You are on page 1of 38

3.

2 Pristine biochar preparation

The watermelon rinds (WMR) were washed with distilled water to remove dirt and

impurities prior to oven drying at 80 ºC. The dried WMR were crushed in an

electric grinder and sieved using 60 to 100 mesh screens with a particle size of

0.20 to 0.40 mm. Then, the WMR was analyzed using thermogravimetric analysis

(TGA, 4000 Perkin Elmer, UK) to determine the temperature range for the

pyrolysis experiment. Thereafter, the powdered WMR were pyrolyzed in a tube

furnace (Carbolite, STF 15/610, UK) at different temperatures (400, 500, 600 and

700 ºC) under nitrogen atmosphere (50 mL/min) for 1 h at the heating rate of 10

ºC/min. The obtained biochar from WMR at different temperatures were labeled

BCW@400, BCW@500, BCW@600, BCW@700 and were stored in an air tight

container until further use.

3.3 Preparation of activated carbon

Herein, BC@700 was activated using zinc nanoparticles and hydrogen peroxide.

Briefly, 90 mL of ((Zn(NO3)2).6H2O (0.1 M) was prepared using deionized water

in a 250 mL Erlenmeyer flask before the addition of 10 mL of Pterocarpus

mildraedii extract from previous study (Egbosiuba et al., 2022). The mixture was

heated to 80 ºC on the hot plate for 2 h under constant stirring at 150 rpm. A color

change from yellow to brown indicate the reduction of zinc nitrate to zerovalent

zinc oxide nanoparticles. The sample was dried in the oven at 105 ºC for 8 h before

calcination at 400 ºC for 6 h to remove contaminants. Particularly, 50 g of


BC@700 was measured into an Erlenmeyer flask of 250 mL containing zinc oxide

nanoparticles (10 g) and H2O2 (100 mL). The mixture was placed in an ultrasonic

bath (SB25-12DT, Ultrasonic Scientz) for sound wave enhanced activation at 50

ºC for 5 h. In the end, the sample was washed severally using distilled water with

the assistance of centrifugation at 5000 rpm for 10 min until a pH close to 7 was

obtained. The activated sample was oven dried overnight and the sample stored in

an airtight bottle labeled AC@ZnPO.

3.4 Characterization of samples

The thermogravimetric analysis (TGA) and derivative thermogravimetric (DTG)

analysis was conducted on the WMR, biochar samples at different temperatures

and the AC@ZnPO in the range of 30℃/5.0(K/min)/900℃ using Al2O3 crucible

and a nitrogen gas environment. Fourier Transform Infrared (FTIR, NICOLET

IS10 spectrometer) was used to determine the surface functional chemistry of the

BC@400, BC@500, BC@600, BC@700, and AC@ZnPO in the range of 500 to

4000 cm-1. The crystallographic structures of the BC@400, BC@500, BC@600,

BC@700, and AC@ZnPO were using X-ray diffraction (XRD, 6000, Shimadzu

scientific, Germany) using a speed rate of 2º/min. Brunauer–Emmett–Teller (BET)

technique was used to determine the surface areas of the samples from the nitrogen

adsorption–desorption isotherm (NOVA4200e, Quantachrome, UK). The surface

morphological characteristics of the samples was examined using SEM (JSM

6390LV, JEOL Inc., Japan). Also, the surface area and pore structure were
analyzed using the Brunauer Emmett Teller (BET) technique. Zeta potential

analyzer (Malvern ZS90) was employed to measure the zeta potential of the

adsorbent at the pH and temperature of 7 and 25 ºC, respectively. The

concentrations of As5+, Cr6+, Cd2+ and Pb2+ in the solution was determined by

atomic adsorption spectrometry (AAS, PG 990, PG Instruments, UK).

3.5 Adsorption experiments

Initially, a simulated metal ions solution was prepared by dissolving As(NO3)3,

Cr(NO3)3, Cd(NO3)2, Pb(NO3)2 in 1000 mL of water to obtain 1000 mL of 1000

mg/L As5+, Cd2+, Cr6+, and Pb2+. Various concentrations of the metal ions including

10, 25, 50, 100, 150, 200 mg/L that are required for the adsorption experiments

were obtained by serial dilution and stored in a fridge at -20 ºC to maintain the

sample volume. Unless stated otherwise, batch adsorption studies were conducted

in triplicate using 200 mL conical flasks covered with a cork at a stirring speed of

150 rpm in a thermostatic shaker. The comparative adsorption capacity

performance of BC@400, BC@500, BC@600, BC@700, and AC@ZnPO towards

the removal of As5+, Cd2+, Cr6+, and Pb2+ was investigated using 2 g/L of the

adsorbent, metal ions concentration of 100 mg/L, pH of 6, and temperature of 25

ºC for 10 min. At the completion of the adsorption experiments, the concentrations

of As5+, Cd2+, Cr6+, and Pb2+in the solution were determined by AAS after filtration

through a Whatman grade 1 filter paper.


Furthermore, the detailed adsorption effect of AC@ZnPO was conducted on the

metal ions at different pH values (1, 2, 3, 4, 5, 6,7), time (10, 20, 30, 40, 50, 60, 90,

120, 150, 180 min), adsorbent dosage (1, 2, 3, 4, 5, 6 g/L), metal ions

concentrations (100, 200, 300, 400, 500, 600 mg/L) and temperatures (25, 30 and

40 ºC) in a batch system. The different pH value adjustment was carried out by 1

mol/L each for HCl and NaOH. The coexistence adsorption experiments were

performed using selected metal ions including As5+, Cr6+, Cd2+, Pb2+, Cu2+, Fe3+,

Ni2+, and Zn2+ at a constant initial metal ions concentration of 100 mg/L. To

evaluate the reusability of AC@ZnPO, the solution containing 0.1 M of HCl and

NaOH was used as an eluent, while the desorption process was performed for 3 h

at 150 rpm. In eight repeated adsorption cycles, the adsorption and desorption

processes were alternated.

3.6 Data analysis and modelling

The adsorption efficiency (η , %) and the uptake capacity (q e , mg/ g) for the removal

of As5+, Cd2+, Cr6+, and Pb2+ by AC@ZnPO were calculated according to mass

balance Eqs. (3.1) and (3.2) (Chen et al., 2023; Xue et al., 2022).

η=(C ¿ ¿ o−C t )/C o ×100 % ¿ (3.1)

q t=(C ¿ ¿ o−Ct )×V /m ¿ (3.2)

whereby qt (mg/g) denotes the adsorption capacity of the metal ions; Co (mg/L)

refers to the initial concentration of the metal ions; C t (mg/L) is the concentration

of the metal ions at time, t (min); V (L) represents the volume of solution; m (g) is
the added mass of ACW@ZnPO and η (%) is the efficiency of the adsorption

process.

The adsorption modelling determines the performance of the adsorbent and the

mechanism of adsorption process. To this end, the equilibrium data was assessed

by Langmuir, Freundlich, Temkin and Dubinin-Radushkevich (D-R) isotherm

models described in Eqs. (3.3), (3.4), (3.5) and (3.6) (P. Ma et al., 2023; Sadhu et

al., 2022; Sulistiyo et al., 2022).

qm K L C e
q e= (3.3)
1+ K L C e

1
q e =K F C e n (3.4)

RT
q e= ln ( k T Ce ) (3.5)
bT

q e =q m e
− K DR ε2
(
; ε=RTIn 1+
1
Ce); E=
1
√2 K DR
(3.6)

whereby q m (mg/g) and C e (mg/L) represents the maximum adsorption capacity and

the equilibrium concentration, while K L (L/mg), K F ((mg/g)/(L/mg)1/n), k T (L/mg)

and K DR (mol2/kJ2) are the Langmuir, Freundlich, Temkin and D-R constants. Other

D-R parameters such as ε and E (kJ/mol) refers to the Polanyi potential and the

free energy of adsorption.

For the analysis of the rate-controlling steps and the heavy metal ions adsorption

mechanism by AC@ZnPO, the pseudo-first-order (Eq. (3.7)), pseudo-second-order


(Eq. (3.8)), Elovich (Eq. (3.9)), and intraparticle diffusion models (Eq. (3.10)) were

applied to fit the kinetic data (Chen et al., 2023; Ho and McKay, 1999, 1998; Yang

et al., 2023).

q t=qe ( 1−e−k t ) 1
(3.7)

2
k 2 qe t
q t= (3.8)
1+k 2 q e t

1
q t= ∈ ( 1+αβt ) (3.9)
β

0.5
q t=k 3 t +C (3.10)

where qt (mg/g) and qe (mg/g) refers to the adsorption capacity at time t (min) and

the adsorption capacity at equilibrium, while k 1 (min-1), k2 (g·mg-1·min-1), and k3

(mg·g-1·min-0.5) are the rate constants associated with the kinetic models of the

pseudo-first-order, pseudo-second-order, and intraparticle diffusion, respectively. α

(mg/g min) and β (mg/g) are the initial rate constant and the desorption constant of

Elovich model. C (mg/g) is the constant of intraparticle diffusion that is related to

the boundary layer thickness.

The thermodynamics parameters of As 5+, Cd2+, Cr6+, and Pb2+ adsorption by

AC@ZnPO was evaluated to determine the change in enthalpy (ΔH°, J/mol),

change in Gibbs free energy (ΔG°, J/mol), and change in entropy (ΔS°, J/(mol⋅K))

using Eqs. (3.11), (3.12), and (3.13).


qe
k c= (3.11)
Ce

ΔG ° =−RTIn k c (3.12)

− ΔG ° ΔS ° ΔH °
Ink c= = − (3.13)
RT R RT

where k c, R (8.314 J/mol⋅K) and T (K) denotes the thermodynamic equilibrium

constant, the universal gas constant and the temperature of the experimental.
CHAPTER FOUR

4.0. RESULTS AND DISCUSSION

4.1 Physicochemical properties of the materials

Table 4.1 illustrates the physicochemical properties of the WMR, BC@400,

BC@500, BC@600, BC@700 and AC@ZnPO. The yield of the biochar decreased

with increasing pyrolysis temperature and may be attributed to the mass

degradation of the biomass materials at elevated temperature (B. Wang et al.,

2023). On the other hand, the yield of the AC@ZnPO increased after activation

due to the incorporated activating materials that provided improved binding sites

for the adsorbent material (Quang et al., 2022). The MC demonstrates the

requirements of the drying energy that significantly influences the process

economics and energy recovery in pretreatment techniques. The VM are the

components of the materials liberated as gas or vapour under high temperature

heating in the absence of oxygen, which are linked to the crystal, and functional

morphology of cellulose that determines the devolatilization, ignition, burning and

biochar combustion (Fakayode et al., 2021). Also, the FC are the non-volatile

fraction and the combustible solid residue remaining after heating, while the ash

content is the uncombustible mineral materials in the mass fraction of the samples.

As shown in Table 4.1, the MC and VM of the biochars decreased, while the ash

content increased at elevated pyrolysis temperature.


Table 4.1 Properties of the WMR, pristine biochar at different temperatures and the
activated carbon.
Parameter WMR BC@400 BC@500 BC@600 BC@700 AC@ZnPH
Yield (%) - 46.2±1.65 42.4±1.48 37.5±1.22 31.8±1.05 75.62±1.80
MC (%) 9.73±0.33 4.40±0.41 3.66±0.30 2.81±0.40 1.35±0.32 1.20±0.35
VM (%) 74.50±1.60 35.62±1.55 29.20±1.03 23.74±1.20 16.22±0.98 15.80±1.00
Ash (%) 0.84±0.10 3.6±0.13 5.30±0.15 6.68±0.20 8.15±0.11 8.12±0.12
FC (%) 14.93±1.25 56.3±1.58 61.10±1.96 65.18±2.10 71.23±2.00 74.88±2.44
C (%) 40.25±1.82 52.12±2.00 57.21±1.74 62.32±2.43 67.75±1.96 70.65±2.25
H (%) 6.11±0.15 5.14±0.10 4.99±0.11 4.86±0.14 4.71±0.11 8.57±0.12
N (%) 3.80±0.26 3.52±0.24 3.40±0.19 3.25±0.15 3.20±0.20 4.30±0.25
S (%) 2.46±0.18 2.00±0.02 1.85±0.45 1.54±0.22 1.30±0.01 1.25±0.13
O (%) 35.04±1.04 34.09±1.22 32.47±0.98 31.11±1.41 29.37±1.82 40.50±1.70
H/C (%) 0.15±0.001 0.098±0.002 0.087±0.001 0.078±0.003 0.070±0.002 0.121±0.001
O/C (%) 0.87±0.004 0.65±0.003 0.57±0.007 0.50±0.004 0.43±0.006 0.57±0.002
Surface area (m2/g) - 119.53±2.45 157.446±3.40 171.497±2.0 194.810±3.21 327.934±4.65
0
Pore size (nm) - 3.413±0.03 3.192±0.02 2.940±0.05 2.226±0.03 2.027±0.04
Pore volume (cm³/g) - 0.033±0.001 0.065±0.002 0.086±0.001 0.105±0.003 0.258±0.002

Similarly, the MC and VM values of AC@ZnPO was observably lower, while the

ash content showed increasing trend compared with the biochar samples. The

higher ash content of the biochar at different temperatures and AC@ZnPO than the

WMR suggests higher amount of inorganic residue, for instance minerals

(Medeiros et al., 2023). Importantly too, the FC of the biochar samples were found

to increase with increasing temperature. For the AC@ZnPO, the FC increased after

the activation process, compared to the biochar samples. The trend of this results

may be attributed to the volatiles devolatilization at high pyrolysis temperature.

Similar findings had been reported by a number of studies on biochar and activated

biochar from watermelon rinds (Fakayode et al., 2021; Li et al., 2019; Sadhu et al.,

2022; B. Wang et al., 2023).


The C, H, N, S and O elemental compositions of the biochar samples declined with

increasing the pyrolysis temperature from 400 to 700 ºC due to the release of

volatiles (Muhammad et al., 2022). Except for S, the compositions of the C, H, N

and O were comparatively higher in the activated AC@ZnPO than the biochar

obtained at the pyrolysis temperature of 700 ºC. Particularly, C and O were the

main elements in the biochars and the activated AC@ZnPO with the later having a

higher C and O values of 70.65±2.25 and 40.50±1.70, respectively. The observed

higher C, H, N and O values in AC@ZnPO may be attributed to the activation

effectiveness of zinc nanoparticles and hydrogen peroxide to enhance the nitrogen

rich surface of the adsorbent with oxygen containing functional groups (Medeiros

et al., 2023). In general, the development of microstructures decreases the atomic

ratios (H/C and O/C) and increases the amorphous carbon. The observed decrease

in the ratios of the H/C and O/C indicate that dehydration of polycondensation and

dehydrogenation of polymerization facilitated enhanced loss of oxygen and

aliphatic hydrogen during the pyrolysis of biomass (B. Wang et al., 2023). Above

all, the H/C representing the degree of carbonization of biomaterials is generally

lesser than 0.5, indicating possible decomposition resistance due to high

aromaticity and carbonization (Muhammad et al., 2022). The lower values of O/C

indicate the availability of less polar groups with higher hydrophilicity (B. Wang et

al., 2023). Notably, zinc nanoparticles and hydrogen peroxide activation of biochar

enhanced the atomic ratios (H/C and O/C) than the biochar at different

temperatures, indicating that the AC@ZnPO had lower aromaticity and higher
affinity towards water soluble compounds compared to the biochar samples. The

activating materials also contributed to oxygen retention during pyrolysis and was

supported by higher O/C ratio for AC@ZnPO than the biochar at different

temperatures. These observations were similar to the results of biochar and

activated carbon obtained from different agricultural wastes (Muhammad et al.,

2022; Xiang et al., 2022).

4.2 Characterization of materials

4.2.1 Structural characteristics

Table 4.1 provides the surface area and porous properties of the biochars

(BC@400, BC@500, BC@600, and BC@700) and the activated carbon

(AC@ZnPO). Among the biochars and the activated carbon, AC@ZnPO possessed

a higher surface area of 327.934±4.65 m2/g, compared to BC@400 (119.53±2.45

m2/g), BC@500 (157.446±3.40 m2/g), BC@600 (171.497±2.00 m2/g) and

BC@700 (194.810±3.21 m2/g). Importantly too, the pore sizes of the biochar

decreased from 3.413 to 2.226 nm, while the pore volume increased from 0.033 to

0.105 cm3/g. After the activation of the biochar, the pore size of AC@ZnPO further

reduced to 2.027 nm, while the pore volume increased to 0.258 cm3/g. Overall, the

superior porous properties of AC@ZnPO is an indication of enhanced

sequestration of pollutants from wastewater.


250
0.20
AC@ZnPO BC@700 AC@ZnPO
BC@600 BC@500 BC@700
3 /g)

200 BC@400 BC@600


0.15
BC@500

3 /g)
N 2 Adsorption quantity (cm

150 BC@400

Pore volume (cm


0.10
100

0.05
50

0.00
0
0.0 0.2 0.4 0.6 0.8 1.0 0 20 40 60 80 100
Relative pressure (P/Po) Pore size (nm)
C-O/C-O-C (d)
C=C
-CH2
O-H (c)
AC@ZnPO
C-O
C-H
AC
-CH3 AC@ZnPO BC@600 BC@700
C=O
BC@500
Transmittance (a.u.)

BC@400
Intensity (a.u.)

BC@700 BC@600 AC@Z


BC@300
Quartz BC@7
BC@600 Graphite BC@6
AC BC@500 BC@5
ZnO
BC@500 BC@600 oC BC@4

BC@500 oC BC@400
BC@400
BC@400 oC
500 1000 1500 2000 2500 3000 3500 4000 4500 20 40 60 80 100
-1
Wavenumber (cm ) BC@300 oC2 Theta (degree)

Figure 4.1 (a) Nitrogen adsorption-desorption isotherms; (b) Pore size distribution curves;
(c) FTIR and (d) XRD of BC@400, BC@500, BC@600, BC@700 and AC@ZnPO.
Additionally, the observed increase on the yield of AC@ZnPO (Table 4.1) after

biochar activation using zinc oxide nanoparticles and hydrogen peroxide is an

indication of the activating materials enhanced effectiveness in carbonaceous

materials decomposition, thereby facilitating polymerization reactions (Medeiros et

al., 2023). The observed decrease in the pore sizes and increase in the pore

volumes with increase in pyrolysis temperature were significantly comparable with

previous studies (Huang et al., 2023; Medeiros et al., 2023; Muhammad et al.,
2022). The N2 physisorption isotherms and pore size distribution representations of

the biochars and AC@ZnPO are shown in Figure 4.1(a,b). Based on the IUPAC

classification, the N2 physisorption isotherms of the biochars and AC@ZnPO were

classified as type IV isotherms (Rambabu et al., 2020). The result indicate that the

materials are typically mesoporous (2 – 50 nm) with a uniform and channelized

pore structures (B. Wang et al., 2023).

4.2.2 Surface functional groups

The influence of pyrolysis temperature and the activating agents on the surface

functional groups of BC@400, BC@500, BC@600, BC@700 and AC@ZnPO was

investigated by FTIR spectrum shown in Figure 4.1c. As can be seen from the

results of the biochar samples and AC@ZnPO, the vibration band at 720 cm-1

correspond to the C-H due to the presence of alkanes (Li et al., 2019). Similarly,

the appearance of bands characteristic of oxygen-containing surface groups was

evident at 1050, 1240, 1580 and 3450 cm-1. The peaks at 1050 and 1240 cm-1

corresponded to the symmetric and asymmetric vibrations of C-O/C-O-C and C-O

associated with the carbonyl and phenolic groups, while the identified peak at 1580

cm-1 may be due to the stretching vibration of C=C due to the aromatic groups

(Rambabu et al., 2020). The peak at 3450 cm-1 confirmed the presence of O-H

broad vibration ascribed to the carboxylic groups (Medeiros et al., 2023; B. Wang

et al., 2023). The observed decrease on the intensity of the oxygen containing

groups of BC@400, BC@500, BC@600 and BC@700 indicate the thermal

decomposition of the samples due to the oxygen containing groups release. This
corroborated with the elemental analysis (Table 4.1) that recorded a reduction in

oxygen content with increasing pyrolysis temperature. Notably, distinct peaks were

visible in Figure 4.1c at the wavenumbers of 1708 cm-1 (C=O), 2850 cm-1 (CH2),

and 2930 cm-1 (CH3), which described the vibrations of carbonyl, hydroxyl,

alkenes and alkynes groups present in the AC@ZnPO (Li et al., 2019; Medeiros et

al., 2023; Rambabu et al., 2020). These results indicate that the activation with zinc

nanoparticles and hydrogen peroxide increased the aromatization of the carbon and

enabled the abundance of oxygen containing functional groups on the surface of

AC@ZnPO. Similar observation have been made on the aromatization of biochar

with sustained pyrolysis and activation process (Rambabu et al., 2020; B. Wang et

al., 2023).

4.2.3 Crystallographic properties

The XRD patterns of BC@400, BC@500, BC@600, BC@700 and AC@ZnPO

were conducted to evaluate the phase transformation and crystal lattice

arrangement as can be seen from the spectra presented in Figure 4.1d. From the

XRD diffractograms, it is evident that the highest diffraction peak occurred at

26.55º and was assigned to the crystal plane (002) due to graphite (Rambabu et al.,

2020). In addition, the intensity of the diffraction peaks 26.55º showed

proportionality to the carbon contents of BC@400, BC@500, BC@600, BC@700

and AC@ZnPO presented in Table 4.1. The occurrence of quartz was identified on

the BC@400, BC@500, BC@600, and BC@700 at the diffraction peaks of

23.40º(111), 50.05º(400), and 69.10º(440). However, the presence of quartz on the


AC@ZnPO was only found at the 2 Theta values of 23.40º and 50.05º after

activation processes shown in Figure 4.1d. The diffraction peak at 43.20º occurred

only on the AC@ZnPO after activation with zinc nanoparticles and hydrogen

peroxide and was assigned to the crystal plane (100) due to graphite carbon (Sadhu

et al., 2022). The graphite structure of biochars and AC@ZnPO may assist the π-π

interaction between the adsorbent surface and the pollutants (B. Wang et al., 2023).

Above all, the XRD spectra results demonstrated that the pyrolysis temperature

increased the crystallinity and the mineralogical phases of the biochar.

Furthermore, the diffraction peak at 75.20º assigned to the crystal plane of (202)

revealed the successful impregnation of zinc particles on the biochar after

activation due to the presence of peak associated with ZnO.

4.2.4 Thermal stability

Herein, the TGA-DTG analysis was conducted on the WMR to determine the best

pyrolysis temperature range for the biomass. The result presented in Figure 4.2a

revealed three remarkable thermal degradation zones in the weight loss of WMR.

The first stage, 10.07 % weight loss was observed in the temperature regions of 50

to 150 ºC and may be ascribed to the loss of moisture and acid functional groups.

The weight loss of 24.03 % in the second temperature range of 150 to 300 ºC could

be due to the decomposition of cellulosic and hemicellulosic contents of the WMR.

In the third degradation stage, a weight loss of 38.18 that occurred at the

temperature zone of 300 to 900 ºC was attributed to the decomposition of lignin.


100 0.5 0.5
100
10.07 % (a)
0.0 TGA7.31 % (b)
DTG 0.0
80 -0.5 90
24.03 %

C
o

o
-1.0

Derivative weight, %/
Derivative weight, %/
-0.5
Weight, %

Weight, %
60 -1.5 80 534.61 ºC
38.18 % -2.0 -1.0
30.03 %
40 -2.5 70 TGA
TGA -1.5
-3.0 DTG
DTG 340.45 ºC
20 DTG 60
0 200 400 600 800 1000 0 200 400 600 800 1000
Temperature, oC Temperature, oC
100 0.5 100 0.5
(c) (d)
8.51 % 11.34 %
0.0 0.0
90 90

C
C

o
o

Derivative weight, %/
Derivative weight, %/

-0.5 -0.5
Weight, %

Weight, %

80 594.19 ºC 80 611.48 ºC
19.78 %
25.37 % -1.0 -1.0
70 70
TGA -1.5 TGA -1.5
DTG DTG
60 60 DTG
0 200 400 600 800 1000 0 200 400 600 800 1000
o
Temperature, C Temperature, oC
100 0.5 100 0.5
(e) 5.64 % (f)
6.69 %
11.55 % 0.0 8.09 % 0.0
90 90 C
C

o
o

695.55 ºC 740.79 ºC
Derivative weight, %/

Derivative weight, %/

-0.5 -0.5
Weight, %

Weight, %

80 80

-1.0 -1.0
70 70
TGA -1.5 TGA -1.5
DTG DTG
60 DTG 60 DTG
0 200 400 600 800 1000 0 200 400 600 800 1000
Temperature, C o
Temperature, oC

Figure 4.2. Curves of the thermogravimetric (TGA) and the initial derivative of
thermogravimetric (DTG) of (a) WMR; (b) BC@400 ºC; (c) BC@500 ºC; (d) BC@600 ºC;
(e) BC@700 ºC and (f) AC@ZnPO.
The peak degradation temperature of 340.45 ºC was evident on the WMR, thereby

influencing the choice of 400, 500, 600 and 700 ºC as the pyrolysis temperature in

this study. On the other hand, the TGA-DTG analysis of the biochar samples at

different temperatures revealed only two weight loss zones as can be seen in Figure

4.2(b-e). The initial weight loss of 8.51, 6.69, 11.34 and 5.64 % were recorded for

BC@400, BC@500, BC@600, and BC@700 at the temperature zones of 50 to

120, 50 to 230, 50 to 250 and 50 to 300 ºC, respectively. From the DTG curve, a

maximum weight loss of 30.03, 25.37, 19.78, and 11.55 % were recorded for

BC@400, BC@500, BC@600, and BC@700 at the temperatures of 534.61,

594.19, 611.48 and 695.55 ºC. In the case of AC@ZnPO presented in Figure 4.2f,

the thermostability significantly increased after activation using zinc oxide

nanoparticles and hydrogen peroxide. According to Figure 4.2f, the DTG curve of

AC@ZnPO showed a weight loss of 5.64 and 8.09 % at the temperature range of

50 to 300 ºC due to loss of physisorbed water (Rambabu et al., 2020) and 300 to

900 ºC due to degradation of volatile and organic compounds (Medeiros et al.,

2023). The peak degradation temperature of AC@ZnPO was obtained as 740.79

ºC. It is evident from the TGA-DTG results that at the final pyrolysis temperature

of 900 ºC, the residual mass for the different materials is WMR (27.72 %),

BC@400 (61.62 %), BC@500 (65.00 %), BC@600 (68.87 %), BC@500 (81.76

%), and AC@ZnPO (85.30 %). From the TGA curves, the thermal stability of the

materials was in the order AC@ZnPO > BC@700 > BC@600 > BC@500 >

BC@400, while the maximum weight loss followed the trend of AC@ZnPO <
BC@700 < BC@600 < BC@500 < BC@400, thereby corresponding with the

result of the FTIR spectra presented in Figure 4.1c. The TGA and DTG results of

this study showed agreement with similar degradation trend as the pyrolysis

temperature increases for WMR derived biochar and activated carbon (Medeiros et

al., 2023; Muhammad et al., 2022; Quang et al., 2022; Rambabu et al., 2020).

4.2.5 Morphological characteristics

The SEM images of BC@400, BC@500, BC@600, BC@700 and AC@ZnPO are

shown in Figure 4.3 at different magnifications. An irregular shaped block

structures of the biochars were obtained after pyrolysis at different temperatures.

As can be seen in Figure 4.3(a,b), a coarse porous structure with uneven surface

characteristics were revealed after pyrolysis at 400 ⁰C due to low temperature

effect. The SEM images revealed that the particle size of the biochars were finer

with increment in the pyrolysis temperature. Noticeably, the surface of AC@ZnPO

exhibited a honeycomb-like morphology after activation using zinc nanoparticles

and hydrogen peroxide. Similarly, the morphology of the AC@ZnPO revealed the

interconnectivity of the porous structure activated by zinc nanoparticles and

hydrogen peroxide. With the activation of biochar, the porous surfaces and edges

are enhanced, thereby improving the number of active sites rich in oxygenated

functional groups for the mesoporous adsorption of As5+, Cr6+, Cd2+ and Pb2+ from

wastewater.
(a)

(c) (d)

(e) (f)

(g) (h)

(i) (j)

Fig. 3 SEM images of (a, b) BC@400 ºC; (c, d) BC@500 ºC; (e, f) BC@600 ºC; (g, h)
BC@700 ºC and (i, j) AC@ZnPO at lower and higher magnifications.
In another study, Wang et al. (B. Wang et al., 2023) reported that activated carbon

pores possess gaps between the microcrystals, resembling cross-links of wedges

producing a cage-like mesoporous structure.

4.3 Adsorption performance

4.3.1 Comparative adsorption studies

Adsorption capacity indicates the quantity of metal ions adsorbed on the adsorbent.

The preliminary tests on the adsorption capacity performance of BC@400,

BC@500, BC@600, BC@700, and AC@ZnPO towards the removal of As5+, Cd2+,

Cr6+, and Pb2+ was investigated at a constant adsorbent dosage of 2 g/L, metal ions

concentration of 100 mg/L, pH of 6, temperature of 25 ºC, and time of 10 min

(Figure 4.4a). Previous studies have reported that pyrolysis temperature influences

the adsorption of pollutants because it enhances the surface area of biochar (Qi et

al., 2022). For the biochar materials, the adsorption capacities of the metal ions

increased as the pyrolysis temperature increased from 400 to 700 ⁰C. In this case,

BC@700 adsorbed 224 mg/g of As5+, 186 mg/g of Cd2+, 197 mg/g of Cr6+ and 162

mg/g of Pb2+ compared to BC@400 with the adsorption capacities of 172, 143, 153

and 132 mg/g for As5+, Cd2+, Cr6+, and Pb2+, respectively. The adsorption of heavy

metals on the biochar can occur through binding to the carbonized fraction and the

partitions of the noncarbonized organic matter fraction (Qi et al., 2022; Wang et

al., 2016). As such, the higher metal ions adsorption onto BC@700 may be

attributed to the enlarged surface area at the elevated pyrolysis temperature.


300 300
BC@400 BC@500
(a) BC@600 BC@700 (b)
250 250
AC@ZnPO
Adsorption capacity (mg/g)

Adsorption capacity (mg/g)


200
200

150
150
As5+
100
100 Cd2+
50 Cr6+
50 Pb2+
0
As^5+
As5+ Cd^2+
Cd2+ Cr^6+
Cr6+
Pb2+ Pb^2+ 0 1 2 3 4 5 6 7 8
Heavy metal adsorption pH
120 300
As5+ Cd2+ (c)
500
(d)
6+ 2+
100 Cr Pb 250 Adsorption capacity (mg/g) 400
80
Adsorption capacity (mg/g)

200
Removal efficiency (%)

300
As5+
60 150
200 Cd2+
40 Cr6+
100
100 Pb2+
20
As5+ Cd2+ 50
6+ 2+ 0
0 Cr Pb
0
0 1 2 3 4 5 6 7 0 20 40 60 80 100 120
Adsorbent dosage (g/L) Time (min)
120 800 550
(e) (f)
500
100
600
Adsorption capacity (mg/g)

450
Adsorption capacity (mg/g)

80
Removal efficiency (%)

400
400
60 350 As5+
Cd2+
40 As5+ Cd2+ 200 300
Cr6+
Cr6+ Pb2+
20
250 Pb2+
As5+ Cd2+ 0
200
Cr6+ Pb2+
0 298 300 302 304 306 308 310 312 314
0 50 100 150 200
Temperature (K)
Initial metal concentration (mg/L)
Figure 4.4 (a) Adsorption capacity performance adsorbents (adsorbent dosage = 2 g/L, metal ions
concentration = 100 mg/L, pH = 5, temperature = 25 ºC, and = 10 min); (b) effect of pH on the metal ions
adsorption using AC@ZnPO; (c) effect of different adsorbent dosage at a constant pH (3 for Cr 6+ and 5 for
As5+, Cd2+ and Pb2+), time (10 min), metal ions concentration (100 mg/L), and temperature (25 ºC); (d)
effect of different time at a constant pH (3 for Cr6+ and 5 for As5+, Cd2+ and Pb2+), adsorbent dosage (5 g/L),
metal ions concentration (100 mg/L), and temperature (25 ºC); (e) effect of different initial heavy metal
ions concentrations at a constant pH (3 for Cr 6+ and 5 for As5+, Cd2+ and Pb2+), adsorbent dosage (5 g/L),
time (60 min), and temperature (25 ºC); (f) effect of different temperature at a constant pH (3 for Cr 6+ and
5 for As5+, Cd2+ and Pb2+), adsorbent dosage (5 g/L), time (60 min), and initial heavy metal ions
concentrations (100 mg/L).
However, the observed adsorption capacity of BC@400 was related to the

combined influence of the surface area and the organic matter fraction.

Furthermore, the activation of the BC@700 using zinc oxide nanoparticles and

hydrogen peroxide revealed increased adsorption capacity compared to the biochar

materials through the improvement of the surface area of AC@ZnPO as can be

seen in Table 4.1. From the results in Figure 4.4a, AC@ZnPO recorded the highest

adsorption capacity of 250 mg/g for As5+, 208 mg/g for Cd2+, 217 mg/g for Cr6+ and

188 mg/g for Pb2+. In all, AC@ZnPO with the surface area of 327.934 m 2/g

demonstrated a higher adsorption capacity compared to the BC@700, BC@600,

BC@500, and BC@400 with the surface areas of 194.810, 171.497, 157.446 and

119.530 m2/g, respectively. Overall, AC@ZnPO was the best adsorbent for the

metal ions adsorption, followed by BC@700, BC@600, BC@500 and BC@400.

The superior performance of AC@ZnPO may be ascribed to the improved binding

sites on the adsorbent surface through zinc oxide nanoparticles and hydrogen

peroxide enhanced activation process. Notably too, H 2O2 introduces oxygen

containing functional groups to the activated carbon surface which increased the

active sites for the adsorption of the metal ions (Qi et al., 2022). This result

highlights the crucial influence of pyrolysis temperature and activation materials

on the biochar and activated carbon, thereby showing that AC@ZnPO is the most

suitable material for the detailed exploration of the adsorption applications.


4.3.2 Effect of pH

The value of the pH is an important parameter for the adsorption process which

effect the ionization degree of the metal ions in solution and the surface charge of

the adsorbent (Wan et al., 2023; Yuan et al., 2023). Additionally, the solution pH

influences the degree of protonation on the surface of AC@ZnPO (F. Ma et al.,

2023). In this study, the impact of pH was conducted on the removal of As 5+, Cd2+,

Cr6+, and Pb2+ using AC@ZnPO at various pH values of 1 to 7, with the results

shown in Figure 4.4b. The experimental range of pH was conducted from 1.0 to

7.0 due to the hydroxide precipitation of As5+ and Cd2+ at pH > 7 and Pb2+ at pH >

5. The results revealed that the adsorption capacity of AC@ZnPO towards As5+,

Cd2+, and Pb2+ rose with increment in pH. Particularly, increase in the pH from 1.0

to 6.0 remarkably increased the adsorption capacity of AC@ZnPO from 125 to 250

mg/g for As5+, 80 to 211 mg/g for Cd2+, and 50 to 186 mg/g for Pb2+, respectively.

On the other hand, the adsorption capacity of Cr 6+ increased from 164 to 278 mg/g

at the pH of 1.0 to 3.0. Beyond this point, the adsorption capacity of Cr 6+ declined

to 155 mg/g as the pH was increased to 7.0. Overall, further batch adsorption study

was examined at pH 5.0 for As5+, Cd2+, and Pb2+ , while the adsorption of Cr6+ was

investigated at pH 3.0 to mitigate possible precipitation of metal ions that may

interfere with the adsorption process (Baby et al., 2023; Egbosiuba et al., 2022).

To further explain the influence of pH on the adsorption of the metal ions, zeta

potential was used to explore the surface charge of the AC@ZnPO at different pH

values (Figure S1). It confirmed that the surface of AC@ZnPO was positively
charged when pH < pHPZC and negatively charged when pH > pHPZC. Therefore, the

adsorption of negatively charged metal ions such as Cr 6+ that occurs largely HCrO4-

and CrO4- in acid, neutral and alkaline pH are favored at pH < pHPZC due to the

protonation of the surface functional groups of AC@ZnPO to increased

concentration of H+ (Baby et al., 2019; Egbosiuba et al., 2022). However, pH >

pHPZC favors the adsorption of positively charged metal ions such as As5+, Cd2+, and

Pb2+ due to increase in the amount of OH- attributed to the negative charge on the

adsorbent (H. Wang et al., 2023). The reason for the observed behavioral pattern of

the adsorption process may be attributed to the protonation of the surface

functional groups of AC@ZnPO due to the increased concentration of H + at pH <

pHPZC. Above all, AC@ZnPO electrostatically adsorbed the metal ions with both

negative and positive valence at low and high pH values. Furthermore, other

studies have reported effective removal of As5+, Cd2+, and Pb2+ at less acidic or

neutral pH (Gao et al., 2023; Li et al., 2023; Liu et al., 2023; Yuan et al., 2023),

whereas a higher removal efficiency have been reported for Cr6+ in a strong acidic

environment (Jyoti et al., 2020; Sulistiyo et al., 2022; H. Wang et al., 2023). For

this study, the adsorption capacity of AC@ZnPO for Cr6+ and As5+ was higher than

that of Cd2+, and Pb2+ which may be attributed to the increased mobility of the

higher valence metal ions to the active sites of the adsorbent.

4.3.3 Effect of adsorbent dosage

The impact of AC@ZnPO dosage on the removal efficiency and adsorption

capacity of the heavy metal ions was investigated in the range of 1 to 6 g/L and the
result presented in Figure 4.4c. It can be seen from the results that the removal

efficiency of the heavy metal ions increased as the adsorbent dosage was increased

and the observed adsorption characteristics may be due to increase in the available

active sites on the surface of AC@ZnPO (Dong et al., 2020; Miao et al., 2021). In

addition, the adsorption capacity of the heavy metal ions decreased with an

increment in the adsorbent dosage which may be ascribed to the increase in the

number of available active sites favorable for the heavy metal ions adsorption on

the surface of AC@ZnPO, thereby declining the adsorption capacity due to the

reduction in the adsorbed mass of species per unit weight of the adsorbent (Ahmed

et al., 2021; Dong et al., 2020; Tang et al., 2022). Figure 4.4c revealed that the

removal efficiency of the heavy metal ions increased from 50.53% to 94.20% for

Cr6+, from 35.80% to 87.10% for As5+, from 31.30% to 77.52% for Cd2+ and from

26.06% to 70.02% for Pb2+ using 1 to 5 g/L of the AC@ZnPO. The increasing

behavioral trend in the removal efficiency of the metal ions as the adsorbent

dosage increased may be due to the abundant availability of active sites for the

removal of the water pollutants (Baby et al., 2019; Jyoti et al., 2020). Further

increment in the adsorbent dosage to 6 g/L did not translate to significant removal

efficiency of the metal ions and may be attributed to the saturation of the binding

sites on the surface of the adsorbent by the metal ions (Li et al., 2019). For the

adsorption capacity, increase in the adsorbent dosage led to a sequential decrease

in the adsorption capacity from 179.00 to 72.63 mg/g for As 5+, from 156.45 to

63.80 mg/g for Cd2+, from 252.65 to 79.17 mg/g for Cr 6+, and from 125.30 to 41.24
mg/g for Pb2+. The results of this study showed that the smaller the quantity of the

adsorbent, the faster the exposure and saturation rate of the adsorption sites,

whereas higher adsorbent dosage produces excess number of unoccupied

adsorption sites during adsorption that result to lower adsorption capacity. Based

on the findings of this study, 5 g/L was selected as the best adsorbent dosage for

the rest of the experiments.

4.3.4 Effect of adsorption time

The evaluation of the equilibrium time which marks the point for the maximum

uptake of the heavy metal ions is very crucial and is conducted at different time

intervals. Herein, the influence of adsorption time on the heavy metal ions

adsorption by AC@ZnPO was evaluated and the results displayed in Figure 4.4d.

The result demonstrated a rapid increase in the adsorption capacity within the first

40 min of the adsorption time and then increased slowly until maximum uptake

was attained at the equilibrium time of 60 min. The observed adsorption trend with

increasing time may be attributed to the vacant active sites largely available at the

initiation of the adsorption process and the high gradient of the solute

concentration (Sadhu et al., 2022). At the equilibrium time of 60 min, the

maximum adsorption capacity evaluated for the heavy metal ions were As 5+

(454.20 mg/g), Cd2+ (412.50 mg/g), Cr6+ (494.10 mg/g), and Pb2+ (377.00 mg/g),

respectively. With further the extension of time to 90 and 120 min, the adsorption

of heavy metal ions by AC@ZnPO exhibited an insignificant change in the

adsorption capacity, which was ascribed to the decrease of free binding sites that
are available for adsorption process (Chen et al., 2023). Apparently, the observed

trend in the adsorption capacity of the heavy metal ions revealed that more of Cr 6+

and As5+ were adsorbed ahead of Cd2+ and Pb2+ which may be attributed to the

ionic radius of the metal ions. From the literature, the ionic radius of As 5+, Cd2+,

Cr6+, and Pb2+ were reported as 0.46, 0.78, 0.44 and 0.98 Å (Egbosiuba et al., 2022;

Pawar et al., 2018). Heavy metal ion with a lower ionic radius diffuse faster to the

surface of the adsorbent compared to the metal ion with a higher ionic radius

(Chen et al., 2019; Huang et al., 2020). Consequently, the removal efficiency of the

heavy metal ions onto the reactive surface binding sites of AC@ZnPO

demonstrated preference towards Cr6+ followed by As5+, Cd2+, and Pb2+. In addition,

the higher valence electron of the Cr 6+ may be linked to its faster adsorption ahead

of As5+ and lastly trailed by Cd2+ and Pb2+, composed of divalent electrons.

4.3.5 Effect of initial metal concentration

Figure 4.4e displayed the results of the removal efficiency and the adsorption

capacity for the removal of As5+, Cd2+, Cr6+, and Pb2+ by AC@ZnPO at different

initial metal ions concentration. It is evident from the result that increase in the

initial heavy metal ions concentration, gradually decreases the adsorption

efficiency and steadily increases the adsorption capacity of AC@ZnPO towards the

heavy metal ions. Particularly, increasing the initial heavy metal ions concentration

from 10 to 200 mg/L significantly decreased the removal efficiency from 92.50 to

65.48% for As5+, 88.60 to 60.40% for Cd2+, 98.84 to 70.68% for Cr6+, and from

74.10 to 50.45% for Pb2+. The observed decrease in the uptake efficiency of the
heavy metal ions may be attributed to the availability of limited number of active

sites on the surface of AC@ZnPO (Dong et al., 2020; Sadhu et al., 2022). In

general, the quantity of heavy metal ions at a high initial concentration exceeds the

number of adsorption sites, thereby not accommodating the binding of all the metal

ions and resulting to a decreased adsorption efficiency sequel to the saturation of

the adsorbent surface (Sulistiyo et al., 2022). Conversely, lower initial metal ions

concentration indicates that the amount of heavy metal ions in the solution is

proportional to the number of the binding sites present on the adsorbent, which

enhances improved removal of the heavy metal ions. On the other hand, the

adsorption capacity of the AC@ZnPO towards the heavy metal ions increased

rapidly from 46.25 to 654.80 mg/g for As5+, from 44.30 to 604.00 mg/g for Cd2+,

from 50.00 to 706.80 mg/g for Cr6+, and from 37.05 to 504.50 mg/g for Pb2+,

respectively as the initial heavy metal ions concentration was increased from 10 to

150 mg/L. Further increment in the initial heavy metal ions concentration to 200

mg/L obviously recorded a slower increase in the adsorption capacity due to the

reduction in the active sites and larger amount of heavy metal ions layer covering

the adsorbent surface (H. Wang et al., 2023). Overall, the adsorption capacity of

Cr6+ and As5+ were significantly higher than that of Cd 2+ and Pb2+ which

corresponds with the results of previous studies on the adsorption of the metal ions

using a different material (Baby et al., 2023).


4.4 Kinetic studies

The investigated kinetic parameters revealed that pseudo-second order kinetic model fitted better

to the kinetic data compared to the pseudo-first-order, Elovich and intraparticle diffusion due to

the high values of the correlation coefficient (R 2) and lower error values of X2 and SSE as can be

seen in Figure 4.5 and Table 4.2.

600 600
(a) (b)
500 500

400 400

300 300 Experimental data


q e (mg/g)

q e (mg/g)

Experimental data
As5+ As5+
200 200
Cd2+ Cd2+
100 Cr6+ 100 Cr6+
Pb2+ Pb2+
0 0 Pseudo-second order curve
Pseudo-first order curve
0 20 40 60 80 100 120 140 0 20 40 60 80 100 120 140
Time (min) Time (min)
600 600
(c) (d)
500 500 As5+
As5+
Cd2+
400 400 Cd2+
6+ Cr6+
Cr
Pb2+
300 300 Pb2+ Intraparticle diffus
Experimental data Experimental data
q e (mg/g)

q e (mg/g)

Elovich curve Intraparticle diffus


5+
200 As 200
Elovich curve As5+ Intraparticle diffus
Elovich curve
Cd2+ Elovich curve Cd2+ Intraparticle diffus
100 Cr6+ 100 Cr6+
Pb2+ Pb2+
0 0
Elovich curve IPD curve
0 20 40 60 80 100 120 140 0 20 40 60 80 100 120 140
Time (min) Time (min)

Figure 4.5 Kinetics curves for (a) pseudo-first-order; (b) pseudo-second-order; (c) Elovich
and (d) intraparticle diffusion model for the adsorption of heavy metals by AC@ZnPO.
Table 4.3. Determined parameters of different kinetic models for metal ions removal by AC@ZnPO.

Kinetic models Parameters As5+ Cd2+ Cr6+ Pb2+


qe, (mg/g) 434.622±13.58
462.558±5.020 8 497.284±4.216 408.495±23.351
Pseudo-first order K1 (1/min) 0.050±0.002 0.038±0.004 0.058±0.002 0.032±0.005
X2 2.403 7.928 1.705 8.239
SSE 14.418 47.568 10.228 49.436
R2 0.960 0.937 0.975 0.903
qe, (mg/g) 549.382±21.31 543.440±40.95
9 1 578.205±17.045 535.143±62.569
Pseudo-second K2 x 10-
order 3
(1/min) 0.109±0.0001 0.072±0.0002 0.128±0.0001 0.054±0.0003
X2 0.409 3.823 0.398 5.143
SSE 2.453 22.937 2.388 30.859
R2 0.993 0.969 0.994 0.939
α (g/mg min) 67.119±26.578 31.742±11.930 110.891±49.259 19.566±7.314
β (mg/g) 0.008±0.0013 0.0072±0.0015 0.008±0.001 0.006±0.002
Elovich X2 6.818 13.249 6.426 11.714
SSE 40.909 79.495 38.559 70.284
R2 0.885 0.894 0.907 0.862
k3 43.772±6.425 42.527±5.822 46.142±7.413 40.558±5.602
C (mg/g) 69.804±43.893 32.068±39.774 93.611±50.637 5.992±3.268
Intraparticle X2 15.592 24.522 15.812 18.634
diffusion SSE 93.549 147.133 94.873 111.802
R2 0.738 0.804 0.770 0.781

4.5 Isotherm studies

The investigated isotherm models revealed that Langmuir isotherm fitted better to the

equilibrium data compared to the Freundlich, Temkin and D-R isotherm due to the high values of

the correlation coefficient (R2) and lower error values of X2 and SSE as can be seen in Figure 4.6

and Table 4.3.


800 800
(a) (b)
600 600
q e (mg/g)

q e (mg/g)
400 Experimental data 400 Experimental data
As 5+
As5+
200 Cd2+ 200 Cd2+
Cr6+ Cr6+
Pb2+ Pb2+
0 Langmuir curve 0 Freundlich curve
0 20 40 60 80 100 0 20 40 60 80 100
Ce (mg/L) Ce (mg/L)
800 800
(c) (d)

600 600
q e (mg/g)

q e (mg/g)

400 400
Experimental data Experimental data
5+
As As5+
200 Cd2+ 200 Cd2+
Cr6+ Cr6+
0 Pb2+ Pb2+
Temkin curve 0 D-R curve
0 20 40 60 80 100 0 20 40 60 80 100
Ce (mg/L) Ce (mg/L)

Figure 4.6 Isotherm curves for (a) Langmuir; (b) Freundlich; (c) Temkin and (d) D-R
isotherm model for the adsorption of heavy metals using AC@ZnPO.
Table 3. Adsorption isotherm parameters determined for the adsorption of heavy metal ions by
AC@ZnPO.
Isotherm Parameters As5+ Cd2+ Cr6+ Pb2+
model
qm(mg/g) 844.452±32.80 974.922±50.706 764.332±39.23 1285.666±116.512
Langmuir 5 2
KL (L/mg) 0.051±0.005 0.029±0.003 0.200±0.040 0.0116±0.0017
X2 1.492 1.184 7.084 1.220
SSE 5.967 4.734 28.338 4.880
R2 0.996 0.997 0.984 0.996
KF 88.936±20.736 60.141±11.001 196.363±36.76 27.556±6.382
(mg/g)/(L/mg)1/n 5
nF 2.059±0.266 1.751±0.148 2.975±0.490 1.395±0.113
Freundlich X2 15.424 5.811 26.461 4.943
SSE 61.697 23.245 105.845 19.776
R2 0.959 0.986 0.943 0.985
k T (L/g) 1.074±0.378 0.715±0.249 6.733±3.065 0.311±0.093
bT 143.268±18.04 151.243±20.538 116.017±13.81 169.818±24.414
8 8
Temkin X2 27.940 33.943 31.155 36.442
SSE 111.760 135.772 124.619 145.767
R2 0.926 0.917 0.933 0.890
qm(mg/mg) 815.812±74.11 1106.265±190.315 853.954±97.33 1065.744±191.655
3 9
KDR (mol2/kJ2) 0.048±0.009 0.031±0.004 0.216±0.037 0.011±0.002
D-R E (kJ/mol) 1.049±0.124 0.905±0.098 0.792±0.130 1.109±0.118
X2 1.879 1.246 4.7872 1.314
SSE 5.639 3.739 14.362 3.942
R2 0.995 0.997 0.990 0.996
CHAPTER FIVE

5.0 CONCLUSION AND RECOMMENDATIONS

5.1 Conclusion

5.2 Recommendations
REFERENCES

Ahmed, W., Núñez-Delgado, A., Mehmood, S., Ali, S., Qaswar, M., Shakoor, A.,
Chen, D.Y., 2021. Highly efficient uranium (VI) capture from aqueous
solution by means of a hydroxyapatite-biochar nanocomposite: Adsorption
behavior and mechanism. Environ. Res. 201.
https://doi.org/10.1016/j.envres.2021.111518
Baby, R., Hussein, M.Z., Zainal, Z., Abdullah, A.H., 2023. Preparation of
Functionalized Palm Kernel Shell Bio-adsorbent for the treatment of heavy
metal-contaminated water. J. Hazard. Mater. Adv. 10, 100253.
https://doi.org/10.1016/j.hazadv.2023.100253
Baby, R., Saifullah, B., Hussein, M.Z., 2019. Palm Kernel Shell as an effective
adsorbent for the treatment of heavy metal contaminated water. Sci. Rep. 9, 1–
11. https://doi.org/10.1038/s41598-019-55099-6
Chen, L., Wu, K., Zhang, M., Liu, N., Li, C., Qin, J., Zhao, Q., Ye, Z., 2023.
Synthesis of carbon disulfide modified chitosan resin and its adsorption
properties for palladium(Ⅱ) in wastewater. Chem. Eng. J. 466, 143082.
https://doi.org/10.1016/j.cej.2023.143082
Chen, Q., Zheng, J., Wen, L., Yang, C., Zhang, L., 2019. A multi-functional-group
modified cellulose for enhanced heavy metal cadmium adsorption:
Performance and quantum chemical mechanism. Chemosphere 224, 509–518.
https://doi.org/10.1016/j.chemosphere.2019.02.138
Dong, W., Lu, Y., Wang, W., Zhang, M., Jing, Y., Wang, A., 2020. A sustainable
approach to fabricate new 1D and 2D nanomaterials from natural abundant
palygorskite clay for antibacterial and adsorption. Chem. Eng. J. 382, 122984.
https://doi.org/10.1016/j.cej.2019.122984
Egbosiuba, T.C., Egwunyenga, M.C., Tijani, J.O., Mustapha, S., Abdulkareem,
A.S., Kovo, A.S., Krikstolaityte, V., Veksha, A., Wagner, M., Lisak, G., 2022.
Activated multi-walled carbon nanotubes decorated with zero valent nickel
nanoparticles for arsenic, cadmium and lead adsorption from wastewater in a
batch and continuous flow modes. J. Hazard. Mater. 423, 126993.
https://doi.org/10.1016/j.jhazmat.2021.126993
Fakayode, O.A., Wang, Z., Wahia, H., Mustapha, A.T., Zhou, C., Ma, H., 2021.
Higher heating value, exergy, pyrolysis kinetics and thermodynamic analysis
of ultrasound-assisted deep eutectic solvent pretreated watermelon rind
biomass. Bioresour. Technol. 332, 125040.
https://doi.org/10.1016/j.biortech.2021.125040
Gao, Z., Shan, D., He, J., Huang, T., Mao, Y., Tan, H., Shi, H., Li, T., Xie, T.,
2023. Effects and mechanism on cadmium adsorption removal by CaCl2-
modified biochar from selenium-rich straw. Bioresour. Technol. 370, 128563.
https://doi.org/10.1016/j.biortech.2022.128563
Ho, Y.S., McKay, G., 1999. Pseudo-second order model for sorption processes.
Process Biochem. 34, 451–465. https://doi.org/10.1016/S0032-
9592(98)00112-5
Ho, Y.S., McKay, G., 1998. Sorption of dye from aqueous solution by peat. Chem.
Eng. J. 70, 115–124. https://doi.org/10.1016/S1385-8947(98)00076-X
Huang, P., Wei, X., Wang, X., Gu, Z., Guo, Y., Zhao, C., 2023. Facile fabrication
of silica aerogel supported amine adsorbent pellets for Low-concentration CO2
removal from confined spaces. Chem. Eng. J. 468, 143629.
https://doi.org/10.1016/j.cej.2023.143629
Huang, X., Zemlyanov, D.Y., Diaz-Amaya, S., Salehi, M., Stanciu, L., Whelton,
A.J., 2020. Competitive heavy metal adsorption onto new and aged
polyethylene under various drinking water conditions. J. Hazard. Mater. 385,
121585. https://doi.org/10.1016/j.jhazmat.2019.121585
Jyoti, Singh, S., Das, S., Srivastava, S., 2020. Comparative study for removal of
toxic hexavalent chromium by zinc oxide nanoparticles, chitosan, chitin and
zinc-chitosan nano-biocomposite. Environ. Technol. Innov. 109231.
https://doi.org/10.1016/j.eti.2023.103310
Li, H., Xiong, J., Xiao, T., Long, J., Wang, Q., Li, K., Liu, X., Zhang, G., Zhang,
H., 2019. Biochar derived from watermelon rinds as regenerable adsorbent for
efficient removal of thallium(I) from wastewater. Process Saf. Environ. Prot.
127, 257–266. https://doi.org/10.1016/j.psep.2019.04.031
Li, H., Yuan, Z., Ding, S., Yuan, J., 2023. Adsorption of lead ions by magnetic
carbon: Comparison of magnetic carbon properties and modification methods.
J. Environ. Chem. Eng. 11, 110136. https://doi.org/10.1016/j.jece.2023.110136
Liu, Y., Cai, L., Wang, X., Chen, Z., Yang, W., 2023. Efficient adsorption of
arsenic in groundwater by hydrated iron oxide and ferromanganese oxide
chitosan gel beads. Sep. Purif. Technol. 315, 123692.
https://doi.org/10.1016/j.seppur.2023.123692
Ma, F., Zhao, H., Zheng, X., Zhao, B., Diao, J., Jiang, Y., 2023. Enhanced
adsorption of cadmium from aqueous solution by amino modification biochar
and its adsorption mechanism insight. J. Environ. Chem. Eng. 11, 109747.
https://doi.org/10.1016/j.jece.2023.109747
Ma, P., Yao, S., Wang, Z., Qi, F., Liu, X., 2023. Preparation of nitrogen-doped
hierarchical porous carbon aerogels from agricultural wastes for efficient
pollution adsorption. Sep. Purif. Technol. 311, 123250.
https://doi.org/10.1016/j.seppur.2023.123250
Medeiros, D.C.C. da S., Chelme-Ayala, P., Gamal El-Din, M., 2023. Sludge-based
activated biochar for adsorption treatment of real oil sands process water:
Selectivity of naphthenic acids, reusability of spent biochar, leaching potential,
and acute toxicity removal. Chem. Eng. J. 463, 142329.
https://doi.org/10.1016/j.cej.2023.142329
Miao, J., Zhao, X., Zhang, Y.X., Lei, Z.L., Liu, Z.H., 2021. Preparation of hollow
hierarchical porous CoMgAl-borate LDH ball-flower and its calcinated
product with extraordinary adsorption capacity for Congo red and methyl
orange. Appl. Clay Sci. 207, 106093.
https://doi.org/10.1016/j.clay.2021.106093
Muhammad, N., Ge, L., Chan, W.P., Khan, A., Nafees, M., Lisak, G., 2022.
Impacts of pyrolysis temperatures on physicochemical and structural
properties of green waste derived biochars for adsorption of potentially toxic
elements. J. Environ. Manage. 317, 115385.
https://doi.org/10.1016/j.jenvman.2022.115385
Pawar, R.R., Lalhmunsiama, Kim, M., Kim, J.G., Hong, S.M., Sawant, S.Y., Lee,
S.M., 2018. Efficient removal of hazardous lead, cadmium, and arsenic from
aqueous environment by iron oxide modified clay-activated carbon composite
beads. Appl. Clay Sci. 162, 339–350.
https://doi.org/10.1016/j.clay.2018.06.014
Qi, G., Pan, Z., Zhang, X., Miao, X., Xiang, W., Gao, B., 2022. Effect of ball
milling with hydrogen peroxide or ammonia hydroxide on sorption
performance of volatile organic compounds by biochar from different
pyrolysis temperatures. Chem. Eng. J. 450, 138027.
https://doi.org/10.1016/j.cej.2022.138027
Quang, H.H.P., Phan, K.T., Ta, P.D.L., Dinh, N.T., Alomar, T.S., AlMasoud, N.,
Huang, C.W., Chauhan, A., Nguyen, V.H., 2022. Nitrate removal from
aqueous solution using watermelon rind derived biochar-supported ZrO2
nanomaterial: Synthesis, characterization, and mechanism. Arab. J. Chem. 15,
104106. https://doi.org/10.1016/j.arabjc.2022.104106
Rambabu, K., Bharath, G., Hai, A., Luo, S., Liao, K., Haija, M.A., Banat, F.,
Naushad, M., 2020. Development of watermelon rind derived activated
carbon/manganese ferrite nanocomposite for cleaner desalination by capacitive
deionization. J. Clean. Prod. 272, 122626.
https://doi.org/10.1016/j.jclepro.2020.122626
Sadhu, M., Bhattacharya, P., Vithanage, M., Padmaja Sudhakar, P., 2022.
Adsorptive removal of fluoride using biochar – A potential application in
drinking water treatment. Sep. Purif. Technol. 278, 119106.
https://doi.org/10.1016/j.seppur.2021.119106
Sulistiyo, C.D., Cheng, K.C., Su’andi, H.J., Yuliana, M., Hsieh, C.W., Ismadji, S.,
Angkawijaya, A.E., Go, A.W., Hsu, H.Y., Tran-Nguyen, P.L., Santoso, S.P.,
2022. Removal of hexavalent chromium using durian in the form of rind,
cellulose, and activated carbon: Comparison on adsorption performance and
economic evaluation. J. Clean. Prod. 380, 135010.
https://doi.org/10.1016/j.jclepro.2022.135010
Tang, D., Li, J., Yang, Z., Jiang, X., Huang, L., Guo, X., Li, Y., Zhu, J., Sun, X.,
2022. Fabrication and mechanism exploration of oxygen-incorporated 1T-
MoS2 with high adsorption performance on methylene blue. Chem. Eng. J.
428, 130954. https://doi.org/10.1016/j.cej.2021.130954
Wan, Y., Luo, H., Cai, Y., Dang, Z., Yin, H., 2023. Selective removal of total Cr
from a complex water matrix by chitosan and biochar modified-FeS: Kinetics
and underlying mechanisms. J. Hazard. Mater. 454, 131475.
https://doi.org/10.1016/j.jhazmat.2023.131475
Wang, B., Dai, X., Ma, Y., Xin, J., Pan, C., Zhai, Y., 2023. Nitrogen-rich porous
biochar for highly efficient adsorption of perchlorate: Influencing factors and
mechanism. J. Environ. Chem. Eng. 11, 110293.
https://doi.org/10.1016/j.jece.2023.110293
Wang, H., Wang, W., Zhou, S., Gao, X., 2023. Adsorption mechanism of Cr(VI)
on woody-activated carbons. Heliyon 9, e13267.
https://doi.org/10.1016/j.heliyon.2023.e13267
Wang, Z., Han, L., Sun, K., Jin, J., Ro, K.S., Libra, J.A., Liu, X., Xing, B., 2016.
Sorption of four hydrophobic organic contaminants by biochars derived from
maize straw, wood dust and swine manure at different pyrolytic temperatures.
Chemosphere 144, 285–291.
https://doi.org/10.1016/j.chemosphere.2015.08.042
Xiang, W., Zhang, X., Luo, J., Li, Y., Guo, T., Gao, B., 2022. Performance of
lignin impregnated biochar on tetracycline hydrochloride adsorption:
Governing factors and mechanisms. Environ. Res. 215, 114339.
https://doi.org/10.1016/j.envres.2022.114339
Xue, H., Wang, X., Xu, Q., Dhaouadi, F., Sellaoui, L., Seliem, M.K., Ben Lamine,
A., Belmabrouk, H., Bajahzar, A., Bonilla-Petriciolet, A., Li, Z., Li, Q., 2022.
Adsorption of methylene blue from aqueous solution on activated carbons and
composite prepared from an agricultural waste biomass: A comparative study
by experimental and advanced modeling analysis. Chem. Eng. J. 430.
https://doi.org/10.1016/j.cej.2021.132801
Yang, L., Jin, X., Lin, Q., Owens, G., Chen, Z., 2023. Enhanced adsorption and
reduction of Pb(II) and Zn(II) from mining wastewater by carbon@nano-zero-
valent iron (C@nZVI) derived from biosynthesis. Sep. Purif. Technol. 311.
https://doi.org/10.1016/j.seppur.2023.123249
Yuan, Q., Wang, P., Wang, X., Hu, B., Wang, C., Xing, X., 2023. Nano-
chlorapatite modification enhancing cadmium(II) adsorption capacity of crop
residue biochars. Sci. Total Environ. 865, 161097.
https://doi.org/10.1016/j.scitotenv.2022.161097

You might also like