Sei sulla pagina 1di 20

Research Article

Adsorption Science & Technology


2018, Vol. 36(5–6) 1366–1385
Removal of nitrobenzene ! The Author(s) 2018
DOI: 10.1177/0263617418771823
from aqueous solution by journals.sagepub.com/home/adt

adsorption onto carbonized


sugarcane bagasse

Dunqiu Wang
Guangxi Key Laboratory of Environmental Pollution Control Theory
and Technology, Guilin University of Technology, Guilin, China; Guangxi
Collaborative Innovation Center for Water Pollution Control and
Water Safety in Karst Area, Guilin University of Technology,
Guilin, China

Huijun Shan
Guangxi Key Laboratory of Environmental Pollution Control Theory
and Technology, Guilin University of Technology, Guilin, China

Xiaojie Sun and Hongxia Zhang


Guangxi Key Laboratory of Environmental Pollution Control Theory
and Technology, Guilin University of Technology, Guilin, China; Guangxi
Collaborative Innovation Center for Water Pollution Control and
Water Safety in Karst Area, Guilin University of Technology,
Guilin, China

Yanhua Wu
Hezhou Solid Waste and Hazardous Chemical Environmental
Management Center, Hezhou, China

Abstract
A sorbent was prepared by charring sugarcane bagasse (SCB) and used to remove nitrobenzene
from aqueous solution. The surface area, morphology, and functional groups of the adsorbent
were characterized by Brunauer–Emmett–Teller method, scanning electron microscopy, and
Fourier transforms infrared spectroscopy. Analysis indicated that oxygen-containing functional

Corresponding author:
Xiaojie Sun, Guilin University of Technology, Guilin Jiangan Road No.12, China, Guilin, 541004, China.
Email: sunxiaojie@glut.edu.cn

Creative Commons CC BY: This article is distributed under the terms of the Creative Commons Attribution
4.0 License (http://www.creativecommons.org/licenses/by/4.0/) which permits any use, reproduction and
distribution of the work without further permission provided the original work is attributed as specified on the SAGE and
Open Access pages (https://us.sagepub.com/en-us/nam/open-access-at-sage).
Wang et al. 1367

groups, such as C ¼ O, –OH, –COOH, and C–O–C, may be involved in the adsorption process.
The adsorption of nitrobenzene was investigated under different operating conditions, including
adsorbent dosage, initial nitrobenzene concentration, pH, and contact duration. Four kinetic
models were applied to describe the adsorption process. Results revealed that the optimal
sorbent mass was 0.3 g/50 mL at pH 5.8 and 25 C. The kinetic data obeyed the pseudo-
second-order kinetic model (R2 > 0.9965). In addition, Langmuir and Freundlich isotherm
models were employed to describe the adsorption equilibrium. The Freundlich model presented
better fitting for the adsorption equilibrium, suggesting that the carbonized SCB surface had a
heterogeneous nature. The maximum adsorption capacities calculated by the Langmuir model
were 38.27, 41.72, and 44.70 mg/g at 25 C, 35 C, and 45 C, respectively. The calculated values of
DG0 and DH0 indicated the spontaneous and exothermic nature of the adsorption process at the
considered temperature range. The adsorption mechanism of nitrobenzene onto carbonized SCB
cannot be described either as physical adsorption or chemisorption. This study demonstrated
that SCB biochar is a potential sorbent for removing nitrobenzene from aqueous solutions.

Keywords
Nitrobenzene, adsorption, carbonization, sugarcane bagasse, kinetic model

Introduction
Nitrobenzene (NB), which is a well-known highly toxic organic compound, can present high
risks to ecological and human health even at low concentrations (Wang et al., 2009), and it is
widely used in the manufacture of dyes, explosives, petrochemicals, and pesticides. After
being used, the NB in solutions is generally discharged into wastewater treatment plants,
where a substantial proportion of it cannot be removed by conventional treatment processes
and is discharged into the surrounding aquatic environment (Qin and Xu, 2016). Given its
poor biodegradability, long-term residue, and accumulation in the environment, many
countries have listed NB as a priority pollutant. In 2014, the Office of Water, Science,
and Technology in the US Environmental Protection Agency (EPA) decreased the tolerance
limit of NB in water and organisms from 17 to 10 lg/L to promote public health and
environmental protection (Dai et al., 2016). Therefore, strategies for removing NB from
aqueous solutions must be developed.
The conventional methods for NB removal from aqueous solutions can be divided into
three main categories: physical, chemical, and biological treatments (Wei et al., 2010).
Adsorption is the most versatile and effective method for NB removal (Jin et al., 2011).
Examples of adsorbents for NB adsorption include nanocrystalline hydroxyapatite (Wei
et al., 2010), MCM-41 (Qin and Ma, 2010), marine sediments (Zhao et al., 2003), carbon
materials released during woody biomass combustion (Dai et al., 2010), activated sludge
modified by cetyltrimethylammonium bromide (Pand and Guan, 2010), and lipoid adsorp-
tion materials (Wen et al., 2012). Activated carbon is a commonly used adsorbent that
possesses a large pore structure and excellent surface chemical adsorption characteristics
(Wei et al., 2017); however, it is not economical. Therefore, low-cost adsorbents, such as
industrial waste, natural ores, and agricultural by-products, are preferable (Sharma and
Kaur, 2011).
1368 Adsorption Science & Technology 36(5–6)

As the main by-product of the sugar industry, sugarcane bagasse (SCB) is the fibrous
residue of sugarcane stems after the extraction of sugar juice by crushing and pressing. SCB
accounts for approximately 25% of the dry weight of sugarcane. The main components of
SCB are cellulose, hemicellulose, and lignin, all of which contain various functional groups,
such as hydroxyl, carboxyl, phenolic hydroxyl, and so on. Given that SCB has a high carbon
content, is available abundantly, and is a non-toxic material, it can serve as a natural
feedstock for manufacturing biocarbon to improve waste management and protect the
environment. The biochar is characterized by a high surface area, a porous structure, and
functional groups characterizing, suggesting that it might be a good alternative for removing
different pollutants from aqueous solution (Abdelhafez and Li, 2016; Bhatnagar and
a€
Sillanp€ a, 2011). In addition, previous biochar adsorption studies have indicated that the
magnitude adsorption capacity of NB on biochar varies at different temperatures. The
reason is that the adsorbability of biochar depends strongly on its physical and chemical
properties, which are affected by temperature. The adsorption isotherm equation can be
used to understand the surface characteristics of the adsorbent and its interaction with the
adsorbate. Thus, studies on equilibrium adsorption are vital to further elucidate the adsorp-
tion mechanism.
This study mainly aimed to examine the adsorption capacity of NB in aqueous solutions
by using carbonized SCB. The adsorbent was characterized using Brunauer–Emmett–Teller
(BET) method, Fourier transform infrared spectroscopy (FTIR), and scanning electron
microscopy (SEM) techniques. The influences of several operating parameters, such as
pH, contact time, temperature, initial NB concentration, and adsorbent dosage, were deter-
mined. The kinetic and equilibrium data of adsorption were analyzed to investigate the
adsorption mechanism of NB.

Materials and methods


Materials
SCB was obtained and selected from a sugar factory in Guangxi, China. NB (98%, without
further purification) was supplied by Shenyang Chemical Company, China. n-Hexane
(C6H14) was acquired from Sinopharm Chemical Reagent Co., Ltd (Shanghai, China)
and used as a chromatographic-grade reagent. All other chemicals used were of analyti-
cal grade.

Preparation of carbonized SCB


The carbonized SCB adsorbent used in this study was prepared as follows: Prior to its use,
SCB was rinsed with running water, soaked in ultrapure water for 24 h, and then dried at
110 C for 24 h. The dried SCB was placed in a carbonization furnace at 500 C for 4 h. After
natural cooling, the sample was grinded to powder (60 mesh). The obtained powder was the
carbonized SCB. The adsorbent was stored in a desiccator for adsorption analyses.

Characteristics of adsorbent
The specific surface area of the carbonized bagasse adsorbent was determined using the
NOVAe1200 automatic surface analyzer (Quantachrome, USA) for nitrogen adsorption–
desorption experiments under nitrogen atmosphere at 77 K. The samples were degassed at
Wang et al. 1369

120 C for 3 h prior to the measurements. The specific surface area was calculated using BET
equation. Total pore volume (Vtotal) and average pore diameter (dav) were derived through
Barrett–Joyner–Halenda (BJH) method.
The FTIR spectra of the samples were recorded on a NEXUS470 FTIR spectrometer
(ThermoNicolet, USA) to characterize the various functional groups present in the carbon-
ized bagasse. Analysis was conducted using KBr pellets made by mixing KBr with the fine
powder of the carbonized bagasse samples (20:1 mass ratio of KBr to SCB). The resolution
was set to 4 cm1, and the range was between 4000 and 500 cm1. The morphology of the
adsorbent was directly observed through S–4800 scanning electron microscopy (Hitachi,
Japan) analysis. The elemental compositions of samples were determined with an EA2400
II elemental analyzer (PerkinElmer, USA).

Adsorption experiments
Effects of adsorption conditions. The effects of the adsorption conditions on the NB adsorption
onto the carbonized SCB were studied at pH 5.8. Desired amounts of carbonized SCB
sorbent powder were placed in 50 mL of NB solution in a set of conical flasks and agitated.
Then, the samples were placed in a water bath oscillator at a constant temperature (25 C,
35 C, or 45 C) and an agitation speed of 200 r/min until the adsorption equilibrium of the
solid–solution mixture was reached. All samples were collected and filtered by a micropore
membrane with a pore size of 0.45 lm, and 1–2 mL of the filtrate was discarded. The
residual concentrations of the NB solution were determined by gas chromatography
(Agilent6890, USA).
Seven initial concentrations of NB (50, 100, 150, 200, 300, 400, and 500 mg/L) were used
to determine the effects of the initial concentration at 25 C, 35 C, or 45 C. The effect of the
adsorbent dosage was studied using different adsorbent doses (0.05, 0.1, 0.2, 0.3, 0.4, 0.5,
0.6, 0.8, and 1.0 g) and 50 mL of 200 mg/L NB solution at 25 C. The NB solution was
studied at different pH (2–10), and the solutions were adjusted using HCl and
NaOH solutions.

Adsorption kinetics. Kinetic studies were performed in a set of conical flasks (100 mL) at a
constant temperature (25 C) by shaking 0.3 g of carbonized SCB sorbent powder in 50 mL
NB solutions (50, 200, and 400 mg/L; initial pH = 5.8). Then, each capped conical flask was
placed in a water bath oscillator at an agitation speed of 200 r/min. Aqueous samples were
obtained from different conical flasks at different time intervals (5, 10, 15, 20, 30, 40, 50, 60,
70, and 90 min). All samples were filtered similarly through a 0.45 lm membrane filter to
remove the carbon fines. The total NB concentration in aqueous solutions was determined
through gas chromatography (Agilent6890, USA). The adsorption capacity of NB at time t,
qt (mgg1) was calculated using the following equations

qt ¼ ðC0  Ct ÞV=W (1)

where C0 (mgL1) is the initial concentration of NB, Ct (mgL1) is the liquid phase con-
centration of NB at any time, V (L) is the volume of the solution, and W (g) is the quality of
the SCB used.
1370 Adsorption Science & Technology 36(5–6)

Adsorption isotherm. The NB adsorption capacity of the carbonized SCB was obtained using
adsorption isotherms. In this experiment, 50 mL of NB solutions at different concentrations
(50, 100, 150, 200, 300, 400, and 500 mg/L) was mixed with 0.3 g carbonized SCB at 25 C,
35 C, or 45 C in a water bath oscillator at an agitation speed of 200 r/min. After 90 min, the
bottles were removed from the oscillator, and the final concentrations of NB in the solution
were analyzed. The amount of adsorbed NB per unit of adsorbent mass, qe (mgg1), was
calculated using equation (2)

qe ¼ ðC0  Ce ÞV=W (2)

where C0 and Ce (mgL1) are the initial and equilibrium concentrations, respectively, of the
adsorbate; V (L) is the experimental volume of the adsorbate in the solution; and W (g) is
the mass of the adsorbent used.

Result and discussion


Characterization of the sorbent
Table 1 shows the porosity of the raw and carbonized SCB. SCB possesses not only a larger
pore size than the diameter of the NB molecule (0.33  0.02 nm) but also a larger specific
surface area. Therefore, SCB can be used as an adsorbent for NB adsorption. The trend of
the BET surface area followed the order of carbonized SCB (191.1 m2g1) > raw SCB
(65.89 m2g1), indicating that the surface area of the carbonized SCB sorbents significantly
increased after modification. The carbonized SCB exhibited a larger total pore volume and
average pore diameter than those of the SCB. After calcination, the components of the SCB
were decomposed, and the volatiles were released. The volume of the pores was increased,
and new pores were formed.
Figure 1 presents the SEM images of the SCB and carbonized SCB sorbent. Figure 1(a)
and (b) shows the SEM diagrams of the SCB and the partially magnified SCB, and Figure 1
(c) and (d) depicts the SEM diagrams of the carbonized SCB and the partially magnified
carbonized SCB. As illustrated in Figure 1(a), the bagasse was in the form of debris, and the
surface of the sample contained few unevenly distributed holes. As shown in Figure 1(b),
numerous micropores were observed in the inner channel of the SCB. After carbonization,
the carbonized SCB exhibited a sponge-like structure, and the number of small pores were
significantly increased compared with that in the SCB.
Figure 2 shows the FTIR spectra of the unmodified and carbonized SCB. The SCB
presented six peaks located at 3439.88, 2921.12, 1630.70, 1383.77, 1058.06, and
670.51 cm1. The band observed at 3439.88 cm1 represented the stretching vibration of
the hydroxyl groups. The band at 2921.21 cm1 was attributed to the CH2 units in the

Table 1. Porosity parameter of different sorbents.

Specific surface Total pore volume Average pore


Sample area (m2g1) (cm3g1) diameter (nm)

Raw SCB 65.89 0.04 1.91


Carbonized SCB 191.1 0.17 1.95
Wang et al. 1371

Figure 1. SEM images of (a) SCB, (b) partially magnified SCB, (c) carbonized SCB, and (d) partially magnified
carbonized SCB.

Figure 2. FTIR spectra for raw SCB and the carbonized SCB.
1372 Adsorption Science & Technology 36(5–6)

biopolymers (Chen et al., 2005). The peak at 1630.70 cm1 was assigned to C = O vibra-
tions in hemicellulose. The band at 1383.77 cm1 was attributed to the bending vibration of
–COOH and the phenol groups (Xu et al., 2008). An adsorption band appeared at
1058.06 cm1, which represented the aliphatic C–O–C and alcohol –OH in cellulose
(Chen et al., 2008). The peak at 670.51 cm1 was assigned to the out-of-plane bending of
–OH. All these bands changed after pyrolysis. In view of the humid climate in March and
April in Guilin, the carbonized SCB was wet, such that the –OH peak (3440.24 cm1)
intensity was too strong. The band intensities at 2923.24 (CH2) and 1121.21 cm1 (aliphatic
C–O–C and alcohol–OH) were decreased, whereas the bands at 1629.90 (C = O) and
1383.94 cm1 (–COOH and phenol group) were preserved. The intense band for –OH
(670.51 cm1) disappeared after carbonization, indicating that the hydrogen bonds among
cellulose, lignin, and hemicellulose were ruptured during pyrolysis. Analysis suggested that
oxygen-containing functional groups (e.g., C = O, –OH, –COOH, and C–O–C) may be
involved in the adsorption process and that they can provide sufficient adsorption sites
for adsorption.
Table 2 presents the elemental analysis results for the SCB and the carbonized SCB.
Carbon and oxygen were the main components of biochar. The carbon content increased
from 44.28 to 56.02% after carbonization, followed by a simultaneous decrease in oxygen
and hydrogen contents. The molar ratios of O/C and (N þ O)/C also decreased because of
the reduction of the surface hydrophilic and polar functional groups. The H/C ratio > 1.0
suggested that the carbonized SCB contained a good amount of original organic residues
(Chen et al., 2008).

Effects of sorption conditions


The effects of contact time, mass of carbonized SCB, and initial concentration on the
adsorption of NB were investigated. All parameters, except the desired one, were set
as constant.

Contact time. Equilibrium time is an important parameter used to assess the efficiency and
feasibility of an adsorbent (Sharma, 2003). Figure 3 shows the effect of contact time on the
adsorption of the carbonized SCB. The NB concentration decreased sharply during the first
period of contact (20 min) possibly due to the abundant available vacant sites on the
adsorbent surface (Kavitha and Namasivayam, 2007). Figure 3 reveals that the amount
of adsorbed NB (mg/g) increased with increasing contact time until it gradually approached
the equilibrium state because of the continuous decrease in driving force and NB molecular

Table 2. Elemental analysis results of samples.

Sample Raw SCB Carbonized SCB

C (%) 44.28 56.02


H (%) 5.83 5.41
O (%) 48.51 21.98
N (%) 0.18 0.45
C/H 1.58 1.16
O/C 0.82 0.29
(NþO)/C 0.83 0.30
Wang et al. 1373

100

80

Removal (%)
60

40

Initial NB=50mg/L
20
Initial NB=200mg/L
Initial NB=400mg/L
0
0 20 40 60 80 100

35

30
NB adsorbed (mg/g)

25

20
Initial NB=50mg/L
15 Initial NB=200mg/L
Initial NB=400mg/L
10

0
0 20 40 60 80 100

Time (min)

Figure 3. Effect of the contact time on NB adsorption by the carbonized SCB (initial pH ¼ 5.8, amount of
sorbent 0.3 g/50 mL, 25 C).

diffusion rate. After adsorption equilibrium, the amount of adsorbed NB and the concen-
tration of NB in the liquid phase remained nearly constant, and the average removal effi-
ciency of NB reached approximately 99.2% (8.25 mg/g adsorption capacity), 71.2% (23.69
mg/g adsorption capacity), and 53.6% (35.62 mg/g adsorption capacity) at 25 C and pH 5.8
at initial NB concentrations of 50, 200, and 400 mg/L, respectively. The removal rate of NB
at the adsorption equilibrium decreased with increasing initial NB concentration. This find-
ing could be due to the limited number of adsorption sites available for the uptake of NB at
a fixed adsorbent dosage.

Adsorbent dose. The adsorption of NB onto the carbonized SCB sorbent was studied by
varying the adsorbent quantity in the test solution while maintaining the contact duration,
initial NB concentration, and initial pH fixed at 25 C. Figure 4 demonstrates that the
removal ratio of NB increased with increasing adsorbent dosage because of the larger sur-
face area and the availability of numerous adsorption sites on the surface (Abdelhafez and
Li, 2016). However, the amount of adsorbed NB per unit mass of the adsorbent decreased
with increasing adsorbent mass. This finding may be attributed to the overlapped or aggre-
gated adsorbent surface area available to NB and to the increased diffusion path length
(Yinian et al., 2011). The percentages of adsorbed NB increased from 40.30% to 88.00% on
1374 Adsorption Science & Technology 36(5–6)

100 100

80

NB adsorbed (mg/g)
80

Removal rate (%)


60

NB adsorbed (mg/g) 60
40
Removal rate (%)

20
40

0
0 5 10 15 20

Adsorbent mass (g/L)

Figure 4. Effect of adsorbent doses on the adsorption of NB on the carbonized SCB (initial concentration
200 mg/L, at pH ¼ 5.8 and 25 C).

the carbonized SCB, whereas the amount of adsorbed NB decreased from 80.28 mg/g to 8.8
mg/g as the dose of the carbonized SCB adsorbent was increased from 0.05 g/50 mL to 1.0 g/
50 mL. A ratio higher than 10 g/L did not linearly improve the adsorption rate, indicating
that the availability of the adsorption surface was not the limiting factor in the process.
Thus, the remaining experiments applied an adsorbent dose of 0.3 g/50 mL for the carbon-
ized SCB.

Initial NB concentration. Kinetic dependencies were measured at various initial NB concen-


trations (50–500 mg/L) at 25 C, 35 C, and 45 C while the other experimental parameters
were kept constant. Figure 5 shows the dependencies of the adsorption capacity and the
removal percentage versus the initial concentration. The overall trend was similar for dif-
ferent temperatures: as the NB concentrations in the test solution increased, the actual
amount of adsorbed NB per unit mass of adsorbent increased from 8.26 mg/g to 36.67
mg/g at 25 C, from 8.29 mg/g to 40.47 mg/g at 35 C, and from 8.32 mg/g to 43.76 mg/g at
45 C. When the initial concentration was less than 400 mg/L, the number of consumed
adsorption sites decreased with increasing initial concentration of NB, resulting in decreased
number of remaining adsorption sites on NB. At this stage, the adsorption capacity con-
stantly increased. As the initial concentration increased, the total available adsorption sites
were completely replaced by NB, and the saturation adsorption capacity was achieved. By
contrast, the adsorption rate decreased continuously from 99.20% to 44.06% at 25 C, from
99.67% to 48.66% at 35 C, and from 99.81% to 52.62% at 45 C as the initial NB concen-
tration increased from 50 to 500 mg/L.

Solution pH. To determine the effect of pH on the adsorption capacity of the carbonized
SCB, solutions were prepared at different pH values ranging from 2 to 10. Figure 6 presents
the dependence of pH on the adsorption of NB. The adsorption effect of the carbonized
SCB on NB in aqueous solution was hardly affected by changing the pH of the solution.
When the pH was increased from 5.8 to 10, the amount of adsorbed NB decreased from
Wang et al. 1375

50 120

45
100
40

35
80
NB adsorbed (mg/g)

Removal rate (%)


30

25 60

20
25 °C 40
15
35 °C
45 °C
10 25 °C 20
35 °C
5 45 °C

0 0
0 100 200 300 400 500

Initial NB concentration (mg/L)

Figure 5. Effect of initial concentration on NB sorption onto the carbonized sorbent prepared from SCB
(initial pH ¼ 5.8, amount of sorbent 0.3 g/50 mL).

40

Carbonized SCB
35
NB adsorbed (mg/g)

30

25

20

15
2 4 6 8 10
pH

Figure 6. Effect of pH on NB adsorption (initial concentration 200 mg/L, amount of sorbent 0.3 g/
50 mL, 25 C).

29.54 to 28.72 mg/g. Given that NB is a nonionizable compound, the change in pH exerted
an insignificant effect on the solubility of NB. Therefore, in this study, the pH was 5.8 under
all experimental conditions (the amount of adsorbed NB was relatively high).

Sorption kinetics
Kinetics is a critical parameter in the design of the sorption process and in evaluating the
suitability of a sorbent because it controls the size of the sorption units (Rezgui et al., 2017).
1376 Adsorption Science & Technology 36(5–6)

12
3 Pseudo-first-order kinetics Pseudo-second-order kinetics
10
2
Initial = 50mg/L
Initial = 200mg/L
8
1 Initial = 400mg/L
ln(qe-qt)

t/qt
6
0

4
-1

Initial = 50mg/L
-2 Initial = 200mg/L 2
Initial = 400mg/L

-3 0
0 10 20 30 40 50 60 70 80 0 20 40 60 80 100

t (min) t (min)
4
Modified pseudo-first-order kinetics 35 Intra-particle diffusion kinetics
Initial = 50mg/L
3 Initial = 200mg/L
30
Initial = 400mg/L

2 25
qt/qe+ln(qe-qt)

qt (mg/g)

20
1

15
0
10
Initial = 50mg/L
-1 Initial = 200mg/L
Initial = 400mg/L 5

-2 0
0 10 20 30 40 50 60 70 80 1 2 3 4 5 6 7 8 9 10

t (min) t1/2 (min1/2)

Figure 7. Four kinetic models for NB sorption onto the carbonized sorbent prepared from SCB (initial
pH ¼ 5.8, amount of sorbent 0.3 g/50 mL, 25 C).

Four famous kinetic models, namely, pseudo-first-order, pseudo-second-order, modified


pseudo-first-order, and intraparticle diffusion, were used to fit the experimental data of
adsorption time and capacity by linear fitting.

Pseudo-first-order kinetic model. The pseudo-first-order kinetic model, which is commonly used
to determine the adsorption rate based on the adsorption capacity (Somasekhara et al.,
2012), can be expressed as follows

Inðqe  qt Þ ¼ In qe  k1 t (3)

where qt (mgL1) is the amount of adsorbed NB at time t (min), qe (mgL1) is the amount
of the adsorbed NB on the adsorbent at time under equilibrium conditions, and k1 is the
pseudo-first-order rate constant (min1).
After the values of ln(qeqt) versus t were plotted, a straight line was obtained with slope
k1 and intercept lnqe (Figure 7). Table 3 lists the values of k1 and qe at different initial NB
concentrations. The correlation coefficient value R2 for NB adsorption onto the carbonized
Wang et al. 1377

Table 3. Kinetic parameters for NB sorption onto the carbonized sorbent prepared from sugarcane.

Initial NB concentration (mg/L) k1/(1/min) qe (mg/g) R2

Pseudo-first-order constants
50 0.0586 4.58 0.9547
200 0.0612 20.69 0.9909
400 0.0289 19.23 0.9559
Modified pseudo-first-order constants
50 0.0515 8.32 0.9738
200 0.0528 34.14 0.9776
400 0.0224 32.82 0.9747
Initial NB concentration (mg/L) k2/(g/(mg min)) qe (mg/g) h/(g/(mg min)) R2

Pseudo-second-order constants
50 0.0180 8.92 1.4304 0.9981
200 0.0038 26.65 2.7013 0.9984
400 0.0030 37.86 4.2985 0.9965
Initial NB concentration (mg/L) Kd/(g/(mg min1/2)) R2

Intraparticle diffusion constants


50 0.5902 0.7262
200 1.9743 0.8924
400 2.5135 0.9139

SCB sorbent changed in the range of 0.9547–0.9909. Thus, the experimental data did not
agree well with the pseudo-first-order kinetic model.

Modified pseudo-first-order kinetic model. The pseudo-first-order equation was modified by


changing the rate constant (Yang and Al-Duri, 2005). By denoting the rate constant in
the modified pseudo-first-order rate equation as K1 (min1), the following equation
was proposed

qe
k1 ¼ K1 (4)
qt

The modified pseudo-first-order equation can be obtained as follows

qt
þ Inðqe  qt Þ ¼ Inðqe Þ  K1 t (5)
qe

If the adsorption process followed the modified pseudo-first-order kinetic model repre-
sented by equation (5), then the plot of qt/qeþln(qeqt) against t should be a straight line
(Figure 7). Table 3 presents the values of K1 and qe at different temperatures. The correla-
tion coefficient values for NB adsorption onto the carbonized SCB sorbent changed in the
range of 0.9738–0.9776. Thus, the experimental data did not fit well with the modified
pseudo-first-order kinetic model.
1378 Adsorption Science & Technology 36(5–6)

Pseudo-second-order kinetic model. The pseudo-second-order kinetic model, which assumes that
sorption is the interaction between the functional groups on the adsorbent surface and the
adsorbate (Somasekhara et al., 2012), can be represented as follows

t 1 1
¼ þ t (6)
qt k2 q2e qe

The initial adsorption rate, h (mg/(g min)), at t = 0 can be defined as follows

h ¼ k2 q2e (7)

where the initial adsorption rate (h), the equilibrium adsorption capacity (qe), and the
pseudo-second-order constants k2 (g/(mg min)) can be determined experimentally from
the slope and intercept of plot t/qt versus t (Figure 7). The calculated correlations were
closer to unity for the pseudo-second-order kinetics model (R2 = 0.9965–0.9984); therefore,
the adsorption kinetics favorably fitted with the pseudo-second-order kinetic model for the
NB adsorption onto the carbonized SCB sorbent. The k2 and h values were calculated from
Figure 7 as 0.0030–0.0180 g/(mgmin) and 1.4304–4.2985 mg/(gmin), respectively (Table 3).
The equilibrium adsorption capacities (qe) were close to the actual values determined by
the experiment.

Intraparticle diffusion kinetic model. The reaction rate can be modeled using the resistance to
intraparticle diffusion if the sorption is not limited by the reaction between the sorbate and
the active sites. A kinetic model was developed by Weber and Morris Rezgui et al. (2017)
and defined as follows

qt ¼ Kd t1=2 þ C (8)

where Kd is the adsorption constant of W-M (mgg1min1/2), and C is constant. According


to equation (8), the plot of qt versus t1/2 should be a straight line with slope Kd and intercept
C if the adsorption mechanism followed the intraparticle diffusion process. The values of the
intercept represented the thickness of the boundary layer; i.e., a larger intercept corre-
sponded to a greater boundary layer effect.
Figure 7 shows the plot of the mass of adsorbed NB per unit mass of adsorbent; qt versus
t1/2 is presented for NB. The values of Kd obtained from the slope of the straight lines were
0.5902–2.5135 g/(mgmin1/2). As shown in Table 3, the correlation coefficient value R2 for
the NB adsorption onto the carbonized SCB sorbent changed in the range of 0.7262–0.9139.
Thus, the experimental data did not agree well with the intraparticle diffusion kinetic model.
Kinetic profiles can be used not only to determine the equilibrium time (as a function of
experimental conditions) but also to identify the controlling (limiting) mechanisms and steps
(Rezgui et al., 2017). Among the four kinetic models, the pseudo-second-order equation
generates the best fit to the experimental data of the three investigated adsorption systems at
initial NB concentrations of 50, 200, and 400 mg/L for the entire adsorption period. The
regression coefficients were higher than 0.9965 for the concentration range used. The cal-
culated qe from the model was close but not equal to the experimental values. Thus, the
Wang et al. 1379

adsorption of NB onto the carbonized SCB was associated with a mechanism involving the
chemical interaction between the sorbent and the solute (Zabihi et al., 2010), implying that
NB adsorption probably occurred via surface exchange reactions until the surface functional
sites were fully occupied. After which, the NB molecules diffused into the adsorbent network
for further interactions.
The other models generated a poor fit to the results of the experiments conducted in this
study. The pseudo-first-order kinetic model and the modified pseudo-first-order kinetic
model exhibited lower correlation coefficients than the pseudo-second-order kinetic
model, and the calculated qe differed from the experimental qe. This finding suggested the
insufficiency of the model fitted the kinetic data for the initial concentrations examined.
After a short period, the experimental data deviated considerably from linearity.
In the intraparticle diffusion model, if intraparticle diffusion is involved in the sorption
process, then the plot of qt against t1/2 should be linear; if these lines pass through the origin
(if the intercept C is equal to zero), then intraparticle diffusion is the rate-controlling step
(Banat et al., 2003). The values of the constant C at different concentrations were not zero,
indicating that particle diffusion was involved but was not the limiting step. Furthermore,
the intraparticle diffusion model was inapplicable to the investigated adsorption systems
(Yang and Al-Duri, 2005).

Sorption isotherm
Adsorption isotherms can describe how adsorbate molecules interact with the adsorbent
surface between the liquid phase and the solid phase at equilibrium (Chai et al., 2016). The
two isotherm models, Langmuir and Freundlich, are widely employed to describe the exper-
imental data of adsorption isotherms.

Langmuir isotherm. The Langmuir equation is applicable to homogeneous adsorption, where


the adsorption of each molecule onto the surface has an equal adsorption activation energy
(Crini, 2008). The linear form of the Langmuir isotherm equation was represented as follows
(Jianlong, 2002)

Ce =qe ¼ 1=ðqm  KL Þ þ Ce =qm (9)

where Ce is the concentration of the NB solution (mgL1) at equilibrium; qm signifies the


maximum adsorption capacity (mgg1); and KL is Langmuir constant related to the free
energy of adsorption (Lmg1). The linear plot of Ce/qe versus Ce showed that the adsorp-
tion followed the Langmuir isotherm (Figure 8). qm and KL were calculated from the slope
and intercept of the linear plot (Table 4). The R2 values were 0.9879, 0.9931, and 0.9955 for
the adsorption at 25 C, 35 C, and 45 C, respectively. Increasing the temperature improved
the adsorption capacities from 38.27 mg/g at 25 C to 41.72 mg/g at 35 C and to 44.70 mg/g
at 45 C. This finding demonstrated that the adsorption process was endothermic; i.e.,
increasing the adsorption temperature can improve the adsorption capacity of
the adsorbent.
The RL parameter is considered a more reliable indicator of adsorption. RL values have
four possibilities: (1) for favorable adsorption, 0 < RL < 1; (2) for unfavorable adsorption,
RL > 1; (3) for linear adsorption, RL = 1; and (4) for irreversible adsorption, RL = 0
(Senturk et al., 2009).
1380 Adsorption Science & Technology 36(5–6)

(a) 8

5
Ce/qe
4

3
25 °C
2 35 °C
45 °C
1

0
0 100 2 00 3 00

Ce (mg/L)

(b) 4.0

3.5

3.0
lnqe

2.5

25 °C
35 °C
2.0 45 °C

1.5
-2 0 2 4 6

lnCe

Figure 8. Isotherm plots for NB adsorption onto the carbonized SCB sorbent. (a) Langmuir isotherm;
(b) Freundlich isotherm.

Table 4. Isotherm parameters for NB sorption onto the carbonized SCB sorbent.

Temperature ( C) qm (mg/g) KL (L/mg) RL R2

Langmuir constants
25 38.27 0.061 0.0317–0.2466 0.9879
35 41.72 0.084 0.0231–0.1915 0.9931
45 44.70 0.129 0.0152–0.1339 0.9955
Temperature ( C) KF (mg1  1/nL1/ng1) 1/n R2

Freundlich constants
25 9.96 0.2351 0.98873
35 12.00 0.2237 0.99366
45 14.18 0.2167 0.99654
Wang et al. 1381

Table 4 shows that the RL near 0 confirmed the favorable uptake of the NB process,
which indicated a favorable adsorption.

Freundlich isotherm. The Freundlich isotherm is an empirical equation describing the adsorp-
tion of an adsorbate onto a heterogeneous surface of an adsorbent as well as multilayer
adsorption (Chai et al., 2016). This equation is expressed as follows

In qe ¼ In KF þ 1=n In Ce (10)

where qe refers to the amount of adsorbed NB (mgg1) at equilibrium; Ce is the equilibrium


concentration of NB in the solution (mgL1); the Freundlich constant KF (mg1  1/nL1/
n 1
g ) is an approximate indicator of the adsorption capacity; and 1/n is the adsorp-
tion intensity.
KF and n were calculated from the intercept and the slope of plots (Figure 8, Table 4).
The obtained R2 values were 0.9887, 0.9937, and 0.9965 for the adsorption at 25 C, 35 C,
and 45 C, respectively. 1/n is a heterogeneity factor that can be used as a measure of the
deviation from linearity of the adsorption (Crini, 2008). In this study, the 1/n values were
calculated from the slope as 0.2167–0.2351, which ranged from 0 to 1 and implied that the
adsorption involved chemical reactions.
The Langmuir and Freundlich equations were well described for NB adsorption onto the
carbonized SCB under isothermal conditions. However, according to the correlation coef-
ficient R2, the Freundlich isotherm could yield a better fit than the Langmuir isotherm,
indicating that the adsorption of the carbonized bagasse on NB was dominated by the multi-
molecular layer and heterogeneous nature of the adsorbent surface. At the same time, the
correlation coefficient R2 of the Langmuir equation was also above 0.98, suggesting that
the adsorption on the carbonized SCB was a monolayer adsorption, which depended on
the diffusion force.

Thermodynamic parameters
Thermodynamic parameters were calculated to describe the sorption of NB onto the car-
bonized SCB and to gain insight into the mechanism involved in the sorption process. The
thermodynamic parameters for the adsorption of DH0 (enthalpy), DS0 (entropy), and DG0
(Gibbs energy) were calculated using equations (11) and (12)

DG0 ¼ RT In KL (11)

DG0 ¼ DH0  TDS0 (12)

where R is the gas constant and has the value of 8.314 J K1mol1, KL is the Langmuir
constant, and T is the experiment temperature expressed in Kelvin (K).
The values of DH0 and DS0 were determined from the slope and intercept of the linear
plot of DG0 versus T. The negative values of the calculated DG0 (5.002, 5.998, and
7.321 kJ/mol) confirmed the spontaneous nature of the NB adsorption at 25 C, 35 C,
and 45 C, and verified the affinity of the sorbent for NB. The positive calculated values of
DH0 (29.63 J/mol) showed the endothermic nature of adsorption. The positive values of DS0
1382 Adsorption Science & Technology 36(5–6)

(0.1160 kJ/(molK)) suggested the increased randomness at the solid/solution interface


during the adsorption of NB onto the adsorbent.

Regeneration
Regeneration studies can reveal the possibility of recycling the adsorbent. Desorption was
performed using EtOH. The adsorption capacity significantly changed after the third reuse
cycle, and the regeneration efficiencies for the three times were 29.56%, 10.38%, and 8.27%.
This result suggested that the carbonized SCB was unsuitable for many cycles.

Comparisons of NB adsorption
To assess the performance of the carbonized SCB as an adsorbent for NB, it was compared
with other adsorbents, and the data are presented in Table 5. As expected, a good adsorbent
should have a large surface area and sufficient active sites to generate a great adsorption
capacity. The maximum adsorption capacity in Table 5 was calculated by the Langmuir
isotherm model. The results suggested that the amount of adsorbed NB onto raw SCB was
insignificant compared to that onto the carbonized SCB. This finding is consistent with the
results obtained by BET method and SEM images. Moreover, the adsorption capacity of the
carbonized SCB for NB was higher than natural or organic synthetic adsorbent but lower
than that of activated carbon. The low adsorption capacity was most probably due to the
small surface area of carbonized SCB (191.1 m2/g) compared to activated carbons (around
1400 m2/g) (Dai et al., 2010; Villaca~
nas et al., 2006). The carbonized SCB adsorbent utilizes
agricultural waste, has a simple preparation process, and is environmentally friendly.
Therefore, the carbonized SCB is a potential biochar material adsorbent for NB removal
from aqueous solutions.

Mechanism of NB adsorption
Given that SCB has a high carbon content, it can pyrolyze the carbonization of SCB under
complete or partial anoxic conditions to prepare bio-carbonaceous materials. In general,
biochar will generate a highly aromatic and insoluble solid material after pyrolysis carbon-
ization. This substance has a similar structure and polarity to NB. As its specific surface area
increases, it will develop a good stability and an improved pore structure. Therefore, car-
bonaceous materials have the potential adsorption capacity of NB in environmental gover-
nance. In this study, the considerable number of surface oxygen-containing functional
groups in the surface of carbonized SCB could provide adequate adsorption sites and
improve adsorption capacity. From a kinetic point of view, the adsorption of NB onto

Table 5. Comparison of the adsorption capacities of NB for various adsorbents.

Adsorbent Temperature ( C) qm (mg/g) References

Carbonized SCB 25 38.27 This work


Raw SCB 25 2.69 This work
Nanocrystalline hydroxyapatite 25 5.754 Wei et al. (2010)
Marine sediment 25 4.755 Zhao et al. (2003)
Wood-based activated carbon 25 263 Dai et al. (2010)
Granular activated carbon 25 234 Villaca~nas et al. (2006)
Wang et al. 1383

carbonized SCB was associated with a mechanism involving chemical interactions.


However, from the adsorption isotherm analysis, the adsorption process presented mono-
layer and multi-molecular layer adsorption processes. To a certain extent, the magnitude
of the enthalpy can be classified as physical adsorption and chemisorption. The range of
5–40 kJ/mol of DH0 indicates to a physisorption mechanism, and the range of 40–800 kJ/
mol suggests a chemisorption mechanism (Gerçel et al., 2007). The positive values of DH0
(29.63 J/mol) indicated the adsorption of NB was an endothermic and physisorption pro-
cess. Therefore, the mechanism of NB sorption from aqueous solution onto carbonized SCB
can be described neither as a physical adsorption nor a chemisorption.

Conclusion
Carbonized SCB can be used as a low-cost sorbent for NB removal from aqueous solutions.
BET, SEM, and FTIR analyses were conducted to characterize the adsorbent. Oxygen-
containing functional groups (e.g., C = O, –OH, –COOH, and C–O–C) may be involved
in the adsorption process. The adsorption capacity was affected by the adsorbent dosage,
contact time, pH, and initial concentration. The amount of adsorbed NB gradually
increased with increasing contact time and was maintained until equilibrium was reached.
At pH 5.8 and 25 C, the adsorption rates were 99.2%, 71.2%, and 53.6%, and the adsorp-
tion capacities were 8.25, 23.69, and 35.62 mg/g at initial NB concentrations of 50, 200, and
400 mg/L, respectively. Furthermore, the optimal adsorbent dose for the carbonized SCB
was 0.3 g/50 mL. The adsorption kinetics and isotherms followed the pseudo-second-order
model and the Freundlich equation. The maximum adsorption capacities qm were 38.27,
41.72, and 44.70 mg/g. The calculated values of DG0 and DH0 indicated the spontaneous
and exothermic nature of the adsorption process at the examined temperature range.
Analysis implied that the mechanism of NB sorption from aqueous solution onto carbon-
ized SCB can be described neither as a physical adsorption nor a chemisorption.

Declaration of Conflicting Interests


The author(s) declared no potential conflicts of interest with respect to the research, authorship, and/or
publication of this article.

Funding
The author(s) disclosed receipt of the following financial support for the research, authorship, and/or
publication of this article: This work was financially supported by the National Natural Science
Foundation of China (no. 51368012, 51668014); the Natural Science Foundation of Guangxi
Province, China (no. 2016GXNSFDA380007); the Special Funding for “BaGui Scholars”
Construction Projects; the Guangxi Scientific Experiment Center of Mining, Metallurgy and
Environment, and the Guangxi Talent Highland for Hazardous Waste Disposal Industrialization.

References
Abdelhafez AA and Li J (2016) Removal of Pb(II) from aqueous solution by using biochars derived
from sugar cane bagasse and orange peel. Journal of the Taiwan Institute of Chemical Engineers
61: 367–375.
Banat F, Al-Asheh S and Al-Makhadmeh L (2003) Evaluation of the use of raw and activated date pits
as potential adsorbents for dye containing waters. Process Biochemistry 39(2): 193–202.
1384 Adsorption Science & Technology 36(5–6)

Bhatnagar A and Sillanp€a€a M (2011) A review of emerging adsorbents for nitrate removal from water.
Chemical Engineering Journal 168: 493–504.
Chai B, Meng H, Zhao Z, et al. (2016) Removal of color compounds from sugarcane juice by modified
sugarcane bagasse: Equilibrium and kinetic study. Sugar Tech 18(3): 317–324.
Chen B, Johnson EJ, Chefetz B, et al. (2005) Sorption of polar and nonpolar aromatic organic
contaminants by plant cuticular materials: The role of polarity and accessibility. Environmental
Science & Technology 39: 6138–6146.
Chen B, Zhou D and Zhu L (2008) Transitional adsorption and partition of nonpolar aromatic
contaminants by biochars of pine needles with different pyrolytic temperatures. Environmental
Science & Technology 42: 5137–5143.
Crini G (2008) Kinetic and equilibrium studies on the removal of cationic dyes from aqueous solution
by adsorption onto a cyclodextrin polymer. Dyes and Pigments 77: 415–426.
Dai Y, Mihara Y, Tanaka S, et al. (2010) Nitrobenzene-adsorption capacity of carbon materials
released during the combustion of woody biomass. Journal of Hazardous Materials 174: 776–781.
Dai Y, Zhang D and Zhang K (2016) Nitrobenzene-adsorption capacity of NaOH-modified spent
coffee ground from aqueous solution. Journal of the Taiwan Institute of Chemical Engineers
68: 232–238.
€ Ozcan
Gerçel O, € A, et al. (2007) Preparation of activated carbon from a renewable bio-plant of
Euphorbia rigida by H2SO4 activation and its adsorption behavior in aqueous solutions. Applied
Surface Science 253: 4843–4852.
Jianlong W (2002) Biosorption of copper (II) by chemically modified biomass of Saccharomyces
cerevisiae. Process Biochemistry 37: 847.
Jin L, He DD and Wei M (2011) Selective adsorption of phenol and nitrobenzene by b-cyclodextrin-
intercalated layered double hydroxide: Equilibrium and kinetic study. Chemical Engineering &
Technology 34: 1559–1566.
Kavitha D and Namasivayam C (2007) Experimental and kinetic studies on methylene blue adsorption
by coir pith carbon. Bioresource Technology 98(1): 14–21.
Pan J and Guan B (2010) Adsorption of nitrobenzene from aqueous solution on activated sludge
modified by cetyltrimethylammonium bromide. Journal of Hazardous Materials 183: 341–346.
Qin QD and Ma J (2007) Adsorption of nitrobenzene from aqueous solution by MCM-41. Journal of
Colloid and Interface Science 315: 80–86.
Qin Q and Xu Y (2016) Enhanced nitrobenzene adsorption in aqueous solution by surface silylated
MCM-41. Microporous and Mesoporous 232: 143–150.
Rezgui A, Guibal E and Boubakera T (2017) sorption of hg(ii) and zn(ii) ions using lignocellulosic
sorbent (date pits). The Canadian Journal of Chemical Engineering 95: 775–782.
Senturk HB, Ozdes D, Gundogdu A, et al. (2009) Removal of phenol from aqueous solutions by
adsorption onto organomodified Tirebolu bentonite: Equilibrium, kinetic and thermodynamic
study. Journal of Hazardous Materials 172: 353–362.
Sharma P and Kaur H (2011) Sugarcane bagasse for the removal of erythrosin B and methylene blue
from aqueous waste. Applied Water Science 1: 135–145.
Sharma YC (2003) Cr(VI) removal from industrial effluents by adsorption on an indigenous low-cost
material. Colloids and Surfaces A: Physicochemical and Engineering Aspects 215(1–3): 155–162.
Somasekhara RCM, Sivaramakrishna L and Varada RA (2012) The use of an agricultural waste
material, Jujuba seeds for the removal of anionic dye (Congored) from aqueous medium.
Journal of Hazardous Materials 203–204: 118–127.
Villaca~
nas F, Pereira MFR, Orfao JJM, et al. (2006) Adsorption of simple aromatic compounds on
activated carbons. Journal of Colloid and Interface Science 293: 128–136.
Wang XL, Li Y, Wang YZ, et al. (2009) New evidence for the importance of Mn oxides contributed to
nitrobenzene adsorption onto the surficial sediments in Songhua River, China. Journal of
Hazardous Materials 172: 755–762.
Wang et al. 1385

Wei J, Lin Z, He Z, et al. (2017) Bagasse activated carbon with TETA/TEPA modification and
adsorption properties of CO2. Water, Air, & Soil Pollution 228: 128.
Wei W, Sun R, Cui J, et al. (2010) Removal of nitrobenzene from aqueous solution by adsorption on
nanocrystalline hydroxyapatite. Desalination 263: 89–96.
Wen Q, Chen Z, Lian J, et al. (2012) Removal of nitrobenzene from aqueous solution by a novel lipoid
adsorption material (LAM). Journal of Hazardous Materials 209–210: 226–232.
Xu Di Tan X, Chen C, et al. (2008) Removal of Pb(II) from aqueous solution by oxidized multiwalled
carbon nanotubes. Journal of Hazardous Materials 154: 407–416.
Yinian Z, Meina L, Rongrong L, et al. (2011) Phosphorus removal from aqueous solution by Fe(III)-
impregnated sorbent prepared from sugarcane bagasse. Fresenius Environmental Bulletin
20: 1288–1296.
Yang XY and Al-Duri B (2005) Kinetic modeling of liquid-phase adsorption of reactive dyes on
activated carbon. Journal of Colloid and Interface Science 287(1): 25–34.
Zabihi M, Asl AH and Ahmadpour A (2010) Studies on adsorption of mercury from aqueous solution
on activated carbons prepared from walnut shell. Journal of Hazardous Materials 174: 251–256.
Zhao XK, Yang GP and Gao XC (2003) Studies on the sorption behaviors of nitrobenzene on marine
sediments. Chemosphere 52: 917–925.

Potrebbero piacerti anche