Numerous personal care products contain Bisphenol-A (BPA), a hormone disruptor that ultimately finds its way into waterways. A combination of experimental investigations and Monte-Carlo simulations (MCS) were used to explore zerovalent iron nanoparticles (nZVI) for their removal. The nZVI exhibited an absorption peak (λmax) at 373 nm with a mesoporous structure (pore size 2.138 nm), 159.419 m2 g−1 surface area, and crystalline peaks. The adsorption processes were positively influenced by batch parameters. BPA adsorption on nZVI varied with temperature as predicted by Freundlich and Langmuir isotherms, achieving a maximum adsorption capacity (qmax) of 982.13 mgg−1 at 308 K and pH 2. The adsorption process at 303 and 308 K was physisorption, whereas, at 313 K, it was chemisorption suitably described by pseudo-second-order kinetics. The exothermic and spontaneous nature of the adsorption processes were demonstrated by the negative values of enthalpy (−ΔH°) and free energy (−ΔG°) that decreased with increasing temperatures (308–328 K). The density function theory and MCS studies showed that BPA's phenyl ring, isopropyl, and hydroxyl groups interacted electrostatically with nZVI, specifically crystal plane 121 with the most negative adsorption energy (ΔEads), enhancing BPA removal. Through optimized adsorption mechanisms, nZVI can effectively remove BPA from wastewater.

  • Crystalline nZVI were mesoporous with 159.419 m2g−1 surface area.

  • Freundlich and Langmuir's isotherms predicted BPA removal on nZVI with a qmax of 982.13 mg/g.

  • nZVI adsorption of BPA was best described by pseudo-second-order kinetics.

  • DFT and MCS simulations explained electrostatic attraction between BPA and nZVI.

  • In both DFT and MCS, crystal plane 121 of nZVI was the most effective for Bisphenol-A adsorption.

Globally, water pollution is a significant environmental issue, particularly regarding water supply since some pollutants such as Bisphenol-A (BPA) have not been considered for removal in many developing countries (Mashabi et al. 2022; Elwakeel et al. 2023). BPA has been detected in several environmental samples, including groundwater, surface water, sediment, sludge, and agricultural soil, largely due to industrial wastewater discharges and improper waste disposal. Its significant effects on the endocrine system and the central nervous system have serious concerns (Mpatani et al. 2021, 2022; Ahmad et al. 2022; Al-Qadri et al. 2022). As a result of indiscriminate disposal of personal care products, plastics, and food cans, among others into water bodies, BPA is released into the environment, resulting in adverse effects on human health and the ecosystem. BPA is a classified carcinogen known for its hormone-disrupting and estrogenic activities, causing hormonal imbalance, high blood pressure, cardiovascular diseases, and breast cancer (Bhatia & Datta 2019; Cao et al. 2022; Chandrasekar et al. 2022; Ding et al. 2022; Porcar-Santos et al. 2022). In addition, BPA has been shown to cause metabolic disorders, kidney and liver diseases, and prostate cancer in humans (G'omez-Toledano 2022; Molangiri et al. 2022). More particularly, BPA can exacerbate its effects in ecosystems via bioaccumulation and biomagnification (Liu et al. 2021; Torres-García et al. 2022).

It is therefore inevitable to remove BPA from wastewater before its discharge into the environment. There have been various methods for removing BPA from wastewater, including membrane separation, enzymatic degradation, chemical oxidation, microbial degradation, adsorption, ozonization, reverse osmosis, and photo-oxidation (Mpatani et al. 2021; Cao et al. 2022; Torres-García et al. 2022; Yan et al. 2022). Due to its cost-effectiveness, efficiency, simplicity, and convenience, adsorption remains the most efficient, effective, and safest approach for removing emerging pollutants. Various adsorbents both expensive and cost-effective have been deployed for BPA remediation including tripartite magnetic montmorillonite composites, fly ash-derived zeolite modified by β-cyclodextrin, graphene oxide functionalized with amine molecules, Fe3O4 loaded eggshell, mineralized sawdust, powdered activated carbon, ultra-high surface area porous activated carbon derived from asphalt and metal-organic frameworks (Bandura et al. 2021; Konzen et al. 2021; Mpatani et al. 2021; Al-Qadri et al. 2022; Chandrasekar et al. 2022; Okon et al. 2022; Yan et al. 2022).

Among the adsorbents previously applied for BPA removal are nanomaterials owing to their large surface area, adaptive characteristics, vast morphology, and reactive surfaces. Nanomaterials as adsorbents for water treatment have been widely applied because they play a significant role in environmental pollution treatment, particularly water pollution (Dehghani et al. 2020; Bucur et al. 2022). Among the nanomaterials previously applied are silica, nanobentonite intercalated with magnetite and sodium alginate, magnetic nanoparticles functionalized with 2-vinylpyridine, amine-modified magnetic multiwalled carbon nanotubes, polymer-based sugarcane bagasse impregnated with silver nanoparticles, multiwalled carbon nanotubes coated with CoFe2O4 nanoparticles, and sugarcane waste ash mesoporous silica nanoparticles (Bhatia & Datta 2019; Rovani et al. 2020; Al-Musawi et al. 2022; Bucur et al. 2022; El-Sharkawy et al. 2022; Mpatani et al. 2022). Due to its unparalleled properties, zerovalent iron nanoparticles (ZVI) are one of the most important categories of iron-based nanomaterials. The environmental remediation advantages of nZVI are their large surface areas, excellent pollutant degradation ability and affordability. Moreover, some studies have applied nZVI in combination with graphene, chitosan metals such as CoFe2O4, ZnFe2O4, and nZVI-GO-Cu (Dehghani et al. 2020; Rani et al. 2020; Yousefinia et al. 2021).

Hence, this study investigated the effective removal of BPA in simulated wastewater using nZVI synthesized with Terminalia catappa extract. This study explored quantum chemical analysis and Monte-Carlo (MC) simulation to provide unique insights into the stable surface interactions between BPA and nZVI. Molecular dynamics calculations further elucidated the adsorption mechanisms of BPA on nZVI, adding considerable new understanding to the effective removal of BPA.

Synthesis and characterization of ZVI

A modified method described by Azeez et al. (2022a) was used to synthesize iron nanoparticles (nZVI) using one-pot green synthesis. Precisely, 1 mL of T. catappa extract was added to 40 mL of 0.1 M Fe(NO3)2 (Sigma Aldrich, Germany) solution. The mediated nZVI were characterized with UV-Visible (Biobase BK-UV1900 PC spectrometer, China) scanning from 200 to 800 nm and for surface functional biomolecules from 400 to 4,000 cm−1 with Fourier transform Infra-Red (SHIMADZU, FTIR 8400S) spectroscopic techniques. Crystallinity and grain size (Debye–Scherrer formula – Equation (1)) were determined using X-ray diffraction (Empyrean X-ray diffractometer, Thermofisher, Switzerland) scanning 2θ from 0 to 70° with Cu-Kα maintained at λ = 1.541 A° while morphology was characterized with scanning electron microscopy coupled with energy-dispersive X-ray (Phenom PRO X, Netherlands). Surface elemental composition was done using energy-dispersive X-ray fluorescence EDXRF (ARL XTRA Thermoscientific, Switzerland). The surface area, porosity, and shape of the isotherm were determined using Brunauer–Emmett–Teller (BET, Quantachrome Autosorb 1 series, USA) techniques:
(1)

Adsorbate characteristics, preparation, and instrumentation

Analar grade Bisphenol-A (BPA), 98% pure with the molecular formula (C15H16O2), molecular mass 228.29 gmol−1, maximum absorption at 280 nm, and pKa of 9.6 were purchased from Loba Chemie, India. A standard stock solution of 1,000 μg mL−1 BPA and other concentrations ranging from 10 to 50 mg L−1 of BPA were prepared and measured at 280 nm. The measurements were further confirmed with Agilent HPLC 1100 Series.

An Agilent HPLC 1100 Series (Germany) with a binary pump, an auto liquid sampler (ALS), a thermostat column compartment, a multiwavelength UV/Vis detector (MWD), and an online degasser was used to analyze the BPA standard solutions and residual concentrations after adsorption. For the standard stock solution analysis, 20 mL of LiChrosolv Methanol (LC Grade, Germany) was added into a flask containing 25.7 mg BPA and agitated for 2 min before filling to the 1 L mark with the solvent. The BPA samples were transferred into HPLC vials with a stationary phase consisting of a Column (Waters Xbridge C18 (150 × 4.6 mm I.D., 3.5 um), mobile phases A and B were 40% water and 60% methanol, respectively, and run for 7 min at a flow rate of 1 mL min−1 and at a wavelength of 230 nm. The R2 for the calibration curves of standard concentrations analyzed was 0.9979. Residual BPA concentration in the adsorption experiment was similarly run and analyzed.

Determination pH point of zero charge (pHpzc)

As reported by Azeez et al. (2022a), pHpzc of nZVI was determined by adding 0.1 g of nZVI to 50 mL of 0.1 M NaCl and adjusting the pH between 2 and 12 with 0.1 M NaOH or 0.1 M HCl. The final pH values were determined after agitating the solutions for 24 h on an orbital shaker (Stuart SSLI, Barlword Scientific Ltd, Britain). The pHpzc was calculated by plotting the final pH against the initial pH and the intersection point indicates pHpzc.

Batch adsorption experiment

To investigate the concentration-dependent adsorption of BPA on nZVI, 1 gL−1 of nZVI was added to 100 mL of various BPA concentrations (10–50 mg L−1) at 303, 308, and 313 K. Other batch adsorption parameters investigated were dosage (1–5 g L−1), pH (2–12), contact time (0–100 min), and temperature (308–328 K). The agitation was done at 300 rpm for 100 min and the determination of residual BPA concentrations in each investigation was measured at 280 nm. The percentage and quantity of BPA adsorbed at time t were determined using the following equations, respectively:
(2)
(3)
where qt is the quantity adsorbed at a time (mg g−1), Ci is the initial BPA concentration (mg L−1), Cf is the final BPA concentration (mgL−1), Ct is the residual equilibrium BPA concentration (mgL−1), V is the volume of BPA solution (L), and M is the quantity of nZVI (g).

The adsorption study was carried out at the Department of Pure and Applied Chemistry Laboratory, Osun State University, Nigeria.

Adsorption isotherms, kinetics, and thermodynamics

Linear equations for describing adsorption isotherms (Langmuir, Freundlich, Temkin, and Dubinin–Radushkevich), kinetics and mechanisms (pseudo-first-order (PFO), pseudo-second-order (PSO), Elovich, and intraparticle diffusion), and thermodynamics are presented in Table 1. Correlation coefficient (R2) and root mean square error (RMSE, Equation (4)) were used to appropriate the best model:
(4)
qe,means is the mean of measured BPA quantity removed, qe,cal is the calculated BPA quantity removed and n is the number of experimental data.
Table 1

Equations describing adsorption isotherms, kinetics and thermodynamics of BPA removal on mesoporous nZVI

IsothermsKineticsThermodynamics
Langmuir  Pseudo first order   
  Pseudo-second-order   
Freundlich  Elovich   
Temkin  Intraparticle diffusion   
     
Dubinin–Radushkevich     
     
     
IsothermsKineticsThermodynamics
Langmuir  Pseudo first order   
  Pseudo-second-order   
Freundlich  Elovich   
Temkin  Intraparticle diffusion   
     
Dubinin–Radushkevich     
     
     

Desorption/regeneration study

This desorption experiment was used to determine the nZVI reusability for more adsorption after the first cycle. It was done by adding 0.1 g nZVI to 50 mL of 50 mgL−1 BPA. The pH was adjusted to 2 and stirred at 303 K for 30 min before centrifuging at 300 rpm for 100 min. Following drying and desorption with deionized-distilled water, the resulting solution was centrifuged, and the residual BPA concentration was determined. Three additional cycles were done. The desorption percentage was calculated with the following equation:
(5)

Molecular docking

Quantum chemical elucidation of BPA–nZVI interaction

To explain the adsorption of BPA by nZVI from a quantum chemical viewpoint and determine the electronic properties of the adsorbate (BPA) that might be responsible for interaction with the nanoparticles, the frontier molecular orbitals (HOMO and LUMO) electron density isosurfaces of BPA were obtained by density functional theory (DFT) calculation using the popularly known B3LYP/6-31 + G(d,p) model which has been adjudged sufficient for accurate determination of molecular geometries (Parr 1980; Becke 1983; Olasunkanmi et al. 2020, 2021; Waziri et al. 2023). The molecular structure of BPA was carefully constructed on GaussView 5.0 modeling and visualization tool (Dennington & Millam 2009) and then subjected to an unconstrained optimization with frequency calculation to obtain its ground state geometry and minimum energy. The obtained geometry was confirmed to be authentic by the absence of an imaginary value in its vibrational frequency data. All calculations were carried out using the Gaussian 09 program (Frisch et al. 2009).

Monte-Carlo simulations study

An MC simulation visualizes the adsorption of a molecule (adsorbate) on a substrate (adsorbent) by using molecular dynamics calculations (Rahmati & Modarress 2009; Basiony et al. 2019; Olasunkanmi et al. 2020, 2021). Additionally, it provides a clear understanding of what changes may occur in the molecular configuration when an adsorbent interacts with a molecule (adsorbate).

Simulation of the adsorption of BPA from water by synthesized nZVI was conducted using the Monte-Carlo method with a single BPA molecule competing with 50 H2O molecules for surface space. This simulation was carried out using the adsorption locator module of Acceryls Material Studio 2013 (Akkermans et al. 2013). As the synthesized nZVI is mainly made up of iron, the adsorption site is represented by cleaved surfaces of iron, in accordance with cleavage planes revealed by XRD measurements (h k l). Having built a well-constructed surface, the DFT-optimized BPA structure was then mounted and allowed to relax into a more favorable configuration within a 10 Å distance. The simulation was then performed at a fine accuracy level using a COMPASS force field with a smart algorithm. This was done by running a simulated annealing task in five cycles where each cycle consisted of 50,000 steps.

Green synthesis, characterization, and doping of adsorbent

An observed color change from pink to reddish-brown with a maximum absorption (λmax) at 373 nm (Figure 1) indicates the formation of plasmonic nZVI mediated with T. catappa leaf extract. The λmax of nZVI in this study occurred within the ranges of their characteristic peak (Jagathesan & Rajiv 2018; Desalegn et al. 2019; Iqbal et al. 2021; Perveen et al. 2022).
Figure 1

UV-visible spectrum of nZVI mediated with T. catappa.

Figure 1

UV-visible spectrum of nZVI mediated with T. catappa.

Close modal
The biomolecules responsible for reducing, capping and stabilizing nZVI in the leaf extract T. catappa as revealed in the FTIR spectrum (Figure 2) show prominent peaks at 3,451, 1,638, and 1,385 cm−1 corresponding to stretching hydrogen-bonded OH, C = C, and bending OH of phenol, respectively, in nZVI. The peaks at 599 and 511 cm−1 are characteristics of the Fe–O band (Bouafia & Laouini 2020; Jain et al. 2021; Selvaraj et al. 2022). nZVI is known to exhibit both metallic Fe and iron oxide nanoparticles (FeONPs) properties such that the peaks (400–800 cm−1) are related to FeONPs (Dehghani et al. 2020; Adebowale & Egbedina 2021). The presence of these peaks is suggestive of the involvement of polyphenols as the biomolecules responsible for the mediation of nZVI reducing Fe2+ to Fe0 by T. catappa leaf extract. This agrees with previously published reports of Jain et al. (2021), Azeez et al. (2022a, 2023), Selvaraj et al. (2022) and Perveen et al. (2022).
Figure 2

FTIR spectrum of nZVI mediated with T. catappa.

Figure 2

FTIR spectrum of nZVI mediated with T. catappa.

Close modal
The SEM pattern of nZVI (Figure 3(a)) shows clustered-wired particles with particle size 91.15 ± 12.31 nm. The irregularly clustered and aggregated pattern of the nZVI is indicative of their magnetic properties as previously reported by Jain et al. (2021). The elemental composition and oxides in the EDX and EDXRF patterns (Figures 3(b) (A and (B)) show Fe (54%), P (35%), Ca (11%) in A and Fe (20.98%), Zr (0.048), Al (1.299%), P (0.469%), Ca (0.375%), Ni (0.119%), Zn (0.014%), and Pb (0.001%) in B. The higher percentage of Fe in both patterns implies the synthesis of nZVI (Desalegn et al. 2019; Bouafia & Laouini 2020; Jain et al. 2021; Isik et al. 2023).
Figure 3

(a) SEM pattern of nZVI mediated with T. catappa. (b) EDX pattern of T. catappa mediated nZVI (A) elemental composition (B) oxides of elements.

Figure 3

(a) SEM pattern of nZVI mediated with T. catappa. (b) EDX pattern of T. catappa mediated nZVI (A) elemental composition (B) oxides of elements.

Close modal
The distinct crystalline structure of nZVI as revealed by the XRD pattern (Figure 4) showed monoclinic-shaped nZVI nanoparticles with an average grain (particle) size of 24.65 ± 6.54 nm and matching JCPDS card number 00-153-5451. The lattice parameters are a = 12.8270 Å, b = 11.9090 Å, c = 13.3690 Å, and β = 90.830° with 11 notable diffraction peak angles at 2θ = 9.40°, 13.40°, 16.00°, 17.80°, 19.80°, 21.20°, 23.00°, 26.40°, 36.80°, 46.60°, and 53.38° These peaks correspond to hkl crystal planes of (10-1), (002), (021), (121), (022), (122), (311), (230) (025), (533), and (317). The crystalline diffraction angle at 46.60° corresponds to the α-phase of nZVI (Dehghani et al. 2020; Jain et al. 2021). These results are consistent with the reports of Bouafia & Laouini (2020) and Perveen et al. (2022).
Figure 4

XRD pattern of T. catappa-mediated nZVI.

Figure 4

XRD pattern of T. catappa-mediated nZVI.

Close modal
As demonstrated in BET characterization at 77 K, nZVI exhibited textural properties of a type I isotherm (Figure 5(a)) and an N2 adsorption–desorption cycle with a H4 hysteresis loop (Figure 5(b)). There are well-developed pores in microporous and mesoporous regions of nZVI with well-developed surfaces that reside almost exclusively inside the microspores, which, once they have filled with adsorbate, leave little or no external surfaces for further adsorption. Surface area (SBET), total volume, average pore diameter, and micropore surface area (Table 2) are all indicative of mesoporous nZVI, especially with an average pore diameter of 2.138 nm that falls within the 2–50 nm range (Konzen et al. 2021; Selvaraj et al. 2022). A larger surface area was found in this study for T. catappa-mediated nZVI than α-BR–Fe2O3NPs of 75.19 m2/g (Selvaraj et al. 2022).
Table 2

Textural properties of ZVI

BET surface area (m2 g−1159.419 
Micropore surface area (m2 g−1162.046 
BJH total pore volume (cc g−10.081 
BJH average pore diameter (nm) 2.138 
BET surface area (m2 g−1159.419 
Micropore surface area (m2 g−1162.046 
BJH total pore volume (cc g−10.081 
BJH average pore diameter (nm) 2.138 
Figure 5

(a) and (b) N2 adsorption/desorption for isotherm (a) and hysteresis loop (b) of nZVI.

Figure 5

(a) and (b) N2 adsorption/desorption for isotherm (a) and hysteresis loop (b) of nZVI.

Close modal

Influence of operating parameters on BPA adsorption

The pH determines the ionic form of contaminants such as BPA in water while pHpzc reveals the surface charge and the ionic nature of nZVI (Al-Musawi et al. 2022; El-Sharkawy et al. 2022). The pHpzc occurred at pH 3.87 (Figure 6). The influence of pH studied between pH 2 and 12 showed a decline in the percentage adsorption of BPA (Figure 7) from 50.9 at pH 2 to 44.6% at pH 10 on mesoporous nZVI with a pHpzc of 3.87. The surface charge of nZVI at pH 2 where the maximum adsorption of BPA occurred was cationic (<pHpzc 3.87) and the BPA occurred as a molecular and neutral species at this pH, hence the highest adsorption. BPA exists in solution in three forms; BPA, molecular (BPA), and BPA2− due to its phenolic rings (Adebowale & Egbedina 2021; Bandura et al. 2021; Konzen et al. 2021). The decreased adsorption percentage at pH 10 could be due to the repulsion between the surface charge of nZVI which was anionic and the ionic forms of BPA occurring as either BPA or BPA2− at basic pH. Similar observations were reported by Al-Musawi et al. (2022), Chandrasekar et al. (2022), Ding et al. (2022), El-Sharkawy et al. (2022), and Mpatani et al. (2022) on maximum removal of BPA occurring between pH 2 and 7 owing to the nature of adsorbent surface and ionic form of BPA whereas a different result was reported at pH > 7 by Dovi et al. (2021) due to the nature of adsorbent used.
Figure 6

The plot of pH point of zero charge of mesoporous T. catappa-mediated nZVI.

Figure 6

The plot of pH point of zero charge of mesoporous T. catappa-mediated nZVI.

Close modal
Figure 7

pH-dependent percentage adsorption of BPA on mesoporous T. catappa-mediated nZVI. 1 g L−1 of nZVI in 100 mL of 50 mg L−1 BPA, 300 rpm for 100 min between pH 2 and 12 at 303 K.

Figure 7

pH-dependent percentage adsorption of BPA on mesoporous T. catappa-mediated nZVI. 1 g L−1 of nZVI in 100 mL of 50 mg L−1 BPA, 300 rpm for 100 min between pH 2 and 12 at 303 K.

Close modal
The percentage of BPA adsorbed increased with increasing dosage (Figure 8) from 43.3% for 1 g to 91.5% for 5 g. This is attributable to the larger surface area and the presence of more pores/sites for adsorption as the dosage increased (Dovi et al. 2021; Al-Musawi et al. 2022). Similar observations were recorded by Li et al. (2018) and Ding et al. (2022) on the use of mineralized sawdust and Fe/Mn/N co-doped biochar for removing BPA.
Figure 8

Effects of dosage on BPA removal on mesoporous T. catappa-mediated nZVI. 1–5 gL−1 of nZVI in 100 mL of 50 mg L−1 BPA, 300 rpm for 100 min at pH 2 at 303 K.

Figure 8

Effects of dosage on BPA removal on mesoporous T. catappa-mediated nZVI. 1–5 gL−1 of nZVI in 100 mL of 50 mg L−1 BPA, 300 rpm for 100 min at pH 2 at 303 K.

Close modal
The influence of initial BPA concentration from 10 to 50 mg L−1 with simultaneous effects of temperature variations from 303 to 313 K (Figure 9) showed the highest percentage adsorption was recorded for the highest concentration (50 mg L−1) although the increase was not linear at the respective temperature. The increase in percentage adsorption with an increase in concentration could be attributed to the concentration gradient that served as the driving force and the extensive adsorption sites on the nZVI surface, thus, the highest adsorption was not limited by the interaction between BPA and nZVI. Additionally, this implies a high ratio of BPA at low concentrations to nZVI's reactive adsorptive sites, which explains the higher percentage of adsorption with increasing concentration (Dehghani et al. 2020; Dovi et al. 2021; El-Sharkawy et al. 2022; Mpatani et al. 2022).
Figure 9

Effects of initial BPA concentration on percentage adsorption of BPA on mesoporous T. catappa-mediated nZVI. 1 gL−1 of nZVI in 100 mL of 50 mg L−1 BPA, 300 rpm for 100 min at pH 2 at 303 K.

Figure 9

Effects of initial BPA concentration on percentage adsorption of BPA on mesoporous T. catappa-mediated nZVI. 1 gL−1 of nZVI in 100 mL of 50 mg L−1 BPA, 300 rpm for 100 min at pH 2 at 303 K.

Close modal
Contact time is an important parameter in determining adsorption equilibrium and partition between BPA and nZVI. There was an initial increase in the qt of BPA adsorbed which got reduced and thereafter increased before it reached equilibrium at 100 min (Figure 10). The reports by Bandura et al. (2021), El-Sharkawy et al. (2022), and Okon et al. (2022) showed that the BPA adsorption on different adsorbents reached equilibrium at 90, 80 and 70 min, respectively, earlier than in this study. The initial increase could be connected to extensive surface reactivity and empty sites on nZVI that got gradually occupied/saturated as time increased before it reached equilibrium (El-Sharkawy et al. 2022). This is similar to the results of Dehghani et al. (2020) and Chandrasekar et al. (2022).
Figure 10

Quantity of BPA adsorbed per unit time on mesoporous T. catappa-mediated nZVI. 1 g L−1 of nZVI in 100 mL of 50 mg L−1 BPA, 300 rpm between 0 and 100 min at pH 2 at 303 K.

Figure 10

Quantity of BPA adsorbed per unit time on mesoporous T. catappa-mediated nZVI. 1 g L−1 of nZVI in 100 mL of 50 mg L−1 BPA, 300 rpm between 0 and 100 min at pH 2 at 303 K.

Close modal
The inverse relationship between the percentage of BPA adsorbed decreasing from 99 to 97% with an increase in temperature from 308 to 328 K (Figure 11) on mesoporous nZVI is an indication of the exothermic nature of the adsorption process as well as its physical nature. Temperature-dependent reductions in adsorption percentage reflect an exothermic process and physisorption (Mpatani et al. 2021, 2022).
Figure 11

Effects of temperature variation on the adsorption of BPA on mesoporous T. catappa-mediated nZVI. 1 g L−1 of nZVI in 100 mL of 50 mg L−1 BPA, 300 rpm for 100 min at pH 2 between 308 and 328 K.

Figure 11

Effects of temperature variation on the adsorption of BPA on mesoporous T. catappa-mediated nZVI. 1 g L−1 of nZVI in 100 mL of 50 mg L−1 BPA, 300 rpm for 100 min at pH 2 between 308 and 328 K.

Close modal

Adsorption isotherms

An isothermal model comprising Langmuir, Freundlich, Temkin, and Dubinin–Radushkevich were used to experimentally model the adsorption data to provide a description of the adsorption processes (Figure 12(a)–(d)). BPA adsorption capacity, energy requirement, and adsorbent affinity for BPA were accurately predicted using these models. Correlation coefficient (R2) and RMSE are used to rank models according to their relative accuracy as measures of relative model adequacy for predicting the adsorption process (Dehghani et al. 2020). The adsorption process of BPA on mesoporous nZVI with respect to temperature variations was described by Freundlich (R2 = 0.976, RMSE = 0.014) > Temkin (R2 = 0.971, RMSE = 0.017) > Dubinin–Radushkevich (R2 = 0.959, RMSE = 0.023) > Langmuir (R2 = 0.944, RMSE = 0.033) at 303 K (Table 3). At 308 K, the adsorption process was described by Langmuir (R2 = 0.966, RMSE = 0.019) > Dubinin–Radushkevich (R2 = 0.958, RMSE = 0.024) > Freundlich (R2 = 0.937, RMSE = 0.036) > Temkin (R2 = 0.934, RMSE = 0.038) while at 313 K, the description follows Langmuir (R2 = 0.793, RMSE = 0.119) > Freundlich (R2 = 0.655, RMSE = 0.339) > Temkin (R2 = 0.313, RMSE = 0.397) > Dubinin–Radushkevich (R2 = 0.299, RMSE = 0.405) (Table 3). The Freundlich isotherm best describes the adsorption at 303 K implying the surface of nZVI was heterogeneously distributed in terms of pores and pore sizes whereas, at 308 and 313 K, Langmuir isotherm was the most accurate fit for predicting the adsorption of BPA on nZVI indicating that the surface became homogeneously distributed as temperature progressed with pores of equal energy, similar affinities and sizes. Langmuir isotherm has been shown to be the most accurate fitness model for the removal of BPA in similar experiments (Dehghani et al. 2020; Bandura et al. 2021; Al-Musawi et al. 2022; Ding et al. 2022; El-Sharkawy et al. 2022). The maximum monolayer adsorption capacities (qmax) of mesoporous nZVI for the removal of BPA are 832.11, 982.13, and 133.47 mg g−1 at 303, 308, and 313 K, respectively. These values increased from 303 to 308 K but declined from 308 to 313 K (Table 3). As obtained in this study, temperature plays a major factor in increasing BPA mobility by increasing its penetration and chemical interactions with nZVI (Mpatani et al. 2022; Azeez et al. 2023). Mesoporous nZVI is substantially more efficient at removing BPA than other adsorbents previously used comparing their qmax (Table 4). Considering its eco-friendliness, adsorption capability, biodegradability, and ease of synthesis, nZVI represents a preferred adsorbent for removing BPA. Since RL < 1 for BPA removal at all temperatures, this suggests a favorable adsorption process (Azeez et al. 2022b). The Freundlich plots showed that the adsorption strength (n) was greater than 1, indicating an efficient adsorption process involving physical adsorption where the nZVI surface is homogeneous (Al-Musawi et al. 2022; El-Sharkawy et al. 2022).
Table 3

Isotherm models' parameters for the removal of BPA on mesoporous nZVI

IsothermsParameters
303 K308 K313 K
Langmuir  (mg g−1832.11 982.13 133.47 
 KL (L mg−10.048 0.192 0.083 
 RL 0.294 0.094 0.193 
 R2 0.944 0.966 0.793 
 RMSE 0.033 0.019 0.119 
Freundlich N 1.690 2.292 4.042 
 Kf 68.886 39.961 21.264 
 R2 0.976 0.937 0.655 
 RMSE 0.014 0.036 0.339 
Temkin B 20.039 10.444 5.886 
 A (L g−118.880 32.309 38.664 
 b (J mol−1125.708 241.185 427.931 
 R2 0.971 0.934 0.313 
 RMSE 0.017 0.038 0.397 
Dubinin–Radushkevich  (mg g−160.009 37.961 21.606 
 β (mol2 kJ−23.97E − 08 3.24E − 08 1.27E − 09 
 E (kJ mol−13.548 3.929 19.829 
 R2 0.959 0.958 0.299 
 RMSE 0.023 0.024 0.405 
IsothermsParameters
303 K308 K313 K
Langmuir  (mg g−1832.11 982.13 133.47 
 KL (L mg−10.048 0.192 0.083 
 RL 0.294 0.094 0.193 
 R2 0.944 0.966 0.793 
 RMSE 0.033 0.019 0.119 
Freundlich N 1.690 2.292 4.042 
 Kf 68.886 39.961 21.264 
 R2 0.976 0.937 0.655 
 RMSE 0.014 0.036 0.339 
Temkin B 20.039 10.444 5.886 
 A (L g−118.880 32.309 38.664 
 b (J mol−1125.708 241.185 427.931 
 R2 0.971 0.934 0.313 
 RMSE 0.017 0.038 0.397 
Dubinin–Radushkevich  (mg g−160.009 37.961 21.606 
 β (mol2 kJ−23.97E − 08 3.24E − 08 1.27E − 09 
 E (kJ mol−13.548 3.929 19.829 
 R2 0.959 0.958 0.299 
 RMSE 0.023 0.024 0.405 

1 gL−1 of nZVI in 100 mL of 50 mg L−1 BPA, 300 rpm for 100 min at pH 2 at 303 K.

Table 4

Comparison of adsorption capacities of adsorbents for BPA removal

Adsorbent (mg g−1)Reference
Zerovalent iron nanoparticles/graphene oxide/copper 21.59 Yousefinia et al. (2021)  
Tripartite magnetic montmorillonite composites 25.44 Okon et al. (2022)  
N-Bent-NFe3O4-Sod.Alg 28.11 El-Sharkawy et al. (2022)  
Fly ash-derived zeolite modified by β-cyclodextrin 32.70 Bandura et al. (2021)  
Fe3O4 loaded eggshell 35.00 Adebowale & Egbedina (2021)  
Mineralized sawdust 53.43 Ding et al. (2022)  
Chitosan immobilized zerovalent iron nanoparticles 65.16 Dehghani et al. (2020)  
Fe/Mn/N co-doped biochar 82.76 Ding et al. (2022)  
2-Vinylpyridine functionalized magnetic nanoparticles 115.87 Li et al. (2018)  
10 AgNPs-SB-βCD 141.2 Mpatani et al. (2022)  
11 Sugarcane bagasse ash mesoporous silica nanoparticles 155.78 Rovani et al. (2020)  
12 Diethylamine functionalized graphene oxide 334.4 Al-Qadri et al. (2022)  
13 PAC Norit® SAE-super 367.88 Konzen et al. (2021)  
14 MWCNTs/CoFe2O4 416.6 Al-Musawi et al. (2022)  
15 Mesoporous nZVI 982.13 This study 
16 AS 1,113 Javed et al. (2018)  
Adsorbent (mg g−1)Reference
Zerovalent iron nanoparticles/graphene oxide/copper 21.59 Yousefinia et al. (2021)  
Tripartite magnetic montmorillonite composites 25.44 Okon et al. (2022)  
N-Bent-NFe3O4-Sod.Alg 28.11 El-Sharkawy et al. (2022)  
Fly ash-derived zeolite modified by β-cyclodextrin 32.70 Bandura et al. (2021)  
Fe3O4 loaded eggshell 35.00 Adebowale & Egbedina (2021)  
Mineralized sawdust 53.43 Ding et al. (2022)  
Chitosan immobilized zerovalent iron nanoparticles 65.16 Dehghani et al. (2020)  
Fe/Mn/N co-doped biochar 82.76 Ding et al. (2022)  
2-Vinylpyridine functionalized magnetic nanoparticles 115.87 Li et al. (2018)  
10 AgNPs-SB-βCD 141.2 Mpatani et al. (2022)  
11 Sugarcane bagasse ash mesoporous silica nanoparticles 155.78 Rovani et al. (2020)  
12 Diethylamine functionalized graphene oxide 334.4 Al-Qadri et al. (2022)  
13 PAC Norit® SAE-super 367.88 Konzen et al. (2021)  
14 MWCNTs/CoFe2O4 416.6 Al-Musawi et al. (2022)  
15 Mesoporous nZVI 982.13 This study 
16 AS 1,113 Javed et al. (2018)  

AgNPs-SB-βCD – impregnated silver nanoparticles on polymer-based sugarcane bagasse. N-Bent-NFe3O4-Sod.Alg – nanobentonite intercalated with magnetite and sodium alginate. MWCNTs/CoFe2O4 – multiwalled carbon nanotubes coated with CoFe2O4 nanoparticles. PAC Norit® SAE-super – powdered activated carbon. AS – ultra-high surface area porous activated carbon derived from asphalt. ZVI – iron nanoparticles.

Figure 12

(a) Langmuir adsorption isotherm plot for BPA on mesoporous T. catappa-mediated nZVI at 313 K. (b) Freundlich adsorption isotherm plot for BPA on mesoporous T. catappa-mediated nZVI at 303 K. (c) Temkin adsorption isotherm plot for BPA on mesoporous T. catappa-mediated nZVI at 308 K. (d) Dubinin–Radushkevich adsorption isotherm plot for BPA on mesoporous T. catappa-mediated nZVI at 303 K.

Figure 12

(a) Langmuir adsorption isotherm plot for BPA on mesoporous T. catappa-mediated nZVI at 313 K. (b) Freundlich adsorption isotherm plot for BPA on mesoporous T. catappa-mediated nZVI at 303 K. (c) Temkin adsorption isotherm plot for BPA on mesoporous T. catappa-mediated nZVI at 308 K. (d) Dubinin–Radushkevich adsorption isotherm plot for BPA on mesoporous T. catappa-mediated nZVI at 303 K.

Close modal

The energies of adsorption (E) are 3.548, 3.929, and 19.829 kJ mol−1 at 303, 308, and 313 K, respectively, for the removal of BPA obtained from Dubinin–Radushkevich plots demonstrating that physisorption was the underlying mechanism at 303 and 308 K since E < 8 kJ mol−1 at lower temperatures while it was chemisorption at 313 K because it falls between 8 < E < 16 kJ mol−1 (Al-Musawi et al. 2022; Azeez et al. 2022a).

Adsorption kinetics and mechanism

For fitting data regarding the rate-controlling step and kinetics for BPA removal, PFO, PSO, Elovich, and intraparticle diffusion models were used. Similarly, R2 and RMSE are used to rank kinetic models' relative accuracy (Dehghani et al. 2020). Based on the comparison between experimental and calculated qe, the appropriateness of PFO and PSO for the description of kinetics was inferred. For both adsorption kinetics, the PSO kinetic model (Figure 13) fitted the data best, with the highest R2 value, the lowest RMSE, and the highest agreement between experimental and calculated qe (Table 5). In similar reports on adsorption of BPA, PSO has been found more appropriate than PFO (Dehghani et al. 2020; Adebowale & Egbedina 2021; Bandura et al. 2021; Konzen et al. 2021; Al-Musawi et al. 2022; Chandrasekar et al. 2022; Ding et al. 2022). BPA removal was influenced by both adsorption and intramolecular diffusion alongside intraparticle diffusion in the rate-determining step, as evident from the two profiles of intraparticle diffusion obtained from Figure 14 that did not pass through the origin.
Table 5

Adsorption kinetics of BPA on mesoporous nZVI

(mg g−1)1.68
Pseudo-first-order  (mg g−10.25 
 K1 (min−10.04 
 R2 0.611 
 RMSE 0.585 
Pseudo-second-order  (mg g−11.71 
 K2 (g mg−1 min−10.315 
 R2 0.999 
 RMSE 9.71 × 10−3 
Elovich β (mg g−1 min−111.96 
 α (g mg−15.27E06 
 R2 0.670 
Intraparticle diffusion  0.030 
  0.003 
 C1 (mg g−11.422 
 C2 (mg g−11.654 
   0.480 
   0.764 
(mg g−1)1.68
Pseudo-first-order  (mg g−10.25 
 K1 (min−10.04 
 R2 0.611 
 RMSE 0.585 
Pseudo-second-order  (mg g−11.71 
 K2 (g mg−1 min−10.315 
 R2 0.999 
 RMSE 9.71 × 10−3 
Elovich β (mg g−1 min−111.96 
 α (g mg−15.27E06 
 R2 0.670 
Intraparticle diffusion  0.030 
  0.003 
 C1 (mg g−11.422 
 C2 (mg g−11.654 
   0.480 
   0.764 

1 gL−1 of nZVI in 100 ml of 50 mg L−1 BPA, 300 rpm between 0 and 100 min at pH 2 at 303 K.

Figure 13

Pseudo-second-order kinetic plot for BPA on mesoporous T. catappa-mediated nZVI.

Figure 13

Pseudo-second-order kinetic plot for BPA on mesoporous T. catappa-mediated nZVI.

Close modal
Figure 14

The plot of intraparticle diffusion of BPA on mesoporous T. catappa-mediated nZVI.

Figure 14

The plot of intraparticle diffusion of BPA on mesoporous T. catappa-mediated nZVI.

Close modal

Adsorption thermodynamics

The thermodynamic parameters of BPA adsorption indicated an exothermic process (ΔH° = −32.419 kJ mol−1), which is below 40 kJ mol−1, suggesting physical adsorption (Azeez et al. 2022a, 2022b, 2023). Moreover, increased temperature resulted in an increased degree of disorderliness (+ΔS°) on the surface of mesoporous nZVI (Table 6). This is in agreement with the results of Al-Musawi et al. (2022). The ΔG° values for BPA removal vary between −34.851 and −35.009 kJ mol−1 implying that the adsorption process was more spontaneous with an increase in temperature (Dehghani et al. 2020).

Table 6

Thermodynamic parameters of BPA on mesoporous nZVI

Temperature (K)ΔG° (kJ mol−1)ΔH° (kJ mol−1)ΔS° (J K−1 mol−1)
308 −34.851 −32.419 7.898 
313 −34.891   
318 −34.930   
323 −34.970   
328 −35.009   
Temperature (K)ΔG° (kJ mol−1)ΔH° (kJ mol−1)ΔS° (J K−1 mol−1)
308 −34.851 −32.419 7.898 
313 −34.891   
318 −34.930   
323 −34.970   
328 −35.009   

1 gL−1 of nZVI in 100 mL of 50 mgL−1 BPA, 300 rpm for 100 min at pH 2 at 303 K.

Desorption/regeneration

The regeneration of mesoporous nZVI for BPA removal (Figure 15) after four cycles of adsorption–desorption was 53%. As cycles increased, reusability decreased. Above 50% regeneration indicates a high reusability of nZVI, which is also an environmentally friendly adsorbent due to its green synthesis.
Figure 15

Reusability performance of nZVI for BPA desorption in water.

Figure 15

Reusability performance of nZVI for BPA desorption in water.

Close modal

Molecular docking mechanism

DFT calculation

The BPA's optimized geometry is displayed alongside its HOMO and LUMO electron density isosurfaces in Figure 16. The HOMO density isosurface reveals where in the BPA molecule electrons can be donated to nZVI, whereas the LUMO density isosurface reveals where electrons can be accepted from nZVI for adsorption. The results (Figure 16) suggest that the entire parts of the BPA molecule (i.e., the isopropyl moiety, the phenyl rings and the hydroxyl substituents) would be responsible for electron donation to nZVI surface while only the phenyl rings of this molecule would be concerned with back donation from nZVI.
Figure 16

Optimized geometry of BPA (a), HOMO electron density map (b), and LUMO electron density map (c).

Figure 16

Optimized geometry of BPA (a), HOMO electron density map (b), and LUMO electron density map (c).

Close modal
The fraction of electrons transferred (ΔN) between BPA and nZVI was calculated as 0.70 using the relation:
(6)
where and , the electronegativity and absolute hardness of iron are assigned values of 7 and 0 eV mol−1, respectively (Basiony et al. 2019) and the electronegativity () and absolute hardness () of BPA were calculated from its HOMO and LUMO energies using (Wahab et al. 2018; Waziri et al. 2023):
(7)
(8)

Since ΔN > 0, the adsorption of BPA onto nZVI involves the movement of electrons from BPA to nZVI while the opposite would have been the case if ΔN < 0 (Basiony et al. 2019).

Monte-Carlo simulations

The XRD data presented in Table 7 show a number of h k l combinations corresponding to different cleavage planes and particle sizes which are all within the nano range. Since each h k l combination represents a different surface morphology of the synthesized nZVI, it would be interesting to compare the different surfaces (i.e., h k l combinations) of the nZVI in terms of adsorption capacity as this will help to predict the h k l combination with the highest potential for the removal of BPA. To achieve this, the calculation was limited to only the h k l combinations that produce a uniform surface with a relatively large atomic density. These are 1 0 −1, 1 2 1, and 0 0 2. The 1 1 0 surface which is widely reported as the most stable Fe surface (Anderson & Mehandru 1984; Arya & Carter 2004; Guo et al. 2017; Basiony et al. 2019; Olasunkanmi et al. 2021) was also included for comparison.

Table 7

ΔEads for adsorption of a single BPA molecule both in isolation and in competition with 50 molecules of water by the selected surfaces

h k lΔEads (kJ/mol)
in gasin water
1 0 −1 −453.03 −1,193.67 
0 0 2 −410.93 −1,811.36 
1 2 1 −439.18 −2,038.76 
1 1 0 −453.02 −1,183.24 
h k lΔEads (kJ/mol)
in gasin water
1 0 −1 −453.03 −1,193.67 
0 0 2 −410.93 −1,811.36 
1 2 1 −439.18 −2,038.76 
1 1 0 −453.02 −1,183.24 

Representative results of MC simulation of adsorption of a single BPA molecule by the 1 2 1 surface both in the absence and presence of 50 molecules of water are shown in Figures 17 and 18. The adsorption energies (ΔEads) associated with each surface are listed in Table 6. A negative ΔEads value suggests that the adsorption process is favorable and the more negative the ΔEads, the stronger the binding and the better the favorability of the process. Based on this interpretation, the results in Table 7 clearly show that surfaces corresponding to h k l 1 0 −1, 1 2 1, and 0 0 2 are better adsorption sites for the removal of BPA from water than the popularly known 1 1 0 surface. The 1 2 1 cleaved surface yielded the most negative ΔEads value and is thus suspected to be the major surface responsible for the adsorption of BPA from water. However, the trend observed in the gas phase suggests that the 1 0 −1 and 1 1 0 surfaces are the best adsorption surfaces for the removal of gaseous molecules of BPA.
Figure 17

Adsorbed molecule of BPA on Fe(1 2 1) surface in absence and presence of 50 molecules of H2O as revealed by Monte-Carlo simulation.

Figure 17

Adsorbed molecule of BPA on Fe(1 2 1) surface in absence and presence of 50 molecules of H2O as revealed by Monte-Carlo simulation.

Close modal
Figure 18

Adsorbed molecule of BPA on surfaces 1 0 −1 (a), 0 0 2 (b), and 1 1 0 (c) in the absence and presence of 50 molecules of H2O as revealed by Monte-Carlo simulation.

Figure 18

Adsorbed molecule of BPA on surfaces 1 0 −1 (a), 0 0 2 (b), and 1 1 0 (c) in the absence and presence of 50 molecules of H2O as revealed by Monte-Carlo simulation.

Close modal

Zerovalent iron nanoparticles (nZVI) mediated with T. catappa extract absorbed maximally at 373 nm, characteristic of iron nanoparticles with FTIR peaks indicating Fe–O, C = O, and -OH bands. Clustered-wired particles with a particle size of 91.15 ± 12.31 nm were observed in the SEM micrograph and EDX showed prominent Fe and FeO presence. For nZVI in this study, seven crystalline planes matching JCPDS card number 00-153-5451 were observed. The T. catappa-mediated nZVI had a surface area of 159.419 m2 g−1 and a pore size of 2.138 nm, suggesting a mesoporous surface. The pHpzc of nZVI was 3.87, and maximum adsorption of BPA occurred at pH 2, decreasing with increasing pH. Other batch adsorption parameters significantly affected BPA removal. BPA adsorption was predicted by Freundlich isotherm at 303 K, and Langmuir isotherm at 308 and 313 K. In this study, BPA adsorption was favorable with n values > 1 and RL < 1. The adsorption at 303 and 308 K was physisorption, whereas the adsorption at 313 K was chemisorption. It was found that pseudo-second-order kinetics was the most accurate for predicting the adsorption rate, whereas intraparticle diffusion, intramolecular diffusion, and adsorption were the mechanisms controlling the BPA removal. The adsorption was spontaneous as demonstrated by –ΔG°, decreasing with increasing temperature; exothermic with decreased order of disorderliness as shown by –ΔH° and ΔS°. The DFT and MSC studies revealed that the isopropyl moiety of BPA, the phenyl rings, and the hydroxyl substituents were involved in electrostatic interaction with nZVI with crystal plane 121 having the most negative ΔEads responsible for most BPA removal. The nZVI mediated with T. catappa is a more promising adsorbent for removing BPA due to its over 50% reusability and adsorption capacity.

All relevant data are included in the paper or its Supplementary Information.

The authors declare there is no conflict.

Adebowale
K. O.
&
Egbedina
A. O.
2021
Facile green synthesis of bio-carbon material from eggshells and its application for the removal of Bisphenol A and 2,4,6-trichlorophenol from water
.
Environ. Nanotechnol. Monit. Manag.
17
,
100622
.
https://doi.org/10.1016/j.enmm.2021.100622
.
Ahmad
A.
,
Priyadarshini
M.
,
Yadav
S.
,
Ghangrekar
M.
&
Surampalli
Y.
2022
The potential of biochar-based catalysts in advanced treatment technologies for efficacious removal of persistent organic pollutants from wastewater: A review
.
Chem. Eng. Res. Des.
187
,
470
496
.
https://doi.org/10.1016/j.cherd.2022.09.024
.
Akkermans
R. L.
,
Spenley
N. A.
&
Robertson
S. H.
2013
Monte Carlo methods in materials studio
.
Mol. Simul.
39
(
14–15
),
1153
1164
.
https://doi.org/10.1080/08927022.2013.843775
.
Al-Musawi
T. J.
,
Mengelizadeh
N.
,
Ganji
F.
,
Wang
C.
&
Balarak
D.
2022
Preparation of multi-walled carbon nanotubes coated with CoFe2O4 nanoparticles and their adsorption performance for Bisphenol A compound
.
Adv. Powder Technol.
33
,
103438
.
https://doi.org/10.1016/j.apt.2022.103438
.
Al-Qadri
Q. A.
,
Drmosh
A.
&
Onaizi
A. Q.
2022
Enhancement of Bisphenol A removal from wastewater via the covalent functionalization of graphene oxide with short amine molecules
.
Case Stud. Chem. Environ. Eng.
6
,
100233
.
https://doi.org/10.1016/j.cscee.2022.100233
.
Anderson
A. B.
&
Mehandru
S. P.
1984
Acetylene adsorption to Fe (100), (110), and (111) surfaces; structures and reactions
.
Surf. Sci.
136
(
2–3
),
398
418
.
https://doi.org/10.1016/0039-6028(84)90620-4
.
Arya
A.
&
Carter
E. A.
2004
Structure, bonding, and adhesion at the ZrC (1 0 0)/Fe (1 1 0) interface from first principles
.
Surf. Sci.
560
(
1–3
),
103
120
.
https://doi.org/10.1016/j.susc.2004.04.022
.
Azeez
L.
,
Adebisi
S. A.
,
Adejumo
A. L.
,
Busari
H. K.
,
Aremu
H. K.
,
Olabode
O. A.
&
Awolola
O.
2022a
Adsorptive properties of rod-shaped silver nanoparticles-functionalized biogenic hydroxyapatite for remediating methylene blue and Congo red
.
Inorg. Chem. Commun.
142
,
109655
.
https://doi.org/10.1016/j.inoche.2022.109655
.
Azeez
L.
,
Adetoro
R. O.
,
Busari
H. K.
,
Aremu
H. K.
,
Adeleke
J. T.
,
Adewinbi
S.
,
Olabode
O. A.
&
Ayandayo
I.
2022b
AgNPs-TiO2NPs doped calcined hydroxyapatite for effective removal of ibuprofen and Acetaminophen
.
Int. J. Environ. Anal. Chem.
https://doi.org/10.1080/03067319.2022.2106434
.
Azeez
L.
,
Adeleke
A. E.
,
Popoola
S. A.
,
Busari
H. K.
,
Agbaje
W. B.
,
Ojewuyi
S. S.
,
Oluwafemi
I. I.
,
Oladeji
R. D.
&
Adetoro
O. R.
2023
Dye degradation and molecular docking mechanism of Terminalia catappa mediated mesoporous titanium dioxide nanoparticles
.
Inorg. Chem. Commun.
153
,
110873
.
https://doi.org/10.1016/j.inoche.2023.110873
.
Becke
A. D.
1983
Numerical Hartree–Fock–Slater calculations on diatomic molecules: Addendum
.
J. Chem. Phys.
78
(
7
),
4787
4788
.
https://doi.org/10.1063/1.445285
.
Bhatia
D.
&
Datta
D.
2019
Removal of Bisphenol-A using amine-modified magnetic multiwalled carbon nanotubes: Batch and column studies
.
J. Chem. Eng.
64
(
6
),
2877
2887
.
https://doi.org/10.1021/acs.jced.9b00240
.
Bucur
S.
,
Diacon
A.
,
Mangalagiu
I.
,
Mocanu
A.
,
Rizea
F.
,
Dinescu
A.
,
Ghebaur
A.
,
Boscornea
A. C.
,
Voicu
G.
&
Rusen
E.
2022
Bisphenol-A adsorption on silica particles modified with beta-cyclodextrins
.
Nanomaterials.
12
,
39
.
https://doi.org/10.3390/nano12010039
.
Cao
T.
,
Zhou
D.
,
Wang
X.
&
Cui
C.
2022
Photocatalytic degradation of Bisphenol A over Co-BiOCl/biochar hybrid catalysts: Properties, efficiency and mechanism
.
J. Mol. Liq.
362
,
119622
.
https://doi.org/10.1016/j.molliq.2022.119622
.
Chandrasekar
R.
,
Rajendran
H. K.
,
Priyan
V. V.
&
Narayanasamy
S.
2022
Valorization of sawdustby mineral acid assisted hydrothermal carbonization for the adsorptive removal of Bisphenol A: A greener approach
.
Chemosphere.
303
,
135171
.
https://doi.org/10.1016/j.chemosphere.2022.135171
.
Dehghani
M. H.
,
Karri
R. R.
,
Alimohammadi
M.
,
Nazmara
S.
,
Zarei
A.
&
Saeedi
Z.
2020
Insights into endocrine-disrupting Bisphenol-A adsorption from pharmaceutical effluent by chitosan immobilized nanoscale zero-valent iron nanoparticles
.
J. Mol. Liq.
https://doi.org/10.1016/j.molliq.2020.113317
.
Dennington
R. T. A. K.
&
Millam
J. M.
2009
GaussView, Version 5.0
.
Semichem Inc.
,
Shawnee Mission, KS
.
Desalegn
B.
,
Megharaj
M.
,
Chen
Z.
&
Naidu
R.
2019
Green synthesis of zero valent iron nanoparticle using mango peel extract and surface characterization using XPS and GC-MS
.
Heliyon.
5
(
5
),
e01750
.
https://doi.org/10.1016/j.heliyon.2019.e01750
.
Ding
T.
,
Cai
M.
,
Wu
C.
,
Bao
L.
&
Li
J.
2022
Distribution profiles of bisphenols in school supplies and implications for human exposure
.
Sci. Total Environ.
849
,
157938
.
http://dx.doi.org/10.1016/j.scitotenv.2022.157938
.
Dovi
E.
,
Kani
A. N.
,
Aryee
A. A.
,
Jie
M.
,
Li
J.
,
Li
Z.
,
Qu
L.
&
Han
R.
2021
Decontamination of Bisphenol A and Congo red dye from solution by using CTAB functionalised walnut shell
.
Environ. Sci. Pollut. Res.
28
,
28732
28749
.
https://doi.org/10.1007/s11356-02112550-4
.
El-Sharkawy
R. M.
,
Allam
E. A.
,
Ali
A. S. M.
&
Mahmoud
M. E.
2022
Adsorption study of Bisphenol A and Chlorpyrifos onto nanobentonite intercalated with magnetite and sodium alginate: Kinetics and isotherm models
.
Int. J. Environ. Sci. Technol.
19
,
9827
9842
.
https://doi.org/10.1007/s13762-021-03815-z
.
Elwakeel
K. Z.
,
Ahmed
M. M.
,
Akhdhar
A.
,
Alghamdi
H. M.
,
Sulaiman
M. G. M.
,
Hamza
M. F.
&
Khan
Z. A.
2023
Effect of the magnetic core in alginate/gum composite on adsorption of divalent copper, cadmium, and lead ions in the aqueous system
.
Int. J. Biol. Macromol.
253
,
126884
.
https://doi.org/10.1016/j.ijbiomac.2023.126884
.
Frisch
M.
,
Trucks
G.
,
Schlegel
H.
,
Scuseria
G.
,
Robb
M.
&
Cheeseman
J.
2009
Gaussian 09, Revision B. 01. Gaussian 09, Revision B. 01
.
Gaussian Inc
,
Wallingford, CT
.
G'omez-Toledano
R. M.
2022
Relationship between emergent BPA-substitutes and renal and cardiovascular diseases in adult population
.
Environ. Pollut.
313
,
120106
.
https://doi.org/10.1016/j.envpol.2022.120106
.
Guo
L.
,
Qi
C.
,
Zheng
X.
,
Zhang
R.
,
Shen
X.
&
Kaya
S.
2017
Toward understanding the adsorption mechanism of large size organic corrosion inhibitors on an Fe (110) surface using the DFTB method
.
RSC Adv.
7
(
46
),
29042
29050
.
https://doi.org/10.1039/C7RA04120A
.
Iqbal
J.
,
Shah
N. S.
,
Sayed
M.
,
Khan
J. A.
,
Imran
M.
,
Khan
Z. U.
,
Niazi
N. K.
,
Al-Taani
A. A.
,
Howari
F.
&
Nazzal
Y.
2021
Exploring the potential of nano-zerovalent copper modified biochar for the removal of ciprofloxacin from water
.
Environ. Nanotechnol. Monit. Manag.
16
,
100604
.
https://doi.org/10.1016/j.enmm.2021.100604
.
Isik
Z.
,
Bouchareb
R.
,
Arslan
H.
,
Özdemir
S.
,
Gonca
S.
,
Dizge
N.
,
Balakrishnan
D.
&
Prasad
S. V. S.
2023
Green synthesis of iron oxide nanoparticles derived from water and methanol extract of Centaurea solstitialis leaves and tested for antimicrobial activity and dye decolorization capability
.
Environ. Res.
219
,
115072
.
https://doi.org/10.1016/j.envres.2022.115072
.
Jagathesan
G.
&
Rajiv
P.
2018
Biosynthesis and characterization of iron oxide nanoparticles using Eichhornia crassipes leaf extract and assessing their antibacterial activity
.
Biocatal. Agric. Biotechnol.
13
,
90
94
.
https://doi.org/10.1016/j.bcab.2017.11.014.
Jain
R.
,
Mendiratta
S.
,
Kumar
L.
&
Srivastava
L.
2021
Green synthesis of iron nanoparticles using Artocarpus heterophyllus peel extract and their application as a heterogeneous Fenton-like catalyst for the degradation of Fuchsin basic dye
.
Curr. Res. Green Sustain. Chem.
4
,
100086
.
https://doi.org/10.1016/j.crgsc.2021.100086
.
Javed
H.
,
Luong
D. X.
,
Lee
C-G.
,
Zhang
D.
,
Tour
J. M.
&
Alvarez
P. J. J.
2018
Efficient removal of bisphenol-A by ultra-high surface area porous activated carbon derived from asphalt
.
Carbon
140
,
441
448
.
https://doi.org/10.1016/j.carbon.2018.08.038
.
Konzen
R. A.
,
Batista
P. R.
,
Ducati
L. C.
,
Souza
T. E. A.
,
Cavalcante
L. C.
,
Santos
C. E.
,
Bassetti
F. J.
,
Rodrigues
P. C.
&
Coral
L. A.
2021
New insights into Bisphenol A removal from water applying experimental and theoretical studies
.
Desalination Water Treat.
227
,
238
249
.
https://doi.org/10.5004/dwt.2021.27302
.
Li
Q.
,
Pan
F.
,
Li
W.
,
Li
D.
,
Xu
H.
,
Xia
D.
&
Li
A.
2018
Enhanced adsorption of Bisphenol A from aqueous solution with 2-vinylpyridine functionalized magnetic nanoparticles
.
Polymer
12
(
10
),
1136
.
https://doi.org/10.3390/polym10101136
.
Liu
J.
,
Zhang
L.
,
Lu
G.
,
Jiang
J.
,
Yan
Z.
&
Li
Y.
2021
Occurrence, toxicity and ecological risk of Bisphenol A analogues in aquatic environment – a review ecotoxicol
.
Environ. Saf.
208
,
111481
.
https://doi.org/10.1016/j.ecoenv.2020.111481
.
Mashabi
R. A.
,
Khan
Z. A.
&
Elwakeel
K. Z.
2022
Chitosan- or glycidyl methacrylate-based adsorbents for the removal of dyes from aqueous solutions: A review
.
Mater. Adv.
3
,
5645
5671
.
https://doi.org/10.1039/D2MA00320A
.
Molangiri
A.
,
Varma
S.
,
Satyavani
M.
,
Kambham
S.
,
Duttaroy
A. K.
&
Basak
K.
2022
Prenatal exposure to Bisphenol S and Bisphenol A differentially affects male reproductive system in the adult offspring
.
Food Chem. Toxicol.
167
,
113292
.
https://doi.org/10.1016/j.fct.2022.113292
.
Mpatani
F. M.
,
Han
A.
,
Aryee
A. A.
,
Kani
N. A.
,
Li
Z.
&
Qu
L.
2021
A comprehensive review adsorption performance of modified agricultural waste materials for removal of emerging micro-contaminant bisphenol
.
Sci. Total Environ.
780
,
146629
.
https://doi.org/10.1016/j.scitotenv.2021.146629
.
Mpatani
F. M.
,
Kombo
U. M.
,
Kheir
K. R.
,
Nassor
N. K.
,
Joshua
D.
,
Mussa
S. S.
,
Mohamed
S. A.
,
Mbarak
S. A.
,
Hamad
A. S.
&
Ahmada
S. A.
2022
Impregnation of silver nanoparticles onto polymers based on sugarcane bagasse for the remediation of endocrine disruptor – Bisphenol A from water
.
Adsorpt. Sci. Technol.
4997205
.
https://doi.org/10.1155/2022/4997205
.
Okon
O. E.
,
Inam
E. J.
,
Offiong
A.-A.
&
Akpabio
U. D.
2022
Aqueous adsorptive removal of Bisphenol A using tripartite magnetic montmorillonite composites
.
Pollutants.
2
(
3
),
363
387
.
https://doi.org/10.3390/pollutants2030025
.
Olasunkanmi
L. O.
,
Idris
A. O.
,
Adewole
A. H.
,
Wahab
O. O.
&
Ebenso
E. E.
2020
Adsorption and corrosion inhibition potentials of salicylaldehyde-based Schiff bases of semicarbazide and p toluidine on mild steel in acidic medium: Experimental and computational studies
.
Surf. Interfaces
21
,
100782
.
https://doi.org/10.1016/j.surfin.2020.100782
.
Olasunkanmi
L. O.
,
Aniki
N. I.
,
Adekunle
A. S.
,
Durosinmi
L. M.
,
Durodola
S. S.
,
Wahab
O. O.
&
Ebenso
E. E.
2021
Investigating the synergism of some hydrazine carboxamides and iodide ions as corrosion inhibitor formulations for mild steel in hydrochloric acid: Experimental and computational studies
.
J. Mol. Liq.
343
,
117600
.
https://doi.org/10.1016/j.molliq.2021.117600
.
Parr
R. G.
,
1980
Density functional theory of atoms and molecules
. In:
Horizons of Quantum Chemistry. Académie Internationale Des Sciences Moléculaires Quantiques/International Academy of Quantum Molecular Science
, vol
3
. (
Fukui
K.
&
Pullman
B.
, eds).
Springer
,
Dordrecht
.
https://doi.org/10.1007/978 94009-9027-2_2
.
Perveen
S.
,
Nadeem
R.
,
ur Rehman
S.
,
Afzal
N.
,
Anjum
S.
,
Noreen
S.
,
Saeed
R.
,
Amami
M.
,
Al-Mijalli
S. H.
&
Iqbal
M.
2022
Green synthesis of iron (Fe) nanoparticles using Plumeria obtusa extract as a reducing and stabilizing agent: Antimicrobial, antioxidant and biocompatibility studies
.
Arab. J. Chem.
15
(
5
),
103764
.
https://doi.org/10.1016/j.arabjc.2022.103764
.
Porcar-Santos
O.
,
Cruz-Alcalde
A.
,
Bayarri
B.
&
Sans
B.
2022
Reactions of Bisphenol F and Bisphenol S with ozone and hydroxyl radical kinetics and mechanisms
.
Sci. Total Environ.
846
,
157173
.
http://dx.doi.org/10.1016/j.scitotenv.2022.157173
.
Rahmati
M.
&
Modarress
H.
2009
Nitrogen adsorption on nanoporous zeolites studied by grand canonical Monte Carlo simulation
.
J. Mol. Struct. Theochem.
901
,
110
116
.
https://doi.org/10.1016/j.theochem.2009.01.010
.
Rani
M.
,
Rachna
&
Shanker
U.
2020
Efficient photocatalytic degradation of Bisphenol A by metal ferrites nanoparticles under sunlight
.
Environ. Technol. Innov.
19
,
100792
.
https://doi.org/10.1016/j.eti.2020.100792
.
Rovani
S.
,
Jonnatan
J.
,
Santos
B.
,
Sabine
N.
,
Guilhen
A.
,
Coriob
P.
&
Fungaro
A. D.
2020
Fast, efficient and clean adsorption of Bisphenol-A using renewable mesoporous silica nanoparticles from sugarcane waste ash
.
RSC Adv.
10
,
27706
27712
.
https://doi.org/10.1039/D0RA05198E
.
Selvaraj
R.
,
Pai
S.
,
Vinayagam
R.
,
Varadavenkatesan
T.
,
Kumar
P. S.
,
Duc
P. A.
&
Rangasamy
G.
2022
A recent update on green synthesized iron and iron oxide nanoparticles for environmental applications
.
Chemosphere.
308
,
136331
.
https://doi.org/10.1016/j.chemosphere.2022.136331
.
Torres-García
J. L.
,
Ahuactzin-Ṕerez
M.
,
Ferńandez
F. J.
&
Cort́es-Espinos
D. V.
2022
Bisphenol A in the environment and recent advances in biodegradation by fungi
.
Chemosphere
303
,
134940
.
https://doi.org/10.1016/j.chemosphere.2022.134940
.
Wahab
O. O.
,
Olasunkanmi
L. O.
,
Govender
K. K.
&
Govender
P. P.
2018
Synergistic effect of opposite polar substituents on selected properties of disperse yellow 119 dye
.
Chem. Phys. Lett.
704
,
55
61
.
https://doi.org/10.1016/j.cplett.2018.05.025
.
Waziri
I.
,
Wahab
O. O.
,
Mala
G. A.
,
Oselusi
S. O.
,
Egieyeh
S. A.
&
Nasir
H.
2023
Zinc (II) complex of (Z)-4-((4-nitrophenyl) amino) pent-3-en-2-one, a potential antimicrobial agent: Synthesis, characterization, antimicrobial screening, DFT calculation and docking study
.
Bull. Chem. Soc. Ethiopia.
37
(
3
),
633
651
.
https://doi.org/10.4314/bcse.v37i3.8
.
Yan
C.
,
Cao
L.
,
Liu
C.
,
Jin
J.
,
Wang
J.
,
Zhang
F.
,
Tian
Y.
,
Zhang
F.
,
Zhou
Y.
&
Han
Q.
2022
Metal–organic frameworks (MOFs) for the efficient removal of contaminants from water: Underlying mechanisms, recent advances, challenges, and future prospects
.
Coord. Chem. Rev.
468
,
214595
.
https://doi.org/10.1016/j.ccr.2022.214595
.
This is an Open Access article distributed under the terms of the Creative Commons Attribution Licence (CC BY 4.0), which permits copying, adaptation and redistribution, provided the original work is properly cited (http://creativecommons.org/licenses/by/4.0/).