You are on page 1of 18

This article was downloaded by: [Rizalinda de Leon]

On: 21 January 2014, At: 10:49


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered
office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

International Journal of Green Energy


Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/ljge20

The Oxidative Desulfurization of Fuels


with a Transition Metal Catalyst: A
Comparative Assessment of Different
Mixing Techniques
a b c
Ming-Chun Lu , Luisa Cyd Charisse Biel , Meng-Wei Wan ,
d e
Rizalinda de Leon & Susan Arco
a
Department of Environmental Resources Management , Chia-Nan
University of Pharmacy and Science , Tainan , Taiwan
b
Environmental Engineering Graduate Program , University of
Philippines-Diliman , Quezon City , Philippines
c
Department of Environment Engineering and Science , Chia-Nan
University of Pharmacy and Science , Tainan , Taiwan, R.O.C.
d
Department of Chemical Engineering , University of Philippines-
Diliman , Quezon City , Philippines
e
Institute of Chemistry, University of Philippines-Diliman , Quezon
City , Philippines
Accepted author version posted online: 08 Aug 2013.Published
online: 17 Jan 2014.

To cite this article: Ming-Chun Lu , Luisa Cyd Charisse Biel , Meng-Wei Wan , Rizalinda de Leon
& Susan Arco (2014) The Oxidative Desulfurization of Fuels with a Transition Metal Catalyst: A
Comparative Assessment of Different Mixing Techniques, International Journal of Green Energy, 11:8,
833-848, DOI: 10.1080/15435075.2013.830260

To link to this article: http://dx.doi.org/10.1080/15435075.2013.830260

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the
Content) contained in the publications on our platform. However, Taylor & Francis,
our agents, and our licensors make no representations or warranties whatsoever as to
the accuracy, completeness, or suitability for any purpose of the Content. Any opinions
and views expressed in this publication are the opinions and views of the authors,
and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content
should not be relied upon and should be independently verified with primary sources
of information. Taylor and Francis shall not be liable for any losses, actions, claims,
proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or
howsoever caused arising directly or indirectly in connection with, in relation to or arising
out of the use of the Content.

This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden. Terms &
Conditions of access and use can be found at http://www.tandfonline.com/page/terms-
and-conditions
Downloaded by [Rizalinda de Leon] at 10:49 21 January 2014
International Journal of Green Energy, 11: 833848, 2014
Copyright Taylor & Francis Group, LLC
ISSN: 1543-5075 print / 1543-5083 online
DOI: 10.1080/15435075.2013.830260

THE OXIDATIVE DESULFURIZATION OF FUELS WITH A


TRANSITION METAL CATALYST: A COMPARATIVE
ASSESSMENT OF DIFFERENT MIXING TECHNIQUES

Ming-Chun Lu1 , Luisa Cyd Charisse Biel2 , Meng-Wei Wan3 ,


Rizalinda de Leon4 , and Susan Arco5
1
Department of Environmental Resources Management, Chia-Nan University of
Pharmacy and Science, Tainan, Taiwan
Downloaded by [Rizalinda de Leon] at 10:49 21 January 2014

2
Environmental Engineering Graduate Program, University of
Philippines-Diliman, Quezon City, Philippines
3
Department of Environment Engineering and Science, Chia-Nan University of
Pharmacy and Science, Tainan, Taiwan, R.O.C.
4
Department of Chemical Engineering, University of Philippines-Diliman,
Quezon City, Philippines
5
Institute of Chemistry, University of Philippines-Diliman, Quezon City,
Philippines

In this study, oxidative desulfurization of sulfur compounds using hydrogen peroxide,


phosphotungstic acid, and phase-transfer agent using high-intensity probe ultrasonication,
and high shear mixer was investigated. The effect of agitation speed (760014,000 rpm),
reaction temperature (5070 C), and treatment time (1030 min) on sulfur conversion was
examined and optimized using response surface methodology. A box-behnken design was
employed to determine the significance of various process parameters and their interac-
tions using analysis of variance. Analytical results for the model sulfur compounds by
ultrasound-assisted oxidative desulfurization and high shear showed comparable results.
Both treatment systems provided a 98% conversion of dibenzothiopene and benzothiopene
to polar sulfones at 70 C in 30 min. In both systems, the experimental data followed
the pseudo-first-order equation with activation energy of 60 kJ mol. Results indicate that
ultrasound energy produce greater reaction rates when compared to mixing with no sig-
nificant difference in activation energy. In addition, the advantages and drawbacks of
ultrasound-assisted extraction with respect to the high shear oxidative desulfurization were
discussed.

Keywords: Benzothiophene; Box-behnken design; Dibenzothiophene; High shear mixer;


Ultrasound-Assisted oxidative desulfurization

INTRODUCTION
Organic sulfur compounds (OSCs) in transportation fuels are a major source of
air pollution. Production of SOx and particulate matter during combustion is an environ-
mental concern due to their toxicity to humans and negative impact on the environment.

Address correspondence to Meng-Wei Wan, Ph.D., Department of Environment Engineering and Science,
Chia-Nan University of Pharmacy and Science, Tainan 71710, Taiwan. E-mail: peterwan@mail.chna.edu.tw

833
834 LU ET AL.

Emissions of SO2 have several anthropogenic sources such as excessive burning of fos-
sil fuels, petroleum refineries, chemical and coal burning power plants, and use of motor
vehicles. Therefore, it is important to produce fuels with low levels of OSCs in diesel and
gasoline (Song 2003; Song and Ma 2003). In addition, production of ultra-low sulfur fuels
could be utilized in the hydrocarbon fuel process for fuel cell applications (Song and Ma
2004).
Currently, oxidative desulfurization (ODS) is one of the technologies used in the
removal of OSCs such as benzothiophene, dibenzothiophene and their alkyl derivatives
from fuel (Riad and Mikhail 2012). In ODS, sulfur compounds are easily oxidized through
the electrophilic addition of oxygen atoms to sulfur, forming sulfoxides (1-oxides) and
sulfones (1,1-dioxides). Due to the strong affinity of sulfur toward oxygen, sulfides, and
thiophene derivatives can be readily converted without breaking any carboncarbon bonds
(Collins, Lucy, and Sharp 1997; Song and Ma 2003). The chemical and physical properties
of the more polar compounds are significantly different from those of hydrocarbons in fuel
Downloaded by [Rizalinda de Leon] at 10:49 21 January 2014

oil, and consequently they can be removed by distillation, solvent extraction, or adsorption
(Etemadi and Yen 2007).
Potential catalytic oxidative routes that could produce low sulfur fuels include vari-
ous types of oxidants such as molecular oxygen (Murata et al. 2003), superoxides (Chan,
Lin, and Yen 2008), hydrogen peroxide (Mei, Mei, and Yen 2003; Noyori, Aoki, and Sato
2003; Wan and Yen 2007), Fenton (Dai et al. 2008), and ozone (Wang, Zhao, and Li
2010). Among the different oxidants, hydrogen peroxide (H2 O2 ) is preferred due to its
low cost and high oxidizing efficiency (Zhang, Yu, and Wang 2008). Furthermore, the use
of H2 O2 in the oxidation of thiophene derivatives could occur in several catalytic systems
namely acetic acid (Yazu et al. 2003), formic acid (Herbstman and Patel 1982), solid bases
(Caero et al. 2005; Cedeno-Caero et al. 2008), phosphotungstic acid (Komintarachat and
Trakarnpruk 2006; Lai and Luo 2009; Trakarnpruk and Rujiraworawut 2009; Wan and Yen
2007; Te, Fairbride, and Ring 2001), potassium superoxide (Chan, Lin, and Yen 2008), and
chromium promoted sulfated zirconia (Kumar, Srivastava, and Badoni 2012).
One of the important considerations when combining H2 O2 with fuel is the formation
of a biphasic layer, which requires good dispersion between the aqueous and organic phases
to accelerate the reaction. Usually, the resulting fluid fluid interfacial area through which
molecular diffusion can take place is greatly enhanced by rapid mixing. With adequate
mixing, the organic phase can break into smaller droplets that could enhance mass transfer
of the oxidant and improve the rate of phase transfer catalysis reactions (Starks, Liotta, and
Halpern 1994; Adams, Dyson, and Tavener 2004). In a study by Huang et al. (2006), results
showed that increasing the mixing speed from 800 to 1770 rpm would cause formation
of more droplets, which provides greater interfacial area for better molecular diffusion,
thereby increasing the conversion of OSCs.
Mixing methods that are currently utilized for ODS are high shear mixer and
ultrasound-assisted oxidative desulfurization (UAOD). The high shear mixer, one of the
conventional methods, utilizes a magnetic stirrer (Huang et al. 2006; Wang, Zhu, and Ma
2009; Jose, Sengupta, and Basu 2011) that provides low sulfur conversion. In a study con-
ducted by Kuznetsova et al. (2007), the maximum conversion of OSCs was only 64% at 5 h.
Therefore, alternative oxidative techniques such as UAOD have been considered, where
additional energy is introduced in order to facilitate the transfer of analysts from sample
to solvent. The main advantage of UAOD is the higher dispersion that leads to increased
yield and or shorter reaction time (Mei, Mei, and Yen 2003; Wan and Yen 2007).
OXIDATIVE DESULFURIZATION OF FUELS 835

Ultrasonic radiation is a powerful tool for the acceleration of various steps in the
analytical process for solid and liquid samples (Priego-Lpez and Luque de Castro 2003;
Aydin, Tor, and Ozcan 2006; Weng et al. 2013). There are several advantages in using ultra-
sonication such as high oxidation efficiency and ease of operation. In ODS, the application
of ultrasonic radiation facilitates the emulsification phenomenon and accelerates the mass-
transfer process between two immiscible phases. This would improve the desulfurization
efficiency, where 97% conversion of benzothiophene was achieved in minimum time in the
study conducted by Wan and Yen (2006). Furthermore, Dai et al. (2008) investigated the
treatment of hydrotreated Middle East diesel fuel using an ultrasonic process at an energy
output of 50200 W. The results indicated that 91% of the sulfur compounds were oxi-
dized after 15 min at 150 W, and with increase in the applied ultrasonic energy further
enhances sulfur conversion. However, the main disadvantage is the difficulty for indus-
trial application where a great deal of energy would be utilized. In addition, the need of
additional devices such as amplifier, sono-reactor, and function generator would lead to
Downloaded by [Rizalinda de Leon] at 10:49 21 January 2014

increased capital costs. In order to commercialize the ODS process, it is essential to find
an alternative strategy in order to replace ultrasonication.
The main objective of the study is to compare the conventional oxidative technique
using the high shear mixer method against UAOD in terms of sulfur conversion. Before
comparing the two systems, the mixing method was optimized and the effect of operating
parameters such as temperature, mixing speed, and time on sulfur conversion was investi-
gated. An experimental factorial design methodology was applied to assess the statistically
important parameters affecting the conversion and for system optimization. Furthermore,
a kinetic model was applied to determine the rate of reaction and activation energy of the
two systems.

MATERIALS AND METHOD


Materials
All chemicals were of analytical grade and used without further purification. Toluene
and hydrogen peroxide (30 vol%) were obtained from Merck with quoted purities of
0.99 mass fractions. Tetraoctylammonium bromide and phosphotungstic acid were pur-
chased from Sigma-Aldrich with quoted purities of 0.99 mass fractions. Dibenzothiophene
(DBT) and benzothiophene (BT) were obtained from Acros. Diesel fuel with a sulfur
content of 1430 ppm was purchased from Kao Hsing Chang Company (Taiwan).
Stock solutions of BT and DBT were prepared within the range of 500 ppm in toluene
for each compound. The solutions prepared by dissolving appropriate amounts of BT and
DBT in toluene.

Experimental Method
High shear mixer oxidative desulfurization procedure. Experiments were
conducted in a glass reactor equipped with a T25 digital Ultra-turrax high shear mixer
with speed of up to 24,000 rpm. An electronic control unit was used to monitor the
temperature throughout the experiment. In an experimental run, an equal amount of
aqueous and organic solution was loaded in the reactor. Then, the mixture was stirred for
1030 min under varying temperature (5070 C) and agitation speed (740014,000 rpm).
After cooling, the mixture was centrifuged for 10 min to break the emulsion in order to
separate the organic phase.
836 LU ET AL.

Ultrasound-assisted oxidative desulfurization procedure. In this study, a


500-W sonicator (Sonic VCX 500, USA) equipped with a titanium probe tip (25 mm in
diameter and 122 mm in length) and operated at 20 kHz was used. The amplitude was
adjusted to 40% (200 W power output) without pulse length set-up. To maintain the tem-
perature, experiments were performed in a glass reactor immersed in a water bath. An equal
amount of the organic and aqueous phase was added into the reactor together with the
catalyst and phase transfer agent. To determine the kinetics of UAOD, the mixture was
irradiated for 20 min where 1 mL samples were taken at predetermined time intervals. The
aqueous phase was separated and the organic phase was kept for analysis.
Application to diesel fuel. The optimized conditions obtained from box-behken
model were applied in the desulfurization of diesel fuel. An equal amount of diesel fuel
and aqueous solution were loaded into the mixer, where the mixture was stirred using the
optimized conditions for agitation speed, temperature, and time.
Downloaded by [Rizalinda de Leon] at 10:49 21 January 2014

Instrument Analysis
The concentration of the reactants was determined using a gas chromatograph (GC,
Agilent 6890N, Agilent Technologies, Palo Alto, CA, USA) equipped with a fused-silica
capillary HP-5 ms column (30 m) having a thickness of 0.25 mm film (J & W Scientific,
Folsom, CA USA). All samples were filtered (0.45 nm, Nylaflo, PALL) prior to analysis.
The initial GC oven temperature was set to 100 C and was heated at the increasing rate
of 20 C min (run time: 10 min). Figures 1 and 2 show the analytical curves of BT, DBT,
DBTO, and BTO in toluene within the concentration range between 1 and 100 ppm. The
correlation coefficients (R2 ) for BT and DBT were 0.999 and 0.999, respectively.
Statistical analysis. Response surface methodology (RSM), specifically box-
behnken design, is a tool that investigates the simultaneous influence of different variables
and studies the optimization of a process response through variation of the input variables
(Bezerra et al. 2008; Myers and Montgomery 2002). One advantage of the RSM over the
conventional is that it accounts the interactions among different variables. For the con-
ventional optimization method, one variable is varied at a time while keeping the other
variables constant. However, the response can be a result of interactions among different
variables (Abramov 1994).
In the present work, a full experimental design methodology has been applied which
includes 17 runs, with each one of the independent variables taking two different val-
ues between a low (1) and a high (+1) level. The variables investigated are presented
in Table 1 and represent the agitation speed (X 1 ), reaction temperature (X 2 ), and treat-
ment time (X 3 ), respectively. Experimental results were analyzed using Design Expert
7.1.4 software (trial version) and a regression model was proposed. The measured and the
model predicted values of the response variable (Y) were used to compute the correlation
coefficient (R2 ).

RESULTS AND DISCUSSION


The results for the UAOD were already presented in the previous study conducted by
Wan and Yen, (2006). In their study, about 99% conversion of DBT and 97% BT conversion
were achieved in 10 min. In this study, only the optimization of the high shear mixer system
was carried out.
OXIDATIVE DESULFURIZATION OF FUELS 837

(a)
Downloaded by [Rizalinda de Leon] at 10:49 21 January 2014

(b)

Figure 1 Calibration curve for (a) BT and (b) DBT.

Experimental Design Methodology


The effects of three process variables that include speed (X 1 ), temperature (X 2 ), and
time (X 3 ) were studied during experimentation. The results of 17 runs using box-behnken
design (BBD) design are presented in Table 2 that include the design, experimental values
and predicted values, where the predicted values were obtained from quadratic model fit-
ting techniques using the Design Expert. BBD with three factors and three levels, including
five replicates at the center point, was used to fit a second-order response surface in order to
optimize the sulfur conversion. The five center point runs were carried out to measure the
process stability and inherent variability. The sulfur conversion was taken as the response
Y and could be expressed by the following second order polynomial equations:
838 LU ET AL.

(a)
Downloaded by [Rizalinda de Leon] at 10:49 21 January 2014

(b)

Figure 2 Calibration curve for (a) BTO and (b) DBTO.

Table 1 Levels of Independent Process Variables and Experimental Range

Independent variables Range and levels

1 0 +1

Agitation speed (rpm), X 1 7,400 10,000 14,000


Temperature (Co ), X 2 50 60 70
Time (min), X 3 10 20 30

Y = 62.69 + 2.47X1 + 26.17X2 + 21.17X3 + 1.67X1 2 + 4.31X1 3 0.21X2 3


+ 1.01X1 2 11.95X2 2 + 2.09X3 2 (1)

Where X 1 denotes agitation speed, rpm; X 2 denotes temperature, C; and X 3 denotes


treatment time, min.
OXIDATIVE DESULFURIZATION OF FUELS 839

Table 2 Experimental Design Matrix and Response

Coded values Real values Response

Std x1 x2 x3 X1 X2 X3 Observed (Y)% Predicted (Y)%

1 1 1 0 7400.00 50 20 19.25 24.78


2 1 1 0 14, 000.00 50 20 26.07 26.38
3 1 1 0 7, 400.00 70 20 74.07 73.76
4 1 1 0 14, 000.00 70 20 87.59 82.06
5 0 0 1 7, 400.00 60 10 49.62 46.46
6 0 0 1 14, 000.00 60 10 40.72 42.78
7 0 0 1 7, 400.00 60 30 82.24 80.18
8 0 0 1 14, 000.00 60 30 90.59 93.75
9 1 1 1 10, 000.00 50 10 8.07 5.70
10 1 1 1 10, 000.00 70 10 54.14 57.61
11 1 1 1 10, 000.00 50 30 51.08 47.62
Downloaded by [Rizalinda de Leon] at 10:49 21 January 2014

12 1 1 1 10, 000.00 70 30 98.00 100.37


13 0 0 0 10, 000.00 60 20 53.52 62.69
14 0 0 0 10, 000.00 60 20 65.54 62.69
15 0 0 0 10, 000.00 60 20 64.96 62.69
16 0 0 0 10, 000.00 60 20 64.92 62.69
17 0 0 0 10, 000.00 60 20 64.51 62.69

Table 3 Analysis of Variance (ANOVA) for Response Surface Quadratic Model

Source Sum of squares df Mean square F value Prob > F

Model 9, 809.72 9 1, 089.97 33.07 < 0.0001


X1 48.94 1 48.94 1.48 0.2625
X2 5, 477.33 1 5, 477.33 166.17 < 0.0001
X3 3, 585.65 1 3, 585.65 108.78 < 0.0001
X 1 2 11.19 1 11.19 0.34 0.5784
X 1 3 74.44 1 74.44 2.26 0.1766
X 2 3 0.18 1 0.18 5.34E 03 0.9438
X1 2 4.31 1 4.31 0.13 0.7283
X2 2 602.06 1 602.06 18.26 0.0037
X3 2 18.42 1 18.42 0.56 0.4791
Residual 230.74 7 32.96
Lack of fit 125.14 3 41.71 1.58 0.3264
Pure error 105.6 4 26.4
Total 10, 040.46 16

Statistical analysis. The response surface quadratic model using ANOVA is pre-
sented in Table 3. The significance of each coefficient and the interaction strength between
each independent variable was determined using p value. A small p value denotes higher
significance of the corresponding coefficient, where values lower than 0.05 are considered
to be significant. The significance of the F value depends on the number of degrees of
freedom (DF) in the model at 95% confidence level. The determination coefficient (R2 =
0.9770) was computed using the quadratic regression model from ANOVA. The value
of the adjusted determination coefficient (adjusted R2 = 0.9475) also confirmed that the
model was highly significant with regards to sulfur conversion. At the same time, the model
was found to be adequate for prediction within the range of experimental variables, with
the predicted determination coefficient (predicted R2 = 0.7841). The lack of fit F value
of 1.52 implied that the lack of fit was not significant relative to the pure error, and there
840 LU ET AL.

is a 32.64 % chance that this value could occur due to noise (Table 3). Thus, this model
was found to be adequate to navigate the design space and further optimization on sulfur
conversion can be carried out. In addition, the temperature (X 2 ) and time (X 3 ) were found
to be extremely significant (p < 0.0001) in sulfur conversion.
Based from Figure 3, the actual and predicted percent values are in good agreement,
which indicates the goodness of fit of the model.
The internally studentized residuals are presented in Figure 4. The residuals deter-
mine the normality and response transformation of the model. In this study, most of the
data points lie along the line, which indicates that the model is an acceptable fit with no
apparent problem in normality and response transformation.
Downloaded by [Rizalinda de Leon] at 10:49 21 January 2014

Figure 3 Plot of the actual and predicted value of sulfur conversion (%).

Figure 4 Internally studentized residuals and normal% probability plot.


OXIDATIVE DESULFURIZATION OF FUELS 841

Analysis of responses. To investigate the effect of time, temperature, and agita-


tion speed on the conversion of BT, RSM was used and 3-D plots were drawn. Based on
ANOVA, both time and temperature were found to have significant effect on sulfur conver-
sion, where agitation speed provided the least effect on the response. The quadratic effects
of X 1 and X 2 as well as the interaction effects between X 1 2 , X 1 3 , and, X 2 3 were con-
sidered moderate. The conversion of sulfur response surface graphs is shown in Figure 5.

The 3-D response surface plot of the combined effect of temperature and time was
illustrated in Figure 5a. A maximum sulfur conversion of >98 % was achieved after 30 min

(a)
Downloaded by [Rizalinda de Leon] at 10:49 21 January 2014

(b)

(c)

Figure 5 Surface Plots on the effect of on the desulfurization efficiency (a) temperature and time; (b) temperature
and agitation speed; and (c) time and agitation speed.
842 LU ET AL.

at 70 C. A sharp increase in the percentage removal was observed with increase in reaction
time. After 20 min, only a gradual increase in the percentage conversion could be seen.
The interactive effect of temperature and agitation speed on percent conversion
of sulfur is shown in Figure 5b. It shows that sulfur conversion increases with increas-
ing temperature, where the rate of reaction increases by lowering the interfacial tension.
In addition, further increase in agitation speed has little effect on the sulfur conversion. The
maximum sulfur conversion of >95% was measured at 70 C.
Figure 5c shows the 3-D response surface of the combined effect of time and tem-
perature on sulfur conversion. An increase in the reaction time from 10 to 30 min caused
an increase in sulfur conversion efficiency, which is due to longer interaction between
the aqueous and organic phase. From the figure, increasing the speed has little effect on the
conversion of sulfur, which implies that the maximum speed has been attained. In addition,
a maximum conversion of >93% was obtained after 30 min of treatment.
Optimization. The optimum experimental conditions in the conversion of sulfur
Downloaded by [Rizalinda de Leon] at 10:49 21 January 2014

compounds were estimated by solving the regression equation and analyzing the response
surface plots. Using the Design Expert software, the predicted sulfur conversion of 98%
could be achieved using an agitation speed of 10,000 rpm, temperature of 70 C and 30 min
reaction time. The optimum conditions were applied to three independent replicates to
verify the prediction from the model. The mean experimental sulfur conversion of 971%
validates that the response model is adequate for the optimization study.

Comparison of the Different Mixing Techniques


Sulfur conversion and selectivity. Figure 6 shows the effect of ultrasonic irra-
diation and mixer method on the sulfur conversion. From the graph, both systems attained
99% conversion of DBT at 10 min and 98% conversion of BT at 30 min. However,
ultrasonicator provided a faster reaction rate where 90 % conversion of BT in 10 min was
achieved in comparison to the mixer with only 60% conversion. This is due to the efficient
mixing between the liquid phases, resulting to a more enhanced interfacial area.
The effect of mixing such as mechanical stirring and 20 kHz ultrasonication on the
dispersion of droplets was examined. Applying ultrasonication created very fine droplets,
leading to an emulsion-like dispersion of the two phases. It was observed that ultrasonicator
caused better emulsification as compared to the high shear mixer, in which the mixture
provided a milky white consistency throughout the experiment. Mei, Mei, and Yen (2003)
have reported similar observations regarding the application of ultrasonicator in a biphasic
system that contained one viscous liquid.
The use of ultrasound provided a better performance, which is attributed to acoustic
cavitation. Acoustic cavitation creates a unique interaction of energy and matter through
the formation, growth, pulsating, and implosive collapse of microscopic bubbles during
ultrasonic irradiation (Casadonte 2000). In the process, a sinusoidal pressure variation of
alternating rarefaction (negative pressure) and compression (positive pressure) is imposed
on the liquid leading to a highly mixed system.
Figure 7 illustrates the effect of temperature on the conversion and selectivity using
a high shear mixer. BT was chosen as the sulfur compound because it is more difficult
to oxidize than DBT. As discussed in the previous study, the use of ultrasonicator showed
high selectivity towards the sulfone compound. Similar results were obtained, where a high
selectivity value of 96% was achieved in using the high shear mixer. It was also determined
that increasing the temperature causes an increase in sulfur conversion with no significant
OXIDATIVE DESULFURIZATION OF FUELS 843
Downloaded by [Rizalinda de Leon] at 10:49 21 January 2014

Figure 6 Comparison on the sulfur conversion between ultrasonicator and mixer.

500 700
BT at 70 C
450
BT at 60 C
600
BT at 50 C
400 BTO at 50 C
BTO at 60 C
500

BTO Concentration (ppm)


350
BT Concentration (ppm)

BTO at 70 C

300
400
250
300
200

150 200

100
100
50

0 0
0 5 10 15 20 25 30
Time (min)

Figure 7 Effect of temperature on the desulfurization efficiency of Benzothiophene.

effect on selectivity. From the result, the use of high shear mixer in the desulfurization of
DBT and BT is comparable to that of ultrasonicator.
Table 4 shows the summary of the advantages and disadvantages of the two mixing
systems, mainly UAOD and the mixing-assisted ODS.
Kinetics of the system. The kinetics of BT was examined since it is more diffi-
cult to oxidize than DBT. In determining the rate of reaction for a biphasic system using
844 LU ET AL.

Table 4 Rate Constants for Benzothiophene

Process Reaction temp ( C) Rate constant (min1 ) Correlation coefficient, R2

Mixer 50 0.0251 0.982


60 0.0448 0.980
70 0.0938 0.993
Ultrasonicator 50 0.0389 0.995
60 0.0681 0.982
70 0.1450 0.991

BT (C8 H6 S), the same process as a single-phase system was applied. The only difference
was the concentration taken into consideration was not the total concentration of the whole
system but the concentration of the reactants and products in the phase where the reaction
took place (Huang et al. 2006). Thus the reaction can be written as:
Downloaded by [Rizalinda de Leon] at 10:49 21 January 2014

k
C8 H6 S + H 2 O2 C8 H6 O2 S + H 2 O (2)

Based from Equation (2), the rate equation for BT is presented below in Equation (3).
 
rBT = k[CBT ] CH2 O2 (3)

where and are the order of reaction with respect to the concentration of BT and H2 O2 ,
respectively.
For simplification, the amount of H2 O2 was in excess such that the change in concen-
tration of H2 O2 compared to BT was negligible. With this assumption, the reaction data
can be fitted to pseudo-first order. Therefore, the rate constant and reaction time (t) can be
expressed using the following equations

rBT = k [CBT ] (4)

where the apparent rate constant (k) is described as:


 
k = k CH2 O2 (5)

Then integrating equation 4

ln(Co /Ct ) = k t (6)

where C0 and Ct are the initial sulfur concentration and equilibrium sulfur concentration
(mg l) at time t, respectively.
A plot of Equation (6) versus reaction time (t) displayed a linear relationship that
confirmed the goodness of fit of the pseudo-first-order reaction kinetics for mixer and
ultrasonicator. Table 5 lists the kinetic rate constants and correlation coefficients of BT
using ultrasound and mixer methods. The apparent rate constant obtained from ultrasound
as the mixing technique, was higher in comparison to that of the mixer. Wherein, the appar-
ent rate constants at 70 C for ultrasonicator and mixer were 0.1450 and 0.0938 min1 ,
respectively. This explains the high percentage conversion of sulfur in short amount of time
OXIDATIVE DESULFURIZATION OF FUELS 845

Table 5 Rate Constants for Benzothiophene

Process Reaction temp ( C) Rate constant (min1 ) Correlation coefficient, R2

Mixer 50 0.0251 0.982


60 0.0448 0.980
70 0.0938 0.993
Ultrasonicator 50 0.0389 0.995
60 0.0681 0.982
70 0.1450 0.991

1.0 Mixer
Ultrasonicator
1.5
Downloaded by [Rizalinda de Leon] at 10:49 21 January 2014

y = 7281.1x + 19.428
2.0
R2 = 0.989

2.5
ln k

3.0
y = 7295.1x + 18.86
3.5 R2 = 0.9924

4.0

4.5
0.00290 0.00295 0.00300 0.00305 0.00310
1/T (K1)

Figure 8 Arrhenius Plot for benzothiophene.

was obtained when using ultrasonicator since a high value of rate constant, the faster is the
reaction. Figure 8 shows the plot of ln k versus 1 T for both systems, where the apparent
activation energy of pseudo-first-order reaction for BT was determined to be 60 kJ mol.
This was derived from Equation (7):

k = AeEa/RT (7)

where k is the rate constant; Ea is the activation energry; A is the preexponential factor; T
is temperature; and R is the universal gas constant. This implies that ultrasonicator does not
decrease the activation energy instead it enhances the mass transfer, causing an increase in
the rate of sulfur conversion.

Application to Real Diesel Fuel


After optimization, desulfurization of diesel fuel with 1430 ppm sulfur content was
performed using a mixer system under the following optimized conditions: 10,000 rpm
agitation speed, 70 C temperature, and 30 min reaction time. Table 6 shows the two dif-
ferent removal methods for sulfones, which are extraction and adsorption. The final sulfur
846 LU ET AL.

Table 6 Removal efficiency of the mixing assisted desulfurization process applied in actual diesel fuel

Initial sulfur concentration (ppm) Final sulfur concentration (ppm) Sulfones removal method

1430 216 Extraction


1430 ND Adsorption
ND = Not detected.

concentration of the diesel was reduced to 216 ppm, with sulfur conversion of 84.90%
using the extraction method, which is lower than the 98% sulfur conversion achieved using
500 ppm BT model compound. However, the adsorption method was able to attain 100%
sulfone removal.
Downloaded by [Rizalinda de Leon] at 10:49 21 January 2014

CONCLUSION
The effect of agitation speed (7400 to 14,000 rpm), temperature (5070 C), and time
(1030 min) on the sulfur conversion was determined using RSM with the three variables
box-behnken design. The regression analysis, statistical significance and response surface
were carried out using Design Expert Software in predicting the responses in all experimen-
tal regions. Models were developed to correlate the adsorption variables to the responses.
Through analysis of the response surfaces, temperature, and time were found to have the
most significant effect on the conversion. Process optimization was carried out and the
experimental values obtained for the sulfur conversion are found to agree satisfactorily
with the values predicted by the models. The optimal condition of 98% sulfur conversion
was attained under an agitation speed, temperature and time of l0,000 rpm, 70 C, and
30 min, respectively.
Using the optimized condition, the efficiency of using a high shear mixer was
evaluated and compared with the ultrasonicator. Ultrasonic irradiation provides effective
emulsification of the mixture, resulting in faster conversion when compared to the high
shear mixer. Although the use of ultrasound in this process would enhance the treatment
efficiency, it also increases the energy consumption. Moreover, both systems attained 99 %
conversion for DBT and 98% conversion for BT, respectively. This study demonstrated the
viability of employing high shear mixer in ODS. Finally, the kinetics of BT in both systems
indicated that degradation of sulfur follows the pseudo first order reaction with activation
energy of 60 kJ mol.

FUNDING
The authors would like to acknowledge the National Science Council, Taiwan (NSC
101-2221-E-041-010-MY3) and Engineering Research and Development for Technology,
Philippines for the financial support.

REFERENCES
Adams, D., P. Dyson, and S. Tavener. 2004. Chemistry in alternative reaction media. Chichester:
John Wiley & Sons, Ltd.
OXIDATIVE DESULFURIZATION OF FUELS 847

Aydin, M.E., A. Tor, and S. Ozcan. 2006. Determination of selected polychlorinated biphenyls
in soil by miniaturized ultrasonic solvent extraction and gas chromatography-mass-selective
detection. Analytica Chimica Acta 577:23237.
Bezerra, M.A., R.E. Santelli, E.P. Oliveira, L.S. Villar, and L.A. Escaleira. 2008. Response surface
methodology (RSM) as a tool for optimization in analytical chemistry. Talanta 76:96577.
Caero, L.C., E. Hernandez, F. Pedraza, and F. Murrieta. 2005. Oxidative desulfurization of synthetic
diesel using supported catalysts Part I. Study of the operation conditions with a vanadium
oxide based catalyst. Catalysis Today 107108:56469.
Casadonte, D.J. 2000. The Sound of Science: The chemical effects of high-intensity ultrasound.
Lubbock Magazine 4043.
Cedeno-Caero, L., H. Gomez-Bernal, A. Fraustro-Cuevas, H.D. Guerra-Gomez, and R. Cuevas-
Garcia. 2008. Oxidative desulfurization of synthetic diesel using supported catalysts (III)
Support effect on vanadium-based catalysts. Catalysis Today 133135:24454.
Chan, N.Y., T. Lin, and T.F. Yen. 2008. Superoxides: Alternative oxidants for the oxidative
desulfurization process. Energy Fuels 22:332628.
Downloaded by [Rizalinda de Leon] at 10:49 21 January 2014

Collins, F., M. Lucy, and C. Sharp. 1997. Oxidative desulphurisation of oils via hydrogen peroxide
and heteropolyanion catalysis. Journal of Molecular Catalysis A: Chemical 117:397403.
Dai, Y., Y. Qia, D. Zhao, and H. Zhang. 2008. An oxidative desulfurization method using
ultrasound Fentons reagent for obtaining low and or ultra-low sulfur diesel fuel. Fuel
Processing Technology 89:92732.
Etemadi, O., and T.F. Yen. 2007. Aspects of selective adsorption among oxidized sulfur compounds
in fossil fuels. Energy Fuels 21:162227.
Herbstman, S., and J. Patel. 1992. Oxidative desulfurization of residual oils. Division Petroleum
Chemistry Preprint 27:826.
Huang, D., Y.J. Wang, L.M. Yang, and G.S. Luo. 2006. Chemical oxidation of dibenzothiophene with
a directly combined amphiphilic catalyst for deep desulfurization. Industrial and Engineering
Chemistry Research 45:188085.
Jose, N., S. Sengupta, and J.K. Basu. 2011. Optimization of oxidative desulfurization of thiophene
using CU titanium silicate-1 by box-behnken design. Fuel 90:62632.
Komintarachat, C., and W. Trakarnpruk. 2006. Oxidative desulfurization using polyoxometalates.
Industrial and Engineering Chemistry Research 45:185356.
Kumar, S., V.C. Srivastava, and R.P. Badoni. 2012. Oxidative desulfurization by chromium promoted
sulfated zirconia. Fuel Processing Technology 93:1825.
Kuznetsova, L.I., L.G. Detusheva, N.I. Kuznetsov, V.K. Duplyakin, and V.A. Likholobov. 2008.
Liquid-Phase oxidation of benzothiophene by cumyl hydroperoxide in the presence of
catalysts based on supported metal oxides. Kinetics and Catalysis 49(5):64452.
Lai, J., and G. Luo. 2009. Zinc-substituted polyoxometalate for oxidative desulfurization of
dibenzothiophene. Journal of Petroleum Science and Technology 27:78187.
Mei, H., B.W. Mei, and T.F. Yen. 2003. A new method for obtaining ultra-low sulfur diesel fuel via
ultrasound assisted oxidative desulfurization. Fuel 82:40514.
Murata, S., K. Murata, K. Kidena, and M. Nomura. 2003. Oxidative desulfurization by molecular
oxygen. Division of Fuel Chemistry 48(2):531.
Myers, R.H., and D.C. Montgomery. 2002. Response surface methodology: Process and product
optimization using designed experiments. New York: Wiley.
Noyori, R., M. Aoki, and K. Sato. 2003. Green oxidation with aqueous hydrogen peroxide. Chemistry
Communications 197786.
Priego-Lpez, E., and M.D. Luque de Castro. 2003. Ultrasound-assisted derivatization of phenolic
compounds in spiked water samples before pervaporation, gas chromatographic separation
and flame ionization detection. Chromatographia 57:51318.
Riad, M., and S. Mikhail, 2012. Oxidative desulfurization of light gas oil using zinc catalysts
prepared via different techniques. Catalysis Science & Technology 2:143746.
848 LU ET AL.

Starks, C.M., C.L. Liotta, and M. Halpern. 1994. Phase-transfer catalysis: fundamentals, applica-
tions, and industrial perspective. New York: Chapman and Hall, Inc.
Song, C. 2003. An overview of new approaches to deep desulfurization for ultra-clean gasoline,
diesel fuel and jet fuel. Catalysis Today 86:21163.
Song, C., and X. Ma. 2003. New design approaches to ultra-clean diesel fuels by deep desulfurization
and deep dearomatization. Applied Catalysis B: Environmental 41:207.
Song, C., and X. Ma. 2004. Ultra-deep desulfurization of liquid hydrocarbon fuels: Chemistry and
process. International Journal of Green Energy 1(2):16791.
Te, M., C. Fairbride, and Z. Ring. 2001. Oxidation reactivities of dibenzothiophenes in
polyoxometalate H2O2 and formic acid H2O2 systems. Applied Catalysis A: General
219:267.
Trakarnpruk, W., and K. Rujiraworawut. 2009. Oxidative desulfurization of gas oil by
Polyoxometalates Catalysts. Fuel Processing Technology 90:41114.
Wan, M.W., and T.F. Yen. 2007. Enhance efficiency of tetraoctylammonium fluoride applied to
ultrasound-assisted oxidative desulfurization (UAOD) process. Applied Catalysis A: General
Downloaded by [Rizalinda de Leon] at 10:49 21 January 2014

319:23745.
Wang, J., D. Zhao, and K. Li. 2010. Oxidative desulfurization of dibenzothiophene using ozone and
hydrogen peroxide in ionic liquid. Energy Fuels 24(4):252729.
Wang, B., J. Zhu, and H. Ma. 2009. Desulfurization from thiophene by SO4 2 ZrO2 catalytic
oxidation at room temperature and atmospheric pressure. Journal of Hazardous Materials
164:25664.
Weng, C. H., Y.T. Lin, C.K. Chang, and N. Liu. 2013. Decolourization of direct blue
15 by Fenton ultrasonic process using a zero-valent iron aggregate catalyst, Ultrasonics
Sonochemistry, DOI: 10.1016 j.ultsonch.2012.09.014
Yazu, K., T. Furuya, K. Miki, and K. Ukegawa. 2003. Tungstophophoric acid-catalyzed oxidative
desulfurization of light oil with hydrogen peroxide in a light oil acetic acid biphasic system.
Chemistry Letters 32:920.
Zhang, G., F. Yu, and R. Wang. 2008. Research advances in oxidative desulfurization technologies
for the production of low sulfur fuel oils. Petroleum and Coal 51(3):196207.

The author has requested enhancement of the downloaded file. All in-text references underlined in blue are linked to publications

You might also like