Previous Article in Journal
Effect of Vibration on Open-Cathode Direct Methanol Fuel Cell Stack Performance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation of Poly(vinylidene fluoride-co-hexafluoropropylene) Doped Cellulose Acetate Films for the Treatment of Calcium-Based Hardness from Aqueous Solution

by
Khaleke Veronicah Ramollo
1,
Lutendo Evelyn Macevele
1,*,
Abayneh Ataro Ambushe
2 and
Takalani Magadzu
1
1
Department of Chemistry, University of Limpopo, Private Bag X 1106, Sovenga, Polokwane 0727, South Africa
2
Department of Chemical Sciences, University of Johannesburg, P.O. Box 524, Auckland Park, Johannesburg 2006, South Africa
*
Author to whom correspondence should be addressed.
Physchem 2025, 5(4), 45; https://doi.org/10.3390/physchem5040045
Submission received: 15 August 2025 / Revised: 18 September 2025 / Accepted: 24 September 2025 / Published: 20 October 2025
(This article belongs to the Section Surface Science)

Abstract

Calcium (Ca2+ ions) is one of the dominant elements that contribute to water hardness, scaling in pipes, bathroom faucets, and kitchen utensils. Herein, we report on the development of poly(vinylidene fluoride-co-hexafluoropropylene) cellulose acetate (PVDF-HFP/CA) films for the treatment of Ca2+ ions as one of the constituents that causes water hardness. CA and PVDF-HFP polymers, and their blend consisting of 3 wt.% PVDF-HFP/CA, were effectively synthesised through the phase inversion technique. Analysis using thermogravimetric analysis (TGA), Fourier transform infrared (FTIR) spectroscopy, and scanning electron microscopy (SEM) confirmed the effective incorporation of 3 wt.% PVDF-HFP into the cellulose acetate film. Parameters such as temperature, initial concentration, pH, adsorbent dosage and contact time were investigated in batch studies during the removal of Ca2+ ions in synthetic water samples. Under optimal conditions (pH 7, adsorbent dosage of 0.5 mg/L, and concentration of 120 mg/L), the 3 wt.% PVDF-HFP/CA film achieved a 99% adsorption efficiency for Ca2+ ions in 90 min. The adsorption process adhered to pseudo-second-order and Freundlich isotherm models, which suggest that the adsorption of Ca2+ ions is heterogeneous. The maximum adsorption efficiency achieved was 56 mg/g, indicating an endothermic physisorption process. The 3 wt.% PVDF-HFP/CA film maintained higher adsorption in the presence of counter ions and in a binary system, and it could be recycled at least three times. Thus, the findings demonstrated that the 3 wt.% PVDF-HFP/CA film could be a valuable material for Ca2+ ions removal to acceptable drinking water levels.

Graphical Abstract

1. Introduction

In rural and arid regions, the majority of water supplies come from both groundwater and river sources [1]. However, these water sources have elevated levels of dissolved salts, including calcium (Ca2+) and magnesium (Mg2+) ions, which contribute to water hardness, particularly when concentrations exceed 120 mg/L [2,3]. Consuming water with higher levels of calcium and magnesium, particularly magnesium, lowers the risk of stroke in postmenopausal women [4]. Although magnesium and calcium present in hard water have some health benefits, calcium causes major challenges in industrial processes, which might cause severe economic impact [5,6]. For instance, the hardness of water contributes to scale buildup in both industrial and household electrical devices, which functions as an insulator and extends the heating duration [7]. In industrial sectors, water hardness is frequently monitored to prevent expensive failures of cooling towers, scaling in boilers, and various other equipment [8]. The use of domestic water containing high levels of calcium and magnesium ions can react with soap anions which reduces cleaning efficiency and increase detergent consumption [3]. Research also indicates that the effectiveness of certain herbicides is diminished by water hardness, especially in areas where the bedrock is predominantly made up of lime and dolomite [9,10]. Although both calcium and magnesium contribute to water hardness, magnesium is mainly found in natural groundwater at low concentrations (below 100 mg/L); hence, calcium-based hardness often predominates [11].
A study conducted by Elumalai et al., 2020 showed that groundwater collected from Luvuvhu catchment area had moderate to hard water levels of calcium [12]. Most water-softening treatments available for water hardness are very costly, directly affecting the rural and low-income community [1]. Despite the preference for ion-exchange and adsorption methods due to their cost-effectiveness, the limited adsorption capacity and specific properties of ion-exchange materials present significant challenges [13].
Although substantial scientific research has been conducted on the elimination of contaminants from water [13,14,15,16,17], there have been limited investigations focused on reducing water hardness, especially through the use of composite films. For instance, a study carried out by Sepehr et al., 2013 using alkaline modified pumices stone achieved 96% Ca2+ and 73% Mg2+ ions adsorption within 300 min from synthetic water [18]. Werkneh et al., 2015 developed an alkali-treated sugarcane bagasse and coffee husk for the adsorption of Ca2+ and Mg2+ ions from water samples [19], achieving 78% and 73% calcium and magnesium adsorption, respectively, within 200 min. In another study, Hailu et al., 2019 reported maximum adsorption of 80% Ca2+ and 85% Mg2+ ions from water samples using natural zeolite as an adsorbent [1].
In recent years, poly(vinylidene fluoride-hexafluoropropylene) (PVDF-HFP) has been employed as membranes in water disinfection as a result of their excellent dielectric properties, mechanical strength, and elevated ionic conductivity [20,21]. However, this polymer has low thermal stability, swellability, water content and porosity which limit its performance in adsorption processes [22]. The PVDF-HFP polymer can be modified with a variety of materials such as other polymers or carbon nanocomposites, to improve its performance in water purification [23,24,25,26]. Incorporating cellulose acetate (CA) into PVDF-HFP polymer membranes and films has been found to notably enhance porosity, electrolyte absorption, ionic conductivity, and hydrophilicity [27,28], owing to its outstanding wettability, cost-effective processing, high porosity, and favourable mechanical attributes [29].
Therefore, this research seeks to develop a CA-modified PVDF-HFP film using phase inversion techniques to remove Ca2+ ions from synthetic water samples. Batch adsorption experiments were conducted to examine the impacts of pH, adsorbent dosage, concentrations of Ca2+ ions, counter ions, and other variables.

2. Materials and Methods

The chemicals utilised were all of analytical grade. Cellulose acetate (CA), poly(vinylidene fluoride-co-hexafluoropropylene) (PVDF-HFP), dimethylformamide (DMF), nitric acid, calcium sulphate, polyethylene glycol (PEG), and sodium hydroxide were all purchased from Merck Life Science (Pty) Ltd., Modderfontein, South Africa.

2.1. Preparation of PVDF-HFP Film

2.1.1. Preparation of CA Film

Cellulose acetate films were fabricated utilising a phase inversion technique as described in the existing literature [22]. A cellulose acetate (CA) polymer solution was prepared by combining CA (0.900 g) and PEG (0.2 g) in 10 mL of DMF within a round-bottom flask, followed by stirring this mixture vigorously at 80 °C for 2 h. Subsequently, the polymer solution was placed in a silica gel desiccator overnight. The next day, it was manually cast onto a glass plate using a 180 µm thick casting knife (Elcometer 3580, Elcometer Limited, Manchester, UK). The resulting films were washed with deionised water and then air-dried on plain paper at room temperature.

2.1.2. Preparation of PVDF-HFP and 3 wt.% PVDF-HFP/CA Film

The PVDF-HFP films were synthesised utilising a phase inversion technique as documented in the prior literature [22]. In summary, PVDF-HFP (1.00 g) and PEG (0.2 g) were dissolved in DMF (10 mL) inside a round-bottom flask, and the mixture was stirred intensely at 80 °C for 2 h. Afterward, the reaction mixture was left to settle overnight in a silica gel desiccator. The solution was then hand-cast onto a glass plate using an Elcometer 3580 casting knife (Elcometer 3580, Elcometer Limited, Manchester, UK) with a 180 µm gap. Once fabricated, the films were dried in a vacuum oven at 80 °C for 30 s to partially evaporate the solvent, followed by immersion in a coagulation bath of deionized water maintained at 5 °C to trigger phase inversion. The final PVDF-HFP film was rinsed with deionized water and air-dried on plain paper towels at room temperature (see the detailed schematic process in Scheme 1).
The 3 wt.% PVDF-HFP/CA film was crafted following the previously mentioned procedure described above. Initially, 0.97 g of cellulose acetate was combined with 9 mL of DMF and 1 mL of PEG (used as a pore-forming agent) while stirring constantly at 600 rpm and maintaining a temperature of 60 °C, resulting in a cellulose acetate solution. Subsequently, 0.03 g of PVDF-HFP polymer was incorporated into the cellulose acetate solution to form a 3 wt.% PVDF-HFP/CA solution. [27]. Lastly, the solution was cast as described above.
The concentration of 3 wt.% PVDF-HFP film was selected based on preliminary trials, as this composition was found to enhance the removal efficiency of Ca2+ ions from water samples while maintaining good film stability and processability.

2.1.3. Characterisation Techniques

The surface morphology of the specimens was analysed by scanning electron microscopy (SEM) with a JSM 5910 model from JEOL, Tokyo, Japan. The functional groups were identified through Fourier transform infrared (FTIR) spectroscopy using a BRUKER-FTIR-ATR spectrometer. This analysis was conducted in the 650 to 4000 cm−1 range, with a resolution of 4 cm−1, employing four scans at a temperature of 25 °C. Additionally, thermogravimetric analysis (TGA) was performed with a Perkin Elmer STA 4000 analyser to assess the mass changes in the films across various temperatures, thereby evaluating the thermal stability of the specimens.

2.1.4. Adsorption Studies of Ca2+ Ions in Synthetic Water Samples

The study of pH was conducted by adjusting the initial pH of a 35 mg/L copper sulphate solution to values of 5, 5.5, 6, 6.5, and 7. The pH levels were modified using 0.1 M nitric acid (HNO3) and 0.1 M sodium hydroxide (NaOH). During each experiment, 25 mg of films were immersed in 50 mL of these prepared solutions and stirred vigorously for 90 min at room temperature (25 °C) [30]. Flame-atomic absorption spectroscopy (F-AAS) was employed to analyse the concentration of Ca2+ ions post-adsorption. Each experiment was conducted in triplicates, with mean values presented. The same procedure was followed to investigate the influence of variables such as dosage (ranging from 0.1 to 0.9 mg/L), time (at intervals from 30 to 180 min), initial concentration (spanning 100 to 1200 mg/L), and temperature (between 25 and 35 °C).
The equation below was used to obtain the amount of the adsorbed salts:
%   R e m o v a l =   ( C o   C e ) C 0 × 100
where Co and Ce denote the initial and final concentrations of Ca2+ ions, respectively, in mg/L [31,32].
The adsorption kinetics were examined by evaluating the kinetic parameters of the Lagergren pseudo-first-order and pseudo-second-order models. This process involved performing nonlinear regression analysis to align the experimental data with these empirical kinetic models [32]. The Lagergren’s pseudo-first-order model and its linearised representation, as detailed in the equations below, were utilised [25,33].
q t = q t ( 1 e k 1 t )
In ( q e q t ) = In q e k 1 t
Furthermore, the Lagergren pseudo-second-order model expressed below was used:
t q t = 1 k 2 q e 2 + t q e
where qe and qt (mg/g) represent the concentrations of Ca2+ ions adsorbed at equilibrium and at time, t. The constants k1 (min−1) and k2 (g/mg·min) correspond to the pseudo-first-order and pseudo-second-order reaction rate constants, respectively.
The experimental data collected from the results were analysed using both Langmuir and Freundlich isotherms. The theoretical adsorption capacity of the PVDF-HFP film was calculated based on the nonlinear Langmuir and Freundlich isotherm models [16,34].
The nonlinear equation for the Langmuir isotherm may be expressed as [35]:
q e = q m a x b C e 1 + b C e
1 q e = 1 q m a x + 1 b q C e
where qe and qmax represents both the equilibrium and maximum adsorption capacity (mg/g), respectively. Ce (mg/L) is the equilibrium concentration of metal adsorbed in the solution and b (L/mg) is the Langmuir equilibrium constant.
The Langmuir adsorption isotherm model includes a significant dimensionless factor RL, which can be determined using the following equation:
R L = 1 1 + b C o
where Co (mg/L) represents the initial concentration of the metal and b (L/mg) represent the Langmuir equilibrium constant for adsorption.
The Freundlich isotherm, another renowned adsorption model, can be represented as [33,36]:
qe = Kf Ce 1/n
In q e = ln K f + 1 n ln C e
where qe represents the adsorption at equilibrium measured in mg/g and Ce (mg/L) is the solute concentration at equilibrium. The Freundlich constants are Kf (mg/g), indicating adsorption capacity, and n (dimensionless) signifying adsorption intensity.
The equations provided below were utilised to perform a thermodynamic analysis [37,38]:
ΔGo = −RT lnKc
ln K c = Δ S ° R Δ H ° R T
K c = C a d C e
where Ce (mg/L) and Cad (mg/L) represent the equilibrium concentration of metal present in the solution and the concentration of the metal within the adsorbent at equilibrium. R denotes the universal gas constant (8.314 J/mol. K). ΔH°, ΔG°, and ΔS° represent changes in standard enthalpy (kJ/mol), the standard Gibbs free energy (kJ/mol), and standard entropy (J/mol/K), respectively. A plot of lnK against 1/T results in a straight line that has a slope of ΔH°/R and an intercept ΔS°/R; from the slope and intercept, the values of ΔS° and ΔH° were obtained.

2.1.5. Effect of Counterions

To evaluate the performance of PVDF-HFP films under field conditions, water samples were prepared with concentrations of sulphate (325 mg/L), nitrate (25 mg/L), and chloride (450 mg/L) in the presence of 120 mg/L calcium. These mixtures were created using sodium sulphate, potassium nitrate, sodium chloride, and analytical-grade calcium sulphate. Each solution provided 12.5 mL, which was then poured into a 100 mL conical flask, culminating a total volume of 50 mL [39,40]. The same procedure was employed to investigate the impact on a binary system, involving approximately 25 mg of the PVDF-HFP film, which was introduced into a 50 mL solution comprising 25 mL of Ca2+ and Mg2+ ions (concentration: 120 mg/L) at a pH of 7 for a duration of 90 min at ambient temperature. Each experiment was conducted in triplicate, and the average results were documented.

2.1.6. Reusability Studies

The most effective polymeric film underwent recycling by monitoring the adsorption and desorption of Ca2+ ions over three cycles. Used PVDF-HFP films (25 mg) were submerged in a 50 mL solution containing Ca2+ ions at a concentration of 120 mg/L for 90 min at ambient temperature. After the 90 min period, the concentration of Ca2+ ions was assessed using F-AAS. The Ca2+-saturated PVDF-HFP films were air-dried and then immersed in a 0.1 M NaOH solution (50 mL) for 2 h at 25 °C for Ca2+ ion removal (desorption process). Afterwards, the films were removed from the solution, rinsed with deionized water, and dried at 60 °C. The recycling protocol for the spent PVDF-HFP films was conducted four times to address the hardness components [32].

3. Results

3.1. Characterisation of Films

3.1.1. The FTIR Profiles of PVDF-HFP, CA, and 3 wt.% PVDF-HFP/CA Films

Figure 1 presents the FTIR spectra for the as-prepared PVDF-HFP, CA, and the PVDF-HFP/CA composite films containing 3 wt.% PVDF-HFP. The peak observed at 840 cm−1 confirms the amorphous phase formation of PVDF-HFP (Figure 1 (a)). Meanwhile, the IR bands at 1402 and 1279 cm−1 correspond to the deformation vibration of CH2 and the stretching vibration of CF2, which are characteristic of the PVDF-HFP polymer, respectively [22,24]. In Figure 1 (b), the carbonyl (C=O) stretching vibration at 1738 cm−1 confirms the existence of cellulose, and in Figure 1 (c), it indicates that CA was successfully blended into the PVDF-HFP polymer. The cellulose addition to the PVDF-HFP strengthened the vibration at 1045 cm−1, while the presence of cellulose was also confirmed by the O−H stretch with a broad peak at 3406 cm−1, as observed in the literature [27,41].

3.1.2. TGA Profiles of PVDF-HFP, CA, and 3 wt.% PVDF-HFP/CA Films

Figure 2 presents the TGA curves for PVDF-HFP, CA, and films composed of 3 wt.% PVDF-HFP/CA. The thermogram of the PVDF-HFP film alone (Figure 2 (a)) displays one significant weight loss occurring approximately between 420 and 470 °C, which can be ascribed to the breakdown of the polymeric backbone, specifically the PVDF matrix [31,42]. The TGA profile of CA (Figure 2 (b)) shows a decomposition at 200 and 350 °C, corresponding to the deacetylation and the collapse of the CA structure [27,43]. After the addition of PVDF-HFP on CA (Figure 2 (c)), the thermal stability of CA decreased. This indicates that the small percentage (3 wt.%) of PVDF-HFP introduced on CA does not improve the stability of CA; instead, it slightly weakens the structure of CA. However, the 3 wt.% PVDF-HFP/CA film has good thermal stability below 200 °C, which is within the water purification operating temperature range.

3.1.3. SEM Images of PVDF-HFP, CA, and 3 wt.% PVDF-HFP/CA Films

Figure 3 presents SEM images that were employed to examine the surface morphology of the PVDF-HFP, CA, and 3 wt.% PVDF-HFP/CA films. The SEM images confirm the dense nature of the films, indicating that Ca2+ ions removal arises mainly from surface interactions rather than from internal porosity. SEM images of the prepared PVDF-HFP film (Figure 3a) shows some interconnected pores distributed uniformly, and a porous honeycomb structure within the films [27]. The SEM image of CA films (Figure 3b) shows a dense surface layer, with agglomerates on the surface [41,44]. This is due to the intermolecular forces from hydrogen bonds found within the CA molecules. The morphology of 3 wt.% PVDF-HFP/CA film (Figure 3c) shows a spongy structure with visible rough surface. This morphology is noted for its extensive surface area and superior adsorption capability [25]. The results demonstrate that the preparation of all the PVDF-HFP films was successful.

3.2. Adsorption Studies

3.2.1. Impact of pH on Ca2+ Ions Removal by PVDF-HFP Film

Figure 4 illustrates how pH influences the adsorption of Ca2+ ions onto a PVDF-HFP film. As depicted, the removal efficiency of Ca2+ ions rises from 74.5% to 79.5% when the solution’s pH increases from 5 to 7. The removal efficiency increased steadily from acidic conditions toward neutral pH, with values ranging from 76.3% at pH 6.0 to 75.7% at pH 6.5 and higher at pH 7.0. The slight decrease at pH 6.5 (0.6%) is within the experimental uncertainty and does not affect the overall trend. Maximum adsorption efficiency of the PVDF-HFP film for Ca2+ ions is observed in the pH range of 6.8 to 7.0, compared to more acidic conditions. The lower percentage removal at the acidic pH is attributed to competition between the H+ and Ca2+ ions for binding to the available active sites of the film. These reduced percentages might also be attributed to the extended protonation of the adsorbent’s functional groups, caused by the introduction of 0.1 M HNO3 during the pH adjustment process. It was also observed that when the pH exceeded 7, the adsorption of the Ca2+ ions diminished, likely due to the generation of OH ions. Consequently, this led to the creation of metal hydroxide (Ca(OH)2), which precipitates as white crystals, thereby resulting in a reduced percentage of ion removal. Similar results were reported at pH of 6.5 by Muqeet et al., 2018 and Macevele et al., 2021 during the adsorption of Ca2+ ions and Cd2+ ions, respectively [25,30]. For the subsequent adsorption tests, a pH of 7.00 was chosen.

3.2.2. Effect of Film Adsorbent Dosage

Figure 5 illustrates the investigation into the influence of CA, PVDF-HFP, and a 3 wt.% PVDF-HFP/CA film adsorbent dosage on the hardness of Ca2+ ions. When the film dosages were varied from 0.1 to 0.9 mg/L, the removal efficiency increased. The CA, PVDF-HFP, and 3 wt.% PVDF-HFP/CA films attained their peak adsorption capacities for Ca2+ ions at 70%, 62%, and 80%, respectively, when an adsorbent dosage of 0.5 mg/L was employed. This rise can be attributed to a greater number of active sites available on the sorbent material. Furthermore, the enhanced Ca2+ removal observed for the PVDF-HFP/CA films can be explained by a synergistic mechanism wherein the polar functional groups of CA (–OH and –C=O) provide active binding sites for Ca2+ ions, while the incorporation of PVDF-HFP improves the stability and hydrophilic–hydrophobic balance of the film, thereby increasing the accessibility and effectiveness of these sites [45,46]. However, upon further increase in dosage, the films yielded a negligible Ca2+ ions adsorption, which is consistent with the data reported by Rolence et al., 2014 [47]. Nonetheless, this contrasted with the findings presented by Sepehr et al., 2013, who observed an increase in Ca2+ ions uptake as the dosage increased [18]. Interestingly, the 3 wt.% PVDF-HFP/CA film achieved much better Ca2+ ion adsorption (80% Ca2+ ions per 0.5 mg/L of adsorbent dosage) in comparison with the data reported by Sepehr et al., 2013 [18] and Werkneh et al., 2015 [19].

3.2.3. Influence of Contact Duration on the Adsorption of Ca2+ Ions onto Polymeric Films

Figure 6 illustrates the impact of time on the adsorption of Ca2+ ions by CA, PVDF-HFP, and 3 wt.% PVDF-HFP/CA films. The results indicate that over 90% removal was attained within the initial 30 min for all polymer films, and this level was maintained nearly stable up to 180 min. Initially, numerous vacant adsorption sites are available for Ca2+ ions; as these sites become saturated, the adsorption rate slowed down, nearing equilibrium [19]. These findings indicate that the PVDF-HFP film independently removed 94% of Ca2+ ions, with a marginal enhancement when CA was utilised. When 3 wt.% PVDF-HFP/CA film was used, roughly 99.7% of Ca2+ ions were removed within 30 min, and the removal rates remained stable for almost 120 min. The results surpass the previous 96% removal rate of Ca2+ ions, which required 120 min and a high dosage of 2 g/L, utilising alkali-modified sugarcane bagasse [19]. The optimal time chosen for the extraction of Ca2+ ions was 90 min. Unless otherwise specified, the adsorption studies presented here and onward were performed at pH 7 with a fixed adsorbent dosage of 0.5 mg/L.

3.2.4. Influence of Ca2+ Ion Concentration on Their Adsorption by Different Polymeric Films

Figure 7 demonstrates how varying the concentration influences the adsorption of Ca2+ ions on CA, PVDF-HFP, and 3 wt.% PVDF-HFP/CA films. As the concentration rose from 100 to 120 mg/L, the adsorption of Ca2+ ions by the PVDF-HFP film increased from 85% to 89%, whereas the CA film exhibited a minor reduction in adsorption from 89% to 87%. The 3 wt.% PVDF-HFP/CA film showed a slight increase from 99 to 99.5% with an increase in concentration of Ca2+ ions, which is slightly higher than the data reported by Sepehr et al., 2013 while using alkali-modified pumice stone [18]. The 3 wt.% PVDF-HFP/CA film reached saturation just above 300 mg/L of Ca2+ ions.

3.2.5. Effect of Temperature

Figure 8 illustrates how temperature affects the adsorption of Ca2+ ions using CA, PVDF-HFP, and 3 wt.% PVDF-HFP/CA films. The findings show that as temperature rises, the efficiency of Ca2+ ion removal by these films also rises, achieving 92%, 96%, and 97%, respectively. The enhanced adsorption rate is attributed to the increased thermal energy of the films at higher temperatures. This rise in temperature leads to swelling of the adsorbent materials, thereby exposing more active sites for the adsorption of Ca2+ ions [47].

3.3. Adsorption Kinetics of Ca2+ Ions Using Polymer-Based Films

Figure 9i,ii illustrate the behaviour of the pseudo-first-order and pseudo-second-order kinetic models in the adsorption of Ca2+ ions across CA, PVDF-HFP, and 3 wt.% PVDF-HFP/CA films. Table 1 outlines the sorption rate parameters (K1, K2, and qe) in addition to the coefficient of determination (R2) values for both kinetic models. The results indicate that the pseudo-second-order kinetic model (represented in Figure 9ii) provides a superior fit compared to the pseudo-first-order model (shown in Figure 9i). According to Table 1, the experimental qe values closely matched those predicted by the pseudo-second-order model, demonstrating that this model accurately represents the data for all film types: CA, PVDF-HFP, and 3 wt.% PVDF-HFP/CA. In addition, the high R2 values further support that the adsorption process adheres to a pseudo-second-order mechanism, as corroborated by the existing literature [18].

3.4. Models for Adsorption Isotherms

The outcomes of the adsorption tests were analysed using the Langmuir and Freundlich adsorption isotherm models. As illustrated in Figure 10, the adsorption isotherms are presented, with the calculated values detailed in Table 2. When attempting to fit the equilibrium data to both the Langmuir and Freundlich models, it was evident that the Freundlich isotherm model provided a superior fit, yielding higher R2 values ranging from 0.99633 to 0.99991, in comparison to the Langmuir R2 values between 0.96464 and 0.99957. This suggests that the Freundlich model, which indicates a surface with nonuniform characteristics and multilayer adsorption, aligns well with observations noted in previous studies [48]. Moreover, the Freundlich isotherm is an empirical model that is highly recommended because of its precision [49]. Each film’s equilibrium parameter values (specifically, RL) are less than one, suggesting a favourable adsorption as documented in the literature [25,50]. The Langmuir isotherm model for the 3 wt.% PVDF-HFP/CA film indicated a maximum adsorption capacity (qmax) of 56 mg/g. This value is higher than the qmax values reported in the literature, although smaller than the qmax of modified pumice adsorbent [18], while the percentage removal by the prepared film was still higher (99%) (Table 3).

3.5. Thermodynamics Studies of Ca2+ Ions Adsorption on 3 wt.% PVDF-HFP/CA

Figure 11 illustrates the linear graph of InKo against 1/T for the adsorption of Ca2+ ions on 3 wt.% PVDF-HFP/CA films. Utilising the slope and intercept from the van’t Hoff plots enabled the computation of the thermodynamic parameters ΔH° and ΔS°, which are detailed in Table 4. The effect of temperature (ranging from 20 to 35 °C) on the adsorption process of Ca2+ ions by 3 wt.% PVDF-HFP/CA films was examined, and the results are presented in Table 4. The adsorption process is endothermic, as evidenced by the positive ΔH° value of 313 KJ/mol (see Table 4). Conversely, the negative ΔG° value, that decreases with a rise in temperature, suggests that the adsorption of Ca2+ ions occurs spontaneously [38]. Furthermore, the sorption was accompanied by an increase in entropy, as supported by the positive value of ΔS°, which arises from the release of the structured hydration shell around Ca2+ ions upon adsorption [53].

3.6. Impact of Counterions

Actual water samples comprise diverse ionic species which could impact the adsorption of hardness-inducing agents. A simulation study using sodium sulphate, potassium nitrate, sodium chloride, and high-purity calcium sulphate (to simulate a real water sample) was undertaken to assess the impact of varying concentrations of metal ions. Figure 12 illustrates how the presence of counterions affects the adsorption of Ca2+ ions by films composed of 3 wt.% PVDF-HFP/CA. As indicated in Figure 12 (a), without counterions, the adsorption of Ca2+ ions by the 3 wt.% PVDF-HFP/CA film exceeded 99%, consistent with the findings in Section 3.2.5. However, in the presence of counter ions (Figure 12 (b)), a 3% drop in Ca2+ ion adsorption was observed. The drop is insignificant when compared to the results reported by Sepehr et al., 2013 [18], who observed about a 33% decrease in Ca2+ ion uptake when exposed to counterions. This is due to competition between the counterions and hardness-causing agents for the available active sites [18].

3.7. Influence of Binary System on the Uptake of Ca2+ Ions

Experiments were conducted under optimal conditions to investigate how Mg2+ ions interfere with the adsorption process of Ca2+ ions (i.e., 0.5 mg/L dosage, run time = 90 min, 120 mg/L Ca2+ ions, and pH = 7) (Figure 13). The 3 wt.% PVDF-HFP/CA films was able to adsorb 91% of Ca2+ ions, which is an 8% drop as compared to a single system. The adsorption of Ca2+ ions on a binary system reported by Sepehr et al., 2013 [18] had a 22% drop. These results suggest that the 3 wt.% PVDF-HFP/CA is highly selective towards Ca2+ ions adsorption and much better as compared to an alkali-modified pumice stone.

3.8. Reutilisation of Films

Reusability trials were performed to investigate the regeneration of used films, an essential factor for evaluating the economic viability of the method developed. A sodium hydroxide solution was utilised to conduct reusability tests on the spent 3 wt.% PVDF-HFP/CA film (refer to Figure 14). The 3 wt.% PVDF-HFP/CA film exhibited an adsorption loss of 0.34% (from 95% to 94.6%) after three adsorption cycles, as opposed to the 20% loss noted by Muqeet et al., 2018 [30]. Notably, even after three consecutive cycles, the adsorption efficiency was retained above 90%, suggesting that the 3 wt.% PVDF-HFP/CA film is reusable for the adsorption of Ca2+ ions.

4. Conclusions

This research focused on the softening of hard water by eliminating Ca2+ ions using films made from PVDF-HFP as adsorbents. Films composed of CA, PVDF-HFP, and 3 wt.% PVDF-HFP/CA were successfully synthesised through the phase inversion method. The FTIR, TGA, and SEM analyses of the PVDF-HFP/CA films confirmed that 3 wt.% PVDF-HFP was effectively incorporated into the CA structure. Concentrations of Ca2+ ions in synthetic water samples were examined through batch studies, optimising parameters like pH, duration, temperature, and adsorbent dosage. The 3 wt.% PVDF-HFP/CA film exhibited around 99% adsorption of Ca2+ ions. Using the Langmuir isotherm, the peak adsorption efficiency was found to be 56 mg/g at an optimal pH of 7, with an adsorbent dose of 0.5 mg/L at a 120 mg/L concentration of Ca2+ ions over 90 min. The adsorption kinetics and isotherm models were in line with the pseudo-second-order model. All films followed the Freundlich isotherm model, indicating that the sorption process involved heterogeneous adsorption. Additionally, the thermodynamic parameters suggested that the adsorption process was physical and endothermic. The presence of counterions and binary systems led to a minor reduction in Ca2+ ion adsorption on the 3 wt.% PVDF-HFP/CA film. Reusability tests showed that the 3 wt.% PVDF-HFP/CA film could be reused at least three times with only a 0.34% loss in adsorption efficiency. Consequently, this study indicates that PVDF-HFP-based films could serve as effective materials for reducing the hardness of Ca2+ ions to acceptable levels.

Author Contributions

K.V.R. prepared all samples, investigated all the data, and wrote the first draft of the manuscript. L.E.M. and A.A.A. co-supervised the research project, supported data analysis, and interpretation from various characterisation techniques. T.M. supervised the project, supported data interpretation, and contributed to the introduction of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

The authors express their gratitude to the National Research Foundation (NRF) for the support through NRF-grant linked bursary (TTK180412320177), NRF Thuthuka (TTK2203311375), the Water Research Commission (WRC) of South Africa, Project Number K5/2515//1, and the Sasol Inzalo Foundation for financial assistance.

Data Availability Statement

Data analysis pertaining to characterisation was conducted at the Council for Scientific and Industrial Research (CSIR) as well as at the University of Johannesburg’s Research Centre for Synthesis and Catalysis. Interested parties can obtain the experimental data by making a reasonable request to the corresponding author.

Acknowledgments

The authors greatly appreciate the support from the Department of Chemistry, University of Limpopo, and the University of Johannesburg for the opportunity to conduct this research. This article is based on the author Khaleke Veronicah Ramollo’s thesis [54].

Conflicts of Interest

The authors declare that there are no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
TGAThermogravimetric analysis
SEMScanning electron microscopy
FTIRFourier transform infrared
PVDF-HFPPoly(vinylidene fluoride-co-hexafluoropropylene)
CACellulose acetate

References

  1. Hailu, Y.; Tilahun, E.; Brhane, A.; Resky, H.; Sahu, O. Ion Exchanges Process for Calcium, Magnesium and Total Hardness from Ground Water with Natural Zeolite. Groundw. Sustain. Dev. 2019, 8, 457–467. [Google Scholar] [CrossRef]
  2. Kadir, N.N.A.; Shahadat, M.; Ismail, S. Formulation Study for Softening of Hard Water Using Surfactant Modified Bentonite Adsorbent Coating. Appl. Clay Sci. 2017, 137, 168–175. [Google Scholar] [CrossRef]
  3. Pourshadlou, S.; Mobasherpour, I.; Majidian, H.; Salahi, E.; Shirani Bidabadi, F.; Mei, C.T.; Ebrahimi, M. Adsorption System for Mg2+ Removal from Aqueous Solutions Using Bentonite/γ-Alumina Nanocomposite. J. Colloid Interface Sci. 2020, 568, 245–254. [Google Scholar] [CrossRef]
  4. Helte, E.; Säve-Söderbergh, M.; Larsson, S.C.; Åkesson, A. Calcium and Magnesium in Drinking Water and Risk of Myocardial Infarction and Stroke—A Population-Based Cohort Study. Am. J. Clin. Nutr. 2022, 116, 1091–1100. [Google Scholar] [CrossRef] [PubMed]
  5. Zhang, W.; Zhang, H.; Zhou, A. Smartphone Colorimetric Detection of Calcium and Magnesium in Water Samples Using a Flow Injection System. Microchem. J. 2019, 147, 215–223. [Google Scholar] [CrossRef]
  6. Lee, P.X.; Liu, B.L.; Show, P.L.; Ooi, C.W.; Chai, W.S.; Munawaroh, H.S.H.; Chang, Y.K. Removal of Calcium Ions from Aqueous Solution by Bovine Serum Albumin (BSA)-Modified Nanofiber Membrane: Dynamic Adsorption Performance and Breakthrough Analysis. Biochem. Eng. J. 2021, 171, 108016. [Google Scholar] [CrossRef]
  7. Li, C.; Liu, C.; Cao, Z.; Shan, M.; Bing, Y. Effect and Mechanism of Induced Crystallization Softening Treatment on Water Quality in Drinking Water Distribution System with High Hardness Water Source. J. Environ. Chem. Eng. 2023, 11, 110474. [Google Scholar] [CrossRef]
  8. Xiang, J.; Liu, T.; Hua, X.; Cheng, P.; Zhang, L.; Wang, S.; Du, W. The Future Prospect of China’s Independent R&D Technology (ITK) in Water Resources Utilization and Wastewater Treatment; Elsevier Inc.: Amsterdam, The Netherlands, 2020; ISBN 9780128183397. [Google Scholar]
  9. Voojodi, S.; Rastgoo, M.; Izadi-Darbandi, E.; Hajmohammadnia Ghalibaf, K.; Hasanfard, A. Tank-Mixing 2,4-D Amine and Sulfosulfuron Can Help Alleviate the Adverse Effects of Water Hardness on Controlling Flixweed [Descurainia sophia (L.) Webb Ex Prantl]. Crop Prot. 2023, 173, 106377. [Google Scholar] [CrossRef]
  10. Rastgoo, M.; Mirzaei, M.; Gherekhloo, J.; Hasanfard, A. Effect of Water Hardness Induced by Bicarbonate and Chloride Forms of Magnesium and Sodium on the Performance of Herbicides for Littleseed Canarygrass Control. J. Crop Prot. 2022, 11, 315–327. [Google Scholar]
  11. Abdullah, A.D. Modelling Approaches to Understand Salinity Variations in a Highly Dynamic Tidal River: The Case of the Shatt Al-Arab River; CRC Press: London, UK, 2017; pp. 1–188. [Google Scholar] [CrossRef]
  12. Elumalai, V.; Nethononda, V.G.; Manivannan, V.; Rajmohan, N.; Li, P.; Elango, L. Groundwater Quality Assessment and Application of Multivariate Statistical Analysis in Luvuvhu Catchment, Limpopo, South Africa. J. Afr. Earth Sci. 2020, 171, 103967. [Google Scholar] [CrossRef]
  13. Rao, G.P.; Lu, C.; Su, F. Sorption of Divalent Metal Ions from Aqueous Solution by Carbon Nanotubes: A Review. Sep. Purif. Technol. 2007, 58, 224–231. [Google Scholar] [CrossRef]
  14. da Silva Alves, D.C.; Healy, B.; Pinto, L.A.d.A.; Cadaval, T.R.S.; Breslin, C.B. Recent Developments in Chitosan-Based Adsorbents for the Removal of Pollutants from Aqueous Environments. Molecules 2021, 26, 594. [Google Scholar] [CrossRef]
  15. Pandiyan, R.; Dharmaraj, S.; Ayyaru, S.; Sugumaran, A.; Somasundaram, J.; Kazi, A.S.; Samiappan, S.C.; Ashokkumar, V.; Ngamcharussrivichai, C. Ameliorative Photocatalytic Dye Degradation of Hydrothermally Synthesized Bimetallic Ag-Sn Hybrid Nanocomposite Treated upon Domestic Wastewater under Visible Light Irradiation. J. Hazard. Mater. 2022, 421, 126734. [Google Scholar] [CrossRef] [PubMed]
  16. Sharma, S.; Bhattacharya, A. Drinking Water Contamination and Treatment Techniques. Appl. Water Sci. 2017, 7, 1043–1067. [Google Scholar] [CrossRef]
  17. Macevele, L.E.; Moganedi, K.L.M.; Magadzu, T. Effects of Poly(Vinylidene Fluoride-Co-Hexafluoropropylene) Nanocomposite Membrane on Reduction in Microbial Load and Heavy Metals in Surface Water Samples. J. Compos. Sci. 2024, 8, 119. [Google Scholar] [CrossRef]
  18. Sepehr, M.N.; Zarrabi, M.; Kazemian, H.; Amrane, A.; Yaghmaian, K.; Ghaffari, H.R. Removal of Hardness Agents, Calcium and Magnesium, by Natural and Alkaline Modified Pumice Stones in Single and Binary Systems. Appl. Surf. Sci. 2013, 274, 295–305. [Google Scholar] [CrossRef]
  19. Ayaliew Werkneh, A. Removal of Water Hardness Causing Constituents Using Alkali Modified Sugarcane Bagasse and Coffee Husk at Jigjiga City, Ethiopia: A Comparative Study. Int. J. Environ. Monit. Anal. 2015, 3, 7. [Google Scholar] [CrossRef]
  20. Shi, L.; Wang, R.; Cao, Y.; Feng, C.; Liang, D.T.; Tay, J.H. Fabrication of Poly(Vinylidene Fluoride-Co-Hexafluropropylene) (PVDF-HFP) Asymmetric Microporous Hollow Fiber Membranes. J. Memb. Sci. 2007, 305, 215–225. [Google Scholar] [CrossRef]
  21. Stephan, A.M.; Kumar, S.G.; Renganathan, N.G.; Kulandainathan, M.A. Characterization of Poly(Vinylidene Fluoride-Hexafluoropropylene) (PVdF-HFP) Electrolytes Complexed with Different Lithium Salts. Eur. Polym. J. 2005, 41, 15–21. [Google Scholar] [CrossRef]
  22. Macevele, L.E.; Moganedi, K.L.M.; Magadzu, T. Investigation of Antibacterial and Fouling Resistance of Silver and Multi-Walled Carbon Nanotubes Doped Poly(Vinylidene Fluoride-Co-Hexafluoropropylene) Composite Membrane. Membranes 2017, 7, 35. [Google Scholar] [CrossRef]
  23. Wang, X.; Xiao, C.; Liu, H.; Huang, Q.; Hao, J.; Fu, H. Poly(Vinylidene Fluoride-Hexafluoropropylene) Porous Membrane with Controllable Structure and Applications in Efficient Oil/Water Separation. Materials 2018, 11, 443. [Google Scholar] [CrossRef]
  24. Gao, C.; Li, X.; Wei, G.; Wang, S.; Zhao, X.; Kong, F. Cellulose Acetate Propionate Incorporated PVDF-HFP Based Polymer Electrolyte Membrane for Lithium Batteries. Compos. Commun. 2022, 33, 101226. [Google Scholar] [CrossRef]
  25. Macevele, L.E.; Lydia, K.; Moganedi, M.; Magadzu, T. Adsorption of Cadmium (II) Ions from Aqueous Solutions Using Poly (Amidoamine)/Multi-Walled Carbon Nanotubes Doped Poly (Vinylidene Fluoride-Co-Hexafluoropropene) Composite Membrane. J. Membr. Sci. Res. 2021, 7, 152–165. [Google Scholar] [CrossRef]
  26. Rajesh, S.; Maheswari, P.; Senthilkumar, S.; Jayalakshmi, A.; Mohan, D. Preparation and Characterisation of Poly (Amide-Imide) Incorporated Cellulose Acetate Membranes for Polymer Enhanced Ultrafiltration of Metal Ions. Chem. Eng. J. 2011, 171, 33–44. [Google Scholar] [CrossRef]
  27. Asghar, M.R.; Zhang, Y.; Wu, A.; Yan, X.; Shen, S.; Ke, C.; Zhang, J. Preparation of Microporous Cellulose/Poly(Vinylidene Fluoride-Hexafluoropropylene) Membrane for Lithium Ion Batteries by Phase Inversion Method. J. Power Sources 2018, 379, 197–205. [Google Scholar] [CrossRef]
  28. Jabbour, L.; Bongiovanni, R.; Chaussy, D.; Gerbaldi, C.; Beneventi, D. Cellulose-Based Li-Ion Batteries: A Review. Cellulose 2013, 20, 1523–1545. [Google Scholar] [CrossRef]
  29. Kuribayashi, I. Characterization of Composite Cellulosic Separators for Rechargeable Lithium-Ion Batteries. J. Power Sources 1996, 63, 87–91. [Google Scholar] [CrossRef]
  30. Muqeet, M.; Khalique, A.; Qureshi, U.A.; Mahar, R.B.; Ahmed, F.; Khatri, Z.; Kim, I.S.; Brohi, K.M. Aqueous Hardness Removal by Anionic Functionalized Electrospun Cellulose Nanofibers. Cellulose 2018, 25, 5985–5997. [Google Scholar] [CrossRef]
  31. Tohdee, K.; Kaewsichan, L.; Asadullah. Enhancement of Adsorption Efficiency of Heavy Metal Cu(II) and Zn(II) onto Cationic Surfactant Modified Bentonite. J. Environ. Chem. Eng. 2018, 6, 2821–2828. [Google Scholar] [CrossRef]
  32. Ahmad, A.; Razali, M.H.; Mamat, M.; Mehamod, F.S.B.; Anuar Mat Amin, K. Adsorption of Methyl Orange by Synthesized and Functionalized-CNTs with 3-Aminopropyltriethoxysilane Loaded TiO2 Nanocomposites. Chemosphere 2017, 168, 474–482. [Google Scholar] [CrossRef]
  33. Ren, X.; Yang, L.; Liu, M. Kinetic and Thermodynamic Studies of Acid Scarlet 3r Adsorption onto Low-Cost Adsorbent Developed from Sludge and Straw. Chin. J. Chem. Eng. 2014, 22, 208–213. [Google Scholar] [CrossRef]
  34. Zhang, C.; Dai, Y.; Wu, Y.; Lu, G.; Cao, Z.; Cheng, J.; Wang, K.; Yang, H.; Xia, Y.; Wen, X.; et al. Facile Preparation of Polyacrylamide/Chitosan/Fe3O4 Composite Hydrogels for Effective Removal of Methylene Blue from Aqueous Solution. Carbohydr. Polym. 2020, 234, 115882. [Google Scholar] [CrossRef]
  35. Kahu, S.S.; Shekhawat, A.; Saravanan, D.; Jugade, R.M. Two Fold Modified Chitosan for Enhanced Adsorption of Hexavalent Chromium from Simulated Wastewater and Industrial Effluents. Carbohydr. Polym. 2016, 146, 264–273. [Google Scholar] [CrossRef]
  36. Alghamdi, M.M.; El-Zahhar, A.A.; Idris, A.M.; Said, T.O.; Sahlabji, T.; El Nemr, A. Synthesis, Characterization, and Application of a Novel Polymeric-Bentonite-Magnetite Composite Resin for Water Softening. Sep. Purif. Technol. 2019, 224, 356–365. [Google Scholar] [CrossRef]
  37. Le, T.T.N.; Le, V.T.; Dao, M.U.; Nguyen, Q.V.; Vu, T.T.; Nguyen, M.H.; Tran, D.L.; Le, H.S. Preparation of Magnetic Graphene Oxide/Chitosan Composite Beads for Effective Removal of Heavy Metals and Dyes from Aqueous Solutions. Chem. Eng. Commun. 2019, 206, 1337–1352. [Google Scholar] [CrossRef]
  38. Mustapha, S.; Tijani, J.O.; Ndamitso, M.M.; Abdulkareem, S.A.; Shuaib, D.T.; Mohammed, A.K.; Sumaila, A. The Role of Kaolin and Kaolin/ZnO Nanoadsorbents in Adsorption Studies for Tannery Wastewater Treatment. Sci. Rep. 2020, 10, 13068. [Google Scholar] [CrossRef]
  39. Nthumbi, R.M.; Adelodun, A.A.; Ngila, J.C. Electrospun and Functionalized PVDF/PAN Composite for the Removal of Trace Metals in Contaminated Water. Phys. Chem. Earth 2017, 100, 225–235. [Google Scholar] [CrossRef]
  40. Istiningrum, R.B.; Tiwow, C.D.; Nuryono; Narsito. Au(III) Selective Adsorption of Quaternary Ammonium-Silica Hybrid in Au/Cu System. Procedia Chem. 2015, 17, 132–138. [Google Scholar] [CrossRef]
  41. El-Gendi, A.; Abdallah, H.; Amin, A.; Amin, S.K. Investigation of Polyvinylchloride and Cellulose Acetate Blend Membranes for Desalination. J. Mol. Struct. 2017, 1146, 14–22. [Google Scholar] [CrossRef]
  42. Sagar, S.; Iqbal, N.; Maqsood, A. Dielectric, Electric and Thermal Properties of Carboxylic Functionalized Multiwalled Carbon Nanotubes Impregnated Polydimethylsiloxane Nanocomposite. J. Phys. Conf. Ser. 2013, 439, 012024. [Google Scholar] [CrossRef]
  43. Das, A.M.; Ali, A.A.; Hazarika, M.P. Synthesis and Characterization of Cellulose Acetate from Rice Husk: Eco-Friendly Condition. Carbohydr. Polym. 2014, 112, 342–349. [Google Scholar] [CrossRef]
  44. Asghar, M.R.; Anwar, M.T.; Rasheed, T.; Naveed, A.; Yan, X.; Zhang, J. Lithium Salt Doped Poly(Vinylidene Fluoride)/Cellulose Acetate Composite Gel Electrolyte Membrane for Lithium Ion Battery. IOP Conf. Ser. Mater. Sci. Eng. 2019, 654. [Google Scholar] [CrossRef]
  45. Zi, X.; Wu, H.; Song, J.; He, W.; Xia, L.; Guo, J.; Luo, S.; Yan, W. Electrospun Sandwich-like Structure of PVDF-HFP/Cellulose/PVDF-HFP Membrane for Lithium-Ion Batteries. Molecules 2023, 28, 4998. [Google Scholar] [CrossRef]
  46. Shi, R.J.; Wang, T.; Lang, J.Q.; Zhou, N.; Ma, M.G. Multifunctional Cellulose and Cellulose-Based (Nano) Composite Adsorbents. Front. Bioeng. Biotechnol. 2022, 10, 891034. [Google Scholar] [CrossRef]
  47. Rolence, C.; Machunda, R.L.; Njau, K.N. Water Hardness Removal by Coconut Shell Activated Carbon. Int. J. Sci. Technol. Soc. 2014, 2, 97–102. [Google Scholar] [CrossRef]
  48. Bibiano-Cruz, L.; Garfias, J.; Salas-García, J.; Martel, R.; Llanos, H. Batch and Column Test Analyses for Hardness Removal Using Natural and Homoionic Clinoptilolite: Breakthrough Experiments and Modeling. Sustain. Water Resour. Manag. 2016, 2, 183–197. [Google Scholar] [CrossRef]
  49. Joshi, P.; Manocha, S. Sorption of Cadmium Ions onto Synthetic Hydroxyapatite Nanoparticles. Mater. Today Proc. 2017, 4, 10460–10464. [Google Scholar] [CrossRef]
  50. Rahman, M.; Gul, S.; Ajmal, M.; Iqbal, A.; Achakzai, A. Removal of Cadmium from Aqueous Solutions Using Excised Leaves of Quetta Pine (Pinus halepensis mill.). Bangladesh J. Bot. 2014, 43, 277–281. [Google Scholar] [CrossRef]
  51. Lestari, A.Y.D.; Malik, A.; Sukirman; Ilmi, M.I.; Sidiq, M. Removal of Calcium and Magnesium Ions from Hard Water Using Modified Amorphophallus Campanulatus Skin as a Low Cost Adsorbent. MATEC Web Conf. 2018, 154, 4–7. [Google Scholar] [CrossRef]
  52. Calgan, E.; Ozmetin, E. Optimization of Hardness Removal Using Response Surface Methodology from Wastewater Containing High Boron by Bigadiç Clinoptilolite. Desalination Water Treat. 2019, 172, 24907. [Google Scholar] [CrossRef]
  53. Snoussi, Y.; Abderrabba, M.; Sayari, A. Removal of Cadmium from Aqueous Solutions by Adsorption onto Polyethylenimine-Functionalized Mesocellular Silica Foam: Equilibrium Properties. J. Taiwan Inst. Chem. Eng. 2016, 66, 372–378. [Google Scholar] [CrossRef]
  54. Ramollo, K.V. Preparation of Poly(vinylidene fluoride-co-hexafluoropropylene) Composite Membranes for Treatment of Water Hardness. Master’s Thesis, University of Limpopo, Polokwane, South Africa, 2022. [Google Scholar]
Scheme 1. Preparation and characterisation of cellulose acetate-based thin films.
Scheme 1. Preparation and characterisation of cellulose acetate-based thin films.
Physchem 05 00045 sch001
Figure 1. FTIR spectra of PVDF-HFP, CA, and 3 wt.% PVDF−HFP/CA films.
Figure 1. FTIR spectra of PVDF-HFP, CA, and 3 wt.% PVDF−HFP/CA films.
Physchem 05 00045 g001
Figure 2. Thermogravimetric analysis profiles of PVDF-HFP, CA, and films containing 3 wt.% PVDF-HFP blended with CA.
Figure 2. Thermogravimetric analysis profiles of PVDF-HFP, CA, and films containing 3 wt.% PVDF-HFP blended with CA.
Physchem 05 00045 g002
Figure 3. SEM images of (a) PVDF-HFP, (b) CA, and (c) films containing 3 wt.% PVDF-HFP blended with CA.
Figure 3. SEM images of (a) PVDF-HFP, (b) CA, and (c) films containing 3 wt.% PVDF-HFP blended with CA.
Physchem 05 00045 g003
Figure 4. Impact of pH on the uptake of Ca2+ ions by PVDF-HFP films.
Figure 4. Impact of pH on the uptake of Ca2+ ions by PVDF-HFP films.
Physchem 05 00045 g004
Figure 5. Impact of film adsorbent dosage on the adsorption of Ca2+ ions by (a) CA, (b) PVDF-HFP, and (c) 3 wt.% PVDF-HFP/CA films.
Figure 5. Impact of film adsorbent dosage on the adsorption of Ca2+ ions by (a) CA, (b) PVDF-HFP, and (c) 3 wt.% PVDF-HFP/CA films.
Physchem 05 00045 g005
Figure 6. Influence of the impact of time on the adsorption of Ca2+ ions by (a) CA, (b) PVDF-HFP, and (c) 3 wt.% PVDF-HFP/CA films.
Figure 6. Influence of the impact of time on the adsorption of Ca2+ ions by (a) CA, (b) PVDF-HFP, and (c) 3 wt.% PVDF-HFP/CA films.
Physchem 05 00045 g006
Figure 7. Impact of starting concentrations of Ca2+ ions on the adsorption performance of (a) CA, (b) PVDF-HFP, and (c) 3 wt.% PVDF-HFP/CA films.
Figure 7. Impact of starting concentrations of Ca2+ ions on the adsorption performance of (a) CA, (b) PVDF-HFP, and (c) 3 wt.% PVDF-HFP/CA films.
Physchem 05 00045 g007
Figure 8. Influence of temperature on the adsorption of Ca2+ ions by (a) CA, (b) PVDF-HFP, and (c) 3 wt.% PVDF-HFP/CA films.
Figure 8. Influence of temperature on the adsorption of Ca2+ ions by (a) CA, (b) PVDF-HFP, and (c) 3 wt.% PVDF-HFP/CA films.
Physchem 05 00045 g008
Figure 9. Pseudo-first-order (i) and pseudo-second-order (ii) kinetic models describing the adsorption of Ca2+ ions onto (a) CA films, (b) PVDF-HFP films, and (c) films composed of 3 wt.% PVDF-HFP/CA.
Figure 9. Pseudo-first-order (i) and pseudo-second-order (ii) kinetic models describing the adsorption of Ca2+ ions onto (a) CA films, (b) PVDF-HFP films, and (c) films composed of 3 wt.% PVDF-HFP/CA.
Physchem 05 00045 g009
Figure 10. Adsorption of Ca2+ ions is described by (i) Langmuir and (ii) Freundlich isotherms for the following films: (a) CA, (b) PVDF-HFP, and (c) 3 wt.% PVDF-HFP/CA.
Figure 10. Adsorption of Ca2+ ions is described by (i) Langmuir and (ii) Freundlich isotherms for the following films: (a) CA, (b) PVDF-HFP, and (c) 3 wt.% PVDF-HFP/CA.
Physchem 05 00045 g010aPhyschem 05 00045 g010b
Figure 11. Thermodynamic parameters for the adsorption of Ca2+ ions on 3 wt.% PVDF-HFP/CA films. Conditions: 0.5 mg/L dosage, duration = 90 min, 120 mg/L concentration of Ca2+ ions, and pH = 7.
Figure 11. Thermodynamic parameters for the adsorption of Ca2+ ions on 3 wt.% PVDF-HFP/CA films. Conditions: 0.5 mg/L dosage, duration = 90 min, 120 mg/L concentration of Ca2+ ions, and pH = 7.
Physchem 05 00045 g011
Figure 12. Impact of counterions on the adsorption of Ca2+ ions by 3 wt.% PVDF-HFP/CA films: (a) in the absence of counterions and (b) in the presence of counterions. Experimental conditions: 0.5 mg/L dosage, duration = 90 min, 120 mg/L concentration of Ca2+ ions, and pH = 7.
Figure 12. Impact of counterions on the adsorption of Ca2+ ions by 3 wt.% PVDF-HFP/CA films: (a) in the absence of counterions and (b) in the presence of counterions. Experimental conditions: 0.5 mg/L dosage, duration = 90 min, 120 mg/L concentration of Ca2+ ions, and pH = 7.
Physchem 05 00045 g012
Figure 13. Impact of (a) isolated and (b) binary systems on the uptake of Ca2+ ions by a 3 wt.% PVDF-HFP/CA film (Conditions: 0.5 mg/L dosage, run time = 90 min, 120 mg/L Ca2+ ions, and pH = 7).
Figure 13. Impact of (a) isolated and (b) binary systems on the uptake of Ca2+ ions by a 3 wt.% PVDF-HFP/CA film (Conditions: 0.5 mg/L dosage, run time = 90 min, 120 mg/L Ca2+ ions, and pH = 7).
Physchem 05 00045 g013
Figure 14. Impact of recycling 3 wt.% PVDF-HFP/CA films on the adsorption of Ca2+ ions (Conditions: 0.5 mg/L dosage, run time = 90 min, 120 mg/L Ca2+ ions, and pH = 7).
Figure 14. Impact of recycling 3 wt.% PVDF-HFP/CA films on the adsorption of Ca2+ ions (Conditions: 0.5 mg/L dosage, run time = 90 min, 120 mg/L Ca2+ ions, and pH = 7).
Physchem 05 00045 g014
Table 1. Adsorption kinetics parameters for Ca2+ ions on the prepared films.
Table 1. Adsorption kinetics parameters for Ca2+ ions on the prepared films.
FilmExperimental qe
(mg/g)
Pseudo-First-Order
Kinetic Model
Pseudo-Second-Order
Kinetic Model
K1
(min−1)
qe
(mg·g−1)
R2K2
(g/mg·min)
qe
(mg·g−1)
R2
CA57.570.041.730.870.1057.601.00
PVDF-HFP57.210.010.680.570.1357.140.99
3 wt.% PVDF-HFP/CA59.940.000.08−0.200.0759.600.99
Table 2. Parameters of Langmuir and Freundlich isotherms for the adsorption of Ca2+ ions by films.
Table 2. Parameters of Langmuir and Freundlich isotherms for the adsorption of Ca2+ ions by films.
Filmqmax
(mg/g)
Langmuir Model Freundlich Model
b (L/mg)R2RLKf
(mg/g)
nR2
CA50.63−1.260.99−0.0116,268.22−11.290.99
PVDF-HFP51.71−1.660.96−0.0115,000.30−13.700.99
3 wt.% PVDF-HFP/CA55.71−6.460.99−0.00112,944.94−26.660.99
Table 3. Evaluation of the adsorption capabilities of different materials for Ca2+ ions.
Table 3. Evaluation of the adsorption capabilities of different materials for Ca2+ ions.
AdsorbentConditions of the ExperimentAdsorption Capacity (mg/g)% RemovalReference(s)
3 wt.% PVDF-HFP/CApH = 7, Temp = 298 K, Dosage = 0.5 mg/L5699This study
Modified pumice adsorbentspH = 6, Temp = 293 K, Dosage = 10 g/L 62.3496[18]
Coconut shell activated carbonpH = 6.30, Temp = 303 K, Dosage = 0.16 g/cm348.5060[47]
Alkali-modified sugarcane bagasse pH = 7.50, Temp = 298 K, Dosage = 2.5 g/L52.978 [19]
Natural and homoionic clinoptiloliteNot specified 10.50Not specified[48]
Surfactant-modified bentonite adsorbent coatingNot specified29.2766.67[2]
Modified Amorphophallus campanulate skin as a low-cost adsorbent Not specified 10.85 85[51]
Bigadic clinoptiloliteTemp = 299 K, Dosage = 20 g/L, Time = 93 min12.3099[52]
Natural zeolitepH = 6.9, Dosage = 50 g/L-80.2[1]
Table 4. Thermodynamic variables of Ca2+ ion adsorption using 3 wt.% PVDF-HFP/CA films.
Table 4. Thermodynamic variables of Ca2+ ion adsorption using 3 wt.% PVDF-HFP/CA films.
FilmTemperature
(K)
Thermodynamic Variables
∆G (KJ/mol)∆H (KJ/mol)∆S (J.mol/K)
3 wt.% PVDF-HFP/CA293.15−7514.85314.331.44
298.15−8037.39
303.15−8642.53
308.15−8918.86
Conditions: 0.5 mg/L dosage, duration = 90 min, 120 mg/L concentration of Ca2+ ions, and pH = 7.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ramollo, K.V.; Macevele, L.E.; Ambushe, A.A.; Magadzu, T. Preparation of Poly(vinylidene fluoride-co-hexafluoropropylene) Doped Cellulose Acetate Films for the Treatment of Calcium-Based Hardness from Aqueous Solution. Physchem 2025, 5, 45. https://doi.org/10.3390/physchem5040045

AMA Style

Ramollo KV, Macevele LE, Ambushe AA, Magadzu T. Preparation of Poly(vinylidene fluoride-co-hexafluoropropylene) Doped Cellulose Acetate Films for the Treatment of Calcium-Based Hardness from Aqueous Solution. Physchem. 2025; 5(4):45. https://doi.org/10.3390/physchem5040045

Chicago/Turabian Style

Ramollo, Khaleke Veronicah, Lutendo Evelyn Macevele, Abayneh Ataro Ambushe, and Takalani Magadzu. 2025. "Preparation of Poly(vinylidene fluoride-co-hexafluoropropylene) Doped Cellulose Acetate Films for the Treatment of Calcium-Based Hardness from Aqueous Solution" Physchem 5, no. 4: 45. https://doi.org/10.3390/physchem5040045

APA Style

Ramollo, K. V., Macevele, L. E., Ambushe, A. A., & Magadzu, T. (2025). Preparation of Poly(vinylidene fluoride-co-hexafluoropropylene) Doped Cellulose Acetate Films for the Treatment of Calcium-Based Hardness from Aqueous Solution. Physchem, 5(4), 45. https://doi.org/10.3390/physchem5040045

Article Metrics

Back to TopTop