ArticlePDF Available

Adsorption of Tetracycline by Magnetic Mesoporous Silica Derived from Bottom Ash-Biomass Power Plant

MDPI
Sustainability
Authors:

Abstract and Figures

In recent years, the contamination of the aquatic environment with antibiotics, including tetracyclines, has drawn much attention. Bottom ash (BA), a residue from the biomass power plant, was used to synthesize the magnetic mesoporous silica (MMS) and was utilized as an adsorbent for tetracycline (TC) removal from aqueous solutions. The MMS was characterized by Fourier transform-infrared (FTIR), X-ray diffraction (XRD) pattern, and scanning electron microscopy (SEM). Optimum conditions were obtained in overnight incubation at 60 °C, a pH of 6–8, and an adsorption capacity of 276.74 mg/g. The isotherm and kinetic equations pointed to a Langmuir isotherm model and pseudo-first-order kinetic optimum fitting models. Based on the very low values of entropy changes (ΔS°), the negative value of enthalpy changes (ΔH°) (−15.94 kJ/mol), and the negative Gibbs free-energy changes (ΔG°), the adsorption process was physisorption and spontaneous.
Content may be subject to copyright.
Citation: Hanh, P.T.H.; Phoungthong,
K.; Chantrapromma, S.; Choto, P.;
Thanomsilp, C.; Siriwat, P.;
Wisittipanit, N.; Suwunwong, T.
Adsorption of Tetracycline by
Magnetic Mesoporous Silica Derived
from Bottom Ash—Biomass Power
Plant. Sustainability 2023,15, 4727.
https://doi.org/10.3390/su15064727
Academic Editors: Ken Kawamoto,
Tomonori Ishigaki and Hoang
Giang Nguyen
Received: 1 January 2023
Revised: 26 February 2023
Accepted: 27 February 2023
Published: 7 March 2023
Copyright: © 2023 by the authors.
Licensee MDPI, Basel, Switzerland.
This article is an open access article
distributed under the terms and
conditions of the Creative Commons
Attribution (CC BY) license (https://
creativecommons.org/licenses/by/
4.0/).
sustainability
Article
Adsorption of Tetracycline by Magnetic Mesoporous Silica
Derived from Bottom Ash—Biomass Power Plant
Phan Thi Hong Hanh 1, Khamphe Phoungthong 1, Suchada Chantrapromma 2, Patcharanan Choto 3,4,
Chuleeporn Thanomsilp 3, Piyanuch Siriwat 3, Nuttachat Wisittipanit 3,5 and Thitipone Suwunwong 3,4 ,*
1Industrial Ecology in Energy Research Center, Faculty of Environmental Management, Prince of Songkla
University, Songkhla 90112, Thailand
2Division of Physical Science, Faculty of Science, Prince of Songkla University, Songkhla 90112, Thailand
3School of Science, Mae Fah Luang University, Chiang Rai 57100, Thailand
4Center of Chemical Innovation for Sustainability, Mae Fah Luang University, Chiang Rai 57100, Thailand
5Department of Materials Engineering, School of Science, Mae Fah Luang University,
Chiang Rai 57100, Thailand
*Correspondence: thitipone.suw@mfu.ac.th; Tel.: +66-869658254
Abstract:
In recent years, the contamination of the aquatic environment with antibiotics, including
tetracyclines, has drawn much attention. Bottom ash (BA), a residue from the biomass power plant,
was used to synthesize the magnetic mesoporous silica (MMS) and was utilized as an adsorbent
for tetracycline (TC) removal from aqueous solutions. The MMS was characterized by Fourier
transform-infrared (FTIR), X-ray diffraction (XRD) pattern, and scanning electron microscopy (SEM).
Optimum conditions were obtained in overnight incubation at 60
C, a pH of 6–8, and an adsorption
capacity of 276.74 mg/g. The isotherm and kinetic equations pointed to a Langmuir isotherm model
and pseudo-first-order kinetic optimum fitting models. Based on the very low values of entropy
changes (
S
), the negative value of enthalpy changes (
H
) (
15.94 kJ/mol), and the negative Gibbs
free-energy changes (G), the adsorption process was physisorption and spontaneous.
Keywords: antibiotic contamination; tetracycline; solid waste; adsorption mechanisms
1. Introduction
Antibiotics have been commonly used for the prevention and treatment of infectious
diseases, both in humans and animals. According to Cycon et al., the situation analyzed
in 75 countries, indicates that the antibiotics usage climbed by 65%, and the prediction is
that in 2030, the consumption of antibiotics will be higher by 200% compared with 2015 [
1
].
Among antibiotics, tetracycline (TC) is a conventional, inexpensive antibiotic having wide-
ranging antibacterial action that is frequently used in both human and veterinary medicine
to prevent infection [
2
]. Additionally, TC antibiotics are utilized as a growth stimulant
for animals and for agricultural purposes [
3
,
4
]. Moreover, TC constitutes one of the most
important antibiotic families, ranking second in production and usage worldwide [
5
,
6
].
Because of its high aqueous solubility, TC has been found in ecological communities such
as surface water and groundwater (ranging from 5.4 to 8.1 ng/L) [
7
], municipal solid waste
(greater than 100 ng/L) [
8
], and soil (ranging from 86.2 to 198.7
µ
g/kg) [
9
], and it can be eas-
ily transferred to other environments via aqueous matrixes [
10
], and like other antibiotics, it
has a structural framework called naphthol ring that makes it difficult to degrade [
11
]. Thus,
Cyco´n et al. [
1
] suggested that “the reason is that antibiotics are not completely metabolized
by humans and animals, and a large proportion of the administered drug is released as
the parent compound through feces and urine, discharging into domestic wastewater and
into the pits where slurries/sludges are deposited”. Gu and Karthikeyan [
12
] also found
that “in the statewide survey of wastewater treatment plants, the compound TC was the
most frequently detected antibiotic (among 25 antibiotics), being present in 80% of the
Sustainability 2023,15, 4727. https://doi.org/10.3390/su15064727 https://www.mdpi.com/journal/sustainability
Sustainability 2023,15, 4727 2 of 13
wastewater influent and effluent samples”. However, standard aqueous solution water
treatment and spontaneous biodegradation are ineffective in removing TC from aqueous
solution [
12
,
13
]. Therefore, techniques for the secure and efficient removal of TC from
liquid have received a great deal of interest. Methods for TC removal include oxidation [
14
],
photo electrocatalytic [
15
], degradation [
16
], membrane processing [
17
], adsorption [
18
],
permeation [
19
], flocculation [
19
], ozonation oxidation [
19
]. Among these, adsorption has
several advantages over other procedures, including ease of operation, inexpensive operat-
ing costs, and significant removal efficiency at extremely low TC content in wastewater and
water [
20
]. Particularly, no generation of toxicity of intermediated and by-products during
adsorption makes it a more attractive appealing method of treating TC [
21
]. Until now,
some studies investigated the adsorption and removal of tetracyclines on several materials
such as graphene oxide [
22
], activated carbon [
23
], kaolinite [
24
], single-walled carbon
nanotubes, and multi-walled carbon nanotubes [
13
]. Carbon nanotubes and graphene,
which have a high graphite structure, have a strong TC elimination capacity [
25
]. However,
the high manufacturing, disposal, and regeneration costs of the aforementioned materials
would pose significant barriers in their practical application for TC adsorption, as well as
the capacity for large-scale application. As a result, there is still a considerable desire for
the development of efficient and affordable, simple-to-operate, high-selectivity devices.
Mesoporous silicas are often employed in a variety of applications, including catalysis,
separation, drug administration, chemical sensing, optic and electrical devices, rubber
reinforcement, and as a template in the production of nanomaterials [
26
]. Because they
possess a large surface area, big pore size, pore volume, and regular channel-type structures,
these properties are potential advantages that suit the adsorbate. A renewed interest in the
design of adsorbents [
27
] and catalysts has been sparked by the discovery of hexagonally
organized mesoporous silica [
28
], which has a distinctively large surface area, well-defined
pore size, and pore-shaped pores. For the production of surfactant-template silica materials,
which need an organic structure-directing agent or template is typically a single surfactant,
such as: the Pluronic-type surfactant; poly(ethylene oxide)-b-poly(propylene oxide)-b-
poly(ethylene oxide)-Pluronic P123 (EO
20
PO
70
EO
20
) or Pluronic F127 (EO
106
PO
60
EO
106
)
and cetyltrimethylammonium bromide (CTMAB) [
29
32
]. Some studies have magnetized
the silicates to allow for rapidly and effectively separating the adsorbents from the aqueous
solution by an external magnetic field, and incorporating magnetic elements within or on
mesoporous silicates has also enhanced the adsorption properties [33].
Some previous researches have investigated magnetic mesoporous silica (MMS) with
a variety of structures: including embedded [
31
], core-shell [
32
], and yolk-shell [
34
]. The
MMS could be synthesized by combining the mesoporous silica (MS) and magnetic particles.
Tetraethyl orthosilicate (TEOS) was used as the silica source, and magnetic particles were
prepared from iron (II) chloride tetrahydrate (FeCl
2
,4H
2
O), iron (III) chloride hexahydrate
(FeCl
3·
6H
2
O) [
35
]. Because TEOS is expensive and magnetic particles require time and
chemicals to make, searches for alternative cheap source are necessary.
Bottom ash (BA)—biomass power plant was a byproduct from the combusting of agri-
cultural waste to produce energy. Approximately 85–95 percent of BA was generated from
biomass power plants [
36
], with SiO
2
making up the majority of this ash [
37
]. Thus, these
residues were an ideal silica source. Currently, Thailand has an abundance of BA—biomass
power plants that may be utilized to create magnetic mesoporous silica as an appropriate
supply. Generally, utilizing the residues from incineration and combustion plant to produce
silica materials is feasible. Some studies have been successful in converting this byproduct
to zeolite in alkali solution [38], synthesizing zeolitic material and successfully separating
SiO
2
from municipal solid waste incineration (MSWI) ash [
39
], MCM-41, SBA-15, and
SBA-16 mesoporous silica from power plant bottom ash were successfully synthesized
for the first time in 2007 [
40
]. Mesoporous silica was prepared by the sol-gel method
from municipal solid waste incineration bottom ash [
41
] and industrial fly ash [
42
]. This
application would increase the usage of BA—biomass power plants, minimize pollution,
and effectively improve the quality of the wastewater process. However, according to our
Sustainability 2023,15, 4727 3 of 13
knowledge, there has not been any published research on the synthesis of magnetic silica
nanoparticles derived from BA for specific applications in the adsorption of tetracycline.
Therefore, the purpose of this study is to apply magnetic mesoporous silica synthesized
from biomass power plant ash to absorb TC from aqueous solutions. The properties of
the MMS including surface area, function groups, and crystal structure were analyzed.
The research was conducted on the isotherm model, kinetics model, and thermodynamic
characteristics of TC’s adsorption onto the MMS. According to the findings of adsorption
tests, the MMS demonstrated promising potential for TC removal from water. This study
provides a detailed technique for achieving the goal of treating trash with waste in addition
to describing how solid waste is utilized as a resource.
2. Materials and Methods
2.1. Materials
In this study, the bottom ash (BA) was obtained from the rubber biomass power
plant of Gulf Yala company, Yala, located in southern Thailand. The composition of
BA was determined by X-ray fluorescence spectrometer (XRF) as listed in Table S1 (in
the Supplementary material). Tetracycline hydrochloride, 96%, Alfa Aesar were used
without further purification. Hexadecyltrimethyl ammonium bromide (CTAB)
98%,
Sigma, Cibolo, TX, USA; Sodium hydroxide (NaOH) 98%, Loba Chemie PVT.LTD, India;
Hydrochloric acid (HCl) 37%, Qrec, Newzealand; Ethanol 99.9%, Qrec, Newzealand.
2.2. Methods
2.2.1. Synthesis of Magnetic Mesoporous Silica
The biomass power plant ash (BA) which includes 55% SiO
2
content as the main
component (Table S1 in the Supplementary Material) was used as raw material for extraction
of SiO
2
compound. Briefly, BA sample was ground into small particles (<45
µ
m) and then
washed with DI water several times and dried in the oven at 90
C for 24 h. The permanent
magnets were employed to separate the magnetic and nonmagnetic ashes. The magnetic
ash was served as the magnetic component in the magnetic mesoporous silica, and the
nonmagnetic ash was utilized to extract silica. The weight ratio between the magnetic and
nonmagnetic ashes was 1:10.
A total of 4 g of nonmagnetic ash was refluxed with 100 mL NaOH solution with the
concentration 4 M at 90
C for 16 h in a round bottom flask. Then, the reaction solution was
filtered through a membrane of 0.45
µ
m and the supernatant was collected for adjusting
pH at 7 by 5 M HCl and aging at the room temperature for 24 h. The precipitated product
was washed with distilled water several times and dried at 90
C for 24 h. To prepare the
sodium silicate solution, the obtained white power and NaOH (4:5 w/w) were precisely
weighed and then dissolved in 250 mL of distilled water at 80
C for 1 day [
43
]. The white
power is SiO
2
and was extracted from BA, which was analyzed by X-ray fluorescence
spectrometer (XRF), as shown in Table S2 (in the Supplementary material).
Deionized water (20 mL) was added to 1.2 g hexadecyltrimethyl ammonium bromide
(CTAB) and then mixed at 40
C until a clear solution was observed. The aqueous CTAB
solution was slowly mixed with sodium silicate solution and continuously stirred for
15 min. This mixture was added slowly into the magnetic ash; the ratio of the mass of
maghemite to Na
2
SiO
3
10% was 1:20, and it was stirred while being heated to 80
C. After
further stirring for 30 min, the pH of the mixture was adjusted to 11 by 5 mol/L HCl
solution and then continuously stirred for 6 h. The mixture was kept in a water bath at
80
C for 72 h. The ethanol 99.9%, 100 mL, was added to the precipitated product. To
eliminate CTAB, the mixture was sonicated for 30 min at 60
C [
44
] and then washed with
DI water several times and dried at 80 C for 24 h.
The experimental procedure was repeated more than three times to verify that the
results are reproducible.
Sustainability 2023,15, 4727 4 of 13
2.2.2. Characterization
The mineral compositions in BA and MMS were determined by X-ray fluorescence
(XRF), (S2175 Ranger, Bruker, Burladingen, Germany). Fourier transform infrared spec-
troscopy FT-IR (Perkin Elmer Model Spectrum GX) was used to analyze the specific func-
tional groups of the adsorbents by compressed samples into KBr pellets and then analyzed
with a Nicole IS10 spectrometer over the wavelength ranged from 400 to 4000 cm
1
. X-ray
diffraction (XRD) pattern (PAN analytical, X’Pert Pro MPD) was performed on a Bruker
AXS Advance instrument for confirmation the structure. The surface morphology (SEM) of
magnetic mesoporous silica was examined by scanning electron microscopy (SEM, Apreo,
FEI, South Moravian Region, Czech Republic) running by Schottky field emission at the
accelerating voltage of 20 kV in high vacuum mode. The elemental mapping analysis
of the sample was recorded by energy-dispersive X-ray spectrometer (EDS)—(X-Max80,
Oxford, UK).
2.2.3. Adsorption Experiments
MMS was used for tetracyclines (TC) adsorption in an aqueous solution. To find the
ideal conditions, the impact of adsorption time and temperature on the unit adsorption
capacity and adsorption rate of TC was investigated. MMS (3 mg) was dispersed in a flask
containing 10.0 mL TC solution with various concentrations (10–100 mg/L), pH 6–8, and
then fully homogenized with a vortex mixer. The suspension was incubated overnight at
25
C, 45
C, and 60
C and covered with aluminum foil to protect TC from the potential
photo degradation. MMS was then separated from the samples through a magnet after
centrifuging at 2000 rpm for 15, 30, 45, 60, and 90 min while maintaining the experiment
temperature including centrifugation. The residual concentration of TC in an aqueous
solution was determined by UV–vis absorbance at 357 nm, using a calibration curve
built. All the experiments were replicated thrice, and the averaged results were reported.
Equation (1) was used to calculate the percentage removal of TC, and Eq
uation (2
) was
used to determine the adsorption capacity:
Removal(%)=(CoCe)
Co
×100 (1)
Qe=(CoCe)×V
m(2)
where:
Qeis the amount of TC adsorption per unit weight of adsorbent (mg/g);
Cois the initial concentration of TCs (mg/mL);
Ce(mg/L) is the equilibrium concentration of TC;
Vis the solution volume (mL);
mis the mass of adsorbent (g).
3. Results and Discussion
3.1. Characterization of MMS
The FT-IR peaks of the MMS before and after adsorption of TC were collected in the
range from 400 to 4000 cm
1
(Figure 1), indicating the chemical bonds and functional
groups in the compound. The O–H stretching vibration was identified as the source of the
big broadband at 3440 to 3443 cm
1
. The absorption peaks at 1640 cm
1
were caused by
the symmetric and asymmetric bending vibration of C=O. Fe–O stretching was assigned to
the band below 700 cm
1
. The characteristic absorption bands of the sample at 692 and
5
83 cm1
were assigned to iron(II) oxide bending, and the band at 459 to 446 cm
1
was
ascribed to the bending vibration mode of Fe
2
O
3
. The band at 1048 cm
1
to 1051 cm
1
,
which is associated with Si-O-Si antisymmetric stretching vibrations, is a sign that silicon
dioxide is present in the sample. The bands at 579 and 583 cm
1
are also an indication of
the presence of Si–O–Fe [
45
]. Most of the adsorbent’s peaks remained constant after TC
Sustainability 2023,15, 4727 5 of 13
adsorption, revealing that the adsorption procedure did not modify the structure of material.
However, several distinctive peaks at 1478, 2852, and 2922 cm
1
were characteristic peaks
of the C=C skeleton and C-H stretching vibration of CH
2
and CH
3
induced by aromatic
groups of TC [
46
,
47
]. The FT-IR spectrum proved that TC was adsorbed onto the adsorbent
and the sample’s structure is mostly stable after adsorption [48].
Sustainability 2022, 14, x FOR PEER REVIEW 5 of 14
groups in the compound. The O–H stretching vibration was identified as the source of the
big broadband at 3440 to 3443 cm—1. The absorption peaks at 1640 cm—1 were caused by
the symmetric and asymmetric bending vibration of C=O. Fe–O stretching was assigned
to the band below 700 cm—1. The characteristic absorption bands of the sample at 692 and
583 cm—1 were assigned to iron(II) oxide bending, and the band at 459 to 446 cm—1 was
ascribed to the bending vibration mode of Fe2O3. The band at 1048 cm—1 to 1051 cm—1,
which is associated with Si-O-Si antisymmetric stretching vibrations, is a sign that silicon
dioxide is present in the sample. The bands at 579 and 583 cm—1 are also an indication of
the presence of Si–O–Fe [45]. Most of the adsorbent’s peaks remained constant after TC
adsorption, revealing that the adsorption procedure did not modify the structure of ma-
terial. However, several distinctive peaks at 1478, 2852, and 2922 cm—1 were characteristic
peaks of the C=C skeleton and C-H stretching vibration of CH2 and CH3 induced by aro-
matic groups of TC [46,47]. The FT-IR spectrum proved that TC was adsorbed onto the
adsorbent and the sample’s structure is mostly stable after adsorption [48].
Figure 1. FTIR spectra of the adsorbent before and after adsorption of TC.
The XRD profile of the MMS in the FT-IR spectrum proved that TC was adsorbed
onto the adsorbent and the sample’s structure is mostly stable after adsorption [48].
Comparisons with those of SiO2 (extracted from biomass power plant Ash) and Fe2O3
peak [49] are shown in Figure 2. The XRD pattern of the MMS exhibited the characteristic
diffraction peaks of Fe2O3 with weak intensity due to the lower concentration of Fe2O3. In
addition, the relatively slightly broad peak observed at 1530° arose from the SiO2,
demonstrating that the crystalline structure of the iron oxide was retained after the encap-
sulation in silica.
Figure 1. FTIR spectra of the adsorbent before and after adsorption of TC.
The XRD profile of the MMS in the FT-IR spectrum proved that TC was adsorbed onto
the adsorbent and the sample’s structure is mostly stable after adsorption [48].
Comparisons with those of SiO
2
(extracted from biomass power plant Ash) and
Fe
2
O
3
peak [
49
] are shown in Figure 2. The XRD pattern of the MMS exhibited the
characteristic diffraction peaks of Fe
2
O
3
with weak intensity due to the lower concentration
of Fe
2
O
3
. In addition, the relatively slightly broad peak observed at 15
30
arose from the
SiO
2
, demonstrating that the crystalline structure of the iron oxide was retained after the
encapsulation in silica.
Sustainability 2022, 14, x FOR PEER REVIEW 6 of 14
Figure 2. XRD patterns of the MMS.
The morphological structures of the MMS were examined using SEM. As shown in
Figure 3a, there are many unobvious spherical particles (aggregated) with a size of around
5080 nm. A rough surface, which mainly consists of notwellcrystallized Fe2O3, seems
not to be efficient to achieve homogeneous coating SiO2, but instead irregular and severely
agglomerated particles are observed in Figure 3b,c.
Figure 3. SEM images of MMS (ac), and EDS curve of MMS (d).
Figure 2. XRD patterns of the MMS.
Sustainability 2023,15, 4727 6 of 13
The morphological structures of the MMS were examined using SEM. As shown in
Figure 3a, there are many unobvious spherical particles (aggregated) with a size of around
50
80 nm. A rough surface, which mainly consists of not
well
crystallized Fe
2
O
3
, seems
not to be efficient to achieve homogeneous coating SiO
2
, but instead irregular and severely
agglomerated particles are observed in Figure 3b,c.
Sustainability 2022, 14, x FOR PEER REVIEW 6 of 14
Figure 2. XRD patterns of the MMS.
The morphological structures of the MMS were examined using SEM. As shown in
Figure 3a, there are many unobvious spherical particles (aggregated) with a size of around
5080 nm. A rough surface, which mainly consists of notwellcrystallized Fe2O3, seems
not to be efficient to achieve homogeneous coating SiO2, but instead irregular and severely
agglomerated particles are observed in Figure 3b,c.
Figure 3. SEM images of MMS (ac), and EDS curve of MMS (d).
Figure 3. SEM images of MMS (ac), and EDS curve of MMS (d).
The samples typically determined the elemental analysis or chemical properties using
energy-dispersive X
ray spectrometry (EDS). The various elements in the sample are rep-
resented by energy peaks. As-prepared nanocomposites were found to include significant
amounts of Fe, O, Si, Ti, Na, Ca, Al, etc. (Figure 3d), confirming the formation of additional
nanocrystals on the surface of Fe2O3@SiO2particles.
3.2. Adsorption Isotherms of TC on MMS
In order to elucidate the interactions between the adsorption ability of the MMS and
TC in solution at adsorption equilibrium, the data were modeled using four adsorption
isotherms, including Langmuir, Freundlich, Temkin, and Sips isotherm models at 25, 45,
and 60 C. The isotherm models were listed below with Equations (3)–(6) as follows.
Qe=QmKLCe
1+KLCe;RL=1
1+KoCo(3)
Qe=KfC
1
n f
e(4)
Qe=RT
bT
ln(KTCe)(5)
Qe=Qm
(KSCe)a
1+ (KSCe)a(6)
where Q
m
= maximum adsorption ability (mg/g), C
e
= equilibrium concentration (mg/L),
Q
e
= adsorption capacity (mg/g) at equilibrium time, K
L
= Langmuir constant, R
L
= sepa-
ration constant, K
f
and n= Freundlich constant, R= universal gas constant (
8.314 J/mol
),
Sustainability 2023,15, 4727 7 of 13
T= tempe
rature in terms of Kelvin, b
T
= Temkin constant, K
T
= equilibrium bond constant
related to the maximum energy of bond, and Ksand aare Sips constants.
According to Figure 4a, as the equilibrium TC concentration climbed, the TC’s capacity
to adsorb on the MMS also dramatically increased. It was determined that the adsorption
process was endothermic because the adsorption capacity of TC increased as the adsorption
temperature rose. This is likely because the higher temperature supported the quantity of
activated functional groups on the surface of the MMS or sped up the diffusion rate of the
tetracycline molecules.
Sustainability 2022, 14, x FOR PEER REVIEW 8 of 14
Figure 4. Adsorption isotherms of TC onto MMS at 25, 45, and 60 °C (a), The nonlinear fits for Lang-
muir, Freundlich, Temkin, Sips models at 25 °C (b), 45 °C (c), 60 °C (d). (adsorbent dosage = 0.03
g/L, pH = 8, V = 10 mL).
Table 1. Parameters of adsorption isotherm model of TC onto MMS at 25, 45, 60 °C.
Temp (°C) 25 °C 45 °C 60 °C
Langmuir
Qm 18.11 ± 2.65 43.92 ± 15.67 276.74 ± 159.62
KL 0.031 ± 0.009 0.009 ± 0.004 0.002 ± 0.001
R2 0.9832 0.9876 0.9815
SSE 1.58 1.74 0.19
χ2 0.53 0.53 0.53
Freundlich
n 0.54 ± 0.003 0.79 ± 0.001 0.96 ± 3.25 × 104
KF 1.31 ± 0.01 0.64 ± 0.003 0.56 ± 6.20 × 104
R2 0.9867 0.9983 0.9999
SSE 138.27 32.07 1.66
χ2 0.14 0.03 0.002
Temkin
AT 0.58 ± 0.01 0.58 ± 0.01 0.88 ± 0.03
BT 790.61 ± 7.53 790.62 ± 7.53 909.52 ± 10.28
R2 0.9181 0.9181 0.8880
SSE 855.40 3806.77 8002.25
χ2 0.86 3.89 8.03
Sips
Qm 12.92 ± 0.11 14.52 ± 0.17 17.07 ± 0.24
KS 0.18 ± 0.02 0.35 ± 0.08 0.47 ± 0.13
A 2.09 ± 0.18 3.79 ± 0.77 3.63 ± 0.87
R2 0.9603 0.9415 0.9218
SSE 512.81 1055.65 2210.77
χ2 0.51 1.06 2.22
Figure 4.
Adsorption isotherms of TC onto MMS at 25, 45, and 60
C (
a
), The nonlinear fits for
Langmuir, Freundlich, Temkin, Sips models at 25
C (
b
), 45
C (
c
), 60
C (
d
). (adsorbent dosage =
0.03 g/L, pH = 8, V= 10 mL).
The starting TC concentration used in the modeling procedure varied from 10 to
1
00 mg/L
, and the adsorption temperature was adjusted at 25, 45, and 60
C. Figure 4b–d
show the plot of the nonlinear fits for the aforementioned four isotherm models, and Table 1
lists the related parameter values. The Langmuir model was more appropriate to explain
the adsorption of TC onto the MMS, as shown by the correlation coefficients, R
2
, which
were consistently close to 1 and the sum of square error (SSE), chi-square (
χ2
) values, which
suggest that the surface of the MMS was homogeneous and the adsorption was monolayer.
Additionally, K
L
is a significant evaluation factor in relation to the binding site affinities.
The value of K
L
was between 0–1 and declined with the rising temperature, and separation
factor (R
L
) <1 reflected that the adsorption of TC onto the MMS was favorable [
50
]. The
nonlinear fitting of the Langmuir model (Figure 4b) showed the adsorption capacity (Q
m
)
of MMS for TC was 276.74 mg/g.
Sustainability 2023,15, 4727 8 of 13
Table 1. Parameters of adsorption isotherm model of TC onto MMS at 25, 45, 60 C.
Temp (C) 25 C 45 C 60 C
Langmuir
Qm18.11 ±2.65 43.92 ±15.67 276.74 ±159.62
KL0.031 ±0.009 0.009 ±0.004 0.002 ±0.001
R20.9832 0.9876 0.9815
SSE 1.58 1.74 0.19
χ20.53 0.53 0.53
Freundlich
n0.54 ±0.003 0.79 ±0.001 0.96 ±3.25 ×104
KF1.31 ±0.01 0.64 ±0.003 0.56 ±6.20 ×104
R20.9867 0.9983 0.9999
SSE 138.27 32.07 1.66
χ20.14 0.03 0.002
Temkin
AT0.58 ±0.01 0.58 ±0.01 0.88 ±0.03
BT790.61 ±7.53 790.62 ±7.53 909.52 ±10.28
R20.9181 0.9181 0.8880
SSE 855.40 3806.77 8002.25
χ20.86 3.89 8.03
Sips
Qm12.92 ±0.11 14.52 ±0.17 17.07 ±0.24
KS0.18 ±0.02 0.35 ±0.08 0.47 ±0.13
A2.09 ±0.18 3.79 ±0.77 3.63 ±0.87
R20.9603 0.9415 0.9218
SSE 512.81 1055.65 2210.77
χ20.51 1.06 2.22
The R
2
values of the Freundlich isotherms were from 0.98 to 0.99, but the 1/nvalues
were more than 1, at 1.85, 1.26, and 1.04, with increased temperatures at 25, 45, 60
C,
respectively, suggesting the adsorption is not prone to occur. Therefore, the Freundlich
isotherm had a hard time accurately explaining the TC adsorption pattern.
The Temkin and Sips isotherms were a poor fit, although heterogeneity was also
assumed. The Temkin isotherms illustrate that the heat of adsorption is negatively pro-
portional to the surface area covered by the adsorbate molecules. The Temkin’s constant,
abbreviated B
T
, represents the heat generated during adsorption. The fact that B
T
was
positive (790.61–909.52 J/mol) for the TC adsorption from the aqueous solution indicates
that TC was endothermically adsorbed on the MMS. The equilibrium binding constant A
T
is the quantity that relates to the greatest binding energy. When the temperature was raised
from 25 to 60
C, the values of A
T
also increased from 0.58 to 0.88 L/mg, indicating that the
adsorption of TC on the MMS was endothermic.
Equations (7) and (8) of the van’t Hoff equation were used to compute the Gibbs
free-energy change
G
enthalpy
H
and entropy
S
during the adsorption process (8).
The results are displayed in Table 2.
Go=RTlnKL(7)
ln(KL)=So
RHo
RT (8)
where Ris the ideal gas constant 8.314 J/(mol
·
K), Tis Kelvin temperature (K), and K
L
is the
Langmuir isotherm equilibrium constant (L/mg) [5053].
Sustainability 2023,15, 4727 9 of 13
Table 2. Thermodynamic parameters of TC adsorption onto MMS.
Temperature (K) G(kJ/mol) H(kJ/mol) S(J/(mol.K)
298 23.61
7.62 15.94
318 21.92
333 18.79
The absolute values of
G
are negative for all the parameter intervals, indicating that
the adsorption behavior is spontaneous.
G
also increased as the adsorption temperature
rose, suggesting that the adsorption process was impeded [
51
]. Moreover, the value of
G
in the range from
20 to
80 (kJ/mol) showed that physisorption was involved in the
process, but a minor chemical action effect could speed up the procedure. Furthermore, the
value of
H
lower than 20 (kJ/mol) revealed the physical adsorption with van der Waals
interaction implied in the process mechanism [
52
]. A very low
S
value
(15.94 J/(mol.K))
proved a little change in entropy occurred during the adsorption of TC by the MMS [53].
3.3. Adsorption Kinetics of TC on MMS
The adsorption kinetics explained the impact of time and temperature on the TC
adsorption rate. The adsorption kinetics test was conducted by adding 0.03 g of adsorbent to
10 mL of 100 mg/L TC solution at 45
C. The adsorption data were fitted using three popular
kinetic models, including the pseudo-first-order, pseudo-second-order, and Elovich models.
The kinetic models are listed in the Equations (9)–(11) respectively, the corresponding
nonlinear curves are shown in Figure 5, and Table 3contains the estimated and shown
kinetic parameters.
Qt=Qe1eK1t(9)
Qt=KeQ2
et
1+K2Qet(10)
Qt=1
bln(ab)+1
bln(t); (11)
Sustainability 2022, 14, x FOR PEER REVIEW 10 of 14
𝑄=(
)𝑙𝑛(𝑎𝑏) + (
)𝑙𝑛(𝑡); (11)
Table 3. Adsorption Kinetic Coefficients.
Model Parameter Value
pseudo-1st-order
K
1 0.14 ± 0.02
Qe 1.44 ± 0.02
R2 0.9971
SSE 0.005
χ2 0.00
pseudo-2nd-order
Qe 1.56 ± 0.003
K2 0.14 ± 0.002
R2 0.9646
SSE 2.44
χ2 0.00
Elovich
β 1.33 ± 0.05
α 3.90 ± 0.03
R2 0.9559
SSE 4.20
χ2 4.20
Figure 5. Adsorption kinetics of TC onto MMS at 45 °C, with initial TC concentration = 100 mg/L.
For the pseudo-second-order and Elovich models, R2 was considerably less than 1,
and SSE, χ2 were higher. It was determined that neither model could adequately explain
the experimental findings. However, the pseudo-first-order kinetic model showed a high
value of R2 (0.9971) and a small value of SSE, χ2.The closer fitting curves imply that the
pseudo-first-order adsorption process was dominant for TC adsorption on the MMS.
As presented in Table 4, the MMS has a high adsorption capacity when compared to
the Qm values of a variety of adsorbents utilized from waste for TC antibiotics in aquatic
environments. From this table, the MMS can be considered as an alternative adsorbent
material for TC adsorption.
Figure 5. Adsorption kinetics of TC onto MMS at 45 C, with initial TC concentration = 100 mg/L.
Sustainability 2023,15, 4727 10 of 13
Table 3. Adsorption Kinetic Coefficients.
Model Parameter Value
pseudo-1st-order
K10.14 ±0.02
Qe1.44 ±0.02
R20.9971
SSE 0.005
χ20.00
pseudo-2nd-order
Qe1.56 ±0.003
K20.14 ±0.002
R20.9646
SSE 2.44
χ20.00
Elovich
β1.33 ±0.05
α3.90 ±0.03
R20.9559
SSE 4.20
χ24.20
For the pseudo-second-order and Elovich models, R
2
was considerably less than 1,
and SSE,
χ2
were higher. It was determined that neither model could adequately explain
the experimental findings. However, the pseudo-first-order kinetic model showed a high
value of R
2
(0.9971) and a small value of SSE,
χ2
.The closer fitting curves imply that the
pseudo-first-order adsorption process was dominant for TC adsorption on the MMS.
As presented in Table 4, the MMS has a high adsorption capacity when compared to
the Q
m
values of a variety of adsorbents utilized from waste for TC antibiotics in aquatic
environments. From this table, the MMS can be considered as an alternative adsorbent
material for TC adsorption.
Table 4. The capacity adsorption of TC with other different adsorbents’ utilization from waste.
Adsorbent Qm(mg g1) Refs.
Alkali modified magnetic biochar (MSBC) 97.962
[54]
Alkali-acid modified magnetic biochar (MSABC) 98.334
The raw biochar (RBC) 37.803
Alfalfa-derived biochar 372 [55]
Pinus taeda-derived activated biochar 274.8 [56]
Waste chicken-feather-derived multilayered
graphene-phase biochar 388.33 [57]
Clay-biochar composites 77.962 [58]
Spent coffee-ground-derived biochar 39.22 [59]
Biomass ash pyrolyzed from municipal sludge 50.75 [60]
Shrimp Shell Waste 229.98 [61]
Magnetic mesoporous silica from BA-Biomass
Power Plant 276.74 In this study
4. Conclusions
Magnetic mesoporous silica that was fabricated from rubber biomass power plant ash
was considered as a low-price adsorbent for tetracycline adsorption in aqueous solution.
The adsorption of tetracycline onto the magnetic mesoporous silica was a monolayer en-
dothermic process. The maximum tetracycline adsorption capacity of magnetic mesoporous
silica in aqueous solutions was 276.74 mg/g at 60
C. With the adsorption, the pseudo-first-
order model accurately reflects the data on adsorption kinetics, while the Langmuir model
for adsorption matches the data on adsorption isotherms well. Tetracycline adsorption
Sustainability 2023,15, 4727 11 of 13
by the MMS was spontaneous due to the negative values of the Gibbs free-energy and
enthalpy change, and at the higher temperature the adsorption was adversely affected.
The adsorbent showed exceptional properties, including tetracycline removal efficiency
and easy separation from aqueous media by a magnet. Therefore, magnetic mesoporous
silica could be used as a potential adsorbent for tetracycline from an aqueous solution.
Furthermore, this adsorbent can be utilized as an environmentally friendly adsorbent for
the treatment of wastewater.
Supplementary Materials:
The following supporting information can be downloaded at: https:
//www.mdpi.com/article/10.3390/su15064727/s1, Table S1: Chemical compositions of Bottom Ash -
Biomass Power Plant by XRF; Table S2: Chemical composition of SiO
2
extracted from Bottom Ash -
Biomass Power Plant by XRF.
Author Contributions:
Conceptualization, K.P., S.C., P.C., C.T., P.S., N.W. and T.S.; methodology,
P.T.H.H., P.S. and N.W.; validation, S.C., P.S. and N.W.; formal analysis, P.T.H.H., K.P., C.T. and P.S.;
investigation, P.C., C.T., P.S., N.W. and T.S.; resources, S.C., C.T., P.S. and N.W.; data curation, P.T.H.H.;
writing—original draft preparation, P.T.H.H.; writing—review and editing, P.T.H.H., K.P., S.C., P.C.,
N.W. and T.S.; visualization, K.P., C.T. and P.S.; project administration, N.W. and T.S. All authors have
read and agreed to the published version of the manuscript.
Funding:
This research was funded by Thailand Science Research and Innovation (TSRI), grant
number 652A01012 and Mae Fah Luang University.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Not applicable.
Acknowledgments:
We would like to acknowledge the Graduate School and Faculty of Environmen-
tal Management, Prince of Songkla University for this study and carrying out this research.
Conflicts of Interest: The authors declare no conflict of interest.
References
1.
Cyco´n, M.; Mrozik, A.; Piotrowska-Seget, Z. Antibiotics in the Soil Environment—Degradation and Their Impact on Microbial
Activity and Diversity. Front. Microbiol. 2019,10, 338. [CrossRef]
2.
Chopra, I.; Roberts, M. Tetracycline Antibiotics: Mode of Action, Applications, Molecular Biology, and Epidemiology of Bacterial
Resistance. Microbiol. Mol. Biol. Rev. 2001,65, 232–260. [CrossRef]
3.
Ötker, H.M.; Akmehmet-Balcıo˘glu, I. Adsorption and degradation of enrofloxacin, a veterinary antibiotic on natural zeolite. J.
Hazard. Mater. 2005,122, 251–258. [CrossRef]
4.
Kim, S.; Shon, H.; Ngo, H. Adsorption characteristics of antibiotics trimethoprim on powdered and granular activated carbon. J.
Ind. Eng. Chem. 2010,16, 344–349. [CrossRef]
5.
Luo, Y.; Xu, L.; Rysz, M.; Wang, Y.; Zhang, H.; Alvarez, P.J.J. Occurrence and Transport of Tetracycline, Sulfonamide, Quinolone,
and Macrolide Antibiotics in the Haihe River Basin, China. Environ. Sci. Technol. 2011,45, 1827–1833. [CrossRef] [PubMed]
6.
Yu, D.; Yi, X.; Ma, Y.; Yin, B.; Zhuo, H.; Li, J.; Huang, Y. Effects of administration mode of antibiotics on antibiotic resistance of
Enterococcus faecalis in aquatic ecosystems. Chemosphere 2009,76, 915–920. [CrossRef]
7.
Javid, A.; Mesdaghinia, A.; Nasseri, S.; Mahvi, A.H.; Alimohammadi, M.; Gharibi, H. Assessment of tetracycline contamination in
surface and groundwater resources proximal to animal farming houses in Tehran, Iran. J. Environ. Health Sci. Eng.
2016
,14, 1–5.
[CrossRef] [PubMed]
8.
Le, T.-H.; Ng, C.; Tran, N.H.; Chen, H.; Gin, K.Y.-H. Removal of antibiotic residues, antibiotic resistant bacteria and antibiotic
resistance genes in municipal wastewater by membrane bioreactor systems. Water Res.
2018
,145, 498–508. [CrossRef] [PubMed]
9.
Hamscher, G.; Sczesny, S.; Höper, H.; Nau, H. Determination of Persistent Tetracycline Residues in Soil Fertilized with Liquid
Manure by High-Performance Liquid Chromatography with Electrospray Ionization Tandem Mass Spectrometry. Anal. Chem.
2002,74, 1509–1518. [CrossRef] [PubMed]
10. Tolls, J. Sorption of Veterinary Pharmaceuticals in Soils: A Review. Environ. Sci. Technol. 2001,35, 3397–3406. [CrossRef]
11.
Zhang, Z.; Li, H.; Liu, H. Insight into the adsorption of tetracycline onto amino and amino–Fe
3+
gunctionalized mesoporous
silica: Effect of functionalized groups. J. Environ. Sci. 2018,65, 171–178. [CrossRef]
12.
Lv, J.-M.; Ma, Y.-L.; Chang, X.; Fan, S.-B. Removal and removing mechanism of tetracycline residue from aqueous solution by
using Cu-13X. Chem. Eng. J. 2015,273, 247–253. [CrossRef]
Sustainability 2023,15, 4727 12 of 13
13.
Ji, L.; Chen, W.; Duan, L.; Zhu, D. Mechanisms for strong adsorption of tetracycline to carbon nanotubes: A comparative study
using activated carbon and graphite as adsorbents. Environ. Sci. Technol. 2009,43, 2322–2327. [CrossRef] [PubMed]
14.
Wang, H.; Yao, H.; Sun, P.; Pei, J.; Li, D.; Huang, C.-H. Oxidation of tetracycline antibiotics induced by Fe(III) ions without light
irradiation. Chemosphere 2015,119, 1255–1261. [CrossRef]
15.
Zhao, C.; Deng, H.; Li, Y.; Liu, Z. Photodegradation of oxytetracycline in aqueous by 5A and 13X loaded with TiO
2
under UV
irradiation. J. Hazard. Mater. 2010,176, 884–892. [CrossRef] [PubMed]
16.
Chang, B.; Hsu, F.; Liao, H. Biodegradation of three tetracyclines in swine wastewater. J. Environ. Sci. Health Part B
2014
,49,
449–455. [CrossRef]
17.
Zazouli, M.A.; Susanto, H.; Nasseri, S.; Ulbricht, M. Influences of solution chemistry and polymeric natural organic matter on the
removal of aquatic pharmaceutical residuals by nanofiltration. Water Res. 2009,43, 3270–3280. [CrossRef]
18.
Xiang, Y.; Xu, Z.; Wei, Y.; Zhou, Y.; Yang, X.; Yang, Y.; Yang, J.; Zhang, J.; Luo, L.; Zhou, Z. Carbon-based materials as adsorbent
for antibiotics removal: Mechanisms and influencing factors. J. Environ. Manag. 2019,237, 128–138. [CrossRef]
19.
Hawker, D.W.; Chimpalee, D.; Vijuksangsith, P.; Boonsaner, M. The pH dependence of the cellulosic membrane permeation of
tetracycline antibiotics. J. Environ. Chem. Eng. 2015,3, 2408–2415. [CrossRef]
20.
Ahmed, M.J. Adsorption of quinolone, tetracycline, and penicillin antibiotics from aqueous solution using activated carbons:
Review. Environ. Toxicol. Pharmacol. 2017,50, 1–10. [CrossRef]
21.
Putra, E.K.; Pranowo, R.; Sunarso, J.; Indraswati, N.; Ismadji, S. Performance of activated carbon and bentonite for adsorption of
amoxicillin from wastewater: Mechanisms, isotherms and kinetics. Water Res. 2009,43, 2419–2430. [CrossRef] [PubMed]
22.
Gao, Y.; Li, Y.; Zhang, L.; Huang, H.; Hu, J.; Shah, S.M.; Su, X. Adsorption and removal of tetracycline antibiotics from aqueous
solution by graphene oxide. J. Colloid Interface Sci. 2012,368, 540–546. [CrossRef] [PubMed]
23.
Choi, K.-J.; Kim, S.-G.; Kim, S.-H. Removal of antibiotics by coagulation and granular activated carbon filtration. J. Hazard. Mater.
2008,151, 38–43. [CrossRef]
24.
Li, Z.; Schulz, L.; Ackley, C.; Fenske, N. Adsorption of tetracycline on kaolinite with pH-dependent surface charges. J. Colloid
Interface Sci. 2010,351, 254–260. [CrossRef]
25.
Yu, F.; Li, Y.; Han, S.; Ma, J. Adsorptive removal of antibiotics from aqueous solution using carbon materials. Chemosphere
2016
,
153, 365–385. [CrossRef] [PubMed]
26.
Mesa, M.; Sierra, L.; Patarin, J.; Guth, J.-L. Morphology and porosity characteristics control of SBA-16 mesoporous silica. Effect of
the triblock surfactant Pluronic F127 degradation during the synthesis. Solid State Sci. 2005,7, 990–997. [CrossRef]
27.
Anbia, M.; Lashgari, M. Synthesis of amino-modified ordered mesoporous silica as a new nano sorbent for the removal of
chlorophenols from aqueous media. Chem. Eng. J. 2009,150, 555–560. [CrossRef]
28.
Kresge, C.T.; Leonowicz, M.E.; Roth, W.J.; Vartuli, J.C.; Beck, J.S. Ordered Mesoporous Molecular Sieves Synthesized by a
Liquid-Crystal Template Mechanism. Nature 1992,359, 710–713. [CrossRef]
29. Wu, S.H.; Lin, H.P. Synthesis of mesoporous silica nanoparticles. Chem. Soc. Rev. 2013,42, 3862–3875. [CrossRef]
30.
Diagboya, P.N.; Olu-Owolabi, B.I.; Adebowale, K.O. Microscale scavenging of pentachlorophenol in water using amine and
tripolyphosphate-grafted SBA-15 silica: Batch and modeling studies. J. Environ. Manag. 2014,146, 42–49. [CrossRef]
31.
Sathe, T.R.; Agrawal, A.; Nie, S. Mesoporous Silica Beads Embedded with Semiconductor Quantum Dots and Iron Oxide
Nanocrystals: Dual-Function Microcarriers for Optical Encoding and Magnetic Separation. Anal. Chem.
2006
,78, 5627–5632.
[CrossRef] [PubMed]
32.
Zuo, B.; Li, W.; Wu, X.; Wang, S.; Deng, Q.; Huang, M. Recent Advances in the Synthesis, Surface Modifications and Applications
of Core-Shell Magnetic Mesoporous Silica Nanospheres. Chem. Asian J. 2020,15, 1248–1265. [CrossRef] [PubMed]
33.
Diagboya, P.N.; Dikio, E.D. Silica-based mesoporous materials; emerging designer adsorbents for aqueous pollutants removal
and water treatment. Microporous Mesoporous Mater. 2018,266, 252–267. [CrossRef]
34.
Yue, Q.; Zhang, Y.; Wang, C.; Wang, X.; Sun, Z.; Hou, X.-F.; Zhao, D.; Deng, Y. Magnetic yolk–shell mesoporous silica microspheres
with supported Au nanoparticles as recyclable high-performance nanocatalysts. J. Mater. Chem. A
2015
,3, 4586–4594. [CrossRef]
35.
Faaliyan, K.; Abdoos, H.; Borhani, E.; Afghahi, S.S.S. Magnetite-silica nanoparticles with core-shell structure: Single-step synthesis,
characterization and magnetic behavior. J. Sol-Gel Sci. Technol. 2018,88, 609–617. [CrossRef]
36.
Cho, B.H.; Nam, B.H.; An, J.; Youn, H. Municipal Solid Waste Incineration (MSWI) Ashes as Construction Materials—A Review.
Materials 2020,13, 3143. [CrossRef]
37.
Zhai, J.; Burke, I.T.; Stewart, D.I. Beneficial management of biomass combustion ashes. Renew. Sustain. Energy Rev.
2021
,151,
111555. [CrossRef]
38. Yang, G.C.; Yang, T.-Y. Synthesis of zeolites from municipal incinerator fly ash. J. Hazard. Mater. 1998,62, 75–89. [CrossRef]
39.
Fan, Y.; Zhang, F.-S.; Zhu, J.; Liu, Z. Effective utilization of waste ash from MSW and coal co-combustion power plant—Zeolite
synthesis. J. Hazard. Mater. 2008,153, 382–388. [CrossRef]
40.
Chandrasekar, G.; You, K.-S.; Ahn, J.-W.; Ahn, W.-S. Synthesis of hexagonal and cubic mesoporous silica using power plant
bottom ash. Microporous Mesoporous Mater. 2008,111, 455–462. [CrossRef]
41.
Chen, D.; Zhang, Y.; Xu, Y.; Nie, Q.; Yang, Z.; Sheng, W.; Qian, G. Municipal solid waste incineration residues recycled for typical
construction materials—A review. RSC Adv. 2022,12, 6279–6291. [CrossRef] [PubMed]
42.
Liu, Z.-S.; Li, W.-K.; Huang, C.-Y. Synthesis of mesoporous silica materials from municipal solid waste incinerator bottom ash.
Waste Manag. 2014,34, 893–900. [CrossRef] [PubMed]
Sustainability 2023,15, 4727 13 of 13
43.
Li, Y.; Wang, R.; Chen, Z.; Zhao, X.; Luo, X.; Wang, L.; Li, Y.; Teng, F. Preparation of magnetic mesoporous silica from rice husk for
aflatoxin B1 removal: Optimum process and adsorption mechanism. PLoS ONE 2020,15, e0238837. [CrossRef]
44.
Jabariyan, S.; Zanjanchi, M.A. A simple and fast sonication procedure to remove surfactant templates from mesoporous MCM-41.
Ultrason. Sonochemistry 2012,19, 1087–1093. [CrossRef] [PubMed]
45.
Ahangaran, F.; Hassanzadeh, A.; Nouri, S. Surface modification of Fe3O4@SiO2 microsphere by silane coupling agent. Int. Nano
Lett. 2013,3, 23. [CrossRef]
46.
Gu, C.; Karthikeyan, K.G. Interaction of Tetracycline with Aluminum and Iron Hydrous Oxides. Environ. Sci. Technol.
2005
,39,
2660–2667. [CrossRef]
47.
Prakasham, R.S.; Devi, G.S.; Laxmi, K.R.; Rao, C.S. Novel Synthesis of Ferric Impregnated Silica Nanoparticles and Their
Evaluation as a Matrix for Enzyme Immobilization. J. Phys. Chem. C 2007,111, 3842–3847. [CrossRef]
48.
Zhang, Z.; Lan, H.; Liu, H.; Li, H.; Qu, J. Iron-incorporated mesoporous silica for enhanced adsorption of tetracycline in aqueous
solution. RSC Adv. 2015,5, 42407–42413. [CrossRef]
49.
Suwunwong, T.; Patho, P.; Choto, P.; Phoungthong, K. Enhancement the rhodamine 6G adsorption property on Fe
3
O
4
-composited
biochar derived from rice husk. Mater. Res. Express 2020, 7. [CrossRef]
50.
Wang, J.; Guo, X. Adsorption isotherm models: Classification, physical meaning, application and solving method. Chemosphere
2020,258, 127279. [CrossRef]
51.
Bruckmann, F.D.S.; Schnorr, C.E.; Salles, T.d.R.; Nunes, F.B.; Baumann, L.; Müller, E.I.; Silva, L.F.O.; Dotto, G.L.; Rhoden, C.R.B.
Highly Efficient Adsorption of Tetracycline Using Chitosan-Based Magnetic Adsorbent. Polymers
2022
,14, 4854. [CrossRef]
[PubMed]
52.
Jiang, Z.; Hu, D. Molecular mechanism of anionic dyes adsorption on cationized rice husk cellulose from agricultural wastes. J.
Mol. Liq. 2018,276, 105–114. [CrossRef]
53. Liu, Y. Is the Free Energy Change of Adsorption Correctly Calculated? J. Chem. Eng. Data 2009,54, 1981–1985. [CrossRef]
54. Dai, J.; Meng, X.; Zhang, Y.; Huang, Y. Effects of modification and magnetization of rice straw derived biochar on adsorption of
tetracycline from water. Bioresour. Technol. 2020,311, 123455. [CrossRef]
55.
Jang, H.M.; Kan, E. A novel hay-derived biochar for removal of tetracyclines in water. Bioresour. Technol.
2018
,274, 162–172.
[CrossRef]
56.
Jang, H.M.; Yoo, S.; Choi, Y.-K.; Park, S.; Kan, E. Adsorption isotherm, kinetic modeling and mechanism of tetracycline on Pinus
taeda-derived activated biochar. Bioresour. Technol. 2018,259, 24–31. [CrossRef]
57.
Li, H.; Hu, J.; Meng, Y.; Su, J.; Wang, X. An investigation into the rapid removal of tetracycline using multilayered graphene-phase
biochar derived from waste chicken feather. Sci. Total. Environ. 2017,603–604, 39–48. [CrossRef] [PubMed]
58.
Premarathna, K.; Rajapaksha, A.U.; Adassoriya, N.; Sarkar, B.; Sirimuthu, N.M.; Cooray, A.; Ok, Y.S.; Vithanage, M. Clay-biochar
composites for sorptive removal of tetracycline antibiotic in aqueous media. J. Environ. Manag. 2019,238, 315–322. [CrossRef]
59.
Nguyen, V.-T.; Nguyen, T.-B.; Chen, C.-W.; Hung, C.-M.; Vo, T.-D.-H.; Chang, J.-H.; Dong, C.-D. Influence of pyrolysis temperature
on polycyclic aromatic hydrocarbons production and tetracycline adsorption behavior of biochar derived from spent coffee
ground. Bioresour. Technol. 2019,284, 197–203. [CrossRef]
60.
Yu, C.; Chen, X.; Li, N.; Chen, J.; Yao, L.; Zhou, Y.; Lu, K.; Lai, Y.; Lai, X. Biomass ash pyrolyzed from municipal sludge and its
adsorption performance toward tetracycline: Effect of pyrolysis temperature and KOH activation. Environ. Sci. Pollut. Res.
2022
,
29, 81383–81395. [CrossRef]
61.
Chang, J.; Shen, Z.; Hu, X.; Schulman, E.; Cui, C.; Guo, Q.; Tian, H. Adsorption of Tetracycline by Shrimp Shell Waste from
Aqueous Solutions: Adsorption Isotherm, Kinetics Modeling, and Mechanism. ACS Omega
2020
,5, 3467–3477. [CrossRef]
[PubMed]
Disclaimer/Publisher’s Note:
The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.
... The enthalpy change (∆H) of the adsorption process, as determined by the pertinent thermodynamic parameters in Table 6, was G = 23.21 kJ/mol > 0, indicating that the reaction involving the spherical adsorption of tetracycline by a spherical adsorbent was a heat-absorption reaction process, and that raising the temperature was advantageous for the adsorption to proceed [33]. The adsorption process' entropy change (∆S) of S = 84.30 ...
Article
Full-text available
The “sol–gel method” was used to prepare spherical chitosan-modified bentonite (SCB) hydrogels in this study. The SCB hydrogels were characterized and used as sorbents to remove tetracycline (TC) from aqueous solutions. The adsorbents were characterized by SEM, XRD, FTIR, TG, and BET techniques. Various characterization results showed that the SCB adsorbent had fewer surface pores and a specific surface area that was 96.6% lower than the powder, but the layered mesoporous structure of bentonite remained unchanged. The adsorption process fit to both the Freundlich model and the pseudo-second-order kinetic model showed that it was a non-monolayer chemical adsorption process affected by intra-particle diffusion. The maximum monolayer adsorption capacity determined by the Langmuir model was 39.49 mg/g. Thermodynamic parameters indicated that adsorption was a spontaneous, endothermic, and entropy-increasing process. In addition, solid–liquid separation was easy with the SCB adsorbent, providing important reference information for the synthesis of SCB as a novel and promising adsorbent for the removal of antibiotics from wastewater at the industrial level.
Article
Full-text available
As a class of broad-spectrum antibiotics, tetracyclines find extensive use in human, veterinary, and aquacultural applications. Releasing tetracycline in the form of parent or derivative compounds into the aquatic environment is extremely dangerous to human health. This study investigates the ability of silica (SiO 2 ) extracted from Beureunut beach sand to remove tetracycline in an aqueous solution by combining adsorption and Fenton-like oxidation. The beach sand was used as a precursor, and it was reacted with a sodium hydroxide solution at 80 °C before being precipitated with sulfuric acid and dried. The extraction yielded 9.22 g of silica, which was then further characterized using Fourier transform infrared (FT-IR) spectroscopy. Prior to the adsorption test, the stability of tetracycline solution was evaluated at two different temperatures (11 °C and 30 °C). The findings of a 6-day stability test performed in water showed that tetracycline was more stable at 11 °C than at 30 °C. The adsorption capacity of silica was found to be 1.68 mg/g (17.00%) at 50 mg/L tetracycline concentration after 3 hours of contact time. Meanwhile, the adsorption method combined with the Fenton-like process increased the percentage of tetracycline removal from 17.00% to 56.32%. In conclusion, combining adsorption and Fenton-like processes provides an option for greatly increasing the ability of beach sand-based silica as a potential adsorbent to remove tetracyclines from water.
Article
Full-text available
Focusing on the great potential of municipal solid waste incineration (MSWI) residues in the construction sector, the applications of recycling MSWI residues in construction materials are discussed in this review. Incineration is a promising method for managing the great quantity of municipal solid waste (MSW). Careful handling of incineration residues including fly ash, air pollution control (APC) residues, and bottom ash is required for this approach. The yield of these residues is large, and they contain many toxic and harmful substances. On the other hand, these residues contain valuable components such as SiO2, CaO, Al2O3, MgO, which are important components of building materials. Therefore, MSWI residues present huge opportunities for potential recycling and reuse in the construction and building industry. This paper summarized and discussed the application of MSWI residues in four typical building materials including cast stone, glass-ceramic, cement, and concrete. Before utilization, three types of pretreatment methods can be used to reduce the toxicity of the residues and improve the performance of the products. In addition, the current issues and the prospects of this field, and the environmental impacts of this application were discussed. It was concluded that MSWI residues can be used to prepare building materials after proper treatment which can improve the mechanical and chemical properties of the residues. The recycling can gain significant economic and environmental benefits at the same time. However, further researches on treatment methods for fine particles are needed.
Article
Full-text available
Herein, tetracycline adsorption employing magnetic chitosan (CS‧Fe3O4) as the adsorbent is reported. The magnetic adsorbent was synthesized by the co-precipitation method and characterized through FTIR, XRD, SEM, and VSM analyses. The experimental data showed that the highest maximum adsorption capacity was reached at pH 7.0 (211.21 mg g −1). The efficiency of the magnetic adsorbent in tetracycline removal was dependent on the pH, initial concentration of adsorbate, and the adsorbent dosage. Additionally, the ionic strength showed a significant effect on the process. The equilibrium and kinetics studies demonstrate that Sips and Elovich models showed the best adjustment for experimental data, suggesting that the adsorption occurs in a heterogeneous surface and predominantly by chemical mechanisms. The experimental results suggest that tetracycline adsorption is mainly governed by the hydrogen bonds and cation-π interactions due to its pH dependence as well as the enhancement in the removal efficiency with the magnetite incorporation on the chitosan surface, respectively. Thermodynamic parameters indicate a spontaneous and exothermic process. Finally, magnetic chitosan proves to be efficient in TC removal even after several adsorption/desorption cycles.
Article
Full-text available
Large amount of municipal sludge is difficult to handle; its resource utilization is an effective measure. In this study, the municipal sludge from sewage treatment plant was pyrolyzed without gas protection at different temperatures and potassium hydroxide (KOH) concentrations for activation. The pyrolysis products, named biomass ash, with higher surface area and enriched pore structures could be obtained at the pyrolysis temperature of 773 K. Moreover, the KOH activation for raw municipal sludge could further increase the surface area of the pyrolysis biomass ash. The maximum specific surface area was 44.71 m²/g, which was obtained under 2 mol/L KOH activation before pyrolysis at 773 K. And in this situation, the obtained pyrolysis biomass ash as adsorbent showed the maximum adsorption capacity of 50.75 mg/g toward tetracycline (TC). Moreover, the TC adsorption onto pyrolysis biomass ash obtained under various conditions followed the pseudo-second-order kinetic model. Adsorption thermodynamics analysis suggested the TC adsorption onto the pyrolysis biomass ash with no pre-activation was mainly due to the multi-molecule heterogeneous adsorption, while the TC adsorption onto pyrolysis biomass ash pretreated through the activation of KOH followed the monomer adsorption mechanism. This different adsorption mechanism was largely related to the pore structure, polarity, and aromaticity of the adsorbent.
Article
Full-text available
The liquid foodstuffs such as edible oil products remain a problem of excessive aflatoxin B1 (AFB1) content. This paper focused on the preparation of magnetic mesoporous silica (MMS) from rice husk ash for the removal of AFB1 in oil system. The MMS preparation process, adsorption conditions, structural characteristics, and adsorption mechanism were investigated. The optimum conditions for MMS preparation were pH 11.0 and 80°C for 24 h. The characterization results showed that magnetic particles were successfully embedded in the MMS and had high responsiveness to a magnetic field, which was advantageous for recyclability. The MMS had ordered uniform channels with a specific surface area of 730.98 m²/g and pore diameter of 2.43 nm. The optimum adsorption conditions were 2 h at 20°C. For AFB1 with an initial concentration of 0.2 μg/mL, the MMS adsorption capacity was 171.98 μg/g and the adsorption rate was 94.59%. The MMS adsorption isotherm fitted the Langmuir model well under the assumption of monolayer AFB1 adsorption with uniformly distributed adsorption sites on the MMS surface. The maximum amount of AFB1 adsorbed according to the Langmuir isotherm was 1118.69 μg/g. A quasi-second-order kinetic model gave a better fit to the process of AFB1 adsorption on MMS. The values of ΔH (−19.17 kJ/mol) and ΔG (−34.09, −34.61, and −35.15 kJ/mol at 283, 293, and 303 K, respectively) were negative, indicating that AFB1 adsorption on MMS was a spontaneous exothermic process. The results indicated that MMS was a promising material for AFB1 removal in oil system, and this study will serve as a guide for practical MMS applications.
Article
Full-text available
Over the past decades, extensive studies on municipal solid waste incineration (MSWI) ashes have been performed to develop more effective recycling and waste management programs. Despite the large amount of research activities and the resulting improvements to MSWI ashes, the recycling programs for MSWI ashes are limited. For instance, although the U.S. generates more MSWI ashes than any other country in the world, its reuse/recycle programs are limited; bottom ash and fly ash are combined and disposed of in landfills. Reuse of MSWI ashes in the construction sectors (i.e., geomaterials, asphalt paving, and concrete products) as replacements for raw materials is one of most promising options because of the large consumption and relatively lenient environmental criteria. The main objective of this study was to comprehensively review MSWI ashes with regard to specific engineering properties and their performance as construction materials. The focus was on (1) the current practices of MSWI ash management (in particular, a comparison between European countries and the U.S.), (2) the engineering properties and performance of ashes when they are used as substitutes of construction materials and for field applications, and (3) the environmental properties and criteria for the use of MSWI ashes. Overall, the asphalt and concrete applications are the most promising, from both the mechanical and leachate viewpoints. However, cons were also observed: high absorption of MSWI ash requires a high asphalt binder content in hot-mix asphalt, and metallic elements in the ash may generate H2 gas in the high-pH environment of the concrete. These side effects can be predicted via material characterization (i.e., chemical and physical), and accordingly, proper treatment and/or modified mix proportioning can be performed prior to use.
Article
Use of biomass for energy production is increasing, so management of the resultant ash is important. This review compares current and future production, chemical composition, and reuse options for ash from common feedstocks (agricultural residues, energy crops, woody biomass, forest residues, recovered wood, paper sludge, sewage sludge and municipal solid waste). Global production is ~170 Mt/yr, but could increase to ~1000 Mt/yr if all available biomass were exploited. Current production is dominated wood and waste derived ashes, but there is capacity to greatly increase use of agricultural residues. Combustion of virgin biomass in modern furnaces can produce ash with negligible persistent organic pollutants and low contaminant metals concentrations, so application to land is possible. Agricultural residue ashes contain abundant potassium and useful phosphate, so could potentially be used as fertiliser. Forestry ashes are rich in CaO, but slightly higher contaminant metals levels may restrict their use to forestry soils. Recovery of potassium from these ashes, and their use in cementitious materials have also been demonstrated. Biomass containing waste ashes potentially contain more persistent organic pollutants and contaminant metals. However, municipal solid waste bottom ash is routinely used as a construction aggregate for prescribed applications. Paper sludge ash is suitable for restricted use as a soil conditioner and possibly as a secondary pozzolan. However, controlled disposal may be required for recovered wood ash and sewage sludge incineration ash. As persistent organic pollutants tend to partition to the flue gases, fly ash and air-pollution control residues are likely to require controlled disposal.
Article
The main objective of this study was to evaluate the effect of different pyrolysis temperatures on the formation of polycyclic aromatic hydrocarbons (PAHs) in biochar originated spent coffee ground (SCG) and the tetracycline (TC) adsorption behavior of biochar in water. The results showed that biochar synthesized at 500°C (SCG 500) contained low PAHs (600 µg kg −1) and the highest TC adsorption efficiency. In addition, the characteristics, influencing factors on TC adsorption, and the related mechanisms of SCG 500 were comprehensively investigated. The results showed that the highest efficiency was observed at pH of 7 and the presence of ions in salinity solution reduced the adsorption capacity of SCG 500. The electrostatic interaction, hydrogen bonding, and π-EDA were the major adsorption mechanisms. Safety PAHs level, low-cost, widely material sources and high TC removal capacity suggested that SCG 500 was a promising environmentally friendly effective absorbent.
Article
Adsorption is widely applied separation process, especially in environmental remediation, due to its low cost and high efficiency. Adsorption isotherm models can provide mechanism information of the adsorption process, which is important for the design of adsorption system. However, the classification, physical meaning, application and solving method of the isotherms have not been systematical analyzed and summarized. In this paper, the adsorption isotherms were classified into adsorption empirical isotherms, isotherms based on Polanyi's theory, chemical adsorption isotherms, physical adsorption isotherms, and the ion exchange model. The derivation and physical meaning of the isotherm models were discussed in detail. In addition, the application of the isotherm models were analyzed and summarized based on over 200 adsorption equilibrium data in literature. The statistical parameters for evaluating the fitness of the models were also discussed. Finally, a user interface (UI) was developed based on Excel software for solving the isotherm models, which was provided in supplemental material and can be easily used to model the adsorption equilibrium data. This paper will provide theoretical basis and guiding methodology for the selection and use of the adsorption isotherms.
Article
The hierarchically structured core-shell magnetic mesoporous silica nanospheres (Mag-MSNs) have attracted extensive attention, particularly in the studies of reliable preparation methods of the multifunctional materials and the diverse applications of these nanomaterials across a wide range of multidisciplinary researches. These Mag-MSNs have been prepared with well-designed synthesis strategies and used as adsorbent materials, biomedicine and catalysis due to their excellent magnetic responsiveness, large specific surface area and readiness in their surface modification. The surface functionalization of the core-shell Mag-MSNs bridges the gap between heterogeneous and homogeneous phases and greatly extends the properties and application prospects of the materials. The introduction of various molecular matrices into the shell of Mag-MSNs enables the combination of functionalization and magnetic separation technology. In some cases, it is an important feature in the sustainable development process to recover the functionalized core-shell Mag-MSNs after the reaction and reuse them without losing activity. In this review, the design strategies and the construction of the core-shell Mag-MSNs are discussed. Various surface functionalization methods of core-shell Mag-MSNs are summarized and the latest applications of these functionalized nanomaterials in the fields of biomedicine, catalysis and wastewater treatment are demonstrated.