Home Oxidation of dibenzothiophene using the heterogeneous catalyst of tungsten-based carbon nanotubes
Article Open Access

Oxidation of dibenzothiophene using the heterogeneous catalyst of tungsten-based carbon nanotubes

  • Nguyen Duc Vu Quyen EMAIL logo , Tran Ngoc Tuyen , Dinh Quang Khieu , Ho Van Minh Hai , Dang Xuan Tin and Kiyoshi Itatani
Published/Copyright: May 16, 2018
Become an author with De Gruyter Brill

Abstract

Highly effective tungsten-based carbon nanotubes (W/CNT) were synthesized and used as a heterogeneous catalyst for the oxidation of dibenzothiophene (DBT) with the oxidant H2O2. The obtained materials were characterized by modern methods. The Langmuir–Hinshelwood kinetics model described the precursor mechanism of the reaction well through an intermediate compound. The low activation energy showed that the reaction was mainly controlled by diffusion. The positive activation enthalpy proved the endothermic nature of the activation process, and this process did not alter the inside structure of the catalyst. The thermodynamic parameters of the reaction were determined, which implied that the oxidation was endothermic and spontaneous at 303 K. The more negative values of the Gibbs free energy from 283 to 323 K confirmed that the reaction was more favorable at high temperatures. The stability and activity of catalyst were retained after three reaction cycles.

1 Introduction

Petroleum constitutes one of the most important fossil fuels of modern society, which is used to generate electricity, for transportation, and for industrial productions. However, the emission of sulfur compounds in petroleum, which mainly consist of thiols, sulfides, disulfides, and thiophenes, are undesirable during the utilization of petroleum because the combustion of these compounds releases sulfur oxides and oxyacids which are not only toxic to humans and contribute to environmental pollution, such as acid rain, but also cause corrosion to parts of internal combustion engines [1], [2]. Dibenzothiophene (DBT) is one of sulfur compounds that is difficult to remove from diesel by conventional desulfurization processes [3]. Therefore, the removal of sulfur compounds from the fuel has aroused the interest of many scientists.

Hydrodesulfurization (HDS) is a catalytic chemical process that is widely used to remove sulfur. This process, however, was limited for treating DBTs, especially those having alkyl substituents on their 4- and/or 6-position [4]. Thus, the petroleum with a low sulfur level requires critical conditions (high pressure and temperature) and special active catalysts [5]. New methods of reduction of the sulfur content in petroleum are needed to meet future legislations on the sulfur content, e.g. <10 ppm [6]. Oxidative desulfurization (ODS) is considered to be one of the promising methods for deep desulfurization of petroleum. In comparison with HDS, ODS can be carried out at room temperature and under atmospheric pressure. By this method, the sulfur compounds will be oxidized into stable and strongly polar compounds, such as sulfone or sulfoxide, that can be separated from petroleum by extraction into a polar solvent [7], [8], [9], [10], [11]. The catalyst used for ODS can be homogeneous such as HCOOH [12] or Na2WO4 [13], or heterogeneous such as Cu/TS-1 [14], a heteropolyacid (HPA) catalyst such as H3PW12O40 (HPW) [15], and, especially, tungsten oxide supported on carriers [3], [16], [17], [18] that shows high effectiveness. With different carriers, the catalytic ability of tungsten oxide can be different based on the interaction between the catalyst and the carrier. In many studies, silica-based carriers were employed, such as SiO2, MCM-41, MIL-101, so on, but thiophene removal was not high at room temperature (~70%) [3], [16], [17], [18].

Nanocarbon is well known as an excellent adsorbent not only of heavy metals and organic pigments but also for sulfur in petroleum when this material is made passive by poly(ethylene glycol) (PEG-200), as shown by Fallah et al. [19]. Nevertheless, the maximum sulfur conversion was ~80% even at high temperature. Zhang et al. [20] employed carbon nanotubes (CNTs) as a good catalyst for DBT conversion only at high temperature (150°C) and with a strong oxidant (oxygen with a flow rate of 150 ml min−1). In development of using CNTs to desulfurize petroleum, no mention has been made of any studies of their use as a carrier for the dispersion of tungsten oxide. The important advantages of CNTs are their high surface area and electron-transport ability, which contribute to the creation of many active catalyst sites. Besides, sulfur-containing aromatic compounds such as DBT can be strongly attracted toward the surface of CNTs because of π–π stacking, which is one of the reasons why the catalyst exhibits high activity for the oxidation of DBT.

In this work, a new material, tungsten dispersed on CNTs (W/CNT), exhibiting high catalytic efficiency for DBT removal in petroleum, was synthesized. The kinetic and thermodynamic studies of DBT conversion using W/CNT catalyst were also carried out.

2 Materials and methods

2.1 Materials

The CNTs were synthesized from liquid petroleum gas (LPG) by chemical vapor deposition (CVD) using Fe/Al2O3 as the catalyst. The tubes of the material were uniform with size in the range 30–50 nm (Figure 1).

Figure 1: SEM image of CNTs.
Figure 1:

SEM image of CNTs.

The resulting CNTs were dispersed in 0.1 m H2SO4 solution by ultrasonic stirring for 30 min, and then Na2WO4 was added to the CNT suspension again under ultrasonication. The W/CNT material was obtained by drying the above mixture at 80°C until its weight remained unchanged and after heating the obtained solid at 500°C for 1 h.

2.2 Methods

2.2.1 Characterization of CNTs and W/CNT:

The phases of the obtained materials were studied using an X-ray diffractometer (XRD) (model RINT2000/PC; Rigaku, Tokyo, Japan) with λCuKα=0.154 nm. The functional groups on the surface of the materials were identified by a Fourier transform infrared (FT-IR) spectrometer (model IRPrestige-21; Shimadzu, Kyoto, Japan). Energy-dispersive spectroscopy (EDS) was employed to determine the elemental composition of the materials. Raman spectroscopy was used to study the presence of defects on the surface of materials, using a Lab RAM HR800 device (Horiba, Tokyo, Japan). The tube morphology was studied using a field-emission scanning electron microscope (FE-SEM; model S-4800; Hitachi, Tokyo, Japan) and a scanning transmission electron microscope (STEM) attached to the FE-SEM. The specific surface area was measured by nitrogen adsorption/desorption instrument at 77 K (model BELSORP-mini, MicrotrackBEL, Osaka, Japan) and calculated on the basis of Brunauer–Emmett–Teller (BET) theory.

2.2.2 Catalytic studies:

Anhydrous DBT (Merck, Germany) was dissolved in n-hexane (Daejung Chemicals & Metals Company, Korea) to prepare model oil samples. Catalytic oxidation was carried out in a reactor containing 20 ml of the model oil, 10 ml of acetonitrile, W/CNT, and H2O2 solution. The concentration of DBT was determined by UV–vis spectroscopy (λmax=235 nm). DBT conversion was calculated using the following expression:

(1)H=(C0C)100C0

where H (%) is the DBT conversion, and C0 and C (mg l−1) are the concentrations of DBT before and after catalysis, respectively.

To determine the effect of the tungsten amount in W/CNT for optimal DBT conversion, samples of W/CNT containing different percentages of tungsten (w/w) were prepared and their DBT conversions were tested. The effect of the ultrasonication time was also studied (0.5–5.0 h).

In catalytic studies, the effect of the molar ratio of O/S (H2O2/S) and catalyst dosage for DBT conversion, kinetic, and thermodynamic investigations were carried out. W/CNT was added to the sample at different dosages at 30°C with different molar ratios of O/S. The catalysis was allowed to take place, and the concentration of DBT before and after the catalytic reaction was determined. The kinetic data were inferred from the study of the catalytic ability of W/CNT for different reaction times.

The effect of temperature on the DBT removal reaction was studied from 283 to 323 K. At each temperature, 20 ml of the model oil sample containing 500 mg l−1 of sulfur and 10 ml of acetonitrile was stirred with 5 g l−1 of W/CNT for different times. Consequently, the Gibbs free energy (ΔG°), enthalpy (ΔH), entropy (ΔS), activation energy (Ea), and activation enthalpy (ΔH*) of adsorption were determined.

The recycling ability of the catalyst was studied. The catalyst was filtered and washed by n-hexane, acetonitrile, and distilled water and then dried at 80°C for 24 h before reuse with the initial sulfur concentration of 491 mg l−1.

3 Results and discussion

3.1 Determination of suitable synthesis conditions of W/CNT

The synthesis of W/CNT was carried out by combining ultrasonic and mechanical stirring of a mixture of 0.1 m H2SO4 solution of CNTs and Na2WO4. During the synthesis, the catalytic efficiency of W/CNT material invariably depends on the tungsten amount (percentage of tungsten in catalyst) and ultrasonication time. Consequently, two groups of samples were prepared with different amounts of tungsten (0.5–10.0%) and different ultrasonic times (0.5–5.0 h) at the same initial sulfur concentration of 520 mg l−1. The dosage of W/CNT, reaction time, and O/S molar ratio were fixed at 5.0 g l−1, 3.0 h, and 13.7, respectively. The W/CNT material was obtained after ultrasonic-assisted mechanical stirring of the mixture of CNTs, H2SO4, and Na2WO4 solution, prior to the filtering and drying in a furnace at 773 K. DBT conversions of samples were determined based on initial and remaining concentrations of DBT, as shown in Figure 2.

Figure 2: Effect of tungsten amount in W/CNT for DBT conversion.
Figure 2:

Effect of tungsten amount in W/CNT for DBT conversion.

Figure 2 presents the effect of the amount of tungsten in W/CNT on DBT conversion at room temperature. As can be seen from the figure, bare CNTs with and without H2O2 showed the same DBT conversion of ~30%, indicating that the conversion of DBT is due to the adsorption and liquid/liquid extraction (20%) and that the CNTs without tungsten did not catalyze any DBT oxidation reaction. Meanwhile, W/CNT catalyzed the DBT oxidation reaction and its catalytic ability depended on the amount of tungsten. DBT conversion increased significantly from 3.8% to 95.9% as the amount of tungsten was increased from 0.5% to 6.0%, and then remained constant at ~95% as the tungsten content further increased. This means that tungsten acts as the catalytic sites, and the more the amount of tungsten, the more the number of catalytic sites, resulting in enhanced DBT conversion. A high amount of tungsten in W/CNT may make the catalytic sites to saturate, leading to unchanged DBT conversion. Therefore, a tungsten content of 6.0% was taken for the optimal catalytic activity of W/CNT. In comparison with some other tungsten-containing catalysts (shown in Table 1), although the tungsten content used in our catalyst was lower, DBT conversion still was higher or equal at room temperature. As can be seen, CNTs can act as a carrier for the effective enhancement of the number of catalytic sites. Also, the π–π interaction between the carbon rings of CNTs and DBT increases the attraction of DBT toward the surface of the catalyst.

Table 1:

Catalytic oxidation conditions of DBT in some studies.

CatalystTungsten amount (%)O/S molar ratioT (°C)DBT conversion (%)Catalyst dosage (g l−1)References
W/CNTs6.013.72895.95.0Present study
W/MCM-4115.05.04069.64.0[16]
7098.0
W/MIL-10143.650.04591.07.5[21]
W/CeO226.75.03091.66.0[22]
W/SiO213.34.36075.51.5[23]
23.599.7
38.0100.0
  1. Initial sulfur concentration=500 mg l−1; MIL-101 is a metal organic framework material; MCM-41 is a mesoporous silica material.

The homogeneous catalyst Na2WO4 without carrier yielded lower DBT conversion (67.9%) than the heterogeneous catalyst W/CNT, which was unresponsive for deep desulfurization. This was also proved in some studies such as those of Al-Shahrani et al. [2], who found that the DBT conversion was only 20% at room temperature even though they used a higher dosage of the catalyst (10.8%) at 60°C as in the study of Chao et al. [13]. The same experiment was carried out with the heterogeneous catalyst WO3 without CNTs. The low DBT conversion (~60%) proved the important role of CNTs as the catalyst carrier.

The effect of ultrasonication time at 40°C during W/CNT synthesis on DBT conversion was investigated. The result showed that DBT conversion increased with an increase in ultrasonication time and only slightly changed when applied for more than 1 h. It is worth noting that the significant increase of DBT conversion of ~78% was obtained for ~30 min of ultrasonication and the DBT conversion reached 96% only after 1 h of treatment compared to 50% with W/CNT prepared without ultrasonication. This means that ultrasonic-assisted mechanical stirring dispersed tungsten efficiently and homogeneously on the CNTs. Therefore, the suitable ultrasonic time for synthesizing W/CNT was taken as 1 h.

3.2 Charaterization of W/CNT

The XRD patterns of CNTs and W/CNT are shown in Figure 3. The crystal phase of carbon in both materials, which was confirmed by the diffraction peak at 2θ of ~26° (JCPDS card files, no 41-1487), corresponded with the (002) lattice face [24], [25]. For W/CNT, the appearance of broad diffraction lines at 2θ from 26° to 30 corresponding to the (002) and (020) lattice faces, as well as two peaks at ~35 corresponding to the (022) and (202) lattice faces, proved the presence of tungsten on the surface of CNTs [26].

Figure 3: XRD patterns of CNTs and W/CNT.
Figure 3:

XRD patterns of CNTs and W/CNT.

EDS analysis of CNTs and W/CNT (Figure 4) provided evidence for the presence of graphite as the main component of the samples (92.36% for CNTs and 84.25% for W/CNT). The weight percent of tungsten in the material inferred from EDS analysis of W/CNT was 4.61%. The weight percent of oxygen in the W/CNT sample was higher than that in the CNT samples, which could be explained by the dispersion of oxygen from WO3 on the surface of the material.

Figure 4: EDX analyses of CNTs and W/CNT.
Figure 4:

EDX analyses of CNTs and W/CNT.

Some organic functional groups of CNTs and W/CNT were detected from the FT-IR spectra (Figure 5). For both of materials, the broad absorption band at ~3447 cm−1 was due to the stretching vibration of the C–OH groups on the surface of the materials. The peaks from 1562 to 1700 cm−1 for W/CNT and the broad band at ~1629 cm−1 for CNTs were attributed to the H–O–H bending vibration of the adsorbed H2O molecules, the C=C groups in C-sp2 of graphite, and C=O groups relating to the oxygen-containing groups such as –COOH. For W/CNT, the peak band at ~855.4 cm−1 was ascribed to the vibration of the W–O–W bond, which was not found in the case of the bare CNTs [27], [28].

Figure 5: FT-IR spectra of CNTs and W/CNT.
Figure 5:

FT-IR spectra of CNTs and W/CNT.

Figure 6 shows a selected region of the Raman spectra of the CNTs and W/CNT. Both CNTs and W/CNT show two peaks at ~1320 cm−1 called the D band (D for disorder) and 1566 cm−1 called the G band (G for graphite). The graphitic tubes were characterized by the G band. Meanwhile, the D band confirmed the presence of defects within the hexagonal graphitic structure. The G′ band at 2642 cm−1 was an overtone of the D band. The density of defects in the CNT structure could be estimated by the ratio of the integrated intensities of the D and G bands (ID/IG) [27], [28], [29]. Thus, a higher value of ID/IG of W/CNT (1.53) is indicative of higher defects and disorders of the graphitized structures in comparison with CNTs (1.42) due to the bonding between WO3 and C-sp2 of the graphitic tubes. Compared to those of CNTs, the Raman vibrations of W/CNT centered at 250 and 757 cm−1 were attributed to the W–O bonds [27], [28]. The presence of these peaks in the Raman spectrum of W/CNT confirmed that WO3 was dispersed on the surface of CNTs.

Figure 6: Raman spectra of CNTs and W/CNT.
Figure 6:

Raman spectra of CNTs and W/CNT.

Figure 7 shows the FE-SEM and STEM images of the W/CNT. These STEM images show that the long tubular structure with tube diameter of ~30–50 nm was still maintained after dispersing tungsten. WO3 particles could not be observed on the SEM and STEM images because of their small size. The surface area based on the BET model was 139 m2 g−1.

Figure 7: FE-SEM and STEM images of W/CNT.
Figure 7:

FE-SEM and STEM images of W/CNT.

3.3 DBT catalytic oxidation

3.3.1 Effect of molar ratio of O/S

The effect of the molar ratio O/S on the DBT oxidation conversion was investigated from 1.7 to 51.2. With initial sulfur concentration of 529 mg l−1, the catalyst dosage of 5.0 g·l−1, and at room temperature, DBT oxidation conversion increased remarkably up to 96% as the O/S molar ratio was increased from 1.7 to 17.0, but varied only unremarkably afterward.

Therefore, the O/S molar ratio of 17.0 was chosen for further experiments. This ratio is higher than that reported by some other authors, as shown in Table 1. However, it is worth noting that the present study was conducted under ambient conditions with low tungsten content.

3.3.2 Effect of W/CNT dosage

An increase in trend of DBT conversion was observed (from 77% to 94%) when the dosage of W/CNT was increased from 1.25 to 5 g l−1. Subsequently, DBT conversion reached 98% when the W/CNT dosage was 6.25 g l−1 and more, with the initial sulfur concentration of 525 mg l−1, O/S molar ratio of 17.0, and at room temperature (28°C). Therefore, the suitable W/CNT dosage was chosen as 5 g l−1 for deep desulfurization.

3.3.3 Catalytic kinetics

Figure 8 presents the reaction kinetics of DBT oxidation with different initial concentrations of DBT. As seen from Figure 8, the DBT conversion increased with increase in contact time. Because of the increase of the initial sulfur concentration without the increase in dosage of the catalyst (5 g l−1) and oxidant amount (O/S molar ratio of 17.0), the efficiency of the reaction decreased from ~95% to 86% corresponding the initial sulfur concentration from 223 to 620 mg l−1, and the equilibrium reaction time increased from 60 to 90 min.

Figure 8: Effect of reaction times in DBT conversion of W/CNT at different initial DBT concentrations.
Figure 8:

Effect of reaction times in DBT conversion of W/CNT at different initial DBT concentrations.

The Langmuir–Hinshelwood (LH) kinetic model is popularly used to explain the mechanism of heterogeneous catalytic reactions [30], [31]. In this model, the reactants are adsorbed on the catalytic centers (S) to form the intermediates (A…S or B…S), and then these intermediates react together to form the products. The reaction can be describes as follows:

(2)A+SAS
(3)B+SB.S
(4)AS+BSproducts

It is assumed that Equation (4) is the rate-limiting step. The LH expression is given by Equation (5):

(5)r=dCdt=krKC1+KC

where r represents the rate of oxidation (mg l−1 min−1), kr is the LH rate constant (mg l−1 min−1), C is the DBT concentration (mg l−1), and K is the equilibrium adsorption constant (l mg−1).

During the catalytic process, an intermediate is formed, which hinders the determination of the reaction kinetics because of the competition between adsorption and catalysis. Therefore, in very early time, the effect of the intermediate is supposed to be unremarkable. Then, the initial rate of reaction, r0, as a function of the initial concentration of sulfur, C0, is calculated by Equation (6):

(6)r0=krKC01+KC0

which can be linearizied as in Equation (7):

(7)1r0=1kr+1krKC0

KC<<1 means that means the DBT is weakly adsorbed on the catalyst, and Equation (5) can be rewritten as Equation (8) and linearized as the first-order kinetic equation (Eq. 9):

(8)r=krKC
(9)lnC0C=k1t

where kr is replaced by the first-order rate constant, k1 (min−1).

Figure 9 shows the linear plots of ln(C0/C) versus the reaction time (t). As seen from Figure 9, the high determination coefficients (0.987–0.999) for all initial concentrations indicate that the kinetic data fit well the first-order kinetics model. The value of k1 was obtained from the slope of the linear regression line shown in Table 2.

Figure 9: First-order kinetic study of the catalytic oxidation of DBT at different initial DBT concentrations.
Figure 9:

First-order kinetic study of the catalytic oxidation of DBT at different initial DBT concentrations.

Table 2:

First-order kinetic parameters of the catalytic oxidation of DBT at different initial DBT concentrations.

C0 (mg l−1)k1 (min−1)r0 (mg l−1 min−1)
2230.0554012.35
2920.0441712.89
3950.0335013.18
5210.0263513.73
6200.0223013.83

As can be seen from Table 2, the higher initial DBT concentration, the higher the initial rate and the lower rate constant. The initial concentration provides an important driving force to overcome all the mass transfer resistance of DBT between the aqueous solution and the W/CNT surface. As a result, a high initial DBT concentration will enhance the initial rate. The mass of the catalyst and oxidant reagent is fixed, so the rate constant is reduced as a result of the decrease in O/S molar ratio and the number of DBT/catalytic sites.

The plot 1/r0 versus 1/C0 following LH kinetic model is shown in Figure 10. The correlation coefficient for the LH kinetic equation was nearly unity (r2=0.991), indicating that the DBT catalytic oxidation data fitted well with LH kinetic model. The LH reaction rate (kr) and equilibrium constant (K) of the adsorption of DBT onto W/CNT were obtained from the slope and the intercept of regression line, which were 14.84 (mg l−1·min−1) and 0.022 (l mg−1), respectively.

Figure 10: Langmuir–Hinshelwood kinetic model of the catalytic oxidation of DBT.
Figure 10:

Langmuir–Hinshelwood kinetic model of the catalytic oxidation of DBT.

The values of the rate constants at different temperatures were obtained from the first-order kinetic equation, as shown in Figure 11 and Table 3.

Figure 11: First-order kinetic studies of the catalytic oxidation of DBT at different temperatures.
Figure 11:

First-order kinetic studies of the catalytic oxidation of DBT at different temperatures.

Table 3:

First-order kinetic parameters of the catalytic oxidation of DBT at different temperatures.

Temperature (K)Correlation coefficient (r2)k1 (min−1)
2830.9920.01272
2930.9910.01932
3030.9990.02635
3130.9890.04079
3230.9890.06476

Table 3 shows that the reaction rate constant (k1) or the reaction rate (r) or DBT conversion increases with temperature. In other words, the reaction is more favorable at high temperature.

The activation energy (Ea) was determined by the Arrhenius equation (Eq. 10)

(10)kT=AeEa/RT

Taking the natural logarithm of both sides of Equation (11), one obtains

(11)lnkT=lnAEaRT

where A is the pre-exponential factor, R is the universal gas constant (8.314 J mol−1 K−1), and T is absolute temperature (K).

Figure 12A shows the linear plot of (ln kT) versus 1/T. The value of Ea could be obtained from the slope (−Ea/R). The obtained Ea value calculated by the Arrhenius equation was 30.38 kJ mol−1. The low activation energy (below 42 kJ mol−1) implies that the movement of the reactants to an external surface of catalyst is controlled by diffusion.

Figure 12: (A) Arrhenius and (B) Eyring equations for DBT catalytic oxidation.
Figure 12:

(A) Arrhenius and (B) Eyring equations for DBT catalytic oxidation.

The thermodynamic parameters of activation were evaluated to estimate whether there is any complex formation prior to the final adsorption. They are obtained by applying the Eyring equation in the linear form, which is expressed as follows:

(12)lnkTT=ΔH#RT+lnkBh+ΔS#R

where kT is the rate constant equal to the rate constant in the first-order model, kB (1.3807×10−23 J K−1) is the Boltzmann constant, and h (6.621 J s) is the Planck constant.

From the linear plot of ln(k/T) versus 1/T, the valued s of ΔS# and ΔH# were calculated from the slope (−ΔH#/T) and the y-intercept [ln(kB/h)+(ΔS#/R)]. The linear plot of ln(k/T) versus 1/T is shown in Figure 12B. The activation parameters for DBT catalytic oxidation are shown in Table 4.

Table 4:

Activation parameters for DBT catalytic oxidation.

Temperature (K)ΔH# (kJ mol−1)ΔS# (J mol−1 K−1)ΔG# (kJ mol−1)
28327.87468.28−104.65
293−109.34
303−114.02
313−118.70
323−123.39

The positive value of ΔS# (468.28 J mol−1 K−1) suggests the possibility of formation of an activated complex (metal peroxide intermediate) between DBT and H2O2 on the catalyst. Also, the positive value of ΔS# normally reflects significant changes occurring in the internal structure of the catalyst during the catalytic process. The value of ΔH# (27.87 kJ mol−1) suggests that the formation of the activated complex was endothermic. The large negative values of ΔG# imply that the formation of the activated complex occurred spontaneously and favorably at high temperature.

The Gibbs free energy (ΔG°) of reaction was determined by Equation (13), while the enthalpy (ΔH°) and entropy (ΔS°) parameters were calculated using the Van’t Hoff equation (Eq. 14 and Figure 13).

Figure 13: Van’t Hoff plot for DBT catalytic oxidation.
Figure 13:

Van’t Hoff plot for DBT catalytic oxidation.

(13)ΔG°=RTlnKC
(14)lnKC=ΔG°RT=ΔS°RΔH°RT

The thermodynamic nature of the reaction was examined through the thermodynamic parameters ΔG°, ΔH°, and ΔS° calculated from Equations (13) and (14), and are shown in Table 5.

Table 5:

Thermodynamic parameters of DBT catalytic oxidation.

Temperature (K)ΔH° (kJ mol−1)ΔS° (J mol−1 K−1)ΔG° (kJ mol−1)
28392.69332.18−1.31
293−4.64
303−7.96
313−11.28
323−146.01

The positive value of standard enthalpy change (ΔH°=92.69 kJ mol−1) confirmed the endothermic nature of catalytic oxidation. The positive value of standard entropy (ΔS° 322.18 J mol−1 K−1) demonstrates that the reaction enhances the randomness at the liquid–solid interface of the system as a result of the increase in the number and kinds of molecules in the mixture after the reaction. The Gibbs free energy (ΔG°) at different temperatures had negative values, indicating that DBT oxidation was spontaneous and favorable at high temperature.

The final product, dibenzothiophene sulfone (DBTS), was analyzed by gas chromatography-mass spectrometry (GC-MS; not shown here). In n-hexane solvent, the retention times of DBT and DBTS were 16.7 and 18.7 min, respectively. After the oxidation, the oxidation number of sulfur atoms in DBT increased to +2 in sulfoxide and +4 in sulfone. Based on the kinetic and thermodynamic studies, the proposed mechanism of DBT oxidation over W/CNT can be described as in Scheme 1 [32], [33].

Scheme 1: Proposed mechanism of the catalytic oxidation of DBT to DBTS.
Scheme 1:

Proposed mechanism of the catalytic oxidation of DBT to DBTS.

The W/CNT with high surface area and active tungsten sites adsorbed the reactants on its surface. This adsorption should be related to the coordination of tungsten toward the free electron pairs of sulfur and the formation of hydroperoxide tungstate through the nucleophilic attack of hydrogen peroxide to tungsten. The π–π stacking hydrophobic interaction between the sp2-carbon of graphite and aromatic rings of DBT was also responsible for the accumulation of the reactant on W/CNT surfaces. The intermediates, including hydroperoxide tungstate and DBT intermediates, react together to provide DBTS.

3.4 Recyclability of catalyst

The possibility to reuse the W/CNT catalyst was evaluated by the determination of DBT conversion and remaining tungsten amount of the catalyst after four reaction cycles. The result showed that the DBT conversion was retained at ~94% after two first cycles and slightly decreased to 80% after the third reaction cycle. This meant that the activity of catalyst was maintained well after three reaction cycles. However, after the fourth cycle, the DBT conversion was strongly reduced to ~25%, which was equivalent to that with bare CNTs (Figure 2). The stability of the catalyst was confirmed by EDS/SEM and XRD analyses, as shown in Figure 14 and Table 6. No remarkable reduction of the amount of tungsten in catalyst was observed, and the WO3 crystal phase still appeared in the XRD pattern of the catalyst during the three reaction cycles. The activity reduction of the used catalyst could result from the poisoning of the active sites on the catalyst by organic compounds.

Figure 14: XRD patterns of reused W/CNT.
Figure 14:

XRD patterns of reused W/CNT.

Table 6:

Element compositions of reused catalyst of W/CNT.

Runwt%
WCOAl
14.2383.1911.211.38
24.1284.129.542.23
33.8785.218.792.13
43.2588.236.781.74

4 Conclusions

W/CNT was found to be an efficient catalyst for the oxidation of DBT to DBTS. More than 95% of DBT was oxidized at ambient temperature with the catalyst W/CNT containing 6.0% tungsten (w/w) and O/S molar ratio 17.0. The DBT conversion reached 95% after 90 min at an initial sulfur concentration of less than 600 mg l−1. The oxidation mechanism of DBT on the W/CNT catalyst followed the LH model. The negative Gibbs free energy variation (ΔG°) and positive enthalpy (ΔH°) in the range of temperature 283–323 K showed that the reaction was spontaneous and an endothermic process.

References

[1] Yang RT, Hernández-Maldonado AJ, Yang FH. Science 2003, 301, 79–81.10.1126/science.1085088Search in Google Scholar

[2] Al-Shahrani F, Xiao T, Llewellyn SA, Barri S, Jiang Z, Shi H, Martinie G, Green MLH. Appl. Catal. B 2007, 73, 311–316.10.1016/j.apcatb.2006.12.016Search in Google Scholar

[3] Ribeiro S, Barbosa ADS, Gomes AC, Pillinger M, Goncalves IS, Cunha-Silva L, Balula SS. Fuel Process. Technol. 2013, 116, 350–357.10.1016/j.fuproc.2013.07.011Search in Google Scholar

[4] Houalla M, Broderick DH, Sapre AV, Nag NK, De Beer VHJ, Gates BC, Kwart H. J. Catal. 1980, 61, 523–527.10.1016/0021-9517(80)90400-5Search in Google Scholar

[5] Shiraishi Y, Naito T, Hirai T, Komasawa I. Chem. Commun. 2001, 14, 1256–1257.10.1039/b103061mSearch in Google Scholar

[6] Kabe T, Ishihara A, Qian W. Hydrodesulfurization and Hydrodenitrogenation, Kodansha Scientific, Wiley/VCH: Tokyo, New York, 1999.Search in Google Scholar

[7] Yazu K, Makino M, Ukegawa K. Chem. Lett. 2004, 33, 1306–1307.10.1246/cl.2004.1306Search in Google Scholar

[8] Campos-Martin JM, Capel-Sanchez MC, Fierro JLG. Green Chem. 2004, 6, 557–562.10.1039/b409882jSearch in Google Scholar

[9] Garcia-Gutierrez JL, Fuentes GA, Hernandez-Teran ME, Murrieta F, Navarrete J, Jimenez-Cruz F. Appl. Catal. A: Gen. 2006, 305, 15–20.10.1016/j.apcata.2006.01.027Search in Google Scholar

[10] Wang D, Qian EW, Amano H, Okata K, Ishihara A, Kabe T. Appl. Catal. A: Gen. 2003, 253, 91–99.10.1016/S0926-860X(03)00528-3Search in Google Scholar

[11] Tam PS, Kittrell JR, Eldridge JW. Ind. Eng. Chem. Res. 1990, 29, 321–324.10.1021/ie00099a002Search in Google Scholar

[12] Yao X, Wang S, Ling F, Li F, Zhang J, Ma B. J. Fuel Chem. Technol. 2004, 32, 318–322.Search in Google Scholar

[13] Chao Y, Li H, Zhu W, Zhu G, Yan Y. Pet. Sci. Technol. 2010, 28, 1242–1249.10.1080/10916460903030441Search in Google Scholar

[14] Jose N, Sengupta S, Basu JK. Fuel 2011, 90, 626–632.10.1016/j.fuel.2010.09.026Search in Google Scholar

[15] Tang L, Luo G, Zhu M, Kang L, Dai B. J. Ind. Eng. Chem. 2013, 19, 620–626.10.1016/j.jiec.2012.09.015Search in Google Scholar

[16] Abdalla ZEA, Li B, Han C, Mustafa B. Jordan J. Chem. 2010, 5, 33–45.Search in Google Scholar

[17] Kong L, Li G, Wang X. Catal. Lett. 2004, 92, 163–167.10.1023/B:CATL.0000014340.08449.d0Search in Google Scholar

[18] Zhang M, Zhu W, Li H, Xun S, Ding W, Liu J, Zhao Z, Wang Q. Chem. Eng. J. 2013, 243, 386–393.10.1016/j.cej.2013.12.093Search in Google Scholar

[19] Fallah RN, Azizian S, Dwivedi AD, Sillanpaa M. Fuel Process. Technol. 2015, 130, 214–223.10.1016/j.fuproc.2014.10.018Search in Google Scholar

[20] Zhang W, Zhang H, Xiao J, Zhao Z, Yu M, Li Z. Green Chem. 2014, 16, 211–220.10.1039/C3GC41106KSearch in Google Scholar

[21] Hu X, Lu Y, Dai F, Liu C, Liu Y. Microporous Mesoporous Mater. 2012, 170, 36–44.10.1016/j.micromeso.2012.11.021Search in Google Scholar

[22] Li Y, Zhang M, Zhu W, Li M, Xiong J, Zhang Q, Weia Y, Li H. RSC Adv. 2016, 6, 68922–68928.10.1039/C6RA06081ASearch in Google Scholar

[23] Zhu W, Gu Q, Hu J, Wu P, Yin S, Zhu F, Zhang M, Xiong J, Li H. J. Porous Mater. 2015, 22, 1227–1233.10.1007/s10934-015-9999-4Search in Google Scholar

[24] Cao A, Xu C, Liand J, Wu D, Wei B. Chem. Phys. Lett. 2001, 344, 13–17.10.1016/S0009-2614(01)00671-6Search in Google Scholar

[25] Zhang J, Li J, Cao J, Qian Y. Mater. Lett. 2008, 62, 1839–1842.10.1016/j.matlet.2007.10.015Search in Google Scholar

[26] Martínez-de la Cruz A, Sánchez Martínez D, López Cuéllar E. Solid State Sci. 2010, 12, 88–94.10.1016/j.solidstatesciences.2009.10.010Search in Google Scholar

[27] Zhou MJ, Zhang N, Hou ZH. Int. J. Photoenergy 2014, 2014, 1–6.10.1155/2014/569763Search in Google Scholar

[28] Guo J, Li Y, Zhu S, Chen Z, Liu Q, Zhang D, Moon W, Song D. RSC Adv. 2012, 2, 1356–1363.10.1039/C1RA00621ESearch in Google Scholar

[29] Sui XM, Giordani S, Prato M, Wagner HD. Appl. Phys. Lett. 2009, 95, 233113.10.1063/1.3272012Search in Google Scholar

[30] Khezrianjoo S, Revanasiddappa HD. Chem. Sci. J. 2012, 2012, 1–7.10.5402/2012/728594Search in Google Scholar

[31] Kumar KV, Porkodi K, Rocha F. Catal. Commun. 2008, 9, 82–84.10.1016/j.catcom.2007.05.019Search in Google Scholar

[32] Long Z, Yang C, Zeng G, Peng L, Dai C, He H. Fuel 2014, 130, 19–24.10.1016/j.fuel.2014.04.005Search in Google Scholar

[33] Zhang M, Zhu W, Li H, Xun S, Li M, Li Y, Wei Y, Li H. Chin. J. Catal. 2016, 37, 971–978.10.1016/S1872-2067(15)61103-2Search in Google Scholar

Received: 2017-12-24
Accepted: 2018-04-03
Published Online: 2018-05-16
Published in Print: 2019-01-28

©2019 Walter de Gruyter GmbH, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 Public License.

Articles in the same Issue

  1. Regular Articles
  2. Studies on the preparation and properties of biodegradable polyester from soybean oil
  3. Flow-mode biodiesel production from palm oil using a pressurized microwave reactor
  4. Reduction of free fatty acids in waste oil for biodiesel production by glycerolysis: investigation and optimization of process parameters
  5. Saccharin: a cheap and mild acidic agent for the synthesis of azo dyes via telescoped dediazotization
  6. Optimization of lipase-catalyzed synthesis of polyethylene glycol stearate in a solvent-free system
  7. Green synthesis of iron oxide nanoparticles using Platanus orientalis leaf extract for antifungal activity
  8. Ultrasound assisted chemical activation of peanut husk for copper removal
  9. Room temperature silanization of Fe3O4 for the preparation of phenyl functionalized magnetic adsorbent for dispersive solid phase extraction for the extraction of phthalates in water
  10. Evaluation of the saponin green extraction from Ziziphus spina-christi leaves using hydrothermal, microwave and Bain-Marie water bath heating methods
  11. Oxidation of dibenzothiophene using the heterogeneous catalyst of tungsten-based carbon nanotubes
  12. Calcined sodium silicate as an efficient and benign heterogeneous catalyst for the transesterification of natural lecithin to L-α-glycerophosphocholine
  13. Synergistic effect between CO2 and H2O2 on ethylbenzene oxidation catalyzed by carbon supported heteropolyanion catalysts
  14. Hydrocyanation of 2-arylmethyleneindan-1,3-diones using potassium hexacyanoferrate(II) as a nontoxic cyanating agent
  15. Green synthesis of hydratropic aldehyde from α-methylstyrene catalyzed by Al2O3-supported metal phthalocyanines
  16. Environmentally benign chemical recycling of polycarbonate wastes: comparison of micro- and nano-TiO2 solid support efficiencies
  17. Medicago polymorpha-mediated antibacterial silver nanoparticles in the reduction of methyl orange
  18. Production of value-added chemicals from esterification of waste glycerol over MCM-41 supported catalysts
  19. Green synthesis of zerovalent copper nanoparticles for efficient reduction of toxic azo dyes congo red and methyl orange
  20. Optimization of the biological synthesis of silver nanoparticles using Penicillium oxalicum GRS-1 and their antimicrobial effects against common food-borne pathogens
  21. Optimization of submerged fermentation conditions to overproduce bioethanol using two industrial and traditional Saccharomyces cerevisiae strains
  22. Extraction of In3+ and Fe3+ from sulfate solutions by using a 3D-printed “Y”-shaped microreactor
  23. Foliar-mediated Ag:ZnO nanophotocatalysts: green synthesis, characterization, pollutants degradation, and in vitro biocidal activity
  24. Green cyclic acetals production by glycerol etherification reaction with benzaldehyde using cationic acidic resin
  25. Biosynthesis, characterization and antimicrobial activities assessment of fabricated selenium nanoparticles using Pelargonium zonale leaf extract
  26. Synthesis of high surface area magnesia by using walnut shell as a template
  27. Controllable biosynthesis of silver nanoparticles using actinobacterial strains
  28. Green vegetation: a promising source of color dyes
  29. Mechano-chemical synthesis of ammonia and acetic acid from inorganic materials in water
  30. Green synthesis and structural characterization of novel N1-substituted 3,4-dihydropyrimidin-2(1H)-ones
  31. Biodiesel production from cotton oil using heterogeneous CaO catalysts from eggshells prepared at different calcination temperatures
  32. Regeneration of spent mercury catalyst for the treatment of dye wastewater by the microwave and ultrasonic spray-assisted method
  33. Green synthesis of the innovative super paramagnetic nanoparticles from the leaves extract of Fraxinus chinensis Roxb and their application for the decolourisation of toxic dyes
  34. Biogenic ZnO nanoparticles: a study of blueshift of optical band gap and photocatalytic degradation of reactive yellow 186 dye under direct sunlight
  35. Leached compounds from the extracts of pomegranate peel, green coconut shell, and karuvelam wood for the removal of hexavalent chromium
  36. Enhancement of molecular weight reduction of natural rubber in triphasic CO2/toluene/H2O systems with hydrogen peroxide for preparation of biobased polyurethanes
  37. An efficient green synthesis of novel 1H-imidazo[1,2-a]imidazole-3-amine and imidazo[2,1-c][1,2,4]triazole-5-amine derivatives via Strecker reaction under controlled microwave heating
  38. Evaluation of three different green fabrication methods for the synthesis of crystalline ZnO nanoparticles using Pelargonium zonale leaf extract
  39. A highly efficient and multifunctional biomass supporting Ag, Ni, and Cu nanoparticles through wetness impregnation for environmental remediation
  40. Simple one-pot green method for large-scale production of mesalamine, an anti-inflammatory agent
  41. Relationships between step and cumulative PMI and E-factors: implications on estimating material efficiency with respect to charting synthesis optimization strategies
  42. A comparative sorption study of Cr3+ and Cr6+ using mango peels: kinetic, equilibrium and thermodynamic
  43. Effects of acid hydrolysis waste liquid recycle on preparation of microcrystalline cellulose
  44. Use of deep eutectic solvents as catalyst: A mini-review
  45. Microwave-assisted synthesis of pyrrolidinone derivatives using 1,1’-butylenebis(3-sulfo-3H-imidazol-1-ium) chloride in ethylene glycol
  46. Green and eco-friendly synthesis of Co3O4 and Ag-Co3O4: Characterization and photo-catalytic activity
  47. Adsorption optimized of the coal-based material and application for cyanide wastewater treatment
  48. Aloe vera leaf extract mediated green synthesis of selenium nanoparticles and assessment of their In vitro antimicrobial activity against spoilage fungi and pathogenic bacteria strains
  49. Waste phenolic resin derived activated carbon by microwave-assisted KOH activation and application to dye wastewater treatment
  50. Direct ethanol production from cellulose by consortium of Trichoderma reesei and Candida molischiana
  51. Agricultural waste biomass-assisted nanostructures: Synthesis and application
  52. Biodiesel production from rubber seed oil using calcium oxide derived from eggshell as catalyst – optimization and modeling studies
  53. Study of fabrication of fully aqueous solution processed SnS quantum dot-sensitized solar cell
  54. Assessment of aqueous extract of Gypsophila aretioides for inhibitory effects on calcium carbonate formation
  55. An environmentally friendly acylation reaction of 2-methylnaphthalene in solvent-free condition in a micro-channel reactor
  56. Aegle marmelos phytochemical stabilized synthesis and characterization of ZnO nanoparticles and their role against agriculture and food pathogen
  57. A reactive coupling process for co-production of solketal and biodiesel
  58. Optimization of the asymmetric synthesis of (S)-1-phenylethanol using Ispir bean as whole-cell biocatalyst
  59. Synthesis of pyrazolopyridine and pyrazoloquinoline derivatives by one-pot, three-component reactions of arylglyoxals, 3-methyl-1-aryl-1H-pyrazol-5-amines and cyclic 1,3-dicarbonyl compounds in the presence of tetrapropylammonium bromide
  60. Preconcentration of morphine in urine sample using a green and solvent-free microextraction method
  61. Extraction of glycyrrhizic acid by aqueous two-phase system formed by PEG and two environmentally friendly organic acid salts - sodium citrate and sodium tartrate
  62. Green synthesis of copper oxide nanoparticles using Juglans regia leaf extract and assessment of their physico-chemical and biological properties
  63. Deep eutectic solvents (DESs) as powerful and recyclable catalysts and solvents for the synthesis of 3,4-dihydropyrimidin-2(1H)-ones/thiones
  64. Biosynthesis, characterization and anti-microbial activity of silver nanoparticle based gel hand wash
  65. Efficient and selective microwave-assisted O-methylation of phenolic compounds using tetramethylammonium hydroxide (TMAH)
  66. Anticoagulant, thrombolytic and antibacterial activities of Euphorbia acruensis latex-mediated bioengineered silver nanoparticles
  67. Volcanic ash as reusable catalyst in the green synthesis of 3H-1,5-benzodiazepines
  68. Green synthesis, anionic polymerization of 1,4-bis(methacryloyl)piperazine using Algerian clay as catalyst
  69. Selenium supplementation during fermentation with sugar beet molasses and Saccharomyces cerevisiae to increase bioethanol production
  70. Biosynthetic potential assessment of four food pathogenic bacteria in hydrothermally silver nanoparticles fabrication
  71. Investigating the effectiveness of classical and eco-friendly approaches for synthesis of dialdehydes from organic dihalides
  72. Pyrolysis of palm oil using zeolite catalyst and characterization of the boil-oil
  73. Azadirachta indica leaves extract assisted green synthesis of Ag-TiO2 for degradation of Methylene blue and Rhodamine B dyes in aqueous medium
  74. Synthesis of vitamin E succinate catalyzed by nano-SiO2 immobilized DMAP derivative in mixed solvent system
  75. Extraction of phytosterols from melon (Cucumis melo) seeds by supercritical CO2 as a clean technology
  76. Production of uronic acids by hydrothermolysis of pectin as a model substance for plant biomass waste
  77. Biofabrication of highly pure copper oxide nanoparticles using wheat seed extract and their catalytic activity: A mechanistic approach
  78. Intelligent modeling and optimization of emulsion aggregation method for producing green printing ink
  79. Improved removal of methylene blue on modified hierarchical zeolite Y: Achieved by a “destructive-constructive” method
  80. Two different facile and efficient approaches for the synthesis of various N-arylacetamides via N-acetylation of arylamines and straightforward one-pot reductive acetylation of nitroarenes promoted by recyclable CuFe2O4 nanoparticles in water
  81. Optimization of acid catalyzed esterification and mixed metal oxide catalyzed transesterification for biodiesel production from Moringa oleifera oil
  82. Kinetics and the fluidity of the stearic acid esters with different carbon backbones
  83. Aiming for a standardized protocol for preparing a process green synthesis report and for ranking multiple synthesis plans to a common target product
  84. Microstructure and luminescence of VO2 (B) nanoparticle synthesis by hydrothermal method
  85. Optimization of uranium removal from uranium plant wastewater by response surface methodology (RSM)
  86. Microwave drying of nickel-containing residue: dielectric properties, kinetics, and energy aspects
  87. Simple and convenient two step synthesis of 5-bromo-2,3-dimethoxy-6-methyl-1,4-benzoquinone
  88. Biodiesel production from waste cooking oil
  89. The effect of activation temperature on structure and properties of blue coke-based activated carbon by CO2 activation
  90. Optimization of reaction parameters for the green synthesis of zero valent iron nanoparticles using pine tree needles
  91. Microwave-assisted protocol for squalene isolation and conversion from oil-deodoriser distillates
  92. Denitrification performance of rare earth tailings-based catalysts
  93. Facile synthesis of silver nanoparticles using Averrhoa bilimbi L and Plum extracts and investigation on the synergistic bioactivity using in vitro models
  94. Green production of AgNPs and their phytostimulatory impact
  95. Photocatalytic activity of Ag/Ni bi-metallic nanoparticles on textile dye removal
  96. Topical Issue: Green Process Engineering / Guest Editors: Martine Poux, Patrick Cognet
  97. Modelling and optimisation of oxidative desulphurisation of tyre-derived oil via central composite design approach
  98. CO2 sequestration by carbonation of olivine: a new process for optimal separation of the solids produced
  99. Organic carbonates synthesis improved by pervaporation for CO2 utilisation
  100. Production of starch nanoparticles through solvent-antisolvent precipitation in a spinning disc reactor
  101. A kinetic study of Zn halide/TBAB-catalysed fixation of CO2 with styrene oxide in propylene carbonate
  102. Topical on Green Process Engineering
Downloaded on 12.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/gps-2017-0189/html
Scroll to top button