You are on page 1of 8

Construction and Building Materials 149 (2017) 378–385

Contents lists available at ScienceDirect

Construction and Building Materials


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

Enhanced properties of cement mortars with multilayer graphene


nanoparticles
Rodrigo Alves e Silva a, Paulo de Castro Guetti a, Mário Sérgio da Luz a, Francisco Rouxinol b,
Rogério Valentim Gelamo a,⇑
a
Federal University of Triângulo Mineiro, Dr. Randolfo Borges Júnior Ave., 1250, Uberaba, MG, Brazil
b
Campinas State University – Unicamp, Gleb Wataghin Institute of Physics, Sérgio Buarque de Holanda St, 777 – Cidade Universitária Zeferino Vaz, Barão Geraldo,
13083-859 Campinas, SP, Brazil

h i g h l i g h t s

 Incorporating MLG into OPC mortars substantially increase their strength.


 Thermal stresses are reduced and hydration reactions increase within the composite matrix.
 Dispersing MLG with isopropanol prevents the flakes from agglomerating very efficiently.
 The addition of MLG has relatively low costs and may be extended to a large scale.

a r t i c l e i n f o a b s t r a c t

Article history: The present work aims to combine multilayer graphene (MLG) nanoparticles with Ordinary Portland
Received 24 February 2017 Cement (OPC) mortar specimens, evaluating possible improvements on their mechanical properties
Received in revised form 15 May 2017 and ultimately coming up with an ideal and effective concentration for future applications. Mortars with
Accepted 17 May 2017
water/cement ratio of 0.4, MLG dosage varying from 0.015 to 0.033% by weight of cement and sand at a
proportion of 3 the weight of cement were prepared. 50  100 mm cylindrical specimens were shaped
and subjected to compressive and splitting tensile strength tests at the ages of 3, 7 and 28 days. In addi-
Keywords:
tion, SEM micrographs and metallographic images were collected and analyzed in order to better char-
Multilayer graphene
Cement mortars
acterize the sample’s morphology and also in attempt of explaining the MLG influence on cement
Mechanical properties paste. The optimal tensile strength was achieved with samples designed with MLG dosage of 0.033%,
Hydration heat for which the corresponding increases were 100.0, 144.4 and 131.6%, after 3, 7 and 28 days, respectively,
compared with samples without MLG. On the other hand, the optimal compressive strength increases
were obtained with samples designed with MLG dosage of 0.021%, with improvements of 63.6, 94.1
and 95.7% after 3, 7 and 28 days, respectively, compared with samples without MLG. The addition of
MLG is believed to accelerate cement hydration reactions, reduce the pore volume and harden cement
properties. An improvement in thermal conductivity is also associated to MLG incorporation, and this
favorable heat transfer in OPC-based mortars probably promoted the observed changes in the composite
chemical structure and mechanical properties herein analyzed.
Ó 2017 Elsevier Ltd. All rights reserved.

1. Introduction ing the properties of well-known materials [1], for exploring their
fundamental influence at microscopic and macroscopic levels [3,4],
Developments in nanotechnology over the last years have and for the development of new materials [5,6]. Indeed, nanoscale
begun to demonstrate the potential applications of nanoscale fil- fillers are now being incorporated to different materials and com-
lers in composites material [1,2]. These developments have been posites to specific applications [7–9]. Additionally, those recent
driven by the prospects of using those nanoscale fillers for modify- developments have also stimulated new ideas for processing tech-
niques in order to improve the properties of many conventional
materials significantly.
⇑ Corresponding author. Over the last decades, many studies have been focusing on
E-mail address: rogelamo@gmail.com (R.V. Gelamo). enhancing the properties of Ordinary Portland Cement (OPC), the

http://dx.doi.org/10.1016/j.conbuildmat.2017.05.146
0950-0618/Ó 2017 Elsevier Ltd. All rights reserved.
R.A. e Silva et al. / Construction and Building Materials 149 (2017) 378–385 379

main ingredient of concrete and mortar, by combining their Fig. 1(a) shows a typical high oriented natural crystalline graphite flake used to
synthetize multilayer graphene sheets (MLG) powder. Large natural graphite flakes,
ceramic matrix with different nanomaterials. In those nanocom-
provided by Nacional de Grafite Ltda. (Brazil), were submitted to an acid bath: sul-
posites, nano-sized particles and fibers, such as nanosilica [10], furic and nitric acids, followed by rapid heat treatment at high temperature for
nanotitanium oxide [11], nanoalumina [12,13] and carbon nan- some seconds, in order to obtain an expanded graphite structure (EGS) [41,42].
otubes [14], could be used for engineering reinforcements and To exfoliate the graphite structure, analytical grade isopropanol alcohol was
for enhancing a wide range of OPC-based materials properties blended with the EGS in a 1-to-1 ratio (ml-to-mg) and the solution was further
ultra-sonicated for 2 h. The combination of MLG and isopropanol yielded excellent
[11–16]. Because of the unique mechanical, electrical, optical and
dispersion conditions to prevent MLG from agglomerating and was directly poured
catalytic properties of nanoparticles, the interaction between the into the dry sand. Fig. 1(b) displays a SEM image of a typical flake present in the
nanoscale filler and the ceramic matrix material could offer the MLG powder used to prepare the OPC nanocomposites reported in this work.
possibility of modifying cement chemical and physical properties, Scanning electron microscopy was conducted to characterize the morphology of
the samples. In order to improve the imaging, the samples were covered by a thin
creating new mechanisms of interaction [17] that could be useful
gold film. All images were obtained in the secondary electron mode, using an elec-
agents on self-cleaning processes [11] and procedures associated tron beam of 5–10 kV and 0.4–2.1 nA. Metallographic characterizations were also
with abrasion resistance [15,16]. It has been anticipated that the made with the aid of a microscope from Inova Optical Systems. Atomic force micro-
application of nanocomposites within the civil construction indus- scope Shimadzu SPM9700 was used in dynamic mode to get the morphology of an
try could decrease the impacts of industrial activities over the envi- isolated MLG flake representative of our samples.
To obtain the Raman spectra, we employed a Renishaw confocal microscope
ronment, since structures with higher durability could be designed
operating at room temperature. The spectrums were obtained in backscattering
using less raw materials [18]. In addition, nanomaterials could be configuration using a 532 nm laser source operating at constant power of 5 mW.
used to reduce and prevent the toxicity of nanoparticles in the The Infrared spectra were obtained using an Agilent Care 640 FTIR Spectrometer,
environment more efficiently [19]. in the diffuse reflection mode. The spectral range measured was 4000–350 cm 1.
All samples were prepared with the same dimensions: 1 mm thick disk with
Carbon-based nanoparticles constitute elements of one impor-
10 mm of diameter, cut directly from cement mortar specimens.
tant class of materials that is applied to cement and mortar based
composites to enhance their tensile and flexural strength, or
2.2. Sample preparation
expand their thermal [20] and electrical [21] characteristics. In
particular, reinforcing nanomaterials such as carbon nanofibers Five different mortar recipes were prepared using a pan type mortar mixer. The
(CNF) and carbon nanotubes (CNT), have been shown to improve water-to-cement ratio was kept at 0.4 and the quantity of sand at a proportion of
mechanical strength and stiffness of cementitious nanocomposites 3 the weight of cement. The MLG powder content varied from 0% to 0.033% by
[22], although, some contradictory results have been reported weight of cement. In order to guarantee MLG dispersion through the composite
we first poured the solution of MLG and isopropanol into the dry sand, and this
[23,24]. One challenge of studying these nanocomposites is the dif-
resulting combination underwent mixing for approximately 10 min. The portion
ficulty in dispersing CNT’s and CNF in the OPC matrix. CNT’s and of sand was put to rest at room temperature, in a way that the isopropanol fully
CNF tends to form agglomerates and adhere together, probably evaporated after 48 h. Only then cement and water were added to produce mortar.
due to strong Van der Waals force, making difficult to archive a It is important to notice that no additives, such as superplasticizer, water-
reducing agents or surfactants, were employed for workability. Mix proportions
desired level of dispersion [25].
of mortars with and without MLG are given in Table 2. All mixes were cast into
As CNTs and CNF, graphene-oxide (GO) sheets have been exten- cylindrical molds (50 mm diameter and 100 mm length), and maintained at
sively used to improve the properties of mortar nanocomposites 23 ± 2 °C and a relative humidity of 50 ± 15% for 24 h, then demolded and sub-
[24,26–28]. In this 2-dimensional material, the oxygen- merged into a tank of water for curing. 12 h before the intended mechanical inves-
containing functional groups attached to the GO sheets improves tigations all samples were removed from the water tank and left to stand at room
temperature. Compressive and splitting tensile strength were investigated at the
its dispersion in water [29–31], and in the cement matrix
ages of 3, 7 and 28 days. In order to verify the reproducibility of the results, we pre-
[24,26–28,32] affecting the mortar hydration and mechanical pared at least two samples for each evaluation.
properties [33]. However, the process of oxidation and exfoliation
might affect GO mechanical properties negatively, lowering the 2.3. Mechanical property tests
elastic modulus and tensile strength [34].
In addition to GO, a few studies have focused on graphene- Compressive and splitting tensile strength tests were performed using a
based mortar composites [35–38]. Graphene exhibits many HD200T Digital Servo-Hydraulic Press (2000 kN maximum load). The compressive
strength tests were determined following the protocols described by the Brazilian
extraordinary properties including a high tensile strength and elas-
Regulatory Agency in ABNT/NBR-7215:1996 ‘‘Portland Cement – Determination of
tic modulus [39,40], high surface area, and also the fact that their compressive strength”, using a loading rate of 0.25 ± 0.05 MPa/s; and the splitting
edges can be functionalized to enhance its dispersion in compos- tensile strength was measured following the procedures described by ABNT/NBR-
ites [36,37]. Therefore, studies evaluating how multilayer graphene 7222:2011 ‘‘Concrete and mortar – Determination of the tensile strength by dia-
influences the parameters of cementitious composites are of great metrical compression of cylindrical test specimens” with a loading rate of
0.05 ± 0.02 MPa/s.
value and can certainly stimulate future research lines within civil
construction technologies. Here, we report investigations on
OPC-based mortar nanocomposites reinforced with multilayer 3. Results and discussion
graphene nanoparticles. We investigated the morphological
properties of MLG and cement mortar specimens using SEM and 3.1. Morphology
metallographic microscopy. Their chemical structure were evalu-
ated by using FTIR and Raman techniques, and their mechanical 3.1.1. MLG
properties by undertaking compressive and splitting tensile AFM analyses (Fig. 2a-b) of the representative flake depicted in
strength tests. Fig. 1(b) shows an average thickness around 10 nm (Fig. 2b) and an
average lateral size around 4 lm, in agreement with values
observed in previous works [43,44] involving the exfoliation of
2. Experimental
natural crystalline graphite flakes to further obtain multilayer gra-
2.1. Materials and synthesis phene flakes with thickness varying between 0.7 and 20 nm. It is
important to observe that the MLG flakes display a low surface
Ordinary Portland Cement (OPC), type CP-II Z-32, which incorporates poz- roughness and sharp corners. These characteristics suggest that
zolanic additions varying from 6 to 14% by weight of cement, and sieved sand with
mass density of 3.57 g/cm3 and fineness modulus equivalent to 2.4 were used to
the synthetized MLG flakes are composed of graphene layers with
prepare the mortar specimens. Table 1 presents the characteristic properties of defects/irregularities at the borders. This was in fact observed
the OPC used in this work along with its chemical composition. through Raman analyses and is discussed in Section 3.2.
380 R.A. e Silva et al. / Construction and Building Materials 149 (2017) 378–385

Table 1
Chemical and physical properties of OPC – CP II Z32.

Material SiO2 Al2O3 Fe2O3 SO3 CaO


Content (% mass) 22.32 6.53 3.09 2.53 53.73
Specific surface area (cm2/g)(Blaine) 3.522
Setting time (min) Start 250
Final 292

3.1.2. Portland cement mortar


Fig. 3(a-b) depicts metallographic images of a plain mortar sam-
ple and a sample prepared using 0033% MLG. From the images, it is
possible to observe a slight decrease in pore size and distribution
with the incorporations of MLG. Fig. 3(a) shows a typical Portland
Cement mortar specimen consisting of (1) hydrated paste with CSH
(calcium silicate hydrate) and Ca(OH)2 (calcium hydroxide), (2)
sand grains, which are partly responsible for determining mortar’s
specific mass, its modulus of elasticity and thermal conductivity,
(3) empty spaces, that said, trapped air pores and capillarity pores,
(4) anhydrous cement. Despite the small concentration of MLG, it is
relatively notable the presence of unfamiliar structures (5) in Fig. 3
(b). We believe that those elements appear to be graphene sheets.
Fig. 4(a–d) shows SEM images for different cement mortar
components with MLG.
A careful analysis has been made on extra SEM images that are
not presented in this work and yet no microcrack formations were
identified on graphene-reinforced samples, which confirms that
MLG incorporation may contribute to the propagation of internal
tensions throughout the mortar microstructure [45]. Also, gra-
phene might be acting as a structural binder, consequently reduc-
ing microcrack occurrences through the composite samples and
increasing both the splitting tensile and compressive strength as
observed through the mechanical tests.

3.2. Raman scattering

A typical Raman spectra of a MLG flake is shown in Fig. 5. The


peak in the region of 1590 cm 1 (G-mode) is related to the CAC
stretching mode in graphene [46], while the peak around
1355 cm 1 (D-mode) corresponds to the breathing mode of sp2
carbon atoms, and is associated to the presence of disorder and
defects in the graphitic structure [47].
The SEM and AFM images suggest that the MLG flakes are
formed by folded layers of small crystalline sheets of graphene,
which indicates that the low intensity D-band (low ID/IG ratio) is
related to defects occurring at the edge of the flakes [48,49] and
not to chemical oxidation of graphene.

3.3. Chemical structure

Fig. 1. a) Image of a typical high oriented natural graphite flake before the FTIR Spectra of OPC-based mortar samples with and without
syntheses process; b) SEM image of a representative MLG flake present in mortars the addition of MLG powder are shown in Fig. 6(a–c). Assignments
prepared herein.
of the main functional groups bands are displayed in Table 3. It can

Table 2
Mix proportions of mortar recipes. All samples have the same water/cement ratio and the same sand/cement proportion; MLG percentage is given by weight of cement.

Materials Pristine Mortar (CM) MLG-0.015% (CM-MLG15) MLG-0.021% (CM-MLG21) MLG-0.027% (CM-MLG27) MLG-0.033% (CM-MLG33)
Cement (g) 910 910 962 962 962
Sand (g) 2730 2730 2886 2886 2886
Water (ml) 413.1 413.1 436.7 436.7 436.7
MLG (%) – 0.015 0.021 0.027 0.033
R.A. e Silva et al. / Construction and Building Materials 149 (2017) 378–385 381

Fig. 2. a) AFM image of a MLG nanosheet and b) the height profile at 3 different
regions of the MLG as indicated in a).

be considered that edge defects of MLG flakes contain carbonyl Fig. 3. Metallographic microscopy of OPC-based mortar without (a) and with (b)
(C@O) and carboxyl (CAOOH) groups, which have been associated MLG. In the figure: (1) hydrated paste with CSH (calcium silicate hydrate), and Ca
with the D-band detection in Raman Spectra from Fig. 5. According (OH)2 (calcium hydroxide); (2) sand grains; (3) empty spaces; (4) anhydrous
cement; (5) MLG.
to a mechanism proposed by Sharma and Kothiyal [50] and Tong
et al. [45], those groups are likely to interact with calcium in CSH
molecules and this can be in fact observed if a thorough evaluation The effect of tricalcium aluminate conversion into CSH crystals
on FTIR spectra is made. Bands associated with Ca-O stretchings can also be linked with the fact that our low-defective MLG
(around 365 cm 1) that indicate the presence of non-hydrated nanoparticles has a remarkable hydrophobic behavior, which con-
Alite and Belite had their relative intensities lowered in the pres- sequently reduces tricalcium aluminate immediate hydration. As
ence of MLG, implying that these elements have been efficiently previously reported by Wang et al. [26], this reduction induces a
turned into crystalline and amorphous CSH structures. There are decrease in hydration heat and contributes to the generation of
indeed intensity and quantity increases in regard to the foregoing CSH crystals, fact that leads to higher resistances in advanced ages.
structures after MLG has been added (shoulder between 1050 and The presence of CSH crystals is notably observed around graphene
850 cm 1). nanoflakes in Fig. 4(d).
By carefully comparing our results with what was presented by
Sharma and Kothiyal [50], it can be inferred that the increase of
aluminate and silicate crystalline phases is strongly related to a 3.4. Mechanical strength of the hardened mortar
subsequent increase in the hydration ratio. Once again, an analysis
of relative intensities of absorption bands around 815, 570, 408 Fig. 7 shows the compressive strength variation along the three
and 390 cm 1 in FTIR spectra (respectively associated to AlAO, investigated ages for mortars samples with and without MLG. It
SiAO, AlAO and AlAO bands) demonstrates that these chemical can be noticed that the incorporation of MLG has substantially
bonds, which are in turn characteristic in aluminate and silicate increased the regarded mechanical property. The highest values
structures, had their intensities increased as MLG percentage was were obtained for mortar samples with 0.021% of MLG, reaching
incremented. Based upon the aforementioned modifications of IR improvements of 63.6, 94.1 and 95.7% at the age of 3, 7 and
bands, it can be ascertained that MLG addition into the composite 28 days, respectively, compared to the plain samples.
structure has stimulated favorable conditions to cement hydration Fig. 8 shows the splitting tensile strength variation along the
with higher tricalcium aluminate consumption, which has three investigated ages for mortars samples with and without
increased the generation of CSH crystals and consequently may MLG. It is possible to see that mortars prepared using MLG dosage
have improved the composite mechanical properties. This mecha- of 0.033% had the largest increases on the referred property at all
nism has also been considered by Tong et al. [45] using graphene ages. We observed remarkable increases of 100.0, 144.4 and
oxide nanoparticles in cementitious materials. 131.6% after 3, 7 and 28 days, respectively.
382 R.A. e Silva et al. / Construction and Building Materials 149 (2017) 378–385

Fig. 4. SEM images for different cement mortar components with MLG. a) Accumulation of several etringite crystals, which have been shaped in the presence of humidity
from the hardened cement paste; b) Empty air space in which secondary etringite crystals seem to have been assembled, along with prismatic calcium hydroxides (CH’s); c)
prismatic calcium hydroxides (CH’s), hexagonal crystals that are assembled in a rose-shaped arrangement and ettringite crystals; d) Graphene sheet immersed in the cement
paste and non-hydrated area.

With the addition of MLG, the relative improvement in splitting when the GO content was respectively 0.03% and 0.05%. The
tensile strength of mortar samples appears to be larger than the water-to-cement ratio was kept at 0.5.
relative compressive strength for all testing ages. Similar results The water-to-cement ratio adopted to prepare our mortar spec-
have been obtained with graphene oxide (GO) additions into imens was well comparable to the ones used in other works
cement mortars [24,26,27]. Wang et al. [26] observed an improve- [24,26], as well as the observed lack of fluidity and the magnitude
ment of 69.4, 106.4 and 70.5% for flexural strength and 43.2, 33 and of the compressive and tensile strength at ages even less than
24.4% for compressive strength of hardened mortar samples pre- 28 days.
pared with a water-to-cement ratio of 0.37. It was pointed out Despite our results presented significantly higher increases in
by the authors that those additions could have accelerated the compressive and splitting tensile strength compared to the forego-
hydration of cement and produced more regular crystals, with ing works, the specimens prepared herein incorporated pure and
fewer defects and contribute to the significant increase on the ana- low defective multilayer graphene sheets, produced with relatively
lyzed mechanical properties. simpler procedures in comparison to those required to generate
Pan et al. [24] have also demonstrated a remarkable increase in graphene oxide nanoparticles.
the same mechanical properties (up to 33% and 59%, respectively, In their works [24,26], the authors suggested that the improve-
for compressive and tensile strength after 56 days) for Portland ment of the composite mechanical properties was a result of strong
cement pastes reinforced with graphene oxide additions of 0.05% interfacial reactions between carboxylic acid groups of GO sheets
by weight of cement and prepared using a water-to-cement ratio and CSH or Ca(OH)2 groups of the cement matrix. That interaction
of 0.36. Shenghua et al. [27] investigated the influence of graphene creates, according to the authors, the mechanism to form a strong
oxide nanosheets on the microstructure and mechanical properties covalent bond on the interface between the GO and the cement
of cement mortars. The authors claimed that some of the major matrix, which collaborates to a more efficient load-transfer from
ingredients of unhydrous cement reacted preferentially with GO the matrix to graphene sheets [51].
functional groups, which has induced the formation of flower- Interestingly, cement hydration, which ends up producing
like hydration crystals. Their results showed a tensile strength hydrated calcium silicates (CSH), is an exothermic reaction
increase of 78.6% and a compressive strength increase of 47.9% (500 J/g), that is to say, it releases heat [52]. The influence of
R.A. e Silva et al. / Construction and Building Materials 149 (2017) 378–385 383

Crystaline Graphite

G
Intensity (A. U.)

Multilayer Graphene
G

D D´

1200 1400 1600 1800


-1
Wavenumber (cm )

Fig. 5. Raman spectra of crystalline graphite (upper graphic) and MLG showing the
typical G and D peaks.

hydration heat on the thermal aspect is directly related to the


cement consumption and also to its type [52]. Because of the poor
thermal conduction of cement, the hydration reaction correlated
with cement paste formation substantially increases the matrix
temperatures during initial ages. This increase in temperature
results in non-uniform thermal expansion inside the material,
which eventually induces the occurrence of unintended tensile
forces and crack formations [52]. We believe that the MLG filler
improve the heat propagation through the mortar specimen, which
leads to a lower temperature inside the sample. This behavior
might be understood in part because of the high thermal conduc-
tivity of non-oxidized graphene, typically 5000 W/mK [53], which
makes it behave very efficiently in terms of heat transportation
when incorporated to cementitious materials.
Although effects on the setting time of the cement mortar spec-
Fig. 6. FTIR spectra obtained by diffuse reflection method for cement mortar
imens have not been addressed in this work, the aforementioned samples without MLG (CM), with 0.021% of MLG (CM-MLG21) and 0.033% of MLG
thermal property of graphene possibly justifies a setting time (CM-MLG33). The spectra were separated in 3 regions for a better visualization.
reduction after MLG incorporation, which has also been suggested
by Wang et al. [26] as a consequence of better thermal conduction
during the hydration reaction in a study using graphene oxide in
Table 3
cement mortar samples.
Infrared assignment for plain cement mortars and with MLG incorporation. The
Thus, by inserting MLG on the composite structure we are able symbols m and d means stretching and bending vibrations. S and AS indexes refers to
to create a favorable heat transporting mechanism through which symmetric and antisymmetric, respectively.
thermal stresses are reduced and the hydration process accelerates 1
Wavenumber (cm ) Assignment References
the formation of conventional CSH gel, as suggested by FTIR anal-
1730–1630 m OAH [48,49]
ysis and observed in SEM image (Fig. 4d). As a beneficial conse-
1525 mAS CAO in aragonite [48]
quence, there is a reduction in pore volume and this reduction is 1160 mAS SiO4 [49]
conclusively accompanied by an enhancement of the composite 1100–850 m SiAOASi in CSH [48–50]
mechanical properties. 880 mS SiAO [49]
Considering the cost of raw material within MLG production, 815 m AlAO [49]
790 m AlAO in Al(OH)3 (Nordstrandite) [48]
chemical reagents and electric power consumption, approximately 699 m CAO [48]
U$ 0.02 would be needed to obtain 0.33 g of MLG, without regard 570 m SiAOASi [48]
to equipment-related expenses. In order to prepare a cubic meter 525 d AlAO [49]
of mortar with MLG proportion of 0.033%, supposing that the 408 m AlAO [49]
390–370 Tricalcium aluminate (Al2O3-3CaO) [49]
water-to-cement and sand-to-cement ratios are kept at the same
363 CaAO [49]
values herein adopted, we would need 85.8 g of MLG. Hence,
384 R.A. e Silva et al. / Construction and Building Materials 149 (2017) 378–385

ating a composite with improved mechanical properties. An


increase in compressive and splitting tensile strength was
observed for all samples prepared with MLG addition, reaching
improvements up to 131.6% and 95.7% for splitting tensile and
compressive strength, respectively, at 28 days. We believe that
due to MLG insertion the specimens ended up having a microstruc-
ture with less defects, which is associated to a more efficient con-
duction of thermal stresses through the composite. This has in turn
increased the generation of components such as hydrated calcium
silicates (CSH) crystals, elements that improve the composite
mechanical properties.
The encouraging results obtained with the addition of MLG to
OPC-based mortars may contribute to several civil construction
applications, since the process used for obtaining the MLG’s
described herein could be used on a large scale and with lower
costs than other previous reports.

Acknowledgment

The authors like to thank to CNPq, CAPES and FAPEMIG for the
financial support for this work. A special thanks to Nacional de
Grafite Ltda (Brazil) for providing the natural graphite used in this
work.
Fig. 7. Compressive strength for mortars with different MLG dosages as a function
of time (upper graphic), and the relative change in relation to the sample without
MLG (rCO ). The symbols s d D j h refers to samples CM, CM-MLG15, CM-MLG21,
References
CM-MLG27 and CM-MLG33 respectively.
[1] V. Viswanathan et al., Challenges and advances in nanocomposite processing
techniques, Mater. Sci. Eng. R: Rep. 54 (2006) 121–285.
[2] R. Roy, R.A. Roy, D.M. Roy, Alternative perspectives on ‘quasi-crystallinity’:
Non-uniformity and nanocomposites, Mater. Lett. 4 (1986) 323–328.
[3] B.C. Trasferetti et al., Nanocomposites of amorphous hydrogenated carbon and
siloxane networks produced by PECVD, Chem. Mater. 16 (2004) 567–569.
[4] R. Gangopadhyay, A. De, Conducting polymer nanocomposites: a brief
overview, Chem. Mater. 12 (2000) 608–622.
[5] X. Cao et al., New nanocomposite materials reinforced with cellulose
nanocrystals in nitrile rubber, Polym. Testing 32 (2013) 819–826.
[6] P.H.C. Camargo, K.G. Satyanarayana, F. Wypych, Nanocomposites: synthesis,
structure, properties and new application opportunities, Mater. Res. 12 (2009)
1–39.
[7] J.M. Garcés et al., Polymeric nanocomposites for automotive applications, Adv.
Mater. 12 (2000) 1835–1839.
[8] M. Khalina et al., Preparation of acrylic/silica nanocomposites latexes with
potential application in pressure sensitive adhesive, Int. J. Adhes. Adhes. 58
(2015) 21–27.
[9] J.D. Nam et al., Novel electroactive, silicate nanocomposites prepared to be
used as actuators and artificial muscles, Sens. Actuators, A 105 (2003) 83–90.
[10] Y. Qing et al., Influence of nano-SiO2 addition on properties of hardened
cement paste as compared with silica fume, Constr. Build. Mater. 21 (2007)
539–545.
[11] M.V. Diamanti, M. Ormellese, M. Pedeferri, Characterization of photocatalytic
and superhydrophilic properties of mortars containing titanium dioxide, Cem.
Concr. Res. 38 (2008) 1349–1353.
[12] Z. Li et al., Investigations on the preparation and mechanical properties of the
nano-alumina reinforced cement composite, Mater. Lett. 60 (2006) 356–359.
[13] I. Campillo et al., Improvement of initial mechanical strength by nanoalumina
in belite cements, Mater. Lett. 61 (2007) 1889–1892.
[14] M.S. Morsy, S.H. Alsayed, M. Aqel, Hybrid effect of carbon nanotube and nano-
clay on physico-mechanical properties of cement mortar, Constr. Build. Mater.
25 (2011) 145–149.
[15] H. Li, M. Zhang, J. Ou, Flexural fatigue performance of concrete containing
nano-particles for pavement, Int. J. Fatigue 29 (2007) 1292–1301.
Fig. 8. Splitting tensile strength for mortars with different MLG dosages as a [16] H. Li, M. Zhang, J. Ou, Abrasion resistance of concrete containing nano-particles
function of time (upper graphic), and the relative change in relation to the sample for pavement, Wear 260 (2006) 1262–1266.
without MLG (rTO ). The symbols s d D j h refers to samples CM, CM-MLG15, [17] L. Raki et al., Cement and concrete nanoscience and nanotechnology, Materials
CM-MLG21, CM-MLG27 and CM-MLG33 respectively. 3 (2010) 918–942.
[18] B. Zhang et al, Environmental impacts of nanotechnology and its products, in:
Proceedings of the 2011 Midwest Section Conference of the American Society
for Engineering Education (2011).
incorporating MLG into a cubic meter of mortar to ultimately [19] D.G. Rickerby, Mark. Morrison, Nanotechnology and the environment: a
obtain a mixture with 0.033% would cost around U$ 5.20. European perspective, Sci. Technol. Adv. Mater. 8 (1) (2007) 19–24.
[20] D.D.L. Chung, Carbon materials for structural self-sensing, electromagnetic
shielding and thermal interfacing, Carbon 50 (2012) 3342–3353.
4. Conclusion [21] H. Li, H. Xiao, J. Ou, Effect of compressive strain on electrical resistivity of
carbon black-filled cement-based composites, Cement Concr. Compos. 28
(2006) 824–828.
We successfully developed a process to combine multilayer gra- [22] E.E. Gdoutos et al., Advanced cement based nanocomposites reinforced with
phene nanoparticles with Ordinary Portland Cement mortars, cre- MWCNTs and CNFs, Front. Struct. Civil Eng. 10 (2016) 142–149.
R.A. e Silva et al. / Construction and Building Materials 149 (2017) 378–385 385

[23] A. Cwirzen et al., SEM/AFM studies of cementitious binder modified by [38] M. Cao, H. Zhang, C. Zhang, Effect of graphene on mechanical properties of
MWCNT and nano-sized Fe needles, Mater. Charact. 60 (2009) 735–740. cement mortars, J. Cent. South Univ. 23 (2016) 919–925.
[24] Z. Pan et al., Mechanical properties and microstructure of a graphene oxide– [39] C. Lee et al., Measurement of the elastic properties and intrinsic strength of
cement composite, Cement Concr. Compos. 58 (2015) 140–147. monolayer graphene, Science 321 (2008) 385–388.
[25] S. Parveen et al., A review on nanomaterial dispersion, microstructure, and [40] W. Wei, X. Qu, Extraordinary physical properties of functionalized graphene,
mechanical properties of carbon nanotube and nanofiber reinforced Small 8 (2012) 2138–2151.
cementitious composites, J. Nanomater. (2013) e710175. [41] P.-H. Chen, D.D.L. Chung, Comparative evaluation of cement-matrix
[26] Q. Wang et al., Influence of graphene oxide additions on the microstructure composites with distributed versus networked exfoliated graphite, Carbon
and mechanical strength of cement, New Carbon Mater. 30 (2015) 349–356. 63 (2013) 446–453.
[27] L. Shenghua et al., Effect of graphene oxide nanosheets of microstructure and [42] D.D.L. Chung, Interface-derived extraordinary viscous behavior of exfoliated
mechanical properties of cement composites, Constr. Build. Mater. 49 (2013) graphite, Carbon 68 (2014) 646–652.
121–127. [43] F.P. Rouxinol et al., Low contact resistivity and strain in suspended multilayer
[28] K. Gong et al., Reinforcing effects of graphene oxide on Portland cement paste, graphene, Appl. Phys. Lett. 97 (2010) 253104.
J. Mater. Civ. Eng. 27 (2) (2014) A4014010. [44] R.T. Khare et al., Enhanced field emission of plasma treated multilayer
[29] J. Kim et al., Graphene oxide sheets at interfaces, J. Am. Chem. Soc. 132 (2010) graphene, Appl. Phys. Lett. 107 (2015) 123503.
8180–8186. [45] T. Tong et al., Investigation of the effects of graphene and graphene oxide
[30] L. Qiu et al., Dispersing carbon nanotubes with graphene oxide in water and nanoplatelets on the micro-and macro-properties of cementitious materials,
synergistic effects between graphene derivatives, Chem – A Eur. J. 16 (2010) Constr. Build. Mater. 106 (2016) 102–114.
10653–10658. [46] A.C. Ferrari et al., Raman spectrum of graphene and graphene layers, Phys. Rev.
[31] W.S. Hummers, R.E. Offeman, Preparation of graphitic oxide, J. Am. Chem. Soc. Lett. 97 (2006) 187401.
80 (1958) 1339. 1339. [47] O. Akhavan, The effect of heat treatment on formation of graphene thin films
[32] S. Chuah et al., Nano reinforced cement and concrete composites and new from graphene oxide nanosheets, Carbon 48 (2010) 509–519.
perspective from graphene oxide, Constr. Build. Mater. 73 (2014) 113–124. [48] C. Casiraghi et al., Raman spectroscopy of graphene edges, Nano Lett. 9 (2009)
[33] S. Lv et al., Effect of graphene oxide nanosheets of microstructure and 1433–1441.
mechanical properties of cement composites, Constr. Build. Mater. 49 (2013) [49] L.G. Cançado et al., Influence of the atomic structure on the raman spectra of
121–127. graphite edges, Phys. Rev. Lett. 93 (2004) 247401.
[34] D.G. Papageorgiou, I.A. Kinloch, R.J. Young, Graphene/elastomer [50] S. Sharma, N.C. Kothiyal, Comparative effects of pristine and ball-milled
nanocomposites, Carbon 95 (2015) 460–484. graphene oxide on physico-chemical characteristics of cement mortar
[35] A. Sedaghat et al., Investigation of physical properties of graphene-cement nanocomposites, Constr. Build. Mater. 115 (2016) 256–268.
composite for structural applications, Open J. Compos. Mater. 04 (2014) 12– [51] G.Y. Li, P.M. Wang, X. Zhao, Mechanical behavior and microstructure of cement
21. composites incorporating surface-treated multi-walled carbon nanotubes,
[36] T. Tong et al., Investigation of the effects of graphene and graphene oxide Carbon 43 (2005) 1239–1245.
nanoplatelets on the micro- and macro-properties of cementitious materials, [52] A.M. Neville, Properties of Concrete, Fourth Ed., Pearson Educational Limited,
Constr. Build. Mater. 106 (2016) 102–114. Harlow, 1995. pp. 16, 17 and 37.
[37] H. Alkhateb et al., Materials genome for graphene-cement nanocomposites, J. [53] A. Dey, A. Chroneos, N.St.J. Braithwaite, Ram P. Gandhiraman, S.
Nanomech. Micromech. 3 (2013) 67–77. Krishnamurthy, Appl. Phys. Rev. 3 (2016) 021301.

You might also like