It has been suggested that moderate to high fluoride ion concentrations in the presence of calcium in groundwater can represent human health hazards resulting in possible lesions in tubular cells, leading to chronic kidney disease of unknown etiology (CKDu). This study focuses on fluoride removal in the presence of calcium using iron oxide nanoparticles (FeONs) as an effective adsorbent in single- and multi-solute batch-studies. Fluoride and calcium removal by FeONs in single-solute batch-studies occurred only by adsorption. Both species followed the Langmuir isotherm model and pseudo second-order kinetics, indicating monolayer adsorption and chemisorption, respectively. In multi-solute batch-studies, fluoride removal was attributed to adsorption and chemical precipitation forming CaF2 in almost similar magnitude. For calcium removal in multi-solute studies, adsorption, co-precipitation, and precipitation were the prominent mechanisms, with co-precipitation dominant. Thus, FeONs are effective in fluoride removal by adsorption in both the presence and absence of calcium. FeONs could also be used successfully year-round in tropical climates, where calcium levels vary substantially, to lessen the risk of CKDu.

People drink groundwater containing substantial levels of fluoride (Chandrajith et al. 2012; Weragoda & Kawakami 2017), exceeding that recommended by the World Health Organisation (1.5 mg/L) (WHO 2017). Fluoride and its interactions with some other ions have been implicated as trigger factors of the ‘chronic kidney disease of unknown etiology’ (CKDu) prevalent in Sri Lanka (Wasana et al. 2016; Dissanayake & Chandrajith 2017; Dharma-wardana 2018). Over-extraction of groundwater, particularly during dry spells in tropical climates is highly likely to result in elevated fluoride and calcium levels. High calcium concentrations instigate a profound synergistic effect with fluoride on CKDu (Wasana et al. 2016). Fluoride exerts direct and calcium dose-dependent cytotoxic effects on human renal proximal tubular epithelial cells, and extracellular calcium chelation markedly attenuates fluoride-induced cell–death (Zager & Iwata 1997). The oxidative stress instigated by fluoride on human kidney cells also induces apoptosis and promotes the production of reactive oxygen species (anion superoxide, hydrogen peroxide, peroxynitrite, hydroxyl radicals), resulting in decreased cellular antioxidant defence mechanisms against oxidative kidney cell damage (Barbier et al. 2010).

Since about 2010, many researchers have studied fluoride removal using adsorption (Bhatnagar et al. 2011; Jayarathna et al. 2015), precipitation (Jadhav et al. 2015), ion exchange (Robshaw et al. 2019), electrodialysis (Belkada et al. 2018), coagulation and microfiltration (Da Conceição et al. 2015), and membrane filtration (Camacho et al. 2013). Among these, defluoridation using nanomaterial-based adsorption has gained considerable attention because of its efficiency. In particular, iron oxide nanoparticles (FeONs) have been widely studied (Raul et al. 2012). FeONs often favour fluoride and calcium adsorption, but the effect of calcium on defluoridation using FeONs does not appear to have been reported in the literature. In this study, the defluoridation potential of FeONs in both calcium-free and -rich groundwater was investigated. The study mainly focused on the different removal mechanisms of fluoride and calcium, adsorption isotherms, and kinetic studies of fluoride and calcium, using single- and multi- solute batch experiments.

Synthesis and characterisation of FeONs

Nano zero-valent iron was synthesised following the modified method of Petala et al. (2013), using 0.18 M FeCl3 and 0.94 M NaBH4. FeONs were synthesised by purging compressed air (3.5 L/min) through the solution containing nano zero-valent iron particles. Finally, they were extracted by centrifuging, rinsing three times with ethanol and oven drying at 80 °C (Gui et al. 2012).

The morphology and elemental composition of the FeONs were analysed, before and after the adsorption experiments, using Environmental Scanning Electron Microscopy (ESEM, Carl Zeiss, EVO 18, Secondary Electron Microscope, Germany) coupled with Energy-Dispersive X-ray Spectroscopy (EDAX, Element EDS system, AMETEK Materials Analysis Division, USA). FeON phase identification was performed by X-ray Powder Diffraction (XRD-D8, ECO, Advance Bruker Diffractometer with filtered Cu Kα radiation, Germany). Fourier transform-infrared spectroscopy (FT-IR, ALPHA Bruker, Germany) was performed in adsorption mode at ambient temperature in the spectral range 500 to 4,000 cm−1 to identify the FeONs' functional groups.

Single-solute batch-studies for fluoride and calcium removal

Single-solute batch experiments were carried out to determine the FeON dosage at which maximum fluoride and calcium removal occurred (optimum FeON dosage). Dosage was varied from 0.04 to 7.20 g/L on the basis of preliminary studies. The operating parameters were – sample volume, 100 mL; contact time, 1 hour; initial fluoride and calcium concentrations, 5 and 300 mg/L, respectively; initial pH, 6.5; temperature, 28 °C; and, stirring speed, 150 rpm. The FeON dosage for maximum removal of fluoride and calcium was determined as 1.25 g/L. Subsequently, a series of batch experiments was conducted using the optimum FeON dosage, to determine the fluoride and calcium adsorption isotherm and kinetic behaviour. In these studies, the initial concentrations of fluoride and calcium were varied from 1 to 20 mg/L, and 100 to 500 mg/L, respectively. All other experimental conditions were similar to those used when determining the optimum FeON dosage. The theoretical amounts of fluoride and calcium adsorbed at equilibrium, (Qe)ad and the actual amounts removed, (Qe)exp, were calculated to determine the extent of chemical precipitation and co-precipitation.

Multi-solute batch-studies for removal of fluoride and calcium

Multi-solute batch experiments were carried out in the presence of both fluoride and calcium to determine the optimum FeON dosage for maximum removal of fluoride and calcium. The operating parameters used were similar to those used in the single-solute batch studies and the optimum FeON dosage was determined as 1.25 g/L.

The differing amounts of fluoride and calcium removed by co-precipitation, chemical precipitation and adsorption were estimated by varying their respective concentration combinations. In CKDu prevalent areas in Sri Lanka, higher concentrations of fluoride and calcium have been reported during the dry season (June to August) than in the wet season. The combinations of fluoride and calcium concentrations considered for the multi-solute studies were based on seasonal variations typically encountered in the dry season. Twenty groundwater samples were collected during the dry season and the fluoride (1.1 to 9.0 mg/L) and calcium (140.0 to 900.0 mg/L) concentrations determined. Based on prevailing conditions in the CKDu affected areas, the experimental fluoride and calcium concentrations were set in the ranges 0.9 to 9.2 and 136.0 to 980.0 mg/L, respectively, and 10 fluoride/calcium concentration combinations were prepared accordingly. The concentration ranges were determined to exceed the natural equivalents slightly.

The first fluoride/calcium combination (0.9 and 136 mg/L, respectively) was prepared without FeONs and kept under quiescent conditions for 24 hours at ambient temperature to allow CaF2 precipitation. After 24 hours, the mixture was filtered (WhatmanTM 1003-110 Grade 3, 6 μm) and the residual fluoride concentration in the filtrate determined. FeONs (1.25 g) were subsequently added to 100 mL of the filtrate, and the mixture stirred continuously (150 rpm) at ambient temperature. The solution's residual fluoride and calcium concentrations were determined after one-hour contact time. The mixture's pH was monitored continuously for 72 hours. It always exceeded 5, confirming that no CaF2 dissolution occurred. The same procedure was followed for all other fluoride/calcium combination solutions.

For different fluoride/calcium combinations without FeON, fluoride supersaturation levels were estimated and compared with the respective residual fluoride concentrations, to determine whether or not chemical precipitation had occurred, and estimate the quantities of CaF2 precipitated. The solubility product (Ksp) of CaF2 [Ksp = (Ca2+) (F)2], which is 3.5 × 10−11 (mol/L)2 at 25 °C (Nasr et al. 2014), was used to determine fluoride supersaturation levels for different calcium concentrations. XRD and ESEM-EDX analyses were performed to confirm CaF2 precipitation, and the quantities used to estimate the corresponding amounts of fluoride and calcium precipitated. The amounts of fluoride and calcium adsorbed onto FeONs were estimated assuming monolayer adsorption phenomena. As the amounts of fluoride and calcium both adsorbed and precipitated were known, the amounts co-precipitated were estimated from the total amounts of these species removed from solution.

Analysis of fluoride and calcium

Residual fluoride concentrations were analysed by ion chromatography (Metrohm, 930 Compact IC Flex, Switzerland) with a mobile phase of 3.2 mmol/L Na2CO3 + 1 mmol/L NaHCO3, and a flow rate of 0.7 m/L/min. Residual calcium concentrations were determined using the EDTA titrimetric method based on USEPA Method No. 130.2 (USEPA 1983). Deionised water (18.2 Ω/cm) from a deioniser (Smart Plus-N, Heal Force, China) was used in all experiments.

Adsorption isotherms and kinetic studies for fluoride and calcium removal

Single-solute batch-studies

The experimental data were used to study fluoride and calcium removal with adsorption isotherms and kinetic models. The amount of fluoride or calcium removed per unit mass of FeONs (Qe in mg/g) for a single solute was calculated using Equation (1):
formula
(1)
where Co and Ce are the initial and equilibrium concentrations of fluoride or calcium in solution (mg/L) respectively, V the volume of solution (L), and W the mass of the adsorbent (g).
The Langmuir adsorption isotherm model was used to study fluoride and calcium removal and describe the amount of adsorbate adsorbed per unit weight of adsorbent at equilibrium. The Langmuir model is based on the assumptions that the adsorbent surface is homogeneous with equal sorption sites, only monolayer adsorption occurs with no interaction between adjacent adsorbed ions, and adsorbate ions tend to either adsorb or desorb. The isotherm (Langmuir 1918) is represented by Equation (2):
formula
(2)
where Ce is the equilibrium concentration of adsorbate (mg/L), Qe the amount of adsorbate adsorbed per unit mass of adsorbent at equilibrium (mg/g), Qm the maximum monolayer coverage capacity (mg/g), and KL the Langmuir isotherm constant (L/mg). The dimensionless constant separation factor or equilibrium parameter, RL, as given in Equation (3) is used to express the essential characteristics of the Langmuir adsorption isotherm:
formula
(3)
where Co is the initial concentration (mg/L) and KL the Langmuir isotherm constant. RL describes the adsorption mechanisms – i.e., unfavourable (RL > 1), linear (RL = 1), favourable (0 < RL > 1), or irreversible (RL = 0) (Weber & Chakravorti 1974).
The Freundlich model was applied for fluoride and calcium removal to ascertain the adsorption characteristics. It is based on an assumption of heterogeneous surfaces where fluoride or calcium could bind by multi-layer formation, with interaction among adsorbed ions. The isotherm (Freundlich 1906) is represented by Equation (4):
formula
(4)
where Kf is the Freundlich isotherm constant (mg/g), n the adsorption intensity, Ce the equilibrium concentration of the adsorbate (mg/L), and Qe the amount of adsorbate adsorbed per gram of adsorbent at equilibrium (mg/g). The Freundlich constant (n) indicates the adsorption mechanism; when 2 < n< 10, adsorption is favourable, between 1 < n< 2 it is moderately difficult and when n < 1 adsorption is poor (Kakavandi et al. 2013).
The experimental data for fluoride and calcium were fitted to the Brunauer-Emmett-Teller (BET) adsorption isotherm model, based on the postulation of adsorption in the form of multilayers but without interaction between adsorbed ions. The isotherm describes the adsorption of multiple ions at each site, energetically homogeneous adsorption sites with no adsorbate interactions between neighboring ions, and immobile adsorbed ions. The BET isotherm (Brunauer et al. 1938) is given by Equation (5):
formula
(5)
where Ce is the equilibrium concentration (mg/L), Cs the adsorbate monolayer saturation concentration (mg/L), Qo the maximum adsorption capacity for forming single layers (mg/g), and KB the BET adsorption isotherm which relates to the energy of surface interaction (L/mg).
The experimental data were fitted to the Dubinin-Radushkevich isotherm, which is generally applied to express the adsorption mechanism with a Gaussian energy distribution onto a heterogeneous surface, distinguishing between chemisorption and physical adsorption (physisorption) (Dubinin 1960) – see Equation (6):
formula
(6)
where Qe is the amount of adsorbate in the adsorbent at equilibrium (mg/g), Qd the theoretical isotherm saturation capacity (mg/g), BD the Dubinin-Radushkevich isotherm constant (mol2/kJ2), and ε the potential energy (estimated using Equation (7)). E is the Dubinin-Radushkevich isotherm constant (kJ/mol) and can be estimated using Equation (8). Physisorption is likely if E< 8 kJ/mol, chemisorption if 8 < E< 16 kJ/mol, and stronger chemical adsorption over ion exchange if E > 16 kJ/mol.
formula
(7)
formula
(8)
The fluoride and calcium data were also fitted to the Temkin adsorption isotherm (Tempkin & Pyzhev 1940), which contains a factor that considers adsorbent-adsorbate interactions explicitly. The isotherm is given by Equations (9) and (10):
formula
(9)
formula
(10)
where AT is the Temkin isotherm equilibrium binding constant (L/g), b the Temkin isotherm constant, R the universal gas constant (8.314 J/mol/K), T temperature (K), and B a constant related to the heat of sorption (J/mol).
The kinetics of fluoride and calcium adsorption were analysed using Lagergren's pseudo-first-order kinetic model (Lagergren 1898), which describes the adsorption of one ion of adsorbate onto one adsorption site, and the pseudo-second-order kinetic model, which describes the adsorption of one ion of adsorbate onto two adsorption sites (Ho & McKay 1999). These two models identify the fluoride and calcium adsorption process kinetics to be fitted to either chemisorption or physisorption. The pseudo-first order and pseudo-second-order kinetic models are expressed in Equations (11) and (12):
formula
(11)
formula
(12)
where Qe is the amount of adsorbate adsorbed on the adsorbent (mg/g) at equilibrium, Qt the amount of adsorbate adsorbed (mg/g) at time t (minutes), K1 the rate constant (min−1) for pseudo-first-order kinetics, and K2 the rate constant (gm/g/min) for pseudo-second-order kinetics.

Multi-solute batch-studies

The extension of the Langmuir equation proposed by Jain & Snoeyink (1973) was applied to determine the monolayer adsorption of binary mixtures consisting of fluoride and calcium by FeONs, with competitive adsorption of the binary mixtures. This development was based on the assumption that only a fraction of the active adsorption sites available for fluoride could also be occupied by calcium. The amounts of fluoride or calcium adsorbed on FeONs are estimated using Equations (13) and (14):
formula
(13)
formula
(14)
where Qm,1 is the adsorption capacity of adsorbate-1 (fluoride) at equilibrium under single-solute study (mg/g); and b1 and c1 the Langmuir isotherm constants for adsorbate-1 under single-solute study (L/mg), and the equilibrium concentration of adsorbate-1 obtained under single-solute study (mg/L), respectively. The symbols incorporating suffix 2 have the same meanings as those with suffix 1 but ascribed to adsorbate-2 (calcium).

Fluoride and calcium supersaturation are a prerequisite for CaF2 precipitation (Nath & Dutta 2010). Thus, fluoride removal depends on both the calcium concentration and the initial fluoride concentration, implying lower likelihood of precipitation for low initial fluoride concentrations.

The average FeON crystallite size was calculated from the XRD spectrum using the Scherrer Formula – Equation (15) (Chekli et al. 2016):
formula
(15)
where D is the average crystallite size, λ the wavelength of the X-ray radiation, β the full width of the half maximum diffraction peak, and ϴ the angle of diffraction.

Characterisation of FeONs

XRD analysis

Figure 1(a) shows the XRD spectrum of synthesised FeONs, exhibiting peaks at 30.4°, 35.8°, 43.7°, 57.6°, and 63.1°, which correspond to diffraction of the (220), (311), (400), (511), and (440) planes, respectively, of magnetite nanoparticles. Others have made similar observations for XRD characterisation (Lin et al. 2005; Xu et al. 2007; Wu et al. 2011; Raul et al. 2012; Mohseni-Bandpi et al. 2015; Shukla et al. 2015). The average crystallite size obtained for this study was 46.3 nm, confirming that a significant fraction of FeONs was within the nano-scale range. XRD analysis [Figure 1(b)] of CaF2 shows peaks at 28.2° (111), 46.9° (220), 55.7° (311), 68.6° (400) and 75.8° (331), confirming the formation of CaF2 precipitate (Pandurangappa et al. 2010; Tahvildari et al. 2012) in the binary mix, with the average particle size 96.3 nm.

Figure 1

XRD spectra of (a) synthesised FeONs and (b) CaF2 formed in the aqueous solution.

Figure 1

XRD spectra of (a) synthesised FeONs and (b) CaF2 formed in the aqueous solution.

Close modal

ESEM–EDX analysis

The ESEM image of synthesised FeONs (Figure 2(a)) shows the presence of interlinked flake-like aggregates, indicating agglomeration, possibly because of magnetic interaction between particles. The EDX analysis (Figure 2(b)) shows that the surface was composed of Fe and O with no impurities, and the narrow diffraction peaks confirm the crystalline structure. Similar observations on the agglomeration and crystalline structure formation of FeONs are reported by Lin et al. (2005) and Azzam et al. (2016). ESEM analysis of FeONs after fluoride adsorption in single-solute batch-studies shows spherical agglomerated particles with sharp edges (Figure 3(a)) and EDX analysis confirms the presence of fluoride adsorbed on FeON surfaces (Figure 3(b)). Figure 4(a) depicts the morphology of the FeON surface after multi-solute batch-studies, where both fluoride and calcium adsorbed and co-precipitated, which are visible as fluffy and irregular shapes. A magnified ESEM image (nm-scale) shows a nodular-like pattern on the FeON surface, which is ascribed to calcium co-precipitation with FeONs (Figure 4(b)). EDX analysis confirmed the presence of both fluoride and calcium on the FeON surface (Figure 4(c)).

Figure 2

(a) ESEM image of FeONs and (b) EDX spectrum of FeONs (Note: EDX was performed on the whole surface area in Figure 2(a)).

Figure 2

(a) ESEM image of FeONs and (b) EDX spectrum of FeONs (Note: EDX was performed on the whole surface area in Figure 2(a)).

Close modal
Figure 3

Single-solute studies with fluoride (a) ESEM image of FeONs, and (b) EDX spectrum of FeONs (Note: EDX was performed on the whole surface area in Figure 3(a)).

Figure 3

Single-solute studies with fluoride (a) ESEM image of FeONs, and (b) EDX spectrum of FeONs (Note: EDX was performed on the whole surface area in Figure 3(a)).

Close modal
Figure 4

Multi-solute studies with fluoride and calcium (a) ESEM image of FeONs, (b) ESEM image of FeON showing nodular-like patterns on the surface with Ca co-precipitation with FeONs (c) EDX spectrum of FeONs performed on the whole surface area in Figure 4(a).

Figure 4

Multi-solute studies with fluoride and calcium (a) ESEM image of FeONs, (b) ESEM image of FeON showing nodular-like patterns on the surface with Ca co-precipitation with FeONs (c) EDX spectrum of FeONs performed on the whole surface area in Figure 4(a).

Close modal

ESEM-EDX analysis of CaF2 (Figure 5(a) and 5(b)) shows that the CaF2 formed was agglomerated and porous, with polycrystalline nanoparticles. The larger particles exhibited different, spherical, surface perturbations (Pandurangappa et al. 2010).

Figure 5

Characterisation of CaF2 formed in the binary mixture (a) ESEM image (b) EDX spectrum of CaF2.

Figure 5

Characterisation of CaF2 formed in the binary mixture (a) ESEM image (b) EDX spectrum of CaF2.

Close modal

FT-IR analysis

The FT-IR spectrum of FeONs shows both the specific peaks reflecting the surface functional groups and the surface's complex nature. Figure 6 shows the FT-IR spectrum of FeONs (a) before adsorption, (b) after fluoride adsorption (single-solute studies), and (c) after fluoride and calcium adsorption (multi-solute studies). Before adsorption, there were broad bands around 1,033, 1,632, and 3,417 cm−1, attributed to N–H and O–H bending, and O–H stretching vibrations, respectively (Sharma & Jeevanandam 2013). N–H functional groups are indicated because of the use of NH4OH solution to keep the pH between 10 and 11 during FeON synthesis. The band at 1,434 cm−1, assigned to the symmetric and asymmetric bending of the C–H bond, may be ascribed to the ethanol added for FeON synthesis and storage (Qiu et al. 2011; Ayob & Abdullah 2012). The broad peak at 928 cm−1 arises from the bending vibrations of O–Fe–O groups (Jayarathna et al. 2015), and the narrow band around 575 cm−1 represents the Fe–O vibrations of magnetite nanoparticles (Xu et al. 2018).

Figure 6

FT-IR spectra of FeONs (a) before adsorption, (b) after fluoride adsorption (single-solute studies), and (c) after fluoride and calcium adsorption (multi-solute studies).

Figure 6

FT-IR spectra of FeONs (a) before adsorption, (b) after fluoride adsorption (single-solute studies), and (c) after fluoride and calcium adsorption (multi-solute studies).

Close modal

After fluoride adsorption, the functional groups on the FeONs' surface had changed, and the broad peak around 3,000 to 4,000 cm−1 had weakened and shifted. The hydrogen-bonding strength was also diminished, and hydroxyl groups became much freer from their initial hydrogen-bonding character (Jayarathna et al. 2015), so that the new, isolated, surface –OH peak appeared around 3,824 cm−1. The band at 1,632 cm−1 was shifted to 1,631 cm−1, which was attributed to the presence of –OH groups even after fluoride adsorption onto FeONs (Jayarathna et al. 2015; Mohseni–Bandpi et al. 2015). The intensity of the band at 1,383 cm−1 was increased sharply, perhaps representing the characteristic peak of fluoride adsorption (Patnaik et al. 2016; Liu et al. 2018). The results generally indicate that the surface hydroxyl groups play a significant role in fluoride adsorption.

After fluoride and calcium adsorption, the broad peak at 3,417 cm−1 was decreased and shifted, with new peaks appearing at 3,406 and 3,205 cm−1, attributed to the symmetric and asymmetric vibration of Fe–F and Fe–O–Ca–F bonds, respectively. A new peak appeared at 3,804 cm−1, and the peak at 1,632 cm−1 was shifted to 1,630 cm−1, indicating the presence of –OH groups after fluoride and calcium adsorption. In multi-solute adsorption, the sharp peak at 1,383 cm−1, attributed to Fe–F bonding, was decreased and shifted, and a broad peak appeared assigned to the asymmetric F–Ca–O–Fe bridge. This might arise because calcium occupies the active sites available for fluoride, showing a new peak around 500 cm−1, ascribed to Ca–O and Ca = O bonds (Galván-Ruiz et al. 2009) formed due to calcium adsorption and co-precipitation. The precipitation of CaF2 in the multi-solute batch experiments could not be verified with FT-IR, as CaF2's Ca–F stretching vibration occurs at 443 cm−1 (Tahvildari et al. 2012) – i.e., beyond the analytical range in this study.

Single-solute batch-studies

The possible mechanism of fluoride and calcium removal in single-solute batch-studies was identified as adsorption by FeONs. Therefore, attributes affecting fluoride and calcium adsorption by FeONs were studied.

Effect of FeON dosage, and initial fluoride and calcium concentrations on adsorption by FeONs

FeON dosage was varied from 0.04 to 7.20 g/L (Figure 7) to study its effect on fluoride and calcium adsorption. Fluoride adsorption showed little variation at FeON dosages exceeding 1.25 g/L, when it was 74%. Maximum calcium removal (39%) was observed at 1.25 g-FeONs/L with minor variation beyond that. For every FeON dosage, the total dissolved iron (Fe) concentration remaining after one-hour contact time with fluoride or calcium was measured by ion chromatography-mass spectrometry (ICP-MS) (Model: Agilent 7900 ICP-MS system, Japan, USA). The data showed that dissolution of Fe from the FeONs was less than 0.001 mg/L when dosage was increased, and that minute amounts of Fe dissolution did not affect FeON adsorption capacity. Increasing FeON dosage did, however, increase particle density per unit volume, increasing interaction between FeON particles and leading to agglomeration. This decreased the FeONs' adsorption capacity by decreasing the number of active sorption sites. Similar results have been reported for defluoridation with different adsorbent dosages of magnetite-chitosan composite (Mohseni-Bandpi et al. 2015).

Figure 7

Fluoride and calcium removal (%) vs FeON dosage in single-and multi- solute studies. The dosage was 1.25 g-FeON/L in all studies.

Figure 7

Fluoride and calcium removal (%) vs FeON dosage in single-and multi- solute studies. The dosage was 1.25 g-FeON/L in all studies.

Close modal

The effect of initial concentration on fluoride adsorption was studied for the range 1 to 20 mg-F/L, over which fluoride adsorption increased from 0.04 to 5.3 mg/g. Calcium adsorption increased from 0.04 to 3.34 mg/g, at constant FeON dosage, for the initial calcium concentration range 100 to 500 mg-Ca/L.

Adsorption isotherms and kinetic models for fluoride and calcium

The equilibrium isotherm and kinetic model parameters for fluoride and calcium adsorption by FeONs are presented in Table 1. The R2 values for the Langmuir isotherm adsorption model for both fluoride and calcium are 0.99, the high values indicating monolayer adsorption onto a surface with a finite number of active adsorption sites. The maximum monolayer adsorption capacities (Qm) of FeONs based on the Langmuir model were 28.98 mg-F/g and 1.13 mg-Ca/g. The equilibrium parameter (RL) was between 0.37 and 0.48 for fluoride, and 0.52 and 0.75 for calcium, indicating that the monolayer adsorption mechanism was favoured. The Freundlich constants (n) for fluoride and calcium were 1.35 and 1.79, respectively, indicating that F and Ca monolayer adsorption were moderately uncertain and weak, respectively. In single-solute batch experiments, monolayer adsorption was, therefore, preferred for both species. The R2 values for the BET adsorption isotherm for fluoride and calcium were 0.85 and 0.91, respectively, implying that the experimental data do not follow the BET isotherm assumption of multilayer formation during adsorption by FeONs. The equilibrium binding constant and adsorption energy variation for the Temkin model are given in Table 1. The Temkin model's R2 values for F and Ca – 0.87 and 0.91, respectively – show that the model does not fit the experimental data well. The Dubinin-Radushkevich model's R2 values for F and Ca were 0.93 and 0.75, respectively, showing that this model also failed to fit the experimental data well.

Table 1

Adsorption isotherm parameters for fluoride and calcium adsorption on FeONs

Isotherm modelParameter (units)Adsorbed ions
FluorideCalcium
Langmuir Qm (mg/g) 28.98 1.13 
KL (L/mg) 0.44 0.002 
R2 0.99 0.99 
Freundlich Kf (L/m) 0.78 0.63 
1/n 0.74 0.56 
R2 0.90 0.91 
BET Qo (mg/g) 0.05 0.001 
Kb (L/mg) 46.95 0.39 
R2 0.85 0.91 
Temkin AT (L/mg) 0.40 7.58E-06 
B (J/mol) 0.10 3.41 
R2 0.87 0.91 
Dubinin-Radushkevich Qe (mg/g) 0.33 0.26 
E (KJ/mol) 1.80 0.01 
BD (mol2/KJ) 0.15 3,682.30 
R2 0.93 0.75 
Isotherm modelParameter (units)Adsorbed ions
FluorideCalcium
Langmuir Qm (mg/g) 28.98 1.13 
KL (L/mg) 0.44 0.002 
R2 0.99 0.99 
Freundlich Kf (L/m) 0.78 0.63 
1/n 0.74 0.56 
R2 0.90 0.91 
BET Qo (mg/g) 0.05 0.001 
Kb (L/mg) 46.95 0.39 
R2 0.85 0.91 
Temkin AT (L/mg) 0.40 7.58E-06 
B (J/mol) 0.10 3.41 
R2 0.87 0.91 
Dubinin-Radushkevich Qe (mg/g) 0.33 0.26 
E (KJ/mol) 1.80 0.01 
BD (mol2/KJ) 0.15 3,682.30 
R2 0.93 0.75 

The adsorption kinetics of fluoride onto FeONs are described well by a pseudo-second-order model (R2 = 0.99 to 1; Table 2). Hence, the rate-limiting factor of fluoride and calcium adsorption by FeONs is chemisorption, confirming covalent bonding by sharing or exchange of electrons between adsorbate and adsorbent (Arshadi et al. 2014). The results show that the experimental Qe values (Qe(Exp)) derived from the mass balance equation (Equation (1)) are approximately equal to the Qe values calculated (Qe(Cal)) with the pseudo-second-order equation (Equation (12)). As the experimental and calculated Qe values are approximately equal, in single-solute batch-studies for both fluoride and calcium, adsorption was pronounced and removal due to co-precipitation could not be observed.

Table 2

Equilibrium constants and thermodynamic parameters for kinetic models in single-solute studies

Pseudo-first-order
Pseudo-second-order
FluorideQe(exp)Qe(Cal)K1Qe(Exp)Qe(Cal)K2
mg/Lmg/gmg/gmin1R2mg/gmg/gg/min/mgR2
0.04 0.23 0.220 0.08 0.05 0.05 0.29 0.99 
0.19 0.47 0.030 0.20 0.20 0.21 0.64 0.99 
0.36 0.69 0.030 0.01 0.37 0.37 0.54 1.00 
10 1.32 0.24 0.001 0.01 1.41 1.44 0.69 1.00 
15 2.24 0.35 0.001 0.13 2.45 2.45 0.41 1.00 
20 3.24 0.98 0.010 0.04 3.29 3.32 0.88 1.00 
CalciumQe(Exp)Qe(Cal)K1Qe(Exp)Qe(Cal)K2
mg/Lmg/gmg/gmin−1R2mg/gmg/gg/min/mgR2
200 0.48 0.25 −0.001 0.12 0.45 0.37 0.03 0.92 
300 1.14 0.33 −0.006 0.33 1.78 1.80 0.17 0.99 
400 2.38 0.05 −0.004 0.13 3.63 3.77 0.23 0.99 
500 4.02 0.06 −0.002 0.25 5.98 6.01 0.25 0.99 
Pseudo-first-order
Pseudo-second-order
FluorideQe(exp)Qe(Cal)K1Qe(Exp)Qe(Cal)K2
mg/Lmg/gmg/gmin1R2mg/gmg/gg/min/mgR2
0.04 0.23 0.220 0.08 0.05 0.05 0.29 0.99 
0.19 0.47 0.030 0.20 0.20 0.21 0.64 0.99 
0.36 0.69 0.030 0.01 0.37 0.37 0.54 1.00 
10 1.32 0.24 0.001 0.01 1.41 1.44 0.69 1.00 
15 2.24 0.35 0.001 0.13 2.45 2.45 0.41 1.00 
20 3.24 0.98 0.010 0.04 3.29 3.32 0.88 1.00 
CalciumQe(Exp)Qe(Cal)K1Qe(Exp)Qe(Cal)K2
mg/Lmg/gmg/gmin−1R2mg/gmg/gg/min/mgR2
200 0.48 0.25 −0.001 0.12 0.45 0.37 0.03 0.92 
300 1.14 0.33 −0.006 0.33 1.78 1.80 0.17 0.99 
400 2.38 0.05 −0.004 0.13 3.63 3.77 0.23 0.99 
500 4.02 0.06 −0.002 0.25 5.98 6.01 0.25 0.99 

Note: Qe(Exp) = experimental Qe derived using the mass balance equation (Equation (1)); Qe(Cal) = calculated Qe derived using the pseudo-second-order model.

Multi-solute batch-studies

The possible fluoride removal mechanisms in multi-solute batch-studies were identified as chemical precipitation as CaF2 and adsorption by FeONs. For calcium removal they were identified as co-precipitation with oxide and hydroxide functional groups of FeONs, chemical precipitation as CaF2 and adsorption by FeONs. Different physicochemical conditions in the mixture governed these mechanisms, so the potential for adsorption, co-precipitation and chemical precipitation was determined separately for different fluoride and calcium combinations.

Adsorption of fluoride and calcium by FeONs

The monolayer adsorption capacity of FeONs for fluoride and calcium removal was determined using constant 1.25 g-FeON/L dosage with different concentration combinations of fluoride and calcium (Figure 8). The fluoride and calcium concentrations were varied between 1.12 and 19.3 mg-F/L, and 136 and 980 mg-Ca/L, respectively. The monolayer adsorption capacity of fluoride for the corresponding combinations of fluoride and calcium, was between 0.07 and 1.63 mg for a 100 mL solution. For calcium it was between 0.29 and 0.49 mg (Figure 9). The results showed that for the combination containing 1.12 mg-F/L and 136 mg-Ca/L, total fluoride removal by FeONs occurred only through monolayer adsorption, with limited CaF2 precipitation due to the low fluoride concentration. For other fluoride and calcium combinations, total fluoride removal was attributed to the formation of CaF2 through chemical precipitation apart from some monolayer adsorption through chemisorption.

Figure 8

Total fluoride removed by chemical precipitation and adsorption for different combinations of F and Ca. Note: For each combination in multi-solute batch-studies, the amounts of F and Ca (both as mg) used with solution volume 100 mL were (a) 0.11, 13.60 (b) 0.31, 22.00 (c) 0.56, 33.60 (d) 0.71, 40.40 (e) 0.91, 57.60 (f) 1.13, 66.00 (g) 1.32, 74.80 (h) 1.52, 83.20 (i) 1.71, 92.80, and (j) 1.93, 98.00.

Figure 8

Total fluoride removed by chemical precipitation and adsorption for different combinations of F and Ca. Note: For each combination in multi-solute batch-studies, the amounts of F and Ca (both as mg) used with solution volume 100 mL were (a) 0.11, 13.60 (b) 0.31, 22.00 (c) 0.56, 33.60 (d) 0.71, 40.40 (e) 0.91, 57.60 (f) 1.13, 66.00 (g) 1.32, 74.80 (h) 1.52, 83.20 (i) 1.71, 92.80, and (j) 1.93, 98.00.

Close modal
Figure 9

Total calcium removed by co-precipitation, chemical precipitation and adsorption for different F and Ca combinations. Note: For each combination in multi-solute batch-studies, the amounts of F and Ca (both as mg) used with solution volume 100 mL were (a) 0.11, 13.60 (b) 0.31, 22.00 (c) 0.56, 33.60 (d) 0.71, 40.40 (e) 0.91, 57.60 (f) 1.13, 66.00 (g) 1.32, 74.80 (h) 1.52, 83.20 (i) 1.71, 92.80, and (j) 1.93, 98.00.

Figure 9

Total calcium removed by co-precipitation, chemical precipitation and adsorption for different F and Ca combinations. Note: For each combination in multi-solute batch-studies, the amounts of F and Ca (both as mg) used with solution volume 100 mL were (a) 0.11, 13.60 (b) 0.31, 22.00 (c) 0.56, 33.60 (d) 0.71, 40.40 (e) 0.91, 57.60 (f) 1.13, 66.00 (g) 1.32, 74.80 (h) 1.52, 83.20 (i) 1.71, 92.80, and (j) 1.93, 98.00.

Close modal

Chemical precipitation of CaF2

Calcium and fluoride tend to form insoluble CaF2 (Ksp = 3.45 × 10−11) via a stoichiometric reaction in the pH range 4.17 to 10 (Nasr et al. 2014). The solubility of CaF2 is independent of pH above 4.17 but increases significantly at lower pH levels.

Measurement of residual fluoride concentration temporal variations showed that much of the CaF2 formed within an hour of mixing provided that the residual fluoride and calcium concentrations exceeded supersaturation levels. Similar observations were made by Nasr et al. (2014). Fluoride removal by chemical precipitation was in the range 0.04 to 0.13 mg for 100 mL of solution with increasing F and Ca levels (Figure 8). Therefore, total fluoride removal due to adsorption and chemical precipitation was determined as between 0.11 and 1.76 mg from corresponding initial fluoride concentrations of between 0.11 and 1.93 mg. The calcium content removed by chemical precipitation and co-precipitation was between 0.14 and 0.05 mg, and 2.76 and 55.46 mg, respectively, when F and Ca concentrations increased (Figure 9). Total calcium removal due to adsorption, chemical precipitation and co-precipitation was, therefore, in the range 3.20 to 56.00 mg, with initial calcium amounts between 13.60 and 98.00 mg. The difference between adsorption and co-precipitation depends mainly on the geometry of the adsorbate surface (Corey 1981), where co-precipitation is considered a 3-dimensional process leading to higher removal capacity than adsorption alone. The formation of mixed metal hydroxides, which is leading to enhanced metal ion removal at high orders of magnitude has been attributed to co-precipitation over adsorption (Crawford et al. 1993).

Fluoride and calcium removal by FeONs in single-solute batch-studies was ascribed only to adsorption, and calcium co-precipitation was not observed. Adsorption studies suggested that both fluoride and calcium follow the Langmuir isotherm model and pseudo-second-order kinetics, indicating monolayer adsorption and chemisorption, respectively. The maximum monolayer adsorption capacities of FeONs were 28.98 and 1.13 mg/g for fluoride and calcium, respectively. In multi-solute batch-studies, fluoride removal was attributed to both adsorption and chemical precipitation as CaF2 in almost similar magnitudes. For calcium removal in multi-solute batch-studies, co-precipitation, precipitation, and adsorption were the removal mechanisms, with co-precipitation the predominant contributor.

The study confirmed that FeONs are effective in removing fluoride by adsorption, in the absence or presence of calcium. Irrespective of the calcium concentration in raw water, the use of FeONs is worthwhile to lessen the combined effect of fluoride and calcium on CKDu.

No potential conflicts of interest were reported by the authors.

The Senate Research Committee (SRC) Grant of the University of Moratuwa (Grant No. SRC/ND/15/01) and the National Research Council (NRC), Sri Lanka (Grant No. 15-056) supported this research study.

Arshadi
M.
,
Amiri
M. J.
&
Mousavi
S.
2014
Kinetic, equilibrium and thermodynamic investigations of Ni(II), Cd(II), Cu(II) and Co(II) adsorption on barley straw ash
.
Water Resources Industry
6
,
1
17
.
doi.org/10.1016/j.wri.2014.06.001
.
Ayob
A.
&
Abdullah
A. Z.
2012
Characterisation of polymer-stabilized nano zero-valent iron particle by ultrasonic irradiation-assisted method
.
Journal of Polymer Materials
29
,
167
179
.
Azzam
A. M.
,
El-Wakeel
S. T.
,
Mostafa
B. B.
&
El–Shahat
M. F.
2016
Removal of Pb, Cd, Cu and Ni from aqueous solution using nanoscale zero-valent iron particles
.
Journal of Environmental Chemical Engineering
4
(
2
),
2196
2206
.
doi.org/10.1016/j.jece.2016.03.048
.
Barbier
O.
,
Arreola-Mendoza
L.
&
Razo
D. L. M.
2010
Molecular mechanisms of fluoride toxicity
.
Chemico-Biological Interaction
188
(
2
),
319
333
.
doi:10.1016/j.cbi.2010.07.011
.
Belkada
F. D.
,
Kitous
O.
,
Drouiche
N.
,
Aoudj
S.
,
Bouchelaghem
O.
,
Abdi
N.
&
Mameri
N.
2018
Electrodialysis for fluoride and nitrate removal from synthesised photovoltaic industry wastewater
.
Separation and Purification Technology
204
,
108
115
.
doi.org/10.1016/j.seppur.2018.04.068
.
Bhatnagar
A.
,
Kumar
E.
&
Sillanpaa
M.
2011
Fluoride removal from water by adsorption – a review
.
Chemical Engineering Journal
171
(
3
),
811
840
.
doi:10.1016/j.cej.2011.05.028
.
Brunauer
S.
,
Emmett
P. H.
&
Teller
E.
1938
Adsorption of gases in multimolecular layers
.
Journal of the American Chemical Society
60
(
2
),
309
319
.
Camacho
L. M.
,
Dumée
L.
,
Zhang
J.
,
Li
J.
,
Duke
M.
,
Gomez
J.
&
Gray
S.
2013
Advances in membrane distillation for water desalination and purification applications
.
Water
5
(
1
),
94
196
.
doi:10.3390/w5010094
.
Chandrajith
R.
,
Padmasiri
J. P.
,
Dissanayake
C. B.
&
Prematilaka
K. M.
2012
Spatial distribution of fluoride in groundwater of Sri Lanka
.
Journal of the Natural Science Foundation of Sri Lanka
40
(
4
),
303
309
.
Chekli
L.
,
Bayatsarmadi
B.
,
Sekine
B. R.
,
Sarkar
B.
,
Shen
A. M.
,
Scheckel
K. G.
,
Skinner
W.
,
Naidu
R.
,
Shon
H. K.
,
Lombi
E.
&
Donner
E.
2016
Analytical characterisation of nanoscale zero-valent iron: a methodological review
.
Analytica Chimica Acta
903
,
13
35
.
doi.org/10.1016/j.aca.2015.10.040
.
Corey
R. B.
1981
Adsorption of Inorganics at Solid-Liquid Interfaces
(
Anderson
M. A.
&
Rubin
A. J.
eds).
Ann Arbor Science Publishers
,
Michigan, MI
.
Da Conceição
V. M.
,
Ugri
M. C. B. A.
,
Silveira
C.
,
Nishi
L.
,
Vieira
M. F.
,
Bassetti
F. J.
,
Vieira
A. M. S.
&
Bergamasco
B.
2015
Removal of excess fluoride from groundwater using natural coagulant Moringa oleifera lam and microfiltration
.
The Canadian Journal of Chemical Engineering
93
(
1
),
37
45
.
Dissanayake
C. B.
&
Chandrajith
R.
2017
Groundwater fluoride as a geochemical marker in the etiology of chronic kidney disease of unknown origin in Sri Lanka
.
Ceylon Journal of Science
46
(
2
),
3
12
.
doi.org/10.4038/cjs.v46i2.7425
.
Freundlich
H. M. F.
1906
Over the adsorption in solution
.
Journal of Physical Chemistry
57
,
385
471
.
Galván-Ruiz
M.
,
Hernandez
J.
,
Banos
L.
,
Noriega-Montes
J.
&
Rodríguez-García
M. E.
2009
Characterisation of calcium carbonate, calcium oxide, calcium hydroxide as starting point to the improvement of lime for their use in construction
.
Journal of Materials in Civil Engineering
DOI: 10.1061/(ASCE)0899-1561(2009)21:11(694)
.
Gui
M.
,
Smuleac
V.
,
Ormsbee
L. E.
,
Sedlak
D. L.
&
Bhattacharyya
D.
2012
Iron oxide nanoparticle synthesis in aqueous and membrane systems for oxidative degradation of trichloroethylene from water
.
Journal of Nanoparticle Research
14
(
5
),
861
.
DOI: 10.1007/s11051-012-0861-1
.
Ho
Y. S.
&
McKay
G.
1999
Pseudo-second order model for sorption processes
.
Process Biochemistry
34
(
5
),
451
465
.
Jadhav
S. V.
,
Bringas
E.
,
Yadav
G. D.
,
Rathod
V. K.
,
Ortiz
I.
&
Marathe
K. V.
2015
Treatment of fluoride concentrates from membrane unit using salt solutions
.
Journal of Environmental Management
162
(
1
),
306
325
.
Jain
S. J.
&
Snoeyink
V. L.
1973
Adsorption from bi-solute systems on active carbon
.
Water Pollution Control Federation
45
(
12
),
2463
2479
.
Jayarathna
L.
,
Bandara
A.
,
Ng
W. J.
&
Weerasooriya
R.
2015
Fluoride adsorption on γ − Fe2O3 nanoparticles
.
Journal of Environmental Health Science and Engineering.
13
,
54
.
DOI: 10.1186/s40201-015-0210-2
.
Kakavandi
B.
,
Jafari
A. J.
,
Kalantary
R. R.
,
Nasseri
S.
,
Ameri
A.
&
Esrafily
A.
2013
Synthesis and properties of Fe3O4-activated carbon magnetic nanoparticles for removal of aniline from aqueous solution: equilibrium, kinetic and thermodynamic studies
.
Iranian Journal of Environmental Health Sciences & Engineering
10
(
1
),
19
.
Lagergren
S.
1898
About the theory of so-called adsorption of soluble substances
.
Kungliga Svenska Vetenskapsakademiens Handlingar
24
,
1
39
.
Langmuir
I.
1918
The adsorption of gases on plane surfaces of glass, mica and platinum
.
Journal of the American Chemical Society
40
(
9
),
1361
1403
.
Lin
C. L.
,
Lee
C. F.
&
Chiu
W. Y.
2005
Preparation and properties of poly (acrylic acid) oligomer stabilised superparamagnetic ferrofluid
.
Journal of Colloid and Interface Science
291
(
2
),
411
420
.
Liu
J.
,
Zhu
R.
,
Xu
T.
,
Laipan
M.
,
Zhu
Y.
,
Zhou
Q.
,
Zhu
J.
&
He
H.
2018
Interaction of polyhydroxy fullerenes with ferrihydrite: adsorption and aggregation
.
Journal of Environmental Sciences
64
,
1
9
.
DOI: 10.1016/j.jes.2017.06.016
.
Mohseni-Bandpi
A.
,
Kakavandi
B.
,
Kalantary
R. R.
,
Azari
A.
&
Keramati
A.
2015
Development of a novel magnetite-chitosan composite for the removal of fluoride from drinking water: adsorption modelling and optimisation
.
RSC Advances
5
(
89
),
73279
73289
.
DOI: 10.1039/C5RA11294 J
.
Nasr
A. B.
,
Walha
K.
,
Puel
F.
,
Mangin
D.
,
Amar
R. B.
&
Charcosset
C.
2014
Precipitation and adsorption during fluoride removal from water by calcite in the presence of acetic acid
.
Desalination and Water Treatment
52
(
10–12
),
2231
2240
.
DOI: 10.1080/19443994.2013.799441
.
Nath
S. K.
&
Dutta
R. K.
2010
Fluoride removal from water using crushed limestone
.
Indian Journal of Chemical Technology
17
(
2
),
120
125
.
Pandurangappa
C.
,
Lakshminarasappa
B. N.
&
Nagabhushana
B. M.
2010
Synthesis and characterisation of CaF2 nanocrystals
.
Journal of Alloys and Compounds
489
(
2
),
592
595
.
Patnaik
S.
,
Mishra
P. C.
,
Nayak
R. N.
&
Giri
A. K.
2016
Removal of fluoride from aqueous solution using chitosan-iron complex
.
Journal of Analytical Bioanalytical Techniques
7
,
326
.
DOI: 10.4172/2155-9872.1000326
.
Petala
E.
,
Dimos
K.
,
Douvalis
A.
,
Bakas
T.
,
Tucek
J.
,
Zboril
R.
&
Karakassides
M. A.
2013
Nanoscale zero-valent iron supported on mesoporous silica: characterization and reactivity for Cr(VI) removal from aqueous solution
.
Journal of Hazardous Materials
261
,
295
306
.
Raul
P.
,
Umlong
I. M.
&
Purkait
M. K.
2012
Removal of fluoride from water using iron oxide-hydroxide nanoparticles
.
Journal of Nanoscience and Nanotechnology
12
(
5
),
3922
3930
.
DOI: 10.1166/jnn.2012.5870
.
Robshaw
T.
,
Dawson
R.
,
Bonser
K.
&
Ogden
M. D.
2019
Towards the implementation of an ion-exchange system for recovery of fluoride commodity chemicals. Kinetic and dynamic studies
.
Chemical Engineering Journal
367
,
149
159
.
doi.org/10.1016/j.cej.2019.02.135
.
Sharma
G.
&
Jeevanandam
P.
2013
Synthesis of self-assembled prismatic iron oxide nanoparticles by a novel thermal decomposition route
.
RSC Advances
3
(
1
),
189
200
.
DOI: 10.1039/c2ra22004 k
.
Shukla
S.
,
Jadaun
A.
,
Arora
V.
,
Sinha
R. K.
,
Biyani
N.
&
Jain
V. K.
2015
In vitro toxicity assessment of chitosan oligosaccharide coated iron oxide nanoparticles
.
Toxicology Report
2
,
27
39
.
Tahvildari
K.
,
Pour
M. E.
,
Ghammamy
S.
&
Nabipour
H.
2012
CaF2 nanoparticles: synthesis and characterisation
.
International Journal of Nano Dimension
2
(
4
),
269
273
.
Tempkin
M. I.
&
Pyzhev
V.
1940
Kinetics of ammonia synthesis on promoted iron catalyst
.
Acta Physico-Chimica Sinica
12
,
327
356
.
USEPA
1983
Methods for chemical analysis of water and waste, Report EPA/600/4-79/020, Environmental Monitoring and Support Laboratory, Office of Research and Development, United State Environmental Protection Agency, Cincinnati, Ohio, Washington
.
Wasana
H. M. S.
,
Aluthpatabendi
D.
,
Kularatne
W. M. T. D.
,
Wijekoon
P.
,
Weerasooriya
R.
&
Bandara
J.
2016
Drinking water quality and chronic kidney disease of unknown etiology (CKDu): synergic effects of fluoride, cadmium and hardness of water
.
Environmental Geochemistry and Health
38
(
1
),
157
168
.
DOI 10.1007/s10653-015-9699-7
.
Weber
T. W.
&
Chakravorti
R. K.
1974
Pore and solid diffusion models for fixed bed adsorbers
.
AIChE J.
20
(
2
),
228
238
.
Weragoda
S. K.
&
Kawakami
T.
2017
Evaluation of groundwater quality in 14 districts in Sri Lanka: a collaboration research between Sri Lanka and Japan
. In:
Trends in Asian Water Environmental Science and Technology
(Kurisu, F., Ramanathan, A. L., Kazmi, A. A. & Kumar, M. eds).
Springer International Publishing, Cham, Switzerland and Capital Publishing Company
,
New Delhi
,
India
, pp.
151
155
.
WHO (World Health Organisation)
.
2017
Guidelines for Drinking Water Quality: 4th Edition Incorporating the 1st Addendum
.
Geneva
,
World Health Organization, CC BY-NC-SA 3.0 IGO
.
Xu
J.
,
Yang
H.
,
Fu
W.
,
Du
K.
,
Sui
Y.
,
Chen
J.
,
Zeng
Y.
,
Li
M.
&
Zou
G.
2007
Preparation and magnetic properties of magnetite nanoparticles by sol-gel method
.
Journal of Magnetism and Magnetic Materials
309
(
2
),
307
311
.
doi.org/10.1016/j.jmmm.2006.07.037
.
Xu
C.
,
Yang
W.
,
Liu
W.
,
Sun
H.
,
Jiao
C.
&
Lin
A.
2018
Performance and mechanism of Cr (VI) removal by zero-valent iron loaded onto expanded graphite
.
Journal of Environmental Sciences
67
,
14
22
.
doi.org/10.1016/j.jes.2017.11.003
.
Zager
R. A.
&
Iwata
M.
1997
Inorganic fluoride divergent effects on human proximal tubular cell viability
.
American Journal of Pathology
150
(
2
),
735
743
.
PMC1858266
.