Fluoride in Drinking - Water - Status - Issues - and - Solutions

Download as pdf or txt
Download as pdf or txt
You are on page 1of 208

Fluoride in

Drinking
Water

Status,
Issues,
and
Solutions

A.K. Gupta
S. Ayoob

www.ebook3000.com
www.ebook3000.com
Fluoride in
Drinking
Water
Status, Issues,
and Solutions

www.ebook3000.com
www.ebook3000.com
Fluoride in
Drinking
Water
Status, Issues,
and Solutions

A.K. Gupta
S. Ayoob

www.ebook3000.com
CRC Press
Taylor & Francis Group
6000 Broken Sound Parkway NW, Suite 300
Boca Raton, FL 33487-2742

© 2016 by Taylor & Francis Group, LLC


CRC Press is an imprint of Taylor & Francis Group, an Informa business

No claim to original U.S. Government works


Version Date: 20160322

International Standard Book Number-13: 978-1-4987-5653-2 (eBook - PDF)

This book contains information obtained from authentic and highly regarded sources. Reasonable efforts
have been made to publish reliable data and information, but the author and publisher cannot assume
responsibility for the validity of all materials or the consequences of their use. The authors and publishers
have attempted to trace the copyright holders of all material reproduced in this publication and apologize to
copyright holders if permission to publish in this form has not been obtained. If any copyright material has
not been acknowledged please write and let us know so we may rectify in any future reprint.

Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, reproduced, transmit-
ted, or utilized in any form by any electronic, mechanical, or other means, now known or hereafter invented,
including photocopying, microfilming, and recording, or in any information storage or retrieval system,
without written permission from the publishers.

For permission to photocopy or use material electronically from this work, please access www.copyright.
com (http://www.copyright.com/) or contact the Copyright Clearance Center, Inc. (CCC), 222 Rosewood
Drive, Danvers, MA 01923, 978-750-8400. CCC is a not-for-profit organization that provides licenses and
registration for a variety of users. For organizations that have been granted a photocopy license by the CCC,
a separate system of payment has been arranged.

Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are used
only for identification and explanation without intent to infringe.
Visit the Taylor & Francis Web site at
http://www.taylorandfrancis.com

and the CRC Press Web site at


http://www.crcpress.com

www.ebook3000.com
Contents

Preface.......................................................................................................................xi
Acknowledgments............................................................................................... xiii
Authors....................................................................................................................xv

1. Fluoride in Drinking Water: A Global Perspective.................................. 1


1.1 Introduction.............................................................................................. 1
1.2 Drinking-Water Scenario........................................................................1
1.3 Geogenic Pollutants................................................................................. 3
1.3.1 Fluorine: The Chemical Profile..................................................3
1.3.2 Sources of Fluoride...................................................................... 3
1.4 Fluoride in Humans................................................................................5
1.5 Genesis of Fluoride in Groundwater....................................................5
1.6 Summary.................................................................................................. 8
References.......................................................................................................... 8

2. Scenario of Fluoride Pollution.................................................................... 11


2.1 Introduction............................................................................................ 11
2.2 Global Scenario...................................................................................... 11
2.2.1 Asian and African Scenario..................................................... 11
2.2.2 Indian Scenario.......................................................................... 20
2.3 Summary................................................................................................ 21
References........................................................................................................22

3. Dental Fluorosis............................................................................................. 27
3.1 Introduction............................................................................................ 27
3.2 Dental Effects of Fluoride..................................................................... 27
3.2.1 Dental Caries (Tooth Decay).................................................... 28
3.2.2 Prevention of Dental Caries by Fluoride................................ 29
3.2.3 Role of Fluoride in Dental Decay............................................ 29
3.3 Dental Fluorosis: History and Occurrence........................................ 30
3.4 Development of Dental Fluorosis........................................................ 31
3.4.1 Physical Symptoms of Dental Fluorosis................................. 32
3.4.2 Issues of Dental Fluorosis......................................................... 33
3.4.3 Prevalence of Dental Fluorosis................................................ 33
3.5 Summary................................................................................................ 35
References........................................................................................................ 36

v
www.ebook3000.com
vi Contents

4. Skeletal Fluorosis.......................................................................................... 39
4.1 Introduction............................................................................................ 39
4.2 Action of Fluoride on Bone.................................................................. 39
4.3 Fluoride Exposure Level and Skeletal Fracture................................ 40
4.4 Skeletal Fluorosis................................................................................... 41
4.4.1 Crippling Skeletal Fluorosis.....................................................43
4.5 Fluoride Level and Effects Related to Skeletal Fluorosis.................44
4.6 Significance of Other Factors............................................................... 45
4.7 Recent Developments............................................................................ 46
4.8 Summary................................................................................................ 47
References........................................................................................................ 48

5. Stress Effects of Fluoride on Humans....................................................... 51


5.1 Introduction............................................................................................ 51
5.2 Nonskeletal Fluorosis............................................................................ 51
5.3 Fluoride and Cancer.............................................................................. 51
5.4 Fluoride and Gastrointestinal System................................................ 53
5.5 Other Health Effects..............................................................................54
5.6 Summary................................................................................................ 55
References........................................................................................................ 56

6. Fluoride in the Environment and Its Toxicological Effects.................. 59


6.1 Introduction............................................................................................ 59
6.2 Sources of Environmental Exposure.................................................. 59
6.3 Environmental Transport, Distribution, and Transformation........ 60
6.4 Environmental Levels and Human Exposure................................... 61
6.4.1 Fluoride from Dental Products................................................ 62
6.4.2 Fluoride from Food and Beverage...........................................63
6.4.3 Fluoride in Soil........................................................................... 68
6.4.4 Fluoride in Tobacco and Pan Masala...................................... 69
6.4.5 Fluoride from Occupational Exposure................................... 69
6.4.6 TF Exposure................................................................................ 71
6.5 Effects of Fluoride on Laboratory Animals and In Vitro
Systems..............................................................................................71
6.6 Effect of Fluoride on Aquatic Organisms.......................................... 76
6.7 Effect of Fluoride on Plants..................................................................77
6.8 Effect of Fluoride on Animals.............................................................. 78
6.9 Guidelines Values and Standards....................................................... 79
6.10 Summary................................................................................................80
References........................................................................................................ 81

7. Defluoridation Techniques: An Overview............................................... 87


7.1 Introduction............................................................................................ 87
7.2 Coagulation............................................................................................ 87
7.2.1 Lime............................................................................................. 88

www.ebook3000.com
Contents vii

7.2.2 Magnesium Oxide..................................................................... 88


7.2.3 Calcium and Phosphate Compounds..................................... 89
7.3 Co Precipitation of Fluoride................................................................. 90
7.3.1 Alum............................................................................................90
7.3.2 Alum and Lime (Nalgonda Technique).................................. 91
7.4 Adsorption.............................................................................................. 93
7.4.1 Bone and Bone Charcoal........................................................... 93
7.4.2 Clays and Soils........................................................................... 94
7.4.3 Carbonaceous and Other Adsorbents.................................... 96
7.4.4 Alumina...................................................................................... 99
7.4.5 Activated Alumina.................................................................. 101
7.4.6 Other Alumina-Based Adsorbents....................................... 103
7.5 Electrochemical Methods................................................................... 106
7.5.1 Electrocoagulation................................................................... 106
7.5.2 Electrosorption......................................................................... 109
7.6 Membrane Processes........................................................................... 109
7.6.1 Reverse Osmosis...................................................................... 110
7.6.2 Nanofiltration........................................................................... 112
7.6.3 Electrodialysis.......................................................................... 113
7.7 Defluoridation Techniques: A Summary......................................... 114
7.8 Summary.............................................................................................. 115
References...................................................................................................... 117

8. Adsorptive Removal of Fluoride: A Case Study................................... 123


8.1 Introduction.......................................................................................... 123
8.2 Materials and Methods....................................................................... 123
8.2.1 Reagents and Adsorbate......................................................... 123
8.2.2 Synthesis of the Adsorbent..................................................... 124
8.2.3 Instrumentation....................................................................... 124
8.2.4 Characterization of the Adsorbent........................................ 124
8.3 Batch Studies........................................................................................ 125
8.3.1 Effect of Process Parameters.................................................. 125
8.3.2 Equilibrium Studies................................................................ 126
8.3.3 Column Studies........................................................................ 127
8.4 Theoretical and Mathematical Formulations.................................. 128
8.4.1 Adsorption Capacity............................................................... 128
8.4.2 Kinetic Modeling..................................................................... 129
8.4.2.1 Pseudo-First-Order Model....................................... 129
8.4.2.2 Pseudo-Second-Order Model.................................. 130
8.4.2.3 Intra Particle Diffusion Model................................ 131
8.4.2.4 Elovich Equation....................................................... 131
8.4.2.5 Arrhenius Equation.................................................. 132
8.4.3 Elucidation of Rate-Limiting Step......................................... 132
8.4.4 Adsorption Equilibrium and Isotherms.............................. 133
8.4.4.1 Langmuir Isotherm.................................................. 134

www.ebook3000.com
viii Contents

8.4.4.2 Freundlich Isotherm................................................. 135


8.4.4.3 Dubinin–Radushkevich (D–R) Isotherm.............. 136
8.4.4.4 Selection of Best-Fitting Isotherm.......................... 137
8.4.4.5 Natural and Synthetic Systems............................... 137
8.4.4.6 Concentration and Dose Variation Studies........... 138
8.4.5 Factors Influencing Adsorption............................................. 138
8.4.5.1 Adsorbent Dose......................................................... 138
8.4.5.2 Contact Time.............................................................. 139
8.4.5.3 Agitation Rate............................................................ 139
8.4.5.4 Effect of pH and Coexisting Ions........................... 139
8.4.5.5 Temperature............................................................... 140
8.4.5.6 Ionic Strength............................................................ 141
8.4.6 Behavior of Adsorption Columns......................................... 141
8.4.7 Analysis and Modeling of Breakthrough Profile................ 143
8.4.7.1 Hutchins BDST Model............................................. 143
8.4.7.2 Thomas Model........................................................... 144
8.4.7.3 Yoon–Nelson Model................................................. 145
8.4.7.4 Clark Model............................................................... 146
8.4.7.5 Wolborska Model...................................................... 146
8.4.7.6 Bohart and Adams Model....................................... 147
8.4.8 Regeneration............................................................................. 147
8.5 Results and Discussions..................................................................... 148
8.5.1 Characterization of the Adsorbent........................................ 148
8.5.2 Kinetics Studies....................................................................... 149
8.5.2.1 Agitation Rate............................................................ 149
8.5.2.2 Adsorbent Dosage..................................................... 150
8.5.3 Kinetic Profile of Fluoride Uptake........................................ 152
8.5.3.1 Pseudo-First-Order Model....................................... 153
8.5.3.2 Pseudo-Second-Order Model.................................. 153
8.5.3.3 Intra Particle Surface Diffusion Model.................. 154
8.5.3.4 Elovich Model............................................................ 154
8.5.3.5 Arrhenius Equation.................................................. 155
8.5.4 Elucidation of Rate-Limiting Step......................................... 156
8.5.5 Fluoride Removal Mechanism............................................... 158
8.5.6 Isotherm Studies...................................................................... 162
8.5.6.1 Effects of Temperature............................................. 163
8.5.7 Performance Evaluation of ALC in Natural and
Synthetic Systems.................................................................... 163
8.5.7.1 Effect of pH, Ionic Strength, and Temperature.... 164
8.5.7.2 Effects of Other Ions................................................. 166
8.5.8 Column Studies����������������������������������������������������������������������� 167
8.5.8.1 Effect of Process Parameters on Breakthrough.... 168

www.ebook3000.com
Contents ix

8.5.9 Application of Sorption Models............................................ 171


8.5.9.1 Comparison of the Applied Models
(Synthetic Water)....................................................... 175
8.5.9.2 Comparison of the Applied Models
(Natural Water).......................................................... 176
8.5.10 Fluoride Desorption Studies.................................................. 177
8.6 Summary of the Case Study.............................................................. 179
8.7 Conclusions of the Case Study.......................................................... 184
References...................................................................................................... 185
Preface

Of late, geogenic pollutants have been responsible for polluting groundwa-


ter. This has resulted in the pollution of groundwater evolving as a distinct
academic discipline in the arena of environmental science and engineering.
So we felt the scope of books related to environmental engineering needs to
be inclusive of a broader range of such key environmental issues of our time.
The genesis of this book stems from such a feeling.
Textbooks on environmental engineering are generally confined to tradi-
tional isolated subjects that are based on various courses and syllabi. Given
the impending issues related to fluoride in drinking water and associated
human health issues, we believe it is pedagogically undesirable to keep the
issues of fluoride outside our syllabi and classrooms.
The main audience of this book is expected to be the research commu-
nity, faculties, scholars, students (undergraduate, graduate, and postgradu-
ate), and technical professionals who are learning or working in the arena
of environmental science and engineering and dealing with drinking water
quality, supply, and management. For practicing environmental and water
supply, engineers, and professionals working in related areas, this book is
tailored as a guide for providing a solid fundamental understanding of the
crux of global fluoride pollution. For anyone who needs a substantial over-
view of the present global water quality and fluoride-related issues, this
book provides a strong, specific, and updated database.
We have structured the chapters with continuity to get a clear vision and
understanding of the magnitude and gravity of the problem. This book is an
illustration of a research effort that is attempted to facilitate a deeper under-
standing of the issue of fluorosis. To have an in-depth understanding of any
related issues, readers are advised to refer to the advanced literature cited
in our publications. It is our hope this book may serve as an inspiration in
the journey of research to a world of fluoride-free drinking water, which we
believe is not a distant dream.

A. K. Gupta and S. Ayoob

xi
Acknowledgments

We remember the love and prayers from our parents and teachers, without
which our lives would not have been as they are now. We appreciate our
family members for preserving a cheerful home and our scholarly friends
and students for their lovable support.

xiii
Authors

A.K. Gupta earned a PhD in environmental science


and engineering at the Indian Institute of Technology
Bombay, India. Currently, he is a professor in the
Environmental Engineering Division of the Civil
Engineering Department at the Indian Institute of
Technology Kharagpur, India, and he is actively
involved in teaching, research, and consultancy. His
research interests are focused on water treatment,
environmental impact assessment, monitoring, and
modeling of air and water pollution, geogenic pollut-
ant scavenging, and so on. He has more than 60 publications in top-ranking
international journals. He is a renowned technical consultant in the arena of
environmental engineering and has more than 30 completed/ongoing proj-
ects of national and international importance to his credit.

S. Ayoob is a professor in the Department of Civil


Engineering and the principal of TKM College of
Engineering, Kerala, India. He graduated in civil
engineering from the TKM college of engineer-
ing, Kollam, India, and earned master’s and doc-
toral degrees at the Indian Institute of Technology
Kharagpur, India. He also served as course leader
and headed the Department of Health, Safety, and
Environmental Management at the International
College of Engineering and Management, the
Sultanate of Oman (affiliated with the University of Central Lancashire,
United Kingdom) for a short period.

xv
1
Fluoride in Drinking Water:
A Global Perspective

1.1 Introduction
Water is life as it aids in nurturing the lives of all biota. The availability
of clean water has become obliquely central to the quality of human life.
However, there are deep currents that affect the water dynamics of today’s
world. Currently, two-thirds of our planet is covered water; however, the
unfortunate paradox is that in the next decade, bulk of the human popula-
tion will lack access to safe drinking water. This acute shortage of drinking
water may change our traditional perceptions about both the quality and
usage of water in the future. Further, the presence of geogenic pollutants in
groundwater makes this issue more complex. Of late, the presence of fluo-
ride in drinking water has attracted much attention in the scientific world
due to the impending issues associated with human health and well-being.

1.2  Drinking-Water Scenario


In 2014, it was reported by the World Health Organization that around 750
million people from the poor and socially deprived sections of society all over
the world do not have access to improved drinking-water sources. Almost
one-fourth of the people living in rural habitations use and drink untreated
surface water. In 2015, it was estimated that around 550 million do not have
access to improved drinking-water sources.1 It is predicted that in 2035, there
will be a one-third reduction in the per capita drinking-water availability.
By 2025, around 34% of the global population will face acute drinking-water
shortage.1 This sorry state of affairs is reflected in the terribly short supply
of good quality water in many parts of the world. As a result, the global
water supply system is beleaguered at both the demand and supply ends.
The acute scarcity of drinking-water sources and increased competition in

1
2 Fluoride in Drinking Water

the early periods of the twenty-first century, the traditional ways of using
and valuing water have taken a dramatic turn. Thus, the scarcity of water
and the limited access to safe drinking-water sources are predicted to be the
most challenging and crucial environmental issues of the future in preserv-
ing and defining the quality of life on earth.2,3
Groundwater has been perceived to be the safest of all the drinking-water
sources available on the surface of the earth. As a result, half of the global
population blindly relies on groundwater sources for both drinking and sur-
vival. Apart from this, in many regions of the world, groundwater sources
turn out to be the single largest source of supply for drinking. Further, in
many communities, these sources appear to be the only economically viable
option for drinking, as they supply reliable quality water and stable quan-
tity water compared with water from surface sources. Since groundwater
plays such a crucial role in the existence of the majority of the global human
population, its availability, safety, and purity have become issues of critical
concern for many habitations across the globe.3,4
Of late, due to rapid urbanization and industrialization, more xenobiotic sub-
stances are getting diffused into different spheres of the earth. A considerable
portion of these substances rests with the biosphere, of which water sources
occupy a considerable share. Water sources act as “sinks” for many of these pol-
lutants, resulting in drinking-water pollution and water scarcity. This situation
is more serious in developing countries, as they are grappling with acute issues
related to both scarcity and contamination of drinking water. Intrinsically, the
excessive groundwater pumping that is disproportionate to recharge will also
lead to the depletion of water, thus posing challenges to drinking-water supply
systems. Plenty of examples are available to validate this point. In many devel-
oping countries such as India, in urban groundwater sources cater to around
one-half of the water requirements. In rural areas, groundwater sources alone
cater to more than two-thirds of the total water demand, thus resulting in the
situation appearing really critical. The indiscriminate tapping of groundwater
created alarmingly low levels of the water table in many parts of the develop-
ing world.3,5,6 Thus, inadequate access to safe drinking water, on the one hand,
and its ever-increasing intimidation from abundant contaminants, on the other
hand, make the global drinking-water scenario more complex. As a result, the
world is heading toward a water crisis. This is a crisis affecting both the quantity
and quality of water. In a little more than half a century, this global water crisis
has evolved, mainly affecting the developing world in and around the arid
and semi arid regions, especially areas where groundwater is the main source
of drinking water. The drilling of tube wells for agricultural purposes is often
unregulated, though it is supplemented by subsidized electricity for pump-
ing. Much of the progress in food production, such as the Green Revolution
in India and other countries, has taken place unsustainably at the expense of
groundwater. This has triggered a severe drawdown of groundwater tables7; it
has also had lasting and often irreversible impacts on groundwater, resulting
in a synergy of water quality issues.

www.ebook3000.com
Fluoride in Drinking Water 3

1.3  Geogenic Pollutants


Of late, the entry of geogenic pollutants such as fluoride and arsenic into
groundwater aquifers has turned out to be a decisive environmental problem
the world over. This situation is extremely critical in developing countries
such as China and India. In India, fluoride has become endemic in approxi-
mately 37,000 habitations, whereas issues of arsenic are diffused into around
3,200 habitations, thus exhibiting the dominance of the issue. The presence
of excess fluoride in drinking water raises a red flag of concern, as it initiates
fluorosis in various proportions, thereby reducing the quality of human life.
It was estimated that people from more than 35 nations across the globe are
under the threat of fluoride attack. The number of people affected with the
“risk of fluorosis” has probably crossed 200 million.8–10 Therefore, fluoride
in groundwater can be treated as a critical driver in defining the quality of
groundwater in many parts of the world. The steady increase in the num-
ber of people falling prey to fluoride pollution, especially in the develop-
ing world, brings the issue under global focus. Thus, it would be interesting
to have a brief overview on the pathways of fluoride into groundwater, the
chemical profile and geo-chemistry of fluoride, and the status of global fluo-
ride pollution while narrating its context and relevance.

1.3.1  Fluorine: The Chemical Profile


Fluorine is the ninth element in the periodic table, with an atomic weight of
18.9984. It belongs to the group VII A. It is rated thirteenth in abundance and
is estimated to be widely distributed at 0.3 g/kg of the earth’s crust. Elemental
fluorine is the most electronegative and reactive of all elements; as a result,
it rarely occurs naturally in the elemental state. Its electronegative nature
demonstrates that it has a strong tendency to acquire a negative charge in
solution, forming fluoride ion (F−). Except inert gases, it can bond with every
other element, thus forming stable electronegative bonds. “Fluorine reacts
with other elements to produce ionic compounds like hydrogen fluoride and
sodium fluoride in water and upon dissociation forms negatively charged
fluoride ion.”3,6,11,12

1.3.2  Sources of Fluoride


Fluoride is an abundant trace element that is found with an average con-
centration of 625 mg/kg of fluorine in the earth’s crust. However, its occur-
rence is found to vary depending on the types of rocks (from 100 mg/kg
in limestones to 2000 mg/kg in volcanic rocks).13 The rich underlain pres-
ence of crystalline igneous and metamorphic rocks in regions of India, Sri
Lanka, Senegal, Ghana, and South Africa along with areas of volcanic and
associated hydrothermal activity contributes to fluoride. Since “fluorine
4 Fluoride in Drinking Water

has a higher affinity for silicate melts than solid phases it is progressively
enriched in magmas and hydrothermal solutions with time due to mag-
matic differentiation.”3 As a result, the hydrothermal vein deposits and
rocks that crystallize from highly evolved magmas often contain fluorite-,
fluorapatite-, and fluoride-enriched micas and/or amphiboles. Based on the
percentages of silica and calcium present in magma, cryolite, villiaumite,
and/or topaz can also be formed. The highest fluoride levels were reported
from regions that were predominately occupied by crystalline igneous and
metamorphic rocks. These rocks are associated with syenites, granites,
quartz monzonites, granodiorites, felsic and biotite gneisses, and alkaline
volcanic types. It is suggested that the presence of biotite alone may pro-
duce dissolved fluoride concentrations in groundwater to a level of more
than 4 mg/L.13
The parent rock serves as the most natural contributor of fluoride into
drinking water. However, fluorite, the only principal mineral of fluorine,
is regarded as an accessory mineral in granitic rocks. Granite rocks are
reported to exhibit fluoride concentrations of 20–3600 mg/L. Apatite, mus-
covite, amphibole, hornblende, pegmatite, mica, biotite, villiaumite, and cer-
tain types of clays are also found to contain fluorine. The reported natural
Indian sources include the following: the hard rock terrains (south of Ganges
valley) in the arid north-western part; fluoride-rich rocks and canal-irrigated
black cotton soils of Karnataka; the dark mineral fraction of gneisses of
Tamil Nadu; granites, minerals such as sepiolite and palygorskite, acid volca-
nic and basic dikes of Rajasthan; soils and clays of Gujarat; granitic rocks of
Andhra Pradesh; tourmaline-bearing pegmatites of Maharashtra; and sodic
soils in irrigated areas of Haryana and Andhra Pradesh.3,14–19 High fluoride
concentrations can also result from “anion exchange (OH− for F−) on certain
clay minerals, weathered micas and oxyhydroxides that are typically found
in residual soils and sedimentary deposits.”3 Areas underlain by alkaline
volcanic rocks and sedimentary formations that contain fluorapatite- and/
or fluoride-enriched clay minerals may contribute to fluoride concentration.
Crystalline basement rocks such as felsic intrusive rocks and their meta-
morphic forms are also prone to fluoride dissolution. Though fluoride is not
readily leached from soils due to its strong associations with the soil compo-
nents (only 5%–10% of the total fluoride in soil is water soluble), its concen-
tration may increase with depth to the tune of 200–300 mg/L.20 The rate of
fluoride dissociation depends on the soil chemistry, chemical form, climate
of the region, and deposition rates. It is reported that in acidic soils with a pH
less than 6, fluorides can form complexes with iron and aluminum; they can
also form a bonding with clay by replacing hydroxide from the clay surface.
pH plays a crucial role in this adsorption process as it turns significant at
a pH of 3–4, whereas it is reduced at a pH higher than 6.5. The application
of fertilizers under intensive irrigation may result in releasing fluoride into
groundwater. Alkalinization may enhance the concentration of fluoride con-
tent in irrigated lands and soils.20,21
Fluoride in Drinking Water 5

1.4  Fluoride in Humans


It was estimated that 99% of the absorbed fluoride in humans gets depos-
ited in bones and teeth. Though fluoride does not get accumulated in most
soft tissues, such as hydrogen fluoride (HF), it can find its way into the
intracellular fluid of soft tissues. It is plausible that within kidney tubules,
fluoride may get concentrated at high levels, even at a higher concentration
than plasma. Due to this relatively high fluoride-level exposure, kidneys are
regarded as the most vulnerable sinks of fluoride, resulting in them becom-
ing an impending target of acute and chronic fluoride toxicity.21,22 Since the
transportation of fluoride from plasma to milk is minimal, the observed
fluoride levels in human milk are only to the tune of 5–10 μg/L.20 The level
of fluoride concentration in saliva reflects the plasma fluoride availability.
Only low concentrations of fluoride are reported in sweat (around one-fifth
of plasma levels). Renal excretion of fluoride may be 35%–70% of intake in
adults. As a result, urine, plasma, or saliva could be used as biomarkers of
fluoride exposure. The average per capita dietary intake range of fluoride
will be 0.020–0.048 mg/kg for adults (living in regions with fluoride concen-
trations of 1.0 mg/L in water). Although a “no-observed-adverse-effect level
(NOAEL) of 0.15 mg fluoride/kg/day and a lowest observed-adverse-effect
level (LOAEL) of 0.25 mg fluoride/kg/day of fluoride in human” are sug-
gested, these levels are still under scientific debate.3,20,23–25

1.5  Genesis of Fluoride in Groundwater


The origin of fluoride in groundwater is due to an interaction between
groundwater and surface water with rocks containing fluoride-rich mineral.
The presence and concentration of fluoride in groundwater are a reflection
of the amount of concentration of fluoride-bearing minerals present in their
parent rock types. The decomposition and dissolution activities of the rock
types that are exhibited through rock–water interactions play a crucial role.26
In the developing countries such as India, since the contribution from drink-
ing water is the most significant source of fluoride entry into the human
body, a thorough understanding of the geo-chemistry of fluoride in ground-
water appears relevant. It is observed that the rainwater falling on the earth
gets charged by different sources of CO2 from the soil and the atmosphere,
in addition to the biochemical reactions between bacteria and organic mat-
ter during its descent. Thus, the rainwater may turn slightly acidic due to
the formation of carbonic acid. As a result, during percolation, the second-
ary salts present in the soil (mixture of varying content of NaHCO3, NaCl,
and Na2SO4) may get leached out. In phosphate fertilizer-applied lands, soils
6 Fluoride in Drinking Water

most likely contain different percentages and proportions of fluoride-rich


materials and compounds. Simultaneously, an ion-exchange reaction takes
place, with exchangeable cations present in the soil–clay complex as follows:27

CaX 2 + 2Na+(aq ) ↔ 2NaX + Ca 2+(aq ) (1.1)

where X is the clay mineral. As demonstrated by the equations cited, the


hydrogen-ion concentration in groundwater is increased due to the dissolu-
tion of CO2. The calcareous minerals, especially CaCO3, also get dissolved as
follows:27–29

CO 2 + H 2O → H 2CO 3 (1.2)

H 2CO 3 → H+ + HCO 3− (1.3)

HCO 3− → H+ + CO 3 2− (1.4)

CaCO 3 + H+ + 2F − → CaF2 + HCO 3− (1.5)

CaF2 → Ca 2+ + 2F − (1.6)

The groundwater charged with alkalinity may mobilize F− from weathered


rocks, soils, and CaF2, resulting in the precipitation of CaCO3 as follows:29

CaF2 + 2HCO 3− → CaCO 3 + 2F − + H 2O + CO 2 (1.7)

The dissolutional and assimilating activity of fluoride gets enhanced with


the excess presence of sodium bicarbonates in groundwater as follows:29

CaF2 + 2NaHCO 3 → CaCO 3 + 2Na+ + 2F − + H 2O + CO 2 (1.8)

The CaF2 has a solubility product (Ksp) as follows:30

2
K sp = ⎡⎣F − ⎤⎦ ⎡⎣Ca 2+ ⎤⎦ = 4.0 ×10−11 (1.9)

The low-solubility product (Equation 1.9) suggests that high fluoride con-
centrations in groundwater are generally associated with low calcium content
(with a negative correlation between the two ions) and high bicarbonate ions
(in some cases with high nitrate ions).3 Also, it was observed that groundwa-
ter is generally undersaturated with respect to fluorite (in some cases, it may
be saturated with both calcite and fluorite).3,27 It has been established that
Fluoride in Drinking Water 7

the pH of groundwater plays a significant role in defining the concentration


of fluoride in groundwater. An alkaline environment (within a pH range
7.6–8.6) with a high bicarbonate concentration is said to be more conducive
for fluoride dissolution. Thus, the weathering of primary minerals in rocks
appears to be the main contributing factor of fluoride into groundwater.
The weathering of rocks results in the leaching of fluoride-containing min-
erals into groundwater. Being the most predominant mineral, mainly the
presence of fluorite clues in about the concentrations of fluoride in ground-
water. As evidenced in Equation 1.9, the low-solubility product is an irrefut-
able proof that low levels of calcium result in high fluoride concentrations
in water. Further, groundwater in the sodium bicarbonate and bicarbon-
ate chloride types exhibits high fluoride concentrations. The ion-exchange
mechanism, as suggested in Equation 1.1, turns significant in the context of
reported excess fluoride in groundwater in regions of high sodicity of soil.
This process and the mechanisms related to it turn relevant in issues per-
taining to reported high fluoride concentrations near the major south Indian
irrigation projects.16,17,26,28,31–33 In 2012, the most recent research conducted in
the rocks of Hangjinhouqi in a fluoride endemic area of China suggests that
due to CaF2 solubility control, Na-predominant water is favorable for fluoride
enrichment with low total dissolved solid (TDS).34 The elevated fluoride con-
centrations in the groundwaters of south-eastern Pakistan region (1.13–7.85
mg/L) are also attributed to the enhanced fluorite solubility due to Ca deple-
tion, high ionic strength, and the release of fluoride from colloid surfaces
under high pH conditions.3,35 The excess fluoride concentration in the Kolar
and Tumkur districts in Karnataka, India, is also attributed to the reduced
levels of calcium-ion concentration in groundwater due to the calcite pre-
cipitation.36 It is also reported that the dry and hot climate enhances fluoride
levels in the groundwater nearer to the surface. In regions with high rainfall,
the dilution effects in groundwater may outweigh the enrichment effects. In
dry regions where precipitation rates are lower than evaporation, the fluo-
ride generated from the dissolution of fluoride-bearing minerals may move
toward the surface as a result of evaporation. This causes an increase in fluo-
ride concentrations in groundwater. The hydrolysis (OH− in water exchanges
for F−) of F-bearing silicates such as muscovites (Equation 1.10) and biotites
(Equation 1.11) in alkaline soda water can be expressed as follows:37,38

KAl 2 ( AlSi 3O10 ) F2 + 2OH −  = KAl 2 ( AlSi 3O10 ) ( OH ) 2 + 2F − (1.10)

KMg3 ( AlSi 3O10 ) F2 + 2OH − = KMg3 ( AlSi 3O10 ) ( OH ) 2 + 2F − (1.11)

Thus, it could be inferred that the enrichment of fluoride in ground-


water results from water–rock interactions of F-bearing silicates. The
8 Fluoride in Drinking Water

dissolution of fluorite may get further triggered by its enrichment through


evapo-transpiration.

1.6 Summary
• The average per capita water availability may get reduced by one-
third over the next two decades. As a result, by 2025, one-third of
humanity will be under the risk of severe water scarcity.
• Inadequate access to safe drinking water has become the most cru-
cial challenge to the sustainable water supply systems of the world.
• The world is heading toward a water crisis. This global water crisis
mainly affects the developing world in and around the arid and semi
arid regions where groundwater is the main source of drinking water.
• The entry of geogenic pollutants such as fluoride and arsenic into
groundwater aquifers has become an issue of global concern, espe-
cially in developing countries such as India and China.
• Around 200 million people from more than 35 nations the world
over are “at risk” of fluorosis.
• The main source of fluoride in soil is obviously the parent rock itself.
The origin of fluoride in groundwater is mainly due to the interac-
tion between groundwater and surface water with rocks containing
fluoride-rich minerals.
• The rate of fluoride dissociation depends on the chemical form, rate
of deposition, soil chemistry, and climate.
• Around 99% of the absorbed fluoride in humans gets deposited in
bones and teeth. The fluoride accumulation within kidney tubules
may be very high compared with the plasma. As a result, the kidney
could be considered an impending sink, site, and target of the acute
fluoride toxicity.

References
1. WHO and UNICEF (2014). World Health Organization and UNICEF, Progress on
Sanitation and Drinking-Water–2014 Update. Geneva, Switzerland: WHO Press.
2. WWC (2003). World Water Council, 3rd World Water Forum, Press Release,
Crucial water issues to be addressed. Tokyo, Japan: Secretariat of the 3rd
World Water Forum, 2003. http://www.worldwatercouncil.org/download/PR
_­curtainraiser_10.03.03.pdf.
Fluoride in Drinking Water 9

3. Ayoob, S. and Gupta, A.K. (2006). Fluoride in drinking water: A review on the
status and stress effects. Critical Rev. Environ. Sci. Technol., 36, 433–487.
4. WHO and UNICEF (2004). WHO and UNICEF, Meeting the MDG Drinking Water
and Sanitation Target: A Midterm Assessment of Progress. New York: WHO Geneva
and UNICEF, 2004.
5. Maria, A. (2003). The Costs of Water Pollution in India, CERNA, Ecole Nationale
Superieure des Mines de Paris, Paris, France. Revised version, Paper Presented
at the Conference on Market Development of Water and Waste Technologies
through Environmental Economics, 30th–31st October 2003, New Delhi, India.
6. WHO (2004). Fluoride in drinking water, Background document for prepa-
ration of WHO Guidelines for drinking water quality. Geneva, Switzerland:
World Health Organization, http://www.who.int/water_sanitation_health
/dwq/guidelines/en/.
7. Edmunds, W.M. (2009). Geochemistry’s vital contribution to solving water
resource problems. Appl. Geochem., 24, 1058–1073.
8. MRD (2004). First report: Standing Committee on Rural Development, Ministry
of Rural Development (Department of Drinking Water Supply), Presented to
the Fourteenth Lok Sabha, Lok Sabha Secretariat, New Delhi, India, p. 33.
9. Daw, R.K. (2004). Experiences with domestic defluoridation in India,
Proceedings of the 30th WEDC International Conference on People-Centred
Approaches  to  Water and Environmental Sanitation, Vientiane, Lao PDR,
pp. 467–473.
10. Ayoob, S., Gupta, A.K. and Bhat, V.T. (2008). A conceptual overview on sustain-
able technologies for the defluoridation of drinking water. Crit. Rev. Environ.
Sci. Technol., 38, 401–470.
11. Mackay, K.M. and Mackay, R.A. (1989). Introduction to Modern Inorganic
Chemistry, 4th edn, p. 339. Englewood Cliffs, NJ: Prentice Hall.
12. Cotton, F.A. and Wilkinson, G. (1988). Advanced Inorganic Chemistry, p. 546. New
York: John Wiley and Sons.
13. Ozsvath, D.L. (2009). Fluoride and environmental health: A review. Rev. Environ.
Sci. Biotechnol., 8, 59–79.
14. WHO (1984). Fluorine and fluorides. Environmental Health Criteria, 36. Geneva,
Switzerland: World Health Organization.
15. Jacks, G., Bhattacharya, P., Chaudhary, V. and Singh, K.P. (2005). Controls on
the genesis of some high fluoride groundwaters in India. Appl. Geochem., 20,
221–228.
16. Umar, R. and Sami Ahmad, M. (2000). Groundwater quality in parts of Central
Ganga Basin, India. Environ. Geol., 39, 673–678.
17. Datta, K.K. (2000). Reclaiming salt-effected land through drainage in Haryana,
India: A financial analysis. Agric. Water Manag., 46, 55–71.
18. Ramamohana Rao, N.V., Suryaprakasa Rao, K. and Schuiling, R.D. (1993).
Fluorine distribution in waters of Nalgonda District, Andhra Pradesh, India.
Environ. Geol., 21, 84–89.
19. Duraiswami, R.A. and Patankar, U. (2011). Occurrence of fluoride in the drinking
water sources from gad river basin, Maharashtra. J. Geo. Soc. India, 77, 167–174.
20. ATSDR (2003). Report on Toxicological Profile For Fluorides, Hydrogen Fluoride
and Fluorine. U.S. Department of Health and Human Services, Public Health
Service Agency for Toxic Substances and Disease Registry.
10 Fluoride in Drinking Water

21. WHO (2002). Fluorides, Environmental Health Criteria Number, WHO


Monograph No. 227. Geneva, Switzerland: World Health Organization.
22. MRC (2002). Working Group Report: Water Fluoridation and Health, MRC:
47. London: Medical Research Council, 2002. http://www.mrc.ac.uk/pdf-
publications-water_fluoridation_report.pdf.
23. Fomon, S.J. and Ekstrand (1999). Fluoride intake by infants. J. Public Health Dent.,
59, 229–234.
24. Oliveby, A., Twetman, S. and Ekstrand, J. (1990). Diurnal fluoride concentration
in whole saliva in children living in a high- and a low-fluoride area. Caries Res.,
24, 44–47.
25. Li, Y., Liang, C.K., Slemenda, C.W., Ji, R., Sun, S., Cao, J., Emsley, C.L., Ma, F.,
Wu, Y., Ying, P., Zhang, Y., Gao, S., Zhang, W., Katz, B.P., Niu, S., Cao, S. and
Johnston, C.C. Jr. (2001). Effect of long term exposure to fluoride in drinking
water on risks of bone fractures. J. Bone Miner. Res., 16, 932–939.
26. Saxena, V.K. and Ahmed, S. (2001). Dissolution of fluoride in ground water: A
water-rock interaction study. Environ. Geol., 40, 1084–1087.
27. Handa, B.K. (1975). Geochemistry and genesis of fluoride-containing ground
waters in India. Groundwater, 13(3).
28. Saxena, V.K. and Shakeel, A. (2003). Inferring the chemical parameters for the
dissolution of fluoride in ground water. Environ. Geol., 43, 731–736.
29. Subba, R.N. and John, D.D. (2003). Fluoride incidence in ground water in an
area of Peninsular India. Environ. Geol., 45, 243–251.
30. Butler, J.N. (1964). Ionic Equilibrium—A Mathematical Approach. Reading, MA:
Addison-Wesley Publishing Co., Inc.
31. Krishnamachari, K.A.V.R. (1976). Further observations on the syndrome of
endemic genu valgum of South India. Ind. J. Med. Res., 64, 284–291.
32. Singh, R.B. (2000). Environmental consequences of agricultural development:
A case study from the green revolution state Haryana, India. Agric. Ecosys.
Environ. 82, 97–103.
33. Apambire, W.B., Boyle, D.R. and Michel, F.A. (1997). Geochemistry, genesis,
and health implications of fluoriferous ground waters in the upper regions of
Ghana. Environ. Geol., 33, 13–24.
34. He, X., Ma, T., Wang, Y., Shan, H. and Deng, Y. (2013). Hydrogeochemistry of
high fluoride ground water in shallow aquifers, Hangjinhouqi, Hetao Plain.
J Geochem. Explor. http://dx.doi.org/10.1016/j.gexplo.2012.11.010.
35. Rafique, T., Naseem, S., Usmani, T.H., Bashir, E., Khan, F.A. and Bhanger, M.I.
(2009). Geochemical factors controlling the occurrence of high fluoride ground-
water in the Nagar Parkar area, Sindh, Pakistan. J. Hazard. Mater., 171, 424–430.
36. Mamatha, P. and Rao, S.M. (2010). Geochemistry of fluoride rich groundwater
in Kolar and Tumkur Districts of Karnataka. Environ. Earth Sci., 61, 131–142.
37. MHC (2010). Ministry of Health of China, China Health Statistical Yearbook 2010.
Beijing: Peking Union Medical College Press.
38. Guo, Q.H., Wang, Y.X., Ma, T. and Ma, R. (2007). Geochemical processes control-
ling the elevated fluoride concentrations in groundwaters of the Taiyuan basin,
Northern China. J. Geochem. Explor., 93, 1–12.
2
Scenario of Fluoride Pollution

2.1 Introduction
Considerable bodies of scientific literature suggest that fluoride pollution has
been spreading its tentacles into many regions of the world. As a result, more
and more human habitations are forced to consume fluoride-rich ground-
water. The situations in China and India, the most populous countries of
the world, are the worst. In many regions of the world, especially those in
the developing countries, the issue has acquired the dimensions of a socio-
economic problem rather than being one pertaining to mere water quality.
Thus, it would be interesting to have a look into the progression of fluoride
invasions over the boundaries of the world.

2.2  Global Scenario


Of late, fluorosis has become endemic in many parts of the world (Figure 2.1),
as it has spread its tentacles into more than 40 nations, particularly into mid-­
latitude regions. Groundwater with high fluoride occurs in large parts of Africa,
China, the Middle East, and southern Asia, including India and Sri Lanka.
The belt from Eritrea to Malawi along the East African Rift is the most popular
fluoride belt on the earth. Other fluoride-rich zones span wide across China,
Northern Thailand, India, Afghanistan, Iran, Iraq, and Turkey. The United
States and Japan have similar fluoride belts, as shown in Figure 2.1.1 India and
China are the worst affected nations due to fluoride attacks (Figures 2.2 and
2.3). The global scenario of the affected countries along with a description on
the reported intensity and severity of excess fluoride is presented in Table 2.1.

2.2.1  Asian and African Scenario


South Asia serves as home to nearly 25% of the world population, including
43% of the world’s poor. Unfortunately, one-fourth of the populations in the
developing world are living under acute water scarcity. The rural populations

11
12 Fluoride in Drinking Water

6 7
1
8
9 10
2 11 28 32
27 31 33
12 29
3 26 30
13 34
14 17 24
25 37
1516 18 23 22
21 35
4
20

38
5 19
36
Excessive concentration of fluoride in groundwater
1 Canada 8 Poland 15 Ivory Coast 22 Kenya 29 Pakistan 36 New Zealand
2 U.S. 9 Germany 16 Ghana 23 Uganda 30 India 37 Sri Lanka
3 Mexico 10 Check Republic 17 Nigeria 24 Sudan 31 China 38 Australia
4 Brazil 11 Spain 18 Cameroon 25 Ehtopia 32 Korea
5 Argentina 12 Algeria 19 South Africa 26 Saudi Arabia 33 Japan
6 Norway 13 Niger 20 Zambia 27 Israel 34 Thailand
7 Finland 14 Senegal 21 Tanzania 28 Turkey 35 Indonesia

FIGURE 2.1
Global fluoride map. (Modified from WHO, Fluoride and Arsenic in Drinking Water, World Health
Organization, Geneva, Switzerland, 2004, http://www.who.int/water_sanitation_health/en
/map08b.jpg.)

Hailar Basin
Erlian Basin Heilongjiang
Junggar Basin
Zhangye Basin
Inner Mongolia Jilin

Xinjiang Songnen Plain


Liaoning
Tarim Basin
Gansu Beijing
Tianjing
Ordos Basin North China Plain
Hebei
Qaidam Basin Ningxin
Shanxi Shandong
Qinghai

Guide Basin Jiangsu Huaihe River Basin


Henan
Tibet Shaanxi
Yuncheng Basin Taiyuan Basin
Anhui
Shanghai
Sichuan Hubei
Chongqing Zhejiang
N
Hunan Jiangxi
Guizhou Fujian

Yunnan Guangdong Taiwan


Guangxi
0 500 1,000 km
Hongkong
Pearl River Delta

Groundwater F > 5 mg/L Hainan


1,000
km
High F groundwater area South China Sea

FIGURE 2.2
Fluoride map of China. (Modified from Wen, D., Zhang, F., Zhang, E., Wang, C., Han, S. and
Zheng, Y., J. Geochem. Explor., 135, 1–21, 2013.)
Scenario of Fluoride Pollution 13

Jammu &
Kashmir

Himachal Pradesh
Punjab
Uttarakhand
Haryana Arunachal
Delhi Pradesh
Uttar Sikkim
Rajasthan Pradesh Assam Nagaland
Bihar Meghalaya
Manipur
Jharkhand Tripura
Madhya West Mizoram
Gujarat
Pradesh Bengal
arhsg
atti

Maharashtra Orissa
Chh

Andhra
Pradesh States Affected
ka

Goa 70%–100%
nata
Kar

40%–70%

1%–40%
Tamil No data
Kerala Nadu

FIGURE 2.3
Fluoride map of India. (Modified from UNICEF, States of the Art Report on the Extent
of Fluoride in Drinking Water and the Resulting Endemicity in India, Fluorosis and Rural
Development Foundation for UNICEF, New Delhi, 1999.)

are very vulnerable to water-borne diseases due to contamination of drink-


ing water; this ultimately leads to the outbreak of endemic diseases. The con-
tinued use of untreated water from shallow groundwater sources increases
the disease burden of these rural habitations. As per the WHO assessment in
2004, two-thirds of people in Asia live without any access to drinking water.
Further, the water quality issues due to the intrusion of geogenic pollutants
such as fluoride into the scarce drinking water sources add to this misery. The
intensity and gravity of the water quality issues that are persisting in these
regions and associated habitations are so huge to quantify and tackle.7, 69,70
TABLE 2.1
14

Global Scenario of the Intensity and Severity of Excess Fluoride in Drinking Water
Affected
S. No Nations Reported Fluoride Levels in Drinking Water and Associated Effects References
1 India In India, fluoride concentrations ranging from 0.5 to around 70 mg/L were reported. Andhra Ayoob et al.,5
Pradesh, Rajasthan, and Gujarat are the severely affected states. Maximum fluoride concentrations Agarwal et al.,6
reported include 69.7 mg/L in Rajasthan, 23 mg/L in Assam, 32 mg/L in New Delhi, and 48 mg/L Ayoob and Gupta,7
in Haryana. Earlier, around 67 million people living in more than 20 states in India were estimated Susheela,8 Susheela
to be “at risk” of fluorosis. Crippling skeletal fluorosis was reported at a low fluoride concentration and Bhatnagar9
of 2.8 mg/L; dental fluorosis, at 0.5 mg/L; and skeletal fluorosis, at 0.7 mg/L.
2 China In 1990, it was reported that around 300 million people in China were exposed to fluoride-rich waters Li and Cao,10
and associated issues. Out of these, 40 million people had been afflicted with dental fluorosis and 3 UNICEF,11 Wang
million had been afflicted with skeletal fluorosis. However, in 1995, it was reported that one-tenth of et al.,12 Wang and
the total Chinese population was exposed to endemic fluorosis. In Kuitan region of Zhuiger basin, Huang,13 WHO,14,15
concentrations to the level of 21.5 mg/L were reported. The data in 2004 suggested that more than Chen,16 MHC,17
26 million people in China suffered from dental fluorosis and 1 million suffered from skeletal MHPRC18
fluorosis. Fluorosis was extensively reported in China from Shanxi, Inner Mongolia, Shandong,
Henan, and Xinjiang provinces due to high fluoride levels in drinking water. As per the endemic
fluorosis control status of China in 2006, more than 1.34 million inhabitants suffer from skeletal
fluorosis and 21.45 million suffer due to dental fluorosis. In 2010, there were 41.76 million fluorosis
cases in 1325 different counties of China, out of which 58.2% were caused by chronic exposure to
high levels of fluoride in drinking water.
3 Tanzania Many regions of Tanzania are the worst affected with fluorosis-related issues. The range of reported Mjengera and
fluoride concentrations varies from 8 to 12.7 mg/L. Severe cases of dental, skeletal, and crippling Mkongo19
fluorosis were reported from Singida, Shinyanga, Mwanza, Kilimanjaro, Mara, and Arusha regions.
(Continued)
Fluoride in Drinking Water
TABLE 2.1 (Continued)
Global Scenario of the Intensity and Severity of Excess Fluoride in Drinking Water
Affected
S. No Nations Reported Fluoride Levels in Drinking Water and Associated Effects References
4 South Africa The reported fluoride levels in South Africa vary from 0.05 to 13 mg/L. Concentrations as high as 30 Coetzee et al.,20
mg/L were reported from Western Bushveld and Pilanesberg. Fluoride concentrations of 3, 0.48 and Grobler and
0.19 mg/L were reported from Lee Gamka, Kuboes, and Sanddrif, respectively. The occurrences of Dreyer,21 Grobler et
dental fluorosis in children belonging to these regions were reported as 95%, 50% and 47%, al.,22 Mothusi23
respectively. In Western Bushveld regions, acute cases of skeletal fluorosis were reported. The
morbidity rate of dental fluorosis in the Northwest province was similarly very high (97%).
Scenario of Fluoride Pollution

5 Kenya In Kenya, the reported fluoride concentration varies from 1 to 8.0 mg/L, with a fluorosis prevalence Kaimenyi,24 Nair and
rate of 44%–77%. Issues and incidences of skeletal fluorosis were reported at fluoride levels of 18 Manji,25 Nair et al.26
mg/L. The highest fluoride levels were reported from the Rift Valley around Naivasha, Mount Kenya,
and Nakuru and regions near the northern frontier, in addition to the peri-urban areas of Nairobi.
Throughout Kenya, the local fluoride concentrations vary from 2 to 20 mg/L. Very high concentrations
of 2800 and 1640 mg/L were reported from lakes of Nakuru and Elmentaita, respectively. In a sample
study consisting of 1000 groundwater samples, more than 600 samples exceeded 1 mg/L, 200 samples
crossed 5 mg/L, and more than 120 samples showed fluoride concentrations higher than 8 mg/L.
6 Ghana In Ghana, 62% of school children were afflicted with dental fluorosis in the Bongo areas. The fluoride Apambire et al.,27
levels were found to be in the range of 0.11–4.6 mg/L. Recently, the presence of excess fluoride in Salifu et al.28
groundwater (to the tune of 11.6 mg/L) is reported from the northern region of Ghana.
7 Sudan In 1953, the reported fluoride concentrations in Abu Deleig and Jebel Gaili were in the range of 0.65–3.2 Ibrahim et al.,29 Smith
mg/L. Incidentally, the dental fluorosis in Abu Deleig was higher than 60%. In 1995, a high prevalence of and Smith30
(91%) dental fluorosis was observed among those children drinking water with 0.25 mg/L of fluoride.
(Continued)
15
TABLE 2.1 (Continued)
16

Global Scenario of the Intensity and Severity of Excess Fluoride in Drinking Water
Affected
S. No Nations Reported Fluoride Levels in Drinking Water and Associated Effects References
8 United The average fluoride concentrations reported in Illinois were 1.06 and 4.07 mg/L, and those in Cohen and
States Texas were 0.3 and 4.3 mg/L. In the hot springs and geysers of the National Park at Yellowstone, Conrad,31 Driscoll
fluoride levels of 25 to 50 mg/L were reported. The range of fluoride concentration in Lakeland at et al.,32 Neuhold
Southern California was 3.6–5.3 mg/L. A range of 5.0–15 mg/L of fluoride was reported in the and Sigler,33
deep aquifers of Western United States. Reardon and
Wang,34 Segreto
et al.35
9 Mexico In Mexico, around 6% of the population (around 5 million people) is affected by high Díaz-Barriga et al.,36
concentrations of fluoride in groundwater. The maximum concentrations reported were in Abasolo UNICEF37
in Guanajuato state (8 mg/L) and in Hermosillo in Sonara state (7.8 mg/L). Average fluoride
levels of 0.9–4.5 mg/L were observed in rural locations and of 1.5–2.8 mg/L were observed in
urban locations.
10 Ethiopia A high occurrence of dental fluorosis was reported in Ethiopian Rift Valley, where the Haimanot et al.,38
concentrations of fluoride range from 1.5 to 177 mg/L. The Wonji-Shoa sugar estates in Ethiopian Kloos et al.,39
Rift Valley recorded the highest occurrence of skeletal and crippling skeletal fluorosis. In the Main Rango et al.40
Ethiopian Rift (MER) Valley (a part of the East African Rift), among 10 million people,
more than 8 million are exposed to an elevated concentration of fluoride. It is estimated that about
1.2 million inhabitants drink groundwater with fluoride contents that exceed international
guideline values.
11 Canada High fluoride concentrations were reported from Alberta (4.3 mg/L), Saskatchewan (2.8 mg/L) and Health Canada,
Quebec (2.5 mg/L) in Canada. The range of fluoride concentrations observed in Rigolet and Priority Substances
Labrador was 0.1–3.8 mg/L. Issues of dental fluorosis were also reported from Rigolet. List Assessment
Report on Inorganic
Fluorides,41 Ismail
and Messer,42
WHO43
(Continued)
Fluoride in Drinking Water
TABLE 2.1 (Continued)
Global Scenario of the Intensity and Severity of Excess Fluoride in Drinking Water
Affected
S. No Nations Reported Fluoride Levels in Drinking Water and Associated Effects References
12 Poland, Fluoride concentrations higher than 3 mg/L were reported from Czech Republic, Finland, and Azbar and
Finland, Poland. In Brazil, higher concentrations of fluoride were reported from Paraiba state (0.1–2.3 Türkman,44 Cortes
Czech mg/L) in the northeast region and from Ceara state (2–3 mg/L). In Indonesia, fluoride levels of et al.,45 Czarnowski
Republic, 0.1–4.2 mg/L were reported in the well waters of the north-eastern part of Java in the Asembagus et al.,46 Heikens et
Brazil, coastal plain. In Israel, natural fluoride concentrations up to 3 mg/L were reported from the al.,47 Milgalter et
Indonesia, Negev desert regions. In Turkey, high fluoride concentrations were reported from the Middle and al.,48 WHO,14,43
Israel, Eastern parts. Denizli-Saraykoy and Caldiran plains exhibited a maximum concentration of 13.7 Oruc,49 Fantong
Scenario of Fluoride Pollution

Turkey, mg/L. Recently, high fluoride-induced dental and skeletal fluorosis have been observed in et al.,50 Fordyce
Cameron, Cameroon, Zambia, and Europe. et al.51
Zambia, Shitumbanuma
and Europe et al.52
13 Ivory coast, The occurrence of fluorosis was reported from all these regions. High prevalence of severe dental Brouwer et al.,53
Senegal, fluorosis among children (30%–60%) was reported in Senegal, from Guinguineo and Darou Paoloni et al.,54
North Rahmane Fall regions with fluoride concentrations of 4.6 and 7.4 mg/L. In Argentina, the Rwenyonyi et al.,55
Algeria, south-east subhumid pampa regions showed fluoride levels of 0.9–18.2 mg/L with an average WHO1
Uganda, value of 3.8 mg/L. In Western Uganda, issues of dental fluorosis were reported from the Rift
and Valley area with fluoride levels of 0.5–2.5 mg/L.
Argentina
(Continued)
17
TABLE 2.1 (Continued)
18

Global Scenario of the Intensity and Severity of Excess Fluoride in Drinking Water
Affected
S. No Nations Reported Fluoride Levels in Drinking Water and Associated Effects References
14 Norway, The occurrence of fluorosis has been reported from all these regions. Issues of severe dental WHO,1 Bardsen
New fluorosis were reported in the county of Hordaland, Norway, where fluoride levels reported in et al.,56 Hardisson et
Zealand, groundwater ranged from 0.02 to 9.48 mg/L. In the Muenster regions of Germany, fluoride al.,57 Queste et al.,58
Germany, concentrations of 8.8 mg/L were reported. The range of fluoride concentration in the Tenerife Shah and
Spain, areas of Spain varies from 2.50 to 4.59 mg/L. The prevalence of skeletal fluorosis was reported Danishwar,59
Niger, from Tibiri in Niger; this was evident among boys who were exposed to fluoride concentrations of Wongdem et al.,60
Nigeria, 4.7–6.6 mg/L. A fluoride exposure level of 0.5–3.96 mg/L in the Langtang town area of Nigeria Poureslami et al.,61
Pakistan, exhibited 26.1% occurrence of dental fluorosis. In Pakistan, fluoride concentrations of 8–13.52 Fekri and Kasmaei62
and Iran mg/L were observed in and around the spring and stream sources of Naranji. High fluoride levels
(8.85 mg/L) in drinking water and high prevalence of dental fluorosis are reported from Iran.
15 Saudi Elevated fluoride levels were reported from Saudi Arabia in Mecca (2.5 mg/L) and Hail regions Akpata et al.,63
Arabia, (2.8 mg/L) along with incidences of fluorosis. In Eritrea, elevated fluoride concentrations of Al-Khateeb et al.,64
Eritrea, 2.02–3.73 mg/L were reported in and around Keren areas. In Sri Lanka, very high concentrations Dissanayake,65 Kim
Sri Lanka, up to 10 mg/L were reported in the North Central Province. In Thailand, at least 1% of the natural and Jeong,66
Thailand, drinking water sources are laced with fluoride concentrations that are more than 2 mg/L, with Srikanth et al.,67
Japan, and higher values exceeding 10 mg/L. In Japan, people exposed to 1.4 mg/L fluoride concentrations Tsutsui et al.68
Korea are afflicted with 15.4% prevalence of dental fluorosis. One-fourth of the total wells in the
south-eastern part of Korea have groundwater with fluoride concentrations greater than 5 mg/L.
Source:  Modified from Ayoob, S., Gupta, A. K. and Bhat, V.T., Critical Rev. Environ. Sci. Technol., 38, 401–470, 2008.
Fluoride in Drinking Water
Scenario of Fluoride Pollution 19

High fluoride concentrations in drinking water and associated fluorosis


issues were reported from African countries such as Tanzania, South Africa,
Kenya, Ghana, and Sudan (Table 2.1). Tanzania is one of the countries in the
world that is severely affected by fluorosis. The crippling skeletal fluorosis
keeps people, especially children, immobile. Severe issues related to crippling
skeletal and dental fluorosis were reported from regions such as Shinyanga,
Singida, Mara, Mwanza, Kilimanjaro, and Arusha (Figure 2.4); however,
Dodoma, Kigoma, Tabora, and Tanga are moderately affected. Reports deal-
ing with excess fluoride in drinking water and associated health problems in
South Africa made their appearance from 1935 onward (Figure 2.5). North-West
provinces, Limpopo, Northern Cape, and major portions of Karoo are badly
affected with issues of excess fluoride. Areas belonging to Western Bushveld
and Pilanesberg reported fluoride concentrations higher than 1 mg/L. Even
very high concentrations (around 30 mg/L) were reported in alkaline waters
with a pH higher than 9. High fluoride concentrations exist in groundwater of
North-West and Kwa-Zulu-Natal provinces, Northern Cape and Limpopo. In
the North-West Province, the morbidity of dental fluorosis was 97%. Kenya in
East Africa is bordered by Somalia, Ethiopia, Sudan, Uganda, and Tanzania.
Fluorosis incidences of varying degrees were observed in Kenya, with a gen-
eral prevalence rate up to 77%. Issues of skeletal fluorosis were identified
among people drinking borehole water with high fluoride concentrations of
18 mg/L. In Ghana, more than 60% of the total populations of school children

Mara
Kagera

Mwanza
Arusha
Shinyanga
Kilimanjaro
S
Kigoma i
n Manyara
Tabora d Tanga
i
g
a Dodoma
Rukwa

Morogom
Iringa
Mbeya

Fluoride-affected areas Lindi

Severely affected
Ruvuma
Moderately affected Mtwara

No data

FIGURE 2.4
Fluorotic map of Tanzania. (Modified from Mjengera, H. and Mkongo, G., Phys. Chem. Earth,
28, 1097–1104, 2003.)
20 Fluoride in Drinking Water

Bela-Bela Polokwane
Pilanesberg
Mmabatho
Botswana
Swaziland

Johannesburg
Namibia
Lesotho Richards Bay
Kimberley Mooiriver
Vredendal Durban
Bloemfontein
Atlantic Indian Ocean
Ocean East London
Port Elizabeth
Cape Town Mossel Bay

Regions with groundwater fluoride


greater than 1.5 ppm

FIGURE 2.5
Fluorosis map of South Africa with groundwater fluoride concentration greater than 1.5 mg/L.
(Modified from McCaffrey, L. P. and Willis, J. P., Distribution of Fluoride-Rich Groundwater in
the Eastern and Mogwase Regions of the Northern and North-West Provinces, W.R.C. Report.
No. 526/1/01, 2001.)

in the Bongo area reported to have dental fluorosis. In Bolgatanga and Bongo
districts, elevated levels in natural groundwater were reported. In Sudan, 91%
prevalence of dental fluorosis was observed among children consuming water
having 0.25 mg/L of fluoride.7

2.2.2  Indian Scenario


India is the seventh largest, second most populous country in the world and
is home to 16% of the global population. Though home to a sixth of human-
ity, India has just 4% of global water resources. The expected population is
1330 million in 2020. Though the bulk majority of the drinking water require-
ments are addressed by groundwater, nearly one-half of Indian villages are
facing issues of acute drinking water shortage. Thus, providing reliable and
potable water to all habitations is still a distant dream. Of late, the advance-
ments made in India in the frontier areas of irrigation and food security were
believed to be at the expense of groundwater. As per an estimate in 2004,
around 3.7 billion bore wells were constructed for irrigation. The uncontrolled
mining of groundwater through these bore wells contributed to an imbalance
in our natural ecosystem. This imbalance is believed to be one of the reasons
for the increased scarcity and pollution of our groundwater sources. Thus,
such unfettered pumping from groundwater sources might have resulted in
Scenario of Fluoride Pollution 21

the decline of the natural water table, which might have triggered the entry
of geogenic pollutants such as fluoride into groundwater aquifers. Further,
geological processes governed by different hydrological and geochemical set-
tings might have accelerated the entry of fluoride into groundwater.
In the 1930s, though fluorosis was reported from only 4 states of India,
as per latest reports, more than 20 states are affected. Rajasthan, Andhra
Pradesh, and Gujarat are the worst affected states (Figure 2.3). Higher fluo-
ride concentrations of 44 mg/L and 23 mg/L, respectively, are reported from
Rajasthan and Assam (Table 2.1). A natural maximum fluoride concentration
of 32 mg/L was reported from the mega city of Delhi.8 A fluoride concentra-
tion of 21.0 mg/L is reported in the groundwater at Kurmapalli watershed in
Nalgonda district of Andhra Pradesh (the worst affected state of India); this
is one of the highest concentrations in the groundwater of granite terrain
in the country.72 The granitic rocks in Nalgonda were identified as the chief
source of excess fluoride in the groundwater that ranges from around 300 to
3200 mg/L. The granitic rocks of Nalgonda appear to have the highest fluo-
ride content in the world with a mean fluoride concentration of 1440 mg/L.73

2.3 Summary
• The fluoride belt from Eritrea to Malawi along the East African Rift is
the most popular fluoride belt on the earth. Other fluoride-rich zones
span wide across China, Northern Thailand, India, Afghanistan,
Iran, Iraq, and Turkey.
• More than 65% of people in Asia have no access to safe drinking
water. Due to the intrusion of fluoride into drinking water sources,
water quality issues could turn the Asian people’s lives miserable.
• In addition to China and India, high fluoride concentrations in
drinking water and associated fluorosis issues were reported from
African countries such as Tanzania, South Africa, Kenya, Ghana, and
Sudan. Tanzania is one of the countries in the world that is severely
affected by fluorosis. The most recent literature suggests that people
from more than 40 nations across the world are under the “risk” of
fluorosis.
• Geological processes facilitating the weathering of rocks and associ-
ated fluoride-bearing minerals under different hydrological and geo-
chemical settings accelerate the entry of fluoride into groundwater.
• Although fluorosis was reported from only 4 states of India in the
1930s, as per the latest reports, more than 20 states are affected.
Rajasthan, Andhra Pradesh, and Gujarat are the worst affected
states. Higher fluoride concentrations of 44 and 23 mg/L, respec-
tively, are reported from Rajasthan and Assam.
22 Fluoride in Drinking Water

References
1. WHO (2005). World Sanitation and Heath, Geneva, Switzerland: World Health
Organization.  http://www.who.int/water_sanitation_health/diseases
/fluorosis/en.
2. WHO (2004). Fluoride and Arsenic in Drinking Water. Geneva, Switzerland:
World Health Organization. http://www.who.int/water_sanitation_health
/en/map08b.jpg.
3. Wen, D., Zhang, F., Zhang, E., Wang, C.L., Han, S. and Zheng, Y. (2013). Arsenic,
fluoride and iodine in groundwater of China. J. Geochem. Explor., 135, 1–21.
4. UNICEF (1999). States of the Art Report on the Extent of Fluoride in Drinking
Water and the Resulting Endemicity in India. New Delhi, India: Fluorosis and
Rural Development Foundation for UNICEF.
5. Ayoob, S., Gupta, A.K. and Bhat, V.T. (2008). A conceptual overview on sustain-
able technologies for the defluoridation of drinking water. Crit. Rev. Environ.
Sci. Technol., 38, 401–470.
6. Agarwal, C.K., Gupta, K.S. and Gupta, B.A. (1999). Development of new low
cost defluoridation technology. Water Sci. Technol., 40, 167–173.
7. Ayoob, S. and Gupta, A.K. (2006). Fluoride in drinking water: A review on the
status and stress effects. Crit. Rev. Environ. Sci. Technol., 36, 433–487.
8. Susheela, A.K. (2003). A Treatise on Fluorosis, 2nd revised edn, pp. 13–14. New
Delhi, India: Fluorosis Research and Rural Development Foundation.
9. Susheela, A.K. and Bhatnagar, M. (1999). Structural aberrations in fluo-
rosed human teeth: Biochemical and scanning electron microscopic studies.
Curr. Sci., 77, 1677–1680.
10. Li, J. and Cao, S. (1994). Recent studies on endemic fluorosis in China. Fluoride,
27, 125–128.
11. UNICEF. Web site: UNICEF statement on fluoride in water, http://www
.fluoride.org.uk/statements/000000unicef.htm.
12. Wang, G.Q., Huang, Y.Z., Xiao, B.Y., Qian, X.C., Yao, H., Hu , Y., Gu, Y.L.,
Zhang, C. and Liu, K.T. (1997). Toxicity from water containing arsenic and fluo-
ride in Xinjiang. Fluoride, 30, 81–84.
13. Wang, L.F. and Huang, J.Z. (1995). Outline of control practice of endemic fluo-
rosis in China. Soc. Sci. Med., 41, 1191–1195.
14. WHO (2006). In: Fawell, J., Bailey, K., Chilton, J., Dahi, E., Fewtrell, L. and
Magara, Y. (Eds.), Fluoride in Drinking Water, pp. 41–75. London, UK: IWA
Publishing.
15. WHO (2004). WHO issues revised drinking water guidelines to help pre-
vent water-related outbreaks and disease, press release WHO/67. Geneva,
Switzerland: World Health Organization.
16. Chen, N., Zhang, Z., Feng, C., Li, M., Chen, R. and Sugiura, N. (2011).
Investigations on the batch and fixed-bed column performance of fluoride
adsorption by Kanuma mud. Desalination, 268, 76–82.
17. MHC (2010). Ministry of Health of China, China Health Statistical Yearbook 2010.
Beijing: Peking Union Medical College Press.
18. MHPRC (2007). Ministry of Health of the People's Republic of China, Chinese
Health Statistical Digest. http://www.moh.gov.cn/open/2007tjts/P50.htm.
Scenario of Fluoride Pollution 23

19. Mjengera, H. and Mkongo, G. (2003). Appropriate deflouridation technology


for use in flourotic areas in Tanzania. Phys. Chem. Earth, 28, 1097–1104.
20. Coetzee, P.P., Coetzee, L.L., Puka, R. and Mubenga, I.S. (2003). Characterisation
of selected South African clays for defluoridation of natural waters. Water SA,
29, 331–338.
21. Grobler, S.R. and Dreyer, A.G. (1988). Variations in the fluoride levels of drink-
ing water in South Africa, Implications for fluoride supplementation. S. Afr.
Med. J., 73, 217–219.
22. Grobler, S.R., Dreyer, A.G. and Blignaut, R.J. (2001). Drinking water in South Africa:
Implications for fluoride supplementation. J. S. Afr. Dent. Assoc., 56, 557–559.
23. Mothusi, B. (1998). Psychological effects of dental fluorosis, Fluoride and
Fluorosis, The Status of South African Research, Pilanesberg National Park,
North West Province, 7, 1995; as cited in Muller, W.J., Heath, R.G.M. and Villet,
M.H. (1998). Finding the optimum: Fluoridation of potable water in South
Africa. Water SA, 24, 1–27.
24. Kaimenyi, T.J. (2004). Oral health in Kenya. Int. Dent. J., 54, 378–382.
25. Nair, K.R. and Manji, F. (1982). Endemic fluorosis in deciduous dentition—
A study of 1276 children in typically high fluoride area (Kiambu) in Kenya.
Odonto-Stomatol. Trop., 4, 177–184.
26. Nair, K.R., Manji, F. and Gitonga, J.N. (1984). The occurrence and distribution of
fluoride in groundwaters in Kenya. East Afr. J. Med., 61, 503–512.
27. Apambire, W.B., Boyle, D.R. and Michel, F.A. (1997). Geochemistry, genesis,
and health implications of fluoriferous ground waters in the upper regions of
Ghana. Environ. Geol., 33, 13–24.
28. Salifu, A., Petrusevski, B., Ghebremichael, K., Buamah, R. and Amy, G. (2012).
Multivariatestatistical analysis for fluoride occurrence in groundwater in the
Northern region of Ghana. J. Contam. Hydrol., 140–141, 34–44.
29. Ibrahim, Y.E., Affan, A.A. and Bjorvatn, K. (1995). Prevalence of dental fluorosis
in Sudanese children from two villages with respectively 0.25 mg/L and 2.56
mg/L F in the drinking water. Int. J. Paediatr. Dent., 5, 223–229.
30. Smith, H.R. and Smith, L.C. (1937). Bone contact removes fluoride. Water Works
Eng., 90, 600.
31. Cohen, D. and Conrad, M.H. (1998). 65,000 GPD fluoride removal membrane
system in Lakeland, California, USA. Desalination, 117, 19–35.
32. Driscoll, W.S., Horowitz, H.S., Meyers, R.J., Heifetz, S.B., Kingman, A. and
Zimmerman, E.R. (1983). Prevalence of dental caries and dental fluorosis in
areas with optimal and above-optimal water fluoride concentrations. J. Am.
Dental Assoc., 107, 42–47.
33. Neuhold, J.M. and Sigler, W.F. (1960). Effects of sodium fluoride on carp and
rainbow trout. Trans. Am. Fish. Soc., 89, 358–370.
34. Reardon, J.E. and Wang, Y. (2000). A limestone reactor for fluoride removal
from wastewaters. Environ. Sci. Technol., 34, 3247–3253.
35. Segreto, V.A., Collins, E.M., Camann, D. and Smith, C.T. (1984). A current study
of mottled enamel in Texas. J. Am. Dental Assoc., 108, 56–59.
36. Díaz-Barriga, F., Navarro-Quezada, A., Grijalva, M.I., Grimaldo, M., Loyola-
Rodriguez, J.P. and Ortz, M.D. (1997). Endemic fluorosis in Mexico. Fluoride, 30,
233–239.
37. UNICEF. Web site: UNICEF statement on fluoride in water, http://www
.fluoride.org.uk/statements/000000unicef.htm.
24 Fluoride in Drinking Water

38. Haimanot, R.T., Fekadu, A. and Bushra, B. (1987). Endemic fluorosis in the
Ethiopian Rift Valley. Tropic. Geogr. Med., 39, 209–217.
39. Kloos, H., Tekle-Haimanot, R., Fluorosis, Kloos, H. and Zein, A.H. (1993). The
Ecology of Health and Disease in Ethiopia, pp. 44–541. Boulder, CO: Westview
Press.
40. Rango, T., Bianchini, G., Beccaluva, L. and Tassinari, R. (2010). Geochemistry
and water quality assessment of central Main Ethiopian Rift natural waters
with emphasis on source and occurrence of fluoride and arsenic, J. Afr. Earth
Sci., 57, 479–491.
41. Health Canada, Priority Substances List Assessment Report on Inorganic
Fluorides. (1993). Canadian Environmental Protection Act, Minister of Supply and
Services Canada, pp. 12–19. Ottawa, Canada: Canada Communication Group-
Publishing. K1A 0S9 1993.
42. Ismail, A.I. and Messer, J.G. (1996). The risk of fluorosis in students exposed to
a higher than optimal concentration of fluoride in well water. J. Public Health
Dent., 56, 22–27.
43. WHO (2002). Fluorides. Environmental Health Criteria. WHO monograph No. 227.
Geneva: World Health Organization.
44. Azbar, N. and Türkman, A. (2000). Defluoridation in drinking water. Water Sci.
Technol., 42, 403–407.
45. Cortes, D.F., Ellwood, R.P., O'Mullane, D.M. and de Magalhaes Bastos, J.R.
(1996). Drinking water fluoride levels, dental fluorosis and caries experience in
Brazil. J. Public Health Dentistry, 56, 226–228.
46. Czarnowski, W., Wrzesniowska, K. and Krechniak, J. (1996). Fluoride in drink-
ing water and human urine in Northern and Central Poland. Sci. Total Environ.,
191, 177–184.
47. Heikens, A., Sumarti, S., vanBergen, M., Widianarko, B., Fokkert, L., van
Leeuwen, K. and Seinen, W. (2005). The impact of the hyperacid Ijen Crater
Lake: Risks of excess fluoride to human health. Sci. Total Environ., 346, 56–69.
48. Milgalter, N., Zadik, D. and Kelman, A.M. (1974). Fluorosis and dental caries in
Yotvata area. Israel Dental J., 23, 104–109.
49. Oruc, N. (2008). Occurrence and problems of high fluoride waters in Turkey: An
overview. Environ. Geochem. Health, 30, 315–323.
50. Fantong, W.Y., Satake, H., Ayonghe, S.N., Suh, E.C., Adelana, S.M.A., Fantong,
E.B.S., Banseka, H.S., Gwanfogbe, C.D., Woincham, L.N., Uehara, Y. and
Zhang, J. (2009). Geochemical provenance and spatial distribution of fluoride
in groundwater of Mayo Tsanaga River Basin, Far North Region, Cameroon:
Implications for incidence of fluorosis and optimal consumption dose. Environ.
Geochem. Health, 32, 147–163.
51. Fordyce, F.M., Vrana, K., Zhovinsky, E., Povoroznuk, V., Toth, G., Hope, B.C.,
Iljinsky, U. and Baker, J. (2007). A health risk assessment for fluoride in Central
Europe. Environ. Geochem. Health, 29, 83–102.
52. Shitumbanuma, V., Tembo, F., Tembo, J.M., Chilala, S. and Van Ranst, E. (2007).
Dental fluorosis associated with drinking water from hot springs in Choma
district in southern province, Zambia. Environ. Geochem. Health, 29, 51–58.
53. Brouwer, I.D., Dirks, O.B., De-Bruin, A. and Hautvast, J.G. (1988). Unsuitability
of World Health Organization guidelines for fluoride concentrations in drink-
ing water in Senegal. Lancet, 30, 223–225.
Scenario of Fluoride Pollution 25

54. Paoloni, J.D., Fiorentino, C.E. and Sequeira, M.E. (2003). Fluoride contamina-
tion of aquifers in the southeast sub humid pampa, Argentina. Environ. Toxicol.,
18, 317–320.
55. Rwenyonyi, C.M., Birkeland, J.M., Haugejorden, O. and Bjorvatn, K. (2000). Age
as a determinant of severity of dental fluorosis in children residing in areas
with 0.5 and 2.5 mg fluoride per liter in drinking water. Clin. Oral Invest., 4,
157–161.
56. Bardsen, A., Klock, K.S. and Bjorvatn, K. (1999). Dental fluorosis among per-
sons exposed to high- and low-fluoride drinking water in western Norway.
Commun. Dentistry Oral Epidemiol., 27, 259–267.
57. Hardisson, A., Rodriguez, M.I., Burgos, A., Flores, L.D., Gutierrez, R. and Varela,
H. (2001). Fluoride levels in publically supplied and bottled drinking waters in
the island of Tenerife, Spain. Bull. Environ. Contamin. Toxicol., 67, 163–170.
58. Queste, A., Lacombe, M., Hellmeier, W., Hillermann, F., Bortulussi, B., Kaup,
M., Ott, K. and Mathys, W. (2001). High concentrations of fluoride and boron in
drinking water wells in the Muenster region—Results of a preliminary investi-
gation. Int. J. Environ. Health, 203, 221–224.
59. Shah, M.T. and Danishwar, S. (2003). Potential fluoride contamination in the
drinking water of Naranji area, Northwest Frontier Province, Pakistan. Environ.
Geochem. Health, 25, 475–481.
60. Wongdem, J.G., Aderinokun, G.A., Sridhar, M.K. and Selkur, S. (2000).
Prevalence and distribution pattern of enamel fluorosis in Langtang town,
Nigeria. Afr. J. Med. Medical Sci., 29, 243–246.
61. Poureslami, H.R., Khazaeli, P. and Nooric, G.R. (2008). Fluoride in food and
water consumed in koohbanan. Fluoride, 41, 216–219.
62. Fekri, M. and Kasmaei, L.S. (2011). Fluoride pollution in soils and waters of
Koohbanan region, southeastern Iran. Arab. J. Geosci., 6, 157–161.
63. Akpata, E.S., Fakiha, Z. and Khan, N. (1997). Dental fluorosis in 12–15-year-old
rural children exposed to fluorides from well drinking water in the Hail region
of Saudi Arabia. Comm. Dentistry Oral Epidemiol., 25, 324–327.
64. Al-Khateeb, T.L., Al-Marasafi, A.I. and O'Mullane, D.M. (1991). Caries preva-
lence and treatment need amongst children in an Arabian community. Commun.
Dentistry Oral Epidemiol., 19, 277–280.
65. Dissanayake, C.B. (1996). Water quality and dental health in the Dry Zone of Sri
Lanka. Environ. Geochem. Health, 113, 131–140.
66. Kim, K. and Jeong, Y.G. (2005). Factors influencing natural occurrence of
fluoride-rich groundwaters: A case study in the southeastern part of the Korean
Peninsula. Chemosphere, 58, 1399–1408.
67. Srikanth, R., Viswanatham, K.S., Kahsai, F., Fisahatsion, A. and Asmellash,
M. (2002). Fluoride in groundwater in selected villages in Eritrea (North East
Africa). Environ. Monit. Assess., 75, 169–177.
68. Tsutsui, A., Yagi, M. and Horowitz, A.M. (2000). The prevalence of dental caries
and fluorosis in Japanese communities with up to 1.4 ppm of naturally occur-
ring fluoride. J. Public Health Dentistry, 60, 147–153.
69. WHO (2004). Ground Water and Public Health, Chapter 1. Geneva, Switzerland:
World Health Organization. www.who.int/entity/water sanitation health
/resources quality/en/groundwater1.pdf.
70. WHO (2004). FACTS Water, Sanitation and Hygiene Links to Health, Facts and
Figures. Geneva, Switzerland: World Health Organization.
26 Fluoride in Drinking Water

71. McCaffrey, L.P. and Willis, J.P. (2001). Distribution of Fluoride-Rich


Groundwater in the Eastern and Mogwase Regions of the Northern and
North-West Provinces. W.R.C. Report. No. 526/1/01.
72. Mondal, N.C., Prasad, R.K., Saxena, V.K., Singh, Y. and Singh, V.S. (2009).
Appraisal of highly fluoride zones in groundwater of Kurmapalli watershed,
Nalgonda district, Andhra Pradesh (India). Environ. Earth Sci., 59, 63–73.
73. Ramamohana Rao, N.V., Rao, N., Rao, S.P. and Schuiling, H.D. (1993). Fluorine
distribution in waters of Nalgonda District, Andhra Pradesh, India. Environ.
Geology, 21, 84–89.
3
Dental Fluorosis

3.1 Introduction
Many studies suggest that fluoride may be an essential element for both
animals and humans. However, it is true that the essentiality of fluoride
for humans has not been demonstrated indisputably. Further, data on the
minimum nutritional requirement are also inadequate. Incidentally, many
epidemiological studies have clearly demonstrated possible adverse effects
and health issues that arise due to the continuous ingestion of fluoride that
is derived mainly through drinking water. These studies clearly show that
fluoride primarily produces effects on skeletal tissues, especially bones and
teeth. However, low concentrations of fluoride provide protection against
dental caries, especially in children. According to the World Oral Health
Report 2003, for a considerable percentage of people, especially children, in
most of the industrialized countries, dental decay (dental caries) still remains
a major public health issue. The changing living conditions and dietary
habits are expected to be reasons for increased incidences of dental decay.
Although considerable advancements have been made in preserving and
improving global oral health issues, many of such issues related to the poor,
marginalized, and disadvantaged groups still persist. Scientific research on
the oral health issues related to fluoride started more than a century ago and
focused on establishing a link between fluoride, dental caries, and fluorosis.
Studies suggest that fluoride toothpastes and mouth rinses can significantly
reduce the occurrence and prevalence of dental decay.1

3.2  Dental Effects of Fluoride


The presence of fluoride in water has dual significance on human health, that
is, the concentration of fluoride in the drinking water defines both its ben-
eficial and harmful effects. Although a minimum level of fluoride in drink-
ing water may reduce dental caries, as stated earlier, higher concentrations

27
28 Fluoride in Drinking Water

may initiate dental fluorosis in various proportions. Though water fluorida-


tion at minimal levels may be beneficial, minimizing the adverse fluorotic
effects on teeth at higher concentrations should be cautioned. Accordingly,
the optimum concentration of fluoride in drinking water is generally limited
to 1 mg/L, as it may ensure maximum dental protection without initiating
any adverse health problems.
Fluoride was considered to improve the crystal lattice stability of enamel
and render it less soluble to acid demineralization. Since fluoride is incor-
porated into enamel as partially fluoridated hydroxyapatite, it is considered
best when it is ingested. However, an increasing body of evidence suggests
that a substantial part of the cariostatic activity of fluoride is due to its effects
on erupted teeth. Further, the mechanism of action is mainly centered on
the presence of fluoride in the fluid phase of dental plaque and saliva. The
fluoride available in saliva and dental plaque reduces the demineralization of
teeth and enhances the remineralization, mainly through an interaction with
the surface of the enamel. Fluoridated toothpastes, mouth rinses, and topi-
cally applied dental treatments such as varnishes, gels, and solutions are also
used in addition to water fluoridation. In addition, fluoridated milk, fluori-
dated salt, and other fluoride supplements are also used in many countries.
Of late, considering the average annual maximum daily air temperature of
the region, the optimum recommended range of fluoride in drinking water
is from 0.5 to 1.2 mg/L.2 The population using fluoride toothpastes (contain-
ing 10 g of fluoride per kg) for preventing dental caries may be twice as those
consuming groundwater with excess fluoride. It is suggested that in many
developed countries, use of these products is considered helpful for the grad-
ual decline in the prevalence of dental decay. However, many studies sug-
gest that fluoride supplements have only a limited role in enhancing dental
health. It is pointed out in some studies that applications of fluoridated mouth
rinses have become popular among school children and kids. However, their
efficacy in preventing dental decay is dependent on the frequency of usage,
level of compliance, and exposure to other fluoride sources. Proper caution
is to be maintained in recommending the usage of mouth rinses for children
younger than 6 years, as they may swallow it during rinsing, thereby increas-
ing the risk of dental fluorosis.2 Thus, it would be appropriate to recommend
a mouth rinse only for the elderly people with elevated risks of dental decay.

3.2.1  Dental Caries (Tooth Decay)


Dental caries (decay) is one of the most prevalent chronic childhood diseases
worldwide. The disease develops in both the crowns and roots of teeth, and
it grows into aggressive tooth decay. It is presumed that there exists a physi-
ological stability between the oral microbial films and teeth minerals. The
alterations in this stable equilibrium may contribute to the initiation of den-
tal caries.3 It could be viewed as an “infectious and multifactorial disease”
that is characterized by demineralization of inorganic components of teeth
Dental Fluorosis 29

and dissolution of organic substances of microbial etiology. Unhygienic oral


cavities invite bacterial growth, resulting in caries, which further leads to
acid production by fermentation. This may etch away enamel, leaving black
spots or cavities on teeth. These microorganisms can damage the soft pulp
tissues by infiltrating through the dentin.3 Dental decay ultimately leads to
a state of acute systemic infection through different phases such as devas-
tating pain, bacterial contagion resulting in pulpal necrosis, reduced dental
function, tooth extraction, and so on. Streptococcus mutans and lactobacilli in
dental plaque are mainly the two types of specific bacteria that are respon-
sible for inducing caries.

3.2.2  Prevention of Dental Caries by Fluoride


The mineral crystals of calcium and phosphate form enamel and dentin,
which are entrenched in an organic protein–lipid medium. Fluoride is uti-
lized by cariogenic bacteria for fermentation, resulting in the production of
acids. The absorbed/ingested fluoride inside the bacterial cell can interfere
and alter the enzymic activities of the bacterial cell. Thus, bacterial activi-
ties and the resulting acid production get disturbed so that demineralization
of dental mineral gets reduced. Incidentally, the adsorbed fluoride attracts
calcium ions in saliva. Thus, fluoride aids the calcium and phosphate ions
and chemically produces a crystal surface that is more resistant against acid
solubility than the natural tooth mineral, thereby enhancing remineraliza-
tion of the teeth.4–6

3.2.3  Role of Fluoride in Dental Decay


The presence of fluoride in water has dual significance. Since fluoride has
both positive and negative impacts on human health, the stressful effects
of fluoride in drinking water constitute a subject of intense scientific delib-
eration. As a result, many research findings in this direction are either con-
tradicting or inconclusive. The population consuming fluoridated drinking
water increased from 210 million in 1994 to 355 million in 2005.2,7 In addi-
tion, it was estimated that more than 50 million people are consuming water
with a naturally fluoridated concentration of 1 mg/L.7 A systematic review
investigating the association between fluoride and dental caries suggested
that fluoridation of drinking water supplies reduces the occurrence of dental
decay. Furthermore, since fluoride was found to be highly valuable when
continually present at lower concentrations in plaque fluid and saliva, the
WHO recommended the use of fluoridated mouth rinses to reduce elevated
risk of dental caries.2
The effectiveness of fluoride in preventing tooth decay is a topic of intense
scientific debate. Dental caries in children is said to be a bacterial disorder.3
Several factors, namely, nutrition, oral bacteria, oral hygiene, and educational
30 Fluoride in Drinking Water

and economic statuses of parents, are cited as reasons for poor dental health.
Large temporal reductions in tooth decay could be attributed to dietary pat-
terns and immune status of populations. It is suggested that dietary con-
trol of caries without the use of fluoride is possible, as even chewing cheese
reduces tooth decay.8 It is also recommended that fluoride need not be treated
as an essential nutrient to humans, as its absence has not proved to result in
any disease. Since the essentiality of fluoride in humans is not established
unambiguously, data on the minimum nutritional requirement of fluoride
are unvailable.6,9 It was also pointed out that majority of teeth decay issues
develop on fissures and pits of the teeth, areas where fluoride is proved inef-
fective. This further demonstrates the topical action of fluoride on the teeth
surface. Thus, many studies reserve apprehensions in actually consuming or
ingesting fluoride.10

3.3  Dental Fluorosis: History and Occurrence


Dental fluorosis may appear as a cosmetic effect that ranges in appear-
ance from scarcely discernible to a marked staining or pitting of the teeth
in severe forms.11 However, it could be treated as an early sign of fluo-
ride attack that is visible to the naked eye; it also induces an irreversible
toxic effect on tooth-forming cells. Although it histologically represents
a hypocalcification, clinically it ranges from barely visible white stria-
tions on the teeth to gross defects and staining of teeth enamel.11 The first
report on dental fluorosis was from Mexico in 1888, where a family from
Durango was identified with “black teeth.” In 1891, cases of dental fluo-
rosis were reported among Italian migrants from Naples to the United
States. From 1900 onward, there was a pouring in of the prevalence and
issues of dental fluorosis from different parts of the world, including the
United States.12 Dental fluorosis was first related to drinking water in 1925,
though it was shown to be “specifically caused by fluoride in drinking
water much before.” It was Dr. Frederick S. McKay who first reported the
development of an unusual permanent stain or “mottled enamel” on the
teeth surface. This initiated and channelized a lot of scientific research
on the relationship between fluoride in drinking water and fluorosis. In
1930, fluorosis was rated and identified as an “occupational disease in
human.” Subsequently, in 1932, in Denmark, the prevalence of skeletal flu-
orosis in cryolite miners was reported. In the 1930s, in India, fluorosis was
first detected among cattle by the farmers of Nellore district of Andhra
Pradesh. In 1937, the first medical report to this effect was published in
the Indian Medical Gazette. In 1978, it was suggested that “fluorosis might
be one of the most widespread endemic health problems” associated with
natural geochemistry.6,13–15
Dental Fluorosis 31

3.4  Development of Dental Fluorosis


The progression of dental fluorosis is believed to take place through mul-
tiple stages. Initially, teeth become opaque and chalky due to “subsurface
hypomineralization.” Subsequently, the teeth lose enamel, which initiates
the development of grooves and pits. Enamel and dentin, the calcium-rich
constituents of teeth, have much affinity for fluoride in the formation and
development of teeth. During mineralization of teeth, fluoride essentially
combines with calcium to form calcium fluorapatite crystals. Due to such
fluoride accumulation, calcium gets lost from teeth. As a result of this loss of
calcium, fluorosed teeth move from a “mild to severe” state. Since calcium
ions are lost to fluoride, teeth become weaker. Further, the severely fluorosed
enamel becomes more discolored, pitted, porous, and prone to wear and
fracture, as the “well mineralized zone is very fragile to mechanical stress.”
Due to the reduction in mineral content, mutilation of enamel mineraliza-
tion, and associated structural alterations coupled with the morphological
abnormalities on the teeth surface, the fluorosed teeth readily and easily get
fractured.3,16
Most of the recent research suggests that “dental fluorosis results from a
fluoride-induced delay in the hydrolysis and removal of amelogenin matrix
proteins during enamel maturation and subsequent effects on crystal
growth.”6 The proteins secreted by ameloblasts are known as amelogenins.
These amelogenins slow down the growth of enamel crystallites. Once ame-
logeninases remove the amelogenins from the enamel matrix, the crystallite
growth is increased in the initial maturation phase of tooth development.
This most crucial enamel maturation phase turns out to be the most sensi-
tive time for a higher level of fluoride exposure.6,17,18 Depending on the level
of exposure and nutritional status of the child, Dean (1934) classified the
fluorosis on a scale from 0 to 4 as follows:19 “class 0, no fluorosis; class 1, very
mild fluorosis (opaque white areas irregularly covering about 25% of the
tooth surface); class 2, mild fluorosis (white areas covering about 50% of the
tooth surface); class 3, moderate fluorosis (all surfaces affected, with some
brown spots and marked wear on surfaces subject to attrition); and class 4,
severe fluorosis (widespread brown stains and pitting).”6 Figure 3.1 shows
the classification of dental fluorosis symptoms into seven categories accord-
ing to Dean’s classification (normal, questionable, very mild, mild, moder-
ate, moderately severe, and severe), and each of these seven categories was
given a numerical weight between 0 and 4.20 Mild dental fluorosis is mak-
ing its presence felt through the manifestation of small white areas in the
enamel. Teeth that are stained and pitted or mottled in appearance denote
the severe form of dental fluorosis. The hypomineralization of the enamel
is the most prominent feature in human fluorotic teeth. The excessive fluo-
ride assimilation into the enamel may interfere with its maturation p ­ rocess,
resulting in alterations in the rheologic structure of the enamel  matrix
32 Fluoride in Drinking Water

0 0.5

1 1.5

2 3

4 4

FIGURE 3.1
Dental fluorosis symptoms as per Dean’s classification. (From Viswanathan, G., Jaswanth, A.,
Gopalakrishnan, S., Ilango, S. S. and Aditya, G., Sci. Total Environ., 407, 5298–5307, 2009.)

and/or effects on cellular metabolic processes that are associated with nor-
mal enamel development.21,22

3.4.1  Physical Symptoms of Dental Fluorosis


As shown in Figure 3.1, the color of teeth may progress from white, yel-
low, and brown to black due to dental fluorosis. This “discoloration may be
in spots, or appear as streaks invariably horizontal in orientation, as new
layers of the matrix are added on horizontally” during the tooth develop-
ment ­process.3,6 Thus, the discoloration will usually appear in pairs based
on developmental patterns and will not appear as a single isolated tooth.
As against normal belief, the discoloration appears away from the gums,
that is, on the enamel. As it may progress, in due course, the enamel will
lose its brightness, luster, and shine. However, the discoloration due to other
factors (e.g., dirty teeth, smoking, tobacco chewing, coffee or tea stains) may
appear along the gums and only on the periphery of the teeth. Because
“enamel lines are laid down in incremental lines during prenatal and post-
natal periods,” dental fluorosis may appear invariably as horizontal lines or
bands (but never as vertical bands) on the teeth surface.3,6

www.ebook3000.com
Dental Fluorosis 33

FIGURE 3.2
A girl from Garhtipli village in Dhar district in Madhya Pradesh suffering from dental fluo-
rosis. (From UNICEF, Photo essay, UNICEF/India/2006/Ruhani kaur, fluorosis-mitigating the
scourge, http://www.unicef.org/india/1425.html.)

3.4.2  Issues of Dental Fluorosis


Teeth are important components of our facial skeleton, as they are vital in
aesthetics, phonation, and speech. An attractive smile is the biggest human
asset; however, a person with dental fluorosis gets deprived of this privilege.
Dental fluorosis may impart psycho-sociological problems that arise due to
distress, self-insolence, and loneliness. Children with dental fluorosis may be
afraid to laugh, fearing embarrassment, and isolation, which may ruin their
one and only childhood and their self-esteem. Such frustrations may even-
tually lead to deep psychological depressions. The fluoride endemic areas of
the developing world agonizingly accommodate many human beings in this
direction.3,23,24 Figure 3.2 shows a girl from a rural Indian village who is suf-
fering from dental fluorosis.

3.4.3  Prevalence of Dental Fluorosis


The correlation between the concentration of fluoride in water and the
prevalence of dental fluorosis is well documented. As a consequence of
34 Fluoride in Drinking Water

the increased fluoride intake through multiple sources, the severity of


dental fluorosis is increasing the world over. Only 50% of absorbed fluo-
ride is retained in adults, whereas 80% is retained in children. Thus, chil-
dren and adolescents become more susceptible to dental caries. Indian
literature showed 100% prevalence of dental fluorosis at a fluoride level
of 3.4–3.8 mg/L.3,26 The “prevalence of dental fluorosis at a water fluoride
level of 1 mg/L was estimated to be 48% in fluoridated areas and 15% in
­nonfluoridated areas. Limiting consideration to aesthetically important
levels of severity, the prevalence of fluorosis is 12.5% in fluoridated areas
and 6.3% in nonfluoridated areas. Increasing the water fluoride level from
0.4 to 1 mg/L, would mean that one additional person for every 22 people
would have fluorosis of aesthetic concern, but with no risk.”6,27 Thus, it
could be undoubtedly inferred that the benefit of reduction in dental decay
due to fluoride is at the expense of the increased prevalence of dental fluo-
rosis. A 100% prevalence of dental fluorosis was reported at a daily total
fluoride intake of 2.78 mg/child/day. Further, at higher levels of daily total
intake of fluoride, the prevalence of dental caries increased. It is impor-
tant to monitor total fluoride exposure of children and excessive fluoride
intake, especially during the years of tooth development.28 The elevated
fluoride content of the surrounding geological environment also imparts
rich prevalence to dental fluorosis. Children in the age group of 3–14 years
are more likely at the risk of fluorosis from consumption of vegetables and
cereals grown in fluoridated areas.29 This suggests the need for revising
the maximum permissible limits of fluoride in drinking water prescribed
by different regulatory bodies where sizable fluoride intake takes place
through the food chain.
On examining more than 480,000 students covering 18 districts of
Gujarat6 (one of the three worst affected states in India), the percentage
prevalence of dental fluorosis was found to vary from 2.6% to 33%. In
Rajasthan (India), “maximum prevalence of dental fluorosis (77.1%) was
observed among 17–22 year age group with severe dental fluorosis with
black staining at 2.6 mg/L” fluoride.6 The highest overall prevalence of
dental fluorosis (77.2%) was observed at 3.2 mg/L. The study in Haryana
(India) also demonstrates the correlation between dental fluorosis and
dental caries in certain fluoride endemic locations. The “prevalence rate
of dental fluorosis increases from 13% to 77% with increase in fluoride
level from 0.64 mg/L to a range of 1.89–3.83 mg/L.”6 Incidentally, there is
a steady increase in dental caries from 65% to 98%. This experimental evi-
dence substantiates the poor correlation of dental caries with ground­water
fluoride concentration levels. However, in Kolar district of Karnataka,
India, genu valgum was prevalent among children with and without den-
tal fluorosis.30 These observations reveal that the prevalence and occur-
rences of fluorosis widely vary in populations at different regions, though
they consume waters with almost the same fluoride concentrations. Other
Dental Fluorosis 35

factors such as climate, individual biological responses, nutritional sta-


tus, individual vulnerability, extent of fluoride exposure, and presence of
other dissolved salts in the groundwater also have significant roles.3,27,31

3.5 Summary
• Literature suggests that fluoride may be an essential element for
both animals and humans. However, for humans, the essentiality
of fluoride has not yet been demonstrated indisputably. Thus, data
indicating the minimum nutritional requirement of fluoride are
unavailable.
• Many epidemiological studies have demonstrated possible adverse
effects and health issues arising due to long-term ingestion of fluo-
ride through drinking water. These studies clearly show that fluo-
ride primarily produces effects on skeletal tissues, especially on
bones and teeth.
• Considerable literature suggests that the usage of different water
fluoridation techniques, applications of mouth rinses, and fluoride
toothpastes significantly reduce the prevalence of dental caries.
• The acceptable and recommended permissible level of fluoride in
drinking water for preventing dental caries ranges from 0.5 to 1.2
mg/L. This range of limit is fixed based on the annual average maxi-
mum daily air temperature of a region. In general, regulatory agen-
cies suggest a general acceptable maximum fluoride concentration of
1 mg/L in drinking water. This is to ensure the maximum level of pro-
tection of dental caries and to avoid the prevalence of dental fluorosis.
• The ecological imbalance between physiological equilibrium of
tooth minerals and oral microbial biofilms is a causative factor for
dental caries.
• The effectiveness of fluoride in preventing tooth decay is a topic of
intense scientific debate, as many studies reserve apprehensions in
actually consuming or ingesting fluoride.
• Dental fluorosis could be treated as an early sign of fluoride attack
that is visible to the naked eye. It induces an irreversible toxic effect
on tooth-forming cells.
• It could be inferred that fluorosis may be regarded as one of the most
pervasive endemic health problems generated and coupled with nat-
ural geochemistry.
• Most of the recent research suggests that “dental fluorosis results
from a fluoride-induced delay in the hydrolysis and removal of
36 Fluoride in Drinking Water

amelogenin matrix proteins during enamel maturation and subse-


quent effects on crystal growth.”
• Dental fluorosis may impart psycho-sociological problems arising
due to distress, self-insolence, and loneliness. Children with den-
tal fluorosis may be afraid to laugh, fearing embarrassment, and
isolation, which may ruin their one and only childhood and their
self-esteem.
• Indian literature demonstrates a 100% prevalence of dental fluorosis
at a fluoride level of 3.4–3.8 mg/L.
• Considerable literature suggests the need for revising the maximum
permissible limits of fluoride in drinking water prescribed by dif-
ferent regulatory bodies where sizable fluoride intake takes place
through the food chain.

References
1. Petersen, P.E. and Lennon, M.A. (2004). Effective use of fluorides for the preven-
tion of dental caries in the 21st century: The WHO approach. Community Dent.
Oral Epidemiol., 32, 319–321.
2. WHO (2002). Fluorides, Environmental Health Criteria Number, 227. Geneva,
Switzerland: World Health Organization.
3. Susheela, A.K. (2003). A Treatise on Fluorosis, 2nd revised edn, pp. 13–14. New
Delhi, India: Fluorosis Research and Rural Development Foundation.
4. Shellis, R.P. and Duckworth, R.M. (1994). Studies on the cariostatic mechanisms
of fluoride. Int. Dent. J., 44, 263–273.
5. Featherstone, J.D. (1999). Prevention and reversal of dental caries: Role of low
level fluoride, Community Dent. Oral Epidemiol. 27, 31–40.
6. Ayoob, S. and Gupta, A.K. (2006). Fluoride in drinking water: A review on the
status and stress effects. Cri. Rev. Environ. Sci. Technol., 36, 433–487.
7. Lennon, M., Whelton, H., O’Mullane, D. and Ekstrand, J. (2005). Nutrients in
drinking water, Chapter 14. In: Fluoride, Geneva, Switzerland: WHO, 2005.
http://www.who.int/water_sanitation_health/dwq/nutrientschap14.pdf.
8. Diesendorf, M. (1995). How science can illuminate ethical debates: A case study
on water fluoridation. Fluoride, 28, 87–104.
9. WHO (2004). Fluoride in drinking-water, Background document for prepa-
ration of WHO Guidelines for drinking water quality. Geneva, Switzerland:
World Health Organization. http://www.who.int/water sanitation health
/dwq/guidelines/en.
10. IAOMT (2003). Report on Policy position on ingested fluoride and fluo-
ridation. In: David, C.K., (Ed.), International Academy of Oral Medicine and
Toxicology, Orlando. http://www.iaomt.org/documents/IAOMT_Fluoridation
_Position.pdf.
Dental Fluorosis 37

11. MRC (2002). Working Group Report: Water fluoridation and health, MRC: 47.
London: Medical Research Council. http://www.mrc.ac.uk/pdf-publication-
swaterfluoridation report.pdf.
12. WHO (2006). Fluoride in Drinking Water. Fawell, J., Bailey, K., Chilton, J., Dahi, E.,
Fewtrell, L. and Magara, Y. (Eds.), pp. 41–75. London, UK: IWA Publishing.
13. Shortt, H.E., Pandit, C.G. and Raghavachari, T.N.S. (1937). Endemic fluorosis in
the Nellore District of South India. Indian Med. Gaz., 72, 396.
14. McKay, F.S. (1928). Relation of mottled enamel to caries. J. Am. Dent. A., 15,
1429–1437.
15. McKay, F.S. and Black, G.V. (1916). An investigation of mottled teeth: An endemic
developmental imperfection of the enamel of the teeth, heretofore unknown in
the literature of dentistry. Dent. Cosmos., 58, 477–484.
16. Susheela, A.K. and Bhatnagar, M. (1999). Structural aberrations in fluorosed
human teeth: Biochemical and scanning electron microscopic studies. Curr.
Sci., 77, 1677–1680.
17. Chilvers, C. (1983). Cancer mortality and fluoridation of water supplies in 35
U.S. cities. Int. J. Epidemiol., 12, 397–404.
18. Whitford, G.M. (1997). Determinants and mechanisms of enamel fluorosis. Ciba
Found. Symp., 205, 226–245.
19. Dean, H.T. (1934). Classification of mottled enamel diagnosis. J. Am. Dent. Assoc.,
21, 1421–1426.
20. Viswanathan, G., Jaswanth, A., Gopalakrishnan, S., ilango, S.S. and Aditya, G.
(2009). Determining the optimal fluoride concentration in drinking water for
fluoride endemic regions in South India. Sci. Total Environ., 407, 5298–5307.
21. WHO (1994). Fluorides and oral health. Report of a WHO Expert Committee
on Oral Health Status and Fluoride Use. WHO Technical Report Series 846.
Geneva: World Health Organization.
22. Aoba, T. (1997). The effect of fluoride on apatite structure and growth. Crit. Rev.
Oral Biol. Med., 8, 136–153.
23. EPA (1985). National primary drinking water regulations: Fluoride, final rule.
Federal Register, 50, 47142–47155.
24. McKnight, C.B., Levy, S.M., Cooper, S.E. and Jakobsen, J.R. (1998). A pilot study
of esthetic perceptions of dental fluorosis vs. selected other dental conditions.
ASDC J. Dent. Child., 65, 233–238.
25. UNICEF. Photo essay, UNICEF/India/2006/Ruhani kaur, fluorosis-mitigating
the scourge. http://www.unicef.org/india/1425.html
26. Choubisa, S.L., Sompura, K., Bhatt, S.K., Choubisa, D.K., Pandya, H., Joshi, S.C.
and Choubisa, L. (1996). Prevalence of fluorosis in some villages of Dungarpur
district of Rajasthan. Indian J. Environ. Health, 38, 119–126.
27. McDonagh, M., Whiting, P., Bradley, M., Cooper, J., Sutton, A., Chestnutt, I.,
Misso, K., Wilson, P., Treasure, E. and Kleijne, J. (2000). A Systematic Review
of Public Water Fluoridation, NHS Centre for Reviews and Dissemination,
University of York, York YO10 5DD, 2000. http://www.nhs.uk/conditions
/­fluoride/documents/crdreport18.pdf.
28. Xianga, Q., Zhoua, M., Wua, M., Zhoub, X., Linb, L., Huangb, J. and Liangc, Y.
(2009). Relationships between daily total fluoride intake and dental fluorosis
and dental caries, JNMU, 23, 33–39.
38 Fluoride in Drinking Water

29. Jha, S.K., Nayak, A.K. and Sharma, Y.K. (2011). Site specific toxicological risk
from fluoride exposure through ingestion of vegetables and cereal crops in
Unnao district, Uttar Pradesh, India. Ecotoxicol. Environ. Saf., 74, 940–946.
30. Arvind, B.A., Isaac, A., Murthy, N.S., Shivaraj, N.S., Suryanarayana, S.P. and
Pruthvishet, S. (2012). Prevalence and severity of dental fluorosis and genu val-
gum among school children in rural field practice area of a medical college.
Asian Pac J. Trop. Dis., 465–469.
31. Choubisa, S.L. (2001). Endemic fluorosis in southern Rajasthan, India, Research
Report. Fluoride, 34, 61–70.
4
Skeletal Fluorosis

4.1 Introduction
The most consistent and the best characterized toxic response to fluoride
is its effect on the human skeleton, as 99% of the ingested fluoride in the
human body gets stored in bones. The highly vascularized soft tissues and
blood store the remaining fluoride. Once absorbed, the fluoride gets readily
accommodated in the active, growing, and cancellous areas than in compact
regions. The concentration of fluoride stored in various bones of the same
skeleton differs with the type of bones. The pelvis accumulates higher fluo-
ride than the limb bones; young and cancellous bones are more receptive to
fluoride than old or cortical bones. Though factors such as sex, age, and type
and specific part of the bone influence the concentration of fluoride in bones,
fluoride accumulation gets slower with age and reaches an “equilibrium”
effect after about 50 years of age.1,2

4.2  Action of Fluoride on Bone


The chemical composition and the physical structure of human bones get
altered with ingested fluoride. The resorption and accretion of bone tis-
sue are altered by fluoride intake, which in turn affects the homeostasis of
bone mineral metabolism. A combination of osteomalacia, osteosclerosis,
and osteoporosis of different degrees illustrates the bone lesions. Fluoride
toxicity in bones imparts impaired bone collagen synthesis, increased meta-
bolic turnover, and increased avidity for calcium. Fluoride toxicity is also
reflected in the structural changes of bones, namely, “increased bone mass
and density, exostosis (bony outgrowth) at bone surfaces, increased osteoid
seam and resorption surfaces, increased osteon diameter and mottling of
osteons, increased trabecular bone volume, cortical porosity and develop-
ment of cartilaginous lesions in the cancellous bones.”3,4 Fluoride can replace
the hydroxyl and bicarbonate ions that are associated with hydroxyapatite

39
40 Fluoride in Drinking Water

(mineral phase during formation of bone), forming hydroxyfluorapatite and


altering the mineral structure of bones. Thus, the fluoride ions occupying
the plane of the calcium ions form structurally compact and electrostatically
stable structures that may shift the mineralization profile to higher levels
of density and hardness. Thus, the remineralization process increases bone
density, making denser and harder bones. However, high-dose administra-
tion of fluoride over a long period reduces the mechanical strength of bones.5
The interface between collagen and the mineral is one of the causative fac-
tors inducing bone strength. During overexposure for a long term, a rapid
exchange mechanism becomes active between fluoride and hydroxyl ions
in the hydroxyl apatite structure of the bone. This process of exposure may
continue irreversibly. Due to this activity, the rate of synthesis of bone mate-
rial (hydroxyl apatite) gets considerably increased. As a result, the develop-
ment of osteosclerosis (hardening and calcifying of the bones) takes place.
This denotes one of the basic symptoms of people suffering from skeletal
fluorosis. The mechanical properties of bones get drastically influenced by
the increasing fluoride dose into bones. This increase in bone fluoride may
affect the bone matrix proteins and mineral–organic interfacial bonding; it
may “interfere with bone crystal growth inhibition on the crystallite faces as
well as bonding between the mineral and organic interface.” It was pointed
out that long-term fluoride exposure and its cumulative accumulation in the
bones produce “heavier and brittle” bones that are more fragile than normal
bones. However, fluoride has been used for the treatment of osteoporosis,
as it stimulates the formation of bones. Further, fluoride may add to bone
mass, thereby inhibiting bone resorption. However, it has been established
recently that the “beneficial increase in spinal bone mass is at the expense of
an increasing risk of hip fractures.”4 It was suggested that though “fluoride is
the single most effective agent for increasing axial bone volume in the osteo-
porotic skeleton, its therapeutic window is very narrow.”4,6–8

4.3  Fluoride Exposure Level and Skeletal Fracture


Effects on the skeleton are the best indicators of the toxic responses to flu-
oride and are considered to have direct public health relevance. However,
studies on the association between exposure to fluoride and incidence of
hip fractures are either contradicting or inconclusive. Many studies ruled
out a correlation between fluoridated water and prevalence of hip fractures,
whereas some studies reported otherwise. The increase in serum alkaline
phosphatase due to fluoride uptake, referred to as a sign of osteoblast activ-
ity in conventional medicine, is actually a reflection of increased mortality
of osteocytes in bones. After analyzing 29 studies (among these, 18 investi-
gations provided data on hip fracture) dealing with bone fracture and bone
Skeletal Fluorosis 41

development problems, “any association between water fluoridation and


bone or hip fracture incidence was clearly ruled out.”4,9

4.4  Skeletal Fluorosis


Due to prolonged accumulation of fluoride, bones may turn fragile and
exhibit low tensile strength. Such a condition of the bone is referred to as
­skeletal fluorosis. This sorry state of affairs is not easily recognizable until it
reaches the advanced stage. Though in the initial stages it may show symp-
toms of arthritis, it may turn out to be a crippling disability in the most
advanced stages. The restrictions in spine movement could be regarded
as an early warning sign of skeletal fluorosis. Symptoms of ill effects are
detected in the knee, spine, neck, pelvi and shoulder joints.3 There is a steady
increase in stiffness until the entire spine becomes one continuous column
of bone, a condition referred to as poker back. This stage refers to a condition
in which the ligaments of the spine become ossified and calcified. Further,
the stiffness associated with the spine rapidly gets transmitted to assorted
joints in the limbs. The attachment of the ribs steadily reduces the progress
of the chest during breathing. As a result, the “chest assumes a barrel shape.”
It is reported that there is an increasing immobilization of joints due to con-
tractures. This induces flexion deformities at knees, hips, and other associ-
ated joints, forcing the patient to be confined to bed. Despite the fact that
the entire bone structure has become affected, the mental faculties remain
unimpaired until the last stage.4,10
The occurrence of endemic skeletal fluorosis has been reported mainly
from the fluorosis endemic regions of the developing world (Figures 4.1
and 4.2). Increased risk of bone effects at total intakes higher than 6.0 mg
fluoride/day and at a fluoride concentration of 1.4 mg/L are reported from
India and China.1 Skeletal fluorosis was reported from India at an average
fluoride concentration as low as 0.7 mg/L that was within the range of 0.4–
1.4 mg/L. Evidence of skeletal fluorosis with severe clinical manifestations
was reported from areas with a concentration range of 0.7–2.5 mg/L.11 The
bone ash fluoride concentrations range from 500–1500 mg/L among people
exposed to fluoride levels of 1.0 mg/L. Symptoms such as joint stiffness, spo-
radic pain, and osteosclerosis of the pelvis are reported at concentrations
of 6000–7000 mg/L. These symptoms may turn out to be increased osteo-
sclerosis, chronic joint pain, and moderate calcification of ligaments at con-
centrations of 7500–9000 mg/L. The most acute form of fluorosis (crippling
fluorosis) may be observed for fluoride bone concentrations greater than
10,000 mg/L.4 Although most of the ingested fluoride gets excreted through
urine, a portion of the absorbed fluoride gets deposited in skeletal tissues.
More than half of the ingested fluoride in humans is excreted through the
42 Fluoride in Drinking Water

FIGURE 4.1
Children from Jhabua district of Madhya Pradesh suffering from skeletal fluorosis. (From
UNICEF, Photo essay, UNICEF/India/2006/Ruhani kaur, fluorosis-mitigating the scourge,
http://www.unicef.org/india/1425.html.)

FIGURE 4.2
A 10-year-old boy from Jhabua district of Madhya Pradesh with a deformed leg and suffer-
ing from skeletal fluorosis. (From UNICEF, Photo essay, UNICEF/India/2006/Ruhani kaur,
fluorosis-mitigating the scourge, http://www.unicef.org/india/1425.html.)
Skeletal Fluorosis 43

most crucial organ, namely the kidney. However, any failure or malfunc-
tion of the kidneys may reduce this excretion of fluoride, resulting in cumu-
lative fluoride deposition in serum and bones. Renal failure may increase
skeletal fluoride content nearly four times. This may also invite more risk of
spontaneous bone fractures and possibilities of skeletal fluorosis, even at low
concentrations.

4.4.1  Crippling Skeletal Fluorosis


Crippling skeletal fluorosis (Figure 4.3) is a significant cause of morbidity
in a number of regions of the world. This stage is a cumulative outcome of
the most advanced and severe form of skeletal fluorosis. Long-term expo-
sure to high levels of fluoride intake coupled with malnutrition, strenu-
ous manual labor, and impaired renal function leads to this stage. It is
reported that “crippling skeletal fluorosis is marked by kyphosis (abnor-
mally increased convexity in the curvature of the thoracic spine as viewed
from the side), scoliosis (lateral curvature of the vertebral column), flexon
deformity (the act of bending or the condition of being bent) of knee
joints, paraplegia (paralysis of the lower part of the body including lugs)
and quadriplegia (paralysis of all the four limbs).”4 Further, the “pressure
caused by osteophytes (bony outgrowth) or narrowing of intervertebral
foramen and increase in size of the body of the vertebrae or narrowing of
the spinal canal” leads to paralysis.3,4 The two important forms of skeletal
fluorosis observed in India are termed genu valgum (knock-knees) and genu
varum (bow legs). Genu valgum, the most acute form of skeletal fluorosis,
has been observed with osteosclerosis of the spine and associated osteo-
porosis of the limb bones. Patients suffering from fluorosis usually experi-
ence difficulty in walking because of the progressive weakness in the lower
limbs. Neurological disabilities may also occur in some cases along with
the spreading of this weakness to upper limbs. Since crippling fluorosis

FIGURE 4.3
People in the state of Rajasthan, India, suffering from crippling skeletal fluorosis. (From
Hussain, J., Hussain, I. and Sharma, K.C., Environ. Monit. Assess., 162, 1–14, 2010.)
44 Fluoride in Drinking Water

develops slowly and relentlessly, the neurological deficits may cause a


slight trauma. The co existence of skeletal crippling deformities at the knee,
hips, and other joints sends a confusing signal in diagnosis. It is difficult to
ascertain whether such disabilities are actually induced by skeletal defor-
mities or by neurological lesions. Such cases present a “wide spectrum of
neurological deficits, which may be found manifesting either the lower
motor neuron or the upper motor neuron defects, or both, which is much
more common.”4 This refers to a stage in which the “anatomical features of
the cervical spine will be affected and in advanced cases marked cachexia
may develop due to disuse of limb and trunk muscles.”4 Another notice-
able feature at this stage is the occurrence of a progressive high-frequency
perceptive deafness. However, total deafness is a remote possibility. It was
suggested that “the compression of the nerve in the sclerosed and nar-
rowed auditory canal may account for the deafness” that is caused due to
fluorosis.4,13,14 It could be inferred that the acute forms of crippling skeletal
fluorosis have severe bearings on the socio-economic aspects of human life,
as they bring in issues such as loss of work and livelihoods, social aloof-
ness, psychological trauma, and above all, the loss of the will to live.

4.5  Fluoride Level and Effects Related to Skeletal Fluorosis


On a rough estimate, a fluoride dose of 10–20 mg/day (equivalent to 5–10
mg/L of fluoride in water, for an individual ingesting 2 L/day) for at least
10 years may induce crippling skeletal fluorosis.7 However, Indian research
substantiates the prevalence of skeletal fluorosis at very low fluoride concen-
trations of 1.35 and 0.7 mg/L, and crippling skeletal fluorosis at or higher
than 2.8 mg/L.11,16–18 A high prevalence (51.1%) of crippling skeletal fluorosis
(genu valgum) at 9–13 mg/L of fluoride has been reported from the state of
Madhya Pradesh in India.19 Interestingly, the prevalence of skeletal fluorosis
for the same fluoride exposure of 3.0 mg/L was found to be different (19.6%
and 42.2%) in two villages of the state of Punjab. In Rajasthan, a change in
fluoride level from 2.5 to 2.6 mg/L resulted in an increase in the prevalence
of skeletal fluorosis from around 7%–16%. At a fluoride level of 6 mg/L, 63%
was the maximum skeletal fluorosis reported.20 A lower prevalence of skel-
etal fluorosis (around 13%) for a fluoride exposure of 3.2 mg/L is reported in
China, whereas comparatively higher values (around 40%–59%) are reported
for typical Indian villages. In Andhra Pradesh state, a study was conducted
on correlating the influence of the duration of exposure and the age of the
people consuming water at fluoride concentrations of 9 mg/L. It was dem-
onstrated that a daily dose of 36–54 mg of fluoride is ingested by consuming
4.6 L of water.21 As a result, on residing in the village, the skeletal fluorosis
Skeletal Fluorosis 45

could start in 10 years and can reach 100% at 20 years.1,7 All these research
findings demonstrate that the occurrence and prevalence of fluorosis can
vary widely among different locations having almost similar fluoride con-
centrations. It means that apart from the fluoride concentration in drinking
water, the prevalence of fluorosis can be affected by a number of other fac-
tors, as demonstrated in fluoride endemic Indian villages.

4.6  Significance of Other Factors


The relatively high prevalence of fluorosis in India at low fluoride levels
needs attention. As cited, Indian research revelations provide ample evi-
dence on the critical role of malnutrition and poverty in the prevalence and
severity of fluorosis. Vitamin C deficiency and poor nutrition are also found
to enhance the prevalence of fluorosis. The crippling skeletal fluorosis “genu
valgum” is found to be more predominant in poorly nourished children
with a low calcium intake. Such children in the endemic areas of fluorosis
may also develop secondary hyperparathyroidism.10,17 As cited earlier, at the
same fluoride levels, a markedly different incidence of skeletal fluorosis was
reported from India. The lower incidence of fluorosis was later associated
with higher hardness of water. Thus, it appears that calcium and magne-
sium components of hard water play a “protective influence” against fluo-
rosis. In China, studies suggested that prevalence of skeletal fluorosis was
found to vary with the nutritional status of the habitations.22 A study carried
out in the Anantapur district of Andhra Pradesh, the worst ­fluoride-affected
state in India, demonstrated a high content of fluoride (0.2–11.0 mg/kg) in
the locally grown agricultural crops.23 The mean fluoride content of the
brewed teas was three to four times higher than the national mean of the
tap water fluoride level in the United States.24 So, it is plausible to infer that
the contribution of fluoride through various food items may also add to the
reported occurrence of skeletal fluorosis at low fluoride levels.1 Thus, the
fluoride content in food stuff also turns significant. However, variations
in the occurrence of different forms of fluorosis are reported at the same
fluoride exposure levels. These variations could be obviously attributed to
the nutritional status of the people who are exposed to fluoride. Based on
the research findings from China and India, a “clear excess risk” of skeletal
adverse effects is predicted by the WHO for a total fluoride intake of 14 mg/
day. Further, an “increased risk” of skeletal effects higher than 6 mg/day is
also predicted. Recent research suggests that a fluoride level higher than 1.33
mg/L in drinking water may turn out to be a health risk. Every increase of
0.5 mg/L of water fluoride level may result in an increase in the 52 mg/kg of
bone fluoride level during 2–3 years. The Agency for Toxic Substances and
46 Fluoride in Drinking Water

Disease Registry (ATSDR) recommended an optimum fluoride dose level of


4 mg/day for adults. However, it was pointed out that the consumption of
water with fluoride concentrations greater than 0.65 mg/L may be enough to
cross this limit. Thus, drinking water with a fluoride concentration range of
0.5–0.65 mg/L is recommended as a “safe range” for preventing habitations
of India residing in fluoride endemic regions from being placed “at risk.”25

4.7  Recent Developments


A study conducted in China in 2012 (on 40 villages in Yuanmou County
region) turns significant as it cites reduction in fluorosis over the past 20
years. From 1984 to 2006, a reduction of around 22% in dental fluorosis was
reported. Further, over this period, the number of skeletal fluorosis cases
was reduced from 327 to 148. This observed improvement in the reduction
of fluorosis cases was credited with the supply of a low fluoride drinking
water supply system, which has been successfully executed by the local
governments in China over the past two decades.26 This research output
from China26 should be an eye-opener, especially to the local governments
of the fluoride endemic regions of the developing world. The high-risk
areas with fluoride concentration above guideline values should be identi-
fied in every region. Dialogue between the government and local villages
will help identify priority areas for intervention and budgetary allocation
for fluorosis treatment. The weight of scientific evidence clearly suggests
that the provision of safe drinking water supply is the only effective way
for eliminating or reducing fluoride hazards. Even if safe water is pro-
vided to the inhabitants of the fluoride endemic areas, it may take a long
time to eliminate at least the visible impacts of fluorosis in humans. No
doubt, groundwater may remain the primary option for drinking in all
the fluoride endemic remote villages where safe drinking water is still a
distant dream. The ever-increasing demands of water for all other activi-
ties should also be properly accounted for when planning for alternative
water sources. If fluoride testing is not carried out, the common practice
of digging wells for additional water supply may increase the health risks.
Reliable facilities available at a reasonable cost for fluoride tests in the fluo-
ride endemic rural areas present another option that local governments
should utilize in future. Surely, increased understanding and awareness
of the relationship among fluoride levels, duration of exposure, and antici-
pated health impacts would help reduce the health risks. The addition of
new areas in our fluoride maps cautions that fluorosis relief requires long-
term monitoring and raising people’s awareness on the effects of fluoride
on human health.
Skeletal Fluorosis 47

4.8 Summary
• The toxic response to fluoride is mostly reflected in its effects on the
human skeleton, as the bulk of the ingested fluoride in the human
body gets stored in bones. Although the concentration of fluoride
in bones varies with age, sex, and type and specific part of bone, its
accumulation gets slower with age and reaches an equilibrium effect
after about 50 years of age.
• The chemical composition and the physical structure of human
bones get altered with the ingested fluoride. Fluoride toxicity in
bones imparts impaired bone collagen synthesis, increased meta-
bolic turnover, and increased avidity for calcium.
• The increased fluoride dose into bones reduces their mechanical
properties. Overexposure to fluoride and its cumulative build-up in
the bones makes bones “heavier and brittle.”
• It has been suggested that the valuable increase in spinal bone mass
may induce the risk of hip fractures. However, most of the studies
on the relationship between fluoride exposure and the occurrence of
hip fractures are either contradicting or inconclusive.
• Due to prolonged accumulation of fluoride, bones may turn fragile
with a low tensile strength. Such a condition of the bone is referred
to as skeletal fluorosis and is not easily recognizable until it reaches
the advanced stage.
• An increased risk of bone effects at total intakes higher than 6.0 mg
fluoride/day and at a fluoride concentration of 1.4 mg/L are reported
from India and China. Skeletal fluorosis was reported from India at
an average fluoride concentration as low as 0.7 mg/L that is within
the range of 0.4–1.4 mg/L.
• It could be inferred that the crippling skeletal fluorosis is an impor-
tant ground of morbidity in many fluoride endemic regions of the
world. This stage is a cumulative outcome of the most advanced and
severe form of skeletal fluorosis.
• The two important forms of skeletal fluorosis observed in India are
termed genu valgum (knock-knees) and genu varum (bow legs). Genu
valgum, the most acute form of skeletal fluorosis, has been observed
with osteosclerosis of the spine and associated osteoporosis of the
limb bones.
• Skeletal and crippling skeletal fluorosis persuades sharp social
impacts on issues such as psychological shock, loss of work and
employment, social remoteness, and above all, the loss of the will
to live.
48 Fluoride in Drinking Water

• As a rough estimate, a fluoride dose of 10–20 mg/day for at least


10 years may induce crippling skeletal fluorosis. However, Indian
research substantiates the prevalence of skeletal fluorosis at very low
fluoride concentrations of 1.35 and 0.7 mg/L and of crippling skel-
etal fluorosis at or higher than 2.8 mg/L.
• The occurrence, prevalence, intensity, and severity of fluorosis may
vary widely among different regions and habitations having simi-
lar fluoride concentrations. Apart from the fluoride concentration
in drinking water, the prevalence of fluorosis can be affected by a
number of other factors.
• The crippling skeletal fluorosis genu valgum is found to be more pre-
dominant in poorly nourished children with a low calcium intake.
Such children in the endemic areas of fluorosis may also develop
secondary hyperparathyroidism.
• The ATSDR recommended an optimum fluoride dose level of 4 mg/
day for adults. However, the consumption of water with fluoride
concentrations greater than 0.65 mg/L may be enough to cross this
limit.
• Drinking water with a fluoride concentration range of 0.5–0.65
mg/L is recommended as a “safe range” for preventing habita-
tions of India residing in fluoride endemic regions from being
placed “at risk.”
• A fluoride level higher than 1.33 mg/L in drinking water may turn
out to be a health risk. Every increase of 0.5 mg/L of water fluoride
level may result in an increase of 52 mg/kg of bone fluoride level
within a few years.

References
1. WHO (2002). Environmental Health Criteria Number, 227, Fluorides. Geneva,
Switzerland: World Health Organization.
2. WHO (1970). Fluorides and Human Health, Monograph Series, 59, 364. Geneva,
Switzerland: World Health Organization.
3. Susheela, A.K. (2003). A Treatise on Fluorosis. 2nd revised edn. New Delhi, India:
Fluorosis Research and Rural Development Foundation.
4. Ayoob, S. and Gupta, A.K. (2006). Fluoride in drinking water: A review on the
status and stress effects. Cri. Rev. Env. Sci. Technol., 36, 433–487.
5. Chachra, D., Turner, C.H., Dunipace, A.J. and Grynpas, M.D. (1999). The effect
of fluoride treatment on bone mineral in rabbits. Calcif. Tissue Int., 64, 345–351.
6. Diesendorf, M. (1996). Fluoridation: Breaking the silence barrier. In: Brian, M.,
(Ed.), Confronting the Experts, pp. 45–75. New York: State University of New York
Press.
Skeletal Fluorosis 49

7. ATSDR (2003). Report on Toxicological Profile For Fluorides, Hydrogen Fluoride


and Fluorine. Atlanta, GA: U.S. Department of Health and Human Services,
Public Health Service Agency for Toxic Substances and Disease Registry.
8. Dequeker, J. and Declerck, K. (1993). Fluor in the treatment of osteoporosis: An
overview of thirty years clinical research, Department of Rheumatology and
Arthritis, K.U. Leuven, U.Z. Pellenberg, Belgium. Schweiz. Med. Wochenschr.,
123, 2228–2234.
9. McDonagh, M., Whiting, P., Bradley, M., Cooper, J., Sutton, A., Chestnutt,
I., Misso, K., Wilson, P., Treasure, E. and Kleijnen, J. (2000). A Systematic
Review of Public Water Fluoridation. York, UK: NHS Centre for Reviews and
Dissemination, University of York. http://fluoride.oralhealth.org/papers/pdf
/yorkreport.pdf.
10. Reddy, D.R. and Deme, S.R. (2000). Skeletal Fluorosis, http://210.210.19.130
/vmu1.2/dmr/dmrdata/cme/fluorosis/Fluorosis.htm.
11. Diesendorf, M. (1991). The health hazards of fluoridation: A re-examination. Int.
Clin. Nutr. Rev., 10, 304–321.
12. UNICEF. Photo essay, UNICEF/India/2006/Ruhani kaur, fluorosis-mitigating
the scourge. http://www.unicef.org/india/1425.html.
13. Reddy, D.R., Ravishankar, E.V., Prasad, V.S. and Muralidhar, B.N. (1994). Minor
trauma causing major cervical myelopathy in fluorosis, ICRAN 94. Gold Coast,
585–586.
14. Rao, A.B.N. and Siddiqui, A.H. (1962). Some observations on eighth nerve func-
tion in fluorosis. J. Laryng. Otol., 74, 94–99.
15. Hussain, J., Hussain, I. and Sharma, K.C. (2010). Fluoride and health hazards:
Community perception in a fluorotic area of central Rajasthan (India): An arid
environment. Environ. Monit. Assess., 162, 1–14.
16. Jolly, S.S, Singh, B.M., Mathur, O.C. and Malhotra, K.C. (1968). Epidemiological,
clinical and biochemical study of endemic dental and skeletal fluorosis in
Punjab. Brit. Med. J., 4, 427–429.
17. Krishnamachari, K.A. (1976). Further observations on the syndrome of endemic
genu valgum of South India. Ind. J. Med. Res., 64, 284–291.
18. Choubisa, S.L. (2001). Endemic fluorosis in southern Rajasthan, India, Research
Report. Fluoride, 34, 61–70.
19. Chakma, T., Rao, P.V., Singh, S.B. and Tiwary R.S. (2000). Endemic Genu valgum
and other bone deformities in two villages of Mandla district in Central India.
Fluoride, 33, 187–195.
20. Choubisa, S.L., Choubisa, D.K., Joshi, S.C. and Choubisa, L. (1997). Fluorosis
in some tribal villages of Dungapur district of Rajasthan, India. Fluoride, 30,
223–228.
21. Saralakumari, D. and Rao, R.P. (1993). Endemic fluorosis in the village Ralla
Ananthapuram in Andhra Pradesh, an epidemiological study. Fluoride, 26,
177–180.
22. Chaoke, L., Rongdi, J. and Shouren, C. (1997). Epidemiological analysis of
endemic fluorosis in China. Environ. Carcinogen. Ecotoxicol. Rev., 15, 123–138.
23. Rao, V and Mahajan, C.L (1991). Fluoride content of some common south Indian
foods and their contribution to fluorosis, J. Sci. Food. Agric., 51, 275–279.
24. Pehrsson, P.R., Patterson, K.Y. and Perry, C.R. (2011). The fluoride content of
select brewed and microwave-brewed black teas in the United States. J. Food
Compos. Anal., 24, 971–975.
50 Fluoride in Drinking Water

25. Viswanathan, G., Jaswanth, A., Gopalakrishnan, S., Siva Ilango, S. and Aditya,
G. (2009). Determining the optimal fluoride concentration in drinking water for
fluoride endemic regions in South India, Sci. Total Environ., 407, 5298–5307.
26. Chen, H.F., Yan, M., Yang, X.F., Chen, Z., Wang, G.A., Schmidt-Vogt, D., Xu, Y.F.
and Xu, J.C. (2012). Spatial distribution and temporal variation of high fluoride
contents in groundwater and prevalence of fluorosis in humans in Yuanmou
County, Southwest China. J. Hazard. Mater., 235–236, 201–209.
5
Stress Effects of Fluoride on Humans

5.1 Introduction
Clinical and epidemiological studies related with human health or stress
effects denote important sources of data. However, the crux of the fluoride-
related problems relies on the extent of coverage of the affected or sensitive
subpopulations. This is significant for deriving conclusions from the toxico-
logical viewpoint.1 Since the range of safety is frequently unknown, clinical
studies fail to identify effect levels. Thus, when such data are extensively
used, it may be impossible to obtain exceptionally rigorous guideline val-
ues based on the application of unsuitable uncertainty parameters or factors.
However, clinical studies and epidemiological observations habitually con-
stitute a precious provider that is used for assigning a weight of evidence for
a meticulous approach.

5.2  Nonskeletal Fluorosis


The interaction of fluoride with soft tissues, organs, and other systems of the
human body induces nonskeletal fluorosis. As a result, the skeletal muscles,
erythrocytes, gastrointestinal mucosa, ligaments, spermatozoa, and thyroid
glands of humans will be either affected or damaged. Destruction of actin
and myosin filaments in the muscle tissues reduces muscle energy. Due to
this muscle weakness and corresponding loss of muscle energy, fluorosed
patients find themselves unfit for normal routine activities.2

5.3  Fluoride and Cancer


Numerous epidemiological and experimental studies conducted world over
raised concern on the impact of fluoride in drinking water and morbidity or
mortality due to cancer. Though positive correlations were reported by some

51
52 Fluoride in Drinking Water

studies that cited statistically significant associations between fluoridation


index and cancer, many reviews have not accepted such findings.3–7 Since
fluoride can have a mitogenic effect on osteoblasts, it may increase the risk
for osteosarcoma.8 The WHO suggest that apprehensions in this direction
“cannot be casually dismissed.”9 An ecological study conducted in 1991 with
follow-up for up to 35 years of fluoridation that dealt with 125,000 cases of
incident cancers and 2.3 million cancer deaths also ruled out any correlation.
In 1999, while summarizing significant research, the Centers for Disease
Control and Prevention (CDC) inferred that studies to date have produced
“no credible evidence of any correlation between fluoridated drinking water
and increased risk for cancer.”10 In a systematic review conducted in 2000,
while considering 26 studies that explored the association between cancer
incidence and water fluoride exposure, no statistically significant associa-
tion was found to exist between water fluoridation and incidence of c­ ancer.11
While summarizing considerable research in this direction, the WHO con-
cluded that “the weight of evidence” does not support the hypothesis that
“fluoride causes cancer in humans.” Though most ecological studies rule
out the hypothesis of an association, their “considerable limitations pre-
clude firm  conclusions from being drawn regarding the carcinogenicity
of fl
­ uoride” in humans.9,12,13 Osteosarcoma, a rare primary malignant bone
tumor, is considered the sixth leading cancer in children younger than the
age of 15. The annual incidence rate in the United States is 5.4 cases per mil-
lion for men and 4.0 per million for women less than 20 years of age. Though
the set of causes of osteosarcoma still mostly remain unknown, many stud-
ies suggest a possible link between fluoride uptake and increased occurrence
of osteosarcoma in children and adolescents. In a study conducted in the
United States, the age- and sex-adjusted osteosarcoma incidence data among
youths between 5 and 19 years of age are compared with the water fluorida-
tion level. The results of the study provide “no evidence that young males
are at greater risk than females of the same age group to osteosarcoma” from
fluoride in drinking water. Such studies suggest that water fluoridation may
not have any influence on the development of osteosarcoma for either sex
or age group during childhood and adolescence.14 The efforts to correlate
fluoridated water and incidence of cancer rates have not been fruitful due to
a number of inherent challenges faced in such studies. Usually, it takes years
to perhaps decades after exposure to the causal factors for cancer to be diag-
nosed. Further, a large diversity of cancers and their potential causal factors
need more time for careful analysis. In general, studies conducted on human
populations demonstrated contradicting and divergent views on this subject.
There are studies that demonstrate a positive correlation between fluoride
ingestion and osteosarcoma (bone cancer). Many of these are elusive, show-
ing no strong relationships, and some depict negative correlations.15 Fluoride
ingestion might increase kidney and bladder cancer rates, as hydrogen fluo-
ride (a caustic and potentially toxic substance) may be formed under the acid
conditions of urine. Workers of cryolite processing plants experience this
Stress Effects of Fluoride on Humans 53

situation due to chronic occupational exposure to fl­ uoride dust.15 Though


some studies established a relationship between the incidence of kidney and
bladder cancer and the usage of drinking water laced with fluoride, univer-
sal scientific acceptance is yet to be established.

5.4  Fluoride and Gastrointestinal System


Fluoride ingestion produces symptoms of gastric irritation, such as nausea,
vomiting, and gastric pain. Further, fluoride toxicity may produce loss of
appetite, gas formation and nagging pain in the stomach, chronic diarrhea,
chronic constipation, and persistent headache. Other symptoms include
unusual fatigue, loss of muscle power and weakness, excessive thirst and
frequent urination, depression, tingling sensation in fingers and toes, aller-
gic manifestations, and so forth. The formation of hydrofluoric acid in
the acidic environment of the stomach may create irritation of the gastric
mucosa.2,10,16 The exposure level is rated as “slight to moderate” when the
fluoride concentration remains below 4 mg/L. It is suggested that at this
stage, a population of less than 1% of those affected may experience signs of
gastrointestinal issues and may be subjected to gastrointestinal hypersensi-
tivities.15,17 Kidneys are generally acknowledged as the major route for fluo-
ride excretion. Fluoride induces fatal chronic kidney diseases.18 In Children,
exposure to fluoride concentration levels higher than 2.0 mg/L may induce
damage to liver and kidney functions.19 Chronic ingestion of fluoride can
have noncarcinogenic effects on kidneys. Hospital admission rates for uroli-
thiasis (kidney stones) were reported to be higher in areas of higher fluoride
concentrations. In India, patients with clear signs of skeletal fluorosis living
in areas of high fluoride concentrations (3.5–4.9 mg/L) were reported to be
4.6 times more likely to develop kidney stones.20 Urinary fluoride can be use-
ful in public health and epidemiological studies for marking fluoride expo-
sure and intake. High levels of fluoride in urine and serum are observed
in fluorotic patients. Around 30%–50% of fluoride is excreted from urine in
children and is a reflection of the total fluoride intake from multiple sources.
Studies conducted on short- and long-term fluctuating patterns of urinary
fluoride concentration after fluoride ingestion demonstrated that higher fluo-
ride levels in urine are associated with higher fluoride exposure.21
Fluoride has an impact on the thyroid-stimulating hormone production
and may affect the functions of the thyroid gland.2 Fluoride can also bind
with serum calcium, thus probably reducing myocardial contractility and
inviting cardiovascular collapses. The nonulcer dyspeptic symptoms, preva-
lence of still and deformed childbirths are reported among fluorotic patients
in endemic areas of India. The established association between increas-
ing fluoride concentrations and decreasing birth rates is suggestive of the
reproductive properties of fluoride. Though the exact reason was not fully
elucidated, high fluoride ingestion may have an impact on males, including
54 Fluoride in Drinking Water

the morphology and mobility of sperm, or the levels of testosterone, follicle-


stimulating hormones, and inhibin-B.15 A considerable reduction in serum
testosterone levels in people diagnosed with skeletal fluorosis was observed
in India.22 Fluoride can cause pathological changes such as lipid peroxida-
tion and DNA damage in humans,23 which may affect our immune system.
Though some genotoxic effects cannot be excluded, the overall evidence sug-
gests that fluoride is neither genotoxic nor allergetic in humans.24

5.5  Other Health Effects


Fluorine can cause functional and biochemical changes in the nervous system
during pregnancy, as it is capable of crossing the blood–brain barrier and gets
accumulated in the brain tissue prior to birth. Such accumulations of fluoride
in the tissues of the brain will disrupt the synthesis of certain receptors and
neurotransmitters in the cells of the nervous system. Fluoride has a specific
effect on protein synthesis in the brain, entailing degenerative changes in the
neurons, varying the degrees of loss of gray matter, and causing changes in
Purkinje cells in the cerebella cortex. These changes indicate that fluoride can
delay cell growth and division in the cortex and that the reduced number of
mitochondria, microtubules, and vesicles in the synaptic terminal may reduce
efficiency in neuronal connections. Such changes might account for some
of the neurological alterations present in patients with skeletal fluorosis.25,26
Many epidemiological studies have consistently demonstrated a reduction in
children’s intellectual ability (IQ) due to their excessive exposure to fluoride in
drinking water. Meta-analyses dealing with the effect of fluoride exposure in
drinking water on the intelligence of children also displayed a strong negative
relationship between fluoride exposure and kids’ IQ performance. Kidneys,
being an active site of metabolism, excrete considerable fluoride (around 80%)
that is ingested through drinking water and other sources. So, it is plausible that
fluoride concentrations in human urine can systematically reflect the burden of
fluoride coverage in drinking water as an internal exposure index.27 Evidence
developed by studies conducted in high fluoride exposure areas revealed that
urine fluoride concentrations were positively associated with dental fluorosis.
Many literatures have shown that exposure to high levels of fluoride in drink-
ing water is associated with deficits in children’s intelligence.28,29 It is known
that the biochemical changes induced by fluoride in proteins and associated
enzymatic systems may interfere with normal brain function. This may lead
to impaired cognition and memory. Fluoride may adversely affect the reac-
tion response times and visuospatial capabilities, which will get manifested
as reduced IQ scores in time-sensitive tests.30 Compared with normal children,
the chance of those with severe dental fluorosis getting lower IQ scores is the
most likely. This clearly indicates that exposure to high levels of fluoride in
drinking water has a negative impact on the dental health and intelligence of
Stress Effects of Fluoride on Humans 55

children. Dose–response relationships exist between urine fluoride concentra-


tions and IQ scores as well as between fluoride levels and dental fluorosis.
Even a small decline in IQ scores or dental fluorosis can induce a profound
influence on individuals.27
Studies suggest abnormalities in the ECG of people living in fluoride
endemic areas that are afflicted with skeletal and dental fluorosis. The elastic
properties of the ascending aorta of such patients are found to be damaged.
Fluorosis is found to have an impact on the cardiovascular system, including
the heart and major vessels originating from the heart. It is also observed
that fluorosis patients have left ventricular diastolic and global dysfunctions.
Acute fluoride toxicity will also induce enhanced oxidative stress in human
beings. Acute or cronic stage of fluorosis may produce reactive oxygen which
induces severe damage or even result in the death of myocardial cells. These
impacts in these cells invite issues related with left ventricular diastolic.31
The study conducted in 2013 in Zhaozhou county from Heilongjiang prov-
ince in China confirmed the relationship between excess fluoride intake
and essential hypertension in adults. It was also demonstrated that high
levels of fluoride exposure in drinking water increase plasma ET-1 levels
in people living in fluoride endemic areas.32 Of late, many scientific stud-
ies clearly underlined the fact that fluoride is capable of inducing oxidative
stress. Further, it may transform intracellular redox homeostasis, protein
carbonyl content, and oxidative degradation of lipids; amend gene expres-
sion; and originate apoptosis. Genes related with metabolic enzymes, cell
cycle, stress response, signal transduction, and cell-to-cell communication
are modulated by the presence of fluoride. Biologically relevant concentra-
tions of fluoride are capable of increasing cell migration in tumor cells and
of stimulating tumor invasion.33 Fluoride causes disturbances in the lipid
metabolism in the blood of patients who are afflicted with skeletal fluorosis.
Due to the inhibition of lipid synthesis by fluoride, the cholesterol content
(both high-density lipoprotein [HDL] and low-density lipoprotein [LDL])
gets reduced.34

5.6 Summary

• The interaction of fluoride with soft tissues, organs, and other sys-
tems of the human body induces nonskeletal fluorosis. As a result,
the skeletal muscles, erythrocytes, gastrointestinal mucosa, liga-
ments, spermatozoa, and thyroid glands of humans will be either
affected or damaged.
• Though the set of causes of osteosarcoma still mostly remain
unknown, many studies suggest a possible link between fluoride
uptake and increased occurrence of osteosarcoma in children and
56 Fluoride in Drinking Water

adolescents. Though most of the ecological studies rules out the


hypothesis of an association, their considerable limitations preclude
firm conclusions from being drawn regarding the carcinogenicity of
fluoride in humans.
• Though some studies established an association between fluoridated
drinking water and the incidence of kidney and bladder cancer, uni-
versal scientific acceptance is yet to be established.
• Fluoride ingestion produces symptoms of gastric irritation, such
as nausea, vomiting, and gastric pain. Chronic ingestion of fluo-
ride can have noncarcinogenic effects on the kidneys. People with
clear signs of skeletal fluorosis living in areas of high fluoride
concentrations are at a high risk of developing kidney stones.
• It is plausible that fluoride concentrations in human urine can sys-
tematically reflect the burden of fluoride coverage in drinking water
as an internal exposure index. Evidence developed by studies in
high fluoride exposure areas revealed that urine fluoride concentra-
tions were positively associated with dental fluorosis.
• Fluoride may adversely affect the reaction response times and
visuospatial capabilities, which will get manifested as reduced IQ
scores in time-sensitive tests. This clearly indicates that exposure to
high levels of fluoride in drinking water has a negative impact on the
dental health and intelligence of children.
• Acute fluoride toxicity will also induce enhanced oxidative stress in
human beings.
• Fluoride causes disturbances in the lipid metabolism in the blood
of patients who are afflicted with skeletal fluorosis. Due to the
inhibition of lipid synthesis by fluoride, the cholesterol content
gets reduced.

References
1. WHO. (2009). World Health Organization Guidelines for Drinking-Water
Quality Policies and Procedures Used in Updating the WHO Guidelines for
Drinking Water Quality Public Health. WHO/HSE/WSH/09.05. Geneva,
Switzerland: Public Health and the Environment, WHO Press.
2. Susheela, A.K. (2003). A Treatise on Fluorosis, Revised 2nd edn. New Delhi, India:
Fluorosis Research and Rural Development Foundation.
3. Cohn, P.D. (1992). An Epidemiologic Report on Drinking Water and Fluoridation,
pp. 1–17. Trenton, NJ: Fluoride and Osteosarcoma in New Jersey, Environmental
Health Service.
4. Freni, S.C. and Gaylor, D.W. (1992). International trends in the incidence of bone
cancer are not related to drinking water fluoridation. Cancer, 70, 611–618.
Stress Effects of Fluoride on Humans 57

5. Takahashi, K., Akiniwa, K. and Narita, K. (2001). Regression analysis of


cancer incidence rates and water fluoride in the U.S.A. based on IACR/
IARC (WHO) data (1978–1992), International Agency for Research on
Cancer. J. Epidemiol., 11, 170–179.
6. IARC. (1987). IARC Monographs on Evaluation of Carcinogenic Risks of Chemicals to
Humans. Supplement 7: Overall Evaluations of Carcinogenicity: An Updating
of IARC Monographs, Vols. 1–42. Lyon: International Agency for Research on
Cancer.
7. Knox, E.G. (1985). Fluoridation of Water and Cancer: A Review of the
Epidemiological Evidence. London, UK: Working Party on the Fluoridation of
Water and Cancer. HMSO.
8. MRC. (2002). Working Group Report: Water Fluoridation and Health, MRC:
47. London: Medical Research Council, 2002. http://www.mrc.ac.uk/pdf
-­publicationswaterfluoridation report.pdf.
9. WHO. (2002). Fluorides, Environmental Health Criteria Number, 227. Geneva,
Switzerland: World Health Organization.
10. Ayoob, S. and Gupta, A.K. (2006). Fluoride in drinking water: A review on the
status and stress effects. Crit. Rev. Environ. Sci. Technol., 36, 433–487.
11. McDonagh, M., Whiting, P., Bradley, M., Cooper, J., Sutton, A., Chestnutt, I.,
Misso, K., Wilson, P., Treasure, E. and Kleijnen, J. (2000). A Systematic Review of
Public Water Fluoridation. York, UK: NHS Centre for Reviews and Dissemination,
University of York. http://fluoride.oralhealth.org/papers/pdf/yorkreport.pdf.
12. ATSDR. (2003). Report on Toxicological Profile for Fluorides, Hydrogen Fluoride
and Fluorine. Atlanta, GA: U.S. Department of Health and Human Services,
Public Health Service Agency for Toxic Substances and Disease Registry.
13. IPCS. (2002). International Programme on Chemical Safety. Fluorides,
Environmental Health Criteria, 227. Geneva, Switzerland: World Health
Organization.
14. Levy, M. and Leclerc, B.S. (2012). Fluoride in drinking water and osteosarcoma
incidence rates in the continental United States among children and adoles-
cents. Cancer Epidemiol., 36, e83–e88.
15. Ozsvath, D.L. (2009). Fluoride and environmental health: A review. Rev.
Environ. Sci. Biotechnol., 8, 59–79.
16. Susheela, A.K. (1989). Fluorosis: Early warning signs and diagnostic test. Bull.
Nutr. Found. India, 10(2).
17. Doull, J., Boekelheide, K., Farishian, B.G., Isaacson, R.L., Klotz, J.B., Kumar, J.V.,
Limeback, H., Poole, C., Puzas, J.E., Reed, N.M.R., Thiessen, K.M. and Webster,
T.F. (2006). Fluoride in Drinking Water: A Scientific Review of EPA’s Standards,
p. 530. Committee on Fluoride in Drinking Water, Board on Environmental
Studies and Toxicology, Division on Earth and Life Sciences, National Research
Council of the National Academies. Washington, DC: National Academies
Press. http://www.nap.edu.
18. Chandrajith, R., Dissanayake, C.B., Ariyarathna, T., Herath, H.M. and Padmasiri,
J.P. (2011). Dose-dependent Na and Ca in fluoride-rich drinking water-Another
major cause of chronic renal failure in tropical arid regions. Sci. Total Environ., 409,
671–675.
19. Xiong, X., Liu, J., He, W., Xia, T., He, P., Chen, X., Yang, K. and Wang, A. (2007).
Dose-effect relationship between drinking water fluoride levels and damage
to liver and kidney functions in children. Environ. Res., 103, 112–116.
58 Fluoride in Drinking Water

20. Singh, P.P., Barjatiya, M.K., Dhing, S., Bhatnagar, R., Kothari, S. and Dhar, V.
(2001). Evidence suggesting that high intake of fluoride provokes nephrolithia-
sis in tribal populations. Urol. Res., 29, 238–244.
21. Liu, H.Y., Chen, J.R., Hung, H.C., Hsiao, S.Y., Huang, S.T. and Chen, H.S. (2011).
Urinary fluoride concentration in children with disabilities following long-
term fluoride tablet ingestion. Res. Dev. Disabil., 32, 2441–2448.
22. Susheela, A.K. and Jethanandani, P. (1996). Circulating testosterone levels in
skeletal fluorosis patients. Clin. Toxicol., 34, 183–189.
23. Wang, A.G., Xia, T., Chu, Q.L., Zhang, M., Liu, F., Chen, X.M. and Yang, K.D.
(2004). Effects of fluoride on lipid peroxidation, DNA damage and apoptosis in
human embryo hepatocytes. Biomed. Environ. Sci., 17, 217–222.
24. NRC. (1993). Health Effects of Ingested Fluoride, Commission on Life Sciences,
pp. 51–72, 125–128. Washington, DC: National Academy Press. http://www.nap
.edu/books/030904975X/html.
25. Valdez-Jiménez, L., Soria Fregozo, C., Miranda Beltrán, M.L., Gutiérrez
Coronado, O., Pérez Vega, M.I., (2011). Effects of the fluoride on the central ner-
vous system. Neurología, 26, 297–300.
26. Du, L., Wan, C., Cao, X. and Liu, J., (1992). The effect of fluorine on the develop-
ing human brain. Chin. J. Pathol., 21, 218–220.
27. Ding, Y., Gao, Y., Sun, H., Han, H., Wang, W., Ji, X., Liu, X. and Sun, D. (2011). The
relationships between low levels of urine fluoride on children’s intelligence,
dental fluorosis in endemic fluorosis areas in Hulunbuir, Inner Mongolia,
China. J. Hazard. Mater., 186, 1942–1946.
28. Trivedi, M.H., Verma, R.J., Chinoy, N.J., Patel, R.S. and Sathawara, N.G. (2007).
Effect of high fluoride water on intelligence of school children in India. Fluoride,
40, 178–183.
29. Wang, S.X., Wang, Z.H., Cheng, X.T., Li, J., Sang, Z.P., Zhang, X.D., Han, L.L.,
Qiao, X.Y., Wu, Z.M. and Wang, Z.Q. (2007). Arsenic and fluoride exposure in
drinking water: children’s IQ and growth in Shanyin county, Shanxi province,
China. Environ. Health Perspect., 115, 643–647.
30. Calderon, J., Blenda, M., Marielena, N., Leticia, C., Deogracias, O.M., Diaz-
Barriga, F. (2000). Influence of fluoride exposure on reaction time and visuospa-
tial organization in children. Epidemiology, 11, S153.
31. Varol, E., Akcay, S., Ersoy, I.H., Koroglu, B.K. and Varol, S. (2010). Impact of
chronic fluorosis on left ventricular diastolic and global functions. Sci. Total
Environ., 408, 2295–2298.
32. Sun, L., Gao, Y., Liu, H., Zhang, W., Ding, Y., Li, B., Li, M. and Sun, D. (2013). An
assessment of the relationship between excess fluoride intake from drinking
water and essential hypertension in adults residing in fluoride endemic areas.
Sci. Total Environ., 443, 864–869.
33. Mendoza-Schulz, A., Solano-Agama, C., Arreola-Mendoza, L., Reyes-Márquez,
B., Barbier, O., Del Razo, L.M. and Mendoza-Garrido, M.E. (2009). The effects
of fluoride on cell migration, cell proliferation, and cell metabolism in GH4C1
pituitary tumour cells. Toxicol. Lett., 190, 179–186.
34. Bhardwaj, M. and Shashi, A. (2012). Dose effect relationship between high fluo-
ride intake and biomarkers of lipid metabolism in endemic fluorosis. Biomed
Prev Nutr., 3, 121–127. http://dx.doi.org/10.1016/j.bionut.2012.10.006.
6
Fluoride in the Environment and
Its Toxicological Effects

6.1 Introduction
Fluorine is the most electronegative and the most reactive element present in
the periodic table. It is always found in the nature with the combination of
some other elements because of its reactivity. Fluoride is a naturally occur-
ring, widely distributed element, and it is found in varying amounts in min-
erals, rocks, gases from volcanoes, and so forth. Anthropogenic sources such
as coal-fired power plants; aluminum smelters; phosphate fertilizer plants;
glass, brick, and tile works; and plastics factories are also responsible for
the increase in the level of the fluorine in the atmosphere.1,2 As stated in pre-
vious chapters, human beings are benefited when the fluoride uptake is of
an appropriate quantity. If the fluoride uptake is lower than the minimum
amount or more than the upper limit, it can adversely affect human health.
Studies have shown that apart from human beings, plants, and animals also
experience toxicological issues when they are exposed to fluoride. Drinking
water serves as the major source of fluoride in humans; however, other envi-
ronmental sources also contribute to it.2

6.2  Sources of Environmental Exposure


Fluorine is present in the environment mostly as fluorides. Fluoride is
released into the environment either naturally or through human activi-
ties. The natural processes that are responsible for fluoride intrusion into
the environment may include weathering and dissolution of minerals, volca-
nic eruption, and contribution of marine aerosols.3 Fluoride is also released
into the environment by manifold human activities, for example, coal com-
bustion; processing of water and waste from various industrial processes,
including steel manufacture; aluminum, copper, and nickel production;

59
60 Fluoride in Drinking Water

phosphate ore processing; phosphate fertilizer production and use; glass,


brick, and ceramic manufacturing; glue and adhesive production; and so
forth. Among these, industrial sources such as phosphate ore production
and use, and aluminum manufacture are the major anthropogenic sources
of fluoride.4,5 Controlled addition of fluoride to a public water supply, food
items, dental products, and pharmaceutical products also contribute to the
release of fluoride into the environment through anthropogenic sources.6 A
number of fluoride compounds that are mostly used for industrial and com-
mercial purposes are also potential sources of fluoride in the environment.
In this direction, some important fluoride compounds are listed as follows:2,4

• Hydrogen fluoride is predominantly used in the production of


aluminum and chlorofluorocarbons (CFCs). It is used in electronic
industries, cleaning glass, brick and tiles work, tanning leather, and
in commercial rust removers and pickling operation in stainless
steel. It has wide application in separating uranium isotopes or as a
catalyst in the petroleum industry.
• Calcium fluoride has multiple uses in the fiber glass, ceramic, weld-
ing rod, glass, and fluorescent lamps industry. It is used in blending
with burned lime and dolomite in the steel industry, and it is also a
fluxing agent for aluminum metallurgy.
• Sodium fluoride has wide use in fluoridation of drinking water. It is
used as a preservative in glues and wood, and as an insecticide. It is
also used as a flux in steel and aluminum production, as well as in
glass and enamel production.
• Sulfur hexafluoride is used as a gaseous dielectric medium in the
electrical industry for high-voltage circuit breakers, switch gear, var-
ious electronic components, and so forth. It has some application in
the production of magnesium and aluminum.
• Fluorosilicic acid and sodium hexafluorosilicate are used for the flu-
oridation of drinking-water supplies.

6.3  Environmental Transport, Distribution, and Transformation


In the atmosphere, fluorides exist in gaseous or particulate form. Gaseous
fluorides can be either transported over long distances by the effect of wind
or atmospheric turbulence or absorbed by the atmospheric water (rain,
clouds, fog, snow), forming an aerosol or a fog of aqueous hydrofluoric acid.
They can be removed from the atmosphere via wet deposition. Fluoride pres-
ent in the particulate form can be removed from the atmosphere and depos-
ited on land or surface water by wet and dry deposition.7 Most of the fluoride
Fluoride in the Environment and Its Toxicological Effects 61

compounds are not expected to remain in the troposphere for long periods or
to migrate to the stratosphere. However, sulfur hexafluoride can reside in the
atmosphere for a period ranging from 500 years to several thousand years.2
Several factors influence the process of the transport, and transformation of
fluoride in water. Some of the important factors are pH, water hardness, and
the presence of ion-exchange materials such as clays. Fluorides usually com-
bine with aluminum during their transportation and transformation process
through water cycle. In soils, the transport and transformation of fluoride are
influenced by pH and the formation of complexes, predominantly those with
aluminum and calcium. When the soil is slightly acidic (pH 5.5–6.5), fluoride
may be adsorbed more strongly by the soil particles. Fluorides strongly bond
with the soil and are not easily leached from it. Terrestrial plants may accu-
mulate fluorides from the atmosphere through the opening of stomata and
from the soil via roots. Some aquatic plants and animals can bioaccumulate
the soluble fluorides from water. The quantity of fluoride uptake depends on
the route of exposure, how fluoride is absorbed by the body and how quickly
it is consumed and excreted.

6.4  Environmental Levels and Human Exposure


The level of fluoride in surface water such as rivers, lakes, and so forth
depends on its location and distance from the emission source. Fluoride
concentration in the range between 0.01 and 0.3 mg/L is found in surface
water,2 whereas sea water contains more fluoride than fresh water and it
ranges between 1.2 and 1.5 mg/L.2 Fluoride concentration in groundwa-
ter is very high in areas where the natural rocks and soil are rich in fluo-
ride. Inorganic fluoride concentration is very high in the regions where
there is geothermal or volcanic activity. Anthropogenic discharges are also
one of the major sources that increase the concentration of fluoride in the
environment.
The fate and transport of gaseous and particulate fluorides in the environ-
ment occurring from natural and anthropogenic sources depend on various
atmospheric conditions. The transportation, distribution, transformation,
and deposition of airborne fluorides are governed by meteorological con-
ditions, particulate size, chemical reactivity, and emission strength of the
source. Fluorides that are released as gaseous and particulate matter are gen-
erally deposited in the vicinity of an emission source, whereas some particu-
lates may react with other atmospheric constituents. It has been found that
the fluoride concentrations in the ambient air in areas that do not have any
nearby emission sources2 are generally less than 0.1 μg/m3. However, even
in the areas that have nearby emission sources, the concentration of levels of
airborne fluoride2 usually do not exceed 2–3 μg/m3.
62 Fluoride in Drinking Water

Fluoride is present in various concentrations in different types of soils. The


fluoride concentrations in the areas without natural phosphate and fluoride
deposits range from 20 to 1000 μg/g, whereas the concentration may increase
to several thousand micrograms per gram for the soil that is enriched with
natural deposits of fluoride.2 Many foodstuffs contain very small amounts
of fluoride; however, very high amounts of fluoride are reported in fish and
tea leaves. The exposure of individuals to fluorides varies considerably. It
depends on the level of fluoride in drinking water or dietary intake, the use
of fluoridated dental/pharmaceutical products and, in some cases, the levels
of fluoride in indoor air. The inhalation of airborne fluoride generally contrib-
utes a minor stake in the total fluoride (TF) intake. Infants who are fed with
formula receive 50–100 times more fluoride than infants who are fed exclu-
sively breast milk. In the case of young children, the ingestion of toothpaste
containing fluoride contributes a significant amount to their TF intake. The
intake of fluoride in adults is mainly through foodstuffs and drinking water.
Generally, 2 mg/day of fluoride finds its way into the body of children and
adolescents.2 Although adults may have a higher absolute daily intake of fluo-
ride in terms of milligrams, the daily fluoride intake of children, expressed
on a milligram-per-kilogram body weight basis, may exceed that of adults.2
Occupational individuals working in industries such as aluminum, fertilizer,
iron, oil refining, semiconductor, phosphate, ore, and steel are more suscepti-
ble to get exposed to fluoride via inhalation or dermal contact. Recent studies2
reported that the concentration of fluoride in the pot rooms generated from
aluminum smelters is found to be in the order of 1 mg/m3.

6.4.1  Fluoride from Dental Products


It has been established that different dental and pharmaceutical products
that are used by people in their daily life for oral hygiene contain fluoride.
These include toothpastes (1.0–1.5 g/kg fluoride), gels (0.25–24.0 g/kg fluo-
ride), fluoride tablets (0.25–1.00 mg fluoride/tablet) and so forth.6 Fluorides in
the form of sodium fluoride (NaF), sodium monofluorophosphate (Na2FPO3)
and tin fluoride (SnF2), or in the form of different amines can be added to
toothpastes. Dental products such as toothpaste and tooth powder may also
contribute to the TF intake. In pharmaceutical products, fluoride can be
added in the form of NaF or Na2FPO3. The intake of fluoride through den-
tal products varies from person to person on the basis of their practice and
exercise of cleaning teeth. Generally, small children are more susceptible to
ingestion of toothpaste while cleaning their teeth. Studies on toothpaste sug-
gest that during cleaning of the teeth, most children ingest approximately
20% of the toothpaste. It has been found that ingesting the toothpastes con-
taining fluoride may result in a contribution of about 0.50–0.75 mg of fluoride
per child per day.7–9 Some of the researchers conducted studies to determine
the fluoride concentration in dental products. Kumar and Yadav (2014) esti-
mated the fluoride concentration in toothpastes and tooth powder available

www.ebook3000.com
Fluoride in the Environment and Its Toxicological Effects 63

in the market of Rampur district, Uttar Pradesh, India.10 The water-soluble


fluorides in the toothpastes were found to vary from 20 to 1100 mg/kg and
those in tooth powders ranged from 9.11 to 22.11 mg/kg.10 Yadav et al. (2007)
collected 15 samples of toothpaste and determined the fluoride concentra-
tion present in them.11 It was found that one gram of toothpaste contributes
approximately 53.5–338.5 µg of fluoride to the human body with a mean
value of 183.78 µg.11

6.4.2  Fluoride from Food and Beverage


Fluorides may find their way into the human body through diet. Generally,
food contains low levels of fluoride, though a trace amount of fluoride may
be present in all foodstuffs. However, fluoride may be present in higher
amounts in the food grown in areas where soils contain higher amounts of
fluorides or where phosphate fertilizers are used for agricultural purposes.
Tea and some seafood are reported to have very high levels of fluoride. A
summary of various studies in which the levels of fluoride in foods have
been assessed is presented in Table 6.1.
The concentration of fluoride in food items mainly depends on the
nature of soil and the quality of water used for irrigation; thus, it varies
from place to place. Several studies were conducted in the past to deter-
mine the intake of fluoride through the food items. Most of the studies
suggested that the concentration of the fluoride in the raw food items is
the function of the fluoride content in the water that is used for irrigation.
Some of the studies related to the fluoride concentration in Indian food
items with reference to the Indian context are interesting.
Rao and Mahajan (1990) conducted a survey in 41 villages of Anantapur
district, Andhra Pradesh, India, to determine the fluoride concentration in 98
different food items.33 In this area, a fluoride concentration up to 4.5 mg/kg
was reported in irrigation water and this fluoride finds its way into most of the
food items that are grown in this area. It was estimated that 32 locally grown
food items generally had a fluoride concentration ranging from 0.2 to 11.0 mg/
kg. The TF intake from both food and drinking water was found to be in the
range of 2.2–7.3 mg/d (0.05–0.32 mg/d/kg body weight): Food contributes a
large fraction to TF intake, ranging from 1.3 to 3.4 mg/d.33 Gautam et al. (2010)
estimated the concentration of fluoride in the food items collected from Nawa
tehsil in Nagaur district of Rajasthan.34 The fluoride content in water was
found to be in the range of 0.92–14.62 mg/L. Leafy vegetables grown in fluo-
ride endemic areas were found to be the most susceptible to fluoride and they
contained fluoride in the range of 8.08–25.70 mg/kg. Fluoride concentrations
in cereal crops were also high (ranging from 1.88 to 18.98 mg/g).34 Bhargava
and Bhardwaj (2009) conducted a study in 10 endemic villages of north-east
Rajasthan to determine the fluoride levels in different local food items.35 Food
items such as vegetables, cereals, fodder, and milk were collected from the
fluoride endemic villages and analyzed for fluoride content. It was reported
64 Fluoride in Drinking Water

TABLE 6.1
Fluoride Concentration in Different Foodstuffs
Fluoride
Type of Concentration
Food Test Insights (mg/kg)a Study Area References
Milk and 30 samples of milk 0.23–1.36 Connersville and Jackson et al.12
milk and milk the Richmond
products products community,
Indiana (USA)
42 different types 0.007–0.068 Houston, Texas, Liu et al.13
and brands of USA
milk
66 cow milk 0.043–0.147 Dindigul district, Amalraj and
samples Tamil Nadu, Pius14
India
Pasteurized milk 0.143–0.157 United kingdom Duff 15
supplied to (non-
households fluoridated
area)
Untreated milk 0.162–0.173 United kingdom Duff 15
samples (where the farm
water supply
was fluoridated
at 1 mg/L level)
68 samples of 0.007–0.086 Canada Dabeka and
market milk McKenzie16
Chocolate-flavored 0.05−1.27 Bauru Buzalaf et al.17
milk for infant municipality,
Brazil
Soy beverages for 0.09−0.29 Bauru Buzalaf et al.17
infant municipality,
Brazil
Meat and Mechanically 0.08–8.63 Corvallis, Fein and
poultry separated chicken Oregon, USA Cerklewski18
and turkey
9 kinds of deboned 0.3–2.7 Poland Jedra et al.19
poultry meat
55 samples of meat 0.03–1.41 Connersville and Jackson et al.12
and poultry the Richmond
community,
Indiana (USA)
25 ready to eat 0.01–8.38 Iowa City, Iowa Heilman et al.20
samples of meats
and chicken for
infant
Fish Range of fluoride 45–1207 – Camargo21
levels in skeletal
bone of saltwater
fish
(Continued)
Fluoride in the Environment and Its Toxicological Effects 65

TABLE 6.1 (Continued)


Fluoride Concentration in Different Foodstuffs
Fluoride
Type of Concentration
Food Test Insights (mg/kg)a Study Area References

Range of fluoride 1.3–26 – Camargo21


levels in muscle
of saltwater fish.
Different variety of 2.10–11.10 Spain Rocha et al.22
fishes samples
included in the
Valencian
Community Total
Diet Study
3 different species 2.35–274.29 Alappuzha Thomas and
of fish (The district, Kerala, James23
fluoride India
concentration in
the water samples
from where fishes
were caught were
in the range of
0.035–0.051
mg/L)
Baked 129 samples of 0.07–1.36 Connersville and Jackson et al.12
goods and uncooked grain the Richmond
cereals products community,
Indiana (USA)
528 staple food 1.16–4.94 Dindigul district, Amalraj and
grain samples Tamil Nadu, Pius14
India
66 cooked rice 0.34–0.73 Dindigul district, Amalraj and
samples Tamil Nadu, Pius14
India
Rice samples 2.20 Spain Rocha et al.22
included in the
Valencian
Community Total
Diet Study
9 ready to eat 0.01–0.31 Iowa City, Iowa Heilman et al.20
samples of cereals
for infant
Cereals for infant 0.20−7.84 Bauru Buzalaf et al.17
municipality,
Brazil
Biscuits for infant 0.34−13.68 Bauru Buzalaf et al.17
municipality,
Brazil
(Continued)
66 Fluoride in Drinking Water

TABLE 6.1 (Continued)


Fluoride Concentration in Different Foodstuffs
Fluoride
Type of Concentration
Food Test Insights (mg/kg)a Study Area References

Vegetables 65 samples of 0.38–5.37 Warsaw market, Sawilska-


vegetables Poland Rautenstrauch
et al.24
660 green leafy 0.58–7.68 Dindigul district, Amalraj and
vegetable Tamil Nadu, Pius14
samples India
Samples of various 0.74–4.76 Spain Rocha et al.22
vegetable
included in the
Valencian
Community Total
Diet Study
48 ready to eat 0.01–0.42 Iowa City, Iowa Heilman et al.20
samples of
vegetables for
infant
78 samples of 0.003–1.930 Connersville and Jackson et al.12
uncooked the Richmond
vegetables community,
Indiana (USA)
Fruits and Different fruit juice 0.07–0.53 Davangere city, Thippeswamy
fruit juices samples India et al.25
105 juice samples 0.67 Mexico City, Jimenez-Farfan
Mexico et al.26
88 ready to eat 0.01–0.49 Iowa City, Iowa Heilman et al.20
samples of fruits
and desserts for
infant
26 samples of 0.01–0.84 Connersville and Jackson et al.12
fruits the Richmond
community,
Indiana (USA)
Fats and 14 samples of fats 0.05–0.62 Connersville and Jackson et al.12
oils and oils the Richmond
community,
Indiana (USA)
Sugars and 15 samples of 0.07–0.60 Connersville and Jackson et al.12
candies sugar and sweets the Richmond
community,
Indiana (USA)
Beverages 32 samples of 0.04–0.93 Connersville and Jackson et al.12
beverages the Richmond
community,
Indiana (USA)
(Continued)
Fluoride in the Environment and Its Toxicological Effects 67

TABLE 6.1 (Continued)


Fluoride Concentration in Different Foodstuffs
Fluoride
Type of Concentration
Food Test Insights (mg/kg)a Study Area References

12 different 0.19–0.42 Davangere city, Thippeswamy


carbonated soft India. et al.25
drinks
57 carbonated 0.43 Mexico City, Jimenez-Farfan
drinks Mexico et al.26
332 soft drinks 0.02–1.28 Iowa, USA Heilman et al.27
samples
Tea 6 different kind 1.97–8.64 Taiwan Lung et al.28
of tea
Different types of 170–878 China Wong et al.29
tea products
Bottled 10 different types 0.06–1.05 Davangere city, Thippeswamy
drinking of bottled India et al.25
water drinking water
20 types of 0.21 Mexico City, Jimenez-Farfan
bottled waters Mexico et al.26

10 types of 0.08–0.30 Poland Opydo-


bottled waters szymaczek
recommended and Opydo30
for use by
infants and
young children
15 randomly 0.5–0.83 Riyadh, Saudi Aldrees and
selected Arabia. Al-Manea31
commercial
brands (12 local
brands + 3
imported
brands) of
bottled water
29 commercially 0.19–1.07 Algeria Bengharez
available et al.32
brands of
bottled waters
a For liquid items, concentrations are in mg/L.

that the fluoride concentrations in vegetables and cereals varied, respectively,


from 3.91 to 29.15 mg/kg and 0.45 to 5.98 mg/kg. Fluoride content in milk
samples was found to be in the range of 0.37–6.85 mg/L. Results suggested
that along with drinking water, food items, and milk also contribute to the TF
68 Fluoride in Drinking Water

intake.35 Raghavachari et al. (2008) conducted a study on the fluoride concen-


tration present in the food items available at Palamau district of Jharkhand,
where food is grown with irrigation water that has a fluoride concentration
0.10–12.30 mg/L.36 The fluoride content of cereals and pulses ranges from 1.5
to 1.78 mg/kg and from 1.46 to 2.28 mg/kg, respectively. The concentration
of fluoride in vegetables is very low (0.14–0.23 mg/kg) compared with that
of cereals and pulses. Fluoride intake through food items alone was found
to be between 0.97 and 1.23 mg/capita/day.36 Ramteke et al. (2007) performed
a study on the fluoride concentration in commonly used food items such as
rice, corn, wheat, and lentils (dal) in Dhar and Jhabua districts of Madhya
Pradesh.37 It was found that the fluoride concentrations in rice, corn, wheat,
and lentils (dal) were in the ranges of 0.51–5.52, 10.2–40, 0.75–9.02, and 1.1–13.42
mg/kg, respectively. The TF consumption was in the range of 10.7–21.21 mg/
capita/person. The maximum consumption of fluoride was 21.21 mg/day in
the age group of 31–45 years.37 Yadav et al. (2012) investigated fluoride concen-
tration in one crop (wheat) and two vegetables (potato and tomato) that were
collected from seven villages of Dausa district in Rajasthan, India.38 The fluo-
ride concentration in groundwater samples obtained from hand pumps and
open wells of these seven villages was found to vary from 5.1 to 14.7 mg/L.
The fluoride accumulation in crop (wheat) ranged from 3.24 to 14.3 mg/kg,
and the fluoride content in vegetables ranged from 1.10 to 4.60 mg/kg.38

6.4.3  Fluoride in Soil


Fluorine concentration in soil generally varies between 150 and 400 mg/L.
Soils that are derived from rocks with a high fluorine content or that are
affected by anthropogenic sources contain a very high concentration (1000
g/kg) of fluorine.39 Fluoride mobility in soil is highly dependent on the
sorption capacity of the soil and varies with pH, types of sorbents, and soil
salinity.40 This is because the fluoride concentration in the soil differs from
place to place. Jha (2012) studied the distribution of the fluoride in the soils
of ­Indo-Gangetic plains.41 The fluoride distribution in soil profiles and sur-
face soil (0–15 cm) samples were analyzed. Results demonstrated that the TF
in the profiles varied from 248 to 786 mg/kg. The CaCl2 extractable soluble
fluoride (FCa) was found to be in the range of 1.68–99.1 mgF/kg soil. While
in surface soils, the TF and FCa ranged from 118 to 436 mg/kg and from 1.01
to 5.05 mg/kg, respectively.41 Mishra et al. (2009) conducted an experiment
to measure the effect of fluoride emission from an aluminum smelter located
at Hirakud in western Orissa on the environment.42 The study was carried
out in a radius of 5 km from the plant. It was found that the concentration of
fluoride in soil varied from 88.30 to 191.20 mg/L.42 Jha et al. (2008) estimated
the fluoride concentration in the soil in the vicinity of a brick field in the
suburb of Lucknow, India.43 It was observed that the water-soluble fluoride
in the surface soil varied from 0.59 to 2.74 mg/L, and the CaCl2 extractable
fluoride ranged from 0.69 to 3.18 mg/L. The mean TF concentration in surface
Fluoride in the Environment and Its Toxicological Effects 69

soil varied from 322 to 456 mg/kg.43 Chaudhary et al. (2009) determined the
fluoride concentration in the soil samples from 60 village sites in the Indira
Gandhi, Bhakra, and Ganga canal catchment areas of north-west Rajasthan,
India.44 Results suggested that the mean water-extractable soil fluoride con-
centrations varied from 0.50 to 3.00 mg/L. It was concluded that the heavy use
of diammonium phosphate (DAP) fertilizer is a possible source of elevated
fluoride in the soil of that area.44 Anbuvel et al. (2014) studied the accumula-
tion of fluoride in the soil of eight villages near the bank of Thovalai Channel
in Kanyakumari district, Tamilnadu.45 The result showed that the fluoride
content of the soil varied from 1.0 to 3.2 mg/L. Since this area is free from
industrial activity, the heavy use of phosphate fertilizers over long periods
may be the reason for the increased concentrations of fluoride in the soil.45

6.4.4  Fluoride in Tobacco and Pan Masala


In India, a large number of people are addicted to tobacco and pan masala.
While calculating the TF intake, generally these items are not taken into
consideration. Generally, tobacco and pan masala are not swallowed, only
some fraction of them is ingested during their consumption, which ulti-
mately becomes available to the body for absorption along with sublingually
absorbed fluoride. Therefore, these items should be considered while calcu-
lating the TF intake by a human body. The exposure of fluoride through pan
masala and tobacco differs from person to person depending on their con-
sumption habits. Yadav et al. (2007) estimated the fluoride content in tobacco
and pan masala that are available in the local market of Delhi.46 They investi-
gated 8 samples of tobacco and 15 samples of pan masala (7 without tobacco
and 8 with tobacco). It was found that the fluoride content varied from 28.0 to
113.0 mg/kg for tobacco. In the case of pan masala without tobacco, the fluo-
ride content varied between 23.5 and 185.0 mg/kg, whereas for pan masala
with tobacco, it was between 16.5 and 306.5 mg/kg. Fluoride ingestion would
be different depending on the intake habits of various people.46 Kumar and
Yadav (2014) estimated the levels of fluoride content in pan masala, chew-
ing tobacco, and betel nuts in the rural and urban areas of Rampur district,
Uttar Pradesh, India.47 Water soluble fluoride content ranged from 23.50 to
42.50 mg/kg in pan masala. In case of chewing tobacco, water-soluble fluo-
ride ranged from 18.10 to 40.00 mg/kg. Water-soluble fluoride in supari (betel
nut) was found to be in the range of 8.8–76.50 mg/kg. The intake of two to four
sachets of pan masala and chewing tobacco by a person would yield between
0.18 and 1.20 mg and between 0.40 and 1.20 mg of fluoride per day, in addition
to the fluoride ingested from food and liquids.47

6.4.5  Fluoride from Occupational Exposure


Fluoride is a common pollutant in the industrial workspace. Workers in
many heavy industries such as aluminum, fertilizer, iron, oil refining,
70 Fluoride in Drinking Water

semiconductor, and steel may be routinely exposed to high levels of fluo-


ride. Fluoride is also used in the welding process; therefore, welders are
also commonly exposed to airborne fluorides. As per current U.S. regula-
tions, industries can have fluorides up to 2.5 mg/m3 in the workspace air,
which produces a fluoride intake of 16.8 mg/day for an 8-h working day.
A long-term air-monitoring study48 demonstrated that the fluorine concen-
tration in the air all around the workplace of a small-scale enamel enter-
prise was in the range of 0.1–3.7 mg/m3. In the United States, the fluoride
concentration in the aluminum production industries for exposed workers
was reported to be 1.25 mg/m3. The average concentration of particulate
fluorides was measured as 1.024 mg/m3, whereas gaseous fluoride (HF)
had a mean level49 of 0.22 mg/m3. In Sweden, the average TF exposure of
the workers in an aluminum plant was calculated as 0.91 mg/m3, of which
34% (approximately 0.31 mg/m3) was gaseous fluoride.50 Electronic indus-
try workers in Japan, where hydrogen fluoride is used for glass etching of
TV picture tubes and as a silicon cleaner for semiconductors, are exposed
to a daily average concentration of up to 5 mg/L of air hydrogen fluoride.51
It was reported52 that the workers in the pot room of an aluminum smelter
in British Columbia, Canada, were exposed to an average total airborne
fluoride concentration of approximately 0.48 mg/m3. In Netherlands, the
concentrations of fluoride range in the workroom air of welding machine
shops and shipyards were reported2 to be 30–16,500 μg/m3. From 1981 to
1983, the National Institute for Occupational Safety and Health (NIOSH)
conducted the National Occupational Exposure Survey (NOES), which
collected data on the effect of occupational exposure of chemical, physi-
cal, and biological agents on the workers. The NOES estimated that about
182,589 workers were affected by inhalation of hydrogen fluoride.53
As pointed out earlier, large numbers of workers are getting exposed to
toxic chemicals while working in industrial areas. Workers in many heavy
industries such as aluminum, fertilizer, iron, oil refining, semiconductor,
and steel may be routinely exposed to high levels of fluoride. Arshad and
Shanavas (2013) studied the effect of fluoride exposure on workers in the
fertilizer industry and the wood industry in Mangalore city, India.54 They
investigated the fluoride concentration in the serum and urine of the 34 work-
ers from the fertilizer industry and the 55 workers from the wood industry.
Urinary fluoride and serum fluoride levels are valid biomarkers for estimat-
ing the levels of occupational exposure to fluoride. The fluoride concentra-
tions in the serum and the urine of the workers employed in the fertilizer
industry were 0.077 ± 0.027 and 3.85 ± 1.66 mg/L, respectively. The work-
ers in the wood industry had fluoride concentrations of 0.037 ± 0.009 and
0.97±0.37 mg/L in their serum and urine, respectively. The study concluded
that the phosphate fertilizer workers in India are at a high risk of exposure
to excessive amounts of fluoride.54 Sharma et al. (1991) conducted a study to
investigate the effect of fluoride in a factory manufacturing inorganic fluoride
compounds.55 The preshift and postshift urinary fluoride levels of workers
Fluoride in the Environment and Its Toxicological Effects 71

were estimated. The preshift urinary fluoride levels ranged from 0.5 to 4.54
mg/L and the postshift levels ranged from 0.5 to 13.00 mg/L. The preshift and
postshift urinary fluoride concentration depends on the nature of work and
the category of workers in each department.55 Susheela et al. (2013) performed
a study on the effect of fluoride exposure on the workers in one of the largest
primary aluminum-producing industries located in the north-eastern part of
the state of Uttar Pradesh, India.56 It was observed that smelter workers had a
significantly higher fluoride concentration in their urine and serum than non-
smelter workers; in addition, the nail fluoride content was higher in smelter
workers than in nonsmelter workers. These studies clearly demonstrate that
industrial emission of fluoride is a major source of fluoride exposure.56

6.4.6  TF Exposure
As discussed in Sections 6.4.1 through 6.4.5, the TF exposure is influenced by
different sources and several factors. The factors that affect the fluoride con-
centration in foodstuffs include fluoride emission sources in the local area,
amount of fertilizers and pesticides applied in agricultural activities, and
use of fluoridated water in the preparation of food and so forth.57 The fluo-
ride concentration in the ambient air is influenced by several factors such as
nature and type of the industrial sources in the area, the distance from the
fluoride sources, the prevailing meteorological conditions and the geological
features of the area defined by its topography.58 The fluoride concentration
in water depends on many factors such as the local geological features and
proximity to emission sources, mineral constitution of the aquifers, seepage
from nearby saline formations, low recharge and dilution rates in the aqui-
fers, peculiarities of the local soil or rock formations, and so forth.59 However,
many scientific studies suggest that the total daily fluoride exposure in a
temperate climate when no fluoride is added to the drinking water is approx-
imately equal to 0.6 mg/adult/day; whereas it is around 2 mg/adult/day in a
fluoridated area.60 A range of estimated fluoride intakes as a consequence of
exposure to a number of different sources is given in Table 6.2.

6.5 Effects of Fluoride on Laboratory


Animals and In Vitro Systems
Considerable research was undertaken to determine the effect of fluoride in
laboratory animals. Effects on the skeleton, organs, and tissues have been
observed in a variety of studies conducted in rats and rabbits. Both short-
term and long-term effects of fluoride exposure for low doses and high doses
were investigated by a number of researchers. Some of the studies related to
fluoride exposure on laboratory animals are listed in Table 6.3.
TABLE 6.2
72

Fluoride intake from different sources


Estimated Fluoride
Sources of Fluoride Age Group of Intake, mg/day (mg/kg
Exposure Exposure Body Weight per day)a Test Insights Study Area References

Foodstuffs Children aged 1–4 (0.05) 457 whole-day meals Poland Jedra et al.61
years
Toothpaste and diet 1- to 3-year old (0.130) Fluoride intake from diet was measured by Brazil de Almedia
children the duplicate plate method, and fluoride et al.62
ingested from dentifrice was determined by
subtracting the amount of fluoride recovered
after brushing from the amount originally
placed onto the child’s toothbrush. Samples
were carried out by analyzing 33 children.
Food group (Grain 3–5 years 0.454 Estimated mean daily fluoride ingestion Connersville Jackson
products, Vegetables, community ( Indiana, et al.12
Fruits, Milk products, USA) having water
Meat and poultry , Nuts fluoride concentration
and seeds, Fats and oils, of 0.16±0.01 mg/L
Sugars and sweets, and 3–5 years 0.535 Estimated mean daily fluoride ingestion Richmond community Jackson
Beverages) ( Indiana, USA), an et al.12
optimally fluoridated
area having water
fluoride concentration
of 0.90±0.05 mg/L
Commercial food for Infants, 3–8 months (0.023–0.029) Estimated mean daily fluoride intake of Japan Tomori
Infant infants from food et al.63
Total diet samples, Slovenian military (0.010–0.035) Range of fluoride intake was calculated by Slovenia Ponikvar
including drinking water personnel assuming the mean weight of Slovenian et al.64
and beverages military personnel as 70 kg. The amount of Vaidya
fluoride was determined in 20 lyophilized et al.65
total diet samples obtained from the
Slovenian Military.

(Continued)
Fluoride in Drinking Water
TABLE 6.2 (Continued)
Fluoride intake from different sources
Estimated Fluoride
Sources of Fluoride Age Group of Intake, mg/day (mg/kg
Exposure Exposure Body Weight per day)a Test Insights Study Area References

Food Diet, Liquid 2–5 years (Moderate (0.0252) Estimated the mean concentration of fluoride Two areas of Japan Nohno
included water, milk, fluoride area) from the diet ingested by children of two age (Moderate fluoride et al.66
ready-made beverages 2–5 years (0.0126) groups susceptible to dental fluorosis area having mean
and beverages made at (Relatively low water fluoride
home (diluted powder fluoride area) concentration of 0.555
and concentrated fruit mg/L and relatively
6–8 years (Moderate (0.0254)
juice, etc., and a low fluoride area
fluoride area)
beverage of tea leaves having fluoride
and wheat ears with tap 6–8 years (0.0144) concentration in the
water). (Relatively low community water in
fluoride area) the range of
0.040–0.131 mg/L)
Solids food, water and 2–6 years (0.017) Mean fluoride intake from food items Nonfluoridated area of Levy et al.67
other beverages Brazil
All drinks (Water, Tea, 4-year-old children 0.413 Dietary fluoride intake in children residing in Iran Zohouri
Milk, Soft drink ) (First area) low, medium and high fluoride areas. (The and
All foods (Fruit, 4-year-old children 0.698 mean fluoride concentrations in drinking Rugg-
Vegetable, Soup and (Second area) water in the three areas were 0.3, 0.6 and 4.0 Gunn.68
gravy, Rice, Bread) mg F/L.)
4-year-old children 3.472
Fluoride in the Environment and Its Toxicological Effects

(Third area)
Food (Enjera, homemade Adults (Village A) 10.5 Daily dietary fluoride intake by adults from Ethiopia Dessalegne
bread, kale stew, potato three rural villages of the Ethiopian Rift et al.69
Adults (Village B) 16.6
stew, shiro stew, fish Valley (Village A uses water with 1.0 mg/L
stew), beverage fluoride, village B uses water with 3.0 mg/L
(including tea, coffee), fluoride, and village C uses water with 11.5
and water (used for Adults (Village C) 35.3 mg/L fluoride both for food preparation and
drinking and cooking). for drinking)

a Data in parentheses are the estimated intakes of fluoride, expressed as mg/kg body weight per day, when presented in the reference cited.
73
TABLE 6.3
74

Effect of Fluoride Exposure on Laboratory Animals


Laboratory Affected
Animal/In Organ/
Vitro System Fluoride Dose and Exposure Time System Toxicological Issues References
Female Wistar 226 mg/L fluoride ion in drinking water from Liver Ingestion of a high amount of fluoride Bouaziz et al.70
mice day 15 of pregnancy until day 14 after through drinking water may lead to
delivery impaired liver function.
Adult female 226 mg/L fluoride ion in drinking water from Brain Fluoride intoxication in the early stage of Bouaziz et al.71
mice of Swiss day 15 of pregnancy until day 14 after life interfered with brain physiology and
Albinos delivery induced neurotoxicity in mice.
strain
Wistar rats 5, 15 or 50 mg/L of fluoride in drinking water Liver and Exposure fluoride doses (15 and 50 mg/L) Iano et al.72
over 60 days kidney caused alterations in the antioxidant
system of liver and kidney of rats.
However, exposure to 5 mg/L of a
fluoride dose causes few changes in the
parameters.
Rats with 5, 15 or 50 mg/L of fluoride in drinking water Bones Fluoridated water of concentrations of 15 Turner et al.73
surgically for a period of 6 months and 50 mg/L caused osteomalacia and
induced renal reduced bone strength in rats, whereas
deficiency water with a low to moderate fluoride
concentration (0 and 5 mg/L) affected
neither bone mineralization nor strength
in rats.
Mature female 100 and 150 mg/L of fluoride in drinking Vertebral bone Fluoridated water causes an increase in Sogaard et al.74
rats water for 90 days bone mass while simultaneously causing
a decrease in bone strength/quality,
thereby suggesting the negative effect of
fluoride on bone quality.
(Continued)
Fluoride in Drinking Water
TABLE 6.3 (Continued)
Effect of Fluoride Exposure on Laboratory Animals
Laboratory Affected
Animal/In Organ/
Vitro System Fluoride Dose and Exposure Time System Toxicological Issues References
Rats 25 mg/L of fluoride/rat/day for 8 and Tissue Drinking of water containing high fluoride Shanthakumari
16 weeks may result in tissue damage and other et al.75
secondary complications.
Male 0.33 and 0.95 mg/L of fluorine ion in drinking Nervous Chronic administration of fluorine in the Varner et al.76
Long–Evans water for 52 weeks system form of AlF3 and NaF in the drinking
rats water of rats caused distinct
morphological alterations in the brain,
including effects on neurons and
cerebrovasculature, and may cause injury
to the brain.
Rabbits Drinking water with fluoride concentrations of Blood Excessive ingestion of fat and fluoride can Sun et al.77
50 and 100 mg/L for 5 months cause an oxidative stress reaction and
increase serum lipid levels either
separately or synergistically, which leads
to hypercholesterolemia in the
experimental rabbits.
Fluoride in the Environment and Its Toxicological Effects

Male Kunming Drinking water with fluoride concentrations of Nervous Chronic exposure of fluoride may impair Han et al.78
mice 11, 22, and 45 mg/L and food with 8.40 mg/kg system the long-term recognition memory of
for 180 days male mice, enhance the excitement of
mice, and upregulate VAMP-2 mRNA
expression, all of which are involved in
object recognition memory of the nervous
system.
Male albino 2 mg of sodium fluoride in 1 ml of distilled Reproductive Fluoride exposure may cause an adverse Ghosh et al.79
rats water per 100 g body weight per day for system effect on the reproductive system.
29 days
75
76 Fluoride in Drinking Water

6.6  Effect of Fluoride on Aquatic Organisms


Industrial applications such as phosphate processing, aluminum smelting,
steel manufacturing, and glasses frosting are capable of producing an efflu-
ent with a high concentration of fluoride. These high fluoride effluents may
find their way to the nearby water bodies, thereby causing damage to the
aquatic animals and plants. Several researchers gathered information on
the effect of fluoride on aquatic organisms. Mishra and Mohapatra (1998)
performed a study to measure the fluoride concentration in bones and to
monitor the haematological characteristics (RBC, haemoglobin, haemato-
crit, mean corpuscular haemoglobin, and mean corpuscular volume) in
amphibians, Bufo melanostictus, collected from fluoride-contaminated and
-­uncontaminated areas of the Hirakud Smelter Plant, Hirakud, India.80 The
average haemoglobin content, total RBC count and haematocrit in blood
samples were significantly reduced, whereas the mean corpuscular concen-
tration and volume were found to significantly increase with respect to the
toads at an uncontaminated site. The average fluoride concentration in bones
was 2736 mg/kg at the contaminated site, which was 11 times greater than
the fluoride concentration in bones of toads from the contaminated areas
(241 mg/kg).80 Hemens and Warwick (1972) performed experiments to deter-
mine the short-term and long-term effects of fluoride on fish and prawns
in an estuary in Zululand, South Africa.81 No toxic effects of fluoride were
noticed on the species of fish and prawns during their exposure to fluoride
up to 100 mg/L for 96 h (short-term exposure). The brown mussel Pernaperna
showed evidence of toxic effects after the fifth day of fluoride exposure at
a concentration of 7.2 mg/L. Long-term (72 days) exposure of fluoride at a
concentration of 52 mg/L was performed in recirculated outdoor labora-
tory estuary models without providing external food and with 20% salinity.
Results of long-term exposure showed physical deterioration and an increase
in mortality in the mullet Mugil cephalus and the crab Tylodiplax blephariskios.
The reproductive processes of the shrimp Palaemon pacificus were found to
be adversely affected due to the long-term exposure of fluoride.81 Johnstone
et al. (1982) conducted an experiment to determine the effect of exposure of
cryolite recovery sludge (CRS, an aluminum smelter waste dumped at sea)
filtrate, which also contains fluoride on salmon fish.82 The effects of exposure
of salmon on CRS filtrate (for up to 1 h) were monitored. It was observed
that Atlantic salmon exposed to aluminum smelter waste (including fluo-
ride) experienced an increase in oxygen consumption and ventilation rates
and a decrease in heart rate.82 Shi et al. (2009) carried out an experiment to
determine the accumulation of fluoride ion in juvenile sturgeon fish.83 In a
growth trial of 90 days, fishes were exposed to concentrations of 4, 10, 25 and
62.5 mg/L of F (added as NaF), along with a control group. Results indicated
that there was a significant inhibition of growth for groups exposed to high
fluoride concentrations (10, 25 and 62.5 mg/L) compared with the control
Fluoride in the Environment and Its Toxicological Effects 77

group. Shi et al. also observed that exposure of fluoride to a concentration of


25 mg/L or more may cause alterations in the fishes’ respiration and violent
erratic movements.83

6.7  Effect of Fluoride on Plants


Fluoride can enter into plants mainly through two pathways: aerial deposi-
tion of gaseous fluoride through stomatal diffusion and passive diffusion
through soil and water into the plant roots. Fluoride in the form of gas enters
into the stomata of the leaf by diffusion. Initially, it accumulates in the sto-
mata from where it moves toward the tip and the margin, causing injury to
the leaf. The injury symptoms are produced only when a critical level of fluo-
ride is attained.84 Fluoride as particulate falls on the leaf from the polluted
atmosphere and gets deposited on the surface of the leaf. Subsequently, these
deposited fluorides on the surface penetrate into the leaf and destroy it.85
The symptoms observed in plants due to the exposure to hydrogen fluoride
depend on a number of factors such as the concentration of HF, time of expo-
sure, type and age of plant, temperature, type of light and intensity, composi-
tion and rate of circulation of air. When exposed to high concentrations of HF
gas for sufficient time under controlled environmental conditions, the plants
that are sensitive to the HF may produce one or more of the following effects:
slight paling of normal green pigment at the tips or margins of the leaf that
may spread to other portions of the leaf; a pale green area at the margins
that may gradually turn into a light buff color and finally, a reddish brown.
All these stated effects of fluoride influence the photosynthesis and respira-
tion process of plants. The exact mechanism of injury to plants by fluorides
is unknown. It has been postulated that they interfere with the function-
ing of certain enzymes such as enolase.86 Fluoride can enter into the plant
system through the soil; it may also be deposited into the soil from several
anthropogenic sources from where it gets accumulated in the plant through
roots. The accumulation of fluoride from the soil is generally very small, and
there is a limited relationship between concentrations in plants and the total
content in soils.87 Plants absorb fluoride from the soil by their roots, and this
gets transported to the transpiratory organs of the plant (mainly the leaves)
via xylematic flow. This transported fluoride from the soil can accumulate in
the leaves where it can cause adverse effects such as tip burning and even
plant death by affecting the photosynthesis and transpiration process.88,89 A
lot of research was undertaken to determine the effect of fluoride on differ-
ent types of plants and tree species.
Kessabi et al. (1984) performed a study to determine the effect of fluorine
emission from the factories processing natural phosphate on plants and ani-
mals of south Safi zone (Morocco).90 These factories were 10 km south of Safi.
78 Fluoride in Drinking Water

Results revealed that the concentrations of fluoride were 4–10 times higher
in contaminated plants than in noncontaminated plants. In a certain study
area, the effect of fluoride pollution is so high and noticeable that grain crops
and vegetables are no longer grown there. In some areas, 30% of the grain
corps showed burn signs at the tips of leaves and their yield was reduced. In
trees, many fruits either fail or are necrosed. But even in the most contami-
nated zones, the grasses were unaffected.90 Zouari et al. (2014) carried out a
pot experiment to investigate the uptake, accumulation, and toxicity effects
of fluoride in olive trees that were grown in a soil spiked with inorganic
fluoride in the form of sodium fluoride.91 NaF was applied through irrigating
water in six different groups of olive with six different concentrations (0, 20,
40, 60, 80 and 100 mM NaF). Symptoms due to fluoride toxicity such as leaf
necrosis and leaf drop appeared only in highly spiked soils (80 and 100 mM
NaF). It was also reported that a significant reduction of biomass took place in
roots, shoots, and leaves of olive plants that are exposed to 60, 80 and 100 mM
NaF in comparison to the control plants. But the biomass reduction was not
significant for both 20 and 40 mM NaF treated soil.91 Singh and Verma (2013)
performed an experiment to examine the influence of fluoride-contaminated
irrigation water having a concentration of fluoride from 100 to 500 mg/L
on poplar seedlings (Populus deltoides L. clone-S7C15).92 Results indicated that
the exposure of the poplar seedlings to 100, 200 and 500 mg/L of fluoride in
the irrigation water for six weeks decreases the physiological characteristics
(growth, leaf expansion, photosynthetic CO2 assimilation, stomatal conduc-
tance, chlorophyll fluorescence yield, and plant biomass). Intervein chlorosis
and leaf-margin necrosis followed by leaf curl were observed even in the
younger leaves after the exposure of the seedling to irrigation water contain-
ing 500 mg/L of fluoride for six weeks. It was also observed that continuous
and prolonged exposure of fluoride-contaminated water results in falling of
the leaves.92

6.8  Effect of Fluoride on Animals


Excessive fluoride injections can affect animals. The impact of fluoride
depends on a number of factors such as dosage or amount of intake, rate
of intake, period of administration, and the presence of interfering sub-
stances. Fluoride effects on animals also depend on the physical param-
eters of the animal, such as age, state of health, and sensitivity. Young
animals are generally more susceptible to harmful effects of fluoride than
older ones. Healthy animals have more resistance to the harmful effects
of fluoride than sick or inadequately nourished animals.86 The intake of
fluoride in an excessive quantity than required by the animal body can
either induce acute toxicosis or cause chronic intoxication depending on
Fluoride in the Environment and Its Toxicological Effects 79

the concentration of fluoride and the time of exposure of the animals. The
symptom that arises in the animals when they inhale large quantities of
fluoride (several grams) in a very short interval of time (few minutes or
hours) is termed acute symptom. The symptoms that usually arise due to
acute toxicosis are an immediate decrease in appetite, high fluoride content
in animals’ blood and urine, rapid loss of weight, reduced milk produc-
tion, weakness, excessive salivation, perspiration, dyspnea, and weakened
pulse.93 Animals grazing on fluoride-affected plants may develop fluoro-
sis, characterized by damage to the musculoskeletal system, including dif-
ficulty in mastication, softening of the teeth, painful gait, and lameness.
Fluorosis occurs in animals grazing in fields near brickworks, aluminum
smelters, and phosphate fertilizer factories.

6.9  Guidelines Values and Standards


The WHO guidelines on fluoride in drinking water stipulate that less than
1 mg/L may give rise to dental fluorosis in some children, and much higher
concentrations (>1.5 mg/L) may eventually result in skeletal damage in both
children and adults. So, in order to prevent dental caries, a large number
of communities supply water with a fluoride concentration that is equal to
approximately 1.0 mg/L. The 1971 International Standards recommended
control limits for fluorides in drinking water for various ranges of the annual
average of maximum daily air temperatures. This limit ranges from 0.6 to
0.8 mg/L for a temperature range of 26.3–32.6°C and from 0.9 to 1.7 mg/L
for a temperature range of 10–12°C. In the first edition of the Guidelines for
Drinking Water Quality, published in 1984, a guideline value of 1.5 mg/L
was recommended by the WHO for fluoride, as mottling of teeth has been
reported very occasionally at higher levels. It was also noted that this guide-
line value is not fixed and local application must take consideration of local
climatic conditions, diet, and water consumption. The 1993 WHO Guidelines
concluded that there was no sufficient evidence to suggest that the guideline
value of 1.5 mg/L set in 1984 needed to be revised. In some countries, par-
ticularly parts of India, Africa, and China, drinking water can contain very
high concentrations of naturally occurring fluoride (in excess of the WHO
guideline value of 1.5 mg/L). So, it was felt that the guideline value may be
difficult to achieve in some circumstances with the treatment technology
available. In 1994, the WHO recommended that the optimal concentration of
fluoride should be in the range of 0.5–1.0 mg/L and should vary according
to climatic conditions, volume of water intake, and intake of fluoride from
other sources. Fluoride effects are best predicted by the dose (mg fluoride/
kg of body weight/day), the duration of exposure, and other factors such
as age. The U.S. National Academy of Sciences Institute of Medicine has
80 Fluoride in Drinking Water

recommended an adequate intake of fluoride from all sources as 0.05 mg F/


kg body weight/day. This amount exhibits a reduction in the occurrence of
dental caries in maximum cases without inducing unwanted side effects,
including moderate dental fluorosis.2,6,94

6.10 Summary
• Natural processes such as weathering and dissolution of minerals,
volcanic eruptions, and marine aerosols are responsible for fluoride
release into the environment. Volcanoes are the main natural persis-
tent source of fluorine.
• The transportation, distribution, transformation, and deposition
of airborne fluoride are governed by meteorological conditions,
particulate size, chemical reactivity, and emission strength of the
source.
• The daily fluoride intake of children, expressed in a milligram-per-
kilogram body weight basis, may exceed that of adults. It has been
found that ingestion of the toothpastes containing fluoride may con-
tribute to 0.50–0.75 mg fluoride per child per day.
• The concentration of the fluoride in the raw food items is the func-
tion of fluoride content in the irrigation water.
• Workers in many heavy industries such as aluminum, fertilizer,
iron, oil refining, semiconductor, and steel get routinely exposed to
high levels of fluoride.
• Urinary fluoride and serum fluoride levels are valid biomarkers for
estimating the levels of fluoride due to occupational exposure of
fluoride.
• Industrial emission of fluoride is a major source of fluoride exposure.
• Studies on toxicological impacts of fluoride on animals suggest
impaired liver function, changes in brain physiology, induced neu-
rotoxicity, alterations in the antioxidant system of liver and kidney,
reproductive systems, reduced level of bone quality, tissue damages,
morphological alterations and injury to brain, oxidative stress reac-
tion, and so forth.
• Fluoride exposures influence the photosynthesis and respiration
process of plants.
• Fluorosis occurs in animals grazing in fields near brickworks, alu-
minum smelters, and phosphate fertilizer factories.
Fluoride in the Environment and Its Toxicological Effects 81

References
1. Ayoob, S. and Gupta, A.K. (2006). Fluoride in drinking water: A review on the
status and stress effects. Crit. Rev. Environ. Sci. Technol., 36, 433–487.
2. WHO. (2002). Environmental Health Criteria 227, Fluorides. Geneva,
Switzerland: World Health Organization.
3. Barnard, W.R. and Nordstrom, D.K. (1982). Fluoride in precipitation-II.
Implications for the geochemical cycling of fluorine. Atmos. Environ., 16, 105–111.
4. ATSDR. (2003). Report on Toxicological Profile For Fluorides, Hydrogen Fluoride
and Fluorine. Atlanta, GA: U.S. Department of Health and Human Services,
Public Health Service Agency for Toxic Substances and Disease Registry.
5. Cape, J.N., Fowler, D. and Davison, A. (2003). Ecological effects of sulfur diox-
ide, fluorides, and minor air pollutants: Recent trends and research needs.
Environ. Int., 29, 201–211.
6. WHO. (2006). In: Farewell, J., Bailey, K., Chilton, J., Dahi, E., Fewtrell, L. and
Magara, Y. (Eds.) Fluoride in Drinking-Water. World Health Organization.
London-Seattle: IWA Publishing.
7. Slooff, W., Eevens, H.C., James, J.A. and Rose, J.R.M. (1989). Integrated Criteria
Document Fluoride. The Netherlands: National Institute of Public Health and
Environment Pollution, Bilthover (Report No. 758474010).
8. Murray J.J. [Ed.] (1986). Appropriate Use of Fluorides for Human Health.
Geneva, Switzerland: World Health Organization.
9. Bralić, M., Buljac, M., Prkić, A., Buzuk, M. and Brinić, S. (2015). Determination
fluoride in products for oral hygiene using flow-injection (FIA) and continuous
analysis (CA) with home-made FISE. Int. J. Electrochem. Sci., 10, 2253–2264.
10. Kumar, R. and Yadav, S.S. (2014). Fluoride content in pan masala, chewing
tobacco, betel nuts (supari), toothpaste and tooth-powder items used and con-
sumed in rural and urban parts of Rampur district, Uttar Pradesh, India. J. Sci.
Technol. Manag., 2, 38–47.
11. Yadav, A.K., Kaushik, C.P., Haritash, A.K., Singh, B., Raghuvanshi, S.P. and
Kansal, A. (2007). Determination of exposure and probable ingestion of fluoride
through tea, toothpaste, tobacco and pan masala. J. Hazard. Mater., 142, 77–80.
12. Jackson, R.D., Brizendine, E.J., Kelly, S.A., Hinesley, R., Stookey, G.K. and
Dunipace, A.J. (2002). The fluoride content of foods and beverages from negli-
gibly and optimally fluoridated communities. Community Dent. Oral Epidemiol.,
30, 382–91.
13. Liu, C., Wyborny, L.E. and Chan, J.T (1995). Fluoride content of dairy milk from
supermarket: A possible contributing factor to dental fluorosis. Int. Soc. Fluoride
Res., 28, 10–16.
14. Amalraj, A. and Pius, A. (2013). Health risk from fluoride exposure of a popu-
lation in selected areas of Tamil Nadu South India. Food Sci. Hum. Wellness, 2,
75–86.
15. Duff, E.J. (1981). Total and ionic fluoride in milk. Caries Res., 15, 406–408.
16. Dabeka, R.W. and McKenzie, A.D. (1987). Lead, cadmium, and fluoride levels
in market milk and infant formulas in Canada. J. Assoc. Off. Anal. Chem., 70,
754–757.
82 Fluoride in Drinking Water

17. Buzalaf, M.A., de Almeida, B.S., Cardoso, V.E., Olympio, K.P. and Furlani, T.
de A. (2004). Total and acid-soluble fluoride content of infant cereals, bever-
ages and biscuits from Brazil. Food Addit. Contam., 21, 210–5.
18. Fein, N.J. and Cerklewski, F.L. (2001). Fluoride content of foods made with
mechanically separated chicken. J. Agric. Food Chem., 49, 4284–4286.
19. Jedra, M., Urbanek-Karłowska, B., Fonberg-Broczek, M., Sawilska-Rautenstrauch,
D. and Badowski, P. (2001). Bioavailable fluoride in poultry deboned meat and
meat products. Rocz. Panstw. Zakl. Hig., 52, 225–230.
20. Heilman, J.R., Kiritsy, M.G., Levy, S.M. and Wefel, J.S. (1997). Fluoride concen-
trations of infant foods. J. Am. Dent. Assoc., 128, 857–863.
21. Camargo, J. (2003). Fluoride toxicity to aquatic organisms: A review. Chemosphere,
50, 251–264.
22. Rocha, R.A., Rojas, D., Ruiz, A., Devesa, V. and Ve, D. (2013). Quantification
of Fluoride in Food by Microwave Acid Digestion and Fluoride Ion-Selective
Electrode. J. Agric. Food Chem., 16, 10708−10713.
23. Thomas, A. and James, R. (2013). Accumulation of Fluoride in Etroplus Suratensis,
Oreochromis Mossambicus and Anabas Testudineus Caught from the Surface Fresh
Water Sources in Alappuzha town, Kerala, India. Int. J. Innov. Res. Sci. Eng.
Technol., 2, 2756–2761.
24. Sawilska-Rautenstrauch, D., Jedra, M., Fonberg-Broczek, M., Badowski, P. and
Urbanek-Karłowska, B. (1998). Fluorine in vegetables and potatoes from the
market in Warsaw. Rocz. Panstw. Zakl. Hig., 49, 341–346.
25. Thippeswamy, H.M., Kumar, N., Anand, S.R., Prashant, G.M. and Chandu, G.N.
(2010). Fluoride content in bottled drinking waters, carbonated soft drinks and
fruit juices in Davangere city, India. Indian J. Dent. Res., 21, 528–530.
26. Jimenez-Farfan, M.D., Hernandez-Guerrero, J.C., Loyola-Rodriguez, J.P. and
Ledesma-Montes, C. (2004). Fluoride content in bottled waters, juices and car-
bonated soft drinks in Mexico City, Mexico. Int. J. Paediatr. Dent.,14, 260–266.
27. Heilman, J.R., Kiritsy, M.C., Levy, S.M. and Wefel, J.S. (1999). Assessing fluoride
levels of carbonated soft drinks. J. Am. Dent. Assoc., 130, 1593–1599.
28. Lung, S.C., Cheng, H.W. and Fu, C.B. (2007). Potential exposure and risk of
fluoride intakes from tea drinks produced in Taiwan. J. Expo. Sci. Environ.
Epidemiol., 18, 158–166.
29. Wong, M., Fung, K. and Carr, H. (2003). Aluminium and fluoride contents of
tea, with emphasis on brick tea and their health implications. Toxicol. Lett., 137,
111–120.
30. Opydo-szymaczek, J. and Opydo, J. (2009). Fluoride content of bottled waters
recommended for infants and children in Poland. Res. Rep. Fluoride, 42, 233–236.
31. Aldrees, A.M. and Al-Manea, S.M. (2010). Fluoride content of bottled drinking
waters available in Riyadh, Saudi Arabia. Saudi Dent. J., 22, 189–93.
32. Bengharez, Z., Farch, S., Bendahmane, M., Merine, H. and Benyahia, M. (2012).
Evaluation of fluoride bottled water and its incidence in fluoride endemic and
non endemic areas. e-SPEN. J., 7, e41–e45.
33. Rao, K.V. and Mahajan, C.L. (1990). Fluoride content of some common South
Indian foods and their contribution to fluorosis. J. Sci. Food. Agric., 51, 215–219.
34. Gautam, R., Bhardwaj, N. and Saini, Y. (2010). Fluoride accumulation by
vegetables and crops grown in Nawa Tehsil of Nagaur district (Rajasthan,
India). J. Phytol., 2, 80–85.
Fluoride in the Environment and Its Toxicological Effects 83

35. Bhargava, D. and Bhardwaj, N. (2009). Study of fluoride contribution through


water and food to human population in fluorosis endemic villages of North-
Eastern Rajasthan. Afr. J. Basic Appl. Sci., 1, 55–58.
36. Raghavachari, S., Tripathi, R.C. and Bhupathi R.K. (2008). Endemic fluorosis in
five villages of the Palamau district, Jharkhand, India. Fluoride, 41, 206–211.
37. Ramteke, D. S., Onkar, R., Pakhide, D. and Sahasrabudhe, S. (2007). Assessment
of Fluoride in Groundwater, Food and Soil and its Association with Risk to
Health. Proceedings of the 10th International Conference on Environmental
Science and Technology, Kos Island, Greece.
38. Yadav, R.K., Sharma, S., Bansal, M., Singh, A., Panday, V. and Maheshwari, R.
(2012). Effects of fluoride accumulation on growth of vegetables and crops in
Dausa district, Rajasthan, India. Adv. Biores., 3, 14–16.
39. Kabata Pendias, A. and Pendias, H. (2001). Trace Elements in Soils and Plants, 3rd
edn, p. 413. Boca Raton, FL: CRC Press.
40. Cronin, S.J., Manohara, V., Hedley, M.J. and Loganathan, P. (2000). Fluoride: A
review of its fate, bioavailability, and risks of fluorosis in grazed-pasture sys-
tems in New Zealand. N. Z. J. Agri. Res., 43, 295–321.
41. Jha, S.K. (2012). Geochemical and spatial appraisal of fluoride in the soils of
indo-gangetic plains of India using multivariate analysis. Clean Soil Air Water,
40, 1392–1400.
42. Mishra, P.C., Meher, K., Bhosagar, D. and Pradhan, K. (2009). Fluoride distri-
bution in different environmental segments at Hirakud Orissa (India). Sci.
Technol., 3, 260–264.
43. Jha, S.K., Nayak, A. K., Sharma, Y.K., Mishra, V.K. and Sharma, D.K. (2008).
Fluoride accumulation in soil and vegetation in the vicinity of brick fields. Bull.
Environ. Contam. Toxicol., 80, 369–373.
44. Chaudhary, V., Sharma, M. and Yadav, B.S. (2009). Elevated fluoride in canal
catchment soils of Northwest Rajasthan, India. Res. Rep. Fluoride, 42, 46–49.
45. Anbuvel, D., Kumaresan, S. and Margret, R.J. (2014). Flouride analysis of soil
in cultivated areas of Thovalai channel in Kanyakumari District, Tamilnadu,
India: Correlation with physico-chemical parameters. Int. J. Basic Appl. Chem.
Sci., 4, 20–29.
46. Yadav, A.K., Kaushik, C.P., Haritash, A.K., Singh, B., Raghuvanshi, S.P. and
Kansal, A. (2007). Determination of exposure and probable ingestion of fluo-
ride through tea, toothpaste, tobacco and pan masala. J. Hazard. Mater., 142,
77–80.
47. Kumar, R. and Yadav, S.S. (2014). Fluoride content in pan masala, chewing
tobacco, betel nuts (supari), toothpaste and tooth-powder items used and con-
sumed in rural and urban parts of Rampur District, Uttar Pradesh, India. J. Sci.
Technol. Manag., 2, 38–47.
48. Viragh, E., Viragh, H., Laczka, J. and Coldea, V. (2006). Health effects of occupa-
tional exposure to fluorine and its compounds in a small-scale enterprise. Ind.
Health, 44, 64–68.
49. Taiwo, O.A., Sircar, K.D., Slade, M.D., Cantley, L.F., Vegso, S.J., Rabinowitz, P.M.,
Fiellin, M.G. and Cullen, M.R. (2006). Incidence of asthma among aluminium
workers. J. Occup. Environ. Med., 48, 275–282.
50. Ehrnebo, M. and Ekstrand, J. (1986). Occupational fluoride exposure and plasma
fluoride levels in man. Int. Arch. Occup. Environ. Health, 58, 179–190.
84 Fluoride in Drinking Water

51. Kono, K., Yoshida, Y. and Yamagata, H. (1987). Urinary fluoride monitoring of
industrial hydrofluoric acid exposure. Environ. Res., 42, 415–520.
52. Chan-Yeung, M., Wong, R., Earnson, D., Schulzer, M., Subbarao, K.,
Knickerbocker, J. and Grzybowski, S. (1983). Epidemiological health study of
workers in an aluminum smelter in Kitimat, B.C. II. Effects on musculoskeletal
and other systems. Arch. Environ. Health, 38, 34–40.
53. NIOSH. (1989). National occupational exposure survey (1980–1983). Cincinnati,
OH: National Institute for Occupational Safety and Health, Department of
Health and Human Services.
54. Arshad, M. and Shanavas, P. (2013). Comparison of serum and urinary fluoride
levels among fertilizer and wood industry workers in Mangalore city, India.
Res. Rep. Fluoride, 46, 80–82.
55. Sharma, Y.K., Kulkarni, P.K., Shah, A.R., Patel, M.D. and Kashyap, S.K. (1991).
Occupational exposure to inorganic fluorides. Indian J. Ind. Med., 37, 13–22.
56. Susheela, A.K., Mondal, N.K. and Singh, A. (2013). Exposure to fluoride in
smelter workers in a primary aluminum industry in India. Int. J. Occup. Environ.
Med., 4, 61–72.
57. Myers, H.M. (1978). Fluorides and dental fluorosis. Monogr. Oral. Sci., 7, 1–76.
58. Davis, W.L. (1972). Ambient air fluorides in Salt Lake County, Rocky Mountain.
Med. J., 69, 53–56.
59. Hudak, P.F. (1999). Fluoride levels in Texas groundwater. J. Environ. Sci. Heal. A,
34, 1659–1676.
60. WHO. (1984). Fluorine and Fluorides, Environmental Health Criteria 36.
Geneva, Switzerland: World Health Organization.
61. Jędra, M., Sawilska-rautenstrauch, D., Gawarska, H. and Starski, A. (2011).
Fluorine content in total diets samples of small children. Rocz. Panstw. Zakl.
Hig., 62, 275–281.
62. de Almeida, B.S., da Silva Cardoso, V.E. and Buzalaf, M.A. (2007). Fluoride
ingestion from toothpaste and diet in 1- to 3-year-old Brazilian children com-
munity. Dent. Oral Epidemiol., 35, 53–63.
63. Tomori, T., Koga, H., Maki, Y. and Takaesu, Y. (2004). Fluoride analysis of foods
for infants and estimation of daily fluoride intake. Bull. Tokyo Dent. Coll., 45,
19–32.
64. Ponikvar, M., Stibilj, V. and Žemva, B. (2007). Daily dietary intake of fluoride
by Slovenian military based on analysis of total fluorine in total diet samples
using fluoride ion selective electrode. Food Chem., 103, 369–374.
65. Vaidya, R., Bhalwar, R. and Bobdey, S. (2009). Anthropometric parameters of
armed forces personnel. Med. J. Armed Forces India, 65, 313–318.
66. Nohno, K., Sakuma, S., Koga, H., Nishimuta, M., Yagi, M. and Miyazaki, H.
(2006). Fluoride intake from food and liquid in Japanese children living in two
areas with different fluoride concentrations in the water supply. Caries Res., 40,
487–493.
67. Levy, F.M., Olympio, K.P., Philippi, S.T. and Buzalaf, M.A. (2013). Fluoride
intake from food items in 2- to 6-year-old Brazilian children living in a non-
fluoridated area using a semiquantitative food frequency questionnaire. Int. J.
Paediatr. Dent., 23, 444–451.
68. Zohouri, F.V. and Rugg-Gunn, A.J. (2000). Sources of dietary fluoride intake in
4-year-old children residing in low, medium and high fluoride areas in Iran.
Int. J. Food Sci. Nutr., 51, 317–26.
Fluoride in the Environment and Its Toxicological Effects 85

69. Dessalegne, M. and Zewge, F., (2013). Daily dietary fluoride intake in rural vil-
lages of the Ethiopian Rift Valley. Toxicol. Environ. Chem., 95, 1056–1068.
70. Bouaziz, H., Ketata, S., Jammoussi, K., Boudawara, T., Ayedi, F., Ellouze, F. and
Zeghal, N. (2006). Effects of sodium fluoride on hepatic toxicity in adult mice
and their suckling pups. Pestic. Biochem. Physiol., 86, 124–130.
71. Bouaziz, H., Amara, I. Ben, Essefi, M., Croute, F. and Zeghal, N. (2010). Fluoride-
induced brain damages in suckling mice. Pestic. Biochem. Physiol., 96, 24–29.
72. Iano, F.G., Ferreira, M.C., Quaggio, G.B., Fernandes, M.S., Oliveira, R.C., Ximenes,
V.F. and Buzalaf, M.A.R. (2014). Effects of chronic fluoride intake on the antioxi-
dant systems of the liver and kidney in rats. J. Fluor. Chem., 168, 212–217.
73. Turner, C.H., Owan, I., Brizendine, E.J., Zhang, W., Wilson, M.E. and Dunipace,
A. J. (1996). High fluoride intakes cause osteomalacia and diminished bone
strength in rats with renal deficiency. Bone, 19, 595–601.
74. Sogaard, C.H., Mosekilde, L., Schwartz, W., Leidig, G., Minne, H.W. and Ziegler,
R. (1995). Effects of fluoride on rat vertebral body biomechanical competence
and bone mass. Bone, 16, 163–169.
75. Shanthakumari, D., Srinivasalu, S. and Subramanian, S. (2004). Effect of fluo-
ride intoxication on lipidperoxidation and antioxidant status in experimental
rats. Toxicology, 204, 219–228.
76. Varner, J., Jensen, K.F., Horvath, W. and Isaacson, R.L. (1998). Chronic
administration of aluminum-fluoride or sodium-fluoride to rats in drinking
water: Alterations in neuronal and cerebrovascular integrity. Brain Res., 784,
284–298.
77. Sun, L., Gao, Y., Zhang, W., Liu, H. and Sun, D. (2014). Effect of high fluoride and
high fat on serum lipid levels and oxidative stress in rabbits. Environ. Toxicol.
Pharmacol., 38, 1000–1006.
78. Han, H., Du, W., Zhou, B., Zhang, W., Xu, G., Niu, R. and Sun, Z. (2014). Effects
of chronic fluoride exposure on object recognition memory and mRNA expres-
sion of SNARE complex in hippocampus of male mice. Biol. Trace Elem. Res., 158,
58–64.
79. Ghosh, D., Das, S., Maiti, R., Jana, D. and Das, U.B. (2002). Testicular toxicity in
sodium fluoride treated rats: Association with oxidative stress. Reprod. Toxicol.,
16, 385–390.
80. Mishra, P.C. and Mohapatra, K. (1998). Haematological characteristics and
bone fluoride content in Bufo melanostictus from an aluminium industrial site.
Environ. Pollut., 99, 421–423.
81. Hemens, J. and Warwick, R.J. (1972). The effects of fluoride on estuarine organ-
isms. War. Res., 6, 1301–1308.
82. Johnstone, A.D.F. and Hawkins, A.D. (1982). The effects of an industrial waste
(cryolite recovery sludge) upon the Atlantic salmon, Salmo salar (L). Water Res.,
16, 1529–1535.
83. Shi, X., Zhuang, P., Zhang, L., Feng, G., Chen, L., Liu, J., Qu, L. and Wang, R.
(2009). The bioaccumulation of fluoride ion (F–) in Siberian sturgeon (Acipenser
baerii) under laboratory conditions. Chemosphere, 75, 376–380.
84. Sharma, M.R. and Gupta, V. (2014). Fluoride and its ecological effects in water:
A review. Global J. Res. Anal., 3, 2277–8160.
85. Bellomo, S., Aiuppa, A., D’Alessandro, W. and Parello, F. (2007). Environmental
impact of magmatic fluorine emission in the Mt. Etna area. J. Volcanol. Geotherm.
Res., 165, 87–101.
86 Fluoride in Drinking Water

86. Greenwood, D.A. (1956). Some Effects of Inorganic Fluoride on Plants, Animals,
and Man, USU Faculty Honor Lectures. Paper 41. http://digitalcommons.usu
.edu/honor_lectures/41.
87. Weinstein, L.H. (1977). Fluoride and plant life. J. Occup. Med., 19, 49–78.
88. Klumpp, A., Klumpp, G., Domingos, M. and Silva, M.D. (1996). Fluoride impact
on native tree species of the Atlantic Forest near Cubatao, Brazil. Water Air Soil
Poll., 78, 57–71.
89. Davison, A. and Weinstein, L.W. (1998). The effects of fluorides on plants. Earth
Island J., 13, 257–264.
90. Kessabi, M., Assimi, B. and Braun, J.P. (1984). The effects of fluoride on animals
and plants in the south Safi zone. Sci. Total Environ., 38, 63–68.
91. Zouari, M., Ben Ahmed, C., Fourati, R., Delmail, D., Ben Rouina, B., Labrousse,
P. and Ben Abdallah, F. (2014). Soil fluoride spiking effects on olive trees (Olea
europaea L. cv. Chemlali). Ecotoxicol. Environ. Saf., 108, 78–83.
92. Singh, M. and Verma, K.K. (2013). Influence of fluoride-contaminated irrigation
water on physiological responses of poplar seedlings (Populus deltoides L. clone-
S7C15). Fluoride, 46, 83–89.
93. Hobbs, C.S., Moorman, R.P, Griffith, J.M., West, J.L., Merriman, G.M., Hansard,
S.L., Chamberlain, C.C., MacIntire, W.M., Hardin, L.J. and Jones, L.S. (1954).
Fluorosis in cattle and sheep. Tenn. Agr. Exp. Sta. Bull., 235.
94. WHO. (2003). Fluoride in Drinking-Water. Background document for prepa-
ration of WHO Guidelines for drinking-water quality. Geneva, Switzerland:
World Health Organization.
7
Defluoridation Techniques: An Overview

7.1 Introduction
The defluoridation techniques generally practiced include (1) coagula-
tion, (2) adsorption (including ion exchange), (3) electrochemical methods,
and (4) membrane processes. Coagulation processes mainly use chemical
reagents such as lime, calcium, or magnesium salts, poly aluminum chloride,
and alum to make precipitation (or co precipitation) with fluoride, necessi-
tating its removal. Adsorption is a popular technique practiced in fluoride
endemic areas of the developing world. In this method, the adsorbent is used
in fixed columns in packed beds and fluoride-laced water is cycled through
it. The pollutant from a relatively bulk liquid volume gets concentrated and
confined onto a small adsorbent mass, which can invariably be regener-
ated, reused, or safely disposed under control.1,2 Electrochemical techniques
mainly include electrocoagulation and other electrosorptive processes.3,4
Electrosorptive techniques basically involve activation of an adsorbent bed
and enhanced removal by application on an electric field. Electrocoagulation
involves the use of aluminum electrodes that release Al3+ ions (by an anodic
reaction) that react with fluoride ions near the anode. In this process, the
removal of fluoride by precipitation is expected to occur at the electrode–
electrolyte interface. Membrane techniques generally include reverse osmo-
sis (RO), nanofiltration (NF), ultrafiltration (UF), electrodialysis, and Donnan
dialysis. A combination of two or more of these membrane techniques for
enhanced removal of fluoride was also reported.1,5,6

7.2 Coagulation
In general, the removal mechanisms that are operative in coagulation include:
(1) charge neutralization of negatively charged colloids by cationic hydroly-
sis products and (2) incorporation of impurities onto an amorphous precipi-
tate of metal hydroxide. The relative importance of these two mechanisms

87
88 Fluoride in Drinking Water

depends on many factors, which are mainly pH and coagulant dosage.7


Defluoridation processes by coagulation include (1) precipitation of fluoride
by a suitable reagent through chemical reactions; (2) co precipitation of fluo-
ride, which involves its simultaneous precipitation with a macro-component
from the same solution through the formation of mixed crystals, by multiple
mechanisms such as adsorption, occlusion, or mechanical entrapment.1

7.2.1 Lime
Precipitation of fluoride in the form of insoluble calcium fluoride (CaF2) is one
of the most commonly adopted precipitation techniques used in defluorida-
tion. For this purpose, either lime [Ca(OH)2] or salts of calcium such as CaSO4
or CaCl2 may be used. The precipitation reaction involves the following:

Ca(OH)2 + 2 F − → CaF2 ↓+ 2OH − (7.1)

If Ca(OH)2 is used as a source of lime, the pH will increase with calcium


dosage as per Equation 7.1, which displays the major limitation of the process.
Hence, proper dosage of lime should be ensured to keep pH within permis-
sible limits prescribed for drinking water.8 It has been reported that liming
usually leaves higher residual fluoride concentrations of 10– 20 mg/L, which
makes drinking water unpalatable.1,9 So, additional defluoridation processes
are to be employed for removing the excess fluoride present. This further
demands additional necessities to be provided, such as usage of more chemi-
cals, reagents, and employing processes that enhance the treatment expenses.
However, these techniques may render large volumes of additional sludge.10,11
Another noticeable limitation with lime precipitation is the poor settling char-
acteristics of the precipitate. As stated earlier, “water treated with lime fre-
quently has much higher concentrations of residual fluoride because of slow
nucleation during precipitation, leading to a high ionic strength” and hard-
ness.1,10,12 The supplementary processes to be employed in removing excess
chemicals lead to the inappropriateness of this method in ensuring the quality
and originality of natural waters.11 Further, this may add to the total process
cost. These constraints that are related to precipitation, limit its usage as a sus-
tainable option for defluoridation of drinking water.

7.2.2  Magnesium Oxide

Magnesium oxide has also been used for defluoridation of drinking water.13
On hydration, magnesium oxide gets converted to magnesium hydroxide,
which combines with fluoride ions and forms practically insoluble magne-
sium fluoride as follows:

MgO + H 2O → Mg(OH)2 (7.2)


Defluoridation Techniques 89


Mg(OH)2 + 2 F − → MgF2 ↓+ 2OH – (7.3)

The use of magnesium oxide is prevalent in fluoride endemic areas14 in


domestic defluoridation units (DDUs). However, as shown in Equation 7.3,
there will be an increase in the pH of treated water to the range of 10 to 11.
However, the addition of small amounts (0.15–0.20 g/L) of sodium bisulfate
can decrease pH within the desirable limits (6.5–8.5). The DDU consists of
two units of 20 L capacity each. The mixing of magnesium oxide is carried
out in the upper unit through a manually operated, geared mechanical stir-
ring device. The lower unit serves as a collection unit. The mixing of the
coagulant is completed within 5 min, and the mixture is clarified for 16 h.
The bottom portion of the container receives the flocs or sludge that settles in
the system. Clear water is poured into the lower collection unit through an
elastic connecting pipe. This pipe is fitted with a fine filter for arresting the
flocs and tiny sludge elements. In the lower unit, sodium bisulfate is added
so as to dissolve it in drinking water before it is supplied for drinking.1

7.2.3  Calcium and Phosphate Compounds


Calcium and phosphate compounds such as calcium chloride (CaCl2 ⋅ 2H2O)
and monosodium phosphate (NaH2PO4 ⋅ H2O) can be used for defluorida-
tion in which fluoride gets precipitated as calcium fluoride or fluorapatite,
as follows:1

CaCl 2 ⋅ 2H 2O = Ca 2+ + 2Cl – + 2H 2O (7.4)

NaH 2PO 4 ⋅ H 2O = PO 4 3– + Na+ + 2H+ + H 2O (7.5)

Ca 2+ + 2F – = CaF2 (7.6)

10Ca 2+ + 6PO 4 3– + 2F – = Ca10 (PO 4 )6 F2 (7.7)

This process can be catalyzed in a contact bed consisting of a saturated


bone charcoal medium that may act as a filter for the precipitate.15 In addi-
tion, gravel or coarse-grained bone charcoal can be used as a supporting
medium for this adsorbent bed in a column. Fluoride-rich water mixed
with calcium chloride and monosodium phosphate is fed to the column
for a contact period of 20–30 min. The treated water flows continuously by
gravity through the bed to the clean water collecting tank. This process
has been practiced in many African countries such as Tanzania and Kenya
for treating groundwater with fluoride concentrations of about 10 mg/L.1,16
90 Fluoride in Drinking Water

7.3  Co Precipitation of Fluoride


7.3.1 Alum
Alum [Al2(SO4)3 ⋅ 18H2O] is a popular coagulant used for the flocculation
of colloidal impurities in water. When aluminum (Al) salts are dissolved
in water, the metal ion Al3+ gets hydrated and forms an aquometalion
Al(H 2 O)6 3+, which upon further hydrolysis forms a series of mononu-
clear, dinuclear, and possibly polynuclear hydroxo complexes, namely,
Al 13 (OH)34 5+, Al 7 (OH)17 4+, Al 8 (OH)20 4+ and Al 6 (OH)15 3+ ultimately precipi-
tating onto the metal hydroxide floc of Al(OH)3. The hydrolytic reactions
can be expressed as follows:1,17–19

Al 3+ + H 2O ↔ Al(OH)2+ + H+ (7.8)

Al 3+ + 2H 2O ↔ Al(OH)2+ + 2H+ (7.9)

7Al 3+ +17H 2O ↔ Al 7 (OH)17 4+ +17H+ (7.10)

Al(H 2O)6 3+ + H 2O ↔ Al(H 2O)5 (OH)2 + H 3O+ (7.11)

Al 3+ + 3H 2O ↔ Al(OH)3 (s)+ 3H+ (7.12)

Al 3+ + 4H 2O ↔ Al(OH)4 – + 4H+ (7.13)

Except the simple aquometal ions, the hydroxometal complexes thus formed
get readily adsorbed at interfaces19 that are responsible for the destabilization
of colloids in water that are treated with aluminum salts through charge neu-
tralization. Also, “polymers of high molecular weight can adsorb simultane-
ously on two or more particles and bind them” together through “polymer
bridging.”1,7 The acidity of treated water may increase due to the release of H+,
as shown in the earlier equations. So, the pH of the system after alum treat-
ment may be influenced by the dosage of alum and the initial alkalinity of the
water treated. In this process, it is plausible that fluoride ions are removed by
forming a part of the gelatinous Al(OH)3 flocs, which, subsequently, gets pre-
cipitated. It is suggested that the mechanism of fluoride removal due to alum
addition may be due to (1) coprecipitation of fluoride and hydroxide (OH)
ions with aluminum (Al) ions, forming a precipitate or floc with the chemi-
cal formula AlnFm(OH)3n−m (Equation 7.14); (2) and/or by adsorption or ligand
Defluoridation Techniques 91

exchange (Equation 7.15). The fluoride removal mechanism by adsorption or


complexation with Al(OH)3(s) can be expressed as follows:10,20

nAl(aq )3+ + (3n-m) OH(aq ) – + mF(aq ) – → Al n Fm (OH)3n-m (s) (7.14)

Al n (OH)3n (s) + m F(aq ) – → Al n Fm (OH)3n-m (s) + m OH(aq )− (7.15)

Though the species Al(OH)4 – aids fluoride removal through a ligand-


exchange mechanism (Equation 7.16), it may release high aluminum residu-
als into the treated water.17

Al(OH)4 – + F – → Al(OH)3 F + OH – (7.16)

The “efficiency of removal of fluoride by a fixed alum dose depends


on pH, alkalinity, the coexisting anions, and other characteristics of the
solution.”1,21 Literature also suggests some major limitations of alum
treatment22,23 due to the increased release of sulfate and aluminum con-
centrations into treated water. Coagulation with alum results in aluminum
residuals of 0.37 mg/L at a pH of 9.8. However, the amount of aluminum
residuals released was only 0.07 mg/L at a lower pH of 7.61. This clearly
demonstrates that the lowering of pH dramatically reduces the residual
aluminum.17 So, the most appropriate pH for defluoridation by coagula-
tion is in the range of 5.5–6.5.24 Incidentally, for practical applications, post-
treatment pH control may be necessary for providing stable water, which
may invite increased initial investments, capital, operational, and mainte-
nance costs, and enhanced hardness of treated water. So, alum treatment is
confined to high dosage requirement, issues of sludge disposal, high pH of
the treated water, and residual alumina in treated water.1,25

7.3.2  Alum and Lime (Nalgonda Technique)


This technique derived its name from Nalgonda, which is a place in India
where the first community defluoridation plant was constructed and is a
district affected by severe fluorosis. Defined quantities of alum, lime, and
bleaching powder are added to raw water; this is followed by rapid mix-
ing, flocculation, sedimentation, filtration, and disinfection.26 The added
sodium aluminate or lime hastens settlement of precipitate, and bleach-
ing powder ensures disinfection. The dose of lime is only 1/20th to 1/25th
that of filter alum. Bleaching powder is added to raw water at the rate of 3
mg/L to ensure disinfection. The process involves coagulation with alum
in an alkaline aqueous environment; this is followed by adsorption and
charge neutralization. “Fluoride may be adsorbed onto the sticky gelatinous
Al(OH)3 flocs during sweep coagulation” and may get co precipitated. The
use of aluminum sulfate and lime for defluoridation has its genesis from
92 Fluoride in Drinking Water

the United States in the 1930s and got popularized in India as the Nalgonda
technique after the 1970s. Over the years, in many developing countries such
as Tanzania, Senegal, Kenya, and India, this technique has been successfully
implemented for community applications and at individual household lev-
els.5,15,16 “Fill and draw type” defluoridation units are basically designed for
community applications for serving around 200 people. The entire opera-
tions of such a unit can be completed within 2–3 h with multiple batch per-
formance in a single day. Of late, an advanced electrically operated model
of this unit has also been developed.27 Community installations using “fill
and draw type” units having capacities up to 20–40 million gallons have
been used in the fluoride endemic areas in India (Figure 7.1). In addition to
defluoridation, the mechanisms involved, namely, coagulation, flocculation,
clarification, and disinfection, also aid in the simultaneous removal of color,
odor, turbidity, bacteria, and organic contaminants from water. The low cost
and ease of handling made this process more preferable.28 However, it is also
reported that this technology had limited field applications both as hand
pump–based units and as smaller domestic units, mainly due to the need for
constant attention. Limitations of this technique also include medium effi-
ciency, high aluminum sulfate dosage, controlling of varying alum and lime
dosages for different sources of raw water with different alkalinity and fluo-
ride concentrations, residual sulfate, salinity, hardness of the treated water,
higher pHs, and high residual aluminum concentrations.1,5,16,29

M.S. channel
450

W.L
ϕ2500
Detachable
ladder
1975

50ϕ clear
50ϕ raw water
water pipe
900

160ϕ drain
pipe

Pump Defluoridation unit Pump Overhead tank Stand post

Bore well Storage tank


Dimensions (mm)

FIGURE 7.1
The fill-and-draw type of defluoridation system for rural water supply for a population of
more than 1500 at 40 lpcd. (From CPHEEO, Central Public Health and Environmental Engineering
Organization, Manual on Water Supply and Treatment, 3rd edn., pp. 289–297, The Controller of
Publications, New Delhi, India, 1991.)
Defluoridation Techniques 93

7.4 Adsorption
Adsorption basically denotes an interface accumulation of substances at a sur-
face or interface. The material adsorbed is termed adsorbate, and the adsorbing
phase is called adsorbent. The mechanisms of adsorption may be “physisorp-
tion,” or “chemisorption,” or both. A physisorbed species is not attached to a
specific site; rather, it is free of any translational motion within the interface.
Physisorption may be significant at low temperatures and it develops a low
energy of adsorption, indicating that the adsorbate is loosely held with the
adsorbent. If the adsorbate develops a chemical interaction with the adsor-
bent, the process may be referred to as chemisorption. The adsorbed molecules
are attached on the surface as they form strong localized bonds at the active
centers of the adsorbent. The “ion exchange” may be treated as an “exchange
adsorption,” in which “ions of one substance concentrate at a surface as a result
of electrostatic attraction to charged sites at the surface.”1,19 Of late, adsorption
or ion exchange is one of the most frequently used methods for defluorida-
tion. Water laced with fluoride is passed through a column packed with an
adsorbent and on saturation, the adsorbent bed is backwashed for reuse. The
adsorption capacity, cost of the adsorbent, ease in operation, and potential for
reuse and regeneration are some of the factors that define the selection of an
adsorbent. It would be interesting to understand the mechanism of fluoride
removal in the applications of some of the most frequently used adsorbents.

7.4.1  Bone and Bone Charcoal


The use of bone in fluoride scavenging was demonstrated from 1930 onward.
The caustic-and-acid–treated bone material was demonstrated successful in
reducing fluoride concentration from 3.5 mg/L to even less than 0.2 mg/L.
The removal mechanism includes an ion exchange of carbonate radical of
the apatite [Ca9(PO4)6 ⋅ CaCO3] in the bone with fluoride (forming insoluble
fluorapatite) as follows:

Ca 9 (PO 4 )6 ⋅ CaCO 3 + 2F – → Ca 9 (PO 4 )6 ⋅ CaF2  + CO 3 2– (7.17)

Although the high cost of bone was an inhibiting factor in the initial period, it
was reported that bone char (obtained by carbonizing bone at 1100°C–1600°C)
has superior defluoridation potential than the original unprocessed bone. As
a result, thereafter, bone char was used in defluoridation operations. Bone
char is obtained by charring animal bones for removing all organics. The
resultant product essentially consists of tricalcium phosphate and carbon.
The adsorption mechanism of bone char is also an ion exchange in which
phosphate in bone char is exchanged with a fluoride ion.30 The removal of
fluoride by hydroxyapatite can be represented as follows:

Ca10 (PO 4 )6 (OH)2 + n F – = Ca10 (PO 4 )6 (OH)2-n Fn + n OH – (7.18)


94 Fluoride in Drinking Water

During the 1940s–1960s, bone charcoal was one of the oldest water defluo-
ridation agents in the United States because of its wide commercial avail-
ability; it was also successfully used in many full-scale installations. The
exchange capacity of the U.S. Public Health Service (USPHS) plant in Britton,
South Dakota, was 102 g fluoride/m3 bone char in treating waters with ini-
tial fluoride concentrations of 5 mg/L.1,31 The simplicity, local availability,
and easy processing facilities made this method more popular for domes-
tic- and community-level applications in many developing countries such as
Tanzania. The combined use of bone char with the Nalgonda technique was
also reported in literature.5 However, a major limitation is its poor regenera-
tion capacity, as on many occasions the used bone char gets discarded rather
than regenerated. Although the bone char method was successful in remov-
ing fluoride to very low levels, it was found to be more expensive and less
stable in a continuous flow system than activated alumina (AA). The quality
of bone charcoal defines its practical applicability. The bone charring process
is very important in this direction, as any failure in this process may result
in poor-quality bone charcoal. This may produce drinking waters with bad
taste that may smell similar to rotten meat and, ultimately, turn out to be
unacceptable to society. Once consumers get exposed to such smell or taste,
they may reject the process forever. Also, practical applications of bone char
demonstrate that its fluoridation capacity (represented in terms of the quan-
tity of fluoride removed by one gram of bone charcoal at the saturation level)
is less than that was being claimed (6 mg/g) in laboratory studies. In actual
applications dealing with water treatment, its defluoridation capacity may
range from one-third to two-third of this claimed capacity.16 Of late, bone
charcoal defluoridation waterworks are found to be replaced by ion-exchange
resins and AA. The domestic-level applications of bone charcoal defluorida-
tion were reported from Tanzania, Thailand, and Africa. In countries such as
India, its use was constrained by religious beliefs of many communities. The
cost of bone charcoal may vary depending on the method of manufacture.
The cost of bone charcoal from the United Kingdom, China, and the United
Republic of Tanzania in 1995 was reported to be US$ 2280, US$ 333, and US$
167 per ton, respectively. However, in many fluorotic areas, it was prepared
at a much lower cost; for instance, in Arusha region of Tanzania, by using
about 120 kg of charcoal per ton of bone.16

7.4.2  Clays and Soils


The potential of clay and soil-based adsorbents for defluoridation has been
under investigation in many fluoride endemic regions of the world. Major
studies include Illinois soils in the United States, Ando soils in Kenya, sodic
soils in India, fired clay chips in Sri Lanka and Africa, and fly ash, Alberta
soil, clay pottery, activated clay, kaolinite and bentonite, and illite–goethite
soils in China. The Ando soils of Kenya belong to porous soils derived from
Defluoridation Techniques 95

volcanic ash, which, in part, has weathered to yield “active aluminum”


in various forms. This type of soil was reported to have a high adsorp-
tion capacity to the tune of 5.51 mg/g.32 Investigations on improving the
adsorption capacity through surface coating of clays and soils were also
reported.33,34 The coating of clays and soils with alumina and iron hydrox-
ides were found to improve their adsorption capacity. Clay, a significant
form of soil, has been traditionally used in many developing countries such
as India, for making potteries to store water. This clay has been modified
with the additions of Al2O3, FeCl3 and CaCO3, leading to an observed reduc-
tion in water fluoride content.33 The defluoridation studies on kaolin clay
reveal that “solution pH, clay surface area, structure, aluminum content,
and the presence of certain exchangeable cations capable of forming fluo-
ride precipitates” are significant in defluoridation.1,35 Though ion-exchange
reactions are believed to be the predominant form of fluoride sorption, they
may also be immobilized through the formation of complexes or precipi-
tates with exchangeable cations such as magnesium, iron, and calcium.1,35
There may be an electrostatic attraction to the clay surface through which
fluoride (as F−) may be retained in the electric double layer.1,35 It was clearly
demonstrated that “disruption of the kaolin crystal structure occurred due
to fluoride uptake and F/OH exchange occurred primarily with Al(OH)3
rather than with –OH from the crystal lattice of clay minerals.” The fluoride
removal mechanism can be represented as follows:1,35

n(kaolin-OH)(s) + nF n– (aq ) ↔ n(kaolin-F)(s) + nOH(aq )n– (7.19)

The initially sorbed fluoride pushes the layers of metal oxides or hydrated
layers on the clay for providing easier access to sorption sites. The hydro-
gen bonding between kaolin sheets may be disrupted by fluoride by either
“attaching themselves to the slightly positive gibbsite surfaces or replacing
the hydroxyl groups on these surfaces.”1,36 Due to electronegativity, fluo-
ride ions acquire a highly negatively charge, which will force the silica and
gibbsite sheets to further move away from each other. This increases the
accessibility and exposure of hydroxyl groups in the gibbsite for fluoride
removal. This may enhance an exponential increase in fluoride sorption.
The kaolin sheets will be separated to their maximum so that any further
separation would not enhance fluoride access to sorption sites. A two-step
ligand exchange model was suggested for fluoride sorption onto goethite.37
Sorption is found to be maximum at around pH 3 in goethite, as fluoride
ions hydrolyze and form the neutral species HF near this pH, thus becoming
unavailable for sorption.1,38
The fluoride-scavenging potential of calcite, quartz, and fluorspar was also
investigated. Fluoride uptake was suggested to be a surface adsorption pro-
cess. The mechanism of fluoride by calcite was considered to work in two
phases. In the first phase, calcium ions may get gradually released into the
96 Fluoride in Drinking Water

solution at a certain pH range. Further, such dissolved calcium ions interre-


late with fluoride ions and form calcium fluoride (CaF2) precipitates.8,39 The
fluoride replaced CO 3 2– from calcite as follows:

2 F – + CaCO 3 (s) = CaF2(s) + CO 3 2− (7.20)

In the initial phase, quartz displayed poor adsorption capacity. The acti-
vation of quartz by the ions of iron (Fe3+) drastically increases the fluoride
adsorption capacity. It was observed that the siloxane groups of quartz (SiO2)
interrelate with water, forming –SiOH group formulations. The adsorption
of fluoride onto quartz is believed to be due to the replacement of F– for OH–
groups on quartz surfaces.39 Adsorption capacities of different soils and clays
are compared in Table 7.1, which indicates their fluoride adsorption trends.40
Studies reveal that “hydrated aluminum oxide and iron oxide surfaces occur-
ring in bauxites and goethites/hematites are useful substrates for fluoride
sorption.”1,41 Multiple removal mechanisms such as ligand exchange (with
surface hydroxyl groups and water molecules), anion exchange, electrostatic
attraction, and precipitation are believed to occur.41 Although applications of
different clays for defluoridation are reported in some African countries and
Sri Lanka,40,42 its use in columns is found to be troublesome due to difficul-
ties in packing the columns, controlling the flow, and regenerating the bed.
Moreover, in most of the cases, it would not be cost effective.16 In general, it
could be inferred that the clay process would be of either no or, at least, much
less use in defluoridation, especially when higher removal efficiencies are
expected or higher concentrations of fluoride exist in water.

7.4.3  Carbonaceous and Other Adsorbents


Though attempts were made to attain defluoridation of drinking water by
activated carbon, its defluoridation potential was found to be poor. This ten-
dency can be ascribed to the fact that metallic solids such as AA or activated
bauxite have an intense affinity for fluoride than nonmetallic solids such as
activated carbons. However, it was reported that the adsorption capacity of
activated carbon was found to have doubled due to aluminum impregna-
tion, with a maximum capacity of 1.07 mg/g.43 The carbonized form of the
biomass of an aquatic weed Eichhornia crassipes after thermal activation at
600°C demonstrated a removal of 4.4 mg/g.44 Further, isotherm studies on
the algal biomass of Spirogyra suggested a maximum adsorption capacity of
1.272 mg/g.45 The reported adsorption potential of different types of carbon
is presented in Table 7.2. It was observed that the fluoride removal mecha-
nism by activated carbon was governed primarily by physical adsorption
depending on the specific surface area. Further, the numbers of phenolic
hydroxide groups or carboxyl groups on the carbonaceous material surface
have no role in fluoride uptake.30
Defluoridation Techniques 97

TABLE 7.1
Comparison of Fluoride Adsorption Capacity of Major Clay Types
Maximum
Initial Fluoride Adsorption
Place Concentration Capacity
Sorbent Type Description pH (mg/L) (mg/g)
Gibbsite
Australia 5–7 10.0 0.40
South Africa 5–7 10.0 0.25–0.40
Goethite
Goethite/Kaolinite South Africa 5–7 10.0 0.20
Goethite/Illite China 5–7 10.0 0.23
Goethite/Kaolinite Sri Lanka 5–7 10.0 0.35
Palygorskite
Palygorskite/ South Africa 5–7 10.0 0.21–0.29
Dolomite
Smectite South Africa 5–7 10.0 0.10
United States 5–7 10.0 Trace
Alkaline soil, 5–7 10.0 0.04–0.08
United States
Kaolinite
Kaolinite South Africa 5–7 10.0 0.03
Acidic soils, 5–7 10.0 0.17–0.25
United States
Acid soils, 5–7 10.0 0.130
Illinois
Pottery clay 5–7 10.0 0.12
Clay pots, 5–7 10.0 0.07
Ethiopia
South Carolina, 6–7 16–660 4.05
Australia
Kaolin clay 6 10–250 3.48

Source: Coetzee, P.P., Coetzee, L.L., Puka, R. and Mubenga, S., Water SA, 29, 331–338, 2003.

The use of hydrous ferric oxide (HFO) for defluoridation revealed that
the sorption of fluoride was pH dependent and was taking place by van der
Waal's interaction and ion exchange.46 At an alkaline pH higher than 6, “HFO
functions as a cation-exchanger and adsorbs sodium ions present in solution
releasing protons,”1,46 which may reduce the final pH. The maximum adsorp-
tion capacity of HFO was found to be 16.50 mg/g. The operating mechanism
for fluoride adsorption could be depicted as follows:46
98 Fluoride in Drinking Water

TABLE 7.2
Removal of Fluoride by Various Carbonaceous Materials (Initial Fluoride
Concentration = 20.0 mg/L, Dose of Adsorbent = 10.0 g/L)
Adsorbent Samples Base Material Removal of Fluoride (%)
AC1 Activated carbon (coal) 17
AC2 Activated carbon (wood) 13
AC3 Activated carbon (coal) 12
AC4 Activated carbon (wood) 16
AC5 Activated carbon (wood) 6.6
AC6 Activated carbon (petroleum coke) 5.4
CB Carbon black 10
CC 1 Charcoal 3.5
CC 2 Charcoal 3.2
CC 3 Charcoal 1.6
CC 4 Charcoal 0.40
Source: Abe, I., Iwasaki, S., Tokimoto, T., Kawasaki, N., Nakamura, T. and Tanada, S., J. Colloid
Interface Sci., 275, 35–39, 2004.

In the pH range of 2.0–5.0:

   Fe 2O 3 ⋅ xH 2O(solid) + F(aq ) ↔ Fe 2O 3 ⋅(x −1)H 2O ⋅ H F(solid) + OH(aq ) (7.21)


− + − −

and for pH > 6:

Fe 2O 3 ⋅ xH 2O(solid) + Na(aq )+ + F(aq )− ↔ Fe 2O 3 · (x −1) H 2O


· OH −Na+F(solid)− + H(aq )+ (7.22)

Adsorbents such as Chitin (refers to the polysaccharides commercially


extracted from shellfish processing wastes and naturally available in
organisms such as bacteria, fungi, etc.), its deacetylated product chitosan
(mainly obtained from crustacean shells of prawn, crab, shrimp, or lob-
ster), and 20% lanthanum-incorporated chitosan were tested in remov-
ing excess fluoride from drinking water.47 It was observed that the 20%
La-chitosan showed a much better fluoride removal capacity with a higher
uptake of 3.1 mg/g at an adsorbent dose of 1.5 g/L.47 However, the capac-
ity reduced drastically in treating natural water (nearly reduced to half)
compared with synthetic systems. In experiments dealing with magnetic-
chitosan particles, adsorption capacities of 3–17 mg/g were reported as
corresponding to initial fluoride concentrations of 5–140 mg/L.48 The
Langmuir-saturated monolayer capacity of Laterite, the geomaterial, was
found to be 0.8461 mg/g with an experimental maximum column capac-
ity of 0.3473 mg/g.49 Of late, MgAl-CO3 -layered double hydroxides (LDHs)
were employed to treat high fluoride concentration solutions.50 LDHs
(known as hydrotalcite-like materials) are a class of naturally occurring
Defluoridation Techniques 99

and synthetic anionic clays, and they display a maximum adsorption


capacity of 319.8 mg/g in isotherm studies.50

7.4.4 Alumina
The surface structure and adsorption play an important role in the use of
alumina as a catalyst and its application in separation processes. The acidity–
basicity properties of alumina are found to be the major determinants in
defining its adsorption behavior. Water may be either physisorbed or chemi-
sorbed onto alumina surfaces. However, the amount of sorption depends on
temperature and vapor pressure. Water may be adsorbed as undissociated
molecules with strong hydrogen bonds at ambient temperatures. Further,
at slightly elevated temperatures, surface hydroxyl groups are formed that
will be expelled as H2O at further higher temperatures. “Chemisorption
of water onto alumina surfaces is generally regarded as a Lewis acid–base
adduct formation with the Al3+ ion acting as an electron pair accepter (Lewis
acid) and the hydroxyl ion acting as an electron pair donor (Lewis base). The
hydroxyl groups on the alumina surface are sources of protons and there-
fore behave as Bronsted acid sites. But oxygen bridges formed on the alu-
mina surface through dehydration of two adjacent OH− groups are active
Lewis acid sites.”1,51,52 Knozinger's model is regarded as a total advancement
for understanding the OH surface groups on alumina. It is assumed that
the “termination of alumina crystallites occurs along three possible crystal
planes. Depending on the coordination properties of surface anions and
the number of Al ions attached to hydroxyl groups,” five types of hydroxyl
groups may be present on the three possible crystal planes.1,53
The acidity of the hydroxyl groups as well as the OH-stretching frequency
is affected by the net charge of alumina. “Hydroxyl groups with highest
frequency possess the highest basicity and those with the lowest frequency
are thought to possess the highest acidity. The surface hydroxyl groups
of hydrated alumina are amphoteric in nature and therefore may ionize
as Bronsted acids or bases depending on the circumstances. This ioniza-
tion is responsible for the surface charging at the aqueous interface of alu-
mina”1 and is believed to be generated through a two-step process. The first
step in this direction involves “surface hydration.” This is an effort by the
exposed surface atoms for completing their respective coordination shells.
Aluminum cations achieve surface hydration by linking to an OH− ion or
a water molecule. However, oxygen ions pull out a proton from water. The
second step is the dissociation of the surface hydroxyls. Both the steps create
surface hydroxyls that can ionize as Bronsted acids or bases, giving rise to
surface charges.1
It is plausible that alumina surfaces may get charged by specific adsorption
of ions other than protons. However, it is suggested that pH has considerable
influence on the nature and amount of the surface charge of alumina. “The
point of zero charge (pHZPC) is the pH value at which net surface charge is
100 Fluoride in Drinking Water

zero. In acidic medium below pHZPC, the hydroxyl groups on the surface are
protonated and therefore the surface has a net positive charge. At basic pH,
the surface is negatively charged as the hydroxyl groups act as Bronsted acids
and give off protons.”1,51 This “surface charging” in alumina is instrumen-
tal in the formation of “an electrical double layer” at the aqueous interfaces
due to electrostatic interactions between the charged surfaces and oppositely
charged ions in the bulk solution. It could be inferred that this pH-dependent
“surface charge” and associated “electrical double layer” formation at aque-
ous interfaces turn out to be instrumental in removing impurities from water.
The adsorption on alumina from aqueous solutions can be expressed
in terms of two models, namely, the ligand-exchange model and the ion-
exchange model. The ability of surface hydroxyls to dissociate or to be
protonated (depending on the pH) defines the ion-exchange properties, as
shown by Equations 7.23 and 7.2452 as follows:

M-OH + H+ ↔ M-OH 2+ (7.23)

M-OH + OH – ↔ M-O – + H 2O (7.24)

The ability of alumina to exchange ligand originates from the “presence


of Lewis acid sites on the surface (Al3+) and the presence of water mole-
cules bonded to the sites which can be exchanged for other Lewis based
molecules.” The following equilibria have been used to describe the ligand-
exchange reactions by Equations 7.25 through 7.29) as follows:52,54,55

M(OH) (H 2O) +L1 – ↔ M(H 2O)L1  +OH – (7.25)

M(OH) (H 2O) +L1 – ↔ M(OH)L1 – + H 2O (7.26)

M(OH) (H 2O) +L2 – ↔ M(H 2O)L2 + OH – (7.27)

M(OH) (H 2O)+ L2 – ↔ M(OH)L2 – + H 2O (7.28)

M(OH)L1 – + L2 – ↔ M(OH)L2 – + L1 – (7.29)

M represents the metal oxide metal, and L1 and L2 represent Lewis bases
present in water and solute, respectively. Equations 7.23 and 7.24 describe the
modifications of the surface sites by Lewis bases present in water, whereas
Equations 7.25 through 7.29 show ligand exchange with a harder base from
Defluoridation Techniques 101

the solute (L2). The “driving force for Ligand exchange in alumina is the
affinity Al3+ (a hard Lewis acid) has for hard Lewis bases.”1

7.4.5  Activated Alumina


AA is a granular porous material with a very high surface area (200–300
m2/g); it consists mainly of aluminum oxide (Al2O3). Of late, the application
of AA in the defluoridation of drinking water is widely accepted world-
wide.56–59 The sorption mechanism of AA could be effectively described by
the surface complex formation model (ligand-exchange model). This model
signifies the active role of surface hydroxo groups in surface complex forma-
tion. It is suggested that amphoteric hydroxo groups develop on the alumina
surface due to hydration, as shown by Equations 7.30 and 7.31 as follows:21,25

AlOH 2 +   = AlOH + H + (7.30)

AlOH = AlO – + H+ (7.31)

where AlO –, AlOH2+ and AlOH are, respectively, the negative, positive, and
neutral surface hydroxo and oxo groups. The adsorption model of fluoride
can be described by Equations 7.32 through 7.3421 as follows:

AlOH + F – = AlF + OH – (pH > 7) (7.32)

As represented in the earlier equation, OH– is released from AA surface


into the bulk phase, resulting in an increase in pH. The extent of OH− release
is the most pronounced at pH > 6–7 and diminishes at pH < 7.25,60,61 However,
at a lower range of pH (3–5), a different adsorption mechanism may be in
operation as follows:

AlOH 2+ + F – = AlF + H 2O (pH < 6) (7.33)

At low pHs, it is plausible that the amount of fluoride adsorbed may exceed
the total available surface sites during higher initial fluoride concentrations.
This could be represented by a polynuclear surface complex formation as
follows:21

AlOH + 2F – = AlF2 – + OH – (7.34)

In the presence of fluoride, AA is rendered soluble by forming alumina-


fluoro complexes that are stable in acidic pH and may turn unstable with a
rise in pH. So, AA sorption systems for defluoridation should be carried out
at pH values where alumina-fluoro complexes are unable to resist alumina
from dissolving. The regeneration of AA is usually carried out with a caustic
102 Fluoride in Drinking Water

solution (usually 1% sodium hydroxide) followed by a dilute acid (usually


0.05 N sulfuric acid) and water rinse.1,62 The following reactions take place
during regeneration:

AlF (s)+ OH − → AlOH (s)+ F − (7.35)

AlOH (s)+ H 2SO 4 → AlHSO 4 (s)+ H 2O (7.36)

The characteristic features of AA, excellent scavenging potential, and spe-


cific affinity for fluoride make it an ideal candidate for defluoridation. Packed
beds of granular AA have been traditionally used for defluoridation of pub-
lic water supplies, and their reported capacity is in the range of 6,750–11,760
g/m3 in a continuous flow system. The minimum interference from counter-
ions and the attractive costs are added advantages for AA-based systems.1
The Department of Science and Technology, Government of India, has sup-
ported the Public Health Engineering Departments of State Governments
of fluoride endemic areas to develop hand pump–attached defluoridation
units. A cylindrical defluoridation unit fabricated in this direction and field
tested in Makkur, Unnao district of the state of Uttar Pradesh, in 1993 had
been reported by Daw (2004).56 The AA (particle size 0.3–0.9 mm) bed of 110-
kg alumina in 55-cm height is packed in a mild steel cylindrical drum that
is 0.5 m in diameter and 1.5 m in height. An elevated hand pump discharge
level and elevated platforms are provided. A bypass line was provided to
draw water directly from the hand pump for uses other than drinking and
cooking. A system was designed for a raw water fluoride concentration of 6–7
mg/L. The two important parameters that define the field application of AA
for defluoridation are its fluoride uptake capacity and reuse potential.1,56,58
The application and adaptation of AA as a defluoridating medium with a
hand pump or DDUs started gaining ground in India. The United Nations
Children's Fund (UNICEF) and Rajiv Gandhi Drinking Water Mission
(RGDWM) are taking leadership in this front. The DDU consists of two
chambers. The upper chamber, fitted with a micro-filter and an orifice to
give a flow rate of about 12 L/h, is charged with 3–5 kg of AA packed to a
depth of 9–17 cm. The perforated stainless-steel plate on the top of the AA
bed ensures uniform distribution of raw water. A lid is provided at the upper
chamber, and a tap is provided at the bottom. The raw water filled in the
upper chamber percolates through the adsorbent bed, and the fluoride in the
influent gets adsorbed by the AA media. Treated water can be collected in
the lower chamber, and it can be drawn as and when required.28,56
Further, novel and user-friendly techniques for the regeneration of
exhausted AA were developed. In UNICEF-assisted projects, the spent AA
beds and sludge are used for making bricks. The practical adaptability and
technical viability of AA in small community-level applications, such as
household DDUs and water taps, has been successfully demonstrated56,58,63
Defluoridation Techniques 103

in many countries such as India, China, and Thailand. However, interna-


tional performance standards are yet to be framed.16 Further, AA poses some
limitations. Regeneration brings in a reduction of about 5%–10% in material,
30%–40% in capacity and an increased presence of aluminum (>0.2 mg/L)
in the effluent. It was suggested that normal pHs may keep the aluminum
residual within permissible limits. Though adjustment of pH can be done at
waterworks, actual pH of the raw water is to be relied on in domestic and
small community treatments. So, the fluoride-scavenging potential of AA
has to be verified through field testing under genuine ground-level condi-
tions. Though a high fluoride scavenging potential of AA (to the range of
4–15 mg/g) has been reported in literature,21 field experiences demonstrate a
sharp reduction in its scavenging potential (of around 1 mg/g).16

7.4.6  Other Alumina-Based Adsorbents


Metal oxyhydroxides possess surface oxygen that is different from the num-
ber of coordinating metal ions. This felicitates the adsorption of different
anions and cations by oxyhydroxides. Successful application of activated
bauxite for defluoridation turns significant in this direction.25 Red mud, the
bauxite waste of alumina manufacturing, is an unwanted by-product of alka-
line leaching of bauxite that is also used for defluoridation.64 The adsorption
of fluoride is found to be maximum at pH 5.5. The observed sharp reduction
in fluoride removal at a pH above 5.5 could be ascribed to the stronger com-
petition of fluoride with hydroxide ions on the adsorbent surface. The pos-
sibilities of the formation of weakly ionized hydrofluoric acid in a high acidic
range may also reduce adsorption. The use of r­ efractory-grade bauxite for
fluoride removal65 demonstrates the significance of pH. Alum sludge (a waste
product obtained during the manufacture of alum from bauxite through sul-
furic acid process) was found to have a fluoride adsorption capacity of 5.394
mg/g.24 It was also found that thermal activation at moderate temperatures
(300°C–450°C) increases the adsorption capacity of titanium-rich bauxite.66
Amorphous alumina-supported carbon nanotubes (CNTs) were reported
successful for defluoridation. The adsorbent was prepared by heating the
composites of Al(NO3)3 and CNTs at 500°C for 2 h under N2 atmosphere.
It was observed that the best fluoride adsorption on Al2O3/CNTs occurs
at a wide pH range of 5.0–9.0. The isoelectric point (IEP) of Al2O3/CNTs is
found to be 7.5. At pH < 7.5, the surface charge of the material is positive,
and, consequently, columbic attraction may trigger an interaction between
the adsorbent and fluoride ions. At pH ≥ IEP, due to the neutral or negatively
charged nature of the adsorbent, its scavenging potential may get reduced.
The excellent fluoride adsorption capacity exhibited by Al2O3/CNTs was
also attributed to the amorphous structure of Al2O3. The removal of fluo-
ride by CNTs may be due to their large surface area or surface reactions,
resulting in the formation of functional groups by oxidation. At pH 6, Al2O3/
CNTs exhibit a fluoride adsorption capacity of 28.7 mg/g at equilibrium for
104 Fluoride in Drinking Water

a fluoride concentration of 50 mg/L, which is 13.5 times higher than that of


alumina-impregnated carbons.1,67
Rare earth oxide-based mixtures and rare earth element impregnation
of porous adsorbents or carrier materials have shown promising results in
defluoridation. In this direction, the use of lanthanum-impregnated silica
gel is significant, as it is found to have an adsorption potential of 3.8 mg/g.
The chemical adsorption reaction between lanthanum and silica gel can be
explained as follows:9

−Si − OH → −Si − O – + H+ (7.37)

−Si − O – + La(OH)2+ → −Si − O − La(OH)2 (7.38)

It was observed that fluoride removal by the original silica gel was neg-
ligible in a wide pH range (4–10). This signifies the active role played by
adsorbed lanthanum in defluoridation. In general, the mechanism for the
removal of anions by chemical adsorption by lanthanum-impregnated silica
gel can be explained as follows:9

−Si − O − La(OH)2 + A(z-n) → −Si − O − La(OH) A(z-n ) + n OH –


(7.39)

where A(z-n) is the anion, for instance, fluoride, phosphate, and arsenate ions.
The observed pH rise (from 3.5 to 7.1) is due to the release of OH– ions from the
adsorbent, as shown in Equation 7.39. However, an adsorbent developed by
impregnating lanthanum on cross-linked gelatin has been reported to have
a high fluoride adsorption potential of 21.28 mg/g.68 Gelatin is a polypeptide
with many functional groups. These functional groups have strong attrac-
tion for metal ions. Glutaraldehyde (GTA) is used for cross-linking of gelatin,
which involves the “reaction of free amino groups of lysine or hydroxyly-
sine amino acid residues of the polypeptide chains with the aldehyde groups
of GTA to form a Schiff's base.”1,69 Impregnation was carried out by using
La(NO3)3 solution at different pH. It was observed that a pH range of 5–8 is
optimum for La3+ impregnation of cross-linked gelatin, indicating the prin-
cipal role of carboxyl groups of protein in binding. The removal mechanism
of fluoride and the regeneration process of lanthanum-impregnated cross-
linked gelatin can be represented1,70,71 as shown in Figure 7.2.
The fluoride-scavenging characteristics of zeolite F-9 of size ranging
from 0.15 to 0.30 mm containing surface-active sites created by exchanging
Na+-bound zeolite with Al3+ or La3+ ions was investigated. It was observed
that the three-dimensional skeletal structure of zeolite has small pores in
which the exchangeable cations are located, facilitating the mechanisms of
ion exchange. The exchange of the trivalent ions La3+ and Al3+ for Na+ ions
attached to the zeolite F-9 can be represented as follows:72

Me3+ (soln)+ 3 Na+ (zeo) ↔ Me3+ (zeo)+ 3 Na+ (soln) (7.40)


Defluoridation Techniques 105

F– H2O

R La-OH2 R La-F –
Adsorption

Acid wash Regeneration Alkali wash

F–
H+

R La-OH–

FIGURE 7.2
The adsorption and regeneration mechanism of lanthanum-impregnated cross-linked gelatin.
(From Zhou, Y., Yu, C. and Shan, Y., Sep. Pur. Technol., 36, 89–94, 2004.)

where Me3+ refers to La3+ or Al3+ and (soln) and (zeo) denote the solution
and zeolite phases. The porosimetric studies (done on samples after sorp-
tion) revealed an increase in porosity of zeolite particles from 25% to 32%,
where Al3+ was loaded onto zeolite. When La3+ was loaded onto zeolite, the
corresponding increase was 38%. It was also pointed out that at around neu-
tral pH, the surface of zeolite turns heterogeneous as the trivalent metals
used for surface modification of zeolite form many complexes (protonated
and nonprotonated). At equilibrium, the pH was found to increase from 4.0
to 6.86 in Al3+-exchanged zeolites. For La3+-exchanged zeolites, the increase
was from 4.0 to 6.36, as depicted by the following equations:72

Zeo-MeOH 2+ + OH – ↔ Zeo-MeOH + H 2O (7.41)

Zeo-MeO – + H+ ↔ Zeo-MeOH (7.42)

For fluoride sorption on aluminol surface sites (Me = Al):

Zeo-MeOH 2+ + F – ↔ Zeo-MeF + H 2O (7.43)

Zeo-MeOH + F – ↔ Zeo-MeF + OH – (7.44)

The interaction between La-exchanged zeolite and fluoride can be repre-


sented as (Me = La) follows:

Zeo-MeOH 2+ + F – ↔ Zeo-MeOH 2+ … F – (7.45)


106 Fluoride in Drinking Water

where Zeo-MeOH, Zeo-MeO− and Zeo-MeOH2+ are the neutral, hydroxyl-


ized, and protonated surface sites of zeolite, respectively. The expression
“Zeo-Me” denotes the zeolite–metal surface, where Me represents either La
or Al. As shown, the sorption of fluoride on aluminol sites is mainly by ion
exchange and inner-sphere complexation, as expressed by Equations 7.41
through 7.44. As shown in Equation 7.45, the mechanism of fluoride sorption
in La-exchanged zeolite is of a physical nature, demonstrating the forma-
tion of an outer-sphere complex. Adsorption capacities of 40–43 mg/g and of
42.0–58.5 mg/g were reported for Al3+-exchanged zeolite and La3+-exchanged
zeolite, respectively.72 Alum-impregnated activated alumina (AIAA) was also
experimented for defluoridation.73 AIAA displayed a highly pH-dependent
removal mechanism with adsorption capacities of 192.65, 40.68 and 19.80 mg/g
at pH values of 4.0, 6.5 and 9.0, respectively.73
The use of ion-exchange resins are also found to be effective in removing
fluoride from water. However, during the removal by ion-exchange resins,
it is most likely that similar anions in water will also be removed. So, fluo-
ride exchange capacity mainly depends on the ratio of fluoride to the total
anions in solution. It is naturally expected that the exchange capacity of resins
increases swiftly with higher values of fluoride to total anion ratios. However,
it is reported that most fluoride waters have comparatively low fluoride con-
centrations. In addition to the relatively lower value of exchange capacity, the
high cost of resins compared with other adsorbents such as AA should also be
considered. This limits the use of synthetic anionic resins. Further, since the
removal also includes the sorption of other anions, the sorption capacity of
resins appears too low to the tune of 0.5 mg/L.74 A study on defluoridation by a
two-way ion-exchange cyclic process using two anion-exchange columns was
also reported.11 The results show that the two-way ion-exchange cyclic process
may remove fluoride within the range limits of the defluoridation potential
of anion-exchange resins. A detailed analysis of the adsorbents so far used
in defluoridation clearly suggests that a universally adaptable defluoridation
medium is still an elusive goal. Given the magnitude and diversity of the prob-
lem, the choice mainly depends on the availability of the adsorbent adjacent to
a fluorotic endemic area, that too at reasonable, affordable, or no cost.

7.5  Electrochemical Methods


7.5.1 Electrocoagulation
Electrocoagulation (EC) employs an electrolytic process for producing a
coagulant in situ through the oxidation of appropriate anodic materials. The
coagulant ions thus released react with the ions (targeted to be removed),
initiating a normal coagulation process. Generally, aluminum and iron are
used as sacrificial electrodes that produce ions continuously in the sys-
tem. Additionally, these sacrificial electrodes protect the normal anode by
Defluoridation Techniques 107

reducing its dissolution potential and the cathode through passivation.


When an electric current passes through the Al electrodes, an anodic reac-
tion releases Al ions that react with OH ions produced at the cathode and
with F ions in solution. As a result of electro-condensation, the defluorida-
tion efficiency of the EC system may be higher than the traditional coagula-
tion process. Since F− ions are attracted to the anode, the concentration of
F− ions near the anode exceeds that in the bulk solution, leading to higher
efficiency through the “condensation effect” (Figure 7.3).10
Of late, the EC process was modified through the introduction of bipolar
electrodes.75 For developing bipolarity, a conductive plate without any elec-
tric connection is placed between two electrodes having opposite charges,
as shown in Figure 7.4. The anodic reaction will commence at the positive
side of the bipolar electrode, and cathodic reactions will commence at the
negative side. The Al3+ ions (produced by the dissolution of Al anode) at
appropriate pHs get initially transformed to aluminum hydroxide Al(OH)3
and ultimately polymerized to Aln(OH)3n, which have a very strong affinity
toward F− ions as follows:76

Al → Al 3+ + 3e – (7.46)

Al 3+ + 3 H 2O → Al(OH)3 + 3H+ (7.47)

nAl(OH)3 → Al n (OH)3n (7.48)

Al anode
+
+ Al3+ F–
3+
+ Al
Cl– Na+
+

+ F
F– Cl–
+
+ Na+ Cl–

+ Cl
Al3+
+ Al3+
+ Al3+
F– F–
+ – Cl–
F
+ Al3+
+ Al3+ Cl–
+ F–
+ Cl– F–
+ F–
+ Al3+ Cl–
F–

FIGURE 7.3
Diagram representing the electro-condensation effect. (From Hu, C.Y., Lo, S.L., and Kuan,
W.H., Water Res., 37, 4513–4523, 2003.)
108 Fluoride in Drinking Water

+ – + –
+ – H2 + –
O2
+ – + – +

+ – F– Al3+ + – H2
O2 + – + –
+ – OH– + –

FIGURE 7.4
Schematic diagram of bipolar electrode system. (Modified from Mameri, N., Yeddou, A.R.,
Lounici, H., Belhocine, D., Grib, H. and Bariou, B., Water Res., 32, 1604–1612, 1998.)

It could also be plausible that the Al3+ ions under high fluoride concen-
trations at the anode may be induced to form AlF63− ions (Equation 7.49).
These may get transformed into an insoluble salt (Na3AlF6) by sodium ions
(Equation 7.50).76

Al 3+ + 6F – → AlF6 3− (7.49)

AlF6 3– + 3Na+ → Na 3 AlF6 (7.50)

The reactions encountered at the cathodic (Equation 7.51) and anodic com-
partments (Equation 7.52) are as follows:

2 H 2O + 2e – → H 2 + 2OH – (7.51)

2OH – → 12 O 2 + H 2O + 2e – (7.52)

The efficiency of this process was nearly 100% at optimum pH (5–7.6). However,
a rise in temperature may reduce the efficiency due to fluoride desorption from
aluminum hydroxide and/or by destruction of fluoro-­aluminum complexes.
Recently, the concept of hydro-fluoro-aluminum complex Al n (OH)m Fk 3n-m-k
has been introduced to describe various chemical species containing fluo-
ride, hydroxide, and Al3+, such as Al-F complexes, Al-OH complexes and
the hydro-fluoro-aluminum colloid flocs.77 The electrical field encourages
Al n (OH)m Fk 3n-m-k to be condensed near the electrodes, making fluoride sorp-
tion mostly on the electrodes and resulting in a higher defluoridation efficiency
in the EC process.1 However, issues such as interference from other anions pres-
ent in water due to the competition effect may reduce the efficiency of the EC
process. For example, the presence of sulfate may significantly reduce the effi-
ciency of the process due its strong affinity toward Al3+ ions. Further, regular
replacement of sacrificial electrodes due to continuous dissolution into solution
because of oxidation and high consumption of electric power during operation
may also be treated as limitations of the process.
Defluoridation Techniques 109

7.5.2 Electrosorption
Electrosorptive techniques could be used to enhance sorption capacity of
the conventional systems. It was demonstrated that the efficiency of the alu-
mina bed got significantly enhanced through a new activation technique
by the application of an electric field. An electrochemical cell is prepared
with two stainless-steel electrodes. These electrodes were introduced into
a 20-cm-long and 2-cm-diameter PVC column to produce an electrical field
in the AA bed. This technique exhibited more efficiency than the con-
ventional activation techniques. The regeneration of an adsorbent plays a
crucial role in its selection, economical evaluation, and field application.
Any reduction in the cost of the regeneration will significantly add to the
efficiency of the adsorbent bed and the process.78 It was observed that dur-
ing regeneration, the electrical field created between two electrodes gave
greater mobility to OH ions in accessing active sites located within the
pores, thereby improving the regeneration of AA. Successive regenera-
tion in three cycles could be done without any reduction in the sorption
capacity of the used adsorbent. Moreover, the volume of the cleaning agent
(NaOH) used for regeneration got drastically reduced. Further, more than
95% recovery of the adsorption capacity was ensured with electrosorption
techniques. It is also pointed out that the amount of water necessary for
the regeneration of the saturated bed was very minimal compared with
conventional techniques.1,78

7.6  Membrane Processes


Membrane may be treated as a selective barrier controlling material trans-
port between two adjacent phases. Thus, it could be regarded as an inter-
phase between the two phases. So, in separation science, the main advantage
of membrane technology is related with the transport selectivity of the
membrane.79 A wide range of technical applications are possible by altering
or changing the membrane structure, cross section, and shape. The differ-
ence in chemical potential due to a concentration or pressure gradient across
the membrane (or by an electric field) acts as the driving force for passive
transport in membrane separations. Depending on the membrane barrier
structures and the trans membrane gradients, the process is characterized
as shown in Table 7.3.
Though different membrane technologies (Table 7.3) have been used over
the years in diverse fields, their application in removing inorganic anions
such as fluoride is of recent origin. The applications of membrane processes
turn attractive in water treatment, as many of the difficulties associated
with precipitation, coagulation, or adsorption can be avoided or minimized.
Recently, many hybrid systems have been developed in which one membrane
110 Fluoride in Drinking Water

TABLE 7.3
Classification of Membranes and Membrane Processes for Separations via Passive
Transport
Trans Membrane Gradient
Membrane Barrier Electrical
S. No. Structure Concentration Pressure Field
1 Nonporous Reverse Electrodialysis
osmosis (RO) (ED)
2 Microporous pore Dialysis (D) Nanofiltration
diameter, dp ≤ 2 nm (NF)
3 Mesoporous pore Dialysis (D) Ultrafiltration Electrodialysis
diameter , dp = 2–50 nm (UF) (ED)
4 Macroporous pore Microfiltration
diameter, dp = 50–500 nm (MF)
Source: Ulbricht, M., Polymer, 47, 2217–2262, 2006.

process has been integrated with another to produce even higher-quality


water.6 The basic principles governing these membrane-based processes are
discussed next.

7.6.1  Reverse Osmosis


Reverse osmosis (RO) is a pressure-driven membrane process. In this pro-
cess, the driving force for transport across the membrane is the pressure
gradient between the water to be treated and the permeate side. The most
popular application of RO is in desalination of drinking water and process
water. The applied operating trans membrane pressure ranges from 20 to
100 bar.6 The type of polymers used and the morphological details of the
barrier are given in Table 7.4. The application of RO is similar to the ordinary
filtration process. The major difference is in the role of osmotic pressure. The
osmotic pressure is relatively insignificant in ordinary filtration; however, it
is a crucial factor in RO. The RO modules of spiral-wound polyamide mem-
branes were successfully employed in treating electronic industry wastewa-
ter. This wastewater has an average fluoride concentration around 400 mg/L
and is generated due to the chemical cleaning of silicon plates with a diluted
hydrofluoric acid.80 It was reported that a bench-scale RO unit with a capac-
ity of 0.15 m3/h removed more than 60% of fluoride from feed water concen-
trations of 1.7 mg/L employing spiral-wound cellulose acetate membranes.81
The implementation of a pilot scale RO plant in Lakeland, California, was
also reported successful in treating fluoride concentration in the range of
3.6–5.3 mg/L.82
The major limitations of RO include the high energy consumption
needed for maintaining the required pressure difference. In addition to
Defluoridation Techniques 111

TABLE 7.4
Polymer-Based Separation Membranes for Different Processes
Morphology
Barrier
Membrane Thickness
Process Polymer Barrier Type Cross Section (μm)
Reverse osmosis Cellulose acetates Nonporous Anisotropic ~0.1
Polyamide, aromatic, Nonporous Anisotropic/ ~0.05
in situ synthesized composite
Polyether, aliphatic Nonporous Anisotropic/ ~0.05
cross-linked, in situ composite
synthesized
Nanofiltration Polyamide, aromatic, Nonporous Anisotropic/ ~0.05
in situ synthesized composite
Polyether, aliphatic Nonporous Anisotropic/ ~0.05
cross-linked, in situ composite
synthesized
Polyimides Nonporous Anisotropic ~0.1
Polysiloxanes Nonporous Anisotropic/ ~0.1<1–10
composite
Ultrafiltration Cellulose acetates Mesoporous Anisotropic ~0.1
Cellulose, regenerated Mesoporous Anisotropic ~0.1
Polyacrylonitrile Mesoporous Anisotropic ~0.1
Polyetherimides Mesoporous Anisotropic ~0.1
Polyethersulfones Mesoporous Anisotropic ~0.1
Polyamide, aromatic Mesoporous Anisotropic ~0.1
Polysulfones Mesoporous Anisotropic ~0.1
Polyvinylidenefluoride Mesoporous Anisotropic ~0.1
Electrodialysis Perfluorosulfonic acid Nonporous Isotropic 50–500
polymer
Poly(styrene-co- Nonporous Isotropic 100–500
divinylbenzene),
sulfonated
Source: Ulbricht, M., Polymer, 47, 2217–2262, 2006.

biological fouling of the membrane (due to natural organic matter [NOM]


and microorganisms), mineral fouling may also occur due to precipitation
of certain salts. Further, interference from divalent metallic cations such
as Ca2+ and Mg2+, and anions such as SO42− may interfere with the separa-
tion of particular toxic anionic species. This may affect water recovery and
reduce the osmotic pressure83 demanding pre treatment of raw water. At
lower pH values, because of strong hydrogen bonding of fluoride in acidic
solution, its permeability and solubility in the membrane may increase.
This will drastically reduce the life of the membrane and force its replace-
ment.84 It was reported that while treating fluoride-rich waters, up to 99%
112 Fluoride in Drinking Water

of the salts in water were getting rejected by the membrane.85 This may
result in unpleasant taste, as treated water may lack the right balance of
minerals.1,86

7.6.2 Nanofiltration
NF is a membrane-separation process that is targeted to remove uncharged
organic species on nanoscale by size exclusion, and ions by charge effects.87
UF is a membrane processes targeting macromolecular substances. RO,
UF, and NF are pressure-driven membrane processes targeting removal
of multivalent ions from monovalent species. Through judicious selec-
tion of membranes and operating conditions, ions of the same valence can
also be separated. NF membranes are essentially low-pressure RO mem-
branes that are capable of removing hardness along with a wide range
of organic (bacteria, viruses, and pesticides) and inorganic components
(nitrates, arsenic, and fluoride) in a single process.1,87 NF membranes pro-
vide higher water fluxes at lower trans membrane pressures6 than con-
ventional RO membranes. So, NF membranes are commonly referred to
as low-pressure RO membranes.6 The applied pressure gradient across the
membrane paves the way for solute transport by convection. Given a nega-
tive charge in a neutral and alkaline environment, NF membranes provide
an asymmetric pattern (Table 7.4). The surface charge on NF membranes
is due to the anions adsorbed onto their surface. In contrast, the surface
charge in ion-exchange membranes is due to various fixed charged groups
that are bonded to the polymer structure.6,88 In addition to diffusion and
convection mechanisms, the repulsion between anions is also significant.
These repulsive actions between surface groups will be higher in the case
of multivalent anions.6,87 In addition to size-based exclusion, the system
offers mechanisms of ion exclusion as well. So, relatively higher degrees
of ion rejections or separations at higher water fluxes, as in the case of RO,
could be achieved through the NF membranes.1 Of late, one of the world's
largest NF plants is operative in Paris, France, with a capacity of 140,000
m3/d and having more than 9,000 Filmtec NF200 membrane modules.86
Another NF plant in Finland (capacity of 380–600 m3/d) was reported
successful in fluoride removal to the tune of 76% with NF 255 (Filmtec).89
Negatively charged commercial thin-film composite (TFC) membranes
were also used for fluoride removal.85 Most recently, hybrid systems using
NF in combination with subsequent adsorption or biodegradation were
also attempted.6 Recent trends in membrane development encompass the
applications of existing RO thin-film composite membranes to the usage of
NF for water treatments. The replacement of multiple treatment processes
and techniques by single-membrane applications is the need of the hour.
Currently, this triggers advanced research in NF.
This trend is gaining ground as one of the best available technologies
in many developed countries. Though NF is a preferred option and an
Defluoridation Techniques 113

increasingly recommended technique globally as against the traditional RO


systems, it is constrained with limitations as well. Among them, the treat-
ment and management of relatively large concentrations of retentate frac-
tion (generated to the tune of nearly 20% of the feed volume) is a matter of
concern.

7.6.3 Electrodialysis
Electrodialysis (ED) is a process in which the transport of ions present in
wastewater is accelerated due to an externally applied electric potential dif-
ference. ED has been conventionally used for desalination and demineraliza-
tion of brackish waters. In an ED cell, cation- and anion-selective membranes
are placed in a parallel fashion across the current path90 (Figure 7.5). By
introducing a current, cations move through the cation-exchange membrane
toward the cathode and anions move through the anion-exchange membrane
to the anode. As a result, salinity may decrease in one space in the alternat-
ing spaces between the membranes. However, salinity may increase in the
next space throughout the stack of parallel membranes. The desired level of
salinity could be achieved once water is made to pass through several such
stacks. The cost of ED is directly proportional to the salinity of the water to
be treated. As in RO, ED is also bound with limitations of membrane foul-
ing. This may largely be ascribed to the presence of colloidal elements and
the precipitation of sparingly soluble salts such as ferric hydroxides, calcium
sulfates, and calcium carbonates. Most negatively charged colloids present in
natural waters may get deposited on the anion-exchange membranes.6

Diluted solution
Concentrated solution
Electrode rinse solution
AEM CEM AEM CEM AEM CEM

Cathode Anode
+

Feeding solution
Anions AEM Anion-exchange membrane
Cations CEM Cation-exchange membrane

FIGURE 7.5
Schematic diagram of a typical ED Cell. (From Marder, L., Sulzbach, G.O., Bernardes, A.M. and
Ferreira, J. Z., J. Braz. Chem. Soc., 14, 610–615, 2003.)
114 Fluoride in Drinking Water

It was reported that an ED cell containing 15 cell pairs of cation- and


anion-exchange membranes of 80-cm2 effective cross-sectional areas was
found to be successful in multiple ways. In addition to defluoridation,
instantaneous desalination of brackish water (at a reasonable energy cost of
2.5–5 kWh/m3) of product water was also reported feasible.91 Membranes
used in this cell were prepared from interpolymeric films based on high-
density polyethylene (HDPE), linear low-density polyethylene (LLDPE), and
styrene-divlnylbenzene, which had an 80:20 ratio of HDPE to LLDPE. The
fluoride in the treated water was found to be reduced from 15 to 1.5 mg/L,
and total dissolved solid (TDS) in brackish water was found to be reduced
from 5000 to 15 mg/L. The ED was found to be successful in removing fluo-
ride from brackish waters of Southern Algeria that had a TDS of 3000 mg/L
and a fluoride concentration of 3 mg/L. The ED cell used for this purpose
comprised 10 pairs of anion- and cation-exchange membranes (0.18 mm,
mono-anion, permselective). However, it was suggested that both RO and
ED require a high degree of pre treatment for application. The removal of
suspended solids and organics is very much essential to prevent membrane
fouling. It is to be noted that proper provision should be made to offset the
losses due to the retent formation in the order of 10%–25% of permeate.1,92
Operations of an ED cell with a capacity of 1.0 m3/h comprising two anion-
exchange membranes using a pre selected membrane in removing fluoride
from Moroccan ground water was also reported. The study demonstrated
that the desired drinking water quality can be easily obtained by ED under
predetermined optimized operating conditions.93 The recovery rates in ED
were higher than in other membrane processes such as RO. The ED can be
made operational over longer periods in response to fluctuations in feed
concentrations. This makes EC suitable for seasonal operations.93 Although
ED is found to be more economical, the quality of the treated water will
be slightly inferior to that of RO. Research suggests that ED is efficient in
removing inorganic anions such as fluoride from drinking water.

7.7  Defluoridation Techniques: A Summary


Although many technologies are available for defluoridation, as discussed
earlier, with each having its own advantages and limitations, a lasting solu-
tion is still at large. Coagulation methods are proved to be unsuccessful in
bringing fluoride to desired concentration levels. Sorption methods (includ-
ing ion-exchange and adsorption techniques) are more effective in reducing
fluoride concentrations below permissible limits compared with precipita-
tion methods. Among the adsorbents used, AA was reported successful at
the implementation level. Regeneration of used adsorbents and associated
treatment costs, reduction in adsorption capacity after regeneration, and
Defluoridation Techniques 115

pre treatment requirement of pH adjustments could be quoted as the major


limitations of most of the adsorbents discussed. Membrane separations are
prone to fouling, membrane degradation, and scaling. Further, these are
comparatively expensive to install and operate. The electrochemical tech-
niques are constrained by high costs, during both operation and mainte-
nance. The high cost of technology is a limiting factor in the adaptation of
treatment systems in many fluoride endemic areas of the developing world.
It is important to consider the social acceptance and sustainability of the
technology. Needless to say, the “best available technology” need not always
be the “most appropriate.” Paradoxically, though defluoridation research
has made significant advancement, any universal sustainable solution still
appears intangible.1

7.8 Summary
• The defluoridation techniques generally practiced include coagula-
tion, adsorption/ion exchange, electrochemical methods, and mem-
brane processes.
• Defluoridation processes by coagulation include precipitation of
fluoride by a suitable reagent through chemical reactions and/or co-
precipitation of fluoride. This involves simultaneous precipitation of
fluoride with a macro-component from the same solution through
the formation of mixed crystals, by multiple mechanisms such as
adsorption, occlusion, or mechanical entrapment.
• Alum treatment is confined to high dosage requirement, issues of
sludge disposal, high pH of the treated water, and residual alumina
in treated water.
• Adsorption or ion exchange is one of the most frequently used meth-
ods for defluoridation. Water laced with fluoride is passed through a
column packed with an adsorbent and on saturation, the adsorbent
bed is be backwashed for reuse. The adsorption capacity, cost of the
adsorbent, ease in operation, and potential for reuse and regenera-
tion are some of the factors that define the selection of an adsorbent.
• The defluoridation potential of activated carbon was found to be
poor. The nonmetallic solids such as activated carbons are found to
have only weaker attractions for fluoride than that of metallic solids
such as activated bauxite or AA.
• AA is a granular porous material with a very high surface area; it
consists mainly of aluminum oxide. The application of AA in the
defluoridation of drinking water is widely accepted worldwide. The
sorption mechanism could be effectively described by the surface
116 Fluoride in Drinking Water

complex formation model depicting the role of surface hydroxo


groups in surface complex formation.
• Though many adsorbents are widely used for defluoridation, none
of these adsorbents is considered a “universal defluoridation media”
given the magnitude and diversity of the problem. The proximity
and availability of these adsorbents that are adjacent to a fluorotic
endemic area at reasonable or no cost may be the most appropriate
choice for consideration.
• The regular replacement of sacrificial electrodes due to continuous
dissolution into solution because of oxidation and the high con-
sumption of electric power during operation may also be treated as
limitations of the electrocoagulation process.
• The application of membrane technologies in removing inorganic
anions such as fluoride is of recent origin. This process turns attrac-
tive in water treatment, as many of the difficulties associated with
precipitation, coagulation, or adsorption can be either avoided or
minimized. Recently, many hybrid systems have been developed in
which one membrane process has been integrated with another to
produce even higher-quality water.
• The major limitations of RO include the high energy consumption
needed for maintaining the required pressure difference. In addi-
tion to biological fouling, membranes are subjected to NOM and
microorganisms. Mineral fouling may also occur due to precipita-
tion of certain salts. Further, interference from divalent metallic cat-
ions such as Ca2+ and Mg2+ and anions such as SO42− may interfere
with the separation of particular toxic anionic species. This may
affect water recovery and reduce the osmotic pressure demanding
pre treatment of raw water.
• The recovery rates in electrodialysis were higher than in other mem-
brane processes such as RO. The ED can be made operational over
longer periods in response to fluctuations in feed concentrations,
making it suitable for seasonal operations. Though ED is found to
be more economical, RO ensures more quality of the treated water
compared with it.
• Membrane separations are prone to membrane degradation, scaling,
and fouling. It is relatively expensive to install and operate. The elec-
trochemical techniques are constrained by high costs, during both
operation and maintenance.
• Although many technologies are available for defluoridation, with
each having its own advantages and limitations, a lasting solution
is still at large. It appears that “though defluoridation research has
made significant advancement, any universal sustainable solution
still appears intangible.”
Defluoridation Techniques 117

References
1. Ayoob, S., Gupta, A.K and Bhat, Venugopal, T. (2008). A conceptual overview
on sustainable technologies for the defluoridation of drinking water. Crit. Rev.
Environ. Sci. Technol., 38, 6, 401–470.
2. Aksu, Z. and Gönen, F. (2004). Biosorption of phenol by immobilized activated
sludge in a continuous packed bed: Prediction of breakthrough curves. Process
Biochem., 39, 599–613.
3. Lounici, H., Belhocine, D., Grib, H., Drouiche, M., Pauss, A. and Mameri, N.
(2004). Fluoride removal with electro-activated alumina. Desalination, 161,
287–293.
4. Hu, C.Y., Lo, S.L. and Kuan, W.H. (2005). Effects of the molar ratio of hydroxide
and fluoride to Al(III) on fluoride removal by coagulation and electrocoagula-
tion. J. Colloid Interface Sci., 283, 472–476.
5. Mjengera, H. and Mkongo, G. (2003). Appropriate deflouridation technology
for use in flourotic areas in Tanzania. Phys. Chem. Earth, 28, 1097–1104.
6. Velizarov, S., Crespo, J.G. and Reis, M.A. (2004). Removal of inorganic anions
from drinking water supplies by membrane bio/processes. Rev. Environ. Sci.
Bio/Technol., 3, 361–380.
7. Gregory, J. and Duan, J. (2001). Hydrolyzing metal salts as coagulants. Pure
Appl. Chem., 73, 2017–2026.
8. Reardon, J.E. and Wang, Y. (2000). A limestone reactor for fluoride removal
from wastewaters. Environ. Sci. Technol., 34, 3247–3253.
9. Wasay, S.A., Haran, M.J. and Tokunaga, S. (1996). Adsorption of fluoride,
phosphate, and arsenate ions on lanthanum-impregnated silica gel. Water
­
Environ. Res., 68, 295–300.
10. Hu, C.Y., Lo, S.L. and Kuan, W.H. (2003). Effects of co-existing anions on fluoride
removal in electrocoagulation (EC) process using aluminum electrodes. Water
Res., 37, 4513–4523.
11. Castel, C., Schweizer, M., Simonnot, M.O. and Sardin, M. (2000). Selective
removal of Fluoride ions by a two-way ion-exchange cyclic process. Chem. Eng.
Sci., 55, 3341–3352.
12. Huang, C.J. and Liu, J.C. (1999). Precipitation flotation of fluoride containing
wastewater from semi-conductor manufacture. Water Res., 33, 3403–3412.
13. Lislie, A.L. (1967). Means and methods of defluoridation of water, United States
Patent No. 3,337,453.
14. Rao, S.M. and Mamatha, P. (2004). Water quality in sustainable water
­management. Curr. Sci., 87, 942–947.
15. Dahi, E., Mtalo, F., Njau, B. and Bregnhj, H. (1996). Defluoridation using the
Nalgonda technique in Tanzania. In: 22nd WEDC Conference, Reaching the
Unreached: Challenges For the 21st Century, New Delhi, India, pp. 266–268.
16. WHO (2006). In: Fawell, J., Bailey, K., Chilton, J., Dahi, E., Fewtrell, L. and Magara,
Y. (Eds.), Fluoride in Drinking Water, pp. 41–75. London, UK: IWA Publishing.
17. Qureshi, N. and Malmberg, R.H. (1985). Reducing aluminum residuals in
­finished water. J. AWWA, 77, 101–108.
18. Peavy, S.H., Rowe, R.D. and Tchobanoglous, G. (1985). Environmental Engineering,
International Edition, pp. 134–135. Singapore: McGraw-Hill Book Co.
118 Fluoride in Drinking Water

19. Weber, W.J. Jr. (1972). Physicochemical Processes for Water Quality Control, pp. 199–
360. New York: A Wiley-Inter science Publication, John Wiley and Sons.
20. Mekonen, A., Kumar, P. and Kumar, A. (2001). Integrated biological and physi-
cochemical treatment process for nitrate and fluoride removal. Water Res., 35,
3127–3136.
21. Hao, J.O. and Huang, C.P. (1986). Adsorption characteristics of fluoride onto
Hydrous Alumina. J. Environ. Eng. (ASCE), 112, 1054–1067.
22. Bulusu, K.R. (1984). Defluoridation of water using combination of aluminum
chloride and aluminum sulfate. J. Inst. Eng. (India), 65, 22–26.
23. Gupta, S.K., Gupta, A.B., Dhindsa, S.S., Seth, A.K., Agrawal, K.C. and Gupta,
R.C. (1999). Performance of domestic filter based on KRASS defluoridation pro-
cess. J. Indian Water Works Assoc., 31, 193–200.
24. Sujana, M.G., Thakur, R.S. and Rao, S.B. (1998). Removal of fluoride from
­aqueous solutions by using alum sludge. J. Colloid Interface Sci., 206, 94–101.
25. Choi, W.W. and Chen, K.Y. (1979). The removal of fluoride from waters by
adsorption. J. AWWA, 71, 562–570.
26. Nawlakhe, W.G., Kulkarni, D.N., Pathak, B.N. and Bulusu, K.R. (1975).
Defluoridation of water by nalgonda technique. Indian J. Environ. Health, 17,
26–65.
27. CPHEEO (1991). Central Public Health and Environmental Engineering Organization,
Manual on Water Supply and Treatment. 3rd edn, pp. 289–297. New Delhi, India:
The Controller of Publications.
28. RGNDWM (2001). Rajiv Gandhi National Drinking Water Mission, Making
water safe; two user friendly methods to deal with the scourage of fluoro-
sis, Jalavani-news letter on rural water and sanitation in India, Published by
RGNDWM and Water and Sanitation Program–South Asia (WSP-SA), 3, p.7.
29. Susheela, A.K. (2003). A Treatise on Fluorosis, revised 2nd edn. New Delhi, India:
Fluorosis Research and Rural Development Foundation.
30. Abe, I., Iwasaki, S., Tokimoto, T., Kawasaki, N., Nakamura, T. and Tanada, S.
(2004). Adsorption of fluoride ions onto carbonaceous materials. J. Colloid
Interface Sci., 275, 35–39.
31. Bhargava, D.S. and Killedar, D.J. (1991). Batch studies of water defluoridation
using fishbone charcoal. J. Water Pollut. Control Fed., 63, 848–858.
32. Zevenbergen, C., van Reeuwijk, L.P., Frapporti, G., Louws, R.J. and Schuiling,
R.D. (1996). Simple method for defluoridation of drinking water at village level
by adsorption on Ando soil in Kenya. Sci. Total Environ., 188, 225–232.
33. Agarwal, M., Rai, K., Shrivastav, R. and Dass, S. (2003). Deflouridation of water
using amended clay. J. Cleaner Prod., 11, 439–444.
34. Zhuang, J. and Yu, G.R. (2002). Effects of surface coatings on electrochemical
properties and contaminant sorption of clay minerals. Chemosphere, 49, 618–629.
35. Kau, P.M.H., Smith, D.W. and Binning, P. (1998). Experimental sorption of fluo-
ride by kaolinite and bentonite. Geoderma, 84, 89–108.
36. Hingston, F.J., Posner, A.M. and Quirk, J.P. (1972). Anion adsorption by goe-
thite and gibbsite I, The role of the protein in determining adsorption envelops.
J. Soil Sci., 23, 177–192.
37. Davis, J.A. and Kent, D.B. (1990). Surface complexation modeling in aque-
ous geochemistry. Reviews in Mineralogy, Vol. 23, pp. 177–260. Mineral-Water
Interface Geochemistry. In: Hochella Jr., M.F., White, A.F. (Eds.), Washington, DC:
Mineralogical Society of America.
Defluoridation Techniques 119

38. Hiemstra, T. and Van Riemsdijk, W.H. (2000). Fluoride adsorption on


Goethite in relation to different types of surface sites. J. Colloid Interface Sci.,
225, 94–104.
39. Fan, X., Parker, D.J. and Smith, M.D. (2003). Adsorption kinetics of fluoride on
low cost materials. Water Res., 37, 4929–4937.
40. Coetzee, P.P., Coetzee, L.L., Puka, R. and Mubenga, S. (2003). Characterisation of
selected South African clays for defluoridation of natural waters. Water SA, 29,
331–338.
41. Harrington, L.F., Cooper, E.M. and Vasudevan, D. (2003). Fluoride sorption and
associated aluminum release in variable charge soils. J. Colloid Interface Sci., 267,
302–313.
42. Padmasiri, J.P. and Dissanayake, C.B. (1995). A simple defluoridator for remov-
ing excess fluorides from fluoride-rich drinking water. Int. J. Environ. Health
Res., 5, 153–160.
43. Ramos, R.L., Ovalle-Turrubiartes, J. and Sanchez-Castillo, M.A. (1999).
Adsorption of fluoride from aqueous solution on aluminum impregnated
­carbon. Carbon, 37, 609–617.
44. Sinha, S., Pandey, K., Mohan, D. and Singh, K.P. (2003). Removal of fluoride from
aqueous solutions by Eichhornia crassipes biomass and its carbonized form. Ind.
Eng. Chem. Res., 42, 6911–6918.
45. Mohan, S.V., Ramanaiah, S.V., Rajkumar, B. and Sarma, P.N. (2007). Biosorption
of fluoride from aqueous phase onto algal Spirogyra IO1 and evaluation of
adsorption kinetics. Bioresour. Technol., 98, 1006–1011.
46. Dey, S., Goswami, S. and Ghosh, C.U. (2004). Hydrous Ferric Oxide (HFO)—A
Scavenger for Fluoride from Contaminated Water. Water Air Soil Pollut., 158,
311–323.
47. Kamble, S.P., Jagtap, S., Labhsetwar, N.K., Thakare, D., Godfrey, S., Devotta, S.
and Rayalu, S.S. (2007). Defluoridation of drinking water using chitin, chitosan
and lanthanum-modified chitosan. Chem. Eng. J., 129, 173–180.
48. Ma, W., Ya, F.Q., Han, M. and Wang, R. (2007). Characteristics of equilibrium,
kinetics studies for adsorption of fluoride on magnetic-chitosan particle.
J. Hazard. Mater., 143, 296–302.
49. Sarkar, M., Banerjee, A., Pramanick, P.P. and Sarkar, A.R. (2007). Design and
operation of fixed bed laterite column for the removal of fluoride from water.
Chem. Eng. J., 131, 329–335.
50. Lv, L., He, J., Wei, M., Evans, D.G. and Zhou, Z. (2007). Treatment of high fluo-
ride concentration water by MgAl-CO3 layered double hydroxides: Kinetic and
equilibrium studies. Water Res., 41, 1534–1542.
51. Kasprzyk-Hordern, B. (2004). Chemistry of alumina, reactions in aqueous solu-
tions and its application in water treatment. Adv. Colloid Interface Sci., 110, 19–48.
52. Nawrocki, J., Dunlap, C., McCormick, A. and Carr, P.W. (2004). Part I.
Chromatography using ultra-stable metal oxide-based stationary phases for
HPLC. J. Chromatogr. A, 1028, 1–30.
53. Knozinger, H. and Ratnasamy, P. (1978). Catalytic aluminas: Surface models
and characterization of the surface sites. Catal. Rev. Sci. Eng., 17, 31–70.
54. Nawrocki, J., Rigney, M.P., McCormick, A. and Carr, P.W. (1993). Chemistry of
zirconia and its use in chromatography. J. Chromatogr. A, 657, 229–282.
55. Blackwell, J.A. and Carr, P.W. (1991). Study of the fluoride adsorption character-
istics of porous microparticulate zirconium oxide. J. Chromatogr., 549, 43–57.
120 Fluoride in Drinking Water

56. Daw, R.K. (2004). Experiences with domestic defluoridation in India. In:
Sam, G. (Ed.), People-Centred Approaches to Water and Environmental Sanitation.
Proceedings of the 30th WEDC International Conference, October 2004,
pp. 467–473. Vientiane, Lao PDR: Lao National Cultural Hall.
57. Ghorai, S. and Pant, K.K. (2005). Equilibrium, kinetics and breakthrough studies
for adsorption of fluoride on activated alumina. Sep. Pur. Technol., 42, 265–271.
58. Chauhan, V.S., Dwivedi, P.K., Iyengar, L. (2007). Investigations on activated alu-
mina based domestic defluoridation units. J. Hazard. Mater., 139, 103–107.
59. Nakkeeran, E. and Sitaramamurthy, D.V. (2007). Removal of fluoride from
ground water. Can. J. Pure Appl. Sci., 1, 79–82.
60. Bishop, P.L. and Sansoucy, G. (1978). Fluoride removal from drinking water by
fluidized activated alumina adsorption. J. AWWA, 70, 554–559.
61. Rubel, F. and Woosley, R.D. (1979). The removal of excess fluoride form drink-
ing water by activated alumina. J. AWWA, 71, 45–49.
62. Schoeman, J.J. and MacLeod, H. (1987). The effect of particle size and interfer-
ing ions on fluoride removal by activated alumina. Water SA, 13, 229–234.
63. Dahi, E. (2000). The State of Art of Small Community Defluoridation of
Drinking Water. In: Dahi, E., Rajchagool, S., Osiriphan, N. (Eds.), Proceedings
of the 3rd International Workshop on Fluorosis Prevention and Defluoridation
of Drinking Water, Chiang Mai, Thailand, pp. 137–167. http://www.icoh.org
/download/3rdproceeding.pdf.
64. Cengeloglu, Y., Kir, E. and Ersoz, M. (2002). Removal of fluoride from aqueous
solution by using red mud. Sep. Purif. Technol., 28, 81–86.
65. Mohapatra, D., Mishra, D., Mishra, S.P., Chaudhury, G.R. and Das, R.P. (2004).
Use of oxide minerals to abate fluoride from water. J. Colloid Interface Sci., 275,
355–359.
66. Das, N., Pattanaik, P. and Das, R. (2005). Defluoridation of drinking water using
activated titanium rich bauxite. J. Colloid Interface Sci., 292, 1–10.
67. Li, H.Y., Wang, S., Cao, A., Zhao, D., Zhang, X., Xu, C., Luan, Z., Ruan, D., Liang,
J., Wu, D. and Wei, B. (2001). Adsorption of fluoride from water by amorphous
alumina supported on carbon nanotubes. Chem. Phys. Lett., 350, 412–416.
68. Zhou, Y., Yu, C. and Shan, Y. (2004). Adsorption of fluoride from aqueous solu-
tion on La 3+ impregnated cross-linked gelatin. Sep. Purif. Technol., 36, 89–94.
69. Olde Damink, L.H.H., Dijkstra, P.J., Van Luyn, M.J.A., Van Wachem, P.B.,
Nieuwenhuis, P. and Feijen, J. (1995). Glutaraldehyde as a crosslinking agent for
collagen based biomaterials. J. Mater. Sci. Mater. Med., 6, 460–472.
70. Xiaoyun, L., Kuanxiu, S., Xiuru,Y. et al. (1999). Development of deflourination
from water by rare earth compound. Chem. Ind. Eng., 16, 286–291, as cited in:
Zhou, Y., Yu, C., Shan, Y. (2004) Adsorption of fluoride from aqueous solution
on La3+ impregnated cross-linked gelatin. Sep. Purif. Technol., 36, 89–94.
71. Xiaoyun, L., Jianping, W., Kuanxiu, S. et al. (2001). Properties of resin adsorbent
loaded Ce(IV) Ion for removing fluoride ions. Ion Exchange Adsorption, 17, 131–137,
as cited in: Zhou, Y., Yu, C., Shan, Y. (2004) Adsorption of fluoride from aqueous
solution on La 3+ impregnated cross-linked gelatin. Sep. Purif. Technol., 36, 89–94.
72. Onyango, M.S., Kojima, Y., Aoyi, O., Bernardo, E.C. and Matsuda, H. (2004).
Adsorption equilibrium modeling and solution chemistry dependence of flu-
oride removal from water by trivalent-cation-exchanged zeolite F-9. J. Colloid
Interface Sci., 279, 341–350.
Defluoridation Techniques 121

73. Tripathy, S.S., Bersillon, J.L. and Gopal, K. (2006). Removal of fluoride from
drinking water by adsorption onto alum-impregnated activated alumina. Sep.
Pur. Technol., 50, 310–317.
74. Veressinina, Y., Trapido, M., Ahelik, V. and Munter, R. (2001). Fluoride in drink-
ing water: The problem and its possible solutions. Proc. Estonian Acad. Sci.
Chem., 50, 81–88.
75. Mameri, N., Yeddou, A.R., Lounici, H., Belhocine, D., Grib, H. and Bariou, B.
(1998). Defluoridation of septentrional Sahara water of North Africa by elec-
trocoagulation process using bipolar aluminum electrodes. Water Res., 32,
1604–1612.
76. Ming, L., Yi, S.R., Hua, Z.J., Yuan, B., Lei, W., Ping, L. and Fuwa, K.C. (1983).
Elimination of excess fluoride in potable water with coacervation by electroly-
sis using aluminum anode. Fluoride, 20, 54–63.
77. Zhu, J., Zhao, H. and Ni, J. (2007). Fluoride distribution in electrocoagulation
defluoridation process. Sep. Pur. Technol., 56, 184–191.
78. Lounici, H., Adour, L., Belhocine, D., Elmidaoui, A., Bariou, B. and Mameri, N.
(2001). Novel technique to regenerate activated alumina bed saturated by fluo-
ride ions. Chem. Eng. J., 81, 153–160.
79. Ulbricht, M. (2006). Advanced functional polymer membranes. Polymer, 47,
2217–2262.
80. Ndiaye, P.I., Moulln, P., Dominguez, L., Millet, J.C. and Charbit, F. (2005).
Removal of fluoride from electronic industrial effluent by RO membrane sepa-
ration. Desalination, 173, 25–32.
81. Schneiter, R.W. and Middtebrooks, E.J. (1983). Arsenic and fluoride removal
from ground water by reverse osmosis. Environ. Int., 9, 289–292.
82. Cohen, D. and Conrad, H.M. (1998). 65,000 GPD fluoride removal membrane
system in Lakeland, California, USA. Desalination, 117, 19–35.
83. Ritchie, S.M.C. and Bhattacharyya, D. (2002). Membrane-based hybrid pro-
cesses for high water recovery and selective inorganic pollutant separation. J.
Hazard. Mater., 92, 21–32.
84. Arora, M., Maheshwari, R.C., Jain, S.K. and Gupta, A. (2004). Use of membrane
technology for potable water production. Desalination, 170, 105–112.
85. Hu, K. and Dickson, J.M. (2006). Nanofiltration membrane performance on fluo-
ride removal from water, J. Membr. Sci., 279, 529–538.
86. Nicoll, H. (2001). Nanofiltration makes surface water drinkable. Filtr. Sep., 38,
22–23.
87. Bruggen, B.V. and Vandecasteele, C. (2003). Removal of pollutants from surface
water and ground water by nanofiltration: Overview of possible applications in
the drinking water industry. Environ. Pollut., 122, 435–445.
88. Hagmeyer, G. and Gimbel, R. (1998). Modelling the salt rejection of nanofiltration
membranes for ternary ion mixtures and for single salts at different pH values.
Desalination, 117, 247–256.
89. Kettunen, R. and Keskitalo, P. (2000). Combination of membrane technology
and limestone filtration to control drinking water quality. Desalination, 131,
271––283.
90. Marder, L., Sulzbach, G.O., Bernardes, A.M. and Ferreira, J.Z. (2003). Removal of
cadmium and cyanide from aqueous solutions through electrodialysis. J. Braz.
Chem. Soc., 14, 610–615.
122 Fluoride in Drinking Water

91. Adhikary, S.K., Tipnis, U.K., Harkare, W.P. and Govindan, K.P. (1989).
Defluoridation during desalination of brackish water by electrodialysis.
Desalination, 71, 301–312.
92. Amor, Z., Malki, S., Taky, M., Bariou, B., Mameri, N. and Elmidaoui, A. (1998).
Optimization of fluoride removal from brackish water by electrodialysis.
Desalination, 120, 263–271.
93. Tahaikt, M., Achary, I., Sahli, M.A.M., Amor, Z., Taky, M., Alami, A., Boughriba,
A., Hafsi, M. and Elmidaoui, A. (2006). Defluoridation of Moroccan ground
water by electrodialysis: Continuous operation. Desalination, 189, 215–220.
8
Adsorptive Removal of
Fluoride: A Case Study

8.1 Introduction
This chapter describes a case study dealing with the application of a novel
adsorbent, alumina cement granules (ALC), in removing fluoride from
groundwater. The sorption capacity of ALC in fluoride uptake was evalu-
ated by various laboratory experiments. The feasibility of its use was first
examined by continuously mixed batch reactor (CMBR) studies, commonly
referred to as batch studies. The kinetics of sorption, equilibrium sorption
capacity, and the mechanisms of fluoride removal were mainly evalu-
ated through batch performances. The field application potential of ALC
for domestic and community uses was evaluated in terms of its sorptive
responses from continuous flow, fixed bed studies, commonly referred to as
column studies.

8.2  Materials and Methods


8.2.1  Reagents and Adsorbate
All chemicals and reagents used in this study were of analytical grade.
NaF (Merck) was used for the preparation of standard fluoride stock
solution in double-distilled water. All synthetic fluoride solutions for
adsorption and analysis were prepared by an appropriate dilution of the
stock solution in de ionized (DI) water. The natural fluoride-rich drink-
ing water was collected from Baliasingh Patna, a fluoride endemic vil-
lage (Kurda district, Orissa state) in India. Only plasticware was used
for handling fluoride solution, and it is neither prepared in nor added
to glass containers. All plasticware and glassware were pre soaked in
a dilute HNO3 acid bath, washed with dilute soap solution, rinsed thor-
oughly with DI water and dried prior to use.1

123
124 Fluoride in Drinking Water

8.2.2  Synthesis of the Adsorbent


The adsorbent ALC, selected for the present research, was prepared
from commercially available high alumina cement. The high proportion
of alumina and calcium, whose potential in fluoride removal was estab-
lished, was instrumental in selection. Initially, slurry was prepared by
adding distilled water to 1 kg of high alumina cement at a water–cement
ratio of 0.3. The slurry was kept at an ambient temperature for 2 days for
setting, drying and hardening. This hardened paste was cured in water
for 5 days. After curing, it was broken, granulated, sieved to a geometric
mean size of around 0.212 mm, and kept in airtight containers for use.1,2

8.2.3 Instrumentation
The elemental composition of ALC was determined by energy-dispersive
x-ray (EDX) analysis (Oxford ISIS-300 model) by the quantitative method in
two iterations by using ZAF correction, at a system resolution of 65 eV, and
results were normalized stoichiometrically. The surface area of the adsor-
bent was determined by the Brunauer, Emmett and Teller (BET) method
at liquid nitrogen temperature by using FlowSorb II 2300 (Micrometrics
Instruments corporation, USA). The chemical composition of ALC was
determined by x-ray diffraction analysis (XRD) by using a Miniflex diffrac-
tometer (30 kV, 10 Maq; Rigaku Corp., Tokyo, Japan) with a Cu Kα source
and a scan rate of 2o/min at room temperature. The NexusTM 870 spec-
trometer (Thermo Nicolet) was used for Fourier transform infrared (FTIR)
analysis. Expandable Ion Analyzer EA 940 with Orion ion plus (96-09)
fluoride electrode (Thermo Electron Corporation, Beverly, Massachusetts),
using TISAB III buffer, was used for fluoride measurement. The pH mea-
surement was done by a Cyber Scan 510 pH meter (Oakton Instruments,
USA). A temperature-controlled orbital shaker (Remi Instruments Ltd.,
Mumbai, India) was used for agitation of the samples in batch studies.
A high-precision electrical balance (Mettler Toledo, Model AG135) was
used for the weight measurement. A flame atomic absorption spectropho-
tometer (FAAS) (Shimadzu, Model AA−6650) was used for quantitative
analysis of elements such as iron, calcium and magnesium. Conductivity
and total dissolved solids (TDS) were measured using Cyber Scan 510,
Eutech instruments, Singapore. Nephelo turbidity meter (Systronics,
Model 131) was used for the measurement of turbidity.2–4

Peristaltic pumps (Miclins, Chennai, India) were used for controlling flow
rates in column studies.5

8.2.4  Characterization of the Adsorbent


The developed adsorbent ALC was characterized by EDX, scanning electron
microscopy (SEM), XRD, FTIR, and various physicochemical analyses. EDX
analysis was used to determine the elemental composition of ALC (combined
with oxygen). SEM photographs of the adsorbent were taken to study the
Adsorptive Removal of Fluoride 125

surface texture of ALC grains. The XRD analyses were carried out to identify
the morphological structure and the extent of crystallinity of the adsorbent.
The FTIR analyses were done to understand the spectroscopic features of the
adsorbent. The bulk density was determined by pouring 5 g of ALC into a
100-mL-stoppered measuring cylinder half filled with water. This was thor-
oughly mixed by inverting the stoppered measuring cylinder several times.
The adsorbent was then allowed to settle to a constant volume. The bulk den-
sity is reported in g/cm3. The pH at zero-point charge of ALC was also deter-
mined. Different quantities of ALC were placed in 10-mL solutions of 0.1 M
NaCl (prepared in pre boiled water) in various bottles and kept in a thermostat
shaker for overnight continuous agitation. The equilibrium pH values of these
mixtures were measured, and the limiting value was reported as pHzpc.3 The
other chemical analyses were conducted as per APHA guidelines.6

8.3  Batch Studies


Batch studies were performed to investigate the sorptive characteristics of
the developed adsorbent ALC. Thus, experiments were conducted on the
adsorption and desorption of fluoride in single-component aqueous sys-
tems, unless otherwise specified. The experimental studies include kinetics,
equilibrium, thermodynamic profiles of sorption, and influence of various
parameters on the sorption of fluoride onto ALC. The batch sorption experi-
ments were conducted using polyethylene bottles (Tarson Co. Ltd., India)
of 150-mL capacity with 50 mL of fluoride solutions of a desired concentra-
tion and pH. No pH adjustments were made unless otherwise mentioned.
ALC was added as per dose requirements; bottles were capped tightly and
shaken in the orbital shaking incubator at 230 ± 10 rpm at room temperature
(300 ± 2 K), unless otherwise mentioned.

The bottles were taken out from the shaker at the desired time interval
and filtered using Whatman No-42 filter paper to separate the sorbent
and filtrate. From the filtered sample of each batch reactor, 10 mL was
taken for the analysis and determination of residual fluoride. All batch
sorption experiments were duplicated with an experimental error limit
±5%, and average values were reported. In order to check for any adsorp-
tion on the walls of the container, blank container adsorption tests were
also carried out.2,3

8.3.1  Effect of Process Parameters


Experiments to study the effects of agitation rate on fluoride removal were
conducted at a fixed dose of ALC (1.5 g/L in synthetic water and 8 g/L in
natural water) at an initial pH of 6.9 ± 0.4 and initial fluoride concentra-
tions of 8.65 mg/L. The samples were shaken at varying agitation rates of
126 Fluoride in Drinking Water

120–280 rpm for 4 h at a temperature of 300 ± 2 K. Experiments to study the


effects of adsorbent dose on the removal of fluoride were conducted by vary-
ing the dose of ALC (0.25–3.0 g/L in synthetic water and 3–12 g/L in natural
water) at an initial pH of 6.9 ± 0.4 with an initial fluoride concentration of
8.65 mg/L. The samples were shaken at an agitation rate of 230 ± 10 rpm for
3 h at a temperature of 300 ± 2 K. Experiments to study the effects of contact
time on fluoride removal were conducted at a fixed dose of ALC (2 g/L in
synthetic water and 10 g/L in natural water) at an initial pH of 6.9 ± 0.4 and
initial fluoride concentrations of 8.65 mg/L (agitation rate of 230 ± 10 at 300
± 2 K). The samples were taken at the first 2, 4, 8, 10, 30, 45, 60, 75, 90, 150,
180, 210, and 240 min and analyzed for remaining fluoride concentrations.
Experiments to study the effects of initial adsorbate concentration were con-
ducted at a fixed dose of ALC (2 g/L in synthetic water) at an initial pH of
6.9 ± 0.4. Initial fluoride concentrations of 2.5, 5, 8.65, 25, and 50 mg/L were
taken and agitated at 230 ± 10 rpm at 300 ± 2 K. The samples were taken at
the first 10, 30, 45, 60, 90, 120, 150, 180, 210, and 240 minutes and analyzed for
remaining fluoride concentrations.

8.3.2  Equilibrium Studies


The equilibrium sorption studies were conducted in natural water with
varying amounts of ALC in the range of 3–12 g/L at its natural concentra-
tion of 8.65 mg/L and pH of 6.9 ± 0.4 at 230 ± 10 rpm for 3 h (equilibrium
time). In synthetic water, equilibrium sorption studies were conducted by
both variations in doses of ALC (dose variation study) and variations in
initial fluoride concentrations (concentration variation study). In the dose
variation study, the dose range of ALC selected was 0.25–3.0 g/L at an initial
fluoride concentration of 8.65 mg/L (pH = 6.9 ± 0.4, agitation rate = 230 ± 10
rpm, equilibrium agitation time = 3 h). In concentration variation studies
in synthetic water, the range of initial fluoride concentration selected was
2.5–100 mg/L at a fixed ALC dose of 2 g/L (pH = 6.9 ± 0.4, agitation rate =
230 ± 10 rpm, equilibrium agitation time = 3 h). The effect of initial pH on
the removal of fluoride in synthetic water was studied over a lower ALC
dose of 1.5 g/L for the initial fluoride concentration of 8.65 mg/L (agitated
at 230 ± 10 rpm for 3 h at 300 ± 2 K). Then, 2M HCl and 2M NaOH solutions
were used to adjust the pH of the synthetic solution. Sorption experiments
were conducted at 290, 300, and 310 K to study the effects of temperature on
the removal of fluoride. The experiments were conducted on both synthetic
and natural water at a fixed ALC dose (2 g/L in synthetic water and 10 g/L
in natural water) and an initial fluoride concentration (8.65 mg/L). The agi-
tation time and rate provided were 3 h and 230 ± 10 rpm, respectively, at all
temperatures of the study. To study the effects of ionic strength or inert elec-
trolyte concentrations on the sorption of fluoride onto ALC, sodium nitrate
solutions of strength ranging from 10−4 to 10−1 M were used in synthetic
water. The dose of ALC was 1.5 g/L for the initial fluoride concentration of
Adsorptive Removal of Fluoride 127

8.65 mg/L (agitation rate = 230 ± 10 rpm, equilibrium agitation time = 3 h,


temperature = 300 ± 2 K). Batch equilibrium experiments were conducted
in synthetic water to elucidate the individual effects of some common ions
such as nitrate, chloride, sulfate, bicarbonate, silicate, calcium, iron, and
phosphate on the sorption of fluoride onto ALC. The effects of humic acid
on fluoride sorption were also investigated. The interference effects of other
ions were conducted in synthetic samples having fluoride concentrations of
8.65 mg/L at an ALC dose of 1.5 g/L (agitation rate = 230 ± 10 rpm, equilib-
rium agitation time = 3 h, temperature = 300 ± 2 K). The salts NaNO3, KCl,
Na2SO4, NaHCO3, Na2SiO3, Ca(NO3)2, Fe(NO3)2, and K2HPO4 (Merck) were
used for preparing the respective ions for interference studies. Humic acid
solutions were prepared from humic acid sodium salt. Batch desorption
studies were carried out by the dilution method7 using the spent ALC that
had been previously used for fluoride sorption. After completing adsorp-
tion up to the equilibrium time of 3 h, 20 mL of the supernatant in the bot-
tle was replaced by an equal volume of 10% NaOH. The batch desorption
experiments were conducted as per the same experimental procedure fol-
lowed in sorption studies.

8.3.3  Column Studies


Continuous flow, fixed bed column experiments were conducted using glass
columns of a length of 550 mm and an internal diameter of 20 mm. The col-
umn was packed with a desired depth of ALC between two layers of glass
wool at the top and bottom ends to prevent the absorbent from floating.
The column was fed continuously with the feed solution (natural ground-
water/synthetic fluoride water) having the desired concentration at the
desired volumetric flow rate by using peristaltic pumps (Miclins, India) in
the up-flow mode. The schematic diagram of the column setup is shown in
Figure 8.1. The effluent samples were collected at pre determined time inter-
vals and analyzed for the remaining fluoride concentration. The breakpoint
concentration of fluoride was taken as 1.00 mg/L according to the permis-
sible limit in drinking water set by the World Health Organization (WHO)8,9
standards. The exhaust concentration of fluoride was considered 90% of the
influent (i.e., 0.9C0). All studies were performed at a constant temperature of
300 K to be representative of the prevailing environmental conditions. No
pH adjustments were made in natural water. The pH of synthetic water was
maintained similar to that of natural water. The effects of various process
parameters such as bed depth and flow rate on the sorption profile of ALC
were investigated in both natural and synthetic waters. The effects of initial
fluoride concentrations in synthetic water were also investigated. The effects
of bed depth on the sorption profile of ALC in both synthetic and natural
water were investigated by varying the bed depths (5–15 cm) while keeping
the influent feed (fluoride) concentration (8.65 mg/L) and influent flow rate
(8 mL/min) fixed.
128 Fluoride in Drinking Water

1: Fluoride feed storage


2: Delivery line 7
3: Peristaltic pump
4: Inlet end
5: Glass wool 5
6: ALC bed
7: Effluent outlet 8
8: Treated water storgae 6

2 4

FIGURE 8.1
Experimental setup for fixed bed column studies. (Modified from Ayoob, S., Gupta, A.K. and
Bhakat, P.B., Sep. Purif. Technol., 52, 430–438, 2007.)

The effects of feed flow rates on the sorption profile of ALC in both synthetic
and natural water were investigated by varying the flow rates (4–12 mL/min)
while keeping the influent feed (fluoride) concentration (8.65 mg/L) and bed
depth (10 cm) fixed. The effects of initial sorbate concentrations on the sorption
profile of ALC in synthetic water were investigated by varying the initial sor-
bate (fluoride) concentrations (4–15 mg/L) while keeping the flow rate (8 mL/
min) and bed depth (10 cm) fixed. The exhausted ALC fixed beds after lengthy
column runs (bed depth = 5 cm; flow rate = 8 mL/min; initial fluoride concen-
tration = 8.65 mg/L) of both natural and synthetic waters were regenerated
using sodium hydroxide (10%, w/v) solution. The desorption studies in col-
umns were conducted in the same manner as sorption studies, replacing the
feed solution with 10% NaOH solutions at a very low flow rate of 0.5 mL/min.

8.4  Theoretical and Mathematical Formulations


8.4.1  Adsorption Capacity
The amount of fluoride adsorbed per unit mass of ALC at any time t (qt,
mg/g) and the adsorption efficiency (%R, determined as the fluoride removal
percentage relative to the initial concentration) of the system were calculated
as follows:3

C0 − Ct
qt = V (8.1)
m
Adsorptive Removal of Fluoride 129

C0 − Ct
%R = ×100
C0 (8.2)

where C0 and Ct are the fluoride concentrations in solution (mg/L) initially


and at any time (t), respectively, m is the mass of ALC (g), and V is the volume
in liters of the solution. In Equation 8.1, when Ct = Ce (fluoride concentrations
remaining in solution at equilibrium in mg/L), qt = qe (equilibrium adsorp-
tion capacity in mg/g).

8.4.2  Kinetic Modeling


Prediction of the rate at which adsorption takes place for a given system
is probably the most important factor in adsorption system design, as the
adsorbate residence time and reactor dimensions are controlled by the kinet-
ics of the system. A number of adsorption processes for pollutants have been
studied in an attempt to find a suitable explanation for the mechanisms and
kinetics for sorting out environment solutions.11 Based on kinetic data, var-
ious models have been suggested that throw light on the mechanisms of
adsorption and potential rate-controlling steps such as mass transport and
chemical reaction processes.12 Kinetic models based on the capacity of the
adsorbents mainly include the Lagergren’s pseudo-first-order equation and
pseudo-second-order equation models.11,13

8.4.2.1  Pseudo-First-Order Model


In 1898, Lagergren presented the first-order rate equation for the adsorption
of solutes from a liquid solution onto charcoal.13 Lagergren’s equation was
the first kinetic equation for the sorption of the liquid/solid system based
on solid capacity, and it was used extensively to describe the sorption kinet-
ics. This model assumes that the rate of change of solute uptake with time
is directly proportional to the difference in saturation concentration and the
amount of solid uptake with time. In order to distinguish a kinetic equation
based on the adsorption capacity of a solid from one based on the concen-
tration of a solution, Lagergren’s first-order rate equation has been called a
pseudo-first-order equation and is expressed as follows:

dqt
= ks 1 (qe − qt ) (8.3)
dt

Integration within the boundary conditions t = 0 to t = t and qt = 0 to qt = qe


gives the linearized form as follows:

ln(qe − qt ) = ln qe − ks 1t (8.4)
130 Fluoride in Drinking Water

where qe is the amount of soluted sorbate sorbed at equilibrium (mg/g), qt


is the amount of soluted sorbate on the surface of the sorbent at any time
t (mg/g), and ks1 is the pseudo-first-order rate constant (min–1). The plot of
ln(qe−qt) versus t gives a straight line for first-order kinetics, which allows
computation of the adsorption rate constant, ks1. When adsorption is pre-
ceded by diffusion through a boundary, the kinetics in most cases follows
Lagergren’s pseudo-first-order rate equation. If the experimental results do
not follow Equations 8.3 and 8.4, the sorption is not diffusion controlled and
the data differ in two important aspects: (1) ks1(qe−qt) does not represent the
number of available sites and (2) ln qe is not equal to the intercept of the plot
of ln(qe−qt) against t.12,14,15 Then, the reaction is not likely to be first order, irre-
spective of the magnitude of the correlation coefficient.

8.4.2.2  Pseudo-Second-Order Model


If the rate of sorption is a second-order mechanism, the pseudo-second-order
chemisorption kinetic rate equation is expressed as follows:12

dqt
= k(qe − qt )2 (8.5)
dt
Rearranging and integration within the boundary conditions t = 0 to t = t
and qt = 0 to qt = qe give the linearized form as follows:

1 1
= + kt (8.6)
qe − qt qe
which is the integrated rate law for a pseudo-second-order reaction.
Rearranging it again, Equation 8.6 is reduced to

t 1 1
= + t (8.7)
qt h qe
where k is the pseudo-second-order rate constant (g/mg/min), and h is the
initial sorption rate (mg/g/min), which is given by

h = kqe 2 (8.8)

The rate of a reaction is defined as the change in the concentration of a


reactant or product per unit time. Concentrations of products do not appear
in the rate law, because the reaction rate is studied under conditions where
the reverse reactions do not contribute to the overall rate. The reaction order
and rate constant must be determined by experiments. In order to distin-
guish the kinetics equation based on the concentration of a solution from
the adsorption capacity of solids, this second-order-rate equation is called
a pseudo-second-order one. The pseudo-second-order rate expression was
Adsorptive Removal of Fluoride 131

used to describe chemisorption involving valency forces through the shar-


ing or exchange of electrons between the adsorbent and adsorbate as cova-
lent forces, and ion exchange. The pseudo-second-order model constants
can be determined experimentally by plotting t/qt against t (Equation 8.7).
Although there are many factors that influence the adsorption capacity,
including initial adsorbate concentration, reaction temperature, solution pH,
adsorbent particle size and dose, and nature of the solute, a kinetic model
is concerned only with the effect of observable parameters on the overall
rate.11 It was also suggested that the pseudo-second-order equation may be
applied to chemisorption processes with a high degree of correlation in sev-
eral literature cases where a pseudo-first-order rate mechanism has been
arbitrarily assumed.12 Recently, the pseudo-second-order rate equations have
been widely applied to the adsorption of pollutants from aqueous solution,
since the model has the following advantages: the adsorption capacity, the
rate constant, and the initial adsorption rate, all of which can be determined
from the equation without knowing any parameter beforehand.11

8.4.2.3  Intra Particle Diffusion Model


The concentration dependence of the rate of sorption is commonly used as
a partial test of hypothesis regarding the nature of the rate-controlling step.
Accordingly, most of the recent studies involving adsorption of fluoride on
metal oxides explored the use of the intra particle surface diffusion model16,17
represented by Equation 8.9 to elucidate its mechanism15,18 as follows:

qt = k p t 1/2 + C (8.9)

where kp is the intra particle diffusion rate constant (mg/g/h1/2). The value C
(mg/g) in this equation is a constant that gives an idea about the thickness of
the boundary layer (the larger the value, the greater is the boundary effect).
When the plot (qt vs. t1/2) does not pass through the origin, it is indicative
of some degree of boundary layer control. This behavior indicates that the
intra  particle diffusion is not the only rate-limiting step, but other kinetic
factors also may control the rate of adsorption, all of which may be operating
simultaneously. The slope of linear plot can be used to derive values for the
rate parameter, kp, for the intra particle diffusion. However, if the data exhibit
multi linear plots, then two or more steps influence the sorption process.

8.4.2.4  Elovich Equation


The Elovich equation has also been successfully applied in aqueous systems
to describe the adsorption and desorption reactions19,20 as follows:
1
qt = ln (1+ αβt) (8.10)
β
132 Fluoride in Drinking Water

where α and β are constants, t is the time, and qt is the surface coverage. The
Elovich equation can be derived from either a diffusion-controlled process or
a reaction-controlled process. If the Elovich equation is based on adsorption
on an energetically heterogeneous surface, the parameter β is related to the
distribution of activation energies. In the diffusion control model, it is a func-
tion of particle and diffusion coefficients. When the term αβt is much greater
than 1, the equation just cited can be simplified to19

1 1 1
qt = ln (αβt) = ln (αβ)+ ln (t) (8.11)
β β β
If the results follow an Elovich equation, the kinetic results will be linear
on a qt versus ln t plot. It was suggested that the diffusion accounted for the
Elovich kinetics pattern21; that conformation to this equation alone might be
taken as evidence that the rate-determining step is diffusion in nature; and
that this equation should apply to conditions where the desorption rate can
be neglected.22

8.4.2.5  Arrhenius Equation


It was suggested that the value of energy of the activation value (Ea),
obtained from Arrhenius equation (Equation 8.12), could also be a useful
kinetic parameter in assessing rate-limiting steps.23 Low Ea values usually
indicate diffusion-controlled transport and physical adsorption processes,
whereas higher Ea values indicate chemical reactions or surface-controlled
processes:

⎛ E ⎞
K = Af exp ⎜− a ⎟ (8.12)
⎝ RT ⎠

where K is the rate coefficient of the sorption reaction (g/mg/min), Af is the


pre-exponential factor or frequency factor (g/mg/min), Ea is the activation
energy of sorption (kJ/mol), R is the gas constant (8.314 J/mol/K), and T is
the thermodynamic temperature (K). Some of the assigned values of Ea (kJ/
mol) include 8–25 to physical adsorption, less than 21 to aqueous diffusion,
20–40 to pore diffusion, and greater than 84 to ion exchange.23 In this study,
the kinetic data of fluoride sorption onto ALC were fitted to all the kinetic
models cited earlier. The activation energy of the sorption process in both
systems was also found out. The experimental investigations as suggested
next were carried out to identify the rate-limiting step.

8.4.3  Elucidation of Rate-Limiting Step


The three important parameters that may determine the kinetics and
rate-limiting process of a sorption system are being identified as pH, con-
centration of inert electrolyte and desorption pattern of the adsorbent.3,19
Adsorptive Removal of Fluoride 133

In solutions, the effective particle sizes are influenced by the con-


centration of inert electrolyte, with larger aggregated particles being
formed in the higher concentrations of inert electrolytes at a given pH.
Increasing the effective particle size will increase the diffusion path
length from the external surface of the aggregate to the reactive sites
located on the internal surface of the adsorbent. As a result, it will take
a longer time for the ions to reach these reactive sites from an external
solution, suggesting the rate-limiting step. It was suggested that the slow
stage sorption, indicative of the rate-limiting step, may be due to two
reasons: (1) diffusion or (2) surface reactions.3,19

As defined earlier, the diffusion may be either inter particle or intra par-
ticle. The surface reactions include surface precipitation and surface site
bonding energy heterogeneity or other surface reactions.

The following procedure was adopted to elucidate the rate-limiting step:

1. If the inter particle diffusion is rate limiting, the


adsorption patterns should be sensitive to inert electrolyte
concentrations or pH.
2. If the intra particle diffusion is rate limiting, the desorption
pattern should follow the same two-stage pattern as that of
sorption.
3. If both inter particle and intra particle diffusion are not rate
controlling, by elimination, the surface reactions will be rate
limiting.3,19

So, in this study, different kinetic models were examined to describe the
sorption kinetics; further, the response of the adsorbent to pH and inert
electrolyte concentration were also examined to elucidate the rate-limiting
mechanism.

8.4.4  Adsorption Equilibrium and Isotherms


In adsorption, in addition to the surface unsaturation of the solid phase, the
intermolecular interactions in the liquid phase also play a dominant role. As
a result, in general, adsorption is viewed as a complex phenomenon. During
the process of adsorption, the molecules of the solute are removed from the
solution and transferred onto the adsorbent. This transfer of the solute from the
solution to the adsorbent continues until the concentration of the solute remain-
ing in solution is in equilibrium with the concentration of the solute adsorbed
by the adsorbent. This state is in a dynamic stability, that is, the amount of
solute migrating onto the adsorbent is counterbalanced by the amount of solute
migrating back into the liquid phase. The position of equilibrium depends on
various parameters such as solute concentration, characteristics of the adsor-
bent, solvent temperature, and pH. The adsorption isotherms are used for
understanding and quantifying this distribution of the adsorbate between the
134 Fluoride in Drinking Water

liquid phase and the solid adsorbent phase that are at equilibrium during the
adsorption process and also for understanding the mechanism of adsorption.
Fitting of experimental data to adsorption isotherm models is an important
step in designing and optimizing an adsorption system.4
The adsorption isotherm function is an important experimental parameter
that measures adsorption as a function of equilibrium solute concentration
(C2), at a given temperature (T). It is expressed quantitatively in terms of the
moles of solute species adsorbed per gram of the adsorbent (n2s) and can be
written as follows:24

n2 s = f (C2 , T ) (8.13)

At constant temperature,

n2 s = fT (C2 ) (8.14)

Various functional forms for f have been proposed either by empirical


observations or in terms of specific models. The Langmuir, Freundlich, and
Dubinin-Radushkevich (D-R) isotherm models are particularly important
examples.25–28

8.4.4.1  Langmuir Isotherm


Though originally derived for the solid–gas interface, the general kinetic fea-
tures of the Langmuir model (Equation 8.15) are equally applicable for any
interface:

qmax bCe
qe =
1+ bCe (8.15)

where qe is the amount of adsorbate adsorbed at equilibrium per unit weight


of adsorbent (mg/g), Ce is the equilibrium solute concentration (mg/L), qmax is
the saturated monolayer adsorption capacity (mg/g), and b is the Langmuir
constant related to the binding energy or affinity parameter (L/mg) of the
sorption system, respectively. The corresponding linear form of the equation
is as follows:

1 1 1
= + (8.16)
qe bqmaxCe qmax

This model assumes a uniform surface with equivalent adsorption sites29


with no lateral interactions between adsorbed species. The adsorption is lim-
ited to a monolayer24; adsorption energy is assumed to be independent of the
occupancy of neighbouring sites and, hence, has a constant value. There is
also no transmigration of the adsorbate in the plane of the surface.
Adsorptive Removal of Fluoride 135

The Langmuir equation includes two constants, each of which has a clear
physical meaning: ‘‘b’’ is the equilibrium constant for the adsorption
process expressed in terms of the ratio of the adsorption and desorp-
tion rate constants and, hence, is directly related to the binding energy
(bαe-ΔH/RT, where ΔH is the net enthalpy change).4 qmax is the adsorption
limit obtained at high solute concentrations when bCe >> 1, and qe shows
a zero-order dependence on the solute concentration. For very low val-
ues of Ce, the term bCe << 1 reduces the hyperbolic equation to a first-
order linear equation in solute concentration as follows4:

qe = k Ce (8.17)

where

k = qmax b (8.18)

The affinity parameter (b) is best estimated from the slope of the adsorp-
tion isotherm at very low concentrations. However, this slope gives the
product qmax b and not just b. In order to separate these two parameters,
it is necessary to know the adsorption maximum qmax, and this can only
be estimated with precision from data at very high concentrations where
the slope of the isotherm approaches zero. If data are restricted to an
intermediate range of concentration, they may be fitted very well, but it
will be difficult to separate qmax and b, the values of which will show a
high negative correlation and correspondingly, high standard errors.4,30

8.4.4.2  Freundlich Isotherm


Similar to the Langmuir model, the Freundlich model (Equation 8.19) also
has an empirical origin, but it is extremely useful for experimentally deter-
mining the adsorption capacity (kf):

qe = kf Ce1/n (8.19)

where n is a constant representing adsorption intensity and is always greater


than unity. The corresponding linear form of the equation is as follows:

1
ln qe = ln kf + ln Ce (8.20)
n
The Freundlich equation is a special case for heterogeneous surface
energies in which the binding energy term b, in the Langmuir equa-
tion, varies as a function of the surface coverage qe, essentially due to
variations in heats of adsorption.24 This model agrees quite well with
the Langmuir isotherm at moderate concentrations. However, unlike the
Langmuir equation, it is not reduced to a linear adsorption expression at
low solute concentrations but remains convex to the concentration axis.
It also does not agree with the Langmuir equation at very high solute
136 Fluoride in Drinking Water

concentrations, since n must reach a finite limiting value when the sur-
face is fully covered.4

The intercept and slope of its linear plot gives a measure of the adsorp-
tion capacity (kf), and an inverse measure of the intensity of adsorption (n),
respectively.

8.4.4.3  Dubinin–Radushkevich (D–R) Isotherm


The D–R isotherm (Equation 8.21) is more general than the Langmuir iso-
therm, since it does not assume a homogeneous surface or constant sorption
potential. It was applied to distinguish between the physical and chemical
adsorption of metal ions and has limited utility28:

qe = Qm exp(−kadε 2 ) (8.21)

where kad is a constant related to adsorption energy, Qm is the theoretical


saturation capacity (mg/g), and ε is the Polanyi potential and is given by

⎛ 1 ⎞
= RT ln ⎜1+ ⎟ (8.22)
⎝ Ce ⎠
where R is the universal gas constant (8.314 × 10-3 kJ/mol/K) and T is the
absolute temperature (K). The corresponding linear form of Equation 8.21 is
as follows:

ln qe = lnQm − kadε 2 (8.23)

The nature of interaction between the adsorbate and adsorbent binding sites
can be evaluated by the mean free energy of sorption per mole of the adsor-
bate calculated by the following equation31:

E = −(2k)−0.5 (8.24)

where k is the constant obtained from the D–R isotherm. This free energy
term is the work done while transferring one mole of the adsorbate to the
surface from infinity in solution. If the magnitude of E is between 8 and 16
kJ/mole, the adsorption process proceeds by the exchange mechanism (ion
exchange); if it is less than 8 kJ/mole, physisorption occurs.32

The Langmuir and Freundlich isotherm models are said to suffer from
two major drawbacks. First, the model parameters obtained are usually
appropriate for a particular set of conditions and these cannot be used
as a prediction model for another set. Second, these models are unable to
provide a fundamental understanding of ion adsorption.4,33
Adsorptive Removal of Fluoride 137

Although many other isotherm models have been developed,34,35 it can be


seen that the Langmuir and Freundlich isotherms still remain the two most
commonly used equilibrium adsorption equations, due to their simplicity
and the ease with which their adjustable parameters can be estimated. This
trend generally holds for defluoridation research in which the Langmuir and
Freundlich isotherm models have found wide applicability in determining
the adsorption capacities of various adsorbents.

8.4.4.4  Selection of Best-Fitting Isotherm


The selection of the best-fit isotherm model is a significant step in adsorption
studies. Generally, the success of the adsorption process hinges on the devel-
opment and application of suitable adsorbents. The criteria for selection of an
adsorbent mainly include its adsorption capacity (mg of fluoride adsorbed/g
of adsorbent) and the cost. The adsorption capacity parameter (qmax, kf, and
Qm by Langmuir, Freundlich, and D–R) can be obtained from the respective
isotherm models.

8.4.4.5  Natural and Synthetic Systems


Generally, the adsorption capacity of an adsorbent is assessed on the
basis of equilibrium sorption data that are generated by batch studies
through the best-fitting of isotherm models. So, the fitting of experimen-
tal data to the isotherm models turns out to be the most important,
as it dictates the optimum model parameters. However, most of the
reported isotherm studies are conducted on concentration variations in
synthetic samples (DI or distilled water laced with fluoride), which lose
the real-life flavors and characteristics of natural samples. In general,
apart from concentrations of adsorbate, the pH of the medium and the
presence of other competing ions drastically alter the adsorption capac-
ity of the adsorbents in aqueous solutions.4,36 So, it is rational to sug-
gest that the adsorption capacity reported in synthetic samples cannot
be a reliable representation of its actual scavenging potential in field
applications.4

This observation becomes significant in the context of the reported reduction


in the adsorption capacities of activated alumina in removing fluoride in the
field studies. Though high fluoride removal capacity of activated alumina
(4–15 mg/g) is reported in literature, field experiences demonstrate that it
is often around 1 mg/g only.37 So, in this study, the adsorption capacity of
ALC was tested in natural fluoride-rich groundwater. The synthetic water
of the same concentration as the natural water (8.65 mg/L) was prepared
to have a reliable comparison of the sorption profile with natural water.
Further, the effects of various ions (naturally expected in groundwater), pH,
ionic strength, and temperature on the sorption behavior of ALC were also
investigated.
138 Fluoride in Drinking Water

8.4.4.6  Concentration and Dose Variation Studies


In defluoridation research involving adsorption, it has also become a common
practice to compare the adsorption capacities of various adsorbents to high-
light their effectiveness and relative trustworthiness. However, of late,

the limited success of adsorbents in field applications raises apprehen-


sions over the use of adsorption capacity (generated from equilibrium
data) as a measure of their effectiveness in drinking water treatment.
The batch studies for equilibrium data can be conducted by changing
either the concentrations of adsorbate (fluoride) or the dose of adsor-
bent (ALC). Generally, the adsorptive capacity of the adsorbent will
increase with an increase in influent concentrations, till it reaches a
maximum. Obviously, the isotherm studies performed for a higher
range of fluoride concentrations will show a higher capacity than that
at moderate or lower ranges. However, this maximum capacity may not
describe the media behavior at typical influent fluoride concentrations.
It is reported that the maximum natural fluoride concentrations in the
groundwaters of India are in the range of 0.5–48 mg/L. However, it is
not uncommon to encounter isotherm studies performed at extremely
higher fluoride concentrations. In isotherm studies of natural samples,
this approach (of concentration variation studies) becomes inappropri-
ate, as the concentration of fluoride in natural groundwater remains
constant. So, in defluoridation studies dealing with natural groundwa-
ter, it is only possible to have isotherm studies with dose variations of
adsorbents. This further (in general) suggests the irrelevance of con-
centration variation studies in the adsorption process for removing
pollutants such as fluoride from aqueous systems. So, in this study, in
addition to concentration variation studies, isotherm studies were also
conducted by dose variations of ALC to have a reliable field defluori-
dating capacity.4

8.4.5  Factors Influencing Adsorption


Adsorption is a complex process that must be carefully analyzed in order
to ensure optimum process design. Many factors influence adsorption reac-
tions and the rate and extent to which adsorption occurs.

8.4.5.1  Adsorbent Dose


Since each adsorbent particle has to purify a certain volume of liquid, at a
constant solute concentration, higher adsorbent dosages will be more effi-
cient in removing the adsorbate from solution. This is due to the greater
availability of sorption sites at higher doses of the adsorbent. So, to estimate
the optimum adsorbent dose for a particular system, the sample solution
with a fixed adsorbate concentration should be made to come into contact
with varying adsorbent doses till equilibrium is reached.
Adsorptive Removal of Fluoride 139

8.4.5.2  Contact Time


Adequate contact time between the adsorbent and the solution containing
the adsorbate is essential for the adsorbent to approach equilibrium with
the adsorbate. Each adsorbent particle has to purify a certain volume of liq-
uid, so that the higher adsorbent dosages with less volume to treat per unit
weight may reach equilibrium somewhat faster than the low dosages. To
better estimate the optimum contact time for a particular system, a single
adsorbent dosage should be made to come into contact with the solution for
a certain period. Measurement of the concentration change over time in this
system will show the effect of contact time.

8.4.5.3  Agitation Rate


There are essentially three consecutive mass transport steps associated with
the adsorption of solutes from the solution by porous adsorbents, namely, bulk
diffusion, film diffusion, and pore diffusion. The rate of adsorption is con-
trolled by either film diffusion or pore diffusion, depending on the amount
of agitation in the system. If relatively little agitation occurs between the fluid
and the adsorbent, the surface film of the liquid around the particle will be
thick and film diffusion will likely be the rate-limiting step. If adequate mix-
ing is provided, the rate of film diffusion will increase to the point that pore
diffusion may become the rate-limiting step. Pore diffusion is generally rate
limiting for batch systems, as it provides a high degree of agitation.34

8.4.5.4  Effect of pH and Coexisting Ions


Generally, the important factors influencing the adsorption process and capac-
ity of the adsorbent include pH, interference effects from other counter-ions in
water, initial adsorbate concentration, and temperature. It was observed that
many of the adsorbents developed for defluoridation have shown a reduc-
tion in their capacity while dealing with natural groundwater38,39 compared
with their laboratory performance in synthetic water; however, reasons for the
same were not fully elucidated. Since nearly all natural waters contain traces
of many chemical elements, these competing adsorbate compounds influence
the solution nature of the sorption system. Only very few adsorbents exhibit
controllable selectivity for specific compounds and so all the adsorbable com-
pounds may compete for adsorption sites. This effect turns significant, as it can
have dramatic consequences on the capacity of the adsorbent. So, in this study,
the reasons for this behavior were also investigated. The pH of the system
derives significance, as it controls the electrostatic interactions within the sys-
tem, thereby affecting the adsorption capacity and removal rate. Experimental
evidence suggests that the removal of fluoride by activated alumina (the popu-
lar adsorbent used for defluoridation) is highly sensitive to pH. In a study, the
maximum adsorption capacity of the adsorbent, alum-impregnated activated
alumina, was reported to be reduced by almost 10 times when the pH was
140 Fluoride in Drinking Water

increased from 4.0 to 9.0.40 The reported maximum adsorption capacities and
removal rates were usually in the acidic pH range of 5–6. However, the reported
pHs of fluoride-rich groundwaters were generally in the alkaline range.41 So,
it can be rationally expected that those adsorbents having optimum pH in the
acidic range will not be at their best in treating natural fluoride-rich waters.
Thus, the effect of pH turns extremely significant in fluoride removal.

Although it may be easy to adjust the pH for maximum removal in labo-


ratory studies and waterworks, it is necessary to depend on the actual pH
of raw water in domestic and small community treatments. So, for design
and practical applications, it becomes necessary2 for sorption profiles of
the adsorbent to be established under prevailing field conditions. Since
a reduction in adsorption capacity was generally expected (as demon-
strated in the limited studies cited earlier) in treating natural groundwa-
ter, this study attempts to elucidate the factors responsible for the same.2

8.4.5.5 Temperature
The temperature of the adsorption process will affect both the rate and the
extent of adsorption. The temperature of a solution has two major effects on
adsorption. First, the rate of adsorption is usually increased at higher tem-
peratures. This is due primarily to the increased rate of diffusion of adsor-
bate molecules through the solution to the adsorbent. Further, since solubility
and adsorption are inversely related, the effect of temperature on solubility
will naturally affect the extent of adsorption (or capacity of the adsorbent)
onto a particular adsorbate. Hence, the temperature effects should be stud-
ied carefully and evaluated for possible effects.
The temperature dependence of equilibrium capacity for adsorption can
be defined by the thermodynamic parameters enthalpy (ΔH°), entropy (ΔS°),
and Gibbs free energy (ΔG°). These parameters are useful tools for delineat-
ing the nature of adsorption mechanisms. The change in the heat content of
a system in which adsorption occurs, the total amount of heat evolved in the
adsorption of a defined quantity of adsorbate on an adsorbent, is termed the
heat of adsorption (ΔH°).29 Standard Gibbs free energy (ΔG°), standard enthalpy
(ΔH°), and entropy (ΔS°) changes for the adsorption process can be calculated
from Equations 8.25 and 8.26:

ΔG 0 = −RT ln K (8.25)

ΔS0 ⎛ ΔH 0 ⎞ ⎛ 1 ⎞
ln K = −⎜ ⎟⎜ ⎟
R ⎝ R ⎠ ⎝ T ⎠ (8.26)

where T is the temperature (K), and R is the universal gas constant. When
any spontaneous process occurs, there is a decrease in the Gibbs free energy.
For significant adsorption to occur, the free energy changes of adsorption,
Adsorptive Removal of Fluoride 141

ΔG°, must be negative. Further, there must also be a decrease in entropy,


because the molecules lose at least one degree of freedom when adsorbed.42

8.4.5.6  Ionic Strength


Apart from pH, another important parameter in adsorption is the ionic
strength. It is often stated in literature that an increase in ionic strength sup-
presses sorbate uptake as a result of the screening of electrostatic charge.43
This fact is relevant to the way that the metal is electrostatically or covalently
bound. Alternatively, other sorbable ions can compete with the cations of
interest for binding with the adsorbent, thereby affecting adsorption. When
direct evidence from microscopic data is absent, studying the influence of
ionic strength is a simple approach to distinguish between the inner-sphere
and outer-sphere surface complexes. If the adsorption is not affected by the
variations of the ionic strength, an inner-sphere surface complexation should
form; whereas if the adsorption is reduced with an increase in ionic strength
(i.e., due to the competitive adsorption with counter-anions), an outer-sphere
surface complexation is more likely.44 In the absence of microscopic techniques
(such as extended x-ray absorption fine structure [EXAFS]), investigation into
the effect of ionic strength has been carried out by macroscopic studies.

8.4.6  Behavior of Adsorption Columns


When the solution containing the solute (i.e., fluoride) is introduced at the bot-
tom of a packed bed containing the adsorbent media (ALC), most of the solute
removal initially occurs in a narrow band at the bottom of the column, which is
referred to as the adsorption zone.10 As the column operation continues, the lower
layers of the adsorbent bed become saturated with the solute and the adsorption
zone progresses upward through the bed. Eventually, as the adsorption zone
reaches the top of the column, the solute concentration in the effluent begins to
increase. The loading behavior of the solute to be removed from the solution in
a fixed bed containing the adsorbent media is shown by breakthrough curves
and is usually expressed in terms of a normalized concentration that is defined
as the ratio of effluent solute concentration to inlet solute concentration (Ct/C0)
as a function of time (t) or bed volumes for a given bed height (Z).
The breakthrough curve, developed from the column studies, basically
portrays the dynamic sorptive responses of the adsorbent. The point at
which the effluent fluoride concentration reaches 1 mg/L is taken as break-
through point and that corresponding to 90% of influent concentration (Ct/C0 =
0.90) is considered point of exhaust. The capacity of the column up to the point
of breakthrough will represent the breakthrough capacity, which is indicative
of the column adsorption capacity in a single-column operation. In series
column operations, the effluent from the last column represents the effluent
of the desired quality. So, all columns except the last one can be run up to the
exhaustion point, as these columns still contain unused adsorbents. In such
142 Fluoride in Drinking Water

cases, breakthrough capacity serves as the minimum capacity of the adsorbent


in the column (qmin,col), and the capacity up to the exhaust will be its total or
maximum capacity (qcol). The value of qmin,col turns useful in the design of
domestic defluoridation units (DDUs), which usually involve a single-cham-
ber use; however, for field-scale community applications, the involvement of
a number of columns qcol will be appropriate.
The total quantity of fluoride adsorbed (Ftot) in the column for a given feed
concentration (C0) and flow rate (Q, l/h) can be found by calculating the area
above the breakthrough curve by integrating the adsorbed fluoride concen-
tration (Cad = C0 –Ct) versus time t (h) plot as follows2:
t=et
Ftot = Q ∫ Cad dt (8.27)
t=0

Similarly, the quantity of fluoride adsorbed up to the breakthrough (Fb) is as


follows:
t=bt
Fb = Q ∫ Cad dt (8.28)
t=0

The minimum and maximum adsorption capacity of ALC in the sorptive


filtration system can be calculated as follows:
Fb
qmin,col = (8.29)
M
Ftot
qcol = (8.30)
M

where M is the mass of the adsorbent in the column.


The prediction of column breakthrough or the shape of the adsorption wave
front, which determines the operation life span of a bed and regeneration
times, is the most important criterion in the successful design of fixed bed
adsorption systems. However, it is innately difficult to develop a model that
accurately describes the dynamic behavior of adsorption in a fixed bed sys-
tem. The process does not operate in a steady state, as the concentration of the
adsorbate changes as the feed moves through the bed. The fundamental trans-
port equations for a fixed bed are those of material balance between the solid
and the fluid. The equation of material mass balance can be stated as follows:
input flow = output flow + flow inside pore + matter adsorbed onto the bed.45
The material mass balance equation for this system can be expressed
mathematically as follows:

dC dq
QCo = QCt +Vp + m (8.31)
dt dt

where (QC0) is the inlet flow of solute to the column (mg/min); (QCt) is the
outlet flow of solute leaving the column (mg/min); Vp is the porous volume
Adsorptive Removal of Fluoride 143

(l) (Vp = V /(1 − ε), where V is the bulk volume (l) and ε is the void fraction in
the bed; Vp (dC / dt) is the flow rate through the column bed depth (mg/min);
and m(dq / dt) is the amount of solute adsorbed onto the sorbent media (mg/
min), where m is the mass of the adsorbent (g) and (dq / dt) is the adsorption
rate (mg/g/min).45
From the relationship just cited (Equation 8.31), it is evident that the lin-
ear flow rate (u = Q / A, where A is the column section area), the initial sol-
ute concentration, and the adsorption potential are the determining factors
of the balance for a given column bed depth. Therefore, it is necessary to
examine these parameters and to estimate their influence in order to opti-
mize the fixed bed column adsorption process. However, these equations
that are derived to model the fixed bed adsorption system with theoretical
vigor are differential in nature and usually require complex numerical meth-
ods to be solved.45 Because of this, various simple numerical models have
been developed to predict the dynamic behavior of the columns and some of
these models have been discussed in this study. The prediction and analy-
sis of the dynamic behavior of the column was carried out with Hutchin’s
bed depth service time (BDST) model, Thomas model, Yoon–Nelson model,
Clark model, Wolborska model, and Bohart and Adams model.

8.4.7  Analysis and Modeling of Breakthrough Profile


For a given bed depth, the service times of a unit plant are correlated with the
initial sorbate concentration, flow rate, and adsorption capacity of the adsorbent
to be used. So, obtaining a meaningful and reliable loading capacity of the adsor-
bent turns crucial in efficient process design and operation. This necessitates a
careful evaluation and analysis of the experimental data, to predict the effect of
variations in operational parameters of the sorption process, through modeling.

8.4.7.1  Hutchins BDST Model


The BDST model proposed by Hutchins in 1973 is based on the assumptions
that intra particular diffusion and external mass resistance are negligible and
that adsorption kinetics is controlled by the surface chemical reaction between
the solute in the solution and the unused adsorbent.46 A linear relationship
between the column depth (Z) and service time (t) was proposed as follows:

N0 1 ⎛C ⎞
t= Z− ln ⎜ 0 −1⎟ (8.32)
C0u KC0 ⎝ Ct ⎠

where C0 is the initial solute concentration (mg/L), Ct is the desired solute


concentration at breakthrough (mg/L), K is the adsorption rate constant (L/
mg/min), N0 is the adsorption capacity (mg/L), Z is the depth of the sorbent
bed (cm), u is the linear flow velocity of feed to bed (cm/min), and t is the
service time of column under the conditions mentioned (min). The dynamic
144 Fluoride in Drinking Water

bed capacity (N0) and the adsorption rate constant (K) can be evaluated by
the linear regression of the following straight-line relationship:

t = aZ + b (8.33)

where slope

N 0 (8.34)
a=
C0u
intercept

1 ⎡C ⎤
b=− ln ⎢ 0 −1⎥ (8.35)
KC0 ⎣ Ct ⎦
The BDST model is a useful tool for comparing the performance of columns
operating under different process variables. If there is a change in the initial
solute concentration C0 to a new value Coʹ, the new values of aʹ and bʹ can be,
respectively, obtained from the slope and the intercept according to the rela-
tions proposed by Hutchins:

⎛C ⎞
aʹ = a ⎜⎜ 0 ⎟⎟ (8.36)
⎝ C0ʹ ⎠

bʹ = b ⎜⎜ 0 ⎟⎟
( )
⎛ C ⎞ ln C0ʹ Cb −1
(8.37)
⎝ C0ʹ ⎠ ln (C0 Cb −1)
When the linear flow rate is changed from u to uʹ, the new gradient aʹ can be
calculated as follows:
⎛u⎞
aʹ = a ⎜ ⎟ (8.38)
⎝ uʹ ⎠
The intercept remains unchanged, because it depends on only the inlet solute
concentration C0. This is useful to scale up the process for other flow rates with-
out further experimental run. Also, at 50% breakthrough (Ct/C0 = 0.5), the term
b in Equation 8.35 becomes zero, and Equation 8.33 is reduced to the following:

t50 = aZ (8.39)

So, if the sorption process follows the BDST model, the plot of t against Z at
50% breakthrough will represent a straight line passing through the origin.

8.4.7.2  Thomas Model


The Thomas model47 is another most general and widely used model in column
performance theory. This model is derived by assuming Langmuir kinetics
Adsorptive Removal of Fluoride 145

of adsorption–desorption with no axial dispersion and that the rate-driving


force obeys second-order reversible reaction kinetics. The data obtained in
fixed bed column studies are used to calculate the maximum solid-phase
concentration of sorbate on the sorbent and the adsorption rate constant. The
expression by Thomas for an adsorption column is given as follows:

Ct 1
= (8.40)
C0 ⎛ kTh qTh M ⎞
1+ exp ⎜ − kTh C0 t ⎟
⎝ Q ⎠
where kTh is the Thomas rate constant (mL/min/mg), qTh is the equilibrium sor-
bent uptake per gram of the adsorbent (mg/g), M is the amount of adsorbent in
the column (g), C0 is the influent sorbate concentration (mg/L), Ct is the effluent
concentration at time t (mg/L), Q is the flow rate (mL/min), and t is the sampling
time. The value of Ct/C0 is the ratio of effluent and influent sorbate concentrations.
The linearized form of the Thomas model is as follows:

⎛C ⎞ k q M
ln ⎜ 0 −1⎟ = Th Th − kTh C0t (8.41)
⎝ Ct ⎠ Q

The kinetic coefficient k Th and the adsorption capacity of column qTh can be
determined from a plot of ln[(C0/Ct) - 1] against t at a given flow rate.

8.4.7.3  Yoon–Nelson Model


The Yoon–Nelson model48 is based on the assumption that the rate of
decrease in the probability of adsorption for each adsorbate molecule is pro-
portional to the probability of adsorbate adsorption and the probability of
adsorbate breakthrough on the adsorbent. This model requires no detailed
data regarding the characteristics of adsorbate, the type of adsorbent, and
physical properties of the adsorption bed and so it is less complicated. The
model is expressed as follows:

Ct 1
= (8.42)
C0 1+ exp ⎡⎣kYN (τ − t)⎤⎦

where k YN is the rate constant (min−1), τ is the time required for 50% adsor-
bate breakthrough (min), and t is the breakthrough (sampling) time (min).
The linearized form of the Yoon–Nelson model is as follows:

⎛ Ct ⎞
ln ⎜ ⎟ = kYN t − τkYN (8.43)
⎝ C0 − Ct ⎠
The parameters k YN and τ may be determined from the plot of ln[Ct/(C0 - Ct)]
versus sampling time (t).
146 Fluoride in Drinking Water

8.4.7.4  Clark Model


The model developed by Clark (1987)49 was based on the use of a mass-
transfer concept in combination with the Freundlich isotherm:

n−1
⎛ C0 ⎞
⎜ ⎟ −1 = Ae −rt (8.44)
⎝ Ct ⎠
where n is the Freundlich parameter, and A and r are the Clark constants:

⎛K N Z⎞
A = exp ⎜ c 0 ⎟ and r = K C (8.45)
⎝ u ⎠ c 0

Linearizing Equation 8.44:

⎛⎡ ⎤n−1 ⎞
C
ln ⎜⎢ 0 ⎥ −1 ⎟ = ln A − rt (8.46)
⎜⎣ Ct ⎦ ⎟
⎝ ⎠
From a plot of ln[(C0/Ct)n-1 – 1] versus time, the values of r and A can be
determined from its slope and intercept, respectively.

8.4.7.5  Wolborska Model


The Wolborska model50 is used for describing adsorption dynamics by using
mass transfer equations for diffusion mechanisms in the low-concentration
ranges of the breakthrough curve. The mass transfer in the fixed bed ­sorption
is described by the following equations:

∂Cb ⎛ ∂C ⎞ ⎛ ∂q ⎞ ⎛ ∂2 C ⎞
+ u ⎜ b ⎟ + ⎜ ⎟ = D ⎜ 2 b ⎟ (8.47)
∂t ⎝ ∂Z ⎠ ⎝ ∂t ⎠ ⎝∂ Z⎠
∂q ⎛ ∂q ⎞
= −v ⎜ ⎟ = β(Cb − Cs ) (8.48)
∂t ⎝ ∂Z ⎠

where Cb is the fluoride concentration in solution (mg/L), Cs is the fluoride


concentration at the solid/liquid interface (mg/L), β is the kinetic coefficient
of the external mass transfer (h−1), v is the migration rate (cm/min), D  is
the axial diffusion coefficient (cm2/min), and q is the fluoride concentration
on the sorbent at any time t (mg/L). With some assumptions previously
described by Wolborska: Cs << Cb, v << u, and axial diffusion is negligible
(D → 0 as t → 0), the solution can be approximated to50

⎛ C ⎞ β C0 t β Z
ln ⎜ t ⎟ = − (8.49)
⎝ C0 ⎠ N0 u
Adsorptive Removal of Fluoride 147

8.4.7.6  Bohart and Adams Model


Bohart and Adams (1920)51 proposed the fundamental equations ­describing
the relationship between Ct/C0 and t for the quantitative description of a
­continuous flow fixed bed system. It was assumed that adsorption rate
is proportional to both the residual capacity of the adsorbent and the
­concentration of the sorbate. The Bohart and Adams model is generally used
only for ­describing the initial part of the breakthrough curve (Ct/C0 ~ 0.5).
The mass transfer rates obey the following equations52:

∂q
= −kABq Cb (8.50)
∂t
∂Cb k
= − AB qCb (8.51)
∂Z u
where kAB is the Adams and Bohart kinetic constant (L/mg/h), q is the flu-
oride concentration in the sorbent at any time t (mg/L), Cb is the fluoride
concentration in solution (mg/L), and u is the linear flow velocity of feed to
bed (cm/min). The solution of the differential equations cited earlier with
the following assumptions of a low concentration field (Ct < 0.15C0) and that
when t → ∞, q → N0 (its saturation concentration) renders a linear relation-
ship between its parameters as follows:

⎛C ⎞ Z
ln ⎜ t ⎟ = kABC0t − kAB N 0 (8.52)
⎝ C0 ⎠ u
From this equation, values describing the characteristic operational parameters
of the column can be determined from a plot of ln(Ct/C0) against t, at a given
bed height and flow rate. In all cases, the average percentage errors (APE)
between the experimental and predicted values were calculated as follows1:

N Ct(exp) − Ct(theo)
∑i=1
Ct(exp)
APE% = ×100 (8.53)
N
where N is the total number of samples.

8.4.8 Regeneration
The regeneration capacity of an adsorbent also plays an important role in
making sorptive systems economical. The exception is where there is very
long adsorption or loading cycles due to very low concentrations of solute
in the inlet feed; this type of system usually uses the adsorbent only once
on a throw away basis and safe disposal becomes a problem. If very large
quantities of adsorbent are involved, regeneration and reuse becomes neces-
sary. Regeneration of adsorbents can be accomplished with heat, chemical
148 Fluoride in Drinking Water

change, or solvent action. Each of these methods has advantages, and dis-
advantages when applied to specific adsorbates, adsorbents, and systems.
The regeneration methods using solvents are relatively straightforward and
are commonly used to determine the working capacity of an adsorbent.
Thermal treatment is more difficult to perform and is related to full-scale
performance. Very few regeneration methods can be economically operated
to 100% efficiency. Generally, there will be a reduction in the capacity of the
adsorbent due to successive regeneration.
Since the solute is bound to the adsorbent by physical and/or chemical
forces, the regeneration procedure must develop conditions that these forces
should overcome. This can be accomplished either by subjecting the sys-
tem to conditions where the attractive forces for the adsorbed solute by the
regenerating medium are greater than the adsorbent attractive forces or by
chemically changing the solute so that the binding forces are neutralized.
An elution curve, which plots the concentration of the sorbate in the regener-
ant as a function of time or regenerant volume, will describe the efficiency
of regeneration. The shape of the regeneration curve can be influenced by
regenerant concentration, temperature, and flow rate. At the end of the regen-
eration, the void spaces in the adsorbent bed are filled with the regenerant
solvent. Before returning the adsorbent bed back into service, the regenerant
must be adequately rinsed from the adsorbent grains.

8.5  Results and Discussion


8.5.1  Characterization of the Adsorbent
The physical and chemical properties of an adsorbent play an important role
in determining its performance in terms of its sorption capacity. The adsorbent
used in this study, hardened granules of high alumina cement, is a composite
mixture of Al–Ca–Si–Fe containing substances. The adsorbent is abbreviated
as ALC. It is a variety of commercially available cement with a high content
of alumina (Al2O3). XRD analysis (Figure 8.2) identified the presence of com-
pounds such as

aluminum oxide (Al2O3), aluminum iron silicon (Al0.7Fe3 Si0.3), mag-


netite (Fe3O4), calcium iron oxide (Ca4Fe9O17), calcium aluminum
oxide (Ca3Al10O18), silicon oxide (SiO2), calcium aluminum silicate
(Ca0.88Al1.77Si2.23O8), and wollastonite (CaSiO3) along with many other
trace compounds.3

The physical parameters and major chemical constituents of ALC media identi-
fied by EDX are shown in Table 8.1. The surface texture of ALC was observed
by SEM studies. Figure 8.3 shows the SEM photograph of ALC at a magnifica-
tion of 500 µm. Figure 8.4 (at 10 µm magnification) shows a rough and a highly
porous surface texture that may be effective for high sorptive removal.
Adsorptive Removal of Fluoride 149

Intensity (a.u.)

ALC

20 30 40 50 60 70 80 90
2θ (degrees)

FIGURE 8.2
The x-ray diffractogram of ALC. (Modified from Ayoob, S., Gupta, A.K., Bhakat, P.B. and Bhat,
V.T., Chem. Eng. J., 140, 6–14, 2008.)

TABLE 8.1
Properties of ALC Media
Properties Quantitative Value

Geometric mean size (mm) 0.212


Bulk density (g/cm3) 2.33
Al2O3 (%) 78.49
CaO (%) 15.82
SiO2 (%) 5.39
Fe2O3 (%) 0.30
pH of the PZC 11.32
BET surface area (m2/g) 4.385
Source: Ayoob, S. and Gupta, A.K., Chem. Eng. J., 150, 485–491, 2009.

8.5.2  Kinetics Studies


The adsorption rate is strongly influenced by several parameters related to the
state of the solid, generally having a very heterogeneous reactive ­surface, and
to the physicochemical conditions under which adsorption is carried out.

8.5.2.1  Agitation Rate


The influence of agitation speed on the sorption of fluoride onto ALC was
studied by changing the speed of agitation from 120 to 280 rpm. Figure 8.5
shows that the sorption was influenced by the rate of agitation. The
removal increases from 64.99% to 78.49% and from 74.04% to 90.09% with
150 Fluoride in Drinking Water

FIGURE 8.3
Scanning electron microscopic (SEM) photograph of ALC particles at 500 µm magnification.

FIGURE 8.4
SEM photograph of ALC particles at 10 µm magnification.

an increase in agitation speed from 120 to 240 rpm in natural and synthetic
waters, respectively. The increasing agitation rate decreases the  boundary
layer resistance to mass transfer in the bulk and increases the driving force
of the fluoride ions. This may indicate that film d ­ iffusion does not domi-
nantly control the overall adsorption process.53 Further sorption studies
were carried out at an agitation rate of 230 ± 10 rpm.

8.5.2.2  Adsorbent Dosage


The response of the adsorbent to different adsorbent dosages is as shown in
Figure 8.6. The observed increase in removal efficiency with an increase in
Adsorptive Removal of Fluoride 151

100

Fluoride removal (%)


90

80

70

60
100 150 200 250 300
Agitation rate (rpm)

FIGURE 8.5
Effect agitation rates on fluoride sorption onto ALC.

100

75
% Fluoride removal

50

25 Natural water
Synthetic water

0
0 2.5 5 7.5 10 12.5
Dose of ALC (g/L)

FIGURE 8.6
Effect of dose variations of ALC on fluoride removal percentage. (From Ayoob, S. and Gupta,
A.K., Chem. Eng. J., 150, 485–491, 2009.)

ALC doses implies that the process is dependent on the availability of sorp-
tive binding sites. This sorption pattern indicates the predominance of sur-
face adsorption, since both the internal and external sorption sites are found
to increase with higher adsorbent dosage.54

It is postulated that at low adsorbent dosages all types of sites are


entirely exposed to adsorption, and the surface may become saturated
faster. However, at higher adsorbent dosages, the availability of higher
energy sites may be reduced and a larger fraction of lower energy sites
may get occupied. This results in an overall decrease in binding energy
of the surface, and a reversible type of process exists between the fluo-
ride ions attached to low energy sites and those present in bulk solution.
This could be the reason for the observed increase in uptake with an
increase in adsorbent dosage up to a certain stage and almost a constant
uptake thereafter.2
152 Fluoride in Drinking Water

8.5.3  Kinetic Profile of Fluoride Uptake


The kinetic curve of fluoride removal sorption (Figures 8.7 and 8.8) displays
a rapid uptake for the initial few minutes and is followed by a slow phase.
The removal process reaches an equilibrium stage after around 150 min
with negligible removal after 3 h. Within the first 10 minutes itself, around
70%–75% fluoride gets removed. The maximum uptake of ALC at equilib-
rium in groundwater (fluoride concentration of 8.65 mg/L) was found to be
0.806 mg/g.2,3 The increase in adsorption capacity in the synthetic system
at higher fluoride concentrations, signifying heterogeneous sorptive sur-
faces, may be ascribed to high intramolecular competitiveness to occupy the
unsaturated lower energetic surface sites.7 This biphasic fluoride sorption
behavior ­suggests the key role of mass transfer in the removal process. The
initial rapid uptake may indicate surface-bound sorption and precipitative
removal. The slow second phase of fluoride sorption may be ascribed to the
long-range diffusion of fluoride ions onto the interior pores of ALC.55

100
% Removal of fluoride

75

50

25

0
0 50 100 150 200 250
Time (min)

FIGURE 8.7
Kinetic curve of fluoride sorption onto ALC. (Modified from Ayoob, S. and Gupta, A.K., Chem.
Eng. J., 133, 273–281, 2007.)

100

75
% Fluoride removal

25 mg/L (Syn)
50
8.65 mg/L (Syn)
5 mg/L (Syn)
25
2.5 mg/L (Syn)

0
0 50 100 150 200 250
Time (min)

FIGURE 8.8
Kinetic curve of fluoride sorption onto ALC at different concentrations.
Adsorptive Removal of Fluoride 153

Kinetic models were applied to examine and describe the dynamics of


sorption of fluoride onto ALC. Kinetic modeling was carried out by the
pseudo-first-order model, pseudo-second-order model, intra particle surface
diffusion model, and the Elovich model; a comparison of the best-fit sorption
mechanisms was also made.

8.5.3.1  Pseudo-First-Order Model


The linearized form of the pseudo-first-order model indicates that a linear fit
between ln(qe – qt) and contact time (t) demonstrates that the reaction may fol-
low a pseudo-first-order reaction.3 The plots of ln(qe – qt) versus t for synthetic
water is shown in Figure 8.9. This reasonable linear fitting of the kinetic data
in most portions of the contact time shows the possibilities of a diffusion-
controlled sorption mechanism,15 though the fitting in the initial portions is
relatively poor. The correlation coefficients were found to be 0.975 and 0.959
in synthetic and natural waters, respectively. The pseudo-first-order rate con-
stant ks1 for synthetic and natural water that was calculated from the slopes
of the linear plots were found to be 0.0215 and 0.0164 min–1, respectively.3

8.5.3.2  Pseudo-Second-Order Model


A linear fit between t/qt and contact time (t) indicates that the reaction is
of pseudo–second order. The plots of t/qt versus t for synthetic water is
shown in Figure 8.10. Linear regression curves show a good ­correlation, as
R 2 is 0.999 in synthetic waters and 0.998 in natural waters. This shows that
the pseudo-second-order model provides a b ­ etter fit to the ­sorption kinet-
ics data of fluoride onto ALC over the whole range of the s­ orption process.3
The pseudo-second-order constants k and h were ­calculated from the inter-
cept and slope of the lines obtained by ­plotting t/qt against t. The initial
sorption rates (h) of sorption were found to be 0.9706 and 0.1877 mg/g/h in
synthetic and natural waters with ­corresponding values of rate constants (k)
as 0.0625 and 0.2891 g/mg/min.

1
0.5
0 y = –0.0215x + 0.2567
–0.5 R2 = 0.9746
In (qe-qt)

–1
–1.5
–2
–2.5
–3
–3.5
0 50 100 150 200
t

FIGURE 8.9
Pseudo-first-order kinetic fit of fluoride sorption onto ALC.
154 Fluoride in Drinking Water

50

40

t/qt (min g/mg)


30

20

10
0
0 50 100 150 200
t

FIGURE 8.10
Pseudo-second-order kinetic fit of fluoride sorption onto ALC.

8.5.3.3  Intra Particle Surface Diffusion Model


The use of the intra particle surface diffusion model has been widely
explored to elucidate the mechanism of the sorption process. It is sug-
gested that if the plot of q versus t1/2 renders a straight line, the sorption
process is controlled by intra particle diffusion. If it does not pass through
the origin, it indicates that the intra particle diffusion is not the only rate-
limiting step. It further suggests that the process is complex, with more
than one mechanism limiting the rate of sorption. In this study, though the
plot of adsorbate uptake versus the square root of time (Figure 8.11) could
be represented by such a linear relationship (R 2 > 0.899), it was found to
not pass through the origin. This indicates that intra particle diffusion is
involved in the sorption process. However, it is not the only rate-limiting
mechanism and some other mechanisms may also get involved. It was
pointed out that such a deviation of the straight line from the origin may
be due to the difference in the rate of mass transfer in the initial stage
of sorption.56,57 The values of the intercept C provide information about
the thickness of the boundary layer, that is, the resistance to the external
2
qt (mg/g)

0
0 5 10 15
t1/2 (min)1/2

FIGURE 8.11
Intra particle diffusion model kinetic fit of fluoride sorption onto ALC.
Adsorptive Removal of Fluoride 155

mass transfer. The larger the intercept is, the higher the external resis-
tance; namely, any increase in the intercept indicates the abundance of
solute adsorbed on the boundary layer. The slope shown in Figure 8.11 was
used to find the rate parameter (kp) of the model. The application of this
model to the natural and synthetic systems of this study rendered values
of R 2, kp, and C as 0.912, 0.1172 mg/g/h1/2, and 2.5261 mg/g respectively, in
synthetic water with corresponding values of 0.899, 0.0164 mg/g/h1/2, and
0.5843 mg/g in natural water.

8.5.3.4  Elovich Model


The Elovich equation has been applied in aqueous systems to describe
adsorption and desorption reactions. The Elovich equation can be derived
from either a diffusion-controlled process or a reaction-controlled process.

If the Elovich equation is based on adsorption on an energetically het-


erogeneous surface, the parameter β is related to the distribution of
activation energies. However, in the diffusion control model, it is a
function of particle and diffusion coefficients.19

As per the model, if it follows the Elovich equation, the kinetic results will
be linear on a qt versus ln t plot. The kinetic curve of sorption demonstrates
excellent fitting (Figure 8.12) in synthetic systems (R 2 = 0.996) and natural
systems (R 2 = 0.838). The fitting to the Elovich kinetics pattern indicates that
the rate-determining step is diffusion in nature.58
On comparing the fitting of the applied kinetic models, it is evident that
the kinetic profile could be best modeled by pseudo–second order, which
is indicative of a chemisorptive rate-limiting step. The model could also
excellently predict the equilibrium adsorption capacity of ALC in natural
and synthetic water as 0.806 and 3.941 mg/g compared with corresponding
experimental values of 0.803 and 3.912 mg/g.

8.5.3.5  Arrhenius Equation


The value of energy of activation (Ea) obtained from Arrhenius equation
could also be a useful kinetic parameter in assessing rate-limiting steps.
Low Ea values indicate diffusion-controlled transport and physical adsorp-
tion processes, whereas higher Ea values indicate chemical reactions or
surface-controlled processes. Some of the assigned values of Ea (kJ/mol)
include 8–25 to physical adsorption, less than 21 to aqueous diffusion,
20–40 to pore diffusion, and greater than 84 to ion exchange.23 The plot of
ln K versus 1/T, which is required to calculate the activation energies, is
shown in Figure 8.13.
The activation energies of the sorption process were calculated by
­multiplying the slope of the curves by the value of R. Accordingly, the values
of Ea were found to be 17.67 and 20.12 kJ/mol in synthetic and natural waters,
156 Fluoride in Drinking Water

3
qt (mg/g)
2

0
0 2 4 6
ln t

FIGURE 8.12
Elovich model kinetic fit of fluoride sorption onto ALC.

–3.2
–2.7
R2 = 0.999
–2.2
ln K

–1.7
–1.2
–0.7
3.2 3.25 3.3 3.35 3.4 3.45 3.5
10–3(1/T) K–1

FIGURE 8.13
The plot of ln K versus 1/T of fluoride sorption onto ALC in synthetic water. (Modified from
Ayoob, S. and Gupta, A.K., Chem. Eng. J., 150, 485–491, 2009.)

respectively.3 This value of activation energies signifies the role of diffusion-


controlled transport and physical adsorption processes in the rate limiting of
fluoride sorption onto ALC.3

8.5.4  Elucidation of Rate-Limiting Step


In this study, the following procedure was adopted to elucidate the
­rate-limiting step19:

(1) The adsorption patterns should be sensitive to inert electrolyte con-


centrations or pH if the inter particle diffusion is rate limiting. (2) The
desorption pattern will follow the same two-stage pattern as that of
sorption if intra particle diffusion is rate limiting. (3) If both inter par-
ticle and intra particle diffusion are not rate controlling, the surface
reactions will be rate limiting.3

So, to ascertain whether inter particle diffusion is rate limiting, the effects
of pH and electrolyte concentrations were investigated. ALC exhibited
Adsorptive Removal of Fluoride 157

consistent removal (Figure 8.14) in the pH range of 3–11.5, and it was reduced
slightly thereafter by around 10% at pH 12.3

The pH dependence of fluoride sorption onto ALC could be well


explained in terms of its pHZPC (11.32). The pHZPC indicates where the
net surface charge on the adsorbent is zero. When pH < pHZPC, the net
surface charge on the solid surface of ALC is positive due to adsorption
of excess H+, which favors adsorption due to coulombic attraction. At
pH > pHZPC, the net surface charge is negative due to desorption of H+
and adsorption must compete with coulombic repulsion. The consistent
fluoride removal in the range of 3–11.5 could be due to the combined
effect of both chemical and electrostatic interactions between the oxide
surfaces and fluoride ion. The observed reduction in fluoride adsorption
above pH 11.5 may suggest that the strong negative surface charge devel-
oped at this pH may cause repulsion for the available adsorption sites.
However, the considerable potential of the adsorbent above its pHZPC of
11.32 may be attributed to the predominance of specific adsorption due
to chemical interactions.3

As shown in Figure 8.14,

fluoride sorption is also found to be unaffected by electrolyte con-


centrations in the range of 10 –4 M NaNO3 to 10 –1 M NaNO3, though
the marginal increase observed at higher concentrations may be due
to compression of the electrostatic double layer.3,59 Since the sorption
process is unaffected by inert electrolyte concentrations and pH over
a wide range, it can be suggested that the fluoride removal percentage,
and hence the surface coverage, approaches almost the same value
within the range of inert electrolyte concentrations tested (10 –4 to 10 –1
M), and the rate of sorption is unaffected by the effective particle size
or diffusive path lengths.3

This response to pH and inert electrolyte concentrations

0.1
Electrolyte concentration (M)

12
0.08
10
0.06
8
pH

6 0.04
pH Conc. of electrolyte (M) 0.02
4
2 0
0 50 100
Fluoride removal (%)

FIGURE 8.14
Effect of pH and inert electrolyte concentrations on the fluoride sorption onto ALC. (Modified
from Ayoob, S., Gupta, A.K., Bhakat, P.B. and Bhat, V.T., Chem. Eng. J., 140, 6–14, 2008.)
158 Fluoride in Drinking Water

adsorbed/desorbed (mg/g)
4

Fluoride 2

1 Adsorption
Desorption
0
0 50 100 150 200
Time (min)

FIGURE 8.15
The sorption and desorption patterns of fluoride by ALC. (Modified from Ayoob, S., Gupta,
A.K., Bhakat, P.B. and Bhat, V.T., Chem. Eng. J., 140, 6–14, 2008.)

clearly demonstrates that inter particle or external diffusion is not the


rate-determining step. Since the possibility of inter particle or exter-
nal diffusion is ruled out, the other possibility is that of intra particle
diffusion. Since the presence and shape of the pores in an adsorbent
does not change with pH or inert electrolyte concentrations, and
only depends on its crystal properties, the desorption behavior of the
adsorbent gives enough indication as to whether intra particle diffu-
sion is rate limiting.3

However, in batch desorption studies, it was observed that concentra-


tions of fluoride increased slowly with time with nearly 55% and 63% of
the fluoride not being desorbed in synthetic and natural systems, respec-
tively. The sorption and desorption patterns are shown in Figure 8.15.

The desorption pattern in column studies also showed a poor response


with much of the fluoride not being desorbed using 10% NaOH, even
after 20 bed volumes. So, it is clear that the sorption and desorption
patterns are not identical, demonstrating that the diffusion into ALC
(i.e., intra particle diffusion) is not the rate-limiting process. Since nei-
ther inter particle nor intra particle diffusion is rate limiting, it indi-
cates that the slow sorption phase in fluoride sorption is not the result
of diffusion. So by elimination, it could be plausible that it is due either
to the heterogeneity of the surface site-binding energy or to other reac-
tions controlling fluoride removal from solution.3

8.5.5  Fluoride Removal Mechanism


The mechanisms of sorptive removal of fluoride may include specific or ­non
specific adsorption or both. Experimental evidence suggests that anions
that form inner-sphere complexes coordinate directly with the oxide sur-
face without getting influenced by the ionic strength.60 Since the sorption of
fluoride onto ALC is relatively independent of the electrolyte concentrations
and pH, the mechanism of removal can be ascribed to inner-sphere complex
formations.
Adsorptive Removal of Fluoride 159

The poor desorption characteristics of ALC suggest stronger adsorption of


fluoride ions, further supporting the formations of inner-sphere complexes.
As cited earlier, the SEM analysis observed that Al2O3 was the most promi-
nent metal oxide in ALC. Further, the XRD analysis (Figure 8.16), which
was carried out to identify the morphological structure and the extent of
crystallinity of the adsorbent, shows multiple peaks, thus indicating the
presence of various oxides and heterogeneous surface sites for sorption.3

On hydration, the metal ions on the oxide surface complete their coordina-
tion shells with OH groups. Depending on the pH, these OH groups can bind
or release H+, resulting in the development of a surface charge as follows:

MOH + H + = MOH 2 + (8.54)

MOH = MO –  + H+ (8.55)

where M is the metal (Al, Si, Fe, etc.) and MOH2+, MOH, and MO – are posi-
tive, neutral, and negative surface hydroxo and oxo groups, respectively.3
The adsorption properties of the metal oxides are due to the presence
of these OH2+, OH, and O – surface functional groups,61 as they dictate the
number of reactive sites. Accordingly, a surface complex formation model is
proposed (ligand-exchange model) to describe fluoride adsorption on metal
oxides as follows3,36:

MOH + F – = MF + OH – (pH > 7) (8.56)

FALC
Intensity (a.u.)

FALCN

ALC

20 30 40 50 60 70 80 90
2θ (degrees)

FIGURE 8.16
XRD pattern of the adsorbent before sorption (ALC) and after sorption of fluoride in natural
(FALCN) and synthetic (FALC) water. (Modified from Ayoob, S., Gupta, A.K., Bhakat, P.B. and
Bhat, V.T., Chem. Eng. J., 140, 6–14, 2008.)
160 Fluoride in Drinking Water

The equation implies that OH– is released from the ALC surface into
the bulk phase, which is confirmed by the rise in pH during fluoride
sorption in both natural and synthetic water and is more prominent at
the initial stages. To experimentally quantify the process represented
by the equations cited earlier and to understand spectroscopic changes
in the adsorbent due to fluoride sorption, FTIR analysis was performed
both before and after adsorption3

in synthetic and natural water. As shown in Figure 8.17, the FTIR spectrum
of the samples presents no significant spectroscopic change due to fluoride
sorption. The broad band corresponding to 3469 cm–1 (range of 3550–3200
cm–1) represents O–H stretching vibrations, that at 1420 cm–1 represents
Al–H stretching, and that at 1015 cm–1 represents the characteristic stretch-
ing bands of Al═O. The band at 570 cm–1 may be ascribed to the stretching of
Al–OH, that at 662 cm–1 represents Si–H, and that at 872 cm–1 indicates Fe–O
stretching.3
On closer examination, it has been observed that the intensity of many of
the peaks shows variations after fluoride sorption. Before adsorption, in the
virgin adsorbent (ALC), this peak height ratio was 1.0256, but after fluoride
sorption, it was found to be 2.1348 in synthetic water (FALC) and 1.7613 in
natural water (FALCN). This shows that the OH– band at 3750 cm–1 is decreas-
ing due to fluoride sorption, confirming the exchange between OH and F
ions, and enhancing fluoride removal. In a similar way, the peak height ratio
of OH– band at 3469 cm–1 to that of Al–OH band at 561 cm–1 is also compared.

FALCN
% Transmittance (a.u.)

FALC
2520

872
662
570
1015
1420

ALC
3469

4000 3500 3000 2500 2000 1500 1000 500


Wave number (cm–1)

FIGURE 8.17
FTIR spectra of the adsorbent before sorption (ALC) and after sorption of fluoride in synthetic
(FALC) and natural (FALCN) water. (Modified from Ayoob, S., Gupta, A.K., Bhakat, P.B. and
Bhat, V.T., Chem. Eng. J., 140, 6–14, 2008.)
Adsorptive Removal of Fluoride 161

The ratio of the peak height was calculated as 0.99268 in ALC, whereas
it is found to be 2.129 and 3.08 in FALC and FALCN, respectively. This
clearly indicates that the OH ions of the Al–OH band are consumed in
fluoride sorption, more prominently for the exchange reactions. Further,
these observations are experimentally supported by the observed
increase in pH (6.9 ± 0.4 to 11.7 ± 0.4) due to fluoride uptake, especially at
the initial stages of sorption.3

So, the predominant mechanism of fluoride removal can be illustrated as


shown in Figure 8.18.

The role of sodium in groundwater deserves special mention, as the


reduction in fluoride sorption at high pH may also be attributed to the
sodium occupying sorption sites and altering the surface structure.
Under basic conditions, surface hydronium ions can dissociate, enabling
the aluminum to serve as a Lewis acid toward cations (such as sodium,
calcium, or iron). This suggests that cations will be sorbed onto the alu-
mina surface under basic conditions,3

involving the formation of an outer-sphere complex as follows:

AlOH + Na+ = AlONa + H + (8.57)

Before sodium sorption occurs, ═Al–OH and ═Al–O− sites are present in an
alkaline solution. It is suggested that the sorption onto alumina (═ALONa)
increases under alkaline conditions when the surface of the adsorbent is
negatively charged.42

This may lead to electrostatic repulsion between sorbed Na+ ions and
adjacent H+ ions, causing some H+ ions to break away from the surface;
thus, fluoride sorption becomes reduced due to a reduction in available
═AlOH sites. Due to the sodium occupying sorption sites, the concen-
trations of ═Al–OH and ═Al–O−, in turn, reduce the total number of
available sites for fluoride sorption. Since these reactions take place only
when the surface charge of ALC is negative, this indicates that they
occur during the late hours of sorption. This further supports the fact

Metal(Al, Si) Metal(Al, Si)

Oxygen OH F— Oxygen F OH

FIGURE 8.18
The ligand-exchange model of fluoride sorption onto ALC. (From Ayoob, S., Gupta, A.K.,
Bhakat, P.B. and Bhat, V.T., Chem. Eng. J., 140, 6–14, 2008.)
162 Fluoride in Drinking Water

that surface reactions are rate limiting. Thus, the fluoride sorption onto
ALC may be viewed as an inner-sphere complexation that is predomi-
nated by a ligand-exchange process. The surface reactions leading to
mixed surface complex precipitative reactions or scavenging reactions
may also be involved.3

8.5.6  Isotherm Studies


The fitting of equilibrium sorption data to the Langmuir, Freundlich, and
D–R isotherm models, for a wide range of fluoride concentrations (2.5–100
mg/L) in synthetic water, is shown in Figure 8.19. The visual inspection of the
equilibrium sorption curves represented by the isotherm models indicate,

A comparable uptake, of fluoride by ALC. An enhanced uptake of fluo-


ride by ALC at a higher concentration is visible for LI and DRI models
(with a tendency to tail off at very high concentrations), whereas FI pro-
vides an identical uptake pattern at all concentrations.4

The shape of the fluoride isotherm data suggests that both FI and LI
models would provide a better fit to the experimental data at all concen-
tration ranges compared with DRI. It can be observed that (Table 8.2) LI
fit the data better at all ranges of concentration with very high consistent
values of R 2 (>0.995). The FI also makes identical fitting, though with
slightly lesser R 2 values. However, the FI fitting becomes better towards
higher concentration ranges, with a steady increase in the value of R 2
from 0.9738 to 0.9955.4

Thus, the sorption of fluoride by ALC derives significance in the concentra-


tion ranges studied, as it deviates from the notion that FI usually fits better at
low concentrations and LI fits better at higher concentrations.62

60
qe(exp)
Fluoride absorbed qe (mg/g)

50 LI
FI
40 DRI
30

20

10

0
0 5 10 15 20 25
Equilibrium concentration Ce (mg/L)

FIGURE 8.19
A comparison of the experimental data and various isotherm models in the concentration
range of 2.5–100 mg/L of fluoride. (Modified from Ayoob, S. and Gupta, A.K. J. Hazard. Mater.,
152, 976–985, 2008.)
Adsorptive Removal of Fluoride 163

TABLE 8.2
Isotherm Model Parameters and R 2 Values in Natural and Synthetic Systems
Range of Fluoride Concentrations
(mg/L) Synthetic Samples at 300 K
Isotherm Model Parameters
Model and R 2 2.5–20 2.5–40 2.5–80 2.5–100

LI R
2 0.9952 0.9955 0.9954 0.9955
b 0.329 0.2958 0.244 0.230
qmax (mg/g) 24.57 27.17 32.57 34.36
FI R2 0.9738 0.9776 0.9843 0.9955
1/n 0.7755 0.7939 0.7839 0.7248
Kf 5.593 5.676 5.682 5.610
DRI R2 0.9588 0.8729 0.7821 0.7536
Kad 0.0791 0.0955 0.1135 0.1215
Qm (mg/g) 7.776 10.46 14.39 16.510
Source: Ayoob, S. and Gupta, A.K., J. Hazard. Mater., 152, 976–985, 2008.

As cited earlier, this traditional approach of determining the isotherm


parameters by linear regression of LI and FI models appears to give very
good fits to the experimental data, as their respective R 2 values are very
high (Table 8.2). Thus, purely from the comparison of R 2 values (being
very close to unity), at all concentration ranges, the linearized Langmuir
isotherm would be expected to provide a better fit to the experimental
data than the linearized Freundlich isotherm.4

8.5.6.1  Effects of Temperature


The effects of temperature on all isotherm models in both natural and synthetic
systems (with dose variations of ALC) are shown in Figure 8.20, with respec-
tive parameter values in Table 8.3. As can be seen, both LI and FI have shown
a reasonably good fitting at all temperature ranges in both systems than DRI,
though values of R2 (Table 8.3) suggest that the Freundlich model fits better.4

8.5.7  Performance Evaluation of ALC in Natural and Synthetic Systems


Though the sorption profiles of fluoride removal were found to be identical in
both synthetic and natural waters, ALC exhibited a reduced fluoride adsorp-
tion capacity in treating natural water compared with synthetic systems.
The optimum dose requirement of ALC in natural water is five times more
than that for synthetic water. The maximum monolayer saturation capacity
of ALC showed identical reduction (with values of 10.214 mg/g in synthetic
waters, and 0.9358 mg/g in natural waters).2 Since the factors influencing the
adsorption process and capacity of the adsorbent include pH, ionic strength,
interference effects from other counter-ions, initial adsorbate concentrations,
and temperature, these effects were investigated to elucidate the reasons for
its reduced sorption capacity in natural water.
164 Fluoride in Drinking Water

6
qe (mg/g)

4
qe (exp)
LI
2 FI
DRI

0
0 0.5 1 1.5 2
Ce (mg/L)

FIGURE 8.20
A comparison of the experimental data and various isotherm models in synthetic samples
under dose variation study. (Modified from Ayoob, S. and Gupta, A.K., J. Hazard. Mater., 152,
976–985, 2008.)

TABLE 8.3
Isotherm Model Parameters and R 2 Values in Natural and Synthetic Systems at
Different Temperatures
Dose Variation Study
Temperature of Observations and Nature of Samples
Model 290 K 300 K 310 K
Isotherm Parameters
Model and R 2 Natural Synthetic Natural Synthetic Natural Synthetic

LI R2 0.8768 0.9877 0.9215 0.9599 0.9672 0.9918


b 1.0336 0.8245 15.22 1.0471 10.696 1.3036
qmax (mg/g) 1.0779 9.09 0.9358 10.215 1.164 12.658
FI R2 0.9275 0.9937 0.9364 0.9734 0.9391 0.9878
1/n 0.3342 0.5569 0.1022 0.5959 0.1553 0.6287
Kf 0.5589 3.980 0.825 5.1924 0.9939 7.5198
DRI R2 0.7440 0.9048 0.8826 0.9278 0.9463 0.4651
Kad 0.2046 0.0864 0.0136 0.1291 0.0171 0.0572
Qm (mg/g) 0.9157 6.9302 0.9157 6.1978 1.140 8.2137
Source: Ayoob, S. and Gupta, A. K., J. Hazard. Mater., 152, 976–985, 2008.

8.5.7.1  Effect of pH, Ionic Strength, and Temperature


Since the sorption studies on synthetic water were also conducted at
the same pH of natural water (6.9 ± 0.4), the reduction in the adsorption
potential in the natural system could not be ascribed to the effect of pH.
However, the sorptive responses of ALC were investigated at different
values of pH to understand the mechanism and to ascertain its field use.
As shown in Figure  8.14, the fluoride removal percentage was almost
consistent in the pH range of 3–11.5 and was reduced thereafter. This
could be readily explained as follows: When pH < pHZPC, the net surface
Adsorptive Removal of Fluoride 165

charge on the solid surface of ALC is positive due to the adsorption of


excess H+, which favors adsorption due to coulombic attraction; whereas
at pH > pHZPC, the net surface charge is negative due to desorption of H+
and adsorption must compete with columbic repulsion.2

The observed reduction in fluoride adsorption above pH 11.5 may suggest that
the strong negative surface charge developed at this pH may prevent fluoride
from occupying available adsorption sites through columbic repulsion.2 The
effect of ionic strength on fluoride sorption onto ALC is shown in Figure 8.14.
Experimental evidence suggests that anions that form inner-sphere com-
plexes coordinate directly with the oxide surface without getting influenced
by ionic strength.60 Since the fluoride sorption is almost unaffected by the
ionic strength, it can be surmised that the removal of fluoride occurs mainly
through the formation of the inner-sphere surface complexation process.

Since sorption studies on both systems were conducted at the same tem-
peratures of 300 K, the reduction in the adsorption potential in natural
water could not be ascribed to the effect of temperature. However, the
temperature effects of sorption were evaluated in both systems within the
range from 290 to 310 K to delineate the nature of sorption mechanisms
in terms of the thermodynamic parameters: Gibbs free energy (ΔG°), stan-
dard enthalpy (ΔH°), and standard entropy changes (ΔS°).2

The ln K versus (1/T) plot is shown in Figure 8.13. The respective param-
eter values are illustrated in Table 8.4.

The increase in fluoride sorption with temperature in both systems


reflects the surface heterogeneity of ALC and its increased activity, which
results in enhanced diffusion of fluoride ions into its pores. It would be
expected that higher temperatures stimulate the surface reactivity of the
bound oxides/hydroxides, which increases the sorption capacity of the
system. The negative values of ΔG° confirm the spontaneity of sorption in
both systems within the conditions applied. The higher negative value at
elevated temperatures assures more energetically favorable adsorptions.

TABLE 8.4
Thermodynamic Parameters of Adsorption of Fluoride onto ALC in Natural and
Synthetic Water
ΔG° (kJ/mol)
ΔH° ΔS°
System (kJ/mol) (kJ/mol/K) 290 K 300 K 310 K

Synthetic 36.569 0.144 –5.191 –6.631 –8.071


Water
Natural 83.810 0.301 –3.480 –6.490 –9.50
Water
Source: Ayoob, S. and Gupta, A.K., Chem. Eng. J., 150, 485–491, 2009.
166 Fluoride in Drinking Water

Further, the decrease in the magnitude with increasing temperature indi-


cates more efficient sorption at elevated temperatures.2

The endothermic nature of the process in both systems is confirmed by


the positive enthalpy values (ΔH°). The positive value of entropy change
(ΔS°) reflects the affinity of ALC toward fluoride, which may also indicate
some structural changes within the adsorbent.2,63,64 Generally, the enhanced
adsorption at elevated temperatures indicates that chemisorption is taking
place in the system.2

8.5.7.2  Effects of Other Ions


The possible interferences from other ions that are commonly present in
natural water were also investigated to examine their influences in the fluo-
ride sorption onto ALC. Nitrates, chlorides, sulfates, and bicarbonates did not
significantly affect the sorption process. The effect, of silicate was insignifi-
cant up to 25 mg/L; thereafter, it slightly reduced the fluoride removal and
at 400 mg/L, the reduction was about 13%. The presence of calcium gradu-
ally enhanced the removal and at 400 mg/L, by around 7%. The presence of
iron also did not affect the removal up to 10 mg/L, whereas the interference
of phosphate was considerable. At a concentration of 4 mg/L, it reduces the
removal by around 6% and at 8 mg/L, by more than 10%. The interference
analyses suggest that high levels of salinity and hardness in water do not
affect the fluoride removal performance of the adsorbent.

Since fluoride-rich groundwaters are generally associated with high


bicarbonate ions in alkaline environments, their applications turn out to
be important. The interference of phosphate and silicate associated with
fluoride sorption was already reported. The reduction in fluoride sorption
in the presence of high silicates may be due to its scavenging of aluminum
ions forming aluminosilicate solute species, especially in an alkaline envi-
ronment. Also, silicic acid is known to inhibit the formation of aluminum
hydroxide precipitates by replacing hydroxylated aluminum(III) ions.
High silica concentrations may also result in its polymerization, resulting
in an increase in negative surface charge. The reduction in the fluoride
removal efficiency in the presence of phosphates may be due to the strong
affinity of aluminum(III) for phosphate, thereby reducing the availability
for fluoride uptake. The absence or negligible competitive effect produced
from most of the ions indicates that fluoride is strongly adsorbed onto
ALC. The interference pattern indicates that the presence of ion chlorides,
nitrates, and bicarbonates may form outer-sphere complexes but sulfates
and silicates form partial inner-sphere complexes with ALC.2

The total organic carbon (TOC) typically quantifies the amount of natural
organic matter (NOM) concentrations in the natural water sources that are
present as a result of adsorption onto aquifer solids or depositional history.65
Aqueous NOM represents a wide range of structurally complex compounds
Adsorptive Removal of Fluoride 167

that are derived from the chemical and biological degradation of plants and
animals, composed mainly of humic substances (humic and fulvic acids),
that are hydrophobic. Humic substances are complex mixtures containing
both aromatic and aliphatic components with mainly carboxylic and phenolic
functional groups; they were found to interfere with anionic adsorption2,66–69
through stable metal complex formations.70,71 In this study, additions of 5, 10,
20, and 40 mg/L of humic acid to fluoride samples (C0 = 8.65 mg/L) reduced
the percentage removal by 5.5%, 11.54%, 14.65%, and 18.03%, respectively.2

As the structurally complex product of biomass decomposition, NOM


molecules possess unique combinations of functional groups, includ-
ing carboxylic, esteric, phenolic, quinone, amino, nitroso, sulfhydryl,
hydroxyl, and other moieties, the majority of which are negatively
charged at neutral pH. Along with this predominant anionic nature
coupled with its high reactivity toward both metals and surfaces, NOM
can compete with fluoride for sorption onto ALC. In sorptions involving
natural groundwater, the adsorption capacity may depend on the acces-
sibility of the organic molecules to the inner surface of the adsorbent.
The small molecules can access micropores, and NOM can access meso-
pores of the adsorbent.2,72

This presorbed NOM reduces/destructs the sorption sites of the adsorbent.


It was suggested that NOM readily forms both aqueous and surface inner-
sphere complexes with cationic metals and metal oxides.2,73

Aqueous NOM-metal complexes may, in turn, associate strongly with


dissolved anions such as fluoride, presumably by metal-bridging mech-
anisms, diminishing the tendencies of fluoride ions to form surface
complexes. This metal bridging appears to be a potential mechanism for
reduced fluoride uptake in natural water. Though phosphates pose inter-
ferences to fluoride sorption in synthetic systems, since their respective
concentrations in natural water are less, they may not be responsible for
the reduced fluoride intake.2

However, the presence of silicate (Table 8.5), much above the average abun-
dance level of 14 mg/L in groundwater,6 may pose a slight interference.

It is plausible that the presence of NOM represented by the high TOC


value may play a role in the reduced capacity of the adsorbent in natural
water. However, further studies are warranted to elucidate the reduced
uptake in natural waters and the synergetic effects of various ions.2

8.5.8  Column Studies


The breakthrough curve developed from column studies basically portrays
the dynamic sorptive responses of the adsorbent used in a continuous flow
fixed bed. The shape of the breakthrough curve and the time for breakthrough
168 Fluoride in Drinking Water

TABLE 8.5
Characteristics of Natural Groundwater (Collected from Baliasingh
Patna, Kurda District, Orissa, India)
Characteristic Parameter Quantitative Value
Fluoride (mg/L) 8.65
pH 6.9 ± 0.4
TDS (mg/L) 463
Acidity (mg/L) 1.5
Alkalinity (mg/L) 260
Chloride (mg/L) 165
Total hardness (mg/L) 145
Total organic carbon (mg/L) 59.08
Total phosphorous (mg/L) 0.032
Silicate as SiO2 (mg/L) 39.22
Boron (mg/L) 0.33
Sodium (mg/L) 14.00
Potassium (mg/L) 2.00
Ammonia nitrogen (mg/L) 0.328
Salinity (PSS) 0.30a,b
Source: Ayoob, S. and Gupta, A. K. Chem. Eng. J., 150, 485–491, 2009.
a Salinity is expressed in practical salinity scale (PSS).

b Minimum detection limit of salinity = 0.1 PSS.

appearance are predominant factors in determining the operation and


dynamic response of an adsorption system. The general position of the break-
through curve along the volume/time axis depends on the capacity of the col-
umn with respect to bed height, feed concentration, and flow rate. The point at
which effluent fluoride concentration reaches 1 mg/L was taken as breakthrough
point and that corresponding to 90% of influent concentration (Ct/C0 = 0.90)
was considered point of exhaust. Fixed bed column studies were undertaken in
the up-flow mode to evaluate the performance of ALC in removing fluoride
under varying operating conditions. During the sorption experiments, it was
observed that the flow rate remained more or less constant, which indicated
that the clogging of pores did not occur and, hence, the sorption sites of ALC
were easily accessible through the inter particle pore network.

8.5.8.1  Effect of Process Parameters on Breakthrough


The breakthrough curves were obtained by varying the depths of the ALC
bed (Z), flow rates (Q), and initial fluoride concentrations (C0). A combined
graph (Figure 8.21) shows the bed volumes of water treated at different stages.
The respective service times, volumes of water treated, and the adsorption
capacity of the columns under various process conditions are illustrated
in Tables 8.6 and 8.7. It was observed that at lower Ct/C0 ranges, the curves
turn less steeper at higher bed depths under the same flow rates. Also, both
breakthrough and exhaust times increase with the corresponding volumes
Adsorptive Removal of Fluoride 169

0.75
Ct /Co

0.5 Co = 8.65 mg/L, Z= 5 cm, Q = 8 mL/min


Co = 8.65 mg/L, Z= 10 cm, Q = 8 mL/min
Co = 8.65 mg/L, Z= 15 cm, Q = 8 mL/min
0.25 Co = 8.65 mg/L, Z= 10 cm, Q = 12 mL/min
Co = 8.65 mg/L, Z= 10 cm, Q = 8 mL/min
Co = 4.0 mg/L, Z= 10 cm, Q = 8 mL/min
0 Co =15 mg/L, Z= 10 cm, Q = 8 mL/min
0 1000 2000 3000 4000 5000 6000

Bed volumes treated

FIGURE 8.21
Experimental breakthrough curves of fluoride sorption onto ALC with bed volumes of water
treated at different bed depths, flow rates, and initial fluoride concentrations in synthetic
water. (From Ayoob, S. and Gupta, A.K., Chem. Eng. J., 133, 273–281, 2007.)

TABLE 8.6
Sorption Data for Fixed Bed of ALC for Fluoride Sorption (Breakthrough) under
Different Process Conditions in Synthetic Water
Volume of
Initial Flow Water Treated
Fluoride ALC Bed Rate in Time for up to Point of Bed Volumes
Concentration Depth in mL/ EBCT Breakthrough Breakthrough up to Point of qmin,col
in mg/L (C 0) cm (Z) min (Q) (min) in h (bt) in L (Vb) Breakthrough (mg/g)

8.65 5 8 4.5749 17 8.16 519.48 1.8466


8.65 10 8 9.1498 33 15.84 504.20 1.8021
8.65 15 8 13.7247 58 27.84 590.78 2.096
8.65 10 4 18.2995 84 20.16 641.71 2.270
8.65 10 12 6.0998 20 14.40 458.37 1.6355
4.00 10 8 9.1498 72 34.56 1100.08 1.722
15.00 10 8 9.1498 22 10.56 336.14 2.087

Source: Modified from Ayoob, S. and Gupta, A.K., Chem. Eng. J., 133, 273–281, 2007.

of water treated, along with an increase in bed depths. The breakthrough


volumes are found to increase around two times by doubling the depth of
the bed from 5 to 10 cm; whereas they are more than three times for 15 cm.

It is naturally expected that the availability of more adsorbent at higher


bed depths offers more surface area and binding sites for sorption, result-
ing in enlarged mass transfer zones. However, the adsorption capacity
of ALC at these different bed depths (5–15 cm) for a particular flow rate
(8 mL/min) shows only marginal variation, indicating its consistency in
affinity for fluoride sorption.1
170 Fluoride in Drinking Water

TABLE 8.7
Sorption Data for Fixed Bed of ALC for Fluoride Sorption (Exhaust) under Different
Process Conditions in Synthetic Water
Flow Volume of Bed
Initial Fluoride ALC Bed Rate in Time for Water Treated Volumes
Concentration Depth in mL/min EBCT Exhaust up to Exhaust up to qcol
in mg/L (C 0) cm (Z) (Q) (min) in h (e t) in L (Ve) Exhaust (mg/g)

8.65 5 8 4.5749 139 66.72 4247.54 6.965


8.65 10 8 9.1498 162 77.76 2475.18 5.849
8.65 15 8 13.7247 184 88.32 1874.21 4.875
8.65 10 4 18.2995 212 50.88 1619.56 4.841
8.65 10 12 6.0998 134 96.48 3071.06 5.616
4.00 10 8 9.1498 336 161.28 5133.71 4.594
15.0 10 8 9.1498 90 43.20 1375.10 6.163

Source: Modified from Ayoob, S. and Gupta, A.K., Chem. Eng. J., 133, 273–281, 2007.

The breakthrough curves developed for the same bed height (10 cm) at
higher flow rates appeared steeper, which may be due to the faster move-
ment of the adsorption zone along the bed. The breakthrough capacity of
the column (qmin,col) showed a consistent increase, though marginal, with
a reduction in flow rates.

The total adsorption capacity of the column (qcol) also showed an identical
response at 8–12 mL/min. An increase in the adsorption capacity at lower
flow rates is usually expected due to better diffusivity of fluoride, resulting
in enhanced sorption. The kinetic curve also reflects similar trends of sorp-
tion in the empty bed contact time (EBCT = volume of the bed/volumetric
flow rate) ranges corresponding to these flow rates.1
It is also observed that the volumes of water treated at both breakthrough
and exhaust increased more than two times, thus corresponding to a reduc-
tion in initial fluoride concentrations from 8.65 to 4.0 mg/L.

Also, the service times of the column indicates that high fluoride con-
centrations aid quick saturation of the bed and enable faster movement
of the adsorption zone. Since the higher concentration gradient between
the solute in solution and the solute on the sorbent results in enhanced
diffusion and sorption, the adsorption capacity of ALC is also found to
increase at higher initial fluoride concentrations.1

The column performance of ALC renders an average adsorption potential


of 5.896 mg/g at the point of exhaust. Though the column study suggests
an adsorption potential of 5.896 mg/g, the maximum Langmuir monolayer
adsorption capacity of ALC in batch studies was found to be much higher
(10.215 mg/g).1 Theoretically, the adsorption capacities from batch studies
may not give accurate scale-up information about the column operation sys-
tem73; whereas in fixed beds, the adsorption media have not been subjected to
Adsorptive Removal of Fluoride 171

equilibrium sorption conditions and are, hence, not getting totally exhausted
as in the batch system. Also, uneven flow patterns throughout the column
may result in an incomplete exhaustion of bed, as cited earlier. However,
in this study, the range of EBCT provided (~4.5–18.3 min) ensures around
75%–80% of the total removal observed in the batch system is represented
by the kinetic curve.1

8.5.9  Application of Sorption Models


All the six models cited were applied to investigate the breakthrough behav-
ior of fluoride sorption onto ALC in natural groundwater.

The characteristic parameters of the models obtained by linear regres-


sion were used to predict the theoretical effluent fluoride concentrations.
In the BDST model applications, the values of the maximum adsorption
capacity parameter No. (Table 8.8) are found to decrease with increased
bed depths, indicating that the adsorption zone is not moving with a
constant speed along the column.1 The adsorption rate constants (K),
characterizing the rate of solute transfer from liquid to solid phase, were
observed to increase with flow rate and initial fluoride concentrations,
indicating the influence of external mass transfer on system kinetics.
The higher values of K are advantageous, as they indicate that even a
short bed will avoid breakthrough. Theoretically, the slope of the BDST
line represents the time required for the adsorption zone to travel a unit
length through the adsorbent bed.1

The predicted breakthrough times for 4 and 12 mL/min are found to be 80


and 23.33 h (using bt = 8.5Z – 5 for 4 mL/min and bt = 2.833Z – 5 for 12 mL/
min), respectively, with corresponding exhaust times of 202.33 and 145.66 h
(using et = 8.5Z + 117.33 for 4 mL/min and et = 2.833Z + 117.33 for 12 mL/min).1
The time required for the adsorption zone to travel a unit length through

TABLE 8.8
Characteristic Parameters Predicted by BDST Model in Synthetic Water
BDST Model
Initial Fluoride
Concentration Flow Rate K N0
(mg/L) Bed Depth (cm) (mL/min) (L/mg/h) (mg/L) R2 APE

8.65 5 8 0.00482 18573.56 0.9048 38.853


8.65 15 8 0.00357 11794.47 0.9613 17.474
8.65 10 4 0.00251 12574.39 0.8916 22.057
8.65 10 12 0.00489 13653.96 0.9519 20.486
4.00 10 8 0.00323 11039.11 0.8794 27.440
15.0 10 8 0.00456 14760.46 0.9778 16.389
Source: Modified from Ayoob, S. and Gupta, A.K., Chem. Eng. J., 133, 273–281, 2007.
172 Fluoride in Drinking Water

the 10-cm adsorbent bed is as follows: 9.191, 4.25, and 2.4508 h for initial fluo-
ride concentrations of 4, 8.65, and 15 mg/L, respectively. The predicted break-
through times for 4 and 15 mg/L are 86.072 and 20.9814 h, respectively (using
bt = 9.191Z – 5.838 for 4 mg/L and bt = 2.4508Z – 3.5266 for 15 mg/L) with corre-
sponding exhaust times of 345.626 and 92.168 h1 (using et = 9.191Z + 253.326 for
4 mg/L and et = 2.4508Z + 67.66 for 15 mg/L). A comparison of these predicted
service times with the experimental service times indicates that the predicted
service times are more than the corresponding experimental values (Table 8.9).
Also, as per Equation 3.40, a curve between service times and bed heights at
the 50% breakthrough point must result in a straight line passing through the
origin if the process follows this model. In this study, the t50 values for 5-, 10-,
and 15-cm bed depths are 50.24, 108.45, and 146.78 h, respectively. The plot of Z
versus t50 offers a straight line plot (R2 = 0.9858) but quite deviates from the ori-
gin (Figure 8.22). Generally, this failure is attributed to the complexity of the
sorption process.1 So, it is rational to believe that intra particle diffusion and
external mass resistance are considerable and rate limiting in this complex

TABLE 8.9
Comparison of Experimental Service Times with Those Predicted by BDST Model
under Different Process Conditions in Synthetic Water
Time for Breakthrough Time for Exhaust
Initial Fluoride ALC Bed Flow Rate in in Hours (bt) in Hours (e t)
Concentration Depth in mL/min
in mg/L (C 0) cm (Z) (Q) Experimental Predicted Experimental Predicted

8.65 5 8 17 16.25 139 138.58


8.65 15 8 58 58.75 184 181.08
8.65 10 4 84 80 212 202.33
8.65 10 12 20 23.33 134 145.66
4.00 10 8 72 86.072 336 345.062
15.0 10 8 22 20.981 90 92.165

150

100 R2 = 0.98
t50 in h

50

0
4 8 12 16
Z (cm)

FIGURE 8.22
Bed depth versus time for 50% breakthrough plot of fluoride sorption onto ALC fixed bed in
synthetic water.
Adsorptive Removal of Fluoride 173

fluoride sorption process and that their kinetics is controlled not alone by the
surface chemical reactions between fluoride and ALC.5

The linear regression of the Thomas model with the experimental flu-
oride sorption data also shows good correlations in most of the cases
(Table  8.10). As shown, both the parameters qTh and KTh of the model
were found to decrease with higher bed depths. For all conditions of
bed depths, flow rates, and initial fluoride concentrations, the model
predicted a marginally higher sorption capacity qTh than experimental
q0 values. The rate constant (KYN) of the Yoon–Nelson model decreased
with an increase in bed depths but increased with flow rates and initial
fluoride concentrations (Table 8.11). From the experimental results and
data regression, it could be said that the Yoon–Nelson model provided a
good correlation of the sorption of fluoride by ALC in most of the cases.1

The prediction of time for 50% breakthrough correlated well in higher flow
rates and higher concentrations but differed much in other cases.5

TABLE 8.10
Characteristic Parameters Predicted by Thomas Model for Synthetic Water
Thomas Model
Initial Fluoride
Concentration Flow Rate KTh q0 q Th
(mg/L) Bed Depth (cm) (mL/min) (L/mg/h) (mg/g) (mg/g) R2 APE

8.65 5 8 0.0048 6.9652 7.9700 0.9048 38.853


8.65 15 8 0.0035 4.8747 5.0620 0.9613 17.474
8.65 10 4 0.0025 4.8413 5.3969 0.8916 22.057
8.65 10 12 0.0049 5.6160 5.8590 0.9519 20.486
4.00 10 8 0.0032 4.5940 4.7380 0.8794 27.440
15.0 10 8 0.0046 6.1630 6.3350 0.9778 16.389
Source: Modified from Ayoob, S. and Gupta, A.K., Chem. Eng. J., 133, 273–281, 2007.

TABLE 8.11
Characteristic Parameters Predicted Yoon–Nelson Model for Synthetic Water
Yoon–Nelson Model
Initial Fluoride
Concentration Flow Rate KYN
(mg/L) Bed Depth (cm) (mL/min) (h-1) τ (h) τcal (h) R2 APE

8.65 5 8 0.0417 50.240 70.256 0.9048 38.853


8.65 15 8 0.0309 146.78 133.86 0.9613 17.474
8.65 10 4 0.0217 191.72 190.29 0.8916 22.057
8.65 10 12 0.0423 64.330 68.870 0.9519 20.486
4.00 10 8 0.0129 148.00 180.63 0.8794 27.440
15.0 10 8 0.0684 65.20 64.40 0.9778 16.389
Source: Modified from Ayoob, S. and Gupta, A.K., Chem. Eng. J., 133, 273–281, 2007.
174 Fluoride in Drinking Water

Since the fluoride sorption by ALC follows the Freundlich isotherm


model, its constant n was used for evaluating the parameters in Clark
model. The model renders a good fit with comparatively higher R 2 and
lower APE values in most of the experimental conditions (Table 8.12).1

The value of r is observed to increase with flow rates and initial fluoride con-
centrations but it decreases with bed depths. The magnitude of r increased
more than two times, thus corresponding to the increase in flow rates from
4 to 12 mL/min. Similarly, the value of r increased more than two times when
the initial fluoride concentration increased from 4 to 8.65 mg/L; whereas it
was more than five times for 8.65–15 mg/L.1
The Wolborska sorption model was applied to experimental data for
describing the initial part of the breakthrough curve (Ct/C0 ~ 0.5). The model
offers a reasonably good fitting with experimental data at lower flow rates
and bed depths. The values of β and N0 that are obtained from linear regres-
sion of the model are shown in Table 8.13. The values of the kinetic constant
β were much influenced by flow rate and sharply increase with an increase
in flow rates. With an increase in initial fluoride concentrations, they exhibit

TABLE 8.12
Characteristic Parameters Predicted by Clark Model for Synthetic Water
Initial Fluoride Clark Model
Concentration Bed Flow Rate
(mg/L) Depth (cm) (mL/min) ln A r (h-1) R2 APE

8.65 5 8 1.9276 0.0362 0.9290 35.783


8.65 15 8 2.9132 0.0258 0.9635 15.676
8.65 10 4 2.8734 0.0177 0.8623 22.736
8.65 10 12 1.986 0.0374 0.9696 17.140
4.00 10 8 1.4816 0.0113 0.9022 25.456
15.0 10 8 3.109 0.0571 0.9694 18.636
Source: Modified from Ayoob, S. and Gupta, A.K., Chem. Eng. J., 133, 273–281, 2007.

TABLE 8.13
Characteristic Parameters Predicted by Wolborska Model for Synthetic Water
Initial Wolborska Model
Fluoride
Concentration Bed Flow Rate β N0
(mg/L) Depth (cm) (mL/min) (h-1) (mg/L) R2 APE

8.65 5 8 116.907 14303.36 0.9424 23.179


8.65 15 8 40.532 13536.95 0.8948 28.891
8.65 10 4 28.818 15979.27 0.9319 18.441
8.65 10 12 79.155 13399.07 0.8654 29.990
4.00 10 8 49.105 9723.76 0.8939 24.480
15.0 10 8 63.013 17601.35 0.9364 20.274
Adsorptive Removal of Fluoride 175

TABLE 8.14
Characteristic Parameters Predicted by Bohart and Adams Model for Synthetic
Water
Initial Bohart and Adams Model
Fluoride
Concentration Bed Flow Rate K AB N0
(mg/L) Depth (cm) (mL/min) (L/mg/h) (mg/L) R2 APE

8.65 5 8 0.00817 14303.36 0.9424 23.179


8.65 15 8 0.00299 13536.95 0.8948 28.891
8.65 10 4 0.00180 15979.27 0.9319 18.441
8.65 10 12 0.00591 13399.07 0.8654 29.990
4.00 10 8 0.00505 9723.76 0.8939 24.480
15.0 10 8 0.00358 17601.35 0.9364 20.274

identical trends. For an increase in flow rate from 4 to 8 mL/min, the per-
centage increase in β is ~50% and further, for an increase from 8 to 12 mL/
min, it is ~28%. The increased turbulences that developed within the bed
at higher flow rates may reduce the film boundary layer surrounding the
adsorbent, thereby increasing the β values. Practically, this results in faster
breakthrough, considerably reducing the service time of columns, as is evi-
denced at higher flow rates and concentrations. This sorption behavior indi-
cates that the system kinetics of fluoride sorption is dominated by external
mass transfer, especially at initial breakthrough.52 As expected, maximum
adsorption capacity values (N0) were reduced by increased flow rates from
4 to 12 mL/min. The Bohart and Adams model reflects similar features to the
Wolborska sorption model in modeling the experimental data within the low
range of breakthrough applied.5 An identical increase in kinetic rates and
adsorption capacities is also observed, as shown in Table 8.14.

8.5.9.1  Comparison of the Applied Models (Synthetic Water)


In general, the linear fittings of Hutchins BDST, Thomas and Yoon–
Nelson models demonstrate that only the characteristic parameters
associated with these models vary. However, all the three models
will predict essentially the same Ct/C0 values for a particular data
set, and they are bound to give the same APE and R 2 values as illus-
trated earlier. But the prominent and unique characteristic features of
the respective models, such as service time (Hutchins BDST model),
adsorption capacity (Thomas model) and time for 50% breakthrough
(Yoon–Nelson model), enable a further comparison to be made. The
APE values on service time, adsorption capacity and time for a 50%
breakthrough are 7.9952%, 6.1998% and 11.64%, respectively. So, the
predictions by the Thomas model on its characteristic parameters turn
more appropriate (as it renders the least error among the three) fol-
lowed by the Hutchins BDST model and the Yoon–Nelson model.1
176 Fluoride in Drinking Water

TABLE 8.15
Comparison of the Coefficient of Regression Values (R 2) Obtained from Linear
Regression of Clark and Bohart and Adams Model (or Wolborska Model) with
Experimental Data at Different Stages of Fluoride Sorption in Synthetic Water
Breakthrough Point 50% Breakthrough
Initial Fluoride (Up to Ct = 1 mg/L) (Ct /C 0 = 0.5)
Concentration Flow Rate
(mg/L) Bed Depth (cm) (mL/min) CM BAWM CM BAWM

8.65 5 8 0.9244 0.9353 0.9769 0.9424


8.65 15 8 0.9638 0.9527 0.9585 0.8948
8.65 10 4 0.9569 0.9464 0.9536 0.9319
8.65 10 12 0.9603 0.966 0.9351 0.8654
4.00 10 8 0.9446 0.9623 0.9394 0.8939
15.0 10 8 0.9662 0.9628 0.9552 0.9364

Note: CM, Clark model; BAWM, Bohart and Adams (or Wolborska model).

On comparing the R2 and APE values of the Thomas model with those of the
Clark model, it could be seen that the latter correlates marginally better. So,
though the Thomas model could also describe the sorption process fairly well,
the Clark model is deemed to be the best fit in terms of its slightly higher R2 and
lower APE values. The Bohart and Adams model and Wolborska model, plot-
ted against the same axial settings of ln(Ct/C0) against t, turn equivalent when
kAB becomes equal to β/N0. As discussed earlier, both these models will also
render similar APE and R2 values and are equally good in correlating the sorp-
tion process in low breakthrough ranges. To have a meaningful comparison
of the Clark model with the Bohart and Adams (or Wolborska) models, their
correlations were also evaluated at lower breakthrough ranges (Table 8.15). It
can be seen that the Clark model correlates better than the Bohart and Adams
model (or Wolborska model) in all process conditions up to a 50% break-
through; whereas both models are equally good in describing the process up
to the point of breakthrough. So, in practical single-column applications or
unit applications such as DDU, the Bohart and Adams model (or Wolborska
model) can also be used, as it can describe the process up to the point of break-
through fairly well. Thus, it can be concluded that since the Clark model could
describe the process with the same vigor in all sorption ranges, it would be the
most appropriate model to describe the sorption of fluoride onto ALC.1,5

8.5.9.2  Comparison of the Applied Models (Natural Water)


The features of the model are compared as in synthetic water studies, as dis-
cussed earlier. In case of fluoride sorption onto ALC in natural water, the APE
values on service time (Hutchins BDST model), adsorption capacity (Thomas
model), and time for 50% breakthrough (Yoon–Nelson model) are 20.57%,
8.29%, and 18.46%, respectively. So, the predictions by the Thomas model on the
characteristic parameters turn more appropriate followed by the Yoon–Nelson
Adsorptive Removal of Fluoride 177

TABLE 8.16
Comparison of the Coefficient of Regression Values (R 2) Obtained from Linear
Regression of Thomas and Bohart and Adams Model (or Wolborska Model) with
Experimental Data at Different Stages of Fluoride Sorption in Natural Water
50% Breakthrough
Breakthrough Point Point
Initial Fluoride (Up to Ct = 1 mg/L) (Ct /C 0 = 0.5)
Concentration Bed Depth Flow Rate
(mg/L) (cm) (mL/min) TM BAWM TM BAWM
8.65 5 8 0.9063 0.8989 0.9745 0.9601
8.65 10 8 0.9375 0.9437 0.9634 0.9775
8.65 15 8 0.8731 0.8812 0.9391 0.9478
8.65 10 4 0.9507 0.9588 0.9405 0.9500
8.65 10 12 0.9145 0.9161 0.9348 0.9658

and Hutchins BDST models. It was observed that the Clark model fails to fit the
sorption data, rendering poor R2 and high APE values. Thus, it appears that the
Bohart and Adams model (or Wolborska model) could describe the sorption
process fairly well up to a 50% breakthrough, whereas the Thomas model could
describe the process up to exhaust. Further, to differentiate between these mod-
els, their applications at different stages of breakthrough are compared. The
correlation of these models with the experimental fluoride sorption data at
these stages is illustrated in Table 8.16. It becomes very obvious that the Thomas
model could describe the sorption of fluoride onto ALC at all these stages of
the sorption process, whereas the Bohart and Adams model (or Wolborska
model) could be applied only up to low breakthrough ranges. Practically, in
single-­column applications or unit applications such as DDU, both models
may describe the process with the same vigor. However, in pilot plants and
field applications involving series of columns, the Thomas model would be the
most appropriate. The conventional practice of comparing the sorption system
responses with predicted values of the model involves plotting the respective
curves against the same axial settings, usually Ct/C0 versus t1,5 (Figure 8.23).

8.5.10  Fluoride Desorption Studies


The desorption properties of the adsorbent ALC were tested in fixed bed
studies by using an eludent of 10% NaOH in synthetic and natural systems.
Exhausted columns of 5-cm ALC media were regenerated after the single
run of fluoride sorption. The desorption profiles of ALC in both synthetic
and natural water are shown in Figure 8.24. The 5-cm ALC bed sorbed 6.965
mg/g of fluoride, depicted during sorption studies conducted on synthetic
water. The concentration fluoride desorbed in the first bed volume was
2350 mg/L, which got reduced thereafter and remained almost constant
at around 4.2 mg/L after 20 bed volumes. On calculation, the total fluo-
ride desorbed was found to be 3.729 mg/g only given that the percentage
178 Fluoride in Drinking Water

8 5 cm, 8 mL/min (exp


––––
Equilibrium concentration Ce (mg/L)

++
7 ++++ –––– 5 cm, 8 mL/min (mod)
+ ––
++ 10 cm, 8 mL/min (exp)
6 + –– 10 cm, 8 mL/min (mod)
–– 15 cm, 8 mL/min (exp)
5 15 cm, 8 mL/min (mod)
+ ––
4 + 10 cm, 12 mL/min (exp)
10 cm, 12 mL/min (mod)
3 –– – 10 cm, 4 mL/min (exp)
+
10 cm, 4 mL/min (mod)
2 ––
point of breakthrough
+ –– ––
1 –
––– point of exhaust
++
– – – ––
0 +
++ – – – –– –
0 20 40 60
t (h)

FIGURE 8.23
Comparison of experimental breakthrough profiles against theoretical values predicted by
Thomas model under different process conditions in natural water. (From Ayoob, S., Gupta,
A.K. and Basheer, A.B., J. Urban Environ. Eng., 3, 17–22, 2009).

2500
2250 Natural water
Fluoride desorbed (mg/L)

2000
1750 Synthetic water
1500
1250
1000
750
500
250
0
0 5 10 15 20
Bed volumes treated

FIGURE 8.24
Fluoride desorption profile of ALC during regeneration with 10% NaOH in column studies.

of fluoride desorbed was 53.54%. Almost identical desorption profiles were


observed for the ALC bed that sorbed 0.7503 mg/g of fluoride from natural
water. Though the fluoride desorbed in the first bed volume was 250 mg/L, it
was 2.15 mg/L after 20 bed volumes. The total fluoride desorbed was found
to be 0.3571 mg/g, rendering the percentage of fluoride desorbed as 47.60%.
In both systems, even after 20 bed volumes, the concentration of fluoride in
the effluent remained high, indicating a very slow desorption profile. The
poor desorptional characteristics may indicate a chemisorptive mechanism
through a strong chemical interaction between fluoride and ALC.
Adsorptive Removal of Fluoride 179

This may indicate the irreversibility of fluoride sorption, which further


suggests that ion exchange is not the only sorption mechanism involved
in fluoride removal. It is also plausible that the fluoride ions that ini-
tially attached to ALC by ligand- or ion-exchange mechanisms may
become more firmly bound by chemisorption. The chemical compounds
formed due to sorption as indicated by the XRD, coupled with these poor
desorption characteristics of ALC, further demonstrate that such trans-
formations might have taken place.3

8.6  Summary of the Case Study


The world is in a water crisis. The presence of geogenic pollutants such as
fluoride in groundwater exacerbates the gravity of this issue, thereby posing
hindrance to the global onward march of humanity toward a water secure
world. Of late, fluoride in drinking water has been wreaking havoc in more
than 35 nations, forcing hundreds of millions of people to live in the shad-
ows of fluorosis. Blame it on different forms of fluorosis, in many countries
such as India and China, its excess presence in water is prioritized as a major
impediment to the sustainable drinking water supply. As a result, defluori-
dation of drinking water has been regarded as one of the key areas of atten-
tion among the global water community, triggering global research. Since
majority of the affected people worldwide live in small communities, more
emphasis should be given to developing treatment technologies tailored
for such habitations. Adsorption is an important technique that is the most
widely used for excess fluoride removal in such habitats of the developing
world. Though many adsorbents had been developed for defluoridation,
only very few of them such as activated alumina were found to be useful in
field applications. So, the development of economically viable and socially
acceptable adsorbents with adequate fluoride removal potential has become
a necessity. The adsorbent ALC developed in this direction was tested for
its applicability as an adsorbent for defluoridation. The experimental inves-
tigations comprised a series of batch and continuous flow experiments
performed under varying operating conditions. The results of such experi-
mental observations and major findings of the research can be summarized
as follows:

• Batch study results demonstrate the feasibility of ALC in removing


fluoride from aqueous systems. It offers comparable defluoridation
potential in both natural and synthetic systems. The kinetic curve of
fluoride sorption exhibited a biphasic uptake with a rapid phase in
the initial few minutes followed by a slow phase in both the systems.
The slow phase, which governs the rate limiting, has been gener-
ally ascribed to diffusion into micropores, retention on sites of lower
reactivity, and could also be due to surface nucleation/precipitation.
180 Fluoride in Drinking Water

• At optimum conditions, an ALC dose of 10 g/L could remove


94.54% of fluoride from natural water with an initial fluoride con-
centration of 8.65 mg/L, whereas a lesser dose of 2 g/L could remove
92.76% fluoride in synthetic waters of the same concentration. The
fluoride removal percentage decreased from 96.76% to 89.56% with
an increase in initial fluoride concentrations from 2.5 to 50 mg/L;
whereas, the adsorption capacity increased from 1.2 to 22.39 mg/g.
This increase in adsorption capacity in the synthetic system at higher
fluoride concentrations, signifying heterogeneous sorptive surfaces,
may be ascribed to high intramolecular competitiveness to occupy
the unsaturated lower energetic surface sites.
• The XRD results reveal multiple oxide surface sites, which are sug-
gestive of sorption occurring on heterogeneous binding sites. This
is further confirmed by an increase in sorption density with higher
fluoride concentrations. Further, the variation in the affinity of ALC
to different adsorbent doses indicates a surface-bound sorption.
Since the adsorbent is a composite, its surface may be composed
of sites with a spectrum of binding energies. It is postulated that
at low adsorbent dosages all types of sites are entirely exposed to
sorption, and the surface may become saturated faster. However,
at higher adsorbent dosages, the availability of higher energy sites
may decrease and a larger fraction of lower energy sites may become
occupied. This results in an overall decrease in binding energy of
the surface, and a reversible type of process may exist between the
fluoride ions attached to low energy sites and those present in bulk
solution. This may be the reason for the observed increase in uptake
with an increase in adsorbent dosages up to a certain stage.
• The fitting of the kinetic data in both systems demonstrates that the
dynamics of sorption could be well described by the pseudo-second-
order model, which is indicative of rate-limiting chemisorption. The
reasonable fitting of the kinetic data to the pseudo-first-order model,
intra particle diffusion model, and Elovich model shows the pos-
sibilities of a diffusion-controlled yet complex sorption mechanism.
The values of activation energy in natural and synthetic systems fur-
ther support the role of diffusion in the sorption process.
• The experimental evidence pertaining to the response of the system
to pH, inert electrolyte concentration, and desorption pattern suggests
that the slow sorption phase in fluoride removal is not the result of dif-
fusion (neither inter particle nor intra particle). Further, the rate limit-
ing is due either to the heterogeneity of the surface site-binding energy
or to other reactions controlling fluoride removal from the solution.
• The analysis of FTIR data together with an observed increase in pH
during sorption suggests that a ligand-exchange mechanism is prom-
inent in the removal process. The poor desorptional characteristics
Adsorptive Removal of Fluoride 181

of ALC indicate the irreversibility of fluoride sorption, which may


further suggest that ion exchange is not the only sorption mecha-
nism involved. The insensitivity of inert electrolyte concentrations to
fluoride uptake suggests the possibilities of fluoride forming inner-
sphere complexes with ALC. It is plausible that the fluoride ions ini-
tially attached to ALC by ligand- or ion-exchange mechanisms may
become more firmly bound by chemisorption. The chemical com-
pounds formed due to sorption (as indicated by the XRD) coupled
with poor desorption characteristics of ALC further demonstrate
that such transformations are plausible. Overall, the removal of fluo-
ride may be viewed as a sorption process in which fluoride is attached
to ALC through an inner-sphere complex formation through strong
bonds (adsorption), diffusion into the crystal structure (absorption),
and various surface precipitation reactions.
• The equilibrium data of fluoride sorption onto ALC conducted at
different concentration ranges of fluoride could fit better with the
Langmuir isotherm model than the Freundlich and DR isotherm
models. At all concentration ranges tested, LI offered an excellent
fit with very high consistent values of R 2 (>0.995). The maximum
saturated monolayer capacity of ALC in synthetic water recorded
an increase from 24.57 to 34.36 mg/g, corresponding to an increase
in the initial fluoride concentration ranges from 2.5 to 20 mg/L and
from 2.5 to 100 mg/L. The dose variation study rendered the max-
imum saturated monolayer capacity of ALC in synthetic water as
10.215 mg/g and in natural water as 0.9358 mg/g. So, ALC exhibited
a reduction in adsorption capacity in treating natural groundwater
compared with synthetic water.
• The adsorption capacity parameters of all isotherm models recorded
an increase at high temperatures, indicating that the sorption of flu-
oride by ALC is endothermic, which is characterized by a chemi-
sorption mechanism. The sorption of fluoride onto ALC was almost
consistent in the pH range of 3–11.5 and was reduced thereafter.
Also, the fluoride removal process was unaffected by ionic strength,
indicating that the removal of fluoride occurs mainly through the
formation of the inner-sphere surface complexation process.
• The endothermic nature of the process in both systems is further
confirmed by the negative enthalpy values. The increase in fluoride
sorption with an increase in temperature in both systems reflects
the surface heterogeneity of ALC and its increased activity, which
results in enhanced diffusion of fluoride ions into its pores. It would
be expected that higher temperatures would stimulate the surface
reactivity of the bound oxides/hydroxides, which increases the sorp-
tion capacity of the system. The positive value of entropy change
reflects the affinity of ALC toward fluoride, which may also indicate
182 Fluoride in Drinking Water

some structural changes within the adsorbent. The negative values


of free energy confirm the spontaneity of sorption in both systems.
• The presence of nitrates, chlorides, sulfates, and bicarbonates did not
significantly affect the sorption process. This may suggest that high
levels of salinity and hardness in water did not affect the fluoride
removal performance of ALC. Since fluoride-rich groundwaters are
generally associated with high bicarbonate ions in alkaline environ-
ments, this may enhance the scope of ALC applications. The effect
of silicate turns significant only after 25 mg/L, and at 400 mg/L it
reduces fluoride removal by around 13%. The presence of iron also did
not affect the removal up to 10 mg/L. At a concentration of 4 mg/L,
the presence of phosphate reduces fluoride removal by around 6%
and at 8 mg/L, by more than 10%. The presence of 40 mg/L of humic
acid reduces fluoride removal by 18.03%. The reduction of fluoride
sorption in the presence of high silicates may be due to its scavenging
of aluminum ions, forming aluminosilicate solute species in the alka-
line environment, and due to its polymerization, thereby resulting
in an increase in negative surface charge. The reduction in fluoride
removal efficiency in the presence of phosphates may be due to the
strong affinity of aluminum for phosphate, reducing its availability
for fluoride uptake. The interference pattern indicates that the pres-
ence of ion chlorides, nitrates, and bicarbonates may form outer-
sphere complexes but sulfates and silicates form partial inner-sphere
complexes with ALC. Aqueous NOM-metal complexes may, in turn,
associate strongly with dissolved anions such as fluoride, presum-
ably by metal-bridging mechanisms, diminishing the tendencies of
fluoride ions to form surface complexes. This metal bridging appears
to be a potential mechanism for reduced fluoride uptake in natural
water. However, further experiments on groundwater samples from
different locations and possessing a different chemistry are needed
to properly elucidate this mechanism.
• The performance of ALC in column studies in synthetic water ren-
dered an average adsorption potential of 5.896 mg/g at the point of
exhaust. Also, the adsorption capacity of ALC at different bed depths
for a particular flow rate showed only marginal variation. The break-
through and total capacities of the column showed a consistent
increase, though marginal, with a reduction in flow rates. The volumes
of water treated at breakthrough and exhaust were reduced consider-
ably with an increase in initial fluoride concentrations. High fluoride
concentrations may aid quick saturation of the bed and enable faster
movement of the adsorption zone. As the higher concentration gradi-
ent between the solute in solution and solute on the sorbent results in
enhanced diffusion and sorption, the adsorption capacity of ALC is
also found to increase at higher initial fluoride concentrations.
Adsorptive Removal of Fluoride 183

• Though the BDST model rendered reasonable fitting with the experi-
mental data, the model fails to make fair predictions of service times
in most of the cases. Further, the predicted service times are gen-
erally more than the corresponding experimental values. Also, the
curve between service times and bed heights at a 50% breakthrough
point was found to not pass through the origin, indicating its fail-
ure in accurately describing the sorption of fluoride onto ALC. This
failure may be attributed to the complexity of the sorption process.
So, it is plausible that the intra particle diffusion and external mass
resistance are considerable and rate limiting in this complex fluoride
sorption process and that its kinetics is controlled not alone by the
surface chemical reactions between fluoride and ALC.
• The linear regression of the Thomas model with the experimental fluo-
ride sorption data showed very good correlations in most of the cases.
Moreover, for all conditions of bed depths, flow rates, and initial fluo-
ride concentrations, the model predicted the sorption capacity fairly
well. Though the Yoon–Nelson model provided good correlations, the
prediction of time for a 50% breakthrough was inaccurate in most of
the cases. The Clark model rendered a good fit with comparatively
higher R2 and lower APE values in most of the experimental condi-
tions. The Bohart and Adams model reflected similar features to the
Wolborska sorption model in modeling the experimental data within
the low range of the breakthrough applied. Though the Thomas model
could also describe the sorption process fairly well, the Clark model
is deemed to be the best fit in terms of slightly higher R2 values and
lower APE values. It was also observed that the Clark model correlated
better than the Bohart and Adams model (or Wolborska model) at all
process conditions up to a 50% breakthrough; whereas both models
were equally good in describing the process up to breakthrough.
• The performance of ALC in column studies with natural water ren-
dered an average adsorption potential of 0.7064 mg/g at the point of
exhaust. Also, as observed in synthetic water, the adsorption capac-
ity of ALC at different bed depths for a particular flow rate showed
only marginal variation. The breakthrough and total capacities of
the column showed a consistent increase, though marginal, with a
reduction in flow rates, thus indicating its consistency in affinity for
fluoride sorption. The breakthrough curves appear steeper at higher
flow rates; this may be due to faster movement of the adsorption
zone along the bed, aiding its quick saturation. The observed reduc-
tion in column adsorption capacity in both natural and synthetic
systems compared with its corresponding batch performance is rea-
sonable, as batch studies may not give accurate scale-up information
about the column operation system. In fixed beds, the adsorption
media have not been subjected to equilibrium sorption conditions
184 Fluoride in Drinking Water

and, hence, are not getting totally exhausted as in the batch sorp-
tion system. Also, uneven flow patterns throughout the column may
result in an incomplete exhaustion of bed.
• The BDST model fairly predicted the service times of the column in
natural water up to breakthrough, but it failed to correlate exhaust
times in a better way. So, the model could fit well only with initial
portions of the experimental breakthrough curve. Also, as in syn-
thetic water, the curve between service times and bed heights at a 50%
breakthrough point was found to not pass through the origin, indi-
cating its failure in accurately describing the sorption process. This
failure may be attributed to the complexity of the sorption process,
indicating that the kinetics of the sorption process in natural water
is controlled not alone by the surface chemical reactions between
fluoride and ALC. The linear regression of the Thomas model with
the experimental fluoride sorption data in natural water also showed
very good correlations in most of the cases. Moreover, for all condi-
tions of bed depths, flow rates and initial fluoride concentrations, the
model predicted the sorption capacity fairly well. Though the Yoon–
Nelson model also provided good correlations, the prediction of time
for a 50% breakthrough was inaccurate in most of the cases. However,
Clark model rendered a poor fit with the fluoride sorption data with
comparatively low R2 and higher APE values in all ranges of experi-
mental conditions. The Bohart and Adams model reflected similar
features to Wolborska sorption models in modeling the experimental
data within the low range of breakthrough applied. Altogether, the
Thomas model was demonstrated successful in describing the sorp-
tion of fluoride onto ALC at all stages of the sorption process, whereas
the Bohart and Adams model (or Wolborska model) could be applied
only up to low breakthrough ranges.

8.7  Conclusions of the Case Study


The synthesis of the ALC sorptive system for defluoridation of water has
been convincingly demonstrated successful, as it showed comparable and
consistent fluoride sorption potential in both batch and column applica-
tions. The capacity of the adsorbent in bringing the fluoride concentrations
within the permissible limits over a wide range of pH without sacrificing its
potential may be useful in treating natural fluoride-rich groundwaters. The
removal may be viewed as a complex sorption process, occurring via multiple
mechanisms, including electrostatic attraction, anion and ligand exchange,
and other chemical reactions. The consistency in fluoride sorption over a
wide pH range, availability and adaptability in the local environment, and
Adsorptive Removal of Fluoride 185

comparable sorption potential of ALC in both synthetic and natural systems


deserve merit. This study describes in detail the systematic procedures and
protocols to be adopted in carrying out a study on the adsorptive removal of
fluoride. The detailed investigations performed through batch and column
studies convincingly demonstrate the feasibility of using ALC for fluoride
removal in household and community applications in the endemic areas.

References
1. Ayoob, S. and Gupta, A.K. (2007). Sorptive response profile of an adsorbent in
the defluoridation of drinking water. Chem. Eng. J., 133, 273–281.
2. Ayoob, S. and Gupta, A.K. (2009). Performance evaluation of alumina cement
granules in removing fluoride from natural and synthetic waters. Chem. Eng.
J., 150, 485–491.
3. Ayoob, S., Gupta, A. K., Bhakat, P.B. and Bhat, V.T., (2008). Investigations on the
kinetics and mechanisms of sorptive removal of fluoride from water using alu-
mina cement granules. Chem. Eng. J., 140, 6–14.
4. Ayoob S. and Gupta A.K. (2008). Insights into isotherm making in the sorptive
removal of fluoride from drinking water. J. Hazard. Mater., 152, 976–985.
5. Ayoob, S., Gupta, A.K. and Basheer, A.B. (2009). A fixed bed sorption system for
defluoridation of ground water. J. Urban Environ. Eng., 3, 17–22.
6. APHA. (1998). American Public Health Association, Standard Methods for the
Examination of Water and Wastewater, 20th edn. Washington, DC: American
Public Health Association.
7. Das, D.P., Das, J. and Parida, K. (2003). Physicochemical characterization and
adsorption behavior of calcined Zn/Al hydrotalcite-like compound (HTlc)
towards removal of fluoride from aqueous solution. J. Colloid Interface Sci., 261,
213–220.
8. UNICEF. (1999). UNICEF Water Front, Fluoride in Water: An Overview, p. 11. New
York: UNICEF programme division, WES section, three United Nations plazas.
9. WHO. (2002). Fluorides, Environmental Health Criteria Number, 227. Geneva,
Switzerland: World Health Organization.
10. Ayoob, S., Gupta, A. K., Bhakat, P.B. (2007). Analysis of breakthrough develop-
ments and modeling of fixed bed adsorption system for As(V) removal from
water by modified calcined bauxite (MCB). Sep. Purif. Technol., 52, 430–438.
11. Ho, Y.S. (2006). Review of second order models for adsorption systems. J. Hazard.
Mater., 136, 681–689.
12. Ho, Y.S., McKay, G. (1998). A comparison of chemisorption kinetic models
applied to pollutant removal on various sorbents. Trans. IChemE, 76, 332–340.
13. Lagergren, S. (1898). About theory of so-called adsorption of soluble substances.
K. Svenska Ventenskapsakad Handlingar, 24, 1–39.
14. Aharoni, C. and Sparks, D.L. (1991). Kinetics of soil chemical reactions: A theo-
retical treatment. In: Sparks, D.L. and Suarez, D.L., (Eds.), Rates of Soil Chemical
Processes, SSSA special publication no. 27, pp. 1–18. Madison, WI: Soil Science
Society of America.
186 Fluoride in Drinking Water

15. Crini, G., Peindy, H.N., Gimbert, F. and Robert, C. (2007). Removal of C.I. Basic
Green 4 (Malachite Green) from aqueous solutions by adsorption using cyclo-
dextrin-based adsorbent: Kinetic and equilibrium studies. Sep. Purif. Technol.,
53, 97–110.
16. Maliyekkal, S.M., Sharma, A.K. and Philip, L. (2006). Manganese-oxide-coated
alumina: A promising sorbent for defluoridation of water. Water Res., 40,
3497–3506.
17. Ghorai, S. and Pant, K.K. (2005). Equilibrium, kinetics and breakthrough stud-
ies for adsorption of fluoride on activated alumina. Sep. Purif. Technol., 42,
265–271.
18. Weber, W.J. and Morris, J.C. (1963). Kinetics of adsorption on carbon from solu-
tion. J. Sanit. Eng. Div. ASCE, 89, 31–39.
19. Zhang, J. and Stanforth, R. (2005). Slow adsorption reaction between arsenic
species and goethite (α-FeOOH): Diffusion or heterogeneous surface reaction
control. Langmuir, 21, 2895–2901.
20. Hingston, F.J. (1981). A review of anion adsorption. In: Anderson, M.A. and
Rubin, A. J., (Eds.), Adsorption of Inorganics at the Solid-Liquid Interface, p. 67. Ann
Arbor, MI: Ann Arbor Science.
21. Aharoni, C., Sparks, D.L., Levinson, S. and Revina, I. (1991). Kinetics of soil
chemical reactions: Relationships between empirical equations and diffusion
models. Soil Sci. Soc. Am. J., 55, 1307–1312.
22. Rudzinksi, W. and Panczyk, P. (1998). In: Schwarz, J.A. and Contescu, C.I., (Eds).,
Surfaces of Nanoparticles and Porous Materials, p. 355. New York: Dekker.
23. Sparks, D.L. (2005). Sorption-desorption, kinetics. In: Hillel, D, Hatfield, J.L.,
Powlson, D.S., Rosenweig, C., Scow, K.M., Singer M.J. and Sparks, D.L. (Eds.),
Encyclopedia of Soils in the Environment. pp. 556–561. Oxford, UK: Elsevier Ltd.
24. Adamson, A.W. and Gast, A.P. (1997). Physical Chemistry of Surfaces, 6th edn.
New York: Wiley-Interscience.
25. Langmuir, I. (1916). The constitution and fundamental properties of solids and
liquids. J. Am. Chem. Soc., 38, 2221–2295.
26. Freundlich, H.M.F. (1906). Über die adsorption in lösungen. Z. Phys. Chem., 57,
385–470.
27. Dubinin, M.M. (1960). The potential theory of adsorption of gases and vapors
for adsorbents with energetically non-uniform surface. Chem. Rev. 60, 235–266.
28. Ceyhan, O. and Baybas, D. (2001). Adsorption of some textile dyes by hexadecy-
ltrimethylammonium bentonite. Turk. J. Chem., 25, 193–200.
29. Wolkenstein, T. (1991). Electronic Processes on Semiconductor Surfaces during,
Chemisorption. New York: Consultants Bureau.
30. Kinnilburgh, D.G. (1986). General purpose adsorption isotherms. Environ. Sci.
Technol., 20, 895–904.
31. Qadeer, R., Hanif, J., Khan, M. and Saleem, M. (1995). Uptake of uranium ions
by molecular sieve. Radiochim. Acta, 68, 197–201.
32. Ozcan, A., Oncu, E.M. and Ozcan, A.S. (2006). Kinetics, isotherm and thermo-
dynamic studies of adsorption of Acid Blue 193 from aqueous solutions onto
natural sepiolite. Colloids Surf. A: Physicochem. Eng. Aspects, 277, 90–97.
33. Kasprzyk-Hordern, B. (2004). Chemistry of alumina, reactions in aqueous solu-
tions and its application in water treatment. Adv. Colloid Interface Sci., 110, 19–48.
34. Weber, W.J. Jr. (1972). Physicochemical Processes for Water Quality Control. New
York: A Wiley-Inter science Publication, John Wiley and Sons.
Adsorptive Removal of Fluoride 187

35. Kumar, K.V. and Sivanesan, S. (2006). Pseudo second order kinetics and pseudo
isotherms for malachite green onto activated carbon: Comparison of linear and
non-linear regression methods. J. Hazard. Mater., 136, 721–726.
36. Hao, J.O. and Huang, C.P. (1986). Adsorption characteristics of fluoride onto
hydrous alumina. J. Environ. Eng. (ASCE), 112, 1054–1069.
37. Ayoob, S., Gupta, A.K and Bhat, V.T. (2008). A conceptual overview on sustain-
able technologies for the defluoridation of drinking water. Crit. Rev. Environ.
Sci. Technol., 38, 401–470.
38. Onyango, M.S., Kojima, Y., Aoyi, O., Bernardo, E.C. and Matsuda, H. (2004).
Adsorption equilibrium modeling and solution chemistry dependence of flu-
oride removal from water by trivalent-cation-exchanged zeolite F-9. J. Colloid
Interface Sci., 279, 341–350.
39. Das, N., Pattanaik, P. and Das, R. (2005). Defluoridation of drinking water using
activated titanium rich bauxite. J. Colloid Interface Sci., 292, 1–10.
40. Tripathy, S.S., Bersillon, J.L. and Gopal, K. (2006). Removal of fluoride from
drinking water by adsorption onto alum-impregnated activated alumina. Sep.
Purif. Technol., 50, 310–317.
41. Handa, B.K. (1975). Geochemistry and genesis of fluoride-containing ground
waters in India. Ground Water, 13, 278–281.
42. Thomas, W.J. and Crittenden, B.D (1998). Adsorption Technology and Design.
Oxford, UK: Reed Educational and Professional Publishing.
43. Schiewer, S. and Volesky, B. (1997). Ionic strength and electrostatic effects in bio-
sorption of divalent metal ions and protons. Environ. Sci. Technol., 31, 2478–2485.
44. Gao, Y. and Mucci, A. (2001). Acid base reactions, phosphate and arsenate com-
plexation, and their competitive adsorption at the surface of goethite in 0.7 M
NaCl solution. Geochim. Cosmochim. Acta, 65, 2361–2378.
45. Kundu, S. and Gupta, A.K. (2007). As(III) removal from aqueous medium in
fixed bed using iron oxide-coated cement (IOCC): Experimental and modeling
studies. Chem. Eng. J., 129, 123–131.
46. Hutchins, R.A. (1973). New simplified design of activated carbon system. Am.
J. Chem. Eng., 80, 133–138.
47. Thomas, H.C. (1944). Heterogeneous ion exchange in a flowing system. J. Am.
Chem. Soc., 66, 1664–1666.
48. Yoon, Y.H. and Nelson, J.H. (1984). Application of gas adsorption kinetics. I. A theo-
retical model for respirator cartridge service life. Am. Ind. Hyg. Assoc. J., 45, 509–516.
49. Clark, R.M. (1987). Evaluating the cost and performance of field-scale granular
activated carbon systems. Environ. Sci. Technol., 21, 573–580.
50. Wolborska, A. (1989). Adsorption on activated carbon of p-nitrophenol from
aqueous solution. Water Res., 23, 85–91.
51. Bohart, G.S and Adams, E.Q. (1920). Some aspects of the behavior of charcoal
with respect to chlorine. J. Am. Chem. Soc., 42, 523–544.
52. Aksu, Z. and Gönen, F. (2004). Biosorption of phenol by immobilized activated
sludge in a continuous packed bed: Prediction of breakthrough curves. Process
Biochem., 39, 599–613.
53. McKay, G., Otterburn, M.S. and Sweeny, A.G. (1981). Surface mass transfer
­processes during colour removal from effluent using silica. Water Res., 15, 327–331.
54. Zhu, M.X., Xie, M. and Jiang, X. (2006). Interaction of fluoride with hydroxyalu-
minum-montmorillonite complexes and implications for fluoride-contaminated
acidic soils. Appl. Geochem., 21, 675–683.
188 Fluoride in Drinking Water

55. Chen, J.P. and Wang, L. (2004). Characterization of metal adsorption kinetic
properties in batch and fixed-bed reactors. Chemosphere, 54, 397–404.
56. Kannan, N. and Meenakshisundaram, M. (2002). Adsorption of Congo red on
various activated carbons. A comparative study. Water Air Soil Pollut., 138, 289–305.
57. Acemioglu, B. (2005). Batch kinetic study of sorption of methylene blue by per-
lite. Chem. Eng. J., 106, 73–81.
58. Pavlatou, A. and Polyzopoulos, N.A. (1988). The role of diffusion in the kinetics
of phosphate desorption: The relevance of the Elovich equation. Eur. J. Soil Sci.,
39, 425–436.
59. Chen, J.P. and Lin, M.S. (2001). Equilibrium and kinetic of metal ion adsorption
onto a commercial h-type granular activated carbon: Experimental and model-
ing studies. Water Res., 35, 2385–2394.
60. Stollenwerk, K.G. (2003). Geochemical processes controlling transport of arse-
nic in ground water: A review of adsorption. In: Welch, A. H. and Stollenwerk,
K. G., (Eds.), Arsenic in Ground Water, pp. 67–100. Boston, MA: Kluwer Academic
Publishers.
61. Sposito, G. (1984). The Surface Chemistry of Soils. New York: Oxford University
Press.
62. Richter, E., Wilfried, S. and Myers, A.L. (1989). Effeect of adsorption equation
on prediction of multicomponent adsorption equilibria by the ideal adsorbed
solution theory. Chem. Eng. Sci., 44, 1609–1616.
63. Genc-Fuhrman, H., Tjell, J.C. and Mcconchie, D. (2004). Adsorption of arse-
nic from water using activated neutralized red mud. Environ. Sci. Technol., 38,
2428–2434.
64. Altundogan, H.S., Altundogan, S., Tumen, F. and Bildik, M. (2000). Arsenic
removal from aqueous solutions by adsorption on red mud. Waste Manage., 20,
761–767.
65. Thurman, E.M. (1985). Organic Geochemistry of Natural Waters, p. 497. Dordrecht,
the Netherlands: Kluwer Publishers.
66. Fan, L., Harris, J.L., Roddick, F.A. and Booker, N.A. (2001). Influence of the
characteristics of natural organic matter on the fouling of microfiltration mem-
branes. Water Res., 35, 4455–4463.
67. Xu, H., Allard, B. and Grimvall, A. (1991). Effects of acidification and natural
organic matter on the mobility of arsenic in the environment. Water Air Soil
Pollut., 57, 269–278.
68. Warwick, P., Inam, E. and Evans, N. (2005). Arsenic’s interaction with humic
Acid. Environ. Chem., 2, 119–124.
69. Zhang, Y., Yang, M. and Huang, X. (2003). Arsenic(V) removal with a Ce(IV)-
doped iron oxide adsorbent. Chemosphere, 51, 945–952.
70. Courtijn, E., Vandecasteele, C. and Dams, R. (1990). Speciation of aluminum in
surface water. Sci. Total Environ., 90, 191–202.
71. Redman, A.D., Macalady, D.L. and Ahmann, D. (2002). Natural organic matter
affects arsenic speciation and sorption onto hematite. Environ. Sci. Technol., 36,
2889–2896.
72. Moreno-Castilla, C. (2004). Adsorption of organic molecules from aqueous
solutions on carbon materials. Carbon, 42, 83–94.
73. Benefield, D.L., Judkins, F.J. and Weand, L.B. (1982). Process Chemistry for Water
and Waste Water Treatment. Englewood Cliffs, NJ: Prentice-Hall, Inc.
Water Science & Engineering

Explore the Health Effects of Fluoride Pollution

Fluoride in Drinking Water: Status, Issues, and Solutions establishes the negative
impacts of naturally occurring fluoride on human health and considers the depth and
scope of fluoride pollution on an international scale. The book discusses current global
water quality and fluoride-related issues and draws overall awareness to the problems
associated with fluoride in drinking water.

Utilizing recent scientific studies to examine the current status of fluoride pollution,
it provides a fundamental understanding of fluorosis, describes health problems
associated with fluorosis, and discusses viable scientific solutions. The book places
special emphasis on India, Africa, China, and other countries deeply affected by
fluoride pollution.

A single, comprehensive source covering health issues related to fluoride and its effect
on humans, this book:

• Compiles information from scientific literature on the state of fluoride pollution


• Characterizes the human impacts of fluorosis
• Provides a comparative evaluation of technologies used for defluoridation
• Gives a comprehensive account of human health effects with appropriate
scientific descriptions and photographs
• Includes detailed descriptions on the geochemistry of fluoride entry into
groundwater aquifers
• Presents a case study that deals with the successful removal of fluoride
from drinking water

A vital resource for environmental and public health officials as well as academic
researchers in the area, Fluoride in Drinking Water: Status, Issues, and Solutions
covers human health issues associated with fluoride-rich water and describes relevant
techniques for defluoridation that can be used to overcome the stress, issues, and
challenges of natural fluoride in drinking water.

K27531
6000 Broken Sound Parkway, NW
Suite 300, Boca Raton, FL 33487 ISBN: 978-1-4987-5652-5
711 Third Avenue 90000
an informa business New York, NY 10017
2 Park Square, Milton Park
www.crcpress.com Abingdon, Oxon OX14 4RN, UK
9 781498 756525

w w w.crcpress.com

You might also like