You are on page 1of 10

Desulfurization

Mohammad Amin Sobati1


Asghar Molaei Dehkordi1

Research Article

Mohammad Shahrokhi1

Extraction of Oxidized Sulfur-Containing


Compounds of Non-Hydrotreated Gas Oil

Department of Chemical and


Petroleum Engineering, Sharif
University of Technology,
Tehran, Iran.

1515

The desulfurization of non-hydrotreated gas oil using oxidation followed by


liquid/liquid extraction has been studied. The solvent extraction of oxidized sulfur-containing compounds of gas oil was carried out by utilizing various types of
solvents, including acetonitrile, methanol, and 96 vol.-% ethanol. The performance of these solvents was carefully evaluated and compared by considering
both desulfurization and gas oil recovery. It was found that their performance follows the order: acetonitrile > methanol > 96 vol.-% ethanol. The influences of operating conditions, such as number of extraction stages and solvent to gas oil ratio, on the desulfurization and recovery of gas oil, were investigated. Over 99 %
desulfurization of gas oil was achieved after liquid/liquid extraction of the oxidized sulfur-containing compounds.
Keywords: Desulfurization, Gas oil, Liquid/liquid extraction, Oxidation, Solvent extraction
Received: March 06, 2010; revised: May 09, 2010; accepted: May 21, 2010
DOI: 10.1002/ceat.200900622

Introduction

Sulfur-containing compounds in middle distillate fuels are


known to have negative impacts on the environment because
of SOx emissions from their combustion exhaust products,
and the associated contribution to acid rain. The SOx compounds also poison the catalysts involved in the conversion of
NOx, CO, and particulate matter for exhaust gas treatment in
vehicles [112]. In addition, desulfurization has recently
become a worldwide challenge because of more stringent environmental regulations. At present, hydrodesulfurization
(HDS) is the most widely used process in oil refineries for
removing thiols, sulfides, and disulfides from middle distillate
fuel oils [79, 1113]. However, the major sulfur-containing
compounds in middle distillate fuels, especially in the gas oil
boiling range, are thiophenic compounds and their alkyl-substituted derivatives. It is hard to remove these refractory sulfur-containing compounds such as dibenzothiophene (DBT)
and its derivatives (methyldibenzothiophene (MDBT), 4,6-dimethyldibenzothiophene (4,6-DMDBT)) by HDS due to their
steric-hindrances even under strong desulfurization conditions
[7, 13]. Therefore, it is desirable to develop supplementary or
alternative processes to HDS for desulfurization [9, 11, 12].
Among different alternative processes, oxidative desulfurization (ODS) has attracted much attention due to its significant
advantages. The greatest advantage of ODS compared to con-

Correspondence: Assoc. Prof. Dr. A. M. Dehkordi (amolaeid@sharif.


edu), Department of Chemical and Petroleum Engineering, Sharif
University of Technology, P.O. Box 11155-9465, Tehran, Iran.

Chem. Eng. Technol. 2010, 33, No. 9, 15151524

ventional HDS is that it can be carried out in the liquid phase


under very mild operating conditions, i.e., near room temperature and under atmospheric pressure. Besides, in the ODS
process, the majority of HDS refractory sulfur-containing
compounds such as 4,6-DMDBT are easily oxidized into more
polar sulfoxides and subsequently sulfones, which can be
removed by conventional separation processes such as solvent
extraction or adsorption [11].
The main separation process of oxidized sulfur-containing
compounds is liquid/liquid extraction. In the separation of
oxidized sulfur-containing compounds by the liquid/liquid
extraction process, both fuel oil loss and sulfur removal should
be considered simultaneously [11]. Thus, the design and operating conditions of the liquid/liquid extraction process, i.e.,
solvent type, number of extraction stages, and the solvent to
fuel oil ratio should be set in such a way to achieve high levels
of desulfurization with acceptable fuel oil recovery.
Several oxidants such as nitric acid (HNO3) [14, 15], tertbutyl-hydroperoxide (TBHP) [1619], superoxides [20], and
ozone [21] have been used in the ODS processes. Among these
oxidants, hydrogen peroxide (H2O2) is preferentially chosen as
the primary oxidant [2236] due to its environmentally benign
properties. However, H2O2 needs to be activated in the presence of a catalyst such as formic acid [23, 25, 26, 2832], acetic
acid [6, 22, 23, 25, 27], acetic acid-sulfuric acid [24], phosphotungstic acid [33], sulfuric acid [34], iron-complexes (TAML,
activators) [35], and activated carbon plus formic acid [36],
etc. Among different oxidation systems, the H2O2/carboxylic
acid system (especially formic acid) has several advantages
such as reaction simplicity, commercial availability of catalyst,
and no requirement to have the solvent and the solid catalyst

2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.cet-journal.com

1516

M. A. Sobati et al.

in the oxidation media. These advantages render the H2O2/formic acid system an interesting option among different oxidation systems for the ODS process. Several studies have been
conducted on the H2O2/formic acid oxidation system. These
studies mainly focused on the investigation of the effects of the
main oxidation parameters on the desulfurization. Otsuki et al.
[29] have studied the ODS of two high sulfur fuel oils, i.e.,
straight run light gas oil (SR-LGO) and vacuum gas oil (VGO)
with total sulfur contents of 1.35 wt % and 2.17 wt %, respectively, by using the H2O2/formic acid oxidation system. However, the authors used very high amounts of oxidant and formic acid in their experiments [29]. In addition, the fuel oil
recovery reported in their studies was very low, i.e., only 40 %
recovery of the initial fuel oil after ten extraction stages by
means of dimethylformamide (DMF) for desulfurization of
VGO up to 100 ppm. Hao et al. [30] have studied desulfurization of non-hydrotreated coker gas oil (CGO) with a total sulfur content of 2.5844 wt % by oxidation using the H2O2-formic acid system followed by liquid/liquid extraction. The
authors have also reported that n-methyl pyrrolidone (NMP)
was the best solvent for the extraction of oxidized sulfur-containing compounds among NMP, DMF, and furfural, and a
90 % desulfurization level has been obtained after oxidation
followed by extraction with NMP. However, the authors did
not report the fuel oil recovery or fuel oil loss after extraction,
which are important parameters for the evaluation of the
desulfurization process involved.
Ali et al. [23] have studied the oxidative desulfurization of
gasoline and hydrotreated diesel fuel using a H2O2/carboxylic
acid, i.e., formic and acetic acid, oxidation system. The authors
found the ODS process to be a promising approach for the
desulfurization of diesel fuel and reported ca. 92 % desulfurization of diesel. However, they have also stated that the ODS is
not successful for gasoline due to the high olefin content that
tends to react with H2O2 to form epoxides. Ali et al. [24] have
studied the oxidative desulfurization of diesel with H2O2 and
acetic acid using sulfuric acid (H2SO4) as a catalyst and reported ca. 90 % desulfurization.
Although there are numerous publications and reports on
the ODS of middle distillate fuel oils using H2O2/formic acid,
comprehensive studies focusing on the extraction of oxidized
sulfur-containing compounds by considering the effects of the
main operating parameters of the extraction process such as
solvent type, number of extraction stages, and solvent to fuel
oil ratio on the final desulfurization and fuel oil recovery is
lacking. Some researchers studied the extraction process of
oxidized sulfur-containing compounds but they did not study
the effects of all operating parameters in details. Ali et al. [25]
have studied the removal of DBT and its alkylated derivatives
as typical sulfur-containing compounds from a model solution
by oxidation and liquid/liquid extraction. Their oxidation system was H2O2-carboxylic acid, i.e., formic and acetic acid.
However, their study focused on the desulfurization of model
fuel rather than real middle distillate fuel oils.
Zannikos et al. [22] studied the extraction process of ODS
of two gas oil feedstocks with high sulfur content. They
detailed the effects of solvent type and solvent to fuel oil ratio,
but they did not investigate the effect of the number of extraction stages on the final desulfurization and fuel oil recovery.

www.cet-journal.com

Shiraishi et al. [6] studied the extractability of sulfones produced during the oxidation of sulfur-containing compounds
of three different fuel oils in the boiling point range of gas oil
by means of an acetonitrile/water azeotropic mixture. The
authors reported that the extractability of sulfones by means of
this solvent increases on increasing the carbon number of the
alkyl substituent of the sulfur-containing compounds. Their
study was comprehensive from fundamental aspects, but they
did not investigate the effects of solvent type and the number
of extraction stages on the final desulfurization and fuel oil
recovery.
The main aim of the present work was to study the effects of
the primary operating parameters, i.e., solvent type, solvent to
fuel oil ratio, and number of extraction stages, on the liquid/
liquid extraction of oxidized sulfur-containing compounds of
non-hydrotreated gas oil.

Experimental

2.1

Materials

Formic acid (> 99 % purity), acetonitrile, methanol, and hydrogen peroxide (30 wt %) were obtained from Merck Co.
(Germany). Ethanol (96 vol.-%) was provided by an Iranian
company, Taghtir Khorasan Co. (Khorasan, Iran). All of the
chemicals were used as received without any further treatment.
Non-hydrotreated gas oil as the feedstock with total sulfur
content of 7990 ppmw and with physical specifications as summarized in Tab. 1, was obtained from the Tehran Oil Refinery
Co. (Tehran, Iran).
Table 1. Properties of gas oil feedstock
Test Method

Value

Property

832

Density [kg m3]

3.98

Viscosity at 20 C [mPa s]

ASTM D-3228

3.3

Total nitrogen [ppmw]

ASTM D-4294

7990

Total Sulfur [ppmw]

0.063

Total acid number


[mg KOH g1]

ASTM D-86

2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Distillation range [C]


163

Initial Boiling Point


(I.B.P)

191

5%

208

10 %

252

30 %

289

50 %

316

70 %

355

90 %

372

95 %

379

Final boiling point (F.B.P)

Chem. Eng. Technol. 2010, 33, No. 9, 15151524

Desulfurization

2.2

Method of Analysis

2.3.2 Extraction Experiments

The total sulfur content of the untreated and treated gas oil
samples with total sulfur content higher than 990 ppmw was
determined by SLFA-20 sulfur in an oil analyzer (Horiba,
USA). The test method is based on ASTM D 4294. The total
sulfur content of the treated gas oil samples with total sulfur
content lower than 990 ppmw was determined using a more
precise analyzer, i.e., an FX-700 sulfur analyzer (Tanka, Japan).
This analyzer is suitable for a total sulfur detection range of
0990 ppmw, with a lower detection limit of 1 ppm. The test
method is based on ASTM D 2622. All total sulfur measurements were performed by the central laboratory of Esfahan Oil
Refinery Co. (Esfahan, Iran).
A Chrompack CP 9000 gas chromatograph equipped with a
flame photometric detector (GC-FPD) was used to compare
the chromatograms of sulfur-containing compounds present
in the untreated gas oil feedstock with those in the treated gas
oil. A fused-silica capillary column (30 m 0.32 mm 0.25 lm,
CP-Sil 8 CB obtained from Supelco Co.) was used for the
separation. The injector and detector temperatures were set
at 270 C and 300 C, respectively, while it was fed with
142 mL min1 of hydrogen, 75 mL min1 of air1 (the primary
air-H2 flame for combustion purposes), and 166 mL min1 of
air2 (the secondary air-H2 flame for photometric detection).
Moreover, nitrogen was used as the carrier gas with a constant
column flow rate of 0.5 mL min1.
An Agilent 6890 series GC system with Agilent 5973 network
mass selective detector and capillary column DB-1ms (100 %
dimethyl polysiloxane, 30 m 0.25 mm 25 lm) was used to
qualitatively detect sulfur-containing compounds of the gas oil
feedstock with an injector temperature set at 250 C. The initial temperature was set at 50 C for 5 min and then the temperature was increased to 275 C at a rate of 5 C min1 before
being maintained constant at 275 C for 10 min.

2.3

1517

Experimental Procedures

2.3.1 Oxidation Experiments


In each oxidation run, 250 mL of the untreated gas oil with a
total sulfur content of 7990 ppmw was introduced into a 1 L
three-necked glass reactor equipped with a condenser and a
thermometer. The reactor was then placed in a constant-temperature water bath and 78.5 mL of formic acid was added to
the reactor. The reaction mixture was continuously stirred and
heated to 60 C. Following this, 26.5 mL of hydrogen peroxide
was added to the reaction mixture. The reaction time was set
to 150 min. After this time, stirring of the reaction mixture
was stopped and the aqueous and hydrocarbon phases were
decanted in a separation funnel. The hydrocarbon phase, i.e.,
oxidized gas oil, was washed three times with 500 mL of distilled water in order to remove any entrained aqueous phase.
This procedure was repeated until ca. 5 L of oxidized gas oil
feedstock was obtained. This oxidized feedstock was used
throughout the extraction experiments.

In each extraction experiment, 30 mL of oxidized gas oil with


an appropriate amount of solvent according to the preset solvent/gas oil ratio were charged to a 250 mL flat-bottomed
flask. This flask was then placed in a constant-temperature
water bath and the mixture was vigorously stirred for 30 min
at 25 C. The dispersion formed was then allowed to separate
into two distinct phases, i.e., aqueous and hydrocarbon phases,
in a separation funnel. The aqueous phase was removed and
the hydrocarbon phase, i.e., treated gas oil, was washed twice
with distilled water and dried over magnesium sulfate. The
treated gas oil was analyzed for the total sulfur content measurements. It is important to note that for each data point, the
experimental run was repeated twice, and thus, each data point
was determined based on the mean value of at least two measurements with a standard deviation of 13 %.

Results and Discussion

As can be observed from the data in Tab. 1, the range of boiling points of the gas oil used in the present work is from 163
to 379 C. According to the literature [7], the main sulfur compounds of fuel oils in the range of gas oil boiling points are
alkylated benzothiophene, dibenzothiophene and its alkylated
derivatives. The GC-MS analysis confirmed the presence of
these compounds in the present gas oil feedstock. Moreover,
the data in Tab. 2 summarizes the sulfur-containing compounds detected in the present gas oil.

3.1

Desulfurization of Gas Oil in the Oxidation Stage

The chemistry involved in the ODS process is such that the divalent sulfur can be oxidized by the electrophilic addition reaction of oxygen atoms to form the hexavalent sulfur in a sulfone
functionality. The reactivity of different sulfur-containing
compounds in the present oxidation system is related to the
electron densities on the sulfur atom of these compounds. The
electron density on the sulfur atom of typical sulfur-containing
compounds present in the gas oil feedstock is high enough to
oxidize these compounds in the formic acid-hydrogen peroxide oxidation system [29]. Therefore, if the operating conditions of the oxidation system, i.e., temperature, hydrogen peroxide to sulfur molar ratio (O/S), formic acid to sulfur molar
ratio (Acid/S), and reaction time, are chosen appropriately, the
sulfur-containing compounds can be completely oxidized1). In
a previous study by the current authors [32], it was shown that
the appropriate operating conditions of ODS for the present
feedstock are a temperature of ca. 60 C, O/S = 5, and acid/
S = 40. Therefore, it was decided to apply these operating conditions in the current oxidation experiments. Moreover, the
run time of the oxidation experiments was set to 150 min.
Furthermore, it was found that with oxidation of the gas oil

1)

Chem. Eng. Technol. 2010, 33, No. 9, 15151524

List of symbols at the end of the paper.

2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.cet-journal.com

1518

M. A. Sobati et al.

Table 2. Different sulfur-containing compounds present in the


gas oil feedstock.
Sulfur-containing Compounds

MW

1,3-Propanedithiol

108

Methylphenyl sulfide

124

Benzothiophene

134

1,5-Pentanedithiol

136

1-Octanethiol

146

5-Methylbenzothiophene

148

3-Methylbenzothiophene

148

4-Methylthiophenol

156

2-Naphtalenethiol

160

3,5-Dimethyl benzothiophene

162

Di-tert-butyl disulfide

178

Dibenzothiophene

184

2,3,4,6-Tetramethyl benzotiophene

190

4-Methyl dibenzotiophene

198

3-Methyl dibenzotiophene

198

1-Methyl dibenzotiophene

198

8-Methylnaphtho[2,1-b]thiophene

198

1-Dodecanethiol

202

2,8-Dimethyl dibenzotiophene

212

1,6-Dimethyl dibenzotiophene

212

4,6-Dimethyl dibenzothiophene

212

3,4-Dimethyl dibenzotiophene

212

4,9-Dimethyl [2,3-b] napthotiophene

212

Thianthrene

214

1-Octadecanethiol

286

1-Docosanethiol

342

feedstock, ca. 39.7 % of the total sulfur is removed from the


original gas oil, i.e., the total sulfur content of the gas oil
decreases from 7990 to 4820 ppmw. This has been achieved by
simultaneous extraction of oxidized sulfur-containing compounds to the aqueous phase in the oxidation media, which is
a mixture of formic acid and hydrogen peroxide. The recovery
of gas oil after oxidation is 97 %. As a result of the oxidation
reaction, the polarity of the sulfur-containing compounds is
increased, and hence, their extractability in the polar aqueous
phase in the oxidation media increases. Therefore, oxidation
alone leads to significant levels of desulfurization, i.e., 39.7 %.
The main target of solvent extraction after oxidation is to
further increase the desulfurization by the extraction of the
remaining oxidized sulfur-containing compounds. In other
words, solvent extraction after oxidation is applied to decrease
the sulfur content of oxidized gas oil from 4820 ppmw to lower

www.cet-journal.com

values. The results in Fig. 1 show the GC-FPD chromatograms


of the gas oil before and after oxidation. As can be seen from
Fig. 1, the peaks of the sulfur-containing compounds in the
oxidized samples are shifted to higher retention times relative
to the untreated samples. In addition, the peak intensity of the
sulfur-containing compounds in the oxidized samples is lower
than that of the untreated samples due to the partial extraction
of oxidized sulfur-containing compounds by the aqueous
phase in the oxidation media. Moreover, as can be also observed, the application of solvent extraction after oxidation
leads to reduced sulfur-containing compounds peaks in the
treated gas oil.

3.2

Preliminary Solvent Selection

There are some guidelines for choosing a suitable solvent for


a liquid/liquid extraction process. However, it should be noted
that these guidelines may be conflicting. Therefore, compromises should be made between important factors affecting the
performance of extraction process. The main factors in solvent
selection that should be considered are selectivity, recoverability, density, interfacial tension, chemical reactivity, mutual solubility with the solution to be extracted, corrosiveness, viscosity, vapor pressure, freezing point, inflammability, toxicity,
and cost [37, 38]. Selectivity is the first criterion that should be
examined for the applicability of solvents in a separation process. The sulfur-containing compounds are more polar than
other accompanying hydrocarbon compounds, and hence, they
can be extracted by means of a polar solvent. However, polarity
alone is not sufficient to select a suitable solvent and the recovery of the gas oil is also seen to be important. From this point
of view, the appropriate solvent would extract the maximum
amount of desired components, i.e., oxidized sulfur-containing
compounds, and the minimum amount of undesired compounds, i.e., different hydrocarbon compounds present in the
gas oil. In the separation of oxidized sulfur-containing compounds, the range of boiling point of oxidized sulfur-containing compounds, i.e., sulfoxides and sulfones, is ca. 270670 C.
If the selected solvent has a high boiling point near the boiling
points of the oxidized sulfur-containing compounds, the solvent may not be recovered by distillation. Another property
that should be considered is the solvent viscosity. The application of low viscosity solvents in industrial-scale extraction
processes is advantageous because of a lower power-input
requirement and higher heat-and mass-transfer rates [38].
In the separation of the oxidized sulfur-containing compounds of gas oil, the solubility of gas oil in the solvent is very
important, because it determines the gas oil recovery after extraction. The dissolved hydrocarbon in the extraction solvent
can be recovered in a solvent recovery stage by distillation.
However, this results in additional costs, and hence, an appropriate solvent for the separation of oxidized sulfur-containing
compounds from gas oil is a solvent that results in the efficient
extraction of oxidized sulfur-containing compounds with a
low gas oil loss. The solvent effectiveness in the separation of
oxidized sulfur-containing compounds of middle distillates
can be easily evaluated by employing a solvent effectiveness
factor, E. This factor includes the effect of desulfurization and

2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Chem. Eng. Technol. 2010, 33, No. 9, 15151524

Desulfurization

1519

ly. In the present work, methanol,


96 vol.-% ethanol, acetonitrile, NMP,
and DMF were chosen in a pre-screening step as potential solvents. The data
in Tab. 3 summarizes the physical
properties of these solvents. As can be
seen from Tab. 3, the boiling points of
NMP and DMF are significantly higher
than those of the other solvents. Therefore, in the cases where NMP or DMF
are used as a solvent in the extraction
stage, the difference between the boiling point of the solvent and the oxidized sulfur-containing compounds is
small. Therefore, in such circumstances, the separation of oxidized sulfur-containing compounds from the
solvent in the solvent recovery stage
may be very difficult. Moreover, as can
be seen from Tab. 3, NMP and DMF
are more toxic than the other potential
solvents, and hence, are more difficult
to handle industrially. As a general
guideline, more toxic solvents should
be avoided whenever possible. Furthermore, Otsuki et al. [29] clearly reported the lowest fuel oil recovery for the
application of DMF among different
solvents for the separation of oxidized
sulfur-containing compounds from
SR-LGO and VGO. Therefore, NMP
and DMF were rejected by this preliminary solvent screening and the
performance capability of methanol,
96 vol.-% ethanol, and acetonitrile as
potential solvents for the extraction
of oxidized sulfur-containing compounds, was experimentally evaluated.
It is important to note that detailed investigations to determine the appropriate conditions such as optimum temperature for the desulfurization and
recovery of gas oil are the targets of
extended studies and will be reported
in the future.

Figure 1. GC-FPD chromatograms of untreated and treated gas oil: (a) Untreated gas oil,
(b) Gas oil after oxidation, and (c) Gas oil after oxidation followed by extraction. Extraction
conditions: Solvent; acetonitrile, T = 25 C, mixing time = 30 min, number of extraction
stages = 3, and solvent/gas oil ratio = 0.5.

middle distillate recovery after extraction and is normally


defined as in Eq. (1):
E = D/(100 R)

(1)

where D and R are the percentage of desulfurization and recovery of gas oil after oxidation followed by extraction, respective-

Chem. Eng. Technol. 2010, 33, No. 9, 15151524

3.3

Detailed Extraction Studies


for the Chosen Solvents

3.3.1 Effect of Solvent/Gas Oil Ratio

The solvent/gas oil ratio is an important factor in the extraction of oxidized sulfur-containing compounds. On the one
hand, an increase in the level of desulfurization can be expected by increasing the solvent/gas oil ratio. However, on the
other hand, an increase in the solvent/gas oil ratio leads to a
decrease in the gas oil recovery. The results in Figs. 24 demonstrate the effects of solvent/gas oil ratio on the desulfuriza-

2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.cet-journal.com

1520

M. A. Sobati et al.

Figure 2. Effect of solvent/gas oil ratio on the desulfurization


and recovery of gas oil. Extraction conditions: acetonitrile solvent, T = 25 C, mixing time = 30 min, and number of extraction
stages = 1.

Figure 3. Effect of solvent/gas oil ratio on the desulfurization


and recovery of gas oil. Extraction conditions: methanol solvent,
T = 25 C, mixing time = 30 min, and number of extraction stages
= 1.

Table 3. Physical properties of potential solvents for the extraction of oxidized sulfur-containing compounds.
Solvent

Density
[kg m3]

Boiling point
[C]

Freezing point
[C]

Viscosity at 20 C
[mPa s]

PELa (TWAb)

Methanol

790

64.7

97

0.59

200 ppm (OSHAc)

96 vol.-% Ethanol

790

78

113.9

1.07

1000 ppm (OSHAc)

Acetonitrile

780

81.6

43.8

0.37

40 ppm (OSHAc)

NMP

1028

203

24

1.7

10 ppm (HESISd)

DMF

950

153

61

0.92

10 ppm (OSHAd)

Permissible Exposure Limit (PEL): is the maximum amount or airborne concentration of a substance to which a worker may be legally
exposed. Most PELs have been defined for substances that are dangerous when inhaled, but some are for substances that are dangerous
when absorbed through the skin or eyes; b Time Weighted Average (TWA): This average concentration must not be exceeded during any
8 h work shift of a 40 h working week; c Occupational Safety and Health Administration (OSHA); d Hazard Evaluation System and Information Service (HESIS)

www.cet-journal.com

2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Chem. Eng. Technol. 2010, 33, No. 9, 15151524

Desulfurization

1521

3.3.2 Effect of Number of Extraction Stages


The number of extraction stages is an important factor in the
extraction of oxidized sulfur-containing compounds. Generally, an increase in the desulfurization level is expected with an
increase in the number of extraction stages due to an improvement in the extraction efficiency of oxidized sulfur-containing
compounds. In the present work, the effect of number of extraction stages on the extraction of oxidized sulfur-containing
compounds was investigated for acetonitrile, methanol, and
96 vol.-% ethanol. The data in Figs 57 show the effects of the
number of extraction stages on the desulfurization and recovery of gas oil for the chosen solvents. In each of Figs. 57, the
trends in desulfurization and recovery of gas oil versus the
number of extraction stages can be observed. As an example, it
can be observed from Fig. 5 for acetonitrile with a solvent/gas
oil ratio of 0.5, that an increase in the number of extraction
stages from 1 to 3 leads to an increase in the final desulfurization level from 84.9 % to 95.8 % and a decrease in the gas oil
recovery from 90 % to 83.3 %. A further increase in the num-

Figure 4. Effect of solvent/gas oil ratio on the desulfurization


and recovery of gas oil. Extraction conditions: 96 vol.-% ethanol
solvent, T = 25 C, mixing time = 30 min, and number of extraction stages = 1.

tion and the recovery of gas oil for acetonitrile, methanol, and
96 vol.-% ethanol, respectively. The trends of desulfurization
and the recovery of gas oil versus solvent to gas oil ratio are
shown in Figs 24. As an example, it can be observed from
Fig. 2 for acetonitrile that an increase in the solvent/gas oil
ratio from 0.25 to 2 results in an increase in the desulfurization
level from 82.3 % to 93.4 % and a decrease in the gas oil recovery from 93.3 % to 85 % . Moreover, with a further increase in
the solvent/gas oil ratio from 2 to 4, the final desulfurization
level does not improve significantly, i.e., only from 93.4 % to
96.4 %, but the gas oil recovery continuously decreases from
85 % to 78.3 %. The same trends are observed in Figs. 3 and 4
for methanol and 96 vol.-% ethanol, respectively.
In summary, by increasing the solvent/gas oil ratio for all
solvents, the increase in the desulfurization approaches a limit
asymptotically but the gas oil recovery continuously decreases
due to a continuous increase in the amount of dissolved gas
oil.

Chem. Eng. Technol. 2010, 33, No. 9, 15151524

Figure 5. Effect of number of extraction stages on the desulfurization and recovery of gas oil. Extraction conditions: acetonitrile
solvent, T = 25 C, mixing time = 30 min, and solvent/gas oil ratio = 0.5.

2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.cet-journal.com

1522

M. A. Sobati et al.

Figure 6. Effect of number of extraction stages on the desulfurization and recovery of gas oil. Extraction conditions: methanol
solvent, T = 25 C, mixing time = 30 min, and solvent/gas oil
ratio = 1.

ber of extraction stages from 3 to 5 leads to a small improvement in the final desulfurization, i.e., from 95.8 % to 98.1 %,
and a decrease in the gas oil recovery from 83.3 % to 80 %.
The same trends are observed for methanol and 96 vol.-% ethanol, as shown in Figs. 6 and 7, respectively.
In summary, for all three solvents, increasing the number of
extraction stages from 1 to 3 leads to a noticeable increase in
the desulfurization level. However, further increasing the extraction stages has no significant effect on the final desulfurization, while the gas oil recovery continuously decreases. The
solvent effectiveness factor, E, defined by Eq. (1) expresses both
the effects of desulfurization and the gas oil recovery for a solvent. The data in Fig. 8 shows the variation of E with the number of extraction stages for the three solvents. As can be seen
from Fig. 8, acetonitrile has the highest E factor, and therefore,
provides a better performance compared to the other two solvents. It should be noted that the applied solvent/gas oil ratio
for acetonitrile in each extraction stage is 0.5, while it is 1 for
methanol and 96 vol.-% ethanol. In other words, for the same
amount of desulfurization, the application of a lower amount

www.cet-journal.com

Figure 7. Effect of number of extraction stages on the desulfurization and recovery of gas oil. Extraction conditions: 96 vol.-%
ethanol solvent, T = 25 C, mixing time = 30 min, and solvent/
gas oil ratio = 1.

Figure 8. Solvent effectiveness factor, E, versus the number of


extraction stages for acetonitrile, methanol, and 96 vol.-% ethanol.

2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Chem. Eng. Technol. 2010, 33, No. 9, 15151524

Desulfurization

of acetonitrile leads to a higher gas oil recovery relative to


methanol and 96 vol.-% ethanol. Moreover, methanol has a
higher value of E, and therefore, a better performance compared to 96 vol.-% ethanol.

3.4

extraction of oxidized sulfur-containing compounds of gas


oil. The E values for the chosen solvents decreases in the
following order: acetonitrile > methanol > 96 vol.-% ethanol.
VI. The level of desulfurization of gas oil by single extraction
is much lower than that obtained by the ODS process.

Effect of Oxidation on the Desulfurization

In order to examine the role of oxidation in the ODS process,


the desulfurization of the gas oil feedstock by a single extraction process was compared with oxidation followed by solvent
extraction, i.e., the ODS process. The extraction conditions
were the same for all experimental runs, i.e., acetonitrile was
used as the solvent, solvent/gas oil ratio = 0.5, number of
extraction stages = 3, T = 25 C, and the mixing time = 30 min.
It was found that the single extraction process leads to 14.7 %
desulfurization, while the oxidation before the extraction process enhances the desulfurization to 95.8 %. In single extraction, the polarity of the sulfur-containing compounds is not
much higher than that of the corresponding hydrocarbon
compounds, and therefore, the extraction is not efficient.
However, in the ODS process, a major desulfurization of gas
oil is obtained by the extraction of oxidized sulfur-containing
compounds into the aqueous phase of oxidation media, and
thus, the polarities of the remaining oxidized sulfur-containing
compounds are changed in such a way that they can be extracted more effectively.

1523

Conclusions

The authors gratefully acknowledge the financial and technical


support provided by the R&D center of Esfahan Oil Refinery
Co. (Esfahan, Iran).
The authors have declared no conflict of interest.

Symbols used
D
E

[]
[]

[mol]

R
S

[]
[mol]

[min]

percentage of desulfurization
solvent effectiveness factor
defined by Eq. (1)
moles of oxidant (i.e., hydrogen
peroxide)
gas oil recovery after extraction
total moles of sulfur atoms
present in gas oil
time

References

In the present investigation, an attempt was made to carefully


study the extraction step after oxidation in the ODS process. It
was found that:
I. A large fraction of desulfurization is obtained during the
oxidation step, i.e., the total sulfur content of gas oil
reached 4820 ppmw, which represents over 39.7 % reduction in the total sulfur content of the gas oil.
II. Acetonitrile, methanol, and 96 vol.-% ethanol were chosen
as the potential solvents in a solvent prescreening step to
extract the remaining oxidized sulfur-containing compounds of gas oil. Over 99 % desulfurization was obtained
by the solvent extraction of oxidized sulfur-containing
compounds of the gas oil.
III. For all three solvents, with an increase in the solvent/gas
oil ratio, the desulfurization of the gas oil increases and
approaches a limit asymptotically. However, the gas oil
recovery continuously decreases due to a continuous increase of the amount of dissolved gas oil in the extraction
solvent.
IV. For all three solvents, the desulfurization of the gas oil
increases as the number of extraction stages increases, but
approaches a limit asymptotically for the number of extraction stages greater than 3. However, the gas oil recovery
decreases continuously due to an increase in the total
amount of dissolved gas oil in the extraction solvent.
V. The solvent effectiveness factor, E, which accounts for both
the desulfurization and recovery of the gas oil is an appropriate criterion for evaluating solvent performance in the

Chem. Eng. Technol. 2010, 33, No. 9, 15151524

Acknowledgment

[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]

Y. Horii et al., US Patent 5 494 572, 1996.


W. Gore, US Patent 6 160 193, 2000.
W. Gore, US Patent 6 274 785, 2001.
J. A. Kocal, T. A. Brandvold, US Patent 6 368 495, 2002.
M. J. Grassman et al., US Patent 5 910 440, 1999.
Y. Shiraishi, K. Tachibana, T. Hirai, I. Komasawa, Ind. Eng.
Chem. Res. 2002, 41, 4362.
Ch. Song, X. Ma, Appl. Catal., B 2003, 41, 207.
C. Song, X. Ma, Int. J. Green Eng. 2004, 1 (2), 167.
A. M. Aitani, M. F. Ali, H. H. Al-Ali, Pet. Sci. Technol. 2000,
18 (56), 537.
T. F. Yen, H. Mei, S. Lu, US Patent 6 402 939, 2002.
E. Ito, J. A. R. van Veen, Catal. Today 2006, 116, 446.
I. Babich, J. Moulijn, Fuel 2003, 82, 607.
M. Macaud et al., J. Catal. 2000, 193, 255.
P. S. Tam, J. R. Kittrell, J. W. Eldridge, Ind. Eng. Chem. Res.
1990, 29, 321.
P. S. Tam, J. R. Kittrell, J. W. Eldridge, Ind. Eng. Chem. Res.
1990, 29, 324.
A. Chica et al., Chem. Eur. J. 2006, 12, 1960.
A. Ishihara et al., Appl. Catal., A 2005, 279, 279.
A. Chica, A. Corma, M. E. Dmine, J. Catal. 2006, 242, 299.
V. V. D. N. Prasad et al., Catal. Commun. 2008, 9, 1966.
N. Y. Chan, T.-Y. Lin, T. F. Yen, Energy Fuels 2008, 22, 3326.
O. S. Nonaka et al., J. Jpn. Pet. Inst. 1999, 42, 315.
F. Zannikos, E. Lois, S. Stournas, Fuel Process. Technol. 1995,
42, 35.

2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.cet-journal.com

1524

M. A. Sobati et al.

[23] M. F. Ali et al., Fuel 2006, 85, 1354.


[24] M. F. Ali, A. Al-Malki, Sh. Ahmed, Fuel Process. Technol.
2009, 90 (4), 536.
[25] S. H. Ali, D. M. Hamad, B. H. Albusairi, M. A. Fahim, Energy
Fuels 2009, 23 (12), 5986.
[26] A. M. Dehkordi, Z. Kiaei, M. A. Sobati, Fuel Process. Technol.
2009, 90, 435.
[27] A. M. Dehkordi, M. A. Sobati, M. A. Nazem, Chin. J. Chem.
Eng. 2009, 17 (5), 869.
[28] M. Te, C. Fairbridge, Z. Ring, Appl. Catal., A 2001, 219, 267.
[29] Sh. Otsuki et al., Energy Fuels 2000, 14, 1232.
[30] L. Hao, S. Benxian, X. Zhou, Pet. Sci. Technol. 2005, 23, 991.
[31] L. Hao, S. Benxian, X. Zhou, Pet. Sci. Technol. 2006, 24, 1043.

www.cet-journal.com

[32] M. A. Nazem, M.Sc. Thesis, Sharif University of Technology,


Tehran 2008.
[33] F. M. Collins, A. R. Lucy, C. Sharp, J. Mol. Catal. A: Chem.
1997, 117, 397.
[34] H. Karaca, Z. Yildiz, Pet. Sci. Technol. 2005, 23, 285.
[35] Y. Hangun et al., J. Prepr. Am. Chem. Soc., Div. Pet. Chem.
2002, 47, 42.
[36] G. Yu, Sh. Lu, H. Chen, Z. Zhu, Carbon 2005, 43, 2285.
[37] J. D. Thornton, Science and Practice of Liquid-Liquid Extraction, Vol. 1, Oxford University Press USA, New York 1992.
[38] R. E. Treyball, Liquid Extraction, McGraw-Hill, New York
1952.

2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Chem. Eng. Technol. 2010, 33, No. 9, 15151524

You might also like