Attempting to reduce issues with dumping and water pollution, bio-based membrane material (MB1000), based on bentonite was elaborated for application in tangential ultrafiltration. For this, morphological properties, textural properties, and chemical structure of the elaborated membrane material were established using Fourier transform infrared spectroscopy, X-ray diffraction, and Brunauer–Emmett–Teller analyses. Water permeability, chemical resistance, as well as point of zero charge of the membrane material were also investigated. The studied membrane material has a mesoporous structure, with a pore size of 7.20 nm and a water permeability of 318.06 L/h.m2.bar. The effect of transmembrane material pressure, pH solution, and concentration on Orange G (OG) and Rhodamine B (RB) dye rejection efficiency was examined and hence optimized. Besides, a mixture of RB and OG dyes was tested for membrane material ultrafiltration in a simultaneous system (RB/OG). Remarkably, an enhancement of the rejection results was noticed for the two dyes (ROG = 94.33%, RRB = 89.38%) resulting from a synergic effect of hydrogen bonding as well as electrostatic interactions generated from functional groups of the molecules dyes.

  • Elaboration of low-cost and soft bio-based membrane using natural clay.

  • Membrane ultrafiltration of Orange G (OG) and Rhodamine B (RB).

  • Membrane ultrafiltration of OG/RB blend.

  • Synergy in the ultrafiltration of the blend OG/RB and.

  • Improvement of the rejection percent of OG and RB in the blend compared to the single systems.

Organic dyes are now widely used across a number of industries like many other sectors that use dyes and pigments, such as paper, leather tanning, cosmetics, and plastics, and the textile industry is notable for its high water use as well as for being one of the main industrial producers of colored effluents (Wang et al. 2019; Seo et al. 2020). Additionally, these dyes are resistant to ultraviolet radiation and do not biodegrade, which results in long-term water pollution (Bouazizi et al. 2017a; Touahra et al. 2022). Reusing water is a major challenge for industrial companies as it allows them to conserve water while maintaining the sustainability of water supplies for future generations (Hamri et al. 2022). Multiple harmful compounds are emitted from industrial sites into water and soil, which is one of the negative repercussions of industrialization, For example, wastewater from the textile sector contains a complex mixture of cationic and anionic dyes, including some hazardous metal ions, and they discharge effluents that pose environmental problems (Atun et al. 2019; Liu et al. 2020). The aforementioned industries make extensive use of Rhodamine B (RB) and Orange G (OG), carcinogen compounds that not only harm the environment but also pose a risk to the health of both humans and animals (USEPA 2014; Mohamed et al. 2019; Tolkachev & Bespamyatnykh 2019). Therefore, it is very important to remove these contaminants from wastewater. Few studies, however, have focused on the simultaneous removal of both pollutants despite the fact that they frequently coexist in wastewater and may have a substantial impact on one another's destiny and travel. As reported in the literature, there are many different technologies to remove dyes from wastewater, such as coagulation (Liang et al. 2014), adsorption (Benhalima et al. 2017; Djamaa et al. 2019), photocatalytic degradation (Bouarroudj et al. 2021; Mekidiche et al. 2021), biological oxidation (Das et al. 2021), chemical oxidation (Oller et al. 2011), and membrane filtration (Bouzerara et al. 2006; Harabi et al. 2014). Unfortunately, these technologies are not always effective at removing soluble dyes because wastewater varies in terms of the type and quantity of dyes present, as well as in complexity and volume, and hence, new adequate and effective alternatives based on natural, abundant, and inexpensive materials have been developed for colored wastewater treatment. Given that sieving is mostly used in membrane separation, which makes it more suitable for the majority of industrial effluent, membrane separation might be seen in this context as an efficient and green technology to remediate dye wastewater (Cao et al. 2020). Membrane procedures are also acknowledged as being reasonably cost-effective methods, primarily because they require few chemical reagents and use little energy (Achiou et al. 2017). Reverse osmosis (RO), microfiltration, ultrafiltration (UF), and nanofiltration (NF) are four pressure-driven membrane technologies that are frequently employed to decontaminate textile wastewater (Dasgupta et al. 2015). It should be noted that additional membrane processes such as vacuum membrane distillation, membrane bioreactors, and electrodialysis (ED) may be applicable (Dasgupta et al. 2015; Parakala et al. 2019). Targeted species, wastewater volume, and operating conditions are just a few of the factors that should be taken into account when choosing an appropriate membrane process. Compared to NF and RO, UF requires lower pressure and might provide excellent separation efficiency in terms of flux and desirable rejection (saving energy) (Benkhaya et al. 2019). Due to these factors, UF is frequently chosen as the best procedure for removing soluble dyes (Bouazizi et al. 2017b; Tran et al. 2019). As described in the literature, many works that used UF membranes to remove dyes were published. Aloulou et al. developed an asymmetric microfiltration membrane from natural zeolite with a permeability of 176 L/m2.h.bar and an average pore size of 0.55 μm. These membranes were characterized by good chemical oxygen demand removal potential (>85%) and applied in industrial wastewater treatment (Aloulou et al. 2020). Sawunyama et al. functionalized ceramic membranes for the removal of pharmaceuticals in wastewater (Sawunyama et al. 2023). On the other hand, a new alumina UF membrane with a permeability of 70 L/m2.h.bar was developed by Zou et al., and this membrane was used for heat treatment of colored wastewater (Zou et al. 2017). Also, Ouaddari et al. elaborated membranes from natural clay for the UF of Direct Red 80 (Ouaddari et al. 2019). In another study, Jiang et al. desalinated several reactive dye species using a tight ceramic UF membrane made of TiO2 with a molecular weight cutoff of 2,410 Da. A large percentage of the reactive dye was rejected by this membrane (>98.12%) (Jiang et al. 2018). Generally, industrial metal oxides like titania, zirconia, silica, and alumina were used to manufacture ceramic membranes that are more expensive (Achiou et al. 2018a), and hence, various materials were used for mineral membrane preparation, such as phosphate (Barrouk et al. 2014; Mouiya et al. 2018), ceramics (Fang et al. 2013; Bouazizi et al. 2016; Abdullayev et al. 2019; Mouiya et al. 2019), and zeolites (Wang & Peng 2010; Visa 2016). Due to their excellent mechanical properties such as thermal stability, good chemical resistance, and long lifetime, mineral membrane-based clays have real potential in the environmental protection field. Therefore, a great deal of inexpensive UF membranes have been effectively proposed based on low-cost materials (Al-Futaisi et al. 2007; Bhatt et al. 2012; Silva et al. 2012; Sharma et al. 2015; Foorginezhad & Zerafat 2017; Ouaddari et al. 2019; Saja et al. 2020; Ouachtak et al. 2021; de Almeida et al. 2022). As an example, for the elimination of bromothymol blue, methyl orange, and murexide from aqueous solutions, Karim et al. produced UF membrane from graphene oxide and natural pozzolan (Karim et al. 2018). In other works, the TiO2-UF membrane was prepared by bentonite/phosphate support and was successfully used for the filtration of Direct Red 80 (Bouazizi et al. 2017a; Touahra et al. 2022). In a review on the development of nanocomposite membranes for water treatment, Ursino et al. recommend incorporating innovative nanoscale materials into polymeric membranes, such as carbon nanotubes, graphene oxide, zinc oxide, titanium dioxide, silver, and copper nanoparticles (Sikorski & Trzybiński 2014; Minitha et al. 2017; Ursino et al. 2018). So, one of the major problems, which can occur in all membrane classes, is membrane fouling, which refers to the accumulation of undesirable substances on the surface or within the pores of a membrane material, leading to such reduction of its efficiency. This membrane fouling can be essentially related to the feed water quality, the inadequacy of the selected filtration process, and/or the operating conditions. In addition, membrane fouling has a clear impact on the membrane performance by reducing the permeate flow, increasing transmembrane pressure and thus shortening membrane lifespan. Referring to these works and based on our previous study, the preparation of low-cost UF membrane-based bentonite clay with optimized conditions is a promising approach to prepare a cylindrical membrane for tangential filtration. In this work, bio-based material was elaborated by physical blending way of bentonite clay under a sintering temperature of 1,000 °C. The microstructure, pore size, chemical stability, point of zero charge, as well as permeability and other properties were investigated. The performance of the elaborated membrane was evidenced by tangential filtration of the cationic dye RB and the anionic dye OG in a single system and a simultaneous one (RB/OG), respectively. For this, the evolution of the dye's rejection was examined according to transmembrane pressure, pH, and concentration solution experimental parameters. The synergy on the simultaneous system (RB/OG) was also described to underline the important rejection results obtained in this system.

Materials

Bentonite used for preparation of UF membrane was kindly supplied by Entreprise nationale des produits miniers non ferreux et des substances utiles (ENOF). Amijel, Methocel, OG (C16H10N2Na2O7S2), and RB (C28H31CIN2O3) dyes were supplied by Sigma Aldrich. Sodium hydroxide pellets (NaOH) and hydrochloric acid (HCl) were purchased from Merck. All reagents were used, as received without further purification. The chemical structure of the studied dyes is illustrated in Figure 1.
Figure 1

Molecular structure of the dyes.

Figure 1

Molecular structure of the dyes.

Close modal

Elaboration of tubular membrane

The UF tubular membrane was elaborated as follows: First, bentonite clay was sieved to a fine powder with a mean particle size of 100 μm. Then, 4 wt. % of both binding and plasticizer agents was dispersed in 50 mL of distilled water. The obtained material was purged for a few minutes before extrusion processing at a screw speed of 0.02 rpm. The material was dried at room temperature for 24 h and then annealed at 1,000 °C for 3 h at a heating rate of 1 °C.min−1. The obtained tubular membrane dimensions are as follows: L: 150 mm, ED: 8 mm, and ID: 5 mm, with an effective filtration area of 33 cm2. It worked inside our membrane module and was noticed as MB1000.

Point of zero charge pHpzc

The point of zero charge of the MB1000 was determined according to the method described in the literature (Djebri et al. 2016; Saavedra-Labastida et al. 2019). Briefly, 0.1 g of materials was placed in aliquots of 0.01 N NaCl solution for 48 h, and then, the pH was varied using HCl or NaOH solutions (0.1N). The materials' pHpzc was calculated by intersecting the bisector curve with the plot of pH final versus pH initial.

Membrane ultrafiltration experiments

Tangential UF experiments were carried out using a laboratory-grade UF pilot made of stainless steel, equipped with a 3 L feed tank, a circulation pump, two manometers, a membrane housing, and a permeate tank (Figure 2). First, the membrane was submerged in water for 24 h before being installed in the housing with a filtering surface area of 3.30 cm2. The membrane received feed solution through tangential pumping. The permeate was collected and analyzed. Water filtered at room temperature (25 ± 2 °C) under various pressures served as the basis for measuring water permeability. Kipping this point in mind, applied transmembrane pressure in all experiments was varied from 0 to 4 bar. The filtering parameters flux (Jw0: L/m2.h) and permeability (Lp: L/m2.h.bar) were calculated according to Equations (1) and (2), respectively.
(1)
(2)
where V (L) is the permeate's volume, A (m2) is the filtration area, t (h) is the filtration time, and ΔP (bar) is the transmembrane pressure. To highlight the performance of the elaborated membrane, UF of RB and OG selected as cationic and anionic dyes, respectively, in a single and then a simultaneous system (RB/OG) was conducted under different experimental conditions. The effects of pressure (1–4 bar), pH solution (1–12), and feed concentration (25–100 ppm) on rejection percent results were established at a wavelength of 554 and 478 nm for RB and OG, respectively. The rejection (R (%)) of dyes was calculated according to Equation (3).
Figure 2

Experimental treatment system.

Figure 2

Experimental treatment system.

Close modal
(3)
where Cfeed and Cpermeate are the concentration of the feed and the permeate, respectively.

After each applied pressure, the UF membrane was cleaned by the clean water backwash method, using hot distilled water and then using acid and basic solutions. Finally, it is washed several times using hot distilled water.

Characterization techniques

Fourier transform infrared spectroscopy (FTIR) (Bruker Alpha) spectrometer was used to analyze the chemical composition of the membrane over the wavelength range 4,000–400 cm−1, with a spectral resolution of 2 cm−1 and 64 scans. The mineralogical composition of clay and modified sintered bentonite at 1,000 °C was carried out using X-ray diffraction (XRD) analysis and a diffractometer (BRUKER D8 ADVANCE) equipped with a Ni-filtered Cu-Kα (λ = 1.542A°) radiation. The apparatus operated at 40 KV and 25 mA in a step scan mode and a diffraction speed of 0.02°s−1. Pore size distribution and surface area of the membrane were determined via N2 adsorption/desorption at 77 K using an automated gas sorption apparatus by a well-known, Brunauer–Emmett–Teller method. The morphological and microstructure of the elaborated membrane were checked by using scanning electron microscopy (SEM) (QUANTA 250) operating at a working distance of 10.0 mm and under an acceleration voltage of 20.00 KV. A spectrophotometer device (Varian Cary 50 Scan) was used for the measurements of RB and OG concentration in the range 200–800 nm, at 2 nm resolution, at a wavelength of 554 and 478 nm for RB and OG, respectively.

Membrane material characterization

FTIR spectra of the elaborated membrane before and after sintering are shown in Figure 3(a). The stretching vibrations of the hydroxyl groups in the interlayer montmorillonite structure are allocated to a broadband in the pure clay spectrum that is focused around 3,550, 3,414, and 1,638 cm−1. The bands at 1,036 cm−1 define the stretching vibrations of the clay's Si-O, Si-O-Si, and Al-O groups. The bands at 521 and 469 cm−1 are attributed to Al-O and Mg-O, deformation vibrations, respectively. Only the distinctive vibrations of silicate groups are recorded after sintering at 1,000 °C, especially, the intense band centered at 1,010 cm−1, assigned to the Si-O stretching vibrations, the band at 787 cm−1 relied on quartz-specific Si-O stretching vibrations, and the band in the range 454 and 490 cm−1 ascribed to O-Si-O stretching vibrations. XRD diffractogram of bentonite elaborated membrane material at 1,000 °C is displayed in Figure 3(b). Spectrum demonstrates the presence of comparable crystalline phases. The quartz silicone oxide phase (JCPDS 76-1906) with hexagonal symmetry of the P3121 space group is represented by the diffraction peaks at 2 = 20.8°, 21.9°, 26.7°, 28.2°, 31.4°, 35.7°, and 36.5°, with triclinic symmetry of the C-0 space group. Calcium alumina silicate Ca 0.88 Al 1.77 Si 2.23 O8 is also identified in the bio-based membrane (JCPDS 00-041-1) (Zhang et al. 2019; Labied et al. 2022). Figure 3(c) depicts the nitrogen adsorption and desorption isotherm of the bio-based membrane. The type IV isotherm that is typical of mesoporous material can be seen according to the International Union of Pure and Applied Chemistry (IUPAC classification). The generated isotherm has hysteresis loops of type H3, which develop plate-shaped particle aggregates that give rise to slit-shaped pores and hysteresis that occurs when equilibrium pressures vary during adsorption and desorption with pores in the 2.5–50 nm range.
Figure 3

Structural characterization of MB1000: FTIR spectra before and after sintering at 1,000 °C (a), XRD pattern after sintering at 1,000 °C (b), and adsorption–desorption isotherm behavior (c).

Figure 3

Structural characterization of MB1000: FTIR spectra before and after sintering at 1,000 °C (a), XRD pattern after sintering at 1,000 °C (b), and adsorption–desorption isotherm behavior (c).

Close modal

Water permeability

Figure 4(a) shows the permeate flux of the MB1000 membrane as a function of transmembrane pressure. It is evident that the curve in the examined pressure range of 0–4 bar is largely linear, demonstrating that Darcy's law is strictly adhered to (Tran et al. 2019). This excellent permeability value of 318.06 L/m2.h.bar can be explained by the homogeneous porosity repartition that agrees with the obtained morphological and pore size distribution results.
Figure 4

Water permeability (a) and point of zero charge (b) of the elaborated bio-based membrane.

Figure 4

Water permeability (a) and point of zero charge (b) of the elaborated bio-based membrane.

Close modal

Point of zero charge (pHpzc)

As the pH of the solution changes, the charge of the material surface may be impacted. Therefore, the material exhibits a positive surface charge when the solution pH value is lower than the pHpzc (Silva et al. 2012; Saavedra-Labastida et al. 2019; El Gaayda et al. 2021). The surface tends to have a negative charge when the pH solution exceeds pHpzc, and pHPZC was determined by using the batch equilibration method described in the previous work (Labied et al. 2022). In our case, the pHpzc value is 7.25 (Figure 4(b)).

Optimal ultrafiltration experimental parameters

The prepared membrane MB1000 was applied for tangential filtration of RB and OG dyes, as organic pollutants. Experiments were conducted in single and simultaneous systems (RB/OG), respectively. For this, the effects of transmembrane pressure, pH solution, and dye feed concentration on permeate flux and rejection percent were studied.

Transmembrane pressure

Figure 5(a) Illustrates the permeate flux of RB and OG as a function of time, at different transmembrane pressures and at feed solution dye concentration of 25 ppm. It has been clearly observed that permeate flux increases by increasing transmembrane pressure whatever the dye pollutant, implying a decrease in the rejection efficiency. According to the obtained results, the transmembrane seems to be ideal for the tangential filtration at 1 bar, in which the rejection progress seems to be more important for OG than RB dyes (Figure 5(b)). Similar results were obtained by Bouazizi et al. (2017b). The polarization phenomenon caused by the build-up of retained dye molecules on the membrane surface can explain the increasing rejection over time (Karim et al. 2018). Furthermore, dye particles may be absorbed onto the inner surface of membrane pores, leading to the increase in the rejection percent.
Figure 5

Evolution of RB and OG flux as a function of transmembrane pressures, at different filtration time, with feed dye concentration of 25 ppm (a) and rejections of dyes at optimized transmembrane pressure of 1 bar (b).

Figure 5

Evolution of RB and OG flux as a function of transmembrane pressures, at different filtration time, with feed dye concentration of 25 ppm (a) and rejections of dyes at optimized transmembrane pressure of 1 bar (b).

Close modal

Solution pH

The influence of feed solution pH on the UF efficiency of the studied dyes as a function of time was also examined. The UF process was increased in the range 0–50 min, to achieve the equilibrium after 70 min of filtration (Figure 6). This equilibrium is generated by the electrostatic interactions between the membrane surface and the dye molecules (Breida et al. 2018). According to Figure 6, maximum rejection values of 65.08 and 75.89% were obtained for RB and OG, respectively. The corresponding pH values are 5.36 and 3.07, respectively. These results may be relied on by the electrostatic interactions between the membrane surface and the dye molecules depending on pHPZC (Chieng et al. 2015; Breida et al. 2018). Therefore, when RB solution pH is inferior to pH = 5, the dye molecule is in cationic form (RBH+) as well as the membrane MB1000. Therefore, there are sufficient repulsive interactions between the membrane surface and cationic dye and hence a few rejection percent values (R = 25.84%). In other words, the protons compete with the positive dye cations, and hence, few RBH+ species are taken up by the clay minerals. Similar results were reported in the literature (Gupta et al. 2000; Arivoli et al. 2009; Anandkumar & Mandal 2011; Hou et al. 2011). However, when RB solution pHPZC >pH >5, RB is in zwitterionic form (RB±) regenerated from = N+, and COO functions and molecules species switched from cationic to anionic form. Consequently, the cationic form was reduced in favor of zwitterionic and anionic forms (Bhattacharyya et al. 2014; Pu et al. 2017; Rao et al. 2020). So, the membrane remains positively charged, implying stronger electrostatic attractions, and hence, a rejection percent value of 65.08% was obtained. For a pH solution greater than pHPZC, RB is in zwitterionic form, RB±, but the membrane is negatively charged promoting an electrostatic repulsion with the anionic predominant species. Moreover, the intermolecular interactions generated from COO and C = N+ functions of the RB molecules lead to some aggregation. As a result, the dye molecules either return to the feed solution under the transmembrane pressure effect or escape from the membrane's large pores, and hence the drop in rejection results (Figure 6). In the case of OG dye solution and as described in our previous work (Labied et al. 2022), the OG molecule presents two sulfonate groups (SO3) governing the UF process by engendered electrostatic interactions and hydrogen bonding one. So, for OG solution pH < pHpzc, the two anionic groups located at the end of the molecule underline a strong electrostatic attraction with the positively protonated material surface and hence an optimal rejection value of 75.89% (Figure 6) (Breida et al. 2018; Dev et al. 2021; Ouachtak et al. 2021). In this range of solution pH and considering the negative charge of the dye molecules, it may be able to compete with (OH) for the binding sites. However, the attraction of the anionic dye with the interlayer cationic minerals is greatly influenced by azo contacts and primarily π–π-dispersive interactions, generated by OG aromatic rings. The variations in the permeate flow of RB and OG dyes at different solutions’ pH values as a function of time are shown in Figure 7. The flow decreases with time whatever the pH solution. This is mainly attributed to the attraction phenomenon for pHpzc > 4, leading to such agglomeration of RB molecules and hence forming a cake layer, which blocks the pores (observation confirmed by the SEM images of the membrane). Besides, the flow increases at pH = 3.07, as a result of the repulsion forces between RB+ and positive membrane. Meanwhile, in the case of OG, the decrease of the flow for pH < pHpzc corresponds to the attraction phenomenon. The surface and the cross-sectional SEM images of the used membrane are illustrated in Figure 8(a) and 8(b). A dense microstructure is observed on the surface with a high porosity more sponge-like macro-voids and more roughness, which can be explained by the fact that not enough vitreous phase was present after sintering at 1,000 °C. Similar results were observed in the literature (Dong et al. 2007). These micrographs demonstrate the lack of any fissures and demonstrate the uniformity of the surface. It is evident that the membrane exhibits surface densification with grain boundary development. Additionally, it has been noted that membrane cavities have been seen. After dye filtration and at optimal solution pH, no agglomeration of RB was noticed. This result suggests the uptake of the molecule on the interlayer membrane matrix. However, for pH values superior to 4 and inferior to pHpzc, agglomerated particles with rounded contours can be seen (Figure 8(c)). The RB aggregation clearly appeared (Figure 8(c)), and a formation of cake and an irreversible fouling could be explained by the agglomeration of RB dye molecules into the inlet surface of membrane pores at these pH values, and this results in blocking of partially permeate transfer (Achiou et al. 2018b). It can be concluded that the richness of the bentonite membrane creates a significant impact on the membrane's microstructure, which is expected to have an impact on the permeability and selectivity of the filtration process (Benkhaya et al. 2019).
Figure 6

Dyes rejection percent as a function of solutions pH (C = 25 ppm and P = 1 bar).

Figure 6

Dyes rejection percent as a function of solutions pH (C = 25 ppm and P = 1 bar).

Close modal
Figure 7

Dyes permeates flows as a function of solution pH (C = 25 ppm, P = 1 bar).

Figure 7

Dyes permeates flows as a function of solution pH (C = 25 ppm, P = 1 bar).

Close modal
Figure 8

SEM images of MB0 1,000 (a) before dye filtration and (b) after dye filtration at pH = 5.36 (Experimental optimized pH value) and (c) at pH = 12.

Figure 8

SEM images of MB0 1,000 (a) before dye filtration and (b) after dye filtration at pH = 5.36 (Experimental optimized pH value) and (c) at pH = 12.

Close modal

Dye solution feed concentration effect

To highlight the effect of initial dye concentration on the flux and rejection parameters, using the optimal pressure of one bar, different filtration tests were conducted at different concentrations. As shown in Figure 9, when feed concentration rises from 25 to 100 ppm, permeate flux decreases from 151.86 to 115.15 L/m2.h for OG and from 65.50 to 51.97 L/m2.h for RB. This phenomenon may be caused by the rise in osmotic pressure, which is directly related to dye concentrations (Banerjee & De 2010). Remarkably, the rejection of OG is increased from 75.89 to 94.33% by increasing the dye concentration from 25 to 100 ppm. That of RB was enhanced from 65.08 to 89.38% (Figure 9).
Figure 9

Evolution of dyes flux (a) and rejection (b) as a function of feed concentration (t = 120 min, p = 1 bar).

Figure 9

Evolution of dyes flux (a) and rejection (b) as a function of feed concentration (t = 120 min, p = 1 bar).

Close modal

Ultrafiltration in simultaneous system (RB/OG)

The simultaneous membrane UF of OG and RB was examined in the following optimal experimental conditions: c = 100 ppm, for each dye, t = 120 min, and p = 1 bar and at different pH values. As described in Figure 10(a) and 10(b), a remarkable improvement of the OG and RB rejections is obtained with the simultaneous system (RB/OG) in comparison to the two systems considered separately. Therefore, optimal rejection values of 95.23 and 99.5% for OG and RB, respectively, were obtained at pH < pHpzc. This result may be related to the hydrogen bonding interactions generated from the azo functions of RB and hydroxyl functions, as well as the sulfonate groups of OG molecules (Deng et al. 2019). Also, molecular complexation may be developed, leading to an enhancement of dye's retention on the membrane surface and hence better rejection values. In other words, such a synergetic effect was created between the different function's dyes as well as the membrane surface, implying efficient membrane filtration (Figure 10(c)). A comparison between the performance of this study and another uptake of RB and OG dyes in different treatment scenarios is presented in Tables 1 and 2. Interestingly, the studied membrane exhibits competitive performances, indicating that it may be efficient for industrial wastewater treatment due to both its high filtration performance and inexpensive cost.
Table 1

Permeate flow and rejection percent of Rhodamine B and Orange G for different treatment scenarios

TreatmentFlux (L/m2.h)Rejection %Reference
Rhodamine B 
Ultrafiltration Water 30.00 – Saja et al. (2020)  
Permeate 80.00 80.10 
Microfiltration Water 200.17 – Tao et al. (2017)  
Permeate – 99.76 
Ultrafiltration: CNs Water 42.20 –  
Permeate 409.60 81.50 Abbas et al. (2020)  
Bentonite ultrafiltration Water 318.06 – 
Permeatea 51.97 89.38 This study 
Permeateb 44.53 99.50 
Orange G 
Nanofiltration Water 80.00 – Benfer et al. (2001)  
Permeate 20.00 30.00 
Nanofiltration:TiO2 Water 122.00 – Benfer et al. (2001)  
Permeate 41.00 47.30 
Ultrafiltration Water 141.00 – Ashok Kumar et al. (2019)  
Permeate 148.70 85.00 
Nanofiltration Water 30.40 – Yi et al. (2022)  
Permeate – 96.30 
Bentonite ultrafiltration Water 318.06 – This study 
Permeatec 115.15 94.33 
Permeated 48.72 95.23 
TreatmentFlux (L/m2.h)Rejection %Reference
Rhodamine B 
Ultrafiltration Water 30.00 – Saja et al. (2020)  
Permeate 80.00 80.10 
Microfiltration Water 200.17 – Tao et al. (2017)  
Permeate – 99.76 
Ultrafiltration: CNs Water 42.20 –  
Permeate 409.60 81.50 Abbas et al. (2020)  
Bentonite ultrafiltration Water 318.06 – 
Permeatea 51.97 89.38 This study 
Permeateb 44.53 99.50 
Orange G 
Nanofiltration Water 80.00 – Benfer et al. (2001)  
Permeate 20.00 30.00 
Nanofiltration:TiO2 Water 122.00 – Benfer et al. (2001)  
Permeate 41.00 47.30 
Ultrafiltration Water 141.00 – Ashok Kumar et al. (2019)  
Permeate 148.70 85.00 
Nanofiltration Water 30.40 – Yi et al. (2022)  
Permeate – 96.30 
Bentonite ultrafiltration Water 318.06 – This study 
Permeatec 115.15 94.33 
Permeated 48.72 95.23 

aRB: single system.

bRB: blend system.

cOG: Single system.

dOG: blend system.

Table 2

Comparison between the performance of current study and other uptake of RB and OG dyes at different condition

DyeMembrane materialWater permeability (L/m2.h.bar)Feed concentration (ppm)Rejection (%)Reference
Rhodamine B Bentonite + perlite 30 100 80.10 Sawunyama et al. (2023)  
Natural silk + nanofibrils + SNFs 13,000 2.39 91.00 Deng et al. (2019)  
Coal carbon + electric field 200.17 100 99.00 Ling et al. (2016)  
Polyvinylidene fluoride + graphitic + carbon nitride (g-C3N4– 10 96.00 Tao et al. (2017)  
Graphene oxide (GO) 399.04 1.5 85.03 Kolesnyk et al. (2020)  
Graphene oxide/polyacrylamide (GO/PAM) 188.89 1.5 95.43 Kolesnyk et al. (2020)  
Reduced graphene oxide (rGO) 85.85 1.5 97.06 Kolesnyk et al. (2020)  
Orange G Natural silk + nanofibrils + SNFs 13,000 271.88 82.00 Deng et al. (2019)  
 ZrO2 20 200 30.00 Cheng et al. (2019)  
TiO2 41 200 47.30 Cheng et al. (2019)  
Polyethersulfone; polyethylene glycol (PES/PEG) 65 40 72.50 Benfer et al. (2001)  
Polyethersulfone; polyethylene glycol (PES/PEG) + Particle-activated carbon – 40 100.00 Benfer et al. (2001)  
Blended polyethersulfone 148.7 100 85.00 Dong et al. (2011)  
Rhodamine B/Orange G Bentonite 318.06 100 89.38a
94.33b
99.50c
95.23d 
This study 
DyeMembrane materialWater permeability (L/m2.h.bar)Feed concentration (ppm)Rejection (%)Reference
Rhodamine B Bentonite + perlite 30 100 80.10 Sawunyama et al. (2023)  
Natural silk + nanofibrils + SNFs 13,000 2.39 91.00 Deng et al. (2019)  
Coal carbon + electric field 200.17 100 99.00 Ling et al. (2016)  
Polyvinylidene fluoride + graphitic + carbon nitride (g-C3N4– 10 96.00 Tao et al. (2017)  
Graphene oxide (GO) 399.04 1.5 85.03 Kolesnyk et al. (2020)  
Graphene oxide/polyacrylamide (GO/PAM) 188.89 1.5 95.43 Kolesnyk et al. (2020)  
Reduced graphene oxide (rGO) 85.85 1.5 97.06 Kolesnyk et al. (2020)  
Orange G Natural silk + nanofibrils + SNFs 13,000 271.88 82.00 Deng et al. (2019)  
 ZrO2 20 200 30.00 Cheng et al. (2019)  
TiO2 41 200 47.30 Cheng et al. (2019)  
Polyethersulfone; polyethylene glycol (PES/PEG) 65 40 72.50 Benfer et al. (2001)  
Polyethersulfone; polyethylene glycol (PES/PEG) + Particle-activated carbon – 40 100.00 Benfer et al. (2001)  
Blended polyethersulfone 148.7 100 85.00 Dong et al. (2011)  
Rhodamine B/Orange G Bentonite 318.06 100 89.38a
94.33b
99.50c
95.23d 
This study 

aRB: single system.

bOG: single system.

cRB: blend system.

dOG: blend system.

Figure 10

Evolution of ultrafiltration in single (a) and simultaneous (RB/OG) (b) systems, under optimal conditions (c = 100 ppm, t = 120 min, p = 1 bar), and scheme of interactions developed in the simultaneous (RB/OG) (c).

Figure 10

Evolution of ultrafiltration in single (a) and simultaneous (RB/OG) (b) systems, under optimal conditions (c = 100 ppm, t = 120 min, p = 1 bar), and scheme of interactions developed in the simultaneous (RB/OG) (c).

Close modal

Bentonite was used to prepare an UF membrane (MB1000), for removing RB and OG dyes from aqueous solutions in single and simultaneous (RB/OG) systems, respectively. The effect of transmembrane pressure, solution pH, and concentration on the dye tangential UF rejection percent was investigated, and the optimized parameters were described. The tangential UF rejection percent of RB and OG in the simultaneous systems (RB/OG) was remarkably important compared to similar single UF systems. So, their coexistence improves their UF in a synergistic manner. Finally, the results obtained in this study confirm that the used bentonite can be considered as a key material for membrane synthesis. Its low cost and soft elaboration make it a promoting bio-based membrane for industrial wastewater treatment.

The authors acknowledge that the financial support from the Ministry of Higher Education and Scientific Research (MESRS) and the General Agency of Scientific Research and Technological Development (DGRSDT), in the frame of the national projects is grateful.

All relevant data are included in the paper or its Supplementary Information.

The authors declare there is no conflict.

Abbas
H. M.
,
Kamel
R. M.
,
Ali
M. E. A.
,
Atta
A. H.
,
Hassan
H. M. A.
&
Shahat
A.
2020
Fabrication of polysulfone/carbon nanospheres ultrafiltration membranes for removing some dyes from aqueous solutions
.
D.W.T.
193
,
57
63
.
Abdullayev
A.
,
Bekheet
M. F.
,
Hanaor
D. A. H.
&
Gurlo
A.
2019
Materials and applications for low-cost ceramic membranes
.
Membranes
9
,
1
31
.
Achiou
B.
,
Elomari
H.
,
Bouazizi
A.
,
Karim
A.
,
Ouammou
M.
,
Albizane
A.
,
Bennazha
J.
,
Alami Younssi
S.
&
El Amrani
I. E.
2017
Manufacturing of tubular ceramic microfiltration membrane based on natural pozzolan for pre-treatment of seawater desalination
.
Desalination
419
,
181
187
.
Achiou
B.
,
Elomari
H.
,
Ouammou
M.
,
Albizane
A.
,
Bennazha
J.
,
Aaddane
A.
&
Alami Younssi
S.
2018a
Study of added starch on characteristics of flat ceramic microfiltration membrane made from natural Moroccan pozzolan
.
J. Mater. Environ. Sci.
9
,
1013
1021
.
Achiou
B.
,
Beqqour
D.
,
Elomari
H.
,
Bouazizi
A.
,
Ouammou
M.
,
Bouhria
M.
,
Aaddane
A.
,
Khiat
K.
&
Alami Younssi
S.
2018b
Preparation of inexpensive NaA zeolite membrane on pozzolan support at low temperature for dehydration of alcohol solutions
.
J. Environ. Chem. Eng.
6
,
4429
4437
.
Al-Futaisi
A.
,
Jamrah
A.
&
Al-Hanai
R.
2007
Aspects of cationic dye molecule adsorption to palygorskite
.
Desalination
214
,
327
342
.
Aloulou
H.
,
Bouhamed
H.
,
Daramola
M. O.
,
Khemakhem
S.
&
Ben Amar
R.
2020
Fabrication of asymmetric ultrafiltration membranes from natural zeolite and their application in industrial wastewater treatment
.
Euro-Mediterr. J. Environ. Integr.
36
(
5
),
1
11
.
Arivoli
S.
,
Thenkuzhali
M.
&
Devaprasath
P. M.
2009
Adsorption of Rhodamine B by acid activated carbon kinetic, thermodynamic and equilibrium studies
.
Orbital
.
1
,
138
155
.
Ashok Kumar
S.
,
Srinivasan
G.
&
Govindaradjane
S.
2019
Development of a new blended polyethersulfone membrane for dye removal from synthetic wastewater
.
Environ. Nanotechnol., Monit. Manag.
12
,
100238
.
Barrouk
I.
,
Alami Younssi
S.
,
Kabbabi
A.
,
Persin
M.
,
Albizane
A.
&
Tahiri
S.
2014
New ceramic membranes from natural Moroccan phosphate for microfiltration application
.
Desalin. Water Treat.
55
,
1
8
.
Benfer
S.
,
Popp
U.
,
Richter
H.
,
Siewert
C.
&
Tomandl
G.
2001
Development and characterization of ceramic nanofiltration membranes
.
Sep. Purif. Technol.
22–23
,
231
237
.
Benkhaya
S.
,
Achiou
B.
,
Ouammou
M.
,
Bennazha
J.
,
Alami Younssi
S.
,
M'rabet
S.
&
El Harfi
A.
2019
Preparation of low-cost composite membrane made of polysulfone/polyetherimide ultrafiltration layer and ceramic pozzolan support for dyes removal
.
Mater. Today Commun.
19
,
212
219
.
Bhatt
A. S.
,
Sakaria
P. L.
,
Vasudevan
M.
,
Pawar
R. R.
,
Sudheesh
N.
,
Bajaj
H. C.
&
Mody
H. M.
2012
Adsorption of an anionic dye from aqueous medium by organoclays: Equilibrium modelling, kinetic and thermodynamic exploration
.
R. Soc. Chem.
2
,
8663
8671
.
Bhattacharyya
K. G.
,
SenGupta
S.
&
Sarma
G. K.
2014
Interactions of the dye, Rhodamine B with kaolinite andmontmorillonite in water
.
Applied Clay Science
99
,
7
17
.
Bouarroudj
T.
,
Aoudjit
L.
,
Djahida
L.
,
Zaidi
B.
,
Ouraghi
M.
,
Zioui
D.
,
Mahidine
S.
,
Shekhar
C.
&
Bachari
K.
2021
Photodegradation of tartrazine dye favored by natural sunlight on pure and (Ce, Ag) co-doped ZnO catalysts
.
Water Sci. Technol.
83
,
2118
2134
.
Bouazizi
A.
,
Saja
S.
,
Achiou
B.
,
Ouammou
M.
,
Calvo
J. I.
,
Aaddane
A.
&
Younssi
S. A.
2016
Elaboration and characterization of a new flat ceramic MF membrane made from natural Moroccan bentonite. Application to treatment of industrial wastewater
.
Appl. Clay Sci.
132–133
,
33
40
.
Bouazizi
A.
,
Breida
M.
,
Karim
A.
,
Achiou
B.
,
Ouammou
M.
,
Calvo
J. I.
,
Aaddane
A.
,
Khiat
K.
&
Alami Younssi
S.
2017a
Development of a new TiO2 ultrafiltration membrane on flat ceramic support made from natural bentonite and micronized phosphate and applied for dye removal
.
Ceram. Int.
43
,
1479
1487
.
Bouazizi
A.
,
Breida
M.
,
Achiou
B.
,
Ouammou
M.
,
Calvo
J. I.
,
Aaddane
A.
&
Younssi
S. A.
2017b
Removal of dyes by a new nano–TiO2 ultrafiltration membrane deposited on low-cost support prepared from natural Moroccan bentonite
.
Appl. Clay Sci.
149
,
127
135
.
Bouzerara
F.
,
Harabi
A.
,
Achour
S.
&
Larbot
A.
2006
Porous ceramic supports for membranes prepared from kaolin and doloma mixtures
.
J. Eur. Ceram. Soc.
26
,
1663
1671
.
Breida
M.
,
Alami Younssi
S.
,
Bouazizi
A.
,
Achiou
B.
,
Ouammou
M.
&
El Rhazi
M.
2018
Nitrate removal from aqueous solutions by γ-Al2O3 ultrafiltration membranes
.
Heliyon.
4
,
e00498
.
Cheng
M. M.
,
Huang
L. J.
,
Wang
Y. X.
,
Zhao
Y. C.
,
Tang
J. G.
,
Wang
Y.
,
Zhang
Y.
,
Hedayati
M.
,
Kipper
M. J.
&
Wichramasinghe
S. R.
2019
Synthesis of graphene oxide/polyacrylamide composite membranes for organic dyes/water separation in water purification
.
J. Mater. Sci.
54
,
252
264
.
Chieng
H. I.
,
Priyantha
N.
&
Lim
L. B.
2015
Effective adsorption of toxic brilliant green from aqueous solution using peat of Brunei Darussalam: Isotherm, thermodynamics, kinetics and regeneration studies
. RSC Adv. 5,
34603–34615
.
Das
P. P.
,
Mondal
P.
,
Sinha Anweshan
A.
,
Biswas
P.
,
Sarkar
S.
&
Purkait
M. K.
2021
Treatment of steel plant generated biological oxidation treated (BOT) wastewater by hybrid process
.
Sep. Purif. Technol.
258
,
118013
118029
.
Dasgupta
J.
,
Sikder
J.
,
Chakraborty
S.
,
Curcio
S.
&
Drioli
E.
2015
Remediation of textile effluents by membrane-based treatment techniques: A state of the art review
.
J. Environ. Manage.
147
,
55
72
.
de Almeida
A. S. V.
,
de Figueiredo Neves
T.
,
da Silva
M. G. C.
,
Prediger
P.
&
Vieira
M. G. A.
2022
Synthesis of a novel magnetic composite based on graphene ox-ide, chitosan and organoclay and its application in the removal of bisphenol A, 17 α-ethinylestradiol and triclosan
.
Powder Technol.
395
,
695
707
.
Deng
H.
,
Mao
Z.
,
Xu
H.
,
Zhang
L.
,
Zhong
Y.
&
Sui
X.
2019
Synthesis of fibrous La feo3 perovskite oxide for adsorption of rhodamine B
.
Ecotoxicol. Environ. Saf.
168
,
35
44
.
Dev
V. V.
,
Wilson
B.
,
Nair
K. K.
,
Antony
S.
&
Krishnan
K. A.
2021
Response surface modeling of orange-G adsorption onto surface tuned ragi husk
.
Colloid Interface Sci. Commun.
41
,
100363
100373
.
Djamaa
Z.
,
Lerari
D.
,
Mesli
A.
&
Bachari
K.
2019
Poly (acrylic acid-co-styrene)/clay nanocomposites: Efficient adsorbent for methylene blue dye pollutant
.
Int. J. Plast. Technol.
23
,
110
121
.
Dong
Y.
,
Feng
X.
,
Dong
D.
,
Wang
S.
,
Yang
J.
,
Gao
J.
,
Liu
X.
&
Meng
G.
2007
Elaboration and chemical corrosion resistance of tubular macro-porous cordierite ceramic membrane supports
.
J. Membr. Sci.
304
,
65
75
.
Dong
Y.
,
Su
Y.
,
Chen
W.
,
Peng
J.
,
Zhang
Y.
&
Jiang
Z.
2011
Ultrafiltration enhanced with activated carbon adsorption for efficient dye removal from aqueous solution
.
Chin. J. Chem. Eng.
19
(
5
),
863
869
.
El Gaayda
J.
,
Ait Akbour
R.
,
Titchou
F. E.
,
Afanga
H.
,
Zazou
H.
,
Swanson
C.
&
Hamdani
M.
2021
Uptake of an anionic dye from aqueous solution by aluminum ox-ide particles: Equilibrium, kinetic, and thermodynamic studies
.
Groundw. Sus-tain. Dev.
12
,
100540
100552
.
Foorginezhad
S.
&
Zerafat
M. M.
2017
Microfiltration of cationic dyes using nano-clay membranes
.
Ceram. Int.
43
,
15146
15159
.
Gupta
V. K.
,
Mohan
D.
,
Sharma
S.
&
Sharma
M.
2000
Removal of basic dyes (Rhodamine B and methylene blue) from aqueous solutions using bagasse fly ash
.
Sep. Sci. Technol.
35
,
2097
2113
.
Harabi
A.
,
Zenikheri
F.
,
Boudaira
B.
,
Bouzerara
F.
,
Guechi
A.
&
Foughali
L.
2014
A new and economic approach to fabricate resistant porous membrane supports using kaolin and CaCO3
.
J. Eur. Ceram. Soc.
34
,
1329
1340
.
Hou
M. F.
,
Mac
C. X.
,
Zhang
W. D.
,
Tang
X. Y.
,
Fan
Y. N.
&
Wan
H. F.
2011
Removal of rhodamine B using iron-pillared bentonite
.
J. Hazard. Mater.
186
,
1118
1123
.
Jiang
M.
,
Ye
K.
,
Lin
J.
,
Zhang
X.
,
Ye
W.
,
Zhao
S.
&
Van der Bruggen
B.
2018
Effective dye purification using tight ceramic ultrafiltration membrane
.
J. Membr. Sci.
566
,
151
160
.
Karim
A.
,
Achiou
B.
,
Bouazizi
A.
,
Aaddane
A.
,
Ouammou
M.
,
Bouziane
M.
,
Bennazha
J.
&
Alami Younssi
S.
2018
Development of reduced graphene oxide membrane on flat Moroccan ceramic pozzolan support. Application for soluble dyes removal
.
J. Environ. Chem. Eng.
6
,
1475
1485
.
Kolesnyk
I.
,
Kujawa
J.
,
Bubela
H.
,
Konovalova
V.
,
Burban
A.
,
Cyganiuk
A.
&
Kujawski
W.
2020
Photocatalytic properties of PVDF membranes modified with g-c3n4 in the process of Rhodamines decomposition
.
Sep. Purif. Technol.
250
,
117231
.
Labied
R.
,
Ouraghi
M.
,
Hazam
S.
,
Touahra
F.
&
Lerari
D.
2022
Effect of porogen agent on bio-based membranes filtration performances: Experimental and theoritical study
.
Journal of Molecular Structure.
1269
,
133863
.
Ling
S.
,
Jin
K.
,
Kaplan
D. L.
&
Buehler
M. J.
2016
Ultrathin free-standing Bombyx mori silk nanofibril membranes
.
ACS Publications.
16
,
3795
3800
.
Liu
Q.
,
Li
Y.
,
Chen
H.
,
Lu
J.
,
Yu
G.
&
Möslang
M.
2020
Superior adsorption capacity of functionalised straw adsorbent for dyes and heavy-metal ions
.
J. Hazard.Mater.
382
,
121040
.
Mekidiche
M.
,
Khaldi
K.
,
Nacer
A.
,
Boudjema
S.
,
Ameur
N.
,
Lerari-Zinai
D.
,
Bachari
K.
&
Choukchou- Braham
A.
2021
Organometallic modified montmorillonite application in the wastewater purification: Pollutant photodegradation and anti-bacterial efficiencies
.
Appl. Surf. Sci.
569
,
151097
151108
.
Minitha
C. R.
,
Lalitha
M.
,
Jeyachandran
Y. L.
,
Senthilkumar
L.
&
Rajendra Kumar
R. T.
2017
Adsorption behaviour of reduced graphene oxide towards cationic and anionic dyes: Co-action of electrostatic and π – π interactions
.
Mater. Chem. Phys.
194
,
243
252
.
Mohamed
A. A. J.
,
Ali Vuai
L.
,
Kombo
M.
&
Chukwuma
O. J.
2019
Removal of selected metal ions using powder of seeds of Ajwaa dates from aqueous solution
.
J Anal Pharm Res.
8
,
228
232
.
Mouiya
M.
,
Abourriche
A.
,
Bouazizi
A.
,
Benhammou
A.
,
El Hafiane
Y.
,
Abou-liatim
Y.
,
Nibou
L.
,
Oumam
M.
,
Ouammou
M.
,
Smith
A.
&
Hannache
H.
2018
Flat ceramic microfiltration membrane based on natural clay and Moroccan phosphate for desalination and industrial wastewater treatment
.
Desalination
427
,
42
50
.
Mouiya
M.
,
Bouazizi
A.
,
Abourriche
A.
,
Benhammou
A.
,
El Hafiane
Y.
,
Ouammou
M.
,
Abouliatim
Y.
,
Alami Younssi
S.
,
Smith
A.
&
Hannache
H.
2019
Fabrication and characterization of a ceramic membrane from clay and banana peel powder: Application to industrial wastewater treatment
.
Mater. Chem. Phys.
227
,
291
301
.
Ouachtak
H.
,
El Guerdaoui
A.
,
Haounati
R.
,
Akhouairi
S.
,
El Haouti
R.
,
Hafid
N.
,
Ait Addi
A.
,
Šljukic
B.
,
Santos
D. M. F.
&
Labd Taha
M.
2021
Highly efficient and fast batch adsorption of orange G dye from polluted water using superb organo montmorillonite: Experimental study and molecular dynamics investigation
.
J. Mol. Liq.
335
,
116560
116575
.
Ouaddari
H.
,
Karim
A.
,
Achiou
B.
,
Saja
S.
,
Aaddane
A.
,
Bennazha
J.
,
El Amrani El Hassani
I.
,
Ouammou
M.
&
Albizane
A.
2019
New low-cost ultrafiltration membrane made from purified natural clays for direct Red 80 dye removal
.
J. Environ. Chem. Eng.
7
,
103268
103278
.
Rao
W.
,
Piliouras
P.
,
Wang
X.
,
Guido
A.
,
Kugler
K.
,
Sieren
B.
,
Wang
L.
,
Lv
G.
&
Li
Z.
2020
Zwitterionic dye rhodamine B (RhB) uptake on different types of clay minerals
.
Appl. Clay Sci.
197
,
105790
.
Saavedra-Labastida
E.
,
Díaz-Nava
M. C.
,
Illescas
J.
&
Muro
C.
2019
Comparison of the removal of an anionic dye from aqueous solutions by adsorption with organically modified clays and their composites
.
Water Air Soil Pollut.
230
,
88
114
.
Saja
S.
,
Bouazizi
A.
,
Achiou
B.
,
Ouaddari
H.
,
Karim
A.
,
Ouammou
M.
,
Aaddane
A.
,
Bennazha
J.
&
Alami Younssi
S.
2020
Fabrication of low-cost ceramic ultrafiltration membrane made from bentonite clay and its application for soluble dyes removal
.
J. Eur. Ceram. Soc.
40
,
2453
2462
.
Sawunyama
L.
,
Oyewo
O. A.
,
Saheri
N.
,
Onjefu
S. A.
&
Onwudiwe
D. C.
2023
Metal oxide functionalized ceramic membranes for the removal of pharmaceuticals in wastewater
.
Surf. Inter Surf.
38
,
102787
.
Seo
Y.-D.
,
Oh
S.-Y.
,
Rajagopal
R.
&
Ryo
K.-S.
2020
FeS-biochar and Zn (0)-biochar for remediation of redox-reactive contaminants
.
RSC. Adv.
10
(
50
),
30203
30213
.
Sharma
P.
,
Boraha
D. J.
,
Dasb
P.
&
Dasa
M. R.
2015
Cationic and anionic dye removal from aqueous solution using montmorillonite clay: Evaluation of adsorption parameters and mechanism
.
Desalin. Water Treat.
57
,
1
17
.
Silva
M. M. F.
,
Oliveira
M. M.
,
Avelino
M. C.
,
Fonseca
M. G.
,
Almeida
R. K. S.
&
Filho
E. C. S.
2012
Adsorption of an industrial anionic dye by modified-KSF-montmorillonite: Evaluation of the kinetic, thermodynamic and equilibrium data
.
Chem. Eng. J.
203
,
259
268
.
Tao
P.
,
Xu
Y.
,
Song
C.
,
Yin
Y.
,
Yang
Z.
,
Wen
S.
,
Wang
S.
,
Liu
H.
,
Li
S.
,
Li
C.
,
Wang
T.
&
Shao
M.
2017
A novel strategy for the removal of rhodamine B (RhB) dye from wastewater by coal-based carbon membranes coupled with the electric field
.
Sep. Purif. Technol.
179
,
175
183
.
Tolkachev
O. V.
&
Bespamyatnykh
E. N.
2019
The new method of rhodamine mark detection and its application possibilities in zoological studies
.
J. Sib. Fed. Univ. Biol.
12
,
352
365
.
Ursino
C.
,
Castro-Muñoz
R.
,
Drioli
E.
,
Gzara
L.
,
Albeirutty
M. H.
&
Figoli
A.
2018
Progress of nano composite membranes for water treatment
.
Membranes
8
(
18
),
1
40
.
Wang
J.
,
Gao
M.
,
Shen
T.
,
Yu
M.
,
Xiang
Y.
&
Liu
J.
2019
Insights into the efficient adsorption of rhodamine B on tunable organo-vermiculites
.
J. Hazard. Mater.
366
,
501
511
.
Zhang
C.
,
Xu
Z.
,
Hu
Y.
,
He
J.
,
Tian
M.
,
Zhou
J.
,
Zhou
Q.
,
Chen
S.
,
Chen
D.
,
Chen
P.
&
Sun
W.
2019
Novel insights into the hydroxylation behaviors of α-quartz (101) surface and its effects on the adsorption of sodium oleate
.
Minerals
9
,
450
568
.
This is an Open Access article distributed under the terms of the Creative Commons Attribution Licence (CC BY-NC-ND 4.0), which permits copying and redistribution for non-commercial purposes with no derivatives, provided the original work is properly cited (http://creativecommons.org/licenses/by-nc-nd/4.0/).