DR - Jibran Paper
DR - Jibran Paper
DR - Jibran Paper
Ali Maged, Jibran Iqbal, Sherif Kharbish, Ismael Sayed Ismael, Amit
Bhatnagar
PII: S0304-3894(19)31274-9
DOI: https://doi.org/10.1016/j.jhazmat.2019.121320
Reference: HAZMAT 121320
Please cite this article as: Maged A, Iqbal J, Kharbish S, Ismael IS, Bhatnagar A, Tuning
tetracycline removal from aqueous solution onto activated 2:1 layered clay mineral:
characterization, sorption and mechanistic studies, Journal of Hazardous Materials (2019),
doi: https://doi.org/10.1016/j.jhazmat.2019.121320
This is a PDF file of an article that has undergone enhancements after acceptance, such as
the addition of a cover page and metadata, and formatting for readability, but it is not yet the
definitive version of record. This version will undergo additional copyediting, typesetting and
review before it is published in its final form, but we are providing this version to give early
visibility of the article. Please note that, during the production process, errors may be
discovered which could affect the content, and all legal disclaimers that apply to the journal
pertain.
Ali Maged1,2,*, Jibran Iqbal3, Sherif Kharbish2, Ismael Sayed Ismael2, Amit Bhatnagar1
1
Department of Environmental and Biological Sciences, University of Eastern Finland, P.O. Box
2
Geology Department, Faculty of Science, Suez University, El Salam City, P.O. Box 43518, Suez
of
Governorate, Egypt
ro
3
College of Natural and Health Sciences, Zayed University, P.O. Box 144534, Abu Dhabi, United
Arab Emirates
-p
re
lP
Graphical abstract
ur
Jo
1
Raw Bentonite Pre-treatment 500 C
of
O O
O O
Si Si
- O
-
O
TC+
ro
After Adsorption Tetracycline
-p
Thermal Activated Bentonite
re
lP
Highlights
na
Egyptian bentonite clays (raw (BC) and modified (TB)) were studied for TC adsorption.
The BET surface area of TB increased more than two-fold after modification.
ur
Adsorption capacity of BC and TB was 156.7 and 388.1 mg g-1, respectively for TC.
Jo
2
Abstract
threat which results in the development of antibiotic-resistant bacteria/resistance genes in the aquatic
environment. Therefore, robust and cost-effective methods are required to address this problem. In
this study, thermal activation was opted for the modification of natural bentonite clay (BC) and
utilized to investigate the adsorptive removal of tetracycline (TC) from aqueous solution. The
physicochemical surface properties of the raw and modified bentonite samples were also investigated.
The BET analysis revealed that the thermally activated bentonite (TB) has better properties than BC.
of
The surface area of TB was found to be more than two-fold higher compared to that of BC. The FTIR
ro
spectra exhibited the existence of Al—OH, Si—O and Si—O—Si functional groups in the samples,
confirming the presence of hydrated aluminosilicate in the clay. The effects of various operating
-p
parameters were analyzed via optimization studies. The maximum monolayer adsorption capacity
estimated by Langmuir model was found to be 156.7 and 388.1 mg g-1 for BC and TB, respectively.
re
Furthermore, fixed-bed column studies were performed to get insights into the adsorption behavior
lP
1. Introduction
3
Currently, more than two billion people do not have accessibility to clean drinking water and more
than four billion people are without safe sanitation (Hokkanen et al., 2018). A large number of health
problems have emerged due to the arbitrary discharge of pollutants into the environment. Therefore,
the demand for sufficient clean water is increasing around the world. World Health Organization
(WHO) has shown concerns about contaminated water with pharmaceutical compounds (WHO,
1997). Pharmaceutical compounds are one of the emerging contaminants, which have existed in low
concentrations in the environment and can have negative effects on the human health and ecosystems,
of
et al., 2019). While the concentrations of pharmaceutical compounds in raw domestic wastewater are
usually reported in the range of 100 ng L-1 to 10 μg L-1, their concentration in hospital and
ro
pharmaceutical industry effluents can reach up to 100-500 mg L-1 (Cetecioglu et al., 2013).
-p
Consequently, the effective control and removal of these compounds would provide greatly beneficial
group of broad-spectrum antibiotics extensively used for human therapy, aquaculture and veterinary
medicines in the treatment of various bacterial infectious diseases (Chang and Ren, 2015), and also
na
as one of the growth promoters in livestock-farming (Nasseri et al., 2017). TCs are poorly absorbed
(up to 50 %) in the humans and digestive tracts of animals, the unadsorbed fraction of TCs are
ur
excreted through urine and stool as an unmetabolized original compound (Sarmah et al., 2006). TCs
are excessively hydrophilic, amphoteric and ionizable compounds (Wang et al., 2008). The attraction
Jo
of TC towards solid surface strongly depends on the solution pH. Moreover, TC molecule contains
different ionizable groups depending on the solution pHs as shown in Fig. 1 (a,b). In strongly acidic
conditions (below 3.3), TC becomes positively charged (TCH3+ ) and exists as a cationic species as
shown in Fig. 1 (b). This behavior could be attributed to the dimethyl ammonium group protonation.
While, TC exists as a zwitterion (TCH2±) due to loss of a proton from the phenolic diketone moiety in
4
the range between pH 3.3 to 7.7. However, TC at pH more than 7.7 exists as monovalent (TCH − ) or
divalent (TC2− ) anions. This is due to the proton loss from tricarbonyl system and phenolic diketone
Many technologies have been developed for removing different pollutants from water including
adsorption (Alqadami et al., 2017), electrochemical technology (Gatsios et al., 2015), membrane
separation (Giacobbo et al., 2017) and hybrid technologies (Hokkanen et al., 2018). Among these
technologies, the adsorption method has gained wide attention for their effectiveness, efficiency, low
operational cost and has a wide range of source materials. Numerous adsorbents, for instance,
of
activated carbon (Choi et al., 2008), carbon nanotubes (Ncibi and Sillanpää, 2015), biochar (P. Liu et
ro
al., 2012), ion exchange resins (Wang et al., 2017) and clay minerals (N. Liu et al., 2012) have been
reported for the removal of antibiotics. However, the separation and discharge of small colloidal
-p
adsorbents from treated water could induce another separation problem which must be addressed. In
re
this regard, clay adsorbents present a good alternative adsorbent material option because of the ease
Clays are considered as important adsorbents for the removal of pharmaceutical compounds from
aqueous medium due to their high specific surface area, sorption capacity and cation exchange
na
capacity (CEC). Several clays and clay minerals such as smectite, illite, zeolite, kaolinite, and
bentonite have been extensively used as adsorbents for the removal of various pollutants from
ur
contaminated water. Bentonite is a 2:1 type layered-structure clay mineral consisting mostly of
montmorillonite. Its basic structure includes one alumina (Al3+) octahedral sheet and two silica (Si4+)
Jo
tetrahedral sheets. Naturally, bentonite possesses negative surface charge owing to the cation’s
substitution (isomorphic) in its crystal lattice, for e.g. substitution of Mg2+ for Al3+ in the alumina
octahedral sheet and Al3+ for Si4+ in the silica tetrahedral sheet. Owing to its properties, bentonite has
been utilized for the removal of both anionic (Chen et al., 2018) and cationic (Sen Gupta and
5
In ancient times, heat-treated bentonite and clays were used as bricks for construction buildings and
soil stabilization. Thermal treatment changes several physicochemical properties of bentonite, such
as particle size, permeability, water content, cohesion and specific gravity (Abu-Zreig et al., 2001).
Dehydration and dehydroxylation processes during the thermal treatment of bentonite could be
attributed to the octahedral cation movements within the octahedral sheet (Odom, 1984).
The aim of this study was to prepare and characterize the Egyptian raw and modified bentonite, and
to evaluate the possibility of using bentonite as an inexpensive adsorbent for the effective
decontamination of water polluted with tetracycline antibiotics. Thermal activation technique was
of
applied for bentonite modification, in order to increase and attain maximum adsorption capacity.
ro
Additionally, investigation of adsorbent’s structural characteristics (raw and modified forms) was
also conducted. In batch mode, the effect of various operating conditions such as adsorbent dosage,
-p
initial solution pH, contact time, initial TC concentration, and ionic strength was studied. The column
experiments were carried out to examine the influence of flow rate, adsorbent loading, and influent
re
TC concentration. The adsorption kinetics and isotherms were determined using widely-used models.
lP
The possible adsorption mechanism of tetracycline on the activated bentonite was also successfully
explored.
na
used as an adsorbent. The bentonite sample was crushed, ground and sieved to a particle size less
than 160 µm. The clay mineral was washed with Milli-Q water and dried.
2.2. Reagents
The pharmaceutical solution was prepared using tetracycline (TC, purity grade ≥98.0 %). Hydrogen
peroxide (H2O2; 30% w/w), acetic acid (Acc), sodium hydroxide (NaOH), silver nitrate (AgNO3),
6
sodium acetate (Na-Ac) and sodium chloride (NaCl) were of analytical grade and purchased from
Sigma-Aldrich in purities of at least 98% and used without further treatment. The TC solutions were
prepared by diluting TC stock solution. Milli-Q water (18.2 MΩ.cm) (MQW) was used to prepare all
solutions throughout the experiments. Fig. 1 (d) summarizes the physicochemical properties of TC
of
In order to remove the organic matters and carbonates, 25 g of bentonite clay was treated with 400
mL of 0.1 N Na-Ac and Acc (pH=5.0) and 100 mL H2O2 was added gradually. The slurry was stirred
ro
with magnetic stirring for 12 h at 70 ± 1 °C followed by 24 h at room temperature. The suspension
-p
was centrifuged at 5000 rpm for 5 min. The residue was washed three times with 0.01 N NaCl
solution. Thereafter, the suspension was centrifuged again at 5000 rpm for 15 min (model 5430,
re
Eppendorf AG, Germany) and filtered. Next, the residue was washed three times with 0.01 N NaCl
solution followed by three times with MQW and then dried at 105 ± 1 °C for 24 h. Thereafter, the
lP
bentonite (labeled as ‘BC’) was ground, sieved using 100 µm sieve and stored in a desiccator until
further use. The BC sample was heated (muffle furnace) at 500 ±1 °C for 4 h. The optimization
na
condition for the thermal activation of bentonite sample has been given in supplementary material
(section S1 and Fig. S1). The aim of the thermal activation was to increase the adsorbent’s stability
ur
and ion exchange capacity (Oliveira et al., 2019). Thereafter, the sample was cooled down to room
Jo
temperature in a desiccator over silica gel overnight. Afterwards, the sample was labeled as TB and
X-ray diffraction (XRD) analysis of BC and TB samples was conducted using Shimadzu, LabX XRD-
6100 with Cu-Kα radiation. The accelerating voltage was set as 40 kV, and the current was 30 mA,
7
with a monochromatic Cu Kα radiation wavelength (λ = 1.5405 Å). The scanning was limited from
3 to 70° on the 2θ angle. In order to determine the interlayer space of bentonite, the d-values was
calculated according to Bragg´s law (Bragg and Bragg, 1913) using the following eq. (1).
𝑛 𝜆 = 2𝑑 sin 𝜃 (1)
Fourier transform infrared (FT-IR) analysis of BC and TB samples was performed using Cary 660
FTIR spectrometer, Agilent technologies, USA, in order to determine the changes in the functional
groups before and after thermal treatment. The FTIR spectra of the samples were obtained within the
range of 4000–600 cm-1 with 256 scans per sample at 4 cm-1 resolution. Scanning electron microscopy
of
(SEM) analysis (TESCAN VEGA (LMU) SEM), was used for the determination of surface
ro
morphology of BC and TB at different magnifications ranging from 500 x to 5 kx. Energy-dispersive
X-ray (EDX) analysis was conducted for investigating the chemical composition of samples and to
-p
determine the alteration in chemical composition before and after the thermal activation using
INCAx-act (Oxford Instruments, UK) EDX operating at 20 kV. X-ray fluorescence (XRF) analysis
re
was done to gain insight into the chemical composition of BC and TB samples using Philips PW
lP
1410. The N2 adsorption/desorption measurements for the evaluation of the textural properties
(surface area, total pore volume, and pore diameter) were performed at -196 °C using a Belsorp Mini
na
II (Japan). The point of zero charge (pHzpc) of BC and TB was determined by pH drift method. The
initial pH of 10 mM NaCl solution was adjusted in the range of 2 – 10. A measured amount of 0.05
ur
g adsorbent was dispersed in 25 mL of pH adjusted solution and agitated at 200 rpm in the shaker
(multi-functional orbital shaker, Grant-Bio, UK), at constant temperature 25 ± 1 °C for 24 h in glass
Jo
bottles. The deviation in solution pH was recorded as the final pH after the aforementioned time. The
pHzpc of the adsorbents was determined by plotting the change in pH versus the initial pH. The cation
exchange capacity (CEC) of the samples was evaluated according to the ASTM C 837-81 (Standard
Test Methods of Methylene Blue (MB) Index of Clay)) (Calábria et al., 2013).
8
2.5. Batch Adsorption studies
Batch experiments were conducted with BC and TB adsorbents to determine their potential towards
TC removal from aqueous solution. Equilibrium studies were carried out using a known volume of
30 mg L-1 TC solution to which 0.4 g L-1 adsorbent was added and the adsorbent-adsorbate solution
was agitated at 200 rpm in a shaker for 300 min at ambient temperature. Optimization studies were
conducted to gain insights into the effect of various experimental conditions on the adsorption of TC
onto BC and TB. The effect of pH (3 - 9), adsorbent dosage (0.1 - 1.0 g L-1), contact time (0 - 300
min), initial TC concentration (5 - 200 mg L-1) and ionic strength (0.05 - 0.5 M) was analyzed during
of
the adsorption process. The solution pH was adjusted by the addition of 0.1 M NaOH and/or 0.1 M
ro
HCl before the adsorbents were added. All the experiments were done in duplicate. After
equilibration, the suspension was filtered with 0.45 μm cellulose acetate membrane syringe driven
-p
filter (Sartorius, Germany). The residual concentration of TC was analyzed by UV-visible
level exhibits a linear relationship (R2 = 0.9998) over 0.05-30 mg L-1 concentration range. The amount
of TC adsorbed per unit mass of adsorbent (𝑞𝑒 ; mg g-1) and removal efficiency (𝑅; %) were obtained
na
𝐶𝑖 −𝐶𝑒
𝑞𝑒 = ∗𝑉 (2)
𝑚
ur
𝐶𝑖 −𝐶𝑒
𝑅 (%) = ∗ 100 (3)
𝐶𝑖
Jo
For regeneration studies, the experiments were performed with BC and TB at the optimum conditions
based on the optimization studies. The adsorbents were added to a solution containing 30 mg L-1 TC
and shaken for 5 h to achieve equilibrium. Then, the separated adsorbents were dried at 60 °C for 4 h.
NaOH (0.1 M) was selected as the regeneration agent. Five cycles of adsorption/desorption were
9
2.6. Theoretical models
Kinetic studies of pharmaceutical removal by an adsorbent are important for evaluating the rate and
mechanism of adsorption. In this study, pseudo-first-order kinetic model (PFO) (Lagergren, 1898)
(Eq. (4)), pseudo-second-order kinetic model (PSO) (Ho and McKay, 1999) (Eq. (5)) and Avrami
kinetic model (AV) (Avrami, 1939) (Eq. (6)) were applied to the experimental data using non-linear
of
regression in MATLAB software (R2018b).
𝑞𝑡 = 𝑞𝑒 (1 − 𝑒 −𝑘1 𝑡 )
ro
(4)
𝑘2 𝑞𝑒2 𝑡
𝑞𝑡 = -p (5)
1+𝑘2 𝑞𝑒 𝑡
𝑛
𝑞𝑡 = 𝑞𝑒 (1 − 𝑒 (−(𝑘𝐴𝑣 𝑡) 𝐴𝑉)
) (6)
re
where, 𝑞𝑒 and 𝑞𝑡 (mg g-1) are the adsorption capacity at equilibrium at time t (min) respectively, 𝑘1
lP
(min-1) is the rate constant of PFO, 𝑘2 (g mg-1min-1) is the rate constant of PSO and 𝑘𝐴𝑣 is the constant
of AV (min-1).
na
The adsorption isotherm represents adsorbent and adsorbate interaction at variable initial
Jo
concentrations. Experimental adsorption data were fitted to Langmuir (Langmuir, 1918), Freundlich
(Freundlich, 1924), Sips (Sips, 1948) and Redlich-Peterson (Redlich and Peterson, 1959) (Eqs 7-11),
𝑞𝑚 𝐾𝐿 𝐶𝑒
𝑞𝑒 = 1+𝐾𝐿 𝐶𝑒
(7)
10
1⁄
𝑛
𝑞𝑒 = 𝐾𝐹 𝐶𝑒 (8)
𝛽𝑠
𝐾𝑠 𝐶𝑒
𝑞𝑒 = 𝛽𝑠 (9)
1+𝑎𝑠 𝐶𝑒
𝐾 𝐶𝑒
𝑞𝑒 = 1+𝛼𝑅 𝑔 (10)
𝑅 𝐶𝑒
where, 𝑞𝑒 (mg g-1) is the adsorption capacity of BC and TB at the equilibrium time, 𝐶𝑒 is the
equilibrium concentration (mg L-1), 𝑞𝑚 (mg g-1) is the Langmuir constant associated with adsorption
capacity at its maximum adsorption capacity (mg g-1), 𝐾𝐿 is the constant of Langmuir (L mg-1), 𝐾𝐹 is
of
the constant of Freundlich (mg g-1), 𝑛 is the exponent of Freundlich related to adsorption intensity
and 𝐾𝑠 (L mg−1) is the Sips constant. 𝐾𝑅 is the Redlich-Peterson constant (L g-1), 𝛼𝑅 is the Redlich-
ro
Peterson constant (mg-1), and g is the Redlich-Peterson exponent.
-p
The correlation coefficient (R2) was used to designate the best fit of the experimental data with the
re
models. R2 and root mean square error (RMSE) were used for comparing the performance of different
models. The lower RMSE value defines the data more precisely for this model. The RMSE can be
lP
∑𝑁
1 ( 𝑞𝑒𝑥𝑝 −𝑞𝑚𝑜𝑑𝑒𝑙 )
2
𝑅𝑀𝑆𝐸 = √ (11)
na
where, 𝑞𝑒𝑥𝑝 and 𝑞𝑚𝑜𝑑𝑒𝑙 are the experimentally measured value and model prediction for TC
ur
adsorption, respectively.
Jo
The fixed-bed column studies were conducted with TB using laboratory-scale glass column (column
length of 130 mm and an internal diameter of 5 mm) assembly. The adsorbent was sandwiched by
glass wool layer to prevent the adsorbent loss, and tightly closed for good distribution of the liquid
11
phase. Two different adsorbent loadings of 25 and 50 mg were employed to study the effect of varying
column beds, respectively. Two different flow rates (1.5 and 3 mL min-1) were operated to elucidate
the effect of influent flow. Similarly, two different solutions having TC concentration of 5 and 10 mg
L-1 were utilized to explicate the effect of influent concentration (Co). During the experiment, effluent
samples were collected at different time intervals. Thereafter, the samples were analyzed in order to
determine the final TC concentration in the effluent solutions. Finally, the column operation was
stopped when there was no further TC adsorption, i.e. when the influent and effluent TC
of
ro
2.7.2. Column data analysis
-p
The effluent samples were analyzed for TC concentration using UV-visible spectrophotometer. The
performance of fixed-bed column is represented by the breakthrough curve. The determination of the
re
operation and dynamic response of the column adsorption depends on the time of breakthrough
appearance and the breakthrough shape. The breakthrough point (tb) is reached when the effluent
lP
concentration (Ct/Co) > 0.05 of Co occurs. The point of column exhaustion (te) is reached when the
effluent concentration reaches 95 % at a constant value of Ct/Co (Kundu et al., 2004). Ct/Co as a
na
function of time is used to elucidate the breakthrough curve. While, the total effluent volume ( 𝑉𝑒𝑓𝑓
(mL)) passed through column can be calculated using Eq. (12). The total quantity of TC adsorbed
ur
(𝑞𝑡𝑜𝑡𝑎𝑙 (mg)) by fixed-bed column, can be estimated using Eq. (13) from the area under the
Jo
breakthrough curve (Chen et al., 2012). From 𝑞𝑡𝑜𝑡𝑎𝑙 value, the experimental maximum uptake
capacity (𝑞𝑏𝑒𝑑 , mg g-1) of the column can be calculated as shown in Eq. (14).
𝑄 𝑡=𝑡𝑜𝑡𝑎𝑙
𝑞𝑡𝑜𝑡𝑎𝑙 = 1000 ∫𝑡=0 𝐶𝑎𝑑 𝑑𝑡 (13)
12
𝑞𝑡𝑜𝑡𝑎𝑙
𝑞𝑏𝑒𝑑 = (14)
𝑚
where, 𝑄 is flow rate (mL min-1), 𝑡𝑡𝑜𝑡𝑎𝑙 is total flow time (min) and 𝐶𝑎𝑑 is the adsorbed TC
The total amount of TC molecules passed throw the column system (𝑚𝑡𝑜𝑡𝑎𝑙 ) is calculated using Eq.
(15). While the removal percentage (𝑅𝐸, %) of TC molecules can be calculated from Eq. (16). At
equilibrium, the unadsorbed TC concentration (𝐶𝑒𝑞 ; mg L-1) during the column process can be
of
𝐶𝑜 𝑄𝑡𝑡𝑜𝑡𝑎𝑙
𝑚𝑡𝑜𝑡𝑎𝑙 = (15)
1000
ro
𝑞
𝑅𝐸(%) = 𝑚𝑡𝑜𝑡𝑎𝑙 × 100 (16)
𝑡𝑜𝑡𝑎𝑙
𝐶𝑒𝑞 =
𝑚𝑡𝑜𝑡𝑎𝑙 − 𝑞𝑡𝑜𝑡𝑎𝑙
𝑉𝑒𝑓𝑓
× 1000
-p (17)
re
lP
The adsorption mechanism of TC onto thermally activated bentonite was studied based on the
na
characterization of the adsorbent before and after TC adsorption process. TB adsorption experiments
were conducted according to the optimum conditions of the batch experiments. The contaminated TB
ur
sample (after sorption) was filtered. Subsequently, the samples were dried overnight without any
further treatment at 70 ± 1 °C in an air oven. The contaminated TB sample then characterized with
Jo
the assist of XRD, FTIR and SEM analysis. Thereafter, the characterization data then compared with
13
3.1. Characterization studies
The chemical composition of the bentonite clay before and after thermal treatment are presented in
Table 1. The chemical analysis showed that SiO2 and Al2O3 were the major constituents of studied
adsorbents. In addition to these, Fe2O3, K2O, CaO, MgO, TiO2, and P2O5 occur as minor metallic
oxides. The ratio of Al2O3/SiO2 was found to be around 0.36 and 0.37 for BC and TB, respectively.
These values indicated that the bentonite used in this study contained mainly montmorillonite mineral
of
(Chinoune et al., 2016). Fe2O3 content was high in BC and TB samples, and this could explain the
brownish and reddish colors in the raw and thermally treated bentonite, respectively (Table 1). After
ro
the thermal treatment, there was a decrease in loss on ignition (L.O.I.) from 9.93 % (BC) to 2.84 %
(TB). The high L.O.I. value in BC could be due to organic matter and anhydrous material content,
-p
which was expelled from the sample during ignition. However, lower L.O.I. value in TB could
re
indicate that high loss on ignition happened during the thermal treatment (Saikia and Parthasarathy,
2010). Overall, the presence of high SiO2, Al2O3, and L.O.I. values in the clay adsorbent proved that
lP
the main constituent of the studied bentonite clay is silica, alumina, and water (Aytas et al., 2009).
na
In this study, XRD diffractograms were obtained to evaluate the changes in the layered structure of
the clay adsorbent before and after thermal activation. The XRD patterns of BC and TB are illustrated
Jo
in Fig. 2 (a). Natural bentonite is a member of the smectite clay family which is mainly composed of
montmorillonite minerals (Choo and Bai, 2016). The XRD peaks confirmed the presence of
montmorillonite and quartz. Kaolinite peak was also found in BC, while this peak disappeared in the
thermally activated bentonite. The main characteristic pattern of the montmorillonite peak was
detected at 2θ value of 5.5°. The montmorillonite peaks also appeared in the natural bentonite at 2θ
14
value of 20° and 36°; similar observations have been previously reported in the literature (Terzić et
al., 2016). Concurrently, some non-clay minerals were also recorded at 2θ value of 22° and 26° for
quartz. After the thermal activation of bentonite (TB), no significant difference was found in XRD
pattern compared to BC pattern (Fig. 2 (a)). However, the appearance of the patterns differed only in
the main montmorillonite peak at 2θ value of 5.5°, where the intensity of this peak was reduced
significantly, and a slight shift occurred towards the right at 2θ value of 8.3°. The shift of 001 lattice
plane led to decrease in the interlayer space of TB. According to Bragg's law, the interlayer distance
decreased from 1.61 nm for BC to 1.07 nm for TB. This result could be attributed to the water
of
molecule removal from the crystal lattice sheets. Moreover, this result also indicates that the high
temperature which was used for the thermal activation process, caused distortions in crystalline
ro
structure (Cantuaria et al., 2016). This phenomenon has been verified in other studies where it has
-p
been reported that during the process of calcination (temperature ≥ 400 °C), the edges of the bentonite
layers get damaged, while the layer structure remains intact (Bertagnolli et al., 2011). Therefore, TB
re
becomes more stable in the aqueous media compared to BC, making TB a better alternative for a
lP
fixed-bed column studies and solving the expansion problem of natural bentonite (Danková and
Dolinská, 2012). Based on the XRD analysis, it can be deduced that thermal activation of bentonite
increases the binding adsorption sites and the specific surface area of bentonite without changing the
na
layer structure.
ur
Fig. 2 (b) shows the FTIR spectra of BC and TB adsorbents. The main characteristic bands of the
Jo
natural bentonite were observed at 3620, 993, 910, and 794 cm-1 (Zazoua et al., 2013). The bands at
3620 and 3386 cm-1 could be attributed to the O–H stretching vibration of the silanol (Si-OH) groups
from the solid and HO–H vibration of the water molecules adsorbed on the silicate, respectively
(Yaming et al., 2016). The Si-O-Si stretching vibration was observed as a strong band at 993 cm-1.
The presence of a band at 910 cm-1 represents Al–OH–Al hydroxyl bending vibration (Tyagi et al.,
15
2006). The band at 794 cm-1 could be attributed to the stretching vibration of Si–O–Fe (Zuo et al.,
2017). The existing band at 1634 cm-1 reflects the bending of HO–H bond of water molecules, which
is retained in the silicate matrix. After the thermal activation, the significant difference between the
two spectra was in the bands around 900, 1600 and 3600 cm-1. The TB spectra peaks at those
wavelengths were significantly decreased compared to the BC spectra peaks. This reduction indicated
the loss of water content in TB. However, the band at 3702 cm-1 completely disappeared after the
thermal activation for TB. The reduced band at 1634 cm-1 indicates the reduction of the H-O-H
deformation group, while the decrease in the bands at 3386 and 3702 cm-1 revealed the reduction of
of
the O-H stretching group. This phenomenon occurred due to the dehydration and dehydroxylation
processes which were triggered during the calcination process (Aytas et al., 2009).
ro
3.1.4. The nitrogen adsorption/desorption measurements
-p
re
The nitrogen adsorption-desorption isotherms of BC and TB are presented in Fig. 2 (c). The BC and
TB adsorbents were of IV-type isotherm (Thommes et al., 2015). This isotherm type is characteristic
lP
of mesoporous materials (average pore diameter 2-50 nm) (Budsaereechai et al., 2012). There is no
significant difference in both curves for BC and TB. This indicates that the mesoporous character is
na
preserved in the natural bentonite after the thermal treatment. In addition, the similar hysteresis loops
were also observed for both BC and TB, which indicates occupation and evacuation of the mesopores
ur
by capillary condensation (Thommes et al., 2015). Table 2 summarizes the textural properties of BC
Jo
and TB using the standard BET method, Langmuir and BJH equations. The obtained results
demonstrated that the thermal activation process has significantly elevated the surface properties of
raw bentonite. The TB was found to have more than two-fold higher surface area than BC (Table 2).
This increase in TB surface area could be attributed to the removal of adsorbed water molecules and
volatile organic compounds attached on the bentonite surface (Toor et al., 2015).
16
3.1.5. pHzpc measurements
The presence of an excess of H+ ions or OH- ions in the solution may alter the surface charge
characteristics of the adsorbent. The adsorbent carries positive surface charge below its point of zero
charge (pHzpc) and attracts negatively charged species. On the contrary, the adsorbent carries negative
surface charge above the pHzpc and is suitable for adsorption of positively charged species (Divband
Hafshejani et al., 2016). The pHzpc as depicted in Fig. 2 (d), of bentonite (BC) and thermally treated
of
bentonite (TB) was found to be 6.01 and 5.82, respectively.
ro
3.1.6. EDX and SEM analysis
-p
As shown in Fig. 3 (a,b), the most abundant constituents in BC and TB are O, Si, and Al; which are
re
the basic constituents of smectite clay group (Araujo et al., 2013). Additionally, the presence of Fe,
K, Ca, Mg and Na could be observed. The presence of these cations in the bentonite samples could
lP
be designated as polycationic bentonite which was favorable for the adsorption process (Bertagnolli
et al., 2011). The results confirmed that there were no significant differences in the adsorbent’s
na
The SEM micrographs at various magnifications are illustrated in Fig. 3 (c) for BC and Fig. 3 (d) for
ur
TB. Both adsorbents exhibit individual particles having recognizable contours, fluff appearance, and
irregular platelets which formed large and thick agglomerates. No other significant difference in the
Jo
surface morphology of the materials could be observed. However, a small increase in the particle
roughness and little loss of foliated structure could be seen after the thermal activation as shown in
Fig. 3 (d). This morphological change after calcination can be attributed to the origin and composition
17
3.2. Tetracycline adsorption studies
The solution pH is one of the significant parameters that affect the TC adsorption and the degree of
ionization of functional groups on clay mineral surfaces. The removal efficiency of TB was higher
than that of BC as presented in Fig. 4 (a). The highest removal efficiency of 98.39 % was observed
at pH 3, whereas only 51.90% removal was achieved with BC for the given dosage and TC
of
concentration at the same pH. The removal efficiency further kept on decreasing less markedly with
the increase in solution pH for both adsorbents. On the other hand, the removal efficiency of TB
ro
decreased steadily with increasing pH throughout the studied pH range. However, the decrease was
-p
considerable at pH ≥ 7 for both adsorbents. The adsorption of TC onto clay minerals is pH-dependent
which is mainly because of the interaction between the clay surface and cationic species of TC
re
(TCH3+ ) (Browne et al., 1980). The concentration of cationic species of TC was found to decrease
with the increase in solution pH (Fig. 1 (b)). In literature, the affinity of different TC species has been
lP
reported as TCH3+ >TCH2± > TCH− (Parolo et al., 2008) which is in agreements with the pH-dependent
adsorption behavior observed in this study. Similar research findings have been reported by other
na
authors for TC adsorption on clay minerals (N. Liu et al., 2012; Parolo et al., 2008). Although, TC
adsorption onto the adsorbents was found better at acidic pH, further experiments were carried out at
ur
original solution pH (pH ~ 5), as it is highly unlikely to have highly acidic pH conditions in real
Jo
The influence of adsorbent dosage on TC removal was studied by utilizing various amounts of the
adsorbents (0.1- 1.0 g L-1). It could be noted that TB attained total removal (95.05 %) at a relatively
18
lower dose of 0.4 g L-1 whereas BC could only attain a removal efficiency of 50.29 % at the same
dosage (Fig. 4(b)). However, after increasing dose beyond 0.4 g L-1, the adsorption sites are higher
than the adsorbate molecules available in the solution for a given concentration resulting in total
removal and leaves the adsorbent unsaturated. Therefore, an adsorbent dosage of 0.4 g L-1 was chosen
as an optimum dose for further studies. Similar research results were also reported by other authors
of
3.2.3. Effect of ionic strength on TC removal
The effect of electrolyte (NaCl) concentration in the aqueous solution has a significant role on TC
ro
adsorption by BC and TB. Fig. 4 (c) shows that there was a decrease in the removal efficiency of BC
-p
and TB for TC with increasing electrolyte concentrations from 0 to 0.5 M. Moreover, the removal
efficiency of BC decreased from 50.29 % (0 M NaCl) to 15.04 % (0.5 M NaCl), and from 95.05 %
re
(0 M NaCl) to 57.46 % (0.5 M NaCl) for TB. The decrease in removal efficiency with increasing
ionic strength indicated that the Na+ competed with TC species for adsorption sites. The NaCl solution
lP
of concentration ≥ 0.01 M has plentiful of Na+ ions to compete with TC species (Parolo et al., 2008).
Therefore, a decrease in the adsorption efficiency with increasing electrolyte concentration would
na
make the bentonite surface less negatively charged, which would decrease the TC adsorption. Similar
Fig. 5 (a, b) shows the effect of contact time on the adsorption efficiency of TC onto BC and TB.
The kinetic studies were conducted for 5–300 min under optimized experimental conditions (pH~ 5,
dose: 0.4 g L-1, initial concentration: 30 mg L-1). The adsorption kinetic modeling of TC onto BC and
TB were assessed with AV, PFO, and PSO kinetic models (Fig. 5 (a, b)). The computed parameters
19
for fitting different models are presented in Table 3. Comparison of the fitted curves and computed
model parameters suggested that PSO describes the adsorption process superior to other models, with
a high coefficient of determination (R2 = 0.99 for both BC and TB) and relatively low RMSE for both
the adsorbents (Table 3). This fact confirms that TC adsorption onto BC and TB is rate-determined
of
The influence of initial TC concentration on the adsorption capacities of the clay adsorbents was
estimated by varying concentration of TC in the range 5 – 200 mg L-1. The non-linear form of
ro
Langmuir, Freundlich, Sips and Redlich-Peterson isotherm models were studied to understand the
-p
adsorption equilibrium. The corresponding model parameters along with the R2, RMSE, and SSE
from the fitting of different isotherm models are listed in Table 4. The maximum monolayer
re
adsorption capacity obtained by Langmuir isotherm was 156.7 mg g-1 (BC) and 388.1 mg g-1 (TB).
The absence of saturation plateau in Fig. 5 (c,d) suggests that the conditions are not fulfilled for
lP
verifying the suitability of this model (Daneshvar et al., 2012). The high R2 (0.989 and 0.997 for BC
and TB, respectively) and n (2.079 and 1.944 for BC and TB, respectively) values of Freundlich
na
isotherm suggest the appropriateness of using Freundlich model to describe the experimental data.
The computed value of adsorption intensity n being within the range of 0 – 10 indicates the
ur
favorability of TC adsorption onto both the adsorbents (Santhana Krishna Kumar and Jiang, 2015).
Jo
Furthermore, the value of 1/n expresses the heterogeneity of the adsorption sites and affinity between
the adsorbate and adsorbent. The estimate of g obtained from Redlich-Peterson model lies within the
range of 0 – 1 indicating favorable adsorption (Nigri et al., 2017). The value of g is also useful in
determining the mechanism of adsorption. If g =1, the mechanism of adsorption is said to be physio-
sorption, while g = 0, indicates linear isotherm. The heterogeneity of adsorption is greater when the
value of g lies closer to [1-(1/n)] (Mane and Babu, 2011). The value of g in Table 4 is nearly the
20
same as [1-(1/n)], thus the adsorption of TC onto the clay adsorbents can be described as
heterogeneous in nature.
Reusability of adsorbent materials is critical for the practical feasibility of adsorbents in water
process for five cycles with a known volume of a 0.1 M NaOH solution. As shown in Fig. 4 (d), BC
of
and TB exhibited good reusability with remarkable regeneration behavior. The sorption percent of
TC onto BC was decreased significantly from 50% to 30% after five cycles. This could be attributed
ro
to the loss of some active adsorption sites during solvent washing. However, the removal efficiency
-p
of TC onto TB was slightly decreased from 95% to 83% after five consecutive adsorption-desorption
cycles. These results suggest that TB can be successfully regenerated by NaOH treatment. The low-
re
cost regeneration of TB definitely would facilitate practical applications in water treatment.
lP
The effect of various adsorbent loadings on the breakthrough curve was studied using fixed-bed
ur
column. Based on the batch results, the adsorbate solution (pH ~ 5) with initial TC concentration of
Jo
10 mg L-1 at 1.5 mL min-1 flow rate was passed through the column system with 50 and 25 mg
adsorbent loadings. The steepness of the breakthrough curve was found to be a function of adsorbent
loadings in the column system (Fig. 6 (a)). The breakthrough time was also found to decrease from
51 to 17 min with a decrease in the adsorbent loading from 50 mg to 25 mg. The column parameters
are presented in Table 5. The results exhibit that an increased adsorbent loading leads to an increase
21
in the removal efficiency (RE%) for TC from 29.44 % (25 mg) to 35.89 % (50 mg). Similarly, the TC
uptake in the column system (qbed) was found to be 48.42 and 71.28 mg g-1, when the adsorbent
loadings were 25 and 50 mg and the corresponding total TC adsorbed were 8.06 and 8.16 mg,
respectively. The increase in total TC adsorbed with the increase of adsorbent loadings in the fixed
bed column could be attributed to the increased surface area of the adsorbent and increased number
of
3.3.2. Effect of flow rate
Fig. 6 (b) shows the effect of different flow rates on the breakthrough curve. The fixed-bed column
ro
adsorption experiments were carried out at two different flow rates (1.5 and 3 mL min-1). During the
-p
experiment, the influent concentrations, solution pH and adsorbent loading were maintained as 10 mg
L-1, 5.0 and 50 mg, respectively. Breakthrough time, exhaustion time and the adsorption efficiency
re
exhibited higher values at lower flow rate (1.5 mL min-1) (Table 5). This behavior could be explained
by the fact that at higher flow rates, the adsorbate molecule does not get enough residence time (to be
lP
in contact) with the adsorbent. As a consequence, the columns operated at high flow rate show
decreased removal efficiency. On the other hand, the low flow rate provides more time for the
na
adsorbate molecule to interact with the adsorbent resulting in higher removal efficiency. Our
interpretations are in agreement with those of previous studies (Chen et al., 2012). Table 5 represents
ur
the experimental parameters calculated using the column data. It was evident that an increase in flow
Jo
rate from 1.5 to 3 mL min-1 led to a decreased removal (RE%) from 35.89 to 34.49 %, and
simultaneously decreased TC uptake in the column system (qbed) from 71.28 to 52.87 mg g-1,
respectively.
22
The column experiments were further conducted with TB to investigate the effect of initial TC
concentrations (5 and 10 mg L-1) on the breakthrough time. For this investigation, other parameters
such as flow rate (1.5 mL min-1), solution pH (pH~5) and adsorbent loading (50 mg) were kept
constant. Breakthrough time reduced with high TC concentration and the exhaustion time was
reached rapidly (Fig. 6 (c)). This behavior could be due to the quick saturation of available active
surface sites with an increase in influent TC concentration. Table 5 also reveals that increasing the
influent concentration from 5 to 10 mg L-1 led to an increase in the equilibrium TC uptake (qbed) from
45.64 to 71.28 mg g-1, and an increase in the total TC adsorbed in the column system (from 2.28 to
of
3.56 mg), respectively. However, it also shows that the removal efficiency decreased from 46.93 to
ro
3.4. Adsorption mechanism
-p
re
The mechanism of TC sorption onto TB was investigated by comparing the TB characterization
results before and after adsorption. Fig. 7 (a) illustrates the XRD patterns of TB before and after TC
lP
adsorption. The most significant differences appeared in the montmorillonite peaks at 2θ = 5.5° and
20° after adsorption. The first strong peak (001) at 2θ = 5.5° was expanded and shifted. While the
na
montmorillonite peak at 2θ = 20° showed a decrease in the intensity. It was probable that TC bonded
strongly with the hydroxyl groups. These results were similar to previous reports of TC adsorption
ur
on montmorillonite (Kulshrestha et al., 2004) and soil clay (Pils et al., 2007). Meanwhile, the original
Jo
thickness of each 2:1 layer in montmorillonite was about 0.96 nm (Zhao et al., 2012) and the size of
the TC molecule (length 1.29 nm, thickness 0.75 nm and height 0.62 nm) (Gambinossi et al., 2004).
The d-spacing expansion after adsorption suggested that TC intercalated as a monolayer between two
of 2:1 montmorillonite layer together with a monolayer of water molecules (Fig. 7 (e)).
23
Fig. 7 (b) shows the FTIR spectra of TB before and after the adsorption. The FTIR bands of the
backbone structure of TB before and after TC adsorption showed no apparent changes. This behavior
indicates that the adsorption of TC did not affect the clay structure. Originally, clay was free from
vibration bands between 3000 and 1300 cm-1, excluding the water bending vibration at around 1630
cm-1, thereby providing an appropriate window for the observation of TC adsorption bands (Chang
et al., 2009). Moreover, the band at 1635 cm-1 is principally due to the direct water coordination to
the exchangeable cations of the clay (Theng, 1974). Consequently, disappearing or shifting of this
band after interaction with tetracycline was a good indication of the replacement of this interlayer
of
cation by TC. For TC, the most characteristic peaks for the potential interacting functional groups
exist in the region between 1700 – 1200 cm-1. It is clear from the spectra of adsorbent (before and
ro
after TC adsorption) (Fig. 7 (b)) that the band at 1635 cm-1 disappeared for TB and appeared as three
-p
new bands at 1626, 1500, 1461 cm-1 after the adsorption of TC. This can be attributed to the
interaction of clay with the amide and amino carbonyl groups in the TC structure. Similar behavior
re
for TC has been reported by other authors (Kulshrestha et al., 2004). The shift in the band at 3619
cm-1 to lower frequency can be attributed to the interaction between TC and TB.
lP
Fig. 7 (c,d) shows the SEM micrographs obtained for TB before and after TC adsorption,
na
respectively. Before adsorption, the porous and irregular structure was observed for TB. However,
after TC adsorption, the roughness in the adsorbent surface was lost. This could be attributed to the
ur
Overall, TC adsorption by TB can be via (i) ion exchange by the intercalation of TC into the interlayer
Jo
space (Fig. 7 (e①)) of TB (which was confirmed by the interlayer expanding after adsorption based
on the XRD analysis (Fig. 7 (a))), (ii) electrostatic interaction between cationic species of TC and
negative surface binding sites (Fig. 7 (e②)) (which was confirmed by the pHzpc measurements and
TC speciation (Fig. 1 (b) and Fig. 2 (c))) and (iii) surface complexation on the TB edge sites (Fig. 7
24
3.5. Comparison with other adsorbents
Table 6 shows a comparison between the obtained maximum adsorption capacities (qmax) for the
adsorption of tetracycline by BC and TB with other adsorbents from the literature. The presented data
in Table 6 reflect the very high adsorption capacities of TB adsorbent compared to other adsorbents.
This can be attributed mainly to the presence of different types of functional groups on TB and its
high surface area and unique porous properties. Therefore, thermal activation of bentonite is one of
of
the best clay modification techniques which can enhance its adsorption capacities for tetracycline.
Some of the adsorbent materials shown in Table 6 display a higher uptake of tetracycline as compared
ro
to the results obtained in this work. This could be due to the difference in structure and specific surface
area of adsorbents and the effective interactions between tetracycline and the adsorbent.
-p
re
4. Conclusions
lP
The natural bentonite clay collected from Al-Hammam area, Egypt was studied in terms of its
adsorption capability to remove tetracycline antibiotics from aqueous solution. Simple pre-treatment
na
and thermal activation technique were successfully applied to modify bentonite to achieve higher
adsorption efficiency for TC. The characterization of BC and TB showed that the structure of the raw
ur
clay was successfully modified without destructing the structural integrity of bentonite adsorbent.
The BET analysis revealed that the specific surface area and pore volume increased more than 2 times
Jo
after thermal activation. The elemental composition confirmed the polycationic nature of bentonite
with the dominance of SiO2 and Al2O3. The sorption ability of bentonite toward TC was significantly
elevated after activation. This sorption improvement is mostly due to the loss of water and hydroxyl
in bentonite structural, which could uncover the binding sites with negative potential and enhance its
activity. The equilibrium data fitted better with the PSO model which governs the rate kinetics of the
25
adsorption process. The Freundlich isotherm model was found to be more appropriate for describing
the adsorption behavior amongst other studied models. Increasing the ionic strength had a detrimental
effect on the adsorption efficiency of the adsorbent. Regeneration studies showed that TB can be
reused for 5 cycles using 0.1 M NaOH as eluent. Analysis of the column data revealed that increased
influent flow rate and influent concentration inversely affected the column efficiency, whereas the
increased amount of adsorbent loading reciprocated the removal efficiency of the column. Overall,
the results of this study demonstrate that TB is a promising adsorbent for TC removal. This also
suggests that the modified bentonite can be assessed as an adsorbent for other emergent pollutants in
of
large scale, which can lead to future environmental benefits.
ro
Declarations of interest -p
None
re
lP
Acknowledgements
First author (A.M.) is grateful to the cultural affairs and missions sector, the Ministry of Higher
na
Education, Egypt, for the financial support toward this study. Also, the first author is grateful to Gijs
Peters, Hogeschool van Arnhem en Nijmegen, Netherlands for his assistance in some lab
ur
experiments.
Jo
References
Abu-Zreig, M.M., Al-Akhras, N.M., Attom, M.F., 2001. Influence of heat treatment on the behavior
26
Ahmed, M.B., Zhou, J.L., Ngo, H.H., Guo, W., 2015. Adsorptive removal of antibiotics from water
and wastewater: Progress and challenges. Sci. Total Environ. 532, 112–126.
https://doi.org/10.1016/j.scitotenv.2015.05.130
Alqadami, A.A., Naushad, M., Abdalla, M.A., Ahamad, T., Abdullah ALOthman, Z., Alshehri,
S.M., Ghfar, A.A., 2017. Efficient removal of toxic metal ions from wastewater using a
Alatalo, S.-M., Daneshvar, E., Kinnunen, N., Meščeriakovas, A., Thangaraj, S.K., Jänis, J., Tsang,
of
D.C.W., Bhatnagar, A., Lähde, A., 2019. Mechanistic insight into efficient removal of
ro
tetracycline from water by Fe/graphene. Chem. Eng. J. 373, 821–830.
https://doi.org/10.1016/J.CEJ.2019.05.118 -p
Araujo, A.L.P. de, Bertagnolli, C., Silva, M.G.C. da, Gimenes, M.L., Barros, M.A.S.D. de, 2013.
re
Zinc adsorption in bentonite clay: influence of pH and initial concentration. Acta Sci. Technol.
35. https://doi.org/10.4025/actascitechnol.v35i2.13364
lP
Antón-Herrero, R., García-Delgado, C., Alonso-Izquierdo, M., García-Rodríguez, G., Cuevas, J.,
na
Eymar, E., 2018. Comparative adsorption of tetracyclines on biochars and stevensite: Looking
for the most effective adsorbent. Appl. Clay Sci. 160, 162–172.
ur
https://doi.org/10.1016/J.CLAY.2017.12.023
Avrami, M., 1939. Kinetics of Phase Change. I General Theory. J. Chem. Phys. 7, 1103–1112.
Jo
https://doi.org/10.1063/1.1750380
Aytas, S., Yurtlu, M., Donat, R., 2009. Adsorption characteristic of U(VI) ion onto thermally
https://doi.org/10.1016/J.JHAZMAT.2009.07.049
27
Bertagnolli, C., Kleinübing, S.J., da Silva, M.G.C., 2011. Preparation and characterization of a
Brazilian bentonite clay for removal of copper in porous beds. Appl. Clay Sci. 53, 73–79.
https://doi.org/10.1016/J.CLAY.2011.05.002
Bragg, W.H., Bragg, W.L., 1913. The Reflection of X-rays by Crystals. Proc. R. Soc. A Math.
Browne, J.E., Feldkamp, J.R., White, J.L., Hem, S.L., 1980. Characterization and adsorptive
of
https://doi.org/10.1002/JPS.2600690719
Budsaereechai, S., Kamwialisak, K., Ngernyen, Y., 2012. Adsorption of lead, cadmium and copper
ro
on natural and acid activated bentonite clay, KKU Res. J.
-p
Calábria, J.A.A., Do Amaral, D.N., Cláudia, A., Ladeira, Q., Cota, S.D.S., Silva, T.S.S., 2013.
HYPERALKALINE FLUID.
Cantuaria, M.L., de Almeida Neto, A.F., Nascimento, E.S., Vieira, M.G.A., 2016. Adsorption of
na
silver from aqueous solution onto pre-treated bentonite clay: complete batch system
Cetecioglu, Z., Ince, B., Gros, M., Rodriguez-Mozaz, S., Barceló, D., Orhon, D., Ince, O., 2013.
https://doi.org/10.1016/J.WATRES.2013.02.053
Chang, P.-H., Li, Z., Jean, J.-S., Jiang, W.-T., Wang, C.-J., Lin, K.-H., 2012. Adsorption of
tetracycline on 2:1 layered non-swelling clay mineral illite. Appl. Clay Sci. 67–68, 158–163.
28
https://doi.org/10.1016/J.CLAY.2011.11.004
Chang, B.-V., Ren, Y.-L., 2015. Biodegradation of three tetracyclines in river sediment. Ecol. Eng.
Chang, P.-H., Li, Z., Yu, T.-L., Munkhbayer, S., Kuo, T.-H., Hung, Y.-C., Jean, J.-S., Lin, K.-H.,
2009. Sorptive removal of tetracycline from water by palygorskite. J. Hazard. Mater. 165,
148–155. https://doi.org/10.1016/J.JHAZMAT.2008.09.113
Chen, S., Yue, Q., Gao, B., Li, Q., Xu, X., Fu, K., 2012. Adsorption of hexavalent chromium from
of
aqueous solution by modified corn stalk: A fixed-bed column study. Bioresour. Technol. 113,
114–120. https://doi.org/10.1016/j.biortech.2011.11.110
ro
Chen, X., Wu, L., Liu, F., Luo, P., Zhuang, X., Wu, J., Zhu, Z., Xu, S., Xie, G., 2018. Performance
-p
and mechanisms of thermally treated bentonite for enhanced phosphate removal from
Chinoune, K., Bentaleb, K., Bouberka, Z., Nadim, A., Maschke, U., 2016. Adsorption of reactive
dyes from aqueous solution by dirty bentonite. Appl. Clay Sci. 123, 64–75.
na
https://doi.org/10.1016/J.CLAY.2016.01.006
Choi, K.-J., Kim, S.-G., Kim, S.-H., 2008. Removal of antibiotics by coagulation and granular
ur
https://doi.org/10.1016/J.JHAZMAT.2007.05.059
Jo
Choo, K.Y., Bai, K., 2016. The effect of the mineralogical composition of various bentonites on
CEC values determined by three different analytical methods. Appl. Clay Sci. 126, 153–159.
https://doi.org/10.1016/J.CLAY.2016.03.010
Daneshvar, E., Kousha, M., Jokar, M., Koutahzadeh, N., Guibal, E., 2012. Acidic dye biosorption
29
onto marine brown macroalgae: Isotherms, kinetic and thermodynamic studies. Chem. Eng. J.
Daneshvar, E., Zarrinmehr, M.J., Hashtjin, A.M., Farhadian, O., Bhatnagar, A., 2018. Versatile
applications of freshwater and marine water microalgae in dairy wastewater treatment, lipid
https://doi.org/10.1016/J.BIORTECH.2018.08.032
Danková, Z., Dolinská, S., 2012. EFFECT OF THERMAL TREATMENT ON THE BENTONITE
of
PROPERTIES. Arh. za Teh. Nauk. https://doi.org/10.5825/afts.2012.0407.049O
Divband Hafshejani, L., Hooshmand, A., Naseri, A.A., Mohammadi, A.S., Abbasi, F., Bhatnagar,
ro
A., 2016. Removal of nitrate from aqueous solution by modified sugarcane bagasse biochar.
-p
Ecol. Eng. 95, 101–111. https://doi.org/10.1016/J.ECOLENG.2016.06.035
Gambinossi, F., Mecheri, B., Nocentini, M., Puggelli, M., Caminati, G., 2004. Effect of the
https://doi.org/10.1016/j.bpc.2004.01.008
Gatsios, E., Hahladakis, J.N., Gidarakos, E., 2015. Optimization of electrocoagulation (EC) process
ur
for the purification of a real industrial wastewater from toxic metals. J. Environ. Manage. 154,
117–127. https://doi.org/10.1016/J.JENVMAN.2015.02.018
Jo
Giacobbo, A., Meneguzzi, A., Bernardes, A.M., de Pinho, M.N., 2017. Pressure-driven membrane
processes for the recovery of antioxidant compounds from winery effluents. J. Clean. Prod.
Giammarco, J., Mochalin, V.N., Haeckel, J., Gogotsi, Y., 2016. The adsorption of tetracycline and
30
vancomycin onto nanodiamond with controlled release. J. Colloid Interface Sci. 468, 253–261.
https://doi.org/10.1016/J.JCIS.2016.01.062
Guo, F., Shi, W., Wang, H., Han, M., Guan, W., Huang, H., Liu, Y., Kang, Z., 2018. Study on
https://doi.org/10.1016/J.JHAZMAT.2018.01.042
Hafshejani, L.D., Tangsir, S., Daneshvar, E., Maljanen, M., Lähde, A., Jokiniemi, J., Naushad, M.,
of
Bhatnagar, A., 2017. Optimization of fluoride removal from aqueous solution by Al2O3
ro
Ho, Y.., McKay, G., 1999. Pseudo-second order model for sorption processes. Process Biochem.
-p
34, 451–465. https://doi.org/10.1016/S0032-9592(98)00112-5
Hokkanen, S., Bhatnagar, A., Srivastava, V., Suorsa, V., Sillanpää, M., 2018. Removal of Cd2+,
re
Ni2+ and PO43− from aqueous solution by hydroxyapatite-bentonite clay-nanocellulose
lP
https://doi.org/10.1016/J.IJBIOMAC.2018.06.095
na
Kulshrestha, P., Giese, R.F., Aga, D.S., 2004. Investigating the Molecular Interactions of
Oxytetracycline in Clay and Organic Matter: Insights on Factors Affecting Its Mobility in
ur
Kundu, S., Kavalakatt, S.S., Pal, A., Ghosh, S.K., Mandal, M., Pal, T., 2004. Removal of arsenic
Jo
using hardened paste of Portland cement: batch adsorption and column study. Water Res. 38,
3780–3790. https://doi.org/10.1016/J.WATRES.2004.06.018
Lagergren, 1898. About the Theory of So-called Adsorption of Soluble Substances. Sven.
31
Langmuir, I., 1918. THE ADSORPTION OF GASES ON PLANE SURFACES OF GLASS, MICA
Li, N., Zhou, L., Jin, X., Owens, G., Chen, Z., 2019. Simultaneous removal of tetracycline and
Liao, P., Zhan, Z., Dai, J., Wu, X., Zhang, W., Wang, K., Yuan, S., 2013. Adsorption of tetracycline
and chloramphenicol in aqueous solutions by bamboo charcoal: A batch and fixed-bed column
of
study. Chem. Eng. J. 228, 496–505. https://doi.org/10.1016/J.CEJ.2013.04.118
Liu, N., Wang, M., Liu, M., Liu, F., Weng, L., Koopal, L.K., Tan, W., 2012. Sorption of
ro
tetracycline on organo-montmorillonites. J. Hazard. Mater. 225–226, 28–35.
https://doi.org/10.1016/J.JHAZMAT.2012.04.060 -p
Liu, P., Liu, W.-J., Jiang, H., Chen, J.-J., Li, W.-W., Yu, H.-Q., 2012. Modification of bio-char
re
derived from fast pyrolysis of biomass and its application in removal of tetracycline from
lP
https://doi.org/10.1016/J.BIORTECH.2012.06.085
na
López-Montilla, J.C., Pandey, S., Shah, D.O., Crisalle, O.D., 2005. Removal of non-ionic organic
pollutants from water via liquid–liquid extraction. Water Res. 39, 1907–1913.
ur
https://doi.org/10.1016/j.watres.2005.02.018
Mane, V.S., Babu, P.V.V., 2011. Studies on the adsorption of Brilliant Green dye from aqueous
Jo
solution onto low-cost NaOH treated saw dust. Desalination 273, 321–329.
https://doi.org/10.1016/J.DESAL.2011.01.049
Nasseri, S., Mahvi, A.H., Seyedsalehi, M., Yaghmaeian, K., Nabizadeh, R., Alimohammadi, M.,
32
peroxydisulfate activated by ultrasound irradiation: Effect of radical scavenger and water
Ncibi, M.C., Sillanpää, M., 2015. Optimized removal of antibiotic drugs from aqueous solutions
using single, double and multi-walled carbon nanotubes. J. Hazard. Mater. 298, 102–110.
https://doi.org/10.1016/J.JHAZMAT.2015.05.025
Nigri, E.M., Bhatnagar, A., Rocha, S.D.F., 2017. Thermal regeneration process of bone char used in
the fluoride removal from aqueous solution. J. Clean. Prod. 142, 3558–3570.
of
https://doi.org/10.1016/J.JCLEPRO.2016.10.112
Odom, I.E., 1984. Smectite clay Minerals: Properties and Uses. Philos. Trans. R. Soc. A Math.
ro
Phys. Eng. Sci. 311, 391–409. https://doi.org/10.1098/rsta.1984.0036
-p
Oliveira, M.F., da Silva, M.G.C., Vieira, M.G.A., 2019. Equilibrium and kinetic studies of caffeine
adsorption from aqueous solutions on thermally modified Verde-lodo bentonite. Appl. Clay
re
Sci. 168, 366–373. https://doi.org/10.1016/J.CLAY.2018.12.011
lP
Parolo, M.E., Savini, M.C., Vallés, J.M., Baschini, M.T., Avena, M.J., 2008. Tetracycline
adsorption on montmorillonite: pH and ionic strength effects. Appl. Clay Sci. 40, 179–186.
na
https://doi.org/10.1016/J.CLAY.2007.08.003
Pils, J.R. V., Laird, D.A., Jutta R. V. Pils, A., Laird, A., D., 2007. Sorption of Tetracycline and
ur
Redlich, O., Peterson, D.L., 1959. A Useful Adsorption Isotherm. J. Phys. Chem. 63, 1024–1024.
https://doi.org/10.1021/j150576a611
Saikia, B.J., Parthasarathy, G., 2010. Fourier Transform Infrared Spectroscopic Characterization of
Kaolinite from Assam and Meghalaya, Northeastern India. J. Mod. Phys. 01, 206–210.
33
https://doi.org/10.4236/jmp.2010.14031
Santhana Krishna Kumar, A., Jiang, S.-J., 2015. Preparation and characterization of exfoliated
6304. https://doi.org/10.1039/C4RA12564A
Sarmah, A.K., Meyer, M.T., Boxall, A.B.A., 2006. A global perspective on the use, sales, exposure
pathways, occurrence, fate and effects of veterinary antibiotics (VAs) in the environment.
of
Sen Gupta, S., Bhattacharyya, K.G., 2014. Adsorption of metal ions by clays and inorganic solids.
ro
Singh, R., Singh, A.P., Kumar, S., Giri, B.S., Kim, K.-H., 2019. Antibiotic Resistance in Major
-p
Rivers in the World: A Systematic Review on Occurrence, Emergence, and Management
https://doi.org/10.1063/1.1746922
Terzić, A., Pezo, L., Andrić, L., Pavlović, V.B., Mitić, V. V, 2016. Optimization of bentonite clay
na
https://doi.org/10.1016/j.ceramint.2016.11.058
ur
Theng, B.K.G., 1974. The Chemistry of Clay-Organic Reactions. Chem. Clay-Organic React.
Jo
Thommes, M., Kaneko, K., Neimark, A. V., Olivier, J.P., Rodriguez-Reinoso, F., Rouquerol, J.,
Sing, K.S.W., 2015. Physisorption of gases, with special reference to the evaluation of surface
area and pore size distribution (IUPAC Technical Report). Pure Appl. Chem. 87, 1051–1069.
https://doi.org/10.1515/pac-2014-1117
Toor, M., Jin, B., Dai, S., Vimonses, V., 2015. Activating natural bentonite as a cost-effective
34
adsorbent for removal of Congo-red in wastewater. J. Ind. Eng. Chem. 21, 653–661.
https://doi.org/10.1016/J.JIEC.2014.03.033
Tyagi, B., Chudasama, C.D., Jasra, R. V., 2006. Determination of structural modification in acid
Wang, J., Hu, J., Zhang, S., 2010. Studies on the sorption of tetracycline onto clays and marine
of
https://doi.org/10.1016/J.JCIS.2010.04.081
Wang, T., Pan, X., Ben, W., Wang, J., Hou, P., Qiang, Z., 2017. Adsorptive removal of antibiotics
ro
from water using magnetic ion exchange resin. J. Environ. Sci. 52, 111–117.
https://doi.org/10.1016/J.JES.2016.03.017 -p
Wang, Y.-J., Jia, D.-A., Sun, R.-J., Zhu, H.-W., Zhou, D.-M., 2008. Adsorption and Cosorption of
re
Tetracycline and Copper(II) on Montmorillonite as Affected by Solution pH. Environ. Sci.
lP
WHO, 1997. The Medical Impact of Antimicrobial Use in Food Animals. Report of a WHO
na
Yaming, L., Mingliang, B., Zhipeng, W., Run, L., Keliang, S., Wangsuo, W., 2016. Organic
ur
modification of bentonite and its application for perrhenate (an analogue of pertechnetate)
removal from aqueous solution. J. Taiwan Inst. Chem. Eng. 62, 104–111.
Jo
https://doi.org/10.1016/J.JTICE.2016.01.018
Zazoua, A., Kazane, I., Khedimallah, N., Dernane, C., Errachid, A., Jaffrezic-Renault, N., 2013.
35
https://doi.org/10.1016/J.MSEC.2013.09.005
Zhao, Y., Gu, X., Gao, S., Geng, J., Wang, X., 2012. Adsorption of tetracycline (TC) onto
https://doi.org/10.1016/J.GEODERMA.2012.03.004
Zuo, Q., Gao, X., Yang, J., Zhang, P., Chen, G., Li, Y., Shi, K., Wu, W., 2017. Investigation on the
thermal activation of montmorillonite and its application for the removal of U(VI) in aqueous
of
https://doi.org/10.1016/J.JTICE.2017.09.016
ro
-p
re
lP
na
ur
Jo
36
f
oo
(a) pKa2=7.7 pKa1=3.3 (c)
pr
e-
Pr
(b)
pKa3=9.7 (d)
l
na
pKa1=3.3 pKa3=9.7 Chemical Tetracycline (TC)
+ ± − − 444.44 g mol-1
ur
Molecular weight
TC TC TC T
λ max 357
Jo
37
Fig. 1. (a) Chemical structure of TC depicting ionizable groups and their dissociation constant (pKa) values, (b) pH dependent TC speciation, (c)
3D structure of TC and (d) Chemical formula and properties of TC.
f
oo
pr
e-
l Pr
na
ur
Jo
38
Jo
ur
na
l Pr
39
e-
pr
oo
f
70
Volume Adsorbed (cm /g STP)
60
(c)
(cm3/g STP)
50
3
40
30 TB
N2 adsorption
20 N2 desorption
10
Volume Adsorbed
of
0
500.0 0.2 0.4 0.6 0.8 1.0
Volume Adsorbed (cm /g STP)
ro
40
3
30
20
BC
-p
N2 adsorption
re
N2 desorption
10
lP
0
0.0 0.2 0.4 0.6 0.8 1.0
Relative pressure (P/P0)
Relative pressure (P/P 0)
na
ur
Jo
40
4 (d)
2
Δ pH
of
0
2 4 6 8 10
ro
-1
pH BC
-p TB
-2
re
Fig. 2. Characterization results of the natural bentonite before and after treatment (a) XRD Pattern,
lP
41
f
oo
pr
e-
l Pr
na
ur
Jo
Fig. 3. EDX spectra results of (a) BC and (b) TB, the micrographs of the samples with different magnifications of (c) BC and (d) TB.
42
(a) BC
100
TB
80
TC sorbed (%)
60
40
of
20
ro
0
3 4 5 6 7 8 9
pH -p
re
100
(b) 120
80 100
TC sorbed (%)
Removal efficiency
na
BC 80
60 TB
Adsorption capacity
BC 60
ur
40 TB
40
Jo
20
20
0 0
0.0 0.2 0.4 0.6 0.8 1.0
Dosage (g L-1)
43
100
(c) BC
TB
80
TC sorbed (%)
60
40
20
of
ro
0
Control 0.1 M 0.2 M 0.3 M 0.4 M 0.5 M
NaCl Concentration (M)
100
(d)
-p BC
TB
re
75
lP
TC sorbed (%)
na
50
ur
25
Jo
0
1 2 3 4 5
Cycle number
Fig. 4. Effect of (a) pH, (b) adsorbent dosage, (c) ionic strength on TC adsorption onto BC and TB,
(d) regeneration studies of TC adsorption onto BC and TB using 0.1 M NaOH as eluent.
44
(a) (b)
f
oo
qt (mg g-1)
pr qt (mg g-1)
e-
Pr
Time (min) Time (min)
(c) (d)
l
na
qe (mg g-1)
qe (mg g-1)
ur
Jo
Fig. 5. (a and b) adsorption kinetic modelling of TC (30 mg L−1) by BC and TB, respectively (0.4 g L−1), (c and d) adsorption
isotherms modelling of TC (concentration: 5–200 mg L−1) by BC and TB, respectively (0.4 g L−1).
45
1.0 (a)
0.8
0.6
Ct/Co
0.4
of
0.2
50 mg
ro
25 mg
0.0
0 100 200 300 400 500
Time (min) -p
(b)
re
1.0
lP
0.8
na
0.6
Ct/Co
ur
0.4
Jo
0.2
1.5 mL min-1
3 mL min-1
0.0
0 100 200 300 400 500
Time (min)
46
1.0 (c)
0.8
0.6
Ct/Co
0.4
of
ro
0.2
5 mg L-1
-p 10 mg L-1
0.0
0 100 200 300 400 500 600 700
re
Time (min)
lP
47
Jo
ur
na
l Pr
48
e-
pr
oo
f
Fig. 7. (a) XRD pattern of TB before and after TC adsorption, (b) FTIR spectra of TB before and after TC adsorption, SEM micrograph of TB (c)
before and (d) after TC adsorption, (e) possible adsorption mechanism of TC adsorption onto TB.
f
oo
pr
e-
l Pr
na
ur
Jo
49
Table 1. The chemical composition of natural (BC) and thermal activated bentonite (TB) sample
(% by weight).
of
Al2O3 19.83 20.22
ro
Fe2O3 8.99 12.01
MgO
CaO
2.46
0.66
-p 6.35
0.61
re
K2O 1.50 1.44
lP
50
Table 2. The textural properties of bentonite samples based on N2 adsorption-desorption isotherms.
Samples
Properties
BC TB
of
Total pore volume (cm3 g-1) 74.27 *10-3 95.76 *10-3
ro
Mean pore diameter (nm) 22.53 14.22
51
Table 3. Pseudo-first-order, pseudo-second-order and Avrami model parameters of TC adsorption
onto BC and TB.
Pseudo-first-order model
Adsorbent
k1 (min−1) qe(cal) (mg g−1) qe(exp) (mg g−1) RMSE R2
Pseudo-second-order model
of
BC 2 *10-5 59.76 45.22 0.60 0.999
ro
Avrami model
52
Table 4. Langmuir, Freundlich, Sips and Redlich–Peterson isotherm model parameters for TC
adsorption onto BC and TB.
of
BC 10.76 2.079 0.989 4.49 161.4
ro
Sips isotherm model
KR (L g−1) aR (mg-1) R2
lP
g RMSE SSE
53
Table 5. The effect of flow rate, adsorbent loading and initial TC concentration on the total adsorbed TC (qtotal), equilibrium uptake (qeq), total
removal efficiency of the column (RE %) and total unadsorbed TC concentration at equilibrium (Ceq).
f
oo
Parameters
Co Q M Veff ttotal qtotal qbed mtotal RE Ceq
pr
Experiments (mg L-1) (mL min-1) (mg) (mL) (min) (mg) (mg g-1) (mg) (%) (mg L-1)
e-
Column 2 10 3 50 660 260 2.64 52.87 7.66 34.49 7.61
Pr
Column 4 5 1.5 50 930 620 2.28 45.64 4.86 46.93 2.78
l
na
ur
Jo
54
Table 6. Comparison of maximum adsorption capacities of various adsorbents reported in the
literature for TC adsorption.
Scenedesmus quadricauda
of
295.3 (Daneshvar et al., 2018)
microalgae
Fe/graphene 442.0 (Alatalo et al., 2019)
ro
Bamboo charcoal 22.7 (Liao et al., 2013)
55