Next Article in Journal
Greenness Assessment and Synthesis for the Bio-Based Production of the Solvent 2,2,5,5-Tetramethyloxolane (TMO)
Previous Article in Journal
Sustainable Chemistry—A New Open Access Journal
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Removing Simultaneously Sulfur and Nitrogen from Fuel under a Sustainable Oxidative Catalytic System

LAQV/REQUIMTE & Department of Chemistry and Biochemistry, Faculty of Sciences, University of Porto, 4169-007 Porto, Portugal
*
Author to whom correspondence should be addressed.
These authors contributed equally for the research work.
Sustain. Chem. 2021, 2(2), 382-391; https://0-doi-org.brum.beds.ac.uk/10.3390/suschem2020022
Submission received: 31 March 2021 / Revised: 4 June 2021 / Accepted: 7 June 2021 / Published: 19 June 2021

Abstract

:
An effective process to remove nitrogen-based compounds from fossil fuels without harming the process of sulfur removal is an actual gap in refineries. A success combination of desulfurization and denitrogenation processes capable of completely removing the most environmental contaminates in diesel under sustainable conditions was achieved in this work, applying polyoxometalates as catalysts, hydrogen peroxide as oxidant, and an immiscible ionic liquid as an extraction solvent. The developed process based in simultaneous oxidative desulfurization (ODS) and oxidative denitrogenation (ODN) involved initial extraction of sulfur and nitrogen compounds followed by catalytic oxidation. Keggin-type polyoxomolybdates revealed much higher reusing capacity than the related polyoxotungstate. Effectively, the first catalysts practically allowed complete sulfur and nitrogen removal only in 1 h of reaction and for ten consecutive cycles, maintaining the original catalyst and ionic liquid samples.

1. Introduction

Long-term energy outlooks agree that fossil fuels will remain the dominant energy for at least more two decades, since the demand for energy is increasing with the growth of the world’s population, economy, and increasing living standards. At present, fossil fuels cover ~88% of the energy requirement in the world [1]. The inevitable drawbacks of the current fossil fuel consumption are related to the environment degradation and public health. In the transportation sector, fossil fuels will be the main energy source for the near future. Motor vehicles powered on fossil fuels are among the major sources of air pollution in urban areas [2,3]. Fuels such as fuel oil, jet fuel, gasoline, diesel, etc., contain various pollutants, with the sulfur- and nitrogen-based organic compounds (SCs and NCs, respectively) as the majority of the contaminants [4]. The utilization of fossil fuels without purification poses a devastating effect on the environment and humans, mainly because of the emission of sulfur oxides (SOx) and nitrogen oxides (NOx) from SCs and NCs, respectively [5,6]. The global impact of SOx emissions resulting from the burning of fossil fuels has led to the imposition of environmental restrictions for the sulfur level in transportation fuels (10 ppm in EU) [7,8].
The standard industrial methods for the removal of sulfur and nitrogen-based compounds from crude oils are hydrodesulfurization (HDS) and hydrodenitrogenation (HDN) [9,10]. HDS is an effective process that needs high temperature (>350 °C) and pressure (up to 6 MPa), and also requires high amounts of hydrogen consumption [11,12,13]. SCs and NCs in fuels are converted to hydrogen sulfide and ammonia, respectively, via the catalytic reduction with H2 [14] In the HDS process, NCs severely affect the catalyst and cause corrosion of the refinery equipment due to their high reactivity [14,15,16]. Therefore, it is essential to remove NCs before the desulfurization treatment of fuel, or preferably, to remove NCs and SCs simultaneously by an alternative process, capable of depth desulfurization and denitrogenation, reducing the impact of fuels on the environment by decreasing NOx and SOx emissions. Oxidative desulfurization (ODS) and oxidative denitrogenation (ODN) processes have been the focus of promising research in recent years due to their efficiency for removal of the most refractory S and N-content present in fuels [10,17,18]. This technology conciliates oxidative catalysis and liquid-liquid extraction steps [15,16]. Consequently, the efficiency of combined ODS/ODN processes depends of the catalyst performance and the ability of the extraction solvent. The sustainability and the cost-effectiveness of the process can be assured by the use of recovered and recyclable catalysts and the application of environmental-friendly solvents, such as ionic liquids [19,20]. The former are able to behave as an extraction solvent and also as an immobilization medium for the homogeneous catalyst, containing catalyst and subtract in the same liquid phase [21,22]. Hydrogen peroxide is generally the preferred oxidant to use in these processes due to its high active oxygen content and the fact that water is the sole by-product [23,24]. Our research group has been dedicated to the design of efficient catalysts for ODS based on polyoxometalates (POMs) [18,25,26,27,28,29] POMs are polyanions incorporating various transition metals and are well-known by their structural diversity and potential in several areas of application [30]. In general, POMs are powerful catalysts for a great variety of oxidative reactions. In particular, Keggin-type POMs ([Xn+M12O40](8-n)- (X: block p or d heteroatom) have been successfully used as efficient catalysts in ODS processes [31,32,33,34,35]. However, the investigation combining ODS and ODN is in a primitive stage and only one publication could be found in the literature demonstrating the ability of POMs as catalysts in this mixed systems [36].
This work reports an innovative combined ODS/ODN system catalyzed by hybrid-POMs and by applying an ionic liquid that acts as extraction solvent (1-butyl-3-methylimidazole hexafluorophosphate ([BMIM]PF6)). The reusability and the stability of the catalyst were further investigated.

2. Materials and Methods

2.1. Reagents and Solvents

Phosphomolybdic acid (H3PMo12O40.xH2O, for microscopy), sodium phosphomolybdate (Na3PMo12O40.xH2O, technical), tetradecane (C14H30, >99%), 1-butyl-3-methylimidazole hexafluorophosphate (C8H15F6N2P, >97%, [BMIM]PF6), hydrogen peroxide (H2O2, aq. 30%), dibenzothiophene (C12H8S, 98%, DBT), 4-methyldibenzothiophene (C13H10S, 96%, MDBT), quinoline (98 %, QUI), and indole (>99%, IND) were obtained from Sigma-Aldrich. Benzothiophene (C8H6S, >95%, BT), phosphotungstic acid (H3PW12O40.xH2O, 98%), and 1-butylpyridinium bromide (>99%, BPyBr) were acquired from Fluka. N-octane (C8H18, >99%) and 4,6-dimethyldibenzothiophene (C14H12S, 95%, DMDBT) were acquired from Acros Organic. None of these were subject to further treatment or purification.

2.2. Characterization Methods and Instrumentation

Fourier-transformed Infrared (FT-IR) spectra were acquired between 400 and 4000 cm−1 on a Bruker Tensor 27 Spectrometer using KBr pellets, and all the representations are shown in arbitrary unities of transmittance. Solution 31P nuclear magnetic resonance (NMR) spectra were recorded in CD3CN or D2O at 162 MHz with a Bruker Avance III 400 spectrometer, and the chemical shifts are given with respect to external 85% H3PO4. Catalytic reactions were periodically monitored by GC-FID analysis carried out in a Bruker 430-GC-FID chromatograph (Germany). Hydrogen was used as carrier gas (55 cm.s−1) and fused silica Supelco capillary columns SPB-5 (30 m × 0.25 mm i.d.; 25 μm film thickness) were used.

2.3. [BPy]3 Salts Preparation

[BPy]3PMo12O40 ([BPy]PMo12) and [BPy]3PW12O40 ([BPy]PW12) were prepared following an experimental procedure adapted from previously reported methods [34,35]. Briefly, [BPy]Br (2.5 mmol) was dissolved in 2.5 mL of deionized water, and an aqueous solution of phosphomolybdic or phosphotungstic acid (0.5 mmol) in deionized water (2.5 mL) was added dropwise under magnetic stirring at room temperature. The mixture was left stirring for 30 min, after which the precipitate was filtrated, washed thoroughly with deionized water, and dried in a desiccator overnight. FT-IR (cm−1, KBr): ν([BPy]PMo12) = 3435 (w), 3126 (w), 3085 (w), 3065 (w), 2964 (w), 2930 (w), 2874 (w), 1632 (m), 1486 (m), 1465 (vw), 1384 (vw), 1314 (vw), 1169 (w), 1063 (s), 956 (s), 878 (m), 796 (s), 683 (m), 596 (vw), 502 (w); ν([BPy]PW12) = 3456 (w), 3127 (w), 3085 (w), 3065 (w), 2964 (w), 2930 (w), 2874 (w), 1632 (m), 1486 (m), 1462 (vw), 1385 (vw), 1319 (vw), 1168 (w), 1080 (s), 978 (s), 896 (m), 804 (s), 683 (m), 596 (vw), 524 (m). 31P NMR (162 MHz, CD3CN, 25 °C): δ([BPy]PMo12) = −2.39 ppm; δ([BPy]PW12) = −13.89 ppm.

2.4. Simultaneous Desulfurization and Denitrogenation Processes (ODS/ODN)

ODS/ODN studies were performed with a multicomponent Model Diesel containing SCs and NCs prepared with BT, DBT, MDBT, DMDBT, Q, and I in n-octane (500 ppm of sulfur or 300 ppm of nitrogen from each compound). The reactions were carried out under air in a closed borosilicate vessel with a magnetic stirrer and were immersed in a thermostatically controlled liquid paraffin bath at 70 °C. ODS/ODN reactions were performed in a biphasic system composed of the Model Diesel and [BMIM]PF6 as extraction solvent. In a representative experiment, a certain amount of the catalyst equivalent to 3 μmol of POM was added to 0.75 mL of [BMIM]PF6 and 0.75 mL of Model Diesel, and this mixture was stirred for 10 min at 70 °C. The oxidative catalytic step was then initiated with the addition of aqueous H2O2 30 % (75 μL) to the reaction mixture. Tetradecane was used as a standard in the periodical monitorization of the sulfur content by gas chromatography (GC) analysis. At the end of each reaction, the processed Model Diesel phase was removed and substituted by a fresh volume. Both catalyst and [BMIM]PF6 solvent were reused in the subsequent catalytic cycle, under identical reaction conditions and with the addition of a new portion of H2O2. The experimental error associated with the combined ODS/ODN efficiency was calculated by performing at least three repeated reactions (see Figure S1 in Supporting Information), and it was found to be approximately 5% to ODS and 4% to ODN.

3. Results and Discussion

3.1. Catalysts Synthesis and Characterization

The catalysts [BPy]PMo12 and [BPy]PW12 are POM salts previously reported by our group [34,35]. These were prepared by replacing the protons of phosphomolybdic or phosphotungstic acid through a simple acid-base reaction with [BPy]Br in water. [Na]3PMo12O40 ([Na]PMo12) is a commercially available salt. Characterization results by FT-IR and 31P NMR spectroscopies obtained for these compounds are displayed in Figure 1. All infrared spectra generally display the expected bond vibration absorption band pattern. For phosphomolybdic salts, strong intensity bond vibration bands around 1064, 960, 880, and 794 cm−1 are registered for [BPy]PMo12 corresponding to νas(P—Oa), νas(Mo—Od), νas(Mo—Ob—Mo), and νas(Mo—Oc—Mo), respectively [37,38,39]. For [Na]PMo12, νas(Mo—Ob—Mo) seems to be expressed in a pair of bands observed at 904 and 862 cm−1, hinting at variable oxidation states for Mo. Phosphotungstate in [BPy]PW12 is similarly expressed by strong intensity νas(P—Oa), νas(W—Od), νas(W—Ob—W), and νas(W—Oc—W) bond vibration absorption bands at 1080, 976, 892, and 802 cm−1, correspondingly. The infrared spectra of both [BPy]3 salts contain weak bands at 3126, 3085, and 3065 cm−1 corresponding to ν(C—H) aromatic bond vibrations, as well as at 2964, 2930, and 2874 cm−1 attributed to the ν(C—H) of the aliphatic chain, a medium intensity band at 1632 cm−1 corresponding to ν(C=N) vibrations that is typical of the quaternary nitrogen atom in a heterocyclic ring. Several weak/medium bands in the range 1580–1100 cm−1 are observed due to aromatic ν(C—N), ν(C—C), and δ(C—H) vibrations. A strong band at 1485 cm−1 is attributable to the conjugation of ν(C=C) and ν(C=N) bonds, typical of pyridinium salts. The strong one at 682 cm−1 corresponds to the aromatic out-of-plane hydrogen deformation [40]. 31P NMR analysis reveals the typical chemical shifts for each structure [34,35]. The spectrum recorded for [Na]PMo12 registers two single peaks attributed to a different content of Na cations in the POM structure.

3.2. ODS/ODN Catalytic Studies

All ODS/ODN studies were performed at 70 °C in a biphasic liquid-liquid system based on [BMIM]PF6 as extraction solvent and an immiscible multicomponent S/N Model Diesel containing the most representative refractory sulfur and nitrogen content present in fuels, namely benzothiophene (BT), dibenzothiophene (DBT), 4-methyldibenzothiophene (MDBT), and 4,6-dimethyldibenzothiophene (DMDBT) from SCs, and indole (IND) and quinoline (QUI) from NCs. Catalytic systems based on [Na]3PMo12O40 [Na]PMo12), [BPy]3PMo12O40 ([BPy]PMo12), and [BPy]3PW12O40 ([BPy]PW12) were composed by equal volumes of Model Diesel and [BMIM]PF6 (750 μL) and an equivalent of 3 μmol of active catalyst. The simultaneous desulfurization and denitrogenation processes occurred in two main steps; at first, an initial extraction of SCs and NCs was transferred from the Model Diesel to the BMIM]PF6 extraction phase (during 10 min at 70 °C under stirring). A time longer than 10 min did not increase the initial extraction of SCs and NCs. In the next step, the oxidative catalytic stage was initiated by the addition of aq. H2O2 (30%). The performance of catalysts [Na]PMo12, [BPy]PMo12, and [BPy]PW12 in simultaneous ODS/ODN reactions was evaluated and the experimental results are displayed in Figure 2. All these catalysts behaved as homogenous catalysts presented only in the [BMIM]PF6 phase, since the 31P NMR analysis did not present any phosphorus signal. The combined desulfurization/denitrogenation profiles demonstrate that [Na]PMo12 and [BPy]PMo12 have similar catalytic performance, achieving near total conversion after 1 h of reaction (98.7 and 97.9%, respectively). On the other hand, using the [BPy]PW12 catalyst, near complete ODS/ODN was achieved faster than the analogous [BPy]PMo12 catalyst. These results indicate that the nature of the cation in the POM catalyst does not seem to have an important influence in its catalytic performance. On the other hand, a distinct combined ODS/ODN profile was found using polyoxotungstate and polyoxomolybdate, where the first showed to be the most effective, guaranteeing total desulfurization and denitrogenation after just 1 h of reaction. The removal rates of each component of the S/N Model Diesel are the following: IND > QUI > DBT > MDBT > DMDBT ~ BT. The removal of NCs is largely achieved during the initial extraction step. The difference in removal rates between IND and QUI relates to the higher basicity of the proton-donor present in IND [41,42]. During the oxidation catalytic step (after the 10 min), the difference of reactivity between the SCs can be explained by the decrease in electron density on the sulfur atom, which hinders oxidation [18]. The nature of the obtained oxidized products was identified by the analysis of the extraction phase, since, in the model diesel phase, no oxidized products were found. The oxidized products of sulfur compounds were the corresponding sulfones. Only vestigial amounts of sulfoxide from BT were found. From the QUI oxidation, the quinoline N-oxide was identified.
Figure 3 displays the denitrogenation and the desulfurization profiles of the combined ODS/ODS processes. These results demonstrated that complete denitrogenation is achieved faster than desulfurization, since higher extraction of NCs occurred during the first 10 min of the process, which is associated to the higher oxidative facility of NCs than SCs. Figure 3 also compares the desulfurization profiles that occurred in a combined ODS/ODN process and a single ODS process, i.e., using a Model Diesel containing only SCs. These experiments were performed using the same reactional conditions. Without duties, the desulfurization profiles are similar in combined S/N and single S processes. These results indicate that the conciliation of extraction and oxidative catalytic steps did not result in a competitive S and N removal process, and these may occur in parallel, sharing the oxidant and the catalyst action. This is a remarkable advantage compared to the combined hydrodesulfuriation/hydrodenitrogenation processes [14].
The recycling ability of each POM/[BMIM]PF6 catalytic system was accessed by consecutive reutilization cycles that consisted in the preservation of the POM/[BMIM]PF6 phase for various consecutive cycles. This reusing process is the desired sustainable procedure of using the same portion of catalyst and ionic liquid extraction solvent for consecutive combined ODS/ODN cycles. After each cycle, the treated S/N Model Diesel is removed from the system and replaced by a novel portion of S/N Model Diesel, as well the H2O2 oxidant. During the reutilization cycles, an interlayer aqueous phase increased in volume between the S/N Model Diesel and [BMIM]PF6 phase due to the successive addition of aqueous oxidizing agent (Figure 4a). Figure 4b display the results obtained from reusing the three different POM/[BMIM]PF6 catalytic systems. These results were achieved after 1 h. The [BPy]PW12/[BMIM]PF6 system presented a decrease of catalytic efficiency, mainly after the 3rd cycle, since its combined ODS/ODN efficiency decreased from 98% from the 3rd cycle to 63% for the 4th cycle. In fact, the efficiency found for the 4th cycle was only attributed to the initial extraction step, indicating the absence of the oxidative catalytic performance of [BPy]PW12 catalyst. In the 5th combined desulfurization/denitrogenation cycle, the catalyst did not promote any oxidative catalysis, and the initial extraction of the S/N compounds decreased to 46%. This decrease of S/N extraction is a consequence of absence of S/N oxidation, i.e., without decreasing non-oxidized SCs and NCs in the [BMIM]PF6 phase, the extraction of more S/N from Diesel did not occur. Reusing experiments using the [BPy]PW12/[BMIM]PF6 system were not performed after the 5th cycle. On the other hand, using the polyoxomolybdate catalytic systems, i.e., [Na]PMo12 and [BPy]PMo12, the loss of the catalytic activity was not verified for ten consecutive combined ODS/ODN cycles and near complete removal of S/N from Model Diesel was achieved after 1 h. A small decrease in activity is only observed during the 10th cycle (93 and 98%, for [Na]PMo12 and [BPy]PMo12, respectively).
After the reusing cycles, the stability of the homogeneous POM catalysts was investigated by 31P NMR. The immiscible aqueous and [BMIM]PF6 phases (Figure 4b) were separated, analyzed, and the results are displayed in Figure 5. In general, all studied catalysts suffered structural transformations during the various reusing cycles. Furthermore, the structural transformations may alter solubility properties, since POM derivative peaks could be found in both the Ionic Liquid and aqueous phase. The structural modification of Na[PMo12] and [BPy]PMo12 was not coincident, and different active POM derivative fragments were found in aqueous and [BMIM]PF6 phases. On the other hand, the analysis of both phases obtained after the 5th cycle using the [BPy]PW12 catalyst revealed similar distribution of POM transformed species. It ends up being implied that the structural transformations associated with [BPy]3PW12O40 catalytic activity ultimately led to its deactivation. In a future work, the stability and the facility of the recovery of these homogeneous catalysts will be treated by their immobilization in appropriate supporting materials.

4. Conclusions

This work reported a novel efficient strategy capable of combining complete desulfurization and denitrogenation to treat a multicomponent S/N Model Diesel containing the most refractory elements present in Diesel. Keggin-type POMs showed to be active catalysts to perform parallel oxidative desulfurization (ODS) and oxidative denitrogenation (ODS) under sustainable conditions, i.e., using hydrogen peroxide as oxidant and an ionic liquid as extraction solvent ([BMIM]PF6). Complete removal of sulfur and nitrogen compounds was achieved only after 1 h. Furthermore, a reusing test was performed by using the homogeneous POM catalyst and [BMIM]PF6 for consecutive ODS/ODN cycles. The polyoxomolybdates [Na]PMo12 and [BPy]PMo12 showed high reusing capacity for ten consecutive cycles without any appreciable reduction of catalytic performance. The kinetic similarity obtained between these two polyoxomolybdates-based compounds indicated that the nature of the cation does not seem to influence its activity. The reusing capacity was drastically decreased when the polyoxotungstate [BPy]PW12 was used and, in this case, only three consecutive ODS/ODN cycles were successfully performed. Lastly, the stability analysis of the POM catalysts after reusing cycles indicates that the polyoxomolybdate is in fact a precursor for the active catalytic species, since its Keggin structures are modified in other active compounds. On the other hand, the polyoxotungstate also suffered structural modification but into inactive species. Therefore, future work is already planned to prepare heterogeneous POM catalysts to avoid their fast structural modification and also to facilitate catalyst recovery.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/suschem2020022/s1, Figure S1, Denitrogenation and desulfurization profiles of a combined ODS/ODN process, repeated three times for accuracy investigation, using a single Model Diesel containing BT, DMDBT, MDBT and DBT, QUI and IND, using [Na]PMo12 catalyst, [BMIM]PF6 extraction solvent, H2O2 as oxidant, at 70 °C.

Author Contributions

Conceptualization, S.S.B.; methodology, S.S.B.; validation, S.S.B. and L.C.-S.; formal analysis, S.S.B.; investigation, D.F.S., A.M.V. and F.M.; writing—original draft preparation, A.M.V. and S.S.B.; writing—review and editing, S.S.B. and L.C.-S.; supervision, B.d.C., S.S.B. and L.C.-S.; project administration, S.S.B. and L.C.-S.; funding acquisition, B.d.C. and S.S.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research work received financial support from Portuguese national funds (FCT/MCTES, Fundação para a Ciência e a Tecnologia and Ministério da Ciência, Tecnologia e Ensino Superior) through the strategic project UIDB/50006/2020 (for LAQV-REQUIMTE). Parts of this work were performed with funds from the European Union (FEDER funds through COMPETE POCI-01-0145-FEDER-031983) and FCT/MCTES by National Funds to the R&D project GlyGold (PTDC/CTM-CTM/31983/2017).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

More data can be obtained by request to the authors.

Acknowledgments

L.C.-S. and S.S.B. thank FCT/MCTES for funding through the Individual Call to Scientific Employment Stimulus (Ref. CEECIND/00793/2018 and Ref. CEECIND/03877/2018, respectively). AMV thanks FCT/MCTES and ESF (European Social Fund) through POCH (Programa Operacional Capital Humano) for his PhD grant (Ref. SFRH/BD/150659/2020).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chow, J.; Kopp, R.J.; Portney, P.R. Energy Resources and Global Development. Science 2003, 302, 1528–1531. [Google Scholar] [CrossRef] [Green Version]
  2. Frey, H.C. Trends in onroad transportation energy and emissions. J. Air Waste Manag. Assoc. 2018, 68, 514–563. [Google Scholar] [CrossRef] [PubMed]
  3. Chossière, G.P.; Malina, R.; Allroggen, F.; Eastham, S.D.; Speth, R.L.; Barrett, S.R.H. Country- and manufacturer-level attribution of air quality impacts due to excess NOx emissions from diesel passenger vehicles in Europe. Atmos. Environ. 2018, 189, 89–97. [Google Scholar] [CrossRef]
  4. Agency, E.E. Effects of air pollution on European ecosystems. Past and future exposure of European freshwater and terrestrial habitats to acidifying and eutrophying air pollutants. EEA Tech. Rep. 2014, 11, 1–42. [Google Scholar] [CrossRef]
  5. Reşitoğlu, İ.A.; Altinişik, K.; Keskin, A. The pollutant emissions from diesel-engine vehicles and exhaust aftertreatment systems. Clean Technol. Environ. Policy 2014, 17, 15–27. [Google Scholar] [CrossRef] [Green Version]
  6. Samokhvalov, A. Desulfurization of Real and Model Liquid Fuels Using Light: Photocatalysis and Photochemistry. Catal. Rev. 2012, 54, 281–343. [Google Scholar] [CrossRef]
  7. Alshammari, Y.M.; Benmerabet, M. Global scenarios for fuel oil utilisation under new sulphur and carbon regulations. OPEC Energy Rev. 2017, 41, 261–285. [Google Scholar] [CrossRef]
  8. Yang, L.; Franco, V.; Mock, P.; Kolke, R.; Zhang, S.; Wu, Y.; German, J. Experimental Assessment of NOx Emissions from 73 Euro 6 Diesel Passenger Cars. Environ. Sci. Technol. 2015, 49, 14409–14415. [Google Scholar] [CrossRef]
  9. Bhadra, B.N.; Song, J.Y.; Uddin, N.; Khan, N.A.; Kim, S.; Choi, C.H.; Jhung, S.H. Oxidative denitrogenation with TiO2@porous carbon catalyst for purification of fuel: Chemical aspects. Appl. Catal. B Environ. 2019, 240, 215–224. [Google Scholar] [CrossRef]
  10. Mirante, F.; Mendes, R.F.; Paz, F.A.A.; Balula, S.S. High Catalytic Efficiency of a Layered Coordination Polymer to Remove Simultaneous Sulfur and Nitrogen Compounds from Fuels. Catalysts 2020, 10, 731. [Google Scholar] [CrossRef]
  11. Prado, G.H.C.; Rao, Y.; de Klerk, A. Nitrogen Removal from Oil: A Review. Energy Fuels 2016, 31, 14–36. [Google Scholar] [CrossRef]
  12. Wang, B.; Zheng, P.; Fan, H.; Meng, Q.; Duan, A.; Chen, Z.; Xu, C. Insights into the effect of solvent on dibenzothiophene hydrodesulfurization. Fuel 2021, 287, 119459. [Google Scholar] [CrossRef]
  13. Paucar, N.E.; Kiggins, P.; Blad, B.; De Jesus, K.; Afrin, F.; Pashikanti, S.; Sharma, K. Ionic liquids for the removal of sulfur and nitrogen compounds in fuels: A review. Environ. Chem. Lett. 2021, 19, 1205–1228. [Google Scholar] [CrossRef]
  14. Angelici, R.J. Hydrodesulfurization & Hydrodenitrogenation. In Encyclopedia of Inorganic Chemistry; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2005. [Google Scholar] [CrossRef]
  15. Li, H.; Jiang, X.; Zhu, W.; Lu, J.; Shu, H.; Yan, Y. Deep Oxidative Desulfurization of Fuel Oils Catalyzed by Decatungstates in the Ionic Liquid of [BMIM]PF6. Ind. Eng. Chem. Res. 2009, 48, 9034–9039. [Google Scholar] [CrossRef]
  16. Shiraishi, Y.; Tachibana, K.; Hirai, T.; Komasawa, I. Desulfurization and Denitrogenation Process for Light Oils Based on Chemical Oxidation followed by Liquid-Liquid Extraction. Ind. Eng. Chem. Res. 2002, 41, 4362–4375. [Google Scholar] [CrossRef]
  17. Yao, Z.; Miras, H.N.; Song, Y.-F. Efficient concurrent removal of sulfur and nitrogen contents from complex oil mixtures by using polyoxometalate-based composite materials. Inorg. Chem. Front. 2016, 3, 1007–1013. [Google Scholar] [CrossRef] [Green Version]
  18. Julião, D.; Gomes, A.C.; Pillinger, M.; Gonçalves, I.S.; Balula, S.S. Desulfurization and Denitrogenation Processes to Treat Diesel Using Mo(VI)-Bipyridine Catalysts. Chem. Eng. Technol. 2020, 43, 1774–1783. [Google Scholar] [CrossRef]
  19. Coletti, A.; Sabuzi, F.; Floris, B.; Galloni, P.; Conte, V. Efficient and sustainable V-catalyzed oxidative desulfurization of fuels assisted by ionic liquids. J. Fuel Chem. Technol. 2018, 46, 1121–1129. [Google Scholar] [CrossRef]
  20. Li, S.-W.; Wang, W.; Zhao, J.-S. The quantity and type of ILs needed to form magnetic-heteropolyacid mesoporous catalysts and their highly performance for DBT removal. Sustain. Energy Fuels 2020, 4, 2422–2437. [Google Scholar] [CrossRef]
  21. Yu, G.; Wu, X.; Wei, L.; Zhou, Z.; Liu, W.; Zhang, F.; Qu, Y.; Ren, Z. Desulfurization of diesel fuel by one-pot method with morpholinium-based Brønsted acidic ionic liquid. Fuel 2021, 296, 120551. [Google Scholar] [CrossRef]
  22. Gao, Y.; Julião, D.; Silva, D.F.; de Castro, B.; Zhao, J.; Balula, S.S. A simple desulfurization process to achieve high efficiency, sustainability and cost-effectivity via peroxotungstate catalyst. Mol. Catal. 2021, 505, 111515. [Google Scholar] [CrossRef]
  23. Juliao, D.; Gomes, A.C.; Cunha-Silva, L.; Valenca, R.; Ribeiro, J.C.; Pillinger, M.; de Castro, B.; Goncalves, I.S.; Balula, S.S. A sustainable peroxophosphomolybdate/H2O2 system for the oxidative removal of organosulfur compounds from simulated and real high-sulfur diesels. Appl. Catal. A-Gen. 2020, 589, 9. [Google Scholar] [CrossRef]
  24. Juliao, D.; Mirante, F.; Ribeiro, S.O.; Gomes, A.C.; Valenca, R.; Ribeiro, J.C.; Pillinger, M.; de Castro, B.; Goncalves, I.S.; Balula, S.S. Deep oxidative desulfurization of diesel fuels using homogeneous and SBA-15-supported peroxophosphotungstate catalysts. Fuel 2019, 241, 616–624. [Google Scholar] [CrossRef]
  25. Nogueira, L.S.; Ribeiro, S.; Granadeiro, C.M.; Pereira, E.; Feio, G.; Cunha-Silva, L.; Balula, S.S. Novel polyoxometalate silica nano-sized spheres: Efficient catalysts for olefin oxidation and the deep desulfurization process. Dalton Trans. 2014, 43, 9518–9528. [Google Scholar] [CrossRef] [PubMed]
  26. Julião, D.; Gomes, A.C.; Pillinger, M.; Cunha-Silva, L.; de Castro, B.; Gonçalves, I.S.; Balula, S.S. Desulfurization of model diesel by extraction/oxidation using a zinc-substituted polyoxometalate as catalyst under homogeneous and heterogeneous (MIL-101(Cr) encapsulated) conditions. Fuel Process. Technol. 2015, 131, 78–86. [Google Scholar] [CrossRef]
  27. Mirante, F.; Gomes, N.; Branco, L.C.; Cunha-Silva, L.; Almeida, P.L.; Pillinger, M.; Gago, S.; Granadeiro, C.M.; Balula, S.S. Mesoporous nanosilica-supported polyoxomolybdate as catalysts for sustainable desulfurization. Microporous Mesoporous Mater. 2019, 275, 163–171. [Google Scholar] [CrossRef]
  28. Gao, Y.; Mirante, F.; de Castro, B.; Zhao, J.; Cunha-Silva, L.; Balula, S.S. An Effective Hybrid Heterogeneous Catalyst to Desulfurize Diesel: Peroxotungstate@Metal-Organic Framework. Molecules 2020, 25, 5494. [Google Scholar] [CrossRef] [PubMed]
  29. Ribeiro, S.; Barbosa, A.D.S.; Gomes, A.C.; Pillinger, M.; Gonçalves, I.S.; Cunha-Silva, L.; Balula, S.S. Catalytic oxidative desulfurization systems based on Keggin phosphotungstate and metal-organic framework MIL-101. Fuel Process. Technol. 2013, 116, 350–357. [Google Scholar] [CrossRef]
  30. Long, D.-L.; Tsunashima, R.; Cronin, L. Polyoxometalates: Building Blocks for Functional Nanoscale Systems. Angew. Chem. Int. Ed. 2010, 49, 1736–1758. [Google Scholar] [CrossRef]
  31. Wang, S.-S.; Yang, G.-Y. Recent Advances in Polyoxometalate-Catalyzed Reactions. Chem. Rev. 2015, 115, 4893–4962. [Google Scholar] [CrossRef]
  32. Zhu, W.; Huang, W.; Li, H.; Zhang, M.; Jiang, W.; Chen, G.; Han, C. Polyoxometalate-based ionic liquids as catalysts for deep desulfurization of fuels. Fuel Process. Technol. 2011, 92, 1842–1848. [Google Scholar] [CrossRef]
  33. Wang, R.; Zhang, G.; Zhao, H. Polyoxometalate as effective catalyst for the deep desulfurization of diesel oil. Catal. Today 2010, 149, 117–121. [Google Scholar] [CrossRef]
  34. Mirante, F.; Gomes, N.; Corvo, M.C.; Gago, S.; Balula, S.S. Polyoxomolybdate based ionic-liquids as active catalysts for oxidative desulfurization of simulated diesel. Polyhedron 2019, 170, 762–770. [Google Scholar] [CrossRef]
  35. Ribeiro, S.O.; Duarte, B.; De Castro, B.; Granadeiro, C.M.; Balula, S.S. Improving the Catalytic Performance of Keggin [PW12O40]3− for Oxidative Desulfurization: Ionic Liquids versus SBA-15 Composite. Materials 2018, 11, 1196. [Google Scholar] [CrossRef] [Green Version]
  36. Bertleff, B.; Haider, M.S.; Claußnitzer, J.; Korth, W.; Wasserscheid, P.; Jess, A.; Albert, J. Extractive Catalytic Oxidative Denitrogenation of Fuels and Their Promoting Effect for Desulfurization Catalyzed by Vanadium Substituted Heteropolyacids and Molecular Oxygen. Energy Fuels 2020, 34, 8099–8109. [Google Scholar] [CrossRef]
  37. Rocchiccioli-Deltcheff, C.; Fournier, M.; Franck, R.; Thouvenot, R. Vibrational investigations of polyoxometalates. 2. Evidence for anion-anion interactions in molybdenum(VI) and tungsten(VI) compounds related to the Keggin structure. Inorg. Chem. 1983, 22, 207–216. [Google Scholar] [CrossRef]
  38. Rocchiccioli-Deltcheff, C.; Aouissi, A.; Bettahar, M.M.; Launay, S.; Fournier, M. Catalysis by 12-Molybdophosphates: 1. Catalytic Reactivity of 12-Molybdophosphoric Acid Related to Its Thermal Behavior Investigated through IR, Raman, Polarographic, and X-ray Diffraction Studies: A Comparison with 12-Molybdosilicic Acid. J. Catal. 1996, 164, 16–27. [Google Scholar] [CrossRef]
  39. Silviani, E.; Burns, R.C. Synthesis and characterization of soluble alkali metal, alkaline earth metal and related Keggin-type [PMo12O40]3−salts for heterogeneous catalysis reactions. J. Mol. Catal. A Chem. 2004, 219, 327–342. [Google Scholar] [CrossRef]
  40. Cook, D. Vibrational Spectra of Pyridinium Salts. Can. J. Chem. 1961, 39, 2009–2024. [Google Scholar] [CrossRef]
  41. Jiao, T.; Zhuang, X.; He, H.; Zhao, L.; Li, C.; Chen, H.; Zhang, S. An ionic liquid extraction process for the separation of indole from wash oil. Green Chem. 2015, 17, 3783–3790. [Google Scholar] [CrossRef]
  42. Zhou, Z.Q.; Li, W.S.; Liu, J. Removal of nitrogen compounds from fuel oils using imidazolium-based ionic liquids. Pet. Sci. Technol. 2017, 35, 45–50. [Google Scholar] [CrossRef]
Figure 1. (a) FT-IR spectra showed in the wavenumber regions 4000–2500 cm−1 and 1800–400 cm−1, and (b) 31P NMR spectra showed in the chemical shifts ranging from 1 to −17 ppm, obtained for [Na]PMo12, [BPy]PMo12, and [BPy]PW12.
Figure 1. (a) FT-IR spectra showed in the wavenumber regions 4000–2500 cm−1 and 1800–400 cm−1, and (b) 31P NMR spectra showed in the chemical shifts ranging from 1 to −17 ppm, obtained for [Na]PMo12, [BPy]PMo12, and [BPy]PW12.
Suschem 02 00022 g001
Figure 2. (a) Combined denitrogenation and desulfuration profiles catalyzed by the various POMs, using a biphasic S/N Model Diesel/[BMIM]PF6 system, H2O2 as oxidant, at 70 °C; (b) denitrogenation and desulfurization profile of each SCs and NCs present in the Model Diesel, using [Na]PMo12 catalyst.
Figure 2. (a) Combined denitrogenation and desulfuration profiles catalyzed by the various POMs, using a biphasic S/N Model Diesel/[BMIM]PF6 system, H2O2 as oxidant, at 70 °C; (b) denitrogenation and desulfurization profile of each SCs and NCs present in the Model Diesel, using [Na]PMo12 catalyst.
Suschem 02 00022 g002
Figure 3. Denitrogenation and desulfurization profiles of a combined ODS/ODN process and also a desulfurization of a single ODS process (using a single Model Diesel containing BT, DMDBT, MDBT, and DBT), using [Na]PMo12 catalyst, [BMIM]PF6 extraction solvent, H2O2 as oxidant, at 70 °C.
Figure 3. Denitrogenation and desulfurization profiles of a combined ODS/ODN process and also a desulfurization of a single ODS process (using a single Model Diesel containing BT, DMDBT, MDBT, and DBT), using [Na]PMo12 catalyst, [BMIM]PF6 extraction solvent, H2O2 as oxidant, at 70 °C.
Suschem 02 00022 g003
Figure 4. (a) Representation of the [Na]PMo12/[BMIM]PF6 system after ten consecutive ODS/ODS cycles. (b) Combined desulfurization/denitrogenation efficiency obtained after 1 h of reaction, when catalyst/[BMIM]PF6 system is reused for ten consecutive cycles, using catalysts [Na]PMo12, [BPy]PMo12, and [BPy]PW12, with H2O2 as oxidant, at 70 °C.
Figure 4. (a) Representation of the [Na]PMo12/[BMIM]PF6 system after ten consecutive ODS/ODS cycles. (b) Combined desulfurization/denitrogenation efficiency obtained after 1 h of reaction, when catalyst/[BMIM]PF6 system is reused for ten consecutive cycles, using catalysts [Na]PMo12, [BPy]PMo12, and [BPy]PW12, with H2O2 as oxidant, at 70 °C.
Suschem 02 00022 g004
Figure 5. 31P NMR spectra acquired in CD3CN or D2O for each [BMIM]PF6 and aqueous phases present in the catalytic systems [Na]PMo12 (10th cycle; a), [BPy]PMo12 (10th cycle; b), and [BPy]PW12 (5th cycle; c).
Figure 5. 31P NMR spectra acquired in CD3CN or D2O for each [BMIM]PF6 and aqueous phases present in the catalytic systems [Na]PMo12 (10th cycle; a), [BPy]PMo12 (10th cycle; b), and [BPy]PW12 (5th cycle; c).
Suschem 02 00022 g005aSuschem 02 00022 g005b
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Silva, D.F.; Viana, A.M.; Mirante, F.; de Castro, B.; Cunha-Silva, L.; Balula, S.S. Removing Simultaneously Sulfur and Nitrogen from Fuel under a Sustainable Oxidative Catalytic System. Sustain. Chem. 2021, 2, 382-391. https://0-doi-org.brum.beds.ac.uk/10.3390/suschem2020022

AMA Style

Silva DF, Viana AM, Mirante F, de Castro B, Cunha-Silva L, Balula SS. Removing Simultaneously Sulfur and Nitrogen from Fuel under a Sustainable Oxidative Catalytic System. Sustainable Chemistry. 2021; 2(2):382-391. https://0-doi-org.brum.beds.ac.uk/10.3390/suschem2020022

Chicago/Turabian Style

Silva, Dinis F., Alexandre M. Viana, Fátima Mirante, Baltazar de Castro, Luís Cunha-Silva, and Salete S. Balula. 2021. "Removing Simultaneously Sulfur and Nitrogen from Fuel under a Sustainable Oxidative Catalytic System" Sustainable Chemistry 2, no. 2: 382-391. https://0-doi-org.brum.beds.ac.uk/10.3390/suschem2020022

Article Metrics

Back to TopTop