Next Article in Journal
Strategy of Coniferous Needle Biorefinery into Value-Added Products to Implement Circular Bioeconomy Concepts in Forestry Side Stream Utilization
Next Article in Special Issue
A Study on the Adsorption of Rhodamine B onto Adsorbents Prepared from Low-Carbon Fossils: Kinetic, Isotherm, and Thermodynamic Analyses
Previous Article in Journal
Design, Synthesis, and Antifungal Activity of Some Novel Phenylthiazole Derivatives Containing an Acylhydrazone Moiety
Previous Article in Special Issue
Identification of the Contamination Sources by PCBs Using Multivariate Analyses: The Case Study of the Annaba Bay (Algeria) Basin
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Sorption Performance of Natural Zeolites Modified with pH-Fractionated Humic Acids for the Removal of Methylene Blue from Water

by
Stefano Salvestrini
1,*,
Jean Debord
2 and
Jean-Claude Bollinger
3
1
Department of Environmental, Biological and Pharmaceutical Sciences and Technologies, University of Campania “Luigi Vanvitelli”, 81100 Caserta, Italy
2
Service de Pharmacologie-Toxicologie, Hôpital Dupuytren, 87042 Limoges, France
3
Laboratoire E2Lim, Faculté des Sciences et Techniques, Université de Limoges, 87060 Limoges, France
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(20), 7083; https://doi.org/10.3390/molecules28207083
Submission received: 21 August 2023 / Revised: 24 September 2023 / Accepted: 11 October 2023 / Published: 14 October 2023

Abstract

:
This work explores the effect of humic acids (HA) fractionation on the sorption ability of a natural zeolite (NYT)—HA adduct. HA were extracted from compost, fractionated via the pH fractionation method, and characterized via UV-Vis spectroscopy and gel permeation chromatography. The HA samples were immobilized onto NYT via thermal treatment. The resulting adducts (NYT-HA) were tested for their ability to remove methylene blue (MB) from an aqueous solution. It was found that the sorption performance of NYT-HA strongly depends on the chemical characteristics of humic acids. Sorption capacity increased with the molecular weight and hydrophobicity degree of the HA fractions. Hydrophobic and π–π interactions are likely the primary mechanisms by which MB interacts with HA. The sorption kinetic data conform to the pseudo-second-order model. The Freundlich isotherm model adequately described the sorption equilibrium and revealed that the uptake of MB onto NYT-HA is endothermic in nature.

1. Introduction

Over the years, the continued release of organic contaminants from anthropogenic activities into natural waters has attracted increased attention owing to its implications for human health and environmental preservation [1].
Various techniques for the abatement of pollutants in water have been investigated. Among them, innovative processes using reactive oxygen species as degrading agents of pollutants (Advanced Oxidation Processes, AOPs) are currently being investigated [2,3,4,5]. Although AOPs are very promising, the well-established process of sorption is still extensively studied in the field of water remediation, mostly because of its effectiveness and ease of use [6].
Different materials have been tested as sorbents in aqueous systems. Activated carbon is probably the most widely applied sorbent for organic compounds because of its high surface area and low water affinity [7]. In the last few decades, the development of alternative sorbents, especially naturally occurring materials, has intensified. Of particular interest are natural minerals coated with an appropriate organic phase that imparts enhanced sorption ability to original materials. Among these minerals, organo-modified natural zeolites are noteworthy due to their pollutant removal efficiency. The organo-modification of zeolites can be accomplished using the following approaches: (i) the impregnation method, in which a zeolite is treated with a solution of a desired compound; (ii) the replacement of external exchangeable elemental cations (e.g., Na+, K+, and Ca2+) with organic cations; and (iii) the direct interaction of the external exchangeable cations with negatively charged molecules. Examples of the first approach include the use of biochar [8], chitosan [9], and biological matter [10,11]. The cation exchange modification of zeolites can be achieved using quaternary ammonium salts such as hexadecyltrimethylammonium chloride [12], cetylpyridinium bromide [13], and stearyldimethylbenzylammonium chloride [14,15]. An example of the third type of modification involves the use of humic acids (HA) [16,17]. Thanks to their chemical heterogeneity (e.g., large mass distribution, presence of various functional groups, and presence of hydrophilic and hydrophobic moieties [18,19,20,21]), HA are able to efficiently interact with a wide variety of chemical compounds, rendering the zeolite-HA adduct a very versatile sorbent [22]. The loading of HA on zeolites is favored in the presence of exchangeable Ca2+ ions, which not only act as a bridge between the negatively charged surface of zeolites and the deprotonated functional groups (mainly carboxylic) of HA [17] but also promote HA self-aggregation [23].
Interestingly, the sorptive performance of the zeolite-HA material can be improved by selectively employing only the more appropriate fraction of a raw HA sample. HA fractionation can be achieved by means of ultrafiltration [24] or pH adjustments [25]. In particular, pH fractionation is preferred over ultrafiltration because it is easier to execute and does not suffer from the progressive fouling of the filter. The various pH-fractionated HA samples are expected to have different chemical compositions [25,26] and possibly different affinities for the sorbable organic contaminants.
Based on the above considerations, in this study, an HA sample was extracted from a commercial vegetable compost and fractionated using the pH fractionation method. The unfractionated HA and their fractions were characterized via UV-Vis spectroscopy and gel permeation chromatography. The samples were immobilized on a natural zeolite (Neapolitan Yellow Tuff, NYT) and tested with respect to their ability to act as sorbents for the removal of methylene blue (MB). MB was chosen as a model compound due to its high occurrence in water worldwide and its harmfulness to human health and the environment [27,28]. The kinetic and thermodynamic aspects of the sorption process were also investigated.

2. Results and Discussion

2.1. HA Characterization

The HA used for this study were extracted from vegetable compost and fractionated via sequential pH solubilization. To gain information on the physico-chemical characteristics of the HA, high-pressure gel permeation chromatography analyses were performed [29]. The results of these experiments are reported in Figure 1. The chromatogram of the unfractionated HA (HAref, Figure 1A) is very broad, typical of polydisperse substances with a wide distribution of molecular weights [30]. In contrast, all the pH-fractionated HA samples (Figure 1B–D) exhibit a slightly more symmetrical and sharp elution peak, indicating the presence of a more homogeneous distribution of molecules with respect to HAref. Moreover, the retention time of the pH-fractionated HA decreases with an increasing pH. This suggests that the HA samples dissolved at higher pH levels have a higher average molecular weight. The latter finding is in agreement with the results obtained by Zhang et al. using pH-fractionated HA derived from Chinese weathered coal [25]. Additional information on the characteristics of the HA fractions were obtained via UV-Vis spectroscopic measurements, specifically by determining the ratio of absorbance at 250 nm to 365 nm (E2/E3) and at 465 nm to 665 nm (E4/E6). These ratios are widely used as descriptors of the chemical properties of HA [31]. The values of E2/E3 and E4/E6 are generally inversely correlated with the aromaticity, degree of condensation, and molecular weight of HA [32,33,34]. In our study, the E2/E3 and E4/E6 ratios decreased with the increase in pH (see Table 1). This finding suggests that sequential pH solubilization progressively leads to HA fractions having higher aromaticity content and, in line with the gel permeation experiments, a higher molecular weight distribution. It is interesting to note that HA exhibiting appreciable aromaticity and sizes are also likely associated with a high degree of hydrophobicity [35].

2.2. Sorption Kinetics of MB

2.2.1. MB Sorption Kinetics on NYT

Figure 2 shows the trend of the total MB concentration in the liquid phase versus time at different initial concentrations of MB and a constant loading of natural zeolite (NYT). For all the kinetic runs, an appreciable decrease in the MB concentration was observed as a result of the sorption process.
In order to model the uptake of MB, the popular pseudo-first-order (PFO), pseudo-second-order (PSO), Boyd, and Vermeulen models were selected.
The PFO model is an empirical model in which the rate of sorption is considered to be proportional to the distance from equilibrium, wherein the latter is expressed as the difference between the sorbed amount at equilibrium and at any time t [36]:
d q d t = k 1 ( q e q )
Here, k 1 (h−1) is the PFO kinetic rate constant, and q (mg g−1) is the amount of MB sorbed per mass of sorbent (see Equation (15)), wherein the subscript e denotes the sorption equilibrium state.
Another empirical model is the PSO model (Equation (2)), which considers the rate of sorption to be proportional to the square of the distance from equilibrium [36]:
d q d t = k 2 ( q e q ) 2
Here, k 2 (g mg−1 h−1) is the PSO kinetic rate constant.
Integrating Equations (1) and (2) for the boundary conditions t = 0 to t and q = 0 to q and replacing q with Equation (15) permits us to derive the dependence of the MB aqueous concentration (mg L−1) on time, according to the PFO (Equation (3)) and PSO (Equation (4)) models, respectively:
C = C 0 X q e ( 1 exp ( k 1 t ) )
C = C 0 X q e 2 k 2 t 1 + q e k 2 t
The Boyd (or Reichenberg) model [37] represents an approximated form of a more complex sorption kinetic model relying on the intraparticle diffusion theory developed, among others, by Boyd himself and Crank [38]. The Boyd approximating model consists of the following two equations.
C = C 0 X q e ( 6 π 3 / 2 B t 3 π 2 B t )         ( for   q / q e < 0.85 )
C = C 0 X q e ( 1 6 π 2 exp ( B t ) )         ( for   q / q e > 0.85 )
Here, B (h−1) is a constant related to the sorbent particle size and the effective diffusion coefficient of the sorbate inside the sorbent.
Another approximation of the intraparticle diffusion model is the Vermeulen equation [39]:
C = C 0 X q e ( 1 exp ( k D t ) )
The parameter k D (h−1) has a physical meaning similar to that of B.
Equations (3)–(7) were used to model the kinetic data for the sorption of MB onto NYT. The Boyd model was applied, switching from Equations (5) to (6) in correspondence with the threshold value q/qe = 0.85. Table 2 and Table S1 show the results of the regression procedure. According to a statistical analysis, the results of which are reported in these tables, in most of the cases, the sorption kinetics conform more closely to the Vermeulen and the Boyd models, as can be inferred from the low AICc values and the low error of the estimated parameters. Moreover, the Vermeulen equation provides a reliable estimate of the amount of MB sorbed at equilibrium.
The result of the application of the Vermeulen model to the kinetic data is presented in Figure 2. The close fitting between the Vermeulen model and the Boyd model could suggest that the sorption of MB onto NYT is controlled by intraparticle diffusion.

2.2.2. MB Sorption Kinetics on NYT-HA

The sorption rate of MB as affected by humic acids immobilization on the NYT surface is shown in Figure 3.
From the figure, it is evident that the sorption kinetics of MB are influenced by both the presence of HA and its pH fractionation. The uptake rate of MB decreases in the following row: NYT-HA7 > NYT-HAref > NYT-HA5 > NYT-HA3 > NYT.
The data in figure were modelled using Equations (3)–(7), and the results of this procedure are displayed in Table 3 and Table S2. Overall, the PSO model provides the best agreement with the experimental data, while the Boyd and Vermeulen intraparticle diffusion models were found to be less accurate than when they were used for modelling the sorption kinetics of MB onto NYT alone. This finding could be explained by the fact that the rate of MB uptake onto NYT-HA is controlled by surface interactions with the HA domain rather than by diffusion phenomena [40,41].

2.3. Sorption Equilibrium

2.3.1. Isotherms of the Sorption of MB onto Natural Zeolite and Natural Zeolite–Humic Acids Adducts: Effect of pH Fractionation

Figure 4 compares the isotherm data at equilibrium for the sorption of MB onto various sorbents.
As can be seen, the raw zeolitic material has the lowest sorption capacity. It is clear that NYT benefits from the presence of HA and, more importantly, from its fractionation. NYT-HA7 outperforms all the other tested sorbents in terms of sorption efficiency, followed by NYT-HAref, NYT-HA5, and NYT-HA3.
In order to shed light on the reasons underlying the different behaviors of the sorbents, their points of zero charge were measured and reported in Table 4.
The point of zero charge (pHPZC) is a useful parameter for gaining information on the net surface charge of a sorbent at a given pH and may help in elucidating mechanisms of sorption [42,43]. Generally, the higher or the lower the pH with respect to pHPZC, the greater the net negative or positive charge surface of the sorbent, respectively. According to the pHPZC values reported in Table 4, NYT bears a net positive charge surface at the experimental pH (7.4). In contrast, all the NYT-HA samples have lower pHPZC values and are negatively charged (pHPZC < pH). The lower pHPZC of the NYT-HA sorbents can be ascribed to the presence of acidic groups, mostly carboxylic and phenolic groups, in the HA moiety [44]. The observed pHPZC decrement is less marked than that found for other materials coated with HA [44,45]. A feasible explanation for this finding is the partial decarboxylation of our HA during the thermal immobilization procedure [46,47].
In many studies, the sorption of MB is ascribed to electrostatic forces established between the negatively charged groups of the sorbent surface and the MB cation [48,49,50]. It is clear that such interactions are favored when pHPZC < pH, and this would explain the lower uptake of MB by NYT as compared to the NYT-HA samples. However, the pHPZC values of the NYT-HA materials are similar to each other and show a poor correlation with sorbent sorption capacity (cfs. Table 4 and Figure 4). This suggests that factors other than pHPZC also control the performance of the sorbents. The results of the UV-Vis and GPC experiments support this hypothesis: in Section 2.1, it was shown that the HA fractions obtained at higher pH levels were larger and had higher aromaticity and hydrophobicity than those isolated at lower pH levels. Moreover, the sorbents containing higher amounts of pH-fractionated HA exhibited higher sorption capacity (Figure 4). Based on the above considerations, it is reasonable to infer that the sorption of MB onto the NYT-HA materials is primarily driven by π–π and hydrophobic effects. Similar interactions were found for MB in the presence of carbon nanotubes [51] and cyclodextrin derivatives [52]. The sorbing MB molecule is expected to orient itself parallel to the sorbent surface, with the positively charged sulfur atom facing the liquid phase [51].
The data in Figure 4 were modelled using the most representative sorption isotherm models, namely, the classical Langmuir model [53] and the Freundlich model [54,55]:
q e = q m K L C e 1 + K L C e
q e = K F C e n
q m (mg g−1) and K L (L mg−1) represent the maximum sorption capacity and the equilibrium constant of the Langmuir model, whereas K F (mg1-n g−1 Ln) and n (dimensionless) are the Freundlich isotherm parameters. The results of the fitting procedure are displayed in Table 5. They show that the NYT isotherm data more closely fit the Langmuir model, while the HYT-HA isotherms are better described by the Freundlich model. These findings may be taken as an indication that NYT-HA, contrary to NYT, exhibits a heterogeneous sorbing surface for MB [56].
In order to evaluate the efficiency of the best sorbent, NYT-HA7, Table 6 reports the predicted amounts of MB sorbed ( q e ) at the highest equilibrium concentration explored (Ce = 1.0 mg L−1) and, for comparison, the predicted amounts of MB sorbed in the same experimental conditions by other sorbents from the literature
The data in Table 6 reveal that the performance of NYT-HA7 is quite good in comparison with that of other materials. The sorption properties of the NYT-HA material could be improved by enhancing the loading content of humic acids and varying their chemical compositions, for example, by using a different source of humic acids or via physicochemical activation. These aspects will be explored in a more systematic manner in future works.

2.3.2. Sorption Thermodynamics

The effect of temperature on the sorption equilibrium of MB was investigated for the best-performing sorbent NYT-HA7. The sorption isotherms at 293, 300, 307, and 313 K are displayed in Figure 5. The curves in the figure were obtained by applying the Freundlich model, and the estimates of the fitted parameters are reported in Table 7. The values of q e and C e in Figure 5 were expressed in mol kg−1 and mol L−1 in accordance with the standard states universally accepted in thermodynamic sorption studies [68].
It can be seen from the figure that the sorption of MB increased with temperature, which suggests that the process is endothermic in nature ( Δ H ° > 0 ). A similar behavior was found for the sorption of MB onto bentonite [69], peanut-shell-based activated carbon [70], and iron-oxide-modified montmorillonite [71]. In contrast, an exothermic sorption of MB was observed, for example, in the presence of bone char [72], citrate-modified pomelo peel [73], and a magnetic-gas-to-liquid-derived biosolid [74].
In order to acquire quantitative information on the thermodynamic parameters of the sorption of MB onto NYT-HA7, the following expressions were used [75]:
Δ G ° = R T ln K °
Δ G ° = Δ H ° T Δ S °
Here, R = 8.314 J K−1 mol−1 is the universal gas constant, and Δ G ° (kJ mol−1), Δ H ° (kJ mol−1), and Δ S ° (J K−1 mol−1) are the standard sorption Gibbs energy, standard sorption enthalpy, and standard sorption entropy, respectively. K ° (dimensionless) is the thermodynamic sorption equilibrium constant [76]. Its value should be derived from the best-fitting isotherm model. In the case of the Freundlich model, K ° can be calculated from its exponent n , as recently shown by Debord et al. [56]. According to the authors, the Freundlich isotherm is consistent with the existence of a heterogeneous sorbent whose sorbing sites exhibit an exponential distribution of binding energies ( Δ G ° ). The sorption equilibrium on each site can be locally described using a Langmuir-type model. The exponent n is related to the mean value of Δ G ° ( Δ G ° 0 ) and to its associated thermodynamic equilibrium constant K ° 0 via the relationships
Δ G ° 0 = R T n
K ° 0 = exp ( 1 n )
from which we can obtain
1 n = Δ H ° R T + Δ S ° R
According to Equation (14), the values of Δ H ° and Δ S ° (which are assumed not to vary appreciably over the temperature range investigated) can be estimated from the slope and the intercept, respectively, with the y-axis of the straight line obtained by plotting 1 / n against 1 / T .
The results of the application of Equation (14) to the sorption data regarding the sorption of MB onto NYT-HA7 are graphically displayed in Figure 6. The fitting results exhibit a good linear relationship, and the resultant values of Δ H ° and Δ S ° were 3.9 ± 0.6 kJ mol−1 and 24 ± 2 J K−1 mol−1, respectively.

3. Materials and Methods

3.1. Chemicals

All reagents were purchased from Sigma-Aldrich (St. Louis, MO, USA).

3.2. Natural Zeolite Sample

The natural zeolite sample came from a Neapolitan yellow tuff quarry located in Marano, some 10 km NW of Naples, Italy. The material was characterized in a previous work [77] as being composed 34% by weight of zeolite (phillipsite). Its chemical composition is as follows, wt%: SiO2 = 52.9, Al2O3 = 14.7, Fe2O3 = 4.0, MgO = 1.1, CaO = 2.1, K2O = 7.6, Na2O = 2.8, and P2O5 = 0.1. The corresponding cation exchange capacity and specific surface area are 1.90 meq g−1 and 23 m2 g−1, respectively.
Before its use, the material (henceforth termed NYT, denoting “Neapolitan yellow tuff”) was crushed and sieved to produce particles with a size (0.5–1 mm) suitable for real industrial applications [78]. Afterwards, the cation exchange surface of the sample was saturated in Ca2+ via treatment with 3 M of CaCl2 for 8 h under gentle stirring at a mass/liquid ratio of 1:100. The procedure was repeated two more times by replacing the supernatant with a fresh CaCl2 solution; finally, the NYT sample was recovered, washed with pure water, and dried at 40 °C in an oven for 24 h.

3.3. Extraction of Humic Acids from Vegetable Compost, pH Fractionation, and Characterization

3.3.1. HA Extraction

A commercial vegetable compost (VC) purchased from Selex (Milan, Italy) was used as source of humic acids (HA). The total organic carbon and the humic + fulvic carbon content of VC were (as declared by the supplier) 22% and 3% (in dry weight), respectively. HA were extracted from VC using NaOH and sodium pyrophosphate (Na4P2O7) [77]. In more detail, 1 kg of VC was placed in contact, for 2 days under stirring, with 10 L of 0.1 M NaOH + 0.1 M Na4P2O7. The sample was centrifuged at 1000 g-force units of relative centrifugal force for 1 h; then, the supernatant was recovered, and its pH was adjusted to 1.5 with a few drops of concentrated HCl in order to promote HA precipitation. The suspension was stored at 4 °C for 2 days and then centrifuged at 1000 g-force units for 1 h. Finally, the precipitate was collected, purified in a dialysis bag (cut off = 1000 Da), and lyophilized.

3.3.2. pH-Fractionation of HA

The HA extracted from VC (unfractionated HA soluble at pH 7, HAref) was separated into three fractions differing in their pH solubility. A total of 1 g of HAref was placed in contact with 1 L of pure water, and the pH was adjusted to 3.0 using NaOH. After 24 h, the soluble part of HA at pH 3 (hereafter referred to as HA3) was recovered from the supernatant via acidic precipitation. The insoluble part was adjusted to pH 5 and then pH 7. The soluble fractions of HA at pH 5 and 7 were recovered through the same procedure as that used for HA3 and denoted as HA5 and HA7, respectively.

3.3.3. HA Characterization

HA were analyzed via gel permeation chromatography (GPC) using a Waters instrument consisting of a 515 HPLC pump and a 2487 dual λ absorbance detector; additionally, the instrument was equipped with a Biosep (Melton Mowbray, UK) Sec-2000 (300 × 7.80 mm2) column. According to a previous report [79], the following operational conditions were selected: mobile phase composition = 75% 10 mM phosphate buffer (pH 7.0) + 25% CH3CN (v/v); flow rate = 0.7 mL min−1; detection wavelength = 280 nm.
UV-Vis spectroscopic characterization of HA solutions was carried out by determining the ratio between the absorbance at 250 nm and 365 nm (E2/E3) and at 465 nm and 665 nm (E4/E6) [80] via a Perkin-Elmer (Waltham, MA, USA) Lambda 40 spectrophotometer.

3.4. Preparation of the Humic Acids—Zeolitic Tuff Sorbents and Point-of-Zero-Charge Measurements

Each HA sample (unfractionated or pH-fractionated) was dissolved in 10 mM Tris buffer (pH ≈ 7.4) to obtain a 100 mg L−1 solution. In total, 40 mL of each solution was then placed with 1 g of NYT pre-treated with CaCl2 in 50 mL conical tubes under continuous stirring using an orbital shaker operating at 30 rpm and under room temperature. At selected times, the absorbance of the supernatant at 450 nm was measured for the quantification of the loaded HA. When the loading of HA onto NYT reached approximatively the value of 2 mg g−1, the suspension was filtered on filter paper with a vacuum pump, and the solid was recovered and heated at 330 °C in an oven to stabilize, via decarboxylation, the NYT-HA adduct [17]. The entire procedure, i.e., HA loading and immobilization, was repeated four more times until reaching, for each type of sorbent, an HA loading of about 10 mg g−1. The obtained materials were named NYT-HAref, NYT-HA3, NYT-HA5, and NYT-HA7.
The point of zero charge (pHPZC) was determined via a pH titration procedure [42,43]. Aliquots of 0.01 M NaCl solution were placed in contact with 0.29 g L−1 of sorbent, and the pH was adjusted to between 2 and 10. The pH was measured again after 2 days. A plot of the final pH against the initial pH was constructed. The pH at which the initial and final pH values were the same corresponded to pHPZC.

3.5. MB Spectrophotometric Measurements and Calibration Curve

During preliminary experiments, we observed that MB is prone to strongly sorbing onto any glass surface, including vials and Pasteur pipettes but also quartz cuvettes, in line with a previous report [81]. This can lead to an overestimation of the sorption capacity of NYT and HA for MB. For this reason, we used only plastic materials to store, handle, and analyze the MB solutions.
The MB solutions were analyzed in polystyrene cuvettes using a Lambda 40 Perkin Elmer spectrophotometer. The absorbance recorded at 665 nm was used for determining the MB concentration.
The absorbance–concentration calibration curve was constructed as follows. A 100 mg L−1 stock solution of MB (methylene blue, also known as methylthioninium chloride, with a molar mass = 319.85 g mol−1) was prepared by dissolving 58.44 mg of methylene blue trihydrate (molar mass = 373.9 g mol−1) into 0.5 L of an aqueous solution buffered at pH = 7.4 with 10 mM Tris buffer. Appropriate volumes of this working solution were diluted with the addition of water in order to obtain a set of MB solutions with concentrations ranging from 0.5 mg L−1 to 14 mg L−1. The absorbance at 665 nm of each solution was then measured and plotted as a function of the MB concentration. The calibration curve (see Figure S1 in the Supplementary Information) was obtained by fitting the absorbance vs. concentration data with Equation (S6). The latter equation considers that MB exists in monomeric and dimeric forms only (a valid assumption for the range of concentrations used). More specific details about the calculation method can be found in the Supplementary Information.

3.6. MB Sorption Experiments

A total of 12 mg of the selected sorbent was placed in 50 mL polypropylene conical tubes. Afterwards, 42 mL of an MB aqueous solution (buffered at pH = 7.4) with a concentration varying between 1 and 12 mg L−1 (a concentration range similar to that reported in other MB sorption studies [60,82,83]) was added. The samples were thermostated at the desired temperature (20, 27, 34, or 40 °C) under continuous stirring and periodically analyzed via spectrophotometric measurements at 665 nm. The amount of MB sorbed per mass of sorbent ( q , mg g−1) was calculated using the following mass balance formula:
q = C 0 C X
Here, C 0 and C (mg L−1) are the initial and actual concentrations in the liquid phase, respectively, whereas X (g L−1) represents the ratio between the mass of sorbent and the volume of the solution.

4. Conclusions

In this work, HA were immobilized onto natural zeolites and successfully applied in the removal of MB from the liquid phase. The sorbing ability of HA is well known in the literature. Here, we showed that the fractionation of HA can greatly enhance their sorption efficiency. This can be easily accomplished via a sequential pH solubilization procedure. Thanks to their chemical heterogeneity, HA can interact with other molecules through different mechanisms involving electrostatic and π–π interactions and hydrophobic effects. Each HA fraction exhibits distinctive chemical characteristics that make it more or less affine towards a specific group of pollutants. Depending on the chemical nature of the target pollutant, the more appropriate fraction of HA can be conveniently selected to prepare an efficient sorbent material.
Based on the above considerations, natural zeolites modified with pH-fractionated humic acids can be considered versatile and promising materials for water remediation processes. Future works will be devoted to exploring the possibility of further implementing the sorption capacity of the HA-based sorbents, for example, by increasing the HA content and using HA of different origins and altering their chemical surface activity.

Supplementary Materials

Please add this part as following format: ‘The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28207083/s1. Refs [84,85] are in the Supplementary Materials.

Author Contributions

Conceptualization, investigation, and writing—original draft preparation, S.S.; writing—review and editing, J.D. and J.-C.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are available from the author on request.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not applicable.

References

  1. Singh, J.; Yadav, P.; Pal, A.K.; Mishra, V. Water Pollutants: Origin and Status. In Sensors in Water Pollutants Monitoring: Role of Material; Pooja, D., Kumar, P., Singh, P., Patil, S., Eds.; Springer: Singapore, 2020; pp. 5–20. ISBN 978-981-15-0670-3. [Google Scholar]
  2. Kang, Z.; Jia, X.; Zhang, Y.; Kang, X.; Ge, M.; Liu, D.; Wang, C.; He, Z. A Review on Application of Biochar in the Removal of Pharmaceutical Pollutants through Adsorption and Persulfate-Based AOPs. Sustainability 2022, 14, 10128. [Google Scholar] [CrossRef]
  3. Gupta, A.D.; Singh, H.; Varjani, S.; Awasthi, M.K.; Giri, B.S.; Pandey, A. A critical review on biochar-based catalysts for the abatement of toxic pollutants from water via advanced oxidation processes (AOPs). Sci. Total Environ. 2022, 849, 157831. [Google Scholar] [CrossRef] [PubMed]
  4. Jiang, T.; Wang, B.; Gao, B.; Cheng, N.; Feng, Q.; Chen, M.; Wang, S. Degradation of organic pollutants from water by biochar-assisted advanced oxidation processes: Mechanisms and applications. J. Hazard. Mater. 2023, 442, 130075. [Google Scholar] [CrossRef] [PubMed]
  5. Deng, Y.; Zhao, R. Advanced Oxidation Processes (AOPs) in Wastewater Treatment. Curr. Pollut. Rep. 2015, 1, 167–176. [Google Scholar] [CrossRef]
  6. Jia, Z.; Wu, L.; Han, C.; Zhang, D.; Li, M.; Wei, R. Preparation of auto-suspending hollow silica microfiber derived from biotemplate and its highly selective adsorption for organic dyes. Mater. Today Commun. 2023, 34, 105437. [Google Scholar] [CrossRef]
  7. Jaria, G.; Calisto, V.; Esteves, V.I.; Otero, M. Overview of relevant economic and environmental aspects of waste-based activated carbons aimed at adsorptive water treatments. J. Clean. Prod. 2022, 344, 130984. [Google Scholar] [CrossRef]
  8. Atugoda, T.; Gunawardane, C.; Ahmad, M.; Vithanage, M. Mechanistic interaction of ciprofloxacin on zeolite modified seaweed (Sargassum crassifolium) derived biochar: Kinetics, isotherm and thermodynamics. Chemosphere 2021, 281, 130676. [Google Scholar] [CrossRef]
  9. Kazemi, J.; Javanbakht, V. Alginate beads impregnated with magnetic Chitosan@Zeolite nanocomposite for cationic methylene blue dye removal from aqueous solution. Int. J. Biol. Macromol. 2020, 154, 1426–1437. [Google Scholar] [CrossRef] [PubMed]
  10. Hamd, A.; Dryaz, A.R.; Shaban, M.; Almohamadi, H.; Abu Al-Ola, K.A.; Soliman, N.K.; Ahmed, S.A. Fabrication and application of zeolite/acanthophora spicifera nanoporous composite for adsorption of congo red dye from wastewater. Nanomaterials 2021, 11, 2441. [Google Scholar] [CrossRef]
  11. Dryaz, A.R.; Shaban, M.; AlMohamadi, H.; Al-Ola, K.A.A.; Hamd, A.; Soliman, N.K.; Ahmed, S.A. Design, characterization, and adsorption properties of Padina gymnospora/zeolite nanocomposite for Congo red dye removal from wastewater. Sci. Rep. 2021, 11, 21058. [Google Scholar] [CrossRef]
  12. Bowman, R.S. Applications of surfactant-modified zeolites to environmental remediation. Microporous Mesoporous Mater. 2003, 61, 43–56. [Google Scholar] [CrossRef]
  13. Zhang, X.; Bai, R. Mechanisms and kinetics of humic acid adsorption onto chitosan-coated granules. J. Colloid Interface Sci. 2003, 264, 30–38. [Google Scholar] [CrossRef]
  14. Milićević, S.; Matović, L.; Petrović, Đ.; Đukić, A.; Milošević, V.; Đokić, D.; Kumrić, K. Surfactant modification and adsorption properties of clinoptilolite for the removal of pertechnetate from aqueous solutions. J. Radioanal. Nucl. Chem. 2016, 310, 805–815. [Google Scholar] [CrossRef]
  15. Cappelletti, P.; Colella, A.; Langella, A.; Mercurio, M.; Catalanotti, L.; Monetti, V.; de Gennaro, B. Use of surface modified natural zeolite (SMNZ) in pharmaceutical preparations Part 1. Mineralogical and technological characterization of some industrial zeolite-rich rocks. Microporous Mesoporous Mater. 2017, 250, 232–244. [Google Scholar] [CrossRef]
  16. Chianese, S.; Fenti, A.; Iovino, P.; Musmarra, D.; Salvestrini, S. Sorption of Organic Pollutants by Humic Acids: A Review. Molecules 2020, 25, 918. [Google Scholar] [CrossRef] [PubMed]
  17. Leone, V.; Canzano, S.; Iovino, P.; Salvestrini, S.; Capasso, S. A novel organo-zeolite adduct for environmental applications: Sorption of phenol. Chemosphere 2013, 91, 415–420. [Google Scholar] [CrossRef]
  18. De Melo, B.A.G.; Motta, F.L.; Santana, M.H.A. Humic acids: Structural properties and multiple functionalities for novel technological developments. Mater. Sci. Eng. C 2016, 62, 967–974. [Google Scholar] [CrossRef]
  19. Yonebayashi, K.; Hattori, T. Chemical and biological studies on environmental humic acids. Soil Sci. Plant Nutr. 2012, 34, 571–584. [Google Scholar] [CrossRef]
  20. Abakumov, E.V.; Cajthaml, T.; Brus, J.; Frouz, J. Humus accumulation, humification, and humic acid composition in soils of two post-mining chronosequences after coal mining. J. Soils Sediments 2013, 13, 491–500. [Google Scholar] [CrossRef]
  21. He, M.; Shi, Y.; Lin, C. Characterization of humic acids extracted from the sediments of the various rivers and lakes in China. J. Environ. Sci. 2008, 20, 1294–1299. [Google Scholar] [CrossRef]
  22. Leone, V.; Iovino, P.; Salvestrini, S.; Capasso, S. Sorption of non-ionic organic pollutants onto a humic acids-zeolitic tuff adduct: Thermodynamic aspects. Chemosphere 2014, 95, 75–80. [Google Scholar] [CrossRef] [PubMed]
  23. Chen, M.; Gowthaman, S.; Nakashima, K.; Kawasaki, S. Influence of humic acid on microbial induced carbonate precipitation for organic soil improvement. Environ. Sci. Pollut. Res. 2023, 30, 15230–15240. [Google Scholar] [CrossRef] [PubMed]
  24. Guo, J.; Liu, H.; Liu, J.; Wang, L. Ultrafiltration performance of EfOM and NOM under different MWCO membranes: Comparison with fluorescence spectroscopy and gel filtration chromatography. Desalination 2014, 344, 129–136. [Google Scholar] [CrossRef]
  25. Zhang, S.; Yuan, L.; Li, W.; Lin, Z.; Li, Y.; Hu, S.; Zhao, B. Characterization of pH-fractionated humic acids derived from Chinese weathered coal. Chemosphere 2017, 166, 334–342. [Google Scholar] [CrossRef]
  26. Baglieri, A.; Vindrola, D.; Gennari, M.; Negre, M. Chemical and spectroscopic characterization of insoluble and soluble humic acid fractions at different pH values. Chem. Biol. Technol. Agric. 2014, 1, 9. [Google Scholar] [CrossRef]
  27. Sivakumar, R.; Lee, N.Y. Adsorptive removal of organic pollutant methylene blue using polysaccharide-based composite hydrogels. Chemosphere 2022, 286, 131890. [Google Scholar] [CrossRef]
  28. Khan, I.; Saeed, K.; Zekker, I.; Zhang, B.; Hendi, A.H.; Ahmad, A.; Ahmad, S.; Zada, N.; Ahmad, H.; Shah, L.A.; et al. Review on Methylene Blue: Its Properties, Uses, Toxicity and Photodegradation. Water 2022, 14, 242. [Google Scholar] [CrossRef]
  29. Grasso, D.; Chin, Y.P.; Weber, W.J. Structural and behavioral characteristics of a commercial humic acid and natural dissolved aquatic organic matter. Chemosphere 1990, 21, 1181–1197. [Google Scholar] [CrossRef]
  30. Xu, P.; Huang, X.; Pan, X.; Li, N.; Zhu, J.; Zhu, X. Hyperbranched Polycaprolactone through RAFT Polymerization of 2-Methylene-1,3-dioxepane. Polymers 2019, 11, 318. [Google Scholar] [CrossRef]
  31. Chang, Y.T.; Lee, C.H.; Hsieh, C.Y.; Chen, T.C.; Jien, S.H. Using Fluorescence Spectroscopy to Assess Compost Maturity Degree during Composting. Agronomy 2023, 13, 1870. [Google Scholar] [CrossRef]
  32. Kloster, N.; Brigante, M.; Zanini, G.; Avena, M. Aggregation kinetics of humic acids in the presence of calcium ions. Colloids Surf. A Physicochem. Eng. Asp. 2013, 427, 76–82. [Google Scholar] [CrossRef]
  33. Doskočil, L.; Burdíková-Szewieczková, J.; Enev, V.; Kalina, L.; Wasserbauer, J. Spectral characterization and comparison of humic acids isolated from some European lignites. Fuel 2018, 213, 123–132. [Google Scholar] [CrossRef]
  34. Kang, K.H.; Shin, H.S.; Park, H. Characterization of humic substances present in landfill leachates with different landfill ages and its implications. Water Res. 2002, 36, 4023–4032. [Google Scholar] [CrossRef] [PubMed]
  35. Zhang, S.; Yuan, L.; Li, Y.; Zhao, B. Characterization of ultrafiltration-fractionated HUMIC acid derived from Chinese weathered coal by spectroscopic techniques. Agronomy 2021, 11, 2030. [Google Scholar] [CrossRef]
  36. Salvestrini, S. Analysis of the Langmuir rate equation in its differential and integrated form for adsorption processes and a comparison with the pseudo first and pseudo second order models. React. Kinet. Mech. Catal. 2018, 123, 455–472. [Google Scholar] [CrossRef]
  37. 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]
  38. Crank, J. The Mathematics of Diffusion; Oxford University Press: Oxford, UK, 1979. [Google Scholar]
  39. Vermeulen, T. Theory for Irreversible and Constant-Pattern Solid Diffusion. Ind. Eng. Chem. 2002, 45, 1664–1670. [Google Scholar] [CrossRef]
  40. Magesh, N.; Renita, A.A.; Siva, R.; Harirajan, N.; Santhosh, A. Adsorption behavior of fluoroquinolone(ciprofloxacin) using zinc oxide impregnated activated carbon prepared from jack fruit peel: Kinetics and isotherm studies. Chemosphere 2022, 290, 133227. [Google Scholar] [CrossRef] [PubMed]
  41. Gamoudi, S.; Srasra, E. Adsorption of organic dyes by HDPy+-modified clay: Effect of molecular structure on the adsorption. J. Mol. Struct. 2019, 1193, 522–531. [Google Scholar] [CrossRef]
  42. Mashentseva, A.A.; Aimanova, N.A.; Parmanbek, N.; Temirgaziyev, B.S.; Barsbay, M.; Zdorovets, M.V. Serratula coronata L. Mediated Synthesis of ZnO Nanoparticles and Their Application for the Removal of Alizarin Yellow R by Photocatalytic Degradation and Adsorption. Nanomaterials 2022, 12, 3293. [Google Scholar] [CrossRef]
  43. Fiol, N.; Villaescusa, I. Determination of sorbent point zero charge: Usefulness in sorption studies. Environ. Chem. Lett. 2009, 7, 79–84. [Google Scholar] [CrossRef]
  44. Peng, L.; Qin, P.; Lei, M.; Zeng, Q.; Song, H.; Yang, J.; Shao, J.; Liao, B.; Gu, J. Modifying Fe3O4 nanoparticles with humic acid for removal of Rhodamine B in water. J. Hazard. Mater. 2012, 209–210, 193–198. [Google Scholar] [CrossRef]
  45. Chen, Q.; Yin, D.; Zhu, S.; Hu, X. Adsorption of cadmium(II) on humic acid coated titanium dioxide. J. Colloid Interface Sci. 2012, 367, 241–248. [Google Scholar] [CrossRef] [PubMed]
  46. Iovino, P.; Leone, V.; Salvestrini, S.; Capasso, S. Sorption of non-ionic organic pollutants onto immobilized humic acid. Desalin. Water Treat. 2015, 56, 55–62. [Google Scholar] [CrossRef]
  47. Seki, H.; Suzuki, A. Adsorption of heavy metal ions onto insolubilized humic acid. J. Colloid Interface Sci. 1995, 171, 490–494. [Google Scholar] [CrossRef]
  48. Do, T.H.; Nguyen, V.T.; Dung, N.Q.; Chu, M.N.; Van Kiet, D.; Ngan, T.T.K.; Van Tan, L. Study on methylene blue adsorption of activated carbon made from Moringa oleifera leaf. Mater. Today Proc. 2021, 38, 3405–3413. [Google Scholar] [CrossRef]
  49. Yao, X.; Ji, L.; Guo, J.; Ge, S.; Lu, W.; Cai, L.; Wang, Y.; Song, W.; Zhang, H. Magnetic activated biochar nanocomposites derived from wakame and its application in methylene blue adsorption. Bioresour. Technol. 2020, 302, 122842. [Google Scholar] [CrossRef] [PubMed]
  50. Jiang, L.; Wen, Y.; Zhu, Z.; Liu, X.; Shao, W. A Double cross-linked strategy to construct graphene aerogels with highly efficient methylene blue adsorption performance. Chemosphere 2021, 265, 129169. [Google Scholar] [CrossRef]
  51. Yan, Y.; Zhang, M.; Gong, K.; Su, L.; Guo, Z.; Mao, L. Adsorption of methylene blue dye onto carbon nanotubes: A route to an electrochemically functional nanostructure and its layer-by-layer assembled nanocomposite. Chem. Mater. 2005, 17, 3457–3463. [Google Scholar] [CrossRef]
  52. Zhang, G.; Shuang, S.; Dong, C.; Pan, J. Study on the interaction of methylene blue with cyclodextrin derivatives by absorption and fluorescence spectroscopy. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2003, 59, 2935–2941. [Google Scholar] [CrossRef]
  53. Langmuir, I. The adsorption of gases on plane surfaces of glass, mica and platinum. J. Am. Chem. Soc. 1918, 40, 1361–1403. [Google Scholar] [CrossRef]
  54. Freundlich, H. Über die Adsorption in Lösungen. Zeitschrift Phys. Chemie 1907, 57U, 385–470. [Google Scholar] [CrossRef]
  55. Vigdorowitsch, M.; Pchelintsev, A.; Tsygankova, L.; Tanygina, E. Freundlich isotherm: An adsorption model complete framework. Appl. Sci. 2021, 11, 8078. [Google Scholar] [CrossRef]
  56. Debord, J.; Chu, K.H.; Harel, M.; Salvestrini, S.; Bollinger, J.C. Yesterday, Today, and Tomorrow. Evolution of a Sleeping Beauty: The Freundlich Isotherm. Langmuir 2023, 39, 3062–3071. [Google Scholar] [CrossRef]
  57. Bouchelkia, N.; Tahraoui, H.; Amrane, A.; Belkacemi, H.; Bollinger, J.C.; Bouzaza, A.; Zoukel, A.; Zhang, J.; Mouni, L. Jujube stones based highly efficient activated carbon for methylene blue adsorption: Kinetics and isotherms modeling, thermodynamics and mechanism study, optimization via response surface methodology and machine learning approaches. Process Saf. Environ. Prot. 2023, 170, 513–535. [Google Scholar] [CrossRef]
  58. Mouni, L.; Belkhiri, L.; Bollinger, J.C.; Bouzaza, A.; Assadi, A.; Tirri, A.; Dahmoune, F.; Madani, K.; Remini, H. Removal of Methylene Blue from aqueous solutions by adsorption on Kaolin: Kinetic and equilibrium studies. Appl. Clay Sci. 2018, 153, 38–45. [Google Scholar] [CrossRef]
  59. Cheikh, S.; Bollinger, J.-C.; Belkhiri, L.; Tiri, A.; Bouzaza, A.; El, J.; Assadi, A.; Amrane, A.; Mouni, A.; Waste, L.Z.; et al. Zeolite Waste Characterization and Use as Low-Cost, Ecofriendly, and Sustainable Material for Malachite Green and Methylene Blue Dyes Removal: Box–Behnken Design, Kinetics, and Thermodynamics. Appl. Sci. 2022, 12, 7587. [Google Scholar] [CrossRef]
  60. Yao, Y.; Xu, F.; Chen, M.; Xu, Z.; Zhu, Z. Adsorption behavior of methylene blue on carbon nanotubes. Bioresour. Technol. 2010, 101, 3040–3046. [Google Scholar] [CrossRef]
  61. Manna, S.; Roy, D.; Saha, P.; Gopakumar, D.; Thomas, S. Rapid methylene blue adsorption using modified lignocellulosic materials. Process Saf. Environ. Prot. 2017, 107, 346–356. [Google Scholar] [CrossRef]
  62. Banat, F.; Al-Asheh, S.; Al-Makhadmeh, L. Evaluation of the use of raw and activated date pits as potential adsorbents for dye containing waters. Process Biochem. 2003, 39, 193–202. [Google Scholar] [CrossRef]
  63. Yan, H.; Tao, X.; Yang, Z.; Li, K.; Yang, H.; Li, A.; Cheng, R. Effects of the oxidation degree of graphene oxide on the adsorption of methylene blue. J. Hazard. Mater. 2014, 268, 191–198. [Google Scholar] [CrossRef]
  64. Juzsakova, T.; Salman, A.D.; Abdullah, T.A.; Rasheed, R.T.; Zsirka, B.; Al-Shaikhly, R.R.; Sluser, B.; Cretescu, I. Removal of Methylene Blue from Aqueous Solution by Mixture of Reused Silica Gel Desiccant and Natural Sand or Eggshell Waste. Materials 2023, 16, 1618. [Google Scholar] [CrossRef]
  65. Bestani, B.; Benderdouche, N.; Benstaali, B.; Belhakem, M.; Addou, A. Methylene blue and iodine adsorption onto an activated desert plant. Bioresour. Technol. 2008, 99, 8441–8444. [Google Scholar] [CrossRef] [PubMed]
  66. Kuang, Y.; Zhang, X.; Zhou, S. Adsorption of Methylene Blue in Water onto Activated Carbon by Surfactant Modification. Water 2020, 12, 587. [Google Scholar] [CrossRef]
  67. Zhao, H.; Zhong, H.; Jiang, Y.; Li, H.; Tang, P.; Li, D.; Feng, Y. Porous ZnCl2-Activated Carbon from Shaddock Peel: Methylene Blue Adsorption Behavior. Materials 2022, 15, 895. [Google Scholar] [CrossRef]
  68. Salvestrini, S.; Ambrosone, L.; Kopinke, F.D. Some mistakes and misinterpretations in the analysis of thermodynamic adsorption data. J. Mol. Liq. 2022, 352, 118762. [Google Scholar] [CrossRef]
  69. Hong, S.; Wen, C.; He, J.; Gan, F.; Ho, Y.S. Adsorption thermodynamics of Methylene Blue onto bentonite. J. Hazard. Mater. 2009, 167, 630–633. [Google Scholar] [CrossRef]
  70. Ahmad, M.A.; Mohamad Yusop, M.F.; Zakaria, R.; Karim, J.; Yahaya, N.K.E.M.; Mohamed Yusoff, M.A.; Hashim, N.H.F.; Abdullah, N.S. Adsorption of methylene blue from aqueous solution by peanut shell based activated carbon. Mater. Today Proc. 2021, 47, 1246–1251. [Google Scholar] [CrossRef]
  71. Cottet, L.; Almeida, C.A.P.; Naidek, N.; Viante, M.F.; Lopes, M.C.; Debacher, N.A. Adsorption characteristics of montmorillonite clay modified with iron oxide with respect to methylene blue in aqueous media. Appl. Clay Sci. 2014, 95, 25–31. [Google Scholar] [CrossRef]
  72. Jia, P.; Tan, H.; Liu, K.; Gao, W. Removal of Methylene Blue from Aqueous Solution by Bone Char. Appl. Sci. 2018, 8, 1903. [Google Scholar] [CrossRef]
  73. Ren, Y.; Cui, C.; Wang, P. Pomelo Peel Modified with Citrate as a Sustainable Adsorbent for Removal of Methylene Blue from Aqueous Solution. Molecules 2018, 23, 1342. [Google Scholar] [CrossRef] [PubMed]
  74. Zuhara, S.; Pradhan, S.; Zakaria, Y.; Shetty, A.R.; McKay, G. Removal of Methylene Blue from Water Using Magnetic GTL-Derived Biosolids: Study of Adsorption Isotherms and Kinetic Models. Molecules 2023, 28, 1511. [Google Scholar] [CrossRef] [PubMed]
  75. Atkins, P.; de Paula, J.; Keeler, J. Atkins’ Physical Chemistry; OUP Oxford: Oxford, UK, 2017; ISBN 978-0198769866. [Google Scholar]
  76. Tran, H.N.; Bollinger, J.-C.; Salvestrini, S.; Chu, K.H.; Juang, R.-S. Critical Review and Discussion of the Nonlinear Form of Radke–Prausnitz Model in Adsorption Solid–Liquid Phases. J. Environ. Eng. 2022, 149, 03122006. [Google Scholar] [CrossRef]
  77. Salvestrini, S. New insights into the interaction mechanism of humic acids with phillipsite. React. Kinet. Mech. Catal. 2017, 120, 735–752. [Google Scholar] [CrossRef]
  78. Cooney, D.O. Adsorption Design for Wastewater Treatment; Lewis Publishers: Ririe, ID, USA, 1999; ISBN 1566703336. [Google Scholar]
  79. Colella, A.; de Gennaro, B.; Salvestrini, S.; Colella, C. Surface interaction of humic acids with natural and synthetic phillipsite. J. Porous Mater. 2015, 22, 501–509. [Google Scholar] [CrossRef]
  80. Csicsor, A.; Tombácz, E. Screening of Humic Substances Extracted from Leonardite for Free Radical Scavenging Activity Using DPPH Method. Molecules 2022, 27, 6334. [Google Scholar] [CrossRef]
  81. Kalmár, J.; Lente, G.; Fábián, I. Kinetics and mechanism of the adsorption of methylene blue from aqueous solution on the surface of a quartz cuvette by on-line UV–Vis spectrophotometry. Dye. Pigment. 2016, 127, 170–178. [Google Scholar] [CrossRef]
  82. Dinh, N.T.; Vo, L.N.H.; Tran, N.T.T.; Phan, T.D.; Nguyen, D.B. Enhancing the removal efficiency of methylene blue in water by fly ash via a modified adsorbent with alkaline thermal hydrolysis treatment. RSC Adv. 2021, 11, 20292–20302. [Google Scholar] [CrossRef]
  83. Sawafta, R.; Shahwan, T. A comparative study of the removal of methylene blue by iron nanoparticles from water and water-ethanol solutions. J. Mol. Liq. 2019, 273, 274–281. [Google Scholar] [CrossRef]
  84. Fernández-Pérez, A.; Valdés-Solís, T.; Marbán, G. Visible light spectroscopic analysis of Methylene Blue in water; the resonance virtual equilibrium hypothesis. Dye. Pigment. 2019, 161, 448–456. [Google Scholar] [CrossRef]
  85. Spencer, W.; Sutter, J.R. Kinetic study of the monomer-dimer equilibrium of methylene blue in aqueous solution. J. Phys. Chem. 1979, 83, 1573–1576. [Google Scholar] [CrossRef]
Figure 1. High-pressure gel permeation chromatograms of extracted HA samples: (A) unfractionated HA; (B) fraction of HA soluble at pH 3; (C) fraction of HA insoluble at pH 3 and dissolved at pH 5; (D) residual fraction of HA insoluble at pH 5 and dissolved at pH 7.
Figure 1. High-pressure gel permeation chromatograms of extracted HA samples: (A) unfractionated HA; (B) fraction of HA soluble at pH 3; (C) fraction of HA insoluble at pH 3 and dissolved at pH 5; (D) residual fraction of HA insoluble at pH 5 and dissolved at pH 7.
Molecules 28 07083 g001
Figure 2. Dependence of the MB aqueous concentration on time in the presence of NYT; C0 = initial MB concentration in the liquid phase; T = 20 °C. The curves in the figure were obtained via non-linear regression of the data using the Vermeulen model.
Figure 2. Dependence of the MB aqueous concentration on time in the presence of NYT; C0 = initial MB concentration in the liquid phase; T = 20 °C. The curves in the figure were obtained via non-linear regression of the data using the Vermeulen model.
Molecules 28 07083 g002
Figure 3. Comparison of the sorption kinetics regarding the sorption of MB onto various sorbents. Initial concentration of MB = 11.9 mg L−1; T = 20 °C. The curves in the figure were obtained via nonlinear regression analysis using the PSO model, with the exception of the NYT data set modelled using the Vermeulen equation.
Figure 3. Comparison of the sorption kinetics regarding the sorption of MB onto various sorbents. Initial concentration of MB = 11.9 mg L−1; T = 20 °C. The curves in the figure were obtained via nonlinear regression analysis using the PSO model, with the exception of the NYT data set modelled using the Vermeulen equation.
Molecules 28 07083 g003
Figure 4. Sorption isotherms regarding the sorption of MB onto various sorbents. T = 20 °C. The curves in the figure were obtained via nonlinear regression analysis using the Freundlich model, with the exception of NYT data set modelled using the Langmuir equation.
Figure 4. Sorption isotherms regarding the sorption of MB onto various sorbents. T = 20 °C. The curves in the figure were obtained via nonlinear regression analysis using the Freundlich model, with the exception of NYT data set modelled using the Langmuir equation.
Molecules 28 07083 g004
Figure 5. Sorption isotherms regarding the sorption of MB onto NYT-HA7 at various temperatures. The curves in the figure were obtained via non-linear regression analysis using the Freundlich isotherm model.
Figure 5. Sorption isotherms regarding the sorption of MB onto NYT-HA7 at various temperatures. The curves in the figure were obtained via non-linear regression analysis using the Freundlich isotherm model.
Molecules 28 07083 g005
Figure 6. Plot of 1/n against 1/T for the estimation of the standard sorption enthalpy ( Δ H ° ) and the standard sorption entropy ( Δ S ° ).
Figure 6. Plot of 1/n against 1/T for the estimation of the standard sorption enthalpy ( Δ H ° ) and the standard sorption entropy ( Δ S ° ).
Molecules 28 07083 g006
Table 1. Optical properties of HA samples as determined via the E2/E3 and E4/E6 indices.
Table 1. Optical properties of HA samples as determined via the E2/E3 and E4/E6 indices.
HA TypeE2/E3E4/E6
HAref1.74.0
HA31.94.4
HA51.84.1
HA71.63.6
Table 2. Kinetic parameters for the sorption of MB onto NYT. Initial aqueous concentration of MB = 11.9 mg L−1; sorbent dosage = 0.29 g L−1; pH = 7.4; T = 20 °C (additional results can be found in Table S1).
Table 2. Kinetic parameters for the sorption of MB onto NYT. Initial aqueous concentration of MB = 11.9 mg L−1; sorbent dosage = 0.29 g L−1; pH = 7.4; T = 20 °C (additional results can be found in Table S1).
Modelk1
(h−1)
k2
(g mg−1
h−1)
B
(h−1)
kD
(h−1)
Res. Sum of
Squares
AICc
PFO0.023
±
0.003
---1.087−8.23
PSO-0.0027
±
0.0005
--0.533−14.63
Boyd--0.0036
±
0.0009
-0.093−30.36
Vermeulen---0.0048
±
0.0008
0.094−30.25
Table 3. Kinetic parameters for the sorption of MB onto NYT-HA7: initial aqueous concentration of MB = 11.9 mg L−1; sorbent dosage = 0.29 g L−1; pH = 7.4; T = 20 °C (additional results are reported in Table S2).
Table 3. Kinetic parameters for the sorption of MB onto NYT-HA7: initial aqueous concentration of MB = 11.9 mg L−1; sorbent dosage = 0.29 g L−1; pH = 7.4; T = 20 °C (additional results are reported in Table S2).
Modelk1
(h−1)
k2
(g mg−1
h−1)
B
(h−1)
kD
(h−1)
Res. Sum of
Squares
AICc
PFO0.108
±
0.008
---1.24123.03
PSO-0.0066
±
0.0003
--0.0859.65
Boyd--0.077
±
0.006
-0.72820.36
Vermeulen---0.070
±
0.005
0.61619.53
Table 4. Points of zero charge (pHPZC) of the sorbent materials.
Table 4. Points of zero charge (pHPZC) of the sorbent materials.
SorbentpHPZC
NYT7.2
NYT-HAref5.0
NYT-HA34.1
NYT-HA54.9
NYT-HA74.8
Table 5. Isotherm sorption parameters for the uptake of MB onto NYT and NYT-HA at 20 °C.
Table 5. Isotherm sorption parameters for the uptake of MB onto NYT and NYT-HA at 20 °C.
Sorbent q m
(mg g−1)
K L
(L mg−1)
Res.
Sum of
Squares
AICc K F
(mg1-n g−1 Ln)
n Res.
Sum of
Squares
AICc
NYT10.9
±
0.5
4
±
1
2.93013.708
±
1
0.18
±
0.08
11.78122.04
NYT-HAref70
±
10
0.14
±
0.03
1.79410.768.8
±
0.1
0.79
±
0.01
0.190−2.73
NYT-HA344
±
8
0.06
±
0.02
0.4772.812.52
±
0.07
0.84
±
0.02
0.078−8.08
NYT-HA570
±
10
0.10
±
0.02
1.1598.146.56
±
0.08
0.79
±
0.01
0.076−8.19
NYT-HA7110
±
10
0.5
±
0.1
1.3198.9136.7
±
0.2
0.77
±
0.01
0.255−0.95
Table 6. Predicted sorption capacity (qe) of various sorbents for MB in correspondence with an equilibrium aqueous concentration of MB equal to 1.0 mg L−1.
Table 6. Predicted sorption capacity (qe) of various sorbents for MB in correspondence with an equilibrium aqueous concentration of MB equal to 1.0 mg L−1.
Sorbent P r e d i c t e d
q e
(mg g−1)
pH aSorbent
Dosage (g L−1)
T
(K)
Fitting ModelRef.
NYT-HA7377.4 (buff.)0.3293FreundlichThis work
Jujube-stone-based activated carbon237.01298Langmuir[57]
Kaolin192.00.5298Langmuir[58]
Zeolite waste8free1298Langmuir[59]
Carbon nanotubes87.00.3298Langmuir[60]
Modified lignocellulosic materials396.010298Langmuir[61]
Row date pits118.05298Langmuir[62]
Graphene oxide1917.00.5298Langmuir[63]
Silica gel/eggshell powder167.00.25298Freundlich[64]
Magnetic activated biochar nanocomposites derived from wakame183n.d.1293Langmuir[49]
Commercial activated carbon246.94297Langmuir[65]
Surfactant-modified activated carbon1065.00.15298Langmuir[66]
ZnCl2-Activated Carbon642free0.5303Langmuir[67]
a (buff.) = buffered solution; free = natural pH of the mixture, not adjusted; n.d. = no data available.
Table 7. Freundlich isotherm and thermodynamic parameters for the sorption of MB onto NYT-HA7 at various temperatures.
Table 7. Freundlich isotherm and thermodynamic parameters for the sorption of MB onto NYT-HA7 at various temperatures.
T
(K)
K F
(mol1−n kg−1 Ln)
n Δ G ° 0
(kJ mol−1)
Δ H °
(kJ mol−1)
Δ S °
(J K−1 mol−1)
2931600
±
300
0.75
±
0.01
−3.25
±
0.04
3.9
±
0.6
24
±
2
3001600
±
400
0.74
±
0.02
−3.37
±
0.09
3071400
±
600
0.71
±
0.03
−3.6
±
0.2
3131200
±
500
0.70
±
0.03
−3.7
±
0.2
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

Salvestrini, S.; Debord, J.; Bollinger, J.-C. Enhanced Sorption Performance of Natural Zeolites Modified with pH-Fractionated Humic Acids for the Removal of Methylene Blue from Water. Molecules 2023, 28, 7083. https://doi.org/10.3390/molecules28207083

AMA Style

Salvestrini S, Debord J, Bollinger J-C. Enhanced Sorption Performance of Natural Zeolites Modified with pH-Fractionated Humic Acids for the Removal of Methylene Blue from Water. Molecules. 2023; 28(20):7083. https://doi.org/10.3390/molecules28207083

Chicago/Turabian Style

Salvestrini, Stefano, Jean Debord, and Jean-Claude Bollinger. 2023. "Enhanced Sorption Performance of Natural Zeolites Modified with pH-Fractionated Humic Acids for the Removal of Methylene Blue from Water" Molecules 28, no. 20: 7083. https://doi.org/10.3390/molecules28207083

Article Metrics

Back to TopTop