You are on page 1of 9

Effects of Nanoalumina and Graphene Oxide on Early-Age

Hydration and Mechanical Properties of Cement Paste


Wengui Li, M.ASCE 1; Xiangyu Li 2; Shu Jian Chen 3; Guangcheng Long 4;
Yan Ming Liu 5; and Wen Hui Duan 6

Abstract: The effects of nanoalumina (NA) and graphene oxide (GO) on the early-age hydration and mechanical properties of portland
Downloaded from ascelibrary.org by University Of Florida on 04/15/17. Copyright ASCE. For personal use only; all rights reserved.

cement pastes were investigated in this study. The hydration heat release rate and cumulative heat of cement pastes incorporating different
dosages of NA and GO were evaluated using an isothermal calorimeter measurement method. Early-age electrical resistivity development was
investigated by a noncontact electrical resistivity technique. The results show that both NA and GO could efficiently accelerate cement
hydration. As a physical filler, NA significantly accelerates the hydration of tricalcium aluminate (C3 A) in cement. On the other hand,
GO is able to obviously reduce the dormant period of cement hydration and shift the heat flow peaks to the left by accelerating the hydration
of tricalcium silicate (C3 S) in cement. Compared to plain cement pastes, both the compressive and flexural strengths of cement pastes incor-
porating NA or GO are significantly increased. However, when NA and GO contents exceed the optimal amounts, improvements in flexural
strength tend to decline, which is probably due to particle agglomeration. NA-cement paste exhibited slightly higher electrical resistivity than
plain cement paste during hydration acceleration and deceleration stages. But GO-cement paste clearly showed lower electrical resistivity,
which might be attributed to iron diffusion caused by GO with large surface areas. DOI: 10.1061/(ASCE)MT.1943-5533.0001926. © 2017
American Society of Civil Engineers.
Author keywords: Cement paste; Nanoalumina (NA); Graphene oxide (GO); Hydration; Electrical resistivity.

Introduction cracks at the initial stage before these cracks develop into meso-/
macro-/microscale cracks (Konsta-Gdoutos et al. 2010). Previous
Cement-based materials are extensively used for building and researches showed that nanoalumina (NA) could significantly in-
construction. However, they usually have limited structural appli- crease the elastic modulus of cement paste by up to 143% at a
cations owing to their poor tensile strength and ductility. For dosage of 5% by weight of cement but had a limited effect on com-
example, weak tensile strength is associated with porous micro- pressive strength (Li et al. 2006). For cement mortars with NA ex-
structures and preexisting flaws. Recent advancements in nanotech- posed to high temperatures up to 800°C, NA enhanced compressive
nology have led to the production of nanosized particle, fibers, and strength by up to 16% and residual compressive strength and im-
sheets [e.g., nanoalumina, carbon nanotubes, and graphene oxide proved the relative elastic modulus and facilitated higher energy
(GO)], which can be used as reinforcements for cementitious com- absorption when 1% NA was added (Farzadnia et al. 2013). More-
posites (Sobolev and Shah 2015; Sanchez and Sobolev 2010). It has over, the frost resistance of concrete containing NA was better than
been reported that nanomaterials in cementitious composites are that containing the same amount of nanosilica. However, the com-
more effective than conventional steel bar and fiber reinforcements pressive strength of concrete containing nanosilica was higher than
at the millimeter scale because they can control nano-/microscale that containing the same amount of NA (Behfarnia and Salemi
2013). The microstructural refinement of cement mortar has also
1
Lecturer, ARC DECRA Fellow, Centre for Built Infrastructure been reported because NA caused a shift from pores greater than
Research, School of Civil and Environmental Engineering, Univ. of 1.0 μm to pores smaller than 0.1 μm (Campillo et al. 2007). The
Technology Sydney, Sydney, NSW 2007, Australia (corresponding author). increase in cement hydration and the faster dissolution of the Belite
E-mail: wengui.li@uts.edu.au phase due to the high reactivity of fine NA particles was probably
2
Research Fellow, Dept. of Civil Engineering, Monash Univ., Clayton, the reason for the refinement of microstructures. At the same time, a
VIC 3800, Australia. compact formation of hydration products and a reduced content of
3
ARC DECRA Fellow, Dept. of Civil Engineering, Monash Univ.,
calcium hydroxide (CH) crystals were also found when NA was
Clayton, VIC 3800, Australia.
4
Professor, School of Civil Engineering, National Engineering Labora- added in cementitious composites (Nazari and Riahi 2011). The
tory for High Speed Railway Construction, Central South Univ., Changsha enhanced mechanical properties of cementitious composites with
410075, Hunan, P.R. China. NA was reported to be mainly the result of their physical effects,
5
Ph.D. Candidate, Dept. of Civil Engineering, Monash Univ., Clayton, such as filling and nucleation effects, as well as the acceleration
VIC 3800, Australia. of early-age hydration due to the high surface area (Hemalatha et al.
6
Professor, ARC Future Fellow, Dept. of Civil Engineering, Monash 2015).
Univ., Clayton, VIC 3800, Australia. E-mail: wenhui.duan@monash Similarly, recent researches have reported that GO exhibits a
.edu
super high tensile strength and aspect ratio and a large surface area
Note. This manuscript was submitted on September 13, 2016; approved
on December 30, 2016; published online on April 12, 2017. Discussion (Xu and Gao 2011). Its extraordinary mechanical properties, com-
period open until September 12, 2017; separate discussions must be sub- bined with an effective dispersibility in water and inexpensive
mitted for individual papers. This paper is part of the Journal of Materials source, can make GO a promising material for enhancing the
in Civil Engineering, © ASCE, ISSN 0899-1561. mechanical properties of composites (Pan et al. 2015). It has been

© ASCE 04017087-1 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., -1--1


shown that carboxyl acid groups can react with calcium silicate hy- The aforementioned studies summarize well the improvements
drate (C-S-H) to form strong covalent bonds and, thus, notably im- to cementitious composites by NA and GO with dense microstruc-
prove the mechanical properties of cementitious composites (Gong tures and enhanced mechanical properties. However, there remains
et al. 2014). On the other hand, GO sheets have a filler effect by a lack of deep understanding on the early-age properties of cement
filling up the voids between hydration products. With the right paste when NA and GO are incorporated. Therefore, the reinforce-
composition, a higher packing density results in a lower water de- ment of NA and GO for cementitious composites needs to be thor-
mand of a cement mixture and contributes to strength enhancement oughly examined in terms of the hydration process and interaction
due to the reduced capillary porosity (Mohammed et al. 2015, between cement particles and NA/GO. In this study, the effect of
2016). Several studies also showed that GO application in cemen- NA and GO on the early-age hydration process, electrical resistiv-
titious systems enhances mechanical strength, especially at the ity, and mechanical strengths of cement paste were investigated by
early age of hydration (Lu et al. 2015). Moreover, GO accelerates thermometric TAM air, noncontact electrical resistivity measure-
cement hydration by a nucleation effect, and the permeability of ments (Cement and Concrete Resistivity-III) and mechanical tests.
cement paste is lower and the microstructure much denser when Therefore, the mechanical strength of cement paste was also exam-
GO is incorporated (Gong et al. 2014). ined and linked with early-age cement hydration processes.
Downloaded from ascelibrary.org by University Of Florida on 04/15/17. Copyright ASCE. For personal use only; all rights reserved.

During the early-age cement hydration, temperature is one of


the key indicators enabling a prediction of dormancy and strength
development based on the cement paste mix design, which is usu-
ally evaluated according to ASTM C1679 (ASTM 2008) using a
thermometric TAM Air (TA Instruments, New Castle, Delaware)
isothermal calorimeter (Zhang and Jahidul 2012). The temperature
gradient can be separated into five stages based on duration. It was
found that when nanosilica was introduced into cement composites,
the heat release of the hydration process obviously increased. The
temperature increase took place following a dormancy period and
into the acceleration period as a result of incorporating nanosilica
(Hou et al. 2013). On the other hand, measurement of electrical
resistivity development also indicated an early-age cement hydra-
tion process that relies on solid hydration products and ionic dif-
fusions through water-filled pores (Li et al. 2003; Wei and Li 2005).
The factors influencing the electrical resistivity of cementitious
composites are mainly controlled by the microstructure and porous
solution conductivity. A noncontact electrical resistivity method
has been used to measure the development of electrical resistivity
of fresh concrete and the pore solution within concrete mixtures
[Zhang et al. 2009; Z. Li and W. Li, “Contactless, transformer-based
measurement of the resistivity of materials.” U.S. Patent No.
6639401 (2003)]. The effects of cement type, water-to-cement ratio,
alkali content, superplasticizer selection, and blended cement on
the hydration process of cement composites were investigated cor-
respondingly (Wei and Li 2005, 2006).

Fig. 1. Morphologies of nanoalumina and graphene oxide: (a) TEM


Table 1. Chemical Composition of Portland Cement
images of nanoalumina (NA); (b) TEM images of graphene oxide (GO)
Composition Percentage
Al2 O3 4.7
SiO2 19.9
CaO 63.9
Fe2 O3 3.4
K2 O 0.5
MgO 1.3
Na2 O 0.2
SO3 2.6
Loss on ignition 3
Other 0.5

Table 2. Elemental Analysis of Graphene Oxide


Element Percentage
Carbon 49–56
Hydrogen 0–1
Nitrogen 0–1
Sulfur 0–2 Fig. 2. Thermometric TAM air isothermal calorimeter (image by
Oxygen 41–50 Wengui Li)

© ASCE 04017087-2 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., -1--1


Materials and Methods (ASTM 2009) standard. The chemical composites of cement are
shown in Table 1. Nanoalumina (nanopowder, 13 nm primary particle
Raw Materials size, 99.8% trace metals basis) was purchased from Sigma-Aldrich
(Castle Hill, New South Wales, Australia) with an average size of
The portland cement was provided by Cement Australia (West less than 50 nm and surface area greater than 40 m2 =g. GO
Footscray, Victoria, Australia) and conforms to the requirements of (4 mg=mL, water dispersion) was obtained from Graphenea (San
Type I general purpose cement, as defined by the ASTM C150 Sebastián, Spain), and the elemental properties of graphene oxide
are shown in Table 2. The morphologies of NA and GO obtained
by transmission electron microscopy (TEM) at the Monash Centre
for Electron Microscopy are shown in Fig. 1. The figure clearly
shows the shapes of NA particles and GO sheets at nano- and micro-
scales. Both the NA powder and GO solutions were diluted with tap
water. The solution was subjected to ultrasonication dispersion using
a Vibra-Cell VCX-500 ultrasonic processor with an energy of 40,000
Downloaded from ascelibrary.org by University Of Florida on 04/15/17. Copyright ASCE. For personal use only; all rights reserved.

J, an amplitude of 80%, and a pulse of 4 s (Vibra-Cell-Sonics &


Materials, Newtown, Connecticut). For the mix design, the water-
to-cement ratio was constant at 0.4. The NA-cement paste samples
were prepared with 2, 4, and 6% NA by weight of cement. The
GO-cement paste samples were cast containing 0.02, 0.04, and
0.06% GO by replacement of cement. Plain cement pastes were also
prepared as reference samples for comparison in all measurements.
A high-speed shear mixer (CTE Model 7000, Cement Test
Equipment, Tulsa, Oklahoma) was used for the sample mixing
to improve the distribution of NA/GO in cement paste. Mixing pro-
Fig. 3. Noncontact electrical resistivity measurement (CCR-III) (image
cedures similar to those in ASTM C1738-11a (ASTM 2011) were
by Wengui Li)
adopted. Add the correct amount of nanomaterials and water to the

Fig. 4. Effect of NA on hydration heat of cement paste during 48 h: (a) heat release rate; (b) difference in heat release rate; (c) cumulative hydration
heat; (d) difference in cumulative hydration heat

© ASCE 04017087-3 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., -1--1


mixing container and premix the solution at low speed [100–200
revolutions per minute (rpm)] for 15 s to homogenize the solution;
add cement powder within 30 s while the mixer is operated at the first
preset speed (4,000 rpm) for 60 s; stop the mixer for 30 s, during
which any paste that may have collected on the sides of the bowl
is scraped down into the hatch, and operate the mixer at the second
preset speed (12,000 rpm) for 30 s, stop the mixer for 15 s, and start
the mixer at the same speed for an additional 30 s. After mixing,
duplicate portions of the mixtures were used for isothermal calorim-
eter, electrical resistivity measurement, and mechanical tests. The
compressive and flexural strengths were measured for cement paste
following the 28 days after sample casting. The cement pastes were
mixed, cast, compacted, and placed in a moist storage room
(20  2°C) for 24 h. Then the samples were demolded and moved
Downloaded from ascelibrary.org by University Of Florida on 04/15/17. Copyright ASCE. For personal use only; all rights reserved.

to a standard curing room up to 28 days to measure the compressive


and flexural strengths. At least three duplicate samples were tested
for each strength measurement to ensure reproducibility of results.

Test Methods
The effects of NA/GO on the early-age hydration process of cement
paste were investigated according to ASTM C 1679 by thermomet-
ric TAM air at a temperature of 25°C (Bentz et al. 2009, 2012;
Onuaguluchi et al. 2014). Temperature was selected to simulate
the standard curing condition. The calorimeter was conditioned at
25°C for a day before experiments, and the amplifier range was set
at 600 mW, as shown in Fig. 2. After the mixing, a sample of 5.0 g
was transferred to a sample ampoule. The calorimeter began to
record heat after the cementitious materials had been in contact
with water for approximately 30 min. Thus, the hydration heat gen-
erated initially during mixing and preparation was not captured.
The heat generated from the hydration was monitored every minute
for 48 h. The measurement was normalized by sample mass. The
normalized power output was then converted to heat generated in
the sample (in joules/gram). In the following results, each curve
Fig. 5. Effects of NA on mechanical strengths of cement paste at
represents the average of three easily reproduced measurements.
28-day: (a) compressive strength; (b) flexural strength
The error of the specifical heat flow is 0.1 mW=g for hydration
durations. On the other hand, the electrical resistivity of cement
paste was measured by a noncontact electrical resistivity measure-
with increasing NA content. For instance, the main heat flow peak
ment device (CCR-III, Hong Kong Brilliant Concept Technologies,
appeared at around 8 h for the reference samples without NA,
Hong Kong, China) as shown in Fig. 3. The samples were cast into
whereas the main peaks were reached at around 5, 6, and 7 h for the
ring-shaped molds (1.67 L), of dimensions 770 mm ðperimeterÞ ×
NA-cement pastes with 2, 4, and 6% NA, respectively. The left shift
55 mm ðheightÞ × 42 mm ðwidthÞ including a cover for preventing
of the peak heat flow can be an indication of NA-promoting hydration
water evaporation during the test period. Data acquisition occurred at
product growth and, therefore, accelerating the cement particles from
5 s intervals over 48 h. The compression and three-point bending
reacting with water at an early age. Based on the results, it seems that
tests of NA/GO reinforced cement paste cylinders and nonnotch
there are two mechanisms by which NA can affect the early-age
beams were performed to evaluate mechanical properties at the
hydration of cement pastes. First, NA particles can function as fine
Concrete Structures Laboratory at Monash University Australia. The
fillers, providing additional surfaces for the nucleation and growth of
compressive and flexural strengths of the cement paste after 28 days
hydration products. In this case, the greatly increased surface area of
of standard curing were determined in accordance with ASTM C109
the NA is expected to accelerate the early-age hydration (Farzadnia
and ASTM C348. Compressive strength was measured using a Shi-
et al. 2013; Heikal et al. 2015). Second, NA probably provided
madzhu (Shimadzu Japan, Kyoto, Japan) universal testing machine
an additional source of alumina ions to the pore solution and signifi-
(AG-X 300 kN) on 25 × 50 mm cylinders with a displacement rate
cantly accelerated the hydration of tricalcium alumina (C3 A) by an
of 0.5 mm=min. Flexural strength was examined using an Instron
increase in the second peak of the heat release flow curve (Bullard
universal testing machine (33R4204) on a 140 × 30 × 15 mm beam
et al. 2011; Taylor 1997). It was also found that the time to reach the
with a displacement rate of 0.05 mm=min.
main peaks of cement hydration decreased as NA increased. This de-
duction still requires more comfirmation by further investigations.
Results and Discussions In comparison to the reference cement paste, the increase in
the maximum heat flow of NA-cement pastes containing different
dosages of NA after the first 48 h was 0.8 mW=g for 2% NA,
Effect of Nanoalumina
2.3 mW=g for 4% NA, and 4.1 mW=g for 6% NA, respectively.
Figs. 4(a and b) present the heat flow curves for the NA-cement Figs. 4(c and d) show the 48 h cumulative heat release of the
pastes with different dosages of NA from 2 to 6% during the first NA-cement pastes with different NA dosages. It was observed that
48 h. It can be observed that the hydration heat release accelerated during the first 24 h, the cumulative heat increased with the NA

© ASCE 04017087-4 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., -1--1


amount. After 24 h hydration, this trend gradually changed until the Moreover, the restriction in crystal size of CH and better distribu-
end of the test (at 48 h), where the cumulative heat curves of cement tion of calcium silicate hydrate (C-S-H) through the high reactivity
pastes approximately overlapped each other. For example, the cu- of NA as ultrafine filler could be another reason for the enhance-
mulative heat release for the cement paste with 6% NA at 48 h was ment in compressive strength (Campillo et al. 2007). On the other
179.9 J=g, which was slightly higher compared to the reference hand, the effects of NA on the flexural strength of cement pastes
cement paste, which was 176.9 J=g for the same duration. This im- were relatively less obvious compared to compressive strength, as
plies that the hydration process of cement paste increases with NA shown in Fig. 5(b). The flexural strengths of NA-cement pastes
content during the first 24 h. It is noteworthy that during the first 24 with dosages of 1, 2, 3, and 4% are higher than that of the respective
h the cumulative heat release shows that an increasing trend with a reference samples. When the dosage of NA is 2%, the flexural
larger amount of NA exhibits more hydration heat. After 24 h hy- strength of cement paste reaches its maximum and is 29.7% higher
dration, the cumulative heat release curves are quite similar to each than that of the reference paste. However, the flexural strengths
other, even with different NA dosages. This trend may be attributed of cement pastes with dosages of 3 and 4% NA are relatively lower
to an inadequate supply of water to maintain hydration, reducing than that of cement with 2% NA but are still 28.7 and 23.8% those
the reaction surface area between cement particles and water (Bentz of the respective reference samples. Therefore, as NA content
Downloaded from ascelibrary.org by University Of Florida on 04/15/17. Copyright ASCE. For personal use only; all rights reserved.

et al. 2009; Lootens and Bentz 2016). As a result, the cement increases to 3 and 4%, the enhancement in flexural strength de-
hydration slowed down in comparison with the surface in the pure creases. This might be explained by particle agglomeration result-
cement system at a late age. ing from an excessive amount of NA, which is likely to surround
The compressive strength of NA-cement pastes incorporating fine cement particles and, hence, loosen the microstructures of ce-
different NA dosages at 28 days is illustrated in Fig. 5(a). With the ment paste (Li et al. 2006). The optimal amount of NA is 2% for
increase of NA dosage, the compressive strength of cement pastes cement paste based on the mechanical strength improvement.
correspondingly increases. The compressive strengths of the ce-
ment pastes with NA dosages of 1, 2, 3, and 4% are 32.1, 35.6,
Effect of Graphene Oxide
47.3 and, 52.3% higher than that of the respective reference sam-
ples. The reason for this is that NA, as a super fine aggregate, can The effect of GO on the rate of hydration heat development and
accelerate a higher rate of hydration and fill porous areas in ce- the corresponding cumulative heat release of cement pastes with
ment pastes to lead to a more compact and denser microstructure. GO dosages from 0.02 to 0.06% are shown in Figs. 6(a and b),

Fig. 6. Effect of GO on hydration heat of cement paste during 48 h: (a) hydration heat release rate; (b) difference in heat release rate; (c) cumulative
hydration heat; (d) difference in cumulative hydration heat

© ASCE 04017087-5 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., -1--1


respectively. The GO reduced the dormant period and shifted the
heat flow peaks to the left. For example, the increase in the first
peak is related to the accelerated hydration of C3 S due to increased
nucleation sites provided by the fine GO sheets with large surface
areas (Pan et al. 2015; Gong et al. 2014). In addition to the effect of
the fine GO particle size, the accelerated cement hydration (the sec-
ond peak) might also be attributed to the increased concentration of
CH in solution due to increased C3 S reaction. However, the in-
creases of the first and second peaks were relatively less significant
for the GO-cement paste than those of NA-cement pastes. This im-
plies that the NA is more active than GO in accelerating cement
hydration. In the GO-cement paste, the induction period (an initial
slow stage of a chemical reaction) started earlier than the reference
sample. The acceleration period of the GO-cement pastes, which
Downloaded from ascelibrary.org by University Of Florida on 04/15/17. Copyright ASCE. For personal use only; all rights reserved.

was associated with the accelerated formation rate of C-S-H, also


began earlier than that of the reference paste (Mostafa and Brown
2005). The first heat flow peaks of GO-cement pastes incorporating
0, 0.02, 0.04, and 0.06% GO were 2.25, 2.39, 2.44, and 2.49 mW=g,
respectively. The subsequent deceleration period correlated to the
depletion of the calcium sulphate phase and was assigned to a sec-
ond C3 A reaction presumably resulting in the formation of ettrin-
gite or AFm (aluminum ion mono-) phase (Deschner et al. 2012).
Moreover, the hydration time needed to reach the main peaks of
cement hydration decreased as the GO dosage increased.
In comparison to the reference sample, the increase in the maxi-
mum heat flow of cement pastes containing a GO dosage was
0.1 mW=g with 0.02% GO, 0.2 mW=g with 0.04% GO, and
0.3 mW=g with 0.06% GO. As shown in Figs. 6(c and d), the initial
reaction in the cement paste increased as the amount of GO in-
creased. The GO-cement pastes yielded much greater heat release
in the first few minutes than the reference paste. This effect of GO
continued through the induction period because the heat release
of cement pastes increases as the GO dosage increases during the
induction period. In the acceleration period, GO-cement pastes
showed a clear acceleration effect on C3 S at around 6–12 h. It
was found that cement pastes incorporating GO released more heat Fig. 7. Effect of GO on mechanical strengths of cement paste at
than the reference sample during the first 14 h. It appears that GO 28-day: (a) compressive strength; (b) flexural strength
contributes significantly to heat release during the first 14 h (Pane
and Hansen 2005; Kawashima et al. 2013). After 48 h of hydration,
the increases in the total heat of GO-cement pastes compared to the hydration kinetics, stimulating the growth of hydration products
reference one became less obvious. This result is in agreement with (Chuah et al. 2014). It is also considered that flexural strength was
previous studies (Lu et al. 2015) that reported that GO was more further enhanced by the high specific surface area of GO, which
effective at increasing the early-age mechanical strength of concrete improved the bond interface between the GO and the paste matrix.
compared to the late age. However, a decrease in the flexural strength of cement paste was
The effect of GO on the compressive strength of cement pastes observed when the GO dosage exceeded 0.04%. The factor that
with GO dosages of 0.02, 0.04, 0.06, and 0.08% at 28 days were mitigated the improvement in mechanical strength as GO content
evaluated as shown in Fig. 7(a). The figure shows that compared to increased is likely the poor dispersion of GO in cement paste. The
the reference sample, the compressive strengths of cement pastes GO agglomeration negatively affected flexural strength enhance-
incorporating 0.02, 0.04, 0.06, and 0.08% were greatly enhanced ment, especially in the case of a 0.08% GO dosage (Li et al. 2016).
by 42.3, 43.4, 48.5, and 56.3%, respectively. A GO sheet contains
carboxylic acid groups that can react with CH, which was also ob-
served in a previous investigation (Li et al. 2005). The increase in Electrical Resistivity Development
compressive strength of the GO-cement pastes is more likely due The development of electrical resistivity of cement paste can be
to the greater contribution of the cement hydration acceleration ef- determined mainly by the porosity and conductive ion concentra-
fect and the strong interfacial adhesion between GO and the cement tion in the pore solution. The hydration process can be generally
paste. Fig. 7(b) shows that GO obviously enhanced the flexural separated into four stages: dissolution, setting, acceleration, and de-
strength of the cement pastes. With GO dosages of 0.02 and 0.04%, celeration (Wei and Li 2006; Skalny and Young 1980). The effects
the increases in the flexural strength of cement paste relative to that of NA and GO on the development of electrical resistivity of ce-
of the reference sample are 4.5 and 14.2%, respectively. For exam- ment pastes are presented in Fig. 8. First, in the dissolution stage,
ple, the maximum flexural strength was achieved by a GO dosage after the cement is mixed with water, ions are rapidly released
of 0.04%. The optimal amount of GO is 0.04% for cement paste from the surface of cement grains and dissolved into the solution.
based on mechanical strength improvement. The improvement in Electrical resistivity decreases due to the increase in ion concen-
flexural strength can be due to the increased degree of hydration tration and the mobility of these ions. Based on improvements in
of the cement paste because GO has a nucleation effect on cement the mechanical properties of cement paste, samples with optimal

© ASCE 04017087-6 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., -1--1


Downloaded from ascelibrary.org by University Of Florida on 04/15/17. Copyright ASCE. For personal use only; all rights reserved.

Fig. 8. Effects of NA and GO on electrical resistivity of cement paste


during 48 h: (a) electrical resistivity development; (b) difference in Fig. 9. Effects of NA and GO on electrical resistivity development of
electrical resistivity development cement paste during 12 h: (a) rate of electrical resistivity development;
(b) difference in rate of electrical resistivity development

amounts of 2% NA and 0.04% GO were compared in terms of


mechanical strength and electrical resistivity. When 2.0% NA or are consistent with the hydration heat results in Fig. 10(a). For in-
0.04% GO was added to cement pastes, the electrical resistivity was stance, cement paste with 2% NA was observed to exhibit higher
generally higher compared to reference samples because of the electrical resistivity compared to cement paste with 0.04% GO.
decrease of free water in pores owing to the high water adsorption Finally, in the deceleration stage, as cement hydration continues,
capacity of NA and GO with large surface areas, especially for GO- hydrates form a thicker barrier that blocks the exposure of the sol-
cement paste. The reason for this is that, compared to the spherical ution to unhydrated cement particles. NA-cement paste exhibits
shape of NA, GO can be regarded as having a two-dimensional higher electrical resistivity compared to both GO-cement paste
planar sheet structure (high aspect ratio), which creates much more and the reference sample because NA seems more active than GO
contact area with free water. Second, in the setting stage, owing to a when it comes to accelerating cement hydration and solid hydration
high dilution effect, a gel layer or membrane forms rapidly over the growth. Moreover, although GO, like NA, also accelerates cement
surfaces of the cement grains. This initial hydration reaction con- hydration, the massive ion diffusion caused by GO sheets with
sumes ions, thereby increasing electrical resistivity. Therefore, ion large surface areas leads to much lower electrical resistivity com-
diffusion affects the development of electrical resistivity of cement pared to NA-cement paste and the reference sample (Xie et al.
paste. Cement pastes with NA or GO exhibit significantly lower 1996; Chen et al. 2004; Park et al. 2012).
electrical resistivity relative to reference samples because of the Fig. 10(b) shows that both NA and GO effectively improve both
ion diffusion caused by NA or GO, as shown in Fig. 9. Third, in the compressive and flexural strengths of cement pastes as a result
the acceleration stage, NA acts as nuclei for cement phases, effi- of accelerated hydration and nucleation effects. Compared to the
ciently promoting cement hydration owing to their high reactivity. reference sample, the compressive strengths of 2% NA-cement
At the same time, GO also provides nucleation sites to accelerate paste and 0.04% GO-cement paste after 28 days of standard curing
cement hydration, leading to rapid hardening. were increased by 35.6 and 43.4%, respectively. Meanwhile, the
Although the absolute values of the electrical resistivity of ce- increases in the flexural strengths of 2% NA-cement paste and
ment pastes with NA or GO were lower compared to reference 0.04% GO-cement paste were 29.7 and 14.2%, respectively. The
pastes, the development rates of electrical resistivity were much results imply that the contents of 2% NA and 0.04% GO by weight
faster. NA was found to be more efficient at accelerating cement of cement are related to optimal improvements in the mechanical
hydration and reaction products growth compared to GO, which strengths of cement paste, respectively.

© ASCE 04017087-7 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., -1--1


owing to ion diffusion caused by NA and GO. Furthermore,
in the acceleration and deceleration stages, the final electrical
resistivity of NA-cement paste is higher than those of both GO-
cement paste and plain cement paste. The relatively lower elec-
trical resistivity of GO-cement paste might be due to the ion
diffusion caused by GO with a large surface area.

Acknowledgments
The authors gratefully acknowledge the financial support of
the Australian Research Council (DE150101751, IH150100006),
Australia. The authors are also grateful for the financial support of
the National Natural Science Foundation of China (51408210) and
Downloaded from ascelibrary.org by University Of Florida on 04/15/17. Copyright ASCE. For personal use only; all rights reserved.

National Engineering Laboratory for High-speed Railway


Construction, Central South University, P.R. China. The construc-
tive comments and suggestions from Professor Surendra P. Shah
at Northwestern University, Evanston, Illinois are also highly
appreciated.

References
ASTM. (2008). “Standard practice for measuring hydration kinetics of
hydraulic cementitious mixtures using isothermal calorimetry.” ASTM
C1679, West Conshohocken, PA.
ASTM. (2009). “Standard specification for portland cement.” ASTM C150,
West Conshohocken, PA.
ASTM. (2011). “Standard practice for high-shear mixing of hydraulic
cement pastes.” ASTM C1738-11a, West Conshohocken, PA.
Behfarnia, K., and Salemi, N. (2013). “The effects of nano-silica and nano-
alumina on frost resistance of normal concrete.” Constr. Build. Mater.,
48, 580–584.
Bentz, D. P., Peltz, M. A., and Winpigler, J. (2009). “Early-age properties
of cement-based materials. II: influence of water-to-cement ratio.”
J. Mater. Civ. Eng., 10.1061/(ASCE)0899-1561(2009)21:9(512),
Fig. 10. Comparison between hydration heat release and mechanical 512–517.
strengths of cement pastes incorporating NA and GO at 28-day: Bentz, D. P., Sato, T., Varga, I., and Weiss, W. J. (2012). “Fine limestone
(a) cumulative hydration heat; (b) compressive and flexural strengths additions to regulate setting in high volume fly ash mixtures.” Cem.
Concr. Compos., 34(1), 11–17.
Bullard, J. W., et al. (2011). “Mechanisms of cement hydration.” Cem.
Concr. Res., 41(12), 1208–1223.
Conclusions Campillo, I., Guerrero, A., Dolado, J. S., Porro, A., Ibáñez, J. A., and
Goñi, S. (2007). “Improvement of initial mechanical strength by nano-
The early-age hydration and mechanical properties of cement alumina in belite cements.” Mater. Lett., 61(8–9), 1889–1892.
pastes incorporating NA and GO were investigated using iso- Chen, B., Wu, K. R., and Yao, W. (2004). “Conductivity of carbon
thermal calorimeter, noncontact electrical resistivity measurement, fiber reinforced cement-based composites.” Cem. Concr. Compos.,
and mechanical experiments. Based on the relevant results, the 26(4), 291–297.
following conclusions can be drawn: Chuah, S., Pan, Z., Sanjayan, J. G., Wang, C. M., and Duan, W. H. (2014).
1. NA can significantly accelerate cement hydration reactions, “Nano reinforced cement and concrete composites and new perspective
especially the C3 A phase, by an increase in the second peak of from graphene oxide.” Constr. Build. Mater., 73, 113–124.
Deschner, F., et al. (2012). “Hydration of portland cement with high replace-
heat release flow. Compared to plain cement paste, NA-cement
ment by siliceous fly ash.” Cem. Concr. Res., 42(10), 1389–1400.
paste exhibits improvements in both compressive and flexural Farzadnia, N., Ali, A., and Demirboga, R. (2013). “Characterization of
strengths. However, when NA dosage exceeds 2%, the enhance- high strength mortars with nano alumina at elevated temperatures.”
ment in flexural strength tends to decline, which could be due to Cem. Concr. Res., 54, 43–54.
NA agglomeration. Gong, K., et al. (2014). “Reinforcing effects of graphene oxide on portland
2. GO can obviously accelerate cement hydration and increase hy- cement paste.” J. Mater. Civ. Eng., 27(2), 1–6.
dration heat release. The accelerated hydration might also be Heikal, M., Ismail, M. N., and Ibrahim, N. S. (2015). “Physico-mechanical,
attributed to increased concentration of CH owing to increased microstructure characteristics and fire resistance of cement pastes con-
C3 S reaction. Both compressive and flexural strengths are found taining Al2 O3 nano-particles.” Constr. Build. Mater., 91, 232–242.
to be significantly improved by GO. However, the enhancement Hemalatha, T., Gunavadhi, M., Bhuvaneshwari, B., Sasmal, S., and Iyer, N. R.
(2015). “Characterization of micro- and nano- modified cementitious
in flexural strength of GO-cement paste tends to decrease by
system using micro analytical techniques.” Cem. Concr. Compos., 58,
agglomeration when the GO dosage exceeds 0.04%. 114–128.
3. In the dissolution stage, the electrical resistivity of cement pastes Hou, P. K., Kawashima, S., Wang, K. J., Corr, D. J., Qian, J. S., and Shah,
incorporating NA/GO is higher compared to plain cement paste S. P. (2013). “Effects of colloidal nanosilica on rheological and
because of the high water adsorption capacity of NA/GO. How- mechanical properties of fly ash-cement mortar.” Cem. Concr. Compos.,
ever, in the setting stage, electrical resistivity decreases perhaps 35(1), 12–22.

© ASCE 04017087-8 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., -1--1


Kawashima, S., Seo, J. T., Corr, D., Hersam, M., and Shah, S. P. (2013). Onuaguluchi, O., Panesar, D. K., and Sain, M. (2014). “Properties of
“Dispersion of CaCO3 nanoparticles by sonication and surfactant treat- nanofibre reinforced cement composites.” Constr. Build. Mater., 63,
ment for application in fly ash-cement systems.” Mater. Struct., 47(6), 119–124.
1011–1023. Pan, Z., et al. (2015). “Mechanical properties and microstructure of
Konsta-Gdoutos, M. S., Metaxa, Z. S., and Shah, S. P. (2010). “Multi-scale a graphene oxide-cement composite.” Cem. Concr. Compos., 58,
mechanical and fracture characteristics and early-age strain capacity 140–147.
of high performance carbon nanotube/cement nanocomposite.” Cem. Pane, I., and Hansen, W. (2005). “Investigation of blended cement hydra-
Concr. Compos., 32(2), 110–115. tion by isothermal calorimetry and thermal analysis.” Cem. Concr. Res.,
Li, G. Y., Wang, P. M., and Zhao, X. (2005). “Mechanical behavior 35(6), 1155–1164.
and microstructure of cement composites incorporating surface-treated Park, J. H., Choudhury, A., Farmer, B. L., Dang, T. D., and Park, S. Y.
multi-walled carbon nanotubes.” Carbon, 43(6), 1239–1245. (2012). “Chemically modified graphene oxide/polybenzimidazobenzo
Li, X. Y., et al. (2016). “Incorporation of graphene oxide and silica fume phenanthroline nanocomposites with improved electrical conductivity.”
into cement paste: A study of dispersion and compressive strength.” Polymer, 53(18), 3937–3945.
Constr. Build. Mater., 123, 327–335. Sanchez, F., and Sobolev, K. (2010). “Nanotechnology in concrete—A
Li, Z., Wang, H., He, S., Lu, Y., and Wang, M. (2006). “Investigations review.” Constr. Build. Mater., 24(11), 2060–2071.
Downloaded from ascelibrary.org by University Of Florida on 04/15/17. Copyright ASCE. For personal use only; all rights reserved.

on the preparation and mechanical properties of the nano-alumina Skalny, J., and Young, J. F. (1980). “Mechanisms of portland cement
reinforced cement composite.” Mater. Lett., 60(3), 356–359. hydration.” Proc., 7th Int. Conf. on the Chemistry of Cement, Paris, 1–45.
Li, Z., Wei, X., and Li, W. (2003). “Preliminary interpretation of hydration
Sobolev, K., and Shah, S. P. (2015). “Nanotechnology in construction.”
process of portland cement using resistivity measurement.” ACI Mater. J.,
Proc., NICOM5, Springer, Cham, Switzerland.
100(3), 253–257.
Taylor, H. F. W. (1997). Cement chemistry, Thomas Telford Services,
Lootens, D., and Bentz, D. P. (2016). “On the relation of setting and early-
London.
age strength development to porosity and hydration in cement-based
materials.” Cem. Concr. Compos., 68, 9–14. Wei, X., and Li, Z. (2005). “Study on hydration of portland cement
Lu, Z. Y., Hou, D. S., Meng, L. S., Sun, G. X., Lu, C., and Li, Z. J. (2015). with fly ash using electrical measurement.” Mater. Struct., 38(3),
“Mechanism of cement paste reinforced by graphene oxide/carbon 411–417.
nanotubes composites with enhanced mechanical properties.” RSC Adv., Wei, X., and Li, Z. (2006). “Early hydration process of portland cement
5(122), 100598–100605. paste by electrical measurement.” J. Mater. Civ. Eng., 10.1061/(ASCE)
Mohammed, A., Sanjayan, J. G., Duan, W. H., and Nazari, A. (2015). 0899-1561(2006)18:1(99), 99–105.
“Incorporating graphene oxide in cement composites: A study of trans- Xie, P., Gu, P., and Beaudoin, J. (1996). “Electrical percolation phenomena
port properties.” Constr. Build. Mater., 84, 341–347. in cement composites containing conductive fibres.” J. Mater. Sci.,
Mohammed, A., Sanjayan, J. G., Duan, W. H., and Nazari, A. (2016). 31(15), 4093–4097.
“Graphene oxide impact on hardened cement expressed in enhanced Xu, Z., and Gao, C. (2011). “Aqueous liquid crystals of graphene oxide.”
freeze-thaw resistance.” J. Mater. Civ. Eng., 10.1061/(ASCE)MT ACS Nano, 5(4), 2908–2915.
.1943-5533.0001586, 04016072. Zhang, J., Qin, J., and Li, Z. J. (2009). “Hydration monitoring of cement-
Mostafa, N. Y., and Brown, P. W. (2005). “Heat of hydration of high based materials with resistivity and ultrasonic methods.” Mater. Struct.,
reactive pozzolans in blended cements: Isothermal conduction calorim- 42(1), 15–24.
etry.” Thermochim. Acta, 435(2), 162–167. Zhang, M. H., and Jahidul, I. (2012). “Use of nano-silica to reduce setting
Nazari, A., and Riahi, S. (2011). “Al2 O3 nanoparticles in concrete and time and increase early strength of concretes with high volumes of fly
different curing media.” Energy Build., 43(6), 1480–1488. ash or slag.” Constr. Build. Mater., 29, 573–580.

© ASCE 04017087-9 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., -1--1

You might also like