Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/275851435

Synthesis, Characterization, and Mechanical Properties of Red Mud-Based


Geopolymers

Article in Transportation Research Record Journal of the Transportation Research Board · December 2010
DOI: 10.3141/2167-01

CITATIONS READS
154 2,931

3 authors, including:

Guoping Zhang Jian He


University of Massachusetts Amherst C&J Energy Services
109 PUBLICATIONS 3,781 CITATIONS 6 PUBLICATIONS 1,077 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Micro-mechanics of Geo-materials View project

All content following this page was uploaded by Jian He on 04 January 2016.

The user has requested enhancement of the downloaded file.


Synthesis, Characterization,
and Mechanical Properties of
Red Mud–Based Geopolymers
Guoping Zhang, Jian He, and Robert P. Gambrell

A pilot study investigates the potential of reusing red mud, an abundant alkalinity and high water content are the two major environmental
industrial waste produced from alumina refining by the Bayer process, concerns for the safe and economical disposal of red mud. Thus, its
by geopolymerization reactions with another solid waste, fly ash, and treatment and disposal are major difficulties for alumina refineries.
sodium silicate. Parameters involved in the synthesis, including red mud In the past, red mud was disposed of from the plant site mainly in
to fly ash ratio (values of 80/20, 50/50, and 20/80), presence of sand filler, two ways: dumping it (a) directly into the sea or else (b) onto the
curing duration (up to 28 days), and sodium silicate solution to solid land in waste ponds, whose monitoring and maintenance are costly.
mixture (consisting of red mud and fly ash) ratio, were examined to Now, however, stricter environmental regulations are being enforced,
understand the extent and degree of geopolymerization. Unconfined com- and new technologies are urgently needed for the disposal or recycling
pression testing was employed to assess the influence of these synthesis and reuse of red mud. Although extensive research on red mud uti-
parameters on the mechanical properties of the end products, red lization has been conducted in past decades (e.g., 3–9), a widely
mud–based geopolymers. The composition and microstructure were accepted technology that can be employed for the recycling of red
characterized by X-ray diffraction and scanning electron microscopy, mud is not available at present.
respectively, which confirm the geopolymerization reactions. The Recently, a new class of materials, inorganic geopolymers, has been
mechanical properties, including strength, stiffness, and failure strain, studied intensively as a viable economical alternative to organic
were analyzed against the chemical compositions of the red mud geopoly- polymers and inorganic cements in diverse applications, including
mers, such as Si/Al and Na/Si molar ratios. For the studied geopolymers, military (10), aircraft (11, 12), high-tech ceramics (13), thermal
the unconfined compressive strength, ranging from 7 to 13 MPa, increases insulating foams (14), fireproof building materials (15), protective
with the Si/Al ratio as in some types of portland cement. A higher Na/Si coatings (16), refractory adhesives (17), and hybrid inorganic–organic
ratio appears to reduce the strength and stiffness but enhance the ductil- composites (18, 19), because of their exceptionally high compressive
ity. The results indicate that red mud geopolymers are a viable cementi- strength, high thermal and chemical stability, and adhesive behavior.
tious material that can be used in roadway construction. The engineering Geopolymers are inorganic aluminosilicates formed by polycondens-
implications are discussed in terms of waste recycling, environmental ing tetrahedral silica (SiO4) and alumina (AlO4) under highly alkaline
benefits, and energy consumption. conditions into a polymeric structure.

Red mud is the major industrial waste produced by the Bayer process { }
M +n - ( SiO 2 )z -AlO 2 - n
⎡⎣ e.g., ( -Si-O-Al-O- )n , ( -Si-O-Al-O-Si-O- ) n ,
for the extraction of alumina from bauxite ores. Depending on the
quality and purity of the bauxite ore, the quantity of red mud generated
(-Si-O-Al-O-Si-O-Si-O-)n ⎤⎦
varies from 55% to 65% of the processed bauxite (1). According to
a recent U.S. Geological Survey report (2), bauxite ore mined globally where
amounts to 202 million tons (MT) in 2007 and 205 MT in 2008. The M+ = alkali cation (K+ or Na+) for charge balance,
processing of these ores generates a huge amount of red mud as n = degree of polymerization, and
the major waste. Red mud is characterized by strong alkalinity even z = Si/Al molar ratio.
with a high water content (up to 95%), owing to the presence of an
excessive amount of dissolved sodium hydroxide used to extract With varying Si/Al ratios (i.e., z = 1 to 15, up to 300), geopolymers
silicates and alumina. Its solid constituents include mainly iron exhibit different properties (20): low ratios result in three-dimensional,
oxides (mostly hematite), alumina, and some heavy metals. Strong cross-linked, rigid networks with stiff and brittle properties and
thus cementitious and ceramic material; ratios >3 and >15 result in
two-dimensional networks and linearly linked polymeric structures
G. Zhang, 3504 Patrick F. Taylor Hall, and J. He, 210 Old Coastal Studies Building,
with adhesive and rubbery properties, respectively.
Department of Civil and Environmental Engineering, and R. P. Gambrell, 3253
Energy, Coast, and Environment Building, Department of Oceanography and Coastal Geopolymer formation consists of two basic steps (14, 21–24):
Sciences, Louisiana State University, Baton Rouge, LA 70803. Corresponding author: (a) generation of reactive species, or alkali activation, which is the
G. Zhang, gzhang@lsu.edu. dissolution of amorphous phases (e.g., aluminosilicates) by alkali to
produce small reactive silica and alumina, and (b) actual setting reac-
Transportation Research Record: Journal of the Transportation Research Board,
No. 2167, Transportation Research Board of the National Academies, Washington,
tion, which is the polycondensation process leading to the formation
D.C., 2010, pp. 1–9. of amorphous to semicrystalline aluminosilicate polymers. In partic-
DOI: 10.3141/2167-01 ular, the presence of reactive alumina [Al(IV)] and silica [Si(IV)] and

1
2 Transportation Research Record 2167

alkali activation are two key requirements for geopolymerization, as TABLE 1 Chemical Composition and Concentrations of Red Mud
illustrated by the following two reactions (25): and Fly Ash

n ( Si 2 O5 , A12 O 2 ) + 2nSiO 2 + 4 nH 2 O + NaOH KOH Material Red Mud Fly Ash Material Red Mud Fly Ash

(Si-Al material ) SiO2 1.2 47.5 MgO — 4.8

→ Na +, K + + n ( OH )3-Si-O-A1− -O-Si- ( OH )3
Al2O3 14.0 15.3 S — 0.3
(1)
Fe2O3 30.9 4.2 K2O — 0.6

( OH )2 NaOH 20.2 — TiO2 4.5 —

( geopolymer precursor )
NaAlO2 23.0 — MnO 1.7 —
CaO 2.5 24.0 Total 98.0 96.7

n ( OH )3 -Si-O-A1− -O-Si- ( OH )3 + NaOH KOH


| phase (Table 1). The slurry was air-dried, homogenized, and pulver-
( OH )2 ized to a powder until all solids passed a No. 60 sieve to facilitate
geopolymerization and minimize the influence of compositional
| | | variation on the geopolymers. The results of hydrometer analysis
→ ( Na + , K + ) - ( -Si-O-A1− -O-Si-O- ) + 4nH 2 O (2) (ASTM D422) show that the red mud contains 33%, 43%, and 24%
| | | clay, silt, and sand-sized particles, respectively. The fly ash mainly
O O O contains silica, alumina, and gypsum (Table 1). A clean white quartz
| | | sand with particle sizes of 0.25 to 0.425 mm was used as an additional
( geopolymer backbone ) aggregate filler in the geopolymer matrix.

Geopolymers are synthetic mineral products that combine the


properties of polymers, ceramics, and cements (22, 26) and thus Methods
possess a series of distinct properties: (a) they are exceptionally
heat- and fire-resistant (stable up to 1200°C); (b) they resist all Geopolymer synthesis started with dry-mixing the powders at a
organic solvents and acids and are highly corrosion resistant; (c) they selected red mud to fly ash (RM/FA) weight ratio, followed by
are nontoxic, “green” materials, because their production saves energy adding 1.5 M sodium trisilicate solution to the powder mixture
and does not produce CO2 emissions unlike traditional portland (containing both red mud and fly ash) again at a selected weight
cement; (d) they can be made from a wide range of low-cost alumino- ratio (solution/solid). The mixture was then stirred for >15 min to
silicate materials, or even wastes, such as metakaolin, fly ash, furnace ensure sufficient reaction (e.g., dissolution) between the powder
slag, and mine tailings (27, 28); and (e) they do not incorporate and the solution, resulting in the formation of geopolymer pre-
hydration water within crystal structures, unlike Ca-based cements, cursor paste with appropriate consistency. If needed, the aforemen-
and hence are impermeable and resistant to water or moisture and tioned sand was then added as an aggregate filler to the precursor
have more robust mechanical properties (e.g., surface hardness, at a weight ratio (3/10) of dry sand to wet geopolymer precursor
compressive strength) than do cement-based materials. These features (i.e., sand/geopolymer ratio = 3/10). The geopolymer precursor was
make geopolymers a promising material for waste utilization and then poured into cylindrical molds with an inner diameter of 2 cm
new materials development and applications. and height of 5 cm (i.e., an aspect ratio of 2.5 to minimize the end
This paper describes a potentially viable technology for the reuse of effects), followed by curing in an ambient environment for 7 days.
red mud via geopolymerization, which takes advantage of its two The specimens were then demolded, followed by prolonged curing
major characteristics: high alkalinity and the presence of alumina. in exposed conditions. A total of four curing durations (i.e., 7, 14, 21,
Reactive silica (usually a finely sized, amorphous phase) absent in red and 28 days) and three ratios of RM/FA (i.e., 80/20, 50/50, and 20/80)
mud will be provided by another industrial waste, fly ash. Sodium sil- and of solution/solid (e.g., 1.0, 1.2, and 1.5) were examined to
icate, the only constituent that is a nonwaste used for geopolymer syn- investigate the influence of geopolymer synthesis and curing on the
thesis, will provide additional Si and Na for geopolymerization of red mechanical properties of the final geopolymeric products. The pur-
mud. The goal was to investigate the potential utilization of red mud pose of selecting the three RM/FA ratios was to examine the range
as a raw material for the production of geopolymers that can be used of red mud quantity used in geopolymer synthesis so that as much
in roadway construction either as pavement, subgrade, or subbase. red mud waste as possible could be used. Although geopolymeriza-
tion requires alkali activation, no NaOH or KOH was added to the
synthesis, owing to the presence of residual NaOH in red mud.
MATERIALS AND METHODS Unconfined compression tests (ASTM C39 and C39M) were
performed on cured cylindrical specimens by using an automated
Materials GeoTAC loading frame (Trautwein Soil Testing Equipment, Inc.) at
a constant strain rate of 0.5%/min. The two ends of the specimens
The raw materials used for geopolymer synthesis include red mud were polished by sandpaper to obtain flat and parallel surfaces. A thin
slurry (Gramercy Alumina, LLC), Class C fly ash (Bayou Ash, Inc.), layer of lubricant coating was then applied to minimize the friction
sodium trisilicate (Na2O  3SiO2) powder (Sigma-Aldrich Co.), and (and hence shear stress development) between the specimen end
deionized water. The red mud slurry has a pH of 11.9 and contains surfaces and polished stainless steel end platens.
mainly water with dissolved Na-aluminate (NaAlO2) and NaOH Further characterization of the geopolymers was performed to
as liquid phase and hematite (Fe2O3) and alumina (Al2O3) as solid understand the strength–composition–microstructure relationships.
Zhang, He, and Gambrell 3

The compositions of the red mud, fly ash, and geopolymers of varied a poorly defined yield point, and a brittle failure mode. A feature
RM/FA ratios cured for 28 days were characterized by X-ray diffrac- common to all curing durations is readily observed: the higher the
tion (XRD) with a Bruker/Siemens D5000 automated X-ray powder red mud content, the lower the compressive strength (σf ) and stiffness
diffractometer. All XRD scans used Cu K α radiation, a step size of (or Young’s modulus E), and the higher the compressive failure
0.02°, a scan speed of 0.02° per 2 s, and a scan range of 2° to 70° 2θ strain (⑀f ). This suggests that geopolymers with a higher RM/FA
(diffraction angle). All scanned sample powders were obtained by wet ratio tend to be more ductile but also weaker than those with a lower
grinding with alcohol in a McCrone micronizing mill (McCrone ratio. Two reasons may account for this: (a) the higher contents of
Accessories and Components) for 3 min, which usually produces a reactive silica and alumina in fly ash result in a higher degree of geo-
fine powder with particle sizes ≤38 μm for most silicate minerals. The polymerization and more geopolymeric binder and (b) the fine particle
microstructure of the geopolymers was examined with a JEOL 840A size and high specific surface area of red mud make geopolymers
scanning electron microscope (SEM) under a high vacuum of 10−6 Torr more ductile.
and electron beam energy of 20 keV. Small pieces of samples selected To show more clearly the influence of curing on the mechanical
from the failed cylindrical specimens were examined, and the fracture properties, Figure 2 replots the stress–strain curves of a geopolymer
surface was a particular focus of SEM observations. All samples were with the same RM/FA ratio but different curing durations. Clearly,
coated with gold before SEM observation. both the strength and stiffness increase with curing duration, and the
failure strain decreases with curing duration. For the geopolymer
with a RM/FA ratio of 50/50 (Figure 2b), the strength is the same
ANALYSIS OF RESULTS for 21 and 28 days of curing, but extended curing can further increase
stiffness and brittleness. The same phenomenon is also observed for
Influence of RM/FA Ratio and Curing Duration the geopolymer with a RM/FA ratio of 20/80 (Figure 2c) except that
a shorter curing time (i.e., 14 days) is needed to reach the stabilized
Figure 1 shows the axial stress–strain curves of the red mud geo- strength.
polymers synthesized at varied RM/FA ratios cured for different To compare the influence of RM/FA ratio on the curing rate and
durations. All curves exhibit a well-defined linear elastic regime, strength of geopolymers, Figure 3 plots the compressive strength

12 14
RM/FA = 20/80 RM/FA = 20/80
10 12
RM/FA = 50/50 RM/FA = 50/50
Axial stress, σ [MPa]
Axial stress, σ [MPa]

RM/FA = 80/20 10 RM/FA = 80/20


8
8
6
6
4
4
2 2

0 0
0 2 4 6 8 10 0 2 4 6 8
Axial strain, ε [%] Axial strain, ε [%]
(a) (b)

14 14
RM/FA = 20/80 RM/FA = 20/80
12 12 RM/FA = 50/50
RM/FA = 50/50
Axial stress, σ [MPa]

Axial stress, σ [MPa]

10 RM/FA = 80/20 10 RM/FA = 80/20

8 8

6 6

4 4

2 2

0 0
0 1 2 3 4 5 0 0.5 1 1.5 2 2.5 3
Axial strain, ε [%] Axial strain, ε [%]
(c) (d)

FIGURE 1 Strain–stress curves of red mud geopolymers of varied RM/FA ratios cured for (a) 7 days, (b) 14 days, (c) 21 days, and
(d) 28 days.
4 Transportation Research Record 2167

8 12
28 days 28 days
7
21 days 10 21 days

Axial stress, σ [MPa]


Axial stress, σ [MPa]

6 14 days 14 days
8
5 7 days 7 days

4 6
3
4
2
2
1
0 0
0 2 4 6 8 10 0 1 2 3 4 5
Axial strain, ε [%] Axial strain, ε [%]
(a) (b)

14
28 days
12
21 days
Axial stress, σ [MPa]

10 14 days
7 days
8

0
0 1 2 3 4
Axial strain, ε [%]
(c)

FIGURE 2 Influence of curing duration on the stress–strain behavior of red mud geopolymers with RM/FA of (a) 80/20,
(b) 50/50, and (c) 20/80.

14
Compressive strength, σf [MPa]

12

10

4 RM/FA = 20/80
RM/FA = 50/50
2 RM/FA = 80/20

0
7 14 21 28
Curing time, t [day]

FIGURE 3 Influence of curing time and RM/FA ratio on the compressive


strength of geopolymers.
Zhang, He, and Gambrell 5

(σf ) against curing time for all geopolymers. In general, σf increases Influence of Solution/Solid Ratio
initially with curing time and then reaches a constant for even pro-
longed curing. Interestingly, the three curves are nearly parallel to The RM/FA ratio = 50/50 geopolymer was again chosen to assess
each other (at least for the available data), indicating that the rate the influence of the solution/solid ratio on the mechanical behavior
of strength increase is the same for all thee geopolymers. Yet two of red mud geopolymers. Figure 4b compares the stress–strain
remarkable differences exist: (a) the final cured strength increases curves of the geopolymers with three different solution/solid ratios
as the RM/FA ratio decreases and (b) the curing time required to reach (i.e., 1.0, 1.2, and 1.5). The higher the ratio, the stronger the geo-
the final constant strength decreases as the RM/FA ratio decreases. polymer and the higher the failure strain. Moreover, a higher solution/
Moreover, for all three geopolymers, the strength stabilizes after solid ratio leads to a better-defined peak strength (i.e., failure occurs
21 days of curing, which suggests that the red mud geopolymers can more abruptly). Young’s modulus of the three geopolymers remains
achieve complete curing in 21 days. To ensure complete curing and nearly unaltered. A higher solution/solid ratio can generate positive
minimize the undesired influence of incomplete curing on the mechan- effects on the mechanical properties of the red mud geopolymers.
ical properties, all geopolymer samples discussed in the following One reason for such influence may be that a higher solution/solid
sections were cured for at least 28 days before testing and XRD and ratio facilitates the movement of reactive Si and Al and hence pro-
SEM characterization. motes a better extent of geopolymerization of the solid mixtures.
However, a higher ratio also introduces more water and sodium
to the reaction process. As discussed later, the overall influence
Influence of Sand Filler needs to be analyzed more systematically against the composition
(e.g., Si/Al and Na/Si ratios).
The geopolymer with an RM/FA ratio of 50/50 was selected to assess
the influence of sand filler on the mechanical properties. Figure 4a
compares the stress–strain curves of the geopolymers without and XRD Analysis
with sand filler at a sand/geopolymer ratio of 3/10. It appears that
the sand filler (at this given ratio) has detrimental influence on strength Figure 5 shows the XRD patterns of fly ash, red mud, and geopolymers
and stiffness, but little effect on ductility, although the mode of failure with three different RM/FA ratios after >28 days of curing. Fly ash
seems less brittle and abrupt and the postfailure stress decreases shows a broad nonsymmetrical hump between 20° and 36° 2θ,
more slowly. The strength (σf ) is reduced from 10.6 to 3.6 MPa, a indicating the presence of noncrystalline phases such as amor-
nearly two-thirds’ reduction. Although the failure strain (⑀f ) remains phous silica and alumina generated via coal combustion. In fact,
the same, the yield strain (⑀y) is reduced by nearly 50%. Young’s it is those amorphous phases that are actively involved in geopoly-
modulus is also reduced by 50% (i.e., from 0.95 to 0.49 GPa). Two merization reactions. A few sharp peaks also indicate the presence of
reasons may account for the influences: (a) the sand does not react crystalline phases, including quartz, anhydrite (CaSO4), and gypsum
with the geopolymer matrix or simply presents as inactive filler (CaSO4  2H2O), the latter probably transformed from anhydrite via
and (b) the geopolymer already contains a high content of other con- absorption of water. By comparison with its chemical composition
stituents (e.g., hematite, quartz, gypsum) as inactive fillers; adding (Table 1), alumina mainly presents as amorphous phase, whereas
more sand results in excessive concentration of inactive filler but silica may be present as both amorphous and crystalline (quartz)
insufficient geopolymer binder. More work is needed to achieve a phases. Only those amorphous phases are reactive and hence par-
better understanding of the influence of sand filler on the mechanical ticipate in geopolymerization reactions. Red mud shows a few sharp
behavior of red mud geopolymers. peaks—mainly from hematite and calcite—but no observable broad

12 14
Solution/solid = 1.5
No sand filler
12 Solution/solid = 1.2
Sand/
10 Solution/solid = 1.0
geopolymer = 3/10
Axial stress, σ [MPa]
Axial stress, σ [MPa]

10
8
8
6
6
4
4

2 2

0 0
0 0.5 1 1.5 2 2.5 3 0.0 0.5 1.0 1.5 2.0
Axial strain, ε [%] Axial strain, ε [%]
(a) (b)

FIGURE 4 Strain–stress curves of the geopolymer with RM/FA ratio of 50/50 influenced by (a) sand filler and
(b) solution/solid ratio.
6 Transportation Research Record 2167

Geopolymer: RM/FA = 20/80

Geopolymer: RM/FA = 50/50

Intensity [arbitrary unit]


Geopolymer: RM/FA = 80/20

H H
H
Q C H
Red Mud

G
Q An An
Fly ash

0 5 10 15 20 25 30 35 40
2θθ [o]

FIGURE 5 XRD patterns of fly ash, red mud, and geopolymers with different RM/FA
ratios (An ⴝ anhydrite, Q ⴝ quartz, G ⴝ gypsum, H ⴝ hematite, C ⴝ calcite).

humps, suggesting that the amorphous phases are not present in tity of microspheres than those with a lower ratio, whereas the
any large quantity, which is consistent with its chemical compo- opposite was observed for platy films. It is believed that the platy
sition (Table 1). Thus, red mud provides mainly Al (via either films are the geopolymer—a cementitious phase bonding those
amorphous Al2O3 or dissolved NaAlO2) and NaOH but little Si to unreactive phases in the matrix—whereas the microspheres are
geopolymerization. mainly hematite particles. Hematite spheres may be generated by
The patterns of the three geopolymers all show the presence of two processes during alumina refining: (a) alumina extraction,
sharp peaks of crystalline phases from parent materials, both red which involves prolonged and intensive grinding, particle colli-
mud and fly ash, with variable peak intensities that change with sion, and alkali dissolution; and (b) transport of the red mud slurry
the RM/FA ratio, as reflected by the peak intensities from quartz to the disposal pond, which involves high-velocity flow inside
and hematite. This confirms that the crystalline phases are not reac- pipelines, where interparticle collision and grinding take place. The
tive or involved in geopolymerization, but simply present as inactive presence of a large quantity of unreacted or insoluble phases such as
fillers in the geopolymer network. For all three geopolymers, how- hematite, quartz, and gypsum (as determined by XRD) may hin-
ever, a weak, broad hump between 18° and 36° 2θ is observed der the homogeneous distribution and transport of dissolved alu-
(it became less clear because of the large vertical axis scale), which mina [Al(IV)] and silica [Si(IV)] which further polycondense to
is regarded as the characteristic peak of geopolymers (e.g., 21–23). form geopolymer (23).
Furthermore, under the same scan conditions, this hump is more The geopolymer matrix appears to exhibit an inhomogeneous
pronounced for the lower RM/FA ratio geopolymers, which indicates structure at the microscale, as reflected by the presence of ran-
that the geopolymer content is higher. The geopolymer made with domly distributed micropores and microcracks. Figure 6c shows
higher fly ash content possesses higher strength and stiffness, a typical semispherical pore 80 to 100 μm in diameter on the
which is consistent with the mechanical testing results. Duxson fractured surface, surrounded by geopolymer matrix with micro-
et al. (23) also reported that a higher concentration of reactive Si cracks. The pores are likely caused by (a) residual air bubbles
in the geopolymerization process typically leads to higher compressive introduced into the geopolymer precursor through mixing or
strength. trapped inside the geopolymer when pouring into the mold;
and (b) the cavity previously occupied by water that evaporated.
Previous studies also observed the presence of micropores in
Characterization of Microstructure geopolymers, especially those without deairing treatment on geo-
polymer precursors (29). The presence of microcracks is the
Figure 6 presents selected SEM micrographs detailing the micro- manifestation of the disturbing effect of quartz sand on the geo-
structure of red mud geopolymers. The fractured surfaces—but polymer matrix (Figure 6d). As discussed earlier, quartz is not
not external surfaces—were observed to minimize the influence involved in a geopolymer reaction. Differential shrinkage of the
of aerial exposure on the microstructure during curing. Typically, geopolymer precursor and sand particles causes interfacial sepa-
the geopolymer matrix consists of spherical particles 2 to 10 μm ration, resulting in the formation of interfacial microcracks after
in diameter and irregular platy films connecting and surrounding curing. This is likely one of the reasons why the sand filler has a
the microspheres (Figures 6a and 6b). It was also observed that remarkably detrimental influence on the mechanical properties of
geopolymers with a higher RM/FA ratio contain a greater quan- the geopolymer.
Zhang, He, and Gambrell 7

8μm 5μm

(a) (b)

Sand

40μm μm
20μ

(c) (d)

FIGURE 6 SEM micrographs showing the microstructure of geopolymers: (a) and (b) typical
geopolymer matrix consisting of spherical microparticles and glassy flakes, (c) a micropore, and
(d) interfacial cracks between a quartz sand particle and geopolymer matrix.

DISCUSSION OF RESULTS former acts as a binder or cementitious material bonding inactive


fillers. The mechanical properties of red mud geopolymers are affected
Geopolymerization of Red Mud by a number of factors: (a) the chemical composition (e.g., Si/Al and
Na/Si ratios) of the pure geopolymer phase; (b) the characteristics
Three parent materials are involved in the synthesis of red mud–based of inactive granular fillers, such as particle size, shape, and strength;
geopolymers: red mud, fly ash, and sodium trisilicate. The pure (c) the interfacial bonding strength between granular fillers and
geopolymer network actually consists mainly of Si, Al(IV), and O, pure geopolymer binder; (d) the relative concentrations of pure
with alkali Na+ or K+ to balance the charge of Al(IV). On the basis geopolymer and inactive fillers; and (e) the extent and degree of
of the above XRD analysis, not all of the mineral phases in parent geopolymerization. Evaluation of the mechanical properties of the
materials participate in geopolymerization. Hematite and calcite final products is complicated by these factors. Quantitative analysis
from red mud, and quartz, gypsum, and anhydrite (transformed to of the composition and concentration of pure geopolymer in the end
gypsum after synthesis) are present in the final products as unreactive products is challenging because of nonreacting Si (e.g., in quartz)
fillers but not in the neoformed geopolymer structure. Reactive phases and incomplete geopolymerization of even reactive Si and Al.
include (a) NaOH (20.2%), alumina (14.0%), and Na-aluminate An attempt was made, however, to correlate one of the mechanical
(23.0%) from red mud; and (b) silica (47.5%) and alumina (15.3%) properties (i.e., compressive strength, failure strain, and Young’s
from fly ash. Sodium trisilicate contributes mainly Si and Na to modulus) with one of the chemical compositions of the geopolymer
geopolymerization reactions. However, certain silica in fly ash is (including Si/Al and Na/Si molar ratios and concentrations of SiO2,
present as crystalline quartz, as reflected by the XRD patterns, and Al2O3, Na2O, and Fe2O3 in the final products), resulting in a total of
is not involved in geopolymerization. The end product, therefore, is 18 plots. Figure 7 presents four selected correlations with the highest
not strictly a pure geopolymer and hence possesses variable, complex coefficients of determination (i.e., R2). The strength increases almost
mechanical properties. linearly with the Si/Al ratio, but decreases with the Na/Si ratio, in
accord with previous findings that geopolymers with Si/Al ratios <3
act as a cementitious material. Young’s modulus decreases with the
Factors Affecting Mechanical Performance Na/Si ratio, but the failure strain shows the opposite trend. For the
examined range of Si/Al ratio (z = 1.7 to 3.2) and Na/Si ratio of
As discussed above, the final synthesized products actually contain 0.4 to 1.7, higher Na content tends to decrease the strength and
neoformed geopolymer and unreacted phases as inactive fillers. The stiffness but enhance the ductility of geopolymers.
8 Transportation Research Record 2167

Na/Si molar ratio not only the expense of waste disposal and long-term monitoring
0.0 0.5 1.0 1.5 2.0 and maintenance of waste-containment facilities but also the cost of
14 14 manufacturing portland cement. Second, recycling the two major
Compressive strength, σf [MPa]

13 13 industrial wastes can minimize the damaging effects of the waste


on the environment and human health. Third, the hematite in the
12 12
geopolymer is actually highly absorptive for heavy metals, and it
11 y = 303.4x + 17.5 11 can act as a reactive barrier to filter the contaminants transported
Si/Al R 2 = 0.9732
10 Na/Si 10 through surface water runoff. Finally, the elimination of portland
9 9 cement use can save the energy associated with cement production
and reduce CO2 emissions caused by firing carbonates (21).
8 8
y = -1344.5x + 2192.6
7 R 2 = 0.8635 7
6 6 CONCLUSIONS
1 2 3 4
Si/Al molar ratio This paper presents an experimental study that investigates the
(a) potential reuse of red mud, an industrial waste from alumina refining,
via geopolymerization reactions with another solid waste, fly ash, and
2.0 1.2
sodium silicate. A variety of parameters involved in the synthesis,
y = -99,960x + 175,055
including RM/FA ratio, presence of sand filler, and curing duration,
R 2 = 0.9808 1.1 were examined to understand the extent and degree of geopoly-
Young’s modulus, E [GPa]
1.8 merization and their influence on the mechanical properties of red
1
Failure strain, εf [%]

mud–based geopolymers.
1.6 Failure strain 0.9 The results show that the mechanical properties of the geopolymers
Young’s modulus
0.8
are influenced by curing duration, RM/FA ratio, and additional sand
filler. In general, the compressive strength increases with curing but
1.4 0.7 reaches a constant after complete curing. The stiffness appears to
continue to increase even after the strength stabilizes, for up to 28 days
0.6
1.2 y = 0.92x + 1.08 of curing. Additional sand filler appears to have a detrimental influence
R 2 = 0.7482 0.5 on mechanical performance. Empirical correlations also suggest that
the strength increases with the Si/Al molar ratio but decreases with the
1.0 0.4
0.0 0.5 1.0 1.5 2.0 Na/Si ratio. Higher Na/Si ratios also decrease the stiffness and enhance
Na/Si molar ratio the ductility. For the studied range of compositions, the red mud
(b) geopolymers possess a compressive strength of 7 to 13 MPa, as do
certain types of portland cement. This study demonstrates that red
FIGURE 7 The correlations of (a) compressive strength with Si/Al mud geopolymers are a viable cementitious material with promis-
and Na/Si ratios and (b) failure strain and Young’s modulus with ing potential for use in roadway construction, which could bring
Na/Si ratio. economic and environmental benefits.

Engineering Implications ACKNOWLEDGMENTS


for Roadway Construction
Red mud samples were provided by Gramercy Alumina, LLC. The
The 28-day unconfined compressive strength of the red mud geo- authors thank Charles L. Preston and David Kirkpatrick for their
polymers ranges from 7 to 13 MPa (excluding the one with sand filler). technical discussion and help.
The highest strength is like that of the portland Type IIA (9 to 12 MPa)
or IA (16 MPa) cement (30), although the former contains a signifi-
cant amount of impurities as inactive fillers (e.g., hematite, quartz, REFERENCES
calcite, and gypsum). Red mud geopolymers can be used as a suitable
adhesive binder or cementitious material to replace cement in certain 1. Paramguru, R. K., P. C. Rath, and V. N. Misra. Trends in Red Mud
Utilization—A Review. Mineral Processing and Extractive Metallurgy,
engineering applications, such as roadway construction. They can Vol. 26, 2005, pp. 1–29.
be used in top surface pavement, or in subgrade, or they can even be 2. Mineral Commodity Summaries. U.S. Geological Survey, Jan. 2009.
mixed with soil as subbase. The red color, an intrinsic characteris- http://minerals.usgs.gov/minerals/pubs/mcs/2009/mcs2009.pdf. Accessed
tic of red mud, may be advantageous in the construction of roadway May 25, 2009.
intersections or in other situations where the purpose is to warn 3. Glanville, J. I., and P. E. Winnipeg. Bauxite Waste Bricks. Inter-
national Development Research Center. 1991. http://www.idrc.ca/library/
drivers. document/099941. Accessed July 25, 2008.
Use of red mud geopolymers in roadway construction has both 4. Singh, M. S. N., and P. M. Prasad. Preparation of Special Cements from
environmental and economic advantages. First, two of the three Red Mud. Waste Management, Vol. 16, 1996, pp. 665–670.
major parent materials, red mud and fly ash, used in the synthesis of 5. Singh, M. S. N., S. N. Upadhayay, and P. M. Prasad. Preparation of Iron
Rich Cements Using Red Mud. Cement and Concrete Research, Vol. 27,
red mud geopolymers are abundant industrial wastes. The red mud 1997, pp. 1037–1046.
requires special disposal because of its high alkalinity and to date is 6. Marabini, A. M., P. Plescia, D. Maccari, F. Burragato, and M. Pelino.
barely reused or recycled. Use of red mud can significantly reduce New Materials from Industrial and Mining Wastes: Glass-Ceramics
Zhang, He, and Gambrell 9

and Glass- and Rock-Wool Fiber. International Journal of Mineral 20. Mackenzie, K. J. D., D. Brew, R. Fletcher, C. Nicholson, R. Vagana, and
Processing, Vol. 53, 1998, pp. 121–134. M. Schmucker. Advances in Understanding the Synthesis Mechanisms
7. Yalcin, N., and V. Sevinc. Utilization of Bauxite Waste in Ceramic Glazes. of New Geopolymeric Materials. Novel Processing of Ceramics and
Ceramics International, Vol. 26, 2000, pp. 485–493. Composites, Vol. 195, 2005, pp. 187–200.
8. Ayres, R. U., J. Holmberg, and B. Andersson. Materials and the Global 21. Davidovits, J. Geopolymers: Inorganic Polymeric New Materials.
Environment: Waste Mining in the 21st Century. Materials Research Journal of Thermal Analysis, Vol. 37, 1991, pp. 1633–1656.
Society Bulletin, 2001, pp. 477–480. 22. Davidovits, J. Geopolymers: Man-Made Rocks Geosynthesis and the
9. Cundi, W., Y. Hirano, T. Terai, R. Vallepu, A. Mikuni, and K. Ikeda. Resulting Development of Very Early High Strength Cement. Journal
Preparation of Geopolymeric Monoliths from Red Mud–PFBC Ash of Material Education, Vol. 16, 1994, pp. 91–139.
Fillers at Ambient Temperature. Proceedings of the World Congress 23. Duxson, P., A. Fernandez-Jimenez, J. L. Provis, G. C. Lukey, A.
Geopolymer 2005, Saint-Quentin, France, 2005, pp. 85–87. Palomo, and J. S. J. van Deventer. Geopolymer Technology: The
10. Malone, P. G., T. Kirkpatrick, and C. A. Randall. Potential Applications Current State of the Art. Journal of Materials Science, Vol. 42, 2007,
of Alkali-Activated Alumino-Silicate Binders in Military Applications. pp. 2917–2933.
Report WES/MP/GL-85-15. U.S. Army Corps of Engineers, Vicksburg, 24. Duxson, P., J. L. Provis, G. C. Lukey, S. W. Mallicoat, W. M. Kriven,
Miss., 1986. and J. S. J. van Deventer. Understanding the Relationship Between
11. Lyon, R. E., P. N. Balaguru, A. Foden, U. Sorathia, J. Davidovits, and Geopolymer Composition, Microstructure and Mechanical Properties.
M. Davidovics. Fire Resistant Aluminosilicate Composites. Fire and Colloids and Surfaces A: Physicochemical and Engineering Aspects,
Materials, Vol. 21, 1997, pp. 67–73. Vol. 269, 2005, pp. 47–58.
12. Giancaspro, J., P. N. Balaguru, and R. E. Lyon. Use of Inorganic Polymer 25. Xu, H., and J. S. J. van Deventer. The Geopolymerization of Alumino-
to Improve the Fire Response of Balsa Sandwich Structures. Journal of Silicate Minerals. International Journal of Mineral Processing, Vol. 59,
Materials in Civil Engineering, Vol. 18, 2006, pp. 390–397.
2000, pp. 247–266.
13. Goretta, K., J. Fuller, and E. Crawley. Geopolymers. Document
26. Lecomte, I., M. Liegeois, A. Rulmont, R. Cloots, and F. Maseri.
OSR-H-05-05. Air Force Office of Scientific Research Report, 2006.
Synthesis and Characterization of New Inorganic Polymeric Com-
14. Buchwald, A., M. Hohmann, C. Kaps, H. Bettzieche, and J. T. Kuhnert.
Stabilized Foam Clay Material with High Performance Thermal Insulation posites Based on Kaolin or White Clay and on Ground-Granulated
Properties. Ceramic Forum International, Vol. 81, 2004, pp. 39–42. Blast Furnace Slag. Journal of Materials Research, Vol. 18, 2003,
15. Valeria, F. F. B., and J. D. M. Kenneth. Synthesis and Thermal Behavior pp. 2571–2579.
of Potassium Sialate Geopolymers. Materials Letters, Vol. 57, 2003, 27. van Jaarsveld, J. G. S., J. S. J. van Deventer, and A. Schwartzman. The
pp. 1477–1482. Potential Use of Geopolymeric Materials to Immobilize Toxic Metals:
16. Balaguru, P. Geopolymer for Protective Coating of Transportation Part II. Material and Leaching Characteristics. Minerals Engineering,
Infrastructures. Final report, New Jersey Department of Transportation, Vol. 12, 1999, pp. 75–91.
Trenton, 1998. 28. Swanepoel, J. C., and C. A. Strydom. Utilization of Fly Ash in a Geo-
17. Bell, J., M. Gordon, and W. M. Kriven. Use of Geopolymeric Cements as polymeric Material. Applied Geochemistry, Vol. 17, 2002, pp. 1143–1148.
a Refractory Adhesive for Metal and Ceramic Joins. Ceramic Engineering 29. Zhao, Q., B. Nair, T. Rahimian, and P. Balaguru. Novel Geopolymer
and Science Proceedings, Vol. 26, 2005, pp. 407–413. Based Composites with Enhanced Ductility. Journal of Materials Science,
18. Zhang, S. Z., K. C. Gong, and H. W. Lu. Novel Modification Method Vol. 42, 2007, pp. 3131–3137.
for Inorganic Geopolymer by Using Water Soluble Organic Polymers. 30. Annual Book of ASTM Standards (C 150–05, C39/C 39M −05, D 422−05).
Materials Letters, Vol. 58, 2004, pp. 1292–1296. ASTM, West Conshohocken, Pa., 2005.
19. Li, Z., Y. Zhang, and X. Zhou. Short Fiber Reinforced Geopolymer
Composites Manufactured by Extrusion. Journal of Materials in Civil The Physicochemical and Biological Processes in Soils Committee peer-reviewed
Engineering, Vol. 17, 2005, pp. 624–631. this paper.

View publication stats

You might also like