Mechanical and Durability Performance of Mortars With Fine Recycled

You might also like

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

Cement and Concrete Composites 105 (2020) 103420

Contents lists available at ScienceDirect

Cement and Concrete Composites


journal homepage: http://www.elsevier.com/locate/cemconcomp

Mechanical and durability performance of mortars with fine recycled


concrete aggregates and reactive magnesium oxide as partial
cement replacement
T. Gonçalves a, R.V. Silva b, *, J. de Brito b, J.M. Ferna
�ndez c, A.R. Esquinas d
a
Instituto Superior T�ecnico, Universidade de Lisboa, Av. Rovisco Pais, 1049-001, Lisboa, Portugal
b
CERIS, Instituto Superior T�ecnico, Universidade de Lisboa, Av. Rovisco Pais, 1049-001, Lisboa, Portugal
c
Department of Inorganic Chemistry and Chemical Engineering, Universidad de C�ordoba, Campus de Rabanales, Edificio Marie Curie, Planta Baja y 1a Planta, 14071,
C�ordoba, Spain
d
Department of Inorganic Chemistry and Chemical Engineering, Engineering School Science of Belmez, 14240, C�
ordoba, Spain

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

Keywords: In this paper, mortar specimens were produced using two types of commercially-available reactive MgO as
Mortar partial cement replacement (10%, 15% and 20%, by weight) and fine recycled concrete aggregate as siliceous
MgO sand substitute (50% and 100%, by volume). The specimens were subjected to thermogravimetric, energy
Recycled concrete aggregate
dispersive X-ray, differential thermal and powder X-ray diffraction analyses. The mechanical and durability
XRD
performance of all specimens was evaluated in terms of their flexural and compressive strength, dynamic
TGA
Shrinkage modulus of elasticity, water absorption by capillary action, carbonation, and shrinkage. The main results indicate
an overall decline in mechanical and durability-related performance with the use of both MgO and fine recycled
concrete aggregates, but enhanced shrinkage behaviour was observed in all MgO-containing specimens.

1. Introduction maximum of 5% of MgO according to EN 197–1 [4]). The reason for this
lies on the low reactivity of the resulting MgO (dead-burnt) [5], as a
It is widely known that the cement industry has one of the highest result of the high temperatures of the clinkerization process. The very
environmental impacts in the World, with about 5–7% of the total global slow rate of hydration of this type of MgO leads to its late expansion in
CO2 emissions [1]. Some of the solutions that have been typically the form of brucite (Mg(OH)2) causing a disruption in an already
adopted to decrease the environmental impact of concrete are the use of hardened and stable cementitious microstructure. The formation of Mg
alternative fuels and/or raw materials, improving the energy efficiency (OH)2 from the hydration reaction of MgO is as follows:
of the calcination process and the carbon-capture ability of cementitious
products by incorporating new types of binders and adopting new curing MgO þ H2O → Mg2þ þ 2OH → Mg(OH)2 (1)
technologies [2]. On the other hand, subjecting magnesium-based components (i.e.
In the light of the reactive magnesium oxide’s (MgO) participation in magnesite and dolomite) to relatively lower temperatures of
the hydration reactions of cement and strength development when used 700 � C–1000 � C yields light-burned MgO, which exhibits greater reac­
as an addition, as well as its improved carbon-capture ability when tivity and expansion at early ages [6–8]. Furthermore, it has been re­
compared to that of the cement’s products of hydration [3], its use in ported that the net CO2 emissions of its production process and the
cementitious products has been progressively considered as a techni­ enhanced carbon-capture ability during its life-cycle [3,9,10] are about
cally viable approach to further reduce the composite’s overall envi­ 70% lower than those of ordinary Portland cement [11].
ronmental impact and improve its performance from certain In parallel, there has been an increasing worldwide demand for
perspectives. However, the incorporation of magnesium-containing raw construction aggregates. Recent statistics indicated a probable increase
feed for the production of cement has been strictly limited (i.e. from 45 billion tonnes, in 2017, to 66 billion tonnes by 2025 [12]. It has

* Corresponding author.
E-mail addresses: tiago.miguel.cleto.goncalves@ist.utl.pt (T. Gonçalves), rui.v.silva@tecnico.ulisboa.pt (R.V. Silva), jb@civil.ist.utl.pt (J. de Brito), um1feroj@
uco.es (J.M. Fern�
andez), p52roesa@uco.es (A.R. Esquinas).

https://doi.org/10.1016/j.cemconcomp.2019.103420
Received 13 September 2018; Received in revised form 28 March 2019; Accepted 19 September 2019
Available online 20 September 2019
0958-9465/© 2019 Elsevier Ltd. All rights reserved.
T. Gonçalves et al. Cement and Concrete Composites 105 (2020) 103420

been estimated that the production of cementitious construction mate­ Table 1


rials is responsible for almost half of the consumption of aggregates Aggregates’ properties and test methods.
[13]. Therefore, there is considerable scope to further reduce the Property Test method
exploration of these natural resources and one of the most viable ap­
Particle size distribution EN 1015–1 [24]
proaches is through the use of the mineral fraction of construction and Bulk density EN 1097–3 [25]
demolition waste (CDW). It was estimated that about 35% of the total Water absorption EN 1097–6 [26]
amount of waste generated in all economic sectors within the European Particle density EN 1097–6 [26]
Union in 2014 are from construction and demolition activities [14].
The use of recycled aggregates in the production of mortars and
concrete has been thoroughly studied and discussed in the literature Table 2
[15] and some of the existing standards for the production of structural Aggregate’s characteristics.
concrete already include specifications for the use of the recycled ma­ Properties Fine sand Coarse sand RCA
terial’s coarse fraction (e.g. EN 206:2013 þ A1 [16]). Nevertheless,
Nominal size (d/D) 0/2 0/4 0/4
some drawbacks from the use of recycled aggregates can still be seen, Bulk density (g/cm3) 1.56 1.63 1.14
limiting their application in construction, namely in elements in ρa (g/cm3) 2.62 2.63 2.54
aggressive environments or that are prone to exhibit deformation due to ρrd (g/cm3) 2.60 2.61 2.14
shrinkage. The magnitude of this decline in performance is usually ρssd (g/cm3) 2.61 2.62 2.30
WA24 (%) 0.23 0.25 7.2
greater when using fine recycled aggregates, which present greater in­
ternal porosity and deformability when compared to the coarse aggre­ Sieve size (mm) Cumulative passing material (%)
gate fraction from the same source material [17,18]. 4.0 100
This study is a preliminary one belonging to a state-funded research 2.0 93.2
project, linking several construction companies within the sector. 1.0 77.4
0.5 50.9
Contrarily to the methodology of other studies, which use lower
0.25 15.6
replacement levels of cement with MgO, in this paper, relatively high 0.125 1.4
contents of 10%, 15% and 20%, by weight, of MgO were used thus 0.063 0.0
constituting one of the novelties. Additionally, such MgO contents also
have the objective of decreasing the environmental impacts of concrete,
hexametaphosphate) to prevent agglomeration of the fine material and
in the light of the notable carbon-capture ability of MgO systems.
potentiate air bubble release during the test. Since the use of aggregates
Furthermore, this component can reduce the cost of cement manufac­
with different particle size distributions can affect the compaction and
ture as it comes from a less energy-intensive process. Concerning the
mechanical performance of the resulting mortars, the RCA used in this
interaction between MgO and RCA, the products of hydration of light-
study were sieved to various sizes fractions and later recombined with
burned MgO exhibit a volume greater than that of the initial constitu­
the objective of producing an aggregate with an equivalent distribution
ents, thus resulting in an expansion of the cementitious matrix. A close
to that of the sand (fine and coarse fraction) as in Table 2.
control of this phenomenon may result in a mitigation of the composite’s
Fig. 1 presents the results of the XRD analysis to fine RCA. The main
deformation due to shrinkage [6]. Moreover, the reaction between of Mg
identified phases were quartz (International Centre for Diffraction Data -
(OH)2 and constituents based on amorphous silica can lead to the for­
ICDD # 33–1161 [50]), calcite (ICDD # 05–0586), microcline (ICCDD #
mation of magnesium silicate hydrates (i.e. MgO–SiO2–H2O or M-S-H)
19–0932) and portlandite (ICDD # 04–0733).
[19–22], which can further improve the composite’s strength develop­
According to the EDX results, around 13.7%, by weight, of the RCA is
ment. The combination of using MgO and incorporating fine recycled
composed of carbon. It is likely that crushing the original concrete
concrete aggregates (RCA), the interaction of which still needs to be
significantly increased the material’s surface area thus potentiating the
researched in depth, is expected to result in cementitious composites
carbonation of Ca(OH)2 into CaCO3. Furthermore, the notable presence
with reduced carbon footprint and improved shrinkage behaviour. Thus,
of calcite is also explained by the presence of limestone aggregates in the
this paper reports the results of an experimental investigation on the
original concrete. The XRD analysis of the siliceous fine and coarse sands
interaction between fine RCA (50% and 100% by volume of natural
used in this study showed that, as expected, the main phase was quartz
sand) and of the presence of two types of commercially available MgO as
(ICDD # 33–1161). Orthoclase (ICDD # 31–0966), albite (ICDD #
partial cement replacement (10%, 15% and 20% by weight of cement).
41–1480), and microcline (ICDD # 19–0926) were also observed.
The cementitious microstructure of the resulting composites was eval­
uated by means of thermogravimetric analysis (TGA), energy dispersive
X-ray analyses (EDX), differential thermal analysis (DTA) and powder
X-ray diffraction (XRD). The specimens’ performance from a
macro-level perspective was appraised in terms of their flexural and
compressive strengths, dynamic modulus of elasticity, water absorption
by capillary action, carbonation, and shrinkage.

2. Materials and methods

2.1. Characterization of aggregates

The RCA and sand were characterized according to the test methods
presented in Table 1. The physical properties and particle size distri­
bution of both fine and coarse sand (0/2 and 0/4 mm, respectively), and
fine RCA are presented in Table 2. The fine RCA were sourced from a
C30/37 concrete, of X0 exposure class, maximum aggregate size of
22 mm and S2 slump class. The RCA’s water absorption and density test
methods were adapted to the procedure presented by Rodrigues et al.
[23]. This method consisted of adding a surface active agent (i.e. sodium Fig. 1. XRD analysis of fine RCA.

2
T. Gonçalves et al. Cement and Concrete Composites 105 (2020) 103420

2.2. Cement and MgO

The chemical composition of the CEM I 42.5 R cement used in this


study is in Table 3. The cement presented initial and final setting times of
161 min and 232 min, respectively. The water used in the production of
all specimens was potable tap water. Two types of commercially avail­
able MgO were used: MgO A was supplied by Styromag, an Austrian
company; and MgO G from Grecian Magnesite S.A., a Greek company.
Their chemical compositions are presented in Table 3. Because of their
relatively low calcination temperature (~800 � C), the two types are
considered as light-burned MgO [7]. The particle size distributions of
cement and both MgO are presented in Fig. 2. The size distributions of
MgO A and cement were similar; most of their particles had a diameter
between 10 μm and 60 μm. Smaller particles were identified in MgO G;
mostly between 10 μm and 30 μm. This addition presented a less
extensive particle size distribution in comparison with MgO A. MgO A
and MgO G were found to have around 52% and 65%, respectively, of
particles within 3–32 μm. 20.2% and 3.5% of the particles of MgO A and
MgO G were above 32 μm, which may be too large to hydrate rapidly.
27.7% and 31.1% of the particles of MgO A and MgO G, respectively,
were below 3 μm and may have a greater water requirement and smaller Fig. 2. Particle size distribution by laser diffraction of MgO A, MgO G
contribution to the materials’ mechanical performance [27–29]. The and cement.
bulk density of cement, MgO A, MgO G and cement were 1040 kg/m3,
653 kg/m3 and 618 kg/m3, respectively.
The BET specific surface area of MgO A and MgO G were very similar
(20.12 m2/g and 18.01 m2/g, respectively). By means of the adsorption-
desorption isotherms of N2 it was possible to analyse the pore size dis­
tribution of the two MgO (Fig. 3). According to the BJH method, the
results showed that both additions exhibited a range of pore diameter of
4–200 nm. Even though both materials showed equivalent distribution
of pore volume, it was observed that the highest quantity of pore volume
of MgO A is centred at a slightly lower pore diameter than for MgO G
(23 nm vs. 33 nm, respectively).
The results of the XRD analysis for MgO A and MgO G are presented
in Fig. 4. The pattern of MgO A shows that the main phase was periclase
(ICDD # 45–0946), but there were also other peaks for calcite (ICDD #
05–0586), dolomite (ICDD # 36–0426) and talc (ICDD # 13–0558). The
XRD pattern of MgO G shows a periclase phase only, indicates lower
degree of impurity in comparison with MgO A.
TGA and DTA results showed that MgO G exhibited a significant
weight loss at temperatures between 300 � C and 400 � C, indicating
dehydroxylation of Mg(OH)2, which suggests partial hydration of the
Fig. 3. Pore volume of MgO A and MgO G.
addition (Fig. 5). The TGA and DTA profiles of MgO A showed a notable
peak at about 700 � C probably due to the decarbonation of calcite
(CaCO3) and dolomite (CaMg(CO3)2), the presence of which is the water content for the same binder content. The mortars’ mix code is
confirmed by the XRD patterns in Fig. 4. the following: M mortar; A and G - MgO A and MgO G, respectively; 0,
50 and 100 - replacement level of sand with fine RCA; 0, 10, 15 and 20 -
replacement level of cement with MgO.
2.3. Mix design

The mortars’ mix design was carried out in accordance with the 2.4. Material characterization
method of Nepomuceno et al. [30]. The mortars’ compositions are
presented in Table 4. The mortars’ constituents and the resulting specimens were analysed
All mortars were produced with an initial volumetric ratio of 1/3 by means of XRD patterns using a Bruker D8 Discover A25 instrument
(binder/aggregates). Cement was partially replaced with the two types with Cu-Kα radiation. Diffraction patterns were obtained by scanning
of MgO (mortars with MgO A and MgO G are MA and MG, respectively) the goniometer from 10� to 80� (2θ) at a rate of 0.05� ∙s 1. TGA was
with ratios of 10%, 15% and 20%, by weight. The sand was replaced by performed in a Setaram Setsys Evolution 16/18 apparatus at a heating
fine RCA with ratios of 50% and 100%, by volume. The consistence of all rate of 5 � C/min. The electron microprobe technique was conducted on
mortars was fixed at 200 � 15 mm. This range was achieved by altering an electron microscope JEOL JSM-6300 using an acceleration voltage of

Table 3
Chemical composition of cement and MgO (% by mass).
Component SiO2 Al2O3 Fe2O3 CaO MgO SO3 CuO ZnO C3S C2S C3A C4AF

CEM I 42.5 R 19.49 5.02 3.32 63.48 1.26 3.26 – – 57.70 16.50 4.30 11.20
MgO A 4.6 – 7.4 5.5 78.6 – 2.8 1.1 – – – –
MgO G – – – 4.4 89.5 – 3.4 2.6 – – – –

3
T. Gonçalves et al. Cement and Concrete Composites 105 (2020) 103420

1015–13 [38].

3. Results and discussion

3.1. Fresh performance

3.1.1. Consistence
The mortars in this study can be considered equivalent to the mortar
phase of concrete using the method proposed by Nepomuceno et al.
[30]. Thus, 200 � 15 mm is considered an acceptable range for target
consistence considering that it would decrease to a S3 slump class ac­
cording to Ref. [16] after the addition of the coarse aggregate fraction.
To maintain the spread within the target and achieve equivalent
compaction energy, as expected, more water content was required with
Fig. 4. XRD patterns for MgO A and MgO G. increasing MgO and RCA content. In the case of MgO, this is due to the
particles’ greater surface area in comparison with that of those of
cement, even though both components present similar particle size
distributions. Also, due to the RCAs’ greater water absorption capacity
when compared to natural sand, which has almost nil water absorption.
Moreover, since the particles of fine RCA also present greater roughness
and angularity, additional water is also needed to wet the increased
surface area and overcome the materials’ interparticle friction [17].
Although less noticeable, replacing the cement with MgO also led to
mortar with slightly worse workability, probably due to the greater
angularity that is typical of MgO particles [39] and thus additional water
content was required to achieve the same consistence. This lower
workability was more prominent in mortars made with MgO G possibly
as a result of the lower compatibility of the particle size distribution with
that of cement and lower particle size, which increased the need for
more water to keep the workability constant. Naturally, the increasing
use of both fine RCA and MgO led to an increased water requirement;
Fig. 5. TGA and DTA of MgO A and MgO G about 138% and 156% more water was required for mixes made with
100% fine RCA and MgO A and MgO G, respectively.
20 kV and a working distance of 15 mm. The X-ray detector was a model
Oxford Instruments ATW2-6699. Particle sizes were measured in a 3.1.2. Bulk density
Mastersizer S laser diffraction particle size analyser (Malvern In­ The results of the fresh density test are presented in Table 4. As ex­
struments) using ethanol as dispersant. pected, there was a decrease in the mortars’ density with increasing fine
Mortars were produced and cast in accordance to EN 1015–2 [31]. RCA content; an average value of 8% was observed for incorporation of
The consistence of fresh mortars was tested in accordance with EN 100% RCA. This is due to their lower density when compared to that of
1015–3 [32], bulk density with EN 1015–6 [33], dynamic Young’s sand. No significant changes in density were observed with increasing
modulus with ASTM E1876 [34], flexural and compressive strength with MgO content regardless of the type.
EN 1015–11 [35], carbonation with LNEC E391 [36], water absorption
by capillary action with EN 1015–18 [37] and shrinkage with EN

Table 4
Composition of mortars.
Mortar Cement (kg/m3) MgO (kg/m3) Water (kg/m3) Sand (kg/m3) RCA (kg/m3) w/b ratio Fresh density (kg/m3)

M-0-0 560.7 – 280.3 1406.9 – 0.50 2294


M-50-0 560.7 – 352.3 703.4 582.8 0.63 2217
M-100-0 560.7 – 498.3 – 1165.5 0.89 2114
MA-0-10 504.6 56.1 280.3 1406.9 0.50 2302
MA-0-15 476.6 84.1 294.3 1406.9 0.52 2289
MA-0-20 448.5 112.1 304.3 1406.9 0.54 2310
MA-50-10 504.6 56.1 376.3 703.4 582.8 0.67 2229
MA-50-15 476.6 84.1 392.3 703.4 582.8 0.70 2238
MA-50-20 448.5 112.1 454.3 703.4 582.8 0.81 2210
MA-100-10 504.6 56.1 574.3 – 1165.5 1.02 2114
MA-100-15 476.6 84.1 616.3 – 1165.5 1.10 2099
MA-100-20 448.5 112.1 668.3 – 1165.5 1.19 2116
MG-0-10 504.6 56.1 280.3 1406.9 0.50 2299
MG-0-15 476.6 84.1 342.3 1406.9 0.61 2290
MG-0-20 448.5 112.1 342.3 1406.9 0.61 2292
MG-50-10 504.6 56.1 400.3 703.4 582.8 0.71 2218
MG-50-15 476.6 84.1 466.3 703.4 582.8 0.83 2210
MG-50-20 448.5 112.1 504.3 703.4 582.8 0.90 2192
MG-100-10 504.6 56.1 602.3 – 1165.5 1.07 2091
MG-100-15 476.6 84.1 654.3 – 1165.5 1.17 2106
MG-100-20 448.5 112.1 716.3 – 1165.5 1.28 2102

4
T. Gonçalves et al. Cement and Concrete Composites 105 (2020) 103420

3.2. Physicochemical characterization of hardened mortars

Fig. 6 presents the results of TGA and DTA of specimens M-0-0, MA-
0-20 and MA-100-20. Up to 200 � C, the mass loss may be attributed to
the dehydration of C–S–H and AFt phases. Higher mass loss was
observed for MA-100-20 mixes because the dehydration of the fine
RCA’s adhered mortar. Even though the thermal decomposition of M-S-
H typically occurs in the same temperature range [40], it is likely that it
did not form since the presence of amorphous SiO2 would be necessary
to react with Mg(OH)2 [41]. Clear mass losses were observed between
300 � C and 400 � C for MA-0-20 and MA-100-20 (and not for M-0-0),
which may be due to the dehydroxylation of Mg(OH)2. All three speci­
mens presented equivalent mass losses between 400 � C and 500 � C due
to the dehydroxylation of Ca(OH)2. The final endothermic peak in the
range of 700–800 � C is attributed to the decarbonation of CaCO3 and
also MgCO3 for MgO-containing mixes. A considerably higher mass loss
can be observed in MA-100-20 specimens. After the concrete’s crushing
process, a higher surface area of the material becomes exposed to at­
mospheric CO2, thereby accelerating carbonation reactions with existing
Ca(OH)2. This ultimately leads to higher CaCO3 content in the resulting
mortar.
Fig. 7. XRD analysis of mortars.
Fig. 7 presents the results of the XRD analysis of M-0-0, MA-0-20 and
MA-100-20 specimens. The main mineral phase of all three patterns was
quartz (ICDD # 33–1161), which is a result of the incorporation of contribution to strength. Furthermore, since Mg(OH)2 is capable of
siliceous sand. Its intensity decreased in the MA-100-20 pattern because forming at the ITZ and because it presents preferential orientation upon
the sand was completely replaced with 100% fine RCA, which contains formation similar to Ca(OH)2 particles, this section may show decreased
siliceous sand to a much lesser extent. resistance and possibly micro-cracks due to excessive expansion.
The presence of portlandite (ICDD # 04–0733) can also be observed Though this is not the case, the strength decrease could be mitigated
in the three mortars. Calcite (ICDD # 05–0586) was also observed in M- with the addition of amorphous SiO2 and Al2O3 (e.g. fly ash) wherein
0-0 and MA-0-20, but with greater intensity in MA-100-20 mortars due improved strength development would be noticed due to pozzolanic
to the higher calcite content of fine RCA in comparison with natural reactions with Mg(OH)2 [20]. The result of this reaction would be M-S-H
sand. Orthoclase (ICDD # 31–0966), Ca3SiO5 (ICDD # 42–0551), and hydrotalcite-like phases, which are the main contributing factor for
microcline (ICDD # 19–0932), dolomite (ICDD # 36–0426), and peri­ the strength development of MgO-containing cementitious products [19,
clase (ICDD # 45–0946) were also observed. 43,44]. The decline in performance was more noticeable for MgO
G-containing mortars, possibly because MgO A exhibited an extensive
particle size distribution and similar to that of cement, thereby
3.3. Flexural and compressive strength improving the packing of specimens. Also, MgO G presented a smaller
particle size and higher surface area and thus produced mortars with
Fig. 8, Fig. 9 and Table 5 contain the results of the 28-day flexural greater water requirement.
and compressive strengths of all mixes. A general strength decrease can Concerning the influence of fine RCA, the results show that there is a
be observed with increasing replacement level of cement with MgO. This strength decrease with increasing replacement level as expected [45].
decrease in performance can be mostly explained by the greater water This is mostly due to the greater water requirement, which led to the
requirement of mixes with increasing MgO content and, of course, the production of mortars with a more porous cementitious microstructure
consequent dilution of cement [41,42]. The first factor, obviously, re­ apart from that imparted by the fine RCA themselves. Furthermore,
sults in a more porous microstructure and thus with lower resistance to since fine RCA are mostly comprised of old mortar, its mechanical per­
loading. In the second factor, since there is a lower quantity of cement, formance is poorer than that of siliceous sand from a microstructural
there is a decrease in the amount of strength-contributing C–S–H phases, level. The greater porosity of fine RCA leads to stress concentration
being replaced with Mg(OH)2 that does not have such an obvious around empty spaces, which result in crack formation and subsequent
propagation until failure.

3.4. Dynamic modulus of elasticity

The results of the dynamic modulus of elasticity test are presented in


Table 5. There is a clear decline in this property for mortars with
increasing fine RCA content; decreases of 30–40% were observed for
mixes with 100% fine RCA when compared to the corresponding control
mixes. This decrease is due to the increased porosity of the specimens as
a result of the incorporation of less dense RCA. These results are in line
with those of the literature [46]. Regarding the influence of MgO, there
was a slight decrease in the mortars’ modulus of elasticity with
increasing MgO A and MgO G contents (more so for the latter). This
decrease is most likely due to the greater water requirement and ne­
cessity to add more water to obtain equivalent workability levels. The
simultaneous incorporation of MgO and fine RCA resulted in a combi­
nation of effects; both resulted in a decrease of the dynamic modulus of
Fig. 6. TGA and DTA results of mortars. elasticity, the decline of which was mostly due to the presence of RCA.

5
T. Gonçalves et al. Cement and Concrete Composites 105 (2020) 103420

Fig. 8. Flexural strength of mortars with (a) MgO A and (b) MgO G.

Fig. 9. Compressive strength of mortars with (a) MgO A and (b) MgO G.

Table 5
Flexural (ffl) and compressive (fc) strength, and dynamic modulus of elasticity (Ec,dyn) of mortars.
Mix code 28-day ffl (MPa) 28-day fc (MPa) Ec,dyn (GPa) Mix code 28-day ffl (MPa) 28-day fc (MPa) Ec,dyn (GPa)

M-0-0 7.66 53.7 32.1 M-0-0 7.66 53.7 32.1


MA-0-10 7.17 43.2 31.5 MG-0-10 6.72 48.4 31.2
MA-0-15 7.07 46.9 31.6 MG-0-15 6.58 43.3 29.2
MA-0-20 5.62 42.3 29.0 MG-0-20 5.40 38.6 27.4
MA-50-0 7.66 54.0 27.7 MG-50-0 7.66 54.0 27.7
MA-50-10 6.49 52.8 27.9 MG-50-10 5.92 42.4 25.9
MA-50-15 6.78 46.3 26.0 MG-50-15 5.53 40.3 24.5
MA-50-20 5.97 33.9 24.7 MG-50-20 5.45 33.9 22.6
MA-100-0 6.04 44.4 21.2 MG-100-0 6.04 44.4 21.2
MA-100-10 6.21 44.1 21.5 MG-100-10 5.23 47.9 19.4
MA-100-15 5.72 39.8 19.4 MG-100-15 5.67 34.8 19.5
MA-100-20 5.39 35.8 20.0 MG-100-20 4.98 28.1 17.2

3.5. Water absorption by capillary action less interconnected porous network [20,47]. This was not the case in this
study probably due to the addition of extra water, which resulted in a
Fig. 10a and b present the water absorption by capillarity of mortars more porous microstructure. As expected, the incorporation of RCA led
with MgO A and MgO G, respectively. Laboratory problems stopped to higher sorptivity as a result of the greater porosity imparted by the
measurements of mixes MG-50-0 and MG-100-0. The results show that fine RCA and that of the microstructure as a result of the addition of
there was a slight increase in sorptivity with increasing MgO content in extra water.
most mixes.
The presence of Mg(OH)2, which exhibits higher volume than the
initial constituents, would be expected to produce a more tortuous and

6
T. Gonçalves et al. Cement and Concrete Composites 105 (2020) 103420

Fig. 10. Water absorption by capillary action after 72 h of mortars with (a) MgO A and (b) MgO G.

3.6. Carbonation Interestingly, the combination of both fine RCA and MgO led to a
mitigation of the mortars’ carbonation rate; while mixes without RCA
Fig. 11 and Table 6 present the carbonation depth of specimens presented average increases in carbonation depths of 1.2 mm, 2.6 mm
containing increasing MgO and fine RCA content, after being exposed and 4.1 mm after adding 10%, 15% and 20% MgO A, respectively, for
for 28 and 91 days in a chamber to 5 � 0.1% CO2, 60 � 5% RH at a mixes with 100% RCA, these increases were of 1.0 mm, 1.7 mm and
temperature of 23 � 3 � C. It is immediately clear that the incorporation 2.3 mm, respectively. A similar trend was observed for MgO G-con­
of fine RCA as sand replacement led to a significant increase of mortars’ taining specimens. It is still unclear what caused this reduction in
carbonation as a result of the microstructure’s greater porosity; apart carbonation depth considering the considerably high w/b ratio of mixes
from that induced by the increased water requirement, the porous containing both 100% RCA and 20% MgO and that this trend was not
adhered mortar of RCA resulted in greater interconnectivity of the observed in the water absorption by capillary action tests.
porous network thus facilitating the diffusion of CO2.
Concerning the effect of MgO incorporation, the results showed that
the carbonation depths increased with increasing content of the addi­ 3.7. Shrinkage
tion. This was expected for a number of reasons. Firstly, the addition of
extra water led to a more porous matrix in comparison to that of the Fig. 12 presents the shrinkage strain of mortars with increasing RCA
control specimens, which means that there was a greater surface area and MgO A contents. As expected, the shrinkage strains increased
within the cementitious microstructure available to carbonate in the significantly with increasing RCA content; incorporation of 50% and
presence of CO2. Apart from this, replacing the cement with MgO led to 100% RCA led to increases of about 50% and 110%. This is due to the
the formation of a lower quantity of C–S–H and Ca(OH)2 phases, the lower stiffness of the fine RCA’s adhered mortar, which resulted in lower
presence and interaction of which would decrease the progression of ability to restrain the shrinkage of the new cementitious matrix. On the
carbonation. Moreover, one should also bear in mind the mortars’ pH other hand, the use of MgO as cement replacement led to notable de­
level as a result of the MgO’s products of hydration and its influence on creases in shrinkage; replacing the cement with 15% MgO A (considered
the phenolphthalein test. The decreased quantity of Ca(OH)2, which as the optimum content for this property) resulted in decreases of
exhibits a pH level of ~12.5, and a higher amount of Mg(OH)2, with a 400–600 μm/m in 91-day shrinkage measurements when compared to
pH of ~10.5, means that the overall pH level lowers, thereby showing those of M-0-0. Although the use of MgO G also resulted in lower
increased rate of carbonation according to the phenolphthalein test. shrinkage readings, it was not as effective as MgO A; the optimum
content of MgO G was of 10% and resulted in reduction in shrinkage of

Fig. 11. 91-day carbonation depths of mortars with increasing RCA and MgO content.

7
T. Gonçalves et al. Cement and Concrete Composites 105 (2020) 103420

Table 6
Carbonation depths of mortars.
Age 28 days 91 days Age 28 days 91 days

Mix code Av. (mm) Std. Dev. (mm) Av. (mm) Std. Dev. (mm) Mix code Av. (mm) Std. Dev. (mm) Av. (mm) Std. Dev. (mm)

M-0-0 1.32 1.44 1.76 1.47 M-0-0 1.32 1.44 1.76 1.47
MA-0-10 1.22 1.31 3.00 1.53 MG-0-10 1.21 1.14 2.04 1.07
MA-0-15 1.39 1.21 4.38 0.88 MG-0-15 1.56 1.52 4.13 1.38
MA-0-20 2.39 1.93 5.83 1.20 MG-0-20 2.74 1.82 – –
MA-50-0 0.68 0.85 2.86 0.95 MG-50-0 0.68 0.85 2.86 0.95
MA-50-10 1.53 0.56 3.25 1.11 MG-50-10 1.86 0.68 3.64 1.20
MA-50-15 2.74 1.02 5.18 1.04 MG-50-15 2.60 1.37 4.17 1.21
MA-50-20 3.54 1.40 5.99 1.91 MG-50-20 2.81 1.27 5.49 1.67
MA-100-0 1.21 1.06 3.92 1.07 MG-100-0 1.21 1.06 3.92 1.07
MA-100-10 2.74 1.41 4.92 1.59 MG-100-10 3.22 1.30 4.72 1.33
MA-100-15 2.85 0.90 5.58 1.48 MG-100-15 3.50 1.23 5.56 1.82
MA-100-20 – – 6.25 1.52 MG-100-20 4.64 1.30 5.82 1.23

Fig. 12. Shrinkage of mortars with (a) 0%, (b) 50% and (c) 100% RCA.

around 35% (Table 7). The lower effectiveness of MgO G may be cement past and aggregate. Since it may present preferential orientation
explained by its partial hydration before its use and thus did not upon formation, this section may weaken and, when subjected to
contribute to the specimens’ expansion as effectively with MgO A. Also, compression, it may become a failure zone due to its excessive expan­
MgO G-containing mortars required more water, thus increasing the sion, thus causing micro-cracks [50]. However, it was observed other­
specimens’ open porosity and decreasing their stiffness. wise by means of SEM analysis of mixes with up to 10% MgO content
The reduction of the mortar’s overall shrinkage can be explained by [51]. The expansion of brucite, homogenously distributed in the mix, is
the fact that Mg(OH)2 presents a greater volume than that of its initial capable of offsetting part of the cement matrix’s initial shrinkage. After
reagents. This leads to the specimens’ expansion at an initial stage of this expansion (reaching up to about 250 μm/m for MA-0-15 speci­
their curing process (i.e. 2–3 days after casting) prompted by the rapid mens), a distinctive drying shrinkage behaviour was observed in all
hydration of light-burned MgO particles [48,49]. The hydrated product specimens. Regarding the combination of both materials, it was
of these particles tends to nucleate and grow at the ITZ between the perceived that the considerable shrinkage normally observed in recycled

8
T. Gonçalves et al. Cement and Concrete Composites 105 (2020) 103420

Table 7 yield mortars with greater porous interconnectivity and surface area
91-day shrinkage readings of mortars with MgO A and MgO G. that can react with atmospheric CO2;
Mix code 91-day εs Δ (μm/ Mix code 91-day εs Δ (μm/ � The products of hydration of MgO resulted in early-age expansion,
(μm/m) m) (μm/m) m) which led to an overall reduction of shrinkage of specimens con­
M-0-0 439 – M-0-0 439 – taining the addition. The use of MgO as partial cement replacement is
MA-0-10 240 199 MG-0-10 362 77 especially appealing in the case of cementitious products containing
MA-0-15 55 384 MG-0-15 377 62 recycled aggregates, because these usually result in considerable
MA-0-20 40 400 MG-0-20 654 214 increases in shrinkage strain. Indeed, the combination of 100% fine
MA-50-0 783 MG-50-0 783
RCA with 15–20% of one of the types of light-burned MgO used in
– –
MA-50- 543 240 MG-50- 514 269
10 10 this study led to a decrease in shrinkage strain of about 600 μm/m
MA-50- 314 470 MG-50- 624 159 after 91 days. The resulting material exhibited a shrinkage behaviour
15 15 comparable to that of the control made with natural sand. These
MA-50- 446 337 MG-50- 590 193
findings demonstrate the technical viability of the use of both fine
20 20
MA-100- 987 – MG-100- 987 – RCA and MgO, and open the possibility of using recycled aggregates
0 0 in elements capable of exhibiting notable deformation due to
MA-100- 754 233 MG-100- 647 340 shrinkage.
10 10
MA-100- 389 598 MG-100- 738 249
15 15
Declarations of interest
MA-100- 372 615 MG-100- 984 3
20 20 None.

Acknowledgements
aggregate mortars and concrete can be mitigated with the incorporation
of MgO. The use of 10%, 15% and 20% MgO A by weight of cement led
The authors gratefully acknowledge the support of the Civil Engi­
to shrinkage reductions of 233 μm/m, 598 μm/m and 615 μm/m,
neering Research and Innovation for Sustainability, IST - University of
respectively, when compared to MA-100-0. The 91-day shrinkage strain
Lisbon and FCT - Foundation for Science and Technology. This work was
of MA-100-15 and MA-100-20 became comparable to that of the control
partly supported by the Andalusian Regional Government (Research
mix without RCA and MgO.
Groups FQM-391) and MECD-Spain FPU-13/04030 (http://www.mecd.
gob.es/educacion-mecd/). The authors also wish to thank Styromag,
4. Conclusions
Grecian Magnesite S.A., Research Plan of the University of Co �rdoba
(2016), the staff at the Electron Microscopy and Elemental Analysis
This paper presented an experimental investigation on the interac­
units of the Central Research Support Service (SCAI) of University of
tion between MgO, as partial cement replacement, and fine RCA, as sand
Co
�rdoba for their technical assistance, and the Fine Chemistry Institute
substitute, and its influence on the properties of cementitious products
of the University of Co
�rdoba for their support.
equivalent to the mortar phase of concrete. The following conclusions
were drawn:
References

� The particle size distribution of reactive MgO already commercially [1] E. Benhelal, G. Zahedi, E. Shamsaei, A. Bahadori, Global strategies and potentials
available presents is comparable to that of cement. This excludes the to curb CO2 emissions in cement industry, J. Clean. Prod. 51 (Supplement C)
(2013) 142–161.
necessity for further grinding and milling processing stages of light-
[2] M. Schneider, M. Romer, M. Tschudin, H. Bolio, Sustainable cement
burned MgO after its manufacture, in order to make it suitable for the production—present and future, Cement Concr. Res. 41 (7) (2011) 642–650.
production of mortar and concrete; [3] C. Unluer, A. Al-Tabbaa, Impact of hydrated magnesium carbonate additives on the
� The results of the thermogravimetric analysis showed that Mg(OH)2 carbonation of reactive MgO cements, Cement Concr. Res. 54 (Supplement C)
(2013) 87–97.
has marginal reaction with the siliceous sand, cement’s products of [4] EN-197-1, Cement - Part 1: Composition, Specifications and Conformity Criteria for
hydration and fine RCA, and thus would require the presence of Common Cements, Comit�e Europ�een de Normalisation (CEN), Brussels, Belgium,
amorphous silica and alumina to be able to form the aforementioned 2011, p. 50.
[5] P. Mehta, History and status of performance tests for evaluation of soundness of
compounds; cements, in: Cement Standards - Evolution and Trends, Symposia Papers & STPs,
� Lower flexural and compressive strengths can be expected from the STP663, ASTM International, 1978.
incorporation of MgO as partial cement replacement due to the [6] L. Mo, M. Deng, M. Tang, Effects of calcination condition on expansion property of
MgO-type expansive agent used in cement-based materials, Cement Concr. Res. 40
dilution of cement and thus formation of lower amount of its prod­ (3) (2010) 437–446.
ucts of hydration. This, allied with the presence of fine RCA, which [7] F. Jin, A. Al-Tabbaa, Strength and hydration products of reactive MgO-silica pastes,
are known to have a lower mechanical performance in comparison Cement Concr. Compos. 52 (2014) 27–33.
[8] C. Du, A review of magnesium oxide in concrete, Concr. Int. 27 (12) (2005) 45–50.
with that of natural sand, resulted in the formation of a more porous
[9] M. Liska, A. Al-Tabbaa, Ultra-green construction: reactive magnesia masonry
and less resistant microstructure; products, Proc. Inst. Civ. Eng. Waste Res. Manag. 162 (4) (2009) 185–196.
� A slight decrease in the dynamic modulus of elasticity can be ex­ [10] C. Unluer, A. Al-Tabbaa, Enhancing the carbonation of MgO cement porous blocks
through improved curing conditions, Cement Concr. Res. 59 (Supplement C)
pected from the inclusion of reactive MgO, assuming that additional
(2014) 55–65.
water was required to attain a specific target consistence. This [11] S. Ruan, C. Unluer, Comparative life cycle assessment of reactive MgO and
decrease is even greater with the incorporation of fine RCA, which Portland cement production, J. Clean. Prod. 137 (Supplement C) (2016) 258–273.
are considerably more porous and deformable when compared to [12] PMR, Global Market Study on Construction Aggregates: Crushed Stone Product
Type Segment Projected to Register High Value and Volume CAGR during 2017 -
natural sand. The greater porosity of the cementitious microstructure 2025, Persistence Market Research, New York, USA, 2017, 235 pp.
affects both water absorption by capillary action and carbonation. [13] EAA, A Sustainable Industry for a Sustainable Europe - Annual Review 2016-2017,
The replacement of cement with MgO results in increased porosity, European Aggregates Association (EAA), Brussels, Belgium, 2018, 32 pp.
[14] Eurostat, Waste Statistics in Europe, 2017 available at: epp.eurostat.ec.europa.eu
thereby affecting both carbonation and water absorption by capillary (last accessed July 2017).
action. A greater influence on the second property is likely, as Mg [15] J. de Brito, R.V. Silva, Current status on the use of recycled aggregates in concrete:
(OH)2 presents a lower pH level than that of Ca(OH)2, and fine RCA where do we go from here? RILEM Tech. Lett. 1 (2016) 1–5.

9
T. Gonçalves et al. Cement and Concrete Composites 105 (2020) 103420

[16] EN-206:2013þA1, Concrete - Specification, Performance, Production and [35] EN-1015-11, Methods of Test for Mortar for Masonry - Part 11: Determination of
Conformity, Comit�e Europ� een de Normalisation (CEN), Brussels, Belgium, 2016, 98 Flexural and Compressive Strength of Hardened Mortar, Comit�e Europ� een de
pp. Normalisation (CEN), Brussels, Belgium, 1999, 12 pp.
[17] R.V. Silva, J. de Brito, R.K. Dhir, Properties and composition of recycled aggregates [36] LNEC-E391, Concrete: Determination of Carbonation Resistance (In Portuguese),
from construction and demolition waste suitable for concrete production, Constr. National Laboratory in Civil Engineering (LNEC - Laborat� orio Nacional de
Build. Mater. 65 (2014) 201–217. Engenharia Civil), Lisbon, Portugal, 1993, 2 pp.
[18] L. Evangelista, J. de Brito, Concrete with fine recycled aggregates: a review, Eur. J. [37] EN-1015-18, Methods of Test for Mortar for Masonry - Part 18: Determination of
Environ. Civ. Eng. 18 (2) (2014) 129–172. Water Absorption Coefficient Due to Capillary Action of Hardened Mortar, Comit� e
[19] D.R.M. Brew, F.P. Glasser, Synthesis and characterisation of magnesium silicate Europ�een de Normalisation (CEN), Brussels, Belgium, 2002, 12 pp.
hydrate gels, Cement Concr. Res. 35 (1) (2005) 85–98. [38] EN-1015-13, Methods of Test for Mortar for Masonry - Part 13: Determination of
[20] S.W. Choi, B.S. Jang, J.H. Kim, K.M. Lee, Durability characteristics of fly ash Dimensional Stability of Hardened Mortars, Comit� e Europ�
een de Normalisation
concrete containing lightly-burnt MgO, Constr. Build. Mater. 58 (2014) 77–84. (CEN), Brussels, Belgium, 1993, 20 pp.
[21] F. Jin, A. Al-Tabbaa, Evaluation of novel reactive MgO activated slag binder for the [39] H.M. Tran, A. Scott, Strength and workability of magnesium silicate hydrate binder
immobilisation of lead and zinc, Chemosphere 117 (2014) 285–294. systems, Constr. Build. Mater. 131 (2017) 526–535.
[22] D. Nied, K. Enemark-Rasmussen, E. L’Hopital, J. Skibsted, B. Lothenbach, [40] T. Zhang, L.J. Vandeperre, C.R. Cheeseman, Formation of magnesium silicate
Properties of magnesium silicate hydrates (M-S-H), Cement Concr. Res. 79 hydrate (M-S-H) cement pastes using sodium hexametaphosphate, Cement Concr.
(Supplement C) (2016) 323–332. Res. 65 (Supplement C) (2014) 8–14.
[23] F. Rodrigues, L. Evangelista, J. de Brito, A new method to determine the density [41] L.J. Vandeperre, M. Liska, A. Al-Tabbaa, Microstructures of reactive magnesia
and water absorption of fine recycled aggregates, Mater. Res. 16 (2013) cement blends, Cement Concr. Compos. 30 (8) (2008) 706–714.
1045–1051. [42] L. Mo, M. Liu, A. Al-Tabbaa, M. Deng, Deformation and mechanical properties of
[24] EN-1015-1, Methods of Test for Mortar for Masonry - Part 1: Determination of the expansive cements produced by inter-grinding cement clinker and MgOs with
Particle Size Distribution (By Sieve Analysis), Comit�e Europ�een de Normalisation various reactivities, Constr. Build. Mater. 80 (Supplement C) (2015) 1–8.
(CEN), Brussels, Belgium, 1999, 8 pp. [43] G. Ye, T. Troczynski, Hydration of hydratable alumina in the presence of various
[25] EN-1097-3, Tests for Mechanical and Physical Properties of Aggregates - Part 3: forms of MgO, Ceram. Int. 32 (3) (2006) 257–262.
Determination of Loose Bulk Density and Voids, Comit� e Europ�een de [44] M.B. Haha, B. Lothenbach, G. Le Saout, F. Winnefeld, Influence of slag chemistry
Normalisation (CEN), Brussels, Belgium, 1998, 10 pp. on the hydration of alkali-activated blast-furnace slag - Part I: effect of MgO,
[26] EN-1097-5, Tests for Mechanical and Physical Properties of Aggregates - Part 5: Cement Concr. Res. 41 (9) (2011) 955–963.
Determination of the Water Content by Drying in a Ventilated Oven, Comit� e [45] R.V. Silva, J. de Brito, R.K. Dhir, Performance of cementitious renderings and
Europ�een de Normalisation (CEN), Brussels, Belgium, 2008, 54 pp. masonry mortars containing recycled aggregates from construction and demolition
[27] S. Tsivilis, S. Tsimas, A. Benetatou, E. Haniotakis, Study on the contribution of the wastes, Constr. Build. Mater. 105 (2016) 400–415.
fineness on cement strength, Zement-Kalk-Gips 43 (1) (1990) 26–29. [46] R.V. Silva, J. de Brito, R.K. Dhir, Establishing a relationship between modulus of
[28] J.M. Pommersheim, Effect of particle size distribution on hydration kinetics, MRS elasticity and compressive strength of recycled aggregate concrete, J. Clean. Prod.
Proc. 85 (2011) 301. 112 (2016) 2171–2186.
[29] D.P. Bentz, C.J. Haecker, An argument for using coarse cements in high- [47] L. Mo, D.K. Panesar, Effects of accelerated carbonation on the microstructure of
performance concretes, Cement Concr. Res. 29 (4) (1999) 615–618. Portland cement pastes containing reactive MgO, Cement Concr. Res. 42 (6) (2012)
[30] M. Nepomuceno, L. Oliveira, S.M.R. Lopes, Methodology for mix design of the 769–777.
mortar phase of self-compacting concrete using different mineral additions in [48] F. Jin, K. Gu, A. Abdollahzadeh, A. Al-Tabbaa, Effects of different reactive MgOs on
binary blends of powders, Constr. Build. Mater. 26 (1) (2012) 317–326. the hydration of MgO-activated GGBS paste, J. Mater. Civ. Eng. 27 (7) (2015).
[31] EN-1015-2, Methods of Test for Mortar for Masonry - Part 2: Bulk Sampling of [49] R. Polat, R. Demirbo� ga, F. Karag€ol, The effect of nano-MgO on the setting time,
Mortars and Preparation of Test Mortars, Comit�e Europ�een de Normalisation autogenous shrinkage, microstructure and mechanical properties of high
(CEN), Brussels, Belgium, 1999, 8 pp. performance cement paste and mortar, Constr. Build. Mater. 156 (2017) 208–218.
[32] EN-1015-3, Methods of Test for Mortar for Masonry - Part 3: Determination of [50] L.Q. Yu, M. Deng, L.W. Mo, J.X. Liu, F.F. Jiang, Effects of lightly burnt MgO
Consistence of Fresh Mortar (By Flow Table), Comit� e Europ�een de Normalisation expansive agent on the deformation and microstructure of reinforced concrete
(CEN), Brussels, Belgium, 1999, 10 pp. wall, Ann. Mater. Sci. Eng. (2019), 1948123, https://doi.org/10.1155/2019/
[33] EN-1015-6, Methods of Test for Mortar for Masonry - Part 6: Determination of Bulk 1948123, 9 pp.
Density of Fresh Mortar, Comit�e Europ�een de Normalisation (CEN), Brussels, [51] K.J. Huang, X.J. Shi, D. Zollinger, M. Mirsayar, A.G. Wang, L.W. Mo, Use of MgO
Belgium, 1999, 8 pp. expansion agent to compensate concrete shrinkage in jointed reinforced concrete
[34] ASTM-E1876, Standard Test Method for Dynamic Young’s Modulus, Shear pavement under high-altitude environmental conditions, Constr. Build. Mater. 202
Modulus, and Poisson’s Ratio by Impulse Excitation of Vibration, American Society (2019) 528–536.
for Testing and Materials, USA, 2015, 17 pp.

10

You might also like