Accepted Manuscript Preparation of magnetic hierarchical porous carbon spheres with graphitic features for high methyl orange adsorption capacity Adisak Siyasukh, Yothin Chimupala, Nattaporn Tonanon PII:
S0008-6223(18)30344-0
DOI:
10.1016/j.carbon.2018.03.093
Reference:
CARBON 13036
To appear in:
Carbon
Received Date: 22 January 2018 Revised Date:
30 March 2018
Accepted Date: 31 March 2018
Please cite this article as: A. Siyasukh, Y. Chimupala, N. Tonanon, Preparation of magnetic hierarchical porous carbon spheres with graphitic features for high methyl orange adsorption capacity, Carbon (2018), doi: 10.1016/j.carbon.2018.03.093. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. 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.
ACCEPTED MANUSCRIPT
Preparation of magnetic hierarchical porous carbon spheres with graphitic features for high methyl orange adsorption capacity
1
RI PT
Adisak Siyasukh*,1, Yothin Chimupala1, Nattaporn Tonanon2
Department of Industrial Chemistry, Faculty of Science, Chiang Mai University, Chiang Mai
2
SC
50200, Thailand
Department of Chemical Engineering, Faculty of Engineering, Chulalongkorn University,
*
AC C
EP
TE D
M AN U
Bangkok 10330, Thailand
Corresponding author. Tel: 66-53-943405. E-mail:
[email protected] (Adisak Siyasukh)
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
Graphical abstract
ACCEPTED MANUSCRIPT -2-
Abstract
2
This report describes the successful use of a template-free method in a microemulsion system
3
for the preparation of magnetic hierarchical porous carbon (MHPC) spheres, which possess
4
high adsorption capability for methyl orange (MO) in aqueous solution, easy separability
5
from water by magnetic force after the MO removal, and reusability that allowed MO to be
6
adsorbed several times. MHPC spheres are produced by water-in-oil emulsification coupled
7
with sol-gel polymerization of resorcinol containing Fe(NO3)3 and formaldehyde to make
8
spherical carbon precursors with a macroporous framework inside, followed by carbonization
9
with N2 or CO2 activation to transform them into MHPC spheres. Carbonization at
10
temperatures higher than 800 °C resulted in not only graphitic feature formation, but also
11
both micro- and mesopores in the macroporous framework. Moreover, CO2 activation at 900
12
°C both remarkably enhanced the graphitic features and increased the mesopore volume to
13
2.16 cm3/g. The magnetic property could be developed by both carbonization and CO2
14
activation at 900 °C. MHPC spheres obtained from CO2 activation developed tremendous
15
adsorption capacities, as high as 1522.6 mg/g under neutral conditions, which were suitable
16
to be reused for MO removal for at least four consecutive cycles without efficiency loss.
EP
TE D
M AN U
SC
RI PT
1
AC C
17
18
Keywords: Hierarchical porous carbon; graphitic carbon; magnetic porous carbon; methyl
19
orange removal
20
ACCEPTED MANUSCRIPT -3-
1. Introduction
2
Water pollution caused by expansion of urbanized areas, industrial sectors, and agricultural
3
activities is a concerning environmental problem, which currently threatens the quality of
4
human life. Dyes employed in various industries contribute to the threats posed to the
5
environment and human health by industrial waste water, especially the azo dyes, which are
6
toxic, mutagenic, and even carcinogenic owing to the presence of amines [1-3]. Methyl
7
orange (MO), which is widely used in the textiles, printing, the food and pharmaceutical
8
industries, and research laboratories, is among the water-soluble azo dyes [1,4]. Adsorption,
9
an effective technology for dye removal, is most often used because of its simple design, low
10
cost, ease of operation, and reusable features [5,6]. Hence, much research has focused on
11
creating new approaches for preparing adsorbents with improved properties to enhance
12
adsorption efficiency. Activated carbons are popularly used as an adsorbent for MO removal
13
from wastewater [7,8]. Nonetheless, there are some disadvantages to using activated carbons.
14
For instance, the majority of pore sizes are in the range of micropores; these possess large
15
specific surface areas but provide low accessibility for the adsorption of large dye molecules,
16
e.g. MO, resulting in low adsorption performance [9-12]. Also, as activated carbons are
17
usually used in the form of powders, they are difficult to separate from wastewater for reuse.
18
To overcome such problems, hierarchical porous carbon-based adsorbents responsive to
19
magnetic force are promising materials. Their distinctive magnetic properties provide an easy
20
method for removing and recycling the adsorbents using an external magnet [13-15]. The
21
hierarchical porous structure is composed of interconnected macroporous frameworks (pore
22
size > 50 nm), whose walls contain micropores (pore size < 2 nm) and/or mesopores (2 ≤
AC C
EP
TE D
M AN U
SC
RI PT
1
ACCEPTED MANUSCRIPT -4-
pore size ≤ 50 nm). Mass transport can be enhanced via the macroporous frameworks, and
2
large specific surface areas are obtained from the existing micro- and mesopores [16-23].
3
This work aims to introduce a convenient method for the preparation of magnetic
4
hierarchical porous carbon (MHPC) spheres, which possess high capability for MO
5
adsorption from aqueous solution, easy separability from water by magnetic force after the
6
MO removal, and reusability to adsorb MO several times.
7
In the past few years, many research groups have created various macroscopic shapes of
8
MHPCs, for instance, preparing MHPC monoliths derived from synthesized polymer rods by
9
carbonization [24,25] or hydrothermal treatment [26], fabricating MHPCs with irregular
10
structures by thermal treatment of commercial activated carbons impregnated with iron
11
compounds [27,28], producing MHPC spheres (or beads) by carbonization of industrial
12
resins containing iron sources [29,30] or by solvothermal methods [31]. Spheres or beads
13
have advantages compared with other shapes, especially for use in adsorption columns,
14
because they are easy to handle on a large scale and have high resistance to attrition [32, 33].
15
Although some of the referenced articles describe interesting methods for making MHPC
16
spheres from commercial macroporous resins [29,30], a disadvantage of these methods is the
17
inability to tailor the macroporous morphology inside the spheres because of the inherent
18
structure of the precursors. To the best of our knowledge, the fabrication of MHPC spheres
19
without using any template materials or without resorting to prefabricated spherical
20
precursors has never been reported [33-38].
21
22
AC C
EP
TE D
M AN U
SC
RI PT
1
ACCEPTED MANUSCRIPT -5-
The main aim of this work is to introduce a convenient template-free method for the
2
preparation of MHPC spheres. The spheres were fabricated by water-in-oil emulsification
3
coupled with sol-gel polymerization of resorcinol containing ferric (III) nitrate (Fe(NO3)3)
4
and formaldehyde to make spherical carbon precursors with a macroporous framework
5
inside. Whereas the emulsification system provides the spherical shape of the precursors,
6
Fe(NO3)3 generates interconnected macroporous frameworks within the spheres. The
7
prepared precursors were further converted to carbon spheres by carbonization with nitrogen
8
or activation with carbon dioxide (CO2 activation), by which either micro- or mesopores
9
could be generated within the macroporous frameworks. The distinctive properties of the
10
prepared MHPC spheres – very high mesopore volumes (2.16 cm3/g) on the macroporous
11
framework, the graphitic features, and magnetic responsivity – were obtained by CO2
12
activation at 900 °C.
13
In addition, to demonstrate that the prepared MHPC spheres can be utilized as an effective
14
adsorbent for MO removal, we also investigated their adsorption performance. The
15
experiments were conducted at neutral pH only, to avoid any requirement to readjust the pH
16
to neutral after MO removal. The prepared MHPC spheres obtained by CO2 activation had
17
tremendous adsorption capacity (1522.6 mg/g), higher than that of many carbon-based
18
adsorbents recently reported elsewhere [39–44]. The results also highlighted the role of
19
mesopores within hierarchical porous structures in enhancing the effective performance of
20
MO removal. Furthermore, the prepared MHPC spheres were reusable several times. This
21
reusability is a significant feature for practical applications.
22
23
AC C
EP
TE D
M AN U
SC
RI PT
1
ACCEPTED MANUSCRIPT -6-
2. Experimental
2
Research-grade resorcinol (C6H6O2, Sigma-Aldrich, ≥ 99%), formaldehyde (CH2O, Loba
3
Chemie, 37%), ferric (III) nitrate nonahydrate (Fe(NO3)3.9H2O, AnalaR BDH, ≥ 99%),
4
cyclohexane (C6H12, RCI Labscan, 99.5%), SPAN80 (nonionic surfactant) (C24H44O6,
5
Sigma-Aldrich), acetone (C3H6O, RCI Labscan, 99.5%), MO (C14H14N3NaO3S, Sigma-
6
Aldrich, dye content 85%), and ethanol (RCI Labscan, 96%) were used.
7
2.1 Preparation of hierarchical porous carbon monoliths and carbon spheres
8
First, resorcinol was dissolved in deionized water with a constant mass ratio of 0.41. When
9
dissolution was complete, 0.1 M of Fe(NO3)3 solution was added into the resorcinol solution
10
with various mass ratios of Fe(NO3)3 to obtain the total masses of resorcinol and the
11
formaldehyde (Fe/RF-ratio) shown in Table 1. The mixture was continuously stirred at
12
ambient temperature until its color turned from light yellow to purple. Next, formaldehyde
13
was added to the mixture with a constant mole ratio of formaldehyde to resorcinol of 2 to
14
initiate sol-gel polymerization, by which the mixture was transformed to resorcinol-
15
formaldehyde (RF) gels. Carbon precursors were prepared for both monolithic and spherical
16
shapes to produce the hierarchical porous carbon monoliths and carbon spheres, respectively.
17
To prepare the carbon monoliths, the mixture was poured into a cylindrical glass mold until
18
the sol-gel reaction was completed, followed by removal from the mold and further drying in
19
a hot-air oven at 75 °C until a constant weight was reached. All monolithic gels were
20
carbonized at 900 °C for 2 h with a heating rate of 10 °C/min under ambient nitrogen
21
conditions. These monolithic carbons were denoted as CF0-900, CF1-900, CF2-900, and
22
CF3-900, as listed in Table 1. Note that these monolithic carbons were synthesized only to
AC C
EP
TE D
M AN U
SC
RI PT
1
ACCEPTED MANUSCRIPT -7-
1
investigate how Fe(NO3)3 influenced the macroporous morphology and size (see discussion
2
in Section 3.1).
RI PT
3
4
5
SC
6
M AN U
7
8
9
11
TE D
10
To prepare the carbon spheres (Fig. 1), the mixture (water phase) was poured into
13
cyclohexane (oil phase) containing SPAN80 (a surfactant) before losing its flow property,
14
resulting in a water-in-oil emulsion system. The volume ratio of RF solution and cyclohexane
15
was fixed at 0.1. The mixture was continuously agitated at 300 rpm until it became a reddish-
16
brown suspension, which was dispersed throughout the cyclohexane solution. Then, it was
17
filtered out and immersed in acetone three times, for 6 h each time. Finally, the carbon
18
precursors with spherical shape were obtained by drying in a hot-air oven at 75 °C until the
19
weight was constant. The precursors were converted to carbon spheres by carbonization
20
under ambient nitrogen conditions. The carbonization was conducted in a tube furnace under
21
nitrogen flow at 50 cm3/min at a heating rate of 10 °C/min until the desired temperature was
AC C
EP
12
ACCEPTED MANUSCRIPT -8-
reached; this temperature was then maintained for 2 h. The precursors without acetone
2
treatment were also carbonized at 900 °C for comparison. CO2 activation was carried out in a
3
similar manner to the carbonization procedure, except for changing the ambient gas from
4
nitrogen to carbon dioxide. The carbon sphere samples were denoted as CF3-500, CF3-600,
5
CF3-700, CF3-800, CF3-850, CF3-900, and ACF3-900, as listed in Table 1.
RI PT
1
SC
6
7
M AN U
8
9
10
12
TE D
11
2.2 Methyl orange adsorption experiment
14
CF3-700, CF3-900, and ACF3-900 were selected for the MO adsorption experiments. Batch
15
adsorption was conducted by adding spherical carbons into the MO solution with carbon
16
dosages and initial concentrations varying from 0.05 to 0.4 g/L and 1 to 600 mg/L,
17
respectively. As this work mainly focused on using the prepared MHPC spheres for
18
removing contaminated MO from aqueous solution under neutral pH only, examination of
19
the effect of pH on adsorption performance was beyond the scope of this work. It should be
20
noted, however, that under the range of initial concentrations of MO used in the experiment
21
(1 to 600 mg/L), the pH of the MO solution remained constant at 7.1 ± 0.2.
AC C
EP
13
ACCEPTED MANUSCRIPT -9-
All samples were shaken at 200 rpm at 30 °C for 48 h. MO solution was then withdrawn to
2
analyze the equilibrium concentration using an ultraviolet-visible light (UV-vis)
3
spectrophotometer at a wavelength of 463 nm. The amount of adsorbate at equilibrium ( qe )
4
was determined by Eq. 1:
5
qe =
6
A removal efficiency or % removal of MO was calculated using Eq. 2:
% r emoval =
( C0 − Ce ) ×100 C0
(2)
SC
(1)
W
M AN U
7
( C 0 − C e )V
RI PT
1
8
where C0 and Ce are the initial and equilibrium concentrations of MO (mg/L), and V and W
9
are the volume of the MO solution (L) and the weight of spherical carbons (mg), respectively.
11
The kinetics study was performed by shaking the MO solution containing the carbon spheres
12
with a carbon dosage of 0.4 g/L. The initial concentration of MO, shaking speed, and
13
temperature were 20 mg/L, 200 rpm, and 30 °C, respectively, for all experiments. MO
14
solution samples were taken at pre-set time intervals and their concentrations were
15
determined. The uptake of MO at time qt (mg/g) was calculated by Eq.3:
16
qt =
17
where Ct is the concentration of MO at any time.
18
In the thermodynamics study, the influence of different temperatures (30, 40, 60, and 80 °C)
19
was investigated. The experiment was conducted by adding spherical carbons into MO
AC C
EP
TE D
10
( C 0 − C t )V W
(3)
ACCEPTED MANUSCRIPT -10-
solution with carbon dosage and initial concentration of 0.4 g/L and 20 mg/L, respectively.
2
All samples were shaken at 200 rpm at the desired temperature for 48 h. Then MO solution
3
was withdrawn to analyze the equilibrium concentration.
4
2.2.1 Equilibrium isotherm models
5
Three equilibrium adsorption isotherm models, Langmuir, Freundlich, and Dubinin-
6
Radushkevich (D-R), were compared to elucidate the adsorption mechanism underlying the
7
experimental equilibrium adsorption data. The Langmuir model [45] is given in Eq. 4:
8
qe =
9
where Q0 and b are the monolayer adsorption capacity (mg/g) and the Langmuir constant
SC
RI PT
1
M AN U
Q0bCe (1 + bCe )
(4)
(L/mg), respectively. The Freundlich model [46] is represented in Eq. 5:
11
qe = K f Ce1 n
12
where Kf (mg/g)(l/mg)1/n and n are Freundlich adsorption constants related to the adsorption
13
capacity and the intensity of adsorption, respectively. The D-R isotherm model [47] is
14
expressed mathematically by Eqs. 6 and 7:
15
q e = q DR exp ( − βε 2 )
16
ε = RT ln 1 +
17
where qDR, β, ε, R, and T are the maximum adsorption capacity, a constant related to the
18
adsorption energy, the Polanyi potential, the gas constant (J/mol K), and temperature (K),
AC C
EP
TE D
10
1 Ce
(5)
(6)
(7)
ACCEPTED MANUSCRIPT -11-
1
respectively. β is useful for obtaining the mean adsorption energy E (kJ/mol), which can be
2
expressed by Eq. 8:
3
E=
4
The parameters of the three above-mentioned isotherms were determined by nonlinear
5
regression analysis, using MS Excel [48]. The correlation coefficient (R2) was calculated as a
6
measure of the compatibility of the experimental data and the adsorption isotherm models,
7
using Eq. 9:
R
SC
RI PT
(8)
∑(q = 1− ∑(q
e e
9
− qe,calc. ) − qe,avg . )
M AN U
8
2
1 2β
2
(9)
2
where qe, qe,calc, and qe,avg are the experimental data, the calculated, and the average values of the amount of MO adsorbed at equilibrium.
11
2.2.2 Kinetics adsorption models
12
Pseudo-first-order (PFO) and pseudo-second-order (PSO) kinetic models were employed to
13
test the consistency of the kinetic data. R2 was calculated using Eq. (9) to determine the
14
compatibility of the experimental data and the adsorption isotherm models.
15
The PFO model [49,50] is generally expressed as Eq. 10:
16
qt = qe1 (1 − exp ( −k1t ) )
17
The PSO introduced by Ho and Mckay [51] can be expressed as Eq. 11:
18
qt =
AC C
EP
TE D
10
k 2 qe22t (1 + k2 qe 2t )
(10)
(11)
ACCEPTED MANUSCRIPT -12-
where k2 is the PSO rate constant and qe2 is the equilibrium adsorption amount determined
2
from the PSO model. At the initial condition, the initial rate of adsorption of the PSO can be
3
expressed by Eq. 12:
4
h2 = k2qe22
RI PT
1
(12)
where h2 is the initial rate constant (mg/g·min).
6
2.2.3 Thermodynamic adsorption model
7
To evaluate the thermodynamic feasibility of the adsorption process, three basic
8
thermodynamic parameters, standard enthalpy (∆H0), standard entropy (∆S0), and standard
9
free energy (∆G0), were calculated using Eqs. 13 and 14 [52]:
M AN U
SC
5
11
∆G 0 = ∆H 0 − T ∆S 0
12
where qe/Ce is the adsorption affinity.
13
2.3 Characterizations
14
The micro- and nanostructures were imaged using a field emission scanning electron
15
microscope (FE-SEM, JSM-6335F, JEOL) equipped with an energy-dispersive X-ray
16
spectroscopy (EDS) detector (Inca, Oxford), operated at 200 kV. The interconnected
17
macroporous size distribution and volume were measured by mercury intrusion (Pore-Sizer-
18
9320, Micromeritics). Nitrogen adsorption-desorption isotherms at -196 °C were carried out
19
by nitrogen adsorption apparatus (Autosorb-1-MP, QuantaChrome). The specific surface area
(13)
(14)
AC C
EP
TE D
10
q ∆H 0 1 ∆S 0 log e = − Ce 2.303R 2.303R T
ACCEPTED MANUSCRIPT -13-
(SBET) was determined by the Brunauer-Emmett-Teller (BET) method. The mesopore width
2
distribution and mesopore volume (Vmeso) were calculated from the desorption branch of the
3
corresponding isotherm using the Barrett-Joyner-Halenda (BJH) method. The Dubinin-
4
Radushkevich (D-R) model was applied to calculate the micropore volume (Vmic). The
5
particle size and distribution were measured by laser particle size analyzer (Mastersizer S,
6
Malvern). Powder X-ray diffraction (PXRD) was performed using an X-ray diffractometer
7
(D8 Advance, Bruker). The diffractometer employs a Cu Kα tube with λ = 1.545 Å as the X-
8
ray source. Raman results were collected at 25 °C and conducted on a Raman spectrometer
9
(Jobin-Yvon T64000, HORIBA) with an excitation wavelength of 532 nm. Images were
10
captured by a transmission electron microscope (TEM, JEM-2010, JOEL). The lattice
11
spacing of the graphitic samples was measured using the ImageJ Software. The magnetic
12
property was tested using a vibrating sample magnetometer (VSM) (Model-73098, Lake
13
Shore). The concentrations of MO were measured with a UV-vis spectrophotometer
14
(SPECORD® 50 PLUS, Analytik Jena).
15
3. Results and discussion
16
3.1 Effect of ferric (III) nitrate on the macroporous morphology
17
The gelation step in the preparation process of CF0-900 (without Fe(NO3)3) took almost 2
18
days at room temperature, whereas the precursors of CF1-900, CF2-900, and CF3-900 (with
19
Fe/RF-ratios of 0.003, 0.010, and 0.023, respectively), require only 45, 25, and 5 min,
20
respectively. This suggests that Fe(NO3)3 in resorcinol solution at the first period can
21
strongly catalyze sol-gel polymerization. However, the Fe/RF ratio should not be greater than
22
0.03, because the sol-gel reaction would be very severe, resulting in sudden complete gel
23
formation that would cause practical difficulties in the preparation process.
AC C
EP
TE D
M AN U
SC
RI PT
1
ACCEPTED MANUSCRIPT -14-
Prior to the preparation of hierarchical porous carbon spheres, in order to clearly understand
2
the effects of Fe(NO3)3 on the microporous morphology, the samples (CF0-900, CF1-900,
3
CF2-900, and CF3-900) were formed as monoliths instead of spherical powders to
4
investigate how Fe(NO3)3 influences the macroporous morphology and size by the mercury
5
intrusion method. This was necessary as the technique has a limitation in that fine particles
6
cannot be measured, because pressure exerts the mercury to flow through spaces between
7
compacted particles before penetration into the macropores [37].
SC
M AN U
8
9
10
15
16
EP
14
AC C
13
TE D
11
12
RI PT
1
17
The macroporous textures of CF0-900, CF1-900, CF2-900, and CF3-900 are shown in Fig. 2.
18
(a). It can be seen that CF1-900, CF2-900, and CF3-900 possess macroporous frameworks,
19
which are interconnected voids of dumbbell-like microparticle networks, while the
20
macroporous morphology of CF0-900 (without the presence of Fe(NO3)3) exhibits
21
interconnected voids of spherical-like microparticles. As shown in Fig. 2. (b), CF1-900, CF2-
ACCEPTED MANUSCRIPT -15-
900, and CF3-900 had narrow macropore size distributions in the ranges of 9.5–32.1, 4.1–
2
15.8, and 4.5–10.9 µm with median pore sizes of 10.9, 8.6, and 6.6 µm, respectively.
3
However, CF0-900 had a wider distribution in a range of 1.5–38 µm with a median pore size
4
of 14.5 µm. It is worth noting that the existence of Fe(NO3)3 plays an important part as a
5
macropore size controller, and thus, increasing Fe(NO3)3 can decrease the macropore size.
RI PT
1
AC C
EP
TE D
M AN U
SC
6
ACCEPTED MANUSCRIPT -16-
1
3.2 Preparation of carbon spheres containing hierarchical porous structure
2
3.2.1 Template-free method for preparation of RF spherical gels and carbon spheres
RI PT
3
4
SC
5
6
M AN U
7
8
9
13
14
15
16
17
18
EP
12
AC C
11
TE D
10
ACCEPTED MANUSCRIPT -17-
As described in section 3.1, carbon monoliths with hierarchical porous structures can be
2
prepared by sol-gel polymerization of RF solution containing Fe(NO3)3, followed by
3
carbonization. This synthesis approach can be adapted to fabricate carbon spheres containing
4
hierarchical porous structures, if the sol-gel-polymerized area is confined within droplets
5
(water phase) made from a water-in-oil emulsion system until the completion of the reaction.
6
Fig. 3 (a) shows the microsphere formation of RF gel droplets (water phase), which are
7
enveloped by emulsifier (SPAN80) to minimize interfacial tension at the boundary of the two
8
immiscible phases. SPAN80 forms microscopic droplets dispersed in cyclohexane (oil phase)
9
and prevents coalescence of the droplets by generating a repulsive force between them. These
10
droplets are thus stably dispersed throughout the oil phase by agitation. As the reaction
11
proceeds, the droplets transform into gel spheres, and the RF spherical gels can be obtained
12
as shown in Fig. 3 (b). After carbonization to convert the gels into carbon spheres, both the
13
spherical morphology and 3D interconnected macroporous structure remained unchanged, as
14
shown in Fig. 3 (c). This confirms that the microemulsion is a key important step to prepare
15
carbon spheres with hierarchical porous structures. To tailor the spherical size, dosages of
16
SPAN80 in cyclohexane were adjusted in the range of 0.2 to 0.4 %Vol. Fig. 3 (d) shows the
17
particle size distributions of RF spherical gels and carbon spheres prepared from different
18
%Vol of SPAN80. The mean diameter of the spherical RF gels became lesser when the %Vol
19
was higher. The mean sizes were 302, 222, and 141 µm for %Vol of 0.2, 0.3, and 0.4,
20
respectively. A possible interpretation of this result is that an increase of SPAN80 reduces
21
the interfacial tension between the two phases, allowing the droplets to adjust their surface
22
area to be as small as possible, and the resulting small surface area causes a decrease in the
23
size of the spherical drops [53]. All the obtained spherical carbons shrunk after carbonization
AC C
EP
TE D
M AN U
SC
RI PT
1
ACCEPTED MANUSCRIPT -18-
compared with their initial sizes. The mean sizes of the carbon spheres are 259, 141, and 103
2
µm for %Vol of 0.2, 0.3, and 0.4, respectively.
3
3.2.2 Effect of acetone treatment
RI PT
1
4
5
SC
6
M AN U
7
8
9
13
14
15
16
17
18
EP
12
AC C
11
TE D
10
ACCEPTED MANUSCRIPT -19-
To investigate the effects of acetone treatment, the RF spherical gels prepared from sol-gel
2
emulsification with 0.4 %Vol of SPAN80 were further suddenly immersed in pure acetone
3
before carbonization at 900 °C (CF3-900). For comparison, the RF spherical gels without
4
immersion in acetone were carbonized under the same conditions. It was observed that the
5
CF3-900 without acetone treatment had a thin-film layer encapsulating its exterior, as shown
6
in Fig. 4 (a), but this disappeared with acetone treatment, as shown in Fig. 4 (b). The
7
formation of the thin film covering is similar to that reported in Yamamoto's work [54], in
8
which mesoporous carbon spheres were synthesized by an emulsification method using
9
SPAN80 as the emulsifier. It may be presumed that copious SPAN80 gathers at the oil/water
10
interface, such that the micelles can construct the thin layer covering the exterior of the
11
droplets. As the SPAN80 was dissolved in acetone immediately after completion of the sol-
12
gel polymerization within the droplets, leaching with acetone could easily break the micelle
13
thin film. This result confirms that acetone treatment is an important step to reveal the true
14
structure of hierarchical porous carbon spheres. Fig. 4 (c) shows a comparison of average
15
diameters and particle size distributions of CF3-900 with and without acetone treatment.
16
These results indicate that acetone treatment does not affect the particle size distribution. The
17
mean diameters were 103 and 102 µm for the samples with and without the acetone
18
treatment, respectively.
SC
M AN U
TE D
EP
AC C
19
RI PT
1
ACCEPTED MANUSCRIPT -20-
1
3.2.3 Effect of carbonization and CO2 activation on the porous properties of spherical
2
carbons
RI PT
3
4
5
SC
6
M AN U
7
8
9
13
14
15
16
17
18
19
EP
12
AC C
11
TE D
10
ACCEPTED MANUSCRIPT -21-
Pore development of the micro/mesoporous structures in the macroporous walls was
2
determined using the N2 adsorption-desorption isotherms, as shown in Fig. 5 (a) and (b). At a
3
temperature for carbonization of 500 °C (CF3-500), the isotherm started to uptake a small
4
amount of nitrogen, indicating that pore development had begun to occur on the wall of the
5
macropores. When the temperature further rose to 600 and 700 °C (CF3-600 and CF3-700,
6
respectively), the isotherms showed a higher uptake of nitrogen adsorption and exhibited the
7
characteristic Type-I shape according to IUPAC classification. This implies that a
8
microporous structure will only form in the macroporous wall if the carbonization
9
temperature is higher than 500 °C. The Vmic values of CF3-600 and CF3-700 were 0.23 and
10
0.24 cm3/g, respectively, as presented in Table 2. This development of micropores results in
11
an increase in SBET from 569 to 584 m2/g for CF3-600 and CF3-700, respectively. Focusing
12
on the isotherms of CF3-800, CF3-850, and CF3-900 in Fig. 5 (b), they can be seen to
13
display a combination of Type-I and Type-IV shapes, including the H3-hysteresis loop that
14
indicates the existence of both micropores and slit-shaped mesopores within the walls of the
15
macropores. This result implies that the mesopores can be generated at carbonization
16
temperatures higher than 800 °C. However, without the presence of Fe(NO3)3, the N2
17
isotherm of CF0-900, as shown in Fig. 5 (b), provides a Type-I shape even if the
18
carbonization temperature is as high as 900 °C. This is evidence of the crucial role of
19
Fe(NO3)3 in generating the mesopores.
20
As shown in Table 2, the Vmeso increased as the temperature of carbonization increased,
21
reaching values of 0.19, 0.21, and 0.21 cm3/g for CF3-800, CF3-850, and CF3-900,
22
respectively. On the other hand, the Vmic decreased when the temperature of carbonization
23
rose, to 0.22, 0.19, and 0.16 cm3/g for CF3-800, CF3-850, and CF3-900, respectively. This
AC C
EP
TE D
M AN U
SC
RI PT
1
ACCEPTED MANUSCRIPT -22-
indicates that the presence of Fe(NO3)3 could destroy the micropore structure, constructing
2
the expanded pore size in a range of mesopores. Likewise, SBET could be reduced by
3
increasing the carbonization temperature, to 553, 467, and 400 m2/g for CF3-800, CF3-850,
4
and CF3-900, respectively, because of micropore decrease.
5
SC
6
M AN U
7
8
9
10
15
16
EP
14
AC C
13
TE D
11
12
RI PT
1
17
The SEM images, as illustrated in Fig. 6 (a) and (b), reveal the spherical morphology of CF3-
18
900 and CF3-700. These results confirm the existence of mesoporosity within the
19
macroporous framework of CF3-900, and the lack of such mesoporosity in CF3-700. This
20
evidence is consistent with the N2 isotherm result as discussed above.
ACCEPTED MANUSCRIPT -23-
1
2
RI PT
3
4
SC
5
6
ACF3-900 was prepared similarly to CF3-900 except for the use of CO2 activation for the
8
former instead of carbonization. The particle size distribution of ACF3-900, as shown in Fig.
9
7 (a), was roughly between 7 and 120 µm, with a median size of 35 µm. One can see that the
10
size distribution and the median size of ACF3-900 were reduced compared with those of
11
CF3-900. The evidence of the mesoporosity improvement by CO2 activation can be seen in
12
the N2 adsorption-desorption isotherm shown in Fig. 7 (b). This isotherm shows the high-
13
volume uptake of liquid nitrogen and exhibits hybridization of type-I and type-IV with a H3-
14
hysteresis loop, which indicates the presence of both micropores and a large amount of slit-
15
shaped mesopores. The mesopore width distribution of the ACF3-900, as shown in Fig. 7
16
(b)-inset, suggests that the mesopore widths were between 2.00 to 4.87 nm, with a peak at
17
2.54 nm. The porosities, as listed in Table 2, indicate that ACF3-900 possessed a very high
18
mesopore volume (Vmeso = 2.16 cm3/g) compared with CF3-900 (Vmeso = 0.21 cm3/g).
19
Meanwhile, ACF3-900 had Vmic and SBET of 0.2411 cm3/g and 603 m2/g, respectively. Fig. 7
20
(c) presents the SEM images of ACF3-900, which depict the spherical morphology (upper
21
image) and high mesoporosity located in the macroporous framework (middle and lower
AC C
EP
TE D
M AN U
7
ACCEPTED MANUSCRIPT -24-
1
images). It is clear that the mesopores were significantly enhanced by CO2 activation, which
2
is consistent with the results of the N2 adsorption-desorption isotherm.
RI PT
3
4
5
SC
6
M AN U
7
8
9
13
14
15
16
17
18
19
EP
12
AC C
11
TE D
10
ACCEPTED MANUSCRIPT -25-
1
3.3 Graphitic features and magnetic properties of the prepared carbon spheres
2
RI PT
3
4
SC
5
6
M AN U
7
8
9
13
14
15
16
EP
12
AC C
11
TE D
10
17
18
The carbon spheres were characterized using Raman spectroscopy, owing to the high
19
sensitivity of this technique for the characterization of graphitic materials. The results are
ACCEPTED MANUSCRIPT -26-
displayed in Fig. 8 (a). A D-band at ∼1360 cm-1 and G-band at ∼1580 cm-1 of the Raman
2
spectra were observed for all samples. The D-band could be attributed to the local defects
3
and disorders at the edges of the graphitic materials [55]. The peak intensities at the D-band
4
were stronger when the samples were treated at higher temperatures, from 500 to 900 °C,
5
indicating phase transformation from amorphous carbon to a graphitic structure. The high
6
intensities of the D-band in CF3-900 and ACF3-900 indicate that a high number of π bonds
7
(sp2 carbon atom hybridization) were replaced with C-O and/or C-C σ bonds (sp3), resulting
8
in structural defects from idiomatically perfect graphite/graphene [55,56]. The higher values
9
of the ID/IG ratios in CF3-900 and ACF3-900 compared with those of other samples indicate
10
the higher degree of defects in the graphitic structure [55]. The Raman-active mode located
11
at ∼2700 cm-1 (2D-band), which was observed only in CF3-900 and ACF3-900, supports the
12
presence of a graphitic phase that was indicated by the XRD results. The two additional
13
active modes of the graphitic structure at approximately 2400 cm-1 of the G* band and 3000
14
cm-1 of the D+D* band appeared clearly only for ACF3-900 [40]. This confirms that the CO2
15
activation process encourages the phase formation of graphitic structure in ACF3-900 more
16
than in the other samples, which were synthesized via the carbonization process. To further
17
confirm the influence of Fe(NO3)3 on structural formation in CF3-900, the Raman spectra of
18
CF3-900 and CF0-900 (controlled conditions of carbonization at 900 °C in the absence of
19
Fe(NO3)3) were also analyzed, showing that the condition without Fe(NO3)3 does not give the
20
high-intensity peak at the D-band or 2D-band. This can be attributed to the fact that Fe(NO3)3
21
promotes the replacement of π bonds (sp2) with σ bonds (sp3) in the graphitic structure.
22
The XRD patterns of the samples are shown in Fig. 8 (b). The XRD spectra of CF3-800,
23
CF3-900, and ACF3-900 showed a sharp peak at 26.57°, which could be assigned to the
AC C
EP
TE D
M AN U
SC
RI PT
1
ACCEPTED MANUSCRIPT -27-
graphite phase (JCPDS 01-089-8487), together with a broad peak at approximately 44°. The
2
calculated crystallite (grain) sizes of the graphite phases derived using XRD line broadening
3
at 26.57° together with Scherrer’s equation were 3.71 and 3.42 nm for CF3-800 and CF3-
4
900, respectively; the crystallite size of ACF3-900 could not be evaluated owing to the
5
overlap between the peak of graphite and the broad pattern of the amorphous carbon.
6
However, CF3-500, CF3-700, and CF0-900 all showed only the broad pattern of amorphous
7
carbon, with a weak diffraction peak of graphite at 26.41° for CF3-700 and CF0-900. This
8
indicates that the graphitic phase formation is supported by the presence of Fe(NO3)3 co-
9
operating with reaction temperatures higher than 800 °C in both CO2 activation and
M AN U
SC
RI PT
1
carbonization.
11
Fig. 8 (c) and (d) shows bright-field TEM images for CF3-900 and ACF3-900. The graphitic
12
samples were tangled with each other by bending and folding of multilayer graphitic sheets,
13
which provided spaces between the folded graphitic sheets to form the mesoporous
14
structures. This confirms the existence of mesopore volumes within both CF3-900 and
15
ACF3-900, as discussed in section 3.2.3. Also, good distribution of multilayer sheets led to
16
the increase of mesoporosities. The enlarged square regions given as insets in the Figure
17
show the multi-layer graphitic sheets with a lattice fringe spacing of 4.21 ± 0.05 Å and 4.33 ±
18
0.05 Å, respectively, consistent with the (001) spacing of expanded graphite structures [58].
19
The lattice spacings of both samples were greater than the general spacing (3.34 Å) of
20
multilayer graphene structures, owing to the structural expansion resulting from the
21
replacement of sp2 carbon atoms with sp3 carbon, as indicated by the Raman spectra.
22 23
AC C
EP
TE D
10
ACCEPTED MANUSCRIPT -28-
1 2 3
RI PT
4 5 6
SC
7 8
M AN U
9 10 11 12
16 17 18 19 20
EP
15
AC C
14
TE D
13
21 22
To confirm the Fe3O4 phase formation, therefore, pure Fe(NO3)3 powders were subjected to
23
the same synthesis conditions as CF3-900 and ACF3-900, resulting in products that were
ACCEPTED MANUSCRIPT -29-
labeled CFe-900 and AFe-900, respectively. Fig. 9 (a) and (b) show comparisons of the XRD
2
patterns of CF3-900 vs. CFe-900 and ACF3-900 vs. AFe-900, respectively, to elucidate the
3
structural change in Fe(NO3)3. The XRD patterns of both CFe-900 and AFe-900 showed the
4
peak pattern of Fe3O4, indicating that Fe(NO3)3 was easily transformed to the Fe3O4 structure
5
after both carbonization and CO2 activation at 900 °C. Moreover, the influence of CO2
6
seemed to promote the phase formation of Fe3O4 with higher crystallinity compared with N2,
7
owing to the higher oxidation firing of CO2. The characteristic peaks of Fe3O4 (JCPDS 00-
8
019-0629) also appeared, as an intense peak at 35.42° with five other weaker peaks at 30.09,
9
43.05, 53.40, 56.94, and 62.57° on ACF3-900, which confirms that ACF3-900 consists of a
10
mixed composition of graphite and Fe3O4. However, the Fe3O4 phase with very low
11
%loading of Fe (III), less than 0.6% weight in the starting mixture (Fe/RF-ratio of 0.023 g/g
12
as listed in Table 1), would normally be obstructed by the domain phase (graphite) and
13
background in the XRD pattern, resulting in the lower relative intensity of Fe3O4 in both
14
CF3-900 and ACF3-900 in the XRD patterns.
17 18 19 20
SC
M AN U
TE D
EP
16
AC C
15
RI PT
1
21 22
The SEM-EDS results, as listed in Table 3, revealed the presence of C, O, and Fe with
23
atomic% of 96.53, 3.12, and 0.35, respectively, on CF3-900; and 95.06, 4.47, and 0.47,
ACCEPTED MANUSCRIPT -30-
respectively, on ACF3-900. Furthermore, the well distributions of Fe3O4 particles in CF3-900
2
and ACF3-900 were measured using SEM-EDS mapping as shown in Fig. 9 (c) and (d),
3
respectively. The results illustrated that both samples contained a large number of carbon
4
atoms with equally good distributions of both iron and oxygen atoms. There were particles of
5
Fe3O4 dispersed randomly in the matrix of the carbon structure of both samples,
6
corresponding to the phase formation of Fe3O4 indicated by the XRD results.
RI PT
1
SC
7
M AN U
8
9
10
14
15
16
17
18
19
EP
13
AC C
12
TE D
11
ACCEPTED MANUSCRIPT -31-
Fig. 10 (a) shows the magnetization curves, in which the magnetization (emu/g) and
2
magnetic field (Oe) are plotted. The observed hysteresis loop corresponds to the
3
ferromagnetism property, which indicates the presence of Fe3O4 in both samples. The
4
saturation magnetization (Ms), coercive force (Hc), and chemical compositions as
5
characterized by EDS are presented in Table 3. The Ms of ACF3-900 was 5.8 emu/g with Fe
6
content of 2.10 (weight%), compared with 0.4 emu/g and 1.57 weight% of Fe for CF3-900.
7
This result suggests substantial improvement in the magnetic property of ACF3-900, in
8
agreement with the phase formation of Fe3O4 having higher crystallinity than that of CF3-
9
900, as previously discussed. It also confirms that our method was successful for fabricating
10
the MHPC spheres, whose inner structures contained the interconnected macroporous
11
framework and whose surface possessed both micro- and mesoporosities, especially in the
12
case of CF3-900 and ACF3-900. Moreover, graphitic features could be found in both
13
samples. Fig. 10 (b) and VDO 1 (video clip) illustrate the performance of ACF3-900 for MO
14
removal from water at an initial concentration of 20 mg/L; the suspended solution rapidly
15
turned transparent, showing an easy separation that confirms the good magnetic response of
16
the MHPC spheres to an outer magnet. This also reveals that the prepared MHPC spheres are
17
a promising adsorbent for MO removal from aqueous solution.
SC
M AN U
TE D
EP
AC C
18
RI PT
1
ACCEPTED MANUSCRIPT -32-
3.4 Application of the prepared carbon spheres for methyl orange adsorption
2
MHPC spheres with different porous characteristics were chosen to adsorb MO at neutral pH.
3
ACF3-900 and CF3-900, which possess Vmeso values of 2.16 and 0.21 cm3/g, respectively,
4
were used as representative MHPC spheres. For comparison, hierarchical porous carbon
5
spheres without mesopores (CF3-700) were also studied.
6
3.4.1 Effects of carbon dosages and initial concentrations on MO removal efficiency
7
The effects of carbon dosage on the removal percentage of MO was studied by varying the
8
amount of carbon in the MO solution from 0.05 to 0.4 g/L, with initial concentration,
9
temperature, stirring rate, and contact time kept constant at 20 mg/L, 30 °C, 200 rpm, and 48
10
h, respectively. The results, presented in Fig. 11 (a), show that CF3-700 (with an absence of
11
mesopores within its structure) had a poor removal efficiency of MO adsorption even when
12
the carbon dosage was as high as 0.4 g/L (2.51% removal), whereas the MHPC spheres
13
containing mesoporous structures (CF3-900 and ACF3-900) provided markedly better MO
14
removal performance. The results also show that the removal percentage of MO was
15
improved by increasing the carbon dosage (from 97.29 to 99.88% and 36.84 to 95.95% for
16
ACF3-900 and CF3-900, respectively), because the available surface area was enhanced,
17
consequently increasing the adsorption. MO adsorption reached its highest removal
18
percentage at a carbon dosage of 0.4 g/L for all MHPCs. Therefore, the remaining
19
experiments were carried out with 0.4 g/L carbon dosage.
20
The effects of initial concentration (varied from 10 to 300 mg/L) on the removal percentage
21
of MO are shown in Fig. 11 (b). Notably, the removal percentage decreased when initial
22
concentrations of MO were increased, from 99.91 to 77.46%, 97.22 to 12.50%, and 3.27 to
AC C
EP
TE D
M AN U
SC
RI PT
1
ACCEPTED MANUSCRIPT -33-
1.46% for ACF3-900, CF3-900, and CF3-700, respectively. The decrease in removal
2
percentage was attributed to there being insufficient adsorption sites on the adsorption
3
surface at the higher initial concentrations [59]. The slight decrease in removal efficiency for
4
ACF3-900 compared with the others, even at higher initial concentrations, was due to its
5
abundant surface adsorption sites, resulting from the tremendous mesoporosities within its
6
hierarchical porous structure.
RI PT
1
SC
7
M AN U
8
9
10
14
15
16
17
18
19
EP
13
AC C
12
TE D
11
ACCEPTED MANUSCRIPT -34-
1
3.4.2 Equilibrium isotherms, kinetics, and thermodynamics of MO adsorption
2
RI PT
3
4
SC
5
6
M AN U
7
8
9
13
14
15
16
17
18
EP
12
AC C
11
TE D
10
ACCEPTED MANUSCRIPT -35-
Adsorption equilibrium isotherms, kinetics, and thermodynamics were adopted in this study
2
to help understand the mechanism of MO adsorption onto MHPC spheres. Fig. 12 (a) – (c)
3
shows the equilibrium adsorption data for MO adsorption, fitted to the Langmuir, Freundlich,
4
and D-R isotherms by a nonlinear regression method at 30 °C and carbon dosage of 0.4 g/L.
5
The R2 values, as well as the related parameters of Langmuir (Q0 and b), Freundlich (Kf and
6
n), and D-R (qDR and E), are summarized in Table 4.
SC
7
RI PT
1
M AN U
8
9
10
TE D
11
To describe the adsorbate-adsorbent interaction mechanism, i.e., mono- or multilayer
13
adsorption, Langmuir and Freundlich isotherms were analyzed by fitting with the
14
experimental data. The equilibrium data were found to be in accordance with both isotherms,
15
as evidenced by the values of R2 being close to unity for all carbon spheres, with notably
16
slightly higher values for Freundlich. Thus, it is possible that adsorption onto heterogeneous
17
sites and multilayer mechanism are the main processes involved, instead of monolayer
18
adsorption onto homogeneous sites [60].
19
To consider the potential for adsorption on carbon surfaces, the monolayer adsorption
20
capacity (Q0) of the Langmuir isotherm should also be discussed. As shown in Table 4, the
21
values of Q0 were 1522.6, 77.1, and 16.2 mg/g for ACF3-900, CF3-900, and CF3-700,
AC C
EP
12
ACCEPTED MANUSCRIPT -36-
respectively. Consequently, ACF3-900 appears to have tremendous capacity for MO
2
adsorption, higher than that of many carbon-based adsorbents recently reported by others, as
3
listed in Table 7. However, while monolayer adsorption is assumed for Langmuir, it is also
4
reasonable to explain the adsorption capacity of MO in terms of the ratio Q0/SBET. As shown
5
in Table 4, the values of this ratio were 2.5 and 0.2 mg/m2 for ACF3-900 and CF3-900,
6
respectively, whereas CF3-700 had a smaller value of 0.3 x 10-1 mg/m2. Clearly, the
7
existence of mesoporosities within hierarchical porous structures can significantly enhance
8
the Q0/SBET of MO adsorption, especially for MO adsorption on ACF3-900. The excellent
9
adsorption capacity of ACF3-900 could possibly be explained by the presence of more
10
available adsorptive sites resulting from the very high mesopore volumes (Vmeso = 2.16
11
cm3/g), whose pore widths are between 2.00 to 4.87 nm, suitable for the molecular size of the
12
adsorbate. Previous reports have given the molecular size of MO as approximately 1.2 nm
13
[61, 62], and it may be 2.6 nm owing to ionic micelles forming in aqueous solution [63].
14
The D-R isotherm can give insight into the adsorption energy, E, which provides information
15
about the adsorption process, i.e., whether it involves physisorption, ion-exchange, or
16
chemisorption. Values of E lower than 8 kJ/mol correspond to physical adsorption, while
17
values between 8 and 16 kJ/mol suggest ion-exchange adsorption, and those higher than 16
18
kJ/mol indicate a chemisorption mechanism [64]. As shown in Table 4, the E values were 4.1
19
and 3.5 kJ/mol for ACF3-900 and CF3-700, respectively, indicating that the adsorption of
20
MO was physisorption, whereas the E value of 12.0 kJ/mol for CF3-900 indicated an ion-
21
exchange adsorption process.
22
23
AC C
EP
TE D
M AN U
SC
RI PT
1
ACCEPTED MANUSCRIPT -37-
1
2
RI PT
3
4
To further explore the thermodynamic properties, three basic properties, ∆H0, ∆S0, and ∆G0,
6
were determined and are presented in Table 5. The negative values of ∆H° obtained for all
7
samples show that the adsorption process was exothermic in nature and favorable at low
8
temperatures [65]. In addition, the value of ∆H° can give an indication of the type of
9
adsorption, i.e., physical or chemical. Values of ∆H° larger than -84 kJ/mol indicate physical
10
adsorption, while those between -80 and -420 kJ/mol are associated with chemisorption [65-
11
67]. The obtained ∆H0 values were -42.4, -72.9, and -27.8 kJ/mol for ACF3-900, CF3-900,
12
and CF3-700, respectively; these results indicate that the MO adsorption onto all carbon
13
spheres was physisorption. However, MO adsorption onto CF3-900 tended to be a relatively
14
strong adsorbate-adsorbent interaction, which is in agreement with the ion-exchange
15
mechanism previously determined from the D-R isotherm. Detailed studies of the role of
16
surface chemistry to understand the in-depth adsorption mechanism of the prepared carbon
17
spheres would be worth considering in further research; however, these are beyond the scope
18
of this work.
19
Furthermore, the negative values of ∆S0 for all carbon spheres indicate a decrease in degrees
20
of freedom of MO after adsorption, resulting from the restriction of molecular movement on
21
the surface of the adsorbent [68]. Finally, the negative values of ∆G0 indicate the spontaneity
22
and feasibility of adsorption, while a decreased value of ∆G0 illustrates more favorable
AC C
EP
TE D
M AN U
SC
5
ACCEPTED MANUSCRIPT -38-
adsorption [69,70]. These results confirm that MO adsorption onto ACF3-900 and CF3-900
2
is spontaneous at all temperatures, and is less favorable at high temperatures. On the
3
contrary, ∆G0 of MO adsorption onto CF3-700 had a positive value at all temperatures,
4
indicating unfavorable adsorption of MO. This explains why the MO removal efficiency by
5
CF3-700 was very low.
RI PT
1
SC
6
7
M AN U
8
9
10
Fig. 12 (d) – (f) shows the kinetic adsorption data for MO onto ACF3-900, CF3-900, and
12
CF3-700 at a temperature of 30 °C, carbon dosage of 0.4 g/L, and initial concentration of 20
13
mg/L. MO adsorption onto both ACF3-900 and CF3-900 reached near-equilibrium within
14
approximately 20 min, implying an excellent kinetic property for MO adsorption, whereas
15
CF3-700 required at least 200 min to reach equilibrium. It is clear that the presence of
16
mesoporosities within hierarchical porous structures can enhance the rate of MO adsorption.
17
PFO and PSO kinetic models were tested to understand the dynamic mechanism of the MO
18
adsorption rate. The experimental data were fitted to PFO and PSO kinetic models by a
19
nonlinear regression method. The R2 values and the related parameters of PFO (k1 and qe1)
20
and PSO (k2, qe2, and h) are summarized in Table 6. PSO fit better with the experimental data
21
compared with PFO for all samples, indicating that MO adsorption onto all carbon spheres
22
followed the pseudo-second-order kinetic mechanism [39]. The initial rate constants (h2), as
AC C
EP
TE D
11
ACCEPTED MANUSCRIPT -39-
listed in Table 6, were 36.8, 10.6, and 0.3 x 10-1 mg/g·min for ACF3-900, CF3-900, and
2
CF3-700, respectively. These results reveal that the existence of mesoporosities within
3
hierarchical porous structures has a crucial role in the enhancement of the adsorption kinetic
4
rate.
5
3.4.3 Reusability of MHPC spheres
RI PT
1
SC
6
M AN U
7
8
9
10
13
14
EP
12
TE D
11
Reusability is an important performance index for assessing the practical applicability of
16
MHPC spheres. In order to study the reusability of the adsorbents, recycling experiments
17
were conducted with each cycle. MHPC spheres were separated by a magnet after adsorption
18
and regenerated by desorption in pure ethanol at 30 °C with shaking at 200 rpm several times
19
until the solution became colorless. Then, the treated carbon spheres were used to re-adsorb
20
MO at the initial concentration and temperature of 20 mg/L and 30 °C, respectively, for each
AC C
15
ACCEPTED MANUSCRIPT -40-
cycle. As shown in Fig. 13, after four consecutive cycles, the MO removal efficiency of
2
ACF3-900 still remained over 90%. On the contrary, the removal percentage of CF3-900
3
drastically decreased after the first run compared with ACF3-900; this was probably owing to
4
more MO residuals covering the surface after regeneration, resulting from a strong adsorbate-
5
adsorbent interaction. Moreover, it is probable that the immense mesoporosities of ACF3-
6
900 could provide abundant adsorption sites for MO removal, thus having more-than-
7
adequate available adsorption sites after regeneration for several cycles, despite the
8
remaining MO residuals. These results suggest that ACF3-900 could be regarded as a high-
9
performance and reusable adsorbent for MO removal from aqueous solution. Moreover, both
10
ACF3-900 and CF3-900 can remain responsive to an external magnet despite being used for
11
four consecutive regeneration cycles.
15
16
17
18
EP
14
AC C
13
TE D
12
M AN U
SC
RI PT
1
19
20
Table 7 gives a comparison of the maximum monolayer adsorption capacities (Q0) of MO for
21
various carbon-based adsorbents, together with their magnetism and reusability. Clearly, the
ACCEPTED MANUSCRIPT -41-
MHPC spheres prepared in this work have tremendous adsorption capacity (1522.6 mg/g),
2
which is higher than that of many carbon-based adsorbents recently reported by others
3
[29,39-44]. This material also features good magnetism and reusability.
4
4. Conclusion
5
The successful preparation of MHPC spheres by a template-free method in a microemulsion
6
system is presented here. The prepared MHPC spheres possess high capability for adsorption
7
of MO from aqueous solution, easy separability from water by magnetic force after use, and
8
reusability to adsorb MO several times. The results can be summarized as follows.
9
(1) Water-in-oil emulsification coupled with sol-gel polymerization of resorcinol containing
10
Fe(NO3)3 and formaldehyde is a crucial factor to make spherical carbon precursors
11
containing a macroporous framework. The prepared precursors can be further converted to
12
MHPC spheres by carbonization or CO2 activation; in this way, either micro- or mesopores
13
can be generated within the macroporous frameworks.
14
(2) Another important step in the synthesis of MHPC spheres not covered by a thin film is
15
treatment with acetone immediately after completion of the sol-gel polymerization.
16
(3) Carbonization at temperatures above 800 °C can not only generate both micro- and
17
mesopores in the macroporous framework, but can also impart graphitic features, whereas
18
carbonization at lower temperatures can only generate micropores, not graphitic features.
19
(4) CO2 activation at 900 °C can result in both high mesopore volume (2.16 cm3/g) within the
20
macroporous framework and remarkable graphitic features. Good magnetic properties can
21
also be obtained by both carbonization and CO2 activation at 900 °C; however, CO2
22
activation can better enhance the magnetic responsivity.
AC C
EP
TE D
M AN U
SC
RI PT
1
ACCEPTED MANUSCRIPT -42-
(5) The MHPC spheres obtained by CO2 activation possess tremendous adsorption capacities,
2
as high as 1522.6 mg/g, for MO removal, and are easily separated from the suspension by
3
magnetic force. Moreover, the spheres are a high-performance and reusable adsorbent that
4
can be used for at least four consecutive cycles of regenerations.
5
Acknowledgement
6
This work is fully supported by CMU Junior Research Fellowship Program.
7
References
8
[1] Yagub MT, Sen TK, Afroze S, Ang HM. Dye and its removal from aqueous solution by
9
adsorption: A review. Adv Colloid Interfac 2014; 209: 172-184.
M AN U
SC
RI PT
1
[2] Wang Y, Zhao L, Peng H, Wu J, Liu Z, Guo X. Removal of anionic dyes from aqueous
11
solutions by cellulose-based adsorbents: equilibrium, kinetics, and thermodynamics. J Chem
12
Eng Data 2016; 61: 3266-3276.
13
[3] Mittal A, Mittal J, Malviya A, Kaur D, Gupta VK. Adsorption of hazardous dye crystal
14
violet from wastewater by waste materials. J Colloid Interf Sci 2010; 343: 463-473.
15
[4] Chung K-T, Cerniglia CE. Mutagenicity of azo dyes: Structure-activity relationships.
16
Mutat Res-Rev Genet 1992; 277: 201-220.
17
[5] Gupta VK, Ali I, Saleh TA, Nayak A, Agarwal S. Chemical treatment technologies for
18
waste-water recycling-an overview. RSC Adv 2012; 2: 6380-6388.
19
[6] Tan KL, Hameed BH. Insight into the adsorption kinetics models for the removal of
20
contaminants from aqueous solutions. J Taiwan Inst Chem E 2017; 74: 25-48.
AC C
EP
TE D
10
ACCEPTED MANUSCRIPT -43-
[7] Ramakrishna KR, Viraraghavan T. Dye removal using low cost adsorbents. Water Sci
2
Technol; 1997: 36: 189-196.
3
[8] Babel S, Kurniawan TA. Low-cost adsorbents for heavy metals uptake from
4
contaminated water: a review. J Hazard Mater; 2003: 97: 219-243.
5
[9] Byun J, Patel HA, Thirion D, Yavuz CT. Charge-specific size-dependent separation of
6
water-soluble organic molecules by fluorinated nanoporous networks. Nat Commun; 2016: 7:
7
13377-13387.
8
[10] Tamai H, Kakii T, Hirota Y, Kumamoto T, Yasuda H. Synthesis of extremely large
9
mesoporous activated carbon and its unique adsorption for giant molecules. Chem Mater;
M AN U
SC
RI PT
1
1996: 8: 454–462.
11
[11] Yutong G, Zhongzhe W, Jing W, Pengfei Z, Haoran L, Yong W. Design and Fabrication
12
of Hierarchically Porous Carbon with a Template-free Method. Sci Rep; 2014: 4: 6349 –
13
6355.
14
[12] Fu R-W, Li Z-H, Liang Y-R, Li F, Xu F, Wu D-C. Hierarchical porous carbons: design,
15
preparation, and performance in energy storage. New Carbon Mater; 2011: 26:171–179.
16
[13] Fuertes AB, Tartaj P. A facile route for the preparation of superparamagnetic porous
17
carbons. Chem Mater 2006; 18: 1675-1679.
18
[14] Lu A-H, Li W-C, Kiefer A, Schmidt W, Bill E, Fink G, Schüth F. Fabrication of
19
magnetically separable mesostructured silica with an open pore system. J Am Chem Soc
20
2004; 126: 8616-8617.
AC C
EP
TE D
10
ACCEPTED MANUSCRIPT -44-
[15] Veerakumar P, Muthuselvam IP, Hung C-T, Lin K-C, Chou F-C, Liu S-B. Biomass-
2
derived activated carbon supported Fe3O4 nanopaticles as recyclable catalysts for reduction
3
of nitroarenes. ACS Sustainable Chem Eng 2016; 4: 6772–6782.
4
[16] Guo X, Wang R, Yu H, Zhu Y, Nakanishi K, Kanamori K, Yang H. Spontaneous
5
preparation of hierarchically porous silica monoliths with uniform spherical mesopores
6
confined in a well-defined macroporous framework. Dalton Trans 2015; 44: 13592-13601.
7
[17] Dutta S, Bhaumik A, Wu K C-W. Hierarchically porous carbon derived from polymers
8
and biomass: effect of interconnected pores on energy applications. Energ Environ Sci 2014;
9
7: 3574-3592.
M AN U
SC
RI PT
1
[18] Nakanishi K. Porous gels made by phase separation: recent progress and future
11
directions. J Sol-Gel Sci Techn 2000; 19: 65-70.
12
[19] Nakanishi K. Sol–gel process of oxides accompanied by phase separation. B Chem Soc
13
Jpn 2006; 79: 673-691.
14
[20] Nakanishi K, Sato Y, Ruyat Y, Hirao K. Supramolecular templating of mesopores in
15
phase-separating silica sol-gels incorporated with cationic surfactant. J Sol-gel Sci Techn
16
2003; 26: 567-570.
17
[21] Roberts AD, Li X, Zhang H. Porous carbon spheres and monoliths: morphology control,
18
pore size tuning and their applications as Li-ion battery anode materials. Chem Soc Rev
19
2014; 43: 4341-4356.
20
[22] Mukai SR, Nishihara H, Tamon H. Porous microfibers and microhoneycombs
21
synthesized by ice templating. Catal Surv Asia 2006; 10: 161-171.
AC C
EP
TE D
10
ACCEPTED MANUSCRIPT -45-
[23] Siyasukh A, Maneeprom P, Larpkiattaworn S, Tonanon N, Tanthapanichakoon W,
2
Tamon H, Charinpanitkul T. Preparation of a carbon monolith with hierarchical porous
3
structure by ultrasonic irradiation followed by carbonization, physical and chemical
4
activation. Carbon 2008; 46: 1309-1315.
5
[24] Sevilla M, Fuertes AB. Fabrication porous carbon monoliths with a graphitic frame
6
work. Carbon 2013; 56: 155-166.
7
[25] Duffy E, He X, Nesterenko PN, Paull B. Hierarchical porous graphitic carbon monoliths
8
with detonation nanodiamonds: synthesis, characterization and adsorptive properties. J Mater
9
Sci 2015; 50: 6245-6259.
M AN U
SC
RI PT
1
[26] Yu Y, Du J, Liu L, Wang G, Zhang H, Chen A. Hierarchical porous nitrogen-doped
11
partial graphitized carbon monoliths for supercapacitor. J Nanopart Res 2017; 19: 119-131.
12
[27] Dai J, Zhang R, Ge W, Xie A, Chang Z, Tian S, Zhou Z, Yan Y. 3D macroscopic
13
superhydrophobic magnetic porous carbon aerogel converted from biorenewable popcorn for
14
selective oil-water separation. Mater Design 2018; 139: 122-131.
15
[28] Yang N, Zhu S, Zhang D, Xu S. Synthesis and properties of magnetic Fe3O4-activated
16
carbon nanocomposite particles for dye removal. Mater Lett 2008; 62: 645-647.
17
[29] Yin C, Wei Y, Wang F, Chen Y, Bao X. Magnetic hierarchical porous carbon sphere
18
prepared for removal of organic pollutants in water. Mater Lett 2003; 104: 64-67.
19
[30] Wang L, Delgado JJ, Frank B, Zhang Z, Shan Z, Su DS, Xiao F-S. Resin-derived
20
hierarchical porous carbon spheres with high catalytic performance in the oxidative
21
dehydrogenation of ethylbenzene. ChemSusChem 2012; 5: 687-693.
AC C
EP
TE D
10
ACCEPTED MANUSCRIPT -46-
[31] Jia Z, Yang L, Liu J, Wang Q, Zhu R. Preparation of magnetic carbon spheres derived
2
form 8-quinoliolato Fe (III) complex and its application in water treatment. J Ind Eng Chem
3
2015; 21: 111-117.
4
[32] Tosheva L, Parmentier J, Saadallah S, Vix-Guterl C, Valtchev V, Patarin J. Carbon and
5
SiC macroscopic beads from ion-exchange resin templates. J Am Chem Sco 2004; 125:
6
13624-13625.
7
[33] Deshmukh AA, Mhlanga SD, Coville NJ. Carbon spheres. Mater Sci Eng R 2010; 70: 1-
8
28.
9
[34] Yang Y, Fang Z, Chen X, Zhang W, Xie Y, Chen Y, Liu Z, Yuan W. An Overview of
10
pickering emulsions: solid-particle materials, classification, morphology, and applications.
11
Front Pharmacol 2017; 8: 1-20.
12
[35] Sun Z, Zhou X, Lu W, Yue Q, Zhang Y, Cheng X, Li W, Kong B, Deng Y, Zhao D.
13
Interfacial engineering of magnetic particles with porous shells: towards magnetic core –
14
porous shell microparticles. Nano Today 2016; 11: 464-482.
15
[36] Shan J, Wang L, Yu H, Ji J, Amer WA, Chen Y, Jing G, Khalid H, Akram M, Abbasi
16
NM. Recent progress in Fe3O4 based magnetic nanoparticles: from synthesis to application.
17
J Mater Sci Technol 2016; 32: 602-614.
18
[37] Shen Y. Carbothermal synthesis of metal-functionalized nanostructures for energy and
19
environmental applications. J Mater Chem A 2015; 3: 13114-13188.
AC C
EP
TE D
M AN U
SC
RI PT
1
ACCEPTED MANUSCRIPT -47-
[38] Yamauchi Y, Suzuki N, Radhakrishnan L, Wang L. Breakthrough and future: nanoscale
2
controls of compositions, morphologies, and mesochannel orientations toward advanced
3
mesoporous materials. Chem Record 2009; 9: 321–339.
4
[39] Wang L, Keb F, Zhu J. Metal–organic gel templated synthesis of magnetic porous
5
carbon for highly efficient removal of organic dyes. Dalton Trans 2016; 45: 4541-4547.
6
[40] Xia K, Wang G, Zhang H, Liu L, Yu Y, Wang L, Chen A. Synthesis of bimodal
7
mesoporous carbon nanospheres for methyl orange adsorption. J Porous Mat 2017; 24: 1605
8
– 1612.
9
[41] Shi W, Guo F, Wang H, Liu C, Fu Y, Yuan S, Huang H, Liu Y, Kang Z. Carbon dots
10
decorated magnetic ZnFe2O4 nanoparticles with enhanced adsorption capacity for the
11
removal of dye from aqueous solution. Appl Surf Sci 2018; 433: 790 – 797.
12
[42] Liu A, Xia Z, Zhou W, Huang S. Well-dispersed hematite nanoparticles decorating
13
graphene nanosheets: Characterization and performance for methyl orange removal. J Env
14
Chem Eng 2017; 5: 6039 – 6044.
15
[43] Li H, Sun Z, Zhang L, Tian Y, Cui G, Yan S. A cost-effective porous carbon derived
16
from pomelo peel for the removal of methyl orange from aqueous solution. Colloid Surface
17
A 2016; 489: 191-199.
18
[44] Zhang Y, Nan Z. Preparation of magnetic ZnLa0.02Fe1.98O4/MWCNTs composites and
19
investigation on its adsorption of methyl orange fromaqueous solution. Mater Res Bull;
20
2015: 66: 176-185.
AC C
EP
TE D
M AN U
SC
RI PT
1
ACCEPTED MANUSCRIPT -48-
[45] Langmuir I. The constitution and fundamental properties of solids and liquids. J Am
2
Chem Soc; 1916: 38:2221–2295.
3
[46] Freundlich HMF. Over the adsorption in solution. Z Phys Chem; 1906: 57:385–470.
4
[47] Dubinin MM, Radushkevich LV, The equation of the characteristic curve of the
5
activated charcoal. Proc Acad Sci USSR Phys Chem Sect; 1947: 55: 331–337.
6
[48] Brown AM. A step-by-step guide to non-linear regression analysis of experimental data
7
using a Microsoft Excel spreadsheet. Comput Meth Prog Bio; 2001: 65: 191-200.
8
[49] Lagergren S. Zur theorie der sogenannten adsorption gelӧster stoffe. K Sven
9
Vetenskapsakad Handl; 1898: 24: 1-39.
M AN U
SC
RI PT
1
[50] Ho YS. Citation review of Lagergren kinetic rate equation on adsorption reactions.
11
Scientometrics; 2004: 59: 171-177.
12
[51] Ho YS. Review of second-order models for adsorption systems. J Hazard Mater; 2006:
13
136: 681-689
14
[52] Zhao XL, Wang JM, Wu FC, Wang T, Cai YQ, Shi YL, Jiang GB. Removal of fluoride
15
from aqueous media by Fe3O4@Al(OH)3 magnetic nanoparticles. J Hazard Mater; 2010: 173:
16
102-109.
17
[53] Hecht LL, Wagner C, Landfester K, Schuchmann HP. Surfactant Concentration Regime
18
in Miniemulsion Polymerization for the Formation of MMA Nanodroplets by High-Pressure
19
Homogenization. Langmuir; 2011: 27: 2279-2285.
20
[54] Yamamoto T, Sugimoto T, Suzuki T, Mukai SR, Tamon H. Preparation and
21
characterization of carbon cryogel microspheres. Carbon 2002; 40: 1345-1351.
AC C
EP
TE D
10
ACCEPTED MANUSCRIPT -49-
[55] Mohandoss M, Gupta SS, Nelleri A, Prodeep T, Maliyekkal SM. Solar mediated
2
reduction of graphene oxide. RSC Adv; 2017: 7: 957-963.
3
[56] Malard LM, Pimenta MA, Dresselhaus G, Dresselhaus MS. Raman spectroscopy in
4
graphene. Phy Rep; 2009: 473: 51-87.
5
[57] Yoon D, Cheong H. Raman spectroscopy for characterization of graphene. In: Kumar C
6
SSR, editor. Raman Spectroscopy for Nanomaterials Characterization, Berlin; Springer, 2012
7
p. 191-214.
8
[58] Comet M, Pichot V, Siegert G, Spitzer D, Moeglin J-P, Boehrer Y. Use of
9
nanodiamonds as a reducing agent in a chlorate-based energetic composition. Propell Explos
M AN U
SC
RI PT
1
Pyrot 2009; 34: 166-173.
11
[59] Pathania D, Sharma S, Singh P. Removal of methylene blue by adsorption onto
12
activated carbon developed from Ficus carica bast. Arab J Chem; 2017: 10: 1445 – 1451.
13
[60] Foo KY, Hameed BH. Insights into the modeling of adsorption isotherm systems. Chem
14
Eng J 2010; 156: 2-10.
15
[61] Huang J. Molecular sieving effect of a novel hyper-cross-linked resin. Chem Eng J
16
2010; 165: 265-272.
17
[62] Wu Q-Y, Liang H-Q, Li M, Liu B-T, Xu Z-K. Hierarchically porous carbon membranes
18
derived from PAN and their selective adsorption of organic dyes. Chinese J Polym Sci 2016;
19
34: 23-33.
AC C
EP
TE D
10
ACCEPTED MANUSCRIPT -50-
[63] Danish M, Hashim R, Ibrahim MNM, Sulaiman O. Characterization of physically
2
activated acacia mangium wood-based carbon for the removal of methyl orange dye,
3
BioResources 2013; 8: 4323-4339.
4
[64] Youssef AM, El-Khouly S, El-Nabarawy TH. Removal of Pb(11) and Cd(11) from
5
aqueous solution using activated carbons from pecan shells. Carbon Lett;2008: 9: 8-13.
6
[65] Emna E, Duplay J, Darragi F, M’Rabet I, Aubert A, Efficient anionic dye adsorption on
7
natural untreated clay: Kinetic study and thermodynamic parameters. Desalination; 2011:
8
275: 74-81.
9
[66] Faust S, Aly O. Adsorption processes for water treatment. Stoneham MA: Butterworth;
M AN U
SC
RI PT
1
1987.
11
[67] Wang JP, Feng HM, Yu HQ. Analysis of adsorption characteristics of 2,4-
12
dichlorophenol from aqueous solutions by activated carbon fiber. J Hazard Mater; 2007: 144:
13
200-207.
14
[68] Kumar U. Thermodynamics of the adsorption of Cd (II) from aqueous solution on
15
NCRH. Int J Environ Sci Dev; 2011: 2: 334-336.
16
[69] Yousef RI, El-Eswed B, Al-Muhtaseb AH. Adsorption characteristics of natural zeolites
17
as solid adsorbents for phenol removal from aqueous solutions: kinetics, mechanism, and
18
thermodynamics studies. Chem Eng J; 2011: 171: 1143−1149.
19
[70] Xue F, Xu Y, Lu S, Ju S, Xing W. Adsorption of cefocelis hydrochloride on
20
macroporous resin: kinetics, equilibrium, and thermodynamic Studies. J Chem Eng Data;
21
2016: 61: 2179 –2185.
AC C
EP
TE D
10
ACCEPTED MANUSCRIPT -51-
AC C
EP
TE D
M AN U
SC
RI PT
1
ACCEPTED MANUSCRIPT -52-
Table captions
2
Table 1. Synthesis conditions of the prepared carbon spheres and monoliths derived from RF
3
gels.
4
Table 2. Type of isotherm, specific surface area, micropore volume, and mesopore volume
5
of prepared carbons obtained from N2 adsorption-desorption isotherms at -196 °C.
6
Table 3. Chemical compositions and magnetic properties of CF3-900 and ACF3-900.
7
Table 4. Langmuir, Freundlich, and D-R isotherm parameters for MO adsorption.
8
Table 5. Thermodynamic properties of MO adsorption onto the prepared carbon spheres.
9
Table 6. Pseudo-second-order (PSO) and pseudo-first-order (PFO) kinetic parameters of MO
M AN U
SC
RI PT
1
adsorption onto the prepared carbon spheres.
11
Table 7. Comparison of maximum monolayer adsorption capacity (Q0) of MO, magnetism,
12
and reusability of various carbon-based adsorbents.
AC C
EP
TE D
10
ACCEPTED MANUSCRIPT -53-
Figure captions
2
Fig. 1. Schematic representation of carbon sphere fabrication.
3
Fig. 2. (a) SEM images showing the macroporous textures; (b) macropore size distributions
4
determined by the mercury intrusion method.
5
Fig. 3. (a) Schematic illustration of the synthetic strategy to fabricate spherical RF gels
6
(carbon precursors) via sol-gel polymerization in a water-in-oil emulsion system; (b) SEM
7
image of the obtained spherical RF gels; (c) SEM image of carbon spheres (CF3-900)
8
derived from carbonization at 900 °C; (d) particle size distributions of RF spherical gels (red)
9
and carbon spheres (black) with different % Vol. of SPAN80 at 0.2 % (rectangular), 0.3 %
M AN U
SC
RI PT
1
(triangle), and 0.4 % (circle).
11
Fig. 4. Demonstration of effect of acetone treatment before carbonization; (a) SEM image of
12
CF3-900 without acetone treatment; (b) SEM image of CF3-900 with acetone treatment; (c)
13
comparison of particle size distributions.
14
Fig. 5. (a), (b) Nitrogen adsorption-desorption isotherms of CF0-900, CF3-500, CF3-600,
15
CF3-700, CF3-800, CF3-850, and CF3-900; (c) mesopore width distributions of CF3-800,
16
CF3-850, and CF3-900.
17
Fig. 6. SEM images of (a) CF3-900 and (b) CF3-700, showing the surface morphology of
18
spherical carbons with and without mesoporous structure, respectively.
19
Fig. 7. Effect of CO2 activation on the physical morphology and porous properties of ACF3-
20
900; (a) particle size distribution; (b) N2 adsorption-desorption isotherm and mesopore width
21
distribution (inset-picture); (c) SEM image of ACF3-900 at different magnifications.
AC C
EP
TE D
10
ACCEPTED MANUSCRIPT -54-
Fig. 8. (a) Raman spectra showing the develop of the graphitic structure with respect to
2
reaction temperature and synthesis conditions, (b) XRD data showing the graphitic phase
3
formation of the samples, (c) and (d) TEM images with HRTEM inset of CF3-900 and
4
ACF3-900, respectively, showing the multilayer sheet of the samples.
5
Fig. 9. (a) and (b) Comparison of XRD patterns of CF3-900 vs CFe-900 and ACF3-900 vs
6
AFe-900 to study the Fe3O4 phase formation; (c) and (d) SEM-EDS mapping of CF3-900 and
7
ACF3-900, respectively.
8
Fig. 10. (a) VSM results showing magnetization curves of CF3-900 and ACF3-900; (b)
9
pictures for illustration of convenient separation of MHPC spheres (ACF3-900) by a magnet
M AN U
SC
RI PT
1
after MO removal.
11
Fig. 11. (a) Effects of carbon dosages and (b) initial concentrations on MO removal
12
efficiency of carbon spheres.
13
Fig. 12. (a)–(c) Isotherm fitting curves for MO adsorption onto carbon spheres; (d)-(f) kinetic
14
models fitting curves for MO adsorption onto carbon spheres.
15
Fig. 13. Reusability of ACF3-900 and CF3-900 at four cycles for MO removal (initial
16
concentration and carbon dosage of 20 mg/L and 0.4 g/L, respectively).
AC C
EP
TE D
10
ACCEPTED MANUSCRIPT
Table 1.
Conditions Carbonized or activated temperature (°C) 900 900 900 900 850 800 700 600 500 900
EP AC C
Macroscopic shape
Monolith Monolith Monolith Monolith and sphere Sphere Sphere Sphere Sphere Sphere Sphere
SC
Carbonization or CO2-activation Carbonization with N2 Carbonization with N2 Carbonization with N2 Carbonization with N2 Carbonization with N2 Carbonization with N2 Carbonization with N2 Carbonization with N2 Carbonization with N2 Activation with CO2
TE D
CF0-900 CF1-900 CF2-900 CF3-900 CF3-850 CF3-800 CF3-700 CF3-600 CF3-500 ACF3-900
Fe/RFratios (g/g) 0.003 0.010 0.023 0.023 0.023 0.023 0.023 0.023 0.023
M AN U
Samples
RI PT
Synthesis conditions of the prepared carbon spheres and monoliths derived from RF gels.
ACCEPTED MANUSCRIPT
Table 2. Type of isotherm, specific surface area, micropore volume, and mesopore volume of prepared carbons obtained from N2 adsorption-desorption isotherms at -196 °C. Type of isotherm
Vmic (cm3/g)
Vmeso (cm3/g)
685 400 467 553 584 569 12 603
0.27 0.16 0.19 0.22 0.24 0.23 0.4 x 10-2 0.24
N/D* 0.21 0.21 0.19 N/D* N/D* N/D* 2.16
AC C
EP
TE D
M AN U
SC
CF0-900 I CF3-900 I + IV CF3-850 I + IV CF3-800 I + IV CF3-700 I CF3-600 I CF3-500 I ACF3-900 I+IV Remark: *N/D = not determined
SBET (m2/g)
RI PT
Samples
ACCEPTED MANUSCRIPT
Table 3. Chemical compositions and magnetic properties of CF3-900 and ACF3-900. Chemical composition* Atomic % Weight % C O Fe C O Fe 96.53 3.12 0.35 94.36 4.07 1.57 95.06 4.47 0.47 92.13 5.77 2.10
Magnetic properties**
RI PT
Samples
AC C
EP
TE D
M AN U
SC
Hc (Oe) CF3-900 134.3 ACF3-900 234.0 Remarks: *The chemical compositions of carbon (C), oxygen (O), and Iron (Fe) are characterized by EDS. **The magnetic properties were tested by VSM.
Ms (emu/g) 0.4 5.8
ACCEPTED MANUSCRIPT
Table 4. Langmuir, Freundlich, and D-R isotherm parameters for MO adsorption.
AC C
EP
TE D
M AN U
SC
ACF3-900 CF3-900 CF3-700
Q0 (mg/g) 1522.6 77.1 16.2
D-R E (kJ/mol) 4.1 12.0 3.5
R2 (-) 0.9874 0.9905 0.9879
RI PT
Samples
Isotherm model constants and correlation coefficients Langmuir Freundlich b Q0/SBET R2 Kf n R2 qDR (L/mg) (mg/m2) (-) (mg/g)(L/mg)1/n (-) (-) (mg/g) 1.0 x 10-2 2.5 0.9941 63.0 1.9 0.9983 3.6 x 10-3 1.5 0.2 0.9940 39.2 6.3 0.9948 2.8 x 10-4 4.6 x 10-3 0.3 x 10-1 0.9977 0.2 1.4 0.9976 2.9 x 10-5
ACCEPTED MANUSCRIPT
Table 5.
Samples
∆S° (J/mol-K) -90.8 -197.2 -106.5
30 °C -14.9 -13.1 4.5
∆G° (kJ/mol) 40 °C 60 °C -14.0 -12.2 -11.2 -7.2 5.5 7.7
AC C
EP
TE D
M AN U
SC
ACF3-900 CF3-900 CF3-700
∆H° (kJ/mol) -42.4 -72.9 -27.8
RI PT
Thermodynamic properties of MO adsorption onto the prepared carbon spheres. 80 °C -10.3 -3.3 9.8
R2 (-) 0.9951 0.9655 0.9578
ACCEPTED MANUSCRIPT
Table 6. Pseudo-second-order (PSO) and pseudo-first-order (PFO) kinetic parameters of MO adsorption onto the prepared carbon spheres.
AC C
EP
TE D
M AN U
SC
ACF3-900 CF3-900 CF3-700
k2 (g/mg· min) 1.5 x 10-2 4.4 x 10-3 1.6 x 10-2
RI PT
Samples
Adsorption kinetics parameters and correlation coefficients Pseudo-second order (PSO) Pseudo-first order (PFO) qe2 h2 R2 k1 qe1 (mg/g) (mg/g· min) (-) (min-1) (mg/g) 49.7 36.8 0.9794 3.8 x 10-1 48.2 49.2 10.6 0.9833 1.2 x 10-1 46.4 1.3 0.3 x 10-1 0.9912 1.8 x 10-2 1.1
R2 (-) 0.9379 0.9718 0.9843
ACCEPTED MANUSCRIPT
Table 7.
Maximum adsorption capacity (mg/g)
References
yes
No report
152.0
[29]
yes no
yes No report
182.8 570.0
[39] [40]
yes
No report
no
yes
no yes yes
yes
EP AC C
SC
Reusability
181.2
[41]
545.5
[42]
680.2
[43]
yes
81.0
[44]
yes
1522.6
This work
TE D
Magnetic hierarchical porous carbon sphere Magnetic porous carbon Carbon nanosphere Carbon dots decorated magnetic nanoparticle Hematite nanoparticles decorating graphene nanosheet Porous carbon derived from pomelo peel MWCNTs Magnetic hierarchical porous carbon (MHPC) sphere
Magnetism
M AN U
Carbon-based adsorbents
RI PT
Comparison of maximum monolayer adsorption capacity (Q0) of MO, magnetism, and reusability of various carbon-based adsorbents.
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT