You are on page 1of 18

materials

Article
Design of Inorganic Polymer Mortar from
Ferricalsialic and Calsialic Slags for Indoor
Humidity Control
Elie Kamseu 1,2 , Isabella Lancellotti 1, , Vincenzo M. Sglavo 3, , Luca Modolo 4 and
Cristina Leonelli 1, *,
1 Department of Engineering Enzo Ferrari, University of Modena and Reggio Emilia, Via P. Vivarelli 10,
Modena 41125, Italy; kamseuelie2001@yahoo.fr (E.K.); isabella.lancellotti@unimore.it (I.L.)
2 Local Materials Promotion Authority, Nkolbikok Yaound 2396, Cameroon
3 Department of Industrial Engineering, University of Trento, Via Sommarive, 9, Trento 38123, Italy;
vincenzo.sglavo@unitn.it
4 Acciaierie Bertoli Safau S.p.A., Via Buttrio 28fraz. Cargnacco, Pozzuolo del Friuli (UD) 33050, Italy;
L.Modolo@absacciai.it
* Correspondence: cristina.leonelli@unimore.it; Tel.: +39-059-2056247
These authors contributed equally to this work.

Academic Editor: Claudio Ferone


Received: 7 March 2016; Accepted: 19 May 2016; Published: 24 May 2016

Abstract: Amorphous silica and alumina of metakaolin are used to adjust the bulk composition of
black (BSS) and white (WSS) steel slag to prepare alkali-activated (AAS) mortars consolidated at room
temperature. The mix-design also includes also the addition of semi-crystalline matrix of river sand
to the metakaolin/steel powders. The results showed that high strength of the steel slag/metakaolin
mortars can be achieved with the geopolymerization process which was particularly affected by the
metallic iron present into the steel slag. The corrosion of the Fe particles was found to be responsible
for porosity in the range between 0.1 and 10 m. This class of porosity dominated (~31 vol %) the
pore network of B compared to W samples (~16 vol %). However, W series remained with the higher
cumulative pore volume (0.18 mL/g) compared to B series, with 0.12 mL/g. The maximum flexural
strength was 6.89 and 8.51 MPa for the W and B series, respectively. The fracture surface ESEM
observations of AAS showed large grains covered with the matrix assuming the good adhesion bonds
between the gel-like geopolymer structure mixed with alkali activated steel slag and the residual
unreacted portion. The correlation between the metallic iron/Fe oxides content, the pore network
development, the strength and microstructure suggested the steel slag's significant action into the
strengthening mechanism of consolidated products. These products also showed an interesting
adsorption/desorption behavior that suggested their use as coating material to maintain the stability
of the indoor relative humidity.

Keywords: steel slag; mix-design; alkali-activated slag (AAS) cements; pore size distribution;
microstructure; moisture control capacity

1. Introduction
Due to their industrial scale of production, fly ash and slag are now presented as low life cycle
CO2 emissions, environmentally-friendly and sustainable solid precursors for alkali-activated cement,
mortar and concrete [1,2]. A wide variety of such industrial wastes is described in the literature with
very variable composition [3,4] thus making their alkali-activation a complex issue. Apart from coal
fly ash, blast furnace slag, calcined kaolins, and few others for which it can be said confidently that are
suitable for EN 197 cements and can also be properly and successfully employed in the production

Materials 2016, 9, 410; doi:10.3390/ma9060410 www.mdpi.com/journal/materials


Materials 2016, 9, 410 2 of 18

of geopolymers or hybrid alkaline cements, few recent works have been aimed at valorization of
poorly reactive/not-recycled waste materials [5,6]. In the present study two different steel slags
were used with metakaolin and sand to generate an aluminosilicatic matrix via alkali activation.
This aluminosilicatic matrix was designed to attain a good chemical stability taking into account the
existing models on mixed CSH (CaOSiO2 H2 O) structure. In fact, the chemical and mechanical
stability of alkali-activated materials is based on some equilibrium between the geopolymer gel phase
and the reacted aggregate that constitutes the alkali activated matrix. The stability of silicates-based
binders, including ordinary Portland cement, has been found to be improved by the presence of Al
ions during the polycondensation reactions [1,7]. Aluminum cations incorporated in CSH gel were
found to improve the resistance to carbonation. It was observed that all aluminum present in the
silicate tetrahedra chains were five-fold coordinated species in interlayer and six-fold coordinated
at the surface and theywere incorporated into the amorphous silica phase forming fully condensed
tetrahedral Al(-SiO)3 sites [1,7]. Myers et al. [1] identified in granulated blast furnace slag three distinct
tetrahedral Al sites: Q3 (1Al), Q4 (3Al) and Q4 (4Al) which are indicative on the cross-linking degree
in the calcium (alkali) aluminosilicate hydrate (C(N)ASH) gels and the presence of additional
highly polymerized aluminosilicate products. For appropriate Si/Al ratio, the amount of non-bridging
oxygen (NBO) is a function of SiO2 /Na2 O molar ratio. Low value of SiO2 /Na2 O is consistent with the
increasing number of NBO sites where SiQ2 and SiQ1 structural units were preferentially formed from
silicate chains, dimers and monomers. High SiO2 /Na2 O molar ratio generates a decreasing number
of NBO sites with structures consisting principally of SiQ3 and SiQ4 structural units which form
silicate 3D frameworks and sheets [7]. When CaO is present, CSH is the principal hydration product
and primary binding phase; it can be described as a single chain structure that faces polymerization
decreasing with the increase of Ca content [8].
While pure geopolymer (NASH or KASH) and ordinary Portland cement (CSH) phases
are generally formed from pure metakaolin and kaolin-calcite respectively, the industrial and natural
wastes (fly ash, slag, incinerator bottom ash, volcanic ash, rice husk ash, etc.) are complex mixes of
geopolymers and OPC forming elements. The high temperature of process, the level of vitrification
and melting, the cooling rate and the initial bulk composition will result in the formation of amorphous
or partially crystalline phases, as mullite, anorthite, plagioclase, pyroxene etc. [911]. The better
crystallized phases will not dissolve even in high alkaline media, thus hindering the polycondensation.
Therefore, the appropriate mode to design the alkali-activated materials should not be generalized
to a simple activation of finely grounded powder with the alkaline solution, but a more complex
mix-design approach is necessary.
Slag from ferrous metallurgy are generally calcium- and iron-rich (40 wt %52 wt % CaO;
70 wt %80 wt % FeO; 20 wt %30 wt % Fe2 O3 [9]) with limited Si and Al content (10 wt %19 wt %
SiO2 ; 1 wt %3 wt % Al2 O3 [9]). The main minerals consist of C3 S, C2 S, C4 AF, C12 A7 , ROss (R = Ca,
Fe, Mn, Mg solid solution) and free CaO [911] embedded in a variable amount of amorphous phase
depending on the slag cooling practice. Steel slag has a large amount of non-active compounds such
as RO and Fe3 O4 . In the high alkaline context, like that of inorganic polymer cement, the non-active
components as well as the others crystalline phases of the slag are dissolved in an incongruent manner
essentially at their surfaces. With their fine particle size, they seem attractive for the mechanical
reinforcement of the matrices through the filler effect and densification. Additionally, they can ideally
act as nucleation sites to optimize the reaction of geopolymerization.
Starting from the considerations listed above, the approach used for this research was to consider
the slag powder as binder precursor, finer and more amorphous particles, as well as fine aggregates,
non-reactive particles. We also operated to obtain a geopolymeric paste with Si/Al close to 2.02.5 and
Na/Al equal to 1, with the addition of a minimum amount of metakaolin. The larger unreacted slag
phases remained embedded in such a matrix as aggregate in a mortar.
Materials 2016, 9, 410 3 of 18

The two different steel slags from Electric Arc Furnace (EAF) steelmaking industry (Italy),
used in this study, were prepared in an industrial process of production of manufactured aggregates.
The presence of slags reactive phases, eventually metals, in alkaline environment were supposed to
produced geopolymeric cements with a relevant porosity suitable for application as passive phase
change humidity control material, which can moderate the indoor moisture. In fact our formulations
were expected to be close to vesuvianite, sepiolite and zeolite which are all used as hygroscopic
base materials in the manufacturing of moisture control materials for indoor wall coatings. Actually,
it is generally observed that materials produced with cold setting processes as cement based mortars,
concretes, etc. is capable to collect up to 8 wt %10 wt % of humidity after complete curing and exposure
to ambient environment. So for, aside the determination of the adsorption/desorption capacity and the
permeable voids assessed with water absorption for these steel slag based inorganic polymer mortars,
we also evaluated the morphology and the microstructure using several techniques, in particular
X-ray diffraction (XRD), Fourier Infrared spectroscopy (FIT-IR), Mercury Intrusion Porosimeter (MIP),
three-point flexural strength and Environmental Scanning Electronic Microscope (ESEM).

2. Materials and Methods

2.1. Characterization of Ferricalsialic and Calsialic Slags


Two steel slags from EAF steelmaking process were provided by Acciaierie Bertoli Safau in Italy
and identified as ABS black (BSS) and ABS white (WSS). The first slag is obtained by the primary
process in production of steel (melting of scrap in EAF furnace), while the second is obtained by the
secondary metallurgical process (final steel alloy in ladle furnace). Additionally, BSS can be classified as
an oxidation slag, while WSS is a reduction slag. Both slags, also indicated as SS, coming out from the
furnaces, after cooling process, constitute the raw materials for production of aggregates. The cooling
of slags is realized in a two-step process: Firstly the hot material is placed underneath water sprinklers
of in order to rapidly reduce its temperature, and second the material is stocked in a pile to conclude
the cooling process by air. Especially for black slag, the cooling process is very important and necessary
to reduce the amount of free lime and to determine a high mechanical resistance for material. The black
slag is generally considered as not suitable for the use in production of cement, while it is considered a
good material to produce aggregates for concrete, bituminous conglomerate and civil constructions.
Due to its high content of free lime and poor mechanical resistance, the white slag is instead suitable for
the production of cement and may substitute the lime itself in stabilization of soils. The steel slags used
in this study were provided as manufactured aggregates, obtained by processing the raw materials
with the following operations: ageing storage, removing of scrap particles by magnetic separation,
reduction of dimension by crushing, calibration of different products by screening.
The chemical compositions of the slag were determined by X-ray fluorescence (ARL QUANT'X
EDXRF, Thermo Fisher Scientific, Waltham, MA, USA) are presented in Table 1. Both slags had
negligible content of C. The S content is 0.3% for BSS and 1.6% for WSS. The sulphur content is
reported as SO3 according to XRF analysis. The chemical dissolution of white ABS slag performed
using concentrated NaOH (8M) (extended description of the test method is reported in Ref. [12])
allowed to determine 35.89 wt % of reactive species, while the remaining 64.11 wt % could be accounted
as of aggregates. For the black slag, 72.96 wt %was reactive species while only 27.04 wt % was identified
as aggregates.
Materials 2016, 9, 410 4 of 18

Table 1. Chemical compositions (wt %) of black and white steel slag (EUROPEAN STANDARD EN
15309 by XRF).

Oxides ABS Black Oxides ABS White


CaO 36.52 CaO 40.31
Fe2 O3 31.87 Fe2 O3 12.28
SiO2 12.78 SiO2 11.33
MnO 5.90 MnO 2.19
Al2 O3 5.04 Al2 O3 15.43
MgO 3.26 MgO 11.54
Cr2 O3 2.15 Cr2 O3 0.73
TiO2 0.41 TiO2 0.31
SO3 0.31 SO3 1.68
V2 O5 0.28 Be...F 0.36
P2 O5 0.50 Others 3.84
Others 0.98

The mineralogical phases as determined by XRD analysis, using X-ray (XPert PROPANAlytical,
Eindhoven, The Netherland) CuK, Ni-filtered radiation ( = 1.54184 ) are given in Figure 1a.
Based on the chemical compositions in Table 1, the SS can be classified as calsialic for the white
and ferricalsialic for the black slag specimens. The particle size distributions which were measured
by a laser particle size instrument (MASTER SIZER 2000, Malvern, UK) arepresented in Figure 1b.
The morphology of the grains as observed by ESEM is presented in Figure 1c. It is evident that the
black slag is finer than the white sample in the range of small particles, up to 32 m (Figure 1b).
Between 32 and 283 m the B specimen is finer and when the particle size was greater than 283 m
the black slag is coarser. The white slag with low iron content and highest CaO presents particles size
similar to that of ordinary Portland cement [2] while the particle size distribution of the black sample
is typical of steel slag. The iron oxide and RO phases constitute the essential of the coarse particles.
The specific surface area as determined using nitrogen sorption method, BET
(Brunauer-Emmet-Teller, method by nitrogen absorption on a GEMINI 2360, Micromeritics
Instrument Corp., Norcross, GA, USA) gave 3.39 m2 /g for the white ABS slag and 2.06 m2 /g for
the black ones. Under thermal analysis (Figure 2), both slags presented endothermic peak around
106 C related to the physico-absorbed water present within the samples. White slag showed peaks of
decomposition at 256, 324, 413, and 487 C. Peaks consecutive to transformation of hydrated for of
Fe2 O3 . xH2 O, y-FeOH(OH) and Mg(OH)2 [4]. These compounds formed during the water granulation
processes through the pouring into larger quantity of water to avoid excessive foaming. The peak
situated at 760 C corresponds to the decomposition of the Ca based phases of the slag. The peak
at 1119 C is related to the crystallization of aluminosilicate type mullite. The weight decrease for
both steel slags (Figure 2b) is indicative for the percentage of volatile materials into the samples
(Ca(OH)2 , H2 O, CO2 , etc.).
The metakaolin (MK) was obtained from a standard kaolin used for glaze formulation in the
ceramic industry after calcinations at 700 C for 4 hand ground finely down to 30 m (D90). The oxide
molar composition of the metakaolin, determined by XRF, gave a value of Si/Al = 1.3 in terms of the
mass ratio [13].
Materials 2016, 9, 410 5 of 18
Materials
Materials2016,
2016,9,9,410
410 55of
of18
18

(a)
(a) (b)
(b)

(c)
(c)
Figure1.1. Characterization
Figure Characterization of of the
the starting
starting two
two steel
steel slags
slags used
used for
for this
this study:
study: (a)
(a) XRD
XRD patterns
patterns
Figure 1. Characterization of the starting two steel slags used for this study: (a) XRD patterns
(C
Ca(OH)2;; CF
(C(C===Ca(OH)
Ca(OH) CF
2;2CF
==xCaO
xCaO
= xCaO
yFe2O3; CS = C2S + C3S; Q = quartz; RO = pyroxenes solid-solution); (b) grain
Fe3;2 O
yFey2O CS3 ; =CS
C2S=+CC23SS;+QC=3quartz;
S; Q = RO
quartz; RO = pyroxenes
= pyroxenes solid-solution);
solid-solution); (b) grain
size
(b) distribution;
grain (c) powder
size distribution; morphology
(c)morphology
powder morphology with ESEM.with ESEM.
size distribution; (c) powder with ESEM.

(a) (b)
(a) (b)
Figure 2. Thermal characterization of the starting two steel slags used for this study: (a) DSC curve
Figure 2. TGA
Thermal characterization of the starting two steel slags used for this study: (a) DSC curve
Figure(b)
and analyses.
2. Thermal characterization of the starting two steel slags used for this study: (a) DSC curve
and (b) TGA analyses.
and (b) TGA analyses.
2.2. Mix-Design and Preparation of Inorganic Polymer Mortars
2.2. Mix-Design and Preparation of Inorganic Polymer Mortars
2.2. Mix-Design and Preparation of Inorganic Polymer Mortars
2.2.1. Mix-Design
2.2.1.
2.2.1. Mix-Design
Mix-Design
The white and black ABS slags presented 40.31 wt % and 36.52 wt % of CaO but the SiO2
The
contents white
of both
The white and
and blackABS
materials
black ABS
were slags
slags11.33presented
wt % and
presented 40.31
40.31 wt %wt
12.78 wt
and% and
%,36.52 36.52 wtCaO
respectively.
wt % of % of
butCaO
These contentsbut result
the SiO the SiOin2
2 contents
contents
Ca/Si of
molar both materials
ratios of 5.45 were
for the 11.33
WSS wt
and% and
4.38 12.78
for BSS.wt
The%, respectively.
main oxide These
content, contents
i.e.,
of both materials were 11.33 wt % and 12.78 wt %, respectively. These contents result in Ca/Si molar SiO 2 + result
Al 2O3 in
+
Ca/Si
Fe 2Omolar ratios
3, are less of
than 5.45
40 wtfor the
%for WSS
the and
white 4.38
ABS for
and BSS.
<50 The
wt % main
for oxide
the content,
black sample,i.e.,
farSiO 2 + Al
from
ratios of 5.45 for the WSS and 4.38 for BSS. The main oxide content, i.e., SiO2 + Al2 O3 + Fe2 O3 , are less the 2O3
~80+
Fewt
2O%3, are less than
proposed 40 wt %for the
by Davidovidts [14]white ABS and <50
as requirement wt formulation
for the % for the black sample, far from the ~80
of polysialates.
wt % proposed by Davidovidts [14] as requirement for the formulation of polysialates.
Materials 2016, 9, 410 6 of 18

than 40 wt % for the white ABS and <50 wt % for the black sample, far from the ~80 wt % proposed by
Davidovidts [14] as requirement for the formulation of polysialates.
We proposed to adjust the bulk chemical composition with addition of metakaolin, MK, in a
mix with semi-crystalline particles of silica sand. The white and black ABS slags, SS, were used with
metakaolin in the following proportions: 1 SS:1 MK(B60 and W60), 2 SS:1 MK (B90 and W90) and
3 SS:1 MK (B120 and W120). The obtained solid precursors were finely ground down to 60 m and
used to form inorganic polymer mortars with sand representing two times the mass of slag-metakaolin,
SS-MK, powder. The sand was also ground to reduce the particles size below 400 m for research sake,
but to improve the economy of the process the natural sand can be replaced with fine gravel.

2.2.2. Sample Preparation


NaOH (laboratory grade of pellets from Sigma Aldrich, Milan, Italy) was dissolved in deionized
water to reach a concentrated 8M solution. This solution, after having been stored for a minimum
of 48 h, was mixed with sodium silicate (SiO2 /Na2 O, molar ratio = 2.98; industrial grade from
Ingessil, Verona, Italy) in volume proportion of 2:3. The final mix was used to prepare the inorganic
polymer pastes with proportions reported in Table 2. The variation in the liquid: solid weight ratio
(liquid = soda plus Na-silicate solution; solid = SS plus MK plus sand) as indicated into the Table 2
as well as the NaOH:Na2 SiO3 volume ratio were chosen considering the behavior of the steel slag
into the alkaline solution and the objective to draw pastes with similar final viscosity. The difference
between the series B and W was dependant on the difference between the potential reactive fraction of
each steel slag. The pastes for the six formulations were mixed in a bench mixer for approximately
10 min and poured in 1 1 14 cm3 polyurethane molds. Soon after pouring, the specimens were
isolated from air into closed plastic films for the first 72 h. The curing continued in ambient conditions,
room temperature around 21 2 C and 54% 1% of humidity (usual indoor environment) for up to
365 days. This curing procedure was considered optimal for the samples prepared in the laboratory
with no defects. The mechanical characterization was performed on the specimens cured for 28
and 365 days.

Table 2. Mix-design of steel slag (SS) inorganic polymer mortars with similar value in the
cone-fluidity test.

Formulation SS:MK 1 (Slag + MK):Sand Solid:Liquid NaOH:Na2 SiO3


B60 1:1 1:2 1.3:0.4 2:3
B90 2:1 1:2 1.3:0.4 2:3
B120 3:1 1:2 1.3:0.4 2:3
W60 1:1 1:2 1:0.4 2:3
W90 2:1 1:2 1:0.4 2:3
W120 3:1 1:2 1:0.4 2:3
1: MKmetakaolin.

2.3. Characterization of the Cured Steel Slag-BasedAlkali-Activated Mortars


The mineralogical analysis of the alkali-activated steel slag, AAS, mortars were carried out with
an X-ray powder diffractometer, XRD (see above) from 5 to 70 , 2theta steps and integrated at the
rate of 2 s per step.
Fourier transformed infrared spectroscopy, FT-IR, (Avatar 330 FTIR, Thermo Nicolet, Waltham,
MA, USA) was performed on each sample analyzing fine powders of ground specimens ( < 80 m)
collected from samples used for the mechanical test after curing for 28 days. A minimum of 32 scans
between 4000 and 500 cm1 were averaged for each spectrum at intervals of 1 cm1 .
An Autopore IV 9500, Micromeritics, 33000 psi (228 MPa) Mercury Intrusion Porosimeter
(MIP, Manchester, UK) covering the pore diameter range from approximately 360 to 0.005 m having
two low-pressure ports and one high-pressure chamber was used for the pores analysis. Pieces were
Materials 2016, 9, 410 7 of 18

prepared from the bulk of each sample with specimens of ~1 cm3 of volume for the MIP. Apart from
the pore volume and size, the equipment software also automatically evaluated the tortuosity as the
ratio of the length of the path described by the pore space to the length of the shortest path across
a porous mass.
Specimens (10 for each composition to reduce the error within 20%) with nominal size of
10 0.10 mm in width, 10 0.10 mm thick and a length of 140 0.10 mm were used for the
mechanical tests by using a three-point bending configuration with span equal to 40 mm according to
ASTM C1161-02c standard. The specimens were loaded by a universal testing machine, type MTS 810,
(MTS Systems Corporation, Eden Prairie, MN, USA) with cross-head speed of 5 mm/min.
The microstructure of the inorganic polymer cement specimens was studied using an
Environmental Scanning Electron Microscope (ESEM, Model Quanta 200, FEI, Hillsboro, OR, USA) at
low vacuum. Both fractured pieces from mechanical testing and etched polished specimens were used.
The specimens were preliminarily coated with 10 nm thick gold layer. The ESEM was equipped with
an Oxford Instruments energy dispersive spectrometer (EDS) for the microanalysis thus facilitating the
investigation of phase distribution in the matrix.
To evaluate the capability to help in the control of indoor humidity in buildings,
the adsorption/desorption behavior of the AAS mortars was evaluated on two specimens aged
28 days per each formulation dried at room temperature for 24 h and then soaked in deionizer water
for 24 h accordingly to previous studies [15]. Specimens of prismatic shape as those used for mechanical
characterization were weighted in their saturated condition and the time of the first measurement was
considered as t0 . After each hourfor the first 10 hthe measurement was repeated, then weight
measurement of the sample continued every 24 h for 1000 h. This measurement was repeated four
times (four cycles, also indicated as water cycles) in the laboratory conditions (T = 21 2 C, R.H.
(Room Humidity) 54% 1%). Finally, water absorption was determined as ratio of the water saturated
specimen with respect to the dry one; variation of the water release after saturation is expressed as
rate of weight change and the humidity content coincides with the water retained by the sample after
the 1000 h test time. Additionally, the shrinkage and expansion were monitored, however, changes in
dimensions were not identified.

3. Results

3.1. Phases Development


The XRD patterns of both white and black slag-based AAS mortars exhibited amorphous structure
and higher peak intensities of crystalline phases (Figure 3a,b). The halo that generally characterizes the
fully reacted and amorphous structure of alkali-activated materials is visible around 2 = 27 for all
formulations. These broad diffuse halos in XRD patterns correspond to three-dimensional networks
for low angle values (2 < 30 ) [5,16]. In our study, these halos were centered at 28.42 , 28.16 and
27.64 respectively for 1:1, 1:2 and 1:3 MK:WSS; and at 27.33 , 27.30 and 27.31 for MK:BSS. Typical
metakaolin based inorganic polymer cement generally shows 2 < 27 [17]. Thus it can be deduced
that the introduction of the steel slag, with low aluminosilicate content, reduced the degree of typical
NASH gel formation and polymerization of aluminosilicates (geopolymerization).The shift toward
higher 2 values by introducing steel slag is also confirmed by Bignozzi et al. [13] for geopolymers
with ladle slag (white slag).
The formation of only amorphous phases peculiar of MK geopolymers is no longer ensured
after the introduction of steel slags. In fact, the as-received steel slag are partially composed of
crystalline phases, such as C3 S, C2 S, C4 AF, and C12 A7 as well as RO (R = Ca, Fe, Mn) solid solution [18],
which contributed to modify the level of disorder and crystallinity into matrix. Their participation into
the geopolymerization process is limited to the surface reactivity that contribute to the densification of
the mortars, nevertheless no quantitative calculation of the crystallinity/amorphicity ratio was done
in this study.
Materials 2016, 9, 410 8 of 18
Materials 2016, 9, 410 8 of 18

(a) (b)

(c) (d)
Figure 3. XRD and FT-IR characterization of the white and black slag-based AAS mortars: (a) XRD
Figure 3. XRD and FT-IR characterization of the white and black slag-based AAS mortars: (a) XRD of
of white; and (b) black ABS; (c) FT-IR of white; and (d) black ABS.
white; and (b) black ABS; (c) FT-IR of white; and (d) black ABS.
From the FT-IR spectra (Figure 3c,d), it is observed that all samples show two bands related to
OH
From thestretching: (i) at 3347
FT-IR spectra cm1 for
(Figure B series
3c,d), it isand 3386 cmthat
observed
1 for W series; and (ii) ~1646 cm1 for W
all samples show two bands related to
series and 1654 cm1 for B
OH stretching: (i) at 3347 cm 1 for B series and 3386 cm1 content
series. The increase in steel slag decrease the intensity of both
for W series; and (ii) ~1646 cm1 for W
peaks related to 1the OH stretching. This is related to the reduction of typical NASH phases
series and 1654 cm for B series. The increase in steel slag content decrease the intensity of both
from metakaolin-based geopolymers that decrease with the increase of the steel slag. The behavior
peaks of
related to thecan
the bands OH stretching.
be correlated This
with theisevolution
related toofthe thereduction
peaks around of typical
14001450NASH phases from
cm1, generally
metakaolin-based geopolymers that decrease
ascribed to three-fold-coordinated aluminum with
thatthe increase
decreases in of the steel
intensity withslag.
the The behavior
increase of the of the
bands steel
can be correlated with the evolution of W60,
the peaks 1 , generally ascribed
cmintensity
slag content (Figure 3c,d). B60 and whicharound
presented14001450
the higher for the
OH-stretching peaks, have
to three-fold-coordinated the principal
aluminum peaks of alkali-activated
that decreases in intensity aluminosilicates
with the increase (SiOSi(Al))
of the steelat slag
~980 cm 1 and 982 cm1 respectively [19].
content (Figure 3c,d). B60 and W60, which presented the higher intensity for the OH-stretching peaks,
By increasing
have the principal peaks theofslag content, the peaks
alkali-activated of B series moves
aluminosilicates to 994 cm1 for
(SiOSi(Al)) B120 and
at ~980 cm995 cm1982
1 and for cm1
W120. The shift of the principal band towards higher wavenumber with the increase of the steel
respectively [19].
slag content indicates that with the introduction of the slag, very limited content of aluminum
By increasing the slag content, the peaks of B series moves to 994 cm1 for B120 and 995 cm1 for
available and the possibility of their incorporation into the geopolymer network decrease [1923].
W120. The
Thebands
shift with
of thesmall
principal band
intensities towards
between 690higher
and 850wavenumber with to
cm1 are attributed thegeopolymer
increase ofproducts.
the steel slag
contentThey
indicates that to
are related with
the the introduction
symmetric of the
stretching slag, very
vibration limited
of SiOSi (orcontent
SiOAl) ofbridges
aluminum available
as well as the and
the possibility of their
cyclosilicates incorporation
vibrations [19,20]. They into
arethe
alsogeopolymer
affected by thenetwork
variation decrease [1923].
of the slag content.The
Thebands
band with
between 880between
small intensities and ~900 690cm1and
, presented
850 cmas 1a are
small shoulder on
attributed the main SiOT
to geopolymer band, is They
products. assigned
aretorelated
the stretching vibration mode of AlO bonds in condensed AlO 4 groups.
to the symmetric stretching vibration of SiOSi (or SiOAl) bridges as well as the cyclosilicates
vibrations [19,20]. They are also affected by the variation of the slag content. The band between 880
and ~900 cm1 , presented as a small shoulder on the main SiOT band, is assigned to the stretching
vibration mode of AlO bonds in condensed AlO4 groups.

3.2. Porosity and Pore Size Distribution

3.2.1. Pore Permeable to Mercury in MIP Tests


The variations of the cumulative pore volume of the white and black slag-based AAS mortars are
summarized in Figure 4. B60 presents 0.160 mL/g of cumulative pore volume. This value decreases
respect to B90. Similarly, the larger capillary pores in W120 are significantly bigger than in W90.
The understanding of the relative decrease in cumulative pore volume with the increase of the slag
content needs two important factors to be considered: (i) both slags contain a fraction of aggregates
(27 wt % for the black slag and 64 wt % for the white one) which in this study are presented as fine
aggregates [17]; and (ii) the role of metallic iron in the dissolution and polymerization reactions
Materials 2016, 9, 410 9 of 18
[23,24]. As described above, the finer particles of the steel slags are essentially composed of ROss (R
= Ca, Fe, Mn, and Mg solid solution) of semi-crystalline nature. These fines undergo to incongruent
to 0.122 mL/g
dissolution, and
i.e., 0.118
they mL/g,
react respectively
essentially on the forsurface,
B90 andduring
B120. Athesimilar trend isthus
dissolution, observed for the W
participating in
series: the cumulative pore
the chemical-physical volume
reactions decreases
typical from 0.180 mL/g
of geopolymeric for W60 to
consolidation, 0.170 and
reducing the0.160 mL/g
volume of for
air
W90
voidsandandW120,
bubbles respectively.
as generallyIn general,
observedtheinpores volume for the
metakaolin-based W seriespolymer
inorganic is more important than
cement [15,19].
for
TheB role
series. However,
of fines in thethe pores volume
reduction in volume
of pores both inorganic polymer
has already beenmortars is lowerbythan
demonstrated in typical
Kamseu et al.
metakaolin
[15]. The fines into the context of inorganic reactions are known to accelerate the hydration acting as
based inorganic polymer cement (~0.300 mL/g) [21,22]. These values are pretty reliable as
far as our experience
nucleation is related
sites for binding since (CSH,
phases we usually operate after NASH,
C(N)ASH, careful calibration of the equipment [17].
etc.) [25].

(a) (b)
Figure 4. Cumulative pore volume of the slag-based AAS mortars: (a) white; and (b) black ABS.
Figure 4. Cumulative pore volume of the slag-based AAS mortars: (a) white; and (b) black ABS.

The small particles of metals still present in the slags are corroded by the alkaline media
enhancing theon
Focusing theofcumulative
heat reaction andpores
thusvolume curves
the kinetic of(Figure 4), the progressive
the reactions. These metallicdecrease in the
particles arepore
also
(larger capillary pores with size >1 m) volume noted has important values for B120 with
responsible for the increase of the fraction of the larger capillary pores due to hydrogen evolution respect to B90.
Similarly,
similarly the larger
to what capillary
shown pores
for the in of
case W120 are significantly
aluminum bigger than inalkaline
metal in concentrated W90. The understanding
solutions [26]. The
of the relative decrease in cumulative pore volume with the increase of the slag
major fraction of pores of the W series is concentrated between 0.01 and 0.1 m (Figure 5). Such content needs two
important factors to be considered: (i) both slags contain a fraction of aggregates
pores correspond to those generally described for the fly ash and metakaolin based inorganic (27 wt % for the
black slagcements
polymer and 64 wt % forThe
[17,21]. the white
coarserone) whichpores
capillary in thiswith
study arebetween
size presented as fine
1 and aggregates
10 m that are[17];
not
and (ii) the role of metallic iron in the dissolution and polymerization reactions
generally observed in the IPC, are ascribed to pores developed from the corrosion reactions [23,24]. As described
of the
above,
metallictheiron
finer particles in
contained of the
the slag.
steel slags are essentially
This class of pores iscomposed of ROss for
more important (R =the
Ca,BFe, Mn, andthan
specimens Mg
solid solution) of semi-crystalline nature. These fines undergo to incongruent dissolution, i.e., they react
essentially on the surface, during the dissolution, thus participating in the chemical-physical reactions
typical of geopolymeric consolidation, reducing the volume of air voids and bubbles as generally
observed in metakaolin-based inorganic polymer cement [15,19]. The role of fines in the reduction
of pores volume has already been demonstrated by Kamseu et al. [15]. The fines into the context of
inorganic reactions are known to accelerate the hydration acting as nucleation sites for binding phases
(CSH, C(N)ASH, NASH, etc.) [25].
The small particles of metals still present in the slags are corroded by the alkaline media enhancing
the heat of reaction and thus the kinetic of the reactions. These metallic particles are also responsible
for the increase of the fraction of the larger capillary pores due to hydrogen evolution similarly to
what shown for the case of aluminum metal in concentrated alkaline solutions [26]. The major fraction
of pores of the W series is concentrated between 0.01 and 0.1 m (Figure 5). Such pores correspond
to those generally described for the fly ash and metakaolin based inorganic polymer cements [17,21].
The coarser capillary pores with size between 1 and 10 m that are not generally observed in the IPC,
are ascribed to pores developed from the corrosion reactions of the metallic iron contained in the slag.
This class of pores is more important for the B specimens than for W ones in agreement with the iron
content accounting for 12.3 wt % and 31.9 wt % for W and B, respectively. In the W series the dominance
of the gel pores (size between 0.01 and 0.1 m) make the cumulative pore curves present the typical
characteristics of the IPC mortars: apart from a small increase in the region of the capillary pores due to
iron residues (110 m), a significant increase is observed at 0.1 m. The fraction of the pores between
Materials 2016, 9, 410 10 of 18

for W ones in agreement with the iron content accounting for 12.3 wt % and 31.9 wt % for W and B,
respectively. In the W series the dominance of the gel pores (size between 0.01 and 0.1 m) make
the cumulative pore curves present the typical characteristics of the IPC mortars: apart from a small
Materials 2016, 9, 410 10 of 18
increase in the region of the capillary pores due to iron residues (110 m), a significant increase is
observed at 0.1 m. The fraction of the pores between 110 m in W series increased from 16.7 vol
% form
110 W60in Wto series
17.6 vol % and 18.8
increased fromvol16.7%volfor%W90 and W120,
for W60 to 17.6 respectively.
vol % and 18.8 From
vol % thefor
behavior
W90 andofW120,
the
cumulative pores
respectively. Fromvolume and the
the behavior particles
of the size distribution
cumulative pores volume (Figures 4 and
and the 5) the size
particles twodistribution
classes of
pores into W series are isolated, hence the interconnection degree is very
(Figures 4 and 5) the two classes of pores into W series are isolated, hence the interconnection limited. In the B series theis
degree
cumulative pore volume curves evidence the effect of the corrosion of metallic
very limited. In the B series the cumulative pore volume curves evidence the effect of the corrosion iron on the fraction
of metallic
of the poresiron>0.1onm.
theOnly ~16 of
fraction volthe
% pores
of pores>0.1have
m.sizeOnly<0.1~16mvolin %
B90of and
poresB120
have specimens,
size <0.1 with
m in
B90 and B120 specimens, with the percentage raising to 18.8 vol % for B60. Pores with size >0.1 of
the percentage raising to 18.8 vol % for B60. Pores with size >0.1 m dominate the total porosity m
the specimens
dominate of theporosity
the total B series.ofThis
the isspecimens
completely ofdifferent in series
the B series. ThisWiswhere most of
completely the pores
different in have
series
Wsize <0.1 most
where m. Consequently,
of the pores have the size
structure
<0.1 m.of the B series materials
Consequently, corresponds
the structure of theto a matrix
B series with
materials
interconnected pores as confirmed in the Figures 4 and 5. The increase of
corresponds to a matrix with interconnected pores as confirmed in the Figures 4 and 5. The increase ofthe fraction of larger
capillary
the fractionpores with the
of larger steel slag
capillary porescontent
with thein both
steelBslag
andcontent
W series in confirms
both B andtheWrole of the
series metallic
confirms the
iron during the alkali-activation of the steel slag mortars. The difference between the two samples
role of the metallic iron during the alkali-activation of the steel slag mortars. The difference between
shows that the Fe particles action on porosity can be tuned using the mix-design.
the two samples shows that the Fe particles action on porosity can be tuned using the mix-design.

(a) (b)
Figure 5. Pores size distribution of the slag-based AAS mortars: (a) white; and (b) black ABS.
Figure 5. Pores size distribution of the slag-based AAS mortars: (a) white; and (b) black ABS.

3.2.2. Pores Permeable to the Moisture and Liquid Water


3.2.2. Pores Permeable to the Moisture and Liquid Water
After 365 days curing at room temperature, the white and black slag-based AAS mortars
After 365 days curing at room temperature, the white and black slag-based AAS mortars release
release 2.5 wt %, 5.9 wt % and 4.1 wt % moisture for W60, W90 and W120, respectively, when
2.5 wt %,at5.9
treated 100wtC% for
and244.1h.wt % moisture
The moisture for W60, W90
released and W120,
in similar respectively,
conditions for B60,when treated
B90 and B120 at 100
wereC
for
6.6 24
wth.%,
The5.7moisture
wt % and released
4.4 wtin%similar conditions
respectively. It isfor B60,that
clear B90moisture
and B120content
were 6.6increases
wt %, 5.7with
wt %theand
4.4 wt % respectively. It is clear that moisture content increases with the
amount of white slag while the behavior is opposite with black slag. W60 specimen with the loweramount of white slag while
the behavior
moisture is opposite
release (2.5 wt %) with black slag.
possesses the W60
higher specimen with the lower
specific cumulative poremoisture release
volume (0.18 (2.5 wt
mL/g). B60%)
possesses
specimen withthe higher specific
the higher cumulative
moisture releasepore volume
(6.6 wt (0.18 mL/g).
%) presents B60volume
the highest specimen with the
of larger higher
capillary
moisture release (6.6 wt %) presents the highest volume of larger capillary
pores with size >1 m. It is observed that the specimens of the series B reabsorbed the moisture pores with size >1 m.
Itwith
is observed
a higherthatratethe specimens
compared to of
thethe series B reabsorbed
specimens of the W series the moisture
(Figure with a higher
6). These rate compared
observations are
to the specimens of the W series (Figure 6). These observations are correlated
correlated to the pore network structure of the two AAS mortars: B series samples present a larger to the pore network
structure
fraction ofof pores
the two AAS
with mortars:
size >0.1 m B series samples
(>70 vol present
%) while a larger
W series fraction ofpossess
specimens pores with
onlysize
~16 >0.1 m
vol %
(>70
poresvol %) size
with while>0.1W m
series
andspecimens
more thanpossess
80 vol %only ~16with
pores vol % pores
size <0.1with
m. size
The >0.1
largem and more
fraction than
of pores
80 vol % pores with size <0.1 m. The large fraction of pores with size >0.1
with size >0.1 m explains the higher degree of pore connectivity in series B with respect to series m explains the higher
degree of pore connectivity
W. Specifically, the fractionin of series B with
pores withrespect to series1W.
size between Specifically,
and 10 m that the are
fraction of pores
supposed to with
be
size between
developed 1 and
from the10Fem that arecorrosion
particles supposedduring
to be developed from the Fe
the alkali-activated particles
reactions corrosion
(Figures during
4 and the
5) are
alkali-activated
more importantreactions
in the series(Figures
B. The4 important
and 5) are volume
more important in thecan
of these pores series B. The important
be considered as the volume
origin
of
of these
porespores can be hence
coalescence, considered as the origin
significantly reducingof pores
the gel coalescence,
pores. B60, hence
B90 andsignificantly
B120 specimensreducingcuredthe
gel pores. B60, B90 and B120 specimens cured at room temperature for 365 days at ambient relative
humidity (R.H. 54% 1%) absorb 5.9 wt %, 4.5 wt % and 3.3 wt % water, respectively,. This decrease
in water absorption with the amount of steel slag for B series can be correlated with the decrease of the
cumulative pore volume (Figure 4).
Water absorption of the W series is 3.5%, 5.5%, 9.7% for W60, W90 and W120, respectively. It is
observed that the presence of steel slag in W samples increases the ability to absorb water, as it can
be appreciated from the variation of the moisture content (Figure 6). These results demonstrate that
the permeability of the steel slag based mortars can be interpreted for B samples with the
cumulative
Materials pore
2016, 9, 410 volume variation: the reduction of the cumulative pore volume implies a reduction
11 of 18
of the water absorption.

(a) (b)
Figure 6. Moisture absorption-desorption performance of the slag-based AAS mortars: (a) moisture
Figure 6. Moisture absorption-desorption performance of the slag-based AAS mortars: (a) moisture
absorption performance at ambient temperature after 24 h in oven at 100 C; and (b) moisture
absorption performance at ambient temperature after 24 h in oven at 100 C; and (b) moisture
desorption performance at ambient temperature after complete saturation in water.
desorption performance at ambient temperature after complete saturation in water.

The reduction of the water absorption is in agreement with the reduction of the volume of the
Water absorption
larger capillary pores of the size
with W series
>0.1 is 3.5%,
m. In 5.5%,
the W9.7% for although
series, W60, W90the andcumulative
W120, respectively. It is
pore volume
observed that the presence of steel slag in W samples increases the ability
reduces with the white steel slag, water absorption increases. Here, a better correlation is made to absorb water, as it can
be appreciated from the variation of the moisture content (Figure 6). These
between water absorption and the volume of the larger capillary pores. This volume increases from results demonstrate that
the permeability
16.7% for W60 of tothe steel slag
17.6% and based
18.8%,mortars
for W90 can be
andinterpreted for B samplesThis
W120, respectively. withmeans
the cumulative
that the
pore volume variation: the reduction of the cumulative pore volume
permeability in W series is mostly affected by the development of larger capillary pores from implies a reduction of the
the
water absorption.
corrosion in the pore solution of the iron metal particles. Finally, it can be mentioned that the
The reduction
presence of the steelof slag
the water
reducesabsorption
the capacityis inofagreement with the
the metakaolin reduction
based of the volume
geopolymer of the
to accumulate
larger capillary pores with size >0.1 m. In the W series, although the cumulative
moisture. The specimens with SS:MK = 1:1 (B60 e W60) remain after 1000 h with ~1 wt % moisture pore volume reduces
with the white
at room steel slag,
temperature water
while absorption
specimens withincreases.
SS:MK =Here, a better
2:1 and correlation
3:1 (B90, is made
B120, W90 andbetween water
W120) reduce
absorption and the volume of
the moisture content down to ~0.25 wt %. the larger capillary pores. This volume increases from 16.7% for W60
to 17.6% and 18.8%, for W90 and W120, respectively. This means that the
The collected results make this class of slag (EAF white and black steel slag) as promisingpermeability in W series is
mostly
additive affected by the
materials fordevelopment
the moistureofcontrol
larger capillary
capacity pores from the corrosion
of geopolymeric mortars. in It
theseems
pore solution
that the
of the iron metal particles. Finally, it can be mentioned that the presence
mechanism of the moisture control capacity enhancement is focused essentially to the formation of the steel slag reduces the of
capacity of the class
the particular metakaolin
of pores based
(110geopolymer
m) in theto accumulate moisture.
geopolymeric matrix. The specimens with SS:MK = 1:1
(B60 e W60) remain after 1000 h with ~1 wt % moisture at room temperature while specimens with
SS:MK = 2:1 and
3.3. Density andFlexural
3:1 (B90,Strength
B120, W90 and W120) reduce the moisture content down to ~0.25 wt %.
The collected results make this class of slag (EAF white and black steel slag) as promising additive
The physical and mechanical properties of the white and black slag-based AAS mortars are
materials for the moisture control capacity of geopolymeric mortars. It seems that the mechanism of
summarized in Table 3. In general, the bulk density of standard metakaolin based IPC is ~1.50 g/cm3
the moisture control capacity enhancement is focused essentially to the formation of the particular
[27,28]. The presence of fine aggregates (64.1 wt % for W and 27.0 wt % for B) is the primary reason
class of pores (110 m) in the geopolymeric matrix.
of the density increase. The presence of sand also plays a significant role together with the presence
of iron
3.3. (12.3and
Density wt Flexural
% for WStrength
and 31.9 wt % for B). The relative increase in density corresponds to higher
flexural strength, raising from ~4 MPa [17,21] to 6.8 MPa for W60 specimen. White ABS contains
The physical and mechanical properties of the white and black slag-based AAS mortars
higher aggregates and this favors the reactivity reduction and polycondensation, thus reducing the
are summarized in Table 3. In general, the bulk density of standard metakaolin based IPC is
flexural strength (Figure 7). In addition, increasing the curing time up to 365 days does not change
~1.50 g/cm3 [27,28]. The presence of fine aggregates (64.1 wt % for W and 27.0 wt % for B) is
the primary reason of the density increase. The presence of sand also plays a significant role together
with the presence of iron (12.3 wt % for W and 31.9 wt % for B). The relative increase in density
corresponds to higher flexural strength, raising from ~4 MPa [17,21] to 6.8 MPa for W60 specimen.
White ABS contains higher aggregates and this favors the reactivity reduction and polycondensation,
thus reducing the flexural strength (Figure 7). In addition, increasing the curing time up to 365 days
does not change this tendency. ForB series the high reactivity of the black slag can explain the
continuous increase of the polycondensation, thus increasing the flexural strength with the steel slag
Materials 2016, 9, 410 12 of 18
Materials 2016, 9, 410 12 of 18
this tendency. ForB series the high reactivity of the black slag can explain the continuous increase of
the polycondensation, thus increasing the flexural strength with the steel slag amount. The
amount. The in
difference difference in thedevelopment
the strength strength development
between thebetween
whitethe white
and andslag
black black
canslag
be can be justified
justified by
bydifferent
differentaggregates
aggregates fractions (27 wt % for B sample and 64 wt % for W). Actually, since
fractions (27 wt % for B sample and 64 wt % for W). Actually, since the two theSS
two
SSpresented
presentedthese
thesetwo
twodifferent
differentsand
sandfractions,
fractions,weweadopted
adopteddifferent
differentpaste/sand
paste/sand ratios
ratios in in order
order to to
optimize the final consolidated product in terms of chemical stability and strength. The
optimize the final consolidated product in terms of chemical stability and strength. The optimumoptimum ratio
corresponds to the best
ratio corresponds ITZ
to the features.
best ITZ features.

Table 3. Microstructural
Table datadata
3. Microstructural as determined by MIP
as determined by ofMIP
all the
of formulations aged 28 daysand
all the formulations aged 28 mechanical
daysand
properties at 1properties
mechanical year aging.
at 1 year aging.

Bulk Density Median Pore Total Tortuosity Flexural Strength


Formulation Bulk Density Median Pore Total Tortuosity Flexural Strength
Formulation (gcm3) 3 Diameter (m) Porosity/% (a.u.) at 365atDays
365 Days
(g cm ) Diameter (m) Porosity/% (a.u.) (MPa)(MPa)
B60 2.13 1.4 34.0 29.6 7.7
B60 2.13 1.4 34.0 29.6 7.7
B90 B90 2.22 2.22 0.80.8 29.0
29.0 39.839.8 6.9 6.9
B120 B120 2.39 2.39 0.90.9 28.1
28.1 39.039.0 8.1 8.1
W60 W60 1.74 1.74 0.10.1 31.9
31.9 38.138.1 6.8 6.8
W90 W90 1.77 1.77 0.10.1 29.8
29.8 42.442.4 6.2 6.2
W120 1.82 0.1 28.6 49.8 5.6
W120 1.82 0.1 28.6 49.8 5.6

Figure Three
7. 7.
Figure Threepoint
pointbending
bendingstrength
strength of
of slag-based AASmortars
slag-based AAS mortarswith
withwhite
whiteand
and black
black ABS
ABS after
after
28 28
and 365
and 365days.
days.

Commentingthethe
Commenting relationship
relationship between
between porosity
porosity and and
flexuralflexural strength,
strength, we canwe saycan
thatsay that
according
according to our experience on geopolymers and ceramics, we noticed that in general when the
to our experience on geopolymers and ceramics, we noticed that in general when the total porosity is
total porosity is determined with the mercury intrusion porosimeter, the values obtained are
determined with the mercury intrusion porosimeter, the values obtained are generally higher when
generally higher when compared to others methods. This is particularly true for geopolymers, due
compared to others methods. This is particularly true for geopolymers, due to the fact that the
to the fact that the nanoporosity of these materials are takeninto account. At the same time we also
nanoporosity of these materials are takeninto account. At the same time we also noticed from literature
noticed from literature data that this nanoporosity does not affect the materials during the
data that this nanoporosity
mechanical solicitation. does not affect the materials during the mechanical solicitation.
Concerning
Concerningthe thechemical
chemicalreactivity
reactivity and chemical bonding
and chemical bondingininrelation
relationtotomechanical
mechanical strength,
strength,
wewe noticed
noticed that although the incongruent dissolution of these aggregates in the pore solution can becan
that although the incongruent dissolution of these aggregates in the pore solution
bepointed
pointedoutoutasaswell
wellas as their
their participation
participationininthe thesurface
surfacephysical-chemical
physical-chemical reactions,
reactions, their
their high
high
proportion limits the strength when amount of binder is not enough
proportion limits the strength when amount of binder is not enough to embed the aggregates. On to embed the aggregates.
Onthetheother
other hand,
hand, both
both white
white andand black
black steelsteel
slagsslags
show show
similarsimilar
silicasilica
(SiO2(SiO 2 ) content.
) content. However,However,
the
thewhite
whitespecimenscontain15
specimenscontain15 wtwt
%% AlAl
2O23,Othree
3 , three times
times larger
larger than than in the
in the black
black onesones (5 wt
(5 wt %) %) (Table
(Table 1). 1).
TheThehigh
hightemperatures
temperaturesatatwhich
whichslags
slagsareare produced (>1600C),
produced (>1600 C),allow
allowthe theformation
formation of of more
more stable
stable
aluminosilicatesasasglass-ceramics
aluminosilicates glass-ceramics in in the
the white
whitespecimens.
specimens. These
Thesephases
phaseshave limited
have dissolution
limited in
dissolution
in alkaline
alkaline media,
media,meaning
meaningrelatively
relativelylow lowsoluble
solublesilica and
silica and alumina
alumina and andlowlow gelgel
formed
formed forfor
thethe
polycondensation optimization. The relative increase in strength in both B and W series with
curing time from 28 to 365 days is related to the additional pozzolanic reactions that take place
Materials 2016, 9, 410 13 of 18
Materials 2016, 9, 410 13 of 18
polycondensation optimization. The relative increase in strength in both B and W series with curing
time from 28 to 365 days is related to the additional pozzolanic reactions that take place with very
with very The
low rate. low rate. The completely
completely dissolved dissolved
metakaolin metakaolin
and steeland slagsteel
finerslag finer particles
particles contribute
contribute to the
to the formation of oligomers, that convert to polymerized small clusters
formation of oligomers, that convert to polymerized small clusters of aluminosilicates or calcium of aluminosilicates or
calcium silicates/calcium-aluminosilicate
silicates/calcium-aluminosilicate hydrated gels hydrated
(510 gels (510 nm)
nm) [17,29]. The[17,29]. The effectiveness
effectiveness of these gels of to
these gelsstrength
enhance to enhance strength
depends depends
on the on the gels/aggregates
gels/aggregates ratio, their
ratio, their interfaces andinterfaces
chemicaland chemical
equilibrium.
equilibrium.
The aggregates Thedissolution
aggregates level
dissolution
and the level and the
strain strain generated
generated by the surrounding
by the surrounding matrix
matrix play a
play a significant
significant role in role in determining
determining the finalthestrength
final strength
since itsince it affects
affects cross-linking
cross-linking amongamongphases.phases.
In the
In the metakaolin
metakaolin based based IPC,reactive
IPC, the the reactive
phasesphases
are are generally
generally NASH
NASH gelsleading
gels leadingtoto aa flexural
strength near 4 MPa [17]. [17]. Increasing
Increasing the the available
available soluble
soluble silica
silica (in
(in the
the range
rangeof ofSi/Al
Si/Al = 22.5 molar
mechanical strength
ratio), the mechanical strength increases
increases as as aa result
result ofof the
the pores
pores volume
volume and and size
size reduction
reduction [17,29].
[17,29].
For both steel slags in W series, an improvement of the flexural strength was
steel slags in W series, an improvement of the flexural strength was observed as the resultobserved as the result of
thethe
of cumulative
cumulative pore
porevolume
volumereduction
reductionalthough
although the the
presence of aof
presence new classclass
a new of pores between
of pores 1 and
between 1
10
andm 10ismformed, (not significant)
is formed, (Figure
(not significant) 5). The
(Figure 5).larger amount
The larger of pores
amount with size
of pores with>0.1
sizem
>0.1inm
B series
in B
limits limits
series the strength increase.
the strength The strengthening
increase. The strengthening mechanism
mechanism in this
in case includes
this case a reduction
includes a reductionof the
of
cumulative
the cumulative porepore
volume.
volume.A possible
A possibleimprovement
improvement of theofmechanical
the mechanical performance couldcould
performance have have
been
reached
been extending
reached the first
extending step
the of curing
first step ofofcuring
72 h. The
of 72laboratory environment
h. The laboratory might havemight
environment brought to
have
excessivetodesaturation
brought of samples of
excessive desaturation in the very in
samples early
theage.
very early age.

3.4. Microstructure
3.4. Microstructure
Figures 88 and
Figures and 99 show
show thethe microstructure
microstructure of of the
the inorganic
inorganic polymer
polymer mortars
mortars with
with ABS
ABS white
white and
and
ABS
ABS black,
black, respectively.
respectively. The
The micrographs,
micrographs, takentaken onon polished
polished samples,
samples, point
point out larger quartz
out larger grains
quartz grains
(dense
(dense and
and compact)
compact)andandthetheaggregates
aggregates form
form thethe
steel slag
steel (highly
slag porous).
(highly TheThe
porous). matrix is compact
matrix and
is compact
homogeneous with well-embedded aggregates. At limited magnification,
and homogeneous with well-embedded aggregates. At limited magnification, no crack can be no crack can be observed.
The fracture
observed. Thesurface of the
fracture specimens
surface with whitewith
of the specimens (Figure 10a)
white and black
(Figure 10a) ABS (Figure
and black 10b)
ABS show 10b)
(Figure that
the aggregates are well covered by the geopolymer matrix composed
show that the aggregates are well covered by the geopolymer matrix composed of of calcium-aluminosilicate gels.
In this case, aggregates are
calcium-aluminosilicate gels.notIn clearly visible,
this case, same as
aggregates areinnot
theclearly
polished samples:
visible, same this
as indemonstrates
the polished
samples: this demonstrates the good capacity of the steel slag based IPC to embed aggregatesbond
the good capacity of the steel slag based IPC to embed aggregates with significant adhesion with
strength. Atadhesion
significant higher magnification
bond strength. (Figures 11 and
At higher 12), the microporous
magnification (Figures 11network structure
and 12), in B series
the microporous
can observed.
network The in
structure matrix is dominated
B series can observed.by homogeneously dispersedby
The matrix is dominated pores with sizes in dispersed
homogeneously the range
of 1 to 10 m. This is the origin of the difference in microstructure between B and
pores with sizes in the range of 1 to 10 m. This is the origin of the difference in microstructure W series. Pores
between 1 and 10 m can also be observed in W specimens; however,
between B and W series. Pores between 1 and 10 m can also be observed in W specimens; the concentration is too low to
be relevant
however, forconcentration
the the mechanical performance
is too of the matrix.
low to be relevant for the mechanical performance of the matrix.

(a) (b)
Figure 8. ESEM
Figure 8. ESEM microstructure
microstructure of
of white slag-based AAS
white slag-based AAS mortars
mortars showing:
showing: (a)
(a) overall
overall microstructure
microstructure
with quartz and iron-rich aggregates (solid solution) embedded into the matrix; (b) interface zone
with quartz and iron-rich aggregates (solid solution) embedded into the matrix; (b) interface zone
demonstrating good adhesion bonds between the various aggregates and the matrix.
demonstrating good adhesion bonds between the various aggregates and the matrix.
Materials 2016, 9, 410 14 of 18
As already mentioned, the importance of the specific pores range (110 m) is directly related
toAs
thealready mentioned,
iron content, the is
which importance
12.3 wt %offor
the the
specific
ABSpores
whiterange
and (110 m)
31.9 wt %isfor
directly
the ABSrelated
black.
to Additionally,
the iron content, which is 12.3 wt % for the ABS white and 31.9 wt % for the ABS
the residues of metallic iron formed in the slag are corroded by the NaOH solution black.
Additionally, thethe
during 2016,
Materials residues ofphase
9,dissolution
410 metallic
of iron formed in the slag
geopolymerization. Theareevolved
corrodedgasbyistheresponsible
NaOH solutionfor
14 ofthe
18
during the dissolution
formation of pores with phase of geopolymerization.
size between 1 and 10 m for W Theseries
evolved gas
and 0.1 andis 10responsible for the
m for B series.
formation of pores with size between 1 and 10 m for W series and 0.1 and 10 m for B series.

(a) (b)
(a) (b)
Figure 9. ESEM microstructure of black slag-based AAS mortars showing: (a) overall microstructure
Figure
Figure 9. ESEM microstructure of black slag-based AAS mortars showing: (a) overall microstructure
with9.quartz
ESEM microstructure
and of black
iron-rich aggregatesslag-based AAS mortars
(solid-solution) showing:
embedded (a) matrix;
into the overall microstructure
(b) interface zone
with quartz and iron-rich aggregates (solid-solution) embedded into the matrix; (b) interface zone
with quartz and iron-rich
demonstrating aggregates
good adhesion bonds(solid-solution) embedded into the matrix; (b) interface zone
between aggregates.
demonstrating good adhesion bonds between aggregates.
demonstrating good adhesion bonds between aggregates.

(a) (b)
(a) (b)
Figure 10. ESEMmicrostructure of fracture surface at low enlargements of the specimens with:
Figure
(a) 10. ESEMmicrostructure
white (b) black ABS. of fracture surface at low enlargements of the specimens with:
Figure 10.ABS;
ESEMmicrostructure of fracture surface at low enlargements of the specimens with: (a) white
(a) white ABS; (b) black ABS.
ABS; (b) black ABS.
The homogeneous dispersion of pores from the exothermic reactions of iron corrosion is due to
theThe homogeneous
efficient dispersionofofsteel
mixing operation pores from
slag, the exothermic
metakaolin and sand reactions of iron
particles corrosion
for the is dueof
preparation tothe
theAAS As already
efficient mixing mentioned,
operation thesteel
importance of the specific poresparticles
range (110 is directly related
m)preparation to
paste. Jointly with of slag, metakaolin
the proper and sand
mixing operation, the presence for theof iron particles of the
in a
AASthe iron content,
paste. Jointly which is 12.3
withsystem wt
the proper% for the ABS white and 31.9 wt % for the ABS black. Additionally,
homogeneous mortar resultedmixing operation,
well distributed the way
in the presence of iron
to produce particles reaction
the corrosion in a
the residues
homogeneous of
mortarmetallic iron formed
systemextensive
resulted well in the slag
distributed are corroded
in the by the NaOH solution during the
at isolated sites avoiding pores coalescence. Theway to produce
limited fractionthe ofcorrosion
the metallic reaction
form of
dissolution
at isolated phase of geopolymerization. The evolved gas is responsible for the formation of pores
the iron sites avoiding
justified extensive
the behavior pores coalescence.
observed for both B and TheWlimited
series, fraction
althoughthe of themicrographs
metallic form of B
of the
thewith
iron size between
justified the 1behavior
and 10 m for W series
observed for andB0.1
both andandW 10 m for
series, B series. micrographs of the B
althoughthe
series although this class of interpores demonstrated a high compactness into the interpores
series The homogeneous
although this class dispersion of pores from the exothermic reactions of iron corrosion is due to
partitions (Figures 11 andof12).
interpores demonstrated
The W series maintaineda the high compactness
global compactness intoof the
the interpores
conventional
the efficient
partitions mixing
(Figures operation of steel slag, metakaolin and sand particles for the preparation of the
IPC mortars as a11result
and 12). The W series
of dissolution of maintained the global
the major fraction compactness
of the slag to highly of thereactive
conventional
oligomer
IPCAAS paste.asJointly
mortars
species.
with
The aoligomers
result of the proper mixing
dissolution
participate in of the operation, the presence
major fraction
formation of newof the of iron
slag
phases: to particles
highly
NASH,
in a homogeneous
reactive oligomer
C(A)SH which
mortar system resulted well distributed in the way to produce the
species. The oligomers participate in the formation of new phases: NASH, C(A)SH corrosion reaction at isolated
which sites
avoiding extensive pores coalescence. The limited fraction of the metallic form of the iron justified
the behavior observed for both B and W series, althoughthe micrographs of the B series although this
class of interpores demonstrated a high compactness into the interpores partitions (Figures 11 and 12).
Materials 2016, 9, 410 15 of 18

The W series maintained the global compactness of the conventional IPC mortars as a result of
Materials 2016, 9, 410 15 of 18
dissolution
Materials 2016, 9,of410
the major fraction of the slag to highly reactive oligomer species. The oligomers 15 of 18
participate in the formation of new phases: NASH, C(A)SH
form, with the residues of non-reacted phases, a compact structure due to the adhesion bonds which form, with the residues of
form, with
non-reacted the residues
phases, a of
compact non-reacted
structure duephases,
to the a compact
adhesion structure
bonds due
capacity
capacity of these particles as well as that of the river sand. The steel slag dissolution allows the of to the
these adhesion
particles as bonds
well as
capacity ofriver 3+ ions
that of the
formation ofthese 3+ particles
Fesand.ionsThe asalso
steel
that well
slag as that of to
dissolution
participate the river
allows
the sand.
the The
formation
formation steel
of Feslag
of geopolymeric dissolution
that also allows
phases the
participate
with their
formation of Fe
to the formation
insertion into
3+ ions
theofnetwork thatstructure
geopolymeric also phases
participate
withto
of NASH the
their formation
orinsertion
C(A)SH intoofresulting
geopolymeric
the network phasesofwith
intostructure
ferrosialtes [14].their
NASH The
insertion
or into
C(A)SH the network
resulting intostructure of
ferrosialtes NASH
[14]. The or C(A)SH
similarity of Fe 3+
resulting
with Al 3+ ion,
into ferrosialtes
explains [14]. Fe
why The3+
similarity of Fe with Al ion, explains why Fe acts as network former that can occupied
3+ 3+ 3+ a
similarity
tetrahedric ofsite.
acts as network Fe3+
Bothwith
former Al
thethat
3+ ion, explains why Fe3+ acts as network former that can occupied a
can occupied
energy a tetrahedric
of the dissolution of thesite. Both the
metallic ironenergy
and the of potential
the dissolution of the
insertion of
tetrahedric
metallic
its iron
cations site.
into andBoth
the the energyinsertion
thegeopolymer
potential of the dissolution
network ofcan
its cations of the
explain themetallic
into the iron
geopolymer
strength and the potential
network
enhancement can insertion
observedexplain of
the
for the
its cations
strength into
enhancement the geopolymer
observed network
for the AAS can explain
mortars, the
mainlyin strength
the BSS
AAS mortars, mainlyin the BSS formulations that shows higher strength, although the considerableenhancement
formulations observed
that shows for the
higher
AAS mortars,
strength,
fraction ofalthoughmainlyin
capillary the
poresthebetween
BSS formulations
considerable fraction
1 and 10ofm that shows
capillary
(Figure higher
pores
7). strength,
between
The 1 andalthough
RO crystalline10 m the considerable
(Figure
particles 7).the
in Thesteel
RO
fraction of
crystalline capillary
particles pores
in the between
steel slag 1 and
refer to10 m (Figure
crystalline Fe 2+
7). The
bearing RO crystalline
mineral phases particles
as in the steel
ferronforsterite,
slag refer to crystalline Fe bearing mineral phases as ferronforsterite, augite, and others pyroxene
2+
slag refer
augite,
type. and toothers
crystalline Fe2+ bearing
pyroxene type. mineral phases as ferronforsterite, augite, and others pyroxene
type.

(a) (b)
(a) (b)
Figure 11. ESEM microstructure of fracture surface at high enlargements of the specimens with
11. ESEM
Figure 11.
Figure
white ABS: ESEM microstructure
(a) 400;
ofoffracture
microstructure
(b) 800 showing fracturesurface at high
surface
the relative
enlargements
at impact
poor high of the
enlargements ofspecimens with white
the specimens
of the low concentration of the with
iron
ABS:
white (a) 400; (b)
ABS: (a)the 800
400; showing the relative poor impact of the low concentration of
(b) 800 showing the relative poor impact of the low concentration of the iron
themetal
iron
metal during process.
during the process.
metal during the process.

(a) (b)
(a) (b)
Figure 12. ESEM microstructure of fracture surface showing the homogeneous distribution of the
Figurewith
pores 12. ESEM microstructure
size between 0.1 and of fracture surface showing the homogeneous irondistribution of the
the
Figure 12. ESEM microstructure of 10 m forming
fracture from
surface showingthethe
corrosion of the distribution
homogeneous metal of
during
the pores
pores with
process size
ofbetween between
black ABS 0.1 and 10 m forming from the corrosion of the iron metal during the
with size 0.1mortars: (a) 400;
and 10 m (b) from
forming 800 SS-IPC.
the corrosion of the iron metal during the process of
process of black ABS mortars: (a) 400; (b) 800 SS-IPC.
black ABS mortars: (a) 400; (b) 800 SS-IPC.
According to Lemougna et al. [23], a portion of these minerals, as augite, react with the alkaline
According
solution to formtonew
Lemougna et al.which
ferric sites [23], amay
portion of these
be located inminerals,
amorphous as augite,
gel-likereact with
phase. thesituation
This alkaline
solution to form new ferric sites which may be located in amorphous gel-like phase. This
may explain why the steel slag to metakaolin mass ratio did not appear as the significant parameter situation
may explain
governing why
the the steel slag to metakaolin
geopolymerization and strengthmass ratio did
behavior of not
theappear as the
steel slag significant
mortars. It is parameter
suggested
governing the geopolymerization and strength behavior of the steel slag mortars.
that in the formulations range adopted in this study, the increase of slag content reduces It is suggested
the
that in the formulations range adopted in this study, the increase of slag content reduces the
Materials 2016, 9, 410 16 of 18

According to Lemougna et al. [23], a portion of these minerals, as augite, react with the alkaline
solution to form new ferric sites which may be located in amorphous gel-like phase. This situation
may explain why the steel slag to metakaolin mass ratio did not appear as the significant parameter
governing the geopolymerization and strength behavior of the steel slag mortars. It is suggested
that in the formulations range adopted in this study, the increase of slag content reduces the alumina
oligomers formed during the dissolution process due to the reduction of metakaolinthat provides
the alumina oligomers inthe mortars. Thus a fully reacted and cured matrix of the steel slag based
inorganic polymer mortars can then be described as a continuous and homogeneous solid solution
(gel-like structure) comparable to that based on metakaolin or fly ash, with the particularity of the pore
network that can be controlled with mix-design and compensated with the strengthening effects of
the iron into the adhesive bonds. The presence of metakaolin and semi-crystalline silica sand has the
effects to increase the homogeneity and the compactness of the microstructure avoiding micro cracks
and unreacted alkali [17,21].

4. Discussion and Conclusions


Astutingsih and Liu [30] indicated that in alkali-activated slag, alkali ions (K+ , Na+ , etc.) are
located at a non-bridging oxygen site [3133]. The use of amorphous alumina and silica from
metakaolin and semi-crystalline silica sand generates structural matrices with high flexural strength
as the result of the cracks stabilization, reduction of the cumulative pore volume and improvement
of the microstructure. The strengthening mechanism is focused on the existence of both NASH
and CN(A)H suitable adhesive bonds to embed the various aggregates in the matrix. The metallic
particles content of the various mix acts at first through its corrosion in alkaline environment to enhance
the reactions heat thus densifying the inter-pores partitions although the porosity between 1 and 10 m
results from the iron activities. Fe3+ and Fe2+ from the corrosive reactions and the dissolution of
some RO minerals react with the alkaline solution to form new ferric sites which may be located at
amorphous gel-like phase [23].
Kotama et al. [34] correlated the moisture control capacity with pore size distribution makethe
difference between the gel pores and the capillary pores. They designed ideal construction materials
with pores distributed between those with size <0.1 m, considered as pores responsible for the
absorption, and capillary pores (0.1 m < < 10 m) responsible for desorption. The detailed analysis
of the pores in the AAS mortars show that the BSS series present 15.6 vol % for B60 and 17 vol %
for B90 and B120 of pores with size 0.1 m. For WSS series, apart from W90, the specimens present
pores having 60% with size <0.1 m. The analysis of moisture absorptiondesorption performance
indicates the potential absorption up to ~6% moisture in the context of maximum humidity. In general,
increasing the gel pores content ( < 0.1 m) will increase moisture absorption. Pores with size
between 0.1 and 10 m contribute to enhance the moisture desorption capacity. In fact, the moisture
absorbed was found to be desorbed almost totally within ~8 h. Less than 2.5 vol % moisture remains
in B60 and W60. The moisture percentage decreases below 0.5 vol % for B90, B120, W90 and W120.
It can be concluded that the presence of the steel slag improves the moisture control capacity leading
to the production of pore network that enable regulation of absorption and desorption within a short
range of time (Figure 6). The above described pore network allows: (i) no accumulation of moisture
within the bulk structure of the mortars and by the way high exchange capacity; and (ii) by using the
geopolymer process technology, the surface thickness described by Kotama et al. [34] can be achieved
with the molding design in relation to the viscosity of the alkali-activated pastes making the surface
thick enough to avoid the ingress of aggressive agents while improving the exchange rate between the
environment and the bulk.
The above-described formulations based on the steel slag with particular class of capillary porosity
have shown that alkali activation accompanied by an appropriate mix design is a procedure capable
of reusing a high amount of end-of-waste materials or byproducts as steel slag. Apart from the low
embodied energy, since the activation of slag and its consolidation required very low amount of
Materials 2016, 9, 410 17 of 18

energy with respect to the energy required for conventional fired bricks and cement clinker, these
AAS mortars also showed an interesting adsorption/desorption behavior that suggested their use as
coating material to maintain the stability of indoor relative humidity.
All of these characteristics are in line with the challenges of the EU Framework Programme
Horizon 2020 [35] for the eco-efficient construction and building materials research.

Acknowledgments: Authors wish to thank Magdalena Gualtieri from DIEF, University of Modena and
Reggio Emilia for the FT-IR characterization of the samples of this study and Chiara Ponzoni from the same
department for the evaluation of the reactive fraction of steel slag.
Author Contributions: Isabella Lancellotti, Vincenzo M. Sglavo, Cristina Leonelli, equally contributed to
this paper. Elie Kamseu did most of the experimental work, elaborated the data and wrote the manuscript.
Luca Modolo contributed for steel slag materials.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Myers, R.J.; Bernal, S.A.; Gehman, J.D.; van Deventer, J.S.J.; Provis, J.L. The role of Al in cross-Linking of
Alkali-Activated Slag cements. J. Am. Ceram. Soc. 2015, 98, 9961004. [CrossRef]
2. Han, F.; Zhang, Z.; Wang, D.; Yan, P. Hydration heat evolution and kinetics of blended cement containing
steel slag at different temperatures. Thermochim. Acta 2015, 605, 4351. [CrossRef]
3. Vassilev, S.V.; Vassileva, C.G. A new approach for the classification of coal fly ashes based on the origin,
composition, properties and behavior. Fuel 2007, 86, 14901512. [CrossRef]
4. Pontikes, Y.; Machiels, L.; Onisei, S.; Pandelaers, L.; Geysen, D.; Jones, P.T.; Blanpain, B. Slags with a high Al
and Fe content as precursors for inorganic polymers. Appl. Clay Sci. 2013, 73, 93102. [CrossRef]
5. Meja, J.M.; Rodrguez, E.; de Gutirrez, R.M.; Gallego, N. Preparation and characterization of a hybrid
alkaline binder based on a fly ash with no commercial value. J. Clean. Prod. 2015, 104, 346352. [CrossRef]
6. Molino, B.; De Vincenzo, A.; Ferone, C.; Messina, F.; Colangelo, F.; Cioffi, R. Recycling of clay sediments for
geopolymer binder production. A new perspective for reservoir management in the framework of Italian
legislation: The Occhito reservoir case study. Materials 2014, 7, 56035616. [CrossRef]
7. Sevelsted, T.F.; Skibsted, J. Carbonation of CSH and CASH samples studied by 13 C, 27 Al and 29 Si MAS
NMR spectroscopy. Cem. Concr. Res. 2015, 71, 5665. [CrossRef]
8. Yu, P.; Kirkpatrick, R.J.; Poe, B.; McMillan, F.P.; Cong, X. Structure of calcium silicate hydrate (CSH): near-,
mid-, and far-infrared spectroscopy. J. Am. Ceram. Soc. 1999, 82, 742748. [CrossRef]
9. Emery, J.J. Slag utilization in pavement construction. In Extending Aggregate Resources; ASTM Internation
Technical Publication: Washington, DC, USA, 1982.
10. Kourounis, S.; Tsivilis, S.; Tsakiridis, P.E.; Papadimitriou, G.D.; Tsibouki, Z. Properties and hydration of
blended cements with steel making slag. Cem. Concr. Res. 2007, 37, 815822. [CrossRef]
11. Shi, C. Characterizatics and cementitious properties of ladle slag fines from steel production. Cem. Concr. Res.
2002, 32, 459462. [CrossRef]
12. Ruiz-Santaquiteria, C.; Skibsted, J.; Fernndez-Jimnez, A.; Palomo, A. Alkaline solution/binder ratio as
a determining factor in the alkaline activation of aluminosilicates. Cem. Concr. Res. 2012, 42, 12421251.
[CrossRef]
13. Bignozzi, M.C.; Manzi, S.; Lancellotti, I.; Kamseu, E.; Barbieri, L.; Leonelli, C. Mix-design and characterization
of alkali activated materials based on metakaolin and ladle slag. Appl. Clay Sci. 2013, 73, 7885. [CrossRef]
14. Davidovidts, J. Geopolymer, Chemistry & Applications, 2nd ed.; The Geopolymer Institute: Saint-Quentin,
France, 2008; p. 508.
15. Kamseu, E.; Ponzoni, C.; Tippayasam, C.; Taurino, R.; Chaysuwan, D.; Bignozzi, M.C.; Barbieri, L.; Leonelli, C.
Influence of fine aggregates on the microstructure, porosity and chemico-mechanical stability of inorganic
polymer concretes. Constr. Build. Mater. 2015, 96, 473483. [CrossRef]
16. Alonso, S.; Palomo, A. Calorimetric study of alkaline activation of calcium hydroxidemetakaolin
solid-mixtures. Cem. Concr. Res. 2001, 31, 2530. [CrossRef]
17. Kamseu, E.; Cannio, M.; Obonyo, E.A.; Tobias, F.; Bignozzi, M.C.; Sglavo, V.M.; Leonelli, C. Metakaolin-based
inorganic polymer composite: Effects of fine aggregate composition and structure on porosity evolution,
microstructure and mechanical properties. Cem. Concr. Compos. 2014, 53, 258269. [CrossRef]
Materials 2016, 9, 410 18 of 18

18. Gan, L.; Wang, H.; Li, X.; Qi, Y.; Zhang, C. Strength activity index of air quenched basic oxygen furnace steel
slag. J. Iron Steel Int. 2015, 22, 219222. [CrossRef]
19. Lecomte, I.; Henrist, C.; Ligeois, M.; Maseri, F.; Rulmont, A.; Cloots, R. (Micro)-structural comparison
between geopolymers, alkali-activated slag cement and Portland cement. J. Eur. Ceram. Soc. 2006,
23, 37893797. [CrossRef]
20. Mozgawa, W.; Deja, J. Spectroscopic studies of alkaline activated slag geopolymers. J. Mol. Struct. 2009,
924926, 434441. [CrossRef]
21. Kamseu, E.; Bignozzi, M.C.; Melo, U.C.; Leonelli, C.; Sglavo, V.M. Design of inorganic polymer cements:
Effects of matrix strength strengthening on microstructure. Constr. Build. Mater. 2013, 38, 11351145.
[CrossRef]
22. Zhang, Z.; Provis, J.L.; Reid, A.; Wang, H. Fly ash-based geopolymers: The relationship between composition,
pore structure and efflorescence. Cem. Concr. Res. 2014, 64, 3041. [CrossRef]
23. Lemougna, P.N.; Mackenzie, K.J.D.; Jameson, G.N.L.; Rahier, H.; Melo, U.C. The role of iron in the formation
of inorganic polymers (geopolymers) from volcanic ash: A 57 Fe Mossbauer spectroscopy study. J. Mater. Sci.
2013, 48, 52805286. [CrossRef]
24. Kamseu, E.; Leonelli, C.; Perera, D.S.; Melo, U.C.; Lemougna, P.N. Investigation of volcanic ash based
geopolymers as potential building materials. Interceram 2009, 58, 136140.
25. Berodier, E.; Scrivener, K. Understanding the filler effect on the nucleation and growth of CSH.
J. Am. Ceram. Soc. 2014, 97, 37643773. [CrossRef]
26. Pyun, S.; Moon, S.-M. Corrosion mechanism of pure aluminum in aqueous alkaline solution. J. Solid State
Electrochem. 2000, 4, 267272. [CrossRef]
27. Kamseu, E.; Leonelli, C.; Chinje, M.U.F.; Perera, D.S.; Lemougna, P.N. Polysialate matrixes from Al-rich and
Si-rich metakaolins: Polycondensation and physicochemical properties. Interceram 2011, 60, 2531.
28. Aredes, F.G.M.; Camposa, T.M.B.; Machado, J.P.B.; Sakane, K.K.; Thima, G.P.; Brunelli, D.D. Effect of cure
temperature on the formation of metakaolinite-based geopolymer. Ceram. Int. 2015, 41, 73027311. [CrossRef]
29. Duxson, P.; Provis, J.L.; Lukey, G.C.; Mallicoat, S.W.; Kriven, W.M.; van Deventer, J.S.J. Understanding the
relationship between geopolymer composition, microstructure and mechanical properties. Colloids Surf. A
Physicochem. Eng. Aspects 2005, 269, 4758. [CrossRef]
30. Astitiningsihand, S.; Liu, Y. Geopolymerization of Australian slag with effective dissolution by
alkali. In Proceedings of the World Congress on Geopolymer, Saint-Quentin, France, 129 June 2005;
Davidovits, J., Ed.; The Geopolymer Institute: Saint-Quentin, France, 2005; pp. 6973.
31. Belitsky, I.V.; Sakata, A.; Goto, S. Kinetics of hydration of slag in the slag-alkaline cements. In Proceedings
of the 3rd Beijing International Symposium on Cement and Concrete, Beijing, China, 2730 October 1993;
International Academic Publishers: San Bernardino, CA, USA, 1993; pp. 10281031.
32. Malolepszy, J. Some aspects of Alkali-activated cementitious materials setting and hardening. In Proceedings
of the 3rd Beijing International Symposium on Cement and Concrete, Beijing, China, 2730 October 1993;
International Academic Publishers: San Bernardino, CA, USA, 1993; pp. 10431046.
33. Wang, S.D. The role of sodium during the hydration of alkali-activated slag. Adv. Cem. Res. 2000, 12, 6569.
[CrossRef]
34. Kotama, M.; Fukumizu, H.; Matsumoto, Y.; Toyama, M.; Yamamoto, K.; Endo, M. Moisture Control
Construction Material. U.S. Patent No. 6472 061 B2, 12 January 2002.
35. Pacheco-Torgal, F. Eco-efficient construction and building materials research under the EU Framework
Programme Horizon 2020. Constr. Build. Mater. 2014, 51, 151162. [CrossRef]

2016 by the authors; licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC-BY) license (http://creativecommons.org/licenses/by/4.0/).

You might also like