This study reports the feasibility of recycled polyvinylidene difluoride (PVDF) beads to decolourize methylene blue (MB) from aqueous streams. The beads were characterized using scanning electron microscopy (SEM), X-ray powder diffraction (XRD), thermogravimetric analysis (TGA), and Fourier transform infrared spectroscopy (FT-IR) for its morphological and structural analysis. The effect of various process parameters such as adsorbent dose, initial concentration, contact time, and pH was studied. The first principle density functional theory (DFT) calculations were performed to investigate the underlying mechanism behind the adsorption process. The MB dye adsorption on recycled PVDF beads followed the pseudo-second-order kinetics and Langmuir isotherm, indicating the adsorption was chemical and monolayer. The maximum adsorption capacity obtained was 27.86 mg g−1. The adsorption energy of MB-PVDF predicted from the DFT study was –64.7 kJ mol−1. The HOMO-LUMO energy gap of PVDF decreased from 9.42 eV to 0.50 eV upon interaction with MB dye due to the mixing of molecular orbitals. The DFT simulations showed that the interaction of the MB dye molecule was from the electronegative N atom of the MB dye molecule, implying that electrostatic interactions occurred between the recycled PVDF beads and the positively charged quaternary ammonium groups in MB dye. The present study demonstrates the potential of recycled PVDF beads for a low-cost dye removal technique from textile wastewater.

  • Experiments and DFT calculations were performed to evaluate the binding of MB dye to PVDF beads synthesized from used membranes.

  • The adsorption of MB dye on PVDF beads was monolayer and chemical in nature, with a maximum adsorption capacity of 27.86 mg g−1.

  • The DFT study predicted the adsorption energy of 64.7 kJ mol−1 for MB-PVDF, indicating the strong interaction between MB dye and PVDF.

  • The study demonstrated the potential of recycled PVDF beads for a low-cost dye removal technique.

Graphical Abstract

Graphical Abstract
Graphical Abstract

In the past few decades, membranes have been used excessively for water and wastewater treatment, leading to the increased disposal of waste membranes (Ezugbe et al. 2020). At the current rate, the disposal of membrane modules exhibits significant and escalating detrimental impacts, leading to the need to limit the direct disposal of these modules. Currently, used membranes are disposed of in landfills or incinerated. However, these methods are not completely dependable because of the environmental impacts. The discarded membranes can be recycled and reused for different separation processes with less demanding specifications with some post-processing. Several methods have been employed to recycle membranes, such as chemical cleaning (Yi et al. 2017), physical cleaning and backwashing (Park et al. 2018), and direct recycling of the various module components (Lawler et al. 2012). Paula et al. (2017) examined the technological viability of desalination membrane recycling by chemical oxidation. The end-of-life membranes deteriorate due to fouling, biofouling, etc. The polymer recovered from the membranes can be used as adsorbents to remove hazardous contaminants from wastewater. However, very few studies are available where polymers recovered from waste membranes have been used for adsorption (Zwain et al. 2014).

Synthetic dyes are found in huge quantities in textile and dye processing industrial effluent, posing a significant environmental threat. Many industries produce these dyes, including textiles, printing, paper, and plastic (Ngulube et al. 2017; Yadav et al. 2021a, 2021b, 2021c). Methylene blue (MB) is a cationic dye found in considerable quantities in textile effluent released into the environment without the significant treatment of its release into the surrounding ecosystem (Yadav et al. 2021a, 2021b, 2021c). At present, biological, physical, and chemical methods are employed for textile dye removal from wastewater (Yaseen & Scholz 2019). However, there are many shortcomings associated with these traditional technologies for dye removal, including chemical release, secondary sludge, high initial cost, and low decolourization efficiency (Piaskowski et al. 2018). Adsorption is one of the most economical, effective, and simple processes for the decolourization of textile water. Furthermore, it shows a wide range of materials such as nanomaterials, carbonaceous materials, bioadsorbents, etc. (Tran et al. 2019).

In recent years, polymeric adsorbents have become popular as an alternative to traditional adsorbents such as clays and activated carbon due to their tunable physicochemical characteristics, structural variability, reusability, and selectivity (Mok et al. 2020; Yadav & Sinha 2021). Zheng et al. (2018) studied the feasibility of polymer-functionalized magnetic nanoparticles for MB removal. Fe3O4-incorporated carboxyl functionalized nanoporous polymer showed high performance for MB dye adsorption from wastewater (Su et al. 2018). Carbon nanotubes-based polymer nanocomposites were found to be suitable for the adsorption of MB from the aqueous solution (Gan et al. 2020). Several researchers have used waste material to remove dyes from textile wastewater. For instance, Kiran et al. (2020) used agro-industrial waste (palm date stones) to adsorb basic violet 3 and basic red 2 from textile wastewater with high removal efficiency (77%- basic red 2, 93% basic violet 3). Wong et al. (2020) used coffee waste modified with polyethyleneimine to adsorb dye from textile wastewater with an adsorption capacity of 34.36 mg g−1 (Congo red) and 77.52 mg g−1 (reactive black 5). Temesgen et al. (2018) used activated orange and banana peel to eliminate reactive red dye from textile industry wastewater with very high removal efficiency (>89%). Vecino et al. (2015) synthesized biocomposites from vineyard waste entrapped in calcium alginate hydrogel beads to eliminate dye from wastewater and achieved 74.6% dye removal efficiency. Polyvinyledienefluoride (PVDF) and its co-polymer have been widely used to treat textile wastewater. Zhang et al. (2019) fabricated a PVDF/GO/ZnO composite to remove MB dye through photocatalytic degradation, achieving a removal efficiency of 86.84%.

A careful literature survey shows that PVDF effectively adsorbs MB dye from an aqueous stream. Hence, in this study, PVDF polymer from a recycled membrane was used as an adsorbent for removing MB dye from textile wastewater. The PVDF beads obtained from recycled membranes were characterized using scanning electron microscopy (SEM), X-ray powder diffraction (XRD), thermogravimetric analysis (TGA), and Fourier transform infrared spectroscopy (FT-IR) for morphological and structural analysis. A parametric study was performed to study the effect of varying adsorbent dose, initial concentration, contact time, and pH on the beads’ adsorption capacity and removal efficiency. Moreover, density functional theory (DFT) calculations, adsorption isotherm, and kinetics were studied to identify the adsorption process' mechanism.

Materials

The recycled PVDF was obtained from used membranes. Dimethylformamide (DMF) solvent (purity ∼99.5%) was purchased from Spectrochem Pvt. Ltd Mumbai, India. MB dye (purity ∼82%) was procured from NICE chemicals Pvt. Ltd, India. Deionized (DI) water (∼18 M Ω) was obtained from the Millipore Q BIOCEL unit, Millipore.

Adsorbent synthesis

The recycled membrane was washed thoroughly, and the PVDF (16 g) was delaminated from the fabric and dissolved in DMF solvent (100 ml) for 6 h at 60 °C under constant stirring. The PVDF solution (16 wt.%) was taken in a clean syringe and added dropwise to a beaker filled with DI water (coagulation bath). The recycled PVDF polymer precipitated in the form of spherical beads. The beads were kept in the coagulation bath overnight to complete the phase inversion process.

Adsorption experiments

Adsorption experiments were performed to investigate recycled PVDF beads’ adsorption capacity and removal efficiency for MB dye. The batch experiments were performed by varying different parameters, such as adsorbent dosage, pH, initial dye concentration, and contact time to assess the optimal conditions for the adsorption process. To examine the effect of the adsorbent dose for MB dye, adsorption was varied from 0.2 to 3 g L−1. To assess the effect of MB dye concentration on the recycled PVDF beads, the solution concentration was varied from 10 to 250 ppm. The MB dye solution's pH was varied from 2 to 12 to assess its effect. The pH adjustment of the dye solution was achieved using 0.1 N HCl or 0.1 N NaOH. To optimize the adsorption time, the contact time of the adsorbent was varied in the range of 5–240 mins. The adsorption capacity and removal efficiency was estimated from the equations shown below:
(1)
(2)
where qe – equilibrium adsorption capacity, Ci – MB dye's initial concentration, Ce – MB dye's equilibrium concentration, m – mass of the recycled PVDF beads, and V – volume of MB dye's solution.

Adsorption kinetics

Time is critical in studying an adsorbent's adsorption kinetics as the adsorption process changes with time. The kinetics of the adsorption process help predict the pathway by which the adsorption process must have taken place (Ivanets et al. 2019). In the case of the adsorption, the physiochemical characteristics of the adsorbent and system parameters such as temperature and contact time determine the nature of the process (Dindorkar et al. 2022a). The amount of dye adsorbed at time t, denoted by qt, was calculated by using the following equation:
(3)
where qt – adsorbed ions amount per unit mass of the adsorbent in time t, m – mass of adsorbent used, and Ct – dye's concentration at instant t. To investigate the adsorption kinetics in this study, the non-linear pseudo-first-order (PFO) and the pseudo-second-order (PSO) kinetic models were used. The non-linear form of the PFO and PSO equation is given below:
(4)
(5)
The integration of Equations (4) and (5) for the boundary conditions (t=0, qt=0 and t=t, qe=qt) results in the following equations
(6)
(7)
where qt – amount of adsorbed ions per unit adsorbent at instant t, k1 – PFO constant, and k2 – rate constant of the PSO.

Adsorption isotherm

The adsorption isotherms provide the equilibrium concentration between the adsorbed and unadsorbed phase at a particular condition. The Langmuir and Freundlich isotherms were chosen for this investigation. The Langmuir adsorption (monolayer adsorption) isotherm assumes that the adsorption occurs within the adsorbent at specific homogeneous sites, while the Freundlich isotherms assume the heterogeneous adsorption sites (multilayer adsorption) (Patel and Yadav, 2022). The non-linear form of the Langmuir isotherm is described below.
(8)
where b – constant for Langmuir isotherm, qm – maximum adsorption capacity, Ce – equilibrium dye concentration. The non-linear form of the equation of Freundlich isotherm is described as:
(9)
where n – intensity and KF – Freundlich coefficient.

Computational details

The Gaussian 09 package was used to perform the first-principle DFT calculations (Wallingford CT 2013). For the geometry optimizations, a hybrid CAM-B3LYP functional was used with an IEF-PCM model to reflect the long-range corrections (Yanai et al. 2004; Caldeweyher et al. 2017). Theoretical FT-IR spectra, HOMO-LUMO distribution and density of states (DOS) plots were obtained using GaussSum (O'Boyle et al. 2008) software. The basis set superposition error (BSSE) was incorporated while calculating adsorption energy using counterpoise correction (Boys & Bernardi 1970). The following equation was used to calculate the adsorption energy.
(10)
where the term, total energy of the MB-PVDF complex cluster, energy of PVDF and energy of the MB dye. The reactivity descriptors (chemical hardness, chemical potential, and electrophilicity index) were evaluated using Koopmans’ theorem (Yadav & Dindorkar 2022a, 2022b).

Characterization techniques

The morphological characteristics of the adsorbent composite beads were studied from SEM (JEOL JSM 7100F). The adsorbent IR spectrography was performed using the Perkin Elmer FTIR spectrometer (USA). The thermogravimetric analysis (TGA) was accomplished with a thermogravimetric analyzer (Mettler Toledo) with nitrogen flow at a heating rate of 10° C min−1. The measurement of MB dye concentrations was assessed using a UV-Vis spectrophotometer by constructing a calibration curve, and for pH monitoring, a Eutech PC2700 multiparameter device (Shimadzu, Japan) was used. The characteristic absorbance of MB dye at 663 nm was chosen to study the decolourization during the adsorption.

Characterization studies

Figure 1 illustrates the surface and cross-sectional morphology of recycled PVDF beads. The surface morphology of the beads was rough and porous. The recycled PVDF beads’ internal porous and spongy cross-section morphology was attributed to the solvent and non-solvent exchange during the phase inversion process (Zahirifar et al. 2019). The recycled PVDF beads’ rough and porous morphology will provide a larger surface area and an adsorption process with more active sites.

Figure 1

SEM images of recycled PVDF beads (a) surface and (b) cross-section.

Figure 1

SEM images of recycled PVDF beads (a) surface and (b) cross-section.

Close modal

TGA was conducted to study the thermal stability of the synthesized recycled PVDF beads (Figure 2(a)). Till 427 °C, there was a loss of ∼3% mass, attributed to the evaporation of humidity entrapped in the beads. When the temperature reached 427 °C, the residual mass declined rapidly to 34%, attributed to the thermal degradation of the PVDF polymer's −(C2H2F2)n− units. Hence, it was concluded that the synthesized beads were stable up to 427 °C, favouring non-isothermal adsorption.

Figure 2

(a) TGA thermogram and (b) XRD spectra of the recycled PVDF beads.

Figure 2

(a) TGA thermogram and (b) XRD spectra of the recycled PVDF beads.

Close modal

Figure 2(b) shows the XRD spectra of the recycled PVDF beads. The sharp peaks in XRD spectra at 2θ = 17.75° and 22.80° corresponded to (100) and (110) α-crystal diffractions recycled PVDF, respectively (Janakiraman et al. 2016). The sharp peaks at 2θ = 26.07° (022) were attributed to the γ- phase of the recycled PVDF phase. The peak at 2θ = 20.03° was due to the combined β (110) and γ (101) phase (Martins et al. 2014; Esterly & Love 2004).

Figure 3 shows the FT-IR spectra of the recycled PVDF beads. The characteristic absorption bands at 1,400 and 1,180 cm−1 were due to C–H bending and C–F stretching, respectively. The absorbance band at 877 cm−1 was attributed to C–H wagging, while the absorbance band at 840 cm−1 was due to C–F bending (Daems et al. 2018; Yadav et al. 2021a, 2021b, 2021c). The absorbance bands at 839 and 870 cm−1 were attributed to the amorphous phase and at 1276 cm−1 to β-phase vibration (Zahirifar et al. 2019). The FT-IR analysis of the recycled PVDF beads confirmed no impurity in the recycled PVDF beads.

Figure 3

FT-IR spectra of the recycled PVDF beads.

Figure 3

FT-IR spectra of the recycled PVDF beads.

Close modal

Adsorption experiments

The batch adsorption experiments were performed by varying one parameter at a time while the other parameters were kept constant. The effect of process parameters has been discussed in the subsequent sub-sections.

Effect of adsorbent dose

To remove MB dye with recycled PVDF beads, the influence of the adsorbent dose (0.2–3.0 g L−1) was examined (Figure 4). The removal efficiency of MB dye increased from 25% to 99.10%, with an increase in adsorbent dose from 0.2 to 3 g L−1. The increment in the removal efficiency was attributed to the increased availability of adsorption sites (Patel & Yadav 2022). After 2.5 g L−1 adsorbent dose, the removal efficiency reached the maximum value and remained constant on further increasing the dose. Although more active sites were available, no dye molecule was available in the solution to bind with the active adsorbent site due. The adsorption capacity of recycled PVDF beads decreased rapidly in the initial stages and steadily afterwards. As the number of recycled PVDF beads increased, the adsorption capacity decreased due to increased attraction forces, which led to the agglomeration of the beads (Mahmoudzadeh et al. 2013). In addition, the adsorption capacity decrement was due to increased excess active sites. The maximum adsorption capacity of active sites on recycled PVDF beads was not achieved because the concentration of the MB dye was fixed. Moreover, the adsorption capacity is inversely proportional to the adsorbent dose (Equation (1)). Similar observations are reported in other studies (Li et al. 2020; Shojaei & Esmaeili 2022).

Figure 4

Removal efficiency and adsorption capacity of recycled PVDF beads with varying adsorbent dosage (initial MB dye concentration: 20 ppm, contact time: 180 min, and pH: 7).

Figure 4

Removal efficiency and adsorption capacity of recycled PVDF beads with varying adsorbent dosage (initial MB dye concentration: 20 ppm, contact time: 180 min, and pH: 7).

Close modal

Effect of pH

The effect of MB dye solution pH on the adsorption capacity and removal efficiency of the recycled PVDF beads was studied by varying the pH of the solution from 2 to 12. With the increased pH of MB dye solution, the removal efficiency increased from 55% to 99.20%, with the maximum removal efficiency at pH 10 (Figure 5). In a similar trend with an increase in MB dye solution pH, the adsorption capacity increased from 4.4 to 7.94 mg g−1. This increase in removal efficiency and adsorption capacity can be explained by the negative charge present on the recycled PVDF beads (Chiao et al. 2020) and cationic MB dye. The removal efficiency increment was three stages. In the first stage (pH from 2 to 4), the removal efficiency increased slowly because the surface charge of the recycled PVDF was positive. In the second stage (pH from 4 to 8), the removal efficiency increased rapidly since the surface charge of the recycled PVDF became more negative. In the third stage (pH from 8 to 12), the removal efficiency increment was almost constant, attributed to the constant surface charge of the recycled PVDF.

Figure 5

Removal efficiency and adsorption capacity of recycled PVDF beads with varying pH (adsorbent dose: 2.5 g L−1, initial MB dye concentration: 20 ppm, and contact time: 180 min).

Figure 5

Removal efficiency and adsorption capacity of recycled PVDF beads with varying pH (adsorbent dose: 2.5 g L−1, initial MB dye concentration: 20 ppm, and contact time: 180 min).

Close modal

Effect of contact time

The effect of recycled PVDF beads' contact time on the adsorption capacity and removal efficiency for MB dye adsorption was studied with varying contact times from 5 to 240 mins. The removal efficiency and adsorption capacity increased rapidly with the contact time of the adsorbent with dye solution and after some point of time, the curve becomes a plateau (Figure 6). The increments in removal efficiency and adsorption capacity were initially because of the availability of more active adsorption sites. As the adsorption process progressed, the active number of sites decreased as the adsorbent dose was fixed. Hence the curve becomes plateau with a maximum removal efficiency of 98.4%. Similar observations are reported in other studies (Geng et al. 2018; Yadav et al. 2022a, 2022b).

Figure 6

Removal efficiency and adsorption capacity of recycled PVDF beads with varying contact time (adsorbent dose: 2.5 g L−1, initial MB dye concentration: 20 ppm, and pH: 7).

Figure 6

Removal efficiency and adsorption capacity of recycled PVDF beads with varying contact time (adsorbent dose: 2.5 g L−1, initial MB dye concentration: 20 ppm, and pH: 7).

Close modal

Effect of the initial MB concentration

The effect initial concentration of MB dye on removal efficiency and adsorption capacity was investigated by varying the initial concentration from 10 to 250 ppm. The removal efficiency decreased as the MB concentration increased (Figure 7). With increased MB dye concentration from 10 to 250 ppm, the removal efficiency reduced from 99.1% to 27.72%. This decrement in the removal efficiency was due to the increased interaction of MB dye with available active sites on recycled PVDF beads. The adsorption capacity increased from 3.96 to 27.72 mg g−1 with the MB dye initial concentration increasing from 10 to 250 ppm. The increased absorption capacity was attributed to the increased MB dye concentration. The adsorption capacity rose to 27.72 mg g−1 and remained constant because the adsorbent sites were fixed and the MB dye concentration increased (Liu et al. 2019; Jahan et al. 2021).

Figure 7

Removal efficiency and adsorption capacity of recycled PVDF beads with varying initial concentrations of MB dye (adsorbent dose: 2.5 g L−1, contact time: 180 min, and pH: 7).

Figure 7

Removal efficiency and adsorption capacity of recycled PVDF beads with varying initial concentrations of MB dye (adsorbent dose: 2.5 g L−1, contact time: 180 min, and pH: 7).

Close modal

The maximum adsorption capacity of different adsorbents (mainly waste or recycled material) reported in the literature for the adsorption of the MB dye are compared with the maximum adsorption capacity of recycled PVDF beads in Table 1. The adsorption capacity of recycled PVDF beads was comparable with the reported adsorbents. This suggested the feasibility of the cost-effective recycled PVDF beads for the MB dye adsorption.

Table 1

Performance comparison of synthesized beads with previous works for adsorption of MB dye

AdsorbentDose (g L−1)Contact time (min)pHAdsorption capacity (mg g−1)Reference
Biochar 2.5 40 7.5 19 Yang et al. (2016)  
Biochar microparticles (waste derived-pig manure) 0.5 – 16.30 Lonappan et al. (2016)  
Mint waste – 7.1 250 Ainane et al. (2014)  
Bio-waste 0.2 120 135 Reddy et al. (2016)  
Waste rice husk 0.2 45 18.7 Reddy et al. (2013)  
Chemically treated cellulosic waste banana 0.8 1,440 5.6 250 Jawad et al. (2018)  
Metroxylon (waste) 45 36.82 Amode et al. (2016)  
Citrus limetta peel waste 180  227.3 Shakoor & Nasar (2016)  
Potato (Solanum tuberosum) plant waste 25 52.6 Gupta et al. (2016)  
Daucus carota (carrot leaves powder) waste 30 66.6 Kushwaha et al. (2014)  
Date stones and palm-trees waste 10 240 6.3 40 Belala et al. (2011)  
PVDF membrane 0.5 – 6.5 58.82 Bangari et al. (2022a, 2022b)  
Recycled PVDF beads 2.5 180 27.86 This study 
AdsorbentDose (g L−1)Contact time (min)pHAdsorption capacity (mg g−1)Reference
Biochar 2.5 40 7.5 19 Yang et al. (2016)  
Biochar microparticles (waste derived-pig manure) 0.5 – 16.30 Lonappan et al. (2016)  
Mint waste – 7.1 250 Ainane et al. (2014)  
Bio-waste 0.2 120 135 Reddy et al. (2016)  
Waste rice husk 0.2 45 18.7 Reddy et al. (2013)  
Chemically treated cellulosic waste banana 0.8 1,440 5.6 250 Jawad et al. (2018)  
Metroxylon (waste) 45 36.82 Amode et al. (2016)  
Citrus limetta peel waste 180  227.3 Shakoor & Nasar (2016)  
Potato (Solanum tuberosum) plant waste 25 52.6 Gupta et al. (2016)  
Daucus carota (carrot leaves powder) waste 30 66.6 Kushwaha et al. (2014)  
Date stones and palm-trees waste 10 240 6.3 40 Belala et al. (2011)  
PVDF membrane 0.5 – 6.5 58.82 Bangari et al. (2022a, 2022b)  
Recycled PVDF beads 2.5 180 27.86 This study 

Adsorption kinetics and isotherm models

The optimization of the contact time of the adsorbent with the adsorbate molecules plays a significant role in adsorption studies to ensure complete equilibrium between the MB dye and recycled PVDF beads. The PFO and PSO models’ adsorption kinetics were investigated to propose a plausible kinetic mechanism. Table 2 depicts different parameters of the PFO and PSO adoption kinetics for adsorption of MB dye on recycled PVDF beads. The correlation coefficient (R2) for PFO and PSO models was 0.955 and 0.991, respectively. The R2 value for PFO was far more unity than the PSO model, indicating the PSO non-linear second-order model fitting for the adsorption of MB dye on recycled PVDF beads. The PSO model fitting confirmed that the adsorption was chemical (Maslova et al. 2021). The value of the residual sum of the square and reduced chi-square was smaller from the PSO model than PFO. This was due to the lesser difference between the experimental and theoretical values for PSO and PFO models (Figure 8).

Table 2

PFO and PSO kinetics’ parameter for MB dye adsorption on recycled PVDF beads

PFOPSO
qe (mg g−17.512 8.57 
k (min−1) 0.039 0.006 
R2 0.955 0.991 
Reduced chi-square 0.235 0.046 
Residual sum of square 1.88 0.372 
PFOPSO
qe (mg g−17.512 8.57 
k (min−1) 0.039 0.006 
R2 0.955 0.991 
Reduced chi-square 0.235 0.046 
Residual sum of square 1.88 0.372 
Figure 8

Non-linear form of adsorption kinetics (a) PFO and (b) PSO.

Figure 8

Non-linear form of adsorption kinetics (a) PFO and (b) PSO.

Close modal

Two adsorption isotherms (Freundlich and Langmuir) were investigated for the adsorption of MB dye on recycled PVDF beads in the present study (Figure 9). The R2, reduced chi-square, and residual sum of the square were calculated to predict the adsorption isotherm for MB adsorption on recycled PNDF beads (Table 3). The R2 value was close to unity in the case for the Langmuir model (Freundlich: 0.843 and Langmuir: 0.991). The R2 value was close to unity for the Langmuir isotherm model, indicating that the experimental data and predicted results obtained for the MB dye adsorption were closer. Moreover, the value of KL was 0.767 (less than 1). If the value of KL is between 0 and 1, the system can be considered suitable for adsorption purposes. In addition, the smaller value of the residual sum of square (6.965) makes this model applicable for the present work. The horizontal asymptote in the Langmuir isotherm indicated saturation after monolayer adsorption. The fitting Langmuir isotherm indicated the adsorption mechanism was chemical in nature of MB dye on recycled PVDF beads (Bangari et al. 2022a, 2022b).

Table 3

Freundlich and Langmuir isotherm's parameter for MB dye adsorption on recycled PVDF beads

Freundlich isotherm KF ((mg g−1) (L mg−1)−1/n1/n R2 Reduced chi-square Residual sum of squares 
13.636 0.161 0.843 14.247 113.978 
Langmuir isotherm qm (mg g−1KL (L mg−1R2 Reduced chi-square Residual sum of squares 
27.863 0.767 0.991 0.871 6.965 
Freundlich isotherm KF ((mg g−1) (L mg−1)−1/n1/n R2 Reduced chi-square Residual sum of squares 
13.636 0.161 0.843 14.247 113.978 
Langmuir isotherm qm (mg g−1KL (L mg−1R2 Reduced chi-square Residual sum of squares 
27.863 0.767 0.991 0.871 6.965 
Figure 9

Adsorption isotherms (a) Freundlich and (b) Langmuir.

Figure 9

Adsorption isotherms (a) Freundlich and (b) Langmuir.

Close modal

Computational analysis

The DFT optimized geometries of the PVDF, MB dye, and MB-PVDF clusters are shown in Figure 10. The relaxed non-planar geometries of PVDF and MB dye agreed with the literature (Bangari et al. 2022a, 2022b; Yadav et al. 2022a, 2022b). Several starting positions and orientations for the dye molecule were tested to find the most effective contact site. The N atom of MB dye was closest to the PVDF at a distance of 3.21 Å. The adsorption energy for the MB-PVDF complex was –64.7 kJ mol−1 which described that the MB-PVDF system had positive interaction, supporting the experimental findings.

Figure 10

DFT optimized geometries of (a) PVDF, (b) MB, and (c) MB-PVDF.

Figure 10

DFT optimized geometries of (a) PVDF, (b) MB, and (c) MB-PVDF.

Close modal

To determine the stability of the post adsorbed MB-PVDF complex, the vibrational frequencies of the PVDF, MB dye before and after adsorption complexes were computed (Figure 11(a)). All the PVDF and MB dye characteristic bands were present in their respective FT-IR spectra. After the interaction between PVDF and MB dye, the PVDF and MB dye bands were visible in the FT-IR of the MB-PVDF complex. As a result, the vibrational spectra indicated that the MB and PVDF interacted during adsorption. The number of electronic states per unit of energy in a material, as a function of energy, is provided by TDOS, which is critical for understanding its properties (Dindorkar et al. 2022b). The TDOS plots for PVDF, MB, and MB-PVDF clusters are shown in Figure 11(b). For the PVDF, the occupied and unoccupied molecular orbitals were widely spaced. After adsorption of MB dye on PVDF, additional energy levels appeared, which were traced to the charge transfer from the HOMO of the PVDF to the MB dye molecule. The adsorption of MB dye on PVDF resulted in a decrease in band gaps, which might be due to the emergence of new energy levels that mix the orbitals of MB dye molecules with the molecular orbitals of PVDF (Bangari et al. 2021). The TDOS plots confirmed the findings from the above sections and gave an account of the variation in the electronic density of states after the adsorption of MB dye.

Figure 11

(a) Theoretical IR spectra (b) TDOS of PVDF, MB, and PVDF-MB cluster.

Figure 11

(a) Theoretical IR spectra (b) TDOS of PVDF, MB, and PVDF-MB cluster.

Close modal

The molecular orbitals (HOMO and LUMO) representation for PVDF, MB and PVDF-MB clusters are shown in Figure 12. The HOMOs and LUMOs were distributed uniformly on PVDF before adsorption. From the HOMO–LUMO, a reasonable charge distribution can be observed on the adsorbent sites on the double bond of the ethylene group. The HOMOs were concentrated on the electronegative N atoms of the MB dye molecule, while LUMOs were delocalized. After adsorption, the HOMOs and LUMOs were concentrated on MB dye molecules. This means that the MB-PVDF attained more nature of MB dye rather than PVDF. The HOMO and LUMO energies are related to chemical parameters that provide the details of the reactivity of the molecule (Dindorkar & Yadav 2022). Higher chemical stability can be attained with higher chemical hardness as they reduce the polarizability of the molecules. Therefore, as MB-PVDF's HOMO-LUMO energy gap (HLG) reduced after adsorption, chemical hardness also reduced, but there was little chemical potential change (Table 4). The HLG value of PVDF before and after adsorption was 9.42 eV and 0.50 eV, respectively. This was because of charge transfer from MB dye to PVDF during adsorption, which led to a change in the density of occupied orbitals. As discussed above that the MB-PVDF complex attained the nature of MB dye, the complex became strong electrophile after absorption.

Table 4

The quantum mechanical descriptors before and after adsorption

PropertiesPVDFMBMB-PVDF
HOMO −9.53 −5.01 −4.95 
LUMO −0.11 −4.54 −4.45 
HLG 9.42 0.47 0.50 
VIP 9.53 5.01 4.95 
VEA 0.11 4.54 4.45 
Chemical hardness 4.71 0.24 0.25 
Chemical potential −4.82 −4.78 −4.70 
Electrophilicity index 2.47 48.51 44.18 
PropertiesPVDFMBMB-PVDF
HOMO −9.53 −5.01 −4.95 
LUMO −0.11 −4.54 −4.45 
HLG 9.42 0.47 0.50 
VIP 9.53 5.01 4.95 
VEA 0.11 4.54 4.45 
Chemical hardness 4.71 0.24 0.25 
Chemical potential −4.82 −4.78 −4.70 
Electrophilicity index 2.47 48.51 44.18 
Figure 12

HOMO-LUMO distributions on (a) PVDF, (b) MB, and (c) MB-PVDF.

Figure 12

HOMO-LUMO distributions on (a) PVDF, (b) MB, and (c) MB-PVDF.

Close modal

Plausible adsorption mechanism

Studying the adsorption mechanism of MB dye on recycled PVDF beads is important to gain insights into the actual process. Usually, if the PSO kinetic model describes an adsorption process better, it was inferred to be a chemisorption process. However, the DFT studies showed that the adsorption energy was −64 kJ mol−1, implying the adsorption process was physical. Moreover, the shortest distance between the PVDF and MB dye molecules was >2 Å, implying that the adsorption was physical (Yadav & Dindorkar 2022a). As a result, the adsorption mechanism in this investigation could not be categorized solely as a chemical or physical response but rather as a mixed one. Other researchers have reported this observation (Zheng et al. 2018). MB dye is a cationic dye that gets adsorbed on the surface of the recycled PVDF beads by electrostatic attraction. The DFT simulations showed that the interaction of the MB dye molecule was from the electronegative N atom of the MB dye molecule. This means that electrostatic interaction occurred between the recycled PVDF beads and the positively charged quaternary ammonium groups in MB dye. The hydrogen bonding was due to the attraction between F atoms of PVDF and the amine group present in the MB dye molecule (Zheng et al. 2018). The adsorption was enhanced by the MB dye's linearity and three parallel aromatic rings, which encourage π-π interactions (Muslim et al. 2021).

In this study, we report the feasibility of the recycled PVDF for the adsorption of MB dye which mitigates two problems simultaneously: (i) disposal of used membranes and (ii) textile wastewater treatment. Batch adsorption experiments indicated that the MB dye adsorption onto the recycled PVDF beads depended on adsorbent dosage, initial concentration, contact time, and pH. The adsorption kinetics study of the MB dye adsorption followed the PSO model, indicating chemisorption. Furthermore, the adsorption isotherm study showed that the adsorption process followed the Langmuir isotherm showing the monolayer adsorption of MB dye on recycled PVDF beads. The PSO model's equilibrium adsorption capacity was 8.56 mg g−1, and the experimental equilibrium adsorption capacity was 8.30 mg g−1. The maximum adsorption capacity was 27.86 mg g−1. The DFT study predicted the adsorption energy of –64.7 kJ mol−1 for MB-PVDF, indicating the strong interaction between MB dye and PVDF. Moreover, the gap between the HOMO and LUMO of PVDF upon interaction with MB dye decreased from 9.42 eV to 0.50 eV due to the mixing of molecular orbitals. The present study reveals the recycled PVDF beads’ potential for textile wastewater dye removal at a low cost.

The CSIR-CSMCRI PRIS number for this manuscript is 109/2022. The authors are grateful for partial funding support from the Council of Scientific and Industrial Research, India (MLP-0043). The authors acknowledge AED&CIF division, CSIR-CSMCRI for providing instrumental facilities. The authors also thank Dr B. Ganguly, CSIR-CSMCRI, for his help in theoretical calculations. The comments from anonymous reviewers and the editor have greatly improved the content.

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

The authors declare there is no conflict.

Amode
J. O.
,
Santos
J. H.
,
Md. Alam
Z.
,
Mirza
A. H.
&
Mei
C. C.
2016
Adsorption of methylene blue from aqueous solution using untreated and treated (Metroxylon spp.) waste adsorbent: equilibrium and kinetics studies
.
International Journal of Industrial Chemistry
7
(
3
),
333
345
.
Bangari
R. S.
,
Yadav
A.
,
Awasthi
P.
&
Sinha
N.
2022a
Experimental and theoretical analysis of simultaneous removal of methylene blue and tetracycline using boron nitride nanosheets as adsorbent
.
Colloids and Surfaces A: Physicochemical and Engineering Aspects
634
,
127943
.
Bangari
R. S.
,
Yadav
A.
,
Bharadwaj
J.
&
Sinha
N.
2022b
Boron nitride nanosheets incorporated polyvinylidene fluoride mixed matrix membranes for removal of methylene blue from aqueous stream
.
Journal of Environmental Chemical Engineering
10
(
1
),
107052
.
Belala
Z.
,
Jeguirim
M.
,
Belhachemi
M.
,
Addoun
F.
&
Trouvé
G.
2011
Biosorption of basic dye from aqueous solutions by date stones and palm-trees waste: kinetic, equilibrium and thermodynamic studies
.
Desalination
271
(
1–3
),
80
87
.
Caldeweyher
E.
,
Bannwarth
C.
&
Grimme
S.
2017
Extension of the D3 dispersion coefficient model
.
The Journal of Chemical Physics
147
(
3
),
034112
.
Chiao
Y.-H.
,
Chen
S.-T.
,
Sivakumar
M.
,
Ang
M. B. M. Y.
,
Patra
T.
,
Almodovar
J.
,
Wickramasinghe
S. R.
,
Hung
W.-S.
&
Lai
J.-Y.
2020
Zwitterionic polymer brush grafted on polyvinylidene difluoride membrane promoting enhanced ultrafiltration performance with augmented antifouling property
.
Polymers
12
(
6
),
1303
.
Coutinho de Paula
E.
,
Gomes
J. C. L.
&
Amaral
M. C. S.
2017
Recycling of end-of-life reverse osmosis membranes by oxidative treatment: a technical evaluation
.
Water Science and Technology
76
(
3
),
605
622
.
Daems
N.
,
Milis
S.
,
Verbeke
R.
,
Szymczyk
A.
,
Pescarmona
P. P.
&
Vankelecom
I. F. J.
2018
High-performance membranes with full pH-stability
.
RSC Advances
8
(
16
),
8813
8827
.
Dindorkar
S. S.
,
Patel
R. V.
&
Yadav
A.
2022a
Adsorptive removal of methylene blue dye from aqueous streams using photocatalytic CuBTC/ZnO chitosan composites
.
Water Science and Technology
85, 2748–2760.
Dindorkar
S. S.
,
Patel
R. V.
&
Yadav
A.
2022b
Quantum chemical study of the defect laden monolayer boron nitride nanosheets for adsorption of pesticides from wastewater
.
Colloids and Surfaces A: Physicochemical and Engineering Aspects
643
,
128795
.
Esterly
D. M.
&
Love
B. J.
2004
Phase transformation to?-poly(vinylidene fluoride) by milling
.
Journal of Polymer Science Part B: Polymer Physics
42
(
1
),
91
97
.
Gan
D.
,
Dou
J.
,
Huang
Q.
,
Huang
H.
,
Chen
J.
,
Liu
M.
,
Qi
H.
,
Yang
Z.
,
Zhang
X.
&
Wei
Y.
2020
Carbon nanotubes-based polymer nanocomposites: bio-mimic preparation and methylene blue adsorption
.
Journal of Environmental Chemical Engineering
8
(
2
),
103525
.
Ivanets
A. I.
,
Prozorovich
V. G.
,
Roshchina
M. Y.
,
Srivastava
V.
&
Sillanpää
M.
2019
Unusual behavior of MgFe2O4 during regeneration: desorption versus specific adsorption
.
Water Science and Technology
80
(
4
),
654
658
.
Jahan
K.
,
Tyeb
S.
,
Kumar
N.
&
Verma
V.
2021
Bacterial cellulose-polyaniline porous mat for removal of methyl orange and bacterial pathogens from potable water
.
Journal of Polymers and the Environment
29
(
4
),
1257
1270
.
Janakiraman
S.
,
Surendran
A.
,
Ghosh
S.
,
Anandhan
S.
&
Venimadhav
A.
2016
Electroactive poly(vinylidene fluoride) fluoride separator for sodium ion battery with high coulombic efficiency
.
Solid State Ionics
292
,
130
135
.
Jawad
A. H.
,
Rashid
R. A.
,
Ishak
M. A. M.
&
Ismail
K.
2018
Adsorptive removal of methylene blue by chemically treated cellulosic waste banana (Musa sapientum) peels
.
Journal of Taibah University for Science
12
(
6
),
809
819
.
Kiran
S.
,
Ahmad
T.
,
Gulzar
T.
,
Ashraf
A.
,
Naqvi
S. A. R.
&
Naz
S.
2020
Agro-waste applications for bioremediation of textile effluents
. In:
Recycling From Waste in Fashion and Textiles
(
Pandit, P., Ahmed, S., Singha, K. & Shrivastava, S. eds.
).
Wiley
,
Beverly, MA 01915, United States
. pp.
391
421
.
Kushwaha
A. K.
,
Gupta
N.
&
Chattopadhyaya
M. C.
2014
Removal of cationic methylene blue and malachite green dyes from aqueous solution by waste materials of Daucus carota
.
Journal of Saudi Chemical Society
18
(
3
),
200
207
.
Lawler
W.
,
Bradford-Hartke
Z.
,
Cran
M. J.
,
Duke
M.
,
Leslie
G.
,
Ladewig
B. P.
&
Le-Clech
P.
2012
Towards new opportunities for reuse, recycling and disposal of used reverse osmosis membranes
.
Desalination
299
,
103
112
.
Lonappan
L.
,
Rouissi
T.
,
Das
R. K.
,
Brar
S. K.
,
Ramirez
A. A.
,
Verma
M.
,
Surampalli
R. Y.
&
Valero
J. R.
2016
Adsorption of methylene blue on biochar microparticles derived from different waste materials
.
Waste Management
49
,
537
544
.
Manoj Kumar Reddy
P.
,
Mahammadunnisa
S.
,
Ramaraju
B.
,
Sreedhar
B.
&
Subrahmanyam
C.
2013
Low-cost adsorbents from bio-waste for the removal of dyes from aqueous solution
.
Environmental Science and Pollution Research
20
(
6
),
4111
4124
.
Martins
P.
,
Lopes
A. C.
&
Lanceros-Mendez
S.
2014
Electroactive phases of poly(vinylidene fluoride): determination, processing and applications
.
Progress in Polymer Science
39
(
4
),
683
706
.
Maslova
M.
,
Mudruk
N.
,
Ivanets
A.
,
Shashkova
I.
&
Kitikova
N.
2021
The effect of pH on removal of toxic metal ions from aqueous solutions using composite sorbent based on Ti-Ca-Mg phosphates
.
Journal of Water Process Engineering
40
,
101830
.
Mok
C. F.
,
Ching
Y. C.
,
Muhamad
F.
,
Abu Osman
N. A.
,
Hai
N. D.
&
Che Hassan
C. R.
2020
Adsorption of dyes using poly(vinyl alcohol) (PVA) and PVA-based polymer composite adsorbents: a review
.
Journal of Polymers and the Environment
28
(
3
),
775
793
.
Ngulube
T.
,
Gumbo
J. R.
,
Masindi
V.
&
Maity
A.
2017
An update on synthetic dyes adsorption onto clay based minerals: a state-of-art review
.
Journal of Environmental Management
191
,
35
57
.
Obotey Ezugbe
E.
,
Rathilal
S.
,
Ezugbe
E. O.
,
Rathilal
S.
,
Obotey Ezugbe
E.
&
Rathilal
S.
2020
Membrane technologies in wastewater treatment: a review
.
Membranes
10
(
5
),
89
.
O'Boyle
N. M.
,
Tenderholt
A. L.
&
Langner
K. M.
2008
Cclib: a library for package-independent computational chemistry algorithms
.
Journal of Computational Chemistry
29
(
5
),
839
845
.
Piaskowski
K.
,
Świderska-Dąbrowska
R.
&
Zarzycki
P. K.
2018
Dye removal from water and wastewater using various physical, chemical, and biological processes
.
Journal of AOAC International
101
(
5
),
1371
1384
.
Reddy
P. M. K.
,
Verma
P.
&
Subrahmanyam
C.
2016
Bio-waste derived adsorbent material for methylene blue adsorption
.
Journal of the Taiwan Institute of Chemical Engineers
58
,
500
508
.
Tran
H. N.
,
Nguyen
H. C.
,
Woo
S. H.
,
Nguyen
T. V.
,
Vigneswaran
S.
,
Hosseini-Bandegharaei
A.
,
Rinklebe
J.
,
Kumar Sarmah
A.
,
Ivanets
A.
,
Dotto
G. L.
,
Bui
T. T.
,
Juang
R.-S.
&
Chao
H.-P.
2019
Removal of various contaminants from water by renewable lignocellulose-derived biosorbents: a comprehensive and critical review
.
Critical Reviews in Environmental Science and Technology
49
(
23
),
2155
2219
.
Wallingford
C. T.
2013
Gaussian, Inc.
.
Wong
S.
,
Ghafar
N. A.
,
Ngadi
N.
,
Razmi
F. A.
,
Inuwa
I. M.
,
Mat
R.
&
Amin
N. A. S.
2020
Effective removal of anionic textile dyes using adsorbent synthesized from coffee waste
.
Scientific Reports
10
(
1
),
2928
.
Yadav
A.
&
Dindorkar
S. S.
2022a
Adsorption behaviour of hexagonal boron nitride nanosheets towards cationic, anionic and neutral dyes: insights from first principle studies
.
Colloids and Surfaces A: Physicochemical and Engineering Aspects
640
,
128509
.
Yadav
A.
&
Sinha
N.
2021
Organic polymers for drinking water purification
. In:
Reference Module in Materials Science and Materials Engineering
.
Elsevier
.
Yadav
A.
,
Patel
R. V. R. V.
,
Labhasetwar
P. K. P. K.
&
Shahi
V. K. V. K.
2021b
Novel MIL101(Fe) impregnated poly(vinylidene fluoride-co-hexafluoropropylene) mixed matrix membranes for dye removal from textile industry wastewater
.
Journal of Water Process Engineering
43
,
102317
.
Yadav
P.
,
Yadav
A.
&
Labhasetwar
P. K.
2022b
Sustainable adsorptive removal of antibiotics from aqueous streams using Fe3O4-functionalized MIL101(Fe) chitosan composite beads
.
Environmental Science and Pollution Research
29, 37204 –37217.
Yanai
T.
,
Tew
D. P.
&
Handy
N. C.
2004
A new hybrid exchange–correlation functional using the Coulomb-attenuating method (CAM-B3LYP)
.
Chemical Physics Letters
393
(
1–3
),
51
57
.
Yang
G.
,
Wu
L.
,
Xian
Q.
,
Shen
F.
,
Wu
J.
&
Zhang
Y.
2016
Removal of Congo red and methylene blue from aqueous solutions by vermicompost-derived biochars
.
PLoS ONE
11
(
5
), 1–18.
Yaseen
D. A.
&
Scholz
M.
2019
Textile dye wastewater characteristics and constituents of synthetic effluents: a critical review
.
International Journal of Environmental Science and Technology
16
(
2
),
1193
1226
.
Yi
X.
,
Wang
Y.
,
Jin
L.
&
Shi
W.
2017
Critical flux investigation in treating o/w emulsion by TiO2/Al2O3-PVDF UF membrane
.
Water Science and Technology
76
(
10
),
2785
2792
.
Zhang
D.
,
Dai
F.
,
Zhang
P.
,
An
Z.
,
Zhao
Y.
&
Chen
L.
2019
The photodegradation of methylene blue in water with PVDF/GO/ZnO composite membrane
.
Materials Science and Engineering: C
96
,
684
692
.
Zheng
X.
,
Zheng
H.
,
Zhao
R.
,
Sun
Y.
,
Sun
Q.
,
Zhang
S.
&
Liu
Y.
2018
Polymer-functionalized magnetic nanoparticles: synthesis, characterization, and methylene blue adsorption
.
Materials
11
(
8
),
1312
.
Zwain
H. M.
,
Vakili
M.
&
Dahlan
I.
2014
Waste material adsorbents for zinc removal from wastewater: a comprehensive review
.
International Journal of Chemical Engineering
2014
,
1
13
.
This is an Open Access article distributed under the terms of the Creative Commons Attribution Licence (CC BY 4.0), which permits copying, adaptation and redistribution, provided the original work is properly cited (http://creativecommons.org/licenses/by/4.0/).