Next Article in Journal
Multiscale Modeling and Simulation of Polymer Blends in Injection Molding: A Review
Next Article in Special Issue
Preparation, Characterisation and Antibacterial Activity of Carvacrol Encapsulated in Gellan Gum Hydrogel
Previous Article in Journal
Self-Healing Hydrogels: Preparation, Mechanism and Advancement in Biomedical Applications
Previous Article in Special Issue
Cytocompatibility of Polymethyl Methacrylate Honeycomb-like Pattern on Perfluorinated Polymer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

An Insight into a Sustainable Removal of Bisphenol A from Aqueous Solution by Novel Palm Kernel Shell Magnetically Induced Biochar: Synthesis, Characterization, Kinetic, and Thermodynamic Studies

by
Kamil Kayode Katibi
1,2,
Khairul Faezah Yunos
3,*,
Hasfalina Che Man
2,
Ahmad Zaharin Aris
4,5,
Mohd Zuhair Mohd Nor
3 and
Rabaah Syahidah Azis
6,7
1
Department of Agricultural and Biological Engineering, Faculty of Engineering and Technology, Kwara State University, Malete 23431, Nigeria
2
Department of Biological and Agricultural Engineering, Faculty of Engineering, University Putra Malaysia, Serdang 43400, Selangor, Malaysia
3
Department of Food and Process Engineering, Faculty of Engineering, University Putra Malaysia, Serdang 43400, Selangor, Malaysia
4
Department of Environment, Faculty of Forestry and Environment, University Putra Malaysia, Serdang 43400, Selangor, Malaysia
5
Material Processing and Technology Laboratory (MPTL), Institute of Advance Technology (ITMA), University Putra Malaysia, Serdang 43400, Selangor, Malaysia
6
Department of Physics, Faculty of Science, University Putra Malaysia, Serdang 43400, Selangor, Malaysia
7
Materials Synthesis and Characterization Laboratory (MSCL), Institute of Advanced Technology (ITMA), University Putra Malaysia, Serdang 43400, Selangor, Malaysia
*
Author to whom correspondence should be addressed.
Polymers 2021, 13(21), 3781; https://doi.org/10.3390/polym13213781
Submission received: 28 September 2021 / Revised: 22 October 2021 / Accepted: 24 October 2021 / Published: 31 October 2021

Abstract

:
Recently Bisphenol A (BPA) is one of the persistent trace hazardous estrogenic contaminants in the environment, that can trigger a severe threat to humans and environment even at minuscule concentrations. Thus, this work focused on the synthesis of neat and magnetic biochar (BC) as a sustainable and inexpensive adsorbent to remove BPA from aqueous environment. Novel magnetic biochar was efficiently synthesized by utilizing palm kernel shell, using ferric chloride and ferrous chloride as magnetic medium via chemical co-precipitation technique. In this experimental study, the influence of operating factors comprising contact time (20–240 min), pH (3.0–12.0), adsorbent dose (0.2–0.8 g), and starting concentrations of BPA (8.0–150 ppm) were studied in removing BPA during batch adsorption system using neat biochar and magnetic biochar. It was observed that the magnetically loaded BC demonstrates superior maximum removal efficiency of BPA with 94.2%, over the neat biochar. The functional groups (FTIR), Zeta potential, vibrating sample magnetometer (VSM), surface and textural properties (BET), surface morphology, and mineral constituents (FESEM/EDX), and chemical composition (XRD) of the adsorbents were examined. The experimental results demonstrated that the sorption isotherm and kinetics were suitably described by pseudo-second-order model and Freundlich model, respectively. By studying the adsorption mechanism, it was concluded that π-π electron acceptor–donor interaction (EAD), hydrophobic interaction, and hydrogen bond were the principal drives for the adsorption of BPA onto the neat BC and magnetic BC.

1. Introduction

Recently, the safety and quality of drinking water has been problematic owing to rapid urban development induced by human activities. The upsurge of some micro-contaminants identified as endocrine disrupting compounds (EDCs) has elicited increasing concerns over the supply of safe and clean drinking water [1,2,3]. Particularly, Bisphenol A (BPA; 2,2-bis(4-hydroxyphenyl)propane) is an intermediate and highly essential raw material highly utilized in the manufacture of several polycarbonate plastics and epoxy resins (plastic component), food cans, polyester fibers, and thermal paper as well as other materials in industry, thus it frequently emerges in numerous products for daily use, such as electronic equipment, water-pipes, toys, or paper [4,5]. Globally, BPA is classified among the most ever-present and most extensively produced synthetic chemical compounds in manufacturing today, with more than three million tons generated annually [6]. Markedly, the manufacture and sale of plastics comprising BPA has been prohibited by the Brazilian government since 2012, owing to its environmental universality, elevated industrial output, and toxicological consequence, and it has been classified as an important contaminant in water purification in many regions and nations [7]. Furthermore, exposure to BPA is through tableware and polycarbonate bottles, including those employed for new-born formula milk, and by means of epoxy resin coatings inside beverage and food containers, dental fissure sealants, adhesives, epoxy-based surface coatings, canned goods, household dust, and printing inks [5,6] and massively released into the water environment [8]. BPA in plastic containers can hydrolyse and pollute groundwater via landfill leachate [9]. Risen plastic industries trigger profuse discharge of BPA into the environment and it is detected in drinking and surface water [10]. More unfortunately, BPA has undesirable ecological impacts that are associated with dysfunction of the hormone system in animals and humans, even at very minuscule concentration. Specifically, an estrogenic action of BPA can provoke cancer [11,12,13], neurological challenges, diabetes, tumors, obesity, immune effects, heart disease in humans, damaged reproductive function, disturbance of the normal hormone functions, and undesired physiological state in animals and humans, biomagnification, and bioaccumulation through food chain or food web in human beings [14,15]. Besides, the dysfunction and damages caused by BPA are long-term [16]. Therefore, releasing water-containing BPA into waterways without adequate treatment poses a damaging impact on the environment (humans and aquatic) [17]. Owing to the frequent extensive utilization of BPA-based products by humans with their associated negative impacts on public health and ecosystem, hence, the removal of BPA became unavoidable. Extensive efforts have been employed towards extenuating the undesirable effects and environmental hazards by reducing the contaminants’ concentration using various treatment technologies, including precipitation, coagulation, activated sludge, biological filter, and constructed wetlands, with inadequate effect on BPA removal [18,19,20,21]. Some relatively advanced treatment techniques, such as ozonation and advanced oxidation [22,23], Fenton oxidation [24], membrane technology [25,26], photocatalysis degradation [27], enzymatic degradation [28,29], and adsorption technique utilizing activated carbon [30,31,32] exhibit a good BPA removal from wastewater and drinking water. Yet, the above treatment approaches still have significant drawbacks, such as high capital and maintenance cost, complex treatment procedures with increased operating expenses, costly capital expenses (CAPEX), generation of toxic by-products which hamper their application in developing nations [33]. Noticeably, among the aforementioned water-treatment approaches, adsorption technique is found more suitable, environmentally friendly, relatively economical, robust, and simple, and it could efficiently be employed in large-scale applications devoid of generating by-products in the environment [34,35]. Adsorption method has attracted wide attention and has been extensively considered as a promising technology for removing BPA pollutant during the past two decades, due to its ease of configuration and application, high efficiency, insensitivity to poisonous compounds, low cost, and comparatively small footprint as compared to other water-treatment methods [31].
Several adsorbents are available, however, carbon materials have been reported to perform remarkably owing to their several benefits, such as enormous surface area, superior stability, and outstanding removal efficiency [36]. Activated carbons generated from biomass wastes are often used as adsorbent during the adsorption process for the removal of emerging pollutants [37,38], since it can efficiently and rapidly remove BPA as a result of its abundant surface area as well as adsorptive capacity; but, its preparation is highly expensive. Thus, various studies have considered reliable resources, low-cost, and environment-friendly materials as alternative adsorbents [38,39]. Thus far, traditional carbonaceous materials, such as BC, activated carbon, graphene oxide together with its derivates, and carbon nanotubes, have been studied as promising adsorbent materials for hydrophobic organic pollutants, including pharmaceutical compounds and EDCs [40]. Furthermore, biochar as an adsorbent has porous structure comparable with activated carbon, which is the most generally utilized and effective sorbent for the removal of various contaminants from water globally [41].
Biochar, a major pyrogenic by-product obtained from the complete or partial pyrolysis of naturally plenteous biomass under oxygen-deficient conditions [42], has several macropores, hence it could attain a superior adsorption capacity [43]. More importantly, BC exhibits superior unique properties, such as augmented surface functional groups, porous structure, abundant specific surface area, eco-friendly, low-cost, ample inherent mineral components, high cation exchange capacity, and, notably, is efficient in the removal of various hydrophobic and hydrophilic organic pollutants owing to its high aromaticity and hydrophobicity [40,44,45]. These properties have made BC to be the best and most promising precursor over other adsorbent materials for various contaminants removal. Synchronically, after pyrolysis, BC can be reapplied to sorb contaminants in water; thus, it has abundant utilization value [46]. However, the minuscule particle size of BC in addition to its lower density makes its regeneration, separation, and recovery more problematic after adsorption, and this could undermine its recycling capacity and industrial applications. In order to subdue these difficulties, few studies have considered synthesis of magnetically recyclable biochar (magnetic biochar) via the implantation of iron oxide (Fe3O4 and Fe2O3) [47]. For instance, Lu et al. [48] studied the removal of BPA using N-doped ulva prolifera (marine macroalgae). Over 90% of BPA was eliminated and the sorption capacity of 9.19 mg/g was attained within 4 hrs. In another application, Heo et al. [49] synthesized CuZnFe2O4 composite biochar using bamboo to enhance the removal of BPA and sulfamethoxazole (SMX) from aqueous solution. The adsorption capacity of 263.2 mg/g was recorded when CZF–biochar adsorbent was applied to remove BPA. It was observed that the sorption capacity was improved for SMX and BPA when CZF nanomaterials were impregnated on the surface of biochar. Furthermore, Wang et al. [16] use grapefruit peel biochar to remove BPA under variable adsorbent dosage, pH, and contact time. The authors reported that pH has a significant influence on the adsorption of BPA and that almost 100% BPA removal was achieved at a pH (6).
In Malaysia, the agro-industrial sector generates a substantial volume of biomass solid wastes, predominantly from palm oil mills which produce the huge expanse of biomass, including palm kernel shells (PKS), empty fruit bunches, oil palm fibers, and palm oil mill effluents [50]. Malaysia recorded the largest export of 19.9 million tonnes (mt) from oil palm biomass residues in 2017 [51]. With the expansion of palm oil production in Malaysia, the volume of residue generated has correspondingly risen. Approximately 50–70 tonnes of biomass residues could be produced from a hectare of oil palm plantation [52]. Thus, the palm oil industry is presently generating around 50 mt of dry oil palm residues annually and attain 100 mt per year by 2020 [53]. Notably, PKS has the maximum commercial consumption value as compared to other oil palm biomass [54]. Besides, the specific characteristics of PKS-BC (palm kernel shell biochar), particularly its porous structure, enhancement of functional groups, its huge specific surface area, and improved mineral component, make it suitable for use as adsorbent material [42]. Despite this huge potential, PKS is considered as biomass waste, generating superfluous waste of resources and environmental nuisance [55].
There is still a lack of reports on the adsorption of BPA using PKS magnetic BC, as well as its adsorption mechanism which is yet to be explored and needs to be investigated. In view of this highlight, PKS biomass waste that is available in huge quantities in Malaysia was applied as a precursor for the synthesis of magnetic BC as a sustainable and inexpensive adsorbent for the removal of BPA from aqueous solutions via adsorption procedure.
Magnetic BC was synthesized using the chemical co-precipitation method. A batch adsorption test was employed to evaluate the maximum adsorption capacity of BC produced from PKS. The effects of essential factors, in particular contact time, adsorbent dosage, ionic strength, initial BPA concentration, and pH, on the adsorption of BPA by NBC (neat biochar) and MBC (magnetic biochar) were studied. The adsorption mechanism of BPA by synthesized BC was examined together with the analysis of the NBC and MBC, to boost the study of the adsorption of BPA by NBC and MBC. The adsorption isotherm was stimulated via Freundlich and Langmuir models. The adsorption kinetics and reusability of the novel synthesized PKS-BC were also investigated.

2. Materials and Methods

2.1. Materials and Chemicals

All chemicals: Bisphenol A (99% purity) reagent, ferric chloride hexahydrate (FeCl36H2O), ferrous sulfate heptahydrate (FeSO47H2O), and sodium hydroxide (NaOH) (AR) utilized in this study were of analytical grade and purchased from Sigma-Aldrich (West Chester, PA, USA) and applied without additional purification. Distilled water (DW) was utilized for the preparation of all the aqueous solutions during the entire experiment. The BPA solution utilized in the experiment was carefully prepared prior to each experiment to avoid possible microbial degradation. The chemical and physical properties of BPA are presented in Table 1. The PKS biochar was obtained from AMR Environmental Sdn Bhd, located at Johor, Bahru, Malaysia.

2.2. Synthesis of Magnetic Biochar

The impregnation of biochar with magnetite using co-precipitation of ferric and ferrous salts on carbonaceous materials method was according to the procedure employed in previous studies [16]. Initially, the PKS-BC sample was grounded and modified using high energy ball milling (HEBM) for 3 hrs to achieve a nano-sized biochar. The obtained ball-milled powdered BC was passed through 50–63 µm sieves and washed with distilled water and ethanol three times. Afterwards, the PKS-BC was oven-dried overnight at 110 °C. The obtained dried BC was collected in a bottle container, sealed, and labelled as NBC and kept in a desiccator for further use. In the next phase, 8.5 g of the dried PKS-BC was soaked in a 100 mL of FeCl3 of 0.25 mol·L−1 and 0.125 mol·L−1 of FeSO4 solution and uniformly mixed via the magnetic stirrer. The mixtures were subjected to heat at 60 °C with constant stirring to ensure that Fe3+ and Fe2+ could saturate into BC and later allowed to cool to 40 °C. Subsequently, 100 mL of 1 mol·L−1 of sodium hydroxide (NaOH) solution was dropwisely added to the solution and stirred until the pH attained the range of (10–11), and the colour of the solution transform from brownish green to black. Fe3O4 precipitation was formed on the surface of carbon materials under this alkaline condition. Upon the completion of the reaction, the NaOH-impregnated biochar solution was adequately stirred for 60 min to achieve better dispersion and homogenous mixture. The resultant mixture was separated by an external magnet and rinsed with distilled water until the solution pH closed to 7 and then oven-dried at 70 °C for 12 h. The final modified BC sample was grounded and collected into a bottle container, sealed, and labelled as MBC and kept in a desiccator thereafter.

2.3. Characterization of Neat and Magnetic Biochar

The magnetic properties and magnetization curve of MBC was determined using a vibrating sample magnetometer (Lakeshore 7404, Westerville, OH, USA) with an applied field between −10,000 and 10,000 Oe at room temperature, from American Quantum.
Surface functional groups of the BC samples (before and after modification) were confirmed using Fourier transform infrared (FTIR) spectroscopy (Perkin Elmer, 1650 Spectrometer) within the scanned range of 400 to 4000 cm−1 using the attenuated total reflection method.
The textural properties such as surface area, pore volume, and pore size distribution of the samples were determined by the standard N2 physisorption procedures using Micrometrics analyzer (Tristar II Plus model) and determined in line with the Brunauere Emmentte Teller (BET) technique with degassed temperature of 350 °C.
X-ray diffraction (XRD) was used to analyse the phase and chemical composition of the prepared BC adsorbents at Cu Kα radiation (2θ spectrum = 20–80°; phase = 0.05° 2θ; time per step = 0.2 s) via X-ray powder diffractometer (Rigaku MiniFlex 600). The compositions and surface morphologies of the samples were detected by field emission scanning electron microscopy (FESEM, Zeiss ULTRA 55) and energy dispersive spectroscopy (EDS, Bruker/Quanta 200), Westerville, OH, USA).
Also, the pH of BC was assessed viz: BC was blended with distilled water at (1:10) mass ratio, agitated magnetically for 30 min, and subsequently kept for 60 min. Afterwards, the pH of BC was determined by a pH5S Spear pH tester (T531009086, Shanghai, China).

2.4. Analysis of Surface Chemistry of Biochar (pHpzc)

The surface chemistry of neat and magnetically-modified biochar was performed experimentally using the pH at the point of zero charge according to the procedure of [56] with slight modifications. Ten samples of varying pH (2.0–11.0) were prepared using a 0.01 M aqueous solution of sodium acetate (C2H3NaO2) as the base electrolyte. The pH of the solution was adjusted using either 0.1 M of NaOH or 0.1 M HCl aqueous solution. Then, 0.1 g of magnetic and neat samples of BC were carefully added to each 20 mL of the prepared solutions in a conical flask placed in a swing agitator at a speed of 120 rpm at 298 K room temperature and stirred for 48 h. Afterwards, the resultant supernatant was then decanted, and its pH was determined. The value of pHpzc was computed from a plot of pH of the initial solution against pH of the supernatant [57,58]. Thus, the zeta potentials of neat biochar (NBC) and magnetic biochar (MBC) were examined via zeta potential analyzer (Zetasizer Nano Plus 3, Zeta/nano particle analyzer, Malvin, Austin, TX, USA).

2.5. Adsorption Experiment

Definite amount of NBC and MBC adsorbents (50.0 mg) and 125.0 mL of BPA solution were weighed and added to a set of 250 mL; Erlenmeyer conical flasks sealed externally using aluminium foil to avoid probable photodegradation. The mixture was then placed on an orbital shaker (HY-8, Shanghai, China) and agitated mechanically at 160 rpm at room temperature for 24 h to achieve equilibrium condition. Upon the completion of the agitation process and at pre-specified times, the solution was rapidly removed and filtered using 0.4-µm membrane and the absorbance of the supernatants were analyzed spectrophotometrically at a wavelength of 276 nm using a double beam UV-visible spectrophotometer (Shimadzu UV-1800, Kyoto, Japan) with a pair of 10 mm matched quartz cells, and the concentrations were converted into the established BPA calibration (standard) curve [59,60].
In the single-variable experiment, the effects of adsorbents (BC) doses (0.2–0.8 g; BPA = 20 ppm), pH range were adjusted by 0.01 mol/L HCl and 0.01 mol/L NaOH solution (3.0–12.0; BPA = 20 ppm), contact time (20–240 min; BPA = 20 ppm), ionic strength (Na+ concentration: 0, 0.2, 0.4, 0.6, 0.8, and 1.0 mol/L, BPA = 20 ppm); initial BPA concentrations (8–150 ppm), and reaction temperature (298.15, 318.15, and 338.15 K), on the adsorption of BPA by NBC and MBC were studied. The amount of BPA adsorbed on BC adsorbent was determined using the following Equation (1):
q t = ( C S C t ) M × V
where, Cs and Ct are the starting and residual concentrations of BPA in mg/L, M is the mass of the adsorbent (g); and V represent the solution volume (L).
For the adsorption thermodynamic tests, 0.5 g of the adsorbent and 100 mL of BPA solution were added to 250 mL conical flask to study the adsorption thermodynamic of MBC for BPA. The mixtures were stirred in a temperature-regulated incubator shaker (Excella E24 Incubator shaker series, USA) at a speed of 150 rpm, pH (6) for 60 min adsorption time, and the adsorption temperature were varied between (298.15 K–338.15 K).

2.6. Statistical Analysis

For the adsorption experiment, each experimental samples and group were carried out in triplicate, and the average value of the data was considered. The removal efficiency of BPA (Re%) and quantity of BPA adsorbed (qe) onto BC adsorbent were determined.
The adsorption isotherms, kinetics (20–240 min), and thermodynamics (298.15, 318.15, and 338.15 K) were studied. Each test was carried out in triplicate, and the average results were considered.

2.7. Reusability and Regeneration of MBC

The reusability studies on the prospect of desorbing of BPA compounds from MBC are indispensable based on environmental and economical perspective, in addition to industrial applicability [61]. With the purpose of achieving regeneration of MBC, an efficient and green eluent was considered. The reusability of MBC at the starting BPA concentration of 20 ppm was assessed by rinsing BPA-ladened MBC with 0.1-M NaOH and distilled water and subjected to ultrasonication, and then followed by oven-drying at 60 °C overnight after magnetic field separation [49]. The MBC was repetitively used five times, and the values of qe was noted accordingly.

2.8. Adsorption Isotherm

Adsorption isotherm was employed to compute the quantity of adsorbed BPA on BC based on BPA concentration at a steady temperature. The application is based on standardizing the amount of BPA adsorbed by the mass of BC adsorbent and this confirms comparison with Isotherm models.

2.9. Langmuir Adsorption Isotherm

The Langmuir model relies on the assumption that uptake of BPA takes place on a homogenous surface via monolayer adsorption with no interaction between adsorbed materials. It also presumes that all sorption sites are ‘correspondingly active’, and the surface is robustly homogeneous [62]. Principally, the larger values of R2 signify the significance of the adsorption model for contaminants removal in water. The value of RL between 0 and 1 implies that the isotherm is advantageous. The isotherm is unsuitable if RL > 1, linear if RL = 1, irreversible if RL = 0, and favorable if RL lies between 0 to 1. Arithmetically, the Langmuir isotherm model is denoted in Equation (2):
1 q e = 1 K L q m a x . · 1 C e + 1 q m a x .
where Ce represent BPA concentration at equilibrium (mg/L); qmax.(mg/g) is the maximum single-layer adsorption capacity of the BC adsorbent, qe (mg/g) is the quantity of BPA adsorbed; KL (mg/g) represent the Langmuir constant closely related to adsorption capacity for overall monolayer coverage; the intercept (1/KL), slope (1/qmax.), and qmax. could be evaluated from a plot of 1/qe versus 1/Ce for BPA adsorption onto NBC and MBC.

2.10. Freundlich Adsorption Isotherm

Freundlich isotherm model was employed to study the uneven distribution on the surface of the absorbent, which is heterogeneousness in the adsorption process [63]. The model indicates that the adsorption energy proportionally declines on the endpoint of the adsorption centres of an adsorbent [64]. The degree of correlation between adsorption and solution concentration relies on the adsorption intensity, n. The adsorption conditions can be chemical (n < 1), favourable physical process (n > 1) and favourable linear (n = 1). The KF and n are Freundlich constants and were found from the graph as the intersection and slope, respectively. Scientifically, the Freundlich isotherm model can be described in Equation (3). The graph of ln qe against ln Ce provide a linear plot with an intercept logKF and slope 1/n, from which R2, KF and n, can be computed, respectively [65].
log q e = log K f + 1 n log C e

3. Results

3.1. Results and Discussion

Characterization and Analysis of Synthesized Biochar Adsorbents

The surface chemistry and morphological characterization of BC sorbents are excellent indicators of adsorption propensity. Hence, VSM, FTIR, FESEM, EDX, and BET surface analyses, together with the point of zero charge of the adsorbents (NBC and MBC) were also evaluated.

3.2. Morphological Analysis of Synthesized Biochar

FESEM analysis was conducted to examine the surface morphology of the BC samples. The presence of abundant pores were observed on the surface of the BC, which was substantiated by the FESEM micrographs (Figure 1a–c). Images of NBC, MBC (before and after adsorption) are described in Figure 1b,c. The FESEM analysis of the MBC (prior to adsorption and after adsorption) was performed to observe the morphological structure along with the particle size and size distribution of MBC, as presented in (Figure 1b–e), respectively. From Figure 1a–c, the images showed the particles obtainable are spherical in shape with a single uniform aggregate and small non-uniform agglomerates. In Figure 1a, larger particles with a porous and rougher surface were noticed for NBC. Also, the particles possess more cleavages with few connected to bulks. Furthermore, it can be observed from the FESEM micrographs (Figure 1a,b) of NBC and MBC, that the surface of NBC is relatively coarse and rough since the spherical shaped particles with particle size more than 28.32 nm were agglomerated and attached to each other and the pore form is not fully developed. Contrastingly, the surface of MBC is comparatively smooth, with substantial porous structure than NBC, and the surface is filled with numerous nano-iron oxide particles. The dispersal of Fe3O4 nanoparticles on the surface of BC is relatively uniform. The surface of MBC developed shinier and smoother surface after adsorbing BPA (Figure 1c). The average particle size of MBC prior to adsorption as depicted in Figure 1d was computed to be 28.32 nm. Conversely, after adsorption of BPA, the average particle size rose to 46.65 nm as showed in Figure 1e, confirming that the adsorption process was taking place.

3.3. Elemental Analysis of Synthesized PKS Biochar

For this procedure, 1.0 g of each of the samples (NBC and MBC) was utilized for the EDX analysis. This test studied different component elements existing in each of the BC samples. The EDX test performed reveal the elemental compositions and distribution both in the NBC and MBC. Figure 2 presents the elemental analysis of the NBC and MBC. As revealed in Figure 2, the existence of peaks matches with the carbon (C), oxygen (O), iron (Fe), and silicon (Si) elements. The intensity of C peak is greater than O, Si, Al, and Fe, demonstrating the elevated content of C in NBC (Figure 2a). Also, three peaks displayed at 0.2, 0.5, and 6.40 keV to substantiate the binding energies of Fe3O4 nanoparticles [66]. The elemental composition of novel MBC was presented in Table 2 with the mass ratio of C, O, and Fe was 32.52, 27.23, and 38.14%, respectively.
The elements of NBC include C, O, Si, Al, and Fe, while the element observed in MBC were C, O, Si, and Fe. Noticeably, from Figure 2b,c, the Al element entirely disappeared after the magnetic modification, and also C, and Si weight components were significantly reduced. This indicates that chemical reactions occur during the modification. Conversely, Fe and O weight composition were substantially increased in modified MBC as compared to NBC. Particularly, the oxidation process considerably influences the obtainability of C active sites, in addition to the structural reformation of the adsorbent [67]. This indicates that the oxidation activity of the carbonaceous material justified the reduction in the weight composition of the trace elements, and also the formation of more acidic oxygen-containing functional groups on the surface [68]. These processes play a significant role in the development of more active sites for effectual adsorption process. The mass fraction of Fe in MBC was 38.14%. The high composition of Fe3O4 further proved the success of the magnetic modification of BC, which agreed with the FTIR and XRD analysis.

3.4. BET Surface Area Analysis of Biochar

The nitrogen adsorption isotherm of NBC and MBC is presented in Figure 3. Essentially, the physical factors of BC samples that can influence the adsorptive removal of organic contaminants comprises of effective surface area, total pore diameter and pore volume. Surface area properties of the NBC and MBC were assessed via BET analysis. The BET results illustrate the adsorbate–adsorbent relationships of that adsorbed molecules which are gathered around on the surface of MBC [69]. The results of effective BET surface area, pore diameter and pores volume for both NBC and MBC are presented in Table 3. Noticeably from Table 3, the NBC exhibited a superior surface area of 536.54 m2/g, while MBC had a surface area of 362.07 m2/g. The decline in the surface area of MBC was possible since the MBC was comprised of a moderate surface area of Fe3O4 and an elevated surface area carbonaceous, and superfluous iron (Fe) nanoparticles loading could capture and clog some of the active sorption sites and pores of the NBC [70,71]. This reduction in surface area of MBC composites as compared to NBC has also been reported in previous studies [72,73]. Similarly, the pore size and pore volume of MBC are higher than that of NBC, which is better beneficial for the sorption of BPA.
Similarly, the total pore volume of NBC and MBC found were, respectively, 0.416260 cm3/g, and 0.442203 cm3/g, while MBC modified by Fe3O4 exhibits a better porosity, and presents a substantial capacity of pollutant adsorption, while after modification with Fe3O4, the existence of Fe3O4 between layers of MBC upsurges the heterogeneity of the adsorbent, thereby resulting in a superior porosity [74]. The impregnation of magnetite on BC has no significant influence on the pore-volume, surface area, and mean aperture. Iron-amendment can either decrease or increase the surface area of an adsorbent and this relies on the initial surface area value of the adsorbent and the proportions of the Fe3O4 particles [75]. The infrequent blocking of surface micropores from magnetite groups may also be responsible for the insignificant reduction in surface area. Thus, the adsorption capacity of modified biochar material is improved. This finding was further corroborated by the FESEM images and EDS spectra. Figure 3 displays the N2 adsorption–desorption isotherms of N2 at 77 K of NBC and MBC. The findings revealed that at moderately high relative pressures, the adsorption isotherm rose relatively than at relatively low pressures. This implies that the major adsorption takes place at moderately increased pressures and suggests that the material is highly porous with a narrow size distribution.

3.5. Magnetic Properties of As-Synthesized Magnetic Biochar

The vibrating sample magnetometer (VSM) procedure was used to determine the hysteresis loop at room temperature. Most importantly it is employed to quantify the magnetic properties of a materials with respect to magnetic field, time, and temperature. Besides, VSM analysis provides information about whether the magnetization is perpendicular or parallel to the plane described by the substrate. The hysteresis loop of the synthesized MBC was obtained by plotting the magnetization (emu/g) against the magnetic field (Oe). Figure 4 reveals the magnetic hysteresis loops used to analyze the magnetic properties of MBC. As illustrated in the loop observed in Figure 4, the magnetization sharply increased with the decrease of the mass ration of BC. The curve signified distinctive super-paramagnetic properties. The synthesized MBC exhibited magnetization M values of 6.4882 emu/g which revealed that the magnetization of MBC can be altered by the mass ration of BC: magnetite (Fe3O4) nanoparticles. The super-paramagnetic properties was triggered by the alteration of BC with magnetite (Fe3O4) nano-particles, which could ensure the MBC can be readily recovered from the suspended solution through external magnetic field, which makes the replicated use of the MBC in the actual wastewater purification system viable [72]. Thus, novel MBC was easily detached using an external magnetic field, as illustrated by the inset. The result of magnetization obtained in this study is consistent with previous studies reported from [76,77,78].

3.6. XRD Analysis of Synthesized Biochar

X-ray diffraction analysis of a material describe the size and the nature of the planes of the synthesized material. The crystalline structures of NBC and MBC were analyzed using XRD as indicated in Figure 5. XRD is an efficient technique to verify the presence of Fe3O4 in the synthesized MBC [79]. Figure 5 reveals the XRD patterns of both NBC and MBC. The diffraction spectra of the synthesized BC samples showed the existence of magnetite (Fe3O4). The diffraction peaks at 2θ of 30.044°, 36.66°, 36.57°, 42.45°, 57.26°, 61.54°, and 62.76° are indexed to the (200), (016), (220), (232), and (040) hkL planes, respectively, which correspond satisfactorily with the database of Fe3O4 standard card Inorganic Crystal Structure Database (ICSD No. 98-007-7864) with a space group of P12/c1 and lattice parameter (a = b = c) of 28.644 Å and confirms the signature peaks of a hexagonal unit cell Fe3O4, respectively. No impurity peak is observed in the XRD pattern, which indicates that the Fe3O4 particles are highly crystalline hexagonal spinel structure. As observed from the spectra, all the diffraction peaks are designated to the magnetic hexagonal structure. There are no other peaks associated with another material detected from the XRD result, which confirmed that the nano-magnetite is pure magnetite (Fe3O4). The XRD spectra of Fe3O4 in this study is analogous to other studies from [80,81], and both Fe3O4 and BC patterns were overlapped in a XRD spectra of MBC, demonstrating successful synthesis of MBC composite.

3.7. Analysis of Functional Group

The surface functional groups are the central chemical variable of BC material that influence the BPA adsorption. FTIR spectroscopy was employed to substantiate the modification process and acquire the information on the existence of different functional groups on the surface of the material [82]. The spectrum results and various bands in the spectra signifying vibration of functional groups were illustrated in Figure 6 for NBC and MBC, respectively. During the Fe3+/Fe2+/NaOH process, the surface of the char aids the nucleation of iron oxide precipitation. As indicated in Figure 6, the existence of functional groups of –COOH and –OH group are responsible for the binding of the iron oxide and iron hydroxide particles in the solution, and then bonded to the char [83]. This could be symbolized as char-O-FexOy (Fe3O4/Fe2O3). In this context, the bold ‘O’ was initially a hydroxyl group on the surface and signify a chemical bond between iron oxide particle and the char phases. This bonding may conceivably be combined with some mechanical interlocking between H-bonding (such as between Fe–OH at the metal oxide surface with C–OH of the char surfaces) and phases in addition to some columbic interfaces. These interactions could strongly bind magnetic iron oxide to the BC materials [84].
Also, as indicated in Figure 6 of the FTIR spectra, there are limited peaks that surfaced for both NBC and MBC. The peaks at 2834, and 2878 cm−1 were distinctive peaks of the C-H bond [82]. Similarly, the peak at 1081 cm−1 was the C-O stretching vibration in the composition of carbohydrate, polysaccharide, or aromatic ether. The peaks close to 1578 and 1579 cm−1 were the stretching vibration peaks of C=C, C-H, and of C=O on the aromatic ring [85]. The results go along with the BC characteristics that not only contain porous structure but also abundant active adsorption sites for BPA removal [86]. Besides, a noticeable peak at 652 cm−1 was observed for MBC, which correspond to the typical stretching peak of Fe-O [87]. This indicated that the Fe3O4 nanoparticles have been efficaciously loaded on the surface of BC as revealed by the FESEM results. The spectra results revealed that magnetic alteration improved the varieties and number of functional groups of the BC, which may influence the adsorption of BPA.

3.8. Determination of Electrokinetic Charge (pHpzc) of Biochar

pHpzc is the pH at which the net surface charge on the surface of an adsorbent is zero. It is a critical variable in examining the efficacy of the adsorption systems. When the pH is greater than the pHpzc, the adsorbent surface acquires negative charge, hence repelling or attracting organic contaminants, in line with their cationic or anionic functional groups [88]. The point of zero charges was investigated to describe the surface chemistry of MBC and NBC. The values of point of zero charge (pHpzc) of neat and MBC in relation to the solution pH were computed and displayed in Figure 7. According to Figure 7, the pHpzc at the point where the change in pH (initial pH-final pH) equal to zero for the MBC and NBC were 5.61 and 4.81, respectively. It is observed that the pHpzc of MBC (5.61) is higher than that of NBC (4.81). This increase may be due to the introduction of iron oxides (Fe3O4) on the surface of NBC [89]. Hence, higher pHpzc is more beneficial for the adsorption of BPA anions, which are often existed in acidic solution [90]. The zeta potential of all samples steadily reduced with rising pH indicating more net negative surface charge at elevated pH. Similarly, the zeta potential of both NBC and MBC was both positive and negative within the whole pH range studied (2–11) as can be seen from Figure 7. Evidently, both MBC and NBC are positively charged when the pH is lower than 6.0. Once the pH is increase, the surface charge of the two biochar materials swiftly transformed from positive to negative. As pH consistently increase, the surface charges of both NBC and MBC risen and, noticeably, the zeta potential of NBC and MBC were −28.56 mV and −22.47 mV, which shows that both biochar samples were negatively charged (see Figure 7). Though, the surface of NBC is slightly more negatively charged, and the electrostatic attraction between the negatively charged NBC and HBPA improves the adsorption. This finding is in good agreement with previous studies [16,91].

3.9. Influence of Working Conditions on the Adsorption of BPA

The operating variables for instance contact time, solution pH, adsorbent dosage, ionic strength, BPA concentration, and other variables can influence the surface properties of the adsorbent surface and its BPA binding ability [88,92]. Hence, a batch adsorption test was carried out on the as-synthesized BC adsorbents to study the influence of these operating variables on the adsorption capacity and removal efficiency for BPA uptake.

3.10. Effect of pH on BPA Adsorption

The pH is one of the main parameters that control the removal of compounds present in aqueous environment using solid adsorbents. The solution pH is among the most essential variables that determine the elimination of various compounds existing in aqueous environment utilizing solid adsorbent materials and the optimization of adsorption process. The effect of pH on adsorption was reliant on the target contaminants and nature of absorbent [42]. It influences not only the speciation of the adsorbate, but also the level of ionization and adsorbent surface charge [90,93,94]. Accordingly, most of the investigations involved in contaminants adsorption onto BC strongly considered the effect of solution pH. To investigate the influence of different solution pH upon adsorption of BPA on the surface of NBC and MBC, the tests were performed in the pH range of 3.0–12.0, BPA concentration 20 ppm, adsorbent dose 0.5 g, and temperature 20 °C, and the results were illustrated in Figure 8a. As indicated in the figure, the sorption of BPA by NBC and MBC is clearly pH reliant. The highest adsorption capacity of NBC to BPA take place at pH (3.0), which is attributable to the development of electron receiver-giver interaction (ERG) between BPA and NBC, together with a robust hydrogen bond [95,96]. When pH rose from 6.0 to 7.0, the adsorption capacity of NBC to BPA slightly increased further, since BPA starts to moderately dissociate, and BPA in the solution is no more in molecular form, nevertheless few HBPA still occurs. The electrostatic attraction between NBC and HBPA enhances the adsorption process due to the positively charged surface of NBC. Further surge in pH led to a subsequent reduction in the adsorption capacity of NBC to BPA, owing to weakness in the interaction of the hydrogen bonding and π-π electron receiver–giver(ERG) between BPA and NBC when the solution pH is higher than the BPA acid dissociation constant [97]. Conversely, the adsorption capacity of BPA by MBC rose initially and after that it declined with the upsurge of pH greater than 6.0, which implied that the adsorption of BPA on BC was largely dependent on the solution pH of the system. Thus, the considerable rise in BPA adsorption on MBC between pH 3.0–6.0 was possibly attributed to the electrostatic interaction between the positively charged HBPA+ species and negatively charged surface of MBC in the solution. The surface of MBC may develop positively charged at low pH because of the protonation reaction (H+(aq) + −ROH(surf) −ROH2+(surf)) on the surface of MBC [98]. When the solution pH is increased, a negatively charged surface of MBC with plentiful active biding sites emerge owing to deprotonation reaction (OH(aq) + −ROH(surf)  H2O + RO(surf)) on the surface of MBC. Hence, the enhanced BPA adsorption by MBC at pH 3.0–6.0 was not only controlled by the electrostatic interaction mechanism [99], but possibly caused by the surface reduction/complexation of BPA species onto MBC, and the decline of BPA adsorption at pH ˃ 6.0 following the repulsion interaction between negatively charged MBC and dissociated bisphenolate anions (BPA2− and HBPA) species. The maximum adsorption capacity take place at pH of 6.0.

3.11. Effect of Ionic Strength on BPA Adsorption

Generally, the water body comprises such a complex system that salts and organic contaminants frequently coexist in wastewater which may influence the removal of the contaminants. Hence, the profile of ionic strength influence on BPA adsorption by the as-synthesized BC was also investigated using 0–1.0 mol/L NaCl, BPA concentration 20 ppm, and the experimental findings are presented in Figure 8b. Initially, it could be noticed in Figure 8b, the increase in NaCl concentration led to a decline in the adsorption capacity of NBC to BPA and later a rise. Noticeably, the surge in high ionic strength (Na+ concentration) considerably increased the adsorption capacity of BPA to MBC (Figure 8b). The BPA adsorption capacity of BC increased when the Na+ concentration rose from 0 to 1.0 mol/L. Analogously, Zhou et al. [100], found that the upsurge in ionic strength with various ionic species (CaCl2 or NaCl) yielded increase in the adsorption capacity of BPA when peat was applied as an adsorbent. Similarly, erstwhile study has also reported that the increased ionic strength could improve the adsorption of organic contaminants into carbonaceous adsorbents, such as BC, owing to the screening effect of the surface charge generated by the addition of salt [101]. The BPA adsorption capacity by BC adsorbents rose steadily with the increase in the NaCl concentration. The plausible explanations for this surge could be due to the penetration of ions into the diffusion dual layer around NBC and MBC surfaces and lessen the repulsion between the adsorbents, thereby stimulating the squeezing-out effect (i.e., nanoparticle aggregation), which caused a decline in the adsorption capacity of BPA [102], enhanced activity coefficient of hydrophobic organic contaminants, which leads to a salting-out effect (reduction in solubility), and, therefore, was favorable to the adsorption of BPA [103]. Thus, the decreased adsorption capacity when the NaCl concentration is at 0.2 mol./L might be caused by the competition between BPA and the low concentration of salt solution for the accessible active sites of adsorbents, corroborating that the squeezing-out effect was stronger than the salting-out effect. Conversely, as the NaCl concentration increases, the adsorption capacity enhanced evidently, suggesting that the salting-out effect improved unceasingly [104].

3.12. Effects of Biochar Dose on the Removal of BPA

Adsorbent dosage (AD) is an important parameter in an adsorption process. It controls the adsorbate–adsorbent equilibrium of the adsorption procedure. The removal efficiency based on the adsorbent dosage was studied at other test conditions, BPA concentration 20 ppm, varying dosages (0.2–0.8 g), pH 6.0, contact time (60 min), and adsorption temperature 20 °C. The influence of BC dose on the adsorption of BPA by NBC and MBC is illustrated in Figure 9. The removal rate of BPA enhanced with the rise in BC dose. When BC dosage was 0.2 g, the removal efficiencies of BPA by NBC and MBC were 48.45% and 60.0%, respectively. Similarly, when the BC dose rose to 0.5 g, the removal efficiency substantially rose to 85.97% and 94.2%, respectively. The increased removal rate is because of the augmented BC, which also expands the operational specific surface area of adsorption, together with increase in the active pore (binding) site of adsorption [105]. This outstanding adsorption performance was attributable to the presence of plentiful active sites and various interactions between BC adsorbent and BPA, specifically ion exchange and complexation electrostatic interaction. Afterward, further increase in BC dosage has no influence on the removal rate. This is because surplus adsorbent overlaps the effective active sites on BC and competes for limited solutes [42].

3.13. Effects of Initial Concentration and Contact Time on the Removal of BPA

The influence of initial BPA concentration and contact time on the adsorption of BPA from aqueous solution was investigated. The experiments were performed in the presence of a constant dose of adsorbent (0.5 g) of BC at room temperature, at pH (6.0) with various initial concentrations of BPA in the stock solution varied between 8.0 to 150.0 ppm, at different time intervals up to 240 min. The result of effect of initial concentration on BPA removal is presented in Figure 10. The initial concentration offers the stimulating force required to overcome the mass transfer wall between the adsorbate and the adsorbent media [106]. Thus, a higher initial concentration may enhance the efficacy of the adsorption process. Evidently, BPA removal is lower at a small concentration due to smaller amount of adsorbates in the solution to dominate active sites on the adsorbent and the amount of BPA adsorbed rises with the upsurge in BPA concentration. Thus, increase in initial BPA concentration, led to gradual rise in the removal efficiency of BPA, since the increased BPA concentration can improve the adsorption drive between the solute and the adsorbent [107,108]. The rise in the concentration of BPA conforms with an increase in BPA removal as shown in Figure 9. When equilibrium is attained, the adsorbent turns out to be saturated. This result is in line with previous study reported by Wang and Zhang [16].
The test conditions evaluating the influence of contact time on the adsorption of BPA at the test conditions, BPA concentration 20 ppm, adsorbent dosage 0.5, and pH 6.0. Similarly, the result of the effects of contact time on the adsorption capacity and the removal efficiency are illustrated in Figure 11. Noticeably, more than 93% of BPA became adsorbed in about 60 min. For contact time beyond 60 min, the per cent removal of BPA remains stable, because the active sorption sites have been saturated on the adsorbent surface.

3.14. BPA Sorption Isotherm Study

Sorption isotherm models are indispensable for recognizing the mechanisms of the adsorption process based on fundamental characteristics and numerical derivations. Also, the isothermal models are used to study the interrelatedness between the adsorbate and the adsorbent. The Freundlich and Langmuir isotherm models are normally employed and were fitted to the adsorption experimental data. The Freundlich isotherm is appropriate to both multilayer and monolayer adsorption and assumes that the adsorbates are adsorbed onto the heterogeneous surface of an adsorbent [109]. The Langmuir isotherm assumes monolayer adsorption on a uniform surface with a limited amount of adsorption sites [110].
The Langmuir isotherm model presumes that every molecule has a stable adsorption enthalpy and activation energy and signifies homogeneous adsorption. Correspondingly, the Freundlich isotherm model is experimental and regards the surface to be heterogeneous [111]. Two regular isotherm models were applied to fit the BPA adsorption isotherms on the as-synthesised BC (Figure 12 and Figure 13). The models were employed to better investigate the BPA adsorption mechanism and performance. The fitting data are illustrated in Figure 12 and Figure 13, and the applicable fitting parameters of these models were summarized in Table 3. The BPA adsorption experimental data were fitted to the Freundlich and Langmuir isotherm models with Origin 8.0.
The fitting of the experimental data into the isotherm models illustrates the adsorption process by the correlation coefficient R2 and constants. To accomplish this, log qe is plotted against log Ce based on the linear form of the Freundlich model (Figure 12) [112]. For the Langmuir model linear expression, 1/qe versus 1/Ce is plotted (Figure 13) [113,114]. The computed qmax values of the Langmuir model from the graph of 1/qe against 1/Ce was lower than the qe values from the experiments (see Table 4), indicating that the experimental adsorption data fit this model. However, the R2 value of Langmuir is lower than that of the Freundlich model.
The variables revealed that the Freundlich model (R2 = 0.88964–0.9195) offers a superior fit to the data than the Langmuir model (R2 = 0.7535–0.85608), as showed by the linear regression values. Additionally, the heterogeneity coefficient (1/n) of the two BC falls above 1 (1/n ˃ 1), which implied satisfactory physical adsorption process [115]. This trend was also corroborated by several studies [116,117] using BC synthesized from eucalyptus forest residues, sugarcane bagasse, castor meal, water hyacinth, green pericarp of coconut, and MBC to remove various contaminants. Hence, BC generated from various biomasses at different temperatures were observed to have various adsorption mechanisms and adsorption abilities [118]. In this work, BPA adsorption onto NBC and MBC was excellently fit using the Freundlich model in the determination of the efficacy of NBC and MBC for the removal of BPA, which demonstrates that the adsorption take place as a heterogeneous surface multilayer. A synopsis of the isotherm variables is presented in Table 4.

3.15. BPA Adsorption Kinetic Studies

Kinetic study offers essential information on the mechanism of adsorption and influencing mechanism of adsorption process as either chemical reaction or mass transfer to attain optimal working conditions for industrial scale [119]. With a view to examine the kinetic mechanism triggering BPA adsorption, experimental data were fitted to the pseudo-first order (PFO) [120] and pseudo-second order (PSO) [121] linearized models using Equations (4) and (5), respectively:
l n ( q e q t ) = l n ( l n q e K 1   t )
1 q t = 1 K 2 q e 2 × 1 t + 1 q e
where qt and qe are the amounts of BPA adsorbed at time t and equilibrium, respectively (mg/g), t is the contact time (min), and K1 and K2 are the rate constants of the PFO and PSO kinetic models, respectively (1/min). In batch process, sorption kinetics is described by different models based on adsorption equilibrium including the PSO (pseudo-first order) and PSO (pseudo-second order) kinetic models. In this context, it is used to model the kinetics of the adsorption of BPA onto BC.
Figure 14a,b illustrates the graphs of (ln(qeqt) against time) and (time/qt against time) for the pseudo-first order and pseudo-second order kinetic models, respectively. The computed variables and the experimental data of the two models are presented in Table 5. Noticeably the coefficient of correlation value (R2 = 0.9937) for the PSO model is of better linearity and higher in comparison with the PFO (R2 =0.8515) for BPA adsorption, which was also noticed for a similar compound (sulfamethoxazole) adsorption by functionalized BC [122]. Similarly, the computed qe (mg/g) for the PSO model is close to the experimental qe value (see Table 5). Hence, PSO model satisfactorily fits better the experimental results than the PFO kinetic model. The PSO kinetic model suggests that the chemisorption can be regarded as a rate-controlling phase during the adsorption procedure [3,123]. Similarly, the chemisorption takes place via electron-exchange or sharing between BPA and BC [124]. The elevated rate constants of the PSO model for BPA can be attributed to rapid interaction with the active adsorption sites of MBC. This finding agrees with previous studies on the BPA adsorption on the following adsorbents: algal BC [48], magnetic CuZnFe2O4–BC composite [125], modified organo-montmorillonites [126].

3.16. Adsorption Thermodynamics

Thermodynamic studies are employed to decipher any reaction in a better approach and reveals whether the adsorption process is an intended process or a spontaneous process [127]. It also indicates the influence of temperature on the adsorption process. Generally, exothermal and endothermal sorption processes are the two common processes. If the sorption declines with increasing temperature, it implies the exothermal sorption process whereas if the sorption increases with rising temperature, it signifies that the sorption is an endothermal process. In this study, the temperature influence on BPA adsorption onto MBC was studied by performing adsorption tests at varying temperatures, viz., 298.15, 318.15, and 338.15 K, at optimum pH (6.0) and adsorbent dosage 0.5 g, respectively. The thermodynamic variables such as enthalpy change (ΔH°), standard Gibb’s free energy change (ΔG°), and change in entropy (ΔS°), were computed and summarized in Table 6 using the following equations:
Δ G O = R T ln K L
K L = q e C e
Δ S ° = Δ H ° Δ G ° / T
where R is the gas constant (8.314 J/mol−1K−1), qe is the adsorption capacity (mg/g), Ce is the equilibrium concentration (mg/L) and T is the actual temperature (K). The y-intercept and the slope of the linear fit following a plot of ln KL versus 1/T were employed to calculate the values of Δ and Δ via the Van’t Hoff plot (Figure 15).
As the sorption temperature rises, the values of Δ turn out to be more negative, which implies that BPA is better effectively adsorbed to MBC at elevated temperatures. The negative Δ indicate that all adsorption processes were spontaneous [128]. The positive Δ (51.23 kJ·mol−1) indicated that the interaction between BPA and MBC was an endothermic process which possibly attributed to the competitive dissolution of BPA in aqueous solution [129]. The positive Δ and negative ΔG0 values suggested the spontaneity of the adsorption process. The higher positive Δ (176.13 J·mol−1·K−1) as indicated in Table 5, further demonstrate increasing randomness at the adsorbate-adsorbent interface as a result of free water molecules [130]. Thus, the determined positive Δ and negative Δ values at experimented temperatures substantiated the BPA adsorption on MBC and indicated the spontaneity of the sorption process, besides the entropy effect should be the primary leading force for the adsorption of BPA on MBC [131].

3.17. Investigation of Biochar Regeneration and Reusability

Recyclability of the adsorbents is strongly essential to preserve process costs down and for the industrial-scale application. Five phases of adsorption–desorption tests were undertaken to examine the reusability of the as-synthesized MBC in accordance with previously adopted experimental procedure [125]. The MBC was repetitively applied five times, and the qe values were recorded, as illustrated in Figure 16. The result indicated that qe was slightly reduced during the adsorption–desorption experiment. Upon the completion of five cycles of reusability test, the removal efficiency of BPA only reduced by 12.85% as compared with the first cycle. The slight reduction in adsorption capacity during multiple adsorption–desorption tests, was due to the partial desorption of BPA in MBC surface, competitive available sorption sites, loss of solid in solution, and the elution of iron oxide nanoparticles and energetic substances on MBC surface [49,132]. This reveals that the MBC could be an effectual, economical, and environmental benign adsorbent with superior re-usability, which can be practically applied in BPA removal procedures. The recyclability test reveals that MBC can be applied repeatedly in wastewater purification as an efficient adsorbent.
Table 7 presents the comparison of specific surface area characteristics, adsorption capacity of BPA, and the magnetic intensity of the synthesized MBC material in the current study in comparison with various studies.
Though few studies reported higher adsorption capacity than the current study despite their low surface area as indicated in Table 7, this is because previous studies have described the performance of MBC to be significantly influenced by the nature of auxiliary and raw materials, starting contaminant concentration, pyrolysis temperature, competitive anions, nature of modifier, reaction temperature, sorption time, and various synthesis methods [86,136,137]. As highlighted in Table 7, the maximum adsorption capacity recorded in the current study is comparable and in line with previous studies using different BC adsorbents. Hence, it can be inferred that the MBC synthesized from palm kernel shell is a promising, efficient, and essential precursor (adsorbent) for the removal of BPA from aqueous environment. The sorption capacity recorded from the current study using magnetic PKS-BC could be stems from highly developed pore structure, superfluous surface area, smaller pore size, considerable surface functional group, high sorptive capacity, and super paramagnetism. It is noteworthy that studies on the adsorption of BPA onto MBC synthesized in-house from palm kernel shell biomass is still very limited. Besides, the adsorbent demonstrated superior regeneration efficacy, and the resultant MBC could be reclaimed numerous times.

3.18. Controlling Mechanism for BPA Removal

The schematic diagram of mechanisms of adsorption of BPA (adsorbate) onto as-synthesized NBC and MBC is depicted in Figure 17. The average pore size of the as-synthesized BC falls below 50 nm, indicating mesoporous. Also, BPA molecules can penetrate the pores of the as-synthesized BC via pore filling. As stated earlier, the adsorption capacity of NBC and MBC to BPA is largely influenced by pH. In acidic environments, the phenolic hydroxyl groups in the chemical structure of BPA protonate, and thereby producing electrostatic repulsion with the positively charged surfaces of both NBC and MBC. Also, since BC adsorbent comprises carbon, silicon, iron, and oxygen groups, these elements C, Si, Fe, and O groups are vastly electronegative because of the sufficiency of available lone pair of electrons, which exhibits binding capabilities toward the BPA as indicated in Figure 2. The carboxyl, hydroxyl, carbonyl, and amine functional groups (as identified in Figure 5) facilitate the affinity of BPA molecules and their adsorption on the surface of BC. Such sorption mechanism is an electron giver–receiver type according to the unbalanced electrons’ supply between BPA compound and the BC functional groups. The hydroxyl functional group (OH) on the BC’s surface also develops potent hydrogen bonds with the C-H and OH on the molecular composition of BPA. Also, the C=O and COOH acidic functional groups of the as-synthesized BC can serves as electron receivers, producing π-π electron giver–receiver interactions (EGR) with BPA. In alkaline environments, the surface of MBC becomes negatively charged, which retains the electrostatic repulsion with the dissociated BPA2− and HBPA, whereas NBC acts oppositely. Similarly, the π-π electron giver-receiver interaction (EGR), as well as hydrogen bonds between BPA and as-synthesized BC, would be severely weakened, thereby making it hard for BC to adsorb BPA. Since, BPA exhibits robust hydrophobicity and can be intermixed with the hydrophobic site on NBC and MBC surfaces. Hence, a hydrophobic interaction is correspondingly a central influential force for the adsorption of BPA onto NBC and MBC.

4. Conclusions

In this work, a novel magnetic palm kernel shell biochar was efficiently synthesized via magnetic modification and applied to remove BPA from aqueous solution. The results of VSM, XRD, FTIR, BET, FESEM/EDX, and Zeta potential revealed that the Fe3O4 nanoparticles was effectually engrafted on the surface of biochar. The experimental findings demonstrated the adsorption isotherm could be better fitted by Freundlich model, whereas the adsorption kinetic data was controlled by pseudo-second order model. The as-synthesized BC also demonstrated a good magnetic strength for facile recovery, superior recyclability, high selectivity, and eco-friendly. It was noticed that BPA removal is greatly influenced by adsorbent dose, pH, and contact time. The BPA adsorption does increases with reduction of pH, with a maximum adsorption at pH 6 for MBC. Thermodynamic study revealed that BPA adsorption on MBC is endothermic process and spontaneous due to +ΔH and −ΔG, respectively. The principal mechanisms for BPA adsorption on the prepared BC comprised of electrostatic interactions (π-π electron acceptor– donor interactions), hydrophobic interaction, and H-bonding. Therefore, it can be deduced from this study that novel MBC is efficient and practical for the removal of BPA from aqueous solution with benefit of being sustainable, and abundantly available.

Author Contributions

K.K.K. designed the project, performed the study, and wrote the manuscript; K.F.Y. supervised the study, and helped acquire funding; R.S.A. contributed to nanoparticles synthesis and characterization; thanks M.Z.M.N., for the software support; A.Z.A. and H.C.M. were principal investigators and helped acquire funding; and A.Z.A. contributed to chemicals. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by Universiti Putra Malaysia under FRGS with grant number 5540204.

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

The Universiti Putra Malaysia is hereby appreciated. The sponsorship received from the Tertiary Education Trust Fund (TETFund) through the Kwara State University, Malete, Nigeria, is also hereby thankfully acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ahsan, M.A.; Islam, M.T.; Hernandez, C.; Kim, H.; Lin, Y.; Curry, M.L.; Gardea-Torresdey, J.; Noveron, J.C. Adsorptive Removal of Sulfamethoxazole and Bisphenol A from Contaminated Water using Functionalized Carbonaceous Material Derived from Tea Leaves. J. Environ. Chem. Eng. 2018, 6, 4215–4225. [Google Scholar] [CrossRef]
  2. Vandenberg, L.N.; Luthi, D.; Quinerly, D.A. Plastic bodies in a plastic world: Multi-disciplinary approaches to study endocrine disrupting chemicals: We dedicate this manuscript in memory of Dr. Theo Colborn: Mentor, colleague and friend. J. Clean. Prod. 2017, 140, 373–385. [Google Scholar] [CrossRef]
  3. Rathnayake, S.I.; Xi, Y.; Frost, R.L.; Ayoko, G.A. Environmental applications of inorganic-organic clays for recalcitrant organic pollutants removal: Bisphenol A. J. Colloid Interface Sci. 2016, 470, 183–195. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Michałowicz, J. Bisphenol A–Sources, toxicity and biotransformation. Environ. Toxicol. Pharmacol. 2014, 37, 738–758. [Google Scholar] [CrossRef]
  5. Hoekstra, E.J.; Simoneau, C. Release of Bisphenol A from Polycarbonate-A Review. Crit. Rev. Food Sci. Nutr. 2013, 53, 386–402. [Google Scholar] [CrossRef]
  6. Mileva, G.; Baker, S.L.; Konkle, A.T.M.; Bielajew, C. Bisphenol-A: Epigenetic reprogramming and effects on reproduction and behavior. Int. J. Environ. Res. Public Health 2014, 11, 7537–7561. [Google Scholar] [CrossRef]
  7. vom Saal, F.S.; Nagel, S.C.; Coe, B.L.; Angle, B.M.; Taylor, J.A. The estrogenic endocrine disrupting chemical bisphenol A (BPA) and obesity. Mol. Cell. Endocrinol. 2012, 354, 74–84. [Google Scholar] [CrossRef] [Green Version]
  8. Su-Hua, W.; Bing-zhi, D.; Yu, H. Adsorption of bisphenol A by polysulphone membrane. Desalination 2010, 253, 22–29. [Google Scholar] [CrossRef]
  9. Wintgens, T.; Gallenkemper, M.; Melin, T. Removal of endocrine disrupting compounds with membrane processes in wastewater treatment and reuse. Water Sci. Technol. 2004, 50, 1–8. [Google Scholar] [CrossRef]
  10. Rajasärkkä, J.; Pernica, M.; Kuta, J.; Lašňák, J.; Šimek, Z.; Bláha, L. Drinking water contaminants from epoxy resin-coated pipes: A field study. Water Res. 2016, 103, 133–140. [Google Scholar] [CrossRef] [Green Version]
  11. Yüksel, S.; Kabay, N.; Yüksel, M. Removal of bisphenol A (BPA) from water by various nanofiltration (NF) and reverse osmosis (RO) membranes. J. Hazard. Mater. 2013, 263, 307–310. [Google Scholar] [CrossRef]
  12. Wee, S.Y.; Aris, A.Z.; Yusoff, F.M.; Praveena, S.M. Occurrence and risk assessment of multiclass endocrine disrupting compounds in an urban tropical river and a proposed risk management and monitoring framework. Sci. Total Environ. 2019, 671, 431–442. [Google Scholar] [CrossRef]
  13. Nazifa, T.H.; Kristanti, R.A.; Ike, M.; Kuroda, M.; Hadibarata, T. Occurrence and distribution of estrogenic chemicals in river waters of Malaysia. Toxicol. Environ. Health Sci. 2020, 12, 65–74. [Google Scholar] [CrossRef]
  14. Shekhar, S.; Sood, S.; Showkat, S.; Lite, C.; Chandrasekhar, A.; Vairamani, M.; Barathi, S.; Santosh, W. Detection of phenolic endocrine disrupting chemicals (EDCs) from maternal blood plasma and amniotic fluid in Indian population. Gen. Comp. Endocrinol. 2017, 241, 100–107. [Google Scholar] [CrossRef]
  15. Colin, A.; Bach, C.; Rosin, C.; Munoz, J.F.; Dauchy, X. Is drinking water a major route of human exposure to alkylphenol and bisphenol contaminants in France? Arch. Environ. Contam. Toxicol. 2014, 66, 86–99. [Google Scholar] [CrossRef]
  16. Wang, J.; Zhang, M. Adsorption characteristics and mechanism of bisphenol a by magnetic biochar. Int. J. Environ. Res. Public Health 2020, 17, 1075. [Google Scholar] [CrossRef] [Green Version]
  17. Gadupudi, C.K.; Rice, L.; Xiao, L.; Kantamaneni, K. Endocrine Disrupting Compounds Removal Methods from Wastewater in the United Kingdom: A Review. Science 2019, 1, 15. [Google Scholar] [CrossRef] [Green Version]
  18. Rodriguez-Narvaez, O.M.; Peralta-Hernandez, J.M.; Goonetilleke, A.; Bandala, E.R. Treatment technologies for emerging contaminants in water: A review. Chem. Eng. J. 2017, 323, 361–380. [Google Scholar] [CrossRef] [Green Version]
  19. Stackelberg, P.E.; Gibs, J.; Furlong, E.T.; Meyer, M.T.; Zaugg, S.D.; Lippincott, R.L. Efficiency of conventional drinking-water-treatment processes in removal of pharmaceuticals and other organic compounds. Sci. Total Environ. 2007, 377, 255–272. [Google Scholar] [CrossRef]
  20. Zielińska, M.; Bułkowska, K.; Cydzik-Kwiatkowska, A.; Bernat, K.; Wojnowska-Baryła, I. Removal of bisphenol A (BPA) from biologically treated wastewater by microfiltration and nanofiltration. Int. J. Environ. Sci. Technol. 2016, 13, 2239–2248. [Google Scholar] [CrossRef] [Green Version]
  21. Gunnarsson, L.; Adolfsson-Erici, M.; Björlenius, B.; Rutgersson, C.; Förlin, L.; Larsson, D.G.J. Comparison of six different sewage treatment processes-Reduction of estrogenic substances and effects on gene expression in exposed male fish. Sci. Total Environ. 2009, 407, 5235–5242. [Google Scholar] [CrossRef]
  22. Hu, Z.; Wen, X.; Si, X. Pre-ultrafiltration or pre-ozonation for EDCs removal in a combined ultrafiltration and ozonation process. J. Chem. Technol. Biotechnol. 2016, 91, 2929–2934. [Google Scholar] [CrossRef]
  23. Fan, X.; Tao, Y.; Wang, L.; Zhang, X.; Lei, Y.; Wang, Z.; Noguchi, H. Performance of an integrated process combining ozonation with ceramic membrane ultra-filtration for advanced treatment of drinking water. Desalination 2014, 335, 47–54. [Google Scholar] [CrossRef]
  24. Li, S.; Zhang, G.; Wang, P.; Zheng, H.; Zheng, Y. Microwave-enhanced Mn-Fenton process for the removal of BPA in water. Chem. Eng. J. 2016, 294, 371–379. [Google Scholar] [CrossRef] [Green Version]
  25. Zahari, A.M.; Shuo, C.W.; Sathishkumar, P.; Yusoff, A.R.M.; Gu, F.L.; Buang, N.A.; Lau, W.J.; Gohari, R.J.; Yusop, Z. A reusable electrospun PVDF-PVP-MnO2 nanocomposite membrane for bisphenol A removal from drinking water. J. Environ. Chem. Eng. 2018, 6, 5801–5811. [Google Scholar] [CrossRef]
  26. Katibi, K.K.; Yunos, K.F.; Man, H.C.; Aris, A.Z.; Zuhair, M.; Syahidah, R. Recent Advances in the Rejection of Endocrine-Disrupting Compounds from Water Using Membrane and Membrane Bioreactor Technologies: A Review. Polymers 2021, 13, 329. [Google Scholar] [CrossRef] [PubMed]
  27. Garg, A.; Singhania, T.; Singh, A.; Sharma, S.; Rani, S.; Neogy, A.; Yadav, S.R.; Sangal, V.K.; Garg, N. Photocatalytic Degradation of Bisphenol-A using N, Co Codoped TiO2 Catalyst under Solar Light. Sci. Rep. 2019, 9, 1–13. [Google Scholar] [CrossRef] [PubMed]
  28. Becker, D.; Rodriguez-mozaz, S.; Insa, S.; Schoevaart, R.; Barcelo, D.; de Cazes, M.; Belleville, M.P.; Marcano, J.S.; Misovic, A.; Oehlmann, J.; et al. Removal of endocrine disrupting chemicals in wastewater by enzymatic treatment with fungal laccases. Org. Process. Res.Dev. 2017, 21, 480–491. [Google Scholar] [CrossRef]
  29. Hou, J.; Dong, G.; Ye, Y.; Chen, V. Enzymatic degradation of bisphenol-A with immobilized laccase on TiO2 sol-gel coated PVDF membrane. J. Memb. Sci. 2014, 469, 19–30. [Google Scholar] [CrossRef]
  30. Bastami, T.R.; Entezari, M.H. Activated carbon from carrot dross combined with magnetite nanoparticles for the efficient removal of p-nitrophenol from aqueous solution. Chem. Eng. J. 2012, 210, 510–519. [Google Scholar] [CrossRef]
  31. Grassi, M.; Kaykioglu, G.; Belgiorno, V.; Lofrano, G. Emerging Compounds Removal from Wastewater. Green Chem. Sustain. 2012, 15–38. [Google Scholar] [CrossRef]
  32. Saber, S.E.M.; Jamil, S.N.A.M.; Abdullah, L.C.; Choong, T.S.Y.; Ting, T.M. Insights into thep-nitrophenol adsorption by amidoxime-modified poly(acrylonitrile-co-acrylic acid): Characterization, kinetics, isotherm, thermodynamic, regeneration and mechanism study. RSC Adv. 2021, 11, 8150–8162. [Google Scholar] [CrossRef]
  33. Katibi, K.K.; Yunos, K.F.; Man, H.C.; Aris, A.Z.; Zuhair, M.; Nor, M.; Azis, R.S.; Umar, A.M. Contemporary Techniques for Remediating Endocrine-Disrupting Compounds in Various Water Sources: Advances in Treatment Methods and Their Limitations. Polymers 2021, 13, 3229. [Google Scholar] [CrossRef]
  34. Haddad, M.; Oie, C.; Vo Duy, S.; Sauvé, S.; Barbeau, B. Adsorption of micropollutants present in surface waters onto polymeric resins: Impact of resin type and water matrix on performance. Sci. Total Environ. 2019, 660, 1449–1458. [Google Scholar] [CrossRef]
  35. Park, C.M.; Han, J.; Chu, K.H.; Al-Hamadani, Y.A.J.; Her, N.; Heo, J.; Yoon, Y. Influence of solution pH, ionic strength, and humic acid on cadmium adsorption onto activated biochar: Experiment and modeling. J. Ind. Eng. Chem. 2017, 48, 186–193. [Google Scholar] [CrossRef]
  36. Hanigan, D.; Zhang, J.; Herckes, P.; Krasner, S.W.; Chen, C.; Westerhoff, P. Adsorption of N-nitrosodimethylamine precursors by powdered and granular activated carbon. Environ. Sci. Technol. 2012, 46, 12630–12639. [Google Scholar] [CrossRef]
  37. Wirasnita, R.; Hadibarata, T.; Yusoff, A.R.M.; Yusop, Z. Removal of bisphenol a from aqueous solution by activated carbon derived from oil palm empty fruit bunch. Water. Air. Soil Pollut. 2014, 225, 2148. [Google Scholar] [CrossRef]
  38. Chang, K.L.; Hsieh, J.F.; Ou, B.M.; Chang, M.H.; Hseih, W.Y.; Lin, J.H.; Huang, P.J.; Wong, K.F.; Chen, S.T. Adsorption Studies on the Removal of an Endocrine-Disrupting Compound (Bisphenol A) using Activated Carbon from Rice Straw Agricultural Waste. Sep. Sci. Technol. 2012, 47, 1514–1521. [Google Scholar] [CrossRef]
  39. Bolong, N.; Ismail, A.F.; Salim, M.R.; Matsuura, T. A review of the effects of emerging contaminants in wastewater and options for their removal. Desalination 2009, 239, 229–246. [Google Scholar] [CrossRef]
  40. Ahmed, M.B.; Zhou, J.L.; Ngo, H.H.; Guo, W. Adsorptive removal of antibiotics from water and wastewater: Progress and challenges. Sci. Total Environ. 2015, 532, 112–126. [Google Scholar] [CrossRef]
  41. 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. [Google Scholar] [CrossRef] [PubMed]
  42. Tan, X.; Liu, Y.; Zeng, G.; Wang, X.; Hu, X.; Gu, Y.; Yang, Z. Application of biochar for the removal of pollutants from aqueous solutions. Chemosphere 2015, 125, 70–85. [Google Scholar] [CrossRef]
  43. Huggins, T.M.; Haeger, A.; Biffinger, J.C.; Ren, Z.J. Granular biochar compared with activated carbon for wastewater treatment and resource recovery. Water Res. 2016, 94, 225–232. [Google Scholar] [CrossRef] [Green Version]
  44. Ahmed, M.B.; Zhou, J.L.; Ngo, H.H.; Guo, W. Insight into biochar properties and its cost analysis. Biomass Bioenergy 2016, 84, 76–86. [Google Scholar] [CrossRef]
  45. Mohan, D.; Sarswat, A.; Ok, Y.S.; Pittman, C.U. Organic and inorganic contaminants removal from water with biochar, a renewable, low cost and sustainable adsorbent-A critical review. Bioresour. Technol. 2014, 160, 191–202. [Google Scholar] [CrossRef]
  46. Wang, S.; Gao, B.; Zimmerman, A.R.; Li, Y.; Ma, L.; Harris, W.G.; Migliaccio, K.W. Removal of arsenic by magnetic biochar prepared from pinewood and natural hematite. Bioresour. Technol. 2015, 175, 391–395. [Google Scholar] [CrossRef]
  47. Liao, T.; Li, T.; Su, X.; Yu, X.; Song, H.; Zhu, Y.; Zhang, Y. La(OH)3-modified magnetic pineapple biochar as novel adsorbents for efficient phosphate removal. Bioresour. Technol. 2018, 263, 207–213. [Google Scholar] [CrossRef]
  48. Lu, J.; Zhang, C.; Wu, J.; Luo, Y. Adsorptive Removal of Bisphenol A Using N-Doped Biochar Made of Ulva prolifera. Water. Air Soil Pollut. 2017, 228, 1–9. [Google Scholar] [CrossRef]
  49. Heo, J.; Yoon, Y.; Lee, G.; Kim, Y.; Han, J.; Park, C.M. Enhanced adsorption of bisphenol A and sulfamethoxazole by a novel magnetic CuZnFe2O4–biochar composite. Bioresour. Technol. 2019, 281, 179–187. [Google Scholar] [CrossRef] [PubMed]
  50. Yahayu, M.; Abas, F.Z.; Zulkifli, S.E.; Ani, F.N. Utilization of Oil Palm Fiber and Palm Kernel Shell in Various Applications. Sustain. Technol. Manag. Agric. Wastes 2018, 45–56. [Google Scholar] [CrossRef]
  51. Loh, S.K. The potential of the Malaysian oil palm biomass as a renewable energy source. Energy Convers. Manag. 2017, 141, 285–298. [Google Scholar] [CrossRef]
  52. Umar, M.S.; Jennings, P.; Urmee, T. Generating renewable energy from oil palm biomass in Malaysia: The Feed-in Tariff policy framework. Biomass Bioenergy 2014, 62, 37–46. [Google Scholar] [CrossRef] [Green Version]
  53. AI, M. National Biomass Strategy 2020: New Wealth Creation for Malaysia’s Palm Oil Industry; Agensi Inovasi Malaysia: Kuala Lumpur, Malaysia, 2011. [Google Scholar]
  54. Kong, S.H.; Loh, S.K.; Bachmann, R.T.; Zainal, H.; Cheong, K.Y. Palm kernel shell biochar production, characteristics and carbon sequestration potential. J. Oil Palm Res. 2019, 31, 508–520. [Google Scholar] [CrossRef]
  55. Zahraee, S.M.; Golroudbary, S.R.; Shiwakoti, N.; Kraslawski, A.; Stasinopoulos, P. An investigation of the environmental sustainability of palm biomass supply chains via dynamic simulation modeling: A case of Malaysia. J. Clean. Prod. 2019, 237, 117740. [Google Scholar] [CrossRef]
  56. Oliveira, L.C.A.; Rios, R.V.R.A.; Garg, V.K.; Lago, R.M. Clay–iron oxide magnetic composites for the adsorption of contaminants in water. Appl. Clay Sci. 2003, 22, 169–177. [Google Scholar] [CrossRef]
  57. Han, Z.; Sani, B.; Akkanen, J.; Abel, S.; Nybom, I.; Karapanagioti, H.K.; Werner, D. A critical evaluation of magnetic activated carbon’s potential for the remediation of sediment impacted by polycyclic aromatic hydrocarbons. J. Hazard. Mater. 2015, 286, 41–47. [Google Scholar] [CrossRef] [Green Version]
  58. Mohan, D.; Sarswat, A.; Singh, V.K.; Alexandre-Franco, M.; Pittman, C.U. Development of magnetic activated carbon from almond shells for trinitrophenol removal from water. Chem. Eng. J. 2011, 172, 1111–1125. [Google Scholar] [CrossRef]
  59. Petrova, B.; Budinova, T.; Tsyntsarski, B.; Kochkodan, V. Removal of aromatic hydrocarbons from water by activated carbon from apricot stones. Chem. Eng. J. 2010, 165, 258–264. [Google Scholar] [CrossRef]
  60. Xie, J.; Wang, Z.; Wu, D.-Y.; Li, C.-J. [Adsorption of phenol chemicals by surfactant-modified zeolites]. Huan Jing Ke Xue Huanjing Kexue 2012, 33, 4361–4366. [Google Scholar]
  61. Cheng, M.; Jiang, J.; Wang, J.; Fan, J. Highly Salt Resistant Polymer Supported Ionic Liquid Adsorbent for Ultrahigh Capacity Removal of p-Nitrophenol from Water. ACS Sustain. Chem. Eng. 2019, 7, 8195–8205. [Google Scholar] [CrossRef]
  62. Abdelnaeim, M.Y.; El Sherif, I.Y.; Attia, A.A.; Fathy, N.A.; El-Shahat, M.F. Impact of chemical activation on the adsorption performance of common reed towards Cu(II) and Cd(II). Int. J. Miner. Process. 2016, 157, 80–88. [Google Scholar] [CrossRef]
  63. Vassileva, P.S.; Radoykova, T.H.; Detcheva, A.K.; Avramova, I.A.; Aleksieva, K.I.; Nenkova, S.K.; Valchev, I.V.; Mehandjiev, D.R. Adsorption of Ag+ ions on hydrolyzed lignocellulosic materials based on willow, paulownia, wheat straw and maize stalks. Int. J. Environ. Sci. Technol. 2016, 13, 1319–1328. [Google Scholar] [CrossRef] [Green Version]
  64. Rangabhashiyam, S.; Anu, N.; Giri Nandagopal, M.S.; Selvaraju, N. Relevance of isotherm models in biosorption of pollutants by agricultural byproducts. J. Environ. Chem. Eng. 2014, 2, 398–414. [Google Scholar] [CrossRef]
  65. Muhammad, M.; Choong, T.S.Y.; Chuah, T.G.; Yunus, R.; Yap, Y.H.T. Adsorption of β-carotene onto mesoporous carbon coated monolith in isopropyl alcohol and n-hexane solution: Equilibrium and thermodynamic study. Chem. Eng. J. 2010, 164, 178–182. [Google Scholar] [CrossRef]
  66. Yew, Y.P.; Shameli, K.; Miyake, M.; Bt Ahmad Khairudin, N.B.; Bt Mohamad, S.E.; Hara, H.; Bt Mad Nordin, M.F.; Lee, K.X. An eco-friendly means of biosynthesis of superparamagnetic magnetite nanoparticles via marine polymer. IEEE Trans. Nanotechnol. 2017, 16, 1047–1052. [Google Scholar] [CrossRef]
  67. Liu, C.; Tang, Z.; Chen, Y.; Su, S.; Jiang, W. Characterization of mesoporous activated carbons prepared by pyrolysis of sewage sludge with pyrolusite. Bioresour. Technol. 2010, 101, 1097–1101. [Google Scholar] [CrossRef] [PubMed]
  68. Ghani, Z.A.; Yusoff, M.S.; Zaman, N.Q.; Zamri, M.F.M.A.; Andas, J. Optimization of preparation conditions for activated carbon from banana pseudo-stem using response surface methodology on removal of color and COD from landfill leachate. Waste Manag. 2017, 62, 177–187. [Google Scholar] [CrossRef]
  69. Sulaiman, S.; Azis, R.S.; Ismail, I.; Man, H.C.; Rosdi, N. Rapid Adsorption of Magnetite Nanoparticles from Recycled Mill Scale Waste as Potential Adsorbent for Removal of Cu(II) Ions. Solid State Phenom. 2021, 317, 270–275. [Google Scholar] [CrossRef]
  70. Dong, H.; Deng, J.; Xie, Y.; Zhang, C.; Jiang, Z.; Cheng, Y.; Hou, K.; Zeng, G. Stabilization of nanoscale zero-valent iron (nZVI) with modified biochar for Cr(VI) removal from aqueous solution. J. Hazard. Mater. 2017, 332, 79–86. [Google Scholar] [CrossRef] [PubMed]
  71. Dong, H.; Zhang, C.; Hou, K.; Cheng, Y.; Deng, J.; Jiang, Z.; Tang, L.; Zeng, G. Removal of trichloroethylene by biochar supported nanoscale zero-valent iron in aqueous solution. Sep. Purif. Technol. 2017, 188, 188–196. [Google Scholar] [CrossRef]
  72. Zhou, X.; Liu, Y.; Zhou, J.; Guo, J.; Ren, J.; Zhou, F. Efficient removal of lead from aqueous solution by urea-functionalized magnetic biochar: Preparation, characterization and mechanism study. J. Taiwan Inst. Chem. Eng. 2018, 91, 457–467. [Google Scholar] [CrossRef]
  73. Reguyal, F.; Sarmah, A.K.; Gao, W. Synthesis of magnetic biochar from pine sawdust via oxidative hydrolysis of FeCl2 for the removal sulfamethoxazole from aqueous solution. J. Hazard. Mater. 2017, 321, 868–878. [Google Scholar] [CrossRef]
  74. Hamedi, A.; Trotta, F.; Zarandi, M.B.; Zanetti, M.; Caldera, F.; Anceschi, A.; Nateghi, M.R. In situ synthesis of MIL-100(Fe) at the surface of fe3o4@ac as highly efficient dye adsorbing nanocomposite. Int. J. Mol. Sci. 2019, 20, 5612. [Google Scholar] [CrossRef] [Green Version]
  75. Kalderis, D.; Kayan, B.; Akay, S.; Kulaksiz, E.; Gözmen, B. Adsorption of 2,4-dichlorophenol on paper sludge/wheat husk biochar: Process optimization and comparison with biochars prepared from wood chips, sewage sludge and HOG fuel/demolition waste. J. Environ. Chem. Eng. 2017, 5, 2222–2231. [Google Scholar] [CrossRef]
  76. Zhang, X.; Lv, L.; Qin, Y.; Xu, M.; Jia, X.; Chen, Z. Removal of aqueous Cr(VI) by a magnetic biochar derived from Melia azedarach wood. Bioresour. Technol. 2018, 256, 1–10. [Google Scholar] [CrossRef]
  77. Rong, X.; Xie, M.; Kong, L.; Natarajan, V.; Ma, L.; Zhan, J. The magnetic biochar derived from banana peels as a persulfate activator for organic contaminants degradation. Chem. Eng. J. 2019, 372, 294–303. [Google Scholar] [CrossRef]
  78. Zhao, T.; Ma, X.; Cai, H.; Ma, Z.; Liang, H. study on the adsorption of CuFe2O4-loaded corncob biochar for Pb(II). Molecules 2020, 25, 3456. [Google Scholar] [CrossRef]
  79. Fe, O.; Lin, Q. RSC Advances Removal of Cu (II) from aqueous solution using. RSC Adv. 2017, 7, 53135–53144. [Google Scholar] [CrossRef] [Green Version]
  80. Wang, L.; Gan, K.; Lu, D.; Zhang, J. Hydrophilic Fe3O4@C for High-Capacity Adsorption of 2,4-Dichlorophenol. Eur. J. Inorg. Chem. 2016, 2016, 890–896. [Google Scholar] [CrossRef]
  81. Jun, B.M.; Kim, Y.; Han, J.; Yoon, Y.; Kim, J.; Park, C.M. Preparation of activated biochar-supported magnetite composite for adsorption of polychlorinated phenols from aqueous solutions. Water 2019, 11, 1899. [Google Scholar] [CrossRef] [Green Version]
  82. Saleh, S.; Kamarudin, K.B.; Ghani, W.A.W.A.K.; Kheang, L.S. Removal of Organic Contaminant from Aqueous Solution Using Magnetic Biochar. Procedia Eng. 2016, 148, 228–235. [Google Scholar] [CrossRef] [Green Version]
  83. Mohan, D.; Kumar, H.; Sarswat, A.; Alexandre-Franco, M.; Pittman, C.U. Cadmium and lead remediation using magnetic oak wood and oak bark fast pyrolysis bio-chars. Chem. Eng. J. 2014, 236, 513–528. [Google Scholar] [CrossRef]
  84. Shan, D.; Deng, S.; Zhao, T.; Wang, B.; Wang, Y.; Huang, J.; Yu, G.; Winglee, J.; Wiesner, M.R. Preparation of ultrafine magnetic biochar and activated carbon for pharmaceutical adsorption and subsequent degradation by ball milling. J. Hazard. Mater. 2016, 305, 156–163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Su, J.; Lyu, T.; Yi, H.; Bi, L.; Pan, G. Superior arsenate adsorption and comprehensive investigation of adsorption mechanism on novel Mn-doped La2O2CO3 composites. Chem. Eng. J. 2020, 391, 123623. [Google Scholar] [CrossRef]
  86. Li, X.; Wang, C.; Zhang, J.; Liu, J.; Liu, B.; Chen, G. Preparation and application of magnetic biochar in water treatment: A critical review. Sci. Total Environ. 2020, 711, 134847. [Google Scholar] [CrossRef]
  87. Fan, H.; Ma, X.; Zhou, S.; Huang, J.; Liu, Y.; Liu, Y. Highly efficient removal of heavy metal ions by carboxymethyl cellulose-immobilized Fe3O4 nanoparticles prepared via high-gravity technology. Carbohydr. Polym. 2019, 213, 39–49. [Google Scholar] [CrossRef]
  88. Kim, E.; Jung, C.; Han, J.; Her, N.; Park, C.M.; Jang, M.; Son, A.; Yoon, Y. Sorptive removal of selected emerging contaminants using biochar in aqueous solution. J. Ind. Eng. Chem. 2016, 36, 364–371. [Google Scholar] [CrossRef]
  89. Zhang, X.; Qian, J.; Pan, B. Fabrication of Novel Magnetic Nanoparticles of Multifunctionality for Water Decontamination. Environ. Sci. Technol. 2016, 50, 881–889. [Google Scholar] [CrossRef]
  90. Li, M.; Liu, Q.; Guo, L.; Zhang, Y.; Lou, Z.; Wang, Y.; Qian, G. Cu(II) removal from aqueous solution by Spartina alterniflora derived biochar. Bioresour. Technol. 2013, 141, 83–88. [Google Scholar] [CrossRef]
  91. Fang, Q.; Chen, B.; Lin, Y.; Guan, Y. Aromatic and hydrophobic surfaces of wood-derived biochar enhance perchlorate adsorption via hydrogen bonding to oxygen-containing organic groups. Environ. Sci. Technol. 2014, 48, 279–288. [Google Scholar] [CrossRef]
  92. Gupta, V.K.; Nayak, A. Cadmium removal and recovery from aqueous solutions by novel adsorbents prepared from orange peel and Fe2O3 nanoparticles. Chem. Eng. J. 2012, 180, 81–90. [Google Scholar] [CrossRef]
  93. Zhang, Z.B.; Cao, X.H.; Liang, P.; Liu, Y.H. Adsorption of uranium from aqueous solution using biochar produced by hydrothermal carbonization. J. Radioanal. Nucl. Chem. 2013, 295, 1201–1208. [Google Scholar] [CrossRef]
  94. Regmi, P.; Garcia Moscoso, J.L.; Kumar, S.; Cao, X.; Mao, J.; Schafran, G. Removal of copper and cadmium from aqueous solution using switchgrass biochar produced via hydrothermal carbonization process. J. Environ. Manag. 2012, 109, 61–69. [Google Scholar] [CrossRef]
  95. Ahmed, M.B.; Zhou, J.L.; Ngo, H.H.; Guo, W.; Johir, M.A.H.; Sornalingam, K.; Belhaj, D.; Kallel, M. Nano-Fe0 immobilized onto functionalized biochar gaining excellent stability during sorption and reduction of chloramphenicol via transforming to reusable magnetic composite. Chem. Eng. J. 2017, 322, 571–581. [Google Scholar] [CrossRef]
  96. Ahmed, M.B.; Zhou, J.L.; Ngo, H.H.; Guo, W.; Johir, M.A.H.; Belhaj, D. Competitive sorption affinity of sulfonamides and chloramphenicol antibiotics toward functionalized biochar for water and wastewater treatment. Bioresour. Technol. 2017, 238, 306–312. [Google Scholar] [CrossRef] [PubMed]
  97. Ahmed, M.B.; Zhou, J.L.; Ngo, H.H.; Johir, M.A.H.; Sornalingam, K. Sorptive removal of phenolic endocrine disruptors by functionalized biochar: Competitive interaction mechanism, removal efficacy and application in wastewater. Chem. Eng. J. 2018, 335, 801–811. [Google Scholar] [CrossRef] [Green Version]
  98. Duan, S.; Liu, X.; Wang, Y.; Shao, D.; Alharbi, N.S.; Alsaedi, A.; Li, J. Highly efficient entrapment of U(VI) by using porous magnetic Ni0.6Fe2.4O4 micro-particles as the adsorbent. J. Taiwan Inst. Chem. Eng. 2016, 65, 367–377. [Google Scholar] [CrossRef]
  99. Sun, Y.; Zhang, R.; Ding, C.; Wang, X.; Cheng, W.; Chen, C.; Wang, X. Adsorption of U(VI) on sericite in the presence of Bacillus subtilis: A combined batch, EXAFS and modeling techniques. Geochim. Cosmochim. Acta 2016, 180, 51–65. [Google Scholar] [CrossRef]
  100. Zhou, Y.; Lu, P.; Lu, J. Application of natural biosorbent and modified peat for bisphenol a removal from aqueous solutions. Carbohydr. Polym. 2012, 88, 502–508. [Google Scholar] [CrossRef]
  101. Fontecha-Cámara, M.A.; López-Ramón, M.V.; Álvarez-Merino, M.A.; Moreno-Castilla, C. Effect of surface chemistry, solution pH, and ionic strength on the removal of herbicides diuron and amitrole from water by an activated carbon fiber. Langmuir 2007, 23, 1242–1247. [Google Scholar] [CrossRef]
  102. Zhou, Z.; Liu, Y.-G.; Liu, S.-B.; Liu, H.-Y.; Zeng, G.-M.; Tan, X.-F.; Yang, C.-P.; Ding, Y.; Yan, Z.-L.; Cai, X.-X. Sorption performance and mechanisms of arsenic(V) removal by magnetic gelatin-modified biochar. Chem. Eng. J. 2017, 314, 223–231. [Google Scholar] [CrossRef]
  103. Li, M.-F.; Liu, Y.-G.; Zeng, G.-M.; Liu, S.-B.; Hu, X.-J.; Shu, D.; Jiang, L.-H.; Tan, X.-F.; Cai, X.-X.; Yan, Z.-L. Tetracycline absorbed onto nitrilotriacetic acid-functionalized magnetic graphene oxide: Influencing factors and uptake mechanism. J. Colloid Interface Sci. 2017, 485, 269–279. [Google Scholar] [CrossRef]
  104. Liu, S.; Li, M.; Liu, Y.; Liu, N.; Tan, X.; Jiang, L. Supported activated magnetic biochar: Adsorption behavior and mechanism. J. Taiwan Inst. Chem. Eng. 2019, 102, 330–339. [Google Scholar] [CrossRef]
  105. Garg, V.K.; Gupta, R.; Yadav, A.B.; Kumar, R. Dye removal from aqueous solution by adsorption on treated sawdust. Bioresour. Technol. 2003, 89, 121–124. [Google Scholar] [CrossRef]
  106. Tuzen, M.; Sari, A. Biosorption of selenium from aqueous solution by green algae (Cladophora hutchinsiae) biomass: Equilibrium, thermodynamic and kinetic studies. Chem. Eng. J. 2010, 158, 200–206. [Google Scholar] [CrossRef]
  107. Lee, C.G.; Park, J.A.; Choi, J.W.; Ko, S.O.; Lee, S.H. Removal and Recovery of Cr(VI) from Industrial Plating Wastewater Using Fibrous Anion Exchanger. Water. Air Soil Pollut. 2016, 227, 287. [Google Scholar] [CrossRef]
  108. Idrees, M.; Batool, S.; Kalsoom, T.; Yasmeen, S.; Kalsoom, A.; Raina, S.; Zhuang, Q.; Kong, J. Animal manure-derived biochars produced via fast pyrolysis for the removal of divalent copper from aqueous media. J. Environ. Manag. 2018, 213, 109–118. [Google Scholar] [CrossRef] [PubMed]
  109. Haghdoost, G.; Aghaie, H.; Monajjemi, M. Investigation of Langmuir and Freundlich Adsorption Isotherm of Co2+ Ion by Micro Powder of Cedar Leaf. Orient. J. Chem. 2017, 33, 1569–1574. [Google Scholar] [CrossRef] [Green Version]
  110. Boparai, H.K.; Joseph, M.; Carroll, D.M.O. Kinetics and thermodynamics of cadmium ion removal by adsorption onto nano zerovalent iron particles. J. Hazard. Mater. 2011, 186, 458–465. [Google Scholar] [CrossRef] [PubMed]
  111. Khan, Z.H.; Gao, M.; Qiu, W.; Islam, M.S.; Song, Z. Mechanisms for cadmium adsorption by magnetic biochar composites in an aqueous solution. Chemosphere 2020, 246, 125701. [Google Scholar] [CrossRef] [PubMed]
  112. Sounthararajah, D.P.; Loganathan, P.; Kandasamy, J.; Vigneswaran, S. Effects of humic acid and suspended solids on the removal of heavy metals from water by adsorption onto granular activated carbon. Int. J. Environ. Res. Public Health 2015, 12, 10475–10489. [Google Scholar] [CrossRef] [Green Version]
  113. Silas, K.; Ghani, W.A.W.A.K.; Choong, T.S.Y.; Rashid, U. Breakthrough studies of Co3O4 supported activated carbon monolith for simultaneous SO2/NOx removal from flue gas. Fuel Process. Technol. 2018, 180, 155–165. [Google Scholar] [CrossRef]
  114. Edet, U.A.; Ifelebuegu, A.O. Kinetics, Isotherms, and Thermodynamic Modeling of the Adsorption of Phosphates from Model Wastewater Using Recycled Brick Waste. Processes 2020, 8, 665. [Google Scholar] [CrossRef]
  115. Radnia, H. Isotherm and Kinetics of Fe(II) Adsorption onto Chitosan in a Batch Process. Iran. J. Energy Environ. 2011, 2, 250–257. [Google Scholar] [CrossRef]
  116. Mubarak, N.M.; Alicia, R.F.; Abdullah, E.C.; Sahu, J.N.; Haslija, A.B.A.; Tan, J. Statistical optimization and kinetic studies on removal of Zn2+ using functionalized carbon nanotubes and magnetic biochar. J. Environ. Chem. Eng. 2013, 1, 486–495. [Google Scholar] [CrossRef]
  117. Doumer, M.E.; Rigol, A.; Vidal, M.; Mangrich, A.S. Removal of Cd, Cu, Pb, and Zn from aqueous solutions by biochars. Environ. Sci. Pollut. Res. 2016, 23, 2684–2692. [Google Scholar] [CrossRef]
  118. Van Hien, N.; Valsami-Jones, E.; Vinh, N.C.; Phu, T.T.; Tam, N.T.T.; Lynch, I. Effectiveness of different biochar in aqueous zinc removal: Correlation with physicochemical characteristics. Bioresour. Technol. Reports 2020, 11, 100466. [Google Scholar] [CrossRef]
  119. Alyüz, B.; Veli, S. Kinetics and equilibrium studies for the removal of nickel and zinc from aqueous solutions by ion exchange resins. J. Hazard. Mater. 2009, 167, 482–488. [Google Scholar] [CrossRef] [PubMed]
  120. Doǧan, M.; Alkan, M.; Demirbaş, Ö.; Özdemir, Y.; Özmetin, C. Adsorption kinetics of maxilon blue GRL onto sepiolite from aqueous solutions. Chem. Eng. J. 2006, 124, 89–101. [Google Scholar] [CrossRef]
  121. Ho, Y.S. Review of second-order models for adsorption systems. J. Hazard. Mater. 2006, 136, 681–689. [Google Scholar] [CrossRef] [Green Version]
  122. Ahmed, M.B.; Zhou, J.L.; Ngo, H.H.; Guo, W.; Johir, M.A.H.; Sornalingam, K. Single and competitive sorption properties and mechanism of functionalized biochar for removing sulfonamide antibiotics from water. Chem. Eng. J. 2017, 311, 348–358. [Google Scholar] [CrossRef]
  123. Park, Y.; Sun, Z.; Ayoko, G.A.; Frost, R.L. Bisphenol A sorption by organo-montmorillonite: Implications for the removal of organic contaminants from water. Chemosphere 2014, 107, 249–256. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Zhou, N.; Chen, H.; Xi, J.; Yao, D.; Zhou, Z.; Tian, Y.; Lu, X. Biochars with excellent Pb(II) adsorption property produced from fresh and dehydrated banana peels via hydrothermal carbonization. Bioresour. Technol. 2017, 232, 204–210. [Google Scholar] [CrossRef]
  125. Yang, Q.; Gao, M.; Luo, Z.; Yang, S. Enhanced removal of bisphenol A from aqueous solution by organo-montmorillonites modified with novel Gemini pyridinium surfactants containing long alkyl chain. Chem. Eng. J. 2016, 285, 27–38. [Google Scholar] [CrossRef]
  126. Saeed, A.A.H.; Harun, N.Y.; Sufian, S.; Bilad, M.R.; Zakaria, Z.Y.; Jagaba, A.H.; Ghaleb, A.A.S.; Mohammed, H.G. Pristine and Magnetic Kenaf Fiber Biochar for Cd2+ Adsorption from Aqueous Solution. Int. J. Environ. Res. Public Health 2021, 18, 7949. [Google Scholar] [CrossRef] [PubMed]
  127. Srivastava, V.; Sharma, Y.C.; Sillanpää, M. Response surface methodological approach for the optimization of adsorption process in the removal of Cr(VI) ions by Cu2(OH)2CO3 nanoparticles. Appl. Surf. Sci. 2015, 326, 257–270. [Google Scholar] [CrossRef]
  128. Chen, H.; Chen, Z.; Zhao, G.; Zhang, Z.; Xu, C.; Liu, Y.; Chen, J.; Zhuang, L.; Haya, T.; Wang, X. Enhanced adsorption of U(VI) and 241Am(III) from wastewater using Ca/Al layered double hydroxide@carbon nanotube composites. J. Hazard. Mater. 2018, 347, 67–77. [Google Scholar] [CrossRef] [PubMed]
  129. Cazetta, A.L.; Pezoti, O.; Bedin, K.C.; Silva, T.L.; Paesano Junior, A.; Asefa, T.; Almeida, V.C. Magnetic Activated Carbon Derived from Biomass Waste by Concurrent Synthesis: Efficient Adsorbent for Toxic Dyes. ACS Sustain. Chem. Eng. 2016, 4, 1058–1068. [Google Scholar] [CrossRef]
  130. Lingamdinne, L.P.; Choi, Y.L.; Kim, I.S.; Yang, J.K.; Koduru, J.R.; Chang, Y.Y. Preparation and characterization of porous reduced graphene oxide based inverse spinel nickel ferrite nanocomposite for adsorption removal of radionuclides. J. Hazard. Mater. 2017, 326, 145–156. [Google Scholar] [CrossRef]
  131. Li, M.; Liu, H.; Chen, T.; Dong, C.; Sun, Y. Synthesis of magnetic biochar composites for enhanced uranium(VI) adsorption. Sci. Total Environ. 2019, 651, 1020–1028. [Google Scholar] [CrossRef]
  132. Nguyen, V.H.; Van, H.T.; Nguyen, V.Q.; Dam, X.V.; Hoang, L.P.; Ha, L.T.; Ha, L.T. Magnetic Fe3O4 Nanoparticle Biochar Derived from Pomelo Peel for Reactive Red 21 Adsorption from Aqueous Solution. J. Chem. 2020, 2020. [Google Scholar] [CrossRef]
  133. Yang, X.J.; Zhan, J.Y.; Yu, J.; Liu, G.; Huang, R.X. Adsorption features and kinetics of bisphenol A onto magnetic composite organic sepiolite. Chin. J. Environ. Eng. 2016, 10, 3597–3602. [Google Scholar]
  134. Zhao, Y.; Zhang, R.; Liu, H.; Li, M.; Chen, T.; Chen, D.; Zou, X.; Frost, R.L. Green preparation of magnetic biochar for the effective accumulation of Pb(II): Performance and mechanism. Chem. Eng. J. 2019, 375, 122011. [Google Scholar] [CrossRef]
  135. Song, X.; Zhang, Y.; Cao, N.; Sun, D.; Zhang, Z.; Wang, Y.; Wen, Y.; Yang, Y.; Lyu, T. Sustainable chromium (VI) removal from contaminated groundwater using nano-magnetite-modified biochar via rapid microwave synthesis. Molecules 2021, 26, 103. [Google Scholar] [CrossRef]
  136. Yi, Y.; Huang, Z.; Lu, B.; Xian, J.; Tsang, E.P.; Cheng, W.; Fang, J.; Fang, Z. Magnetic biochar for environmental remediation: A review. Bioresour. Technol. 2020, 298, 122468. [Google Scholar] [CrossRef]
  137. Wang, J.; Liao, Z.; Ifthikar, J.; Shi, L.; Chen, Z.; Chen, Z. One-step preparation and application of magnetic sludge-derived biochar on acid orange 7 removal via both adsorption and persulfate based oxidation. RSC Adv. 2017, 7, 18696–18706. [Google Scholar] [CrossRef] [Green Version]
Figure 1. FESEM (field emission scanning electron microscopy) images of NBC (a), MBC before adsorption (b), magnetic biochar after adsorption (c), average particle size distribution before (d), and after adsorption (e).
Figure 1. FESEM (field emission scanning electron microscopy) images of NBC (a), MBC before adsorption (b), magnetic biochar after adsorption (c), average particle size distribution before (d), and after adsorption (e).
Polymers 13 03781 g001
Figure 2. EDX spectra of the as-synthesized NBC (a), and MBC prior to adsorption (b), and after adsorption (c).
Figure 2. EDX spectra of the as-synthesized NBC (a), and MBC prior to adsorption (b), and after adsorption (c).
Polymers 13 03781 g002
Figure 3. Nitrogen adsorption–desorption isotherm.
Figure 3. Nitrogen adsorption–desorption isotherm.
Polymers 13 03781 g003
Figure 4. Magnetization curve for MBC.
Figure 4. Magnetization curve for MBC.
Polymers 13 03781 g004
Figure 5. X-ray diffraction (XRD) patterns of NBC (a) and MBC (b).
Figure 5. X-ray diffraction (XRD) patterns of NBC (a) and MBC (b).
Polymers 13 03781 g005
Figure 6. Fourier transform infrared spectra of NBC (a) and MBC (b).
Figure 6. Fourier transform infrared spectra of NBC (a) and MBC (b).
Polymers 13 03781 g006
Figure 7. Zeta potential of NBC and MBC at different pH.
Figure 7. Zeta potential of NBC and MBC at different pH.
Polymers 13 03781 g007
Figure 8. Effect of solution pH (a), and ionic strength (b) on the adsorption of BPA by NBC and MBC. (BPA concentration: 20 ppm; contact time: 60 min; biochar dose: 0.5 g; and temperature: 293.15 K).
Figure 8. Effect of solution pH (a), and ionic strength (b) on the adsorption of BPA by NBC and MBC. (BPA concentration: 20 ppm; contact time: 60 min; biochar dose: 0.5 g; and temperature: 293.15 K).
Polymers 13 03781 g008
Figure 9. Effects of BC (a: NBC, b: MBC) dose on the removal of BPA. BPA removal efficiency by BC under varying doses: 0.2–0.8 g; BPA concentration: 20 ppm; temperature: 293.15 K; pH: 6.0; and contact time: 60 min.
Figure 9. Effects of BC (a: NBC, b: MBC) dose on the removal of BPA. BPA removal efficiency by BC under varying doses: 0.2–0.8 g; BPA concentration: 20 ppm; temperature: 293.15 K; pH: 6.0; and contact time: 60 min.
Polymers 13 03781 g009
Figure 10. Effects of initial BPA concentrations on the removal efficiency under varying initial concentrations (8–150 ppm) at constant pH (6.0), dosage (0.5 g), and varying contact times (20–240 min).
Figure 10. Effects of initial BPA concentrations on the removal efficiency under varying initial concentrations (8–150 ppm) at constant pH (6.0), dosage (0.5 g), and varying contact times (20–240 min).
Polymers 13 03781 g010
Figure 11. Effects of contact time (20–240 min) on the per cent removal of BPA, at 0.5 g dosage, BPA concentration 20 ppm, temperature (293.15 K), and pH (6.0).
Figure 11. Effects of contact time (20–240 min) on the per cent removal of BPA, at 0.5 g dosage, BPA concentration 20 ppm, temperature (293.15 K), and pH (6.0).
Polymers 13 03781 g011
Figure 12. Freundlich isotherm for BPA adsorption onto NBC (a) and MBC (b) at pH 6.0, dose 0.5 g, contact time 240 min, and initial BPA concentration 50 ppm.
Figure 12. Freundlich isotherm for BPA adsorption onto NBC (a) and MBC (b) at pH 6.0, dose 0.5 g, contact time 240 min, and initial BPA concentration 50 ppm.
Polymers 13 03781 g012
Figure 13. Langmuir isotherm on BPA adsorption onto NBC (a) and MBC (b) at 0.5 g dose, pH 6.0, contact time 240 min, and initial BPA concentration 50 ppm.
Figure 13. Langmuir isotherm on BPA adsorption onto NBC (a) and MBC (b) at 0.5 g dose, pH 6.0, contact time 240 min, and initial BPA concentration 50 ppm.
Polymers 13 03781 g013
Figure 14. (a) Pseudo-first order model for adsorption of BPA by MBC, and (b) pseudo-second order model for adsorption of BPA by MBC.
Figure 14. (a) Pseudo-first order model for adsorption of BPA by MBC, and (b) pseudo-second order model for adsorption of BPA by MBC.
Polymers 13 03781 g014
Figure 15. Thermodynamic plot of lnKL versus 1/T for adsorption of BPA on MBC.
Figure 15. Thermodynamic plot of lnKL versus 1/T for adsorption of BPA on MBC.
Polymers 13 03781 g015
Figure 16. Regeneration of magnetic biochar for BPA removal. Experimental conditions: m = 0.5 g; [BPA] = 20 ppm; and temperature (293.15 K).
Figure 16. Regeneration of magnetic biochar for BPA removal. Experimental conditions: m = 0.5 g; [BPA] = 20 ppm; and temperature (293.15 K).
Polymers 13 03781 g016
Figure 17. Schematic coordination of BPA removal mechanism using biochar.
Figure 17. Schematic coordination of BPA removal mechanism using biochar.
Polymers 13 03781 g017
Table 1. Properties of Bisphenol A (BPA).
Table 1. Properties of Bisphenol A (BPA).
Compound NameLipid-Water Partition Coefficient (Log Kow)Molecular Mass (g/mol.)Chemical StructureMolecular FormulapKa
Bisphenol A3.32228 Polymers 13 03781 i001C15H16O210.1
Table 2. Elemental constituents of magnetic biochar.
Table 2. Elemental constituents of magnetic biochar.
ElementsCOSiFeTotal
Mass ratio (%)32.5227.232.1138.14100.00
Atom ratio (%)52.4032.931.4513.22100.00
Table 3. Physico-chemical properties of neat biochar and magnetic biochar.
Table 3. Physico-chemical properties of neat biochar and magnetic biochar.
MaterialsPore Diameter (Å)Pore Volume (cm3/g)BET Specific Surface Area (m2/g)pHpzc
NBC15.5160.416260536.53984.829
MBC24.4270.442203362.06735.612
Table 4. Adsorption Isotherms variables.
Table 4. Adsorption Isotherms variables.
AdsorbentsFreundlich ModelLangmuir Model
1/nKF(L·mg−1)R2*qmax(mg/g)KL(L·mg−1)RLR2**
NBC1.029749.7189640.889645.4380336073.952070.005090.7535
MBC1.0567911.629470.91954.727014894.016520.0050.85608
Note: 1/n, KF, and R2* were computed from a plot of logqe versus logCe (Equation (5)), while KL, qmax, and R2** were calculated by plotting 1/qe against 1/Ce (Equation (4)), respectively, using the data from Figure 10 and Figure 11.
Table 5. Kinetic study result for BPA adsorption.
Table 5. Kinetic study result for BPA adsorption.
Kinetic ModelsParametersValueLinear Regression
PSO
1 q t = 1 K 2 q e 2 × 1 t + 1 q e
where K2 is rate constant (g·mg−1·min−1)
K2 (g·mg−1·min−1)0.005289 y = 1.556 + 0.0583 x
Comp. qe (mg/g)47.15266
Exp. qe (mg/g)37.75
R20.99376
PFO
ln(qeqt) = ln(qe) − K1t
K1 (min−1)3.50 × 10−6 y = 0.00383 + 8.4 x
Comp. qe (mg/g)12.003837
Exp. qe (mg/g)37.75
R20.8515
NB: PSO is pseudo-second order; PFO is pseudo-first order.
Table 6. Computed thermodynamic variables for BPA adsorption on MBC at varying temperature.
Table 6. Computed thermodynamic variables for BPA adsorption on MBC at varying temperature.
AdsorbateTemperature (K)lnKLΔ (kJ·mol−1)Δ (kJ·mol−1)Δ (J·mol·K−1)
BPA298.150.526093278−1.30409006551.22768066176.1349999
318.151.803593997−4.770684859
338.152.971634746−8.354391819
Table 7. Comparison of specific surface area, magnetic strength, and adsorption capacity of BPA on different adsorbents from previous literature with current study.
Table 7. Comparison of specific surface area, magnetic strength, and adsorption capacity of BPA on different adsorbents from previous literature with current study.
AdsorbentsSurface Area (m2/g)Magnetic Strength (emu/g)Regeneration (%)Adsorption Capacity (mg/g)Reference
ulva prolifera (marine macroalgae)25.43 NDND84.19 [48]
pomelo peel889.8NDND26.25 [133]
magnetic composite sepioliteNA14.1NA36.30[134]
bamboo61.537.6 and 32.6 7.6 and 8.2263.2 [125]
sewage sludge
wheat straw65.03ND 196.91 [135]
dried pineapple84.8912.8334.93101.16 [47]
corn straw313.88 14.5ND46.90 [111]
local reed biomass154.79NDND9.92 [136]
grapefruit peel20.73230.6020229.19[16]
magnetic biochar palm kernel shell362.06736.488212.8537.64Current study
NA = Not Available; ND = Not determined.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Katibi, K.K.; Yunos, K.F.; Man, H.C.; Aris, A.Z.; Mohd Nor, M.Z.; Azis, R.S. An Insight into a Sustainable Removal of Bisphenol A from Aqueous Solution by Novel Palm Kernel Shell Magnetically Induced Biochar: Synthesis, Characterization, Kinetic, and Thermodynamic Studies. Polymers 2021, 13, 3781. https://doi.org/10.3390/polym13213781

AMA Style

Katibi KK, Yunos KF, Man HC, Aris AZ, Mohd Nor MZ, Azis RS. An Insight into a Sustainable Removal of Bisphenol A from Aqueous Solution by Novel Palm Kernel Shell Magnetically Induced Biochar: Synthesis, Characterization, Kinetic, and Thermodynamic Studies. Polymers. 2021; 13(21):3781. https://doi.org/10.3390/polym13213781

Chicago/Turabian Style

Katibi, Kamil Kayode, Khairul Faezah Yunos, Hasfalina Che Man, Ahmad Zaharin Aris, Mohd Zuhair Mohd Nor, and Rabaah Syahidah Azis. 2021. "An Insight into a Sustainable Removal of Bisphenol A from Aqueous Solution by Novel Palm Kernel Shell Magnetically Induced Biochar: Synthesis, Characterization, Kinetic, and Thermodynamic Studies" Polymers 13, no. 21: 3781. https://doi.org/10.3390/polym13213781

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop