1
Introduction
Water pollution is a significant environmental issue that is faced worldwide.[1] The world is experiencing a clean water crisis due to increased industries, the
citification of rural areas, and globalization.[2] On the other hand, as the world’s population increases rapidly, so do these necessities
of clean water for human existence.[3] According to the latest United Nations prediction, the world population has reached
8.12 billion as of July 1, 2024, and will rise to 9 billion in 2037 and 10 billion
in 2058.[4] Water scarcity affected 933 million urban people in 2016.[5] Around 1.693–2.373 billion people will live in water-scarce regions by 2050, which
is a great number.[5] Water purification is essential for getting clean drinking water and for pharmaceutical,
industrial, and medical, chemical purposes.[6] Wastewater contains a variety of wastes, including inorganic pollutants like heavy
metals (Pb, Hg, Cd, Cu, etc.), nitrates, phosphates, and colloidal pollutant particles,
as well as organic pollutants including dyes, pesticides, herbicides, human faces,
oil, animal wastes, drugs, and phenols. Dye is one of the significant pollutants affecting
human health, as it is extremely toxic and carcinogenic. About 107 kilograms of dyes are utilized by textile industries with total synthetic dye manufacturing
of 7 × 1010 worldwide annually.[7] Approximately 1.8 × 108 kilograms of dye are released into wastewater annually due to textile dyeing.[8] Besides textiles, dyes are used in paper printing, printing, tanning, leather, and
food-coloring industries. So these industries also expel a considerable amount of
dye into water. This persistent throwing of dye wastewater into the environment without
proper treatment has hazardous consequences.[9] The most immediate effect of dye wastewater is the disturbance in the aquatic ecosystems.
It may kill aquatic plants, fishes, and other wildlife, reducing biodiversity. Chemicals
and heavy metals in dyes can toxify the food chain causing severe contamination in
potable water. Congo red (CR) (C32H22N6Na2O6S2) is a synthetic azo anionic dye and a pH indicator. It has a brownish-red appearance.
It is widely used in the textile industry, paper industry, biological and medical
applications, etc. CR dye is potentially considered to be carcinogenic under certain
conditions, particularly due to its breakdown products. Its presence in wastewater
can harm aquatic ecosystems.
Dyes are very stable and synthesized in a way to combat the degradation, making their
repair a major challenge for scientists.[10] Furthermore, nowadays, most of the dyestuff industries make dyes with high levels
of fastness, so that the stability and resilience of dyes improve. Dyes with high
fastness have high resistance to fade or bleed their color under certain conditions
like light, water, rubbing, washing, and perspiration. While the high color fastness
of dyes is good for aesthetic and functional purposes, it makes it difficult to treat
these dyes from wastewater. There are many methods for the treatment of dyes from
wastewater like membrane filtration, biological and chemical oxidation, electrolysis,
photocatalysis, adsorption, ion exchange, coagulation/flocculation, and catalysis.[11]
[12]
[13]
[14]
[15]
[16] The adsorption technique has been proven superior to other techniques owing to its
low cost, simplicity, flexible design, lack of harmful by-products, effectiveness
even in the presence of toxic pollutants, ease of operation, and eco-friendly approach.[17] Other techniques have limitations such as pore occlusion and membrane clogging as
in membrane filtration; production of intrinsic sludge and waste-management issues
in coagulation/flocculation as in chemical methods; constraints like spatial demands,
poor removal efficiency, and ineffectiveness in addressing stubborn dye components
as in biological methods; plant’s pollutant tolerance capacity and the need for extensive
land for establishing treatment facilities as in the phytoremediation technique of
dye removal; use of costly organic solvents as in ion exchange methods; and high energy
expenditures and the generation of secondary products as in the oxidation process
of dye removal, etc.[18]
[19] Adsorption exhibits significant efficacy in adsorbing dye molecules onto solid surfaces.
Additionally, the diverse selection of adsorbents tailored for different water contaminants
enhances its versatility. The significant advantage of adsorption is that the adsorbent
can be recycled and further used. Simultaneously, through regeneration, we can minimize
secondary pollution and also reuse it in alternative applications, which promotes
cost-effectiveness and resource restoration.[20] Various adsorbent surfaces are modified by effective chemical and physical methods,
which increase the affinity of adsorbate molecules toward adsorbents and increase
the surface area of adsorbents. This in turn increases the adsorption capacity of
adsorbents.[21] Therefore, adsorption is widely used for wastewater treatment.
Activated carbon is a commonly utilized adsorbent, but its effectiveness is often
constrained by challenges such as high production costs, low selectivity, environmental
concerns, and complications related to regeneration and reuse. Recently, activated
carbon has been widely substituted by biochar due to cost-effectiveness, sustainability,
low energy requirements, adequate adsorption for specific applications, versatility
in usage, and reduced environmental footprint for adsorption purposes. One study showed
that the q
max value for methylene blue dye removal was greater for biochar derived from rice straw
than its activated carbon.[22] Biochar is a carbon-rich material produced by the pyrolysis of organic biomass (such
as agricultural residues, forestry waste, or animal manure) under limited or no oxygen
conditions.
In recent times, the incorporation of ferrite spinels onto biochar is an advanced
technique that enhances the adsorption efficiency of biochar in wastewater treatment,
i.e., wheat straw biochar–zirconium ferrite nanocomposite (BC–ZrFe2O5 NCs) showed higher adsorption capacity for tartrazine dye than the pristine wheat
straw biochar.[23]
[24] Ferrite spinels (MFe2O4, where M = divalent atoms like Cu, Zn, Ni, Co, Mg, Mn) have a face-centered cubic
(FCC) structure. They crystallize in either normal or inverse forms, with M2+ ions in the tetrahedral sites and Fe3+ ions in the octahedral sites in the normal form, and vice versa in the inverse form.
They have unique properties like high magnetism, chemical stability, surface active
sites, high surface area, mechanical strength, electrical resistivity, and ease of
modification. Therefore, they are extensively used in magnetic materials, electronics,
catalytic reactions, energy storage, photocatalysis, and wastewater treatment. Recently,
biochar–ferrite spinel composites have gained attention as a powerful adsorbent in
wastewater treatment due to their excellent combination of desirable properties. Biochar
assists in enhancing the durability and stability of spinels, while spinels increase
the mechanical and chemical stability of biochar thereby resulting in the improvement
of the adsorption performance of the adsorbent. Ferrite spinels enhance adsorption
mechanisms through ion exchange, surface complexation, hydrogen bonding, and oxidation–reduction,
owing to the presence of their metal ions and -OH groups. Their magnetic properties
simplify the recovery process of used adsorbent after multiple cycles.
S. monoica is a succulent plant of the family Chenopodium growing wild in coastal areas, salt marshes, and saline desert regions of the Middle
East and South Asia, Africa, the Arabian Peninsula, and Central Asia. It is widely
distributed in the coastal areas of India. It is rich in salt. It is a shrubby or
herbaceous plant, typically growing up to 1 m in height. Its leaves are fleshy and
succulent and are adapted to conserve water. As a halophyte, S. monoica is adapted to high salinity conditions, making it capable of growing where most other
plants cannot survive. Some parts of the plants are believed to have medicinal properties
including anti-inflammatory and antibacterial. Many research studies have been done
on its medicinal properties. Some studies have also been done on its role in soil
stabilization and land reclamation, especially in saline areas. Seemingly, no studies
have been done on the use of S. monoica in water remediation.
An artificial neural network (ANN) is a group of simple algorithms that are connected
to one another and assess the data in response to external input. In a few ways, it
is based on biological neural networks. It models complex nonlinear relationships
between input and output variables. It assesses the performance using coefficient
of determination (R
2) and mean squared error (MSE), where higher R
2 and lower error values indicate better prediction accuracy.[25] In adsorption studies, ANNs are trained using experimental input–output data to
model the relationship between process variables and adsorption capacity/removal percentage.
By minimizing the difference between actual and predicted values, ANNs effectively
forecast adsorption behavior without the need for explicit mechanistic equations.[26]
In the present work, we have synthesized biochar (BC300) from the S. monoica leaves powder (LP) followed by its pretreatment to yield base-treated biochar (BTBC).
A ferrite composite of biochar (FCOB) was synthesized by incorporation of ferrite
spinel NiCuZnFe2O4 into BTBC. FCOB was tested as an adsorbent for the removal of CR as a model anionic
dye. ANN modeling was performed to analyze and predict the adsorption behavior. It
is important to highlight that this is the first research study on the application
of modified S. monoica biochar in toxic CR dye removal. So the synthesized adsorbent in the present work
is novel and unique, representing a significant contribution to the field of adsorbent
materials. The physicochemical characteristics of FCOB were investigated using powder
X-ray diffraction (XRD), Fourier-Transform Infrared (FTIR) Spectroscopy, Scanning
Electron Microscopy (SEM) with Energy Dispersive Spectroscopy (EDS), Atomic Force
Microscopy (AFM), Zeta potential and particle size analysis by Dynamic Light Scattering
(DLS), Thermogravimetry (TG) analysis, and the Brunauer–Emmett–Teller (BET) surface
analyzer.
2
Results and Discussion
2.1
Characterizations of Samples
2.1.1
Scanning Electron Microscopy (SEM)
[Fig. 1a,b] represent SEM images of BTBC with higher and lower magnification, respectively.
[Fig. 1c,d] represent SEM images of FCOB with higher and lower magnification, respectively.
It is apparent from SEM images of BTBC that their surface was not flat and smooth,
but it was filled with small cavities and rough.[27] FCOB has a smooth appearance. This was because of the incorporation of ferrite spinel
particles into the small cavities of BTBC after its modification. The SEM image of
FC-CR is shown in Fig. S1. From Fig. S1, it is evident that there is an agglomeration
of some small spherical and irregularly shaped particles after CR adsorption onto
FCOB. There is a strong deposition of large molecules of CR dye on FCOB through strong
interaction, which may yield an agglomerated structure. Thus, before adsorption, FCOB
had a smoother surface, and after adsorption, FC-CR has a rougher surface.[28]
Fig. 1 SEM images of BTBC (a, b) at higher (×5000), lower (×350) magnification, SEM images of FCOB (c, d) at higher (×1000), lower (×350) magnification (FEI Nova NanoSEM 450, 20 kV, det
= LVD)
The elemental compositions of FCOB and FC-CR were confirmed by EDS analysis ([Table 1]) of FCOB and FC-CR. Their EDS spectra are given in Fig. S2. The presence of Fe,
Ni, Cu, and Zn confirms the incorporation of ferrite spinel (NiCuZnFe2O4) in BTBC. FCOB shows other minerals like Na, Mg, Si, P, and Ca because halophyte
S. monoica accumulates these minerals from the soil.[29] The concentrations of O, Ca, and Fe were decreased, and Cu was diminished from the
surface of FCOB after adsorption, which is an indication of their contribution to
the adsorption process of CR. The decrease in the percentage of Ca may be due to the
cation exchange interaction between Ca ions and CR dye, which may also be the cause
of the diminishment of Na, Mg, Al, and Si on FCOB after adsorption.[30] The greater C concentration of FC-CR was due to the binding of the CR dye organic
framework on FCOB. The displacement of N-containing functional groups from FCOB by
CR dye due to strong interaction led to the diminishment of N on the FC-CR surface.
Table 1
EDS elemental composition (%wt) before (FCOB) and after (FC-CR) CR adsorption
|
Percentage (%)
|
|
Elements
|
Before adsorption (FCOB)
|
After adsorption (FC-CR)
|
|
C
|
59.50
|
76.9
|
|
N
|
10.96
|
–
|
|
O
|
23.98
|
21.8
|
|
Na
|
0.27
|
–
|
|
Mg
|
0.46
|
–
|
|
Al
|
0.06
|
–
|
|
Si
|
0.07
|
–
|
|
P
|
0.14
|
–
|
|
Ca
|
0.77
|
0.4
|
|
Fe
|
2.32
|
1.0
|
|
Ni
|
0.43
|
–
|
|
Cu
|
0.60
|
0.0
|
|
Zn
|
0.44
|
–
|
2.1.2
Thermal Stability Investigation-TGA
The TGA curves of BC300, BTBC, and FCOB are shown in [Fig. 2a]. As shown in the TGA curve of BC300, in the first stage of thermal decomposition,
BC300 shows around 4.08% mass loss from 24 to 184 °C, attributed to the evaporation
of water from it.[31] In the second stage, from 185 to 400 °C, mass loss of 4.10% was observed. It corresponds
to the thermal degradation of hemicellulose and cellulose. In the third stage, from
401 to 691 °C, a mass loss of 29.21% was observed. This is attributed to lignin decomposition.[32] At higher temperatures, aromatic carbon in BC300 begins to oxidize releasing CO
and CO2, which leads to mass loss. Residual volatiles are also released at higher temperatures.
Some of the undecomposed minerals, organic complexes (e.g., residual cellulose, hemicellulose,
lignin, and bioactive compounds) also degrade at the last stage.
Fig. 2 (a) TGA curves of BC300, BTBC, and FCOB (SDT Q600, RT – 700 °C, 5 °C/min, N₂ flow
150 mL/min) and (b) FT-IR spectra plot of FCOB and FC-CR (Perkin Elmer IR spectrometer,
KBr pellet, 400–4000 cm−1, resolution 4 cm−1, 32 scans).
As shown in the TGA curve of BTBC, in the first stage (from 27 to 150 °C), 13.38%
mass loss is attributed to moisture loss, which is greater than that of BC300. This
is because BTBC is an alkaline form of BC300. The alkaline treatment introduces hydrophilic
functional groups (–OH, –C=O, –COOH, and –OC=O) onto BTBC, resulting in a higher number
of such groups compared to BC300. Alkaline treatment may remove a portion of cellulose,
hemicellulose, and lignin but also leave behind some intermediate organic compounds
that are still volatile and prone to degradation at moderate temperatures from 151
to 400 °C causing mass loss of 13.38%. In the third stage, from 401 to 691 °C, a high
mass loss of 66.43% occurs. This mass loss during this last stage is attributed to
the degradation of lignin and residual organic complexes, alkali and other salts,
and residual hydrophilic groups.
As shown in the TGA curve of FCOB, in the first stage, from 21 to 150 °C, 12.21% of
mass loss occurs due to moisture degradation. Here, in the first stage, the mass loss
is slightly lower than that of BTBC because the incorporation of ferrite particles
in the BTBC may have blocked its pores inhibiting its ability to retain water through
capillary forces. This results in a lower overall water-retention capacity. The second
stage, which occurs between 151 and 400 °C, has a mass loss of around 14.68%. Degradation
of residual organic intermediate compounds (hemicellulose, cellulose, bioactive compounds)
may occur in this stage. Ferrite spinel (NiCuZnFe2O4) contributes minimally to the mass loss as they are thermally stable and do not decompose
at these temperatures but may influence the degradation of organic compounds due to
their catalytic activity. The third and last stage (from 401 to 691 °C) has a mass
loss of around 47.24%. This stage is attributed to the decomposition of lignin, ferrite
spinel, and residual components (like organic complexes, moisture, salts, and hydrophilic
groups), which were undecomposed during lower temperature ranges. At these elevated
temperatures, ferrite spinels also oxidize the carbon matrix of biochar into CO and
CO2, yielding mass loss.
2.1.3
Fourier Transform Infrared Spectroscopy (FT-IR)
[Fig. 2b] shows the FTIR spectra plot of FCOB and FC-CR. The FT-IR spectra of LP, BC300, and
BTBC are shown in Fig. S3. The FT-IR spectra of BC300, BTBC, and FCOB closely resemble
that of the LP, indicating the preservation of inherent properties of S. monoica leaves.
LP shows a broad peak at 3455 cm−1 because of stretching vibrations of the OH group, which justifies the presence of
lipids in the leaves. In LP, the peak at 2346 cm−1 is due to atmospheric CO2 interference.[33] While the same peak in BC300 may also be due to residual CO2 from the pyrolysis process. The peak at 1641 cm−1 is due to amides containing C=O bond, N-H bending, C=O stretching, or aromatic/conjugated
nonaromatic C=C stretching vibration. The earlier studies show that this peak identifies
the presence of proteins in the leaves. The peak at 1334 cm−1 corresponds to SO2 stretching, O–H bending (carboxylic acids, alcohols), or N=O stretching (nitro compounds),
which identifies the presence of carbohydrates in the leaves. The peak at 1396 cm−1 is due to COO− symmetric stretching in carboxylate groups from organic acids or their salts, which
indicates the presence of fatty acids or amino acids in leaves.[34] The peak at 1110 cm−1 is attributed to C–N stretching (aliphatic amines), C–O stretching (ethers, alcohols),
or O–H bending (carboxylic acids), and the peak at 780 cm−1 corresponds to C–N stretching (amines), =C–H bending (benzene) or C–Br stretching
vibration.[35] Earlier studies indicate that these two peaks are associated with cell wall components
like cellulose, hemicellulose, pectin, lignin, and proteins. Only one peak at 1396
cm−1 is absent in BC300 because COO− groups are removed or transformed to CO2 and CO gases during the pyrolysis process to form biochar. But all other peaks are
similar in LP as well as BC300 because BC300 was prepared at a relatively low temperature
of 300 °C, which might preserve most of the functional groups found in LP in it. Thus,
both LP and BC300 exhibit similar characteristic peaks in the IR spectrum.
BTBC exhibits all the peaks similar to those of LP, except for the one at 1396 cm−1, which is identical to that of BC300. BTBC lacks a peak at 1396 cm−1 (COO− stretching) because base treatment can also convert carboxyl groups to other forms,
such as phenolic hydroxyls or other oxygen-containing groups.
In FCOB, a very broad peak at 3270 cm−1 is due to OH stretching vibration. This peak has a frequency shift to a lower wavenumber
compared to LP, BC300, and BTBC. Hydrogen bonding occurs between the hydroxyl group
and the NiCuZnFe2O4. It causes a shift in the O–H stretching frequency toward lower wavenumbers (e.g.,
from 3455 to 3270 cm−1). The interaction of hydroxyl groups with Fe/Ni/Cu/Zn ions in ferrite could also
involve a change in the electron density around the O–H bond, leading to the downshift
in the O–H stretching vibration. The peak at 1594 cm−1 is due to the C=C stretching vibration (aromatic and nonaromatic)/ C=O bond of amide/C=O
stretching or N–H bending vibration. This peak is shifted to a lower wavenumber in
FCOB than LP, BC300, and BTBC (1641 cm−1) because the functional groups on the biochar (such as C=O, N–H, and C=C) can coordinate
or interact with the metal ions of Fe/Ni/Cu/Zn from the ferrite. These interactions
often lead to a shift in the vibrational frequencies of the functional groups. Specifically,
the C=O, N–H bending, and C=C stretch vibrations may undergo a redshift (lower wavenumber)
due to changes in the electron density and bonding characteristics resulting from
metal–ligand interactions. The frequency shift is due to biochar–ferrite spinel interaction.
The interaction with the ferrite can also affect the C=C stretching vibration of aromatic
and nonaromatic structures. If the ferrite composite induces changes in the conjugation
or electronic structure of the aromatic system, this can lead to a shift in the stretching
vibration of C=C bonds. These two peak shifts further confirm the successful incorporation
of ferrite spinel onto BTBC. The peak at 1396 cm−1 contributes to the symmetric stretching vibration of the COO− group. The peak at 1110 cm−1 is likely due to the C–O stretching/C–N stretching or O–H bending vibration. This
peak may also reflect the interaction between the composite (NiCuZnFe₂O₄) and the
biochar surface.
In FC-CR, the broad peak at 3455 cm−1 is due to OH stretching vibration. The peak at 1641cm−1 indicates the stretching vibration of aromatic ring C=C or nonaromatic C=C/N–H bending/C=O
of amide/C=O stretching vibration. This peak is more prominent in FC-CR than in LP,
BC300, BTBC, and FCOB. When CR dye adsorbs onto the biochar surface, this peak can
become more prominent due to the interaction between the dye and the biochar, particularly
at sites with aromatic or oxygenated functional groups on the biochar surface. The
peak at 1396 cm−1 in FC-CR is due to C–N stretching vibration or the symmetric bending vibration of
C–H bonds in the aromatic structure of the CR dye, which confirms the interaction
of CR dye with FCOB through the adsorption phenomenon. This peak may also be due to
COO- stretching vibration due to carboxylic acid groups present on FCOB. The peak
at 1110 cm−1 is may be due to the sulfonic acid group (–SO₃−) of CR dye because, during adsorption, these groups may interact with the ferrite
spinel or the biochar’s surface, leading to vibrations characteristic of sulfonic
groups in this range (1000–1200 cm−1).[36] This peak may also be due to the C–O stretching/C–N stretching or O-H bending vibration,
which may also justify a successful adsorption of CR onto FCOB.
2.1.4
XRD
The crystalline substance exhibits strong, sharp, and well-defined peaks, while the
amorphous substance exhibits broad hump-like peaks in the XRD spectrum.[37] The XRD spectrums of BC300, BTBC, and FCOB are shown in [Fig. 3]. In a manner, BC300 contains the crystalline inorganic phase-NaCl ([Fig. 3]). The S. monoica is a halophytic plant that stores high amounts of salt (NaCl) in its tissues. As
the plant in the present study was collected from the saline coastal region of Kandla,
NaCl was anticipated in BC300. The peaks at 27.36°, 31.67°, 45.51°, 56.54°, 66.23°,
and 75.24° (2θ) correspond to crystal planes of NaCl (1 1 1), (2 0 0), (2 2 0), (2
2 2), (4 0 0), and (4 2 0), respectively.[38] NaCl has an FCC crystal structure with an Fm–3m space group. There is an absence of a broad amorphous carbon peak in BC300 because
biomass S. monoica has high concentrations of salt (as mentioned above), which may mask the amorphous
carbon halo. The base treatment of BC300 dissolves the inorganic crystalline minerals
and these minerals are leached out, reducing or eliminating sharp crystalline peaks
in the XRD spectrum (XRD spectrum of BTBC). Moreover, NaOH activates the carbon structure
of BC300 by creating pores, which in turn increases disorder leading to an increase
in amorphous carbon structure. There is a broad peak between 2θ in the BTBC XRD spectrum. This indicates the presence of an amorphous carbon structure
with randomly distributed aromatic sheets.[39] The peak around represents the (002) plane of graphitic carbon. This peak is due to the removal of
impurities by NaOH, which sharpens the graphitic carbon peak. As seen from [Fig. 3], in FCOB, a broad peak is there between 2θ which again indicates the presence of amorphous carbon. The XRD pattern of FCOB clearly
shows the two most intense characteristic peaks of ferrite spinel NiCuZnFe2O4 approximately at 2θ ≈ 35.5° and 62.9°, corresponding to (3 1 1) and (4 4 0) crystal
planes, respectively.[40] This further demonstrates that NiCuZnFe2O4 was successfully incorporated onto BTBC to yield FCOB. As FCOB has a mole ratio of
BTBC: NiCuZnFe2O4 2:1, the absence of additional distinctive peaks of NiCuZnFe2O4 in FCOB may be explained by its low concentration in comparison to BTBC. The crystalline
size of the materials was calculated using Scherrer’s [Eq. (1)].
Fig. 3 XRD plots of BC300, BTBC, and FCOB(PANalytical X’Pert Pro, Cu Kα λ = 0.154 nm, 2θ
= 5–80°, 40 mA, 40 kV).
where D, λ, β, and θ represent crystalline size (nm), wavelength (Cu-Kα = 0.15406 nm), full-width at half maxima, and Bragg’s angle (2θ/2), respectively. The
average crystalline sizes of BC300, BTBC, and FCOB were 17, 3, and 2 nm, respectively.
The crystallographic information is given in Section S1 of the supplementary file.
2.1.5
BET Surface Area Analysis
A Brunauer–Emmett–Teller (BET) graph, also known as a BET plot, is used to measure
the surface area of porous solid materials. The nitrogen adsorption–desorption isotherm
and BJH adsorption cumulative pore volume curve of FCOB are given in [Fig. 4a,b], respectively. A hysteresis loop in adsorption–desorption isotherms refers to the
phenomenon where the amount of gas adsorbed during adsorption differs from the amount
desorbed at the same relative pressure (P/P
0). This loop appears as a mismatch between the adsorption and desorption branches
of the isotherm and is characteristic of certain types of porous materials. [Fig. 4a] reveals that FCOB has type IV(a) isotherm with hysteresis type H3. In type IV(a) isotherm, capillary condensation is accompanied by hysteresis. This
isotherm exhibits mesoporous solids in which capillary condensation takes place at
higher pressures of adsorbate in addition to multilayer adsorption at lower pressures.[41] The surface area, pore size, and pore volume of FCOB were 44.64 m2 g−1 ± 0.2396 m2/g, 7.04 nm, and 0.078 cm3 g−1, respectively. The pore size between 2 nm to 50 nm suggests the mesoporous nature
of the FCOB adsorbent. FCOB has a 44.64 m2 g−1 surface area, which is in a moderate range. The BET surface area value was calculated
following the recommended procedures and avoiding the common pitfalls outlined in
the guidelines reported by Johannes W. M. Osterrieth et al. (2022), including appropriate
selection of the relative pressure range, verification of the linearity of the BET
plot, and adherence to the Rouquerol criteria.[42] The BET surface area was calculated from the linear region of the BET plot in the
relative pressure range of 0.05–0.30, following the Rouquerol criteria. The measurement
of BET surface area showed high precision, with a standard deviation of only 0.2396,
confirming the accuracy and reliability of the obtained BET surface area values.
Fig. 4 (a) BET isotherm linear plot (b) BJH adsorption dP/dV pore volume a,b-Micromeritics 3Flex 3500, degassed at 400 °C, 5 h, 77.29 K), and (c) Zeta potential curve (HORIBA SZ-100, 0.01 wt% in distilled water, pH adjusted to
2–10, 25 °C).
2.1.6
Zeta(ζ) Potential and Particle Size Analysis
The Zeta potential of biochar is the measure of the electrostatic potential at the
boundary between the charged particles of biochar and the surrounding bulk solution.
It is the result of a build-up of charges near the biochar–bulk solution interface
leading to the formation of an electric double layer. pH is an important factor that
affects the zeta potential value. The pH at which zeta (ζ) potential value is zero
is called as point zero charge (pHpzc)/isoelectronic point. The curve of the zeta
(ζ) potential values of FCOB against pH is given in [Fig. 4c]. The curve depicts that FCOB has a negative surface charge in both acidic and alkaline
conditions. FCOB showed the least negative value of zeta potential (−1.9 mV) at pH
2. Notably, the highest adsorption of the anionic dye CR also occurred at this pH
(Section 2.2.1). It can be attributed to the reduced electrostatic repulsion between
the anionic CR molecules and the negatively charged FCOB adsorbent surface due to
increased H+ ion concentrations at pH 2.
2.1.7
Atomic Force Microscopy (AFM)
While SEM provides detailed information on the overall surface morphology, AFM was
further employed to quantify surface roughness at the nanoscale. The surface roughness
parameters were calculated using Gwyddion (64-bit) software. The surface roughness
parameters like mean roughness (Sa), RMS roughness (Sq), and skew (Ssk) of LP, BC300,
BTBC, FCOB, and FC-CR are given in Table S1. The roughness parameter values of FCOB
are lower than those of BTBC. This is because of the fact that ferrite spinel particles
get embedded into cavities of BTBC, making the structure of FCOB smoother. The SEM
imaging (Section 2.1.1) also supports this observation, as BTBC exhibits a channelled
morphology with embedded cavities, whereas FCOB is smoother. The AFM images (2D and
3D) of LP, BC300, BTBC, FCOB, and FC-CR are given in [Fig. 5a–e], respectively.
Fig. 5 2D and 3D images of (a) LP, (b) BC300, (c) BTBC, (d) FCOB, and (e) FC-CR (Bruker Dimension Icon, tapping mode, scan area 10 × 10 μm).
2.2
Effect of Different Parameters on the Adsorption Performance of FCOB
2.2.1
Effect of pH
pH is a key parameter in the adsorption process as it changes the structures of dye
(CR) and also elevates the surface charge density of the adsorbent material (FCOB).
As shown in [Fig. 6a], the adsorption of CR (C
0 = 50 mg L−1; m = 0.02 g; V = 0.04 L; T = 27 °C; t = 360 min) onto FCOB increased with increase in pH from 2 to 10. The highest dye
removal was observed at pH 2, which was ≅ 80.5%. From pH 4 to 8, there is a slight
decrease in the R% of CR dye. However, pH 10 shows the lowest removal of CR dye (≅50.20%).
Fig. 6 (a) Effect of pH [m = 0.02 g, t = 360 min, T = 27 °C, V = 0.04 L, C
o = 50 mg L−1], (b) FCOB adsorbent dose [pH = 2, t = 360 min, T = 27 °C, V = 0.04 L, C
o = 50 mg L−1], (c) CR dye concentration [pH = 2, m = 0.02 g, t = 360 min, T = 27 °C, V = 0.04 L], and (d) effect of equilibrium time at temperature 85 °C [pH = 2, m = 0.02 g, T = 85 °C, V = 0.04 L, C
o = 50 mg L−1].
As shown in [Fig. 4c], FCOB has negative zeta potential values in both acidic and alkaline conditions,
which may reduce the adsorption of anionic CR dye, but this is not the case here.
At pH 2, FCOB has the lowest negative value of zeta potential, which is −1.90 mV.
Therefore, there is a greater possibility of adsorption of anionic CR dye at pH 2
rather than higher pH values, because at higher pH, zeta potentials of FCOB have higher
negative values leading to higher repulsion between anionic dye and highly negatively
charged FCOB surface. Moreover, there is a high number of H+ ions at pH 2, which ultimately reduce the repulsive force between the negatively
charged FCOB and anionic CR dye. This increases the adsorption of CR onto FCOB through
electrostatic interaction.[43]
Another reason for comparatively higher adsorption at lower pH is hydrogen bonding
between the protonated functional groups of FCOB and the hydrogen bond acceptor functional
groups present in CR dye.
R% of CR decreases with increases in pH. As seen in [Fig. 4c], from pH 7 to 14, there is a decreasing trend in the negative values of the zeta
potential of FCOB. So as seen in lower pH conditions, R% should increase with an increase in pH according to zeta pt values but this is not
the case here. The increase in the concentration of OH− ions in alkaline conditions overcompensates the decrease in the negative values of
zeta potentials in basic pHs.
2.2.2
Effect of Adsorbent (FCOB) Dosage
As seen in [Fig. 6b], FCOB dose has great effect on CR dye adsorption (pH = 2; V = 0.04 L; C
o = 50 mg L−1; T = 27 °C; t = 360 min). The R% effectively increases with an increase in FCOB dose, while adsorption capacity decreases
with an increase in FCOB dose. With an increase in FCOB dose, there are more available
active adsorption sites and hence more surface area for adsorption of CR dye. Hence,
from 20 mg to 35 mg, R% increases from ≅80.5% to ≅94%. Hence, there is no accumulation of FCOB particles
with an increase in its dose, which may increase the available active surface area
for CR dye adsorption.[44] On the other hand, with an increase in FCOB dose, the same concentration of CR dye
is dispersed over a greater mass or area of FCOB, which decreases the number of CR
dye molecules adsorbed on the unit mass of FCOB. Hence, adsorption capacity decreases
(from 84.52 to 30.53 mg/g) with an increase in FCOB dose from 20 to 70 mg.
At the 50 and 70 mg FCOB dose, R% is 100%, which is the highest dye removal % any adsorbent can have. This means that
after the 50 mg FCOB dose, increasing the FCOB dose does not affect the R% of CR dye because of the saturation point. 20 mg of FCOB dose was further optimized.
2.2.3
Effect of Adsorbate (CR) Concentration
As seen from [Fig. 6c], CR concentration has a significant effect on its removal by FCOB (pH = 2; V = 0.04 L; m = 0.02 g; T = 27 °C; t = 360 min). Varying the CR concentrations from 52.48 to 600 mg/L, there is a substantial
decrease in R% from 80.20% to 18.30%. This is due to the fact that with an increase in CR concentration,
there is a saturation of active adsorption sites on FCOB making it difficult for all
CR dye molecules to get adsorbed. Moreover, at higher CR concentrations, there is
rivalry between CR molecules for available active sites of FCOB, which makes some
CR molecules unadsorbed.
Adsorption capacity (q
e) substantially increases (from 84 to 226 mg/g) with increasing CR concentration (50–600
mg/L). High CR dye concentration provides a higher concentration gradient or driving
force between bulk solution and FCOB surface, which ultimately increases q
e.[45] Moreover, at high concentrations of CR, the available active adsorption sites are
more efficiently used, which enhances the amount of CR adsorbed per unit weight of
FCOB.
As seen from the graph in [Fig. 6c], FCOB shows good R% and qe at 52.48 mg/L CR concentration. For 600 mg/L, q
e is high but R% is very low and for other CR concentrations, q
e is comparatively higher than 52.48 mg/L, but R% is very low compared to that of 52.48 mg/L CR concentration. As a result, 52.48
mg/L CR concentration was optimized for further adsorption experiments.
2.2.4
Effect of Time and Temperature
[Fig. 6d] depicts the effect of contact time between FCOB and CR dye at temperatures 85 °C
having conditions (pH = 2; V = 0.04 L; m = 0.02 g).
The adsorption experiments were also performed at 27 and 50 °C. The detailed information
and plots are given in Section S2 and Fig. S4 of the supplementary file, respectively.
The equilibrium time for 85 °C was 320 min ([Fig. 6d]). After 320 min, R% and q values remained constant, which were 99.75% and 104.70 mg/g, respectively.
The equilibrium time for CR adsorption decreases with an increase in temperature because
the kinetic energy of CR dye molecules increases at high temperatures leading to an
increase in turnover number between the CR dye and the FCOB adsorbent active sites,
which ultimately slows down the equilibrium time of adsorption. Moreover, we can see
that the R% and qe
significantly increased with an increase in temperature from 27 to 85 °C. This indicates
the endothermic nature of the adsorption of FCOB on CR.[46]
2.2.5
Leaching Behavior
The UV–visible spectra of leachates collected at various pH values after 320 min (equilibrium
time), 700 min and 5 days of contact are shown in Fig. S16. At equilibrium time, no
detectable peaks were observed in the pH range of 2–8, confirming the stability of
the FCOB composite and the absence of leaching. However, slight leaching was detected
only under highly basic conditions (pH 10–13). After 700 min, minor leaching was observed
at highly acidic conditions (pH = 2). After prolonged contact (5 days), slight leaching
was observed even in the pH 3–8 range. The leaching under extreme pH is likely due
to partial degradation of the biochar matrix and/or detachment of NiCuZnFe₂O₄ nanoparticles
from the composite surface. Overall, the FCOB composite demonstrates excellent stability
for repeated adsorption–desorption cycles within the practical pH range.
2.2.6
ANN Training
Regression plots are valuable tools for assessing the performance of ANN regression
modeling. These plots compare the predicted output values generated by the ANN with
the actual (target) values from the dataset. A high close R-value (close to 1) indicates
that ANN has effectively learned the mapping between inputs and outputs, and the regression
plot shows most data points clustered near the diagonal line. For both models (ANN-LM,
ANN-BR), MSEs versus the epochs number plot (performance plot) is shown in Figs. S12
and S14, which signifies that method performance did not change notably after 44 and
255 epochs for LM and BR algorithms, respectively. A reduction in the MSE for the
training set indicates effective learning from the training data. Optimal performance
is achieved at the point where the MSE value is minimized. The regression plots of
ANN-LM and ANN-BR models are shown in Figs. S11 and S13. The LM algorithm is faster
than the BR algorithm as the LM model requires less epochs to achieve convergence
than the BR model. The respective MSE and R-squared values are given in supplementary file (Table S5). The ANN was successfully
trained on the experimental dataset to precisely predict the adsorption capacity (q) and estimate the removal efficiency (%) for CR uptake using FCOB. Both ANN-LM and
ANN-BR models demonstrate the capability to process new inputs under diverse conditions,
including variations in pH, temperature, time (t), FCOB mass, and CR concentration.
The comparison plots for actual and ANN predicted data at these conditions are given
in Fig. S8. The ANN model developed in Simulink, along with its training dataset,
is available on GitHub and has been appropriately referenced.[47]
2.3
Adsorption Isotherm Models and Thermodynamics Parameters
The interaction between adsorbates and adsorbents is described using adsorption isotherm
models, which also aid in understanding the adsorption process, surface characteristics,
and capacity. In the present work, four adsorption isotherm models were examined:
(1) Langmuir isotherm, (2) Freundlich isotherm, (3) Temkin isotherm, and (4) Redlich
Peterson isotherm.[48] The specific equations for each model are given in [Eqs. (2)–(5)], respectively.
where, q
m = Maximum adsorption capacity (mg/g)
qt
= Adsorption capacity at time t (mg/g)
C
e = Equilibirum concentration of adsorbate (dye) solution (mg/L)
K
L = (L/mg)
K
F = (L1/n
mg1−1/n
g−1)
n
F = Represents the heterogeneity of the adsorption surface and the adsorption intensity.
It explains the degree to which the adsorption process is favorable. It is dimensionless.
B
1 = Heat of adsorption constant (J/mol). Moreover, B
1=…where, b
T = Temkin constant, which is related to the heat of adsorption (J/mol), R = Universal gas constant = 8.31 (J mol−1 K−1), T = Absolute temperature (K)
A
T = Temkin equilibrium binding constant (L g−1)
B (β) = indicates whether the adsorption process is closer to Langmuir or Freundlich
behavior. It is dimensionless.
A = It is a constant related to the adsorption capacity and energy of adsorption, obtained
from the intercept. It has unit of . Its unit depends on the value of B.
These parameters of the four adsorption isotherms were calculated from the slope and
intercept values of C
e/q
e vs. C
e (Langmuir isotherm plot, [Fig. 7a]), log q
e vs. log C
e (Freundlich isotherm plot, [Fig. 7b]), q
e vs ln C
e (Temkin isotherm plot [Fig. 7c]), and lnC
e/q
e vs. lnC
e (Redlich Peterson isotherm, plot-[Fig. 7d]). The calculated parameters for the three models are given in [Tables 2]. R
2 value (0.99) is the highest for the Redlich Peterson model than the Langmuir, Freundlich,
and Temkin models, which explains that the adsorption of CR on FCOB can be best described
by the Redlich Peterson model. The value of B for the Redlich Peterson model was 0.7468, indicating that the adsorption mechanism
is a mix that combines features of both Langmuir and Freundlich models. This suggests
a heterogeneous adsorption surface with partial monolayer coverage. The maximum adsorption
capacity (q
max) was 239.80 mg/g, which was calculated using the Langmuir model. A considerably high
q
max value indicates that FCOB has a high number of adsorption sites available to bind
CR dye molecules and FCOB is highly efficient in adsorbing target CR dye molecules.
R
L factor (Eqs. (6)) known as the separation factor is a parameter that is derived from
the Langmuir model. It helps in determining if the adsorption process is favorable
or not. It is a dimensionless quantity. The favorable, unfavorable, linear, and irreversible
nature of the adsorption process is identified by 0 < R
L <1, R
L > 1, R
L = 1, and R
L = 0 values. The R
L values for each CR dye concentration were between 0.07 and 0.5 (0 < R
L < 1) indicating favorable adsorption of CR on FCOB.
Fig. 7 (a) Langmuir, (b) Freundlich, (c) Temkin, (d) Redlich Peterson, and (e) thermodynamic parameter plots.
Table 2
Adsorption isotherm parameters for adsorption of CR on FCOB and ANN predicted data
|
Model
|
Parameter
|
Experimental
|
LM
|
BR
|
|
Langmuir
|
q (mg/g)
|
239.80 ± 0.0003
|
239.80 ± 0.0003
|
236.40 ± 0.0003
|
|
K
L (L/mg)
|
0.021 ± 0.095
|
0.02 ± 0.095
|
0.02 ± 0.088
|
|
R
2
|
0.97
|
0.97
|
0.97
|
|
Freundlich
|
K
F (mg. g−1 (L mg−1)1/n
)
|
44.77 ± 0.052
|
43.21 ± 0.052
|
42.16 ± 0.046
|
|
1/n
|
0.25 ± 0.025
|
0.25 ± 0.025
|
0.26 ± 0.022
|
|
R
2
|
0.96
|
0.97
|
0.96
|
|
Temkin
|
B
T (J/mol)
|
36.80 ± 6.05
|
36.76 ± 6.04
|
37.28 ± 6.12
|
|
A
T (L g−1)
|
0.63 ± 28.93
|
0.64 ± 28.88
|
0.57 ± 29.34
|
|
Redlich Peterson
|
R
2
|
0.89
|
0.89
|
0.90
|
|
B
|
0.74 ± 0.028
|
0.74 ± 0.024
|
0.73 ± 0.028
|
|
A
|
44.56 ± 0.13
|
44.56 ± 0.11
|
41.14 ± 0.13
|
|
R
2
|
0.99
|
0.99
|
0.99
|
The value of nF for CR adsorption was 3.88, which indicates favorable CR adsorption as n
F = 1, n
F > 1, and n
F < 1 corresponding to linear, favorable, and unfavorable adsorption. The higher n
F value (3.88) aligns with the high q
max value (239.80 mg/g). The high values of both these parameters suggest that adsorbent
FCOB has great affinity toward anionic CR dye leading to higher removal percentage
and adsorption capacity. The R
2 (0.89) value is the lowest for the Temkin model, suggesting that CR adsorption by
FCOB cannot be well explained by this model. The comparison plots between actual and
ANN predicted data for Langmuir, Freundlich, and Redlich Peterson models are given
in Fig. S9.
Thermodynamic parameters like free energy change (ΔG°), enthalpy change(ΔH°), and entropy change (ΔS°) were studied to determine feasibility-spontaneity, nature, and randomness of adsorption
reaction. ΔG° values for each temperature were derived using the following [Eq. (7)], while ΔH° and ΔS° were determined using the slope and intercept of plot of ln K
e vs. 1/T ([Fig. 7d]), respectively, as in [Eq. (8)]. The thermodynamic parameter plot is shown in [Fig. 7d].
ΔH° for CR adsorption was positive with a value of 71.02 ± 1.41 kJ/mol indicating the
endothermic nature of adsorption. This is also consistent with the increasing R% with increasing temperature. ΔS° was also positive with a value of 0.256 ± 4.34 kJ mol−1 K−1. This indicates increased randomness during CR adsorption on FCOB. The values of
ΔG° for 27, 50, and 85 °C were −5.77, −11.66, and −20.62 kJ/mol, respectively. The negative
value of ΔG° signifies the spontaneity of CR adsorption on FCOB. This means that adsorption occurs
on its own (naturally) without the need for outside energy.
2.4
Adsorption Kinetics
Adsorption kinetics models explain the rate of adsorbate release or retention from
an aqueous solution to a solid phase boundary. In the present work, four adsorption
kinetics models were examined: (1) pseudo-first-order (PFO), (2) pseudo-second-order
(PSO), (3) intra-particle diffusion (IPD), and (4) Elovich model. The equations for
each model are given in the supplementary file Section S3.
The kinetic parameters were calculated by the slopes and intercepts of the plots t/qt
vs t (PSO), log (q
e − qt
) vs. t (PFO), qt
vs. t
0.5 (IPD), and qt
vs. lnt (EM). The CR adsorption kinetics result at temperature 85 °C with ANN data
is summarized in Table S2. The PSO model had a higher R
2 value (R
2 = 0.99) than that of the PFO model (R
2 = 0.82), and the calculated q
e,cal value using the PSO model was 111.11 mg g−1, closer to the experimental q
e = 104.70 mg g−1 value, proving the better fit for the PSO model for adsorption of CR on FCOB rather
than the PFO model. From the obtained data, it was concluded that the CR adsorption
on FCOB was chemisorption and the rate of adsorption was proportional to the square
of the difference between q
e and qt
.[49] The IPD model does not exhibit a single linear line but three discrete linear lines,
which indicate that multiple processes controlled CR adsorption on FCOB, each governed
by different mechanisms. Moreover, the line did not pass through the origin indicating
that intraparticle diffusion was not the sole rate-controlling step. The first initial
linear line was observed in the first 50 min of contact between CR and FCOB. Around
64% of CR was adsorbed in this initial stage. In this stage, CR molecules diffuse
from the bulk solution to an external surface of the FCOB adsorbent. The second linear
line was observed between 51 min and 265 min of the adsorption experiment, and %CR
adsorbed in this stage was around 23%, which is less than that of the first stage
(K
p1 > K
p2). In the second stage, the rate is controlled by intraparticle diffusion. The third
or last stage corresponds to equilibrium and it is between the contact time of 266–500
min. Around 13% of the CR was adsorbed in this stage. The correlation coefficient
for CR adsorption on FCOB (R
2 = 0.93) was less for the Elovich model than the PSO model indicating that the Elovich
model was not the best fit for explaining the CR adsorptionon on FCOB. The plots for
PFO, PSO, IPD, and Elovich models at 85 °C temperature are shown in Fig. S5. Their
comparison plots with ANN predicted data are given in Fig. S10. All model plots for
27 and 50 °C are given in Figs. S6 and S7, respectively.
2.5
Plausible Mechanisms for the Adsorption of CR by FCOB
The efficient adsorption of CR (q
max = 239.80 mg/g) by FCOB results from the integrated physicochemical properties of
BTBC and NiCuZnFe2O4 ferrite. From zeta potential measurements, it is evident that FCOB has a negative
charge in both acidic and alkaline conditions. Moreover, FCOB showed the highest CR
dye removal in highly acidic condition (pH = 2) because CR is an anionic dye and there
is an electrostatic attraction between the –SO₃− group of CR dye and the protonated biochar surface at pH = 2.[50] Similar results have been reported by Nitish Semwal et al. The surface hydroxyl
groups present on NiCuZnFe₂O₄, which were formed during the coprecipitation synthesis
method from aqueous solutions, become positively charged, attracting the anionic CR
dye through electrostatic attraction. OH groups present on the FCOB surface form hydrogen
bonds with nitrogen atoms (from protonated –NH2 and –N=N– groups) and the -SO₃− group of CR. The aromatic moieties of biochar have π-electrons that interact with
the aromatic ring of CR dye through π–π stacking interactions assisting in adsorption.
The FCOB is rich in carboxyl (–COOH), hydroxyl (–OH), phenolic (–OH) groups, metal
salts, and metal ions due to the presence of S. monoica biochar and NiCuZnFe₂O₄ ferrite spinel. These functional groups and metal salts/ions
of FCOB can coordinate with the N=N and –SO₃− group of CR increasing the adsorption efficiency of it. FCOB surface may also exchange
ions (e.g., H+, Na+, K+) with anion (–SO₃−) of CR in acidic conditions where the functional groups of FCOB are protonated. This
also facilitates CR adsorption on FCOB. As observed from BET analysis, the FCOB surface
is mesoporous having a pore size of 7.0429 nm. So CR dye molecules may diffuse into
these mesopores leading to physical adsorption. The adsorption mechanism diagram is
shown in [Fig. 8].
Fig. 8 Adsorption mechanism of CR by FCOB.
2.6
Recycling Studies of FCOB
FCOB was regenerated and reused five times to analyze its stability and recyclability
following its initial use. Firstly, it was regenerated by stirring with 40 mL 0.1
NaOH for 30 min. CR dye was removed immediately after stirring it with the NaOH. Rapid
removal of CR from the FCOB surface was attributed to the fact that it showed the
least CR adsorption (50%) in highly alkaline pH ([Fig. 6a]). The deprotonation of the FCOB surface in high OH− concentration yielded more negative charge on it, which reduced the electrostatic
attraction between FCOB and CR dye; hence, the dye desorbed rapidly from FCOB. Thus,
the surface charge of the adsorbent and pH of the highest dye removal influence the
choice of solvent used for the desorption of dye from the adsorbent surface.
As shown in the [Fig. 9], the R% of CR reduced drastically from 99.7% to 77% in the first recycling process. Progressively,
the R% decreased in the second cycle to 71.6%, the third cycle to 68.5%, the fourth cycle
to 67.1%, and the fifth cycle to 64.9%. This was due to a strong attractive force
between the CR dye and FCOB surface that some of the CR dye molecules were firmly
attached to active sorption sites of FCOB even after regenerating FCOB with an efficient
desorbing solvent. Desorption efficiency (%) decreases after multiple cycles because
the FCOB adsorbent’s active sites become less available due to incomplete removal
of the CR dye or gradual structural degradation. Additionally, some CR dye molecules
may irreversibly bind to the FCOB adsorbent, reducing its regeneration efficiency
over time. The FCOB showed removal after five regeneration–recycling studies making it a fairly good adsorbent
material considering the fact that a readily accessible and cheaper desorbing solvent
was used for regeneration of FCOB. The adsorption efficiency (%), desorption efficiency
(%), and q
e were calculated from [Eqs. (11), (9), and (12)].
Fig. 9 Adsorption performance of FCOB after recycling (adsorption: 20 mg adsorbent, 50 ppm
CR, 40 mL, pH 2, 85 °C, 320 min, 250 rpm; desorption: 40 mL 0.1 N NaOH, pH 13, 27
°C, 30 min, 250 rpm).
where, q
d = amount of adsorbate desorbed (mg/g)
q
a = amount of adsorbate initially adsorbed during the adsorption cycle (mg/g).
2.7
Comparison with Other Biochar Adsorbents
Many researchers have investigated the removal of CR by biochar adsorbents. The adsorbent
prepared by modifying the orange peel biochar with hexadecyl trimethyl ammonium bromide
(CTAB) had a high BET surface area of 618.44 m2/g and a high Langmuir adsorption capacity of 609.8 mg g−1 for CR dye as depicted in [Table 3]. The mesoporous nano-zerovalent manganese (nZVMn) and Phoenix dactylifera leaves
biochar (PBC) composite had a very low BET surface area of 57.74 cm2/g. Despite having a low BET surface area, it removed CR efficiently with a q
max of 117.647 mg g−1 ([Table 3]). Thus, the advantages and limitations of some of the biochar adsorbents compared
to FCOB are reported in [Table 3].
Table 3
Comparison of various biochar adsorbents utilized in the literature for the removal
of CR
|
Biomass
|
Adsorbent
|
q
max (mg/g)
|
Reference
|
|
S. monoica leaves
|
FCOB
|
239.80
|
Present work
|
|
Orange peel
|
NOBC
|
609.80
|
[43]
|
|
Phoenix dactylifera leaves
|
nZVMn/PBC
|
117.64
|
[51]
|
|
Corncobs agriculture waste
|
AMBC
|
89.30
|
[52]
|
|
Green pea peels
|
ZnO/GPBC
|
114.94
|
[53]
|
|
Pine needles
|
PEI-BC
|
294.11
|
[54]
|
|
Breadfruit leaf
|
Biochar made from breadfruit leaves
|
17.81
|
[55]
|
|
Musa acuminata stem
|
Biochar prepared from Musa acuminata stem
|
135.15
|
[56]
|
|
Rice husk
|
Biochar derived from rice husk
|
42.91
|
[57]
|
|
Green pea husk
|
CGPH
|
2.69
|
[58]
|
|
Sugarcane bagasse
|
Fe3O4/SBB
|
74.07
|
[59]
|
4
Experimental Section
4.1
Materials
The information about chemicals used in the experiment is given in supplementary file
(Section S3). The halophyte S. monoica (Genus: Suaeda, Species: Monoica, and Family: Amaranthaceae) was collected from Nh141, Kandla Port Road, Gandhidham, Gujarat 370210, India (23.026038°N,
70.192009°E). This saline halophyte is widely located in the intertidal zones of Kandla.
Images of biomass from the collection site, dried plant, and LP are shown in [Fig. 10a–c].
Fig. 10 Images of (a) biomass from collection site, (b) dried plant, and (c) leaves powder.
4.2
Synthesis of BC300
The collected biomass was thoroughly washed with distilled water thrice to eliminate
all the dust particles and impurities. It was then dried naturally in sunlight for
seven days to remove the moisture. Sun drying was preferred over oven drying because
it preserves plant tissues’ cellular structure and bioactive compounds with its gentler
drying process. In contrast, oven drying may cause the deformation of plant tissues
with its rapid drying process, which could impact the research studies on it. The
leaves, twigs, stems, and roots were manually separated from the dried plant and cut
into small pieces. The small pieces were pulverized into fine powder by a mortar pestle.
Afterward, the raw powders of leaves, twigs, stems, and roots were stored in an airtight
container. The twigs, stems, and roots were stored for future research. In the present
work, LP was further utilized to synthesize biochar (BC300). About 18 g of LP was
pyrolyzed at 300 °C for 2 h at a heating rate of 2.50 °C/min without oxygen in a muffle
furnace to obtain BC300. The yield of BC300 was 11.90 g. Here the yield of BC was
less than the reactant biomass due to loss of moisture, volatile light gases, and
breaking of bioactive compounds like protein, resins, tannins, cardiac glycosides,
terpenoids, flavonoids, phenols, acidic compounds, and glycosides present in succulent
leaves.[60] BC300 was stored in an oven overnight to remove moisture. The BC300 yield can be
calculated from [Eq. (10)].[61]
4.3
Synthesis of BTBC
About 11.90 g BC300 was immersed in 1.19 L of 0.10 N NaOH (pH = 13) under continuous
mechanical stirring for 4 h as a pretreatment of BC300. Then, the immersion was filtered
and washed with distilled water to remove impurities. It was followed by drying the
precipitates in the oven at 65 °C overnight. The resulting BTBC was stored in an airtight
container. The alkaline pretreatment of BC300 was performed to improve porosity and
unclog the partially blocked pores of BC300, which increased the total surface area
of biochar and hence the adsorption capacity.[62]
4.4
Synthesis of FCOB
FCOB was prepared by the same method as reported in our previous work.[63] A total of 4 ferrite composites of biochar (FCOB) having different ratios of BTBC:
NiCuZnFe2O4 were prepared and tested for CR dye adsorption (See supplementary file, Section S6).
The best-performing FCOB (F3) (Table S4) had a 2:1 ratio of BTBC: NiCuZnFe2O4, so it was selected for bulk synthesis of FCOB. Briefly, 50 mL of distilled water
was added to 9 g BTBC, and the mixture was irradiated with an ultrasonic system for
20 min. Totally, 0.03, 0.005, 0.005, and 0.005 moles of Fe(NO₃)₃·9H₂O, Ni(NO3)2·6H2O, Cu(NO3)2·3H2O, and Zn(NO3)2·6H2O metal salts were respectively mixed with 10 mL of distilled water and stirred for
half an hour to make a homogenous mixture. Further coprecipitation method was used
to make a composite of NiCuZnFe2O4 – ferrite spinel and BTBC.[64] The BTBC solution was added to the homogenous mixture of metal salts. NaOH was added
dropwise into the mixture until its pH reached 11. Then, the mixture was stirred for
2 h with heating at 60 °C. Afterward, the mixture was filtered and washed with distilled
water. The black-colored precipitates were dried in an oven overnight at 60 °C. Then
it was calcined at 300 °C for 2 h and dried in an oven. The schematic diagram for
the synthesis of FCOB from LP is shown in [Fig. 11].
Fig. 11 Schematic diagram for synthesis of FCOB from LP.
4.5
Characterizations
All the characterizations of LP, BC300, BTBC, and FCOB were done before adsorption,
and characterizations of FC-CR were done after adsorption. The information about the
characterizations performed is given in the supplementary file (SectionS5).
4.6
Adsorption and Regeneration Experiments of FCOB
Throughout all the adsorption and regeneration experiments, the concentration of adsorbate
CR dye was measured using UV–vis spectroscopy (TCC-240A UV–Vis spectrophotometer,
by Shimadzu) in the λ range of 200–800 nm. The absorbance of the CR dye solution was
measured at λ
max = 497.5nm.
The aqueous solution of CR with 600 mg L−1 was prepared and used as a stock solution throughout the experiment. It was prepared
accurately by weighing 0.600 g of CR dye and dissolving it in a small volume of distilled
water using a beaker. Once it was fully dissolved, the solution was transferred to
a 1 L volumetric flask and diluted to the mark with distilled water and mixed thoroughly
to ensure a homogeneous solution. It was stored in a properly labeled container for
further use.
A 100 mL conical flask was used for each parameter of the adsorption experiment. The
effect of different pH (2, 4, 6, 8, 10) on adsorption was studied at fixed parameters
like 20 mg FCOB dose, 50 mg L−1 CR dye concentration, 450 min contact time, and 27 °C temperature. All the necessary
pH changes were done using 0.10 N HCl and 0.10 N NaOH and pH was measured using a
pH meter (EI Deluxe). The effect of varying amounts of FCOB (20, 30, 35, 50, 70 mg)
was also studied at fixed parameters like 2 pH, 50 mg L−1 CR dye concentration, 450 min contact time, and 27 °C temperature. The effect of
different CR dye concentrations (52.48, 80, 184, 517, 600 mg L−1) was also studied at fixed parameters of 2 pH, 20 mg FCOB dose, 450 min contact time
and 27 °C temperature. The effects of temperature were studied at fixed parameters
like 2 pH, 20 mg FCOB dose, and 50 mg L−1 CR dye concentration. The temperature effect study was performed at three different
temperatures of 27, 50, and 85 °C for the contact time of 450 min. FCOB was separated
from CR dye using an electrical centrifuge machine. The stirring speed of magnetic
stirrer was kept constant at 250 rpm throughout all the adsorption experiments.
The initial concentration of the CR dye (C
o, mg L−1) and the final concentration of the CR dye (C
e, mg L−1) were used to calculate the removal percentage (R%) of CR as shown in [Eq. (11)]:
where C
o and C
e are in mg L−1 unit.
C
o, C
e, V (volume of CR dye, L), and m (FCOB dose, g) were used in the calculation of adsorption capacity at equilibrium
(q
e, mg g−1) as shown in [Eq. (12)]:
Regeneration of the adsorbent is essential because of its sustainability, resource
conservation, cost-effectiveness, and environmental control. For regeneration studies,
0.1 N NaOH was used as the desorbing agent. In each cycle, 40 mL of NaOH solution
was added to 20 mg of dye-loaded adsorbent (FC-CR). Thus, the desorption was carried
out at a pH of approximately 13, as 1 N NaOH was used. The mixture was stirred on
a magnetic stirrer (250 rpm) at a fixed temperature of 27 °C for 30 min. After desorption,
the adsorbent was filtered, washed thoroughly with distilled water, dried, and reused
for the adsorption of CR dye under the same optimized conditions (pH = 2, FCOB adsorbent
dose = 20 mg, CR dye concentration = 50 ppm, 85 °C temperature, contact time = 320
min, stirring speed = 250 rpm). Further, CR dye was desorbed from FC-CR (ferrite composite
of biochar after CR adsorption) using the same desorption method and reused for CR
adsorption. After each adsorption and desorption cycle, FCOB was washed with distilled
water. Thus, FCOB adsorbent was regenerated and reused for five cycles.
4.7
Adsorbent (FCOB) Leaching
To assess the potential leaching of particles or matrix components from the FCOB adsorbent
composite during the adsorption–desorption process, UV–vis spectroscopy was employed.
Solutions with a composite-to-water mass ratio of 0.00075:1 (0.75 mg mL−−1) were prepared across a pH range of 2–12, adjusted using 0.1 M HCl or 0.1 M NaOH.
The mixtures were stirred for a predetermined period, followed by filtration. UV–vis
measurements were conducted with baseline corrections applied to eliminate any interference
from peaks originating from the pH-adjusting agents.
4.8
Artificial Neural Networks (ANNs)
ANNs are a machine learning technique that mimic the structure of neurons in the human
brain. They consist of input layers, hidden layers, and output layers. By adjusting
connection weights during training, an ANN model is created that identifies patterns
and trends in the input data.[65] With proper training, ANNs can efficiently capture complex multivariate functions
without extensive computations. This capability allows them to quickly interpret intricate,
nonlinear relationships within data. The hidden layers are responsible for processing
the input data and using it to make predictions. The number of hidden layers varies
depending on the complexity of the model.
In this study, the ANN architecture includes five input variables (FCOB mass, CR dye
concentration, temperature, time, and pH), 6 hidden layers for predicting the R% output, and adsorption capacity (q). There are two output layers for the adsorption capacity (q) and R%. The relationship between the experimental dataset and the ANN is crucial; the dataset
forms the foundation that allows the ANN to learn complex correlations between the
inputs and outputs. During training, the ANN understands these correlations, enabling
it to predict outcomes for new input values. This results in faster predictions for
specific experimental conditions.
The experimental dataset used in this study consists of 44 data points, which are
divided as follows: 15% for validation, 70% for training, and 15% for testing. This
division ensures that the model learns effectively, provides accurate predictions,
and improves its generalization ability. The ANN architecture is represented in Fig.
S15.
4.9
Statistical Analysis
All batch adsorption experiments were performed in triplicate, and the results are
presented as mean values ± standard deviation (SD) in [Fig. 6] with error bars. Statistical analysis, including calculation of error bars and regression
coefficients, was carried out using OriginPro (Version 8.5.) and Microsoft Excel (Version
2507). The isotherm and kinetic model parameters were fitted based on the linear regression
method, and the goodness-of-fit was assessed using R
2
adj (adjusted determination coefficient) values. The formulas for R
2
adj and SD are given in the [Eqs. (13) and (14)], respectively. A small value of SD and a value of R
2
adj close to unity indicate a reliable, good curve fit by a model.[66]
where, q
i,exp = An individual, i, data obtained from the batch experiment,
q
i,model = The estimation of the corresponding q
i,exp generated by each model,
n = Number of data points, and
p = Number of parameters in the model.