Next Article in Journal
Targeting Cancer Hallmarks with Epigallocatechin Gallate (EGCG): Mechanistic Basis and Therapeutic Targets
Previous Article in Journal
Structural Characterization and Immunoenhancing Properties of Polysaccharide CPTM-P1 from Taxus media
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Magnetite/MXene (Fe3O4/Ti3C2) Nanocomposite as a Novel Adsorbent for Environmental Remediation of Malachite Green Dye

by
Amal M. Alkhudaydi
1,2,
Ekram Y. Danish
1 and
Mohamed Abdel Salam
1,*
1
Department of Chemistry, Faculty of Science, King Abdulaziz University, P.O. Box 80200, Jeddah 21589, Saudi Arabia
2
Department of Chemistry, Faculty of Science, Taif University, P.O. Box 11099, Taif 21944, Saudi Arabia
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(6), 1372; https://doi.org/10.3390/molecules29061372
Submission received: 26 February 2024 / Revised: 13 March 2024 / Accepted: 15 March 2024 / Published: 19 March 2024

Abstract

:
In this work, a novel adsorbent called magnetite/MXene (Fe3O4/Ti3C2) nanocomposite was prepared, characterized, and applied for the removal of organic dye, malachite green dye (MG), from both real water and model solutions. Numerous techniques were used to characterize the prepared Fe3O4/Ti3C2 nanocomposite: XRD, SEM, TEM, FTIR, and surface area analysis. The outcomes showed that the Al layer had been selectively etched, that the MAX phase (Ti3AlC2) had been transformed into layered Ti3C2 MXene, that the cubic Fe3O4 phase had been prepared, and that the prepared Fe3O4 NPs had been evenly distributed on the MXene surface. Also, SEM pictures showed the successful etching of the MAX phase and the formation of the ultrathin multi-layered MXene, which the Fe3O4 NPs covered upon forming the Fe3O4/Ti3C2 nanocomposite at the surface and inside the ultrathin multi-layered MXene. The effect of different operational parameters affecting the removal process was explored and optimized. The MG dye was removed mostly within 60 min, with a 4.68 mg/g removal capacity using 5 mg of the Fe3O4/Ti3C2 nanocomposite. The removal was examined from both kinetic and thermodynamic perspectives, and the findings demonstrated the spontaneity of the removal process as well as the applicability of fractal-like pseudo-first-order and fractal-like pseudo-second-order kinetics when compared to other kinetics models. The Fe3O4/Ti3C2 nanocomposite was used to remove MG dye from real spiked environmental water samples, and the results revealed the successful remediation of the real samples from the organic dye by the Fe3O4/Ti3C2 nanocomposite. Accordingly, Fe3O4/Ti3C2 nanocomposite could be considered a potential adsorbent for the environmental remediation of polluted water.

1. Introduction

Most consumer items, including textiles, paper, leather, food, and cosmetics, are colored using synthetic organic dyes that, when combined with other auxiliary chemicals, form large discharges that threaten aquatic life and humans with poisonous, carcinogenic, and mutagenic effects, in addition to their great susceptibility to chelate metal ions, increasing their toxicity to fish and other creatures [1]. Additionally, synthetic organic dyes in water bodies absorb light and prevent light penetration, decreasing photosynthetic activity and preventing the growth of aquatic vegetation [2].
The removal of organic dyes can be accomplished by utilizing a variety of treatment methods, including ion exchange, precipitation, coagulation–flocculation, filtration, liquid–liquid extraction, electrochemical methods, and adsorption [3], especially with carbon-based materials for adsorption and separation [4,5]. Adsorption is considered one of the most efficient removal techniques as it has numerous advantages, including a simple design, reusability of adsorbents/adsorbate, low cost, ease of operation, fewer chemical requirements, high efficiency, and a short time of removal [6].
One of the great challenges facing the application of the adsorption process is the search for new adsorbents with high efficacy. Therefore, exploring novel adsorbents is still a prodigious task for the scientific community. Nothing excites the community of material scientists more than the discovery of a new family of two-dimensional (2D) nanomaterials, such as transition metal dichalcogenides (e.g., MoS2 and WS2), black phosphorous layered double hydroxides (LDHs), g-C3N4, boron nitride, graphene, graphene oxide, metal–organic frameworks (MOFs), and MXenes, which are characterized by a wide range of potential compositions and tailorable features for varied uses [7], especially in applied environmental cleaning [8]. Moreover, the layered 2D materials can easily attach to the nano-sized encapsulated particles through strict layer spaces for various applications [9].
MXene is the most recent member of the 2D nanomaterials family; due to their strong negative surface charge, high chemical stability, adjustable chemistry, and high hydrophilicity, MXenes can be employed in a wide range of applications, such as absorbents, energy storage, supercapacitors, catalysis, gas sensors, polymer-based composites, electromagnetic interference (EMI) shielding antibiosis, and lubrication [10]. Many surface modification techniques have been cleverly developed recently to increase the potential of MXenes in various sectors, including surface-initiated polymerization and single-heteroatom doping [11]. MXenes were recently applied in dye removal [12,13]; for instance, they were used for eliminating methylene blue dye from water [14] and were also used for the removal of malachite green dyes [15]. MXene nanosheets were directly deposited on commercially available filters for fast and efficient dye removal, and the results indicated outstanding performance [16]. This demonstrates the possibility of using MXenes in the adsorption and environmental remediation of dyes from water. Recently, MXene modification was explored to enhance extraordinary properties [17,18,19]. Most recently, MXene/carbon foam hybrid aerogel (MCF) was synthesized and used for methylene blue (MB) and Congo red (CR) [20], MXene/NiFeMn-layered double hydroxide decorated gelatin for the removal of Congo red [21], and porous heterostructured MXene/biomass-activated carbon composites were synthesized and used for the removal of three anionic azo dyes: allure red, Congo red, and sunset yellow [22]. In the recent past, the field of environmental remediation has paid considerable attention to magnetite nanoparticles (Fe3O4 NPs) due to their tendency to aggregate and form clusters [23]. Magnetite nanoparticles (Fe3O4) are among the most promising transition metal oxides that are used for the modification of various adsorbents [24,25,26] because of their advantages of having excellent theoretical unique capacity, affordability, and nontoxicity.
The novelty of the current research is to develop a unique, efficient, and reliable solid hybrid adsorbent based on magnetite and MXene for the environmental remediation of polluted water with organic dyes such as malachite green dye.
Accordingly, in this study, magnetite nanoparticles were used for the modification of MXene to form the magnetite/MXene (Fe3O4/Ti3C2) nanocomposite and used to remove malachite green (MG) dye, as an example of synthetic organic dye, from a simulated dye aqueous solution. The adsorption process was investigated based on the operational factors adsorbent mass, interaction time, temperature, solution pH, and ionic strength, and the mechanism was explored using several adsorption kinetics models, in addition to thermodynamics.

2. Results and Discussion

2.1. Characterization

The XRD patterns of the Ti3AlC2 MAX, exfoliated Ti3C2 MXene, Fe3O4 nanoparticles, and Fe3O4/Ti3C2 nanocomposite are shown in Figure 1. At 2θ = 39°, the peak attributed to Ti3AlC2′s (1 0 4) peak disappeared after the procedure of HF exfoliation. Additionally, the peak at (0 0 2) for 2D Ti3C2 changed from 2θ = 9.7° to 2θ = 9.1°, indicating that the Ti3AlC2 phase transformed into layered Ti3C2 MXene and that the Al layer was carefully etched. There are four diffraction peaks for the generated Fe3O4, at 30.2°, 35.5°, 57.2°, and 62.8°, which belong to the cubic Fe3O4 phase’s (2 2 0), (3 1 1), (5 1 1), and (4 4 0) planes, respectively (JCPDS card no. 19-0629). By applying the Scherrer Formula and utilizing the (3 1 1) peak’s half-peak width, it is expected that the produced Fe3O4 crystallite size is about 27.5 nm. There are no discernible XRD peaks in response to new crystalline phases in the Fe3O4/Ti3C2 nanocomposite; instead, the XRD patterns are composed solely of cubic Fe3O4 and pure Ti3C2. This means that the nanocomposite is a true combination of Ti3C2 MXene and Fe3O4 nanoparticles, which is in line with previously published work [27].
The MAX phase, prepared Ti3C2 MXene, prepared Fe3O4 NPs, and Fe3O4/Ti3C2 nanocomposite microstructures were examined using scanning electron microscopy, and the micrographs are presented in Figure 2. The images of the MAX phase (Ti3AlC2) before (Figure 2a) and following an HF etching procedure that eliminated the Al atoms to produce the exfoliated 2D MXene nanosheets (Ti3C2) (Figure 2b) possessing a few nanometers’ thickness are shown. The MAX phase was characterized by an uneven surface and a closely packed structure, whereas the resulting exfoliated 2D MXene nanosheets were ultrathin and multi-layered, characterized by clear sharp edges, confirming the success of the etching process because of the weak Ti–Al bonds compared to the Ti–C bonds within the MAX phase. Also, due to the extreme reactivity of the Al atoms, they could be easily etched and removed from the Ti3AlC2 structure and form the 2D MXene nanosheets (Ti3C2). Moreover, the ultrathin sheets are parallel to one another, which suggests that the basal planes and Ti3AlC2 etching directions are parallel, which is similar to previous studies [28,29,30]. Figure 2c shows the Fe3O4 NPs, which are characterized by a large quantity of agglomerated nanoparticles with a very small size. Figure 2d displays the uniform dispersion of Fe3O4 nanoparticles both on the exterior and within the incredibly thin multi-layered MXene. The Fe3O4 nanoparticles could be easily inserted between the layers of Ti3C2 MXene or adsorb on their surface.
Further examination of the MAX phase, Ti3C2 MXene, Fe3O4 NPs, and Fe3O4/MXene nanocomposite was performed using TEM. Figure 3a,b show the MAX phase and Ti3C2 MXene nanosheets, which demonstrated overlapped and nearly transparent sheets with a few nanometers’ diameter. It is evident that the diameter of the MXene is smaller than the MAX phase due to the etching process, which decreased the size of MXene compared to the MAX phase: 78 nm and 122 nm, respectively. Additionally, the aggregation of Fe3O4 NPs with an average size of particles of 11 nm was visible in the TEM image, and the homogeneous distribution of the Fe3O4 NPs over the MXene nanosheets is displayed in Figure 3d. FT-IR analyses of the MAX phase, Ti3C2 MXene, Fe3O4 NPs, and Fe3O4/MXene nanocomposite were explored to provide details on the produced nanomaterials’ composition and surface functionality. The FTIR spectrum of the Ti3AlC2 MAX phase is displayed in Figure 4, with pronounced bands at 474, 617, 1087, and 1541 cm−1 corresponding to the TiO2, Ti–O, Al–O, and CH groups, respectively, in addition to the strong vibration band of the –OH group at 3500 cm−1 due to the adsorbed external water. In the case of MXene, the vibrational stretching of the highly hydrogen-bonded –OH group and the bending vibration of the C–OH bond, respectively, could be the cause of the absorption peaks at 1638 and 3686 cm−1 because MXene’s hydrophilic nature causes it to absorb external water molecules or strongly coordinated H2O. Such distinct –OH groups’ stretching vibrational patterns may also be associated with Ti4+ cations. Moreover, Ti–O–Ti linkage and Ti–O bonding deformation vibrations account for the maxima at 877 and 617 cm−1, respectively. Furthermore, the band at 617 cm−1 might be a representation of rocking or twisting Ti–C out-of-plane vibrations [31,32]. The FT-IR spectrum of Fe3O4 nanoparticles showed absorption bands at 550–1650 cm−1 due to Fe–O group flexural vibrations and strong vibration bands of the –OH group at 1638 and 3500 cm−1 due to adsorbed external water. The FTIR spectrum of the Fe3O4/MXene nanocomposite showed most of the characteristic absorption bands of both MXene and the Fe3O4 NPs. The specific surface area values of the samples were calculated from the nitrogen gas adsorption/desorption isotherms at 77 K, and the values were 10.40, 40.25, 59.89, and 32.86 m2/g for the MAX phase, Ti3C2 MXene, Fe3O4 NPs, and Fe3O4/MXene nanocomposite, respectively. There was a significant increase in the MAX phase’s specific surface area following HF etching and the formation of exfoliated 2D nanosheets of MXene, from 10.40 to 40.25 m2/g. Upon the modification of MXene with the Fe3O4 NPs and the formation of the Fe3O4/MXene nanocomposite, this BET surface area was slightly reduced from 40.25 to 32.86 m2/g due to the distribution of the Fe3O4 NPs over the MXene nanosheets, as was presented earlier by the SEM (Figure 2) and TEM (Figure 3) analyses.

2.2. Adsorption Studies

The effects of several operating factors were investigated to ascertain the efficacy of Fe3O4/Ti3C2 nanocomposite solid adsorbent concerning the MG dye’s adsorption under diverse conditions, and they are detailed in the following parts. We investigated the amount of absorption of prepared nanomaterials for MG dye at a concentration of 10 mg/l in deionized water. The highest removal ratio (47.72%) of this dye was achieved by the Fe3O4/Ti3C2 nanocomposite. Figure 5a shows how the dosage results for the Fe3O4/Ti3C2 nanocomposite affect the removal efficiency and its UV–Vis. absorbance spectra (inset). Using 100 mg of Fe3O4/Ti3C2 nanocomposite, the maximum removal capacity was found to be 0.993 mg/g of MG, with a removal efficacy of 99.28%. While utilizing 5 mg of Fe3O4/Ti3C2 nanocomposite, the optimum removal capacity was found to be 4.68 mg/g, with a removal efficacy of 23.38%. MG clearance using Fe3O4/Ti3C2 nanocomposite increased with increasing Fe3O4/Ti3C2 nanocomposite dose, possibly due to the improved accessibility to extra adsorption active sites as a result of the increased adsorbent dose. Furthermore, 30 mg of Fe3O4/Ti3C2 nanocomposite was used for further experiments.
Temperature’s effect on the adsorption of MG utilizing Fe3O4/Ti3C2 nanocomposite was examined, and Figure 5b displays the findings. The figure makes it evident that the endothermic nature of the adsorption process is present because the percentage of MG adsorbed on the Fe3O4/Ti3C2 nanocomposite increases as temperature rises, which will be evaluated thoroughly regarding the thermodynamic parts. This is explained by the fact that more MG molecules were able to obtain enough energy to engage with the Fe3O4/Ti3C2 nanocomposite surface, which increased MG molecule mobility and diffusion. Furthermore, because of active diffusion, a temperature increase would cause pore sizes to grow, adding additional surface area for adsorption [33].
The duration of interaction between the solid adsorbents, Fe3O4/Ti3C2 nanocomposite, the adsorbate, and MG dye molecules is one of the critical operating parameters influencing the removal process. The results of this investigation on the impact of interaction time are shown in Figure 5c. The Fe3O4/Ti3C2 nanocomposite showed extremely rapid adsorption; full saturation happened in less than 1 h, and it had a removal capacity (qt) of (3.14 mg MG each g of Fe3O4/Ti3C2 nanocomposite) because of the increase in active sites associated with increasing the mass of the Fe3O4/Ti3C2 nanocomposite and consequent electrostatic interaction between the negative Fe3O4/Ti3C2 nanocomposite surface and the MG dye increased. Therefore, the amount of MG adsorbed on the Fe3O4/Ti3C2 nanocomposite rises with increasing contact time. An increase in interaction duration is correlated with a small increase in removal efficiency and capacity. Furthermore, a 60 min duration of the interaction between MG dye and Fe3O4/Ti3C2 nanocomposite was used for further experiments.
Any adsorption/removal process that involves controlling surface charges of adsorbents, which changed according to pH value, depends heavily on the solution pH value. Surface charge is affected principally by the solution pH, which modifies the electrostatic conditions between adsorbent and adsorbate. The current results show that the adsorption of MG dye molecules onto Fe3O4/Ti3C2 nanocomposite increases when pH increases, indicating the degree of the surface charge that is negative. The information is consistent with the theory that adsorption is mostly caused by electrostatic attraction. The Fe3O4/Ti3C2 nanocomposite exhibited a noteworthy improvement in MG dye removal efficiency (87.63–99.7%) in Figure 5d when the pH of the solution was raised gradually from 2.0 to 12.0. This could be explained by electrostatic attraction between the MG neutral and cationic dye molecules (pKa value 10.3) and the negatively charged Fe3O4/Ti3C2 nanocomposite surface (zeta potential value of −21 mV). In the meantime, the Fe3O4/Ti3C2 nanocomposite dramatically increased its ability to remove MG dye at a pH of 6.0, 80.88%, which in turn increased more through raising the pH to 12.0 (99.28%). The electrostatic attraction between the positively charged MG molecules and the negatively charged Fe3O4/Ti3C2 nanocomposite surface may be the cause of this phenomenon. This result demonstrates that the Fe3O4/Ti3C2 nanocomposite solid adsorbent’s mechanism for removing the MG molecules is electrostatic in nature. The existence of this electrostatic mechanism was demonstrated and confirmed by investigating the Fe3O4/Ti3C2 nanocomposite’s electrostatic interaction mechanism for the adsorption and removal of MG dye through varying the KNO3 concentration in the investigated solution and performing the removal experiment at different ionic strength values. Figure 5e illustrates how, when the (KNO3) concentration in the solution was increased to 5.0 mM, the percentage of MG dye removed dropped sharply to 76.89%, and further addition of KNO3 slightly retarded the adsorption mechanism; 0.1 M KNO3 was associated with 75.09% removal. The reason for this could be that the Fe3O4/Ti3C2 nanocomposite surface is surrounded by a double layer of positive ions (K+) and negative ions (NO3) of (KNO3), which creates electrostatic repulsion with the MG dye molecules. This discovery supports the mechanism of electrostatic contact for the MG dye that had been removed by Fe3O4/Ti3C2 nanocomposite, which strongly corroborated earlier research that established the organic dyes’ electrostatic mechanism through solid clay-based adsorbents [34].

2.3. Kinetics Study

Information regarding the kinetics interaction between the adsorbent and adsorbate in the solid/solution system provides a comprehensive understanding of the possible interaction mechanism of adsorption in addition to the reaction pathways [35]. Several kinetics models assuming that the kinetic constants are time-invariant (a time-dependent system function that is not a direct function of time), such as pseudo-first-order (PFO) and pseudo-second-order (PSO), have been used for adsorption kinetics prediction at solid/solution interfaces. This assumption has been unacceptable when reactants (adsorbent and adsorbate) were spatially constrained by walls or phase boundaries [36]. Based on experimental and theoretical examinations, it was presented that the reaction rate coefficient is time-dependent (fractal-like kinetics), leading to the fractal-like pseudo-first-order (FL-PFO) and fractal-like pseudo-second-order (FL-PSO) kinetics models.
Therefore, the kinetic adsorption data were evaluated by using four nonlinear models: pseudo-first-order (PFO) [37], pseudo-second-order (PSO) [37], fractal-like pseudo-first-order (FL-PFO) [38], and fractal-like pseudo-second-order (FL-PSO) [38]. The mathematical equations of these respective models are shown in Equations (1)–(4).
q t = q e × [ 1 e k 1 t ]
q t = k 2 × q e 2 × t 1 + q e k 2 × t
q t = q e × [ 1 e ( k 1 , 0 × t ) n ]
q t = k 2 , 0 × q e 2 × t n 1 + q e k 2 , 0 × t n
where t is the interaction time (min); qt, and qe are the amount of MG dye adsorbed at time t and at equilibrium, respectively (mg g−1); k1 is the pseudo-first-order rate constant (min−1); k2 is the pseudo-second-order rate constant (g mg−1 min−1); k1,0 is the fractal-like pseudo-first-order constant rate (min−1), k2,0 is the fractal-like pseudo-second-order rate constant (g mg−1 min−1), and n is the fractional-like exponent (n > 0). The kinetic and equilibrium models’ appropriateness was then examined and evaluated using different statistical error functions: the residual sum of squares (RSS), Equation (5); the determination coefficient (R2), Equation (6); the adjusted determination coefficient (R2adj), Equation (7); the standard deviation of residues (SD), Equation (8); and the Bayesian Information Criterion (BIC), Equation (9) [37,39].
R S S = i n ( q i , e x p q i , m o d e l ) 2
R 2 = ( i n ( q i , e x p q ¯ i , e x p ) 2 i n ( q i , e x p q ¯ i , e m o d e l ) 2 i n ( q i , e x p q ¯ i , e x p ) 2 )
R a d j 2 = 1 ( 1 R 2 ) ( n 1 n p 1 )
S D = ( 1 n p ) i n ( q i , e x p q ¯ i , e m o d e l ) 2
B I C = n L n R S S n + p L n ( n )
In the above equations, qi,model is the individual theoretical q value predicted by the model; qi,exp is individual experimental q value; q ¯ i , e x p is the average of all experimental q values measured; n is the number of experiments; and p is the number of parameters in the fitting model. In this work, the R a d j 2 , SD, and BIC values will be used to examine the appropriateness of different kinetics models, taking into consideration that the best-fitted model would have R a d j 2 value closer to unity, in addition to lower values of SD and BIC. Figure 6 shows the fitting of the four kinetics models with the experimental adsorption capacities (qt, mg/g) variation with interaction time (min), and Table 1 shows the kinetic parameters for each model. The four kinetics models were evaluated statistically based on the R a d j 2 , SD, and BIC values to evaluate the most suitable model. Surprisingly, both the FL-PFO and FL-PSO kinetics models showed the best R a d j 2 values of 0.9944 and 0.9945, and the lowest standard deviation (SD) of residue values of 0.0790 and 0.0784, and the lowest BIC values of −31.67 and −31.78 for MG dye using Fe3O4/Ti3C2 nanocomposite as adsorbent (Table 1), respectively.
Furthermore, the process of adsorption/removal process is controlled/governed by diffusion kinetics too as the adsorption process occurs typically in four separate stages: mass action, bulk diffusion, liquid film diffusion, and intra-particle diffusion, which is sometimes referred to as pore and surface diffusion. The identification of the adsorption/diffusion process’s rate-determining step is the main focus of this research segment; based on the liquid film diffusion model, it predicts the rate processes’ kinetics by assuming that the adsorbate molecules’ slowest-moving stage in the adsorption process is their flow through a liquid film that surrounds the solid adsorbent, and intra-particle diffusion model assumes that the adsorbate molecules’ slowest-moving stage in the adsorption process is their flow through the solid adsorbent pores.
The following equation [40] describes the liquid film diffusion model:
L n 1 F = k f d t
where the film diffusion rate coefficient is denoted by kfd (min−1), and F denotes the fractional attainment of equilibrium (F = qt/qe). Figure 7a demonstrates how the liquid film diffusion model plot of [ Ln (1  F)] against t converged when applied to the MG dye adsorption by Fe3O4/Ti3C2 nanocomposite, and a correlation coefficient of 0.856 was obtained, resulting in a straight line through the origin for the first 3 min. This suggests that the rate-determining step could be for the liquid film diffusion model within the first 3 min.
The following is the equation of the intra-particle diffusion model that Weber and Morris presented in 1963 [41]:
q t = k i d t 1 / 2 + C
C is a constant inversely proportional to the boundary layer thickness (mg/g), and kid is the intra-particle diffusion rate constant (mg/g·min1/2). Upon utilizing the intra-particle diffusion model on the experimental data, the adsorption of the MG dye by the Fe3O4/Ti3C2 nanocomposite was plotted as qt versus t; the plot did not converge satisfactorily considering all the points, and, therefore, it was possible to draw a straight line through the origin, as observed in Figure 7b, obtained following the 3 initial minutes, with a correlation coefficient value of 0.973, indicating that the process that influences the rate may be the intra-particle diffusion model. Accordingly, both liquid film and intra-particle diffusion may have an impact on the removal diffusion kinetics in this work at the same time, as presented in Figure 7c.
Figure 7. (a) Liquid film diffusion kinetics model, (b) intra-particle diffusion kinetics model, and (c) change over time in the quantity of MG dye adsorbed per unit mass of the Fe3O4/Ti3C2 nanocomposite.
Figure 7. (a) Liquid film diffusion kinetics model, (b) intra-particle diffusion kinetics model, and (c) change over time in the quantity of MG dye adsorbed per unit mass of the Fe3O4/Ti3C2 nanocomposite.
Molecules 29 01372 g007

2.4. Environmental Applications

Five separate real water samples were utilized to further confirm the effectiveness of the Fe3O4/Ti3C2 nanocomposite as an adsorbent for the removal of MG dye, and Figure 8 shows the results. After the real water sample collection, analysis was conducted on the MG dye concentrations, and it was discovered that they were below the detection limit; they were regarded as zero. Therefore, to simulate pollution and explore the possibility of using Fe3O4/Ti3C2 nanocomposite as a solid material for water pollution treatment, concentrated MG dye was added to real water samples until a final concentration of 10.0 mg/L was obtained. Afterward, the spiked real water samples were mixed with the Fe3O4/Ti3C2 nanocomposite. Regarding the MG dye, the samples of distilled water, mineral water, well water, wastewater, and seawater had removal percentages of 98.40%, 97.49%, 96.13%, 94.27%, and 93.14%, in that order. It is noteworthy that the minimal percentage of MG dye removal observed in the well water, wastewater, and seawater through the Fe3O4/Ti3C2 solid nanocomposite was mainly caused by the elevated concentrations of Na+, K+, Mg2+, and Ca2+ found in the selected real water samples as compared to the other real water samples. Because of these ions’ screening and shielding properties at such high concentrations, the amount of MG dye removed by the solid Fe3O4/Ti3C2 nanocomposite was significantly reduced. According to these findings, the Fe3O4/Ti3C2 nanocomposite was able to remove the majority of the MG dye from real environmental water samples.

2.5. Reusability Studies

For practical application, a solid adsorbent’s reusability is essential. From environmental and economic points of view, it is a highly crucial component of adsorption. Thus, the Fe3O4/Ti3C2 nanocomposite’s recycling effectiveness for MG dye adsorption was examined during four consecutive cycles. After each adsorption cycle, the MG dye molecules were adsorbed from the Fe3O4/Ti3C2 nanocomposite using a concentrated solution of NaCl; after that, the Fe3O4/Ti3C2 adsorbent was washed with acetone several times, and in the last step it was washed with deionized water and a centrifuge was used to separate the Fe3O4/Ti3C2 nanocomposite and then dried at room temperature before being used in another absorption cycle.
We discovered that, even after the four cycles with a high removal efficiency, the proportion of MG that was removed using the Fe3O4/Ti3C2 nanocomposite had somewhat decreased (Figure 9). It is clear from this that the Fe3O4/Ti3C2 nanocomposite is a reusable adsorbent for environmental cleanup.

2.6. Studies on Thermodynamics of the Removal

The three essential thermodynamic functions are entropy change (ΔS), enthalpy change (ΔH), and Gibbs free energy change (ΔG), which must be explored and measured to estimate thermodynamic viability and spontaneity using any physical or chemical procedure, such as using a solid adsorbent Fe3O4/Ti3C2 nanocomposite to remove MG dye from an aqueous solution. To determine the thermodynamic parameters, the following formulas were employed [42]:
K d = q t C t × 1000
l n K d = Δ S R Δ H R T
ΔG = ΔH – TΔS
where Kd, the dimensionless thermodynamic distribution coefficient, is calculated at equilibrium concerning the amount of MG dye removed (mg of MG dye for each gram of Fe3O4/Ti3C2 nanocomposite), and the equilibrium concentration of MG dye in the solution is represented by Ct (mg/L). Figure 10 shows that the plot of ln Kd vs. 1/T revealed a straight line, and the ΔH and ΔS were calculated using the straight-line slope and intercept, in that order. Fe3O4/Ti3C2 nanocomposite’s thermodynamic characteristics for MG dye elimination were calculated, where the corresponding values for ∆H, ∆S, and ∆G were +49.07 kJ/mole, +234.29 J/mole.K, and −20.75 kJ/mole (at 298 K). There was a positive value obtained for the entropy change, demonstrating an increase in the degrees of freedom at the boundary between liquid and solid due to the MG dye attaching itself to the Fe3O4/Ti3C2 nanocomposite surface, promoting the removal process’ spontaneity, whereas the enthalpy change had a positive value, indicating the endothermic nature of the removal process, which unfavored the removal process. As a result, a negative change in the free energy would be anticipated for a process of spontaneous elimination. Conversely, the removal of MG dye from aqueous solution by Fe3O4/Ti3C2 nanocomposite is an entropy-driven process, as indicated by the positive values of ∆H, ∆S, and the negative value of ∆G [43].
Based on the above results, the removal/adsorption mechanism of the MG dye from aqueous solution by Fe3O4/Ti3C2 nanocomposite was found to be electrostatic in nature due to the attraction forces arising between the MG neutral and cationic dye molecules to the negatively charged Fe3O4/Ti3C2 nanocomposite surface, which were confirmed by studying the effect of the solution pH and ionic strength. Also, the removal process was well-described by both the FL-PFO and FL-PSO kinetics models and controlled by both liquid film and intra-particle diffusion kinetics. In addition, the removal process was entropy-driven, as indicated by the positive values of ∆H, ∆S, and the negative value of ∆G.
Based on the above-mentioned results, the maximum removal capacity of MG dye (10.0 mg/L), obtained using 0.5 g/L loading of Fe3O4/Ti3C2 nanocomposite (5 mg/10 mL) within 60.0 min, was found to be 4.68 mg/g, which is very competitive with other removal methods based on removal capacity and treatment time, as presented in Table 2.

3. Experimental

3.1. Substances

All studies were conducted using chemicals of analytical grade: MAX phase powder of Ti3AlC2 (NANOSHEL, Dera Bassi, India, purity: 99%); 48.051% hydrofluoric acid (HF, Parmcac Applichem IT Reagents, Darmstadt, Germany); polyethylene glycol (PEG 4000, BOC Sciences, New York, NY, USA); and 99.5% acetone (Abuljadayel & Sons Co., Jeddah, Saudi Arabia).

3.2. Fe3O4 Nanoparticles and MXene Nanosheet Synthesis

Powdered Ti3AlC2 (>98 weight percent pure) was offered commercially by Forsman Com., Ltd. (Shanghai, China) to produce Fe3O4 and Ti3C2 MXene nanoparticles. In order to remove the Al layers completely, 2 g of the Ti3AlC2 powder was stirred for 24 h at 60 °C while submerged in HF solutions (20 mL; >40 wt%). After removing the HF, the solution was centrifuged and repeatedly washed with deionized water until the pH reached 5–6. The generated Ti3C2 powdered form was dried under a vacuum at 80 °C for a whole day. The process of hydrothermal synthesis was used to create Fe3O4 nanoparticles. First, under vigorous stirring conditions for 30 min, 0.43 g of NaOH, 1.35 g of FeCl3.6H2O, and 10 g of polyethylene glycol (PEG 4000, BOC Sciences, New York, NY, USA) were dissolved in 40 mL of ethylene glycol. The solution was then instantly put into a 100 mL Teflon-lined autoclave and heated at 200 °C for 8 h. In order to get rid of the contaminant, the sediment was repeatedly cleaned with ethanol and distilled water. The Fe3O4 nanoparticles that were produced were then dried in a vacuum for 6 h at 80 °C.

3.3. Synthesis of Fe3O4/Ti3C2 Nanocomposite

For 30 min, deionized water was used to ultrasonically disperse 100 mg of Ti3C2 and 40 mg of Fe3O4 nanoparticles in 80 mL and 20 mL, respectively. Both were then combined, and they underwent 6 h ultra-sonification therapy after that. The Ti3C2/Fe3O4 nanocomposite was then obtained by filtering the suspension. The finished nanocomposite was dried in a vacuum for 24 h at 80 °C.

3.4. Characterizations

The crystalline structures of all the samples were analyzed using powder X-ray diffraction (XRD, Cu K radiation, Bruker D2 Phaser X-ray diffractometer) with a step scan of 0.02° per step. Scanning electron microscopy (SEM, Gemini, Zeiss-Ultra 55, Chiyoda, Tokyo) and transmission electron microscopy (TEM, Sirion 200FEI, Eindhoven, The Netherlands) were used to examine the microstructure and morphology of produced samples. The FT-IR spectra of the samples were recorded using an FT-IR spectrophotometer (Spectrum 100, Perkin Elmer, Waltham, MA, USA). NOVA3200e (Quantachrome, Boynton Beach, FL, USA) was utilized to determine the BET specific surface area of the materials by utilizing the nitrogen adsorption/desorption isotherm at 77 K.

3.5. Adsorption Experiment

First, the prepared nanomaterials studied the adsorption of 10 mL of 10 mg/L of malachite green dye for a certain period. After the experiment began, at a wavelength of 617 nm, the SHIMADZU CPS-240A UV–Vis (Kyoto, Japan) was employed to measure the quantity of adsorbed MG dye. The effects of experimental parameters such as Fe3O4/Ti3C2 nanocomposite mass, solution pH, initial MG dye concentration, solution temperature, and ionic strength were examined to estimate the efficacy of the nanocomposite adsorbent toward the adsorption of MG dye under various conditions in model and real water samples. The efficiency of the Fe3O4/Ti3C2 nanocomposite towards the removal/adsorption of MG dye was calculated using Equation (15), and the removal capacity (qt, mg/g) was estimated using Equation (16).
%   a d s o r p t i o n = C o C t C o
q t = V C o C t m
where V is the volume of the solution (L), m is the dose (g) of the Fe3O4/Ti3C2 nanocomposite, and C0 and Ct are the concentrations of the MG dye in the solution at time zero and any time thereafter.

3.6. Collecting Real Water Samples

Several real water samples were gathered to confirm the Fe3O4/Ti3C2 nanocomposite’s suitability as a solid adsorbent for the environmental cleanup of contaminated real water using MG dye. A wastewater sample was obtained from King Abdelaziz University (KAUWW), Jeddah City’s Membrane Bio-Reactor Technology Wastewater Treatment Plant (MBR 6000 STP) (Latitude deg. North 21.487954, Longitude deg. East 39.236748). At Abdraboh AL-Beladi Farm in the Khulais Province of Saudi Arabia, a deep well (latitude deg. North 22.124376, longitude deg. East 39.500618) was used to gather the well water sample. A sample of distilled water from a distillation unit from King Abdelaziz University’s laboratories also included a Row mineral water sample (commercial water in bottles) and a sample from the Red Sea. A 45.0 µm Millipore filter paper was used to filter all real water samples, which were then stored in Teflon® bottles at 5 °C in the dark.

4. Conclusions

This study explored the preparation of the modified magnetite/MXene nanocomposite and its application for the removal/adsorption of malachite green dye from model solutions and environmental real water samples. The prepared Fe3O4/Ti3C2 nanocomposite was physically and chemically characterized using XRD, SEM, TEM, FT-IR, and surface area analysis, and the results demonstrated the successful preparation of MXene and magnetite, in addition to the homogenous distribution of Fe3O4 over the MXene surface. The removal/adsorption was explored, and the results revealed maximum removal achieved using 100 mg of the Fe3O4/Ti3C2 nanocomposite within 60 min under ambient conditions, with an optimum removal capacity of 4.676 mg/g using 5 mg only. Also, removal was studied kinetically, and the results revealed that both the fractal-like pseudo-first-order and fractal-like pseudo-second-order kinetics models could better explain the removal of MG dye by the Fe3O4/Ti3C2 nanocomposite. Also, the Fe3O4/Ti3C2nanocomposite was used for the MG dye removal/adsorption from spiked real environmental water samples collected from different places, and the results demonstrated extraordinary efficacy of the Fe3O4/Ti3C2 nanocomposite and the prospective application of the Fe3O4/Ti3C2 nanocomposite for the environmental remediation of polluted water. The recycling and reuse of the Fe3O4/Ti3C2 nanocomposite for the successive removal of the MG dye was studied, and the results showed the possibility of reusing the Fe3O4/Ti3C2 nanocomposite for four consecutive times with high efficiency. The thermodynamics of the removal/adsorption process were also explored and the results demonstrated the spontaneity of the process: a negative Gibbs free energy change value of −20.75 kJ/mole, a positive entropy change value of +234.29 J/mole.K, and a positive enthalpy change value of +49.07 kJ/mole, indicating that the removal/adsorption of the MG dye by the Fe3O4/Ti3C2nanocomposite is an entropy-driven process.

Author Contributions

Methodology, A.M.A. and M.A.S.; Software, E.Y.D.; Validation, A.M.A. and E.Y.D.; Formal analysis, A.M.A.; Resources, M.A.S.; Data curation, A.M.A.; Writing—original draft, A.M.A.; Writing—review & editing, E.Y.D. and M.A.S.; Supervision, E.Y.D. and M.A.S.; Funding acquisition, M.A.S. All authors have read and agreed to the published version of the manuscript.

Funding

The Deanship of Scientific Research (DSR) at King Abdulaziz University (KAU), Jeddah, Saudi Arabia has funded this project, under grant no. (KEP-PhD: 92-130-1443).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Al-Tohamy, R.; Ali, S.S.; Li, F.; Okasha, K.M.; Mahmoud, Y.A.-G.; Elsamahy, T.; Jiao, H.; Fu, Y.; Sun, J. A critical review on the treatment of dye-containing wastewater: Ecotoxicological and health concerns of textile dyes and possible remediation approaches for environmental safety. Ecotoxicol. Environ. Saf. 2022, 231, 113160. [Google Scholar] [CrossRef] [PubMed]
  2. Jiang, F.; Dinh, D.M.; Hsieh, Y.-L. Adsorption, and desorption of cationic malachite green dye on cellulose nanofibril aerogels. Carbohydr. Polym. 2017, 173, 286–294. [Google Scholar] [CrossRef] [PubMed]
  3. Chouli, F.; Ezzat, A.O.; Sabantina, L.; Benyoucef, A.; Zehhaf, A. Optimization Conditions of Malachite Green Adsorption onto Almond Shell Carbon Waste Using Process Design. Molecules 2024, 29, 54. [Google Scholar] [CrossRef] [PubMed]
  4. Nayl, A.A.; Abd-Elhamid, A.I.; Arafa, W.A.A.; Ahmed, I.M.; El-Shanshory, A.A.; Abu-Saied, M.A.; Soliman, H.M.A.; Abdelgawad, M.A.; Ali, H.M.; Bräse, S. Chitosan-Functionalized-Graphene Oxide (GO@CS) Beads as an Effective Adsorbent to Remove Cationic Dye from Wastewater. Polymers 2022, 14, 4236. [Google Scholar] [CrossRef] [PubMed]
  5. Zhang, H.; Zhang, T.; Ding, S.; Wang, X. Development of loose thin film nanofibrous composite nanofiltration membrane with modified g-C3N4 nanosheets barrier layer for efficient separation of salt/dye mixtures. Sep. Purif. Technol. 2023, 306 Pt A, 122661. [Google Scholar] [CrossRef]
  6. Akpomie, K.G.; Conradie, J. Advances in application of cotton-based adsorbents for heavy metals trapping, surface modifications and future perspectives. Ecotoxicol. Environ. Saf. 2020, 201, 110825. [Google Scholar] [CrossRef]
  7. He, H.; Wang, X.; Yu, Q.; Wu, W.; Feng, X.; Kong, D.; Ren, X.; Gao, J. In Situ Growth of Ti3C2/UiO-66-NH2 Composites for Photoreduction of Cr (VI). Catalysts 2023, 13, 876. [Google Scholar] [CrossRef]
  8. Yang, G.; Li, S.; Wang, X.; Ding, B.; Li, Y.; Lin, H.; Tang, D.; Ren, X.; Wang, Q.; Luo, S.; et al. A universal strategy boosting photoelectrochemical water oxidation by utilizing MXene nanosheets as hole transfer mediators. Appl. Catal. B Environ. 2021, 297, 120268. [Google Scholar] [CrossRef]
  9. Naguib, M.; Kurtoglu, M.; Presser, V.; Lu, J.; Niu, J.; Heon, M.; Hultman, L.; Gogotsi, Y.; Barsoum, M.W. Two-Dimensional Nanocrystals Produced by Exfoliation of Ti3AlC2. Adv. Mater. 2011, 23, 4248–4253. [Google Scholar] [CrossRef]
  10. Wu, Z.; Shen, J.; Li, C.; Zhang, C.; Wu, C.; Li, Z.; An, X.; He, L. Niche Applications of MXene Materials in Photothermal Catalysis. Chemistry 2023, 5, 492–510. [Google Scholar] [CrossRef]
  11. Wu, Z.; Shen, J.; Li, C.; Zhang, C.; Feng, K.; Wang, Z.; Wang, X.; Meira, D.M.; Cai, M.; Zhang, D.; et al. Mo2TiC2 MXene-Supported Ru Clusters for Efficient Photothermal Reverse Water–Gas Shift. ACS Nano 2023, 17, 1550–1559. [Google Scholar] [CrossRef] [PubMed]
  12. Chen, S.-y.; Deng, Y.-f.; Huang, T.; Zhang, N.; Wang, Y. Polydopamine-assisted MXene decoration on electrospun polylactide fibers toward oil/water separation and organic dye adsorption. Sep. Purif. Technol. 2024, 328, 125040. [Google Scholar] [CrossRef]
  13. Yang, F.; Li, J.; Dong, J.; Chen, S.; Hu, W.; Zhang, Y.; Wang, H.; Li, Z.; Wang, Z. MX@MIL-125(Ti)-mediated sonocatalytic degradation for the dyes and microplastics. Sep. Purif. Technol. 2024, 337, 126488. [Google Scholar] [CrossRef]
  14. Tran, N.M.; Ta, Q.T.H.; Sreedhar, A.; Noh, J.-S. Ti3C2Tx MXene playing as a strong methylene blue adsorbent in wastewater. Appl. Surf. Sci. 2021, 537, 148006. [Google Scholar] [CrossRef]
  15. Katheresan, V.; Kansedo, J.; Lau, S.Y. Efficiency of various recent wastewater dye removal methods: A review. J. Environ. Chem. Eng. 2018, 6, 4676–4697. [Google Scholar] [CrossRef]
  16. Zhang, S.; Liao, S.; Qi, F.; Liu, R.; Xiao, T.; Hu, J.; Li, K.; Wang, R.; Min, Y. Direct deposition of two-dimensional MXene nanosheets on commercially available filter for fast and efficient dye removal. J. Hazard. Mater. 2020, 384, 121367. [Google Scholar] [CrossRef] [PubMed]
  17. Wu, Z.; Li, C.; Li, Z.; Feng, K.; Cai, M.; Zhang, D.; Wang, S.; Chu, M.; Zhang, C.; Shen, J.; et al. Niobium and Titanium Carbides (MXenes) as Superior Photothermal Supports for CO2 Photocatalysis. ACS Nano 2021, 15, 5696–5705. [Google Scholar] [CrossRef] [PubMed]
  18. Tawalbeh, M.; Mohammed, S.; Al-Othman, A.; Yusuf, M.; Mofijur, M.; Kamyab, H. MXenes and MXene-based materials for removal of pharmaceutical compounds from wastewater: Critical review. Environ. Res. 2023, 228, 115919. [Google Scholar] [CrossRef]
  19. Gopalram, K.; Kapoor, A.; Kumar, P.S.; Sunil, A.; Rangasamy, G. MXenes and MXene-Based Materials for Removal and Detection of Water Contaminants: A Review. Ind. Eng. Chem. Res. 2023, 62, 6559–6583. [Google Scholar] [CrossRef]
  20. Wang, X.; Xu, Q.; Zhang, L.; Pei, L.; Xue, H.; Li, Z. Adsorption of methylene blue and Congo red from aqueous solution on 3D MXene/carbon foam hybrid aerogels: A study by experimental and statistical physics modeling. J. Environ. Chem. Eng. 2023, 11, 109206. [Google Scholar] [CrossRef]
  21. Elgarhy, G.S.; El-Subruiti, G.M.; Omer, A.M.; Eltaweil, A.S. 2D/3D MXene/NiFeMn-layered double hydroxide decorated gelatin for removal of Cr (VI) and Congo red: Performance and mechanism. J. Mol. Liq. 2024, 396, 123889. [Google Scholar] [CrossRef]
  22. Xue, H.; Gao, X.; Seliem, M.K.; Mobarak, M.; Dong, R.; Wang, X.; Fu, K.; Li, Q.; Li, Z. Efficient adsorption of anionic azo dyes on porous heterostructured MXene/biomass activated carbon composites: Experiments, characterization, and theoretical analysis via advanced statistical physics models. Chem. Eng. J. 2023, 451, 138735. [Google Scholar] [CrossRef]
  23. Liu, D.; Li, T.; Sun, W.; Zhou, W.; Zhang, G. Magnetic Ti3C2 MXene Nanomaterials for Doxorubicin Adsorption from Aqueous Solutions: Kinetic, Isotherms, and Thermodynamic Studies. ACS Omega 2022, 7 Pt 3, 31945–31953. [Google Scholar] [CrossRef] [PubMed]
  24. Sharifi, M.J.; Nouralishahi, A.; Hallajisani, A. Fe3O4-chitosan nanocomposite as a magnetic biosorbent for removal of nickel and cobalt heavy metals from polluted water. Int. J. Biol. Macromol. 2023, 248, 125984. [Google Scholar] [CrossRef] [PubMed]
  25. Elgamal, A.M.; El-Ghany, N.A.A.; Saad, G.R. Synthesis and characterization of hydrogel-based magnetite nanocomposite adsorbents for the potential removal of Acid Orange 10 dye and Cr (VI) ions from aqueous solution. Int. J. Biol. Macromol. 2023, 227, 27–44. [Google Scholar] [CrossRef] [PubMed]
  26. Zawrah, M.F.; El-Gammal, M.I.; Salem, M.; El-Sonbati, M.A.; Ahmed, M. Recycling of Rice Husk for Preparation of Activated Carbon/Magnetite Nanocomposites for Removal of Methylene Blue from Wastewater. Int. J. Environ. Res. 2023, 17, 51. [Google Scholar] [CrossRef]
  27. Wang, Y.; Li, Y.; Qiu, Z.; Wu, X.; Zhou, P.; Zhou, T.; Zhao, J.; Miao, Z.; Zhou, J.; Zhuo, S. Fe3O4@Ti3C2 MXene hybrids with ultrahigh volumetric capacity as an anode material for lithium-ion batteries. J. Mater. Chem. A 2018, 6, 11189–11197. [Google Scholar] [CrossRef]
  28. Yu, H.; Wang, Y.; Jing, Y.; Ma, J.; Du, C.-F.; Yan, Q. Surface Modified MXene-Based Nanocomposites for Electrochemical Energy Conversion and Storage. Small 2019, 15, 1901503. [Google Scholar] [CrossRef]
  29. Rakhi, B.; Ahmed, B.; Hedhili, M.N.; Anjum, D.H.; Alshareef, H.N. Effect of Postetch Annealing Gas Composition on the Structural and Electrochemical Properties of Ti2CTx MXene Electrodes for Supercapacitor Applications. Chem. Mater. 2015, 27, 5314–5323. [Google Scholar] [CrossRef]
  30. Scheibe, B.; Kupka, V.; Peplińska, B.; Jarek, M.; Tadyszak, K. The Influence of Oxygen Concentration during MAX Phases (Ti3AlC2) Preparation on the α-Al2O3 Microparticles Content and Specific Surface Area of Multilayered MXenes (Ti3C2Tx). Materials 2019, 12, 353. [Google Scholar] [CrossRef]
  31. Mahmood, M.; Rasheed, A.; Ayman, I.; Rasheed, T.; Munir, S.; Ajmal, S.; Agboola, P.O.; Warsi, M.F.; Shahid, M. Synthesis of ultrathin MnO2 nanowire-intercalated 2D-MXenes for high-performance hybrid supercapacitors. Energy Fuels 2021, 35, 3469–3478. [Google Scholar] [CrossRef]
  32. Kiran, N.U.; Deore, A.B.; More, M.A.; Late, D.J.; Rout, C.S.; Mane, P.; Chakraborty, B.; Besra, L.; Chatterjee, S. Comparative Study of Cold Electron Emission from 2D Ti3C2TX MXene Nanosheets with Respect to Its Precursor Ti3SiC2 MAX Phase. ACS Appl. Electron. Mater. 2022, 4, 2656–2666. [Google Scholar] [CrossRef]
  33. Woo, S.-Y.; Lee, H.-S.; Kim, J.-S.; Kim, K.-H.; Ji, H.; Kim, Y.-D. Applicability assessment of functional adsorption zeolite materials in adsorption desalination cum cooling systems driven by low-grade heat source. Chem. Eng. J. 2022, 430, 131375. [Google Scholar] [CrossRef]
  34. Zhang, T.; Wang, W.; Zhao, Y.; Bai, H.; Wen, T.; Kang, S.; Song, G.; Song, S.; Komarneni, S. Removal of heavy metals and dyes by clay-based adsorbents: From natural clays to 1D and 2D nanocomposites. Chem. Eng. J. 2021, 420, 127574. [Google Scholar] [CrossRef]
  35. Gupta, S.S.; Bhattacharyya, K.G. Kinetics of adsorption of metal ions on inorganic materials: A review. Adv. Colloid Interface Sci. 2011, 162, 39–58. [Google Scholar] [CrossRef] [PubMed]
  36. Kopelman, R. Fractal Reaction Kinetics. Science 1988, 241, 1620–1626. [Google Scholar] [CrossRef] [PubMed]
  37. Lima, E.C.; Dehghani, M.H.; Guleria, A.; Sher, F.; Karri, R.R.; Dotto, G.L.; Tran, H.N. Adsorption: Fundamental aspects and applications of adsorption for effluent treatment. In Green Technologies for the Defluoridation of Water; Dehghani, M.H., Karri, R., Lima, E.C., Eds.; Elsevier: Amsterdam, The Netherlands, 2021; pp. 41–88. [Google Scholar] [CrossRef]
  38. Haerifar, M.; Azizian, S. Fractal-like kinetics for adsorption on heterogeneous solid surfaces. J. Phys. Chem. C 2014, 118, 1129–1134. [Google Scholar] [CrossRef]
  39. 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 Van’t Hoof equation for calculation of thermodynamic parameters of adsorption. J. Mol. Liq. 2019, 273, 425–434. [Google Scholar] [CrossRef]
  40. Yao, C.; Chen, T. A film-diffusion-based adsorption kinetic equation and its application. Chem. Eng. Res. Des. 2017, 119, 87–92. [Google Scholar] [CrossRef]
  41. An, B. Cu (II) and As (V) Adsorption Kinetic Characteristic of the Multifunctional Amino Groups in Chitosan. Processes 2020, 8, 1194. [Google Scholar] [CrossRef]
  42. Owija, N.Y.; Salam, S.K.M.A. Removal of cadmium ions from aqueous solution by Zero valent iron nanoparticles: Equilibrium and thermodynamic studies. J. Mol. Liq. 2021, 342, 117462. [Google Scholar] [CrossRef]
  43. Li, C.; Kong, D.; Yao, X.; Ma, X.; Wei, C.; Wang, H. Adsorption Characteristics and Molecular Simulation of Malachite Green onto Modified Distillers’ Grains. Water 2022, 14, 171. [Google Scholar] [CrossRef]
  44. Gupta, V.; Mittal, A.; Krishnan, L.; Gajbe, V. Adsorption kinetics and column operations for the removal and recovery of malachite green from wastewater using bottom ash. Sep. Purif. Technol. 2004, 40, 87–96. [Google Scholar] [CrossRef]
  45. Chen, J.; Mao, J.; Mo, X.; Hang, J.; Yang, M. Study of adsorption behavior of malachite green on polyethylene glycol micelles in cloud point extraction procedure. Colloids Surfaces A Physicochem. Eng. Asp. 2009, 345, 231–236. [Google Scholar] [CrossRef]
  46. Khan, A.A.; Ahmad, R.; Khan, A.; Mondal, P.K. Preparation of unsaturated polyester Ce (IV) phosphate by plastic waste bottles and its application for removal of malachite green dye from water samples. Arab. J. Chem. 2013, 6, 361–368. [Google Scholar] [CrossRef]
  47. Kaith, B.S.; Jindal, R.; Sharma, R. Synthesis of a Gum rosin alcohol-poly(acrylamide) based adsorbent and its application in removal of malachite green dye from waste water. RSC Adv. 2015, 5, 43092–43104. [Google Scholar] [CrossRef]
  48. Kaith, B.S.; Sukriti Sharma, J.; Kaur, T.; Sethi, S.; Shanker, U.; Jassal, V. Microwave-assisted green synthesis of hybrid nanocomposite: Removal of malachite green from waste water. Iran. Polym. J. 2016, 25, 787–797. [Google Scholar] [CrossRef]
  49. Khattri, S.; Singh, M. Removal of malachite green from dye wastewater using neem sawdust by adsorption. J. Hazard. Mater. 2009, 167, 1089–1094. [Google Scholar] [CrossRef]
  50. Abou-Gamra, Z.M.; Ahmed, M.A. TiO2 nanoparticles for removal of malachite green dye from waste water. Adv. Chem. Eng. Sci. 2015, 5, 373–388. [Google Scholar] [CrossRef]
  51. Lv, G.-Y.; Cheng, J.-H.; Chen, X.-Y.; Zhang, Z.-F.; Fan, L.-F. Biological decolorization of malachite green by Deinococcus radiodurans R1. Bioresour. Technol. 2013, 144, 275–280. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of MAX phase (Ti3AlC2), MXene (Ti3C2), magnetite (Fe3O4), and Fe3O4/Ti3C2 nanocomposite.
Figure 1. XRD patterns of MAX phase (Ti3AlC2), MXene (Ti3C2), magnetite (Fe3O4), and Fe3O4/Ti3C2 nanocomposite.
Molecules 29 01372 g001
Figure 2. SEM micrographs of (a) MAX phase, (b) MXene, (c) magnetite, and (d) Fe3O4/Ti3C2 nanocomposite.
Figure 2. SEM micrographs of (a) MAX phase, (b) MXene, (c) magnetite, and (d) Fe3O4/Ti3C2 nanocomposite.
Molecules 29 01372 g002
Figure 3. TEM micrographs of (a) MAX phase, (b) MXene, (c) magnetite, and (d) Fe3O4/Ti3C2 nanocomposite.
Figure 3. TEM micrographs of (a) MAX phase, (b) MXene, (c) magnetite, and (d) Fe3O4/Ti3C2 nanocomposite.
Molecules 29 01372 g003
Figure 4. FTIR spectra of MXene, magnetite NPs, and Fe3O4/Ti3C2 nanocomposite.
Figure 4. FTIR spectra of MXene, magnetite NPs, and Fe3O4/Ti3C2 nanocomposite.
Molecules 29 01372 g004
Figure 5. (a) Effect of Fe3O4/Ti3C2 nanocomposite dose on the MG removal; inset of the UV–Vis. spectra of the removal process of MG dye by different doses from Fe3O4/Ti3C2 nanocomposite; (b) effect of solution temperature on the removal of MG by Fe3O4/Ti3C2 nanocomposite; (c) impact of interaction time on MG elimination by Fe3O4/Ti3C2 nanocomposite removal efficiency; (d) influence of pH of the solution on the Fe3O4/Ti3C2 nanocomposite’s efficiency to remove MG, and inset of the Zeta potential value of Fe3O4/Ti3C2 nanocomposite; and (e) impact of solution ionic strength on removal of MG dye using Fe3O4/Ti3C2 nanocomposite.
Figure 5. (a) Effect of Fe3O4/Ti3C2 nanocomposite dose on the MG removal; inset of the UV–Vis. spectra of the removal process of MG dye by different doses from Fe3O4/Ti3C2 nanocomposite; (b) effect of solution temperature on the removal of MG by Fe3O4/Ti3C2 nanocomposite; (c) impact of interaction time on MG elimination by Fe3O4/Ti3C2 nanocomposite removal efficiency; (d) influence of pH of the solution on the Fe3O4/Ti3C2 nanocomposite’s efficiency to remove MG, and inset of the Zeta potential value of Fe3O4/Ti3C2 nanocomposite; and (e) impact of solution ionic strength on removal of MG dye using Fe3O4/Ti3C2 nanocomposite.
Molecules 29 01372 g005
Figure 6. Kinetic curves for the adsorption removal of MG dye by Fe3O4/Ti3C2 nanocomposite adsorbent (conditions for the experiment: 10 mL, pH of 6.5, 30.0 mg Fe3O4/Ti3C2 nanocomposite, and 10.0 mg/L of MG dye).
Figure 6. Kinetic curves for the adsorption removal of MG dye by Fe3O4/Ti3C2 nanocomposite adsorbent (conditions for the experiment: 10 mL, pH of 6.5, 30.0 mg Fe3O4/Ti3C2 nanocomposite, and 10.0 mg/L of MG dye).
Molecules 29 01372 g006
Figure 8. Fe3O4/Ti3C2 nanocomposite used to remove MG from water in the environment (conditions for the experiment: 10 mL solution, pH 8, 120 min, 50 mg Fe3O4/Ti3C2 nanocomposite, and 10 mg/L of MG dye).
Figure 8. Fe3O4/Ti3C2 nanocomposite used to remove MG from water in the environment (conditions for the experiment: 10 mL solution, pH 8, 120 min, 50 mg Fe3O4/Ti3C2 nanocomposite, and 10 mg/L of MG dye).
Molecules 29 01372 g008
Figure 9. Reusability of the adsorbent Fe3O4/Ti3C2 nanocomposite in the extraction of MG dye from the solution (conditions for the experiment: 10 mL, pH 8, 120 min, 50 mg Fe3O4/Ti3C2 nanocomposite, and 10 mg/L of MG dye).
Figure 9. Reusability of the adsorbent Fe3O4/Ti3C2 nanocomposite in the extraction of MG dye from the solution (conditions for the experiment: 10 mL, pH 8, 120 min, 50 mg Fe3O4/Ti3C2 nanocomposite, and 10 mg/L of MG dye).
Molecules 29 01372 g009
Figure 10. The calculations of thermodynamic parameters for the adsorption of MG dye from a model solution of Fe3O4/Ti3C2 nanocomposite.
Figure 10. The calculations of thermodynamic parameters for the adsorption of MG dye from a model solution of Fe3O4/Ti3C2 nanocomposite.
Molecules 29 01372 g010
Table 1. Parameters of the kinetics models for the removal of MG dye from model solutions using Fe3O4/Ti3C2 nanocomposite.
Table 1. Parameters of the kinetics models for the removal of MG dye from model solutions using Fe3O4/Ti3C2 nanocomposite.
Pseudo-First-Order Kinetics Model (PFO)
qe,exp (mg g−1)qe,cal (mg g−1)k1 (min−1)R2 R a d j 2 SDRSSBIC
3.1402.7761.0800.95730.94880.23920.2861−16.544
Pseudo-second-order kinetics model (PSO)
qe,exp (mg g−1)qe,cal (mg g−1)k2 (g mg−1 min−1)R2 R a d j 2 SDRSSBIC
3.1402.9170.59650.98130.97750.15840.1254−22.315
Fractal-like Pseudo-first-order Kinetics model (FL-PFO)
qe,exp (mg g−1)qe,cal (mg g−1)k1,0 (min−1)R2 R a d j 2 SDRSSBIC
3.1404.4460.04880.99630.99440.07900.0249−31.668
Fractal-like Pseudo-second-order Kinetics model (FL-PSO)
qe,exp (mg g−1)qe,cal (mg g−1)k2,0 (g mg−1 min−1)R2 R a d j 2 SDRSSBIC
3.1405.3580.10980.99630.99450.07840.0246−31.783
Table 2. Removal capacity of MG dye by various removal techniques.
Table 2. Removal capacity of MG dye by various removal techniques.
MaterialRemoval Capacity (mg/g)Removal Time Reference
Bottom ash0.7183 min[44]
Polyethylene glycol micelles1.0010 min[45]
Unsaturated Polyester Ce(IV) phosphate1.0135 min[46]
Fum rosin alcoholpoly(acrylamide)1.40300 min[47]
Hybrid nanocomposite3.2128 h[48]
Neem sawdust4.3514 min[49]
Fe3O4/Ti3C2 nanocomposite4.6860 min[Current study]
TiO2 nanoparticles6.2540 min[50]
Deinococcus radiodurans7.6330 min[51]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Alkhudaydi, A.M.; Danish, E.Y.; Abdel Salam, M. Magnetite/MXene (Fe3O4/Ti3C2) Nanocomposite as a Novel Adsorbent for Environmental Remediation of Malachite Green Dye. Molecules 2024, 29, 1372. https://doi.org/10.3390/molecules29061372

AMA Style

Alkhudaydi AM, Danish EY, Abdel Salam M. Magnetite/MXene (Fe3O4/Ti3C2) Nanocomposite as a Novel Adsorbent for Environmental Remediation of Malachite Green Dye. Molecules. 2024; 29(6):1372. https://doi.org/10.3390/molecules29061372

Chicago/Turabian Style

Alkhudaydi, Amal M., Ekram Y. Danish, and Mohamed Abdel Salam. 2024. "Magnetite/MXene (Fe3O4/Ti3C2) Nanocomposite as a Novel Adsorbent for Environmental Remediation of Malachite Green Dye" Molecules 29, no. 6: 1372. https://doi.org/10.3390/molecules29061372

Article Metrics

Back to TopTop