banner
Home / News / Design of silver-zinc-nickel spinel-ferrite mesoporous silica as a powerful and simply separable adsorbent for some textile dye removal | Scientific Reports
News

Design of silver-zinc-nickel spinel-ferrite mesoporous silica as a powerful and simply separable adsorbent for some textile dye removal | Scientific Reports

Oct 14, 2024Oct 14, 2024

Scientific Reports volume 14, Article number: 16481 (2024) Cite this article

741 Accesses

2 Altmetric

Metrics details

Silver-zinc-nickel spinel ferrite was prepared by the co-precipitation procedure with the precise composition Ag0.1Zn0.4Ni0.5Fe2O4 for bolstering pollutant removal effectiveness while upholding magnetic properties and then coated with a mesoporous silica layer. The surface characteristics and composition of Ag0.1Zn0.4Ni0.5Fe2O4@mSiO2 were confirmed using EDX, FT-IR, VSM, XRD, TEM, SEM, and BET methods. The surface modification of Ag-Zn-Ni ferrite with a silica layer improves the texture properties, where the specific surface area and average pore size of the spinel ferrite rose to 180 m2/g and 3.15 nm, respectively. The prepared spinel ferrite@mSiO2 has been utilized as an efficient adsorbent for eliminating methyl green (MG) and indigo carmine (IC) as models of cationic and anionic dyes from wastewater, respectively. Studying pH, Pzc, adsorbent dosage, dye concentration, and temperature showed that efficient removal of MG was carried out in alkaline media (pH = 12), while the acid medium (pH = 2) was effective for IC removal. Langmuir isotherm and pseudo-second-order kinetics were found to be good fits for the adsorption data. Both dyes were adsorbed in a spontaneous, endothermic process. A possible mechanism for dye removal has been proposed. The adsorbent was effectively recovered and reused.

The rapid industrialization of recent times has resulted in a significant escalation in the release of diverse toxic materials into our marine environment, presenting a critical hazard to environmental preservation1. The textile industry is one of the anthropogenic activities that most consume water and pollute water bodies2. Synthetic dyes, which originate from the chemical and textile sectors, are particularly worrisome among these pollutants on account of their enduring nature, toxicological properties, and ability to cause environmental damage3. The removal of these harmful dye pollutants has emerged as a critical area of research, given the dual implications for safeguarding human health and the environment.

Adsorption is commonly used to remove organic pollutants and dyes from wastewater. Although there are many methods for treating wastewater, including oxidation, flocculation, coagulation, membrane filtration, chemical precipitation, ion exchange, and biological treatment4,5,6,7, adsorption is widely regarded as a straightforward, low-cost, risk-free, and effective method for removing dyes. Bentonite8, clay minerals9 and agricultural waste-based adsorbents10 are only a few of the adsorbents that have been used for dye removal from aqueous solutions.

One effective method for removing dye is by using magnetic materials, specifically metal ferrites. Due to their magnetic properties, metal ferrites have shown significant potential for absorbing and eliminating different dye compounds from wastewater. Metal ferrites possess distinctive attributes, including elevated surface area, heightened magnetic susceptibility, and chemical stability, which make them well-suited contenders for environmental cleanup11,12,13. Nevertheless, improving the efficiency of mixed ferrites for the purpose of dye removal continues to be a difficult task, but it can be overcome by implementing surface modification techniques.

Metal ferrites possess the desirable characteristic of being adjustable, which greatly enhances their appeal for a wide range of uses such as catalysis14, sensors15, dielectric16, biomedical17, and environmental remediation18. Due to their magnetic properties, these substances may be easily separated from the treated water, making them very suitable for magnetic-based systems designed to remove dyes19.

Numerous variables influence the properties of ferrite, such as cation distribution20,21, preparation conditions, size, microstructure, thermal treatment, and synthesis method. Ferrites have been produced through a variety of processes21,22,23. Co-precipitation is regarded as an exceptional synthesis technique owing to its numerous advantages, which encompass minute particle size24, exceptional homogeneity, and the possibility of producing crystalline and high-yield nanoparticles of ultrafine size may be produced via this process25. Spinel ferrites have become a significant group of ferrites that belong to the metal ferrite family. These compounds exhibit a cubic crystal structure where magnetic cations are evenly distributed in octahedral and tetrahedral positions26, leading to their remarkable magnetic characteristics27.

Nickel-ferrite nanoparticles NiFe2O4 is a spinel ferrite with high magnetic, chemical, optical, electrical, and dielectric properties28,29. Which is used for adsorption and other application18,30. Zn2+ and Ag+ ions were often selected among various dopants. Zinc ferrites are ordinary spinel ferrite with chemical, optical, electrical, and dielectric properties. There are numerous potential applications for zinc ferrites, including hyperthermia medication for cancer31, photocatalysis32, organic transformation catalysts33, wastewater remediation34, and sensors35. The incorporation of silver into metal ferrite nanoparticles has shown significant promise in enhancing different elements of wastewater treatment36,37. The addition of silver improves the adsorption efficiency of magnetic ferrite by introducing more active sites on its surface and enhances its photocatalytic degradation and antibacterial properties38,39. As a result, these composite materials are highly efficient at removing pollutants and ensuring the quality and safety of water resources.

Although mixed ferrites possess inherent characteristics that make them effective for dye removal, there is still potential for enhancing their efficiency and selectivity. This is when the importance of silica coating becomes relevant. Silica is a favorable material for surface coating due to its stability, chemical inertness, lack of toxicity, large surface area, high rigidity, thermal stability, low cost, being environmentally friendly, and facile modification40,41,42. Silica coating improves the adsorption capacity over a wide range of pH values, which makes it suitable for adsorption under varying solution conditions43. The thermal stability of the silica layer allows for regeneration and reuse of the adsorbent after adsorption cycles. This versatility makes silica coating suitable for a wide range of adsorption applications.

In the current study, a novel silver-zinc-nickel ferrite is synthesized and coated with mesoporous silica Ag0.1Zn0.4Ni0.5Fe2O4@mSiO2 as a powerful adsorbent to remove methyl green (MG) and indigo carmine (IC) as models of cationic and anionic dyes, respectively, from wastewater.

Pure and high grade ferric nitrate nonahydrate (Fe (NO3)3.9H2O) (98%), zinc nitrate hexahydrate (Zn (NO3)2.6H2O) (98%), nickel nitrate hexahydrate (Ni(NO3)2.6H2O) (97%), silver nitrate (AgNO3) (99%), cetyl trimethyl ammonium bromide (CTAB) (98%), tetraethyl ortho silicate (TEOS)(98%) were obtained from Sigma-Aldrich chemicals, methyl green (C27H35BrClN3) (85%) and indigo carmine (C₁₆H₈N₂Na₂O₈S₂) (85%) were obtained from Merck, and Fig. 1 showed their molecular structures.

Structure of the investigated dyes.

Mixed ferrite with different compositions, AgxZn(0.5-x)Ni0.5Fe2O4, where x = (0.1, 0.3, 0.5), was prepared by the co-precipitation method44. Briefly, Zn (NO3)2.6H2O (0.4 M), AgNO3 (0.1 M), Ni (NO3)2.6H2O (0.5 M), and Fe (NO3)3.9H2O (2 M) were dissolved in distilled water and mixed together with a molar ratio of 1:2, and then NaOH (3 M) was added until a precipitate was formed. The reaction was refluxed at 80°C for two hours. The brownish precipitate of spinel ferrite was then magnetically decanted. The precipitate was washed with distilled water and then ethanol. The sample was dried overnight at 80°C and then characterized using a variety of analytical techniques.

Ag0.1Zn0.4Ni0.5Fe2O4@mSiO2 (SFS) was synthesized according to Liu et al45 with modifications as shown in Scheme 2: 0.06 g of previously prepared mixed ferrite and 2 mL of chloroform were mixed and exposed to ultrasonic radiation of frequency 20 kHz (power of 70 W) for 60 min. Then add a mixture of 36 ml ethanol and 25 ml of dist. water to the solution, stirring for 10 min at 30 °C. After dispersing the solution, 0.16 g of CTAB was added and swirled at 65 °C for 30 min. The mixture was agitated for three hours after TEOS (1 ml) and 25% NH4OH (1 mL) were added. The gathered product, which is a light brown colour, is rinsed with distilled water and dried for 18 h. The product is removed from CTAB by being refluxed in 200 cc ethanol for 24 h, as shown in Fig. 2.

Illustrative steps for the synthesis of SFS.

Powder X-ray diffractometry (XRD) was used to examine the crystalline structure of the prepared samples (GNR APD 2000 PRO diffractometer, CuK α radiation, operated at 40 kV, 30 mA, and 2θ in the range of 10–70°). The molecular structure was studied using FTIR (JASCO, model 4100) with KBr pellets. The elemental compositions were investigated using SEM connected with EDX (IT100LA) operating at 20 kV. The magnetic properties of the samples were measured using VSM (Lake Shore, model 7410). SEM (SU8000 Type II, Hitachi) was used to analyse surface morphology. TEM (JEOL, model JEM-2100) analysis was used to determine particle size and shape. Surface area was determined by the BET method (Quantachrome Touch Win™, version 1.21), and both metal ferrite and metal ferrite@silica were analyzed. Prior to the measurements, the samples were degassed for 12 h at 120 °C. The BJH (Barrett-Joyner-Halenda) method was used to analyse pore size distribution curves and derive the pore volume and pore size (diameter).

To investigate how pH affects spinel ferrite (SF) adsorption, the pH of zero charge (pHpzc) was investigated. Rivera et al.46, report adjusting the initial pH of 50 ml of 0.01 M NaCl solutions in Erlenmeyer flasks by adding 0.1 M HCl or 0.1 M NaOH solutions to a pH range of 2–12. 0.1 g of adsorbent was added to each flask. After 24 h of agitation at room temperature, the pH of the resulting solutions was determined. By contrasting the initial and final pH values (not shown), the pHPZC was calculated46,47. The obtained pHPZC was 6.80 and 6.45 for Ag0.1Zn0.4Ni0.5Fe2O4 (SF) and Ag0.1Zn0.4Ni0.5Fe2O4@mSiO2 (SFS), respectively.

To assess the adsorption capabilities of the investigated ferrite, a variety of adsorption experiments have been performed. In each experiment, several 100-mL Erlenmeyer flasks were filled with a known amount of adsorbent (5–30 mg) and 25 mL of a dye solution with a known concentration. The containers were placed in a water shaker thermostat and agitated at 120 rpm at a working temperature. After specified intervals of time, the adsorbent was extracted magnetically. The concentrations of MG and IC in the supernatant were determined by measuring the absorbance at 632 nm for MG and 610 nm for IC by using UV–Vis spectrophotometry (Varian Cary 400). The following Eqs. (1–3) were utilized to determine the dye removal percentage and the amount of dye adsorbed at time t (qt) and at equilibrium (qe)48 \(.\)

where Co is the initial concentration, Ct is the concentration at time t, and Ce is the concentration of the dyes at equilibrium. The volume of the working solution is denoted by V (L), and the mass of the nanocomposite is m (g). Using universal buffer and either HCl (0.1 M) or NaOH (0.1 M), the pH of the solution was maintained within the desired range (2–12).

The phase composition and crystallinity of spinel ferrite AgxZn(0.5-x)Ni0.5Fe2O4, where x = (0.1, 0.3, 0.5) (SF) and Ag0.1Zn0.4Ni0.5Fe2O4@mSiO2 (SFS) were analyzed by XRD (Fig. 3). All prepared ferrite AgxZn(0.5-x)Ni0.5Fe2O4 exhibited characteristic peaks of the spinel structure of the face-centered cubic (fcc), according to the database of JCPDS (code 00-052-0279)49. These peaks appeared at 2θ = 29.8°, 35.1°, 38.07°, 42.65°, 53.09°, 56.56°, and 62.09°, which could be imputed to (220), (3 1 1), (2 2 2), (4 0 0), (4 2 2), (333), and (4 4 0), respectively. The spinel phase’s development is shown by the strong peak visible at the (311) plane50. The highest peaks slightly migrated towards lower angles as the Ag+ content rose (when x = 0.5), confirming the Ag+ substitution for Zn2+. Additionally, the peaks widened, indicating that the ferrite nanostructure had formed51. The broad peak in the range from 20° to 25° is attributed to the existence of amorphous SiO2 in the coated layer52.

XRD pattern of spinel ferrite AgxZn(0.5-x)Ni0.5Fe2O4 where x = (0.1, 0.3, 0.5) and Ag0.1Zn0.4 Ni0.5Fe2O4@mSiO2.

Scherrer's formula53, was used for calculating the size of the crystallite (d) from the strongest peak.

where β is the full width at half-maximum for the (311) peak in radians, λ is wavelength, and θ is diffraction angle. The size of the crystallites increased from 5.6 to 10 nm when the Ag+ content rose (Table 1). The difference in ionic radii between Ag+ (1.26 A°) and Zn2+ (0.74 A°) may be the cause of this. It was found that the average crystallite size of silica-coated ferrite increased from 5.6 nm for uncoated ferrite to 8.35 nm. This shows that the silica layer was successfully coated.

FTIR spectra of prepared metal ferrites are shown in Fig. 4; they span 400–4000 cm−1. The distinctive band of all metal ferrite AgxZn(0.5-x)Ni0.5Fe2O4, where x = (0.1, 0.3, 0.5), are nearly similar, indicating the same structure of all prepared metal ferrite. There are two prominent absorption peaks observed at 585 cm−1 and 410 cm−1. These peaks correspond to the vibrational modes of tetrahedral metal–oxygen bonds and octahedral metal–oxygen bonds, respectively54. Moreover, when the silver doping ratio is increased, particularly at x = 0.5, both the bands related to octahedral and tetrahedral sites display a displacement towards higher wavenumbers. This shift happens due to the replacement of Ag+ ions with Zn2+ ions, which have a higher atomic weight55.

FT-IR spectra of spinel ferrite AgxZn(0.5-x)Ni0.5Fe2O4 where x = (0.1, 0.3, 0.5) and Ag0.1Zn0.4 Ni0.5Fe2O4@mSiO2.

The strong band at approximately 3432 cm−1 and the small band at 1638 cm−1 refer to the stretching and bending vibrations of OH groups56. The spectrum of Ag0.1Zn0.4Ni0.5Fe2O4@mSiO2 (Fig. 4) confirms the presence of the same bands as in mixed ferrite as well as new bands at 1079.54 cm−1, 965.52 cm−1, 798.08 cm−1 and 455.8 cm−1. The stretching Si–O–Si, bending Si–O–Si, streching Si–O, and bending Si–O bands are presented here in the correct order57. All these bands indicate that SiO2 was successfully loaded onto Ag0.1Zn0.4Ni0.5Fe2O4.

EDX was used to analyse synthetic samples for elemental makeup (Fig. 5). The atomic percentages of Ag, Zn, and Ni in addition to Fe and O components in prepared samples indicate that spinel ferrite has been successfully prepared. The main elements of the all spinel ferrite AgxZn(0.5-x)Ni0.5Fe2O4, where x = (0.1, 0.3, 0.5), were Ag, Zn, Ni, Fe, and O (Fig. 5a and Table 2). In addition to the previously mentioned elements, Si appeared in the spectrum of Ag0.1Zn0.4Ni0.5Fe2O4@mSiO2 (Fig. 5b and Table 2), which is largely related to SiO2. This reflects the effective loading of silica on the ferrite surface of the composition Ag0.1Zn0.4Ni0.5Fe2O4.

The EDX spectra and atomic percentage of uncoated spinel ferrite Ag0.1Zn0.4Ni0.5Fe2O4 (a) and Ag0.1Zn0.4Ni0.5Fe2O4@mSiO2 (b).

At room temperature, VSM confirmed the samples' magnetic characteristics. The magnetic permeability profiles are shown in Fig. 6. All the different compositions of mixed ferrite AgxZn(0.5-x)Ni0.5Fe2O4, where x = (0.1, 0.3, 0.5), are superparamagnetic, as evidenced by the presence of a slight hysteresis. In Table 3, the values for saturation magnetization, remanence, and coercivity are recorded.

Magnetization curves of spinel ferrite Agx Zn(0.5-x)Ni0.5Fe2O4 where x = (0.1, 0.3, 0.5) and Ag0.1Zn0.4 Ni0.5Fe2O4@mSiO2.

The saturation magnetization values of mixed ferrite AgxZn(0.5-x)Ni0.5Fe2O4, where x = (0.1, 0.3, 0.5), decrease as the ratio of Ag increases, mainly due to the presence of diamagnetic silver components in the Ag-based nanoferrite samples58,59. In the case of Ag0.1Zn0.4Ni0.5Fe2O4@mSiO2, the nonmagnetic silica shell is responsible for the drop in saturation magnetization, coercivity, and remanence of Ag0.1Zn0.4Ni0.5Fe2O4@mSiO260,61. However, the saturation magnetization of 14.1 emu/g was sufficient for facile separation of Ag0.1Zn0.4Ni0.5Fe2O4@mSiO2 from aqueous solutions with a magnet.

SEM and TEM techniques were employed to examine the morphological characteristics of the samples. In the SEM image of Ag0.1Zn0.4Ni0.5Fe2O4@mSiO2 (Fig. 7a), a smooth surface composed of spherical crystals was observed. TEM images of Ag0.1Zn0.4Ni0.5Fe2O4 and Ag0.1Zn0.4Ni0.5Fe2O4@mSiO2 illustrated the spherical morphology of the particles, with an average particle size of approximately 1 µm (Fig. 7b,c). Figure 7c exhibited the successful encapsulation of the ferrite within a thin core–shell structure of silica. Additionally, Fig. 7d displayed the presence of a thin mesoporous silica layer with silica aggregations at various locations on the surface of the ferrite, highlighting its porous nature in HRTEM.

SEM image of SFS (a) and TEM images of SF (b) and SFS (c,d).

The adsorbent surface area is a key factor in raising adsorption efficiency. According to BET and porosity measurements, the specific surface area and pore volume of spinel ferrite were determined to be 69.79 m2/g and 0.404 cm3/g, respectively. Coating the prepared spinel ferrite with a silica layer increased the specific surface area and pore volume to 180.81 m2/g and 0.771 cm3/g, respectively. The synthesized materials are porous, as shown by the average pore size (Table 4). According to N2-adsorption/desorption isotherm Fig. 8, both the SF and the SFS displayed a type IV with a hysteresis loop of the H2(b) type, indicating a mesoporous nature with a wide range of pore size distribution62. This takes place in porous materials characterized by a network of interconnected pores as well as in pores characterized by a broad body and a thin neck63,64.

Adsorption–desorption of N2 gas by SF (a) and SFS (b).

This research sets out to develop an efficient adsorbent for the elimination of harmful anionic and cationic dyes that pollute water and endanger human health. As a result, the adsorption abilities of different compositions of spinel ferrite AgxZn(0.5-x)Ni0.5Fe2O4, where x = 0.1, 0.3, and 0.5, to remove MG and IC from simulated wastewater were studied and compared. It was found that Ag0.1Zn0.4Ni0.5Fe2O4 (SF) is more efficient in dye removal, as shown in Fig. 9. Therefore, Ag0.1Zn0.4Ni0.5Fe2O4 was chosen for more adsorption studies, especially because it has the highest saturation magnetization for easy separation. Subsequently, the coating of the chosen spinel ferrite (x = 0.1) with a silica layer increased in surface area from 69.79 to 180.08 m2/g, as proven by BET. As shown in Fig. 10, the absorbance of MG and IC dyes diminishes over time after adsorption by SFS.

The comparison of the removal efficiency of MG (4 mg/L) and IC (19 mg/L) utilising 20 mg of various adsorbents AgxZn(0.5−x)Ni0.5Fe2O4 at pH = 7 and 25 °C.

The absorbance of MG (13 mg/L) (a) and IC (37 mg/L) (b) against time during adsorption onto SFS (20 mg), pH = 7, 25 °C.

Significant effects of initial dye concentration, nanocomposite quantity, temperature, and solution pH on the adsorption of various dyes by spinel ferrite@mSiO2 (SFS) were observed. The results of varying each variable were examined carefully.

The effectiveness of SFS in eliminating MG and IC was studied by adjusting the dosage from 5 to 30 mg. As illustrated in Fig. 11, the removal percentage for dyes increases as the number of adsorbents increases. This is primarily due to the greater availability of adsorption sites on the adsorbent surface, which allows for more effective interaction between the dye molecules and the adsorption sites65,66,67.

Removal efficiencies of MG (13 mg/L) and IC (37 mg/L) using SFS as a function of adsorbent dosage.

The pH level plays a crucial role in influencing the adsorption process. Both adsorbate dyes (MG or IC) and absorbent SFS possess various surface functional groups that gain or lose protons (H+) in response to the pH of the medium68. To optimize the adsorption process for pollutant removal, the adsorption process for MG and IC was examined across the pH range (2–12), as shown in Fig. 12.

The removal efficiency of MG (13 mg/L) and IC (37 mg/L) utilizing SFS as a function of medium pH molecules are now blocking the active sites of the remaining molecules in solution75.

Pzc measurements show that the point of zero charge (pHpzc) of SFS is 6.45. At a pH > 6.45, SFS's surface turns negatively charged due to the ionization of H+ from the active groups leaving negative charges; as a result, MG removal efficiencies increase with increasing pH above 6.45, where they reached 96.6% at a pH of 10 due to electrostatic interactions between cationic methyl green and negatively charged adsorbents69.

However, IC is negatively charged at low pH and has a pKa of 12.670,71. So, the removal efficiencies of IC decrease at pH > 6.45 due to the increased electrostatic repulsion with negatively charged adsorbent surfaces. This result agrees with the work of Yazdi et al.72. As a result, the removal efficiency of IC increases with decreasing pH value until it reaches 93.8% at pH = 273 (Fig. 12), due to the electrostatic attraction between the cation surface of spinel ferrite and the negative IC, and furthermore, the formation of a hydrogen bond. As a result, the alkaline medium was the right choice for removing methyl green dye, and the acidic medium is efficient for removing IC.

The influence of dye concentration on the removal efficiency of pollutant dyes was investigated, as shown in Fig. 13. The removal percentages of MG and IC reached 91% within 90 min below the concentrations of 7 and 19 mg/L for MG and IC, respectively. As the dye concentration increases within 90 min, the percentage of MG and IC that is removed decreases. With an increase in dye concentration, the availability of adsorption sites became restricted, resulting in a decrease in dye removal prior to achieving equilibrium74. At higher concentrations, however, the removal effectiveness noticeably drops as a larger number of dye molecules compete for a smaller number of adsorption sites. This decrease is because the previously adsorbed dye molecules are now blocking the active sites of the remaining molecules in solution75.

The removal efficiency of MG (3.2–16.3 mg/L) and IC (18.6–56 mg/L) utilizing SFS as a function of initial dye concentration.

The time-dependent behaviour of adsorption is depicted by the linear and non-linear adsorption kinetics of the pseudo-first and second-order models in, Eqs. (5)–(8)76. For the water treatment to be successful and economical, the adsorption process must be completed quickly. The contact time is the duration of time it takes for the maximum amount of dye adsorbed to reach equilibrium with the adsorbent surface in an adsorption experiment. Figure 14 shows how the quantity of MG and IC adsorbed (qt) varies with contact time in a nonlinear kinetic model. After 110 min of interaction, the qt had climbed to an equilibrium level. Also, the mechanism of adsorption of MG or IC dyes onto synthesized ferrite is determined by evaluating the adsorption kinetics in the intra-particle diffusion model, Eq. (9)77.

Non-linear kinetic model for adsorption of MG (a) and IC (b) on SFS.

Adsorbed dye amounts are given in terms of mg/g at equilibrium (qe) and contact time (qt) in minutes. Pseudo-first order diffusion has a rate constant of k1 (min−1), pseudo-second order diffusion has a rate constant of k2 (g/mg min), and intra-particle diffusion has a rate constant of ki (mg/g min1/2). The values of the correlation coefficient (R2) in Table 5, indicate that the linear and nonlinear equations forms of the pseudo-second-order approach, Eqs. (7, 8), respectively, are a better fit and more suited for the experimental data than the linear and nonlinear pseudo-first-order model Eqs. (5, 6), respectively. Also, for each dye under examination, the qe, exp, and the qe estimated from the pseudo-second-order model are all quite close to one another, within the experimental errors. Adsorption kinetics may be altered by a number of different processes. Most constrained are the processes of diffusion, which can only occur via (a) extracellular diffusion, (b) boundary layer diffusion, and (c) intra-particle diffusion77,78. Therefore, an intra-particle diffusion kinetic model is used to further evaluate the adsorption data and forecast the rate-limiting stage in the process. If the qt vs. t0.5 linear plot is through the origin, then intra-particle diffusion is supposed to be in charge of the adsorption process, as predicted by this theory. Figure 15 shows that qt is linearly related to t0.5. Here we see examples of both the outer and inner surface diffusions; the former refers to the surface's exterior while the latter describes its interior. According to Table 6, there is a completely quick step of diffusion from the outside onto the adsorbent surface, followed by a completely sluggish step of diffusion into the inner surface75.

Intra-particle diffusion kinetics model of the of MG and IC on SFS.

The surface properties, maximum capacity of the adsorbent, and adsorption mechanism were all stated by the values of certain parameters determined from adsorption isotherm models used to examine the produced ferrite's potency and capacity as an adsorbent. Different adsorption isotherm models were employed to compare the results; these included the Langmuir, Freundlich, Temkin, and Dubinin–Radushkevich (D–R) models. Figure 16 shows nonlinear adsorption isotherm of MG and IC on SFS as a compared with SF adsorbent.

Nonlinear adsorption isotherm of MG (a,b) and IC (c,d) on SF and SFS, respectively.

Adsorption of dye molecules onto an adsorbent surface has the same activation energy according to the Langmuir isotherm, as shown in Fig. 16, which assumes monolayer adsorption above a homogenous adsorbent surface with a set number of identical sites79. This led to the adsorption data of MG and IC being fitted into the nonlinear form of Langmuir's equation80, the determination and quantification of the maximum adsorption capacity (qmax) were performed and tabulated in Table 7.

where Ce is the concentration of dye at equilibrium in mg/l, qe is the adsorption capacity of dye at equilibrium in mg/g, qmax is the maximum possible adsorption capacity in mg/g, and KL is the Langmuir constant in L/mg.

Table 7 shows the results. Adsorption results agreed with the Langmuir model, which postulated the existence of homogenous adsorption sites on SFS substrate, as indicated by high values of the correlation coefficient (R2) for the isotherm plots. The IC dye has sulfonated, amine, and hydroxyl groups in its structure that support electrostatic binding with mesoporous silicate Si–OH through hydrogen bonds, which may explain why the IC dye has a higher Langmuir monolayer coverage capacity value (qmax) than the MG dye and why the value of the same dye increases when using SFS as an absorbent. The results show that ultimate capacity qmax and KL values increased as the temperature of adsorption increased, thus, the affinity of the adsorbent to the investigated dyes increased with increasing temperature.

The maximum monolayer adsorption capacity (qmax) for methyl green and indigo carmine adsorption is compared across a variety of adsorbent surfaces in Table 8. We found that our SFS as an adsorbent is effective in removing methyl green and indigo carmine from aqueous solutions, when compared to data from the relevant literature.

An experimental expression that accounts for surface heterogeneity and an exponential distribution of site and energy energies is the Freundlich isotherm81. The Freundlich adsorption Eq. (11) describes this equilibrium.

The Freundlich isotherm constant is denoted by kF ((mg/g) (mg/L)−1/nF)82, where n is the dimensionless exponent of the Freundlich. KF and n are the determined values. Estimates of nF in this work ranged from 4 to 7, indicating that the adsorption processes are close to irreversible83. When the KF value is large, the adsorption is driven with greater efficiency. This means that the adsorption driving force for IC dye is greater than that for MG dye, and the driving force grows when SFS is employed as the adsorbent. The favourable nature of the adsorption process is confirmed at higher temperatures by an increase in KF value84.

Adsorption data for MG and IC were analyzed using the Temkin isotherm model. This model supports limited interactions between the adsorbent and adsorbate and demonstrates that all molecules in the surface layer have decreased adsorption energies at the cover surface. Additionally, it was believed that adsorbate-adsorbent interactions directly reduced the adsorption heat with coverage85. Equation (12) represents the Temkin isotherm model.

where KT (L/mol) represents the equilibrium binding constant, R is the perfect gas constant, T is the absolute temperature, and b (kJ/mol) represents the heat of adsorption. Table 7's b values match up with an adsorption mechanism involving electrostatic interactions and hydrogen bond formation. Temperature has a beneficial influence on the binding energy of dyes with ferrite surfaces, as evidenced by the rise in KT as the temperature rises.

To identify the adsorption mechanism (physical or chemical), we can use Dubinin–Radushkevich isotherm Eqs. (13–15) to determine the activation energy E (KJ/mol). Adsorption has been explained by ion exchange adsorption if E ranges from 8 to 16 kJ/mol; if E is 8, the process has been validated physically. Adsorption in this study may occur by chemisorption or ion exchange, where E ranges from 8 to 16 kJ/mol.

where qs (mol/kg) is the theoretical capacity of adsorption calculated from Eq. (13) of the D-R model, ε (kJ/mol) is the Polanyi potential, and β (mol2/kJ2) is the mean free energy of adsorption for each molecule adsorbed as given by Eq. (14).

The Langmuir model is the best-fitted model at the different temperatures according to the R2 values in Table 7. The effect of temperature on the adsorption process according to the Langmuir isotherm is studied as shown in Fig. 17, where qmax and KL (L/mg) at different temperatures are shown in Table 9. To determine the thermodynamic parameters, the obtained equilibrium constant must become dimensionless before being applied to the Vant´Hoff equation using Eq. 1695.

where γ is the activity coefficient (dimensionless) and [Dye]° is the standard concentration of dye (1 mol/L).

Nonlinear Langmuir adsorption isotherm of MG (a,b) and IC (c,d) on SF and SFS, respectively at different temperatures.

The entropy ΔS° and enthalpy ΔH° of adsorption are determined according to the Eqs. (17, 18) from the intercept and slope of the relationship between \(ln{K}_{e}^{0}\) vs. 1000/T, respectively, where R is the perfect gas constant (8.314 J/mol K), and T is the absolute temperature96. The adsorption of MG or IC dye onto the synthesized ferrite is endothermic due to the positive values of ΔH° as shown in Table 9. Also, the measured ΔH° values varied from 25 to 48 kJ/mol, suggesting that both physisorption and chemisorption may be involved in the adsorption of the dyes of interest97. Adsorption is characterized by an increase in the degree of disorder at the solid-solution interface, as measured by an increase in the entropy parameter, ΔS°98. This is the normal outcome of electrostatic force interactions that lead to the phenomenon known as physical adsorption. The spontaneous process for MG and IC sorption is revealed by negative values of ΔG°, and as the temperature is raised, the value of ΔG° decreases even further, showing that the adsorption process is more favoured at a higher temperature99.

A Comparison of the FTIR spectra of the used SFS before and after adsorption of IC as a model of the investigated dye showed the appearance of additional peaks after their contact with the IC dye (Fig. 18). Peaks appearing at 2930 cm were due to CH stretching, at 1374 cm−1, it corresponds to the C–H bending vibration or C–N stretching vibration of amines100. A comparison between newly appeared peaks and those of IC showed that the new peaks were probably the characteristic of this textile pollutant. The band broadening about 3400 cm−1 is assigned to the vibration formed by hydrogen-bonded –O–H groups in the absorbent and adsorbate molecules. The band broadening at 3390, 3433 cm−1 indicate to the hydrogen bonding in comparison to the FTIR analysis before the adsorption process as well as existence of (-NH) stretching bond of dye101.

FTIR of SFS before (a) and after adsorption (b) of IC.

The N–H bond between the dye's N atom and the H–O–H bond resulted in a noticeable shift of the H–O–H peak from 1638 to 1627 and an increase in its width. This effect resulted from the establishment of oxygen-hydrogen bonds between the dye and the adsorbent102. The adsorption of molecules onto the SFS nanoparticles in this study is consistent with physical adsorption mechanisms, involving hydrogen bonding and electrostatic forces. The low adsorption heat indicated by the Temkin isotherm model, the formation of a monomolecular layer according to the Langmuir isotherm, and FTIR analysis collectively support the physical adsorption process on the SFS nanoparticle surface.

Scanning electron microscopy revealed a substantial change in surface shape following adsorption (Fig. 19). SEM–EDX analysis was used to analyse the morphological aspects of SFS. The dye molecules appear on the surface of spinel ferrite. The principal components of ferrite/IC are Fe, Ni, Zn, O, Si, and C, N from dyes, according to the EDX analysis, confirming the efficient adsorption of IC.

SEM/EDX of SFS after adsorption of IC.

The regeneration and reusability of adsorbent surfaces are critical aspects of determining their true value. The use of adsorbents with high adsorption capacity and high desorption assets reduces additional environmental pollution and overall costs. As a result, desorption studies on SFS are carried out to determine its recyclable accessibility. The adsorbent was recovered by treating the used sample with 0.1 M HCl for 2 h, washing it three times with distilled water, and finally testing the filtrate with a silver nitrate solution to make sure the HCl was gone before drying it at 45 °C for 18 h. For four cycles, 20 mg of adsorbent was used. The elimination percentage changed with each cycle, as seen in Fig. 20103.

The reuse of dyed SFS samples.

Based on the provided information and the characteristics of the adsorption process, here is a suggested mechanism for the adsorption of MG and IC dyes onto the synthesized ferrite material. From pH and PZC studies, when the pH level goes above 6.45, active groups ionize and make the surface of the synthesised SFS material negatively charged. This negative charge makes it easier for the positively charged cationic methyl green (MG) dye molecules to stick to the negatively charged surface of the adsorbent. This electrostatic interaction is a dominant driving force for MG adsorption. On the other hand, indigo carmine (IC) dye is negatively charged at low pH. The IC dye contains sulfonated, amine, and hydroxyl groups in its structure. These functional groups can form hydrogen bonds with the mesoporous silicate (Si–OH) on the adsorbent surface. The presence of hydrogen bonding sites enhances the adsorption of IC dye onto the adsorbent. All these were confirmed by Temkin, Freundlich, and Dubinin-Radushkevich. It is thought that an increase in temperature gives more thermal energy, which helps both MG and IC dyes stick to the adsorbent surface because the adsorption process is endothermic. This increased energy facilitates the interactions between the dye molecules and the adsorbent. The positive values of ΔH°, which range from 25 to 48 kJ/mol, suggest that MG and IC dyes may bind by both physisorption and chemisorption mechanisms, using hydrogen bonds and electrostatic interactions. The second order and intra-particle diffusion kinetics indicate that the adsorption process involves a rapid initial step of diffusion from the external solution to the adsorbent surface, followed by a slower step of diffusion into the inner surface of the adsorbent as shown in Fig. 21.

Proposed mechanism of the adsorption of MG and IC dyes on SFS.

Spinel ferrite@SiO2 (SFS) is a powerful adsorbent for cleaning textile processing wastewater containing cationic and anionic dyes. This assertion has been supported by the high surface area provided by the coated silica and the ease of separation provided by the magnetite core. The adsorptive activity of SFS against pollutant dyes is due to electrostatic interactions and hydrogen bond formation via Si–OH and Si=O, including wrinkles, flaws, and residual functional groups. Notably, the adsorption efficiency for anionic dye removal is significantly higher than that for cationic dye. Therefore, the adsorbate's ionic characteristics and the nature of the adsorbent's active sites determine the adsorption mechanism. As an anionic dye, IC's molecular structure includes –SO3−, OH−, and NH− groups, all of which are capable of interacting with the absorbent. The aforementioned energetic sites explain why they are so effective at attracting ICs. MG, a cationic dye, should have a lower adsorption capability since it contains a positive canter and a smaller active group capable of forming hydrogen bonds than IC.

It is speculated that the ionic nature of both MG and IC is responsible for their adsorption onto the SFS. Langmuir isotherm model fitting confirmed the development of a dye monolayer across the adsorbent surface. The dramatic rise in qe, cal is reflected in the qmax that was attained. Since physisorption results in lower free energy values ΔG°, it is the adsorption mechanism of choice.

No ethical issues were violated in this study.

All the authors agreed to participate in this work.

All data generated or analyzed during this study are included in this article.

Bhatia, S. & Devraj, S. Pollution control in textile industry, https://doi.org/10.1201/9781315148588 (WPI publishing, 2017).

Lellis, B., Fávaro-Polonio, C. Z., Pamphile, J. A. & Polonio, J. C. Effects of textile dyes on health and the environment and bioremediation potential of living organisms. Biotechnol. Res. Innov. 3, 275–290. https://doi.org/10.1016/j.biori.2019.09.001 (2019).

Article Google Scholar

Hajati, S. et al. Competitive adsorption of Direct Yellow 12 and Reactive Orange 12 on ZnS: Mn nanoparticles loaded on activated carbon as novel adsorbent. J. Ind. Eng. Chem. 20, 564–571. https://doi.org/10.1016/j.jiec.2013.05.015 (2014).

Article CAS Google Scholar

Fan, S. et al. Removal of methylene blue from aqueous solution by sewage sludge-derived biochar: Adsorption kinetics, equilibrium, thermodynamics and mechanism. J. Environ. Chem. Eng. 5, 601–611. https://doi.org/10.1016/j.jece.2016.12.019 (2017).

Article CAS Google Scholar

Mahmoud, D. K., Salleh, M. A. M., Karim, W. A. W. A., Idris, A. & Abidin, Z. Z. Batch adsorption of basic dye using acid treated kenaf fibre char: Equilibrium, kinetic and thermodynamic studies. Chem. Eng. J. 181, 449–457. https://doi.org/10.1016/j.cej.2011.11.116 (2012).

Article CAS Google Scholar

Aboelfetoh, E. F., Aboubaraka, A. E. & Ebeid, E.-Z.M. Binary coagulation system (graphene oxide/chitosan) for polluted surface water treatment. J. Environ. Manage. 288, 112481. https://doi.org/10.1016/j.jenvman.2021.112481 (2021).

Article CAS PubMed Google Scholar

Aboelfetoh, E. F., El-Attar, H. G. & Okba, E. A. Facile synthesis of magnetic and porous zeolite/SnFe2O4 nanocomposite for cationic and anionic dyes deterioration. Microp. Mesop. Mater. 357, 112611. https://doi.org/10.1016/j.micromeso.2023.112611 (2023).

Article CAS Google Scholar

Bergaoui, M. et al. Novel insights into the adsorption mechanism of methylene blue onto organo-bentonite: Adsorption isotherms modeling and molecular simulation. J. Mol. Liquids 272, 697–707. https://doi.org/10.1016/j.molliq.2018.10.001 (2018).

Article CAS Google Scholar

Youcef, L. D., Belaroui, L. S. & López-Galindo, A. Adsorption of a cationic methylene blue dye on an Algerian palygorskite. Appl. Clay Sci. 179, 105145. https://doi.org/10.1016/j.molliq.2018.10.001 (2019).

Article CAS Google Scholar

Kadhom, M., Albayati, N. & Alalwan, H. & Al-Furaiji, M. https://doi.org/10.1016/j.scp.2020.100259 (2020).

Article Google Scholar

Bayantong, A. R. B. et al. Adsorptive removal of dye in wastewater by metal ferrite-enabled graphene oxide nanocomposites. Chemosphere 274, 129518. https://doi.org/10.1016/j.chemosphere.2020.129518 (2021).

Article CAS PubMed Google Scholar

Novoselova, L. Y. Nanoscale magnetite: New synthesis approach, structure and properties. Appl. Surf. Sci. 539, 148275. https://doi.org/10.1016/j.apsusc.2020.148275 (2021).

Article CAS Google Scholar

Panda, S. K. et al. Magnetite nanoparticles as sorbents for dye removal: a review. 19, 2487–2525. https://doi.org/10.1007/s10311-020-01173-9 (2021).

Article CAS Google Scholar

Kharisov, B. I., Dias, H. R. & Kharissova, O. V. Mini-review: Ferrite nanoparticles in the catalysis. Arab. J. Chem. 12, 1234–1246. https://doi.org/10.1016/j.molliq.2018.10.001 (2019).

Article CAS Google Scholar

Wu, K., Li, J. & Zhang, C. Zinc ferrite based gas sensors: A review. Ceram. Int. 45, 11143–11157. https://doi.org/10.1016/j.ceramint.2019.03.086 (2019).

Article CAS Google Scholar

Kefeni, K. K., Msagati, T. A. & Mamba, B. B. Ferrite nanoparticles: synthesis, characterisation and applications in electronic device. Mater. Sci. Eng. B 215, 37–55. https://doi.org/10.1016/j.mseb.2016.11.002 (2017).

Article CAS Google Scholar

Amiri, M., Salavati-Niasari, M. & Akbari, A. Magnetic nanocarriers: Evolution of spinel ferrites for medical applications. Adv. Colloid Interface Sci. 265, 29–44. https://doi.org/10.1016/j.cis.2019.01.003 (2019).

Article CAS PubMed Google Scholar

Kefeni, K. K., Mamba, B. B. & Msagati, T. A. Application of spinel ferrite nanoparticles in water and wastewater treatment: A review. Sep. Purif. Technol. 188, 399–422. https://doi.org/10.1016/j.seppur.2017.07.015 (2017).

Article CAS Google Scholar

Fu, F., Dionysiou, D. D. & Liu, H. The use of zero-valent iron for groundwater remediation and wastewater treatment: A review. J. Hazard. Mater. 267, 194–205. https://doi.org/10.1016/j.jhazmat.2013.12.062 (2014).

Article CAS PubMed Google Scholar

Kulkarni, A. B. & Mathad, S. N. Effect of cadmium doping on structural and magnetic studies of Co-Ni ferrites. Sci. Sinter. 53, 407–418. https://doi.org/10.2298/SOS2103407K (2021).

Article CAS Google Scholar

Rao, P. A. et al. Cadmium substitution effect on structural, electrical and magnetic properties of Ni-Zn nano ferrites. Res. Phys. 19, 103487. https://doi.org/10.1016/j.rinp.2020.103487 (2020).

Article Google Scholar

Amin, N. et al. Transport properties of CE-doped Cd ferrites CdFe 2–x Ce x O 4. J. Superconduct. Novel Magn. 34, 2945–2955. https://doi.org/10.1007/s10948-021-06053-z (2021).

Article CAS Google Scholar

Patil, M., Rendale, M., Mathad, S. & Pujar, R. Structural and IR study of Ni 0.5–x Cd x Zn 0.5 Fe 2 O 4. Int. J. Self-Propag. High-Temp. Synth. 24, 241–245. https://doi.org/10.3103/S1061386215040081 (2015).

Nabi, M. A. U. et al. Study of electrical transport properties of cadmium-doped Zn–Mn soft ferrites by co-precipitation method. J. Superconduct. Novel Magn. 34, 1813–1822. https://doi.org/10.1007/s10948-020-05588-x (2021).

Article CAS Google Scholar

Abdolmohammad-Zadeh, H. & Ayazi, Z. Zinc (II)-doped manganese ferrite nanoparticles as an efficient magnetic adsorbent for cadmium extraction from water samples. Anal. Bioanal. Chem. Res. 9, 209–220. https://doi.org/10.22036/ABCR.2021.293274.1652 (2022).

Article CAS Google Scholar

Satheeshkumar, M. et al. Structural and magnetic properties of CuFe2O4 ferrite nanoparticles synthesized by cow urine assisted combustion method. J. Magn. Magn. Mater. 484, 120–125. https://doi.org/10.1016/j.jmmm.2019.03.128 (2019).

Article ADS CAS Google Scholar

Ghazi, N., Chenari, H. M. & Ghodsi, F. Rietveld refinement, morphology analysis, optical and magnetic properties of magnesium-zinc ferrite nanofibers. J. Magn. Magn. Mater. 468, 132–140. https://doi.org/10.1016/j.jmmm.2018.07.084 (2018).

Article ADS CAS Google Scholar

Debnath, S. & Das, R. Cobalt doping on nickel ferrite nanocrystals enhances the micro-structural and magnetic properties: Shows a correlation between them. J. Alloys Compounds 852, 156884. https://doi.org/10.1016/j.jallcom.2020.156884 (2021).

Article CAS Google Scholar

Debnath, S., Das, A. & Das, R. Effect of cobalt doping on structural parameters, cation distribution and magnetic properties of nickel ferrite nanocrystals. Ceram. Int. 47, 16467–16482. https://doi.org/10.1016/j.ceramint.2021.02.095 (2021).

Article CAS Google Scholar

Osagie, C. et al. Dyes adsorption from aqueous media through the nanotechnology: A review. J. Mater. Res. Technol. 14, 2195–2218. https://doi.org/10.1016/j.jmrt.2021.07.085 (2021).

Article CAS Google Scholar

Mozaffari, M. & Masoudi, H. Zinc ferrite nanoparticles: New preparation method and magnetic properties. J. Superconduct. Novel Magn. 27, 2563–2567. https://doi.org/10.1007/s10948-014-2625-x (2014).

Article CAS Google Scholar

Hou, Y., Li, X. Y., Zhao, Q. D., Quan, X. & Chen, G. H. Electrochemical method for synthesis of a ZnFe2O4/TiO2 composite nanotube array modified electrode with enhanced photoelectrochemical activity. Adv. Funct. Mater. 20, 2165–2174. https://doi.org/10.1002/adfm.200902390 (2010).

Article CAS Google Scholar

Singh, A. S., Patil, U. B. & Nagarkar, J. M. Palladium supported on zinc ferrite: A highly active, magnetically separable catalyst for ligand free Suzuki and Heck coupling. Catal. Commun. 35, 11–16. https://doi.org/10.1016/j.catcom.2013.02.003 (2013).

Article CAS Google Scholar

Amin, A. M. M., Rayan, D. A., Ahmed, Y. M. Z., El-Shall, M. S. & Abdelbasir, S. M. Zinc ferrite nanoparticles from industrial waste for Se (IV) elimination from wastewater. J. Environ. Manag. 312, 114956. https://doi.org/10.1016/j.jenvman.2022.114956 (2022).

Article CAS Google Scholar

Mukherjee, K. & Majumder, S. Analyses of response and recovery kinetics of zinc ferrite as hydrogen gas sensor. J. Appl. Phys. 106, 1. https://doi.org/10.1063/1.3225996 (2009).

Article CAS Google Scholar

Tahir, W. et al. Impact of silver substitution on the structural, magnetic, optical, and antibacterial properties of cobalt ferrite. Sci. Rep. 13, 15730. https://doi.org/10.1038/s41598-023-41729-7 (2023).

Article ADS CAS PubMed PubMed Central Google Scholar

Vankdoth, S., Velidandi, A., Sarvepalli, M. & Vangalapati, M. Poly-extract synthesized silver nanoparticles catalysed rhodamine-B and methyl orange dye degradation: Influence of physicochemical parameters and their recyclability. J. Nanoworld 8, 42–54. https://doi.org/10.17756/nwj.2022-099 (2022).

Article CAS Google Scholar

Hosseini, S. M., Hosseini-Monfared, H. & Abbasi, V. Silver ferrite–graphene nanocomposite and its good photocatalytic performance in air and visible light for organic dye removal. Appl. Organometal. Chem. 31, e3589. https://doi.org/10.1002/aoc.3589 (2017).

Article CAS Google Scholar

Khan, M. A. M. et al. Silver-decorated cobalt ferrite nanoparticles anchored onto the graphene sheets as electrode materials for electrochemical and photocatalytic applications. ACS Omega 5, 31076–31084. https://doi.org/10.1021/acsomega.0c04191 (2020).

Article CAS PubMed PubMed Central Google Scholar

Beagan, A. M. Investigating methylene blue removal from aqueous solution by cysteine-functionalized mesoporous silica. J. Chem. 1–12, 2021. https://doi.org/10.1155/2021/8839864 (2021).

Article CAS Google Scholar

Reich, S.-J. et al. Hindered diffusion in ordered mesoporous silicas: insights from pore-scale simulations in physical reconstructions of SBA-15 and KIT-6 silica. J. Phys. Chem. C 122, 12350–12361. https://doi.org/10.1021/acs.jpcc.8b03630 (2018).

Article CAS Google Scholar

Hashem, A. A., Mahmoud, S. A., Geioushy, R. A. & Fouad, O. A. Adsorption of malachite green dye over synthesized calcium silicate nanopowders from waste materials. Mater. Sci. Eng. B 295, 116605. https://doi.org/10.1016/j.mseb.2023.116605 (2023).

Article CAS Google Scholar

Nicola, R. et al. Highly efficient and fast removal of colored pollutants from single and binary systems, using magnetic mesoporous silica. Chemosphere 261, 127737. https://doi.org/10.1016/j.chemosphere.2020.127737 (2020).

Article CAS PubMed Google Scholar

Tatarchuk, T., Bououdina, M., Judith Vijaya, J. & John Kennedy, L. in Nanophysics, Nanomaterials, Interface Studies, and Applications. (eds Olena Fesenko & Leonid Yatsenko) 305–325, https://doi.org/10.1007/978-3-319-56422-7_22 (Springer International Publishing).

Liu, T., Liu, L., Liu, J., Liu, S. & Qiao, S. Z. Fe3O4 encapsulated mesoporous silica nanospheres with tunable size and large void pore. Front. Chem. Sci. Eng. 8, 114–122. https://doi.org/10.1007/s11705-014-1413-2 (2014).

Article ADS CAS Google Scholar

Rivera-Utrilla, J., Bautista-Toledo, I., Ferro-García, M. A. & Moreno-Castilla, C. Activated carbon surface modifications by adsorption of bacteria and their effect on aqueous lead adsorption. J. Chem. Technol. Biotechnol. 76, 1209–1215. https://doi.org/10.1002/jctb.506 (2001).

Article CAS Google Scholar

Khormaei, M., Nasernejad, B., Edrisi, M. & Eslamzadeh, T. Copper biosorption from aqueous solutions by sour orange residue. J. Hazard. Mater. 149, 269–274. https://doi.org/10.1016/j.jhazmat.2007.03.074 (2007).

Article CAS PubMed Google Scholar

Arabkhani, P., Asfaram, A. & Ateia, M. Easy-to-prepare graphene oxide/sodium montmorillonite polymer nanocomposite with enhanced adsorption performance. J. Water Process Eng. 38, 101651. https://doi.org/10.1016/j.jwpe.2020.101651 (2020).

Article Google Scholar

Alijani, H. Q. et al. Bimetallic nickel-ferrite nanorod particles: greener synthesis using rosemary and its biomedical efficiency. 48, 242–251. https://doi.org/10.1080/21691401.2019.1699830 (2020).

Article CAS Google Scholar

Tanveer, M. et al. Sol-gel extended hydrothermal pathway for novel Cd-Zn co-doped Mg-ferrite nano-structures and a systematic study of structural, optical and magnetic properties. J. Magn. Magn. Mater. 553, 169245. https://doi.org/10.1016/j.jmmm.2022.169245 (2022).

Article CAS Google Scholar

Mangood, A. H., Gemeay, A. H., Abdel-Galeil, M. M., Salama, E. S. & El-Shater, R. E. Evaluation of synergistic approach of spinel cadmium–copper nanoferrites as magnetic catalysts for promoting wastewater decontamination: Impact of Ag ions doping. Environ. Sci. Pollut. Res. 1, 1–18. https://doi.org/10.1007/s11356-023-27170-3 (2023).

Article CAS Google Scholar

Ghani, N., Saeed, M. A. & Hashim, I. H. J. Thermoluminescence (TL) response of silica nanoparticles subjected to 50 Gy gamma irradiation. Malays. J. Fund. Appl. Sci. 13, 178–180 (2017).

Google Scholar

Scherrer, P. Determination of the size and internal structure of colloidal particles using X-rays. Nachr. Ges. Wiss. Göttingen 2, 98–100 (1918).

Google Scholar

Abu-Elsaad, N. I. & Abdel Hameed, R. E. Copper ferrite nanoparticles as nutritive supplement for cucumber plants grown under hydroponic system. J. Plant Nutr. 42, 1645–1659. https://doi.org/10.1080/01904167.2019.1630428 (2019).

Article CAS Google Scholar

Ega, C. S., Babu, B. R., Ramesh, K., Sreenivasulu, M. & Purushotham, Y. Correlation between structural, magnetic and dielectric properties of microwave sintered Ni-Zn-Al nanoferrites. J. Superconduct. Novel Magn. 32, 3525–3534. https://doi.org/10.1007/s10948-019-5097-1 (2019).

Article CAS Google Scholar

Abkenar, S. D. et al. Fast removal of methylene blue from aqueous solution using magnetic-modified Fe 3 O 4 nanoparticles. J. Environ. Eng. 141, 04014049. https://doi.org/10.1002/jctb.506 (2015).

Article CAS Google Scholar

Purnawira, B. et al. in IOP Conference Series: Materials Science and Engineering. (IOP Publishing).

El Shater, R. E., El-Desoky, H. S., Meaz, T. M., Kumar, R. & Abdel-Galeil, M. M. Electrochemical investigation of Ag mixed Cd–Cu nanoferrite mixed reduced graphene oxide as improved platform for supercapacitor application. J. Mater. Sci. Mater. Electron. 34, 842. https://doi.org/10.1007/s10854-023-10023-7 (2023).

Article CAS Google Scholar

Zhang, D.-H., Li, H.-B., Li, G.-D. & Chen, J.-S. Magnetically recyclable Ag-ferrite catalysts: General synthesis and support effects in the epoxidation of styrene. Dalton Trans. 1, 10527–10533. https://doi.org/10.1039/B915232F (2009).

Article Google Scholar

Lee, Y. et al. Degradation of polychlorinated dibenzo-p-dioxins and dibenzofurans in real-field soil by an integrated visible-light photocatalysis and solvent migration system with pn heterojunction BiVO4/Bi2O3. J. Hazard. Mater. 344, 1116–1125. https://doi.org/10.1002/jctb.506 (2018).

Article ADS CAS PubMed Google Scholar

Ulu, A., Ozcan, I., Koytepe, S. & Ates, B. Design of epoxy-functionalized Fe3O4@ MCM-41 core–shell nanoparticles for enzyme immobilization. Int. J. Biol. Macromol. 115, 1122–1130. https://doi.org/10.1016/j.ijbiomac.2018.04.157 (2018).

Article CAS PubMed Google Scholar

Aboelfetoh, E. F., Elabedien, M. E. Z. & Ebeid, E.-Z.M. In situ anchoring of iron and zinc oxides nanoparticles onto rice husk cellulose for efficient wastewater remediation. Int. J. Biol. Macromol. 123, 562. https://doi.org/10.1016/j.ijbiomac.2023.123562 (2023).

Article CAS Google Scholar

Bae, J., Kim, S., Kim, K. S., Hwang, H.-K. & Choi, H. Adsorptive removal of arsenic by mesoporous iron oxide in aquatic systems. Water 12, 3147. https://doi.org/10.3390/w12113147 (2020).

Article CAS Google Scholar

Cychosz, K. A. & Thommes, M. Progress in the physisorption characterization of nanoporous gas storage materials. Engineering 4, 559–566. https://doi.org/10.1016/j.eng.2018.06.001 (2018).

Article CAS Google Scholar

Tsai, W.-T. et al. The adsorption of cationic dye from aqueous solution onto acid-activated andesite. J. Hazard. Mater. 147, 1056–1062. https://doi.org/10.1016/j.jhazmat.2007.01.141 (2007).

Article CAS PubMed Google Scholar

Gupta, V. K., Mittal, A. & Gajbe, V. Adsorption and desorption studies of a water soluble dye, Quinoline Yellow, using waste materials. J. Colloid Interface Sci. 284, 89–98. https://doi.org/10.1016/j.jcis.2004.09.055 (2005).

Article ADS CAS PubMed Google Scholar

Belgacem, A., Ould Brahim, I., Belmedani, M. & Hadoun, H. Removal of methyl green dye from aqueous solutions using activated carbon derived from cryogenic crushed waste tires. Iran. J. Chem. Chem. Eng. 41, 207–219 (2022).

CAS Google Scholar

Ayodele, O. et al. Experimental and theoretical studies of the adsorption of Cu and Ni ions from wastewater by hydroxyapatite derived from eggshells. Environ. Nanotechnol. Monit. Manag. 15, 100439. https://doi.org/10.1016/j.enmm.2021.100439 (2021).

Article CAS Google Scholar

Adebowale, K. O., Olu-Owolabi, B. I. & Chigbundu, E. C. Removal of safranin-O from aqueous solution by adsorption onto kaolinite clay. J. Encapsul. Adsorp. Sci. 4, 89. https://doi.org/10.4236/jeas.2014.43010 (2014).

Article Google Scholar

Adam, F. A. et al. Enhanced adsorptive removal of indigo carmine dye by bismuth oxide doped MgO based adsorbents from aqueous solution: Equilibrium, kinetic and computational studies. RSC Adv. 12, 24786–24803. https://doi.org/10.1039/D2RA02636H (2022).

Article ADS CAS PubMed PubMed Central Google Scholar

Palma-Goyes, R. E., Silva-Agredo, J., González, I. & Torres-Palma, R. A. Comparative degradation of indigo carmine by electrochemical oxidation and advanced oxidation processes. Electrochim. Acta 140, 427–433. https://doi.org/10.1016/j.electacta.2014.06.096 (2014).

Article CAS Google Scholar

Yazdi, M. G., Ivanic, M., Mohamed, A. & Uheida, A. Surface modified composite nanofibers for the removal of indigo carmine dye from polluted water. RSC Adv. 8, 24588–24598. https://doi.org/10.1039/C8RA02463D (2018).

Article ADS CAS PubMed PubMed Central Google Scholar

Hendaoui, K., Trabelsi-Ayadi, M. & Ayari, F. Optimization and mechanisms analysis of indigo dye removal using continuous electrocoagulation. Chin. J. Chem. Eng. 29, 242–252. https://doi.org/10.1016/j.cjche.2020.07.065 (2021).

Article CAS Google Scholar

Ahmad, A. et al. Removal of cationic dyes by iron modified silica/polyurethane composite: Kinetic, isotherm and thermodynamic analyses, and regeneration via advanced oxidation process. Polymers 14, 1. https://doi.org/10.3390/polym14245416 (2022).

Article CAS Google Scholar

Aboelfetoh, E. F., Elhelaly, A. A. & Gemeay, A. H. Synergistic effect of Cu (II) in the one-pot synthesis of reduced graphene oxide (rGO/CuxO) nanohybrids as adsorbents for cationic and anionic dyes. J. Environ. Chem. Eng. 6, 623–634. https://doi.org/10.1016/j.jece.2017.12.047 (2018).

Article CAS Google Scholar

dos Reis, G. S. et al. Preparation of highly porous nitrogen-doped biochar derived from birch tree wastes with superior dye removal performance. Colloids Surf. A Physicochem. Eng. Aspects 669, 131493. https://doi.org/10.1016/j.colsurfa.2023.131493 (2023).

Article CAS Google Scholar

Jing, H.-Y. et al. Efficient adsorption/photodegradation of organic pollutants from aqueous systems using Cu 2 O nanocrystals as a novel integrated photocatalytic adsorbent. J. Mater. Chem. A 2, 14563–14570. https://doi.org/10.1039/C4TA02459A (2014).

Article CAS Google Scholar

Liu, L. et al. Simultaneous removal of cationic and anionic dyes from environmental water using montmorillonite-pillared graphene oxide. J. Chem. Eng. Data 60, 1270–1278. https://doi.org/10.1021/je5009312 (2015).

Article ADS CAS Google Scholar

Zhang, P. & Wang, L. Extended Langmuir equation for correlating multilayer adsorption equilibrium data. Sep. Purif. Technol. 70, 367–371. https://doi.org/10.1016/j.seppur.2009.10.007 (2010).

Article CAS Google Scholar

Benderdouche, N., Bestani, B. & Hamzaoui, M. The use of linear and nonlinear methods for adsorption isotherm optimization of basic green 4-dye onto sawdust-based activated carbon. J. Mater. Environ. Sci. 9, 1110–1118. https://doi.org/10.26872/jmes.2018.9.4.122 (2018).

Article CAS Google Scholar

Freundlich, H. Über die adsorption in lösungen. Z. Phys. Chem. 57, 385–470. https://doi.org/10.1515/zpch-1907-5723 (1907).

Article CAS Google Scholar

Dotto, G. L. et al. Selective adsorption of gadolinium from real leachate using a natural bentonite clay. J. Environ. Chem. Eng. 11, 109748. https://doi.org/10.1016/j.jece.2023.109748 (2023).

Article CAS Google Scholar

Naiya, T. K., Bhattacharya, A. K. & Das, S. K. Adsorption of Cd (II) and Pb (II) from aqueous solutions on activated alumina. J. Colloid Interface Sci. 333, 14–26. https://doi.org/10.1016/j.jcis.2009.01.003 (2009).

Article ADS CAS PubMed Google Scholar

Aljeboree, A. M., Alshirifi, A. N. & Alkaim, A. F. Kinetics and equilibrium study for the adsorption of textile dyes on coconut shell activated carbon. Arab. J. Chem. 10, S3381–S3393. https://doi.org/10.1016/j.arabjc.2014.01.020 (2017).

Article CAS Google Scholar

Bayramoglu, G., Akbulut, A., Liman, G. & Arica, M. Y. Removal of metal complexed azo dyes from aqueous solution using tris (2-aminoethyl) amine ligand modified magnetic p (GMA-EGDMA) cationic resin: Adsorption, isotherm and kinetic studies. Chem. Eng. Res. Des. 124, 85–97. https://doi.org/10.1016/j.cherd.2017.06.005 (2017).

Article CAS Google Scholar

Gupta, T. et al. Adsorption of Indigo Carmine Dye by Acacia nilotica sawdust activated carbon in fixed bed column. 12, 15522. https://doi.org/10.1038/s41598-022-19595-6 (2022).

Article CAS Google Scholar

Banerjee, P., Barman, S. R., Mukhopadhayay, A. & Das, P. Ultrasound assisted mixed azo dye adsorption by chitosan–graphene oxide nanocomposite. Chem. Eng. Res. Des. 117, 43–56. https://doi.org/10.1016/j.cherd.2016.10.009 (2017).

Article CAS Google Scholar

Odogu, A. N., Daouda, K., Desiré, B. B. P., Nsami, N. J. & Mbadcam, K. J. Removal of indigo carmine dye (ic) by batch adsorption method onto dried cola nut shells and its active carbon from aqueous medium. Int. J. Eng. Sci. Res. Tech. 5, 874–887 (2016).

CAS Google Scholar

Ramesh, T., Kirana, D., Mohana Kumari, T. & Ashwini, A. J. Adsorption studies of indigo caramine dye by magnesium oxide. Int. J. Sci. Res. 1, 1 (2013).

Google Scholar

Alahiane, S. et al. Adsorption et photodégradation du colorant indigo carmine en milieu aqueux en présence de TiO2/UV/O2 (Adsorption and photocatalytic degradation of indigo carmine dye in aqueous solutions using TiO2/UV/O2). J. Mater. Environ. Sci. 4, 239 (2013).

CAS Google Scholar

Belaib, F. & Meniai, A.-H. The removal of cationic dye (Methyl green) dye by adsorbant based Silica gel/Polymer.

Sharma, P., Saikia, B. K. & Das, M. R. Removal of methyl green dye molecule from aqueous system using reduced graphene oxide as an efficient adsorbent: Kinetics, isotherm and thermodynamic parameters. Colloids Surf. A Physicochem. Eng. Aspects 457, 125–133. https://doi.org/10.1016/j.colsurfa.2014.05.054 (2014).

Article CAS Google Scholar

Dos Reis, L. G. T., Robaina, N. F., Pacheco, W. F. & Cassella, R. J. Separation of Malachite Green and Methyl Green cationic dyes from aqueous medium by adsorption on Amberlite XAD-2 and XAD-4 resins using sodium dodecylsulfate as carrier. Chem. Eng. J. 171, 532–540. https://doi.org/10.1016/j.cej.2011.04.024 (2011).

Article CAS Google Scholar

Rytwo, G., Tropp, D. & Serban, C. Adsorption of diquat, paraquat and methyl green on sepiolite: Experimental results and model calculations. Appl. Clay Sci. 20, 273–282. https://doi.org/10.1016/S0169-1317(01)00068-0 (2002).

Article CAS Google Scholar

Lima, E. C., Hosseini-Bandegharaei, A., Moreno-Piraján, J. C. & Anastopoulos, I. A critical review of the estimation of the thermodynamic parameters on adsorption equilibria: Wrong use of equilibrium constant in the Vant Hoof equation for calculation of thermodynamic parameters of adsorption. J. Mol. Liquids 273, 425–434. https://doi.org/10.1016/j.molliq.2018.10.048 (2019).

Article CAS Google Scholar

Goswami, M. & Phukan, P. Enhanced adsorption of cationic dyes using sulfonic acid modified activated carbon. J. Environ. Chem. Eng. 5, 3508–3517. https://doi.org/10.1016/j.jece.2017.07.016 (2017).

Article CAS Google Scholar

Shao, Y. et al. Application of Mn/MCM-41 as an adsorbent to remove methyl blue from aqueous solution. J. Colloid Interface Sci. 429, 25–33. https://doi.org/10.1016/j.jcis.2014.05.004 (2014).

Article ADS CAS PubMed Google Scholar

Minitha, C., Lalitha, M., Jeyachandran, Y. & Senthilkumar, L. Adsorption behaviour of reduced graphene oxide towards cationic and anionic dyes: Co-action of electrostatic and π–π interactions. Mater. Chem. Phys. 194, 243–252. https://doi.org/10.1016/j.matchemphys.2017.03.048 (2017).

Article CAS Google Scholar

Chowdhury, S. & Balasubramanian, R. Recent advances in the use of graphene-family nanoadsorbents for removal of toxic pollutants from wastewater. Adv. Colloid Interface Sci. 204, 35–56. https://doi.org/10.1016/j.cis.2013.12.005 (2014).

Article CAS PubMed Google Scholar

Tiotsop Kuete, I.-H. et al. Adsorption of indigo carmine onto chemically activated carbons derived from the cameroonian agricultural waste garcinia cola nut shells and desorption studies. https://doi.org/10.1155/2022/1236621 (2022).

Article Google Scholar

Dastgerdi, Z. H., Meshkat, S. S. & Esrafili, M. D Enhanced adsorptive removal of Indigo carmine dye performance by functionalized carbon nanotubes based adsorbents from aqueous solution: Equilibrium, kinetic, and DFT study. J. Nanostruct. Chem. 9, 323–334. https://doi.org/10.1007/s40097-019-00321-0 (2019).

Article CAS Google Scholar

Satlaoui, Y., Trifi, M., Fkih Romdhane, D., Charef, A. & Azouzi, R. J. Removal properties, mechanisms, and performance of methyl green from aqueous solution using raw and purified sejnane clay type. https://doi.org/10.1155/2019/4121864 (2019).

Dalvand, A. et al. Modeling of Reactive Blue 19 azo dye removal from colored textile wastewater using L-arginine-functionalized Fe3O4 nanoparticles: Optimization, reusability, kinetic and equilibrium studies. J. Magn. Magn. Mater. 404, 179–189. https://doi.org/10.1016/j.jmmm.2015.12.040 (2016).

Article CAS Google Scholar

Download references

Open access funding provided by The Science, Technology & Innovation Funding Authority (STDF) in cooperation with The Egyptian Knowledge Bank (EKB).

Chemistry Department, Faculty of Science, Tanta University, Tanta, 31527, Egypt

Ehab A. Okba, Moamen F. Rabea, Mohamed Y. El-Sheikh & Eman F. Aboelfetoh

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

Ehab A. Okba: Conceptualization, Methodology, Visualization, Validation, Investigation, Writing, review and editing, supervision. Moamen F. Rabea: Conceptualization, Methodology, Visualization, Validation, Investigation, Writing original draft, Mohamed. Y. El-Sheikh: Conceptualization, Visualization, Validation, Investigation, review and editing, supervision. Eman F. Aboelfetoh: Conceptualization, Visualization, Validation, Investigation, review and editing, supervision. All authors agree for publication.

Correspondence to Ehab A. Okba.

The authors declare no competing interests.

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

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

Okba, E.A., Rabea, M.F., El-Sheikh, M.Y. et al. Design of silver-zinc-nickel spinel-ferrite mesoporous silica as a powerful and simply separable adsorbent for some textile dye removal. Sci Rep 14, 16481 (2024). https://doi.org/10.1038/s41598-024-66457-4

Download citation

Received: 17 November 2023

Accepted: 01 July 2024

Published: 17 July 2024

DOI: https://doi.org/10.1038/s41598-024-66457-4

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative