PreprintPDF Available

Preparation of porous silica materials using a eucalyptus template method and its efficient adsorption of methylene blue

Authors:
Preprints and early-stage research may not have been peer reviewed yet.

Abstract and Figures

Mesoporous silica has become one of the primary adsorbent materials for solving dye wastewater pollution due to its high specific surface area and good adsorption properties. However, the high cost of the traditional chemical synthesis method limits its wide application. In this thesis, low-cost and high-efficiency porous silica adsorbent materials (PSAM) were successfully prepared by dissolving quartz powder in NaOH solution and depositing and growing in the pores of eucalyptus wood under hydrothermal conditions using eucalyptus wood as a templating agent. The experimental results showed the prepared materials have a loose, porous slit pore structure and many active adsorption sites. The adsorption efficiency of methylene blue was high, reaching more than 85% within 10 min, and the maximum adsorption amount was 90.01 mg/g. The adsorption process was by the pseudo-first-order,pseudo-second-order, and Langmuir models. The analysis of thermodynamic data showed that the adsorption of methylene blue by PSAM was a heat-absorbing process and spontaneous. Therefore, PSAM can be effectively used for the application of methylene blue dye removal in water.
Content may be subject to copyright.
Page 1/27
Preparation of porous silica materials using a
eucalyptus template method and its ecient
adsorption of methylene blue
Wenxin Zhu
Nanning Normal University
Leping Liu ( 130078@nnnu.edu.cn )
Nanning Normal University
YuanXia Lao
Nanning Normal University
Yan He
Guangxi University
Research Article
Keywords: Quartz powder, Eucalyptus, Adsorption, Organic dye
Posted Date: April 26th, 2023
DOI: https://doi.org/10.21203/rs.3.rs-2844761/v1
License: This work is licensed under a Creative Commons Attribution 4.0 International License. 
Read Full License
Page 2/27
Abstract
Mesoporous silica has become one of the primary adsorbent materials for solving dye wastewater
pollution due to its high specic surface area and good adsorption properties. However, the high cost of
the traditional chemical synthesis method limits its wide application. In this thesis, low-cost and high-
eciency porous silica adsorbent materials (PSAM) were successfully prepared by dissolving quartz
powder in NaOH solution and depositing and growing in the pores of eucalyptus wood under
hydrothermal conditions using eucalyptus wood as a templating agent. The experimental results showed
the prepared materials have a loose, porous slit pore structure and many active adsorption sites. The
adsorption eciency of methylene blue was high, reaching more than 85% within 10 min, and the
maximum adsorption amount was 90.01 mg/g. The adsorption process was by the pseudo-rst-order
pseudo-second-order, and Langmuir models. The analysis of thermodynamic data showed that the
adsorption of methylene blue by PSAM was a heat-absorbing process and spontaneous. Therefore,
PSAM can be effectively used for the application of methylene blue dye removal in water.
1. Introduction
Water is a vital material resource for human health and development. Water pollution from industrial
production is an increasingly severe threat to freshwater resources, causing toxic damage to plants,
animals, and humans. Industrial organic dyes are one of the leading causes of surface and groundwater
pollution. Organic dyes typically have complex structures and are not readily biodegradable. Most contain
elements such as nitrogen and sulfur, which are highly toxic and carcinogenic [1, 2]. One such dye is
methylene blue (MB). This cationic thiazine dye can remain in aquatic systems long if left untreated,
endangering marine organisms' survival and destroying the ecosystem's balance [3]. Industrial
wastewater must be effectively treated before discharge to reduce pollution of inowing water sources.
Current methods for treating industrial wastewater and organic dyes can be classied into physical,
biological, and chemical treatments. Commonly used methods include electrochemical degradation,
photocatalytic degradation, Fenton oxidation, adsorption, and biological anaerobic degradation [4–6].
Each method has its advantages and disadvantages. The adsorption method has received extensive
research attention owing to its low cost, convenience, low energy consumption, and low temperature and
pressure requirements [7].
In recent years, the commonly used adsorbents for treating organic dyes in wastewater by adsorption
include commercial activated carbon, but commercial activated carbon has a high cost, which limits its
application [8]. So other materials have better research and application in adsorption, such as biomass
carbon materials and nano metal oxides, which are widely used in dye adsorption and degradation.
Biomass carbon materials (e.g., lychee shell, sugarcane peel, wood chips, etc.) and nano oxides (e.g.,
TiO2, ZnO, SiO2, etc.) possess the advantages of large specic surface area and pores, which make them
excellent adsorbent materials. [9–13].
Page 3/27
Mesoporous silica is usually prepared using the sol-gel method with surfactants as soft templating
agents; the high preparation cost limits its large-scale application. Karine et al. [14] prepared
mesostructured silica materials via a synergistic self-assembly and liquid crystal templating mechanism
using polyoxyethylene uoroalkyl ether and quaternary ammonium surfactants as substrates. Su et al.
[15] used ZnO nanorod arrays as macroporous templates and CTAB as a mesoporous template to
synthesize graded mesoporous silica lms. However, most surfactants are petrochemicals, which are
costly and do not meet the requirements of a low-carbon environment. Current applications of
mesoporous silica include the adsorption of heavy metal ions and organic pollutants. Zhang et al. [16]
used an improved oil–water biphasic layered coating to synthesize silica with a brous structure. The
resulting particles were modied with CPBA and studied as cis-diol adsorbents for adsorption
performance. The simple, green, and ecient preparation of mesoporous silica materials with good
adsorption properties is still a problem worthy of further investigation.
Eucalyptus is the world's most widely distributed hardwood tree and a relatively abundant timber resource
in Guangxi [17]. Due to its low cost and abundance, it is commonly used in papermaking and building
materials for home use and as a carbon source for fuel and other materials [18–20]. Eucalyptus wood
has many pore structures, and some lignin is dissolved under robust alkaline solutions. The pore size
increases, providing a suitable environment for SiO2 deposition. Quartz has the most stable and
abundant crystal structure of SiO2. Quartz powder can be quickly dissolved in a NaOH solution in a
hydrothermal environment at a specic temperature and pressure.
This paper aims to obtain a low-cost and ecient mesoporous silica adsorbent material using eucalyptus
wood as a templating agent. Under hydrothermal conditions, quartz powder is rapidly dissolved in NaOH
solution and deposited and grown in the pores of eucalyptus wood after calcination to remove the
templating agent. Moreover, the prepared mesoporous silica material showed highly ecient adsorption
performance for methylene blue dye in water. In addition, the lignin-containing waste solution and the
precipitated SiO2 can be recycled as a bio-binder after concentration. For example, Juan et al. [21]
prepared high-density berboard (HDF) from wheat straw under alkaline conditions using lignin as a
natural adhesive, Yang et al. [22] rst Lignin was catalyzed by sodium hydroxide and phenolized by
phenol at high temperature. Lignin was washed by ether and precipitated by acid to prepare phenolic
resin by replacing part of the phenol with this product. In the present study, a report on preparing
mesoporous silica using eucalyptus wood as a template has yet to be prepared. This thesis focuses on
the structure of PSAM and its adsorption performance and adsorption mechanism on methylene blue dye
to provide a new idea for preparing and applying ecient dye adsorbents.
2. Materials And Methods
2.1. materials
The quartz powder in this experiment was purchased from Chengde Group Company in Beihai, Guangxi,
and the eucalyptus wood was from Chongzuo City, Guangxi. Sodium hydroxide is an analytically pure
Page 4/27
reagent purchased from Chongqing Chuandong Chemical Group Co. Methylene blue is also an
analytically pure reagent, purchased from Tianjin Zhiyuan Chemical Reagent Co. Deionized water was
homemade in the laboratory and used for the preparation of chemical solutions. The XRD spectrum of
raw quartz powder is shown in Fig.1.
2.2 Synthesis of PSAM
Sodium hydroxide (30 g) and quartz powder (15 g) with a mass ratio of 2:1 were mixed and stirred well,
and the concentrations of sodium hydroxide were 2 mol/L (2M), 4 mol/L (4M), 6 mol/L (6M) and 8 mol/L
(8M), respectively. After the hydrothermal treatment, the waste solution and precipitate were concentrated
and recovered for the bioadhesive. The eucalyptus slices were sonicated for 5 min in an ultrasonic
cleaner with an operating frequency of 37 kHz three times. 80°C oven drying and calcination at 600°C for
2 h. The calcined material was ground to obtain a black powder and collected for subsequent testing.
2.3 Characterization
The crystal structures of the prepared samples were characterized using X-ray diffraction (XRD, Rigaku
MiniFlex 600 diffractometer, Japan) with Cu Kα radiation at a scanning rate of 5°/min at a voltage of 40
kV and a current of 15 mA at 5–80°. The structure of the substances was analyzed by recording Fourier
transform infrared spectrograms (FTIR, Thermo Scientic Nicolet IS50, USA) using the KBr particle
method. Scanning electron microscopy (SEM) images of the studied catalysts were acquired on a eld
emission scanning electron microscope (FE-SEM, Hitachi SU8220, Japan) device. The specic surface
area and the prepared samples' pores were tested by adsorption and desorption in N2 with a surface area
porosity analyzer (BET, Micromeritics ASAP2460, USA). The samples were degassed at 300°C for 6 h
before measurement. The zeta potential measurements were carried out using a nanoparticle analyzer
(Horiba, SZ-100, Japan).
2.4 Adsorption performance
The adsorption kinetics were performed at room temperature (20 ± 2°C) in the dark. Adsorbent (50 mg)
was added to 50 mL of a 20 mg/L MB solution. The supernatant was shaken at 230 rpm and ltered
through a needle lter (4 mL) at predetermined time intervals before measurement with a double-beam
UV spectrophotometer at a wavelength of 664 nm.
The adsorption isotherms were also obtained at room temperature (20 ± 2°C) in the dark. Adsorption
equilibrium was reached by adding 50 mg of adsorbent to 50 mL of MB solution at different
concentrations (20–60 mg) and shaking at 230 rpm for 24 h. The supernatant (4 mL) was ltered and
measured at 664 nm using a double-beam UV-Vis spectrophotometer (Shimadzu UV2600, Japan).
Isothermal adsorption curves were tted from Langmuir (5) and Freundlich (6) models.
The thermodynamics of attachment was measured at 20°C, 30°C, and 40°C by adding 50 mg of
adsorbent to a solution of MB at a concentration of 20 mg/L. After shaking at 230 rpm for 24 h, 4 mL of
the supernatant was ltered, and the absorbance was measured at a wavelength of 664 nm using a dual-
Page 5/27
beam UV-Vis spectrophotometer (Shimadzu UV2600, Japan). The thermodynamic parameters were
obtained by tting Eqs.(8), (9), and (10).
The adsorbent's effect on MB solution's adsorption was investigated at different temperatures (20–60)
and pH (pH = 2–8).
2.4 Related formulas
The adsorption amount (qt, mg/g) and equilibrium adsorption capacity (qe, mg/g) of MB for the samples
at different adsorption times (t) are expressed as
1
2
where C0 is the starting concentration of the adsorbent solution (MB solution) (mg/L); Ct is the
concentration of the adsorbent (MB solution) at a particular moment t (mg/L); Ce is the concentration of
the adsorbent (MB solution) at the moment of equilibrium (mg/L); V is the volume of the adsorbent
solution (MB solution) (L), and W is the mass of the adsorbent (g).
· Adsorption kinetic
Pseudo-rst order [23] (3)
Pseudo-second order [24] (4)
where k1 is the adsorption rate constant (min− 1) for the pseudo-rst-order model; k2 is the adsorption rate
constant (g/mg/min) for the pseudo-second-order model, and qe and qt are the MB uptakes (mg/g) at
equilibrium and at time t (min), respectively.
·Adsorption isotherm
qt= (C0-Ct)V
m
qe= (C0-Ce)V
m
qt=qe(1-e-k1t)
qt=k2qet
1+k2qet
Page 6/27
Freundlich [25] (5)
Langmuir [26] (6)
(7)
where KL (L/mg) and KF [(mg/g)/(mg/L)n] represent the Langmuir and Freundlich adsorption coecients,
respectively; n is the exponential coecient, and RL (L/mg) is the Langmuir constant.
·Adsorption thermodynamics [13]
(8)
(9)
(10)
where KL is the partition coecient, S (J/K mol) is the standard entropy change; H (kJ/mol) is the
standard enthalpy; G (kJ/mol) is the standard free energy change; R is the ideal gas constant (8.314
J/mol/ K), and T is the temperature (K).
3. Results And Discussion
3.1. Characterization
3.1.1 XRD
Figure 2 shows the XRD patterns of specimens hydrothermally heated at 120°C for 24 h with different
NaOH solution and quartz powder concentrations. Peaks at 2θ = 20.86°, 26.64°, and 50.05° were observed
for all specimens with complete coincidence with the characteristic peaks of quartz crystals (PDF#89-
1961), corresponding to crystallographic planes (100), (011), and (112). As the NaOH concentration
increased, the intensity of the characteristic peaks of the quartz crystals decreased. After the
hydrothermal reaction, the surface of the quartz powder particles reacted with sodium hydroxide solution
to form many silica hydroxyl groups. Quartz particles with silica hydroxyl groups on the surface were
deposited in the pore structure of eucalyptus chips via a condensation reaction between them. The PSAM
was obtained after calcination to remove the template. With an increase in the NaOH solution
concentration, many non-bridging oxygen active sites were formed in the pore channels of PSAM to
balance its charge with Na+. The calcined PSAM formed Na2CO3 crystals after the adsorption of CO2
logqe= logKF+ logCe
1
n
= +
Ce
qe1
KLqm
Ce
qm
RL=1
1+RL+C0
KL= mqe
Ce
ΔG = -RTlnKL
lnKL= +
ΔS
R-ΔH
R
Page 7/27
from the air. With NaOH concentrations of 2M and 4M, Na2CO3 crystals were not detected in the XRD
spectra; the presence of Na2CO3 crystals was determined from later FTIR analysis. With NaOH
concentrations of 6M and 8M, characteristic peaks of Na2CO3 were observed at 2θ = 20.06°, 34.19°,
35.16°, 37.93°, 39.85°, 40.99°, 46.28°, and 48.10° in the spectra of the PSAM specimens, in addition to the
characteristic peaks of quartz crystals (PDF#72–0628).
3.1.2 FTIR
Figure 3 shows the FTIR spectra of wood specimens treated with different concentrations of NaOH
solutions. The gure shows that all four materials correspond to the anti-integrated stretching vibration of
H-O-H at 3440 cm− 1 and the bending vibration peak of O-H at 1590 cm− 1. [27]. As can be seen from the
gure, the characteristic peaks are weak mainly because the water is removed from PSAM after
calcination at 600°C. The PSAM specimens correspond to the antisymmetric stretching vibration of Si-O-
Si and the bending vibration of Si-O in quartz crystals at 1090 cm− 1 and 461 cm− 1, respectively. The
characteristic peaks of quartz crystals gradually weakened with increasing NaOH concentration, which is
consistent with the trend of the characteristic peaks of quartz crystals in the XRD patterns. The
specimens corresponded to the antisymmetric stretching vibration and symmetric stretching vibration of
CO32− at 1440 cm− 1 and 877 cm− 1 [28, 29]. Respectively, the peak of the stretching vibration of CO32−
gradually increased with increasing NaOH concentration, indicating that the Na2CO3 crystals gradually
increased in PSAM, which is consistent with the XRD results.
3.1.3 SEM
Figure 4 shows SEM images of the materials treated with different NaOH concentrations after
magnication at different multiples. It is observed in the gure that all four materials have a very loose
internal structure with many widely spaced voids, providing more active sites for dye adsorption. The pore
size of the PSAM decreased with increasing NaOH concentration. The NaOH solution had two leading
roles in preparing the materials: (1) dissolving part of the lignin in the pore channels of eucalyptus chips
and (2) reacting with the surface of quartz particles. The surface hydroxylated particles were condensed
and deposited in the pore channels of eucalyptus chips. From the microstructural analysis, the second
effect was dominant; the number of quartz particles deposited into the pore channels by dissolution–
condensation increased with increasing NaOH concentration, and the pore size formed after calcination
to remove the template were smaller. Analysis of XRD and FTIR data indicates the mechanism of action
and microstructure change pattern of the NaOH solution during the preparation of the materials. However,
all four materials generally had more pores and better adsorption performance for MB.
3.1.4 BET
Page 8/27
Table 1
Surface area and pore characteristics of PSAM
samples.
Samples SBET(m2/g) Vtotal(cm3/g) Dave(nm)
2M 88.05 0.082 3.707
4M 69.17 0.065 3.786
6M 54.07 0.041 3.017
8M 61.75 0.041 2.645
Figure 5(a) shows the four materials' N2 adsorption–desorption isotherms. According to the IUPAC
classication, the prepared materials' N2 adsorption–desorption isotherms are type II isotherms with an
H3 lagging loop [12]. As observed in the gure, the relative pressure increased gently with increasing
pressure between 0 and 0.8. When the relative pressure was more signicant than 0.8, the adsorption
increased rapidly with increasing pressure, mainly due to capillary condensation in the mesopores and
macropores. Different types of hysteresis loops represent different types of pore structures. The H3
hysteresis loop in the gure indicates the presence of mesopores and large slit pores in the prepared
material [30]. Figure 5(b) shows the pore size distribution of the four materials; the range of the pore
centers of the four materials was between 3–4 nm, and the pore size was approximately 3.7 nm. Table 1
shows the four materials' specic surface area and pore size; the specic surface area, cumulative pore
volume, and pore size showed an overall decreasing trend with increasing sodium hydroxide
concentration. Formation of the pore size of PSAM is based on two factors: (1) the surface hydroxylation
of quartz particles and the residual pores of the unlled eucalyptus pore channels formed by the
condensation reaction deposited in the pore channels of eucalyptus chips and the pores between the
piles of particles, and (2) the slit pores left by the removal of the eucalyptus chip template. When the
NaOH concentration is low, there are fewer hydroxyl groups on the surface of the quartz particles; the
quartz deposited in the pore channels is not lled, and there are pores between the pile of particles. The
eucalyptus pore channels dissolve less under the action of NaOH solution, and more slit holes remain
after calcination. When the concentration of NaOH increased, the surface of the quartz particles dissolved
and produced a large number of hydroxyl groups, and the amount deposited in the structure of
eucalyptus pore channels through the condensation reaction increased. The accumulation between
particles was denser, whereas the pore walls of the eucalyptus chips were thinner under the action of a
high NaOH concentration. Smaller slit holes were formed after calcination. Overall, the specic surface
areas and pore sizes of all four materials were not signicantly different, and they showed good
adsorption properties for MB. The variation pattern of the pore structure can be determined from
microstructure analysis.
3.1.5 Zeta potential
Page 9/27
Figure 6. As the pH of the solution increases from 2 to 10, the zeta potential of PSAM shows a decreasing
trend, and the value changes from positive to negative. The data in the gure shows that the zero point
charges (pHpzc) of PSAM are 2.12, 2.09, 2.11, 2.12. Since MB is a cationic dye, when the pH of the
solution is lower than the pHpzc value (pH < pHpzc), the surface of PSAM is positively charged, which is
not favorable for the adsorption of MB; on the contrary, when the pH of the solution is higher than the
pHpzc value (pH > pHpzc), the surface of PSAM is negatively charged, which will be favorable for MB
adsorption [31].
3.2. Adsorption of MB
Figure 7(a) shows the adsorption rates of different materials on MB. It can be observed from the gure
that the adsorption rates of the four prepared materials on MB increased sharply in the rst 10 min, and
the adsorption amount could reach 17.64 mg/g in the 10th min. With the increase of time, the adsorption
rates increased slowly and gradually stabilized to reach the equilibrium, and the adsorption amount at the
equilibrium was 20 mg/g. When the NaOH concentration was 2 M and 4 M, the pore size of PSAM was
more signicant, and there were more slit pores in When methylene blue adsorption was carried out, more
dyes entered the internal active site dyes through internal diffusion, in addition to those adsorbed on the
surface of PSAM. When the concentration of sodium hydroxide is 6M and 8M, the pore size of PSAM is
small, the slit pores are small, more active sites on the surface, and more dyes are adsorbed.
Meanwhile, due to the signicant molecular weight of dyes, the spatial site resistance is considerable,
and the amount of active sites entering inside and outside through internal diffusion is reduced. The total
active adsorption sites of PSAM prepared from different NaOH solutions were consistent; therefore, the
adsorption rates of methylene blue were the same. The trend analysis of PSAM adsorption rates was
consistent with the previous XRD, FTIR, and SEM analysis that PSAM is a porous material with abundant
active sites. The main reason for the different trends from the BET data is that the adsorption properties
of the dye are related to the active adsorption sites in the pore channels and the active adsorption sites
on the surface of the material. In this study, as the specic surface area and pore size decreased with the
increase of sodium hydroxide concentration, the amount of dye entering the internal pore channels
decreased due to the spatial location. However, the active adsorption sites on the material surface
increased, so the total adsorption rate was unchanged. Figure 7(b) shows the effect of different
temperatures on MB adsorption. With the increase in temperature, the adsorption rate of MB increased for
the four materials. The adsorption process is heat absorption, and the temperature increase accelerates
the binding rate of dye molecules and materials [32]. From the following thermodynamic analysis, this
speculation can be proved. Figure 7(c) shows the effect of different pH on the adsorption of MB, from
which it can be seen that the adsorption of MB was not satisfactory under highly acidic conditions (pH = 
2). However, between pH 4–10, PSAM was suitable for the adsorption of MB on the four materials
because the pH was higher than pHpzc, which was favorable for adsorption, indicating that the material
can be applied to the treatment of weakly acidic, neutral and alkaline dye wastewater, which is consistent
with the previous zeta potential analysis.
Page 10/27
3.2.1. Adsorption kinetics
Table 2
Kinetics parameters for MB Adsorption on PSAM samples.
Samples pseudo-rst-order pseudo-second-order
qe(mg g1) k1(min1) R2qe(mg/g) k2(mg1 g1 L1) R2
2M 19.77 0.83 0.998 19.88 0.32 0.998
4M 20.06 0.67 0.995 20.26 0.15 0.997
6M 19.49 0.71 0.994 19.66 0.18 0.996
8M 18.93 0.43 0.986 19.34 0.06 0.995
Figure 8 shows a t to the adsorption kinetics of PSAM on MB, where (a) is a pseudo-rst-order kinetic
model and (b) is a pseudo-second-order kinetic model. Table 2 presents the parameters of the kinetic ts
for the four materials. The kinetic t curves show that the adsorption of MB by the materials increased
gradually with time, with a faster increase in the rst 10 min, followed by a attening until equilibrium.
This is because the material provided more adsorption sites at the beginning of the adsorption process.
However, with increasing time, the adsorption sites on the material were occupied by dye molecules. With
an increase in adsorption time, the dye concentration decreased; the adsorption eciency decreased, and
the adsorption reached saturation [11]. The tted data show that the R2 values of the pseudo-rst-order
and pseudo-second-order kinetic models are close to 1; thus, both models apply to the adsorption of MB
by PSAM [33, 34]. In a previous study, the same adsorption kinetic trends were observed for the synthesis
of nano-silica-coated magnetic carbonaceous adsorbents for the adsorption of MB in water using a low-
temperature hydrothermal carbonization technique (HCT), indicating that the adsorption of MB by PSAM
is controlled by multiple processes and not by a single process [13, 35].
3.2.2. Adsorption isotherms
Page 11/27
Table 3
Adsorption isotherm parameters for MB Adsorption on PSAM samples.
Samples Langmuir Freundlich
qm
(mg g1)
KL
(L mg1)
RL
(L mg1)
R2n KF
(L mg1)1/n(mg g1)
R2
2M 90.01 1.56 0.03 0.981 1.74 55.74 0.946
4M 80.97 1.02 0.05 0.990 1.92 38.57 0.988
6M 64.52 0.49 0.09 0.983 2.48 23.19 0.977
8M 61.16 1.01 0.05 0.978 2.85 29.40 0.917
Table 4
Comparison of the adsorption capacity of MB by different adsorbents
Adsorbent qm(mg/g) Reference
Acrylated composite hydrogel (ACH) 56.61 [36]
Activated lignin-chitosan extruded pellets (ALiCE) 36.25 [4]
Carboxymethyl chitosan-modied magnetic-cored dendrimers (CCMDs) 96.31 [37]
Mesoporous silicon carbon (MSC) 156.56 [38]
Magnetic starch-based composite hydrogel microspheres (SCHMs) 88.33 [39]
Hydrophobic (surface modied) silica aerogel (MSA) 65.74 [40]
Paintosorp 44.64 [41]
Sago-grafted silica 80:20 10.31 [42]
PSAM 90.01 This work
Figure 9 shows the results of the PSAM t to the MB adsorption isotherm, where Fig. (a) and Fig. (b) are
the Langmuir model t curves, and Fig. (c) is the Freundlich model t curve. Table 3 shows the
parameters of the adsorption isotherm t. From the data of the graphs, it can be seen that the R2 of the
Langmuir model is close to 1, so the Langmuir model is more applicable to the adsorption of PSAM on
MB [43]. The Langmuir model applies to the adsorption of the monomolecular layer, so the adsorption of
PSAM on MB is of monomolecular layer adsorption. From the data in Table 3, the maximum adsorption
amounts of the four materials are not signicantly different. However, in comparison, the adsorption
amount of 2M PSAM reaches 90.01 mg/g, the most considerable adsorption amount among the four
Page 12/27
materials, consistent with the previous BET data. KL denotes the Langmuir constant, which indicates the
anity of the adsorbate for the binding site. The separation factor RL can be calculated by Eq. (7) when 0 
< RL < 1 favors the adsorption, as observed from the Langmuir t data, so all four materials prepared are
favorable for MB adsorption [44]. Moreover, the tted data from the Freundlich model also reveals that
the n values of all four materials are between 1 and 10, n is a parameter describing the adsorption
strength, and when 1  n  10, the adsorption is favored, so all the four prepared PSAMs show good
adsorption on MB [34, 45]. Meanwhile, the maximum adsorption of different adsorbents on MB dyes is
listed in Table 4, which shows that PSAM can effectively remove organic dyes from water.
3.2.3. Adsorption thermodynamics
Table 5
Various thermodynamics parameters for the Adsorption of MB onto 2M PSAM.
Samples G0(kJ mol1)H0(kJ mol1)S0(kJ mol1 K1) R2
2M -1.128 21.364 76.765 0.946
Adsorption thermodynamics is an essential basis for studying whether adsorption occurs spontaneously.
Figure 10 shows the lnKL-1/T curve for 2M PSAM; the relevant thermodynamic parameters can be
calculated from the curves in the gure, as shown in Table 5. From the data in the table, H > 0 indicates
that the adsorption of PSAM on MB is heat-absorbing. If the temperature of the reaction is increased, the
adsorption rate of PSAM on MB can be enhanced, which is consistent with the effect of temperature on
PSAM adsorption. Based on a value of H between 0 kJ mol-1 and 84 kJ mol-1, it can be judged that the
adsorption is mainly a physical adsorption process [46]. A positive value of S indicates that the
adsorption process is disorderly, increasing at the solid–liquid interface. The value of G is less than 0,
indicating that the adsorption of MB by PSAM is spontaneous [47, 48].
3.3. Adsorption mechanism of MB on PSAM
The adsorption effect of adsorbents on adsorbates is not only determined by the physical structure and
chemical properties of the adsorbent material itself, such as the specic surface area and pore size of the
material but also by other inuences, such as the interaction between the adsorbent and the adsorbate
and the charge on the adsorbent surface [13]. When adsorption occurs, the adsorbent and adsorbent are
bound in different ways, usually through ion exchange, hydrogen bonding, electrostatic interactions,
dipole-dipole interactions, hydrophobic interactions, and surface metal cation coordination between
adsorbent and adsorbent. Figure 11 shows the FTIR spectra of 2M PSAM before and after the adsorption
of MB. It can be observed that the peaks at 1440 cm− 1 and 877 cm− 1 disappear before and after the
adsorption, and the peaks at these two locations are the characteristic peaks of CO32−, thus indicating
that MB+ can be adsorbed with CO32− by electrostatic gravitation. At the same time, the characteristic
peaks of SiO2 also changed slightly, indicating that the material's structure did not change before and
Page 13/27
after adsorption, but MB+ could also bind to SiO2 by electrostatic interaction [13, 49]. Therefore, the dye
molecules are adsorbed on the surface of PSAM by electrostatic interaction. Moreover, PSAM can also
bind to the N atom in MB by hydrogen bonding [3, 13]. The adsorption mechanism of MB is shown in
Scheme 1. Moreover, the added wood is structurally increasing the specic surface area of the material
and providing more dye adsorption sites for the material itself.
4. Conclusion
This study successfully prepared a low-cost and high-eciency porous silica adsorbent material via a
hydrothermal synthesis method using quartz powder as the raw material and eucalyptus wood chips as
the template with the template different concentrations of sodium hydroxide. The prepared material was
used for the adsorption of MB; the experimental results are presented as follows:
1. Sodium hydroxide reacted with the surface of quartz particles. Hydroxylated quartz particles were
deposited in the pore channels of eucalyptus wood chips through a condensation reaction, and pore
channels were formed between the stacked quartz particles. After removing the template, sodium
hydroxide dissolved part of the lignin in the pore channels and formed slit pores.
2. As the concentration of NaOH increased, the characteristic peaks of quartz crystals of PSAM
weakened; the number of active sites formed on the surface increased, and the internal pore size
decreased.
3. Adsorption of MB by the four PSAMs was divided into surface and intra-pore adsorption. The pore
sizes of the PSAM materials prepared with 2M and 4M sodium hydroxide were larger, indicating
mainly intra-pore adsorption. The pore sizes of the PSAM materials prepared with 6M and 8M
sodium hydroxide were smaller. Diffusion of MB into the pores was reduced owing to the spatial site
resistance, indicating mainly surface adsorption. An adsorption rate of 85% was achieved in 10 min
for the four PSAM materials, all of which showed good adsorption performance and rate.
4. The adsorption kinetics and isotherms were tted for the four materials. The tted data shows that
they all t the pseudo-rst-order kinetic, pseudo-second-order kinetic, and Langmuir models, which
indicates that a single process controls the adsorption of MB by PSAM and that the adsorption is a
single molecular layer adsorption. Thermodynamic data show that the adsorption of MB by PSAM is
a spontaneous heat-absorbing process. PSAM has potential applications due to its low preparation
cost, wide availability of raw materials, and high eciency in the adsorption of MB from wastewater.
Environmental Implications
A new porous silica adsorbent material was synthesized and tested for its adsorption performance on
methylene blue using eucalyptus wood as a template as a simulated pollutant for industrial wastewater.
The tests showed that the porous material has high adsorption eciency for methylene blue in water and
can reach more than 85% adsorption rate in ten minutes. Moreover, the adsorption process is
spontaneous. The material can effectively remove toxic organic dyes in water.
Page 14/27
Declarations
CRediT authorship contribution statement
Wenxin Zhu: Conceptualization, Methodology, Software, Validation, Investigation, Data curation, Writing -
original draft; Leping Liu: Conceptualization, Supervision, Writing - review & editing; YuanXia Lao:
Investigation, Writing - review & editing. Yan He: Conceptualization, Supervision, Writing - review & editing.
Declaration of Competing Interest
The authors declare that they have no known competing nancial interests or personal relationships that
could have appeared to inuence the work reported in this paper.
Data availability
Data will be made available on request.
Acknowledgments
This work was supported by the National Natural Science Foundation of China (51962024).
References
1. X. Liu, J. Tian, Y. Li, et al., Enhanced dyes adsorption from wastewater via Fe3O4 nanoparticles
functionalized activated carbon. Journal of Hazardous Materials. 373, 397-407, (2019).
https://doi.org/https://doi.org/10.1016/j.jhazmat.2019.03.103.
2. Y. Yu, N. Qiao, D. Wang, et al., Fluffy honeycomb-like activated carbon from popcorn with high
surface area and well-developed porosity for ultra-high eciency adsorption of organic dyes.
Bioresource Technology. 285, 121340, (2019).
https://doi.org/https://doi.org/10.1016/j.biortech.2019.121340.
3. C. Djilani, R. Zaghdoudi, F. Djazi, et al., Adsorption of dyes on activated carbon prepared from apricot
stones and commercial activated carbon. Journal of the Taiwan Institute of Chemical Engineers. 53,
112-121, (2015). https://doi.org/https://doi.org/10.1016/j.jtice.2015.02.025.
4. A. B. Albadarin, M. N. Collins, M. Naushad, et al., Activated lignin-chitosan extruded blends for
ecient adsorption of methylene blue. Chemical Engineering Journal. 307, 264-272, (2017).
https://doi.org/https://doi.org/10.1016/j.cej.2016.08.089.
5. S. Mahalingam, J. Ramasamy and Y.-H. Ahn, Enhanced Photocatalytic Degradation of Synthetic
Dyes and Industrial Dye Wastewater by Hydrothermally Synthesized G–CuO–Co3O4 Hybrid
Page 15/27
Nanocomposites Under Visible Light Irradiation. Journal of Cluster Science. 29(2), 235-250, (2018).
https://doi.org/10.1007/s10876-017-1329-3.
. T. Ochiai, Y. Iizuka, K. Nakata, et al., Ecient electrochemical decomposition of peruorocarboxylic
acids by the use of a boron-doped diamond electrode. Diamond and Related Materials. 20(2), 64-67,
(2011). https://doi.org/https://doi.org/10.1016/j.diamond.2010.12.008.
7. W. Xiang, X. Zhang, J. Chen, et al., Biochar technology in wastewater treatment: A critical review.
Chemosphere. 252, 126539, (2020).
https://doi.org/https://doi.org/10.1016/j.chemosphere.2020.126539.
. M. Rafatullah, O. Sulaiman, R. Hashim, et al., Adsorption of methylene blue on low-cost adsorbents: A
review. Journal of Hazardous Materials. 177(1), 70-80, (2010).
https://doi.org/https://doi.org/10.1016/j.jhazmat.2009.12.047.
9. L. G. da Silva, R. Ruggiero, P. d. M. Gontijo, et al., Adsorption of Brilliant Red 2BE dye from water
solutions by a chemically modied sugarcane bagasse lignin. Chemical Engineering Journal. 168(2),
620-628, (2011). https://doi.org/https://doi.org/10.1016/j.cej.2011.01.040.
10. V. S. Mane and P. V. Vijay Babu, Kinetic and equilibrium studies on the removal of Congo red from
aqueous solution using Eucalyptus wood (Eucalyptus globulus) saw dust. Journal of the Taiwan
Institute of Chemical Engineers. 44(1), 81-88, (2013).
https://doi.org/https://doi.org/10.1016/j.jtice.2012.09.013.
11. H. Mittal and S. S. Ray, A study on the adsorption of methylene blue onto gum ghatti/TiO2
nanoparticles-based hydrogel nanocomposite. International Journal of Biological Macromolecules.
88, 66-80, (2016). https://doi.org/10.1016/j.ijbiomac.2016.03.032.
12. H. Mittal, S. M. Alhassan and S. S. Ray, Ecient organic dye removal from wastewater by magnetic
carbonaceous adsorbent prepared from corn starch. Journal of Environmental Chemical Engineering.
6(6), 7119-7131, (2018). https://doi.org/https://doi.org/10.1016/j.jece.2018.11.010.
13. H. Mittal, R. Babu, A. A. Dabbawala, et al., Low-Temperature Synthesis of Magnetic Carbonaceous
Materials Coated with Nanosilica for Rapid Adsorption of Methylene Blue. ACS Omega. 5(11), 6100-
6112, (2020). https://doi.org/10.1021/acsomega.0c00093.
14. K. Assaker, M.-J. Stebe and J.-L. Blin, Mesoporous Mesoporous silica materials from diluted and
concentrated solutions of nonionic uorinated and ionic hydrogenated surfactants mixtures. Colloids
and Surfaces a-Physicochemical and Engineering Aspects. 536, 242-250, (2018).
https://doi.org/10.1016/j.colsurfa.2017.02.056.
15. X. Su, J. Tao, S. Chen, et al., Uniform hierarchical silica lm with perpendicular macroporous
channels and accessible ordered mesopores for biomolecule separation. Chinese Chemical Letters.
30(5), 1089-1092, (2019). https://doi.org/10.1016/j.cclet.2019.01.022.
1. W. Zhang, S. Li, J. Zhang, et al., Synthesis and adsorption behavior study of magnetic brous
mesoporous silica. Microporous and Mesoporous Materials. 282, 15-21, (2019).
https://doi.org/10.1016/j.micromeso.2019.03.022.
Page 16/27
17. S. Suganya and P. S. Kumar, Evaluation of environmental aspects of brew waste-based carbon
production and its disposal scenario. JOURNAL OF CLEANER PRODUCTION. 202, 244-252, (2018).
https://doi.org/10.1016/j.jclepro.2018.08.143.
1. K. Elyounssi, F. X. Collard, J. A. N. Mateke, et al., Improvement of charcoal yield by two-step pyrolysis
on eucalyptus wood: A thermogravimetric study. FUEL. 96(1), 161-167, (2012).
https://doi.org/10.1016/j.fuel.2012.01.030.
19. G. Kumar, S. Shobana, W.-H. Chen, et al., A review of thermochemical conversion of microalgal
biomass for biofuels: chemistry and processes. Green Chemistry. 19(1), 44-67, (2017).
https://doi.org/10.1039/C6GC01937D.
20. M. E. Vallejos, J. Kruyeniski and M. C. Area, Second-generation bioethanol from industrial wood
waste of South American species. Biofuel Research Journal. 4(3), 654-667, (2017).
https://doi.org/10.18331/BRJ2017.4.3.4.
21. J. Dominguez-Robles, Q. Tarres, M. Delgado-Aguilar, et al., Approaching a new generation of
berboards taking advantage of self lignin as green adhesive. International Journal of Biological
Macromolecules. 108, 927-935, (2018). https://doi.org/10.1016/j.ijbiomac.2017.11.005.
22. S. Yang, J.-L. Wen, T.-Q. Yuan, et al., Characterization and phenolation of biorenery technical lignins
for lignin-phenol-formaldehyde resin adhesive synthesis. Rsc Advances. 4(101), 57996-58004,
(2014). https://doi.org/10.1039/c4ra09595b.
23. S. K. Lagergren, About the theory of so-called Adsorption of soluble substances. Kungliga Svenka.
Vetenspsakademiens Handlingar. 24, 1-39, (1898). https://doi.org/.
24. Y. S. Ho and G. McKay, Sorption of dye from aqueous solution by peat. Chemical Engineering
Journal. 70(2), 115-124, (1998). https://doi.org/https://doi.org/10.1016/S0923-0467(98)00076-1.
25. H. Freundlich, Über die adsorption in L¨osungen. Ztschrift Phys. Chemistry. 57(1), 385–470, (1907).
https://doi.org/.
2. I. Langmuir, THE ADSORPTION OF GASES ON PLANE SURFACES OF GLASS, MICA AND PLATINUM.
Journal of the American Chemical Society. 40(9), 1361-1403, (1918).
https://doi.org/10.1021/ja02242a004.
27. X. Sun, L. Yang, T. Dong, et al., Removal of Cr(VI) from aqueous solution using amino-modied
Fe3O4-SiO2-chitosan magnetic microspheres with high acid resistance and adsorption capacity.
Journal of Applied Polymer Science. 133(10), (2016). https://doi.org/10.1002/app.43078.
2. P. Yu, R. J. Kirkpatrick, B. Poe, et al., Structure of calcium silicate hydrate (C-S-H): Near-, mid-, and far-
infrared spectroscopy. Journal of the American Ceramic Society. 82(3), 742-748, (1999).
https://doi.org/.
29. M. A. Trezza and A. E. Lavat, Analysis of the system 3CaO·Al2O3–CaSO4·2H2O–CaCO3–H2O by FT-
IR spectroscopy. Cement and Concrete Research. 31(6), 869-872, (2001).
https://doi.org/https://doi.org/10.1016/S0008-8846(01)00502-6.
30. M. Thommes, K. Kaneko, A. V. Neimark, et al., Physisorption of gases, with special reference to the
evaluation of surface area and pore size distribution (IUPAC Technical Report). Pure and Applied
Page 17/27
Chemistry. 87(9-10), 1051-1069, (2015). https://doi.org/10.1515/pac-2014-1117.
31. C. Ni, J. Wang, Y. Guan, et al., Self-powered peroxi-coagulation for the ecient removal of p-arsanilic
acid: pH-dependent shift in the contributions of peroxidation and electrocoagulation. Chemical
Engineering Journal. 391, 123495, (2020). https://doi.org/https://doi.org/10.1016/j.cej.2019.123495.
32. N. Hassan, A. Shahat, I. M. El-Deen, et al., Synthesis and characterization of NH2-MIL-88(Fe) for
ecient adsorption of dyes. Journal of Molecular Structure. 1258, (2022).
https://doi.org/10.1016/j.molstruc.2022.132662.
33. Y. S. Ho and G. McKay, A Comparison of Chemisorption Kinetic Models Applied to Pollutant Removal
on Various Sorbents. Process Safety and Environmental Protection. 76(4), 332-340, (1998).
https://doi.org/https://doi.org/10.1205/095758298529696.
34. T. A. Khan, S. Dahiya and I. Ali, Use of kaolinite as adsorbent: Equilibrium, dynamics and
thermodynamic studies on the adsorption of Rhodamine B from aqueous solution. Applied Clay
Science. 69, 58-66, (2012). https://doi.org/10.1016/j.clay.2012.09.001.
35. Y. Li, A. R. Zimmerman, F. He, et al., Solvent-free synthesis of magnetic biochar and activated carbon
through ball-mill extrusion with Fe3O4 nanoparticles for enhancing adsorption of methylene blue.
Science of the Total Environment. 722, (2020). https://doi.org/10.1016/j.scitotenv.2020.137972.
3. J. Wang, X. Meng, Z. Yuan, et al., Acrylated Composite Hydrogel Preparation and Adsorption Kinetics
of Methylene Blue. Molecules. 22(11), 1824, (2017).
https://doi.org/https://doi.org/10.3390/molecules22111824.
37. H.-R. Kim, J.-W. Jang and J.-W. Park, Carboxymethyl chitosan-modied magnetic-cored dendrimer as
an amphoteric adsorbent. Journal of Hazardous Materials. 317, 608-616, (2016).
https://doi.org/https://doi.org/10.1016/j.jhazmat.2016.06.025.
3. S. Zhang, Y. Gao, H. Dan, et al., Effect of washing conditions on adsorptive properties of mesoporous
silica carbon composites by in-situ carbothermal treatment. Science of The Total Environment. 716,
136770, (2020). https://doi.org/https://doi.org/10.1016/j.scitotenv.2020.136770.
39. W. Li, H. Wei, Y. Liu, et al., Fabrication of novel starch-based composite hydrogel microspheres
combining Diels-Alder reaction with spray drying for MB adsorption. Journal of Environmental
Chemical Engineering. 9(5), 105929, (2021).
https://doi.org/https://doi.org/10.1016/j.jece.2021.105929.
40. H. Han, W. Wei, Z. Jiang, et al., Removal of cationic dyes from aqueous solution by adsorption onto
hydrophobic/hydrophilic silica aerogel. Colloids and Surfaces A: Physicochemical and Engineering
Aspects. 509, 539-549, (2016). https://doi.org/https://doi.org/10.1016/j.colsurfa.2016.09.056.
41. Momina, S. Mohammad and I. Suzylawati, Study of the adsorption/desorption of MB dye solution
using bentonite adsorbent coating. Journal of Water Process Engineering. 34, 101155, (2020).
https://doi.org/https://doi.org/10.1016/j.jwpe.2020.101155.
42. Y. Rajan, Z. Ngaini and R. Wahi, Novel adsorbent from sago-grafted silica for removal of methylene
blue. International Journal of Environmental Science and Technology. 16(8), 4531-4542, (2019).
https://doi.org/10.1007/s13762-018-2043-x.
Page 18/27
43. K. Y. Foo and B. H. Hameed, Insights into the modeling of adsorption isotherm systems. Chemical
Engineering Journal. 156(1), 2-10, (2010). https://doi.org/10.1016/j.cej.2009.09.013.
44. Q. Wang, Z. Lai, C. Luo, et al., Honeycomb-like activated carbon with microporous nanosheets
structure prepared from waste biomass cork for highly ecient dye wastewater treatment. Journal of
Hazardous Materials. 416, (2021). https://doi.org/10.1016/j.jhazmat.2021.125896.
45. H. Mittal, V. Kumar, Saruchi, et al., Adsorption of methyl violet from aqueous solution using gum
xanthan/Fe304 based nanocomposite hydrogel. International Journal of Biological Macromolecules.
89, 1-11, (2016). https://doi.org/10.1016/j.ijbiomac.2016.04.050.
4. H. Li, F. Zheng, J. Wang, et al., Facile preparation of zeolite-activated carbon composite from coal
gangue with enhanced adsorption performance. Chemical Engineering Journal. 390, (2020).
https://doi.org/10.1016/j.cej.2020.124513.
47. L. Hu, Z. Yang, L. Cui, et al., Fabrication of hyperbranched polyamine functionalized graphene for
high-eciency removal of Pb(II) and methylene blue. Chemical Engineering Journal. 287, 545-556,
(2016). https://doi.org/10.1016/j.cej.2015.11.059.
4. A. Nasrullah, B. Saad, A. H. Bhat, et al., Mangosteen peel waste as a sustainable precursor for high
surface area mesoporous activated carbon: Characterization and application for methylene blue
removal. Journal of Cleaner Production. 211, 1190-1200, (2019).
https://doi.org/10.1016/j.jclepro.2018.11.094.
49. M. Liu, L.-a. Hou, B. Xi, et al., Synthesis, characterization, and mercury adsorption properties of hybrid
mesoporous aluminosilicate sieve prepared with y ash. Applied Surface Science. 273, 706-716,
(2013). https://doi.org/10.1016/j.apsusc.2013.02.116.
Scheme
Scheme 1 is available in the Supplementary Files section.
Figures
Page 19/27
Figure 1
XRD patterns of the raw material and standard card of SiO2 (PDF#89-1961).
Page 20/27
Figure 2
XRD patterns of PSAM prepared with different NaOH concentrations and standard card for SiO2
(PDF#89-1961).
Page 21/27
Figure 3
FTIR patterns of PSAM prepared with different NaOH concentration conditions.
Page 22/27
Figure 4
SEM images of PSAM prepared with different NaOH concentration conditions.
Page 23/27
Figure 5
N2 adsorption-desorption isotherms (a), pore-size distribution (b).
Page 24/27
Figure 6
Zeta potential of 2M PSAM (a) 4M PSAM (b) 6M PSAM (c) 8M PSAM (d).
Figure 7
Page 25/27
Effects of different samples (a) experiment temperature (b) and solution pH (c) on adsorption eciency
with MB (adsorbent dose, 50 mg; the volume of the medium, 50 mL; initial concentration, 20 mg/L;
contact time, 24 h and temperature, 20).
Figure 8
Adsorption kinetics of MB on PSAM samples:
pseudo-rst-order (a); pseudo-second-order (b) (adsorbent dose, 50 mg; volume of the medium, 50 mL;
initial concentration, 20 mg/L; pH, 6 and temperature, 20).
Figure 9
Adsorption isotherm of MB on PSAM samples: Langmuir (a, b); Freundlich (c) (adsorbent dose, 50 mg;
the volume of the medium, 50 mL; initial concentration, 20-60 mg/L; pH, 6; contact time, 24 h and
temperature, 20).
Page 26/27
Figure 10
Vants Hoffplot for the determination of different thermodynamics parameters for the Adsorption of MB
onto 2M PSAM (adsorbent dose, 50 mg; the volume of the medium, 50 mL; initial concentration, 20 mg/L;
pH, 6; contact time, 24 h and temperature, 20-40).
Page 27/27
Figure 11
FTIR patterns before and after adsorption of 2M PSAM on MB.
Supplementary Files
This is a list of supplementary les associated with this preprint. Click to download.
oatimage1.png
oatimage13.png
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
This work reports the synthesis of nanosilica-coated magnetic carbonaceous adsorbents (MCA@SiO2) using low-temperature hydrothermal carbonization technique (HCT) and the feasibility to utilize it for methylene blue (MB) adsorption. Initially, a carbon precursor (CP) was synthesized from corn starch under saline conditions at 453 K via HCT followed by the magnetization of CP again via HCT at 453 K. Subsequently, MCA was coated with silica nanoparticles. MCA and MCA@SiO2 were characterized using X-ray diffraction, Fourier transform infrared, scanning electron microscopy/energy-dispersive spectroscopy, transmission electron microscopy, and Brunauer–Emmett–Teller (BET) N2 adsorption–desorption isotherms. The BET surface area of MCA and MCA@SiO2 were found to be 118 and 276 m2 g–1, respectively. Adsorption of MB onto MCA@SiO2 was performed using batch adsorption studies and in the optimum condition, MCA@SiO2 showed 99% adsorption efficiency with 0.5 g L–1 of MCA@SiO2 at pH 7. Adsorption isotherm studies predicted that MB adsorption onto MCA@SiO2 was homogeneous monolayer adsorption, which was best described using a Langmuir model with the maximum adsorption capacity of 516.9 mg g–1 at 25 °C. During adsorption kinetics, a rapid dye removal was observed which followed pseudo-first- as well as pseudo-second-order models, which suggested that MB dye molecules were adsorbed onto MCA@SiO2 via both ion exchange as well as the chemisorption process. The endothermic and spontaneous nature of the adsorption of MB onto MCA@SiO2 was established by thermodynamics studies. Mechanism of dye diffusion was collectively governed by intraparticle diffusion and film diffusion processes. Furthermore, MB was also selectively adsorbed from its mixture with an anionic dye, that is, methyl orange. Column adsorption studies showed that approximately 500 mL of MB having 50 mg L–1 concentration can be treated with 0.5 g L–1 of MCA@SiO2. Furthermore, MCA@SiO2 was repeatedly used for 20 cycles of adsorption–desorption of MB. Therefore, MCA@SiO2 can be effectively utilized in cationic dye-contaminated wastewater remediation applications.
Article
Many modern technologies are persistent to improve and adapt dyes removal. Metal-organic Frameworks (MOFs) as the significant porous and multi-dimensional networks has been concerned for water pollutant elimination. In this research solvothermal technique has been used to prepare hexahedral NH2-MIL-88(Fe), which was subsequently characterized using various techniques. The prepared MOF with large surface area (865.59 m²/g) showed perfect adsorption for methylene blue (MB) as cationic dye and acid red 57 (AR57) as anionic dye in water. Several studies were produced like; adsorption technique, isotherm of adsorption, kinetics of adsorption, adsorption energy, and other thermodynamic characteristics. The Langmuir isotherm model was found to be most excellent model for fitting the adsorption of both MB and AR57 dyes onto NH2-MIL-88(Fe). The pseudo‐second‐order model fitted the data on adsorption and a comprehensive description of the adsorption method. Both dyes' adsorption reactions were endothermic, adsorption increasing as temperature is raised. In fact, the as-prepared NH2-MIL-88(Fe) proved that it was a vital material to remove MB and AR57 from water.
Article
Herein, magnetic starch-based composite hydrogel microspheres (SCHMs) were produced via a Diels-Alder reaction during spray drying. The adsorption behavior of SCHMs for methylene blue (MB) was investigated systematically. The adsorption process could be described by Langmuir isothermal model and pseudo-first-order kinetic model with R² > 0.95 and R² > 0.98, respectively, resulting from its dominating physical adsorption. The thermodynamic studies showed the adsorption was spontaneous and the nature of adsorption was endothermic due to that the free energy ∆G was a negative value and the enthalpy ∆H was a positive value. The absorbent had good reusability with 90.5% of removal efficiency after 5 cycles. In addition, the effects of concentration, pH and temperature on the adsorption of MB by SCHMs were optimized using the Box-Behnken Design and response surface methodology. The order of three factors affecting the adsorption capacity of SCHMS is initial pH > initial concentration > temperature. Due to their safe, non-toxic and easily recycled advantages, SCHMs have the potential to be applied to water treatment in the future.
Article
Cork, a porous biomass material, is consist of thin-walled hollow prismatic cells arranged into a compact and orderly honeycomb-like structure and could be applied as an adsorption material. Here, cork-activated carbons (CACs) with a fluffy honeycomb-like structure were synthesized by two-step pyrolysis with solid KOH chemical activation to rapidly and efficiently adsorb methylene blue (MB) (maximum wavelength: 664 nm). The structure, morphology and surface functional groups of the CACs were characterized using BET, SEM, and FTIR analysis. The results show that the CACs have a well-developed hierarchical porous structure and an ultra-high specific surface area of 2864.9 m²/g, which would facilitate the efficient diffusion and adsorption of MB molecules onto CACs. MB adsorption performance results show that the CACs have an outstanding maximum MB adsorption capacity (1103.68 mg/g) and fast adsorption kinetics (800 mg/L, 99.8% in 10 min), indicating that CACs possess significant advantages compared with most other adsorbents previously reported. The adsorption mechanism was studied by various kinetic models, isothermal models and thermodynamic models. Langmuir model is the most adapted to describe the adsorption process, indicating that the MB molecules are uniformly adsorbed on CAC’s surface in a single layer. Moreover, MB adsorption by the CACs was an endothermic, spontaneous and randomly increasing adsorption. The regeneration test showed that the uptake of MB onto CACs can still reached 580 mg/g after three adsorption-desorption cycles, demonstrating the excellent reusability of CACs. The continuous adsorption performance of MB onto CACs was evaluated by a packed column test, which further confirmed its potential as an adsorbent for dye wastewater purification.
Article
Biochar is a promising agent for wastewater treatment, soil remediation, and gas storage and separation. This review summarizes recent research development on biochar production and applications with a focus on the application of biochar technology in wastewater treatment. Different technologies for biochar production, with an emphasis on pre-treatment of feedstock and post treatment, are succinctly summarized. Biochar has been extensively used as an adsorbent to remove toxic metals, organic pollutants, and nutrients from wastewater. Compared to pristine biochar, engineered/designer biochar generally has larger surface area, stronger adsorption capacity, or more abundant surface functional groups (SFG), which represents a new type of carbon material with great application prospects in various wastewater treatments. As the first of its kind, this critical review emphasizes the promising prospects of biochar technology in the treatment of various wastewater including industrial wastewater (dye, battery manufacture, and dairy wastewater), municipal wastewater, agricultural wastewater, and stormwater. Future research on engineered/designer biochar production and its field-scale application is discussed. Based on the review, it can be concluded that biochar technology represents a new, cost effective, and environmentally-friendly solution for the treatment of wastewater.
Article
Magnetic carbonaceous adsorbents were synthesized by ball-milling biochar (BC) or activated carbon (AC) with Fe3O4 nanoparticles, and their capacities to sorb methylene blue (MB) from water were evaluated and compared. Ball milling with magnetite not only improved the surface properties of the carbonaceous adsorbents, especially BC, but also introduced magnetic properties through mechanical extrusion. Furthermore, ball-mill extrusion increased the MB adsorption capacity of BC at all pH values by 14-fold, on average, but BC ball milled with magnetite had even greater MB adsorption capacity (27-fold, greater, on average). While ball milling of AC also improved its MB adsorption capacity (by almost 3-fold, on average), ball milling with magnetite did not further improve its MB adsorption capacity. All the magnetic adsorbents showed fast MB adsorption kinetics, reaching equilibrium within about 8 h. The Langmuir maximum MB adsorption capacity of the magnetic ball-milled BC (MBM-BC) was the highest (500.5 mg/g) among all the samples including the ones derived from AC. After five adsorption-desorption cycles, MBM-BC maintained about 80% MB removal capacity. The high MB adsorption capacity of MBM-BC was attributed to its increased surface area, opened pore structure, functional groups and aromatic CC bonds, which promoted π-π and electrostatic interactions. Findings from this study indicate that the magnetic ball-milled BC is a promising adsorbent due to its environmentally friendly synthesis, high efficiency, low cost, and convenience in operation.
Article
A composite structure of the zeolite and activated carbon is successfully synthesized through CO2 activation followed by the hydrothermal synthesis method using coal gangue as the raw material. The zeolite-activated carbon composite is used as an adsorbent for the removal of heavy metal ions and macromolecular organics. The integrative adsorption property of the composite could be improved by adding a certain amount of powdered coal into coal gangue as the extra carbon source. The specific surface area of a so-prepared composite is 669.4 m²/g, which is much larger than that of the pure NaA type zeolite (249.3 m²/g). As a result, the zeolite-activated carbon shows high adsorptive efficiencies for Cu²⁺ (92.8%) and Rh-B (94.2%). The thermodynamic and kinetic processes for adsorption of Cu²⁺ and Rh-B onto pure zeolite and zeolite-activated carbon are also investigated systematically. The mechanistic study of zeolite-activated carbon indicates that the uniform micropores in zeolite are suitable for the adsorption of heavy metal ions and the multilevel porous structure of activated carbon could accommodate macromolecular organics. The high value-added product prepared from coal gangue is a good alternative to treat industrial waste water.
Article
Mesoporous silicon carbon (MSC) composites were prepared by in-situ carbothermal treatment of cetyl trimethyl ammonium bromide derived from mesoporous silica. MSC composites were used as adsorbent for methylene blue (MB) removal. Research was focused on investigating effect of washing conditions (deionized water or alkali) on surface properties and adsorption capacity of MSC composites. Results showed that best adsorption performance was given by the MSC composite washed by alkali at optimum pH 13.0. Adsorption mechanism for MB removal was systematically investigated by X-ray photoelectron spectrometer analysis, adsorption isotherms, adsorption kinetics and adsorption thermodynamics. MB adsorption by the MSC composites was found to be driven by three possible schemes: physical exothermic reaction, hydrogen bonds and π-interaction. The maximum adsorption capacity was 156.56 mg/g for MSCa13. The negative values of ΔG0, ΔH0 and ΔS0 indicated that the adsorption process was a spontaneous exothermic reaction. The adsorbent can be regenerated for reuse.
Article
In this study, the performance of the peroxi-coagulation fuel cell (PCFC) process for p-arsanilic acid (p-ASA) removal and power generation were evaluated at initial pH of 3.0–6.0. It was found that >94% of p-ASA could be removed within 60 min at pH of 3.0–6.0. The fastest p-ASA removal could be obtained at pH of 4.0, whereas the optimum power density was 98.3 mW m⁻² at 745 mA m⁻² achieved at pH of 3.0. In the PCFC system, p-ASA removal was ascribed to the combined action of the electrostatic attraction between p-ASA⁻ and positively charged iron precipitant and the oxidation of p-ASA by OH. At pH of 4.0–6.0, approximately 45%–50% of p-ASA could be oxidatively converted to inorganic arsenic and the residual inorganic arsenic concentration in solution was below 10 µg L⁻¹. Contrastly, although 92% of p-ASA was oxidized into As(V) at pH of 3.0, approximately 11 µM of As(V) was still remained in the solution. In this case, about 81% of p-ASA was oxidized to p-aminophenol by hydroxylation of OH and less than 10% of p-ASA was oxidized into p-benzoquinone after 60 min. Notably, phosphate exerted the significantly inhibited effect on arsenic species immobilization via competing for the adsorptive sites on the iron precipitant. The kinetics study revealed that PCFC process had superior p-ASA removal efficiency as compared to electrocoagulation fuel cell process under the same condition. Generally, the PCFC process holds promise as a feasible option for p-ASA removal at rural/remote area where electricity is not easily available.