Sei sulla pagina 1di 20

Environmental Technology

ISSN: 0959-3330 (Print) 1479-487X (Online) Journal homepage: http://www.tandfonline.com/loi/tent20

A review of the use of red mud as adsorbent for


the removal of toxic pollutants from water and
wastewater
Amit Bhatnagar , Vtor J.P. Vilar , Cidlia M.S. Botelho & Rui A.R. Boaventura
To cite this article: Amit Bhatnagar , Vtor J.P. Vilar , Cidlia M.S. Botelho & Rui A.R.
Boaventura (2011) A review of the use of red mud as adsorbent for the removal of toxic
pollutants from water and wastewater, Environmental Technology, 32:3, 231-249, DOI:
10.1080/09593330.2011.560615
To link to this article: http://dx.doi.org/10.1080/09593330.2011.560615

Published online: 26 Mar 2011.

Submit your article to this journal

Article views: 5213

View related articles

Citing articles: 53 View citing articles

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=tent20
Download by: [Louisiana State University]

Date: 11 October 2016, At: 09:30

Environmental Technology
Vol. 32, No. 3, February 2011, 231249

REVIEW ARTICLE
A review of the use of red mud as adsorbent for the removal of toxic pollutants
from water and wastewater
Amit Bhatnagar*, Vtor J.P. Vilar, Cidlia M.S. Botelho and Rui A.R. Boaventura
LSRE Laboratory of Separation and Reaction Engineering, Departamento de Engenharia Qumica, Faculdade de Engenharia
da Universidade do Porto (FEUP), Rua Dr. Roberto Frias, 4200465 Porto, Portugal
(Received 5 August 2010; Accepted 2 February 2011 )

Taylor and Francis Ltd

10.1080/09593330.2011.560615

Red mud (an aluminium industry waste) has received wide attention as an effective adsorbent for water pollution
control, showing significant adsorption potential for the removal of various aquatic pollutants. In this review, an
extensive list of red-mud-based adsorbents has been compiled and their adsorption capacities (maximum uptake
value of the adsorbent for the pollutant or adsorbate being removed) for various aquatic pollutants (metal ions, dyes,
phenolic compounds, inorganic anions) are presented. The review provides a summary of recent information obtained
using batch studies and deals with the adsorption mechanisms involved. It is evident from the literature survey that
red mud has been found to be efficient for the removal of various aquatic pollutants, especially arsenic and phosphate.
However, there is still a need to investigate the practical utility of these adsorbents on a commercial scale.
Keywords: water treatment; adsorption; industrial waste; red mud; adsorption capacity

1. Introduction
Increased industrial and agricultural activities have
resulted in the generation of various types of toxic
pollutants, which are the main cause of water pollution
on a global scale. The type of pollutants present in
wastewater mainly depends on the nature of the industry. However, some of the common pollutants generally
present in effluents are metal ions, dyes, phenols, insecticides, pesticides, detergents and a wide spectrum of
aromatics. Pollutants present in wastewaters can be
toxic to aquatic life and can cause natural waters to be
unfit as potable water sources. A number of processes
have been applied, with varying degree of success, to
the treatment of water and wastewater. Some of these
processes are coagulation [1], foam flotation [2], filtration [3], ion exchange [4], aerobic and anaerobic treatment [5,6], advanced oxidation processes [7], solvent
extraction [8], adsorption [9], electrolysis [10], microbial reduction [11], and activated sludge [12]. However,
these technologies have shown some significant disadvantages, which include insufficient removal of pollutants, high capital costs, high reagents and/or energy
requirements, and generation of toxic sludge or other
waste products that require further safe disposal.
Amongst several water and wastewater treatment
technologies, adsorption is considered as the most
versatile process. Activated carbon has been found to be
a very promising adsorbent and is commonly used for

the removal of diverse types of pollutants from water


and wastewater [13]. However, its widespread use in
water treatment is sometimes restricted owing to its
high cost. As such, for quite some time, efforts have
been made to develop inexpensive adsorbents using
various materials [1419]. However, the search is still
going on to find better alternatives to activated carbon.
It is well known that solid waste materials (byproducts) generated from various industrial activities
pose one of societys most vexing problems. In many
cities of developing countries, the lack of adequate treatment of solid wastes, including industrial wastes,
remains one of the major problems to be solved. An
interesting and beneficial utilization of solid wastes
(wherever possible) is to convert them into low-cost
adsorbents for the treatment of water and wastewater
discharged from various industries. If the solid wastes
could be used as low-cost adsorbents, they will provide
a two-fold advantage for environmental pollution
control. Firstly, the volume of waste materials could be
partly reduced and, secondly, the low-cost adsorbent, if
developed, could reduce the pollution of wastewaters at
a reasonable cost. Various industrial wastes, e.g. fly ash,
slag, red mud and different types of sludge, have been
explored as adsorbents for the removal of diverse types
of pollutants from water and wastewater.
Among various industrial by-products, red mud is a
solid waste residue formed after the caustic digestion of

*Corresponding author. Email: dr.amit10@gmail.com or amit_b10@yahoo.co.in


ISSN 0959-3330 print/ISSN 1479-487X online
2011 Taylor & Francis
DOI: 10.1080/09593330.2011.560615
http://www.informaworld.com

232

A. Bhatnagar et al.

bauxite ores during the production of alumina. Each


year, about 90 million tonnes of red mud are produced
globally [20]. Depending on the origin, quality and
composition of the bauxite, the amount of red mud left
over from alumina refining can vary widely. For every
tonne of alumina produced, the process can leave
behind a third of a tonne to more than two tonnes of red
mud. Red mud is a highly alkaline waste material with
a pH of 1013 because of the sodium hydroxide solution used in the refining process. Red mud is mainly
composed of fine particles containing aluminium, iron,
silicon, titanium oxides and hydroxides. The red colour
is caused by the oxidized iron present, which can make
up to 60% of the mass of the red mud. Because of the
alkaline nature and the chemical and mineralogical
species present in red mud, this solid waste causes a
significant impact on the environment, and proper
disposal of waste red mud presents a huge challenge
where alumina industries are installed. In October 2010,
approximately one million cubic metres of red mud
from an alumina plant near Kolontr in Hungary were
accidentally released into the surrounding countryside
in the Ajka alumina plant accident, killing nine people
and contaminating a large area. Currently, most red
mud produced from alumina plants is disposed in landfills or dumped at sea. The disposal cost is high,
accounting for about 5% of alumina production [20].
Red mud has been used successfully by a number of
workers as a stabilizing agent for in situ fixation of
heavy metals in contaminated soils, mine tailings and
other wastes [2122]. Red mud has found some applications in making different construction materials and
ceramic products [2326]. Furthermore, metal recovery
from red mud is also being developed [27]. Readers
interested in a detailed discussion of applications of red
mud for various purposes should refer to the comprehensive review by Wang et al. [28].
Red mud can be classified as hazardous waste
because of its caustic/saline/sodic nature, and, before its
use as an adsorbent, red mud needs to be neutralized
[29]. Different neutralization methods have been
reported in the literature and some are discussed in this
article. Neutralization of red mud results in a residue
with a pH of 8.08.5. This review attempts to discuss
the adsorbent properties of red mud in water treatment.
Although many review articles are available discussing
the importance of low-cost adsorbents in water remediation [3036], many of them are generally either adsorbate-specific (metals, dyes, phenols, etc.) or adsorbentspecific. One of the aims of the present review is to
compile and present the adsorption potential of red mud
as a low-cost adsorbent for the removal of different
pollutants from water and wastewater. A summary of
relevant published data (in terms of adsorption capacities of red-mud-based adsorbents for the removal of

various pollutants), with some of the latest important


findings, and a source of up-to-date literature are
presented and the results have been discussed. For
information pertaining to detailed experimental methodology and conditions, readers are referred to the full
articles listed in the references.

2. Red mud as an adsorbent for water treatment


2.1. Application of red mud for the removal of metal
ions from water
Metal ions are one of the important classes of aquatic
pollutants. Arsenic is one of the elements found in natural waters, posing serious threat to human health in
many parts of the world. Arsenic occurrence in the environment, its toxicity, health hazards and the techniques
used for speciation analysis are well known and have
been reviewed [3740]. The adsorption process has
been widely applied for the removal of arsenic from
water and wastewater [40].
The use of red mud for arsenic removal from water
has been widely explored by various researchers.
Altundogan et al. [41] used red mud for the removal of
As(III) and As(V) from water. The equilibrium was
achieved within 45 and 90 min for As(III) and As(V),
respectively, at 25 C, 133.5 mol/L (10 mg/L) concentration and 20 g/L red mud dosage. For As(III) and
As(V), favourable adsorptions took place at pH 9.5 and
3.2, respectively. The As(III) and As(V) adsorption
capacities of red mud at 25 C, estimated from the
Langmuir isotherm, were 8.86 and 6.86 mol/g, respectively. Based on thermodynamic studies, the authors
stated that As(III) adsorption is exothermic whereas the
adsorption of As(V) is endothermic, and concluded that
the nature of As(III) adsorption is physical and that of
As(V) is chemical.
In another study by the same workers, heat treatment (200, 400, 600 and 800 C for 4 h) and acid treatment (1 L of 0.252.00 M HCl solutions mixed with
red mud stirred for 2 h) methods were applied to red
mud to increase its arsenic adsorption capacity [42].
The results indicated that the adsorptive capacity of red
mud could be increased by acid treatment. It was
reported that an increase in the concentration of acid
used in acid treatment up to 1.0 M caused an increase
in adsorption efficiency of red mud; thereafter a
decrease was observed. An increase in adsorption efficiency after acid treatment was explained by the leaching out of sodalite compounds, which possibly blocked
the active sites of the raw adsorbent. The removal of
sodalites was confirmed by the X-ray diffraction analyses of acid-treated red mud. A decrease in adsorption
efficiency of red mud treated by acid solutions with
concentrations of more than 1.0 M was observed,
gevb[r]

Environmental Technology

which might be attributed to the dissolution of some


small particles that caused a decrease in surface area.
On the other hand, it was observed that mixtures of red
mud and acid solution of 0.25 and 0.50 M exhibited
colloidal properties during treatment. Red mud treated
by these acid solutions exhibited low adsorptivity due
to covering the silicic acid of the active oxidic sites.
The As(III) and As(V) adsorption characteristics of
activated red mud (ARM) are similar to those of raw
red mud. Batch adsorption studies have shown that
ARM in dosages ranging from 20 to 100 g/L could be
used effectively to remove arsenic from aqueous solutions [42]. The process was found to be pH dependent,
the optimum range being 5.87.5 for As(III) and 1.8
3.5 for As(V). The maximum removals were 96.52%
for As(V) and 87.54% for As(III) for solutions with a
final pH of 7.25 and 3.50, respectively, for the initial
arsenic concentration of 133.5 mol/L (10 mg/L),
ARM dosage of 20 g/L, contact time of 60 min and
temperature of 25 C. The adsorption data followed the
first-order rate expression and fitted well with the
Langmuir isotherm. The Langmuir monolayer capacities were 11.80 mol/g for As(III) and 12.57 mol/g
for As(V) at 25 C. Adsorption of As(III) was exothermic, whereas As(V) adsorption was endothermic.
In the previous studies by Altundogan et al. [41,42]
arsenic removal by the solid phase of red mud and acidactivated red mud was investigated. To further continue
their studies, the liquid phase of red mud (LPRM), by
coprecipitation with aluminium hydroxide, was
explored for arsenate removal [43]. First, neutralization
of LPRMarsenical solution mixtures with acid solution
accompanied by air-agitation was done and then
neutralization of those mixtures with CO2 gas was
performed. The effect of the volumetric ratio of
LPRM:As(V)-solution on the removal of As(V) by
coprecipitation of As together with Al present as aluminate in the LPRM was studied. The results showed that
As(V) was removed effectively by LPRM with a volumetric LPRM:As(V)-solution ratio of 0.1 from an
arsenical solution with As(V) concentration of 20 mg/L.
The use of red mud was investigated for its arsenic
removal effectiveness in aqueous solutions under various conditions [44]. Red mud efficiently removed
As(III) in the pH range between 7.6 and 9.0 and As(V)
in the pH range between 5.5 and 6.0. Pre-washing of the
red mud with salt water considerably improved its
arsenic adsorption capabilities. With appropriate red
mud dosage, the residual arsenic in solution was
decreased below the regulated acceptable arsenic limit
(0.1 mg/L) from aqueous industrial wastes. The arsenic
adsorption followed first-order rate kinetics and fitted
the Langmuir isotherm model.
As a possible cost-effective process, red mud was
treated or neutralized with seawater until a pH within
gevb[r]

233

the range 8.48.8 was reached, and the neutralized red


mud was named Bauxsol [45]. In the same study, the
Bauxsol was further used as an adsorbent for removing
arsenate from water. The authors noted that the arsenate
adsorption on Bauxsol showed no dependence on ionic
strength, and they suggested that arsenate might be
adsorbed on Bauxsol by strongly binding chemical
bonds, i.e. a largely covalent bond forming inner-sphere
complexes with little competitive adsorption of counteranions. Inner-sphere complex formation was suggested
for arsenate adsorption on Bauxsol. It was found that
the adsorption of arsenate decreased in the presence of
HCO3, whereas Cl had little effect, and Ca2+ increased
the adsorption. A toxicity characteristic leaching procedure (TCLP) test revealed that the used adsorbent was
not toxic. The adsorption capacity of Bauxsol for arsenate was 6.0814.43 mol/g at pH 6.310.0.
Activated seawater-neutralized red mud (activated
Bauxsol, AB), was used as a novel adsorbent for removing inorganic arsenic from water [46]. The AB was
prepared using the combined acid and heat treatment
methods, which involved refluxing the Bauxsol in HCl,
adding ammonia for complete precipitation, filtering,
washing with distilled water, and calcining at 500 C for
2 h. Kinetic data indicated that the process pseudo
equilibrated in 3 and 6 h for As(V) and As(III), respectively, and followed pseudo-first-order kinetics. Within
the range tested, the optimal pH for As(V) adsorption
was 4.5, and close to 100% removal could be achieved
irrespective of the initial As(V) concentration. Desorption of As(V) was greatest at pH 11.6 where a maximum of 40% could be achieved. In contrast, the
optimum pH for As(III) removal was 8.5, and the
removal efficiency changed with the initial As(III)
concentration. The adsorption data fitted the Langmuir
isotherm well, with thermodynamic data indicating the
spontaneous and endothermic nature of the process. The
FITEQL (v.4) and PHREEQC (v.2) computer programs
were used to predict As(V) adsorption at various pH
values (based on diffuse double layer models). The
models fitted the experimental results well and
indicated that surface complexation modelling is useful
in describing the complex AB surface during the
adsorption process.
The same workers also tested the possibility of
increasing arsenate adsorption by seawater-neutralized
red mud (Bauxsol) through acid treatment, combined
acid and heat treatment, and the addition of ferric
sulphate (Fe2(SO4)37H2O) or aluminium sulphate
(Al2(SO4)318H2O) was also investigated [47]. It was
found that the arsenate adsorption from water on to
Bauxsol (89%) was significantly increased by acid
treatment (95%) and combined acid and heat treatment
(roughly 100%), at the conditions considered. It was
also observed that the addition of aluminium sulphate

234

A. Bhatnagar et al.

considerably reduced the efficiency, whereas the


suppression was relatively small when ferric sulphate
was added. This observation was explained by the
formation of watery gels, which were covering the
available sorption sites for arsenate, and to the fact that
sulphate competes with arsenate for the available
adsorption sites. Furthermore, it was found that the
adsorption process using AB was not accompanied by
the release of unwanted contaminants, and TCLP results
confirmed that the spent AB was not hazardous.
Based on the promising results obtained in batch
tests, the authors further extended their study to investigate the possibility of using Bauxsol and AB under
continuous flow conditions [48]. An extensive
laboratory investigation was carried out to evaluate the
arsenate adsorption capacity of Bauxsol-coated sand
(BCS) and activated Bauxsol-coated sand (ABCS)
grains prepared from Bauxsol and AB, respectively.
Both batch and column tests were carried out using
BCS and ABCS for arsenate removal from water. The
batch experiments indicated that fast adsorption of
arsenate on to BCS took place over about 4 h, but slow
adsorption continued for at least 21 days. The observed
adsorption data fitted well with the Langmuir model,
and the adsorption capacities calculated for ABCS at
pH = 7.1 were 1.642.14 mg/g, and were 3.32 mg/g for
BCS at pH = 4.5. Higher sorption capacities were
obtained from the column experiments than from the
batch tests. The higher arsenate sorption capacity in
column experiments was explained by many reasons
such as underestimation of pseudo-equilibration times
in the batch trials, coagulation and crystal growth, as
well as due to granulation that commonly occurs
during aging. Breakthrough curves were affected by
the flow rate, inflow arsenate concentration and the
presence of the competing anions. Increasing flow rate
or initial arsenate concentration, or adding competing
anions were found to suppress the arsenate removal,
with inflow arsenate concentration having the greatest
effect.
Red mud (RM) was modified with FeCl3 for the
removal of arsenate from water [49]. After modification, the content of iron in RM increased and calcium
decreased. Equilibrium time for arsenate removal was
24 h. Solution pH significantly affected the adsorption,
and the adsorption capacity increased with the decrease
in pH. Langmuir and Freundlich isotherms were used to
fit the adsorption isotherms. The Langmuir isotherm
was the best-fit adsorption isotherm model for the
experimental data. Adsorption capacity of modified red
mud was found to be 68.5 mg/g, 50.6 mg/g and
23.2 mg/g at pH 6, 7 and 9, respectively. Nitrate ion had
little effect on the adsorption, Ca2+ enhanced the
adsorption, whereas HCO3 decreased the adsorption.
The modified red mud could be regenerated with

NaOH, and the regeneration efficiency reached 92.1%


when the concentration of NaOH was 0.2 mol/L.
The adsorption capacity of red mud for arsenate
removal was evaluated at different pH values (4, 7 and
10) [50]. Red mud samples were artificially enriched in
batch tests with solutions containing increasing concentrations of As(V). The arsenate sorption in RM samples
increased as the pH decreased from 10 to 4. This could
be due to the formation of chemical bonds and
unspecific electrostatic bonds between RM at pH 4 and
arsenate, since at this pH value the red mud surfaces
were positively charged. The interaction mechanisms
between As(V) and RM were explained on the basis of
several types of mechanisms, e.g. electrostatic attraction/repulsion, chemical interaction and ion exchange.
The results of sequential extraction showed that a
higher percentage of As(V) in RM at pH 4 was
exchanged with H2O and (NH4)2SO4 than was extracted
in RM at pH 7 and 10. Therefore it was possible to
hypothesize that the As(V) in RM at pH 7 and 10 was
mainly sorbed through a ligand exchange mechanism
that involved the formation of inner-sphere complexes.
It was further suggested that the absolute total As
content in RM enriched at pH 4 was much higher than
at pH 7 and 10, and therefore the absolute value which
remained adsorbed after the extraction with water and
(NH4)2SO4 was also higher than that obtained in the
case of RMAs(V) at pH 7 and 10. The FTIR spectra of
RMAs(V) showed that the oxides and oxyhydroxides
of Fe and Al (haematite, boehmite and gibbsite) of RM
were probably the mineralogical phases involved in the
sorption processes. However, this spectroscopic technique did not differentiate the different role of FeAl
oxides and oxyhydroxides in the arsenate polyhedra
sorption. At the same time it was difficult to determine
the type of coordination of the adsorbed As(V) because
both protonated and unprotonated arsenate ions were
present. The infrared band due to adsorbed arsenate
species was observed at 861865 cm1, and its intensity
and broadness increased as the pH decreased because of
the higher As(V) sorption at pH 4. After the sequential
extraction steps, the FTIR spectra of RMAs(V) at pH
4 showed that the band at 865 cm1 had completely
disappeared after the extraction with NH4+-oxalate.
Consequently, in agreement with the results of the
sequential extraction, it was concluded that most of the
As(V) sorbed in RM (about 80%) was strongly and
specifically associated with the FeAl oxides and
oxyhydroxide phases.
The ability of activated CO2-neutralized red mud
(ANRM) for the removal of arsenate from aqueous solutions was tested [51]. The percentage removal was
found to increase gradually with a decrease in pH, and
maximum removal was achieved at approximately pH 4.
Adsorption kinetic studies revealed that the adsorption

Environmental Technology

process followed pseudo-second-order kinetics and


equilibration was achieved within 24 h. The FTIR spectra of ANRM before and after adsorption revealed the
binding of arsenate to the adsorbent. It was suggested
that the hydroxyl surfaces of a mixture of Fe, Al and Ti
oxides of ANRM provide strong adsorption affinity for
arsenate by forming inner-sphere complexes. Arsenateadsorbed ANRM could be regenerated using NaOH
solution at pH 12.0.
The possibility of utilizing mixed adsorbent materials of red mud with haematite and china clay, china clay
with fly ash, and red mud alone for the removal of
As(III) from solution was investigated in a batch-type
configuration [52]. The effect of solution pH on the
adsorption process using these low-cost mixed adsorbents was studied and it was observed that favourable
conditions were attained at an equilibrium time of 140
min and a pH of 8, and the maximum removal of As(III)
was in the range of 7986% [52]. Adsorption kinetics
for removal of arsenic was determined for red mud and
its mixtures with haematite, china clay and fly ash
besides china clay and fly ash mixed adsorbents at:
adsorbate concentration, 5.0 mg/L; particle size of
adsorbent, <53 m; agitation rate, 220 rpm; pH, 8.0;
and temperature, 30, 40 and 50 C [53]. The Lagergren
model fitted well with the experimental data. The
As(III) removal by adsorbents was found to be diffusion
controlled.
Besides arsenic, chromium is also one of the
elements that need immediate attention, as Cr(VI)
compounds are toxic owing to their high water solubility
and mobility. Some researchers explored the possibility
of using red mud as adsorbent for the removal of chromium from aqueous solutions. Adsorption by ARM was
investigated for Cr(VI) removal from aqueous synthetic
solutions and industrial effluents [54]. Activated red
mud was prepared by simple acid dissolution followed
by ammonia precipitation and then drying at 110 C.
The best conditions for adsorption were found to be at
pH 5.2 and at temperature 303 K in the concentration
range of 230 mg/L with a solid:liquid ratio of 1:500.
The Langmuir monolayer capacity of ARM for Cr(VI)
was found to be 30.74 mmol/g. The affinity sequence for
the adsorption of competing anions on ARM was in the
order, PO43 > SO42 > NO3. Red mud activated with
HCl was also used for chromate removal by Dursun
et al. [55]. About 70% chromate removal efficiency was
obtained by the optimum red mud dose and pH value.
The removal of trivalent chromium from synthetic
wastewater using red mud was studied by Rajkumar
et al. [56]. The maximum adsorption efficiency of
99.9% was achieved when 1.5 g of red mud was used to
remove chromium from a test solution containing
150 mg Cr(III)/100 mL. The idle pH and agitation time
recorded during the study were 6 and 10 h, respectively.

235

Zouboulis and Kydros [57] examined the potential


of red mud for nickel (Ni) removal from aqueous solution. Red mud acted simultaneously as an alkalinity
regulator, causing precipitation of Ni as the insoluble
hydroxide, as an adsorbent of the formed nickel
hydroxide and as a flocculant of the resultant fine
particulate matter. Furthermore, sedimentation was
subsequently considered as a possible solidliquid
separation technique. The authors obtained promising
results under various parameters, such as dispersion
pH, red mud and nickel concentrations and zeta potential. The removal of Ni(II) from aqueous solution by red
mud with some other adsorbents was also investigated
by Hannachi et al. [58]. A contact time of 4 h was
needed to establish equilibrium. Maximum sorption of
Ni(II) from aqueous solution was found at approximately pH 5. The adsorption capacity of red mud for
Ni(II) was found to be 13.69 mg/g. The removal of
nickel ions was attributable to a chemisorption reaction
at the surface of the oxide components that principally
constitute red mud. The adsorption process followed a
first-order rate mechanism. Langmuir and Freundlich
adsorption isotherms fitted well with the experimental
data. Recently, Bosnian red mud was evaluated as an
economical, composite sorbent for the removal of aqueous Ni2+ ions [59]. The investigated mineral mixture
exhibited a high acid-neutralizing capacity, and its most
important role in cation immobilization was observed
in the initial pH range 28. The initial metal ion
concentration strongly influenced the sorption kinetics
and equilibrium times. The addition of 5 g/L of the red
mud caused 100% removal from the solutions of 1
104 to 5 104 mol/L, whereas with a further increase
in Ni2+ concentration to 8 103 mol/L the removal
efficiency decreased to 26%. The maximum sorption
capacity of 0.372 mmol/g, at initial pH 5, was found
using the Langmuir model. The possibility of improving sorption efficiency by annealing the Bosnian red
mud powder was investigated in the temperature range
200900 C, and the relationships between temperature, red mud physicochemical and sorption properties
were established. The optimum heating temperature
was found to be 600 C, as a result of water exclusion
from gibbsite and bayerite phases, leading to improved
porosity and surface area, as well as increased pH and
sorption efficiency. The stability of the sorbed cation
was assessed by leaching experiments in distilled water
and acidic TCLP2 solution.
The adsorption behaviour of copper (Cu) on RM
with two other solid waste materials sea nodule
residue and fly ash was investigated [60]. Maximum
adsorption of copper was found at pH 5.5 in the case of
RM. The Langmuir and Freundlich models did not fit
well with the equilibrium data because there was no
appreciable effect of temperature on the metal removal

236

A. Bhatnagar et al.

by RM. Under the optimized conditions, the adsorption


capacity of RM for Cu was found to be 2.28 mg/g.
Copper desorption of 73% could be achieved with 0.05
M HCl. Recently, Nadaroglu et al. [61] also examined
the potential of red mud for copper removal from water.
It was found that the adsorption of copper increased
with increasing pH and maximum adsorption of copper
was at initial solution pH 5.5. This was explained by the
surface complexation reactions, which are mostly influenced by the electrostatic force of attraction between
copper and the surface of the adsorbent.
Granular red mud (GRM) has also been evaluated
for its potential to remove cadmium from aqueous solutions [62]. The pseudo-second-order model was fitted to
kinetic data at an initial pH of 6.0 and 3.0. Particularly
for the mass transfer process in the GRMcadmium
system at initial pH 6.0, the processes were found to be
mainly controlled by intraparticle-diffusion, through the
analysis of external mass transfer coefficient and
effective particle diffusion coefficients with regards to
different initial cadmium concentration of 50 mg/L,
100 mg/L and 200 mg/L. The maximum adsorption
capacities for GRM were determined to be 38.2 mg/g at
20 C, 43.4 mg/g at 30 C and 52.1 mg/g at 40 C. The
cadmium-loaded GRM adsorbents were regenerated
after each cycle of adsorption, using 0.1 M HCl, up to
four times. The feasibility of RM for wastewater treatment was assessed by batch method by Lpez et al.
[63]. The aggregates were prepared using red mud and
8% (w/w) CaSO4 and their adsorption potential was
examined by batch and column experiments. The RM
aggregates showed maximum adsorption capacities for
Cu2+, Zn2+, Ni2+ and Cd2+ of 19.72, 12.59, 10.95 and
10.57 mg/g, respectively, with a contact time of 48 h. In
continuous adsorption experiments in which secondary
effluent from an urban sewage treatment plant was
percolated through RM aggregates packed into
columns, purification efficiencies for P, Ni2+, Cu2+ and
Zn2+ were 100%, 100%, 68% and 56%, respectively. It
was suggested that aggregated RM is suitable for treatment of wastewaters, in particular those whose principal
contaminants are P or heavy metals.
Metallurgical solid wastes (red muds and fly ashes)
were converted into unconventional sorbents for heavy
metal removal from contaminated water [64]. The
heavy metal (Pb, Cd and Cu) removal capacity as well
as sorption modelling of red muds and fly ashes was
examined. The batch adsorption capacities for red muds
after different types of treatments were in the range of
46.966.8 mg/g for Cd(II), 35.275.2 mg/g for Cu(II)
and 117.3165.8 mg/g for Pb(II). The adsorption capacities of red muds were found to be lower than those of
fly ashes for the removal of the same metal ions.
Red mud has also been used for the removal of lead
and chromium from aqueous solutions [65]. The prod-

uct, activated in air, showed promising adsorption characteristics. The effects of various factors (e.g. pH,
adsorbent dose, adsorbate concentration, temperature,
particle size) on the removal of these metal ions from
water were studied. The material exhibited a good
adsorption capacity and the Freundlich and Langmuir
models were able to fit the data. The adsorption
capacities of red mud were 64.79 mg/g for lead and
35.66 mg/g for chromium by batch method. The column
studies showed that the product could also be used on an
industrial scale. Metal ions adsorbed on the column of
this material could be quantitatively eluted with 1%
HNO3. The exhausted column could be chemically
regenerated by treating it with 1 M HNO3, and no
dismantling was required. Other salts present in the
effluents did not cause any disturbing effects.
Red mud was also examined for the removal of
cadmium and zinc from aqueous solutions [66]. The
crude form of red mud waste showed poor adsorption
properties; therefore, this material was first treated with
hydrogen peroxide at room temperature for 24 h to
oxidize adhering organic matter and was washed
several times with double-distilled water, followed by
drying at 100 C and cooling. The resulting material
was then activated in air in a muffle furnace at 500 C
for 3 h. The final product exhibited the best adsorption
capacity and optimum surface area. The product
obtained at temperatures higher than 500 C exhibited
poor adsorption capacity, which was due to the collapse
of surface functional groups on the adsorbent. The
removal of Cd2+ and Zn2+ was almost complete at low
concentrations, whereas it was 6065% at higher
concentrations at optimum pHs of 4.0 and 5.0, respectively, with 10 g/L of adsorbent in an 810 h equilibration time. The adsorption decreased with increase in
temperature. Chemical regeneration of the columns was
achieved with 1% HNO3. The Langmuir monolayer
capacity was found to be 1.16 104 mol/g for Cd(II)
and 2.22 104 mol/g for Zn(II) at 30 C.
Pellet-type adsorbents were made from red mud [67].
Adsorption of heavy metal ions (Pb2+, Cu2+, Cd2+) in
aqueous solutions by the pellet-type red mud adsorbents
were studied under various experimental conditions. It
was found that pellet-type red mud adsorbents made
from a mixture of 58.7 wt% red mud, 25.2 wt% kaolin,
11.7 wt% sodium silicate solution, 2.9 wt% fly ash and
1.5 wt% magnesium chloride at 600 C exhibited the
highest removal efficiency for the heavy metal ions. The
removal efficiency for Pb2+, Cu2+ and Cd2+ after 24 h of
operation was more than 95%. The Langmuir model was
able to fit with the experimental data. A continuous
adsorption experiment showed that the pellet-type red
mud adsorbents were effective in removing Pb2+.
Red mud has also been explored as a potential
sorbent for the removal of Cd, Zn, Cu and Pb from

Environmental Technology

aqueous solutions in the presence of 0.01 M NaNO3


[68]. The red mud showed a relatively high uptake of
cadmium and zinc from near-neutral aqueous solutions
(maximum uptake capacity for cadmium: 68 mg/g at pH
6 and ca.133 mg/g for zinc at pH 7). A significant
uptake was also observed for copper and lead at pH 6
and 7, respectively, which was attributed to precipitation of the respective insoluble hydroxides.
Heavy metal adsorption of non-treated (RMnt) and
acid-treated red muds (RMa) was investigated in order
to evaluate their efficiency in reducing metal solubility
and bioavailability in polluted soils [69]. Red mud
samples were artificially polluted with solutions
containing increasing concentrations of Pb, Cd and Zn.
Cancrinite and haematite were the main phases of the
red muds and were also the components that adsorbed
the most heavy metals. The results showed that the
RMnt adsorption capacity for the three heavy metals
was Zn Pb > Cd. Acid treatment with HCl decreased
the red muds capacity to adsorb the heavy metals by
30%. In order to study the different interaction mechanisms of heavy metal with RM, all samples after
artificial contamination were treated with solutions with
gradually increasing extraction capacity. Treatments
with H2O and Ca(NO3)2 only extracted very low
concentrations of Pb, Cd and Zn, whereas EDTA
treatment extracted the highest amount of adsorbed
heavy metals from the sorbent particles. In particular,
the water-soluble and exchangeable metal fractions
were higher in the RMa than they were in the RMnt,
whereas the concentrations of Pb, Cd and Zn extracted
with EDTA were lower. The results showed that red
mud can be used successfully to reduce the solubility
and bioavailability of heavy metals in polluted soils.
Red mud was calcined at different temperatures, and
the calcined RM was used as adsorbent for removal of
phosphate and heavy metals from the effluent of swine
wastewater treated with a sequencing batch reactor
(SBR) [70]. The Langmuir isotherm fitted well with the
data. Calcination enhanced the adsorption capacity
greatly. The adsorption capacity of RM calcined at
900 C for phosphate, copper, zinc and arsenic
increased from 46.26, 18.18, 15.45 and 18.83 mg/g to
149.00, 65.17, 99.20 and 27.51 mg/g, respectively. The
pH showed an obvious effect on the removal, and high
pH favoured the removal of phosphate, copper, zinc and
arsenic. The adsorption mechanism was suggested to
include surface complexation reactions, and the
mechanism of phosphate and arsenic removal might
include coprecipitation.
The effectiveness of using thermally activated
hydrotalcite materials has been investigated for the
removal of arsenate, vanadate and molybdate in individual and mixed solutions [71]. The results showed that
increasing the Mg:Al ratio to 4:1 caused an increase in

237

the percentage of anions removed from solution. The


order of increasing affinity of the three anions analyzed
in this investigation was arsenate, vanadate and molybdate. By comparison with several synthetic hydrotalcite
materials, the hydrotalcite structure in seawater-neutralized red mud (SWN-RM) was determined to consist of
magnesium and aluminium with a ratio between 3.5:1
and 4:1. Thermally activated SWN-RM removed at
least twice the concentration of anionic species than
thermally activated red mud alone, as a result of the
formation of 4060% Bayer hydrotalcite during the
neutralization process.
The adsorptive removal of boron from aqueous
solution using neutralized red mud was studied in a
batch equilibration technique [72]. The experiments
demonstrated that boron removal fluctuated little in the
pH range 27 and it took 20 min to attain equilibrium.
The Langmuir monolayer sorption capacity was found
to be 30.12 mg/g. Besides these studies, several other
researchers also explored the potential of red mud as
adsorbent for the removal of toxic metals from aqueous
solutions [7374].
Zhou and Haynes [29] discussed the different mechanisms of sorption on red mud via specific adsorption of
heavy metal cations. They also explained that the sorption of anions occurred on the variable charge surfaces
of oxides, as red mud is mainly composed of iron oxides
and some portion of aluminium oxides. Higher concentrations of metals in solution will lead to the surface
precipitation/coprecipitation on the surface. However,
the relative importance of adsorption versus surface
precipitation may differ for different metals. Additionally, neutralized red mud which often has a pH of 7.0
8.5, favours the adsorption/precipitation of heavy metal
cations but disfavours that of anions such as chromate,
selenite, selenate and arsenate [29]. It is evident from
the vast literature survey that red mud has been proven
to be a very promising adsorbent for the removal of
metal ions (especially arsenic) from water and wastewater. A summary of the adsorption capacity of red mud
for different metal ions is presented in Table 1.
2.2. Application of red mud for the removal
of dyes from water
Coloured dye effluents are generally considered to be
highly toxic to aquatic biota [75]. Many health-related
problems such as allergy, dermatitis, skin irritation,
cancer and mutations in humans are associated with dye
pollution in water [7677]. Red mud was also used for
the removal of different dyes from water and wastewater. Gupta et al. [78] utilized red mud for the removal of
Rhodamine B, Fast Green and Methylene Blue dyes
from wastewater. The adsorbent was developed as
described in an earlier paper [66]. The percentage

238
Table 1.

A. Bhatnagar et al.
Adsorption capacity of red mud for the removal of different metal ions from water.

Adsorbent
Red mud
Red mud
Activated red mud
Activated red mud
Seawater-neutralized red mud (Bauxsol)
Bauxsol-coated sand
Activated-Bauxsol-coated sand
FeCl3-coated sand
CO2-neutralized red mud (ANRM)
Activated red mud
Red mud
Red mud
Granular red mud
Red mud
Red mud
Red mud
Red mud
Treated red mud
Treated red mud
Treated red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Calcined red mud
Calcined red mud
Calcined red mud
Neutralized red mud

Adsorbate

Amount adsorbed

Reference

As(III)
As(V)
As(III)
As(V)
As(V)
As(V)
As(V)
As(V)
As(V)
Cr(VI)
Ni(II)
Cu(II)
Cd(II)
Cu(II)
Zn(II)
Ni(II)
Cd(II)
Cd(II)
Cu(II)
Pb(II)
Pb(II)
Cr(VI)
Cd(II)
Zn(II)
Cd(II)
Zn(II)
Copper
Zinc
Arsenic
Boron

8.86 mol/g
6.86 mol/g
11.80 mol/g
12.57 mol/g
6.0814.43 mol/g
1.643.32 mg/g
2.14 mg/g
23.268.5 mg/g
55.55 mg/g
30.74 mmol/g
13.69 mg/g
2.28 mg/g
38.252.1 mg/g
19.72 mg/g
12.59 mg/g
10.95 mg/g
10.57 mg/g
46.966.8 mg/g
35.275.2 mg/g
117.3165.8 mg/g
64.79 mg/g
35.66 mg/g
1.16 104 mol/g
2.22 104 mol/g
68 mg/g
133 mg/g
18.1865.17 mg/g
15.4599.20 mg/g
18.8327.51 mg/g
30.12 mg/g

[41]
[41]
[42]
[42]
[45]
[48]
[48]
[49]
[51]
[54]
[58]
[60]
[62]
[63]
[63]
[63]
[63]
[64]
[64]
[64]
[65]
[65]
[66]
[66]
[68]
[68]
[70]
[70]
[70]
[72]

removals of Rhodamine B, Fast Green and Methylene


Blue on this adsorbent were 92.5, 94.0 and 75.0, respectively. Optimum pHs for the removal were 1.0, 7.0, and
8.0 for rhodamine B, fast green and methylene blue,
respectively. A decrease in the adsorption of the dyes in
the presence of different concentrations of cationic and
anionic surfactants, cetyltrimethylammonium bromide
(CTAB) and manoxol 1B, was observed. During
column operations, removals of about 9597% for
Rhodamine B, Fast Green, and Methylene Blue were
achieved at a flow rate of 0.5 mL/min, but the removal
decreased with increasing flow rate. The column capacity was found to be higher than the adsorption capacity
of batch experiments, and this was explained by the fact
that a large concentration gradient was continuously
present at the interface zones as the dye-containing
sample passed through the column, whereas the concentration gradient decreased with time in batch experiments. Desorption of Rhodamine B, Fast Green and

Methylene Blue was tried with a number of eluents


(methanol, ethanol, acetone, sodium hydroxide, sulphuric acid, hydrochloric acid, nitric acid, etc.) and it was
found that the desorption of dyes occurred easily with
acetone.
Waste red mud was recycled for the adsorption of
Congo Red from aqueous solution [79]. The first rate
expression fitted with the adsorption kinetics. Langmuir
and Freundlich isotherms were able to fit with the equilibrium adsorption data. The adsorption capacity of the
red mud for the dye was 4.05 mg/g. Adsorption was
found to be nearly quantitative at pH 2.0. Effect of pH
and desorption studies suggested that the mechanism of
adsorption was mainly ion exchange.
The ability of waste red mud to remove procion
orange dye was investigated at different initial dye
concentrations, agitation time, adsorbent dosage and
pH [80]. It was found that when the initial pH of the
solution was increased from 2.0 to 11.0, the per cent

Environmental Technology

removal decreased from 82 to 0. The decrease in


adsorption with an increase in pH was explained on
the basis of aqua complex formation and subsequent
acidbase dissociation at the solidsolution interface.
The Freundlich isotherm model fitted well with the
experimental data. A maximum removal of 82% of the
dye was observed at pH 2.0. Desorption studies
showed that maximum desorption occurred at a pH of
11. Red mud was also used for the removal of acid
violet dye from wastewater [81]. Freundlich and Langmuir models were able to fit well with the experimental data. The Langmuir adsorption capacity was found
to be 1.37 mg/g. Quantitative dye removal occurred at
pH 4.1. Desorption studies showed that ion exchange
was the main mechanism involved in the sorption
process.
Red mud activated by HCl was used for the removal
of Congo Red dye from water in batch adsorption
experiments [82]. The pH of the dye solution strongly
affected the chemistry of both the dye molecules and
the ARM in an aqueous solution. The effective pH was
7.0 for adsorption on ARM. It was found that the
equilibrium was achieved in 90 min. The maximum
monolayer sorption capacity for the adsorption of
Congo Red by ARM was found to be 7.08 mg/g.
Fly ash and red mud have been employed as adsorbents for the removal of Methylene Blue dye from
aqueous solution. Heat treatment (800 C for overnight) and chemical treatment (1 N HNO3 solution for
24 h) have also been applied to the as-received fly ash
and red mud samples [83]. Heat treatment reduced the
adsorption capacity for both fly ash and red mud, but
acid treatment by HNO3 induced a different effect on
fly ash and red mud. Nitric acid treatment resulted in
an increase in adsorption capacity of fly ash, whereas
it decreased the adsorption capacity of red mud. It was
explained that the higher temperature treatment of red
mud resulted in the decomposition of some organics
and hydroxyl groups, which were mainly effective
sites for adsorption, leading to the lower adsorption
capacity. On the other hand, acid treatment dissolved
the minerals in carbon and thus increased the pore
volume and surface area of the fly ash sample, leading
to an increase in adsorption. But, for red mud, acid
Table 2.

239

treatment neutralized the hydroxide ions on the basic


surface, which favours the adsorption of basic dye.
The RedlichPeterson model provided the best correlation of the experimental data. A summary of adsorption capacity of red mud for different dyes is presented
in Table 2.
It is evident from the literature survey that red mud
has shown little effectiveness for dye removal. It is
therefore necessary to continue more research on the
removal of different classes of dyes using red mud as
adsorbent. Further research is also needed to gain a
better understanding of the mechanism of dye adsorption on red mud, as little information/discussion is
available on the possible mechanism of dye sorption by
red mud.
2.3. Application of red mud for the removal of
phenolic pollutants from water
Phenol and substituted phenols are considered as priority pollutants [84]. The discharge of effluents containing phenolic pollutants from various industries into
natural water bodies is an ongoing and serious threat
to human health and natural water quality. The ubiquitous nature of phenols, their toxicity even in trace
amounts and the stricter environmental regulations
make it necessary to develop processes for the removal
of phenols from wastewaters. Red mud has also
been explored as potential adsorbent for the removal
of phenols from water and wastewater. The removal of
1-butanethiol from diesel oil at different concentrations and temperatures was investigated using red
mud [85]. Low concentration and high temperature
favoured the percentage removal of 1-butanethiol from
diesel oil. The first-order adsorption rate expression
fitted well with kinetic data and was controlled by
intraparticle transport. The Langmuir model fitted well
with the equilibrium data at different temperatures.
Thermodynamic parameters indicated the spontaneous
and endothermic nature of the adsorption process with
positive entropy change.
The ability of waste red mud to adsorb 2chlorophenol was investigated [86]. The first-order rate
expression fitted with the kinetic data and the Freundlich

Adsorption capacity of red mud for the removal of different dyes from water.

Adsorbent
Red mud
Red mud
Red mud
Red mud
Red mud
Acid-activated red mud

Adsorbate
Rhodamine B
Fast Green
Methylene Blue
Congo Red
Acid Violet
Congo Red

Amount adsorbed
5

(1.011.16) 10 mol/g
(7.259.35) 106 mol/g
(4.355.23) 105 mol/g
4.05 mg/g
1.37 mg/g
7.08 mg/g

Reference
[78]
[78]
[78]
[79]
[81]
[82]

240

A. Bhatnagar et al.

model fitted with the equilibrium sorption data. A maximum removal of 83% was observed at pH 8.0. Desorption of phenol from the spent adsorbent was only 41%
with water at pH 11.0, indicating that both physisorption
and chemisorption occurred in the adsorption process.
Red mud has also been used for the removal of
phenol, 2-chlorophenol, 4-chlorophenol, and 2,4dichlorophenol from wastewater [87]. The maximum
adsorption of phenol and 2-chlorophenol occurred at
pH 6.0, whereas the maximum adsorption of 4-chlorophenol and 2,4-dichlorophenol was achieved at pH
5.0 and 4.0, respectively. 2,4-Dichlorophenol and 4chlorophenol were sorbed by the developed adsorbent
by 9497%, while the removal of 2-chlorophenol and
phenol was 5081%. Adsorption of mixtures of phenols
was also tried by the researchers, and it was observed
that 2,4-dichlorophenol had the maximum adsorption in
comparison with the other phenols. The order of
adsorption was 2,4-dichlorophenol > 4-chlorophenol >
2-chlorophenol > phenol, and the adsorption achieved
was 97%, 93%, 80%, and 51%, respectively, with a
concentration of 2 105 M of each phenol. The authors
suggested that this sort of behaviour of red mud shows
non-specificity with respect to phenols, and thus it
could be used for the uptake of phenols individually or
in their mixtures. The removal of phenol and its derivatives was up to 98% by column experiments at a flow
rate of 0.5 mL/min. The order of removal was 2,4dichlorophenol > 4-chlorophenol > 2-chlorophenol >
phenol, and the removal took place through a particle
diffusion mechanism. The adsorption was found to be
endothermic in nature and followed both the Langmuir
and Freundlich models. Red mud was also used for the
dynamic uptake of 2,4-dinitrophenol, which was sorbed
by the developed adsorbent up to 95% [88]. The
removal of this pollutant reached 96% by column
experiments at the flow rate of 0.5 mL/min. The
adsorption process was found to be endothermic in
nature.
Tor et al. [89] conducted a study using neutralized
red mud for the removal of phenol from aqueous solution. To neutralize red mud, first the alkaline red mud
was suspended in distilled water with a liquid to solid
ratio of 2:1 on a weight basis, then stirred until the
Table 3.

equilibrium pH reached 8.08.5, and then dried. The


experiments demonstrated that maximum phenol
removal was obtained in a wide pH range of 19 and it
took 10 h to attain equilibrium. The maximum monolayer sorption capacity for the adsorption of phenol by
neutralized red mud was found to be 4.12 mg/g. The
Freundlich isotherm model fitted well with the data.
The effect of competitive ions like chloride, sulphate
and nitrate was also examined. It was found that phenol
adsorption decreased from 66% to 32% and 15% in the
case of nitrate and sulphate, respectively. However,
increasing the chloride ion concentration did not show
an adverse effect on the phenol adsorption on to red
mud.
In another study by Tor et al. [90], the removal of
phenol from aqueous solution by using HCl-activated
red mud was examined. It was found that the maximum removal was obtained at a pH below 8 and the
adsorption equilibrium time was 10 h. The pseudosecond-order model fitted well with the kinetic data.
The Langmuir isotherm represented the adsorption
data better than the Freundlich isotherm. The phenol
adsorption capacity of the ARM (8.156 mg/g) was
found to be higher than that of the neutralized red mud
(4.127 mg/g) at pH 6 and 25 1 C. It was stated by
the authors that chemisorption is the rate-determining
step during the process. A summary of adsorption
capacity of red mud for different phenolic pollutants is
presented in Table 3.
It is evident from the available literature that,
compared with metal ions, there are fewer reports available investigating the potential of red mud for the
removal of phenolic pollutants. Most of these studies
are limited only to phenol or its chloro- substitutes.
Additionally, red mud showed low affinity for phenol
removal and there is a strong need to conduct extensive
research to enhance the removal efficiencies/adsorption
capacity of red mud for different classes of phenols
after appropriate treatment. Furthermore, the mechanism of phenol adsorption on red mud also needs to be
studied in detail as most of the articles focused only on
the adsorption potential (adsorption capacity) of red
mud for phenol removal, and little effort has been made
to elucidate the sorption mechanism.

Adsorption capacity of red mud for the removal of phenolic pollutants from water.

Adsorbent

Adsorbate

Amount adsorbed

Reference

Red mud
Red mud
Red mud
Red mud
Neutralized red mud
Acid-activated red mud

Phenol
2-chlorophenol
4-chlorophenol
2,4-dichlorophenol
Phenol
Phenol

0.630.74 mol/g
0.720.79 mol/g
0.780.82 mol/g
0.800.85 mol/g
4.12 mg/g
8.16 mg/g

[87]
[87]
[87]
[87]
[89]
[90]

Environmental Technology

2.4. Application of red mud for the removal of


inorganic anions from water
Inorganic anions are one of the important classes of
aquatic pollutants, and various inorganic anions have
been found in potentially harmful concentrations in
numerous drinking water sources. The removal of these
pollutants from drinking water supplies is an emerging
issue. Red mud was examined for the removal of
different anions from water and wastewater. Red mud
with the addition of commercial peat for phosphorus
removal was examined [91]. Phosphorus removal was
found to increase from 1721% on peat alone to over
95% with red mud treatment of peat in column study.
The use of red mud could reduce phosphorus concentration in effluent to below 0.15 mg/L.
The uptake of phosphate by red mud activated by
heat treatment and acid-heat treatment was investigated
[92]. The result showed that the red mud sample treated
using the acid-heat method at 80 C with 0.25 mol/L
HCl for 2 h achieved the highest phosphate removal.
For the heat-activated red mud, the sample heated at
700 C for 2 h preformed better than the other heat treatment. Phosphate removal by the ARM was significantly
pH dependent, and pH 7 was the optimal pH for phosphate removal. The Langmuir isotherm model fitted
well with the data and the maximum adsorption
capacities of the acid-heat-activated red mud and the
heat-activated samples were 202.9 mg /g and 155.2 mg
P/g, respectively.
Huang et al. [93] investigated the effects of acid
treatment using different acids, such as HCl and HNO3,
combined with heat treatment on the properties and
adsorptive behaviour of the modified red mud, with the
aim of understanding the chemical changes of different
pretreatments of red mud for phosphate removal. It was
found that all ARM samples showed higher surface area
and total pore volume as well as a higher phosphate
adsorption capacity as compared to the untreated RM.
The red mud with HCl treatment showed the highest
adsorption capacity (0.58 mg P/g at pH 5.5 and 40 C)
among all the red mud samples. The adsorption capacity
of the red mud adsorbents decreased with increase in
pH. At pH 2, the red mud with HCl treatment exhibited
adsorption of 0.8 mg P/g, whereas the adsorption
decreased to 0.05 mg P/g at pH 10. However, the
adsorption was improved by 25% by increasing the
temperature from 30 to 40 C. The kinetic studies of
phosphate adsorption on to red mud indicated that the
adsorption mainly followed the parallel first-order
kinetics because of the presence of two acidic phosphorus species, H2PO4 and HPO42.
The removal of orthophosphate from aqueous
solution at a concentration range of 1 to 5 mg/L was
also tested using ARM [94]. The relationships between
activation conditions (including concentration of acid

241

and treatment temperature), pH of contact solution and


adsorptive capacity were checked and discussed. The
thermal pretreatment for acid-activated red mud was
detrimental to the adsorptive capacity. Acid-activated
red mud was found to be efficient for phosphate
removal from aqueous solution as Alumina F1 (one of
the best phosphate adsorbents).
Koumanova et al. [95] treated red mud with concentrated sulphuric acid, filtered the acid suspension,
washed (pH 7) and dried the residue, ground it to a
powder and then applied it for phosphate removal. The
influence of acid to mud ratio, and contact time between
them, on the extent of phosphate removal was studied.
Red mud treated for 2 h with concentrated H2SO4 (20
mL conc. H2SO4 per gram of mud) was used. It was
postulated that P3O10 ions could combine with four or
six bonds on the surface.
The effect of acidification and heat treatment of raw
red mud and fly ash on the sorption of phosphate was
studied in parallel experiments [96]. The uptake of phosphate by the red mud and fly ash, thermally activated at
various temperatures and acid-activated at various HCl
concentrations, was investigated by batch method. The
sorption capacity was found to increase by activation.
The sample which was prepared by stirring the red mud
with 0.25 M HCl for 2 h, as well as another sample
prepared by heating the red mud at 700 C for 2 h,
showed the maximum removal of phosphate (99%
removal of phosphate) at pH 7.0, 25 C and initial PO43
concentration of 155 mg P/L. Solution pH significantly
influenced the sorption. Each sample achieved the maximal removal of phosphate at pH 7.0. The amount of
phosphate removal increased with the solute concentration. The Langmuir model showed a better correlation
with the experimental data than the Freundlich model.
Red mud was converted into granular adsorbent
(RMGA) with bentonite and starch as the main raw
materials and was further evaluated for phosphate
removal [97]. The important parameters, which greatly
affect the characteristics of RMGA, e.g. the mass ratio
of three raw materials, preheating time, preheating
temperature, sintering time and sintering temperature
(ST), were investigated. It was concluded that RM ratio
and ST affected the characteristics of RMGA greatly,
and that the optimum parameters, under which the largest adsorption capacities could be achieved, varied with
different aquatic temperatures (AT). The optimum ST
was 1080 C, 1050 C and 1030 C for RMGA-85%
(RM:bentonite:starch=85:10:5) under an AT of 17 C,
27 C and 37 C, respectively, and for RMGA-90%
(RM:bentonite:starch=90:5:5) it was 1050 C, 1010 C
and 980 C, respectively, for the same ATs.
Treated red mud was found to effectively adsorb
phosphorus from dilute aqueous solution [98]. The pH
value of 4.5 was found to be optimum for maximum

242

A. Bhatnagar et al.

removal. The equilibrium was attained within 60 min. A


lower adsorbent dose and a higher initial phosphorus
concentration favoured higher loading capacity. The
Freundlich isotherm model fitted well with the equilibrium adsorption data. Lagergrens model fitted with the
sorption kinetics. The effect of different anions on
phosphorus removal could be explained on the basis of
the changing affinity of anions for the surface and their
relative concentrations.
Phosphate removal by Bauxsol was investigated as
a function of time, pH, ionic strength, adsorbent dosage,
competing ions and initial phosphate concentration
[99]. The results of adsorption and desorption studies
indicated that adsorption of PO43 by Bauxsol was
based on a ligand-exchange mechanism, although the
low reversibility, pH-independent desorption observed
in acid-treated Bauxsol indicated a dominance of
chemisorption. It was found that PO43 adsorption on to
both Bauxsol and acid-treated Bauxsol followed the
Langmuir isotherm model, with adsorption capacities of
0.21 and 0.48 mmol/g at pH 9.0 and 5.2, respectively.
Adsorption of PO43 by Bauxsol increased with decreasing pH, with maximum adsorption efficiencies obtained
at pH 5.2 0.1 (the lowest pH investigated), higher
Bauxsol to initial phosphate concentration ratios and
increased time. Studies of the effects of competing ions
on the adsorption of PO43 by Bauxsol indicated that
adsorption decreased in the presence of HCO3 ions,
whereas SO42 and Cl ions had little effect, and Ca2+
and Mg2+ ions increased adsorption.
The removal of fluoride from aqueous solution
using the original and HCl-activated red mud forms has
also been studied [100]. The fluoride adsorption capacity of the activated form was found to be higher than
that of the original form. The maximum removal of
fluoride ion was obtained at pH 5.5. The Langmuir
isotherm was found to fit well with the experimental
data. It was found that the required time for adsorption
equilibrium of fluoride ions was 2 h. The removal of
fluoride ion using red mud was explained on the basis
of the chemical nature and specific interaction with
metal oxide surfaces, and the results were interpreted in
terms of pH variations.
Recently, Tor et al. [101] also reported the feasibility of granular red mud (GRM) for the removal of fluoride from water. The experiments demonstrated that
maximum fluoride removal (0.644 mg/g) was obtained
at pH 4.7 and it took 6 h to attain equilibrium. Also,
equilibrium time did not depend upon the initial fluoride concentration. The pseudo-second-order model
fitted well with the kinetic data. The RedlichPeterson
and Freundlich isotherm models better represented the
adsorption data in comparison with the Langmuir
model. The capacities of the breakthrough and exhaustion points were found to decrease with increase in the

flow rate. The Thomas model was applied to the


experimental results. The results showed that the
sorption capacities of the columns were higher (0.773
1.274 mg/g) than their respective batch capacities
(0.644 mg/g) for the same initial fluoride concentration
(5 mg/L). The column adsorption was reversed and the
regeneration operation was accomplished by pumping
0.2 M of NaOH through the loaded GRM column.
Red mud was also modified with AlC13 (MRMA)
and by heat activation (MRMAH) and tested for the
removal of fluoride from water [102]. The results
showed that the adsorption capacities of MRMA and
MRMAH were 68.07 and 91.28 mg/g, respectively,
which were much higher than that of RM 13.46 mg/g.
The Langmuir isotherm was the best-fit adsorption
isotherm model for the experimental data. The solution
pH values affected the removal efficiency significantly,
and the highest removal efficiency was achieved at
pH 78.
Cengeloglu et al. [103] studied the removal of
nitrate from aqueous solution by using original and
HCl-activated red mud in a batch adsorption technique.
Adsorption equilibrium was achieved in 60 min. The
nitrate adsorption capacity of the ARM was found to be
higher than that of the original form and decreased
above pH 7. The adsorption capacity of the original red
mud and the ARM was found to be 1.859 and 5.858
mmol nitrate/g red mud, respectively. The increase in
sorption capacity in red mud after acid treatment was
attributed to the leaching of the sodalite compounds
during acid treatment, which are expected to hinder the
adsorption by blocking the available adsorption sites for
nitrate in untreated red mud. The mechanism for nitrate
removal was explained by the chemical nature of the red
mud and the interaction between metal oxide surfaces
and nitrate ions. A summary of the adsorption capacity
of red mud for different anions is provided in Table 4.
Recent studies have shown that red mud can be
successfully applied for the removal of some anions,
especially phosphate and fluoride. However, it is necessary to conduct more research on the application of red
mud for other anions also. Additionally, mechanistic
studies with anions need to be performed in detail to
understand the binding mechanism. There is, as yet,
little information in the literature on this topic.
Figure 1 summarizes the utilization of waste red
mud as an adsorbent for the removal of various pollutants from water and wastewater.
Tables 5 and 6 represent the results of adsorption
isotherms and thermodynamic studies of adsorption of
different pollutants on to red mud. In most studies,
adsorption systems are well represented by the Freundlich and Langmuir isotherm models. A negative Gibbs
free energy change (G) indicates the feasibility and
spontaneous nature of the adsorption process, whereas a
Figure 1.

Red mud as adsorbent for the removal of aquatic pollutants from water and wastewater.

Environmental Technology
Table 4.

243

Adsorption capacity of red mud for the removal of different anions from water.

Adsorbent

Adsorbate

Amount adsorbed

Reference

Heat-activated red mud


Acid-heat-activated red mud
HCl-treated red mud
Bauxsol
Red mud
Red mud modified with AlC13 (MRMA)
Red mud modified with heat activation (MRMAH)
Red mud
Activated red mud

Phosphate
Phosphate
Phosphate
Phosphate
Fluoride
Fluoride
Fluoride
Nitrate
Nitrate

155.2 mg/g
202.9 mg/g
0.58 mg/g
0.210.48 mmol/g
13.46 mg/g
68.07 mg/g
91.28 mg/g
1.859 mmol/g
5.858 mmol/g

[92]
[92]
[93]
[99]
[102]
[102]
[102]
[103]
[103]

Figure 1.

Red mud as adsorbent for the removal of aquatic pollutants from water and wastewater.

positive enthalpy change (H) represents the endothermic nature of an adsorption. In some cases, negative
values of enthalpy change were reported indicating the
exothermic nature of the process. A positive entropy
change (S) denotes the affinity of the red mud and
increasing randomness at the solidsolution interface
during the sorption of adsorbates on active sites of the
adsorbent, whereas a negative value of entropy change
in some cases indicates that the degree of freedom
decreases at the solidliquid interface during the
adsorption.
3. Conclusions and future perspectives
In this review, the sorption properties of red mud, as an
adsorbent for the removal of diverse type of pollutants
from water and wastewater, have been reviewed based

on a substantial number of relevant published articles.


As can be seen from the literature reviewed in this
paper, red mud has been found to be efficient for the
removal of different types of metal ions and inorganic
anions from water and wastewater; however, less work
has been conducted on the removal of dyes and phenols.
The neutralization or modification of raw red mud with
acid or heat treatment was found to considerably
improve the sorption capacity in many studies. Among
various process parameters, pH was found to be one of
the important factors affecting the sorption process.
There are still several issues that need more attention
in future studies, such as enhancement of sorption
capacity through modification, assessment of sorbent
under multi-component pollutants, mechanistic modelling to correctly understand the sorption mechanisms,
investigation of these materials with real industrial

Adsorbate

As(III)
As(III)
As(III)
As(III)
As(V)
As(V)
As(V)
As(V)
As(III)
As(III)
As(III)
As(III)
As(V)
As(V)
As(V)
As(V)
As(V)
As(V)
As(V)
As(V)
As(V)
As(V)
As(III)
As(V)
Cu(II)
Cd(II)
Cd(II)
Cd(II)
Pb(II)
Pb(II)
Pb(II)
Cr(VI)
Cr(VI)

Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Activated red mud
Activated red mud
Activated red mud
Activated red mud
Activated red mud
Activated red mud
Activated red mud
Activated red mud
Bauxsol
Activated Bauxsol
Activated Bauxsol
Activated Bauxsol
Activated Bauxsol
Bauxsol
Activated Bauxsol
Modified red mud
Activated red mud
Granular red mud
Granular red mud
Granular red mud
Red mud
Red mud
Red mud
Red mud
Red mud

25
40
55
70
25
40
55
70
25
40
55
70
25
40
55
70
23
15
20
23
50
23
23
20
30
20
30
40
30
40
50
30
40

Temperature
(C)
b

8.86 mol/g
0.025 L/mol
7.93 mol/g
0.018 L/mol
7.17 mol/g
0.016 L/mol
6.18 mol/g
0.017 L/mol
6.86 mol/g
0.123 L/mol
7.73 mol/g
0.128 L/mol
9.60 mol/g
0.134 L/mol
10.80 mol/g
0.135 L/mol
11.80 mol/g
0.073 L/mol
8.85 mol/g
0.070 L/mol
7.93 mol/g
0.059 L/mol
4.49 mol/g
0.050 L/mol
12.57 mol/g
0.208 L/mol
14.99 mol/g
0.280 L/mol
17.15 mol/g
0.273 L/mol
17.71 mol/g
0.410 L/mol
6.0814.43 mol/g 0.010.76 L/mol
25.91 mmol/g
0.41 L/mol
28.90 mmol/g
0.52 L/mol
39.84 mmol/g
0.97 L/mol
40.98 mmol/g
1.01 L/mol
6.0814.43 mmol/g 0.010.76 L/mol
7.22 mmol/g
0.032 L/mol
23.268.5 mg/g
0.0900.142 L/mg
5.349 mg/g
0.2045 L/mg

15.90 103 L/mol


3.44 104 mol/g
4.80 103 L/mol
3.39 104 mol/g
4
2.05 103 L/mol
3.23 10 mol/g
4
0.40 103 L/mol
4.36 10 mol/g
4
0.371 103 L/mol
4.15 10 mol/g

qm

Langmuir constants

Results of adsorption isotherm studies of different pollutants on red mud adsorbents.

Adsorbent

Table 5.

5.93 L/mol
6.71 L/mol
17.08 L/mol
16.85 L/mol
1.066.70 L/mol
0.37 L/mol
6.718.2
1.4086
9.86 mg/g
10.02 mg/g
10.10 mg/g
1.45 104

2.24 104
0.31

KF

0.51 (M)/(M)n
0.59 (M)/(M)n
0.58 (M)/(M)n
0.42 (M)/(M)n
0.390.80 (M)/(M)n
0.59 (M)/(M)n
0.2790.365
2.501
0.25
0.27
0.31
0.20

1/n

Freundlich constants

(Continued)

[41]
[41]
[41]
[41]
[41]
[41]
[41]
[41]
[42]
[42]
[42]
[42]
[42]
[42]
[42]
[42]
[45]
[46]
[46]
[46]
[46]
[46]
[46]
[49]
[61]
[62]
[62]
[62]
[65]
[65]
[65]
[65]
[65]

Reference

244
A. Bhatnagar et al.

Adsorbate

Cr(VI)
Cd(II)
Cd(II)
Cd(II)
Zn(II)
Zn(II)
Zn(II)
Rhodamine B
Rhodamine B
Rhodamine B
Fast Green
Fast Green
Fast Green
Methylene Blue
Methylene Blue
Methylene Blue
Congo Red
Acid Violet
Methylene Blue
Phenol
Phenol
Phenol
2-chlorophenol
2-chlorophenol
2-chlorophenol
4-chlorophenol
4-chlorophenol
4-chlorophenol
2,4-dichlorophenol
2,4-dichlorophenol
2,4-dichlorophenol
Phenol
Phenol

Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Neutralized red mud
Activated red mud

(Continued).

Adsorbent

Table 5.

30
40
50
30
40
50
30
40
50
30
40
50

50
30
40
50
30
40
50
30
40
50
30
40
50
30
40
50

Temperature
(C)
b
0.316 103 L/mol
8.55 103 L/mol
2.88 103 L/mol
1.04 103 L/mol
1.07 103 L/mol
2.75 103 L/mol
1.54 103 L/mol
14.28 103 L/mol
7.76 103 L/mol
4.08 103 L/mol
4.88 104 L/mol
1.66 104 L/mol
0.61 104 L/mol
1.79 102 L/mol
1.28 102 L/mol
0.84 102 L/mol
0.0360 L/mg
6.91 102 L/mg
8.88 104 L/mol
39.00 L/mol
49.20 L/mol
68.70 L/mol
40.00 L/mol
50.00 L/mol
71.00 L/mol
42.00 L/mol
51.80 L/mol
72.00 L/mol
44.10 L/mol
53.10 L/mol
72.21 L/mol
0.051 L/mg
0.101 L/mg

qm
4.05 104 mol/g
1.16 104 mol/g
1.06 104 mol/g
1.00 104 mol/g
2.22 104 mol/g
2.00 104 mol/g
1.81 104 mol/g
1.16 105 mol/g
1.06 105 mol/g
1.01 105 mol/g
9.35 106 mol/g
8.77 106 mol/g
7.25 106 mol/g
5.23 105 mol/g
4.81 105 mol/g
4.35 105 mol/g
4.05 mg/g
1.3724 mg/g
7.8 106 mol/g
0.63 mol/g
0.68 mol/g
0.74 mol/g
0.72 mol/g
0.75 mol/g
0.79 mol/g
0.78 mol/g
0.80 mol/g
0.82 mol/g
0.80 mol/g
0.82 mol/g
0.85 mol/g
4.127 mg/g
8.156 mg/g

Langmuir constants

0.15 104 mol/g

0.23 104 mol/g

1.92 105 [(mg/g) (L/mg)1/n]


1.62 105 [(mg/g) (L/mg)1/n]
1.38 105 [(mg/g) (L/mg)1/n]
1.63 105 [(mg/g) (L/mg)1/n]
1.19 105 [(mg/g) (L/mg)1/n]
9.88 106 [(mg/g) (L/mg)1/n]
4.57 105 [(mg/g) (L/mg)1/n]
3.55 105 [(mg/g) (L/mg)1/n]
2.82 105 [(mg/g) (L/mg)1/n]

0.579
3.90
4.55
4.60
4.60
4.65
4.70
4.72
2.74
4.80
4.82
4.83
4.85
0.513 mg/g
1.211 mg/g

0.51

1.785
1.333
1.282
1.754
1.369
1.333
4.166
3.703
1.754

3.65 103 mol/g


0.081
0.078
0.057
0.064
0.063
0.055
0.062
0.061
0.050
0.059
0.056
0.046
0.411
0.421

KF

0.58

1/n

Freundlich constants

[65]
[66]
[66]
[66]
[66]
[66]
[66]
[78]
[78]
[78]
[78]
[78]
[78]
[78]
[78]
[78]
[79]
[81]
[83]
[87]
[87]
[87]
[87]
[87]
[87]
[87]
[87]
[87]
[87]
[87]
[87]
[89]
[90]

Reference

Environmental Technology
245

246
Table 6.

A. Bhatnagar et al.
Results of thermodynamic studies of different pollutants on to red mud adsorbents.

Adsorbent

Adsorbate

Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Activated red mud
Activated red mud
Activated red mud
Activated red mud
Activated red mud
Activated red mud
Activated red mud
Activated red mud
Granular red mud
Granular red mud
Granular red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud
Red mud

As(III)
As(III)
As(III)
As(III)
As(V)
As(V)
As(V)
As(V)
As(III)
As(III)
As(III)
As(III)
As(V)
As(V)
As(V)
As(V)
Cd(II)
Cd(II)
Cd(II)
Pb(II)
Pb(II)
Pb(II)
Cr(VI)
Cr(VI)
Cr(VI)
Cd(II)
Cd(II)
Cd(II)
Zn(II)
Zn(II)
Zn(II)
Rhodamine B
Rhodamine B
Rhodamine B
Fast Green
Fast Green
Fast Green
Methylene Blue
Methylene Blue
Methylene Blue
2,4-dichlorophenol
2,4-dichlorophenol
2,4-dichlorophenol

Temperature
(C)

G (kJ/mol)

25
40
55
70
25
40
55
70
25
40
55
70
25
40
55
70
20
30
40
30
40
50
30
40
50
30
40
50
30
40
50
30
40
50
30
40
50
30
40
50
30
40
50

25.11
25.58
26.41
27.81
29.06
30.62
32.22
33.71
27.75
29.06
30.00
30.90
30.36
32.67
34.16
36.88
4.546
5.620
5.933
22.82
20.74
18.67
23.39
20.62
19.72
22.8
20.7
18.7
23.9
20.6
19.7
24.1
23.3
22.3
27.2
25.3
23.4
13.1
12.6
11.9
21.98
23.82
26.48

H
(kJ/mol)
12.83

1.85

6.88

7.08

15.822

S
(kJ/(molK))
0.0412
0.0407
0.0415
0.0438
0.1037
0.1034
0.1038
0.1034
0.0699
0.0708
0.0703
0.0699
0.1256
0.1269
0.1256
0.1281
0.069

85.94

0.208

78.50

0.183

85.9

0.21

78.5

0.18

67.1

140.1

84.6

189.6

31.0

59.1

46.50

0.23

Reference
[41]
[41]
[41]
[41]
[41]
[41]
[41]
[41]
[42]
[42]
[42]
[42]
[42]
[42]
[42]
[42]
[62]
[62]
[62]
[65]
[65]
[65]
[65]
[65]
[65]
[66]
[66]
[66]
[66]
[66]
[66]
[78]
[78]
[78]
[78]
[78]
[78]
[78]
[78]
[78]
[87]
[87]
[87]

Environmental Technology

effluents, recovery of metal ions, regeneration studies


and continuous flow studies.

[17]

Acknowledgements
Financial support for this work was in part provided by LSRE
financing by FEDER/POCI/2010. Amit Bhatnagar acknowledges his post-doctoral scholarship (DFRH-SFRH/BPD/
62889/2009) supported by the Portuguese Foundation for
Science and Technology (FCT).

[18]

[19]

References
[1] B.H. Tan, T.T. Teng, and A.K.M. Omar, Removal of
dyes and industrial dye wastes by magnesium chloride, Water Res. 34 (2000), pp. 597601.
[2] P. Mavros, A.C. Danilidou, N.K. Lazaridis, and L. Stergiou, Color removal from aqueous solutions. Part I.
Flotation, Environ. Technol. 15 (1994), pp. 601616.
[3] A.I. Zouboulis, N.K. Lazaridis, and A. Grohmann,
Toxic metals removal from waste waters by upflow
filtration with floating filter medium. I. The case of
zinc, Sep. Sci. Technol. 37 (2002), pp. 403416.
[4] B. Bolto, D. Dixon, R. Eldridge, S. King, and K.
Linge, Removal of natural organic matter by ion
exchange, Water Res. 36 (2002), pp. 50575065.
[5] T.M. LaPara, A. Konopka, C.H. Nakatsu, and J.E.
Alleman, Thermophilic aerobic wastewater treatment
in continuous-flow bioreactors, J. Environ. Eng.
ASCE 126 (2000), pp. 739744.
[6] J. Bell, J.J. Plumb, C.A. Buckley, and D.C. Stuckey,
Treatment and decolorization of dyes in an anaerobic
baffled reactor, J. Environ. Eng. ASCE 126 (2000),
pp. 10261032.
[7] S. Esplugas, J. Gimenez, S. Contreras, E. Pascual, and
M. Rodriguez, Comparison of different advanced
oxidation processes for phenol degradation, Water
Res. 36 (2002), pp. 10341042.
[8] S.H. Lin and R.S. Juang, Removal of free and chelated
Cu(II) ions from water by a nondispersive solvent extraction process, Water Res. 36 (2002), pp. 36113619.
[9] S.D. Faust and O.M. Aly, Adsorption Process for
Water Treatment, Butterworths, Boston, 1987.
[10] L. Szpyrkowicz, J. Naumczyk, and F. Zilio-Grandi,
Electrochemical treatment of tannery wastewater
using Ti/Pt and Ti/Pt/Ir electrodes, Water Res. 29
(1995), pp. 517524.
[11] H. Shen and Y.-T. Wang, Biological reduction of
chromium by E. coli, J. Environ. Eng. ASCE 120
(1994), pp. 560571.
[12] A. Pala and E. Tokat, Color removal from cotton
textile industry wastewater in an activated sludge
system with various additives, Water Res. 36 (2002),
pp. 29202925.
[13] R.C. Bansal and M. Goyal, Activated Carbon Adsorption, CRC Press, Boca Raton, FL, 2005.
[14] W.K. Buah and P.T. Williams, Activated carbons
prepared from refuse derived fuel and their gold
adsorption characteristics, Environ. Technol. 31
(2010), pp. 125137.
[15] B.K. Vu, O. Snisarenko, H.S. Lee, and E.W. Shin,
Adsorption of tetracycline on La-impregnated MCM-41
materials, Environ. Technol. 31 (2010), pp. 233241.
[16] S.-M. Lee, W.-G. Kim, J.-K. Yang, and D. Tiwari,
Sorption behaviour of manganese-coated calcined-

[20]

[21]

[22]

[23]
[24]
[25]
[26]

[27]

[28]

[29]

[30]

[31]

[32]

247

starfish and manganese-coated sand for Mn(II), Environ. Technol. 31 (2010), pp. 445453.
B. Cheknane, M. Baudu, J.-P. Basly, and O. Bouras,
Adsorption of basic dyes in single and mixture systems
on granular inorganic-organic pillared clays, Environ.
Technol. 31 (2010), pp. 815822.
M. Cui, M. Jang, S.-H. Cho, and J. Khim, Kinetic and
thermodynamic studies of the adsorption of heavy
metals on to a new adsorbent: Coal mine drainage
sludge, Environ. Technol. 31 (2010), pp. 12031211.
N. Boujelben, J. Bouzid, Z. Elouear, and M. Feki,
Retention of nickel from aqueous solutions using iron
oxide and manganese oxide coated sand: Kinetic and
thermodynamic studies, Environ. Technol. 31 (2010),
pp. 16231634.
S. Kumar, R. Kumar, and A. Bandopadhyay, Innovative methodologies for the utilization of wastes from
metallurgical and allied industries, Resour. Conserv.
Recycl. 48 (2006), pp. 301314.
V.P. Gadepalle, S.K. Ouki, R. van Herwijnen, and T.
Hutchings, Immobilization of heavy metals in soil
using natural and waste materials for vegetation
establishment on contaminated sites, Soil Sediment
Contam. 16 (2007), pp. 233251.
E. Lombi, F.J. Zhao, G.Y. Zhang, B. Sun, W. Fitz, H.
Zhang, and S.P. McGrath, In situ fixation of metals in
soils using bauxite residue: Chemical assessment,
Environ. Pollut. 118 (2002), pp. 435443.
M. Singh, S.N. Upadhayay, and P.M. Prasad, Preparation of iron rich cement from red mud, Cem. Concr.
Res. 27 (1997), pp. 10371046.
V.A. Mymrin, Red mud of aluminium production
waste as basic component of new construction materials, Waste Manage. Res. 19 (2001), pp. 465469.
J. Yang, D. Zhang, J. Hou, B. He, and B. Xiao, Preparation of glass-ceramics from red mud in the aluminium industries, Ceram. Int. 34 (2008), pp. 125130.
W. Liu, J. Yang, and B. Xiao, Application of Bayer
red mud for iron recovery and building material
production from alumosilicate residues, J. Hazard.
Mater. 161 (2009), pp. 474478.
Y. Cengeloglu, E. Kir, M. Ersoz, T. Buyukerkek, and
S. Gezgin, Recovery and concentration of metals from
red mud by Donnan dialysis, Colloids Surf. A 223
(2003), pp. 95101.
S. Wang, H.M. Ang, and M.O. Tad, Novel applications of red mud as coagulant, adsorbent and catalyst
for environmentally benign processes, Chemosphere
72 (2008), pp. 16211635.
Y.-F. Zhou, and R.J. Haynes, Sorption of heavy metals
by inorganic and organic components of solid wastes:
Significance to use of wastes as low cost adsorbents
and immobilizing agents, Crit. Rev. Environ. Sci.
Technol. 40 (2010), pp. 909977.
S.J.T. Pollard, G.D. Fowler, C.J. Sollars, and R. Perry,
Low-cost adsorbents for waste and wastewater
treatment: A review, Sci. Total Environ. 116 (1992),
pp. 3152.
G.S. Miguel, S.D. Lambert, and N.J.D. Graham, A
practical review of the performance of organic and
inorganic adsorbents for the treatment of contaminated waters, J. Chem. Technol. Biotechnol. 81
(2006), pp. 16851696.
G. Crini, Non-conventional low-cost adsorbents for
dye removal: A review, Bioresour. Technol. 97 (2006),
pp. 10611085.
gevb[r]

248

A. Bhatnagar et al.

[33] D. Mohan and C.U. Pittman Jr, Activated carbons and


low cost adsorbents for remediation of tri- and
hexavalent chromium from water, J. Hazard. Mater.
137 (2006), pp. 762811.
[34] M. Ahmaruzzaman, Adsorption of phenolic
compounds on low-cost adsorbents: A review, Adv.
Colloid Interface Sci. 143 (2008), pp. 4867.
[35] S.B. Wang and H.W. Wu, Environmental-benign utilization of fly ash as low-cost adsorbents, J. Hazard.
Mater. 136 (2006), pp. 482501.
[36] A. Bhatnagar and M. Sillanp, Utilization of agroindustrial and municipal waste materials as potential
adsorbents for water treatment: A review, Chem. Eng.
J. 157 (2010), pp. 277296.
[37] M. Bissen and F.H. Frimmel, Arsenic a review. Part
I. Occurrence, toxicity, speciation, mobility, Acta
Hydrochim. Hydrobiol. 31 (2003), pp. 918.
[38] J.M. Azcue and J.O. Nriagu, Arsenic: Historical
perspectives. In J.O. Nriagu, ed., Arsenic in the Environment. Part 1: Cycling and Characterization. John
Wiley, New York, 1994, pp. 115.
[39] L. Wang and J. Huang, in Arsenic in the Environment.
Part II. Human Health and Ecosystem Effects, J.O.
Nriagu, ed., Wiley, New York, 1994, pp. 159172.
[40] D. Mohan and C.U. Pittman Jr, Arsenic removal from
water/wastewater using adsorbents: A critical review,
J. Hazard. Mater. 142 (2007), pp. 153.
[41] H.S. Altundogan, S. Altundogan, F. Tmen, and M.
Bildik, Arsenic removal from aqueous solutions by
adsorption on red mud, Waste Manage. 20 (2000),
pp. 761767.
[42] H.S. Altundogan, S. Altundogan, F. Tmen, and M.
Bildik, Arsenic adsorption from aqueous solutions by
activated red mud, Waste Manage. 22 (2002),
pp. 357363.
[43] H.S. Altundogan and F. Tmen, As(V) removal from
aqueous solutions by coagulation with liquid phase of
red mud, J. Environ. Sci. Health. A 38 (2003), pp.
12471258.
[44] L.C. White, E. Paling, P. Singh, and W. Zhang,
Removal of arsenic by red mud from contaminated waste water, Proceedings of the TMS Fall
Extraction and Processing Conference, 2, 2003,
pp. 19511957.
[45] H. Gen, J.C. Tjell, D. McConchie, and O. Schuiling,
Adsorption of arsenate from water using neutralized red
mud, J. Colloid Interface Sci. 264 (2003), pp. 327334.
[46] H. Gen-Fuhrman, J.C. Tjell, and D. McConchie,
Adsorption of arsenic from water using activated
neutralized red mud, Environ. Sci. Technol. 38 (2004),
pp. 24282434.
[47] H. Gen-Fuhrman, J.C. Tjell, and D. McConchie,
Increasing the arsenate adsorption capacity of
neutralized red mud (Bauxsol), J. Colloid Interface
Sci. 271 (2004), pp. 313320.
[48] H. Gen-Fuhrman, J.C. Tjell, and D. McConchie,
Arsenate removal from water using sand-red mud
columns, Water Res. 39 (2005), pp. 29442954.
[49] S. Zhang, C. Liu, Z. Luan, X. Peng, H. Ren, and J.
Wang, Arsenate removal from aqueous solutions using
modified red mud, J. Hazard. Mater. 152 (2008),
pp. 486492.
[50] P. Castaldi, M. Silvetti, S. Enzo, and P. Melis, Study
of sorption processes and FT-IR analysis of arsenate
sorbed onto red muds (a bauxite ore processing
waste), J. Hazard. Mater. 175 (2010), pp. 172178.
gevb[r]

gevb[r]

gevb[r]

gevb[r]

gevb[r]

[51] R.C. Sahu, R. Patel, and B.C. Ray, Utilization of activated CO2-neutralized red mud for removal of arsenate from aqueous solutions, J. Hazard. Mater. 179
(2010), pp. 10071013.
[52] A.P. Singh, K.K. Srivastava, and H. Shekhar, Effect
of pH on removal of arsenic(III) by mixed adsorbents, Proc. Natl. Acad. Sci. India A 78 (2008), pp.
283286.
[53] A.P. Singh, K.K. Srivastava, and H. Shekhar, Kinetic
characterization of removal of As(III) by mixed adsorbents, J. Sci. Ind. Res. 66 (2008), pp. 952956.
[54] J. Pradhan, S.N. Das, and R.S. Thakur, Adsorption of
hexavalent chromium from aqueous solution by using
activated red mud, J. Colloid Interface Sci. 217
(1999), pp. 137141.
[55] S. Dursun, D. Guclu, A. Berktay, and T. Guner,
Removal of chromate from aqueous system by activated
red-mud, Asian J. Chem. 20 (2008), pp. 64736478.
[56] M. Rajkumar, N. Nagendran, and S.S. Sasikumar,
Removal of trivalent chromium from wastewater
using red mud, Ind. J. Environ. Prot. 21 (2001), pp.
97100.
[57] A.I. Zouboulis and K.A. Kydros, Use of red mud for
toxic metals removal the case of nickel, J. Chem.
Technol. Biotechnol. 58 (1993), pp. 95101.
[58] Y. Hannachi, N.A. Shapovalov, and A. Hannachi,
Adsorption of nickel from aqueous solution by the use
of low-cost adsorbents, Korean J. Chem. Eng. 27
(2010), pp. 152158.
[59] S. Smiljanic, I. Smiciklas, A. Peric-Grujic, B. Loncar,
and M. Mitric, Rinsed and thermally treated red mud
sorbents for aqueous Ni2+ ions, Chem. Eng. J. 162
(2010), pp. 7583.
[60] A. Agrawal, K.K. Sahu, and B.D. Pandey, A comparative adsorption study of copper on various industrial
solid wastes, AIChE J. 50 (2004), pp. 24302438.
[61] H. Nadaroglu, E. Kalkan, and N. Demir, Removal of
copper from aqueous solution using red mud, Desalination 251 (2010), pp. 9095.
[62] C. Zhu, Z. Luan, Y. Wang, and X. Shan, Removal of
cadmium from aqueous solutions by adsorption on
granular red mud (GRM), Sep. Purif. Technol. 57
(2007), pp. 161169.
[63] E. Lpez, B. Soto, M. Arias, A. Nnez, D. Rubinos,
and M.T. Barral, Adsorbent properties of red mud and
its use for wastewater treatment, Water Res. 32
(1998), pp. 13141322.
[64] R. Apak, E. Ttem, M. Hgl, and J. Hizal, Heavy metal
cation retention by unconventional sorbents (red muds
and fly ashes), Water Res. 32 (1998), pp. 430440.
[65] V.K. Gupta, M. Gupta, and S. Sharma, Process development for the removal of lead and chromium from
aqueous solutions using red mud an aluminium
industry waste, Water Res. 35 (2001), pp. 11251134.
[66] V.K. Gupta and S. Sharma, Removal of cadmium and
zinc from aqueous solutions using red mud, Environ.
Sci. Technol. 36 (2002), pp. 36123617.
[67] S.W. Han, D.K. Kim, I.G. Hwang, and J.H. Bae,
Development of pellet-type adsorbents for removal of
heavy metal ions from aqueous solutions using red
mud, J. Ind. Eng. Chem. 8 (2002), pp. 120125.
[68] M. Vaclavikova, P. Misaelides, G. Gallios, S. Jakabsky, and S. Hredzak, Removal of cadmium, zinc,
copper and lead by red mud, an iron oxides containing
hydrometallurgical waste, Stud. Surf. Sci. Catal. 155
(2005), pp. 517525.
ca[eu]t

ocanr[]

ca[eu]t

ca[eu]t

ca[eu]t

coa[rn]

Environmental Technology
[69] L. Santona, P. Castaldi, and P. Melis, Evaluation of the
interaction mechanisms between red muds and heavy
metals, J. Hazard. Mater. B 136 (2006), pp. 324329.
[70] L. Shi, X. Peng, Z. Luan, N. Wei, Q. Wang, and Y. Zhao,
Use of activated red mud to remove phosphate and heavy
metals from the effluent of biologically treated swine
wastewater, Huanjing Kexue Xuebao/Acta Scientiae
Circumstantiae 29 (2009), pp. 22822288.
[71] S.J. Palmer, M. Nothling, K.H. Bakon, and R.L. Frost,
Thermally activated seawater neutralised red mud
used for the removal of arsenate, vanadate and molybdate from aqueous solutions, J. Colloid Interface Sci.
342 (2010), pp. 147154.
[72] Y. Cengeloglu, A. Tor, G. Arslan, M. Ersoz, and S.
Gezgin, Removal of boron from aqueous solution by
using neutralized red mud, J. Hazard. Mater. 142
(2007), pp. 412417.
[73] C. Brunori, C. Cremisini, P. Massanissso, V. Pinto,
and L. Torricelli, Reuse of a treated red mud bauxite
waste: Studies on environmental compatibility, J.
Hazard. Mater. B 117 (2005), pp. 5563.
[74] Y-F. Zhou and R.J. Haynes, A comparison of
inorganic solid wastes as adsorbents of heavy metal
cations in aqueous solution and their capacity for
desorption and regeneration, Water Air Soil Pollut., in
press, doi: 10.1007/s11270-010-0659-7.
[75] G.E. Walsh, L.H. Bahner, and W.B. Horning, Toxicity
of textile mill effluents to freshwater and estuarine
algae, crustaceans and fishes, Environ. Pollut. A 21
(1980), pp. 169179.
[76] P.K. Ray, Environmental pollution and cancer, J. Sci.
Ind. Res. 45 (1986), pp. 370371.
[77] K.G. Bhattacharyya and A. Sharma, Azadirachta
indica leaf powder as an effective biosorbent for dyes:
A case study with aqueous Congo red solutions, J.
Environ. Manage. 71 (2004), pp. 217229.
[78] V.K. Gupta, Suhas, I. Ali, and V.K. Saini, Removal of
rhodamine B, fast green, and methylene blue from
wastewater using red mud, an aluminum industry
waste, Ind. Eng. Chem. Res. 43 (2004), pp. 17401747.
[79] C. Namasivayam and D.J.S.E. Arasi, Removal of
congo red from wastewater by adsorption onto waste
red mud, Chemosphere 34 (1997), pp. 401417.
[80] C. Namasivayam, R.T. Yamuna, and D.J.S.E. Arasi,
Removal of procion orange from wastewater by
adsorption on waste red mud, Sep. Sci. Technol. 37
(2002), pp. 24212431.
[81] C. Namasivayam, R.T. Yamuna, and D.J.S.E. Arasi,
Removal of acid violet from wastewater by adsorption on waste red mud, Environ. Geol. 41 (2001),
pp. 269273.
[82] A. Tor and Y. Cengeloglu, Removal of congo red from
aqueous solution by adsorption onto acid activated
red mud, J. Hazard. Mater. 138 (2006), pp. 409415.
[83] S. Wang, Y. Boyjoo, A. Choueib, and Z.H. Zhu,
Removal of dyes from aqueous solution using fly ash
and red mud, Water Res. 39 (2005), pp. 129138.
[84] F.A. Banat, B. Al-Bashir, S. Al-Asheh, and O. Hayajneh, Adsorption of phenol by bentonite, Environ.
Pollut. 107 (2000), pp. 391398.
[85] A.P. Singh, P.C. Singh, and V.N. Singh, Removal of
1-butanethiol from diesel oil by red mud, Ind. Eng.
Chem. Res. 27 (1988), pp. 21012104.
[86] C. Namasivayam and K. Thamaraiselvi, Adsorption of
2-chlorophenol by waste red mud, Fresenius Environ. Bull. 7 (1998), pp. 314319.

249

[87] V.K. Gupta, I. Ali, and V.K. Saini, Removal of


chlorophenols from wastewater using red mud: An
aluminum industry waste, Environ. Sci. Technol. 38
(2004), pp. 40124018.
[88] V.K. Gupta and I. Ali, Removal of 2,4-dinitrophenol
from wastewater by adsorption technology: A batch
and column study, Int. J. Environ. Pollut. 27 (2006),
pp. 104120.
[89] A. Tor, Y. Cengeloglu, M.E. Aydin, and M. Ersoz,
Removal of phenol from aqueous phase by using
neutralized red mud, J. Colloid Interface Sci. 300
(2006), pp. 498503.
[90] A. Tor, Y. Cengeloglu, and M. Ersoz, Increasing the
phenol adsorption capacity of neutralized red mud by
application of acid activation procedure, Desalination
242 (2009), pp. 1928.
[91] G. Roberge, J.F. Blais, and G. Mercier, Phosphorus
removal from wastewater treated with red mud-doped
peat, Can. J. Chem. Eng. 77 (1999), pp. 11851194.
[92] C-J. Liu, Y-Z. Li, Z-K. Luan, Z.Y. Chen, Z-G. Zhang,
and Z-P. Jia, Adsorption removal of phosphate from
aqueous solution by active red mud, J. Environ. Sci.
19 (2007), pp. 11661170.
[93] W. Huang, S. Wang, Z. Zhu, L. Li, X. Yao, V.
Rudolph, and F. Haghseresht, Phosphate removal from
wastewater using red mud, J. Hazard. Mater. 158
(2008), pp. 3542.
[94] S.J. Shiao and K. Akashi, Phosphate removal from
aqueous solution from activated red mud, J. Water
Pollut. Control Fed. 49 (1977), pp. 280285.
[95] B. Koumanova, M. Drame, and M. Popangelova,
Phosphate removal from aqueous solutions using red
mud wasted in bauxite Bayers process, Resour.
Conserv. Recycl. 19 (1997), pp. 1120.
[96] Y. Li, C. Liu, Z. Luan, X. Peng, C. Zhu, Z. Chen,
Z. Zhang, J. Fan, and Z. Jia, Phosphate removal
from aqueous solutions using raw and activated red
mud and fly ash, J. Hazard. Mater. 137 (2006), pp.
374383.
[97] Q. Yue, Y. Zhao, Q. Li, W. Li, B. Gao, S. Han, Y. Qi,
and H. Yu, Research on the characteristics of red mud
granular adsorbents (RMGA) for phosphate removal,
J. Hazard. Mater. 176 (2010), pp. 741748.
[98] S. Mohanty, J. Pradhan, S.N. Das, and R.S. Thakur,
Removal of phosphorus from aqueous solution using
alumized red mud, Int. J. Environ. Stud. 61 (2004),
pp. 687697.
[99] D.J. Akhurst, G.B. Jones, M. Clark, and D. McConchie,
Phosphate removal from aqueous solutions using neutralised bauxite refinery residues (Bauxsol), Environ.
Chem. 3 (2006), pp. 6574.
[100] Y. engeloglu, E. Kir, and M. Ersz, Removal of
fluoride from aqueous solution by using red mud, Sep.
Purif. Technol. 28 (2002), pp. 8186.
[101] A. Tor, N. Danaoglu, G. Arslan, and Y. Cengeloglu,
Removal of fluoride from water by using granular red
mud: Batch and column studies, J. Hazard. Mater. 164
(2009), pp. 271278.
[102] N. Wei, Z-K. Luan, J. Wang, L. Shi, Y. Zhao, and J-W.
Wu, Preparation of modified red mud with aluminum
and its adsorption characteristics on fluoride removal,
Chin. J. Inorg. Chem. 25 (2009), pp. 849854.
[103] Y. Cengeloglu, A. Tor, M. Ersoz, and G. Arslan,
Removal of nitrate from aqueous solution by using red
mud, Sep. Purif. Technol. 51 (2006), pp. 374378.
gevb[r]

Potrebbero piacerti anche