You are on page 1of 9

Chemical Engineering Journal 369 (2019) 775–783

Contents lists available at ScienceDirect

Chemical Engineering Journal


journal homepage: www.elsevier.com/locate/cej

Synthesis and characterization of PWMn/NiO/PAN nanosphere composite T


with superior catalytic activity for oxidative desulfurization of real fuel

Mohammad Ali Rezvani , Omid Feghhe Miri
Department of Chemistry, Faculty of Science, University of Zanjan, 451561319 Zanjan, Iran

H I GH L IG H T S G R A P H I C A L A B S T R A C T

• APANnewwasnanocomposite PWMn/NiO/
synthesized and employed as
a nanocatalyst in the ODS process.
• PWMn/NiO/PAN was shown be able
to remove sulfur compounds in real
fuel.
• Deep oxidative desulfurization of real
gasoline could be achieved with this
system.
• The nanocatalyst could be separated
and recycled easily after five runs.

A R T I C LE I N FO A B S T R A C T

Keywords: A new organic-inorganic hybrid material (PWMn/NiO/PAN) was successfully synthesized based on cesium salt
Heteroply phosphotungstate of mono Mn (II)-substituted phosphotungstate Cs5PW11MnO39 (PWMn), nickel oxide (NiO), and polyaniline
Nickel oxide (PAN). The samples were characterized using FT-IR, UV-vis, XRD, SEM, and EDX techniques. Taking into account
Polyaniline the analysis results, the phosphotungstate species were highly dispersed on the metal oxide and polymer matrix.
Oxidative desulfurization
Additionally, the PWMn/NiO/PAN composite was found in a nanoscale spherical shape with the main range
Nanocomposite
particle size of 50–60 nm. Oxidative desulfurization (ODS) experiments were carried out for removal of sulfur-
containing compounds in gasoline and model fuels. Compared with the catalytic activity of the bulk materials,
the prepared PWMn/NiO/PAN nanosphere composite showed superior performance in the ODS system.
According to the designed process, the sulfur compounds were oxidized in the presence of the catalyst by oxidant
and then extracted from the fuel phase by extraction solvent. The ODS results revealed that the concentration of
sulfur content of typical gasoline was lowered from 4989 to 133 ppm with 97% efficiency at 35 °C after 1 h. Also,
the PWMn/NiO/PAN hybrid catalyst could be reused for five successive runs without any appreciable loss in its
catalytic activity. The current research suggests that the PWMn/NiO/PAN nanosphere composite can represent a
new promising class of organic-inorganic hybrid catalysts for ODS system.

1. Introduction oxygen clusters with a variety of chemical composition and definite


structures [1]. They have exposed the potential for applications in
Polyoxometalates (POMs) are displayed as a vast class of metal- electrochemistry [2], catalysis [3], medicine [4], and materials design


Corresponding author.
E-mail address: marezvani@znu.ac.ir (M.A. Rezvani).

https://doi.org/10.1016/j.cej.2019.03.135
Received 27 December 2018; Received in revised form 13 March 2019; Accepted 16 March 2019
Available online 18 March 2019
1385-8947/ © 2019 Elsevier B.V. All rights reserved.
M.A. Rezvani and O.F. Miri Chemical Engineering Journal 369 (2019) 775–783

[5]. Among a large diversity in geometrical structures of POMs, the Hence, the use of catalysts has drawn increasing attention for lowering
Keggin-type heteropolyanions with the general formula of the oxidant consumption and increasing the reaction efficiency [39]. In
XnM12O40(8−n)- (where X = P, Si, As and M = W, Mo) have received a this regard, the POM-based heterogeneous catalyst exhibited superior
great deal of attention in catalysis [6,7]. Moreover, they are found rapid catalytic activity and stability in the ODS reactions [40–42].
development in catalytic oxidative reactions due to their fascinating The present work focuses on the development of the ODS process by
architectures, strong Brønsted acidity, and fast reversible multi-electron synthesis of a new organic-inorganic hybrid material, aiming at pro-
redox transformations [8]. The acid form of heteropoly compounds has duction of clean gasoline. Further, the PWMn/NiO/PAN as a hetero-
limited number of catalytic applications since they have associated with geneous catalyst was adopted for the oxidation removal of Th, BT, and
certain obstacles such as low surface areas (less than10 m2/g) and high DBT compounds in the model fuels. The mixture of hydrogen peroxide/
solubility in polar solvents (e.g., water, alcohols, and ethers) [9,10]. To acetic acid (in v/v ratio of 2:1) was used as an oxidant and acetonitrile
confront this limitation, many approaches were proposed to architect applied as an extraction solvent. The major factors that influence the
the POM-based heterogeneous catalysts. The modification of heteropoly desulfurization efficiency and the kinetic study of the ODS reactions
acids with cesium salt and immobilization of them on solid support were fully detailed and discussed. The probable ODS pathway was
allow obtaining active recyclable catalysts [11–13]. Polyaniline (PAN) proposed on the basis of the formation of the peroxometal (abbreviated
has been of particular interest because of its interesting redox proper- M(O2)n) complex.
ties, chemical stability, easy polymerization, relatively high con-
ductivity, and low cost of the monomer [14]. The incorporation of PAN 2. Experimental
and inorganic metal oxides have been adopted as an effective method in
order to improve physical, mechanical, and electrical properties [15]. 2.1. Materials and methods
Besides, nickel oxide (NiO) particles are very popular in the coatings for
the catalysis because of its high specific capacitance, environmental The following chemicals and reagents were purchased from com-
friendly inherent, and practical accessibility [16]. All these properties mercial suppliers and used as received: phosphotungstic acid hydrate
make the organic PAN polymer and inorganic NiO particles as pro- (H3PW12O40·nH2O), manganese (II) chloride (MnCl2), cesium chloride
spective catalytic materials and supports for immobilization of hetero- (CsCl), aniline, ammonium peroxydisulfate ((NH4)2S2O8), hydrogen
poly compounds. peroxide (H2O2, 30 vol%), glacial acetic acid (CH3COOH), thiophene
Petroleum products, from petrol (or gasoline) and kerosene to as- (Th), benzothiophene (BT), dibenzothiophene (DBT), and n-heptane
phalt, naturally abundant organic and inorganic sulfur compounds. The were obtained from Sigma-Aldrich. Also, nickel nitrate hexahydrate (Ni
high toxicity and corrosive nature of sulfur-containing impurities pose (NO3)2·6H2O) and citric acid monohydrate (C6H8O7·H2O) were ob-
serious environmental problems during the combustion of the trans- tained from Merck.
portation fuel [17–19]. In view of the strict legislation about fuel The Fourier transform infrared (FT-IR) spectra of the samples were
quality, a lot of substantial research efforts have been focused on re- recorded using a Thermo-Nicolet-iS10 spectrometer between 400 and
ducing the sulfur concentration of refined petroleum products in many 4000 cm−1 from KBr pellets. Ultraviolet-visible (UV–vis) spectra were
countries [20–22]. Several desulfurization technologies have been de- measured with a double beam Thermo-Heylos spectrometer. The
veloped such as hydro-desulfurization (HDS) [23], selective adsorption powder X-ray diffraction (XRD) of the materials was performed by a
desulfurization (ADS) [24], extractive desulfurization (EDS) [25], bio- Bruker D8 Advance X-ray diffractometer with monochromatic CuKα
desulfurization (BDS) [26], and oxidative desulfurization (ODS) [27] in radiation in the 2θ range of 10–80°. For SEM analysis, samples were
response to the need for clean fuel. Each of these techniques has its sputter coated with gold layers and images were taken by scanning
specific advantages and disadvantages. The HDS as a conventional electron microscopy (SEM) on LEO 1455 VP equipped with an energy
procedure is applied in many industries to reduce the sulfur content of dispersive X-ray (EDX) spectroscopy apparatus. The ultrasonic water
fuel to a level of ∼500 ppm [28]. This technology has been proven to bath was used a Bandelin Sonorex Digitec frequency 35 kHz mains
be a robust method in reducing hydrogen sulfides, thiols, mercaptans, connection 230 V. The content of total sulfur in gasoline and model fuel
and disulfides [29]. Nevertheless, it is unsuccessful to deep desulfur- were determined using X-ray fluorescence (XRF) with a TANAKA X-ray
ization because of the poor effect on refractory sulfur species such as fluorescence spectrometer RX-360 SH.
thiophene (Th), benzothiophene (BT), dibenzothiophene (DBT), and
their alkyl-substituted derivatives [30]. In addition, high temperature 2.2. Synthesis of PWMn
and high pressure conditions with large quantities of hydrogen are re-
quired during the HDS, which result in high costs and energy con- Amount of H3PW12O40·nH2O (2.88 g) was dissolved in 30 mL of
sumption [31]. The ADS is a non-invasive approach based on the ability distilled water. The pH was adjusted to 4.8 using the aqueous solution
of a solid sorbent to selectively adsorb the sulfur-containing compounds of NaOH (1 M), and then the solution was stirred at 90 °C for 1 h. After
(mostly DBT) from fuel. However, the results indicated that the hin- completing the reaction, a solution of MnCl2 (0.20 g) in 10 mL of dis-
dered derivatives of DBT such as 4,6-dimethyl dibenzothiophene cannot tilled water was added dropwise to the above solution with vigorous
be significantly removed [32]. Also, some interesting reviews have been stirring. Then, 10 mL of saturated solution of CsCl was prepared and
published on biocatalytic desulfurization due to its green processing added slowly to the mixture. The orange precipitate of Cs5PW11MnO39
[33]. Indeed, a series of bacterial species has been used to catalyze the (abbreviated as PWMn) was filtrated, washed thoroughly with water,
sulfur removal reactions through the consumption of DBT as their en- and dried at 80 °C for 2 h.
ergy source. One of the important aspects of BDS is the accumulation of
reaction products, which inhibit the microbial growth of bacteria and 2.3. Synthesis of PWMn/NiO
the removal of sulfur compounds [34]. Therefore, there is an urgent
need for inexpensive and energy-saving alternative methods to achieve In a typical procedure, Ni(NO3)2·6H2O (2.90 g) and C6H8O7·H2O
ultra-deep desulfurization. The application of oxidative desulfurization (1.90 g) were dissolved in 50 mL of distilled water, separately. Then,
(ODS) of liquid fuels has extensively studied in recent decades [35–37]. the Ni(NO3)2·6H2O solution was dripped in the C6H8O7·H2O solution.
It can be carried out under mild environmental conditions (low tem- The mixed solution was heated to 80 °C with stirring for 1 h. Then, the
perature and pressure) and economically viable. Also, the ODS capable synthesized PWMn (0.10 g) was dissolved in 5 mL boiling distilled
to easily remove of refractory sulfur compounds [38]. However, this water and added into the reaction vessel. After the removal of water
method still cannot overcome the challenge in terms of the high cost through evaporation, a green gel was formed. To prepare PWMn/NiO
due to the employment of a large amount of oxidants (e.g., H2O2). powder, the obtained gel was aged at 90 °C for 2 h and calcined at

776
M.A. Rezvani and O.F. Miri Chemical Engineering Journal 369 (2019) 775–783

400 °C for 4 h.

2.4. Synthesis of PWMn/NiO/PAN catalyst

The PWMn/NiO/PAN catalyst was synthesized by in situ chemical


oxidation polymerization of aniline monomer in the presence of
PWMn/NiO colloidal solution. In a typical synthesis process, 0.5 mL of
aniline and the powder of PWMn/NiO (0.50 g) were added to 30 mL
distilled water under ultrasonic conditions using an ultrasonic water
bath (frequency 35 kHz mains connection 230 V) for 30 min. The
polymerization reaction was initiated immediately after the addition of
(NH4)2S2O8 (1.10 g) to the above solution. The mixture was allowed to
react for 10 h under constant stirring. The dark green precipitate
(PWMn/NiO/PAN) was collected by filtration, washed with the water
and ethanol until the filtrate became clear, and finally dried at 70 °C for
4 h.

2.5. ODS process of model fuels and gasoline

In a typical reaction run, the catalysis ODS system was prepared as


follows: the different types of thiophenic compounds such as Th, BT,
and DBT as a source of sulfur were dissolved in n-heptane to represent
model fuels, separately. The initial concentration of each sulfur com-
pound in the model fuels was 500 ppm by weight. Then, prepared
model fuel (50 mL) was placed in a two-necked round-bottom flask
coupled in a temperature-controlled water bath. The appropriate
amounts of PWMn/NiO/PAN heterogeneous catalyst and 3 mL of aqu-
eous solution of hydrogen peroxide/acetic acid (in v/v ratio of 2:1) as
the oxidant were added to the flask. The reaction was proceeded under
vigorous stirring (700 rpm) at various temperatures (25, 30, 35, and
40 °C) and times (0–60 min) to evaluate the catalytic activity of the Fig. 1. FT-IR spectra of (a) PWMn, (b) NiO, (c) PAN, and (d) PWMn/NiO/PAN
catalyst. Afterward, the solution was cooled to room temperature, fol- nanocomposite.
lowed by addition of 10 mL of acetonitrile to extract the oxidation
products. The formed immiscible biphasic mixture was decanted by a [PW12O40]n−, the band of P–Oa is split into mainly two components,
separation funnel to separate the oil phase from the water phase. The one with a peak maximum at 1060 cm−1 and the other with a double
oil phase (n-heptane) was analyzed using the XRF spectrometer ac- band at 1050 cm−1 [43]. The results strongly imply that the structure of
cording to D-4294 and D-3227 ASTM analysis methods. the primary heteropoly phosphotungstate anion remains intact after
The ODS test of gasoline was conducted in the same manner as in assembling with the Mn2+ cation. In Fig. 1(b), the band at 496 cm−1
the way described above. In brief, 50 mL of gasoline was added to the corresponds to the vibration of NieO. The advent of the peaks at 1384
flask and heated to 35 °C. The catalyst (0.10 g) and oxidant (3 mL) were and 1031 cm−1 is indicated the formation of NiO particles [16]. The
poured into the reaction vessel. The mixture was stirred magnetically at bands at 1384 cm−1 may also be ascribed to physically adsorbed CO32−
the predefined temperature for 1 h. After cooling the solution, 10 mL of group on the surface of the NiO calcined powder [44]. The remarkable
acetonitrile was utilized as the extraction solvent. The gasoline phase reduction in the intensity of the O–H peak around 3448 cm−1 is due to
was separated from the aqua phase and analyzed by XRF for the de- the transformation of most of the hydroxyl groups to oxide. According
termination of the sulfur amount. The sulfur removal efficiency (%) was to the spectrum of PAN, the bands at 1576 and 1490 cm−1 are exhibited
calculated according to the following equation: the quinonoid (Q) and benzenoid (B) ring-stretching vibrations, re-
spectively. The absorption band at 1307 cm−1 indicates the CeN
C⎤
ΔC (\%) = ⎡1 − × 100 stretching vibration of a secondary aromatic amine. Moreover, the

⎣ C0 ⎥
⎦ (1)
broad and strong band located at 1148 cm−1 has been appointed to the
where C0 and C represent the concentration sulfur compounds before vibration mode of eNH+] structure, and it is associated with the vi-
and after ODS process, respectively. brations of Q]NH+eB or BeNH+%eB charged polymer units [45].
This indicated the presence of positive charges on the polymeric chain
3. Characterization of PAN, which make it as an ideal candidate for immobilization of
heteropolyanions. In the FT-IR spectrum of PWMn/NiO/PAN catalyst,
3.1. Characterization of the catalyst the observed bands at 1587, 1490, 1144, and 499 cm−1 are revealed
that the structure of PAN and NiO retained after incorporation
The FT-IR spectroscopic analysis was carried out to demonstrate the (Fig. 1(d)). Further, the peaks of PWMN in the range of 800–1100 cm−1
successful preparation of the PWMn/NiO/PAN catalyst. Fig. 1 displays are discernable maybe by covering with the bands of PAN. Never-
the sequence FT-IR spectra of (a) PWMn, (b) NiO, (c) PAN, and (d) theless, some shifts that are seen in the spectrum of the PWMn/NiO/
PWMn/NiO/PAN in the 400–4000 cm−1 wavenumber range. According PAN catalyst (at the wavenumbers of 499, 799, 880, 1144, and
to Fig. 1(a), the absorption peaks of the synthesized Keggin-type PWMn 1587 cm−1) compared to the pure samples confirmed the interaction
are shown bands at 1060, 954, 890, 830, and 510 cm−1, which are between materials and the successful preparation of the nanocomposite
assigned to νas(P–Oa), νas(W = Ot), νas(W–Ob–W), νas(W–Oc–W), and [46].
νas(W–O–Mn), respectively [18]. As the symmetry decrease of the PO4 The absorption spectra of (a) PWMn, (b) NiO, (c) PAN, and (d)
tetrahedron in the structure of [PW11MnO39]n− compared with PWMn/NiO/PAN are presented in Fig. 2. According to Fig. 2(a), the

777
M.A. Rezvani and O.F. Miri Chemical Engineering Journal 369 (2019) 775–783

Fig. 2. UV-vis spectra of (a) PWMn, (b) PAN, (c) NiO, and (d) PWMn/NiO/PAN
nanocomposite.

charge-transfer (CT) transition of O2− b/c → W


6+
in the structure of het-
eropoly compound induced an absorption peak at 323 nm. In the lower-
energy region, there is an absorption band at 205 nm which is assigned
to the O → P transition. Note that the Mn(II) in the PWMn does not
show remarkable absorption peak owing to its CT transitions emerge in
the higher energy region [47]. As can be seen in Fig. 2(b), the spectrum
of PAN polymer displays two characteristic absorption bands which are
placed at 240 nm and 317 nm. The absorption band around 240 nm is Fig. 3. XRD patterns of (a) PWMn, (b) NiO, (c) PAN, and (d) PWMn/NiO/PAN
assigned to π → π* transition of the benzenoid ring and the second one nanocomposite.
is corresponded to n → π* excitation transition of the quinonoid ring
[48]. The UV–vis spectrum of NiO particles is shown in Fig. 2(c). It is where L is the average crystallite size (nm), λKα1 is the X-ray wave-
believed that the bands at 320 and 355 nm are caused by CT transitions length (0.15406 nm), β(2θ) is full peak width at half max (FWHM) in
from the valence to the conduction band, O2− (2p) → Ni(3d), in the NiO
2+
radians, and θmax is the half of the diffraction angle.
[49]. In Fig. 2(d), the observed slight shifts in the spectrum of PWMn/ SEM as a surface imaging method was accomplished to provide
NiO/PAN indicated the interactions between the inorganic metal oxides morphological details of the materials. The rod-shaped particles of
and organic polymer as well as the confirmation of FT-IR studies [50]. PWMn with the lengths of 1–3 μm and diameters of 50–100 nm can be
The successful synthesis of PWMn/NiO/PAN nanocomposite was seen in Fig. 4(a). Regarding the electronic microscopy studies, the
further proved by the XRD method. The XRD pattern of PWMn was image of NiO illustrates a rough surface which composed of some small
matched well with the Keggin structure as indexed with the JCPDS 00- spherical particles (Fig. 4(b)). In Fig. 4(c), the SEM image of PAN
050-0654 file. The typical peaks at 2θ = 28.5° and 40.6° in XRD pattern clearly implies that the polymer has a small sponge-like morphology.
of PWMn indicate that the synthesized mono-substituted phospho- Compared to the morphology of the bulk materials, the PWMn/NiO/
tungstate has retained its Keggin structure (Fig. 3(a)). Also, the for- PAN composite indicates the well-uniform nanoscale spherical parti-
mation of [PW11MnO39]n− compound instead of [PW12O40]n− leads to cles. The corresponding particle size distribution of the nanosphere
the additional peaks [51]. Notably, the diffraction peak at 2θ = 28.50° composite is presented in Fig. 5(a). It is clear that the size of PWMn/
with the highest intensity indicated that H+ was partially substituted NiO/PAN particles is in the range of 50–60 nm. According to the EDX
with Cs+ [52]. According to Fig. 3(b), the diffraction peaks at analysis, the successful preparation of PWMn/NiO/PAN is confirmed by
2θ = 37.3, 43.2, 62.8, 75.3, and 79.17° can be indexed to (1 0 1), the existence of C, O, P, Mn, Ni, and W elements with an approximated
(0 1 2), (1 1 0), (1 1 3), and (2 0 2) crystallographic plane of NiO (JCPDS percentage quantity of 70.65, 22.42, 0.30, 0.42, 2.27, and 3.94%, re-
04-0835), respectively [16]. The XRD pattern of PAN exhibited only a spectively (Fig. 5(b)).
broad peak at 2θ = 25°, indicating that the amorphous nature of the
precursor (Fig. 3(c)) [14,15]. After immobilization of PWMn/NiO on
3.2. ODS results of model fuels and gasoline
the PAN, the XRD amorphous profile of PAN remained almost the same
(Fig. 3(d)). In addition to the diffraction peaks of NiO, no diffraction
In this work, the ODS experiments were investigated using the
peak of PWMn substance is clearly detected in the profile of PWMn/
synthesized catalyst for removal of sulfur compounds from model fuels
NiO/PAN composite. It is probably due to the fine dispersion of het-
and gasoline. About 0.10 g of the PWMn/NiO/PAN nanosphere com-
eropoly compound on the surface of NiO and PAN. This observation is
posite and 50 mL of model fuel (500 ppmw of Th, BT, and DBT in n-
consistent with the reports in the literature [53]. The average crystallite
heptane) were mixed with 3 mL of oxidant under the conditions men-
size of the PWMn/NiO/PAN was estimated to be about 14.50 nm using
tioned in the experimental section. In this regard, the desulfurization
the Debye-Scherrer formula as follows:
efficiency of Th, BT, and DBT could reach 97, 98, and 98% at 35 °C after
1 h, respectively. Also, it can be concluded from the Table 1 that the
0.9λ K α1
L= removal efficiency (%) of sulfur content and mercaptans in gasoline
β(2θ) cos θmax (2) reached 97%, demonstrating that PWMn/NiO/PAN acted as an

778
M.A. Rezvani and O.F. Miri Chemical Engineering Journal 369 (2019) 775–783

Fig. 4. SEM image of (a) PWMn, (b) NiO, (c) PAN, and (d) PWMn/NiO/PAN nanocomposite.

Fig. 5. (a) The particle size distribution histogram and (b) EDX analysis of the PWMn/NiO/PAN nanocomposite.

Table 1
ODS results of gasoline by PWMn/NiO/PAN catalyst.
a
Entry Properties of gasoline Unit Method Before ODS After ODS

1 Total Sulfur content by XRF wt% ASTM D 4294 0.4989 0.0133


2 Mercaptan compounds ppm ASTM D 3227 97 3
3 Density by hydrometer @ 15 °C g/ml ASTM D 1298 0.7996 0.7995
4 Salt ptb ASTM D 3230 18 18
5 Water content by distillation vol.% ASTM D 4006 Nil. Nil.
6 Distillation IBP °C ASTM D 86 49.7 49.6
FBP 208.5 208.4
10 vol.% 68.7 68.6
50 119.8 119.6
90 187.4 187.3
95 206.9 206.8

a
Condition of ODS process: 50 mL of gasoline, 0.10 g of catalyst, 3 mL of oxidant, 10 mL of acetonitrile, reaction temperature = 35 °C, and reaction time = 1 h.

779
M.A. Rezvani and O.F. Miri Chemical Engineering Journal 369 (2019) 775–783

Table 2
Effect of different catalysts on the ODS of gasoline and model fuels.a
Entry Catalyst ODS efficiency (%)

Th BT DBT Gasoline Ref.

1 PWMn/NiO/PAN 97 98 98 97 In this work


2 [Bmim]PW/HMS – 79 98 – [51]
3 HPW-TiO2-SiO2 – – 96 – [52]
4 CTAB-PTA@CS 92 93 95 94 [39]
5 [C16mim]3PW12O40/SiO2 – 85 100 – [53]
6 PWMn 68 69 71 67 In this work
7 K3[PW12O40] 58 59 60 58 [22]
8 K4[SiW12O40] 56 56 58 55 [22]
9 NiO 48 46 45 49 In this work
10 PAN 38 34 35 34 In this work
11 None (blank test) 13 15 16 13 In this work

a
Condition of ODS process: 50 mL of gasoline or model fuel (500 ppmw of
Th, BT, and DBT in n-heptane), 0.10 g of catalyst, 3 mL of oxidant, 10 mL of
acetonitrile, reaction temperature = 35 °C, and reaction time = 1 h.

Fig. 6. Effect of catalyst dosage on the ODS of gasoline and model fuels by
PWMn/NiO/PAN catalyst. Condition of ODS process: 50 mL of gasoline or
model fuel (500 ppmw of Th, BT, and DBT in n-heptane), 0.10 g of catalyst,
3 mL of oxidant, 10 mL of acetonitrile, reaction temperature = 35 °C, and re-
action time = 1 h.

effective heterogeneous catalyst for deep desulfurization. It is noted


that the other specifications of gasoline (e.g. density or water content)
were remained largely unchanged after ODS treatment compared to
their corresponding initial values.
In order to investigate the effect of different catalysts, a series of
ODS reactions were carried out. The comparative results of catalytic
activity of PWMn, NiO, PAN, PWMn/NiO/PAN, and some different
types of catalysts were listed in Table 2. According to the blank test Fig. 7. Effect of reaction temperature and reaction time on the ODS of (a) Th,
(without catalyst), the results clearly indicated that the removal effi- (b) BT, and (c) DBT by PWMn/NiO/PAN catalyst.
ciencies of Th, BT, and DBT were examined 13, 15, and 16%, respec-
tively (Table 2, Entry 11). Also, 13% of the sulfur content of real ga- Table 3
soline was lowered. It is shown that the hydrogen peroxide/acetic acid Effect of different oxidation systems on the ODS of gasoline and model fuelsa.
oxidant has a weak influence in the oxidation step. The PWMn and NiO
Entry Acid used ODS efficiency %
catalysts were more active than PAN within the same reaction period.
Apparently, the PWMn/NiO/PAN was the most effective catalyst among Th BT DBT Gasoline
the reported ones [54–56]. The results suggested that the immobiliza-
1 Acetic acid 97 98 98 97
tion of heteropoly compound on metal oxide within a polymer matrix
2 Benzoic acid 90 91 91 88
offer attractive routes for combining properties stemming from metal 3 Oxalic acid 85 87 89 84
oxides and from that of polymers. 4 Sulfuric acid 69 70 72 68
The optimization of dosage of the catalyst was executed, essentially 5 Carbonic acid 67 69 71 66
by varying the amount of PWMn/NiO/PAN from 0 to 0.12 g in the ODS a
Condition of ODS process: 50 mL of gasoline or model fuel (500 ppmw of
systems. Fig. 6 displays the comparison of desulfurization efficiency
Th, BT, and DBT in n-heptane), 0.10 g of PWMn/NiO/PAN catalyst, the mixture
data obtained for the Th, BT, DBT, and the sulfur content of gasoline
of hydrogen peroxide/acid in v/v ratio of 2:1, 10 mL of acetonitrile, reaction
using different catalyst dosage. It is noted that the sulfur level of model temperature = 35 °C, and reaction time = 1 h.

780
M.A. Rezvani and O.F. Miri Chemical Engineering Journal 369 (2019) 775–783

Table 4 oxidizing agents. So, they are capable to oxidize facilely the sulfur-
Reusability performance of PWMn/NiO/PAN cat- containing compounds without forming a considerable amount of re-
alyst for the ODS of DBT.a sidual [38].
Run ODS efficiency (%) The reusability performance of PWMn/NiO/PAN catalyst was car-
ried out under optimized conditions for the ODS of DBT as the model
1 98 component. The catalyst was separated from the reaction media at the
2 97
end of the oxidation process by simple filtration, washed with the di-
3 97
4 96 chloromethane solvent, and then dried at 70 °C for 2 h. The recycled
5 95 catalyst was processed for the consequent reaction using the fresh
oxidant and extraction solvents. After each catalytic run, the oil phase
a
Condition of ODS process: 50 mL of model fuel (n-heptane) was analyzed to determine the DBT concentration. The data
(500 ppmw of DBT in n-heptane), 0.10 g of cata- listed in Table 4 revealed that there is no obvious decrease in the ac-
lyst, 3 mL of oxidant, 10 mL of acetonitrile, reac-
tivity of the catalyst even after five successive runs. It could be clarified
tion temperature = 35 °C, and reaction time = 1 h.
by the electrostatic interaction between the functional groups (oxygen
and amine groups) of support materials and [PW11MnO39]n− ions that
fuels and gasoline was reduced drastically (from less than 20% to up to
it prevents from leaching of PWMn species [58]. The slight decrease in
95%) with increasing catalyst dosage. Although the efficiency of the
the removal efficiency of DBT from 98 to 95% may be related to the
process did not increase when the dosage of PWMn/NiO/PAN increased
adsorption amounts of dibenzothiophene sulfoxide (DBTO) and/or
to 0.11 and 0.12 g. No increase in the ODS efficiency at high dosage
sulfone (DBTO2) on the surface of the catalyst. On the basis of the above
(> 0.10 g) might be associated with the accumulation of the catalyst
evidence, it can be concluded that the designed PWMn/NiO/PAN na-
particles in the oil media that hinders the adsorption and oxidation of
nosphere composite is a promising heterogeneous catalyst in terms of
sulfides. The optimum PWMn/NiO/PAN dosage was found to be 0.10 g
catalytic activity and reusability.
for the forthcoming ODS experiments.
The influence of temperature on the ODS process was also studied
(Fig. 7). When the processing temperature was adjusted at 25 °C, the
3.3. Probable mechanism of ODS process
desulfurization efficiencies of Th, BT, and DBT were approximately
75% after 1 h. It could be found that the sulfur compounds were almost
The probable two-step mechanism of ODS process catalyzed by the
eliminated when the temperature was raised. However, increasing the
PWMn/NiO/PAN catalyst has been proposed as follows: before the
temperature from 35 to 40 °C did not have a significant effect on the
ODS, the sulfur-containing compounds are in the non-polar oil phase (n-
rate of reactions. This may explain the phenomena that the high tem-
heptane or gasoline). After the addition of the heterogeneous catalyst
perature will cause the decomposition of hydrogen peroxide, and the
and oxidant, the oxidation reactions are performed in the intermediate
decrease of the concentration of M(O2)n complex [57]. On the other
phase. The catalytic sulfur oxidation mechanism using POM-based
hand, the highest sulfur removal efficiency was recorded after 1 h. In
catalysts was well studied [40–42]. In all cases, the peroxometal in-
view of economizing the energy, reaction temperature and time were
termediate complexes, M(O2)n, were introduced as responsible species
set to 35 °C and 1 h, respectively, as optimum conditions of ODS sys-
for oxidation of sulfides. As shown in Scheme 1, the reaction of the
tems.
peroxide oxygen in peroxyacids with terminal oxygen-metal groups
The effect of different oxidation systems on the ODS rate of sulfur
(M = Ot) in the structure of Keggin-type PWMn led to the formation of
compounds was investigated using the PWMn/NiO/PAN catalyst at
the M(O2)n complexes (step 1). The sulfur atom of organic compounds,
35 °C after 1 h. Hence, the various oxidation reactions were carried out
which is quite nucleophilic, can attack the M(O2)n species to produce
in the presence of different organic and inorganic acids such as acetic
the corresponding sulfoxide and sulfones (step 2 and 3). These reaction
acid, oxalic acid, benzoic acid, sulfuric acid, and carbonic acid. The
products accumulate in the water phase due to their nature of polarity,
results in Table 3 indicated that poor efficiency was caused by carbonic
which can be considered to be a benefit for the extraction step by
acid, while, the deep desulfurization achieved by acetic acid. In fact,
acetonitrile solvent (step 4).
using the organic acids can enhance the ODS efficiency due to the in situ
production of peroxyacids (also known as peracids), RCO3H, as strong

Scheme 1. Schematic illustration of the probable ODS mechanism of DBT as a sulfur-containing compound catalyzed by PWMn/NiO/PAN catalyst.

781
M.A. Rezvani and O.F. Miri Chemical Engineering Journal 369 (2019) 775–783

Fig. 9. Arrhenius plots for the ODS of (a) Th, (b) BT, and (c) DBT.
Fig. 8. Plots of C/C0 against time for the ODS of (a) Th, (b) BT, and (c) DBT.

3.4. Kinetic study of ODS process


Table 5
Pseudo-first-order rate constants and correlation factors of the ODS of sulfur
compounds. To evaluate the kinetics of the oxidation of sulfur compounds, the
2
values of the apparent reaction rate constant k (min−1) were de-
Temperature (°C) Rate constant (k) Correlation factor (R )
termined on the basis of the pseudo-first-order kinetic model as follows:
Th BT DBT Th BT DBT
d [C ]
r= = −k [C ]
25 0.020 0.020 0.022 0.9279 0.9439 0.9513 dt (3)
30 0.025 0.026 0.030 0.8818 0.9151 0.9576
35 0.045 0.047 0.049 0.9313 0.9174 0.9048 If [C]0 and [C]t are respectively assumed as concentration of sulfur
40 0.050 0.049 0.054 0.9432 0.9477 0.9533
compound at t = 0 and t = t, the k can be found by using integrated
first-rate law:

782
M.A. Rezvani and O.F. Miri Chemical Engineering Journal 369 (2019) 775–783

Ct d [C ] t Polymers 6 (2014) 2435–2450.


∫C 0 [C ]
=− ∫0 kdt
(4) [15] S.K. Pillalamarri, F.D. Blum, A.T. Tokuhiro, M.F. Bertino, Chem. Mater. 17 (2005)
5941–5944.
[16] M.A. Rezvani, S. Khandan, M. Aghmasheh, Taiwan. Inst Chem. Eng. J. 77 (2017)
C C
ln ⎡ t ⎤ = −kt ⇒ ⎡ t ⎤ = e−kt 321–328.
⎣ C0 ⎥
⎢ ⎦ ⎣ C0 ⎥
⎢ ⎦ (5) [17] H. Ji, J. Sun, P. Wu, B. Dai, Y. Chao, M. Zhang, W. Jiang, W. Zhu, H. Li, J. Mol.
Catal. A: Chem. 423 (2016) 207–215.
The variations of Th, BT, and DBT concentration [C]/[C0] with t [18] M.A. Rezvani, Z. Shokri Aghbolagh, H.H. Monfared, S. Khandan, J. Ind. Eng. Chem.
were shown in Fig. 8. In this study, the correlation coefficients were 52 (2017) 42–50.
[19] Y. Li, H. Yi, X. Tang, X. Liu, Y. Wang, B. Cui, S. Zhao, Chem. Eng. J. 304 (2016)
found close to unity (Table 5). It was clearly observed that ODS systems 89–97.
followed pseudo-first-order kinetics in all cases. Also, the apparent ac- [20] S.Y. Haruna, U.Z. Faruq, A.Y. Zubairu, M.G. Liman, M.L. Riskuwa, Am. J. Appl.
tivation energies (Ea) for the ODS of sulfur molecules were calculated Chem. 6 (2018) 15–24.
[21] Z. Hasan, S.H. Jhung, A.C.S. Appl, Mater. Interfaces 7 (2015) 10429–10435.
from Arrhenius plots (Eq. (6) and Fig. 9). The Ea values were 51.80, [22] M.A. Rezvani, M. Shaterian, F. Akbarzadeh, S. Khandan, Chem. Eng. J. 333 (2018)
50.98 and 47.40 kJ/mol for Th, BT, and DBT, respectively. 537–544.
[23] J. Kou, C. Lu, W. Sun, L. Zhang, Z. Xu, ACS Sustain. Chem. Eng. 3 (2015)
−Ea −Ea 1 3053–3061.
k = Ae RT ⇒ ln k = ⎛ ⎞ + ln A
R ⎝T ⎠ (6) [24] I. Ahmed, S. Hwa Jhung, J. Hazard. Mater. 301 (2016) 259–276.
[25] K.H. Almashjary, M. Khalid, S. Dharaskar, P. Jagadish, R. Walvekar,
T.C.S. Manikyam Gupta, Fuel 234 (2018) 1388–1400.
4. Conclusion [26] S. Labana, G. Pandey, R.K. Jain, Lett. Appl. Microbiol. 40 (2005) 159–163.
[27] D. Wang, N. Liu, J. Zhang, X. Zhao, W. Zhang, M. Zhang, J. Mol. Catal. A: Chem.
393 (2014) 47–55.
In summary, the PWMn/NiO/PAN hybrid material was prepared [28] Y. Qin, S. Xun, L. Zhan, Q. Lu, M. He, W. Jiang, H. Li, M. Zhang, W. Zhu, H. Li, New
successfully and developed to catalyze the ODS treatments. The nano- J. Chem. 41 (2017) 569–578.
[29] X.D. Tang, T. Hu, J.J. Li, F. Wang, D.Y. Qing, RSC Adv. 5 (2015) 53455–53461.
sphere composite exhibited outstanding performance for the oxidation [30] F. Mirante, S.O. Ribeiro, B. De Castro, C.M. Granadeiro, S.S. Balula, Top. Catal. 60
of organic sulfur-containing molecules in gasoline and model fuels (2017) 1140–1150.
under mild conditions. The removal efficiency of the sulfur content and [31] M.A. Rezvani, S. Khandan, N. Sabahi, Energy Fuels 31 (2017) 5472–5481.
[32] A. Srivastav, V.C. Srivastava, J. Hazard. Mater. 170 (2009) 1133–1140.
the mercaptan compounds of gasoline were achieved to 97% within 1 h
[33] K.A. Gray, G.T. Mrachko, C.H. Squires, Curr. Opin. Microbiol. 6 (2003) 229–235.
at 35 °C. The results showed that the hydrogen peroxide/acetic acid was [34] M. Soleimani, A. Bassi, A. Margaritis, Biotechnol. Adv. 25 (2007) 570–596.
likely to play a role as the oxidant in the activation of heteropoly [35] E. Rafieea, N. Nobakht, J. Mol. Catal. A: Chem. 398 (2015) 17–25.
[36] M.A. Rezvani, M. Ali Nia Asli, M. Oveisi, R. Babaei, K. Qasemi, S. Khandan, RSC
compound and formation of M(O2)n complex. Further, the aromatic
Adv. 6 (2016) 53069–53079.
sulfurs such as Th, BT, and DBT could be removed from n-heptane with [37] L. Shen, G. Lei, Y. Fang, Y. Cao, X. Wang, L. Jiang, Chem. Commun. 54 (2018)
the yield of 97, 98, and 98%, respectively. The report indicated that the 2475–2478.
kinetics of oxidative processes followed the pseudo-first-order model. [38] A. Fallah Shojaei, M.A. Rezvani, M.H. Loghmani, Fuel Process. Technol. 118
(2014) 1–6.
Moreover, the PWMn/NiO/PAN heterogeneous catalyst showed ex- [39] P. Paniv, S. Pysh’ev, V. Gaivanovich, O. Lazorko, Chem. Tech. Fuels Oils 42 (2006)
cellent reusability properties for five oxidation runs. Despite the results 159–166.
attended in this research, we envision that intensified investigation [40] R. Frenzel, D. Morales, G. Romanelli, G. Sathicq, M. Blanco, L. Pizzio, J. Mol. Catal.
A: Chem. 420 (2016) 124–133.
would open a new chapter of research that brings more reactions to the [41] J. Xiao, L. Wu, Y. Wu, B. Liu, L. Dai, Z. Li, Q. Xia, H. Xi, Appl. Energy 113 (2014)
benefit of organic–inorganic hybrid catalysis for ODS process. 78–85.
[42] M.A. Rezvani, M. Alinia Asli, S. Khandan, H. Mousavi, Z. Shokri Aghbolagh, Chem.
Eng. J. 312 (2017) 243–251.
References [43] K. Patel, P. Shringarpure, A. Patel, Supramol. Chem. 24 (2012) 149–156.
[44] A. Adekunle, J. Oyekunle, O.S. Abiodun, A.O. Joshua, W.O. Makinde,
[1] H. Zheng, C. Wang, X. Zhang, Y. Li, H. Ma, Y. Liu, Appl. Catal. B: Environ. 234 A. Ogunfowokan, M.A. Eleruja, E. Ebenso, Int. J. Electrochem. Sci. 9 (2014)
(2018) 79–89. 3008–3021.
[2] Y. Han, K. Choi, H. Oh, C. Kim, D. Jeon, C. Lee, J.H. Lee, J. Ryu, J. Catal. 367 (2018) [45] M. Trchová, J. Stejskal, Pure Appl. Chem. 83 (2011) 1803–1817.
212–220. [46] A.G. Yavuz, A. Uygun, V.R. Bhethanabotla, Carbohydr. Polym. 75 (2009) 448–453.
[3] M.A. Rezvani, M. Oveisi, M. Ali Nia Asli, J. Mol. Catal. A: Chem. 410 (2015) [47] P. Shringarpure, K. Patel, A. Patel, J. Clust. Sci. 22 (2011) 587–601.
121–132. [48] P. Bober, P. Humpolíček, T. Syrový, Z. Capáková, L. Syrová, J. Hromádková,
[4] D.Y. Fu, S. Zhang, Z. Qu, X. Yu, Y. Wu, L. Wu, A.C.S. Appl, Mater. Interfaces 10 J. Stejskal, Synth. Met. 232 (2017) 52–59.
(2018) 6137–6145. [49] A.A. Ahmed, M. Devarajan, N. Afzal, Sens. Actuators, A. 262 (2017) 78–86.
[5] X.M. Luo, N.F. Li, Z.B. Hu, J.P. Cao, C.H. Cui, Q.F. Lin, Y. Xu, Inorg. Chem. 58 [50] P. Patil, S. Kondawar, Procedia Mater. Sci. 10 (2015) 195–204.
(2019) 2463–2470. [51] P. Shringarpure, B.K. Ripuramallu, K. Patel, A. Patel, J. Coord. Chem. 64 (2011)
[6] A.M. Escobar Caicedo, J.A. Rengifo-Herrera, P. Florian, M.N. Blanco, 4016–4028.
G.P. Romanelli, L.R. Pizzio, J. Mol. Catal. A: Chem. 425 (2016) 266–274. [52] G. Liu, J. Xu, J. Jiang, B. Peng, X. Wang, Int. J. Hydrogen Energy 39 (2014)
[7] M.A. Rezvani, S. Khandan, Appl. Organometal. Chem. e4524 (2018) 1–13. 1914–1923.
[8] J.M. Brégeault, M. Vennat, L. Salles, J.Y. Piquemal, Y. Mahha, E. Briot, P. Célestin [53] X. Zhang, G. Luo, M. Zhu, L. Kang, F. Yu, B. Dai, RSC Adv. 5 (2015) 76182–76189.
Bakala, A. Atlamsani, R. Thouvenot, J. Mol. Catal. A: Chem. 250 (2006) 177–189. [54] L. Tang, G. Luo, L. Kang, M. Zhu, B. Dai, J. Chem. Eng. 30 (2013) 314–320.
[9] Y. Zhou, G. Chen, Z. Long, J. Wang, RSC Adv. 4 (2014) 42092–42113. [55] X.M. Yan, P. Mei, L. Xiong, L. Gao, Q.F. Yang, L.J. Gong, Catal Sci. Technol. 3
[10] Y.F. Song, R. Tsunashima, Chem. Soc. Rev. 41 (2012) 7384–7402. (2013) 1985–1992.
[11] H. Rezaei Ghalebi, S. Aber, A. Karimi, J. Mol. Catal. A: Chem. 415 (2016) 96–103. [56] M. Li, M. Zhang, A. Wei, W. Zhu, S. Xun, Y. Li, H. Li, H. Li, J. Mol. Catal. A: Chem.
[12] A. Patel, N. Narkhede, S. Singh, S. Pathan, Catal. Rev. Sci. Eng. 58 (2016) 337–370. 406 (2015) 23–30.
[13] Z.E.A. Abdalla, B. Li, Chem. Eng. J. 200–202 (2012) 113–121. [57] L. Tang, G. Luo, M. Zhu, L. Kang, B. Dai, J. Ind. Eng. Chem. 19 (2013) 620–626.
[14] A.M. Meftah, E. Gharibshahi, N. Soltani, W. Mahmooud Mat Yunus, E. Saion, [58] A.K. Dizaji, H.R. Mortaheb, B. Mokhtarani, Chem. Eng. J. 335 (2018) 362–372.

783

You might also like