P. H. P. Panapitiyaa,
Tharindra Weerakoon
a,
M. Shanika Fernandoa,
A. K. D. V. K. Wimalasiria,
K. M. Nalin de Silva
*a and
Rohini M. de Silva
*ab
aCentre for Advanced Materials and Devices (CAMD), Department of Chemistry, University of Colombo, Colombo 03, Sri Lanka. E-mail: rohini@chem.cmb.ac.lk
bDepartment of Life Sciences, Faculty of Science, NSBM Green University, Mahenwatta, Pitipana, Homagama 10206, Sri Lanka
First published on 5th August 2025
A new, non-toxic, environmentally friendly nanocomposite, based on graphene oxide and biopolymers was developed as an effective adsorbent for water hardness removal. Graphene oxide was synthesized by a modified Hummers' method, whereas crosslinked carboxymethyl cellulose (CMC) and chitosan were used as biopolymers. Montmorillonite (MMT) was utilized as an additive to enhance the adsorbent's performance. The solubility and adsorption behaviours of the prepared materials were investigated in respect to calcium ions (Ca2+) and magnesium ions (Mg2+). Among the prepared materials, a film containing graphene oxide (GO), crosslinked carboxymethyl cellulose (CMC), and montmorillonite (MMT), denoted as GO-CMC-MMT-3, exhibited the highest water softening capacity. The successful synthesis of the materials was confirmed through scanning electron microscopy (SEM), Fourier-transform infrared spectroscopy (FT-IR), X-ray diffraction (XRD), and ultraviolet-visible spectroscopy (UV-Vis). The optimal pH for the adsorption process was around 6–7. Kinetic studies showed that second-order kinetic model described the adsorption process. The thermodynamic analysis indicated that the process was endothermic in nature and showed a reduced degree of spontaneity at the given conditions. The Langmuir isotherm model was the best fit, and the adsorption capacities were 6.46 mg g−1 for Mg2+ and 7.98 mg g−1 for Ca2+, which indicated the formation of a monolayer of cations on homogeneous adsorption sites. Further, an investigation on the reusability of gravity filtration was carried out, demonstrating the practical utility of the GO-CMC-MMT-3 membrane in real-life water treatment.
Water hardness is primarily determined by the concentrations of dissolved calcium (Ca2+) and magnesium (Mg2+) ions. The issues caused by water hardness range from an unpleasant taste to excessive mineral scaling, which obstructs water flow in pipes, damages equipment such as boilers, and shortens the lifespan of fabrics. Additionally, it reduces soap's effectiveness by forming soap scum through the complexation of metal ions with soap molecules.
Hard and very hard water, with Ca2+ and Mg2+ levels of 120 ppm or above can significantly impact on human health. Although the body can regulate the absorption of these ions in the intestine up to a certain level, excessive intake will cause serious health complications.1 In Sri Lanka, the chronic kidney disease of unknown etiology (CKDu) is attributed to various chemical factors related to water hardness.2,3 Studies have shown a strong correlation between the geographical distribution of CKDu patients and the areas with high ground water hardness, where as 96% of CKDu patients reside in areas with hard or very hard water.4 Due to these risk factors associated with water hardness, it is recommended to soften the water before consumption. Various methods are applied on both industrial and domestic scales to achieve water softening.
The temporary hardness of water can easily be removed through methods like, boiling, adding slaked lime (Clark's method),5 or using washing soda.6 However, the permanent hardness requires more advanced methods such as Gan's permutate method,7 Calgon's process, carbonation,8 electrocoagulation,9 capacitive deionization (CDI),10 electrolysis,11 use of ion exchange resins,12,13 ultrasounds with ion exchange,14 precipitation using pellet reactor,15 membrane separation systems, nanofiltration membranes,16 freezing with dry gas,17 electrodialysis,18 electro deionization(EDI),19 and various adsorption methods.20,21 Among these techniques, absorption stands out due to its simplicity and cost effectiveness, which can be applicable at both industrial and domestic scales, making it the focus of interest in this study.
According to the literature different materials have been studied as absorbents. Some of them are activated carbon,22,23 different lignocellulose materials,24,25 plant-based ashes,26 synthetic and derived natural compounds,27,28 synthetic and natural zeolites,29,30 clay-based materials,31 polymer-based materials,27 and ion exchange resins.32,33 Nanomaterials such as carbon nanotubes,34 graphene-based materials35 have also been given significant attention as adsorbents due to their high absorption efficiency. However, concerns about high cost and toxicity, make nanomaterials less favourable. As a result, this study mainly focuses nanocomposite, which offers advantages like enhanced adsorption and mechanical properties, cost reduction, high characterization potential, and minimized toxicity concerns. Among nanocomposites, biopolymer-based materials have given a considerable interest in water purification, particularly bio nanocomposites (nanocomposites consist of biopolymers) due to their non-toxicity,36,37 excellent chelation properties and biodegradability.38,39
According to Dimiev et al. (2013) and Lakshmi et al. (2017), graphene-based nanocomposites have shown an excellent adsorption capacity toward a wide range of contaminants,40 and also water hardness.41,42 Carboxymethyl cellulose (CMC) and chitosan have also been widely investigated for water purification due to their biocompatibility, abundance, and high chelating properties.5,43–46 However, there is a lack of well-rounded exploration of the effectiveness of bio-nanocomposites on the removal of water hardness. Hence, this study will help to examine their potential in improving the adsorption of Ca2+ and Mg2+ ions. Furthermore, cross-linking of the biopolymer CMC with citric acid and glycerol were studied as these components are simple, cheap, and non-toxic crosslinking agents. This approach is expected to improve the characteristics of the composite, including its negative charge, mechanical strength, antifouling properties, chelation ability toward negatively charged groups, cation exchange capacity, and ability to entrap MMT clay and GO layer particles.41
The layers of MMT are held together mainly by van der Waals forces and electrostatic interactions. The intercalation of the MMT layers is readily attained with different polymers by means of isomorphous substitution, which imparts a negative charge on adjacent oxygen atoms. This negative charge enhances the ability of the clay to attract positively charged molecules and cations. With its high surface area, excellent mechanical stability, and notable swelling capacity, this polymer demonstrates superior adsorption properties and high cation exchangeability.47,48
This work was mainly focused on developing an effective biopolymer nanocomposite that can be used to remove water hardness by the adsorption of Ca2+ and Mg2+ ions via different interactions. In this research, melt intercalation/melt mixing process is used to form the nanocomposite which is more compatible with industrial polymer process.49,50 For this study, GO was taken as the nanomaterial and it was synthesized using Sri Lankan vein graphite which is known as the best graphite in the world.
The initial comparative adsorption studies indicated that GO-CMC-MMT-3 membrane is the best material for the removal of Ca2+ and Mg2+ from water. After the formation of nanocomposite, SEM images, FT-IR data, and XRD data was used to determine the intercalation and alternations of GO in cross-liked CMC matrix. FT-IR analysis was also performed for pure CMC powder, MMT powder, and crosslinked CMC membrane in the wavelength range between 400 cm−1 and 4000 cm−1. The XRD analysis was performed using Cu-Kα radiation over a 2θ range of 2–80°, with a step size of 0.02° for crosslinked CMC membrane, GO-CMC-MMT-3 before and after the adsorption Ca2+ and Mg2+ cations. The SEM analysis was performed with 10.00 KV voltage and 5.00 KX, 25.00 KX magnifications.
Langmuir adsorption isotherm
![]() | (1) |
Freundlich adsorption isotherm
log Qe = log Kf + n log Ce | (2) |
ln (qe − qt) = ln qe + k1t… | (3) |
![]() | (4) |
![]() | (5) |
ΔG° = −RT ln kd | (6) |
![]() | (7) |
ΔG° = ΔH° − TΔS° | (8) |
![]() | (9) |
![]() | ||
Fig. 1 The FT – IR spectrum of (a) Aqueous GO sample, (b) GO membrane; solid GO membrane formed by using vacuum oven. |
The other form of GO was prepared as a membrane, and its FT-IR spectrum is given in Fig. 1(b). The reduced peak size in the spectrum indicates the less water content in the membrane after drying in the vacuum oven. The peaks due to carbonyl stretching (1700–1750 cm−1) and double bond stretching in graphene structure (1600–1650 cm−1) also can be seen in the spectrum.55 The broad peak at 1630 cm−1 in Fig. 1 a results due to overlapping of peaks in aqueous GO. The small peaks around 1000–1200 cm−1 in Fig. 1(b) are attributed to epoxy groups in the membrane which are not clearly visible in 1(a).55 The overlapping of the peaks in the fingerprint region (400–1500 cm−1) make it difficult to identify individual peaks.
The XRD spectrum was taken in the range of 2θ from 2° to 80°. This XRD spectrum (Fig. 2) of the thin layer of GO shows a sharp peak at 2θ around 9.5° and also there is a small peak at 2θ around 19°. According to the literature, the GO shows its (002) diffraction at the 2θ = 9.70°. Reduced graphene Oxide (rGO) shows its (002) diffraction at 2θ = 23.7°, graphite powder and graphene show its (002) diffraction at 2θ = 26°.55,56
In this study, graphite powder was used as the starting material. Therefore the XRD of Graphite has two characteristic peaks at 2θ = 26° (002) and 2θ = 55° (004)57 and the resulted XRD spectrum of GO sample shows two peaks at the position of 2θ = 9.5° (002) and 2θ = 19°. The d space was calculated using Bragg's law equation, nλ = 2d sin θ where n is an integer for constructive interferences, λ = wavelength of the x rays, d = spacing between layers of atoms and θ = angle between the incident rays and the surface of the crystal. The calculated values were tabulated in the (Table S1†). According to the inter layer spacing values of products and reactants, it can be concluded that 2θ = 19° peak appears due to GO sheets that are not fully exfoliated to GO.
After oxidation, the characteristic peaks of graphite were shifted from 26° to 9.5° indicating an increase of inter planer distance of graphene sheets. This serves as strong evidence for formation of GO. Also, the absence of characteristic peaks of graphite and graphene in the spectrum suggests that their presence is negligible in the sample.
GO samples were used in diluted concentrations in UV-visible spectrometry, as high concentrations can lead to scattering of the radiation. According to UV-visible spectrum (Fig. 3), peak at 300 nm can be attributed to n – π* transition of the carbonyl carbon. The expected peak that is correspond to π – π* transition of the double bond between carbon atoms is absent in the spectrum.55
Following the oxidation, 3.00 g of graphite, produced a total of 5.06 g of GO. However, according to the literature, 15–25% of water can be trapped even after the oven drying process. Therefore, the real value of GO can be slightly lower than that of the reported value.
To remove water molecules formed during crosslinking process, the initial solution was heated to 180 °C and graphene oxide was added only after the crosslinks had formed. To avoid the disturbance of the GO structure, the solution was allowed to cool down to about 100 °C before introducing the GO. According to the FT-IR spectrum (Fig. 4b) of CMC powder, there is a wide absorption peak between 3000–3400 cm−1 due to stretching vibrations of – OH groups. Peak at 2900 cm−1 is correspond to CH2– stretching vibrations, and peak at 1590 cm−1 is due to stretching vibrations of carbonyl groups (–COO–). Also, the peaks at 1420 cm−1, 1325 cm−1, and 1050 cm−1 can be attributed to –CH2– scissoring, –OH bending, and –C–O–C– stretching frequencies respectively.58
![]() | ||
Fig. 4 The FT – IR spectrum of (a) cross-linked CMC and CMC powder, (b) the expanded FT- IR spectrum of CMC powder. |
Peaks at 3000 cm−1, 3400 cm−1 and 2900 cm−1 were almost unchanged after the crosslink formation, while absorbance values were increased compared to CMC powder. That may be due to increased concentration of –OH groups and –CH2-(Fig. 4a). And also due to the stretching vibrations of protonated carboxyl groups in citric acid, a novel peak at 1718 cm−1 was visible apart from the peak at 1590 cm−1 which is due to the stretching vibrations of carbonyl groups. Peaks at 1408 cm−1 and 1320 cm−1 are correspond to –CH2– scissoring and –OH bending vibrations respectively. Peak at 1220 cm−1 can be due to –C–OH groups on the citric acid units.
Peak at 1050 cm−1 in CMC powder has been split into three peaks as at 1100 cm−1 (due to –O–C– stretching) 1033 cm−1, and 919 cm−1 (due to –C–O–C– stretching) in cross linked CMC spectrum and these peaks have been shifted from the original position, mainly due to the stability of H bonds formed in the polysaccharide molecules.58,59 Based on the above data, high efficiency of cross link formation can be observed.1
The results of the solubility studies and comparative adsorption studies are provided and explained in the ESI using Tables S2, S3,† and Fig. 6.60,61
![]() | ||
Fig. 6 (a) The graph of percentage of adsorption of Ca2+, (b) the graph of percentage of adsorption of Mg2+. |
According to the results, the adsorption percentage of Ca2+ ion is comparatively higher in the membrane than that of Mg2+ ion. This can be explained by the hydration of metal ions in aqueous solutions and hydration effect of the membrane functional groups. In aqueous solutions, the metal ions are existing as hydrated ions in which the metal ions are surrounded by H2O molecules. This occurs due to dipole–dipole interactions between the positive charge metal ion and the negative charge oxygen atom in the H2O molecules. These H2O molecules form a structured shell around the metal ion.
The Mg2+ which has the higher charge density compared to Ca2+, forms a larger hydration shell due to strong interactions between water molecules and metal ions and conversely, larger metal ions with low charge density like Ca2+ have smaller hydration shells due to weak interactions between water molecules and metal ions. Because of these weak interactions, during shaking, weak shells with weak interactions undergo rearrangements or loss partially or completely. The metals ions with lower charge density easily undergo these rearrangements and lost its hydration shell. This will lead to a difference in ion selectivity in the same membrane. When considering Ca2+ and Mg2+ ions, the ionic radius of Ca2+ is larger than that of Mg2+. Therefore, the Mg2+ has a strong and larger hydration shell and Ca2+ has a soft and smaller hydration shell which leads to higher interactions between Ca2+ and negative charge on the membrane compared to Mg2+.38 Adsorption properties are also affected by the hydration effect of the membrane functional groups. Once the functional groups on the membrane which have the higher negative charge, interact with the positive polarity of the water molecules, membrane gets hydrated and the affinity of membrane for strongly hydrated cations such as Mg2+ get decrease and less hydrated cations interactions increase due to electrostatic interactions.38
During the competitive uptake of Ca2+ and Mg2+ ions from an aqueous solution, the membrane should exhibit simultaneous adsorption behaviour for both ions.62 But according to the results, Ca2+ ions in GO-CMC membrane gives highest adsorption percentage (61.50%) with in a contact time of 60 minutes (Table S3† and Fig. 6) while Mg2+ ions give the highest adsorption capacity of 53.00% in contact time of 45 minutes. When evaluating the adsorption percentages of both Ca2+ and Mg2+, GO-CMC-MMT-3 emerges as the most effective material, achieving the highest removal percentage of total hardness. Consequently, this material was deemed the best and subjected to further adsorption studies.
This membrane specifically includes 2.50 g of MMT clay in addition to CMC and GO. During the formation of nanocomposite membranes, three membranes containing varying amounts of MMT were produced. It was observed, by increasing the amount of MMT enhanced Mg2+ adsorption; however, exceeding 2.50 g of MMT caused the membrane to dissolve in water. This is because each layer consists of two types of structural sheets: octahedral and tetrahedral. The tetrahedral sheets are composed of silicon–oxygen tetrahedra, while the octahedral sheets are situated between two tetrahedral sheets. The octahedral structure is formed by aluminium or magnesium ions in six-fold coordination with oxygen atoms and hydroxyl groups from the adjacent tetrahedral sheets. Multiple layers in the MMT clay are held together by weak van der Waals forces or interlayer cations.63
Isomorphous substitution of clay mineral generates a charge. In this process, silicon in the tetrahedral coordination places, replace by aluminium, as aluminium is small enough to fit into these sites. Similarly, aluminium in octahedral coordination sites can replace by large ions such as magnesium or iron. Larger cations such as Ca2+, Cs+, K+ which are also referred as interlayer cations, can be seen in between layers. Mg2+ ions can occupy octahedral coordination places and also in between layers while Ca2+ can only occupy in between layers. Therefore, MMT clay has higher potential to intake Mg2+ compared to Ca2+. Consequently, the absorption of Mg2+ ions increase with an increase in MMT content in the composite.63
As there is higher concentration gradient of adsorbent and also large number of vacant sites on the adsorbent, the contact time of the membrane at room temperature in neutral pH, increase the absorption percentage. Once the equilibrium concentration achieved, and all the vacant sites are occupied, adsorption percentage becomes constant. To achieve the equilibrium point, Mg2+ ions take approximately 45 minutes while Ca2+ take less than 10 minutes. Therefore, it can be concluded, the optimum contact time for the highest absorption of Ca2+ and Mg2+ ions by the membrane is 45 minutes.
![]() | ||
Fig. 7 (a) The expanded FT – IR spectrum of GO-CMC-MMT-3 membrane, (b) the FT – IR spectrums of constituent materials such as GO, MMT powder (MMT std) and cross-linked CMC membrane. |
The broad peak around 3300 cm−1 corresponds to the OH− stretching vibrations of H2O present in GO, CMC, and the interlayer water of MMT particles in the composite. The small peaks at 2920 cm−1 and 2853 cm−1 arise from the stretching vibrations of –CH2– groups in the cross-linked CMC structure. The peak at 1715 cm−1 represents the protonated carbonyl stretching vibrations of GO and the citric acid component of CMC in the composite. Peaks at 1625 cm−1 and 1591 cm−1 are attributed to carbon–carbon double bonds in GO, H2O bending vibrations in MMT, and the stretching vibrations of carbonyl groups in CMC, respectively.
Additionally, the peaks at 1407 cm−1, 1374 cm−1, and 1315 cm−1 correspond to –CH2– scissoring and OH− bending vibrations in CMC particles in the composite membrane.53,57,58,61 The peaks at 1220 cm−1 and 1095 cm−1 arise due to vibrations of – C–OH groups in citric acid and epoxy groups in GO. The peaks at the position of 1068 cm−1 and 921 cm−1 on the composite membrane arise due to –C–O–C– stretching vibrations of CMC matrix and Si–O stretching of MMT particles. Also, there are large number of peaks in the fingerprint region of GO-CMC-MMT-3 membrane due to MMT particles in the composite.55,58,59,64 These characteristic peaks confirm the effective composite formation from constituent materials.
![]() | ||
Fig. 8 (a) The XRD spectrums of constituent materials of the GO-CMC-MMT-3 membrane, (b) The XRD spectrum of GO-CMC-MMT-3 membrane. |
This shifting result mainly due to increment of interplanar distance of constituent materials during the composite formation. The peak at 2θ = 8.06° in the Fig. 8b, is due to peaks of 7.1° of MMT powder, 9.4° CMC powder, 9.5° of GO after the intercalation.48 Peak 10.06° in the XRD spectrum of GO-CMC-MMT-3 membrane is due to the shift of 11° peak of MMT clay. Also, the peaks at 16.46°, 17.84°, and 19.62° resulted by the shift of peaks of CMC 20.1° and GO 19°. The shifts in peaks and the presence of characteristic peaks from different constituent materials provide evidence for the successful formation of the nanocomposite with effective intercalation. The peaks at higher 2θ values cannot be distinctly identified, as they represent overlapping from both CMC and MMT.
According to the SEM images of GO-CMC-MMT-3 membrane (Fig. 10(a) and (b)), there are clear sheet/layer structures with thickness lower than 100 nm (and roughly in 200–400 nm range in length and width), that was not observed in the image of CMC-MMT membrane (Fig. 9). This observation is attributed to the intercalation of GO particles within the GO-CMC-MMT-3 membrane. Additionally, Fig. 9b and 10b clearly show GO particles embedded in the cross-linked CMC membrane.
The SEM images of the nanocomposite provide strong evidence of the successful intercalation and incorporation of GO and MMT particles67 into the cross-linked CMC polymer matrix.
![]() | ||
Fig. 11 The changes of the percentage of adsorption for 45 minutes time duration with different pH values. |
As the pH increases, the competition between Ca2+, Mg2+ and H+ decreases, leading to higher percentage of adsorption. The adsorption percentage increased to its maximum value around pH 6, which is the natural pH of Ca2+ and Mg2+ solutions. Therefore, the optimal pH range for GO-CMC-MMT-3 membrane is considered to be 6–7. At higher pH values, such as above 8, Ca2+ and Mg2+ ions tend to precipitate as hydroxides due to their reaction with OH−. Consequently, the concentration of Ca2+ and Mg2+ ions decreases at higher pH levels.62
Langmuir model | |||||
Ce/Q0 = 1/Q0. KL + Ce/Q0 | |||||
y = C + mx | |||||
Ion | R2 | 1/Q0. KL | 1/Q0 | Q0 (mg g−1) | KL (min −1) |
Mg2+ | 0.9802 | 0.3337 | 0.1546 | 6.4683 | 2.1584 |
Ca2+ | 0.9618 | 0.4998 | 0.1252 | 7.9872 | 3.9920 |
Freundlich model | |||||
Log Qe = log K + (1/n) log Ce (y = C + mx) | |||||
Ion | R2 | 1/n (m) | n | Log k (c) | K (g mg−1 min−1) |
Mg2+ | 0.7044 | 0.1399 | 7.14 | 0.5680 | 3.6983 |
Ca2+ | 0.9227 | 0.2700 | 3.70 | 0.4574 | 2.8668 |
Ca2+ adsorption | Mg2+ adsorption | ||||||||
---|---|---|---|---|---|---|---|---|---|
20 °C | 30 °C | 40 °C | 60 °C | 20 °C | 30 °C | 40 °C | 60 °C | ||
Experimental qe | 9.50 | 9.10 | 8.58 | 10.07 | 4.57 | 13.29 | 3.66 | 3.16 | |
GO-CMC-MMT-3 Lagergren first order model | qe (mg g−1) | 1.97 | 2.01 | 2.57 | 1.51 | 1.61 | 5.32 | 0.77 | 0.59 |
k1 (min−1) | −0.006 | −0.016 | −0.010 | −0.001 | 0.003 | −0.037 | 0.004 | 0.012 | |
R2 | 0.196 | 0.350 | 0.219 | 0.003 | 0.026 | 0.974 | 0.098 | 0.569 | |
GO-CMC-MMT-3 Ho's pseudo-second-order model | qe (mg g−1) | 8.02 | 9.32 | 8.52 | 8.35 | 2.02 | 13.77 | 2.82 | 1.89 |
k1 (min−1) | −0.32 | 0.03 | 0.015 | −0.093 | −0.100 | 0.017 | −0.264 | −0.112 | |
R2 | 0.987 | 0.993 | 0.872 | 0.964 | 0.866 | 0.994 | 0.942 | 0.858 |
According to Table 2, the adsorption process of the GO-CMC-MMT-3 membrane shows a higher correlation coefficient (the calculated and experimental qe are close to each other) with Ho's pseudo second order kinetic model.65 Therefore, this Ho's pseudo second order kinetic model is the best model to explain adsorption of Ca2+ and Mg2+ ions on the surface of the nanocomposite GO-CMC-MMT-3. The pseudo-second-order kinetic model is widely regarded as the most suitable model for describing adsorption processes involving chemisorption between the adsorbate and the adsorbent. This suggests that a chemical interaction, such as an exchange or replacement reaction, occurs during the process. In case of replacement, Ca2+ and Mg2+ ions from the bulk solution can replace the water molecules or H+ ions on the surface of the nanocomposite.
In addition to that, the effect of temperature on the adsorption rate can be explained by using kinetic data. According to the results, higher adsorption capacity of 13.3 mg g−1 exhibit in Mg2+ ions at 30 °C, once the nanocomposite reaches its equilibrium capacity after 45 minutes. Even in other temperatures (20 °C, 40 °C, 60 °C) nanocomposite reaches its equilibrium capacity after 45 minutes, but the equilibrium capacity (also highest capacity) gets a lower value (3.7 mg g−1). Same pattern was observed in Ca2+ adsorption as well. Ca2+ also reached its equilibrium temperature after around 45 minutes with 9.0 mg g−1 equilibrium adsorption capacity.46 The kinetic data clearly demonstrate that the adsorption rate increases with rising temperature. This behaviour can be attributed to the temperature-dependent nature of the rate constant and the activation energy of the adsorption process.
Thermodynamic parameters for the adsorption (Table 3) can be calculated using thermodynamic equations (eqn (5)–(9)) and plots of “ln Kap vs. 1/T” (Arrhenius activation energy plot, (Fig. S13 in ESI†), “ln Kd vs. 1/T” (Van't Hoff plot, Fig. S14 in ESI†). The data indicate that all apparent activation energies for the adsorption processes are positive. As the temperature increases, more adsorbate and adsorbent particles overcome the activation energy barrier, thereby enhancing the reaction rate. As presented in Table 3, the overall standard Gibbs free energy during the adsorption is negative, and that values gradually decrease upon increasing the temperature indicating that the adsorption is more favorable at high temperatures. Therefore, the degree of feasibility of the ion adsorption increases with increasing temperature. Additionally, the negative standard Gibbs free energy values found in the adsorption system indicated that the ion adsorption process is spontaneous. Furthermore, the enthalpy change of the system is also positive, and it indicates the endothermic behavior of the adsorption. The entropy change of the system was positive, and it indicates the randomness at the solid-solution interface increased during the adsorption process.
T (K) | ΔG° (kJ mol−1) | ΔH° kJ mol−1) | ΔS° (J/K mol) | Activation energy (kJ mol−1) | |
---|---|---|---|---|---|
3rd GO/CMC/MMT Ca2+ adsorption | 293.15 | −0.37 | 6.79 | 24.50 | 10.55 |
303.15 | −0.70 | ||||
313.15 | −0.85 | ||||
3rd GO/CMC/MMT Mg2+ adsorption | 293.15 | −0.10 | 6.0 | 20.88 | 26.67 |
303.15 | −0.45 | ||||
313.15 | −0.50 |
Adsorption of Ca2+ and Mg2+ ions are influenced by their hydration properties and interaction with the membrane. Ca2+ forms a weaker hydration shell than Mg2+, leading to higher adsorption. In contrast, Mg2+ has a stronger hydration shell that limits its adsorption. Further membrane hydration also causes to adsorb less hydrated ions like Ca2+. In this situation by adding montmorillonite (MMT) enhances Mg2+ adsorption, as Mg2+ can occupy both octahedral and interlayer sites in the clay structure, while Ca2+ is limited to interlayer sites. Thus, Ca2+ shows higher adsorption on the membrane alone, whereas Mg2+ adsorption increases when increasing the MMT content.
Apart from that, the overall negative surface charges are normally neutralized by counterions (H+) or molecules, forming two distinct layers in colloidal chemistry: the inner layer and the diffusion layer. The inner layer consists of a monolayer of counterions, and molecules firmly bound to the surface through chemical interactions and can be considered as an inherent part of the nanocomposite surface.
The counter ions in the diffusion layer are loosely associated with the surface by Coulomb interactions and actively exchangeable with ions from the bulk solution depending on the concentration gradient. This major mechanism was used in this study to remove Ca2+ and Mg2+ ions from the bulk solution.68
The important parameters that would influence the adsorption process, including optimal pH and contact time, were studied. Accordingly, it was found that the optimal pH for both Ca2+ and Mg2+ ions is at around pH 6–7, while the optimal contact time is 45 minutes. Thermodynamic analysis depicts that the adsorption process is endothermic and less spontaneous. Reusability studies showed the nanocomposite membrane was efficient for reuse up to two cycles. At the same time satisfactory results were given by the gravity filtration study. Therefore, based on these findings, it can be concluded that the prepared GO-CMC-MMT-3 nanocomposite is a potential adsorbent for the removal of Ca2+ and Mg2+ ions from hard water.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ra00562k |
This journal is © The Royal Society of Chemistry 2025 |