Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

pubs.acs.

org/jced Article

Batch and Continuous Adsorptive Desulfurization of Model Diesel


Fuels Using Graphene Nanoplatelets
Divyam Jha, Mohd Belal Haider, Rakesh Kumar, Wang Geun Shim,
and Balathanigaimani Marriyappan Sivagnanam*

Cite This: J. Chem. Eng. Data 2020, 65, 2120−2132 Read Online

ACCESS
Downloaded via RAJIV GANDHI INST PETROLEUM TECHLGY on July 9, 2020 at 03:49:11 (UTC).
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Metrics & More Article Recommendations

ABSTRACT: An experimental investigation on adsorptive de-


sulfurization for model diesel fuels (MDFs) is described using
graphene nanoplatelets (GNPs) as an adsorbent. Batch experi-
ments for a single component as well as a multicomponent system
were conducted. The single-component adsorption isothermal
experimental results for the batch process were well represented by
the Freundlich isotherm (thiophene (T), 2-methylthiophene (2-
MT)) and the Langmuir isotherm (dibenzothiophene (DBT))
models. The adsorption process kinetics fits with the pseudo-
second-order model for T, 2-MT, and DBT on GNPs. Both surface
and pore diffusions controlled the sorption process. A process
design of a single-stage batch-adsorber for the adsorption of T, 2-
MT, and DBT onto GNPs was also studied using the calculated
adsorption isotherm parameters. In addition, a fixed-bed adsorber was used for studying the continuous system at ambient
conditions for multicomponent MDFs. The main objective of continuous studies was to investigate the process variables’ effect on
the desulfurization. The effect of flow rate, bed height, and initial sulfur concentration on the adsorptive capacity of the adsorbent for
T, 2-MT, and DBT were evaluated. The breakthrough time and adsorptive capacity increased with increasing the bed height and
decreasing the flow rate and initial sulfur concentration. For continuous studies, more than 90% of the bed was saturated for all thio-
compounds within 100 min for the feed flow rate of 3 mL/min and adsorbent weight of 3 g. The Adam−Bohart model was used to
check the performance of the adsorption breakthrough curves. The Adam−Bohart model rate constant and adsorption capacity were
found to be 0.006 (mg/kg)−1 h−1 and 117.21 mg S/kg, respectively.

1. INTRODUCTION In order to increase the hydrodesulphurization activity of the


The presence of high levels of sulfur compounds in diesel fuels catalyst, researchers modified the CoMo catalyst using
is a major source of SOx emission, which contributes to air alumina-supported carbon nanofiber, boron, and bismuth.
pollution. Sulfur oxide emission also causes acid rain and Their results showed that these catalysts are promising in
deterioration of human health.1−4 Stringent rules have been obtaining ultraclean fuels.4−10 According to an estimation, to
implemented to regulate the sulfur content of fuels by most of reduce the sulfur level in diesel from 300 to less than 10 ppm,
the countries using desulfurization of liquid fuels. The emission the HDS reactor volume needed to be increased by a factor in
standards like European emission and Bharat Stage (BS) between 5 and 15, where the operating pressure condition
emission standards have been framed to control the various would be 1000 and 600 psi, respectively. Thus, HDS is not an
pollutants including sulfur emission to the <10 ppmw in the appropriate solution for treating the whole (100 wt %) fuel to
fuel at present.2,3 convert the fuel mass of less than 0.3 wt % (sulfur
Various processes have been used for desulfurization of compounds); it is not an efficient method for removing all
liquid fuels. Hydrodesulfurization (HDS) is a conventional kinds of sulfur compounds from the liquid hydrocarbon fuels
process for desulfurization of liquid fuels and is efficient in
removing mercaptans, thiophenes (T), and sulfides that are Received: December 30, 2019
present in diesel, that is, linear compounds but less effective for Accepted: February 25, 2020
the refractory sulfur compounds such as benzothiophene (BT), Published: March 10, 2020
dibenzothiophene (DBT), and 4,6-dimethyl dibenzothiophene
(4,6-DMDBT) present in the fuels and the probable reason is
the existing high steric hindrance.

© 2020 American Chemical Society https://dx.doi.org/10.1021/acs.jced.9b01204


2120 J. Chem. Eng. Data 2020, 65, 2120−2132
Journal of Chemical & Engineering Data pubs.acs.org/jced Article

because of high demand of temperature, pressure, and Table 1. List of Chemicals


hydrogen for the removal of refractory sulfur compounds.11
chemical name source purity (%) CAS-number
Therefore, to meet the stringent environmental norms,
alternative technologies12−16 have been developed and N-heptane Spectrochem >99 142-82-5
implemented. Among all techniques, adsorption desulfuriza- DBT Spectrochem ≥98 132-65-0
tion has gained huge attention17−21 as it does not require 2-MT Alfa Aesar ≥98 554-14-3
severe operating conditions and hydrogen consumption. T Alfa Aesar >98 110-02-1
However, the main challenge in adsorptive desulfurization is
the choice of the adsorbents which should have high specific surface area of the adsorbent. Furthermore, the pore
desulfurization capacities and should be thio-selective over size distribution of the adsorbent was obtained based on the
olefinic and aromatic compounds at the same time. Zeolites density functional theory (DFT) method.
emerged as one of the possible alternative adsorbents for
adsorptive desulfurization.22−25 They have high mechanical 3. METHODS
and thermal stabilities, and are size-selective for adsorptive 3.1. Preparation of MDFs. Three single-component (T, 2-
desulfurization. Several studies on adsorption desulfurization MT, and DBT) and a multicomponent MDFs named as MDFs
for liquid fuels have been performed with different materials
such as activated carbon, activated aluminas, meso-silicas,
silica−zirconia cogel, zeolites, ion exchange resins, Ti-HMSs,
Ni-based adsorbents and NiMos, mesoporous materials (MSU-
S) and cobalt, iron, and chromium (Fe2O3-MSU-S and CrO2-
MSU-S), cerium, phosphotungstic acid (HPW) and nickel
oxide-HPW (NiO-HPW) modified (MSU-S), and acid-treated
activated carbon. These aluminate silicate mesostructured
adsorbent showed significant sulfur removal.26−33
Graphene is a new captivating adsorbent that has gained the
researchers’ attention in recent years. The carbon in graphene
is sp2-bonded,is one-atom-thick, and has a two-dimensional
layer. Graphene also has remarkable properties, such as
excellent electrical, thermal, mechanical, optical properties
and has very high specific surface area. In addition, because of a
delocalized p-electron system which can form a strong
interaction with S-compounds, graphene has shown potential Figure 1. Schematic experimental setup for the fixed-bed adsorption
for use as an adsorbent for desulfurization.34−37 Graphene process.
nanoplatelets (GNPs) have been used for water pollutant
removal and HDS processes.38,39 Adsorptive desulfurization Table 2. Experimental Conditions for Fixed Bed Column
over GNPs has not been investigated as of now. Experiments
In this work, we have investigated the adsorptive adsorbent GNPs
desulfurization using GNPs as the adsorbent for the removal temperature ambient 25 °C
of T, 2-methylthiophene (2-MT) and DBT from model diesel pressure 1 atm
fuels (MDFs). To understand the fundamental adsorption adsorbent dose 3−10 g
process, adsorptive desulfurization was conducted in batch as MDFm flow rate 0.5−3.0 mL/min
well as continuous modes. The adsorption capacities of GNPs inlet sulfur content 1500 ppmw
for single-component MDFs and multicomponent MDFs were diameter of column 1 cm
studied in detail. Isotherm modeling was done for equilibrium height of bed 3−10 cm
studies. Kinetic and thermodynamic studies were conducted
for single-component MDFs and later it was compared with
the multicomponent system. The process design for the single-
batch adsorber process has also been studied. Further, the and MDFm were prepared. MDFsD, MDFsT, MDFsM contain n-
study of adsorption of DBT, 2-MT, and T was continued in a heptane and DBT, T, and 2-MT as a sulfur component,
fixed bed column for the continuous process to obtain respectively, whereas MDFm consists of n-heptane and T, 2-
breakthrough capacity of the adsorbent bed for different MT, and DBT as sulfur compounds. The single-component
variables. In addition, the Adam−Bohart model was used for MDFs with an initial concentration of 500 ppmw were
continuous studies. prepared by separately mixing 0.1982 g of DBT, 0.1370 g of 2-
MT, and 0.1046 g of T with 100 mL of n-heptane in three
2. EXPERIMENTAL SECTION different containers. The multicomponent MDFs (MDFm)
2.1. Materials. N-Heptane (purity > 99%), T, 2-MT, and were prepared with a total concentration of 1500 ppm by
DBT were purchased as given in Table 1. GNPs were mixing each 500 ppmw of T, 2-MT, and DBT in 100 mL of n-
purchased from Sigma-Aldrich. Adsorbents were kept in an heptane to study the effects of flowrate and bed height on
oven and dried overnight before using them for the adsorption continuous adsorption. MDFm were also prepared by varying
studies. The physical textural characteristics of GNPs were the initial concentrations of T, 2-MT, and DBT from 200 to
obtained using nitrogen adsorption/desorption (Micromeritics 600 ppmw to investigate the effect of initial concentrations of S
ASAP 2010) isotherm data measured at 77 K. The Brunauer− compounds on continuous adsorption. Further, because of the
Emmett−Teller equation was considered for the calculation of volatile nature of n-heptane, MDFs and MDFm were stored in
2121 https://dx.doi.org/10.1021/acs.jced.9b01204
J. Chem. Eng. Data 2020, 65, 2120−2132
Journal of Chemical & Engineering Data pubs.acs.org/jced Article

amber vials to avoid change in the sulfur concentration as


prepared.
3.2. Batch Adsorption Experiments. The adsorption
studies of DBT, 2-MT, and T on GNPs were conducted at 30
°C and 400 rpm to investigate the effect of dosage by varying
the adsorbent amounts from 0.025 to 0.4 g. A fixed volume of
20 mL of MDFs was used for all these batch adsorption
studies. All adsorption experiments were performed in an
Incubator Shaker (BR Biochem) for 60 min, except for the
kinetic studies. In case of kinetic studies, the experiments were
carried out to the specified time and then the specific sample
container itself was taken out for sample collection. This
phenomenon was chosen to not disturb the fuel to adsorbent
ratio in the middle of experiments. After each adsorption study,
the adsorbent was filtered out using a Whatman paper.
The concentrations of T, 2-MT, and DBT in the
desulfurized samples were analyzed by a gas chromatograph
equipped with a flame ionization detector (PerkinElmer Claurs
580 GC-FID detector). GC is equipped with an Elite-1 column
having a length of 50 m and an inner diameter of 200 μm.
Helium gas was used as carrier with a constant 1 cm3/min flow
rate. The injector temperature was kept at 340 °C with a
sample injection volume of 1 μL. The oven temperature was
varied from 50 to 350 °C, increasing the heating rate at 3 °C/
min.
The amount of sulfur adsorbed, qe (mg S/g A), is given by
eq 1
V
qe = × (C0 − Ce) Figure 2. (a) Nitrogen adsorption−desorption isotherm for GNPs at
W (1)
77 K; (b) DFT pore size distribution curve.
where Ce and C0 are the equilibrium and initial concentrations
of sulfur for W gram of adsorbent and V (L) volume of the
MDF. The % sulfur removal is given by the equation
(C0 − Ce)
% Sulfur removal = × 100
C0 (2)

3.3. Continuous Adsorption Experiments. In order to


discover the adsorption capacity of GNPs, dynamic adsorption
experiments were carried out by flowing MDFm through a fixed
bed adsorber of 1 cm diameter at 30 °C. To conduct the
dynamic adsorption experiments, a vertical custom-made glass
adsorber was used. A glass adsorption column loaded with 3−
10 g of GNPs was used for this continuous study. The
adsorbent sample was loaded with the support of glass wool.
The bottom section of the column was connected with a Figure 3. Effect of adsorbent weight on sulfur removal. (Temperature
peristaltic pump and a feed tank. MDFm was fed from the = 30 °C, initial sulfur concentration = 500 ppm, time = 60 min).
column in an upward direction continuously at different flow
rates (0.5−3.0 mL/min) as given in Figure 1. The MDFm
passed through the adsorbent bed and the samples were
collected at regular intervals of 5 min till the bed was saturated. values of GNPs are 635 m2/g, 3.34 nm, and 0.24 cm3/g
The sulfur present in the MDF was analyzed using GC-FID respectively.
after desulfurization. Further, the adsorption breakthrough 4.2. Equilibrium Studies and Adsorption Isotherms.
curves were obtained by normalizing the plot of amount of The adsorption of T, 2-MT, and DBT on GNPs was studied in
sulfur adsorbed versus cumulative time. Table 2 summarizes batch mode. The effect of adsorbent dosage on the
the experimental conditions used for these experiments. desulfurization capacity was studied by mixing 500 mg/L
sulfur solution at 30 °C. The adsorbent dosage was varied from
0.025 to 0.4 g for 20 mL of MDF shown in Figure 3 and Table
4. RESULTS AND DISCUSSIONS 3. To obtain the optimum condition, equilibrium studies were
4.1. Physical Properties of Adsorbents. The isotherms conducted for removal of T, 2-MT, and DBT over GNPs. It
for adsorption/desorption of nitrogen on GNPs at 77 K are can be observed that at equilibrium, the adsorption capacity
given in Figure 2a. The DFT result of GNPs is shown in Figure reaches maximum after which there was no further change in
2b. The surface area, average pore size, and total pore volume the sulfur concentration. The Langmuir equation was used to
2122 https://dx.doi.org/10.1021/acs.jced.9b01204
J. Chem. Eng. Data 2020, 65, 2120−2132
Journal of Chemical & Engineering Data pubs.acs.org/jced Article

Table 3. Sulfur Removal Efficiency for Different Adsorbent/ Table 4. Adsorption Isotherm Parameters
MDF Ratios
adsorption parameters
% sulfur removal isotherms DBT 2-MT T
adsorbent/20 mL of MDF (wt g) DBT 2-MT T Langmuir qm = 181.650 qm = 268.500 qm = 360.10 × 10−2
0.025 55.64 45.65 35.01 (mg/g) (mg/g) (mg/g)
0.050 66.37 51.47 37.22 b = 0.004 b = 0.002 b = 4.79 × 10−7
(g/mg) (g/mg)
0.075 72.41 55.29 40.21
R2 = 0.986 R2 = 0.943 R2 = 0.955
0.100 76.74 67.28 42.02
χ2 = 6.770 χ2 = 4.680 χ2 = 3.760
0.125 85.00 71.18 49.20
RMSE = 5.790 RMSE = 5.278 RMSE = 2.690
0.200 90.00 79.49 50.45
Freundlich KF = 3.010 KF = 0.151 KF = 0.088
0.250 94.50 81.46 64.89
nF = 1.61 nF = 0.835 nF = 0.893
0.300 95.00 88.13 70.45
R2 = 0.948 R2 = 0.989 R2 = 0.995
0.400 95.72 90.28 71.93
χ2 = 8.270 χ2 = 1.090 χ2 = 1.670
RMSE = 7.520 RMSE = 2.760 RMSE = 0.965
correlate the data obtained by experimental adsorption. The
Langmuir isotherm equation is given below40 With increasing time, DBT, 2-MT, and T removal was also
qmbCe increased till 60 min and then became constant. Therefore,
qe = optimum time for desulfurization is 60 min. In addition, with
1 + qmbCe (3) increasing adsorbent dosage from 0.025 to 0.4 g the adsorption
capacity also increases.22 The adsorptions by GNPs were
where qm and b signify the maximum sulfur adsorbed (mg/g) observed to be 95.7, 90.,1 and 75% for DBT, 2-MT, and T and
at saturation and affinity of adsorption, respectively. adsorptive capacities were 98, 80, and 58 mg/g, respectively.
The Freundlich isotherm is applied for multilayer adsorption These results are comparable with the desulfurization
and for adsorbents with heterogeneous surfaces. It is given by efficiency previously reported.8−10 Figure 4 shows the
the equation41 adsorption equilibrium data for DBT, 2-MT, and T on
qe = KFCe1/ nf GNPs. As shown in the figure, the experimental results of DBT
(4)
were correlated well with the Langmuir isotherm. The
nf and KF are the Freundlich constant and adsorption goodness of fit of predicted and experimental data was
coefficient (mg/g) (L/mg), respectively. The higher the analyzed using R2, χ2, and root mean square error (RMSE) and
adsorbent heterogeneity, the larger will be the value of nf. the value so obtained is given in Table 4.42

Figure 4. Adsorption isotherms on GNPs of (a) DBT, (b) 2-MT, and (c) T.

2123 https://dx.doi.org/10.1021/acs.jced.9b01204
J. Chem. Eng. Data 2020, 65, 2120−2132
Journal of Chemical & Engineering Data pubs.acs.org/jced Article

increase in the adsorption. This can be explained by the large


number of active sites available for adsorption of sulfur
compounds initially, which decrease with time because of
adsorption of adsorbate on the active sites, which leads to a
reduction in the sulfur removal rate. Further, the initial stage of
the adsorption of the molecules happens at the macropores
and mesopores, followed by the deeper penetration of
molecules into the micropores of the adsorbent. Therefore, it
is observed in this study that sulfur removal rate was reduced
mainly because of the increase in the resistance to the molecule
diffusion.43−45 Therefore, 60 min has been chosen as the
equilibrium contact time.
The adsorption rate slows down after the S-compound
saturates the meso- and micropores, which may be because the
Figure 5. Effect of time on sulfur removal. (Temperature = 30 °C, microspores experience a larger resistance in the later stage
initial sulfur concentration = 500 ppm, adsorbent weight = 0.01 g). when the S molecules move further.46
For interpreting the experimental data, the first-order
equation, the pseudo-second-order equation, and an intra-
particle diffusion equation were studied. The pseudo-first-order
kinetic is given by47,48
qt = qe(1 − e−k1t ) (5)
49
Similarly, the pseudo-second-order kinetic is given by
(k 2qe 2 ·t )
qt =
1 + (k 2qe ·t ) (6)
where k1 and k2 are the t pseudo-first-order and second-order
rate constants and qt is the adsorption at time t.
qt = k idt 1/2 + C (7)
Boyd’s surface diffusion model and Weber’s intraparticle
diffusion were used to predict the diffusion mechanism.50
The intraparticle controlled diffusion mechanism is given,
where C and Kid are the boundary layer thickness and
intraparticle diffusion coefficient (mg/g (min)0.5), respectively.
Boyd’s kinetic expression is given by
Bt = −0.4977 − ln(1 − F ) (8)

6
F=1− exp(−Bt )
π2 (9)
qt
F=
Figure 6. Kinetics for DBT, 2-MT, & T (a) pseudo-first-order and qe
(b) pseudo-second-order models.
where F is the equilibrium attained in the fraction.
MDFs containing DBT, 2-MT, and T in a concentration of
4.3. Kinetic Studies. The effect of contact time on the 500 ppm were kept in contact with GNPs for a range of time of
removal of T, 2-MT, and DBT adsorbed was investigated at a 10−90 min. The adsorption rates attain equilibrium at 60 min.
constant temperature of 30 °C. It can be observed that The adsorption rate increases rapidly initially because of
adsorption increases with increasing contact time and also the available free sites for adsorbates to adsorb on the surface.
maximum adsorption was obtained at 60 min as shown in Adsorption slows down at equilibrium because of the few
Figure 5. On further increasing of time, there was no significant available sites and the mutual repulsive forces between the

Table 5. Parameters for Kinetics of the Pseudo-First-Order Model for the Adsorption of DBT, 2-MT, & T on GNPs

Qe (mg/g)
k1 (L/min) calculated experimental R2 NSD ARE
DBT 0.044 6.504 153.970 0.860 68.540 1.654
2-MT 0.039 5.220 107.250 0.986 45.430 8.384
T 0.042 6.615 119.250 0.946 11.650 2.712

2124 https://dx.doi.org/10.1021/acs.jced.9b01204
J. Chem. Eng. Data 2020, 65, 2120−2132
Journal of Chemical & Engineering Data pubs.acs.org/jced Article

Table 6. Parameters for Kinetics of Pseudo-Second-Order Model for the Adsorption of DBT, 2-MT, & T on GNPs
Qe (mg/g)
k2 (g/mg min) calculated experimental R2 NSD ARE
−4
DBT 9.920 × 10 163.930 153.970 0.997 21.720 0.005
2-MT 8.792 × 10−4 125.010 107.250 0.998 29.690 1.923
T 5.495 × 10−4 114.940 119.250 0.993 5.920 0.884

Figure 7. Intraparticle diffusion of DBT, 2-MT, & T by the (a) Weber


and Morris model and the (b) Boyd kinetic model. Figure 8. (a) Effect of temperature on DBT, 2-MT, and T adsorption
on GNPs; (b) modeling for thermodynamic behavior of DBT, 2-MT,
Table 7. Intraparticle Diffusion Parameters for Adsorption and T on GNPs using the Van’t Hoff equation. (Temperature = 30
°C, initial sulfur concentration = 500 ppm, adsorbent weight = 0.01
of DBT, 2-MT, & T on GNPs by the Weber and Morris
g).
Model
kid (mg/g min0.5) C (mg/g) R2
DBT 9.085 79.480 0.976
2-MT 6.249 70.910 0.992 MT, and T plots are multilinear, concluding that both
T 8.116 37.285 0.967 intraparticle transport and surface diffusion control the
adsorption process.52
The slowest step between surface and intraparticle diffusion
adsorbate present in the solution and the adsorbent surface.51 was predicted by the Boyd kinetic expression. The linearity of
The kinetics involved to remove DBT, 2-MT, and T by GNPs the plot in Figure 7b is confirmed by high R2. This means that
was investigated using linear plots of log (qe − qt) versus t for for DBT, 2-MT, and T adsorption surface diffusion is the
the pseudo-first-order and t/qt versus t for the pseudo-second- controlling factor. Further, Table 7 shows that boundary layer
order kinetic models as shown in Figure 6. The kinetic thickness of DBT is greater than those of 2-MT and T; thus
parameters obtained are given in Tables 5 and 6. The kinetic adsorption of DBT on GNPs possesses more surface
controlling behavior of the system was analyzed using controlling compared to 2-MT and T.
regression coefficient (R2), average relative error (ARE), and 4.4. Error Data Analysis. The error functions were
normalized standard deviation (NSD). These error analysis evaluated using residual RMSE, coefficient of regression
shows that DBT, 2-MT, and T removal by GNPs follows (R2), and chi-square test (χ2), which were employed to
pseudo-second-order kinetics. check the fitness of the models. The empirical equations are as
The Kid and C values are the slope and intercept, follows
respectively, of the plot between qt and t0.5 as shown in Figure n
7a. The figure shows that the plot is multilinear, which means 1 − ∑n = 1 (qe, n − qm, n)2
2
that the overall adsorption process is controlled by two or R = n
∑n = 1 (qe, n − qe, n )2 (10)
more steps. It can be observed from Figure 7a that DBT, 2-
2125 https://dx.doi.org/10.1021/acs.jced.9b01204
J. Chem. Eng. Data 2020, 65, 2120−2132
Journal of Chemical & Engineering Data pubs.acs.org/jced Article

Table 8. Thermodynamic Parameters for the Adsorption of DBT, 2-MT, & T on GNPs
Kd ΔG0 (kJ/mol) ΔH0 (kJ/mol) ΔS0 (kJ/(mol K))
T (°C) DBT 2-MT T DBT 2-MT T DBT 2-MT T DBT 2-MT T
40 1.648 1.395 1.238 −1.302 −0.868 −0.558
50 1.586 1.358 1.203 −1.241 −0.824 −0.497 −8.751 −4.169 −3.150 −0.024 −0.011 −0.008
60 1.358 1.324 1.133 −0.849 −0.779 −0.346
70 1.252 1.203 1.124 −0.641 −0.527 −0.334

n 2-MT (−4.169 kJ/mol), and T (−3.15 kJ/mol) indicates that


1
RMSE = ∑ (q − qm,n)2 the adsorption follows a physisorption behavior. The lower
n − 1 n = 1 e, n (11) ΔH0 value implies that the adsorbate molecules and the
adsorbent surface are weak.54−56
n (qe, n − qm, n)2 4.6. Design of Batch Sorption from Isotherm Data.
χ2 = ∑ For designing a single-batch system, the adsorption isotherm
qe, n

ÅÄÅ ÑÉÑ 2
n=1 (12) was used.57−59 The single-stage batch adsorption system

Å
Å ÑÑ
schematic is shown in Figure 9a. Consider MDF of volume

∑ ÅÅÅÅÅ t ,e t ,m ÑÑÑÑÑ
N q − q V (L) and the DBT concentration that reduces from C0 to C1
N − 1 i = 1 ÅÅ ÑÑÖ
1

Ç
NSD = 100 × (mg/L). The adsorbent dosage is M (mg) and the s-loading
qt ,e changes from q0 to q1. Initially, q0 = 0 and as time passes, the
Ä É
N Å ÅÅ q − q ÑÑÑ
i (13)
mass balance equates the S-compound removed from the MDF
ÅÅ t ,e t ,m Ñ
ÑÑ
∑ ÅÅ ÑÑ
by adsorption. The mass balance for Figure 9a is given by

N i = 1 ÅÅÅÅ ÑÑ
100

Ç ÑÖ
ARE = V (C0 − C1) = M(q0 − q1) = M q
qt ,e 1 (17)
i (14)
At equilibrium, C1 approaches Ce and q1 approaches qe.
where qt,m and qt,e are the model predicted and experimental As adsorption of DBT onto GNPs follows the Langmuir
amount of sulfur adsorbed on GNPs. isotherm, and 2-MT &T follow the Freundlich isotherm, the
4.5. Thermodynamic Parameters. The spontaneity of adsorber design equation can be written as
the process was determined by taking consideration of energy
and entropy. These parameters give the practical viability of M (C0 − Ce) (C0 − Ce) C −C
the process. To determine the thermodynamic parameters, = = = 0q bC e
V q1 qe m e
adsorptions at different temperatures (20, 30, 35, and 40 °C) 1 + bCe (18)
were conducted for the adsorption system, DBT, 2-MT, and T.
The effect of temperature on the adsorption is shown in M (C0 − Ce) (C0 − Ce) C −C
Figure 8a. It can be observed that with increasing temperature, = = = 0 1/ ne
V q1 qe qmK f f (19)
the sulfur removal decreases. This shows the physisorption
behavior of DBT, 2-MT, and T adsorption on GNPs. Using eq Figure 9b−d shows the predicted GNP dosage required to
15, the change in Gibbs free energy was calculated. Van’t Hoff remove DBT, 2-MT, and T from MDFs having initial
plot was generated to determine the change in enthalpy and concentrations of 500 mg/L for 90, 80, 70, and 60% with
entropy as shown in Figure 8b. The Gibbs free energy (ΔG0) different solution volumes of 1, 2, 3, 4, 5, 6, and 7 L. The
change was calculated by amount of GNPs required to remove DBT from the MDF of
concentration 500 mg/L was 0.013, 0.027, 0.041, and 0.055 g,
ΔG 0 = −RT ln Kd (15) for 2-MT was 0.027, 0.055, 0.085, and 0.011 g, and for T was
concentration of S‐compound adsorbed 0.011, 0.022, 0.033, and 0.445 g, respectively, for MDF
Kd = volumes of 1, 2, 3, and 4 L, respectively.
concentration of S‐compound remained in solution 4.7. Comparison with the Multicomponent System.
where ΔG0 is in (kJ/mol), R is in kJ/kmol·K. Kd is the The multicomponent system was investigated to observe its
distribution coefficient and T is the temperature in K. The adsorptive capacity on GNPs. MDFm with a total sulfur
concentration of 1500 ppmw was mixed with GNPs in

ij ΔS 0 yz ij ΔH 0 yz
Van’t Hoff equation is given by

ln Kd = jjj zzz − jjj zz


different dosages and it was observed that the % sulfur removal

z
for each sulfur component decreased as compared to a single-
k R { k RT { (16)
component system. It was seen that the removal of DBT, 2-
MT, and T was decreased from 95.72, 90.21, and 75.42 to
Thermodynamic parameters of the system are shown in 93.24, 82.32, and 64.88%, respectively, shown in Figure 10.
Table 8. The value of ΔG0 lies in between −20 and 0 kJ mol−1 The reason for the reduction in sulfur removal is competitive
for physisorption and −80 to −400 kJ mol −1 for behavior of each sulfur compound when all present together in
chemisorption.53 The feasibility of the system is checked by a single MDF. Further, there were no significant changes in the
the negative value of ΔG0.31 The exothermic behavior of the equilibrium time observed in both cases shown in Figure 11a.
adsorption process is confirmed by the negative values of ΔH0. In addition, the multicomponent system was also following
The negative value of ΔS0 indicates the decrease in the degree pseudo-second-order kinetics as shown in Figure 11b, similar
of freedom of the adsorbed DBT, 2-MT, and T molecules on to single-component MDFs. However, the values for the
GNPs. Hence, the systems show nonrandom adsorption second-order rate constants given in Table 9 indicate that the
behavior. The change in enthalpy for DBT (−8.751 kJ/mol), adsorption process kinetic gets slower for the multicomponent
2126 https://dx.doi.org/10.1021/acs.jced.9b01204
J. Chem. Eng. Data 2020, 65, 2120−2132
Journal of Chemical & Engineering Data pubs.acs.org/jced Article

Figure 10. Effect of adsorbent weight on sulfur removal for the


multicomponent system. (Temperature = 30 °C, initial sulfur
concentration = 500 ppm, time = 60 min).

similar trend observed for all DBT, 2-MT, and T for


temperature variation. The process possesses an exothermic
behavior similar to a single-component system as shown in
Figure 12.

5. CONTINUOUS ADSORPTIVE EXPERIMENTS


The continuous adsorptive desulfurization experiments were
conducted in a fixed bed with the adsorbent. The effects of bed
height, flow rate, and initial sulfur concentration were analyzed.
All the experiments were repeated for three runs and each time
the error was found less than 5%.
5.1. Effect of Bed Height and Flow Rate. The empty
bed contact time (EBCT) plays an important role in
determining the adsorbent performance in continuous
operation. Different adsorbent dosages and different inlet
flow rates were used to investigate the EBCT effect on the
breakthrough curve performance. The flow rates of the MDF
were varied from 0.2 to 2.0 mL/min in the column. As
observed in Figure 13, with increasing flow rate, sulfur removal
decreases. This is due to decreased contact time between the
adsorbent and the MDFm. The product yield also decreases in
higher flow rates with the residence time in the time range
from 0.1 to 2.4 s.60,61 Also, it can be observed from Figure 13
that with increasing flow rate, bed saturation also increases. For
example, in the case of DBT with increasing feed flow rate
from 0.5 to 3 mL/min, the value of Ct/C0 increases from 0.29
to 0.89. This behavior is also observed in MT and T. This can
be explained based on the number of active sites present that
got saturated as a low amount of feed is present. In addition,
with increasing flow rate, saturation time also increases.
Further, Figure 14 shows the bed height effect on sulfur
removal. As the bed height increases, the sulfur removal also
increases and finally gets saturated. This is because with
increasing adsorbent height, the adsorbent mass also increased
and therefore the corresponding adsorption sites also increases.
It can also be observed from Figure 14 that with increasing
Figure 9. (a) Single-stage batch-adsorber design; (b−d) adsorbent
adsorbent dosage the breakthrough curve shifted toward higher
mass (M) against the volume of solution treated (L) for DBT, 2-MT,
and T. (Temperature = 30 °C, initial sulfur concentration = 500 saturation. For instance, in the case of DBT on increasing
ppm). adsorbent dosage from 1 to 5 g, the value of Ct/C0 increases
from 0.254 to 0.68. This can be ascribed to more adsorbent
dosage having more active sites to accommodate the thio-
system as compared to the single-component MDFs. The compounds.
presence of other S-compounds is the primary reason for 5.2. Effect of Initial Sulfur Concentration. Initial
slowing down the reaction kinetics slightly. Also, there was a concentration effect of sulfur dosage on column performance
2127 https://dx.doi.org/10.1021/acs.jced.9b01204
J. Chem. Eng. Data 2020, 65, 2120−2132
Journal of Chemical & Engineering Data pubs.acs.org/jced Article

Figure 12. Effect of temperature on DBT, 2-MT, & T adsorption on


GNPs in the multicomponent system. (Temperature = 30 °C, initial
sulfur concentration = 500 ppm, adsorbent weight = 0.01 g).

concentration gradient developed in the system. With


increasing initial feed concentration of the breakthrough
curve, the slope become steeper and the breakthrough time
will also reduce. Therefore, initial concentration has a
significant effect on the saturation rate and breakthrough
time.62 With increasing feed concentration, the driving force
for mass transfer also increases. The overall effect will increase
the adsorption capacity as shown in Figure 15.63 It can be
observed from Figure 15 that with increasing concentration of
sulfur compounds, the saturation curve gets diminished. For
example, the value of Ct/C0 decreases from 0.76 to 0.32 on
increasing the concentration from 200 to 600 ppmw.
5.3. Modeling of Column Study Results: Bohart−
Adams Model. To facilitate the design and analysis of a full-
scale system, various mathematical models have been
developed. In addition, various adsorption models must be
investigated by the designer along with their applications to
investigate the adsorber performance for desulfurization. The
breakthrough curve is affected by the flow rates, bed height,
and feed concentration. There is a critical minimum bed height
below which the solute concentration in the effluent increases
rapidly from the first appearance of the effluent. At
Figure 11. (a) Effect of time on (b) pseudo-first-order and (c) equilibrium, the area between the curve and a line at Ct/C0
pseudo-second-order kinetics for DBT, 2-MT, & T adsorption on = 1.0 is proportional to the sulfur adsorbed. The Adam−
GNPs in the multicomponent system. (Temperature = 30 °C, initial Bohart model was used to predict the dynamic behavior of the
sulfur concentration = 500 ppm, adsorbent weight = 0.01 g). column. It is based on the assumption that the adsorption rate
is directly proportional to adsorbent residual capacity and the
was studied using different amounts of DBT, 2-MT, and T. initial concentration of feed. The present analysis is based on
The concentration was varied from 200 to 600 ppmw of DBT, the estimation of maximum adsorption capacity and kinetic
2-MT, and T each, keeping the concentration of other two S- constants from the Adam−Bohart model.
compounds constant at 500 ppmw each, and it was observed The bed-depth service time (BDST) model was based on
for all the S-compounds when the highest sulfur concentration the fact that the service time (t) and bed height (Z) are
was found to be the maximum. This is due to the correlated with the process parameters by the equation. Bed

Table 9. Parameters for Kinetics of Pseudo−First and Second-Order Models for the Adsorption of DBT, 2-MT, and T on
GNPs for the Multicomponent System

k1 (L/min) k2 Qe (mg/g)
(g/mg min) calculated experimental calculated experimental R2 NSD ARE R2 NSD ARE
1st order 2nd order 1st order 2nd order
DBT 0.052 6.651 × 10−4 7.287 125.270 140.840 125.270 0.887 5.670 5.876 0.987 1.960 0.856
2-MT 0.049 5.761 × 10−4 7.128 97.550 113.630 97.550 0.838 18.560 8.654 0.981 6.450 3.765
T 0.042 4.128 × 10−4 5.446 64.730 76.330 64.730 0.986 9.760 4.923 0.995 2.980 1.456

2128 https://dx.doi.org/10.1021/acs.jced.9b01204
J. Chem. Eng. Data 2020, 65, 2120−2132
Journal of Chemical & Engineering Data pubs.acs.org/jced Article

Figure 14. Effect of bed height for the multicomponent MDF ((a)
DBT, (b) 2-MT, and (c) T) on GNPs. (Flow rate of MDF = 1 mL/
min, temperature = 30 °C, initial sulfur concentration = 1500 ppm).

depth depends directly on the time service as given by the


following equation

t=
N0Z

t
(
ln 0 C t
C −C
)
C0V K aC 0 (20)
where C and C0 are the effluent and initial concentration (mg/
L), v is the velocity (cm/min), N0 is the adsorption capacity
Figure 13. Effect of flow rate for the multicomponent MDF ((a) (mg adsorbed/liter of solution), Ka is the adsorption rate
DBT, (b) 2-MT, and (c) T) on GNPs. (Adsorbent weight = 3 g, constant (L/(mg h)), t is the time (min), and Z is the bed
temperature = 30 °C, initial sulfur concentration = 1500 ppm). height (cm). The above equation can be written in the form of
a straight line.
Table 10. Constants of the BDST Equation

flow rate (mL/min) initial concentration (mg/L) breakthrough (%) N0 (mg/g) Ka (mg/g)−1 min−1 R2
DBT 1.0 500 30 144.900 0.00042 0.981
2-MT 1.0 500 30 124.900 0.00021 0.957
T 1.0 500 30 117.400 0.00013 0.992

2129 https://dx.doi.org/10.1021/acs.jced.9b01204
J. Chem. Eng. Data 2020, 65, 2120−2132
Journal of Chemical & Engineering Data pubs.acs.org/jced Article

t = m′Z − B (21)

where m′ is the slope and represents the time required by the


adsorption zone to travel a unit length through the adsorbent.
Figure 16 shows the plot between the bed depth and service
time. The adsorption capacity was found to be 496 mg S/kg of
adsorbent and the rate constant is 0.006 (mg/kg)−1 h−1,
respectively, as shown in Table 10.

6. CONCLUSIONS
In this work, adsorptions of MDFs containing sulfur
compounds over GNPs were conducted in both batch and
continuous adsorption modes. The individual studies of DBT,
2-MT, and T removal by GNPs from the batch studies showed
that the removal capacity of GNPs follows the order DBT > 2-
MT > T. The Langmuir adsorption isotherms model was fitted
best to represent the adsorption equilibrium experimental data
of DBT on GNPs, whereas the Freundlich isotherm fit well
with the experimental adsorption data of both 2-MT and T on
GNPs. The batch adsorption of DBT, 2-MT, and T on GNPs
followed the pseudo-second-order kinetic. Each batch
adsorption process was well interpreted by a multistage
diffusion model. The diffusion of DBT, 2-MT, and T on
GNPs follows the complex diffusion model, which means that
both intraparticle and surface diffusions were involved in the
batch adsorption mechanism. The negative values of ΔG, ΔH,
and ΔS from the thermodynamic studies of DBT, 2-MT, and T
removal by GNPs indicated that the adsorption was
spontaneous in nature with an exothermic behavior and with
less randomness on the GNPs’ surface. Investigation of a
continuous desulfurization study shows that the adsorption
breakthrough curve strongly depends on the bed height, flow
rate, and initial S-compounds’ concentration. The determined
adsorption capacities and the rate constants from the
continuous adsorption studies showed that the adsorptive
removal of DBT by GNPs is relatively predominant compared
to both 2-MT and T removal. These results suggest that the
GNPs have potential to be used industrially for the adsorptive
desulfurization process because of their high adsorptive
capacities.
Figure 15. Effect of initial concentration of S-compounds in the
multicomponent MDF ((a) DBT, (b) 2-MT, and (c) T) on GNPs.
(Temperature = 30 °C, flow rate of MDF = 1 mL/min, adsorbent
weight = 3 g).
■ AUTHOR INFORMATION
Corresponding Author
Balathanigaimani Marriyappan Sivagnanam − Department of
Chemical Engineering, Rajiv Gandhi Institute of Petroleum
Technology, Jais 229304, India; orcid.org/0000-0001-
5862-6215; Phone: +91-535-27004502; Email: msbala@
rgipt.ac.in
Authors
Divyam Jha − Department of Chemical Engineering, Rajiv
Gandhi Institute of Petroleum Technology, Jais 229304, India;
orcid.org/0000-0001-7303-8913
Mohd Belal Haider − Department of Chemical Engineering,
Rajiv Gandhi Institute of Petroleum Technology, Jais 229304,
India; orcid.org/0000-0002-3228-391X
Rakesh Kumar − Department of Chemical Engineering, Rajiv
Gandhi Institute of Petroleum Technology, Jais 229304, India;
orcid.org/0000-0001-9926-6156
Wang Geun Shim − Department of Polymer Science and
Figure 16. Plot of bed depth service time vs bed height (Adam− Engineering, Sunchon National University, Suncheon-si 57922,
Bohart Model). Jeollanam-do, Republic of Korea
Complete contact information is available at:
2130 https://dx.doi.org/10.1021/acs.jced.9b01204
J. Chem. Eng. Data 2020, 65, 2120−2132
Journal of Chemical & Engineering Data pubs.acs.org/jced Article

https://pubs.acs.org/10.1021/acs.jced.9b01204 Carbon for Hydrodesulfurization of Liquid Fuels. J. Mol. Liq. 2018,


272, 715−721.
Notes (18) Danmaliki, G. I.; Saleh, T. A.; Shamsuddeen, A. A. Response
Surface Methodology Optimization of Adsorptive Desulfurization on
The authors declare no competing financial interest.


Nickel/Activated Carbon. Chem. Eng. J. 2017, 313, 993−1003.
(19) Danmaliki, G. I.; Saleh, T. A. Effects of Bimetallic Fe−Ce
ACKNOWLEDGMENTS Nanoparticles on the Desulfurization of Thiophenes Using Activated
The research work was funded by the Rajiv Gandhi Institute of Carbon. Chem. Eng. J. 2017, 307, 914−927.
(20) Saleh, T. A.; Siddiqui, M. N.; Al-arfaj, A. A. Kinetic and
Petroleum Technology (RGIPT), Jais. The authors express Intraparticle Diffusion Studies of Carbon Nanotubes-Titania for
their sincere gratitude to the Director, RGIPT, Jais, for his kind Desulfurization of Fuels. Pet. Sci. Technol. 2016, 34, 1468−1474.
support. (21) Saleh, T. A.; Danmaliki, G. I. Adsorptive Desulfurization of

■ REFERENCES
(1) Verma, R. Fundamentals of Hydro-Processing from Indian Oil
Dibenzothiophene from Fuels by Rubber Tyres-Derived Carbons:
Kinetics and Isotherms Evaluation. Process Saf. Environ. Prot. 2016,
102, 9−19.
(22) Mikhail, S.; Zaki, T.; Khalil, L. Desulfurization by an
Publication, 2005.
Economically Adsorption Technique. Appl. Catal., A 2002, 227,
(2) Gense, N. L. J.; Riemersma; Such, C.; Ntziachristos, L. TNO
265−278.
Report: Euro VI Technologies and Costs for Heavy Duty Vehicles The
(23) Ma, X.; Sun, L.; Song, C. A New Approach to Deep
Expert Panels Summary of Stakeholders Responses, 2006; pp 30−34.
(3) Chaudhuri, S. Bharat Emission Standards; Alternative Fuels Pros Desulfurization of Gasoline, Diesel Fuel and Jet Fuel by Selective
& Cons, 2014. Adsorption for Ultra-Clean Fuels and for Fuel Cell Applications.
(4) Xue, M.; Chitrakar, R.; Sakane, K.; Hirotsu, T.; Ooi, K.; Catal. Today 2002, 77, 107−116.
Yoshimura, Y.; Toba, M.; Feng, Q. Preparation of Cerium-Loaded Y- (24) Ibrahim, A.; Xian, S. B.; Wei, Z. Desulfurization of FCC Gas
Zeolites for Removal of Organic Sulfur Compounds from Hydro- Oil by Solvent Extraction, Photooxidation, and Oxidizing Agents. Pet.
desulfurizated Gasoline and Diesel Oil. J. Colloid Interface Sci. 2006, Sci. Technol. 2004, 22, 287−301.
298, 535−542. (25) Shan, G.; Liu, H.; Xing, J.; Zhang, G.; Wang, K. Separation of
(5) Song, C.; Ma, X. New Design Approaches to Ultra-Clean Diesel Polycyclic Aromatic Compounds from Model Gasoline by Magnetic
Fuels by Deep Desulfurization and Deep Dearomatization. Appl. Alumina Sorbent Based on π-Complexation. Ind. Eng. Chem. Res.
Catal., B 2003, 41, 207−238. 2004, 43, 758−761.
(6) Farag, H. Selective Adsorption of Refractory Sulfur Species on (26) Ahmadi, M.; Mohammadian, M.; Khosravi-nikou, M. R.;
Active Carbons and Carbon Based CoMo Catalyst. J. Colloid Interface Baghban, A. Experimental , Kinetic , and Thermodynamic Studies of
Sci. 2007, 307, 1−8. Adsorptive Desulfurization and Denitrogenation of Model Fuels
(7) Hernández-Maldonado, A. J.; Yang, F. H.; Qi, G.; Yang, R. T. Using Novel Mesoporous Materials. J. Hazard. Mater. 2019, 374,
Desulfurization of Transportation Fuels by π-Complexation Sorbents: 129−139.
Cu(I)-, Ni(II)-, and Zn(II)-Zeolites. Appl. Catal., B 2005, 56, 111− (27) Mohammadian, M.; Khosravi-Nikou, M. R.; Shariati, A.;
126. Aghajani, M. Model Fuel Desulfurization and Denitrogenation Using
(8) Al-jamimi, H. A.; Bagudu, A.; Saleh, A. An Intelligent Approach Copper and Cerium Modified Mesoporous Material ( MSU - S )
for the Modeling and Experimental Optimization of Molecular through Adsorption Process. Clean Technol. Environ. Policy 2018, 20,
Hydrodesulfurization over AlMoCoBi Catalyst. J. Mol. Liq. 2019, 278, 95−112.
376−384. (28) Sasanipour, J.; Shariati, A.; Aghajani, M. Dibenzothiophene
(9) Al-hammadi, S. A.; Al-amer, A. M.; Saleh, T. A. Alumina-Carbon Removal from Model Fuel Using an Acid Treated Activated Carbon.
Nanofiber Composite as a Support for MoCo Catalysts in Pet. Sci. Technol. 2017, 35, 2066−2073.
Hydrodesulfurization Reactions. Chem. Eng. J. 2018, 345, 242−251. (29) Mohammadian, M.; Ahmadi, M.; Khosravi-Nikou, M. R.
(10) Saleh, T. A.; Al-hammadi, S. A.; Al-amer, A. M. Effect of Boron Adsorptive desulfurization and denitrogenation of model fuel by
on the Efficiency of MoCo Catalysts Supported on Alumina for the mesoporous adsorbents (MSU-S and CoO-MSU-S). Pet. Sci. Technol.
Hydrodesulfurization of Liquid Fuels. Process Saf. Environ. Prot. 2019, 2017, 35, 608−614.
121, 165−174. (30) Montazerolghaem, M.; Rahimi, A.; Seyedeyn-Azad, F.
(11) Yang, Y.; Lv, G.; Deng, L.; Lu, B.; Li, J.; Zhang, J.; Shi, J.; Du, S. Equilibrium and Kinetic Modeling of Adsorptive Sulfur Removal
Ultra-Deep Desulfurization of Diesel Fuel via Selective Adsorption from Gasoline by Synthesized Ce-Y Zeolite. Appl. Surf. Sci. 2010, 257,
over Modified Activated Carbon Assisted by Pre-Oxidation. J. Clean. 603−609.
Prod. 2017, 161, 422−430. (31) Ahmadi, M.; Anvaripour, B.; Khosravi-nikou, M. R.;
(12) Monticello, D. J. Biodesulfurization and the Upgrading of Mohammadian, M. Selective denitrogenation of model fuel through
Petroleum Distillates. Curr. Opin. Biotechnol. 2000, 11, 540−546. iron and chromium modified microporous materials (MSU-S). J.
(13) Gray, K. A.; Mrachko, G. T.; Squires, C. H. Biodesulfurization Environ. Chem. Eng. 2017, 5, 849−860.
of Fossil Fuels. Curr. Opin. Microbiol. 2003, 6, 229−235. (32) Ban, L.-L.; Liu, P.; Ma, C.-H.; Dai, B. Deep Extractive
(14) Collins, F. M.; Lucy, A. R.; Sharp, C. Oxidative Desulphurisa- Desulfurization of Diesel Fuels by FeCl3 / Ionic Liquids. Chin. Chem.
tion of Oils via Hydrogen Peroxide and Heteropolyanion Catalysis. J. Lett. 2013, 24, 755−758.
Mol. Catal. A: Chem. 1997, 117, 397−403. (33) Rashidi, S.; Khosravi Nikou, M. R.; Anvaripour, B. Adsorptive
(15) Jha, D.; Haider, M. B.; Kumar, R.; Balathanigaimani, M. S. Desulfurization and Denitrogenation of Model Fuel Using HPW and
Extractive Desulfurization of Dibenzothiophene Using Phosphonium- NiO-HPW Modified Aluminosilicate Mesostructures. Microporous
Based Ionic Liquid: Modeling of Batch Extraction Experimental Data Mesoporous Mater. 2015, 211, 134−141.
and Simulation of Continuous Extraction Process. Chem. Eng. Res. (34) Wang, S.; Sun, H.; Ang, H. M.; Tadé, M. O. Adsorptive
Des. 2016, 111, 218−222. Remediation of Environmental Pollutants Using Novel Graphene-
(16) Jha, D.; Mubarak, N. M.; Haider, M. B.; Kumar, R.; Based Nanomaterials. Chem. Eng. J. 2013, 226, 336−347.
Balathanigaimani, M. S.; Sahu, J. N. Adsorptive Removal of (35) Yuan, X.; Wang, Y.; Wang, J.; Zhou, C.; Tang, Q.; Rao, X.
Dibenzothiophene from Diesel Fuel Using Microwave Synthesized Calcined Graphene/MgAl-Layered Double Hydroxides for Enhanced
Carbon Nanomaterials. Fuel 2019, 244, 132−139. Cr(VI) Removal. Chem. Eng. J. 2013, 221, 204−213.
(17) Saleh, T. A.; AL-Hammadi, S. A.; Abdullahi, I. M.; Mustaqeem, (36) Wang, H.; Yuan, X.; Wu, Y.; Huang, H.; Zeng, G.; Liu, Y.;
M. Synthesis of Molybdenum Cobalt Nanocatalysts Supported on Wang, X.; Lin, N.; Qi, Y. Adsorption Characteristics and Behaviors of

2131 https://dx.doi.org/10.1021/acs.jced.9b01204
J. Chem. Eng. Data 2020, 65, 2120−2132
Journal of Chemical & Engineering Data pubs.acs.org/jced Article

Graphene Oxide for Zn(II) Removal from Aqueous Solution. Appl. (59) Crittenden, B.; Thomas, W. J. Adsorption Technology and
Surf. Sci. 2013, 279, 432−440. Design; Elsevier, 1998; p 288.
(37) Ai, L.; Jiang, J. Removal of Methylene Blue from Aqueous (60) Saleh, T. A.; Sulaiman, K. O.; AL-Hammadi, S. A.; Dafalla, H.;
Solution with Self-Assembled Cylindrical Graphene-Carbon Nano- Danmaliki, G. I. Adsorptive Desulfurization of Thiophene, Benzo-
tube Hybrid. Chem. Eng. J. 2012, 192, 156−163. thiophene and Dibenzothiophene over Activated Carbon Manganese
(38) Al-Khateeb, L. A.; Almotiry, S.; Salam, M. A. Adsorption of Oxide Nanocomposite: With Column System Evaluation. J. Clean.
Pharmaceutical Pollutants onto Graphene Nanoplatelets. Chem. Eng. Prod. 2017, 154, 401−412.
J. 2014, 248, 191−198. (61) Fox, B. R.; Brinich, B. L.; Male, J. L.; Hubbard, R. L.; Siddiqui,
(39) Guo, K.; Gu, M.; Yu, Z. Carbon Nanocatalysts for M. N.; Saleh, T. A.; Tyler, D. R. Enhanced Oxidative Desulfurization
Aquathermolysis of Heavy Crude Oil: Insights into Thiophene in a Film-Shear Reactor. Fuel 2015, 156, 142−147.
Hydrodesulfurization. Energy Technol. 2017, 5, 1228−1234. (62) Goel, J.; Kadirvelu, K.; Rajagopal, C.; Kumar Garg, V. Removal
(40) Foo, K. Y.; Hameed, B. H. Insights into the Modeling of of Lead(II) by Adsorption Using Treated Granular Activated Carbon:
Adsorption Isotherm Systems. Chem. Eng. J. 2010, 156, 2−10. Batch and Column Studies. J. Hazard. Mater. 2005, 125, 211−220.
(41) Sentorun-Shalaby, C.; Saha, S. K.; Ma, X.; Song, C. (63) Ko, D. C. K.; Porter, J. F.; Mckay, G. M. Fixed Bed Studies for
Mesoporous-Molecular-Sieve-Supported Nickel Sorbents for Adsorp- the Sorption of Metal Ions onto Peat. Process Saf. Environ. Protect.
tive Desulfurization of Commercial Ultra-Low-Sulfur Diesel Fuel. 2003, 81, 73−86.
Appl. Catal., B 2011, 101, 718−726.
(42) Srivastav, A.; Srivastava, V. C. Adsorptive Desulfurization by
Activated Alumina. J. Hazard. Mater. 2009, 170, 1133−1140.
(43) Taylor, P.; Saleh, T. A. Carbon Nanotubes/Silica Nanoparticles
and Their Use for Adsorption of Pb (II): From Surface Properties to
Sorption Mechanism. Desalin. Water Treat. 2016, 57, 10730.
(44) Saleh, T. A. Mercury Sorption by Silica/Carbon Nanotubes and
Silica/Activated Carbon: A Comparison Study. J. Water Supply: Res.
Technol.–AQUA 2015, 64, 892−903.
(45) Saleh, T. A. Isotherm, Kinetic , and Thermodynamic Studies on
Hg (II) Adsorption from Aqueous Solution by Silica- Multiwall
Carbon Nanotubes. Environ. Sci. Pollut. Res. 2015, 22, 16721−16731.
(46) Suresh, S.; Srivastava, V. C.; Mishra, I. M. Studies of Adsorption
Kinetics and Regeneration of Aniline, Phenol, 4-Chlorophenol and 4-
Nitrophenol by Activated Carbon. Chem. Ind. Chem. Eng. Q. 2013, 19,
195−212.
(47) Lagergren, S. About the Theory of So-Called Adsorption of
Soluble Substances. K. Sven. Vetenskapsakad. Handl. 1898, 24, 1−39.
(48) Ishaq, M.; Sultan, S.; Ahmad, I.; Ullah, H.; Yaseen, M.; Amir, A.
Adsorptive Desulfurization of Model Oil Using Untreated, Acid
Activated and Magnetite Nanoparticle Loaded Bentonite as
Adsorbent. J. Saudi Chem. Soc. 2017, 21, 143−151.
(49) Ho, Y. S.; McKay, G. Comparative Sorption Kinetic Studies of
Dye and Aromatic Compounds onto Fly Ash. J. Environ. Sci. Health,
Part A: Toxic/Hazard. Subst. Environ. Eng. 1999, 34, 1179−1204.
(50) Ahmed, M. J. K.; Ahmaruzzaman, M. Adsorptive Desulfuriza-
tion of Feed Diesel Using Chemically Impregnated Coconut Coir
Waste. Int. J. Environ. Sci. Technol. 2015, 12, 2847−2856.
(51) Chen, M.; Ding, Y.; Liu, Y.; Wang, N.; Yang, B.; Ma, L.
Adsorptive Desulfurization of Thiophene from the Model Fuels onto
Graphite Oxide/Metal-Organic Framework Composites. Pet. Sci.
Technol. 2018, 36, 141−147.
(52) Ofomaja, A. E. Intraparticle Diffusion Process for Lead(II)
Biosorption onto Mansonia Wood Sawdust. Bioresour. Technol. 2010,
101, 5868−5876.
(53) Jaycock, M. J.; Parfitt, G. D. Chemistry of Interfaces; Ellis
Horwood Ltd.: Onichester, 1981; p 1981.
(54) Murzin, D. S. Chemical Kinetics, 1st ed.; Elsevier: Amsterdam,
2005.
(55) Samaniego, M. L.; De Luna, M. D. G.; Ong, D. C.; Wan, M.-
W.; Lu, M.-C. Isotherm and Thermodynamic Studies on the Removal
of Sulfur from Diesel Fuel by Mixing-Assisted Oxidative-Adsorptive
Desulfurization Technology. Energy Fuels 2019, 33, 1098−1105.
(56) Ismaiel, A. A.; Aroua, M. K.; Yusoff, R. Palm Shell Activated
Carbon Impregnated with Task-Specific Ionic-Liquids as a Novel
Adsorbent for the Removal of Mercury from Contaminated Water.
Chem. Eng. J. 2013, 225, 306−314.
(57) Doğan, M.; Alkan, M.; Onganer, Y. Adsorption of Methylene
Blue From Aqueous Solution Onto Perlite. Water, Air, Soil Pollut.
2000, 120, 229−248.
(58) McKay, G.; Otterburn, M. S.; Aga, J. A. Fuller’s Earth and Fired
Clay as Adsorbents for DyestuffsEquilibrium and Rate Studies.
Water, Air, Soil Pollut. 1985, 24, 307−322.

2132 https://dx.doi.org/10.1021/acs.jced.9b01204
J. Chem. Eng. Data 2020, 65, 2120−2132

You might also like