SlideShare une entreprise Scribd logo
1  sur  13
Télécharger pour lire hors ligne
Journal of Contaminant Hydrology 247 (2022) 103977
Available online 14 February 2022
0169-7722/© 2022 Published by Elsevier B.V.
New insights into the selective adsorption mechanism of cationic and
anionic dyes using MIL-101(Fe) metal-organic framework: Modeling and
interpretation of physicochemical parameters
Mohamed Shakly a
, Laila Saad b
, Moaaz K. Seliem c,*
, Adrián Bonilla-Petriciolet d
,
Nabila Shehata a
a
Environmental Science and Industrial Development Department, Faculty of Postgraduate Studies for Advanced Science (PSAS), Beni-Suef University, P.O. Box 62511,
Beni-Suef, Egypt
b
Renewable Energy Science and Engineering Department, Faculty of Postgraduate Studies for Advanced Science (PSAS), Beni-Suef University, P.O. Box 62511, Beni-
Suef, Egypt
c
Faculty of Earth Science, Beni-Suef University, 62511, Egypt
d
Instituto Tecnológico de Aguascalientes, Aguascalientes 20256, Mexico
A R T I C L E I N F O
Keywords:
MIL-101(Fe)
Methylene blue
Methyl orange
Adsorption mechanism
Advanced modeling
A B S T R A C T
In the current study, iron-based metal organic framework (MOF) MIL-101(Fe) was successfully prepared via a
facile solvothermal method. The as–synthesized MIL-101(Fe) was characterized by XRD, FE-SEM, FTIR, TGA and
zeta potential techniques, and then employed as an adsorbent for methyl orange (MO) and methylene blue (MB)
dyes. The adsorbed quantities of MO (1067 to 831 mg/g) were higher than those of MB (402 to 353 mg/g)
indicating the high selectivity of MIL-101(Fe) towards the anionic dye at all temperatures (20–60 ◦
C). Adsorption
processes of MO and MB followed the pseudo-second order kinetics and the Langmuir equilibrium model. The
interaction mechanism at a molecular level was analyzed and deeply interpreted via the advanced multilayer
adsorption model. Steric parameters indicated that MO molecular aggregation (n) was 0.95–1.33 thus signifying
the presence of multi–docking and multi–interactions mechanisms. The aggregated number of MB was superior
to unity (i.e., n = 1.17–1.78) suggesting a vertical adsorption position and a multi-interactions mechanism at all
operating temperatures. The density of MIL-101(Fe) active sites (DM = 77.33–52.38 mg/g for MB and
149.91–107.07 for MO) and the total adsorbed dye layers (Nt = 3.12–2.49 for MB and 5.36–3.67 for MO) resulted
in improving the adsorption capacities of MO dye. The adsorption energies ranged from 8.89 to 33.73 kJ/mol
and they displayed that MO and MB uptake processes were exothermic controlled by physical interactions at all
temperatures. Regeneration results indicated that this adsorbent can be reutilized without a significant loss in its
removal efficiency after five adsorption-desorption cycles. Overall, the adsorption capacity, chemical stability,
and regeneration performance of MIL-101(Fe) support its application as a very promising adsorbent for the
removal of organic hazardous pollutants from water.
1. Introduction
Pollution of water bodies due to industrial toxic-containing effluents
is considered a worldwide problem (Homaeigohar, 2020). Commonly,
significant quantities of dyed wastewater are generated from various
industries including textile, leather, paper, printing, and dyestuff sec­
tors. According to previous studies, methylene blue (MB) and methyl
orange (MO) dyes (the studied organic pollutants in this paper) are
recognized as chemical compounds that can persist for long times in
aqueous solutions due to their non-biodegradability properties. Besides,
the improper management of wastewater-containing MO and MB dyes,
which have carcinogenic and toxic properties, causes a major threat for
human beings (Haque et al., 2010). Consequently, different methods as
adsorption, coagulation, biological, advanced oxidation, or precipita­
tion have been utilized in cleaning the contaminated water. Based on the
benefits and drawbacks associated to each treatment method, adsorp­
tion is recommended as low-cost, simple, and more effective technique
(Homaeigohar, 2020). Especially, the adsorption processes using
* Corresponding author.
E-mail address: Moaaz.korany@science.bsu.edu.eg (M.K. Seliem).
Contents lists available at ScienceDirect
Journal of Contaminant Hydrology
journal homepage: www.elsevier.com/locate/jconhyd
https://doi.org/10.1016/j.jconhyd.2022.103977
Received 27 October 2021; Received in revised form 13 January 2022; Accepted 10 February 2022
Journal of Contaminant Hydrology 247 (2022) 103977
2
nanoparticles–based adsorbents that have different active adsorption
sites and great surface areas are very promising for water purification
(Ahmadijokani et al., 2020).
Metal-organic frameworks (MOFs), a new group of porous crystalline
materials, consist of organic ligands and the desired metal clusters
(Younas et al., 2020). Recently, MOFs have been employed in various
applications as catalysis, storage of hydrogen, separation of gases, drug
delivery, CO2 capture, sensing, and wastewater purification (Younas
et al., 2020). The suitability of MOFs for different applications is due to
the appropriate pore size, high surface area, simplicity of preparation,
and the high stability of these porous coordination polymers (Mandal
et al., 2018; Jhung et al., 2012; Cavka et al., 2008). Moreover, the
simplicity for modification of MOFs structure can change their textural
properties thus contributing to increase and widen the acceptability of
these materials in numerous fields. Generally, MOFs are fabricated via
bridging metal ion clusters with metal ions using organic linkers and the
experimental parameters as reaction time, pressure, temperature, solu­
tion pH, and the used solvent are considered (Safaei et al., 2019). Defects
in the preparation conditions (i.e., metal ion, organic ligand, metal-
ligand coordination geometry, pore surface hydrophobicity) have
contributed to decrease the stability of MOFs for industrial applications
(Adegoke et al., 2020; Burtch et al., 2014; Wang et al., 2016; Yuan et al.,
2018). A water–stable MOFs type, MIL-101 (MIL: Material Institute
Lavoisier), was combined with different metal precursors as chromium
(Cr) (MIL-101-Cr), aluminum (Al) (MIL-101-Al), and iron (Fe) (MIL-101-
Fe) and finally used in water purification (Li et al., 2019a; Zhao et al.,
2018).
Different varieties of MOFs such as an amino-functionalized Zr-based
MOFs (Lv et al., 2019), RGO/ MOFs (Liu et al., 2019), MIL-101-Cr, MIL-
53-Al and ZIF-8 (Zhang et al., 2020), Cu-BTC-1 (Mantasha et al., 2020),
Cu-BTC (Kaur et al., 2019), MOF-235 (Haque et al., 2011), UiO-66
(Molavi et al., 2018), and Fe-MIL-101 (Konik et al., 2019) were used
in the adsorption of MB and MO. In several adsorption systems, the
experimental data modeling was conducted with the Langmuir and
Freundlich equations. The fundaments and hypothesis of these classical
models cannot provide a scientific meaning and reliable explanation for
the influence of operating parameters as the adsorbate concentration
and temperature in the removal mechanism (Ramadan et al., 2021;
Sharib et al., 2021). To obtain a better understanding of the MO and MB
removal mechanisms by MIL-101(Fe), it is mandatory to use the
Fig. 1. Characterization results of MIL-101(Fe) adsorbent: (a) XRD pattern, (b, c) SEM images, and (d) FTIR spectrum.
M. Shakly et al.
Journal of Contaminant Hydrology 247 (2022) 103977
3
advanced adsorption models based on the statistical physics theory
(Mohamed et al., 2020; Seliem et al., 2020). This theory is appropriate to
provide additional steric and energetic parameters for the analysis and
explanation of the adsorption mechanism. The interpretation of the
physicochemical factors associated to MB and MO adsorption onto MIL-
101 (Fe) can contribute to clarify the adsorption mechanisms at the
molecular level (Ramadan et al., 2021; Sharib et al., 2021; Mohamed
et al., 2020). The main objectives of the current study were to (a)
characterize MIL-101(Fe) metal organic framework prepared through a
facile solvothermal method, (b) determine the removal performance of
the prepared MIL-101(Fe) for MB and MO dyes in single and binary
systems, (c) use different models (i.e., kinetics and isotherms) to
describe the adsorption processes, (d) provide new results and a deeper
understanding of the interactions between dyes molecules and MIL-101
(Fe) active sites via statistical physics calculations, and (e) evaluate the
possibility of the multiple use of MIL-101(Fe) for the adsorption of
cationic and anionic dyes. Overall, the results of this study contributed
with new insights into MO and MB adsorption performance and mech­
anism using MIL-101(Fe) as an adsorbent.
2. Materials used and methods
2.1. Materials
Ferric chloride hexahydrate (FeCl3.6H2O: Alpha Chemica India),
terephthalic acid (Benzene-1,4-dicarboxylic acid BDC: Merck KGaA
Germany) and N, N-dimethyl formamide (DMF: Techno Pharm chem
India) were used in this study in the preparation of MIL-101(Fe)
adsorbent. MO and MB dyes and absolute ethyl alcohol (99.9%) were
supplied from Merck KGaA, Germany and the two dyes were tested as
adsorbates.
2.2. Synthesis of MIL-101(Fe)
MIL-101(Fe) was synthesized by a hydrothermal method following a
similar procedure reported by (Hu et al., 2019). Ferric chloride hexa­
hydrate (FeCl3.6H2O) and Benzene-1,4-dicarboxylic acid was dissolved
in DMF with 2:1 ratio, the suspended solution was sonicated for 20 min
until was fully homogenous, and then it was transferred into a Teflon-
lined stainless-steel autoclave and heated at 110 ◦
C for 20 h. The light
orange product was separated via centrifugation and purified by
washing with DMF for 3 h followed by hot ethanol for 2 h. This step was
repeated for three times in order to remove the unreacted BDC. The
purified product was dried at 70 ◦
C for 30 min, and then activated at
150 ◦
C for 10 h before carrying out the dye adsorption studies.
2.3. Characterization of MIL-101 (Fe)
MIL-101 (Fe) was characterized using different techniques to inves­
tigate its physical and chemical properties. X–ray diffraction (XRD)
pattern of this adsorbent was recorded in the 2θ range from 5.02–79.98◦
using a Panalytical Philips diffractometer operated at 40 kV and 35 mA
under Cu–kα radiation (λ = 0.154 nm). The surface morphology of MIL-
101 (Fe) was observed using (Zeiss Sigma 500 VP analytical FE-SEM).
The samples were coated with gold before analysis to enhance the sur­
face conductivity. The functional groups of MOF surface were investi­
gated with FTIR spectrum (Bruker optics -vertex 70 equipment) at
constant ambient temperature of 25 ◦
C by accumulating 10 scans at 1
cm− 1
resolution in the 4000–400 cm− 1
region. The Iso-ionic point and
the particles size range of MIL-101 (Fe) were obtained from Zeta po­
tential and sizer (Malevern Zeta sizer-nano series ZS 90). The ther­
mogravimetric analysis (TGA) and differential scanning calorimetry
(DSC) were performed to analyze the thermal stability of MIL-10(Fe) in
the temperature range of 25–1200 ◦
C with a heating rate of 10 ◦
C/min
(Labsys evo Setaram, France).
2.4. Adsorption experiments of MB and MO dyes
Stock solutions (1000 mg/L of MB and 2000 mg of MO) were pre­
pared, and the desired initial concentrations of the tested dyes were
obtained using dilutions with high purity distilled water. To avoid any
possible reactions associated to the photo catalytic process, all adsorp­
tion experiments were conducted in dark glasses in the absence of light.
After each adsorption experiment, the solutions containing MO and MB
dyes were separated from MIL-101 (Fe) adsorbent using syringe filters.
The concentrations of the studied dyes in the filtrated solutions were
quantified using a UV double beam spectrophotometer at 664 and 464
nm for MB and MO dyes, respectively. All MO and MB removal experi­
ments were performed in triplicates and the results were averaged for
Fig. 2. TGA (a) and DSC (b) analysis of MIL-101(Fe).
Fig. 3. Effect of pH on the removal of MO and MB dyes by MIL-101(Fe)
adsorbent. The error bars were obtained from triplicate experiments.
M. Shakly et al.
Journal of Contaminant Hydrology 247 (2022) 103977
4
data evaluation obtaining standard deviations below 5.2%.
2.5. MO and MB adsorption kinetics
To study the adsorption kinetics, the experiments were performed at
the following conditions: 0.15 g of MIL-101 (Fe) mass, pH 4.0 for MO
and 9.5 for MB, initial concentrations of 30 mg/L for MB and 100 mg/L
for MO, adsorption temperature of 20 ◦
C, and different contact times
ranging from 15 min to 14 h. The adsorbed amounts (qt) of MB and MO
after each time interval were calculated using Eq. (1).
qt (mg/g) = (C0–Ct)
V
m
(1)
Where C0 (mg/L) is the initial concentrations of MO and MB, Ct (mg/
Fig. 4. The adsorption kinetics of MB and MO onto MIL-101(Fe) adsorbent at 20 ◦
C. (a, b) contact time effect, (c, d) Pseudo-First Order, (e, f) Pseudo-Second Order
and (g, h) intra-particle diffusion.
M. Shakly et al.
Journal of Contaminant Hydrology 247 (2022) 103977
5
L) is the final concentration of the tested dyes after time (t), V (L) is the
solution volume, and m (g) is the mass of MIL-101 (Fe) adsorbent (g). To
study the MO and MB adsorption onto MIL-101(Fe) and to identify the
possible rate–controlling step, the kinetic results were fitted to the
pseudo-first order (Lagergren, 1898), the pseudo-second order (Ho and
McKay, 1999), and intra-particle diffusion (Weber and Morris, 1962)
models as given below:
ln(qe − qt) = lnqe − k1t Pseudo − first − order (2)
t
/
qt =
(
1
/
k2q2
e
)
+ t
/
qe Pseudo − second − order (3)
qt = kp t1/2
+ C Intra − particle diffusion (4)
where k1(min− 1
) and k2(g/mg min) indicate the adsorption rate
constants of the first order and the second-order kinetics, respectively.
For the intra-particle diffusion equation, kp(mg/g min) and C (mg/g)
represent the rate constant and the intercept value of this model,
respectively.
2.6. MO and MB adsorption isotherms
Equilibrium studies related to MO and MB adsorption onto MIL-101
(Fe) were performed at the next conditions: Initial dye concentrations
(25–500 mg/L for MB and 25–2000 mg/L), three temperatures (20, 40
and 60 ◦
C), 0.01 g of MIL-101 (Fe), and solution pH 4 for MO and 9.5 for
MB. In all these experiments, 10 mg of the tested adsorbent was mixed
with 30 mL of each dye solution at 300 rpm using digital orbital shaker
(SHO–2D, Germany). The liquid phases containing MO and MB were
Table 1
Parameters of Kinetic Models for the Adsorption of MO and MB onto MIL101-Fe.
Co (mg/L) K1 (min 1
) qe exp (mg/g) qe cal (mg/g) R2
Dye Pseudo-first order
MO 100 0.316 80.238 0.909
MB 28 0.073 2.5 0.903
Dye Pseudo-second order
MO 100 0.01 104 108.7 0.993
MB 28 4.538 27.1 27.1 1
Fig. 5. Modeling of adsorption isotherms of MB and MO dyes on MIL-101 (Fe) MOF using Langmuir and Freundlich equations.
M. Shakly et al.
Journal of Contaminant Hydrology 247 (2022) 103977
6
separated and the concentrations of these dyes in solutions were
determined. The equilibrium adsorption quantities of MO and MB dyes
(qe) were calculated using Eq. (1) assuming that t corresponded to the
equilibrium time.
2.6.1. Classical modeling analysis
The experimental data of MB and MO were fitted to the Langmuir
(1916) and Freundlich (Freundlich, 1906) equations to find the suitable
classical model that can describe the adsorption of dyes onto MIL-101
(Fe) at equilibrium. The non-linear relations defined the Langmuir
(Eq. (5)) and Freundlich (Eq. (6)) isotherm models are given below:
qe =
qmax KLCe
(1 + KLCe )
(Langmuir model) (5)
qe = KF Ce
1/n
(Freundlich model) (6)
where qmax (mg/g) is the maximum adsorption capacity and KL (L/
mg) represents the constant of the Langmuir model. On the other hand,
KF((mg/g)(mg/L)− 1/n
)) signifies the adsorption capacity and n is the
heterogeneity factor of the Freundlich model.
2.6.2. Advanced modeling analysis
Herein, three advanced statistical physics models (i.e., single layer,
double layer, and multilayer) were employed to investigate the dye–
MIL-101 (Fe) interaction (Ramadan et al., 2021; Sharib et al., 2021).
• Firstly, the single layer model assumes that MO and MB removal by
MIL-101 (Fe) was in the form of one layer directed by a distinctive
adsorption energy (ΔE) (i.e., the adsorption of each tested dye mol­
ecules via MIL-101 (Fe) active sites is nearly governed by a constant
energy). The mathematical expression of this advanced single layer
adsorption model is given by Eq. (7) (Sharib et al., 2021).
qe =
nDM
1 +
(
c1/2
c
)n (7)
where n is the number of MO and MB dye molecules captured by active
site of MIL-101(Fe) adsorbent, C1/2 signifies the concentration at half-
saturation related to the formed adsorbate layer, and DM is the density
of adsorption sites.
• Secondly, the double layer model suggests that these adsorption
systems are characterized by the formation of two layers of dyes
molecules (MB or MO) with two dissimilar adsorption energies (i.e.,
ΔE1 for dye– MIL-101 (Fe) interaction and ΔE2 for MB–MB or
MO–MO interaction). The mathematical expression of the double
layer model is denoted by Eq. (8) (Sharib et al., 2021).
qe = nDM
(
c
c1
)n
+ 2
(
c
c2
)2n
1 +
(
c
c1
)n
+
(
c
c2
)2n
(8)
where C1 and C2 describe the two concentrations at half saturation
attributed to the first and second removed dye layers, respectively.
• Finally, the multilayer model suggests that the MO and MB adsorp­
tion onto MIL-101(Fe) results in the formation of an adaptable but
limited number of adsorbed dye (i.e., MO or MB) layers. Conse­
quently, the total adsorbed layers number (1 + N2) of each tested dye
is related to a single layer (N2 = 0), double layer (N2 = 1), and
multilayer process (N2 > 1) (Mohamed et al., 2020). Similar to the
double layer model, two energies (ΔE1 and ΔE2) were associated to
this multilayer model (i.e., dye–MIL-101(Fe) and dye-dye
Table 2
Parameters of isotherm equations for the MB adsorption on MIL-101 (Fe) MOF.
Isotherm Model 20 ◦
C 40 ◦
C 60 ◦
C
Langmuir
qmax (mg/g) 402.00 391.67 352.96
kL (L/mg) 0.0365 0.0227 0.0165
R2
0.9613 0.9618 0.9660
Freundlich
kF ((mg/g)(mg/L)− 1/n
) 75.51 64.76 49.26
1/n 0.2881 0.3617 0.3283
R2
0.9408 0.9443 0.9554
Multiple layer
N2 3.12 3.08 2.49
Qsat (mg/g) 373.74 348.07 325.08
R2
0.9742 0.9962 0.9844
Table 3
Parameters of isotherm equations for the MO adsorption on MIL-101 (Fe) MOF.
Isotherm Model 20 ◦
C 40 ◦
C 60 ◦
C
Langmuir
qmax (mg/g) 1066.96 934.17 830.97
kL (L
mg
) 0.0077 0.0071 0.0068
R2
0.9449 0.9447 0.9398
Freundlich
kF ((mg/g)(mg/L)− 1/n
) 67.60 53.36 45.27
1/n 0.4019 0.4141 0.4190
R2
0.8834 0.8700 0.8537
Multiple layer
N2 5.36 5.01 3.76
Qsat (mg/g) 908.39 777.01 677.92
R2
0.9631 0.9655 0.9632
Table 4
Comparison of the adsorption capacities of several Metal Organic Frameworks
for the removal of MB & MO dyes.
Adsorbent Dye
adsorbate
Adsorption
capacity (mg/
g)
Optimum
pH
T
(◦
C)
Ref
Amine-
MOF-Fe
MB 312 9 20 (Paiman
et al.,
2020)
UiO-66 MB 90 5.5 25 (Chen
et al.,
2015)
NH2-UiO-
66
MB 96 5.5 25 (Luo et al.,
2017)
ZIF-67 MB 57.14 10 – (Haque
et al.,
2010)
MIL-101
(Fe)
MB 149 9 20 (Paiman
et al.,
2020)
MIL-101
(Fe)
MB 402 9.5 20 This study
MIL-101 MO 140 3.5 45 (Haque
et al.,
2010)
UiO-66 MO 39 5.5 25 (Chen
et al.,
2015)
NH2-UiO-
66
MO 28 5.5 25 (Chen
et al.,
2015)
Cu-BDC MO 86.7 4 25 (Salama
et al.,
2018)
ZIF-67 MO 75.5 4 – (Luo et al.,
2017)
MIL-101
(Fe)
MO 1067 4 20 This study
M. Shakly et al.
Journal of Contaminant Hydrology 247 (2022) 103977
7
interactions). Generally, the adsorption energy (ΔE1) was related to
the interaction between the first adsorbed dye layer (unchanging
number and equal to unity) and MIL-101(Fe) adsorption sites, while
the other energy (ΔE2) was linked to the dye–dye interaction and
thus, ΔE1 is found to be greater than ΔE2. The multilayer adsorption
model is given by Mohamed et al. (2020); Li et al. (2019b):
qe = n DM
F1(c) + F2(c) + F3(c) + F4(c)
G(c)
(9)
F1(c) = −
2
(
c
c1
)2n
1 −
(
c
c1
)n +
(
c
c1
)n
(
1 −
(
c
c1
)2n
)
(
1 −
(
c
c1
)n )2
, (10)
F2(c) =
2
(
c
c1
)n(
c
c2
)n
(
1 −
(
c
c2
)n N2
)
1 −
(
c
c2
)n , (11)
F3(c) = − N2
(
c
c1
)n(
c
c2
)n(
c
c2
)n N2
1 −
(
c
c2
)n , (12)
F4(c) =
(
c
c1
)n(
c
c2
)2n
(
1 −
(
c
c2
)n N2
)
(
1 −
(
c
c2
)n )2
, (13)
Fig. 6. Modeling of adsorption isotherms of MO and MB dyes on MIL-101 (Fe) MOF using a multilayer statistical physics model.
M. Shakly et al.
Journal of Contaminant Hydrology 247 (2022) 103977
8
G(c) =
(
1 −
(
c
c1
)2n
)
1 −
(
c
c1
)n +
(
c
c1
)n(
c
c2
)n
(
1 −
(
c
c2
)n N2
)
(
1 −
(
c
c2
)n )2
, (14)
where N2 is the layers of removed dye molecules, C1 is the half saturation
concentration connected to the first layer, and C2 is the corresponding
one correlated to N2 adsorbed dye layers.
2.7. Regeneration of MIL-101(Fe) adsorbent
Recycling and utilization of the MIL-101(Fe) adsorbent were inves­
tigated via the solvent method using ethanol as an eluent. The MIL-101
(Fe) loaded with the tested dyes (MB and MO) was continuously agitated
on a rotatory shaker at 200 rpm for 4 h at 25 ◦
C to achieve the complete
removal of the adsorbed dyes. This adsorption/desorption test was
repeated five times under the same conditions. At the end of each
desorption round, the MIL-101(Fe) was washed several times by distilled
water and dried at 65 ◦
C for 12 h before the next desorption cycle.
3. Results and discussion
3.1. Characterization of MIL-101 (Fe) adsorbent
X-ray diffraction (XRD) was carried out to identify the crystallinity
and structure of the synthesized MIL-101 (Fe) adsorbent. Different peaks
related to the MIL-101 (Fe) were observed at 2 thetas = 5.3◦
, 8.82◦
,
9.18◦
and 18.5◦
, see Fig. 1a. These detected peaks were comparable with
the previous studies (Hu et al., 2019; Wang et al., 2018; Liu et al., 2018a;
Li et al., 2016).
FE-SEM images (Fig. 1b and c) display the morphological features of
the MIL-101 (Fe) surface. It was observed that the MIL-101(Fe) had a
special octahedral structure with diameters from 0.7 to 2 μm, which was
consistent to the other reported studies (Hu et al., 2019; Wang et al.,
2018).
FTIR spectrum (Fig. 1 d) shows a strong band at 1574 cm− 1
that
indicated (C=O) of the carboxylate group (Hu et al., 2019). The broad
band detected at 3376 cm− 1
corresponded to the hydroxyl group of the
adsorbed water molecules (Li et al., 2019a). The strong band at 1392
cm− 1
was attributed to the vibration of (C–C) in benzene ring; however,
the parent bands at 545 and 748 cm− 1
were related to (Fe-OH) vibration
MB dye MO dye
n
DM
(mg/g)
Qsat
(mg/g)
20 °C 40 °C 60 °C 20 °C 40 °C 60 °C
0.0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
1.6
1.8
2.0
0.0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
0
10
20
30
40
50
60
70
80
90
0
20
40
60
80
100
120
140
160
0
50
100
150
200
250
300
350
400
0
100
200
300
400
500
600
700
800
900
1,000
Fig. 7. Statistical physics parameters for the MB and MO adsorption on MIL-101 (Fe) MOF.
M. Shakly et al.
Journal of Contaminant Hydrology 247 (2022) 103977
9
(Wang et al., 2018). These FTIR results also agreed to the earlier studies
(Li et al., 2019a; Hu et al., 2019; Wang et al., 2018) and thus confirmed
the preparation of MIL-101(Fe).
Particles size range of MIL-101(Fe) was obtained using zetasizer
measurement and results indicated that it ranged from 470 to 603 nm.
Also, the surface charge value of the prepared MIL-101(Fe) was − 3 at
pH 7.5.
TGA analysis displayed the presence of two main weight losses
within the temperature range of 25–650 ◦
C (Fig. 2a). The first weight
loss (about 16%) below 400 ◦
C could be related to the elimination of
water molecules and free terephthalates in the pores of MIL-101(Fe). In
addition, the significant second weight loss (38%) from 400 to 650 ◦
C
could be due to the decomposition of organic ligands that result in the
collapse of MIL-101(Fe). TGA result of this study is similar to that of
other UIO-66 and MIL-101(Cr) metal organic frameworks (Hu et al.,
2015; Yang and Yan, 2011). Furthermore, the DCS result displays the
existence of two sharp endothermic peaks (Fig. 2b). The first endo­
thermic sharp peak observed at 250–300 ◦
C could be attributed to the
evaporation of H2O molecules and free terephthalate in pores of the MIL-
101(Fe), while the second one at 600–700 ◦
C could be associated to the
decomposition of MIL-101(Fe). Consequently, TGA and DSC results
indicated that MIL-101(Fe) is characterized by high thermal stability up
to 650 ◦
C.
3.2. pH effect on the adsorption of MB and MO onto MIL-101(Fe)
Adsorption experiments were conducted at a wide range of solution
pH (i.e., pH 2–12) to find the optimum value of this parameter for the
removal of MB and MO (100 mg/L) using 0.01 g of MIL-101(Fe) at 20 ◦
C.
Fig. 3 displays that the removal efficiency for MB increased with
improving pH value where the highest efficiency was achieved at pH
9.5. This behavior for MB adsorption was related to the electrostatic
attraction between the positively charged MB molecule and negatively
charged MIL-101(Fe) surface, whose pH at zero point of charge was
pHPZC = 7.1. Thus, the surface of MIL-101 (Fe) was negatively charged
at pH > 7.1 and it was more effective to attractive and remove the
cationic MB dye. On the contrary case, the maximum percentage of MO
adsorption was attained at pH 4 due to the electrostatic attraction force
caused by the negatively charged anionic MO dye. This indicated that
the MIL-101(Fe) was found to be an excellent adsorbent for the tested
dyes and thus, all adsorption experiments were conducted at pH 9.5 and
pH 4 for MB and MO, respectively.
3.3. Contact time effect on MB and MO adsorption using MIL-101(Fe)
Very rapid adsorption of both MB and MO dyes were observed in the
first 15 min with removal percentages of 52.7 and 76% for MO and MB
respectively, see Fig. 4a and b. It was assumed that the adsorption
process of MB and MO molecules was very fast at the beginning due to
the availability of numerous active sites on the MIL-101(Fe) surface.
After that, the interaction time between dyes and MIL-101(Fe) started to
be slower (i.e., a gentle slope was observed) due to the limitation of the
available adsorption sites. Finally, the removed amounts of both dyes
were nearly constant after contact times of 2 h for MB and 10 h for MO,
which indicated that equilibrium was achieved after the saturation of
active sites of the MIL-101(Fe) surface.
3.4. MB and MO adsorption kinetics
As indicated, the kinetics of dyes adsorption were investigated using
different models as the Pseudo-first order (Lagergren, 1898), Pseudo
-second order (Ho and McKay, 1999), and intra-particle diffusion
models (Weber and Morris, 1962). The Pseudo-first order considers that
the adsorption process is strongly depends on the diffusion, while the
Pseudo-second order assumes that a chemical adsorption can be
involved in the removal process (Ahmadijokani et al., 2020). The
experimental data of MO and MB were fitted to these different kinetic
models and the best fit model was identified via the R2
value, see Fig. 4.
All the kinetics parameters were calculated and compared with the
experimental data as shown in Table 1. The values of R2
for MO
adsorption on MIL-101(Fe) indicating that the best fit model was the
Pseudo-Second-order equation (R2
= 0.993) in comparison to the
Pseudo-First order equation (R2
= 0.909). Also, the best fitted model (R2
= 1) for MB was the Pseudo-Second order kinetics equation followed by
the Pseudo-first order (R2
= 0.903). These results indicated that the
adsorption kinetics of both MB and MO on MIL-101 (Fe) were described
by the Pseudo-Second-order model and thus, the adsorption of MO and
MB could be related to chemisorption process. The intra-particle diffu­
sion analysis attributed to each dye molecule displayed dissimilar linear
stages over the whole-time range (Fig. 4). The initial sharp stage can be
associated to the external mass transfer of MB and MO molecules from
the solution to the outer surface of the MIL-101(Fe) adsorbent. The
second and third linear sections characterized the control of pore-
diffusion and equilibrium stages, respectively. Consequently, the
adsorption of MB and MO was mostly directed by the film diffusion
during the first stage, followed by pore-diffusion and equilibrium at
subsequent stages.
3.5. Classical modeling of MB and MO adsorption isotherms
Modeling of MB and MO adsorption data at equilibrium with the
applied traditional models (Langmuir and Freundlich equations) are
displayed in Fig. 5 and the corresponding adjusted parameters of these
models are listed in Tables 2 and 3. Langmuir model was the best
alternative to fit MB and MO adsorption data where their R2
values were
highest at all tested temperatures. Therefore, the removal of these dyes
molecules was related to homogenous adsorption sites of MIL-101(Fe)
and the adsorbates–adsorbent interaction resulted in the formation of
a single dye layer. The maximum adsorption capacities (qmax) for MB
E
(kJ/mol)
E
(kJ/mol)
20 °C 40 °C 60 °C
0
5
10
15
20
25
30
35
40
E1
E2
0
2
4
6
8
10
12
14
16
18
20
E1
E2
MB
MO
Fig. 8. Adsorption energies for MB and MO adsorption on MOF MIL-101
(Fe) MOF.
M. Shakly et al.
Journal of Contaminant Hydrology 247 (2022) 103977
10
were 402, 391.67, and 352.96 mg/g and 1066.96, 934.17, and 830.97
mg/g for MO at 20, 40, and 60 ◦
C, as given in Table 2. The decrease of
qmax values with the increment of solution temperature suggested that
the adsorption of MB and MO molecules by MIL-101(Fe) was exothermic
(i.e., the dyes–adsorbent interaction was more energetic at low tem­
perature) (Sharib et al., 2021; Mohamed et al., 2020). Moreover, the
adsorption capacities associated to MO molecules were higher than
those of MB dye, which indicated the selectivity of the as–synthesized
Fig. 9. Adsorption selectivity of MIL-101(Fe) for the adsorption of MB and MO dyes in binary systems.
Fig. 10. MB and MO removal after the regeneration of MIL-101 (Fe) adsorbent.
M. Shakly et al.
Journal of Contaminant Hydrology 247 (2022) 103977
11
MIL-101(Fe) active sites for this azo dye. This variation in the removal
efficiency of MIL-101(Fe) could be related to the chemical structures (i.
e., C14H14N3NaO3S for MO and C16H18ClN3S for MB) and the molecular
size (i.e., 1.2 nm for MO and 1.38 nm for MB) of these dye molecules
(Seliem et al., 2020). The molecular characteristics of MO and MB
played a relevant role for transfer phenomena in the interface between
these dyes and MIL-101(Fe) adsorbent surface. Furthermore, the
decrease of KF values from 75.51 to 49.26 for MB and from 67.60 to
45.27 (mg/g)(mg/L)− 1/n
with temperature also supported the
exothermic adsorption of these dyes. Calculated 1/n values by the
Freundlich equation were lower than unity thus signifying an effective
dye–MIL-101(Fe) interaction even at low concentrations of these
organic pollutants (Sharib et al., 2021; Mohamed et al., 2020). The
maximum adsorption capacities for MO and MB by the synthetic MIL-
101(Fe) and other MOFs are listed in Table 4. It can be observed that
the qmax values of MIL-101(Fe) were higher than those reported for
different MOFs. Therefore, it can be concluded that the MIL-101(Fe) is a
promising adsorbent for removal of dyes-bearing solutions.
3.6. Advanced statistical modeling of MB and MO adsorption isotherms
The multilayer model (Fig. 6) presented the highest R2
values
compared to the single- and double-layer models at all solution tem­
peratures (i.e., 20, 40, and 60 ◦
C). Therefore, the steric (e.g., n, DM, and
Qsat) and energetic (ΔE) parameters of this multilayer adsorption model
were deeply analyzed to obtain a better understanding of the adsorption
mechanism of MB and MO dyes on MIL-101(Fe).
3.6.1. The n parameter
The determination of steric n parameter provides information asso­
ciated to the removal mechanism of the tested adsorbates (MB and MO
dyes) on MIL-101(Fe) especially in terms of adsorption orientation and
molecular aggregation phenomenon. In particular, the adsorption ge­
ometry (vertical or horizontal) of the removed MB and MO molecules on
the adsorbent can be analyzed using this parameter (Ramadan et al.,
2021; Mohamed et al., 2020; Barakat et al., 2020). Based on the value of
n parameter, two main scenarios can be recognized to describe the ge­
ometry and mechanism of dyes adsorption onto MIL-101(Fe) surfaces
(Sharib et al., 2021; Barakat et al., 2020).
• The first scenario (n > 1): The adsorption of MO and MB molecules
can take place in a vertical position through multi–interactions
mechanism (i.e., one active site of MIL-101(Fe) can capture several
dye molecules).
• The second scenario (n < 1): MO and MB adsorption occurs in a
horizontal location via a multi–docking mechanism (i.e., numerous
MIL-101(Fe) functional groups can remove one dye molecule).
Fig. 7 exhibits the values of n parameter related to MB and MO
adsorption on MIL-101(Fe) as a function of solution temperature. For
MO dye, the n parameter values were 0.95, 1.02, and 1.33 at 20, 40, and
60 ◦
C, respectively. As a result, horizontal position and multi–docking
mechanism could be expected for MO adsorption at 20 ◦
C, while vertical
position and multi–interactions mechanism could occur at 40 and 60 ◦
C.
The n values were higher than unity at 40 and 60 ◦
C and this result
indicated the effect of aggregation phenomenon with increasing tem­
perature. On the other hand, the n values associated to MB adsorption
were 1.17, 1.68, and 1.78 and thus, vertical orientation and multi­
–interactions mechanism were involved in the removal of this dye by the
MIL-101(Fe) active sites at all temperatures. Consequently, the aggre­
gation of MB molecules (i.e., MB-MB binding) in solution was identified
at all tested temperatures and this phenomenon resulted in the existence
of the non-parallel adsorption geometry and a multi–molecular mech­
anism. The high degree of MB and MO aggregations with temperature
suggested that these adsorption processes were energetically activated
in solutions before the interaction between dyes and MIL-101(Fe)
surface (Li et al., 2020).
3.6.2. The DM parameter
Fig. 7 displays the number of MIL-101(Fe) active sites (i.e., DM
parameter) with respect the adsorption temperature. From 20 to 60 ◦
C,
the DM parameter decreased from 77.33 to 52.38 mg/g for MB and from
149.91 to 107.07 mg/g for MO dye. The decrease of DM value with
temperature displayed a reverse trend as compared to the steric n pa­
rameters (i.e., the increment of n parameter caused a reduction of the
occupied active sites number of the MIL-101(Fe) surface). Thus, the
trend of DM parameter assessed the accumulation of the removed dyes
molecules at high temperature. Moreover, the aggregated MO and MB
molecules could be more selective to definite functional groups of the as-
synthesized MIL-101(Fe) (Mobarak et al., 2019). Also, it can be noticed
that the DM values associated to MO dye were higher than those of MB
dye at all temperatures, which could be considered as an important
variable of the enhanced adsorption capacity of MIL-101(Fe) for MO
dye.
3.6.3. Total number of the adsorbed MB and MO layers (Nt = 1 + N2)
To understand the adsorption mechanism of MB and MO molecules
on MIL-101(Fe), the determination of the total layers of adsorbed dyes is
necessary (Barakat et al., 2020). The calculated Nt values were 3.12,
3.08, and 2.49 for MB and 5.36, 5.03, and 3.76 for MO at 20, 40, and
60 ◦
C, respectively, see Fig. 7. The reduction of Nt value with temper­
ature increments could be related to the influence of thermal agitation,
which resulted in a disordered movement of the adsorbed MB and MO
molecules. The disturbed state associated to dye-dye interaction caused
a decrease of the adsorbed dye layers (Seliem et al., 2020; Barakat et al.,
2020). Like the DM parameter, the Nt values associated to MO dye were
higher than those of MB dye at all adsorption temperatures (i.e., the N2
parameter presented a significant impact to improve the MO adsorption
capacity).
3.6.4. The adsorbed MB and MO quantities at saturation (Qsat = n. DM. Nt)
To estimate the maximum MIL-101(Fe) performance for the
adsorption of MO and MB dyes, the calculation of Qsat values was done.
Fig. 7 exhibits the trend of Qsat versus temperature, and the values of this
parameter were 373.74, 348.07, and 325.08 mg/g for MB and 908.39,
777.01, and 677.92 mg/g for MO at 20, 40, and 60 ◦
C, respectively.
These Qsat values confirmed the exothermic interactions associated to
MB and MO adsorption on MIL-101(Fe) surface. For the investigated
dyes, it was identified that the DM and Nt parameters preserved the same
trend of the Qsat parameter with respect to improving temperature.
According to the calculated steric parameters, it can be concluded that
the removal efficiency of MIL-101(Fe) for MO and MB dyes was
controlled by the DM and Nt parameters. Furthermore, the high values of
the Qsat for MO as compared to that of MB confirmed the important roles
of DM and Nt parameters to determine the preference of MIL-101(Fe) for
the adsorption of this azo dye.
3.7. Energetic parameters
To obtain an appropriate interpretation for the interactions between
MB and MO molecules and MIL-101(Fe) active sites, the adsorption
energies were calculated (Seliem et al., 2020; Barakat et al., 2020) These
adsorption energies (ΔE) were estimated as follows (Barakat et al.,
2020).
C1 = Cse−
ΔE1
RT (15)
C2 = Cse−
ΔE2
RT (16)
where c1and c2 are the half-saturation concentrations and cs is the sol­
ubility of the tested MB and MO adsorbates.
Fig. 8 shows the ΔE values as a function of the adsorption
M. Shakly et al.
Journal of Contaminant Hydrology 247 (2022) 103977
12
temperature. The negative ΔE values confirmed the exothermic in­
teractions between the dyes molecules and the MIL-101(Fe) adsorbent.
This performance agreed with the influence of solution temperature to
reduce the MO and MB adsorption capacities. Besides, the ΔE values of
MO and MB adsorption were < 40 kJ/mol at 20, 40, and 60 ◦
C, which
suggested physical adsorption processes (Sharib et al., 2021; Barakat
et al., 2020). ΔE1 was associated to MIL-101(Fe)–dye (MB or MO)
interaction, while ΔE2 described the dye–dye interface. Therefore, ΔE1
values were higher than ΔE2 all tested temperatures. Although the ΔE
values associated to MB adsorption were higher than those of MO at all
temperature, the adsorbed MB dye quantities were low and, conse­
quently, the selectivity of MIL-101(Fe) was not governed by the ΔE
parameter.
3.8. Selectivity of MIL-101 (Fe) for the adsorption of MB and MO dyes in
binary systems
The adsorption selectivity is an important parameter for studying the
behavior of an adsorbent for the removal of water pollutants (Liu et al.,
2018b). For instance, graphene like metal organic framework (BUC-17)
was found to be more selective to anionic dyes from an organic dye
mixture (Li et al., 2017). Moreover, Eu-based MOF (BUC-88) displayed
high selectivity to many pharmaceuticals and personal care products
(Wang et al., 2021). In this study, the adsorption selectivity of the MIL-
101(Fe) for the dye removal was investigated using a mixture of MB and
MO dyes at different pH values, keeping the other operating parameters
constant. Fig. 9 shows that MIL-101 (Fe) surface was selective to MO and
MB dyes at pH 4 and 9.5, respectively. Besides, the adsorption capacities
of the tested dyes were slightly decreased in the binary system as
compared to the single-component one at optimum pH values due to the
competition between MO and MB molecules to interact with MIL-101
(Fe) active sites (i.e., weak antagonism interaction). At pH 7, both
dyes were slightly adsorbed without a significant preference for any dye
molecule, and the adsorption processes could be associated to the π- π
stacking effect. These results suggested that the MIL-101 (Fe) MOF could
be a more selective adsorbent for anionic and cationic dyes depending
on the solution pH value.
3.9. Reusability study
Fig. 10 shows the removal efficiency of MIL-101(Fe) after five
regeneration cycles for both MB and MO dyes. It is clear that MIL-101
(Fe) presented more than 95% for MB removal and 90% for MO
removal after all regeneration rounds. For MB dye, increment of the
removal efficiency in the last cycles (i.e., from cycle 3 to 5) could be
attributed to the high surface area of MIL-101(Fe)-MB interaction,
which required frequent washing and drying under vacuum to achieve
the complete desorption of the attached MB molecules (Paiman et al.,
2020). Based on the regeneration results, the investigated adsorbent can
be reutilized numerous times to remove the tested dyes without a sig­
nificant loss of its removal efficiency, thus suggesting an economic
benefit and high stability of this adsorbent in the purification of dyes-
bearing water.
4. Conclusions
MIL-101(Fe) (MOF) was successfully prepared through a facile sol­
vothermal method, which was characterized and employed as an
adsorbent for MO and MB at different experimental conditions. In single
and binary adsorption systems, MIL-101(Fe) was more selective to MO
dye as compared to MB dye. The adsorption of both MO and MB fol­
lowed the pseudo-second order and the Langmuir equations. A multi­
layer model from statistical physics theory was utilized to understand
the MO and MB adsorption mechanisms at a molecular level. MO
adsorption was governed by multi–docking and multi–interactions
mechanisms, while MB adsorption was controlled only by
multi–interactions mechanism. The density of MIL-101(Fe) active sites
and the total adsorbed dye layers formed on MOF surface were the main
parameters that determined the adsorption capacity of MO dye. The
adsorption of MB and MO molecules by MIL-101(Fe) was mainly caused
by physical interactions at all solution temperatures. The reusability
study demonstrated that the MIL-101(Fe) can maintain an outstanding
performance after four adsorption/desorption cycles. The current study
clearly proved that the MIL-101(Fe) can be utilized as an adsorbent for
efficient removal of anionic and cationic dyes from polluted solutions.
Declaration of Competing Interest
The authors declare that they have no known competing financial
interests or personal relationships that could have appeared to influence
the work reported in this paper.
References
Adegoke, K.A., Agboola, O.S., Ogunmodede, J., Araoye, A.O., Bello, O.S., 2020. Metal-
organic frameworks as adsorbents for sequestering organic pollutants from
wastewater. Mater. Chem. Phys. 253, 123246.
Ahmadijokani, F., Mohammadkhani, R., Ahmadipouya, S., Shokrgozar, A.,
Rezakazemi, M., Molavi, H., Arjmand, M., 2020. Superior chemical stability of UiO-
66 metal-organic frameworks (MOFs) for selective dye adsorption. Chem. Eng. J.
399, 125346.
Barakat, M.A., Selim, A.Q., Mobarak, M., Kumar, R., Anastopoulos, I.,
Giannakoudakis, D.A., Bonilla-Petriciolet, A., Mohamed, E.A., Seliem, M.K.,
Komarneni, S., 2020. Experimental and theoretical studies of methyl orange uptake
by Mn–rich synthetic mica: Insights into manganese role in adsorption and
selectivity. Nanomaterials 10, 1–19.
Burtch, N.C., Jasuja, H., Walton, K.S., 2014. Water stability and adsorption in
metal–organic frameworks. Chem. Rev. 114 (20), 10575–10612.
Cavka, J.H., Jakobsen, S., Olsbye, U., Guillou, N., Lamberti, C., Bordiga, S., Lillerud, K.P.,
2008. A new zirconium inorganic building brick forming metal organic frameworks
with exceptional stability. J. Am. Chem. Soc. 130 (42), 13850–13851.
Chen, Q., He, Q., Lv, M., Xu, Y., Yang, H., Liu, X., Wei, F., 2015. Selective adsorption of
cationic dyes by UiO-66-NH2. Appl. Surf. Sci. 327, 77–85.
Freundlich, H.M.F., 1906. Over the adsorption in solution. J. Phys. Chem. 57 (385471),
1100–1107.
Haque, E., Lee, J.E., Jang, I.T., Hwang, Y.K., Chang, J.S., Jegal, J., Jhung, S.H., 2010.
Adsorptive removal of methyl orange from aqueous solution with metal-organic
frameworks, porous chromium-benzenedicarboxylates. J. Hazard. Mater. 181 (1–3),
535–542.
Haque, E., Jun, J.W., Jhung, S.H., 2011. Adsorptive removal of methyl orange and
methylene blue from aqueous solution with a metal-organic framework material,
iron terephthalate (MOF-235). J. Hazard. Mater. 185 (1), 507–511.
Ho, Y.S., McKay, G., 1999. Pseudo-second order model for sorption processes. Process
Biochem. 34 (5), 451–465.
Homaeigohar, S., 2020. The nanosized dye adsorbents for water treatment.
Nanomaterials 10 (2), 295.
Hu, Z., Khurana, M., Seah, Y.H., Zhang, M., Guo, Z., Zhao, D., 2015. Ionized Zr-MOFs for
highly efficient post-combustion CO2 capture. Chem. Eng. Sci. 124, 61–69.
Hu, H., Zhang, H., Chen, Y., Ou, H., 2019. Enhanced photocatalysis using metal–organic
framework MIL-101 (Fe) for organophosphate degradation in water. Environ. Sci.
Pollut. Res. 26 (24), 24720–24732.
Jhung, S.H., Khan, N.A., Hasan, Z., 2012. Analogous porous metal–organic frameworks:
synthesis, stability and application in adsorption. CrystEngComm 14 (21),
7099–7109.
Kaur, R., Kaur, A., Umar, A., Anderson, W.A., Kansal, S.K., 2019. Metal organic
framework (MOF) porous octahedral nanocrystals of Cu-BTC: synthesis, properties
and enhanced adsorption properties. Mater. Res. Bull. 109, 124–133.
Konik, P.A., Berdonosova, E.A., Savvotin, I.M., Klyamkin, S.N., 2019. The influence of
amide solvents on gas sorption properties of metal-organic frameworks MIL-101 and
ZIF-8. Microporous Mesoporous Mater. 277, 132–135.
Lagergren, S., 1898. Zur theorie der sogenannten adsorption geloster stoffe.
Langmuir, I., 1916. The constitution and fundamental properties of solids and liquids.
Part I. Solids. J. Am. Chem. Soc. 38 (11), 2221–2295.
Li, X., Guo, W., Liu, Z., Wang, R., Liu, H., 2016. Fe-based MOFs for efficient adsorption
and degradation of acid orange 7 in aqueous solution via persulfate activation. Appl.
Surf. Sci. 369, 130–136.
Li, J.J., Wang, C.C., Fu, H.F., Cui, J.R., Xu, P., Guo, J., Li, J.R., 2017. High-performance
adsorption and separation of anionic dyes in water using a chemically stable
graphene-like metal–organic framework. Dalton Trans. 46 (31), 10197–10201.
Li, Z., Liu, X., Jin, W., Hu, Q., Zhao, Y., 2019a. Adsorption behavior of arsenicals on MIL-
101 (Fe): the role of arsenic chemical structures. J. Colloid Interface Sci. 554,
692–704.
Li, Z., Sellaoui, L., Dotto, G.L., Lamine, A.B., Bonilla-Petriciolet, A., Hanafy, H., Erto, A.,
2019b. Interpretation of the adsorption mechanism of Reactive Black 5 and Ponceau
4R dyes on chitosan/polyamide nanofibers via advanced statistical physics model.
J. Mol. Liq. 285, 165–170.
M. Shakly et al.
Journal of Contaminant Hydrology 247 (2022) 103977
13
Li, Z., Hanafy, H., Zhang, L., Sellaoui, L., Netto, M.S., Oliveira, M.L., Li, Q., 2020.
Adsorption of congo red and methylene blue dyes on an ashitaba waste and a walnut
shell-based activated carbon from aqueous solutions: experiments, characterization
and physical interpretations. Chem. Eng. J. 388, 124263.
Liu, J., Ye, J., Chen, Y., Li, C., Ou, H., 2018a. UV-driven hydroxyl radical oxidation of tris
(2-chloroethyl) phosphate: intermediate products and residual toxicity.
Chemosphere 190, 225–233.
Liu, A., Wang, C.C., Wang, C.Z., Fu, H.F., Peng, W., Cao, Y.L., Du, A.F., 2018b. Selective
adsorption activities toward organic dyes and antibacterial performance of silver-
based coordination polymers. J. Colloid Interface Sci. 512, 730–739.
Liu, Y., Zhu, M., Chen, M., Ma, L., Yang, B., Li, L., Tu, W., 2019. A polydopamine-
modified reduced graphene oxide (RGO)/MOFs nanocomposite with fast rejection
capacity for organic dye. Chem. Eng. J. 359, 47–57.
Luo, X.P., Fu, S.Y., Du, Y.M., Guo, J.Z., Li, B., 2017. Adsorption of methylene blue and
malachite green from aqueous solution by sulfonic acid group modified MIL-101.
Microporous Mesoporous Mater. 237, 268–274.
Lv, S.W., Liu, J.M., Ma, H., Wang, Z.H., Li, C.Y., Zhao, N., Wang, S., 2019. Simultaneous
adsorption of methyl orange and methylene blue from aqueous solution using amino
functionalized Zr-based MOFs. Microporous Mesoporous Mater. 282, 179–187.
Mandal, T.N., Karmakar, A., Sharma, S., Ghosh, S.K., 2018. Metal-organic frameworks
(MOFs) as functional supramolecular architectures for anion recognition and
sensing. Chem. Rec. 18 (2), 154–164.
Mantasha, I., Saleh, H.A., Qasem, K.M., Shahid, M., Mehtab, M., Ahmad, M., 2020.
Efficient and selective adsorption and separation of methylene blue (MB) from
mixture of dyes in aqueous environment employing a Cu (II) based metal organic
framework. Inorg. Chim. Acta 511, 119787.
Mobarak, M., Mohamed, E.A., Selim, A.Q., Mohamed, F.M., Sellaoui, L., Bonilla-
Petriciolet, A., Seliem, M.K., 2019. Statistical physics modeling and interpretation of
methyl orange adsorption on high–order mesoporous composite of MCM–48 silica
with treated rice husk. J. Mol. Liq. 285, 678–687.
Mohamed, E.A., Selim, A.Q., Ahmed, S.A., Sellaoui, L., Bonilla-Petriciolet, A., Erto, A.,
Seliem, M.K., 2020. H2O2-activated anthracite impregnated with chitosan as a novel
composite for Cr (VI) and methyl orange adsorption in single-compound and binary
systems: Modeling and mechanism interpretation. Chem. Eng. J. 380, 122445.
Molavi, H., Hakimian, A., Shojaei, A., Raeiszadeh, M., 2018. Selective dye adsorption by
highly water stable metal-organic framework: long term stability analysis in aqueous
media. Appl. Surf. Sci. 445, 424–436.
Paiman, S.H., Rahman, M.A., Uchikoshi, T., Abdullah, N., Othman, M.H.D., Jaafar, J.,
Ismail, A.F., 2020. Functionalization effect of Fe-type MOF for methylene blue
adsorption. J. Saudi Chem. Soc. 24 (11), 896–905.
Ramadan, H.S., Mobarak, M., Lima, E.C., Bonilla-Petriciolet, A., Li, Z., Seliem, M.K.,
2021. Cr (VI) adsorption onto a new composite prepared from Meidum black clay
and pomegranate peel extract: Experiments and physicochemical interpretations.
J. Environ. Chem. Eng. 9 (4), 105352.
Safaei, M., Foroughi, M.M., Ebrahimpoor, N., Jahani, S., Omidi, A., Khatami, M., 2019.
A review on metal-organic frameworks: synthesis and applications. TrAC Trends
Anal. Chem. 118, 401–425.
Salama, R.S., El-Hakama, S.A., Samraa, S.E., El-Dafrawya, S.M., Ahmeda, A.I., 2018.
Adsorption, equilibrium and kinetic studies on the removal of methyl orange dye
from aqueous solution by using of copper metal organic framework (Cu-BDC). Int. J.
Modern Chem 10 (2), 195–207.
Seliem, M.K., Mobarak, M., Selim, A.Q., Mohamed, E.A., Halfaya, R.A., Gomaa, H.K.,
Dotto, G.L., 2020. A novel multifunctional adsorbent of pomegranate peel extract
and activated anthracite for Mn (VII) and Cr (VI) uptake from solutions: experiments
and theoretical treatment. J. Mol. Liq. 311, 113169.
Sharib, A.S.A., Bonilla-Petriciolet, A., Selim, A.Q., Mohamed, E.A., Seliem, M.K., 2021.
Utilizing modified weathered basalt as a novel approach in the preparation of Fe3O4
nanoparticles: experimental and theoretical studies for crystal violet adsorption.
J. Environ. Chem. Eng. 9 (6), 106220.
Wang, C., Liu, X., Demir, N.K., Chen, J.P., Li, K., 2016. Applications of water stable
metal–organic frameworks. Chem. Soc. Rev. 45 (18), 5107–5134.
Wang, D., Jia, F., Wang, H., Chen, F., Fang, Y., Dong, W., Yuan, X., 2018. Simultaneously
efficient adsorption and photocatalytic degradation of tetracycline by Fe-based
MOFs. J. Colloid Interface Sci. 519, 273–284.
Wang, C.Y., Wang, C.C., Zhang, X.W., Ren, X.Y., Yu, B., Wang, P., Fu, H., 2021. A new
Eu-MOF for ratiometrically fluorescent detection toward quinolone antibiotics and
selective detection toward tetracycline antibiotics. Chin. Chem. Lett. https://doi.
org/10.1016/j.cclet.2021.08.095.
Weber, W.J., Morris, J.C., 1962, September. Advances in water pollution research. In:
Proceedings of the First International Conference on Water Pollution Research, Vol.
2. Pergamon Press, Oxford, p. 231.
Yang, C.X., Yan, X.P., 2011. Metal–organic framework MIL-101 (Cr) for high-
performance liquid chromatographic separation of substituted aromatics. Anal.
Chem. 83 (18), 7144–7150.
Younas, M., Rezakazemi, M., Daud, M., Wazir, M.B., Ahmad, S., Ullah, N.,
Ramakrishna, S., 2020. Recent progress and remaining challenges in post-
combustion CO2 capture using metal-organic frameworks (MOFs). Prog. Energy
Combust. Sci. 80, 100849.
Yuan, S., Feng, L., Wang, K., Pang, J., Bosch, M., Lollar, C., Zhou, H.C., 2018. Stable
metal–organic frameworks: design, synthesis, and applications. Adv. Mater. 30 (37),
1704303.
Zhang, X., Qian, L., Yang, S., Peng, Y., Xiong, B., Li, J., He, C., 2020. Comparative studies
of methyl orange adsorption in various metal-organic frameworks by nitrogen
adsorption and positron annihilation lifetime spectroscopy. Microporous
Mesoporous Mater. 296, 109993.
Zhao, T., Yang, L., Feng, P., Gruber, I., Janiak, C., Liu, Y., 2018. Facile synthesis of nano-
sized MIL-101 (Cr) with the addition of acetic acid. Inorg. Chim. Acta 471, 440–445.
M. Shakly et al.

Contenu connexe

Similaire à New insights into the selective adsorption mechanism of cationic and anionic dyes using MIL-101(Fe) metal-organic framework Modeling and interpretation of physicochemical parameters.pdf

Estimating the Biodegradation Kinetics by Mixed Culture Degrading Pyrene (Pyr)
Estimating the Biodegradation Kinetics by Mixed Culture Degrading Pyrene (Pyr)Estimating the Biodegradation Kinetics by Mixed Culture Degrading Pyrene (Pyr)
Estimating the Biodegradation Kinetics by Mixed Culture Degrading Pyrene (Pyr)
AZOJETE UNIMAID
 
A critical review on the recent progress in application of electro-Fenton pro...
A critical review on the recent progress in application of electro-Fenton pro...A critical review on the recent progress in application of electro-Fenton pro...
A critical review on the recent progress in application of electro-Fenton pro...
plalak6330
 
IC Catalysis by FeSqMOF with pg nos
IC Catalysis by FeSqMOF with pg nosIC Catalysis by FeSqMOF with pg nos
IC Catalysis by FeSqMOF with pg nos
Soumyabrata Goswami
 
Metal-OrganicFramework-BasedEngineered.pdf
Metal-OrganicFramework-BasedEngineered.pdfMetal-OrganicFramework-BasedEngineered.pdf
Metal-OrganicFramework-BasedEngineered.pdf
huu_trinh
 
Sonocatalytic Performance of Fe3O4 Cluster Microspheres Gratiphic Carbon Comp...
Sonocatalytic Performance of Fe3O4 Cluster Microspheres Gratiphic Carbon Comp...Sonocatalytic Performance of Fe3O4 Cluster Microspheres Gratiphic Carbon Comp...
Sonocatalytic Performance of Fe3O4 Cluster Microspheres Gratiphic Carbon Comp...
ijtsrd
 
A review of imperative technologies for for waste water tratament 1
A review of imperative technologies for for waste water tratament 1A review of imperative technologies for for waste water tratament 1
A review of imperative technologies for for waste water tratament 1
Edgy Rod
 

Similaire à New insights into the selective adsorption mechanism of cationic and anionic dyes using MIL-101(Fe) metal-organic framework Modeling and interpretation of physicochemical parameters.pdf (20)

Estimating the Biodegradation Kinetics by Mixed Culture Degrading Pyrene (Pyr)
Estimating the Biodegradation Kinetics by Mixed Culture Degrading Pyrene (Pyr)Estimating the Biodegradation Kinetics by Mixed Culture Degrading Pyrene (Pyr)
Estimating the Biodegradation Kinetics by Mixed Culture Degrading Pyrene (Pyr)
 
Performance of a two Chambers Reactor for the Treatment of an Oily Effluent b...
Performance of a two Chambers Reactor for the Treatment of an Oily Effluent b...Performance of a two Chambers Reactor for the Treatment of an Oily Effluent b...
Performance of a two Chambers Reactor for the Treatment of an Oily Effluent b...
 
20120140504007 2
20120140504007 220120140504007 2
20120140504007 2
 
REMEDIATION STRATEGY
REMEDIATION STRATEGYREMEDIATION STRATEGY
REMEDIATION STRATEGY
 
A critical review on the recent progress in application of electro-Fenton pro...
A critical review on the recent progress in application of electro-Fenton pro...A critical review on the recent progress in application of electro-Fenton pro...
A critical review on the recent progress in application of electro-Fenton pro...
 
Fabrication and characterization of graphene oxide nanoparticles incorporated...
Fabrication and characterization of graphene oxide nanoparticles incorporated...Fabrication and characterization of graphene oxide nanoparticles incorporated...
Fabrication and characterization of graphene oxide nanoparticles incorporated...
 
Kinetics of Substituted Bis- and Mono-azo Dyes as Corrosion Inhibitors for Al...
Kinetics of Substituted Bis- and Mono-azo Dyes as Corrosion Inhibitors for Al...Kinetics of Substituted Bis- and Mono-azo Dyes as Corrosion Inhibitors for Al...
Kinetics of Substituted Bis- and Mono-azo Dyes as Corrosion Inhibitors for Al...
 
Walton-2015-Nature
Walton-2015-NatureWalton-2015-Nature
Walton-2015-Nature
 
IC Catalysis by FeSqMOF with pg nos
IC Catalysis by FeSqMOF with pg nosIC Catalysis by FeSqMOF with pg nos
IC Catalysis by FeSqMOF with pg nos
 
2014_CP
2014_CP2014_CP
2014_CP
 
Dr.C.Kavitha (1).pdf
Dr.C.Kavitha (1).pdfDr.C.Kavitha (1).pdf
Dr.C.Kavitha (1).pdf
 
Dr.C.Kavitha.pdf
Dr.C.Kavitha.pdfDr.C.Kavitha.pdf
Dr.C.Kavitha.pdf
 
Dr.P.Vijayasarathi.pdf
Dr.P.Vijayasarathi.pdfDr.P.Vijayasarathi.pdf
Dr.P.Vijayasarathi.pdf
 
Metal organic framework(MOF)
Metal organic framework(MOF)Metal organic framework(MOF)
Metal organic framework(MOF)
 
Metal-OrganicFramework-BasedEngineered.pdf
Metal-OrganicFramework-BasedEngineered.pdfMetal-OrganicFramework-BasedEngineered.pdf
Metal-OrganicFramework-BasedEngineered.pdf
 
66 song2012.pdf
66 song2012.pdf66 song2012.pdf
66 song2012.pdf
 
Facultative methylotrophic bacterium вrevibacterium
Facultative methylotrophic bacterium вrevibacteriumFacultative methylotrophic bacterium вrevibacterium
Facultative methylotrophic bacterium вrevibacterium
 
Sonocatalytic Performance of Fe3O4 Cluster Microspheres Gratiphic Carbon Comp...
Sonocatalytic Performance of Fe3O4 Cluster Microspheres Gratiphic Carbon Comp...Sonocatalytic Performance of Fe3O4 Cluster Microspheres Gratiphic Carbon Comp...
Sonocatalytic Performance of Fe3O4 Cluster Microspheres Gratiphic Carbon Comp...
 
COMPARATIVE STUDY BETWEEN REMOVERS AGENTS OF SILICON INTO THE SYNTHESIS OF MI...
COMPARATIVE STUDY BETWEEN REMOVERS AGENTS OF SILICON INTO THE SYNTHESIS OF MI...COMPARATIVE STUDY BETWEEN REMOVERS AGENTS OF SILICON INTO THE SYNTHESIS OF MI...
COMPARATIVE STUDY BETWEEN REMOVERS AGENTS OF SILICON INTO THE SYNTHESIS OF MI...
 
A review of imperative technologies for for waste water tratament 1
A review of imperative technologies for for waste water tratament 1A review of imperative technologies for for waste water tratament 1
A review of imperative technologies for for waste water tratament 1
 

Dernier

CNv6 Instructor Chapter 6 Quality of Service
CNv6 Instructor Chapter 6 Quality of ServiceCNv6 Instructor Chapter 6 Quality of Service
CNv6 Instructor Chapter 6 Quality of Service
giselly40
 
IAC 2024 - IA Fast Track to Search Focused AI Solutions
IAC 2024 - IA Fast Track to Search Focused AI SolutionsIAC 2024 - IA Fast Track to Search Focused AI Solutions
IAC 2024 - IA Fast Track to Search Focused AI Solutions
Enterprise Knowledge
 
Artificial Intelligence: Facts and Myths
Artificial Intelligence: Facts and MythsArtificial Intelligence: Facts and Myths
Artificial Intelligence: Facts and Myths
Joaquim Jorge
 

Dernier (20)

How to Troubleshoot Apps for the Modern Connected Worker
How to Troubleshoot Apps for the Modern Connected WorkerHow to Troubleshoot Apps for the Modern Connected Worker
How to Troubleshoot Apps for the Modern Connected Worker
 
CNv6 Instructor Chapter 6 Quality of Service
CNv6 Instructor Chapter 6 Quality of ServiceCNv6 Instructor Chapter 6 Quality of Service
CNv6 Instructor Chapter 6 Quality of Service
 
🐬 The future of MySQL is Postgres 🐘
🐬  The future of MySQL is Postgres   🐘🐬  The future of MySQL is Postgres   🐘
🐬 The future of MySQL is Postgres 🐘
 
08448380779 Call Girls In Friends Colony Women Seeking Men
08448380779 Call Girls In Friends Colony Women Seeking Men08448380779 Call Girls In Friends Colony Women Seeking Men
08448380779 Call Girls In Friends Colony Women Seeking Men
 
IAC 2024 - IA Fast Track to Search Focused AI Solutions
IAC 2024 - IA Fast Track to Search Focused AI SolutionsIAC 2024 - IA Fast Track to Search Focused AI Solutions
IAC 2024 - IA Fast Track to Search Focused AI Solutions
 
Finology Group – Insurtech Innovation Award 2024
Finology Group – Insurtech Innovation Award 2024Finology Group – Insurtech Innovation Award 2024
Finology Group – Insurtech Innovation Award 2024
 
Automating Google Workspace (GWS) & more with Apps Script
Automating Google Workspace (GWS) & more with Apps ScriptAutomating Google Workspace (GWS) & more with Apps Script
Automating Google Workspace (GWS) & more with Apps Script
 
Powerful Google developer tools for immediate impact! (2023-24 C)
Powerful Google developer tools for immediate impact! (2023-24 C)Powerful Google developer tools for immediate impact! (2023-24 C)
Powerful Google developer tools for immediate impact! (2023-24 C)
 
08448380779 Call Girls In Civil Lines Women Seeking Men
08448380779 Call Girls In Civil Lines Women Seeking Men08448380779 Call Girls In Civil Lines Women Seeking Men
08448380779 Call Girls In Civil Lines Women Seeking Men
 
Artificial Intelligence: Facts and Myths
Artificial Intelligence: Facts and MythsArtificial Intelligence: Facts and Myths
Artificial Intelligence: Facts and Myths
 
GenAI Risks & Security Meetup 01052024.pdf
GenAI Risks & Security Meetup 01052024.pdfGenAI Risks & Security Meetup 01052024.pdf
GenAI Risks & Security Meetup 01052024.pdf
 
08448380779 Call Girls In Diplomatic Enclave Women Seeking Men
08448380779 Call Girls In Diplomatic Enclave Women Seeking Men08448380779 Call Girls In Diplomatic Enclave Women Seeking Men
08448380779 Call Girls In Diplomatic Enclave Women Seeking Men
 
[2024]Digital Global Overview Report 2024 Meltwater.pdf
[2024]Digital Global Overview Report 2024 Meltwater.pdf[2024]Digital Global Overview Report 2024 Meltwater.pdf
[2024]Digital Global Overview Report 2024 Meltwater.pdf
 
How to Troubleshoot Apps for the Modern Connected Worker
How to Troubleshoot Apps for the Modern Connected WorkerHow to Troubleshoot Apps for the Modern Connected Worker
How to Troubleshoot Apps for the Modern Connected Worker
 
04-2024-HHUG-Sales-and-Marketing-Alignment.pptx
04-2024-HHUG-Sales-and-Marketing-Alignment.pptx04-2024-HHUG-Sales-and-Marketing-Alignment.pptx
04-2024-HHUG-Sales-and-Marketing-Alignment.pptx
 
The 7 Things I Know About Cyber Security After 25 Years | April 2024
The 7 Things I Know About Cyber Security After 25 Years | April 2024The 7 Things I Know About Cyber Security After 25 Years | April 2024
The 7 Things I Know About Cyber Security After 25 Years | April 2024
 
Boost Fertility New Invention Ups Success Rates.pdf
Boost Fertility New Invention Ups Success Rates.pdfBoost Fertility New Invention Ups Success Rates.pdf
Boost Fertility New Invention Ups Success Rates.pdf
 
presentation ICT roal in 21st century education
presentation ICT roal in 21st century educationpresentation ICT roal in 21st century education
presentation ICT roal in 21st century education
 
What Are The Drone Anti-jamming Systems Technology?
What Are The Drone Anti-jamming Systems Technology?What Are The Drone Anti-jamming Systems Technology?
What Are The Drone Anti-jamming Systems Technology?
 
A Domino Admins Adventures (Engage 2024)
A Domino Admins Adventures (Engage 2024)A Domino Admins Adventures (Engage 2024)
A Domino Admins Adventures (Engage 2024)
 

New insights into the selective adsorption mechanism of cationic and anionic dyes using MIL-101(Fe) metal-organic framework Modeling and interpretation of physicochemical parameters.pdf

  • 1. Journal of Contaminant Hydrology 247 (2022) 103977 Available online 14 February 2022 0169-7722/© 2022 Published by Elsevier B.V. New insights into the selective adsorption mechanism of cationic and anionic dyes using MIL-101(Fe) metal-organic framework: Modeling and interpretation of physicochemical parameters Mohamed Shakly a , Laila Saad b , Moaaz K. Seliem c,* , Adrián Bonilla-Petriciolet d , Nabila Shehata a a Environmental Science and Industrial Development Department, Faculty of Postgraduate Studies for Advanced Science (PSAS), Beni-Suef University, P.O. Box 62511, Beni-Suef, Egypt b Renewable Energy Science and Engineering Department, Faculty of Postgraduate Studies for Advanced Science (PSAS), Beni-Suef University, P.O. Box 62511, Beni- Suef, Egypt c Faculty of Earth Science, Beni-Suef University, 62511, Egypt d Instituto Tecnológico de Aguascalientes, Aguascalientes 20256, Mexico A R T I C L E I N F O Keywords: MIL-101(Fe) Methylene blue Methyl orange Adsorption mechanism Advanced modeling A B S T R A C T In the current study, iron-based metal organic framework (MOF) MIL-101(Fe) was successfully prepared via a facile solvothermal method. The as–synthesized MIL-101(Fe) was characterized by XRD, FE-SEM, FTIR, TGA and zeta potential techniques, and then employed as an adsorbent for methyl orange (MO) and methylene blue (MB) dyes. The adsorbed quantities of MO (1067 to 831 mg/g) were higher than those of MB (402 to 353 mg/g) indicating the high selectivity of MIL-101(Fe) towards the anionic dye at all temperatures (20–60 ◦ C). Adsorption processes of MO and MB followed the pseudo-second order kinetics and the Langmuir equilibrium model. The interaction mechanism at a molecular level was analyzed and deeply interpreted via the advanced multilayer adsorption model. Steric parameters indicated that MO molecular aggregation (n) was 0.95–1.33 thus signifying the presence of multi–docking and multi–interactions mechanisms. The aggregated number of MB was superior to unity (i.e., n = 1.17–1.78) suggesting a vertical adsorption position and a multi-interactions mechanism at all operating temperatures. The density of MIL-101(Fe) active sites (DM = 77.33–52.38 mg/g for MB and 149.91–107.07 for MO) and the total adsorbed dye layers (Nt = 3.12–2.49 for MB and 5.36–3.67 for MO) resulted in improving the adsorption capacities of MO dye. The adsorption energies ranged from 8.89 to 33.73 kJ/mol and they displayed that MO and MB uptake processes were exothermic controlled by physical interactions at all temperatures. Regeneration results indicated that this adsorbent can be reutilized without a significant loss in its removal efficiency after five adsorption-desorption cycles. Overall, the adsorption capacity, chemical stability, and regeneration performance of MIL-101(Fe) support its application as a very promising adsorbent for the removal of organic hazardous pollutants from water. 1. Introduction Pollution of water bodies due to industrial toxic-containing effluents is considered a worldwide problem (Homaeigohar, 2020). Commonly, significant quantities of dyed wastewater are generated from various industries including textile, leather, paper, printing, and dyestuff sec­ tors. According to previous studies, methylene blue (MB) and methyl orange (MO) dyes (the studied organic pollutants in this paper) are recognized as chemical compounds that can persist for long times in aqueous solutions due to their non-biodegradability properties. Besides, the improper management of wastewater-containing MO and MB dyes, which have carcinogenic and toxic properties, causes a major threat for human beings (Haque et al., 2010). Consequently, different methods as adsorption, coagulation, biological, advanced oxidation, or precipita­ tion have been utilized in cleaning the contaminated water. Based on the benefits and drawbacks associated to each treatment method, adsorp­ tion is recommended as low-cost, simple, and more effective technique (Homaeigohar, 2020). Especially, the adsorption processes using * Corresponding author. E-mail address: Moaaz.korany@science.bsu.edu.eg (M.K. Seliem). Contents lists available at ScienceDirect Journal of Contaminant Hydrology journal homepage: www.elsevier.com/locate/jconhyd https://doi.org/10.1016/j.jconhyd.2022.103977 Received 27 October 2021; Received in revised form 13 January 2022; Accepted 10 February 2022
  • 2. Journal of Contaminant Hydrology 247 (2022) 103977 2 nanoparticles–based adsorbents that have different active adsorption sites and great surface areas are very promising for water purification (Ahmadijokani et al., 2020). Metal-organic frameworks (MOFs), a new group of porous crystalline materials, consist of organic ligands and the desired metal clusters (Younas et al., 2020). Recently, MOFs have been employed in various applications as catalysis, storage of hydrogen, separation of gases, drug delivery, CO2 capture, sensing, and wastewater purification (Younas et al., 2020). The suitability of MOFs for different applications is due to the appropriate pore size, high surface area, simplicity of preparation, and the high stability of these porous coordination polymers (Mandal et al., 2018; Jhung et al., 2012; Cavka et al., 2008). Moreover, the simplicity for modification of MOFs structure can change their textural properties thus contributing to increase and widen the acceptability of these materials in numerous fields. Generally, MOFs are fabricated via bridging metal ion clusters with metal ions using organic linkers and the experimental parameters as reaction time, pressure, temperature, solu­ tion pH, and the used solvent are considered (Safaei et al., 2019). Defects in the preparation conditions (i.e., metal ion, organic ligand, metal- ligand coordination geometry, pore surface hydrophobicity) have contributed to decrease the stability of MOFs for industrial applications (Adegoke et al., 2020; Burtch et al., 2014; Wang et al., 2016; Yuan et al., 2018). A water–stable MOFs type, MIL-101 (MIL: Material Institute Lavoisier), was combined with different metal precursors as chromium (Cr) (MIL-101-Cr), aluminum (Al) (MIL-101-Al), and iron (Fe) (MIL-101- Fe) and finally used in water purification (Li et al., 2019a; Zhao et al., 2018). Different varieties of MOFs such as an amino-functionalized Zr-based MOFs (Lv et al., 2019), RGO/ MOFs (Liu et al., 2019), MIL-101-Cr, MIL- 53-Al and ZIF-8 (Zhang et al., 2020), Cu-BTC-1 (Mantasha et al., 2020), Cu-BTC (Kaur et al., 2019), MOF-235 (Haque et al., 2011), UiO-66 (Molavi et al., 2018), and Fe-MIL-101 (Konik et al., 2019) were used in the adsorption of MB and MO. In several adsorption systems, the experimental data modeling was conducted with the Langmuir and Freundlich equations. The fundaments and hypothesis of these classical models cannot provide a scientific meaning and reliable explanation for the influence of operating parameters as the adsorbate concentration and temperature in the removal mechanism (Ramadan et al., 2021; Sharib et al., 2021). To obtain a better understanding of the MO and MB removal mechanisms by MIL-101(Fe), it is mandatory to use the Fig. 1. Characterization results of MIL-101(Fe) adsorbent: (a) XRD pattern, (b, c) SEM images, and (d) FTIR spectrum. M. Shakly et al.
  • 3. Journal of Contaminant Hydrology 247 (2022) 103977 3 advanced adsorption models based on the statistical physics theory (Mohamed et al., 2020; Seliem et al., 2020). This theory is appropriate to provide additional steric and energetic parameters for the analysis and explanation of the adsorption mechanism. The interpretation of the physicochemical factors associated to MB and MO adsorption onto MIL- 101 (Fe) can contribute to clarify the adsorption mechanisms at the molecular level (Ramadan et al., 2021; Sharib et al., 2021; Mohamed et al., 2020). The main objectives of the current study were to (a) characterize MIL-101(Fe) metal organic framework prepared through a facile solvothermal method, (b) determine the removal performance of the prepared MIL-101(Fe) for MB and MO dyes in single and binary systems, (c) use different models (i.e., kinetics and isotherms) to describe the adsorption processes, (d) provide new results and a deeper understanding of the interactions between dyes molecules and MIL-101 (Fe) active sites via statistical physics calculations, and (e) evaluate the possibility of the multiple use of MIL-101(Fe) for the adsorption of cationic and anionic dyes. Overall, the results of this study contributed with new insights into MO and MB adsorption performance and mech­ anism using MIL-101(Fe) as an adsorbent. 2. Materials used and methods 2.1. Materials Ferric chloride hexahydrate (FeCl3.6H2O: Alpha Chemica India), terephthalic acid (Benzene-1,4-dicarboxylic acid BDC: Merck KGaA Germany) and N, N-dimethyl formamide (DMF: Techno Pharm chem India) were used in this study in the preparation of MIL-101(Fe) adsorbent. MO and MB dyes and absolute ethyl alcohol (99.9%) were supplied from Merck KGaA, Germany and the two dyes were tested as adsorbates. 2.2. Synthesis of MIL-101(Fe) MIL-101(Fe) was synthesized by a hydrothermal method following a similar procedure reported by (Hu et al., 2019). Ferric chloride hexa­ hydrate (FeCl3.6H2O) and Benzene-1,4-dicarboxylic acid was dissolved in DMF with 2:1 ratio, the suspended solution was sonicated for 20 min until was fully homogenous, and then it was transferred into a Teflon- lined stainless-steel autoclave and heated at 110 ◦ C for 20 h. The light orange product was separated via centrifugation and purified by washing with DMF for 3 h followed by hot ethanol for 2 h. This step was repeated for three times in order to remove the unreacted BDC. The purified product was dried at 70 ◦ C for 30 min, and then activated at 150 ◦ C for 10 h before carrying out the dye adsorption studies. 2.3. Characterization of MIL-101 (Fe) MIL-101 (Fe) was characterized using different techniques to inves­ tigate its physical and chemical properties. X–ray diffraction (XRD) pattern of this adsorbent was recorded in the 2θ range from 5.02–79.98◦ using a Panalytical Philips diffractometer operated at 40 kV and 35 mA under Cu–kα radiation (λ = 0.154 nm). The surface morphology of MIL- 101 (Fe) was observed using (Zeiss Sigma 500 VP analytical FE-SEM). The samples were coated with gold before analysis to enhance the sur­ face conductivity. The functional groups of MOF surface were investi­ gated with FTIR spectrum (Bruker optics -vertex 70 equipment) at constant ambient temperature of 25 ◦ C by accumulating 10 scans at 1 cm− 1 resolution in the 4000–400 cm− 1 region. The Iso-ionic point and the particles size range of MIL-101 (Fe) were obtained from Zeta po­ tential and sizer (Malevern Zeta sizer-nano series ZS 90). The ther­ mogravimetric analysis (TGA) and differential scanning calorimetry (DSC) were performed to analyze the thermal stability of MIL-10(Fe) in the temperature range of 25–1200 ◦ C with a heating rate of 10 ◦ C/min (Labsys evo Setaram, France). 2.4. Adsorption experiments of MB and MO dyes Stock solutions (1000 mg/L of MB and 2000 mg of MO) were pre­ pared, and the desired initial concentrations of the tested dyes were obtained using dilutions with high purity distilled water. To avoid any possible reactions associated to the photo catalytic process, all adsorp­ tion experiments were conducted in dark glasses in the absence of light. After each adsorption experiment, the solutions containing MO and MB dyes were separated from MIL-101 (Fe) adsorbent using syringe filters. The concentrations of the studied dyes in the filtrated solutions were quantified using a UV double beam spectrophotometer at 664 and 464 nm for MB and MO dyes, respectively. All MO and MB removal experi­ ments were performed in triplicates and the results were averaged for Fig. 2. TGA (a) and DSC (b) analysis of MIL-101(Fe). Fig. 3. Effect of pH on the removal of MO and MB dyes by MIL-101(Fe) adsorbent. The error bars were obtained from triplicate experiments. M. Shakly et al.
  • 4. Journal of Contaminant Hydrology 247 (2022) 103977 4 data evaluation obtaining standard deviations below 5.2%. 2.5. MO and MB adsorption kinetics To study the adsorption kinetics, the experiments were performed at the following conditions: 0.15 g of MIL-101 (Fe) mass, pH 4.0 for MO and 9.5 for MB, initial concentrations of 30 mg/L for MB and 100 mg/L for MO, adsorption temperature of 20 ◦ C, and different contact times ranging from 15 min to 14 h. The adsorbed amounts (qt) of MB and MO after each time interval were calculated using Eq. (1). qt (mg/g) = (C0–Ct) V m (1) Where C0 (mg/L) is the initial concentrations of MO and MB, Ct (mg/ Fig. 4. The adsorption kinetics of MB and MO onto MIL-101(Fe) adsorbent at 20 ◦ C. (a, b) contact time effect, (c, d) Pseudo-First Order, (e, f) Pseudo-Second Order and (g, h) intra-particle diffusion. M. Shakly et al.
  • 5. Journal of Contaminant Hydrology 247 (2022) 103977 5 L) is the final concentration of the tested dyes after time (t), V (L) is the solution volume, and m (g) is the mass of MIL-101 (Fe) adsorbent (g). To study the MO and MB adsorption onto MIL-101(Fe) and to identify the possible rate–controlling step, the kinetic results were fitted to the pseudo-first order (Lagergren, 1898), the pseudo-second order (Ho and McKay, 1999), and intra-particle diffusion (Weber and Morris, 1962) models as given below: ln(qe − qt) = lnqe − k1t Pseudo − first − order (2) t / qt = ( 1 / k2q2 e ) + t / qe Pseudo − second − order (3) qt = kp t1/2 + C Intra − particle diffusion (4) where k1(min− 1 ) and k2(g/mg min) indicate the adsorption rate constants of the first order and the second-order kinetics, respectively. For the intra-particle diffusion equation, kp(mg/g min) and C (mg/g) represent the rate constant and the intercept value of this model, respectively. 2.6. MO and MB adsorption isotherms Equilibrium studies related to MO and MB adsorption onto MIL-101 (Fe) were performed at the next conditions: Initial dye concentrations (25–500 mg/L for MB and 25–2000 mg/L), three temperatures (20, 40 and 60 ◦ C), 0.01 g of MIL-101 (Fe), and solution pH 4 for MO and 9.5 for MB. In all these experiments, 10 mg of the tested adsorbent was mixed with 30 mL of each dye solution at 300 rpm using digital orbital shaker (SHO–2D, Germany). The liquid phases containing MO and MB were Table 1 Parameters of Kinetic Models for the Adsorption of MO and MB onto MIL101-Fe. Co (mg/L) K1 (min 1 ) qe exp (mg/g) qe cal (mg/g) R2 Dye Pseudo-first order MO 100 0.316 80.238 0.909 MB 28 0.073 2.5 0.903 Dye Pseudo-second order MO 100 0.01 104 108.7 0.993 MB 28 4.538 27.1 27.1 1 Fig. 5. Modeling of adsorption isotherms of MB and MO dyes on MIL-101 (Fe) MOF using Langmuir and Freundlich equations. M. Shakly et al.
  • 6. Journal of Contaminant Hydrology 247 (2022) 103977 6 separated and the concentrations of these dyes in solutions were determined. The equilibrium adsorption quantities of MO and MB dyes (qe) were calculated using Eq. (1) assuming that t corresponded to the equilibrium time. 2.6.1. Classical modeling analysis The experimental data of MB and MO were fitted to the Langmuir (1916) and Freundlich (Freundlich, 1906) equations to find the suitable classical model that can describe the adsorption of dyes onto MIL-101 (Fe) at equilibrium. The non-linear relations defined the Langmuir (Eq. (5)) and Freundlich (Eq. (6)) isotherm models are given below: qe = qmax KLCe (1 + KLCe ) (Langmuir model) (5) qe = KF Ce 1/n (Freundlich model) (6) where qmax (mg/g) is the maximum adsorption capacity and KL (L/ mg) represents the constant of the Langmuir model. On the other hand, KF((mg/g)(mg/L)− 1/n )) signifies the adsorption capacity and n is the heterogeneity factor of the Freundlich model. 2.6.2. Advanced modeling analysis Herein, three advanced statistical physics models (i.e., single layer, double layer, and multilayer) were employed to investigate the dye– MIL-101 (Fe) interaction (Ramadan et al., 2021; Sharib et al., 2021). • Firstly, the single layer model assumes that MO and MB removal by MIL-101 (Fe) was in the form of one layer directed by a distinctive adsorption energy (ΔE) (i.e., the adsorption of each tested dye mol­ ecules via MIL-101 (Fe) active sites is nearly governed by a constant energy). The mathematical expression of this advanced single layer adsorption model is given by Eq. (7) (Sharib et al., 2021). qe = nDM 1 + ( c1/2 c )n (7) where n is the number of MO and MB dye molecules captured by active site of MIL-101(Fe) adsorbent, C1/2 signifies the concentration at half- saturation related to the formed adsorbate layer, and DM is the density of adsorption sites. • Secondly, the double layer model suggests that these adsorption systems are characterized by the formation of two layers of dyes molecules (MB or MO) with two dissimilar adsorption energies (i.e., ΔE1 for dye– MIL-101 (Fe) interaction and ΔE2 for MB–MB or MO–MO interaction). The mathematical expression of the double layer model is denoted by Eq. (8) (Sharib et al., 2021). qe = nDM ( c c1 )n + 2 ( c c2 )2n 1 + ( c c1 )n + ( c c2 )2n (8) where C1 and C2 describe the two concentrations at half saturation attributed to the first and second removed dye layers, respectively. • Finally, the multilayer model suggests that the MO and MB adsorp­ tion onto MIL-101(Fe) results in the formation of an adaptable but limited number of adsorbed dye (i.e., MO or MB) layers. Conse­ quently, the total adsorbed layers number (1 + N2) of each tested dye is related to a single layer (N2 = 0), double layer (N2 = 1), and multilayer process (N2 > 1) (Mohamed et al., 2020). Similar to the double layer model, two energies (ΔE1 and ΔE2) were associated to this multilayer model (i.e., dye–MIL-101(Fe) and dye-dye Table 2 Parameters of isotherm equations for the MB adsorption on MIL-101 (Fe) MOF. Isotherm Model 20 ◦ C 40 ◦ C 60 ◦ C Langmuir qmax (mg/g) 402.00 391.67 352.96 kL (L/mg) 0.0365 0.0227 0.0165 R2 0.9613 0.9618 0.9660 Freundlich kF ((mg/g)(mg/L)− 1/n ) 75.51 64.76 49.26 1/n 0.2881 0.3617 0.3283 R2 0.9408 0.9443 0.9554 Multiple layer N2 3.12 3.08 2.49 Qsat (mg/g) 373.74 348.07 325.08 R2 0.9742 0.9962 0.9844 Table 3 Parameters of isotherm equations for the MO adsorption on MIL-101 (Fe) MOF. Isotherm Model 20 ◦ C 40 ◦ C 60 ◦ C Langmuir qmax (mg/g) 1066.96 934.17 830.97 kL (L mg ) 0.0077 0.0071 0.0068 R2 0.9449 0.9447 0.9398 Freundlich kF ((mg/g)(mg/L)− 1/n ) 67.60 53.36 45.27 1/n 0.4019 0.4141 0.4190 R2 0.8834 0.8700 0.8537 Multiple layer N2 5.36 5.01 3.76 Qsat (mg/g) 908.39 777.01 677.92 R2 0.9631 0.9655 0.9632 Table 4 Comparison of the adsorption capacities of several Metal Organic Frameworks for the removal of MB & MO dyes. Adsorbent Dye adsorbate Adsorption capacity (mg/ g) Optimum pH T (◦ C) Ref Amine- MOF-Fe MB 312 9 20 (Paiman et al., 2020) UiO-66 MB 90 5.5 25 (Chen et al., 2015) NH2-UiO- 66 MB 96 5.5 25 (Luo et al., 2017) ZIF-67 MB 57.14 10 – (Haque et al., 2010) MIL-101 (Fe) MB 149 9 20 (Paiman et al., 2020) MIL-101 (Fe) MB 402 9.5 20 This study MIL-101 MO 140 3.5 45 (Haque et al., 2010) UiO-66 MO 39 5.5 25 (Chen et al., 2015) NH2-UiO- 66 MO 28 5.5 25 (Chen et al., 2015) Cu-BDC MO 86.7 4 25 (Salama et al., 2018) ZIF-67 MO 75.5 4 – (Luo et al., 2017) MIL-101 (Fe) MO 1067 4 20 This study M. Shakly et al.
  • 7. Journal of Contaminant Hydrology 247 (2022) 103977 7 interactions). Generally, the adsorption energy (ΔE1) was related to the interaction between the first adsorbed dye layer (unchanging number and equal to unity) and MIL-101(Fe) adsorption sites, while the other energy (ΔE2) was linked to the dye–dye interaction and thus, ΔE1 is found to be greater than ΔE2. The multilayer adsorption model is given by Mohamed et al. (2020); Li et al. (2019b): qe = n DM F1(c) + F2(c) + F3(c) + F4(c) G(c) (9) F1(c) = − 2 ( c c1 )2n 1 − ( c c1 )n + ( c c1 )n ( 1 − ( c c1 )2n ) ( 1 − ( c c1 )n )2 , (10) F2(c) = 2 ( c c1 )n( c c2 )n ( 1 − ( c c2 )n N2 ) 1 − ( c c2 )n , (11) F3(c) = − N2 ( c c1 )n( c c2 )n( c c2 )n N2 1 − ( c c2 )n , (12) F4(c) = ( c c1 )n( c c2 )2n ( 1 − ( c c2 )n N2 ) ( 1 − ( c c2 )n )2 , (13) Fig. 6. Modeling of adsorption isotherms of MO and MB dyes on MIL-101 (Fe) MOF using a multilayer statistical physics model. M. Shakly et al.
  • 8. Journal of Contaminant Hydrology 247 (2022) 103977 8 G(c) = ( 1 − ( c c1 )2n ) 1 − ( c c1 )n + ( c c1 )n( c c2 )n ( 1 − ( c c2 )n N2 ) ( 1 − ( c c2 )n )2 , (14) where N2 is the layers of removed dye molecules, C1 is the half saturation concentration connected to the first layer, and C2 is the corresponding one correlated to N2 adsorbed dye layers. 2.7. Regeneration of MIL-101(Fe) adsorbent Recycling and utilization of the MIL-101(Fe) adsorbent were inves­ tigated via the solvent method using ethanol as an eluent. The MIL-101 (Fe) loaded with the tested dyes (MB and MO) was continuously agitated on a rotatory shaker at 200 rpm for 4 h at 25 ◦ C to achieve the complete removal of the adsorbed dyes. This adsorption/desorption test was repeated five times under the same conditions. At the end of each desorption round, the MIL-101(Fe) was washed several times by distilled water and dried at 65 ◦ C for 12 h before the next desorption cycle. 3. Results and discussion 3.1. Characterization of MIL-101 (Fe) adsorbent X-ray diffraction (XRD) was carried out to identify the crystallinity and structure of the synthesized MIL-101 (Fe) adsorbent. Different peaks related to the MIL-101 (Fe) were observed at 2 thetas = 5.3◦ , 8.82◦ , 9.18◦ and 18.5◦ , see Fig. 1a. These detected peaks were comparable with the previous studies (Hu et al., 2019; Wang et al., 2018; Liu et al., 2018a; Li et al., 2016). FE-SEM images (Fig. 1b and c) display the morphological features of the MIL-101 (Fe) surface. It was observed that the MIL-101(Fe) had a special octahedral structure with diameters from 0.7 to 2 μm, which was consistent to the other reported studies (Hu et al., 2019; Wang et al., 2018). FTIR spectrum (Fig. 1 d) shows a strong band at 1574 cm− 1 that indicated (C=O) of the carboxylate group (Hu et al., 2019). The broad band detected at 3376 cm− 1 corresponded to the hydroxyl group of the adsorbed water molecules (Li et al., 2019a). The strong band at 1392 cm− 1 was attributed to the vibration of (C–C) in benzene ring; however, the parent bands at 545 and 748 cm− 1 were related to (Fe-OH) vibration MB dye MO dye n DM (mg/g) Qsat (mg/g) 20 °C 40 °C 60 °C 20 °C 40 °C 60 °C 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 0 10 20 30 40 50 60 70 80 90 0 20 40 60 80 100 120 140 160 0 50 100 150 200 250 300 350 400 0 100 200 300 400 500 600 700 800 900 1,000 Fig. 7. Statistical physics parameters for the MB and MO adsorption on MIL-101 (Fe) MOF. M. Shakly et al.
  • 9. Journal of Contaminant Hydrology 247 (2022) 103977 9 (Wang et al., 2018). These FTIR results also agreed to the earlier studies (Li et al., 2019a; Hu et al., 2019; Wang et al., 2018) and thus confirmed the preparation of MIL-101(Fe). Particles size range of MIL-101(Fe) was obtained using zetasizer measurement and results indicated that it ranged from 470 to 603 nm. Also, the surface charge value of the prepared MIL-101(Fe) was − 3 at pH 7.5. TGA analysis displayed the presence of two main weight losses within the temperature range of 25–650 ◦ C (Fig. 2a). The first weight loss (about 16%) below 400 ◦ C could be related to the elimination of water molecules and free terephthalates in the pores of MIL-101(Fe). In addition, the significant second weight loss (38%) from 400 to 650 ◦ C could be due to the decomposition of organic ligands that result in the collapse of MIL-101(Fe). TGA result of this study is similar to that of other UIO-66 and MIL-101(Cr) metal organic frameworks (Hu et al., 2015; Yang and Yan, 2011). Furthermore, the DCS result displays the existence of two sharp endothermic peaks (Fig. 2b). The first endo­ thermic sharp peak observed at 250–300 ◦ C could be attributed to the evaporation of H2O molecules and free terephthalate in pores of the MIL- 101(Fe), while the second one at 600–700 ◦ C could be associated to the decomposition of MIL-101(Fe). Consequently, TGA and DSC results indicated that MIL-101(Fe) is characterized by high thermal stability up to 650 ◦ C. 3.2. pH effect on the adsorption of MB and MO onto MIL-101(Fe) Adsorption experiments were conducted at a wide range of solution pH (i.e., pH 2–12) to find the optimum value of this parameter for the removal of MB and MO (100 mg/L) using 0.01 g of MIL-101(Fe) at 20 ◦ C. Fig. 3 displays that the removal efficiency for MB increased with improving pH value where the highest efficiency was achieved at pH 9.5. This behavior for MB adsorption was related to the electrostatic attraction between the positively charged MB molecule and negatively charged MIL-101(Fe) surface, whose pH at zero point of charge was pHPZC = 7.1. Thus, the surface of MIL-101 (Fe) was negatively charged at pH > 7.1 and it was more effective to attractive and remove the cationic MB dye. On the contrary case, the maximum percentage of MO adsorption was attained at pH 4 due to the electrostatic attraction force caused by the negatively charged anionic MO dye. This indicated that the MIL-101(Fe) was found to be an excellent adsorbent for the tested dyes and thus, all adsorption experiments were conducted at pH 9.5 and pH 4 for MB and MO, respectively. 3.3. Contact time effect on MB and MO adsorption using MIL-101(Fe) Very rapid adsorption of both MB and MO dyes were observed in the first 15 min with removal percentages of 52.7 and 76% for MO and MB respectively, see Fig. 4a and b. It was assumed that the adsorption process of MB and MO molecules was very fast at the beginning due to the availability of numerous active sites on the MIL-101(Fe) surface. After that, the interaction time between dyes and MIL-101(Fe) started to be slower (i.e., a gentle slope was observed) due to the limitation of the available adsorption sites. Finally, the removed amounts of both dyes were nearly constant after contact times of 2 h for MB and 10 h for MO, which indicated that equilibrium was achieved after the saturation of active sites of the MIL-101(Fe) surface. 3.4. MB and MO adsorption kinetics As indicated, the kinetics of dyes adsorption were investigated using different models as the Pseudo-first order (Lagergren, 1898), Pseudo -second order (Ho and McKay, 1999), and intra-particle diffusion models (Weber and Morris, 1962). The Pseudo-first order considers that the adsorption process is strongly depends on the diffusion, while the Pseudo-second order assumes that a chemical adsorption can be involved in the removal process (Ahmadijokani et al., 2020). The experimental data of MO and MB were fitted to these different kinetic models and the best fit model was identified via the R2 value, see Fig. 4. All the kinetics parameters were calculated and compared with the experimental data as shown in Table 1. The values of R2 for MO adsorption on MIL-101(Fe) indicating that the best fit model was the Pseudo-Second-order equation (R2 = 0.993) in comparison to the Pseudo-First order equation (R2 = 0.909). Also, the best fitted model (R2 = 1) for MB was the Pseudo-Second order kinetics equation followed by the Pseudo-first order (R2 = 0.903). These results indicated that the adsorption kinetics of both MB and MO on MIL-101 (Fe) were described by the Pseudo-Second-order model and thus, the adsorption of MO and MB could be related to chemisorption process. The intra-particle diffu­ sion analysis attributed to each dye molecule displayed dissimilar linear stages over the whole-time range (Fig. 4). The initial sharp stage can be associated to the external mass transfer of MB and MO molecules from the solution to the outer surface of the MIL-101(Fe) adsorbent. The second and third linear sections characterized the control of pore- diffusion and equilibrium stages, respectively. Consequently, the adsorption of MB and MO was mostly directed by the film diffusion during the first stage, followed by pore-diffusion and equilibrium at subsequent stages. 3.5. Classical modeling of MB and MO adsorption isotherms Modeling of MB and MO adsorption data at equilibrium with the applied traditional models (Langmuir and Freundlich equations) are displayed in Fig. 5 and the corresponding adjusted parameters of these models are listed in Tables 2 and 3. Langmuir model was the best alternative to fit MB and MO adsorption data where their R2 values were highest at all tested temperatures. Therefore, the removal of these dyes molecules was related to homogenous adsorption sites of MIL-101(Fe) and the adsorbates–adsorbent interaction resulted in the formation of a single dye layer. The maximum adsorption capacities (qmax) for MB E (kJ/mol) E (kJ/mol) 20 °C 40 °C 60 °C 0 5 10 15 20 25 30 35 40 E1 E2 0 2 4 6 8 10 12 14 16 18 20 E1 E2 MB MO Fig. 8. Adsorption energies for MB and MO adsorption on MOF MIL-101 (Fe) MOF. M. Shakly et al.
  • 10. Journal of Contaminant Hydrology 247 (2022) 103977 10 were 402, 391.67, and 352.96 mg/g and 1066.96, 934.17, and 830.97 mg/g for MO at 20, 40, and 60 ◦ C, as given in Table 2. The decrease of qmax values with the increment of solution temperature suggested that the adsorption of MB and MO molecules by MIL-101(Fe) was exothermic (i.e., the dyes–adsorbent interaction was more energetic at low tem­ perature) (Sharib et al., 2021; Mohamed et al., 2020). Moreover, the adsorption capacities associated to MO molecules were higher than those of MB dye, which indicated the selectivity of the as–synthesized Fig. 9. Adsorption selectivity of MIL-101(Fe) for the adsorption of MB and MO dyes in binary systems. Fig. 10. MB and MO removal after the regeneration of MIL-101 (Fe) adsorbent. M. Shakly et al.
  • 11. Journal of Contaminant Hydrology 247 (2022) 103977 11 MIL-101(Fe) active sites for this azo dye. This variation in the removal efficiency of MIL-101(Fe) could be related to the chemical structures (i. e., C14H14N3NaO3S for MO and C16H18ClN3S for MB) and the molecular size (i.e., 1.2 nm for MO and 1.38 nm for MB) of these dye molecules (Seliem et al., 2020). The molecular characteristics of MO and MB played a relevant role for transfer phenomena in the interface between these dyes and MIL-101(Fe) adsorbent surface. Furthermore, the decrease of KF values from 75.51 to 49.26 for MB and from 67.60 to 45.27 (mg/g)(mg/L)− 1/n with temperature also supported the exothermic adsorption of these dyes. Calculated 1/n values by the Freundlich equation were lower than unity thus signifying an effective dye–MIL-101(Fe) interaction even at low concentrations of these organic pollutants (Sharib et al., 2021; Mohamed et al., 2020). The maximum adsorption capacities for MO and MB by the synthetic MIL- 101(Fe) and other MOFs are listed in Table 4. It can be observed that the qmax values of MIL-101(Fe) were higher than those reported for different MOFs. Therefore, it can be concluded that the MIL-101(Fe) is a promising adsorbent for removal of dyes-bearing solutions. 3.6. Advanced statistical modeling of MB and MO adsorption isotherms The multilayer model (Fig. 6) presented the highest R2 values compared to the single- and double-layer models at all solution tem­ peratures (i.e., 20, 40, and 60 ◦ C). Therefore, the steric (e.g., n, DM, and Qsat) and energetic (ΔE) parameters of this multilayer adsorption model were deeply analyzed to obtain a better understanding of the adsorption mechanism of MB and MO dyes on MIL-101(Fe). 3.6.1. The n parameter The determination of steric n parameter provides information asso­ ciated to the removal mechanism of the tested adsorbates (MB and MO dyes) on MIL-101(Fe) especially in terms of adsorption orientation and molecular aggregation phenomenon. In particular, the adsorption ge­ ometry (vertical or horizontal) of the removed MB and MO molecules on the adsorbent can be analyzed using this parameter (Ramadan et al., 2021; Mohamed et al., 2020; Barakat et al., 2020). Based on the value of n parameter, two main scenarios can be recognized to describe the ge­ ometry and mechanism of dyes adsorption onto MIL-101(Fe) surfaces (Sharib et al., 2021; Barakat et al., 2020). • The first scenario (n > 1): The adsorption of MO and MB molecules can take place in a vertical position through multi–interactions mechanism (i.e., one active site of MIL-101(Fe) can capture several dye molecules). • The second scenario (n < 1): MO and MB adsorption occurs in a horizontal location via a multi–docking mechanism (i.e., numerous MIL-101(Fe) functional groups can remove one dye molecule). Fig. 7 exhibits the values of n parameter related to MB and MO adsorption on MIL-101(Fe) as a function of solution temperature. For MO dye, the n parameter values were 0.95, 1.02, and 1.33 at 20, 40, and 60 ◦ C, respectively. As a result, horizontal position and multi–docking mechanism could be expected for MO adsorption at 20 ◦ C, while vertical position and multi–interactions mechanism could occur at 40 and 60 ◦ C. The n values were higher than unity at 40 and 60 ◦ C and this result indicated the effect of aggregation phenomenon with increasing tem­ perature. On the other hand, the n values associated to MB adsorption were 1.17, 1.68, and 1.78 and thus, vertical orientation and multi­ –interactions mechanism were involved in the removal of this dye by the MIL-101(Fe) active sites at all temperatures. Consequently, the aggre­ gation of MB molecules (i.e., MB-MB binding) in solution was identified at all tested temperatures and this phenomenon resulted in the existence of the non-parallel adsorption geometry and a multi–molecular mech­ anism. The high degree of MB and MO aggregations with temperature suggested that these adsorption processes were energetically activated in solutions before the interaction between dyes and MIL-101(Fe) surface (Li et al., 2020). 3.6.2. The DM parameter Fig. 7 displays the number of MIL-101(Fe) active sites (i.e., DM parameter) with respect the adsorption temperature. From 20 to 60 ◦ C, the DM parameter decreased from 77.33 to 52.38 mg/g for MB and from 149.91 to 107.07 mg/g for MO dye. The decrease of DM value with temperature displayed a reverse trend as compared to the steric n pa­ rameters (i.e., the increment of n parameter caused a reduction of the occupied active sites number of the MIL-101(Fe) surface). Thus, the trend of DM parameter assessed the accumulation of the removed dyes molecules at high temperature. Moreover, the aggregated MO and MB molecules could be more selective to definite functional groups of the as- synthesized MIL-101(Fe) (Mobarak et al., 2019). Also, it can be noticed that the DM values associated to MO dye were higher than those of MB dye at all temperatures, which could be considered as an important variable of the enhanced adsorption capacity of MIL-101(Fe) for MO dye. 3.6.3. Total number of the adsorbed MB and MO layers (Nt = 1 + N2) To understand the adsorption mechanism of MB and MO molecules on MIL-101(Fe), the determination of the total layers of adsorbed dyes is necessary (Barakat et al., 2020). The calculated Nt values were 3.12, 3.08, and 2.49 for MB and 5.36, 5.03, and 3.76 for MO at 20, 40, and 60 ◦ C, respectively, see Fig. 7. The reduction of Nt value with temper­ ature increments could be related to the influence of thermal agitation, which resulted in a disordered movement of the adsorbed MB and MO molecules. The disturbed state associated to dye-dye interaction caused a decrease of the adsorbed dye layers (Seliem et al., 2020; Barakat et al., 2020). Like the DM parameter, the Nt values associated to MO dye were higher than those of MB dye at all adsorption temperatures (i.e., the N2 parameter presented a significant impact to improve the MO adsorption capacity). 3.6.4. The adsorbed MB and MO quantities at saturation (Qsat = n. DM. Nt) To estimate the maximum MIL-101(Fe) performance for the adsorption of MO and MB dyes, the calculation of Qsat values was done. Fig. 7 exhibits the trend of Qsat versus temperature, and the values of this parameter were 373.74, 348.07, and 325.08 mg/g for MB and 908.39, 777.01, and 677.92 mg/g for MO at 20, 40, and 60 ◦ C, respectively. These Qsat values confirmed the exothermic interactions associated to MB and MO adsorption on MIL-101(Fe) surface. For the investigated dyes, it was identified that the DM and Nt parameters preserved the same trend of the Qsat parameter with respect to improving temperature. According to the calculated steric parameters, it can be concluded that the removal efficiency of MIL-101(Fe) for MO and MB dyes was controlled by the DM and Nt parameters. Furthermore, the high values of the Qsat for MO as compared to that of MB confirmed the important roles of DM and Nt parameters to determine the preference of MIL-101(Fe) for the adsorption of this azo dye. 3.7. Energetic parameters To obtain an appropriate interpretation for the interactions between MB and MO molecules and MIL-101(Fe) active sites, the adsorption energies were calculated (Seliem et al., 2020; Barakat et al., 2020) These adsorption energies (ΔE) were estimated as follows (Barakat et al., 2020). C1 = Cse− ΔE1 RT (15) C2 = Cse− ΔE2 RT (16) where c1and c2 are the half-saturation concentrations and cs is the sol­ ubility of the tested MB and MO adsorbates. Fig. 8 shows the ΔE values as a function of the adsorption M. Shakly et al.
  • 12. Journal of Contaminant Hydrology 247 (2022) 103977 12 temperature. The negative ΔE values confirmed the exothermic in­ teractions between the dyes molecules and the MIL-101(Fe) adsorbent. This performance agreed with the influence of solution temperature to reduce the MO and MB adsorption capacities. Besides, the ΔE values of MO and MB adsorption were < 40 kJ/mol at 20, 40, and 60 ◦ C, which suggested physical adsorption processes (Sharib et al., 2021; Barakat et al., 2020). ΔE1 was associated to MIL-101(Fe)–dye (MB or MO) interaction, while ΔE2 described the dye–dye interface. Therefore, ΔE1 values were higher than ΔE2 all tested temperatures. Although the ΔE values associated to MB adsorption were higher than those of MO at all temperature, the adsorbed MB dye quantities were low and, conse­ quently, the selectivity of MIL-101(Fe) was not governed by the ΔE parameter. 3.8. Selectivity of MIL-101 (Fe) for the adsorption of MB and MO dyes in binary systems The adsorption selectivity is an important parameter for studying the behavior of an adsorbent for the removal of water pollutants (Liu et al., 2018b). For instance, graphene like metal organic framework (BUC-17) was found to be more selective to anionic dyes from an organic dye mixture (Li et al., 2017). Moreover, Eu-based MOF (BUC-88) displayed high selectivity to many pharmaceuticals and personal care products (Wang et al., 2021). In this study, the adsorption selectivity of the MIL- 101(Fe) for the dye removal was investigated using a mixture of MB and MO dyes at different pH values, keeping the other operating parameters constant. Fig. 9 shows that MIL-101 (Fe) surface was selective to MO and MB dyes at pH 4 and 9.5, respectively. Besides, the adsorption capacities of the tested dyes were slightly decreased in the binary system as compared to the single-component one at optimum pH values due to the competition between MO and MB molecules to interact with MIL-101 (Fe) active sites (i.e., weak antagonism interaction). At pH 7, both dyes were slightly adsorbed without a significant preference for any dye molecule, and the adsorption processes could be associated to the π- π stacking effect. These results suggested that the MIL-101 (Fe) MOF could be a more selective adsorbent for anionic and cationic dyes depending on the solution pH value. 3.9. Reusability study Fig. 10 shows the removal efficiency of MIL-101(Fe) after five regeneration cycles for both MB and MO dyes. It is clear that MIL-101 (Fe) presented more than 95% for MB removal and 90% for MO removal after all regeneration rounds. For MB dye, increment of the removal efficiency in the last cycles (i.e., from cycle 3 to 5) could be attributed to the high surface area of MIL-101(Fe)-MB interaction, which required frequent washing and drying under vacuum to achieve the complete desorption of the attached MB molecules (Paiman et al., 2020). Based on the regeneration results, the investigated adsorbent can be reutilized numerous times to remove the tested dyes without a sig­ nificant loss of its removal efficiency, thus suggesting an economic benefit and high stability of this adsorbent in the purification of dyes- bearing water. 4. Conclusions MIL-101(Fe) (MOF) was successfully prepared through a facile sol­ vothermal method, which was characterized and employed as an adsorbent for MO and MB at different experimental conditions. In single and binary adsorption systems, MIL-101(Fe) was more selective to MO dye as compared to MB dye. The adsorption of both MO and MB fol­ lowed the pseudo-second order and the Langmuir equations. A multi­ layer model from statistical physics theory was utilized to understand the MO and MB adsorption mechanisms at a molecular level. MO adsorption was governed by multi–docking and multi–interactions mechanisms, while MB adsorption was controlled only by multi–interactions mechanism. The density of MIL-101(Fe) active sites and the total adsorbed dye layers formed on MOF surface were the main parameters that determined the adsorption capacity of MO dye. The adsorption of MB and MO molecules by MIL-101(Fe) was mainly caused by physical interactions at all solution temperatures. The reusability study demonstrated that the MIL-101(Fe) can maintain an outstanding performance after four adsorption/desorption cycles. The current study clearly proved that the MIL-101(Fe) can be utilized as an adsorbent for efficient removal of anionic and cationic dyes from polluted solutions. Declaration of Competing Interest The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper. References Adegoke, K.A., Agboola, O.S., Ogunmodede, J., Araoye, A.O., Bello, O.S., 2020. Metal- organic frameworks as adsorbents for sequestering organic pollutants from wastewater. Mater. Chem. Phys. 253, 123246. Ahmadijokani, F., Mohammadkhani, R., Ahmadipouya, S., Shokrgozar, A., Rezakazemi, M., Molavi, H., Arjmand, M., 2020. Superior chemical stability of UiO- 66 metal-organic frameworks (MOFs) for selective dye adsorption. Chem. Eng. J. 399, 125346. Barakat, M.A., Selim, A.Q., Mobarak, M., Kumar, R., Anastopoulos, I., Giannakoudakis, D.A., Bonilla-Petriciolet, A., Mohamed, E.A., Seliem, M.K., Komarneni, S., 2020. Experimental and theoretical studies of methyl orange uptake by Mn–rich synthetic mica: Insights into manganese role in adsorption and selectivity. Nanomaterials 10, 1–19. Burtch, N.C., Jasuja, H., Walton, K.S., 2014. Water stability and adsorption in metal–organic frameworks. Chem. Rev. 114 (20), 10575–10612. Cavka, J.H., Jakobsen, S., Olsbye, U., Guillou, N., Lamberti, C., Bordiga, S., Lillerud, K.P., 2008. A new zirconium inorganic building brick forming metal organic frameworks with exceptional stability. J. Am. Chem. Soc. 130 (42), 13850–13851. Chen, Q., He, Q., Lv, M., Xu, Y., Yang, H., Liu, X., Wei, F., 2015. Selective adsorption of cationic dyes by UiO-66-NH2. Appl. Surf. Sci. 327, 77–85. Freundlich, H.M.F., 1906. Over the adsorption in solution. J. Phys. Chem. 57 (385471), 1100–1107. Haque, E., Lee, J.E., Jang, I.T., Hwang, Y.K., Chang, J.S., Jegal, J., Jhung, S.H., 2010. Adsorptive removal of methyl orange from aqueous solution with metal-organic frameworks, porous chromium-benzenedicarboxylates. J. Hazard. Mater. 181 (1–3), 535–542. Haque, E., Jun, J.W., Jhung, S.H., 2011. Adsorptive removal of methyl orange and methylene blue from aqueous solution with a metal-organic framework material, iron terephthalate (MOF-235). J. Hazard. Mater. 185 (1), 507–511. Ho, Y.S., McKay, G., 1999. Pseudo-second order model for sorption processes. Process Biochem. 34 (5), 451–465. Homaeigohar, S., 2020. The nanosized dye adsorbents for water treatment. Nanomaterials 10 (2), 295. Hu, Z., Khurana, M., Seah, Y.H., Zhang, M., Guo, Z., Zhao, D., 2015. Ionized Zr-MOFs for highly efficient post-combustion CO2 capture. Chem. Eng. Sci. 124, 61–69. Hu, H., Zhang, H., Chen, Y., Ou, H., 2019. Enhanced photocatalysis using metal–organic framework MIL-101 (Fe) for organophosphate degradation in water. Environ. Sci. Pollut. Res. 26 (24), 24720–24732. Jhung, S.H., Khan, N.A., Hasan, Z., 2012. Analogous porous metal–organic frameworks: synthesis, stability and application in adsorption. CrystEngComm 14 (21), 7099–7109. Kaur, R., Kaur, A., Umar, A., Anderson, W.A., Kansal, S.K., 2019. Metal organic framework (MOF) porous octahedral nanocrystals of Cu-BTC: synthesis, properties and enhanced adsorption properties. Mater. Res. Bull. 109, 124–133. Konik, P.A., Berdonosova, E.A., Savvotin, I.M., Klyamkin, S.N., 2019. The influence of amide solvents on gas sorption properties of metal-organic frameworks MIL-101 and ZIF-8. Microporous Mesoporous Mater. 277, 132–135. Lagergren, S., 1898. Zur theorie der sogenannten adsorption geloster stoffe. Langmuir, I., 1916. The constitution and fundamental properties of solids and liquids. Part I. Solids. J. Am. Chem. Soc. 38 (11), 2221–2295. Li, X., Guo, W., Liu, Z., Wang, R., Liu, H., 2016. Fe-based MOFs for efficient adsorption and degradation of acid orange 7 in aqueous solution via persulfate activation. Appl. Surf. Sci. 369, 130–136. Li, J.J., Wang, C.C., Fu, H.F., Cui, J.R., Xu, P., Guo, J., Li, J.R., 2017. High-performance adsorption and separation of anionic dyes in water using a chemically stable graphene-like metal–organic framework. Dalton Trans. 46 (31), 10197–10201. Li, Z., Liu, X., Jin, W., Hu, Q., Zhao, Y., 2019a. Adsorption behavior of arsenicals on MIL- 101 (Fe): the role of arsenic chemical structures. J. Colloid Interface Sci. 554, 692–704. Li, Z., Sellaoui, L., Dotto, G.L., Lamine, A.B., Bonilla-Petriciolet, A., Hanafy, H., Erto, A., 2019b. Interpretation of the adsorption mechanism of Reactive Black 5 and Ponceau 4R dyes on chitosan/polyamide nanofibers via advanced statistical physics model. J. Mol. Liq. 285, 165–170. M. Shakly et al.
  • 13. Journal of Contaminant Hydrology 247 (2022) 103977 13 Li, Z., Hanafy, H., Zhang, L., Sellaoui, L., Netto, M.S., Oliveira, M.L., Li, Q., 2020. Adsorption of congo red and methylene blue dyes on an ashitaba waste and a walnut shell-based activated carbon from aqueous solutions: experiments, characterization and physical interpretations. Chem. Eng. J. 388, 124263. Liu, J., Ye, J., Chen, Y., Li, C., Ou, H., 2018a. UV-driven hydroxyl radical oxidation of tris (2-chloroethyl) phosphate: intermediate products and residual toxicity. Chemosphere 190, 225–233. Liu, A., Wang, C.C., Wang, C.Z., Fu, H.F., Peng, W., Cao, Y.L., Du, A.F., 2018b. Selective adsorption activities toward organic dyes and antibacterial performance of silver- based coordination polymers. J. Colloid Interface Sci. 512, 730–739. Liu, Y., Zhu, M., Chen, M., Ma, L., Yang, B., Li, L., Tu, W., 2019. A polydopamine- modified reduced graphene oxide (RGO)/MOFs nanocomposite with fast rejection capacity for organic dye. Chem. Eng. J. 359, 47–57. Luo, X.P., Fu, S.Y., Du, Y.M., Guo, J.Z., Li, B., 2017. Adsorption of methylene blue and malachite green from aqueous solution by sulfonic acid group modified MIL-101. Microporous Mesoporous Mater. 237, 268–274. Lv, S.W., Liu, J.M., Ma, H., Wang, Z.H., Li, C.Y., Zhao, N., Wang, S., 2019. Simultaneous adsorption of methyl orange and methylene blue from aqueous solution using amino functionalized Zr-based MOFs. Microporous Mesoporous Mater. 282, 179–187. Mandal, T.N., Karmakar, A., Sharma, S., Ghosh, S.K., 2018. Metal-organic frameworks (MOFs) as functional supramolecular architectures for anion recognition and sensing. Chem. Rec. 18 (2), 154–164. Mantasha, I., Saleh, H.A., Qasem, K.M., Shahid, M., Mehtab, M., Ahmad, M., 2020. Efficient and selective adsorption and separation of methylene blue (MB) from mixture of dyes in aqueous environment employing a Cu (II) based metal organic framework. Inorg. Chim. Acta 511, 119787. Mobarak, M., Mohamed, E.A., Selim, A.Q., Mohamed, F.M., Sellaoui, L., Bonilla- Petriciolet, A., Seliem, M.K., 2019. Statistical physics modeling and interpretation of methyl orange adsorption on high–order mesoporous composite of MCM–48 silica with treated rice husk. J. Mol. Liq. 285, 678–687. Mohamed, E.A., Selim, A.Q., Ahmed, S.A., Sellaoui, L., Bonilla-Petriciolet, A., Erto, A., Seliem, M.K., 2020. H2O2-activated anthracite impregnated with chitosan as a novel composite for Cr (VI) and methyl orange adsorption in single-compound and binary systems: Modeling and mechanism interpretation. Chem. Eng. J. 380, 122445. Molavi, H., Hakimian, A., Shojaei, A., Raeiszadeh, M., 2018. Selective dye adsorption by highly water stable metal-organic framework: long term stability analysis in aqueous media. Appl. Surf. Sci. 445, 424–436. Paiman, S.H., Rahman, M.A., Uchikoshi, T., Abdullah, N., Othman, M.H.D., Jaafar, J., Ismail, A.F., 2020. Functionalization effect of Fe-type MOF for methylene blue adsorption. J. Saudi Chem. Soc. 24 (11), 896–905. Ramadan, H.S., Mobarak, M., Lima, E.C., Bonilla-Petriciolet, A., Li, Z., Seliem, M.K., 2021. Cr (VI) adsorption onto a new composite prepared from Meidum black clay and pomegranate peel extract: Experiments and physicochemical interpretations. J. Environ. Chem. Eng. 9 (4), 105352. Safaei, M., Foroughi, M.M., Ebrahimpoor, N., Jahani, S., Omidi, A., Khatami, M., 2019. A review on metal-organic frameworks: synthesis and applications. TrAC Trends Anal. Chem. 118, 401–425. Salama, R.S., El-Hakama, S.A., Samraa, S.E., El-Dafrawya, S.M., Ahmeda, A.I., 2018. Adsorption, equilibrium and kinetic studies on the removal of methyl orange dye from aqueous solution by using of copper metal organic framework (Cu-BDC). Int. J. Modern Chem 10 (2), 195–207. Seliem, M.K., Mobarak, M., Selim, A.Q., Mohamed, E.A., Halfaya, R.A., Gomaa, H.K., Dotto, G.L., 2020. A novel multifunctional adsorbent of pomegranate peel extract and activated anthracite for Mn (VII) and Cr (VI) uptake from solutions: experiments and theoretical treatment. J. Mol. Liq. 311, 113169. Sharib, A.S.A., Bonilla-Petriciolet, A., Selim, A.Q., Mohamed, E.A., Seliem, M.K., 2021. Utilizing modified weathered basalt as a novel approach in the preparation of Fe3O4 nanoparticles: experimental and theoretical studies for crystal violet adsorption. J. Environ. Chem. Eng. 9 (6), 106220. Wang, C., Liu, X., Demir, N.K., Chen, J.P., Li, K., 2016. Applications of water stable metal–organic frameworks. Chem. Soc. Rev. 45 (18), 5107–5134. Wang, D., Jia, F., Wang, H., Chen, F., Fang, Y., Dong, W., Yuan, X., 2018. Simultaneously efficient adsorption and photocatalytic degradation of tetracycline by Fe-based MOFs. J. Colloid Interface Sci. 519, 273–284. Wang, C.Y., Wang, C.C., Zhang, X.W., Ren, X.Y., Yu, B., Wang, P., Fu, H., 2021. A new Eu-MOF for ratiometrically fluorescent detection toward quinolone antibiotics and selective detection toward tetracycline antibiotics. Chin. Chem. Lett. https://doi. org/10.1016/j.cclet.2021.08.095. Weber, W.J., Morris, J.C., 1962, September. Advances in water pollution research. In: Proceedings of the First International Conference on Water Pollution Research, Vol. 2. Pergamon Press, Oxford, p. 231. Yang, C.X., Yan, X.P., 2011. Metal–organic framework MIL-101 (Cr) for high- performance liquid chromatographic separation of substituted aromatics. Anal. Chem. 83 (18), 7144–7150. Younas, M., Rezakazemi, M., Daud, M., Wazir, M.B., Ahmad, S., Ullah, N., Ramakrishna, S., 2020. Recent progress and remaining challenges in post- combustion CO2 capture using metal-organic frameworks (MOFs). Prog. Energy Combust. Sci. 80, 100849. Yuan, S., Feng, L., Wang, K., Pang, J., Bosch, M., Lollar, C., Zhou, H.C., 2018. Stable metal–organic frameworks: design, synthesis, and applications. Adv. Mater. 30 (37), 1704303. Zhang, X., Qian, L., Yang, S., Peng, Y., Xiong, B., Li, J., He, C., 2020. Comparative studies of methyl orange adsorption in various metal-organic frameworks by nitrogen adsorption and positron annihilation lifetime spectroscopy. Microporous Mesoporous Mater. 296, 109993. Zhao, T., Yang, L., Feng, P., Gruber, I., Janiak, C., Liu, Y., 2018. Facile synthesis of nano- sized MIL-101 (Cr) with the addition of acetic acid. Inorg. Chim. Acta 471, 440–445. M. Shakly et al.