Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

Two–three parameters isotherm modeling, kinetics with statistical validity, desorption and thermodynamic studies of adsorption of Cu(II) ions onto zerovalent iron nanoparticles

Abstract

Adsorption of problematic copper ions as one of the endocrine disruptive substances from aqueous solution onto nanoscale zerovalent iron (nZVI) was studied. The high pore size 186.9268 Å, pore diameter 240.753 Å, and BET surface area 20.8643 m2 g−1 and pH(pzc) enlisted nZVI as an efficient nano-adsorbent for treatment of heavy metals from synthetic wastewater. SEM and EDX revealed the morphology and elemental distribution before and after adsorption. 98.31% removal efficiency was achieved at optimum adsorption operational parameters. Of all the thirteen isotherm models, equilibrium data were well fitted to Langmuir. Kinetics and mechanism data across the concentrations from 10 to 200 mg L−1 were analyzed by ten models. PSO best described kinetics data as confirmed by various statistical error validity models. The intraparticle diffusion model described that the intraparticle diffusion was not the only rate-limiting step. The adsorption mechanism was diffusion governed established by Bangham and Boyd models. Feasible, spontaneous, endothermic, and degree of randomness were reveal by the thermodynamic studies. Better desorption index and efficiency were obtained using HCl suggesting multiple mechanism processes. The performance of ZVI suggested it has a great potential for effective removal of endocrine disruptive cationic contaminant from wastewater.

Introduction

Endocrine disruptive compounds (EDC) have been among emerging contaminants whose adverse effects in the environments have received the attention of various researchers. Heavy metal ions have been identified to have hazardous effects on the endocrine system. Copper, as heavy metal ions, has been listed among endocrine disruptive compounds. Some of the endocrine disruptive effects of copper ions are: Increased levels of plasma cortisol associated with protein synthesis, cell proliferation, and apoptose in gill cells. It can also lead to an increase in Catecholamines which promotes metabolic and haematopoietic responses1,2. Several anthropogenic activities and natural phenomena release copper ions into the environment3. The largest threat to human lives and aquatic organisms arises from the soluble form of copper4. These soluble copper ions get released into the environment via different agricultural applications. Research has shown that high uptakes of copper may cause liver and kidney damage and even death5,6,7. Research on the utilization of nanoparticles is on the increase due to their special characteristics. Nanotechnology is a trending research area for the science and technology of functional structures at the molecular scale. This covers current research work in chemical, physical, biological, medical, material sciences, and engineering. Nano-materials have been reported to be applicable in environmental remediation, catalysis, development of optical devices, and medicine8. Nanoparticles are the new trend of effective adsorbents used in the decontamination of water and immobilization of heavy metal ions from their solutions9. A study conducted by the U.S Environmental Protection Agency showed that zerovalent iron nanoparticle (nZVI) is environmentally benign and effective in soil and water remediation8,10,11,12. nZVI has found relevance in the adsorption of problematic toxicants such as EDC heavy metal ions. Of all various conventional approaches13 described for heavy metal ions removal, adsorption via adsorption is much more favored since it is low cost, efficient, available, and easy to operation14,15,16.

The most vital quantity for comprehending the adsorption process is gotten from adsorption isotherm models. The parameters of isotherm modeling are essential factors relevant to the design of an effluent treatment reactor. More so, extensive isotherm models were investigated to predict and compare adsorption performances. Most common isotherm models are Langmuir and Freundlich, Temkin and Dubinin–Kaganer–Raduskevich (DKR). Other isotherm models used in this study are Halsey, Jovanovic, Elovich, Jossen, Flory–Huggins, Kiselev, Harkins–Jura, Fowler–Guggenheim, and Redlich–Peterson. In most adsorption studies carried out using nZVI, no detailed investigation has been reported for mathematical isotherm parameters that could be utilized for treatment reactor design. This has not been given a priority hence a research gap of global interest has been created. The energy of the adsorption process for uptake of endocrine disruptive copper ions via thermodynamic studies was examined to determine the feasibility, spontaneity, energy content, and degree of disorderliness of the process. More so, the mechanism vis-à-vis desorption studies were investigated using three desorbing agents. The reality of the adsorption process was assessed by post-adsorption characterization using Fourier Transform Infrared Spectroscopy (FTIR), Scanning Electron Microscopy (SEM), and Energy Dispersive X-ray (EDX).

Materials and methods

All through this work, analytical grade reagents were used without further purification. Double-Distilled-Deionized water, Copper sulphate (CuSO4·5H2O, Breckland Scientific Batch No. 6688), Isopropyl alcohol (BDH, Min. Assay 99%, Prd No. 29694 6). Other chemicals purchased from Sigma Aldrich, USA are Sodium borohydride (NaBH4), Iron (III) chloride (FeCl3.6H2O), Hydrochloric acid (HCl), Sodium hydroxide (NaOH), Sodium nitrate (NaNO3).

Synthesis of zerovalent iron nanoparticles (nZVI)

The synthesis of nZVI for the removal of endocrine disruptive heavy metal ions was undertaken by following the procedure described in our prior studies6,17. Under an anaerobic environment, a resulting black coloration of core–shell zerovalent iron nanoparticles (nZVI) was obtained from the reaction between 0.023 M solution of FeCl3·6H2O and 0.125 M solution of NaBH4 in ratio 1:5. Detailed synthetic procedure is presented in the supplementary document associated with this study. The synthesis equation is depicted in Eq. (1):

$$4{\text{Fe}}^{3 + } + 3{\text{BH}}_{4}^{ - } + 9{\text{H}}_{2} {\text{O}} \to 4{\text{Fe}}^{0} \downarrow + 3{\text{H}}_{2} {\text{BO}}_{3}^{ - } + 12{\text{H}}^{ + } + 6{\text{H}}_{2} \uparrow$$
(1)

Surface charge (pHpzc), BET surface area, surface morphology and elemental distribution

Following the salt addition and pH variation method, the point of zero charge was determined as presented in the supplementary materials18. Surface area by BET, pore width, and volume were determined using Micrometritics AutoChem II Chemisorption Analyzer. The surface morphological characterization and elemental analysis were carried out using a Scanning Electron Microscopy (SEM) integrated with Energy Dispersive X-ray (EDX) analyzer. SEM images and EDX spectra were obtained using a TESCAN Vega TS 5136LM typically at 20 kV at a working distance of 20 mm. Samples for SEM analysis were prepared by coating them in gold using a Balzers’ Spluttering device.

Effect of stirring speed, pH, and co-existing ions

In order to optimize the stirring speed, 160–240 rpm speed was studied at optimum conditions. Effect of pH was studied by regulating the solution to the desired pH value using 0.1 M NaOH and 0.1 M HNO3 solutions. Effect of Co-existing ions/Ionic strength varying the concentration of NaCl introduced into Cu2+ solution from 0.001 to 1.0 M.

Batch isotherm, kinetics, and thermodynamic studies

A typical batch adsorption study was carried out following procedure reported in our previous study19,20. 1000 ppm Cu2+ stock solution was prepared by dissolving 2.5 g of CuSO4.5H2O in 1000 mL of distilled-deionized water. Study on initial Cu2+ ion concentration was examined by adding 100 mg nZVI at different Cu (II) ions concentrations (10–200 ppm) and residual concentration determined by using AAS model AA320N. The quantity adsorbed and percentage removal efficiency were calculated utilizing Eqs. (2) and (3)21,22,23:

$$Q = \frac{{(c_{o} - c_{e} )V}}{m}$$
(2)
$$\% \;RE = \frac{{C_{i} - C_{e} }}{{C_{i} }} \times 100$$
(3)

The characteristics and mechanism of the adsorption process were investigated from the study of the Ce-dependent changes of Qe applied to thirteen isotherm models. Similarly, the batch adsorption kinetic experiments were conducted at optimum conditions for contact time ranging from 10 to 120 min. Adsorption capacities at contact time were obtained using Eq. (4)5,23:

$$Q_{t} = \frac{{(C_{o} - C_{e} )V}}{W}$$
(4)

Kinetic data were fitted to ten kinetics and mechanism models.

From the thermodynamics studies, the effect of temperature at optimum conditions was investigated at five different temperatures (298 K, 303 K, 318 K, 328 K, 333 K) for adsorption of endocrine disruptive Cu2+ onto nZVI following our previously reported procedure24,25. The study was carried out in a temperature-controlled water bath. Data obtained were fitted to Van’t Hoff equation depicted in Eqs. (5) and (6).

$$logK_{C} = \frac{\varDelta S^\circ }{{2.303R}} - \frac{\varDelta H^\circ }{{2.303RT}}$$
(5)
$$\varDelta G = - 2.303RT\ell ogKc$$
(6)

R is the gas constant (8.314 J mol−1 K)−1), T the absolute temperature (K), Kc (qe/Ce) an equilibrium constant at various temperature. Standard enthalpy change ∆H° (kJ mol−1) and standard entropy change ∆S° (J mol−1 K1) were determined from the slope and intercept of the Van’t Hoff plot of log Kc versus 1/T26,27.

Desorption studies

The desorption is a means of regenerating the adsorbent capacity for reusability and cost effectiveness determination. Desorption studies were investigated using the following eluents: deionized (DI) water, 0.2 M HCl and 0.2 M CH3COOH, at pre-determined optimum conditions. The desorption capacity, percentage desorbed, desorption efficiency and desorption index were determined using Eqs. (7)–(10)28:

$$q_{des} = C_{des} \frac{V}{W}$$
(7)
$$\% \;Desorption\; = \;\frac{{C_{des} }}{{C_{i} }} \times 100$$
(8)
$$\% \;Desorption\;Efficiency\; = \;\frac{{q_{des} }}{{q_{e} }} \times 100$$
(9)
$$Desorption\;index\; = \;\frac{\% \;total\;metal\;removed\;after\;adsorption}{{\% \;total\;metal\;remained\;on\;the\;adsorbent\;after\;desorption}}$$
(10)

where qdes is the quantity of metal ion desorbed (mg g−1), qe is the quantity of metal adsorbed after sorption (mg g−1), Cdes is the concentration of metal ion left after desorption (mg L−1), V is the volume of the metal ion solution (mL) while W is the weight of adsorbents (mg).

Results and discussion

nZVI physicochemical characterization: pH(pzc), BET surface area, pore volume and pore size

Summarized in Table 1 are pH, point of zero charge (PZC), BET surface area, and other physicochemical parameters describing the core–shell. The point of zero charge finds relevance in surface and nanoscience.

Table 1 Physicochemical parameters of nZVI.

Figure S1 (from the supplementary document associated with this study) shows the pH(pzc) of nZVI. It revealed that adsorption of Cu2+ would take place at a pH > pHpzc as a result of more active binding being available due to deprotonation and low electrostatic repulsion. This finding shows that nZVI was positive at pH < pHpzc and negative at pH > pHpzc. Thus this signposts the suitability of nZVI for effective adsorption. The BET surface area 20.86 m2 g−1 and the external surface area 16.4503 m2 g−1 being greater than their corresponding micropores further support the suitability of nZVI for adsorption. Higher surface area enhances the adsorption process as supported. Therefore, it can be deduced that nZVI nano-adsorbent would utilize its external surfaces for heavy metal uptake than its micropore areas29.

SEM/EDX characterization

Percolation of EDC-Cu2+ into the pores ad matrix of nZVI was proved by the SEM/EDX depicted in Fig. 1A–D. Figure 1A revealed the SEM image before adsorption while Fig. 1B depicted the EDX with an intense peak of a zerovalent iron nanoparticle. Before adsorption spherical, chain-like aggregated morphology of nZVI was revealed SEM. The core–shell nature of zerovalent iron with intense peaks between 0.6–6.4 and 7.0 keV was revealed from the EDX result in Fig. 1B. Presented in Fig. 1C is the SEM micrograph showing swollen and robust nature of the surface of the nZVI nano-adsorbent after adsorption suggesting that the nZVI surface had been Cu-loaded up. More so, corroborating the result from SEM, Fig. 1D revealed the EDX spectrum showing the presence of Cu(II) as evidence of Cu adsorption onto core–shell nZVI. This is supported by finding in the literature30.

Figure 1
figure1

(A) SEM micrograph of nZVI before adsorption. (B) EDX spectrum of nZVI before adsorption. (C) SEM micrograph of Cu-loaded-nZVI after adsorption. (D) EDX of Cu-loaded-nZVI after adsorption.

Effect of operational parameters

Effect of initial Cu2+ concentration

A major role in the adsorption study is played by the factor of initial concentration at optimum conditions presented in supplementary document as seen in Figure S2 This showed that at optimum conditions, 85.04% RE and 81.04 mgg−1 quantity of Cu2+ was adsorbed. The extent of removal of Cu(II) ion cation increased based on the availability of more active sites at lower concentrations until the pore sizes were saturated at an advanced concentration (150–200 mg L−1). Concentration gradient was built up in the Cu-nZVI system due to intensification of drive force as concentration increased from 10 to 200 mg L−1. This is supported by the findings in the literature31.

Effect of contact time

The Build-up of Cu(II) ions at the solid–liquid interfaces are controlled by the contact time. From this study, optimization of the contact time was investigated from 10 to 120 min. Fast kinetics from the bulk to the outer and inner surface of the nano-material (nZVI) identified by a short contact time to reach equilibrium was observed in the supplementary document as seen in Figure S3 Quantity of Cu(II) adsorbed increase from 4.96 to 82.82 mg g−1 as the initial Cu2+ concentration increased from 10 to 200 ppm. A similar trend was observed by Baby et al.32 on the adsorption of heavy metals.

Effect of initial solution pH

The key to the adsorption of heavy metal ions is the solution pH because it affects the surface chemistry of the system. A plot of the effect of initial concentration is presented in the supplementary document as seen in Figure S4 portrayed the effect of pH at optimum conditions. Coined from the understanding of the isoelectric point of the pH(pzc), nZVI is suitable for the uptake of cationic pollutants such as Cu2+ since the pH > pH(pzc). At low solution pH, the system is protonated leading to electrostatic competition among Cu2+ and other cationic species such as H+, Cu(OH)+, Cu(OH)2. However, at solution pH > pH(pzc), the system is negative, deprotonation occurs, there is less competition between Cu2+ and other anionic species (Cu(OH)3 and Cu(OH)4 2−). Effective adsorption occurs at pH > pH(pzc). Optimum adsorption was achieved at pH 6 with 98.31% removal efficiency and quantity adsorbed 73.73 mg g−1 indicating effective binding of Cu2+ onto nZVI surface. This is corroborated by the findings of other researchers33.

Effect of ionic strength

Analysis of Figure S5 (presented in the supplementary document associated with this study) showed the effect of ionic strength on Cu2+ adsorption. Pollution of the water system is not limited to heavy metal ions only, some co-existing ions increase the salinity and ionic strength of the water body as investigated in this study. Co-existing ions polluted waste system increases the salinity and background electrolyte of the water body. A decrease in the percentage of Cu2+ removed from 81.99 to 79.73% with a reduction in quantity adsorbed from 61.49 to 59.79 mg g−1 was observed in Figure S5. The decrease in Cu(II) ions uptake may also be due to a decrease in the electrostatic attraction arising from compressed electrical diffuse double layer supporting the findings of Advantageously, the removal efficiency of 81.99% shows that nZVI is an effective nano-sorbent in treatment industrial discharge containing co-existing ions. This is supported by findings in the literature34.

Effect of stirring speed

This study demonstrated as shown in Figure S6 (supplementary document) that at 200 rpm maximum adsorption of Cu2+ onto nZVI was attained. Stirring speed is also one of the important parameters in adsorption studies because it promotes turbulence, frequency of collision and improves mass transfer in the medium between the two phases. At 200 rpm, the percentage Cu2+ removal efficiency and quantities adsorbed are 96.98% and 72.73 mg g−1 for nZVI. Stirring speed increases the retention of Cu2+ and it encourages a better transfer of Cu2+ between solid–liquid interfaces (Cu2+-nZVI system)18. No appreciable percentage removal efficiency was observed after 200 rpm and all other study was carried out at this stirring speed.

Two–three parameters adsorption isotherm modelings

One of the important and significant aspects of adsorption studies is mathematical isotherm modeling. Isotherm modeling is important in order to observe the relationship between nZVI and Endocrine disruptive Cu(II) ions at equilibrium conditions. A good understanding of this would significantly enhance the design of the adsorption system, effluent treatment reactor, and the pattern describing adsorbate-adsorbent interaction. Equilibrium data obtained from initial concentration were analyzed using thirteen mathematical isotherm models. All mathematical isotherm models used in this study were presented in Table 2 together with their non-linear, linear equations and parameters' description. The estimated parameters are portrayed in Table 3. The plots in the isotherm studies are presented in the supplementary document associated with this article from Figure S7A–S7M.

Table 2 Adsorption isotherm and kinetics models35,43,47.
Table 3 Isotherm models and their various evaluated parameters for the adsorption of Cu2+ onto nZVI.

Langmuir isotherm model (Figure S7A) assumes no interaction of the neighboring sites, monolayer surface, identical active sites, uniformity in adsorption energy14. The non-linear and linear Langmuir equations are presented in Eq. (11). In this study, the Langmuir isotherm model has the highest correlation coefficient (R2 > 0.97) indicating the appropriateness and best fitting of equilibrium data to the Langmuir model. The Langmuir essential feature, as well as the separation factor or dimensionless constant (RL), was calculated using Eq. (12)20. Values of calculated characteristics parameters are presented in Table 3. The values of RL (1 > RL > 0) portrayed in Table 3 supported favorable adsorption process35.

Freundlich isotherm model is presented in Eq. (13) (Table 2) and the plot is as depicted in Figure S7B. The characteristic parameters are represented in Table 3. The values 12.54 and 1.83 indicated Freundlich capacity (KF) and intensity (nF) of adsorption respectively. The values of nF also measure whether the adsorption is favorable or not. The value of 1/nF (0.5457) less than unity and nF greater than unity and less than 10 indicated a normal and favorable adsorption36,37. Temkin model (Eq. 14) fits the experimental data (R2 = 0.95) as depicted in Figure S7C. The positive value of B (14.678) and high bT (168.794 J mol−1) revealed the binding of Cu2+ onto nZVI as well as the endothermic nature of the system. A report from other researchers corroborated this36.

Equations (15)–(17) defined the Dubinin–Kaganer–Raduskevich (DKR) model, Polanyi potential, and adsorption free energy of DRK. DRK plot is presented in Figure S7D and Table 3 shows evaluated parameters. The DKR free energy (E = 1581.14 J mol−1) lower than 8 kJ mol−1 supported that electrostatic interaction between Cu2+-nZVI system is Physisorption mechanism38. Halsey isotherm model (Eq. 18 and plot in Figure S7E) with parameters of KH and nH (0.0097 and -1.8325) further supported normal and favorable adsorption indicated by the Freundlich isotherm model. However, the negative value of nH couple with the low R2 value of Harkin-Jura (Eq. 19, Figure S7F) showed that the adsorption nature of the nZVI surface is not multilayer and heterogeneous39.

Combination of both the Langmuir and Freundlich isotherm attribute could be assessed in Redlich–Peterson Isotherm model (Eq. 20, Figure S7G). Redlich–Peterson correlation value (R2 = 0.9475) shows its versatility and fitting to equilibrium data40,41. Figure S7H depicts the Jovanovic isotherm (Eq. 21) model plot. Elovich isotherm model (Eq. 22, Figure S7I) has a foundation on the kinetic principle with the assumption of an increase in the adsorption sites exponentially42. It also takes into consideration the maximum monolayer capacity (Qmax). Based on R2 value, the equilibrium data were fitted to the Elovich model but it was not as better described as compared to the Langmuir model. Also, its estimated Qmax = 32.573 mg g−1 (Table 3) being less than that of Langmuir signposted Langmuir as a better model. Jossen’s isotherm model (Eq. 23, Figure S7J) is based on a distribution of the energy of interactions between the system solid–liquid system43. Jossen’s fit equilibrium data with R2 > 0.97 (Table 3). Both Kiselev Isotherm Model (Eq. 24, Figure S7K) and the Flory–Huggins isotherm model (Eq. 25, Figure S7L) take into consideration the surface coverage (θ) of the Cu2+ adsorbate on the nZVI. Jovanovic isotherm model corresponds to another approximation for monolayer localized adsorption without lateral interactions which ought to be similar to the Langmuir isotherm model. However, a lower correlation coefficient (R2 = 0.6105) obtained in this study indicated that there is a lateral interaction and thus this model lower approach towards saturated compared to Langmuir adsorption isotherm as reported by Al-Ghouti et al.44. This is supported by the fit and parameters obtained from Fowler–Guggenheim (F–G) isotherm model (Eq. 27, Figure S7M). Taken into consideration is the lateral interaction of adsorption of EDC-Cu2+ onto nZVI by the FG- isotherm model45. As reported by the literature, the interaction between the adsorbed molecule is attractive, if W is positive; repulsive interaction if W is negative and no interaction between the adsorbed molecules will be observed if W = 042,46. In this study, the fit of the Fowler–Guggenheim isotherm model (R2 = 0.9487) and positive value of W (W = 881.02 J mol−1) indicated that there is positive contact in the Cu2+-nZVI system, hence the adsorption heat increased with loading confirming endothermic adsorption process as observed in the thermodynamics studies.

The screening and arrangement are based on the understanding of the important parameters (Qmax and R2). With regards to Qmax (in descending order) Langmuir > DKR > Elovich > Jovanovic. With respect to R2 (in descending order): Langmuir > Jossen = Elovich > Freundlich = Halsey > Temkin > Fowler–Guggenheim > DKR > Flory–Huggins Kiselev > Harkin–Jura > Jovanovic. Presented in Table 4 is the comparison of maximum monolayer adsorption capacities of adsorption of Cu2+ onto various nano-adsorbents and nZVI used in this study. It is obvious that nZVI exceedingly surpassed other existing adsorbents reported. This indicated that nZVI is an excellent potential nano-adsorbent for effective removal of endocrine disruptive heavy metal ions.

Table 4 Comparison of adsorption capacity nZVIf other nano-adsorbents used for Cu2+ removal.

Adsorption kinetic with statistical error validity modeling

A kinetic study was undertaken to understand the controlling pathway, the rate of surface adsorption of the contaminant to the adsorbent, and the quantity of the adsorption capacity. The kinetics equations vis-à-vis pseudo-first-order (PFO), pseudo-second-order (PSO), Elovich, Avrami, and Power Function (Fractional power) are represented on Eqs. (28)–(34)47.

$${\text{Pseudo}}\;{\text{first-order}}\left( {{\text{Lagergren's}}\;{\text{rate}}\;{\text{equation}}} \right)\quad \ell og(q_{e} - q_{t} ) = \ell og\;q_{e} - \frac{{k_{1} t}}{2.303}$$
(28)
$${\text{h}}_{{1}} \;{\text{initial}}\;{\text{pseudo}}\;{\text{first-order}}\;{\text{adsorption}}\;{\text{rate}} \left( {{\text{mg}}\;{\text{g}}^{{ - 1}} \;{\text{min}}^{{ - 1}} } \right)\quad h_{1} = k_{1} q_{e}$$
(29)
$${\text{Pseudo}}\;{\text{second-order}}\;{\text{rate}}\;{\text{equation:}}\quad \frac{t}{{q_{t} }} = \frac{1}{{k_{2} q_{e}^{2} }} + \frac{1}{{q_{e} }}t$$
(30)
$${\text{h}}_{{2}} \;{\text{is}}\;{\text{the}}\;{\text{initial}}\;{\text{pseudo}}\;{\text{second-order}}\;{\text{adsorption}}\;{\text{rate:}}\quad h_{2} = k_{2} q_{e}^{2}$$
(31)
$${\text{Elovich}}\;{\text{model:}}\quad q_{t} = \frac{1}{\beta }\ell n\left( {\alpha \beta } \right) + \frac{1}{\beta }\ell n\left( t \right)$$
(32)
$${\text{Avrami}}\;{\text{model:}}\quad \ln \left( {\ln \left( {\frac{{q_{e} }}{{q_{e} - q_{t} }}} \right)} \right) = n\ln k_{Av} + n\ln t$$
(33)
$${\text{Power}}\;{\text{Function:}}\quad \ell og\;(q_{t} ) = \log (k) + v\log (t)$$
(34)

The kinetic plots are presented in Fig. 2A–E with error bars indicating the application of error models and the evaluated parameters are presented in Table 5. The kinetic constant k1of Pseudo first order (PFO), its adsorption rate constant h1, disagreement between qe, exp and qe, cal and low correlation coefficient, R2 < 0.90, demonstrated that PFO is not applicable in this study. A similar low trend in the R2 value was observed in the Avrami model demonstrating that it is not applicable in this study. A good agreement between the experimental quantity adsorbed (qe, exp) and the calculated quantity adsorbed (qe, cal) was observed in PSO, Elovich, and Power function. From the Elovich model, the values of  (adsorption rate) increased with an increase in concentration as a result of an increase in the number of sites. The values of 1/β at 10 ppm, 50 ppm, 100 ppm, and 150 ppm are 5.882, 11.764, and 17.123 respectively. These values reflect the number of sites available for adsorption30. Kinetic parameters from Power Function in Table 5 indicated time-dependent of Cu(II) onto nZVI with the value of constant v less than 1 across all the concentrations. Of all these kinetic models, PSO best described the Cu(II) adsorption process and this was supported by the statistical error validity model presented in Table 5. The PSO initial adsorption rate (h2) increases with increase in concentration from 33.67 to 238.095 mg g−1 min−1. R2 values range from 0.99 to unity demonstrating the best fitting by PSO suggesting chemisorption mechanism.

Figure 2
figure2

(AE): Linearized plots of (A) pseudo-first-order rate equation, (B) pseudo-second-order rate equation, (C) Elovich rate equation, (D) Avrami kinetics models, (E) fractional power rate equation models for adsorption of Cu2+ onto nZVI at 10 ppm, 50 ppm, 100 ppm, 150 ppm and 200 ppm.

Table 5 Kinetic data for adsorption of Cu2+ onto nZVI at different initial concentrations.

Statistical validity of the kinetic models

Assessment on the best kinetic fitting model that is always based on linear regression coefficient could be biased inherent, hence the need for statistical validity model. The suitability, agreement, and best fit among the kinetic models are judged not only by regression coefficient (R2) but also with the use of statistical error validity models. Validity of kinetic data was fitted to statistical error models namely; Average relative error (ARE), Normalized Standard Deviation Δqt (%), Hybrid fractional error function (HYBRD), Derivative of Marquardt's percent standard deviation (MPSD), Standard deviation of relative Error (SRE). The various statistical functions are presented in Table 6. Presented in Table 7 are the statistical error validity data of the kinetic models. Five statistical tools were used for the validity of these kinetic models. It is observed that the closer the agreement between the experimental quantity adsorbed (qe, exp) and calculated quantity adsorbed (qe, cal), the lower the values of these statistical tools, the better the model. In order to justify and juxtapose the best model, a reference was made to the coefficient of regression (R2). The higher the R2 values, the closer the values of qe, exp, and qe, cal, the lower the values of ∆q, HYBRID, MPSD, ARE, and SRE, the better the kinetic models in describing the sorption process48,49,50. The values in Table 4 vividly show that pseudo-second-order at various initial Cu2+ concentrations (10 ppm, 50 ppm, 100 ppm, 150 ppm, and 200 ppm) best describe the sorption process. the model can be arranged in descending order with respect to R2: pseudo-second-order > Elovich > fractional power > Avrami > pseudo-first-order.

Table 6 Adsorption statistical error validity models (ASEVM)24,43,44,46.
Table 7 Statistical Error validity data on kinetics models of adsorption of Cu(II) onto nZVI.

Adsorption mechanisms for sorption of Cu2+ onto nanoscaled zerovalent iron (nZVI)

Figure 3A–E show the linear plots of intraparticle diffusion, liquid diffusion, external diffusion, Bangham and Boyd models. Adequate understanding of the adsorption mechanism is enhanced by the determination of the rate-controlling/determining step. The three definite steps that could be used to describe the adsorption rate are51: (1) Intraparticle or pore diffusion, where adsorbate molecules percolate into the interior of adsorbent particles, (2) Liquid film or surface diffusion where the adsorbate is transported from the bulk solution to the external surface of the adsorbent, and (3) adsorption on the interior sites of the sorbent. Since the plot of Intraparticle diffusion (Fig. 3A) did not pass through the origin, it is demonstrated that it is not the only rate-determining step52. Other mechanisms such as surface diffusion and external diffusion also participated in the mechanism of Cu(II) removal. However, the higher R2 values of intraparticle diffusion from the evaluated parameters presented in Table 8 demonstrated that the mechanism is pore diffusion dependent which was confirmed by Bangham and scattered plot of Boyd models53,54. The intercept of intraparticle diffusion which is the thickness of the surface gives information about the contribution of the surface sorption in the rate-determining step. The larger the intercept, the greater the contribution of nZVI in adsorption of Cu2+ as observed from the trend across the concentrations investigated.

Figure 3
figure3

Linearized plots of (A) intraparticle diffusion, (B) liquid film diffusivity, (C) external diffusion, (D) Bangham and (E) Boyd models for adsorption of Cu2+ onto nZVI at 10 ppm, 50 ppm, 100 ppm, 150 ppm and 200 ppm.

Table 8 Adsorption Mechanism models for Sorption of Cu2+ onto nZVI.

Thermodynamics analysis

Thermodynamics analysis is imperative to determine the (enthalpy change) heat content (ΔH); entropy change (degree of randomness, ΔS), possibility, and spontaneity (Gibbs free energy change, ΔG) in every adsorption process. The plots in thermodynamics studies are presented in the supplementary document associated with this article. As observed in Figure S8, intensification in the percentage removal efficiency was attained with an increase in temperature of the system supporting the endothermic process. This is due to a decrease in the mass transfer resistance and boundary layer thickness of nZVI55. Van’t Hoff’s linear plot of log Kc against 1/T was portrayed in Figure S9 and the result obtained was presented in Table 9. The positive value of ΔH (+ 50.6059 kJ mol−1) confirmed that the adsorption process is endothermic in nature56. The positive value of ΔS (+ 174.679 J mol−1 K−1) shows an increase in the degree of randomness of the lateral interaction during the adsorption of Cu2+ at the solid/liquid interface. This could be enhanced by the appropriate stirring speed. The feasibility and spontaneity of the adsorption process are confirmed by the negative values of ΔG (− 1.6765 to 7.9602 kJ mol−1).

Table 9 Thermodynamic parameters for adsorption of Cu2+ onto nZVI.

Desorption mechanism

Figure 4 shows the comparative effect of different eluents in the desorption of Cu2+ from Cu2+-loaded nZVI. The opportunity to investigate regeneration and reusability of loaded adsorbent is enhanced by desorption studies. The effectiveness of three different eluents and desorbing agents (HCl, CH3COOH, and H2O) was investigated. The basic desorption mechanisms are ion exchange, complexation, and precipitation depending on the most effective desorbing agent57. The exact mechanism involved in the adsorption process is revealed by the performance of the most effective desorbing agent. The maximum percentage of Cu(II) desorbed from Cu(II) loaded-nZVI using HCl was 79.89% showing the best desorption index of 3.39. The effectiveness of HCl as the best desorbing agent among the three eluents used is supported by the findings of Reddiar et al. (2019)50 Acetic acid also performed averagely while distilled-deionized water was a poor desorbing agent in the desorption of Cu(II) from Cu(II)-loaded-nZVI. Thus, the adsorption of Cu(II) onto nZVI is routed by ion exchange. Ion-exchange, electrostatic and physiochemical mechanistic nature of the adsorption supported by the previous studies50,58,59.

Figure 4
figure4

Comparative effect of different eluents in the desorption of Cu2+ from Cu2+-loaded nZVI.

Conclusion

This study revealed the effectiveness of nZVI as an auspicious nano sorbent for the efficient elimination of endocrine disruptive heavy metal ions. The quality physicochemical properties of nZVI gave it an edge among the list of other nano-adsorbents compared. Evidence of the adsorption of Cu2+ onto nZVI was revealed by a change in morphology and elemental distribution by SEM and EDX respectively from post adsorption characterization. The adsorption of Cu2+ onto nZVI was well influenced by operational parameters. Optimum adsorption was achieved at pH 6 with 98.31% removal efficiency, 73.73 mg g−1 quantity adsorbed and 200 rpm stirring speed. Thermodynamics parameters ΔH° (+ 50.6059 kJ mol−1), ΔS° (174.6790 J mol−1 K−1), ΔG° (− 1.6765 kJ mol−1 to − 7.9602 kJ mol−1). Indicated random, feasible, spontaneous, and endothermic nature of the adsorption process. The adsorption behavior was well explained by the Langmuir isotherm model and it followed the following order: Langmuir > Jossen/Elovich > Freundlich/Halsey > Temkin > Fowler–Guggenheim > Redlich–Peterson > DKR > Flory–Huggins > Kiselev > Harkin–Jura > Jovanovic. Langmuir best described equilibrium data. The Langmuir monolayer adsorption capacity (90.09 mg g−1) surpassed other nano-adsorbents utilized for the adsorption of Cu(II) ion. The Pseudo-second-order (PSO) best described the kinetics model based on R2 values greater than 0.99, close agreement between qe, exp and qe, cal and lower values of the five rigorous statistical validity models (Δqt, ARE, HYBRD, MPSD, and SRE). The mechanism model was pore diffusion dependent. Best desorption capacity and the index was portrayed by HCl indicating that ion-exchange, electrostatic, and physisorption mechanism. Based on the capacity displayed by nZVI in adsorption of EDC Cu2+, it could be recommended for effective industrial treatment of heavy metal ions.

Data availability

Available data are presented in the study and no other data were used to support the study.

References

  1. 1.

    Tapia-Orozco, N. et al. Removal strategies for endocrine-disrupting chemicals using cellulose-based materials as adsorbents: A review. J. Environ. Chem. Eng. 4, 3122–3142 (2016).

    CAS  Article  Google Scholar 

  2. 2.

    Handy, R. D. Chronic effects of copper exposure versus endocrine toxicity: Two sides of the same toxicological process?. Comp. Biochem. Physiol. A Mol. Integr. Physiol. 135, 25–38 (2003).

    PubMed  Article  CAS  Google Scholar 

  3. 3.

    Saravanan, A. et al. Simultaneous removal of Cu(II) and reactive green 6 dye from wastewater using immobilized mixed fungal biomass and its recovery. Chemosphere 271, 129519 (2021).

    ADS  CAS  PubMed  Article  Google Scholar 

  4. 4.

    Saravanan, A., Senthil Kumar, P. & Mugilan, R. Ultrasonic-assisted activated biomass (fishtail palm Caryota urens seeds) for the sequestration of copper ions from wastewater. Res. Chem. Intermed. 42, 3117–3146 (2016).

    CAS  Article  Google Scholar 

  5. 5.

    Dada, A. O., Latona, D. F., Ojediran, O. J. & Nath, O. O. Adsorption of Cu (II) onto bamboo supported manganese (BS-Mn) nanocomposite: Effect of operational parameters, kinetic, isotherms, and thermodynamic studies. J. Appl. Sci. Environ. Manag. 20, 409–422 (2016).

    CAS  Google Scholar 

  6. 6.

    Dada, A. O., Adekola, F. A. & Odebunmi, E. O. Kinetics, mechanism, isotherm and thermodynamic studies of liquid-phase adsorption of Pb2+ onto wood activated carbon-supported zerovalent iron (WAC-ZVI) nanocomposite. Cogent Chem. 68, 1351653 (2017).

    Article  CAS  Google Scholar 

  7. 7.

    Saravanan, A. et al. Effective adsorption of Cu(II) ions on sustainable adsorbent derived from mixed biomass (Aspergillus campestris and agro-waste): Optimization, isotherm and kinetics study. Groundw. Sustain. Dev. 11, 100460 (2020).

    Article  Google Scholar 

  8. 8.

    Das, S., Sen, B. & Debnath, N. Recent trends in nanomaterials applications in environmental monitoring and remediation. Environ. Sci. Pollut. Res. 22, 18333–18344 (2015).

    Article  Google Scholar 

  9. 9.

    Sivaranjanee, R. & Kumar, P. S. A review on cleaner approach for effective separation of toxic pollutants from wastewater using carbon Sphere’s as adsorbent: Preparation, activation, and applications. J. Clean. Prod. 291, 125911 (2021).

    CAS  Article  Google Scholar 

  10. 10.

    Keane, E. & Di Giulio, D. R. Fate, transport, and toxicity of nanoscale zero-valent iron (nZVI) used during superfund remediation. Nicholas Sch. Environ. 52 (2009). https://dukespace.lib.duke.edu/dspace/bitstream/handle/10161/2172/Keane_MP_042610.pdf?sequence=1.

  11. 11.

    Mukherjee, R., Kumar, R., Sinha, A., Lama, Y. & Saha, A. K. A review on synthesis, characterization, and applications of nano zero valent iron (nZVI) for environmental remediation. Crit. Rev. Environ. Sci. Technol. 46, 443–466 (2016).

    CAS  Article  Google Scholar 

  12. 12.

    Singh, R., Misra, V. & Singh, R. P. Removal of hexavalent chromium from contaminated groundwater using zero-valent iron nanoparticles. Environ. Monit. Assess 184, 3643–3651 (2012).

    CAS  PubMed  Article  Google Scholar 

  13. 13.

    Saravanan, A. et al. Phytoremediation of Cr(VI) ion contaminated soil using Black gram (Vigna mungo): Assessment of removal capacity. J. Environ. Chem. Eng. 7, 103052 (2019).

    CAS  Article  Google Scholar 

  14. 14.

    Ahmad, M. A., Ahmad Puad, N. A. & Bello, O. S. Kinetic, equilibrium and thermodynamic studies of synthetic dye removal using pomegranate peel activated carbon prepared by microwave-induced KOH activation. Water Resour. Ind. 6, 18–35 (2014).

    Article  Google Scholar 

  15. 15.

    Gong, J. et al. Shellac-coated iron oxide nanoparticles for removal of cadmium(II) ions from aqueous solution. J. Environ. Sci. (China) 24, 1165–1173 (2012).

    CAS  Article  Google Scholar 

  16. 16.

    Prabu, D., Parthiban, R., Senthil Kumar, P., Kumari, N. & Saikia, P. Adsorption of copper ions onto nano-scale zero-valent iron impregnated cashew nutshell. Desalin. Water Treat. 57, 6487–6502 (2016).

    CAS  Article  Google Scholar 

  17. 17.

    Dada, O. A., Adekola, F. A. & Odebunmi, E. O. Kinetics and equilibrium models for sorption of Cu(II) onto a novel manganese nano-adsorbent. J. Dispers. Sci. Technol. 37, 119–133 (2016).

    CAS  Article  Google Scholar 

  18. 18.

    Mtshatsheni, K. N. G., Ofomaja, A. E. & Naidoo, E. B. Synthesis and optimization of reaction variables in the preparation of pine-magnetite composite for removal of methylene blue dye. S. Afr. J. Chem. Eng. 29, 33–41 (2019).

    Google Scholar 

  19. 19.

    Dada, A. O. et al. Kinetics and thermodynamics of adsorption of rhodamine B onto bentonite supported nanoscale zerovalent iron nanocomposite. J. Phys. Conf. Ser. 1299, 012106 (2019).

    CAS  Article  Google Scholar 

  20. 20.

    Bello, O. S. et al. Rhodamine B dye sequestration using Gmelina aborea leaf powder. Heliyon 6, e02872 (2020).

    PubMed  PubMed Central  Article  Google Scholar 

  21. 21.

    Bhatti, H. N., Jabeen, A., Iqbal, M., Noreen, S. & Naseem, Z. Adsorptive behavior of rice bran-based composites for malachite green dye: Isotherm, kinetic and thermodynamic studies. J. Mol. Liq. 237, 322–333 (2017).

    CAS  Article  Google Scholar 

  22. 22.

    Olalekan, A. P., Dada, A. O. & Adesina, O. A. Review: Silica aerogel as a viable absorbent for oil spill remediation. J. Encapsul. Adsorpt. Sci. https://doi.org/10.4236/jeas.2014.44013 (2014).

    Article  Google Scholar 

  23. 23.

    Heidari-Chaleshtori, M. & Nezamzadeh-Ejhieh, A. Clinoptilolite nano-particles modified with aspartic acid for removal of Cu(II) from aqueous solutions: Isotherms and kinetic aspects. New J. Chem. 39, 9396–9406 (2015).

    CAS  Article  Google Scholar 

  24. 24.

    Dada, A. O., Adekola, F. A. & Odebunmi, E. O. Liquid phase scavenging of Cd (II) and Cu (II) ions onto novel nanoscale zerovalent manganese (nZVMn): Equilibrium, kinetic and thermodynamic studies. Environ. Nanotechnol. Monit. Manag. 8, 63–72 (2017).

    Google Scholar 

  25. 25.

    Dada, A. O., Adekola, F. A. & Odebunmi, E. O. A novel zerovalent manganese for removal of copper ions: Synthesis, characterization and adsorption studies. Appl. Water Sci. https://doi.org/10.1007/s13201-015-0360-5 (2015).

    Article  Google Scholar 

  26. 26.

    Ngueagni, P. T. et al. Adsorption of Cu(II) ions by modified horn core: Effect of temperature on adsorbent preparation and extended application in river water. J. Mol. Liq. 298, 112023 (2020).

    CAS  Article  Google Scholar 

  27. 27.

    Rajabi, M. et al. Adsorption of malachite green from aqueous solution by carboxylate group functionalized multi-walled carbon nanotubes: Determination of equilibrium and kinetics parameters. J. Ind. Eng. Chem. 34, 130–138 (2016).

    CAS  Article  Google Scholar 

  28. 28.

    Adekola, F. A., Hodonou, D. S. S. & Adegoke, H. I. Thermodynamic and kinetic studies of biosorption of iron and manganese from aqueous medium using rice husk ash. Appl. Water Sci. 6, 319–330 (2016).

    ADS  CAS  Article  Google Scholar 

  29. 29.

    Guan, H. et al. A facile method to synthesize magnetic nanoparticles chelated with copper(II) for selective adsorption of bovine hemoglobin. Korean J. Chem. Eng. 37, 1097–1106 (2020).

    CAS  Article  Google Scholar 

  30. 30.

    Şentürk, İ & Yıldız, M. R. Highly efficient removal from aqueous solution by adsorption of Maxilon Red GRL dye using activated pine sawdust. Korean J. Chem. Eng. 37, 985–999 (2020).

    Article  CAS  Google Scholar 

  31. 31.

    Liu, Z., Li, X., Zhan, P., Hu, F. & Ye, X. Removal of cadmium and copper from water by a magnetic adsorbent of PFM: Adsorption performance and micro-structural morphology. Sep. Purif. Technol. 206, 199–207 (2018).

    CAS  Article  Google Scholar 

  32. 32.

    Baby, R., Saifullah, B. & Hussein, M. Z. Palm Kernel Shell as an effective adsorbent for the treatment of heavy metal contaminated water. Sci. Rep. 9, 1–11 (2019).

    Article  CAS  Google Scholar 

  33. 33.

    Medhi, H. & Bhattacharyya, K. G. Kinetic and mechanistic studies on adsorption of Cu(II) in aqueous medium onto montmorillonite K10 and its modified derivative. New J. Chem. 41, 13533–13552 (2017).

    CAS  Article  Google Scholar 

  34. 34.

    Liu, G., Ma, J., Li, X. & Qin, Q. Adsorption of bisphenol A from aqueous solution onto activated carbons with different modification treatments. J. Hazard. Mater. 164, 1275–1280 (2009).

    CAS  PubMed  Article  Google Scholar 

  35. 35.

    Kampalanonwat, P. & Supaphol, P. The study of competitive adsorption of heavy metal ions from aqueous solution by aminated polyacrylonitrile nanofiber mats. Energy Proc. 56, 142–151 (2014).

    CAS  Article  Google Scholar 

  36. 36.

    Del Bubba, M. et al. Physicochemical properties and sorption capacities of sawdust-based biochars and commercial activated carbons towards ethoxylated alkylphenols and their phenolic metabolites in effluent wastewater from a textile district. Sci. Total Environ. 708, 135217 (2020).

    ADS  PubMed  Article  CAS  Google Scholar 

  37. 37.

    Dahri, M. K., Kooh, M. R. R. & Lim, L. B. L. Water remediation using low-cost adsorbent walnut shell for removal of malachite green: Equilibrium, kinetics, thermodynamic and regeneration studies. J. Environ. Chem. Eng. 2, 1434–1444 (2014).

    CAS  Article  Google Scholar 

  38. 38.

    Zhao, Y. et al. Adsorptive removal of endocrine-disrupting compounds and a pharmaceutical using activated charcoal from aqueous solution: Kinetics, equilibrium, and mechanism studies. Environ. Sci. Pollut. Res. 26, 33897–33905 (2019).

    CAS  Article  Google Scholar 

  39. 39.

    Song, C. et al. Adsorption studies of coconut shell carbons prepared by KOH activation for removal of lead(II) from aqueous solutions. Sustainability 6, 86–98. https://doi.org/10.3390/su6010086 (2014).

    CAS  Article  Google Scholar 

  40. 40.

    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. 2, 398–414 (2014).

    CAS  Article  Google Scholar 

  41. 41.

    Dada, A. O. et al. Article ID: IJCIET_09_13_159 Adsorption of rhodamine B dye onto chitosan supported zerovalent iron nanocomposite (C-nZVI). Int. J. Civ. Eng. Technol. 9, 1591–1605 (2018).

    Google Scholar 

  42. 42.

    Hamdaoui, O. & Naffrechoux, E. Modeling of adsorption isotherms of phenol and chlorophenols onto granular activated carbon. Part I. Two-parameter models and equations allowing determination of thermodynamic parameters. J. Hazard. Mater. 147, 381–394 (2007).

    CAS  PubMed  Article  Google Scholar 

  43. 43.

    Ayawei, N., Ebelegi, A. N. & Wankasi, D. Modelling and interpretation of adsorption isotherms. J. Chem. 2017, 1–11 (2017).

    Article  CAS  Google Scholar 

  44. 44.

    Al-Ghouti, M. A. & Da’ana, D. A. Guidelines for the use and interpretation of adsorption isotherm models: A review. J. Hazard. Mater. 393, 122383 (2020).

    CAS  PubMed  Article  Google Scholar 

  45. 45.

    Rangabhashiyam, S. & Balasubramanian, P. Performance of novel biosorbents prepared using native and NaOH treated Peltophorum pterocarpum fruit shells for the removal of malachite green. Bioresour. Technol. Rep. 3, 75–81 (2018).

    Article  Google Scholar 

  46. 46.

    Dada, A. O. Modeling of biosorption of Pb (II) and Zn (II) ions onto PAMRH: Langmuir, Freundlich, Temkin, Dubinin–Raduskevich, Jovanovic, Flory–Huggins, Fowler–Guggenheim and Kiselev comparative isotherm studies. Int. J. Mech. Eng. Technol. (IJMET) 10, 1048–1058 (2019).

    Google Scholar 

  47. 47.

    Dada, A. O. et al. For biosorption of indigo carmine dye: Kinetics, isotherm, and thermodynamic studies. Int. J. Phytoremediat. 1–14 (2020).

  48. 48.

    Sivarajasekar, N. & Baskar, R. Adsorption of basic red 9 onto activated carbon derived from immature cotton seeds: Isotherm studies and error analysis. Desalin. Water Treat. 52, 7743–7765 (2014).

    CAS  Article  Google Scholar 

  49. 49.

    Abdul, J. M., Vigneswaran, S., Shim, W. G. & Kandasamy, J. Removal of metsulfuron methyl by granular activated carbon adsorption. Desalin. Water Treat. 21, 247–254 (2010).

    CAS  Article  Google Scholar 

  50. 50.

    Hemavathy, R. R. V., Kumar, P. S., Suganya, S., Swetha, V. & Varjani, S. J. Modelling on the removal of toxic metal ions from aquatic system by different surface modified Cassia fistula seeds. Bioresour. Technol. 281, 1–9 (2019).

    CAS  PubMed  Article  Google Scholar 

  51. 51.

    Boparai, H. K., Joseph, M. & O’Carroll, D. M. Kinetics and thermodynamics of cadmium ion removal by adsorption onto nano zerovalent iron particles. J. Hazard. Mater. 186, 458–465 (2011).

    CAS  PubMed  Article  Google Scholar 

  52. 52.

    Suganya, S. An investigation of adsorption parameters on ZVI-AC nanocomposite in the displacement of Se(IV) ions through CCD analysis. J. Ind. Eng. Chem. 75, 211–223 (2019).

    Article  CAS  Google Scholar 

  53. 53.

    Azmier, M. & Norhidayah, A. Modified durian seed as adsorbent for the removal of methyl red dye from aqueous solutions. Appl. Water. Sci. https://doi.org/10.1007/s13201-014-0208-4 (2014).

    Article  Google Scholar 

  54. 54.

    Senthil Kumar, P., Senthamarai, C., Sai Deepthi, A. S. L. & Bharani, R. Adsorption isotherms, kinetics and mechanism of Pb(II) ions removal from aqueous solution using chemically modified agricultural waste. Can. J. Chem. Eng. 91, 1950–1956 (2013).

    CAS  Article  Google Scholar 

  55. 55.

    Hao, Y. M., Man, C. & Hu, Z. B. Effective removal of Cu (II) ions from aqueous solution by amino-functionalized magnetic nanoparticles. J. Hazard. Mater. 184, 392–399 (2010).

    ADS  CAS  PubMed  Article  Google Scholar 

  56. 56.

    Kaveeshwar, A. R. et al. Adsorption properties and mechanism of barium (II) and strontium (II) removal from fracking wastewater using pecan shell based activated carbon. J. Clean. Prod. 193, 1–13 (2018).

    CAS  Article  Google Scholar 

  57. 57.

    Lezcano, J. M. et al. Sorption and desorption of Cd, Cu and Pb using biomass from an eutrophized habitat in monometallic and bimetallic systems. J. Environ. Manag. 92, 2666–2674 (2011).

    CAS  Article  Google Scholar 

  58. 58.

    Petrović, M. et al. Mechanism of adsorption of Cu2+ and Zn2+ on the corn silk (Zea mays L.). Ecol. Eng. 99, 83–90 (2017).

    Article  Google Scholar 

  59. 59.

    Türkmen, D., Yilmaz, E., Öztürk, N., Akgöl, S. & Denizli, A. Poly(hydroxyethyl methacrylate) nanobeads containing imidazole groups for removal of Cu(II) ions. Mater. Sci. Eng. C 29, 2072–2078 (2009).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

The research enabling environment provided by the Management of Landmark University is appreciated. The assistance rendered by Dr. Ogunlaja Adeniyi in Rhodes University, South Africa for morphology characterization is thankfully acknowledged.

Author information

Affiliations

Authors

Contributions

A.O.D., FA.A., E.O.O., A.S.O. carried out the experiments, analyzed the results for the adsorption’ data. A.O.D., FA.A., E.O.O., O.S.B. fitted the thermodynamics and kinetics of adsorption curves. A.O.D., FA.A., and E.O.O., drafted the manuscript. All authors reviewed and approved the manuscript.

Corresponding author

Correspondence to Adewumi O. Dada.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Publisher's note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and Permissions

About this article

Verify currency and authenticity via CrossMark

Cite this article

Dada, A.O., Adekola, F.A., Odebunmi, E.O. et al. Two–three parameters isotherm modeling, kinetics with statistical validity, desorption and thermodynamic studies of adsorption of Cu(II) ions onto zerovalent iron nanoparticles. Sci Rep 11, 16454 (2021). https://doi.org/10.1038/s41598-021-95090-8

Download citation

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing