Sunteți pe pagina 1din 12

RSC Advances

PAPER

Efficient non-catalytic oxidative and extractive


desulfurization of liquid fuels using ionic liquids†
Cite this: RSC Adv., 2016, 6, 103606
Omar U. Ahmed,a Farouq S. Mjalli,*a Talal Al-Wahaibi,a Yahya Al-Wahaibia
and Inas M. AlNashefb

Oxidative desulfurization (ODS) is one of the promising alternative and heavily researched desulfurization
technologies. This is partly due to its ability to preferentially oxidize and ease the removal of refractory
sulfur compounds with the aid of a suitable solvent. Despite its long list of advantages, challenges in
different research areas within ODS technology still exist. In this work, an effort was made to bridge the
gap that exists in terms of the selection of suitable oxidant and strategy. A preliminary kinetic modeling
of the experimental data showed that the non-catalytic conversion of dibenzothiophene (DBT) and
benzothiophene (BT) to their corresponding sulfones using the electrophilic meta-chloroperoxybenzoic
acid (mCPBA) can be considered a bimolecular and a trimolecular reaction respectively. Using an ionic
liquid (IL) as an extraction solvent in a simultaneous oxidation and extraction setup (EODS), >78% BT was
removed at optimum experimental conditions. Using the post-oxidation extractive desulfurization setup
(OEDS), 99% removal of BT was achieved at milder optimum experimental conditions. Also using the
Received 2nd September 2016
Accepted 21st October 2016
OEDS strategy, >99% of DBT removal was achieved after only 15 min at 60  C, with a mass fraction of
$0.5 and O/S of 3/1 for all the ILs tested. Finally, the sulfur content of a commercial diesel fuel was
DOI: 10.1039/c6ra22032k
reduced to 15.6 ppm using the OEDS strategy with tetrabutylphosphonium methanesulfonate as
www.rsc.org/advances a suitable extraction solvent, which can be readily regenerated.

crude oil products in 2010 and is expected to reach 33.3% by


1 Introduction 2035,5 has been a subject of these international regulations. For
The presence of sulfur compounds in petroleum products is of instance, the maximum allowable sulfur content in diesel fuel
great concern to the processing industries and governments was pegged at 10 ppm and 15 ppm in Europe and US respec-
around the world. In a petroleum renery, certain sulfur tively.6 Hydrodesulfurization (HDS) is conventionally used for
compounds originating from crude oil react with metal surfaces efficient removal of sulfur compounds. However, owing to high
at temperatures exceeding 260  C to form metal suldes and concentrations of a class of sulfur compounds (refractory sulfur
other products. This is known as suldic corrosion1 and it compounds) present in the heavier and sourer crude oils
occurs on process equipment. Furthermore, sulfur compounds produced in recent years, the efficiency of the once reliable HDS
present in certain reactor feed streams in the renery may lead has declined. This is especially true in the case of deep desul-
to catalyst poisoning.2 Oxides of sulfur (SOx) generated during furization.4 For this reason, the potentials of other desulfur-
combustion of sulfur-containing hydrocarbon fuels not only ization techniques have been under intense investigation.
lead to environmental and health hazards but also reduce the Adsorptive,7–9 extractive,4,10,11 oxidative4,12,13 and bio-desul-
efficiency of the advanced emission control devices installed in furization14,15 are among the promising desulfurization tech-
modern cars. Increased particulate matter, CO2 and NOx emis- niques that have received a great deal of attention from the
sions are some of the consequences of the reduction in the scientic community. Extractive desulfurization (EDS) using
efficiency of these devices.3,4 As a result, stringent regulations ionic liquids (ILs) has been considered promising in terms of
on the maximum level of sulfur allowable in transportation both environmental and technical efficiency. This is because it
fuels have been imposed to varying degrees by governments in can be carried out under mild operating conditions and without
most parts of the world. Diesel fuel, which made up 29.0% of any material consumption. Therefore, the major challenge that
may exist against the employment of EDS using ILs is nding
a
Petroleum and Chemical Engineering Department, Sultan Qaboos University, 123,
the most suitable solvent. The fact that ILs have been catego-
Sultanate of Oman. E-mail: farouqsm@yahoo.com rized based on their constituent ions has simplied the
b
Department of Chemical and Environmental Engineering, Masdar Institute for Science screening and application of ILs in EDS. The potential of the
and Technology, Masdar City, Abu Dhabi, United Arab Emirates oxidative desulfurization (ODS) technique lies in the fact that it
† Electronic supplementary information (ESI) available. See DOI: preferentially attacks refractory sulfur compounds owing to the
10.1039/c6ra22032k

103606 | RSC Adv., 2016, 6, 103606–103617 This journal is © The Royal Society of Chemistry 2016
Paper RSC Advances

presence of high electron density on the sulfur atom.16 It can, of an electrophilic oxidant in the non-catalytic oxidation of
therefore, serve as a complementary desulfurization technique sulfur compounds available in diesel fuel. The ionic liquids
to be placed downstream of the conventional HDS unit in (ILs) whose extractive desulfurization (EDS) efficiencies were
reneries. Furthermore, the ease with which ODS can be inte- studied by our group29 were used as extraction solvents in
grated with other techniques such as extractive and adsorptive a simultaneous extraction and oxidation of the sulfur
desulfurization makes it more attractive. In fact, ODS can serve compounds (EODS) and/or post-oxidation extraction of the
as an improvement to the performance of ionic liquid-based sulfur compounds (OEDS) from the diesel fuel. The EDS
EDS using techniques such as extractive oxidative desulfuriza- performance of these ILs with and without an oxidation step
tion (EODS) and extractive catalytic oxidative desulfurization will be compared.
(ECODS).
Research activities in ODS can be categorized in terms of
catalytic/non-catalytic, type of oxidant and type of process, 2 Materials and methods
among others. Using hydrogen peroxide (H2O2) as an oxidant, 2.1 Chemicals
various heterogenous17–19 and homogenous16,20,21 catalysts have The ionic liquids tetrabutylphosphonium methanesulfonate
been used in ODS studies. H2O2 has been used extensively in [P4444][MeSO3] (98%) and triisobutyl(methyl)phosphonium p-
studies related to ODS partly because the by-product of the toluenesulfonate [Pi444,1][Tos] (98%) were supplied by Cytec and
oxidation reaction is water. However, this same advantage Sigma-Aldrich respectively. Acetonitrile (99.9%, HPLC grade)
negatively affects the performance of the oxidant as the pres- was supplied by Merck. Hexadecane (99%) was supplied by
ence of water results in formation of biphasic systems and BDH. Iso-dodecane (80%, mixture of isomers, tech.) and
a consequent mass transfer problem. Furthermore, the pres- toluene (ACS) were supplied by Alfa Aesar and Honeywell
ence of water during the IL-based EDS of oxidized fuel will result respectively. 1-Benzothiophene for synthesis was supplied by
in lower desulfurization efficiency. The potentials of other Merck. Dibenzothiophene (98%), dibenzothiophene sulfone
oxidants have also been studied. These oxidants include (97%) and meta-chloroperoxybenzoic (77%) acid were supplied
molecular oxygen,22 hydroperoxide generated in situ from air23 by Aldrich. Acetone (99.5%) was supplied by Sigma-Aldrich. All
and alkyl hydroperoxides, among others. chemicals were used as supplied without further purication.
Despite the elimination of the use of expensive catalysts, only Scheme 1 shows the structures of the cations and anions of
a few studies have been carried out on non-catalytic ODS.24,25 the ILs used in this work.
This can be attributed to the fact that results reported in the
literature for the non-catalytic ODS are not as impressive as
those reported for the catalytic systems. The apparently poor 2.2 Experimental procedures
performance of the non-catalytic ODS systems can be addressed A mixture of iso-dodecane (51%), hexadecane (39%) and
by using a systematically selected oxidant. In our previous toluene (10%) was prepared to represent the sulfur-free simu-
studies on the use of superoxide ion for oxidation of sulfur lated fuel. Appropriate amounts of benzothiophene (BT) and
compounds, it was observed that the superoxide ion, which acts dibenzothiophene (DBT) were dissolved in the aforementioned
as a nucleophile, inefficiently oxidized thiophene while leaving mixture in order to obtain a mixture with 500 ppm each of
dibenzothiophene unreacted.26 This was understood to be as benzothiophene (BT) and dibenzothiophene (DBT). For the
a result of the relatively higher electron density of dibenzo- kinetic studies, the BT- and DBT-doped simulated fuels were
thiophene as compared with thiophene, which is more prepared separately so that the kinetic studies carried out using
susceptible to the nucleophilic attack. This, therefore, implies one particular sulfur compound are not affected by the presence
that either a very weak nucleophilic or a strong electrophilic of the other.
oxidant may react with dibenzothiophene to produce its corre- 2.2.1 Preliminary kinetic studies. To establish the
sponding sulfone. meta-Chloroperoxybenzoic acid (mCPBA) optimum oxidant-to-sulfur molar ratio (O/S) to be employed for
seems to be a plausible option. In a typical reaction, an oxygen the preliminary kinetic studies, the effect of molar ratio on the
atom is transferred from the weak O–O bond of mCPBA to an % sulfur conversion (SC) of BT- and DBT-doped simulated fuels
electron-rich substrate, thereby making it an electrophilic was studied at 30  C, 30 min and a mixing rate of 600 rpm. The
oxidant. mCPBA may also serve as a nucleophilic oxidant in the O/S molar ratios studied were in the range 2.5 to 20 mol mol1.
presence of certain aldehydes.27 Although ammable and An appropriate amount of meta-chloroperbenzoic acid
hygroscopic, mCPBA is easy to handle and the concentration of (mCPBA) was added directly to about 2000 mg of BT- or DBT-
its commercial grade can reach up to 77 wt%. At 77 wt%, it doped simulated fuel to make up a predetermined (optimum)
possesses an active oxygen concentration of 7.14% and has molar ratio with the sulfur compound (O/S) present in the
been used in organic syntheses such as Baeyer–Villiger oxida- selected simulated fuel. The mixture of the oxidant and the
tion, Cope elimination and even oxidation of suldes.28 Despite simulated fuel was agitated in a screw-capped vial using
its application in the oxidation of suldes, work on oxidative a thermomixer. Aer a designated time, a sample of about
desulfurization of fuels using mCPBA as an oxidant is almost 60 mg was withdrawn and diluted with acetone, and the
non-existent. concentration of sulfur compound (DBT or BT) was analyzed by
The aim of this work is therefore to demonstrate the HPLC. The mixing rate was xed at 600 rpm throughout the
importance of oxidant selection and particularly the efficiency studies while the temperature was varied in the range 30  C to

This journal is © The Royal Society of Chemistry 2016 RSC Adv., 2016, 6, 103606–103617 | 103607
RSC Advances Paper

Scheme 1 Cation and anion used in this work.

45  C. The temperature was restricted to 45  C to allow sulfur (O/S) molar ratio was calculated and introduced into the
reasonable time of reaction. High temperatures resulted in bottle. Finally, the amount of IL required to make up the desired
faster reaction. As a result, sampling and analysis of DBT in MR was added at the appropriate time. Agitation of the reaction
particular becomes more complicated. To calculate the % SC (% mixture was carried out using a magnetic stirrer. For the anal-
X), eqn (1) (see below) was used. ysis of commercial diesel, energy dispersive X-ray uorescence
2.2.2 Extractive and oxidative desulfurization experiments. (EDXRF) was used to measure the concentration of elemental
Screw-capped Chromacol vials (20 mL) were used for both sulfur before and aer each experiment.
simultaneous extractive oxidative desulfurization (EODS) and 2.2.4 Regeneration and recovery. Regeneration was carried
post-oxidative extractive desulfurization (OEDS) studies. First, out by mixing distilled water with the spent IL at a ratio of 5 : 1
preliminary runs using the EODS system were carried out. (water : IL). Cellulose nitrate lter paper (0.2 mm pore, Sartorius
During these runs, the oxidant-to-sulfur ratio (O/S) was kept at Stedim) was used to lter out the precipitated sulfur
3 : 1 and the fuel-to-solvent mass ratio (MR) of 1 : 1 was retained compounds. The mixture of water and IL was introduced into
while the temperature and time were varied between 30 and the ask of a rotary evaporator (IKA RV 10), where water was
80  C and 30 and 120 min respectively. At the end of each evaporated under vacuum (25 mbar) at 95  C for 2 h using tap
experiment, the mixture was allowed to settle, aer which water at room temperature as the cooling uid. The regenerated
a sample of fuel was withdrawn for analysis. HPLC was used for IL was then used for another extraction stage. The advance
determining the sulfur content of the simulated fuel before and regeneration method was carried out in a similar fashion. The
aer the experiment. The relationship shown as eqn (1) was only difference was that n-hexane was mixed with the water–IL
used to determine the percentage sulfur removal. mixture at a ratio of 1 : 1 for 30 min prior to water removal using
C0  Ct the rotary evaporator. The degree of IL recovery was calculated
% SR or % SC ¼  100 (1) aer each extraction–regeneration stage using eqn (2) below:
C0
ðmIL ÞE  ðmIL ÞR
where % SR is % sulfur removal, % SC is sulfur conversion in % recovery ¼  100 (2)
ðmIL ÞE
the absence of an IL and the presence of mCPBA, C0 is initial
sulfur concentration (ppm), and Ct is sulfur concentration aer where (mIL)E is the mass of IL used in the extraction experiment,
t min of the experiment. and (mIL)R is the mass of IL recovered aer regeneration.
A response surface methodology was used for the optimiza-
tion of % SR of BT and DBT in the studied ILs. Design Expert 7.0 2.3 Optimized experimental data models
soware was used for the design, optimization and analysis of
the experiments. The four independent factors studied were 2.3.1 Response surface quadratic model. The following
time (15–80 min), temperature (30–60  C), mass fraction (0.5– quadratic equation (eqn (3)) was used for the tting and
0.83) and oxidant to sulfur ratio (3/1–9/1). All experiments were subsequent optimization of the experimental data obtained
carried out at a mixing speed of 600 rpm. A comparison between following the procedure outlined in Section 2.2.2.
the optimum EODS and OEDS was carried out. P P P
Y ¼ b0 + biXi + biiXi2 + bijXiXj (3)
2.2.3 Commercial diesel fuel desulfurization experiments.
To study the deep desulfurization of diesel fuel, an appropriate where Y is the response (% SR), Xi and Xj are the optimized
amount of mCPBA was added into a 50 mL screw-capped bottle. factors and b0, bi, bii and bij are coefficients of the intercept,
The amount of diesel required to make up a desired oxidant-to- linear, square and interaction effects respectively.

103608 | RSC Adv., 2016, 6, 103606–103617 This journal is © The Royal Society of Chemistry 2016
Paper RSC Advances

2.3.2 Kinetic reaction models. For the conversion of 3 min, then heated to 250  C at a heating rate of 6  C min, which
dibenzothiophene (DBT) and benzothiophene (BT) to their temperature was held for 10 min.
corresponding sulfones using meta-chloroperbenzoic acid Rigaku NexQC+ energy dispersive X-ray uorescence
(mCPBA) as an oxidant, the following stoichiometric reactions (EDXRF) was used for the elemental sulfur analysis of
were assumed: commercial diesel fuel. In a typical run, about 5 g of diesel fuel
was poured into the specialized sample cups (SC 3332), which
DBT + 2mCPBA / DBTO2 + 2mCBA (R1) were wrapped on one side with a thin lm (Prolene from
Chemplex) and inserted into the machine. Helium gas
BT + 2mCPBA / BTO2 + 2mCBA (R2) (99.9992%) was passed through the machine at 1 bar. Each
measurement took 300 seconds and was repeated three times.
The reactions with stoichiometry similar to (R1) and (R2) can
The minimum coefficient of determination (R2) for all calibra-
be assumed to be elementary irreversible bimolecular and tri-
tion curves for different sulfur ranges, 500–100 ppm, 100–
molecular reactions, respectively.30
20 ppm and 25–5 ppm, was 0.994867. The intensity extraction
The equation for the rate of an irreversible bimolecular
parameters were Ka (line), 2.24–2.37 (region) and low z (condi-
reaction is of the form:
tion). Acquisition conguration was set at low z, 6.5k, 100 mA
rDBT/BT ¼ kCDBT/BTCmCPBA (4) and open lter.

The reaction is rst order with respect to both DBT/BT and


mCPBA but second order overall.
3 Results and discussion
The linearized form is: 3.1 Preliminary kinetic studies
  3.1.1 Effect of O/S molar ratio. Fig. 1 shows the effect of O/S
M  2XA
ln ¼ CAO ðM  2Þkt; Ms2 (5) molar ratio on % SC. DBT, owing to its high electron density,
Mð1  XA Þ
reacts faster than BT. At a molar ratio of 5 : 1 (O/S), >85% SC was
where XA is the conversion and M is the initial molar ratio of achieved for DBT-containing simulated fuel while 11% SC was
reactants, which is 3 in this case. achieved in the case of BT. This is not very discouraging
The equation for the rate of an irreversible trimolecular considering the fact that typical hydrotreated diesel contains
reaction is of the form: 14% BTs (including their derivatives) and 86% DBT (and
derivatives) of total sulfur compounds31 with 4,6-dimethyldi-
rDBT/BT ¼ kCDBT/BTCmCPBA2 (6)
benzothiophene being the most refractory. However, dime-
thyldibenzothiophene possesses higher electron density. It is
The reaction is rst order with respect to DBT/BT and second
therefore safe to assume that its reaction rate will be faster than
order with respect to mCPBA but third order overall.
that of DBT if steric hindrance due to the alkyl group does not
The linearized form is:
  come into play.
ð2CAO  CBO ÞðCBO  CB Þ CAO CB
þ ln ¼ ð2CAO  CBO Þ2 kt; 3.1.2 Effect of mixing time and temperature. A molar ratio
CBO CB CA CBO of 3 : 1 (O/S) was used for studying the concentration–time
Ms2 (7) prole of DBT and BT. For DBT, the reaction time of 4200
seconds to 6600 seconds was used (Fig. 2). Because the reaction

2.4 Chemical analysis of samples


An Agilent high performance liquid chromatograph (HPLC)
1260 Innity equipped with a reversed-phase column and
variable wavelength detector was used. The column specica-
tion is 150  4.6 mm, 5 mm, 100A (Ecosil C18-extend). The
temperature of the column was set at 25  C. The mobile phase
consists of a mixture of 70 vol% acetonitrile and 30 vol% water.
Flow rate was kept at 1 mL min1 while the injection volume
was set to 1 mL. Detection of dibenzothiophene was made at
234 nm while that of benzothiophene was made at 290 nm.
Finally, the method R2 was determined as 0.99994 and 0.99972
for dibenzothiophene and benzothiophene respectively.
A Shimadzu GCMS-QP2010 Ultra equipped with HP-5MS
column was used for the determination of oxidation products.
The injection and oven temperatures were set to 250  C and
100  C respectively. Helium as the carrier gas was maintained at
a total ow rate of 14 mL min1 with column ow rate of 1 mL Fig. 1 Effect of oxidant/sulfur compound molar ratio on % sulfur
min1. The temperature program was as follows: 100  C held for conversion of DBT and BT at 30  C, 30 min and 600 rpm.

This journal is © The Royal Society of Chemistry 2016 RSC Adv., 2016, 6, 103606–103617 | 103609
RSC Advances Paper

of BT with the oxidant is relatively slow, the reaction time of Similar to the case of DBT, both bimolecular and trimo-
28 800 seconds was used (Fig. 3). In both DBT and BT oxidation, lecular rate orders were tted to the data obtained from the
the concentration of the reactant decreases with time as experiment using BT-containing simulated fuel. From visual
expected. observation of Fig. S2A and B,† the trimolecular reaction ts
3.1.3 Kinetic model tting. Both the bimolecular and tri- more satisfactorily than the bimolecular. Although not as
molecular reaction mechanisms were used in tting the data pronounced as in the gures, the R2 values presented in Table
obtained for DBT in an effort to determine the most suitable S2† also indicate this conclusion. Table S2† shows the rate
rate equation. Fig. 4A and B show the result of these ts. By constant for both bimolecular and trimolecular reactions and
visual observation, the bimolecular reaction seems to be their corresponding R2 values. As expected, the kinetic rate
a better t to the data. This can be conrmed by observing the constants for both the bi- and trimolecular reactions were lower
regression coefficients (R2) in Table S1 (ESI†). For the bimolec- than their corresponding values for DBT. If a bimolecular
ular reaction model, all the R2 values were greater than 0.9988 reaction is assumed in both cases, the ratio of the rate constants
while the best R2 value obtained for the trimolecular reaction (kDBT/kBT)1 becomes 20 at 30  C. Similarly, if a trimolecular
model was 0.9916. The bimolecular reaction model was chosen reaction is assumed for both cases, the ratio of the rate
as the model that satises the reaction stoichiometry. constants (kDBT/kBT)2 becomes 40 at 30  C. This is roughly
A plot of ln[k] against 1/T (Fig. S1†) was used to determine the double as compared to the bimolecular reaction constant,
activation energy for the bimolecular reaction. The activation which was found to be the same for all temperatures measured.
energy was determined as 49.98 kJ mol1. The nal form of the In either case, this means that the reaction rate of DBT is far
rate constant was determined as eqn (8) greater than that of BT. The trimolecular reaction mechanism
was selected to describe the stoichiometry of the reaction of the
rDBT ¼ kDBTCDBTCmCPBA, kDBT ¼ 2.82  1010 e49 978/RT (8) oxidant with BT. A plot of ln[k] against 1/T (Fig. S3†) was used to
determine activation energy for the trimolecular reaction. The
activation energy was determined as 41.90 kJ mol1. The nal
form of the rate constant was determined as presented in eqn
(S1).†
The implication of the results obtained from the kinetic
studies is that in the case of DBT the reaction only requires an
effective collision of a DBT molecule and an mCPBA molecule to
proceed. According to collision theory, an effective collision
takes place when the colliding molecules are in the proper
orientation and possess the right amount of energy. One would
think then that this reaction does not go directly to sulfone and
meta-chlorobenzoic acid (mCBA) as products. An intermediary,
possibly sulfoxide, can be assumed to be generated and then
subsequently converted to sulfone. In the case of BT, the reac-
tion requires a molecule of BT and two molecules of mCPBA to
collide effectively before it proceeds. The chance of a DBT
molecule and an mCPBA molecule colliding effectively is far
Fig. 2 Concentration–time profile of DBT at O/S ¼ 3/1, 600 rpm. greater than that of a molecule of BT and two molecules mCPBA
colliding at the same time, considering that both of these
compounds are at the ppm level. Furthermore, the trimolecular
reaction is more sensitive to concentration change than the
bimolecular reaction. This, alongside the difference in the
electron density on the sulfur atom of these compounds, is
perhaps the reason why DBT is more reactive than BT.
Although sulfones of DBT and BT are the expected products
of the reaction, a GC-MS analysis was carried out to conrm
this. Fig. 5 shows the GC-MS result aer oxidation. Dibenzo-
thiophene sulfone (DBTO2) and benzothiophene sulfone peaks
could be clearly seen aer the oxidation reaction at 24.8 min
(mass ¼ 216) and 15.3 min (mass ¼ 137), respectively. meta-
Chlorobenzoic acid, which is the by-product of the oxidation
reaction can be seen at 10.0 min (mass ¼ 156). The oxidation of
DBT to DBTO2 can be assumed to follow the route presented
schematically in Fig. 6. It is therefore important to remove the
sulfones and mCBA as well as the unreacted mCPBA to ensure
Fig. 3 Concentration–time profile of BT at O/S ¼ 3/1, 600 rpm. effective oxidative and extractive desulfurization.

103610 | RSC Adv., 2016, 6, 103606–103617 This journal is © The Royal Society of Chemistry 2016
Paper RSC Advances

Fig. 4 (A) Plot of ln[(M  2XA)/(M(1  XA)] versus time at different temperatures (bimolecular), M ¼ 3. (B) Plot of (2CAO  CBO)(CBO  CB)/(CBOCB) +
ln[CAOCB)/CACBO] against time at different temperatures (trimolecular) for DBT.

Fig. 5 Products of oxidation of DBT and BT in simulated fuel using mCPBA.

3.2 Oxidative and extractive desulfurization An IL that allows the oxidant to preferentially reside in the
fuel phase can be used in extractive oxidative desulfurization
Studies related to oxidative and extractive desulfurization can
(EODS) experiments, where extraction and oxidation occur
be kinetics-limiting or mass-transfer-limiting depending on
simultaneously. Because both oxidized and unoxidized sulfur
where the oxidant preferentially resides. If the oxidant resides
compounds simultaneously migrate to the IL phase, this
in the IL phase, then sulfur compounds have to migrate into the
process is expected to reduce the mixing time required for the
IL phase before they can be converted. When the reaction is fast
same IL. However, it must be ensured that unreacted mCPBA
relative to the rate of migration of the sulfur compounds, then
and mCBA either get extracted along with the sulfones or
the process is mass-transfer-limiting. Although a high reaction
subsequently removed. Thus, in this kinetics-limiting system,
rate, as in the case of DBT, will result in a concentration
ensuring high reaction rate is the main goal. There are two
gradient for DBT and a consequent faster migration into the IL
options for ILs that do not allow the oxidant to reside in the fuel
phase, the overall sulfur removal could still be low. On the other
phase. If the rate of the oxidation reaction in the IL phase
hand, if the oxidant preferentially resides in the fuel phase,
signicantly promotes the migration of the unoxidized sulfur
assuming the more polar sulfones produced from the oxidation
compounds, then an EODS process is suitable. Otherwise, post-
reaction migrate into the IL phase faster than the oxidation
oxidation extractive desulfurization (OEDS) is preferred despite
reaction occurs, the process becomes rate-limiting.
the advantages of EODS. In the OEDS process, the oxidant will

This journal is © The Royal Society of Chemistry 2016 RSC Adv., 2016, 6, 103606–103617 | 103611
RSC Advances Paper

Fig. 6 Non-catalytic oxidation route of DBT in the presence of mCPBA.

have maximum contact with the sulfur compounds, with of both EODS and OEDS strategies was carried out and
a consequent high reaction rate. The treated fuel can then be compared.
transferred into an EDS unit where the sulfones will be removed 3.2.1 Optimization of the EODS of simulated fuel using
by the IL. Generally, these ILs would also simultaneously [Pi444,1][Tos] IL. Response surface methodology using a central
remove the product of the oxidation reactions and the unreac- composite experimental design was employed for the analysis,
ted oxidant. It should be noted that the above discussion is tting and optimization of the factors affecting the EODS of the
applicable because the oxidant can dissolve in the fuel. simulated fuel. The details of this analysis can be found in the
Oxidants such as the 30 wt% H2O2 do not offer such exibility ESI† le. The coefficients of the reduced model (quadratic
except in the presence of a phase-transfer agent. model) terms are presented in Table S11 (ESI†). It was observed
A preliminary study carried out using the tri- that while DBT can be completely removed, only 78.7% of BT
isobutylphosphonium tosylate ([Pi444,1][Tos]) IL for the simul- can be removed (Tables 1 and 2). Although the % SR obtained is
taneous extractive and oxidative desulfurization (EODS) of the higher than that obtained prior to optimization, BT concen-
simulated fuel showed relative improvement as compared with tration is still beyond the acceptable limit of #15 ppm.
solely EDS of the same fuel. This was true at all temperatures It was also observed that mass fraction is the most signi-
and times considered (Fig. S4 and S5†). Although an improve- cant term in the ANOVA table (Table S4†) for the EODS of BT
ment was observed using the EODS strategy, the improvements using an IL. This is similar to the results obtained from the EDS
obtained were not as much as expected and this was attributed experiments, indicating mass transfer is an important factor in
to the preferential residence of the oxidant as discussed above. the % SR of BT using an IL.29 It also indirectly indicates that the
Therefore, the post-oxidative extractive desulfurization (OEDS) oxidant preferentially resides in the IL phase and the migration
might be a better strategy in this case. As a result, optimization of BT into the IL phase is responsible for the relatively low
improvements. Furthermore, going back to the preliminary
kinetic studies presented, DBT possesses high reaction rate,
Table 1 Validation of the predicted optimum of performance of the which resulted in 60% conversion even at 30  C, aer 30 min.
respective ILs in removing DBT But as seen in Fig. S5,† 90% SR (DBT) was achieved at 60  C aer
the same interval of time. It can be concluded that the sulfur
[Pi444,1][Tos] OEDS [P4444][MeSO3] OEDS
compounds in the presence of an IL do not have as much
Time (X1, min) 15 15 contact with oxidant as they have in its absence, which also
Temperature (X2,  C) 60 60 points toward mass transfer resistance. It is therefore safe to
Mass fraction (X3, w/w) 0.5(1 : 1) 0.67(2 : 1) assume that the process is mass-transfer-limiting and that the
O/S (X4, mol mol1) 3/1 3/1
rate of migration of the sulfur compounds even in the presence
Prediction (SR, %) 100 100
Validation (SR, %) >99 >99 of the oxidation step is not favorable. This calls for the adap-
tation of the OEDS strategy previously mentioned.

Table 2 Validation of the predicted optimum of performance of the respective ILs in removing BT

[Pi444,1][Tos] EODS [Pi444,1][Tos] OEDS [Pi444,1][Tos] OEDS [P4444][MeSO3] OEDS

Time (X1, min) 80 60 72 60


Temperature (X2,  C) 60 60 60 60
Mass fraction (X3, w/w) 0.5(1 : 1) 0.5(1 : 1) 0.5(1 : 1) 0.5(1 : 1)
O/S (X4, mol mol1) 7.5/1 6.5/1 8/1 6/1
Prediction (SR, %) 97 97 100 97
Validation (SR, %) 78.7 94.5 99 95

103612 | RSC Adv., 2016, 6, 103606–103617 This journal is © The Royal Society of Chemistry 2016
Paper RSC Advances

3.2.2 Optimization of the OEDS of simulated fuel using Since very high DBT removal (>99%) was achieved using the
[Pi444,1][Tos] and [P4444][MeSO3] IL. In the OEDS, oxidation and OEDS strategy as compared with the EODS strategy, BT was the
extraction take place in separate stages thereby allowing the limiting reactant and the design of the experiment and opti-
oxidant to come into maximum contact with the sulfur mization of the responses were carried out on BT in a similar
compounds. The oxidized sulfur compounds, which become fashion to the foregoing section. The ranges in the variables are
more polar, can be removed easily in the extraction unit. The the same, except for total time, which was made to be between
variables used for the optimization of the EODS of BT using 15 min and 81 min. This is to avoid fractions of either oxidation
[Pi444,1][Tos] can be assumed to be equally important and used or extraction time. Details of the experimental design and
for the optimization of OEDS of BT. However, in OEDS, one optimization can be found in the ESI.†
more variable has to be considered. This is the ratio of oxidation- Aer carrying out the optimization, it was evident that the
to-extraction times. This variable can be split into two variables, OEDS was superior to the EODS (Table 2). It can be seen that at
namely oxidation time and extraction time, thereby removing 60  C, MR of 1 : 1, O/S of 6.5 and total time of 60 min, a % SR
“time” as an independent variable. In both cases, there will be (BT) of 97.4%, (which was validated as 94.5%), was achieved. In
a total of ve variables. This results in a large number of exper- comparison with the EODS strategy, this requires 20 min less
iments, thereby tending to produce more errors. Hence, an time and 1/1 lower O/S to achieve a better BT removal of 94.5%
experiment was carried out in order to retain the number of as against 78.7%. Despite the improvement, the set target had
variables at four while keeping the ratio of oxidation-to-extraction not yet been achieved. To achieve the target of #15 ppm, at least
(oxidation/extraction) times constant. This is shown in Fig. 7. It 97.0% of sulfur must be removed from a fuel containing
should be noted that oxidation still takes place in the extraction 500 ppm sulfur. The conditions under which this was achieved
portion of the time and therefore an oxidation/extraction time of are presented in Table 2.
zero corresponds to an EODS strategy. The OEDS strategy was employed to study the improvement
It can be seen that as the oxidation/extraction time was in the performance of the [P4444][MeSO3] IL upon the intro-
decreased from 20/10 to 0/30, DBT decreased from 96% to 61%. duction of an oxidation step. The OEDS was considered because
Unexpectedly, BT increased from 42% to 50%. This implies this IL is hydrophilic and is expected to have similar behavior to
that, whereas the assumption made that the relatively poor DBT [Pi444,1][Tos] in this context. Therefore an experiment was
removal was due to inefficient contact with the oxidant holds, designed and optimized using the same ranges of variables as
the same cannot be said for the removal of BT under these in the OEDS using [Pi444,1][Tos]. The only difference was that the
conditions. This can be attributed to the low reaction rate of BT temperature was kept constant at 60  C and this brought down
with the oxidant. Therefore, a possibility exists that upon the number of experimental runs to 17. The experimental
increase in temperature, the increase in rate of oxidation of BT design matrix of the OEDS of BT using this IL is presented in
in an OEDS setup will most likely be higher than that in an Table S9.†
EODS setup owing to higher collision rate, thereby resulting in 3.2.3 Summary of the performances of the ILs. Having
a better BT removal overall. The oxidation/extraction time of 20/ carried out the preliminary investigation and optimization of
10 (two-thirds of total time) was selected for the purpose of the the factors affecting the EODS and OEDS processes, the coeffi-
optimization. This is due to the fact that it showed highest DBT cients of the tted quadratic model representing the response
removal and a possibility that a change in one of the indepen- surface in EODS/OEDS of the BT and DBT using the respective
dent variables or interaction between the variables might result ILs are presented in Table S11.† Values of terms with negative
in higher BT removal than in an EODS setup. coefficients need to be reduced while values of terms with
positive coefficients need to be increased to achieve high % SR.
The 3D response surface plot for the all the systems studied can
be found in the ESI.†
3.2.3.1 Dibenzothiophene removal. Table 1 presents the
validated result for the optimized OEDS of DBT using the
respective ILs. The fact that a better result was obtained using
[P4444][MeSO3] indicates that this IL is superior to the other IL.
Comparing the results obtained in this work with others avail-
able in the literature shows that the choice of mCPBA as an
oxidant is valid. This is because most of the reported studies
showed that >99% DBT removal can be achieved but either at
a higher time of up to 6 h, with a high O/S ratio or in the
presence of a catalyst.4 The results obtained in this work are
promising and clearly better than most of the results available
in the literature for DBT removal, to the best of the authors'
knowledge.
3.2.3.2 Benzothiophene removal. Table 2 presents the vali-
Fig. 7 % SR as a function of ratio of oxidation-to-extraction time using dation of EODS/OEDS optimum operating conditions for BT
[Pi444,1][Tos], 30  C, 30 min, 600 rpm, 3/1 (O/S). using the respective ILs. The validations are in reasonable

This journal is © The Royal Society of Chemistry 2016 RSC Adv., 2016, 6, 103606–103617 | 103613
RSC Advances Paper

agreement with the predicted results, which indicates the


model, coefficients of which are presented in Table S11,†
reasonably describes the studied systems. It can be concluded
that the target of #15 ppm can be achieved by optimizing the
factors affecting the process. It can be seen (Table 2) that to
increase BT removal from 95% to 99%, time and O/S molar ratio
have to be increased from 60 to 72 min and from 6.5/1 to 8.0/1
respectively. This is because as more BT is removed its
concentration in the fuel phase becomes very low and conse-
quently the probability that it will engage in effective collisions
with oxidant becomes low. It should be recalled that the reac-
tions of BT and mCPBA were assumed to be trimolecular, and
therefore sensitive to changes in concentration. This means
that the rate of reaction becomes very low at low concentrations
Fig. 8 % SR for the OEDS of commercial diesel (iS ¼ 275 ppm) using
of BT in the fuel phase. the [Pi444,1][Tos] IL as a function of water-washing time, at 60  C, MR
The expected products of reaction are DBTO2, BTO2 and 1 : 1, O/S 6/1, for 60 min.
mCBA. At the optimum conditions, it was observed that neither
the aforementioned products of the reaction nor the unreacted
mCPBA were present in the fuel aer treating it with the It should be recalled that similar operating conditions when
hydrophilic phosphonium-based ILs. Furthermore, it should be applied on the simulated fuel resulted in SR of 95%. The rela-
noted that the OEDS process used was based on the oxidation tively low performance of the OEDS process using the same IL
time being kept at two-thirds of the total time. can be attributed to the possible presence of other similar
compounds such as aromatic nitrogen-based compounds that
may compete with the sulfur compounds in the oxidation or
3.3 Commercial diesel desulfurization
even extraction step. The washing time of 30 min at 1 : 1 diesel-
OEDS of commercial diesel fuel obtained from a local lling to-water ratio seemed to be the optimum option and therefore
station was carried out using the ILs selected in this work. XRF was used in studying the effect of O/S molar ratio, shown in
analysis on the diesel fuel samples indicated sulfur content of Fig. 9.
275 ppm. An SR of $94% is enough to ensure deep desulfur- The O/S molar ratio of zero indicates EDS of the diesel fuel
ization (#15 ppm) of this diesel fuel. To achieve this value using for 60 min. It can be seen that the O/S molar ratio has virtually
the [Pi444,1][Tos] IL in an OEDS strategy, it was shown that the no impact on the % SR at O/S $ 4/1. This implies that adding
temperature, mass ratio, O/S ratio and total time should be kept more oxidant will not result in any improvement in the process.
at 60  C, 1 : 1, 6.5/1 and 60 min (40 min oxidation and 20 min It also means that the possibility that the oxidation is the
extraction). Using the same parameters with the exception of O/ limiting step of the process is slim. Keeping the O/S molar ratio
S, which was kept at 6/1, the results presented in Fig. 8 were at 4/1 and increasing the total time to 90 min (60 min oxidation
obtained. The negative % SR actually indicates an increase in and 30 min extraction) did not improve the result (SR ¼ 74%)
the sulfur content in the diesel fuel. This is the result of the rst either.
experiment carried out and it was observed that the IL ([Pi444,1]
[Tos]), which contains sulfur in its anion, dissolved in the fuel.
It should be noted that the solubility of the IL in the diesel is
still relatively small considering the fact that a fuel-to-IL ratio of
1 : 1 is used while the sulfur content of diesel fuel only
increased from 275 ppm to 451 ppm. One of the advantages of
the hydrophilic phosphonium ILs is that they can easily be
removed from the fuel phase with water. While the IL can be
removed by water-washing the diesel fuel, the sulfur
compounds are not expected to dissolve in water. This implies
that any decrease in the sulfur content of the diesel fuel by
water-washing is as a result of the removal of the sulfur-
containing IL.
By adding water to the treated diesel fuel at a diesel-to-water
mass ratio of 1 : 1 and stirring for 30 min, an increase in the SR
to 74% (71 ppm) was observed. Increasing the washing time to
60 min did not affect the result, which indicates either complete
removal of the IL or a very low concentration of the IL in the Fig. 9 Effect of O/S molar ratio on % SR for the OEDS of commercial
diesel fuel phase. Increasing the diesel-to-water mass ratio to diesel using [Pi444,1][Tos], 60  C, MR 1 : 1, 60 min, washing time 30 min,
1 : 3 did not signicantly improve the result (SR ¼ 75%) either. diesel-to-water ratio 1 : 1.

103614 | RSC Adv., 2016, 6, 103606–103617 This journal is © The Royal Society of Chemistry 2016
Paper RSC Advances

If the extraction step is the limiting step, then using [P4444] Using [P4444][MeSO3] under the conditions shown in Table 3,
[MeSO3] will likely improve the result as it is a better extraction the sulfur content can be reduced from 56 ppm to 31 ppm by
solvent. Therefore, keeping the temperature at 60  C, MR at addition of a new EDS stage and to 24 ppm by addition of yet
1 : 1, O/S at 4 : 1, total OEDS time at 60 min, washing time at another EDS stage (Fig. 11). This therefore becomes a process
30 min and diesel-to-water mass ratio at 1 : 1, the performance composed of a 40 min oxidation stage and three 20 min
of the two hydrophilic phosphonium-based ILs was compared extraction stages. Even with the three extraction stages the % SR
while introducing stage-wise water mixing. This result is pre- does not reach the target. It can be deduced that the sulfur
sented in Fig. 10. removal is becoming more difficult as the target is approached.
It can be seen that the decrease in the sulfur content of the This is due to very low concentration of the sulfur compounds,
diesel from the second stage to the third stage is not very which affects their distribution. Maintaining the three extrac-
signicant and therefore it can be assumed that adding another tion stages while doubling the extraction time from 20 min to
stage is not worthwhile. Secondly, it can be seen that, at the 40 min resulted in a sulfur content of 15.6 ppm.
third stage, the sulfur content of the diesel treated using the
[P4444][MeSO3] IL (56 ppm) is less than that of the diesel treated
with the [Pi444,1][Tos] IL (61 ppm). Three-stage water washing 3.4 Regeneration and recovery
was adopted for all the subsequent experiments. An effort was Having carried out the deep oxidative desulfurization of
made to improve the performance of [P4444][MeSO3] by doubling commercial diesel using the [P4444][MeSO3] IL, the regeneration
the O/S molar ratio to 8 : 1 using the current operating condi- of the same IL was carried out. Fig. 12 shows the result of this
tions. However, the sulfur content obtained was 55 ppm, indi- study. The change in the % SR of the EDS of commercial diesel
cating that the oxidation step may not be the limiting step, using the regenerated IL as compared with the fresh IL is pre-
similar to the case of the [Pi444,1][Tos] IL. Comparison between sented. The % SR dropped slightly aer the [P4444][MeSO3] IL
the OEDS and EODS strategy using the two ILs is presented in was used four times (third regeneration). The % SR obtained
Table 3. aer the third regeneration stage was 95% that of the fresh IL.
However, using the advance regeneration method (Adv) the
original % SR of the IL was restored.
Fig. 13 shows the % recovery of the IL aer each regeneration
stage. The % recovery was 90% for the rst regeneration stage
and 86% for the advanced regeneration stage. It should be
noted that some of the ILs can be recovered from the water used
for the three-stage diesel water-washing process.
Therefore, a suitable system for the deep desulfurization of
the commercial diesel involves the use of [P4444][MeSO3] as
a solvent for the extraction of sulfur compounds from diesel fuel
treated with mCPBA at an O/S molar ratio of 4/1 at 60  C. The
extraction is carried out for 40 min in three stages at 60  C using
a fuel-to-solvent mass ratio (MR) of 1 : 1. Finally the treated
diesel is separated from dissolved IL by water-washing it in

Fig. 10 Sulfur content in treated diesel fuel as a function of water-


washing stages, 60  C, MR 1 : 1, 60 min, washing time 30 min, diesel-
to-water ratio 1 : 1.

Table 3Performance of all selected ILs for the EODS/OEDS of


commercial diesela

Sulfur content aer EODS Sulfur content aer OEDS


IL treatment (ppm) treatment (ppm)

[Pi444,1][Tos] 145.39 60.99


[P4444][MeSO3] 144.69 55.63
a
Temperature 60  C, MR 1 : 1, O/S 4 : 1 (*6/1), total OEDS or EODS time
60 min, washing time 30 min, diesel-to-water mass ratio 1 : 1, three- Fig. 11 A plot of sulfur content of commercial diesel (initial concen-
stage water washing. tration of 275 ppm) as a function of number of post-oxidation
extraction stages.

This journal is © The Royal Society of Chemistry 2016 RSC Adv., 2016, 6, 103606–103617 | 103615
RSC Advances Paper

three stages using a diesel-to-water mass ratio of 1 : 1. This is


followed by a set of regeneration stages as presented in Fig. 14.

4 Conclusion
In this work, the feasibility of an electrophilic oxidant (meta-
chloroperoxybenzoic acid) as a suitable and efficient reagent for
the non-catalytic conversion of refractory sulfur compounds to
their corresponding sulfones has been presented. The oxidant
possesses a number of advantages over the commonly used
hydrogen peroxide by virtue of its electrophilicity, solubility in
the fuel phase and lower water content. A preliminary kinetic
study carried out employing benzothiophene (BT) and diben-
zothiophene (DBT) in a simulated fuel showed that the latter,
owing to the high electron density on its sulfur atom, is ex-
Fig. 12 Regeneration of [P4444][MeSO3]. pected to be more susceptible to electrophilic attack by the
oxidant. A bimolecular reaction kinetic model correlates well
with the experimental data obtained using DBT as the
compound of interest while a trimolecular reaction kinetic
model correlates well with that of BT. This partly explains the
reason why the DBT conversion is much faster.
Application of meta-chloroperoxybenzoic acid as oxidant in
an ionic liquid-based oxidative desulfurization of the simulated
fuel revealed noticeable improvements as compared with the
ionic liquid-based extractive desulfurization of the simulated
fuel using the same solvent and experimental conditions. It was
observed that the optimum performance of the ILs varies
depending on whether the extraction and oxidation stages were
carried out simultaneously as in extractive oxidative desulfur-
ization (EODS) or separately as in post-oxidation extractive
desulfurization (OEDS). This in turn was understood to be
dependent on the phase in which the oxidant preferentially
resides. It can then be concluded that while EODS may be
a suitable strategy when using a hydrophobic IL, which allows
the oxidant to preferentially reside in the fuel phase, the OEDS
Fig. 13 A plot of IL recovery against regeneration stage. strategy is suitable for the hydrophilic phosphonium-based ILs
used in this work. This is to allow maximum contact between
the oxidant and sulfur compounds.

Fig. 14 Schematic of the proposed post-oxidative extractive desulfurization of commercial diesel. ULSF, ultra-low-sulfur fuel; WS, waste solid
compounds; clean-up, 3-stage water-washing and drying.

103616 | RSC Adv., 2016, 6, 103606–103617 This journal is © The Royal Society of Chemistry 2016
Paper RSC Advances

Employing the OEDS strategy, deep desulfurization of 13 H. Zhao and G. A. Baker, Front. Chem. Sci. Eng., 2015, 9, 262–
a commercial diesel fuel was achieved using the tetrabutyl- 279.
phosphonium methanesulfonate IL at optimized experimental 14 S. Escobar, A. Rodriguez, E. Gomez, A. Alcon, V. E. Santos
conditions. The IL was successfully regenerated using a combi- and F. Garcia-Ochoa, Bioprocess Biosyst. Eng., 2016, 1–10,
nation of water and n-hexane as primary and secondary regen- DOI: 10.1007/s00449-016-1536-6.
eration solvents respectively. 15 I. Martinez, V. E. Santos, E. Gomez and F. Garcia-Ochoa,
J. Chem. Technol. Biotechnol., 2016, 91, 184–189.
Acknowledgements 16 S. Otsuki, T. Nonaka, N. Takashima, W. Qian, A. Ishihara,
T. Imai and T. Kabe, Energy Fuels, 2000, 14, 1232–1239.
The authors appreciate the nancial support of The Research 17 O. Gonzalez-Garcıa and L. Cedeno-Caero, Catal. Today, 2009,
Council and Sultan Qaboos University, Muscat, Oman, under 148, 42–48.
the project RC/ENG/PCED/12/02. 18 K. G. Haw, W. A. W. AbuBakar, R. Ali, J. F. Chong and
A. AbdulKadir, Fuel Process. Technol., 2010, 91, 1105–1112.
References 19 J. M. Fraile, C. Gil, J. A. Mayoral, B. Muel, L. Roldán, E. Vispe,
S. Calderón and F. Puente, Appl. Catal., B, 2016, 180, 680–
1 L. Garverick, Corrosion in the petrochemical industry, http:// 686.
books.google.com/books?id¼XblTAAAAMAAJ. 20 F. Al-Shahrani, T. Xiao, S. A. Llewellyn, S. Barri, Z. Jiang,
2 M. D. Argyle and C. H. Bartholomew, Catalysts, 2015, 5, 145– H. Shi, G. Martinie and M. L. H. Green, Appl. Catal., B,
269. 2007, 73, 311–316.
3 A. Stanislaus, A. Mara and M. S. Rana, Catal. Today, 2010, 21 S. H. Ali, D. M. Hamad, B. H. Albusairi and M. A. Fahim,
153, 1–68. Energy Fuels, 2009, 23, 5986–5994.
4 F. S. Mjalli, O. U. Ahmed, T. Al-Wahaibi, Y. Al-Wahaibi and 22 N. Tang, Z. Jiang and C. Li, Green Chem., 2015, 17, 817–820.
I. M. AlNashef, Rev. Chem. Eng., 2014, 30, 337–430. 23 R. Sundararaman, X. Ma and C. Song, Ind. Eng. Chem. Res.,
5 OPEC, World Oil Outlook, Vienna, 2011. 2010, 49, 5561–5568.
6 Z. Jiang, H. Lu, Y. Zhang and C. Li, Chin. J. Catal., 2011, 32, 24 R. Javadi and A. d. Klerk, Energy Fuels, 2011, 26, 594–602.
707–715. 25 S. Pysh'yev, Chem. Chem. Technol., 2012, 6, 229–235.
7 Y.-X. Li, W.-J. Jiang, P. Tan, X.-Q. Liu, D.-Y. Zhang and 26 O. U. Ahmed, F. S. Mjalli, T. Al-Wahaibi, Y. Al-Wahaibi and
L.-B. Sun, J. Phys. Chem. C, 2015, 119, 21969–21977. I. M. AlNashef, Ind. Eng. Chem. Res., 2015, 54, 2074–2080.
8 S. S. Shah, I. Ahmad and W. Ahmad, J. Hazard. Mater., 2016, 27 S. Kamijo, S. Matsumura and M. Inoue, Org. Lett., 2010, 12,
304, 205–213. 4195–4197.
9 I. Shimoyama and Y. Baba, Carbon, 2016, 98, 115–125. 28 H. Hussain, A. Al-Harrasi, I. R. Green, I. Ahmed, G. Abbas
10 P. Nancarrow, N. Mustafa, A. Shahid, V. Varughese, U. Zaffar, and N. U. Rehman, RSC Adv., 2014, 4, 12882–12917.
R. Ahmed, N. Akther, H. Ahmed, I. Alzubaidy, S. Hasan, 29 O. U. Ahmed, F. S. Mjalli, T. Al-Wahaibi, Y. Al-Wahaibi and
Y. Elsayed and Z. Sara, Ind. Eng. Chem. Res., 2015, 54, I. M. AlNashef, Ind. Eng. Chem. Res., 2015, 54, 6540–6550.
10843–10853. 30 O. Levespiel, Chemical Reaction Engineering, Wiley, New
11 K. Zhao, Y. Cheng, H. Liu, C. Yang, L. Qiu, G. Zeng and York, 1999.
H. He, RSC Adv., 2015, 5, 66013–66023. 31 M. F. Ali, A. Al-Malki and S. Ahmed, Fuel Process. Technol.,
12 Z. Ismagilov, S. Yashnik, M. Kerzhentsev, V. Parmon, 2009, 90, 536–544.
A. Bourane, F. M. Al-Shahrani, A. A. Hajji and
O. R. Koseoglu, Catal. Rev., 2011, 53, 199–255.

This journal is © The Royal Society of Chemistry 2016 RSC Adv., 2016, 6, 103606–103617 | 103617

S-ar putea să vă placă și