You are on page 1of 10

Removing arsenic from groundwater for the developing

Water Science and Technology Vol 44 No 6 pp 89–98 © IWA Publishing 2001


world – a review
J.-Q. Jiang
Department of Civil Engineering, University of Surrey, Guildford, Surrey GU2 7XH, UK

Abstract This paper is principally concerned with summarising the experience to date of treating arsenic
containing ground/surface water by oxidation, coagulation/precipitation and adsorption processes. Arsenic
(As) has been verified through epidemiological evidence as one of the most carcinogenic and toxic
substances in surface and ground water. Oxidation, coagulation/precipitation, and adsorption have been
widely used in arsenic removal and the study results demonstrated that these technologies can remove
arsenic from ground/surface water efficiently; the residual arsenic concentration in the effluent could be in
the range of 5–10 µg/l, against the influent arsenic concentration in the range of 10–500 µg/l. However,
these technologies need to be surveyed in order to validate the efficiency, cost and maintenance requirments
by considering social and economic situations and the availability of the local resources in the developing
world.
Keywords Adsorption; arsenic (As); As chemistry; As removal technologies; developing world; oxidation;
precipitation–coagulation

Introduction
Although arsenic is of long legendary toxicity, the presence of trace (micrograms per litre)
arsenic levels in public drinking water supplies has not previously cause alarm. Arsenic is
both toxic and carcinogenic. Inorganic forms of arsenic dissolved in drinking water are the
most significant forms of natural exposure. Organic forms of arsenic that may be present in
food are much less toxic to humans. Clinical evidences of arsenic poisoning begin with var-
ious forms of skin disease, and proceed via damage to internal organs ultimately to cancer
and death. The symptoms of chronic arsenic poisoning may take between five and fifteen
years to reveal themselves. Recent studies have indicated that arsenic in drinking water is
more dangerous than previously suspected (Smith et al., 1992). The World Health
Organisation (WHO, 1996), and the US Environmental Protection Agency (USEPA, 1993)
have set up more stringent arsenic regulations to minimise these risks.
International concern has also focused on the problem of arsenic in groundwater in
Bangladesh and Bengal, India, because of the large scale of the population who are affected
adversely by high arsenic concentrations. In Bangladesh, where drinking water is mainly
from groundwater, geochemical conditions favour arsenic dissolution. High concentra-
tions (0.5–5 mg/l) are reported from many districts, and British Geological Survey (BGS)
estimates that 25–30 million are now using this water for drinking, cooking and other
domestic purposes (BGS, 1999). Similarly, in Bengal, India, nearly 30 million people in the
state are exposed to arsenic in the concentration range of 0.2–2.0 mg/l due to the consump-
tion of groundwater for drinking purposes (PHED, 1991). The Bangladesh and India
Standard for arsenic in drinking water is 0.05 mg/l, which was based on WHO advice at the
time when the regulations were drafted, and the new WHO provisional guideline for As
(0.01 mg/l, WHO, 1996) has not been adopted in either Bangladesh or India.
A variety of treatment processes has been used for arsenic removal from water. The
most commonly used technologies for the developing world include oxidation, coagula-
tion–precipitation, and adsorption with activated carbon, activated alumina, and 89

Downloaded from https://iwaponline.com/wst/article-pdf/44/6/89/39044/89.pdf


by Tufts University user
on 01 August 2018
iron-oxide-coated materials (e.g., Song and Logsdon, 1978). These processes are most
promising but require validation for efficiency, sustainability, adaptation and adoption
using locally available resources. This paper therefore aims to compare the effectiveness
and limitations of these arsenic removal technologies and investigates their practicality for
the developing world.

Arsenic chemistry
J.-Q. Jiang

In natural waters, the soluble arsenic species are As3+ (arsenite) and As5+ (arsenate).
Although both organic and inorganic forms have been detected, organic forms (methylated
arsenic) are seldom present at concentration greater than 1 µg/l and are generally consid-
ered of less toxicity in comparison with the inorganic species in drinking water treatment.
Thermodynamic predictions provide a useful understanding of the equilibrium
chemistry of the inorganic arsenic species in water (Figure 1). In oxygenated waters As5+ is
dominant, existing in anionic forms of either H2AsO4–, HAsO42–, or AsO43– over the pH
range typically encountered in water treatment.
Under anoxic conditions, As3+ is stable, with arsenous acid (H3AsO3) and anionic
(H2AsO–) species dominant below and above pH 9, respectively. In the presence of sul-
phides, precipitation of AsS (realgar) or As2S3 (orpiment) may remove soluble As3+ and
exert considerable control over trace arsenic concentrations (Ferguson and Garvis, 1972).
Although thermodynamics can provide an accurate prediction of possible changes in a
given non-equilibrium condition, they give no insight to the rate at which those conditions
will occur. In general As3+ and As5+ acid-base reactions can be assumed to occur instanta-
neously, whereas time periods for the changes between oxidation states are indeterminate
in natural waters. For instance, the conversion of As3+ to As5+ in oxygenated water is
thermodynamically favoured, yet the rate of the transformation may take days, weeks or
months, depending on the specific conditions. Strong acidic or alkaline solutions, the
presence of copper salts, carbon, unknown catalysts and higher temperatures can increase
the oxidation rate (Johnson and Pilson, 1972). The existence of manganese oxide, chlorine,
permanganate and other oxidants can directly transform As3+ to As5+ in the absence of
oxygen (e.g., Frank and Clifford, 1986). The reduction of As5+ to As3+ in the absence
of oxygen is also chemically slow and may require bacterial mediation.
Arsenic can be immobilized through adsorption-coprecipitation with iron and
manganese hydroxides, mobilised when such solids are dissolved under reducing condi-
tions, or released from the precipitate surfaces in the event of competition (for sorptive sur-

Figure 1 Diagram EH-pH for arsenic (25°C) EH = oxidation potential, [As] = 10–5 M, [S]=10–3 M, () = solid
90 state, (Reproduced from Dojilido and Best, 1993)

Downloaded from https://iwaponline.com/wst/article-pdf/44/6/89/39044/89.pdf


by Tufts University user
on 01 August 2018
face sites) in the presence of orthophosphate and natural organic matter (NOM) (e.g.,
Peryea, 1991; Xu et al., 1991). The latter factors can explain why arsenic may be correlated
to high Mn (II) (dissolved and reduced manganese oxide), Fe (II) (dissolved and reduced
Fe(OH)3), and orthophosphate (competition with arsenic for adsorption sites) in certain
waters. (e.g., Peryea, 1991).
The above reactions can occur in both groundwater and surface water sources. For each
water source, there are three distinct zones, namely, aerobic, anoxic without sulphide,

J.-Q. Jiang
anoxic with sulphides. Within oxygenated zones, As5+ is stable and may remain soluble or
sorb-coprecipitate with iron and manganese oxides if present. High concentrations of
orthophosphate may compete with As5+ for adsorption sites in the zone, increasing soluble
arsenic concentrations and mobility. In anoxic regimes without sulphides, As3+ is stable,
and dissolved forms of iron and manganese are favoured. Arsenic mobility (or solubility) is
highest in this zone because (a) As3+ is believed to sorb less strongly onto oxides than As5+
and (b) coprecipitated-sorbed arsenic is released upon dissolution of arsenic-containing Fe
and Mn oxides. In anoxic zones with sulphides, As3+ becomes immobilised because of the
formation of orpiment (As2S3), realgar (As2S2), or is coprecipitated with iron pyrite.

Evaluating arsenic removal technologies


Arsenic removal technologies generally fall into three major classes: chemical coagula-
tion–precipitation, adsorption and membrane separation. The purpose of the oxidation is to
oxidize the As3+ to As5+ and this favours coagulation–precipitation and adsorption.
Membrane technologies have been used for arsenic removal in the developed countries
(e.g., Song and Logsdam, 1978) but have not been shown to be feasible for the developing
world. This section thus compares the technologies of oxidation, coagulation–precipitation
and adsorption only for the performance of arsenic removal, and discusses whether these
technologies could be adopted in the developing world.

Oxidation of As3+ to As5+


Since As5+ is readily removed by chemical coagulation and adsorption, complete oxidation
of As3+ to As5+ is required if these technologies are adapted. There are numerous methods
of oxidation available, but important factors such as the residual concentrations, availabili-
ty and running cost of oxidants and oxidation by-products should be considered when
selecting an oxidant. Ozone and UV radiation oxidise As3+ normally within about 30–60
seconds. However, it requires a high energy input, which may not be cost effective for small
water treatment plants in the developing world. Chlorine is commonly used in the water
industry as a disinfectant and is a suitable oxidant. However, the drawback in the use of
chlorine is that it will react with natural organic matter to form chlorinated by-products,
which are believed to have health concern effects. In addition, it is not so effective for As3+
oxidation if the raw water has high turbidity. The most feasible oxidants to date are either
potassium permanganate or Fenton’s reagent. These oxidants produce no harmful by-
products, and are easily applied.

Coagulation–precipitation process
Coagulation with metal coagulants like alum (Al2(SO4)3) and ferric chloride (FeCl3) has
long been recognized as an effective method for removal of arsenic. The efficiency of
arsenic removal with FeCl3 and alum was examined in bench-scale studies (Hering et al.,
1996; 1997). As5+ removal by either FeCl3 or alum was relatively insensitive to variations
in source water composition below pH 8. At pH 8 and 9, the efficiency of As5+ removal by
ferric chloride was decreased in the presence of natural organic matter. The pH range for
arsenic removal with alum was more restricted than with FeCl3. For source waters spiked 91

Downloaded from https://iwaponline.com/wst/article-pdf/44/6/89/39044/89.pdf


by Tufts University user
on 01 August 2018
with 20 mg/l As5+, final dissolved As5+ concentrations in the product water of less than 2 mg /l
were achieved with both coagulants at neutral pH. Removal of As3+ from source waters by
FeCl3 was both less efficient and more strongly influenced by source water composition
than removal of As5+. The effect of coagulation pH on the As removal percentage over wide
FeCl3 doses is shown in Figure 2, where the influent As concentration was 0.1 mg/l.
Increasing coagulation pH decreased As removal percentage and the optimum coagulation
pH for As removal with ferric chloride was 7.
J.-Q. Jiang

Water samples collected from six sources in West Bengal were studied extensively for
arsenic removal with coagulation–flocculation (Pande et al., 1997). The results have
shown that a dose of 3.0 mg/l of chlorine (for pre-chlorination) followed by 50 mg/l of
FeCl3 was able to remove arsenic from the raw waters. Arsenic removal can also be facili-
tated by a variety of solids formed during softening including CaCO3, Mg(OH)2, Mn(OH)2,
and Fe(OH)3 (e.g., Hering et al., 1997). The extent of As5+ removal is decreased in the pres-
ence of orthophosphate and carbonate. As3+ removal is much lower than As5+ removal. At
typical solids concentrations, arsenic removal followed a linear isotherm for CaCO3,
Mg(OH)2, and Fe(OH)3 with constant percentage arsenic removal regardless of initial
arsenic concentrations. However, for Mg(OH)2 solids arsenic removal was sensitive to
arsenic concentrations.
The possible use of enhanced coagulation (the term is defined as the optimum coagula-
tion performance could be achieved by modifying the coagulation pH and coagulant dose)
for arsenic removal was examined at bench, pilot, and demonstration scales with two
source waters (Cheng et al., 1994). Alum and FeCl3, with cationic polymer, were investi-
gated at various influent arsenic concentrations. For the source waters tested, enhanced
coagulation could be effective for arsenic removal and that less FeCl3 than alum, on a
weight basis, is needed to achieve the same removal percentage.
Arsenic removal efficiencies from the full-scale plants were evaluated (e.g., Scott et al.,
1995). Raw water with As3+ 0.6–2 mg/l was prechlorinated with 3–5 mg/l of chlorine and
coagulated with 3–10 mg/l of FeCl3 or 6–10 mg/l of Al2(SO4)3. Arsenic removal was
81–96% with FeCl3 and 23–71% with Al2(SO4)3. The best results were obtained with
4 mg/l of FeCl3 at a filtration rate 150 m3/m2/day and up to 70 runs. The full-scale study
results also suggest that arsenic removal is best achieved by FeCl3 instead of Al2(SO4)3
coagulation. This is consistent with that from the bench-scale results.

9.5
60%
70%
80%
9 90%

8.5

8
pH

7.5
100%

6.5

6
5 10 15 20 25 30 35 40
Dose (mg/l as FeCl 3)

92 Figure 2 The effect of coagulation pH on the As removal

Downloaded from https://iwaponline.com/wst/article-pdf/44/6/89/39044/89.pdf


by Tufts University user
on 01 August 2018
Arsenic removal by adsorption
Adsorption is a traditional process to separate solutes from solvent or gases, where the
solute increases on the surface of the adsorbent and decreases in the solvent. Adsorption
equilibrium is achieved when the number of molecules leaving the surface of the adsorbent
is equal to the number of molecules being adsorbed on the surface of the adsorbent.
Effective adsorbents have a highly porous structure so that their surface area to volume
ratio is very large. Adsorbents used for arsenic removal include activated alumina, activat-

J.-Q. Jiang
ed carbon, and iron-oxide-coated sand.

Activated alumina (AA)


Granular activated alumina is a commercial porous aluminium oxide with specific surface
areas of 200 to 300 m2/g, has high pH of zero-point-of-charge (zpc) (pHzpc = 8.2) and has
the following selectivity sequence:

OH– > H2AsO4 > H3SiO4– > F– > SO42– > HCO3– >Cl– >NO3–

AA is a promising adsorbent for arsenic removal. It can be used in the fixed bed technique
and can be regenerated with diluted NaOH and H2SO4. The sorption equilibria of As5+ on
activated alumina were measured by Rosenblum and Clifford (1984). The best pH value for
As5+ adsorption was 5.5–6 because the alumina is protonate, but the anions of the acid
added (to lower the pH) are not yet competitive in arsenic adsorption. The adsorption
capacities were about 5–15 mg/g at equilibrium concentrations of 0.02–0.1 mg/l. The exis-
tence of sulphate and chloride ions can reduce the adsorption capacity by up to 50%. As3+ is
difficult to be removed by AA, but As5+ is readily to be adsorbed and its breakthrough may
occur at 14,000 or 24,000 times the bed volumes (pH 6) against the effluent arsenic concen-
tration of 20 mg/l and 50 mg/l, respectively (Figure 3).
The above findings were confirmed at pilot-scale studies (Rubel and Hathaway, 1987),
where the best arsenic removal was achieved at pH 5.5 in the down-flow mode; with up to
15,000 times the bed volumes against the breakthrough concentration of 20 mg/l. Using 4 to
5% NaOH cannot achieve the completed regeneration, and 20% of the arsenic remained in
the adsorbed. Operation under non-optimal conditions reduced the run time of the adsorber
by about 50%.
The competitive effects of other anions, such as phosphate or fluoride, will reduce the
capacity for As5+ adsorption considerably. This may be of particular concern if the com-
petitors are present in the range 0.1–2 mg/l and arsenic has to be removed down to the mg/l
range.

50 µg/l

20 µg/l

Figure 3 Comparison of breakthrough curves for As5+ and As3+ in filters of activated alumina (reproduced
from Rosenblum and Clifford, 1984) 93

Downloaded from https://iwaponline.com/wst/article-pdf/44/6/89/39044/89.pdf


by Tufts University user
on 01 August 2018
The available study results suggest that AA can be applied successfully for removing
arsenic if the influent pH is slightly acidic (e.g., pH 6) and the competing anions are present
in small concentrations (e.g., < 0.1 mg/l). The great advantage of AA is its simple operation
over one to three months before regeneration is required, making it more feasible for small-
scale requirements. However, after regeneration, the run length of AA is only at 75% of the
virgin run length (Simms and Azizian, 1997). Another disadvantage of AA adsorption is
both sodium hydroxide and sulphuric acid should be used in regeneration process and
J.-Q. Jiang

the spent arsenic contaminated alkaline and acid wastewaters require treatment and
disposal of.

Granular activated carbon (GAC)


GAC is very common used for drinking water treatment and wastewater treatment in some
cases. GAC has the following characteristics, particle size is 0.5–2 mm, surface area is
around 1,000 m2/g and apparent density is 0.4–0.5 g/m3. Factors influencing the GAC per-
formance for arsenic removal could be pH, the type and properties of GAC, and pretreat-
ment of GAC with various metals.
The effect of pH on arsenic adsorption capacities of the activated carbon has been inves-
tigated. The maximum As5+ loading could be achieved at pH values of 5 to 7 (Figure 4), and
pH adjustment can control arsenic adsorption and arsenic stripping from GAC (Gupta and
Chen, 1978; Lee and Rosehart, 1972). The type of activated carbon also influences the
arsenic adsorption capacities. Table 1 shows the arsenic loading of three types of activated
carbon, and the contact period with the arsenic solution was 24 hours. Both the peat-based
carbon and the coal-based carbon performed better than coconut-shell carbon due to their
high ash content. Higher arsenic loading by carbon with high ash content were contributed
to interact between the mineral matter (metal oxides and metal ions) and arsenic in the car-
bon. The results in Table 1 also show that with an increase in the specific surface area of the
carbon, there is a decrease in arsenic loading, indicating that surface area of the carbon
plays minor role on arsenic adsorption.
The adsorption of arsenic by GAC can also be improved by pre-treating the carbon with
metal ions (e.g., Rajakoviv, 1992). This can be performed by contacting the activated car-
bon with various metal ion solutions. In order to increase metal ion loading on the carbon,
the alkalinity of the solution can be raised with the addition of solution hydroxide until the
metal hydroxide precipitates onto the carbon surface. A large number of metal ions have
been selected for pretreatment. The arsenic adsorption capacities of various pretreated acti-
vated carbons are shown in Table 2. The pH values presented are that at which the maxi-
mum arsenic adsorption occurred.

6
Untreated C
As loading onto GAC

5
Pretreated C
(mg As/mg C)

0
0 2 4 6 8 10 12 14
pH
94 Figure 4 The effect of influent pH on the GAC performance for As removal

Downloaded from https://iwaponline.com/wst/article-pdf/44/6/89/39044/89.pdf


by Tufts University user
on 01 August 2018
Table 1 The effect of carbon type on the adsorption of arsenic onto activated carbon

The type of activated carbon Ash content (%) Surface area (m2/g) Arsenic loading (mg/g carbon)

Coconut-shell carbon 3.0 1,200 2.40


Peat-based extruded carbon 5.0 975 4.91
Coal-based carbon 5.5 1,125 4.09

J.-Q. Jiang
Table 2 The effect of the pretreatment of coconut-shell carbon with metal ions on arsenic loading

Pretreatment metal pH Arsenic loading (mg As/g C) Expected compound

None 2.4
Ba 8.0 1.87 BaHAsO4.H2O
Ca(CN) 7.7 1.60 Ca3(AsO4)2
Co 8.0 2.51 Co3(AsO4)2 8H2O
Cu 6.0 5.79 Cu2AsO4OH
Mg 9.0 1.88 Mg3(AsO4)2 10H2O
Ni 8.0 2.80 Ni3(AsO4)2 8H2O
Pb 6.5 3.19 Pb(AsO4)3 Cl
Fe(II) 5.0 3.69 Fe3(AsO4)2
Fe(III) 5.0 4.53 Fe(AsO4)2 x Fe(OH)3
Zn 8.0 2.56 Zn2AsO4OH

From Table 2, it can be seen that pretreatment with copper and ferric salts resulted in
greater arsenic loading than that with untreated activated carbon. Arsenic loading follow-
ing copper pretreatment with and without precipitation of the metal hydroxide was identi-
cal (5.79 mg As/g carbon). In the case of Fe(III), arsenic loading dropped from 4.53 mg
As/g carbon with precipitation to 4.09 mg As/g carbon without precipitation. The effect of
different anions in the copper salts used for the carbon pretreatment was also investigated.
Cuprous chloride resulted in the greatest arsenic loading. Arsenic loading by pretreated
peat-based carbon was also higher than that of pretreated coconut-shell carbon (7.41 mg/g
carbon for the former and 5.79 mg/g carbon for the latter). In addition, arsenic desorption
can be achieved using either strong acidic or alkaline solutions. Acidic solutions are more
effective and the copper is also eluted.

Iron-oxide-coated sands (IOCS)


In recent years, iron-oxide-coated sand (IOCS) (e.g., Joshi and Chuadhuri, 1996) has been
prepared and investigated for its performance to remove arsenic. Its characteristics are
shown in Table 3. It can be seen that IOCS has high surface area and PZC, and the better
adsorption performance is expected in comparison with the plain sand.
Column tests have been undertaken using 75 g of iron-oxide-coated sand in an 11 mm
internal diameter glass column (bed depth 525 mm; porosity 0.36) and a flow rate of
1 ml/min to assess the potential of iron-oxide-coated sand as a adsorbent for removing
arsenic from ground water. Background batch sorption kinetic tests indicated that the bulk
of the arsenic sorption on iron-oxide-coated sand occurred during 30–60 min of contact. A
column was run until the effluent arsenic level was well above 0.01 mg/l (the WHO provi-
sional guideline value for arsenic in drinking water). The medium was then regenerated by
backwashing with 2 litres of a 0.2N sodium hydroxide solution. This was followed by back-
washing with distilled water until the influent and effluent pH was comparable. This consti-
tuted one cycle of the column test. The spent regenerant and backwash water were analyzed
for the recovered arsenic.
Breakthrough empty-bed volumes at arsenic concentration of 0.01 mg/l were in
the range 163–184 and 149–165 per cycle for As3+ and As5+, respectively. During 95

Downloaded from https://iwaponline.com/wst/article-pdf/44/6/89/39044/89.pdf


by Tufts University user
on 01 August 2018
Table 3 Characteristics of plain sand and IOCS (Benjamin et al. 1996)

Plain sand IOCS

Fe salt used for coating None Fe(NO3)3


% Fe by weight 0 2.4
Surface area by BET (m2/g) 0.04 5.8
pH of the PZC 0.7 9.3
J.-Q. Jiang

regeneration, 94–99% of the removed arsenic was recovered in each cycle. Leaching of
iron from the medium was not detected in the effluent. The results indicated promise of
IOCS as an adsorbent for use in small systems for removing arsenic from groundwater.
Spent catalysts are generated in various industries (e.g., petrochemical industry). After
2–3 regenerations, they are to be disposed of as solid wastes with significant amount, and
have been studied as adsorbents as they consist of porous silica and alumina. The effective-
ness of pretreating spent catalysts with iron solutions to improve arsenic removal has been
studied (Huang and Liu, 1997). The results indicated that iron-coated spent catalysts could
be used as adsorbents for removing As5+ from aqueous solutions. The arsenic adsorption
capability of spent catalysts was improved by coating iron with 0.01 M Fe(III) solution, and
the arsenic adsorption increases with decreasing pH (the optimum pH is at 5). The As5+ is
proposed to be specifically adsorbed onto the iron-coated spent catalysts and the adsorption
capability is comparable to those of other adsorbents.

Arsenic removal technologies for developing world


In many developing countries, increase in population, poverty and economic stagnation have
left the society with economic problems. In these countries, strategies for development
including opening up of new lands, development of new industries and agriculture as means
to solve these problems have increased the demand of water resources for agricultural, indus-
trial and drinking water supply. Many countries promoted the use of groundwater, which usu-
ally has advantages over surface water in various aspects, however, the quality of
groundwater is largely determined by geologocal conditions and sometimes it turns out to
containing various susbstances harmful to human health. Arsenic (As) in groundwater
has become a serious problem in a number of developing countries. A lot of technologies have
been developed to remove As from groundwaters. Among which, pre-oxidation,

Table 4 Comparative arsenic removal performance

Treatment technology Influent arsenic Effluent arsenic Percentage


(µg/l) (µg/l) removal

Coagulation–precipitation
Oxidation/iron coagulation 110 10–85 23–91
Oxidation/iron coagulation 200 80 60
Oxidation/iron coagulation 1100 5 99
Lime softening 220 30 86
Lime softening (pH>10.5) 75 5 95
Lime softening 100 5 95
Iron coagulation 50 5 90
Iron coagulation 21 2 90
Iron coagulation 377 11 95
Adsorption
Activated alumina 88 <50 >43
Activated alumina 103 <50 >51
Activated carbon 300 10 96
Iron oxide coated materials 250 10 96
96

Downloaded from https://iwaponline.com/wst/article-pdf/44/6/89/39044/89.pdf


by Tufts University user
on 01 August 2018
coagulation–precipitation, and adsorption could be the cost-effective technologies for
removing arsenic for the developing world. The comparative As removal performances have
been summarised in Table 4, where the arsenic concentration in the effluent could be less than
5–10 mg/l by treatment with pre-oxidation and iron coagulation/precipitation, and adsorption.
A future study should be concerned with a survey of various household technologies for
As removal in developing world, in order to validate their efficiency, cost and maintenance
requirments by considering social and economy situations and the availability of the local

J.-Q. Jiang
resources. Such household technology would provide a low-cost technology for use, prin-
cipally by the rural poor; this must therefore be most efficient and robust and require mini-
mal maintenance.

Conclusions
International concern has been focused on the problem of arsenic in groundwater in some
developing countries because of the large scale of the population who are adversely affect-
ed by high arsenic concentrations.
The inorganic form of arsenic in water is the result of dissolution from the solid phase
and is present in stable oxidation states of As5+ (arsenate) and As3+ (arsenite). Arsenate
predominates in oxygenated surface waters in three forms: H2AsO4–, HAsO42–, or AsO43–.
Arsenite is favoured under reducing conditions (anaerobic groundwater), and its species is
a function of pH and redox potential.
Pre-oxidation, coagulation–precipitation, and adsorption could be the cost-effective
technologies for removing arsenic for the developing world. By treatment with pre-oxida-
tion and iron coagulation/precipitation, the arsenic concentration in the effluent could be
less than 5 mg/l. Adsorption with AA, GAC and IOCS can also remove arsenic efficiently
and the arsenic concentration in the treated water could be less than 10 µg/l. However, using
these technologies in groundwater remediation has not been consistently demonstrated.
Therefore, a survey to evaluate various household technologies for the arsenic removal and
for use by the rural poor is necessary and such technologies should be most efficient with
low cost, sustainability, adaptation and adoption using locally available reagents, and
requires minimum maintenance.

References
Benjamin, M.M., Sletten, R.S., Bailey, R.P. and Bennett, T. (1996). Sorption and filtration of metals using
iron-oxide-coated sand. Water Research, 30, 2609–20.
British Geological Survey (1999). Groundwater studies for arsenic contamination in Bangladesh, Phase 1
report. BGS, London.
Cheng, R.C., Wang, C.C., Beuhler, M.D. (1994). Enhanced coagulation for arsenic removal. Jour. AWWA,
86, 79.
Dojilido, J.R. and Best, G.A. (1993). Chemistry of Water and Water Pollution. Ellis Horwood (London),
p. 1461.
Ferguson, J.F. and Garvis, J.A. (1972). Review of the arsenic cycle in natural waters. Water Resources, 6,
1259.
Frank, P. and Clifford, D. (1986). Arsenic (III) oxidation and removal from drinking water. USEPA (REP)
EPA-600-52-86/021.
Gupta, S.K. and Chen, K.Y. (1978). Arsenic adsorption by adsorption. Jour. Wat. Pollut. Control Fed., 50,
493.
Hering, J.G., Chen, P.Y., Wilkie, J.A. and Elimelech, M. (1997). Arsenic removal from drinking water
during coagulation. Jour. Environmental Eng., 123, 155.
Hering, J.G., Chen, P.Y., Wilkie, J.A., Elimelech, M., and Liang, S. (1996). Arsenic removal by ferric
chloride. Jour. AWWA, 88, 155.
Huang, J.G. and Liu, J.C. (1997). Enhanced removal of As(V) from water with iron-coated spent catalyst.
Sep. Sci. Technol., 32, 1557.
97

Downloaded from https://iwaponline.com/wst/article-pdf/44/6/89/39044/89.pdf


by Tufts University user
on 01 August 2018
Johnson, D.L. and Pilson, M.E.Q. (1972). Spectrophotometric determination of arsenic and phosphatye in
natural waters. Analytical Chim. Acta, 58, 289.
Joshi, A. and Chuadhuri, M. (1996). Removal of arsenic from ground water by iron oxide-coated sand. Jour.
Environ. Eng., ASCE, 122, 769.
Korte, N. (1991). Naturally occurring arsenic in groundwaters of the Midwestern United States.
Environmental Geology Water Sci., 18, 137.
Lee, J.Y. and Rosehart, R.G. (1972). Effective methods of arsenic removals from gold mine wastes. 4th
J.-Q. Jiang

Annual Meeting of the Canadian Mineral Processes. Ottawa, Canada, 107.


USEPA Life Systems (1993). Draft Drinking Water Criteria Documents on As. USEPA Contract 68-C8-
0033.
Pande, S.P., Deshpande, L.S., Patni, P.M. and Lutade, S.L. (1997). Arsenic removal studies in some ground
waters of West Bengal, India. Jour. Environ. Sci. Health Part A – Environ. Sci. and Eng. and Toxic and
Hazardous Substance Control, 32, 1981.
Peryea, F.J. (1991). Phosphate-induced release of arsenic from soils contaminated with lead arsenate. Soil
Science Society American Jour., 55, 1301.
Public Health Engineering Department (PHED), India (1991). Final Report of Arsenic Investigation Project.
Government of West Bengal, India.
Rajakoviv, L.V. (1992). The sorption of arsenic onto activated carbon impregnated with metallic silver and
copper. Sep. Sci. Technol., 27, 1423.
Rosenblum, E. and Clifford, D. (1984). The equilibrium arsenic capacity of activated alumina. US EPA
Report, EPA-600/S2-83-107.
Rubel, F. and Hathaway, S.W. (1987). Pilot study for the removals of arsenic from drinking water at the
Fallon, Nevada and Naval Air Station. Jour. AWWA. 79, 61.
Scott, K.N., Green, J.F., Do, H.D. and Mclean, S.J. (1995). Arsenic removal by coagulation. Jour. AWWA,
87, 114.
Simms, J. and Azizian, F. (1997). Pilot-plant trials on removal of arsenic from potable water using activated
alumina. Proceedings of 1997 AWWA Water Quality Technology conference, Denver.
Smith, A.H. et al. (1992). Cancer risks from As in drinking water. Environmental Health Persp., 97, 259.
Sorg, T.J. and Logsdon, G.S. (1978). Treatment technology to meet the interim primary drinking water regu-
lations for inorganics: Part 2. Jour. AWWA, 80, 56.
Waychunas, G.A. et al. (1993). Surface chemistry of ferrihydrite: Part 1, EXAFS studies of the geometry of
coprecipitated and adsorbed arsenate. Geochim. Cosmochim. Acta, 57, 2251.
WHO (1996). Guildlines for Drinking Water Quality, Vol. 2, Second Edition: Health Criteria and Other
Supporting Information. WHO, Geneva.
Xu, H., Allard, B. and Grimvall, A. (1991). Effects of acidification and natural organic materials on the
mobility of arsenic in the environment water. Air Soil Pollution, 57–58, 269.

98

Downloaded from https://iwaponline.com/wst/article-pdf/44/6/89/39044/89.pdf


by Tufts University user
on 01 August 2018

You might also like