Isotermas C0 Vs Qe, Qe

Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Journal of Environmental Chemical Engineering 9 (2021) 106214

Contents lists available at ScienceDirect

Journal of Environmental Chemical Engineering


journal homepage: www.elsevier.com/locate/jece

Facile fabrication of PVB-PVA blend polymer nanocomposite for


simultaneous removal of heavy metal ions from aqueous solutions: Kinetic,
equilibrium, reusability and adsorption mechanism
Hajar Azad a, Mohsen Mohsennia a, b, *, Chun Cheng c, Abbas Amini d, e, **
a
Thermodynamic Research Lab., Department of Chemistry, University of Kashan, Kashan, Iran
b
Department of Chemical Engineering, University of Kashan, Kashan, Iran
c
Department of Materials Science and Engineering, Southern University of Science and Technology, Shenzhen, People’s Republic of China
d
Department of Mechanical Engineering, Australian College of Kuwait, Mishref, Safat 13015, Kuwait
e
Centre for Infrastructure Engineering, Western Sydney University, Penrith, NSW 2751, Australia

A R T I C L E I N F O A B S T R A C T

Editor: Dr. G.L. Dotto The available poly (vinyl butyral) (PVB)-based materials are potential media for wastewater treatment, however,
their high butyral contents and low quantity of hydroxyl groups cause their low adsorption efficiency for heavy
Keywords: metals removal. Here, a new polymer blend nanocomposite of PVB-poly (vinyl alcohol), (PVB-PVA), incorpo­
Blend nanocomposite rated with different contents of stearic acid-grafted epsilon-manganese oxide nanoflowers, (PVB-PVA/SA-
Simultaneous removal
ε-MnO2), was prepared through blending, solution casting and solvent evaporation techniques. According to the
Adsorption kinetics
results, PVB-PVA/SA-ε-MnO2 nanocomposite containing 6 wt% SA-ε-MnO2 showed the maximum efficiency of
Adsorption isotherm
Reusability 99.88%, 94.26% and 87.12%, respectively, for the adsorptive removal of Cu2+, Cd2+ and Pb2+ metallic ions from
Electrostatic interaction aqueous solutions. The binary and ternary adsorption of the heavy metal ions in aqueous solutions indicated a
high adsorption affinity towards Cu2+. Kinetic modeling revealed that the adsorption of Cu2+, Cd2+ and Pb2+
ions follow the diffusion–chemisorption model (R2 = 0.99, RMSE = 0.31–0.56). The adsorption isotherms
modeling confirmed that the Langmuir isotherm well-fitted the adsorption data of the target ions on the nano­
composite surface (R2 = 0.96–0.99, RMSE = 3.20–9.68), suggesting a homogeneous monolayer surface
adsorption with the maximum adsorption capacity of 172.54, 112.89 and 111.23 mg/g for Cu2+, Cd2+ and Pb2+
ions, respectively. The reusability of PVB-PVA/SA-ε-MnO2 nanocomposite presented the removal efficiency of
94.50%, 92.05% and 88.60%, respectively, for Cu2+, Cd2+ and Pb2+ ions after five successive sorption/
desorption cycles. The great adsorption performance of PVB-PVA/SA-ε-MnO2 towards the heavy metal ions could
be attributed to the important role of various types of physicochemical interactions (e.g., electrostatic in­
teractions and surface complexation) in the adsorptive removal process.

1. Introduction body leading to infertility and dysfunction of kidneys, liver and lung [7].
The long-term accumulation of Cd2+ in human body can be responsible
Different industrial activities, such as mining, metal finishing, bat­ for high blood pressure, kidney damage, osteoporosis and destruction of
tery manufacturing, textile, solid wastes, petroleum refining, computer red blood cells [8,9]. And, the acute Cu2+ poisoning can cause Menkes’,
and mobile phone trashes etc., insert toxic organic and inorganic con­ Wilson’s, Alzheimer’s diseases or even death [10]. All these facts
taminants into water that cause serious threats to the ecosystem [1–4]. confirm the demands for developing an efficient setup for the removal of
Among various inorganic pollutants, heavy metal ions, i.e., Cu2+, Cd2+ such hazardous heavy metal ions from drinking and groundwater re­
and Pb2+, have been considered as one of the greatest concerns to sources [11].
human health as they are inclined to accumulate in living organisms [5, So far, various methods have been developed for heavy metals
6]. Continuing exposure to Pb2+ can cause serious damage to the human removal from water and wastewater including ion exchange [12],

* Corresponding author at:Thermodynamic Research Lab., Department of Chemistry, University of Kashan, Kashan, Iran.
** Corresponding author at: Department of Mechanical Engineering, Australian College of Kuwait, Mishref, Safat 13015, Kuwait.
E-mail addresses: [email protected] (M. Mohsennia), [email protected], [email protected], [email protected] (A. Amini).

https://doi.org/10.1016/j.jece.2021.106214
Received 12 July 2021; Received in revised form 8 August 2021; Accepted 12 August 2021
Available online 14 August 2021
2213-3437/© 2021 Elsevier Ltd. All rights reserved.
H. Azad et al. Journal of Environmental Chemical Engineering 9 (2021) 106214

coagulation-flocculation [13], reduction-precipitation [14], electrodi­ agglomeration tendency of ε-MnO2 nanoflowers, while it improves their
alysis, photocatalysis [15], chemical precipitation [16], membrane adsorption performance through the chemical interactions between
separation [17], and adsorption [18]. Among them, the adsorption carboxylic (-COOH) functional groups and heavy metal ions [44].
process is considered as an economical and potential technique for This work aims at improving the adsorption performance of PVB-
efficient removal of heavy metals from aqueous solutions due to its ef­ based materials for efficient removal of heavy metal ions from
ficiency, simplicity, cost-effectiveness, low energy consumption and aqueous solutions. A blended polymer composite of PVB and PVA
recyclability [19,20]. polymers was prepared through a facile blending technique and then
Up to date, a number of effective nanostructured materials, including functionalized with different contents (3, 6, 9 wt%) of stearic acid-
carbon nanotubes (CNTs) [21], graphene materials [22], metal oxides grafted ε-MnO2 (SA-ε-MnO2) nanoflowers. The (PVB-PVA)/SA-ε-MnO2
[23–27], metal organic frameworks (MOFs) [28], and polymer materials nanocomposites were fabricated by a solution casting method for the
[29], have received great attentions as adsorbents for the heavy metals sole and simultaneous removal of Cu2+, Cd2+ and Pb2+ ions from
removal purposes from aqueous solutions [30]. These adsorbents often aqueous solutions. The effects of adsorption parameters, such as initial
suffer from the lack of chelating groups, low adsorption efficiency, concentration, pH, contact time, and temperature were investigated
possible particle agglomeration and facile regeneration. The using batch adsorption experiments. The kinetic and diffusion models
polymer-based adsorbents are actually suitable for effective adsorption (including pseudo-first-order, pseudo-second-order, intra-particle
of heavy metals due to their superior structural strengths, modifiable diffusion, Elovich and diffusion-chemisorption models) and equilib­
surface functional groups, cost-effectiveness and eco-friendly properties rium adsorption isotherm models (including Langmuir and Freundlich
[31]. However, their low adsorption efficiency, low selectivity and poor models) were used to analyze the data.
mechanical stability are disadvantages limiting their practical applica­
tions in wastewater treatment. The surface functionalization with one or 2. Experimental
more different polymers through irreversible dispersion of nano­
structured materials have been considered as an effective technique to 2.1. Materials
improve the adsorption performance of the polymeric materials [31]. By
combining these materials, the resultant polymeric nanocomposites not The following materials were obtained from Merck Company (Ger­
only retain their inherent properties, but also present excellent me­ many): Potassium permanganate (KMnO4), manganese(II) sulfate mon­
chanical strength with high adsorption performance [32]. Besides, the ohydrate (MnSO4⋅H2O), lead(II) chloride (PbCl2), cadmium chloride
functional groups such as OH, -COOH and -NH2 from the nanostructured (CdCl2), copper(II) chloride dehydrate (CuCl2.2H2O), hydrochloric acid
materials incorporated into the polymer matrix, are beneficial for the 37% (HCl), sodium hydroxide (NaOH), butyraldehyde (C4H8O), pure
elimination of the heavy metals [31,32]. ethanol (C2H6O), stearic acid (SA, CH3(CH2)16COOH), and N, N-dime­
Among various polymer materials, poly (vinyl butyral) (PVB) has thylformamide (DMF, C3H7NO). PVA (C2H4O)n with a high molecular
been considered as one of the attractive materials for preparing the weight and degree of hydrolysis (87.0–89.0%) was received from Kur­
polymer nanocomposites for wastewater treatment because of its flex­ aray Co. Ltd. (Tokyo, Japan). All chemicals had the highest purity grades
ible molecular chain, low glass-transition temperature, excellent me­ and used with no further purifications. De-ionized (DI) water was used
chanical strength, good bio-compatibility, hydrophilic nature and strong for the preparation of all stock solutions in the experiments.
adhesive property with glass and metal [33,34]. The perfect structure of
PVB is composed of polar and hydrophilic vinyl alcohol and vinyl ace­ 2.2. Preparation of PVB
tate units and hydrophobic vinyl butyral units [35]. In recent years,
owing to the hydrophilic nature of the hydroxyl groups, PVB has been A condensation polymerization reaction between PVA and butyral­
utilized in ultrafiltration membranes for wastewater treatment [36–38]. dehyde was carried out in the presence of hydrochloric acid (HCl) as an
However, the systematic study on the fabrication and characterization of acidic catalyst for preparing the PVB polymer [17]. In brief, 5.5 g of
the PVB-based nanocomposite adsorbents is rather scarce. Currently, butyraldehyde and 90 mL of pure ethanol were added to a 500-mL
PVB with high contents of butyral groups (up to 84%) has been proposed three-necked glass batch reactor equipped with a stirrer, condenser,
[39], however, it seems that the PVB-based adsorbents with high butyral thermometer, and ultrasonic bath. PVA polymer (5 g) swelled with 5 mL
contents suffer from low adsorption efficiency for heavy metals removal DI water in a reactor while HCl was added dropwise under mechanical
purposes as per their low contents of hydroxyl groups. To improve the stirring to adjust the pH value. By raising the reaction temperature to
adsorption performance of PVB-based adsorbents, material blending 40 ◦ C after 30 min, the swollen PVA was entirely dissolved to attain a
technique can be used due to its simplicity, low cost, eco-friendliness homogeneous solution. Then, the reaction was carried out at 70 ◦ C for
and convenient in operation. This technique has been used to blend 4.5 h, and eventually, hot water (with the same temperature) was added
few numbers of materials, e.g., polyacrylonitrile (PAN) and poly­ to the solution to obtain PVB through precipitation. The obtained white
vinylidene fluoride (PVDF), with PVB membranes to modify their precipitant of PVB was washed with DI water and dried at 50 ◦ C in an
adsorption behavior for Cr(VI) removal purposes [40,41]. oven.
In the present study, to obtain a PVB-based adsorbent with high
hydroxyl contents, PVB was blended with polymeric nanomaterials 2.3. Preparation and surface functionalization of ε-MnO2 nanoflowers
containing high contents of polar hydroxyl groups, like poly (vinyl
alcohol) (PVA). Notably, through this blended polymeric hybrids (PVB- ε-MnO2 nanoflowers were synthesized through a hydrothermal
PVA), the drawbacks, such as relatively low adsorption capacity and method [45]. In brief, MnSO4⋅H2O and KMnO4 aqueous solutions were
decomposition in alkaline solutions, can be effectively eliminated [31]. stirred for 30 min at ambient temperature. The solution was poured into
Further improvement in the adsorption performance was achieved by a Teflon-lined stainless-steel autoclave, then the temperature was
the functionalization of PVB-PVA with metal oxides. For this purpose, maintained at 180 ◦ C for 12 h. After cooling down to ambient temper­
nanostructured manganese oxide (MnO2) presents distinctive physico­ ature, the brown solid slurry was obtained after centrifuging and
chemical properties, high surface area, negative surface charge, and low washing several times with pure ethanol and DI water. A facile grafting
point of zero charge [42]. With abundant porous sites, flower-like technique through a solvothermal process was carried out for modifying
nanostructure of epsilon-MnO2 (ε-MnO2) can be used as a superior the as-synthesized ε-MnO2 nanoflowers. 0.2 g of the as-synthesized
adsorbent for the removal of heavy metals [43]. On the other hand, ε-MnO2 nanoflowers were dispersed in 10 mL of ethanol and stirred
ε-MnO2 nanoflowers tend to agglomerate because of their high surface vigorously for 40 min under ultrasonic irradiation at 25 ◦ C. Then, an
energy. Grafting ε-MnO2 nanoflowers with stearic acid decreases the ethanol solution (30 mL) containing 0.6 g stearic acid was added to the

2
H. Azad et al. Journal of Environmental Chemical Engineering 9 (2021) 106214

solution which contained the prepared ε-MnO2 nanoflowers. The nanocomposites (PPSMs) was examined using field emission scanning
mixture was poured into a Teflon-lined stainless-steel autoclave and electron microscopy (FE-SEM, TESCAN BRNO-Mira3 LMU, Czech Re­
kept at 120 ◦ C for 6 h. After cooling down to ambient temperature, the public). Fourier transform infrared spectroscopy (FTIR, Magana 550,
mixture was centrifuged and the obtained brown color solids, stearic Nicolet, USA) was employed to examine the chemical structures of the
acid-grafted ε-MnO2 (SA-ε-MnO2) nanoflowers, were washed with prepared materials to confirm the formation of functional groups. X-ray
ethanol and DI water several times and dried at 60 ◦ C in air for 12 h. The diffraction (XRD) patterns were recorded using X-ray diffractometer (X-
fabrication process of SA-ε-MnO2 nanoflowers is schematically pre­ pert pro, Philips, Netherlands) with monochromatic CuKα radiation
sented in Scheme 1. (λ = 1.54 Å) at 2θ range of 10–80◦ . Brunauer-Emmett-Teller (BET)
(BELSORP-mini II analyzer) and Barrett-Joyner-Halenda (BJE) methods
2.4. Preparation of (PVB-PVA)/SA-ε-MnO2 nanocomposites were used to analyze the specific surface area and average pore diam­
eter, respectively. The equilibrium concentrations of the heavy metals
The fabrication process of (PVB-PVA)/SA-ε-MnO2 nanocomposites is were measured by a flame atomic absorption spectrometer (FAAS,
schematically presented in Scheme 2. The (PVB-PVA) (1:1) blended Varian Spectra AA 220FS, USA).
polymer composite incorporated with SA-ε-MnO2 nanoflowers, was
prepared through solution casting and solvent evaporation techniques. 2.6. Batch adsorption experiments
First, 0.1 g of as-synthesized PVB was dissolved in 5 mL ethanol and
5 mL DMF at 50 ◦ C for 6 h under continuous stirring and 20 min ultra­ The adsorption experiments were performed using a batch adsorp­
sonication to obtain a clear solution. Then, 0.1 g of PVA was dissolved in tion process to investigate the adsorption quantity (qe, mg/g) and the
7 mL DI water and 3 mL ethanol at 85 ◦ C and then added to the as- removal efficiency (RE%), which are expressed by Eqs. (1) and (2),
prepared PVB solution and sonicated for 20 min. Next, certain respectively [46,47]:
amounts of SA-ε-MnO2 nanoflowers (3, 6 and 9 wt%) were magnetically
dispersed in 6 mL of ethanol using ultrasonication for 20 min. The SA- (C0 − Ce )
qe = .V (1)
ε-MnO2 nanoflowers contained suspension was added to the (PVB-PVA) m
blended solution and sonicated for 1 h. The obtained mixture was stirred
(C0 − Ce )
for 24 h at ambient temperature and sonicated again for 1 h. The (PVB- RE% = (2)
C0
PVA)/SA-ε-MnO2 nanocomposites containing 3, 6 and 9 wt% of SA-
ε-MnO2 nanoflowers, respectively labeled as PPSM-3, PPSM-6 and where C0 and Ce respectively are the initial and equilibrium concen­
PPSM-9, were prepared by pouring the obtained homogenous mixture trations (mg/L), m is the mass of adsorbents (g), and V is the volume of
into glass Petri dishes, dried at ambient temperature for 48 h, and stored aqueous solutions (L). Stock aqueous solutions were prepared by dis­
for further experiments. solving certain amounts of Cu2+, Cd2+ and Pb2+ ions in DI water
(60 mg/L). The adsorption experiments were carried out by the pre­
2.5. Characterizations pared adsorbents (PPSMs) (0.02 g) and 20 mL of the heavy metals
contained solutions in 250 mL glass beaker. The initial concentrations
The morphology of the as-prepared SA-ε-MnO2 nanoflowers and were within the range of 20–180 mg/L. The effect of pH value on the

Scheme 1. The schematic of the preparation route of SA-ε-MnO2 nanoflowers.

3
H. Azad et al. Journal of Environmental Chemical Engineering 9 (2021) 106214

Scheme 2. Schematic of the preparation route of (PVB-PVA)/SA-ε-MnO2 nanocomposite.

adsorption capacity was investigated at pH range of 1–7 using 0.1 M adsorbent was used for the next sorption/desorption cycles [11].
HCl/NaOH solutions. The effect of contact time on the adsorption
quantity was studied at the intervals of 0.5–30 h at ambient temperature 3. Results and discussion
(25 ± 0.1 ◦ C). In each step, the equilibrium concentrations were
measured using FAAS. 3.1. Material characterization

2.7. Simultaneous adsorption The FTIR spectra of ε-MnO2 and SA-ε-MnO2 nanoflowers are shown
in Fig. 1a and b, respectively. As shown in Fig. 1a, the strong band
The equimolar binary solutions (Cu2+-Cd2+, Cu2+-Pb2+ and Cd2+- centered at 3408 cm− 1 is associated with the stretching vibration of H-
Pb ) and ternary solutions (Cu2+-Cd2+-Pb2+) were employed to
2+ bonding of hydroxyl groups. The Mn-O vibration of SA-ε-MnO2 nano­
investigate the simultaneous adsorption of the heavy metal ions under flowers is observed at 442 and 507 cm− 1. The comparison between the
the same operating conditions of the single-metal adsorption process. spectra in Fig. 1a and b reveals the shift of broad peak at 3408 cm− 1 to
The initial concentration of each metal ion in the binary and ternary
solutions was set as 60 mg/L. The adsorption quantity of the prepared
adsorbent for each metal ion was calculated for the binary and ternary
solutions. Also, the selectivity coefficient, αij , of PPSM-6 adsorbent for
the adsorption of heavy metal ions in the binary systems can be found
through Eq. (3) [48]:
qe,i ce,j
αij = (3)
qe,j ce,i

where parameters qe,i and qe,j are the adsorption quantity at equilibrium
(mg/g), and Ce,i and Ce,j are the equilibrium concentrations (mg/L) of
two metal ions i and j in the binary metal solutions.

2.8. Regeneration experiments

Reusability performance of the prepared adsorbent was investigated


during five sequential cycles. The experiments were carried out using
aqueous solutions of Cu2+, Cd2+ and Pb2+ ions (60 mg/L) at room
temperature and the optimum pH values (5, 6 and 4 for Cu2+, Cd2+ and
Pb2+, respectively). When the mixture reached the equilibrium/satura­
tion conditions, the adsorbent was taken out, rinsed with DI water and
kept in HCl solution (0.1 M) for 20 min, followed by washing with DI Fig. 1. FTIR spectra of: (a) ε-MnO2, (b) SA-ε-MnO2 nanoflowers, (c) pristine
water and drying in a vacuum oven at 65 ◦ C. Then, the regenerated PVB, (d) (PVB-PVA) blended composite, and (e) PPSM-6 nanocomposite.

4
H. Azad et al. Journal of Environmental Chemical Engineering 9 (2021) 106214

higher wavenumber (3428 cm− 1) which may be due to the -OH vibra­ broad peak of (PVB-PVA) blended composite.
tion in SA-ε-MnO2 nanoflowers. The absorption peaks at 2849 and The surface morphology of ε-MnO2, SA-ε-MnO2 nanoflowers, (PVB-
2918 cm− 1 are assigned to the stretching vibration of C− H bonds in PVA) blended composite, PPSM-3, PPSM-6 and PPSM-9 were investi­
alkyl groups, and 441 and 511 cm− 1 to the Mn-O bond. The spectrum in gated via FE-SEM analysis as shown in Fig. 3(a-f). Fig. 3(a-1) and 3(b-1)
Fig. 1c shows a strong peak in pure PVB at 1001 cm− 1 because of C− O respectively confirm the successful synthesis of ε-MnO2 and SA-ε-MnO2
stretching. The absorption peaks at 1132, 1382, 1436 and 1738 cm− 1 nanoflowers. The flower-like nanostructures of as-synthesized ε-MnO2
are respectively attributed to the stretching vibration of C− O− C of and SA-ε-MnO2 with a diameter of ~135 and 125 nm were, respectively,
butyral group, CH3 stretching mode, CH2 bending mode, and C– – O vi­ obtained by the magnified FE-SEM images (Fig. 3(a-2) and 3(b-2)). The
bration. Two relatively strong absorption peaks at 2925 and 2872 cm− 1 SA-ε-MnO2 nanoflowers showed lower agglomeration compared to the
are corresponded to the stretching mode of aliphatic C− H. Another ε-MnO2 nanoflowers. According to the FE-SEM results, the flower-liked
relatively strong absorption peak centered at 3445 cm− 1 is referred to morphology of ε-MnO2 was perfectly maintained after grafting with SA.
the − OH absorption. The FTIR results confirm the backbone structure of According to FE-SEM images indicated in Fig. 3c and d, a uniform
the PVB polymer which is composed of vinyl butyral, vinyl alcohol and morphology with smooth surface and low porosity can be observed for
vinyl acetate units. the pristine PVB and (PVB-PVA) blended composite. The surface
The functional groups of PVB-PVA blended composite are verified by morphology of PPSM-3 and PPSM-6 adsorbents indicated a uniform
the FTIR spectrum in Fig. 1d. The absorption bands around 3353, 2940 dispersion of SA-ε-MnO2 nanoflowers throughout the (PVB-PVA)
and 1737 cm− 1 are respectively attributed to -OH stretching, C-H blended matrix (Fig. 3e and f). This might be due to a decrease in the
stretching and C– – O stretching vibrations. Fig. 1e displays the FTIR surface energy and H-bonding interactions between SA-ε-MnO2 nano­
spectra of PPSM-6 nanocomposite. Compared to Fig. 1d, the absorption flowers and hydroxyl functional groups in the (PVB-PVA) blend com­
peaks of -OH stretching around 3434 cm− 1 as well as the carbonyl group posite. However, PPSM-9 possessed a nonuniform morphology which
at 1732 cm− 1 show a shift to the lower wavenumber. This can be orig­ was attributed to the aggregation of SA-ε-MnO2 nanoflowers throughout
inated from the strong H-bonding interactions between SA-ε-MnO2 the (PVB-PVA) blended matrix (Fig. 3g).
nanoflowers and hydroxyl functional groups in the backbone of (PVB- To further characterize the specific surface area and porosity of the
PVA) blended composite. The presence of weak peaks at fabricated adsorbents, the hysteresis of N2 adsorption-desorption iso­
~400–500 cm− 1, related to the Mn-O group, indicates the successful therms and pore diameter distribution of the prepared adsorbents
embedding of SA-ε-MnO2 nanoflowers within the (PVB-PVA) blended (PPSM-3, PPSM-6, PPSM-9) were evaluated. As indicated in Fig. 4a, all
composite. the PPSM adsorbents showed distinct N2 adsorption-desorption
The crystalline structure of the synthesized ε-MnO2 and SA-ε-MnO2 isotherm hysteresis loops in the range of 0.4–1.0 p/p0, indicating a
nanoflowers was analyzed using the XRD characterization at 2θ from 10◦ mesoporous structure. The BET surface area, average pore diameter, and
to 80◦ (Fig. 2a). As shown in Fig. 2, the broad peaks located at pore volume of PPSM adsorbents are tabulated in Table 1. The values of
2θ = 19.0, 37.4, 41.7, 55.9 and 66.05◦ were assigned to (0 0 1), specific surface area for PPSM-3, PPSM-6 and PPSM-9 were found to be
(1 0 0), (1 0 1), (1 0 2), and (1 1 0) planes of ε-MnO2 (JCPDS card No. 15.414, 35.289 and 13.439 m2/g, respectively. In addition, the di­
00–030–0820) before SA grafting. The diffraction spectrum of SA- ameters of the adsorbents were in the range of 3.86–4.85 nm, confirm­
ε-MnO2 nanoflowers is depicted in Fig. 2b indicating some extra peaks ing their mesoporous structure. From Table 1, it is clear that the values
related to stearic acid appeared at 2θ = 20.8, 21.6, 23.0, 24.0, 24.7 of the specific surface area for the PPSM adsorbents were enhanced upon
and 41.1◦ . After SA grafting, the crystallin phase of ε-MnO2 nanoflowers increasing the SA-ε-MnO2 contents. However, it was reduced for PPSM-9
was maintained. According to Fig. 2c, (PVB-PVA) blended composite has which may be due to the aggregation of SA-ε-MnO2 throughout the
a broad and amorphous diffraction peak at 2θ = 20◦ . The characteristic (PVB-PVA) blended matrix as proved by FE-SEM image in Fig. 3g. By
diffraction peaks related to SA-ε-MnO2 nanoflowers are not observed in considering the positive effect of larger specific surface area on the
the XRD pattern of PPSM-6 (Fig. 2d), which can be due to the low adsorption efficiency, the prepared PPSM-6 containing 6 wt% SA-
content of SA-ε-MnO2 nanoflowers as well as the high intensity of the ε-MnO2 with the highest specific surface area (35.289 m2/g) was chosen
background and covering the peaks of SA-ε-MnO2 nanoflowers with as the best adsorbent for further adsorption studies.

3.2. Adsorption studies

The adsorption performance of the as-synthesized (PVB-PVA)/SA-


ε-MnO2 nanocomposites towards Cu2+, Cd2+, and Pb2+ ions were eval­
uated. As shown in Fig. 5a, the pristine PVB had a small adsorption
quantity for three heavy metal ions with a maximum adsorption quan­
tity towards Pb2+ ions. After blending PVA polymer, the adsorption
quantity of pristine PVB increased substantially for all heavy metal
cations. This substantial enhancement in the adsorption quantity was
attributed to an increase in the content of the hydroxyl functional groups
on the surface of (PVB-PVA) blended composite, that played a leading
role in the adsorption of heavy metals. As indicated in Fig. 5a, the (PVB-
PVA) blended composite had a higher adsorption quantity for Cd2+ and
Cu2+, respectively, rather Pb2+. Furthermore, the effect of SA-ε-MnO2
contents on the adsorption quantity of PPSM-3, PPSM-6 and PPSM-9 for
adsorption of Cu2+, Cd2+, and Pb2+ ions was evaluated as indicated in
Fig. 5a. As expected, after embedding different amounts of SA-ε-MnO2
nanoflowers in the (PVB-PVA) blended polymer matrix, a significant
increase in the adsorption quantity of the PVB-based nanocomposite
adsorbents (PPSMs) was achieved for all three metal ions especially for
Cu2+. This significant increase in the adsorption quantity can be
Fig. 2. XRD patterns of: (a) ε-MnO2, (b) SA-ε-MnO2 nanoflowers, (c) (PVB- ascribed to the synergistic effect of both (PVB-PVA) blended composite
PVA) blended composite, and (d) PPSM-6 nanocomposite. and SA-ε-MnO2 nanoflowers on the adsorption performance of the

5
H. Azad et al. Journal of Environmental Chemical Engineering 9 (2021) 106214

Fig. 3. FE-SEM images of: (a–1 and 2) ε-MnO2 nanoflowers, (b–1 and 2) SA-ε-MnO2 nanoflowers, (c) pristine PVB, (d) (PVB-PVA) blended composite, (e) PPSM-3, (f)
PPSM-6, and (g) PPSM-9.

6
H. Azad et al. Journal of Environmental Chemical Engineering 9 (2021) 106214

Fig. 4. (a) N2 adsorption-desorption isotherms, and (b) distribution of pore diameter of PPSM-3, PPSM-6 and PPSM-9.

continuously enhanced by growing the pH level till the optimal pH levels


Table 1 appeared for the adsorption quantities at pH = 4, 5 and 6 for Pb2+, Cu2+
Physical properties of (PVB-PVA)/SA-ε-MnO2 nanocomposites.
and Cd2+, respectively. The maximum adsorption quantities were ach­
Adsorbent Surface area Average pore diameter Pore volume ieved and after that a decreasing trend was observed as the pH level
(m2/g) (nm) (cm3/g)
increased to 7. According to the results in Fig. 5d, a relatively low
PPSM-3 15.414 4.851 0.0187 adsorption quantity was achieved at acidic pH, which could be attrib­
PPSM-6 35.289 4.580 0.0404 uted to not only the occupancy of the active adsorptive sites on the
PPSM-9 13.439 3.863 0.0129
surface of PPSM-6 by H+ ions, but also to the strongly competitive
adsorption between heavy metal ions and H+ ions. By increasing the pH
adsorbents. In this regard, high amounts of the hydroxyl and carboxyl level to neutral pH, the adsorption quantity was gradually increased due
functional groups on the surface of (PVB-PVA) blended composite and to the availability of active surface sites for the heavy metal ions, which
SA-ε-MnO2 as well as the more structural faults and pores on the surface was accompanied with decreasing the number of H+ ions on PPSM-6
of SA-ε-MnO2 nanoflowers played the main role in capturing the heavy through dissociating and binding with hydroxide anions. At the higher
metals ions. The adsorption quantity was also enhanced with the in­ pH (higher than the optimal pH values for each metal ion), the decrease
crease in SA-ε-MnO2 contents up to 6 wt% because of high specific in the adsorption quantity could be attributed to the formation of hy­
surface area of the adsorbent (PPSM-6). By further increase in the SA- droxyl complexes by the excess OH- ions. At pH > 7, the adsorption
ε-MnO2 loading up to 9 wt% (PPSM-9), a gradual reduction in the examination was not possible due to the precipitation of the metal hy­
adsorption of the metal ions was observed especially for Pb2+. This was droxide complexes in the form of Cu(OH)2, Cd(OH)2 and Pb(OH)2.
probably due to the aggregation of SA-ε-MnO2 nanoflowers which Fig. 5e displays the effect of contact time on the adsorption quantity
decreased the available active adsorptive sites on the surface of PPSM-9. (qt ) of PPSM-6 for the heavy metal ions. The qt values, specifically for
Fig. 5b illustrates the adsorption quantity of PPSM-6 for Cu2+, Cd2+ Cu2+, were increased quickly in the first 6 h. Thereafter, it was increased
and Pb2+ as a function of initial concentrations. As shown, the adsorp­ with a slower rate because of the diffusion of metal ions across the
tion quantity was increased by increasing the initial concentration up to PPSM-6 pores with the equilibrium state after 12 h for Cu2+, 18 h for
the saturation point, where the dynamic equilibrium was established Cd2+, and 24 h for Pb2+. It was found that the adsorption performance of
and further increase in the adsorbate concentration had no significant the prepared adsorbent greatly depended on the electrostatic in­
effect on the adsorption. The adsorption quantity of PPSM-6 gradually teractions between the heavy metal ions and the functional groups on
increased with increasing the initial concentration up to 140 mg/L for the surface. The adsorption quantity of PPSM-6 for the studied ions had
Cd2+ and Pb2+ and 180 mg/L for Cu2+. The maximum saturation ad­ the order of Cu2+ > Cd2+ > Pb2+, which might be related to the ionic
sorptions were found to be 155.02, 98.32 and 87.74 for Cu2+, Cd2+ and radius and electronegativity of the metal ions. In fact, the adsorption
Pb2+, respectively (Fig. 5b). As shown in Fig. 5c, the RE% value for Pb2+, quantity was inversely and directly proportional to the ionic radius
Cd2+ and Cu2+ decreased significantly upon increasing the initial con­ (Cu2+ < Cd2+ < Pb2+) and the electronegativity of the metal ions (Cu2+
centration, which can be attributed to the decrease in the active surface > Pb2+ > Cd2+) (Table 2) [14]. As a result, Cu2+ with the smaller ionic
sites. The maximum value of RE% was obtained as 99.88%, 94.26% and radius and higher electronegativity was adsorbed quickly by the surface
87.12% for Cu2+, Cd2+ and Pb2+, respectively, with the initial concen­ adsorption active sites, resulted in the equilibrium/saturation condi­
tration of 20 mg/L. At higher initial concentration (180 mg/L), the tions at a shorter time.
values of RE% were 93.32%, 70.23% and 62.67% for Cu2+, Cd2+ and The effect of temperature (298, 303 and 318 K) on the adsorption of
Pb2+, respectively. Also, the obtained results confirmed that PPSM-6 Cu2+, Cd2+ and Pb2+ ions onto PPSM-6 was studied at the initial con­
showed stronger adsorption affinity towards Cu2+ than Cd2+ and centration of 60 mg/L (Fig. 5f). As shown, the adsorption quantity of
Pb2+. This can be rationalized by the lowest ionic radius of Cu2+ than the Cu2+, Cd2+ and Pb2+ ions onto PPSM-6 was increased from 58.73 to
ones for Cd2+ and Pb2+ [49]. 87.26 mg/g for Cu2+, from 53.81 to 62.15 mg/g for Cd2+, and from
The effect of pH level on the adsorption quantity of PPSM-6 towards 50.94 to 64.02 mg/g for Pb2+ by increasing temperature from 298 to
Cu2+, Cd2+, and Pb2+ ions was evaluated at various initial pH levels 318 K. Therefore, the order of adsorption quantity at the scale of three
(pH = 1–7) at the concentration of 60 mg/L for all three metal ions at studied temperatures was Cu2+> Cd2+> Pb2+ (Fig. 5f). The adsorption
room temperature. As shown in Fig. 5d, the adsorption quantity was quantity enhancement of the studied ions on PPSM-6 could be due to the

7
H. Azad et al. Journal of Environmental Chemical Engineering 9 (2021) 106214

Fig. 5. The effect of different parameters on the adsorption of Pb2+, Cd2+ and Cu2+ ions versus; (a) SA-MnO2 nanorods content (wt%), (b) the initial concentrations
on the adsorption quantity (qe), (c) removal efficiency (%), (d) pH values, (e) contact time, and (f) temperature.

ion mobility and their tendency to interact with the active surface sites 3.3. Simultaneous adsorption
as well as the intense activity of the active adsorptive sites at higher
temperatures. The finding verified the endothermic nature of adsorption Generally, wastewater contain multi-heavy metal ions. Table 3 elu­
process of Cu2+, Cd2+ and Pb2+ ions onto PPSM-6. cidates the adsorption quantity and selectivity coefficient of binary
(Cu2+-Cd2+, Cu2+-Pb2+ and Cd2+-Pb2+) and ternary (Cu2+-Cd2+-Pb2+)

8
H. Azad et al. Journal of Environmental Chemical Engineering 9 (2021) 106214

Table 2 3.4. Adsorption kinetics


Ionic properties of the heavy metal ions.
Metal ion Ionic radius (Å) Hydrated radius (Å) Electronegativity 3.4.1. Pseudo-first-order and pseudo-second-order models
2+ To inspect the mechanism of the target heavy metals adsorption on
Cu 0.69 4.19 1.90
Cd2+ 1.19 4.01 1.69 PPSM-6, the experimental data was scrutinized by the non-linear forms
Pb2+ 0.97 4.26 1.87 of the Lagergren pseudo-first-order (Eq. (4)) and pseudo-second-order
kinetic (Eq. (5)) models as follow [52]:

qt = qe (1 − e− k1 t
) (4)
Table 3
Competitive adsorption of single, binary and ternary solutions. k2 q2e t
qt = (5)
Solution Components Adsorption capacity Selectivity 1 + k2 qe t
(mg/g) coefficient

Cu2+ Cd2+ Pb2+ αCu αCd αPb where qe and qt respectively are the adsorption quantity (mg/g) of the
adsorbent at equilibrium and time t (min); k1 (1/min) and k2 (g/mg min)
Single 58.73 53.81 50.94 –
Binary Cu2++ Cd2+ 51.28 45.36 – 1.90 0.53 – are the rate constants related to the pseudo-first-order and pseudo-
Cu2++ Pb2+ 53.61 – 42.96 3.33 – 0.30 second-order kinetics models, respectively. The results related to the
Cd2++ Pb2+ – 48.96 36.65 – 2.82 0.35 fitting curves of the pseudo-first-order and pseudo-second-order kinetic
Ternary Cu2++ Cd2++ 44.45 39.54 25.23 – models are respectively illustrated in Fig. 6a and b. The kinetic experi­
Pb2+
mental data obtained at the initial concentration of 60 mg/L for three
metal ions were used for fitting; and, the related kinetic parameters, root
solutions by PPSM-6. As seen in Table 3, the adsorption quantities for the mean squared error (RMSE) and correlation coefficients (R2) were
binary and ternary solutions were lower than those of the single-metal computed and summarized in Table 4. As seen, the R2 values of the
ion solutions. Compared to a single metal-contained solution, the pseudo-first-order model were 0.64, 0.74 and 0.66, for Cu2+, Cd2+ and
adsorption of the ions in a solution containing multi-metal ions by the Pb2+, respectively. Obviously, the values of R2 for the adsorption of
adsorbent was expected to show very different behavior due to simul­ Cu2+, Cd2+ and Pb2+ ions fit poorly with the pseudo-first-order model.
taneous adsorption and selectivity. Competition between metal ions Also, the calculated qe (calc.) values are lower than the qe (exp.) ones.
decreased the adsorption quantity of each metal ion. Inversely, the R2 values suit better with the pseudo-second-order kinetic
The simultaneous competitive adsorption of Pb2+, Cd2+ and Cu2+ model which were approximately 0.91, 0.96 and 0.91 for Cu2+, Cd2+
ions on the prepared PPSM-6 were investigated in the binary and ternary and Pb2+, respectively. With this model, the calculated adsorption
solutions. This competitive adsorption occurs between co-existing heavy quantities were found to be 57.99, 53.45 and 47.43 for Cu2+, Cd2+ and
metal ions on the available adsorptive sites [50,51]. By comparing the Pb2+, respectively, highly consistent with the experimental adsorption
single, binary and ternary solutions, the largest adsorption quantity of quantities. As a result, the pseudo-second-order kinetic model was more
PPSM-6 was acquired towards Cu2+. The Cu2+ adsorption quantity was applicable to describe the kinetics of the adsorption process of the target
51.28, 53.61 and 44.45 mg/g for Cu2+-Cd2+, Cu2+-Pb2+ and Cu2+- heavy metal ions by PPSM-6, indicating the dominant role of chemi­
Cd2+- Pb2+ solutions, respectively. According to the results, it can be sorption in the adsorption process. It is suggested that the adsorption of
found that in the binary and ternary solutions, the adsorption process of metal ions on the surface of PPSM-6 occurred by the chemical in­
Cd2+ and Pb2+ was suppressed in the presence of Cu2+. Moreover, the teractions, such as surface complexation and electrostatic interactions
adsorption amount of Cd2+ was 48.96, 45.36 and 39.54 mg/g in [53–55].
Cd2+-Pb2+, Cu2+-Cd2+ and Cu2+-Cd2+-Pb2+ solutions, respectively.
Similarly, the adsorption amount of Pb2+ was significantly reduced to 3.4.2. Intra-particle diffusion
42.96, 36.65 and 25.23 mg/g in Cu2+-Pb2+, Cd2+-Pb2+ and Cu2+- Cd2+- The experimental data was also modeled by intra-particle diffusion
Pb2+ solutions. Interestingly, the adsorption affinity of three heavy as expressed by Eq. (6):
metal ions in ternary metal solutions was ordered as Cu2+ > Cd2+ >
qt = kip t1/2 + C (6)
Pb2+. This fact also indicated that Cu2+ had a stronger affinity than Cd2+
and Pb2+. Pb2+ had a less pronounced decrease in simultaneous
where kip is the intra-particle diffusion rate constant (g/mg⋅min0.5) and
competitive adsorption compared to other heavy metal ions (Table 3).
C is the intercept of the plot (mg/g) that is the thickness of the boundary
Increasing tendency of PPSM-6 towards Cu2+ confirmed the important
layer. Fig. 6c indicates the fitting curves of the intra-particle diffusion for
role of electronegativity, ionic and hydrated radius on the adsorption
the adsorption of Cu2+, Cd2+ and Pb2+ ions on PPSM-6. The related
process [51]. Nevertheless, the adsorption quantity showed the order of
constants, kip and C, are briefly tabulated in Table 4. The R2 values for
Cu2+ > Cd2+ > Pb2+, indicating more importance role allocated to ionic
Cu2+, Cd2+ and Pb2+ ions were obtained as 0.91, 0.86 and 0.93,
radius than electronegativity. This result was in a good agreement with
respectively (RMSE = 0.79, 1.05, and 0.64). It suggests that the intra-
the adsorption quantities obtained in the single heavy metals solutions
particle diffusion model was not applicable for fitting the experi­
(Section 3.2).
mental data. According to Fig. 6c, the plots did not pass the origin, and
The selectivity coefficient, αij of the heavy metal ions onto PPSM-6 the intercept value (C) was not zero. This demonstrated that the intra-
were also tabulated in Table 3, where the obtained αij for the Cu2+- particle diffusion model was not the sole rate-controlling step in the
Cd2+ and Cu2+-Pb2+ binary solutions indicated the preference of PPSM- adsorption of Cu2+, Cd2+ and Pb2+ ions on PPSM-6.
6 for the adsorption of Cu2+ ions [48]. Moreover, for the Cd2+-Pb2+
binary solution, αij > 1 implied the greater adsorption affinity of PPSM-6 3.4.3. Elovich model
towards Cd 2+ 2+
than Pb . It was found that the αij values for the ions Elovich model is regarded as one of the most valuable models for
predicting the kinetic mechanism of heterogeneous chemisorption of the
followed such trend: Cu2+ > Cd2+ > Pb2+. In the view of the αij data, it
heavy metal ions on the adsorbent surface which is expressed by Eq. (7):
was concluded that Cu2+ adsorption onto the PPSM-6 had an
outstanding selectivity. So, a promising application is offered for the 1
qt = ln(1 + αβt) (7)
removal of Cu2+ from wastewater containing Cu2+, Cd2+ and Pb2+ β
multi-heavy metal ions.

9
H. Azad et al. Journal of Environmental Chemical Engineering 9 (2021) 106214

Fig. 6. The fitting curves of adsorption kinetic and diffusion models: (a) Pseudo-first order, (b) Pseudo-second-order, (c) Intra-particle diffusion, (d) Elovich, and (e)
Diffusion-chemisorption for the adsorption of Cu2+, Cd2+ and Pb2+ ions onto PPSM-6.

here, α (mg/g. min) and β (g/mg) are related to the adsorption rate at The values of the correlated parameters (e.g., α, β, and R2) are listed in
initial time, the correlation between the activation energy, and the de­ Table 4. As indicated in Fig. 6d and summarized in Table 4, the non-
gree of surface coverage. Fig. 6d displays the non-linear fitting of Elo­ linear curves all fit well with Elovich model, indicating that the
vich model for the adsorption of Cu2+, Cd2+ and Pb2+ ions on PPSM-6. adsorption of Cu2+, Cd2+ and Pb2+ ions on PPSM-6 can be described by

10
H. Azad et al. Journal of Environmental Chemical Engineering 9 (2021) 106214

Table 4 valuable insight into dominant adsorption behavior of PPSM-6 towards


Kinetic parameters of pseudo-first-order, pseudo-second-order, Intra-particle Cu2+, Cd2+ and Pb2+ ions, schematic diagrams of the linear forms of
diffusion, Elovich, and Diffusion-chemisorption models for the adsorption of Langmuir and Freundlich isotherm models are shown in Fig. 7 along
Cu2+, Cd2+ and Pb2+ ions onto PPSM-6. with parameters in Table 5. Obviously, the higher value of R2 for
Kinetic model Parameters Cu2þ Cd2þ Pb2þ Langmuir isotherm model indicated the superiority of the Langmuir
pseudo-first order qe (exp.) (mg/g) 58.73 53.81 48.54 model for modeling the adsorption experimental data over Freundlich
qe (calc.) (mg/g) 55.66 51.42 45.46 model for all three metal ions. This confirmed a homogenous adsorbent
k1 (1/min) 0.0401 0.0388 0.0385 with a monolayer coverage [58–60]. The maximum adsorption capac­
R2 0.64 0.74 0.66 ities (qm) calculated using Langmuir isotherm model was a little higher
RMSE 3.94 3.01 3.25
pseudo-second order qe (calc.) (mg/g) 57.99 53.45 47.43
that those obtained by experimental, qe (exp.). In Table 6, the values
k2 × 10− 3 (g/mg⋅min) 1.26 1.37 1.46 related to qm of Cu2+, Cd2+ and Pb2+ onto PPSM-6 were compared with
R2 0.91 0.96 0.91 those of the other adsorbents presented in literature. As observed, the qm
RMSE 1.98 1.13 1.67 of Cu2+, Cd2+ and Pb2+ onto the PPSM-6 ((PVB-PVA)/SA-ε-MnO2) is in
Intra-particle diffusion kip (g/mg⋅min0.5) 0.56 0.30 0.34
the range of qm obtained by other adsorbents for these heavy metal ions.
C 43.51 44.34 37.26
R2 0.97 0.89 0.97 This fact indicates the superior adsorption performance of
RMSE 0.79 1.05 0.64 (PVB-PVA)/SA-ε-MnO2 nanocomposite with respect to other adsorbents.
Elovich α (mg/g⋅min) 4.54E3 6.30E3 2.04E3
β (g/mg) 0.24 0.27 0.28
R2 0.97 0.94 0.98 3.6. Adsorption thermodynamics
RMSE 1.15 1.42 0.74
Diffusion-chemisorption qe (calc.) (mg/g) 63.13 57.90 51.89 Thermodynamic behavior of the adsorption process by PPSM-6 was
KDC (mg/g⋅min0.5) 22.24 21.12 17.17
evaluated through Eq. (11):
R2 0.99 0.99 0.99
RMSE 0.56 0.41 0.31 ∆G◦ = ∆H◦ − T∆S◦ = − RTlnKd (11)

where ΔG◦ , ΔH◦ , and ΔS◦ are changes in Gibbs free energy (kJ/mol),
Elovich model. From such trend, it is discovered that the adsorption via
enthalpy (kJ/mol), and entropy (kJ/mol.K), respectively. R, T and Kd are
chemical interactions between the active adsorptive sites on the surface
gas constant (8.314 J/mol.K), absolute temperature (K), and equilib­
of PPSM-6 and the target heavy metal ions is uniquely the governing
rium constant. Kd was obtained as a dimensionless parameter by
factor.
multiplying kL (Langmuir constant) by the atomic mass of the adsorbed
species (Cu2+, Cd2+ and Pb2+), 55.5 and then 1000 [71]. The values of
3.4.4. Diffusion-chemisorption model
ΔH◦ and ΔS◦ of the adsorption process were, respectively, obtained from
In order to provide a good fit of kinetic data and predict the
the slope and intercept of Van’t Hoff equation deduced by plotting
adsorption mechanism, the diffusion-chemisorption kinetic model was
ln Kd versus 1/T, Eq. (12):
used [56]. This empirical kinetic model can be used to describe the
adsorption of the heavy metals onto the adsorbents with heterogeneous ∆S◦ ∆H ◦
lnKd = − (12)
surfaces. The non-linear form of the diffusion-chemisorption model is R RT
expressed as Eq. (8):
The obtained values of the thermodynamic parameters are listed in
qe kDC t1/2 Table 7 and Van’t Hoff plots are presented in Fig. S1. The positive values
qt = (8) of ∆H◦ for Cu2+, Cd2+ and Pb2+ ions reveal that the adsorption process
kDC t1/2 + qe
of three metal ions by PPSM-6 is an endothermic in nature. The value of
where kDC is the diffusion-chemisorption constant (mg/g.min0.5). Fig. 6e ΔS◦ can be used to indicate whether the sorption reaction is ascribed to
exhibits the diffusion-chemisorption model for the sorption of Cu2+, the associative or dissociative mechanism. The positive values of ΔS◦ for
Cd2+ and Pb2+ ions on PPSM-6 and the related parameters are listed in Cu2+, Cd2+ and Pb2+ ions are an indication of increase in the random­
Table 4. As revealed in Fig. 6e and Table 4, for three heavy metals, the R2 ness of the adsorbed ions during the adsorption process. The sponta­
values are 0.99 implying that the curves fit well with the dif­ neous nature of the adsorption process is specified by the negative
fusion–chemisorption model. This outcome suggests that the adsorption values of ΔG◦ at the studied temperature range (Table 7).
process of Cu2+, Cd2+ and Pb2+ ions on PPSM-6 can be well predicted
through diffusion-chemisorption model. As a result, the adsorption of 3.7. Adsorption mechanism
the studied heavy metal ions on PPSM-6 is verified as the heterogeneous
process. Thus, the diffusion-chemisorption model is an effective way to The simultaneous adsorption of ternary setup of metal ions (Cu2+,
predict the trend of adsorption behaviors of Cu2+, Cd2+ and Pb2+ ions on Cd2+
and Pb2+) onto PPSM-6 was investigated before and after
PPSM-6. adsorption using the high-resolution XPS spectroscopy. Before the
adsorption, main elements of C, O and Mn were verified by the XPS
3.5. Adsorption isotherms spectra (Fig. S2a). After the adsorption process, the new apparent peaks
in the XPS spectra of PPSM-6 were assigned to Cu 2p, Cd 3d, Cd 3p, Pb
The equilibrium adsorption data were fitted by the nonlinear forms 4d, and Pb 4f. The double-peaks of Cu 2p3/2 and Cu 2p1/2 (932.85 and
of Langmuir (Eq. (9)) and Freundlich isotherm (Eq. (10)) models [57]: 953.47 eV) (Fig. S2b), Cd 3d5/2 and Cd 3d3/2 (405.29 and 411.94 eV)
qm kL Ce (Fig. S2c), and Pb 4f7/2 and Pb 4f5/2 (139.98 and 142.04 eV) (Fig. S2d)
qe = (9) were seen in the XPS spectra. From these results, the successful simul­
1 + k L Ce
taneous adsorption of Cu2+, Cd2+ and Pb2+ species onto the surface of
qe = kF C1/n
e (10) PPSM-6 adsorbent was confirmed throughout the adsorption process.
The adsorption performance of (PVB-PVA)/SA-ε-MnO2 nano­
where Ce , qe , qm and KL are, respectively, the equilibrium concentrations composite as an adsorbent towards the heavy metal ions can be due to
(mg/L), equilibrium adsorption capacity (mg/g), maximum adsorption the combination of both physical and chemical adsorption mechanisms
capacity (mg/g), Langmuir constant (L/mg), KF (mg/g)(mg/L)− 1/nF is where both (PVB-PVA) blended composite and SA-ε-MnO2 nanoflowers
Freundlich constant and 1/nF is the heterogeneity factor. To gain play a significant role in the adsorption process. As represented in

11
H. Azad et al. Journal of Environmental Chemical Engineering 9 (2021) 106214

Fig. 7. (a) Langmuir and (b) Freundlich adsorption isotherm models for the heavy metal ions adsorption by PPSM-6.

Table 5
Isotherm parameters of PPSM-6 for the adsorption of Cu2+, Cd2+ and Pb2+.
Metal ion qe (exp.) Langmuir Freundlich
(mg/g) 2 − 1/n
qm (mg/g) kL (L/mg) R RMSE kf (mg/g)(mg/L) F n R2 RMSE
2+
Cu 155.02 172.54 0.34 0.96 9.68 58.21 3.02 0.96 9.90

Cd2+ 98.32 112.89 0.17 0.98 4.18 27.30 2.73 0.94 7.37
Pb2+ 87.74 111.23 0.09 0.99 3.20 19.57 2.46 0.90 8.60

interactions between these metal ions and carboxylic groups in the SA


Table 6
chains grafted on ε-MnO2 nanoflowers. Moreover, ε-MnO2 nanoflowers
Comparative adsorption capacity of (PVB-PVA)/SA-ε-MnO2 nanocomposite
with a porous tunnel-like structure expose abundant vacant active sur­
films with other adsorbents for the removal of heavy metal ions.
face sites to Cu2+, Cd2+ and Pb2+ ions and therefore the adsorption
Adsorbent Adsorption capacity (mg/g) Ref. process of the heavy metals occurs through the diffusion and occupation
Cu2þ Cd2þ Pb2þ of these adsorptive sites [43]. The surface complexation and electro­
Powdered activated carbon – 3.37 26.9 [61] static interactions are more appropriate mechanisms for the adsorption
Fe3O4-MnO2 111.90 169.90 208.17 [62] of metal ions because of the presence of negatively deprotonated func­
Dumbbell MnO2/gelatin composites – 89.2 314.5 [63] tional groups on the surface of PPSM-6 in the forms of Sur-COO- and
Iron oxide nanoparticles – 18.59 29 [64] Sur-O- (4<pH<7) (Sur- denotes the surface of PPSM-6). This negatively
OMWCNT/Ppy 24.39 – 26.32 [65]
Fe3O4/cyclodextrin polymer – 27.70 64.50 [66]
charged functional groups can strongly interact with the metal cations,
Activated alumina – 35.07 83.33 [67] leading to a very distinct improvement in the adsorption performance of
Multi-wall carbon nanotubes – 10.86 97.08 [50] the adsorbent towards Cu2+, Cd2+ and Pb2+ ions [17,72].
Modified Al2O3 16.3 83.33 100 [68]
MnFe2O4/biochar composites 94.7 79.9 175.4 [69]
PVA/PVP/MnO2 NC – 47.00 – [70] 3.8. Reusability test
(PVB-PVA)/SA-ε-MnO2 166.94 112.74 110.01 This work
To evaluate the chemical stability and practical performance of the
fabricated PPSM-6 adsorbent, the reusability experiments were carried
Scheme 3, the possible physicochemical mechanisms for the adsorption
out by conducting five successive adsorption cycles for the adsorptive
of the heavy metal ions (Cu2+, Cd2+ and Pb2+) on the surface of (PVB-
removal of Cu2+, Cd2+ and Pb2+ from their aqueous solutions. After the
PVA)/SA-ε-MnO2 nanocomposite may be attributed to the surface
first cycle, the calculated RE% for Cu2+, Cd2+ and Pb2+ ions were
complexation and Lewis acid-base interactions between heavy metal
98.63%, 97.44% and 97.15%, respectively, while the corresponding
ions and hydroxyl functional groups in the backbone of PVB-PVA
values were reduced to 94.50%, 92.05% and 88.60% in the fifth cycle
blended matrix through sharing lone-pair electrons of oxygen atoms.
(Fig. 8). The slight reduction in RE% may be due to the loss of some
Also, a physical adsorption phenomenon can occur through electrostatic
active adsorptive sites during the acid wash of PPSM-6 in the

Table 7
Thermodynamic parameters of Cu2+, Cd2+ and Pb2+ adsorption process on PPSM-6.
parameters ΔH◦ (kJ/mol) ΔS◦ (kJ/mol. K) ΔG◦ (kJ/mol)
2þ 2þ 2þ 2þ 2þ 2þ
Cu Cd Pb Cu Cd Pb T (K) Cu2þ Cd2þ Pb2þ

Values 27.81 124.86 33.41 0.21 0.53 0.23 298 -34.68 -34.37 -34.31
308 -36.32 -42.80 -36.94
318 -38.88 -42.40 -38.86

12
H. Azad et al. Journal of Environmental Chemical Engineering 9 (2021) 106214

Scheme 3. Schematic presentation of physical and chemical adsorption mechanisms of (PVB-PVA)/SA-ε-MnO2 nanocomposite for the adsorption of heavy
metal ions.

regeneration process. It was concluded that (PVB-PVA)/SA-ε-MnO2 the characterization, the prepared PVB-PVA/SA-ε-MnO2 nano­
nanocomposite possessed a reasonably good reusability without any composites were used for adsorptive removal of Pb2+, Cd2+ and Cu2+
significant loss in the adsorption efficiency during the adsorptive heavy metal ions from their aqueous solutions. The batch adsorption
removal of Cu2+, Cd2+ and Pb2+ ions from aqueous solutions. studies showed a good adsorption performance for the PVB-PVA/SA-
ε-MnO2 nanocomposite containing 6 wt% of SA-ε-MnO2 nanoflowers
4. Conclusions (PPSM-6). The simultaneous adsorption of Pb2+, Cd2+ and Cu2+ ions on
PPSM-6 in the binary and ternary solutions indicated the preferential
A new PVB-PVA polymer blend nanocomposite containing different elimination of Cu2+ ions from the aqueous solutions due to its lower
contents (i.e., 3, 6 and 9 wt%) of SA-ε-MnO2 nanoflowers (PVB-PVA/SA- ionic radius and higher electronegativity. Langmuir isotherm model was
ε-MnO2) was synthesized through blending, solution casting and solvent shown to be the most appropriate model for describing the adsorption
evaporation techniques. The XRD results indicated the crystalline process suggesting a homogenous adsorbent with a monolayer coverage
structure of as-synthesized ε-MnO2 and SA-ε-MnO2 nanoflowers. The FE- of the metal ions. The maximum adsorption capacity (qmax) was found as
SEM images revealed a uniform dispersion of SA-ε-MnO2 nanoflowers 172.54, 112.89 and 111.23 mg/g for Cu2+, Cd2+ and Pb2+ ions,
throughout the PVB-PVA polymer blend. The chemical structure and respectively. Further thermodynamic study confirmed the spontaneous
large specific surface area of the prepared polymer blend nano­ and endothermic nature of Cu2+, Cd2+ and Pb2+ adsorption on PPSM-6
composites were respectively confirmed by FTIR and BET analyses. After with a higher randomness at the solid-solution interface. In addition,
after five successive adsorption cycles, the fabricated PPSM-6 showed
the removal efficiency of 94.50%, 92.05% and 88.60% for Cu2+, Cd2+
and Pb2+ ions, respectively, implying a very good reusability for the
adsorptive removal of heavy metal ions. The confirmed good adsorption
performance of PVB-PVA/SA-ε-MnO2 nanocomposite towards the metal
ions by XPS spectra was attributed to the combination of both physical
and chemical adsorption mechanisms, through the surface complexa­
tion, Lewis acid-base, and electrostatic interactions. The outcomes of
this study verify that the fabricated PPSM-6 nanocomposite can be uti­
lized as a potential adsorbent for highly efficient adsorptive removal of
heavy metal ions from industrial wastewater.

CRediT authorship contribution statement

Hajar Azad: Conceptualization, Methodology, Experiments, Soft­


ware, Resources. Mohsen Mohsennia: Supervision, Data curation,
Validation, Analysis, Editing and revising. Chun Cheng: Visualization,
Investigation, Conceptualization. Abbas Amini: Supervision, Analysis,
Validation, Justification, Editing and revising.

Declaration of Competing Interest

Fig. 8. Obtained RE% after five successive adsorption cycles by the prepared The authors declare that they have no known competing financial
PPSM-6. interests or personal relationships that could have appeared to influence

13
H. Azad et al. Journal of Environmental Chemical Engineering 9 (2021) 106214

the work reported in this paper. [18] U. Habiba, A.M. Afifi, A. Salleh, B.C. Ang, Chitosan/( polyvinyl alcohol)/zeolite
electrospun composite nanofibrous membrane for adsorption of Cr6+, Fe3+ and Ni2
+
, J. Hazard. Mater. 322 (2017) 182–194, https://doi.org/10.1016/j.
Acknowledgments jhazmat.2016.06.028.
[19] F.L. Long, C.G. Niu, N. Tang, H. Guo, Z.W. Li, Y.Y. Yang, L.S. Lin, Highly efficient
removal of hexavalent chromium from aqueous solution by calcined Mg/Al-layered
The authors gratefully acknowledge to University of Kashan, Iran for double hydroxides/polyaniline composites, Chem. Eng. J. 404 (2021), 127084,
partial financial support for this research study. BioRender is grateful for https://doi.org/10.1016/j.cej.2020.127084.
helping in illustrations. [20] W. Yu, J. Hu, Y. Yu, D. Ma, W. Gong, H. Qiu, Z. Hu, H. wen Gao, Facile preparation
of sulfonated biochar for highly efficient removal of toxic Pb(II) and Cd(II) from
wastewater, Sci. Total Environ. 750 (2021), 141545, https://doi.org/10.1016/j.
Appendix A. Supporting information scitotenv.2020.141545.
[21] R. Chen, T. Sheehan, J.L. Ng, M. Brucks, X. Su, Capacitive deionization and
electrosorption for heavy metal removal, Environ. Sci. Water Res. Technol. 6
Supplementary data associated with this article can be found in the (2019) 258–282, https://doi.org/10.1039/c9ew00945k.
online version at doi:10.1016/j.jece.2021.106214. [22] G. Zhao, J. Li, X. Ren, C. Chen, X. Wang, Few-layered graphene oxide nanosheets as
superior sorbents for heavy metal ion pollution management, Environ. Sci.
Technol. 45 (2011) 10454–10462, https://doi.org/10.1021/es203439v.
References [23] P. Zito, H.J. Shipley, Inorganic nano-adsorbents for the removal of heavy metals
and arsenic: a review, RSC Adv. 5 (2015) 29885–29907, https://doi.org/10.1039/
C5RA02714D.
[1] S. Amiri, A. Asghari, V. Vatanpour, M. Rajabi, Fabrication and characterization of a
[24] L.P. Lingamdinne, K.R. Vemula, Y.Y. Chang, J.K. Yang, R.R. Karri, J.R. Koduru,
novel polyvinyl alcohol-graphene oxide-sodium alginate nanocomposite hydrogel
Process optimization and modeling of lead removal using iron oxide
blended PES nanofiltration membrane for improved water purification, Sep. Purif.
nanocomposites generated from bio-waste mass, Chemosphere 243 (2020),
Technol. 250 (2020), 117216, https://doi.org/10.1016/j.seppur.2020.117216.
125257, https://doi.org/10.1016/j.chemosphere.2019.125257.
[2] G.K.R. Angaru, Y.L. Choi, L.P. Lingamdinne, J.S. Choi, D.S. Kim, J.R. Koduru, J.
[25] H. Cui, Q. Li, S. Gao, J.K. Shang, Strong adsorption of arsenic species by amorphous
K. Yang, Y.Y. Chang, Facile synthesis of economical feasible fly ash–based
zirconium oxide nanoparticles, J. Ind. Eng. Chem. 18 (2012) 1418–1427, https://
zeolite–supported nano zerovalent iron and nickel bimetallic composite for the
doi.org/10.1016/j.jiec.2012.01.045.
potential removal of heavy metals from industrial effluents, Chemosphere 267
[26] S. Recillas, J. Colón, E. Casals, E. González, V. Puntes, A. Sánchez, X. Font,
(2021), 128889, https://doi.org/10.1016/j.chemosphere.2020.128889.
Chromium VI adsorption on cerium oxide nanoparticles and morphology changes
[3] A.L.D. da Rosa, E. Carissimi, G.L. Dotto, H. Sander, L.A. Feris, Biosorption of
during the process, J. Hazard. Mater. 184 (2010) 425–431, https://doi.org/
rhodamine B dye from dyeing stones effluents using the green microalgae Chlorella
10.1016/j.jhazmat.2010.08.052.
pyrenoidosa, J. Clean. Prod. 198 (2018) 1302–1310, https://doi.org/10.1016/j.
[27] D. Zhu, Y. Chen, H. Yang, S. Wang, X. Wang, S. Zhang, H. Chen, Synthesis and
jclepro.2018.07.128.
characterization of magnesium oxide nanoparticle-containing biochar composites
[4] L.N. Côrtes, S.P. Druzian, A.F.M. Streit, T.R. Sant’anna Cadaval Junior, G.
for efficient phosphorus removal from aqueous solution, Chemosphere 247 (2020),
C. Collazzo, G.L. Dotto, Preparation of carbonaceous materials from pyrolysis of
125847, https://doi.org/10.1016/j.chemosphere.2020.125847.
chicken bones and its application for fuchsine adsorption, Environ. Sci. Pollut. Res.
[28] M. Nazari, A. Amini, N.T. Eden, M.C. Duke, C. Cheng, M.R. Hill, Highly-efficient
26 (2019) 28574–28583, https://doi.org/10.1007/s11356-018-3679-2.
sulfonated UiO-66 (Zr) optical fiber for rapid detection of trace levels of Pb2+, Int.
[5] G.L. Dotto, G. McKay, Current scenario and challenges in adsorption for water
J. Mol. Sci. 22 (2021) 6053, https://doi.org/10.3390/ijms22116053.
treatment, J. Environ. Chem. Eng. 8 (2020), 103988, https://doi.org/10.1016/j.
[29] P. Wei, H. Lou, X. Xu, W. Xu, H. Yang, W. Zhang, Y. Zhang, Preparation of PP non-
jece.2020.103988.
woven fabric with good heavy metal adsorption performance via plasma
[6] Q.U. Ain, M.U. Farooq, M.I. Jalees, Application of magnetic graphene oxide for
modification and graft polymerization, Appl. Surf. Sci. 539 (2021), 148195,
water purification: heavy metals removal and disinfection, J. Water Process Eng.
https://doi.org/10.1016/j.apsusc.2020.148195.
33 (2020), 101044, https://doi.org/10.1016/j.jwpe.2019.101044.
[30] S.K. Pradhan, V. Pareek, J. Panwar, S. Gupta, Synthesis and characterization of
[7] H. Zhao, F. Song, F. Su, Y. Shen, P. Li, Removal of cadmium from contaminated
ecofriendly silver nanoparticles combined with yttrium oxide (Ag-Y2O3)
groundwater using a novel silicon/aluminum nanomaterial: an experimental study,
nanocomposite with assorted adsorption capacity for Cu(II) and Cr(VI) removal: a
Arch. Environ. Contam. Toxicol. 80 (2021) 234–247, https://doi.org/10.1007/
mechanism perspective, J. Water Process Eng. 32 (2019), 100917, https://doi.org/
s00244-020-00784-1.
10.1016/j.jwpe.2019.100917.
[8] M. Rahimi, A. Amini, H. Behmadi, Novel symmetric Schiff-base benzobisthiazole-
[31] F. Niu, X. Wang, G. Zhao, X. Huang, Z. Tang, Q. Huang, Polymer-based
salicylidene derivative with fluorescence turn-on behavior for detecting Pb2+ ion,
nanocomposites for heavy metal ions removal from aqueous solution: a review, in:
J. Photochem. Photobiol. A: Chem. 388 (2020), 112190, https://doi.org/10.1016/
RSC Polym. Chem., 2018, pp. 3562–3582, https://doi.org/10.1039/C8PY00484F.
j.jphotochem.2019.112190.
[32] V. Srivastava, E. Nazarzadeh, P. Makvandi, X. Zheng, S. Iftekhar, A. Wu, V.V.
[9] E. Nazarzadeh, A. Motahari, M. Sillanpää, Nanoadsorbents based on conducting
T. Padil, B. Mokhtari, R.S. Varma, F.R. Tay, M. Sillanpaa, Cytotoxic aquatic
polymer nanocomposites with main focus on polyaniline and its derivatives for
pollutants and their removal by nanocomposite-based sorbents, Chemosphere 258
removal of heavy metal ions / dyes: a review, Environ. Res. 162 (2018) 173–195,
(2020), 127324, https://doi.org/10.1016/j.chemosphere.2020.127324.
https://doi.org/10.1016/j.envres.2017.12.025.
[33] W. Liu, J. Zhou, L. Ding, Y. Yang, T. Zhang, MIL-88/PVB nanofiber as recyclable
[10] H. Gomaa, M.A. Shenashen, A. Elbaz, H. Yamaguchi, M. Abdelmottaleb, S.A. El-
heterogeneous catalyst for photocatalytic and Fenton process under visible light
Safty, Mesoscopic engineering materials for visual detection and selective removal
irradiation, Chem. Phys. Lett. 749 (2020), 137431, https://doi.org/10.1016/j.
of copper ions from drinking and waste water sources, J. Hazard. Mater. 406
cplett.2020.137431.
(2021), 124314, https://doi.org/10.1016/j.jhazmat.2020.124314.
[34] J. Ma, T. Xiao, N. Long, X. Yang, The role of polyvinyl butyral additive in forming
[11] W. Qin, J. Li, J. Tu, H. Yang, Q. Chen, H. Liu, Fabrication of porous chitosan
desirable pore structure for thin film composite forward osmosis membrane, Sep.
membranes composed of nanofibers by low temperature thermally induced phase
Purif. Technol. 242 (2020), 116798, https://doi.org/10.1016/j.
separation, and their adsorption behavior for Cu2+, Carbohydr. Polym. 178 (2017)
seppur.2020.116798.
338–346, https://doi.org/10.1016/j.carbpol.2017.09.051.
[35] Y.R. Qiu, N.A. Rahman, H. Matsuyama, Preparation of hydrophilic poly(vinyl
[12] M.M. Alam, Z.A. ALOthman, M. Naushad, T. Aouak, Evaluation of heavy metal
butyral)/Pluronic F127 blend hollow fiber membrane via thermally induced phase
kinetics through pyridine based Th(IV) phosphate composite cation exchanger
separation, Sep. Purif. Technol. 61 (2008) 1–8, https://doi.org/10.1016/j.
using particle diffusion controlled ion exchange phenomenon, J. Ind. Eng. Chem.
seppur.2007.09.014.
20 (2014) 705–709, https://doi.org/10.1016/j.jiec.2013.05.036.
[36] P. He, F. Wu, M. Yang, W. Jiao, X. Yin, Y. Si, J. Yu, B. Ding, Green and
[13] A.M. Nasir, P.S. Goh, M.S. Abdullah, B.C. Ng, A.F. Ismail, Adsorptive
antimicrobial 5-bromosalicylic acid/polyvinyl butyral nanofibrous membranes
nanocomposite membranes for heavy metal remediation: Recent progresses and
enable interception-sterilization-integrated bioprotection, Compos. Commun. 25
challenges, Chemosphere 232 (2019) 96–112, https://doi.org/10.1016/j.
(2021), 100720, https://doi.org/10.1016/j.coco.2021.100720.
chemosphere.2019.05.174.
[37] H.T. Kahraman, A. Avcı, E. Pehlivan, Novel sandwiched composite electro-spun
[14] Q. Feng, D. Wu, Y. Zhao, A. Wei, Q. Wei, H. Fong, Electrospun AOPAN/RC blend
mats based on polyacrylonitrile and polyvinyl butyral for fast oil–water separation,
nanofiber membrane for efficient removal of heavy metal ions from water,
Iran. Polym. J. (Engl. Ed. 28 (2019) 445–453, https://doi.org/10.1007/s13726-
J. Hazard. Mater. 344 (2018) 819–828, https://doi.org/10.1016/j.
019-00713-7.
jhazmat.2017.11.035.
[38] Z. Cui, W. Li, H. Zeng, X. Tang, J. Zhang, S. Qin, N. Han, J. Li, Fabricating PVDF
[15] S.Z.N. Ahmad, W.N. Wan Salleh, A.F. Ismail, N. Yusof, M.Z. Mohd Yusop, F. Aziz,
hollow fiber microfiltration membrane with a tenon-connection structure via the
Adsorptive removal of heavy metal ions using graphene-based nanomaterials:
thermally induced phase separation process to enhance strength and permeability,
toxicity, roles of functional groups and mechanisms, Chemosphere 248 (2020),
Eur. Polym. J. 111 (2019) 49–62, https://doi.org/10.1016/j.
126008, https://doi.org/10.1016/j.chemosphere.2020.126008.
eurpolymj.2018.12.009.
[16] H. Alyasi, H.R. Mackey, K. Loganathan, G. McKay, Adsorbent minimisation in a
[39] B. Yang, R. Liu, J. Huang, H. Sun, Reverse dissolution as a route in the synthesis of
two-stage batch adsorber for cadmium removal, J. Ind. Eng. Chem. 81 (2020)
poly ( vinyl butyral) with high butyral contents, ACS Ind. Eng. Chem. Res. 52
153–160, https://doi.org/10.1016/j.jiec.2019.09.003.
(2013) 7425–7431.
[17] H. Azad, M. Mohsennia, A novel free-standing polyvinyl butyral-polyacrylonitrile/
[40] X. Wang, T. Wang, J. Ma, H. Liu, P. Ning, Synthesis and characterization of a new
ZnAl-layered double hydroxide nanocomposite membrane for enhanced heavy
hydrophilic boehmite-PVB/PVDF blended membrane supported nano zero-valent
metal removal from wastewater, J. Membr. Sci. 615 (2020), 118487, https://doi.
org/10.1016/j.memsci.2020.118487.

14
H. Azad et al. Journal of Environmental Chemical Engineering 9 (2021) 106214

iron for removal of Cr(VI), Sep. Purif. Technol. 205 (2018) 74–83, https://doi.org/ adsorbent, Carbohydr. Polym. 131 (2015) 255–263, https://doi.org/10.1016/j.
10.1016/j.seppur.2018.05.010. carbpol.2015.06.031.
[41] H.T. Kahraman, E. Pehlivan, A. Avcı, Preparation and adsorption behavior of novel [58] A.s. Abdulhameed, N.N.M. Firdaus Hum, S. Rangabhashiyam, A.H. Jawad, L.
sandwiched composite electro-spun aminated membrane for hexavalent chromium D. Wilson, Z.M. Yaseen, A.A. Al-Kahtani, Z.A. Alothman, Statistical modeling and
removal, Int. J. Chem. Eng. Appl. 9 (2018) 180–183, https://doi.org/10.18178/ mechanistic pathway for methylene blue dye removal by high surface area and
ijcea.2018.9.5.723. mesoporous grass-based activated carbon using K2CO3 activator, J. Environ. Chem.
[42] N.A. Syed, M. Husnain, U. Asim, A. Yaqub, F. Shahzad, Recent trends of MnO2- Eng. 9 (2021) 105530, https://doi.org/10.1016/j.jece.2021.105530.
derived adsorbents for water treatment: a review, N. J. Chem. 44 (2020) [59] A.H. Jawad, N.S.A. Mubarak, A.S. Abdulhameed, Hybrid crosslinkedchitosan-
6096–6120, https://doi.org/10.1039/C9NJ06392G.Volume. epichlorohydrin/TiO2 nanocomposite for reactive red 120 dye adsorption: Kinetic,
[43] M. Lin, Z. Chen, A facile one-step synthesized epsilon-MnO2 nanoflowers for isotherm, thermodynamic, and mechanism Study, J. Polym. Environ. 28 (2020)
effective removal of lead ions from wastewater, Chemosphere 250 (2020), 126329, 624–637, https://doi.org/10.1007/s10924-019-01631-8.
https://doi.org/10.1016/j.chemosphere.2020.126329. [60] A.H. Jawad, A.S. Abdulhameed, N.N.A. Malek, Z.A.A.L. Othman, Statistical
[44] S. Mallakpour, F. Motirasoul, Bio-functionalizing of a-MnO2 nanorods with natural optimization and modeling for color removal and COD reduction of reactive blue
L-amino acids: a favorable adsorbent for the removal of Cd (II) ions, Mater. Chem. 19 dye by mesoporous chitosan-epichlorohydrin/kaolin clay composite, Int. J. Biol.
Phys. 191 (2017) 188–196, https://doi.org/10.1016/j.matchemphys.2017.01.040. Macromol. 164 (2020) 4218–4230, https://doi.org/10.1016/j.
[45] A. Hazarika, B.K. Deka, K. Kong, D. Kim, Y. Nam, J. Choi, C. Kim, Y.-B. Park, H. ijbiomac.2020.08.201.
W. Park, Microwave absorption and mechanical performance of α-MnO2 [61] Z. Reddad, C. Gerente, Y. Andres, P.L.E. Cloirec, Adsorption of several metal ions
nanostructures grown on woven Kevlar fiber/reduced graphene oxide-polyaniline onto a low-cost biosorbent: kinetic and equilibrium studies, Environ. Sci. Technol.
nanofiber array-reinforced polyester resin composites, Compos. Part B Eng. 140 36 (2002) 2067–2073, https://doi.org/10.1021/es0102989.
(2018) 123–132, https://doi.org/10.1016/j.compositesb.2017.12.003. [62] J. Zhao, J. Liu, N. Li, W. Wang, J. Nan, Z. Zhao, F. Cui, Highly efficient removal of
[46] G.L. Dotto, L.A.A. Pinto, Adsorption of food dyes onto chitosan: optimization bivalent heavy metals from aqueous systems by magnetic porous Fe3O4-MnO2:
process and kinetic, Carbohydr. Polym. 84 (2011) 231–238, https://doi.org/ adsorption behavior and, Chem. Eng. J. 304 (2016) 737–746, https://doi.org/
10.1016/j.carbpol.2010.11.028. 10.1016/j.cej.2016.07.003.
[47] S. Chakraborty, S. Chowdhury, P. Das Saha, C. Violet, Adsorption of crystal violet [63] X. Wang, K. Huang, Y. Chen, J. Liu, S. Chen, J. Cao, S. Mei, Preparation of
from aqueous solution onto NaOH-modified rice husk, Carbohydr. Polym. 86 dumbbell manganese dioxide/gelatin composites and their application in the
(2011) 1533–1541, https://doi.org/10.1016/j.carbpol.2011.06.058. removal of lead and cadmium ions, J. Hazard. Mater. 350 (2018) 46–54, https://
[48] W. Liu, T. Wang, A.G.L. Borthwick, Y. Wang, X. Yin, X. Li, J. Ni, Adsorption of Pb2 doi.org/10.1016/j.jhazmat.2018.02.020.
+
, Cd2+, Cu2+ and Cr3+ onto titanate nanotubes: Competition and effect of [64] N.N. Nassar, Rapid removal and recovery of Pb (II) from wastewater by magnetic
inorganic ions, Sci. Total Environ. 456-457 (2013) 171–180, https://doi.org/ nanoadsorbents, J. Hazard. Mater. 184 (2010) 538–546, https://doi.org/10.1016/
10.1016/j.scitotenv.2013.03.082. j.jhazmat.2010.08.069.
[49] X. Tao, Q. Feng, H. He, Preparation of calixarene-PI nanofibers and application as a [65] W.N. Nyairo, Y.R. Eker, C. Kowenje, H. Bingol, A. Tor, D.M. Ongeri, Efficient
selective adsorbent for heavy metal ions, J. Eng. Fiber Fabr. 13 (2018), https://doi. adsorption of lead (II) and copper (II) from aqueous phase using oxidized
org/10.1177/155892501801300101. multiwalled carbon nanotubes/polypyrrole composite, Sep. Sci. Technol. 53
[50] Y. Li, J. Ding, Z. Luan, Z. Di, Y. Zhu, C. Xu, D. Wu, B. Wei, Competitive adsorption (2018) 1–13, https://doi.org/10.1080/01496395.2018.1424203.
of Pb2+, Cu2+ and Cd2+ ions from aqueous solutions by multiwalled carbon [66] A. Zayed, M. Badruddoza, Z. Bin, Z. Shawon, T. Wei, J. Daniel, K. Hidajat,
nanotubes, Carbon 41 (2003) 2787–2792, https://doi.org/10.1016/S0008-6223 M. Shahab, Fe3O4/cyclodextrin polymer nanocomposites for selective heavy metals
(03)00392-0. removal from industrial wastewater, Carbohydr. Polym. 91 (2013) 322–332,
[51] J. Goel, K. Kadirvelu, C. Rajagopal, E. Safety, Competitive sorption of Cu (II), Pb https://doi.org/10.1016/j.carbpol.2012.08.030.
(II) and Hg (II) ions from aqueous solution using coconut shell-based activated, [67] T.K. Naiya, A.K. Bhattacharya, S.K. Das, Adsorption of Cd (II) and Pb (II) from
Carbon 22 (2004) 257–273, https://doi.org/10.1260/0263617041503453. aqueous solutions on activated alumina, J. Colloid Interface Sci. 333 (2009) 14–26,
[52] J. Simonin, On the comparison of pseudo-first order and pseudo-second order rate https://doi.org/10.1016/j.jcis.2009.01.003.
laws in the modeling of adsorption kinetics, Chem. Eng. J. 300 (2016) 254–263, [68] A. Afkhami, M. Saber-tehrani, H. Bagheri, Simultaneous removal of heavy-metal
https://doi.org/10.1016/j.cej.2016.04.079. ions in wastewater samples using nano-alumina modified with 2,4-
[53] A.H. Jawad, A.S. Abdulhameed, M.S. Mastuli, Mesoporous crosslinked chitosan- dinitrophenylhydrazine, J. Hazard. Mater. 181 (2010) 836–844, https://doi.org/
activated charcoal composite for the removal of thionine cationic dye: 10.1016/j.jhazmat.2010.05.089.
comprehensive adsorption and mechanism study, J. Polym. Environ. 28 (2020) [69] K. Jung, S.Y. Lee, Y.J. Lee, Facile one-pot hydrothermal synthesis of cubic spinel-
1095–1105, https://doi.org/10.1007/s10924-020-01671-5. type manganese ferrite/biochar composites for environmental remediation of
[54] A.H. Jawad, N.S.A. Mubarak, A.S. Abdulhameed, Tunable Schiff’s base-cross- heavy metals from aqueous solutions, Bioresour. Technol. 261 (2018) 1–9, https://
linked chitosan composite for the removal of reactive red 120 dye: adsorption and doi.org/10.1016/j.biortech.2018.04.003.
mechanism study, Int. J. Biol. Macromol. 142 (2020) 732–741, https://doi.org/ [70] S. Mallakpour, F. Motirasoul, Ultrasonication synthesis of PVA/PVP/α-MnO2-
10.1016/j.ijbiomac.2019.10.014. stearic acid blend nanocomposites for adsorbing CdII ion, Ultrason. Sonochem. 40
[55] A. Reghioua, D. Barkat, A.H. Jawad, A.S. Abdulhameed, A.A. Al-Kahtani, Z. (2018) 410–418, https://doi.org/10.1016/j.ultsonch.2017.07.034.
A. Alothman, Parametric optimization by Box-Behnken design for synthesis of [71] H.N. Tran, S.J. You, H.P. Chao, Thermodynamic parameters of cadmium
magnetic chitosan-benzil/ZnO/Fe3O4 nanocomposite and textile dye removal, adsorption onto orange peel calculated from various methods: a comparison study,
J. Environ. Chem. Eng. 9 (2021), 105166, https://doi.org/10.1016/j. J. Environ. Chem. Eng. 4 (2016) 2671–2682, https://doi.org/10.1016/j.
jece.2021.105166. jece.2016.05.009.
[56] C. Sutherland, Removal of heavy metals from water using low-cost adsorbents: [72] C. Zheng, Q. Wu, X. Hu, Y. Wang, Y. Chen, S. Zhang, H. Zheng, Adsorption
process development, Dr. Diss (2009). behavior of heavy metal ions on a polymer-immobilized amphoteric biosorbent:
[57] F. Mi, S. Wu, Y. Chen, Combination of carboxymethyl chitosan-coated magnetic surface interaction assessment, J. Hazard. Mater. 403 (2021), 123801, https://doi.
nanoparticles and chitosan-citrate complex gel beads as a novel magnetic org/10.1016/j.jhazmat.2020.123801.

15

You might also like