You are on page 1of 10

Case Studies in Construction Materials 11 (2019) e00266

Contents lists available at ScienceDirect

Case Studies in Construction Materials


journal homepage: www.elsevier.com/locate/cscm

Short communication

Role of iron in the enhanced reactivity of pulverized Red mud:


Analysis by Mössbauer spectroscopy and FTIR spectroscopy
Smita Singha , M.U. Aswathb , Rahul Das Biswasb , R.V. Ranganathc ,
Harish K. Choudharyd , Rajeev Kumard , Balaram Sahood
a
Dept. of Civil Engineering, Amrita School of Engineering, Bengaluru, Amrita Vishwa Vidyapeetham, India
b
Dept. of Civil Engineering Bangalore Institute of Technology, Bangalore, India
c
Dept. of Civil Engineering BMS College of Engineering, Bangalore, India
d
Materials Research Centre, Indian Institute of Science, Bangalore, India

A R T I C L E I N F O A B S T R A C T

Article history: The present study is carried out to investigate the contribution of iron in the hematite
Received 28 March 2019 phase in enhancing the reactivity of red mud. Compressive strength test was performed on
Received in revised form 19 June 2019 the binder synthesized from both unpulverized and pulverized red mud. The results
Accepted 20 June 2019
indicated an increase in strength of the pulverized red mud based binder. Test results of
Mössbauer Spectroscopy showed transition of metastable ferrihydrite phase to hematite
Keywords: phase when the red mud was processed. An increase in line-width of the Moessbauer sextet
Red mud
and an increase in the area of doublet of pulverized red mud based binder hinted at the
Pulverization
FTIR
increase in Fe3+ ions which could have replaced Al3+ ions in the geopolymer network.
Mössbauer spectroscopy Results of Fourier-Transform Infrared Spectroscopy (FTIR) test validated the Mössbauer
Compressive strength Spectroscopy results. The participation of Fe3+ ions in geopolymerisation was confirmed
from the reduction of hematite peak (wavenumber at 462 cm1).
© 2019 The Authors. Published by Elsevier Ltd. This is an open access article under the CC
BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).

1. Introduction

Red mud is produced as a byproduct in the aluminium industry during the process of extraction of Alumina through
Bayer’s method. The estimated quantity of its production is approximately 120 million tons annually [1]. The quantum of red
mud produced and its hazardous nature is a threat to the environment and society as a whole [2]. Utilizing this industrial
waste in sustainable applications is the need of the hour. Presence of silica and alumina in red mud makes it viable as source
material for geopolymerisation.
The past investigations have identified red mud as less pozzolanic material that primarily acts as a filler in a
geopolymerised matrix due to the crystalline nature of compounds present [3–7]. The reactivity of a material could be
enhanced by chemical activation, thermal activation and mechanical activation. The drawback of chemical activation is the
need of the unit to recover chemical used for impregnation [8,9]. Thermal activation accelerates the rate of gain of early
strength but adversely affects the long time strength. Mechanical activation refers to grinding of material to increase its
specific surface area. Grinding is the most effective method to regulate the particle size distribution, reduce porosity and to
alleviate the negative effect of crystalline materials on its reactivity. In prior investigations it is discussed that grinding of the
raw material enhanced the reactivity of any binder material. The improved reactivity could be due to increase in the surface
area of the material as the particle size is diminished, change in the crystal structure or phase transformation [10,11].
Felekoglu [12] used mechanical activation to improve the cementitious property of fly ash. Fly ash procured in the original
formed comprised of many coarser particles and exhibited poor cementitious property. Pulverisation enhanced the strength
of the binder and enabled it to be used as a high performance cement. In another investigation, Kumar et al. [13] used

https://doi.org/10.1016/j.cscm.2019.e00266
2214-5095/© 2019 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/
licenses/by-nc-nd/4.0/).
2 S. Singh et al. / Case Studies in Construction Materials 11 (2019) e00266

mechanical activation of slag. He concluded that complete hydration of the mechanically activated slag was possible even
without the use of any chemical and the early strength development of cement improved. Hence, it has been acknowledged
from the past research that grinding enhances a material reactivity and the same applies to red mud too. Singh et al. have
found out that red mud’s reactivity increases on pulverization and the binder synthesized with it gives a very high strength
when used in combination with fly ash and GGBS [14]. As hematite is the primary constituent of red mud, it deems necessary
to understand the effect of pulverization on the reactivity of iron. Mössbauer Spectroscopy is an important method probing
any phase transition or change in valance state of iron [15,16]. Choudhary et al. studied the role of the iron oxide phase during
hydration of cement using Moessbauer Spectroscopy and found that the haematite phase in cement prolongs the setting
time [17]. Yong Hu et al. [18] studied the role of iron phases in a geopolymer matrix through Mössbauer Spectroscopy. He
used thermally treated red mud and fly ash for synthesizing geopolymer binder. According to him, there was an increase in
doublet that suggested the replacement of some of Al3+ ions by Fe3+ in the geopolymer network. In another investigation,
Lemougna et al. [19] studied the participation of iron in the geopolymer formed from volcanic ash. Mössbauer Spectroscopy
results revealed that there was no change in the original ferrroanforsterite phase but new ferric sites were found due to
reaction of alkali with augite.
Hence, we know from the past study that pulverization enhances the reactivity of red mud but the participation of
hematite phase in improving the reactivity of pulverized red mud binder needs to be studied. Mössbauer Spectroscopy is one
of the most reliable test pertaining to the study of iron phases. The present study focusses on study of the contribution of iron
in geopolymerisation by Mössbauer Spectroscopy and FTIR.

2. Materials and methods

2.1. Materials

Red mud was obtained in the form of boulder from Hindalco, Belgaum, India. The plant generates red mud during the
process of extraction of alumina from Bauxite through Bayer’s process. The boulders were coarsely ground and passed
through 300 mm sieve to attain uniformity in test results. The red mud thus obtained is being referred to as “unpulverized”
red mud (RM) in the present study. The “pulverized” red mud (PRM) was obtained by further grinding of red mud in a ball
mill for 15–20 min. The ball mill operated at a speed of 300 rpm and steel balls of 10 mm diameter with ball to powder ratio
as 10:1. The size of the red mud particles thus obtained was found to be below 5 mm.

2.2. Methods

2.2.1. Particle size analysis and microscopic studies


Particle size analysis was conducted using laser particle size analyser, Mastersizer of Malvern Instruments capable of
measuring particle size ranging from 0.05 mm to 900 mm. Mineralogical characterization of red mud was assessed using
X'Pert PRO, PANalytical X- ray Diffractometer. Scanning was done by exposing the red mud to Cu-Kα X-ray radiation at a
speed of 1 deg/min ranging over of 0 to 80 2u. The morphology of red mud particles was obtained by TEM and HRTEM
microscope, FEI, FEG TECHNAI F30 at a voltage of 300 kV. The diffraction pattern and the images were captured by a Gatan
2k2k CCD camera. Sample for studies were prepared by transferring a drop of suspension of red mud in distilled water on a
carbon coated copper grid and allowing it to dry for several hours.

2.2.2. Compressive strength of red mud binder


The geopolymer binder was prepared from red mud (RM), fly ash (FA) and ground granulated blast furnace slag (GGBS).
Red mud varied from 30% to 90% and fly ash from 0%–50%. GGBS was added at 10% for all variations for enabling faster setting.
The various mixture proportions are represented in Table 1. Ambient curing effects superior quality of geopolymer concrete
products as compared to thermal curing [20]. Hence, the geopolymer binder was cured at ambient temperature for 7 days
and tested for compressive strength as per the guidelines specified in IS: 4031 (Part 6) 1988 [21].

Table 1
Quantity of constituents per m3in the geopolymer binder.

Specimen RM (kg) GGBS (kg) FA (kg) NaOH solution (kg) Na2SiO3 (kg)
6M_RM90 1203 134 0 161 402
6M_PRM90 1203 134 0 161 402
6M_RM70 936 134 267 161 402
6M_PRM70 936 134 267 161 402
6M_RM50 669 134 535 161 402
6M_PRM50 669 134 535 161 402
6M_RM30 401 134 802 161 402
6M_PRM30 401 134 802 161 402
S. Singh et al. / Case Studies in Construction Materials 11 (2019) e00266 3

Table 2
Chemical composition of red mud, fly ash and GGBS.

Sample SiO2 Al2O3 Fe2O3 CaO Na2O MgO TiO2 P2 O 5 V2O5 SO3 L.O.I
RM 9.93 18.1 42.9 2.3 5.58 – 9.03 0.35 0.31 – 10.5
FA 61.54 25.37 6.73 3.1 0.97 0.73 – – – 0.62 0.39
GGBS 37.73 14.42 1.11 37.34 – 8.71 – – – 0.39 1.41

2.2.3. FTIR
FTIR of RM, PRM, 6M_RM30 and 6M_PRM30 was done at room temperature by means of Bruker Tensor 27 spectrometer
with 0.4 cm1 resolution. DigiTect Detector system was used to detect the intensity of the IR beam and the detected signal
was digitized and Fourier transformed by the OPUS Software to get the IR spectrum of the sample in the range of 400-
4000 cm1 wave number.

2.2.4. Mössbauer spectoscopy


Around 25–30 mg of powder sample was taken for Mössbauer measurement. For measurement, the samples were
sandwiched in PMMA polymer holders. 57Co source in Rh matrix (Rhodium) was used prior to the experiments; the
calibration of the instrument was performed using alpha Fe foil. The measurement was taken in transmission geometry. The
Mössbauer spectral parameter was fitted using “NORMOS” computer program.

3. Results and discussion

3.1. Characterization of red mud

The primary constituent of red mud is iron followed by significant quantity of alumina and some amount of silica. Class F
fly ash was used for the research. It mainly consists of silica followed by alumina. GGBS has high percentage of calcium
present in the form of calcium oxide (37.34%), Table 2.Loss of ignition of red mud is higher as reflected in Table 2. This might
not necessarily be due to incineration of organic carbon. Oxidation of iron based compounds and desorption of physically or
chemically trapped water might also be the reason behind it [22]. Result of particle size distribution is illustrated in Fig. 1. A
reduction in the range of particle size is observed after pulverization. The range of particle size for unpulverized red mud
(RM) varies from 0.05 mm to 110 mm whereas in the pulverized red mud (PRM), particles are present in 0.05 mm to 2.5 mm
size ranges (Table 3). A vast difference in D90 values of RM and PRM is perceived. It is important to note that pulverization
leads to transition of most of the mesoporous particles to microporous ones.
Smita Singh et al., [14] have explained the enhancement in the reactivity of red mud upon pulverization. One of the factor
contributing to this was the reduction in size of all phases. This is illustrated in Fig. 2 which shows the TEM of red mud before
and after pulverization. The encircled area in the figure shows hematite particles which tranforms into very fine nano
particles post pulverization. Also, increase in lattice spacing was observed after pulverization(HRTEM), which improved the

Fig. 1. Sieve analysis of processed and unprocessed red mud, fly ash and GGBS.
4 S. Singh et al. / Case Studies in Construction Materials 11 (2019) e00266

Table 3
Particle size analysis of red mud.

Sample D10 (mm) D50 (mm) D90 (mm)


RM 0.35 2.5 70
PRM 0.12 0.45 2.3

Fig. 2. TEM of a) unprocessed and b) processed red mud.

ease of protonation and deprotonation of ions thereby further increasing its reactivity. The HRTEM image in Fig. 3 shows the
increase in lattice spacing. The enhancement of lustrous appearance indicates increase in crystallinity. The SAED and XRD
images illustrated in Fig. 4 and Fig. 5, also confirm the increase in crystallinity of hematite phases. In the XRD image, the
hematite peaks have become sharper in pulverized red mud. As reported by Hermanek et al. [23], apart from particle size,
crystallinity too played a key role in the influencing the reactivity of iron. According to their studies, catalytic efficiency of
Iron(III) Oxide nano powders increased as they became more crystalline. The effect of enhanced catalytic efficiency upon
crystallisation was also perceived in transition of propane performed with Zn/Ga/H-ZSM-5-based catalysts [24] or with Mo-
V-Te-Nb mixed oxides synthesized at various pH values. An increasing catalytic activity was witnessed with increasing
degree of crystallinity of the as-prepared catalysts in both the cases [25].

3.2. Compressive strength of red mud based binder

Results of the 7-day compressive strength test is represented in Table 4. It is evident from the results that grinding of
red mud enhanced the mechanical performance of red mud based binder. An increase in the compressive strength is

Fig. 3. HRTEM and lattice spacing of a) unprocessed and b) processed red mud [14].
S. Singh et al. / Case Studies in Construction Materials 11 (2019) e00266 5

Fig. 4. SAED of a) unprocessed and b) processed red mud.

Fig. 5. XRD of red mud [14].

Table 4
Compressive strength of red mud binder.

Specimen Molarity (M) Red mud (%) Compressive strength (MPa) % Increase in 7-day strength due to pulverisation

7-day 28-day
6M_RM90 6 90 5.3 6.1 221
6M_PRM90 6 90 17 19.8
6M_RM70 6 70 9.3 10.6 144
6M_PRM70 6 70 22.7 25.1
6M_RM50 6 50 16.7 18.4 59
6M_PRM50 6 50 26.5 29.9
6M_RM30 6 30 23.7 26.4 71
6M_PRM30 6 30 40.5 43.8

observed for all percentages of red mud, particularly for higher red mud percentages. For instance, strength of binder
containing 90% red mud was 5.3 MPa before pulverization (6M_RM90) that increased to 17 MPa post grinding
(6M_PRM90). Maximum increase in strength of 221% is observed for this variation. Increase in the reactivity of red mud
is the probable reason for greater increase in strength of geopolymer binder for all the variation particularly those with
greater red mud percentage.
6 S. Singh et al. / Case Studies in Construction Materials 11 (2019) e00266

3.3. Mössbauer spectoscopy

Mössbauer Spectroscopy test is used in mineralogy for examining the different iron phases present along with their
magnetic and structural properties. A Mössbauer spectrum is a plot of intensity of absorption of g-ray versus energy for a
particular resonant nucleus such as 57Fe or 119Sn.
Fig. 6 shows the fitted Mössbauer spectra of unprocessed red mud (RM), pulverized red mud (PRM), 6M_RM30 and
6M_PRM30 samples at room temperature. As shown in Fig. 6, the Mössbauer spectrum of each sample is fitted with two sub
spectrum; one sextet (magenta colour) and one central doublet (blue colour). The Mössbauer fitting parameters are
tabulated in Table 5.
The presence of sextet in Mössbauer spectrum of RM sample is attributed to the hematite phase [26]. Fe3+ ions present in
hematite phase gave rise to a sextet with a hyperfine field of 50.5 T. The spectra also reveal that the hematite is the dominant
iron carrying species and there is no substantial trace of goethite phase.

Fig. 6. Room temperature Mössbauer spectra of RM, PRM, 6M_RM30 and 6M_PRM30 sample. The data points are shown in black dots.
S. Singh et al. / Case Studies in Construction Materials 11 (2019) e00266 7

Table 5
Mössbauer spectral fitting parameters of RM, PRM, RM30 and GRM30 sample. BHf = Hyperfine field, I.S = Isomer shift, Q.S = Quadruple splitting, D = Width of
spectra.

Sample code Sub-spectrum BHf(T) I.S (mm/s) Q.S (mm/s) D (mm/s) Area (%)
RM S1 50.40 0.264 0.214 0.493 61.47
D1 0.250 0.598 0.450 38.53
PRM S1 50.51 0.255 0.250 0.428 63.73
D1 0.260 0.578 0.424 36.27
GRM30 S1 50.49 0.263 0.202 0.511 60.22
D1 0.255 0.587 0.478 39.78
GPRM30 S1 50.58 0.265 0.233 0.497 63.09
D1 0.243 0.589 0.462 36.91

The relative percentage area of sextet and doublet was about 63% and 36%. The doublet in Mössbauer spectrum of RM
sample could be assigned to the paramagnetic phases present in the sample. One of these paramagnetic phases consists of
ferrihydrite that exists as a fine grained nanomaterial [27] with bulk of particles (>70%) less than 10 mm in size [26].
Ferrihydrite present in minerals (5Fe2O39H2O) exists as a metastable mineral which acts as a precursor to several crystalline
phases as hematite and goethite [28–31]. The atmospheric CO2 reacts with ferrihydrite forming surface adsorbed species of
carbonate [32].
Pulverization of red mud would have caused the partial transformation of paramagnetic ferrihydrite phases into anti-
ferromagnetic hematite phase. This was evident from the increase in percentage relative area of sextet of pulverized red mud
sample (PRM).
From the Table 5, it is observed that polymerisation of red mud and pulverized red mud has led to increase in the
percentage area of doublets. This increase in area percentage might be due to replacement of aluminium with iron in the
geoplymerised sample (6M_RM30 and 6M_PRM30). As we know that geopolymer is an inorganic polymer with silicate
and aluminate linkages (Si-O-Si and Si-O-Al). Gepolymerisation of red mud would cause the partial replacement of Al3+
ions with Fe3+ ions into the alumina-silicate linkage. The presence of Fe3+ in the silicate linkage in the paramagnetic form
would result in increment of the doublet in Mössbauer spectra and a broadening of the Mössbauer lines of the sextet. The
hyperfine parameters of the doublet are approximately consistent with structural tetrahedral Fe3+ in alumina-silicate
structure [33]. This observation might suggest an increase of the iron in tetrahedral sites of alumina-silicate structure due
to alkaline activation. The isomer shifts for most compounds with Fe3+ in tetrahedral coordination fall into the range of 0.2
to 0.3 mm/sec [34].

3.4. FTIR

Fourier Transform Infrared (FTIR) Spectroscopy used infrared light to make quantitative and qualitative analysis for
organic and inorganic materials. It identifies chemical bonds in a molecule by creating a molecular fingerprint of the tested
sample [35–38].
FTIR spectra of RM and PRM is shown in Fig. 7. The main absorption bands in RM were observed at 3437, 2928, 2858, 1751,
1639, 1452, 999, 549 and 462 cm1 whereas the spectra for PRM were dominated by peaks at 3445, 2922, 2856, 1747, 1639,

Fig. 7. FTIR spectra of unprocessed red mud (RM) and pulverized red mud (PRM).
8 S. Singh et al. / Case Studies in Construction Materials 11 (2019) e00266

1458, 1008, 540 and 468 cm1 wavenumbers. The band at 3437 and 3445 cm1 wavenumbers could be assigned to stretching
bonds of surface water molecules or to an envelope formed by hydrogen bonded surface OH groups [39]. The intensity of the
peak increased upon pulverization, possibly due to adsorption of moisture on the surface of the pulverized particles. Peaks at
2928, 2922, 2858 and 2856 cm1 correspond to CH2 symmetric or asymmetric stretch [40] which could be attributed to the
organic hydrocarbon compound present as impurity in red mud. A sharp peak at 2360 cm1 could be attributed to
deformation vibration of HO H group associated with weakly bond water molecules [41]. Small intensity bands at 1751
and 1747 cm1 represent double bond of C¼O group whereas the peak at band 1639 cm1 represent the bending vibration of
H O H [31] from structural water of aluminosilicate material. Peaks at 1452 and 1458 cm1 are the characteristic of
carbonate component possibly from calcite [41]. The IR bands at 999–1008, 549-540 and 462-468 cm1 could be identified as
quartz and hematite phases which are typical of a red mud spectrum [42].
The main difference between the spectra of RM and PRM is the significant reduction of peak at 2360 cm1 in PRM. The IR
at 2360 cm1 possibly represent the water molecules associated with the ferrihydrite and other hydrate phases of red mud
which reduced considerably after pulverization due to phase transition. The result supported the finding from Mössbauer
Spectroscopy that part of paramagnetic ferrihydrite phases transform into anti-ferromagnetic hematite phase. In general
narrowing of peaks was observed after pulverization. One of the factors causing the decrease in the peak width was the
particle size; smaller the particle size, narrower would be the corresponding IR peaks [43].
Fig. 8 compares the IR spectra of 6M_RM30 and 6M_PRM30. The pattern of spectra is similar from wave number 4000-
1500 cm1 while the fingerprint region (1500-400 cm1) of the geopolymer differs from Fig. 7. The band at 1458 in PRM has
shifted to 1452 in its corresponding geopolymer binder (6M_PRM 30), whereas the band at 1452 in RM is replaced by two

Fig. 8. FTIR spectra of 6M_RM30 and 6M_PRM30.

Fig. 9. Combined FTIR spectra of RM, PRM, 6M_RM30 and 6M_PRM30.


S. Singh et al. / Case Studies in Construction Materials 11 (2019) e00266 9

peaks at 1429 and 1481 in 6M_RM30. This could be due to the carbonation of excess alkali (Na+) resulting in the formation of
new carbonate phases [44]. The quartz peak in RM and PRM is observed to be changing its shape and shifting in the
corresponding geopolymer binders suggesting the formation of Si-O-T geopolymer, where T = Si or Al. It is important to note
that there is an increase in intensity of the hematite peaks at 461 and 555 cm1 of unpulverized red mud after
geopolymerisation, possibly due to decrease in size. But 6M_PRM 30 had only one hematite peak at 462 cm1 and that too
with diminished intensity, this indicated the dissolution of more Fe atoms from the crystalline hematite phase and their
participation in the geopolymeric network. The experimental observation supported the aforementioned assertion that
pulverization caused Fe ions to participate in geopolymerisation and impregnate into geopolymeric network. The combined
FTIR spectra of all the four samples is shown in Fig. 9.

4. Conclusion

The present study investigated the contribution of hematite nano particles in the enhanced reactivity of pulverized red
mud. The results of FTIR and Mössbauer Spectroscopy were analyzed and the study could be concluded as below:

(1) The strength of red mud based binder increased when pulverized red mud was used in its synthesis.
(2) Pulverization of red mud results in the transition of metastable ferrihydrite phase to hematite phase. This is reflected in
the Mössbauer Spectroscopy of RM and PRM where a decrease in the area of doublet (associated with ferrihydrate phase)
and an increase in area of sextet (associated with hematite phase) is observed in the pulverized red mud. Also, an
increase in the area of doublet of the red mud based geopolymer suggests the participation of hematite phase in
geopolymerisation. Some of the Al3+ ions are replaced with Fe3+ ions into the alumina-silicate which leads to increase in
the paramagnetic phase.
(3) The results of FTIR complements the outcomes of Mössbauer Spectroscopy. The reduction if ferrihydrate phase upon
pulverization is revealed by the reduction of peak at 2360 cm1 in PRM. Moreover, the reduction in the hematite peak in
the binder synthesized from pulverized red mud suggests its participation in geopolymerisation.

References

[1] C. Klauber, M. Gräfe, G. Power, Bauxite residue issues: II. Options for residue utilization, Hydrometallurgy 108 (2011) 11–32, doi:http://dx.doi.org/
10.1016/j.hydromet.2011.02.007.
[2] V. Mymrin, K. Alekseev, O.M. Fortini, Y.K. Aibuldinov, C.L. Pedroso, A. Nagalli, E. Winter, R.E. Catai, E.B.C. Costa, Environmentally clean materials from
hazardous red mud, ground cooled ferrous slag and lime production waste, J. Clean. Prod. 161 (2017) 376–381, doi:http://dx.doi.org/10.1016/j.
jclepro.2017.05.109.
[3] J. He, Y. Jie, J. Zhang, Y. Yu, G. Zhang, Synthesis and characterization of red mud and rice husk ash-based geopolymer composites, Cem. Concr. Compos.
37 (2013) 108–118, doi:http://dx.doi.org/10.1016/j.cemconcomp.2012.11.010.
[4] G. Zhang, J. He, R. Gambrell, Synthesis, characterization, and mechanical properties of red mud-based geopolymers, Transp. Res. Rec. J. Transp. Res.
Board 2167 (2010) 1–9, doi:http://dx.doi.org/10.3141/2167-01.
[5] I. Giannopoulou, D. Dimas, I. Maragkos, D. Panias, Utilization of metallurgical solid by-products for the development of inorganic polymeric
construction materials, Glob. Nest J. 11 (2009) 127–136.
[6] N. Ye, Y. Chen, J. Yang, S. Liang, Y. Hu, J. Hu, S. Zhu, W. Fan, B. Xiao, Transformations of Na, Al, Si and Fe species in red mud during synthesis of one-part
geopolymers, Cem. Concr. Res. 101 (2017) 123–130, doi:http://dx.doi.org/10.1016/j.cemconres.2017.08.027.
[7] S. Singh, M.U. Aswath, R.V. Ranganath, Durability of red mud based geopolymer paste in acid solutions, Mater. Sci. Forum. 866 (2016) 99–105.
doi:10.4028/ www.scientific.net/MSF.866.99.
[8] D. Lozano-Castelló, M.A. Lillo-Ródenas, D. Cazorla-Amorós, A. Linares-Solano, Preparation of activated carbons from Spanish anthracite: I. Activation by
KOH, Carbon 39 (2001) 741–749, doi:http://dx.doi.org/10.1016/S0008-6223(00)00185-8.
[9] S.U. Balcı, Kinetics of Activated Carbon Production From Almond Shell, Hazelnut Shell, and Beech Wood and Characterization of Products, (1992) .
[10] M. Senna, Recent development of materials design through mechanochemicalroute, Int. J. Inorg. Mater 3 (2001) 509–514.
[11] G. Mucsi, S. Kumar, B. Cs}oke, R. Kumar, Z. Molnár, Á. Rácz, F. Mádai, Á. Debreczeni, Control of geopolymer properties by grinding of land filled fly ash,
Int. J. Miner. Process. 143 (2015) 50–58.
[12] B. Felekoglu, S. Türkel, H. Kalyoncu, Optimization of fineness to maximize thestrength activity of high-calcium ground fly ash – portland
cementcomposites, Constr. Build. Mater. 23 (5) (2009) 2053–2061.
[13] S. Kumar, R. Kumar, A. Bandopadhyay, T.C. Alex, B. Ravi Kumar, S.K. Das, S.P. Mehrotra, Mechanical activation of granulated blast furnace slag and its
effect on the properties and structure of Portland slag cement, Cem. Conr. Compos. 30 (8) (2008) 679–685.
[14] S. Singh, M.U. Aswath, R.V. Ranganath, Effect of mechanical activation of red mud on the strength of geopolymer binder, Constr. Build. Mater. 177 (2018)
91–101, doi:http://dx.doi.org/10.1016/j.conbuildmat.2018.05.096.
[15] A.V. Anupama, W. Keune, B. Sahoo, Thermally induced phase transformation in multi-phase iron oxide nanoparticles on vacuum annealing, J. Magn.
Magn. Mater. 439 (2017) 156–166.
[16] S.K. Boda, A.V. Anupama, B. Basu, B. Sahoo, Structural and magnetic phase transformations of hydroxyapatite-magnetite composites under inert and
ambient sintering atmospheres, J. Phys. Chem. C 119 (2015) 6539–6555.
[17] H.K. Choudhary, A.V. Anupama, R. Kumar, M.E. Panzi, S. Matteppanavar, Baburao N. Sherikar, B. Sahoo, Observation of phase transformations in cement
during hydration, Constr. Build. Mater. 101 (2015) 122–129.
[18] Y. Hu, S. Liang, J. Yang, Y. Chen, N. Ye, Y. Ke, S. Tao, K. Xiao, J. Hu, H. Hou, W. Fan, S. Zhu, Y. Zhang, B. Xiao, Role of Fe species in geopolymer synthesized
from alkali-thermal pretreated Fe-rich Bayer red mud, Constr. Build. Mater. 200 (2019) 398–407, doi:http://dx.doi.org/10.1016/j.
conbuildmat.2018.12.122.
[19] P. Lemougna, K. Mackenzie, G.N.L. Jameson, H. Rahier, U. Melo, The Role of Iron in the Formation of Inorganic Polymers (geopolymers) From Volcanic
Ash: a 57Fe Mössbauer Spectroscopy Study, (2013), doi:http://dx.doi.org/10.1007/s10853-013-7319-4.
[20] S. Singh, Effect of curing methods on the property of red mud based geopolymer, Int. J. Civil Eng. Technol. 8 (2017) 1481–1489.
[21] IS 4031(Part 6), Method of Physical Test for Hydraulic Cement- Determination of Compressive Strength of Hydraulic Cement Other Than Masonry
Cement, India, Bureau of Indian Standards, (2005) .
10 S. Singh et al. / Case Studies in Construction Materials 11 (2019) e00266

[22] A. Atasoy, An investigation on characterization and thermalanalysis of the aughinish red mud, J. Therm. Anal. Calorim. 81 (2005) 357–361, doi:http://
dx.doi.org/10.1007/s10973-005-0792-5.
[23] M. Hermanek, R. Zboril, I. Medrik, J. Pechousek, C. Gregor, Catalytic Efficiency of Iron(III) Oxides in Decomposition of Hydrogen Peroxide: Competition
between the Surface Area and Crystallinity of Nanoparticles, J. Am. Chem. Soc. 129 (2007) 10929–10936, doi:http://dx.doi.org/10.1021/ja072918x.
[24] C.P. Nicolaides, N.P. Sincadu, M.S. Scurrell, Aliphatic-aromatic Hydrocarbon Interconversion Over Zn and Ga Modified Zeolites: Effect of Zeolite
Crystallinity and Method Used for Modification on Performance, Studies in Surface Science and Catalysis (2004), Elsevier, 2004, pp. 2347–2352, doi:
http://dx.doi.org/10.1016/S0167-2991(04)80496-8.
[25] J.M. Oliver, J.M. López Nieto, P. Botella, A. Mifsud, The effect of pH on structural and catalytic properties of MoVTeNbO catalysts, Appl. Catal. A Gen. 257
(2004) 67–76, doi:http://dx.doi.org/10.1016/S0926-860X(03)00632-X.
[26] E. Cristina de Resende, Ido R.G. Carvalho, M. Schlaf, M.C. Guerreiro, Red Mud waste from the Bayer process as a catalyst for the desulfurization of
hydrocarbon fuels, RSC Adv. 4 (2014) 47287–47296, doi:http://dx.doi.org/10.1039/C4RA07635D.
[27] W. Salama, M. El Aref, R. Gaupp, Spectroscopic characterization of iron ores formed in different geological environments using FTIR, XPS, Mössbauer
spectroscopy and thermoanalyses, Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 136 (2015) 1816–1826, doi:http://dx.doi.org/10.1016/j.
saa.2014.10.090.
[28] U. Schwertmann, E. Murad, EffectofpHon the formation ofGoethiteand s’ z laJ trl, I, Clays Clay Miner. 31 (1983) 277–283.
[29] U. Schwertmann, J. Friedl, H. Stanjek, From Fe (III) ions to Ferrihydrite and then to hematite from Fe (III) ions to Ferrihydrite and then to hematite, J.
Colloid Interface Sci. 223 (2014) 215–223, doi:http://dx.doi.org/10.1006/jcis.1998.5899.
[30] H. Stanjek, H.-H. Becher, U. Schwertmann, Long-term in vitro transformation of 2-line ferrihydrite to goethite/hematite at 4, 10, 15 and 25 C, Clay
Miner. 39 (2004) 433–438, doi:http://dx.doi.org/10.1180/0009855043940145.
[31] Y. Cudennec, A. Lecerf, The transformation of ferrihydrite into goethite or hematite, revisited, J. Solid State Chem. 179 (2006) 716–722, doi:http://dx.doi.
org/10.1016/j.jssc.2005.11.030.
[32] D.B. Hausner, N. Bhandari, A.M. Pierre-Louis, J.D. Kubicki, D.R. Strongin, Ferrihydrite reactivity toward carbon dioxide, J. Colloid Interface Sci. 337 (2009)
492–500, doi:http://dx.doi.org/10.1016/j.jcis.2009.05.069.
[33] O. Castelein, L. Aldon, J. Olivier-Fourcade, J.C. Jumas, J.P. Bonnet, P. Blanchart, 57Fe Mössbauer study of iron distribution in a kaolin raw material:
influence of the temperature and the heating rate, J. Eur. Ceram. Soc. 22 (2002) 1767–1773, doi:http://dx.doi.org/10.1016/S0955-2219(01)00496-4.
[34] J.B. Choi, Seoul National University, Mössbauer Spectroscopy and Crystal Chemistry of Aenigmatites, (1983) .
[35] V. Jagadeesha Angadi, A.V. Anupama, R. Kumar, H.K. Choudhary, S. Matteppanavar, H.M. Somashekarappa, B. Rudraswamy, B. Sahoo, Composition
dependent structural and morphological modifications in nanocrystalline Mn-Zn ferrites induced by high energy gamma-irradiation, Mater. Chem.
Phys. 199 (2017) 313–321.
[36] V. Jagadeesha Angadi, A.V. Anupama, Harish K. Choudhary, R. Kumar, H.M. Somashekarappa, M. Mallappa, B. Rudraswamy, B. Sahoo, Mechanism of
g-irradiation induced phase transformations in nanocrystalline Mn0.5Zn0.5Fe2O4 ceramics, J. Solid State Chem. 246 (2017) 119–124.
[37] V. Rathod, A.V. Anupama, R. Vijaya Kumar, V.M. Jali, B. Sahoo, Correlated vibrations of the tetrahedral and octahedral complexes and splitting of the
absorption bands in FTIR spectra of Li-Zn ferrites, Vib. Spectr. 92 (2017) 267–272.
[38] A.V. Anupama, V. Rathod, V.M. Jali, B. Sahoo, Composition dependent elastic and thermal properties of Li-Zn ferrites, J. Alloys. Compd. 728 (2017)
1091–1100.
[39] M. Goti c, S. Musi
c, Mössbauer, FT-IR and FE SEM investigation of iron oxides precipitated from FeSO4solutions, J. Mol. Struct. 834–836 (2007) 445–453,
doi:http://dx.doi.org/10.1016/j.molstruc.2006.10.059.
[40] C. Christou, A. Agapiou, R. Kokkinofta, Use of FTIR spectroscopy and chemometrics for the classification of carobs origin, J. Adv. Res. 10 (2018) 1–8, doi:
http://dx.doi.org/10.1016/j.jare.2017.12.001.
[41] M.M. Al Bakri Abdullah, K. Hussin, M. Bnhussain, K.N. Ismail, Z. Yahya, R.A. Razak, Fly ash-based geopolymer lightweight concrete using foaming agent,
Int. J. Mol. Sci. 13 (2012) 7186–7198 doi:10.3390/ijms13067186.
[42] D. Dodoo-Arhin, D.S. Konadu, E. Annan, F.P. Buabeng, A. Yaya, Fabrication and characterization of Ghanaian bauxite red mud-clay composite bricks for
construction applications, Am. J. Mater. Sci. 3 (2013) 110–119, doi:http://dx.doi.org/10.5923/j.materials.20130305.02.
[43] R. Ruppin, R. Englman, Optical phonons of small crystals, Rep. Prog. Phys. 33 (1970) 149–196, doi:http://dx.doi.org/10.1088/0034-4885/33/1/304.
[44] J.N. Yankwa Djobo, A. Elimbi, H.K. Tchakouté, S. Kumar, Mechanical activation of volcanic ash for geopolymer synthesis: effect on reaction kinetics, gel
characteristics, physical and mechanical properties, RSC Adv. 6 (2016) 39106–39117.

You might also like