of the P 2p electron binding energy at 129.87 eV and 133.98 eV can be assigned to oxidized and elemental P, respectively.


image file: c5ra05364a-f3.tif
Fig. 3 XPS spectra for Ni2P, spent Ni2P/CNTs and fresh Ni2P/CNTs: (a) Ni 2p core level spectra and (b) P 2p core level spectra. The two stable binding sites for Ni2P on CNTs: (c) atop–atop site and (d) bridge–bridge site.

The peaks in the XPS spectrum for the Ni2P catalyst have been assigned previously.16,17 For Ni2P, the binding energies at 857.9 and 134.8 eV are assigned to Ni2+ and P5+ species. The XPS spectrum for the 30 wt% Ni2P/CNTs catalyst commonly mirrors that of the Ni2P catalyst. The fresh Ni2P/CNTs sample displays two distinct peaks, whose positions correspond to the Ni–P and Ni–Ni distances in bulk Ni2P. The P (2p) region shows one significant difference: the XPS spectrum of the Ni2P catalyst has a very intense peak at 132.04 eV but there are two peaks in the XPS spectrum of Ni2P/CNTs that are consistent with the binding energy for phosphorus in the CNTs. Some of the phosphorus impregnated onto the CNTs apparently reacts with the support to form Ni–P-CNTs. The spectrum of fresh Ni2P/CNTs differs from the Ni2P sample, indicating that the Ni+ peak shifted from 855.4 to 853.31 eV (Fig. 3a), and the binding energy assigned to the P5+ species shifted from 131.96 to 129.87 eV (Fig. 3b). The shift of binding energies of the surface Ni+ and P5+ species suggests the existence of an electronic effect caused by the surface –OH or –COOH, or CNTs species, which may result in the electron enrichment of the surface Ni and P species. As it is revealed in Fig. 3c and d, the Ni atom is placed on an atop site instead of at a hole site, there are two stable binding sites for a Ni2 dimer on a carbon nanotube wall: an atop–atop site and a bridge–bridge site. The Ni2 dimer when placed on a nanotube wall has a tendency to separate in a reaction mediated by the underlying carbon atoms. Thus, the number of Ni2 configurations on a tube varies with the curvature and Ni2 exhibits a larger range of Ni–Ni bond lengths (2.34–2.78 Å).18 A significant shift in binding energy peaks assigned to the spent phosphorus species in Ni2P/CNTs was detected, illustrating that the addition of functionalized CNTs could affect the electronic properties of bulk Ni2P.19

The hydrotreatment reaction results for Ni2P/CNTs in the HYD of naphthalene and 1-methyl naphthalene are summarized in Tables 2 and 3. Mass balance is 100 ± 5% for each reaction type. The reaction of naphthalene on 30 wt% Ni2P/CNTs (Ni/P = 1.25, mol) occurs with a very high and stable conversion of 99%, which is much higher than for the other samples or that of the Ni2P/SiO2 and NiP/C catalysts.8,20 This may be attributed to 30 wt% Ni2P/CNTs having a higher BET surface area which could provide more Ni2P active sites on the surface of the CNTs (Table 1), furthermore, the Ni2P particles of the 30 wt% samples with a small size were well dispersed on the CNTs (Fig. 2). There are three major products formed from naphthalene on these catalysts: cis-decalin, trans-decalin and tetralin, compared with the Ni2P/SiO2 catalysts, Ni2P/CNTs gave a higher trans-decalin selectivity of 69% and the proportion of cis-decalin and trans-decalin is nearly 1[thin space (1/6-em)]:[thin space (1/6-em)]4, indicating that it favours the hydrogenation pathway. In comparison with naphthalene, tetralin, the final product of HYD of 1-methyl naphthalene is less than 50%, this is due to the steric hindrance effect caused by the methyl group. The Ni/CNTs catalyst exhibits less activity with a conversion of 84%. Obviously, the existence of P contributes to the HYD; the joint action of the Ni–P-CNTs effect could be explained by the XPS results. Liu et al. also confirmed that the P sites of the phosphide of Ni2P play a complex and significant role. First, the Ni–P bonds produce a weak ligand effect that allows a reasonably high activity of HYD. Second, the amount of Ni active sites on the surface decreases owing to an ensemble effect of P, which prevents the reaction system from deactivating. Third, the P sites are spectators and provide moderate bonding and the H adatoms are essential for HYD.21

Table 2 Catalytic hydrogenation of naphthalene and 1-methyl naphthalenea
Samples Raw materials Products Selectivity [%]
a Reaction conditions: reactions were carried out with naphthalene or 1-methyl naphthalene (5 wt% in decane), at 613 K, 4.0 MPa, vapour–liquid ratio = 600, LHSV = 4.0 h−1. The Ni loading of 20 wt% Ni/CNTs is equal to the Ni loading of 30 wt% Ni2P/CNTs. Reactions were carried out in a continuous-flow microreactor.
30 wt% Ni2P/CNTs Naphthalene cis-Decalin 14
trans-Decalin 69
Tetralin 17
30 wt% Ni2P/CNTs 1-Methyl naphthalene 1-Methyl tetralin 17
5-Methyl tetralin 36
1-Methyl decalin 47
20 wt% Ni/CNTs Naphthalene cis-Decalin 0
trans-Decalin 28
Tetralin 61


Table 3 Hydrotreatment performance of the Ni2P/CNTs catalystsa
Entry Ni2P (wt%) Ni/P (atomic) LHSV (h−1) Conversion [%]
a Reaction conditions: reactions were carried out with naphthalene (5 wt%), at 613 K, 4.0 MPa.
1 40 1.25 4 91
2 30 1.25 4 >99
3 20 1.25 4 65
4 10 1.25 4 30


We next investigated the durability and the resistance to sulfur and nitrogen solutions of the catalysts. As revealed in Fig. 4, after contact with the sulfur and nitrogen solutions, the initial HYD conversion dropped from 99% to 85%, however when the solution was switched to naphthalene, the catalyst can return to its former level, demonstrating superior tolerance towards potential catalyst poisons, such as DBT and quinoline.


image file: c5ra05364a-f4.tif
Fig. 4 Effect of adding DBT or quinolone to HYD. The content of HDS and HDN intermediates of DBT and quinoline was 1.0 wt% in 5.0 wt% naphthalene reactant.

In summary, a novel Ni2P/CNTs catalyst was developed for HYD. The catalyst exhibited excellent activity and stability, and was successfully used for HDS or HDN, showing superior performance to a Pt–Pd catalyst. As aromatic compounds are important components of fuels, this work is a significant step towards the development of more active and practical catalysts for the upgrading of coal-tar and bio-oil.

Experimental section

Multi-carbon nanotubes were purchased from CAS (Chengdu, China). Concentrated nitric acid (70%, Aladdin) was used to functionalize the CNTs at 60 °C for 6 h. Nickel phosphate compounds loaded with CNTs were prepared by the following route: nickel nitrate (60 mg, 99%, Aladdin) and ammonium hydrogen phosphate (20 mg, 99%, Aladdin) were dissolved in deionized water. Functionalized CNTs (100 mg) were then mixed into the solution. Ultrasonication was carried out for 6 h and followed by impregnation at room temperature. The sample was then collected and calcined in N2 at 793 K. Reduction was conducted in 95% H2 mixed with 5% He with a total flow of 100 mL min−1 at 793 K. Subsequently, the calcined sample was pelletized and sieved for the HYD reaction.

Acknowledgements

The authors thank the financial support of the Major Project of the National Energy Administration (NY20130302513-1).

Notes and references

  1. H. Song, J. Wang, Z. Wang, H. Song, F. Li and Z. Jin, J. Catal., 2014, 311, 257 CrossRef CAS PubMed.
  2. S. T. Oyama, J. Catal., 2003, 216, 343 CrossRef CAS.
  3. A. Infantes-Molina, J. A. Cecilia, B. Pawelec, J. L. G. Fierro, E. Rodríguez-Castellón and A. Jiménez-López, Appl. Catal., A, 2010, 390, 253 CrossRef CAS PubMed.
  4. J. Chang, L. Feng, C. Liu, W. Xing and X. Hu, Angew. Chem., Int. Ed., 2014, 53, 122 CrossRef CAS PubMed.
  5. Y. Shu and S. T. Oyama, Carbon, 2005, 43, 1517 CrossRef CAS PubMed.
  6. Y. Lu, J. Tu, Q. Xiong, Y. Qiao, J. Zhang, C. Gu, X. Wang and S. X. Mao, Chem.–Eur. J., 2012, 18, 6031 CrossRef CAS PubMed.
  7. Y. Lu, X. Wang, Y. Mai, J. Xiang, H. Zhang, L. Li, C. Gu, J. Tu and S. X. Mao, J. Phys. Chem. C, 2012, 116, 22217 CAS.
  8. A. C. Dillon, K. M. Jones, T. A. Bekkedahl, C. H. Kiang, D. S. Bethune and M. J. Heben, Nature, 1997, 386, 377–379 CrossRef CAS PubMed; S. T. Oyama, X. Wang, Y. K. Lee and W. J. Chun, J. Catal., 2004, 221, 263 CrossRef.
  9. E. G. Rodrigues, S. A. C. Carabineiro, J. J. Delgado, X. Chen, M. F. R. Pereira and J. J. M. Órfão, J. Catal., 2012, 285, 83–91 CrossRef CAS PubMed.
  10. F. Zhu, G. Ma, Z. Bai, R. Hang, B. Tang, Z. Zhang and X. Wang, J. Power Sources, 2013, 242, 610–620 CrossRef CAS PubMed.
  11. D. Z. Mezalira and M. Bron, J. Power Sources, 2013, 231, 113–121 CrossRef CAS PubMed.
  12. V. Lordi, N. Yao and J. Wei, Chem. Mater., 2001, 13, 733–737 CrossRef CAS.
  13. S. M. Bachilo, L. Balzano, J. E. Herrera, F. Pompeo, D. E. Resasco and R. B. Weisman, J. Am. Chem. Soc., 2003, 125, 11186–11187 CrossRef CAS PubMed.
  14. M. Moula, S. Suzuki, W. Chun, S. Otani, S. T. Oyama and K. Asakura, Surf. Interface Anal., 2006, 38, 1611 CrossRef CAS PubMed.
  15. M. Zhou, H. Zhu and L. Niu, et al., Catal. Lett., 2014, 144(2), 235–241 CrossRef CAS PubMed.
  16. D. Kanama, Surf. Sci. Spectra, 2001, 8, 220 CrossRef CAS.
  17. P. Liu, J. A. Rodriguez, Y. Takahashi and K. Nakamura, J. Catal., 2009, 262, 294 CrossRef CAS PubMed.
  18. M. Menon, A. N. Andriotis and G. E. Froudakis, Chem. Phys. Lett., 2000, 320, 425 CrossRef CAS.
  19. D. S. Misra, A. Misra, P. K. Tyagi and M. K. Singh, Diamond Relat. Mater., 2006, 15, 385 CrossRef PubMed.
  20. S. Yang and R. Prins, Chem. Commun., 2005, 4178 RSC.
  21. P. Liu, J. A. Rodriguez and T. Asakura, et al., J. Phys. Chem. B, 2005, 109(10), 4575–4583 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2015