You are on page 1of 19

Article

Adsorption of Arsenic from Water Using Aluminum-Modified


Food Waste Biochar: Optimization Using Response
Surface Methodology
Sayed Q. Hashimi 1, Seung-Hee Hong 2, Chang-Gu Lee 3 and Seong-Jik Park 2,4,5,*

1 Department of Chemical Engineering, Hankyong National University, Anseong 17579, Korea


2 Department of Integrated System Engineering, Hankyong National University, Anseong 17579, Korea
3 Department of Environmental and Safety Engineering, Ajou University, Suwon 16499, Korea

4 Department of Bioresources and Rural System Engineering, Hankyong National University,

Anseong 17579, Korea


5 Institute of Agricultural Environmental Sciences, Hankyong National University, Anseong 17579, Korea

* Correspondence: parkseongjik@hknu.ac.kr; Tel.: +82-31-670-5131; Fax: +82-31-670-5139

Abstract: Aluminum-impregnated food waste was selected as a filter medium for removing As(III)
from aqueous solutions. The modification of food waste and its carbonization conditions were op-
timized using the Box–Behnken model in the response surface methodology. Pyrolysis temperature
and Al content significantly influenced the As(III) adsorption capacity of aluminum-modified food
waste biochar (Al-FWB), but the pyrolysis time was insignificant. Several factors affecting the ad-
sorption capacity of the Al-FWB, including the pH, contact time, dosage, competitive anions, and
reaction temperature, were studied. The low solution pH and the presence of HCO3−, SO42−, and
PO43− reduced the As(III) adsorption onto Al-FWB. The pseudo-second order model showed a better
Citation: Hashimi, S.Q.; Hong, S.-H.;
Lee, C.-G.; Park, S.-J. Adsorption of
fit for the experimental data, indicating the dominance of the chemisorption process for As(III) ad-
Arsenic from Water Using sorption. Langmuir and Freundlich isotherm models fit the adsorption data, but the Langmuir
Aluminum-Modified Food Waste model with a higher (R2) value showed a better fit. Hence, As(Ⅲ) was adsorbed onto Al-FWB as a
Biochar: Optimization Using monolayer, and the maximum As(Ⅲ) adsorption capacity of Al-FWB was 52.2 mg/g, which is a
Response Surface Methodology. good value compared with the other porous adsorbents. Thus, Al-FWB is a promising low-cost ad-
Water 2022, 14, 2712. https://doi.org/ sorbent for removing As(III) from aqueous solutions and managing food waste.
10.3390/w14172712

Academic Editors: Yanshan Cui and


Keywords: arsenic; response surface methodology; food waste; biochar; optimization
Margaritis Kostoglou

Received: 26 July 2022


Accepted: 29 August 2022
1. Introduction
Published: 31 August 2022
The contamination of natural water bodies and wastewater by arsenic is a serious
Publisher’s Note: MDPI stays neu-
problem that originates from natural and human activities. The composition of natural
tral with regard to jurisdictional
ores that undergo mining activities and erosion and the use of pesticides are known
claims in published maps and institu-
sources of As [1]. As(V) predominantly exists in surface water bodies, but As(III), with a
tional affiliations.
lower redox potential, is a dominant species in groundwater [2,3]. As(III) is more toxic
and mobile than As(V), and it is difficult to remove the As(III) present in wastewater [4].
As is a carcinogen and phytotoxic element with harmful effects on human health and eco-
Copyright: © 2022 by the authors. Li-
systems [5]. Considering the adverse health and environmental effects, as well as the vast
censee MDPI, Basel, Switzerland.
natural exposure to As [6], it is important to investigate and develop methods and tech-
This article is an open access article nologies to reduce As to the threshold limit (0.01 mg/L) recommended by the World
distributed under the terms and con- Health Organization [7].
ditions of the Creative Commons At- Recent investigations have led to the introduction of various technologies for remov-
tribution (CC BY) license (http://crea- ing As from water and wastewater, including oxidation-coagulation [8], electrocoagula-
tivecommons.org/licenses/by/4.0/). tion [9], ion exchange [10], precipitation [11,12], ultrafiltration [13], reverse osmosis [14],

Water 2022, 14, 2712. https://doi.org/10.3390/w14172712 www.mdpi.com/journal/water


Water 2022, 14, 2712 2 of 19

and adsorption [15]. However, owing to the high cost, the requirement for more equip-
ment and different chemicals, and the complexity of these techniques and operating pro-
cedures, less attention has been paid to some of these technologies [16]. Consequently, the
most utilized and promising technology is adsorption because of its low cost and energy
consumption for removing contaminants from water and wastewater. Adsorption tech-
nology is widely applicable for environmental remediation and has simple application
procedures with a higher efficiency than other conventional technologies [17–19]. Using
adsorption technology for environmental remediation minimizes environmental impacts
and waste production [20]. The commonly employed adsorbents for removing As from
aqueous solutions are iron oxyhydroxides [21], iron oxide-modified adsorbents [22], acti-
vated alumina [23], and iron-modified activated carbon [24].
Biochar is a low-cost adsorbent produced by pyrolyzing organic material in the ab-
sence of oxygen; it is highly efficient at removing contaminants from water and is there-
fore widely used in environmental remediation [25–27]. The properties of various organic
residues, including sewage sludge [28], wood [29], rice husk [30], corn cob and stalk [31],
Rhodes grass [32], spent mushroom compost [33], pomelo peel [34], orange peels [35],
palm kernel shell [36], and palm waste [37] have been studied to produce biochar and
have demonstrated different levels of utilization capability [38]. Previous studies have re-
vealed that organic waste materials can be a great source of raw materials for producing
biochar through pyrolysis [39].
Many studies have investigated the efficiency of biochar in removing contaminants
consisting of cations and polar organic molecules, such as phenolics [40], halogenated
compounds [41], and solvents [42], from aqueous solutions. Scientific studies have stated
that the ability of biochar to remove positively charged species is dependent on both its
high surface area and the negative charge on its surface; however, the negative charge on
the biochar surface decreases its tendency for the sorption of anions and oxyanions, such
as As [43]. To enhance the negative ion adsorption capacity of biochar, several methods
have been developed to modify biochar, such as steam activation [44], alkali activation
[45], metal oxide-modified biochar [46], and clay mineral and biochar composites [47].
In this study, aluminum-modified food waste biochar (Al-FWB) was prepared and
utilized to remove As(III) from aqueous solutions. No artificially selected single source of
food waste was used, but instead the actual food waste collected from food waste treat-
ment plants was used. The effects of a set of experimental conditions, such as solution pH,
competitive anions, adsorbent dose, reaction time, initial concentration, and reaction tem-
perature, were studied to evaluate the arsenic removal capability of Al-FWB. Equilibrium,
kinetic, and thermodynamic adsorption model analyses were performed further to char-
acterize As(III) adsorption to Al-FWB. Consequently, the objectives of the present study
were to (1) assess the As(III) removal ability of Al-FWB and optimize the adsorbent prep-
aration using response surface methodology (RSM), which enables the analysis of the var-
ious independent variables influencing the dependent variables in one system with mul-
tiple experimental runs [48]; (2) study the effects of solution pH, ionic competition, adsor-
bent dosage, contact time, initial As(III) concentration, and temperature on the As(III) re-
moval capacity of Al-FWB; and (3) gain insight into the adsorption capability of the bio-
char and the mechanism of As(III) adsorption.

2. Materials and Methods


2.1. Preparation and Characterization of Adsorbent
The food waste (FW) used in this study is described in our previous study [49]. The
FW was collected from households and transported to an FW treatment plant. The water
in the FW was partially removed by squeezing it through a screw press and then drying
it at 150 °C using a steam boiler. The impurities in the dried FW were separated using a
magnet and trommel separator.
Water 2022, 14, 2712 3 of 19

The modification process was performed by adding 100 g of raw FW to 300 mL of the
solutions prepared by separately dissolving 0.25 M, 0.5 M, and 0.75 M of AlCl3 in deion-
ized (DI) water, which resulted in 2, 4, and 6 wt% Al contents in a mixture, respectively.
The AlCl3 used for preparing Al-FWB was purchased from Sigma Aldrich (Burlington,
MA, USA). The mixture of solution and FW was constantly stirred at 100 rpm overnight.
The mixture was then dried at 100 °C for 24 h. Subsequently, the modified FW was car-
bonized at different temperatures (300, 450, and 600 °C) for various durations (0.5, 2, and
3.5 h) under nitrogen gas. The FW-derived biochar was then crushed, sieved (40 mesh),
and stored in a desiccator.
A field emission scanning electron microscope (FE-SEM, S-4700, Hitachi, Japan) was
used to investigate the morphologies of the Al-modified biochar. The elemental composi-
tion of Al-FWB was analyzed using an X-ray fluorescence (XRF) spectrometer (XRF-1700,
Shimadzu, Japan) and an energy dispersive spectrometer (EDS) attached to the FE-SEM.
The crystalline structure of impregnated Al and other inorganic minerals were analyzed
using an X-ray diffractometer (XRD) (SmartLab; Rigaku, Japan). The functional groups on
the surface of Al-FWB were identified using Fourier-transform infrared spectroscopy
(FTIR) (Nicolet iS10 spectrometer, Thermo Scientific, Waltham, MA, USA) in transmit-
tance mode. The specific surface area and porosity of the Al-FWB were measured by ana-
lyzing nitrogen adsorption with a Brunauer–Emmett–Teller (BET) surface analyzer
(Quandrasorb SI, Quantachrome Instrument, Boynton Beach, FL, USA).

2.2. Response Surface Methodology (RSM)


The Box–Behnken model in the RSM was used to optimize the conditions for prepar-
ing and modifying Al-FWB for removing As(III) from aqueous solutions. Three independ-
ent variables, temperature (300, 450, and 600 °C), modified Al dosage (2, 4, and 6 w/w %),
and contact time (0.5, 2, and 3.5 h), were selected, and the arsenic removal percentage was
considered as the response variable. Overall, 17 trial combinations were analyzed using
the Design-Expert statistical software (version 7.0.0; STAT-EASE Inc., Minneapolis, MN,
USA). A quadratic model generated from the experimental data was also used for the
analysis of variance (ANOVA) to determine the significance of the achieved model. The
recommended highest-order polynomial model of the statistical software was not aliased
and had a significant additional term. A quadratic model was generated from the experi-
mental data based on the design and according to the following polynomial equation:

𝑦𝑦 = 𝑎𝑎0 + 𝑎𝑎1 𝐴𝐴 + 𝑎𝑎2 𝐵𝐵 + 𝑎𝑎3 𝐶𝐶 + 𝑎𝑎12 𝐴𝐴𝐴𝐴 + 𝑎𝑎13 𝐴𝐴𝐴𝐴 + 𝑎𝑎23 𝐵𝐵𝐵𝐵 + 𝑎𝑎11 𝐴𝐴2 + 𝑎𝑎22 𝐵𝐵2 + 𝑎𝑎33 𝐶𝐶 2 (1)

where (y) is the predicted amount of As(III) adsorbed onto the unit mass of Al-FWB bio-
char (mg/g), and (a0) is the constant coefficient. A, B, and C are the independent variables,
namely, temperature (°C), reaction time (h), and impregnated Al dosage (w/w %), respec-
tively. In addition, ai is a model coefficient parameter. The response variables and design
of the present experiment are listed in Table 1. The goodness of fit of the model, regression
coefficient (R2), lack of fit, and adjusted and predicted R2 were also evaluated.

Table 1. Box–Behnken design: list of independent variables and levels.

Values of the Coded Factor Levels


Independent Variables Symbol
−1 0 +1
Temperature (°C) A 300 450 600
Pyrolysis time (h) B 0.5 2 3.5
Al content (w/w %) C 2 4 6
Water 2022, 14, 2712 4 of 19

2.3. Batch Adsorption Experiments


For batch adsorption, a set of conical tubes (50 mL), each containing 30 mL arsenic
contaminated solution with an initial concentration of 300 mg/L As(III) and 0.1 g Al-FWB,
were used. An As(III) stock solution of 1000 mg/L was prepared by dissolving 2.6379 g of
As(III) oxide (Sigma Aldrich, Burlington, MA, USA) in DI water. As(III) solutions of vari-
ous concentrations were prepared by diluting the As(III) stock solution with DI water.
The pH of the experimental solution was adjusted by adding 0.1 M NaOH or HCl solution.
Adsorption was performed using a digital shaking incubator to maintain a 100 rpm agi-
tation rate and 25 °C temperature for 24 h. Thereafter, the solution was filtered through a
5–10 μm pore size qualitative filter paper (Advantec, Japan) and was filtered using a 0.45
μm syringe filter (25HP045AN, Advantec, Japan). Subsequently, the concentration of
As(III) in the experimental solutions was determined using inductively coupled plasma-
optical emission spectrometry (ICP-OES; Agilent 5100; Agilent Technologies, USA) after
adsorption. The experiment was performed in triplicate.
The effect of the solution pH was examined at a pH range of 3–11, with 100 mg/L
initial As(III) concentration, 3.33 g/L adsorbent dosage, 25 °C temperature, and 100 rpm
agitation for 24 h. The influence of the co-existing anions on As(III) adsorption was tested
in the presence of NO3−, HCO3−, SO42−, and PO43− anions at initial concentrations of 1 mM
and 10 mM, with 100 mg/L initial concentration As(III), 3.33 g/L adsorbent dosage, 25 °C
temperature, and 100 rpm agitation for 24 h. The 1 mM and 10 mM stock solutions of
NaNO3, Na HCO3, and Na2SO4 were prepared by mixing a 1:1 volume ratio of 2 mM and
20 mM of each chemical stock solution with 200 mg/L of As(III) stock solution. To test the
effect of the adsorbent dosage on arsenic adsorption, a series of adsorbent dosages (1.67,
3.33, 6.67, 10.00, and 13.33 g/L) were used with initial As(III) concentrations of 100 and 300
mg/L. The experiment was performed at 25 °C with agitation at 100 rpm for 24 h.
Kinetic, equilibrium, and thermodynamic studies of As(III) adsorption were also con-
ducted. The adsorption kinetics of As(III) in the batch experiment was similar to that in
the equilibrium test. However, the batch experiment was conducted for a series of 21 sam-
ples, each containing 30 mL of As(III) solution and 0.1 g adsorbent in a 50 mL conical tube,
with 0.5, 1, 2, 3, 6, 12, and 24 h contact time. In contrast, for the equilibrium experiment,
the initial As(III) concentration varied between 10–1000 mg/L, and the contact time was
24 h. Thermodynamic experiments were performed using 0.1 g FW biochar with 30 mL
As(III)-contaminated solution at different temperatures (15 °C, 25 °C, and 35 °C).

2.4. Analysis of the Experimental Data


The adsorption kinetics data were analyzed using pseudo-first order (Equation (2))
and pseudo-second order (Equation (3)) models:
𝑞𝑞𝑡𝑡 = 𝑞𝑞𝑒𝑒 (1 − 𝑒𝑒 −𝑘𝑘1 𝑡𝑡 ) (2)

𝑘𝑘2 𝑞𝑞𝑒𝑒2 𝑡𝑡
𝑞𝑞𝑡𝑡 = (3)
1 + 𝑘𝑘2 𝑞𝑞𝑒𝑒 𝑡𝑡
where qt is the amount of As(III) adsorbed at time t (mg As(III)/g), qe is the amount of
As(III) removed at equilibrium (mg As(III)/g), k1 is the pseudo-first order rate constant
(1/h), and k2 is the pseudo-second order rate constant (g/mg As(III)/h).
The adsorption equilibrium data were analyzed using the Langmuir model (Equa-
tion (4)) and the Freundlich model (Equation (5)):
𝑄𝑄𝑚𝑚 𝐾𝐾𝐿𝐿 𝐶𝐶𝑒𝑒
𝑞𝑞𝑒𝑒 = (4)
1 + 𝐾𝐾𝐿𝐿 𝐶𝐶𝑒𝑒
1
𝑞𝑞𝑒𝑒 = 𝐾𝐾𝐹𝐹 𝐶𝐶𝑒𝑒𝑛𝑛 (5)
Water 2022, 14, 2712 5 of 19

where C0 and Ce are the concentration of As(III) in the aqueous solution at initial and equi-
librium (mg As(III)/L), respectively; KL is the Langmuir constant related to the binding
energy (L/mg As(III)); Qm is the maximum amount of As(III) removed per unit mass of Al-
FWB (mg-As(III)/g); KF is the distribution coefficient (L/g); and n is the Freundlich con-
stant.
The adsorption thermodynamics data were analyzed using the following equations:

∆𝐺𝐺 0 = ∆𝐻𝐻0 − 𝑇𝑇∆𝑆𝑆 0 (6)

∆𝐺𝐺 0 = −𝑅𝑅𝑅𝑅𝑅𝑅𝑅𝑅𝐾𝐾𝑒𝑒 (7)

𝛥𝛥𝛥𝛥 0 ∆𝐻𝐻0
𝑙𝑙𝑙𝑙 𝐾𝐾𝑒𝑒 = − (8)
𝑅𝑅 𝑅𝑅𝑅𝑅
𝛼𝛼𝑞𝑞𝑒𝑒
𝐾𝐾𝑒𝑒 = (9)
𝐶𝐶𝑒𝑒
where ∆𝐺𝐺 0 is the change in Gibbs free energy (kJ/mol), ∆𝑆𝑆 0 is the change in entropy
(J/mol K), ∆𝐻𝐻0 is the change in enthalpy (kJ/mol), R is the universal gas constant (=8.314
J/K·mol), T is the absolute temperature (K), Ke is the equilibrium constant, and 𝛼𝛼 is the
amount of the adsorbent (g/L).

3. Results and Discussion


3.1. Response Surface Methodology (RSM) Study
In this study, a quadratic model comprising 17 trials (experimental results) was gen-
erated using the Box–Behnken design of RSM. The desired range and levels of all three
variables, namely, pyrolysis temperature, time, and Al content, are shown in Table 2.
ANOVA was used to perform an F-value test in order to calculate the significance of the
employed RSM model of the Design Expert statistical software (version 7.0.0. STAT-EASE
Inc., MN, USA). The highest-order polynomial quadratic model recommended by the re-
sults of the dependent variable (YA) indicated that the model was not aliased, the addi-
tional terms were significant, and the criteria were satisfied. The predicted responses were
obtained using the following quadratic model equation of the Box–Behnken design:

𝑌𝑌𝐴𝐴 = 25.16 + 2.36𝐴𝐴 + 1.25𝐵𝐵 + 4.79𝐶𝐶 − 0.29𝐴𝐴𝐴𝐴 + 0.81𝐴𝐴𝐶𝐶 + 0.88𝐵𝐵𝐵𝐵 − 11.855𝐴𝐴2 + 1.24𝐵𝐵2 − 2.44𝐶𝐶 2 (10)
Terms with a positive sign in front of them show the synergistic effect of the term,
while those with a negative sign indicate the antagonistic effect. The significance of the
values showing the quality of the model developed and the amount of As(III) adsorbed
were determined using 𝐹𝐹-value, 𝑅𝑅2 , adjusted 𝑅𝑅2 , lack-of-fit, and adequate precision ex-
aminations. As shown by the ANOVA analysis in Table 2, the F-value of the model, cal-
culated by dividing the mean square of each variable’s effect by the mean square, was
14.72, implying that the model was significant. There is only a 0.09% chance that a “Model
F-value” this large could occur because of noise. The model probability was 0.0009, show-
ing a value of less than 0.05, indicating that the calculated values for all model terms were
significant. The 𝑅𝑅2 value showing goodness-of-fit was 0.9498. The 𝑅𝑅2 indicated that
94.98% of the total variation in the adsorbent capability was ascribed to the independent
variables in the adsorbent preparation process. Furthermore, the standard deviation in
Equation (10) was 2.58. The closer R2 is to one, along with a small standard deviation, the
better the model fit will be in providing predicted values closer to the actual values for
the response. The adjusted R2 value of the model was 0.89. The adequate precision value
was 13.34. Adequate precision measures the signal to noise ratio; a ratio greater than 4 is
desirable and indicates an adequate signal. According to Table 2, the result of the lack-of-
fit test was significant, with a high value of 14.48. Lack-of-fit examination defines model
adequacy, and a significant lack-of-fit shows poor fit.
Water 2022, 14, 2712 6 of 19

Table 2. ANOVA results obtained from the response surface model for analyzing the amount of
As(III) adsorbed per unit mass of Al-FWB.

Sum of Mean F p-Value


Source
Squares df Square Value Prob > F
Model 878.9 9 97.7 14.7 0.0009
A: Temp 44.4 1 44.4 6.7 0.0362
B: Time 12.4 1 12.4 1.9 0.2136
C: Al Content 183.9 1 183.9 27.7 0.0012
AB 0.34 1 0.34 0.05 0.8281
AC 2.64 1 2.64 0.4 0.5479
BC 3.2 1 3.2 0.5 0.5132
A2 591.7 1 591.7 89.2 <0.0001
B2 6.5 1 6.5 0.97 0.3565
C2 25.03 1 25.03 3.8 0.0932
Residual 46.5 7 6.64
Lack of Fit 43.5 3 14.5 19.3 0.0076
Pure Error 3.0 4 0.75
Cor Total 925.4 16

The interaction of the independent variables, namely pyrolyzing temperature (A),


time (B), and aluminum content (C), affected the adsorption capacity of the biochar, as
illustrated by the three-dimensional response surface curves (Figure 1a–c). Each curve
was plotted for the combined effect of a pair of factors, whereas the remaining parameter
was set at the center point. The negative and positive signs of the factor coefficient in the
quadratic model equation (Equation (10)) indicated whether the effect of a factor on As(III)
adsorption onto Al-FWB was synergistic or antagonistic. ANOVA for the response surface
quadratic model indicated the significance of the first-order and second-order effects of
the variables (p < 0.05). As seen in Table 2, the first-order effect of the temperature and
aluminum content showed significant values, while only the Al content was significant in
the case of the second-order effect of the variables. There was no significant interaction
effect between each pair of variables. The insignificant effect (p > 0.05) of pyrolysis time
(B) on the adsorption capacity of metal-impregnated biochar has also been observed in
previous studies [50–52].
From the positive signs of A and C in the regression model equation, it was inferred
that the amount of As(III) adsorbed by Al-FWB increased with the increasing temperature
(from 300 °C to 600 °C) and impregnated Al content (from 2 to 6 w/w %). The positive sign
of A and the negative sign of A2 indicated that the maximum As(III) adsorption capacity
of Al-FWB could be obtained between the minimum (300 °C) and maximum (600 °C) val-
ues of A. This is consistent with the curvature of the estimated response surface of the
As(III) adsorption as a function of A (pyrolysis temperature), as shown in Figure 1a,c. The
positive sign of the first and second order of the C factor indicated that the As(III) adsorp-
tion capacity of Al-FWB could be enhanced by increasing the Al content. Another study
reported the effect of a modified aluminum fraction for improving the adsorption capacity
of agricultural biochar [53]. They explained that the capacity of the biochar for As(V) ad-
sorption increased from 0.051 mg/g to 0.48 mg/g as the Al content was raised from 0.00 to
0.5 M, and the enhanced efficiency was attributed to the presence of Al oxide/hydroxide
on the surface of the biochar. Furthermore, as illustrated by the curvature of each respec-
tive three-dimensional response surface plot presented in Figure 1b,c, the As(III) adsorp-
tion onto Al-FWB was dependent on the Al content and was enhanced by the increase in
Al content.
Water 2022, 14, 2712 7 of 19

Figure 1. Estimated response surface for the amount of As(III) adsorbed onto a unit mass of Al-
modified food waste biochar (AL-FWB), showing the effect of (a) contact time (h) and temperature
(°C), (b) Al content (w/w %) and contact time (h), and (c) Al content (w/w %) and temperature (°C).
Water 2022, 14, 2712 8 of 19

The optimum conditions for the synthesis of the with maximized As(III) adsorption
were tracked using the Box–Behnken model of the RSM. The highest As(III) adsorption
(31.1 mg/g) occurred under the conditions of 3.5 h pyrolysis time, 468.36 °C pyrolysis tem-
perature, and 6 w/w % Al content. Further experiments were performed using Al-FWB
synthesized under optimized conditions.

3.2. Physical and Chemical Characteristics of Al-FWB


The surface morphology of Al-FWB was analyzed using FE-SEM, as shown in Figure
2. Al-FWB showed a rough surface with the clumps of Al (hydr)oxide formed by Al im-
pregnation on the Al-FWB. The presence of Al (hydr)oxide on the surface of the biochar
is identified by the results of energy dispersive spectroscopy (EDS). As shown in Table 3,
9.2% of Al was present on the surface of Al-FWB. The Al amount of Al-FWB analyzed
using XRF also showed a similar result. The Al amount analyzed using EDS and XRF was
higher than the Al amount mixed with the food waste (6 w/w%). This result can be ex-
plained by the fact that the carbon and water were lost from Al-FWB during the pyrolysis
while the Al amount was maintained. The Cl on the surface of the Al-FWB originated from
the AlCl3 and was used to modify the adsorbent. The XRD analysis of Al-FWB performed
at 5°–90° did not show any distinct peaks related to Al (Figure 3), because Al is present as
an amorphous form on the surface of Al-FWB. Previous studies [53,54] also reported the
metals impregnated in the biochar were present in the amorphous form via metallic com-
plexes on the biochar, and the XRD analysis showed no distinct peaks but broad humps.
The BET-calculated surface area and pore volume of the Al-FWB were 3.74 m2/g and
0.0087 cm3/g, respectively. The average pore diameter of the Al-FWB was 9.32 nm, corre-
sponding to a mesopore (2 < diameter < 50 nm) classified by the International Union of
Pure and Applied Chemistry [55]. The functional groups on the surface of the Al-FWB
were investigated by FTIR spectra (Figure 4). A single bond such as O–H (3350 cm−1) and
C–O–C (1000–1100 cm−1) stretching vibrations is attributed to functional groups of hemi-
cellulose and cellulose [56]. The peaks at 1918 and 670 cm−1 were attributed to the C-H
bending of the aromatic compound, and the peak at 1605 cm−1 corresponded to C=C-C of
the aromatic ring stretch [57].

Figure 2. Surface morphologies of Al-FWB analyzed using a field emission scanning electron micro-
scope (FE-SEM) under the conditions of accelerating voltage of 15.0 kV, magnification of ×10,000,
and scale bar and 5.00 μm.
Water 2022, 14, 2712 9 of 19

Table 3. Elemental composition (weight%) of Al-FWB analyzed using an energy dispersive spec-
trometer (EDS) and an X-ray fluorescence (XRF) spectrometer.

C O Al Cl Ca P K Na
EDS 46.8 ± 12.0 15.9 ± 6.7 9.2 ± 3.2 16.3 ± 10.2 8.6 ± 4.4 1.5 ± 0.9 1.8 ± 1.6 0.4 ± 0.1
XRF 41.2 21.5 11.0 12.4 8.6 1.6 1.6 1.3

Figure 3. Fourier transform infrared spectroscopy spectra of Al-FWB.

Figure 4. X-ray diffraction patterns of Al-FWB.


Water 2022, 14, 2712 10 of 19

3.3. Effect of Solution pH


The influence of the solution pH on As(III) removal by Al-FWB (Figure 5) was evi-
dent as the initial pH increased from 3 to 11. The effect of the solution pH on As(III) ad-
sorption on Al-FWB was studied at an initial As(III) concentration of 100 mg/L and a pH
ranging from 3–11. As the initial solution pH increased from 3 to 11, the final pH increased
from 6.0 to 9.4, showing that Al-FWB had a buffering capacity during the adsorption pro-
cess. This buffering effect can be due to the amphoteric nature of the Al-FWB [52]. As
depicted in Figure 5, Al-FWB showed a higher efficiency for As(III) adsorption at higher
pH values than at lower pH values. Adsorption increased significantly at pH 3 and
reached a maximum at pH 11. The adsorption of As(III) onto the biochar surface increased
from 13.5 to 16.2 mg/g through the increase of the initial pH from pH 3 to pH 5. A subse-
quent increase in pH revealed an even higher adsorption capacity, approaching the max-
imum of 18.5 mg/g at pH 11. As(III) species exist predominantly as H3AsO30 in the pH
range of 2.0–9.2, but H2AsO3− is the dominant form at pH > 9. Therefore, the low adsorp-
tion observed in the acidic pH range was attributed to the As(III) species present at pH <
9; there were no electrostatic interactions between H3AsO30 and Al-FWB, making the ad-
sorption very weak [58]. The pH of the solution could significantly alter the speciation of
As and the composition of the adsorbent surface. H2AsO3−, the dominant form above pH
9, was preferred for ionic exchange, favoring greater As(Ⅲ) adsorption [59,60].
Owing to the heterogeneous surface of the Al-modified adsorbent, the active sites of
the adsorbent for As(III) adsorption had a partially positive charge. Furthermore, H2AsO3−
could be coordinated to the surface of adsorbents, and the adsorption energy was suffi-
ciently large to overcome acid dissociation [61]. Again, as the pH of the solution increased,
the adsorption capacity also increased, and the maximum As(III) uptake occurred at pH
11, proving the variation of As(III) from neutral to negatively charged species. This find-
ing is consistent with those of previous studies investigating the adsorption of As(III) [62–
64].

Figure 5. Effect of initial solution pH on the As(III) adsorption at 100 mg/L As(III) concentration in
different pH conditions.

3.4. Effect of Competitive Anions


In real water systems, some anions might be present in the surface water and compete
with As(III) for the active sites during adsorption [65,66]. In order to study the influence
Water 2022, 14, 2712 11 of 19

of competing anions on As(III) adsorption onto Al-FWB, the co-existence of anions such
as NO3−, HCO3−, SO42−, and PO43−, which are suspected of having significant impacts on
arsenic adsorption [58], were investigated by mixing 30 mL As(III) solution with the initial
concentration of 100 mg/L and 1 mM and 10 mM of different anions in 50 mL conical
tubes. As shown in Figure 6, the effect of 1 mM NO3− on the adsorption of As(III) onto Al-
FWB was insignificant, irrespective of the initial concentration, as As(III) is known to form
weak bonds with NO3−. This finding is consistent with a previous study that reported an
insignificant effect of NO3− on As(III) adsorption [67]. In the case of HCO3−, PO43−, and
SO42−, the amount of As(III) adsorbed decreased from 16.3 mg/g to 13.8 mg/g, 14.4 mg/g,
and 15.3 mg/g at an anionic concentration of 1 mM and 7.6 mg/g, 13.7 mg/g, and 14.5 mg/g
at an anionic concentration of 10 mM, respectively. Therefore, HCO3−, PO43−, and SO42−
considerably affected As(III) adsorption in the following order: HCO3− > PO43− > SO42− >
NO3− (Figure 6). Kong, et al. [68] reported similar results and attributed the effect to the
formation of arseno-carbonate complexes, such as As(CO3)2−, As(CO3)(OH)22−, and AsCO3⁺,
which prevents As(III) adsorption onto the adsorbent surface. The PO43− and SO42− inhib-
ited the As(III) adsorption by forming a complex with Al present on the surface of Al-
FWB [69]. In our previous study [52], the impact of anions on fluoride adsorption by Al-
FWB followed the order of PO43− > SO42− > HCO3− > NO3−. The higher impact of HCO3− on
the As(III) adsorption by Al-FWB than fluoride adsorption also supports the formation of
arseno-carbonate complexes.

Figure 6. The effect of different concentrations of co-existing anions on As(III) adsorption onto Al-
FWB at a 100 mg/L initial As(III) concentration.

3.5. Effect of Adsorbent Dosage


The influence of the adsorbent dosage on As(III) adsorption was examined by vary-
ing the amount of adsorbent (1.7–13.33 g/L) at initial arsenic concentrations of 100 and 300
mg/L. The maximum As(III) capacities of 1.7 g/L Al-FWB were 29.99 mg/g and 130.84
mg/g for 100 and 300 mg/L As(III), respectively (Figure 7a,b). The As(III) removal percent-
age increased with the increasing adsorbent dosage, and the maximum removal percent-
ages were 92.5% and 88.1% for 100 and 300 mg/L of As(III) initial concentrations, respec-
tively, at 13.3 g/L adsorbent dose (Figure 7a,b). The increase in the As(III) removal per-
centage with the addition of more adsorbents can be explained by the increase in the num-
ber of adsorption sites for a fixed As(Ⅲ) concentration [4,70]. The lower adsorption capac-
ity of the adsorbent at higher adsorbent dosages is probably due to the presence of un-
saturated adsorption active sites resulting from the increased adsorbent amount at a par-
ticular initial As(III) concentration and the interaction between adsorbent molecules (in-
termolecular attraction) [4,71].
Water 2022, 14, 2712 12 of 19

Figure 7. The effect of adsorbent dosage on As(III) adsorption on Al-FWB at (a) 100 mg/L and (b)
300 mg/L initial As(III) concentrations.

3.6. Adsorption Kinetics


The effect of the contact time on As(III) adsorption was studied over a time range of
0.5–168 h and an initial As(III) concentration of 100 mg/L. Thereafter, the kinetic data were
modeled using the pseudo-first order and pseudo-second order kinetic models (Equations
(2) and (3)). Insights into the nature of the adsorption process and kinetic parameters are
presented in Table 4 and Figure 8. Figure 8 shows that As(III) adsorption reached 90% of
the equilibrium concentration within 48 h at the 100 mg/L initial concentration. The deter-
mination coefficients (R2) for the pseudo-first order (R2 = 0.815) and pseudo-second order
(R2 = 0.925) models revealed that the experimental data of As(III) adsorption fit better to
the pseudo-second order model, suggesting that the chemisorption, which involves the
sharing of electrons between the absorbent and As(III), determines the rate of As(III) ad-
sorption [72–74]. The pseudo-second order model predicted the amount of As(III) ad-
sorbed onto Al-FWB at equilibrium to be 20.5 mg/g (Table 4), which is similar to the ex-
perimental As(III) adsorption amount at equilibrium (19.9 mg/g).
Water 2022, 14, 2712 13 of 19

Table 4. Kinetic model parameters of As(III) adsorption onto Al-FWB at a 100 mg/L initial concen-
tration.

Model Parameters
qe (mg/g) k1 (/h) R2
Pseudo-first order
19.1 0.298 0.815
qe (mg/g) k2 (g/mg/h) R2
Pseudo-second order
20.5 0.019 0.925

Figure 8. As(III) adsorption onto Al-FWB with respect to the reaction time and the kinetic model fits
for the experimental data.

3.7. Adsorption Equilibrium


Figure 9 shows the As(III) adsorption by Al-FWB under different initial As(III) con-
centrations (10–1000 mg/L). The adsorption increased considerably with increasing the
initial As(III) concentration up to 500 mg/L and almost reached a constant value above
500 mg/L. A decrease in the As(III) adsorption amount above 500 mg/L indicates the sat-
uration of the As(III) ion adsorption sites on the surface of Al-FWB [4,75]. The commonly
used equilibrium models, namely, the Langmuir and Freundlich models (Equations (4)
and (5)), were used to model the experimental data. Table 5 shows that both equilibrium
models fit the experimental data fairly well, as their corresponding determination coeffi-
cients (R2) were 0.990 and 0.964 for the Langmuir and Freundlich models, respectively.
However, the Langmuir model with a higher R2 value (0.990) showed a better fit than the
Freundlich model, indicating that As(Ⅲ) was adsorbed onto the surface of the Al-FWB as
a homogeneous monolayer [4,76].

Table 5. Equilibrium models fitted As(III) adsorption onto Al-FWB.

Model Parameter
Qm (mg/g) KL (L/mg) R2
Langmuir
52.2 0.0186 0.990
KF (L/g) 1/n R2
Freundlich
6.1475 0.325 0.964
Water 2022, 14, 2712 14 of 19

Figure 9. Equilibrium models of As(III) adsorption onto Al-FWB with initial As(III) concentrations
of 10–1000 mg/L at 25 °C and 100 rpm agitation for 24 h.

As a waste-derived adsorbent, Al-FWB showed a high adsorption capacity compared


with other adsorbents reported in the literature (Table 6). The maximum As(III) adsorp-
tion capacity of Al-FWB was 52.2 mg/g. The granular nature of Al-FWB is also an im-
portant property for As (Ⅲ) removal because it allows for easy separation after adsorption
and for use as a filter medium in a fixed bed reactor. Furthermore, Al-FWB is a good sub-
stitute for removing As(III) in real water systems by utilizing waste as a resource for pro-
ducing new products to maximize environmental sustainability.

Table 6. Comparison of the adsorption capacity of different adsorbents for As(III) removal.

Adsorption Capacity
Adsorbent pH Temperature Reference
(mg/g)
Fe-Mn binary oxide
296.23 7 25 °C [68]
zeolite
Fe impregnated
119.5 7.0 25.0 [51]
food waste biocahr
α—Fe2O3/MCM-41 102.1 8 25 °C [77]
Al-FWB 52.2 7 25 °C This study
Fe coated empty
fruit bunch and rice 30.7–31.4 8 25 ± 1 °C [64]
husk biochar
ZnCl2-biogas resi-
27.67 7±1 25 ± 1 °C [78]
due biochar
Magnetic pine cone
18.02 7 25.85 °C [79]
biomass
Fe-Mn-La-biochar
14.9 3–7 25 °C [76]
composite
Almond shell bio-
4.86 ~7.2 20 ± 2 °C [4]
char
Cerium-loaded cat-
2.5297 6 25 °C [65]
ion exchange resin
Synthetic Kan grass
2.004 7 25 °C [16]
biochar
Ca-Jatrofa charcoal 0.002 10 28 ± 5 °C [80]
GAC-Cu 0.0144 9–11 29 ± 1 °C [62]
Water 2022, 14, 2712 15 of 19

3.8. Adsorption Thermodynamics


Changes in free energy (∆G°), enthalpy (∆H°), and entropy (∆S°) during the adsorp-
tion of As(III) on Al-FWB are presented in Table 7. The negative value of (∆G°) for all three
temperatures indicated that the adsorption of As(III) on Al-FWB was spontaneous and
feasible [81]. The positive value of the enthalpy (∆H°) indicated that the process was en-
dothermic [82], while the positive value of entropy (∆S°) indicated that the randomness at
the solid–solution interface increased during the adsorption of As(III) on Al-FWB.

Table 7. The estimated thermodynamic parameters for the adsorption of As(III) on Al-FWB.

Temperature ∆H° ∆S° ∆G°


°C (kJ/mol) (J/K mol) (kJ/mol)
15 5.1 31.6 −3.97
25 −4.29
35 −4.61

4. Conclusions
RSM was employed to optimize the synthesis conditions, including the pyrolysis
time, temperature, and Al content, for preparing Al-FWB. The pyrolysis temperature and
Al content remarkably influenced the As(III) adsorption capacity of Al-FWB, but the effect
of the pyrolysis time was not statistically significant. Al-FWB showed the highest adsorp-
tion capacity under the conditions of 3.5 h pyrolysis time, 468.36 °C pyrolysis temperature,
and 6 w/w % of Al content. The characteristics of As(III) adsorption on the optimized Al-
FWB were investigated by varying the solution pH, presence of competitive anions, reac-
tion time, initial As(III) concentration, and reaction temperature. The increase in solution
pH enhanced As (Ⅲ) adsorption to Al-FWB because of the change in As(III) species to
H2AsO3− at pH > 9.2. As(Ⅲ) adsorption was reduced in the presence of anions; their im-
pacts followed the order: HCO3− > HPO42− > SO42− > NO3−. Kinetic adsorption experiments
using the model analysis showed that As(III) adsorption was mainly controlled by chem-
isorption. The better fit of the Langmuir model for As(III) adsorption than the Freundlich
model revealed monolayer As(Ⅲ) adsorption to Al-FWB, and its maximum adsorption
capacity was 52.2 mg/g, which is superior to that of other adsorbents reported in the liter-
ature. The adsorption of As(III) to Al-FWB was endothermic and spontaneous under ex-
perimental conditions. Al-FWB is a promising adsorbent for removing As(III) from aque-
ous solutions and managing FW.

Author Contributions: S.Q.H.: methodology and writing—original draft preparation; S.-H.H.:


methodology and formal analysis; C.-G.L.: writing, review, and editing; S.-J.P.: conceptualization,
data curation, and writing—original draft preparation, review, and editing. All authors have read
and agreed to the published version of the manuscript.
Funding: This work was supported by Korea Environment Industry and Technology Institute
(KEITI) through the Aquatic Ecosystem Conservation Research Program, funded by the Korea Min-
istry of Environment (grant number: RE202201970).
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: The data presented in this study are available on request from the
corresponding author.
Acknowledgments: This work was supported by Korea Environment Industry & Technology Insti-
tute (KEITI) through Aquatic Ecosystem Conservation Research Program, funded by Korea Minis-
try of Environment (grant number: RE202201970)
Conflicts of Interest: The authors declare no conflict of interest.
Water 2022, 14, 2712 16 of 19

References
1. Meharg, A.A.; Hartley-Whitaker, J. Arsenic uptake and metabolism in arsenic resistant and nonresistant plant species. New
Phytol. 2002, 154, 29–43. https://doi.org/10.1046/j.1469-8137.2002.00363.x.
2. Aredes, S.; Klein, B.; Pawlik, M. The removal of arsenic from water using natural iron oxide minerals. J. Clean. Prod. 2013, 60,
71–76.
3. Hao, L.; Zheng, T.; Jiang, J.; Zhang, G.; Wang, P. Removal of As (III) and As (V) from water using iron doped amino function-
alized sawdust: Characterization, adsorptive performance and UF membrane separation. Chem. Eng. J. 2016, 292, 163–173.
4. Ali, S.; Rizwan, M.; Shakoor, M.B.; Jilani, A.; Anjum, R. High sorption efficiency for As(III) and As(V) from aqueous solutions
using novel almond shell biochar. Chemosphere 2020, 243, 125330. https://doi.org/10.1016/j.chemosphere.2019.125330.
5. Wang, S.; Mulligan, C.N. Natural attenuation processes for remediation of arsenic contaminated soils and groundwater. J. Haz-
ard. Mater. 2006, 138, 459–470. https://doi.org/10.1016/j.jhazmat.2006.09.048.
6. Oremland, R.S.; Stolz, J.F. The Ecology of Arsenic. Science 2003, 300, 939. https://doi.org/10.1126/science.1081903.
7. Holm, T. Effects of Co 32-/bicarbonate, Si, and Po 43- on arsenic sorption to HFO. J. Am. Water Work. Assoc. 2002, 94, 174–181.
8. Bordoloi, S.; Nath, S.K.; Gogoi, S.; Dutta, R.K. Arsenic and iron removal from groundwater by oxidation–coagulation at opti-
mized pH: Laboratory and field studies. J. Hazard. Mater. 2013, 260, 618–626. https://doi.org/10.1016/j.jhazmat.2013.06.017.
9. Kumar, N.S.; Goel, S. Factors influencing arsenic and nitrate removal from drinking water in a continuous flow electrocoagula-
tion (EC) process. J Hazard Mater 2010, 173, 528–533. https://doi.org/10.1016/j.jhazmat.2009.08.117.
10. Choong, T.S.Y.; Chuah, T.G.; Robiah, Y.; Gregory Koay, F.L.; Azni, I. Arsenic toxicity, health hazards and removal techniques
from water: An overview. Desalination 2007, 217, 139–166. https://doi.org/10.1016/j.desal.2007.01.015.
11. Harper, T.R.; Kingham, N.W. Removal of arsenic from wastewater using chemical precipitation methods. Water Environ. Res.
1992, 64, 200–203. https://doi.org/10.2175/WER.64.3.2.
12. Cui, J.; Jing, C.; Che, D.; Zhang, J.; Duan, S. Groundwater arsenic removal by coagulation using ferric (III) sulfate and polyferric
sulfate: A comparative and mechanistic study. J. Environ. Sci. 2015, 32, 42–53.
13. Glass, S.; Mantel, T.; Appold, M.; Sen, S.; Usman, M.; Ernst, M.; Filiz, V. Amine-terminated PAN membranes as anion-adsorber
materials. Chem. Ing. Tech. 2021, 93, 1396–1400.
14. Yoon, J.; Amy, G.; Chung, J.; Sohn, J.; Yoon, Y. Removal of toxic ions (chromate, arsenate, and perchlorate) using reverse osmo-
sis, nanofiltration, and ultrafiltration membranes. Chemosphere 2009, 77, 228–235. https://doi.org/10.1016/j.chemo-
sphere.2009.07.028.
15. Navarathna, C.M.; Karunanayake, A.G.; Gunatilake, S.R.; Pittman, C.U.; Perez, F.; Mohan, D.; Mlsna, T. Removal of Arsenic(III)
from water using magnetite precipitated onto Douglas fir biochar. J. Environ. Manag. 2019, 250, 109429.
https://doi.org/10.1016/j.jenvman.2019.109429.
16. Baig, S.A.; Zhu, J.; Muhammad, N.; Sheng, T.; Xu, X. Effect of synthesis methods on magnetic Kans grass biochar for enhanced
As(III, V) adsorption from aqueous solutions. Biomass Bioenergy 2014, 71, 299–310. https://doi.org/10.1016/j.biom-
bioe.2014.09.027.
17. Alaei Shahmirzadi, M.A.; Hosseini, S.S.; Luo, J.; Ortiz, I. Significance, evolution and recent advances in adsorption technology,
materials and processes for desalination, water softening and salt removal. J. Environ. Manag. 2018, 215, 324–344.
https://doi.org/10.1016/j.jenvman.2018.03.040.
18. Gupta, V.K.; Ali, I. Chapter 2—Water Treatment for Inorganic Pollutants by Adsorption Technology. In Environmental Water,
Gupta, V.K., Ali, I., Eds.; Elsevier: Amsterdam, The Netherlands, 2013; pp. 29-91.
19. Skoczko, I.; Szatyłowicz, E. Removal of heavy metal ions by filtration on activated alumina-assisted magnetic field. Desalination
Water Treat. 2018, 117, 345–352.
20. Maiti, A.; DasGupta, S.; Basu, J.K.; De, S. Batch and Column Study:  Adsorption of Arsenate Using Untreated Laterite as Adsor-
bent. Ind. Eng. Chem. Res. 2008, 47, 1620–1629. https://doi.org/10.1021/ie070908z.
21. Usman, M.; Zarebanadkouki, M.; Waseem, M.; Katsoyiannis, I.A.; Ernst, M. Mathematical modeling of arsenic (V) adsorption
onto iron oxyhydroxides in an adsorption-submerged membrane hybrid system. J Hazard Mater 2020, 400, 123221.
22. Medpelli, D.; Sandoval, R.; Sherrill, L.; Hristovski, K.; Seo, D.-K. Iron oxide-modified nanoporous geopolymers for arsenic re-
moval from ground water. Resour. -Effic. Technol. 2015, 1, 19–27. https://doi.org/10.1016/j.reffit.2015.06.007.
23. Lin, T.-F.; Wu, J.-K. Adsorption of Arsenite and Arsenate within Activated Alumina Grains: Equilibrium and Kinetics. Water
Res. 2001, 35, 2049–2057. https://doi.org/10.1016/S0043-1354(00)00467-X.
24. Chen, W.; Parette, R.; Zou, J.; Cannon, F.S.; Dempsey, B.A. Arsenic removal by iron-modified activated carbon. Water Res. 2007,
41, 1851–1858. https://doi.org/10.1016/j.watres.2007.01.052.
25. Asere, T.G.; Stevens, C.V.; Du Laing, G. Use of (modified) natural adsorbents for arsenic remediation: A review. Sci. Total Envi-
ron. 2019, 676, 706–720. https://doi.org/10.1016/j.scitotenv.2019.04.237.
26. Rahim, M.; Mas Haris, M.R.H. Application of biopolymer composites in arsenic removal from aqueous medium: A review. J.
Radiat. Res. Appl. Sci. 2015, 8, 255–263. https://doi.org/10.1016/j.jrras.2015.03.001.
27. Rajapaksha, A.U.; Chen, S.S.; Tsang, D.C.W.; Zhang, M.; Vithanage, M.; Mandal, S.; Gao, B.; Bolan, N.S.; Ok, Y.S. Engineered/de-
signer biochar for contaminant removal/immobilization from soil and water: Potential and implication of biochar modification.
Chemosphere 2016, 148, 276–291. https://doi.org/10.1016/j.chemosphere.2016.01.043.
Water 2022, 14, 2712 17 of 19

28. Faria, W.M.; Figueiredo, C.C.d.; Coser, T.R.; Vale, A.T.; Schneider, B.G. Is sewage sludge biochar capable of replacing inorganic
fertilizers for corn production? Evidence from a two-year field experiment. Arch. Agron. Soil Sci. 2018, 64, 505–519.
https://doi.org/10.1080/03650340.2017.1360488.
29. Dong, C.-D.; Chen, C.-W.; Kao, C.-M.; Chien, C.-C.; Hung, C.-M. Wood-Biochar-Supported Magnetite Nanoparticles for Reme-
diation of PAH-Contaminated Estuary Sediment. Catalysts 2018, 8, 73.
30. Dunnigan, L.; Ashman, P.J.; Zhang, X.; Kwong, C.W. Production of biochar from rice husk: Particulate emissions from the
combustion of raw pyrolysis volatiles. J. Clean. Prod. 2018, 172, 1639–1645. https://doi.org/10.1016/j.jclepro.2016.11.107.
31. Liu, X.; Zhang, Y.; Li, Z.; Feng, R.; Zhang, Y. Characterization of corncob-derived biochar and pyrolysis kinetics in comparison
with corn stalk and sawdust. Bioresour. Technol. 2014, 170, 76–82. https://doi.org/10.1016/j.biortech.2014.07.077.
32. Jouiad, M.; Al-Nofeli, N.; Khalifa, N.; Benyettou, F.; Yousef, L.F. Characteristics of slow pyrolysis biochars produced from
rhodes grass and fronds of edible date palm. J. Anal. Appl. Pyrolysis 2015, 111, 183–190. https://doi.org/10.1016/j.jaap.2014.10.024.
33. Chen, G.-J.; Peng, C.-Y.; Fang, J.-Y.; Dong, Y.-Y.; Zhu, X.-H.; Cai, H.-M. Biosorption of fluoride from drinking water using spent
mushroom compost biochar coated with aluminum hydroxide. Desalination Water Treat. 2016, 57, 12385–12395.
https://doi.org/10.1080/19443994.2015.1049959.
34. Wang, J.; Chen, N.; Li, M.; Feng, C. Efficient Removal of Fluoride Using Polypyrrole-Modified Biochar Derived from Slow
Pyrolysis of Pomelo Peel: Sorption Capacity and Mechanism. J. Polym. Environ. 2018, 26, 1559–1572.
https://doi.org/10.1007/s10924-017-1061-y.
35. Chen, B.; Chen, Z. Sorption of naphthalene and 1-naphthol by biochars of orange peels with different pyrolytic temperatures.
Chemosphere 2009, 76, 127–133. https://doi.org/10.1016/j.chemosphere.2009.02.004.
36. Kyi, P.P.; Quansah, J.O.; Lee, C.-G.; Moon, J.-K.; Park, S.-J. The removal of crystal violet from textile wastewater using palm
kernel shell-derived biochar. Appl. Sci. 2020, 10, 2251.
37. Usman, A.R.A.; Abduljabbar, A.; Vithanage, M.; Ok, Y.S.; Ahmad, M.; Ahmad, M.; Elfaki, J.; Abdulazeem, S.S.; Al-Wabel, M.I.
Biochar production from date palm waste: Charring temperature induced changes in composition and surface chemistry. J.
Anal. Appl. Pyrolysis 2015, 115, 392–400. https://doi.org/10.1016/j.jaap.2015.08.016.
38. Zhao, L.; Cao, X.; Mašek, O.; Zimmerman, A. Heterogeneity of biochar properties as a function of feedstock sources and pro-
duction temperatures. J. Hazard. Mater. 2013, 256–257, 1–9. https://doi.org/10.1016/j.jhazmat.2013.04.015.
39. Gupta, S.; Kua, H.W.; Koh, H.J. Application of biochar from food and wood waste as green admixture for cement mortar. Sci.
Total Environ. 2018, 619–620, 419–435. https://doi.org/10.1016/j.scitotenv.2017.11.044.
40. Lee, C.-G.; Hong, S.-H.; Hong, S.-G.; Choi, J.-W.; Park, S.-J. Production of Biochar from Food Waste and its Application for
Phenol Removal from Aqueous Solution. Water Air Soil Pollut. 2019, 230, 70. https://doi.org/10.1007/s11270-019-4125-x.
41. Oh, S.-Y.; Seo, Y.-D. Sorption of halogenated phenols and pharmaceuticals to biochar: Affecting factors and mechanisms. Envi-
ron. Sci. Pollut. Res. 2016, 23, 951–961. https://doi.org/10.1007/s11356-015-4201-8.
42. Essandoh, M.; Kunwar, B.; Pittman, C.U.; Mohan, D.; Mlsna, T. Sorptive removal of salicylic acid and ibuprofen from aqueous
solutions using pine wood fast pyrolysis biochar. Chem. Eng. J. 2015, 265, 219–227. https://doi.org/10.1016/j.cej.2014.12.006.
43. Sizmur, T.; Fresno, T.; Akgül, G.; Frost, H.; Moreno-Jiménez, E. Biochar modification to enhance sorption of inorganics from
water. Bioresour. Technol. 2017, 246, 34–47. https://doi.org/10.1016/j.biortech.2017.07.082.
44. Lou, K.; Rajapaksha, A.U.; Ok, Y.S.; Chang, S.X. Pyrolysis temperature and steam activation effects on sorption of phosphate
on pine sawdust biochars in aqueous solutions. Chem. Speciat. Bioavailab. 2016, 28, 42–50.
https://doi.org/10.1080/09542299.2016.1165080.
45. Jin, H.; Capareda, S.; Chang, Z.; Gao, J.; Xu, Y.; Zhang, J. Biochar pyrolytically produced from municipal solid wastes for aque-
ous As(V) removal: Adsorption property and its improvement with KOH activation. Bioresour. Technol. 2014, 169, 622–629.
https://doi.org/10.1016/j.biortech.2014.06.103.
46. Agrafioti, E.; Kalderis, D.; Diamadopoulos, E. Ca and Fe modified biochars as adsorbents of arsenic and chromium in aqueous
solutions. J. Environ. Manag. 2014, 146, 444–450. https://doi.org/10.1016/j.jenvman.2014.07.029.
47. Chen, L.; Chen, X.L.; Zhou, C.H.; Yang, H.M.; Ji, S.F.; Tong, D.S.; Zhong, Z.K.; Yu, W.H.; Chu, M.Q. Environmental-friendly
montmorillonite-biochar composites: Facile production and tunable adsorption-release of ammonium and phosphate. J. Clean.
Prod. 2017, 156, 12–659. https://doi.org/10.1016/j.jclepro.2017.04.050.
48. Kang, J.-K.; Seo, E.-J.; Lee, C.-G.; Jeong, S.; Park, S.-J. Application of response surface methodology and artificial neural network
for the preparation of Fe-loaded biochar for enhanced Cr (VI) adsorption and its physicochemical properties and Cr (VI) ad-
sorption characteristics. Environ. Sci. Pollut. Res. 2022, 1–15.
49. Kim, Y.S.; Jang, J.Y.; Park, S.J.; Um, B.H. Dilute sulfuric acid fractionation of Korean food waste for ethanol and lactic acid
production by yeast. Waste Manag. 2018, 74, 231–240. https://doi.org/10.1016/j.wasman.2018.01.012.
50. Kang, J.-K.; Seo, E.-J.; Lee, C.-G.; Park, S.-J. Fe-loaded biochar obtained from food waste for enhanced phosphate adsorption
and its adsorption mechanism study via spectroscopic and experimental approach. J. Environ. Chem. Eng. 2021, 9, 105751.
51. Lyonga, F.N.; Hong, S.-H.; Cho, E.-J.; Kang, J.-K.; Lee, C.-G.; Park, S.-J. As (III) adsorption onto Fe-impregnated food waste
biochar: Experimental investigation, modeling, and optimization using response surface methodology. Environ. Geochem. Health
2021, 43, 3303–3321.
52. Meilani, V.; Lee, J.-I.; Kang, J.-K.; Lee, C.-G.; Jeong, S.; Park, S.-J. Application of aluminum-modified food waste biochar as
adsorbent of fluoride in aqueous solutions and optimization of production using response surface methodology. Microporous
Mesoporous Mater. 2021, 312, 110764.
Water 2022, 14, 2712 18 of 19

53. Liu, Q.; Wu, L.; Gorring, M.; Deng, Y. Aluminum-impregnated biochar for adsorption of arsenic (V) in urban stormwater runoff.
J. Environ. Eng. 2019, 145, 04019008.
54. Kim, J.; Song, J.; Lee, S.-M.; Jung, J. Application of iron-modified biochar for arsenite removal and toxicity reduction. J. Ind. Eng.
Chem. 2019, 80, 17–22.
55. Pradhan, B.K.; Sandle, N. Effect of different oxidizing agent treatments on the surface properties of activated carbons. Carbon
1999, 37, 1323–1332.
56. Saqib, N.U.; Baroutian, S.; Sarmah, A.K. Physicochemical, structural and combustion characterization of food waste hydrochar
obtained by hydrothermal carbonization. Bioresour Technol 2018, 266, 357–363.
57. Coates, J. Interpretation of infrared spectra, a practical approach. In Encyclopedia of Analytical Chemistry; John Wiley & Sons Ltd,:
Hoboken, NJ, USA, 2000.
58. Shakoor, M.B.; Niazi, N.K.; Bibi, I.; Murtaza, G.; Kunhikrishnan, A.; Seshadri, B.; Shahid, M.; Ali, S.; Bolan, N.S.; Ok, Y.S.; et al.
Remediation of arsenic-contaminated water using agricultural wastes as biosorbents. Crit. Rev. Environ. Sci. Technol. 2016, 46,
467–499. https://doi.org/10.1080/10643389.2015.1109910.
59. Sharma, V.K.; Sohn, M. Aquatic arsenic: Toxicity, speciation, transformations, and remediation. Environ. Int. 2009, 35, 743–759.
https://doi.org/10.1016/j.envint.2009.01.005.
60. Wei, Z.; Liang, K.; Wu, Y.; Zou, Y.; Zuo, J.; Arriagada, D.C.; Pan, Z.; Hu, G. The effect of pH on the adsorption of arsenic(III)
and arsenic(V) at the TiO2 anatase [101] surface. J. Colloid Interface Sci. 2016, 462, 252–259.
https://doi.org/10.1016/j.jcis.2015.10.018.
61. Pierce, M.L.; Moore, C.B. Adsorption of arsenite and arsenate on amorphous iron hydroxide. Water Res. 1982, 16, 1247–1253.
62. Mondal, P.; Mohanty, B.; Balomajumder, C. Treatment of simulated arsenic contaminated groundwater using GAC-Cu in batch
reactor: Optimization of process parameters. Can. J. Chem. Eng. 2009, 87, 766–778. https://doi.org/10.1002/cjce.20214.
63. Polowczyk, I.; Bastrzyk, A.; Ulatowska, J.; Szczałba, E.; Koźlecki, T.; Sadowski, Z. Influence of pH on arsenic(III) removal by fly
ash. Sep. Sci. Technol. 2016, 51, 2612–2619. https://doi.org/10.1080/01496395.2016.1163610.
64. Samsuri, A.W.; Sadegh-Zadeh, F.; Seh-Bardan, B.J. Adsorption of As(III) and As(V) by Fe coated biochars and biochars produced
from empty fruit bunch and rice husk. J. Environ. Chem. Eng. 2013, 1, 981–988. https://doi.org/10.1016/j.jece.2013.08.009.
65. He, Z.; Tian, S.; Ning, P. Adsorption of arsenate and arsenite from aqueous solutions by cerium-loaded cation exchange resin.
J. Rare Earths 2012, 30, 563–572. https://doi.org/10.1016/S1002-0721(12)60092-1.
66. Medina-Ramirez, A.; Gamero, P.; Ruiz-Camacho, B.; Minchaca-Mojica, J.; Romero, R.; Gamero-Vega, K. Adsorption of Aqueous
As (III) in Presence of Coexisting Ions by a Green Fe-Modified W Zeolite. Water 2019, 11, 281. https://doi.org/10.3390/w11020281.
67. Bekkouche, S.; Baup, S.; Bouhelassa, M.; Molina-Boisseau, S.; Petrier, C. Competitive adsorption of phenol and heavy metal ions
onto titanium dioxide (Dugussa P25). Desalination Water Treat. 2012, 37, 364–372.
68. Kong, S.; Wang, Y.; Zhan, H.; Yuan, S.; Yu, M.; Liu, M. Adsorption/Oxidation of arsenic in groundwater by nanoscale Fe-Mn
binary oxides loaded on zeolite. Water Env. Res 2014, 86, 147–155. https://doi.org/10.2175/106143013x13807328849170.
69. Sparks, D.L. Sorption phenomena on soils. Environ. Soil Chem. 1995, 99–115.
70. Lalhmunsiama; Tiwari, D.; Lee, S.-M. Activated Carbon and Manganese Coated Activated Carbon Precursor to Dead Biomass
in the Remediation of Arsenic Contaminated Water. Environ. Eng. Res. 2012, 17, 41–48. https://doi.org/10.4491/eer.2012.17.S1.S41.
71. Kazi, T.G.; Brahman, K.D.; Baig, J.A.; Afridi, H.I. A new efficient indigenous material for simultaneous removal of fluoride and
inorganic arsenic species from groundwater. J. Hazard. Mater. 2018, 357, 159–167. https://doi.org/10.1016/j.jhazmat.2018.05.069.
72. Zhang, G.; Ren, Z.; Zhang, X.; Chen, J. Nanostructured iron(III)-copper(II) binary oxide: A novel adsorbent for enhanced arsenic
removal from aqueous solutions. Water Res. 2013, 47, 4022–4031. https://doi.org/10.1016/j.watres.2012.11.059.
73. Bakshi, S.; Banik, C.; Rathke, S.J.; Laird, D.A. Arsenic sorption on zero-valent iron-biochar complexes. Water Res. 2018, 137, 153–
163. https://doi.org/10.1016/j.watres.2018.03.021.
74. Lin, L.; Qiu, W.; Wang, D.; Huang, Q.; Song, Z.; Chau, H.W. Arsenic removal in aqueous solution by a novel Fe-Mn modified
biochar composite: Characterization and mechanism. Ecotoxicol. Environ. Saf. 2017, 144, 514–521.
https://doi.org/10.1016/j.ecoenv.2017.06.063.
75. Bhatti, I.; Qureshi, K.; Kazi, R.A.; Ansari, Q. Preparation and characterization of chemically activated almond shells by optimi-
zation of adsorption parameter for the removal of Cr (VI) from aqueous solution. Int. J. Chem. Biomol. Eng. 2008, 1, 150–155.
76. Lin, L.; Song, Z.; Khan, Z.H.; Liu, X.; Qiu, W. Enhanced As(III) removal from aqueous solution by Fe-Mn-La-impregnated bio-
char composites. Sci. Total Environ. 2019, 686, 1185–1193. https://doi.org/10.1016/j.scitotenv.2019.05.480.
77. Boojari, H.; Pourafshari Chenar, M.; Pakizeh, M. Experimental Investigation of Arsenic (III, V) Removal from Aqueous Solution
Using Synthesized α-Fe2O3/MCM-41 Nanocomposite Adsorbent. Water Air Soil Pollut. 2016, 227, 290.
https://doi.org/10.1007/s11270-016-2989-6.
78. Xia, D.; Tan, F.; Zhang, C.; Jiang, X.; Chen, Z.; Li, H.; Zheng, Y.; Li, Q.; Wang, Y. ZnCl2-activated biochar from biogas residue
facilitates aqueous As(III) removal. Appl. Surf. Sci. 2016, 377, 361–369. https://doi.org/10.1016/j.apsusc.2016.03.109.
79. Pholosi, A.; Naidoo, E.B.; Ofomaja, A.E. Enhanced Arsenic (III) adsorption from aqueous solution by magnetic pine cone bio-
mass. Mater. Chem. Phys. 2019, 222, 20–30. https://doi.org/10.1016/j.matchemphys.2018.09.067.
80. Khatoon, N. Application of Calcium Impregnated Activated charcoal prepared from Jatropha Seed Residue for the removal of
Arsenic (iii) from Water. J. Environ. Res. Dev. 2013, 8, 196–205.
Water 2022, 14, 2712 19 of 19

81. Zhu, N.; Yan, T.; Qiao, J.; Cao, H. Adsorption of arsenic, phosphorus and chromium by bismuth impregnated biochar: Adsorp-
tion mechanism and depleted adsorbent utilization. Chemosphere 2016, 164, 32–40. https://doi.org/10.1016/j.chemo-
sphere.2016.08.036.
82. Zama, E.F.; Zhu, Y.-G.; Reid, B.J.; Sun, G.-X. The role of biochar properties in influencing the sorption and desorption of Pb(II),
Cd(II) and As(III) in aqueous solution. J. Clean. Prod. 2017, 148, 127–136. https://doi.org/10.1016/j.jclepro.2017.01.125.

You might also like