Defluoridation of drinking water using adsorption processes

Defluoridation of drinking water using adsorption processes

Accepted Manuscript Title: Defluoridation of drinking water using adsorption processes Authors: Paripurnanda Loganathan, Saravanamuthu Vigneswaran, Ja...

337KB Sizes 4 Downloads 315 Views

Accepted Manuscript Title: Defluoridation of drinking water using adsorption processes Authors: Paripurnanda Loganathan, Saravanamuthu Vigneswaran, Jaya Kandasamy, Ravi Naidu PII: DOI: Reference:

S0304-3894(12)01211-3 doi:10.1016/j.jhazmat.2012.12.043 HAZMAT 14806

To appear in:

Journal of Hazardous Materials

Received date: Revised date: Accepted date:

2-8-2012 18-12-2012 26-12-2012

Please cite this article as: P. Loganathan, S. Vigneswaran, J. Kandasamy, R. Naidu, Defluoridation of drinking water using adsorption processes, Journal of Hazardous Materials (2010), doi:10.1016/j.jhazmat.2012.12.043 This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

1

Research Highlights

2 3 4

! ! Comprehensive and critical literature review on various adsorbents used for defluoridation ! ! pH, temperature, kinetics and co-existing anions effects on F adsorption

6

! ! Choice of adsorbents for various circumstances

7

! ! Adsorption thermodynamics and mechanisms

8

! ! Future research on efficient, low cost adsorbents which are easily regenerated

us

cr

ip t

5

9

Ac ce

pt

ed

M

an

10

1

Page 1 of 62

10

Defluoridation of drinking water using adsorption processes

11

13

Paripurnanda Loganathan1, Saravanamuthu Vigneswaran1*, Jaya Kandasamy1, Ravi Naidu2 1

Faculty of Engineering and Information Technology, University of Technology, Sydney, NSW, 2007, Australia

14

16

2

CRC CARE, University of South Australia, Adelaide, SA 5095, Australia

*Corresponding author: ph: 612 9514 2641, E-mail: [email protected]

cr

15

ip t

12

us

17

ABSTRACT

19

Excessive intake of fluoride (F), mainly through drinking water, is a serious health hazard

20

affecting humans worldwide. There are several methods used for the defluoridation of

21

drinking water, of which adsorption processes are generally considered attractive because of

22

their effectiveness, convenience, ease of operation, simplicity of design, and for economic

23

and environmental reasons. In this paper, we present a comprehensive and a critical literature

24

review on various adsorbents used for defluoridation, their relative effectiveness, mechanisms

25

and thermodynamics of adsorption, and suggestions are made on choice of adsorbents for

26

various circumstances. Effects of pH, temperature, kinetics and co-existing anions on F

27

adsorption are also reviewed. Because the adsorption is very weak in extremely low or high

28

pHs, depending on the adsorbent, acids or alkalis are used to desorb F and regenerate the

29

adsorbents. However, adsorption capacity generally decreases with repeated use of the

30

regenerated adsorbent. Future research needs to explore highly efficient, low cost adsorbents

31

that can be easily regenerated for reuse over several cycles of operations without significant

32

loss of adsorptive capacity and which have good hydraulic conductivity to prevent filter

33

clogging during the fixed-bed treatment process.

Ac ce

pt

ed

M

an

18

2

Page 2 of 62

1. Introduction

36

2. Adsorption mechanisms

37

3. Factors influencing adsorption

38

3.1. pH

39

3.2. Co-existing anions

40

3.3. Temperature

41

3.4. Adsorption kinetics

42

cr

35

ip t

Keywords: adsorption, defluoridation, fluoride, fluoride toxicity, water treatment

us

34

4. Adsorbents 4.1. Metal oxides and hydroxides

44

4.2. Layered double hydroxides

45

4.3. Ion exchange resins and fibres

46

4.4. Zeolites

47

4.5. Carbon materials

48

4.6. Natural materials

49

4.7. Industrial by-products

pt

ed

M

an

43

5. Adsorption thermodynamics

51

6. Fluoride desorption and adsorbent regeneration

52

7. Conclusions

53

8. Acknowledgements

54

9. References

Ac ce

50

55 56 57

1. Introduction

58 3

Page 3 of 62

Fluoride (F) has beneficial effects on teeth at low concentrations in drinking water (0.4-

60

1.0 mg/L), especially for young children in that it promotes calcification of dental enamel and

61

protects teeth against tooth decay. Excessive levels of F on the other hand can cause a

62

number of problems ranging from mild dental fluorosis to crippling skeletal fluorosis as the

63

level and period of exposure to F increases [1]. The World Health Organisation [2] has

64

recommended a guideline value of 1.5 mg/L as the concentration above which dental

65

fluorosis is likely. However, it is also important to consider climatic conditions and quantity

66

of water intake and other factors such as F intake from certain diets when establishing F

67

limits. Water consumption in hot humid regions is generally higher than in temperate regions

68

and therefore the F concentration limit for likely fluorosis should be lower. For example, the

69

US Public Health Service [3] has recommended that the upper limit for F concentration in

70

drinking water should be decreased from 1.7 to 0.8 mg/L with increases of the average

71

maximum daily air temperature.

M

an

us

cr

ip t

59

High F intake has been suspected being involved in a range of adverse health problems in

73

addition to fluorosis, including cancer, impaired kidney function, digestive and nervous

74

disorders, reduced immunity, Alzheimer’s disease, nausea, adverse pregnancy outcomes,

75

respiratory problems, lesions of the endocrine glands, thyroid, liver and other organs [1,4-9].

76

However, there appears to be no convincing evidence for F being directly involved in causing

77

these conditions [1].

Ac ce

pt

ed

72

78

Elevated concentration of F in drinking water is due to its natural occurrence or industrial

79

activities. Many rocks and minerals in the earth’s crust contain F [10,11] which can be

80

leached out by natural weathering and rainwater, causing F contamination of surface and

81

ground waters. Besides this natural source, F also enters the water bodies from waste waters

82

produced from industries such as aluminium, steel, glass, semiconductors, electronic, tooth

83

paste, fertiliser and insecticide manufacturing plants [9,12-15]. 4

Page 4 of 62

Fluorosis due to excessive concentration of F has been reported in at least 28 countries

85

from South Asia, Africa, the Middle East, North, Central and South America, and Europe [1].

86

In India, it was estimated that 56.2 million people were affected by fluorosis and this problem

87

was prevalent in 17-19 out of the 32 States [16,17]. The major source of F in a majority of

88

countries is rocks and minerals such as fluorospar, cyyolite, and fluorapatite containing F

89

[17]. For example, Choi and Chen [18] reported extremely high F concentration (> 1000

90

mg/L) in surface water in areas with F-rich volcanic rocks.

cr

ip t

84

One method of reducing excessive concentrations of F in water is to blend water having a

92

high F concentration with water that has a low F concentration from an alternative source. If

93

such a source is not readily available, defluoridation is the only means remaining to prevent

94

fluorosis [1]. For contaminants other than F, water treatment methods are used to remove

95

contaminants to below the maximum level permissible but defluoridation is special in that the

96

treated water should have an optimum F concentration to derive the beneficial effects of F.

M

an

us

91

The main methods of defluoridation are precipitation/coagulation, adsorption, ion

98

exchange, reverse osmosis, and electrodialysis. Of these, precipitation/coagulation and

99

adsorption are convenient methods and are widely used, especially in developing countries’

100

rural areas. The scale and treatment site differ between industrialised countries and

101

developing countries. In industrialised countries the treatment of water is generally performed

102

at water treatment plants close to the water source but in developing countries it is carried out

103

at a village community level or at a household level [1] using simple inexpensive locally

104

available adsorptive media [9,19-21]. Industrialised countries generally use more efficient but

105

more costly adsorption media including synthetic ion exchange resins as well as advanced

106

techniques such as reverse osmosis and electrodialysis.

Ac ce

pt

ed

97

107

The advantages and shortcomings of the various methods of defluoridation are presented

108

in Fig. 1. Of these various methods that of adsorption is generally considered attractive 5

Page 5 of 62

because of its effectiveness, convenience, ease of operation, simplicity of design, and

110

economic and environmental considerations, provided low-cost adsorbents which can

111

effectively remove F around the neutral pH of drinking water are used. The

112

precipitation/coagulation method where lime and Al salts are used to remove F as a CaF2

113

precipitate followed by removal of left over F in solution by co-precipitation with and

114

adsorption on to the precipitated Al(OH)3 is further developed into a technique called

115

Nalgonda process in India [17,24,25]. This process is extensively used in India and Africa

116

[1]. However, the main drawback of this technique is the low effectiveness of F removal and

117

production of toxic AlF complexes in solution [8].

us

cr

ip t

109

Previous reviews on defluoridation of water presented timely information on several

119

adsorbents used for defluoridation, but they generally did not focus on the chemical

120

mechanisms and solution factors influencing the adsorption processes, and regeneration of

121

the adsorbents [1,17,26,27]. The objective of this paper is to compile and present current

122

information on the potential of the various adsorbents used for defluoridation of drinking

123

water, their relative effectiveness, mechanisms and thermodynamics of adsorption, factors

124

influencing adsorption and methods of adsorbent regeneration for reuse. Based on the review

125

adsorbents are selected for specific circumstances.

127 128

M

ed

pt

Ac ce

126

an

118

2. Adsorption mechanisms

129

The capacity, energy and kinetics of adsorption of F by adsorbents are controlled by the

130

mechanism of adsorption. Understanding these mechanisms can provide useful information

131

on the optimisation of the adsorption process in water treatment plants and the subsequent

132

desorption/adsorbent regeneration process for reuse. There are mainly five mechanisms of F

133

adsorption, namely: (1) van der Waals forces (outer-sphere surface complexation, (2) ion 6

Page 6 of 62

exchange (outer-sphere surface complexation), (3) hydrogen bonding (H-bonding) (inner-

135

sphere surface complexation), (4) ligand exchange (inner-sphere surface complexation), and

136

(5) chemical modification of the adsorbent surface. The first two mechanisms are governed

137

by weak physical adsorption and are non-specific to F, whereas the third and fourth

138

mechanisms are governed by strong chemical adsorption specific to F. The fifth mechanism

139

is governed by both specific and non-specific adsorption. In the presence of the other anions

140

in water, F cannot be easily removed by adsorbents using the first two mechanisms. Fluoride

141

adsorption resulting from the third and fourth mechanisms selectively removes F from water

142

in the presence of most types of anions; only anions (e.g. phosphate) which also adsorb

143

specifically on the adsorbent compete with F for adsorption sites.

an

us

cr

ip t

134

Van der Waals forces are weak short range forces acting between two atoms. The larger

145

the adsorbate size the greater the force of attraction. Therefore adsorbates with high

146

molecular weights such as dissolved organic matter are adsorbed on adsorbents having high

147

surface area through van der Waals forces. This is the reason for the weak adsorption of F

148

and strong adsorption of dissolved organic matter on activated carbon [28]. Fluoride was

149

considered to be adsorbed on manganese oxide-coated alumina by van der Waals forces at

150

high pH [29].

Ac ce

pt

ed

M

144

151

Ion exchange is a stoichiometric process where any counter ion leaving the ion exchanger

152

surface is replaced by an equivalent number of moles of another counter ion to maintain

153

electro-neutrality of the ion exchanger. The ions are adsorbed physically by fully retaining

154

their inner hydration shell and the adsorption is due to electrostatic or Coulombic attraction.

155

The ion exchange process is rapid and reversible. Fluoride removal by ion exchange resins

156

[30] and ion exchange fibres [31] is mainly governed by the ion exchange mechanism as

157

illustrated in Fig. 2a. Ion exchange tends to prefer counter ions of higher valency, higher

158

concentration and ions of smaller hydrated equivalent volume [32]. Therefore F removal 7

Page 7 of 62

159

using ion exchange resins is difficult because the order of selectivity for anions by anion

160

exchange resins is as follows: citrate > SO42- , oxalate > I- > NO3- > CrO42- > Br- > SCN- > Cl-

161

> acetate > F- [32]. H-bonding is a strong dipole-dipole attractive force between bonding of the strong

163

electropositive H atom in a molecule in an adsorbent or adsorbate and a strong

164

electronegative atom such as oxygen or fluorine in another molecule [33]. The energy of

165

adsorption in H-bonding is stronger than in van der Waals forces and ion exchange but

166

weaker than in the ligand exchange process discussed in the next paragraph. H-bonding

167

occurs in the adsorption of F on ion exchange resins [30,34] and coal-based adsorbents [35]

168

as illustrated in Fig. 2b.

an

us

cr

ip t

162

In a ligand exchange mechanism, the adsorbing anion such as F- forms a strong covalent

170

chemical bond with a metallic cation at the adsorbent surface resulting in the release of other

171

potential determining ions such as OH- ions previously bonded to the metallic cation (Fig.

172

2c). Thus, F is said to form an inner sphere complex or is specifically adsorbed on the

173

adsorbent surface. The adsorption of F on several multivalent metal oxides near neutral pH

174

was reported to have increased the pH of solutions as a result of release of OH- ions from the

175

adsorbents by ligand exchange of OH groups on the adsorbent surface with F in solution

176

[36,37]. Adsorbents with a ligand exchange mechanism have the particular advantage of

177

combining high adsorption capacity with high selectivity for anions. These adsorbents can

178

remove large proportions of anions having higher selectivity for adsorption from very dilute

179

solutions of the anions even in the presence of higher concentrations of competing anions of

180

lower selectivity. Adsorption of F by ligand exchange is illustrated in Fig. 2c [23,37,38].

Ac ce

pt

ed

M

169

181

The adsorption capacity of F on adsorbents can be increased by chemical modification of

182

adsorbent surfaces (Fig. 2d). This is particularly of advantage in the case of adsorbents

183

possessing negative surface charges which tend to repel the similarly charged F- ions. In such 8

Page 8 of 62

adsorbents, positively charged multivalent cations such as Al3+, La4+, Zr4+, Fe3+, and Ce3+ are

185

impregnated onto the adsorbent to create positive charges on the adsorbent surface for

186

attracting F- by coulombic forces as well as producing adsorption sites capable of chemical

187

interaction with F [13,23,39-44] (Fig. 2d). These metallic cations act as a bridge in adsorbing

188

F onto the adsorbent.

ip t

184

Adsorption of F on many microporous adsorbents is recognised as a 2-step process; an

190

initial rapid adsorption (mostly within an hour) at the outer surface of the adsorbent that

191

reaches a pseudo-equilibrium at the solid-solution interface followed by a much slower

192

process (hours to days) where the F moves by intra-particle diffusion into the interior pores

193

and channels of the adsorbent [45]. Intra-particle diffusion rate is directly related to the

194

square root of time of adsorption. Therefore if a straight line relationship is obtained between

195

the rate of adsorption and square root of time with the line passing through the origin, it can

196

be inferred that the diffusion process controls the adsorption, especially at longer times. Such

197

a relationship was obtained in many studies for the adsorption of F on different adsorbents

198

(granular ferric hydroxide [7], manganese oxide-coated alumina [29], fly ash [46], activated

199

alumina [47], and alum sludge [48]).

pt

ed

M

an

us

cr

189

It could be concluded that ligand exchange is the predominant mechanism of F adsorption

201

for inorganic adsorbents having high adsorption capacities. For organic adsorbents having

202

high adsorption capacities, H-bonding seems to be the predominant mechanism.

203 204

Ac ce

200

3. Factors influencing adsorption

205 206

3.1. pH

207

As for other contaminants, the removal of F from water by adsorbents is influenced

208

by several factors including pH, co-existing ions, temperature, adsorption kinetics, and 9

Page 9 of 62

adsorbent particle size and activation. Of these, pH is generally considered to be the most

210

important factor [16]. Fluoride adsorption is low at both very low and very high pH. The pH

211

at which the maximum amount of F is removed depends on the adsorbent characteristics but

212

generally it is between 4 and 8 (Table 1). One of the important properties of the adsorbent

213

influencing the extent of F adsorption is the pH at the point of zero charge (PZC). For

214

example, Fe and Al oxides having a PZC at around 7-8 remove the maximum amount of F at

215

pH 6-8 [7,47,49] and activated carbon with a PZC of 3.9-4.7 was reported to have removed

216

the maximum amount of F at pH 3-4 [42]. At pH values above the PZC the surface of

217

adsorbents is negatively charged, therefore the negatively charged F ions are not attracted to

218

the adsorbent surface. At pH values lower than the PZC the surface is positively charged,

219

therefore F ions are adsorbed. In some situations, at low pH values, F exist as positively

220

charged AlF complexes and this can reduce F adsorption [11,49,75].

M

an

us

cr

ip t

209

The main reason for a reduction in F adsorption below pH 4 is that F forms weakly

222

ionised HF at these pH values [13,47,48,50]. At a pH above 7-8 the removal of F decreases

223

not only because the adsorbent surface becomes negatively charged but also because the

224

concentrations of hydroxyl, bicarbonate and silicates ions increase so that they compete with

225

F for adsorption. In addition to increased number of positive surface charges at low pH values

226

increasing F adsorption, surface protonation at a pH less than the pH of PZC provides an

227

increased number of H atoms at the adsorbent surface leading to an increased number of H-

228

bonding between the H atoms and F in solution resulting in increased F adsorption [76].

229

Adsorption by a ligand exchange mechanism is also favoured at low pH because of a stronger

230

attractive force between F and the adsorbent surface and the presence of more hydroxylated

231

sites for exchange with F than at a high pH [4,29,61].

Ac ce

pt

ed

221

232

It is apparent from the literature review that F adsorption is generally lowest at

233

extremely low and high pH values. The adsorption is highest around the neutral pH that is 10

Page 10 of 62

234

commonly found in natural water. Therefore prior pH adjustment is not normally required for

235

effective removal of F in treatment plants.

236 237

3.2. Co-existing anions In natural water, several anions including PO43-, Cl-, SO42-, Br-, NO3- are

239

simultaneously present with F at different concentrations which can compete with F for

240

removal by adsorbents. The extent of the competition depends on the relative concentrations

241

of the ions and their affinity for the adsorbent. Meenakshi and Viswanathan [30] reported that

242

increased concentrations of Cl-, SO42-, Br-, NO3- , and HCO3- decreased F adsorption by an

243

anion exchange resin which adsorbs anions by an ion exchange mechanism, whereas these

244

anions had no effect on F adsorption by a chelating resin which adsorbed F selectively by a

245

H-bonding mechanism. The preferential order of adsorption of anions by the chelating resin

246

was reported to be F- > Cl- > NO3- > SO42-.

M

an

us

cr

ip t

238

Solangi et al. [34] studied the adsorption of F by a thio-urea incorporated amberlite

248

resin in the presence of PO43-, Cl-, SO42-, Br-, NO3-, NO2-, HCO3-, and CO32- at five times the

249

molar concentration of F and observed that Br-, NO2-, and PO43- had little interference with F

250

adsorption; the other ions had no interference. This was explained as due to the strong H-

251

bonding of F with the amide groups in the resin.

pt

Ac ce

252

ed

247

Multivalent metal oxides are known to adsorb F selectively by the ligand exchange

253

specific adsorption mechanism. An alum sludge containing a high percentage of Al, Ti and

254

Fe oxides adsorbed F selectively in the presence of SO42- and NO3- [48]. At 50 mg/L, SO42-

255

and NO3- reduced F adsorption from 85% to 40% and 62%, respectively from a solution

256

containing 20 mg F/L. In contrast to these anions, PO43- and selenate concentrations at 20

257

mg/L reduced F adsorption to 25% of F in solution. Based on these results, Sujana et al. [48]

258

proposed that the decreased order of competition of anions for F adsorption to be PO43- ≥ 11

Page 11 of 62

259

selenate > SO42- > NO3-. Das et al. [55] also found that PO43- interfered more than SO42- in F

260

removal by a calcined Zn/Al layered double hydroxide (LDH) consisting of Zn/Al oxide. Kumar et al. [7] investigated the adsorption of F by a granular ferric hydroxide in the

262

presence of competing anions such as Cl-, SO42-, BrO3-, NO3-, CO32- and PO43-, each having

263

concentrations of 20 to 100 mg/L with an initial F concentration of 20 mg/L. There was no

264

significant influence of competing anions on F removal when the adsorbent dose was 10 g/L,

265

which was attributed to the availability of plenty of sorption sites. However, when the

266

adsorbent concentration was reduced to 5g/L, PO43-, CO32-, and SO42- , each at concentrations

267

of 100 mg/L, this reduced the F adsorption capacity by 35, 25, and 20% of F in solution,

268

respectively. The other anions did not significantly reduce the F adsorption capacity.

269

Similarly, Raichur and Basu [50] found that F adsorption by a mixture of naturally occurring

270

rare earth oxides (oxides of La, Ce, Pr, Nd, Sm and Y) was not significantly affected by the

271

presence of SO42- or NO3- in water at a concentration equal to that of F (100 mg/L).

M

an

us

cr

ip t

261

It can be concluded that the non-specifically adsorbing anions (e.g. nitrate, chloride)

273

do not compete with F for adsorption on adsorbents that adsorb F using specific adsorption.

274

Only anions that are adsorbed by specific adsorption (e.g. phosphate, selenate, arsenate)

275

compete with F for adsorption. When F is adsorbed by non-specific adsorption the non-

276

specifically adsorbing anions can compete with F.

278 279

pt

Ac ce

277

ed

272

3.3. Temperature

Temperature had no consistent effect on F adsorption. Adsorption on many adsorbents

280

increased with temperature showing an endothermic nature of adsorption (granular ferric

281

hydroxide [7], fly ash [46], calcined Mg/Al/CO3 LDH [54], LDH/chitosan [63], spent catalyst

282

[75]). In contrast with many others it decreased with temperature showing an exothermic

283

nature of adsorption (chelating resin [30], alum sludge [48], calcined Zn/Al LDH [55], 12

Page 12 of 62

284

modified activated carbon [59], LDH/chitin [65], geo-materials [68]). Temperature has also

285

been reported to have no significant effect on adsorption by some adsorbents (trivalent

286

cations/zeolite [40]). The reasons for the differences in the effect of temperature were not clearly stated in

288

the studies. It may depend on the temperature range studied, the nature of the adsorbent, and

289

the conditions used in the studies. For example, at extremely low temperatures (5, 10oC) the

290

rate of adsorption was reported to be low because the rate of movement of F to the adsorption

291

sites is low. Lai and Liu [75] reported that F adsorption on spent catalyst increased

292

significantly when the temperature increased from 5oC to 25oC but very little further increase

293

in adsorption was observed when the temperature rose to 50oC. Sujana et al. [48] reported

294

that the exothermic nature of adsorption of F on alum sludge existed because the rising

295

temperature increased the tendency for F to escape from the adsorbent. Another reason given

296

was an increase in thermal energy of adsorbed F at higher temperatures, causing increased

297

desorption.

298

3.4. Adsorption kinetics

pt

299

ed

M

an

us

cr

ip t

287

Fluoride adsorption studies have shown that the rate of removal of F by adsorbents is

301

high in the initial 5-120 min where generally more than 90% of F is adsorbed, but thereafter

302

the rate significantly levels off and eventually approaches zero denoting the attainment of

303

equilibrium. This is because initially the adsorption sites are vacant and the F concentration

304

gradient between solution and adsorbent surface is high. Subsequently the rate decreases

305

because of the decrease in vacant sites. A fast rate of adsorption helps the adsorbent to treat

306

large quantities of water [54]. A slow rate causes operational, control, and maintenance

307

problems in the adsorption process of the filter bed [31].

Ac ce

300

13

Page 13 of 62

The rate of F adsorption increases with an increase in concentration of adsorbent [54]

309

and a decrease in initial F concentration [29,47,54]. It also depends on the structural

310

properties of the adsorbents and the interaction between F and the sites of adsorption.

311

Adsorption kinetics has been described by pseudo-first order and pseudo-second order kinetic

312

models and the diffusion model and rate constants have been determined (Table 1). These

313

models have also provided information on adsorption mechanisms.

4. Adsorbents

us

315

cr

314

ip t

308

316

The effectiveness of F adsorption from drinking water by various adsorbents, the

318

methods used for the assessment, and the kinetics and equilibrium models that best explained

319

the adsorption process are presented in Table 1. Caution needs to be exercised in comparing

320

the adsorption capacities of adsorbents because of the inconsistencies in data presentation

321

including differences in methodology and parameters used in the studies (pH, temperature, F

322

concentration range, competing ions, etc.). An ideal adsorbent that can be used to remove F

323

must have the following characteristics: low-cost, a high F adsorption capacity, rapid

324

adsorption of F, easily regenerated after its removal capacity is exhausted and good physical

325

characteristics (rapid water flow without filter clogging).

327 328

M

ed

pt

Ac ce

326

an

317

4.1. Metal oxides and hydroxides Oxides and hydroxides, also called hydrous oxides or oxyhydroxides, of trivalent and

329

tetravalent metals such as Fe, Al, La, Mn, and Zr are used to remove both anionic and

330

cationic contaminants from water and wastewaters because of their strong ability to adsorb

331

these ions [37,77]. The predominant mechanism of adsorption of F on oxides and hydroxides

332

is ligand exchange by the formation of inner sphere complexes (specific adsorption) as 14

Page 14 of 62

333

discussed previously. Most metal oxides and hydroxides have their PZC above the natural

334

water pH of 7 (granulated ferric hydroxide pH 7.5-8.0 [7], activated alumina pH 8.25 [36], γ-

335

alumina pH 8 [49]). Therefore, at the neutral pH of natural water, these adsorbents have

336

positive surface charge which is favourable for the adsorption of the negatively charged F. Of the oxides and hydroxides of metals, Al oxide, especially the activated form

338

(activated alumina) has been the most commonly used adsorbent for the removal of F

339

[9,29,36,37,42]. Activated alumina is produced by thermal degradation of aluminium

340

hydroxide to obtain materials with high specific surface area and a distribution of micro- and

341

macro-pores [78]. The specific surface area (m2/g) of activated alumina has been reported to

342

be 160 [49], 297 [29], and > 200 [26]. The adsorption capacity of activated alumina varies

343

with the structure of the alumina. For example, γ-Al2O3 has a much higher adsorption

344

capacity than α-Al2O3 [42] and therefore this form of activated alumina is commonly used for

345

defluoridation of water. The maximum F adsorption capacity of activated γ-Al2O3 (mg/g) has

346

been reported to be 1.1 [36], 12.0 [41], 2.41 [47], and 16.3 [49].

ed

M

an

us

cr

ip t

337

Aluminium oxides have poor adsorption capacity for F in acidic conditions because of

348

their tendency to dissolve and form positively charged AlF complexes (AlF2+, AlF2+) which

349

are repelled by the positively charged surfaces of Al oxides at these conditions (PZC of Al

350

oxides > pH 7) [49]. In alkaline conditions, Al oxides have negatively charged surfaces

351

which repel the negatively charged F ions present at these pHs. Also, as stated previously, the

352

increase in the concentrations of OH- in alkaline condition competes with F- for adsorption.

353

Therefore the optimum pH for F adsorption is considered to be near neutral pH [47,49].

Ac ce

pt

347

354

Recent studies have shown that surface coatings of adsorbents with other materials

355

have enhanced the adsorption of F [29]. Because manganese oxides have a high specific

356

surface area, a micro-porous structure [79] and a high adsorption capacity towards anions

15

Page 15 of 62

357

[80], activated alumina was coated with these oxides and F adsorption capacities have been

358

studied [29,36]. Maliyekkal et al. [36] compared the F adsorption capacity and rate of adsorption of

360

activated alumina (AA) and a granular manganese oxide coated AA (MOCAA) at pH 7 and

361

found that in the batch study, most of the adsorption was complete in 3 h in the case of

362

MOCAA compared to 10 h for AA. The pseudo-first order and pseudo-second order rate

363

constants were also higher for MOCAA. The Langmuir adsorption capacity of MOCAA was

364

2.85 mg/g compared to 1.08 mg/g for AA. Maximum F adsorption was found to occur at a

365

wider pH range of 4-7 for MOCAA compared to that for AA (4-6). The column study also

366

showed that MOCAA had higher F adsorption capacity. The superiority of MOCAA over AA

367

in the adsorption of F was reported to be not due to the surface area difference because the

368

specific surface area of AA was 204 m2/g and that of MOCAA was 170 m2/g. Maliyekkal et

369

al. [36] suggested that the reason for this could be the increased zeta potential (surface

370

charge) of the MOCAA, although no supporting data were presented.

ed

M

an

us

cr

ip t

359

Teng et al. [29] used a redox process to coat AA with an amorphous MnO2. The

372

granular MOCAA produced had a higher surface area (316 m2/g) compared to that for AA

373

(297 m2/g) and a significantly rougher surface with plenty of pores. Maximum F adsorption

374

was obtained at the pH range of 4-6. One of the two mechanisms of F adsorption at a pH less

375

than 6 was considered to be chemical adsorption by a ligand exchange of surface OH groups

376

in MOCAA by F in solution resulting in an increase of pH due to the release of OH-. The

377

other mechanism was intra-particle diffusion of F. The Langmuir adsorption maximum at pH

378

5.2 was 7.09 mg/g which was much higher than the value of 1.08 mg/g reported for AA by

379

Maliyekkal et al. [36] and values reported for many other granulated adsorbents. For initial F

380

concentrations less than 21 mg/L, most of the adsorption was completed within 30 min

Ac ce

pt

371

16

Page 16 of 62

381

compared to 10 h for AA. Faster adsorption by MOCAA was considered to be due to the

382

larger surface area and porous surface of this adsorbent. Iron oxides are also known to have large capacity to remove anions from water by

384

mechanisms similar to those operating in Al oxide adsorbents [81]. Kumar et al. [7] studied

385

the adsorptive removal of F by a highly porous and poorly crystalline granulated ferric

386

hydroxide (GFH) (β-FeOOH) with a specific surface area of 250-300 m2/g, a PZC of pH 8,

387

and a granular size of 0.32-2.0 mm. Nearly 95% adsorption of F from solution was achieved

388

within the first 10 min of agitation of 10-20 mg F/L with 10 g GFH/L at pH 6-7. The

389

maximum F adsorption was observed at the pH range of 3-8. A sharp decrease of F

390

adsorption was observed above the PZC pH of 8 as the GFH surface became more negatively

391

charged causing electrostatic repulsion of the negatively charged F- ions in addition to

392

increasing concentration of OH- ions which competed with F- for adsorption.

M

an

us

cr

ip t

383

Hydroxyapatite, the most abundant of phosphate minerals, has been used for

394

defluoridation of water. Fan et al. [52] compared the F adsorption capacities of several

395

natural minerals and found that the adsorption capacities at pH 6 followed the order:

396

hydroxyapatite > fluorspar > quartz activated using ferric ions > calcite > quartz. The highest

397

F adsorption capacity of hydroxyapatite was explained as owing to F exchanging with a OH

398

group at the surface and inside the apatite mineral. The F adsorption on the other minerals

399

was deemed to be a surface adsorption process.

pt

Ac ce

400

ed

393

It is evident from the literature review that Fe and Al oxides and hydroxides are the

401

commonly used adsorbents for defluoridation. They have a moderate level of F adsorption

402

capacity (1-16 mg/g) and if locally available at low cost, can potentially be employed in rural

403

areas, especially in developing countries.

404 405

4.2. Layered double hydroxides 17

Page 17 of 62

The majority of clay minerals such as kaolinite, mica, montmorillonite, vermiculite

407

and zeolite, carry predominantly negative charges and therefore adsorb very small amounts of

408

anions. Layered double hydroxide (LDH) or hydrotalcite (HTlc) is another type of clay

409

mineral, but has positive charges and therefore adsorb significant quantities of anions and

410

oxyanions (e.g. fluoride, arsenite, arsenate, chromate, phosphate, selenite, selenate, nitrate,

411

etc) and monoatomic anions (e.g. fluoride, chloride) from aqueous solutions [82].

412

Structurally, LDHs are composed of positively charged brucite-like sheets compensated by a

413

large number of exchangeable charge-balancing anions in the hydrated interlayer regions

414

[55,56,82]. Charges can also be produced by the ionisation of the OH groups at the surface of

415

the LDH particles. The PZC of LDHs is in the region of pH 9-12 [82] and therefore at the

416

neutral pH of natural water, LDHs carry positive charges and act as anion exchangers.

an

us

cr

ip t

406

Calcined LDHs have higher F adsorption capacities than the uncalcined LDHs [53].

418

The optimum temperature of calcination is generally considered to be 450-500oC [53-56]. Lv

419

et al. [54] reported that the F adsorption capacity of a Mg/Al LDH increased from 65 to 70

420

and then to 80 mg/g when the calcination temperature was increased from 200oC to 400oC

421

and then to 500oC, respectively, but decreased to 62 and 50 mg/g when the temperature was

422

increased to 600oC and 800oC, respectively. Wang et al. [53] suggested that the increase in F

423

adsorption capacity of LDH as a result of calcination was due to the higher specific surface

424

area, porosity, and surface reactivity of the Mg/Al oxide produced by calcination. Another

425

reason reported was the incorporation of F into the structure of Mg/Al oxide resulting in the

426

formation of the original structure of the LDH. The decrease in F adsorption capacity at

427

calcining temperatures higher than 500oC was considered to be due to the transformation of

428

the Mg/Al oxides into a spinel structure that does not exhibit the property of LDH structural

429

reconstruction [54].

Ac ce

pt

ed

M

417

18

Page 18 of 62

The adsorptive property of LDH depends on the metallic constitution of the LDH

431

structure. Lv et al. [54] showed that the F adsorption capacity of calcined Mg/Al LDH was

432

higher than that of calcined Ni/Al LDH and calcined Zn/Al LDH because of the higher

433

atomic weights of Ni and Zn compared to Mg. Among the calcined Mg/Al LDHs, the one

434

having Mg/Al molar ratio of 2 was found to have the highest F adsorption capacity. The

435

maximum F adsorption capacity of 213 mg/g was obtained at pH 6. This adsorption capacity

436

is the highest of all adsorbents listed in Table 1. However, Das et al. [55] reported a lower

437

Langmuir adsorption capacity of 17 mg/g at pH 6 for a calcined Zn/Al LDH. The F

438

adsorption data of Wang et al. [53] on a calcined Mg/Al LDH did not fit the Langmuir

439

adsorption model but the amount of F adsorbed continued to increase with equilibrium F

440

concentration with an adsorption capacity of 35 mg/g at the highest tested F equilibrium

441

concentration of 70 mg/L. The rate of F adsorption on calcined LDH is variable [53,55,56].

M

an

us

cr

ip t

430

The literature review reveals that LDHs can have very high F adsorption capacity (17-

443

213 mg/g) if calcined to 500oC. The adsorption capacity varies depending on the type and

444

proportion of the metals in the LDH structure. Because LDHs have high adsorption capacity,

445

they are useful adsorbents for defluoridation of waters with high F concentrations.

447 448

pt

Ac ce

446

ed

442

4.3. Ion exchange resins and fibres Ion exchange resins and fibres are an important class of adsorbents used to remove

449

anionic and cationic pollutants from water and wastewater. Their framework or matrix

450

consists of irregular, macromolecular, three dimensional network of hydrocarbon chain [32].

451

The cation exchange resins and fibres have negatively charged functional groups whereas the

452

anion exchange resins and fibres have positively charged functional groups such as -NH3+,

453

NH2+, ≡N+, ≡S+. Therefore the cation exchangers adsorb cations and the anion exchangers

454

adsorb anions such as F-. The cation exchangers can also be made to adsorb anions if they are 19

Page 19 of 62

455

impregnated with positively charged metallic cations that have strong affinity for anions

456

[39,57]. Ku et al. [57] used an Al incorporated cation exchange resin (Amberlite IR-120) to

458

remove F from water and found that the Langmuir maximum F adsorption capacity at pH 4

459

was 4.6 mg/g. Within the pH range of 4-9 tested, F adsorption emerged as the greatest at pH

460

5-7. Metals other than Al have also been incorporated into cation exchangers to enhance the F

461

adsorption capacity. For example, Luo and Inoue [39] compared the F adsorption capacities

462

of the cation exchange resin, Amberlite 200 CT modified by the incorporation of a number of

463

trivalent metal ions, La, Ce, Y, Fe, and Al. The F adsorption capacity of the metals

464

incorporated resins in the pH range of 4-7 was in the order: La ≤ Ce > Y > Fe ~ Al. More

465

than 80% of F was adsorbed by 20 g La/L resin from a 15 mL solution containing 15 mg F/L.

466

The resin without any metal incorporation adsorbed only about 5% of F from the solution.

M

an

us

cr

ip t

457

The majority of anion exchange materials are not effective in adsorbing F from

468

natural water containing other anions because F- affinity to anion exchange materials is the

469

least of all anions (citrate > SO42- > oxalate > I- > NO3- > CrO42- > Br- > SCN- > Cl- > formate

470

> acetate > F- [32]). Consistent with this order of affinity, a strong base anion exchanger was

471

shown to effectively adsorb NO3-, Br-, and SO42-, but not Cl- and F- [83].

pt

Ac ce

472

ed

467

The F adsorption capacity of anion exchangers can be enhanced by modifying the

473

adsorbing sites on the anion exchangers. Generally, anion exchange resins impregnated with

474

chelating agents that can form H-bonding with F are used. Meenakshi and Viswanathan [30]

475

compared the F adsorption capacity of a chelating resin having sulfonic acid functional group

476

and an anion exchange resin and showed that 1 g of the chelating resin adsorbed 95% of F

477

from a 50 mL solution containing 3 mg F/L in 40 min compared to 65% adsorption by the

478

anion exchange resin.

20

Page 20 of 62

Solangi et al. [34] reported that 100 mg of the ion exchange resin, Amberlite XAD-4

480

modified by incorporation with thio-urea binding sites removed 90% of the F from a 10 mL

481

solution containing 16 mg F/L compared to 30% by the unmodified resin at pH 7. The higher

482

F adsorption capacity of the modified resin was explained as being due to H-bonding between

483

the amide groups in thio-urea and F. The Langmuir maximum F adsorption capacity of the

484

modified resin at pH 7 was much higher than the F adsorptive capacities of many other

485

adsorbents in the literature (Table 1). Interference of other co-existing anions present at the

486

concentration ratio of 1:5 (F:other ions) was insignificant. Therefore the authors concluded

487

that the modified resin can be effectively used for defluoridation of water.

us

cr

ip t

479

Fibrous adsorbents, due to their physico-chemical structure, generally have rapid rates

489

of adsorption of ions. If their adsorptive capacity and affinity can also be enhanced, they can

490

be a useful class of adsorbents for removal of ions from water. Ruixia et al. [31] introduced

491

functional groups into a polyacrylonitrile ion exchange fibre and studied its F adsorption

492

behaviour. They found that the F adsorption reached equilibrium in 5 min with 40% of F

493

adsorption from a 500 mL solution containing 5, 34, and 50 mg/L of F, As(V), and P,

494

respectively when 1 g fibre modified with functional groups was added. The adsorption by

495

the unmodified fibre was less than 5% of solution F. The rate of adsorption was considered to

496

be faster than that observed in many other adsorbents. Maximum adsorption was obtained at

497

pH 3. The mechanism of adsorption was considered to be ion exchange.

M

ed

pt

Ac ce

498

an

488

It could be concluded from this review that ion exchange resins and fibres have poor

499

F adsorption capacities but the adsorption capacities and selectivity for F adsorption in the

500

presence of other ions can be significantly increased (up to 61 mg/g) by surface modification

501

of the adsorbents by loading with organic functional groups and metals. These adsorbents are

502

relatively expensive and therefore they are useful only for water treatment in industrial

503

countries. 21

Page 21 of 62

504 505

4.4. Zeolites Because zeolites have negative surface charges at all pH values, they have high

507

adsorption capacity for cations, but have low adsorption capacity for anions. Nonetheless

508

their adsorption capacity for anions can be increased by modifying the zeolite surface with

509

cationic surfactants or multivalent metallic cations [85]. For F adsorption, only metallic

510

cations incorporated with zeolites have been used. There appear to be no studies conducted

511

on using surfactant- or organic compounds-modified zeolites on defluoridation.

us

cr

ip t

506

Onyango et al. [40] modified the surface of a synthetic zeolite by exchanging Na+ in

513

the zeolite with Al3+ and La3+ to create active sites for F adsorption. The introduction of Al

514

and La opened up the pores in zeolite leading to increased porosity. The Langmuir adsorption

515

capacity is slightly larger for La-zeolite than for Al-zeolite (Table 1). However, within the F

516

concentration range studied (10-80 mg/L), Al-zeolite had twice as much adsorption capacity

517

as the La-zeolite. The adsorption capacities obtained for these zeolites were reported to be

518

much higher than many other adsorbents including the commonly used activated alumina

519

(Table 1). This suggested that the mechanism of adsorption of F onto Al-zeolite was mostly

520

by a chemical adsorption process (ligand exchange) and adsorption onto La-zeolite was

521

mostly by a physical adsorption process (coulombic attraction).

M

ed

pt

Ac ce

522

an

512

The Na-zeolite had no PZC because it was negatively charged at all pHs between 3

523

and 13 and therefore the adsorption of F- is poor due to charge repulsion. In contrast, the Al-

524

zeolite and the La-zeolite had PZC at pHs of 8.15 and 4-5.25, respectively indicating that

525

below these pHs these zeolites were positively charged. In accordance with the surface

526

charge on La-zeolite the F adsorption was the highest at pH 5; and decreased above and

527

below this pH. In the case of the Al-zeolite, the increase of pH increased F adsorption with a

528

plateau forming above pH 5. The F adsorption on Al-zeolite at pHs above 5 did not decrease 22

Page 22 of 62

529

as observed for La-zeolite up to the highest pH 9 tested because F was reported to be

530

adsorbed by chemical adsorption. The authors stated that, over all, Al-zeolite was superior to

531

La-zeolite for defluoridation. In a subsequent study by Onyango et al. [41] on F adsorption by four types of Al pre-

533

treated low-silica synthetic zeolite it was observed that F adsorption on all four types

534

increased from pH 2 to 4 and remained constant up to pH 8, before decreasing from pH 8 to

535

11. The pH effect on F adsorption was explained by the following reactions.

cr

ip t

532

an

us

536

537

M

538

At low pH, part of the F is removed as weakly ionised HF and therefore F adsorption is

540

reduced (first equation). Another reason could be that some of the F is complexed to Al

541

solubilised at low pHs to form AlF+ complex which has low tendency to adsorb on the

542

positively charged adsorbent at the low pHs. At high pH, the negatively charged adsorbent

543

(last equation) repels the negatively charged F- as well as increased competition of OH- for

544

adsorption.

pt

Ac ce

545

ed

539

Samatya et al. [23] using a natural zeolite from Turkey pre-treated with La, Al, and Zr

546

to study the removal of F from tap water spiked with NaF. They showed that the F adsorption

547

capacities at the equilibrium F concentration range of 0-12 mg/L were largest for Zr-zeolite

548

and smallest for La-zeolite. The values for the La and Al zeolites were lower than the values

549

reported by Onyango et al. [40], probably because the metal loadings were much lower and

550

the zeolite used was natural compared to the synthetic zeolite of Onyango et al. [40]. 23

Page 23 of 62

551

However, all three metal zeolites removed 95% of F from an aqueous solution containing 2.5

552

mg F/L at an adsorbent dose of 6 g/L. In contrast to the above studies which showed high F removal capacities of

554

multivalent metal incorporated zeolites, Diaz-Nava et al. [85] found that La and Eu treated

555

natural zeolite from Mexico had only slightly higher F removal capacities than when this

556

zeolite was treated with Na and Ca. The weight percentages of the metals in the zeolite were

557

1.7, 3.20, 0.32, and 1.49 for Na, Ca, La, and Eu, respectively.

cr

ip t

553

The review revealed that zeolites on their own had very low adsorption capacities for

559

F but when they were loaded with multivalent metallic cations they produced moderate to

560

very high F adsorptive capacities (2-45 mg/g). The adsorptive capacity depends on the type

561

and amount of metal loading. Zeolites can also provide good physical properties for practical

562

use.

M

an

us

558

564

4.5. Carbon materials

ed

563

Activated carbon (AC) is an important carbon material commonly used as an

566

adsorbent for the removal of a wide range of aquatic pollutants due to its exceptionally high

567

surface area (500-1500 m2/g), highly developed internal microporosity, presence of a range of

568

functional groups, low cost and ready availability [86,87]. However, it displays poor

569

adsorption capacity towards anionic pollutants because of its low PZC (pH 1.6-3.5, [88]).

Ac ce

570

pt

565

The amount of anions adsorbed onto AC depends on the pore size distribution

571

because the adsorption occurs mainly in the pores. Abe et al. [60] reported that F adsorption

572

capacity increased with the specific surface area in 11 out of the 12 carbonaceous materials

573

studied by them. The adsorption of F on bone char did not fit this pattern because the

574

mechanism of adsorption was different from the rest. In bone char, F is adsorbed chemically

575

by ligand exchange with the OH group in the hydroxyapatite compound [1]. The 24

Page 24 of 62

ineffectiveness of carbon in adsorbing F was also shown by Srimurali et al. [70] who found

577

that lignite coal and char fines, a by-product obtained during the making of coke, adsorbed

578

only 7.9 and 19% of F, respectively from a 50 mL solution containing 5.0 mg F/L and when

579

100 mg of adsorbents were added and mixed for 5 h. In comparison, bentanite clay removed a

580

much larger percentage of 33% of solution F.

ip t

576

A new type of carbon material called aligned carbon nanotubes (ACNT) made up of

582

needle-like cylindrical tubules of concentric graphitic carbon capped by fullerene-like

583

hemispheres was developed as a promising adsorbent material for the adsorption of

584

contaminants in water [12]. Li et al. [12] found that ACNT had higher F adsorption capacity

585

than AC (Table 1), despite the ACNT having a much lower specific area and pore volume

586

than the AC. Li et al. [12] stated that though the adsorption capacity of the ACNT was high,

587

the cost too was high which may limit its full commercial utilisation.

M

an

us

cr

581

The surface of the carbon particles can be modified to improve the F adsorptive

589

properties. Such modifications have been brought about by creating new functional groups

590

which have strong affinity towards F. Daifullah et al. [59] modified the structure of an AC

591

product by steam pyrolysis of rice straw and oxidised the product using HNO3, H2O2, and

592

KMnO4. Of these treatments, the material obtained by KMnO4 oxidation produced the

593

highest F adsorption. The adsorption mechanism was considered to be ligand exchange of

594

OH groups on the carbon surface with F. The presence of MnO2 on the carbon surface caused

595

by reduction of the KMnO4 may have also participated in the removal of F. Thermodynamic

596

studies showed that the adsorption was chemical in nature. The Langmuir adsorption

597

maximum was 15.9 mg F/g at the natural pH of water. This value was considered to be higher

598

than the values reported in the literature for many other adsorbents.

Ac ce

pt

ed

588

599

Ramos et al. [42] studied the F adsorption behaviour of Al-impregnated AC (AlAC)

600

produced from coconut shells followed by calcination to 300oC. The AlAC had a F 25

Page 25 of 62

adsorption capacity more than four times higher than the plain carbon at an equilibrium

602

concentration of 2-8 mg F/L. The Langmuir adsorption capacity of AlAC (1.07 mg/g) at

603

natural water pH, though higher than the value obtained for plain carbon (0.49 mg F/g), was

604

lower than the values obtained for some other adsorbents especially activated alumina. The F

605

adsorption reached its maximum at pH 3 because the PZC of the AlAC was 4.1-4.8 (above

606

this pH, AlAC had negative surface charges which repel F-). However, there was appreciable

607

adsorption of F at pH 6-7 (0.6 mg F/g at the equilibrium F concentration of 6 mg F/L) due to

608

chemical adsorption of F onto Al which does not involve electrostatic attractive forces.

us

cr

ip t

601

In a later study, Li et al. [61] reported that the adsorption capacity of aligned carbon

610

nanotubes (ACNT) at pH 6 increased from 2.3 mg F/g to 9.6 mg F/g when Al was loaded

611

onto ACNT and calcined to 450oC at an optimum Al2O3 rate of 30% by weight. The

612

adsorption capacity fell to less than 6 mg F/g when the loading rate was 20 and 40%. The

613

maximum adsorption capacity was obtained at pH 6.0-9.0. The mechanism of F adsorption at

614

pH less than the PZC of 7.5 was considered to be by electrostatic attraction of negatively

615

charged F- ions onto the positively charged adsorbent as well as by ligand exchange of the

616

OH group by F. At the PZC and slightly above this pH the adsorption was reported to be

617

mainly by ligand exchange. At pHs much higher than the PZC (pH > 9), OH- ions in solution

618

competed with F- ions for adsorption and therefore F adsorption decreased.

M

ed

pt

Ac ce

619

an

609

It can be concluded from the review that AC is a poor adsorbent for F. However, its

620

adsorption capacity can be increased to moderate levels (up to 16 mg/g) by chemical

621

modification of the carbon surface. Specialised carbon materials such as ACNT have higher

622

adsorption capacity than AC. Their adsorption capacity can also be increased by surface

623

modification but this will incur additional cost.

624 625

4.6. Natural materials 26

Page 26 of 62

626

Several natural inorganic materials (soils, clays, minerals, and building materials)

627

have been used in the defluoridation of water. The adsorption capacities of these materials

628

have been discussed in previous reviews (17, 26, 27) and also presented in Table 1. Bioadsorbents contain a variety of functional groups such as carboxyl, imidazole,

630

sulphydryl, amino, thioether, phenol, carbonyl, amide, and hydroxyl moieties capable of

631

adsorbing the pollutants, especially the metal cations [78]. Considering this feature of

632

bioadsorbents, metals having strong affinity towards F were loaded onto selected

633

bioadsorbents (Zr4+ on collagen fibres [13], La3+ on geletin [44]) for enhancing the adsorption

634

of F from drinking water.

us

cr

ip t

629

Chitin and chitosan have been shown to constitute another effective group of

636

bioadsorbents for removing a range of aquatic pollutants due to their low cost and high

637

contents of amino and hydroxyl functional groups [89]. Chitin is a polysaccharide found in a

638

wide range of organisms but most commonly extracted from shellfish processing waste,

639

while chitosan is a copolymer of glucosamines derived from chitin by deacetylation in hot

640

alkali [67]. Sahli et al. [22] reported that the adsorption of F on chitosan from a synthetic

641

water was rapid and reached a maximum at 5 min. Increase of pH from 2 to 6 increased

642

adsorption followed by a decrease up to pH 10. The F adsorption from a brackish water

643

containing 3.25 mg F/L and several other anions at much higher concentrations (Cl- 1083

644

mg/L, SO42- 215 mg/L, HCO3- 171 mg/L) showed that the selectivity of chitosan towards

645

various anions was as follows: F- > HCO3- >NO3- > Cl- > SO42-. In a batch adsorption study of

646

15 min duration the F concentration fell to 1.82 mg/L, but with a successive batch adsorption

647

the concentration reduced to 0.9 mg/L which was lower than the WHO standard of 1.5 mg

648

F/L.

Ac ce

pt

ed

M

an

635

649

Ma et al. [66] prepared a magnetic-chitosan adsorbent by co-precipitation of chitosan

650

with Fe2+ and Fe3+ salts and used it to remove F from a synthetic F solution in a batch study. 27

Page 27 of 62

The adsorbent was shown to have a higher F adsorption capacity (Langmuir adsorption

652

capacity 22.49 mg F/g) than a commercial activated alumina (Table 1). Adsorption capacity

653

was highest in the pH range 5-9. The FTIR data of the adsorbent showed that the main groups

654

involved in the adsorption process were amine and iron oxides in the magnetic-chitosan. The

655

main advantage of having magnetic properties in the adsorbent is to easily separate the

656

adsorbent from the treated water after its use using an external magnetic field so that the

657

adsorbent can be reused.

cr

ip t

651

A research group in India has reported that the F adsorption capacity of nano-

659

hydroxyapatite (nHAp) can be increased by incorporating chitin or chitosan onto nHAp [63-

660

65,90]. In a batch experiment conducted by Sundaram et al. [65], it was observed that mixing

661

0.25 g each of chitosan, nHAp, and nHAp/chitosan composite with 50 mL solution of NaF

662

containing 10 mg F/L, removed 0.05, 1.30, and 1.56 mg F/g adsorbent, respectively after 30

663

min of shaking the suspensions. Defluoridation of water from a fluoride-endemic village

664

using nHAp-chitosan composite was found to be higher compared to using nHAp. Similar

665

results were obtained from a subsequent study where a composite of nHAp and chitin was

666

used [64]. However, the adsorption capacity of this composite was higher than that of the

667

composite made with chitosan (Table 1).

an

M

ed

pt

Ac ce

668

us

658

The review revealed that although the majority of natural materials have low

669

effectiveness in removing F (< 1 mg/g), they are inexpensive and consequently can be used in

670

less developed countries, especially in rural areas. Furthermore, their F removal capacities

671

can be increased through the process of surface modification, such as loading with

672

multivalent metallic cations, as reported previously for other adsorbents.

673 674

4.7. Industrial by-products

28

Page 28 of 62

675

Several types of industrial by-products have been used for the adsorptive removal of

676

pollutants including fluoride from water [27,28,89,91,92]. Most of the adsorbents fall into the

677

categories of by-products from the mining industry (e.g. red mud), steel industry (e.g. slag

678

materials), and power plant industry (e.g. fly ash). Cengeloğlu et al. [4] investigated the adsorption of F from aqueous solutions using a

680

red mud containing 18.7% Al2O3, 39.7% Fe2O3, and 14.5% SiO2 with (ARM) and without

681

(RM) activation by 20% HCl treatment in a batch study. The amount of F removed from a

682

solution (pH 5.5) containing 3.8-28.5 mg F/L by ARM at a dose of 0.2 g/50 mL was at least

683

three times higher than that by RM at the same dose. The main mechanism of F removal was

684

considered to be that of specific adsorption caused by ligand exchange of F and OH groups

685

on the metal oxide surfaces in red mud.

an

us

cr

ip t

679

Red mud particles are too fine for use in filter beds because of their poor hydraulic

687

conductivity. This problem can be overcome either by mixing red mud with coarse-size

688

particles such as sand [93] or by granulation [94] before use. Tor et al. [72] used the method

689

of Zhu et al. [94] for the granulation of red mud by mixing it with fly ash, sodium carbonate,

690

quicklime and sodium silicate. The maximum F adsorption (2.2 mg/g) from a NaF solution

691

containing 15 mg F/L and granulated red mud (GRM) of 5 g/L was obtained at a pH of 4.7 in

692

the pH test range of 2.5-7.3. The adsorption capacity determined from a column study was

693

2.05 mg F/g for a flow rate of 2 mL/min compared to 0.644 mg F/g obtained in the batch

694

study at the same GRM dosage (5g/L) and initial F concentration (5 mg/L).

ed

pt

Ac ce

695

M

686

Fly ash is a major by-product that is produced from the combustion of coal in power

696

stations. It consists of fine and powdery materials (1.0-100 µm) made up of a mixture of

697

amorphous and crystalline alumina-silicates and several compounds of Si, Al, Fe, Ca, and Mg

698

[95] and therefore a good candidate material for F adsorption from water. Chaturvedi et al.

699

[46] reported that the adsorption of F from water by a fly ash containing 56.0% SiO2, 25.9% 29

Page 29 of 62

Al2O3, 2.22% CaO, and 1.26% Fe2O3 had a maximum Langmuir adsorption capacity of 20

701

mg/g at pH 6.5. The F adsorption increased from 79 to 94% when the pH of the F solution

702

(10 mg/L) increased from 2.0 to 6.5 and then decreased with further increase in pH up to 9.5.

703

However, Nemade et al. [62] reported that F adsorption by a fly ash decreased continuously

704

from pH 2 to 12. The difference in the pH effects between the two studies could be due to the

705

difference in chemical characteristics of the fly ashes and experimental conditions used.

ip t

700

In a column study, Piekos and Paslawska [74] found that an alkaline fly ash (pH ≥ 10,

707

9.1% CaO) effectively removed F from water containing 1-100 mg F/L when the solution

708

was passed through a column (400 mm length) packed with 450 g fly ash at a flow rate of ≤ 2

709

mL/h. Complete adsorption of F was obtained after 120 h. The F adsorption mechanism was

710

suggested to be chemical binding of F onto Ca(OH)2 and physical adsorption onto the

711

residual carbon particles in the fly ash.

M

an

us

cr

706

Steel industry by-products such as blast furnace slag, electric arc furnace slag, basic

713

oxygen furnace slag and converter slag have been shown to adsorb many pollutants,

714

especially phosphate, from water due to the presence of high contents of Ca, Fe, Al, Mg, and

715

Mn oxides [28]. However, limited studies have been conducted in their use in defluoridation

716

of water [16,27]. Islam and Patel [71] showed that a basic oxygen furnace slag (BOFS)

717

containing 46.5% CaO, 16.7% Fe oxides, and 13.8% SiO2 had a good capacity to adsorb F.

718

Thermal activation of the BOFS by heating to 1000oC for 24 h increased the porosity and

719

surface area leading to increased F adsorption. The percentage adsorption of F (initial F

720

concentration 10 mg/L, 0.5 g/100 mL adsorption dosage) was higher for the thermally

721

activated BOFS (TABOFS) (93%) than for the BOFS (70%). Unlike the case with most

722

adsorbents, increase of pH from 2 to 10 increased F adsorption on TABOFS. The increased

723

adsorption at high pH values was considered to be due to F adsorption onto Ca(OH)2 in the

Ac ce

pt

ed

712

30

Page 30 of 62

724

slag. Presence of other anions reduced F adsorption in the order, PO43- > HCO3- > CO32- >

725

SO42- > Cl- > NO3-. Large quantities of alum sludge, a waste product, are generated during the

727

manufacture of alum from bauxite [48] and kaolin [96] by the sulphuric acid process in many

728

countries. As this sludge mainly consists of oxides of Al, it has a high adsorption capacity to

729

remove many aquatic pollutants. Sujana et al. [48] used a sludge (47.2% Al2O3 and 20.7%

730

TiO2) produced from an alum manufacturing plant using bauxite raw material for the removal

731

of F from water. The sludge after washing with water and calcining to 400oC for 3 h adsorbed

732

more than twice the amount of F that was adsorbed by the uncalcined sludge from a solution

733

containing 10 mg F/L of pH 6 (adsorbent dosage 20-300 mg/g). The higher adsorption

734

capacity of the calcined sludge was considered to be due the increased surface area produced

735

by calcination. The F was considered to be specifically adsorbed according to a two-step

736

ligand exchange mechanism. Within a pH range of 3 to 9, the maximum F adsorption was

737

observed at pH 6; and at both above and below pH 6 the adsorption decreased, as also

738

reported for activated alumina by others. The presence of other anions in solution reduced F

739

adsorption according to the order, PO43- > silicate > SO42- > NO3-.

pt

ed

M

an

us

cr

ip t

726

In contrast to F adsorption on sludge produced during the manufacture of alum from

741

bauxite discussed in the previous paragraph, Nigussie et al. [96] reported that F adsorption by

742

a sludge produced during the manufacture of alum from kaolin raw material containing

743

primarily SiO2 did not increase when the sludge was heated to 300oC. At temperatures above

744

300oC up to 700oC, the adsorption decreased. Also, F adsorption was nearly constant from

745

pH 3 to 8 and significantly decreased above pH 10. The difference in the adsorption

746

behaviour of the two types of sludge may be due to the differences in the physico-chemical

747

characteristics of the sludge types. However, the F adsorption capacities of the sludge in the

748

two studies were not too different. The presence of HCO3- at concentrations higher than those

Ac ce

740

31

Page 31 of 62

749

of F, decreased F adsorption efficiency, while other anions (PO43-, SO42- , Cl- , NO3-) had no

750

significant effect [96]. Lai and Liu [75] used a by-product of the petrochemical industry called spent catalyst

752

consisting mainly of porous silica and alumina (specific surface area 130 m2/g, PZC 5.2) to

753

remove F from aqueous solutions. In a batch adsorption experiment on pH effect, they

754

observed that F adsorption decreased with increase in pH from 2 to 9.5, with a plateau in the

755

pH range of 4 to 7. The maximum adsorption capacity at pH 4 was found to be 28 mg F/g at

756

25 and 50oC. This value was considered to be comparable to that of activated alumina.

757

Because the activation energy calculated using the Arrhenius equation was very low (3.2

758

J/mol) the mechanism of adsorption was considered to be non-specific adsorption involving

759

coulombic forces.

an

us

cr

ip t

751

Many of the industrial by-products are wastes that need to be disposed of. Their

761

beneficial use can save disposal cost, prevent environmental pollution arising from the

762

disposal sites, and release disposal lands for alternative uses. In this respect, though these

763

materials have low F adsorption capacities, they are cost-effective and therefore can

764

potentially be used in developing countries where cost of operation is a major factor in the

765

choice of adsorbents. Furthermore, many of these materials have been shown to be good

766

adsorbents for other aquatic pollutants and therefore it is possible to simultaneously remove

767

many pollutants using these materials.

769

ed

pt

Ac ce

768

M

760

5. Adsorption thermodynamics

770

The strength and spontaneous nature of adsorption and information on whether the

771

adsorption process is exothermic or endothermic are provided by determining thermodynamic

772

parameters such as changes in Gibbs free energy (ΔG0), enthalpy (ΔH0), and entropy (ΔS0).

32

Page 32 of 62

773

These thermodynamic parameters are calculated from data on adsorption at different

774

temperatures using standard thermodynamics equations (see references in Table 2). A negative value for ΔG0 indicates spontaneous and thermodynamically favourable

776

adsorption, while a negative ΔH0 value indicates an exothermic adsorption process. An

777

exothermic reaction means that the amount of adsorption decreases with increasing

778

temperature. An endothermic reaction (positive ΔH0 value) has been explained as due to

779

enlargement of pore sizes and/or activation of the adsorbent surface [97,98]. Very low ΔH0

780

values are generally associated with physical adsorption and very high values with chemical

781

adsorption [99]. However, no definite value for distinguishing the two forms of adsorption

782

exists. ΔG0 has also been used to distinguish between the two forms of adsorption. For

783

example, Meenakshi and Viswanathan [30] reported that ΔG0 values of up to -20 kJ/mol were

784

indicative of physical adsorption, while values less than -40 kJ/mol involved chemical

785

adsorption. Positive ΔS0 values indicate good affinity of adsorbed species towards the

786

adsorbent and increased randomness at the solid-solution interface associated with structural

787

changes at the adsorption sites during the adsorption process [99,100].

ed

M

an

us

cr

ip t

775

All studies, except the one on F adsorption at the highest temperature of 55oC on

789

activated carbon modified by H2O2/KMnO4 oxidation [59] presented in Table 2 had negative

790

ΔG0 values. The negative values indicate good affinity of F for the adsorbents and the

791

adsorption process was spontaneous. The small positive value obtained for the activated

792

carbon was attributed to the disruption of the MnO2 formed at the carbon surfaces and

793

increase F solubility at high temperature [59]. The ΔS0 values were mostly positive, again

794

indicating a strong affinity of F towards the adsorbents. The ΔH0 values, however, were

795

positive and negative indicating the endothermic and exothermic nature of adsorption,

796

respectively.

Ac ce

pt

788

797 33

Page 33 of 62

798

6. Fluoride desorption and adsorbent regeneration A suitable adsorbent for F removal should not only have high F adsorption capacity

800

and cost-effectiveness but also be amenable to easy desorption of the adsorbed F and capable

801

of efficient regeneration for multiple reuse of the adsorbent. Also, the desorbing agent should

802

not produce any damage to the adsorbent that causes a reduction in its adsorption capacity.

803

Only adsorbents that can be reused have practical value in real systems for economic and

804

environmental reasons. However, inexpensive adsorbents, such as industrial by-products and

805

natural materials, may be used only once and they need not be regenerated for reuse because

806

usually the costs of desorbing chemicals and regeneration process are higher than that of

807

these adsorbents.

an

us

cr

ip t

799

Desorption of F is carried out by leaching of adsorbed F by acids, bases and salts

809

(Table 3). The selection of a desorbing agent largely depends on the influence of pH on F

810

adsorption and the strength of adsorption. Adsorption on most of the adsorbents decreases at

811

high pHs, and therefore bases having high pHs are commonly used for desorption of F from

812

such adsorbents (Table 3). As explained previously, the desorption of F at high pHs is due to

813

the increased repulsive forces between the negatively charged adsorbent surface and

814

negatively charged F- ions in solution as well as the competition between increased

815

concentration of OH- at high pH values and F- for adsorption. In materials where F adsorption

816

increases with pH, acids have been found to be more effective as desorbing agents. For

817

example, in a hydrous manganese oxide coated alumina [29] and a thio-urea modified

818

amberlite resin [34], F adsorption decreased at lower pH values. Consistent with this

819

adsorption pattern, the percentage of F desorption was found to be higher at low pH values.

820

Adsorption of F by most adsorbents is strong and not easily reversible, partly due to the

821

chemical nature of the adsorption process. Therefore, stronger acids and bases, as well as a

Ac ce

pt

ed

M

808

34

Page 34 of 62

822

longer time of their interaction with the adsorbent are required for efficient F desorption

823

[36,41,50,55]. Water and neutral salts are generally found to be less effective [31,41].

824 825

7. Conclusions Adsorption and precipitation/coagulation methods are widely used for the

827

defluoridation of water. The precipitation/coagulation method has the drawback of not being

828

efficient in waters having low F concentration, and the process produces toxic AlF complexes

829

in the treated water as well as large volumes of waste. The adsorption processes are generally

830

considered attractive because of their effectiveness, convenience, ease of operation,

831

simplicity of design, and for economic and environmental reasons.

an

us

cr

ip t

826

Multivalent metal oxides and hydroxides and layered double hydroxides generally

833

have high F adsorption capacities. Ion exchange resins and fibres, zeolites, and carbon

834

materials have low adsorption capacities on their own, but when their surfaces are modified

835

by incorporating organic functional groups or multivalent metallic cations, the adsorption

836

capacity increased. Some of the adsorbents (e.g. layered double hydroxides, alumina) had to

837

be activated by calcining at high temperatures to increase the adsorption capacity. The above

838

materials have potential for use in industrial countries and in areas where the F concentration

839

in water is very high. Natural and industrial by-products have low adsorption capacities but

840

because they are inexpensive, they have potential for use in rural areas, especially in

841

developing countries.

ed

pt

Ac ce

842

M

832

The pH of water is a dominant factor influencing F adsorption. Generally, F

843

adsorption increases from acidic to near neutral pH and then decreases with increase in pH.

844

Consistent with this pH effect, acids and alkali have been successfully used to desorb F and

845

consequently regenerate the adsorbent for reuse. However, multiple regeneration and reuse

35

Page 35 of 62

846

reduce the adsorptive capacity. Another important factor influencing F adsorption is the type

847

and concentration of other ions present in water. Future research needs to explore highly efficient, low cost adsorbents that can be

849

easily regenerated for reuse over several operational cycles without significant loss of

850

adsorptive capacity and have good hydraulic conductivity to prevent filters clogging during a

851

fixed-bed treatment process. Surface modifications of the adsorbents can be explored to

852

increase the capacity and strength of adsorption without significantly increasing the cost. The

853

majority of studies reported have been conducted in batch trials on synthetic waters. These

854

trials need to be extended using continuous mode column trials which have more relevance to

855

real operating systems on natural waters containing other ions as well.

an

us

cr

ip t

848

856

Acknowledgements

M

857

This study was funded by cooperative Research Centre for Contamination Assessment

859

and Remediation of the Environment (CRC CARE) (Project Number 4.1.12.11/12). We thank

860

Phil Thomas for editing this manuscript.

861

863 864 865

References

Ac ce

862

pt

ed

858

[1] J. Fawell, K. Bailey, E. Chilton, E. Dahi, L. Fewtrell, Y. Magara, Fluoride in Drinking Water, World Health Organization, IWA Publishing, UK, 2006.

866

[2] World Health Organisation, Guidelines for Drinking-water Quality, Volume 2. Health

867

Criteria and Other Supporting Information, 2nd Edition, World Health Organisation,

868

Geneva, 1996.

36

Page 36 of 62

869

[3] US Public Health Service, US Public Health Service Drinking Water Standards, US

870

Government Printing Office, Department of Health Education and Welfare, Washington

871

DC, 1962.

876 877 878

ip t

875

[5] N.J. Chinoy, Effects of fluoride on physiology of animals and human beings, Indian J. Environ. Toxicol. 1 (1991), 17-32.

cr

874

mud, Sep. Purif. Technol. 28 (2002) 81-86.

[6] P.T.C. Harrison, Fluoride in water: a UK perspective, J. Fluor. Chem. 126 (2005) 1448-

us

873

[4] Y. Cengeloğlu, E. Kir, M. Ersöz, Removal of fluoride from aqueous solution by using red

1456.

[7] E. Kumar, A. Bhatnagar, J. Minkyu, W. Jung, S. Lee, S. Kim, G. Lee, H. Song, J. Choi, J.

an

872

Yang, B. Jeon, Defluoridation from aqueous solutions by granular ferric hydroxide

880

(GFH), Water Res. 43 (2009), 490-499.

884 885 886

ed

883

Mater. B137 (2006) 456-463.

[9] S. Vigneswaran, C. Visvanathan, Water Treatment Processes. Simple Options, CRC Press, Inc., Florida, 1995.

pt

882

[8] Meenakshi, R.C. Maheshwari, Fluoride in drinking water and its removal, J. Hazard.

[10] J.J. Murray, Appropriate Use of Fluorides for Human Health, World Health

Ac ce

881

M

879

Organisation, Geneva. 1986.

887

[11] P. Loganathan, M.J. Hedley, N.D. Grace, J. Lee, S.J. Cronin, N.S. Bolan, J.M. Zanders,

888

Fertiliser contaminants in New Zealand grazed pasture with special reference to

889

cadmium and fluorine - a review, Aust. J. Soil Res. 41 (2003) 501-532.

890

[12] Y.H. Li, S. Wang, X. Zhang, J. Wei, C. Xu, Z. Luan, D. Wu, Adsorption of fluoride

891

from water by aligned carbon nanotubes, Materials Res. Bull. 38 (2003) 469-476.

892

[13] X. Liao, B. Shi, Adsorption of fluoride on zirconium(IV)-impregnated collagen fiber.

893

Environ. Sci. Technol. 39 (2005) 4628-4632. 37

Page 37 of 62

894

[14] S.P. Mishra, M. Das, U.N. Dash, Review on adverse effects of water contaminants like

895

arsenic, fluoride and phosphate and their remediation, J. Scientific Industrial Res. 69

896

(2010) 249-253. [15] E.G. Paulson, Reducing fluoride in industrial wastewater, Chem. Eng. 84 (1977) 89-94.

898

[16] S.S. Babu, S. Kumar, P.D. Nemade, T. Roychowdhury, Impacts of fluoride in drinking

ip t

897

water and role of pH in its removal efficiency: a critical review, Crit. Rev. Environ. Sci.

900

Technol. (2012) (in press).

904 905

us

903

defluoridation of water, Chem. Rev. (2012) 2454-2466.

[18] W.W. Choi, K.Y. Chen, The removal of fluoride from waters by adsorption, J. Amer.

an

902

[17] S. Jagtap, M.K. Yenkie, N. Labhsetwar, S. Rayalu, Fluoride in drinking water and

Water Works Assoc. 71 (1979) 562-570.

[19] K.B.P.N. Jinadasa, S.W.R. Weerasooriya, C.B. Dissanayake, A rapid method for the

M

901

cr

899

defluoridation of fluoride-rich drinking waters at village level, Inter. J. Environ. Studies.

907

31 (1988) 305-312.

ed

906

[20] J.P. Padmasiri, C.B. Dissanayake, A simple defluoridator for removing excess fluorides

909

from fluoride-rich drinking water, Intern. J. Environ. Health Res. 5 (1995) 153-160.

910

[21] C. Zevenbergen, L.P. van Reeuwijk, G. Frapporti, R.J. Louws, R.D. Schuiling, A simple

Ac ce

pt

908

911

method for defluoridation of drinking water at village level by adsorption on Ando soil

912

in Kenya, Sci. Total Environ. 188 (1996) 225-232.

913

[22] M.A.M. Sahli, S. Annouar, M. Tahaikt, M. Mountadar, A. Soufiane, A. Elmidaoui,

914

Fluoride removal for underground brackish water by adsorption on the natural chitosan

915

and by electrodialysis, Desalination 21 (2007) 37-45.

916

[23] S. Samatya, U. Yüksel, M. Yüksel, N. Kabay, Removal of fluoride from water by metal

917

ions (Al3+, La3+ and ZrO2+) loaded natural zeolite, Sep. Sci. Technol. 42 (2007) 2033-

918

2047. 38

Page 38 of 62

919

[24] K.R. Bulusu, B.B. Sundaresan, B.N. Pathak, W.G. Nawlakhe, D.N. Kulkarni, V.P.

920

Thergaonkar, Fluorides in water, defluoridation methods and their limitations, J. Inst.

921

Engineers (India) Environ. Engineer. Div. 60 (1979) 1-25.

926 927 928

ip t

925

[26] M. Mohapatra, S. Anand, B.K. Mishra, D.E. Giles, P. Singh, Review of fluoride removal from drinking water, J. Environ. Manage. 91 (2009) 67-77.

cr

924

alum by Nalgonda technique, Indian. J. Environ. Health 17 (1975) 26-65.

[27] A. Bhatnagar, E. Kumar, M. Sillanpää, Fluoride removal from water by adsorption-A

us

923

[25] W.G. Nawlakhe, D.N. Kulkarni, B.N. Pathak, K.R. Bulusu, Defluoridation of water with

review, Chem. Eng. J. 171 (2011) 811-840.

[28] V.K. Gupta, J.M. Carrott, M.M.L.R. Carrott, Suhas, Low-cost adsorbents: Growing

an

922

approach to wastewater treatment – a review. Crit. Rev. Environ. Sci. Technol. 39

930

(2009) 783-842.

M

929

[29] S. Teng, S. Wang, W. Gong, X. Liu, B. Gao, Removal of fluoride by hydrous manganese

932

oxide-coated alumina: Performance and mechanism, J. Hazard. Mater. 168 (2009), 1004-

933

1011.

936 937

pt

935

[30] S. Meenakshi, N. Viswanathan, Identification of selective ion-exchange resin for fluoride sorption, J. Colloid Interface Sci. 308 (2007) 438-450.

Ac ce

934

ed

931

[31] L. Ruixia, G. Jinlong, T. Hongxiao, Adsorption of fluoride, phosphate, and arsenate ions on a new type of ion exchange fiber, J. Colloid Interface Sci. 248 (2002) 268-274.

938

[32] F. Helferich, Ion Exchange, Dover Publication Inc., USA, 1995.

939

[33] E.R. Weiner, Applications of Environmental Aquatic Chemistry. CRC Press, Taylor and

940 941 942

Francis Group, Florida, 2008. [34] I.B. Solangi, S. Memon, M.I. Bhanger, An excellent fluoride sorption behaviour of modified amberlite resin, J. Hazard. Mater. 176 (2010) 186-192.

39

Page 39 of 62

943

[35] A. Sivasamy, K.P. Singh, D. Mohan, M. Maruthamuthu, Studies on defluoridation of

944

water by coal-based sorbents, J. Chem. Technol. Biotechnol. 76 (2001) 717-722.

946 947

[36] S.M. Maliyekkal, A.K. Sharma, L. Philip, Manganese-oxide-coated alumina: A promising sorbent for defluoridation of water, Water Res. 40 (2006) 3497-3506. [37] S. Tokunaga, S.A. Haron, S.A. Wasay, K.F. Wong, K. Laosangthum, A. Uchiumi,

ip t

945

Removal of fluoride ions from aqueous solutions by multivalent metal compounds, Inter.

949

J. Environ. Studies 48 (1995) 17-28.

953 954

us

952

of drinking water, Appl. Geochem. 16 (2001) 531-539.

[39] F. Luo, K. Inoue, The removal of fluoride ion by using metal (III)-loaded Amberlite

an

951

[38] Y. Wang, E.J. Reardon, Activation and regeneration of a soil sorbent for defluoridation

resins, Solvent Extract. Ion Exch. 22 (2004) 305-322.

[40] M.S. Onyango, Y. Kojima, O. Aoyi, E.C. Bernardo, H. Matsuda, Adsorption equilibrium

M

950

cr

948

modelling and solution chemistry dependence of fluoride removal from water by

956

trivalent-cation-exchanged zeolite F-9, J. Coloid Interface Sci. 279 (2004) 341-350.

957

ed

955

[41] M.S. Onyango, Y. Kojima, A. Kumar, D. Kuchar, M. Kubota, H. Matsuda, Uptake of fluoride by Al3+ pretreated low-silica synthetic zeolites: adsorption equilibrium and rate

959

studies, Sep. Sci. Technol. 41 (2006) 683-704.

961

Ac ce

960

pt

958

[42] R.L. Ramos, J. Ovalle-Turrubiartes, M.A.Sanchez-Castillo, Adsorption of fluoride from aqueous solution on aluminium-impregnated carbon, Carbon 37 (1999) 609-617.

962

[43] S.A. Wasay, M.J. Haron, S. Tokunaga, Adsorption of fluoride, phosphate, and arsenate

963

ions on lanthanum-impregnated silica gel, Water Environ. Res. 68 (1996) 295-300.

964 965 966 967

[44] Y. Zhou, C. Yu, Y. Shan, Adsorption of fluoride from aqueous solution on La3+impregnated cross-linked gelatin, Sep. Purif. Technol. 36 (2004) 89-94. [45] P. Trivedi, L. Axe, Long-term fate of metal contaminants in soils and sediments: role of intraparticle diffusion in hydrous metal oxides, in: R. Hamon, M. McLaughlin, E. Lombi 40

Page 40 of 62

968

(Eds.), Natural Attenuation of Trace Element Availability in Soils, CRC Taylor and

969

Francis Group, New York, 2006, pp. 57-71.

975 976 977 978 979

ip t

974

fluoride on activated alumina, Sep. Purif. Technol. 42 (2005) 265-271.

[48] M.G. Sujana, R.S. Thakur, S.B. Rao, Removal of fluoride from aqueous solution by

cr

973

[47] S. Ghorai, K.K. Pant, Equilibrium, kinetics and breakthrough studies for adsorption of

using alum sludge, J. Colloid Interface Sci. 206 (1998) 94-101.

us

972

adsorption on fly ash, Water Air Soil Pollut. 49 (1990) 51-61.

[49] Y. Ku, H. Chiou, The adsorption of fluoride ion from aqueous solution by activated alumina, Water Air Soil Pollut. 133 (2002) 349-360.

an

971

[46] A.K. Chaturvedi, K.P. Yadava, K.C. Pathak, V.N. Singh, Defluoridation of water by

[50] A.M. Raichur, M.J. Basu, Adsorption of fluoride onto mixed rare earth oxides, Sep. Purif. Tech. 24 (2001) 121-127.

M

970

[51] L.E.L. Hammari, A. Laghzizil, P. Barboux, K. Lahlil, A. Saoiabi, Retention of fluoride

981

ions from aqueous solution using porous hydroxyapatite. Structure and conductive

982

properties, J. Hazard. Mater. B114 (2004) 41-44.

pt

984

[52] X. Fan, D.J. Parker, M.D. Smith, Adsorption kinetics of fluoride on low cost materials, Water Res. 37 (2003) 4929-4937.

Ac ce

983

ed

980

985

[53] H. Wang, J. Chen, Y. Cai, J. Ji, L. Liu, H.H. Teng, Defluoridation of drinking water by

986

Mg/Al hydrotalcite-like compounds and their calcined products, Appl. Clay Sci. 35

987

(2007) 59-66.

988

[54] L. Lv, J. He, M. Wei, D.G. Evans, X. Duan, Factors influencing the removal of fluoride

989

from aqueous solution by calcined Mg-Al-CO3 layered double hydroxides, J. Hazard.

990

Mater. B133 (2006) 119-128.

41

Page 41 of 62

991

[55] D.P. Das, J. Das, K. Parida, Physicochemical characterization and adsorption behaviour

992

of calcined Zn/Al hydrotalcite-like compound (HTlc) towards removal of fluoride from

993

aqueous solution, J. Colloid Interface Sci. 261 (2003) 213-220. [56] C. Diaz-Nava, M. Solache-Rios, M.T. Olguin, Sorption of fluoride ions from aqueous

995

solutions and well drinking water by thermally treated hydrotalcite, Sep. Sci. Technol.

996

38 (2003) 131-147.

cation synthetic resin, Sep. Sci. Technol. 37 (2002) 89-103.

cr

998

[57] Y. Ku, H. Chiou, W. Wang, The removal of fluoride ion from aqueous solution by a

us

997

ip t

994

[58] K.M. Popat, P.S. Anand, B.D. Dasare, Selective removal of fluoride ions from water by

1000

the aluminium form of the aminomethylphosphonic acid-type ion exchanger, React.

1001

Polym. 23 (1994) 23-32.

[59] A.A.M. Daifullah, S.M. Yakout, S.A. Elreefy, Adsorption of fluoride in aqueous

M

1002

an

999

solutions using KMnO4-modified activated carbon derived from steam pyrolysis of rice

1004

straw, J. Hazard. Mater. 147 (2007) 633-643.

1007

fluoride ions onto carbonaceous materials, J. Colloid Interface Sci. 275 (2004) 35-39.

pt

1006

[60] I. Abe, S. Iwasaki, T. Tokimoto, N. Kawasaki, T. Nakamura, S. Tanada, Adsorption of

[61] Y.H. Li, S. Wang, X. Zhang, J. Wei, C. Xu, Z. Luan, D. Wu, B. Wei, Removal of

Ac ce

1005

ed

1003

1008

fluoride from water by carbon nanotube supported alumina, Environ. Technol. 24 (2003)

1009

391-398.

1010 1011 1012 1013

[62] P.D. Nemade, A.V. Rao, B.J. Alappat, Removal of fluorides from water using low cost adsorbents, Water Sci. Tech. Water Supply 2 (2002) 311-317. [63] N. Viswanathan, S. Meenakshi, Selective fluoride adsorption by a hydrotalcite/chitosan composite, Appl. Clay Sci. 48 (2010) 607-611.

1014

[64] C.S. Sundaram, N. Viswanathan, S. Meenakshi, Fluoride sorption by nano-

1015

hydroxyapatite/chitin composite, J. Hazard Mater. 172 (2009) 147-151. 42

Page 42 of 62

1016

[65] C.S. Sundaram, N. Viswanathan, S. Meenakshi, Uptake of fluoride by nano-

1017

hydroxyapatite/chitosan, a bioinorganic composite, Bioresource Technol. 99 (2008)

1018

8226-8230.

1019

[66] W. Ma, F. Ya, M. Han, R. Wang, Characteristics of equilibrium, kinetics studies for adsorption of fluoride on magnetic-chitosan particle, J. Hazard. Mater. 143 (2007) 296-

1021

302.

[67] S.P. Kamble, S. Jagtap, N.K. Labhsetwar, D. Thakare, S. Godfrey, S. Devotta, S.S.

cr

1022

ip t

1020

Rayalu, Defluoridation of drinking water using chitin, chitonsan and lanthanum-

1024

modified chitosan, Chem. Eng. J. 129 (2007) 173-180.

us

1023

[68] M.G. Sujana, H.K. Pradhan, S. Anand, Studies on sorption of some geomaterials for

1026

fluoride removal from aqueous solutions, J. Hazard. Mater. 161 (2009) 120-125.

1027

[69] M. Agarwal, K. Rai, R. Shrivastav, S. Dass, Defluoridation of water using amended

1031 1032 1033

M

ed

1030

[70] M. Srimurali, A. Pragathi, J. Karthikeyan, A study on removal of fluorides from drinking water by adsorption onto low-cost materials, Environ. Pollut. 99 (1998) 285-289. [71] M. Islam, R. Patel, Thermal activation of basic oxygen furnace slag and evaluation of its

pt

1029

clay, J. Cleaner Production 11 (2003) 439-444.

fluoride removal efficiency, Chem. Eng. J. 169 (2011) 68-77.

Ac ce

1028

an

1025

[72] A. Tor, N. Danaoglu, G. Arslan, Y. Cengeloglu, Removal of fluoride from water by

1034

using granular red mud: Batch and column studies, J. Hazard. Mater. 164 (2009) 271-

1035

278.

1036 1037 1038 1039

[73] V.K. Gupta, I. Ali, V.K. Saini, Defluoridation of wastewaters using waste carbon slurry, Water Res. 41 (2007) 3307-3316. [74] R. Piekos, S. Paslawska, Fluoride uptake characteristics of fly ash, Fluoride 32 (1999) 14-19.

43

Page 43 of 62

1040 1041 1042

[75] Y.D. Lai, J.C. Liu, Fluoride removal from water with spent catalyst, Sep. Sci. Technol. 31 (1996) 2791-2803. [76] Y. Tang, X. Guanc, J. Wang, N. Gaob, M.R. Mcphaild, C.C. Chusuei, Fluoride adsorption onto granular ferric hydroxide: effects of ionic strength, pH, surface loading,

1044

and major co-existing anions, J. Hazard. Mater. 171 (2009) 774-779.

ip t

1043

[77] Y. Zhou, R.J. Haynes, Sorption of heavy metals by inorganic and organic components of

1046

solid wastes: Significance to use of wastes as low-cost adsorbents and immobilising

1047

agents, Crit. Rev. Environ. Sci. Technol. 40 (2011) 909-977.

us

1048

cr

1045

[78] N. Chubar, Physico-chemical treatment of micropollutants: Adsorption and ion exchange, in: J. Virkutyte, R.S. Varma, V. Jegatheesan, (Eds.), Treatment of

1050

Micropollutants in Water and Wastewater, IWA publishing, London, 2010, pp. 165-203. [79] W.H. Zou, R.P. Han, Z.Z. Chen, J. Shi, H.M. Liu, Characterization and properties of

M

1051

an

1049

manganese oxide coated zeolite as adsorbent for removal of copper (II) and lead (II) ions

1053

from solution, J. Chem. Eng. Data 51 (2006) 534-541.

1056 1057 1058 1059 1060 1061 1062 1063

the accumulation of arsenic in lake sediments, Water Res. 9 (1985) 1029-1032.

pt

1055

[80] T. Takamatsu, M. Kawashima, M. Koyama, The role of Mn2+ rich manganese oxide in

[81] E.N. Peleka, E.A. Deliyanni, Adsorptive removal of phosphates from aqueous solutions,

Ac ce

1054

ed

1052

Desalination 245 (2009) 357-371. [82] K. Goh, T. Lim, Z. Dong, Application of layered double hydroxides for removal of oxyanions: A review, Water Res. 42 (2008) 1343-1368. [83] K. Vaaramma, J. Lehto, Removal of metals and anions from drinking water by ion exchange, Desalination 155 (2003) 157-170. [84] S. Wang, Y. Peng, Natural zeolites as effective adsorbents in water and wastewater treatment, Chem. Eng. J. 156 (2010) 11-24.

44

Page 44 of 62

1064 1065 1066 1067

[85] C. Diaz-Nava, M.T. Olguin, M. Solache-Rios, Water defluoridation by Mexican heulandite-clinoptilolite, Sep. Sci. Technol. 37 (2002) 3109-3128. [86] P. Chingombe, B. Saha, R.J. Wakeman, Surface modification and characterisation of a coal-based activated carbon, Carbon 43 (2005) 3132-3143. [87] C.Y. Yin, M.K. Aroua, W.M.A.W. Daud, Review of modifications of activated carbon

1069

for enhancing contaminant uptakes from aqueous solutions, Sep. Purif. Technol. 52

1070

(2007) 403-415.

1074 1075

cr

us

1073

Purif. Technol. 70 (2010) 329-337.

[89] A. Bhatnagar, M. Sillanpää, A review of emerging adsorbents for nitrate removal from water. Chem. Eng. J. 168 (2011) 493-504.

an

1072

[88] R. Mahmudov, C.P. Huang, Perchlorate removal by activated carbon adsorption. Sep.

[90] C.S. Sundaram, N. Viswanathan, S. Meenakshi, Defluoridation chemistry of synthetic

M

1071

ip t

1068

hydroxyapatite at nano scale: equilibrium and kinetic studies, J. Hazard Mater. 155

1077

(2008) 206-215.

ed

1076

[91] A. Bhatnagar, V.J.R. Vilar, C.M.S. Botelho, R.A.R. Boaventura, A review of the use of

1079

red mud as adsorbent for the removal of toxic pollutants from water and wastewater,

1080

Environ. Technol. 32 (2011) 231-249.

1082 1083 1084

Ac ce

1081

pt

1078

[92] D. Mohan, C.U. Pittman Jr, Arsenic removal from water/wastewater using adsorbents-A critical review. J. Hazard. Mater. 142 (2007) 1-53. [93] H. Genc-Fuhrman, H. Bregnhoj, D. McConchie, Arsenate removal from water using sand-red mud columns, Water Res. 39 (2005) 2944-2954.

1085

[94] C. Zhu, Z. Luan, Y. Wang, X. Shan, Removal of cadmium from aqueous solutions by

1086

adsorption on granular red mud (GRM), Sep. Purif. Technol. 57 (2007) 161-169.

45

Page 45 of 62

1087

[95] I.A.M. Yunusa, P. Loganathan, S.P. Nissanka, V. Manoharan, M.D. Burchett, C.G.

1088

Skilbert, D. Eamus, Application of coal fly ash in agriculture: a strategic perspective,

1089

Crit. Rev. Environ. Sci. Technol.42 (2012) 559-600. [96] W. Nigussie, F. Zewge, B.S. Chandravanshi, Removal of excess fluoride from water

1091

using waste residue from alum manufacturing process, J. Hazard. Mater. 147 (2007)

1092

954-963.

ip t

1090

[97] R.H. Masue, T.A. Loeppert, T.A. Kramer, Arsenate arsenite adsorption and desorption

1094

behaviour on coprecipitated aluminium:iron hydroxides, Environ. Sci. Technol. 41

1095

(2007) 837-842.

us

cr

1093

[98] L. Yan, Y. Xu, H. Yu, X. Xin, Q. Wei, B. Du, Adsorption of phosphate from aqueous

1097

solution by hydroxyl-aluminum, hydroxyl-iron and hydroxyl-iron-aluminium pillared

1098

bentonites, J. Hazard. Mater. 179 (2010) 244-250.

1102

M

ed

1101

1989.

[100] N.Y. Mezenner, A. Bensmaili, Kinetics and thermodynamic study of phosphate adsorption on iron hydroxide-eggshell waste, Chem. Eng. J. 147 (2009) 87-96.

pt

1100

[99] S.D. Faust, O.M. Aly, Adsorption Processes for Water Treatment, Butterworths, Boston,

Ac ce

1099

an

1096

46

Page 46 of 62

ip t cr

an

Initial (I), pH; Adsorption (Ads) capacity (mg/g) and Temperat- Equilibrium other results. Maximum (max) o (E) ure ( C) concentration (mg/L); Adsorbent concentration (g/L); column ht, height; d, diameter; FL, flow rate Metal oxides and hydroxides 3-12; I, 1-100; Max ads pH 3-8, then decreased with 10, 25 2 increased pH to 12; Langmuir ads max (pH 6-7) 10oC, 3.68; 25oC, 5.97

Best kinetic model to fit data;

Best equilibrium model to fit data

Reference

Pseudo-first order, Bangham model Pseudo-first, pseudosecond, diffusion

Langmuir

[7]

Langmuir

[29]

Pseudo-first, pseudosecond, diffusion

Langmuir

[36]

Equilibrium (equil)

M

Adsorption method: batch (B), column (C); Water type: synthetic (S), wastewater (W)

B; S

Manganese oxide coated alumina

B, C; S

Manganese oxide coated alumina

B, C; S

B, 3-12; 25 C, 5.2; 25

Ac c

Granular ferric hydroxide

ep te

d

Adsorbent

us

Table 1. Characteristics of adsorptive removal of fluoride from water by different adsorbents

B, 3-12; 30 C, 7; 30

B, I, 6.242.1; C, I 5; 140; 50cm ht, 2.4 cm d, FL 2.39 m3/m2h B, I, 2.5-30 C, 3.56; 0.5x0.028m d, 0.3mht, FL 2.19 m3/m2h

B, Max ads pH 4-6, decreased with increased pH from 6 to 12. Langmuir ads max 7.1 at pH 5.2 C, breakthrough point 669 bed volume B, max ads pH 4-7, lowest at pH 12; Langmuir ads max 2.85 at pH 7 C, Bed saturation F concentration 1.25 g/L 47

Page 47 of 62

ip t cr

Table 1. Characteristics of adsorptive removal of fluoride from water by different adsorbents (cont.d)

C, 3.56; 0.5x0.028m d, 0.3mht, FL 2.19 m3/m2h B, I, 2.5-14; 4-40 C, 5; 200 550 mm ht, 51mm d, FL 20,30mL/min -; 12.5

Pseudo-first, pseudosecond, diffusion

Langmuir

[36]

Surface adsorption, diffusion

Langmuir and Freundlich

[47]

Max ads pH 5-7 Langmuir ads max 16.3 at pH 6

Pseudosecond order

[49]

First 5 min most F ads; 40 min equil -

Langmuir and Freundlich Langmuir -

[51]

Langmuir

[52]

us

B, max ads pH 4-7, lowest at pH 12; Langmuir ads max 1.08 at pH 7

B, I, 2.5-30

C, Bed saturation F concentration 0.47 g/L max ads pH 7 Langmuir ads max 2.41 at pH 7

B; S

Rare earth oxides

B; S

3-11; 29

E 1-30; 1-8

Max ads pH 6-6.5 Langmuir ads max 196 at pH 6.5

Hydroxyapatite (HA)

B; S

I, 19-19000; 10

Hydroxyapatite (HA)

B; S

-; Room temp 6; Room temp

At initial F concentration of ≤190 mg/L, 60-80% F removed by porous HA and 30-35% by crystalline HA >90% F adsorption; Langmuir ads max 4.54 compared to fluorspar 1.79, activated quartz 1.16, calcite 0.39, quartz 0.19

M

Activated alumina

B, 4-10; Room temp C, 7; room temp 4-11; 30

96% removal

d

ep te

B, C; S

B, 3-12; 30 C, 7; 30

Ac c

Activated alumina

B, C; S

an

Activated alumina

I;2.5 x 10-5-6 x 10-2; 17

Pseudosecond order

[50]

48

Page 48 of 62

ip t cr

Table 1. Characteristics of adsorptive removal of fluoride from water by different adsorbents (cont.d)

7; 25

B; S

5-10; 30

Zn/Al LDH calcined (450oC) Na/Mg/Al LDH calcined (500oC)

B; S

4-10; 30

B; S, well water

S: 5,7,9; Room temp W:8.4,8.5 Room temp

us

B; S

an

Mg/Al LDH calcined and uncalcined Mg/Al LDH calcined (200-800oC)

B, I, 10.4-104 Ads max pH 4.7-7.3, Langmuir ads max 0.1 3.8 at pH 7 C, I, 72; 4; 9 cm ht, 1 cm Column ads capacity 3.4 d, FL 0.5 mL/min Layered double hydroxides (LDH) E, 0-70; Calcination increased F ads (130oC 2.5 optimum). Ads capacity 35 at equilibrium F concentration 70 mg/L E, 0-250; Max ads pH 6.0; Langmuir ads max 213 1 at pH 6. Ads highest for calcination temp 500oC. Ads Mg/Al LDH>Ni/Al, Zn/Al LDHs, Mg/Al molar ratio 2:1 best E, 2-60; Max ads pH 6; Langmuir ads max 17 at 1 pH 6. Max ads 13.43 for adsorbent dose 0.2g/L, F concentration 10 mg/L, pH 6 S: I, 5; Distribution coefficient between solid 10 and solution, S: pH 5, 3501; pH 7, W: I, 5.9, 6.9; 1047; pH 9, 1155 10 W: 414-437

M

B 4-9; 20 C 7.5; 20

Ac c

Pseudo-firast order

Langmuir

[43]

adsorption reached max at 15 min -

-

[53]

Freundlich

[54]

ads reached equil in 4 h

Langmuir

[55]

80-97% F removal in 1h

-

[56]

d

B; W C; W

ep te

La impregnated silica gel

49

Page 49 of 62

ip t cr

Chelating resin (CR), anion exchange resin (AER) Metals loaded Amberlite resin

B; S, field water

3-11; 30

B, C; Field water

B: 1-8; 30 C: 6.0; 30

Al-Amberlite resin

B, C; S

B: E, 2-13; 10 C: I, 16; 2; 0.8 mm d, 50 mm ht, FL,1 mL/min I, 2-10; 20

ep te

B: E, 0-60; 1.6 C: I, 7.9; 2; 8 mm d, FL 6 mL/h -; B: 4-9.1; 30 C: 5.5, 6.7 C: I, 40; 30 20-50; 2cm d, 16cm ht, FL 280700 mL/h

Ac c

an

B: 1-10; 25 C: 7; 25

M

B, C; S

Ion exchange resins and fibres B: ads max at pH 7 for TUA and pH 910 for amberlite (A). At pH 7, ads capacity of TUA three times that of A. Langmuir max for TUA at pH 7, 61. C: column ads capacity 50 pH had no significant effect on ads; Langmuir ads max at pH 7: CR 1.3, AER 1.5. At low F concentration CR removed 30% more F than AER. Field water: CR had higher F removal B: Langmuir ads max at pH 7, resin with La, Ce, Al, 25; Fe, 49; Y, 19. pH for ads max: Ce 4-7, Fe 3, Al 5-9, L 3-7 C: column ads capacity for La resin at pH 6, 20. Bed saturation 300 bed vol. B: max ads at pH 4-7, low ads pH 9, Langmuir ads max 4.6 at pH 4 C: column ads capacity at pH 5.5 and 6.7 were 1.13 and 0.72 (FL 460 mL/h)

d

Thio-urea modified Amberlite (TUA)

us

Table 1. Characteristics of adsorptive removal of fluoride from water by different adsorbents (cont.d)

Ads reached equil in 30 min

Langmuir

[34]

Pseudosecond order

Langmuir

[30]

Ads equil time: La 2 h; Ce 1 h; Fe, Al, Y > 5 h

Langmuir

[39]

Pseudosecond order

Langmuir

[57]

50

Page 50 of 62

ip t cr

B: 2-6; 25 C: -; 25

B: E, 0-60; 4 C: I, 10; 1 1.5cm d, 5cm ht, FL 2.5-3.5 mL/min E, 0-100; 5

B, Tap water

2-9; 30

Al, La, Zr loaded natural zeolites (Z) Al loaded four synthetic zeolites Al, La loaded synthetic zeolites

B; S, tap water (T)

-; 30

B; S

2-11; 25

S: I, 1-20; 2 T: I, 2.9; 6 I, 5-80; 0.05-0.2

B; S, field water (F)

S: 3.5-9; 20-40 F: 7.4; 30

S: I, 10-80; 2 F: I, 3.3, 4; 1-4

ads max at pH 2.6-6.9. Ads capacity 30 for 5 g adsorbent at E = 100 mg/L Zeolites S: Langmuir ads max : Zr (Z) 3.4-4.1, La(Z) 2.4-2.6, Al(Z) 2.0-2.4 T: F removal(%): Zr(Z) 91.1, La(Z) 90.4, Al(Z) 89.7 Ads max pH 4-8; F ads at equilibrium concentration 40 mg F/L and pH 4-6 was 6-16 S: ads max at pH 6-9; Langmuir ads max 20oC, Al(Z) 34, La(Z) 45.At 40 mg/L equilibrium conc., ads capacity 16 vs 8 for Al(Z) vs La(Z). (F): F conc reduction more by Al(Z) than La(Z)

ep te

d

Al-chelating porous anion exchanger

Ac c

B: ads max at pH 3.0 and decreased with increased pH to 6; Freundlich adsorption constant PO43- > AsO43- > FC: column ads capacity for PO43, AsO43, F- 156, 96, 45, respectively

an

B, C; S

M

Ion exchange fibre

us

Table 1. Characteristics of adsorptive removal of fluoride from water by different adsorbents (cont.d) Pseudosecond order

Freundlich

[31]

1 h equil time (95% adsorption)

-

[58]

-

Langmuir

[23]

Elovich

RedlichPeterson

[41]

-

RedlichPeterson

[40]

Equil reached in 5 min

51

Page 51 of 62

ip t cr

Table 1. Characteristics of adsorptive removal of fluoride from water by different adsorbents (cont.d)

4.6-9.2; 25

E, 0-20

B; S

3-11; 25

E, 0-16

B; S

3-11; 25

-

B; S

2-12; -

I, 4-30; -

B; S

us

B; S

ep te

3-7; 25

Carbon materials Ads max at pH 2, then decreased with pH. Langmuir ads max at 25oC, 15.9. Ads decreased with temperature increase Ads capacity of bone char increased with pH. Freundlich adsorption constant highest for bone char and lowest for carbon black Ads max at pH 7; ads capacity at equilibrium concentration 10 mg F/L for activated carbon, γ Al2O3, and ACNT: 0.32, 3.7, and 4.1, respectively Ads max at pH 6-9; ads max of 9.6 at pH 6 and calcined temp 450oC for 0.5 mg/L adsorbent and initial concentration 6 mg F/L Ads decreased from pH 2 to pH 12; Fish bone charcoal had the highest F ads

an

E, 5-20; -

M

2-10; 25, 45, 55

d

B; S

Ac c

KMnO4 modified activated carbon Bone char, activated carbon, carbon black Aligned carbon nanotubes (ACNT) Alumina loaded CNT (calcined, 250-1050oC) Wood, animal , fish bone activated charcoal Al impregnated activated carbon

I, 0.5-15 E, 0-6.5; -

Ads decreased from pH 3 to 7. Calcining at 300oC gave the highest ads among 300-1000oC. Langmuir ads max 1.07, plain carbon 0.49

Pseudosecond order

LangmuirFreundlich

[59]

-

Freundlich

[60]

-

Freundlich

[12]

Equil reached at 20 h

Freundlich

[61]

Most of F removed at 2 h

Langmuir, Freundlich

[62]

-

Langmuir, Freundlich

[42]

52

Page 52 of 62

ip t cr

Table 1. Characteristics of adsorptive removal of fluoride from water by different adsorbents (cont.d)

Chitin/nanohydroxyapatite composite Chitosan/nano hydroxyapatite composite

B; S

3-11; 30

I, 6-12; 2

B; S

3-11; 30

I, 9-15; 2

Magnetic (Fe)-chitosan

B; S

S: 5-9; 30

I, 5-140; 1

La loaded chitosan

B; S, field water (F)

S: 5-9; 30 F: 7; 30

S: E, 0-15; F: I, 10.2; -

Natural chitosan Zr loaded collogen

B; S B; S

2-10; 20 3.5-11; 30

-

Natural materials Ads max at pH 3, decreased to pH 11. At pH 7, 10 mg F/L, 2 g adsorbent/L, composite ads 1.3; LDH alone 1.0, chitosan alone 0.05; Langmuir ads max 1.9 for composite Ads max at pH 3, decreased with increased pH to pH 11; Langmuir ads max 8.4 at pH 7 Ads max at pH 3, decreased with increased pH to pH 11; At pH 7, 10 mg F/L, adsorbent 5 mg/L, composite ads capacity 1.56, hydroxyapatite 1.30, chitosan 0.05; Langmuir ads max 2.04 pH no significant effect. Ads higher than activated alumina. Langmuir twosite max ads 24

us

I, 9-15; 2

an

3-11; 30

M

B; S

Ac c

ep te

d

Chitosan/ LDH composite

I, 19-95; 1

S: Ads max pH 6.7; max ads 5.5 at pH 6.7 and equilibrium conc.15 mg F/L F: ads capacity 1 compared to 2 for distilled water at equilibrium conc. 8 mg F/L Max ads at pH 6; Langmuir ads max 1.39 at pH 6 Max ads pH 5-8, drastic decrease from pH 9 onwards. Langmuir ads max 2.18 at pH 5-8

PseudoLangmuir second order, diffusion

[63]

PseudoLangmuir second order, Freundlich diffusion PseudoLangmuir second order, diffusion

[64]

Langmuir Equil at 90 min; Pseudo- (one and second order two sites)’ Bradley Pseudo-first order, diffusion

[66]

Equil at 5 min Equil at 500 min

[22]

Langmuir, Freundlich Langmuir

[65]

[67]

[13]

53

Page 53 of 62

ip t cr

Table 1. Characteristics of adsorptive removal of fluoride from water by different adsorbents (cont.d)

B; S, field water

Al, Fe, Ca loaded silty clay (C)

B; S

S: E, 5-40; 2 F: I, 10.25; 14 I, 10; 100

Soil with high Fe oxide (calcined 500oC)

B, C; S

S: 2.5-10; 32 F: 7.75; 32 Al, Ca-C, C alone: 7-8; Fe-C 2.7-3.0; 30 B: 7; 25 C: 7; 25

Volcanic ash soil

B; S

2.5-9.5; 32

I, 5; 1-8

Increased pH, decreased ads, Be: 46 to 29%, Ka: 38 to 5% for 2g/L adsorbent

ep te

B: max ads 1 at equilibrium concentration 60 mg/L

-; 18

B:I, 4.8-95; 33 C: I, 4; 47; 50 mL column, FL 0.2 mL/min I, 0-50; 40

Ac c

Bentonite B; (Be), S kaolinite (Ka)

pH 2-5, > 90% ads, then decreased with increased pH to pH 12 (17%); Langmuir ads max 21.3 at pH 5-7 S: ads max at pH 5-6 and 3-5; Langmuir ads max 12-15 at pH 5 F: 3 repeated stages of ads reduced F concentration to <0.5 mg/L % F removal: C alone 25, Al-C 78-94’ Ca-C 32-52, Fe-C 94-98

us

E: 0-50; 4

an

2-12; 29

M

B; S

d

La loaded cross-linked gelatin Laterites with Ni

C: F concentration reached 1 mg/L after 120 pore volumes. Column ads capacity 0.15 At equilibrium concentration of 19 mg/L, ads capacity was 2.9 Langmuir ads max 5.5

Equil at 40 min; pseudofirst order Pseudo-first order, diffusion

Langmuir

[44]

Langmuir

[68]

-

-

[69]

-

Freundlich

[38]

Equil in ~2h Langmuir at low conc, >24 h at high conc -

[21]

[70]

54

Page 54 of 62

ip t cr

Table 1. Characteristics of adsorptive removal of fluoride from water by different adsorbents (cont.d)

us

B; S C; S

B:2.5-7.3; 25 C: 4.7; 25

B; S

1-10; -

B: I, 5-150; 2.5 C: I, 5;10, 0.635 cm2 area,15cm ht, FL 2 mL/min I, 100-1000; 1-8.4

B; S C; W

B:2- 11; 25 C: -; 25

Fly ash (class F, 9.1% CaO)

C; S

Fly ash (2.22 % CaO) Alum sludge (calcined and uncalcined)

B; S B; S

10.1; 20

2-9.5; 30 3.5-8.8; 32

Industrial by-products Ads increased from 21 to 93% as pH increased from 2 to 7, then nearly constant; Langmuir ads max at 25oC, 4.6 and 45oC, 8.1 B: ads max at pH 4.7; Langmuir ads max 8.92 at pH 4.7 C: column total ads capacity 2.05 (0.64 by batch trial for initial F, 5 mg/L

an

I, 1-50; 5

M

2-10; 25

Ads max at pH 5.5; ads capacity A 4.8, UA 1.0 at equilibrium concentration 20 mg/L; Langmuir ads max A 6.3, UA 3.1at pH 5.5 B: ads max at pH 7.58; Langmuir ads max 4.3 at pH 7.5

d

ep te

Activated red mud (A), unactivated (UA) Waste carbon slurry 450oC (activated)

B; S

Ac c

Basic oxygen furnace slag (BOFS) heated1000oC Granular red mud

I, 1-11;1 C: I, 11; 0.5; 0.9 cm2 area, 3.1 cm ht, FL 1.5 mL/min I, 0-100; 450; 40 cm ht, 4.5 cm d, FL 2 mL/h E, 0-3; 20 E, 0-15; 0.5-16

C: breakthrough column ads capacity 4.16 F concentration in effluent reached 0 mg/L after 120-168 h Ads max at pH 6.5; Langmuir ads max 20 at pH 6.5 Ads max at pH 6; calcined higher ads capacity than uncalcined; Langmuir ads max 5.39 at pH 6

Equil at 35 min; pseudofirst order

Langmuir

[71]

Equil at 6 h; pseudosecond order

Freundlich, RedlichPeterson

[72]

Equil at 2 h

Langmuir

[4]

Equil at 1 h; Pseudo-first order

RedlichPeterson

[73]

-

-

[74]

pseudo-first order pseudo-first order; diffusion

Langmuir

[46]

-

[48]

55

Page 55 of 62

ip t cr

E, 0-40; 10

Ads max (1.6 for initial F 20 mg/L, pseudo-first adsorbent 10 g/L) at pH 2, then continue order; 70% to decrease. Max ads 28 ads in 10min

-

[75]

d

M

an

us

2-9.5; 25

ep te

B; S

Ac c

Spent catalyst

56

Page 56 of 62

ΔH0 (kJ/mol)

ΔS0 (J/mol/0C)

Reference

22.7

8.4

[71]

8.23

0.1

[63]

Granular ferric hydroxide Geomaterials

-39.2 25 -5.8 to -15.9 32 to 62

7.34

15.8

[7]

-0.1 to -30

-3 to -10

Nanohydroxyapatite/ chitin composite Nanohydroxyapatite/ chitosan composite

-7.1 30

-7.0 40

-6.8 50

11.8

-3.8 30

-3.4 40

-3.4 50

8.6

Modified activated carbon Waste carbon slurry

-3.6 -0.9 25 45 -25.4 -26.5 25 35

Chelating ion exchange resin

-5.7 30

Calcined Zn/Al LDH

-25.0 -25.6 30 40

Fly ash

-1.3 30

-1.6 40

-26.3 50

-2.0 50

us

[64] [65]

-48.6

-15.1

[59]

7.3

11

[73]

-2.4

4.2

[30]

-5.7

6.4

[55]

6.5-6.9

17.1

[46]

an

-5.6 50

1.5

[68]

1.6

M

0.9 55 -27.6 45

ed

-5.5 40

Ac ce

Basic oxygen furnace slag LDH/chitosan composite

pt

Adsorbent

ip t

ΔG0 (kJ/mol) ----------------------Temperature (oC) -0.38 -0.72 -1.94 25 35 45 -6.81 -6.74 -6.72 30 40 50

cr

Table 2. Thermodynamic parameters for fluoride adsorption by different adsorbents

57

Page 57 of 62

Table 3. Fluoride desorption and adsorbent regeneration Desorption/regeneration reagent (BV, bed volume; d and ht, column internal diameter and height; FL, flow rate

Results

Reference

Amberlite modified with thio-urea (ATU) (B) and (C)

(B) 100 g F-loaded ATU shaken 40 min with 5 mL HNO3, H2SO4, HCl, NaOH (0.01, 0.001 M), 25oC. (C) 2 g ATU, 0.8 mm d, 50 mm ht, FL 1 mL/min, 25oC

(B) maximum desorption [34] (99.9%) with 0.01 M HCl. Next best 0.01 M H2SO4 (77.6%). (C) maximum recovery 92.3% (0.01 M HCl), 35 mL volume used

Manganese oxide coated alumina (MOCA) (B)

0.25 g MOCA adsorbed with F, 50 85% desorption at pH 12. Little mL water with pH adjusted to 3.0- F desorbed at pH < 10 13.5, shaken 24 h, 25oC

Granular red mud (GRM) (C)

10 g GRM, 0.635 cm2 cross sectional area, 10 cm ht, 10 mL 0.2 M NaOH, FL 1 ml/min, 25oC

MOCA and activated alumina (AA) (C)

0.028 m d, 0.1, 0.2 and 0.3 m ht, 2.5% NaOH. 3 cycles of regeneration/adsorption

Al treated zeolite (B)

0.1 g F loaded adsorbent, 50 mL acetic acid, water, NaHCO3, NaOH shaken for 1 d. 4 desorption steps. 25oC

AA (C)

50.8 mm d, 550 mm ht, repeated cycles of regeneration/reactivation

ip t

Adsorbent (batch (B), column (C) method used)

us

cr

[29]

[72]

94.6%F desorbed for MOCA, 91%F for AA in Ist cycle. 2nd cycle no reduction in adsorption, 3rd cycle 5% reduction

[36]

NaHCO3 most effective (63%desorption), Acetic acid least effective (5%). 4 zeolites with NaHCO3 desorption, 1st step 61-67%, total 75-95%

[41]

Loss of 2% F uptake capacity after 5 cycles of regeneration

[47]

Ac ce

pt

ed

M

an

Adsorption (mg/g) decreased from 2.05 to 0.82 in 4 regeneration/adsorption cycles. %desorption from 87 to 46

Zr loaded collagen fibre (B)

0.1 g, 100 mL water with pH adjusted with NaOH, HNO3, 0.5 h shaking, 30oC

pH < 9 very little desorption, pH [13] 11.5 97% desorption

La loaded gelatin (B)

1st wash 1 M NaOH, then water wash or acid wash with 1:1 HNO3 to neutral pH

After adsorption/regeneration process for 3 times, maximum adsorption capacity reduced from 98.5 to 82.3%

[44]

Calcined Zn/Al LDH (B)

NaOH (0.001-0.04 M), 1 g adsorbent treated with maximum F/L of NaOH for 6 h, 30oC.

F desorption increased from 5.12 to 8.55 mg/g (100%). Desorption increased with NaOH concentration

[55]

Ion exchange fibre (B), (C)

(B) 0.1 g adsorbent F loaded, mixed with 25 mL 0.1 M NaCl, NaOH, HCl for 2 h, 25oC. (C) 1.5 cm d, 5

(B) Desorption 20% for HCl, 60% for NaCl, 80% for NaOH. (C) 100% desorbed with 5

[31]

58

Page 58 of 62

mLNaOH

Rare earth oxides (B)

0.2 g adsorbent, 100 mL water, pH adjusted, shaken for 30 min. 29oC

pH<6, F desorbed to ~ 0. pH ~ 12 >95%. Regeneration decreased adsorption 98 to 91%

[50]

Activated alum sludge (B)

F loaded sludge 4g/L, water pH adjusted, 4 h shaking, 30oC

pH 2-7 desorption ~ 0. > pH 7 desorption increased, highest at pH 12

[48]

La loaded silica gel (C)

Adsorbent loaded with F eluted with dilute NaOH (pH 8.5). 1 cm d, 9 cm h, FL 0.5 mL/min, 20oC.

Column was regenerated

[43])

Ac ce

pt

ed

M

an

us

cr

ip t

cm ht, eluted with 0.5 M NaOH, FL 1 mL/min, 25oC

59

Page 59 of 62

List of figures

ip t

Fig. 1. Common technologies for defluoridation of drinking water [8,22,23]

Ac ce

pt

ed

M

an

us

cr

Fig. 2. Mechanisms of F adsorption (, Adsorbent; Me, multivalent metallic cation)

60

Page 60 of 62

Adsorption/Ion Exchange

! ! Most widely used ! ! Medium cost ! ! Low effectiveness; cannot remove F below 5 mg/L because of high solubility product of CaF2; need secondary treatment ! ! Large amounts of chemicals required ! ! Precise control of chemicals additions (frequent testing of feed and treated water) ! ! Costs of chemicals, chemical storage and feeding system ! ! Large volumes of waste sludge; disposal problem ! ! Acid neutralisation of treated water required ! ! Toxic chemicals left in treated water (AlF complexes, SO4)

! ! Most widely used ! ! Can be costly (especially ionexchange resins), but can use low-cost adsorbent (including certain waste materials) ! ! Effective even at low F concentration ! ! Simplicity and flexibility of design ! ! Ease of operation ! ! No waste production ! ! Low selectivity against some anions for adsorption/all anions for ion exchange competing ions ! ! Frequent adsorbent regeneration or replacement required ! ! Granular adsorbent better for good hydraulic flow ! ! Effective, mostly at pH < 7 for adsorption

cr

us

an M

ed

Reverse osmosis

ip t

Precipitation/coagulation

Ac ce

pt

! ! Excellent removal ! ! Very high capital cost. Very high operational (energy) cost ! ! No chemicals required ! ! No waste production ! ! No ion selectively; beneficial nutrients and other contaminants removed together with F ! ! Some membranes pH sensitive ! ! F concentrated residue disposal problem ! ! Water wasted ! ! Clogging, scaling and fouling problems

Electrodialysis

! ! Excellent removal ! ! High capital cost. High operational (energy) cost ! ! No chemicals required ! ! No waste production ! ! No ion selectively; beneficial nutrients and other contaminants removed together with F ! ! Skilled labour required ! ! Polarization problem

Fig. 1.

61

Page 61 of 62

(b)

Hydrogen Bonding

ip t

Ion Exchange

an

us

cr

(a)

Ligand Exchange

(d)

Adsorbent surface chemical modification

Ac ce

pt

ed

M

(c)

Fig 2.

62

Page 62 of 62