Chamalki
Madhusha
a,
Kavindya
Weerasinghe
a,
Imalka
Munaweera
ab,
Chandani
Perera
c,
Gayan
Wijesinghe
d,
Manjula
Weerasekera
ef,
Yugantha
Idangodage
g,
C. S.
Kalpage
h and
Nilwala
Kottegoda
*a
aDepartment of Chemistry, Faculty of Applied Sciences, University of Sri Jayewardenepura, Gangodawila, Nugegoda, Sri Lanka. E-mail: nilwala@sjp.ac.lk
bInstrument Center, Faculty of Applied Sciences, University of Sri Jayewardenepura, Gangodawila, Nugegoda, Sri Lanka
cDepartment of Chemistry, University of Peradeniya, Peradeniya, Sri Lanka
dDepartment of Medical Laboratory Sciences, Faculty of Allied Health Sciences, University of Sri Jayewardenepura, Gangodawila, Nugegoda, Sri Lanka
eDepartment of Microbiology, Faculty of Medical Sciences, University of Sri Jayewardenepura, Gangodawila, Nugegoda, Sri Lanka
fSri Lanka Institute of Biotechnology (SLIBTEC), Pitipana, Homagama, Sri Lanka
gDepartment of Chemical Process and Engineering, University of Moratuwa, Sri Lanka
hDepartment of Chemical and Process Engineering, Faculty of Engineering, University of Peradeniya, Peradeniya, Sri Lanka
First published on 15th November 2022
This study reports the synthesis and characterization of a phosphate functionalized porous activated carbon filter material suitable for sachet filters to remove fluoride, hardness and bacterial pathogens in drinking water. Phosphate functionalized activated carbon is prepared by treating waste coconut coir dust with phosphoric acid/water vapor followed by pyrolysis under an inert environment at 500 °C for 1 h. Unlike other reported chemical functionalization methods, the current approach allows controlling the amount of surface functional groups, thus facilitating the maintenance of pH and conductivity of treated water. Morphological studies and BET surface area analysis of phosphoric acid vapor activated carbon (FAC) suggest the presence of a highly porous network structure with pore sizes ranging from the micro to nano-range (2.2 μm–3.23 nm). Fourier transform infrared and X-ray photoelectron spectroscopy and the shifts in the D and G bands in Raman spectroscopy analysis confirm the functionalization of the activated carbon matrix with phosphate and hydroxyl groups. The isoelectric point of the filter material is 5.9. The hardness and fluoride removal efficiencies of 1.0 g of FAC in a sachet for a contact time of 20 min under static conditions are found to be approximately 45–60% and 45–50%, respectively, for natural water samples. Adsorption complies with the Freundlich isotherm model, and the adsorption kinetics is well described by the pseudo-second-order model. The laboratory scale method has been scaled up and the process parameters have been optimized.
Water impactThis study reports the synthesis and characterization of a functionalized porous activated carbon filter material suitable for sachet filters to remove fluoride, hardness and bacterial pathogens in drinking water. Simply, these domestic water filter/sachet bags can be used anytime, anywhere without the need for any high-tech involvement and they could be an alternative to plastic portable water bottles used worldwide, thus minimizing the use of plastics on the planet. |
Currently, there is an increased interest in exploring the potential of advanced technology to solve or ameliorate many of the technical challenges involved in providing potable water to the world population. Most of the attempts focus on ensuring long term water quality, availability and viability. Currently, the main focus of water treatment is establishment of regional level treatment plants, in particular, most of the techniques are focused on removing turbidity, solid matter, and pathogens. The available technologies are mainly flocculation and coagulation, filtration, and disinfection.7 Flocculation and coagulation remove the turbidity in water, which reduces the supporting structure of microorganisms. Filtration removes microorganisms by size exclusion, whereby microorganisms larger than the pore size of the filter will be retained within the system. UV disinfection and chlorine further destroy the microorganisms in the system to a safe level.
Membrane filtration is a mature technology and is one of the most effective drinking water treatment processes. It provides an absolute barrier for microorganisms, retaining them within the water source. However, membrane filters require an external driving force such as electrical pumping. In the past decade, innovative designs and optimization allowed manufacturers to produce off-grid or non-electric driven membrane filter systems.8 Companies such as Wateroam, Icon Lifesaver, and Villagepump integrated a manual pump within the membrane system, making the system grid-independent. Grid-independent water filtration systems are crucial because many places around the world do not have continuous electricity supply. In South Asia, more than 50% of people living in rural areas do not have electricity.9 Biosand filtration (BSF) is a promising point-of-use (POU) treatment process with more than 500000 people using it worldwide.10 It can be easily constructed using raw materials that are locally sourced. The concept of BSF is similar to that of a conventional slow sand filter. However, BSF experiences varying flow rates and intermittent filtration through the sand layer. The outlet of the POU is located higher than the sand layer, allowing the sand layer to be saturated with water throughout the operation. This allows for microbial growth, developing biofilms around the filter media and sand particles.
However, these techniques do not offer an efficient platform to remove fluoride, hardness and heavy metals, simultaneously. Over 70% of the water is inaccessible due to the presence of these ions and the most effective practically used technique to remove all these ions is reverse osmosis (RO). However, due to the complexity of the RO system, requirement for stable electricity supply, and maintenance, RO systems are not commonly employed in rural areas. In addition, the use of RO also increases the operation cost of the water treatment system.11
Functionalized activated carbon has been used as a water filtering medium for purification of drinking water for many years. It is widely used for the removal of contaminants in water due to its high capacity for adsorption of such compounds, arising from its large surface area and porosity. Generally, activated carbon can remove the total suspended solids (TSS) and the biological oxygen demand (BOD) effectively over 99%, to 1 mg l−1 and can also improve the tastes and odors of the drinking water.12 This is the reason for the use of activated carbon as a major filter medium in most of the home water filtration systems. However, activated carbon does not effectively remove water hardness, fluoride and pathogens.
The present study reveals a sachet filter based on phosphate-functionalized activated carbon derived from coconut coir dust. The requirement for such portable and convenient water purification systems particularly exists in remote/underprivileged areas where potable water supply is not readily available. Similarly, portable water systems are often required in the aftermath of disasters such as floods and droughts where the local water supply has been contaminated or disrupted. In addition, portable water availability is practically needed for field workers, school children, aerospace activities, recreational activities such as camping and hiking and any other activities which last for extended periods. Simply, this domestic water filter/sachet bag can be used anytime, anywhere without the need for any high-tech involvement and it could be an alternative to plastic portable water bottles used worldwide, thus minimizing the use of plastics on the planet.
Pyrolyzed coconut coir dust without acid functionalization (AC) was prepared under the same conditions as a control material.
Thermo Avantage (version-5.982) software was used to process the obtained spectra. The spectra were corrected for charge compensation effects by offsetting the binding energy relative to the intrinsic aliphatic component of the C1s spectrum to 284.8 eV, which is a known carbon status for the activated carbon sample. High-resolution spectra were resolved by fitting each peak with a combined Gaussian/Lorentzian function after subtracting the background (maximum iterations – 100, convergence – 1 × 10−5, fitting algorithm – Powell).
The thermal stability of AC and FAC was investigated via a thermogravimetric analyzer (TA Instruments SDT 650) in the thermogravimetric analyzer temperature range from room temperature to 700 °C under a nitrogen atmosphere at 10 °C min−1 ramp.
The carbon particle leaching into water was tested by examining the water samples (after treating with FAC sachet filters) under a high resolution light microscope at different time intervals, up to 24 h.
The linear form of the PFO model is given by the following equation,
ln(qe − qt) = lnqe − K1t |
The linear form of the PSO model is expressed by the following equation,
The linear form of the Langmuir isotherm equation is given as follows,
The logarithmic form of the Freundlich isotherm is given by the following equation,
P2O5(l) + H2O(l) ↔ 2H3PO4(l) |
Fig. 1 Schematic representation of the synthesis procedure of FAC and its application on water purification. |
Type of activated carbon material | Iodine number (mg g−1) | Methylene blue number (mg g−1) | ||||
---|---|---|---|---|---|---|
300 °C | 500 °C | 700 °C | 300 °C | 500 °C | 700 °C | |
AC | 501 ± 2 | 572 ± 1 | 588 ± 1 | 138 ± 2 | 197 ± 1 | 199 ± 1 |
FAC | 915 ± 1 | 918 ± 1 | 900 ± 2 | 154 ± 1 | 230 ± 2 | 212 ± 2 |
Based on the above results, further investigations of adsorption characteristics were carried out using activated carbon by selecting FAC obtained after 1 h vapor treatment and pyrolyzed at 500 °C.
The morphological and structural characteristics of FAC were studied using SEM, TEM, PXRD, Raman spectroscopy, FT-IR and XPS.
The micrograph of AC without acid treatment (Fig. 2(A)) shows that the surface is heterogeneous and porous forming a network structure. Functionalization with phosphoric acid catalyzes the oxidation and bond cleavage reactions of the interior structure of coir dust where most of the organic volatiles are removed creating an abundance of pores with different sizes. This is clearly demonstrated by the development of the well-defined porous structure with tunnel-shaped pores of different sizes as shown in the SEM image of FAC (Fig. 2(B)). Such observations were reported by Xu et al., 2014, and Oginni et al., 2019 for activated carbon prepared from reedy grass leaves by chemical activation with H3PO4 and activated carbon from Kanlow switchgrass and public miscanthus by two-step H3PO4 activation, respectively.27,28 The inner structure of activated carbon with and without acid functionalization is viewed by TEM. Fig. 2(C) reveals a layered like carbon morphology probably due to the formation of micropores in activated carbon. The carbon frame/pore walls and pores of activated carbon are visualized as dark and light areas, respectively, in the TEM image.29,30 Light shading from the TEM results indicates the development of large amounts of pores after phosphoric acid functionalization, in comparison with the virgin AC. It also confirms the defective nature of graphene layers indicating the non-planarity and microporosity.31
Fig. 2 SEM images of (A) AC and (B) FAC at 1000× magnification, and TEM images of (C) AC and (D) FAC. |
The change of crystallinity or extent of graphitization of AC and FAC is revealed by the PXRD characterization. As depicted in Fig. 3(A), two typical diffraction peaks of activated carbon appear at 2θ = 23.04° and 2θ = 41.44° for AC, corresponding to the (002) and (100) planes, respectively, which is in good agreement with literature data.32,33 According to Xu and coworkers, the strong and broad diffraction peak indicates the presence of several disordered graphitic crystals whereas the weak peak (100 plane) reflects the partial graphitization of activated carbon and the random turbostratic stacking of layers.27,34 Functionalization of activated carbon with phosphate groups has resulted in PXRD patterns with relatively increased 2θ values (25.66° and 43.94°) compared to the AC. Xie et al. suggested that the degree of graphitization can be related to the interplanar spacing d002 of carbon materials so that the degree of graphitization becomes higher with lowering the value of the diffraction angle.35 Therefore, functionalization of the carbon network has led to a higher degree of disorder.
Raman spectroscopy reveals the information on the structure and the bonding types of different carbon based materials. The Raman spectra of AC and FAC are shown in Fig. 3(B). Both spectra are dominated by the two intense bands, D and G, which are the two signature bands for typical graphitic carbon materials.36 The G band corresponds to the stretching vibrations of sp2 carbon pairs in the chain and the ring structures, whereas the D band is attributed to the vibrations of the breathing mode in the aromatic carbon ring.37 As shown in Fig. 3(B), the presence of D and G bands clearly demonstrates the heterogeneous microstructure of the activated carbon material. Accordingly, the D band is observed at 1336 cm−1 and 1319 cm−1 for AC and FAC, respectively, while the G band appears at 1573 cm−1 and 1586 cm−1 for AC and FAC, respectively. Since the D mode becomes active only in the presence of disorder, it confirms the disordered graphitic structure of both AC and FAC.38 Moreover, there is a blue shift in the G band of FAC compared to AC (1573 to 1586 cm−1) which is probably due to merging of the G peak with a second peak D′ (around 1620 cm−1) resulting in a net increase of the G band position. According to Elcey and Manoj, such peaks around 1620 cm−1 become well resolved as the extent of disorder in the crystal lattice increases.39 In addition, the blue shift is attributed to the consequences of the electron transfer between carbon and phosphate groups inducing strong modifications to the carbon material.40 On the other hand, the red shift in the D band of FAC compared to AC (from 1336 to 1319 cm−1) also demonstrates the structural changes to the aromatic carbon ring with the introduction of phosphate groups. The Raman spectrum of FAC consists of an additional broad band in the region of 300–500 cm−1 which is less broadened. According to Gupta and co-workers, such peaks are assigned to PO43− in a triply degenerate bending mode.41 However, observations by Sun et al. suggested that the region around 300–500 cm−1 indicates the presence of phosphorus bonds and probably P–C bonds in the activated carbon.42 The relative ratio of intensities of D and G bands (ID/IG) represents the degree of structural disorder43 or inversely, the degree of graphitization where a lower ID/IG ratio indicates a higher degree of graphitization.32 Therefore, the graphitization degree has decreased in FAC (ID/IG = 0.82) compared to AC (ID/IG = 0.84) attributed to the effect of phosphoric acid functionalization on the carbon material, thus resulting in increased defects on the graphitic structure.28
FTIR spectra (Fig. 3(C)) were obtained to further characterize the surface groups as they provide information about the chemical structure of activated carbon. Assigning of functional groups is mentioned in Table S2.† The broad absorption band at 3600–3200 cm−1 is characteristic of the stretching vibration of hydrogen bonded –OH groups. AC shows low intensity bands at 2923 and 2860 cm−1 corresponding to stretching vibrations of aliphatic C–H, which have disappeared in the spectrum of FAC. This indicates the increased aromaticity of activated carbon upon functionalization. The low intense band around 1700 cm−1 in AC is assigned to CO stretching vibrations of ketones, aldehydes, lactones or carboxyl groups. The weak intensity of this peak in FAC suggests that phosphate functionalized carbon contains only a small amount of carboxyl groups.44,45 The spectrum of AC also shows a strong band at 1600–1580 cm−1 attributed to CC vibrations in aromatic rings. The red shift in FAC (from 1593 to 1575 cm−1) suggests an enlargement of the aromatic ring system after acid activation.44 The broad band at around 1240 cm−1 in AC is assigned to the C–O asymmetric stretching of aromatic ethers, esters and phenols.46 Interestingly, this band has shifted from 1240 to 1195 cm−1 in FAC due to the overlapping of the C–O asymmetric stretching band with the newly appeared band of stretching vibrations of hydrogen bonded PO groups from phosphates or polyphosphates, O–C stretching vibration in the P–O–C (aromatic) linkage and POOH.47 Furthermore, the presence of the weak and overlapped band located at 1080 cm−1 in FAC is caused by the stretching vibrations of P–O–P in polyphosphates or the ionized linkage P+–O− in acid phosphate esters.48 Even though the assignment of peaks in this region is difficult due to overlapped absorption bands, similar observations have been reported in previous research work.44,49 Thus, this confirms the existence of chemically bonded phosphorus containing groups on activated carbon after impregnating with phosphoric acid, successfully supporting the previously explained Raman spectral data. In addition, some weak bands also appeared in AC in the range of 600–900 cm−1 associated with the out of plane bending mode of the C–H or O–H groups.49 The disappearance of the C–H band in FAC at 750 cm−1 suggests the substitution of C–H bonds in the aromatic system to form new C–R bonds.44 This observation is in line with the strong band around 665 cm−1 probably due to increased aromatic ring vibrations.17 Although, according to Puziy et al., the C–P bond also falls into the region of 670 cm−1, it is not confirmed by the other characterization data obtained in the present study.50 This is further supported by the observations by Lee and Radovic showing that the structures with C–O–P bonding are more stable than the structures with C–P–O bonding.51
XPS is used to analyze the surface functional groups of the activated carbon and phosphoric acid functionalized activated carbon. The survey scan spectra of AC and FAC are shown in Fig. S2.† They indicate the presence of C 1s, O 1s and P 2p in FAC at the peak positions of approximately 285 eV, 532 eV and 133 eV, respectively, thus confirming the carbon, oxygen and phosphorus containing functional groups in synthesized phosphoric acid activated carbon.
The high resolution XPS spectra of C 1s excitation for both AC and FAC showed a complex envelope exhibiting several carbon species at the carbon surface (Fig. 4(A) and (B)). The most intense band of AC, located at around 286.25 eV, is assigned to carbon species in alcohol, phenol, and ether groups, and C–O–P and/or C–O–C linkages and the peak at 284.75 eV is attributed to graphitic sp2 carbon.43 These peaks have shifted to higher binding energies of 286.45 eV and 284.8 eV, respectively, after phosphoric acid functionalization (Table S3†). Such shifts to higher binding energies arise due to the interaction of phosphorus species with the graphitic carbon matrix. The peak at 288.3 eV is usually assigned to carbon species in carboxylic groups or esters. More interestingly, a shake-up satellite due to π–π* transitions in aromatic rings (291.4 eV) has appeared on the carbon surface of FAC.52 In addition, the contribution of the 286.45 eV peak is considerably higher than the 288.3 eV peak, reflecting the significant presence of C–O–P and/or C–O–C bonds.27 However, according to the literature, phosphorus compounds cannot be clearly distinguished from the C 1s spectral region because the binding energy of C–O–P bonding is similar to the binding energy in alcohol and ether groups. Moreover, the C 1s electron binding energy of phosphonates (compounds with direct C–P bonding) lies between graphitic and oxygenated carbon.50
Fig. 4 High-resolution XPS spectra of (A) C 1s of AC, (B) C 1s of FAC, (C) O 1s of AC, (D) O 1s of FAC and (E) P 2p of FAC. |
The O 1s spectra of AC and FAC give further evidence to support the conclusion regarding the changes in carbon surface composition. Accordingly, the O 1s spectra of both AC and FAC are shown in Fig. 4(C) and (D), indicating broad peaks as a result of different chemical states of oxygen present. In FAC, the peak at 533.25 eV is attributed to singly bonded oxygen in the C–OH, C–O–C and/or C–O–P groups. The peak at 534.8 eV is ascribed to non-carbonyl oxygen in carboxylic groups. These bands are positioned at higher binding energies compared to AC implying changes in electron density around oxygen atoms due to interaction of phosphorus with the surface functional groups.50
Fig. 4(E) shows the high-resolution spectrum of P 2p in FAC. The major peak at 133.54 eV is attributed to the formation of phosphates, pyrophosphates and phosphonates, while the peak at 134.2 eV is assigned to metaphosphates and phenyl-phosphate.53 The P 2p peak was fitted with the two components P1/2 and P3/2, corresponding to spin–orbit coupling. The P 2p peak of P2O5 has been reported to have a binding energy of 133.54 eV, while unoxidized P has a binding energy of 134.2 eV. The peaks were fitted with a spin–orbit coupling of 0.66 eV. It can be suggested that the interaction of phosphoric acid with the surface phenolic and carbonyl groups may be the reason for the formation of P-containing carbonaceous species. Thus, the high resolution spectrum of P 2p proves the impregnation of the carbon precursor by phosphoric acid supporting the FT-IR spectra results and successful chemical activation of coir dust.
Furthermore, the profiles of thermogravimetric analysis of AC and FAC are shown and explained in Fig. S3.† Three stages of weight loss have taken place in both activated carbon materials. The thermal stability of FAC is high up to 450 °C compared to that of AC.
Therefore, in the preliminary step, the optimization of the FAC dosage and contact time period of FAC sachet filters with water samples was carried out. According to Fig. 5(A) and (B), when the dosage of FAC was increased, the hardness and fluoride removal was increased. This can be explained by the availability of adsorption sites. When there is a higher amount of adsorbent, there are more adsorption sites for cations/anions. In this regard, the optimized dosage was 1000 mg for 250 ml of the water sample. The time optimization was carried out for both hardness and fluoride removal and 20–60 min showed better removal results. Here we are going to use FAC sachet filters; therefore it should have a minimum contact time. As longer contact times are not practical, we concluded that the optimized contact time is 20 min.
Natural water samples collected from different areas in Sri Lanka were used in adsorption experiments after testing for their initial pH, hardness and fluoride levels. In the first step, the hardness and fluoride removal efficiency in the water samples of the FAC loaded sachet filters was determined and compared with that of AC which was synthesized without acid activation. According to the results (Table 2), FAC demonstrates better hardness and fluoride removal efficiency than the AC material for all the samples tested. Therefore, it can be concluded that FAC has a higher adsorption capacity than AC, with the potential to be used in water softening applications and reduction of fluoride levels.
Selected sample location | Latitudes and longitudes | Initial pH (±0.01) | Final pH (±0.01) | Hardness removal (mg L−1) (±0.01) | Fluoride removal (mg L−1) (±0.001) | ||||
---|---|---|---|---|---|---|---|---|---|
Initial | Final (AC treated) | Final (FAC treated) | Initial | Final (AC treated) | Final (FAC treated) | ||||
Ipalogama | 8°06′03.4′′N 80°30′35.0′′E | 7.56 | 7.52 | 413.50 | 305.34 | 170.80 | 0.598 | 0.442 | 0.305 |
Galkulama | 8°16′18.5′′N 80°29′54.8′′E | 7.88 | 7.93 | 446.22 | 367.43 | 209.96 | 2.430 | 1.921 | 1.584 |
Mahailluppallama | 8°06′10.5′′N 80°28′19.0′′E | 7.59 | 7.63 | 556.59 | 417.01 | 250.23 | 1.590 | 1.398 | 1.052 |
Kiribathwehera | 8°23′57.1′′N 80°23′59.6′′E | 7.69 | 7.60 | 327.87 | 285.32 | 142.23 | 0.808 | 0.665 | 0.419 |
Galpalama | 8°23′56.4′′N 80°24′21.6′′E | 6.91 | 6.84 | 485.13 | 401.50 | 313.82 | 1.710 | 1.691 | 1.512 |
Alapathwewa | 8°23′32.8′′N 80°48′14.0′′E | 7.24 | 7.21 | 338.82 | 302.87 | 251.69 | 1.010 | 0.981 | 0.530 |
Hadagaswewa | 8°29′42.1′′N 80°11′19.9′′E | 7.55 | 7.58 | 295.97 | 223.45 | 132.64 | 1.390 | 1.276 | 1.010 |
Thuruwila | 8°13′27.1′′N 80°26′21.0′′E | 8.39 | 8.32 | 579.33 | 442.21 | 258.62 | 1.312 | 1.213 | 0.650 |
Kahatagasdigiliya | 8°35′12.4′′N 80°41′10.4′′E | 8.27 | 8.24 | 253.83 | 242.40 | 136.56 | 1.480 | 1.282 | 0.821 |
Additionally, a statistical analysis was carried out using ANOVA to evaluate the performance of FAC in the present study and activated carbon synthesized by the phosphoric acid soaking method (liquid phase) in the previous study done by Hettiarachchi et al., 2016.54 All the experiments were conducted under similar conditions and detailed results obtained for hardness and fluoride removal using activated carbon synthesized by the phosphoric acid soaking method are shown in Table S4.† The data reveal a statistically significant correlation between the two types of activated carbon materials for hardness removal (p < 0.05) and a significant difference for fluoride removal (p < 0.05). This indicates that the adsorption behavior of both types of activated carbon materials is similar irrespective of the synthesis method, in terms of hardness removal. However, there is a significant difference in fluoride removal. Overall, the phosphoric acid vapor functionalization method can be suggested as a potential economically suitable procedure which uses a lower amount of phosphoric acid, fewer washing steps, and easy maintenance of water pH and conductivity reaching the ASTM standards and being more towards a greener process.
When FAC sachets are used, after 20 minutes of contact time, a hardness removal percentage of 45–60% was achieved for several samples while a fluoride removal percentage of 45–50% was achieved except for certain higher initial fluoride levels (Table 2). These values are well within the WHO recommended levels of hardness and fluoride in drinking water.55 The use of handy sachet bags containing FAC would be useful for consumers in need of clean water as they are convenient and small in size. Rural areas lacking water treatment facilities may benefit from the use of these water treatment sachets containing FAC to remove excess hardness and fluoride levels in their drinking water. On the other hand, there were not any carbon particles leaching into the water samples confirming the suitability of FAC sachet filters in purification of drinking water (Fig. S4†). The percentage of acidic phosphate functional group concentration was 23.3% which was determined by the Boehm method. It was found that 0.001 ± 0.001 mg L−1 phosphate ions leached into water after 24 h at neutral pH values (pH 6.5–7). It was below the threshold value (0.1 mg L−1) of recommended phosphate ions in drinking water. Under acidic (pH = 3) conditions, phosphate leaching into water was 0.023 ± 0.001 mg L−1 compared to neutral pH values. At basic pH values (pH = 9), phosphate leaching into water was 0.012 ± 0.001 mg L−1. In this work, the recommended time period for FAC sachet filters is 20 min. During this 20 min, phosphate leaching was not observed at basic and neutral pH values while at highly acidic pH values, 0.001 ± 0.01 mg phosphate leaching was observed. In this regard, FAC proved its suitability to be used in drinking water purification at acidic, basic or neutral pH values. A pre-packed sachet technique using a controlled dose would facilitate maintaining an exact amount of adsorbent for treatment. An appropriate dose of adsorbent in multiple sachets could be used if necessary. Thus, FAC in sachet filters would be ideal in delivering the activated carbon material into water for the adsorption of cations and anions present in water.
According to the experimental results, the pHpzc value obtained for FAC is 5.9 ± 0.1 while that for AC is 7.4 ± 0.1 (Fig. 6). As the pH values of the tested water samples (Table 2) are higher than the pHpzc of phosphoric acid functionalized activated carbon, FAC possesses a net negative charge on its surface under the conditions adopted in the adsorption experiments. Thus, this favors the higher adsorption ability of FAC towards the cations by electrostatic interactions compared to AC. In addition, the presence of oxygen and phosphorus containing functional groups further facilitates the adsorption of cations into the large number of pores present in FAC.54
Fig. 6 Determination of the point of zero charge for activated carbon materials (I) FAC and (II) AC. |
However, the fluoride removal may become hindered in solutions with higher pH due to the formation of aqua complexes followed by the adsorbent surface becoming net negatively charged.58 Therefore, anions like fluoride will have electrostatic repulsive forces with the activated carbon surface. Furthermore, there would be a competition between hydroxide and fluoride ions for the adsorption onto the activated carbon surface. But some positively charged sites due to the presence of protonated groups may still be available at the pH values of the tested water samples.56 Therefore, the adsorption of fluoride under these conditions into FAC is likely related to the lesser extent of electrostatic interaction, hydrogen bonding and anion exchange. A similar explanation has been given by Araga et al., 2017 supporting such interactions.59 In addition, the highly porous nature of FAC facilitates the trapping and adsorption of fluoride ions into the pores present.
More interestingly, Souza et al., 2014 have shown and supported another hypothesis describing the synergistic effect of Ca2+ ions on adsorption of negatively charged ions onto the activated carbon surface.56 It is proposed that the interaction of Ca2+ ions with acid groups available on the activated carbon (especially carboxyl and phenol groups) generates an excess of positive charges on the surface of the adsorbent, decreasing the overall negative charge of carbon particles and, therefore, creating favorable sites for adsorption of the negatively charged species, even when the net surface charge is negative. A similar model could be applied to fluoride adsorption in natural water samples analyzed in this study where the solution pH exceeds the pHpzc of FAC. However, further experimental data would be necessary to confirm such mechanisms.
After calcium, magnesium and fluoride adsorption, the FTIR bands of the functional groups in FAC have been shifted, especially for CC vibrations, PO groups from phosphates or polyphosphates, O–C stretching vibration in the P–O–C (aromatic) linkage and POOH (Fig. 7(B)). Probably, these functional groups provide the active sites for adsorption of ions. The shifts in the functional group bands suggest the successful adsorption of adsorbents on the active sites of FAC.
Fig. 8(A) depicts the PXRD analysis of FAC after adsorption experiments. The 2θ values of the two major diffraction peaks of activated carbon have decreased after calcium, magnesium and fluoride ion adsorption. The degree of graphitization has also increased slightly, probably due to structural changes in activated carbon after the adsorption process and this phenomenon is consistent with the Raman spectral analysis of the carbon materials.
According to XPS analysis, calcium, magnesium and fluoride ion adsorption has led to the emergence of new peaks related to Ca 2p, Mg 1s and F 1s, respectively, along with shifts in binding energies of existing peaks relevant to C 1s, O 1s and P 2p in FAC (Fig. 8(B), Tables S5 and S6†) indicating the adsorption behavior of activated carbon. The spectrum of FAC-Ca exhibits two peaks at 351.3 eV and 347.65 eV assigned to Ca 2p1/2 and Ca 2p3/2, respectively, indicating the presence of calcium ions on the activated carbon surface. The only peak observed at 1306.1 eV in FAC-Mg is attributed to Mg 1s which indicates the presence of magnesium on activated carbon after the adsorption process. The spectrum of FAC-F depicts two peaks related to F 1s at 685.7 eV and 686.65 eV suggesting the fluoride adsorption onto activated carbon. Binding energy shifts in the regions of C 1s, O 1s and P 2p in FAC are due to the alterations in bonding and changes in electron density of atoms in the surface functional groups after the adsorption process. A change in each region reveals the interaction of calcium, magnesium and fluoride ions with phosphorus containing functional groups in addition to carbon oxygen species.
Organism | Treatment | |
---|---|---|
AC | FAC | |
E. coli (ATCC 25922) | − | + |
S. typhi (ATCC 19430) | − | + |
S. flexneri (ATCC 9199) | − | + |
The FAC MIC values ranged from 2.5 to 10.0 mg mL−1 whereas the MBC values ranged from 10.0 to 40.0 mg mL−1 for the three test strains (Table 4). Importantly, the MBC/MIC ratio indicates the bactericidal activity of the analyzed antibacterial compound FAC. In the current study, FAC exerted a bactericidal effect against E. coli, S. typhi and S. flexneri since the MBC/MIC ratio was equal to or lower than 4.0 (MBC/MIC ≤ 4.0). Furthermore, how these water pathogens destroyed with the treatment of FAC was examined by SEM and the images are shown in Fig. S5.† This can be explained that the initial adhesion of the bacteria on the activated carbon surface is considered to be a physicochemical process described by colloid chemical theories. In the presence of different cations/anions in water the repulsive electrostatic free energy would be reduced, and depending on the extent of this reduction, the van der Waal's attractive interaction may exceed the repulsive electrostatic interaction, therefore favoring the bacterial adsorption.61
Organism | Treatment (mg mL−1) | ||
---|---|---|---|
FACFAC | |||
MIC | MBC | MBC/MIC | |
E. coli (ATCC 25922) | 5.0 | 20.0 | 4.0 |
S. typhi (ATCC 19430) | 2.5 | 10.0 | 4.0 |
S. flexneri (ATCC 9199) | 10.0 | 40.0 | 4.0 |
Fig. 9 Kinetic study curves of the PSO model for (A) calcium, (B) magnesium, and (C) fluoride and of the PFO model for (D) calcium, (E), magnesium, and (F) fluoride. |
Adsorption isotherms offer explanation about how the adsorbate interacts with the adsorbent and describe the nature of the adsorption process. They indicate the distribution of adsorbate molecules between the solid phase and the liquid phase when the adsorption process reaches the equilibrium state.64 In the present work, Langmuir and Freundlich isotherms were employed to investigate the adsorption behavior by evaluating the applicability of isotherm equations to the adsorption process. The fitting of the experimental data for the two isotherm models and related parameters are shown in Fig. 10 and Table S8.†
Fig. 10 Isotherm curves of the Langmuir model for (A) calcium, (B) magnesium, and (C) fluoride and of the Freundlich model for (D) calcium, (E) magnesium, and (F) fluoride. |
The results indicate that the adsorption process is more fitted to the Freundlich isotherm than the Langmuir isotherm, as it possesses higher R2 values of 0.9989, 0.9988 and 0.9980 for calcium, magnesium and fluoride, respectively. This implies multi-layer adsorption on the heterogeneous surface of the activated carbon. KF is a Freundlich constant that shows the adsorption capacity on heterogeneous sites with non-uniform distribution of the energy level, and the n value shows the intensity between the adsorbate and adsorbent. According to Hettiarachchi et al., the adsorption becomes favorable if n is between 1 and 10.54 The calculated values of n prove that the adsorption of all three ions on FAC is favorable as the magnitude lies between 1 and 10.
AC | Activated carbon without phosphoric acid vapor activation |
FAC | Phosphoric acid vapor activated carbon |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d2ew00623e |
This journal is © The Royal Society of Chemistry 2023 |