You are on page 1of 31

This article was downloaded by: [141.85.66.

12]
On: 31 January 2012, At: 02:24
Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954
Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH,
UK

Mineral Processing and


Extractive Metallurgy Review:
An International Journal
Publication details, including instructions for
authors and subscription information:
http://www.tandfonline.com/loi/gmpr20

UTILIZATION OF ALUMINA
RED MUD FOR SYNTHESIS
OF INORGANIC POLYMERIC
MATERIALS
a a
Dimitrios D. Dimas , Ioanna P. Giannopoulou &
a
Dimitrios Panias
a
National Technical University of Athens, School
of Mining and Metallurgical Engineering, Zografou,
Athens, Greece

Available online: 19 May 2009

To cite this article: Dimitrios D. Dimas, Ioanna P. Giannopoulou & Dimitrios Panias
(2009): UTILIZATION OF ALUMINA RED MUD FOR SYNTHESIS OF INORGANIC POLYMERIC
MATERIALS, Mineral Processing and Extractive Metallurgy Review: An International
Journal, 30:3, 211-239

To link to this article: http://dx.doi.org/10.1080/08827500802498199

PLEASE SCROLL DOWN FOR ARTICLE

Full terms and conditions of use: http://www.tandfonline.com/page/terms-


and-conditions

This article may be used for research, teaching, and private study purposes.
Any substantial or systematic reproduction, redistribution, reselling, loan,
sub-licensing, systematic supply, or distribution in any form to anyone is
expressly forbidden.

The publisher does not give any warranty express or implied or make any
representation that the contents will be complete or accurate or up to
date. The accuracy of any instructions, formulae, and drug doses should be
independently verified with primary sources. The publisher shall not be liable
for any loss, actions, claims, proceedings, demand, or costs or damages
whatsoever or howsoever caused arising directly or indirectly in connection
with or arising out of the use of this material.
Downloaded by [141.85.66.12] at 02:24 31 January 2012
Mineral Processing & Extractive Metall. Rev., 30: 211239, 2009
Copyright Taylor & Francis Group, LLC
ISSN: 0882-7508 print/1547-7401 online
DOI: 10.1080/08827500802498199

UTILIZATION OF ALUMINA RED MUD FOR


SYNTHESIS OF INORGANIC POLYMERIC
MATERIALS
Downloaded by [141.85.66.12] at 02:24 31 January 2012

DIMITRIOS D. DIMAS, IOANNA P. GIANNOPOULOU,


AND DIMITRIOS PANIAS
National Technical University of Athens, School of
Mining and Metallurgical Engineering, Zografou,
Athens, Greece

Red mud is a residue coming from the metallurgical treatment of


bauxite with the Bayer process. Million of tons of red mud are
produced annually worldwide and disposed of on land, degrading
vast areas. Therefore, red mud utilization is a first-priority issue
for any alumina plant. In the present work, the potential use of
red mud for synthesis of inorganic polymeric materials through
geopolymerization process was studied. The main focus was the
production of inorganic polymeric materials that could be used
in the construction sector as artificial structural elements such as
massive bricks. The geopolymerization process involves a chemical
reaction between red mud and alkali metal silicate solution under
highly alkaline conditions. The product of this reaction is an
amorphous to semi-crystalline polymeric structure, which binds the
individual particles of red mud transforming the initial granular
material to a compact and strong one. The effect of main synthesis
parameterslike solid-to-liquid ratio, caustic soda as well as
soluble silica concentrations, and metakaolin additionon the
properties of red mud-based inorganic polymeric materials was
investigated. The results showed that the produced materials have
high compressive strength, very low water absorption, satisfactory
apparent density, and excellent fire resistance. Therefore, this work

Address correspondence to Dimitrios Panias, National Technical University of


Athens, School of Mining and Metallurgical Engineering, Laboratory of Metallurgy,
Zografou, Athens 15780, Greece. E-mail: panias@metal.ntua.gr
212 D. D. DIMAS ET AL.

proved that the red mud-based inorganic polymeric materials have


promising properties and have the potential to be used as artificial
structural elements in the construction sector.

Keywords: geopolymers, inorganic polymers, red mud

1. INTRODUCTION
Red mud is a solid residue resulting from the metallurgical treatment of
bauxite with the Bayer process in order to produce alumina. Alumina
Downloaded by [141.85.66.12] at 02:24 31 January 2012

red mud is generated in large quantities worldwide. Producing 1 ton


of alumina requires 1.52 tons of bauxite. This means that about
0.7 tons of red mud is generated per ton of produced alumina with
the Bayer process. Taking into account that the world metallurgical
alumina production in 2005 was 51,627,000 tons (Mineral Commodities
Production and Trade Statistics 2005), it is estimated that the world
production of red mud is roughly 35,000,000 tons per year in dry
basis. Although intense research work on utilization of red mud was
performed during the previous decades (Glanville and Winnipeg 1991;
Singh and Prasad 1996; Singh et al. 1997; Marabini et al. 1998; Yalcin
and Sevinc 2000; Ayres et al. 2001; Sagoe-Crentsil and Brown 2005;
Cundi et al. 2005), there is not a generally accepted technology that
could be industrially applied. Therefore, currently alumina red mud
remains a nonexploitable solid residue, and industries must dispose of
it directly into the environment. In the past, red mud was disposed
of either directly into the sea (a solution preferred by France, Greece,
Japan, and the United States) or straight onto the land, creating huge
ponds. At present, sea disposal is ineligible because of strict environ-
mental constraints worldwide, thus land disposal becomes the only alter-
native for the management of red mud. Land disposal increases alumina
producers costs (Ayres et al. 2001) which, combined with high pressure
exerted by environmentalists and environmental policies worldwide, is
the driving force for development of new technologies utilizing red
mud as a raw material for manufacturing high added-value products.
The most important feature of all these technologies is the complete
utilization of red mud for the production of marketable materials for
the construction sector, which uses billion of tons annually worldwide.
Special cements (Singh and Prasad 1996; Singh et al. 1997), bricks and
tiles (Glanville and Winnipeg 1991), ceramic glasses (Marabini et al.
UTILIZATION OF ALUMINA RED MUD 213

1998; Yalcin and Sevinc 2000) have been developed from alumina red
mud during the last 2 decades, but none of these materials has been
introduced efficiently in the construction market as alternatives for
ceramic and cement commercial products for economic reasons and due
to insufficient properties. Therefore, the field for development of a new
technology that will utilize alumina red mud efficiently in producing
construction materials is still open. Towards this end, geopolymerization
technology (Sagoe-Crentsil and Brown 2005; Cundi et al. 2005) seems
to be an attractive one.
Geopolymerization is a very promising innovative technology that
Downloaded by [141.85.66.12] at 02:24 31 January 2012

can transform useless industrial solid wastes of aluminosilicate compo-


sition into useful products competitive with many known construction
materials (Palomo et al. 1999; Cheng 2003; Bortnovsky et al. 2005;
Sagoe-Crentsil and Brown 2005; Cundi et al. 2005; Pacheco-Torgal
et al. 2005; Panias et al. 2007; Maragkos et al. 2009). Geopolymer-
ization includes a chemical heterogeneous reaction between alumi-
nosilicate compounds and alkali silicate solutions under strong basic
conditions and mild temperatures (Davidovits 2005), which leads to
the creation of polymeric SiOAl and SiOSi frameworks. The
creation of these polymeric chains facilitates the conjunction of the
solid granules together, yielding solid compact materials that are called
geopolymers.
Although the hardening of geopolymers appears to be very similar
to the hardening of Portland cement and pozzolans, a big difference
in the mechanism of the two processes has to be mentioned. The
mechanism of Portland cement or pozzolans hardening is based on the
continuous hydration reactions of calcium silicate, which is dissolved in
water directly from cement or formed through the chemical reaction of
pozzolans with calcium hydroxide solutions (Kawatra and Ripke 2002).
The hydration process has as a result the precipitation of an amorphous
calcium silicate hydrate hydrogel with variable composition in terms of
Ca/Si and H2 O/SiO2 ratios, which hardens with time and acts as the
gluing agent in the case of concrete. The hardening of geopolymers is
based on polycondensation reactions of alkali (Na or K) oligo-(sialate-
siloxo) precursors that have been formed from the dissolution of active
silicates and/or aluminosilicate solid raw materials in alkali (Na or K)
hydroxide solutions (Davidovits 2008). The polycondensation process
has as a result the formation of an alkali (Na or K) poly(sialate-siloxo)
cross-linked polymeric network, which hardens rapidly and acts as the
214 D. D. DIMAS ET AL.

gluing agent in the case of geopolymeric concrete. Water molecules are


not reactive species in the case of geopolymerization, and simply create
the necessary environment in which polycondensation phenomena are
taking place. On the contrary, water molecules are reactive species in
the case of Portland cement and pozzolans reactions and therefore are
consumed in hydration reactions.
Geopolymers are inorganic polymeric materials that have very
good physicochemical and mechanical properties (Barbosa et al. 1999;
Barbosa and MacKenzie 2003). They exhibit high mechanical strength,
low apparent porosity or nanoporosity, high Mohs hardness, low
thermal expansion at high temperatures (up to 800 C), good resis-
Downloaded by [141.85.66.12] at 02:24 31 January 2012

tance under freezing-thawing cycles, and excellent stability in acidic


and alkaline environments. Due to their properties, geopolymers can be
used in a broad field of applications and especially in the construction
sector. A variety of industrial minerals (Barbosa et al. 1999; Xu 2002;
Barbosa and MacKenzie 2003; Davidovits 2005)such as kaolinite and
feldsparsand industrial solid residues or wastes (Palomo et al. 1999;
Cheng et al. 2003; Bortnovsky et al. 2005; Sagoe-Crentsil and Brown
2005; Cundi et al. 2005; Pacheco-Torgal et al. 2005; Panias et al. 2007;
Maragkos et al. 2009)such as fly ashes, metallurgical slags, and mine
wasteshave been used for the synthesis of geopolymers. However,
alumina red mud has not been extensively studied yet as an aluminosil-
icate raw material for the synthesis of inorganic polymers through the
geopolymerization technology.
The present work investigates the potential utilization of alumina
red mud as a raw material for the production of geopolymers that
could be used in the construction sector as artificial structural elements,
such as massive bricks. The sample of red mud that was used in this
work comes from the metallurgical plant of Aluminium of Greece S.A.
that treats Greek bauxites with the Bayer process. The main param-
eters affecting the geopolymerization of red mud were studied in detail
and their effects on the compressive strength of synthesized inorganic
polymeric materials were determined. The physicochemical properties
of produced materials under the optimum conditions were determined
and compared with those of the commercial materials. The results
showed that red mud can be geopolymerized and that the synthe-
sized inorganic polymers had properties comparable with those of the
commercial materials.
UTILIZATION OF ALUMINA RED MUD 215

2. EXPERIMENTAL
2.1. Materials
The materials used for the synthesis of inorganic polymers are divided
in two categories; solid aluminosilicate materials and strong alkaline
aqueous phase.

2.1.1. Solid Aluminosilicate Materials. The solid aluminosilicate


materials used in this work were alumina red mud and metakaolin.
The sample of red mud was originated from the metallurgical plant
of Aluminium of Greece S.A. and is a residue generating during the
Downloaded by [141.85.66.12] at 02:24 31 January 2012

treatment of Greek bauxites with the Bayer process. The chemical


analysis of red mud in dry basis is given in Table 1. The main
chemical components of red mud are iron, aluminum, calcium, titanium,
and silicon, while sodium, potassium, and magnesium are minor
constituents. Therefore, alumina red mud can be considered as an iron-
rich aluminosilicate raw material. The mineralogical analysis of red
mud, that was determined by x-ray diffractometry (XRD) on a Siemens
D5000 diffractometer using Cu K radiation ( = 15418 ) operating
at 40 kV, 30 mA, at a 2 range from 2 to 80 with a 0.02 sec1 step,
is presented in Figure 1. Iron occurs in red mud mainly as hematite
(Fe2 O3 ) and secondarily as goethite (FeOOH). The main mineralogical
phase of aluminum is diaspore (AlOOH), while aluminum also occurs as
gibbsite [Al(OH)3 ], cancrinite [Na6 Ca2 (Al6 Si6 O24 )(CO3 )2 2H2 O], and
katoite [Ca3 Al2 (SiO4 )(OH)8 ]. The presence of diaspore in red mud

Table 1. Chemical analysis (% ww) of red mud and


metakaolin samples

Oxide Red mud Metakaolin

Fe2 O3 4143 184


Al2 O3 2061 4098
CaO 892 018
SiO2 736 5266
TiO2 1028 142
Na2 O 043 056
MgO 025 034
K2 O 005 118
L.O.I. 991
Total 9924 9916
216 D. D. DIMAS ET AL.
Downloaded by [141.85.66.12] at 02:24 31 January 2012

Figure 1. XRD diagrams of red mud and metakaolin samples. H, hematite; F, goethite;
D, diaspore; G, gibbsite; A, cancrinite; K, katoite; C, calcite; I, illite; Q, quartz; N,
anatase; R, ferrocarpholite; L, clinochlore.

indicates the fraction of bauxite that was not dissolved during caustic
leaching under high temperatures in autoclaves. Gibbsite represents
the small amount of aluminum that was precipitated during pregnant
solution/red mud separation, while cancrinite and katoite are the
common disilication products during bauxite digestion. Silicon exists
in red mud mainly as cancrinite and katoite. Finally, the main miner-
alogical phases of calcium were calcite (CaCO3 ) and katoite. Titanium
was not detected in XRD diagrams because the main peaks of rutile
(TiO2 ) are hidden by those of hematite and cancrinite. The BET specific
surface area of red mud, as determined by nitrogen absorption in a
Quantachrome Nova 1200 instrument, was 3.02 m2 /g and the mean
particle size (d50 ) was 0.49 m, measured in a Malvern Laser Particle
Analyzer. Finally, the density of the red mud measured with a water
pycnometer according to the ASTM D854-06 standard was 2520 kg/m3 .
The sample of metakaolin (LG Metakaolin AGSBPM/White)
used in this work was a commercial product supplied by Degussa
and its chemical analysis in dry basis is also given in Table 1.
Metakaolin is the dehydroxylation product of the industrial mineral
kaolin in the temperature range 8001000 C. The thermal dehydrox-
ylation process of kaolin increases its solubility in alkaline media,
making metakaolin an excellent raw material for production of inorganic
polymers through the geopolymerization process. Metakaolin is an
aluminosilicate material rich in both silicon and aluminum oxides.
UTILIZATION OF ALUMINA RED MUD 217

Alkalis (Na, K), alkaline earths (Ca, Mg), iron, and titanium are among
the most important minor constituents. XRD analysis (Figure 1) showed
that metakaolin consists of an amorphous silicate and/or aluminosil-
icate phase indicated by the broad band registered between 2 = 17
and 2 = 30 , as well as crystalline phases such as quartz (SiO2 ), illite
[K05 (Al, Fe, Mg)3 (Si, Al)4 O10 (OH)2 ], anatase (TiO2 ), ferrocarpholite
[(Fe, Mg)Al2 Si2 O6 (OH)4 ], and clinochlore [(Mg, Al)6 (Si, Al)4 O10 (OH)8 ].
The mean particle size (d50 ) of metakaolin was 6.09 m and had a
density of 2454 kg/m3 , which is directly comparable with that of alumina
red mud.
Downloaded by [141.85.66.12] at 02:24 31 January 2012

The dissolution of aluminosilicate phases from alumina red mud


and metakaolin in strong alkaline environments was determined exper-
imentally through leaching tests. The tests were performed in sodium
hydroxide solutions with concentration varied from 0.5 M to 7 M, under
ambient temperature (25 C), mild agitation keeping solids in suspension
and pulp density 1% (w/v) so that a big excess in leaching agent
is achieved. The duration of tests was 24 h. The results are shown
in Figure 2, where the % dissolution of silicon and aluminum from
red mud and metakaolin is plotted versus the caustic concentration
in solution. It was clearly observed that the metakaolin solubility was
substantially higher than that of alumina red mud. This was expected
because metakaolin had a substantial amorphous aluminosilicate phase,
which in general is easily dissolved in sodium hydroxide solutions.

Figure 2. Solubility of alumina red mud (RM) and metakaolin (MK) in sodium hydroxide
solutions at 25 C. (Solid:liquid = 10 g/L, 25 C, 24 h.)
218 D. D. DIMAS ET AL.

In case of metakaolin, silicon and aluminum showed the same disso-


lution trend, which was expected because both metals occur in almost
the same crystalline and amorphous mineralogical phases. On the
contrary, substantial differentiation in dissolution behavior of silicon
and aluminum was observed in the case of alumina red mud. Silicon
dissolution was always greater than for aluminum, which seemed to
be invariant with respect to the caustic concentration in solution. It
was concluded that although red mud is an aluminum-rich alumi-
nosilicate material, it cannot act as a good source of soluble silicon
and aluminum during the geopolymerization process according to the
Downloaded by [141.85.66.12] at 02:24 31 January 2012

performed leaching tests. Therefore, it was decided to use a mixture of


red mud with metakaolin, which was proven as an efficient source of
soluble silicon and aluminum during the geopolymerization process, for
the synthesis of inorganic polymeric materials. Therefore, alumina red
mud in this mixture acted primarily as filler and secondarily as a minor
source of soluble silicon and aluminum.

2.1.2. Strong Alkaline Aqueous Phase. Geopolymerization includes


a chemical heterogeneous reaction between a solid aluminosilicate raw
material and a strong alkaline aqueous phase, which is called the
activator. There is a dual role for the activator: On the one hand, it
transfers soluble silicon and aluminum from the solid aluminosilicate
raw material into the aqueous phase through dissolution; on the other
hand, it accelerates the polymerization process in the aqueous phase.
The efficient dissolution of silicon and aluminum from the solid raw
material is facilitated by a strong alkaline solution, as shown by the
chemical equation (1):

(SiO2 , Al2 O3 ) + 2NaOH + 5H2 O Si(OH)4 + 2Al(OH) +


4 + 2Na (1)

In general, the chemical dissolution of aluminosilicate materials is


favored in the range of high pH values, given that the dissolution rate
of these materials increases significantly as the solution pH is increased
(Stumm 1990; Xu and Van Deventer 2000).
As the Si and Al content of the aqueous phase increases due to
dissolution, the formation of the geopolymer precursors takes place.
The precursors are oligomer species consisting of polymeric bonds of
SiOSi and SiOAl type, as described by chemical equations (2)(4).

Si(OH)4 + Si(OH)4  (OH)3 SiOSi(OH)3 + H2 O (2)


UTILIZATION OF ALUMINA RED MUD 219

Si(OH)4 + Al(OH)
4  (OH)3 SiOAl
()
(OH)3 + H2 O (3)
2Si(OH)4 + Al(OH)
4  (OH)3 SiOAl
()
(OH)2 OSi(OH)3 + 2H2 O
(4)

The above reactions are accelerated by the existence of soluble silicates


in the alkaline aqueous phase. Soluble silicates in the aqueous phase
increase essentially the concentration of Si, shifting mainly equation (2)
to the direction of SiOSi species formation, as well as equations (3)
and (4) to the direction of SiOAl oligomer formation.
Hence, the activator must consist of a strong alkaline solution
Downloaded by [141.85.66.12] at 02:24 31 January 2012

containing soluble silicates. For this reason, the activator was prepared
by dissolving sodium hydroxide pellets (Merck, 99.5% purity) in
deionized water in order to control alkalinity and adding water
glass in the form of a sodium silicate solution (Merck, Na2 O = 8%,
SiO2 = 27%, H2 O = 65%, d = 1346 g/L) to control the initial silicon
concentration.

2.2. Experimental Procedure


The preparation of the specimens of inorganic polymeric materials
followed a 3-stage procedure. Initially, a viscous paste was prepared
by mixing mechanically the solid phase, consisting of red mud and
metakaolin with the activator solution for 5 minutes until a homoge-
neous mixture was obtained. In the second stage, the paste was molded
in plastic cubic molds (50 50 50 mm), which were firmly closed by
a plastic cap. At the last stage, the closed molds were cured at 60 C
for 6 hours in order to allow the paste to set, then the specimens were
demolded and their curing procedure was continued at 60 C and relative
humidity 70% for 66 hours. After this period of time, the specimens were
left 4 additional days at ambient temperature. Afterwards, the properties
of the materials were measured by the performance of the relevant
characterization tests.

2.3. Structure Analysis and Characterization Tests


The structure of the produced inorganic polymeric materials was studied
by means of XRD on a Siemens D5000 diffractometer, scanning
electron microscopy (SEM) on a JEOL JSM-type scanning electron
220 D. D. DIMAS ET AL.

microscope, and Fourier-transform infrared spectroscopy (FTIR) on a


Perkin-Elmer 880 FTIR spectrometer.
The following characterization tests were performed in order to
determine the properties of the produced materials. The uniaxial
compressive strength was measured on a testing machine from the Struc-
tural Behavior Engineering Laboratories Inc. (PTL-10 model) using
cubic specimens with dimensions 50 50 50 mm. Two samples were
tested and their average compressive strength was reported. The flexural
strength was determined on an ELE International Limited 50 KN Max
Load testing machine, using rectangular specimens with dimensions
Downloaded by [141.85.66.12] at 02:24 31 January 2012

300 150 10 mm. The cold water absorption and water permeability
were determined according to the EN 771-1: 2003 and ASTM C 1167-03
standards, respectively. The resistance to freezing-thawing cycles was
determined according to the ASTM C 6703a standard. The thermal
behavior of the materials was determined through the following tests:

a. Pyrosis at high temperatures (400 C and 1000 C). Cubic specimens


with dimensions 50 50 50 mm were suddenly inserted into a kiln,
where the temperature was kept constant at 400 C or 1000 C, respec-
tively, and remained under that temperature for 1 hour.
b. Exposure to high-temperature flame. A round disc with 210 mm
diameter and 10 mm thickness was placed horizontally on a metallic
tripod and its front-side surface was exposed to a propane flame with
a temperature >1100 C for 75 minutes. The front-side and reverse-
side temperatures were measured every 5 minutes through a laser,
high-performance infrared thermometer (RAYTEK, Raynger MX4).

3. RESULTS AND DISCUSSION


The synthesis of inorganic polymeric materials took place through the
mixing of the two basic system components: the solid aluminosilicate
phase and the aqueous activator. Therefore, the first parameter affecting
the formation of the inorganic polymers is the solid-to-liquid ratio (S/L).
Provided that the activator is a strong alkaline aqueous phase consisting
of sodium hydroxide solution containing soluble silica, it implies that the
initial sodium hydroxide and silica concentrations were two additional
parameters affecting the synthesis of inorganic polymeric materials.
Taking into account that the solid aluminosilicate phase in this work was
a mixture of red mud and metakaolin, it can be concluded that the last
UTILIZATION OF ALUMINA RED MUD 221

parameter affecting the properties of the inorganic polymeric materials


was the red mud-to-metakaolin ratio in the used solid phase. The effect
of all the above parameters was studied through four independent sets
of experiments.

3.1. Effect of Synthesis Parameters on Compressive Strength


3.1.1. Effect of Solid to Liquid Ratio. The effect of S/L ratio on the
synthesis of inorganic polymeric materials was studied in the region
from 2.0 to 3.1 g/mL. The concentrations of NaOH and SiO2 in the
aqueous phase were kept constant at 3 M, respectively. The solid phase
Downloaded by [141.85.66.12] at 02:24 31 January 2012

was composed from a mixture of red mud and metakaolin with a mass
ratio of 85:15. The results are shown in Figure 3, where the compressive
strength of the specimens is plotted versus the S/L ratio. The results
showed a linear increase of the compressive strength as the function
of S/L in the studied region between 2.03.1 g/mL. The compressive
strength at S/L ratio 2.0 g/mL was 3.80 MPa, while at 3.1 g/mL it was
9.54 MPa, indicating an increase of almost 2.5-fold. However, even if
the decreasing liquid phase content of the paste resulted in higher
compressive strength, it was not feasible to use pastes with an S/L ratio
>3.1 g/mL, as the pastes inadequacy rendered the molding procedure
extremely difficult and inefficient. At the lowest S/L ratios, 2.0 and

Figure 3. Effect of solid-to-liquid ratio on the compressive strength of produced


inorganic polymeric materials. (Activator composition: [NaOH] = 3 M, [SiO2 ] = 3 M.
Solid phase: Red mud-to-metakaolin mass ratio, 85:15. Curing temperature: 60 C.)
222 D. D. DIMAS ET AL.

2.4 g/mL, the aqueous phase was at a high level, therefore the hydration
of solid grains was excellent, resulting in highly efficient pastes. On the
contrary, the excess water was removed easily during the first stages
of the curing procedure, creating visible cracks on the surface of the
specimens. As the S/L ratio increased, the excess of the aqueous phase
decreased, resulting in more viscous, ineffective pastes that became
inadequate at S/L ratios >3.1 g/mL. At the same time, the surface
cracks gradually decreased and at S/L ratios >2.8 g/mL were not easily
visible.
The results showed that an optimum region for the S/L ratio
Downloaded by [141.85.66.12] at 02:24 31 January 2012

has to be defined so that enough wetting of solid grains is achieved


without significant water excess. This region was determined to be
between 2.8 and 3.0 g/mL in which the paste had adequate worka-
bility, ensuring good molding, and the produced materials had very low
linear shrinkage during curing, as is seen in Figure 4, ensuring negligible
surface cracking.
The increased compressive strengths under the higher S/L values
were attributed to two reasons. As the S/L ratio increases, the number
of red mud grains, which work mainly as a filler, per volume of produced
material increases and the amount of polymeric binder decreases. These
changes were reflected on the apparent density of produced materials,
which was substantially increased from 1500 kg/m3 to 1840 kg/m3 in the
S/L ratio region between 2.0 and 3.1 g/mL, as is shown in Figure 5.

Figure 4. Percent linear shrinkage of specimens versus solid-to-liquid ratio during


synthesis of inorganic polymeric materials. (Activator composition: [NaOH] = 3 M,
[SiO2 ] = 3 M. Solid phase: Red mud-to-metakaolin mass ratio, 85:15. Curing temper-
ature: 60 C.)
UTILIZATION OF ALUMINA RED MUD 223
Downloaded by [141.85.66.12] at 02:24 31 January 2012

Figure 5. Apparent density of materials versus solid-to-liquid ratio during synthesis of


inorganic polymeric materials. (Activator composition: [NaOH] = 3 M, [SiO2 ] = 3 M.
Solid phase: Red mud-to-metakaolin mass ratio, 85:15. Curing temperature: 60 C.)

In addition, it was proved from the experiments that the polymeric


binder had lower strength in relation to the red mud grains. For this
reason, the materials failed during compressive strength tests due to
polymeric binder cracking, as is clearly seen in Figure 6. Taking into
account the above experimental observation and assuming that the
strength of materials is a combination of the strength of their individual
componentsthe red mud grains and the polymeric binderit can
be concluded that an increased S/L ratio causes an increase to the
compressive strength of the materials. Additionally, an increased S/L

Figure 6. SEM photo indicating the cracks occurred in the inorganic polymeric binder
at the failure surface during compressive strength tests.
224 D. D. DIMAS ET AL.

ratio causes an increase in metakaolin per volume of the activator


solution. Provided that metakaolin is the main source of easily soluble
silicon and aluminum, as is seen in Figure 2, the concentration of
both metals in the aqueous phase increases substantially after mixing
the solids with the activator as the S/L ratio increases. This in turn
increases the precursors concentration according to reactions (2)(4)
and therefore accelerates the polycondensation process, resulting in a
polymeric binder with higher strength and thus in inorganic polymeric
materials with higher compressive strength.
Downloaded by [141.85.66.12] at 02:24 31 January 2012

3.1.2. Effect of Initial NaOH Content. The effect of initial NaOH


concentration in the activator phase on the compressive strength of
inorganic polymeric materials was studied in the region from 3 to
10 M. The SiO2 content of the activator phase was kept constant at
3 M while the red mud-to-metakaolin mass ratio of the solid phase
was also kept constant at 85:15. Finally, the S/L ratio of 2.9 g/mL
was selected optimal from the results of the first experimental series.
The results are shown in Figure 7, where the compressive strength of
the produced materials is plotted versus the initial concentration of
NaOH in the activator solution. Three regions were clearly observed in
Figure 7. In the first region, which is located between 3 and 6 M initial
NaOH concentration, the materials compressive strength was almost
invariable at 7.7 MPa. As the NaOH concentration increased from

Figure 7. Compressive strength versus NaOH concentration. (Activator composition:


[SiO2 ] = 3M. Solid phase: Red mud-to-metakaolin mass ratio 85:15. S/L = 2.9 g/mL.
Curing temperature: 60 C.)
UTILIZATION OF ALUMINA RED MUD 225

6 M to 8 M, the compressive strength of materials increased substan-


tially, reaching the maximum value of 14.64 MPa. Then, an increase
of the NaOH concentration from 8 M to 10 M caused substantial
reduction of the compressive strength of the materials. The results
revealed a nonmonotonous effect of NaOH concentration on the
materials compressive strength. An explanation can be given based on
the polycondensation phenomena taking place in the aqueous phase.
Generally speaking, the increase of initial NaOH concentration results
in the increase of silicon and aluminium dissolution from the solid phase
(Figure 2), which normally promotes the polycondensation phenomena
Downloaded by [141.85.66.12] at 02:24 31 January 2012

due to the increase of the SiO2 /Na2 O mass ratios in the aqueous
phase, as seen in Figure 8. On the other hand, the silicon and
aluminum dissolution from the solid phase normally reaches a plateau
under very high initial NaOH concentrations. This affects negatively
the polycondensation phenomena because the amount of dissolved
silicon and aluminum remains almost constant while the free NaOH
increases, resulting in lower SiO2 /Na2 O mass ratios in the aqueous
phase. Therefore, the polycondensation phenomena are optimized under
intermediate NaOH concentration, which in this case seems to happen
at 8 M initial NaOH concentration.

Figure 8. Qualitative interpretation of soluble silicate species in aqueous silicate solution.


226 D. D. DIMAS ET AL.

3.1.3. Effect of Initial Silica Content. The effect of initial soluble silica
concentration in the activator phase on the compressive strength of
inorganic polymeric materials was studied in the region from 3 M to
5.5 M. The NaOH content of the activator phase was kept constant
at 8 M, which was the optimum value obtained from the previous
experimental series. The S/L ratio was 2.9 g/mL and the red mud-to-
metakaolin mass ratio of the solid phase was constant at 85:15 during
all experiments of this series. The results are shown in Figure 9, where
it can be clearly seen that an increase of soluble silica concentration
from 3 M to 3.5 M caused a substantial increase in the compressive
Downloaded by [141.85.66.12] at 02:24 31 January 2012

strength of materials from almost 15 MPa to 20 MPa. A further increase


of soluble silica concentration did not appear to seriously affect the
achieved compressive strength. Principally, the soluble silica provides
the liquid phase with the necessary monosilicate species, Si(OH)o4 , which
are responsible for the formation of the oligomers (reactions 2, 3,
and 4) and accordingly for their polycondensation. The higher the initial
silica concentration, the higher the oligomer precursors concentration
(Figure 8) and therefore the higher the polycondensation rate as well as
the compressive strength of the materials produced. This was observed
only in the initial region 33.5 M SiO2 in the aqueous phase. Under
higher initial silica concentrations, cracks were formed on the surface
of materials immediately after their demolding, which grew bigger as the
initial silica concentration increased (Figure 10). In spite of the surface
defects, the materials had very good mechanical properties, indicating in

Figure 9. Compressive strength versus initial silica concentration. (Activator compo-


sition: [NaOH] = 8. Solid phase: Red mud-to-metakaolin mass ratio, 85:15. S/L =
2.9 g/mL. Curing temperature: 60 C.)
UTILIZATION OF ALUMINA RED MUD 227
Downloaded by [141.85.66.12] at 02:24 31 January 2012

Figure 10. Cracks on the surface of inorganic polymeric materials.

this way a very good degree of polymerization. Probably, the materials


might have had better mechanical properties in the absence of surface
cracks. A reliable explanation for the surface defects does not exist. It is
highly probable that the high polycondensation rate might have formed
an impermeable surface membrane that entrapped the free water of the
aqueous phase. Under curing conditions, the expansion of the water
vapor pressure broke the membrane, creating the surface cracks and
thus affecting the mechanical properties of the materials.

3.1.4. Effect of Metakaolin. This experimental series studied the effect


of the substitution of a part of red mud with metakaolin on the
compressive strength of produced materials. The aqueous phase was
composed of 8 M sodium hydroxide and 3.5 M soluble SiO2 , the S/L
ratio was 2.9 g/mL, and the metakaolin content of the solid phase
varied from 0 to 20%. The results are shown in Figure 11, where an
almost linear dependence of the compressive strength on the metakaolin
percentage of the solid phase in the region 015% was revealed. Higher
substitution of red mud with metakaolin gave no better mechanical
properties to the produced materials under the studied conditions. The
228 D. D. DIMAS ET AL.
Downloaded by [141.85.66.12] at 02:24 31 January 2012

Figure 11. Compressive strength versus percentage of metakaolin in the solid phase.
(Activator composition: [NaOH] = 8 M, [SiO2 ] = 3.5 M. S/L = 2.9 g/mL. Curing temper-
ature: 60 C.)

substitution of red mud, a slightly soluble material, with metakaolin,


a highly soluble aluminosilicate material in alkaline media, caused an
increase of the silicon and aluminum content in the aqueous phase,
and thus made the polycondensation process more efficient and the
produced materials stronger. Moreover, the addition of aluminum in
the aqueous phase due to dissolution of metakaolin improved the
cross-linking between the polysilicate chains and sheets producing
materials with higher compressive strengths, as is seen in Figure 11. The
compressive strength of materials was quadrupled when 15% of red mud
was substituted with metakaolin in the solid phase. A larger substitution
did not improve the mechanical properties probably due to the lack of
sufficient amount of NaOH for the dissolution of the extra amount of
metakaolin.

3.2. Structure of Inorganic Polymeric Materials


The changes in the mineralogical phases of red mud and metakaolin,
which occurred during the geopolymerization process, are shown in
Figure 12, in which the XRD diagrams of initial solid materials are
compared with those of the produced inorganic polymers under several
experimental conditions. It can be clearly observed that all miner-
alogical phases existing initially in red mud have not undergone any
substantial change during the geopolymerization process, indicating the
UTILIZATION OF ALUMINA RED MUD 229
Downloaded by [141.85.66.12] at 02:24 31 January 2012

Figure 12. Comparison of XRD diagrams of initial solid materials (red mud and
metakaolin) and produced inorganic polymers under several experimental conditions.

negligible solubility of red mud in caustic solutions. On the other hand,


metakaolin has undergone substantial changes during the geopolymer-
ization process. As seen in Figure 12, the most important crystalline
phase of metakaolinnamely the illitehas totally dissolved while the
broad band registered between 2 = 17 and 2 = 30 has disappeared,
indicating the dissolution of the metakaolin amorphous aluminosil-
icate phase. The results revealed the role of the solid materials during
geopolymerization. Metakaolin played the role of aluminum and silicon
donor during the initial strong alkaline aqueous phase. Red mud had
principally the role of filler, increasing in this way the final compressive
strength of the inorganic polymeric materials. In addition to the afore-
mentioned composition changes, a new peak is observed around 2 =
18 , as is seen in Figure 12. This peak does not exist in the initial red
mud and metakaolin, but it appeared in almost every inorganic polymer
produced. This peak was attributed to a zeolite-type phase that was
formed during geopolymerization as the result of the polycondensation
phenomena in the alkaline aqueous aluminosilicate phase, and is direct
evidence for the formation of an inorganic binder that conglomerates
the insoluble particles of red mud.
230 D. D. DIMAS ET AL.

The above-stated conclusions were strengthened by the results


obtained with the examination of the initial solid materials (red mud
and metakaolin) as well as the produced inorganic polymers with FTIR
spectroscopy. The results are presented in Figure 13, in which a direct
comparison of FTIR spectra of red mud, metakaolin, and synthe-
sized inorganic polymer is shown. The FTIR spectra are divided into
three regions. The first is located at wavenumbers bigger >1600 cm1
and the bands appeared in it were assigned to stretching (OH) and
bending (HOH) vibrations of bound water molecules (Panias et al.
2007). All spectra in this region showed a lot of similarities. The
Downloaded by [141.85.66.12] at 02:24 31 January 2012

second region is between 1300 and 1600 cm1 , where red mud and
inorganic polymer had an absorption band at 14101430 cm1 assigned
to stretching vibrations of OCO bond (Panias et al. 2007), indicating
the presence of carbonates in the materials. The third region, located
at wavenumbers <1300 cm1 , is very important, including absorption
bands assigned to asymmetric stretching of SiOSi (9901090 cm1 ),
symmetric stretching of SiOSi and AlOSi (550 cm1 ), and bending
of SiOSi and OSiO (470480 cm1 ) vibrations (Panias et al. 2007).
From the above bands, the one referring to asymmetric stretching vibra-
tions of SiOSi (9901090 cm1 ) is very informative during geopoly-
merization (Lee and Van Deventer 2002; Fernandez-Jimenez and

Figure 13. Comparison of FTIR spectra of initial solid materials (red mud and
metakaolin) and produced inorganic polymer under optimum experimental conditions.
([NaOH] = 8 M, [SiO2 ] = 3.5 M, S/L = 2.9 g/mL, 60 C.)
UTILIZATION OF ALUMINA RED MUD 231

Palomo 2005; Panias et al. 2007). The shift of this band to lower
wavenumbers indicates the dissolution of the initial solid materials in the
strong alkaline activating solutions and the formation of a new alumi-
nosilicate phase during geopolymerization. As observed in Figure 13,
red mud and produced polymer adsorbed IR radiation exactly at the
same wavenumber (1000 cm1 ), indicating once again the negligible
solubility of red mud in caustic solutions. Initial metakaolin adsorbed IR
radiation at the wavenumber (1087 cm1 ) showing that during geopoly-
merization its amorphous aluminosilicate phase has dissolved in the
activating solutions, forming a new aluminosilicate phase that conglom-
Downloaded by [141.85.66.12] at 02:24 31 January 2012

erates the insoluble particles of red mud, which works principally


as a filler. Finally, the absorption band observed in metakaolin at
862 cm1 attributed to bending vibrations of SiOHhad disappeared
during geopolymerization, indicating the dissolution not only of its
amorphous aluminosilicate phase but also of the crystalline hydroxylated
phases such as illite.
The macrostructure of the synthesized inorganic polymeric
materials is seen in the typical SEM photo in Figure 14a. The materials

Figure 14. Photos of inorganic polymeric materials taken by scanning electron


microscopy. (a) Macrostructure 40 magnification; (b) Microstructure 2000
magnification; (c) Microstructure 3000 magnification; (d) Microstructure 6000
magnification.
232 D. D. DIMAS ET AL.

seem compact and cohesive without discontinuities, confirming their


satisfactory mechanical properties. Pores were observed inside the
materials with diameter substantially <100 m. The vast majority were
closed. In addition, big spherical voids were observed, attributed to
entrapped air bubbles formed during the molding of the materials. The
microstructure is seen in the typical SEM photos in Figure 14bd. The
granular internal structure of the materials can be clearly observed.
The materials consist of solid particles, mainly red mud and secon-
darily nondissolved metakaolin particles <2 m while the majority are
about 0.5 m, which is almost the same as the mean particle size of
Downloaded by [141.85.66.12] at 02:24 31 January 2012

initial red mud grains. The particles are very well conglomerated by a
new amorphous phase, as seen in Figure 14d, taken under high magni-
fication. In this photo, a gelatinous-like phase is observed in which
very fine particles are embodied, <1 m. Additionally, the SEM photos
(Fig. 14b, c) revealed the layered internal structure of materials. The
inorganic polymers seem to have been developed in two-dimensional
structures (layers) that are packed together at a distance <23 m.

3.3. Properties of Inorganic Polymeric Materials


The experimental work described above defined a set of optimum
parameters ([NaOH] = 8M, [SiO2 ] = 35 M; S/L = 2.9 g/mL; red mud
to metakaolin mass ratio 85:15; curing temperature, 60 C) for the
synthesis of the inorganic polymeric materials from a mixture of alumina
red mud and metakaolin. The properties of the produced materials are
shown in Table 2, where they are compared with the specifications for
clay bricks and roof tiles. The red mud-based inorganic polymers are
high-density materials (apparent density 2180 kg/m3 ) that are charac-
terized by relatively high compressive strength (20.5 MPa), very low
cold water absorption (1.28%), and excellent water impermeability
(0 cm3 /cm2 per day) in relation to the relevant specifications for clay
bricks and roof tiles. In addition, the results revealed two important
drawbacks, the inefficient flexural strength and the absence of resistance
in freezing-thawing cycles. The insufficient behavior of the materials
in both tests was attributed to their layered internal structure. As it
is seen in Figure 15, the materials were compact after the end of the
first freezing-thawing cycle but the beginning of their destruction was
obvious. The destruction process had started from the surface layers of
the materials and after the end of the second freezing-thawing cycle the
materials divided in two parts. The one was compact and rigid but the
Downloaded by [141.85.66.12] at 02:24 31 January 2012

Table 2. Properties of inorganic polymeric materials

Measured Specifications Specifications


Property value for clay bricks for clay roof tiles

Compressive strength 20.5 MPa SW > 20.7 MPa


MW > 17.2 MPa
NW > 10.3 MPa
Flexural strength 351 N >600 N
Cold water absorption 1.28% <8% Category 1 < 8%

233
Category 2 < 13%
Category 3 < 15%
Impermeability 0 cm3 /cm2 day Category 1 05 cm3 /cm2 day
Category 2 08 cm3 /cm2 day
Freezing-thawing cycles 2 cycles 50 cycles 50 cycles
Apparent density 2180 kg/m3 Low density 1000 kg/m3
High density > 1000 kg/m3

SW, severe weathering; MW, moderate weathering; NW, negligible weathering.


234 D. D. DIMAS ET AL.
Downloaded by [141.85.66.12] at 02:24 31 January 2012

Figure 15. The materials after the end of the first and second freezing-thawing cycle.

the other that was submerged in the ice consisted of several successive
layers weakly interconnected. The picture of the material after the
end of the second freezing-thawing cycle revealed again the layered
internal structure of the red mud-based inorganic polymers. This mode
of materials destruction can be explained by the absorption of water
during the wetting phase of the freezing-thawing test, which during the
freezing phase becomes ice. As the density of ice is lower than the
density of water, the entrapped ice in between the surface layers exerts
strong forces that break the bonds that cross-linked the several layers.
This could be an indication of weak cross-linking between the layers,
which could also explain the inefficient flexural strength of synthe-
sized inorganic polymers. Taking into account that four-coordinated
aluminum is a well known cross-linking agent in inorganic polymers
(Carraher and Pittman 2005), it can be concluded that the amount
of aluminum dissolved in the aqueous phase during the synthesis of
the materials was not enough to develop 3D structures that could be
stronger than the weak cross-linked 2D structures actually produced.
The red mud inorganic polymers demonstrated excellent thermal
behavior. They exhibited surprisingly high thermal stability when they
were heated suddenly at 400 C and 1000 C. The materials remained
compact and rigid after heating, although some surface cracks developed
due to high percentage mass loss (17% at 400 C and 22% at 1000 C)
attributed to the geopolymeric water and to the thermal decomposition
of gibbsite, diaspore, and calcite of red mud. The linear shrinkage of the
materials after heating was in any case <2%. Substantial mineralogical
UTILIZATION OF ALUMINA RED MUD 235

changes took place during heating related to the recrystallization of


amorphous or semi-crystalline silicate and/or aluminosilicate phases
that were formed during the geopolymerization process. In addition,
the synthesized inorganic polymers exhibited excellent behavior when a
round disc with 210 mm diameter and 10 mm thickness was exposed to
a high-temperature propane flame (>1100 C), as is seen in Figure 16.
In this figure, the front- and reverse-side temperatures of inorganic
polymeric discs are plotted versus the exposure time. The system
reached thermal equilibrium after 20 minutes, establishing a temperature
gradient of 330 C between the front- (760 C) and reverse-side (430 C)
Downloaded by [141.85.66.12] at 02:24 31 January 2012

of the disc, indicating insulating properties. After the end of the test, the
materials were compact and rigid without deformations and changes in
their geometry with the exception of a small shrinkage (<2%). A few
cracks developed inevitably in both sides due to high percentage mass
loss under those temperatures. In Figure 16, the reverse-side temper-
ature of a marketable fire-resistant calcium silicate board with 10 mm
thickness that was exposed to propane flame under the same condi-
tions (Cheng and Chiu 2003) as those of inorganic polymers is plotted
for comparison reasons. The results showed that the red mud inorganic
polymers have similar resistance to high-temperature flame with existing
materials in the market. Moreover, the necessary time to achieve thermal
equilibrium was almost 4 times longer in the case of the red mud
inorganic polymers than in calcium silicate board. This was attributed
to the endothermic nature of inorganic polymers mainly due to the
geopolymeric water they contain. Thus, the red mud inorganic polymers

Figure 16. Exposure of red mud inorganic polymeric materials shaped in round discs
to high-temperature propane flame (Diameter: 210 mm, thickness: 10 mm, flame temper-
ature 1100 C).
236 D. D. DIMAS ET AL.

are excellent fire-resistant materials that could be used in engineering


applications for protection of construction.

4. CONCLUSIONS
The work performed in this paper proved that red mud, a solid residue
produced in millions of tons annually worldwide in the metallurgical
treatment of bauxite with the Bayer process, can be used as raw material
for the synthesis of inorganic polymeric materials with geopolymer-
ization technology.
Geopolymerization is an innovative technology that can transform
Downloaded by [141.85.66.12] at 02:24 31 January 2012

aluminosilicate solid materials (e.g., industrial minerals, useless indus-


trial solid wastes, demolition wastes) into useful products compet-
itive with many known construction materials. Geopolymerization
includes a chemical heterogeneous reaction between solid aluminosil-
icate compounds (solid phase) and alkali silicate solutions (aqueous
phase or activator) under strong basic conditions and mild temperatures,
which leads to the creation of polymeric SiOAl and SiOSi frame-
works. The creation of these polymeric chains facilitates the conjunction
of the solid granules together yielding solid compact materials that are
called geopolymers or inorganic polymers.
The optimum conditions for synthesis that were determined exper-
imentally in this work included the preparation of a viscous paste by
mixing mechanically the solid phase (consisting of 85% red mud and
15% metakaolin) with the aqueous phase (consisting of [SiO2 ] = 35 M,
[NaOH] = 8 M) at S/L ratio 2.9 g/mL until a homogeneous mixture is
obtained. The paste was molded in firmly closed plastic molds and
cured at 60 C for 6 hours in order to harden. Then the specimens were
demolded and their curing procedure was continued at 60 C and relative
humidity 70% for 66 hours.
The red mud-based inorganic polymers that were produced in this
work had relatively high compressive strength (20.5 MPa), very low
cold water absorption (1.28%), and excellent water impermeability
(0 cm3 /cm2 per day), making them comparable or substantially better
than the conventional construction ceramics such as clay bricks and
roofing tiles. Their bulk density (2180 kg/m3 ) was within the specifica-
tions of conventional construction ceramics. In addition, the materials
demonstrated excellent thermal stability at extremely high temperatures
(4001000 C) as well as fire resistance making them proper for use in
engineering applications for fire protection of construction.
UTILIZATION OF ALUMINA RED MUD 237

The two most important drawbacks of these materials were their


inefficient flexural strength and the absence of resistance in freezing-
thawing cycles. Both disadvantages were attributed to the layered
internal structure of the synthesized inorganic polymers. The materials
were composed from 2D structures (layers) that were packed together
at a distance <23 m. The results proved that the layers were cross-
linked weakly, probably due to an insufficient amount of aluminum in
the polymeric phase, therefore the materials failed during the flexural
strength and freezing-thawing tests.
Taking into account these results, it can be concluded that the
Downloaded by [141.85.66.12] at 02:24 31 January 2012

produced inorganic polymeric materials can substitute for ordinary


clay bricks as inner-wall bricks or backing bricks. For the aforemen-
tioned applications, there are no special requirements in regard to
flexural strength and resistance to weathering. The materials at this
stage of development cannot substitute for facing bricks. Although they
have surface and color suitable for facing bricks, they are not able to
withstand frost. The red mud inorganic polymers at this stage of devel-
opment cannot substitute for clay roofing tiles due to inefficient flexural
strength and resistance to weathering. Finally, the synthesized polymers
are promising materials for fire protection of construction due to their
thermal stability and fire resistance.

ACKNOWLEDGMENT
The authors gratefully acknowledge the financial support of the
company Aluminium of Greece for the research project entitled
Feasibility Study for Geopolymerization of Dry Red Mud.

REFERENCES
Ayres, R. U., Holmberg, J., and Andersson, B., 2001, Materials and the Global
Environment: Waste Mining in the 21st Century. Materials Research
Society Bulletin, pp. 477480.
Barbosa, V. F. F. and MacKenzie, K. J. D., 2003, Thermal behaviour of
inorganic geopolymers and composites derived from sodium polysialate.
Materials Research Bulletin, 38, pp. 319331.
Barbosa, V. F. F., Mackenzie, K. J. D., and Thaumaturgo, C., 1999, Synthesis
and characterization of sodium polysialate inorganic polymer based on
alumina and silica. Proceedings of the 2nd International Conference
Geopolymer 99, Saint-Quentin, France, pp. 6578.
238 D. D. DIMAS ET AL.

Bortnovsky, O., Sobalik, Z., Tvaruzkova, Z., Dedecek, J., Roubicek, P.,
Prudkova, Z., and Svoboda, M., 2005, Structure and stability of
geopolymers synthesised from kaolinitic and shale clay resisues.
Proceedings of the World Congress Geopolymer 2005, Saint-Quentin, France,
pp. 8184.
Carraher, C. E. and Pittman, C. U., 2005, Inorganic Polymers. Ullmanns
Encyclopaedia of Industrial Chemistry. Hoboken, NJ: John Wiley & Sons.
Cheng, T. W., 2003, Fire-resistant geopolymer produced by waste serpentine
cutting. Proceedings of the 7th International Symposium on East Asian
Resources Recycling Technology, Tainan, Taiwan, pp. 431434.
Cheng, T. W. and Chiu, J. P., 2003, Fire-resistant geopolymer produced by
granulated blast furnace slag. Minerals Engineering, 16, pp. 205210.
Downloaded by [141.85.66.12] at 02:24 31 January 2012

Cundi, W., Hirano, Y., Terai, T., Vallepu, R., Mikuni, A., and Ikeda, K., 2005,
Preparation of geopolymeric monoliths from red mudPFBC ash fillers at
ambient temperature. Proceedings of the World Congress Geopolymer 2005,
Saint-Quentin, France, pp. 8587.
Davidovits, J., 2005, Geopolymer chemistry and sustainable development.
The poly(sialate) terminology: A very useful and simple model for the
promotion and understanding of green-chemistry. Proceedings of the World
Congress Geopolymer 2005, Saint-Quentin, France, pp. 915.
Davidovits, J., 2008, Geopolymer Chemistry and Applications, 2nd ed. Institut
Gopolymre, Saint-Quentin, France.
Fernandez-Jimenez, A. and Palomo, A., 2005, Mid-infrared spectroscopic
studies of alkali activated fly ash structure. Microporous and Mesoporous
Materials, 86, pp. 207214.
Glanville, J. I. and Winnipeg, P. E., 1991, Bauxite waste bricks. International
Development Research Center Report. Available at: http://www.idrc.ca/
library/document/099941, 2008.
Kawatra, S. K. and Ripke, S. J., 2002, Pelletizing steel mill desulphurization
slag. International Journal of Mineral Processing, 65, pp. 165175.
Lee, W. K. W. and Van Deventer, J. S. J., 2002, Structural reorganization
of class F fly ash in alkaline silicate solutions. Colloids and Surfaces A:
Physicochemical and Engineering Aspects, 211, pp. 4966.
Marabini, A. M., Plescia, P., Maccari, D., Burragato, F., and Pelino, M., 1998,
New materials from industrial and mining wastes: Glass-ceramics and
glass- and rock-wool fibre. International Journal of Mineral Processing, 53,
pp. 121134.
Maragkos, I., Giannopoulou, I. P., and Panias D., 2009, Synthesis
of ferronickel slag based geopolymers. Minerals Enginnering, 22,
pp. 196203.
Mineral Commodities Production and Trade Statistics, Bauxite and
Alumina, Index Mundi. Available at: http://www.indexmundi.com/en/
commodities/minerals/bauxite_and_alumina/
Pacheco-Torgal, F., Castro-Gomes, J. P., and Jalali, S., 2005, Geopolymeric
binder using tungsten mine waste: Preliminary investigation. Proceedings
of the World Congress Geopolymer 2005, Saint-Quentin, France, pp. 9398.
UTILIZATION OF ALUMINA RED MUD 239

Palomo, A., Grutzeck, M. W., and Blanco, M. T., 1999, Alkali-activated


fly ashesa cement for the future. Cement and Concrete Research, 29,
pp. 13231329.
Panias, D., Giannopoulou, I. P., and Perraki, T., 2007, Effect of synthesis
parameters on the mechanical properties of fly ash-based geopolymers.
Colloids and Surfaces A: Physicochemical and Engineering Aspects, 301,
pp. 246254.
Sagoe-Crentsil, K. and Brown, T., 2005, Bayer process waste stream as
potential feedstock material for geopolymer binder systems. Proceedings
of the 7th International Alumina Quality Workshop, Perth, Australia,
pp. 214217.
Singh, M. S. N. and Prasad, P. M., 1996, Preparation of special cements from
Downloaded by [141.85.66.12] at 02:24 31 January 2012

red mud. Waste Management, 16, pp. 665670.


Singh, M. S. N., Upadhayay, S. N., and Prasad, P. M., 1997, Preparation
of iron rich cements using red mud. Cement and Concrete Research, 27,
pp. 10371046.
Stumm, W., 1990, Aquatic Chemical KineticsReaction Rates of Processes in
Natural Waters. New York: John Wiley & Sons.
Xu, H., 2002, Geopolymerization of Aluminosilicate Minerals, PhD Thesis,
Department of Chemical Engineering, University of Melbourne.
Xu, H. and Van Deventer, J. S. J., 2000, The geopolymerization of alumino-
silicate minerals. International Journal of Mineral Processing, 59, pp.
247266.
Yalcin, N. and Sevinc, V., 2000, Utilization of bauxite waste in ceramic
glazes. Ceramics International, 26, pp. 485493.

You might also like