Next Article in Journal
Pd-HPW/SiO2 Bi-Functional Catalyst: Sonochemical Synthesis, Characterization, and Effect on Octahydroquinazolinone Synthesis
Previous Article in Journal
Performance Study of Zn-Co-Ni/AC Catalyst in Acetylene Acetylation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hydrotreating of Waste Tire Pyrolysis Oil over Highly Dispersed Ni2P Catalyst Supported on SBA-15

1
Green Carbon Research Center, Korea Research Institute of Chemical Technology (KRICT), 141 Gajeongro, Yuseoung, Daejeon 34114, Korea
2
Climate Change Research Division, Korea Institute of Energy Research, 152 Gajeong-ro, Yuseong-gu, Daejeon 34129, Korea
3
Laboratory of Advanced Catalysis for Energy and Environment, Department of Chemical Engineering, Dankook University, 126 Jukjeon, Yongin 448-701, Korea
*
Author to whom correspondence should be addressed.
Submission received: 16 September 2021 / Revised: 18 October 2021 / Accepted: 19 October 2021 / Published: 22 October 2021
(This article belongs to the Special Issue State-of-the-Art in Catalysts for the Hydroconversion of Heavy Oils)

Abstract

:
A highly dispersed nickel phosphide catalyst supported on SBA-15 was prepared and tested for the hydrotreating of tire pyrolysis oil (TPO). Physicochemical properties of the prepared catalyst were analyzed by CO uptake chemisorption, BET, TEM, and X-ray diffraction (XRD). An advanced technique with gas chromatography equipped with mass spectrometry and atomic emission detector was applied to investigate carbon-, sulfur-, and nitrogen-containing compounds in TPO. Hydrotreating tests were carried out in a fixed-bed continuous flow reactor at 350 °C, 3.0 MPa, and LHSV of 0.5 h−1. The Ni2P/SBA-15 exhibited an HDS conversion of 89.3% and an HDN conversion of 60.7%, which was comparable to the performance of a commercial NiMoS catalyst under the same conditions.

1. Introduction

Global waste generation has grown steadily around the world, with 2 billion tons of municipal solid waste produced annually [1]. In particular, as concerns of massive waste tires rise, a chemical recycling technology for them has been extensively studied to overcome the environmental and economic issues [2,3]. Among the chemical recycling processes of waste tires, pyrolysis gives a high yield of oil, up to 38%, at moderate temperatures of around 500 °C [4,5]. The tire pyrolysis oil (TPO) has been regarded as a sustainable energy resource owing to a large amount of aromatics, olefins, and other substances, which can be used as valuable chemicals [6,7,8]. In particular, its high HHV (37–44 MJ/kg) led the way in the use of alternative fuels such as gasoline and diesel. However, TPO cannot be directly used as a fuel due to the high content of sulfur and nitrogen compounds derived from the synthesis processes of rubbers [3,8,9,10].
Catalytic hydrotreating is a promising process to produce usable liquid fuels with simultaneous eliminations of sulfur and nitrogen in the presence of H2. Transition metal sulfide catalysts (NiMoS and CoMoS) have been commercially applied due to their hydrogenation ability in the hydrodesulfurization (HDS) and hydrodenitrogenation (HDN) processes [11,12,13]. Recently, numerous studies of transition metal phosphide catalysts have been considered as a potential candidate, showing high stability and resistance to aromatics and nitrogen-containing species [14,15,16,17]. However, metal phosphide catalysts have not yet been applied to the hydrotreating of TPO. Furthermore, although a number of studies for the hydrotreating of TPO have been reported [18,19,20], they did not provide a molecular-level qualitative analysis due to the complexity of TPO’s components, which require high-performance analysis equipment.
In the present study, a highly dispersed nickel phosphide supported on SBA-15 was prepared, and its activity of TPO hydrotreating was investigated for the first time. The activity was compared with a commercially available NiMoS catalyst, demonstrating Ni2P as a potential catalyst for hydrotreating TPO. Furthermore, an advanced technique was applied to investigate the carbon-, sulfur-, and nitrogen-containing compounds in TPO and HYD products.

2. Results and Discussion

2.1. Characterizations of the Supported Nickel Phosphide Catalysts

Table 1 summarizes the physicochemical properties of the prepared samples. The BET surface areas of Ni2P/SBA-15 and Ni2P/SiO2 were 687 and 124 m2 g−1. After loading, the existence of metal phosphide and remaining phosphorus on the support led to a decrease in the surface area. The CO uptakes of Ni2P/SBA-15 and Ni2P/SiO2 were 84.3 and 71.2 μmol g−1.
Figure 1 shows the XRD patterns of Ni2P/SBA-15 and Ni2P/SiO2 catalysts. The broad feature centered at 2θ = 22° is originated from amorphous silica. Both patterns show three main peaks at 40.5°, 44.8°, and 47.5°, corresponding to the characteristic peaks of hexagonal Ni2P (space group: P .62m, D2 3 h, Struckturbericht notation: revised C22) [21]. The crystallite sizes were calculated by Scherrer’s equation and presented sizes of 3.8 nm for Ni2P/SBA-14 and 9.1 nm for Ni2P/SiO2, indicating that the Ni2P particles on the SBA-15 support were more highly dispersed than those on the SiO2 support.
Figure 2 shows the TEM images and particle size distributions of Ni2P/SBA-15 and Ni2P/SiO2 catalysts. The particle size distributions over 100 particles showed that the Ni2P particles on SBA-15 had a narrow size distribution with an average size of around 5.3 nm, while the Ni2P particles on SiO2 had a broad size distribution with an average size of around 9.7 nm. This result is in reasonable agreement with the crystallite sizes of Ni2P determined by XRD in Figure 1.

2.2. Characterizations of the Tire Pyrolysis Oil (TPO)

Tire pyrolysis oil (TPO) distilled at 320 °C was supplied from a refinery in the Republic of Korea. The specifications of the TPO used in this study are summarized in Table 2. The TPO contains 14,500 ppm of sulfur, 4070 ppm of nitrogen, and 41 wt.% of aromatic compounds (27 wt.% of mono-aromatics, 8 wt.% of di-aromatics, and 6 wt.% of tri+- aromatics).
To identify each heteroatom-containing compound of TPO, GC/AED (atomic emission detector) and GC/MS were used. In particular, GC/AED is a powerful technique in the analysis to assign the heteroatoms in complex oil mixtures because it selectively analyzes each element. The analyzed results of TPO and assigned compounds are demonstrated in Figure 3. The major compounds are benzene, toluene, xylene, and their derivatives. It is especially noteworthy that limonene is the most abundant compound in TPO. These results are compatible with previous reports [4,8], in which TPOs were observed to contain a significant amount of naphtha-like fractions along with about 15% limonene. Additionally, it is clear that the abundant sulfur-containing compounds of TPO are thiophene derivatives and benzothiazole, as shown in Figure 3b. In AED results of nitrogen-containing compounds (Figure 3b), the three major peaks were observed at the retention times of 7, 15, and 21 min, corresponding to aniline, benzothiazole, and dimethylquinoline, respectively. Note that the benzothiazole corresponded to both S- and N-selective analysis. It is reported that benzothiazole is derived from the vulcanizing accelerator in the form of benzothiazole derivatives [7,22]. It is a representative S and N compound for the hydrotreating of TPO [23].

2.3. Activity Test

Figure 4 shows the AED chromatographs of hydrotreating products at 3.0 MPa, 350 °C, LHSV of 0.5 h−1, and time on stream of 12 h on NiMoS, Ni2P/SBA-15, and Ni2P/SiO2 catalysts. In the AED analysis of carbon-containing compounds (Figure 4a), the overall intensity of peaks was maintained after the hydrotreating process, but its products exhibited increased intensity in the peaks at 5–10 min, which are assigned as mono- or di-aromatics. These results indicate that the di- and tri-aromatics were converted to mono-aromatics. In the AED analysis of sulfur and nitrogen-containing compounds (Figure 4b), the hydrotreating products presented a substantial decrease in overall intensity compared to TPO. In the hydrotreating products, the peak assigned as benzothiazole was completely diminished, whereas the peaks at around 20–30 min were maintained. In the AED analysis of nitrogen-containing compounds (Figure 4c), the hydrotreating products displayed a substantial decrease in overall intensity compared to TPO. The peak assigned to benzothiazole was completely reduced, while those of aniline and dimethylquinoline remained.
Table 3 displays a comparison of reactivity of sulfur and nitrogen-containing compounds in TPO at 3.0 MPa, 350 °C, LHSV of 0.5 h−1, and time on stream of 12 h on NiMoS, Ni2P/SBA-15, and Ni2P/SiO2 catalysts. Total HDS and HDN conversions were obtained by the ASTM method, and the conversions of main sulfur and nitrogen compounds were calculated based on the AED intensity in Figure 4. The sulfur and nitrogen compounds were classified based on the retention time (light S and N (B.P. < 230 °C, R.T. < 15.3 min), benzothiazole (R.T. = 15.3 min), and heavy S and N (B.P. > 230, R.T. > 15.3 min)) and each conversion is summarized in Table 3. Additionally, the HDN conversions of three main N compounds (aniline, benzothiazole and dimethylquinoline) are listed in Table 3. The overall activities of HDS and HDN followed the order of NiMoS ≥ Ni2P/SBA-15 > Ni2P/SiO2 at the applied conditions, and the commercial NiMoS catalyst exhibited the highest activity of around 94.5% total sulfur conversion, and 73% total nitrogen conversion. Furthermore, there is no significant change in the HDS and HDN conversion at the applied reaction conditions, as shown in Figure 5. The Ni2P/SBA-15 showed comparable activity with commercial NiMoS, and higher activity than Ni2P/SiO2, suggesting that highly dispersed Ni2P particles as confirmed by the results of CO uptakes, TEM, and XRD, led to the enhancement of the hydrotreating ability. This result can be explained by the structure of the Ni2P catalyst that consists of two Ni sites, tetrahedral Ni(1) sites and square pyramidal Ni(2) sites. The latter sites responsible for the high catalytic activity in the hydrogenation are predominant in highly dispersed samples [24,25].
In comparing reactivity in sulfur and nitrogen compounds, benzothiazole was the most reactive compound, with 100% conversion. This result is in accordance with previous work of hydrotreating scrap tire pyrolysis oil on NiMo catalysts [10]. They reported that benzothiazole was removed entirely, while substituted DBTs substantially remained in the hydrotreating products. As expected, light sulfur compounds such as thiophene derivatives showed higher reactivity than heavy sulfur compounds [26]. On the other hand, the nitrogen compounds had an opposite tendency compared to the light nitrogen compounds, including aniline, which showed the lowest reactivity among the nitrogen compounds. Thus, the reactivity of the main nitrogen compounds followed the order of benzothiazole > dimethylquinoline > aniline, suggesting that the basicity of the nitrogen compounds significantly affected their reactivity (benzothiazole = 2.5 pKa, aniline = 4.58 pKa and quinoline = 4.92 pKa). The nitrogen compounds with high basicity would be strongly adsorbed on the active sites and then limit the surface reaction, resulting in low reactivity. Apart from the basicity, it is presumable that aniline would be one of the intermediate products of benzothiazole, resulting in the low conversion of aniline.
Figure 6 shows the aromatic compound distributions of TPO and hydrotreating products, which were analyzed by high-pressure liquid chromatography with fluorescence detection (HPLC-FD). The TPO used in this study contains 41 wt.% of total aromatics consisting of 27, 8, and 6 wt.% of mono-, di-, and tri+-aromatics, respectively. NiMoS exhibited similar total aromatics, but a higher concentration of mono-aromatics and lower concentration of tri+-aromatics than those of TPO, indicating the transformation of poly-aromatics to mono-aromatics. The supported Ni2P catalysts presented a substantial decrease in total aromatics and tri+-aromatic resulting from the full hydrogenation of aromatics to cycloalkane derivatives under the given reaction condition.

3. Experimental Section

3.1. Synthesis and Characterization of Supported Nickel Phosphide Catalysts

The supported nickel phosphide catalysts were synthesized by the incipient wetness impregnation method, as reported in [27]. The SiO2(EH-5) was commercially obtained, and SBA-15 was prepared by the method reported in [17]. The impregnating solution was obtained by dissolving Ni(NO3)2 6H2O (Alfa Aesar, 98%) and (NH4)2HPO4 (Samchun, 99%) in deionized water. The amount of nickel was 1.5 mmol per 1 g of support, and the initial Ni/P ratio was fixed at 1/2. The supported nickel phosphate precursor was dried at 120 °C overnight and then calcined in the air at 500 °C for 4 h. The resulting precursor was pelletized and sieved to a size of 650–1180 μm, followed by reduction at 600 °C for 2 h in H2 flow (1000 mL min−1 per 1 g of sample). The temperature of reduction was determined by an H2-TPR measurement, as reported previously [17]. After reduction, the sample was cooled in He and then passivated in 0.5% O2/He flow at 25 °C for 4 h. A commercial NiMoS catalyst (TK-607 BRIM) was supplied as a comparison.
X-ray diffraction patterns (XRD) of the prepared phosphide catalysts were measured with a diffractometer (Bruker D8) operated at 40 kV and 100 mA, using a Cu-Kα monochromatic X-ray source. HR-TEM images were collected using a JEM-2100F, 200 kV microscope. Chemisorption uptakes of CO were measured by a pulse injection method using a re-reduced sample at 550 °C for 3 h in H2 flow (100 mL min−1). A Micromeritics ASAP 2010 micropore (Norcross, GA, USA) was used to measure the specific surface area of the passivated samples using N2 adsorption isotherms at 77 K from the linear portion of BET plots (P/P0 = 0.01–0.20).

3.2. Activity Test of Hydrotreating of Tire Pyrolysis-Oil (TPO)

The catalytic activity was tested in a fixed-bed continuous flow reactor at a 3.0 MPa, 350 °C, and liquid hour space velocity of 0.5–1.0 h−1. Prior to the activity test, the passivated Ni2P catalysts were re-reduced at 350 °C for 2 h under H2 flow 100 cm3 min−1, and NiMoS catalyst was pretreated at 300 °C for 2 h under 100 cm3 min−1 of 10% DMDS (dimethyl disulfide)/H2. The tire pyrolysis oil was fed by means of a liquid pump along with hydrogen flow (100 cm3 min−1), and liquid products were collected at 4 h intervals. The total contents of sulfur and nitrogen in the liquid product were analyzed in an Antek 9000 NS analyzer using the combustion/fluorescence technique. Polycyclic aromatic hydrocarbons were monitored and quantified by high-pressure liquid chromatography with fluorescence detection (HPLC-FD). Total sulfur and nitrogen contents were analyzed in an Antek 9000 NS analyzer. A Hewlett-Packard gas chromatograph (HP6890) equipped with an atomic emission detector (G2350A) was used to identify the carbon, sulfur, and nitrogen-containing compounds in TPO and hydrotreating products.

4. Conclusions

A highly dispersed nickel phosphide catalyst supported on SBA-15 was successfully prepared and tested for the hydrotreating (HYD) of tire pyrolysis oil (TPO). The activity tests reveal that Ni2P/SBA-15 displayed comparable activity with commercial NiMoS catalyst and superior activity to Ni2P/SiO2 in HDS and HDN. Furthermore, it is noted that benzothiazole, which contains both sulfur and nitrogen, is the most reactive compound in TPO. A substantial decrease in tri+-aromatic hydrocarbons was observed in all catalysts. In particular, nickel phosphide catalysts exhibited superior hydrogenation activity for aromatic hydrocarbons. The analysis by GC-MS and AED demonstrated the carbon-, sulfur-, and nitrogen-containing compounds in TPO and hydrotreating products.
Overall results of catalytic activity gave an insight into the application of TPO in the field of chemical industry. The similarity of physical and chemical properties between hydrotreated TPO and commercial fuel oil would lead to its various practical applications in the petroleum industry. In particular, the finding of limonene as a major compound has the potential to expand the use of TPO.
Furthermore, the advanced analysis techniques provide a fundamental aspect in the hydrogenation of TPO. This can contribute to atomic-level catalyst design to enhance the selective conversion of a certain compound in TPO.

Author Contributions

Y.-K.L. conceived and designed the experiments; G.-N.Y. and K.-D.K. performed the experiments; G.-N.Y. and Y.-K.L. wrote the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Research Foundation of Korea (NRF-2019R1A2C2009999); the Ministry of Trade, Industry & Energy of Korea (MOTIE-10082582); and the University Innovation Support Program of Dankook University in 2021.

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflict of interest. The funding sponsors had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, and in the decision to publish the results.

References

  1. Chen, D.M.C.; Bodirsky, B.L.; Krueger, T.; Mishra, A.; Popp, A. The world’s growing municipal solid waste: Trends and impacts. Environ. Res. Lett. 2020, 15, 074021. [Google Scholar] [CrossRef]
  2. Palos, R.; Gutiérrez, A.; Vela, F.J.; Olazar, M.; Arandes, J.M.; Bilbao, J. Waste Refinery: The Valorization of Waste Plastics and End-of-Life Tires in Refinery Units. A Review. Energy Fuels 2021, 35, 3529–3557. [Google Scholar] [CrossRef]
  3. Gamboa, A.R.; Rocha, A.M.A.; dos Santos, L.R.; de Carvalho, J.A. Tire pyrolysis oil in Brazil: Potential production and quality of fuel. Renew. Sustain. Energy Rev. 2020, 120, 109614. [Google Scholar] [CrossRef]
  4. Laresgoiti, M.F.; Caballero, B.M.; De Marco, I.; Torres, A.; Cabrero, M.A.; Chomón, M.J. Characterization of the liquid products obtained in tyre pyrolysis. J. Anal. Appl. Pyrolysis 2004, 71, 917–934. [Google Scholar] [CrossRef]
  5. Williams, P.T.; Besler, S. Pyrolysis-thermogravimetric analysis of tyres and tyre components. Fuel 1995, 74, 1277–1283. [Google Scholar] [CrossRef]
  6. Williams, P.T. Pyrolysis of waste tyres: A review. Waste Manag. 2013, 33, 1714–1728. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Dos Santos, R.G.; Rocha, C.L.; Felipe, F.L.S.; Cezario, F.T.; Correia, P.J.; Rezaei-Gomari, S. Tire waste management: An overview from chemical compounding to the pyrolysis-derived fuels. J. Mater. Cycles Waste Manag. 2020, 22, 628–641. [Google Scholar] [CrossRef]
  8. Cheng, Z.; Li, M.; Li, J.; Lin, F.; Ma, W.; Yan, B.; Chen, G. Transformation of nitrogen, sulfur and chlorine during waste tire pyrolysis. J. Anal. Appl. Pyrolysis 2020, 153, 104987. [Google Scholar] [CrossRef]
  9. Mirmiran, S.; Pakdel, H.; Roy, C. Characterization of used tire vacuum pyrolysis oil: Nitrogenous compounds from the naphtha fraction. J. Anal. Appl. Pyrolysis 1992, 22, 205–215. [Google Scholar] [CrossRef]
  10. Hita, I.; Gutiérrez, A.; Olazar, M.; Bilbao, J.; Arandes, J.M.; Castaño, P. Upgrading model compounds and Scrap Tires Pyrolysis Oil (STPO) on hydrotreating NiMo catalysts with tailored supports. Fuel 2015, 145, 158–169. [Google Scholar] [CrossRef]
  11. Hwang, Y.-H.; Lee, Y.-K. Structure and activity of unsupported NiWS2 catalysts for slurry phase hydrocracking of vacuum residue: XAFS studies. J. Catal. 2021, in press. [Google Scholar] [CrossRef]
  12. Dujardin, C.; Lélias, M.A.; van Gestel, J.; Travert, A.; Duchet, J.C.; Maugé, F. Towards the characterization of active phase of (Co)Mo sulfide catalysts under reaction conditions-Parallel between IR spectroscopy, HDS and HDN tests. Appl. Catal. A 2007, 322, 46–57. [Google Scholar] [CrossRef]
  13. Ho, T.C. Deep HDS of diesel fuel: Chemistry and catalysis. Catal. Today 2004, 98, 3–18. [Google Scholar] [CrossRef]
  14. Yun, G.N.; Ahn, S.J.; Takagaki, A.; Kikuchi, R.; Oyama, S.T. Hydrodeoxygenation of Γ-valerolactone on bimetallic NiMo phosphide catalysts. J. Catal. 2017, 353, 141–151. [Google Scholar] [CrossRef]
  15. Yun, G.N.; Takagaki, A.; Kikuchi, R.; Ted Oyama, S. Hydrodeoxygenation of gamma-valerolactone on transition metal phosphide catalysts. Catal. Sci. Technol. 2017, 7, 281–292. [Google Scholar] [CrossRef]
  16. Yun, G.N.; Cho, K.S.; Kim, Y.S.; Lee, Y.K. A new approach to deep desulfurization of light cycle oil over Ni2P catalysts: Combined selective oxidation and hydrotreating. Catalysts 2018, 8, 102. [Google Scholar] [CrossRef] [Green Version]
  17. Yun, G.-N.; Lee, Y.-K. Dispersion effects of Ni2P catalysts on hydrotreating of light cycle oil. Appl. Catal. B 2014, 150–151, 647–655. [Google Scholar] [CrossRef]
  18. Djandja, O.S.; Wang, Z.; Duan, P.; Wang, F.; Xu, Y. Hydrotreatment of pyrolysis oil from waste tire in tetralin for production of high-quality hydrocarbon rich fuel. Fuel 2021, 285, 119185. [Google Scholar] [CrossRef]
  19. Han, Y.; Stankovikj, F.; Garcia-Perez, M. Co-hydrotreatment of tire pyrolysis oil and vegetable oil for the production of transportation fuels. Fuel Process. Technol. 2017, 159, 328–339. [Google Scholar] [CrossRef] [Green Version]
  20. Dębek, C.; Walendziewski, J. Hydrorefining of oil from pyrolysis of whole tyres for passenger cars and vans. Fuel 2015, 159, 659–665. [Google Scholar] [CrossRef]
  21. Rundqvist, S. X-Ray Investigations of Mn3P, Mn2P, and Ni2P. Acta Chem. Scand. 1962, 16, 992–998. [Google Scholar] [CrossRef] [Green Version]
  22. Pan, Y.; Yang, D.; Sun, K.; Wang, X.; Zhou, Y.; Huang, Q. Pyrolytic transformation behavior of hydrocarbons and heteroatom compounds of scrap tire volatiles. Fuel 2020, 276, 118095. [Google Scholar] [CrossRef]
  23. Hita, I.; Aguayo, A.T.; Olazar, M.; Azkoiti, M.J.; Bilbao, J.; Arandes, J.M.; Castaño, P. Kinetic Modeling of the Hydrotreating and Hydrocracking Stages for Upgrading Scrap Tires Pyrolysis Oil (STPO) toward High-Quality Fuels. Energy Fuels 2015, 29, 7542–7553. [Google Scholar] [CrossRef]
  24. Oyama, S.T.; Lee, Y.-K. The active site of nickel phosphide catalysts for the hydrodesulfurization of 4,6-DMDBT. J. Catal. 2008, 258, 393–400. [Google Scholar] [CrossRef]
  25. Lee, Y.K.; Shu, Y.; Oyama, S.T. Active phase of a nickel phosphide (Ni2P) catalyst supported on KUSY zeolite for the hydrodesulfurization of 4,6-DMDBT. Appl. Catal. A 2007, 322, 191–204. [Google Scholar] [CrossRef]
  26. Kim, J.; Jang, J.G.; Lee, Y.K. Reactivity of sulfur compounds in FCC decant oils for hydrodesulfurization over CoMoS2/Al2O3 catalysts. Korean J. Chem. Eng. 2021, 38, 1179–1187. [Google Scholar] [CrossRef]
  27. Jang, J.-G.; Lee, Y.-K. Promotional effect of Ga for Ni2P catalyst on hydrodesulfurization of 4,6-DMDBT. Appl. Catal. B 2019, 250, 181–188. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of Ni2P/SBA-15 and Ni2P/SiO2 catalysts.
Figure 1. XRD patterns of Ni2P/SBA-15 and Ni2P/SiO2 catalysts.
Catalysts 11 01272 g001
Figure 2. TEM images and particle size distributions of Ni2P/SBA-15 and Ni2P/SiO2.
Figure 2. TEM images and particle size distributions of Ni2P/SBA-15 and Ni2P/SiO2.
Catalysts 11 01272 g002
Figure 3. Characterization of tire pyrolysis oil by (a) GC/MS and (b) AED.
Figure 3. Characterization of tire pyrolysis oil by (a) GC/MS and (b) AED.
Catalysts 11 01272 g003
Figure 4. AED chromatographs of hydrotreating products on NiMoS, Ni2P/SBA-15, and Ni2P/SiO2 catalysts at 3.0 MPa, 350 °C, LHSV of 0.5 h−1, and time on stream of 12 h. (a) AED analysis-carbon, (b) AED analysis-sulfur, and (c) AED analysis-nitrogen.
Figure 4. AED chromatographs of hydrotreating products on NiMoS, Ni2P/SBA-15, and Ni2P/SiO2 catalysts at 3.0 MPa, 350 °C, LHSV of 0.5 h−1, and time on stream of 12 h. (a) AED analysis-carbon, (b) AED analysis-sulfur, and (c) AED analysis-nitrogen.
Catalysts 11 01272 g004
Figure 5. (a) HDS and (b) HDN conversion as a function of time on stream on NiMoS, Ni2P/SBA-15, and Ni2P/SiO2 catalysts at 3.0 MPa, 350 °C, and LHSV of 0.5 h−1.
Figure 5. (a) HDS and (b) HDN conversion as a function of time on stream on NiMoS, Ni2P/SBA-15, and Ni2P/SiO2 catalysts at 3.0 MPa, 350 °C, and LHSV of 0.5 h−1.
Catalysts 11 01272 g005
Figure 6. Aromatic compound distributions on NiMoS, Ni2P/SBA-15 and Ni2P/SiO2 catalysts at 3.0MPa, 350 °C, LHSV of 0.5 h−1, and time on stream of 12 h.
Figure 6. Aromatic compound distributions on NiMoS, Ni2P/SBA-15 and Ni2P/SiO2 catalysts at 3.0MPa, 350 °C, LHSV of 0.5 h−1, and time on stream of 12 h.
Catalysts 11 01272 g006
Table 1. Physicochemical properties of supported Ni2P catalysts.
Table 1. Physicochemical properties of supported Ni2P catalysts.
SampleBET Surface Area
/m2 g−1
CO Uptake
/mol g−1
Crystallite Size a
/nm
SBA-15687--
Ni2P/SBA-1534084.33.8
SiO2(EH-5)200--
Ni2P/SiO212471.29.1
a Crystallite size was calculated by Scherrer equation with 111 plane.
Table 2. Composition and properties of tire pyrolysis oil (TPO).
Table 2. Composition and properties of tire pyrolysis oil (TPO).
Physical PropertiesTPO
S/ppm14,500
N/ppm4070
Aromatics/wt.%Total41
Mono27
Di8
Tri+6
Distillation/°CIBP/20/31
44/58/72
80/EP
110/150/180
210/240/270
290/320
Table 3. Comparison of the reactivity of sulfur and nitrogen-containing compounds.
Table 3. Comparison of the reactivity of sulfur and nitrogen-containing compounds.
NiMoSNi2P/SBA-15Ni2P/SiO2
HDS
conversion/%
Total S94.589.373.0
Light S90.882.557.4
Benzothiazole100100100
Heavy S47.979.943.0
HDN
conversion/%
Total N72.960.741.0
Light N (Aniline)78.4 (72.5)63.8 (58.2)70.2 (56.9)
Benzothiazole99.5100100
Heavy N (Dimethylquinoline)93.0 (84.3)90.7 (84.3)96.9 (93)
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yun, G.-N.; Kim, K.-D.; Lee, Y.-K. Hydrotreating of Waste Tire Pyrolysis Oil over Highly Dispersed Ni2P Catalyst Supported on SBA-15. Catalysts 2021, 11, 1272. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11111272

AMA Style

Yun G-N, Kim K-D, Lee Y-K. Hydrotreating of Waste Tire Pyrolysis Oil over Highly Dispersed Ni2P Catalyst Supported on SBA-15. Catalysts. 2021; 11(11):1272. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11111272

Chicago/Turabian Style

Yun, Gwang-Nam, Ki-Duk Kim, and Yong-Kul Lee. 2021. "Hydrotreating of Waste Tire Pyrolysis Oil over Highly Dispersed Ni2P Catalyst Supported on SBA-15" Catalysts 11, no. 11: 1272. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11111272

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