SlideShare a Scribd company logo
Citation: Boutaleb, Y.; Zerdoum, R.;
Bensid, N.; Abumousa, R.A.; Hattab,
Z.; Bououdina, M. Adsorption of
Cr(VI) by Mesoporous Pomegranate
Peel Biowaste from Synthetic
Wastewater under Dynamic Mode.
Water 2022, 14, 3885. https://
doi.org/10.3390/w14233885
Academic Editor: Dipti
Prakash Mohapatra
Received: 29 October 2022
Accepted: 22 November 2022
Published: 28 November 2022
Publisher’s Note: MDPI stays neutral
with regard to jurisdictional claims in
published maps and institutional affil-
iations.
Copyright: © 2022 by the authors.
Licensee MDPI, Basel, Switzerland.
This article is an open access article
distributed under the terms and
conditions of the Creative Commons
Attribution (CC BY) license (https://
creativecommons.org/licenses/by/
4.0/).
water
Article
Adsorption of Cr(VI) by Mesoporous Pomegranate Peel
Biowaste from Synthetic Wastewater under Dynamic Mode
Yassira Boutaleb 1, Radia Zerdoum 2, Nadia Bensid 1, Rasha A. Abumousa 3,* , Zhour Hattab 1
and Mohamed Bououdina 3
1 Laboratory of Water Treatment and Valorisation of Industrial Wastes, Department of Chemistry, Faculty of
Sciences, Badji Mokhtar University, B.P.12, Annaba 23000, Algeria
2 Science and Technology Laboratory of Water and Environment, Department of Material Sciences, Faculty of
Science and Technology, University Mohammed-Cherif Messadia, Souk Ahras 41000, Algeria
3 Department of Mathematics and Sciences, College of Humanities and Sciences, Prince Sultan University,
Riyadh 11586, Saudi Arabia
* Correspondence: rabumousa@psu.edu.sa
Abstract: This study aims to eliminate hexavalent chromium Cr(VI) ions from water using pomegranate
peel (PGP) powder. Dynamic measurements are carried out to examine the influence of the operating
factors on the adsorption efficiency and kinetics. The analyzed PGP is found to be amorphous with
relatively high stability, contains hydroxyl and carboxyl functional groups, a pH of zero charge of
3.9, and a specific surface-area of 40.38 m2/g. Adsorption tests indicate that PGP exhibits excel-
lent removal effectiveness for Cr(VI) reaching 50.32 mg/g while the adsorption process obeys the
Freundlich model. The thermodynamic study favors the exothermic physical adsorption process.
The influence of operating parameters like the flow rate (1 to 3 mL/min), bed height (25 to 75 mm),
concentration (10 to 30 mg/L), and temperature (298 to 318 K) on the adsorption process are investi-
gated in column mode. To assess the performance characteristics of the column adsorption data, a
non-linear regression has been used to fit and analyze four different kinetic and theoretical models,
namely, Bohart-Adams, Thomas model, Clark, and Dose response. The obtained experimental results
were found to obey the Dose Response model with a coefficient of regression R2 greater than 0.977.
This study proved the excellent efficiency in the treatment of chemical industry effluents by using
cost-effect abundant biowaste sorbent. This research demonstrated great efficacy in the treatment of
chemical industrial effluents by using an abundant, cost-effective biowaste sorbent, thereby achieving
the UN SDGs (UN Sustainable Development Goals) primary objective.
Keywords: pomegranate peel; adsorption; fixed bed; batch; modeling
1. Introduction
The contamination of wastewater by chromium Cr(VI) is a major problem due to its
high toxicity and harmful effects on humans and animals [1]. The major effluents containing
such toxic and hazardous element arises from industrial processes such as metallurgy,
dyeing, and finishing industries [2]. Chromium exists in two major forms: Cr(III) and
Cr(VI). The first form has low solubility and low reactivity resulting in low mobility in the
environment and low toxicity in living organisms. However, Cr(VI) is a highly toxic [3]
and carcinogenic [4] heavy metal when its concentration exceeds the standard limit fixed
at 0.05 mg/L [5]. Its compounds include chromate (CrO4
2−), dichromate (Cr2O7
2−) and
chromic acid (H2CrO4). CrO4
2− is predominant in basic solutions, H2CrO4 is predominant
at pH < 1, while HCrO4
2−, and Cr2O7
2− are predominant at pH 2–6 [6].
Extensive studies were devoted toward the development of new cost-effective tech-
nologies for the removal of Cr(VI) metal ions from aquatic environments, like precipita-
tion [7], ion exchange [8], electro-coagulation [9], and adsorption [10–14]. Nevertheless,
most of the technique’s present drawbacks and disadvantages, primarily high energy
Water 2022, 14, 3885. https://doi.org/10.3390/w14233885 https://www.mdpi.com/journal/water
Water 2022, 14, 3885 2 of 20
consumption and cost. In contrast, adsorption is an eco-friendly, cost-effective, and widely
used technique for heavy metals remediation from aqueous solutions, which meets the UN
SDGs (UN Sustainable Development Goals) main aim [15–17].
In recent decades, green chemistry has emerged as a potential fabrication route and
attracted considerable attention in scientific research. Specifically, fruit wastes are being
investigated as potential natural sorbents for heavy metal ions, thanks to their biodegrad-
ability and cost-effectiveness, besides the presence of functional groups on adsorbent
particles’ surface that are responsible for the reduction and elimination of contaminants
and pollutants from aqueous media within a ‘sustainable development’ framework [18,19].
Valency, concentration, the ionic nature of substances in solution, pH of solution, and
binding groups of sorbent material are key sorption parameters. The removal mechanism
relies on the sorbate molecule’s size, concentration, affinity for the sorbent, bulk diffusion
coefficient, and pore size distribution [20,21]. Sorbent’s physicochemical characteristics and
sorption capacity determine the sorption’s efficacy.
Pomegranates (Punica granatum) are widely cultivated in north/tropical Africa, and
the Middle East, the Mediterranean Basin, the Caucasus region, the Indian subcontinent,
Central Asia, and the drier parts of Southeast Asia [22].
Punica granatum presents a great variety of therapeutic bioactive species such as
ellagitannins, proanthocyanidin compounds, phenolics, and flavonoids [23], which are
well-known to facilitate the adsorption process of metal ions. Several studies have been
conducted to highlight the biological and functional characteristics of pomegranate peel
as it contains large amounts of polyphenols such as ellagic acid, gallic acid, and tan-
nin acid [24]. Also, it possesses two important hydroxycinnamic acids and derivatives
of flavones [25]. It has been widely studied as an antioxidant [26], antimicrobial [27],
preparation of nanoparticles [28], anticancer activities [24], and extraction of bioactive
compounds [29]. Additionally, several works have proposed the use of low-cost wastes for
water and effluent treatment [30], particularly for the removal and biosorption of dyes [31],
uranium [32], and heavy toxic metal ions especially chromium [33,34].
Extensive studies were performed on the effectiveness of biomass and its regenera-
tion for Cr(VI) ions removal using pomegranates peel. Salam and Narayanan (2019) [19]
reported a maximum removal of 100% achieved in 3 min for a low concentration of 20
mg/L and an amount of biosorbent (RPP with 125 µm size) of 0.5 g/L. Yi et al. (2017) [35]
studied a modified Litchi peel as a biosorbent to eliminate Cr(VI), but the achieved capacity
is relatively low reaching only 7.05 mg/g for a longer contact time of 100 min at 303 K for a
concentration of 30 mg/g and an amount of biosorbent was 8 g/L. Rafiaee et al. (2020) [36]
used pomegranate peel (PGP) adsorbent powder functionalized by polymeric coatings
such as polypyrrole (PPy) and polyaniline (PANI) to reduce/remove Cr(VI)metal ions from
wastewater. The highest achieved capacity was 59.47% in 90 min obtained in the batch sys-
tem for 50 mg/L and biosorbent dose of 10 g/L. A new composite Anthracite/pomegranate
peel synthesized by Seliem et al. exhibited a relatively high efficiency of 275 mg/g but for a
longer time of 120 min at 25 ◦C and pH 3, the concentration of 70 mg/L and an amount of
biosorbent 0.05 g.
In this study, pomegranate peel (PGP) adsorbent has been prepared by low-cost
and eco-friendly physical and chemical processes. Structural, morphological, and physic-
ochemical studies are investigated. This research aims to study the surface chemistry
characteristics and evaluate the performance of this biomaterial for the elimination of toxic
and hazardous heavy metal ions. Indeed, particular emphasis is devoted to investigate the
adsorption of hexavalent chromium ions Cr(VI) from an aqueous solution under dynamic
mode rather than batch. Several operating parameters are examined, including the flow
rate, the solution pH, adsorbent dose, contact time, bed height, as well as the solution
temperature. To confirm the experimental data, several models have been applied for both
batch and dynamic measurements, subsequently a plausible mechanism is proposed.
Water 2022, 14, 3885 3 of 20
2. Materials and Methods
The pomegranate (Punica granatum) peels (PGP) were collected from fresh juice ven-
dors at a local market. To remove the impurities, the collected peels were thoroughly
washed with tap water and rinsed using distilled water. Afterward, the obtained PGP
product was dried at 100 ◦C for 24 h, grinded into a powdered sample then sieved to a
particle size <500 µm.
The used chemicals were purchased from Merck Company. A stock solution of
1000 mg/L was prepared using potassium dichromate K2Cr2O7, then different concentra-
tions of chromium solution were obtained by consecutive dilution. Adsorption experi-
ments in batch and dynamic modes were performed using a UV-vis spectrophotometer
JENWAL 7315 (England) at maximum absorption of 540 nm after complexation with
1,5-diphenylcarbazide.
2.1. Characterization of the Powdered PGP
To evaluate the morphology and chemical composition of PGP powder, a Quanta 200
FEI scanning electron microscope (SEM) coupled with energy-dispersive X-ray spectroscopy
(EDX) was used. Fourier-transform infrared (FTIR) spectroscopy was used to identify
the functional groups of PGP. FTIR spectrometer (IR Affinity-1S) equipped with a single
reflection ATR, operating in the wave number range 400–4000 cm−1 acquired over 50 scans
with a resolution of 4 cm−1. The phase stability of PGP was checked by X-ray diffraction
(XRD) using a Rigaku Ultima IV instrument equipped with a CPS detector and CuKα
radiation (λ = 1.5418 Å). A Thermogravimetric analyzer TGA-2050 equipment was used to
analyze the thermal degradation profile of the PGP adsorbent up to 600 ◦C with a heating
rate of 10 ◦C/min. The PGP the Brunauer Emmett Teller (BET) specific surface area and the
average pore size were computed by recording Nitrogen adsorption-desorption curves at
70 K using a Quanta chrome NOVA 2200 E BET analyzer.
2.2. Biosorption of Cr(VI) on PGP—Batch and Dynamic Experiments
2.2.1. Batch Studies
The adsorption isotherms were performed at varying Cr(IV) concentrations and the
following operating conditions: m = 1.0 g/L, pH = 2 ± 0.2, T = 20 ◦C, stirring speed 50 rpm,
contact time 2 h. After equilibrium, the adsorbent was separated from the suspension by a
membrane filter (0.45 µm).
The concentration of PGP was determined by analyzing the supernatant. The amount
of the adsorbate, qe (mg/g), was estimated using Equation (1):
qe =
C0 − Ce
m
× V (1)
where C0/Ce represents the initial/equilibrium concentrations of Cr (VI) (mg/L), V is the
solution volume (L), and m is the PGP sorbent mass(g).
Langmuir (Equation (2)) and Freundlich (Equation (3)) models were used to fit the
equilibrium experimental adsorption data by means of the following expressions:
qe =
qmKLCe
1 + KLCe
(2)
qe = KfCe
1
n (3)
where qe is the amount of Cr(VI) adsorbed onto PGP at equilibrium (mg/g), Ce is the Cr(VI)
concentration in solution (mg/g), qm is the PGP capacity of the adsorbed Cr(VI) expressed
per unit mass of adsorbent (mg/g), KL is the Langmuir constant (L/mg), while Kf and n
correspond to Freundlich adsorption constant (mg 1−1/n L1/n) and intensity, respectively.
Water 2022, 14, 3885 4 of 20
2.2.2. Dynamic Studies
The tests were conducted in dynamic mode. The performance of a fixed column was
investigated using breakthrough curves, presented by the concentrations of pollutants vs.
the time profile curve in a fixed bed column.
Packed-bed adsorption experiments were carried out in a cylindrical glass column
(height = 35.0 cm, internal diameter = 11.0 cm) at 20 ◦C and an initial pH of 2 ± 0.2. Em-
ploying a peristaltic pump (ISMATEC A39494), the PGP dose (0.105, 0.210, and 0.315 g) cor-
responding to (25, 50, and 75 mm) bed heights and varying flow rates (1, 2, and 3 mL/min)
were used for the removal of Cr(VI) (10, 20, and 30 mg/L). Using the breakthrough curves
(breakthrough time—tb and exhaustion time—te), it is possible to obtain various informa-
tion about the system such as the treated effluent volume (Veff in mL; Equation (4)), the
total quantity of the adsorbed Cr(VI) (qtot in mg; Equation (5)), the amount of adsorbate that
passes through the glass column(Wtotal in mg; Equation (6)), the percentage of adsorbate
uptake(R in%; Equation (7)), and the uptake (qeq(exp) in mg/g; Equation (8)):
Veff = Fte (4)
qtot =
FA
1000
=
F
1000
Z t=total
t=0
(C0 − Ct)dt (5)
Wtot =
C0Dttotal
1000
(6)
R% =
qtotal
Wtotal
× 100 (7)
qeq(exp) =
qtotal
m
(8)
where F (mL/min) represents the flow rate, te (in min) the exhaustion time, C0 (mg/L) the
initial adsorbate concentration, Ct (mg/L) the column exit concentration of Cr(VI) at time t,
and m (g) the biosorbent quantity.
The area under the breakthrough curve determined by integrating the Cad vs. t
curve will be used to calculate the total amount of PGP adsorbed (maximum column
capacity) [37]. The flow rate indicates the column’s empty bed contact time (EBCT), as
given by Equation (9):
EBCT =
Bed Volume (mL)
Flow Rate(mL/min)
(9)
Herein, it is important to mention that once the PGP sorbent becomes saturated, the
fixed bed operation was stopped. The adsorption experiments were repeated at least in
two replicates in order to estimate the error bars of the measurements.
2.3. Biosorption Modeling in Continuous Systems
The Bohart-Adams (BA), Clark, modified Dose response, and Thomas models were
adopted to characterize the breakthrough curves obtained in the continuous biosorption
studies. All related information is reported in Table 1.
Water 2022, 14, 3885 5 of 20
Table 1. The models used for modelling the adsorption data.
Model Equation and Corresponding Plot Parameters References
Bohart-Adams
Ct
C0
= exp

KBA C0t − KBA N0
Z
U

Ct
C0
vs. t
(10)
Ct
C0
: final adsorbent concentration
initial adsorbent concentration
KBA (L/mg.min): sorption rate coefficient
N0 (mg/L): fixed bed sorption capacity per unit volume
Z(mm): bed–depth in the column
U (mm/min): linear velocity,
t(min): contact time
[38]
Thomas
Ct
C0
= 1
1+exp

Kthqthm
D −kthC0t

Ct
C0
vs. t
(11)
Ct
C0
: final adsorbent concentration
initial adsorbent concentration
m(g): adsorbent mass
Kth (L/mg.min): kinetic constant of Thomas model
qth (mg/g): adsorption capacity
t (min): contact time
[39]
Modified Dose Response
Ct
C0
= 1 − 1
1+

Veff
b
a
q0 = b C0
m
Ct
C0
vs. t
(12) Ct
C0
: final adsorbent concentration
initial adsorbent concentration
Veff (mL): volume of the effluent
q0 (mg/g): adsorption capacity
a and b 6= 0: constant values in Dose Response model
m (g): adsorbent mass
[40]
(13)
Clarck
Ct
C0
=

1
1+Ae−rt
( 1
n−1 )
Ct
C0
vs. t
(14)
A and r (min−1): constant of the Clarck model.
t(min): contact time
n: constant value of Freundlich model
[41]
Water 2022, 14, 3885 6 of 20
3. Results and Discussion
3.1. Characterizations
The SEM images of PGP powder at the magnification (×10,000), as shown in Figure 1a,b,
manifest flat and smooth surface texture of the PGP particles before Cr(VI) uptake. The
corresponding elemental composition as determined by EDX reveals the main constituent’s
carbon and oxygen with minor amounts of Al, Si, Cl, P, K, Ca, Mg, Cu, and Fe, see
Figure 1d. This indicates that the two major elements forming PGP are carbon (51.40%) and
oxygen (46.21%).
Water 2022, 14, x FOR PEER REVIEW 6 of 24
3. Results and Discussion
3.1. Characterizations
The SEM images of PGP powder at the magnification (×10,000), as shown in Figure
1a,b, manifest flat and smooth surface texture of the PGP particles before Cr(VI) uptake.
The corresponding elemental composition as determined by EDX reveals the main con-
stituent’s carbon and oxygen with minor amounts of Al, Si, Cl, P, K, Ca, Mg, Cu, and Fe,
see Figure 1d. This indicates that the two major elements forming PGP are carbon (51.40%)
and oxygen (46.21%).
Figure 1. SEM images (a,b) before and (c) after Cr (VI) adsorption; EDX spectra (d) before and (e)
after adsorption by PGP.
The TGA analysis has been carried out to investigate the thermal properties of PGP,
see Figure 2A. The mass loss reflects three main stages. The first stage (30–100 °C) corre-
sponds to a weight loss of 10 wt.% owing to the evaporation of physisorbed water onto
PGP. The second stage (150–250 °C) demonstrates a weight loss of 45 wt.% and is associ-
ated with the thermal decomposition of hemicellulose and pectin [25]. While the third
stage (250–600 °C) manifests an exothermic peak with a considerable weight loss of 30
wt.%, most probably originating from the separation of lignin and decomposition of the
cellulose material [42,43]. The total mass loss at 600 °C is extremely high reaching 85 wt.%,
hence confirming the organic nature of the PGP.
Figure 1. SEM images (a,b) before and (c) after Cr (VI) adsorption; EDX spectra (d) before and (e); after
adsorption by PGP.
The TGA analysis has been carried out to investigate the thermal properties of PGP, see
Figure 2A. The mass loss reflects three main stages. The first stage (30–100 ◦C) corresponds
to a weight loss of 10 wt.% owing to the evaporation of physisorbed water onto PGP. The
second stage (150–250 ◦C) demonstrates a weight loss of 45 wt.% and is associated with the
thermal decomposition of hemicellulose and pectin [25]. While the third stage (250–600 ◦C)
manifests an exothermic peak with a considerable weight loss of 30 wt.%, most probably
originating from the separation of lignin and decomposition of the cellulose material [42,43].
The total mass loss at 600 ◦C is extremely high reaching 85 wt.%, hence confirming the
organic nature of the PGP.
Water 2022, 14, 3885 7 of 20
Water 2022, 14, x FOR PEER REVIEW 7 of 24
Figure 2. (A) TGA/DTA curves; (B) X−ray diffraction pattern; (C) FTIR spectra before and after ad-
sorption; (D) N2 adsorption−desorption isotherms of PGP.
The crystalline structure of PGP has been investigated by X−ray diffraction. As
shown in Figure 2B, no characteristic pattern of crystalline phase(s) can be observed while
a broad halo appears in the 2θ range (10−50°). This reflects the amorphous structure of
PGP powder, in good agreement with the literature [44].
Figure 2C displays the FTIR spectra of PGP powder before and after Cr(VI) absorp-
tion. The bands in the range 2800–3420 cm−1 are ascribed to −OH of carboxylic acid found
in the benzene compound [28]. The bands located at 2850 and 2930 cm−1 correspond to
C−H aliphatic (CH3) or (−CH2), while the band at 1593 cm−1 is assigned to the aromatic
rings’ stretching of phenolic compounds (C−C and C=C aromatic vibrations). In addition,
hydroxyl functional groups are defined by the absorption peak located at 1480 cm−1 and
attributed to the OH of alcohols or in-plane vibration of the OH bond in carboxylic groups.
Moreover, the absorption band at 1050 cm−1 indicates the C-O stretching of primary alco-
hol groups. Furthermore, after the adsorption of Cr(VI), a relatively significant shift of the
main band from 3200 to 3400 cm−1 is observed, hence confirming the occurrence of elec-
trostatic interaction between PGP adsorbent and adsorbate. The band at 900 cm−1 indicates
that the Cr-sorbet PGP could be due to the formation of Cr(OH)3, which suggests the re-
duction of Cr(VI) to Cr(III).
Figure 2. (A) TGA/DTA curves; (B) X−ray diffraction pattern; (C) FTIR spectra before and after
adsorption; (D) N2 adsorption−desorption isotherms of PGP.
The crystalline structure of PGP has been investigated by X−ray diffraction. As shown
in Figure 2B, no characteristic pattern of crystalline phase(s) can be observed while a broad
halo appears in the 2θ range (10−50◦). This reflects the amorphous structure of PGP
powder, in good agreement with the literature [44].
Figure 2C displays the FTIR spectra of PGP powder before and after Cr(VI) absorption.
The bands in the range 2800–3420 cm−1 are ascribed to −OH of carboxylic acid found
in the benzene compound [28]. The bands located at 2850 and 2930 cm−1 correspond to
C−H aliphatic (CH3) or (−CH2), while the band at 1593 cm−1 is assigned to the aromatic
rings’ stretching of phenolic compounds (C−C and C=C aromatic vibrations). In addition,
hydroxyl functional groups are defined by the absorption peak located at 1480 cm−1 and
attributed to the OH of alcohols or in-plane vibration of the OH bond in carboxylic groups.
Moreover, the absorption band at 1050 cm−1 indicates the C-O stretching of primary alcohol
groups. Furthermore, after the adsorption of Cr(VI), a relatively significant shift of the main
band from 3200 to 3400 cm−1 is observed, hence confirming the occurrence of electrostatic
interaction between PGP adsorbent and adsorbate. The band at 900 cm−1 indicates that the
Cr-sorbet PGP could be due to the formation of Cr(OH)3, which suggests the reduction of
Cr(VI) to Cr(III).
As shown in Figure 2D, the PGP N2 adsorption-desorption isotherm exhibits a type IV
characteristic, in accordance with the International Union of Pure and Applied Chemistry
(IUPAC) classification. The characteristic BET parameters given in Table 2, indicate a
specific surface of 40.38 m2/g, an average pore diameter of 32.13 Å, and a pore volume of
Water 2022, 14, 3885 8 of 20
0.00277 cm3/g. The distribution of pore diameter reveals that the adsorbent is mesoporous
in nature, in good agreement with the literature [45].
Table 2. BET characteristics of the PGP powder.
BET Surface Area
(m2/g)
Total Pore Volume
(cm3/g)
Average Pore Diameter
(Å)
Particle Density
(g/cm3)
PGP 40.38 2.77 × 10−3 32.13 0.25
The point of zero charge (pHpzc) of the PGP has been also determined. NaCl (0.01 N)
solution is added to a series of flasks, and the initial pH is adjusted by the addition of
NaOH (0.01 N, 0.1 N) and/or HCl (0.01 N, 0.1 N). After the pH adjustment, an amount of
adsorbent (PGP) is mixed with aliquots of 15 mL. The containers have been placed and
subjected to rigorous magnetic stirring for 48 h at 25 ◦C. The pHpzc value is then obtained
by intersecting the (pHfinal−pHinitial) vs. pHinitial curve with the bisector (Figure 3a). The
pHpzc value of PGP powder is found to be 3.9. This value suggests the presence of acid
functional groups on the PGP particles’ surface, which corroborates with FTIR analysis,
and indicates that PGP can adsorb Cr(VI) by electrostatic interaction.
Water 2022, 14, x FOR PEER REVIEW 8 of 24
As shown in Figure 2D, the PGP N2 adsorption-desorption isotherm exhibits a type
IV characteristic, in accordance with the International Union of Pure and Applied Chem-
istry (IUPAC) classification. The characteristic BET parameters given in Table 2, indicate
a specific surface of 40.38 m2/g, an average pore diameter of 32.13 Å , and a pore volume
of 0.00277 cm3/g. The distribution of pore diameter reveals that the adsorbent is mesopo-
rous in nature, in good agreement with the literature [45].
Table 2. BET characteristics of the PGP powder.
BET Surface
Area
(m2/g)
Total Pore
Volume
(cm3/g)
Average Pore
Diameter
(Å)
Particle Density
(g/cm3)
PGP 40.38 2.77 × 10−3 32.13 0.25
The point of zero charge (pHpzc) of the PGP has been also determined. NaCl (0.01
N) solution is added to a series of flasks, and the initial pH is adjusted by the addition of
NaOH (0.01 N, 0.1 N) and/or HCl (0.01 N, 0.1 N). After the pH adjustment, an amount of
adsorbent (PGP) is mixed with aliquots of 15 mL. The containers have been placed and
subjected to rigorous magnetic stirring for 48 h at 25 °C. The pHpzc value is then obtained
by intersecting the (pHfinal−pHinitial) vs. pHinitial curve with the bisector (Figure 3a). The
pHpzc value of PGP powder is found to be 3.9. This value suggests the presence of acid
functional groups on the PGP particles’ surface, which corroborates with FTIR analysis,
and indicates that PGP can adsorb Cr(VI) by electrostatic interaction.
Water 2022, 14, x FOR PEER REVIEW 9 of 24
Figure 3. (a) Plot for the determination of point zero charge of the PGP powder, (b) effect of pH on
the adsorption capacity of Cr (VI) by PGP powder and (c) speciation diagram of Cr vs. pH (T = 20 ±
2°C, C0 = 20 mg/L, H= 2 cm, D= 2 mL/min).
3.1.1. Effect of pH Value
The pH effect has been investigated between 2 and 12. It is observed that a marked
removal rate of Cr(VI) ions is achieved at lower pH 2, then declines gradually with in-
creasing the pH value. At acidic medium, the predominant species of chromium are di-
chromate (Cr2O72−) and hydrogen chromate (HCrO4-) meanwhile the surface of PGP parti-
cles becomes highly protonated, thereby favoring the adsorption and elimination of
Cr(VI) metal ions. Under this condition, the phenolic groups (−OH) of the adsorbent easily
exchange with the chromium species. Furthermore, the removal efficiency decreases sig-
nificantly by increasing the pH to more than 4. The effect of pH value on the adsorption
mechanism may be attributed to the presence of different functional carboxylic COO− and
hydroxylic OH- groups onto the surface of PGP particles, hence favoring different possible
interactions with Cr(VI) ions and other ionic species present in the solution. These results
corroborate with zeta potential measurements which revealed a positive surface charge of
PGP for pH smaller than 3.9 and a negative charge at pH greater to 3.9.
The rise in pH value causes a reduction in the positive surface of the adsorbent [46].
In fact, this decrease is directly associated with the electrostatic forces existing between
PGP and chromium solution, as consequence a marked reduction in the sorption capacity
is observed [6]. Furthermore, as the pH value increases, a competition between OH- and
2− −
Figure 3. (a) Plot for the determination of point zero charge of the PGP powder, (b) effect of pH on the
adsorption capacity of Cr (VI) by PGP powder and (c) speciation diagram of Cr vs. pH (T = 20 ± 2 ◦C,
C0 = 20 mg/L, H = 2 cm, D = 2 mL/min).
3.1.1. Effect of pH Value
The pH effect has been investigated between 2 and 12. It is observed that a marked
removal rate of Cr(VI) ions is achieved at lower pH 2, then declines gradually with increas-
ing the pH value. At acidic medium, the predominant species of chromium are dichromate
(Cr2O7
2−) and hydrogen chromate (HCrO4
−) meanwhile the surface of PGP particles be-
comes highly protonated, thereby favoring the adsorption and elimination of Cr(VI) metal
ions. Under this condition, the phenolic groups (−OH) of the adsorbent easily exchange
with the chromium species. Furthermore, the removal efficiency decreases significantly by
Water 2022, 14, 3885 9 of 20
increasing the pH to more than 4. The effect of pH value on the adsorption mechanism may
be attributed to the presence of different functional carboxylic COO− and hydroxylic OH−
groups onto the surface of PGP particles, hence favoring different possible interactions
with Cr(VI) ions and other ionic species present in the solution. These results corroborate
with zeta potential measurements which revealed a positive surface charge of PGP for pH
smaller than 3.9 and a negative charge at pH greater to 3.9.
The rise in pH value causes a reduction in the positive surface of the adsorbent [46].
In fact, this decrease is directly associated with the electrostatic forces existing between
PGP and chromium solution, as consequence a marked reduction in the sorption capacity
is observed [6]. Furthermore, as the pH value increases, a competition between OH−
and CrO4
2− as predominant ions occurs in the basic medium [47]. It was concluded that
HCrO4
− was adsorbed through electrostatic interaction to reduce to trivalent chromium
Cr(III) [34]. The optimum pH of 2 will be utilized for the following experiments. The
as-obtained results are found to be close to the results reported in a previous study using
pomegranate peel for the elimination of hexavalent chromium [12]. The predominant form
of Cr(VI) vs. pH (Figure 3a) is found to be HCrO4
− at the pH of 3.0 and would change to
CrO4
2− and Cr2O7
2− when the pH  6.5. Compared with CrO4
2−, HCrO4
− is known to be
more easily adsorbed on the surface active site owing to its low adsorption free energy [45].
3.1.2. Effect of Contact Time
Figure 4a shows the effect of contact time on Cr(VI) removal from aqueous solution
onto PGP adsorbent. Based on the plot, the adsorption processes are almost accomplished
during the first time and reach equilibrium within 2 h (removal percentage = 92%). After a
certain period of time, the reaction slows down as the process reaching the steady state,
because the number of vacant sites decreases due to the repulsive forces occurring between
the solute molecules and the solid adsorbent phase [20].
Water 2022, 14, x FOR PEER REVIEW 10 of 24
3.1.2. Effect of Contact Time
Figure 4a shows the effect of contact time on Cr(VI) removal from aqueous solution
onto PGP adsorbent. Based on the plot, the adsorption processes are almost accomplished
during the first time and reach equilibrium within 2 h (removal percentage = 92%). After
a certain period of time, the reaction slows down as the process reaching the steady state,
because the number of vacant sites decreases due to the repulsive forces occurring be-
tween the solute molecules and the solid adsorbent phase [20].
Figure 4. (a) effect of contact time on Cr(VI) removal, (b) Langmuir and Freundlich plots for the
adsorption on PGP at different concentration of Cr(VI) (pH = 2 ± 0.2, T = 20 ± 2 °C, m = 1.0 g/L,
agitation speed = 50 rpm).
3.2. Adsorption Isotherms
The non-linear Freundlich and Langmuir isotherms are shown in Figure 4b. It is no-
ticed that the equilibrium data are well described by Freundlich isotherm compared to
Langmuir isotherm. The model parameters, R2 values and 𝜒2
are presented in Table 3. The
obtained results indicate that the regression coefficients for Cr(VI) metal ions adsorption
with Freundlich isotherm are relatively high (0.994) compared to Langmuir (0.990) iso-
Figure 4. (a) effect of contact time on Cr(VI) removal, (b) Langmuir and Freundlich plots for the
adsorption on PGP at different concentration of Cr(VI) (pH = 2 ± 0.2, T = 20 ± 2 ◦C, m = 1.0 g/L,
agitation speed = 50 rpm).
Water 2022, 14, 3885 10 of 20
3.2. Adsorption Isotherms
The non-linear Freundlich and Langmuir isotherms are shown in Figure 4b. It is
noticed that the equilibrium data are well described by Freundlich isotherm compared to
Langmuir isotherm. The model parameters, R2 values and χ2 are presented in Table 3. The
obtained results indicate that the regression coefficients for Cr(VI) metal ions adsorption
with Freundlich isotherm are relatively high (0.994) compared to Langmuir (0.990) isotherm.
Table 3. Langmuir and Freundlich isotherm models constants and determination coefficients for the
adsorption of Cr(VI) by PGP powder.
Langmuir Freundlich
qm(exp)
(mg/g)
qm(theo)
(mg/g)
KL
(L/mg)
R2 χ2 n
Kf
(mg1−1/n.L1/n)
R2 χ2
50.32 38.29 ± 2.58 0.08 ± 0.01 0.990 0.528 3.00 ± 0.07 4.49 ± 0.28 0.994 0.33
The value of 1/n for all adsorption curves is found to be less than unity, which reflects
the favorable exchange characteristics between Cr(VI) and PGP. The presence of hydroxyls
and carboxyls on the surface of PGP particles produces an irregularity in the distribution
energy over the surface of the adsorbent and therefore confirms the suitability of the
Freundlich isotherm.
Furthermore, the batch mode’s isotherm constants are significantly greater than the fixed
bed mode. This may be due to the influence of the solution flow rate which is equal to zero in
the batch mode, i.e., the contact time between Cr(VI) and PGP approximates the infinity [48].
3.3. Influence of Parameters on Fixed Bed
3.3.1. Flow Rate
The breakthrough curves have been recorded at varying flow rates (1, 2, and 3 mL/min),
while the remaining operating parameters are kept constant, i.e., inlet Cr(VI) concentration
20 mg/Land bed-depth 50 mm, see Figure 5 and the computed parameters are summarized
in Table 4. It is observed that the sorption rate is reduced with the increase in the flow rate.
The empty bed contact time (EBCT) decreases significantly from 108 up to 38 min and the
exhaust time (equivalent to 90% of the Cr (VI) concentration) also decreases markedly from
550 to 168 min. The flow rate is found to have a relatively moderate effect on the adsorption
capacity; as the flow rate increases, the amount of adsorbed Cr(VI) (qexp) decreases slightly
from 28.67 to 24.85 mg/g. The observed decline in the adsorption capacity can be attributed
to the shorter contact time for the adsorbent to interact with the solution, which is in
agreement with the previous research findings [39].
Water 2022, 14, x FOR PEER REVIEW 11 of 24
The value of 1/n for all adsorption curves is found to be less than unity, which reflects
the favorable exchange characteristics between Cr(VI) and PGP. The presence of hydrox-
yls and carboxyls on the surface of PGP particles produces an irregularity in the distribu-
tion energy over the surface of the adsorbent and therefore confirms the suitability of the
Freundlich isotherm.
Furthermore, the batch mode’s isotherm constants are significantly greater than the
fixed bed mode. This may be due to the influence of the solution flow rate which is equal
to zero in the batch mode, i.e., the contact time between Cr(VI) and PGP approximates the
infinity [48].
3.3. Influence of Parameters on Fixed Bed
3.3.1. Flow Rate
The breakthrough curves have been recorded at varying flow rates (1, 2, and 3
mL/min), while the remaining operating parameters are kept constant, i.e., inlet Cr(VI)
concentration 20 mg/Land bed-depth 50 mm, see Figure 5 and the computed parameters
are summarized in Table 4. It is observed that the sorption rate is reduced with the in-
crease in the flow rate. The empty bed contact time (EBCT) decreases significantly from
108 up to 38 min and the exhaust time (equivalent to 90% of the Cr (VI) concentration)
also decreases markedly from 550 to 168 min. The flow rate is found to have a relatively
moderate effect on the adsorption capacity; as the flow rate increases, the amount of ad-
sorbed Cr(VI) (qexp) decreases slightly from 28.67 to 24.85 mg/g. The observed decline in
the adsorption capacity can be attributed to the shorter contact time for the adsorbent to
interact with the solution, which is in agreement with the previous research findings [39].
Figure 5. Cont.
Water 2022, 14, 3885 11 of 20
Water 2022, 14, x FOR PEER REVIEW 12 of 24
Figure 5. Comparison of the experimental and predict breakthrough curves obtained at different
(A) flow rights, (B) bed heights, and (C) concentrations according to the studied models for Cr(VI)
adsorption by PGP powder (T = 20 ± 2 °C, pH = 2 ± 0.2).
Table 4. Total removal percentage of Cr(VI), column uptake capacity, and column data parameters
obtained at different bed heights, concentrations, and flow rates (T = 20 ± 2 °C, pH = 2 ± 0.2).
F
(mL/min)
H
(mm)
C0
(mg/L)
tb
(min)
te
(min)
Veff
(L)
Area 𝐀′
(mg/min.L)
𝐪𝐭𝐨𝐭
(mg)
𝐖𝐭𝐨𝐭
(mg)
Total
Removal
(%)
𝐪𝐞𝐪(𝐞𝐱𝐩)
(mg/g)
1 50 20 108 700 0.70 6087 6.08 14.00 43.47 28.67
2 50 20 45 320 0.64 2803 5.60 12.80 43.75 26.73
3 50 20 36 204 0.61 1746 5.23 12.24 42.79 24.85
2 25 20 24 201 0.40 1083 2.16 8.04 20.57 20.62
2 75 20 78 430 0.86 4187 8.37 17.20 48.66 26.62
2 50 10 72 385 0.77 1924 3.84 7.70 49.87 18.28
2 50 30 27 270 0.54 3151 6.30 16.20 38.88 30.00
3.3.2. Bed Height
The influence of varying the bed-height from 25 to 75 mm on the behavior of the
breakthrough curves has been investigated at a fixed adsorbent flow rate of 2 mL/min and
an inlet Cr(VI) concentration of 20 mg/L. The recorded breakthrough curves are displayed
in Figure 5B and the computed values are given in Table 4. It is clearly observed that as
the bed height increases, the Cr(VI) removal efficiency and qeq(exp) increase simultane-
ously, from 20.57 to 48.66% and 20.62 to 26.62 mg/g. Consequently, while investigating
the effect of PGP bed height on the adsorption of 20 mg/L Cr(VI) ions, the breakpoint
value has been fixed at the point where
Ct
C0
is equal to 0.076, 0.090, and 0.100 for a fixed
bed height of 25, 50 and 75 mm respectively. It is found that with the increase of the bed
height, the breakthrough curve slope decreases, resulting in a broader mass transfer [39].
For this reason, the specific surface area of the PGP biosorbent increases, consequently
providing a large number of binding active sites thereby a higher adsorption rate of the
column [49]. Also, it is observed that the breakthrough time increases significantly from
24 to 78 min with the increase in the bed height. This can be associated with the occurrence
of rapid and larger mass transfer in the fixed bed mode [49].
Figure 5. Comparison of the experimental and predict breakthrough curves obtained at different
(A) flow rights, (B) bed heights, and (C) concentrations according to the studied models for Cr(VI)
adsorption by PGP powder (T = 20 ± 2 ◦C, pH = 2 ± 0.2).
Table 4. Total removal percentage of Cr(VI), column uptake capacity, and column data parameters
obtained at different bed heights, concentrations, and flow rates (T = 20 ± 2 ◦C, pH = 2 ± 0.2).
F
(mL/min)
H
(mm)
C0
(mg/L)
tb
(min)
te
(min)
Veff
(L)
Area A’
(mg/min.L)
qtot
(mg)
Wtot
(mg)
Total Re-
moval
(%)
qeq(exp)
(mg/g)
1 50 20 108 700 0.70 6087 6.08 14.00 43.47 28.67
2 50 20 45 320 0.64 2803 5.60 12.80 43.75 26.73
3 50 20 36 204 0.61 1746 5.23 12.24 42.79 24.85
2 25 20 24 201 0.40 1083 2.16 8.04 20.57 20.62
2 75 20 78 430 0.86 4187 8.37 17.20 48.66 26.62
2 50 10 72 385 0.77 1924 3.84 7.70 49.87 18.28
2 50 30 27 270 0.54 3151 6.30 16.20 38.88 30.00
3.3.2. Bed Height
The influence of varying the bed-height from 25 to 75 mm on the behavior of the
breakthrough curves has been investigated at a fixed adsorbent flow rate of 2 mL/min and
an inlet Cr(VI) concentration of 20 mg/L. The recorded breakthrough curves are displayed
in Figure 5B and the computed values are given in Table 4. It is clearly observed that as the
bed height increases, the Cr(VI) removal efficiency and qeq(exp) increase simultaneously,
from 20.57 to 48.66% and 20.62 to 26.62 mg/g. Consequently, while investigating the effect
of PGP bed height on the adsorption of 20 mg/L Cr(VI) ions, the breakpoint value has
been fixed at the point where Ct
C0
is equal to 0.076, 0.090, and 0.100 for a fixed bed height
of 25, 50 and 75 mm respectively. It is found that with the increase of the bed height, the
breakthrough curve slope decreases, resulting in a broader mass transfer [39]. For this
reason, the specific surface area of the PGP biosorbent increases, consequently providing a
large number of binding active sites thereby a higher adsorption rate of the column [49].
Also, it is observed that the breakthrough time increases significantly from 24 to 78 min
with the increase in the bed height. This can be associated with the occurrence of rapid and
larger mass transfer in the fixed bed mode [49].
3.3.3. Initial Concentration
Figure 5C illustrates the breakthrough curves of different inlet Cr(VI) concentra-
tions onto PGP adsorbent obtained at a constant bed depth (50 mm) and fixed flowrate
Water 2022, 14, 3885 12 of 20
(2 mL/min), the corresponding data are given in Table 4. It can be observed that the value
of Ct
C0
reaches 0.010, 0.039, and 0.067 min in the intervals 365, 300, and 230 min when
the Cr(VI) inlet concentration varies from 10 to 30 mg/L. This indicates that the value of
the breakthrough time (tb) is slightly reduced with the rise in the inlet concentration of
Cr(VI). The maximum adsorption capacity at 10, 20, and 30 mg/L Cr(VI) is found to be
18.32, 26.73 and 30.01 mg/g, respectively. In contrast, the adsorption efficiency is reduced
by almost 22% (from 49.87 to 38.88%). Furthermore, it is noted that when the Cr(VI) con-
centration in the solution is high, the rate of Cr(VI) ions for the adsorptive sites has a more
pronounced trend. Indeed, a higher concentration of Cr(VI) in the solution provides a
random diffusion path [50] for sorbate ions onto the surface of PGP adsorbent. This may
be explained by the enhanced driving force and the high mass-transfer flux from Cr(VI)
solution toward PGP [51]. However, at lower inlet Cr(VI) concentration, the breakthrough
curves become dispersed, and the binding sites tend to be slowly saturated. Indeed, similar
results were reported for the biosorption of Cr(VI) by chemically surface-modified Lantana
Camara sorbent [39].
3.3.4. Temperature
The influence of the medium temperature has been also investigated. The break-
through curves recorded for different bed heights at various temperatures are shown in
Figure 6, and the results of Dose Response Model are given in Table 5. It is noted that
with the rise of temperature, the adsorption rate of Cr(VI) ions decreases whereas the rate
constant increases. Also, it is noticed that the adsorption rate is clearly variable at the
same bed height. This is maybe due to the desorption process occurring following the rise
in the thermal energy and the weakening of the interactions of the intermolecular bonds
between solution ions and active sites of the PGP adsorbent. This signifies that the PGP
particles’ surface is exothermic in nature and that the adsorption mechanism is favored at
low temperatures. The above results are in accordance with the results obtained during the
thermodynamic study in batch mode; the exothermic reaction of Cr(VI) ions adsorption onto
PGP will be confirmed. This result has been previously reported by M. Zamouch et al. [21].
Water 2022, 14, x FOR PEER REVIEW 14 of 24
Figure 6. Effect of bed height on the breakthrough curves at different temperatures: (a) 303 K and
(b) 313 K (C0 = 20 mg/L, H = 50 mm, D = 2 mL/min, pH = 2 ± 0.2).3.4. Adsorption Thermodynamics.
The adsorption Cr(VI) ions by PGP thermodynamic parameters in the temperature
range 298–318 K, have been determined using Van’t Hoff equations [52]:
∆G°
= −RTLnKC (15)
∆G°
= ∆H°
− T∆S° (16)
KC =
qe
Ce
(17)
LnK =
∆S°
−
∆H°
(18)
Figure 6. Effect of bed height on the breakthrough curves at different temperatures: (a) 303 K and
(b) 313 K (C0 = 20 mg/L, H = 50 mm, D = 2 mL/min, pH = 2 ± 0.2).3.4. Adsorption Thermodynamics.
Water 2022, 14, 3885 13 of 20
Table 5. Dose response parameters for Cr(VI) adsorption by PGP in fixed bed (C0 = 20 mg/L,
pH = 2 ± 0.2, D = 2 mL/min).
T (K)
Z
(mm)
a
b
(mL)
q0
(mg/g)
qexp
(mg/g)
Total Removal
(%)
R2 χ×104 SD
303
25 2.24 ± 0.04 52.85 ± 0.51 10.06 10.15 38.00 0.997 2.76 0.01
50 2.75 ± 0.07 138.32± 1.47 13.17 13.38 44.00 0.996 5.94 0.02
75 2.92± 0.08 194.12 ± 2.12 13.98 13.93 44.12 0.996 5.20 0.02
313
25 1.28± 0.10 20.96 ± 1.67 3.99 3.89 30.06 0.927 38.50 0.06
50 2.20 ± 0.14 51.85 ± 1.65 4.93 4.90 33.09 0.977 27.80 0.05
75 3.00 ± 0.11 98.13 ± 1.38 6.23 6.20 34.10 0.994 9.17 0.03
The adsorption Cr(VI) ions by PGP thermodynamic parameters in the temperature
range 298–318 K, have been determined using Van’t Hoff equations [52]:
∆G
◦
= −RTLnKC (15)
∆G
◦
= ∆H
◦
− T∆S
◦
(16)
KC =
qe
Ce
(17)
LnKC =
∆S
◦
R
−
∆H
◦
RT
(18)
where ∆G
◦
represents the Gibbs free energy (kJ/mol), R the gas constant (8.314 J/mol.K),
T the absolute temperature (K), KC the sorption distribution coefficient (L/g), ∆H◦ the
enthalpy (kJ/mol.), and ∆S
◦
the entropy (kJ/mol.K).
To better elucidate the adsorption process mechanism of Cr(VI) by PGP biosorbent,
the respective thermodynamic plots are depicted in Figure 7 while the computed fitting
parameters are given in Table 6. The negative values of ∆H◦, ∆S◦, and ∆G◦ demonstrate
that the adsorption process is spontaneous and exothermic, without the need for any
external power source. Also, it signifies the randomness of adsorption of Cr(VI) onto
PGP. The negative value of ∆S◦ corresponds to a decrease in the freedom of the adsorbed
chemical species and describes the randomness of chemical species movement at the
adsorbent-adsorbate interface during the sorption process. The negative values of ∆G◦ and
high negative values of ∆S◦ with the rise of temperature manifest the favorable nature of
the process. Whereas the observed decrease in ∆G◦ values with temperature signifies that
the adsorption process occurs favorably at lower temperatures [53]. Similar results were
also observed in the literature for Cr(VI) adsorption onto different adsorbents [6,54].
Water 2022, 14, x FOR PEER REVIEW 15 of 24
Figure 7. Plots of LnKC vs. 1/T for the estimation of the thermodynamic parameters for the adsorp-
tion of Cr(VI) on PGP (pH = 2 ± 0.2, m = 1.0 g.L−1, C0 = 20 mg/L).
Table 6. Thermodynamic parameters calculated for the adsorption of Cr(VI) onto PGP.
Sample
∆𝐇°
(kJ/mol)
∆𝐒°
(J/mol.K)
∆𝐆°
(kJ/mol)
298 K 308 K 318 K
PGP −44.69 −139.28 −3.98 −2.30 −0.48
3.4. Modeling of Column Data
The fitted curves are shown in Figure 5 and the models’ corresponding calculated
parameters are presented in Table 7. The nonlinear plot
Ct
vs. t has been fitted in the ini-
Figure 7. Plots of LnKC vs. 1/T for the estimation of the thermodynamic parameters for the
adsorption of Cr(VI) on PGP (pH = 2 ± 0.2, m = 1.0 g.L−1, C0 = 20 mg/L).
Water 2022, 14, 3885 14 of 20
Table 6. Thermodynamic parameters calculated for the adsorption of Cr(VI) onto PGP.
Sample
∆H
◦
(kJ/mol)
∆S
◦
(J/mol.K)
∆G
◦
(kJ/mol)
298 K 308 K 318 K
PGP −44.69 −139.28 −3.98 −2.30 −0.48
3.4. Modeling of Column Data
The fitted curves are shown in Figure 5 and the models’ corresponding calculated
parameters are presented in Table 7. The nonlinear plot Ct
C0
vs. t has been fitted in the
initial part of the curve (0–0.2) to compute the Bohart–Adams KBA and sorption capacity N0
parameters. It is found that the value of KBA increases from 4.16 × 10−3 to 7.92 × 10−3 with
increasing the flow rate but decreases from 11.28 × 10−3 to 3.28 × 10−3 with the increase
in the initial Cr(VI) concentration. Also, the value of N0 of the model is found to increase
with increasing both Cr(VI) concentration and bed depth. This can be associated with the
prevalence of external mass transfer activities occurring during the adsorption process
within the column at the initial stage [55]. This corroborates with the results published
by Han et al. [45] when employing green synthesized nanocrystalline chlorapatite for the
sorption of Cr(VI).
The Thomas model parameters namely kinetic coefficient Kth and sorption capacity
qth, have been determined using a nonlinear regression for bed height, flow rate, and Cr(VI)
concentration. From Table 7, it is noted that the value of Kth increases with increasing
both bed height and flow rate. Also, the value of qth increases markedly (from 15.75 to
25.39 mg/g) with the rise of Cr(VI) concentration (from 10 to 30 mg/L), which is attributed
to the high driving force between Cr(VI) ions and PGP sorbent. However, the increase in
the Kth value with the increase in the flow rate, signifies that the kinetics of the overall
process is dominated by the external mass transfer mechanism [56]. This suggests that
the Thomas model is more favorable for describing the adsorption processes, because the
limiting steps do not proceed through external/internal diffusion [37]. Similar results are
found in the literature when using different sorbate-sorbent systems [54].
According to the results given in Table 7, the rate of mass transfer (r) increases gradu-
ally with the rise of both flow rate and Cr(VI) concentration but decreases with the increase
in bed height. Furthermore, the increase in the initial Cr(VI) concentration results in the
enhancement of the driving force, favoring a better mass transfer of the adsorbate onto PGP
particles’ surface. Nevertheless, with the rise in the bed height, an increase in the particles’
number alongside a reduction in the mass transfer rate occur concurrently [57].
The fitting results of the modified Dose Response parameters are also examined. The
predicted breakthrough curves illustrated in Figure 5 evidenced a good agreement with
the measurements. The obtained high value of the regression coefficient (R2) satisfies the
validity of the model. The values of R2 obtained for different operating conditions are
higher than 0.977 and χ2 smaller than 2.98 × 10−3, thus providing a better fitting for the
Dose Response model compared to Thomas, Clarck, and Bohart-Adams models.
Water 2022, 14, 3885 15 of 20
Table 7. Parameters of various models for Cr(VI) adsorption by PGP using non-linear regression. Experimental fixed conditions: T = 20 ± 2 ◦C, pH = 2 ± 0.2.
Parameters Thomas Model Dose Response Model
F
(mL/min)
H
(mm)
C0
(mg/L)
qecal
(mg/g)
Total Removal
(%)
KThX103
(L/mg)
R2 χ2
×103
±SD
× 103 a
b
(mL)
q0
(mg/g)
R2 χ2
×103
±SD
×103
1 50 20 25.69 ± 0.37 43.47 0.59 ± 0.02 0.976 3.16 56.22 2.73 ± 0.07 269.78 ± 2.98 27.79 0.988 1.47 38.38
2 50 20 21.59 ± 0.54 43.75 1.16± 0.07 0.953 5.85 76.50 2.46 ± 0.07 226.74 ± 2.73 23.26 0.988 1.45 38.05
3 50 20 21.52 ± 0.34 42.79 2.04 ± 0.09 0.976 3.11 55.80 2.94 ± 0.06 226.04± 1.80 22.86 0.994 0.71 26.73
2 25 20 16.63 ± 0.44 20.57 3.70 ± 0.28 0.970 4.50 67.05 3.07 ± 0.11 87.34± 1.12 17.57 0.993 1.02 31.89
2 75 20 22.77 ± 0.56 48.66 0.72 ± 0.03 0.956 5.72 75.63 2.49± 0.08 358.64 ± 5.95 26.02 0.979 2.71 52.09
2 50 10 15.75 ± 0.42 49.87 1.53 ±0.09 0.952 6.34 79.63 2.50 ± 0.09 330.88 ± 5.75 18.02 0.977 2.98 54.57
2 50 30 25.39 ± 0.45 38.88 0.95 ±0.04 0.972 3.19 56.48 2.35 ± 0.06 177.37 ± 1.84 27.22 0.991 0.99 31.53
Clark model Bohart-Adams model
D
mL.min−1
H
(mm)
C0
(mg/L)
A
r × 103
(min−1)
R2 χ2
×103
±SD
×103
KBA×103
(mL/mg.min)
N0
(mg/L)
R2 χ2
×103
±SD
×103
1 50 20 535.29 ± 139.02 17.31 ± 0.85 0.964 4.72 68.69 4.16 ± 0.36 6024.43 ± 88.76 0.900 0.26 16.32
2 50 20 184.20 ± 56.80 33.24 ± 2.64 0.933 8.44 91.87 4.24 ± 0.11 6689.07 ± 218.38 0.914 0.32 17.99
3 50 20 298.95 ± 89.87 56.11 ± 3.49 0.962 4.99 70.65 7.92 ±0.11 6839.09 ± 209.58 0.923 0.40 20.21
2 25 20 383.93 ± 188.25 103.12 ± 10.26 0.933 6.66 81.63 10.30± 1.52 6222.88 ± 213.18 0.946 0.37 19.39
2 75 20 206.66 ± 52.18 19.67 ± 1.16 0.941 7.76 88.12 3.30± 1.52 6721.11 ± 91.34 0.940 0.20 14.32
2 50 10 186.00 ± 42.24 20.76 ± 1.31 0.936 8.52 92.29 11.28 ± 1.16 4208.87 ± 73.41 0.923 0.27 16.49
2 50 30 154.18 ± 37.66 40.68 ± 2.51 0.957 4.91 70.10 3.28 ± 0.43 7191.59 ± 323.61 0.922 0.48 21.94
Water 2022, 14, 3885 16 of 20
3.5. Adsorption Mechanism
The SEM image of PGP after the adsorption of Cr(VI) metal ions (Figure 1d) reveals
that the particles’ surface becomes more rough and uneven. This confirms that the pores of
the PGP sorbent are occupied by the ions of the adsorbate. Furthermore, the EDX spectrum
(Figure 1e) of PGP after adsorption reveals the presence of Cr ions in the adsorbent, where
its content reaches 3.64 wt.%. The presence of 1.5 wt.% K, maybe a residue belonging to the
potassium dichromate (K2Cr2O7) used during the preparation of Cr(VI) aqueous solutions.
Through the analytical results of EDX (Figure 1e) and FTIR (Figure 2C) spectra, as well
as the data obtained from modeling of the equilibrium isotherms and kinetics, a plausible
mechanism related to the adsorption of Cr(VI) onto PGP is illustrated in Figure 8. It is
noted that the process is primarily controlled by electrostatic interaction between negatively
charged functional groups present at the PGP particles’ surface and the positively charged
Cr(VI) metal ions. Meanwhile, the surface adsorption via diffusion to the pores of PGP
improves the mass transfer and provides an adsorption pathway besides an area responsible
for the reduction of Cr(VI) metal ions to less toxic Cr(III).
Water 2022, 14, x FOR PEER REVIEW 19 of 24
Figure 8. Plausible mechanism for the adsorption of Cr(VI) by PGP powder.
3.6. Comparison with Other Sorbents in the Literature
To evaluate the sorption efficacy of PGP toward Cr(VI), a comprehensive comparison
with other materials such as solvent impregnated resin (SIR), pomegranate peel powder
(PGP) subjected to polyaniline (PANI), and polypyrrole (PPy) surface functionalization
Figure 8. Plausible mechanism for the adsorption of Cr(VI) by PGP powder.
Water 2022, 14, 3885 17 of 20
3.6. Comparison with Other Sorbents in the Literature
To evaluate the sorption efficacy of PGP toward Cr(VI), a comprehensive comparison
with other materials such as solvent impregnated resin (SIR), pomegranate peel powder
(PGP) subjected to polyaniline (PANI), and polypyrrole (PPy) surface functionalization and
modified Lantana Camara, is performed as illustrated in Table 8. It can be observed that the
proposed biosorbent exhibits a higher adsorption capacity than the previously reported
sorbents; it is even higher than pomegranate peel activated with H2SO4. In addition, the
obtained batch capacity is more significant than that of SIR and PGP for Cr(VI) removal.
Nevertheless, the existence of some differences is more probably attributed to several
factors, primarily the active sites and functional groups formed at the particles’ surfaces of
the sorbents, besides the microstructure nature (surface area, porosity, agglomeration level)
and the chemical formulation (intrinsic properties playing key role in terms of selectivity)
of the prepared materials.
Table 8. A comparison of adsorption capacity or removal percentage of Cr(VI) with different biosorbents.
Adsorbent Qmax (mg/g)/R (%) Operating Conditions References
Pomegranate peel
(Batch study)
22.87 mg/g pH 2, dose 300 mg/L, contact time 30 min [12]
Solvent impregnated resin (SIR)(Batch study) 28.2 0 mg/g pH (0.3–2), dose 1 mg/L, contact time 30 min [54]
Pomegranate peel (PGP) powder modified by
polyaniline (PANI) and polypyrrole (PPy) (Batch study)
47.59% pH 1, dose 10 mg/L, contact time 90 min [36]
Nanocrystalline chlorapatite (ClAP)
(Batch study)
63.47 mg/g pH 3, dose 0.1 mg/L, contact time 25 min [45]
Pomegranate peel (PGP) activated (PGP-1N H2SO4)
(Batch study)
28.28 mg/g pH 3, dose 4 mg/L, contact time 3 h [6]
Modified Lantana Camara
(Column study)
362.80 mg/g
pH 1.5, adsorbent bed height 40 mm,
Breakthrough time 1250 min
[39]
Pomegranate peel
(Column study)
(Batch study)
30.00 mg/g
50.32 mg/g
pH 2, adsorbent bed height 50 mm,
Breakthrough time 27 min
pH 2, dose 1 mg/L, contact time 24 h
This study
4. Conclusions
In the present study Pomegranate peel has been used as cost-effective biosorbent
for the removal of hazardous Cr(VI). The adsorption mechanism of Cr(VI) adsorbate by
PGP nanosorbent is determined and the physicochemical proprieties of PGP are presented.
Both batch and dynamic studies are performed, and the isotherms and thermodynamic
results are analyzed. It is found that the adsorption process occurs through an exothermic
mechanism. Several operating parameters (pH, initial Cr(VI) concentration, temperature,
bed depth, and flow rate) are found to significantly affect the Cr(VI) uptake through a fixed-
bed column. The maximum uptake of 43.78 mg/g is achieved at pH 2 with a breakthrough
time. The experimental data in dynamic mode has been fitted using four theoretical models
namely Bohart-Adams, Clark, Dose Response, and Thomas. The Dose Response model
successfully predicts the breakthrough curves for Cr(VI) removal under optimum operating
conditions (flow rate, bed depth, Cr(VI) concentration) with R2  0.97. The equilibrium data
were found to obey Freundlich isotherm, thereby confirming the heterogeneous sorption of
Cr(VI) onto PGP. The negative values of ∆H◦, ∆G◦, and ∆S◦ indicate an exothermic and
spontaneous process.
Author Contributions: Conceptualization: Z.H.; Methodology: Y.B., N.B.; Formal analysis and
investigation: R.Z.; Writing—original draft preparation: R.Z., M.B.; Writing—review and editing:
R.Z., M.B., R.A.A.; Resources: R.Z.; Supervision: M.B., R.A.A. All authors have read and agreed to
the published version of the manuscript.
Funding: This research was funded by Algerian Ministry of Higher Education, grant number
B00L01UN230120200002.
Water 2022, 14, 3885 18 of 20
Institutional Review Board Statement: Not applicable.
Data Availability Statement: Data is available upon formal request to authors.
Acknowledgments: The authors are thankful to Prince Sultan University for their support for paying
the article processing charge of this publication.
Conflicts of Interest: The authors have no competing interest to declare that are relevant to the
content of this article.
References
1. Laxmi, V.; Kaushik, G. Toxicity of Hexavalent Chromium in Environment, Health Threats, and Its Bioremediation and Detoxifica-
tion from Tannery Wastewater for Environmental Safety. In Bioremediation of Industrial Waste for Environmental Safety; Springer:
Berlin/Heidelberg, Germany, 2020; pp. 223–243.
2. Ina, S. Chromium, an Essential Nutrient and Pollutant: A Review. African J. Pure Appl. Chem. 2013, 7, 310–317.
3. Sharma, P.; Singh, S.P.; Parakh, S.K.; Tong, Y.W. Health Hazards of Hexavalent Chromium (Cr (VI)) and Its Microbial Reduction.
Bioengineered 2022, 13, 4923–4938. [CrossRef] [PubMed]
4. Kazemi, A.; Esmaeilbeigi, M.; Sahebi, Z.; Ansari, A. Health Risk Assessment of Total Chromium in the Qanat as Historical
Drinking Water Supplying System. Sci. Total Environ. 2022, 807, 150795. [CrossRef] [PubMed]
5. Bakshi, A.; Panigrahi, A.K. A Comprehensive Review on Chromium Induced Alterations in Fresh Water Fishes. Toxicol. Rep. 2018,
5, 440–447. [CrossRef]
6. Abdel-Galil, E.A.; Hussin, L.M.S.; El-Kenany, W.M. Adsorption of Cr (VI) from Aqueous Solutions onto Activated Pomegranate
Peel Waste. Desalin. WATER Treat. 2021, 211, 250–266. [CrossRef]
7. Minas, F.; Chandravanshi, B.S.; Leta, S. Chemical Precipitation Method for Chromium Removal and Its Recovery from Tannery
Wastewater in Ethiopia. Chem. Int. 2017, 3, 291–305.
8. Qu, J.; Wang, Y.; Tian, X.; Jiang, Z.; Deng, F.; Tao, Y.; Jiang, Q.; Wang, L.; Zhang, Y. KOH-Activated Porous Biochar with High
Specific Surface Area for Adsorptive Removal of Chromium (VI) and Naphthalene from Water: Affecting Factors, Mechanisms
and Reusability Exploration. J. Hazard. Mater. 2021, 401, 123292. [CrossRef]
9. Wang, Z.; Shen, Q.; Xue, J.; Guan, R.; Li, Q.; Liu, X.; Jia, H.; Wu, Y. 3D Hierarchically Porous NiO/NF Electrode for the Removal
of Chromium (VI) from Wastewater by Electrocoagulation. Chem. Eng. J. 2020, 402, 126151. [CrossRef]
10. Fotsing, P.N.; Woumfo, E.D.; Mezghich, S.; Mignot, M.; Mofaddel, N.; Le Derf, F.; Vieillard, J. Surface Modification of Biomaterials
Based on Cocoa Shell with Improved Nitrate and Cr (vi) Removal. RSC Adv. 2020, 10, 20009–20019. [CrossRef]
11. Dehghani, M.H.; Karri, R.R.; Lima, E.C.; Mahvi, A.H.; Nazmara, S.; Ghaedi, A.M.; Fazlzadeh, M.; Gholami, S. Regression
and Mathematical Modeling of Fluoride Ion Adsorption from Contaminated Water Using a Magnetic Versatile Biomaterial 
Chelating Agent: Insight on Production  Experimental Approaches, Mechanism and Effects of Potential Interferers. J. Mol. Liq.
2020, 315, 113653.
12. Giri, R.; Kumari, N.; Behera, M.; Sharma, A.; Kumar, S.; Kumar, N.; Singh, R. Adsorption of Hexavalent Chromium from Aqueous
Solution Using Pomegranate Peel as Low-Cost Biosorbent. Environ. Sustain. 2021, 4, 401–417. [CrossRef]
13. Zhu, S.; Khan, M.A.; Kameda, T.; Xu, H.; Wang, F.; Xia, M.; Yoshioka, T. New Insights into the Capture Performance and
Mechanism of Hazardous Metals Cr3+ and Cd2+ onto an Effective Layered Double Hydroxide Based Material. J. Hazard. Mater.
2022, 426, 128062. [CrossRef] [PubMed]
14. Bouchelkia, N.; Mouni, L.; Belkhiri, L.; Bouzaza, A.; Bollinger, J.C.; Madani, K.; Dahmoune, F. Removal of Lead(II) from Water
Using Activated Carbon Developed from Jujube Stones, a Low-Cost Sorbent. Sep. Sci. Technol. 2016, 51, 1645–1653. [CrossRef]
15. Imessaoudene, A.; Cheikh, S.; Bollinger, J.C.; Belkhiri, L.; Tiri, A.; Bouzaza, A.; El Jery, A.; Assadi, A.; Amrane, A.; Mouni, L.
Zeolite Waste Characterization and Use as Low-Cost, Ecofriendly, and Sustainable Material for Malachite Green and Methylene
Blue Dyes Removal: Box–Behnken Design, Kinetics, and Thermodynamics. Appl. Sci. 2022, 12, 7587. [CrossRef]
16. Zhu, S.; Xia, M.; Chu, Y.; Khan, M.A.; Lei, W.; Wang, F.; Muhmood, T.; Wang, A. Applied Clay Science Adsorption and Desorption
of Pb (II) on L -Lysine Modi Fi Ed Montmorillonite and the Simulation of Interlayer Structure. Appl. Clay Sci. 2019, 169, 40–47.
[CrossRef]
17. Chestnut, A.; Pigment, S. Cu (II) Adsorption from Aqueous Solution onto Poly (Acrylic Acid/Chestnut Shell Pigment) Hydrogel.
Water 2022, 14, 3500. [CrossRef]
18. Dai, Y.; Sun, Q.; Wang, W.; Lu, L.; Liu, M.; Li, J.; Yang, S.; Sun, Y.; Zhang, K.; Xu, J. Utilizations of Agricultural Waste as Adsorbent
for the Removal of Contaminants: A Review. Chemosphere 2018, 211, 235–253. [CrossRef]
19. Salam, F.A.; Narayanan, A. Biosorption-a Case Study of Hexavalent Chromium Removal with Raw Pomegranate Peel. Desalin.
Water Treat. 2019, 156, 278–291. [CrossRef]
20. Chedri Mammar, A.; Mouni, L.; Bollinger, J.C.; Belkhiri, L.; Bouzaza, A.; Assadi, A.A.; Belkacemi, H. Modeling and Optimization
of Process Parameters in Elucidating the Adsorption Mechanism of Gallic Acid on Activated Carbon Prepared from Date Stones.
Sep. Sci. Technol. 2020, 55, 3113–3125. [CrossRef]
21. Zamouche, M.; Mouni, L.; Ayachi, A.; Merniz, I. Use of Commercial Activated Carbon for the Purification of Synthetic Water
Polluted by a Pharmaceutical Product. Desalin. Water Treat. 2019, 172, 86–95. [CrossRef]
Water 2022, 14, 3885 19 of 20
22. Morton, J. Pomegranate. In Fruits of Warm Climates; J. F. Morton: Miami, FL, USA, 1987; pp. 352–355.
23. Ravikumar, K.V.G.; Sudakaran, S.V.; Ravichandran, K.; Pulimi, M.; Natarajan, C.; Mukherjee, A. Green Synthesis of NiFe Nano
Particles Using Punica Granatum Peel Extract for Tetracycline Removal. J. Clean. Prod. 2019, 210, 767–776. [CrossRef]
24. Khwairakpam, A.D.; Bordoloi, D.; Thakur, K.K.; Monisha, J. Possible Use of Punica Granatum (Pomegranate) in Cancer Therapy.
Pharmacol. Res. 2018, 133, 53–64. [CrossRef]
25. Ben-Ali, S. Application of Raw and Modified Pomegranate Peel for Wastewater Treatment: A Literature Overview and Analysis.
Int. J. Chem. Eng. 2021, 2021, 8840907. [CrossRef]
26. Benabderrahim, M.A.; Yahia, Y.; Bettaieb, I.; Elfalleh, W.; Nagaz, K. Antioxidant Activity and Phenolic Profile of a Collection of
Medicinal Plants from Tunisian Arid and Saharan Regions. Ind. Crops Prod. 2019, 138, 111427. [CrossRef]
27. Elbatanony, M.M.; El-Feky, A.M.; Hemdan, B.A.; El-Liethy, M.A. Assessment of the Antimicrobial Activity of the Lipoidal and
Pigment Extracts of Punica Granatum L. Leaves. Acta Ecol. Sin. 2019, 39, 89–94. [CrossRef]
28. Naeem, G.A.; Muslim, R.F.; Rabeea, M.A.; Owaid, M.N.; Abd-Alghafour, N.M. Punica Granatum L. Mesocarp-Assisted Rapid
Fabrication of Gold Nanoparticles and Characterization of Nano-Crystals. Environ. Nanotechnol. Monit. Manag. 2020, 14, 100390.
[CrossRef]
29. Tan, W.-K.; Lee, S.-Y.; Lee, W.-J.; Hee, Y.-Y.; Abedin, N.H.Z.; Abas, F.; Chong, G.-H. Supercritical Carbon Dioxide Extraction of
Pomegranate Peel-Seed Mixture: Yield and Modelling. J. Food Eng. 2021, 301, 110550. [CrossRef]
30. Msaadi, R.; Sassi, W.; Hihn, J.-Y.; Ammar, S.; Chehimi, M.M. Valorization of Pomegranate Peel Balls as Bioadsorbents of Methylene
Blue in Aqueous Media. Emergent Mater. 2021, 5, 381–390. [CrossRef]
31. Abbas, M.; Harrache, Z.; Trari, M. Mass-Transfer Processes in the Adsorption of Crystal Violet by Activated Carbon Derived from
Pomegranate Peels: Kinetics and Thermodynamic Studies. J. Eng. Fiber. Fabr. 2020, 15, 1558925020919847. [CrossRef]
32. Noli, F.; Avgerinou, A.; Kapashi, E.; Kapnisti, M. Uranium and Thorium Retention onto Sorbents from Raw and Modified
Pomegranate Peel. Water Air Soil Pollut. 2021, 232, 437. [CrossRef]
33. Kenessova, A.K.; Seilkhanova, G.A.; Rakhym, A.B.; Mastai, Y. Composite Materials Based on Orange and Pomegranate Peels for
Cu (II) and Zn (II) Ions Extraction. Int. J. Biol. Chem. 2020, 13, 154–160. [CrossRef]
34. Seliem, M.K.; Mobarak, M.; Selim, A.Q.; Mohamed, E.A.; Halfaya, R.A.; Gomaa, H.K.; Anastopoulos, I.; Giannakoudakis, D.A.;
Lima, E.C.; Bonilla-Petriciolet, A. 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. 2020, 311, 113169. [CrossRef]
35. Yi, Y.; Lv, J.; Liu, Y.; Wu, G. Synthesis and Application of Modified Litchi Peel for Removal of Hexavalent Chromium from
Aqueous Solutions. J. Mol. Liq. 2017, 225, 28–33. [CrossRef]
36. Rafiaee, S.; Samani, M.R.; Toghraie, D. Removal of Hexavalent Chromium from Aqueous Media Using Pomegranate Peels
Modified by Polymeric Coatings: Effects of Various Composite Synthesis Parameters. Synth. Met. 2020, 265, 116416. [CrossRef]
37. Chen, S.; Yue, Q.; Gao, B.; Li, Q.; Xu, X.; Fu, K. Dsorption of Hexavalent Chromium from Aqueous Solution by Modified Corn
Stalk: A Fixed-Bed Column StudyA. Bioresour. Technol. 2012, 113, 114–120. [CrossRef]
38. Hu, Q.; Pang, S.; Wang, D.; Yang, Y.; Liu, H. Deeper Insights into the Bohart–Adams Model in a Fixed-Bed Column. J. Phys. Chem.
B 2021, 125, 8494–8501. [CrossRef] [PubMed]
39. Nithya, K.; Sathish, A.; Kumar, P.S. Packed Bed Column Optimization and Modeling Studies for Removal of Chromium Ions
Using Chemically Modified Lantana Camara Adsorbent. J. Water Process Eng. 2020, 33, 101069. [CrossRef]
40. Vera, M.; Juela, D.M.; Cruzat, C.; Vanegas, E. Modeling and Computational Fluid Dynamic Simulation of Acetaminophen
Adsorption Using Sugarcane Bagasse. J. Environ. Chem. Eng. 2021, 9, 105056. [CrossRef]
41. Hu, Q.; Liu, H.; Zhang, Z.; Pei, X. Development of Fractal-like Clark Model in a Fixed-Bed Column. Sep. Purif. Technol. 2020, 251, 117396.
[CrossRef]
42. Turkmen Koc, S.N.; Kipcak, A.S.; Moroydor Derun, E.; Tugrul, N. Removal of Zinc from Wastewater Using Orange, Pineapple
and Pomegranate Peels. Int. J. Environ. Sci. Technol. 2021, 18, 2781–2791. [CrossRef]
43. Brebu, M.; Vasile, C. Thermal Degradation of Lignin—A Review. Cellul. Chem. Technol. 2010, 44, 353.
44. Bidhendi, M.E.; Poursorkh, Z.; Sereshti, H.; Nodeh, H.R.; Rezania, S.; Kamboh, M.A. Nano-Size Biomass Derived from
Pomegranate Peel for Enhanced Removal of Cefixime Antibiotic from Aqueous Media: Kinetic, Equilibrium and Thermo-
dynamic Study. Int. J. Environ. Res. Public Health 2020, 17, 4223. [CrossRef] [PubMed]
45. Han, X.; Zhang, Y.; Zheng, C.; Yu, X.; Li, S.; Wei, W. Enhanced Cr(VI) Removal from Water Using a Green Synthesized
Nanocrystalline Chlorapatite: Physicochemical Interpretations and Fixed-Bed Column Mathematical Model Study. Chemosphere
2021, 264, 128421. [CrossRef]
46. Liang, X.; Fan, X.; Li, R.; Li, S.; Shen, S.; Hu, D. Efficient Removal of Cr (VI) from Water by Quaternized Chitin/Branched
Polyethylenimine Biosorbent with Hierarchical Pore Structure. Bioresour. Technol. 2018, 250, 178–184. [CrossRef] [PubMed]
47. Lahmar, A.; Hattab, Z.; Zerdoum, R.; Boutemine, N.; Djellabi, R.; Filali, N.; Guerfi, K. Dynamic Sorption of Hexavalent Chromium
Using Sustainable Low-Cost Eggshell Membrane. Desalin. Water Treat. 2020, 181, 289–299. [CrossRef]
48. Djelloul, C.; Hamdaoui, O. Dynamic Adsorption of Methylene Blue by Melon Peel in Fixed-Bed Columns. Desalin. Water Treat.
2015, 56, 2966–2975. [CrossRef]
49. Dovi, E.; Aryee, A.A.; Kani, A.N.; Mpatani, F.M.; Li, J.; Qu, L.; Han, R. High-Capacity Amino-Functionalized Walnut Shell for
Efficient Removal of Toxic Hexavalent Chromium Ions in Batch and Column Mode. J. Environ. Chem. Eng. 2022, 10, 107292.
[CrossRef]
Water 2022, 14, 3885 20 of 20
50. Luo, Z.; Guo, M.; Jiang, H.; Geng, W.; Wei, W.; Lian, Z. Plasma Polymerization Mediated Construction of Surface Ion-Imprinted
Polypropylene Fibers for the Selective Adsorption of Cr(VI). React. Funct. Polym. 2020, 150, 104552. [CrossRef]
51. Lin, S.-H.; Juang, R.-S. Adsorption of Phenol and Its Derivatives from Water Using Synthetic Resins and Low-Cost Natural
Adsorbents: A Review. J. Environ. Manage. 2009, 90, 1336–1349. [CrossRef]
52. Lima, E.C.; Gomes, A.A.; Tran, H.N. Comparison of the Nonlinear and Linear Forms of the van’t Hoff Equation for Calculation of
Adsorption Thermodynamic Parameters (∆ S◦ and ∆H◦). J. Mol. Liq. 2020, 311, 113315. [CrossRef]
53. Singh, S.; Anil, A.G.; Naik, T.S.S.K.; Basavaraju, U.; Khasnabis, S.; Nath, B.; Kumar, V.; Subramanian, S.; Singh, J.; Ramamurthy, P.C.
Mechanism and Kinetics of Cr (VI) Adsorption on Biochar Derived from Citrobacter Freundii under Different Pyrolysis Tempera-
tures. J. Water Process Eng. 2022, 47, 102723. [CrossRef]
54. Van Nguyen, N.; Lee, J.c.; Jeong, J.; Pandey, B.D. Enhancing the Adsorption of Chromium(VI) from the Acidic Chloride Media
Using Solvent Impregnated Resin (SIR). Chem. Eng. J. 2013, 219, 174–182. [CrossRef]
55. Chowdhury, Z.Z.; Zain, S.M.; Rashid, A.K.; Rafique, R.F.; Khalid, K. Breakthrough Curve Analysis for Column Dynamics Sorption of
Mn(II) Ions from Wastewater by Using Mangostana Garcinia Peel-Based Granular-Activated Carbon. J. Chem. 2013, 2013, 959761.
[CrossRef]
56. Egbosiuba, T.C.; Abdulkareem, A.S.; Kovo, A.S.; Afolabi, E.A.; Tijani, J.O.; Bankole, M.T.; Bo, S.; Roos, W.D. Adsorption of Cr(VI),
Ni(II), Fe(II) and Cd(II) Ions by KIAgNPs Decorated MWCNTs in a Batch and Fixed Bed Process. Sci. Rep. 2021, 11, 75. [CrossRef]
[PubMed]
57. Unuabonah, E.I.; El-Khaiary, M.I.; Olu-Owolabi, B.I.; Adebowale, K.O. Predicting the Dynamics and Performance of a Polymer-
Clay Based Composite in a Fixed Bed System for the Removal of Lead (II) Ion. Chem. Eng. Res. Des. 2012, 90, 1105–1115.
[CrossRef]

More Related Content

Similar to 2022 - Adsorption of Cr(VI) by Mesoporous Pomegranate Peel Biowaste from Synthetic Wastewater under Dynamic Mode.pdf

A comparative study and kinetics for the removal of hexavalent
A comparative study and kinetics for the removal of hexavalentA comparative study and kinetics for the removal of hexavalent
A comparative study and kinetics for the removal of hexavalent
Alexander Decker
 
Defence PPT.pptx
Defence PPT.pptxDefence PPT.pptx
Defence PPT.pptx
T-Munesh
 
Karim712015IRJPAC16163
Karim712015IRJPAC16163Karim712015IRJPAC16163
Karim712015IRJPAC16163
Ankit Singh
 
Evaluation of Collagen-Polyurethane-Chitosan Hydrogels for Lead Ions Removal ...
Evaluation of Collagen-Polyurethane-Chitosan Hydrogels for Lead Ions Removal ...Evaluation of Collagen-Polyurethane-Chitosan Hydrogels for Lead Ions Removal ...
Evaluation of Collagen-Polyurethane-Chitosan Hydrogels for Lead Ions Removal ...
Associate Professor in VSB Coimbatore
 
Sorption kinetics and equilibrium studies on the removal of toxic cr(vi) ions...
Sorption kinetics and equilibrium studies on the removal of toxic cr(vi) ions...Sorption kinetics and equilibrium studies on the removal of toxic cr(vi) ions...
Sorption kinetics and equilibrium studies on the removal of toxic cr(vi) ions...
Alexander Decker
 

Similar to 2022 - Adsorption of Cr(VI) by Mesoporous Pomegranate Peel Biowaste from Synthetic Wastewater under Dynamic Mode.pdf (20)

A comparative study and kinetics for the removal of hexavalent
A comparative study and kinetics for the removal of hexavalentA comparative study and kinetics for the removal of hexavalent
A comparative study and kinetics for the removal of hexavalent
 
F027035040
F027035040F027035040
F027035040
 
Statistical optimization of adsorption variables for biosorption of chromium ...
Statistical optimization of adsorption variables for biosorption of chromium ...Statistical optimization of adsorption variables for biosorption of chromium ...
Statistical optimization of adsorption variables for biosorption of chromium ...
 
Kinetic, thermodynamic and equilibrium studies on removal of hexavalent chrom...
Kinetic, thermodynamic and equilibrium studies on removal of hexavalent chrom...Kinetic, thermodynamic and equilibrium studies on removal of hexavalent chrom...
Kinetic, thermodynamic and equilibrium studies on removal of hexavalent chrom...
 
Defence PPT.pptx
Defence PPT.pptxDefence PPT.pptx
Defence PPT.pptx
 
Research article
Research articleResearch article
Research article
 
R0460498105
R0460498105R0460498105
R0460498105
 
Karim712015IRJPAC16163
Karim712015IRJPAC16163Karim712015IRJPAC16163
Karim712015IRJPAC16163
 
Removal of Hexavalent Chromium by Adsorption using low-cost Adsorbents and Ac...
Removal of Hexavalent Chromium by Adsorption using low-cost Adsorbents and Ac...Removal of Hexavalent Chromium by Adsorption using low-cost Adsorbents and Ac...
Removal of Hexavalent Chromium by Adsorption using low-cost Adsorbents and Ac...
 
Removal of Cu(II) Ions from Aqueous Solutions by Adsorption Onto Activated Ca...
Removal of Cu(II) Ions from Aqueous Solutions by Adsorption Onto Activated Ca...Removal of Cu(II) Ions from Aqueous Solutions by Adsorption Onto Activated Ca...
Removal of Cu(II) Ions from Aqueous Solutions by Adsorption Onto Activated Ca...
 
Evaluation of Collagen-Polyurethane-Chitosan Hydrogels for Lead Ions Removal ...
Evaluation of Collagen-Polyurethane-Chitosan Hydrogels for Lead Ions Removal ...Evaluation of Collagen-Polyurethane-Chitosan Hydrogels for Lead Ions Removal ...
Evaluation of Collagen-Polyurethane-Chitosan Hydrogels for Lead Ions Removal ...
 
Gjesm148351451593800
Gjesm148351451593800Gjesm148351451593800
Gjesm148351451593800
 
Treatment of Effluent from Granite Cutting Plant by Using Natural Adsorbents ...
Treatment of Effluent from Granite Cutting Plant by Using Natural Adsorbents ...Treatment of Effluent from Granite Cutting Plant by Using Natural Adsorbents ...
Treatment of Effluent from Granite Cutting Plant by Using Natural Adsorbents ...
 
E041013242
E041013242E041013242
E041013242
 
Adsorption of mercury metal using modified vermicompost biochar
Adsorption of mercury metal using modified vermicompost biocharAdsorption of mercury metal using modified vermicompost biochar
Adsorption of mercury metal using modified vermicompost biochar
 
Vol. 1 (1), 2014, 12–22
Vol. 1 (1), 2014, 12–22Vol. 1 (1), 2014, 12–22
Vol. 1 (1), 2014, 12–22
 
Removal of Cr (VI) Using Low Cost Activated Carbon Developed By Agricultural ...
Removal of Cr (VI) Using Low Cost Activated Carbon Developed By Agricultural ...Removal of Cr (VI) Using Low Cost Activated Carbon Developed By Agricultural ...
Removal of Cr (VI) Using Low Cost Activated Carbon Developed By Agricultural ...
 
Parametric Studies on Detergent Using Low Cost Sorbent
Parametric Studies on Detergent Using Low Cost SorbentParametric Studies on Detergent Using Low Cost Sorbent
Parametric Studies on Detergent Using Low Cost Sorbent
 
Sorption kinetics and equilibrium studies on the removal of toxic cr(vi) ions...
Sorption kinetics and equilibrium studies on the removal of toxic cr(vi) ions...Sorption kinetics and equilibrium studies on the removal of toxic cr(vi) ions...
Sorption kinetics and equilibrium studies on the removal of toxic cr(vi) ions...
 
Sorption kinetics and equilibrium studies on the removal of toxic cr(vi) ions...
Sorption kinetics and equilibrium studies on the removal of toxic cr(vi) ions...Sorption kinetics and equilibrium studies on the removal of toxic cr(vi) ions...
Sorption kinetics and equilibrium studies on the removal of toxic cr(vi) ions...
 

More from DrChimie

More from DrChimie (7)

présl.pptx
présl.pptxprésl.pptx
présl.pptx
 
2018 - Flexible and porous cellulose aerogels-zeolitic imidazolate framework ...
2018 - Flexible and porous cellulose aerogels-zeolitic imidazolate framework ...2018 - Flexible and porous cellulose aerogels-zeolitic imidazolate framework ...
2018 - Flexible and porous cellulose aerogels-zeolitic imidazolate framework ...
 
2020 - A systematic study of hexavalent chromium adsorption and removal from ...
2020 - A systematic study of hexavalent chromium adsorption and removal from ...2020 - A systematic study of hexavalent chromium adsorption and removal from ...
2020 - A systematic study of hexavalent chromium adsorption and removal from ...
 
2017 - Cr(VI) Reduction and Immobilization by Core-Double-Shell Structured Ma...
2017 - Cr(VI) Reduction and Immobilization by Core-Double-Shell Structured Ma...2017 - Cr(VI) Reduction and Immobilization by Core-Double-Shell Structured Ma...
2017 - Cr(VI) Reduction and Immobilization by Core-Double-Shell Structured Ma...
 
2022 - Sequential modifications of chitosan biopolymer for enhanced confiscat...
2022 - Sequential modifications of chitosan biopolymer for enhanced confiscat...2022 - Sequential modifications of chitosan biopolymer for enhanced confiscat...
2022 - Sequential modifications of chitosan biopolymer for enhanced confiscat...
 
2020 - Adsorption and kinetic study of Cr(VI) on ZIF-8 based composites.pdf
2020 - Adsorption and kinetic study of Cr(VI) on ZIF-8 based composites.pdf2020 - Adsorption and kinetic study of Cr(VI) on ZIF-8 based composites.pdf
2020 - Adsorption and kinetic study of Cr(VI) on ZIF-8 based composites.pdf
 
2021 - Application of zeolitic imidazolate framework for hexavalent chromium ...
2021 - Application of zeolitic imidazolate framework for hexavalent chromium ...2021 - Application of zeolitic imidazolate framework for hexavalent chromium ...
2021 - Application of zeolitic imidazolate framework for hexavalent chromium ...
 

Recently uploaded

一比一原版SDSU毕业证圣地亚哥州立大学毕业证成绩单如何办理
一比一原版SDSU毕业证圣地亚哥州立大学毕业证成绩单如何办理一比一原版SDSU毕业证圣地亚哥州立大学毕业证成绩单如何办理
一比一原版SDSU毕业证圣地亚哥州立大学毕业证成绩单如何办理
psavhef
 
一比一原版BC毕业证波士顿学院毕业证成绩单如何办理
一比一原版BC毕业证波士顿学院毕业证成绩单如何办理一比一原版BC毕业证波士顿学院毕业证成绩单如何办理
一比一原版BC毕业证波士顿学院毕业证成绩单如何办理
amvovau
 
The Future of Autonomous Vehicles | civilthings.com | Detailed information
The Future of Autonomous Vehicles | civilthings.com |  Detailed informationThe Future of Autonomous Vehicles | civilthings.com |  Detailed information
The Future of Autonomous Vehicles | civilthings.com | Detailed information
gettygaming1
 

Recently uploaded (8)

Essential Maintenance Tips For Commercial Vans.
Essential Maintenance Tips For Commercial Vans.Essential Maintenance Tips For Commercial Vans.
Essential Maintenance Tips For Commercial Vans.
 
一比一原版SDSU毕业证圣地亚哥州立大学毕业证成绩单如何办理
一比一原版SDSU毕业证圣地亚哥州立大学毕业证成绩单如何办理一比一原版SDSU毕业证圣地亚哥州立大学毕业证成绩单如何办理
一比一原版SDSU毕业证圣地亚哥州立大学毕业证成绩单如何办理
 
一比一原版BC毕业证波士顿学院毕业证成绩单如何办理
一比一原版BC毕业证波士顿学院毕业证成绩单如何办理一比一原版BC毕业证波士顿学院毕业证成绩单如何办理
一比一原版BC毕业证波士顿学院毕业证成绩单如何办理
 
Advanced Technology for Auto Part Industry Inventory Solutions
Advanced Technology for Auto Part Industry Inventory SolutionsAdvanced Technology for Auto Part Industry Inventory Solutions
Advanced Technology for Auto Part Industry Inventory Solutions
 
gtyccccccccccccccccccccccccccccccccccccccccccccccccccccccc
gtycccccccccccccccccccccccccccccccccccccccccccccccccccccccgtyccccccccccccccccccccccccccccccccccccccccccccccccccccccc
gtyccccccccccccccccccccccccccccccccccccccccccccccccccccccc
 
Core technology of Hyundai Motor Group's EV platform 'E-GMP'
Core technology of Hyundai Motor Group's EV platform 'E-GMP'Core technology of Hyundai Motor Group's EV platform 'E-GMP'
Core technology of Hyundai Motor Group's EV platform 'E-GMP'
 
The Future of Autonomous Vehicles | civilthings.com | Detailed information
The Future of Autonomous Vehicles | civilthings.com |  Detailed informationThe Future of Autonomous Vehicles | civilthings.com |  Detailed information
The Future of Autonomous Vehicles | civilthings.com | Detailed information
 
Understanding BDA and BBMP: Key Players in Bangalore’s Urban Development
Understanding BDA and BBMP: Key Players in Bangalore’s Urban DevelopmentUnderstanding BDA and BBMP: Key Players in Bangalore’s Urban Development
Understanding BDA and BBMP: Key Players in Bangalore’s Urban Development
 

2022 - Adsorption of Cr(VI) by Mesoporous Pomegranate Peel Biowaste from Synthetic Wastewater under Dynamic Mode.pdf

  • 1. Citation: Boutaleb, Y.; Zerdoum, R.; Bensid, N.; Abumousa, R.A.; Hattab, Z.; Bououdina, M. Adsorption of Cr(VI) by Mesoporous Pomegranate Peel Biowaste from Synthetic Wastewater under Dynamic Mode. Water 2022, 14, 3885. https:// doi.org/10.3390/w14233885 Academic Editor: Dipti Prakash Mohapatra Received: 29 October 2022 Accepted: 22 November 2022 Published: 28 November 2022 Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affil- iations. Copyright: © 2022 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https:// creativecommons.org/licenses/by/ 4.0/). water Article Adsorption of Cr(VI) by Mesoporous Pomegranate Peel Biowaste from Synthetic Wastewater under Dynamic Mode Yassira Boutaleb 1, Radia Zerdoum 2, Nadia Bensid 1, Rasha A. Abumousa 3,* , Zhour Hattab 1 and Mohamed Bououdina 3 1 Laboratory of Water Treatment and Valorisation of Industrial Wastes, Department of Chemistry, Faculty of Sciences, Badji Mokhtar University, B.P.12, Annaba 23000, Algeria 2 Science and Technology Laboratory of Water and Environment, Department of Material Sciences, Faculty of Science and Technology, University Mohammed-Cherif Messadia, Souk Ahras 41000, Algeria 3 Department of Mathematics and Sciences, College of Humanities and Sciences, Prince Sultan University, Riyadh 11586, Saudi Arabia * Correspondence: rabumousa@psu.edu.sa Abstract: This study aims to eliminate hexavalent chromium Cr(VI) ions from water using pomegranate peel (PGP) powder. Dynamic measurements are carried out to examine the influence of the operating factors on the adsorption efficiency and kinetics. The analyzed PGP is found to be amorphous with relatively high stability, contains hydroxyl and carboxyl functional groups, a pH of zero charge of 3.9, and a specific surface-area of 40.38 m2/g. Adsorption tests indicate that PGP exhibits excel- lent removal effectiveness for Cr(VI) reaching 50.32 mg/g while the adsorption process obeys the Freundlich model. The thermodynamic study favors the exothermic physical adsorption process. The influence of operating parameters like the flow rate (1 to 3 mL/min), bed height (25 to 75 mm), concentration (10 to 30 mg/L), and temperature (298 to 318 K) on the adsorption process are investi- gated in column mode. To assess the performance characteristics of the column adsorption data, a non-linear regression has been used to fit and analyze four different kinetic and theoretical models, namely, Bohart-Adams, Thomas model, Clark, and Dose response. The obtained experimental results were found to obey the Dose Response model with a coefficient of regression R2 greater than 0.977. This study proved the excellent efficiency in the treatment of chemical industry effluents by using cost-effect abundant biowaste sorbent. This research demonstrated great efficacy in the treatment of chemical industrial effluents by using an abundant, cost-effective biowaste sorbent, thereby achieving the UN SDGs (UN Sustainable Development Goals) primary objective. Keywords: pomegranate peel; adsorption; fixed bed; batch; modeling 1. Introduction The contamination of wastewater by chromium Cr(VI) is a major problem due to its high toxicity and harmful effects on humans and animals [1]. The major effluents containing such toxic and hazardous element arises from industrial processes such as metallurgy, dyeing, and finishing industries [2]. Chromium exists in two major forms: Cr(III) and Cr(VI). The first form has low solubility and low reactivity resulting in low mobility in the environment and low toxicity in living organisms. However, Cr(VI) is a highly toxic [3] and carcinogenic [4] heavy metal when its concentration exceeds the standard limit fixed at 0.05 mg/L [5]. Its compounds include chromate (CrO4 2−), dichromate (Cr2O7 2−) and chromic acid (H2CrO4). CrO4 2− is predominant in basic solutions, H2CrO4 is predominant at pH < 1, while HCrO4 2−, and Cr2O7 2− are predominant at pH 2–6 [6]. Extensive studies were devoted toward the development of new cost-effective tech- nologies for the removal of Cr(VI) metal ions from aquatic environments, like precipita- tion [7], ion exchange [8], electro-coagulation [9], and adsorption [10–14]. Nevertheless, most of the technique’s present drawbacks and disadvantages, primarily high energy Water 2022, 14, 3885. https://doi.org/10.3390/w14233885 https://www.mdpi.com/journal/water
  • 2. Water 2022, 14, 3885 2 of 20 consumption and cost. In contrast, adsorption is an eco-friendly, cost-effective, and widely used technique for heavy metals remediation from aqueous solutions, which meets the UN SDGs (UN Sustainable Development Goals) main aim [15–17]. In recent decades, green chemistry has emerged as a potential fabrication route and attracted considerable attention in scientific research. Specifically, fruit wastes are being investigated as potential natural sorbents for heavy metal ions, thanks to their biodegrad- ability and cost-effectiveness, besides the presence of functional groups on adsorbent particles’ surface that are responsible for the reduction and elimination of contaminants and pollutants from aqueous media within a ‘sustainable development’ framework [18,19]. Valency, concentration, the ionic nature of substances in solution, pH of solution, and binding groups of sorbent material are key sorption parameters. The removal mechanism relies on the sorbate molecule’s size, concentration, affinity for the sorbent, bulk diffusion coefficient, and pore size distribution [20,21]. Sorbent’s physicochemical characteristics and sorption capacity determine the sorption’s efficacy. Pomegranates (Punica granatum) are widely cultivated in north/tropical Africa, and the Middle East, the Mediterranean Basin, the Caucasus region, the Indian subcontinent, Central Asia, and the drier parts of Southeast Asia [22]. Punica granatum presents a great variety of therapeutic bioactive species such as ellagitannins, proanthocyanidin compounds, phenolics, and flavonoids [23], which are well-known to facilitate the adsorption process of metal ions. Several studies have been conducted to highlight the biological and functional characteristics of pomegranate peel as it contains large amounts of polyphenols such as ellagic acid, gallic acid, and tan- nin acid [24]. Also, it possesses two important hydroxycinnamic acids and derivatives of flavones [25]. It has been widely studied as an antioxidant [26], antimicrobial [27], preparation of nanoparticles [28], anticancer activities [24], and extraction of bioactive compounds [29]. Additionally, several works have proposed the use of low-cost wastes for water and effluent treatment [30], particularly for the removal and biosorption of dyes [31], uranium [32], and heavy toxic metal ions especially chromium [33,34]. Extensive studies were performed on the effectiveness of biomass and its regenera- tion for Cr(VI) ions removal using pomegranates peel. Salam and Narayanan (2019) [19] reported a maximum removal of 100% achieved in 3 min for a low concentration of 20 mg/L and an amount of biosorbent (RPP with 125 µm size) of 0.5 g/L. Yi et al. (2017) [35] studied a modified Litchi peel as a biosorbent to eliminate Cr(VI), but the achieved capacity is relatively low reaching only 7.05 mg/g for a longer contact time of 100 min at 303 K for a concentration of 30 mg/g and an amount of biosorbent was 8 g/L. Rafiaee et al. (2020) [36] used pomegranate peel (PGP) adsorbent powder functionalized by polymeric coatings such as polypyrrole (PPy) and polyaniline (PANI) to reduce/remove Cr(VI)metal ions from wastewater. The highest achieved capacity was 59.47% in 90 min obtained in the batch sys- tem for 50 mg/L and biosorbent dose of 10 g/L. A new composite Anthracite/pomegranate peel synthesized by Seliem et al. exhibited a relatively high efficiency of 275 mg/g but for a longer time of 120 min at 25 ◦C and pH 3, the concentration of 70 mg/L and an amount of biosorbent 0.05 g. In this study, pomegranate peel (PGP) adsorbent has been prepared by low-cost and eco-friendly physical and chemical processes. Structural, morphological, and physic- ochemical studies are investigated. This research aims to study the surface chemistry characteristics and evaluate the performance of this biomaterial for the elimination of toxic and hazardous heavy metal ions. Indeed, particular emphasis is devoted to investigate the adsorption of hexavalent chromium ions Cr(VI) from an aqueous solution under dynamic mode rather than batch. Several operating parameters are examined, including the flow rate, the solution pH, adsorbent dose, contact time, bed height, as well as the solution temperature. To confirm the experimental data, several models have been applied for both batch and dynamic measurements, subsequently a plausible mechanism is proposed.
  • 3. Water 2022, 14, 3885 3 of 20 2. Materials and Methods The pomegranate (Punica granatum) peels (PGP) were collected from fresh juice ven- dors at a local market. To remove the impurities, the collected peels were thoroughly washed with tap water and rinsed using distilled water. Afterward, the obtained PGP product was dried at 100 ◦C for 24 h, grinded into a powdered sample then sieved to a particle size <500 µm. The used chemicals were purchased from Merck Company. A stock solution of 1000 mg/L was prepared using potassium dichromate K2Cr2O7, then different concentra- tions of chromium solution were obtained by consecutive dilution. Adsorption experi- ments in batch and dynamic modes were performed using a UV-vis spectrophotometer JENWAL 7315 (England) at maximum absorption of 540 nm after complexation with 1,5-diphenylcarbazide. 2.1. Characterization of the Powdered PGP To evaluate the morphology and chemical composition of PGP powder, a Quanta 200 FEI scanning electron microscope (SEM) coupled with energy-dispersive X-ray spectroscopy (EDX) was used. Fourier-transform infrared (FTIR) spectroscopy was used to identify the functional groups of PGP. FTIR spectrometer (IR Affinity-1S) equipped with a single reflection ATR, operating in the wave number range 400–4000 cm−1 acquired over 50 scans with a resolution of 4 cm−1. The phase stability of PGP was checked by X-ray diffraction (XRD) using a Rigaku Ultima IV instrument equipped with a CPS detector and CuKα radiation (λ = 1.5418 Å). A Thermogravimetric analyzer TGA-2050 equipment was used to analyze the thermal degradation profile of the PGP adsorbent up to 600 ◦C with a heating rate of 10 ◦C/min. The PGP the Brunauer Emmett Teller (BET) specific surface area and the average pore size were computed by recording Nitrogen adsorption-desorption curves at 70 K using a Quanta chrome NOVA 2200 E BET analyzer. 2.2. Biosorption of Cr(VI) on PGP—Batch and Dynamic Experiments 2.2.1. Batch Studies The adsorption isotherms were performed at varying Cr(IV) concentrations and the following operating conditions: m = 1.0 g/L, pH = 2 ± 0.2, T = 20 ◦C, stirring speed 50 rpm, contact time 2 h. After equilibrium, the adsorbent was separated from the suspension by a membrane filter (0.45 µm). The concentration of PGP was determined by analyzing the supernatant. The amount of the adsorbate, qe (mg/g), was estimated using Equation (1): qe = C0 − Ce m × V (1) where C0/Ce represents the initial/equilibrium concentrations of Cr (VI) (mg/L), V is the solution volume (L), and m is the PGP sorbent mass(g). Langmuir (Equation (2)) and Freundlich (Equation (3)) models were used to fit the equilibrium experimental adsorption data by means of the following expressions: qe = qmKLCe 1 + KLCe (2) qe = KfCe 1 n (3) where qe is the amount of Cr(VI) adsorbed onto PGP at equilibrium (mg/g), Ce is the Cr(VI) concentration in solution (mg/g), qm is the PGP capacity of the adsorbed Cr(VI) expressed per unit mass of adsorbent (mg/g), KL is the Langmuir constant (L/mg), while Kf and n correspond to Freundlich adsorption constant (mg 1−1/n L1/n) and intensity, respectively.
  • 4. Water 2022, 14, 3885 4 of 20 2.2.2. Dynamic Studies The tests were conducted in dynamic mode. The performance of a fixed column was investigated using breakthrough curves, presented by the concentrations of pollutants vs. the time profile curve in a fixed bed column. Packed-bed adsorption experiments were carried out in a cylindrical glass column (height = 35.0 cm, internal diameter = 11.0 cm) at 20 ◦C and an initial pH of 2 ± 0.2. Em- ploying a peristaltic pump (ISMATEC A39494), the PGP dose (0.105, 0.210, and 0.315 g) cor- responding to (25, 50, and 75 mm) bed heights and varying flow rates (1, 2, and 3 mL/min) were used for the removal of Cr(VI) (10, 20, and 30 mg/L). Using the breakthrough curves (breakthrough time—tb and exhaustion time—te), it is possible to obtain various informa- tion about the system such as the treated effluent volume (Veff in mL; Equation (4)), the total quantity of the adsorbed Cr(VI) (qtot in mg; Equation (5)), the amount of adsorbate that passes through the glass column(Wtotal in mg; Equation (6)), the percentage of adsorbate uptake(R in%; Equation (7)), and the uptake (qeq(exp) in mg/g; Equation (8)): Veff = Fte (4) qtot = FA 1000 = F 1000 Z t=total t=0 (C0 − Ct)dt (5) Wtot = C0Dttotal 1000 (6) R% = qtotal Wtotal × 100 (7) qeq(exp) = qtotal m (8) where F (mL/min) represents the flow rate, te (in min) the exhaustion time, C0 (mg/L) the initial adsorbate concentration, Ct (mg/L) the column exit concentration of Cr(VI) at time t, and m (g) the biosorbent quantity. The area under the breakthrough curve determined by integrating the Cad vs. t curve will be used to calculate the total amount of PGP adsorbed (maximum column capacity) [37]. The flow rate indicates the column’s empty bed contact time (EBCT), as given by Equation (9): EBCT = Bed Volume (mL) Flow Rate(mL/min) (9) Herein, it is important to mention that once the PGP sorbent becomes saturated, the fixed bed operation was stopped. The adsorption experiments were repeated at least in two replicates in order to estimate the error bars of the measurements. 2.3. Biosorption Modeling in Continuous Systems The Bohart-Adams (BA), Clark, modified Dose response, and Thomas models were adopted to characterize the breakthrough curves obtained in the continuous biosorption studies. All related information is reported in Table 1.
  • 5. Water 2022, 14, 3885 5 of 20 Table 1. The models used for modelling the adsorption data. Model Equation and Corresponding Plot Parameters References Bohart-Adams Ct C0 = exp KBA C0t − KBA N0 Z U Ct C0 vs. t (10) Ct C0 : final adsorbent concentration initial adsorbent concentration KBA (L/mg.min): sorption rate coefficient N0 (mg/L): fixed bed sorption capacity per unit volume Z(mm): bed–depth in the column U (mm/min): linear velocity, t(min): contact time [38] Thomas Ct C0 = 1 1+exp Kthqthm D −kthC0t Ct C0 vs. t (11) Ct C0 : final adsorbent concentration initial adsorbent concentration m(g): adsorbent mass Kth (L/mg.min): kinetic constant of Thomas model qth (mg/g): adsorption capacity t (min): contact time [39] Modified Dose Response Ct C0 = 1 − 1 1+ Veff b a q0 = b C0 m Ct C0 vs. t (12) Ct C0 : final adsorbent concentration initial adsorbent concentration Veff (mL): volume of the effluent q0 (mg/g): adsorption capacity a and b 6= 0: constant values in Dose Response model m (g): adsorbent mass [40] (13) Clarck Ct C0 = 1 1+Ae−rt ( 1 n−1 ) Ct C0 vs. t (14) A and r (min−1): constant of the Clarck model. t(min): contact time n: constant value of Freundlich model [41]
  • 6. Water 2022, 14, 3885 6 of 20 3. Results and Discussion 3.1. Characterizations The SEM images of PGP powder at the magnification (×10,000), as shown in Figure 1a,b, manifest flat and smooth surface texture of the PGP particles before Cr(VI) uptake. The corresponding elemental composition as determined by EDX reveals the main constituent’s carbon and oxygen with minor amounts of Al, Si, Cl, P, K, Ca, Mg, Cu, and Fe, see Figure 1d. This indicates that the two major elements forming PGP are carbon (51.40%) and oxygen (46.21%). Water 2022, 14, x FOR PEER REVIEW 6 of 24 3. Results and Discussion 3.1. Characterizations The SEM images of PGP powder at the magnification (×10,000), as shown in Figure 1a,b, manifest flat and smooth surface texture of the PGP particles before Cr(VI) uptake. The corresponding elemental composition as determined by EDX reveals the main con- stituent’s carbon and oxygen with minor amounts of Al, Si, Cl, P, K, Ca, Mg, Cu, and Fe, see Figure 1d. This indicates that the two major elements forming PGP are carbon (51.40%) and oxygen (46.21%). Figure 1. SEM images (a,b) before and (c) after Cr (VI) adsorption; EDX spectra (d) before and (e) after adsorption by PGP. The TGA analysis has been carried out to investigate the thermal properties of PGP, see Figure 2A. The mass loss reflects three main stages. The first stage (30–100 °C) corre- sponds to a weight loss of 10 wt.% owing to the evaporation of physisorbed water onto PGP. The second stage (150–250 °C) demonstrates a weight loss of 45 wt.% and is associ- ated with the thermal decomposition of hemicellulose and pectin [25]. While the third stage (250–600 °C) manifests an exothermic peak with a considerable weight loss of 30 wt.%, most probably originating from the separation of lignin and decomposition of the cellulose material [42,43]. The total mass loss at 600 °C is extremely high reaching 85 wt.%, hence confirming the organic nature of the PGP. Figure 1. SEM images (a,b) before and (c) after Cr (VI) adsorption; EDX spectra (d) before and (e); after adsorption by PGP. The TGA analysis has been carried out to investigate the thermal properties of PGP, see Figure 2A. The mass loss reflects three main stages. The first stage (30–100 ◦C) corresponds to a weight loss of 10 wt.% owing to the evaporation of physisorbed water onto PGP. The second stage (150–250 ◦C) demonstrates a weight loss of 45 wt.% and is associated with the thermal decomposition of hemicellulose and pectin [25]. While the third stage (250–600 ◦C) manifests an exothermic peak with a considerable weight loss of 30 wt.%, most probably originating from the separation of lignin and decomposition of the cellulose material [42,43]. The total mass loss at 600 ◦C is extremely high reaching 85 wt.%, hence confirming the organic nature of the PGP.
  • 7. Water 2022, 14, 3885 7 of 20 Water 2022, 14, x FOR PEER REVIEW 7 of 24 Figure 2. (A) TGA/DTA curves; (B) X−ray diffraction pattern; (C) FTIR spectra before and after ad- sorption; (D) N2 adsorption−desorption isotherms of PGP. The crystalline structure of PGP has been investigated by X−ray diffraction. As shown in Figure 2B, no characteristic pattern of crystalline phase(s) can be observed while a broad halo appears in the 2θ range (10−50°). This reflects the amorphous structure of PGP powder, in good agreement with the literature [44]. Figure 2C displays the FTIR spectra of PGP powder before and after Cr(VI) absorp- tion. The bands in the range 2800–3420 cm−1 are ascribed to −OH of carboxylic acid found in the benzene compound [28]. The bands located at 2850 and 2930 cm−1 correspond to C−H aliphatic (CH3) or (−CH2), while the band at 1593 cm−1 is assigned to the aromatic rings’ stretching of phenolic compounds (C−C and C=C aromatic vibrations). In addition, hydroxyl functional groups are defined by the absorption peak located at 1480 cm−1 and attributed to the OH of alcohols or in-plane vibration of the OH bond in carboxylic groups. Moreover, the absorption band at 1050 cm−1 indicates the C-O stretching of primary alco- hol groups. Furthermore, after the adsorption of Cr(VI), a relatively significant shift of the main band from 3200 to 3400 cm−1 is observed, hence confirming the occurrence of elec- trostatic interaction between PGP adsorbent and adsorbate. The band at 900 cm−1 indicates that the Cr-sorbet PGP could be due to the formation of Cr(OH)3, which suggests the re- duction of Cr(VI) to Cr(III). Figure 2. (A) TGA/DTA curves; (B) X−ray diffraction pattern; (C) FTIR spectra before and after adsorption; (D) N2 adsorption−desorption isotherms of PGP. The crystalline structure of PGP has been investigated by X−ray diffraction. As shown in Figure 2B, no characteristic pattern of crystalline phase(s) can be observed while a broad halo appears in the 2θ range (10−50◦). This reflects the amorphous structure of PGP powder, in good agreement with the literature [44]. Figure 2C displays the FTIR spectra of PGP powder before and after Cr(VI) absorption. The bands in the range 2800–3420 cm−1 are ascribed to −OH of carboxylic acid found in the benzene compound [28]. The bands located at 2850 and 2930 cm−1 correspond to C−H aliphatic (CH3) or (−CH2), while the band at 1593 cm−1 is assigned to the aromatic rings’ stretching of phenolic compounds (C−C and C=C aromatic vibrations). In addition, hydroxyl functional groups are defined by the absorption peak located at 1480 cm−1 and attributed to the OH of alcohols or in-plane vibration of the OH bond in carboxylic groups. Moreover, the absorption band at 1050 cm−1 indicates the C-O stretching of primary alcohol groups. Furthermore, after the adsorption of Cr(VI), a relatively significant shift of the main band from 3200 to 3400 cm−1 is observed, hence confirming the occurrence of electrostatic interaction between PGP adsorbent and adsorbate. The band at 900 cm−1 indicates that the Cr-sorbet PGP could be due to the formation of Cr(OH)3, which suggests the reduction of Cr(VI) to Cr(III). As shown in Figure 2D, the PGP N2 adsorption-desorption isotherm exhibits a type IV characteristic, in accordance with the International Union of Pure and Applied Chemistry (IUPAC) classification. The characteristic BET parameters given in Table 2, indicate a specific surface of 40.38 m2/g, an average pore diameter of 32.13 Å, and a pore volume of
  • 8. Water 2022, 14, 3885 8 of 20 0.00277 cm3/g. The distribution of pore diameter reveals that the adsorbent is mesoporous in nature, in good agreement with the literature [45]. Table 2. BET characteristics of the PGP powder. BET Surface Area (m2/g) Total Pore Volume (cm3/g) Average Pore Diameter (Å) Particle Density (g/cm3) PGP 40.38 2.77 × 10−3 32.13 0.25 The point of zero charge (pHpzc) of the PGP has been also determined. NaCl (0.01 N) solution is added to a series of flasks, and the initial pH is adjusted by the addition of NaOH (0.01 N, 0.1 N) and/or HCl (0.01 N, 0.1 N). After the pH adjustment, an amount of adsorbent (PGP) is mixed with aliquots of 15 mL. The containers have been placed and subjected to rigorous magnetic stirring for 48 h at 25 ◦C. The pHpzc value is then obtained by intersecting the (pHfinal−pHinitial) vs. pHinitial curve with the bisector (Figure 3a). The pHpzc value of PGP powder is found to be 3.9. This value suggests the presence of acid functional groups on the PGP particles’ surface, which corroborates with FTIR analysis, and indicates that PGP can adsorb Cr(VI) by electrostatic interaction. Water 2022, 14, x FOR PEER REVIEW 8 of 24 As shown in Figure 2D, the PGP N2 adsorption-desorption isotherm exhibits a type IV characteristic, in accordance with the International Union of Pure and Applied Chem- istry (IUPAC) classification. The characteristic BET parameters given in Table 2, indicate a specific surface of 40.38 m2/g, an average pore diameter of 32.13 Å , and a pore volume of 0.00277 cm3/g. The distribution of pore diameter reveals that the adsorbent is mesopo- rous in nature, in good agreement with the literature [45]. Table 2. BET characteristics of the PGP powder. BET Surface Area (m2/g) Total Pore Volume (cm3/g) Average Pore Diameter (Å) Particle Density (g/cm3) PGP 40.38 2.77 × 10−3 32.13 0.25 The point of zero charge (pHpzc) of the PGP has been also determined. NaCl (0.01 N) solution is added to a series of flasks, and the initial pH is adjusted by the addition of NaOH (0.01 N, 0.1 N) and/or HCl (0.01 N, 0.1 N). After the pH adjustment, an amount of adsorbent (PGP) is mixed with aliquots of 15 mL. The containers have been placed and subjected to rigorous magnetic stirring for 48 h at 25 °C. The pHpzc value is then obtained by intersecting the (pHfinal−pHinitial) vs. pHinitial curve with the bisector (Figure 3a). The pHpzc value of PGP powder is found to be 3.9. This value suggests the presence of acid functional groups on the PGP particles’ surface, which corroborates with FTIR analysis, and indicates that PGP can adsorb Cr(VI) by electrostatic interaction. Water 2022, 14, x FOR PEER REVIEW 9 of 24 Figure 3. (a) Plot for the determination of point zero charge of the PGP powder, (b) effect of pH on the adsorption capacity of Cr (VI) by PGP powder and (c) speciation diagram of Cr vs. pH (T = 20 ± 2°C, C0 = 20 mg/L, H= 2 cm, D= 2 mL/min). 3.1.1. Effect of pH Value The pH effect has been investigated between 2 and 12. It is observed that a marked removal rate of Cr(VI) ions is achieved at lower pH 2, then declines gradually with in- creasing the pH value. At acidic medium, the predominant species of chromium are di- chromate (Cr2O72−) and hydrogen chromate (HCrO4-) meanwhile the surface of PGP parti- cles becomes highly protonated, thereby favoring the adsorption and elimination of Cr(VI) metal ions. Under this condition, the phenolic groups (−OH) of the adsorbent easily exchange with the chromium species. Furthermore, the removal efficiency decreases sig- nificantly by increasing the pH to more than 4. The effect of pH value on the adsorption mechanism may be attributed to the presence of different functional carboxylic COO− and hydroxylic OH- groups onto the surface of PGP particles, hence favoring different possible interactions with Cr(VI) ions and other ionic species present in the solution. These results corroborate with zeta potential measurements which revealed a positive surface charge of PGP for pH smaller than 3.9 and a negative charge at pH greater to 3.9. The rise in pH value causes a reduction in the positive surface of the adsorbent [46]. In fact, this decrease is directly associated with the electrostatic forces existing between PGP and chromium solution, as consequence a marked reduction in the sorption capacity is observed [6]. Furthermore, as the pH value increases, a competition between OH- and 2− − Figure 3. (a) Plot for the determination of point zero charge of the PGP powder, (b) effect of pH on the adsorption capacity of Cr (VI) by PGP powder and (c) speciation diagram of Cr vs. pH (T = 20 ± 2 ◦C, C0 = 20 mg/L, H = 2 cm, D = 2 mL/min). 3.1.1. Effect of pH Value The pH effect has been investigated between 2 and 12. It is observed that a marked removal rate of Cr(VI) ions is achieved at lower pH 2, then declines gradually with increas- ing the pH value. At acidic medium, the predominant species of chromium are dichromate (Cr2O7 2−) and hydrogen chromate (HCrO4 −) meanwhile the surface of PGP particles be- comes highly protonated, thereby favoring the adsorption and elimination of Cr(VI) metal ions. Under this condition, the phenolic groups (−OH) of the adsorbent easily exchange with the chromium species. Furthermore, the removal efficiency decreases significantly by
  • 9. Water 2022, 14, 3885 9 of 20 increasing the pH to more than 4. The effect of pH value on the adsorption mechanism may be attributed to the presence of different functional carboxylic COO− and hydroxylic OH− groups onto the surface of PGP particles, hence favoring different possible interactions with Cr(VI) ions and other ionic species present in the solution. These results corroborate with zeta potential measurements which revealed a positive surface charge of PGP for pH smaller than 3.9 and a negative charge at pH greater to 3.9. The rise in pH value causes a reduction in the positive surface of the adsorbent [46]. In fact, this decrease is directly associated with the electrostatic forces existing between PGP and chromium solution, as consequence a marked reduction in the sorption capacity is observed [6]. Furthermore, as the pH value increases, a competition between OH− and CrO4 2− as predominant ions occurs in the basic medium [47]. It was concluded that HCrO4 − was adsorbed through electrostatic interaction to reduce to trivalent chromium Cr(III) [34]. The optimum pH of 2 will be utilized for the following experiments. The as-obtained results are found to be close to the results reported in a previous study using pomegranate peel for the elimination of hexavalent chromium [12]. The predominant form of Cr(VI) vs. pH (Figure 3a) is found to be HCrO4 − at the pH of 3.0 and would change to CrO4 2− and Cr2O7 2− when the pH 6.5. Compared with CrO4 2−, HCrO4 − is known to be more easily adsorbed on the surface active site owing to its low adsorption free energy [45]. 3.1.2. Effect of Contact Time Figure 4a shows the effect of contact time on Cr(VI) removal from aqueous solution onto PGP adsorbent. Based on the plot, the adsorption processes are almost accomplished during the first time and reach equilibrium within 2 h (removal percentage = 92%). After a certain period of time, the reaction slows down as the process reaching the steady state, because the number of vacant sites decreases due to the repulsive forces occurring between the solute molecules and the solid adsorbent phase [20]. Water 2022, 14, x FOR PEER REVIEW 10 of 24 3.1.2. Effect of Contact Time Figure 4a shows the effect of contact time on Cr(VI) removal from aqueous solution onto PGP adsorbent. Based on the plot, the adsorption processes are almost accomplished during the first time and reach equilibrium within 2 h (removal percentage = 92%). After a certain period of time, the reaction slows down as the process reaching the steady state, because the number of vacant sites decreases due to the repulsive forces occurring be- tween the solute molecules and the solid adsorbent phase [20]. Figure 4. (a) effect of contact time on Cr(VI) removal, (b) Langmuir and Freundlich plots for the adsorption on PGP at different concentration of Cr(VI) (pH = 2 ± 0.2, T = 20 ± 2 °C, m = 1.0 g/L, agitation speed = 50 rpm). 3.2. Adsorption Isotherms The non-linear Freundlich and Langmuir isotherms are shown in Figure 4b. It is no- ticed that the equilibrium data are well described by Freundlich isotherm compared to Langmuir isotherm. The model parameters, R2 values and 𝜒2 are presented in Table 3. The obtained results indicate that the regression coefficients for Cr(VI) metal ions adsorption with Freundlich isotherm are relatively high (0.994) compared to Langmuir (0.990) iso- Figure 4. (a) effect of contact time on Cr(VI) removal, (b) Langmuir and Freundlich plots for the adsorption on PGP at different concentration of Cr(VI) (pH = 2 ± 0.2, T = 20 ± 2 ◦C, m = 1.0 g/L, agitation speed = 50 rpm).
  • 10. Water 2022, 14, 3885 10 of 20 3.2. Adsorption Isotherms The non-linear Freundlich and Langmuir isotherms are shown in Figure 4b. It is noticed that the equilibrium data are well described by Freundlich isotherm compared to Langmuir isotherm. The model parameters, R2 values and χ2 are presented in Table 3. The obtained results indicate that the regression coefficients for Cr(VI) metal ions adsorption with Freundlich isotherm are relatively high (0.994) compared to Langmuir (0.990) isotherm. Table 3. Langmuir and Freundlich isotherm models constants and determination coefficients for the adsorption of Cr(VI) by PGP powder. Langmuir Freundlich qm(exp) (mg/g) qm(theo) (mg/g) KL (L/mg) R2 χ2 n Kf (mg1−1/n.L1/n) R2 χ2 50.32 38.29 ± 2.58 0.08 ± 0.01 0.990 0.528 3.00 ± 0.07 4.49 ± 0.28 0.994 0.33 The value of 1/n for all adsorption curves is found to be less than unity, which reflects the favorable exchange characteristics between Cr(VI) and PGP. The presence of hydroxyls and carboxyls on the surface of PGP particles produces an irregularity in the distribution energy over the surface of the adsorbent and therefore confirms the suitability of the Freundlich isotherm. Furthermore, the batch mode’s isotherm constants are significantly greater than the fixed bed mode. This may be due to the influence of the solution flow rate which is equal to zero in the batch mode, i.e., the contact time between Cr(VI) and PGP approximates the infinity [48]. 3.3. Influence of Parameters on Fixed Bed 3.3.1. Flow Rate The breakthrough curves have been recorded at varying flow rates (1, 2, and 3 mL/min), while the remaining operating parameters are kept constant, i.e., inlet Cr(VI) concentration 20 mg/Land bed-depth 50 mm, see Figure 5 and the computed parameters are summarized in Table 4. It is observed that the sorption rate is reduced with the increase in the flow rate. The empty bed contact time (EBCT) decreases significantly from 108 up to 38 min and the exhaust time (equivalent to 90% of the Cr (VI) concentration) also decreases markedly from 550 to 168 min. The flow rate is found to have a relatively moderate effect on the adsorption capacity; as the flow rate increases, the amount of adsorbed Cr(VI) (qexp) decreases slightly from 28.67 to 24.85 mg/g. The observed decline in the adsorption capacity can be attributed to the shorter contact time for the adsorbent to interact with the solution, which is in agreement with the previous research findings [39]. Water 2022, 14, x FOR PEER REVIEW 11 of 24 The value of 1/n for all adsorption curves is found to be less than unity, which reflects the favorable exchange characteristics between Cr(VI) and PGP. The presence of hydrox- yls and carboxyls on the surface of PGP particles produces an irregularity in the distribu- tion energy over the surface of the adsorbent and therefore confirms the suitability of the Freundlich isotherm. Furthermore, the batch mode’s isotherm constants are significantly greater than the fixed bed mode. This may be due to the influence of the solution flow rate which is equal to zero in the batch mode, i.e., the contact time between Cr(VI) and PGP approximates the infinity [48]. 3.3. Influence of Parameters on Fixed Bed 3.3.1. Flow Rate The breakthrough curves have been recorded at varying flow rates (1, 2, and 3 mL/min), while the remaining operating parameters are kept constant, i.e., inlet Cr(VI) concentration 20 mg/Land bed-depth 50 mm, see Figure 5 and the computed parameters are summarized in Table 4. It is observed that the sorption rate is reduced with the in- crease in the flow rate. The empty bed contact time (EBCT) decreases significantly from 108 up to 38 min and the exhaust time (equivalent to 90% of the Cr (VI) concentration) also decreases markedly from 550 to 168 min. The flow rate is found to have a relatively moderate effect on the adsorption capacity; as the flow rate increases, the amount of ad- sorbed Cr(VI) (qexp) decreases slightly from 28.67 to 24.85 mg/g. The observed decline in the adsorption capacity can be attributed to the shorter contact time for the adsorbent to interact with the solution, which is in agreement with the previous research findings [39]. Figure 5. Cont.
  • 11. Water 2022, 14, 3885 11 of 20 Water 2022, 14, x FOR PEER REVIEW 12 of 24 Figure 5. Comparison of the experimental and predict breakthrough curves obtained at different (A) flow rights, (B) bed heights, and (C) concentrations according to the studied models for Cr(VI) adsorption by PGP powder (T = 20 ± 2 °C, pH = 2 ± 0.2). Table 4. Total removal percentage of Cr(VI), column uptake capacity, and column data parameters obtained at different bed heights, concentrations, and flow rates (T = 20 ± 2 °C, pH = 2 ± 0.2). F (mL/min) H (mm) C0 (mg/L) tb (min) te (min) Veff (L) Area 𝐀′ (mg/min.L) 𝐪𝐭𝐨𝐭 (mg) 𝐖𝐭𝐨𝐭 (mg) Total Removal (%) 𝐪𝐞𝐪(𝐞𝐱𝐩) (mg/g) 1 50 20 108 700 0.70 6087 6.08 14.00 43.47 28.67 2 50 20 45 320 0.64 2803 5.60 12.80 43.75 26.73 3 50 20 36 204 0.61 1746 5.23 12.24 42.79 24.85 2 25 20 24 201 0.40 1083 2.16 8.04 20.57 20.62 2 75 20 78 430 0.86 4187 8.37 17.20 48.66 26.62 2 50 10 72 385 0.77 1924 3.84 7.70 49.87 18.28 2 50 30 27 270 0.54 3151 6.30 16.20 38.88 30.00 3.3.2. Bed Height The influence of varying the bed-height from 25 to 75 mm on the behavior of the breakthrough curves has been investigated at a fixed adsorbent flow rate of 2 mL/min and an inlet Cr(VI) concentration of 20 mg/L. The recorded breakthrough curves are displayed in Figure 5B and the computed values are given in Table 4. It is clearly observed that as the bed height increases, the Cr(VI) removal efficiency and qeq(exp) increase simultane- ously, from 20.57 to 48.66% and 20.62 to 26.62 mg/g. Consequently, while investigating the effect of PGP bed height on the adsorption of 20 mg/L Cr(VI) ions, the breakpoint value has been fixed at the point where Ct C0 is equal to 0.076, 0.090, and 0.100 for a fixed bed height of 25, 50 and 75 mm respectively. It is found that with the increase of the bed height, the breakthrough curve slope decreases, resulting in a broader mass transfer [39]. For this reason, the specific surface area of the PGP biosorbent increases, consequently providing a large number of binding active sites thereby a higher adsorption rate of the column [49]. Also, it is observed that the breakthrough time increases significantly from 24 to 78 min with the increase in the bed height. This can be associated with the occurrence of rapid and larger mass transfer in the fixed bed mode [49]. Figure 5. Comparison of the experimental and predict breakthrough curves obtained at different (A) flow rights, (B) bed heights, and (C) concentrations according to the studied models for Cr(VI) adsorption by PGP powder (T = 20 ± 2 ◦C, pH = 2 ± 0.2). Table 4. Total removal percentage of Cr(VI), column uptake capacity, and column data parameters obtained at different bed heights, concentrations, and flow rates (T = 20 ± 2 ◦C, pH = 2 ± 0.2). F (mL/min) H (mm) C0 (mg/L) tb (min) te (min) Veff (L) Area A’ (mg/min.L) qtot (mg) Wtot (mg) Total Re- moval (%) qeq(exp) (mg/g) 1 50 20 108 700 0.70 6087 6.08 14.00 43.47 28.67 2 50 20 45 320 0.64 2803 5.60 12.80 43.75 26.73 3 50 20 36 204 0.61 1746 5.23 12.24 42.79 24.85 2 25 20 24 201 0.40 1083 2.16 8.04 20.57 20.62 2 75 20 78 430 0.86 4187 8.37 17.20 48.66 26.62 2 50 10 72 385 0.77 1924 3.84 7.70 49.87 18.28 2 50 30 27 270 0.54 3151 6.30 16.20 38.88 30.00 3.3.2. Bed Height The influence of varying the bed-height from 25 to 75 mm on the behavior of the breakthrough curves has been investigated at a fixed adsorbent flow rate of 2 mL/min and an inlet Cr(VI) concentration of 20 mg/L. The recorded breakthrough curves are displayed in Figure 5B and the computed values are given in Table 4. It is clearly observed that as the bed height increases, the Cr(VI) removal efficiency and qeq(exp) increase simultaneously, from 20.57 to 48.66% and 20.62 to 26.62 mg/g. Consequently, while investigating the effect of PGP bed height on the adsorption of 20 mg/L Cr(VI) ions, the breakpoint value has been fixed at the point where Ct C0 is equal to 0.076, 0.090, and 0.100 for a fixed bed height of 25, 50 and 75 mm respectively. It is found that with the increase of the bed height, the breakthrough curve slope decreases, resulting in a broader mass transfer [39]. For this reason, the specific surface area of the PGP biosorbent increases, consequently providing a large number of binding active sites thereby a higher adsorption rate of the column [49]. Also, it is observed that the breakthrough time increases significantly from 24 to 78 min with the increase in the bed height. This can be associated with the occurrence of rapid and larger mass transfer in the fixed bed mode [49]. 3.3.3. Initial Concentration Figure 5C illustrates the breakthrough curves of different inlet Cr(VI) concentra- tions onto PGP adsorbent obtained at a constant bed depth (50 mm) and fixed flowrate
  • 12. Water 2022, 14, 3885 12 of 20 (2 mL/min), the corresponding data are given in Table 4. It can be observed that the value of Ct C0 reaches 0.010, 0.039, and 0.067 min in the intervals 365, 300, and 230 min when the Cr(VI) inlet concentration varies from 10 to 30 mg/L. This indicates that the value of the breakthrough time (tb) is slightly reduced with the rise in the inlet concentration of Cr(VI). The maximum adsorption capacity at 10, 20, and 30 mg/L Cr(VI) is found to be 18.32, 26.73 and 30.01 mg/g, respectively. In contrast, the adsorption efficiency is reduced by almost 22% (from 49.87 to 38.88%). Furthermore, it is noted that when the Cr(VI) con- centration in the solution is high, the rate of Cr(VI) ions for the adsorptive sites has a more pronounced trend. Indeed, a higher concentration of Cr(VI) in the solution provides a random diffusion path [50] for sorbate ions onto the surface of PGP adsorbent. This may be explained by the enhanced driving force and the high mass-transfer flux from Cr(VI) solution toward PGP [51]. However, at lower inlet Cr(VI) concentration, the breakthrough curves become dispersed, and the binding sites tend to be slowly saturated. Indeed, similar results were reported for the biosorption of Cr(VI) by chemically surface-modified Lantana Camara sorbent [39]. 3.3.4. Temperature The influence of the medium temperature has been also investigated. The break- through curves recorded for different bed heights at various temperatures are shown in Figure 6, and the results of Dose Response Model are given in Table 5. It is noted that with the rise of temperature, the adsorption rate of Cr(VI) ions decreases whereas the rate constant increases. Also, it is noticed that the adsorption rate is clearly variable at the same bed height. This is maybe due to the desorption process occurring following the rise in the thermal energy and the weakening of the interactions of the intermolecular bonds between solution ions and active sites of the PGP adsorbent. This signifies that the PGP particles’ surface is exothermic in nature and that the adsorption mechanism is favored at low temperatures. The above results are in accordance with the results obtained during the thermodynamic study in batch mode; the exothermic reaction of Cr(VI) ions adsorption onto PGP will be confirmed. This result has been previously reported by M. Zamouch et al. [21]. Water 2022, 14, x FOR PEER REVIEW 14 of 24 Figure 6. Effect of bed height on the breakthrough curves at different temperatures: (a) 303 K and (b) 313 K (C0 = 20 mg/L, H = 50 mm, D = 2 mL/min, pH = 2 ± 0.2).3.4. Adsorption Thermodynamics. The adsorption Cr(VI) ions by PGP thermodynamic parameters in the temperature range 298–318 K, have been determined using Van’t Hoff equations [52]: ∆G° = −RTLnKC (15) ∆G° = ∆H° − T∆S° (16) KC = qe Ce (17) LnK = ∆S° − ∆H° (18) Figure 6. Effect of bed height on the breakthrough curves at different temperatures: (a) 303 K and (b) 313 K (C0 = 20 mg/L, H = 50 mm, D = 2 mL/min, pH = 2 ± 0.2).3.4. Adsorption Thermodynamics.
  • 13. Water 2022, 14, 3885 13 of 20 Table 5. Dose response parameters for Cr(VI) adsorption by PGP in fixed bed (C0 = 20 mg/L, pH = 2 ± 0.2, D = 2 mL/min). T (K) Z (mm) a b (mL) q0 (mg/g) qexp (mg/g) Total Removal (%) R2 χ×104 SD 303 25 2.24 ± 0.04 52.85 ± 0.51 10.06 10.15 38.00 0.997 2.76 0.01 50 2.75 ± 0.07 138.32± 1.47 13.17 13.38 44.00 0.996 5.94 0.02 75 2.92± 0.08 194.12 ± 2.12 13.98 13.93 44.12 0.996 5.20 0.02 313 25 1.28± 0.10 20.96 ± 1.67 3.99 3.89 30.06 0.927 38.50 0.06 50 2.20 ± 0.14 51.85 ± 1.65 4.93 4.90 33.09 0.977 27.80 0.05 75 3.00 ± 0.11 98.13 ± 1.38 6.23 6.20 34.10 0.994 9.17 0.03 The adsorption Cr(VI) ions by PGP thermodynamic parameters in the temperature range 298–318 K, have been determined using Van’t Hoff equations [52]: ∆G ◦ = −RTLnKC (15) ∆G ◦ = ∆H ◦ − T∆S ◦ (16) KC = qe Ce (17) LnKC = ∆S ◦ R − ∆H ◦ RT (18) where ∆G ◦ represents the Gibbs free energy (kJ/mol), R the gas constant (8.314 J/mol.K), T the absolute temperature (K), KC the sorption distribution coefficient (L/g), ∆H◦ the enthalpy (kJ/mol.), and ∆S ◦ the entropy (kJ/mol.K). To better elucidate the adsorption process mechanism of Cr(VI) by PGP biosorbent, the respective thermodynamic plots are depicted in Figure 7 while the computed fitting parameters are given in Table 6. The negative values of ∆H◦, ∆S◦, and ∆G◦ demonstrate that the adsorption process is spontaneous and exothermic, without the need for any external power source. Also, it signifies the randomness of adsorption of Cr(VI) onto PGP. The negative value of ∆S◦ corresponds to a decrease in the freedom of the adsorbed chemical species and describes the randomness of chemical species movement at the adsorbent-adsorbate interface during the sorption process. The negative values of ∆G◦ and high negative values of ∆S◦ with the rise of temperature manifest the favorable nature of the process. Whereas the observed decrease in ∆G◦ values with temperature signifies that the adsorption process occurs favorably at lower temperatures [53]. Similar results were also observed in the literature for Cr(VI) adsorption onto different adsorbents [6,54]. Water 2022, 14, x FOR PEER REVIEW 15 of 24 Figure 7. Plots of LnKC vs. 1/T for the estimation of the thermodynamic parameters for the adsorp- tion of Cr(VI) on PGP (pH = 2 ± 0.2, m = 1.0 g.L−1, C0 = 20 mg/L). Table 6. Thermodynamic parameters calculated for the adsorption of Cr(VI) onto PGP. Sample ∆𝐇° (kJ/mol) ∆𝐒° (J/mol.K) ∆𝐆° (kJ/mol) 298 K 308 K 318 K PGP −44.69 −139.28 −3.98 −2.30 −0.48 3.4. Modeling of Column Data The fitted curves are shown in Figure 5 and the models’ corresponding calculated parameters are presented in Table 7. The nonlinear plot Ct vs. t has been fitted in the ini- Figure 7. Plots of LnKC vs. 1/T for the estimation of the thermodynamic parameters for the adsorption of Cr(VI) on PGP (pH = 2 ± 0.2, m = 1.0 g.L−1, C0 = 20 mg/L).
  • 14. Water 2022, 14, 3885 14 of 20 Table 6. Thermodynamic parameters calculated for the adsorption of Cr(VI) onto PGP. Sample ∆H ◦ (kJ/mol) ∆S ◦ (J/mol.K) ∆G ◦ (kJ/mol) 298 K 308 K 318 K PGP −44.69 −139.28 −3.98 −2.30 −0.48 3.4. Modeling of Column Data The fitted curves are shown in Figure 5 and the models’ corresponding calculated parameters are presented in Table 7. The nonlinear plot Ct C0 vs. t has been fitted in the initial part of the curve (0–0.2) to compute the Bohart–Adams KBA and sorption capacity N0 parameters. It is found that the value of KBA increases from 4.16 × 10−3 to 7.92 × 10−3 with increasing the flow rate but decreases from 11.28 × 10−3 to 3.28 × 10−3 with the increase in the initial Cr(VI) concentration. Also, the value of N0 of the model is found to increase with increasing both Cr(VI) concentration and bed depth. This can be associated with the prevalence of external mass transfer activities occurring during the adsorption process within the column at the initial stage [55]. This corroborates with the results published by Han et al. [45] when employing green synthesized nanocrystalline chlorapatite for the sorption of Cr(VI). The Thomas model parameters namely kinetic coefficient Kth and sorption capacity qth, have been determined using a nonlinear regression for bed height, flow rate, and Cr(VI) concentration. From Table 7, it is noted that the value of Kth increases with increasing both bed height and flow rate. Also, the value of qth increases markedly (from 15.75 to 25.39 mg/g) with the rise of Cr(VI) concentration (from 10 to 30 mg/L), which is attributed to the high driving force between Cr(VI) ions and PGP sorbent. However, the increase in the Kth value with the increase in the flow rate, signifies that the kinetics of the overall process is dominated by the external mass transfer mechanism [56]. This suggests that the Thomas model is more favorable for describing the adsorption processes, because the limiting steps do not proceed through external/internal diffusion [37]. Similar results are found in the literature when using different sorbate-sorbent systems [54]. According to the results given in Table 7, the rate of mass transfer (r) increases gradu- ally with the rise of both flow rate and Cr(VI) concentration but decreases with the increase in bed height. Furthermore, the increase in the initial Cr(VI) concentration results in the enhancement of the driving force, favoring a better mass transfer of the adsorbate onto PGP particles’ surface. Nevertheless, with the rise in the bed height, an increase in the particles’ number alongside a reduction in the mass transfer rate occur concurrently [57]. The fitting results of the modified Dose Response parameters are also examined. The predicted breakthrough curves illustrated in Figure 5 evidenced a good agreement with the measurements. The obtained high value of the regression coefficient (R2) satisfies the validity of the model. The values of R2 obtained for different operating conditions are higher than 0.977 and χ2 smaller than 2.98 × 10−3, thus providing a better fitting for the Dose Response model compared to Thomas, Clarck, and Bohart-Adams models.
  • 15. Water 2022, 14, 3885 15 of 20 Table 7. Parameters of various models for Cr(VI) adsorption by PGP using non-linear regression. Experimental fixed conditions: T = 20 ± 2 ◦C, pH = 2 ± 0.2. Parameters Thomas Model Dose Response Model F (mL/min) H (mm) C0 (mg/L) qecal (mg/g) Total Removal (%) KThX103 (L/mg) R2 χ2 ×103 ±SD × 103 a b (mL) q0 (mg/g) R2 χ2 ×103 ±SD ×103 1 50 20 25.69 ± 0.37 43.47 0.59 ± 0.02 0.976 3.16 56.22 2.73 ± 0.07 269.78 ± 2.98 27.79 0.988 1.47 38.38 2 50 20 21.59 ± 0.54 43.75 1.16± 0.07 0.953 5.85 76.50 2.46 ± 0.07 226.74 ± 2.73 23.26 0.988 1.45 38.05 3 50 20 21.52 ± 0.34 42.79 2.04 ± 0.09 0.976 3.11 55.80 2.94 ± 0.06 226.04± 1.80 22.86 0.994 0.71 26.73 2 25 20 16.63 ± 0.44 20.57 3.70 ± 0.28 0.970 4.50 67.05 3.07 ± 0.11 87.34± 1.12 17.57 0.993 1.02 31.89 2 75 20 22.77 ± 0.56 48.66 0.72 ± 0.03 0.956 5.72 75.63 2.49± 0.08 358.64 ± 5.95 26.02 0.979 2.71 52.09 2 50 10 15.75 ± 0.42 49.87 1.53 ±0.09 0.952 6.34 79.63 2.50 ± 0.09 330.88 ± 5.75 18.02 0.977 2.98 54.57 2 50 30 25.39 ± 0.45 38.88 0.95 ±0.04 0.972 3.19 56.48 2.35 ± 0.06 177.37 ± 1.84 27.22 0.991 0.99 31.53 Clark model Bohart-Adams model D mL.min−1 H (mm) C0 (mg/L) A r × 103 (min−1) R2 χ2 ×103 ±SD ×103 KBA×103 (mL/mg.min) N0 (mg/L) R2 χ2 ×103 ±SD ×103 1 50 20 535.29 ± 139.02 17.31 ± 0.85 0.964 4.72 68.69 4.16 ± 0.36 6024.43 ± 88.76 0.900 0.26 16.32 2 50 20 184.20 ± 56.80 33.24 ± 2.64 0.933 8.44 91.87 4.24 ± 0.11 6689.07 ± 218.38 0.914 0.32 17.99 3 50 20 298.95 ± 89.87 56.11 ± 3.49 0.962 4.99 70.65 7.92 ±0.11 6839.09 ± 209.58 0.923 0.40 20.21 2 25 20 383.93 ± 188.25 103.12 ± 10.26 0.933 6.66 81.63 10.30± 1.52 6222.88 ± 213.18 0.946 0.37 19.39 2 75 20 206.66 ± 52.18 19.67 ± 1.16 0.941 7.76 88.12 3.30± 1.52 6721.11 ± 91.34 0.940 0.20 14.32 2 50 10 186.00 ± 42.24 20.76 ± 1.31 0.936 8.52 92.29 11.28 ± 1.16 4208.87 ± 73.41 0.923 0.27 16.49 2 50 30 154.18 ± 37.66 40.68 ± 2.51 0.957 4.91 70.10 3.28 ± 0.43 7191.59 ± 323.61 0.922 0.48 21.94
  • 16. Water 2022, 14, 3885 16 of 20 3.5. Adsorption Mechanism The SEM image of PGP after the adsorption of Cr(VI) metal ions (Figure 1d) reveals that the particles’ surface becomes more rough and uneven. This confirms that the pores of the PGP sorbent are occupied by the ions of the adsorbate. Furthermore, the EDX spectrum (Figure 1e) of PGP after adsorption reveals the presence of Cr ions in the adsorbent, where its content reaches 3.64 wt.%. The presence of 1.5 wt.% K, maybe a residue belonging to the potassium dichromate (K2Cr2O7) used during the preparation of Cr(VI) aqueous solutions. Through the analytical results of EDX (Figure 1e) and FTIR (Figure 2C) spectra, as well as the data obtained from modeling of the equilibrium isotherms and kinetics, a plausible mechanism related to the adsorption of Cr(VI) onto PGP is illustrated in Figure 8. It is noted that the process is primarily controlled by electrostatic interaction between negatively charged functional groups present at the PGP particles’ surface and the positively charged Cr(VI) metal ions. Meanwhile, the surface adsorption via diffusion to the pores of PGP improves the mass transfer and provides an adsorption pathway besides an area responsible for the reduction of Cr(VI) metal ions to less toxic Cr(III). Water 2022, 14, x FOR PEER REVIEW 19 of 24 Figure 8. Plausible mechanism for the adsorption of Cr(VI) by PGP powder. 3.6. Comparison with Other Sorbents in the Literature To evaluate the sorption efficacy of PGP toward Cr(VI), a comprehensive comparison with other materials such as solvent impregnated resin (SIR), pomegranate peel powder (PGP) subjected to polyaniline (PANI), and polypyrrole (PPy) surface functionalization Figure 8. Plausible mechanism for the adsorption of Cr(VI) by PGP powder.
  • 17. Water 2022, 14, 3885 17 of 20 3.6. Comparison with Other Sorbents in the Literature To evaluate the sorption efficacy of PGP toward Cr(VI), a comprehensive comparison with other materials such as solvent impregnated resin (SIR), pomegranate peel powder (PGP) subjected to polyaniline (PANI), and polypyrrole (PPy) surface functionalization and modified Lantana Camara, is performed as illustrated in Table 8. It can be observed that the proposed biosorbent exhibits a higher adsorption capacity than the previously reported sorbents; it is even higher than pomegranate peel activated with H2SO4. In addition, the obtained batch capacity is more significant than that of SIR and PGP for Cr(VI) removal. Nevertheless, the existence of some differences is more probably attributed to several factors, primarily the active sites and functional groups formed at the particles’ surfaces of the sorbents, besides the microstructure nature (surface area, porosity, agglomeration level) and the chemical formulation (intrinsic properties playing key role in terms of selectivity) of the prepared materials. Table 8. A comparison of adsorption capacity or removal percentage of Cr(VI) with different biosorbents. Adsorbent Qmax (mg/g)/R (%) Operating Conditions References Pomegranate peel (Batch study) 22.87 mg/g pH 2, dose 300 mg/L, contact time 30 min [12] Solvent impregnated resin (SIR)(Batch study) 28.2 0 mg/g pH (0.3–2), dose 1 mg/L, contact time 30 min [54] Pomegranate peel (PGP) powder modified by polyaniline (PANI) and polypyrrole (PPy) (Batch study) 47.59% pH 1, dose 10 mg/L, contact time 90 min [36] Nanocrystalline chlorapatite (ClAP) (Batch study) 63.47 mg/g pH 3, dose 0.1 mg/L, contact time 25 min [45] Pomegranate peel (PGP) activated (PGP-1N H2SO4) (Batch study) 28.28 mg/g pH 3, dose 4 mg/L, contact time 3 h [6] Modified Lantana Camara (Column study) 362.80 mg/g pH 1.5, adsorbent bed height 40 mm, Breakthrough time 1250 min [39] Pomegranate peel (Column study) (Batch study) 30.00 mg/g 50.32 mg/g pH 2, adsorbent bed height 50 mm, Breakthrough time 27 min pH 2, dose 1 mg/L, contact time 24 h This study 4. Conclusions In the present study Pomegranate peel has been used as cost-effective biosorbent for the removal of hazardous Cr(VI). The adsorption mechanism of Cr(VI) adsorbate by PGP nanosorbent is determined and the physicochemical proprieties of PGP are presented. Both batch and dynamic studies are performed, and the isotherms and thermodynamic results are analyzed. It is found that the adsorption process occurs through an exothermic mechanism. Several operating parameters (pH, initial Cr(VI) concentration, temperature, bed depth, and flow rate) are found to significantly affect the Cr(VI) uptake through a fixed- bed column. The maximum uptake of 43.78 mg/g is achieved at pH 2 with a breakthrough time. The experimental data in dynamic mode has been fitted using four theoretical models namely Bohart-Adams, Clark, Dose Response, and Thomas. The Dose Response model successfully predicts the breakthrough curves for Cr(VI) removal under optimum operating conditions (flow rate, bed depth, Cr(VI) concentration) with R2 0.97. The equilibrium data were found to obey Freundlich isotherm, thereby confirming the heterogeneous sorption of Cr(VI) onto PGP. The negative values of ∆H◦, ∆G◦, and ∆S◦ indicate an exothermic and spontaneous process. Author Contributions: Conceptualization: Z.H.; Methodology: Y.B., N.B.; Formal analysis and investigation: R.Z.; Writing—original draft preparation: R.Z., M.B.; Writing—review and editing: R.Z., M.B., R.A.A.; Resources: R.Z.; Supervision: M.B., R.A.A. All authors have read and agreed to the published version of the manuscript. Funding: This research was funded by Algerian Ministry of Higher Education, grant number B00L01UN230120200002.
  • 18. Water 2022, 14, 3885 18 of 20 Institutional Review Board Statement: Not applicable. Data Availability Statement: Data is available upon formal request to authors. Acknowledgments: The authors are thankful to Prince Sultan University for their support for paying the article processing charge of this publication. Conflicts of Interest: The authors have no competing interest to declare that are relevant to the content of this article. References 1. Laxmi, V.; Kaushik, G. Toxicity of Hexavalent Chromium in Environment, Health Threats, and Its Bioremediation and Detoxifica- tion from Tannery Wastewater for Environmental Safety. In Bioremediation of Industrial Waste for Environmental Safety; Springer: Berlin/Heidelberg, Germany, 2020; pp. 223–243. 2. Ina, S. Chromium, an Essential Nutrient and Pollutant: A Review. African J. Pure Appl. Chem. 2013, 7, 310–317. 3. Sharma, P.; Singh, S.P.; Parakh, S.K.; Tong, Y.W. Health Hazards of Hexavalent Chromium (Cr (VI)) and Its Microbial Reduction. Bioengineered 2022, 13, 4923–4938. [CrossRef] [PubMed] 4. Kazemi, A.; Esmaeilbeigi, M.; Sahebi, Z.; Ansari, A. Health Risk Assessment of Total Chromium in the Qanat as Historical Drinking Water Supplying System. Sci. Total Environ. 2022, 807, 150795. [CrossRef] [PubMed] 5. Bakshi, A.; Panigrahi, A.K. A Comprehensive Review on Chromium Induced Alterations in Fresh Water Fishes. Toxicol. Rep. 2018, 5, 440–447. [CrossRef] 6. Abdel-Galil, E.A.; Hussin, L.M.S.; El-Kenany, W.M. Adsorption of Cr (VI) from Aqueous Solutions onto Activated Pomegranate Peel Waste. Desalin. WATER Treat. 2021, 211, 250–266. [CrossRef] 7. Minas, F.; Chandravanshi, B.S.; Leta, S. Chemical Precipitation Method for Chromium Removal and Its Recovery from Tannery Wastewater in Ethiopia. Chem. Int. 2017, 3, 291–305. 8. Qu, J.; Wang, Y.; Tian, X.; Jiang, Z.; Deng, F.; Tao, Y.; Jiang, Q.; Wang, L.; Zhang, Y. KOH-Activated Porous Biochar with High Specific Surface Area for Adsorptive Removal of Chromium (VI) and Naphthalene from Water: Affecting Factors, Mechanisms and Reusability Exploration. J. Hazard. Mater. 2021, 401, 123292. [CrossRef] 9. Wang, Z.; Shen, Q.; Xue, J.; Guan, R.; Li, Q.; Liu, X.; Jia, H.; Wu, Y. 3D Hierarchically Porous NiO/NF Electrode for the Removal of Chromium (VI) from Wastewater by Electrocoagulation. Chem. Eng. J. 2020, 402, 126151. [CrossRef] 10. Fotsing, P.N.; Woumfo, E.D.; Mezghich, S.; Mignot, M.; Mofaddel, N.; Le Derf, F.; Vieillard, J. Surface Modification of Biomaterials Based on Cocoa Shell with Improved Nitrate and Cr (vi) Removal. RSC Adv. 2020, 10, 20009–20019. [CrossRef] 11. Dehghani, M.H.; Karri, R.R.; Lima, E.C.; Mahvi, A.H.; Nazmara, S.; Ghaedi, A.M.; Fazlzadeh, M.; Gholami, S. Regression and Mathematical Modeling of Fluoride Ion Adsorption from Contaminated Water Using a Magnetic Versatile Biomaterial Chelating Agent: Insight on Production Experimental Approaches, Mechanism and Effects of Potential Interferers. J. Mol. Liq. 2020, 315, 113653. 12. Giri, R.; Kumari, N.; Behera, M.; Sharma, A.; Kumar, S.; Kumar, N.; Singh, R. Adsorption of Hexavalent Chromium from Aqueous Solution Using Pomegranate Peel as Low-Cost Biosorbent. Environ. Sustain. 2021, 4, 401–417. [CrossRef] 13. Zhu, S.; Khan, M.A.; Kameda, T.; Xu, H.; Wang, F.; Xia, M.; Yoshioka, T. New Insights into the Capture Performance and Mechanism of Hazardous Metals Cr3+ and Cd2+ onto an Effective Layered Double Hydroxide Based Material. J. Hazard. Mater. 2022, 426, 128062. [CrossRef] [PubMed] 14. Bouchelkia, N.; Mouni, L.; Belkhiri, L.; Bouzaza, A.; Bollinger, J.C.; Madani, K.; Dahmoune, F. Removal of Lead(II) from Water Using Activated Carbon Developed from Jujube Stones, a Low-Cost Sorbent. Sep. Sci. Technol. 2016, 51, 1645–1653. [CrossRef] 15. Imessaoudene, A.; Cheikh, S.; Bollinger, J.C.; Belkhiri, L.; Tiri, A.; Bouzaza, A.; El Jery, A.; Assadi, A.; Amrane, A.; Mouni, L. Zeolite Waste Characterization and Use as Low-Cost, Ecofriendly, and Sustainable Material for Malachite Green and Methylene Blue Dyes Removal: Box–Behnken Design, Kinetics, and Thermodynamics. Appl. Sci. 2022, 12, 7587. [CrossRef] 16. Zhu, S.; Xia, M.; Chu, Y.; Khan, M.A.; Lei, W.; Wang, F.; Muhmood, T.; Wang, A. Applied Clay Science Adsorption and Desorption of Pb (II) on L -Lysine Modi Fi Ed Montmorillonite and the Simulation of Interlayer Structure. Appl. Clay Sci. 2019, 169, 40–47. [CrossRef] 17. Chestnut, A.; Pigment, S. Cu (II) Adsorption from Aqueous Solution onto Poly (Acrylic Acid/Chestnut Shell Pigment) Hydrogel. Water 2022, 14, 3500. [CrossRef] 18. Dai, Y.; Sun, Q.; Wang, W.; Lu, L.; Liu, M.; Li, J.; Yang, S.; Sun, Y.; Zhang, K.; Xu, J. Utilizations of Agricultural Waste as Adsorbent for the Removal of Contaminants: A Review. Chemosphere 2018, 211, 235–253. [CrossRef] 19. Salam, F.A.; Narayanan, A. Biosorption-a Case Study of Hexavalent Chromium Removal with Raw Pomegranate Peel. Desalin. Water Treat. 2019, 156, 278–291. [CrossRef] 20. Chedri Mammar, A.; Mouni, L.; Bollinger, J.C.; Belkhiri, L.; Bouzaza, A.; Assadi, A.A.; Belkacemi, H. Modeling and Optimization of Process Parameters in Elucidating the Adsorption Mechanism of Gallic Acid on Activated Carbon Prepared from Date Stones. Sep. Sci. Technol. 2020, 55, 3113–3125. [CrossRef] 21. Zamouche, M.; Mouni, L.; Ayachi, A.; Merniz, I. Use of Commercial Activated Carbon for the Purification of Synthetic Water Polluted by a Pharmaceutical Product. Desalin. Water Treat. 2019, 172, 86–95. [CrossRef]
  • 19. Water 2022, 14, 3885 19 of 20 22. Morton, J. Pomegranate. In Fruits of Warm Climates; J. F. Morton: Miami, FL, USA, 1987; pp. 352–355. 23. Ravikumar, K.V.G.; Sudakaran, S.V.; Ravichandran, K.; Pulimi, M.; Natarajan, C.; Mukherjee, A. Green Synthesis of NiFe Nano Particles Using Punica Granatum Peel Extract for Tetracycline Removal. J. Clean. Prod. 2019, 210, 767–776. [CrossRef] 24. Khwairakpam, A.D.; Bordoloi, D.; Thakur, K.K.; Monisha, J. Possible Use of Punica Granatum (Pomegranate) in Cancer Therapy. Pharmacol. Res. 2018, 133, 53–64. [CrossRef] 25. Ben-Ali, S. Application of Raw and Modified Pomegranate Peel for Wastewater Treatment: A Literature Overview and Analysis. Int. J. Chem. Eng. 2021, 2021, 8840907. [CrossRef] 26. Benabderrahim, M.A.; Yahia, Y.; Bettaieb, I.; Elfalleh, W.; Nagaz, K. Antioxidant Activity and Phenolic Profile of a Collection of Medicinal Plants from Tunisian Arid and Saharan Regions. Ind. Crops Prod. 2019, 138, 111427. [CrossRef] 27. Elbatanony, M.M.; El-Feky, A.M.; Hemdan, B.A.; El-Liethy, M.A. Assessment of the Antimicrobial Activity of the Lipoidal and Pigment Extracts of Punica Granatum L. Leaves. Acta Ecol. Sin. 2019, 39, 89–94. [CrossRef] 28. Naeem, G.A.; Muslim, R.F.; Rabeea, M.A.; Owaid, M.N.; Abd-Alghafour, N.M. Punica Granatum L. Mesocarp-Assisted Rapid Fabrication of Gold Nanoparticles and Characterization of Nano-Crystals. Environ. Nanotechnol. Monit. Manag. 2020, 14, 100390. [CrossRef] 29. Tan, W.-K.; Lee, S.-Y.; Lee, W.-J.; Hee, Y.-Y.; Abedin, N.H.Z.; Abas, F.; Chong, G.-H. Supercritical Carbon Dioxide Extraction of Pomegranate Peel-Seed Mixture: Yield and Modelling. J. Food Eng. 2021, 301, 110550. [CrossRef] 30. Msaadi, R.; Sassi, W.; Hihn, J.-Y.; Ammar, S.; Chehimi, M.M. Valorization of Pomegranate Peel Balls as Bioadsorbents of Methylene Blue in Aqueous Media. Emergent Mater. 2021, 5, 381–390. [CrossRef] 31. Abbas, M.; Harrache, Z.; Trari, M. Mass-Transfer Processes in the Adsorption of Crystal Violet by Activated Carbon Derived from Pomegranate Peels: Kinetics and Thermodynamic Studies. J. Eng. Fiber. Fabr. 2020, 15, 1558925020919847. [CrossRef] 32. Noli, F.; Avgerinou, A.; Kapashi, E.; Kapnisti, M. Uranium and Thorium Retention onto Sorbents from Raw and Modified Pomegranate Peel. Water Air Soil Pollut. 2021, 232, 437. [CrossRef] 33. Kenessova, A.K.; Seilkhanova, G.A.; Rakhym, A.B.; Mastai, Y. Composite Materials Based on Orange and Pomegranate Peels for Cu (II) and Zn (II) Ions Extraction. Int. J. Biol. Chem. 2020, 13, 154–160. [CrossRef] 34. Seliem, M.K.; Mobarak, M.; Selim, A.Q.; Mohamed, E.A.; Halfaya, R.A.; Gomaa, H.K.; Anastopoulos, I.; Giannakoudakis, D.A.; Lima, E.C.; Bonilla-Petriciolet, A. 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. 2020, 311, 113169. [CrossRef] 35. Yi, Y.; Lv, J.; Liu, Y.; Wu, G. Synthesis and Application of Modified Litchi Peel for Removal of Hexavalent Chromium from Aqueous Solutions. J. Mol. Liq. 2017, 225, 28–33. [CrossRef] 36. Rafiaee, S.; Samani, M.R.; Toghraie, D. Removal of Hexavalent Chromium from Aqueous Media Using Pomegranate Peels Modified by Polymeric Coatings: Effects of Various Composite Synthesis Parameters. Synth. Met. 2020, 265, 116416. [CrossRef] 37. Chen, S.; Yue, Q.; Gao, B.; Li, Q.; Xu, X.; Fu, K. Dsorption of Hexavalent Chromium from Aqueous Solution by Modified Corn Stalk: A Fixed-Bed Column StudyA. Bioresour. Technol. 2012, 113, 114–120. [CrossRef] 38. Hu, Q.; Pang, S.; Wang, D.; Yang, Y.; Liu, H. Deeper Insights into the Bohart–Adams Model in a Fixed-Bed Column. J. Phys. Chem. B 2021, 125, 8494–8501. [CrossRef] [PubMed] 39. Nithya, K.; Sathish, A.; Kumar, P.S. Packed Bed Column Optimization and Modeling Studies for Removal of Chromium Ions Using Chemically Modified Lantana Camara Adsorbent. J. Water Process Eng. 2020, 33, 101069. [CrossRef] 40. Vera, M.; Juela, D.M.; Cruzat, C.; Vanegas, E. Modeling and Computational Fluid Dynamic Simulation of Acetaminophen Adsorption Using Sugarcane Bagasse. J. Environ. Chem. Eng. 2021, 9, 105056. [CrossRef] 41. Hu, Q.; Liu, H.; Zhang, Z.; Pei, X. Development of Fractal-like Clark Model in a Fixed-Bed Column. Sep. Purif. Technol. 2020, 251, 117396. [CrossRef] 42. Turkmen Koc, S.N.; Kipcak, A.S.; Moroydor Derun, E.; Tugrul, N. Removal of Zinc from Wastewater Using Orange, Pineapple and Pomegranate Peels. Int. J. Environ. Sci. Technol. 2021, 18, 2781–2791. [CrossRef] 43. Brebu, M.; Vasile, C. Thermal Degradation of Lignin—A Review. Cellul. Chem. Technol. 2010, 44, 353. 44. Bidhendi, M.E.; Poursorkh, Z.; Sereshti, H.; Nodeh, H.R.; Rezania, S.; Kamboh, M.A. Nano-Size Biomass Derived from Pomegranate Peel for Enhanced Removal of Cefixime Antibiotic from Aqueous Media: Kinetic, Equilibrium and Thermo- dynamic Study. Int. J. Environ. Res. Public Health 2020, 17, 4223. [CrossRef] [PubMed] 45. Han, X.; Zhang, Y.; Zheng, C.; Yu, X.; Li, S.; Wei, W. Enhanced Cr(VI) Removal from Water Using a Green Synthesized Nanocrystalline Chlorapatite: Physicochemical Interpretations and Fixed-Bed Column Mathematical Model Study. Chemosphere 2021, 264, 128421. [CrossRef] 46. Liang, X.; Fan, X.; Li, R.; Li, S.; Shen, S.; Hu, D. Efficient Removal of Cr (VI) from Water by Quaternized Chitin/Branched Polyethylenimine Biosorbent with Hierarchical Pore Structure. Bioresour. Technol. 2018, 250, 178–184. [CrossRef] [PubMed] 47. Lahmar, A.; Hattab, Z.; Zerdoum, R.; Boutemine, N.; Djellabi, R.; Filali, N.; Guerfi, K. Dynamic Sorption of Hexavalent Chromium Using Sustainable Low-Cost Eggshell Membrane. Desalin. Water Treat. 2020, 181, 289–299. [CrossRef] 48. Djelloul, C.; Hamdaoui, O. Dynamic Adsorption of Methylene Blue by Melon Peel in Fixed-Bed Columns. Desalin. Water Treat. 2015, 56, 2966–2975. [CrossRef] 49. Dovi, E.; Aryee, A.A.; Kani, A.N.; Mpatani, F.M.; Li, J.; Qu, L.; Han, R. High-Capacity Amino-Functionalized Walnut Shell for Efficient Removal of Toxic Hexavalent Chromium Ions in Batch and Column Mode. J. Environ. Chem. Eng. 2022, 10, 107292. [CrossRef]
  • 20. Water 2022, 14, 3885 20 of 20 50. Luo, Z.; Guo, M.; Jiang, H.; Geng, W.; Wei, W.; Lian, Z. Plasma Polymerization Mediated Construction of Surface Ion-Imprinted Polypropylene Fibers for the Selective Adsorption of Cr(VI). React. Funct. Polym. 2020, 150, 104552. [CrossRef] 51. Lin, S.-H.; Juang, R.-S. Adsorption of Phenol and Its Derivatives from Water Using Synthetic Resins and Low-Cost Natural Adsorbents: A Review. J. Environ. Manage. 2009, 90, 1336–1349. [CrossRef] 52. Lima, E.C.; Gomes, A.A.; Tran, H.N. Comparison of the Nonlinear and Linear Forms of the van’t Hoff Equation for Calculation of Adsorption Thermodynamic Parameters (∆ S◦ and ∆H◦). J. Mol. Liq. 2020, 311, 113315. [CrossRef] 53. Singh, S.; Anil, A.G.; Naik, T.S.S.K.; Basavaraju, U.; Khasnabis, S.; Nath, B.; Kumar, V.; Subramanian, S.; Singh, J.; Ramamurthy, P.C. Mechanism and Kinetics of Cr (VI) Adsorption on Biochar Derived from Citrobacter Freundii under Different Pyrolysis Tempera- tures. J. Water Process Eng. 2022, 47, 102723. [CrossRef] 54. Van Nguyen, N.; Lee, J.c.; Jeong, J.; Pandey, B.D. Enhancing the Adsorption of Chromium(VI) from the Acidic Chloride Media Using Solvent Impregnated Resin (SIR). Chem. Eng. J. 2013, 219, 174–182. [CrossRef] 55. Chowdhury, Z.Z.; Zain, S.M.; Rashid, A.K.; Rafique, R.F.; Khalid, K. Breakthrough Curve Analysis for Column Dynamics Sorption of Mn(II) Ions from Wastewater by Using Mangostana Garcinia Peel-Based Granular-Activated Carbon. J. Chem. 2013, 2013, 959761. [CrossRef] 56. Egbosiuba, T.C.; Abdulkareem, A.S.; Kovo, A.S.; Afolabi, E.A.; Tijani, J.O.; Bankole, M.T.; Bo, S.; Roos, W.D. Adsorption of Cr(VI), Ni(II), Fe(II) and Cd(II) Ions by KIAgNPs Decorated MWCNTs in a Batch and Fixed Bed Process. Sci. Rep. 2021, 11, 75. [CrossRef] [PubMed] 57. Unuabonah, E.I.; El-Khaiary, M.I.; Olu-Owolabi, B.I.; Adebowale, K.O. Predicting the Dynamics and Performance of a Polymer- Clay Based Composite in a Fixed Bed System for the Removal of Lead (II) Ion. Chem. Eng. Res. Des. 2012, 90, 1105–1115. [CrossRef]