Sie sind auf Seite 1von 75

Accepted Manuscript

Title: A review of technologies for manganese removal from


wastewaters

Author: Deepti S. Patil Sanjay M. Chavan John U. Kennedy


Oubagaranadin

PII: S2213-3437(15)30073-7
DOI: http://dx.doi.org/doi:10.1016/j.jece.2015.11.028
Reference: JECE 866

To appear in:

Received date: 2-9-2015


Revised date: 28-10-2015
Accepted date: 19-11-2015

Please cite this article as: Deepti S.Patil, Sanjay M.Chavan, John
U.Kennedy Oubagaranadin, A review of technologies for manganese
removal from wastewaters, Journal of Environmental Chemical Engineering
http://dx.doi.org/10.1016/j.jece.2015.11.028

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
A review of technologies for manganese removal
from wastewaters

Deepti S. Patil a, Sanjay M. Chavan a, John U. Kennedy Oubagaranadin b,*

a
Department of Chemical Engineering, Sinhgad College of Engineering,
Pune – 411041, Maharashtra, India
b
Department of Ceramic and Cement Technology, PDA College of Engineering,
Gulbarga – 585102, Karnataka, India

* Corresponding author. E-mail: ju_kennedy@yahoo.co.in


Tel.: +91 9449638617 Fax: +91 8472 255685

1
ABSTRACT
Today, heavy metal pollution is of prime concern from the environmental point of view.
Among the heavy metals, manganese is a very common contaminant of wastewater and ground
water, as it is a significant raw material for industry. This paper reviews various technologies
that have been used to treat manganese containing wastewater and ground water, such as
chemical precipitation, coagulation, flotation, ion-exchange, oxidation/filtration, electrochemical
treatment, adsorption, and membrane filtration. Recovery of manganese is also very imperative
due to its economical and environmental aspects.
Keywords: Manganese, Removal efficiency, Recovery, Wastewater.

2
1. Introduction
Manganese (Mn), a reddish-gray metal is the twelfth most abundant element with 0.095%
estimated concentration in the Earth’s crust. The name for this element is derived from the Latin
word, magnes which means magnet, referring to the magnetic properties of its ore pyrolusite.
Manganese has an atomic number of 25 with atomic weight 54.938 and electron configuration
4s23d7. It belongs to 4th period, d-block and the sub group VIIB of the periodic table. It exists in
four allotropic modifications: alpha-, beta-, gamma- and delta forms [1].
Speciation of manganese: Speciation of an element is the determination of the individual
physicochemical forms of that element which together make up its total concentration in a
sample. Speciation is sometimes necessary to understand the fate, toxicity, and biological
activity of trace metals and their counter ions in natural waters [2-3]. The key significance of Mn
speciation in relation to drinking water is that both Mn(II) and Mn(VII) are soluble in water
while Mn(IV) is not. Techniques used for speciation of Mn may be divided into three main
groups: physical separation, chemical speciation, and computer modeling. Physical separation
techniques are used to separate ‘soluble’ from ‘insoluble’ forms of Mn. The procedures for
physically dividing the forms of Mn into operationally-defined size classes are by ultrafiltration
of water through different filter sizes and fractionation using two filtration steps. By this method,
three relevant fractions of Mn species can be quantified: particulate, colloidal, and truly
dissolved. Chemical speciation measurements of a metal require selective analytical techniques.
Very few techniques can respond to only one particular species of an element in solution, and
analytical techniques for chemical speciation of Mn are limited. Ion selective electrodes may be
constructed that respond only to the Mn(II)(aq). The sensitivity is limited (~0.5 mg/L) and the
method is subject to interferences, and such electrodes are not generally available commercially.
Anodic stripping voltametry is used to determine the complexation capacity of natural waters for
manganese. Computer modeling to determine Mn speciation based on solubility relationships is
possible, but may not bear resemblance to reality. Modeling is the most useful when applied to
speciation of Mn under equilibrium conditions.
Oxidation states of manganese: Manganese is a transition metal. Although the elemental
(metal) form of manganese does not occur naturally in the environment, manganese is a
component of over 100 minerals. The most common mineral forms include manganese dioxide,
manganese carbonate, and manganese silicate [4]. It can exist in a range of oxidation states, +2,

3
+3, +4, +6 and +7. The divalent form, Mn(II), predominates in most waters at pH 4-7, but more
highly oxidized forms may occur at higher pH values or result from microbial oxidation [5]. If
the pH and redox potential of the water are increased (by the addition of lime and chlorine), then
Mn(II) may be oxidized to Mn(IV). Although Mn can exist in up to 11 oxidation states, Mn(II)
and Mn(III) are the most relevant for biological systems. The oxidation state of Mn appears to be
a key determinant of its distribution, accumulation and excretion. While Mn(II) is rapidly cleared
from the blood and efficiently excreted in the bile, Mn(III) is transported across membranes and
has a slower elimination rate. So the exchange and equilibrium of states from Mn(II) (which is
absorbed) to Mn(III) is an important factor in the distribution of Mn in the body and hence for
any neurotoxic effect. Other oxidation states are possible but Mn(III) is only stable as a complex,
Mn(V) and Mn(VI) are not stable in neutral solutions and Mn(VII), in the form of the purple
permanganate ion (MnO4-), is strongly oxidizing and does not form in most natural waters [6].
Uses of manganese: Manganese is mostly used in ferrous metallurgy apart from
chemical, electrochemical, food and pharmaceutical applications. Ferro-manganese alloys are
used in steel manufacturing. Manganese serves as a deoxidizer of molten steel and controls its
sulfur content. It also enhances strength and hardness of the alloy and its resistance to corrosion.
It is used in high-temperature, stainless and manganese steels and in nickel-chromium and
manganese-aluminum alloys. It is also used as a catalyst. Its salts have various applications in
oxidation, catalysis and medicine. It is an essential element for plants and animals. Its deficiency
can cause deformity of bones in animals [1].
Environmental effects and standards of manganese: The excessive use of manganese
containing products, mine waters, steel manufacturing industries lead to environmental pollution.
Although it is essential for human life, at levels exceeding 0.1 mg/L, manganese in water
supplies causes an undesirable taste in beverages and stains sanitary ware and laundry. The
presence of manganese in drinking water may lead to its bio accumulation [1]. The primary
target of manganese toxicity is the nervous system, and common symptoms of toxic exposure
include ataxia, dementia, anxiety, a ‘mask-like’ face, and manganism, a syndrome similar to
Parkinson’s disease. These effects, when observed, are generally the result of very high
exposures through inhalation, as might occur in an industrial setting, and are not seen among the
general population exposed to low or moderate manganese levels [4,5,7,8]. Even at a
concentration of 0.2 mg/L, manganese will often form a layer on pipes, which may slough off as

4
a black precipitate [9]. According to Indian Standards for Drinking Water Specification IS
10500- 2012, it also gives a taste and forms a brownish color [10]. The health-based value of 0.4
mg/L for manganese is higher than this acceptability threshold of 0.1 mg/L. According to
European Community Directive 98/83/EC Annex 1 Part C-indicator parameters, parametric
value for Mn is 50 μg/L [9]. US Environmental Protection Agency and the EU Directive have
established 0.005 mg/L as the maximum manganese concentration level in domestic water
supplies [5]. Indian standard code IS 10500 has set the maximum permissible limit of Mn in
effluents as 2.0 mg/L [12]. The drinking water standards for various countries are given in Table
1 [11].

Due to the natural discharge of manganese into the environment by the weathering of
manganese-rich rocks and sediments, manganese occurs universally at low levels in soil, water,
air, and food. Recycling of Mn is a decent alternative to recover and reduce environmental
pollution. The objective of this review is to provide a recent literature survey relating Mn
removal from wastewater / groundwater and recovery from various wastes.
2. Treatment techniques for manganese removal
Various recent and comprehensive review papers have presented different
physicochemical treatment techniques for the removal of heavy metals from wastewaters [13-
22]. This review article deals with various techniques used over the period for the removal of
manganese ions from wastewaters and wastes. These techniques are precipitation, coagulation /
flocculation, ion-exchange, oxidation-filtration, electrochemical operations, biological
operations, adsorption and membrane processes. A comprehensive approach has been followed
to discuss all significant work done in this field to date, and a final evaluation has been made to
conclude on the most efficient technique to remove manganese ions.
2.1. Precipitation
Chemical precipitation is a very efficient and widely used method for heavy metals from
various wastewaters. This technique involves conversion of soluble manganous ion to an
insoluble precipitate and to achieve significant manganese reduction by precipitation, a reaction
pH above 9.4 is required [23].

5
2.1.1. Hydroxide precipitation
The precipitation of metals in solution as metal hydroxides is the most common way to
remove metals from solutions in hydrometallurgical processes. Based on the stability constants,
the selectivity of Mn2+ from Mg2+ and Ca2+ can be expressed by reactions (1) and (2):
Mg(OH)2 + Mn2+ → Mn(OH)2 + Mg2+ Log K = 1.44 (1)
Ca(OH)2 + Mn2+ → Mn(OH)2 + Ca2+ Log K = 7.36 (2)
The values of the equilibrium constants K indicate that the selectivity for manganese hydroxide
is thermodynamically more favorable with respect to Ca than Mg [25].
Zhang et al. [26] performed hydroxide precipitation at 60°C with NaOH solutions for
removal and recovery of manganese from synthetic laterite waste solution containing 2g/L Mn2+,
15 g/L Mg2+ and 0.5 g/L Ca2+ (Table 2). At pH 8.2, 71% precipitation of Mn2+occurred along
with 5% Mg co-precipitation. At pH >9, the concentration of Mn2+ obtained in the solution was
less than 10 mg/L with initial concentration as 534 mg/L [26]. This suggested that hydroxide
precipitation alone is not a favorable option for the removal of manganese to very low levels (5
mg/L) due to its poor selectivity, need for higher pH above 8.5, and substantial co-precipitation
of magnesium. It causes high consumption of base for neutralization and therefore, increased
operating cost [24-26]. Also, hydroxide sludge quantities can be substantial and are generally
difficult to dewater due to the amorphous particle structure [22].
2.1.2. Carbonate precipitation
Carbonate precipitation is an important, effective and practical process for manganese
recovery [25,27]. The selectivity for manganese is thermodynamically more favorable with
respect to Mg than Ca [25]. This is indicated in reactions (3) and (4):
MgCO3 + Mn2+ → MnCO3 + Mg2+ Log K = 5.48 (3)
CaCO3 + Mn2+ → MnCO3 + Ca2+ Log K = 2.35 (4)
It is carried out by Na2CO3 and limestone [26,27]. Using Na2CO3 solution at 60°C,
manganese precipitation occurred at pH >7.5 and reached 90% at pH 8 with about 13% Mg and
43% Ca co-precipitation [26]. When compared with hydroxide precipitation, carbonate
precipitation is observed to be more selective for manganese over magnesium than calcium [25-
26]. It has been shown that the precipitation of manganese carbonate is more thermodynamically
favorable than the precipitation of calcium carbonate. It is also observed that limestone-sodium
carbonate blends are beneficial for manganese removal from mine waters [27,28]. Sodium

6
carbonate provided carbonate ions to precipitate high manganese contents in mine water and
industrial effluents. However, powdered limestone induced heterogeneous nucleation of
manganese carbonate and was found as a cost-effective, simple alternative to remove manganese
from mining-affected waters [27,30]. It is also, reported that, above pH 8.5, manganese removal
efficiency is 99.9% [27].
Bamforth et al. [29] used four experimental reactors, comparing the efficiency of
quartzite, high purity lime-stone, dolomite and magnesite as substrates for manganese mineral
precipitation. It has been observed that kutnahorite and aragonite precipitated exclusively on
reactor plastics and only on the quartzite substrate. Magnesian calcite precipitated on dolomite,
magnesite and limestone, and calcite only on dolomite and magnesite. Manganese oxy-
hydroxides precipitated on all substrates.
Hence, carbonate precipitation is a fast and effective method for the recovery of
manganese with relatively good quality from bulk solutions. Subsequent leaching of metal
carbonate is also easier and consumes less acid than hydroxide or sulfide precipitates [31]. Also,
it offers better selectivity for manganese over magnesium for partial recovery of manganese.
However, it is not appropriate for the removal of manganese to very low levels due to extensive
co-precipitation of magnesium impurity at pH above 8 [26,27] (Table 3).
2.1.3. Oxidative precipitation
Oxidative precipitation offers the best selectivity for recovery of manganese as
MnO2/Mn2O3 over other metals including Zn, Ni, Co, Mg and Ca in a pH range of 3 to 6 [25].
Zhang et al. [26] performed oxidative precipitation with air, O2 and SO2/air. However,
kinetics of the oxidative precipitation is a critical factor for a practical application. Therefore,
they investigated the rates of manganese oxidative precipitation at different pHs using both air
and oxygen for evaluation. At pH<6, the manganese precipitation with air was very insignificant.
At pH 7, about 10% Mn(II) precipitated in 30 min, but the precipitation eventually stopped
afterwards. With a further increase in pH, the rate of precipitation was increased which was
apparently initiated by the hydroxide precipitation of manganese. O2 as an oxidant considerably
enhanced the rate of manganese precipitation over the use of air and 99.5% manganese
precipitation was obtained at lower pH range 8 to 8.5 and thus less co-precipitation of
magnesium with the use of expensive oxygen. However, oxidative precipitation with SO2/air
(O2) has revealed much faster kinetics in a neutral pH range of 6 to 7 with 99.5% manganese

7
recovery, less than 10 mg/L Mn(II). It is very favorable for both removal and recovery of
manganese from nickel laterite and other waste solutions [25-26]. Also, with an appropriate
reactor design, agitation and SO2/O2 ratio, it is an effective oxidant for removing manganese
from a neutral leach zinc solution ([Zn] = 150 g/L) at 80 °C down to ppm level. They found the
precipitate in the form of rounded particles of about 10 μm diameter consisting of poorly
crystalline complex manganese oxide hydrate considered of birnessite type [32] (Table 4).
The mechanism is given by following reactions (Eqs.5-9). SO2/air (O2) ratios could be
improved to obtain the maximum rate of Mn(II) oxidation and minimum consumption of SO2
and base reagents for neutralization.
Oxidation:
2Mn2+ + SO2 + O2 + 3H2O → 2MnO(OH) + SO42- + 6H+ (5)
2Mn2+ + SO2 + O2 + 2H2O → MnO2 + SO42- + 4H+ (6)
Reduction:
2MnO(OH) + SO2 + 4H+ → 2Mn2+ + SO42-+ 2H2O (7)
MnO2 + SO2 → Mn2+ + SO42- (8)
Catalyzed side reaction
SO2 + 1/2O2 + H2O → SO42- + 2H+ (9)
Umetsu et al. [33] studied oxidative precipitation using ozone for the removal of arsenic
with manganese from dilute acidic sulfate solutions under the conditions as the initial
concentrations of arsenic(III) and arsenic(V) = 1–100 mg/L, the initial mole ratio of Mn/As =
10–100, pH = 0.4–5.0 and temperatures 15–80oC. They found that when an O3 – O2 gas mixture
was supplied to solutions containing manganese(II) and arsenic(III), the oxidation of arsenic(III)
to arsenic(V) took place prior to the oxidation of manganese(II).
The distinctive feature of the oxidative precipitation in the separation of manganese from
other valuable metals such as zinc, cobalt and nickel is that manganese can be selectively
separated at lower pH range 3–6 [25]. It is also observed that a combination of the hydroxide or
carbonate method for the removal or recovery of a large proportion of Mn(II) with oxidative
precipitation to very low level may be the finest alternative in terms of efficiency and economics
[26].
The primary concern for the use of oxidative precipitation method is the cost of some
oxidants such as ozone, Caro's acid, and peroxy-disulfuric acid. It may not validate its use for

8
recovery of manganese from a solution containing a low manganese concentration, e.g. laterite
waste solutions containing 1–5 g/L manganese. But the SO2/air oxidizing mixture is a relatively
cheap oxidant that is considered to be the most suitable oxidant for recovery of manganese from
waste solutions [25].
2.1.4. Sulfide precipitation
Metal sulfide precipitation is an important process in hydrometallurgical treatment of
ores and effluents [34]. The precipitation of metal sulfides and separation is based on different
sulfide solubilities of metals at a certain pH and temperature. Gaseous hydrogen sulfide (H2S) or
sodium sulfides (Na2S) or ammonium sulfide (NH4)2S is usually employed in the precipitation of
metal sulfides. Manganese sulfide is more soluble than most other metal sulfides. This offers a
theoretical basis for separation of Mn2+ from other metals such as Cu2+, Zn2+, Co2+, Ni2+, and
Fe2+ in hydrometallurgical processes, where other metal ions are precipitated as metal sulfides
while Mn2+ ions remain in solution [25].
Bryson et al. [35] precipitated manganese and cobalt sulfides from a sulphate solution in
a batch reactor by the addition of ammonium sulfide at 35 °C and a pH of 7. They showed that
first-order kinetics simulate the precipitation of manganese sulfide, and oxygen reacts with the
freshly prepared manganese sulfide to form an insoluble higher oxide (e.g. Mn3O4, which tends
to darken the appearance of the precipitate) releasing the sulfide or hydrogen sulfide ion to react
further. They found manganese sulfide as a very light pink precipitate almost immediately after
the addition of ammonium sulfide was begun. It is shown that the rate of precipitation may be
modeled by first-order kinetics with rate constant of 0.04 min-1. The chemical composition of the
bulk of the solids is probably either MnS.H2O or MnOH.SH. The presence of air in the system
has negative effects on the process (Table 5).
Mishra et al. [36] discussed novel information obtained on the precipitation of manganese
sulfide, and the co-precipitation of copper-manganese sulfide, in hot-rolled strips of
commercially produced grain-oriented silicon steel. It was observed, that an increase in copper
concentration in the sulfide particles was associated with (i) a change in shape from spherical to
parallelopipedal to cuboidal, and (ii) a simultaneous decrease in size. Once the co-precipitation
of copper-manganese sulfides was established, it was thought appropriate to study the effect of
copper on the shape and size of the sulfide particles. They found that the size of sulfide particles
decreased with a decrease in manganese-to-copper ratio, i.e. with an increase in copper. This

9
indicates that as the nature of the sulfide changed from pure manganese sulfide to complex
copper-manganese sulfide to pure copper sulfide, the size of the sulfide particles decreased and
associated with the change in shape from spherical to parallelopipedal to cuboidal.
Although hydroxide precipitation is widely used in industry for metal removal, there are
some advantages to sulfide precipitation, including the lower solubility of metal sulfide
precipitates, potential for selective metal removal, fast reaction rates, better settling properties
and potential for reuse of sulfide precipitates by smelting. However, sulfide precipitation is not
used as widely as it could be because the dosing of sulfide is seen to be difficult to control (due
to the very low solubility of the metal sulfides and thus the sensitivity of the process to the dose)
and because of concerns about the toxicity and corrosiveness of excess sulfide [34]. However,
sulfide precipitation is not considered for use for recovery of manganese from industrial waste
solutions because of unfavorable attributes as, need of pollution control and management, higher
stages to separate from other impurities and manganese sulfide is not a favorable product for the
manganese industry, requiring further conversion [25]. Also, it is found that there is no selective
separation of manganese over iron and measured redox potentials had too high values for the
formation of metal sulfides [31]. However, it provides a useful means for optional purification of
small amounts of metal impurities such as Cu, Zn, Ni and Co before the step for production of
highly pure manganese products. It has found wide applications for purification of solutions from
leaching polymetallic manganese nodules, Zn–Mn bearing sludge [25].
2.2. Coagulation / Flocculation
Coagulation / Flocculation is an effective way to control heavy metal pollution
with/without other water treatment technologies. Flocculation is a physicochemical process that
reassures the aggregation of coagulated, colloidal and finely divided suspended matter by
physical mixing or chemical coagulant aids [37].
Due to the use of wide range of pH and large surface area of the resulting flocs, the iron-
based coagulants have better performance compare to aluminum-based coagulants. The factors
that significantly influence the performance of the coagulation process are dosage, pH,
temperature, and valence state of heavy metals [38]. Fu-wang et al. [39] studied the effect of
AlCl3 to remove manganese(II) by coagulation. They found that it can be enhanced by using
K2MnO4 oxidation and coagulation aid, as it was a relatively efficient technology for
manganese(II) removal. Charemtanyarak [40] used synthetic wastewater of pH 1.9 with

10
concentration of manganese as 1085 mg/L and found that the optimum pH for chemical
coagulation and precipitation by lime treatment was more than 9.5 (Table 6).
Improved sludge settling, dewatering characteristics, bacterial inactivation capability,
sludge stability are reported to be the major advantages of lime-based coagulation. In spite of its
advantages, coagulation–flocculation has limitations such as high operational cost due to high
chemical utilization. The increased volume of sludge generated from coagulation–flocculation
may hinder its adoption as a global strategy for wastewater treatment [14].
2.3. Flotation
Dissolved air flotation, ion flotation and precipitate flotation are the main techniques for
removal of heavy metal ions from wastewater.
Dissolved air flotation (DAF) is not in itself a process for removing manganese.
However, like sedimentation, it does remove the precipitates formed following the oxidation of
the dissolved manganese by agents such as air, Cl2, KMnO4, and O3. DAF will remove low levels
of manganese precipitates. It sometimes encounters difficulty in removing high manganese
concentrations, as the precipitates are not always easy to float [41].
Ion flotation may be an alternative method to remove heavy metal ions from wastewaters.
The process of ion flotation is based on imparting the ionic metal species in wastewaters by the
use of surface active agents (surfactants). It is followed by subsequent removal of these
hydrophobic species by air bubbles [42].
Precipitate flotation is another efficient method in which small bubbles stabilized with
surface-active agents carry the precipitated particles, attached by coloumbic forces, to a surface
where the particles concentrate for easy removal. Separation efficiency is determined by the
surface area available for solute-particle attachment and by charge compatibility of the bubbles
with the solute. It is observed that in a batch process, 80% manganese separation was complete
within 5 min, making this process much more attractive than other conventional batch processes
[43].
Stoicaet et al. [44] investigated a Mn(II) recovery method from aqueous systems by
precipitate flotation (DAF variant) using collector reagents as the izoalkyl carboxylic acids type
species with 8 carbon atoms in the structure (molar ratio 1:2.7). These species were not used
before in precipitate flotation. Mn(II) aqua species were insolubilized as basic manganese
carbonate and then interacted with the collector reagent. The results showed the possibility of

11
Mn(II) recovery as percentage recovery ≥ 99. In this method, a tensioactive reagent having a
nonlinear chain with a relative small number of carbon atoms is used. Good results from the
ecological point of view were obtained for Mn(II) initial and final concentrations in the range of
25 mg/L and less than 1 mg/L, respectively (Table 7).
It offers several advantages such as high metal selectivity, high removal efficiency, high
over flow rates, low detention periods, low operating cost and production of more concentrated
sludge. But, the disadvantages involve high initial capital cost, high maintenance and operation
costs [14].
2.4. Ion Exchange
Ion exchange is a physical treatment process in which ions dissolved in a liquid or gas
interchange with ions on a solid medium. The ions on the solid medium are associated with
functional groups that are attached to the solid medium, which is immersed in the liquid or gas.
Typically, ions in dilute concentrations replace ions of like charge that are of lower valence state,
but ions in high concentration replace all other ions of like charge. The divalent or trivalent ions
move from the bulk solution to the surface of the ion exchange medium, where they replace ions
of lesser valence state, which, in turn, pass into the bulk solution. The ion exchange material can
be solid or liquid, and the bulk solution can be a liquid or a gas [24].
The free cations of iron and manganese can be removed using zeolite processes. As water
flows through the zeolite medium, cationic exchange occurs. Backwashing the zeolite, typically
with a brine solution prepared from sodium chloride (NaCl), removes the iron, manganese, and
hardness cations that have accumulated. After regeneration, the zeolite medium is rinsed with
clean water to remove residual brine prior to returning the process to service [2].

This process can be used in wastewater treatment plants to swap one ion for another for
the purpose of demineralization. There are two types of ion exchange systems, the anion
exchange resins and the cation exchange resins. It can be used for removal of salt, ammonium
ion from wastewaters, heavy metals from electroplating wastewaters and other industrial
processes and for polished wastewater before discharging. It is also used to purify acids and
bases for reuse or removal of radioactive contaminants in the nuclear industry [37].
White et al. [45] used synthetic manganese dioxide as a material for the removal of Mn2+
and Fe2+ from drinking water. They prepared three different types of MnO2. One involved the
production of the dioxide from permanganate and hydrogen peroxide, the second involved the

12
use of the permanganate and manganous ions and the third used a formulation the included all
three reagents. Tests showed that the first type of dioxide was by far the best for both iron and
manganese removal. Other tests indicated that the dioxide exchanger could be regenerated
partially by contact with chlorine gas or hypochlorite which is commonly available as a process
chemical at water treatment works.
Kononova et al. [46] investigated sorption recovery of toxic ions such as chromium(VI)
and manganese(II) from aqueous solutions with different acidity (0.001–0.5 M HCl) on cation
and anion exchangers. These exchangers are synthesized with long-chained cross-linking agents
(LCA) like divinyl esters of diethylene-glycol (DVEDEG), triethylene-glycol (DVETEG),
propylene-glycol (DVEPG) and tetravinyl ester of pentaerythritol (TVEPE). The initial
concentrations of Cr(VI) and Mn(II) were 1 g/L and 5 g/L, respectively. It was shown that the
resins with LCA possess high ionic permeability due to the elasticity of polymeric skeleton.
They found that these sorbents have high selectivity and good kinetic properties. It is favorable
to achieve quantifiable (~100%) recovery and separation of manganese(II) and chromium(VI) in
counter-current columns (Table 8).
It is the most energy efficient and economical technology of all the recovery techniques.
This system is the only one that can be efficiently used for treating a very dilute solution in parts
per million (ppm) levels on a once-through basis [45]. Also, it is a useful method for the
purification and separation of manganese from solution containing metals like Cu, Fe, Co, Ni
and Pb. Ion exchange is more environment friendly and easy to control as compared to sulfide
precipitation method. However, a resin has an inadequate capacity for adsorption of particular
metals and therefore, more suitable for removal of trace amounts of metal impurities for the
preparation of highly pure manganese solutions [25].
However, ion-exchange resins must be regenerated by large volume of chemical reagents
when they are exhausted and the regeneration can cause serious secondary pollution. It is also
expensive, especially when treating a large amount of wastewater containing heavy metal in low
concentration and so they cannot be used at large scale [14]. It offers nonselective removal of
other ions which increases operation cost [23]. Also, it requires neutral to alkaline pH conditions
[6].

13
2.5. Oxidation / Filtration
Oxidation / Filtration is the most suitable method for rural areas as these methods can be
applied without electricity [47]. It is the most common method to remove Mn(II). Such a method
is based on the Mn(II) oxidation to its insoluble manganic dioxide, followed by clarification or
filtration [48].
2.5.1. Oxidation
Iron and manganese often naturally occur in normal groundwater pH ranges, in a
dissolved form as Fe2+and Mn2+. In the presence of an oxidant, these ions are oxidized to form
insoluble products, namely Fe(OH)3 and MnO2 that can be filtered as shown in the following
examples:
MnO2 (Media)
Mn2+ + NaOCl + H2O → MnO2(s) + Cl- + 3H+ (10)
Oxidation can be accomplished by aeration or by using strong oxidizing agents such as
potassium permanganate, hypochlorite, chlorine dioxide or ozone [48]. Potassium permanganate
(KMnO4) is used as an oxidant to precipitate Mn(II) and 96% removal of Mn(II) from ground
water is possible [32,34]. Also, some investigators have removed Mn(II) from ground water by
oxidation with KMnO4 followed by microfiltration [50] and KMnO4 oxidation followed by
flocculation, settling and filtration and achieved 95% removal of Mn(II) [48]. Kan et al. [51]
used synthetic groundwater for removal of Mn(II) by oxidation using potassium permanganate.
Also, they studied the effect of parameters such as the oxidant dose, presence of co-ions (Ca2+,
Mg2+) and alum addition on the removal of Mn2+ and Fe2+. The partial removal of Mn2+ using
aeration in single and dual metal system was 30.6% and 37.2%, respectively. To reduce Mn2+
below its maximum contaminant level, a minimum oxidant dose of 0.603mg/L KMnO4 was
required. Oxidation of Mn2+ using KMnO4 is illustrated in Eqs. (11) and (12):
3Mn2+(aq) + 2KMnO4(aq) + 2H2O(l) → 5MnO2(s) + 2K+(aq) + 4H+(aq) (11)
KMnO4(aq) + 8H+(aq) + 5e-(eq) ↔ Mn2+(aq) + K+(aq )+ 4H2O(l) (12)
Lesley et al. [52] used constructed wetlands to treat mine water drainage and achieved
76.8% removal of manganese. It was concluded that bacterially-mediated oxidation might be
involved.

14
2.5.2. Filtration
Filters are physical treatment devices, and the mechanisms of removal include one or
more of physical entrapment, adsorption, gravity settling, impaction, straining, interception, and
flocculation. Slow sand filtration accomplishes removal within only the first few millimeters of
depth from the surface of the sand. Both deep bed granular filters and pre-coat filters (when body
feed is used) make use of a filter medium, and (or other) filtration effectively removes mostly
those particles of metal precipitate that will not settle [24]. Enhanced filtration can be carried out
by conditioning the filter media with permanganate to form a manganese oxide coating on the
surface of the media or by using green-sand [48]. In the case of a manganese greensand filter, the
filter media is coated with potassium permanganate that oxidizes the dissolved iron and
manganese and then filters them out of the water. Manganese greensand filters require significant
regeneration with a potassium permanganate solution as it is consumed during oxidation of
dissolved metals [41].
Treatment of mine water using manganese sand modified by KMnO4 exhibited greater
iron and manganese absorption compared to ordinary manganese sand. The optimum
concentration of the modifier (KMnO4) was found to be 5% with Mn(II) concentration up to
0.01mg/L [53]. Piispanen et al. [54] used manganese oxide coated sand / anthracite filter medium
to treat ground water and found 91% removal of Mn(II) with less than 0.02 mg/L remains in
treated water. Aziz et al. [55] discussed filtration through a low-cost coarse media like limestone.
They found better Mn(II) removal as 95% compared to the gravel (60%), crushed brick (82%) or
with no media addition (less than 15%). Pacini et al. [56] studied the performance of up-flow
roughing filter technology in a biological treatment process. They found 95% Mn(II) removal
efficiency without the use of chemical agents and under natural pH conditions. It is also,
investigated that a particular area plays a crucial role in biological manganese oxidation, and a
model is developed which incorporates heterogeneous autocatalytic and biological manganese
oxidation [55-56].
Biological trickling filter is also, used to remove Mn(II) from potable water and found
94% removal caused by both biological and chemical oxidation. It increases filter’s efficiency by
about 6% and reduces the required filter depth by about 40% [57-59]. In biological aerated filter
(BAF), to remove simultaneously ammonia and manganese for drinking water pretreatment, a
mixed culture of heterotrophic bacteria is used. It was observed for the first time that BAF

15
system could simultaneously bio-oxidize micro-pollutants, ammonia and manganese [60-61].
Gagnon et al. [62] used bio-filters inoculated with L. discophora sp-6 and indigenous mixed bio-
film and achieved 90% removal of Mn(II).
Carriere et al. [63] used point-of-use devices to remove Mn(II) from drinking water and
found greater than 60% Mn(II) removal using pour-through filters. Huang et al. [64] found that
manganese ore constructed wetland is a viable and cost-efficient treatment technology for
wastewater reclamation. Compared with single gravel wetland, manganese ore wetland showed
better removal of all target pollutants (COD, turbidity, ammonia, nitrogen and total phosphorus),
especially for iron and manganese. The effluent concentrations of both species were
continuously below 0.1mg/L under unfavorable conditions for chemical oxidation.
Oxidation / filtration is an efficient and affordable technology for rural areas as it is
applicable without electricity and hence, a low-cost method. The major drawbacks are difficulty
in storage and transport of the oxidant, corrosion of system parts, and formation of solid
manganese compounds that may interfere with system operations [47] (Table 9).

2.6. Electrochemical treatment


Electro-coagulation (EC) is based on the in-situ production of a coagulant by anodic
dissolution and, subsequently, produces aluminium or iron hydroxides in the contaminated water.
Besides, cathodic reactions allow for the pollutant removal either by deposition on cathode
electrode or by flotation (evolution of hydrogen at the cathode).
The electrolytic dissolution of the anode produces cationic monomeric species such as
Al3+ and Fe3+. A range of coagulant species and hydroxides is formed which destabilize and
coagulate the suspended particles or precipitate and adsorb the dissolved contaminants. It is well
known that, in an EC process, the primary reactions occurring at the aluminium electrodes during
electrolysis are:
(a): Anodic reactions:
Al → Al3+ + 3e− (13)
2H2O → O2(gas) + 4H+(aq) + 4e− (14)
(b): Cathodic reactions:
2H2O + 2e− → H2(gas) + (OH)−(aq) (15)

16
Shafaei et al. [65] used EC as an efficient method for the removal of a variety of
pollutants, especially heavy metal ions, from water and wastewater effluents. In this study, the
removal of Mn2+ ions from solutions has been investigated by EC process with aluminium
electrodes. The effect of influential parameters such as initial pH, current density, electrolysis
time, solution conductivity and initial metal concentration on the performance of EC process has
been investigated. It was found that, to remove Mn2+ ions, the optimum initial pH was 7.0. Also,
the results indicated that Mn2+ removal efficiency improves with increase in the current density
and electrolysis time. The removal of Mn2+ ions was not influenced by the solution conductivity,
but the electrical energy consumption decreased with an increase in the solution conductivity.
Besides, it is shown that with increase in concentration of the contaminant, Mn2+ removal rate
decreased.
Ince et al. [66] investigated the effective performance of the EC technique to remove
manganese, phosphate, and iron from rinse water of a Mn-PO4 (MPO) coating plant. They used
sacrificial aluminum electrodes in original pH. It was found that the speed of the removal
enhanced significantly, with an increase in the current. However, a simultaneous increase of
electrode and energy consumption was observed. The optimum current density with the fastest
treatment was found to be 20 Am-2. With this current density, the process removes 97.95% of
manganese, 99.96% of the phosphate, and 99.78% of iron, after 30 min (Table 10).
Thus, these treatment techniques are regarded as rapid and well-controlled that require
fewer chemicals, provide good reduction yields and produce less sludge. However,
electrochemical technologies involve high initial capital investment and expensive electricity
supply and this restricts its development [14].
2.7. Adsorption
Adsorption can be defined as the accumulation of one substance on the surface of
another. The mechanism of adsorption can be one or a combination of several phenomena,
including chemical complex formation at the surface of the adsorbent, electrical attraction (a
phenomenon involved in almost all chemical mechanisms, including complex formation), and
exclusion of the adsorbate from the bulk solution. It results from stronger intermolecular bonding
between molecules of solvent (hydrogen bonding in the case of water) than existed between
molecules of solvent (water) and the solute. In this case, the driving force is explained by the

17
second law of thermodynamics. The total energy of all bonds is greater after adsorption has taken
place than before [24].
Adsorption process is used to remove color and other soluble organic pollutants from the
effluents. The process also removes toxic chemicals such as pesticides, phenols, cyanides and
natural dyes that cannot be treated by conventional treatment methods. Using various adsorbents,
dissolved organics are adsorbed on the surface when wastewater is made to pass through it [37].
Numerous literature examples are summarized in a review paper by Hilal et al. [67].
2.7.1. Activated carbon as adsorbents
Most commonly used adsorbent for treatment of wastewater is activated carbon (AC).
Activated carbon once saturated needs replacement or regeneration. The chemical regeneration
can be done within a column either with acid or other oxidizing chemicals [37]. Many recent
papers describe the use of activated carbon as an adsorbent for removal of Mn(II) [68-72].
Anbia et al. [68] impregnated mesoporous carbons with the anionic surfactants and
cationic surfactants to remove Hg(II) and Mn(II) from aqueous solution. They found mercury
removal by cationic surfactant cetyltrimethyl ammonium bromide (CTAB), anionic surfactant
sodium dodecyl sulfate (SDS) modified mesoporous carbon and unmodified mesoporous carbon
as 94%, 81.6% and 54.5%, respectively while the manganese removal for these adsorbents were
82.2%, 70.5% and 56.8%, respectively. They found that initial concentration and pH are
determinant factors for the removal of these adsorbates and sorption of Hg (II) and Mn (II)
decreases with increase in initial concentration and increases with increase in pH. Also,
modification with surfactants can significantly enhance the adsorption capacity of the
mesoporous carbons and Langmuir adsorption isotherm better fitted than Freundlich isotherm.
Jusoh et al. [69] studied the potential and the effectiveness of granular activated carbon
(GAC) in removal of ferrous and manganese from water. They found Langmuir isotherm to fit
and the adsorption capacity of Fe(II) and Mn(II) were 3.6010 and 2.5451 mg/g, respectively.
Also, they related this capacity onto GAC to ionic radius and electro-negativity of metal ions as
smaller ionic radius and higher electro-negativities corresponded to higher adsorption.
Motojima et al. [70] used packed column with oxine-impregnated activated charcoal to
remove manganese from solution. They found that the amount of the ion adsorbed decreased
exponentially with the depth of the column and the distribution follows a first order kinetic plot.

18
Goher et al. [71] studied the removal of aluminum, iron and manganese using two
different sorbents; granular activated carbon (GAC) and Amberlite IR-120H (AIR-120H). They
showed that the two sorbents have maximum removal efficiency for aluminum and iron at pH 5
and 10 min contact time in room temperature, while pH 7 and 30 min were the most appropriate
for manganese removal. Dosage of 2 g/L for both GAC and AIR-120H gave the maximum
removal capacity. At optimum conditions, the removal trend was in the order of Al+3> Fe+2>
Mn+2 with 99.2, 99.02 and 79.05 and 99.55, 99.42 and 96.65% of metal removal with GAC and
AIR-120H, respectively. For the three metals, Langmuir and Freundlich isotherms showed
higher R2 values, with a slightly better fitting for the Langmuir model. Also, found that removal
percentage of metal ions increases as the pH values increase and decreases with increase in
adsorbent dosage.
Okoniewska et al. [72] used impregnated activated carbon to remove manganese, iron
and ammonium nitrogen. It was found that the surface of the activated carbon studied after
impregnation decreased from 725 to 658 m2/g (about 9%). Both pH and initial concentration of
manganese, iron and ammonium nitrogen in treated water influenced the adsorption quantity,
that is, with increase in pH and decrease in initial concentration, adsorption increases. In all
cases, manganese was best adsorbed on impregnated activated carbon from compounds studied,
independently on the initial concentration and filtration speed. For high removal of manganese,
selection of process parameters is necessary.
2.7.2. Polymer as adsorbents
Various polymers have been used as adsorbents for removal of manganese [73-79].
Moawed et al. [73] used the polyhydroxyl-polyurethane foam (PPF) as a new sorbent for
separation of manganese and iron ions in natural samples. They found that maximum sorption of
Mn(II) and Fe(III) was in the pH range of 6–8. The kinetics of sorption of the Mn(II) and Fe(III)
by the PPF was fast with an average value of half-life of sorption (t1/2) of 11.7 min. The sorption
capacity of PPF was 8.7 μmol/g and achieved the recovery of tested ions as 99–100%.
Al-Wakeel et al. [74] studied removal of Mn(II) ions on glycine modified Chitosan and
found adsorption capacity as 1.3 mmol/g at pH 6 and 25 oC. They showed that adsorption
increases with increase in pH and contact time.
Z. Abdeen et al. [75] explored the ability of polyvinyl alcohol / chitosan (PVA/CS)
binary dry blend as an adsorbent for removal of Mn(II) ion from aqueous solutions. The results

19
revealed that the removal of Mn(II) ion is pH, adsorbent dosage and contact time dependent and
the optimum pH was 6.0. Also, the adsorption equilibrium and kinetics of adsorption were fitted
well by Freundlich isotherm model and by the pseudo-second-order kinetic model, respectively.
The thermodynamic study indicated that the adsorption of Mn(II) ion onto PVA/CS was
spontaneous and endothermic in nature.
Tang et al. [76] used a poly (sodium acrylate)-graphene oxide (PSA-GO) double network
hydrogel adsorbent to remove Cd(II) and Mn(II) from aqueous solution. The adsorption
behaviors in different conditions (e.g., pH, contact time, ionic concentration and existence of
fulvic acid) as well as the adsorption mechanism were studied. The maximum sorption capacities
were found up to 238.3 mg/g and 165.5 mg/g at pH 6 and at a temperature of 303K, for Cd(II)
and Mn(II), respectively, estimated from the Langmuir model. Also, for selective adsorption
study, they prepared a mixed solution of metal ions containing Pb(II), Cu(II), Cd(II), Zn(II),
Mn(II) and Ni(II) at pH 5.0 ± 0.1, and all the heavy metal ions were set at ~ 40 mg/L. It was
found that the uptakes of Pb(II), Cd(II) and Mn(II) ions were higher than others. This may be due
to the difference of chelation ability to different metal species, and Cd(II), Pb(II) and Mn(II)
could be favored to adsorb onto the surface of PSA-GO gel. Since Pb(II) is easily precipitated at
pH>6 under actual conditions, this hydrogel should be a selective adsorbent of Cd(II) and Mn(II)
in a wide range of pH.
Qomi et al. [77] investigated adsorption of manganese ions from aqueous solution by
polyaniline / sawdust (PAn/SD) nano-composite. They evaluated the effect of various
experimental parameters i.e., pH, adsorbent dosage and contact time on the removal efficiency.
They showed that competition of H+ ions with metal ions at low pH affects the adsorption
capacity. Also, increase in adsorbent dosage leads to increase in adsorption due to increase in
number of adsorption sites. They found optimum conditions as pH of 10, adsorbent dosage of 10
g/L and equilibrium contact time of 30 min. The adsorption isotherm fitted with Freundlich
isotherm with monolayer adsorption capacity as 58.824 mg/g.
Reiad et al. [78] prepared microporous chitosan (CS) membranes by extraction of
poly(ethylene glycol) (PEG) from CS/PEG blend membrane and examined for iron and
manganese ions removal from aqueous solutions. They found the affinity of CS/PEG blend
membrane to adsorb Fe(II) ions was higher than that of Mn(II) ions, with adsorption equilibrium
achieved after 60 min for Fe(II) and Mn(II) ions. By increasing CS/PEG ratio in the blend

20
membrane, the adsorption capacity of metal ions increased. Among all parameters like contact
time, dosage of adsorbent, initial metal ion concentration in the feed solution, pH had the most
significant effect on the adsorption capacity, particularly in the range of 2.9–5.9. The increase in
CS/PEG ratio was found to enhance the adsorption capacity of the membranes. The experimental
data were better fitted to Freundlich equation than Langmuir.
Guan et al. [79] removed Mn(II) and Zn(II) in wastewater from dual-alkali flue gas
desulfurization (FGD) system by water-soluble chitosan. The pH investigation revealed that at
the pH ranging from 5 to 9, there were three stages for different actions: chelation of chitosan for
metal ion, precipitation of metal hydroxide and co-precipitation of metal hydroxide and
chitosan–metal complex. The selective chelation of chitosan for Mn(II) and Zn(II) mixture
solution was also studied. The results showed that the chelation of chitosan for Mn(II) was prior
to Zn(II) in multiple component solution. The precipitation of heavy metal hydroxide could have
effect on the selectivity of chitosan for Mn(II) over Zn(II). On the other hand, co-precipitation of
the complicate heavy metal in the FGD wastewater enhanced the heavy metal removal of
chitosan chelation.
2.7.3. Biosorption and natural adsorbents
Although AC is a widely used adsorbent, it is relatively expensive. So searching low-cost
adsorbents is a primary focus of the research [80-84]. Natural minerals, industrial wastes and
bio-sorbents are inexpensive, abundantly available and efficient adsorbents for removal of
manganese.
2.7.3.1. Natural minerals, industrial wastes as natural adsorbents
It has been reported that natural minerals like surfactant modified alumina [85-86],
natural zeolite [87-90], Slovakian natural zeolite [91], synthetic zeolite [92], manganese oxide
coated zeolite [93], clinoptilolite [94,101], Mexican clinoptilolite [95], fly ash [92,96], Nigerian
kaolinite Clay [97], kaolinite [98], coal [99], Calcium hydroxyapatite [100], vermiculite [101-
102], alkaline-modified montmorillonite [103], lignite [104], Al-Zeolite, NH4-Zeolite [90] are
effective adsorbents for Mn(II) removal.
Palet al. [85-86] studied the removal of Mn(II) from aqueous solution on prepared
surfactant-modified alumina (SMA). The bilayer of sodium dodecyl sulfate (SDS; an anionic
surfactant), formed on alumina surface, can adsolubilize Mn(II) in a rapid process. The effects of
contact time, initial concentration of adsorbate, adsorbent dose, pH, temperature, agitation speed

21
were discussed. They showed that with increase of pH from 4.04 to 8.05 and agitation speed
from 110 to 150 rpm, the capacity increased from 43.48 to 62.02% and < 10%, respectively.
Also, with increase in temperature, adsorbent dosage, the removal of Mn(II) increased due to
endothermic process and availability of higher adsorption sites, respectively. They found that the
adsorption followed Freundlich isotherm model and the process to be endothermic, favorable and
spontaneous.
Taffarel et al. [87] employed a Chilean zeolite (Ch-zeolite) as adsorbent for manganese
ions from aqueous solutions. They investigated that the medium pH influenced significantly the
Mn2+ ions adsorption capacity and best results were obtained at pH 6–6.8. The adsorption onto
the activated zeolite followed Langmuir isotherm model and the pseudo-second-order as kinetic
model.
Zamzow et al. [88] analyzed twenty-four zeolite samples. These included clinoptilolite,
mordenite, chabazite, erionite, and phillipsite. The 24 zeolites and an ion-exchange resin were
tested for the uptake of Cd2+, Cu2+, and Zn2+. Of the natural zeolites, phillipsite proved to be the
most efficient, while the mordenites had the lowest uptakes. They used a multi-ion wastewater
from copper mine and found that the metal ions Fe2+, Cu2+, and Zn2+ were removed to below
drinking water standards, but Mn2+ ions were removed only from 22 mg/L to 0.35 mg/L.
Rajic et al. [89] studied the natural zeolite tuff from the Vranjska Banja deposit (Serbia)
as sorbent for Mn(II) ions from aqueous solutions. They found that the zeolite sample containing
mainly clinoptilolite (more than 70%) removed Mn(II) ions by ion-exchange process. The
sorption isotherm was best described by the Langmuir–Freundlich or Sips model. The kinetics
followed the pseudo-second-order model.
Ates [90] used the natural zeolite (NZ) modified by ion-exchange (NH4NO3), alkali
treatment (NaOH) and addition of aluminum (Al2(SO4)3). They found that ion-exchange with
NH4+ of NZ results in the exchange of Na+ and Ca2 + cations and the partial exchange of
Fe3 + and Mg2 + cations. Also, all modifications, ion exchange, alkali treatment, and aluminum
introduction, increased the manganese adsorption capacity of natural zeolites by two times. The
Freundlich isotherm model was best fitted to the isotherm data obtained, due to a heterogeneous
surface existence.
Shavandi et al. [91] investigated the adsorption capacity of natural zeolite for the removal
of heavy metal ions, zinc Zn(II), manganese Mn(II) and iron Fe(III), found in palm oil mill

22
effluent. They evaluated the effects of contact time, agitation speed, pH, and sorbent dosage on
the sorption of heavy metals. They found more than 50% of Zn(II) and Mn(II) and about 60% of
Fe(III) removal and the sorption was fast with equilibrium reached within 180 min. Also, the
metal sorption increased with pH, and adsorption capacities ranged between 0.015 and
1.157 mg/g of zeolite. Equilibrium data and kinetic data followed the Langmuir isotherm model
and the pseudo-second-order model, respectively.
Belviso et al. [92] explored fly ash and zeolite synthesised from fly ash as high-efficiency
adsorbents to removes Mn2+. The results indicated that both materials were effective for the
removal of Mn from aqueous solution by precipitation due to the high pH of the solid / liquid
mixtures. However, the leaching tests revealed that the amount of Mn(II) removed from the fly
ash was greater than that leached from the zeolite, thereby indicating that the metal was partially
sequestrated by zeolite.
Taffarel et al. [93] investigated adsorption properties of Mn2+ by manganese oxide coated
zeolite (MOCZ). They studied adsorption as a function of solution pH, adsorbent concentration
and contact time. They showed that binding of Mn2+ ions onto MOCZ was highly pH dependent
with an increase in the extent of adsorption with the pH of the media. The pseudo-second-order
model fitted better among all the kinetic models suggesting the adsorption mechanism as a
chemisorption process. They found that the equilibrium data fitted excellently for both Langmuir
and Freundlich isotherm model with both monolayer adsorption and a heterogeneous surface
existence. At pH = 6, the Mn2+ uptake by MOCZ was as high as 1.1 meq Mn2+ g−1 at
equilibrium.
Doula [94] used Zeolites especially clinoptilolite and its Fe-modified form as low-cost
adsorbent for simultaneous removal of Cu2+, Mn2+ and Zn2+ from drinking water. It was found
that the Clin–Fe oxide system was capable of adsorbing significantly higher heavy metal
concentrations than untreated clinoptilolite. Effectiveness of clinoptilolite, of the Clin–Fe
system, metal–sorbent chemical behavior and the adsorption selectivity were studied. It was
found that the order of selectivity was Cu2+>Zn2+>Mn2+ and with increase in initial metal
concentration, progressive decrease in Mn2+adsorption happened.
Olguin et al. [95] used a Mexican clinoptilolite-rich tuff and its modified form in sodium
to remove iron and manganese from aqueous solutions. Maximum iron and manganese removal
was found at initial pH value of 6.0. Iron was removed better from the aqueous solutions than

23
manganese by the sodium modified zeolitic material in single systems, in binary systems, iron
and manganese compete for the active sites of the zeolitic material, decreasing the removal of
both metals. It was found that the ion exchange mechanism was responsible for manganese
removal and precipitation for iron removal at pH 6.0. The kinetic results were best adjusted to
the pseudo-second order model and the isotherms to the Freundlich isotherm model. Sodium
modified zeolitic tuff removed more manganese and iron from aqueous solution than the
unmodified material.
Sharma et al. [96] investigated fly ash, a waste material from thermal power plants for the
removal of manganese from aqueous solutions and wastewaters. The removal was found to be
highly concentration dependent and higher removal was obtained at low concentrations of Mn(II)
in the solutions. The removal decreased from 74.2 to 47.2% by increasing the Mn(II)
concentration from 1.5 to 5.0 mg/L at 298 K, pH 8.0, and 1.0 × 10−2 M NaClO4 ionic strength.
Removal, however, decreased from 51.3 to 7.2% by increasing the adsorbent particle size from
100 to 250 μm.
Dawodu et al. [97] used an unmodified Nigerian kaolinite clay (UAK) as a low-cost
adsorbent for the removal of Ni(II) and Mn(II) ions from a binary solution of both metal ions.
They evaluated the effect of solution pH, initial metal ion concentration, sorbent dose, particle
size, contact time, temperature and ligand on adsorption. They found that the Freundlich
isotherm model was the best fit. The Langmuir monolayer maximum adsorption capacities for
Ni(II) and Mn(II) ions were 166.67 mg/g and 111.11 mg/g, respectively. The Elovich equation
gave the best fit to the experimental data for both metal ions.
Yavuza et al. [98] studied the removal of heavy metals such as Mn(II), Co(II), Ni(II), and
Cu(II) from aqueous solution using a raw kaolinite. They found the Langmuir adsorption
equation as the best fit and the order of selectivity as Cu(II) > Ni(II) > Co(II) > Mn(II). They
showed that adsorption of heavy metal on kaolinite was an endothermic process and the process
of adsorption was favored at high temperatures.
Vistuba et al. [99] investigated coal as an adsorbent for the removal of iron and
manganese from aqueous solutions. They found that pseudo-second order model to fit the kinetic
data and selectivity for Fe(II) ions, as Fe(II) ions are smaller than the Mn(II) ions causing the
ferrous ions to penetrate more easily into the micro pores of the adsorbents. The results showed

24
that the adsorbent coal could be used efficiently for the removal of iron and manganese from
aqueous solutions.
Gogoi et al. [100] explored calcium hydroxyapatite (HAP) as an adsorbent for the
removal of manganese(II) from an aqueous solution as a function of contact time, pH, ionic
strength and temperature. It was found that the sorption of manganese on HAP was pH
independent ranging from 4 to 6, due to the buffering properties of HAP. The sorption of
manganese was rapid, and the percentage of manganese sorption was 99% during the initial 20–
30 min of the contact time, when the initial concentration of manganese ions was 1 mg/L at a pH
of 5. The sorption kinetics fitted with pseudo-second-order model. The maximum sorption
capacity of HAP for manganese calculated based on the Langmuir model was ~58.99 mg/g at pH
5.1 and at a temperature of 300 K. The thermodynamic parameters indicated that the manganese
sorption on HAP was an endothermic and spontaneous process. It was observed that in the
presence of bivalent ions like Co2+, the manganese sorption by HAP decreases, whereas it
remained same in the presence of monovalent ions like Na+.
Inglezakis et al. [101] used natural clinoptilolite and vermiculite as well as their Na-forms
for simultaneous removal of Fe (1.5 ppm) and Mn(II) (0.5 ppm) from underground water
samples. They found that vermiculite exhibited higher removal levels than clinoptilolite for both
Fe(II) and Mn(II). In general, Fe removal was higher than Mn(II) for vermiculite and the
opposite holds for clinoptilolite. In particular, Fe and Mn(II) removal levels were between 88–
94% and 65–100% for vermiculite and 22–90% and 61–100% for clinoptilolite, respectively.
They showed that pretreatment as well as the use of smaller particle size increased the removal
of both metals.
Arakaki et al. [102] applied vermiculite, a 2:1 clay mineral as an adsorbent for removal of
cadmium(II), zinc(II), manganese(II), and chromium(III) from aqueous solutions. They
investigated parameters such as time of reaction, effect of pH and cation concentration. The
adsorbent showed good sorption potential for these cations. The experimental data was analyzed
by Langmuir isotherm model. The quantity of adsorbed cations were 0.50, 0.52, 0.60, and
0.48 mmol g−1 of Cd2+, Mn2+, Zn2+, and Cr3+, respectively.
Akpomiea et al. [103] investigated alkaline-activated montmorillonite as a low-cost
adsorbent for simultaneous removal of Ni(II) and Mn(II) ions from solution. The effects of pH,
initial metal ion concentration, particle size, adsorbent dose, contact time and temperature on

25
adsorption were evaluated. The adsorption capacity of the montmorillonite for Ni(II) and Mn(II)
ions increased with alkaline modification. They found that equilibrium data fitted best with the
Freundlich model and the kinetic data with Elovich equation. Thermodynamic studies revealed a
spontaneous, endothermic physical adsorption process.
Mohan et al. [104] studied the removal and recovery of metal ions from acid mine
drainage (AMD) by using lignite, a low-cost sorbent. Sorption of ferrous, ferric, manganese, zinc
and calcium in multi-component aqueous systems was investigated. They found the selectivity as
Fe(II) > Zn(II) >Mn(II) due to displacement of previously adsorbed Mn(II) and Zn(II) ions by
highly retained Fe(II) ions .
2.7.3.2. Agro / plant derived adsorbents
Different agro-plant derived adsorbents have been used as adsorbents. They are
Ziziphusspina-christi seeds (ZSAC) [105], tea waste [106], bean pods waste [107], black carrot
residues [108], thermally decomposed leaf [109], sugarcane bagasse [110,111,115], T. latifolia
roots, S. americanus roots [112], sunflower stems [113], maize cob, Palm fruit bunch [114], Rice
husk [115-116], formaldehyde modified green tomato husk [117], Spirodelapolyrhiza (L.)
Schleiden [118], empty fruit bunch [119].
Omri et al. [105] used activated carbon prepared from Ziziphus spina-christi seeds
(ZSAC) to remove Mn(II) from aqueous solutions. They examined the effects of pH, initial metal
ion concentration and temperature on the adsorption performance of ZSAC for Mn(II) ions. It
was found that maximum adsorption capacity of manganese calculated from Langmuir isotherm
was about 172 mg/g.
Khajeh et al. [106] developed solid-phase tea waste as a low-cost adsorbent to remove
manganese and cobalt from water samples and found 99% and 92% removal of manganese and
cobalt, respectively. They used response surface methodology (RSM) and hybrid of artificial
neural network-particle swarm optimization (ANN-PSO) to develop predictive models for
simulation and optimization of tea waste extraction process.
Budinova et al. [107] investigated a low-cost activated carbon from bean pods waste for
the removal of As (III) and Mn(II) from aqueous solutions. They found that adsorption for both
ions followed Langmuir-type isotherm, the maximum loading capacities for arsenic (III) and
Mn(II) ions being 1.01 and 23.4 mg/g, respectively. They showed that the basic nature of the
surface favors adsorption of ions. Also, the removal of manganese ionic species increased

26
sharply between pH 2 and 4. For pH >4, the maximum uptake was attained and remained
constant as the pH of the solution becomes more basic. They found that manganese adsorption
with this carbon presented higher uptake than that of activated carbons reported in the literature.
Guzel et al. [108] studied the effect of temperature on the adsorption of Mn(II), Ni(II),
Co(II) and Cu(II) from aqueous solution by modified carrot residues (MCR). They showed that
kinetic data obtained for each heavy metal by MCR at different temperatures applied to the
Lagergren equation and isothermal fit to the Langmuir adsorption isotherm. Also, found the
order of selectivity as Cu(II) > Ni(II) > Co(II) > Mn(II).
Imaizumi et al. [109] studied the performance of Mn(II) adsorption by a novel adsorbent,
natural leaf that was partially decomposed at moderate temperature and found Mn(II) adsorption
capacity as 61–66 mg/g. They investigated various factors including adsorbent dosage, pH, and
temperature and equilibration time on Mn(II) adsorption. It was shown that a rapid equilibration
within 30 min could be achieved at pH values as low as 4.
Elwakeel et al. [110] used sugarcane bagasse based activated carbon impregnated with
phosphoric acid at 500 ˚C activation temperature and 2 h activation time. They found that the
adsorbent had high affinity for the removal of both Fe(II) or Mn(II) from aqueous medium with
uptake values of 7.01 and 5.40mg/g for Fe(II) and Mn(II), respectively, at 25˚C. Various
parameters such as pH, agitation time and speed, adsorbent dose, metal ion concentration,
temperature, and ionic strength had been studied. The kinetic and thermodynamic behavior of the
adsorption reaction indicated pseudo-second-order model and the adsorption was endothermic in
nature and mainly physical.
Esfandiar et al. [111] used sugarcane bagasse (SCB) as a low-cost biosorbent and
activated carbon (AC) to remove Mn(II). The effects of three parameters, such as pH, adsorbent
dosage and initial metal concentration on the adsorption of Mn(II) were investigated using
response surface methodology (RSM) based on Box–Behnken design. They obtained maximum
experimental Mn(II) removal efficiencies of 63% and 97% for SCB and AC, respectively.
Additionally, a chemical treatment by hydrochloric acid (HCl) had a significant effect on the
adsorption, as the maximum removal increased up to 99%.
Santos-Diaz et al. [112] studied the ability of in vitro roots cultures of Typha latifolia and
Scirpus americanus to remove metals. They found that T. latifolia roots were able to uptake 68.8
µg Pb/g, 22.1 µg Cr/g, and 1680 µg Mn/g, while the S. americanus roots removed 148.3 µg Pb/g,

27
40.7 µg Cr/g, and 4037 µg Mn/g. About 80–90% of Pb and Cr were absorbed in both cultures.
On the contrary, the Mn(II) removal was due mainly to an adsorption process (82–86%). In
comparison to the T. latifolia cultures, S. americanus cultures were two fold more efficient to
remove Pb and Cr, and three fold more efficient to remove Mn(II). Both plant species adsorbed
metals in the following order: Cr > Pb > Mn.
Sher et al. [113] used sunflower stems for the removal of Mn(II) and phenol from water.
Factors affecting the grafting reaction, such as grafting time and temperature as well as initiator
and monomer concentration, were studied. The maximum removal capacity for Mn(II) was
found with the alkali and acidic hydrolyzed grafted sunflower stem, while the charred sunflower
stem had the highest efficiency for removal of phenol from water. The removal percentage
decreased with increasing Mn(II) concentration.
Nassar et al. [114] studied the removal of toxic metals (iron and manganese) ions from
aqueous solution. The adsorbents used were low-cost materials namely; palm fruit bunch and
maize cobs. They found removal of iron ion by adsorption on palm fruit bunch and maize cobs
was in the range of 80–57%, and for manganese ion in the range of 79–50% for initial
concentration ranging between 1 and 10 ppm. In case of mixture of both metals, removal of iron
from the mixture was in the range 79–54% and for manganese was in range of 76–54%.
Dalai et al. [115] evolved low-cost methods for the removal of iron and manganese from
ground water using Rice Husk based Activated Carbon (RHAC) and Sugarcane Baggase based
Activated Carbon (SBAC). The iron and manganese were 100% removed after passing through
the filter material in both the cases.
Tavlieva et al. [116] used White rice husk ash (WRHA) to remove Mn(II) ions from
aqueous solutions. The obtained results showed that the removal of Mn(II) ions from aqueous
solutions using WRHA was a chemisorption process, where the OH- groups of SiO2 take part,
releasing H+ ions in the solution. The adsorption process was well described by the Langmuir
isotherm. Pseudo-first-order kinetic equation was found as the most appropriate kinetic model. It
was concluded that WRHA was effective and could be a cheap adsorbent for removing Mn(II)
ions from water solutions.
Olguín et al. [117] investigated the biosorption properties of the green tomato husk to
remove iron and manganese from single and binary aqueous systems. They found that the iron,
manganese or iron–manganese adsorption kinetic results best adjusted to the pseudo-second-

28
order kinetic and the Freundlich or Langmuir–Freundlich isotherm models depending on the
iron, manganese or iron–manganese aqueous systems. Modified Langmuir and Freundlich multi-
component isotherm models were also used to describe the simultaneous sorption of iron and
manganese from binary aqueous solutions. Iron and manganese were removed efficiently by the
formaldehyde modified green tomato husk in single systems; iron and manganese did not
compete for the sorption active sites of the biomass and the removal of both metals from binary
systems did not decrease. The sorption mechanisms proposed for Mn2+ were ion exchange and
complexation, while the sorption mechanisms proposed for Fe3+ were precipitation and ion
exchange.
Meitei et al. [118] investigated the potential of the free floating freshwater
macrophyte Spirodela polyrhiza (L.) as an adsorbent to remove Cu(II), Mn(II) and Zn(II) ions
from single, binary and ternary metal solution system. They found that metal adsorption was fast
and equilibrium was attained within 120 min. Langmuir isotherm best described the equilibrium
with maximum adsorption capacities of 52.6 mg/g for Cu(II), 35.7 mg/g for Mn(II) and
28.5 mg/g for Zn(II) ions. Adsorption of the specific metal ions from binary and ternary metal
solution system showed antagonistic nature due to screening effect and competition between the
metal ions for active sites on the biomass.
Amosa [119] explored a low-cost powdered activated carbon (PAC) from solid waste
such as empty fruit bunch (EFB) of palm oil mill effluent (POME) to remove Mn(II) and H 2S.
Adsorption efficiency of 95% and 90% removals for Mn(II) and H2S, respectively, was
observed. Optimization studies revealed that 5 g PAC/100 mL biologically treated POME, 200
rpm agitation speed and 60 minutes of sorption time brought about the best removal of both
Mn(II) and H2S. Equilibrium studies also revealed that Langmuir isotherm model fitted better
with the adsorption data. Also, it was concluded that the presence of other contaminant species in
the multi-component effluent can affect the final concentration obtained, especially if the affinity
of these species are greater than those of interest. EFB-based PAC could be utilized as a standard
adsorbent which could sustainably replace the expensive adsorbents usually procured for
environmental abatement processes.
2.7.3.3. Microbial and animal biomass as biosorbents
Several biomaterials including microorganisms like Pseudomonas aeruginosa AT18
[120], Trichosporoncutaneum R57 [121], animal biomass as crab shell [122] and fungi like

29
Pleurotusostreatus [123] and Saccharomyces cerevisiae [124] are potential candidates for
removal of Mn(II).
Silva et al. [120] studied the sorption of Cr, Cu, Mn and Zn by Pseudomonas aeruginosa
AT18. They found that the solution pH and ionic strength were very important factors in the
metal biosorption performance and the biosorption capacity of P. aeruginosa AT18 for Cr3+,
Cu2+, Mn2+ and Zn2+. In aqueous solution, the biosorption increased with increasing pH in the
range 5.46–7.72. It was shown that P. aeruginosa AT18 has the capacity for biosorption of the
metallic ions Cr3+, Cu2+ and Zn2+ in solutions, although its capacity for the sorption of
manganese is low (22.39 mg Mn2+/g of biomass) in comparison to the Cr3+, Cu2+ and Zn2+ ions.
However, 20% of the manganese was removed from an initial concentration of 49.0 mg/L. The
chromium level sorbed by P. aeruginosa AT18 biomass was higher than that for Cu, Mn and Zn,
with 100% removal in the pH range 7.00–7.72. The removal of Cr, Cu and Zn was also a result
of precipitation processes.
Georgieva et al. [121] prepared new hybrid materials based on polyvinyl alcohol (PVA),
3-mercaptopropyltriethoxysilane (MPTEOS) and tetraethoxysilane (TEOS), using a sol-gel
process. MPTEOS, as a precursor, ensured the presence of SH-groups, thus providing additional
binding places for cell immobilization. Based on good adhesive behavior, the hybrid materials
were tested as matrices for immobilization of Trichosporon cutaneum R57 for removal of
manganese ions from aqueous solutions. Maximum Mn2+ removal was achieved by cells
immobilized onto materials with higher MPTEOS concentration. The synthesized
PVA/TEOS/MPTEOS hybrid materials proved to be efficient for use in biosorption applications.
Vijayaraghavan et al. [122] explored the potential of crab (Portunus sanguinolentus) shell
particles for the removal of manganese(II) and zinc(II) ions from aqueous solutions. The removal
of metal ions by crab shell was found to be pH dependent, with optimum sorption occurring at
pH 6 for both Mn(II) and Zn(II). The mechanism of metal removal by crab shell was identified
as micro-precipitation of metal carbonates followed by adsorption onto chitin at the surface of
crab shell. The process of metal biosorption was rapid (90% removal in 120 min for Mn(II) and
90% removal in 90 min for Zn(II)) at an initial metal concentration of 500 mg/L. Modeling
results revealed that Mn(II) and Zn(II) kinetics data were successfully described using pseudo-
first and second order models. Furthermore, isotherm experiments revealed that crab shell
possesses high uptake capacities of 69.9 and 123.7 mg/g for Mn(II) and Zn(II), respectively.

30
Ma et al. [123] investigated the use of Pleurotus ostreatus (P. ostreatus) nano-particles
(PONP) as a new nano-biosorbent to remove Mn(II) from aqueous solution. They studied the
effects of different experiment parameters including pH of the solution, adsorbent dose, initial
Mn(II) ion concentration and contact time on adsorption capacity of PONP. The adsorption
equilibrium study exhibited that Mn(II) adsorption of PONP was better fitted by Langmuir
isotherm model. The maximum Mn(II) adsorption capacity of PONP was 130.625 mg/g at
298.15 K, which was higher than many other adsorbents. Pseudo-second-order kinetic model
was the best one to predict the sorption kinetics with a maximum adsorption capacity of PONP
attained within 30 min. Thus, they concluded that PONP has a great potential in wastewater
treatment due to its high adsorption capacity.
Fadel et al. [124] used microorganisms Saccharomyces cerevisiae as metal bioadsorbents.
S. cerevisiae F-25 in a live form was found to be good for Mn2+ and biosorbed 22.5 mg Mn2+/g
yeast biomass. They found the optimum concentrations for maximum Mn2+ biosorption by S.
cerevisiae F-25 in a live form as 4.8 mg Mn2+/g after 30 min at pH 7, agitation 150 rpm and yeast
biomass concentration 0.1 gm/L at 30 oC. Addition of other heavy metals affected the percent of
biosorbed Mn2+.
The adsorption capacities of different adsorbents used for manganese removal are
described in Table 11.
Adsorption is now the most effective and economic method for the treatment of
wastewaters containing heavy metals because it is inexpensive and locally / naturally available
material can be efficiently used for the removal of heavy metals from aqueous solutions. The
adsorption process offers flexibility in design and operation and in many cases will produce a
high-quality of treated effluent. Moreover, adsorption is sometimes reversible and the adsorbent
can be regenerated by a suitable desorption process. However, regeneration and the high cost of
activated carbon limits its use in adsorption. Also, the adsorption efficiency depends on the type
of adsorbents. Biosorption of heavy metals from aqueous solutions is a relatively new and
promising process for the removal of heavy metal from wastewaters [14].
2.8. Membrane filtration
Membrane technology is a widely accepted technology for making several qualities of
water from surface water, well water, brackish water, seawater and wastewater [125]. In
membrane filtration, a solvent is passed through a semi-permeable membrane. The size of the

31
pores in the membrane determines the membrane's permeability. It has to be carefully calculated
to reject undesirable particles, and the size of the membrane has to be designed for optimal
operating efficiency. The result is a clean and filtered fluid on one side of the membrane, with
the removed solute on the other side. Microfiltration, ultrafiltration, nanofiltration, reverse
osmosis and electrodialysis are the types of membrane filtration techniques [37]. The choice of
the suitable membrane process depends on the size of the removed contaminants and admixtures
from the water [125].
Reverse osmosis is a relatively new technology that has been used for about ten years in
water treatment. In this technique, water is forced through a semi-permeable membrane to
remove or to separate solutes from solutions. As this separation is accomplished without a phase
change, the energy requirements are low, which makes this technology extremely attractive. The
major problems in the use of these systems are associated with the delicacy of the membranes.
However, the development of improved polymers and membrane systems are overcoming this
problem rapidly. The useful life of membranes is about two years, and can be extended by using
soft water and filters, and operating at the appropriate pressure and pH ranges. Reverse osmosis
can be used positively in heated plating operations or accompanying other systems.
Unlike reverse osmosis membranes that retain all solutes including salts, ultrafiltration
membranes retain only macromolecules and suspended solids. Thus salts, solvents, and low
molecular weight organic solutes pass through ultrafiltration membrane with the permeate water.
The pressure differences across ultrafiltration membrane are negligible as salts are not retained
by the membrane. Flux rates through the membranes are relatively high; therefore lower
pressures can be used.
Nanofiltration (NF) can be positioned between reverse osmosis and ultrafiltration. It is
capable of removing hardness elements such as calcium or magnesium together with bacteria,
viruses, and color. Also, it is operated at lower pressure than reverse osmosis and as such
treatment cost is lower than reverse osmosis treatment. NF is preferred when permeate with
some residual TDS but without color, COD and hardness is acceptable. Feed water to NF should
be of similar qualities as in the case of reverse osmosis with low turbidity and colloids.
Decontamination of feed may also be necessary to remove micro-organisms [37] (Table 12).
Electrodialysis systems are based on cells built like stacks of alternately placed cationic
and anionic selective membranes across which an electric potential is applied. Cations and

32
anions migrate in different directions and become ‘trapped’ between the membranes forming
alternate cells of ion depleted and ion concentrated solutions. Like reverse osmosis, it is a new
technology that uses membranes and has similar limitations. However, it is more energy efficient
and produces a more concentrated solution. The membranes last longer because they do not work
under pressure [37].
Limited studies have been done on the removal of manganese using nanofiltration
technique [67]. Hilal et al. [129] investigated the removal of manganese using a commercial
nanofiltration membrane (NF270) and found 89% rejection. Choo et al. [130] used water-soluble
chelating polymers such as polyacrylic acid (PAA) in combination with ultrafiltration (UF), to
remove manganese from groundwater and the removal efficiency was modeled considering the
chemical equilibria. Teng et al. [131] used self-catalytic oxidation by Mn sand before micro
filtration and found 95% removal of Mn.
Choo et al. [132] investigated the removal of various levels of manganese along with
chlorine dosages from lake water using different ultrafiltration (UF) systems in conjunction with
an in-line prechlorination step. In particular, membrane fouling, caused by oxidized iron and
manganese particles, was assessed in depth with visualization of the membrane surfaces. For
feed water containing 0.5mg/L of Mn(II), only negligible amounts of manganese removal
occurred in the absence of chlorine, but with a dose of chlorine, manganese removal efficiency
increased markedly. It reached above 80% (corresponding to less than 0.1mg/L Mn(II)) at a
chlorine dosage of 3mg/L. Also, oxidized manganese is more responsible for membrane fouling
during UF with chlorination. This phenomenon was related with the kinetics of manganese
oxidation and its oxidized particles deposition inside the pores during backwashing, rather than
its accumulation on top of the membrane skin layer.
Qiu et al. [133] investigated the removal of manganese from waste water by
complexation – ultrafiltration using copolymer of maleic acid and acrylic acid combined with
polyvinyl butyral (PVB) ultrafiltration membrane. They found the rejection rate of Mn(II) was
99.6%. Bouchard et al. [134] investigated the removal of iron and manganese from groundwater
by a process that combines oxidation and microfiltration (MF), especially at high and variable
concentrations of these metals. Watanabe et al. [135] studied the performance of a hybrid MF
membrane system with circulating powdered activated carbon (PAC) and sludge. They found

33
removal efficiency of soluble manganese was almost 100% while raw water temperatures were
higher than 5°C.
Melnyk et al. [136] studied electrodialysis of manganese(II) containing solutions. It was
shown that Mn(II) cations themselves do not render any negative impact on the ion-exchange
membranes. These ions are highly movable within the cationite membrane phase and may be
easily removed during the electrodialysis desalination up to the maximum permissible
concentration allowed for the potable water. Manganese transfer across the cation-exchange
membrane essentially depends on the solution. It was concluded, that there was no need for the
thorough removal of the Mn(II) ions from solutions before their electro-membrane desalination.
Also, the manganese content gradually decreased in the solution subjected to electrodialysis. It
reaches 0.1 ppm and even less in the dialyzate.
Sadyrbaeva [137] studied a novel method of manganese(II) extraction from sulphuric
acid solutions. The method involves electrodialysis through bulk liquid membranes containing
di(2-ethylhexyl)phosphoric acid (D2EHPA) and tri-n-octylaminein1,2-dichloroethane. The
effects of the main electrodialysis parameters as well as of the composition of the liquid
membranes and aqueous solutions on the manganese(II) transportation were determined. It was
demonstrated that a practically complete removal of Mn(II) from the feed solution containing
0.01 mol/L of MnSO4 was achieved during 0.5–2.0 h of electrodialysis. Maximum stripping
degree of 88% was obtained under optimal conditions. It was shown that the liquid membranes
ensure an effective one-stage extraction of manganese(II) into diluted solutions of sulphuric,
nitric, hydrochloric, perchloric acids and water. The manganese(II) transport rate increased with
the increase in current density and initial feed concentration of Mn2+. The increase of sulphuric
acid concentration in the feed solution as well as of tri-n-octylamine (TOA) content in the liquid
membrane exerted negative influence on the manganese(II) flux. Change in D2EHPA
concentration poorly affected the metal transport rate. By varying the ratio of the current, the
time and the concentrations of the components, a practically complete removal of Mn(II) from
the feed solution and a maximum stripping degree of ~88% were achieved in the systems studied
(Table 13).
This technology has few advantages like no need for chemicals (coagulants, flocculants,
disinfectants, pH adjustment), good and constant quality of the treated water, process and plant
compactness and simple automation [47]. It can remove heavy metal ions with high efficiency,

34
but its problems such as high cost, process complexity, membrane fouling and low permeate flux
have limited their use in heavy metals removal [14].
2.9. Waste recovery
The increasing demand for manganese is prominent to a rapid depletion of its natural
sources, which has caused great interest in the recovery of this metal from both economic and
environmental aspects. Manganese recovery from industrial ore processing waste using leaching
with sulfuric acid was investigated and found 95.4% recovery from leachate [138].
Recovery from natural manganese resources (ore) and waste sources (batteries), the
necessary condition of activation of this element and extraction of it to the solution, is the change
of the oxidation level to 2+ (which is specific to MnSO4). However, the most problematic is the
change from the 4+ oxidation level (which is specific for MnO2). To change the oxidation level,
the application of a reducer that is efficient in an extremely acidic environment of the leaching
solution is required. Batteries waste, depending on the type, contains various heavy and toxic
metals being dangerous for soil and groundwater. Most types of batteries can be recycled based
on pyro- and hydrometallurgical processes.
Extraction of manganese from ores or battery waste involves the use of reductive
reagents for the transformation of MnO2 to Mn2+ ions. A series of 18 reducers has been studied in
a research [139]. The experiments of manganese extraction from the paramagnetic fraction of
Zn-C and Zn-Mn battery waste in the laboratory scale have been described for three reducers of
different origin. Oxalic acid (75%, with the lowest redox potential) attained the best result and
urea (with typical redox potential) appeared inactive. A moderate yield (50%) resulted with
extraction supported by hydrogen peroxide.
Currently, there are three methods of spent battery processing, i.e. separative
(mechanical), pyrometallurgical and hydrometallurgical (both chemical in origin). The separative
methods are frequently used for large batteries (industrial type) and as preliminary operations in
other processing technologies. They consist of mechanical loosening of the structure (body) of
the battery and separation of components of characteristic physical properties (i.e. density, size,
magnetic properties). Pyrometallurgical methods are based on material recovery (metals in
particular) by their alteration into specific condensed phases (including metallic alloys) or into
the gas phase with condensation latter. Hydrometallurgical methods are frequently based on
acidic or alkaline leaching of appropriately prepared battery wastes (after stages of mechanical

35
processing). It is followed by a series of physical and chemical operations that leads to the
separation and concentration of valuable or troublesome components between appropriate
phases. It gives commercial products or intermediates for separate technological processes
(pyrometallurgical or hydrometallurgical) or wastes. Generally, Hydrometallurgical processes
are less energy-consuming than pyrometallurgical ones. However, the waste generated in these
processes is more troublesome. The advantage of hydrometallurgical processes is that they allow
simultaneous processing of mixtures of various types of batteries.
Xiao-ming et al. [140] studied bioleaching of manganese from manganese-electrolyzed
slag using a fungus Fusarium sp. that was isolated from manganese-electrolyzed slag and
attained 71.6% bioleaching efficiency. Peluso et al. [141] recovered manganese in the form of
oxide from spent alkaline and zinc-carbon batteries employing a bio-hydrometallurgy process.
Different manganese oxide phases were obtained depending on the method used to recover the
manganese. The oxide obtained by calcination of the bio-lixiviation residue offered the lowest
specific area and the Mn3O4 phase. The oxides obtained by precipitation with KMnO4 and
electrolysis of the batteries leachate (CMO and EMO) gave similar specific areas. Blais et al.
[142] developed a process for selectively leaching zinc and manganese from pyrometallurgical
sludge produced in the steel manufacturing industry. The optimum conditions for Mn(II)
leaching were found to be a sulfuric acid concentration as 0.25 mol/L, the Na2S2O5/Mn
stoichiometry as 1, and a leaching time of 120 min with two leaching steps (Table 14).
Table 15 gives a comparison of the various treatment technologies discussed above.
3. Conclusion
Although manganese is an essential element for human life, its concentration above a
certain level in the environment can pose danger. Hence, it is necessary to remove or recover
manganese from wastewater and various wastes. A number of technologies for manganese
removal from wastes / wastewaters have been discussed and the literature survey indicates that
chemical precipitation (oxidative precipitation with SO2/air (O2) as oxidant in neutral pH range
of 6-7 with 99.5% manganese recovery), oxidation / filtration (95% Mn(II) removal with
limestone as filtration media), and adsorption (99.42% Mn(II) removal with Granular activated
carbon as adsorbent) can be effectively used for the purpose. Also, adsorption by low-cost
adsorbent (65-100% Mn(II) removal with vermiculite as adsorbent) and biosorption
(Pleurotusostreatus with capacity of 130.625 mg/g) can be an alternative. Membrane filtration,

36
which is less studied for Mn(II) removal, may also emerge as an efficient technology (100 %
Mn(II) removal efficiency with MF membrane).

Conflict of interest
No conflict of interest

37
REFERENCES

[1] P. Patnaik, Handbook of inorganic chemicals, McGraw-Hill, New York, 2002, p 538-540.

[2] P. M. Kohl, S. J. Medlar, Occurrence of Manganese in Drinking Water and Manganese


Control, IWA Publishing, USA, 2006.

[3] G. F. Pearson, G. M. Greenway, Recent developments in manganese speciation, Trends


Analyt Chem. 24 (9) (2005) 803-809.

[4] USEPA, Health Effects Support Document for Manganese, U.S. Environmental Protection
Agency Office of Water, Washington, DC EPA 822-R-03-003, 2003.

[5] USEPA, Drinking Water Health Advisory for Manganese, U.S. Environmental Protection
Agency Office of Water, Washington, DC EPA-822-R-04-003, 2004.

[6] P. Rumsby, H. Clegg, J. Jonsson, V. Benson, M. Harman, T. Doyle, L. Rushton, P. Warwick,


D. Wilkinson, Speciation of Manganese in Drinking Water, UC9780, WRc plc, UK, 2014.

[7] M. Aschner, T. R. Guilarte, J. S. Schneider, W. Zheng, Contemporary Issues in Toxicology:


Manganese: Recent advances in understanding its transport and neurotoxicity, Toxicol Appl
Pharmacol. 221 (2007) 131–147.

[8] G. N´adask´a, J. Lesn´y, I. Michal´ik, Environmental aspect of manganese chemistry, HEJ:


ENV-100702-A

[9] World Health Organization, Guidelines for drinking water quality, 2011, p.226.

[10] Indian Standard for Drinking Water -Specification IS 10500- 2012.

38
[11] A. Postwa, C. Hayes, Best Practice Guide On The Control Of Iron And Manganese In Water
Supply, IWA Publishing, London, UK, 2013.

[12] General Standards For Discharge Of Environmental Pollutants Part-A: Effluents, The
Environment (Protection) Rules, 1986, p.548.

[13] Barakat M.A., New trends in removing heavy metals from industrial wastewater, Arab. J.
Chem. 4 (2011) 361-377.

[14] F. Fu, Qi Wang, Removal of heavy metal ions from wastewaters: A review, J. Environ.
Manage. 92 (2011) 407-418.

[15] T. A. Kurniawan, G.Y.S. Chan, Wai-Hung Lo, S. Babel, Physico–chemical treatment


techniques for wastewater laden with heavy metals, Chem. Eng. J. 118 (2006) 83–98.

[16] D. Purkayastha, U. Mishra, S. Biswas, A comprehensive review on Cd(II) removal from


aqueous solution, J. Water Process Eng. 2 (2014) 105–128.

[17] V. Coman, B. Robotin, P. Ilia, Nickel recovery/removal from industrial wastes: A review,
Resour. Conserv. Recy. 73 (2013) 229– 238.

[18] D. Mohan, C.U. Pittman Jr., Arsenic removal from water/wastewater using adsorbents—A
critical review, J. Hazard. Mater. 142 (2007) 1–53.

[19] R. Molinari, P. Argurio, T. Poerio, Comparison of polyethylenimine, polyacrylic acid and


poly(dimethylamine-co-epichlorohydrin-co-ethylenediamine) in Cu2+ removal from wastewaters
by polymer-assisted ultrafiltration, Desalination 162 (2004) 217-228.

[20] T.M. Alslaibi, I. Abustan, M.A. Ahmad, A.A. Foul, Comparison of activated carbon
prepared from olive stones by microwave and conventional heating for iron (II), lead (II), and
copper (II) removal from synthetic wastewater, Environ. Prog. Sustainable Energy 33 (4) (2014)

39
1074-1085.

[21] T.M. Alslaibi, I. Abustan, M. A. Ahmad, A.A. Foul, Comparative studies on the olive stone
activated carbon adsorption of Zn2+, Ni2+, and Cd2+ from synthetic wastewater, Desalin. Water
Treat. 54 (1) (2015) 166-177.

[22] R.W. Peters, Y. Ku, D. Bhattacharyya, Evaluation of recent treatment techniques for
removal of heavy metals from industrial wastewaters, AlChE Symp. Ser. 81(243) (1985) 165–
203.

[23] W. W. Eckenfelder, Industrial Water Pollution Control, McGraw-Hill, 2000, p.151.

[24] F. Woodard, Industrial Waste Treatment Handbook, Butterworth–Heinemann, 2001, p.219.

[25] W. Zhang, C. Cheng, Manganese metallurgy review. Part II: Manganese separation and
recovery from solution, Hydrometallurgy 89 (2007) 160–177.

[26] W. Zhang, C. Cheng, Yoko Pranolo, Investigation of methods for removal and recovery of
manganese in hydrometallurgical processes, Hydrometallurgy 101 (2010) 58–63.

[27] A.M. Silva, E.C. Cunha, F.D.R. Silva, V.A. Leão, Treatment of high-manganese mine water
with limestone and sodium carbonate, J. Clean. Prod. 29-30 (2012) 11-19.

[28] K.L. Johnson, A. Baker, D.A.C. Manning, Passive treatment of Mn-rich mine water: Using
fluorescence to observe microbiological activity, Geomicrobiol. J. 22 (2005) 141–149.

[29] S.M. Bamforth, D.A.C. Manning, I. Singleton, P.L. Younger, K.L. Johnson, Manganese
removal from mine waters–investigating the occurrence and importance of manganese
carbonates, Appli. Geochem. 21 (2006) 1274–1287.

40
[30] A.M. Silva, F.L.S. Cruz, R.M.F. Lima, M.C. Teixeira, V.A. Leão, Manganese and limestone
interactions during mine water treatment, J. Hazard. Mater. 181 (2010) 514–520.

[31] J. Pakarinen, E. Paatero, Recovery of manganese from iron-containing sulfate solutions by


precipitation, Miner. Eng.24 (2011) 1421-1429.

[32] V. Menard, G.P. Demopoulos, Gas transfer kinetics and redox potential considerations in
oxidative precipitation of manganese from an industrial zinc sulfate solution with SO2/O2,
Hydrometallurgy 89(2007) 357-368.

[33] T.Nishimura, Y. Umetsu, Oxidative precipitation of arsenic(III) withmanganese (II) and


iron(II) in dilute acidic solution by ozone, Hydrometallurgy 62 (2001) 83–92.

[34] A.E. Lewis, Review of metal sulfide precipitation, Hydrometallurgy 104 (2010) 222–234.

[35] A.W. Bryson, C.H. Bijsterveld, Kinetics of the precipitation of manganese and cobalt
sulfides in the purification of a manganese sulphate electrolyte, Hydrometallurgy, 27 (1991) 75-
84.

[36] S. Mishra, V. Kumar, Co-precipitation of copper-manganese sulfide in Fe-3%Si steel, Mat.


Sci. Eng. B32 (1995) 177-184.

[37] Global good practices in industrial wastewater treatment and disposal/reuse, with special
reference to common effluent treatment plants, Central Pollution Control Board (Ministry of
Environment & Forests, Govt. of India) 7-10.

[38] X. Tang, H. Zheng, H.Teng, Y. Sun, J.Guo, W.Xiea , Q. Yang, W.Chen, Chemical
coagulation process for the removal of heavy metals from water: a review,Desalin. Water Treat.
(2014) 1–16.

41
[39] Z. Fu-wang, Li Xing, Y. Yan-ling, Study on the effect of manganese(II) removal with
oxidation and coagulation aid of potassium manganate, Bioinformatics and Biomedical Eng.,
2009. ICBBE 2009. 3rd International Conference1-4, doi: 10.1109/ICBBE.2009.5163512.

[40] L. Charemtanyarak, Heavy metals removal by chemical coagulation and precipitation, Wat.
Sci. Technol. 39(10-11) (1999)135-138.

[41] E.O. Sommerfeld, Iron and Manganese Removal Handbook, Am. Wat. Works Assoc.,
Denver, 1999.

[42] H. Polata, D. Erdogan, Heavy metal removal from waste waters by ion flotation, J. Hazard.
Mater. 148 (2007) 267–273.

[43] O. J. Gregory, S. M. Barnett, F. J. Deluise, Manganese removal from water by a precipitate


flotation technique, Sep. Sci. Technol. 15 (8) (1980) 1499-1512.

[44] L. Stoica, M. Dinculescu, C.G. Plapcianu, Mn (II) recovery from aqueous systems by
flotation, Wat. Res. 32 (10) (1998) 3021-3030.

[45] D. A. White,A. Asfar-Siddique, Removal of manganese and iron from drinking water using
hydrous manganese dioxide, Solvent Extr. Ion Exc. 15 (6) (1997) 1133-1145.

[46] O.N. Kononova, G.L. Bryuzgina, O.V. Apchitaeva, Y.S. Kononov, Ion exchange recovery
of chromium (VI) and manganese (II) from aqueous solutions, Arab. J. Chem. (2015), In Press,
Corrected Proof, doi: http:// dx.doi.org/10.1016/j.arabjc.2015.05.021.

[47] S.Chaturvedi, P.N. Dave, Removal of iron for safe drinking water, Desalination 303 (2012)
1–11.

[48] P. Roccaro, C. Barone, G. Mancini, F.G.A. Vagliasindi, Removal of manganese from water
supplies intended for human consumption: a case study, Desalination 210 (2007) 205–214.

42
[49] M.G. Salema, M. El-Awady, E. Amin, Enhanced removal of dissolved iron and manganese
from nonconventional water resources in Delta District, Egypt, Energy Procedia 18 (2012) 983 –
993.

[50] D. Ellis, C. Bouchard, G. Lantagne, Removal of iron and manganese from groundwater by
oxidation and microfiltration, Desalination 130 (2000) 255-264.

[51] P. Phatai, J. Wittayakun, W. Chen, C.M. Futalan, N. Grisdanurak, C. Kan, Removal of


manganese(II) and iron(II) from synthetic groundwater using potassium permanganate, Desalin.
Water Treat. 52 (2014) 5942–5951.

[52] B. Lesley, H. Daniel, Y. Paul, Iron and manganese removal in wetland treatment systems:
Rates, processes and implications for Management, Sci. Total Environ. 394 (2008) 1–8.

[53] H.E. Xuwen, Y. Huimin, H.E. Yong, Treatment of mine water high in Fe and Mn by
modified manganese sand, Mining Sci. Technol. 20 (2010) 0571–0575.

[54] J. K. Piispanen, J.T. Sallanko, Mn(II) removal from groundwater with manganese oxide-
coated filter media, J. Environ. Sci. Health 45 (2010) 1732–1740.

[55] H.A. Aziz, P.G. Smith, Removal of manganese from water using crushed dolomite filtration
technique, Wat. Res. 30 (2) (1996) 489-492.

[56] V.A. Pacini, A.M. Ingallinella, G. Sanguinetti, Removal of iron and manganese using
biological roughing up flow filtration technology, Wat. Res. 39 (2005) 4463–4475.

[57] A. Gouzinis, N. Kosmidis, D.V. Vayenas, G. Lyberatos, Removal of Mn and simultaneous


removal of NH3, Fe and Mnfrom potable water using a trickling filter, Wat. Res. 32 (8) (1998)
2442-2450.

43
[58] A.G. Tekerlekopoulou, D.V. Vayenas, Ammonia, iron and manganese removalfrom potable
water using trickling filters, Desalination 210 (1–3) (2007) 225-235.

[59] A.G. Tekerlekopoulou, I.A. Vasiliadou, D.V. Vayenas, Biological manganese removal from
potable water using trickling filters, Bio. Chem. Eng. J. 38 (2008) 292-301.

[60] M. Han, Z. Zhao, W.Gao, F.Cui, Study on the factors affecting simultaneous removal of
ammonia and manganese by pilot-scale biological aerated filter (BAF) for drinking water pre-
treatment, Bioresour. Technol. 145 (2013) 17–24.

[61] H.A. Hasan, S.R.S. Abdullah, S.K. Kamarudin, N.T. Kofli, N. Anuar, Kinetic evaluation of
simultaneous COD, ammonia and manganese removal from drinking water using a biological
aerated filter system, Sep. Purif. Technol. 130 (2014) 56–64.

[62] M. S. Burger, S.S. Mercer, G.D. Shupe, G. A. Gagnon, Manganese removal during bench-
scale biofiltration, Wat. Res. 42 (19) (2008) 4733-4742.

[63] A. Carrière, M. Brouillon, S. Sauvé, M.F. Bouchard, B. Barbeau, Performance of point-of-


use devices to remove manganese from drinking water, J. Environ. Sci. Health A 46 (2011) 601–
607.

[64] J. Xu, G. Chen, X. Huang, G. Li, J. Liu, N. Yang, S. Gao, Iron and manganese removal by
using manganese ore constructed wetlands in the reclamation of steel wastewater, J. Hazard.
Mater. 169 (2009) 309–317.

[65] A. Shafaei, M. Rezayee, M. Arami, M. Nikazar, Removal of Mn2+ ions from synthetic
wastewater by electrocoagulation process, Desalination 260 (2010) 23–28.

[66] M. Ince, Treatment of manganese-phosphate coating wastewater by electrocoagulation, Sep.


Sci. Technol. 48 (2013) 515–522.

44
[67] B. Al-Rashdi, C. Somerfield, N. Hilal, Heavy metals removal using adsorption and
nanofiltration techniques, Sep. Purif. Rev. 40 (2011) 209–259.

[68] M.Anbia, S.Amirmahmoodi, Removal of Hg (II) and Mn (II) from aqueous solution using
nanoporous carbon impregnated with surfactants, Arab. J.Chem.(2011),In Press, Corrected
Proof,doi:10.1016/j.arabjc.2011.04.004.

[69] A. Jusoh, W.H. Cheng, W.M. Low, A. Nora’aini, M.J. Megat, M. Noor, Study on the
removal of iron and manganese in groundwater by granular activated carbon, Desalination 182
(2005) 347–353.

[70] K. Motojima, E. Tachikawa, H. Kamiyama, Elimination of manganese-54 in waste water by


oxine-impregnated activated charcoal, J. Nucl. Sci. Technol. 16 (5) (1979) 356-362.

[71] M.E. Goher , A.M. Hassan , I.A. Abdel-Moniem, A.H. Fahmy , M.H. Abdo , S.M. El-sayed,
Removal of aluminum, iron and manganese ions from industrial wastes using granular activated
carbon and Amberlite IR-120H, Egypt. J. Aquat. Res.41(2) (2015) 155-164.

[72] E.Okoniewska, J. Lach, M. Kacprzak, E. Neczaj, The removal of manganese, iron, and
ammonium nitrogen on impregnated activated carbon, Desalination 206 (2007) 251–258.

[73] E.A. Moawed, N. Burham, M.F. El-Shahat, Separation and determination of iron and
manganese in water using polyhydroxyl polyurethane foam, J. Assoc. Arab Univ. Basic Appli.
Sci. 14 (2013) 60 –66.

[74] K.Z. Al-Wakeel, H. Abd El Monem, M.M.H. Khalil, Removal of divalent manganese from
aqueous solution using glycine modified chitosan resin, J. Environ. Chem. Eng. 3 (2015) 179–
186.

45
[75] Z. Abdeen, S.G. Mohammad, M.S. Mahmoud, Adsorption of Mn (II) ion on polyvinyl
alcohol/chitosan dry blending from aqueous solution, Environ. Nanotechnol. Monit. Manage. 3
(2015) 1–9.

[76] R. Xu, G. Zhou, Y. Tang, L. Chu, C. Liu, Z. Zeng, S. Luo, New double network hydrogel
adsorbent: Highly efficient removal of Cd(II) and Mn(II) ions in aqueous solution, Chem. Eng. J.
275 (2015) 19-188, doi: http://dx.doi.org/10.1016/j.cej.2015.04.040.

[77] M.H. Qomi, H. Eisazadeh, M. Hosseini, H.A. Namaghi, Manganese removal from aqueous
media using polyanilinenanocomposite coated on wood sawdust, Synthetic Met. 194 (2014)
153–159.

[78] N.A. Reiad, O.E.A. Salam, E.F. Abadir, F.A. Harraz, Adsorptive removal of iron and
manganese ions from aqueous solutions with microporous chitosan/polyethylene glycol blend
Membrane , J. Environ. Sci. 24 (8) (2012) 1425–1432.

[79] B. Guan, W. Ni, Z. Wu, Y. Lai, Removal ofMn(II) and Zn(II) ions from flue gas
desulfurization wastewater with water-soluble chitosan, Sep. Purifi. Technol. 65 (2009) 269–274.

[80] V. K. Gupta, P.J.M. Carrott, M.M.L. Ribeiro Carrott, Suhas, Low-Cost Adsorbents:
Growing Approach to Wastewater Treatment—a Review, Environ. Sci. Technol., 39:10 (2009)
783-842.

[81] T.A. Kurniawan, G.Y.S. Chan, Wai-Hung Lo, S. Babel, Review-Comparisons of low-cost
adsorbents for treating wastewaters laden with heavy metals, Sci. Total Environ. 366 (2006)
409–426.

[82] T.M. Alslaibi, I. Abustan, M.A. Ahmad, A.A. Foul, A review: production of activated
carbon from agricultural byproducts via conventional and microwave heating, J. Chem. Technol.
Biotechnol. 88 (7) (2013) 1183-1190.

46
[83] T.M. Alslaibi, I. Abustan, M.A. Ahmad, A.A. Foul, Preparation of activated carbon from
olive stone waste: optimization study on the removal of Cu2+, Cd2+, Ni2+, Pb2+, Fe2+, and Zn2+
from aqueous solution using response surface methodology, J. Dispersion Sci. Technol. 35 (7)
(2014) 913-925.

[84] J.U.K. Oubagaranadin, N. Sathyamurthy, Z.V.P. Murthy, Evaluation of Fuller’s earth for the
adsorption of mercury from aqueous solutions: A comparative study with activated carbon, J.
Hazard. Mater.142 (2007) 165–174.

[85] M.U. Khobragade, A. Pal, Investigation on the adsorption of Mn(II) on surfactant-modified


alumina: Batch and column studies, J. Environ. Chem. Eng. 2 (2014) 2295–2305.

[86] M.U. Khobragade, A. Pal, Adsorptive removal of Mn(II) from water and wastewater by
surfactant-modified alumina, Desalin. Water Treat. (2014) 1–12.

[87] S.R. Taffarel, J. Rubio, On the removal of Mn2+ ions by adsorption onto natural and
activated Chilean zeolites, Miner. Eng. 22 (2009) 336–343.

[88] M.J. Zamzow, J.E. Murphy, Removal of metal cations from water using zeolites, Sep. Sci.
Technol. 27 (14) (1992) 1969-1984.

[89] N. Rajic, D. Stojakovic, S. Jevtic, N.Z. Logar, J. Kovac, V. Kaucic, Removal of aqueous
manganese using the natural zeolitic tuff from the VranjskaBanja deposit in Serbia, J. Hazard.
Mater. 172 (2009) 1450–1457.

[90] A. Ates, Role of modification of natural zeolite in removal of manganese from aqueous
solutions, Powder Technol. 264 (2014) 86 –95.

[91] M.A. Shavandi, Z. Haddadian, M.H.S. Ismail, N. Abdullah, Z.Z. Abidin, Removal of
Fe(III), Mn(II) and Zn(II) from palm oil mill effluent (POME) by natural zeolite,J. Taiwan Inst.
Chem. Eng. 43 (2012) 750–759.

47
[92] C. Belviso, F. Cavalcante, S.D. Gennaro, A. Lettino, A. Palma, P. Ragone, S. Fiore,
Removal of Mn from aqueous solution using fly ash and its hydrothermal synthetic zeolite, J.
Environ. Manage. 137 (2014) 16-22.

[93] S.R. Taffarel, J. Rubio, Removal of Mn2+ from aqueous solution by manganese oxide coated
zeolite, Miner. Eng. 23 (2010) 1131–1138.

[94] M. K. Doula, Simultaneous removal of Cu, Mn and Zn from drinking water with the use of
clinoptilolite and its Fe-modified form, Wat. Res. 43 (2009) 3659– 3672.

[95] A. García-Mendieta, M. Solache-Ríos, M.T. Olguín, Evaluation of the sorption properties of


a Mexican clinoptilolite-rich tuff for iron, manganese and iron–manganese systems, Micropor.
Mesopor. Mat. 118 (2009) 489–495.

[96] Y.C. Sharma, Uma, S.N. Singh, Paras, F. Gode, Fly ash for the removal of Mn(II) from
aqueous solutions and wastewaters, Chem. Eng. J. 132 (2007) 319–323.

[97] F.A. Dawodu, K.G. Akpomie, Simultaneous adsorption of Ni(II) and Mn(II) ions from
aqueous solution onto a Nigerian kaolinite clay, J. of Mater. Res. Technol. 3 (2) (2014) 129–
141.

[98] O. Yavuza, Y. Altunkaynak, F.G. Uzel, Removal of copper, nickel, cobalt and manganese
from aqueous solution by kaolinite, Wat. Res. 37 (2003) 948–952.

[99] J.P. Vistuba, M. E. Nagel-Hassemer, F.R. Lapolli, M. Á. L. Recio, Simultaneous adsorption


of iron and manganese from aqueous solutions employing an adsorbent coal, Environ. Technol.
34 (2) (2013) 275–282.

48
[100] D. Gogoi, A.G. Shanmugamani, S.V.S. Rao, T. Kumar, S. Velmurugan, Study of removal
process of manganese using synthetic calcium hydroxyapatite from an aqueous solution, Desalin.
Water Treat. (2015) 1–8.

[101] V.J. Inglezakis, M.K. Doula, V. Aggelatou, A.A. Zorpas, Removal of iron and manganese
from underground water by use of naturalMinerals in batch mode treatment, Desalin. Water
Treat. 18 (2010) 341–346.

[102] M.G. da Fonseca, M.M. de Oliveira, L.N.H. Arakaki, Removal of cadmium, zinc,
manganese and chromium cations from aqueous solution by a clay mineral, J. Hazard. Mater.
B137 (2006) 288–292

[103] K.G. Akpomiea, F.A. Dawodu, Efficient abstraction of nickel(II) and manganese(II) ions
from solution onto an alkaline-modified montmorillonite, J. Taibah Univ. Sci. 8 (2014) 343–356.

[104] D. Mohan, S. Chande, Removal and recovery of metal ions from acid mine drainage using
lignite-A low-cost sorbent, J.Hazard.Mater. B137 (2006) 1545–1553.

[105] A.Omri, M.Benzina, Removal of manganese(II) ions from aqueous solutions by adsorption
on activated carbon derived a new precursor: Ziziphusspina-christi seeds, Alexandria Eng. J.
51(2012) 343–350.

[106] M. Khajeh, A. Sarafraz-Yazdi, A.F. Moghadam, Modeling of solid-phase tea waste


extraction for the removal of manganese and cobalt from water samples by using PSO-artificial
neural network and response surface methodology, Arab.J. Chem. (2013), In Press, Corrected
Proof, doi:10.1016/j.arabjc.2013.06.011.

[107] T. Budinova, D. Savova, B. Tsyntsarski, C.O. Ania, B. Cabal, J.B. Parra, N. Petrov,
Biomass waste-derived activated carbon for the removal of arsenic and manganese ions from
aqueous solutions, Appli. Surf. Sci. 255 (2009) 4650–4657.

49
[108] F. Guzel, H. Yakut, G. Topal, Determination of kinetic and equilibrium parameters of the
batch adsorption of Mn(II), Co(II), Ni(II) and Cu(II) from aqueous solution by black carrot
(Daucuscarota L.) residues, J. Hazard. Mater. 153 (2008) 1275–1287.

[109] Z. Li, S. Imaizumi, T. Katsumi, T. Inui, X. Tang, Q. Tang, Manganese removal from
aqueous solution using a thermally decomposed leaf, J. Hazard. Mater. 177 (2010) 501–507.

[110] K.Z. Elwakeel, G.O. El-Sayed, S.M. Abo El-Nassr, Removal of ferrous and manganous
from water by activated carbon obtained from sugarcane bagasse,Desalin. Water Treat. (2014)
1–13.

[111] N. Esfandiar, B. Nasernejad, T. Ebadi, Removal of Mn(II) from groundwater by sugarcane


bagasse and activated carbon (a comparative study): Application of response surface
methodology (RSM), J. Ind. Eng. Chem. 20 (5) (2014) 3726-3736,
doi: 10.1016/j.jiec.2013.12.072.

[112] M. del Socorro Santos-Díaz, M. del Carmen Barrón-Cruz, Lead, chromium and manganese
removal by in vitro root cultures of two aquatic macrophytes species: Typha Latifolia L. and
Scirpus Americanus Pers, Int. J. Phytoremediation 13 (2011) 538–551.

[113] Samir K., Ibrahim M.B.M., A.K. Sher, Decontamination of manganese and phenol from
aqueous media by sunflower stems, Int. J. Polym. Mater. 58(2009)533–547.

[114] M.M. Nassar, K.T. Ewida, E.E. Ebrahiem, Y.H. Magdy, M.H. Mheaedi, Adsorption of iron
and manganese using low-cost materials as adsorbents, J. Environ. Sci. Health 39 (2) (2004)
421–434.

[115] C. Dalai, R. Jha, V.R. Desai, Rice husk and sugarcane bagasse based activated carbon for
iron and manganese removal, Aquat. Procedia 4 (2015) 1126 – 1133.

50
[116] M.P. Tavlieva , S.D. Genieva , V.G. Georgieva , L.T. Vlaev, Thermodynamics and kinetics
of the removal of manganese(II) ions from aqueous solutions by white rice husk ash, J. Mol. Liq.
211 (2015) 938–947.

[117] A. García-Mendieta, M.T. Olguín, M. Solache-Ríos, Biosorption properties of green


tomato husk (Physalisphiladelphica Lam) for iron, manganese and iron–manganese from
aqueous systems, Desalination 284 (2012) 167 –174.

[118] M.D. Meitei, M.N.V. Prasad, Adsorption of Cu (II), Mn (II) and Zn (II)
by Spirodelapolyrhiza (L.) Schleiden: Equilibrium, kinetic and thermodynamic studies, Ecol.
Eng. 71 (2014) 308–317.

[119] M. K. Amosa, Process optimization of manganese and H2S removals from palm oil mill
effluent (POME) using enhanced empty fruit bunch (EFB)-based powdered activated carbon
(PAC) producedfrom pyrolysis,Environ. Nanotechnol. Monit. Manage. (2015),
doi:http://dx.doi.org/doi:10.1016/j.enmm.2015.09.002.

[120] R.M.P. Silva, A.Á. Rodríguez , J.M.G.M. DeOca , D.C. Moreno, Biosorption of
chromium, copper, manganese and zinc by Pseudomonas aeruginosa AT18 isolated from a site
contaminated with petroleum, Bioresour. Technol. 100 (2009) 1533–1538.

[121] N.Georgieva, R. Bryaskova, N. Lazarova, R. Racheva, Immobilization of


TrichosporonCutaneum R57 on PVA/TEOS/MPTEOS hybrid matrices for removal of
manganese ions, Biotechnol. Biotec. Eq. 27 (5) (2013) 4078-4081.

[122] K. Vijayaraghavan, H.Y.N. Winnie, R. Balasubramanian, Biosorption characteristics of


crab shell particles for the removal of manganese(II) and zinc(II) from aqueous solutions,
Desalination 266 (2011) 195–200.

[123] L. Ma, Y. Peng, Bo Wu, D. Lei, H. Xu, Pleurotusostreatus nanoparticles as a new nano-
biosorbent for removal of Mn(II) from aqueous solution, Chem. Eng. J. 225 (2013) 59–67.

51
[124] M. Fadel , N.M. Hassanein, M.M. Elshafei, A.H. Mostafa, M.A. Ahmed, H.M. Khater,
Biosorption of manganese from groundwater by biomass of Saccharomyces cerevisiae, HBRC J.
(2015), In Press, Corrected Proof, doi: 10.1016/j.hbrcj.2014.12.006.

[125] M. Bodzek, K. Konieczny, A. Kwiecińska, Application of Membrane processes in drinking


water treatment–state of art, Desalin. Water Treat. 35 (2011) 164–184.

[126] H.A. Qdaisa, H. Moussa, Removal of heavy metals from wastewater by membrane
processes: a comparative study, Desalination 164 (2004) 105-110.

[127] N. Hilal, H. A1-Zoubi, N.A. Darwish , A.W. Mohammad, M. Abu Arabi,A comprehensive
review of nanofiltration membranes: Treatment, pretreatment, modelling, and atomic force
microscopy, Desalination 170 (2004) 281-308.

[128] S. Jamaly, N.N. Darwish, I. Ahmed, S.W. Hasan, A short review on reverse osmosis
pretreatment technologies, Desalination 354 (2014) 30–38.

[129] B.A.M. Al-Rashdi, D.J. Johnson, N. Hilal, Removal of heavy metal ions by nanofiltration,
Desalination 315 (2013) 2 –17.

[130] S. Han, K. Choo, S. Choi, M.M. Benjamin, Modeling manganese removal in chelating
polymer-assistedMembrane separation systems for water treatment, J. Membr. Sci. 290 (2007)
55–61.

[131] Z. Teng, J.Y. Huang, K. Fujita, S. Takizawa, Manganese removal by hollow fiber micro-
filter. Membrane separation for drinking water. Desalination 139 (2001) 411-418.

[132] K. Choo, H. Lee, S. Choi, Iron and manganese removal and membrane fouling during UF
in conjunction with prechlorination for drinking water treatment, J. Membr. Sci. 267 (2005) 18–
26.

52
[133] Y. Qiu, U.L. Mao, W. Wang, Removal of manganese from waste water by
complexation−ultrafiltration using copolymer of maleic acid and acrylic acid, Trans. Nonferrous
Met. Soc. China 24 (2014) 1196−1201.

[134] D. Ellis, C. Bouchard, G. Lantagne, Removal of iron and manganese from groundwater by
oxidation and microfiltration, Desalination 130 (2000) 255-264.

[135] T. Suzuki, Y. Watanabe, G. Ozawa, S. Ikeda, Removal of soluble organics and


manganese by a hybrid MF hollow fiber membrane system, Desalination 117 (1998) 119-130.

[136] L. Melnyk, V. Goncharuk, Electrodialysis of solutions containing Mn (II) ions,


Desalination 241 (2009) 49-56.

[137] T. Zh. Sadyrbaeva, Hybrid liquid membrane — Electrodialysis process for extraction of
manganese(II), Desalination 274 (2011) 220–225.

[138] O. de S.H. Santos, C. de F.Carvalho, G.A. da Silva, C.G. d. Santos, Manganese ore tailing:
Optimization of acid leaching conditions and recovery of soluble manganese, J. Environ.
Manage. 147 (2015) 314-320.

[139] A. Sobianowska–Turek, W łodzimierz Szczepaniak, M. Zabłocka-Malicka,


Electrochemical evaluation of manganese reducers - Recovery of Mn from Zn-Mn and Zn-C
battery waste, J. Power Sources 270 (2014) 668-674.

[140] C. Jian-bing, Li Xiao-ming, O. Yu-zhu, Z. Wei, W.g Dong-bo, S. Ting-ting, Y. Xiu, Y. Qi,
Manganese-electrolysed slag treatment: bioleaching of manganese by Fusarium sp., Environ.
Technol. 33(11)(2012) 1307–1312.

[141] M.V. Gallegos, L.R. Falco, M.A. Peluso, J.E. Sambeth, H.J. Thomas, Recovery of
manganese oxides from spent alkaline and zinc–carbon batteries. An application as catalysts for
VOCs elimination, Waste Manage. 33 (2013) 1483–1490.

53
[142] J. Mocellin, G. Mercier, J.L. Morel, J.F. Blais, M.O. Simonnot, Factors influencing the Zn
and Mn extraction from pyrometallurgical sludge in the steel manufacturing industry, J. Environ.
Manage. 158 (2015) 48-54.

54
Table 1
Standards for Manganese levels in drinking water in selected countries
Country Standard value in mg/L
Canada 0.05
Australia 0.05
Japan 0.05
South Africa 0.1
Taiwan 0.05
India 0.1
Brazil 0.1

55
Table 2
Manganese removal by Hydroxide precipitation
pH Precipitant Initial Residual Removal Comments References
Mn(II) Mn(II) Efficiency
conc. conc. (%)
mg/L mg/L
8.2 NaOH 1791 534 71.4 Synthetic 26
laterite waste
solution
9 NaOH 1791 10 99.5 Synthetic 26
laterite waste
solution
11 NaOH 1791 1 99.96 Synthetic 26
laterite waste
solution

56
Table 3
Manganese removal by carbonate precipitation
pH Precipitant Initial Residual Removal Comments References
Mn(II) Mn(II) Efficiency
conc. conc. (%)
mg/L mg/L
8 Na2CO3 1764 164 90.7 Synthetic laterite 26
waste solution
8.5 Na2CO3 1764 9 99.5 Synthetic laterite 26
waste solution
9.5 Na2CO3 1764 1 99.9 Synthetic laterite 26
waste solution
>8.5 Na2CO3+ 140 0.14 99.9 Mine water 27
limestone
7 Lower Hart 15-30 NA 94 Mine water 28
dolomite
8.8 Calcite 16.5 <1 NA Mine water 30
limestone

57
Table 4
Manganese removal by oxidative precipitation
pH Precipitant Initial Residual Removal Comments References
Mn(II) Mn(II) Efficiency
conc. conc. (%)
mg/L mg/L
6-7 SO2/air(O2) NA <10 99.5 Synthetic laterite 26
waste solution

<4 SO2/air(O2) NA NA 98 Zinc sulphate 32


solution

58
Table 5
Manganese removal by sulfide precipitation
pH Initial Residual Removal Comments References
Mn(II) conc. Mn(II) conc. Efficiency
mg/L mg/L (%)

7 32.8 NA 75 Precipitant- ammonium 35


sulfide
Temp.: 35 °C

59
Table 6
Manganese removal by coagulation / flocculation
pH Initial Residual Removal Comments References
Mn(II) conc. Mn(II) conc. Efficiency
mg/L mg/L (%)
9.2 NA NA 82 Synthetic Manganese 39
sulphate solution

11 1085 1.8 99.83 Synthetic wastewater 40


Aluminum chloride as
coagulant with
KMnO4 as coagulant
aid

60
Table 7
Manganese removal by flotation
pH Initial Residual Removal Comments References
Mn(II) conc. Mn(II) conc. Efficiency(
mg/L mg/L %)
8-12 5.5 1.5 96 Precipitate flotation 43

10.5 500 NA 99.2 Precipitate flotation 44


0.1 M NaOH
solution as reagent

61
Table 8
Manganese removal by ion exchange
pH Initial Mn(II) Residual Removal Comments References
conc. Mn(II) Efficiency (a) /
mg/L conc. recovery (b) %
mg/L
NA 5000 NA >90 (a) MnCl2Feed solution 46
(20 ± 1) oС
resin masses of 0.1– 0.2 g
time: 24h
Exchangers synthesized
with long-chained cross-
linking agents (LCA) like
divinyl esters of diethylene-
glycol (DVEDEG),
triethylene-glycol
(DVETEG), propylene-
glycol (DVEPG) and
tetravinyl ester of
pentaerythritol (TVEPE).

NA 5000 NA >99.8 (b) Simultaneous sorption of 46


chromium (VI) and
manganese (II)
Cr(VI) initial conc.: 1g/L
Recovery of Cr(VI):
>99.9%.

62
Table 9
Manganese removal by oxidation / filtration
pH Initial Residual Mn(II) Removal Comments References
Mn(II) conc. conc. mg/L Efficiency(
mg/L %)
8.5 1.810 NA 95 KMnO4 oxidation 48
followed by flocculation,
settling and filtration.
Oxidant dose:
1.74 mg/L, Batch
experiment

NA 0.5 - 1.45 0.05 96 NaOCl oxidation 49


Simultaneous removal of
Fe: 92%
8.0 NA NA 30.6% Synthetic ground water 51
KMnO4 oxidation
Oxidant dose:
0.603 mg/L
Dual system(With Fe):
37.2% removal
7.79 NA ≤0.4 76.8 Wetland system 52
For Fe: 98.4 removal
NA 1.97 0.01 NA Mine water, Manganese 53
sand modified by KMnO4

>9 0.59-0.67 <0.02 >91 Manganese oxide coated 54


sand/antracite
7.27 1 NA 95 Filter media: limestone 55

7-7.3 0.18-0.37 0.05 95 Biological roughing up 56


flow filters
7-7.3 1-2 <0.02 94 Biological trickling filter 57

7-7.3 0.6-2.0 NA 100 Biological trickling filter 58

6-7 1-6 0.32-4.18 94.7 Biological aerated filter 61


6.5 0.3 NA >90 bio-filters inoculated with 62
L. discophora sp-6 and
indigenous mixed bio-film

8.5 1 <0.1 >60 Pour-through filters 63


7.3- 0.11-2.23 <0.05 81 Manganese ore 64
9.7 constructed wetland

63
Table 10
Manganese removal by electrochemical treatment
Method Current Initial Removal Optimum References
Density Mn(II) conc. Efficiency (%) pH
mg/L

EC 1.5-9.4 100 87.9 7.0 65


mA/cm2

EC 20 Am-2 115 97.95 3.12 66

64
Table 11
Adsorbents used for manganese removal from wastewater
Adsorbent pH Initial Mn(II) Removal Experimental conditions References
Conc. Efficiency % or
Adsorption
Capacity
Carbon Impregnated with 7 50 82.2% Temp.: 70 ± 0.5 oC 68
cetyltrimethyl ammonium 43 mg/g Time: 420 min
bromide (CTAB)
Carbon Impregnated with 7 50 70.5% Temp.: 70 ± 0.5 oC 68
sodium dodecyl sulfate 47 mg/g Time: 420 min
Unmodified mesoporous 7 50 56.8% Temp.: 70 ± 0.5 oC 68
carbon 40 mg/g Time: 420 min
Granular activated carbon NA NA 2.5451 mg/g Room temp. 69
Time: 6 h
Adsorbent dosage:
0.2-0.9 gm
Granular activated carbon 7 2 79.05 %, 2 mg/g Room temp. 71
Time: 30 min
Adsorbent dosage: 2 g/L
Amberlite IR-120H 7 2 96.65 %, Room temp. 71
4.9 mg/g Time: 30 min
Adsorbent dosage: 2 g/L
Polyhydroxyl-polyurethane 6-8 5 99-100%, Temp.: 25 oC 73
foam 8.13 μmol/g Time: 30 min
Adsorbent dosage:
0.1 g/25 ml
Glycine modified Chitosan 6 5x10-3M 1.3mmol/g Temp.: 25 oC 74
Time: 150 min
Polyvinyl alcohol/chitosan 5 5-100 84.5%, Temp.: 30 ± 1◦C 75
(PVA/CS) hydrogel 10.515 mg/g Time: 5-120 min
Adsorbent dosage:
0.1-1 g/100mL
Poly (sodium acrylate)- 6 20-600 84%, Temp.: 303K 76
graphene oxide (PSA-GO) 165.5 mg/g Adsorbent dosage:
double network hydrogel 1mg/mL
adsorption capacity for
Cd(II): 238.3mg/g
Polyaniline/sawdust composite 10 100 90.28%, Time: 30 min 77
58.824 mg/g Adsorbent dosage: 10 g/L
Chitosan/polyethylene glycol 2.9- 2-10 18 mg/g Temp.: 27 ± 2◦C 78
blend membrane 5.9 Time: 60 min
Agitation speed: 300 rpm
Water-soluble chitosan 5-9 0.45-363 94.59% Temp.: 20◦C 79
mmol/L Adsorbent dosage:
400 mg/L

65
Agitation speed: 300 rpm

Surfactant Modified Alumina 4.04- 20-100 74%, Temp.: 303K 85


8.05 2.04 mg/g Agitation speed:
110-150rpm,
Contact time: 30 min
Natural Zeolite 6-6.8 3.86 meq/L 0.26 meq/g Temp.: 20oC 87
Contact time: 120 min
Adsorbent dosage: 2.5g/L
Natural Zeolite activated with 6-6.8 3.64 meq/L 0.77 meq/g Temp.: 20oC 87
NaCl Contact time: 120 min
Adsorbent dosage: 2.5 g/L
Natural Zeolite 6 25-250 7.68 mg/g Temp.: 25 oC 90
Contact time: 240 min
Agitation speed: 200rpm

Al- Zeolite 6 25-250 25.12 mg/g Temp.: 25 oC 90


Contact time: 240 min
Agitation speed: 200rpm
NH4-Zeolite 6 25-250 24.33 mg/g Temp.: 25 oC 90
Contact time: 24 h
Agitation speed: 200rpm
Slovakian natural zeolite 3-9 14.447 52.877%, Temp.: 25 oC 91
0.075 mg/g, Contact time: 10-240 min
Agitation speed:
90-200rpm
Adsorbent dosage:
1.25-30 mg/250mL
Synthetic Zeolite NA 10 97% Temp.: 25 oC 92
Contact time: 5-240 min
Adsorbent dosage:
1g/50mL
Manganese oxide-coated 6.0 25-600 1.1 meq/g Temp.: 20oC 93
zeolite Contact time: 120 min
Adsorbent dosage: 6 g/L
Mexican Clinoptilolite 6.0 10 97.3% Temp.: 293K 95
0.0350 ±0.0002 Contact time: 5-30 min
meq/g Agitation speed:
90-200 rpm
Adsorbent dosage:
1.25-30 mg/250mL
Fly Ash 8.0 1.5 74.2 % Temp.: 298K 96
Agitation speed: 100rpm

66
Nigerian Kaolinite Clay 2-8 100-500 111.11 mg/g Temp.: 300K 97
Contact time: 180- min
Agitation speed:
90-200 rpm
Adsorbent dosage:
0.1g/100mL
Kaolinite NA 1029 0.446 mg/g Temp.: 25 oC 98
Agitation speed: 3500rpm
Adsorbent dosage:
1 g/100mL
Coal 6.85- 0.05-100 63% Temp.: 25 oC 99
7.35 Contact time:120 min
Adsorbent dosage:
0.2g/50mL
Calcium hydroxyapatite 4-6 1.0 58.9 g/g, Temp.: 300K 100
99% Agitation speed: 50rpm
Contact time:10 -2500min
Clinoptilolite 7.1± 0.5 61-100% Temp.: 25 oC 101
2
Vermiculite 7.1± 0.5 65-100% Temp.: 25 oC 101
2
Vermiculite, Clay Mineral NA 0.01 mol/L 0.52 mmol/g Temp.: 298±1K 102
Contact time: 10-100 min
Alkaline-modified 2-8 100-500 111.95 mg/g Temp.: 300K 103
montmorillonite Contact time: 20-30 min
Adsorbent dosage:
0.1-0.5 g/50mL
Activated carbon from 2.5- 20-140 69.52-87.69% Temp.: 25-40oC 105
Ziziphus spina-christi seeds 6.5 172.413 mg/g Contact time: 180 min
(ZSAC) Agitation speed:
200 rpm
Adsorbent dosage:
75 mg/100mL
Tea waste 9.5 1-100 99% Temp.: 25 oC 106
Adsorbent dosage: 0.5g/L
Activated carbon from bean 5-6 NA 23.4 mg/g Contact time: 30 min 107
pods waste Adsorbent dosage:
100 mg/100mL

Black carrot residues 5.25 NA >70%, Temp.: 20-60oC 108


3.871 mg/g Contact time: <1 h
Thermally decomposed leaf 1.8- 100 61-66 mg/g Temp.: 25 oC 109
3.84 Contact time: 3-300 min
Adsorbent dosage: 10g/L

67
Activated carbon from 6.5 10 5.4 mg/g Temp.: 25 oC 110
sugarcane bagasse Contact time: 30 min
Agitation speed: 4rpm
Sugarcane bagasse 2-6 12-24 63% Temp.: 23 ± 2◦C 111
Contact time: 60 min
Agitation speed: 200 rpm
Adsorbent dosage: 5-15g/L
Sugarcane bagasse with HCl 2-6 12-24 99% Temp.: 23 ± 2◦C 111
treatment Contact time: 60 min
Agitation speed: 200rpm
Adsorbent dosage:
5-15g/L
Activated carbon 2-6 12-24 97% Temp.: 23 ± 2◦C 111
Contact time: 60 min
Agitation speed: 200rpm
Adsorbent dosage:
5-15 g/L
T. latifolia roots 5.7 1.8 82-86% Temp.: 25 oC 112
1.680 mg/g Contact time: 60 min
S. americanus roots 5.7 1.68 82-6%, Temp.: 25 oC 112
4.037 mg/g Contact time: 180min
Maize cob NA 1-10 50% Temp.: 23 ± 2◦C 114
2.28 mg/g,
Palm fruit bunch NA 1-10 79%, Temp.: 23 ± 2◦C 114
2.21 mg/g
Rice husk based activated NA NA 100% NA 115
carbon

Sugarcane Bagasse based NA NA 100% NA 115


activated carbon
Formaldehyde modified green 6.0 NA 84.8 % Temp.: 20oC 117
tomato husk 0.028 meq/g Contact time: 30 min
Spirodelapolyrhiza (L.) 2-7 30-90 61.66% Temp: 20-40 oC 118
Schleiden 35.7 mg/g Contact time: 5-120 min
Agitation speed: 180rpm
Adsorbent dosage:
0.05-0.25 g/100mL
Pseudomonas aeruginosa 5.46- 49 22.39 mg/g Agitation speed: 150 rpm 120
AT18 7.72
Crab Shell 2-6 500 90% Temp.: 23 ± 2◦C 122
69.9 mg/g, Contact time: 120 min
Adsorbent dosage:
0.5 g/100 mL

68
Pleurotusostreatus 2-7 200 130.625 mg/g Temp.: 298.15K 123
Contact time: 30 min
Agitation speed: 3500 rpm
Adsorbent dosage: 0.4 g/L
Saccharomyces cerevisiae 7 4.8 22.5 mg/g Temp.: 30 oC 124
Contact time: 30 min,
Agitation: 150rpm,
Yeast conc.: 0.1 g/L

69
Table 12
Manganese removal by membrane filtration
Membrane Initial Mn(II) Pressure Removal pH References
conc. efficiency
mg/L (%)
NF270 1000 4 bar 89 1.5±0.2 129
Polymer 1 0.7 bar 100 4-9 130
enhanced ultra-
filtration(PEUF)
Polyethylene MF 0.5 NA >95 9.3 131
Cellulose acetate 0.5 1 80 132
UF
Copolymer of NA 40 kPa 99.6 6 133
maleic acid and
acrylic acid (PMA-
100) with polyvinyl
butyral (PVB) UF
complexation−ult
rafiltration
Polymeric 2 6-100 95 8.5-8.7 134
polyethersulfone kPa
MF
Hybrid 0.31 NA 100 NA 135
polyethylene MF

70
Table 13
Manganese removal by electrodialysis
Anode Cathode Initial Mn(II) Removal Current pH Reference
conc. efficiency density
(%)

Platinum Platinum 10 mg/L NA 0.75 A/dm2 2.4-9.7 136

Platinum Platinum 0.01 mol/L >98 2.1-8.5 NA 137


mA/cm2

71
Table 14
Manganese recovery from spent batteries
Battery Acid Temperature Reducer Time Recovery( Mn recovery References
o
type leaching ( C) (h) %) form

Zn-C and H2SO4 80 C2H2O4 1 74-77 Mn in 139


Zn-Mn solution
Zn-C and H2SO4 80 H2O2 1 50 Mn in 139
Zn-Mn solution
Zn-C and H2SO4 80 CO(NH2) 1 Inactive - 139
Zn-Mn 2

Alkaline H2SO4 25 NA 3 NA Mn in oxide 141


and Zn-C form

72
Table 15
Comparison of various treatment technologies
No. Type of treatment Advantages Disadvantages References

1 Hydroxide precipitation Widely used method, Sludge generation, extra 24-27


Low capital cost, simple operational cost for
operation sludge disposal, high pH
requirement, not
appropriate for low
concentration of metal
due to co-precipitation of
Magnesium

2 Coagulation/flocculation Improved sludge Sludge production, extra 14


stability, improved operational cost for
sludge settling sludge disposal

3 Flotation Low operating cost, low High maintenance and 14


detention periods, high capital cost
metal selectivity

4 Ion exchange No sludge generation, Not all ion exchange 14, 25, 45
less time consuming, resinsare suitable for
can be used for dilute metal removal, high
systems, economical, capital cost, only used
energy efficient, easy to for trace amount of
control and environment metals in water due to
friendly clogging, regeneration of
resins causes secondary
pollution, cannot be used
for large scale

5 Oxidation/filtration The majority of iron and The oxidant is difficult to 47


manganese treatment store or transport,
system employ this degradation of system
method, Low-cost parts due to corrosion
method, No requirement
of electricity, suitable
for rural areas

6 Electrochemical treatment Rapid, required low High capital cost, high


chemicals, production energy consumption 14
of less sludge
7 Adsorption Effective and High cost of AC limits 14,67
economical method, the method, low cost
offers flexibility in adsorbents should be

73
design and operation, used, Many natural/low-
high-quality of treated cost adsorbents show
effluent, regeneration ofpoor adsorption capacity,
the adsorbents is a low efficiency/cost
possible ratio, and ineffectiveness
at high metal
concentration
8 Membrane filtration High efficiency, no need High operational cost, 14,15,47,67
of chemicals, no prone to membrane
production of solid fouling, low permeate
waste, compact, simple flux
automation

74

Das könnte Ihnen auch gefallen