Next Article in Journal
The Coriolis Effect on Thermal Convection in a Rotating Sparsely Packed Porous Layer in Presence of Cross-Diffusion
Previous Article in Journal
Research Progress of Biomimetic Memristor Flexible Synapse
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Outstanding Performance of a New Exfoliated Clay Impregnated with Rutile TiO2 Nanoparticles Composite for Dyes Adsorption: Experimental and Theoretical Studies

1
Department of Chemistry, Faculty of Science, King Khalid University, Abha 62224, Saudi Arabia
2
Faculty of Earth Science, Beni-Suef University, Beni Suef 62511, Egypt
3
School of Engineering, Federal University of Rio Grande do Sul (UFRGS), Av. Bento Gonçalves 9500, Porto Alegre 91501-970, RS, Brazil
4
Institute of Chemistry, Federal University of Rio Grande do Sul (UFRGS), Av. Bento Goncalves 9500, Porto Alegre 91501-970, RS, Brazil
5
Laboratoire de Physique et Chimie Théoriques, UMR 7019—CNRS, Université de Lorraine, 54500 Vandoeuvre-lès-Nancy, France
6
College of Life Sciences, College of Chemistry and Chemical Engineering, State Key Laboratory of Bio-Fibers and Eco-Textiles, Institute of Biomedical Engineering, Qingdao University, Qingdao 266071, China
7
Instituto Tecnológico de Aguascalientes, Aguascalientes 20256, Mexico
8
Department of Agriculture, University of Ioannina, UoI Kostakii Campus, 47040 Arta, Greece
*
Authors to whom correspondence should be addressed.
Coatings 2022, 12(1), 22; https://doi.org/10.3390/coatings12010022
Submission received: 3 December 2021 / Revised: 14 December 2021 / Accepted: 21 December 2021 / Published: 27 December 2021

Abstract

:
Pure rutile TiO2 nanoparticles (Rt) were combined with exfoliated black clay (BC) to prepare a new composite for water decontamination, in particular, for the uptake of methylene blue (MB) and methyl orange (MO) dyes. The as-prepared Rt/BC was characterized by X-ray diffraction (XRD), Fourier transform infrared spectroscopy (FTIR), and Field emission scanning electron microscopy (FESEM) techniques, and the dyes’ adsorption isotherms at three temperatures (i.e., 25, 40, and 50 °C) were studied. The results indicated that Rt/BC displayed a high removal performance for MO (96.7%) and MB (91.4%) at pH 3.0 and 8.0, respectively. Adsorption data of MB and MO were adjusted by a double layer model at all temperatures. The theoretical parameters of this statistical physics model were interpreted to understand the MO and MB adsorption mechanisms at the molecular level. The removed molecules per active site (n) of Rt/BC ranged from 1.12 to 1.29 for MB and 1.47 to 1.85 for MO, thus representing parallel orientation and multi-interactions mechanisms (i.e., van der Waals forces, hydrogen bonding, and electrostatic interactions were involved). The Rt/BC composite had a density of surface adsorption sites of 100 mg/g. The aggregation of MO molecules was high and increased their adsorption capacities (Qsat = 294–370 mg/g) compared to that of MB (Qsat = 214–249 mg/g). Adsorption energies were 9.70–20.15 kJ/mol, and these values indicated that MO and MB adsorption processes were endothermic and occurred via physical interactions. Overall, the low cost, high regeneration performance, and stability of Rt/BC support its application as a promising adsorbent for organic pollutants from wastewaters.

Graphical Abstract

1. Introduction

Different industrial activities, such as textiles, printing, food, and cosmetics, use organic dyes as coloring materials to supply products with distinctive colors [1,2,3,4]. In particular, anionic methyl orange (MO) and cationic methylene blue (MB) are documented as stable dyes under a wide range of light, heat, and chemical reagents, and thus, MO and MB are involved in varied industries [1,5,6]. The discharge of effluents containing MB and MO into water resources without any treatment generates ecological impacts and human health risks. Accordingly, it is critical to find an effective manner to clean wastewaters containing MB and MO for attaining a healthy natural environment.
Numerous procedures, including biological treatment, advanced oxidation, adsorption, coagulation, ultrafiltration, or precipitation, were used to remove organic dyes from aqueous solutions [7,8,9]. The adsorption method is most common and preferred in water remediation due to its simple design, high efficiency, non-toxicity, and low cost [10,11,12,13,14,15]. However, the high cost of various adsorbents (e.g., carbon nanotubes and graphene oxide) usually inhibits their broad applications, and consequently, finding an eco-friendly adsorbent with high removal efficiency and low cost has become a required target.
Clays as accessible natural materials are characterized by high thermal and chemical stabilities, non-toxicity, and low cost [5]. Thus, clay-based adsorbents, such as smectite [16], kaolin [17], red mud [18], fibrous clay minerals [19], and Fe3O4/serpentine [6], were utilized to MB uptake from contaminated water. Conversely, activated volcanic mud [7], bentonite-supported zero-valent iron [20], and surfactant-modified clay [5] were utilized for the uptake of MO dye. Furthermore, the activation of black clay using hydrogen peroxide resulted in being an adequate strategy to facilitate the insertion of different organic and/or inorganic modifiers in its structure, generating new effective adsorbents [5,21]. Moreover, TiO2 nanoparticles are widely employed in decreasing the concentrations of dyes through degradation and adsorption techniques because of their chemical inertness, non-toxicity, and high capability of interaction with organics in waters [21,22,23,24,25,26,27]. Furthermore, the utilized TiO2 nanoparticles were related to rutile (Rt) extracted with high purity by thermo-chemical activation of ilmenite ore (FeTiO3) [28]. Consequently, the interface between the treated black clay and Rt nanoparticles is promising to prepare a new eco-friendly composite with abundant active adsorption sites for different environmental applications.
Analysis of the uptake data via equilibrium models is mandatory in understanding the physicochemical parameters that could control the adsorption mechanism [1,9,12,13,29,30,31]. Generally, the adsorbents-adsorbates interactions are commonly considered by applying classical models such as Langmuir, Freundlich, Sips, and others [1]. However, the parameters of these traditional models cannot offer a complete description and understanding of the adsorption mechanism of the adsorption system at hand [1,13,31]. For instance, the traditional classical isotherm models are inadequate to define the horizontal or vertical geometry of the removed chemical species [12,30].
Conversely, the energetic and steric parameters related to the advanced statistical physics models (ASPMs) can delineate the adsorption interactions between the investigated dyes and the adsorbent surface at a molecular scale [9,13]. Furthermore, the adsorption modeling with ASPMs can provide different parameters related to the removal mechanism, such as the number of molecules removed by one active site of the adsorbent (n), the active sites density of the adsorbent (DM), the number of the removed layers of the adsorbate (Nt) and the adsorption capacity at saturation (Qsat). Additionally, the adsorption energy (ΔE) can be determined via the ASPMs [1,13,31]. Therefore, in-depth and interesting results can be obtained from applying these ASPMs to fit the experimental data. The novelty of this research article was to prepare a promising adsorbent for the removal of MB and MO dyes via the combination between rutile (TiO2) nanoparticles (Rt) and H2O2–activated black clay (BC). Numerous techniques (e.g., XRD, FTIR, and FESEM) were used to study the physicochemical characteristics of this Rt/BC composite. In addition, the steric and energetic parameters from the advanced statistical physics models were employed to clarify the performance of Rt/BC for MB and MO dyes’ removal and provide new insights on the adsorption mechanisms considering the molecular level.

2. Experimental

2.1. Reactants and Clay

The used black clay (BC) sample was collected from the Qalmasha area in Fayum governorate, Egypt. A mass (200 g) of the representative crushed BC was ground to obtain a sample size of less than 100 µm. Rutile nanoparticles (Rt) were attained via a thermo-chemical process, with at least 96% purity [28]. The tested adsorbates were: Methylene blue (MB, C16H18ClN3S, λmax 664 nm, 319.85 g/mol (Merck, Darmstadt, Germany), methyl orange (MO, C14H14N3NaO3S, λmax 464 nm, 327.33 g/mol (Merck, Darmstadt, Germany). In addition, the next reactants were employed: H2O2 (30%), aqueous solutions of HCl, and NH4OH. The molecular structures of studied dyes (MB and MO) are shown in Scheme 1.

2.2. Preparation of Rt/BC Adsorbent

The following process was used to modify the investigated clay with H2O2: 3 g of the BC was added to a beaker containing 30 mL of deionized water with continuous stirring for 60 min. Then, 20 mL of 30% H2O2 was added to this BC/H2O mixture, stirring for 120 min at 40 °C. First, H2O2 was used in oxidizing the organic carbon of the black clay, thus producing purified and exfoliated BC with open pits and cavities. Then, 1.0 g of the white Rt nanoparticles powder was introduced to the BC slurry, and the contents were magnetically agitated for 3 h at 45 °C. Subsequently, centrifugation at 6000 rpm was used to separate the solid phase from the liquid phase. Next, the material was washed with distilled water before drying for 24 h at 65 °C. Finally, the dried powder Rt/BC adsorbent was homogenized using a mortar.

2.3. Adsorbent Characterization

The Rt/BC adsorbent structure was determined by X-ray diffraction analysis (XRD) in the 2θ range of 5°–80° using a Philips diffractometer (APD-3720, Amsterdam, The Netherlands).
The functional groups related to the Rt/BC composite surface were identified in the 400–4000 cm−1 range via Fourier transmission infrared spectroscopy (FTIR) analysis using a Bruker spectrophotometer (Billerica, MA, USA).
Field-emission scanning electron microscopy (FESEM, Carl Zeiss, Jena, Germany) images of the Rt/BC material were recorded using a Sigma 500 VP, FESEM microscope.
The pH of the point of zero charge (pHPZC) was determined as follows [9]: 50 mL of 0.1 M KCl was taken in a clean glass Erlenmeyer. The initial pHs (pHi) of these solutions were adjusted at pH values ranging from 2.0 to 10.0. Then, an amount of 100 mg of Rt/BC material was introduced into 50 mL of 0.1 M KCl with previously adjusted pH values, and the slurries were stirred at 150 rpm for 24 h. Finally, the final solution pH (pHf) was determined, and the relation between pHf–pHi versus pHi was used to determine the pHPZC (i.e., ΔpH = 0).

2.4. Isotherm Studies of MB and MO Adsorption

Standard 25–250 mg/L MB and MO solutions were prepared from stock dye solutions (1000 mg/L). Isotherm studies of MB and MO adsorption were performed at pH 8.0 and 3.0, respectively, and three temperatures (25, 40, and 50 °C) using 25 mg of Rt/BC adsorbent. In the equilibrium experiments, the dye–Rt/BC suspensions were shaken at 150 rpm for 4 h using an orbital shaker. The following equation was used to obtain the dye adsorption capacities ( q e ,   mg/g) at equilibrium.
q e = ( C 0 C e )   V   m
where m is the Rt/BC material mass (g), V is the dye (MB or MO) solution volume (L),   C e and C 0 (mg/L) are the equilibrium and initial dye concentrations, respectively. All MO and MB removal experiments were performed by duplicate, and the results were averaged for data analysis finding standard deviations less than 5.0%.

2.5. MB and MO Adsorption Modeling

This research paper applied different advanced statistical physics models (i.e., monolayer, double layer, and multilayer) to assess the interactions between dyes and Rt/BC. The appropriateness of these advanced models to fit MB and MO adsorption data was compared, and the best one was identified according to the determined coefficient R 2 values [13]. These advanced models are described as follows [9,13,31]:
  • Advanced monolayer model (AMM)
In the AMM, the dyes adsorption on Rt/BC results in the formation of a single layer linked to a given interaction energy (ΔE) (i.e., Rt/BC adsorption sites have the same energy).
Q e = n D M 1 + ( C 1 / 2 c ) n     ( advanced   monolayer   model )
where C1/2 is the concentration at half-saturation of the removed MB or MO layer.
  • Advanced double layer model (ADM)
This model suggests that the removal of the tested dyes occurs due to the formation of two layers with two dissimilar adsorption energies (i.e., ΔE1 for dye–Rt/BC interaction and ΔE2 for dye–dye interface).
Q e = n D M ( c c 1 ) n + 2 ( c c 2 ) 2 n 1 + ( c c 1 ) n + ( c c 2 ) 2 n   ( advanced   double   model )
For ADM, two layers of each investigated dye were involved with two dye concentrations at half-saturation that were recognized as C1 and C2, respectively.
  • A multilayer model, as the general form of ASPMs, was also implemented in this study to fit the MB and MO adsorption processes. The following equations (Equations (4)–(9) are used to determine the parameters of the multilayer model [9]:
    Q = n   D M F 1 ( c ) + F 2 ( c ) + F 3 ( c ) + F 4 ( c ) G ( c )      
    F 1 ( c ) = 2 ( c c 1 ) 2 n 1 ( c c 1 ) n + ( c c 1 ) n ( 1 ( c c 1 ) 2 n ) ( 1 ( c c 1 ) n ) 2
    F 2 ( c ) = 2 ( c c 1 ) n ( c c 2 ) n ( 1 ( c c 2 ) n   N 2 ) 1 ( c c 2 ) n
    F 3 ( c ) = N 2 ( c c 1 ) n ( c c 2 ) n ( c c 2 ) n   N 2 1 ( c c 2 ) n
    F 4 ( c ) = ( c c 1 ) n ( c c 2 ) 2 n ( 1 ( c c 2 ) n   N 2 ) ( 1 ( c c 2 ) n ) 2
    G ( c ) = ( 1 ( c c 1 ) 2 n ) 1 ( c c 1 ) n + ( c c 1 ) n ( c c 2 ) n ( 1 ( c c 2 ) n   N 2 ) ( 1 ( c c 2 ) n ) 2
    where the parameter N2 designates the number of removed dye layers with specified energy of adsorption. Numerous situations were anticipated depending on the multilayer adsorption model as given in Figure 1 [13,31,32]:
  • Langmuir adsorption model: n parameter was fixed at 1, and N2 was fixed at 0;
  • Monolayer adsorption model: n was the fitted parameter, and N2 was fixed at 0;
  • Double-layer adsorption model: n was the fitted parameter, and N2 was fixed at 1;
  • Triple-layer adsorption model: n was the fitted parameter, and N2 was fixed at 2;
  • Multilayer adsorption model: n and N2 were the fitted parameters.
The multilayer model suggests that the dye-Rt/BC interaction was related to dissimilar adsorption energies [13]. Thus, the primary energy of adsorption corresponds to the interaction between the first adsorbed dye layer and the active sites of the Rt/BC adsorbent. Conversely, the additional energy was linked to the dye‒dye interaction. Therefore, the whole number of the removed dye layers is calculated by Nt = 1 + N2 [31].

2.6. Regeneration of Rt/BC Adsorbent

In industrial processes, the use of an adsorbent numerous times is considered an essential concern for decreasing the water purification costs [21]. Therefore, in order to test the recycling of Rt/BC adsorbent, the adsorption-desorption cycle was performed 4 times. Firstly, the Rt/BC (50 mg) loaded adsorbent was thoroughly washed using distilled water and subsequently oven-dried for 24 h at 75 °C. Next, the desorption of the studied dyes was performed using 50 mL of 0.5 M NaOH eluent solution at 25 °C. Finally, a rotatory shaker (SHO–2D shaker, Germany) agitated Rt/BC adsorbent loaded with MB and MO molecules for 240 min at 120 rpm.

3. Results and Discussion

3.1. Characterization of the Rt/BC Material

The XRD pattern of Rt/BC composite displays diffracting lines corresponding to black clay and titanium oxide rutile nanoparticles (Figure 2a). Montmorillonite (JCPDS 10-0357), in addition to kaolinite (JCPDS 78-2110), were identified as clay minerals [14]. Quartz (SiO2), with its descriptive diffraction peaks at 3.37 and 4.26 Å (JCPDS 46-1045), was recognized as a common non-clay mineral (Figure 2a). The minor variations in the montmorillonite and kaolinite peaks intensities could be associated with the interface between rutile nanoparticles and H2O2-activated black clay. Moreover, strong peaks were observed in the 2θ range 25°–70° (Figure 2a), which agreed with TiO2 Rt [21]. In particular, the three strong peaks detected at nearly 27°, 36°, and 55° indicated TiO2 in the Rt phase (JCPDS 88–1175) [21].
The vibrational infrared spectrum of the Rt/BC adsorbent displayed strong absorption bands at nearly 3625 and 3406 cm−1 (Figure 2b). These observed bands could be attributed to the stretching vibrations of Al–OH and –OH functional groups [14]. The observed band at 2377 cm−1 could be ascribed to the C≡C of the organic carbon in the studied black clay [33]. The C=O stretching band was observed at 1636 cm−1, and the presence of this functional group in the BC sample could improve its removal efficiency for inorganic/organic compounds [34]. The observed band at 1035 cm−1 was assigned to the stretching vibrations of Si–O, while the detected absorption band at 923 cm−1 was related to the bending vibrations of –OH groups in the Rt/BC composite [9]. Quartz mineral was identified by the vibration band detected at 790 cm−1 [14]. In addition, the broad and intense bands observed at 469, 533, and 679 cm−1 could be associated with the Ti–O and Ti–O–Ti stretching and bending vibrations [14]. These functional groups (i.e., active adsorption sites) agreed with the successful preparation of the Rt/BC composite.
The images of FESEM images of the Rt/BC material (Figure 3) displayed the existence of holes and cavities with different sizes and shapes. The observed cracks and fractures in the Rt/BC could be related to the activation of BC via H2O2 as an initial stage in the preparation of the Rt/BC adsorbent. This H2O2-modification not only resulted in oxidizing the organic matter but also enhanced the separation of the clay minerals into thin sheets, mainly montmorillonite mineral due to its T–O–T structure (i.e., exfoliated clay); see Figure 3a–d. Furthermore, titanium oxide Rt was observed as an aggregate of several nanoparticles with dissimilar sizes below 100 nm (Figure 3e–f). Obviously, the physical combination between TiO2 Rt and BC was clearly observed wherein the Rt nanoparticles were connected to the outer surface of BC clay and filled the cracks and holes of the BC sample (Figure 3a–f). Therefore, the existence of spherical-like Rt nanoparticles could be considered as a support to enhance the uptake of MO and MB species because of the improvement of the surface area of Rt/BC [21].

3.2. pH Effect on MB and MO Adsorption by Rt/BC

The pHPZC value of the Rt/BC composite is an important parameter to study the influence of solution pH in the removal of tested dyes. This pHPZC was 5.4 (Figure 4), and therefore, Rt/BC adsorbent displayed different performances towards MO and MB dyes (i.e., at pH > 5.5, MO uptake% increased, while MB removal % decreased). Therefore, MB and MO dyes uptake by Rt/BC adsorbent as a function of pH is shown in Figure 4. At pH 2.0–4.0, the Rt/BC material was protonated because of the high concentrations of H+ in pH solutions < 4.0 [5], and thus, the MO uptake % increased, presenting the maximum value (96.7%) at pH 3.0; see Figure 4. On the contrary, the MB removal % improved at pH > 6, reaching the maximum removal value (91.4% at pH 8) due to the deprotonation of Rt/BC functional groups, which could be negatively charged at a pH range of 8.0–10.0. Accordingly, the electrostatic forces (attraction or repulsion) between the Rt/BC functional groups (positively or negatively charged) and dyes were the main factors guiding the MO and MB adsorption mechanisms. Nevertheless, the Rt/BC composite can capture MB molecules at pH 2.0–6.0 and MO molecules at pH 6–10, which could be related to the interactions with oxygen-functional groups present on the Rt/BC adsorbent [21]. Regarding the components of the investigated adsorbent (i.e., BC and Rt nanoparticles), the presence of n–π and π–π interactions in black clay resulted in increasing its adsorption capacities as compared to Rt TiO2 nanoparticles because of the involvement of C=O and C=C groups in BC structure; see Figure 4. Therefore, the equilibrium MB and MO uptake experiments were carried out based on the pH results at solutions pH 8.0 and 3.0, respectively.

3.3. MB and MO Adsorption Modeling and Mechanism Interpretation

After checking the advanced statistical physics model parameters, the double layer model was the best choice (i.e., R2 ranged from 0.98 to 0.99) to describe the adsorption of MB and MO on the Rt/BC adsorbent as shown in Figure 5. Therefore, the adsorption mechanisms of MB and MO on Rt/BC were analyzed via this advanced model. Consequently, the removal efficiency and the adsorption mechanisms associated with dyes-Rt/BC interactions were analyzed by considering the steric and energetic fundamentals of the double layer adsorption model.

3.3.1. Steric Parameters

The steric parameter n represents the connected dyes (MB or MO) molecule numbers per functional group (active site) of the Rt/BC adsorbent. This parameter was utilized to recognize MB and MO molecules’ geometry (i.e., horizontal or vertical) on the Rt/BC surface and their aggregation in solutions (i.e., before adsorption). Moreover, the dyes removal mechanism can be clearly identified by determining the value of the n parameter [12,30,31,35]. Thus, this steric parameter is employed to correct the assumption of the classical Langmuir equation, wherein n is equal to unity [12,30]. Overall, numerous scenarios related to the interface geometry between dyes molecules and the Rt/BC active sites can occur [12,30]:
  • In the first scenario: n ≤ 0.5, the molecules of each tested dye were linked to two or more active sites of the Rt/BC adsorbent (i.e., a horizontal adsorption geometry occurred).
  • The second scenario: 0.5 < n < 1, this behavior indicated that Rt/BC could remove MB or MO molecules with both adsorption orientations (i.e., horizontal and vertical) but with various proportions.
  • In the last scenario: n ≥ 1, this suggested that MB or MO dye can be adsorbed on Rt/BC through an entire vertical geometry.
Furthermore, the MB and MO adsorption mechanisms on Rt/BC could be multi-molecular (i.e., one functional group of Rt/BC can remove numerous dye molecules if n > 1) or multi-docking (i.e., several adsorption sites of Rt/BC can capture one dye molecule when n < 1) [12,30,31,35]. Figure 6 shows the values of the n parameter at temperatures of 25, 40, and 50 °C, and the results are listed in Table 1. The n values were further than unity at all temperatures (i.e., 1.1–1.3 for MB and 1.5–1.8 for MO). Consequently, the vertical geometry (non-parallel orientation) and the multi-interactions mechanism were implicated in these adsorption systems. Molecules of organic water contaminants such as MB and MO can accumulate in solutions before the adsorption process (i.e., aggregation phenomenon) [21].
The increment of the n parameter with increasing temperature indicated that the solution’s aggregation of MO and MB molecules was endothermic [1]. Additionally, the free movement of MO dye (molecular dimension of 1.2 nm) favored the MO-MO interface, and thus, its molecular aggregation was higher than that of MB dye [36]. Overall, the aggregation of MB and MO with temperature reflected that the uptake processes were energetically activated in aqueous solutions [30].
The DM parameter signifies the active occupied active sites of the Rt/BC composite by MO and MB dye molecules. The influence of solution temperature on this parameter is described in Figure 5. The temperature effect in changing the value of DM can be ignored (i.e., DM was close to 100 mg/g for MB and MO) at all temperatures; see Figure 6 and Table 1. Thus, it can be concluded that the Rt/BC had a given density of surface functionalities, and the number of these adsorption sites did not change with temperature. Additionally, the addition of new active sites in the Rt/BC adsorbent cannot occur with increasing temperature (i.e., the number of adsorption sites was constant).
The adsorption capacities reflect the efficiency of Rt/BC to remove dye molecules at different experimental conditions. The impact of solution temperature on this steric parameter (Qsat =2 × n × DM) is elucidated in Figure 6. The values of adsorption capacities define the performance of Rt/BC adsorbent to remove MB and MO molecules at 25, 40, and 50 °C. These values were 214.5 (25 °C), 235.9 (40 °C), and 249.0 mg/g (50 °C) for MB. On the other hand, the corresponding Qsat values were 294.5 (25 °C), 330.1 (40 °C), and 370.4 mg/g (50 °C) for MO dye (Table 1). Generally, the increment of Qsat values with temperature suggested that the interactions between the considered dyes (MB or MO) and Rt/BC adsorption sites were endothermic processes.
Thus, the temperature supported the adsorption capacities of MB and MO dyes, and this behavior can be associated with the increasing mobility of these dyes with the enhancing temperature, particularly the MO dye. The DM values were nearly equal at all temperatures for adsorption of MB and MO dyes on Rt/BC; see Figure 6 and Table 1. Accordingly, the activated dyes molecules can interact with identified receptor sites (i.e., selective adsorption positions) in the investigated adsorbent. Therefore, the difference in the adsorption capacity could be related to the number of dye molecules that interact with each functional group. In addition, the diffusion of MO molecules in the Rt/BC pores can play a significant role in increasing the removed amounts of MO dye. This result could be associated with the different molecule sizes between MO and MB dyes and the functional groups of the tested dyes [37]. Moreover, both of the Qsat and n parameters showed an equivalent trend (i.e., the two parameters increased with temperature). Consequently, the n parameter was the key steric factor that enhanced the Rt/BC adsorption performance towards MO dye.
The monolayer adsorption capacities (Qmax) resulted from the removal of MB and MO by numerous adsorbents, and their comparison with the performance of the Rt/BC composite is listed in Table 2. Note that MO and MB adsorption capacities were higher than those reported for other materials, such as natural, modified, and fabricated adsorbents. Therefore, the studied Rt/BC composite material presented outstanding adsorption properties for being utilized to remove dyes in aqueous solutions.
Table 2. MB and MO adsorption capacities of different adsorbents.
Table 2. MB and MO adsorption capacities of different adsorbents.
AdsorbentDyeQmax (mg/g)Reference
GrapheneMB153[38]
KaolinMB45[39]
Zeolite 4AMB22[39]
Polydopamine microspheresMB90.7[40]
Activated rice huskMB65[41]
Ball clayMB25[42]
Fe3O4/montmorilloniteMB69[43]
Rt/BCMB214.52This study
Activated clayMO16.78[44]
Treated coal powderMO18.52[45]
Banana peelMO21[46]
Nano-composite filmsMO29.41[47]
Activated wheat strawMO50.4[48]
CTAB/H2O2-clayMO194.3[5]
Rt/BC compositeMO294.50The present study

3.3.2. Energetic Parameters for MB and MO Adsorption

Calculation of adsorption energy is recommended to understand the removal mechanism of MB and MO on Rt/BC composite. At 25, 40, and 50 °C, the adsorption energies (ΔE1 and ΔE1) associated with the formation of the two layers on this adsorbent were calculated using the next expressions [30,49,50].
C 1 = C s e Δ E 1 R T
C 2 = C s e Δ E 2 R T
where C 1   and C 2 are the half-saturation concentrations, and C s is the solubility of the utilized dyes. Figure 7 displays the ΔE versus solution temperature, and the results are summarized in Table 3. Surface adsorption energies were positive (i.e., both ΔE1 and ΔE2 were positive), indicating that the removal of MB and MO was endothermic, which agreed with the temperature effect on the adsorption capacities. Additionally, the adsorption energies associated with the removal of the investigated dyes were less than 30 kJ/mol representing physical forces (e.g., hydrogen bindings and van der Waals interactions) [12]. At all adsorption temperatures, values of ΔE1 were higher than that of ΔE2 for MB and MO; see Figure 7 and Table 3. This observation is because the ΔE1 characterized the Rt/BC-dye interaction, while ΔE2 characterized the dye-dye interaction (i.e., MB-MB or MO-MO) and, therefore, ΔE1 values were always the highest.

3.4. Regeneration of Rt/BC Adsorbent

After all regeneration cycles, the as-synthesized Rt/BC composite presented removal percentages (%) more than 86% and 75% for MO and MB, respectively (Figure 8). Therefore, this adsorbent can be reused many times to remove both MO and MB dyes without significantly losing its performance, particularly for MO dye molecules. Overall, using accessible and low-cost raw materials to prepare effective adsorbents for the removal of organic/inorganic water pollutants is recommended. Concerning Rt/BC, pure Rt TiO2 nanoparticles were obtained from Egyptian ilmenite and combined with black clay utilizing a facile and low-cost method. According to the adsorption/desorption results, it can be suggested that Rt/BC is a prominent and highly stable adsorbent for the remediation of dyes-containing solutions.

4. Conclusions

A composite of rutile TiO2 nanoparticles (Rt) and exfoliated black clay (BC) was prepared, characterized, and utilized for the removal of dyes (methylene blue and methyl orange). Results of adsorption processes indicated that Rt/BC displayed high efficiency for MO removal compared to MB at the optimum pH value of each tested dye. The theoretical double layer model satisfactorily fitted the experimental data (R2 >0.98) at 25, 40, and 50 °C. The parallel orientation and multi-interactions mechanisms were involved in these adsorption systems. The aggregation of MO molecules was high, and this behavior was the main reason for enhancing the adsorption capacities of MO dye. MO and MB adsorption energies were less than 30 kJ/mol, suggesting endothermic adsorption processes governed by physical interactions. Overall, the interpretation of the theoretical parameters provided new insights into the MB and MO adsorption mechanisms. Decoration of H2O2-activated clay by rutile TiO2 nanoparticles can be considered a prominent adsorbent for treating water and wastewater contaminated with dyes.

Author Contributions

Methodology, I.A.A. and M.K.S.; software A.B.-P.; writing—original draft, M.B. and M.K.S.; writing—review and editing, Z.L., E.C.L., A.B.-P. and I.A. All authors have read and agreed to the published version of the manuscript.

Funding

This work was assisted financially by the Dean of Science and Research at King Khalid University via the General Research Project (Grant No. R.G.P.1/355/42).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing is not applicable to this article.

Acknowledgments

The authors are grateful to the Dean of Science and Research at King Khalid University for making financial support available.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Xue, H.; Wang, X.; Xu, Q.; Dhaouadi, F.; Sellaoui, L.; Seliem, M.K.; Ben Lamine, A.; Belmabrouk, H.; Bajahzar, A.; Bonilla-Petriciolet, A.; et al. Adsorption of methylene blue from aqueous solution on activated carbons and composite prepared from an agricultural waste biomass: A comparative study by experimental and advanced modeling analysis. Chem. Eng. J. 2022, 430, 132801. [Google Scholar] [CrossRef]
  2. Kaledin, V.I.; Ilnitskaya, S.I.; Ovchinnikova, L.P.; Popova, N.A.; Bogdanova, L.A.; Morozkova, T.S. Mutagenic activation and carcinogenicity of aminoazo dyes ortho aminoazotoluene and 3′ methyl-4-dimethyl-amino-azobenzene in experiments on suckling mice. Biophysics 2014, 59, 431–435. [Google Scholar] [CrossRef]
  3. Alderete, B.L.; da Silva, J.; Godoi, R.; da Silva, F.R.; Taffarel, S.R.; da Silva, L.P.; Garcia, A.L.H.; Mitteregger, H., Jr.; de Amorim, H.L.N.; Picada, J.N. Evaluation of toxicity and mutagenicity of a synthetic effluent containing azo dye after advanced oxidation process treatment. Chemosphere 2021, 263, 128291. [Google Scholar] [CrossRef] [PubMed]
  4. Köktürk, M.; Altindağ, F.; Ozhan, G.; Çalimli, M.H.; Nas, M.S. Textile dyes Maxilon blue 5G and Reactive blue 203 induce acute toxicity and DNA damage during embryonic development of Danio rerio. Comp. Biochem. Physiol. Part C Toxicol. Pharmacol. 2021, 242, 108947. [Google Scholar] [CrossRef]
  5. Mobarak, M.; Selim, A.Q.; Mohamed, E.A.; Seliem, M.K. A superior adsorbent of CTAB/H2O2 solution−modified organic carbon rich-clay for hexavalent chromium and methyl orange uptake from solutions. J. Mol. Liq. 2018, 259, 384–397. [Google Scholar] [CrossRef]
  6. Seliem, M.K.; Mobarak, M.; Selim, A.; Mohamed, E.; Halfaya, R.A.; Gomaa, H.K.; Anastopoulos, I.; Giannakoudakis, D.A.; Lima, E.C.; Bonilla-Petriciolet, A.; et al. A novel multifunctional adsorbent of pomegranate peel extract and activated anthracite for Mn(VII) and Cr(VI) uptake from solutions: Experiments and theoretical treatment. J. Mol. Liq. 2020, 311, 113169. [Google Scholar] [CrossRef]
  7. Jalil, A.A.; Triwahyono, S.; Adam, S.H.; Rahim, N.D.; Aziz, M.A.A.; Hairom, N.H.H.; Razali, N.A.M.; Abidin, M.A.; Mohamadiah, M.K.A. Adsorption of methyl orange from aqueous solution onto calcined Lapindo volcanic mud. J. Hazard. Mater. 2010, 181, 755–762. [Google Scholar] [CrossRef] [Green Version]
  8. Mohammadi, N.; Khani, H.; Gupta, V.K.; Amereh, E.; Agarwal, S. Adsorption process of methyl orange dye onto mesoporous carbon material–kinetic and thermodynamic studies. J. Colloid Interface Sci. 2011, 362, 457–462. [Google Scholar] [CrossRef] [PubMed]
  9. Barakat, M.A.; Selim, A.Q.; Mobarak, M.; Kumar, R.; Anastopoulos, I.; Giannakoudakis, D.; Bonilla-Petriciolet, A.; Mohamed, E.A.; Seliem, M.K.; Komarneni, S. Experimental and Theoretical Studies of Methyl Orange Uptake by Mn–Rich Synthetic Mica: Insights into Manganese Role in Adsorption and Selectivity. Nanomaterials 2020, 10, 1464. [Google Scholar] [CrossRef] [PubMed]
  10. Sun, P.; Hui, C.; Khan, R.A.; Du, J.; Zhang, Q.; Zhao, Y.-H. Efficient removal of crystal violet using Fe3O4-coated biochar: The role of the Fe3O4 nanoparticles and modeling study their adsorption behavior. Sci. Rep. 2015, 5, 12638. [Google Scholar] [CrossRef] [Green Version]
  11. Seliem, M.K.; Komarneni, S.; Byrne, T.; Cannon, F.S.; Shahien, M.G.; Khalil, A.A.; AbdEl-Gaid, I.M. Removal of perchlorate by synthetic organosilicas and organoclay: Kineticsand isotherm studies. Appl. Clay Sci. 2013, 71, 21–26. [Google Scholar] [CrossRef]
  12. Li, Z.; Sellaoui, L.; Dotto, G.L.; Ben Lamine, A.; Bonilla-Petriciolet, A.; Hanafy, H.; Belmabrouk, H.; Netto, M.S.; Erto, A. Interpretation of the adsorption mechanism of Reactive Black 5 and Ponceau 4R dyes on chitosan/polyamide nanofibers via advanced statistical physics model. J. Mol. Liq. 2019, 285, 165–170. [Google Scholar] [CrossRef]
  13. Mohamed, E.A.; Selim, A.Q.; Ahmed, S.A.; Sellaoui, L.; Bonilla-Petriciolet, A.; Erto, A.; Li, Z.; Li, Y.; Seliem, M.K. H2O2-activated anthracite impregnated with chitosan as a novel composite for Cr(VI) and methyl orange adsorption in single-compound and binary systems: Modeling and mechanism interpretation. Chem. Eng. J. 2020, 380, 122445. [Google Scholar] [CrossRef]
  14. Ramadan, H.; Mobarak, M.; Lima, E.C.; Bonilla-Petriciolet, A.; Li, Z.; Seliem, M.K. Cr(VI) adsorption onto a new composite prepared from Meidum black clay and pomegranate peel extract: Experiments and physicochemical interpretations. J. Environ. Chem. Eng. 2021, 9, 105352. [Google Scholar] [CrossRef]
  15. Barakat, M.A.; Kumar, R.; Lima, E.C.; Seliem, M.K. Facile synthesis of muscovite–supported Fe3O4 nanoparticles as an adsorbent and heterogeneous catalyst for effective removal of methyl orange: Characterisation, modelling, and mechanism. J. Taiwan Inst. Chem. Eng. 2021, 119, 146–157. [Google Scholar] [CrossRef]
  16. Hajjaji, M.; Alami, A. Influence of operating conditions on methylene blue uptake by a smectite rich clay fraction. Appl. Clay Sci. 2009, 44, 127–129. [Google Scholar] [CrossRef]
  17. Tehrani, A.; Nikkar, H.; Mahmoodi, N.M.; Markazi, M.; Menger, F. The sorption of cationic dyes onto kaolin: Kinetic, isotherm and thermodynamic studies. Desalination 2011, 266, 274–280. [Google Scholar] [CrossRef]
  18. Wang, S.; Boyjoo, Y.; Choueib, A.; Zhu, Z. Removal of dyes from aqueous solution using fly ash and red mud. Water Res. 2005, 39, 129–138. [Google Scholar] [CrossRef] [PubMed]
  19. Hajjaji, M.; Alami, A.; El Bouadili, A. Removal of methylene blue from aqueous solution by fibrous clay minerals. J. Hazard. Mater. 2006, 135, 188–192. [Google Scholar] [CrossRef]
  20. Chen, Z.X.; Jin, X.Y.; Chen, Z.; Megharaj, M. Naidu, Removal of methyl orange from aqueous solution using benton-ite-supported nanoscale zero-valent iron. J. Colloid Interface Sci. 2011, 363, 601–607. [Google Scholar] [CrossRef]
  21. Ramadana, H.S.; Alia, R.A.M.; Mobarak, M.; Badawi, M.; Selim, A.Q.; Mohamed, E.A.; Bonilla-Petriciolet, A.; Seliem, M.K. One-step fabrication of a new outstanding rutile TiO2 nanoparticles/anthracite adsorbent: Modeling and physicochemical interpretations for malachite green removal. Chem. Eng. J. 2021, 426, 131890. [Google Scholar] [CrossRef]
  22. Wang, R.; Cai, X.; Shen, F. TiO2 hollow microspheres with mesoporous surface: Superior adsorption performance for dye removal. Appl. Surf. Sci. J. 2014, 305, 352–358. [Google Scholar] [CrossRef]
  23. Banerjee, S.; Benjwal, P.; Singh, M.; Kar, K.K. Graphene oxide (rGO)-metal oxide (TiO2/Fe3O4) based nanocomposites for the removal of methylene blue. Appl. Surf. Sci. 2018, 439, 560–568. [Google Scholar] [CrossRef]
  24. Quiñones, C.; Ayalaa, J.; Vallejo, W. Methylene blue photoelectrodegradation under UV irradiation on Au/Pd-modified TiO2 films. Appl. Surf. Sci. 2010, 257, 367–371. [Google Scholar] [CrossRef]
  25. Nawi, M.A.; Zain, S.M. Enhancing the Surface Properties of the Immobilized Degussa P-25 TiO2 for the Efficient Photocatalytic Removal of Methylene Blue from Aqueous Solution. Appl. Surf. Sci. 2012, 258, 6148–6157. [Google Scholar] [CrossRef]
  26. Devi, L.G.; Kumar, S.G. Exploring the critical dependence of adsorption of various dyes on the degradation rate using Ln3+-TiO2 surface under UV/solar light. Appl. Surf. Sci. 2012, 261, 137–146. [Google Scholar] [CrossRef]
  27. Li, J.; Feng, J.; Yan, W. Excellent adsorption and desorption characteristics of polypyrrole/TiO2 composite for Methylene Blue. Appl. Surf. Sci. 2013, 279, 400–408. [Google Scholar] [CrossRef]
  28. Shahien, M.G.; Khedr, M.M.; Maurice, A.E.; Farghali, A.; Ali, R.A. Synthesis of high purity rutile nanoparticles from medium-grade Egyptian natural ilmenite. Beni-Suef Univ. J. Basic Appl. Sci. 2015, 4, 207–213. [Google Scholar] [CrossRef] [Green Version]
  29. Selim, A.Q.; Sellaoui, L.; Ahmed, S.A.; Mobarak, M.; Mohamed, E.A.; Ben Lamine, A.; Erto, A.; Bonilla-Petriciolet, A.; Seliem, M.K. Statistical physics-based analysis of the adsorption of Cu2+ and Zn2+ onto synthetic cancrinite in single-compound and binary systems. J. Environ. Chem. Eng. 2019, 7, 103217. [Google Scholar] [CrossRef]
  30. Li, Z.; Hanafy, H.; Zhang, L.; Sellaoui, L.; Netto, M.S.; Oliveira, M.L.S.; Seliem, M.K.; Dotto, G.L.; Bonilla-Petriciolet, A.; Li, Q. Adsorption of congo red and methylene blue dyes on an ashitaba waste and a walnut shell-based activated carbon from aqueous solutions: Experiments, characterization and physical interpretations. Chem. Eng. J. 2020, 388, 124263. [Google Scholar] [CrossRef]
  31. Abu Sharib, A.A.A.; Bonilla-Petriciolet, A.; Selim, A.Q.; Mohamed, E.A.; Seliem, M.K. Utilizing modified weathered basalt as a novel approach in the preparation of Fe3O4 nanoparticles: Experimental and theoretical studies for crystal violet adsorp-tion. J. Environ. Chem. Eng. 2021, 9, 106220. [Google Scholar] [CrossRef]
  32. Amrhar, O.; El Gana, L.; Mobarak, M. Calculation of adsorption isotherms by statistical physics models: A review. Environ. Chem. Lett. 2021, 19, 4519–4547. [Google Scholar] [CrossRef]
  33. Tran, H.N.; Fen, Y.; You, W.S.J.; Chao, H.P. Insights into the mechanism of cationic dye adsorption on activated charcoal: The importance of π–π interactions. Process Saf. Environ. Prot. 2017, 107, 168–180. [Google Scholar] [CrossRef]
  34. Shi, X.; Fu, H.; Li, Y.; Mao, J.; Zheng, S.; Zhu, D. Impact of coal structural heterogeneity on the nonideal sorption of organic contaminants. Environ. Toxicol. Chem. 2011, 30, 1310–1319. [Google Scholar] [CrossRef]
  35. Sellaoui, L.; Saha, B.; Wjihi, S.; Ben Lamine, A. Physicochemical parameters interpretation for adsorption equilibrium of ethanol on metal organic framework: Application of the multilayer model with saturation. J. Mol. Liq. 2017, 233, 537–542. [Google Scholar] [CrossRef]
  36. Hua, P.; Sellaoui, L.; Franco, D.; Netto, M.S.; Dotto, G.L.; Bajahzar, A.; Belmabrouk, H.; Bonilla-Petriciolet, A.; Li, Z. Adsorption of acid green and procion red on a magnetic geopolymer based adsorbent: Experiments, characterization, and theo-retical treatment. Chem. Eng. J. 2020, 383, 123113. [Google Scholar] [CrossRef]
  37. Gong, R.; Ye, J.; Dai, W.; Yan, X.; Hu, J.; Hu, X.; Li, S.; Huang, H. Adsorptive Removal of Methyl Orange and Methylene Blue from Aqueous Solution with Finger-Citron-Residue-Based Activated Carbon. Ind. Eng. Chem. Res. 2013, 52, 14297–14303. [Google Scholar] [CrossRef]
  38. Liu, T.; Li, Y.; Du, Q.; Sun, J.; Jiao, Y.; Yang, G.; Wang, Z.; Xia, Y.; Zhang, W.; Wang, K.; et al. Adsorption of methylene blue from aqueous solution by graphene. Colloids Surf. B Biointerfaces 2012, 90, 197–203. [Google Scholar] [CrossRef]
  39. Rida, K.; Bouraoui, S.; Hadnine, S. Adsorption of methylene blue from aqueous solution by kaolin and zeolite. Appl. Clay Sci. 2013, 83–84, 99–105. [Google Scholar] [CrossRef]
  40. Fu, J.; Chen, Z.; Wang, M.; Liu, S.; Zhang, J.; Zhang, J.; Han, R.; Xu, Q. Adsorption of methylene blue by a high-efficiency adsorbent (polydopamine microspheres): Kinetics, isotherm, thermodynamics and mechanism analysis. Chem. Eng. J. 2015, 259, 53–61. [Google Scholar] [CrossRef]
  41. Franco, D.; Tanabe, E.H.; Bertuol, D.; Reis, G.; Lima, E.C.; Dotto, G.L. Alternative treatments to improve the potential of rice husk as adsorbent for methylene blue. Water Sci. Technol. 2016, 75, 296–305. [Google Scholar] [CrossRef] [PubMed]
  42. Auta, M.; Hameed, B.H. Modified mesoporous clay adsorbent for adsorption isotherm and kinetics of methylene blue. Chem. Eng. J. 2012, 198–199, 219–227. [Google Scholar] [CrossRef]
  43. 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]
  44. Ma, Q.; Shen, F.; Lu, X.; Bao, W.; Ma, H. Studies on the adsorption behavior of methyl orange from dye wastewater onto activated clay. Desalination Water Treat. 2013, 51, 3700–3709. [Google Scholar] [CrossRef]
  45. Liu, Z.; Zhou, A.; Wang, G.; Zhao, X. Adsorption Behavior of Methyl Orange onto Modified Ultrafine Coal Powder. Chin. J. Chem. Eng. 2009, 17, 942–948. [Google Scholar] [CrossRef]
  46. Annadurai, G.; Juang, R.-S.; Lee, D.-J. Use of cellulose-based wastes for adsorption of dyes from aqueous solutions. J. Hazard. Mater. 2002, 92, 263–274. [Google Scholar] [CrossRef]
  47. Jiang, R.; Fu, Y.Q.; Zhu, H.Y.; Yao, J.; Xiao, L. Removal of Methyl Orange from Aqueous Solutions by Magnetic Maghemite/Chitosan Nanocomposite Films: Adsorption Kinetics and Equilibrium. J. Appl. Polym. Sci. 2012, 125, 540–549. [Google Scholar] [CrossRef]
  48. Su, Y.; Jiao, Y.; Dou, C.; Han, R. Biosorption of methyl orange from aqueous solutions using cationic surfactant-modified wheat straw in batch mode. Desalination Water Treat. 2014, 52, 6145–6155. [Google Scholar] [CrossRef]
  49. Mohamed, E.A.; Mobarak, M.; Kumar, R.; Barakat, M.; Bonilla-Petriciolet, A.; Seliem, M.K.; Selim, A.Q. Novel hybrid multifunctional composite of chitosan and altered basalt for barium adsorption: Experimental and theoretical studies. Colloids Surf. A Physicochem. Eng. Asp. 2020, 593, 124613. [Google Scholar] [CrossRef]
  50. Selim, A.Q.; Mohamed, E.A.; Seliem, M.K. Deep insights into the organic carbon role in selectivity and adsorption mechanism of phosphate and crystal violet onto low–cost black limestone: Modelling and physicochemical parameters interpretation. Colloids Surf. A Physicochem. Eng. Asp. 2019, 580, 123755. [Google Scholar] [CrossRef]
Scheme 1. The molecular structures of the studied dyes.
Scheme 1. The molecular structures of the studied dyes.
Coatings 12 00022 sch001
Figure 1. The expected situations based on the multilayer adsorption model.
Figure 1. The expected situations based on the multilayer adsorption model.
Coatings 12 00022 g001
Figure 2. (a) XRD pattern and (b) FTIR spectrum of Rt/BC material.
Figure 2. (a) XRD pattern and (b) FTIR spectrum of Rt/BC material.
Coatings 12 00022 g002
Figure 3. FESEM images of the Rt/BC material at different magnification scales. (a) exfoliated BC; (b,c) Rt/BC composite; (df) the aggregated Rt nanoparticles decorated BC sample.
Figure 3. FESEM images of the Rt/BC material at different magnification scales. (a) exfoliated BC; (b,c) Rt/BC composite; (df) the aggregated Rt nanoparticles decorated BC sample.
Coatings 12 00022 g003
Figure 4. pHPZC of the Rt/BC composite and pH effect on MB and MO dyes adsorption on Rt/BC adsorbent. Rt/BC composite displayed a high removal performance for MO (96.7%) and MB (91.4%) at pH 3.0 and 8.0, respectively.
Figure 4. pHPZC of the Rt/BC composite and pH effect on MB and MO dyes adsorption on Rt/BC adsorbent. Rt/BC composite displayed a high removal performance for MO (96.7%) and MB (91.4%) at pH 3.0 and 8.0, respectively.
Coatings 12 00022 g004
Figure 5. Advanced statistical physics modeling of adsorption isotherms of MO and MB dyes on Rt/BC composite. The double layer model describe well (i.e., R2 > 0.98) the adsorption of MB and MO on the Rt/BC adsorben.
Figure 5. Advanced statistical physics modeling of adsorption isotherms of MO and MB dyes on Rt/BC composite. The double layer model describe well (i.e., R2 > 0.98) the adsorption of MB and MO on the Rt/BC adsorben.
Coatings 12 00022 g005
Figure 6. Physicochemical parameters n, DM, and Qsat as a function of temperature for MB and MO adsorption on the Rt/BC adsorbent.
Figure 6. Physicochemical parameters n, DM, and Qsat as a function of temperature for MB and MO adsorption on the Rt/BC adsorbent.
Coatings 12 00022 g006
Figure 7. Adsorption energies for (a) MB and (b) MO adsorption on Rt/BC adsorbent.
Figure 7. Adsorption energies for (a) MB and (b) MO adsorption on Rt/BC adsorbent.
Coatings 12 00022 g007
Figure 8. Percentage of MB and MO removal during the adsorption/desorption cycles.
Figure 8. Percentage of MB and MO removal during the adsorption/desorption cycles.
Coatings 12 00022 g008
Table 1. Steric parameters of the double layer model for the adsorption of MB and MO on the Rt/BC composite.
Table 1. Steric parameters of the double layer model for the adsorption of MB and MO on the Rt/BC composite.
T (°C)DyenDM (mg/g)Qsat (mg/g)
25MB1.1295.76214.52
40MB1.18100235.9
50MB1.2996.66249.03
25MO1.47100294.5
40MO1.65100330.1
50MO1.85100370.37
Table 3. Energetic parameters of the double layer model for the MB and MO uptake on Rt/BC composite.
Table 3. Energetic parameters of the double layer model for the MB and MO uptake on Rt/BC composite.
T (°C)DyeC1C2ΔE1 (kJ/mol)ΔE2 (kJ/mol)
25MB47.1399.9916.9315.07
40MB36.5799.4718.4415.84
50MB24.0799.5320.1516.342
25MO27.7610012.879.7
40MO20.499.3914.3210.2
50MO12.7889.6116.0310.8
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ahmed, I.A.; Seliem, M.K.; Lima, E.C.; Badawi, M.; Li, Z.; Bonilla-Petriciolet, A.; Anastopoulos, I. Outstanding Performance of a New Exfoliated Clay Impregnated with Rutile TiO2 Nanoparticles Composite for Dyes Adsorption: Experimental and Theoretical Studies. Coatings 2022, 12, 22. https://doi.org/10.3390/coatings12010022

AMA Style

Ahmed IA, Seliem MK, Lima EC, Badawi M, Li Z, Bonilla-Petriciolet A, Anastopoulos I. Outstanding Performance of a New Exfoliated Clay Impregnated with Rutile TiO2 Nanoparticles Composite for Dyes Adsorption: Experimental and Theoretical Studies. Coatings. 2022; 12(1):22. https://doi.org/10.3390/coatings12010022

Chicago/Turabian Style

Ahmed, Inas A., Moaaz K. Seliem, Eder C. Lima, Michael Badawi, Zichao Li, Adrián Bonilla-Petriciolet, and Ioannis Anastopoulos. 2022. "Outstanding Performance of a New Exfoliated Clay Impregnated with Rutile TiO2 Nanoparticles Composite for Dyes Adsorption: Experimental and Theoretical Studies" Coatings 12, no. 1: 22. https://doi.org/10.3390/coatings12010022

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