AKKARI-2023 [55].

Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Sustainable Chemistry and Pharmacy 31 (2023) 100924

Contents lists available at ScienceDirect

Sustainable Chemistry and Pharmacy


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

A sustainably produced hydrochar from pomegranate peels for the


purification of textile contaminants in an aqueous medium
Imane Akkari a, Lucas Spessato b, **, Zahra Graba a, Nacer Bezzi a,
Mohamed Mehdi Kaci c, *
a Materials Technology and Process Engineering Laboratory (LTMGP), University of Bejaia, 06000 Bejaia, Algeria
b Department of Chemistry, State University of Maringá, Av. Colombo 5790, Maringá, Paraná, Brazil
c Laboratory of Reaction Engineering, Faculty of Mechanical and Process Engineering (USTHB), BP 32, 16111, Algiers, Algeria

ARTICLE INFO ABSTRACT

Keywords: Hydrothermal Carbonization presents an easy, inexpensive, and eco-friendly method to convert
Hydrochar waste from natural resources into sustainable materials. This study used the Hydrothermal Car-
Pomegranate peels bonization (HTC) approach to make hydrochar from locally available pomegranate peels. Several
Adsorption analytical tools were employed for its characterization, including elemental analysis, X-ray Dif-
Basic red 46 dye
fraction, micro-Raman, Fourier Transform Infrared Spectroscopy, Scanning Electron Microscopy,
Textual analysis, pHpzc, and Boehm titration. The produced hydrochar performed well in remov-
ing cationic dye Basic Red 46 (BR46) from an aqueous medium, adsorbing 286.90 and
367.72 mg g−1 within only 5 and 60 min, respectively, at ideal circumstances. The Freundlich
model was the most appropriate for the equilibrium data, whilst the pseudo-second-order model
was found closest to the kinetic data (R2 = 0.980; Δqe = 7.934%). Per the thermodynamic
study, the dye adsorption was spontaneous (ΔG° < 0) and endothermic
(ΔH° = 32.777 kJ mol−1). Furthermore, the prepared hydrochar (HCPP) has revealed good
reusability, as it can be reused up to five times without losing its effectiveness. In summary, the
outcomes obtained revealed that HCPP could be an effective and eco-friendly adsorbent for the
treatment of colored wastewater.

1. Introduction
The rising industrialization of countries has resulted in massive waste output in agriculture and chemical areas, resulting in conta-
mination of soil and water resources (Gong et al., 2013). Cationic dyes are used in various sectors, including textiles, leather, and pa-
per (Necibi et al., 2021); their outpourings in the aquatic environment are a substantial pollutant source. Poisonous materials were
formed because of their decomposition, which absorbed the oxygen in the water. This has increased to the point where it now poses a
grave danger to people and the environment. Dyes provide an unpleasant hue to water because of their nature and the intricacy of
their organic structure (Akkari et al., 2022b), and they resist photochemical and biological attacks, reducing sunlight penetration.
Furthermore, their breakdown products, such as aromatic amines, are hazardous, mutagenic, and carcinogenic, resulting in increased
COD and BOD levels in aquatic resources (Atmani et al., 2022; Crini 2006; Fu and Viraraghavan 2001; Mahmoodi and Arami 2009;
Shabaan et al., 2020).

* Corresponding author. .
** Corresponding author. .
E-mail addresses: [email protected] (L. Spessato), [email protected] (M.M. Kaci).

https://doi.org/10.1016/j.scp.2022.100924
Received 11 October 2022; Received in revised form 11 November 2022; Accepted 26 November 2022
Available online 13 December 2022
2352-5541/© 2022 Published by Elsevier B.V.
I. Akkari et al. Sustainable Chemistry and Pharmacy 31 (2023) 100924

Dyes are required to be discarded from wastewater due to their toxicity. A range of approaches has been utilized to eliminate them
(Kaci et al. 2021, 2022; Ma et al., 2019). However, because of its affordability, simplicity, high efficiency, and adsorbent reusability,
the adsorption approach is one of the best procedures for treating dye effluents (El-Zahhar et al., 2014; Fayazi et al., 2015; Zhang et
al., 2016). As a result, it is critical to developing innovative adsorbents for removing organic dyes (Saleh 2022) like poly (ethylene di-
amine-trimesoyl chloride)-modified diatomite (Saleh et al., 2021), magnetic activated carbon from chestnut shell (Altintig et al.,
2018), magnetic AC/CeO2 nanocomposite (Tuzen et al., 2018), magnetic activated carbon from acorn shell (Altıntıg et al., 2017), zinc
oxide nanoparticles loaded activated carbon (Altıntıg et al., 2021) and biochar/PSF mixed matrix membrane (He et al., 2018). How-
ever, the search for other adsorbents with a cheaper initial cost of investing, a large capacity to adsorb, and good renewability re-
mains a hot topic and takes considerable attention.
Under subcritical water conditions, low oxygen, temperatures between 180 and 250 °C, and saturated pressure, biomass could be
thermochemically transformed into a solid product (hydrochar) that resembled coal over several hours. Hydrothermal carbonization
is the term for this process (HTC) (Funke and Ziegler 2010). This latter is said to be capable of treating lignocellulosic feedstock and
high moisture residues, like algae, cow and swine slurry, and sewage sludge, in addition to converting starch, glucose, lignin, and cel-
lulose into functional carbon compounds (Cao et al., 2011). Furthermore, due to minimizing energy consumption, moderate car-
bonization conditions, minimum hazardous by-products, and exothermic character, HTC is regarded as an environmentally accept-
able technique for the speedy and secure treatment and reuse of wastewater compared to pyrochar production (Chen et al., 2017).
Hydrochar is comparable to pyrochar since it has a high carbon content, porous character, multiple oxygenated functional groups
on the surface, a strong affinity for polar and nonpolar compounds, and chemical and biological resistance (Naisse et al., 2013). HTC's
technology has demonstrated the benefits of a fast, secure, inexpensive, and valued processing of liquid media. For the synthesis of
hydrochars, various materials have been employed as precursors, including coffee husks, cassava slag, banana peels, corn stalk, straw
mushroom, sugarcane bagasse and grape seed (Diaz et al., 2019; Jais et al., 2021; Lei et al., 2016; Ronix et al., 2017; Wu et al., 2020;
Zhou et al., 2017; Zulfajri et al., 2021). These hydrochars have been utilized to eliminate various water contaminants, and their effi-
cacy has been excellent.
Basic Red 46, which makes up around 10% of the world's output of dyes for the textile industry, is regarded as a stable harmful or-
ganic contamination in wastewater (Şentürk and Yıldız 2020). Several adsorbents have been recently investigated to remove it, such
as synthesized graphene oxide nanoadsorbent (Shoushtarian et al., 2020), activated carbon from Ziziphus lotus stones (Boudechiche
et al., 2019), activated pine sawdust (Şentürk and Yıldız 2020), Fe@graphite core-shell magnetic nanocomposite (Konicki et al.,
2018), Raw pomegranate peel (Akkari et al., 2021), Raw cactus fruit peel (Akkari et al., 2022a), activated clay (Mekatel et al., 2021)
and biochar prepared from Chrysanthemum morifolium Ramat straw (Yang et al., 2021). These materials have some drawbacks due
to their high cost and complex synthesis processes.
Our prior study described the usefulness of pomegranate peel as a biosorbent. In addition, it demonstrated good performance
(Akkari et al., 2021), which is why it was chosen as a precursor for the production of hydrochar in this work. This is especially impor-
tant considering that only a single recent work in the literature has reported the use of hydrochar based on a specified precursor for
removing Cu (II) ions (Sayğılı and Sayğılı 2021). Nevertheless, no prior study has been published that involves the removal of organic
contaminants over hydrochar made from pomegranate peels.
The preparation of a hydrochar derived from pomegranate peels via hydrothermal carbonization was taken out in this work to im-
prove the usage performance of natural resources. The suggested approach was simple, inexpensive, and environmentally friendly.
The prepared hydrochar was thoroughly characterized using elemental analysis, XRD, micro-Raman, FTIR, SEM, BET/BJH measure-
ments, pHpzc, and Boehm titration. It was then used as a promising adsorbent for removing Basic Red 46 dye under different experi-
mental conditions, including pH solution, adsorbent amount, stirring speed, temperature, starting concentration, and contact time, to
describe the optimal circumstances for the adsorption process. These also include isotherm and kinetic investigations, thermody-
namic parameters, and regeneration cycles.

2. Experimental part
2.1. Dye solution
Basic red 46 (BR46), a cationic dye delivered by a textile mill and utilized without additional purifying. First, a stock solution
(1.0 g.L−1) was intended. After that, it was diluted to the desired concentrations for test solutions. A pH meter was employed to adjust
pH using HCl or NaOH (0.1 mol.L−1) (BOECO BT-675). Fig. 1 shows the molecular dimensions of BR46, its molecular electrostatic po-
tential surfaces and the HOMO-LUMO orbitals, which were calculated operating General Atomic and Molecular Electronic Structure
System (GAMESS) software. The BR46 geometry was optimized with Density Functional Theory (DFT) using non-local hybrid func-
tional B3LYP and a basis set of 6-31 + G (d,p). The electronic structure was calculated using 3C Hartree-Fock approximation, and the
plot was obtained by Gabedit graphical interface.

2.2. Preparation of hydrochar


10 g of powdered pomegranate peels was drenched in 100 g of distilled water for 2 h. Hydrothermal carbonization (HTC) was per-
formed in an autoclave made of stainless steel that was warmed up to 220 °C for 20 h under autogenous pressure. Such operational
conditions were chosen based on prior hydrochar production methods from agricultural by-products already published in the litera-
ture (Román et al. 2012, 2018). The sample was then immediately cooled to ambient temperature, and the hydrochar (HCPP) was
gathered and rinsed before being dried at 100 °C overnight. A summary of the preparation of HCPP is schematized in Fig. 2.

2
I. Akkari et al. Sustainable Chemistry and Pharmacy 31 (2023) 100924

Fig. 1. Molecular dimensions of BR46 dye (C18H21N6O4S), its molecular electrostatic potential surfaces and the HOMO-LUMO orbitals.

2.3. Characterization of hydrochar


The elementary composition of HCPP was identified using an X-ray fluorescence instrument (Rigaku ZSX Primus IV).
The crystallographic analyses were brought out by powder X-ray diffraction (PANalytical Empyrean) with a Cu Kα radiation
source (= 0.15418 nm) and 2θ range from 10 to 80°.
To investigate the degree of structural disorder, micro-Raman spectroscopy was performed at 25 °C using a Senterra Bruker spec-
trometer with an argon laser (532 nm) as the radiation source and an acquisition rate of 60 scans s−1 among 4000 and 400 cm−1.
The surface chemical groups were scrutinized utilizing Fourier-transform infrared spectroscopy (FTIR). The FTIR spectrum was
obtained using a Nicolet IS5 Spectrophotometer (resolution: 0.4 cm−1; a scan rate: 40 scans. min−1).

3
I. Akkari et al. Sustainable Chemistry and Pharmacy 31 (2023) 100924

Fig. 2. A schematic representative for the preparation of HCPP.

The material's morphology was analyzed by Scanning Electron Microscopy through FEI Quanta 250 equipment.
The textural properties of the material were determined by N2 physisorption at 77 K using a Micrometrics ASAP surface area and a
pore size analyzer. The specific surface area of the material (SBET) was obtained by Brunauer-Emmett-Teller (BET) equation. The total
pore volume (Vt) was calculated by Barrett-Joyner-Halenda (BJH) method. The micropore (Vμ) and mesopore (Vm) volumes were ob-
tained from α-plot and BJH methods, respectively. The pore size distribution was computed via the BJH theory, while the average
pore size (APS) was obtained by the ratio 4SBET/VT.
To obtain the pH at the point of zero charges (pHPZC), several solutions of 50 mL of NaCl 0.01 mol-1 had their pH adjusted with
NaOH or HCl (0.01 mol.L−1) from 2 to 12. Then, the solutions were stirred with 0.15 g of the material HCPP in Erlenmeyer flasks for
24 h. After stirring time had elapsed, the final pH of each system was measured, and the results were plotted in a curve of pHfinal vs.
pHinitial. The pHPzc was determined as the intersection of the results with the bisector.
The quantification of surface functional groups was conducted by Boehm titration (Boehm 2002). For 48 h, 50 mL of NaOH,
Na2CO3, NaHCO3, and HCl (0.01 mol.L−1) were mixed with 0.5 g of HCPP. After that, these solutions were normalized using HCl or
NaOH (0.01 mol.L−1). On the basis that NaHCO3 solely titers carboxylic groups, Na2CO3 titers lactonic and carboxylic groups, while
NaOH titers phenolic, lactonic, and carboxylic groups, the quantity of acidic surface groups could be approximated. The quantity of
HCl and NaOH consumed by HCPP was used to quantify the surface basic and acid molecular groups.

2.4. Adsorption tests


The adsorption tests were accomplished using a 1.0 L batch system containing 500 mL of dye solution while studying the experi-
mental parameters (starting dye concentration = 200 mg.L−1 for 60 min). In addition, duration and starting dye concentration ef-
fects, as well as isotherms, kinetics, and thermodynamics, were investigated at optimal settings (pH 6, adsorbent amount of 1.2 g.L−1,
400 rpm, at 25 °C).
After each test, the systems were spun (6000 rpm; 10 min) and the remaining BR46 concentrations were quantified via a UV–vis
spectrophotometer (SHIMADZU UV-1800) at the maximum absorption wavelength (λmax = 531 nm).
Equations (1) and (2) were employed to calculate the adsorption uptake (q) and removal percentage (R):
( )
C0 − C (e,t) V
q(e,t) = (1)
M
( )
C0 − C (e,t)
R(e,t) = × 100 (2)
C0

Co (mg.L−1) is the dye concentration at the beginning, Ce and Ct (mg.L−1) represent, respectively, dye concentrations at the equilib-
rium and at time t, V (L) is the volume of the solution, and M (g) is the material dried mass.

2.5. Fitting of equilibrium and kinetic models


Equilibrium tests were performed for BR46 concentrations from 20 to 500 mg.L−1 at (pH: 6.0, 1.2 g.L−1 of HCPP, 400 rpm at 25 C
within 1 h). The exploration of the interactions among BR46 and HCPP's surface is possible by several isotherm models, including
Langmuir (1918), Freundlich (Freundlich (1906), Temkin (Temkin and Pyzhev 1940), and Dubinin-Radushkevich (D-R) (Dubinin
1960). Equations (3)–(6) present these models, respectively:

4
I. Akkari et al. Sustainable Chemistry and Pharmacy 31 (2023) 100924

qm KL Ce
qe = (3)
1 + KL C e

(4)
1
qe = KF . Cen

(5)
( )
qe = Bln ACe

1
( 𝐶𝑒 )
(6)
2
𝑒 −𝛽ԑ
qe = 𝑞𝐷𝑅 ; ԑ = 𝑅𝑇 𝑙 𝑛 1 +

In which qe (mg.g−1) presents the adsorbed quantity at equilibrium, the final adsorbate's concentration is Ce (mg.L−1), the monolayer
adsorption capacity is qm (mg.g−1), the Langmuir and Freundlich constants are KL (L.mg−1) and KF, the adsorption intensity is n, A is
the Temkin isotherm constant (L.mg−1) representing maximum binding energy, B is Temkin adsorption heat constant (J.mol−1), qDR is
the monolayer capacity of Dubinin-Radushkevich (mg.g−1), β is sorption energy constant, and ԑ is the equilibrium concentration
(Polanyi potential).
The PFO, PSO and Elovich kinetic models (Lagergren 1898; Ho and McKay 1999; Peers 1965) have been investigated to character-
ize the reaction order of the BR46 adsorption onto ACPP. Equations (7)–(9) yield these models, respectively:

qt = qe 1 − e−k1 t (7)
( )

k2 q2e t
qt = (8)
1 + k2 qe t

1
qt = (ln 𝛽 + lnt) (9)
𝛽

qt and qe represent the quantity of BR46 (mg.g−1) adsorbed respectively at time and at equilibrium, k1 represents of the PFO rate
constant (min−1) and k2 represents PSO's rate constant (g.mg−1. min−1). α (mg.g−1. min−1) is initial adsorption and β (g.mg−1) is
desorption rate constant.
The models' applicability was determined using the determination coefficient values (R2) and normalized standard deviations
(Δqe). Equation (10) was used to calculate the Δqe values (%):

   qe exp − qe cal 2

(10)

qe exp


Δqe = 100 x
N−1

qe exp and qe cal represent the experimental and calculated adsorption capacities (mg.g−1), and N is the number of adsorption essays.
The intraparticle diffusion model was utilized to explore the rate-limiting step of the process (Weber and Morris 1963); its linear
version is provided by equation (11):

qt = ki t1∕2 + C (11)

ki represents intraparticle diffusion's rate constant (mg.g−1. min−1/2).

2.6. Thermodynamics of adsorption


The computation of the thermodynamic settings like ΔG°, ΔH°, and ΔS° by equations (12) and (13) allows understanding of the
process:

ΔG° = −RTlnKe0 (12)

(13)
ΔS° ΔH°
ln Ke0 = −
R RT

Ke0 is the dimensionless thermodynamic equilibrium constant (Lima et al., 2019), R means the universal constant of gases
(∼8314 J mol−1. K−1), and T (K) represents the absolute temperature.

2.7. Desorption and reuse tests


It is necessary to evaluate the system's stability prior to implementing HCPP. This was accomplished with the help of 0.1 M hy-
drochloric acid, which is used to get rid of BR46 and regenerate HCPP in order to be ready for more adsorption cycles under ideal cir-
cumstances (pH 6.0, 1.2 g.L−1of HCPP, and 400 rpm at 25 °C within 1 h).

5
I. Akkari et al. Sustainable Chemistry and Pharmacy 31 (2023) 100924

3. Results and discussion


3.1. Characterization
Using HTC process, high carbon content hydrochars are produced. The carbon content in raw pomegranate peels was 30.6%
(Akkari et al., 2021), however, it increased to 65.4% after HTC (Table 1).
Based on the XRD pattern (Fig. 3), the peaks at around 2θ = 22° and 38° correspond to the positions of (002) and (100) graphitic
planes, implying that the prepared hydrochar is made of graphitic-like carbon (Tomul et al., 2020; Zhang et al., 2021). Cellulose has a
stereotyped crystal structure with 2θ peaks at 14, 24, and 30°, which correspond to crystal plane structures of (101), (002), and (040),
respectively (Pala et al., 2014; Singh et al., 2015).
Raman spectrum is shown in Fig. 4. It has been shown that the lattice vibration is primarily accountable for the appearance of two
bands: D (at 1372 cm−1) and G (1568 cm−1). The G-peak is credited to the crystalline graphite feature of the high-frequency phonons
of E2g symmetry at Brillouin zones. In contrast, the D peak comes from defects assigned to the A1g vibrational modes (Spessato et al.,
2019). The obtained ID/IG value (0.67) evidences that the HTC process can create defects on the material's surface. However, the
amount of such defects is low compared to the graphitic portion (Spessato et al., 2022).
The hydrochar's FTIR spectrum prior to the BR46 adsorption (Fig. 5) indicates the following peaks: The OH and N–H stretching
modes have a band at 3090 cm−1 (Liu et al., 2017), While the asymmetric stretching vibration of C–H peaks at 2940 and 2690 cm−1
(Niazi et al., 2018). At 2080 cm−1, –NH2 peak appears (Arief et al., 2008). (C=O) is connected with the peak at 1700 cm−1 (Li et al.,
2020). The CN stretching/NH bending modes are related to the peaks at 1510 and 1210 cm−1 (Baláž et al., 2016). The 1600 and
1040 cm−1 are associated with C=C and C–O (Saleh 2015; Li et al., 2017), while the peaks between 908 and 777 cm−1 are linked to
C–H deformation vibration. Meanwhile, almost all peaks have shifted positions upon adsorption, revealing that distinct functional
groups affect dye uptake.

Table 1
HCPP's characteristics.

Parameter Value

C (wt. %) 65.4
O (wt. %) 28.2
Others (wt. %) 6.4
BET surface area (m2.g−1) 5.0089
Total pore volume (cm3.g−1) 0.023913
Vμ (cm³.g−1) 0.000

Vm (cm³.g−1) 0.023913

APS (nm) 20.00


pHpzc 5.47
Carboxylic groups (mmol.g−1) 0.91
Lactonic groups (mmol.g−1) 0.52
Phenolic groups (mmol.g−1) 0.04
Total Surface acidity (mmol.g−1) 1.47
Total Surface basicity (mmol.g−1) 0.81

Fig. 3. XRD pattern of HCPP.

6
I. Akkari et al. Sustainable Chemistry and Pharmacy 31 (2023) 100924

Fig. 4. Raman spectrum of HCPP.

Fig. 5. FTIR spectra of HCPP before and after adsorption of BR46.

SEM images (Fig. 6) were used to assess the morphology of HCPP. The HCPP particles were uneven in size and shape before ad-
sorption, with a heterogeneous and rough surface suitable for trapping adsorbent molecules. However, the adsorbent particles
clumped together and the surface smoothed down, indicating that BR46 was taken up.
Fig. 7(a and b) manifests the HCPP's N2 adsorption-desorption isotherm and BJH pore size distribution. The isotherm could be
classed as type IV, referring to mesoporous material (Sing 1982; Wu et al., 2020). After hydrothermal treatment, the pore volume
and the surface area (Table 1) of pomegranate peel increased, respectively (0.001–0.023 cm3 g−1 and 1.046–5.008 m2 g−1). The
pore size distribution of the adsorbent has dimensions from 2.0 to 70.0 nm, with 35.67 nm as the maximum value of the distribu-
tion and APS equal to 20 nm, which can be accessible for adsorbate molecules.
The material's surface charge is neutral at pHpzc. HCCP's pHpzc was seen to be 5.47. (Fig. 7 (c) and Table 1). This means that at
pH < 5.47, the surface functional groups are protonated, resulting in a positively charged surface. However, because those functional
groups deprotonate at pH > 5.47, the surface becomes negatively charged. Other works in the literature have indicated that hy-
drochars generally have low pH values (Sevilla and Fuertes 2009). Hydrochar's minor acidic characteristic is supported based on the
results of the Boehm titration (Table 1). HCPP has 1.47 mmol g−1 total acidity, divided into 0.04 mmol g−1 phenolic, 0.52 mmol g−1
lactonic and 0.91 mmol.g−1carboxylic groups; and 0.81 mmol g−1 basicity of the surface.

3.2. Adsorption parameters


3.2.1. Solution pH
pH impacts the adsorbent's surface charge and the ionization of adsorbed compounds, which influences the whole process. The in-
fluence of solution pH (2-10) was studied and depicted in Fig. 8 (a). Adsorption capacity advanced from 142.16 mg g−1
(R = 71.08%) to 168.63 mg g−1 (R = 84.37%) whenever the pH climbed from 2 to 6 and then dropped. This can be clarified based
on pHpzc (5.47) and BR46 pKa value. At pH < pHpzc, HCPP's surface obtains a positive charge because of the protonation of their func-
tional groups of HCPP; this prevents picking up the BR46 cations by electrostatic repulsion. As the pH solution rises, so does the num-
ber of OH- ions, favouring the deprotonation of HCPP's functional groups. This causes the surface of HCPP to become negatively

7
I. Akkari et al. Sustainable Chemistry and Pharmacy 31 (2023) 100924

Fig. 6. SEM image of HCPP before (a) and after adsorption (b).

Fig. 7. N2 adsorption/desorption isotherm of HCPP at 77 K (a), BJH pore size distribution (b) and pHpzc of HCPP (c).

8
I. Akkari et al. Sustainable Chemistry and Pharmacy 31 (2023) 100924

Fig. 8. The influence of different parameters on BR 46 adsorption onto HCPP: solution pH (a), adsorbent dosages (b), stirring speed (c), temperature (d) and contact
time and initial dye concentration (e).

charged and electrostatically attract dye cations, which grows the adsorption capacity (Konicki and Pełech, 2019; Lawal et al., 2019;
Shabaan et al., 2020). The overabundance of OH− ions surrounds the dye molecule at higher pH levels (pH > pka), preventing it from
being uptaken on the negatively charged surface of HCPP.

3.2.2. HCPP dosage


Fig. 8 (b) reveals the impact of HCPP's dose on BR46's uptake. As HCPP's mass increases (0.5–6 g.L−1), the elimination rate
gains from 77.80 to 91.25% due to the adsorption surface and active site availability (Lawal et al., 2019). After equilibrium, un-
saturated adsorption sites can explain the adsorption capacity drop from 311.21 to 30.41 mg g−1 (Bayuo et al., 2019; Değermenci
et al., 2019).

3.2.3. Stirring speed


Fig. 8 (c) displays the effect of stirring speed (100–700 rpm) on the BR46 adsorption. With the increase in agitation
(100–400 rpm), adsorption capacity increased from 145.60 mg g−1 (R = 87.36%) to 150.58 mg g−1 (R = 90.35%), and there-
after dropped. Increasing agitation speed ensures that the adsorbent particles and dye solution are homogeneous (Saxena et al.,
2020). Meanwhile, when agitation rates increase, vortices form, and a part of the adsorbent gets discarded on the flask borders,
causing the mix to become less homogenous. As a result of this, the adsorption capacity is reduced.

9
I. Akkari et al. Sustainable Chemistry and Pharmacy 31 (2023) 100924

3.2.4. Temperature
It was examined from 25 to 50 °C, and the outcomes are displayed in Fig. 8 (d). Temperature increase (25–50 °C) increased BR46's
uptake from 150.58 mg g−1 (R = 90.35%) to 160.54 mg g−1 (R = 96.32%), implying that the process is endothermic. Furthermore,
due to the lower viscosity of the solution, the strength of the adsorbate molecule scatters rate through the outside limit coating, and
the inside pores of the adsorbent's particles increase when the temperature rises. Therefore, increasing the temperature generally
boosts adsorption quantity because the mobility of the adsorbate molecule rises with the temperature (Güzel et al., 2015; Miyah et al.,
2018).

3.2.5. Time and initial BR46 concentration


Various starting dye concentrations (20–500 mg.L−1) were investigated at different span gaps ranging from 0 to 60 min, as de-
picted in Fig. 8 (e). It is worth mentioning that the adsorption pace was fast for the first 5 min, then reconciled and achieved equilib-
rium at around 60 min. A significant sum of vacant uptake sites is attainable initially, gradually decreasing this number (Chanzu et
al., 2012; Mahmoodi 2013; Muinde et al., 2020).
The increase in the starting BR46's concentration (20–500 mg.L−1) raised the uptake from 16.45 to 367.72 mg g−1. Raising the ini-
tial dye concentration raises the concentration gradient and overcomes mass transfer resistance (Magdy and Altaher 2018).

3.3. Adsorption isotherms


Fig. 9 depicts experimental adsorption data and the isotherm model's nonlinear fitting. Per Giles classification (Giles et al., 1960),
the isotherm profile is of type L (Langmuir), which implies that adsorbate molecules are horizontally adsorbed on the surface of the
adsorbent and that these molecules encounter minimal competition from solvent molecules.
The Langmuir, Freundlich, Temkin, and D-R models were adjusted to the experimental data to better understand the interactions
among BR46 molecules and the HCPP surface. The adsorption parameters, (R2), and (Δqe) are listed in Table 2.
The Freundlich model had an R2 value of 0.997 and the lowest Δqe value (13.489%), which was the most refined fit compared to
other models. This model describes multilayer adsorption processes on heterogeneous surfaces with varying adsorption intensities.
The active sites' occupancy degree is inversely proportional to the reduction in the affinity forces. Limited locations are available, and
priority is given to occupying those with the most potent attractions (Syafiuddin et al., 2018).
The Freundlich model's heterogeneity factor (n) could determine whether the process is linear (n = 1.0), chemical (n < 1.0), or
physical (n > 1.0). Furthermore, 1/n < 1.0 and 1/n > 1.0 mean normal Langmuir isotherms and cooperative adsorption, respec-
tively (Foo and Hameed 2010). The values of n and 1/n are 1.776 and 0.563, according to the results (Table 2), indicating both a
physical process and the favourable conditions of a typical Langmuir isotherm (Zanella et al., 2021).

3.4. Adsorption kinetics


Fig. 10 depicts the kinetic results for several initial BR46 concentrations, while Table 3 lists the estimated values. The PSO
model provides the best fit (R2 = 0.980 with the lowest Δqe value (7.934%)). Furthermore, the experimental and estimated ad-
sorption capacities that use this model are highly nigh, revealing that PSO kinetic model well describes BR46's uptake on HCPP.
The PSO model assumes that the quantity of BR46 adsorbed remains the same throughout time and that the overall number of
binding sites depends on the amount of adsorbate deposited at equilibrium (Beltrame et al., 2018). Furthermore, it could be con-
cluded from the kinetic results that adsorption is caused by chemisorption in addition to multilayer physical interactions discussed
in the equilibrium study.

Fig. 9. Adsorption isotherm of BR46 onto HCPP.

10
I. Akkari et al. Sustainable Chemistry and Pharmacy 31 (2023) 100924

Table 2
The values of parameters for each isotherm model used in the study.
Langmuir qm (mg.g−1) 471.698

KL (L.mg−1) 0.043

R2 0.879
Δqe (%) 42.981
Freundlich KF 31.195
1/n 0.563
n 1.776
R2 0.993
Δqe (%) 13.489
Temkin A (L.mg−1) 1.874
B (J.mol−1) 58.006
R2 0.860
Δqe (%) 198.924
Dubinin-Radushkevich qDR (mg.g−1) 122.004

β 1,338E-7
R2 0.763
Δqe (%) 82.697

The outcomes of the intraparticle diffusion model are reported in Fig. 11 (a) and Table 4. Three straight-line sections represent the
intraparticle diffusion model with various slopes in its most generic form. The steep section refers to the phase in which the solute and
the solution move toward the adsorbent's outer edge coating due to external surface adsorption (k1). The boundary layer control is cir-
cumvented whenever a straight line travels via the origin at the rate-limiting stage (Fig. 11 (a)), making it possible for complete pore
diffusion. The gradual section, on the other hand, refers to the gradual spreading of the solute onto the adsorbent (k2). Finally, the sta-
ble section signifies the point at which adsorption approaches equilibrium, and the amount of residual solute in the solution is so
small that it slows down the spread (k3) (Cheung et al., 2007; Poots et al., 1978). There were areas with a slight slope and flat sections,
as shown in Fig. 11 (a) and Table 4, implying that border movement outside the adsorbent was instantaneous (Shin and Kim 2016;
Tan et al., 2008). Instead, intraparticle diffusion occurred at the rate-limiting step (Kim and Kim 2019).
It is worth highlighting that the molecular dimensions of the adsorbate are shallow compared to the dimensions of the pores of the
adsorbent. For instance, the dye BR46 has (1.55 × 0.59 × 0.49) nm as molecular dimensions, while the pore sizes distribution of the
adsorbent has dimensions from 2.0 to 70.0 nm with 35.67 nm as a maximum value of the distribution, and APS equal to 20 nm. The
influence of molecular and pore dimensions on the diffusion mechanism can be understood in terms of the Knudsen number (Kn), as
reported by Spessato et al., (2021). The Kn is defined as the ratio between the mean free path of the molecule (λ) and the average pore
size obtained from the N2 physisorption results (20 nm). The λ for this study was calculated to be 3.8042 × 10−9 m. Thus, the Kn
value is determined to be 0.1902 (dimensionless). It is reported that if Kn < 1.0, the diffusion of the adsorbate into the pores of the
adsorbent is controlled by viscous diffusion, meaning that the fluid's velocity at any point of the system is constant or varies regularly.

3.5. Thermodynamic study


To set ΔH° and ΔS°, lnKe0 vs 1/T plot was utilized (Fig. 11 (b)), and Table 5 contains the estimated parameters. The negative val-
ues of (ΔG° < 0) suggest the spontaneity of the procedure and their values (<−20 kJ mol−1) confirm that the adsorption process
could be driven by chemical forces (Deniz and Saygideger 2011; Tan et al. 2016; Weng and Pan 2007). In addition, the positive value
of ΔH° (32.777 kJ mol−1) means that the process was endothermic, whereas ΔS° (0.184 kJ mol−1. K−1) means that disorder at the
solid-liquid contact surface is increasing.
To sum up, per the obtained results, the adsorption mechanism of BR46 onto HCPP could be governed by chemisorption in addi-
tion to multilayer physical sorption, including pore filling and electrostatic interactions. Moreover, intraparticle diffusion occurs at
the rate-limiting step, and the adsorbate's diffusion into the adsorbent's pores is controlled by viscous diffusion.

3.6. Reuse study


When developing applications on a large scale, it is indispensable to take into account the adsorbents' capacity to be long-lasting
and sustainable during the process. The adsorption capacity lowered from 367.72 mg g−1 (R = 88.25%) to 317.92 mg g−1
(R = 76.30%) after the fifth cycle (Fig. 11 (c)), which indicates that it is safe to reuse material in five cycles without adversely affect-
ing its effectiveness.
Compared to other adsorbents (Table 6), HCPP exhibited an excellent adsorption capacity of 367.72 mg g−1 with a disposal effi-
ciency of 88.25%, two outstanding performance indicators that hydrochar appear suitable for adsorbing several contaminants.

4. Conclusion
This study reports the valorization of pomegranate peels into a sustainable adsorbent of textile contaminant via the HTC proce-
dure, which presents a fast, secure, inexpensive, and green technique. The synthesized material was then characterized through ele-

11
I. Akkari et al. Sustainable Chemistry and Pharmacy 31 (2023) 100924

Fig. 10. Kinetic adsorption results for initial dye concentrations of 20 (a), 50 (b), 100 (c), 200 (d) and 500 mg.L−1 (e).

Table 3
Kinetic parameters.

Concentration qe exp Pseudo first-order model Pseudo second-order model Elovich model
(mg.L−1) (mg.g−1)
qe cal k1 R2 Δqe qe cal k2 (g.mg R2 Δqe α β R2 Δqe
(mg.g−1) (min−1) (%) (mg.g−1) −1.min−1) (%) (mg.g−1.min1) (g.mg−1) (%)

20 16.453 32.318 0.276 0.972 94.769 16.781 0.073 0.980 7.934 535.189 0.566 0.958 11.304
50 40.101 57.546 0.207 0.971 45.354 40.883 0.032 0.981 10.781 824.995 0.220 0.958 11.811
100 78.212 34.268 0.087 0.961 73.965 79.808 0.018 0.981 16.390 908.535 0.104 0.958 89.914
200 150.589 81.177 0.106 0.967 64.107 154.798 0.011 0.984 27.880 786.931 0.048 0.960 13.962
500 367.722 196.69 0.101 0.962 64.530 380.228 0.005 0.973 49.422 1128.956 0.017 0.939 24.553

12
I. Akkari et al. Sustainable Chemistry and Pharmacy 31 (2023) 100924

Fig. 11. Plot of intraparticle diffuion model (a), plot of Van't Hoff equation (b) and adsorption/desorption cycles for BR46 onto HCPP (c).

Table 4
Intraparticle diffusion rate constants.

Concentration (mg.L−1) Intraparticle diffusion rate constants (mg.g−1.min−1/2)

k1 R2 k2 R2 k3 R2

20 9.029 0.991 0.774 0.998 0.135 0.889


50 21.295 0.989 1.986 0.995 0.331 0.990
100 39.589 0.985 4.326 0.998 0.965 0.999
200 69.956 0.979 9.793 0.995 1.376 0.975
500 163.613 0.942 24.425 0.994 3.533 0.992

Table 5
Thermodynamic parameters at different temperatures.

T (K) lnKe0 ΔG° (kJ.mol−1) ΔH° (kJ.mol−1) ΔS° (kJ.mol−1.K−1)

298 8.962 −22.194 32.777 0.184


303 9.108 −22.933
313 9.466 −24.622
323 9.992 −26.822

mental analysis, XRD, micro-Raman, FTIR, SEM, BET/BJH, pHpzc, and Boehm titration. Batch tests were performed to find the suit-
able parameters for pH 6, 1.2 g. L−1 HCPP's mass and 400 rpm at 25 °C, acquiring an adsorption capacity of 286.90 mg g−1 in merely
5 min and 367.72 mg g−1 within 60 min. The Freundlich model described well the equilibrium curves, whereas the pseudo-second-
order model accurately fitted the findings. Thermodynamic investigations revealed that the adsorption process was spontaneous (ΔG°
<0) and endothermic (ΔH° = 32.777 kJ mol−1). Last, the manufactured HCPP was re-generable for five straight cycles without per-

13
I. Akkari et al. Sustainable Chemistry and Pharmacy 31 (2023) 100924

Table 6
Comparison of BR46 dye adsorption onto various adsorbents.

Adsorbent Adsorption capacity (mg.g−1) Reference

Synthesize graphene oxide nanoadsorbent 370.4 Shoushtarian et al. (2020)


Activated carbon from ziziphus lotus stones 307 Boudechiche et al. (2019)
Activated pine sawdust 312.5 Şentürk and Yıldız (2020)
Fe@graphite core-shell magnetic nanocomposite 46.7 Konicki et al. (2018)
Raw pomegranate peel 86.13 Akkari et al. (2021)
Raw cactus fruit peel 82.58 (Akkari et al., 2022a)
Activated clay 175 Mekatel et al. (2021)
Activated carbon from wild olive cores 781.25 (Kaouah et al. 2012)
biochar prepared from Chrysanthemum 53.19 Yang et al. (2021)
morifolium Ramat straw
Hydrochar from pomegranate peels 367.72 The present study

ceptible flop of effectiveness. Thus, we evaluate the prospect of valuing pomegranate peels by developing a hydrochar that is both
economical and eco-friendly that might be used for removing BR46 and treating mixtures that comprise other dyes.

Authors contributions statement


Imane Akkari: Data curation, Methodology, Formal analysis, Writing—Original Draft Preparation.
Zahra Graba and Nacer Bezzi: investigation, software, and visualization.
Lucas Spessato and Mohamed Mehdi Kaci: conceptualization, investigation, writing—reviewing, validation, editing, and super-
vision.

Declaration of competing interest


The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Data availability
The data that has been used is confidential.

Acknowledgments
The authors would like to thank the Algerian Ministry of High Education and the University of Bejaia for the financial support.

References
Akkari, I., Graba, Z., Bezzi, N., Merzeg, F.A., Bait, N., Ferhati, A., 2021. Raw pomegranate peel as promise efficient biosorbent for the removal of Basic Red 46 dye:
equilibrium, kinetic, and thermodynamic studies. Biomass Conversion and Biorefinery 1–14.
Akkari, I., Graba, Z., Bezzi, N., Merzeg, F.A., Bait, N., Ferhati, A., Kaci, M.M., 2022a. Biosorption of Basic Red 46 using raw cactus fruit peels: equilibrium, kinetic and
thermodynamic studies. Biomass Conversion and Biorefinery 1–12.
Akkari, I., Graba, Z., Bezzi, N., Kaci, M.M., Merzeg, F.A., Bait, N., et al., 2022b. Effective removal of cationic dye on activated carbon made from cactus fruit peels: a
combined experimental and theoretical study. Environ. Sci. Pollut. Control Ser. 1–18.
Altintig, E., Onaran, M., Sarı, A., Altundag, H., Tuzen, M., 2018. Preparation, characterization and evaluation of bio-based magnetic activated carbon for effective
adsorption of malachite green from aqueous solution. Mater. Chem. Phys. 220, 313–321.
Altıntıg, E., Altundag, H., Tuzen, M., Sarı, A., 2017. Effective removal of methylene blue from aqueous solutions using magnetic loaded activated carbon as novel
adsorbent. Chem. Eng. Res. Des. 122, 151–163.
Altıntıg, E., Yenigun, M., Sarı, A., Altundag, H., Tuzen, M., Saleh, T.A., 2021. Facile synthesis of zinc oxide nanoparticles loaded activated carbon as an eco-friendly
adsorbent for ultra-removal of malachite green from water. Environ. Technol. Innovat. 21, 101305.
Arief, V.O., Trilestari, K., Sunarso, J., Indraswati, N., Ismadji, S., 2008. Recent progress on biosorption of heavy metals from liquids using low cost biosorbents:
characterization, biosorption parameters and mechanism studies. Clean: Soil, Air, Water 36 (12), 937–962.
Atmani, F., Kaci, M.M., Yeddou-Mezenner, N., Soukeur, A., Akkari, I., Navio, J.A., 2022. Insights into the physicochemical properties of Sugar Scum as a sustainable
biosorbent derived from sugar refinery waste for efficient cationic dye removal. Biomass Conversion and Biorefinery 1–15.
Baláž, M., Ficeriová, J., Briančin, J., 2016. Influence of milling on the adsorption ability of eggshell waste. Chemosphere 146, 458–471.
Bayuo, J., Pelig-Ba, K.B., Abukari, M.A., 2019. Adsorptive removal of chromium (VI) from aqueous solution unto groundnut shell. Appl. Water Sci. 9 (4), 1–11.
Beltrame, K.K., Cazetta, A.L., de Souza, P.S., Spessato, L., Silva, T.L., Almeida, V.C., 2018. Adsorption of caffeine on mesoporous activated carbon fibers prepared from
pineapple plant leaves. Ecotoxicol. Environ. Saf. 147, 64–71.
Boehm, H.P., 2002. Surface oxides on carbon and their analysis: a critical assessment. Carbon 40 (2), 145–149.
Boudechiche, N., Fares, M., Ouyahia, S., Yazid, H., Trari, M., Sadaoui, Z., 2019. Com-parative study on removal of two basic dyes in aqueous medium by adsorption
using activated carbon from Ziziphus lotus stones. Microchem. J. 146, 1010–1018.
Cao, X., Ro, K.S., Chappell, M., Li, Y., Mao, J., 2011. Chemical structures of swine-manure chars produced under different carbonization conditions investigated by
advanced solid-state 13C nuclear magnetic resonance (NMR) spectroscopy. Energy Fuels 25 (1), 388–397.
Chanzu, H.A., Onyari, J.M., Shiundu, P.M., 2012. Biosorption of malachite green from aqueous solutions onto polylactide/spent brewery grains films: kinetic and
equilibrium studies. J. Polym. Environ. 20 (3), 665–672.
Chen, X., Lin, Q., He, R., Zhao, X., Li, G., 2017. Hydrochar production from watermelon peel by hydrothermal carbonization. Bioresour. Technol. 241, 236–243.
Cheung, W.H., Szeto, Y.S., McKay, G., 2007. Intraparticle diffusion processes during acid dye adsorption onto chitosan. Bioresour. Technol. 98 (15), 2897–2904.
Crini, G., 2006. Non-conventional low-cost adsorbents for dye removal: a review. Bioresour. Technol. 97 (9), 1061–1085.
Değermenci, G.D., Değermenci, N., Ayvaoğlu, V., Durmaz, E., Çakır, D., Akan, E., 2019. Adsorption of reactive dyes on lignocellulosic waste; characterization,

14
I. Akkari et al. Sustainable Chemistry and Pharmacy 31 (2023) 100924

equilibrium, kinetic and thermodynamic studies. J. Clean. Prod. 225, 1220–1229.


Deniz, F., Saygideger, S.D., 2011. Removal of a hazardous azo dye (Basic Red 46) from aqueous solution by princess tree leaf. Desalination 268 (1–3), 6–11.
Diaz, E., Manzano, F.J., Villamil, J., Rodriguez, J.J., F Mohedano, A., 2019. Low-cost activated grape seed-derived hydrochar through hydrothermal carbonization and
chemical activation for sulfamethoxazole adsorption. Appl. Sci. 9 (23), 5127.
Dubinin, M., 1960. The potential theory of adsorption of gases and vapors for adsorbents with energetically nonuniform surfaces. Chem. Rev. 60 (2), 235–241.
El-Zahhar, A.A., Awwad, N.S., El-Katori, E.E., 2014. Removal of bromophenol blue dye from industrial waste water by synthesizing polymer-clay composite. J. Mol. Liq.
199, 454–461.
Fayazi, M., Afzali, D., Taher, M.A., Mostafavi, A., Gupta, V.K., 2015. Removal of Safranin dye from aqueous solution using magnetic mesoporous clay: optimization
study. J. Mol. Liq. 212, 675–685.
Foo, K.Y., Hameed, B.H., 2010. Insights into the modeling of adsorption isotherm systems. Chem. Eng. J. 156 (1), 2–10.
Freundlich, H.M.F., 1906. Over the adsorption in solution. J. Phys. Chem. 57 (385471), 1100–1107.
Fu, Y., Viraraghavan, T., 2001. Fungal decolorization of dye wastewaters: a review. Bioresour. Technol. 79 (3), 251–262.
Funke, A., Ziegler, F., 2010. Hydrothermal carbonization of biomass: a summary and discussion of chemical mechanisms for process engineering. Biofuels, Bioproducts
and Biorefining 4 (2), 160–177.
786. Studies in adsorption. Part XIGiles, C.H., MacEwan, T.H., Nakhwa, S.N., Smith, D., 1960. A system of classification of solution adsorption isotherms, and its use in
diagnosis of adsorption mechanisms and in measurement of specific surface areas of solids. J. Chem. Soc. 3973–3993.
Gong, R., Ye, J., Dai, W., Yan, X., Hu, J., Hu, X., Li, S., Huang, H., 2013. Adsorptive removal of methyl orange and methylene blue from aqueous solution with finger-
citron-residue-based activated carbon. Ind. Eng. Chem. Res. 52 (39), 14297–14303.
Güzel, F., Sayğılı, H., Sayğılı, G.A., Koyuncu, F., 2015. New low-cost nanoporous carbonaceous adsorbent developed from carob (Ceratonia siliqua) processing industry
waste for the adsorption of anionic textile dye: characterization, equilibrium and kinetic modeling. J. Mol. Liq. 206, 244–255.
He, J., Cui, A., Deng, S., Chen, J.P., 2018. Treatment of methylene blue containing wastewater by a cost-effective micro-scale biochar/polysulfone mixed matrix hollow
fiber membrane: performance and mechanism studies. J. Colloid Interface Sci. 512, 190–197.
Ho, Y.S., McKay, G., 1999. Pseudo-second order model for sorption processes. Process Biochem. 34 (5), 451–465.
Jais, F.M., Ibrahim, S., Chee, C.Y., Ismail, Z., 2021. High removal of crystal violet dye and tetracycline by hydrochloric acid assisted hydrothermal carbonization of
sugarcane bagasse prepared at high yield. Sustainable Chemistry and Pharmacy 24, 100541.
Kaci, M.M., Nasrallah, N., Atmani, F., Kebir, M., Guernanou, R., Soukeur, A., Trari, M., 2021. Enhanced photocatalytic performance of CuAl2O4 nanoparticles spinel for
dye degradation under visible light. Res. Chem. Intermed. 47 (9), 3785–3806.
Kaci, M.M., Nasrallah, N., Djaballah, A.M., Akkari, I., Belabed, C., Soukeur, Atmani, F., Trari, M., 2022. Insights into the optical and electrochemical features of
CuAl2O4 nanoparticles and it use for methyl violet oxidation under sunlight exposure. Opt. Mater. 126, 112198.
Kim, Y.S., Kim, J.H., 2019. Isotherm, kinetic and thermodynamic studies on the adsorption of paclitaxel onto Sylopute. J. Chem. Therm. 130, 104–113.
Konicki, W., Hełminiak, A., Arabczyk, W., Mijowska, E., 2018. Adsorption of cationic dyes onto Fe@ graphite core–shell magnetic nanocomposite: equilibrium, kinetics
and ther-modynamics. Chem. Eng. Res. Des. 129, 259–270.
Konicki, W., Pełech, I., 2019. Removing cationic dye from aqueous solutions using as-grown and modified multi-walled carbon nanotubes. Pol. J. Environ. Stud. 28 (2),
717–727.
Lagergren, S.K., 1898. About the theory of so-called adsorption of soluble substances. Sven. Vetenskapsakad. Handingarl 24, 1–39.
Langmuir, I., 1918. The adsorption of gases on plane surfaces of glass, mica and platinum. J. Am. Chem. Soc. 40 (9), 1361–1403.
Lawal, I.A., Lawal, M.M., Azeez, M.A., Ndungu, P., 2019. Theoretical and experimental adsorption studies of phenol and crystal violet dye on carbon nanotube
functionalized with deep eutectic solvent. J. Mol. Liq. 288, 110895.
Lei, Y., Su, H., Tian, R., 2016. Morphology evolution, formation mechanism and adsorption properties of hydrochars prepared by hydrothermal carbonization of corn
stalk. RSC Adv. 6 (109), 107829–107835.
Li, Y., Wang, A., Bai, Y., Wang, S., 2017. Evaluation of a mixed anionic–nonionic surfactant modified eggshell membrane as an advantageous adsorbent for the solid-
phase extraction of Sudan I–IV as model analytes. J. Separ. Sci. 40 (12), 2591–2602.
Li, Z., Hanafy, H., Zhang, L., Sellaoui, L., Netto, M.S., Oliveira, M.L., Siliem, M.K., Dotto, G.L., Bonilla-Petriciolet, A., Li, Q., 2020. Adsorption of Congo red and
methylene blue dyes on an ashitaba waste and a walnut shell-based activated carbon from aqueous solutions: experiments, characterization and physical
interpretations. Chem. Eng. J. 388, 124263.
Lima, E.C., Hosseini-Bandegharaei, A., Moreno-Piraján, J.C., Anastopoulos, I., 2019. A critical review of the estimation of the thermodynamic parameters on adsorption
equilibria. Wrong use of equilibrium constant in the Van’t Hoof equation for calculation of thermodynamic parameters of adsorption. J. Mol. Liq. 273, 425–434.
Liu, M., Luo, G., Wang, Y., Xu, R., Wang, Y., He, W., Tan, J., Xing, M., Wu, J., 2017. Nano-silver-decorated microfibrous eggshell membrane: processing, cytotoxicity
assessment and optimization, antibacterial activity and wound healing. Sci. Rep. 7 (1), 1–14.
Ma, H., Kong, A., Ji, Y., He, B., Song, Y., Li, J., 2019. Ultrahigh adsorption capacities for anionic and cationic dyes from wastewater using only chitosan. J. Clean. Prod.
214, 89–94.
Magdy, Y.H., Altaher, H., 2018. Kinetic analysis of the adsorption of dyes from high strength wastewater on cement kiln dust. J. Environ. Chem. Eng. 6 (1), 834–841.
Mahmoodi, N.M., Arami, M., 2009. Numerical finite volume modeling of dye decolorization using immobilized titania nanophotocatalysis. Chem. Eng. J. 146 (2),
189–193.
Mahmoodi, N.M., 2013. Zinc ferrite nanoparticle as a magnetic catalyst: synthesis and dye degradation. Mater. Res. Bull. 48 (10), 4255–4260.
Mekatel, E., Dahdouh, N., Sa-mira, A., Nibou, D., Trari, M., 2021. Removal of maxilon red dye by adsorption and photoca-talysis: optimum conditions, equilibrium, and
kinetic studies. Iran. J. Chem. Chem. Eng. (Int. Engl. Ed.) 40 (1), 93–110.
Miyah, Y., Lahrichi, A., Idrissi, M., Khalil, A., Zerrouq, F., 2018. Adsorption of methylene blue dye from aqueous solutions onto walnut shells powder: equilibrium and
kinetic studies. Surface. Interfac. 11, 74–81.
Muinde, V.M., Onyari, J.M., Wamalwa, B., Wabomba, J.N., 2020. Adsorption of malachite green dye from aqueous solutions using mesoporous chitosan–zinc oxide
composite material. Environmental Chemistry and Ecotoxicology 2, 115–125.
Naisse, C., Alexis, M., Plante, A., Wiedner, K., Glaser, B., Pozzi, A., Carcaillet, C., Criscuoli, I., Rumpel, C., 2013. Can biochar and hydrochar stability be assessed with
chemical methods? Org. Geochem. 60, 40–44.
Necibi, M.C., Amar, I., Draoui, K., Mahjoub, B., 2021. Current situation and future prospects for the production and utilization of sorbing materials for water depollution
in North Africa. In: Sorbents Materials for Controlling Environmental Pollution. Elsevier, pp. 49–71.
Niazi, L., Lashanizadegan, A., Sharififard, H., 2018. Chestnut oak shells activated carbon: preparation, characterization and application for Cr (VI) removal from dilute
aqueous solutions. J. Clean. Prod. 185, 554–556.
Pala, M., Kantarli, I.C., Buyukisik, H.B., Yanik, J., 2014. Hydrothermal carbonization and torrefaction of grape pomace: a comparative evaluation. Bioresour. Technol.
161, 255–262.
Peers, A.M., 1965. Elovich adsorption kinetics and the heterogeneous surface. J. Catal. 4 (4), 499–503.
Poots, V.J.P., McKay, G., Healy, J.J., 1978. Removal of basic dye from effluent using wood as an adsorbent. Journal (Water Pollution Control Federation) 926–935.
Román, S., Nabais, J.M.V., Laginhas, C., Ledesma, B., González, J.F., 2012. Hydrothermal carbonization as an effective way of densifying the energy content of biomass.
Fuel Process. Technol. 103, 78–83.
Román, S., Ledesma, B., Álvarez, A., Herdes, C., 2018. Towards sustainable micro-pollutants’ removal from wastewaters: caffeine solubility, self-diffusion and
adsorption studies from aqueous solutions into hydrochars. Mol. Phys. 116 (15–16), 2129–2141.
Ronix, A., Pezoti, O., Souza, L.S., Souza, I.P., Bedin, K.C., Souza, P.S., et al., 2017. Hydrothermal carbonization of coffee husk: optimization of experimental parameters
and adsorption of methylene blue dye. J. Environ. Chem. Eng. 5 (5), 4841–4849.
Saleh, T.A., 2022. Global trends in technologies and nanomaterials for removal of sulfur organic compounds: clean energy and green environment. J. Mol. Liq. 119340.
Saleh, T.A., Tuzen, M., Sarı, A., 2021. Evaluation of poly (ethylene diamine-trimesoyl chloride)-modified diatomite as efficient adsorbent for removal of rhodamine B
from wastewater samples. Environ. Sci. Pollut. Control Ser. 28 (39), 55655–55666.

15
I. Akkari et al. Sustainable Chemistry and Pharmacy 31 (2023) 100924

Saleh, T.A., 2015. Isotherm, kinetic, and thermodynamic studies on Hg (II) adsorption from aqueous solution by silica-multiwall carbon nanotubes. Environ. Sci. Pollut.
Control Ser. 22 (21), 16721–16731.
Saxena, M., Sharma, N., Saxena, R., 2020. Highly efficient and rapid removal of a toxic dye: adsorption kinetics, isotherm, and mechanism studies on functionalized
multiwalled carbon nanotubes. Surface. Interfac. 21, 100639.
Sayğılı, H., Sayğılı, G.A., 2021. A sustainable generated hydrochar from pomegranate residues for remediation of process water contaminated with Cu (II) ions. Adv.
Powder Technol. 32 (12), 4814–4824.
Şentürk, İ., Yıldız, M.R., 2020. Highly efficient removal from aqueous solution by ad-sorption of Maxilon Red GRL dye using activated pine sawdust. Kor. J. Chem. Eng.
37 (6), 985–999.
Sevilla, M., Fuertes, A.B., 2009. Chemical and structural properties of carbonaceous products obtained by hydrothermal carbonization of saccharides. Chem.--Eur. J. 15
(16), 4195–4203.
Shabaan, O.A., Jahin, H.S., Mohamed, G.G., 2020. Removal of anionic and cationic dyes from wastewater by adsorption using multiwall carbon nanotubes. Arab. J.
Chem. 13 (3), 4797–4810.
Shin, H.S., Kim, J.H., 2016. Isotherm, kinetic and thermodynamic characteristics of adsorption of paclitaxel onto Diaion HP-20. Process Biochem. 51 (7), 917–924.
Shoushtarian, F., Moghaddam, M.R.A., Kowsari, E., 2020. Efficient regeneration/reuse of graphene oxide as a nanoadsor-bent for removing basic Red 46 from aqueous
solutions. J. Mol. Liq. 312, 113386.
Sing, K.S.W., 1982. Reporting physisorption data for gas/solid systems with special reference to the determination of surface area and porosity (Provisional). Pure Appl.
Chem. 54 (11), 2201–2218.
Singh, K., Sinha, T.J.M., Srivastava, S., 2015. Functionalized nanocrystalline cellulose: smart biosorbent for decontamination of arsenic. Int. J. Miner. Process. 139,
51–63.
Spessato, L., Bedin, K.C., Cazetta, A.L., Souza, I.P., Duarte, V.A., Crespo, L.H., et al., 2019. KOH-super activated carbon from biomass waste: insights into the
paracetamol adsorption mechanism and thermal regeneration cycles. J. Hazard Mater. 371, 499–505.
Spessato, L., Duarte, V.A., Viero, P., Zanella, H., Fonseca, J.M., Arroyo, P.A., Almeida, V.C., 2021. Optimization of Sibipiruna activated carbon preparation by simplex-
centroid mixture design for simultaneous adsorption of rhodamine B and metformin. J. Hazard Mater. 411, 125166.
Spessato, L., Duarte, V.A., Fonseca, J.M., Arroyo, P.A., Almeida, V.C., 2022. Nitrogen-doped activated carbons with high performances for CO2 adsorption. J. CO2 Util.
61, 102013.
Syafiuddin, A., Salmiati, S., Jonbi, J., Fulazzaky, M.A., 2018. Application of the kinetic and isotherm models for better understanding of the behaviors of silver
nanoparticles adsorption onto different adsorbents. J. Environ. Manag. 218, 59–70.
Tan, I.A.W., Ahmad, A.L., Hameed, B.H., 2008. Adsorption of basic dye on high-surface-area activated carbon prepared from coconut husk: equilibrium, kinetic and
thermodynamic studies. J. Hazard Mater. 154 (1–3), 337–346.
Temkin, M.J., Pyzhev, V., 1940. Recent Modifications to Langmuir Isotherms.
Tomul, F., Arslan, Y., Kabak, B., Trak, D., Kendüzler, E., Lima, E.C., Tran, H.N., 2020. Peanut shells-derived biochars prepared from different carbonization processes:
comparison of characterization and mechanism of naproxen adsorption in water. Sci. Total Environ. 726, 137828.
Tuzen, M., Sarı, A., Saleh, T.A., 2018. Response surface optimization, kinetic and thermodynamic studies for effective removal of rhodamine B by magnetic AC/CeO2
nanocomposite. J. Environ. Manag. 206, 170–177.
Weber, Jr, W.J., Morris, J.C., 1963. Kinetics of adsorption on carbon from solution. J. Sanit. Eng. Div. 89 (2), 31–59.
Weng, C.H., Pan, Y.F., 2007. Adsorption of a cationic dye (methylene blue) onto spent activated clay. J. Hazard Mater. 144 (1–2), 355–362.
Wu, J., Yang, J., Huang, G., Xu, C., Lin, B., 2020. Hydrothermal carbonization synthesis of cassava slag biochar with excellent adsorption performance for Rhodamine B.
J. Clean. Prod. 251, 119717.
Yang, X., Zhu, W., Song, Y., Zhuang, H., Tang, H., 2021. Removal of cationic dye BR46 by biochar prepared from Chrysanthemum morifolium Ramat straw: a study on
adsorption equilibrium, kinetics and isotherm. J. Mol. Liq. 340, 116617.
Zanella, H.G., Spessato, L., Lopes, G.K., Yokoyama, J.T., Silva, M.C., Souza, P.S., et al., 2021. Caffeine adsorption on activated biochar derived from macrophytes
(Eichornia crassipes). J. Mol. Liq. 340, 117206.
Zhang, Z., Wang, W., Kang, Y., Zong, L., Wang, A., 2016. Tailoring the properties of palygorskite by various organic acids via a one-pot hydrothermal process: a
comparative study for removal of toxic dyes. Appl. Clay Sci. 120, 28–39.
Zhang, S., Sheng, K., Yan, W., Liu, J., Shuang, E., Yang, M., Zhang, X., 2021. Bamboo derived hydrochar microspheres fabricated by acid-assisted hydrothermal
carbonization. Chemosphere 263, 128093.
Zhou, N., Chen, H., Xi, J., Yao, D., Zhou, Z., Tian, Y., Lu, X., 2017. Biochars with excellent Pb (II) adsorption property produced from fresh and dehydrated banana peels
via hydrothermal carbonization. Bioresour. Technol. 232, 204–210.
Zulfajri, M., Kao, Y.T., Huang, G.G., 2021. Retrieve of residual waste of carbon dots derived from straw mushroom as a hydrochar for the removal of organic dyes from
aqueous solutions. Sustainable Chemistry and Pharmacy 22, 100469.

16

You might also like