Molecules 29 02489 v2

Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

molecules

Article
Adsorption of Pb, Cu, and Ni Ions on Activated Carbon
Prepared from Oak Cupules: Kinetics and
Thermodynamics Studies
Dima Khater 1, * , Manal Alkhabbas 2, * and Alaa M. Al-Ma’abreh 2

1 Department of Chemistry, Faculty of Science, Applied Science Private University, Amman 11937, Jordan
2 Department of Chemistry, Faculty of Science, Isra University, Amman 11622, Jordan;
alaa.almaabreh@iu.edu.jo
* Correspondence: d_khater@asu.edu.jo (D.K.); manal.khabbas@iu.edu.jo (M.A.); Tel.: +962-791441467 (D.K.)

Abstract: Agricultural residue-activated carbon and biochar, inexpensive and environmentally


friendly adsorbent materials, have recently received significant research attention. This study in-
vestigated the potential use of oak cupules in activated carbon form to remove widespread heavy
metals (Pb2+ , Cu2+ , and Ni2+ ) from wastewater. The oak-activated carbon was prepared from oak
cupules and activated with phosphoric acid. Oak-activated carbon was characterized using FTIR,
BET analysis, energy-dispersive X-ray spectrometry (EDS), thermogravimetric analysis (TGA), and
differential scanning calorimetry (DSC). The Freundlich, Langmuir, and Temkin isotherm models
were used to assess the equilibrium data. The impact of various parameters, including pH effect,
temperature, adsorbent dose, and contact time, was estimated. The Freundlich model was the most
agreeable with Pb2+ adsorption by oak-based activated carbon, and Langmuir was more compatible
with Cu2+ and Ni2+ . Under optimum conditions, the average maximum removal was 63% Pb2+ , 60%
Cu2+ , and 54% Ni2+ when every ion was alone in the aqueous solution. The removal was enhanced
to 98% Pb2+ , 72% Cu2+ , and 60% Ni2+ when found as a mixture. The thermodynamic model revealed
that the adsorption of ions by oak-based activated carbon is endothermic. The pseudo-second-order
kinetic best describes the adsorption mechanism in this study; it verifies chemical sorption as the
rate-limiting step in adsorption mechanisms. The oak-activated carbon was effective in removing
Citation: Khater, D.; Alkhabbas, M.;
Pb2+ , Cu2+ , and Ni2+ from wastewater and aqueous solutions.
Al-Ma’abreh, A.M. Adsorption of Pb,
Cu, and Ni Ions on Activated Carbon Keywords: adsorption; heavy metal removal; oak-based activated carbon; thermodynamics; kinetics
Prepared from Oak Cupules: Kinetics
and Thermodynamics Studies.
Molecules 2024, 29, 2489. https://
doi.org/10.3390/molecules29112489 1. Introduction
Academic Editor: Eleni Deliyanni In many dry countries, treated wastewater has emerged as a potential remedy to
alleviate seasonal water shortages brought on by drought. Treated wastewater is crucial for
Received: 14 April 2024 irrigation, industrial uses (cooling, processing), toilet flushing, firefighting, etc. [1]. Local
Revised: 9 May 2024
water treatment systems remove organic and heavy metal pollutants using different pro-
Accepted: 17 May 2024
cesses. It is frequently accomplished using challenging and expensive techniques, including
Published: 24 May 2024
ion exchange, membrane filtration [2], and electrochemical deposition [3]. Adsorption and
biosorption are among the most efficient methods for removing pollutants, including heavy
metals, from wastewater due to their affordability [4], simplicity, safety from secondary
Copyright: © 2024 by the authors.
contamination [5], effectiveness, and ease of management [6].
Licensee MDPI, Basel, Switzerland. The most frequently used adsorbent is activated charcoal (AC). However, AC is
This article is an open access article expensive because of the processes involved in its regeneration and reactivation [7]. More
distributed under the terms and research has been focused on biochar and agricultural residue-activated carbon because
conditions of the Creative Commons they are inexpensive, renewable, and widely available. Biochar and agricultural residue-
Attribution (CC BY) license (https:// activated carbon are carbonaceous substances produced from wood, corn straw, rice husk,
creativecommons.org/licenses/by/ pomelo peel, coconut shell, and other agricultural waste [8]. Biochar has textural properties
4.0/). and surface chemical complexity, and it is produced from the thermal decomposition

Molecules 2024, 29, 2489. https://doi.org/10.3390/molecules29112489 https://www.mdpi.com/journal/molecules


Molecules 2024, 29, 2489 2 of 18

of biomass in the absence or presence of a limited amount of oxygen [9]. Agricultural


residue-activated carbon and biochar have been extensively used as adsorbents to remove
water pollutants, such as heavy metals [10]. Heavy metals, including lead (II), copper
(II), and nickel (II), are among the water supply’s most released and dangerous pollutants
and come from industrial processes (e.g., batteries, electroplating, tanneries, textile dyes)
and excessive fertilizer use in agriculture [11,12]. These heavy metals are of high concern
nowadays. They contribute to bioaccumulation, even at low concentrations, and lead to
severe health problems such as cancer and damage to the central nervous system and
kidney [13,14].
Many published studies have used agricultural residue-activated carbon and biochar
as adsorbents to remove organic and inorganic pollutants, including heavy metals, from
water. For instance, Boudrahem et al. [15] developed biochar from coffee residue to remove
Cd(II) and Pb(II) from an aqueous solution. Agricultural residue-activated carbon and
biochar are more favored by many over synthetic adsorbents, even when their efficiency
and selectivity are less than the synthetic ones [16]. They have low production costs, a
large surface area, are environmentally friendly, and pave the way for eliminating tons of
agricultural residues [17]. Recent improvements in the adsorption capacities of activated
carbon have been reported by applying chemical, physical, inorganic, and organic loading
strategies [18].
In this study, we prepared and investigated porous oak-based activated carbon as
a bio-adsorbent for removing lead (II), copper (II), and nickel (II) from wastewater. Our
group previously demonstrated the ability of this adsorbent to remove anionic and cationic
dyes [19]. Oak-based activated carbon was derived from oak cupules, an abundant agricul-
tural waste product in northern Jordan. We assessed the impact of various factors, such as
solution pH, contact time, temperature, adsorbent dose, and initial metal ion concentration,
on the adsorption of metal ions. Our investigation of various adsorption isotherms and
kinetic models provides valuable insights into the adsorption process of metal ions.

2. Results and Discussion


2.1. Characterization of Oak-Activated Carbon
Table 1 lists the prepared oak-activated carbon’s surface area, pore size, and volume
measurements. Based on the IUPAC classification, the oak-activated carbon’s N2 adsorp-
tion/desorption isotherm resembles the type IA isotherm at a relative P/P0 less than 0.4,
and the type IV and H3 hysteresis loops at a relative P/P0 greater than 0.4 (Figure 1a). This
type of sorption isotherm is characteristic of a combined system of microporous (pore diam-
eter < 2 nm) and mesoporous adsorbents (pore diameter = 2–50 nm) [20,21], demonstrating
the presence of a blended structure containing mainly supermicropores and low-range
mesopores. Figure 1b presents a histogram representing the extracted pore size distribution
(PSD), and it shows that the majority of oak-activated carbon pores falling within the
micropore range have a mean pore half width of 0.77 nm (or a pore width of 1.54 nm).

Table 1. Textural properties for oak-based activated carbon obtained from N2 adsorption at 77 K.

Parameter Value
BET surface area (m2 /g) 1088.345
Single point surface area (m2 /g) 1413.518
Pore diameter (nm) 1.54
Total pore volume (cm3 /g) 0.985
Vmicro /Vtotal 0.724
Vmeso /Vtotal 0.196
Vmacro /Vtotal 0.065

Figure 2a,b display the FTIR spectra of the raw oak cupules and oak-activated carbon
(before and after adsorption). A broad band extended from 3200 cm−1 to 3650 cm−1 is
attributed to the hydroxyl group due to lignin and cellulose [22,23]. The hydroxyl group
Molecules 2024, 29, 2489 3 of 18

band waned in oak-based activated carbon due to condensation reactions, carbonization,


adsorbed water molecule release, and cellulose and hemicellulose decomposition in the
raw biomass [24]. The peaks (2924 cm−1 in raw oak cupules, 2983 and 2880 cm−1 in oak
activated carbon) correlate to CH stretching in the CH2 and CH3 groups. The CH stretching
band for the unloaded oak-based activated carbon sample was weak due to the reduced
aliphatic compounds after carbonization [25]. The vibration bands with wavenumber
1723 cm−1 and 1608 cm−1 in the raw oak cupules are related to C=O stretching in uncon-
jugated groups and aromatic C=C stretching vibration in lignin [26]. The
Molecules 2024, 29, x FOR PEER REVIEW
small peak at
3 of 20
1200 cm−1 shows C-O stretching vibration and may refer to C-O phenol vibration. The IR
spectrum reflects various functional groups in raw oak cupules (carbonyls, ester groups,
and ketones) [27].properties
Table 1. Textural An intensive
for oak-basedband in oak-activated
activated carbon
carbon obtained from appeared
N2 adsorption at 1587 cm−1 .
at 77 K.
This peak may refer to the oxygen–aromatic bonding in the aromatic ether [28], the C=O
Parameter Value
carbonyl group stretching BET surfacein area
the (m quinone
2/g) structure, and aromatic 1088.345C=C vibration [29,30].
The extended absorption − 1
Single pointband
surfaceat area
1587(mcm2/g) may be attributed 1413.518
to transforming the organic-
dominated phases into Porebetter-organized
diameter (nm) aromatic structures. The 1.54peak at 1075 cm−1 may
be caused by ionized Total linkage
pore volume (P+(cmO-)g) in phosphoric acid [31].0.985
3/ After the adsorption, a
slight shift from 1587 cmV −micro
1 to/V1578
total cm−1 and from 1180 cm−1 to0.724 1156 cm−1 was observed
for all ions. New peaks appeared Vmeso/Vtotal at 1486 cm−1 and 812 cm−1 after 0.196 ions were loaded on
Vmacro/Vtotal 0.065
oak-activated carbon.

Molecules 2024, 29, x FOR PEER REVIEW 4 of 20

Figure 3a,b shows the scanning electron microscope (SEM) images of oak-activated
carbon after adsorption. The EDS spectrum (Figure 4a) shows that the bulk of oak-
activated carbon before adsorption is mostly carbon (61.41% of its mass), oxygen (21.74%
(a)its mass), and phosphorus (11.90% of its mass), with minor
of (b) amounts of F, Na, Al, Si,
and Ca. After the heavy metal mixture sorption, the characteristic
Figure 1. (a) Adsorption isotherms of nitrogen at 77 K on oak-activated peakscarbons;
of Pb, Cu,
(b) and
poreNisize
Figure 1. (a) in
appeared Adsorption
distribution the EDS
(PSD) isotherms
spectrum,
in the of nitrogen
indicating
oak-activated carbon. at 77metal
that heavy K onionsoak-activated carbons; (b) pore size
had been successfully
distribution
adsorbed(PSD)
(Figure in4b).
the oak-activated carbon.
Figure 2a,b display the FTIR spectra of the raw oak cupules and oak-activated carbon
(before and after adsorption). A broad band extended from 3200 cm−1 to 3650 cm−1 is
attributed to the hydroxyl group due to lignin and cellulose [22,23]. The hydroxyl group
band waned in oak-based activated carbon due to condensation reactions, carbonization,
adsorbed water molecule release, and cellulose and hemicellulose decomposition in the
raw biomass [24]. The peaks (2924 cm−1 in raw oak cupules, 2983 and 2880 cm−1 in oak
activated carbon) correlate to CH stretching in the CH2 and CH3 groups. The CH stretching
band for the unloaded oak-based activated carbon sample was weak due to the reduced
aliphatic compounds after carbonization [25]. The vibration bands with wavenumber 1723
cm-l and 1608 cm−1 in the raw oak cupules are related to C=O stretching in unconjugated
groups and aromatic C=C stretching vibration in lignin [26]. The small peak at 1200 cm−1
shows C-O stretching vibration and may refer to C-O phenol vibration. The IR spectrum
reflects various functional groups in raw oak cupules (carbonyls, ester groups, and
ketones) [27]. An intensive band in oak-activated carbon appeared at 1587 cm−1. This peak
may refer to the oxygen–aromatic bonding in the aromatic ether [28], the C=O carbonyl
group stretching in the quinone structure, and aromatic C=C vibration [29,30]. The
extended absorption band at 1587 cm−1 may be attributed to transforming the organic-
dominated phases into better-organized aromatic structures. The peak at 1075 cm−1 may
(a)be caused by ionized linkage (P+ O-) in phosphoric acid [31]. (b) After the adsorption, a slight
shift 2.
Figure from
FTIR1587 cm−1forto(a)
spectra 1578
rawcm oak−1 and from
cupules and1180 cm−1 to
unloaded 1156 cm−1 was
oak-activated observed
carbon and (b) for
oak-all
Figure 2. FTIR
activated
ions. New spectra
carbon loaded
peaks for
with(a)
appeared raw
Pb(II), oakcm
at Cu(II),
1486 cupules
Ni(II),
−1 and aand
and 812 unloaded
heavy
cm metal
−1 oak-activated
were loadedcarbon
aftermixture.
ions on oak-and (b) oak-
activated carbon
activated loaded with Pb(II), Cu(II), Ni(II), and a heavy metal mixture.
carbon.
Molecules 2024, 29, 2489 4 of 18

Figure 3a,b shows the scanning electron microscope (SEM) images of oak-activated
carbon after adsorption. The EDS spectrum (Figure 4a) shows that the bulk of oak-activated
carbon before adsorption is mostly carbon (61.41% of its mass), oxygen (21.74% of its mass),
(a)and phosphorus (11.90% of its mass), with minor amounts (b) of F, Na, Al, Si, and Ca. After the
heavy metal mixture sorption, the characteristic peaks of Pb, Cu, andcarbon
Figure 2. FTIR spectra for (a) raw oak cupules and unloaded oak-activated Ni appeared
and (b) in the
oak-
EDS spectrum, indicating that heavy metal ions had been successfully
activated carbon loaded with Pb(II), Cu(II), Ni(II), and a heavy metal mixture. adsorbed (Figure 4b).

(a)
Molecules 2024, 29, x FOR PEER REVIEW (b) 5 of 20

Figure
Figure3.3.SEM
SEMimage
imageofofoak-activated
oak-activatedcarbon
carbonloaded
loadedwith
witha aheavy
heavymetal
metalmixture.
mixture.

(a) (b)
Figure 4. EDS Spectrum for oak-activated carbon (a) before adsorption and (b) after adsorption.
Figure 4. EDS Spectrum for oak-activated carbon (a) before adsorption and (b) after adsorption.

Figure
Figure 55displays
displaysthethe thermal
thermal degradation
degradation of raw
of raw oak cupules
oak cupules and oak-activated
and oak-activated carbon
carbon
before andbefore
afterand after adsorption.
adsorption. We observed
We observed that losses
that the mass the mass lossesinoccurred
occurred in three
three stages. The
stages. Theoffirst
first stage stagedegradation
thermal of thermal degradation
was found atwas
50–125 ◦ C at
found 50–125 °C (dehydration).
(dehydration). It has
It has approximate
approximate
percentage valuespercentage valueslosses:
of the mass of the8.95%
mass for
losses:
raw8.95% for raw20.31%
oak cupules, oak cupules, 20.31% for
for oak-activated
oak-activated carbon before adsorption, and 11.59% for oak-activated
carbon before adsorption, and 11.59% for oak-activated carbon loaded with heavy metals.carbon loaded with
heavy
The water content of oak-activated carbon is at least two times higher than that ofthan
metals. The water content of oak-activated carbon is at least two times higher raw
that
oak of raw oak
cupules cupules
because thebecause the developed
developed porosity atporosity at theofsurface
the surface of theleads
the carbon carbonto leads
more
to more agglomerates
agglomerates of water molecules.
of water molecules. The water
The water content content
decreases decreases
after after biochar
biochar adsorbs heavy
adsorbs
metal ions,heavy metalthe
repelling ions, repelling
water the water
molecules molecules
occupied occupied by the
by the oak-activated oak-activated
carbon [32].
carbonThe second stage of thermal degradation in raw oak cupules (around 300 ◦ C) is associated
[32].
with The second stage of hemicellulose,
the decomposition thermal degradation
cellulose,inand
raw oak into
lignin cupules (around
phenolic 300 °C)and
compounds is
organic acids.
associated withThis
thestage leads to a considerable
decomposition loss of mass
of hemicellulose, (47.45%)
cellulose, and[33,34]. However,
lignin into the
phenolic
second thermal
compounds anddegradation
organic acids. of oak-activated carbon
This stage leads to a before and after
considerable adsorption
loss showed
of mass (47.45%)
[33,34]. However, the second thermal degradation of oak-activated carbon before and after
adsorption showed intriguing differences from that of raw oak cupules. The mass loss was
significantly smaller (8.24%) and extended from 140 °C to 500 °C, indicating its high
thermal stability and low volatile matter and cellulose content [35].
The third thermal degradation stage involves weight losses above 400 °C for raw oak
Molecules 2024, 29, 2489 5 of 18

intriguing differences from that of raw oak cupules. The mass loss was significantly smaller
Molecules 2024, 29, x FOR PEER REVIEW 6 of 20
(8.24%) and extended from 140 ◦ C to 500 ◦ C, indicating its high thermal stability and low
volatile matter and cellulose content [35].

Molecules 2024, 29, x FOR PEER REVIEW 6 of 20

Figure
Figure5.5.TGA
TGAprofiles
profilesofofraw
rawoak
oakcupules
cupulesand
andoak-activated
oak-activatedcarbon.
carbon.

The third thermal degradation stage involves weight losses above 400 ◦ C for raw
oak cupules, at which carbonization and graphene sheets grow, and above 500 ◦ C for
oak-activated carbon before and after adsorption. It is linked to the breakdown of lignin
and the devolatilization of residual char [15,36].
Figure 6 displays the measured heat flow (in mW) during the thermal decomposi-
tion of raw oak cupules within a temperature range of 30 to 400 ◦ C at a heating rate of
10 ◦ C/min. Negative heat flow values indicate endothermic processes. The prominent
endothermic peak around 86.3 ◦ C relates to the evaporation of volatile components and
water [37]. Two endothermic peaks were detected at 215.8 and 335.05 ◦ C. These two peaks
may be due to simultaneous lignin condensation, plasticization, and cellulose thermal
Figure 5. TGA profiles of raw oak cupules and oak-activated carbon.
decompositions [38,39].

Figure 6. Differential scanning calorimetry (DSC) scans of raw oak cupules.

2.2. PH Effect
The tendency of heavy metal ions to bind to surfaces was investigated at a range of
pH values (3.0 to 11.0) using the optimum biochar dosage at initial concentrations of 50
mg/L Pb(II) and 30 mg/L Cu(II) and Ni(II). The pH and biochar’s surface charge
distribution influenced the biochar’s heavy metal adsorption. Figure 7a shows the effect
of solution pH on the adsorption capacity, and Figure 7b shows the effect of solution pH
on the removal (%) of heavy metal ions. The absorption was at a minimum at low pH (pH
3) due to electrostatic repulsion between protonated surfaces and positively charged metal
ions (Qe = 17.92 mg/g for Pb(II), 17.11 mg/g for Cu, and 9.07 mg/g for Ni(II)). Sorption
capacity gradually increased with rising pH and reached its maximum for Pb (II) at pH 7.
At the same time, Cu(II) and Ni(II) showed the most significant sorption capacity at pH 5.
Figure 6. Differential
Figure 6. Differentialscanning
scanning calorimetry
calorimetry (DSC)
(DSC) scans
scans of raw
of raw oak cupules.
oak cupules.
Hydronium ions are abundant at low pH and compete with heavy metal ions on active
sorption sites [40]. A gradual increase in pH causes an increase in the amount of negative
2.2. PH Effect
2.2. PH Effect
chargeTheontendency
the surface of the metal
of heavy biochar, which
ions improves
to bind the was
to surfaces biochar’s ability at
investigated to aadsorb until
range of pH
pH The
reaches tendency
7 for of
Pb(II) heavy
and 5 metal
for the ions
other to bind
ions. Theto surfaces
pH effect was
is investigated
consistent
values (3.0 to 11.0) using the optimum biochar dosage at initial concentrations of 50 mg/L with at a
similarrange of
pH values
studies (3.0 toapplying
involving 11.0) using the optimum
another agriculturalbiochar
wastedosage at initial carbon
as an activated concentrations
source of 50
mg/L Pb(II)
[41,42]. The orderandof 30 mg/L
metal Cu(II)
removal wasand
Pb(II)Ni(II). The
> Cu(II) pH and
> Ni(II). biochar’s rise
The additional surface
in pH charge
leads to the precipitation of metal hydroxide. This leads to high
distribution influenced the biochar’s heavy metal adsorption. Figure 7a shows metal uptake at a PH
the effect
range of 8 to pH
of solution 11 [43].
on the adsorption capacity, and Figure 7b shows the effect of solution pH
on the removal (%) of heavy metal ions. The absorption was at a minimum at low pH (pH
Molecules 2024, 29, 2489 6 of 18

Pb(II) and 30 mg/L Cu(II) and Ni(II). The pH and biochar’s surface charge distribution
influenced the biochar’s heavy metal adsorption. Figure 7a shows the effect of solution pH
on the adsorption capacity, and Figure 7b shows the effect of solution pH on the removal
(%) of heavy metal ions. The absorption was at a minimum at low pH (pH 3) due to
electrostatic repulsion between protonated surfaces and positively charged metal ions
(Qe = 17.92 mg/g for Pb(II), 17.11 mg/g for Cu, and 9.07 mg/g for Ni(II)). Sorption capacity
gradually increased with rising pH and reached its maximum for Pb (II) at pH 7. At
the same time, Cu(II) and Ni(II) showed the most significant sorption capacity at pH 5.
Molecules 2024, 29, x FOR PEER REVIEW 7 of 20
Hydronium ions are abundant at low pH and compete with heavy metal ions on active
sorption sites [40]. A gradual increase in pH causes an increase in the amount of negative
charge on the surface of the biochar, which improves the biochar’s ability to adsorb until
pH Our
reaches 7 for Pb(II)
previous studyand 5 for
shows thethe
that other
pHions.
at zeroThe pH charge
point effect is(pH
consistent
pzc) valuewith
for similar
the oak-
studies involving applying another agricultural waste as an activated carbon source
based activated carbon is around 7.0 ± 0.1 [19]. It is consistent with the pH effect study [41,42].in
Thecurrent
the order of metalAbove
study. removalpHwas Pb(II)
pzc, the > Cu(II) >
adsorbent Ni(II). would
surface The additional rise in charged,
be negatively pH leads and
to
the precipitation of metal hydroxide. This leads to high metal
the amounts of heavy metal attached to the adsorbent would increase. uptake at a PH range of 8 to
11 [43].

(a) (b)
Figure
Figure7.7. Effect of pH
Effect of pHon
on(a)
(a)the
the adsorption
adsorption capacity
capacity of heavy
of heavy metalmetal ions(b)
ions and and
the(b) the removal
removal of heavyof
heavy metal ions.
metal ions.

2.3. Effect
Ourofprevious
Adsorbent Doseshows that the pH at zero point charge (pHpzc ) value for the
study
oak-based activated carbon
By equilibrating 0.01 tois0.12
around
g of 7.0 ± 0.1 material
biochar [19]. It is with
consistent with the
50.0 mg/L pHmL
in 50 effect
of astudy
metal
solution, the effect of adsorbent dosage on Pb(II), Cu(II), and Ni(II) removal and
in the current study. Above pH pzc , the adsorbent surface would be negatively charged, was
the amounts (Figure
investigated of heavy8a).
metal
The attached to thedose
adsorbent adsorbent would
is crucial, asincrease.
it determines the system’s
sorbent–sorbate equilibrium. As the amount of adsorbent was increased, the removal
2.3. Effect of Adsorbent Dose
percentage increased until it reached equilibrium at a dosage level of 1.4 g/L for all ions.
By equilibrating
The initial 0.01 percentage
rise in removal to 0.12 g ofcomes
biochar material
from with 50.0biosorbent
the increased mg/L in 50 mL ofarea
surface a
metal solution, the effect of adsorbent dosage on Pb(II), Cu(II), and Ni(II) removal was
and the availability of more adsorption sites [44]. After that, the removal percentage
investigated (Figure 8a). The adsorbent dose is crucial, as it determines the system’s
became almost constant, which may be attributed to the aggregation, causing a reduction
sorbent–sorbate equilibrium. As the amount of adsorbent was increased, the removal
in the overall surface area of the adsorbent [45,46]. Figure 8b shows that increasing oak-
percentage increased until it reached equilibrium at a dosage level of 1.4 g/L for all ions.
based activated carbon dosages decreases the adsorption capacity (Qe) for Pb(II) and
The initial rise in removal percentage comes from the increased biosorbent surface area and
Cu(II).
the availability of more adsorption sites [44]. After that, the removal percentage became
almost constant, which may be attributed to the aggregation, causing a reduction in the
overall surface area of the adsorbent [45,46]. Figure 8b shows that increasing oak-based
activated carbon dosages decreases the adsorption capacity (Qe ) for Pb(II) and Cu(II).
(a) (b)
Figure 7. Effect of pH on (a) the adsorption capacity of heavy metal ions and (b) the removal of
heavy metal ions.
Molecules 2024, 29, 2489 7 of 18
2.3. Effect of Adsorbent Dose

Molecules 2024, 29, x FOR PEER REVIEW 8 of 20


(a) (b)
Figure 8. Effect of oak-based activated carbon dosage on (a) heavy metal removal and (b) heavy
Figure 8. Effect of oak-based activated carbon dosage on (a) heavy metal removal and (b) heavy
(a) metal adsorption capacity. (b)
metal adsorption capacity.
Figure 8. Effect of oak-based 2.4. Contact
activated Timedosage
carbon and Temperature
on (a) heavy metal removal and (b) heavy
metal2.4. Contact
adsorption Time and Temperature
capacity.
The effect of contact time and temperature were estimated within a contact time
2.4. ContactThe
Timeeffect range oftime
of contact
and Temperature 20–160 mintemperature
and at 25, 35, and 45 °C with
were initial concentrations
estimated within a contact of 50 mg/L
timeofrange
Pb (II)
of 20–160 min at and35,
25, 30 mg/L
and of Cu(II)
45 and Ni
◦ C with (II). Their
initial results are presented
concentrations of 50 in Figureof9. Pb
mg/L The(II)
optimum
and
The effect of contact time
time ofand temperature
adsorption waswere
60 minestimated
for Pb(II)within
and 80a contact
min for time
Cu(II) and Ni(II). After that,
range30ofmg/L
20–160ofminCu(II)
at 25, and
35, Ni
and (II).
45 °C Their
with results
initial are presented
concentrations of 50 in Figure
mg/L of Pb 9. The
(II) optimum
saturation was obtained, and the adsorption capacity remained constant. Furthermore,
time of
and adsorption was and
30 mg/L of Cu(II) 60 min forTheir
Ni (II). Pb(II) and are
results 80 min for Cu(II)
presented and9. Ni(II).
in Figure After that, saturation was
The optimum
timeobtained,
of adsorption
andwasthe60adsorption
min for Pb(II) and 80 min
capacity for Cu(II)
remained and Ni(II).
constant. After that, the experimental
Furthermore,
saturation was obtained, and the adsorption capacity remained constant. Furthermore,
results showed that the amount of adsorbed material increased as temperature increased,
the experimental results showed that the amount of adsorbed material increased as
indicating
temperature that the
increased, medium’s
indicating that thetemperature is a significant
medium’s temperature factor
is a significant in adsorption
factor in efficiency,
and this
adsorption process and
efficiency, is endothermic.
this process The temperature
is endothermic. Theeffect is interpreted
temperature effect is by increased collision
and contact
interpreted among
by increased adsorbates
collision and contactand surface
among sites on
adsorbates andthe adsorbent
surface sites on thethat are available for
adsorbent that are [47].
adsorption available for adsorption [47].

Figure 9. Cont.
Molecules 2024, 29, x FOR PEER REVIEW 9 of 20

Molecules 2024,29,
2024,
Molecules 29, 2489
x FOR PEER REVIEW 9 of 20 8 of 18

Figure 9. Effect of contact time and temperature on (a) Pb(II) adsorption capacity, (b) Cu(II)
Figure 9. Effect of contact time and temperature on (a) Pb(II) adsorption capacity, (b) Cu(II)
adsorption capacity,
Figure 9. Effect and (c)
of contact Ni(II)
time and adsorption
temperaturecapacity.
adsorption capacity, and (c) Ni(II) adsorption capacity.
on (a) Pb(II) adsorption capacity, (b) Cu(II) adsorption
capacity, and (c) Ni(II) adsorption capacity.
2.5.2.5. Sorption
Sorption of Mixed
of Mixed Heavyfrom
Heavy Metals Metals from
Aqueous Aqueous
Phase Phase
2.5. Sorption
Figure
of Mixed
10 illustrates
Heavy
the removal
Metals from Aqueous Phase capacity using oak-
Figure 10 illustrates the of heavy metals
removal and adsorption
of heavy metals and adsorption capacity using oak-
activatedFigure 10 illustrates
carbon from an aqueous the removal
solution of heavy
containing mixturesmetals and
of heavy adsorption
metals, includingcapacity using oak-
activated
Pb(II),
carbon
Cu(II), and
from
Ni(II).from
an aqueous
Oak-activated
solution
carbon
containing
showed acontaining
mixtures
powerful affinity
of heavy
for heavy
metals, including
metal metals, including
activated carbon an aqueous solution mixtures of heavy
Pb(II),The
mixtures. Cu(II), andremoval
highest Ni(II).was
Oak-activated carbon
for Pb(II) (98.4%), showed
followed a powerful
by Cu(II) affinity
(72.4%) and Ni for heavy metal
Pb(II), Cu(II), and Ni(II). Oak-activated carbon showed a powerful affinity for heavy metal
(II)mixtures.
(60.0%) at 1.The
4 g/Lhighest removal was
dosage application. for Pb(II)
Removing (98.4%),
heavy metals fromfollowed
the aqueousbyphase
Cu(II) (72.4%) and Ni
may
mixtures.surface
The highest removal was for Pb(II) (98.4%), followed by Cu(II) (72.4%) and Ni (II)
(II)involve
(60.0%) at 1. electrostatic
4 g/L dosage interaction, precipitation,
application. Removing and inner
heavy and outer surface
metals from the aqueous phase
(60.0%) at[48].
complexation 1.4 g/L dosage application. Removing heavy metals from the aqueous phase
may involve surface electrostatic interaction, precipitation, and inner and outer surface
may involve surface electrostatic interaction, precipitation, and inner and outer surface
complexation [48].
complexation [48].

Figure 10. Heavy metal removal and Qe of oak-activated carbon from an aqueous solution
containing a mixture of heavy metals.

2.6. Adsorption Isotherms and Kinetics


Analyzing the adsorption isotherm data and comprehending the adsorption
mechanisms is crucial. Oak-activated carbon was used to study the behavior of heavy
Figure 10. Heavy
Figure 10. Heavymetal
metal removal
removal andand
Qe ofQe of oak-activated
oak-activated carbon carbon
from an from an solution
aqueous
metal adsorption using three isotherm models: Langmuir, Freundlich, and Temkin. The
aqueouscontaining
solution
containing
a mixture
Langmuir a mixture
of heavy
model implies of heavy
metals. metals.
monolayer adsorption over an energetically homogeneous
adsorbent surface. It disregards interactions between ions that have been adsorbed [49].
2.6. Adsorption
2.6. Adsorption
Freundlich’s Isotherms
modelIsotherms and
is based and Kinetics
on Kinetics
the adsorbate forming multiple layers on the
Analyzing the
Analyzing the adsorption
adsorptionisotherm
isothermdata andand
data comprehending
comprehendingthe adsorption mech-
the adsorption
anisms is crucial.
mechanisms Oak-activated
is crucial. carbon
Oak-activated was used
carbon to study
was used the behavior
to study of heavy
the behavior metal
of heavy
metal adsorption using three isotherm models: Langmuir, Freundlich, and Temkin.Lang-
adsorption using three isotherm models: Langmuir, Freundlich, and Temkin. The The
muir modelmodel
Langmuir implies monolayer
implies adsorption
monolayer over an energetically
adsorption homogeneous
over an energetically adsorbent
homogeneous
surface. It disregards interactions between ions that have been adsorbed [49]. Freundlich’s
adsorbent surface. It disregards interactions between ions that have been adsorbed [49].
model is based on the adsorbate forming multiple layers on the heterogeneous solid surface
Freundlich’s model is based on the adsorbate forming multiple layers on the
of the adsorbent, and the binding strength decreases with increasing site usage [50]. The
Temkin model emulates the impact of a few indirect adsorbate/adsorbate interactions. The
heat of adsorption would decrease linearly with coverage due to adsorbate/adsorbate
interactions [51]. The variables and equations for the chosen isotherm models are described
Molecules 2024, 29, 2489 9 of 18

in Section 3.4. Table 2 presents the calculated data based on the Langmuir, Freundlich,
and Temkin isotherms. These models were chosen because they comprehensively analyze
the adsorption process. Each metal ion was measured individually with 10, 20, 30, 50,
and 90 mg/L concentrations at optimum pH. R2 values of isotherms are compared to
demonstrate how well all isotherms fit the experimental data. The metal ion uptakes on
oak-activated carbon experimental data were successfully analyzed. It could be seen that
the Freundlich model’s correlation coefficient (R2 ) values were 0.9461 for Pb(II), 0.9476 for
Cu(II), and 0.9721 for Ni(II) at 25 ◦ C.

Table 2. Fitting parameters for the Langmuir, Freundlich, and Temkin models.

Metal Ion
Parameter
Pb(II) Cu(II) Ni(II)
Langmuir
Qm (mg/g) 7.8 58.5 52.4
KL (L/mg) 0.214 0.054 0.041
R2 0.9182 0.9652 0.9934
RL 0.085 0.38 0.90
Freundlich
n 1.9 1.2 1.5
KF (mg/g) 1.2 1.3 1.6
R2 0.9461 0.9467 0.9712
Temkin
KT (L/mg) 1.08 0.22 0.71
bT (J/mol) 1046 241 299
R2 0.8438 0.7600 0.9603
B 2.37 10.30 8.28

In contrast, Langmuir’s correlation coefficients were 0.9182 for Pb (II), 0.9652, and
0.9934 for Cu(II) and Ni(II), respectively. The Freundlich model was the most appropriate
for the adsorption of Pb(II), which means that this ion showed the most adsorption hetero-
geneity at the bending [52]. Langmuir was the most appropriate for the adsorption of Ni(II),
implying the homogeneous distribution of Ni(II) on the active site of the adsorbent [53].
The Langmuir parameter, the dimensionless separation factor (RL ), is essential in this
study, as it is crucial in predicting the affinity between the adsorbate and the adsorbent. RL
is computed using the following formula [54,55]:

1
RL = (1)
KL C0 +1

where C0 represents the adsorbate’s highest initial heavy metal concentration in the solution
(mg/L). Depending on the value of RL , the isotherm’s shape can be classified as favorable
(0 < RL < 1), irreversible (RL = 0), linear (RL = 1), or unfavorable (RL > 1) [56]. The values of
RL are offered in Table 2. All RL values are between 0.085 and 0.90, which suggests that
heavy metal adsorption onto oak-based activated carbon is highly favorable.
The Freundlich constant (KF ) is related to adsorption capacity, whereas the constant
n deals with the degree of heterogeneity, reflecting the adsorption intensity. n represents
favorable chemical adsorption when its value extends from 1 to 10 [57]. In contrast, if n is
less than 1, the sorption is considered a physical process. The n range in Table 2 confirms
that Pb(II) adsorption is the most favored chemical process.
In the Temkin model, the maximum binding constant (KT ) value was found for Pb(II),
which indicates the most incredible energy for this ion equilibrium adsorption onto the
oak-activated carbon.
Molecules 2024, 29, 2489 10 of 18

The correlation between the experimental data and the Langmuir, Freundlich, and
Temkin models suggests that metal ion uptakes on oak-activated carbon surfaces are
complex and involve multiple simultaneous mechanisms.
Adsorption kinetics are essential to investigating the adsorption mechanism and
the adsorbent’s efficiency. The variables and equations for the selected kinetic studies
are reported in Section 3.4. The rate of the adsorption process and the most likely rate-
controlling step are well-informed by the adsorption kinetics collected at three different
temperatures: 25 ◦ C, 35 ◦ C, and 45 ◦ C. The adsorption kinetics (at optimum pH and
initial concentration) were modeled using pseudo-first-order and pseudo-second-order
models (Table 3). Figure 11 shows that the adsorption process complies more with the
pseudo-second-order kinetic model (R2 = 0.9915–0.9999) than the pseudo-first-order model
(R2 = 0.6273–0.9304). The calculated values of metal ion adsorption capacity from pseudo-
second-order kinetics were 19.7 mg/g for Pb(II), 14.7 mg/g Cu(II), and 11.5 mg/g for Ni(II).
These values are consistent with experimental values (Qe-experimental (Pb) = 22.9 mg/g,
Qe-experimental (Cu) = 14.2 mg/g, and Qe-experimental (Ni) = 11.1 mg/g). The pseudo-
second-order model complies with the chemisorption process. The rate-limiting mechanism
was the chemisorption of all heavy metal; assuming that chemical interactions such as ion
exchange and the chelating reaction caused all heavy metal ions to be adsorbed on the
surface of oak-activated carbon [58].

Table 3. Kinetic parameters for the adsorption of Pd(II), Cu(II), and Ni(II) by oak-activated carbon.

Heavy Metal
Kinetic Model Parameters
Pb(II) Cu(II) Ni(II)
qe , cal (mg/g) 5.0 10.35 8.01
Pseudo-first order k1p (min−1 ) 28 × 10−4 17 × 10−4 41 × 10−4
R2 0.6531 0.6273 0.9314
qe , cal (mg/g) 19.72 14.71 11.53
Pseudo-second order k2p (g/mg min) 1.96 × 10−2 1.41 × 10−2 7.45 × 10−3
R2 0.9999 0.9993 0.9914
Kid (mg/g min1/2 ) 0.4139 0.8228 0.5599
Intraparticle C 22.3 11.98 11.47
R2 0.9288 0.9864 0.964
α (mg/g min) 3.64 × 1013 16.57 18.93
Elovich β (mg/g) 1.90 0.305 0.338
R2 0.9015 0.9798 0.9962

Biochar materials are unique to other materials. Their surfaces can be acidic, basic,
hydrophilic, or hydrophobic and can be moved in and out between [59]. Therefore, biochars
show complicated mechanisms when interacting with pollutants, including heavy metal
ions [60]. This study kept oak-activated carbon at 450 ◦ C for 1 h during pyrolysis. This
situation releases some phenolic substances and prevents a fraction of lignin from being
biodegraded, causing hydroxyl moieties in polyphenols to bind strongly with metal ions
and form a metal–phenolic network (MPN) [61].
The intraparticle diffusion model describes the sorption process and represents a linear
plot of Qt versus t1/2 [62]. R2 and kid are tabulated in Table 3. The predominant controlling
step is intraparticle diffusion if this line passes through the coordinate’s origin [63]. All
graphs in this study do not pass through the origin point, suggesting that the bound-
ary layers are controlled, and a higher intercept is related to a higher boundary layer’s
thickness [64]. Other processes, like adsorption on the external surface (film diffusion),
control the rapid initial uptake rate besides the late intraparticle [65]. The Elovich equation
describes the second-order kinetics, assuming that the actual solid surfaces are energetically
heterogeneous [66]. The value of R2 based on Elovich’s model showed linearity with
Pseudo-second order k2p (g/mg.min) 1.96 × 10−2 1.41 × 10−2 7.45 × 10−3
R 2 0.9999 0.9993 0.9914
Kid (mg/g.min1/2) 0.4139 0.8228 0.5599
Intraparticle C 22.3 11.98 11.47
Molecules 2024, 29, 2489
R2 0.9288 0.9864 0.964 11 of 18
α (mg/g.min) 3.64 × 1013 16.57 18.93
Elovich β (mg/g) 1.90 0.305 0.338
R2 0.9015 0.9798 0.9962
high fitting, suggesting that chemical adsorption may be the mechanism controlling the
adsorption process [67]. The α value was the highest for Pb(II) (3.64 × 1013 mg/g min).

(a) (b)
Figure 11. Kinetics models of (a) pseudo-second-order and (b) pseudo-first-order.
Figure 11. Kinetics models of (a) pseudo-second-order and (b) pseudo-first-order.
Molecules 2024, 29, x FOR PEER REVIEW 13 of 20
2.7.
2.7.Adsorption
AdsorptionThermodynamic
ThermodynamicStudies
Studies
The
The thermodynamic
thermodynamic parameters
parametersfor for the
the adsorption
adsorption of of metal
metal ions
ions on
on oak-activated
oak-activated
carbon
carbonareare estimated
estimated in in Figure
Figure 12.12. Evaluating
Evaluatingthe theintercept
interceptand
andslope
slope of
of the
the plot
plot between
between
Table
lnK 4. Thermodynamic
lnKdd and
and 1/Tyielded
1/T yieldedthe parameters
the values of
values offor
ΔS° ◦ and
∆Sthe and ∆H◦ , respectively
adsorption
ΔH°, of Pd(II), Cu(II),
respectively [68].and
[68]. Ni(II)
Table
Table by oak-activated
44 lists
lists the
the
carbon.
thermodynamic parameter
thermodynamic parameter values.
values. The Thenegative
negativevalues ∆G◦ at
valuesofof ΔG° at most
most temperatures
temperatures
demonstrated the
demonstrated the viability
viability and
and spontaneity
spontaneity of of the
the adsorption
adsorption process,
process, andand spontaneity
spontaneity
increases with anan increase
increase inin temperature
temperature The positive ∆G°
positive value (J/mol)
value of ∆H° ◦
∆H demonstrated
demonstrated
Metal Ion ∆H° (J/mol)increases
∆S° with
(J/mol.K) R2 [69]. The
the endothermic
the endothermic nature nature ofof the
the adsorption
adsorption process, 25 °Cand
process, and higher 35 °C
higher temperatures
temperatures are more
are 45 °C
more
Pb(II) favorable for
31833 favorable for sorption (Figure
105sorption (Figure
0.9998 12). The
12). The observation
observation further proves
543 further proves
−507 that the
that the sorbent’s
sorbent’s
−1557
sorption capability increases as temperature The positive value of ∆S ◦ showed
Cu(II) 39081 sorption capability
132 increases as
0.9956 temperature rises.
−275 The positive value
−1575 of ∆S° showed−2915
increased randomness
increased randomness due due to to increasing
increasing the the percentage
percentage of of active
active sites
sites at
at the
the solution
solution
Ni(II) 77804 interface as257 the adsorption
0.9965
adsorption progressed
progressed [70].[70]. The
1170
The increased
increased randomness
−1390may be attributed
randomness may
−3961
interface as the be attributed
towater
to watermolecule
moleculerelease
release at
at the
the interface.
interface.

12.Thermodynamic
Figure 12.
Figure Thermodynamicstudy for for
study Pb(II), Ni(II),
Pb(II), and Cu(II)
Ni(II), onto oak-activated
and Cu(II) carbon. carbon.
onto oak-activated

2.8. Comparison
The performance of oak-activated carbon was compared with that of some synthetic
and natural adsorbents derived from agricultural residues. Table 5 shows the results of
this comparison. It can be concluded that synthetic adsorbents are more effective than
Molecules 2024, 29, 2489 12 of 18

Table 4. Thermodynamic parameters for the adsorption of Pd(II), Cu(II), and Ni(II) by oak-
activated carbon.

∆G◦ (J/mol)
Metal Ion ∆H◦ (J/mol) ∆S◦ (J/mol K) R2
25 ◦ C 35 ◦ C 45 ◦ C
Pb(II) 31,833 105 0.9998 543 −507 −1557
Cu(II) 39,081 132 0.9956 −275 −1575 −2915
Ni(II) 77,804 257 0.9965 1170 −1390 −3961

2.8. Comparison
The performance of oak-activated carbon was compared with that of some synthetic
and natural adsorbents derived from agricultural residues. Table 5 shows the results of this
comparison. It can be concluded that synthetic adsorbents are more effective than those
prepared as agricultural waste-based biochars. Their capacities as adsorbents are higher
with faster equilibrium times.

Table 5. Comparison of the adsorption capacity of some synthetic and natural adsorbents for Pb(II),
Cu(II), and Ni(II).

Adsorption Capacities (mg/g) Equilibrium Time Adsorbent


Adsorbents Materials Reference
Pb(II) Cu(II) Ni(II) (min)
Poly-chloromethyl Synthetic
207 167 55 60 [71]
styrene chelating resin polymer
Titanium oxide Synthetic metal
686 835 757 28 [72]
(TiO2 ) nanofibers oxide
Worn
- 98 93 40 Synthetic waste [73]
tire-activated carbon
Hydrous Virgin hydrous
- 24 30 10 [74]
manganese oxide manganese oxide
Thiol-functionalized
Modified hydrous
- 31 25 10 hydrous [74]
manganese oxide
manganese oxide
Sugarcane-based
19 13 3 - Agricultural waste [42]
activated carbon
Corn stalk biochar 41 - - 240 Agricultural waste [40]
Ragweed and horse
124–359 - - 120 Invasive plant species [75]
weed biochars
Peanut
56.5 - - 180 min Agricultural waste [76]
shell-based biochar

3. Materials and Methods


3.1. Sample Collection and Preparation of Oak-Based Activated Carbon
The oak cupules were gathered from the Jerash province in Jordan, cleaned of leftover
impurities with distilled water, and dried in an air oven at 100 ◦ C for a full day. Then, they
were ground and sieved to obtain particle sizes less than 0.5 mm. Oak cupules were mixed
with H3 PO4 (85%) at a 3:1 ratio (g H3 PO4 /g oak cupules). After being left overnight, the
generated slurry was dried at 120 ◦ C for 4 h. The mixture was then cooked for one hour at
450 ◦ C in a muffle furnace. After washing with a 1.0 M NaOH solution to pH 7, distilled
water was used to rinse the activated carbon. The prepared oak-based activated carbon
was dried at 110 ◦ C for 6 h.
Molecules 2024, 29, 2489 13 of 18

3.2. Characterization of Oak-Based Activated Carbon


An Autosorb IQ surface analyzer (Quantachrome, Boynton Beach, FL, USA) was used
to investigate textural properties. Nitrogen adsorption and desorption were performed at
77 K. Samples were outgassed at 573 K for 3 h under a 10−3 Pa vacuum. The Brunauer–
Emmett–Teller equation was used to calculate the BET surface area, whereas the total pore
volume Vtotal was estimated at a P/P0 relative pressure of 0.99. The nonlocal density func-
tional theory was used to evaluate the pore size distribution (PSD) in the form of slits [77].
Infrared spectra were collected using a Brucker (Billerica, MA, USA) attenuated total re-
flectance Fourier transform infrared (ATR-FT-IR) spectrophotometer in the 4000–480 cm−1
wave number range. The spectra of each sample were recorded at room temperature. An
energy-dispersive X-ray spectrometer (EDS) interfaced with the Apreo 2 S LoVac SEM
(Thermo Scientific, Waltham, MA, USA) was used to determine the elemental composition
of oak-activated carbon before and after adsorption. Thermogravimetric analysis (TGA)
was performed using a NETZSCH TG 209F1 thermogravimetric analyzer (Iris, Munich,
Germany). A sample of raw oak cupules (7.9 mg), oak-activated carbon before adsorption
(11.4 mg), and oak-activated carbon was used to adsorb a mixture of heavy metals (3.1 mg)
that had been heated to 110 ◦ C to evolve physically sorbed water. These samples were
heated under a nitrogen atmosphere from 25 ◦ C to 800 ◦ C at a heating rate of 10 ◦ C/min. A
Shimadzu DSC-60 differential scanning calorimeter (Tokyo, Japan) was used to examine
the thermal properties of raw oak cupules. A sealed, empty aluminum pan served as a
reference. The sample (3.2 mg) was heated to 110 ◦ C to evolve physically sorbed water.
The sample, enclosed in a covered aluminum pan, was heated at 10 ◦ C/min from 30 ◦ C to
400 ◦ C in a nitrogen gas environment.

3.3. Adsorption Experiments (Optimum Parameters, Metal Removal Efficiency, and


Adsorption Capacities)
Nitrate salts of Cu(NO3 )2 , Pb(NO3 )2 , and Ni(NO3 )2 , all of purity greater than 99%,
were dissolved in distilled water to create stock solutions of 1000 mg/L of metal ions that
were used in the experiments. Dilutions from the stock solution were applied to have
additional concentrations. A calibrated pH meter was used to track the pH changes made
to the solutions.
The test solutions were prepared by diluting the stock standard solution with 1000 mg/L
Pb(II), Cu(II), and Ni(II) and adding a weighed amount of oak-based activated carbon.
The flask’s contents were shaken in an electrically controlled reciprocating shaker for a
specified period. After that, they were filtered through a 0.45 µm Millipore filter. Batch-
mode adsorption studies of Pb(II), Cu(II), and Ni(II) with oak-activated carbon were
conducted in 250 mL stoppered flasks containing 50 mL of the test solution to determine
the effect of various parameters, including pH, contact time, adsorbent dosage level, and
temperature effect.
Dosage effect was performed at 25 ± 1 ◦ C, agitation of 200 rpm, a contact time of
80 min, and at optimum pH (7 for Pb(II) and 5 for Ni(II) or Cu(II)). The dosage was varied
between 0.01 and 0.12 g of biochar material with 50.0 g/L in 50 mL of a metal solution. For
the effect of contact time and temperature, the experiments were conducted at 25, 35, and
45 ◦ C, with initial concentrations of 50 mg/L of Pb(II), and 30 mg/L of Cu(II) and Ni(II),
agitation of 200 rpm, at optimum pH and optimum dosage.
The heavy metals concentrations were measured using ContraAA800 flame atomic
absorption spectrometry (Analytik, Jena, Germany). The following equation determined
the percentage of heavy metal ions removed:

(C0 −Ce )
Removal% = ∗ 100% (2)
Ce

where C0 is the initial metal ion concentration (mg/L) and Ce is the equilibrium metal ion
concentration (mg/L).
Molecules 2024, 29, 2489 14 of 18

Equilibrium sorption capacity (Qe ), the amount of metal adsorbed per gram of biosor-
bent, can be calculated in mg/g using the following equation:

( C0 − Ce )V
Qe = (3)
m
where V is the volume of the aqueous solution (L) and m is the adsorbent weight (in g) used.
Equation (4) determines the metal ion adsorption capacity Qt (mg/g) at time t:

( C 0 − Ct ) V
Qt = (4)
m
where Ct (mg/L) is the metal ion concentration at time t.

3.4. Equilibrium Isotherm and Kinetics Studies


Table 6 lists the variables and equations for the chosen isotherm and kinetic studies
models used in this study.

Table 6. Equations and parameters of equilibrium isotherms and kinetic models used in this study.

Model Name Nonlinear Equation Linear Equation Parameters Ref.


Qm : Maximum sorption
KL Ce 1 capacity (mg/g).
Langmuir Qe = 1+KL Ce Qm Qe = KL Q1m Ce + 1
Qm
[49,78]
KL : Langmuir
constant (L/mg).
1
Freundlich 1 1/n, KF : Freundlich constants [79]
Qe = kf Cen ln Qe = ln kf + n ln Ce
KT : Temkin constant related to
equilibrium binding
Qe =B ln (KT Ce ) energy (L/g).
Temkin Qe =BlnKT +BlnCe [80]
B= RT/bT bT : Temkin constant related to
the sorption heat (J/mol)
B: Temkin constant
k1p : Pseudo-first-order rate
Pseudo-first order Qt = Qe (1 − ek1p t ) ln(Qe − Qt )= lnqQe − k1p t constant (min−1 ). [81]
t: Contact time (min).
k2p : Rate constant of
Pseudo-second order k2p Q2e t t t 1 pseudo-second-order [32]
Qt = Qt = Qe +
1+k2p Qe k2p Q2e
adsorption (g/mg min )
kint : Intraparticle diffusion rate
1
Intraparticle diffusion Qt = kint t 2 + C constant. [82]
C: Intercept
α: Initial adsorption rate
(mg/g min).
β: Constant related to the
Elovich ln αβ ln t extent of surface coverage and [83]
Qt = β + β
activation energy for
chemisorptions
processes (mg/g).

4. Conclusions
Simple and inexpensive decontamination processes were applied to remove three
widespread and non-biodegradable heavy metals from wastewater, namely, Pb(II), Cu(II),
and Ni(II), using oak-activated carbon. Oak-activated carbon proved a promising and easily
accessible adsorbent for removing heavy metals from wastewater with low concentrations
of Pb(II), Cu(II), and Ni(II), offering a low-cost and renewable method. The effectiveness of
oak-based activated carbon in the simultaneous co-adsorption of Pb(II), Cu(II), and Ni(II)
Molecules 2024, 29, 2489 15 of 18

was also investigated. Contact times of 60 min for Pb(II) and 80 min for Cu(II) and Ni(II)
were enough to achieve the adsorption equilibrium condition. The efficacy of heavy metal
removal was the best at oak-active carbon dosage levels of 1.4 g/L. The removal efficiency
was the best for Pb(II). A pseudo-second-order model was well adapted to the adsorption
kinetics of oak-activated carbon, suggesting heavy metal ions have chemisorption and
chelation-directed mechanisms on oak-activated carbon surfaces. Using the pseudo-second-
order kinetics model, the calculated Qe agreed well with the experimental Qe .

Author Contributions: Conceptualization, D.K. and M.A.; methodology, D.K., M.A. and A.M.A.-M.;
analysis, D.K., M.A. and A.M.A.-M.; investigation, D.K.; resources, M.A.; data curation, A.M.A.-M.;
writing—original draft preparation, D.K.; writing—review and editing, D.K. and M.K; project admin-
istration, D.K. and M.A. All authors have read and agreed to the published version of the manuscript.
Funding: This research received no external funding.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Data are contained within this article.
Conflicts of Interest: The authors declare no conflicts of interest.

References
1. Bouwer, H. Integrated water management: Emerging issues and challenges. Agric. Water Manag. 2000, 45, 217–228. [CrossRef]
2. Saravanan, A.; Kumar, P.S.; Jeevanantham, S.; Karishma, S.; Tajsabreen, B.; Yaashikaa, P.; Reshma, B. Effective water/wastewater
treatment methodologies for toxic pollutants removal: Processes and applications towards sustainable development. Chemosphere
2021, 280, 130595. [CrossRef] [PubMed]
3. Barakat, M.A. New trends in removing heavy metals from industrial wastewater. Arab. J. Chem. 2011, 4, 361–377. [CrossRef]
4. Bhatnagar, A.; Sillanpää, M. Utilization of agro-industrial and municipal waste materials as potential adsorbents for water
treatment—A review. Chem. Eng. J. 2010, 157, 277–296. [CrossRef]
5. Zhu, Y.; Fan, W.; Zhou, T.; Li, X. Removal of chelated heavy metals from aqueous solution: A review of current methods and
mechanisms. Sci. Total Environ. 2019, 678, 253–266. [CrossRef]
6. Fu, F.; Wang, Q. Removal of heavy metal ions from wastewaters: A review. J. Environ. Manag. 2011, 92, 407–418. [CrossRef]
[PubMed]
7. Prahas, D.; Kartika, Y.; Indraswati, N.; Ismadji, S. Activated carbon from jackfruit peel waste by H3PO4 chemical activation: Pore
structure and surface chemistry characterization. Chem. Eng. J. 2008, 140, 32–42. [CrossRef]
8. Ouyang, J.; Zhou, L.; Liu, Z.; Heng, J.Y.Y.; Chen, W. Biomass-derived activated carbons for the removal of pharmaceutical
mircopollutants from wastewater: A review. Sep. Purif. Technol. 2020, 253, 117536. [CrossRef]
9. Grutzmacher, P.; Puga, A.P.; Bibar, M.P.S.; Coscione, A.R.; Packer, A.P.; de Andrade, C.A. Carbon stability and mitigation of
fertilizer induced N2O emissions in soil amended with biochar. Sci. Total Environ. 2018, 625, 1459–1466. [CrossRef]
10. Suhas; Carrott, P.J.M.; Ribeiro Carrott, M.M.L. Lignin—From natural adsorbent to activated carbon: A review. Bioresour. Technol.
2007, 98, 2301–2312. [CrossRef]
11. Esposito, A.; Pagnanelli, F.; Lodi, A.; Solisio, C.; Veglio, F. Biosorption of heavy metals by Sphaerotilus natans: An equilibrium
study at different pH and biomass concentrations. Hydrometallurgy 2001, 60, 129–141. [CrossRef]
12. Wang, J.; Chen, C. Biosorption of heavy metals by Saccharomyces cerevisiae: A review. Biotechnol. Adv. 2006, 24, 427–451.
[CrossRef]
13. Vijayaraghavan, K.; Rangabhashiyam, S.; Ashokkumar, T.; Arockiaraj, J. Mono-and multi-component biosorption of lead (II),
cadmium (II), copper (II) and nickel (II) ions onto coco-peat biomass. Sep. Sci. Technol. 2016, 51, 2725–2733. [CrossRef]
14. Lee, J.-C.; Son, Y.-O.; Pratheeshkumar, P.; Shi, X. Oxidative stress and metal carcinogenesis. Free Radic. Biol. Med. 2012, 53, 742–757.
[CrossRef] [PubMed]
15. Boudrahem, F.; Soualah, A.; Aissani-Benissad, F. Pb (II) and Cd (II) removal from aqueous solutions using activated carbon
developed from coffee residue activated with phosphoric acid and zinc chloride. J. Chem. Eng. Data 2011, 56, 1946–1955.
[CrossRef]
16. Liu, C.; Zhang, H.-X. Modified-biochar adsorbents (MBAs) for heavy-metal ions adsorption: A critical review. J. Environ. Chem.
Eng. 2022, 10, 107393. [CrossRef]
17. Srivatsav, P.; Bhargav, B.S.; Shanmugasundaram, V.; Arun, J.; Gopinath, K.P.; Bhatnagar, A. Biochar as an eco-friendly and
economical adsorbent for the removal of colorants (dyes) from aqueous environment: A review. Water 2020, 12, 3561. [CrossRef]
18. Mariana, M.; HPS, A.K.; Mistar, E.; Yahya, E.B.; Alfatah, T.; Danish, M.; Amayreh, M. Recent advances in activated carbon
modification techniques for enhanced heavy metal adsorption. J. Water Process Eng. 2021, 43, 102221. [CrossRef]
Molecules 2024, 29, 2489 16 of 18

19. Alkhabbas, M.; Al-Ma’abreh, A.M.; Edris, G.; Saleh, T.; Alhmood, H. Adsorption of anionic and cationic dyes on activated carbon
prepared from oak cupules: Kinetics and thermodynamics studies. Int. J. Environ. Res. Public Health 2023, 20, 3280. [CrossRef]
20. Thommes, M.; Kaneko, K.; Neimark, A.V.; Olivier, J.P.; Rodriguez-Reinoso, F.; Rouquerol, J.; Sing, K.S. Physisorption of gases,
with special reference to the evaluation of surface area and pore size distribution (IUPAC Technical Report). Pure Appl. Chem.
2015, 87, 1051–1069. [CrossRef]
21. Storck, S.; Bretinger, H.; Maier, W.F. Characterization of micro-and mesoporous solids by physisorption methods and pore-size
analysis. Appl. Catal. A Gen. 1998, 174, 137–146. [CrossRef]
22. Hadjiivanov, K. Identification and characterization of surface hydroxyl groups by infrared spectroscopy. In Advances in Catalysis;
Elsevier: Amsterdam, The Netherlands, 2014; Volume 57, pp. 99–318.
23. De Celis, J.; Amadeo, N.; Cukierman, A. In situ modification of activated carbons developed from a native invasive wood on
removal of trace toxic metals from wastewater. J. Hazard. Mater. 2009, 161, 217–223. [CrossRef] [PubMed]
24. Köseoğlu, E.; Akmil-Başar, C. Preparation, structural evaluation and adsorptive properties of activated carbon from agricultural
waste biomass. Adv. Powder Technol. 2015, 26, 811–818. [CrossRef]
25. Asadullah, M.; Asaduzzaman, M.; Kabir, M.S.; Mostofa, M.G.; Miyazawa, T. Chemical and structural evaluation of activated
carbon prepared from jute sticks for Brilliant Green dye removal from aqueous solution. J. Hazard. Mater. 2010, 174, 437–443.
[CrossRef] [PubMed]
26. Moosavinejad, S.M.; Madhoushi, M.; Vakili, M.; Rasouli, D. Evaluation of degradation in chemical compounds of wood in
historical buildings using FT-IR and FT-Raman vibrational spectroscopy. Maderas Cienc. Tecnol. 2019, 21, 381–392. [CrossRef]
27. Yi, W.; Nadeem, F.; Xu, G.; Zhang, Q.; Joshee, N.; Tahir, N. Modifying crystallinity, and thermo-optical characteristics of Paulownia
biomass through ultrafine grinding and evaluation of biohydrogen production potential. J. Clean. Prod. 2020, 269, 122386.
[CrossRef]
28. Nasir, S.; Hussein, M.Z.; Zainal, Z.; Yusof, N.A.; Mohd Zobir, S.A. Electrochemical energy storage potentials of waste biomass: Oil
palm leaf-and palm kernel shell-derived activated carbons. Energies 2018, 11, 3410. [CrossRef]
29. Fuente, E.; Menéndez, J.; Díez, M.; Suárez, D.; Montes-Morán, M. Infrared spectroscopy of carbon materials: A quantum chemical
study of model compounds. J. Phys. Chem. B 2003, 107, 6350–6359. [CrossRef]
30. Sevilla, M.; Maciá-Agulló, J.A.; Fuertes, A.B. Hydrothermal carbonization of biomass as a route for the sequestration of CO2 :
Chemical and structural properties of the carbonized products. Biomass Bioenergy 2011, 35, 3152–3159. [CrossRef]
31. Liu, H.; Zhang, J.; Bao, N.; Cheng, C.; Ren, L.; Zhang, C. Textural properties and surface chemistry of lotus stalk-derived activated
carbons prepared using different phosphorus oxyacids: Adsorption of trimethoprim. J. Hazard. Mater. 2012, 235, 367–375.
[CrossRef]
32. Ammari, T.G.; Al-Hadidi, M.; Al-Kharabsheh, N.; Khater, D.; Abu-Romman, S. Dry Biosolids Reuse as Costless Biodegradable
Adsorbent for Cadmium Removal from Water Systems. J. Ecol. Eng. 2021, 22, 1–12. [CrossRef]
33. Carrier, M.; Loppinet-Serani, A.; Denux, D.; Lasnier, J.-M.; Ham-Pichavant, F.; Cansell, F.; Aymonier, C. Thermogravimetric
analysis as a new method to determine the lignocellulosic composition of biomass. Biomass Bioenergy 2011, 35, 298–307. [CrossRef]
34. Yazdani, M.R.; Duimovich, N.; Tiraferri, A.; Laurell, P.; Borghei, M.; Zimmerman, J.B.; Vahala, R. Tailored mesoporous biochar
sorbents from pinecone biomass for the adsorption of natural organic matter from lake water. J. Mol. Liq. 2019, 291, 111248.
[CrossRef]
35. Demirbas, A. Combustion characteristics of different biomass fuels. Prog. Energy Combust. Sci. 2004, 30, 219–230. [CrossRef]
36. Duman, G.; Onal, Y.; Okutucu, C.; Onenc, S.; Yanik, J. Production of activated carbon from pine cone and evaluation of its
physical, chemical, and adsorption properties. Energy Fuels 2009, 23, 2197–2204. [CrossRef]
37. Naderi, B.; Keramat, J.; Nasirpour, A.; Aminifar, M. Complex coacervation between oak protein isolate and gum Arabic:
Optimization & functional characterization. Int. J. Food Prop. 2020, 23, 1854–1873.
38. Miranda, M.; Bica, C.; Nachtigall, S.; Rehman, N.; Rosa, S. Kinetical thermal degradation study of maize straw and soybean hull
celluloses by simultaneous DSC–TGA and MDSC techniques. Thermochim. Acta 2013, 565, 65–71. [CrossRef]
39. Mehrotra, R.; Singh, P.; Kandpal, H. Near infrared spectroscopic investigation of the thermal degradation of wood. Thermochim.
Acta 2010, 507, 60–65. [CrossRef]
40. Yang, W.; Lu, C.; Liang, B.; Yin, C.; Lei, G.; Wang, B.; Zhou, X.; Zhen, J.; Quan, S.; Jing, Y. Removal of Pb (II) from Aqueous
Solution and Adsorption Kinetics of Corn Stalk Biochar. Separations 2023, 10, 438. [CrossRef]
41. Abbaszadeh, S.; Alwi, S.R.W.; Webb, C.; Ghasemi, N.; Muhamad, I.I. Treatment of lead-contaminated water using activated
carbon adsorbent from locally available papaya peel biowaste. J. Clean. Prod. 2016, 118, 210–222. [CrossRef]
42. Van Tran, T.; Bui, Q.T.P.; Nguyen, T.D.; Le, N.T.H.; Bach, L.G. A comparative study on the removal efficiency of metal ions (Cu2+ ,
Ni2+ , and Pb2+ ) using sugarcane bagasse-derived ZnCl2 -activated carbon by the response surface methodology. Adsorpt. Sci.
Technol. 2017, 35, 72–85. [CrossRef]
43. Gunatilake, S. Methods of removing heavy metals from industrial wastewater. Methods 2015, 1, 14.
44. Abraham, T.E. Biosorption of Cr (VI) from aqueous solution by Rhizopus nigricans. Bioresour. Technol. 2001, 79, 73–81.
45. Semerjian, L. Equilibrium and kinetics of cadmium adsorption from aqueous solutions using untreated Pinus halepensis sawdust.
J. Hazard. Mater. 2010, 173, 236–242. [CrossRef] [PubMed]
46. Huang, Y.; Li, S.; Chen, J.; Zhang, X.; Chen, Y. Adsorption of Pb (II) on mesoporous activated carbons fabricated from water
hyacinth using H3 PO4 activation: Adsorption capacity, kinetic and isotherm studies. Appl. Surf. Sci. 2014, 293, 160–168. [CrossRef]
Molecules 2024, 29, 2489 17 of 18

47. Kılıç, M.; Kırbıyık, Ç.; Çepelioğullar, Ö.; Pütün, A.E. Adsorption of heavy metal ions from aqueous solutions by bio-char, a
by-product of pyrolysis. Appl. Surf. Sci. 2013, 283, 856–862. [CrossRef]
48. Yang, X.; Wan, Y.; Zheng, Y.; He, F.; Yu, Z.; Huang, J.; Wang, H.; Ok, Y.S.; Jiang, Y.; Gao, B. Surface functional groups of
carbon-based adsorbents and their roles in the removal of heavy metals from aqueous solutions: A critical review. Chem. Eng. J.
2019, 366, 608–621. [CrossRef]
49. Kumar, K.V.; Sivanesan, S. Isotherms for Malachite Green onto rubber wood (Hevea brasiliensis) sawdust: Comparison of linear
and non-linear methods. Dye. Pigment. 2007, 72, 124–129. [CrossRef]
50. Khambhaty, Y.; Mody, K.; Basha, S.; Jha, B. Kinetics, equilibrium and thermodynamic studies on biosorption of hexavalent
chromium by dead fungal biomass of marine Aspergillus niger. Chem. Eng. J. 2009, 145, 489–495. [CrossRef]
51. Behnamfard, A.; Salarirad, M.M. Equilibrium and kinetic studies on free cyanide adsorption from aqueous solution by activated
carbon. J. Hazard. Mater. 2009, 170, 127–133. [CrossRef]
52. Srivastava, V.C.; Mall, I.D.; Mishra, I.M. Competitive adsorption of cadmium (II) and nickel (II) metal ions from aqueous solution
onto rice husk ash. Chem. Eng. Process. Process Intensif. 2009, 48, 370–379. [CrossRef]
53. Huang, L.; Sun, Y.; Yang, T.; Li, L. Adsorption behavior of Ni (II) on lotus stalks derived active carbon by phosphoric acid
activation. Desalination 2011, 268, 12–19. [CrossRef]
54. Hall, K.R.; Eagleton, L.C.; Acrivos, A.; Vermeulen, T. Pore-and solid-diffusion kinetics in fixed-bed adsorption under constant-
pattern conditions. Ind. Eng. Chem. Fundam. 1966, 5, 212–223. [CrossRef]
55. Chowdhury, S.; Saha, P. Sea shell powder as a new adsorbent to remove Basic Green 4 (Malachite Green) from aqueous solutions:
Equilibrium, kinetic and thermodynamic studies. Chem. Eng. J. 2010, 164, 168–177. [CrossRef]
56. Amin, N.K. Removal of reactive dye from aqueous solutions by adsorption onto activated carbons prepared from sugarcane
bagasse pith. Desalination 2008, 223, 152–161. [CrossRef]
57. Al-Wahbi, A.A.M.; Dammag, H.A.Q. Removal of methylene blue from aqueous solutions using Yemen bentonite. Diyala J. Eng.
Sci. 2011, 4, 30–53.
58. Arshadi, M.; Amiri, M.J.; Mousavi, S. Kinetic, equilibrium and thermodynamic investigations of Ni (II), Cd (II), Cu (II) and Co (II)
adsorption on barley straw ash. Water Resour. Ind. 2014, 6, 1–17. [CrossRef]
59. Liu, W.-J.; Jiang, H.; Yu, H.-Q. Development of biochar-based functional materials: Toward a sustainable platform carbon material.
Chem. Rev. 2015, 115, 12251–12285. [CrossRef] [PubMed]
60. Qin, C.; Wang, H.; Yuan, X.; Xiong, T.; Zhang, J.; Zhang, J. Understanding structure-performance correlation of biochar materials
in environmental remediation and electrochemical devices. Chem. Eng. J. 2020, 382, 122977. [CrossRef]
61. Yang, X.; Li, Y.; Wu, D.; Yan, L.; Guan, J.; Wen, Y.; Bai, Y.; Mamba, B.B.; Darling, S.B.; Shao, L. Chelation-directed interface
engineering of in-place self-cleaning membranes. Proc. Natl. Acad. Sci. USA 2024, 121, e2319390121. [CrossRef]
62. Qiu, H.; Lv, L.; Pan, B.-C.; Zhang, Q.-J.; Zhang, W.-M.; Zhang, Q.-X. Critical review in adsorption kinetic models. J. Zhejiang Univ.
Sci. A 2009, 10, 716–724. [CrossRef]
63. Gong, R.; Zhu, S.; Zhang, D.; Chen, J.; Ni, S.; Guan, R. Adsorption behavior of cationic dyes on citric acid esterifying wheat straw:
Kinetic and thermodynamic profile. Desalination 2008, 230, 220–228. [CrossRef]
64. Valderrama, C.; Gamisans, X.; De las Heras, X.; Farran, A.; Cortina, J. Sorption kinetics of polycyclic aromatic hydrocarbons
removal using granular activated carbon: Intraparticle diffusion coefficients. J. Hazard. Mater. 2008, 157, 386–396. [CrossRef]
[PubMed]
65. Vadivelan, V.; Kumar, K.V. Equilibrium, kinetics, mechanism, and process design for the sorption of methylene blue onto rice
husk. J. Colloid Interface Sci. 2005, 286, 90–100. [CrossRef] [PubMed]
66. Plazinski, W.; Rudzinski, W.; Plazinska, A. Theoretical models of sorption kinetics including a surface reaction mechanism: A
review. Adv. Colloid Interface Sci. 2009, 152, 2–13. [CrossRef] [PubMed]
67. Wu, F.-C.; Tseng, R.-L.; Juang, R.-S. Characteristics of Elovich equation used for the analysis of adsorption kinetics in dye-chitosan
systems. Chem. Eng. J. 2009, 150, 366–373. [CrossRef]
68. Iftikhar, A.R.; Bhatti, H.N.; Hanif, M.A.; Nadeem, R. Kinetic and thermodynamic aspects of Cu (II) and Cr (III) removal from
aqueous solutions using rose waste biomass. J. Hazard. Mater. 2009, 161, 941–947. [CrossRef]
69. Al-Othman, Z.A.; Ali, R.; Naushad, M. Hexavalent chromium removal from aqueous medium by activated carbon prepared from
peanut shell: Adsorption kinetics, equilibrium and thermodynamic studies. Chem. Eng. J. 2012, 184, 238–247. [CrossRef]
70. Rao, R.A.K.; Khan, U. Adsorption of Ni (II) on alkali treated pineapple residue (Ananas comosus L.): Batch and column studies.
Groundw. Sustain. Dev. 2017, 5, 244–252. [CrossRef]
71. Zou, B.; Zhang, S.; Sun, P.; Ye, Z.; Zhao, Q.; Zhang, W.; Zhou, L. Preparation of a novel Poly-chloromethyl styrene chelating resin
containing heterofluorenone pendant groups for the removal of Cu (II), Pb (II), and Ni (II) from wastewaters. Colloid Interface Sci.
Commun. 2021, 40, 100349. [CrossRef]
72. Karapinar, H.S.; Kilicel, F.; Ozel, F.; Sarilmaz, A. Fast and effective removal of Pb (II), Cu (II) and Ni (II) ions from aqueous
solutions with TiO2 nanofibers: Synthesis, adsorption-desorption process and kinetic studies. Int. J. Environ. Anal. Chem. 2023,
103, 4731–4751. [CrossRef]
73. Abbasi, S.; Foroutan, R.; Esmaeili, H.; Esmaeilzadeh, F. Preparation of activated carbon from worn tires for removal of Cu (II), Ni
(II) and Co (II) ions from synthetic wastewater. Desalin. Water Treat. 2019, 141, 269–278. [CrossRef]
Molecules 2024, 29, 2489 18 of 18

74. Hezarjaribi, M.; Bakeri, G.; Sillanpää, M.; Chaichi, M.J.; Akbari, S.; Rahimpour, A. New strategy to enhance heavy metal
ions removal from synthetic wastewater by mercapto-functionalized hydrous manganese oxide via adsorption and membrane
separation. Environ. Sci. Pollut. Res. 2021, 28, 51808–51825. [CrossRef] [PubMed]
75. Lian, W.; Yang, L.; Joseph, S.; Shi, W.; Bian, R.; Zheng, J.; Li, L.; Shan, S.; Pan, G. Utilization of biochar produced from invasive
plant species to efficiently adsorb Cd (II) and Pb (II). Bioresour. Technol. 2020, 317, 124011. [CrossRef] [PubMed]
76. Taşar, Ş.; Özer, A. A Thermodynamic and Kinetic Evaluation of the Adsorption of Pb (II) Ions Using Peanut (Arachis Hypogaea)
Shell-Based Biochar from Aqueous Media. Pol. J. Environ. Stud. 2020, 29, 293–305. [CrossRef] [PubMed]
77. Yorgun, S.; Vural, N.; Demiral, H. Preparation of high-surface area activated carbons from Paulownia wood by ZnCl2 activation.
Microporous Mesoporous Mater. 2009, 122, 189–194. [CrossRef]
78. Langmuir, I. The adsorption of gases on plane surfaces of glass, mica and platinum. J. Am. Chem. Soc. 1918, 40, 1361–1403.
[CrossRef]
79. Mustafa, G.; Singh, B.; Kookana, R.S. Cadmium adsorption and desorption behaviour on goethite at low equilibrium concentra-
tions: Effects of pH and index cations. Chemosphere 2004, 57, 1325–1333. [CrossRef]
80. Krishna, B.; Murty, D.; Prakash, B.J. Thermodynamics of chromium (VI) anionic species sorption onto surfactant-modified
montmorillonite clay. J. Colloid Interface Sci. 2000, 229, 230–236. [CrossRef]
81. Amin, M.; Alazba, A.; Shafiq, M. Removal of copper and lead using banana biochar in batch adsorption systems: Isotherms and
kinetic studies. Arab. J. Sci. Eng. 2018, 43, 5711–5722. [CrossRef]
82. Wang, J.; Guo, X. Rethinking of the intraparticle diffusion adsorption kinetics model: Interpretation, solving methods and
applications. Chemosphere 2022, 309, 136732. [CrossRef] [PubMed]
83. Pan, M.; Lin, X.; Xie, J.; Huang, X. Kinetic, equilibrium and thermodynamic studies for phosphate adsorption on aluminum
hydroxide modified palygorskite nano-composites. RSC Adv. 2017, 7, 4492–4500. [CrossRef]

Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.

You might also like