Next Article in Journal
Kinetic Understanding of the Enhanced Electroreduction of Nitrate to Ammonia for Co3O4–Modified Cu2+1O Nanowire Electrocatalyst
Previous Article in Journal
Electrodeposited Co Crystalline Islands Shelled with Facile Spontaneously Deposited Pt for Improved Oxygen Reduction
Previous Article in Special Issue
Analysis of CeO2-Supported Ru Catalysts for an Efficient Catalytic Wet Air Oxidation of Reconstituted Tobacco Wastewater
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Antimony- and Bismuth-Based Ionic Liquids as Efficient Adsorbents for the Removal of Dyes

1
Department of Chemistry, Quaid-i-Azam University, Islamabad 45320, Pakistan
2
Department of Applied Chemistry, Government College University, Faisalabad 38000, Pakistan
3
Department of Energy Technology, Tallinn University of Technology, 19086 Tallinn, Estonia
4
Department of Chemistry, Allama Iqbal Open University, Islamabad 44000, Pakistan
5
Department of Physics, College of Science, Qassim University, Buraydah 51452, Saudi Arabia
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Catalysts 2025, 15(5), 492; https://doi.org/10.3390/catal15050492
Submission received: 2 April 2025 / Revised: 7 May 2025 / Accepted: 14 May 2025 / Published: 19 May 2025

Abstract

:
A series of ionic liquids consisting of anilinium cations with varying alkyl chains and metallic (Sb and Bi) halides as anions have been synthesized and thoroughly characterized by using multinuclear (1H and 13C) NMR, FT-IR, Raman and XPS techniques. They have been exploited as adsorbents for the dye’s removal, such as malachite green, rhodamine B and Sudan II, from the aqueous solution. Various parameters like the effect of stirring rate, pH, reaction time, adsorbent amount and initial dye concentration have been optimized. Both antimony- and bismuth-based ionic liquids exhibit high adsorption efficiencies and have comparable performance for each dye. Kinetic data have been analyzed by applying kinetic models, and the best-fitted model was found to be pseudo-second order with an R2 value greater than 0.98. Adsorption capacity has been determined by analyzing the sorption data using the Langmuir and Freundlich equations, and the Langmuir isotherm model has been found to be the best fitting. The maximum adsorption capacities (qmax) derived from the Langmuir isotherm for malachite green, Sudan II and rhodamine B by M-Sb ILs were 217.36, 162.10 and 62.94 mg·g−1, whereas by M-Bi ILs, the adsorption capacities were slightly higher, at 230.18, 170.00 and 64.21 mg·g−1, respectively. Kinetic studies indicated pseudo-second-order behavior (R2 > 0.98), while thermodynamic analysis demonstrated an endothermic adsorption, and a spontaneous reaction was carried out by a physisorption process. These findings accentuate the potential of Sb- and Bi-based ionic liquids as efficient and reusable adsorbents for removing dyes from wastewater.

1. Introduction

Generally, organic dyes are essential chemical compounds that are significantly used as the coloring agents in various industrial sectors, including textiles, paper, printing, cosmetics, leather, plastics, and pharmaceuticals [1]. Thousands of different types of pigments and dyes are being used globally, and approximately 700 kilotons of dyes are produced in industries annually [2]. One kilogram of textile consumes nearly 200 L of water, while a standard textile mill requires approximately 1.6 million liters of water per day [3]. So, the wastewater contains a huge amount of dye pollutants that are a major threat not only to the biosphere but can also bring harmful impacts to both human [4] and aquatic life by reducing the transmission of sunlight [5,6].
Dye pollutants frequently contain hazardous aromatic substances and heavy metals [7] (e.g., lead, cadmium, chromium, etc.) that lead to the carcinogenicity, mutagenicity and malfunctioning of the brain, kidney, liver [8], central nervous system and reproductive system of human beings [9]. So, developing an economical and efficient approach to treating these contaminant organic dyes from the soil and water is a dire necessity. Numerous chemical, physical and biological methods have been used in the past [10], but due to the stable and resistant nature of dyes under several conditions, such as heat, light, and oxidizing agents, the adsorption technique has been adopted because it is highly proficient, sustainable and economic for wastewater treatment [11,12].
Numerous adsorbents like zeolites [13], nanomaterials [14], activated carbon [15], biosorbents [16] and miscellaneous materials [17] have been frequently used for removing organic dyes [18]; however, due to significant limitations and challenges, including hazardous by-products [19], agglomeration [20], lower efficiency [21], small surface area [21] and clearance of sludge [22], there is an utter requirement for an economical substitute with enhanced adsorbing and recyclability to remove organic dyes/pollutants from wastewater [23]. Ionic liquids (ILs) have emerged as promising materials for wastewater treatment, offering exceptional properties like non-volatility, less toxicity, a broad range of constancy, low liquefying temperature, a tendency to adsorb a variety of materials and eco-friendliness that have enabled them to be potential contenders as adsorbents for the removal of dyes from wastewater [24]. ILs are salts typically composed of different organic cations with inorganic or organic anions and can be altered with relative ease by a variety of suitable anion and cation combinations; thus, they can be designed for task-specific purposes [25,26].
Metal-substituted ionic liquids, also written as halometallate ionic liquids, have the structural building block [MX4]2− (M = Fe, Bi, Sb, Pd, Zn, Co, Cd, Ni, Cu; X = Cl, Br, I) and are a safer alternative to common volatile organic solvents [27,28] because they are thermally [29,30] and chemically stable [31,32], have limited (if any) volatility [33] and have strong catalytic activity [34]. Metallic salts, complexes and organometallic compounds can easily dissolve in ILs, providing a polar medium for a catalytic reaction [35,36]. Metal incorporation in ILs increases the hydrophobic character [37], which is estimated to be effective in regeneration [38], separation [39] and reuse after the catalysis [40].
In previous studies, phosphonium- and imidazolium-based ILs have been effectively used for dye adsorption from aqueous media. However, traditional ILs and adsorbents face significant limitations, including poor recyclability, decreased adsorption proficiency owing to aggregation and complications in separation after adsorption. Moreover, conventional adsorbents, such as zeolites and activated carbon, lack selectivity for specific dye pollutants and may possibly produce secondary waste during disposal processes [41,42]. So, considering these facts and in continuation of our prior work ([40,43,44] and [45]), four novel antimony- and bismuth-containing anilinium-based hydrophobic ILs have been synthesized and used as adsorbents in the extraction and sorption of three organic dyes (malachite green (MG), rhodamine B (RhB) and Sudan II). These synthesized ILs offer enhanced hydrophobicity, which improves their separation efficiency and recyclability after adsorption. Additionally, their strong interactions with various organic dyes significantly improve the performance of dye removal compared to conventional adsorbents and traditional ILs. They have been characterized comprehensively, including by UV–Visible spectroscopy, and their adsorption performance has been assessed and substantiated with detailed kinetic and thermodynamic analyses.

2. Results and Discussion

2.1. 1H NMR

The 1HNMR spectrum for (M-IL) shows the singlet peak at 3.6 ppm with an intensity of 9, representing the protons from three methyl groups, and the 1HNMR for (E-IL) shows the peak for –CH2 at 1.18–1.21 ppm and for the methyl group –CH3 at 3.36 ppm; thus, this confirms the alkylation of the nitrogen atom. The signal was initially at 3.06 ppm with an intensity of 6 in N,N-dimethylaniline. The downfield shifting of the methyl peak is the result of the quaternization of nitrogen, which withdraws the electron density toward itself due to positive charge formation. The aromatic H shifts occur at 6.96 and 7.27 ppm for the aromatic hydrogen in N,N-dimethylaniline. The formation of a fourth bond to the nitrogen atom gives a deficiency of electron density, for which nitrogen withdraws electron density toward itself from neighboring atoms, due to which the aromatic hydrogen shifts downfield to 8.01 and 7.60 ppm, which provides evidence for the formation of iodide-based ILs (Figures S2 and S3).

2.2. 13C NMR

The 13CNMR spectra for (M-IL) and (E-IL) have different carbon environments in DMSO. The methyl carbon in (M-IL) shows a chemical shift most upfield, around 56.89 ppm, which was originally at 41.3 ppm in N,N-dimethylaniline. Similarly, the signals at 9.34 and 55.0 ppm for the -CH3 and -CH2 groups, respectively, are observed for (E-IL). The shifting of the signal to downfield confirms the formation of the N-C bond. The quaternization of nitrogen causes the de-shielding of methyl carbon, thus shifting it from 40 to 56.89 ppm. While the aromatic carbon located on the benzene ring adjacent to the nitrogen shows the shift at 151 ppm, this could be explained based on the shift of electron density from methyl groups to the benzene ring through the nitrogen atom. The other aromatic carbons show the signal at 120.9 ppm and 130.5 ppm, respectively (Figures S4 and S5).

2.3. Raman Spectra

1H and 13C NMR were used to confirm the intermediate structures, and they can indicate the proton and carbon environments. However, in an anion exchange step, the cations are not changing; it is the anions, which possess no carbon or hydrogen atom. Therefore, repeating the NMR step would not give any more useful information. Synthesized ionic liquids were characterized by Raman spectroscopy for the confirmation of metal chlorides. The Raman spectra of antimony- and bismuth-containing ionic liquids are shown in Figure 1. The Raman bands in the range of 0–500 cm−1 are primarily designated to internal modes of metal chlorides. The Raman bands in the range of 220 cm−1 are assigned to the angular distortion mode of Cl-Bi-Cl, whereas the Bi-Cl symmetric stretch was observed at 256 cm−1, and the asymmetric stretch mode was observed at 296 cm−1, respectively. Similarly, the Raman spectrum of Sb-ILs shows a band at 231 cm−1 for Cl-Sb-Cl angular deformations. The band in the range of 270–300 cm−1 is attributed to the symmetric stretching of Sb-Cl, while the band at 344–356 cm−1 is attributed to the asymmetric vibration of [SbCl4]. These values correspond very well with the literature values [46,47].

2.4. X-Ray Photoelectron Spectroscopy

The X-ray photoelectron spectroscopic (XPS) analysis was carried out to investigate the oxidation state, composition and chemical environment of synthesized ILs. The survey spectra in Figure 2a and Figure 3a authenticated the existence of N, C, Cl and corresponding metals (Sb and Bi) in them. The peak fitting of Sb 3d spectra (Figure 2b) clearly consists of doublet peaks at binding energy values of 530.5 and 539.8 eV, with a spin–orbit separation of 9.3 eV. However, the overlapped signal Sb 3d shows the existence of O 1s at 530 eV. There is no significant shift in binding energies, which signifies the occurrence of Sb+3, and the values align well with previous studies [48]. Meanwhile, high-resolution spectra of Bi 4f in Figure 3b also illustrate two peaks at the binding energies of 160.4 and 165.7 eV, with a spin–orbit separation of 5.3 eV. No significant fluctuations in peak values indicated that Bi is in the +3 oxidation state, and this is well in agreement with the literature [49]. The high-resolution spectra for chlorine demonstrated the spin–orbit doublets at 2p core levels as 2p3/2 and 2p1/2, as shown in Figure 2 and Figure 3c. Similarly, the carbon core level spectrum can be fitted by three peaks (lower to higher BE values) that are denoted as Caromatic (remaining five carbons within the aromatic head group), sp2 C-N (phenyl carbon with nitrogen atom) and sp3 C-N (sp3 hybridized carbons attached with nitrogen outside the ring), respectively, as displayed in Figure 2 and Figure 3d. Moreover, the peak at 401.5 eV can be assigned to the N 1s of aniline (R-N+), as represented in Figure 2 and Figure 3e; however, an additional signal was observed in the case of M-Sb at 402.5 of C=NH+, possibly due to a side reaction [50].

2.5. Adsorption Studies for MG, RhB and Sudan II Dyes

To check the adsorption efficiency of prepared ILs, three different dyes were used: MG, RhB and Sudan II, using a 10 mgL−1 dye concentration, a time of 40 min, 0.01 g of adsorbent dose and a 100 rpm stirring rate. It is observed that the Bi-ILs show better efficiency than the corresponding Sb-ILs, and the ethyl chain-containing ILs have better performance than the methyl chain-ILs, as shown in Figure 4.

2.6. Optimization of Parameters for Adsorption Process

Parameters like stirring rate, pH, time, dose of adsorbents and dye concentration have a great effect on adsorption efficiencies. The effects of these parameters are studied in detail and optimized to obtain the maximum adsorption efficiencies. The adsorption behavior of ILs on these dyes and their optimization parameters are illustrated in Figure 5.

2.7. Stirring Rate Effect

The effect of stirring rate was checked out by varying the stirring rate to 100, 150, 200, 250 and 300 rpm in a solution of 25 mgL−1 dye concentration for a time of 40 min and with an adsorbent dose of 0.01 g. Figure 5a illustrates the effect of stirring rate on the adsorption efficiencies of malachite green, Sudan II and rhodamine B. The adsorption efficiencies were enhanced with a varying stirring rate. It was found that with increasing the stirring rate, the adsorption was quicker than with slow stirring, which could be explained by the fact that fast stirring keeps the dye particles suspended and allows more contact of solid–liquid, favoring the transfer of dye molecules to the adsorbent material [51]. However, the stirring speed was optimized at 200 rpm because a higher stirring rate causes the splashing of solution, which may result in loss of material, and the faster stirring rate did not increase adsorption significantly [52]. To run the experiment smoothly, the stirring rate was set at an intermediate rate of 200 rpm. The maximum adsorption efficiencies for MG were found to be about 73% for bismuth-based ionic liquids (M-Bi) and around 55% for antimony-based ionic liquids (M-Sb), which were higher than Sudan II, followed by rhodamine B, respectively.

2.8. Effect of Time

The time effect on the percentage removal of dyes was studied by taking an initial dye concentration of 25 mgL−1, a 0.01 g adsorption dose and a 200 rpm stirring rate. The UV spectra were taken at constant intervals of time to study the constant effect. The removal of dye was quicker at earlier points of contact times and gradually declined until equilibrium was reached. A high rate of adsorption was observed in the initial contact time, corresponding to the presence of a greater number of adsorption sites on the synthesized materials. The optimum time required for the maximum adsorption of maximum dyes with a 25 mgL−1 solution was found to be 40 min for both types of bismuth- and antimony-based ionic liquids, whereas the sorption efficiency did not increase significantly on further increasing the time [53]. The faster and higher adsorption capacities for bismuth-based ionic liquids are due to the larger size of bismuth, providing more adsorption sites. Figure 5b indicates the actual variation of adsorption with changing time. Malachite green was removed 10 times faster than rhodamine B with M-Bi, and M-Sb removed the malachite green only two times faster than rhodamine B. Sudan II showed significantly faster adsorption than rhodamine B. Hence, 40 min was optimal for removing the dyes.

2.9. Effect of Adsorbent Dose

The effect of adsorbent amount was assessed by taking 40 mL of the initial dye concentration of 25 mgL−1, a 200 rpm stirring rate, and 40 min as the equilibrium time. The UV spectra of all the dyes at different dye adsorbances were taken at optimized conditions with changing doses. The adsorbent dosage varied from 10 to 50 mg for both types of ionic liquids, and the adsorption efficiency was enhanced with the amount of adsorbent. This may be attributed to the fact that the adsorption sites increase with an increase in adsorbents, leading to elevated adsorption efficiency [41]. The adsorbent dose of 30 mg was found to be the optimum for both types of ILs for removing the MG, RhB and Sudan II dyes (Figure 5c).

2.10. Effect of Dye Concentration

The effect of initial dye concentration on the adsorption properties of adsorbents is illustrated in Figure 5d. To assess the impact of dye, solutions of dyes from 1 ppm to 40 ppm were prepared. Each solution was agitated under optimized dose, pH and time. The experiments showed that adsorption efficiencies decreased with the increase in dye concentration for all dyes, whereas the adsorption capacities increased. This decrease in efficiency may be attributed to the fact that the number of adsorption sites is limited in adsorbents, so further sorbate molecules could not find a site to get attached. The increase in adsorption capacities is due to the maximum coverage of the adsorption sites [54,55].

2.11. Effect of Metal and Alkyl Chain Length

The UV spectra results indicate that Bi-ILs demonstrate better adsorption efficiency than Sb-ILs, which may be attributed to the larger atomic radius and higher polarizability of bismuth, leading to enhanced electrostatic interactions with organic dye molecules. Additionally, Bi-ILs may exhibit greater Lewis acidity, strengthening the adsorption process [56]. Furthermore, the ethyl-substituted ILs displayed slightly higher adsorption capacity than their methyl counterparts, which can be ascribed to the number of carbon atoms, probably providing more active sites for adsorption, increased hydrophobic interactions and improved molecular orientation [55]. However, the increase in adsorption capacity remains insignificant due to the small difference in alkyl chain length between methyl and ethyl groups [Figure 4]. The absorption spectrum of Sudan II dye undergoes not only a reduction in intensity but also a noticeable change in shape when Bi-based adsorbents (M-Bi and E-Bi) are introduced. The shift in peak positions or the emergence of new spectral features could indicate chemical interactions, such as complexation or changes in the electronic environment of the dye molecules [57].

2.12. Kinetic Studies

The mechanism of molecule adsorption onto adsorbent can be investigated by several kinetic models. To examine the sorption mechanism, characteristic constants of adsorption were determined by using the Lagergren equation and a pseudo-second-order equation based on solid phase adsorption, respectively.

2.13. Pseudo-First-Order Kinetics

The equation of Lagergren or pseudo-first-order kinetics is represented in Equation (1).
d q t / d t = k 1 ( q e q t )
This equation is used to check the rate and kinetics of adsorption. The integrated form of the equation can be written as
l o g ( q e q t ) = l o g q e k 1 2.303 t
where qe and qt are the amount adsorbed at equilibrium and time ‘t’, respectively; and k is the rate constant. k values and qe values can be calculated from the slope and intercept, respectively. If the sorption system occurred via the pseudo-first-order kinetics, the experimental values should fit the linear plot at all concentrations. The important constants calculated from slope and intercept, respectively, are shown in Table 1. These values correspond to the regression of 0.92 to 0.97.

2.14. Pseudo-Second-Order Kinetics

The pseudo-second-order reaction can be expressed in non-linear form as Equation (3):
d q t / d t = k 2 ( q e q t ) 2
where qt is the adsorbed amount (mg·g−1) at a given time, qe is the adsorbed amount at equilibrium and k is the rate constant. To explain the adsorption process, the equation is used in its linear form as given (Equation (4)):
t q t = 1 k 2 q e 2 + 1 q e t
In a normal sorption experiment, initially, due to the large number of vacant sorption sites being available, the adsorbed amount increases rapidly and gradually decreases with time until equilibrium is reached.
The correlation coefficient values were higher than those of pseudo-first-order kinetics, indicating that adsorption on adsorbent occurred through second-order kinetics, based on the correlation with pseudo-second-order kinetics. Although the adsorption followed pseudo-second-order kinetics, which often implies chemisorption, the low enthalpy values and good reusability suggest that the rate-limiting step is more likely governed by physisorption involving weak electrostatic interactions or van der Waals forces [58]. The rate constants, regression correlation and adsorption capacity at equilibrium calculated from the plots are tabulated in Table 1.

2.15. Elovich Plot

The Elovich equation also provides evidence for the nature of adsorption systems, describes the kinetics on heterogeneous surfaces and gives insights about the mechanism of adsorption. The integrated form of the equation is written as
q t = 1 β l n α β + 1 β l n ( t )
The constant β (mg·g−1) is an important parameter of this equation, which is called an adsorption constant, and is associated with the activation energy of chemisorption and the extent of surface coverage. α is associated with the initial adsorption rate in the Elovich equation. Their values depend on temperature; with an increase in temperature, the constants α and β increase, showing an enhanced rate of adsorption and desorption. The Elovich model is typically linked with chemisorption; it can also represent certain physisorption processes, especially those that involve heterogeneous surfaces. In this study, malachite green and rhodamine B have shown a good fit to the Elovich model, and their α and β values are listed in Table 1. However, the adsorption process, due to low values of enthalpy and the higher recyclability of ILs, is described predominantly as physisorption that possibly involves weak interactions.
From the experimental values and plot regression values, it is seen that the adsorption process followed pseudo-second-order kinetics for M-Bi and M-Sb for malachite green, rhodamine B and Sudan II, supporting the involvement of surface interactions as the rate-limiting step.

2.16. Adsorption Isotherm Models

Adsorption isotherms were used to assess the interaction between adsorbate particles and adsorbent surfaces that play a vital role in the optimization of adsorbents. The Langmuir and Freundlich models were preferred to examine the adsorption of dye on ILs.

2.17. Langmuir Isotherm

The Langmuir model proposes that intermolecular forces are indirectly related to the distance between sorbate and sorbent molecules and consequently assumes the formation of a monolayer on the surface of the adsorbent. The Langmuir isotherm further predicts that each active site on the adsorbent molecule is occupied by only one molecule. Once the site is occupied, no further molecules can bind to that site. Moreover, the adsorption capacity of an adsorbent is limited for sorbate, beyond which it cannot adsorb additional molecules; this point is referred to as saturation, and the corresponding value is known as the maximum adsorption capacity of the adsorbent. The linear form of the Langmuir isotherm is given by Equation (6).
C e q e = 1 K L q m + C e q m
Here, Ce is the equilibrium concentration or final concentration, qe is the adsorbed amount at equilibrium, qm is the maximum adsorption capacity of the adsorbent and KL is the Langmuir constant (dm3/mg), related to the affinity of the adsorbent for an adsorbate.
The plot of Ce vs. Ce/qe gives a straight line, and KL and qm are derived from the slope and intercept, respectively. The Langmuir isotherm can also be analyzed using the non-linear form of the following equation:
q e = q m a x .   K L . C e 1 + K L .   C e
Therefore, non-linear regression was performed to directly estimate qmax and KL, ensuring accurate parameter estimation and minimizing error distortion. The goodness-of-fit metrics (R2) confirmed the suitability of the Langmuir model. The Langmuir model has been studied at room temperature, and both adsorbents fit the Langmuir isotherm model. R2, qm and KL values are given in Table 2. Adsorption values of bismuth-incorporated ionic liquids are found to be higher than those of antimony-containing ionic liquids.

2.18. Freundlich Isotherm

The adsorption mechanism of heterogeneous systems is explained by the Freundlich isotherm. The Freundlich empirical equation can be expressed as
qe = KFC1/n.
where qe is the equilibrium adsorption capacity, and C is the concentration at equilibrium. A linear form of the Freundlich equation can be written by taking the natural log of the following equation:
l n q e = l n K F + 1 n l n C e
where KF is the Freundlich constant, which gives the measure of adsorption capacity; and 1/n is the adsorption intensity of the adsorbent. The value of the Freundlich constant and the heterogeneity factor can be measured from the intercept and slope of the isotherm plot ln of Ce vs. ln qe.
Based on R2 values, the adsorption mechanism can be better described by the Langmuir isotherm, which represents the homogeneity of the adsorbent surface. The 1/n values represent the favorability of the adsorbent–adsorbate system. The values of the Freundlich plot are given in Table 2 for both the adsorbent and its adsorption properties for dyes. The intercept KF obtained from the Freundlich isotherm gives roughly the idea of adsorption capacity, and the slope, which is the inverse of n, gives the adsorption intensity. The value of n = 1 suggests homogeneous adsorption and predicts that interactions are absent between the adsorbed species. The value of 1/n less than 1 verifies the favorability of adsorption, and if the value of 1/n is higher than 1, the adsorption is not favorable, in which case the adsorption capacity decreases because of a weaker bond. In the current study, it was found that the 1/n value is less than 1 in the Langmuir plot, which means that adsorption is favorable for the Langmuir isotherm model.

2.19. Thermodynamic Studies

To calculate the thermodynamic parameters, the Van’t Hoff equation is used: [59]
l n q e / q o C e / C 0 = Δ G o = Δ H o R T + Δ S o R
Δ G ° , Δ H ° and Δ S o of adsorption can be calculated from linear and non-linear forms of the Van’t Hoff equation, as described in the literature, and apparently, no remarkable difference could be observed either in sign or in magnitude [60].
The standard enthalpy of reaction shows that adsorption was endothermic, and further values below 50 kJ/mol represent that the adsorption was the physisorption type in both cases. However, the values of entropy and standard Gibbs free energy show that adsorption was spontaneous. The calculated thermodynamic parameters are shown in Table 3.
Comparing the Sb- and Bi-based adsorbents, both have almost similar characteristics of adsorption. Bi-IL- and Sb-IL-based adsorbents both followed the Langmuir isotherm and pseudo-second-order kinetics. The rates are faster for bismuth-based ionic liquids, which could be ascribed to the size and electronegativity difference. The rate constant values are listed in Table 1. The adsorption capacities of both the adsorbents are comparable and have higher capacities for malachite green than Sudan II, followed by rhodamine B. The enthalpy of reaction suggested that both followed endothermic adsorption, and their adsorption was spontaneous, with Gibbs free energy, entropy and enthalpy values listed in Table 3.

2.20. Comparison with the Literature

The adsorption properties and capacities are compared with the previously available adsorbents for malachite green and rhodamine B in Table 4. Most of the adsorbents have followed pseudo-second-order adsorption kinetics and Langmuir isotherms, with few exceptions. The rate constants of prepared adsorbents have shown comparable rate constants and have great adsorption capacities. Similar materials, i.e., imidazolium, Chitosan, and phosphonium ILs, have shown only 84, 9.20 and 150 mg·g−1, respectively, for malachite green, whereas the metal-based ionic liquids in this work showed adsorption capacities of 217 mg·g−1 and 230 mg·g−1 by M-Sb and M-Bi, respectively.
The other materials used for dye removal by adsorption are mainly carbonaceous materials and usually ashes. Rice husk has exhibited an adsorption capacity of 17.98 mg·g−1 for malachite green following pseudo-second-order kinetics and the Freundlich isotherm, i.e., multilayer adsorption. Waste apricot has given the high adsorption capacities for malachite green of about 116.3 mg·g−1, and activated slag adsorbed about 74 mg·g−1 of malachite green following the Langmuir isotherm.
The adsorption capacities by mango leaf powder and fly ash were used by Sharma et al. and Chang et al. [61,62] respectively, and their adsorption capacities were a maximum of 3.31 mg·g−1 and 10.0 mg·g−1, respectively; both followed a single-layer adsorption. Whereas ionic liquids have not been used as adsorbents for rhodamine B removal, other materials, like fly ash and mango leaf powder, have shown adsorption capacities of only 10 mg·g−1 and 3.3 mg·g−1, respectively. Most adsorbents do not adsorb rhodamine B, for it is a hard dye. The sago waste activated carbon has an adsorption capacity of 28 mg·g−1 for the removal of rhodamine B. Duo Lite Resin C-20 has shown a 28 mg·g−1 adsorption capacity for rhodamine B, while antimony- and bismuth-based ionic liquids have shown adsorption capacities of 62 and 64 mg·g−1, respectively.
Table 4. Comparison with the previously reported studies.
Table 4. Comparison with the previously reported studies.
AdsorbentAdsorbateIsotherm ModelReaction KineticsAdsorption Capacity (mg·g−1)References
Activated slagMGLangmuir-74.2[63]
Waste apricotMGLangmuir Model-116.3[64]
Self-assembled ionic liquid-based organosilica (SAIBO)MGLangmuir ModelPseudo-2nd Order
k2 = 0.01
19.23[65]
[BMIM][PF6] MG-Pseudo-2nd Order
k2 = 0.056
84[41]
Chitosan-based ionic liquidsMGLangmuir ModelPseudo-2nd Order
k2 = 0.055
9.20[66]
[PC6C6C6C14] [Tf2N]MG--150[67]
Rice HuskMGFreundlich ModelPseudo-2nd Order
k2 = 0.032
17.98[68]
M-SbMGLangmuir ModelPseudo-2nd Order k2 = 0.00014217.36This work
M-BiMGLangmuir ModelPseudo-2nd Order k2 = 0.00077230.18This work
Sago waste activated carbonRhodamine BLangmuir ModelPseudo-2nd Order16.12[69]
Hyper cross-linked polymeric adsorbentRhodamine BFreundlich ModelPseudo-2nd Order
k2 = 0.0002
2.1[70]
Fly ashRhodamine BLangmuir Model-10.0[61]
Mango leaf powderRhodamine BLangmuir ModelPseudo-1st Order k1 = 0.0533.31[62]
Duo-lite C-20 resinRhodamine BLangmuir ModelPseudo-1st Order k1 = 0.06928.57[71]
M-SbRhodamine BLangmuir ModelPseudo-2nd Order k2 = 0.0000662.21This work
M-BiRhodamine BLangmuir ModelPseudo-2nd Order k2 = 0.0001364.94This work
Organoclays
MCNTS
Sudan IVFreundlich ModelPseudo-2nd Order22.7[72]
Coastal soilSudan IV--0.43–4.6[73]
M-SbSudan IILangmuir ModelPseudo-2nd Order162.3This Work
M-BiSudan IILangmuir ModelPseudo-2nd Order170.1This Work

2.21. Regeneration and Recyclability of Adsorbents/ILs

The regeneration and reusability of adsorbents are crucial constraints to evaluate their overall activity and practical applications [67]. After the complete separation of dyes with the hydrophobic antimony- and bismuth-based ILs (M-Sb & M-Bi), ethanol was used to wash the ILs. In this way, ILs can be recovered and reused in a new cycle after filtration. Interestingly, Sudan II has a 2.953 mg/kg saturation limit in water [53], suggesting that the solubility of these hydrophobic dyes increases in the presence of ILs in aqueous solution, and the same aqueous phase of ILs could be used several times without reaching saturation, as shown in Figure 6.

3. Experimental

3.1. Materials

N,N-Dimethylaniline, methyl iodide, ethyl iodide, antimony(III) chloride, dimethyl sulfoxide (DMSO), bismuth(III) chloride, methanol, acetonitrile (ACN), tetrahydrofuran (THF), hydrochloric acid (HCl), malachite green (MG), rhodamine B (RhB) and Sudan II were purchased from Sigma-Aldrich Chemicals (Co., Gillingham, UK) and used without any purification.

3.2. Instrumentations

The melting points of synthesized ILs were measured by using an electrothermal melting point apparatus (model MP-D Mitamura Riken, Kogyo, Tokyo, Japan). The Bio-Rad Excalibur (model FTS 3000 MX, Bio-Rad, Hercules, CA, USA) and Bruker Advance Digital FT-NMR spectrometer (Fällanden, Switzerland) with 300 MHz were employed to record the IR spectra and multinuclear (1H and 13C) NMR spectra, respectively. DMSO and deuterated ethanol were used as solvents for 1H and 13C NMR. A Renishaw in Via Raman microscope (London, UK) was used to obtain the Raman spectra, having a 10 s acquisition time, around a 1 µm2 laser spot and an excitation wavelength of 785 nm (130 mW, 10%). Optical studies were carried out using the UV-1700 Shimadzu (Kyoto, Japan).

3.3. Synthesis of Adsorbent

Antimony- and bismuth-based ionic liquids were prepared in a two-step reaction using a reported procedure [25], described in Scheme 1. In the first step, the alkylation of N,N-dimethylaniline was performed, which was followed by an anion exchange reaction. An equimolar ratio of alkyl (methyl and ethyl) iodide (0.025 moles) and N,N-dimethylaniline (0.025 moles) was prepared in 25 mL of THF separately and was stirred in a 100 mL round-bottom flask. The temperature of the reaction mixture was maintained at 70–80 °C and allowed to react for 24 h. Pale-green solid products (M & E-ILs) were obtained in the flask, which was filtered, washed with ethyl acetate and allowed to dry. The filtrate was stirred for an additional hour to ensure the complete precipitation of the product. An above 90% yield was obtained after the complete precipitation of the product. In a second step, metal was incorporated by a metathesis reaction. The synthesized N,N,N-trialkylanilinium iodide (M & E-ILs) and metal (antimony and bismuth) chlorides in the equimolar ratio of 1:1 were separately dissolved in acetonitrile and agitated for 6 h, although the reaction was rapid, and solid products were precipitated immediately at room temperature when the reactants were mixed. The light-yellow antimony and orange-colored bismuth-based solid products (Sb & Bi-ILs) were separated, washed with n-hexane and diethyl ether and allowed to air dry. The structures of these synthesized ILs are shown in Scheme 2.

3.4. Ionic Liquids Having Iodide as Anion (M- and E-ILs)

N,N,N-Trimethylanilinium iodide (M-IL): 1H NMR (400 MHz, CD3Cl), d: 7.69 (d, 2JHH = 8.0 Hz, 2H), 7.54 (m, 3H), 3.50 (s, 9H). 13C NMR (150 MHz, CD3Cl), d: 147.70, 130.51, 120.92, 56.88. FT-IR (cm−1); nN–C (1006), sp3 C–H (2852), sp2 C–H (2983), nC=C(1589).
N,N-Dimethyl-N-ethylanilinium iodide (E-IL): 1H NMR (400 MHz, CD3Cl), d: 8.01 (d, 2JHH = 10.8 Hz 2H), 7.61 (m, 5H), 3.64 (s, 2H), 3.37 (s, 6H). 13C NMR (75 MHz, CD3Cl), d: 143.53, 130.78, 120.13, 65.14, 55.06, 9.34. FT-IR (cm−1); nN–C (1006), sp3 C–H (2852), sp2 C–H (2983), nC=C (1589).
Antimony-based ionic liquids (Sb-ILs)
N,N,N-Trimethylanilinium antimony(IV) tetrachloride (M-Sb): Quantities used: N,N,N-Trimethylanilinium iodide (2 mmol), Antimony(III) tetrachloride (2 × 10⁻3 mol); m.p.: 217 °C; Pale yellow powder; anal. calcd%: C, 27.04; H, 3.53; Cl, 35.47; N, 3.50; Sb, 30.46; found %: C, 26.99; H, 3.51; N, 3.48; Cl, 35.46; Sb; 30.44, FT-IR (cm−1); nN–C (1001), sp3 C–H (2858), sp2 C–H (2987), nC=C (1589).
N,N-Dimethy-N-ethylanilinium antimony(IV)tetrachloride (E-Sb): Quantities used: N,N-Dimethyl-N-ethylanilinium iodide (2 mmol), Antimony(III) tetrachloride (2 × 10⁻3 mol); m.p.: 211 °C; Pale yellow powder; anal. calcd%: C, 35.48; H, 5.95; Cl, 29.92; N, 2.96; Sb, 25.69; found %: C, 35.50; H, 5.94; N, 2.97; Cl, 29.93; Sb; 25.7, FT-IR (cm−1) nN–C (1000), sp3 C–H (2855), sp2 C–H (2972), nC=C (1589).
Bismuth-based ionic liquids (Bi-ILs)
N,N-N-Trimethylanilinium Bismuth(IV) tetrachloride (M-Bi): Quantities used: N,N-Dimethyl-N-ethylanilinium iodide (2 mmol), Bismuth(III) tetrachloride (2 mmol); m.p.: 215 °C; Pale yellow powder; anal. calcd (%) C, 22.20; H, 2.90; Bi, 42.91; Cl, 29.12; N, 2.88; found %: C, 22.18; H, 2.94; N, 2.93; Cl, 29.09; Bi; 42.90, FT-IR (cm−1) nN–C (1001), sp3 C–H (2875), sp2 C–H (2970), nC=C (1478).
N,N-Dimethy-N-ethylanilinium Bismuth(IV) tetrachloride (E-Bi): Quantities used: N,N-Dimethyl-N-ethylanilinium iodide (2 mmol), Bismuth(III) tetrachloride (2 mmol); m.p.: 201 °C; Pale yellow powder; anal. calcd(%): C, 23.97; H, 3.22; Bi, 41.71; Cl, 28.30; N, 2.80; found %: C, 23.96; H, 3.20; Bi, 41.71; Cl, 28.31; N, 2.81, FT-IR (cm−1) nN–C (1003), sp3 C–H (2851), sp2 C–H (2969), nC=C (1478).

3.5. Dye Adsorption

Three dyes—malachite green, Sudan II and rhodamine B—were adsorbed on a prepared adsorbent according to the following procedure: A stock solution of 1000 ppm was prepared by dissolving a weighed amount of 100 mg of dye in 100 mL of deionized water (50% ethanol was used for Sudan II). The solutions were diluted to an appropriate concentration for conducting the experiments. A weighed amount of adsorbent was added to 40 cm3 of 25 ppm dye solution in water and stirred for different intervals of time (10% ethanolic solution was used for Sudan II). The following four parameters were varied to find the effect on the rate of adsorption and are discussed in detail below. The adsorbate and synthesized adsorbent (antimony- and bismuth-based ILs) were separated by centrifugation, followed by filtration. The lingering concentration of dye was determined by using a calibration curve attained from a UV/Vis spectrophotometer at its particular maximum wavelength.
A d s o r p t i o n   E f f e c i e n c y ( S % ) = ( C i C e ) C i x 100
A d s o r p t i o n   c a p a c i t y ( q e ) = ( C i C e ) m V
where Ci = initial concentration of the dyes (mg L−1);
Ce = the equilibrium concentration of the dye (mg L−1);
m = the mass of adsorbent material (g);
V= the volume (dm3).

3.6. Adsorption Kinetics

The kinetic studies were carried out at various temperatures (293, 303, 313, 323 and 333 K). Adsorbates at initial concentrations of 25 ppm were agitated with optimized amounts of adsorbents for the optimized intervals of time, and the final concentrations were recorded spectrophotometrically.

3.7. Adsorption Isotherm

The equilibrium points of RhB, MG and Sudan II were determined by carrying out the agitation experiment of the optimized dose of adsorbents (Sb- and Bi-ILs) for each adsorbate with varying concentrations (5, 10, 15, 25 and 50 ppm) of all three adsorbates at varying pH values. Approximately 5 mL of each solution was taken out after intervals of time and centrifuged, and then final concentrations were determined spectrophotometrically at 555 nm for RhB, 617 nm for MG and 495 nm for Sudan II.

4. Conclusions

The Sb- and Bi-based ILs were successfully synthesized by a metathesis reaction. All of them were characterized by 1HNMR, 13CNMR, FTIR, Raman and X-ray photoelectron spectroscopy. They showed great adsorption efficiencies for the removal of rhodamine B and malachite green. The parameters affecting adsorption efficiencies (stirring rate, pH, time and adsorbent dosage) were optimized. The adsorption efficiencies increased with an increase in stirring rate, pH, time and adsorbent dosage. With the optimization of factors, the adsorption efficiencies increased above 90%. Bi- and Sb-based ILs showed comparable adsorption capacities. The maximum adsorption capacity for malachite green, Sudan II and rhodamine B by M-Bi was 230.18 mg·g−1, 170 mg·g−1 and 64.21 mg·g−1, respectively, whereas by M-Sb it was 217.36 mg·g−1, 162.1 mg·g−1 and 62.94 mg·g−1, respectively. The adsorption was homogeneous in nature, occurred through monolayer adsorption and followed pseudo-second-order kinetics. The thermodynamic parameters showed that the adsorption was endothermic and occurred spontaneously by a physisorption process.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal15050492/s1, Figure S1: FT-IR spectra for ILs, Figure S2: 1H NMR spectrum for (M-IL), Figure S3: 13 CNMR spectrum for (M-IL), Figure S4: 1 HNMR spectrum for (E-IL), Figure S5: 13 CNMR spectrum for (E-IL).

Author Contributions

Conceptualization, M.S.; Methodology, A.Z. and N.R.; Validation, S.B.; Formal analysis, N.R.; Investigation, A.Z.; Writing—original draft, A.Z. and N.R.; Writing—review and editing, S.B., M.S. and A.A.; Supervision, A.W.; Project administration, A.W.; Funding acquisition, A.A. All authors have read and agreed to the published version of the manuscript.

Funding

The researchers would like to thank the Deanship of Graduate Studies and Scientific Research at Qassim University for financial support (QU-APC-2025).

Data Availability Statement

The authors confirm that the data supporting the findings of this study are available within the article or in supplementary information.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Dutta, S.; Gupta, B.; Srivastava, S.K.; Gupta, A.K. Recent Advances on the Removal of Dyes from Wastewater Using Various Adsorbents: A Critical Review. Mater. Adv. 2021, 2, 4497–4531. [Google Scholar] [CrossRef]
  2. Petcu, A.R.; Lazar, C.A.; Rogozea, E.A.; Olteanu, N.L.; Meghea, A.; Mihaly, M. Nonionic Microemulsion Systems Applied for Removal of Ionic Dyes Mixtures from Textile Industry Wastewaters. Sep. Purif. Technol. 2016, 158, 155–159. [Google Scholar] [CrossRef]
  3. Zhou, Y.; Lu, J.; Zhou, Y.; Liu, Y. Recent Advances for Dyes Removal Using Novel Adsorbents: A Review. Environ. Pollut. 2019, 252, 352–365. [Google Scholar] [CrossRef]
  4. Yusuf, M. Synthetic Dyes: A Threat to the Environment and Water Ecosystem. In Textiles and Clothing; Scrivener Publishing LLC: Austin, TX, USA, 2019; pp. 11–26. [Google Scholar]
  5. Holkar, C.R.; Jadhav, A.J.; Pinjari, D.V.; Mahamuni, N.M.; Pandit, A.B. A Critical Review on Textile Wastewater Treatments: Possible Approaches. J. Environ. Manag. 2016, 182, 351–366. [Google Scholar] [CrossRef]
  6. Yadav, A.K.; Jain, C.K.; Malik, D.S. Toxic Characterization of Textile Dyes and Effluents in Relation to Human Health Hazards. J. Sustain. Environ. Res. 2014, 3, 95–102. [Google Scholar]
  7. Aigbe, U.O.; Ukhurebor, K.E.; Onyancha, R.B.; Okundaye, B.; Pal, K.; Osibote, O.A.; Esiekpe, E.L.; Kusuma, H.S.; Darmokoesoemo, H. A Facile Review on the Sorption of Heavy Metals and Dyes Using Bionanocomposites. Adsorpt. Sci. Technol. 2022, 2022, 8030175. [Google Scholar] [CrossRef]
  8. Hassaan, M.A.; El Nemr, A.; Hassaan, A. Health and Environmental Impacts of Dyes: Mini Review. Am. J. Environ. Sci. Eng. 2017, 1, 64–67. [Google Scholar]
  9. Manzoor, J.; Sharma, M. Impact of Textile Dyes on Human Health and Environment. In Impact of Textile Dyes on Public Health and the Environment; IGI Global: Hershey, PA, USA, 2020; pp. 162–169. [Google Scholar]
  10. Bal, G.; Thakur, A. Distinct Approaches of Removal of Dyes from Wastewater: A Review. Mater. Today Proc. 2022, 50, 1575–1579. [Google Scholar] [CrossRef]
  11. Kandisa, R.V.; Saibaba, K.V.N.; Shaik, K.B.; Gopinath, R. Dye Removal by Adsorption: A Review. J. Bioremediation Biodegrad. 2016, 7, 371. [Google Scholar] [CrossRef]
  12. Kaykhaii, M.; Sasani, M.; Marghzari, S. Removal of Dyes from the Environment by Adsorption Process. Chem. Mater. Eng. 2018, 6, 31–35. [Google Scholar] [CrossRef]
  13. Madan, S.; Shaw, R.; Tiwari, S.; Tiwari, S.K. Adsorption Dynamics of Congo Red Dye Removal Using ZnO Functionalized High Silica Zeolitic Particles. Appl. Surf. Sci. 2019, 487, 907–917. [Google Scholar] [CrossRef]
  14. Areeb, A.; Yousaf, T.; Murtaza, M.; Zahra, M.; Zafar, M.I.; Waseem, A. Green Photocatalyst Cu/NiO Doped Zirconia for the Removal of Environmental Pollutants. Mater. Today Commun. 2021, 28, 102678. [Google Scholar] [CrossRef]
  15. Saeed, M.; Munir, M.; Nafees, M.; Shah, S.S.A.; Ullah, H.; Waseem, A. Synthesis, Characterization and Applications of Silylation Based Grafted Bentonites for the Removal of Sudan Dyes: Isothermal, Kinetic and Thermodynamic Studies. Microporous Mesoporous Mater. 2020, 291, 109697. [Google Scholar] [CrossRef]
  16. Tee, G.T.; Gok, X.Y.; Yong, W.F. Adsorption of Pollutants in Wastewater via Biosorbents, Nanoparticles and Magnetic Biosorbents: A Review. Environ. Res. 2022, 212, 113248. [Google Scholar] [CrossRef]
  17. Ullah, R.; Iftikhar, F.J.; Ajmal, M.; Shah, A.; Akhter, M.S.; Ullah, H.; Waseem, A. Modified Clays as an Efficient Adsorbent for Brilliant Green, Ethyl Violet and Allura Red Dyes: Kinetic and Thermodynamic Studies. Pol. J. Environ. Stud. 2020, 29, 3831–3839. [Google Scholar] [CrossRef]
  18. Yousaf, T.; Areeb, A.; Murtaza, M.; Munir, A.; Khan, Y.; Waseem, A. Silane-Grafted MXene (Ti3C2TX) Membranes for Enhanced Water Purification Performance. ACS Omega 2022, 7, 19502–19512. [Google Scholar] [CrossRef]
  19. Wang, H.; Li, Z.; Yahyaoui, S.; Hanafy, H.; Seliem, M.K.; Bonilla-Petriciolet, A.; Dotto, G.L.; Sellaoui, L.; Li, Q. Effective Adsorption of Dyes on an Activated Carbon Prepared from Carboxymethyl Cellulose: Experiments, Characterization and Advanced Modelling. Chem. Eng. J. 2021, 417, 128116. [Google Scholar] [CrossRef]
  20. Lee, C.H.; Tang, A.Y.L.; Wang, Y.; Kan, C.W. Effect of Reverse Micelle-Encapsulated Reactive Dyes Agglomeration in Dyeing Properties of Cotton. Dye. Pigment. 2019, 161, 51–57. [Google Scholar] [CrossRef]
  21. Chikri, R.; Elhadiri, N.; Benchanaa, M.; Maguana, Y. Efficiency of Sawdust as Low-Cost Adsorbent for Dyes Removal. J. Chem. 2020, 2020, 2217. [Google Scholar] [CrossRef]
  22. Pandey, D.; Savio, N.; Srivastava, R.K. A Comprehensive Review on the Production and Application of Sludge Biochar for Water and Soil Remediation. Int. J. Res. Eng. Sci. Manag. 2021, 4, 118–122. [Google Scholar]
  23. Velusamy, S.; Roy, A.; Sundaram, S.; Kumar Mallick, T. A Review on Heavy Metal Ions and Containing Dyes Removal through Graphene Oxide-based Adsorption Strategies for Textile Wastewater Treatment. Chem. Rec. 2021, 21, 1570–1610. [Google Scholar] [CrossRef] [PubMed]
  24. Chaudhary, S.; Kaur, Y.; Umar, A.; Chaudhary, G.R. Ionic Liquid and Surfactant Functionalized ZnO Nanoadsorbent for Recyclable Proficient Adsorption of Toxic Dyes from Waste Water. J. Mol. Liq. 2016, 224, 1294–1304. [Google Scholar] [CrossRef]
  25. Zafar, A.; Imtiaz-Ud-Din.; Ahmed, S.; Bučar, D.K.; Tahir, M.N.; Palgrave, R.G. Synthesis, Structural Analysis, Electrochemical and Magnetic Properties of Tetrachloroferrate Ionic Liquids. New J. Chem. 2021, 45, 13429–13440. [Google Scholar] [CrossRef]
  26. Deng, H.N.; Xing, Y.L.; Xia, C.L.; Sun, H.M.; Shen, Q.; Zhang, Y. Ionic iron (III) complexes of bis (phenol)-functionalized imidazolium cations: Synthesis, structures and catalysis for aryl Grignard cross-coupling of alkyl halides. Dalton Trans. 2012, 38, 11597–11607. [Google Scholar] [CrossRef] [PubMed]
  27. de Jesus, S.S.; Filho, R.M. Are Ionic Liquids Eco-Friendly? Renew. Sustain. Energy Rev. 2022, 157, 112039. [Google Scholar] [CrossRef]
  28. Flieger, J.; Flieger, M. Ionic Liquids Toxicity—Benefits and Threats. Int. J. Mol. Sci. 2020, 21, 6267. [Google Scholar] [CrossRef]
  29. Khoo, Y.S.; Tjong, T.C.; Chew, J.W.; Hu, X. Techniques for Recovery and Recycling of Ionic Liquids: A Review. Sci. Total Environ. 2024, 922, 171238. [Google Scholar] [CrossRef]
  30. Gujjala, L.K.S.; Kundu, D.; Dutta, D.; Kumar, A.; Bal, M.; Kumar, A.; Singh, E.; Mishra, R.; Kumar, S.; Vo, D.-V.N. Advances in Ionic Liquids: Synthesis, Environmental Remediation and Reusability. J. Mol. Liq. 2024, 396, 123896. [Google Scholar] [CrossRef]
  31. Ezzat, A.O.; Atta, A.M.; Al-Lohedan, H.A. Demulsification of Stable Seawater/Arabian Heavy Crude Oil Emulsions Using Star-like Tricationic Pyridinium Ionic Liquids. Fuel 2021, 304, 121436. [Google Scholar] [CrossRef]
  32. Piper, S.L.; Kar, M.; MacFarlane, D.R.; Matuszek, K.; Pringle, J.M. Ionic Liquids for Renewable Thermal Energy Storage-a Perspective. Green Chem. 2022, 24, 102–117. [Google Scholar] [CrossRef]
  33. Anceschi, A.; Riccardi, C.; Patrucco, A. The Role of Ionic Liquids in Textile Processes: A Comprehensive Review. Molecules 2025, 30, 353. [Google Scholar] [CrossRef]
  34. Mero, A.; Mezzetta, A.; Nowicki, J.; Łuczak, J.; Guazzelli, L. Betaine and L-Carnitine Ester Bromides: Synthesis and Comparative Study of Their Thermal Behaviour and Surface Activity. J. Mol. Liq. 2021, 334, 115988. [Google Scholar] [CrossRef]
  35. Qiao, Y.; Shi, E.; Wei, X.; Hou, Z. Ionic Liquid-Stabilized Metal Oxoclusters: From Design to Catalytic Application. Green Chem. 2024, 26, 5127–5149. [Google Scholar] [CrossRef]
  36. Dupont, J.; Leal, B.C.; Lozano, P.; Monteiro, A.L.; Migowski, P.; Scholten, J.D. Ionic Liquids in Metal, Photo-, Electro-, and (Bio) Catalysis. Chem. Rev. 2024, 124, 5227–5420. [Google Scholar] [CrossRef]
  37. Liu, F.; Yu, J.; Qazi, A.B.; Zhang, L.; Liu, X. Metal-Based Ionic Liquids in Oxidative Desulfurization: A Critical Review. Environ. Sci. Technol. 2021, 55, 1419–1435. [Google Scholar] [CrossRef] [PubMed]
  38. Fernandez, J.F.; Neumann, J.; Thoming, J. Regeneration, Recovery and Removal of Ionic Liquids. Curr. Org. Chem. 2011, 15, 1992–2014. [Google Scholar] [CrossRef]
  39. Mai, N.L.; Ahn, K.; Koo, Y.-M. Methods for Recovery of Ionic Liquids—A Review. Process Biochem. 2014, 49, 872–881. [Google Scholar] [CrossRef]
  40. Zafar, A.; Palgrave, R.G.; Muhammad, H. Physico-Chemical Properties of Magnetic Dicationic Ionic Liquids with Tetrahaloferrate Anions. ChemistryOpen 2023, 12, e202200229. [Google Scholar] [CrossRef]
  41. Ullah, Z.; Bustam, M.A.; Man, Z.; Khan, A.S. Phosphonium-Based Ionic Liquids and Their Application in Separation of Dye from Aqueous Solution. ARPN J. Eng. Appl. Sci. 2016, 11, 1653–1659. [Google Scholar]
  42. Pei, Y.; Liu, J.; Yan, Z.; Li, Z.; Fan, J.; Wang, J. Association of Ionic Liquids with Cationic Dyes in Aqueous Solution: A Thermodynamic Study. J. Chem. Thermodyn. 2012, 47, 223–227. [Google Scholar] [CrossRef]
  43. Zafar, A.; Evans, T.; Palgrave, R.G.; Imtiaz-ud-Din. An XPS Study of Bridged Dicationic Ionic Liquids. J. Chem. Res. 2022, 46, 174751982210929. [Google Scholar] [CrossRef]
  44. Jabar, G.; Saeed, M.; Khoso, S.; Zafar, A.; Saggu, J.I.; Waseem, A. Development of Graphitic Carbon Nitride Supported Novel Nanocomposites for Green and Efficient Oxidative Desulfurization of Fuel Oil. Nanomater. Nanotechnol. 2022, 12, 18479804221106321. [Google Scholar] [CrossRef]
  45. Zafar, A.; Batool, S.; Palgrave, R.G.; Yousuf, S. Deep Extractive and Catalytic Oxidative Desulfurization of Liquid Fuels by Using Iron (Iii) Based Dication Ionic Liquids. New J. Chem. 2023, 47, 15731–15747. [Google Scholar] [CrossRef]
  46. Rowe, R.; Lovelock, K.R.J.; Hunt, P.A. Bi (III) Halometallate Ionic Liquids: Interactions and Speciation. J. Chem. Phys. 2021, 155, 14501. [Google Scholar] [CrossRef]
  47. Persaud, A. The Aqueous Chloride Complexes of Antimony (III) and Lanthanum (III) from 25 C up to 300 C by Raman Spectroscopy; University of Guelph: Guelph, Canada, 2021. [Google Scholar]
  48. Li, T.; Cai, T.; Hu, H.; Li, X.; Wang, D.; Zhang, Y.; Cui, Y.; Zhao, L.; Xing, W.; Yan, Z. Multivalent Cationic and Anionic Mixed Redox of an Sb 2 S 3 Cathode toward High-Capacity Aluminum Ion Batteries. J. Mater. Chem. A 2022, 10, 10829–10836. [Google Scholar] [CrossRef]
  49. Singh, S.; Sahoo, R.K.; Shinde, N.M.; Yun, J.M.; Mane, R.S.; Chung, W.; Kim, K.H. Asymmetric Faradaic Assembly of Bi2O3 and MnO2 for a High-Performance Hybrid Electrochemical Energy Storage Device. RSC Adv. 2019, 9, 32154–32164. [Google Scholar] [CrossRef]
  50. Stevens, J.S.; Byard, S.J.; Seaton, C.C.; Sadiq, G.; Davey, R.J.; Schroeder, S.L.M. Proton Transfer and Hydrogen Bonding in the Organic Solid State: A Combined XRD/XPS/ssNMR Study of 17 Organic Acid–Base Complexes. Phys. Chem. Chem. Phys. 2014, 16, 1150–1160. [Google Scholar] [CrossRef]
  51. Abd Malek, N.N.; Jawad, A.H.; Ismail, K.; Razuan, R.; ALOthman, Z.A. Fly Ash Modified Magnetic Chitosan-Polyvinyl Alcohol Blend for Reactive Orange 16 Dye Removal: Adsorption Parametric Optimization. Int. J. Biol. Macromol. 2021, 189, 464–476. [Google Scholar] [CrossRef]
  52. Bagotia, N.; Sharma, A.K.; Kumar, S. A Review on Modified Sugarcane Bagasse Biosorbent for Removal of Dyes. Chemosphere 2021, 268, 129309. [Google Scholar]
  53. Ferreira, A.M.; Coutinho, J.A.P.; Fernandes, A.M.; Freire, M.G. Complete Removal of Textile Dyes from Aqueous Media Using Ionic-Liquid-Based Aqueous Two-Phase Systems. Sep. Purif. Technol. 2014, 128, 58–66. [Google Scholar] [CrossRef]
  54. Archana, V.; Begum, K.M.M.S.; Anantharaman, N. Studies on Removal of Phenol Using Ionic Liquid Immobilized Polymeric Micro-Capsules. Arab. J. Chem. 2016, 9, 371–382. [Google Scholar] [CrossRef]
  55. Gharehbaghi, M.; Shemirani, F. A Novel Method for Dye Removal: Ionic Liquid-Based Dispersive Liquid–Liquid Extraction (IL-DLLE). Clean Soil Air Water 2012, 40, 290–297. [Google Scholar] [CrossRef]
  56. Ranjan, M.; Singh, P.K.; Srivastav, A.L. A Review of Bismuth-Based Sorptive Materials for the Removal of Major Contaminants from Drinking Water. Environ. Sci. Pollut. Res. 2020, 27, 17492–17504. [Google Scholar] [CrossRef]
  57. Oladipo, A.A.; Mustafa, F.S. Bismuth-Based Nanostructured Photocatalysts for the Remediation of Antibiotics and Organic Dyes. Beilstein J. Nanotechnol. 2023, 14, 291–321. [Google Scholar] [CrossRef]
  58. Ullah, H.; Nafees, M.; Iqbal, F.; Awan, M.S.; Shah, A.; Waseem, A. Adsorption Kinetics of Malachite Green and Methylene Blue from Aqueous Solutions Using Surfactant-Modified Organoclays. Acta Chim. Slov. 2017, 64, 449–460. [Google Scholar] [CrossRef]
  59. Savara, A. Standard States for Adsorption on Solid Surfaces: 2D Gases, Surface Liquids, and Langmuir Adsorbates. J. Phys. Chem. C 2013, 117, 15710–15715. [Google Scholar] [CrossRef]
  60. Lima, E.C.; Gomes, A.A.; Tran, H.N. Comparison of the Nonlinear and Linear Forms of the van’t Hoff Equation for Calculation of Adsorption Thermodynamic Parameters (∆ S° And∆ H°). J. Mol. Liq. 2020, 311, 113315. [Google Scholar] [CrossRef]
  61. Chang, S.-H.; Wang, K.-S.; Li, H.-C.; Wey, M.-Y.; Chou, J.-D. Enhancement of Rhodamine B Removal by Low-Cost Fly Ash Sorption with Fenton Pre-Oxidation. J. Hazard. Mater. 2009, 172, 1131–1136. [Google Scholar] [CrossRef]
  62. Khan, T.A.; Sharma, S.; Ali, I. Adsorption of Rhodamine B Dye from Aqueous Solution onto Acid Activated Mango (Magnifera Indica) Leaf Powder: Equilibrium, Kinetic and Thermodynamic Studies. Toxicol. Environ. Health Sci. 2011, 3, 286–297. [Google Scholar]
  63. Mittal, A.; Mittal, J.; Kurup, L.; Singh, A.K. Process Development for the Removal and Recovery of Hazardous Dye Erythrosine from Wastewater by Waste Materials—Bottom Ash and de-Oiled Soya as Adsorbents. J. Hazard. Mater. 2006, 138, 95–105. [Google Scholar] [CrossRef]
  64. Başar, C.A. Applicability of the Various Adsorption Models of Three Dyes Adsorption onto Activated Carbon Prepared Waste Apricot. J. Hazard. Mater. 2006, 135, 232–241. [Google Scholar] [CrossRef] [PubMed]
  65. Elhamifar, D.; Shojaeipoor, F.; Roosta, M. Self-Assembled Ionic-Liquid Based Organosilica (SAILBO) as a Novel and Powerful Adsorbent for Removal of Malachite Green from Aqueous Solution. J. Taiwan Inst. Chem. Eng. 2016, 59, 267–274. [Google Scholar] [CrossRef]
  66. Naseeruteen, F.; Hamid, N.S.A.; Suah, F.B.M.; Ngah, W.S.W.; Mehamod, F.S. Adsorption of Malachite Green from Aqueous Solution by Using Novel Chitosan Ionic Liquid Beads. Int. J. Biol. Macromol. 2018, 107, 1270–1277. [Google Scholar] [CrossRef]
  67. Thasneema, K.K.; Dipin, T.; Thayyil, M.S.; Sahu, P.K.; Messali, M.; Rosalin, T.; Elyas, K.K.; Saharuba, P.M.; Anjitha, T.; Hadda, T. Ben Removal of Toxic Heavy Metals, Phenolic Compounds and Textile Dyes from Industrial Waste Water Using Phosphonium Based Ionic Liquids. J. Mol. Liq. 2021, 323, 114645. [Google Scholar] [CrossRef]
  68. Chowdhury, S.; Mishra, R.; Saha, P.; Kushwaha, P. Adsorption Thermodynamics, Kinetics and Isosteric Heat of Adsorption of Malachite Green onto Chemically Modified Rice Husk. Desalination 2011, 265, 159–168. [Google Scholar] [CrossRef]
  69. Kadirvelu, K.; Karthika, C.; Vennilamani, N.; Pattabhi, S. Activated Carbon from Industrial Solid Waste as an Adsorbent for the Removal of Rhodamine-B from Aqueous Solution: Kinetic and Equilibrium Studies. Chemosphere 2005, 60, 1009–1017. [Google Scholar] [CrossRef]
  70. Huang, J.-H.; Huang, K.-L.; Liu, S.-Q.; Wang, A.-T.; Yan, C. Adsorption of Rhodamine B and Methyl Orange on a Hypercrosslinked Polymeric Adsorbent in Aqueous Solution. Colloids Surf. Physicochem. Eng. Asp. 2008, 330, 55–61. [Google Scholar] [CrossRef]
  71. Al-Rashed, S.M.; Al-Gaid, A.A. Kinetic and Thermodynamic Studies on the Adsorption Behavior of Rhodamine B Dye on Duolite C-20 Resin. J. Saudi Chem. Soc. 2012, 16, 209–215. [Google Scholar] [CrossRef]
  72. Huang, L.; Zhou, Y.; Guo, X.; Chen, Z. Simultaneous Removal of 2, 4-Dichlorophenol and Pb (II) from Aqueous Solution Using Organoclays: Isotherm, Kinetics and Mechanism. J. Ind. Eng. Chem. 2015, 22, 280–287. [Google Scholar] [CrossRef]
  73. Teng, Y.; Zhou, Q. Adsorption Behavior of Sudan I-IV on a Coastal Soil and Their Forecasted Biogeochemical Cycles. Environ. Sci. Pollut. Res. 2017, 24, 10749–10758. [Google Scholar] [CrossRef]
Figure 1. Raman spectra for Sb- and Bi-ILs.
Figure 1. Raman spectra for Sb- and Bi-ILs.
Catalysts 15 00492 g001
Figure 2. High-resolution XP spectra of M-Sb showing (a) full survey scan, (b) Sb 3d, (c) Cl 2p, (d) C 1s and (e) N 1s.
Figure 2. High-resolution XP spectra of M-Sb showing (a) full survey scan, (b) Sb 3d, (c) Cl 2p, (d) C 1s and (e) N 1s.
Catalysts 15 00492 g002
Figure 3. High-resolution XP spectra of M-Bi showing (a) full survey scan, (b) Bi 4f, (c) Cl 2p, (d) C 1s and (e) N 1s.
Figure 3. High-resolution XP spectra of M-Bi showing (a) full survey scan, (b) Bi 4f, (c) Cl 2p, (d) C 1s and (e) N 1s.
Catalysts 15 00492 g003
Figure 4. Adsorption behavior of dyes on various ILs.
Figure 4. Adsorption behavior of dyes on various ILs.
Catalysts 15 00492 g004
Figure 5. Optimization studies. Effect of (a) stirring rate, (b) contact time, (c) adsorption dose, (d) initial dye concentration.
Figure 5. Optimization studies. Effect of (a) stirring rate, (b) contact time, (c) adsorption dose, (d) initial dye concentration.
Catalysts 15 00492 g005aCatalysts 15 00492 g005b
Figure 6. Effects of adsorbate on recycling.
Figure 6. Effects of adsorbate on recycling.
Catalysts 15 00492 g006
Scheme 1. Route for synthesis of Sb- and Bi-ILs.
Scheme 1. Route for synthesis of Sb- and Bi-ILs.
Catalysts 15 00492 sch001
Scheme 2. Structures of Sb- and Bi-ILs.
Scheme 2. Structures of Sb- and Bi-ILs.
Catalysts 15 00492 sch002
Table 1. Kinetic parameters.
Table 1. Kinetic parameters.
AdsorbentDyesPseudo-1st-Order KineticsPseudo-2nd-Order KineticsElovich Plot
k1
(min−1)
qe
(mg·g−1)
R2k2
(min−1)
qe
(mg·g−1)
R2α
(mg·g−1 .min−1)
β
(g.mg−1)
R2
M-SbMG0.2258.10.9720.00077146.410.9911.0070.0290.989
RhB0.1147.640.9460.0013956.490.9981.030.0840.995
Sudan II0.030642.520.750.00012165.560.9818.770.0380.965
M-BiMG0.20436.590.9290.0001556.080.9892.260.0330.984
RhB0.054.550.966.1E-0585.470.9855.480.1930.983
Sudan II0.020214.530.910.0009134.950.9939.160.0350.97
Table 2. Adsorption parameters.
Table 2. Adsorption parameters.
AdsorbentAdsorbateLangmuirFreundlich
qe
(mg·g−1)
qm
(mg·g−1)
KL
(L.mg−1)
R2nKF
(mg·g−1)
R2
M-SbMG217.1246.20.2780.9951.994.090.98
RhB62.9466.450.1760.991.491.280.97
Sudan II162.1176.670.7610.922.4663.430.982
M-BiMG230.12612.630.9972.484.710.98
RhB64.2178.570.0780.991.843.680.97
Sudan II170.4187.961.190.8662.0851.410.942
Table 3. Thermodynamic parameters.
Table 3. Thermodynamic parameters.
AdsorbentAdsorbateΔS°
(J·mol−1·K−1)
ΔG°
(J·mol−1)
ΔH°
(J·mol−1)
M-SbMG566.5−1920.0−46,660.0
RhB651.9−3350.0−45.34
Sudan II478.7−3580.0−16,840.0
M-BiMG−1.619−3420.0−20,340.0
RhB471.4−4110.0−3.13
Sudan II475.1−3420.0−16,480.0
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

Zafar, A.; Rafique, N.; Batool, S.; Saleem, M.; Alhodaib, A.; Waseem, A. Antimony- and Bismuth-Based Ionic Liquids as Efficient Adsorbents for the Removal of Dyes. Catalysts 2025, 15, 492. https://doi.org/10.3390/catal15050492

AMA Style

Zafar A, Rafique N, Batool S, Saleem M, Alhodaib A, Waseem A. Antimony- and Bismuth-Based Ionic Liquids as Efficient Adsorbents for the Removal of Dyes. Catalysts. 2025; 15(5):492. https://doi.org/10.3390/catal15050492

Chicago/Turabian Style

Zafar, Anham, Nouman Rafique, Saadia Batool, Muhammad Saleem, Aiyeshah Alhodaib, and Amir Waseem. 2025. "Antimony- and Bismuth-Based Ionic Liquids as Efficient Adsorbents for the Removal of Dyes" Catalysts 15, no. 5: 492. https://doi.org/10.3390/catal15050492

APA Style

Zafar, A., Rafique, N., Batool, S., Saleem, M., Alhodaib, A., & Waseem, A. (2025). Antimony- and Bismuth-Based Ionic Liquids as Efficient Adsorbents for the Removal of Dyes. Catalysts, 15(5), 492. https://doi.org/10.3390/catal15050492

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

Article Metrics

Back to TopTop