Superoxide Ion as Oxidative Desulfurizing Agent for Aromatic Sulfur

Dec 27, 2016 - The generation of superoxide (O2•–) in ionic liquids (ILs) has been used to destroy hazardous materials. In this study, O2•– wa...
2 downloads 3 Views 1MB Size
Subscriber access provided by UNIV OF CALIFORNIA SAN DIEGO LIBRARIES

Article

Superoxide Ion as Oxidative Desulfurizing Agent for Aromatic Sulfur Compounds in Ionic Liquid Media Maan Hayyan, Abdulkader M. Alakrach, Adeeb Hayyan, Mohd Ali Hashim, and Hanee Farzana Hizaddin ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.6b02573 • Publication Date (Web): 27 Dec 2016 Downloaded from http://pubs.acs.org on December 30, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Sustainable Chemistry & Engineering is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Superoxide Ion as Oxidative Desulfurizing Agent for Aromatic Sulfur Compounds in Ionic Liquid Media Maan Hayyan1,2*, Abdulkader M. Alakrach1, Adeeb Hayyan1,2, Mohd Ali Hashim1,3, Hanee F. Hizaddin1,3 1

University of Malaya Centre for Ionic Liquids (UMCiL), University of Malaya, Kuala Lumpur, Petaling Jaya 50603, Malaysia 2 Institute of Halal Research University of Malaya (IHRUM), University of Malaya, 50603, Kuala Lumpur, Petaling Jaya 50603, Malaysia 3 Department of Chemical Engineering, University of Malaya, Kuala Lumpur, Petaling Jaya 50603, Malaysia * E-mail: [email protected], [email protected]; Tel/Fax No.: +6-03-7967-5311

ABSTRACT The generation of superoxide (O2−) in ionic liquids (ILs) has been used to destroy hazardous materials. In this study, O2− was used for the oxidative desulfurization of benzothiophene (BT) and dibenzothiophene (DBT) in two ILs as reaction media. The O2− was released by dissolving potassium superoxide in 1-butyl-1-methylpyrrolidinium bis(trifluoromethylsulfonyl)imide [BMPyrr][TFSI] and n-methoxyethyl-n-methylmorpholinium bis(trifluoromethylsulfonyl)imide [MOEMMor][TFSI]. The highest conversion percentages for BT in [MOEMMor][TFSI] and [BMPyrr][TFSI] were 99.4 and 96.6%, respectively, and the highest conversion percentages for DBT in [MOEMMor][TFSI] and [BMPyrr][TFSI] were 98.3 and 94.3%, respectively. As the temperature was increased, the reaction of the O2− with BT and DBT was enhanced significantly. The BT was converted easier than DBT due to the nucleophilic characteristic of the O2− and the higher electron density of the sulfur atom in DBT than that in BT. In addition, the ILs had a clear effect on the oxidative desulfurization reaction in which the O2− demonstrated a

1 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

higher conversion percentage in [MOEMMor][TFSI] than in [BMPyrr][TFSI]. This was the result of the stability of the O2− in the ILs that were used in the study. In addition, the σ-potential obtained using the ‘Conductor-like Screening Model for Real Solvent’ (COSMO-RS) indicated that [MOEMMor]+ was a superior hydrogen bond donor and had a stronger affinity for the hydrogen bond acceptor than [BMPyrr]+. The reaction mechanism was reported, and no toxic byproducts were detected. This is the first study in which O2− in ILs was used for the conversion of BT and DBT to non-hazardous materials, and this technique can be used for other hazardous materials. Keywords: oxidative desulfurization; reactive oxygen species; reaction engineering; green solvent; benzothiophene; dibenzothiophene; COSMO-RS; deep eutectic solvent.

1. INTRODUCTION Sulfur compounds are among the most undesirable heteroatomic components of petroleum. The high sulfur content in the fuel causes many environmental problems, since the combustion of these compounds generates toxic SOx gases (i.e., SO2 and SO3), which are major sources of acid rain and air pollution.1, 2 Consequently, the European Union issued new regulations in 2010 to reduce the sulfur content of European gasoline and diesel fuels to 10 ppm or less.3 Among several approaches, the hydrodesulfurization (HDS) process is a conventional industrial method to remove sulfur compounds from petroleum. However, this method has the disadvantage of severe operation requirements, e.g., temperatures of 300 to 400 °C, high H2 pressure ranging from 30 to 130 atm, and high consumption of catalysts.4 In addition, the benzothiophene (BT), dibenzothiophene (DBT), and their alkylated derivatives are difficult to remove using the HDS method.4,

5

For these reasons, other methods are needed as alternatives to overcome the

disadvantages of the HDS process. One recent approach is the purification of crude oil using 2 ACS Paragon Plus Environment

Page 2 of 29

Page 3 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

various techniques based on oxidative,6-8 adsorptive,9,

10

extractive,11 and biocatalytic

desulfurization.12 Oxidative desulfurization (ODS) has become one of the most attractive desulfurization methods due to its ability to oxidize (BT) and (DBT) to the corresponding sulfoxides and sulfones,13 which, after reaction, can be separated by polar solvents.14 ODS requires lower temperatures and pressure than HDS for the reactions, and it does not require the use of H2,15 making the ODS process a promising alternative for the removal of sulfur from crude oil.16-18 Currently, hydrogen peroxide (H2O2) is the most popular ODS agent because it has comparatively low cost and is environmentally friendly, i.e., it is non-polluting and not strongly corrosive. In addition, experiments can be conducted at mild conditions, and this is considered to be one of the most favorable advantages of using H2O2 as an oxidative agent.19-21 However, the use of H2O2 also has some disadvantages, such as the production of water after reaction, handling issues, and storage problems.22 Superoxide (O2−), which is good oxidizing agent, is generated by the reduction of O2 to produce an oxygen-centered radical.23 O2− was first reported in 1934 by Haber and Weiss; they generated the ion through the decomposition of H2O2 and during the oxidation reactions of ferrous ions (Fe2+) by oxygen in aqueous solutions.24, 25 However, due to O2− reaction tendency towards H2O (Eq. 1), O2− was generated in non-aqueous solutions, such as aprotic solvents, e.g., dimethyl sulfoxide (DMSO), acetonitrile (AcN), propylene carbonate, pyridine, and acetone.26, 27 Though, the utilization of aprotic solvents in the generation process of O2− in industrial applications was limited due to their volatility, low boiling points, and the negative ecological effects.28 2O2  H2O  O2  HOO  HO

1

3 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Recently, ILs, as neoteric solvents, have attracted the attention of people in both academia and industry because they have many advantages over conventional solvents. For instance, ILs have negligible vapor pressure, high thermal stability from below ambient to well over 300–400 °C,29, 30

wide electrochemical windows (4.5–6 V),31 and the ability to dissolve several inorganic and

organic compounds. In addition, they are recyclable.29, 32, 33 Currently, ILs and their analogs deep eutectic solvents are used extensively in research concerned with the extractive desulfurization of liquid fuels.34-43 For instance, Domańska et al. (2014) investigated the utilization of ILs in the extraction of sulfur compounds from gasoline and diesel oil.3 They used pyridinium-, pyrrolidinium-, morpholinium-, imidazolium-, piperidinium-, phosphonium-, and ammonium-based ILs. Pyrrolidinium provided the best extraction efficiency, and it was followed by morpholinium and pyridinium, which showed high selectivity and high distribution coefficient for the extraction of sulfur compounds. However, the sulfur removal selectivity from fuels is of significant importance as it is directly linked to the removal mechanisms. As a result, the selectivity is rather difficult to obtain unless the mechanism is fully understood.44 The ODS using ILs has also been explored recently.7, 34, 45-49 IL may act as a multi-task fluid since it can be used as extractant to sulfurs, catalyst to accelerate conversion reaction, or it can serve as both extractant and catalyst.50 In addition, the research on oxidizing agents’ stability in ILs as reaction media is of great importance.34 AlNashef et al. (2001, 2002) first reported the generation of a stable O2− in 1-methyl-3-nbutylimidazolium hexafluorophosphate.51,

52

Consequently, numerous studies have been

dedicated to investigating the stability of O2− in ILs.24, 27, 53-57

4 ACS Paragon Plus Environment

Page 4 of 29

Page 5 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

In this work, O2− was examined as a potential oxidizing agent for the ODS of BT and DBT. The ILs, 1-butyl-1-methylpyrrolidinium bis(trifluoromethylsulfonyl)imide [BMPyrr][TFSI] and nmethoxyethyl-n-methylmorpholinium bis(trifluoromethylsulfonyl)imide [MOEMMor][TFSI], were used as media to release O2− by dissolving potassium superoxide (KO2). These ILs demonstrated that the O2− had good stability, as reported in previous studies.53,

58

The

superoxide salts, particularly KO2, can reduce the required reactor volume considerably, reduce the weight of the oxidant dose, and provide flexible storage, thereby facilitating the transport and shipment requirements.59

2. RESULTS AND DISCUSSION 2.1. Process of converting sulfur compounds using O2− in ILs The process of refining crude petroleum has two stages. The first stage is the extraction of sulfur compounds by ILs followed by conversion of sulfur compounds into nontoxic byproducts using O2− in the second stage. Hence, the conversion was studied for sulfur compounds that were dissolved in ILs. The purpose of this reaction was to convert the toxic chemicals to nonhazardous compounds. As a result, the process would be practical for industrial scale applications. The tendency for superoxide to react with sulfur compounds is considerably high taking into account the nucleophilic characteristic of O2−. Figure 1 and Table S1 show the sulfur conversion percentages of BT and DBT, and they clearly show that O2− exhibited significant capability for converting the aromatic sulfur compounds. The percentages of the original sulfur compounds removed in the process were 99.4 and 98.3% for BT and DBT, respectively. It took 2.532 mmol of O2− to oxidize 0.112 and 0.109 mmol of BT and DBT, respectively. This

5 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

indicated that the ratio of sulfur to KO2 was 1:22.7 for BT and 1:23.3 for DBT. Figure S1 shows

100

the chromatograms of BT and DBT before and after reaction with O2− in ILs.

40

60

80

BT in [BMPyrr] DBT in [BMPyrr] BT in [MOEMMor] DBT in [MOEMMor]

0

20

Conversion (%)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 29

0

20

40

60

80

100

120

140

160

180

o

Temperature ( C)

Figure 1. Effect of temperature on the conversion percentage of BT and DBT by O2− as oxidizing agent.

The selection of suitable ILs as reaction media is an important aspect of the process because the ILs may have a direct effect on the stability of the O2− generated in the ILs. The stability of the O2− in [MOEMMor][TFSI] and [BMPyrr][TFSI] was studied by Hayyan et al. (2012, 2015), and they demonstrated that the O2− ions had high stability in both IL systems.53, 60 However, they were more stable in [MOEMMor][TFSI] than in [BMPyrr][TFSI], with consumption rates 0.651×10–3 and 4.049 ×10–3 mM/min, respectively. Figure 2 shows the effect of the structure of the ILs on the conversion results. However, it is noteworthy that the conversion of BT in [BMPyrr][TFSI] was higher under mild conditions (i.e., temperatures below 50 °C ) than the 6 ACS Paragon Plus Environment

Page 7 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

conversion in [MOEMMor][TFSI]. This was thought to be due to the viscosity of the ILs, since highly-viscous ILs will slow the dissolution rate of the superoxide salt. The viscosity of [MOEMMor][TFSI] (257.6 mPa s)61 was higher than that of [BMPyrr][TFSI] (68.16 mPa s)62 at room temperature (25 °C). However, the viscosity of [MOEMMor][TFSI] was reduced significantly as the temperature increased (i.e., 77.98 mPa s at 50 °C)61, and this facilitated the dissolution and diffusion of the reaction products. A linear trend was observed between the conversation rate and increase in the temperature. Ultimately, at 180 °C, the percentages of conversion of BT in [MOEMMor][TFSI] and [BMPyrr][TFSI] reached 99 and 96%, respectively, and the percentages of conversion of DBT in [MOEMMor][TFSI] and [BMPyrr][TFSI] reached 98 and 94%, respectively. These results indicated that the conversion of BT was easier than that of DBT, and this difference was attributed to two factors, i.e., the electron density around the sulfur atom and the steric hindrance. This is in agreement with a recent study conducted on the conversion of thiophene and 2-methylthiophene using O2−.63 The electron densities around the sulfur atoms in BT and DBT were calculated by Otsuki et al. (2002) as 5.739 and 5.758, respectively.64 Many previous studies have reported that the desulfurization performance decreased as the electron density decreased, so this is why DBT can be converted easier than BT.65-67 However, in this study, DBT was more difficult to convert than BT, and this was thought to be attributable to the nucleophilicity of the O2− and the electron density of the S atom. This is in accordance with the study of Roberts et al. (1981) in which it was observed that O2− ions are nucleophilic in solvents that do not provide protons.68 In addition, the solubility of BT is higher than that of DBT in all ILs (Table S2). The structures of the cations and anions of the ILs and their sizes affect the solubility of the sulfur compounds.69 This result indicated that, in addition to the ILs’ main role as the media in which these reactions

7 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

occur, the solubility of the sulfur compounds in the ILs also affected the reactions between the various sulfur compounds and the O2− ions. ILs can be regenerated by washing with organic solvents such as diethyl ether.70, 71 There are several methods can be used to extract K+ following superoxide release from its salt. One of which is the use of crown ethers as metal ion extractant to extract K+. Since the discovery of this class in 1967, crown ethers have been widely explored for metal ion extraction in liquid/liquid separation processes.72-74 It is also noteworthy that crown ethers are used to stimulate the solubility of superoxide salts for the O2− reaction with organic molecules.23

100 BT DBT 80

Conversion (%)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

60

40

20

0 [BMPyrr]

[MOEMMor]

ILs

Figure 2. Effect of the structure of the ILs on the conversion percentage of BT and DBT after reaction with the O2−.

A higher conversion percentage was observed for [MOEMMor][TFSI] than for [BMPyrr][TFSI], but the difference was small, i.e., 3% for BT and 4% for DBT. The COSMO-RS calculation was performed to investigate the effect of the structure of the ILs on the conversion percentage using the σ-profile and the σ-potential analysis, as illustrated in Figure 3. Based on the σ-profile 8 ACS Paragon Plus Environment

Page 8 of 29

Page 9 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

(Figure 3a), both cations had profiles that were complementary to those of BT and DBT, which explains the high percentage of conversion. However, based on σ-potential (Figure 3b), the [MOEMMor] cation is a better hydrogen bond donor and has a better affinity for the hydrogen bond acceptor than the [BMPyrr] cation. In addition, [MOEMMor] cation also can function as a hydrogen bond acceptor despite being a weak one. Probably, the tendency of the [MOEMMor] cation to function as both a hydrogen bond donor and hydrogen bond acceptor (weak one), and also the fact that [MOEMMor] is a better HBD than [BMPyrr], may contribute to the higher conversion percentage of the sulfur compounds. This supported the catalytic activity role of IL as reported earlier.63

9 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

35.0

(a) 30.0

25.0

[MOEMMor] cation

p (σ)

20.0

[BMPyrr] cation 15.0

[TFSI] anion BT

10.0

DBT

5.0

0.0 -0.03

-0.02

-0.01

0.0

0.01

0.02

0.03

σ (e Å-2) 1.0

(b) 0.8

0.6

µ(σ) (kcal mol-1 Å-2)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 29

0.4

[MOEMMor] [BMPyrr]

0.2

[TFSI] BT

0.0

DBT

-0.2

-0.4 -0.03 -0.6

-0.02

-0.01

0.0

0.01

0.02

0.03

σ (e Å-2)

Figure 3. (a) σ-profile and (b) σ-potential of sulfur compounds and ions comprising the ILs that were used in this study. 10 ACS Paragon Plus Environment

Page 11 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

The HOMO and LUMO orbitals also can be rough indicators of the interaction between species in the complex. Ideally, we want the HOMO and LUMO to come from opposite molecules to indicate a favorable interaction. For example, if the HOMO (or LUMO) comes from either the IL cation or anion, and LUMO (or HOMO) comes from the sulfur compound, it indicates favorable interaction between the IL and the sulfur compound. But, if both HOMO and LUMO come from the IL or both come from the sulfur compound, it cannot be concluded whether there is a favorable interaction between the two species (i.e., IL-sulfur compound). For the complex [MOEMMor][TFSI]-BT, HOMO comes from BT, whereas LUMO comes from the cation. This indicates that there is a strong interaction between the IL and BT. In contrast, for the complex [BMPyrr][TFSI]-BT, both HOMO and LUMO come from BT. This is indicative of a weak interaction between the IL and BT (Scheme 1). However, for complexes that contained DBT, both HOMO and LUMO came from the DBT in the complexes [MOEMMor][TFSI]-DBT and [BMPyrr][TFSI]-DBT (Scheme 2). In this case, the result is inconclusive, i.e., the HOMO and LUMO location does not indicate which IL would have the stronger interaction with DBT.

11 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 29

(a) [MOEMMor][TFSI]-BT HOMO

LUMO

(b) [BMPyrr][TFSI]-BT HOMO

LUMO

Scheme 1. HOMO and LUMO for BT interaction with (a) [MOEMMor][TFSI] and with (b) [BMPyrr][TFSI].

12 ACS Paragon Plus Environment

Page 13 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

(a) [MOEMMor][TFSI]-DBT HOMO

LUMO

(b) [BMPyrr][TFSI]-DBT HOMO

LUMO

Scheme 2. HOMO and LUMO for DBT interaction with (a) [MOEMMor][TFSI] and with (b) [BMPyrr][TFSI].

To study the effect of the reaction rate and temperature, the reaction process was investigated as a function of time (i.e., 0, 3, 6, 9, 12, and 15 h) with different temperatures (i.e., 27, 40, 50, 60, 13 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

80, 100, 120, 140, 160, and 180 °C). Figure 4 shows the effect of the reaction rate on the conversion process. The conversion of sulfur compounds at different temperatures increased for both BT and DBT as the temperature increased. This result can be explained by the increase in the solubility of KO2 in the ILs and by the decrease in the viscosities of the ILs as the temperature increased. In contrast, under mild conditions, Ahmed et al. (2015) did not obtain high conversion of sulfur compounds when they investigated the use of O2− in five phosphonium-based ILs.67 They used the O2− to convert two sulfur compounds (i.e., thiophene and DBT). The results were only 15% conversion of thiophene, and DBT did not react at all with the O2− over a period of 2 h. In addition, numerous studies have pointed out the effect of increasing the temperature on the conversion of sulfur compounds. For instance,

carbon

nanotubes were used as catalysts to activate oxygen for ODS of three sulfur compounds, i.e., BT, DBT, and 4,6-dimethyldibenzothiophene.22 A complete conversion of the sulfur compounds to the corresponding sulfones was achieved at 150 °C with a maximum time of 180 min.

100 BT in [BMPyrr] DBT in [BMPyrr] BT in [MOEMMor] DBT in [MOEMMor]

80

Conversion(%)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 29

60

40

20

0 0

3

6

9

12

15

Time (h)

Figure 4. Effect of time on the percentage of conversion of BT and DBT by O2− at 180 °C. 14 ACS Paragon Plus Environment

Page 15 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

2.2. Reaction mechanism Scheme 3 (below) illustrates the proposed mechanism. Several studies have reported similar mechanisms for the oxidation of thiophenes and aromatic sulfides.16,

75, 76

Attar et al. (1978)

reported that aromatic sulfur compounds are oxidized to the corresponding sulfoxides and then to sulfones (Eq. 2). The oxidation products begin to decompose as the temperature increases (Eq. 3).4

15 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 29

Scheme 3. Proposed reaction mechanisms of BT and DBT with O2−.

[O] [O] R-S-R'   R-SO-R'   R-SO2-R'

(2)

(heat) R-SO2-R'   hydrocarbons+SO2

(3)

16 ACS Paragon Plus Environment

Page 17 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

In addition, Chan (2010) reported that SO2, which is produced when sulfur compounds are oxidized by O2− in ILs, can be oxidized to sulfate (SO4−2).77 Sulfates can be precipitated by adding BaCl2 solution, and a white precipitate of barium sulfate (BaSO4) is produced. In this study, a similar result was obtained as it was verified by adding barium chloride to the sample. Zhang et al. (2014) studied the catalytic effect of carbon nanotubes to convert molecular oxygen to active oxygen (i.e., O2−), thereby degrading aromatic sulfur compounds.22 Our results were compatible with their results, in that DBT was oxidized to the corresponding sulfones. FTIR analysis was used to identify the new functional groups in the products of the reactions of BT and DBT with O2− ions in the IL media. Figure 5 shows the FTIR spectra of BT and DBT before and after reaction with O2− ions at 100 °C, and several new peaks were present in the spectra after the reactions. For instance, there is a sharp peak centered at ~ 1060 cm−1 in the spectra after reaction with the O2− that was not present in the spectra of pure BT. This peak resulting from the stretching of the S=O bond in sulfoxide. Furthermore, two other peaks, centered at about 1350 cm−1 and 1140 cm−1, may represent the asymmetric and symmetric stretches of the O=S=O of sulfones.78 Hence, the FTIR analysis suggested the presence of the S=O of sulfoxides and the O=S=O of sulfones as products of the conversion of the sulfur compounds. The GC/MS chromatograms of non-converted and converted sulfur compounds are shown in Figure S2. The GC peaks of sulfur compounds in the oxidized samples were smaller than the peaks in the nonoxidized samples. Consequently, the chromatograms indicated that the sulfur compounds were preferentially converted by this reaction. A new peak of biphenyl in the chromatogram after DBT conversion process is observed in Figure 6, which is in accordance with the results of the FTIR analysis, as shown in Figure 5.

17 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 29

BEFORE

AFTER

O=S=O

S=O

Figure 5. FTIR chromatograms for DBT before and after conversion by O2− in [MOEMMor][TFSI].

18 ACS Paragon Plus Environment

Page 19 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Figure 6. Mass spectra of the oxidation product of DBT and biphenyl.

3. CONCLUSIONS In this work, the conversion of BT and DBT by O2− was investigated in [MOEMMor][TFSI] and [BMPyrr][TFSI]. The O2− was generated by dissolving KO2 in ILs. The results showed that BT was converted easier than DBT. The ILs had a clear effect on the conversion process. The O2− produced higher conversion percentages in [MOEMMor][TFSI] than in [BMPyrr][TFSI] for both of the sulfur compounds. Increasing the temperature substantially affected the conversion process by reducing the viscosity of IL and increasing the solubility of superoxide salt, thereby

19 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 29

accelerating the reaction rate. The ILs showed promise as media for the generation of O2− and for the conversion of sulfur compounds.

4. EXPERIMENTAL SECTION 4.1. Materials Scheme 4 illustrates the chemical structures of the ILs, i.e., [BMPyrr][TFSI] (98%) and [MOEMMor][TFSI] (synthesis grade with a purity of 98%). The sulfur compounds, i.e., BT (99%) and DBT (99%) were supplied by Merck. Potassium superoxide (KO2) (99.9%) was purchased from Sigma-Aldrich, HPLC grade acetonitrile (AcN) (99.9%) was purchased from UNICHROM, and diethyl ether (99%) was supplied by Merck.

[BMPyrr]+

[MOEMMor]+

[TFSI]–

Scheme 4. Chemical structures of the ILs ions and sulfur compounds.

20 ACS Paragon Plus Environment

Page 21 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

4.2. Oxidative desulfurization of BT and DBT in ILs The ILs were dried overnight in a vacuum oven at 50 °C, as described previously.28,

79

Appropriate amounts of the sulfur compounds were added to a labeled vial containing the ILs. The IL-sulfur mixture was stirred for 30 min by a magnetic stirrer. A sample was then withdrawn and diluted with AcN, and analyzed using HPLC, as described before.71 Next, KO2 was added gradually to the vial that contained the IL-sulfur mixture with vigorous stirring. Samples were taken before and after the addition of KO2 by dissolving 0.1 g of the IL-sulfur mixture in 1 g of AcN. This procedure was repeated, and more KO2 was added until either the sulfur peak was not detected or did not change. This reaction was conducted at room temperature and at 40, 50, 60, 80, 100, 120, 140, 160, and 180 °C for both BT and DBT. This temperature range was selected because the temperatures were lower than the boiling points of BT (221 °C) and DBT (332 °C).80 The FTIR spectra were identified by FTIR spectrophotometer (Bruker TENSOR 27) using KBr disks at room temperature. The spectrophotometer was used to characterize the new products in the range of 650-4000 cm−1. The gas-chromatograph/mass spectrometer (GCMS-QP 2008, Shimadzu QP-5050A) with a DB-5 column (30 m × 0.32 mm × 1 µm) was used to identify BT and DBT before and after reaction with the O2− at 180 °C and to identify the products of the reaction. The injector temperature was set at 275 °C. The initial temperature was set at 35 °C for a 10-min holding time, and, then, the temperature was ramped to 250 °C at a heating rate of 10 °C min−1.

21 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 29

4.3. Solubility of sulfur compounds in the ILs The solubility of BT and DBT in the ILs was measured at room temperature by adding an excess amount of the sulfur compound to a glass vial that contained the IL with vigorous stirring. Then, the mixture settled to form a two-phase system. After reaching equilibrium, a sample was withdrawn carefully from the saturated IL phase and diluted in AcN for analysis using HPLC.

5. COMPUTATIONAL DETAILS 5.1

COSMO-RS Model

The ‘Conductor-like Screening Model for Real Solvents’ (COSMO-RS) is a dielectric model in which molecules are placed in a conductor as the reference state. The interaction energy in COSMO-RS is represented in terms of its polarization charge density, σ and σ’, which include the electrostatic misfit energy (Emisfit), hydrogen bond interaction (Ehb), and van der Waals interaction (EvdW).81 The most important descriptor in the COSMO-RS model is the local screening charge density, σ, because it is the only descriptor that is used to determine the interaction energies and, thus, all other statistical thermodynamic properties, such as chemical potential and activity coefficient. This descriptor can be calculated using any quantum chemical program in the form of “.cosmo file” output. In the .cosmo file, the surface of the species is divided into segments with a certain surface charge density. By applying a local averaging algorithm on the surface charge densities over effective contact segments, a probability function (σ-profile) can be plotted. The details of the local averaging algorithm are available in the reference.81

22 ACS Paragon Plus Environment

Page 23 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

5.2

Geometry Optimization and COSMO-RS Implementation

The job of optimizing the geometry was performed for each individual species (each sulfur compound, each cation, and each anion), for the ionic liquid complex, and for the complexes of the sulfur compounds with the ILs. Initial structures for each job were drawn, and the job was executed to perform geometry optimization at the Hartree-Fock level, and the 6-31G* basis was set using Turbomole version 4.0.1 software.82 The HOMO and LUMO orbitals were generated from the results of the calculations. Then, the optimized geometries of the compounds were used for a single point calculation to generate the .cosmo files using the density functional theory combined with the Becke-Perdew functional and the triple zeta valence potential (TZVP) basis set using Turbomole 4.0.1 software. Then, COSMOthermX software with the parameterization file BP_TZVP_C30_1401.ctd was used to generate the σ-profile and the σ-potential graphs.83

Acknowledgements The financial support of the University of Malaya HIR MOHE (D000003-16001) and UMRG (RP037B-15AET) is gratefully acknowledged. Supporting Information Available: This information is available free of charge via the Internet at http://pubs.acs.org/. HPLC and GCMS chromatograms are included.

23 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

5. REFERENCES (1) Ávila, B. M.; Pereira, V. B.; Gomes, A. O.; Azevedo, D. A. Speciation of organic sulfur compounds using comprehensive two-dimensional gas chromatography coupled to time-of-flight mass spectrometry: A powerful tool for petroleum refining. Fuel 2014, 126, 188-193. (2) Ma, X.; Zhou, A.; Song, C. A novel method for oxidative desulfurization of liquid hydrocarbon fuels based on catalytic oxidation using molecular oxygen coupled with selective adsorption. Catal. Today 2007, 123, 276-284. (3) Domańska, U.; Wlazło, M. Effect of the cation and anion of the ionic liquid on desulfurization of model fuels. Fuel 2014, 134, 114-125. (4) Attar, A.; Corcoran, W. H. Desulfurization of organic sulfur compounds by selective oxidation. 1. Regenerable and nonregenerable oxygen carriers. Ind. Eng. Chem. Prod. Res. Dev. 1978, 17, 102-109. (5) Wang, W.; Ma, T.; Lian, K.; Zhang, Y.; Tian, H.; Ji, K.; Li, G. Genetic analysis of benzothiophene biodesulfurization pathway of gordonia terrae strain C-6. PLoS One 2013, 8, e84386. (6) Qiu, J.; Wang, G.; Zeng, D.; Tang, Y.; Wang, M.; Li, Y. Oxidative desulfurization of diesel fuel using amphiphilic quaternary ammonium phosphomolybdate catalysts. Fuel Process. Technol. 2009, 90, 1538-1542. (7) Jiang, W.; Zhu, W.; Chang, Y.; Li, H.; Chao, Y.; Xiong, J.; Liu, H.; Yin, S. Oxidation of aromatic sulfur compounds catalyzed by organic hexacyanoferrates in ionic liquids with a low concentration of H2O2 as an oxidant. Energy Fuels 2014, 28, 2754-2760. (8) Wang, C.; Zhu, W.; Xu, Y.; Xu, H.; Zhang, M.; Chao, Y.; Yin, S.; Li, H.; Wang, J. Preparation of TiO2/g-C3N4 composites and their application in photocatalytic oxidative desulfurization. Ceram. Int. 2014, 40, 11627-11635. (9) Jung, B. K.; Jhung, S. H. Adsorptive removal of benzothiophene from model fuel, using modified activated carbons, in presence of diethylether. Fuel 2015, 145, 249-255. (10) Xiong, J.; Yang, L.; Chao, Y.; Pang, J.; Zhang, M.; Zhu, W.; Li, H. Boron nitride mesoporous nanowires with doped oxygen atoms for the remarkable adsorption desulfurization performance from fuels. ACS Sustainable Chem. Eng. 2016, 4, 4457-4464. (11) Kulkarni, P. S.; Afonso, C. A. Deep desulfurization of diesel fuel using ionic liquids: Current status and future challenges. Green Chem. 2010, 12, 1139-1149. (12) Soleimani, M.; Bassi, A.; Margaritis, A. Biodesulfurization of refractory organic sulfur compounds in fossil fuels. Biotechnol. Adv., 2007, 25, 570-596. (13) Wan, M.-W.; Biel, L. C. C.; Lu, M.-C.; de Leon, R.; Arco, S. Ultrasound-assisted oxidative desulfurization (UAOD) using phosphotungstic acid: effect of process parameters on sulfur removal. Desalin. Water Treat. 2012, 47, 96-104. (14) Capel-Sanchez, M. C.; Campos-Martin, J. M.; Fierro, J. L. Removal of refractory organosulfur compounds via oxidation with hydrogen peroxide on amorphous Ti/SiO2 catalysts. Energy Environ. Sci. 2010, 3, 328-333. (15) Ghaedian, M.; Bazmi, M.; Shafeghat, A.; Mohammadi, A. K.; Rabiei, Z.; Naderi, F. An investigation on oxidative desulfurization of gas condensate. Pet. Coal. 2013, 55, 361-365. (16) Zhu, W.; Wang, C.; Li, H.; Wu, P.; Xun, S.; Jiang, W.; Chen, Z.; Zhao, Z.; Li, H. Onepot extraction combined with metal-free photochemical aerobic oxidative desulfurization in deep eutectic solvent. Green Chem. 2015, 17, 2464-2472.

24 ACS Paragon Plus Environment

Page 24 of 29

Page 25 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

(17) Mamaghani, A. H.; Fatemi, S.; Asgari, M. Investigation of influential parameters in deep oxidative desulfurization of dibenzothiophene with hydrogen peroxide and formic acid. Int. J. Chem. Eng. 2013, 2013, 1-10. (18) Carnaroglio, D.; Gaudino, E. C.; Mantegna, S.; Moreira, E. M.; Vicente de Castro, A.; Flores, E. M. M.; Cravotto, G. Ultrasound-assisted oxidative desulfurization/denitrification of liquid fuels with solid oxidants. Energy Fuels 2014, 28, 1854-1859. (19) García-Gutiérrez, J. L.; Fuentes, G. A.; Hernández-Terán, M. E.; Murrieta, F.; Navarrete, J.; Jiménez-Cruz, F. Ultra-deep oxidative desulfurization of diesel fuel with H2O2 catalyzed under mild conditions by polymolybdates supported on Al2O3. Appl. Catal., A., 2006, 305, 1520. (20) Jiang, X.; Li, H.; Zhu, W.; He, L.; Shu, H.; Lu, J. Deep desulfurization of fuels catalyzed by surfactant-type decatungstates using H2O2 as oxidant. Fuel 2009, 88, 431-436. (21) Capel-Sanchez, M.; Perez-Presas, P.; Campos-Martin, J.; Fierro, J. Highly efficient deep desulfurization of fuels by chemical oxidation. Catal. Today 2010, 157, 390-396. (22) Zhang, W.; Zhang, H.; Xiao, J.; Zhao, Z.; Yu, M.; Li, Z. Carbon nanotube catalysts for oxidative desulfurization of a model diesel fuel using molecular oxygen. Green Chem. 2014, 16, 211-220. (23) Hayyan, M.; Hashim, M. A.; AlNashef, I. M. Superoxide ion: Generation and chemical implications. Chem. Rev. 2016, 116, 3029-3085. (24) Hayyan, M.; Mjalli, F. S.; AlNashef, I. M.; Hashim, M. A. Generation and stability of superoxide ion in tris(pentafluoroethyl)trifluorophosphate anion-based ionic liquids. J. Fluorine Chem. 2012, 142, 83-89. (25) Haber, F.; Weiss, J. The catalytic decomposition of hydrogen peroxide by iron salts. Proceed. of the Royal Soc. of London. Series A, Math. and Phys. Sci. 1934, 147, 332-351. (26) Sawyer, D. T.; Sobkowiak, A.; Roberts, J. L., Electrochemistry for Chemists. Wiley1995. (27) Hayyan, M.; Mjalli, F. S.; AlNashef, I. M.; Hashim, M. A. Stability and kinetics of generated superoxide ion in trifluoromethanesulfonate anion-based ionic liquids. Int. J. Electrochem. Sci. 2012, 7, 9658-9667. (28) Hayyan, M.; Mjalli, F. S.; Hashim, M. A.; AlNashef, I. M.; Tan, X. M. Electrochemical reduction of dioxygen in Bis (trifluoromethylsulfonyl) imide based ionic liquids. J. Electroanal. Chem. 2011, 657, 150-157. (29) Brennecke, J. F.; Maginn, E. J. Ionic liquids: Innovative fluids for chemical processing. AlChE J. 2001, 47, 2384-2389. (30) Villanueva, M.; Coronas, A.; García, J.; Salgado, J. Thermal stability of ionic liquids for their application as new absorbents. Ind. Eng. Chem. Res. 2013, 52, 15718-15727. (31) Hayyan, M.; Mjalli, F. S.; Hashim, M. A.; AlNashef, I. M.; Mei, T. X. Investigating the electrochemical windows of ionic liquids. J. Ind. Eng. Chem. 2013, 19, 106-112. (32) Singh, G.; Kumar, A. Ionic liquids: Physico-chemical, solvent properties and their applications in chemical processes. Indian J. Chem., Sect A 2008, 47, 495. (33) Zhang, S.; Sun, N.; He, X.; Lu, X.; Zhang, X. Physical properties of ionic liquids: database and evaluation. J. Phys. Chem. Ref. Data 2006, 35, 1475-1517. (34) Ibrahim, M. H.; Hayyan, M.; Hashim, M. A.; Hayyan, A. The role of ionic liquids in desulfurization of fuels: A review. Renew. Sustainable Energy Rev. 2016, http://dx.doi.org/10.1016/j.rser.2016.11.194. (35) Yu, F.; Liu, C.; Yuan, B.; Xie, P.; Xie, C.; Yu, S. Energy-efficient extractive desulfurization of gasoline by polyether-based ionic liquids. Fuel 2016, 177, 39-45. 25 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(36) Ren, T.-J.; Zhang, J.; Hu, Y.-H.; Li, J.-P.; Liu, M.-S.; Zhao, D.-S. Extractive desulfurization of fuel oil with metal-based ionic liquids. Chin. Chem. Lett. 2015, 26, 1169-1173. (37) Chen, X.; Yuan, S.; Abdeltawab, A. A.; Al-Deyab, S. S.; Zhang, J.; Yu, L.; Yu, G. Extractive desulfurization and denitrogenation of fuels using functional acidic ionic liquids. Sep. Purif. Technol. 2014, 133, 187-193. (38) Zhao, H.; Baker, G. A.; Wagle, D. V.; Ravula, S.; Zhang, Q. Tuning task-specific ionic liquids for the extractive desulfurization of liquid fuel. ACS Sustainable Chem. Eng. 2016, 4, 4771-4780. (39) Zhang, S.; Zhang, Q.; Zhang, Z. C. Extractive desulfurization and denitrogenation of fuels using ionic liquids. Ind. Eng. Chem. Res. 2004, 43, 614-622. (40) Huang, C.; Chen, B.; Zhang, J.; Liu, Z.; Li, Y. Desulfurization of gasoline by extraction with new ionic liquids. Energy Fuels 2004, 18, 1862-1864. (41) Li, J.-j.; Xiao, H.; Tang, X.-d.; Zhou, M. Green carboxylic acid-based deep eutectic solvents as solvents for extractive desulfurization. Energy Fuels 2016, 30, 5411-5418. (42) Liu, W.; Jiang, W.; Zhu, W.; Zhu, W.; Li, H.; Guo, T.; Zhu, W.; Li, H. Oxidative desulfurization of fuels promoted by choline chloride-based deep eutectic solvents. J. Mol. Catal. A: Chem. 2016, 424, 261-268. (43) Hadj-Kali, M. K.; Mulyono, S.; Hizaddin, H. F.; Wazeer, I.; El-Blidi, L.; Ali, E.; Hashim, M. A.; AlNashef, I. M. Removal of thiophene from mixtures with n-heptane by selective extraction using deep eutectic solvents. Ind. Eng. Chem. Res. 2016, 55, 8415-8423. (44) Li, H.; Zhu, W.; Zhu, S.; Xia, J.; Chang, Y.; Jiang, W.; Zhang, M.; Zhou, Y.; Li, H. The selectivity for sulfur removal from oils: An insight from conceptual density functional theory. AlChE J. 2016, 62, 2087-2100. (45) Ding, W.; Zhu, W.; Xiong, J.; Yang, L.; Wei, A.; Zhang, M.; Li, H. Novel heterogeneous iron-based redox ionic liquid supported on SBA-15 for deep oxidative desulfurization of fuels. Chem. Eng. J. 2015, 266, 213-221. (46) Zhu, W.; Huang, W.; Li, H.; Zhang, M.; Jiang, W.; Chen, G.; Han, C. Polyoxometalatebased ionic liquids as catalysts for deep desulfurization of fuels. Fuel Process. Technol. 2011, 92, 1842-1848. (47) Yang, H.; Jiang, B.; Sun, Y.; Hao, L.; Huang, Z.; Zhang, L. Synthesis and oxidative desulfurization of novel lactam-based Brønsted-Lewis acidic ionic liquids. Chem. Eng. J. 2016, 306, 131-138. (48) Bhutto, A. W.; Abro, R.; Gao, S.; Abbas, T.; Chen, X.; Yu, G. Oxidative desulfurization of fuel oils using ionic liquids: A review. J Taiwan Inst. Chem. Eng. 2016, 62, 84-97. (49) Zhao, H.; Baker, G. A.; Zhang, Q. Design rules of ionic liquids tasked for highly efficient fuel desulfurization by mild oxidative extraction. Fuel 2017, 189, 334-339. (50) Chen, X.; Guan, Y.; Abdeltawab, A. A.; Al-Deyab, S. S.; Yuan, X.; Wang, C.; Yu, G. Using functional acidic ionic liquids as both extractant and catalyst in oxidative desulfurization of diesel fuel: An investigation of real feedstock. Fuel 2015, 146, 6-12. (51) AlNashef, I. M.; Leonard, M. L.; Matthews, M. A.; Weidner, J. W. Superoxide electrochemistry in an ionic liquid. Ind. Eng. Chem. Res. 2002, 41, 4475-4478. (52) AlNashef, I. M.; Leonard, M. L.; Kittle, M. C.; Matthews, M. A.; Weidner, J. W. Electrochemical generation of superoxide in room-temperature ionic liquids. Electrochem. SolidState Lett 2001, 4, D16-D18.

26 ACS Paragon Plus Environment

Page 26 of 29

Page 27 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

(53) Hayyan, M.; Mjalli, F. S.; AlNashef, I. M.; Hashim, M. A. Chemical and electrochemical generation of superoxide ion in 1-butyl-1-methylpyrrolidinium bis (trifluoromethylsulfonyl) imide. Int J Electrochem Sci 2012, 7, 8116-8127. (54) Katayama, Y.; Sekiguchi, K.; Yamagata, M.; Miura, T. Electrochemical behavior of oxygen/superoxide ion couple in 1-butyl-1-methylpyrrolidinium bis (trifluoromethylsulfonyl) imide room-temperature molten salt. J. Electrochem. Soc. 2005, 152, E247-E250. (55) Evans, R. G.; Klymenko, O. V.; Saddoughi, S. A.; Hardacre, C.; Compton, R. G. Electroreduction of oxygen in a series of room temperature ionic liquids composed of group 15centered cations and anions. J. Phys. Chem. B 2004, 108, 7878-7886. (56) Hayyan, M.; Mjalli, F. S.; Hashim, M. A.; AlNashef, I. M. An investigation of the reaction between 1-butyl-3-methylimidazolium trifluoromethanesulfonate and superoxide ion. J. Mol. Liq. 2013, 181, 44-50. (57) AlNashef, I. M.; Hashim, M. A.; Mjalli, F. S.; Ali, M. Q. A.-h.; Hayyan, M. A novel method for the synthesis of 2-imidazolones. Tetrahedron Lett. 2010, 51, 1976-1978. (58) Hayyan, M.; Mjalli, F. S.; Hashim, M. A.; AlNashef, I. M. Generation of superoxide ion in pyridinium, morpholinium, ammonium, and sulfonium-based ionic liquids and the application in the destruction of toxic chlorinated phenols. Ind. Eng. Chem. Res. 2012, 51, 10546-10556. (59) Chan, N. Y.; Lin, T.-Y.; Yen, T. F. Superoxides: Alternative oxidants for the oxidative desulfurization process. Energy Fuels 2008, 22, 3326-3328. (60) Hayyan, M.; Hashim, M. A.; AlNashef, I. M. Kinetics of superoxide ion in dimethyl sulfoxide containing ionic liquids. Ionics 2015, 21, 719-728. (61) Ibrahim, M. H.; Hayyan, M.; Hashim, M. A.; Hayyan, A.; Hadj-Kali, M. K. Physicochemical properties of piperidinium, ammonium, pyrrolidinium and morpholinium cations based ionic liquids paired with bis(trifluoromethylsulfonyl)imide anion. Fluid Phase Equilib. 2016, 427, 18-26. (62) Zarrougui, R.; Dhahbi, M.; Lemordant, D. Effect of temperature and composition on the transport and thermodynamic properties of binary mixtures of ionic liquid n-butyl-nmethylpyrrolidinium bis(trifluoromethanesulfonyl)imide and propylene carbonate. J. Solution Chem. 2010, 39, 921-942. (63) Hayyan, M.; Ibrahim, M. H.; Hayyan, A.; AlNashef, I. M.; Alakrach, A. M.; Hashim, M. A. Facile route for fuel desulfurization using generated superoxide ion in ionic liquids. Ind. Eng. Chem. Res. 2015, 54, 12263-12269. (64) Otsuki, S.; Nonaka, T.; Takashima, N.; Qian, W.; Ishihara, A.; Imai, T.; Kabe, T. Oxidative desulfurization of light gas oil and vacuum gas oil by oxidation and solvent extraction. Energy Fuels 2000, 14, 1232-1239. (65) Akbari, A.; Omidkhah, M.; Darian, J. T. Facilitated and selective oxidation of thiophenic sulfur compounds using MoOx/Al2O3–H2O2 system under ultrasonic irradiation. Ultrason. Sonochem. 2015, 23, 231-237. (66) Ogunlaja, A. S.; Khene, S.; Antunes, E.; Nyokong, T.; Torto, N.; Tshentu, Z. R. The development of catalytic oxovanadium (IV)-containing microspheres for the oxidation of various organosulfur compounds. Appl. Catal., A 2013, 462, 157-167. (67) Ahmed, O. U.; Mjalli, F. S.; Al Wahaibi, T.; Al-Waheibi, Y. M.; AlNashef, I. M. Stability of superoxide ion in phosphonium-based ionic liquids. Ind. Eng. Chem. Res. 2015, 54, 2074–2080.

27 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(68) Roberts Jr, J. L.; Sawyer, D. T. Facile degradation by superoxide ion of carbon tetrachloride, chloroform, methylene chloride, and p, p'-DDT in aprotic media. J. Am. Chem. Soc. 1981, 103, 712-714. (69) Kowsari, E., Recent Advances in the Science and Technology of Desulfurization of Diesel Fuel Using Ionic Liquids. INTECH 2013. (70) Pozo-Gonzalo, C.; Howlett, P. C.; Hodgson, J. L.; Madsen, L. A.; MacFarlane, D. R.; Forsyth, M. Insights into the reversible oxygen reduction reaction in a series of phosphoniumbased ionic liquids. Phys. Chem. Chem. Phys. 2014, 16, 25062-25070. (71) Hayyan, M.; Mjalli, F. S.; Hashim, M. A.; AlNashef, I. M.; Al-Zahrani, S. M.; Chooi, K. L. Generation of superoxide ion in 1-butyl-1-methylpyrrolidinium trifluoroacetate and its application in the destruction of chloroethanes. J. Mol. Liq. 2012, 167, 28-33. (72) Kolthoff, I. M. Ionic strength effect on extraction of potassium complexed with crown ether 18-crown-6. Preliminary communication. Can. J. Chem. 1981, 59, 1548-1551. (73) Visser, A. E.; Swatloski, R. P.; Reichert, W. M.; Griffin, S. T.; Rogers, R. D. Traditional extractants in nontraditional solvents:  Groups 1 and 2 extraction by crown ethers in roomtemperature ionic liquids. Ind. Eng. Chem. Res. 2000, 39, 3596-3604. (74) Luo, H.; Dai, S.; Bonnesen, P. V. Solvent extraction of Sr2+ and Cs+ based on roomtemperature ionic liquids containing monoaza-substituted crown ethers. Anal. Chem. 2004, 76, 2773-2779. (75) Li, F.-t.; Liu, Y.; Sun, Z.-m.; Chen, L.-j.; Zhao, D.-s.; Liu, R.-h.; Kou, C.-g. Deep extractive desulfurization of gasoline with x Et3NHCl· FeCl3 ionic liquids. Energy Fuels 2010, 24, 4285-4289. (76) Ismagilov, Z.; Yashnik, S.; Kerzhentsev, M.; Parmon, V.; Bourane, A.; Al-Shahrani, F.; Hajji, A.; Koseoglu, O. Oxidative desulfurization of hydrocarbon fuels. Catal. Rev. 2011, 53, 199-255. (77) Chan, N. Y. Superoxide Radical and UV Irradiation in Ultrasound Assisted. Superoxide Radical and UV Irradiation in Ultrasound Assisted, PhD thesis, University of Southern California 2010. (78) Smith, B. C., Infrared Spectral Interpretation: A Systematic Approach. CRC press 1998. (79) Hayyan, M.; Hashim, M.; AlNashef, I.; Tan, X.; Chooi, K. Generation of superoxide ion in trihexyl (tetradecyl) phosphonium bis (trifluoromethylsulfonyl) imide room temperature ionic liquid. J. Applied Sci. 2010, 10, 1176-1180. (80) Seredych, M.; Bandosz, T. J. Selective adsorption of dibenzothiophenes on activated carbons with Ag, Co, and Ni species deposited on their surfaces. Energy Fuels 2009, 23, 37373744. (81) Klamt, A.; Eckert, F. COSMO-RS: A novel and efficient method for the a priori prediction of thermophysical data of liquids. Fluid Phase Equilib. 2000, 172, 43-72. (82) TURBOMOLE Users' Manual, 2014. (83) COSMOthermX. A Graphical User Interface to the COSMOtherm Program. User Guide., D. t. B. S., D-51381 Leverkusen, Germany.

28 ACS Paragon Plus Environment

Page 28 of 29

Page 29 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

For Table of Contents Only

Superoxide Ion as Oxidative Desulfurizing Agent for Aromatic Sulfur Compounds in Ionic Liquid Media Maan Hayyan, Abdulkader M. Alakrach, Adeeb Hayyan, Mohd Ali Hashim, Hanee F. Hizaddin

Synopsis Superoxide ion generated in ionic liquids can be used as a potential and sustainable oxidizing agent in desulfurization process.

29 ACS Paragon Plus Environment