Open access peer-reviewed chapter

Water Defluoridation Methods Applied in Rural Areas over the World

Written By

Enos Wamalwa Wambu, Franco Frau, Revocatus Machunda, Lilliane Pasape, Stephen S. Barasa and Giorgio Ghiglieri

Submitted: 23 December 2021 Reviewed: 29 April 2022 Published: 23 June 2022

DOI: 10.5772/intechopen.105102

From the Edited Volume

Fluoride

Edited by Enos Wamalwa Wambu, Grace J. Lagat and Ayabei Kiplagat

Chapter metrics overview

514 Chapter Downloads

View Full Metrics

Abstract

Overexposure to fluoride (F) through drinking water is the most widespread water problem in the world, but it has now exacerbated due to rapid population growth rates, adverse climatic changes, and increasing levels of water scarcity. Thus, despite the large amounts of data, which has accrued on mitigation methods of high F is still the primary impediment to drinking water programs among many developing nations. The current review chapter on F mitigation techniques applied world-over is aimed at providing a succinct overview of water defluoridation techniques and strategies being used to combat the impact of human F overexposure. It represents a starting point to understand the prospects of reducing the global F impact. It is anticipated that this work will lay a strong foundation for this and also inform strategies for safeguarding public health and the environment from F pollution.

Keywords

  • defluoridation technologies
  • drinking water
  • fluoride
  • fluorosis
  • literature review

1. Introduction

The beneficial and detrimental effects of fluoride (F) were established in the early 1940s. Low levels of drinking water F (< 0.1 mg/L) were linked to the occurrence of dental caries, whereas elevated levels of F in water were associated with incidences of dental fluorosis among the communities [1]. Then some countries began artificial fluoridation of drinking water to control teeth decay [2]. Soon the widespread use of F in drinking water and oral products to control teeth decay resulted in a drastic decline in incidences of dental caries with a concomitant rise in dental fluorosis among the communities [3]. The severity of fluorosis increases with increasing F concentrations greater than 1.5 mg/L in drinking water [4] and data over the last few decades indicate trends toward more fluorosis around the world [5].

The new surge in the prevalence of fluorosis around the world has been attributed to, among other the rise in water fluoridation programs [5]; indiscriminate use of fluoridated products [6]; inadequate F legislation in the affected countries [7]; lack of technology and capacity for sustainable F surveillance [2] and widespread water security problems [8]. The WHO’s “guidelines” of 1.5 mg/L as the allowable standards of drinking water F have also come under scrutiny in regard to its success in controlling the adverse F effects among the communities [9]. Severe fluorosis has been recorded among communities using household waters with F levels well within these guidelines [10]. So, there have been efforts to control the adverse F effects in the communities. The strategies that are employed include instituting community health risk assessments and management programs [11]; high water F surveillance [12]; prospecting for safer water [13]; development of water defluoridation strategies; F awareness creation and behavior change campaigns [14]; developing water policies and legislation for F mitigation [9].

It is apparent, however, that there is an urgent need to reconsider the current approaches. The desired approaches for effective control of F impact among the communities should not only be effective but also be affordable, holistic, and comprehensive. In the current work, a review of the previous strategies and methodologies that have been applied in water defluoridation is presented [15]. The chapter aims to provide an update on F mitigation technologies being deployed worldwide and it is expected that this will enhance scientific understanding of the available technologies and the prospect of reducing the global F impact.

Defluoridation of existing waters is the main option where alternate safe water in high F areas is not available. However, the available water defluoridation approaches differ in scale, efficacy, sustainability, affordability, and acceptability. Therefore, the security of supply is heterogeneous. In general, the available options for water defluoridation can be classified as chemical, membrane-based, physical, or adsorption-based.

Advertisement

2. Chemical methods for water defluoridation

Chemical methods of water defluoridation involve the addition of reactive species, which can react with and facilitate water F removal through phase separation steps [16]. Here, five methods of which two are based on precipitation, two on coagulation, and one on electrocoagulation processes will be explained in further detail.

2.1 Precipitation processes

They basically include lime-softening and contact precipitation. The former is a precipitation chemical method of water defluoridation used to remove calcium and magnesium ions from hard water and it is also used to reduce F levels in potable waters [17]. In the traditional “lime-softening” technique, lime (Ca(OH)2) reacts with soluble Ca(HCO3)2 to precipitate insoluble CaCO3 [18] according to Eq. (1) as follows:

CaHCO32aq+CaOH2aq2CaCO3s+2H2OlE1

In presence of water F, however, part of Ca(OH)2 precipitates and removes insoluble fluorite, CaF2, according to Eq. (2) as:

CaOH2aq+2FaqCaF2s+OHaqE2

The removal of F by lime-softening is enhanced in presence of dissolved magnesium salts, which precipitate as Mg(OH)2 according to Eq. (3) below [19].

Mg2+aq+2OHaqMgOH2sE3

This also helps to eliminate excess alkalinity in treated water [20]. Defluoridation by lime is achieved by surface precipitation of fluoride onto the Mg(OH)2 formed but the process is temperature dependent and the F adsorption onto Mg(OH)2 increases at high temperatures.

Lime softening is the most common water defluoridation method in developing countries [21]. Based on this technique, the Indian Institute of Science in Bangalore, developed a simple defluoridation technique, which uses magnesium oxide, lime, and sodium bisulphate [22]. Due to the presence of MgO, the pH of treated water has to be adjusted to desirable levels (6.5 to 8.5) by adding 0.15 to 0.2 g per liter of sodium bisulphate. Even so, water defluoridation based on lime softening is inefficient and the technique requires large amounts of reagents leading to high volumes of F-laden sludge.

2.2 Contact precipitation

Contact precipitation employs the simultaneous addition of soluble calcium and phosphate compounds to brackish water. These react with F ions to precipitate CaF2 and fluorapatite (Ca5(PO4)3F2) [23] catalyzed by a saturated bone char medium. The process has been applied in Tanzania on raw water with 13 mg/L F resulting in F removal efficiency of 97.9% [24]. Contact precipitation has been floated as being more efficient and reliable than lime-softening, but it also generates large volumes of F-enriched sludge [25] and it can impart bad taste and smell to the treated water compromising its palatability.

2.3 Coagulation techniques

Coagulation is a procedure in which soluble metal cations [26] or commercial polyelectrolytes [27] with a large charge-to-volume ratio are added to water to attract and react with organics and other insoluble aggregates to flocculate and sediment and phase them out of the water. The flocculates formed to provide high sorbent surfaces for F ions. Alum (Al2(SO4)3.18H2O) is the usual flocculant in cases where F removal is also desired [26]. The salt reacts with OH ions to form Al(OH)3 flocs according to Eq. (4) as:

Al2SO4314H2O+3CaHCO322AlOH3+3CaSO4+14H2O+6CO2E4

The flocs sorb F from the water. A little lime is added controllably to replenish the OH ions. The amount of alum is controlled to prevent the initiation of complexation of Al3+ with F. Also, coagulation is not efficient and complete F removal is not achieved. Large amounts of coagulants are used leading to large volumes of sludge [25] and the residual coagulants in the water must be monitored to meet the drinking water standards.

Nonetheless, several authors have reported the application of this method with varying degrees of success [28, 29]. On the other hand, some authors have reported investigations aimed at improving upon the technique. Atia et al. [30], for example, compared the coagulants and demonstrated that the use of Al2(SO4)3.18H2O as flocculants was superior to Fe2(SO4)3.H2O. However, F removal by coagulation using poly-aluminum chloride (PAC) has also been reported [31] and compared with other polyelectrolyte coagulants [32]. Most recently some workers have utilized inorganic polymeric coagulant and indicated that 80% defluoridation of 6 mg/L F polluted water [33]. An alternate precipitation technique based on induced crystallization under extreme pH levels and carbonate/bicarbonate equilibriums has recently been applied with high F removal efficiencies [30]. Also, a facile approach to calcium co-precipitation has been reported [34].

2.4 Nalgonda technique

The Nalgonda technique is a modified coagulation protocol, which was developed by National Environmental Engineering Research Institute, Nagpur, India [35]. It takes advantage of the synergy between precipitation using lime and alum flocculation to facilitate F adsorption on Al(OH)3 flocs formed in the solution [36]. The F adsorbed is removed with the flocs by sedimentation. The Nalgonda technique has now been introduced in many countries but it requires high doses of flocculants leading to large sludge transfers and it may release excess Al3+ into the treated water [37].

2.5 Electrocoagulation

Electrocoagulation (EC) uses an electrolytic cell (Figure 1) [38] to supplies coagulant Al3+ ions in a controlled manner [26].

Figure 1.

Schematic representation of an electrocoagulation cell (adapted from Ref. [38]).

As the electric current passes through the cell, the Al anode gets oxidized to Al3+ ions, which are transformed into polymeric species and reacted with hydroxyl ions in solution to form the Al(OH)3 flocs, facilitate F removal from water as in the other coagulation techniques [39].

Even though the EC technique has been associated with demerits including high running costs, need for reliable electric power, and for specialized personnel to operate [25], the technique is efficient, reliable, and produces high-quality water. The sludge generation is low, and the method is used in different water conditions. In Algeria [40], for example, EC was applied as an efficient affordable water defluoridation technique, and, elsewhere, it has been used to reduce borehole F level from 3.5–4.8 to 0.8–1.0 mg/L [41]. Also, Khatibikamala et al. [42] reported that F concentration was reduced, based on an EC cell, from 4.0–6.0 mg/L in raw water to lower than 0.5 mg/L. Further, EC was applied to treat groundwater from Shivdaspura (Rajasthan): Sinha et al. [43], reported initial F levels of 5.0 mg/L were reduced to 0.2 mg/L plus. Then, Emamjomeh and Sivakumar [40] confirmed the technique as an effective electro protocol for domestic and industrial water defluoridation.

On their part, Vasudevan et al. [44] compared the performance of different electrodes in an EC protocol and found that F removal efficiency reached 96% with a magnesium alloy anode and a stainless steel cathode at a current density of 0.2 A/dm2 and pH of 7.0 more recently Khan et al. [45], confirmed earlier findings by Takdastan et al. [46] that aluminum electrodes were more efficient and economical in F removal than iron electrodes. In related analyzes, Hu et al. [47] showed that the efficacy of an EC system in water defluoridation was controlled by the molar ratio of hydroxide and F to Al(III) and related that optimum activity coefficients for defluoridation in coagulation and electrocoagulation are both close to 3.

Advertisement

3. Membrane methods

Membrane methods are those that employ the use of a casing that selectively separates a component in water. They include electrodialysis, reverse osmosis, and nanofiltration [48].

3.1 Electrodialysis

Electrodialysis (ED) technique uses an electric field to separate ions of one charge from the counter ions [49]. A typical ED unit (Figure 2) consists of about 400 alternating cation- and anion-exchange membranes, which are 0.5–2.0 mm wide, sandwiched between an anode and a cathode in a cell [25].

Figure 2.

Schematic representation of an electrodialysis unit.

The membranes have charged groups bound into polymeric substrates which attract and adsorb mobile counter ions. The anionic-exchange membranes permeate cations only, while the cation-exchange membranes permeate anions but trap the cations. Under an electric field, cations and anions move in opposite directions and the membranes capture respective ions resulting in alternating cells of ion-concentrated solutions called concentrates and ion-depleted solutions referred to as dilutes [50]. ED is an efficient defluoridation protocol and the sludge volume generated is low. However, the overall protocol is costly, complex, and requires reliable source power and specialized personnel to operate. The process is also non-selective and removes essential ions required for quality drinking water [51].

The technique has been applied to defluoridation of saline water with 3000 mg/L total dissolved salts (TDS) and 3.0 mg/L F [52]. Elazhar et al. [53] compared the performance of ED and nanofiltration (NF). Kabay et al. [54], on the other hand, was able to optimize a water defluoridation process and evaluated its mass transfer and energy use efficiency. ED was applied in Brazil with 97% defluoridation efficiency [49]. In India, ED was applied to saline water with high TDS of 5000 mg/L and 10 mg/L F levels [51] and it has been reported that ED was used to treat brackish water with 2.9 mg/L to just 0.4 mg/L [F] [55].

3.2 Reverse osmosis

Reverse osmosis (RO) is a membrane process in which dissolved pollutants are removed by applying pressure on raw water to force it through a semi-permeable membrane against the osmotic pressure (Figure 3) [56].

Figure 3.

Schematic representation of reverse osmosis.

The level and rate of contaminants removal depend on the sizes and electrical charge of the polluting ions [57]. It is found that RO is efficient and generates little sludge, but it is expensive to install and to run—it requires specialized personnel and reliable electric power to generate necessary pressures [48]. Some of these limitations of RO can, however, be bypassed. A study in Tanzania [58], for example, applied nanofiltration (NF) and reverse osmosis (RO), with an autonomous membrane system, which was powered by solar energy and the tested membranes could achieve the WHO drinking water standards [59]. The process reached 1000–2500 L daily total permeate volume of portable water with an additional 3500–5000 L of non-potable water fit irrigation and washing. However, the integration of such advanced technologies has not proven successful in many rural areas of developing nations where the necessary power is always available.

3.3 Nanofiltration

Nanofiltration (NF) operates on the same principle as reverse osmosis but the membranes have larger pores [48] offering less resistance to the flow of solvent and solute particles. The procedure is, therefore, able to operate at much lower pressures reducing energy costs. The retention of solutes is ascribed to steric and charge effects and the procedure is selective and considered to be suited to defluoridation of brackish waters. RO/NF has been applied to the treatment of various groundwater contaminants in India [60] and in the efficient removal of F and salinity from high-F brackish water at a village scale in Senegal [61]. In Finland, Kymenlaakso Water Limited, which is a public company, has operated a 6000 m3/day water RO plant with a permeate [F] of <0.03 mg/L since 2003 [62]. Richards et al. [63] evaluated the effect of speciation on the retention of F by NF and RO and found that F retention was independent of pH. In a study realized in Tunisia, F removal from water and from wastewater by NF was found to be controlled by trans-membrane pressure, feed water concentration, ionic strength, type of counter-ions, and pH and higher retentions were linked to pH values and vice versa [64]. Some researchers have reported F retention efficiencies of an NF process of 70% [65] mark above pH 7 but another team of researchers in France reported an RO F rejection efficiency greater than 98% [66].

Nonetheless, the protocol continues to attract the interest of researchers from around the globe [67]. Furthermore, hybrid treatments with sequential use of two or more simple techniques have become common in the recent past. For instance, filtration and ultrafiltration as subsequent treatment of coagulation have been recently tested for water defluoridation by a team of workers in India [68].

Advertisement

4. Distillation

4.1 Solar distillation

Distillation is a physical procedure in which water is converted to steam, and then the steam is condensed back into liquid water. The dissolved salts remain in the brackish water that is left behind. Solar distillation takes advantage of abundant sun rays and is the most applicable technique to circumvent the high costs of electricity. The method is simple, clean, and effective but the resulting wastewater must be removed to prevent encrustation of the vessels and be disposed of with care due to their high salinity. Solar distillation can be used from household scale to large industrial scale. Otherwise, it is inexpensive to run but the initial installation costs are big.

Solar distillation units have been used world-over to treat brackish water [69]. A pilot project of Solar driven membrane distillation has been operated on a small village level at Robanda in Tanzania [70]. Also, a similar solar water defluoridation unit has been built using local materials and successfully operated in Bongo District, Ghana, [71].

4.2 Membrane distillation

Membrane distillation uses a hydrophobic membrane with air-filled pores. The surface tension of the feed water and distillate prevents the water from entering the membrane pores keeping it out of the membrane. Water vapor pressure difference is then generated by applying sufficient temperature difference across the membrane. This is accomplished by heating the feed water and cooling the distillate at the other side of the membrane to cause a flow of water vapor through the membrane and result in distillate condensation [70].

Naidu et al. [71] evaluated the applicability of a modified design vacuum enhanced-multi-effect membrane distillation for drinking water and projected a 70% recovery ratio for a scaled-up unit. The feasibility of a direct contact membrane distillation (DCMD) process to recover F contaminated waters was also tested and up to 99% rejection of F was reported [72]. Boubakri et al. [73], using a similar DCMD process based on polyvinylidene F membrane, observed high thermal efficiency and high permeate flux favored by elevated temperatures.

Advertisement

5. Adsorption

In adsorption, raw water is passed through a bed containing a material that is able to retain F by physical, chemical, or ion exchange mechanisms. Adsorption of F onto solid adsorbents occurs through [74]: external mass transfer; surface adsorption; and, intra-particle diffusion processes. The technique has gained popularity in water defluoridation because it offers satisfactory results; it is simple, affordable, and eco-friendly. Many adsorbents, including alumina, clays, polymeric ion-exchange resin, activated carbons, biosorbents, and layered double hydroxides have been studied for water defluoridation [48].

5.1 Activated alumina

Activated alumina, Al2O3, is dehydrated Al(OH)3, which is prepared by heating Al(OH)3 at 300–600°C. It was first used in water defluoridation in the US in 1952, later in many other countries including China, Thailand, India [75], South Africa, and Ethiopia [76] by the 1980s. It is now widely used in many other countries of the world [77]. Alumina has one of the highest water defluoridation efficacies. In one study, aluminum hydroxide, which is a form of hydrated alumina was reported with an exceptional F adsorption capacity of 116.75 mg/g [78]. Recently hydrated alumina modified NaA zeolite was reported to have high F adsorption capacities of 104 mg/g [79]. The use of alumina in water defluoridation is, thus, widespread [39] but it is costly, requires frequent adsorbent regeneration, and the adsorbent gets fouled easily from the dissolved solids in the water [25].

5.2 Clay adsorbents

Soils and clays present high prospects of application in water defluoridation. This is mainly because they are almost always: (1) available in natural abundance; (2) stable and usable in different water conditions; (3) have high adsorption capacities; (4) easy to prepare; and, (5) are eco-friendly [80]. The specific reactions F adsorptions at the soil surfaces are heterogeneous and the particular choice of soil adsorbent for water defluoridation is controlled by its known adsorption capacities, availability, and the desired physicochemical properties. Consequently, soils are among the most studied matrices for water defluoridation.

Nonetheless, minerals, which have attracted the highest attention for water defluoridation research include: apatite, calcareous minerals, diatomite, attapulgites, and ferric minerals. The apatite minerals because are known to control the natural exchange of F in soil-water solutions in the environment [81]. Fan et al. [82] evaluated the capacity of hydroxyapatite, fluorspar, calcite, and quartz for water defluoridation. They found that F adsorption capacities for the minerals decreased from the apatite to quartz thus: hydroxyapatite > fluorspar > quartz activated using ferric ions > calcite > quartz. o, many workers have studied the capacity of apatite to enhance limestone, for example, and reported a maximum F adsorption capacity of 3.83 mg/g [83]. Else, it is often found that many calcareous minerals exhibit limited adsorption capacities for F [84]. In a study conducted by Kumar and Gupta [85], the authors investigated fluoride adsorption onto activated diatomite and found that the maximum defluoridation capacity of the mineral was 71.97 mg/kg. Other researchers have, however, reported a more enhanced defluoridation capacity of 51.1 mg/g for the [86]. These have also been collaborated most recently by Taabu et al. [87].

The adsorption of F onto modified attapulgite has been studied widely [88, 89]. The F adsorption capacity for the mineral approximates 24.55 mg/g. Hamdi and Srasra [90] found that water defluoridation capacities for some Tunisian soils was 55.8071.94 mg/g. However, other soils including ferrihydrite and kaolinite-ferrihydrite associate [91], ferric polymineral [92], lateritic minerals [93], clays [94], zeolites [95] and siliceous minerals [96] have been evaluated. Clearly, the capacity of clays to sorb F is greatly varied between the minerals and it is controlled mainly by their mineralogy and the operative conditions [80].

5.3 Ion-exchange resins

Defluoridation protocols based on the ion-exchange technique use charged anion resins that substitute anions in the substrate structure (normally chlorides) for F ions in the water [97]. The resin exchange sites are made of adsorbed cations (usually calcium) [98]. The natural polymeric organic resins, chitin/cellulose composites, are among the adsorbents with the greatest potential for water defluoridation. The use of natural polymeric materials has additional advantages because they are readily available in nature. Subsequently, natural polymers have been studied with varying adsorption efficiencies such as for chitosan (8.10 mg/g) [99], nanocellulose/polyvinyl alcohol composite, agglutinin derived from Strychnos potatorum L. seed (11.363 mg/g) [100] and chitosan-zirconia-ferrosoferric oxide composites (17.81 mg/g) [101]. Much higher adsorption capacities of 45.45 mg/g and 52.63 mg/g have, however, been reported for gamma degraded chitosan-Fe(III) beads [102] and for zirconia modified chitosan beads [103], respectively. Furthermore, a sorption potential of 48.78 mg/g has been reported for β-cyclodextrin grafted upon nanoscale titania surfaces [104].

The main challenge of the use of natural polymers in water purification is their liability to chemical and biological degradation. Also, the F ions tend to bind irreversibly into the exchange sites of the resins degrading the membranes. Then it is found that F removal using ion-exchange resins is often limited by low ionic selectivity [105]. Plus, commercial resins are expensive and require continuous regeneration and the spent adsorbents are non-biodegradable, they persist in the environment and must be disposed of very carefully.

5.4 Metal: organic frameworks

There is an emerging class of F adsorbents in fabricated metal–organic frameworks. This are basically multivalent metal ions intercalated in large organic molecules to enhance the synergic F uptake by the metal ions and the organic substrates. Jeyaseelan et al. [106] investigated defluoridation capacities of three MOF’s including fumaric acid–based metal–organic frameworks (MOFs) using Zr4+, La3+, and Fe3+ metal ions and found that they had comparable defluoridation capacities of 4.920, 4.925, and 4.845 mg/g, respectively.

5.5 Carbonaceous adsorbents

Carbonaceous F adsorbents include activated carbons from plants and animal biomasses, carbonaceous mineral adsorbents, and graphene.

5.5.1 Plant biomass-derived carbonaceous F adsorbents

Many activated carbons have been studied for water defluoridation. Hanumantharao et al. [107], for example, evaluated Acacia farnesiana carbon and reported a low defluoridation capacity of 0.268 mg/g. Similar limited water defluoridation capacities (< 1.5 mg/g) have also been reported for carbons of: Neem [108], Tamarindus indica fruit shells [109], family fruit [110], zirconium-impregnated coconut shell [111], rice straw [112], zirconium impregnated cashew nutshell [113], pine cone [114], and zirconium impregnated coconut fiber [115]. In contrast, studies by Mondal et al. [116] using sugarcane charcoal revealed F uptake capacities of 7.33 mg/g. Similarly, investigations using Pithacelobium dulce, Ipomoea batatas, and Peltophorum ferrugineum carbons showed defluoridation capacities of 78.96, 76.62, and 74.48, respectively [117] but much higher defluoridation capacities of 142.86 mg/g and 230.61 meq/g have been reported for Delonix regia pod carbon [118], certain carbon nanostructures [119] and for activated coffee husks carbons [55], respectively. In general water defluoridation using activated carbons has been shown to be pH-dependent and most carbon adsorbents have the highest F removal at acidic pH < 3 values [80]. Plus, the adsorbent particle appears to play a leading role in controlling the adsorption efficiency – high sorption occurs for the lowest size.

5.5.2 Borne char and activated animal charcoal

Bone char is the oldest known water defluoridation agent, and it was first used in USA from 1940s to the 1960s. The technique was later introduced into other countries and it is now among the most used methods in the developing countries of the world. The method involves the use of animal charcoal which is packed into columns and water percolated through the charcoal media [120]. Mutheki et al. [121] compared to the field and laboratory performance of bone char filters and other filters based on a combination of bone char and calcium-phosphate pellets. They found average uptake F capacities to be 1.2 ± 0.3 mg/g and 3.0 ± 1.0 mg/g, respectively. A study to explore activated carbon from fish bladder showed a maximum F removal of 1.43 mg/g [122]. In related work, Kawasaki et al. [123] who investigated four types of animal biomass and Singanan [124] who studied certain bone char adsorbents for F removal, reported more or less similar defluoridation potential. However, some authors reported 99% F removal from borehole water containing 11 mg/L F based on a cartridge bone char affixed onto a domestic faucet as a flow-through defluoridizer [125]. Therefore, some authors contend that F removal bone char is an efficient protocol for defluoridation of brackish water [22]. However, care must be taken about the preparation of the bone char, related to the taste and odor released into the treated water [126]. Besides, large amounts of organic materials are needed for gasification by expensive treatment at 800–1400 K in an inert atmosphere to obtain the adsorbents.

5.5.3 Graphene

Graphene is an emerging carbon material with sp2-hybridized single-carbon atom-layer structure [127] that is also a promising adsorbent for water defluoridation [128]. Some workers have studied amine grafted graphene oxide encapsulated chitosan hybrid beads for water defluoridation and found a defluoridation capacity of 4.65 mg/g [129]. However, Li et al. [127], while studying graphene samples obtained from exfoliating graphite materials, reported high F adsorption capacities of 17.65 mg/g. Also, a team investigating the F adsorption by graphene-aluminum-silver-carbon quantum dots reported an adsorption capacity of 12.04 mg/g [130]. Nonetheless, some workers have shown that enhancing graphene oxide using cupric oxide, improved its F uptake capacity to 34 mg/g [131]. The results collaborate with those of a team of researchers, which used aluminum modified graphene oxide (GO) and showed it to have superior F removal efficacies of 38.31 mg/g [132]. This shows the high potential use of graphene in F water remediation.

Besides activated carbons and graphene, the application of carbonaceous minerals to water defluoridation has been reviewed elsewhere [80]. However, Abe et al. [133] reported that the water defluoridation capacities of various carbons follow the order: bone char > coal charcoal > wood charcoal > carbon black > petroleum coke.

5.6 Biosorption

Biosorption utilizes animal and plant remains in the pulverized form “as is” without prior gasification or charring. A study realized in Tanzania, reported F removal efficiencies, which was 4.1–47.3%, for several biosorbents [134]. In a particular defluoridation study, which was conducted by Yadav et al. [135] using three agricultural-based biomasses as adsorbents tested on groundwater containing 5 mg/L F, the authors reported F removal efficacies of 40–58%. Elsewhere, Gandhi and Sekhar [136], found that the F biosorption capacity of Strychnos potatorum seed powder was 0.9945–1.052 mg/g. Further recently, the F removal of palm kernel shell-based adsorbent was evaluated and found to be 2.35 mg/g [137]. It is found, therefore, that, plant biomasses exhibit limited F uptake capacities when compared to other adsorbents unless they are formidably treated.

Also, defluoridation studies utilizing algal biosorbents derived from Spirogyra IO1 [138], Ulva fasciata sp. [139], polyalthia longifolia [140], and Spirogyra IO2 to adsorb F [141] have been reported. In the defluoridation evaluation of Spirogyra IO2, [141], for example, sorption capacity of 1.272 mg/g was reported. Also, the defluoridation capacities of fungal biosorbents have been studied. In evaluating Fusarium oxysporum to remove F from water, low defluoridation capacity of just 19% was reported [142]. These similar findings were collaborated with those by Ramanaiah et al. [143] fungal biosorbent of Pleurotus ostreatus 1804. However, other researchers have shown that the F biosorption potential of Saccharomyces sp. biomass reached 91% of F removal [144]. Similarly, Aspergillus and Calcium treated Aspergillus biosorbents revealed F adsorption capacities of 8.09 mg/g [145]. Differences in the capacity of various biomasses to sequestrate F from water are related to, among other factors, differences in active functional groups in the biomasses [134].

Thus, even though Mukkanti and Tembhurkar [146] have recently reported a high F adsorption capacity of 26.31 mg/g for an adsorbent developed from clamshell waste, it is apparent that the usual water defluoridation capacities of untreated biomasses are low when compared to the other adsorbents [90, 118]. Furthermore, the source biota for the prized F biosorbents may be non-existent in the regions where they are needed for easy defluoridation of water. Plus, untreated biomasses degrade easily under chemical and biological attacks.

5.7 Layer double hydroxides

Layered double hydroxides (LDHs), also called anionic clay and hydrotalcite-like compounds, are a “host-guest” layered materials, which have the general formula [M2+1-xM3+x(OH)2]x+(An−x/n).mH2O, where M2+ and M3+ are metal cations that occupy octahedral positions in hydroxide layers; x is the molar ratio M3+/(M2+ + M3+) and A denotes interlayer charge-compensating anions [147]. LDHs have attracted a lot of attention as F adsorbents in the recent past. Lu et al. [148] assessed F removal based on NiAl layered double hydroxides (NiAl-LDHs) and reported a low equilibrium F concentration of just 0.2388 mg/L in the treated water. Sadik et al. [149] also reported high F removal rates of 99.2% for calcined LDHs synthesized from seawater (LDHsw). New data have provided the maximum F adsorption capacity of 6.67 mg/g for Fe3O4/Al(OH)3 [150] and 12.63 mg/g for tri-metal Mg/Ce/Mn oxide-modified diatomaceous matrix [151]. However, the mechanism of F adsorption onto an LDH and calcined layered double hydroxide (CLDH) had been earlier evaluated with maximum defluoridation capacities of 1.3 mg/g and 20 mg/g, respectively [152]. The F sorption quantities were somehow similar to the 22.78 mg/g and 20.28 mg/g that have recently been reported for Ce-Ti and Ce-Ti/Fe3O4 hybrid oxides, respectively [153].

However, several studies have reported enhanced defluoridation capacities for LDHs. In a study involving calcined Mg–Al–CO3 LDHs, for example, competitive F adsorption was evaluated and water defluoridation capacity of 1.94 mmol/g was reported [154]. Then, Kang et al. [155] reported water defluoridation capacity of 50.91 mg/g for Mg/Fe l CLDHs. Furthermore, other studies have documented a high F adsorption capacity of 146.6 mg/g for Ca-Al LDHs [156] and 270.3 mg/g for Fe–Mg–La triple-metal hydroxide composite [157]. Consequently, other authors have focused on the optimization of defluoridation conditions for LDHs. Elhalil et al. [158], as such, while evaluating showed that optimum adsorbent dosages of water defluoridation using calcined Mg/Al LDH were in the range of 0.29–0.8 g/L. Also, the suggested F adsorption equilibrium time [149, 156] and solution acidity [149, 155] of LDHs are within 1 h and pH 6–7, respectively.

Advertisement

6. Phytoremediation

This is a technique of defluoridation and removal of other contaminants from the environment, which uses plants to absorb and accumulate excess F from soil and water through their roots into their systems. The plants are then removed at the predetermined time and disposed of safely. Several researchers evaluated technology phytoremediation with varying degrees of success. The tolerance capacity of Solanum tuberosum to accumulate F, has been tested and found that after 87 days, the F levels in the leaves, root, shoot, and potato tuber of the plants had increased to 3.96, 3.02, 2.8, and 1.56 mg, respectively [159]. Some researchers also tested the uptake of Al and F by four green algae species and found that Pseudokirchneriella subcapitata showed the highest aluminum and fluoride absorption under the test conditions [160]. Sirisha et al. [161] studied phytoremediation of Cr and F in industrial wastewater using the aquatic plant ipomoea aquatica plant. They found that the F removal rapidly reached 37% in just 10 min and similar results have been found by researchers in related tests [162] indicating that phytoremediation id a promising green technology for use in environmental fluoride remediation [163].

Advertisement

7. Conclusions

The current work, which was based on a systematic collection of literature data aimed at providing, in a concise and precise manner, an update on the techniques employed to combat the detrimental effects of human overexposure to F so as to focus the attention of stakeholders to the direction of science in the field of F mitigation in the world. From the foregoing discussions covered in this paper, the following conclusions are made:

  1. Water defluoridation techniques applied to the world-over can be classified broadly as chemical, physical, membrane-based, or adsorptive.

  2. The different defluoridation protocols differ in applicability and feasibility, depending on the desired levels of water defluoridation and resource availability.

  3. Adsorptive methods appear to present greater prospects in water defluoridatiore because they are simple, efficient, and cost-effective.

  4. LDH’s and soil adsorbents are the most studied adsorbents for water defluoridation, but the latter has a competitive edge over the former as they are readily available, easier to prepare and use, and more environmentally convenient.

  5. The efficacy of other adsorbents, such as activated carbons, is greatly controlled by the mode of adsorbent preparation and the adsorbent particle sizes.

Advertisement

Acknowledgments

This publication is part of the FLOWERED project that received funding from the EU’s Horizon 2020 research and innovation program under grant agreement No 690378. It is also in fond memories of Prof. Giorgio Ghiglieri, Department of Chemical and Geological Sciences, University of Cagliari, Sardinia, Italy, who was the Coordinator for the FLOWERED Project.

References

  1. 1. Dean HT, Elvove E. Some epidemiological aspects of chronic endemic dental fluorosis. American Journal of Public Health and the Nation's Health. 1936;26:567-575
  2. 2. Waugh DT, Potter W, Limeback H, Godfrey M. Risk assessment of fluoride intake from tea in the Republic of Ireland and its implications for public health and water fluoridation. International Journal of Environmental Research and Public Health. 2016;13:259-280
  3. 3. Levy SM. An update on fluorides and fluorosis. Journal of the Canadian Dental Association. 2003;69:286-291
  4. 4. Ayoob S, Gupta K. Fluoride in drinking water: A review on the status and stress effects. Critical Reviews in Environmental Science and Technology. 2006;36:433-487
  5. 5. Mascarenhas AK. Risk factors for dental fluorosis : A review of the recent literature. Pediatric Dentistry. 2000;269(22):269-277
  6. 6. Browne D, Whelton H, O'Mullane D. Fluoride metabolism and fluorosis. Journal of Dentistry. 2005;33:177-186
  7. 7. Li Y, Liang C, Slemenda CW, Ji R, Sun S, Cao J, et al. Effect of long-term exposure to fluoride in drinking water on risks of bone fractures. Journal of Bone and Mineral Research. 2001;16:932-939
  8. 8. UNICEF, WHO, Progress on sanitation and drinking water. 2010 Update, 2010. Available from: http://www.unwater.org/downloads/JMP_report_2010.pdf
  9. 9. Susheela A. Fluorosis in developing countries: Remedial measures and approaches. Proceedings-Indian National Science Academy Part B. 2002;68:389-400
  10. 10. Manji F, Baelum V, Fejerskov O. Dental fluorosis in an area of Kenya with 2 ppm fluoride in the drinking water. Journal of Dental Research. 1986;65:659-662
  11. 11. Erdal S, Buchanan SN. A quantitative look at fluorosis, fluoride exposure, and intake in children using a health risk assessment approach. Environmental Health Perspectives. 2005;113:111-117
  12. 12. Jacintha TGA, Rawat KS, Mishra A, Singh SK. Hydrogeochemical characterization of groundwater of peninsular Indian region using multivariate statistical techniques. Applied Water Science. 2017;7:3001-3013
  13. 13. Ghiglieri G, Balia R, Oggiano G, Pittalis D. Prospecting for safe (low fluoride) groundwater in the eastern African rift : The Arumeru District (northern Tanzania). Hydrology and Earth System Sciences. 2010;14:1081-1091
  14. 14. Gumbo FJ, Mkongo G. Water Defluoridation for rural fluoride affected communities in Tanzania. In: Spittle B, editor. 1 St Work. Fluorosis Defluoridation Held Ngurdoto, Tanzania, October 18-22, 1995. Dunedin, New Zealand: The International Society for Fluoride Research; 1995. pp. 109-114
  15. 15. Kanduti D, Sterbenk P, Artnik B. Fluoride: A review if use and effects on health. Mater Sociomed. 2016;28:133-137. DOI: 10.5455/msm.2016.28.133-137
  16. 16. Turner BD, Binning P, Stipp BDS. Fluoride removal by calcite: Evidence for fluorite precipitation and surface adsorption. Environmental Science & Technology. 2005;29:9561-9568
  17. 17. Hendrickson K, Vik EA. Adsorption in Water Treatment: Fluoride Removal. Oslo, Norway: Norwegian Institute for Water Research; 1984
  18. 18. Tebbutt THY. Principles of Water Quality Control. 5th ed. Amsterdam: Butterworht-Heinemann; 1998
  19. 19. Federal-Provincial Committee on Drinking Water. Guidelines for Canadian Drinking Water Quality: Guideline Technical Document-Fluoride. Ottawa: Canada; 2010
  20. 20. Opinya GN, Pameijer CH. Simple defluoridation procedures for Kenyan borehole water. Community Dentistry and Oral Epidemiology. 1987;15:60-62
  21. 21. Obasi Onuoha U. Evaluation of Alumina Sorptiom System for Removal of Fluoride from Water. Doctorate Dissertation, Texas Tech University; 1983
  22. 22. Renuka P, Pushpanjali K. Review on Defluoridation techniques of water. International Journal of Engineering Science. 2013;2:86-94
  23. 23. Korir H, Mueller K, Korir L, Kubai J, Wanja E, Wanjiku N, et al. The development of bone char—based filters for the removal of fluoride from drinking water. In: 34th WEDC International Conference, Addis Ababa, Ethiopia, 2009 Water, Sanitation and Hygiene: Sustainable Development and Multisectoral Approaches. Addis Ababa, Ethiopia; 2009. pp. 1-6
  24. 24. Dahi E. Contact precipitation for defluoridation of water. In: 22nd WEDC Conference Reaching the Unreached Challenges for 21st Century. New Delhi, India; 1996. pp. 262-265
  25. 25. Consultative Committee of the Kenya Bureau of Standards (KEBS), Excessive Fluoride in Water in. Kenya, Nairobi, Kenya; 2010
  26. 26. Hu CY, Lo SL, Kuan WH. Effects of the molar ratio of hydroxide and fluoride to Al(III) on fluoride removal by coagulation and electrocoagulation. Journal of Colloid and Interface Science. 2005;283:472-476
  27. 27. Bolto B, Gregory J. Organic polyelectrolytes in water treatment. Water Research. 2007;41:2301-2324
  28. 28. Aoudj S, Drouiche N, Hecini M, Ouslimane T, Palaouane B. Coagulation as a post-treatment method for the Defluoridation of photovoltaic cell manufacturing wastewater. Procedia Engineering. 2012;33:111-120
  29. 29. Stehouwer M. Defluoridation and Natural Organic Matter Removal in Drinking Waters by Alum Coagulation. Doctoral Dissertation, The University of Texas; 2014
  30. 30. Deng L, Wang Y, Zhou J, Huang T, Sun X. Impact of acid-base conditions on de fl uoridation by induced crystallization. Journal of Industrial and Engineering Chemistry. 2019;83:35-45
  31. 31. Ingallinella AM, Pacini VA, Fernández RG, Vidoni RM, Sanguinetti G. Simultaneous removal of arsenic and fluoride from groundwater by coagulation-adsorption with polyaluminum chloride. Journal of Environmental Science and Health. Part A, Toxic/Hazardous Substances & Environmental Engineering. 2011;46:1288-1296
  32. 32. Linzhi Z, Yihan SUN, Cheng H. Research on coagulation/sedimentation process for simulation of fluorine-containing wastewater treatment. Applied Mechanics and Materials. 2013;361-363:755-759
  33. 33. Solanki YS, Agarwal M, Maheshwari K, Gupta S. Removal of fluoride from water by using a coagulant (inorganic polymeric coagulant). Environmental Science and Pollution Research. 2020;1:3897-3905
  34. 34. Wang Z, Wang Z, Su J, Ali A, Zhang R, Yang W, et al. Synergistic removal of fluoride from groundwater by seed crystals and bacteria based on microbially induced calcium precipitation. Science of The Total Environment. 2021;806:1-9
  35. 35. Bulusu KR, Nawlakhe WG. Defluoridation of water with activated alumina: Batch operations. Indian Journal of Environmental Health. 1988;30:262-299
  36. 36. Nawlakhe WG, Paramasivam R. Defluoridation of potable water by Nagonda technique. Current Science. 1993;65:743-748
  37. 37. Agarwal KC, Gupta SK, Gupta AB. Development of nwe low-cost defluoridation technology. Water Science and Technology. 1999;40:167-173
  38. 38. Shankar R, Singh L, Mondal P, Chand S. Removal of lignin from wastewater through electro-coagulation, world. Journal of Environmental Engineering. 2013;1:16-20
  39. 39. Rao N. Fluoride and environment- a review. In: Bunch MJ, Suresh VM, Kumaran TV, editors. Proceedings of the Third International Conference Environmental Health Chennai, India, 15-17 December, 2003. York University, Chennai, India: Department of Geography, University of Madras and Faculty of Environmental Studies; 2003. pp. 386-399
  40. 40. Emamjomeh MM, Sivakumar M. Review of pollutants removed by electrocoagulation and electrocoagulation/flotation processes. Journal of Environmental Management. 2009;90:1663-1679
  41. 41. Emamjomeh MM, Sivakumar M, Schafer AI. Fluoride removal by using a batch electrocoagulation reactor. In: 7th Annual Environmental Engineering Research Event (EERE) Conference, Marysville, Victoria, Australia. 2003. P. 143-152
  42. 42. Khatibikamala V, Torabiana A, Janpoora F, Hoshyaripour G. Fluoride removal from industrial wastewater using electrocoagulation and its adsorption kinetics. Journal of Hazardous Materials. 2010;179:276-280
  43. 43. Sinha R, Khazanchi I, Mathur S. Fluoride removal by a continuous flow electrocoagulation reactor from groundwater of Shivdaspura. International Journal of Engineering Research and Applications. 2012;2:1336-1341
  44. 44. Vasudevan S, Lakshmi J, Sozhan G. Studies on a Mg-Al-Zn alloy as an anode for the removal of fluoride from drinking water in an electrocoagulation process. CLEAN – Soil, Air, Water. 2009;37:372-378
  45. 45. Khan SU, Asif M, Alam F, Khan NA, Farooqi IH. Optimizing fluoride removal and energy consumption in a batch reactor using electrocoagulation : A smart treatment technology. In: Ahmed S, Abbas S, Zia H, editors. Smart Cities - Opportumities and Challenges. Singapore: Springer; 2020. pp. 767-778
  46. 46. Takdastan A, Tabar SE, Islam A, Bazafkan MH, Naisi AK. The effect of the electrode in fluoride removal from drinking water by electro coagulation process. In: International Conference on Chemical and Biological Sciences, March 18-19th, 2015. 2015. pp. 39-44
  47. 47. Hu S, Yan L, Chan T, Jing C. Molecular insights into ternary surface complexation of arsenite and cadmium on TiO2. Environmental Science & Technology. 2015;49:5973-5979
  48. 48. Mohapatra M, Anand S, Mishra BK, Giles DE, Singh P. Review of fluoride removal from drinking water. Journal of Environmental Management. 2009;91:67-77
  49. 49. Zeni M, Riveros R, Melo K, Primieri R, Lorenzini S. Study on fluoride reduction in artesian well—Water from electrodialysis process. Desalination. 2005;185:241-244
  50. 50. Reuther CG. Saline solutions: The quest for fresh water. Environmental Health Perspectives. 2000;108:A78-A80
  51. 51. Adhikary SK, Tipnis UK, Harkare WP, Govindan KP. Defluoridation during desalination of brackish water by Electrodialysis. Desalination. 1989;71:301-312
  52. 52. Sahli MA, Annouar A, Tahaikt S, Mountadar M, Soufiane A, Elmidaoui A, et al. Fluoride removal for underground brackish water by adsorption on the natural chitosan and by electrodialysis. Desalination. 2007;212:37-45
  53. 53. Elazhar F, Tahaikt M, Zouahri A, Taky M, Hafsi M, Elmidaoui A. Defluoridation of Moroccan groundwater by Nanofiltration and Electrodialysis : Performances and cost comparison. World Applied Sciences Journal. 2013;22:844-850
  54. 54. Kabay N, Ara O, Samatya S, Yüksel Ü, Yüksel M. Separation of fluoride from aqueous solution by electrodialysis: Effect of process parameters and other ionic species. Journal of Hazardous Materials. 2008;153:107-113
  55. 55. Ali MBS, Hamrouni B, Dhahbi M. Electrodialytic Defluoridation of brackish water: Effect of process parameters and water characteristics. CLEAN – Soil, Air, Water. 2010;38:623-629
  56. 56. Rajeshwar K, Ibanez JG. Environmental Electrochemistry: Fundamentals and Applications in Pollution Sensors. Amsterdam: Elsevier science & Technology books; 1997
  57. 57. Shen J, Schäfer AI. Removal of fluoride and uranium by nanofiltration and reverse osmosis : A review. Chemosphere. 2014;117:679-691
  58. 58. Shen J, Mkongo G, Abbt-Braun G, Ceppi SL, Richards BS, Schäfer AI. Renewable energy powered membrane technology: Fluoride removal in a rural community in northern Tanzania. Separation and Purification Technology. 2015;149:349-361
  59. 59. WHO. Guidelines for Drinking-Water Quality, Third Edition, Incorporating First and Second Addenda. Geneva; 1984
  60. 60. Gedam VV, Patil JL, Kagne S, Sirsam RS, Labhasetwar P. Performance evaluation of polyamide reverse osmosis membrane for removal of contaminants in ground water collected from Chandrapur District. Journal of Membrane Science & Technology. 2012;2:117-122
  61. 61. Diawara CK, Diop SN, Diallo MA, Farcy M, Deratani A, Corporation P, et al. Performance of Nanofiltration (NF) and low pressure reverse osmosis (LPRO) membranes in the removal of fluorine and salinity from brackish drinking water. Journal of Water Resource and Protection. 2011;3:912-917
  62. 62. Sehn P. Fluoride removal with extra low energy reverse osmosis membranes: Three years of large scale field experience in Finland. Desalination. 2008;223:73-84
  63. 63. Richards LA, Richards BS, Rossiter HMA, Schäfer AI. Impact of speciation on fluoride, arsenic and magnesium retention by nanofiltration/reverse osmosis in remote australian communities. Desalination. 2009;248:177-183
  64. 64. Bejaoui I, Mnif A, Hamrouni B. Performance of reverse osmosis and Nanofiltration in the removal of fluoride from model water and metal packaging industrial effluent performance of reverse osmosis and Nanofiltration in the removal of fluoride from model water and metal packaging Industr, Saparation. Science and Technology. 2014;49:1-11
  65. 65. Richards LA, Vuachère M, Schäfer AI. Impact of pH on the removal of fluoride, nitrate and boron by nanofiltration/reverse osmosis. Desalination. 2010;261:331-337
  66. 66. Ndiaye PI, Moulln P, Dominguez L, Millet JC. Removal of fluoride from electronic industrial effluent by RO membrane separation. Desalination. 2005;173:25-32
  67. 67. Grzegorzek M, Majewska-nowak K, Ahmed AE. Science of the Total environment removal of fl uoride from multicomponent water solutions with the use of monovalent selective ion-exchange membranes. Science of The Total Environment. 2020;722:137681
  68. 68. Dubey S, Agarwal M, Gupta AB. Experimental evaluation of sand filtration and ultrafiltration as subsequent treatment of coagulation for fluoride removal. Environmental Progress & Sustainable Energy. 2021;e13790:1-11
  69. 69. Antwi E, Cudjoe E, Cudjoe J. Use of solar water distiller for treatment of fl uoride-contaminated water : The case of bongo district of Ghana. Desalination. 2011;278:333-336
  70. 70. van Tongeren WGJM, Assink WJ, Appelman WAJ. Pilot Test Solar Driven Membrane Distillation for Drinking Water Production at Robanda Tanzania NWASH Project Result 2: Innovative Fluoride Removal Technology. The Netherlands: Zeist; 2016
  71. 71. Naidu G, Jeong S, Choi Y, Jang E, Hwang T, Vigneswaran S. Application of vacuum membrane distillation for small scale drinking water production. Desalination. 2014;354:53-61
  72. 72. Yarlagadda S, Gude VG, Camacho LM, Pinappu S, Deng S. Water recovery from As, U, and F contaminated ground waters by direct contact membrane distillation process. Journal of Hazardous Materials. 2011;192:1388-1394
  73. 73. Boubakri A, Bouchrit R, Hafiane A. Fluoride removal from aqueous solution by direct contact membrane distillation: Theoretical and experimental studies. Environmental Science & Technology. 2014;21:10493-10501
  74. 74. Habuda-Stanić M, Ravančić ME, Flanagan A. A review on adsorption of fluoride from aqueous solution. Materials. 2014;7:6317-6366
  75. 75. Dahi E. The state of art of small community Defluoridation of drinking water. In: Dahi E, Rajchagool S, Osiriphan N, editors. 3rd International Workshop on Fluorosis Prevention and Defluoridation of Water, Chiang Mai, Thailand, November 20-24, 2000. Dunedin, New Zealand: International Society of Fluoride Research; 2000. pp. 141-170
  76. 76. Feenstra L, Vasak L, Griffioen J. Fluoride in Groundwater: Overview and Evaluation of Removal Methods. International Groundwater Resources Assessment Centre Utrecht, Utrecht, The Netherlands; 2007. Available from: www.igrac.nl
  77. 77. Venkobachar C, Iyengar L, Mudgal A. Household defluoridation of drinking water using activated alumina technology, Nazreth, Ethiopia November 19-25th, 1997. In: Dahi E, Nielsen JM, editors. Proceedings of 2nd International Work Fluorosis Defluoridation Water. Dunedin, New Zealand: International Society for Fluoride Research; 1997. pp. 138-145
  78. 78. Sequeira A, Solache-Ríos M, Balderas-Hernández P. Modification effects of hematite with Aluminum hydroxide on the removal of fluoride ions from water. Water, Air, & Soil Pollution. 2011;223:319-327
  79. 79. Naskar MK. Preparation of colloidal hydrated alumina modified NaA zeolite derived from rice husk ash for effective removal of fluoride ions from water medium. Journal of Asian Ceramic Societies. 2020;8:437-447
  80. 80. Wambu EWEW, Ambusso WOW, Onindo CO, Muthakia GKGK, Wambu EWEW. Review of fluoride removal from water by adsorption using soil adsorbents -an evaluation of the status. Journal of Water Reuse and Desalination. 2016;6:1-29
  81. 81. Willard RL, Campell TJ, Rapp RG. Encyclopaedia of Minerals. 2nd ed. New York: Van Nostrand Reinhold Company; 1990
  82. 82. Fan X, Parker DJ, Smith MD. Adsorption kinetics of fluoride on low cost materials. Water Research. 2003;37:4929-4937
  83. 83. Mohan R, Dutta RK. Continuous fixed - bed column assessment for defluoridation of water using HAp - coated - limestone. Journal of Environmental Chemical Engineering. 2020;8(4):7-8
  84. 84. Yapo NZS, Gouessé B, Briton H, Aw S, Reinert L, Drogui P, et al. Toxic/hazardous substances and environmental engineering bivalve shells (Corbula trigona) as a new adsorbent for the defluoridation of groundwater by. Journal of Environmental Science and Health, Part A. 2021;56:694-704
  85. 85. Kumar P, Gupta P. Defluoridation of domestic waste water by using activated diatomaceous earth in fixed mattress column adsorption system. Oriental Journal of Chemistry. 2021;37:594-601
  86. 86. Wambu EW, Onindo CO, Ambusso WJ, Muthakia GK. Fluoride adsorption onto acid-treated diatomaceous mineral from Kenya. Materials Sciences and Applications. 2011;2:1654-1660
  87. 87. Taabu SM, Wanjala NF, Bernard A, Godwin M, Alex O, Gloria M. Use of organic binders to enhance Defluoridation and pathogen removal efficiency of diatomaceous earth-based ceramic filters. Africa Journal of Physical Sciences. 2021;6:57-64
  88. 88. Zhang J, Xie S, Ho Y. Removal of fluoride ions from aqueous solution using modified attapulgite as adsorbent. Journal of Hazardous Materials. 2009;165:218-222
  89. 89. Zhang G, He Z, Xu W. A low-cost and high efficient zirconium- modified-Na-attapulgite adsorbent for fluoride removal from aqueous solutions. Chemical Engineering Journal. 2012;183:315-324
  90. 90. Hamdi N, Srasra E. Retention of fluoride from industrial acidic wastewater and NaF solution by three tunisian clayey soils. Fluoride. 2009;42:39-45
  91. 91. Wei S, Xiang W. Surface properties and adsorption characteristics for fluoride of kaolinite, ferrihydrite and kaolinite-ferrihydrite association. Journal of Food, Agriculture and Environment. 2012;10:923-929
  92. 92. Wambu EW, Onindo CO, Ambusso W, Gerald K. Equilibrium studies of fluoride adsorption onto a ferric poly − mineral from Kenya. Journal of Applied Sciences and Environmental Management. 2012;16:69-74
  93. 93. Wambu EW, Onindo CO, Ambusso WJ, Muthakia GK, Box PO. Fluoride adsorption onto an acid treated lateritic mineral from Kenya: Equilibrium studies, African. Journal of Environmental Science and Technology. 2012;6:160-169
  94. 94. Obijole O, Wilson G, Mudzielwana R, Ndungu P, Samie A, Babatunde A. Hydrothermally treated aluminosilicate clay (HTAC) for remediation of fluoride and pathogens from water : Adsorbent characterization and adsorption modelling. Water Resources and Industry. 2021;25:100144
  95. 95. Desalegn Y, Melak F, Yitbarek M, Astatkie H. Groundwater for sustainable development Aluminum coated natural zeolite for water defluoridation : A mechanistic insight. Groundwater for Sustainable Development. 2021;12:100525
  96. 96. Wambu EW, Onindo CO, Ambusso W, Muthakia GK. Removal of fluoride from aqueous solutions by adsorption using a siliceous mineral of a Kenyan origin. CLEAN – Soil, Air, Water. 2013;41:340-348
  97. 97. Alexandratos SD. Ion-exchange resins : A retrospective from industrial and engineering chemistry research. Industrial and Engineering Chemistry Research. 2009;48:388-298
  98. 98. Muraviev D, Torrado A, Valiente M. Kinetics of release of calcium and fluoride ions from ion-exchange resins in Artificail saliva. Solvent Extraction and Ion Exchange. 2007;18:345-374
  99. 99. Huang R, Yang B, Liu Q , Ding K. Removal of fluoride ions from aqueous solutions using protonated cross-linked chitosan particles. Journal of Fluorine Chemistry. 2012;141:29-34
  100. 100. Sowmiya BR, Vasanthy M, Rajakannan V, Ravindran B, Soon WC, Chandrasekaran M, et al. Defluoridation of water with a coagulant, Strychnos potatorum L. seed – Agglutinin. Environmental Technology and Innovation. 2021;24:101983
  101. 101. Korde S, Tandekar S, Jugade RM. Journal of environmental chemical engineering novel mesoporous chitosan-zirconia-ferrosoferric oxide as magnetic composite for de fl uoridation of water. Journal of Environmental Chemical Engineering. 2020;8:104360
  102. 102. Tandekar S, Saravanan D, Korde S, Jugade R. Materials today : Proceedings gamma degraded chitosan-Fe (III) beads for defluoridation of water. Materials Today: Proceedings. 2020;24:1-7
  103. 103. Tandekar S, Saravanan D, Jugade R. Zirconia-chitosan beads as highly efficient adsorbent for defluoridation of water. Indian Journal of Chemistry. 2020;59A:1067-1075
  104. 104. Fallah Z, Isfahani HN, Tajbakhsh M. Removal of fluoride ion from aqueous solutions by titania-grafted β -cyclodextrin nanocomposite. Environmental Science and Pollution Research. 2019;27(3):3281-3294
  105. 105. Karthikeyan S, See SW, Balasubramanian R. Simultaneous determination of inorganic anions and selected organic acids in airborne particulate matter by ion chromatography. Analytical Letters. 2007;40:793-804
  106. 106. Jeyaseelan A, Naushad M, Viswanathan N. Development of multivalent metal-ion-fabricated Fumaric acid-based development of multivalent metal-ion-fabricated Fumaric acid-based metal − organic frameworks for De fl uoridation of water. Journal of Chemical & Engineering Data. 2020;65:2990-3001
  107. 107. Hanumantharao Y, Kishore M, Ravindhranath K. Preparation and development of adsorbent carbon from acacia farnesiana for defluoridation. International Journal of Plant, Animal and Environmental Sciences. 2011;1:209-223
  108. 108. Chakrabarty S, Sarma HP. Defluoridation of contaminated drinking water using neem charcoal adsorbent : Kinetics and equilibrium studies. International Journal of ChemTech Research. 2012;4:511-516
  109. 109. Sivasankar V, Rajkumar S, Murugesh S, Darchen A. Tamarind (Tamarindus indica) fruit shell carbon: A calcium-rich promising adsorbent for fluoride removal from groundwater. Journal of Hazardous Materials. 2012;225-226:164-172
  110. 110. Medikondu K. Potable water defluoridation by lowcost adsorbents from Mimosideae family fruit carbons: A comparative study. International Letters of Chemistry, Physics and Astronomy. 2016;56:71-81
  111. 111. Sathish RSS, Raju NSRSR, Raju GSS, Rao GNN, Kumar KA, Janardhana C, et al. Equilibrium and kinetic studies for fluoride adsorption from water on zirconium impregnated coconut Shell carbon fluoride adsorption from water on. Separation Science and Technology. 2007;42:769-788
  112. 112. Daifullah AAM, Yakout SM, Elreefy SA. Adsorption of fluoride in aqueous solutions using KMnO4-modified activated carbon derived from steam pyrolysis of rice straw. Journal of Hazardous Materials. 2007;147:633-643
  113. 113. Alagumuthu G, Rajan M. Equilibrium and kinetics of adsorption of fluoride onto zirconium impregnated cashew nut shell carbon. Chemical Engineering Journal. 2010;158:451-457
  114. 114. Singh R, Surjit T, Katoch S, Modi A. Assessment of pine cone derived activated carbon as an adsorbent in defluoridation. SN Applied Sciences. 2020;2:1-12
  115. 115. Janardhana C, Rao GN, Sathish RS, Kumar PS, Kumar VA, Madhav MV. Study on defluoridation of drinking water using zirconium ion impregnated activated charcoals. Indian Journal of Chemical Technology. 2007;14:350-354
  116. 116. Mondal NK, Bhaumik R, Roy P, Das B, Datta JK. Investigation on fixed bed column performance of fluoride adsorption by sugarcane charcoal. Journal of Environmental Biology. 2013;34:1059-1064
  117. 117. Emmanuel KA, Ramaraju KA, Rambabu G, Rao AV, Autonomous SCRR, Pradesh A. Removal of fluoride from drinking water with activated carbons prepared from HNO3 activation - a comparative study. Rasayan Journal of Chemistry. 2008;11:802-818
  118. 118. Ajisha MAT, Rajagopal K. Fluoride removal study using pyrolyzed Delonix regia pod, an unconventional adsorbent. International journal of Environmental Science and Technology. 2015;12:223-236
  119. 119. Faghihian H, Atarodi H, Kooravand M. Synthesis, treatment, and application of a novel carbon nanostructure for removal of fluoride from aqueous solution. Desalination and Water Treatment. 2014;0:1-9
  120. 120. Mjengera H, Mkongo G. Appropriate defluoridation technology for use in fluoritic areas in Tanzania. Physics and Chemistry of the Earth. 2003;28:1097-1104
  121. 121. Mutheki PM, Osterwalder L, Kubai J, Korir L, Wanja E, Wambui E, et al. Comparative performance of bone char-based filters for the removal of fluoride from drinking water. In: 35th Water Engineerig and Development Centre (WEDC). International Conference. Loughbrgh, UK. 2011. pp. 1-4
  122. 122. Karuga J, Jande Y, Kim H, King C. Fish swim bladder-derived porous carbon for Defluoridation at potable water pH. Advances in Chemical Engineering and Science. 2016;6:500-514
  123. 123. Kawasaki N, Ogata F, Tominaga H, Yamaguchi I. Removal of fluoride ion by bone char produced from animal biomass. Journal of Oleo Science. 2009;535:529-535
  124. 124. Singanan M. Defluoridation of drinking water using metal embedded biocarbon technology. International Journal of Environmental Engineering; 2013;5(2):150-160
  125. 125. Nyanchaga N, Bailey T. Fluoride contamination in drinking water in the Rift Valley, Kenya and evaluation of the efficiency of a locally manufactured Defluoridation filter. Journal of Civil Engineering, JKUAT. 2003;8:79-88
  126. 126. Larsen MJ, Pearce EIF. Defluoridation of drinking water by boiling with Brushite and calcite, caries research;36:341-346. Caries Research. 2002;36:341-346
  127. 127. Li Y, Zhang P, Du Q , Peng X, Liu T, Wang Z, et al. Adsorption of fluoride from aqueous solution by graphene. Journal of Colloid and Interface Science. 2011;363:348-354
  128. 128. Kumar A, Garima VSK, Share R, Gupta S. Defluoridation studies using graphene oxidebased nanoadsorbents. In: Dehghani ÉH, Karri RR, editors. Green Chemistry and Water Remediation: Research and Applications. Elsevier. 2021. pp. 35-57
  129. 129. Jeyaseelan A, Salem N, Mohammedsaleh K, Katubi M, Naushad M, Viswanathan N. International journal of biological macromolecules design and synthesis of amine grafted graphene oxide encapsulated chitosan hybrid beads for de fl uoridation of water. International Journal of Biological Macromolecules. 2021;182:1843-1851
  130. 130. Singh N, Kumari S, Khan S. Improved fluoride removal efficiency using novel defluoridation pencil. Materials Today Communications. 2021;28:102521
  131. 131. Mohan S, Basavaiah K. Defluoridation in aqueous solution by a composite of reduced graphene oxide decorated with cuprous oxide via sonochemical. Arabian Journal of Chemistry. 2020;13:7970-7977
  132. 132. Rajput A, Raj SK, Sharma PP, Yadav V, Sarvaia H, Gupta H, et al. Synthesis and characterization of aluminium modified graphene oxide: An approach towards defluoridation of potable water. Journal of Dispersion Science and Technology. 2019;40:1101
  133. 133. Abe I, Iwasaki S, Tokimoto T, Kawasaki N, Nakamura T, Tanada S. Adsorption of fluoride ions onto carbonaceous materials. Journal of Colloid and Interface Science. 2004;275:35-39
  134. 134. Mwakabona HT, Machunda RL, Njau KN. The influence of stereochemistry of the active compounds on fluoride adsorption efficiency of the plant biomass. American Journal of Chemical Engineering. 2014;2:42-47
  135. 135. Yadav AK, Abbassi R, Gupta A, Dadashzadeh M. Removal of fluoride from aqueous solution and groundwater by wheat straw, sawdust and activated bagasse carbon of sugarcane. Ecological Engineering. 2013;52:211-218
  136. 136. Gandhi N, Sekhar KBC. Bioremediation of waste water by using Strychnos potatorum seeds (clearing nuts) as bio adsorbent and natural coagulant for removal of fluoride and chromium. Journal of International Academic Research for Multidisciplinary. 2014;2:253-272
  137. 137. Bashir M, Salmiaton A, Nourouzi M, Azni I, Harun R. Fluoride removal by chemical modification of palm kernel Shell-based adsorbent: A novel agricultural waste utilization approach. Asian Journal of Microbiology, Biotechnology & Environmental Sciences. 2015;17:533-542
  138. 138. Mohan SV, Ramanaiah SV, Rajkumar B, Sarma PN. Biosorption of fluoride from aqueous phase onto algal Spirogyra IO1 and evaluation of adsorption kinetics. Bioresource Technology. 2007;98:1006-1011
  139. 139. Kalyani G, Rao GB, Saradhi BV, Kumar YP. Biosorption isotherms of fluoride from aqueous solution on Ulva Fasciata SP.-a waste material. International Journal of Applied Environmental Sciences. 2009;4:173-182
  140. 140. Bharali RK, Bhattacharyya KG. Kinetic and thermodynamic studies on fluoride biosorption by devdaru (polyalthia longifolia) leaf powder, Octa. Journal of Environmental Research. 2014;2:22-31
  141. 141. Mohan SV, Ramanaiah SVV, Rajkumar B, Sarma PNN. Removal of fluoride from aqueous phase by biosorption onto algal biosorbent Spirogyra sp.-IO2: Sorption mechanism elucidation. Journal of Hazardous Materials. 2007;141:465-474
  142. 142. Merugu R, Rao K, Garimella S, Medi NK, Prashanthi Y. Factors affecting the defluoridation of water using Fusarium oxysporum bioadsorbent. International Journal of Environmental Biology. 2013;3:12-14
  143. 143. Ramanaiah SV, Venkata Mohan S, Sarma PN. Adsorptive removal of fluoride from aqueous phase using waste fungus (Pleurotus ostreatus 1804) biosorbent: Kinetics evaluation. Ecological Engineering. 2007;31:47-56
  144. 144. Chhipa H, Acharya R, Bhatnagar M, Bhatnagar A. Determination of sorption potential of fermentation industry waste for fluoride removal. International Journal of Bioassays. 2013;2:568-574
  145. 145. Mondal NK, Kundu M, Das K, Bhaumik R, Datta JK, Bengal W. Biosorption of fluoride from aqueous phase onto aspergillus and its calcium-impregnated biomass and evaluation of adsorption kinetics. Fluoride. 2013;46:239-245
  146. 146. Mukkanti VB, Tembhurkar AR. Taguchi’s experimental design for the optimization of the defluoridation process using a novel biosorbent developed from the clamshell waste. Journal of Dispersion Science and Technology. 2022;0:1-10
  147. 147. Yan H, Wei M, Ma J, Li F, Evans DG, Duan X. Theoretical study on the structural properties and relative stability of M (II)-Al layered double hydroxides based on a cluster model. The Journal of Physical Chemistry. A. 2009;113:6133-6141
  148. 148. Lu H, Li Q , Xiao H, Wang R, Xie D. Effect of non-thermal plasma modified NiAl layered double hydroxides on the removal of fluoride from aqueous solution. American Journal of Analytical Chemistry. 2014;5:547-558
  149. 149. Sadik N, Mountadar M, Sabbar E. Defluoridation by calcined layered double hydroxides synthesized from seawater. Journal of Materials and Environmental Science. 2015;6:2239-2246
  150. 150. Girma M, Zewge F, Chandravanshi BS. Fluoride removal from water using magnetic Iron oxide/aluminium hydroxide composite. SINET: Ethiopian Journal of Science. 2020;43:32-45
  151. 151. Gitari WM, Izuagie AA, Gumbo JR. Synthesis, characterization and batch assessment of groundwater fluoride removal capacity of trimetal Mg/Ce/Mn oxide-modified diatomaceous earth. Arabian Journal of Chemistry. 2020;13:1-16
  152. 152. Dessalegne M, Zewge F, Pfenninger N, Johnson CA, Diaz I. Layered double hydroxide and its calcined product for fluoride removal from groundwater of Ethiopian rift. Water, Air, & Soil Pollution. 2016;227:381-391
  153. 153. Dang-bao T, Lam H, Dang T. Defluoridation of water by Ce-Ti hybrid oxide nanoparticles. IOP Conference Series: Earth and Environmental Science. 2021;947:1-6
  154. 154. Cai P, Zheng H, Wang C, Ma H, Hu J, Pu Y, et al. Competitive adsorption characteristics of fluoride and phosphate on calcined Mg-Al-CO3 layered double hydroxides. Journal of Hazardous Materials. 2012;213-214:100-108
  155. 155. Kang D, Yu X, Tong S, Ge M, Zuo J, Cao C, et al. Performance and mechanism of Mg/Fe layered double hydroxides for fluoride and arsenate removal from aqueous solution. Chemical Engineering Journal. 2013;228:731-740
  156. 156. Sun Z, Park J, Kim D. Synthesis and adsorption properties of Ca-Al layered double hydroxides for the removal of aqueous fluoride. Water, Air, & Soil Pollution. 2017;228:23-30
  157. 157. Chen JP, Yu Y, Yu L, Chen JP. Adsorption of fluoride by Fe – Mg – La triple- metal composite: Adsorbent preparation, illustration of performance and study of … preparation, illustration of performance and study of mechanisms. Chemical Engineering Journal. 2015;262:839-846
  158. 158. Elhalil A, Qourzal S, Mahjoubi FZ, Elmoubarki R, Farnane M, Tounsadi H, et al. Defluoridation of groundwater by calcined Mg/Al layered double hydroxide. Emerging Contaminants. 2016;2:42-48
  159. 159. Das C, Dey U, Chakraborty D, Datta JK, Mondal NK. Fluoride toxicity effects in potato plant (solanum solanum tuberosum L.) grown in contaminated soils, Octa. Journal of Environmental Research. 2015;3:136-143
  160. 160. Pitre D, Boullemant A, Fortin C. Uptake and sorption of aluminium and fluoride by four green algal species. Chemistry Central Journal. 2014;8:2-8
  161. 161. Gandhi N, Sirisha D, Sekhar KBC. Phytoremediation of chromium and fluoride in industrial waste water by using aquatic plant Ipomoea aquatica. South pacific Journal of Pharma and Bio Science. 2013;1:1-4
  162. 162. Parikh PS, Mazumder SK. Capacity of Azolla pinnata var. imbricata to absorb heavy metals and fluorides from the wastewater of oil and petroleum refining industry at Vadodara. International Journal of Allied Practice, Research and Review. 2015;2:37-43
  163. 163. Braga AF, Borges AC, Vaz LRL, De Souza TD, Rosa AP. Phytoremediation of fluoride-contaminated water by Landoltia punctata. Engenharia Agrícola. 2021;41:171-180

Written By

Enos Wamalwa Wambu, Franco Frau, Revocatus Machunda, Lilliane Pasape, Stephen S. Barasa and Giorgio Ghiglieri

Submitted: 23 December 2021 Reviewed: 29 April 2022 Published: 23 June 2022