Download as pdf or txt
Download as pdf or txt
You are on page 1of 53

Journal Pre-proof

Synthesis of boron nitride nanorod and its performance as a metal-free catalyst for
oxidative desulfurization of diesel fuel

Tanaz Ghanadi, Gholamreza Moradi, Alimorad Rashidi

PII: S1004-9541(24)00034-X
DOI: https://doi.org/10.1016/j.cjche.2023.08.013
Reference: CJCHE 3119

To appear in: Chinese Journal of Chemical Engineering

Received Date: 21 November 2022


Revised Date: 18 May 2023
Accepted Date: 14 August 2023

Please cite this article as: T. Ghanadi, G. Moradi, A. Rashidi, Synthesis of boron nitride nanorod and
its performance as a metal-free catalyst for oxidative desulfurization of diesel fuel, Chinese Journal of
Chemical Engineering, https://doi.org/10.1016/j.cjche.2023.08.013.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2024 The Chemical Industry and Engineering Society of China, and Chemical Industry Press Co., Ltd.
All rights reserved.
Synthesis of boron nitride nanorod and its performance as a metal-free catalyst
for oxidative desulfurization of diesel fuel

Tanaz Ghanadi a
a
Catalyst Research Center, Faculty of Chemical and Petroleum Engineering, Razi University,
Kermanshah, Iran
E-mail: tanazghanadi@yahoo.com

Gholamreza Moradi a*

f
a
Catalyst Research Center, Faculty of Chemical and Petroleum Engineering, Razi University,

oo
Kermanshah, Iran
E-mail: gmoradi@razi.ac.ir, moradi_m@yahoo.com

r
-p
Alimorad Rashidi b
re
b
Nanotechnology Research Center, Research Institute of Petroleum Industry (RIPI), Tehran,
lP

Iran
Email:Rashidiam@ripi.ir
na
ur

Corresponding author: Gholamreza Moradi


Postal address: Catalyst Research Center, Faculty of Chemical and Petroleum Engineering,
Jo

Razi University, Kermanshah, Iran


E-mail: gmoradi@razi.ac.ir, moradi_m@yahoo.com (Prof. G.R. Moradi)
Synthesis of boron nitride nanorod and its performance as a metal-free catalyst for oxidative
desulfurization of diesel fuel
Tanaz Ghanadi 1, Gholamreza Moradi 1,*, Alimorad Rashidi2
1
Catalyst Research Center, Faculty of Chemical and Petroleum Engineering, Razi University, Kermanshah, Iran
2
Nanotechnology Research Center, Research Institute of Petroleum Industry (RIPI), Tehran, Iran

Abstract

In order to reduce the sulfur compounds in diesel fuel, boron nitride (BN) has been employed
as a novel metal free catalyst in the present research. This nanocatalyst was synthesized via
template-free approach followed by heating treatment at 900 ℃ in nitrogen atmosphere that the
characteristics of the sample were identified by the XRD, FTIR, Raman Spectroscopy, FESEM
TEM, AFM and N2 adsorption–desorption isotherms. The results of structural and

f
morphological analysis represented that BN has been successfully synthesized. The efficacy of

oo
the main operating parameters on the process was studied by using response surface
methodology based on the Box–Behnken design (BBD) method. The prepared catalyst showed

r
high efficiency in oxidative desulfurization of diesel fuel with initial sulfur content of 8040
-p
ppm S. From statistical analysis a significant quadratic model was obtained to predict the sulfur
removal as a function of efficient parameters. The maximum efficiency of 72.4% was achieved
re
under optimized conditions at O/S molar ratio of 10.2, temperature of 71 ℃, reaction time of
lP

113 min and catalyst dosage 0.36 g. Also, the reusability of the BN was studied and the result
showed little reduction in activity of the catalyst after 10 times regeneration. Moreover, a
plausible mechanism was proposed for oxidation of sulfur compounds on the surface of the
na

catalyst. The present study shows that BN materials can be selected as promising metal-free
catalysts for desulfurization process.
ur

Graphical abstract
Jo

DBT
Diesel Fuel

H2O2
Clean Diesel Fuel

B B B
N N N N
B B B B
N N N N
B B B B
N N N

DBTO2

BN Catalyst

10 μm

Keywords: Desulfurization, BN nanostructure, Experimental design, Box–Behnken

1
1. Introduction

Combustion of the fuels with high sulfur content makes the environmental risks due to emission
of the noxious sulfur oxides into the atmosphere, which can cause the air pollution. Moreover,
it raises other problems such as catalyst deactivation during oil processing and corrosion
problems in refining equipment [1]. So, the removing of these materials is essential for
production of green fuels and this issue creates a challenge to the refinery industry [2,3]. In
order to remove sulfur compounds from crude oil and commercial fuels several techniques like
hydrodesulfurization (HDS), oxidation desulfurization (ODS), biological desulfurization
(BDS), adsorption desulfurization (ADS) and extraction desulfurization (EDS) have been
applied in the oil industries [4].

HDS is the common industrial process to remove organosulfur compounds from the fuels but
this method has some problems such as high energy consumption, severe operating conditions

f
oo
and also it's not efficient for removing refractory sulfur compounds like benzothiophene (BT),
dibenzothiophene (DBT) and their derivatives [5-7]. Therefore, developing the alternative or
complementary method for deep desulfurization of fuels is of great interest [8]. In recent years,

r
-p
among the alternative approaches, oxidative desulfurization is one of the best methods due to
its mild reaction condition, no hydrogen requirement, low cost and high efficiency in
re
eliminating refractory sulfur compounds [9-12]. Through this process, the sulfur-bearing
molecules are oxidized to sulfones with a suitable oxidant in the presence of the catalyst and
lP

after that they can be easily removed from oil phase by solvent extraction or absorption
approach [13]. Selecting an appropriate oxidant is a key factor in terms of the cost and
na

environmental aspects. It has been shown that hydrogen peroxide is a powerful oxidant with
high active oxygen content and environmentally safe properties. Also, the best result can be
obtained by using it with heterogeneous catalysts [14-16].
ur

In primary investigations, homogenous catalysts were extensively employed in the ODS


Jo

process due to their high efficiency, but these catalysts have several defects including
deactivation during the reaction, difficulty in recovering and recycling expensive reagents
[17,18]. Accordingly, using the metal containing species as heterogeneous catalysts has been
extended in the ODS process, which include transition metals and metal oxides such as Mo,
Ni, Co, W, Cr, Ag, Ti, MoOx, ZrO2, TiO2, V2O5. Based on the studies, these materials
demonstrate high activity in the oxidation of sulfur compounds [19-21]. Whiles, employing
these metal-based catalysts has some inevitable deficiencies in terms of economic,
environmental protection and so on.

Metal-free catalysts are considered as promising classes of catalytic materials because of


having the advantages of the excellent catalytic activity, high stability, and environmentally
friendly characteristics [22]. Thus, developing metal-free catalysts like graphene oxide, carbon
nanotube, polymeric carbon nitride and etc., is the main issue to overcome the problems of
metal based catalysts [23,24]. In a research, reduced graphene oxide (rGO) was used as a
catalyst for removing dibenzothiophene (DBT) from the model fuel. The results showed that
carbonyl groups are proved to be active sites of catalyst, which caused the high catalytic activity
in the reaction [25]. Boron nitride, as the other metal free material, has been attracted great

2
attention because of its analogy with carbon materials [26]. For instance, it's cubic shape named
as c-BN presents the same structure as diamond crystal arrangement and the hexagonal form
of BN (h-BN) is similar to graphene [27,28]. The unique properties of BN materials such as
oxidative resistance, chemical and thermal stability, wide band gap and high specific surface
area, represent these compounds as useful and efficient 2D-materials [29,30]. Therefore, they
have been used in a broad range of fields such as adsorption, energy storage, gas separation
and catalysis [31-34]. For example, Oh et al. illustrated that h-BNNSs have a good adsorption
capacity towards Cu(II) and Ni(II) ions and as a result they are great adsorbent for efficient
heavy metal removal in aqueous system [35]. In another study, Zhang et al. reported effective
catalytic activity of boron nitride in oxidative dehydrogenation of propane and studied the
reaction mechanism. Their findings recommended highly selective catalysts for the alkane
ODH process [36].

f
In this context, a number of researches revealed that BN-based porous materials exhibit great

oo
activity in desulfurization from model oil [37]. In a work, using BN as a catalyst in removing
DBT from decalin was investigated. The h-BNNs have been prepared by employing various

r
solvents to verify the influence of solvent boiling point on the catalyst structure. BN materials
-p
synthesized with low boiling point solvents showed high catalytic performance in operation
conditions of reaction temperature 150°C; the air flow rate of 100 ml·min-1, catalyst dosage of
re
0.1g and the process time of 180 min [38].
lP

In another research, gas exfoliation was applied as a practical approach to generate the layered
structures of BN and create the edge sites. This method can enhance the active sites of the
na

catalyst by reducing the lateral size and making porosity which causes elevating desulfurization
performance of the BN. The number of exfoliation cycles has a positive effect on the sulfur
removal, so that the efficiency of BN NSs under the same reaction conditions after 5, 10, and
ur

20 cycles of exfoliation were 49.5, 89.3, and 98.2%, respectively. Additionally, it was exhibited
that high efficiency was obtained by employing BN NSs for eliminating the bulky RS
Jo

compounds such as 4,6-DMDBT and 4-MDBT from the model oil after 8 h [39].

It can be noted that structural features like defects have expected to play an important role in
determining the activity of BN for the ODS. Lv et al. performed a theoretical investigation to
explore the effect of possible defect sites of BN on Th oxidation. To achieve this goal, various
defect sites comprising zigzag B, zigzag N, armchair edge sites, and B- or N-monovacancy
sites were considered. The calculated results exhibited that the order of O2 activation on the
defect sites was discussed as follows: Zigzag-B > N-vacancy > zigzag-N > armchair > B-
vacancy. According to the kinetic and thermodynamic analysis, zigzag-N edge sites were
indicated to be the most favorable active sites for aerobic ODS. Theoretically, the order of
desulfurization of Th on different defect sites of BN is as follows: zigzag-N > armchair >
zigzag-B > N-vacancy and B-vacancy [40].

In order to construct the BN materials, various synthesis methods have been applied which are
comprising chemical vapor deposition (CVD), ball milling, mechanical stripping, template-
free approach and etc. [41-44]. Among of these methods, non-template synthesis approaches
are more suggested because of having the advantages like facile manufacturing process, cost

3
effective and also requiring fewer steps. Xue et al. prepared BN-based porous monoliths
(BNPMs) via template free method and applied them in purification of an oil/water system.
The results showed great adsorption capacity of BNPMs for removing rhodamine B (RB) dyes
and heavy-metal cations Cd(II) pollutants from an aqueous solution [45].

The prime novelty of the study lies in using BN for desulfurization of the real fuel with high
sulfur content to make a clean fuel and evaluate the industrial application of this catalyst. So in
the present research, BN nanorods are provided via solid state method as a facile, non- template
and efficient approach to investigate the performance of this material in oxidative
desulfurization process. The physiochemical properties of this catalyst have been studied
through various analysis techniques to illustrate the overall capability of these compounds.
Moreover, the impact of the most significant parameters on the ODS process such as
temperature, reaction time, catalyst dosage and the amount of oxidant was evaluated by using

f
an experimental design based on response surface methodology. Also, the recyclability of this

oo
catalyst for deep desulfurization of the fuel under the optimized conditions was examined.

2.Experimental

r
2.1. Materials
-p
re
In present research, H2O2 (35% (mass)), acetonitrile (CH3CN, 99.9%) and boric acid (H3BO3)
were purchased from Merck and melamine (C3N6H6, 99.8%) was prepared from Urmia
lP

Petrochemical Company. All the chemicals were used without further purification.
na

2.2. Synthesis of boron nitride (BN)


ur

To prepare BN catalyst, 4 g of boric acid was initially dissolved in deionized water and heated
at 75 °C until it was completely resolved. Then, certain amount of melamine (with molar ratio
Jo

1:1) was added to solution under vigorous mixing for 30 min. The mixture was dried at 60 °C
for 12 h for obtaining a white precursor. Subsequently the white solid was placed in a furnace
under nitrogen atmosphere and heated up to 900 °C with a heating rate of 5 °C·min−1 for 3h.
After cooling to room temperature, the remaining solid was gathered as BN catalyst. The
schematic of preparation procedure of BN catalyst was depicted in Fig. 1.

4
Boric acid

melamine 900℃, 3h

10 μm
Followed by Calcined
Boron drying at 60 ℃
Nitrogen
Precursor BN nanorods
Carbon
paste
Oxygen
Hydrogen

Fig. 1 Fabrication method of BN catalyst with nanorods structure

f
oo
2.3. Characterization methods

X-ray diffraction (XRD) patterns were taken by employing Philips PW 1730 with Cu Kα

r
irradiation in the wavelength of 0.1540 nm, current intensity of 30 mA and 40 kV voltage to
-p
illuminate the crystal structure of the catalyst. Fourier transform infrared spectroscopy (FT-IR)
re
was carried out using Thermo AVATAR instrument to identify the chemical composition and
functional groups of the BN catalyst. The morphology and structure of the sample were
lP

analyzed using FESEM on TESCAN MIRA. Brunauer-Emmett-Teller (BET) analysis was


performed by BELSORP MINI II at 77 K under N2 flow to investigate the specific surface area,
pore diameter and pore volume of the sample. Degassing process was conducted at a
na

temperature of 120 °C during the overnight (8 h). Also, the nitrogen adsorption data was used
to calculate pore size distribution of the catalyst by Barrett–Joyner–Halenda (BJH) method.
ur

Eventually, Raman spectra were collected on a Almega Thermo Nicolet Dispersive Raman
Spectrometer for further verification of BN structure. The transmission electron microscopy
Jo

(TEM, Philips CM120) analysis was utilized for better assessment the microstructure of the
sample which was done under 200 kV acceleration voltage. The atomic force microscope
(AFM, Bruker, JPK nanowizard II) was implemented in non-contact mode to observe the
topography and thickness of the catalyst. Gas chromatograph system (Agilent
Technologies7890 GC) coupled with Agilent 5975 network mass selective detector (GC-Mass)
was applied to detect the hydrocarbon components in diesel feedstock. The initial temperature
was set to 45 °C and raised to 250 °C at the ramp of 2 °C·min−1. Then the temperature increased
to 280 °C at the rate of 10 °C·min−1 and kept constant at this temperature for 15 min and the
injector temperature was set at 250 °C.

2.4. Oxidative desulfurization process

Desulfurization process was performed in 50 ml flask equipped with the condenser and
magnetic stirrer in a temperature-controlled water batch. In a typical run, diesel fuel (20 ml)
with characteristics summarized in Table 1, and certain amounts of catalyst were added to the
reactor. Then, H2O2 (35% (mass)) as an oxidant was added and the solution was heated to the

5
desired temperature with stirring in a constant speed. After completing the reaction, the catalyst
was separated from the solution and the diesel oil was mixed with acetonitrile (2:1 v/v) as a
solvent to extract the sulfur compounds and the mixture was settled in separatory funnel for 10
min. The fuel phase was separated and the value of the sulfur content was analyzed by Energy
Dispersive X-ray fluorescence analysis (XRF) based on ASTM D4294 standard method by
RIGAKU NEX QC instrument. The conversion was estimated according to the following
equation:

𝐶𝑜 −𝐶𝑡
Y(%)= × 100 (1)
𝐶𝑜

Where Y is conversion, 𝐶𝑜 is initial sulfur concentration in the fuel and 𝐶𝑡 is remaining sulfur
concentration after the ODS process at time t.

f
Table 1 Properties of diesel fuel

oo
Characteristic Method Content
Density at 15 °C /kg·m−3 ASTM D4052 830

r
Kinematic viscosity/cSt ASTM D445 2.55
Sulfur content/ppm
Flash point/°C ASTM D93
-p
ASTM D4294 8040
60
re
Water content/ppm ASTM D6304 43
Cetan index ASTM D976 50
lP

Distillation range/°C
IBP ASTM D86 168
5% 180
na

10% 188
20% 202.5
30% 216.5
ur

40% 232
50% 250
60% 264.5
Jo

70% 281.5
80% 302
90% 331
95% 355
FBP 376

2.5. Experimental design and statistical analysis

To investigate the effect of the most significant parameters on the process and also find out the
optimum value of independent variables, response surface methodology (RSM) as an
economical, accurate and practical way was broadly used in the research. Box-Behnken design
(BBD) as one of the experimental design methods in RSM was employed in the present study
to analyze the important parameters in the ODS process which is including the amount of
oxidant O/S, temperature T, time t and catalyst dosage by using Design Expert software. In this
method, three levels are required and they are coded as −1, 0, 1, which mentioned the low,
middle and high level, respectively. The levels and experimental range of these independent

6
variables are represented in Table 2. Second-order polynomial equation as a function of
independent variables and their interactions is illustrated in the following equation [46]:

Y=𝛽0 + ∑ 𝛽𝑖 𝑥𝑖 + ∑ 𝛽𝑖𝑖 𝑥𝑖2 + ∑ 𝛽𝑖𝑗 𝑥𝑖 𝑥𝑗 (2)

Where Y is the response, 𝛽0 is the constant coefficient, 𝛽𝑖 , 𝛽𝑖𝑖 and 𝛽𝑖𝑗 are linear, quadratic and
second order interaction coefficient, respectively. 𝑥𝑖 and 𝑥𝑗 are independent variables that
specify Y. It should be noted that the conversion was statistically analyzed through analysis of
variance (ANOVA) while the significant of the polynomial model was expressed by regression
analysis. Additionally, the interaction of the parameters was evaluated by making the response
surface and contour plots [47].
Table 2 Range and levels of process parameters in the BBD

f
Variables Code Range and levels

oo
−1 0 1
Oxidant/sulfur (O/S) molar ratio X1 3 7.5 12
Temperature/℃ X2 50 65 80

r
Time/min X3 30 75 120
Catalyst dosage/g X4 -p 0.1 0.3 0.5
re
3. Results and discussion
lP

3.1. Catalyst characterization


na

Powder XRD data is applied to determine the crystal structure of BN that the result is shown
in Fig. 2. Two major peaks located at 26.2° and 42.2° corresponded to (0 0 2) and (1 0 0)
planes of BN and they were in accordance with the standard of pure h-BN based on JCPDS
ur

Card no. 34.0421. The value of interlayer distance for the plane of (0 0 2) was 0.340 nm.
Jo

Scherrer equation was used to calculate the average crystallite size (D) of the material:

D= k𝜆/𝛽 cos𝜃 (3)

Where k is the shape factor (0.89), 𝜆 is the wavelength of X-rays, 𝛽 is the full width half
maximum of the 2𝜃 peak and 𝜃 is the angle of diffraction for (0 0 2) peak. The average
crystallite size estimated for BN is equal to 3.8 nm which demonstrates the nanocrystalline
structure of the prepared catalyst. As indicated in the curve, no impurities were seen that it
expressed the appropriate synthesis of BN. Also, the results are in accordance with other reports
[48,49].

7
Fig. 2 XRD pattern of BN powder synthesized

The FTIR spectrum was measured to reveal the functional groups of BN which was displayed
in Fig. 3(a). The absorption peaks appeared at 796 and 1384 cm−1 are assigned to out-of-plane

f
oo
bending vibration of B—N—B bonds and in-plane B—N stretching vibration bonds,
respectively. The weak peak at 863 cm−1 can be corresponded to stretching vibration of

r
tetrahedral BO4− units and the broad peak in 3417 cm−1 can be attributed to presents of O—H
-p
and N—H bonds [50,51]. These functional groups have a significant impact on the catalytic
performance of BN for removing the sulfur compounds. Raman analysis of BN synthesized
re
was used for further perceiving the crystalline structure of the sample. As demonstrated in Fig.
3(b), the peak placed around the wavelength number of 1366 cm−1 is assigned to symmetric
lP

vibration of h-BN in an E2g mode [50]. These results confirmed that the h-BN was properly
constructed.
na
ur
Jo

8
Fig. 3 (a) FTIR (b) Raman spectroscopy of BN powder

The surface morphology of the prepared catalyst was investigated by FESEM micrographs,
which are illustrated in Fig. 4(a)−(c). The images of the sample revealed that the catalyst
consists of the rod-like and two-dimensional sheet-like structure. It can be noticed that some
rods joined together and make larger structures. The morphological features of the sample were
more evaluated by TEM image and the corresponding results are depicted in Fig. 4(d). No
clumping of the synthesized particles in the catalyst structure is manifested and also the
nanoparticles are evenly dispersed that the observation supports FESEM results. To identify
elemental composition of the BN, the elemental mapping analysis was performed and the
results in Fig. 4(e) and (f) represent the homogeneous distribution of the B and N elements on
the surface.

f
oo
(a) (b) (c)

r
-p
re
lP
na
ur

(d) (e) (f)


Jo

Fig. 4. (a−c) FESEM, (d) TEM micrographs and (e, f) MAP images of BN synthesized

The surface topography and AFM micrographs of BN in three-dimensional and two-


dimensional schemes are highlighted in Fig. 5(a), (b). Based upon AFM measurement, the root
mean square (RMS) roughness (Rq) is 1.043 nm for BN sample. Additionally, the micrographs
affirm the uniform distribution of BN particles with a smooth surface which can be noticed
from AFM images. This causes to promote the catalytic activity of BN in the ODS process.

9
(a) (b)

Fig. 5. (a) 2-D and (b)3-D AFM micrographs of the BN catalyst

f
The adsorption and desorption isotherms of nitrogen for BN are given in Fig. 6(a). The presence

oo
of hysteresis loop (type H4) at P/P0 between 0.3−0.96 with the type IV isotherm showed the
mesoporous structure of the synthesized powder. Also, the isotherm at high relative pressure

r
shows the existence of macroporous structure. The pore size distribution curve displayed in
-p
Fig. 6(b) was consistent with the N2 adsorption−desorption curve and confirmed the catalyst
re
possessed macroporous and mesoporous in its structure. The average diameter of the pores
based on the BJH method was 3.2 nm. The results of the textural properties investigation of the
lP

sample including surface area, pore diameter, and pore volume are summarized in Table 3.
Table 3 Textural properties of the synthesized catalyst
na

Catalyst Surface area/m2·g−1 Pore volume/cm3·g−1 Pore diameter/nm


BN 74.82 0.06 3.2
ur
Jo

Fig. 6 (a) N2 adsorption and desorption isotherms curve (b) the corresponding pore size distribution curve

10
3.2. Investigation of ODS reaction

3.2.1. Statistical analysis

According to Box-Behnken method, 27 experiments were proposed to carry out for optimizing
the ODS reaction. The conversion value as an experimental result and the independent
variables are indicated in Table 4.

The second order polynomial equations for conversion of diesel fuel based on the coded and
actual values were exhibited in Eqs. 4 and 5 which explain the mathematical relation between
sulfur removal with the independent variables. These equations were obtained as follows:

Conversion (%) = 69.67 + 1.75 X1 + 2.16 X2 + 1.93 X3 + 2.13 X4 + 1.15 X1 X2+ 1.15 X1 X3+ 1.8
X1 X4 + 0.725 X2 X3− 2.85 X2 X4 − 0.125 X3 X4 − 2.85 X12 − 3.44 X22 − 1.37 X32 − 3.69 X42 (4)

f
Conversion (%) = −31.195+ 0.367 X1 + 2.207 X2 + 0.037 X3 + 113.729 X4 + 0.017 X1X2+ 0.006

oo
X1X3 + 2.0 X1X4 + 0.001 X2X3 – 0.950 X2X4 – 0.014 X3X4 – 0.141 X12 – 0.015 X22 – 0.001 X32 –
92.188 X42 (5)

r
-p
Table 4 Box-Behnken design matrix and the obtained experimental results
re
Actual values Sulfur removal/%
Run. X1 X2 X3 X4 Experimental Predicted
order
lP

1 12 65 75 0.5 69.2 68.80


2 7.5 65 75 0.3 70 69.67
3 7.5 50 120 0.3 63 63.90
na

4 7.5 50 75 0.5 66.3 65.36


5 3 65 30 0.3 61.5 62.91
6 7.5 50 75 0.1 55.7 55.41
ur

7 3 50 75 0.3 60.6 60.62


8 7.5 65 120 0.5 67.2 68.54
Jo

9 7.5 65 30 0.5 65.4 64.92


10 7.5 80 30 0.3 65.4 64.35
11 12 50 75 0.3 61 61.82
12 3 65 75 0.1 60.8 61.05
13 12 65 30 0.3 62.6 64.11
14 7.5 65 120 0.1 63.6 64.54
15 7.5 65 75 0.3 70 69.67
16 7.5 80 120 0.3 69.3 69.67
17 7.5 65 30 0.1 61.3 60.42
18 7.5 80 75 0.1 64.8 65.42
19 3 65 120 0.3 66.3 64.47
20 12 65 120 0.3 72 70.27
21 12 80 75 0.3 68 68.44
22 12 65 75 0.1 61.6 60.95
23 7.5 65 75 0.3 69 69.67
24 7.5 80 75 0.5 64 63.97
25 7.5 50 30 0.3 62 61.49
26 3 65 75 0.5 61.2 61.70
27 3 80 75 0.3 63 62.64

11
The comparison between the experimental data and predicted values is shown in Fig. 7(a) to
confirm the proper fitness of the model response and experimental data. The results showed
that the data points are distributed relatively near the predicted values and also demonstrated
great accuracy of the proposed model. As exhibited in Fig. 7(b), a normal distribution of the
points around the straight line explained the satisfaction of the normality assumption. To find
a possible non-constant variance across the predicted range values, the residual versus fitted
response plot was depicted. In Fig. 7(c), the response values scattered randomly in the desired
range and placed in a range of ±3 which are verified correctness of the fit as well as indicated
the constant variance assumption was not violated. The analysis of the variances (ANOVA)
was employed to show the validity and statistical significance of the model. For this purpose,
the Fisher test (F-test) was performed via dividing the mean square of the factor by its residual
mean square. The results of the ANOVA test are summarized in the Table 5.

f
oo
(a) (b)

r
-p
re
lP
na
ur
Jo

(c)

Fig. 7 (a) Predicted versus actual values plot, (b) normal probability versus internally studentized residuals, (c)
residual versus predicted plot

12
If the F-value becomes larger and the probability value (P-value) is to be smaller, the model
and corresponding variables would be more significant. It should be noted that the parameters
with the P-value lower than 0.05 are regarded to be significance [52,53]. In Table 5, the model
F-value is 15.14 which indicates there is only 0.01% chance that such a large F-value could
arise because of the noise. Moreover, the model P-value with an amount of less than 0.0001
showed that the model was significant with further 95% confidence level. It is noteworthy to
state that the temperature and catalyst dosage have a high F-value and the lowest p-value
compared to the other parameters, therefore these factors had more impact on ODS reaction.
According to this table, the model terms with P-value less than 0.05 are statistically significant.
Also, the lack-of–fit was insignificant (P-value greater than 0.05), which is appropriate for
verifying the significance of the model.

The high value of determination coefficient (R2=0.95) displayed the good fitting quality of the
quadratic model. The adjusted R2 was determined as 0.88 which is in suitable agreement with

f
oo
predicted R2 because their difference is less than 0.2. Moreover, these values showed the
consistency of the coefficients. As shown in Table 5, the model can be considered reproducible

r
because the coefficient of variation (CV) is not greater than 10%. Additionally, the calculated
-p
adequate precision value showed an appropriate signal to noise ratio and also confirmed the
model is suitable for explaining the desulfurization process.
re
Table 5 Analysis of variance (ANOVA) results for sulfur removal efficiency
lP

Source SS DF MS F-value P-value


Model 361.54 14 25.82 15.14 < 0.0001 Significant
X1 36.75 1 36.75 21.55 0.0006
na

X2 55.90 1 55.90 32.78 < 0.0001


X3 44.85 1 44.85 26.30 0.0002
ur

X4 54.19 1 54.19 31.77 0.0001


X1X2 5.29 1 5.29 3.10 0.1036
X1X3 5.29 1 5.29 3.10 0.1036
Jo

X1X4 12.96 1 12.96 7.60 0.0174


X2X3 2.10 1 2.10 1.23 0.2886
X2X4 32.49 1 32.49 19.05 0.0009
X3X4 0.0625 1 0.0625 0.0366 0.8514
X12 43.32 1 43.32 25.40 0.0003
X22 63.02 1 63.02 36.95 < 0.0001
X32 10.08 1 10.08 5.91 0.0316
X42 72.52 1 72.52 42.52 < 0.0001
Residual 20.47 12 1.71
Not
Lack of Fit 19.80 10 1.98 5.94 0.1526
significant
Pure Error 0.6667 2 0.3333
Total 382.01 26
R2= 0.9464, R2adj= 0.8839, CV%=2.02, adequate precision = 15.273

13
3.2.2. Effect of the main parameters on ODS process

To investigate the effect of main parameters on the yield of sulfur removal based on the
experimental design, Fig. 8(a)−(d) was illustrated.

The effect of O/S molar ratio on desulfurization process is shown in Fig. 8(a) in which the
temperature, time and catalyst dosage are preserved at the center point values of 65 ℃, 75 min
and 0.3 g, respectively. By considering the figure, the enhancement of the O/S molar ratio up
to 9 has a positive effect on the reaction as a result of the formation of the reactive hydroxyl
functionalized boron nitride species [54]. With further increasing the O/S molar ratio, the
efficiency of the desulfurization decreased. This may be due to the decomposition of H2O2 to
ineffective components consisting of H2O and O2 that it has a negative effect on desulfurization
reaction [55].

f
Fig. 8(b) demonstrates the effect of temperature on removing the sulfur compounds in the

oo
center point values of O/S, time and catalyst dosage. It can be seen that with increasing the
temperature up to 70 ℃ , the efficiency of the process enhanced which is attributed to

r
improving the mass transfer in the porous structure and also the speed of movement and
-p
collision of molecules were raised. More increments in temperature lead to reduction in
desulfurization rate because of thermal decomposition of oxidant at the higher temperatures,
re
according to the following reaction [55]:
lP

2H2O2 → 2H2O + O2 (6)

In Fig. 8(c), the sulfur compounds elimination as a function of reaction time was depicted while
na

the other parameters were in a constant value. By increasing the reaction time, the yield of the
reaction has been improved. It can be related to the enhancement of opportunity for a contact
ur

time of organosulfur compounds, oxidant and the catalyst. Further increasing in reaction time
has a negligible impact on the conversion due to occupying the active sites with products and
Jo

also the number of desorption from the surface reduced [56]. Also, performing the reaction in
a long time causes the loss of oxidant and changes the reaction media from oxidizing to
reducing one and it has a negative effect on the efficiency of the process [57].

The effect of catalyst dosage on the ODS reaction was shown in Fig. 8(d). As indicated, the
conversion increased by enhancing the amount of catalyst up to 0.36 g in the center point of
the other parameter. It can be due to more interaction of oxidant, sulfur compounds and catalyst
which causes easier oxidation of organosulfur compounds in diesel fuel. Further increasing in
catalyst dosage might supply more active phases, but it can lead to agglomeration, pore
blocking and decline the available surface area of catalyst for reaction and as a result, it reduces
the efficiency [56,58].

14
(a) (b)

f
oo
(c) (d)

r
-p
re
lP
na
ur
Jo

Fig. 8 Effect of the main parameters on the sulfur removal conversion in ODS reaction

3.2.3. Interaction effect of the main parameters

Response contour and 3-D response surface plots are displayed in Fig. 9(a)−(d) to evaluate the
effect of interaction terms on the sulfur removal. Fig. 9(a) shows the interaction of oxidant
amount and temperature on the response in which the other parameters were settled in the
middle level. The analysis of the plot shows that as the temperature increases, the reaction rate
first enhances and then reduces as a result of the decomposition of oxidant at high temperatures
and it has a negative effect on desulfurization. As it can be observed, the efficiency of the
process decreased in the high amount of oxidant to sulfur molar ratio since the excess oxidant
value can lead to the activation of side reactions and reduce the sulfur removal [59].

15
The interaction effect of O/S molar ratio and catalyst dosage on the sulfur removal was
exhibited in Fig. 9(b) while the temperature and time were in the middle values. From the
result, it was found that the increment of the catalyst dosage increases the conversion
considering the existence of more active sites on the surface of the catalyst for the oxidation
process. Nonetheless, excessive amounts of the catalyst may lead to reducing the contact
surface of the reactants, agglomeration and decreasing the catalytic activity for ODS reaction
[60].

The binary interaction of temperature and catalyst dosage on the reaction in the central value
of two other parameters, was indicated in Fig. 9(c). According to the plot, enhancement of the
temperature has a positive effect on the reduction of sulfur compounds. But, as illustrated in
this graph, at the higher temperature and catalyst dosage the conversion of sulfur removal was
slightly decreased because of the side reaction at high temperatures which can neutralize active

f
site of the catalyst.

oo
Fig. 9(d) exhibits the combined interaction of the time and catalyst dosage. As in this figure
represented, increasing the time of process improves complete of the oxidation reaction.

r
-p
However, further increasing the reaction time up to a certain amount has a negligible effect and
the sulfur removal from the fuel was a little change. Also, it was illuminated that the higher
re
amount of catalyst dosage causes reducing the yield by reason of accumulation of catalytic
particles and decreasing in the available surface area for the reactants [61].
lP

(a)
na
ur
Jo

(b)

16
(c)

f
oo
(d)

r
-p
re
lP
na
ur

Fig. 9 Response surface and contour plots of desulfurization


Jo

3.2.4 Optimization of desulfurization from diesel fuel

To achieve the optimum conditions of the main parameters for ODS process, the response
optimizer was applied. Based on the implemented experiments, different combinations of
parameters were suggested by the proposed model to obtain maximum conversion. From the
results, an optimized value with 72.03% sulfur removal was selected which can be achieved
under the following conditions: O/S molar ratio of 10.2, temperature of 71 ℃, reaction time of
113 min and catalyst dosage of 0.36 g. To verify the accuracy of optimization and the model,
the experiment was performed under optimum conditions. The result indicated that the
experimental value with 72.4% is in close agreement with predicted sulfur removal. This is
confirmed the validity of the proposed model. Complete desulfurization of diesel fuel is very
difficult and relatively impossible due to the presence of complex compounds in the real fuel.
However, approximately high conversion of sulfur compounds was achieved in the present
study in comparison with other researches performed on desulfurization from real fuel [62−64].

17
For instance, Chen et al. studied the ODS of real FCC diesel fuel in the presence of H2O2 as an
oxidant and [Hnmp]Clx/(ZnCl2)y as a catalyst and extractant. They observed that maximum
sulfur removal reached to 38.2% at operating conditions of temperature 75 ˚C, O/S molar ratio
50, 20 min reaction time and mass ratio of catalyst/oil 1/3. It was found that the low efficiency
value for FFC diesel compared with model fuel can be due to existing the more complex sulfur
species in a real fuel [62].

In a study, Vedachalam et al. used a dual function NiMo/γ–Al2O3 catalyst for desulfurization
from the heavy fuel oil (HFO) feedstock by hydrotreating under moderate process conditions
and then the ODS process was performed. Hydrotreating of fuel was conducted under industrial
conditions, which can reduce sulfur content from 3.4% to 1.14% (mass). Subsequently, ODS
was utilized to obtain the fuel with sulfur level of 0.5% (mass). It was indicated that acetonitrile
was more effective than methanol in the solvent extraction stage. The maximum amount of

f
removing sulfur compounds by ODS reaction was 66.5% achieved at reaction conditions of

oo
O/S molar ratio 5.8, temperature 68 ˚C, catalyst to feed ratio 0.05 and 96 min reaction time
[64]. So, these results show that the BN catalyst has a great activity in the ODS process.

r
3.3. Mechanism of oxidative desulfurization -p
Various hydrocarbons and sulfur compounds exist in diesel fuel in which the major percentage
re
of sulfur compounds are including DBT and its derivatives. The possible reaction mechanism
lP

for ODS process using BN as a catalyst is shown in Fig. 10. According to the pervious study,
during the decomposition reaction of H2O2 on the BN surface, hydroxyl radicals were
generated as the active species that they diffused towards the catalyst surface. Several states
na

may occur for interaction and absorbing of hydroxyl radicals with catalyst depending on the
sites and kind of surface atoms. It was indicated that absorption of .OH on B surface atoms
ur

created the most stable site due to strong interaction between them [65]. These results were in
agreement with the calculated bond strength for the hydroxyl group attached to B atoms [66].
Jo

Subsequently, these active species encounter sulfur components on the catalyst surface and
initiate the oxidation reaction to produce sulfoxide or sulfone components. So, the oxidation
reaction can be conducted based on the following reactions:

BN + H2O2 → B—OH + .OH (7)

Sulfur component + B—OH → Oxidized reaction products (8)

18
R1, R2, R3=H,Me,Et R1
S

OH OH OH
B B B R2
N N N N
.
OH B
N
B
N
B
N
B
N
R3

B B B B O R1
N N N
H2O2 S
.
[ OH]
R2
R3
B B B
N N N N

B B B B . O
N N N N [ OH] R1
O O
B B B B R1 S
N N N O
S
R2
R2 R3

f
BN Particle R3

oo
Organic phase Polar phase

r
-p
Fig. 10 The proposed mechanism for desulfurization of diesel fuel using BN catalyst
re
3.4. Effect of ODS process on diesel characteristics
lP

To investigate the impact of desulfurization on the fuel characteristics, the distribution of


different hydrocarbons was measured before and after the ODS process by GC-Mass analysis.
na

The results indicated that various hydrocarbons and components exist in the diesel structure,
which are classified into four major types comprising parrafins, naphtenes, olefins and
ur

aromatics. Based on the Fig.11(a), (b), the sharp peaks are assigned to paraffin hydrocarbons
(from C7 to C25) where the intensity of these peaks increases after the ODS reaction. Fig. 11
Jo

and the results in Table 6 confirm enhancement the amounts of parrfins in diesel structure after
the treatment. Table 6 illustrates the aromatic compounds including toluene, benzene,
naphthalene and their derivatives were remarkably decline after the reaction which can be
ascribed to dearomatization during the process and this is in accordance with previous studies
[67,68]. Actually, the aromatic oxidation performs competitively with sulfur compounds
oxidation and the oxidized products are removed from diesel in the extraction process [68]. It
should be noted that high aromatic content in fuels reduces the fuel quality and it has negative
effects on the environment [69]. Thus, this process can improve the combustion characteristics
and properties of diesel. Increasing in value of other types of hydrocarbons can be attributed to
the dearomatization of fuel.
Table 6 The comparison of composition of diesel fuel before and after ODS process

Hydrocarbon type Before ODS After ODS


Composition
Paraffins 60.42 79.24
Naphtenes 7.55 9.3
Aromatics 29.45 7.1
Olefins 2.58 4.4

19
A b und a nc e

C11
1550000
C10 T IC : G C M S D O N L IN E _ 2 0 0 9 .D \ d a ta .m s

C12
(a) 1500000

1450000

1400000
C13 C14
1350000

1300000
C15
1250000

1200000

1150000
C9 C16
1100000
A
1050000

1000000
C17
950000

900000

850000 C18
800000

750000

700000

650000
C19
600000
C20

f
550000
C8

oo
500000

450000 A
400000 C21
350000

300000
C22

r
250000
C23
200000

150000

100000

50000
C7
A
-p C24
C25
re
1 0 .0 0 2 0 .0 0 3 0 .0 0 4 0 .0 0 5 0 .0 0 6 0 .0 0 7 0 .0 0 8 0 .0 0 9 0 .0 0 1 0 0 .0 0 1 1 0 .0 0
T im e -->
lP

Abundance

C12 C13 TIC: GCMSD ONLINE_2007.D\data.ms


na

(b) 2200000 C11


C14
2100000 C10 C15
2000000
C16
ur

1900000

1800000

1700000
C17
Jo

1600000

1500000

1400000 C18
1300000
C9
1200000
A C19
1100000

1000000
C20
900000

800000

700000
C21
600000
C22
500000

400000
C8 C23
300000
A C24
200000 C7
C25
100000

10.00 20.00 30.00 40.00 50.00 60.00 70.00 80.00 90.00 100.00 110.00
Time-->

Fig. 11 GC-Mass spectrum of diesel fuel before (a) and after (b) the ODS treatment

20
3.5. Reusability of catalyst

Recycling of the catalyst is an important factor to investigate the industrial application of the
catalyst. The reusability test of BN catalyst was performed under the optimum conditions and
the results are revealed in Fig. 12. After each test, the catalyst in the reaction mixture was
separated and washed with acetonitrile to remove the sulfur compounds which were absorbed
on the surface of the catalyst. Then the catalyst was dried at 80 ℃ in an oven and used for a
new catalytic run. From the result, it can be noticed that the catalyst has a great stability with
less reduction in its activity for removing the sulfur compounds. Decreasing efficiency may
relate to the active sites which have been blocked on the surface of the catalyst.

f
r oo
-p
re
lP
na

Fig. 12 The reusability test of the BN catalyst in ODS reaction

4. Conclusion
ur

Oxidative desulfurization of diesel fuel was carried out by employing the BN as a metal-free
Jo

catalyst which has been prepared through a nontemplate method by using boric acid and
melamine as boron and nitrogen sources. In this regard, various characterization techniques
were used to confirm that BN catalyst has been successfully constructed. According to the
results the prepared catalyst has BN nanorods characteristics and properties. By using the BN
in removing the sulfur compounds from the fuel, high catalytic performance was obtained. The
role of main variables on catalytic desulfurization was studied by the BBD method and
exhibited that they have a significant influence on sulfur removal. As illuminated, the catalyst
dosage and temperature are the most impressive parameters among the other factors. As it was
shown the predicted values by the model were in good agreement with experimental data.
Optimum operating conditions to reach the maximum sulfur removal were as follow: O/S molar
ratio of 10.2, temperature of 71 ℃, reaction time of 113 min and catalyst dosage 0.36 g.
Additionally recyclability of catalyst was investigated and it was found that BN can be reused
10 times with fewer reduction of its activity. The results reveal that BN can be utilized and
selected as a metal-free catalyst with great performance for industrial application in the
oxidative desulfurization process. This research opens up a way for preparing the economical
and ecofriendly catalyst with the outstanding activity for ODS reaction.

21
References
[1] M.D.G. de Luna, M.L. Samaniego, D.C. Ong, M.W. Wan, M.C. Lu, Kinetics of sulfur
removal in high shear mixing-assisted oxidative-adsorptive desulfurization of diesel, J. Clean.
Prod. 178 (2018) 468–475.
[2] M. Taghizadeh, E. Mehrvarz, A. Taghipour, Polyoxometalate as an effective catalyst for
the oxidative desulfurization of liquid fuels: a critical review, Rev. Chem. Eng. 36 (7) (2020)
831–858.
[3] M. Ja’fari, S.L. Ebrahimi, M.R. Khosravi-Nikou, Ultrasound-assisted oxidative
desulfurization and denitrogenation of liquid hydrocarbon fuels: a critical review, Ultrason.
Sonochem. 40 (Pt A) (2018) 955–968.
[4] M. Hossain, H. Park, H. Choi, A comprehensive review on catalytic oxidative
desulfurization of liquid fuel oil, Catalysts 9 (3) (2019) 229.

f
oo
[5] L.W. Hao, T. Su, D.M. Hao, C.L. Deng, W.Z. Ren, H.Y. Lü, Oxidative desulfurization of
diesel fuel with caprolactam-based acidic deep eutectic solvents: tailoring the reactivity of

r
DESs by adjusting the composition, Chin. J. Catal. 39 (9) (2018) 1552–1559.
-p
[6] T. Ganguly, A. Das, M. Jana, A. Majumdar, Cobalt(II)-mediated desulfurization of
re
thiophenes, sulfides, and thiols, Inorg. Chem. 57 (18) (2018) 11306–11309.
[7] X.S. Wang, L. Li, J. Liang, Y.B. Huang, R. Cao, Boosting oxidative desulfurization of
lP

model and real gasoline over phosphotungstic acid encapsulated in metal–organic frameworks:
the window size matters, ChemCatChem 9 (6) (2017) 971–979.
na

[8] C. Sentorun-Shalaby, S.K. Saha, X.L. Ma, C.S. Song, Mesoporous-molecular-sieve-


supported nickel sorbents for adsorptive desulfurization of commercial ultra-low-sulfur diesel
ur

fuel, Appl. Catal. B Environ. 101 (3–4) (2011) 718–726.


[9] F.T. Li, B. Wu, R.H. Liu, X.J. Wang, L.J. Chen, D.S. Zhao, An inexpensive N-methyl-2-
Jo

pyrrolidone-based ionic liquid as efficient extractant and catalyst for desulfurization of


dibenzothiophene, Chem. Eng. J. 274 (2015) 192–199.
[10] S.T. Du, X.X. Chen, Q.M. Sun, N. Wang, M.J. Jia, V. Valtchev, J.H. Yu, A non-
chemically selective top-down approach towards the preparation of hierarchical TS-1 zeolites
with improved oxidative desulfurization catalytic performance, Chem. Commun. 52 (17)
(2016) 3580–3583.
[11] T.A.G. Duarte, S.M.G. Pires, I.C.M.S. Santos, M.M.Q. Simões, M.G.P.M.S. Neves,
A.M.V. Cavaleiro, J.A.S. Cavaleiro, A Mn(iii) polyoxotungstate in the oxidation of
organosulfur compounds by H2O2 at room temperature: an environmentally safe catalytic
approach, Catal. Sci. Technol. 6 (9) (2016) 3271–3278.
[12] S.H. Xun, C.Z. Hou, H.P. Li, M.Q. He, R.L. Ma, M. Zhang, W.S. Zhu, H.M. Li, Synthesis
of WO3/mesoporous ZrO2 catalyst as a high-efficiency catalyst for catalytic oxidation of
dibenzothiophene in diesel, J. Mater. Sci. 53 (23) (2018) 15927–15938.
[13] K. Chen, N. Liu, M.H. Zhang, D.H. Wang, Oxidative desulfurization of dibenzothiophene
over monoclinic VO2 phase-transition catalysts, Appl. Catal. B Environ. 212 (2017) 32–40.

22
[14] C.N. Dai, J. Zhang, C.P. Huang, Z.G. Lei, Ionic liquids in selective oxidation: catalysts
and solvents, Chem. Rev. 117 (10) (2017) 6929–6983.
[15] R. Limvorapitux, H.Y. Chen, M.L. Mendonca, M.T. Liu, R.Q. Snurr, S.T. Nguyen,
Elucidating the mechanism of the UiO-66-catalyzed sulfide oxidation: activity and selectivity
enhancements through changes in the node coordination environment and solvent, Catal. Sci.
Technol. 9 (2) (2019) 327–335.
[16] X.M. Yu, P.F. Han, Y. Li, Oxidative desulfurization of dibenzothiophene catalyzed by α-
MnO2 nanosheets on palygorskite using hydrogen peroxide as oxidant, RSC Adv. 8 (32) (2018)
17938–17943.
[17] T. Guo, W. Jiang, Y.J. Ruan, L. Dong, H. Liu, H.P. Li, W.S. Zhu, H.M. Li,
Superparamagnetic Mo-containing core-shell microspheres for catalytic oxidative
desulfurization of fuel, Colloids Surf. A Physicochem. Eng. Aspects 537 (2018) 243–249.

f
oo
[18] S. Houda, C. Lancelot, P. Blanchard, L. Poinel, C. Lamonier, Oxidative desulfurization of
heavy oils with high sulfur content: a review, Catalysts 8 (9) (2018) 344.

r
[19] L. Kang, H.Y. Liu, H.J. He, C.P. Yang, Oxidative desulfurization of dibenzothiophene
-p
using molybdenum catalyst supported on Ti-pillared montmorillonite and separation of
sulfones by filtration, Fuel 234 (2018) 1229–1237.
re
[20] L. Chen, Z.P. Hu, J.T. Ren, Z. Wang, Z.Y. Yuan, Efficient oxidative desulfurization over
lP

highly dispersed molybdenum oxides supported on mesoporous titanium phosphonates,


Microporous Mesoporous Mater. 315 (2021) 110921.
na

[21] X.Y. Liu, X.P. Li, R.X. Zhao, H. Zhang, A facile sol–gel method based on urea–SnCl2
deep eutectic solvents for the synthesis of SnO2/SiO2 with high oxidation desulfurization
activity, New J. Chem. 45 (35) (2021) 15901–15911.
ur

[22] P.W. Wu, Q.D. Jia, J. He, L.J. Lu, L.L. Chen, J. Zhu, C. Peng, M.Q. He, J. Xiong, W.S.
Jo

Zhu, H.M. Li, Mechanical exfoliation of boron carbide: a metal-free catalyst for aerobic
oxidative desulfurization in fuel, J. Hazard. Mater. 391 (2020) 122183.
[23] P.W. Wu, L.J. Lu, J. He, L.L. Chen, Y.H. Chao, M.Q. He, F.X. Zhu, X.Z. Chu, H.M. Li,
W.S. Zhu, Hexagonal boron nitride: a metal-free catalyst for deep oxidative desulfurization of
fuel oils, Green Energy Environ. 5 (2) (2020) 166–172.
[24] I. Ahmed, P. Puthiaraj, Y.M. Chung, W.S. Ahn, Metal-free aerobic oxidative
desulfurization over a diethyltriamine-functionalized aromatic porous polymer, Fuel Process.
Technol. 215 (2021) 106741.
[25] Q.Q. Gu, G.D. Wen, Y.X. Ding, K.H. Wu, C.M. Chen, D.S. Su, Reduced graphene oxide:
a metal-free catalyst for aerobic oxidative desulfurization, Green Chem. 19 (4) (2017) 1175–
1181.
[26] D.V. Shtansky, K.L. Firestein, D.V. Golberg, Fabrication and application of BN
nanoparticles, nanosheets and their nanohybrids, Nanoscale 10 (37) (2018) 17477–17493.

23
[27] A.I. Epishin, T. Link, B. Fedelich, I.L. Svetlov, E.R. Golubovskiy, Hot isostatic pressing
of single-crystal nickel-base superalloys: mechanism of pore closure and effect on Mechanical
properties, MATEC Web Conf. 14 (2014) 08003.
[28] V. Sharma, H.L. Kagdada, P.K. Jha, P. Śpiewak, K.J. Kurzydłowski, Thermal transport
properties of boron nitride based materials: a review, Renew. Sustain. Energy Rev. 120 (2020)
109622.
[29] P.W. Wu, W.S. Zhu, A.M. Wei, B.L. Dai, Y.H. Chao, C.F. Li, H.M. Li, S. Dai,
Controllable fabrication of tungsten oxide nanoparticles confined in graphene-analogous boron
nitride as an efficient desulfurization catalyst, Chemistry 21 (43) (2015) 15421–15427.
[30] G. Xie, K. Zhang, B. Guo, Q. Liu, L. Fang, J.R. Gong, Graphene-based materials for
hydrogen generation from light-driven water splitting, Adv. Mater. 25 (28) (2013) 3820–3839.
[31] W.W. Lei, D. Liu, Y. Chen, Highly crumpled boron nitride nanosheets as adsorbents:

f
oo
scalable solvent-less production, Adv. Mater. Interfaces 2 (3) (2015) 1400529.
[32] J.Y. Zhu, H. Wang, J.W. Liu, L.Z. Ouyang, M. Zhu, Exfoliation of MoS 2 and h-BN

r
nanosheets by hydrolysis of LiBH4, Nanotechnology 28 (11) (2017) 115604.
-p
[33] D. Liu, M.W. Zhang, W.J. Xie, L. Sun, Y. Chen, W.W. Lei, Efficient photocatalytic
re
reduction of aqueous Cr(vi) over porous BNNSs/TiO2 nanocomposites under visible light
irradiation, Catal. Sci. Technol. 6 (23) (2016) 8309–8313.
lP

[34] A. Seif, A. Rashidi, S. Scheiner, K. Azizi, T. Kar, Theoretical insight into a feasible
strategy of capturing, storing and releasing toxic HCN at the surface of doped BN-sheets by
na

charge modulation, Appl. Surf. Sci. 496 (2019) 143714.


[35] W.D. Oh, M.G.H. Lee, W.D. Chanaka Udayanga, A. Veksha, Y.P. Bao, A. Giannis, J.W.
ur

Lim, G. Lisak, Insights into the single and binary adsorption of copper(II) and nickel(II) on
hexagonal boron nitride: performance and mechanistic studies, J. Environ. Chem. Eng. 7 (1)
Jo

(2019) 102872.
[36] X.Y. Zhang, R. You, Z.Y. Wei, X. Jiang, J.Z. Yang, Y. Pan, P.W. Wu, Q.D. Jia, Z.H. Bao,
L. Bai, M.Z. Jin, B. Sumpter, V. Fung, W.X. Huang, Z.L. Wu, Radical chemistry and reaction
mechanisms of propane oxidative dehydrogenation over hexagonal boron nitride catalysts,
Angew. Chem. Int. Ed Engl. 59 (21) (2020) 8042–8046.
[37] A. Rajendran, H.X. Fan, J. Feng, W.Y. Li, Desulfurization on boron nitride and boron
nitride-based materials, Chem. Asian J. 15 (14) (2020) 2038–2059.
[38] P.W. Wu, W.S. Zhu, Y.H. Chao, J.S. Zhang, P.F. Zhang, H.Y. Zhu, C.F. Li, Z.G. Chen,
H.M. Li, S. Dai, A template-free solvent-mediated synthesis of high surface area boron nitride
nanosheets for aerobic oxidative desulfurization, Chem. Commun. 52 (1) (2016) 144–147.
[39] Y.C. Wu, P.W. Wu, Y.H. Chao, J. He, H.P. Li, L.J. Lu, W. Jiang, B.B. Zhang, H.M. Li,
W.S. Zhu, Gas-exfoliated porous monolayer boron nitride for enhanced aerobic oxidative
desulfurization performance, Nanotechnology 29 (2) (2018) 025604.

24
[40] N. Lv, L. Sun, L. Chen, Y. Li, J. Zhang, P. Wu, H. Li, W. Zhu, H. Li, The mechanism of
thiophene oxidation on metal-free two-dimensional hexagonal boron nitride, Phys. Chem.
Chem. Phys. 21(39), (2019) 21867–21874.

[41] R. Han, M.H. Khan, A. Angeloski, G. Casillas, C.W. Yoon, X.D. Sun, Z.G. Huang,
Hexagonal boron nitride nanosheets grown via chemical vapor deposition for silver protection,
ACS Appl. Nano Mater. 2 (5) (2019) 2830–2835.

[42] W.W. Lei, V.N. Mochalin, D. Liu, S. Qin, Y. Gogotsi, Y. Chen, Boron nitride colloidal
solutions, ultralight aerogels and freestanding membranes through one-step exfoliation and
functionalization, Nat. Commun. 6 (2015) 8849.

[43] C.P. Yu, Q.C. Zhang, J. Zhang, R.J. Geng, W. Tian, X.D. Fan, Y.G. Yao, One-step in situ
ball milling synthesis of polymer-functionalized few-layered boron nitride and its application

f
in high thermally conductive cellulose composites, ACS Appl. Nano Mater. 1 (9) (2018) 4875–

oo
4883.

r
[44] S. Marchesini, A. Regoutz, D. Payne, C. Petit, Tunable porous boron nitride: investigating
-p
its formation and its application for gas adsorption, Microporous Mesoporous Mater. 243
(2017) 154–163.
re
[45] Y.M. Xue, P.C. Dai, X.F. Jiang, X.B. Wang, C. Zhang, D.M. Tang, Q.H. Weng, X. Wang,
lP

A. Pakdel, C.C. Tang, Y. Bando, D. Golberg, Template-free synthesis of boron nitride foam-
like porous monoliths and their high-end applications in water purification, J. Mater. Chem. A
na

4 (4) (2016) 1469–1478.

[46] L. Qiu, Y. Cheng, C.P. Yang, G.M. Zeng, Z.Y. Long, S.N. Wei, K. Zhao, L. Luo,
ur

Oxidative desulfurization of dibenzothiophene using a catalyst of molybdenum supported on


modified medicinal stone, RSC Adv. 6 (21) (2016) 17036–17045.
Jo

[47] M.F. Majid, H.F. Mohd Zaid, C. Fai Kait, K. Jumbri, J.W. Lim, A.N. Masri, S.M. Mat
Ghani, H. Yamagishi, Y. Yamamoto, B. Yuliarto, Liquid polymer eutectic mixture for
integrated extractive-oxidative desulfurization of fuel oil: an optimization study via response
surface methodology, Processes 8 (7) (2020) 848.

[48] F. Hojatisaeidi, M. Mureddu, F. Dessì, G. Durand, B. Saha, Metal-free modified boron


nitride for enhanced CO2 capture, Energies 13 (3) (2020) 549.

[49] W.W. Lei, D. Portehault, D. Liu, S. Qin, Y. Chen, Porous boron nitride nanosheets for
effective water cleaning, Nat. Commun. 4 (2013) 1777.

[50] C. Yang, J.F. Wang, Y. Chen, D. Liu, S.M. Huang, W.W. Lei, One-step template-free
synthesis of 3D functionalized flower-like boron nitride nanosheets for NH3 and CO2
adsorption, Nanoscale 10 (23) (2018) 10979–10985.

[51] M. Öz, Ç. Bozkurt, B.K. Yılmaz, G. Yıldırım, Effect of borates on the synthesis of
nanoscale hexagonal boron nitride by a solid-state method, Micr osc. Res. Tech. 84 (11) (2021)
2677–2684.
25
[52] X.X. Han, L.X. Zhou, Optimization of process variables in the synthesis of butyl butyrate
using acid ionic liquid as catalyst, Chem. Eng. J. 172 (1) (2011) 459–466.

[53] F. Bibak, G. Moradi, Oxidative desulfurization of model oil and oil cuts with MoO3/SBA-
15: experimental design and optimization by Box–Behnken method, React. Kinet. Mech. Catal.
131 (2020) 935–951.

[54] H.P. Li, Y.J. Li, L.H. Sun, S.H. Xun, W. Jiang, M. Zhang, W.S. Zhu, H.M. Li, H 2O2
decomposition mechanism and its oxidative desulfurization activity on hexagonal boron nitride
monolayer: a density functional theory study, J. Mol. Graph. Model. 84 (2018) 166–173.

[55] F. Wang, K. Xiao, L. Shi, L.C. Bing, D.Z. Han, G.J. Wang, Catalytic oxidative
desulfurization of model fuel utilizing functionalized HMS catalysts: characterization, catalytic
activity and mechanistic studies, React. Chem. Eng. 6 (2) (2021) 289–296.

f
oo
[56] M.A. Shadmehri, M.R. Housaindokht, A. Nakhaei Pour, Oxidative desulfurization of
dibenzothiophene via layered graphitic carbon nitride-coordinated transition metal as a

r
catalyst, New J. Chem. 45 (36) (2021) 16773–16783.
-p
[57] A. Haghighat Mamaghani, S. Fatemi, M. Asgari, Investigation of influential parameters
re
in deep oxidative desulfurization of dibenzothiophene with hydrogen peroxide and formic acid,
Int. J. Chem. Eng. 2013 (2013) 951045.
lP

[58] S. Subhan, A. Ur Rahman, M. Yaseen, H. Ur Rashid, M. Ishaq, M. Sahibzada, Z.F. Tong,


Ultra-fast and highly efficient catalytic oxidative desulfurization of dibenzothiophene at
na

ambient temperature over low Mn loaded Co-Mo/Al2O3 and Ni-Mo/Al2O3 catalysts using
NaClO as oxidant, Fuel 237 (2019) 793–805.
ur

[59] M.R. Jalali, M.A. Sobati, Intensification of oxidative desulfurization of gas oil by
ultrasound irradiation: Optimization using Box–Behnken design (BBD), Appl. Therm. Eng.
Jo

111 (2017) 1158–1170.

[60] Y. Cao, H.X. Wang, R.M. Ding, L.C. Wang, Z. Liu, B.L. Lv, Highly efficient oxidative
desulfurization of dibenzothiophene using Ni modified MoO3 catalyst, Appl. Catal. A Gen. 589
(2020) 117308.

[61] S. Hasannia, M. Kazemeini, A. Seif, A. Rashidi, Oxidative desulfurization of a model


liquid fuel over an rGO-supported transition metal modified WO3 catalyst: experimental and
theoretical studies, Sep. Purif. Technol. 269 (2021) 118729.

[62] X.C. Chen, H.S. Guo, A.A. Abdeltawab, Y.W. Guan, S.S. Al-Deyab, G.R. Yu, L. Yu,
Brønsted–lewis acidic ionic liquids and application in oxidative desulfurization of diesel fuel,
Energy Fuels 29 (5) (2015) 2998–3003.

[63] I. Shafiq, M. Hussain, S. Shafique, R. Rashid, P. Akhter, A. Ahmed, J.K. Jeon, Y.K. Park,
Oxidative desulfurization of refinery diesel pool fractions using LaVO4 photocatalyst, J. Ind.
Eng. Chem. 98 (2021) 283–288.

26
[64] S. Vedachalam, P. Boahene, A.K. Dalai, Oxidative desulfurization of heavy gas oil over
a Ti–TUD-1-supported keggin-type molybdenum heteropolyacid, Energy Fuels 34 (12) (2020)
15299–15312.

[65] A. Quintanilla, G. Vega, J. Carbajo, J.A. Casas, Y. Lei, K. Fujisawa, H. Liu, R. Cruz-
Silva, M. Terrones, P. Miranzo, M.I. Osendi, M. Belmonte, J. Fernández Sanz, Understanding
the active sites of boron nitride for CWPO: an experimental and computational approach,
Chem. Eng. J. 406 (2021) 126846.

[66] Y.S. Al-Hamdani, D. Alfè, O.A. von Lilienfeld, A. Michaelides, Tuning dissociation using
isoelectronically doped graphene and hexagonal boron nitride: water and other small
molecules, J. Chem. Phys. 144 (15) (2016) 154706.

[67] A. Mortezaee, M.A. Sobati, S. Movahedirad, S. Shahhosseini, An experimental

f
investigation on the oxidative desulfurization of a mineral lubricant base oil, J. Environ. Health

oo
Sci. Eng. 19 (2) (2021) 1951–1968.

r
[68] E.B. Krivtsov, A.K. Golovko, The kinetics of oxidative desulfurization of diesel fraction
-p
with a hydrogen peroxide-formic acid mixture, Pet. Chem. 54 (1) (2014) 51–57.
re
[69] M.M. Awad, Y.M. El-Toukhee, E.A. Hassan, K.K. Taha, Dearomatization of diesel by
solvent extraction: influence of the solvent ratio and temperature on diesel raffinate properties,
lP

Pet. Chem. 58 (5) (2018) 444–450.


na
ur
Jo

27
Jo
ur
na
lP
re
-p
ro
of
Jo
ur
na
lP
re
-p
ro
of
Jo
ur
na
lP
re
-p
ro
of
Jo
ur
na
lP
re
-p
ro
of
Jo
ur
na
lP
re
-p
ro
of
Jo
ur
na
lP
re
-p
ro
of
Jo
ur
na
lP
re
-p
ro
of
Jo
ur
na
lP
re
-p
ro
of
Jo
ur
na
lP
re
-p
ro
of
Jo
ur
na
lP
re
-p
ro
of
Jo
ur
na
lP
re
-p
ro
of
Jo
ur
na
lP
re
-p
ro
of
Jo
ur
na
lP
re
-p
ro
of
Jo
ur
na
lP
re
-p
ro
of
Jo
ur
na
lP
re
-p
ro
of
Jo
ur
na
lP
re
-p
ro
of
Jo
ur
na
lP
re
-p
ro
of
Jo
ur
na
lP
re
-p
ro
of
Jo
ur
na
lP
re
-p
ro
of
Jo
ur
na
lP
re
-p
ro
of
Jo
ur
na
lP
re
-p
ro
of
Jo
ur
na
lP
re
-p
ro
of
Jo
ur
na
lP
re
-p
ro
of
Declaration of interests

☒ The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.

☐The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:

of
ro
-p
re
lP
na
ur
Jo

You might also like