You are on page 1of 29

Subscriber access provided by University of Glasgow Library

Separations
Combined Extractive Dearomatization, Desulfurization and Denitrogenation
of Oil Fuels Using Deep Eutectic Solvents: A Parametric Study
Samah E. E. Warrag, Ahmad S. Darwish, farah abuhatab,
Idowu A Adeymi, Maaike C. Kroon, and Inas M. Alnashef
Ind. Eng. Chem. Res., Just Accepted Manuscript • DOI: 10.1021/acs.iecr.0c01360 • Publication Date (Web): 02 Jun 2020
Downloaded from pubs.acs.org on June 3, 2020

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 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 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.

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 28 Industrial & Engineering Chemistry Research

1
2
3
4 Combined Extractive Dearomatization, Desulfurization and
5
6 Denitrogenation of Oil Fuels Using Deep Eutectic Solvents: A
7
8 Parametric Study
9
10
11 Samah E. E. Warrag*, Ahmad S. Darwish*, Farah O.S. Abuhatab, Idowu A. Adeyemi, Maaike C. Kroon,
12
and Inas M. AlNashef**
13
14
15 Department of Chemical Engineering, Khalifa University of Science and Technology,
16 Center for Membrane and Advanced Water Technology (CMAT),
17
18 P.O. Box 127788, Abu Dhabi, United Arab Emirates (UAE)
19
20
21 *The authors have equally contributed to this work
22
23 **Corresponding Author: Inas M. AlNashef, enas.nashef@ku.ac.ae
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 Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 2 of 28

1
2
3
4
Abstract
5 Industrially, deep dearomatization of oil fuels is achieved via catalytic hydrodearomatization
6
7 (HDA). However, this process suffers from several drawbacks. The most pronounced disadvantages are
8
9
the intensive energy consumption and the low efficiency towards some aromatic species. With the aim of
10 lowering energy consumption as well as improving the removal efficiency of this process, selective
11
12 liquid-liquid extraction was proposed in this work. A phosphonium-based deep eutectic solvent (DES)
13 composed of methyltriphenylphosphonium bromide (MTPPBr) and triethylene glycol (TEG) in a molar
14
15 ratio equal to (1 (MTPPBr):4 (TEG)) was selected for this investigation. The DES was characterized by
16
its water content, density, viscosity, and degradation temperature. Toluene, thiophene, and quinoline were
17
18 selected to represent the aromatic species in the oil. While the oil fuel was represented by n-heptane.
19
20 Next, the solubility of toluene, thiophene, quinoline, and n-heptane in the pure TEG and MTPPBr: TEG
21 was measured at 298.2 K and 1.01 bar. To assess the selectivities and the solute distribution coefficients
22
23 of the DES for each compound, liquid-liquid equilibrium (LLE) data for the systems {toluene + n-heptane
24 + MTPPBr:TEG}, {thiophene + n-heptane + MTPPBr:TEG}, and {quinoline + n-heptane +
25
26 MTPPBr:TEG} were reported at 298.2 K and 1.01 bar.
27
28
Afterwards, a parametric study on an arbitrary oil model of {20% toluene + 2% thiophene + 2%
29 quinoline + 76% n-heptane} was conducted by firstly testing the single-stage liquid-liquid extraction
30
31 efficiency for each impurity “toluene, thiophene, and quinoline” at 298.2 K and 1.01 bar. Then, the effects
32 of various operating parameters including the influence of the extraction temperature, the effect of
33
34 solvent-to-feed ratio (S:F), and the effect of the initial concentration of the impurity were investigated.
35
Moreover, the number of extraction stages was estimated . Lastly, the effect of the repetitive use of DES
36
37 as well as the possibility of DES regeneration was studied.
38
39
40 1 Introduction
41
42 It is well known that crude oil is a complex mixture of hydrocarbons including paraffins,
43 naphthenes and a range of aromatics (e.g.: alkylbenzenes, polyaromatics, and sulfur/nitrogen aromatics).1
44
45 The combustion of oil fuels results in a substantial amount of gaseous pollutants such as SOx, NOx, COx,
46
unburnt hydrocarbons including polycyclic aromatic hydrocarbons (PAHs) and particulate matters (PM).
47
48 Thus, the oil fuel composition could highly impact the environment as well as human health.
49
50 Accordingly, aiming to improve the air quality, stringent environmental regulations have been introduced
51 worldwide to reduce aromatics, sulfur-, and nitrogen-containing aromatics content of oil fuel.2 In other
52
53 words, emissions are to be controlled by the production of clean-burn oil fuels. Moreover, sulfur- and
54 nitrogen-containing aromatics are problematic in oil and gas processing as they are causes of catalyst
55
56
57
58
59
60 ACS Paragon Plus Environment
Page 3 of 28 Industrial & Engineering Chemistry Research

1
2
3 poisoning or deactivation, equipment corrosion, and gum formation. While high aromatic content fuels
4
5 suffer from low cetane number, and consequently, from poor fuel combustion.3,4
6
Conventionally, deep combined dearomatization “reducing the aromatics and sulfur/nitrogen
7
8 aromatics content” of fuels is achieved by hydrotreatment.5 In this process, the aromatic compound is
9
10 hydrogenated using an ultra-high activity noble metal catalyst. However, this process has many
11 drawbacks such as (i) harsh operating conditions, (ii) the use of expensive catalysts, (iii) large hydrogen
12
13 consumption, (iv) the need for a hydrogen production unit, and unit to treat the hydrogenation products
14 (i.e: hydrogen sulfide and ammonia).6 Furthermore, it has been reported that the hydrotreatment process is
15
16 efficient ‘to some extent’ for the removal of some heteroatomic compounds “sulfur-/nitrogen- containing
17
aromatic” while aromatic hydrocarbons are inefficiently hydrogenated.6 And in most cases, achieving low
18
19 aromatic fuels “~ < 10 wt%” via hydrotreatment might require two- or multi-stage hydroprocessing, e.g.
20
21 the Shell middle distillate hydrogenation process,7 or might require the application of high pressure in a
22 range of 10 – 15 MPa for single-stage treatment.7
23
24 Owing to the drawbacks of the hydrogenation process, the development of alternative methods
25
for improving the quality of oil fuels has become a hot research topic. Many new processes3,8,9 including
26
27 biodesulfurization, oxidative desulfurization, adsorption, and liquid-liquid extraction, have been
28
29 proposed in the literature to overcome the drawbacks of the hydrogenation process as being more
30 environmentally and economically feasible. However, combined extractive dearomatization,
31
32 desulfurization and denitrogenation appears to be especially promising. In this liquid-liquid extraction
33 process, it is proposed to selectively remove the aromatic, sulfur- and nitrogen-containing aromatics from
34
35 oil beforehand (and thus preventing the production of H2S and NH3 during hydrotreatment) using a highly
36
selective solvent as an extractive agent. This allows removal of (sulfur- and nitrogen-containing) aromatic
37
38 compounds from oil at mild conditions and without any hydrogen consumption, making the process much
39
40 more energy- and cost-effective. Moreover, another advantage is that the structure of the aromatic
41 compound does not change and it could be utilized as valuable raw materials for other industries. This
42
43 been said, an extracting solvent must primarily exhibit high capacity, high solute distribution ratio, high
44 selectivity, easy recoverability from the hydrocarbon products, environmental friendliness, and low cost.
45
46 Solvents like sulfolane,10 N-formylmorpholine,11 oxidative extractants like dimethyl sulfoxide (DMSO),12
47
48
and glycols13 have been used for dearomatization, however, solvent volatility, solvent recoverability,
49 extraction performance, and/or thermal stability limited their deployment.
50
51 In line with the global direction towards the application of “green solvents” in processing, ionic
52 liquids (ILs) have been widely proposed as solvents for dearomatization of oil.14–17 Due to the
53
54 low/negligible volatility, thermal stability and the limited miscibility of oil in ILs, their use may offer
55
many advantages over the conventional solvent. Excellent recent research paper focused on the combined
56
57
58
59
60 ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 4 of 28

1
2
3 extractive dearomatization, desulfurization and denitrogenation using ILs citing remarkable
4
5 extraction efficiencies for aromatic and hetero-aromatics from a wide varity of pyrolysis and
6
7 reformer gasoline.18,19 Also in some cases a mixture of ILs has been proposed to further enhance
8
9
the solvent performance.20,21 However, the environmental credentials of ILs have been widely
10 misinterpreted, and severe toxicity and the poor biodegradability of many ILs has been reported.22–24
11
12 Moreover, the high synthesis cost of ILs lowers their economical value.
13 In the last two decades, a new generation of sustainable solvents was discovered and was coined
14
15 as deep eutectic solvents (DESs).25 They are commonly described as designer solvents that results from
16
mixing one or more hydrogen bond acceptors (HBA) and one or more hydrogen bond donors (HBD) that
17
18 are able to undergo hydrogen bonding leading to the formation of a eutectic mixture showing a melting
19
20 point far below that of its constituents.26–29 DESs are arising as promising alternatives of ILs that provide
21 similar properties (i.e. low vapor pressure, wide liquid range, thermal stability, and low flammability)
22
23 while avoiding some drawbacks including simple preparation from low-cost, naturally occurring and
24
often biodegradable depending on the choice of the starting materials.
25
26 Regarding the separation of aromatics and hetero-aromatics from oil using DESs, several research
27
28 papers appeared in the literature showing a good extraction performance.30 For example, Gonzalez et al.31
29 featured DESs that exhibit higher selectivities than the benchmark solvent “Sulfolane”. However, most of
30
31 these studies investigated the removal of only either aromatic (e.g. toluene), or sulfur-containing
32 aromatic, or nitrogen-containing aromatic from oil hydrocarbons. Only very few studies looked at the
33
34 removal of a mixture of aromatics with different natures using DESs.32–34 Larriba et. al.35 investigated
35
36 the application of DESs for the extraction benzene, toluene, ethylbenzene, and xylene (BETX)
37
from reformer and pyrolysis gasoline using choline chloride-based DESs. Their results showed
38
39 that almost complete separation of benzene and substantial reduction of the aromatic was
40
41 achieved. Thus, the main objective of this work was to simultaneously remove an aromatic, a sulfur-
42
containing aromatic, and a nitrogen-containing aromatic from oil using a phosphonium-based DES. A key
43
44 aim was to provide information about the DES extraction performance in a process that mimics the
45
46 conventional combined dearomatization process. An arbitrary oil model was selected as {20% toluene +
47 2% thiophene + 2% quinoline + 76% n-heptane}. Phosphonium-based DESs have been extensively
48
49 studied 36–41 for the removal of either aromatics or sulfur-/nitrogen-containing aromatics. They showed
50
good extraction efficiencies, distribution coefficients as well as selectivities. Building on the literature
51
52 findings, a DES consisting of methyltriphenylphosphonium bromide (MTPPBr) and triethylene glycol
53
54 (TEG) in molar ratio equals to (1 (MTPPBr):4 (TEG)) was chosen for this application. And since glycols
55
56
57
58
59
60 ACS Paragon Plus Environment
Page 5 of 28 Industrial & Engineering Chemistry Research

1
2
3 have been frequently used as extractants for aromatics, the extraction performance of the pure TEG was
4
5 compared to that of the MTPPBr:TEG.
6
First, the DES was characterized for its water content, density, viscosity, and degradation
7
8 temperature. Then, the solubility of toluene, thiophene, quinoline, and n-heptane in the pure TEG and the
9
10 MTPPBr: TEG was measured at 298.2 K and 1.01 bar. To provide insights about the selectivities and the
11 solute distribution coefficients of the DES, the liquid-liquid equilibrium (LLE) data for the systems
12
13 {toluene + n-heptane + MTPPBr:TEG}, {thiophene + n-heptane + MTPPBr:TEG}, and {quinoline + n-
14 heptane + MTPPBr:TEG} were reported at 298.2 K and 1.01 bar. Next, single-stage liquid-liquid
15
16 extraction experiments for the impurities “toluene, thiophene and quinoline” from the oil model using the
17
DES and the pure TEG were conducted at 298.2 K and 1.01 bar. Then, the effects of various operating
18
19 parameters including the influence of the extraction temperature, the effect of solvent to feed ratio (S:F),
20
21 and the effect of the initial concentration of the impurity were investigated. Furthermore, the extraction
22 efficiency for each impurity in 9 extraction stages was evaluated. Finally, the effect of the repetitive use
23
24 of DES as well as the possibility of DES regeneration was assessed.
25
26
27 2 Experimental Procedures
28
29 2.1 Materials
30 Table 1 lists the chemicals used in this work and their respective CAS numbers, purity (as stated by
31
32 the suppliers) and sources. All chemicals were used as obtained.
33
34 Table 1: List of chemicals used and their specifications.
35
36 Chemical CAS number Purity (wt%) Source
37 n-Heptane 142-82-5 ≥99.0 Sigma-Aldrich
38
39 Toluene 108-88-3 ≥ 99.5 Sigma-Aldrich
40
41 Thiophene 110-02-1 ≥ 99.0 Sigma-Aldrich
42
Triethylene glycol 112-27-6 ≥99.0 Sigma-Aldrich
43
44 Quinoline 91-22-5 ≥98.0 Sigma-Aldrich
45
46 Ethanol 64-17-5 ≥ 99.8 Sigma-Aldrich
47
48 Methyltriphenylphosphonium bromide 1779-49-3 ≥98.0 Sigma-Aldrich
49
50
51
2.2 DES Preparation and Characterization
52
53 Precisely weighed hydrogen bond acceptor (HBA) and hydrogen bond donor (HBD) were mixed
54
55 in a screw-capped bottle using Shimadzu balance AUX220 with uncertainty in the measurement of
56
57
58
59
60 ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 6 of 28

1
2
3 ±0.0002 g. The mixture was then kept in a temperature-controlled incubated shaker (IKA KS 4000) i-
4
5 control with temperature stability of ± 0.1 K; at a temperature of 338.2 K; while being stirred at around
6
~300 rpm until a clear homogeneous liquid was formed. The DES prepared was composed of
7
8 methyltriphenylphosphonium bromide as HBA and triethylene glycol as HBD with a molar ratio of 1:4
9
10 (MTPPBr:TEG). The water content of the DES was measured using the Karl-Fischer titration (GRS
11 Scientific/ Aquamax KF Coulometric) method. Then, the solvent degradation was assessed using a
12
13 dynamic thermogravimetric analyzer (TGA) ( Netzsch STA 449 F Jupiter, at a heating rate of 10 K.min-1
14 from 298.2 K to 723 K). Also, the viscosity of the DES was measured using Thermo Scientific’s HAAKE
15
16 Rheo Stress 6000 rheometer at a shear rate of 240 s-1 for 120 s with an average standard deviation of 8
17
mPa.s -1. Furthermore, the DES density was measured at a temperature range from 293.2 – 368.2 K using
18
19 an Anton Paar (DMA 5000 M) with an uncertainty of 0.00001 g.cm-3.
20
21 2.3 Determination of the Solubility and the pseudo-Ternary LLE Data
22
23 The solubility of toluene, thiophene, quinoline, and n-heptane in the pure TEG and the
24
25
MTPPBr:TEG was determined at 298.2 K and 1.01 bar according to the equilibrium cell method. Equal
26 amounts of each compound and the TEG or MTPPBr:TEG were placed in 10 mL vials. The quantity of
27
28 the added components was measured using Shimadzu balance AUX220 with uncertainty in the
29 measurement of ±0.0002 g. The vials were stirred for 4 h at 1000 rpm using an Eppendorf thermomixer at
30
31 a temperature of 298.2 K and kept overnight at a controlled temperature of 298.2 K to reach equilibrium.
32
Thereafter, a sample of the DES-rich phase was taken using a needled syringe. It should be mentioned
33
34 that no phase separation was observed for the systems {thiophene + TEG or MTPPBr:TEG } and
35
36 {quinoline + TEG or MTPPBr:TEG }, which implies that both thiophene and quinoline exhibits full
37 solubility in the pure TEG and MTPPBr:TEG.
38
39
Similarly, the pseudo-ternary LLE data of the systems: {toluene + n-heptane + MTPPBr:TEG},
40
41 {thiophene + n-heptane + MTPPBr:TEG}, and {quinoline + n-heptane + MTPPBr:TEG}were obtained at
42
43 298.2 K and 1.01 bar using the equilibrium method. In those experiments, a pseudo-ternary mixture “6g”
44 of the n-heptane, toluene or thiophene or quinoline and the DES were prepared. Where, two samples one
45
46 from the n-alkane-rich phase and the DES-rich phase were withdrawn using a needled syringe. Here, the
47 term “pseudo” was introduced to indicate that the DES was assumed to behave as one species. This
48
49 assumption was later validated by analyzing a sample of the n-alkane-rich phase “after extraction” via
50
Fourier Transform Infrared Spectrometer (FTIR, Perkin Elmer, USA). The analysis was conducted in
51
52 transmittance mode and the range of the spectrum was 4000-400 cm-1.
53
54 All the prepared samples were diluted using ethanol and analyzed via gas chromatography (GC;
55
56 Agilent 6890 N). The GC method used for the analysis is described in Table 2. Samples of known
57
58
59
60 ACS Paragon Plus Environment
Page 7 of 28 Industrial & Engineering Chemistry Research

1
2
3 composition were analyzed using the GC in order to validate the method. The obtained root-mean-square
4
5 deviation (RMSD) was equal to 0.005. During the experiments, all samples were analyzed in triplicates
6
and the statistical uncertainty in the measurement was found to be less than <0.005. Owing to the low
7
8 volatility of the DES, it cannot be quantified via GC; therefore, its concentration was determined via mass
9
10 balance calculation.
11
12 Table 2: GC methods and specifications
13 Column Agilent J&W HP-5 (30 m × 0.32 mm × 0.25 μm)
14
15
Detector Flame ionization detector (FID)
16 Injector temperature 548.2 K
17 Oven temperature profile 323.2 – 523.2 K at 20 K.min-1
18 Detector temperature 473.2 K
19
Carrier gas Helium
20
21 Gas flow rate 2 mL.min-1
22 Injection volume 1 μL
23 Retention time repeatability 0.003
24
Method verification (Max RMSD) 0.005
25
26 Method verification (statistical uncertainty) 0.004
27
28 2.4 Liquid-Liquid Extraction Experiment and Parametric Study
29 Single-stage liquid-liquid extraction for toluene, thiophene, and pyridine from n-heptane using
30
31 the pure TEG and MTPPBr:TEG was conducted at 298.2 K and 1.01 bar. The starting mixture was
32
composed of {20% toluene + 2% thiophene + 2% quinolone + 76% n-heptane} on mass basis. The
33
34 hydrocarbon mixture was added to the TEG or the DES at 1:1 solvent-to-feed mass ratios using Shimadzu
35
36 balance AUX220 with an uncertainty of ±0.0002 g. Then the mixtures were stirred for 2 h at 1000 rpm
37 using an Eppendorf thermomixer and kept to settle for 2 h at 298.2 K. A sample from the n-alkane-rich
38
39 phase “ top phase” was taken using a syringe without disturbing the phases coexistence interface. The
40
samples were analyzed via GC following the method described in section 2.3.
41
42
43 The stirring time was varied from 0 min (no mixing) to 90 min at 298.2 K and 1.01 bar, followed
44 by 1 h of settling. As can be observed from Figure S.1, The efficiency increased rapidly within the initial
45
46 20 min for the three impurities. Afterward, a slight increase in efficiency was observed between 20 and 60
47 min of the stirring time. The efficiency became rather constant between the 60 to 90 min. It can be
48
49 concluded that 60 min of stirring is sufficient to reach an equilibrium of interactions between the DES and
50
51
the oil mixture. Thereafter, the stirring time was fixed to 60 min for the subsequent experiments.
52
53 The effect of several process parameters including the the effect of extraction temperature, the
54 effect solvent-to-feed ratio, and the effect of feed initial concentration on the DES extraction performance
55
56 was investigated according to the procedure described in the previous paragraph. However, the stirring
57
58
59
60 ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 8 of 28

1
2
3 step was carried out for 1 h followed by settling for 1 h at 298.2 K. The extraction efficiency was reported
4
5 for 9 extraction stages. Furthermore, the repetitive use of DES was reported for 6 cycles. Here, the same
6
DES was used for the extraction of fresh feed in every cycle. It should be mentioned that all the
7
8 experiments were done in duplicates and average STD was 0.8% and max STD is 4%.
9
10 Finally, the DES was regenerated under vacuum using a rotary evaporator (B▪U▪CHI, Rotavapor
11
12 R-215). The pressure was controlled at ~ 35 mbar using a vacuum controller (B▪U▪CHI, Vacuum
13
controller V-850). Also, the sample temperature was kept at 303.2 K in a B▪U▪CHI heating bath B-491.
14
15
16
3 Results and Discussion
17
18 3.1 DES Characterization
19
20 The water content of MTTPBr:TEG was measured and found to be 0.355wt% (= 3.561 mol%).
21
22
Also, since the formation of a DES leads to depression of the freezing point of the resultant liquid, it is
23 very crucial to report the freezing point of the DES. The freezing point of the pure MTPPBr, TEG, and
24
25 the MTPPBr:TEG can be found in Table 3. As it can be seen from Table 3, MTPPBr:TEG exhibits lower
26 freezing point compared to its constituents. Nonetheless, it could be observed that the freezing point
27
28 depression is only 10K which might not qualify the to be coined as “deep” eutectic solvent, however, the
29
solvent “MTPPBr:TEG” was adopted from previous literature.42 It should be mentioned that in many
30
31 studies the terminology is used for granted and further studies about the phase behavior of the DES are of
32
33 utmost importance.
34
35 Table 3: Freezing points of MTPPBr:TEG, MTPPBr, and TEG
36 Compound Freezing point
37
38 MTPPBr:TEG 1:4 254.3 ± 1 K42
39
40 MTPPBr 503 K
41 TEG 266 K
42
43
44
45 Furthermore, the thermal degradation temperature of MTPPBr:TEG was measured and found to
46
47
be 491.5 K. The degradation temperatures determined via dynamic TGA should be taken with coution. It
48 has been experimentally proven by Delgado-Mellado et. al. 43 that the onset temperatures determined with
49
50 this method are generally overestimated. However, in the present study the extraction experiment was
51 conducted at 298.2 K that is far below the onset temperature of the DES. The degradation thermograph of
52
53 DES is shown in Figure 1. As can be observed the DES exhibits two-step degradation. This is attributed
54
55
to the decomposition of one DES constituent prior to the other. According to a study by Ghaedi et. al.44,
56 the polyols “TEG” decompose earlier compared to salts. Moreover, looking to the degradation
57
58
59
60 ACS Paragon Plus Environment
Page 9 of 28 Industrial & Engineering Chemistry Research

1
2
3 thermograph for TEG by the same group,44 it could be concluded that the DES is much more thermally
4
5 stable compared to the pure TEG.
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 Figure 1: Thermal degradation of MTPPBr:TEG measured at a heating rate of 10 K/min from 298.2 K to 723 K
32
33
34
35
36 The DES viscosity at 298.2 K was found to be 138.7 ± 0.5 mPa.s. Also, the temperature
37
38
dependence (293.2 ≤ T ≤ 368.2) of density, ρ for MTPPBr:TEG is shown in Figure 2 (the numerical data
39 are provided in the Supporting Information, Table S.1). Over this temperature range, the specific densities
40
41 are well described by linear (R2 = 0.9999) fits as follows:
42
43 𝝆 [𝒈 𝒄𝒎𝟑] = ―𝟎.𝟎𝟎𝟎𝟕 𝑻 + 𝟏.𝟒𝟎𝟓𝟐 Eq. (1)
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 10 of 28

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23 Figure 2: Density of MTPPBr:TEG in the temperature range of 293.2 K to 368.2 K
24
25
26 3.2 Solubility Data
27 The measured binary solubility for each component is listed in Table 4. Both systems of {thiophene
28
29 + TEG or MTPPBr:TEG (1:4)} and {quinoline + TEG or MTPPBr:TEG (1:4)} showed no phase
30
separation and no turbidity which indicates the high/full solubility of thiophene and quinoline in pure
31
32 TEG and MTPPBr:TEG. On the other hand, the binary systems of {n-heptane + TEG or MTPPBr:TEG
33
34 (1:4)} and {toluene + TEG or MTPPBr:TEG (1:4)} exhibited partial solubility which can be explained by
35 the low polarity of n-heptane and toluene. The solubility of n-heptane in MTPPBr:TEG (1:4) was almost
36
37 half of the that in pure TEG; thus the addition of MTPPBr to TEG is expected to improve the solvent’s
38 recoverability of n-alkane (the raffinate phase) as there will be less/no loss of n-alkane to the DES phase,
39
40 increasing the profitability of the extraction process. However, it can also be observed that the addition of
41
MTPPBr to TEG reduced the solubility of toluene compared to pure TEG, and this can be attributed to the
42
43 increased polarity of the solvent upon the addition of the salt. Based on the results obtained, the selected
44
45 DES could be considered as a potential solvent for simultaneous desulfurization, denitrogenation, and
46 dearomatization.
47
48
Table 4: Solubilities in “weight fractions wi” measured at 298.2 K and 1.01 bar of the solvent-rich phase for the
49
50 systems {solute (1) +solvent (2)}, where the DES (in mole fraction) consists of methyltriphenylphosphonium
51 bromide (xMTPPBr = 0.1927), triethyleneglycol (xTEG = 0.7707), and water (xH2O = 0.0365)a
52
53 System w1 in MTPPBr:TEG (1:4) w1 in TEG
54 {thiophene (1) + solvent (2)} No phase separation observed No phase separation observed
55 {quinoline (1) + solvent (2)} No phase separation observed No phase separation observed
56
57
58
59
60 ACS Paragon Plus Environment
Page 11 of 28 Industrial & Engineering Chemistry Research

1
2
3 {n-heptane (1) + solvent (2)} 0.33 ± 0.01% 0.68 ± 0.01%
4
5 {toluene (1) + solvent (2)} 13.76 ± 0.14% 22.71 ± 0.10%
aStandard uncertainty in temperature u (T) = ± 0.1K, u(p) = 0.04 bar, and for the DES (in mole fraction) u(xTEG) = u(xMTTPBr) =
6
7 u(xH2O) = 0.0002.
8
9
10 3.3 LLE Data
11
12 The LLE data for the pseudo-ternary systems {n-heptane (1) + thiophene/quinoline/toluene (2) +
13
14 MTPPBr:TEG (3)} were determined experimentally via single-stage liquid-liquid extraction at
15 temperature of 298.2 K and pressure of 1.01 bar. This was done in order to understand the phase behavior
16
17 of the selected DES with each oil impurity.
18
19 As observed from Figure 3: (i) the immiscibility window decreased in the following order: toluene
20
21 > thiophene > quinoline, (ii) the slope was found to be positive for quinoline which indicates that only
22 small amount of solvent is required for high extraction. While, negative slopes were observed when
23
24 extracting thiophene and toluene from n-heptane indicating that more solvent is needed to achieve
25 separation. (iii) also, the n-alkane-rich phase was found to be free of DES (w3 =1-w1-w2 = 0.000) implying
26
27 that no further solvent recovery column is needed after the extraction process, which will help to reduce
28
29
the operational cost of the extraction process. This result was confirmed by FTIR analysis shown in
30 Figure S.1 (available in the Supporting Information). From Figure S.2, the spectra of the fresh oil and the
31
32 extracted oil looked similar. It could be deduced that the DES is absent from the raffinate phase and
33 stayed intact in the Extract phase. This results also validuate our assumption that the DES behaves as
34
35 pseudo-pure compound and the {n-heptane + impurity + DES} is a pseudo-ternary system.
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 Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 12 of 28

1
2
3
4 (a) (b)
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23 (c)
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42 Figure 3: Experimental tie lines (●, solid line) in weight fractions for the pseudo-ternary systems {n-heptane +
43 thiophene + MTPPBr:TEG}, {n-heptane + toluene + MTPPBr:TEG}, and {n-heptane + quinoline + MTPPBr:TEG}
44 measured at 298.2 K and 1.01 Bar. The experimental initial in weight fractions compositions are shown as: (●) for
45 thiophene, (■) for toluene, and (▲) for quinoline. The red lines correspond to 2% thiophene, 20% toluene, and 2%
46 quinoline
47
48
49
The distribution ratio (β), selectivity (S) and extraction efficiency (E) were calculated for the
50
51 purpose of evaluating the performance of each selected DES.The numerical values of the equilibrium
52
53 compositions along with the calculated values of β, S, and E at a 1:1 solvent to feed ratio for the three
54 LLE systems can be found in the Supporting Information, Table S.2.
55
56
57
58
59
60 ACS Paragon Plus Environment
Page 13 of 28 Industrial & Engineering Chemistry Research

1
2
3 The distribution ratio (β) can be defined as the ratio of the weight fraction of the solute in the
4
5 extract phase (DES-rich phase) to its weight fraction in the raffinate phase (n-alkane-rich phase). It can be
6
calculated from the experimental LLE data as follows:
7
8
9
𝒘𝟐,𝑬 Eq. (2)
𝜷=
10 𝒘𝟐,𝑹
11
12
13 where 𝑤2,𝐸 refers to the mass fraction of thiophene/quinoline/toluene in the extract phase and 𝑤2,𝑅 counts
14 for the mass fraction of thiophene/quinoline/toluene in the raffinate phase. The higher the distribution
15
16 ratio, the less the solvent-to-feed ratio that is required to achieve high extraction efficiency. Figure S.3
17
18
shows the distribution ratio of thiophene, quinoline and toluene as a function of weight fraction in the
19 raffinate phase. The distribution ratios for {n-heptane (1) + quinoline (2) + MTPPBr:TEG (3)} ranged
20
21 between 9.5 to 4.1, which are much higher distribution ratios compared to thiophene and toluene systems.
22 For the {n-heptane (1) + thiophene (2) + MTPPBr:TEG (3)} system, the distribution ratio was found to be
23
24 less than unity for all compositions. This was also clear from Figure 3, where all the tie lines showed
25
negative slopes. The highest β value was found to be 0.64 and the lowest was found to be 0.40. Moreover,
26
27 the lowest β values were observed for the toluene system ranging from 0.12 to 0.11, which similarly
28
29 showed a negative slope for all its tie lines. (see Figure 3).
30
31
32
33 The selectivity of thiophene/quinoline/ toluene over n-heptane can be calculated by the separation
34
35 factor which is defined as:
36
37 𝜷𝟐 Eq. (3)
38 𝑺=
39 𝜷𝟏
40
41
42 where 𝛽2 is the distribution ratio of thiophene/quinoline/ toluene and 𝛽1 is the distribution ratio of n-
43
44 heptane. The calculated selectivity values of thiophene/quinoline/ toluene are presented in Figure S.4. All
45
the calculated selectivity values were found to be greater than unity (S > 1) thereby implying that
46
47 extraction using MTPPBr:TEG is feasible. High selectivity values are favorable as it means that smaller
48
49 equipment size (i.e., less equilibrium stages) is enough for the targeted separation.
50
51 Both {n-heptane (1) + quinoline (2) + MTPPBr:TEG (3)} and {n-heptane (1) + thiophene (2) +
52 MTPPBr:TEG (3)} systems showed a sharp decrease in the selectivity (from 2327.9 to 235.8 and from
53
54 161.4 to 20.8, respectively) as the concentration of quinoline and thiophene in the raffinate increases.
55
56
However, the selectivity of the {n-heptane (1) + toluene (2) + MTPPBr:TEG (3)} system showed less
57
58
59
60 ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 14 of 28

1
2
3 steep decrease as the concentration of toluene in the raffinate increases, ranging between 30.6 and 12.1.
4
5 Figures S.3 and S.4 5 also show that the observed quinoline’s weight fractions in the raffinate were much
6
smaller (because of the much higher selectivities and distribution ratios), only ranging between 0.002 and
7
8 0.096, while thiophene’s and toluene’s weight fractions were ranging between 0.014 – 0.581 and 0.092 –
9
10 0.773, respectively. This behavior corresponds to the decrease in the immiscibility window and the higher
11 extraction of quinoline compared to thiophene and toluene.
12
13
The following equation was applied to calculate extraction efficiency (E):
14
15
16
𝒘𝒊 ― 𝒘𝒇 Eq. (4)
𝑬 (%) =
17 𝒘𝒊
18
19 where 𝑤𝑖 is the initial weight fraction of the oil impurity in fuel and 𝑤𝑓 is the final weight fraction in the
20
21 raffinate phase after extraction with the DES.
22
23 The calculated extraction efficiency for each tie line is shown in Figure 4. It can be observed from
24
25 Figure S.5 that increasing the initial concentration gradually decreases the extraction efficiency of all
26
impurities. For single extraction stage, the highest efficiencies were found for quinoline (89.9- 86.3%)
27
28 followed by thiophene and toluene efficiencies ranging between (32.7%- 16.7%) and (9.5- 3.2%)
29
30
respectively.
31
32 3.4 Single Stage Liquid-Liquid Extraction from Oil Mixture
33
As mentioned in section 2.4, in those experiments the starting oil mixture was composed of {20%
34
35 toluene + 2% thiophene + 2% quinoline + 76% n-heptane} on mass basis. The three oil impurities, i.e.
36
37 toluene, thiophene and quinoline, were extracted using the pure TEG as well as the DES. The extraction
38 results are expressed as extraction efficiency E calculated as in Eq. 4. As it can be observed from Figure
39
40 4, the MTPPBr:TEG showed about ~ 5% higher extraction efficiency for the thiophene and quinoline
41 compared to TEG. Despite the fact that thiophene and quinoline were found to be fully soluble in both
42
43 MTPPBr:TEG and TEG (see Table 4), the higher extraction capacity of the DES could be attributed to the
44
electrostatic interactions (i.e: π-π) between the thiophene and quinoline and the salt-part ‘MTPPBr’ of the
45
46 DES. Moreover, the existence of the sulfur ‘S in the thiophene’ and the electronegative nitrogen ‘N in the
47
48 quinoline’ increased their reactivity. This justifies the DES usage in comparison to the pure solvent.
49 Nevertheless, the opposite phenomena was observed for the extraction of toluene. Where the extraction
50
51 efficiency of toluene using TEG was found to be almost double the extraction efficiency of the DES.
52
This result was expected based on the binary solubility data (see Table 4). Presumably, the solvation of
53
54 toluene in the polyol were the dominant interactions. This finding also indicates that the MTPPBr:TEG is
55
56 not suitable for the extraction of toluene. And since toluene was the least extracted it’s considered as the
57
58
59
60 ACS Paragon Plus Environment
Page 15 of 28 Industrial & Engineering Chemistry Research

1
2
3 “limiting compound” for this study case. Therefore, it could be said that TEG is a better solvent for this
4
5 separation. Nevertheless, other advantages of the usage of DES such as good solvent recovery will be
6
verified in later paraghaps. It should be mentioned that the solvent selection is a matter of process
7
8 optimization that include several factors.
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 Figure 4: Single-stage extraction efficiency. The gray column is for the pure TEG and the white column is for the
35
36 MTPPBr:TEG. (Conditions: S:F = 1:1, T = 298.2 K, P = 1.01 bar; Stirring time = 2 h, and settling time = 2 h)
37
38 On the contrary, the extraction efficiency of each oil impurity from the n-heptane (see section
39
40
2.4) at the same initial concentration to that of the oil model {20% toluene + 2% thiophene + 2%
41 quinoline + 76% n-heptane} was found to be slightly higher in all cases. The extraction efficiency values
42
43 are listed in Table 5. This implies that the DES extraction efficiency of each component is not
44 independent of mixing effects. This, in fact, indicates that LLE investigations might not be enough for
45
46 drawing conclusions on the DES suitability for oil purification and more realistic oil models should be
47
considered in this field of research.
48
49
50 Table 5: Extraction efficiency of each oil impurity from n-heptane and the extraction efficiency of the oil impurity
51 from the oil model measured at 298.2 K and 1.01 bar*
52
Component wi (%) E (%) in n-heptane E (%) in oil mixture
53
54 Toluene 20 7.5 7.2
55
56
57
58
59
60 ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 16 of 28

1
2
3 Thiophene 2 32.7 28.4
4
5 Quionline 2 89.9 82.6
6
aStandard uncertainty in temperature u(T) = 0.1K, u(p) = 0.04 bar, u(wi) = 0.002 and for the DES (in mole fraction) u(xTEG) =
7
8 u(xMTTPBr) = u(xH2O) = 0.0002.
9
10
11
12 3.5 Parametric Study
13
3.5.1 Effect of Extraction Temperature
14
15 The effect of temperature on the extraction efficiency is shown in Figure 5. It can be seen that the
16
17 increase of extraction temperature influenced each impurity in a different manner. As for the toluene, its
18 extraction efficiency was found to increase by 5% upon the increase of temperature between 298.2 K to
19
20 343.2 K. The extraction efficiency of thiophene showed very slight fluctuations with temperature increase
21
reaching a maximum value of 30% at 313.2 K. On the other hand, quinoline was found be sensitive the
22
23 temperature. The extraction efficiency decreased by the temperature increase. A maximum decrease of
24
25 14% was reported at 338.2 K. However, and due to the possible component evaporation at high
26 temperatures, and to maintain mild operating conditions, the extraction temperature of all the experiments
27
28 was fixed at 298.2 K.
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 Figure 5: Effect of extraction temperature. Symbols: (▲) E for quinoline, (●) E for thiophene, and (■) E for
54 toluene. (Conditions: S:F = 1:1, P = 1.01 bar; stirring time = 1 h and settling time = 1 h)
55
56
57
58
59
60 ACS Paragon Plus Environment
Page 17 of 28 Industrial & Engineering Chemistry Research

1
2
3 3.5.2 Effect of Solvent-to-Feed Ratio
4
5 In this section, the effect of changing the solvent-to-feed mass ratio (S:F) was studied. The S:F
6
mass ratio was varied from 9:1 to 1:9. From Figure 6, it can be observed that the extraction efficiency of
7
8 the three impurities substantially decreases with the decrease of the solvent’s mass, which was expected
9
10 as this is a general trend. Thus, it’s clear that the extraction efficiency is sensitive to variation in the
11 solvent’s mass fraction. Nonetheless, minimizing solvent consumption in any extraction process is indeed
12
13 economically desirable. Therefore, the S:F mass ratio was fixed at 1:1 in all the extraction experiments.
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35 Figure 6: Effect of solvent-to-feed ratio. Symbols: (▲) E for quinoline, (●) E for thiophene, and (■) E for toluene.
36
(Conditions: T = 298.2 K, P = 1.01 bar; stirring time = 1 h and settling time = 1 h)
37
38
39 3.5.3 Effect of Initial Concentration of the Oil Impurity
40
In those experiments, the initial composition of thiophene and quinoline in the oil mixture was
41
42 varied as 4%, 3%, 2%, 1%, and 0.5% (wt.). The composition of toluene was fixed at 20% since only an
43
44 extraction efficiency of 7.2 % could be obtained using this DES. Also, with reference to Figure 6, the
45 extraction efficiency of toluene would become even lower with concentration increase.
46
47 It was found that the DES extraction capability for thiophene and quinoline is not sensitive to
48 their initial concentration “at least for the tested concentrations” as shown in Figure 7. That is to say that
49
50 the removal efficiency is relatively independent of initial concentration. This trend is beneficial from an
51
industrial point of view where fuels of varying sulfur- and nitrogen-aromatic concentrations could exist.
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 18 of 28

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23 Figure 7: Effect of initial concentration. Symbols: (▲) E for quinoline, and (●) E for thiophene. (Conditions: S:F =
24
1:1, T = 298.2 K, P = 1.01 bar; stirring time = 1 h and settling time = 1 h)
25
26
27
28
29 3.5.4 The Number of Extraction Stages
30
31
The extraction of toluene, thiophene, and quinoline using the MTPPBr:TEG in a single extraction
32 stage yielded a reduction of 7% in toluene, 28% thiophene, and 82% quinoline contents (see Table 5).
33
34 However, achieving high extraction efficiencies would require a large amount of solvent as it was shown
35 in section 3.5.3, which will not be economically sound. Instead, a multistage batch extraction process was
36
37 proposed. These experiments were conducted in such a way that fresh solvent was added to the
38
hydrocarbon-rich phase of the previous stage. Nine extraction stages were assessed and the results are
39
40 shown in Figure 8. As can be seen, the extraction efficiency of quinoline reached 100% in only 2 stages,
41
42 while a 100% extraction efficiency of thiophene was achieved at 9 stages. However, an efficiency of 48%
43 was achieved for toluene at 9 stages. In the most ideal case, where the DES ignoring the effect/interaction
44
45 of other compound, while considering that solubility of toluene is 14 wt.% (see Table 4) and the intial
46 composition of toluene is 20 wt.%, the maximum extraction efficiency would be 70%. So then,
47
48 theoretically, let’s target an extraction efficiency of 70% for toluene. The number of extraction stages can
49
50
be estimated as:
51
52
53 𝑬𝒕𝒂𝒓𝒈𝒆𝒕𝒆𝒅 = 𝟏 ― (𝟏 ― 𝑬𝒔𝒕𝒂𝒈𝒆 )𝒏 (Eq. 5)
54
55
56
57
58
59
60 ACS Paragon Plus Environment
Page 19 of 28 Industrial & Engineering Chemistry Research

1
2
3 where Etargeted is the targeted extraction, Estage is the single-stage extraction efficiency, and n is the
4
5 theoretical number of stages. Solving Eq. 5 for n (Etargeted = 70% , and Estage = 7%); the number of
6
extraction stages to achieve 70% extraction efficiency of toluene is ~ 17 stages. However, it was
7
8 impractical to verify that result experimentally, considering an approximately 17 extraction stages are
9
10 required. Still, it could be concluded that the DES “MTPPBr:TEG” was not optimal for the removal of
11 toluene, but quinoline and thiophene can be successfully extracted. Thus, denitrogenation and
12
13 desulfurization are easier than dearomatization with the chosen DES and further research is still needed
14 for a better choice of the DES for the dearomatization of oil fuels.
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 Figure 8: Number of extraction stages. Columns: black for toluene, gray for thiophene, and white for quinoline.
40 (Conditions: S:F = 1:1, T = 298.2 K, P = 1.01 bar; stirring time = 1 h and settling time = 1 h)
41
42
43
3.5.5 The Effect of Repetitive Use of DES
44
45 The repetitive usage of the DES was assessed via performing extraction using the same DES
46
47 added to a fresh feed in every cycle. The DES was applied for six consecutive cycles. As shown in Figure
48 9, the extraction efficiency of each oil impurity decreases with the increase of the extraction cycles.
49
50 Presumably, the solvent loses its extraction capacity due to the presence of the residual oil impurity in the
51
DES phase. In this work the repetitive use of the DES was tested for six cycles, however, it could be
52
53 anticipated that the DES reaches a saturation state in more extraction cycles. Despite the fact that the DES
54
55 loses its extraction capacity in the consecutive cycles, this drawback can be overcome via solvent
56 regeneration, which is highly preferred industrially. This is investigated in the following section.
57
58
59
60 ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 20 of 28

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 Figure 9: Repetitive use of DES. Columns: black for toluene, gray for thiophene, and white for quinoline.
26 (Conditions: S:F = 1:1, T = 298.2 K, P = 1.01 bar; stirring time = 1 h and settling time = 1 h)
27
28
29
30
3.6 DES Regeneration
31
32 For solvent regeneration, a single-stage extraction experiment was done for the oil model {20%
33
34 toluene + 2% thiophene + 2% quinoline + 76% n-heptane}. Thereafter, a sample of the extract phase
35 “DES-rich phase” with a composition {2.4% toluene + 0.7% thiophene + 3.3% quinoline + 0.5% n-
36
37 heptane + 93.1% MTPPBr:TEG} was taken and placed in a vacuum rotary evaporator at a temperature of
38
303.2 K and vacuum pressure of 35 mbar for 12 h. The regenerated DES was then analyzed using the GC.
39
40 The results of a single regeneration cycle showed that the DES was free of toluene, thiophene and
41
42 heptane. However, 3% of the quinoline remained in the DES. This essentially means that the vacuum
43 evaporator was not the ideal regeneration method for a high boiling point impurity such as quinolone
44
45 (Tboiling = 510.3 K). Also, the high extraction efficiency of quinoline indicates the strong interactions with
46 the DES, consequently, the difficultiy of breaking those inteactions for the purpose of regeneration could
47
48 be jusitified and anticipated.
49
50 Furthermore, the regenerated DES {3% quinoline + 97% MTPPBr:TEG} was used for one cycle
51
52 of extraction on a fresh oil model. Figure 10 shows the extraction results. As can be observed the DES
53 efficiency was reduced for all the three impurities due to the accumulated quinoline. Another reason for
54
55 the efficiency reduction could also be due to the loss of some DES constituents via evaporation. The
56
57
58
59
60 ACS Paragon Plus Environment
Page 21 of 28 Industrial & Engineering Chemistry Research

1
2
3 results of this section suggested that further research on economical and efficient DES regeneration
4
5 methods is required.
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29 Figure10 : Effect of solvent regeneration of the DES extraction performance. Columns: black for fresh DES
30 performance, and Gray for generated DES performance. (Conditions: S:F = 1:1, T = 298.2 K, P = 1.01 bar;
31
stirring time = 1 h and settling time = 1 h)
32
33
34
4 Conclusions
35
36 In this work a phosphonium-based DES was assessed for combined dearomatization,
37
38 desulfurization and denitrogenation of oil fuels via the liquid-liquid extraction approach. The selected
39 DES was methyltriphenylphosphonium bromide (MTPPBr) and triethylene glycol (TEG) in molar ratio
40
41 equals (1 (MTPPBr):4 (TEG)). An arbitrary oil model as {20% toluene + 2% thiophene + 2% quinoline +
42
76% n-heptane} was chosen for this study. Initially, the DES was characterized for its water content,
43
44 density, viscosity, and degradation temperature. Next, the solubility of toluene, thiophene, quinoline, and
45
46 n-heptane in the MTPPBr: TEG was determined at 298.2 K and 1.01 bar. The solubilities in the DES
47 were compared to the solubilities in the pure TEG. The results showed that in spite of the workability of
48
49 the pure TEG as a solvent for the intended applications, the addition of the salt “MTPPBr” to form the
50 DES enhanced the extraction performance as well as reduced the product loss (i.e., n-alkane). Other
51
52 aspects, such as the higher thermal stability of the DES over the pure TEG, were also an advantage.
53
54
Furthermore the liquid-liquid equilibrium (LLE) data for the systems {toluene + n-heptane +
55 MTPPBr:TEG}, {thiophene + n-heptane + MTPPBr:TEG} and {quinoline + n-heptane + MTPPBr:TEG}
56
57
58
59
60 ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 22 of 28

1
2
3 were reported at 298.2 K and 1.01 bar. The calculated distribution coefficients and selectivities showed a
4
5 remarkable ability of the DES as extractant for quinoline; however, a lower performance for thiophene
6
and especially toluene was observed. Interestingly, the raffinate phase was found to free of the DES,
7
8 which will facilitate both the solvent and the n-alkanes recoveries.
9
10 In addition, single-stage liquid-liquid extraction experiments for combined removel of the
11 impurities “toluene, thiophene and quinoline” from the oil model using the DES and the pure TEG were
12
13 conducted at 298.2 K and 1.01 bar. The results showed better extraction ability of the DES compared to
14 the pure TEG for both the quinoline and the thiophene, but not for the toluene.
15
16 The influence of the extraction time and the extraction temperature were also investigated. It was
17
found that 1 h of stirring is enough to reach an equilibrium of the interactions between the oil impurities
18
19 and the MTPPBr:TEG, while a temperature of 298.2 K was found to be optimal in consideration of the
20
21 extraction efficiency and the application of mild operating conditions. It was also found the extraction
22 efficiency of the three impurities is very sensitive to the variation in the solvent-to-feed ratio (S:F), but
23
24 independent of the initial concentrations of quinoline and thiophene. Furthermore, the extraction
25
efficiency of quinoline reached 100% in only 2 stages, while a 100% extraction efficiency of thiophene
26
27 and 48 % for toluene was achieved at 9 stages. Thus, full denitrogenation and desulfuziation were easier
28
29 to achieve than full dearomatization with the chosen DES. It was found that the DES gradually loses its
30 extraction ability if used without regeneration. For that reason, we also studied the DES regeneration via
31
32 vacuum evaporation. The results showed that the vacuum evaporation was successful for fully removing
33 n-heptane, toluene, and thiophene, but not ful removal of quinoline from the DES. Further research is still
34
35 utmost needed for finding a DES that could work better for the extraction of toluene, and ultimately better
36
for the overall objective of combined dearomatization, desulfurization, and denitrogenation of oil fuels.
37
38 Moreover, efficient regeneration techniques for the recovery of the DES from a range of different species
39
40 with a different nature should be studied. However, the large degree of freedom for designing an
41 objective-oriented DES opens a wide window for the optimization of DES for oil purification.
42
43
44 Supporting Information
45
46 (i) Effect of the stirring time, (ii) FTIR analysis of a fresh oil mixture sample and an extracted oil mixture
47
"raffinate", (iii) Numerical values of the DES density, (iv) Experimental LLE data for the pseudo-ternary
48
49 systems {n-heptane + thiophene + MTPPBr:TEG}, {n-heptane + quinoline + MTPPBr:TEG}, and{n-
50
51 heptane + toluene + MTPPBr:TEG}, and (v) Figures describing the distribution coefficients, selectivity
52 and the extraction efficiency based on the LLE data. This information is available free of charge via the
53
54 Internet at http://pubs.acs.org/
55
56
57
58
59
60 ACS Paragon Plus Environment
Page 23 of 28 Industrial & Engineering Chemistry Research

1
2
3 5 Author Information
4
5 Corresponding Author: Inas M. Nashef, enas.nashef@ku.ac.ae
6
7
8
6 Notes
9 The authors declare no competing financial
10
11
12 7 Acknowledgments
13
14 The authors gratefully acknowledge the financial support of Khalifa University of Science and
15
Technology (award No. CIRA-2018-023) and the support of Center for Membrane and Advanced Water
16
17 Technology (CMAT).
18
19
20 8 References
21
22 (1) Barker, C. Origin, Composition and Properties of Petroleum. Dev. Pet. Sci. 1985, 17, 11.
23 https://doi.org/10.1016/S0376-7361(08)70564-8.
24
25
26 (2) EPA, U. S. Control of Air Pollution From Motor Vehicles: Tier 3 Motor Vehicle Emission and
27 Fuel Standards. Fed. Regist. 2014, 79 , 23414. https://doi.org/10.2753/RSH1061-1983310140.
28
29
(3) Prado, G. H. C.; Rao, Y.; De Klerk, A. Nitrogen Removal from Oil: A Review. Energy and Fuels
30
31 2017, 31 , 14. https://doi.org/10.1021/acs.energyfuels.6b02779.
32
33 (4) Orr, W. L. Sulfur in Heavy Oils, Oil Sands and Oil Shales. Oil sand oil shale Chem. New York,
34
35 Verlag Chemie 1978, 223.
36
37 (5) Stanislaus, A.; Barry, H. C. Aromatic Hydrogenation Catalysis: A Review. Catal. Rev. 1994, 36,
38
39 75. https://doi.org/10.1080/01614949408013921.
40
41 (6) Sharma, M.; Sharma, P.; Kim, J. N. Solvent Extraction of Aromatic Components from Petroleum
42
Derived Fuels: A Perspective Review. RSC Adv. 2013, 3 , 10103.
43
44 https://doi.org/10.1039/c3ra00145h.
45
46 (7) Oballa, M. C.; Shih, S. S. Catalytic Hydroprocessing of Petroleum and Distillates; Marcel Dekker,
47
48 Inc.: New york, 1994.
49
50 (8) Bacha, J.; Freel, J.; Gibbs, A.; Gibbs, L.; Hemighaus, G.; Hoekman, K.; Horn, J.; Ingham, M.;
51
52 Jossens, L.; Kohler, D.; Lesnini, D.; McGeehan, J.; Nikanjam, M.; Olsen, E.; Organ, R.; Scott, B.;
53 Sztenderowicz, M.; Tiedemann, A.; Walker, C.; Lind, J.; Jones, J.; Scott, D.; Mills, J. Diesel Fuels
54
55 Technical Review. Chevron Glob. Mark. 2007, 1. https://doi.org/10.1063/1.3575169.
56
57
58
59
60 ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 24 of 28

1
2
3 (9) Dwamena, A. K. Recent Advances in Hydrophobic Deep Eutectic Solvents for Extraction.
4
5 Separations 2019, 6, 9. https://doi.org/10.3390/separations6010009.
6
7 (10) Zoretić, Z.; Adžamić, T.; Sertić Bionda, K. Desulfurization of FCC Gasoline by Extraction with
8
9 Sulfolane and Furfural. Nafta 2009, 60 , 485. https://hrcak.srce.hr/41444
10
11 (11) Shen, H.; Shen, B.; Ling, H. Desulfurization of Fluid Catalytic Cracking Gasoline by Extractive
12
Distillation Coupled with Hydrodesulfurization of Heavy Fraction. Energy and Fuels 2013, 27,
13
14 5153. https://doi.org/10.1021/ef401075x.
15
16 (12) Walter G. Method of Desulfurization of Hydrocarbon. US Patent 6274785 B1, 2001.
17
18
19 (13) Kianpour, E.; Azizian, S. Polyethylene Glycol as a Green Solvent for Effective Extractive
20 Desulfurization of Liquid Fuel at Ambient Conditions. Fuel 2014, 137, 36.
21
22 https://doi.org/10.1016/j.fuel.2014.07.096.
23
24 (14) Kulkarni, P. S.; Afonso, C. A. M. Deep Desulfurization of Diesel Fuel Using Ionic Liquids:
25
26 Current Status and Future Challenges. Green Chem. 2010, 12 , 1139.
27 https://doi.org/10.1039/c002113j.
28
29
(15) Abro, R.; Abdeltawab, A. A.; Al-Deyab, S. S.; Yu, G.; Qazi, A. B.; Gao, S.; Chen, X. A Review of
30
31 Extractive Desulfurization of Fuel Oils Using Ionic Liquids. RSC Adv. 2014, 4 , 35302.
32
33 https://doi.org/10.1039/C4RA03478C.
34
35 (16) Jess, A.; Eßer, J. Deep Desulfurization of Fuels by Extraction with Ionic Liquids. ACS Symp. Ser.
36
2005, 902, 83. https://doi.org/10.1021/bk-2005-0902.ch007.
37
38
39 (17) Ibrahim, M. H.; Hayyan, M.; Hashim, M. A.; Hayyan, A. The Role of Ionic Liquids in
40 Desulfurization of Fuels: A Review. Renew. Sustain. Energy Rev. 2015, 76 (November 2016),
41
42 1534. https://doi.org/10.1016/j.rser.2016.11.194.
43
44 (18) Larriba, M.; Delgado-Mellado, N.; Navarro, P.; Alcover, R.; Moya, C.; Palomar, J.; García, J.;
45
46 Rodríguez, F. Novel Process to Reduce Benzene, Thiophene, and Pyrrole in Gasoline Based on
47 [4bmpy][TCM] Ionic Liquid. Energy and Fuels 2018, 32 , 5650.
48
49 https://doi.org/10.1021/acs.energyfuels.8b00529.
50
51 (19) Navarro, P.; Larriba, M.; Delgado-Mellado, N.; Sánchez-Migallón, P.; García, J.; Rodríguez, F.
52
53 Extraction and Recovery Process to Selectively Separate Aromatics from Naphtha Feed to
54 Ethylene Crackers Using 1-Ethyl-3-Methylimidazolium Thiocyanate Ionic Liquid. Chem. Eng.
55
56 Res. Des. 2017, 120, 102. https://doi.org/10.1016/j.cherd.2017.02.008.
57
58
59
60 ACS Paragon Plus Environment
Page 25 of 28 Industrial & Engineering Chemistry Research

1
2
3 (20) Larriba, M.; Navarro, P.; González, E. J.; García, J.; Rodríguez, F. Dearomatization of Pyrolysis
4
5 Gasolines from Mild and Severe Cracking by Liquid-Liquid Extraction Using a Binary Mixture of
6
[4empy][Tf2N] and [Emim][DCA] Ionic Liquids. Fuel Process. Technol. 2015, 137, 269.
7
8 https://doi.org/10.1016/j.fuproc.2015.03.009.
9
10 (21) Larriba, M.; Navarro, P.; García, J.; Rodríguez, F. Liquid-Liquid Extraction of BTEX from
11
12 Reformer Gasoline Using Binary Mixtures of [4empy][Tf2N] and [Emim][DCA] Ionic Liquids.
13
Energy and Fuels 2014, 28 , 6666. https://doi.org/10.1021/ef501671d.
14
15
16 (22) Peric, B.; Sierra, J.; Martí, E.; Cruañas, R.; Garau, M. A.; Arning, J.; Bottin-Weber, U.; Stolte, S.
17 (Eco)Toxicity and Biodegradability of Selected Protic and Aprotic Ionic Liquids. J. Hazard.
18
19 Mater. 2013, 261, 99. https://doi.org/10.1016/j.jhazmat.2013.06.070.
20
21 (23) Quijano, G.; Couvert, A.; Amrane, A.; Darracq, G.; Couriol, C.; Le Cloirec, P.; Paquin, L.; Carrié,
22
23 D. Toxicity and Biodegradability of Ionic Liquids: New Perspectives towards Whole-Cell
24 Biotechnological Applications. Chem. Eng. J. 2011, 174 , 27.
25
26 https://doi.org/10.1016/j.cej.2011.07.055.
27
28 (24) Romero, A.; Santos, A.; Tojo, J.; Rodríguez, A. Toxicity and Biodegradability of Imidazolium
29
30 Ionic Liquids. J. Hazard. Mater. 2008, 151 , 268. https://doi.org/10.1016/j.jhazmat.2007.10.079.
31
32 (25) Abbott, A. P.; Capper, G.; Davies, D. L.; Rasheed, R. K.; Tambyrajah, V. Novel Solvent
33
Properties of Choline Chloride/Urea Mixtures. Chem. Commun. (Camb). 2003, No. 1, 70.
34
35 https://doi.org/10.1039/b210714g.
36
37 (26) Garcia, G.; Aparicio, S.; Ullah, R.; Atilhan, M. Deep Eutectic Solvents: Physicochemical
38
39 Properties and Gas Separation Applications. Energy & Fuels 2015, 29, 2616.
40
https://doi.org/10.1021/ef5028873.
41
42
43 (27) Zhang, Q.; De Oliveira Vigier, K.; Royer, S.; Jérôme, F. Deep Eutectic Solvents: Syntheses,
44 Properties and Applications. Chem. Soc. Rev. 2012, 41 , 7108. https://doi.org/10.1039/c2cs35178a.
45
46
47
(28) Francisco, M.; Van Den Bruinhorst, A.; Kroon, M. C. Low-Transition-Temperature Mixtures
48 (LTTMs): A New Generation of Designer Solvents. Angew. Chemie - Int. Ed. 2013, 52 , 3074.
49
50 https://doi.org/10.1002/anie.201207548.
51
52 (29) Smith, E. L.; Abbott, A. P.; Ryder, K. S. Deep Eutectic Solvents (DESs) and Their Applications.
53
54
Chem. Rev. 2014, 114 , 11060. https://doi.org/10.1021/cr300162p.
55
56 (30) Warrag, S. E. E.; Peters, C. J.; Kroon, M. C. Deep Eutectic Solvents for Highly Efficient
57
58
59
60 ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 26 of 28

1
2
3 Separations in Oil and Gas Industries. Curr. Opin. Green Sustain. Chem. 2017, 5, 55.
4
5 https://doi.org/10.1016/j.cogsc.2017.03.013.
6
7 (31) Gonzalez, A. S. B.; Francisco, M.; Jimeno, G.; De Dios, S. L. G.; Kroon, M. C. Liquid-Liquid
8
9 Equilibrium Data for the Systems {LTTM+benzene+hexane} and {LTTM+ethyl Acetate+hexane}
10 at Different Temperatures and Atmospheric Pressure. Fluid Phase Equilib. 2013, 360, 54.
11
12 https://doi.org/10.1016/j.fluid.2013.09.010.
13
14 (32) Rogošić, M.; Kučan, K. Z. Deep Eutectic Solvents Based on Choline Chloride and Ethylene
15
16 Glycol as Media for Extractive Denitrification/Desulfurization/Dearomatization of Motor Fuels. J.
17 Ind. Eng. Chem. 2019, 72, 87. https://doi.org/10.1016/j.jiec.2018.12.006.
18
19
(33) Zagajski Kučan, K.; Rogošić, M. Purification of Motor Fuels by Means of Extraction Using Deep
20
21 Eutectic Solvent Based on Choline Chloride and Glycerol. J. Chem. Technol. Biotechnol. 2019, 94
22
23 (4), 1282. https://doi.org/10.1002/jctb.5885.
24
25 (34) Kučan, K. Z.; Perković, M.; Cmrk, K.; Načinović, D.; Rogošić, M. Betaine + (Glycerol or
26
Ethylene Glycol or Propylene Glycol) Deep Eutectic Solvents for Extractive Purification of
27
28 Gasoline. ChemistrySelect 2018, 3 , 12582. https://doi.org/10.1002/slct.201803251.
29
30 (35) Larriba, M.; Ayuso, M.; Navarro, P.; Delgado-Mellado, N.; Gonzalez-Miquel, M.; García, J.;
31
32 Rodríguez, F. Choline Chloride-Based Deep Eutectic Solvents in the Dearomatization of
33
Gasolines. ACS Sustain. Chem. Eng. 2018, 6 , 1039.
34
35 https://doi.org/10.1021/acssuschemeng.7b03362.
36
37 (36) Kareem, M. A.; Mjalli, F. S.; Hashim, M. A.; AlNashef, I. M. Liquid-Liquid Equilibria for the
38
39 Ternary System (Phosphonium Based Deep Eutectic Solvent-Benzene-Hexane) at Different
40
Temperatures: A New Solvent Introduced. Fluid Phase Equilib. 2012, 314, 52.
41
42 https://doi.org/10.1016/j.fluid.2011.10.024.
43
44 (37) Kareem, M. A.; Mjalli, F. S.; Hashim, M. A.; Hadj-Kali, M. K. O.; Bagh, F. S. G.; Alnashef, I. M.
45
46 Phase Equilibria of Toluene/Heptane with Tetrabutylphosphonium Bromide Based Deep Eutectic
47
Solvents for the Potential Use in the Separation of Aromatics from Naphtha. Fluid Phase Equilib.
48
49 2012, 333, 47. https://doi.org/10.1016/j.fluid.2012.07.020.
50
51 (38) Hizaddin, H. F.; Hadj-Kali, M. K.; Ramalingam, A.; Ali Hashim, M. Extractive Denitrogenation
52
53 of Diesel Fuel Using Ammonium- and Phosphonium-Based Deep Eutectic Solvents. J. Chem.
54
Thermodyn. 2016, 95, 164. https://doi.org/10.1016/j.jct.2015.12.009.
55
56
57
58
59
60 ACS Paragon Plus Environment
Page 27 of 28 Industrial & Engineering Chemistry Research

1
2
3 (39) Warrag, S. E. E.; Adeyemi, I.; Rodriguez, N. R.; Nashef, I. M.; Van Sint Annaland, M.; Kroon, M.
4
5 C.; Peters, C. J. Effect of the Type of Ammonium Salt on the Extractive Desulfurization of Fuels
6
Using Deep Eutectic Solvents. J. Chem. Eng. Data 2018, 63 , 1088.
7
8 https://doi.org/10.1021/acs.jced.7b00832.
9
10 (40) Warrag, S. E. E.; Alli, R. D.; Kroon, M. C. Liquid–Liquid Equilibrium Measurements for the
11
12 Extraction of Pyridine and Benzothiazole from n -Alkanes Using Deep Eutectic Solvents. J. Chem.
13
Eng. Data 2019, 64, 4882. https://doi.org/10.1021/acs.jced.9b00413.
14
15
16 (41) Hadj-Kali, M. K.; Mulyono, S.; Hizaddin, H. F.; Wazeer, I.; El-Blidi, L.; Ali, E.; Hashim, M. A.;
17 AlNashef, I. M. Removal of Thiophene from Mixtures with n -Heptane by Selective Extraction
18
19 Using Deep Eutectic Solvents. Ind. Eng. Chem. Res. 2016, 55 , 8415.
20
https://doi.org/10.1021/acs.iecr.6b01654.
21
22
23 (42) Shahbaz, K.; Mjalli, F. S.; Hashim, M. A.; AlNashef, I. M. Using Deep Eutectic Solvents Based
24 on Methyl Triphenyl Phosphunium Bromide for the Removal of Glycerol from Palm-Oil-Based
25
26 Biodiesel. Energy and Fuels 2011, 25 , 2671. https://doi.org/10.1021/ef2004943.
27
28 (43) Delgado-Mellado, N.; Larriba, M.; Navarro, P.; Rigual, V.; Ayuso, M.; García, J.; Rodríguez, F.
29
30 Thermal Stability of Choline Chloride Deep Eutectic Solvents by TGA/FTIR-ATR Analysis. J.
31 Mol. Liq. 2018, 260, 37. https://doi.org/10.1016/j.molliq.2018.03.076.
32
33
(44) Ghaedi, H.; Ayoub, M.; Sufian, S.; Lal, B.; Uemura, Y. Thermal Stability and FT-IR Analysis of
34
35 Phosphonium-Based Deep Eutectic Solvents with Different Hydrogen Bond Donors. J. Mol. Liq.
36
37 2017, 242, 395. https://doi.org/10.1016/j.molliq.2017.07.016.
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 28 of 28

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 Paragon Plus Environment

You might also like