You are on page 1of 33

nanomaterials

Review
The Influence of Nanomaterials on the Thermal
Resistance of Cement-Based Composites—A Review
Pawel Sikora 1,2, * ID
, Mohamed Abd Elrahman 1,3 and Dietmar Stephan 1 ID

1 Building Materials and Construction Chemistry, Technische Universität Berlin, Berlin,


Gustav-Meyer-Allee 25, 13355 Berlin, Germany; abdelrahman@tu-berlin.de (M.A.E.);
stephan@tu-berlin.de (D.S.)
2 Faculty of Civil Engineering and Architecture, West Pomeranian University of Technology, Szczecin,
Al. Piastow 50, 70-311 Szczecin, Poland
3 Structural Engineering Department, Mansoura University, Elgomhouria St., Mansoura 35516, Egypt
* Correspondence: pawel.sikora@zut.edu.pl; Tel.: +49-30-314-72104

Received: 7 June 2018; Accepted: 21 June 2018; Published: 26 June 2018 

Abstract: Exposure to elevated temperatures has detrimental effects on the properties of cementitious
composites, leading to irreversible changes, up to total failure. Various methods have been used to
suppress the deterioration of concrete under elevated temperature conditions. Recently, nanomaterials
have been introduced as admixtures, which decrease the thermal degradation of cement-based
composites after exposure to high temperatures. This paper presents a comprehensive review of
recent developments related to the effects of nanoparticles on the thermal resistance of cementitious
composites. The review provides an updated report on the effects of temperature on the properties
of cement-based composites, as well as a detailed analysis of the available literature regarding the
inclusion of nanomaterials and their effects on the thermal degradation of cementitious composites.
The data from the studies reviewed indicate that the inclusion of nanoparticles in composites protects
from strength loss, as well as contributing to a decrease in disruptive cracking, after thermal exposure.
From all the nanomaterials presented, nanosilica has been studied the most extensively. However,
there are other nanomaterials, such as carbon nanotubes, graphene oxide, nanoclays, nanoalumina
or nano-iron oxides, that can be used to produce heat-resistant cementitious composites. Based on
the data available, it can be concluded that the effects of nanomaterials have not been fully explored
and that further investigations are required, so as to successfully utilize them in the production of
heat-resistant cementitious composites.

Keywords: nanomaterials; review; cement-based composites; concrete; elevated temperature; fire


resistance; mechanical properties; microstructure

1. Introduction
Concrete is a well-understood material and for most application purposes has a satisfying
endurance to elevated temperatures. It has been used successfully in various engineering structures,
providing satisfactory dimensional stability to structures in cases of fire [1], although its properties
deteriorate when exposed to elevated temperatures. Alteration of the microstructure of cement-based
composites occurs from the very beginning of the heating process, with cement paste already exhibiting
some damage even after heating up to 105 ◦ C (the standard temperature for the drying of materials) [2].
However, temperatures up to 200–300 ◦ C can still be considered relatively harmless to the stability of
cementitious composites [3]. Nonetheless, exceeding certain temperatures, as well as certain exposure
times, can lead to irreversible changes in concrete, even leading to structural collapse. Resistance
of concrete to high temperatures is highly related to the thermal conductivity and specific heat that

Nanomaterials 2018, 8, 465; doi:10.3390/nano8070465 www.mdpi.com/journal/nanomaterials


Nanomaterials 2018, 8, 465 2 of 33

are influenced by many factors including the aggregate type, mixture composition, and moisture
content [1]. Concrete provides lasting protection to reinforcing steel (which is sensitive to high
temperature), however, due to its thermal conductivity non-linear temperature distribution in the
concrete element occurs [4]. As a result, thermal transients propagate progressively within the
concrete element, leading to significant thermal gradients and thus, causing undesirable damage to
the concrete [4].
Concrete is widely used as a material for high-rise buildings, tunnels, drilling platforms, as
well as nuclear facilities (having both satisfactory thermal and shielding properties) [5–8]. However,
recent accidents involving existing structures have revealed that there is still a strong need to continue
studies, in order to understand the effects of high temperature on cement-based materials, as well as
to find methods for improving their thermal resistance. However, understanding the performance
of concrete under elevated temperature conditions is a complex issue, due to its heterogeneous
nature (consisting of cement paste and aggregates); thus, response to stress is not only dependent
on the response of individual components, but also on the interaction between those components.
For this reason, deterioration in the mechanical and physical properties of concrete depends on
individual physicochemical changes in the cement paste and aggregates, as well as on (thermal)
incompatibility between the aggregate and surrounding cement paste [9]. Therefore, to provide
satisfactory thermal resistance for cement-based composites working under or exposed to elevated
temperature conditions, mix design, including proper proportions and constituents, is a key factor
when designing the material [9]. Nevertheless, the compatibility of the constituents should be provided
not only when unheated materials are tested, but also within a range of heating temperatures.
The choice of proper aggregate plays a crucial role in providing a satisfactory thermal resistance to
mortars and concretes, as aggregate makes up 60 to 80 vol. % of the concrete [2,9]. As such, aggregates
with low mass loss and a low thermal strain coefficient of expansion, along with negligible residual
strains, are clearly preferred in the production of fire-resistant concretes [2,10]. Moreover, the rough
angular surfaces of aggregates (which improve the physical cement paste-aggregate bond), as well as
the presence of reactive silica (which improves chemical bonds), are desirable features of aggregates [9].
In general, aggregates usually remain stable up to 500–600 ◦ C, although some siliceous aggregates,
such as flint [9,11], are stable only up to 300–350 ◦ C [2,12]. It has been reported that siliceous aggregates
(i.e., quartzite, granite), have worse thermal resistance than carbonatic aggregates (such as limestone or
dolomite) [3,5,11]. This is attributable to the higher thermal expansion of carbonatic aggregates, as well
as to the quartz crystal transition, which takes place at around 573 ◦ C (low to high quartz), resulting in
volume expansion [11]. In the case of calcareous aggregates, a temperature range of 700 ◦ C to 900 ◦ C
seems to be disruptive, as a result of the decomposition of calcium carbonate; CaCO3 decomposes into
CaO (lime) and CO2 (carbon dioxide) [2,5,13].
In distinction to the expansion process of aggregates, when heated, cement paste generally exhibits
shrinkage, due to the chemical and physical loss of water, when exposed to elevated temperatures,
although up to 200 ◦ C, a slight expansion might occur. Hydrated cement paste is mainly composed
of calcium-silicate-hydrate (C–S–H) gel (the primary binding phase), calcium hydroxide (CH) and
calcium sulfoaluminates (ettringite and monosulphate), which consist of 50–60 vol. %, 20–25 vol. %
and 15–20 vol. % of solids, respectively [5].
The thermal resistance of cement paste is governed by many factors, in which the most important
are the water to cement (w/c) ratio, the C/S (calcium oxide/silicon dioxide ratio), as well as the
quantity of portlandite (CH). A paste containing a low C/S ratio and thus a low CH content is more
desirable for obtaining heat-resistant cement paste [3,9,13].
In general, the degradation process of cement paste occurs from the very beginning of the heating
process, although up to 300 ◦ C the changes are considered to be relatively small and reversible, being
thus recoverable by the so-called “rehydration process”. At 80 ◦ C, the decomposition of ettringite
begins, as well as the evaporation of physically bound water. Firstly, the capillary and free water not
influenced by the Van der Walls attraction forces, evaporate; afterward, the loss of physically and
Nanomaterials 2018, 8, 465 3 of 33

Nanomaterials 2018, 8, x FOR PEER REVIEW 3 of 32


chemically combined water, associated with the C–S–H phase, occurs [3,13]. The C–S–H decomposition
the loss
process of physically
starts and beginning
at the very chemically combined water,process,
of the heating associated with thein
resulting C–S–H phase, occurs
a reduction of its [3,13].
volume,
The C–S–H decomposition process starts at the very beginning of the heating process, resulting in a
which turns to an increment in total porosity as well as a coarsening of the pore structure [2,13].
reduction of its volume, which turns to an increment in total porosity as well as a coarsening of the
However, up to 300 ◦ C, no dramatic changes in porosity are observed, but when the temperature
pore structure [2,13]. However, up to 300 °C, no dramatic changes in porosity are observed, but when
exceeds 300 ◦ C, the porosity increases significantly [13]. Figure 1 presents the change in total porosity
the temperature exceeds 300 °C, the porosity increases significantly [13]. Figure 1 presents the change
(up to 800 ◦ C) and the area percentage of each phase of heated Portland cement paste when heated
in total porosity (up to 800 °C) and the area percentage of each phase of heated Portland cement paste
◦ C. A similar relation between temperature increment and the coarsening of the pore
upwhen
to 1000heated up to 1000 °C. A similar relation between temperature increment and the coarsening of
structure
the porecan be observed
structure can be in the caseinofthe
observed concrete. Using the
case of concrete. mercury
Using intrusion
the mercury porosimetry
intrusion method
porosimetry
(MIP), Haridharan et al. [14] have analyzed the porosity of concretes, which were exposed
method (MIP), Haridharan et al. [14] have analyzed the porosity of concretes, which were exposed to to elevated
temperature and then cooled
elevated temperature quickly
and then or slowly.
cooled quicklyWith a slow cooling
or slowly. process,
With a slow porosity
cooling increased
process, from
porosity
an increased
initial 19.24fromvol.
an % to 22.34
initial 19.24vol.
vol. %, 25.32
% to vol.
22.34 vol.%,
%,and 32.54
25.32 vol. vol. % after
%, and 32.54 exposure toexposure
vol. % after 200, 400 toand
◦ C, respectively. Likewise, after quick cooling, the porosity increased to 22.17 vol. %, 32.87 vol. %,
600200, 400 and 600 °C, respectively. Likewise, after quick cooling, the porosity increased to 22.17 vol.
and ◦ C, respectively.
%,44.54
32.87vol.
vol.%%,after
and exposure to after
44.54 vol. % 200, 400 and 600
exposure to 200, 400 and 600 °C, respectively.

Figure 1. The effect of temperature on the change of total porosity (a) and the area percentage of each
Figure 1. The effect of temperature on the change of total porosity (a) and the area percentage of
phase (b) of heated Portland cement paste after cooling. Reproduced with permission from [15].
each phase (b) of heated Portland cement paste after cooling. Reproduced with permission from [15].
Copyright Elsevier, 2013.
Copyright Elsevier, 2013.
In addition to the alternation of the pore structure while heating, when the temperature exceeds
In
200 °Caddition
the firstto the alternation
cracks in the cement of matrix
the pore canstructure while
be noticed; heating,
however, when the increment
a significant temperature exceeds
in cracks

200is observable
C the first cracks in the
only when thecement
temperaturematrixexceeds
can be 400
noticed; however,
°C. Above a significant
a temperature increment
of 400 in cracks
°C, the number
of cracks increases
is observable only when drastically with temperature
the temperature exceeds[3,5]. ◦ C. Abovewhen
400 Moreover, of 400 ◦rises,
the temperature
a temperature thermal
C, the number
incompatibilities
of cracks between aggregate
increases drastically and cement
with temperature [3,5].paste arise, when
Moreover, leading thetotemperature
propagationrises,of micro-
thermal
cracking in the interfacial transition zone (ITZ) between the cement paste
incompatibilities between aggregate and cement paste arise, leading to propagation of micro-cracking and the aggregate [3],
resulting
in the in significant
interfacial transition strength
zone loss.
(ITZ) between the cement paste and the aggregate [3], resulting in
A turning point
significant strength loss. in the properties of cement-based composites occurs in the range of 400 °C to
550A°C. At these temperatures, decomposition
turning point in the properties of cement-based of CH occurs [2,3,13]. This
composites occursprocess
in theisrange
not itself critical
of 400 ◦ C to
regarding
◦ the strength loss of concretes; however, as a result of a rapid
550 C. At these temperatures, decomposition of CH occurs [2,3,13]. This process is not itself criticallime rehydration process
(when samples are water cooled or exposed to humidity), portlandite is formed again, resulting in
regarding the strength loss of concretes; however, as a result of a rapid lime rehydration process (when
expansion, which turns to severe cracking and strength loss in samples [3]. Therefore, cement blends
samples are water cooled or exposed to humidity), portlandite is formed again, resulting in expansion,
produced with lower C/S ratios possess lower CH contents, thus decreasing the severe cracking
which turns to severe cracking and strength loss in samples [3]. Therefore, cement blends produced
phenomenon [9]. As reported by Arioz [16], exposure to a temperature of 500 °C causes irreversible
with lower C/S ratios possess lower CH contents, thus decreasing the severe cracking phenomenon [9].
changes in concretes. A further temperature increment leads to the second phase of C–S–H gel
Asdecomposition
reported by Arioz [16], exposure to a temperature of 500 ◦ C causes irreversible changes in concretes.
at 560 °C and formation of β-C2S at 600–800 °C, as well as the decomposition of poorly
A further temperature
crystalized CaCO3 to increment
CaO and CO leads to the second phase of C–S–H gel decomposition at 560 ◦ C and
2 [2,3,5,13]. A temperature of 600 °C seems to be critical for concrete,
formation ◦
resultingof inβ-C 2 S at 600–800
an extreme increaseC, inas wellobserved
cracks as the decomposition
on the surface of of the
poorly crystalized
concrete, as well CaCO 3 to CaO
as significant
and CO [2,3,5,13]. A temperature of 600 ◦ C seems to be critical for concrete, resulting in an extreme
growth 2 of porosity, resulting in a dramatic strength loss in the concrete [5,9]. After exposure to 600
increase in cracks
°C, concrete observed
loses on the surface
its load-bearing capacity,of the
andconcrete,
can thusasbe well as significant
considered growthuseless.
structurally of porosity,
A
resulting
graphicalin arepresentation
dramatic strength of the loss
changein the concrete
of phase [5,9]. After
composition in exposure
hydrated to 600 ◦ C,
cement concrete
paste (w/c = loses
0.67),its
load-bearing
as a function capacity,
of heatingandtemperature,
can thus be considered
is presentedstructurally
in Figure 2. useless. A graphical the
Table 1 summarizes representation
effects of
of the change of
temperature onphase composition
the properties in hydrated cement paste (w/c = 0.67), as a function of heating
of concrete.
Nanomaterials 2018, 8, 465 4 of 33

temperature, is presented in Figure 2. Table 1 summarizes the effects of temperature on the properties
ofNanomaterials
concrete. 2018, 8, x FOR PEER REVIEW 4 of 32

Figure
Figure 2. 2.
TheThe effect
effect ofof temperature
temperature onon
thethe change
change of of phase
phase composition
composition of of cement
cement paste
paste with
with w/cw/c
= =0.67.
0.67. Reproduced with permission from [17]. Copyright Elsevier, 2009.
Reproduced with permission from [17]. Copyright Elsevier, 2009.

Table 1. The effects of temperature on physico-chemical, visual, and mechanical properties of cement-
Table 1. The effects of temperature on physico-chemical, visual, and mechanical properties of
based composites exposed to elevated temperatures. Based on: [1–3,5,9,10,13].
cement-based composites exposed to elevated temperatures. Based on: [1–3,5,9,10,13].
Temperature
Physico-Chemical Changes Effect on Cracking Effect on Strength
Range Range
Temperature Physico-Chemical Changes Effect on Cracking Effect on Strength
• Ettringite dehydration (from 80 to 150 °C) ◦ No cracking is No effect or slight
20–100 °C • Ettringite dehydration (from 80 to 150 C
20–100 ◦ C • Evaporation

of capillary and free water
Evaporation of capillary and free water
observed
No cracking is observed
No effect or slight
strength strength
improvement
improvement
• Dehydration of calcium aluminosulphate hydrates
• Dehydration
(120–140 °C) of calcium aluminosulphate
hydrates (120–140 C) ◦ No cracking is No effect or slight
100–200 °C • Gypsum decomposition (150–170 °C)
• Gypsum decomposition (150–170 ◦ C) observed strength improvement

100–200 C • Beginning of loss of stability of cement paste No cracking is observed
No effect or slight strength
• Beginning of loss of stability of improvement
• Loss ofcement
C–S–Hpaste
interlayer water
• Loss of C–S–H interlayer water <300 °C no effect or
slight strength
• Break
• of some
Break ofsiliceous aggregates
some siliceous (approx. 350 °C)
aggregates First noticeable cracks
200–400 °C <300 ◦ C no
improvement.
effect or slight
200–400 ◦C • Dehydration of 350
(approx. ◦
C–S–H C) phase are cracks
First noticeable developed
strength
>300 improvement.
°C strength loss
• Dehydration of C–S–H phase are developed
>300 ◦ C strength
by loss by 15–40%
15–40%
• •
Decomposition of hydration
Decomposition products
of hydration products
• • Decomposition
Decomposition of Ca(OH)
of Ca(OH) 2 to +
2 to CaO CaO
H2O+ (450–550
H2 O °C) Intensification of

(450–550 C) Strength loss by
400–600 °C • Destruction of C–S–H gel cracking in
Intensification of aggregate,
• Destruction of C–S–H gel 40–60%
400–600 ◦ C • Quartz phase transformation (573 °C) from cracking
α-quartz in aggregate,
cement paste and ITZ Strength loss by 40–60%
• Quartz phase transformation (573 ◦ C) cement paste and ITZ
to β-quartz
from assisted
α-quartzwith increment
to β-quartz of volume
assisted with
• Loss ofincrement
load-bearing capacity
of volume Severe cracking, ITZ
• Second phase of C–S–H decomposition and formation cracks, starting of
• Loss of load-bearing capacity
600–800 °C of β-C2S at 600-800 °C collapsing of Strength loss by 80%
• Second phase of C–S–H decomposition
• Decomposition of poorly crystalized
and formation of β-C2 S at 600-800 CCaCO
◦ 3 Severe cracking,
structural ITZ
integrity of
cracks, starting of
600–800 ◦ C • Decarbonization
• of limestone
Decomposition of poorlyaggregate
crystalized concrete Strength loss by 80%
collapsing of structural
CaCO Loss of compressive
• Breakdown of 3
C–S–H phase integrity of concrete
• Decarbonization of limestone aggregate Loss of bond between strength or almost all
800–1000 °C • Total loss of water of hydration
aggregate and paste of compressive
• Ceramic
• Breakdown
binding of C–S–H phase Loss of compressive
strength strength
is lost or
• Total loss of water of hydration· Loss of bond between
800–1000 ◦ C almost all of compressive strength
Ceramic binding aggregate and paste
is lost
Various methods for improving the thermal resistance of cement-based composites exist: the use
of proper aggregates [11], the use of various fibers (especially polypropylene fibers—PP) [18,19], the
Various methods for improving the thermal resistance of cement-based composites exist: the use
application of surface coatings [20], and the incorporation of supplementary cementitious materials
of(SCMs)
proper [3,5,21].
aggregatesIt has[11],
beenthereported
use of various
that the fibers (especially
inclusion of SCMs,polypropylene
such as fly ashfibers—PP)
(FA) or ground[18,19],
thegranulated
applicationblast-furnace
of surface coatings [20], and the incorporation of supplementary cementitious
slag (GGFBS), can contribute to strength improvements at lower materials
(SCMs) [3,5,21].(upIt to
temperatures has
400been reported
°C), as that
well as to the inclusion
minimizing of SCMs,
strength such
loss up to 600 as
°C,fly ash (FA) or
as compared ground
to plain
granulated blast-furnace slag (GGFBS), can contribute to strength improvements at lower
Portland cement concrete. In addition, due to the pozzolanic activity of SCMs resulting in an extra temperatures
(up to 400of◦ C),
amount C–S–Has well as to minimizing
produced, along with astrength
reducedloss
amountup toof600

CH, aC,reduction
as compared
in thetosurface
plain cracks
Portland
cement concrete.
of concretes In addition,
is observed. Thus,duethetototal
the pozzolanic
porosity andactivity
averageofpore
SCMs resulting
volume in an
become extrathan
lower amount
that of
of plain cement concrete [3,5]. Furthermore, the abovementioned SCMs contribute to a decrease in
the risk of explosive spalling. Silica fume (SF) is a type of SCM which leads to an increase in the risk
Nanomaterials 2018, 8, 465 5 of 33

C–S–H produced, along with a reduced amount of CH, a reduction in the surface cracks of concretes is
observed. Thus, the total porosity and average pore volume become lower than that of plain cement
concrete [3,5]. Furthermore, the abovementioned SCMs contribute to a decrease in the risk of explosive
spalling. Silica fume (SF) is a type of SCM which leads to an increase in the risk of spalling. This
effect is attributable to the filler effect and the pozzolanic activity of SF, resulting in the production
of a cement matrix with low permeability. Hence, as a result of high vapor pressure during concrete
heating, the risk of thermally-induced explosive spalling increases [3,5,21,22].
In the last decade, the use of nanomaterials, as admixtures for improving the thermal resistance of
cementitious composites, has gathered the noticeable attention of researchers. The use of nanomaterials
has been widely promoted due to their superior reactivity, as compared to similar materials available at
the micro-scale [23–25]. Nanomaterials affect the properties of cementitious materials in two ways: by
their hydration seeding effect and their filling effect [23,26,27]. Due to their small size, nanomaterials
act as seeds during the precipitation process, accelerating the hydration process and enabling C–S–H
gel formation around the nanomaterials. In addition to their chemical effect, nanomaterials exhibit
a particular physical effect: the so-called nano-filling effect, which leads to densification of the pore
system [23]. As a result, the pore structure becomes more compact and homogenous, which translates
to a significant improvement in mechanical properties and durability. However, the incorporation
of nanomaterials might lead to noticeable consistency decrement [24] and undesirable compaction
(as a result of high surface area to volume ratio of nanoparticles) of cement matrix resulting in cement
matrix of low permeability, thus increasing susceptibility to cracking during thermal load (as in case of
SF-incorporated concrete) [28].
Even though there are numerous review papers related to the effects of nanoparticles on
cement-based composites in the ambient (unheated) state [23,24,29–33], there is a lack of review
papers summarizing the effects of nanomaterials on the performance of cementitious composites under
elevated temperature conditions. As such, this paper aims to summarize recent developments in this
field, as well as to address the possible directions which are necessary for developing heat-resistant
nanomodified cementitious composites.

2. The Effect of Nanosilica on the Thermal Resistance of Cement-Based Composites


Among all the nanomaterials used to modify the properties of cement-based composites,
nanosilica (NS) has most attracted researchers’ attention. This interest can be attributed to NS’s
superior reactivity, as compared to other SCMs. NS fills the spaces between the particles of C–S–H
gel, acting as a nano-filler and refining its microstructure. Also, due to its high pozzolanic activity,
nanosilica reacts with CH, producing more C–S–H, which results in a denser cement matrix and thus
improved strength and durability of cement-based composites [29]. The effect of nanosilica on the
mechanical properties and durability of cementitious composites has been comprehensively revised
by many authors [23,24,29–31,33,34].
Studies undertaken by various researchers have shown that nanosilica can be utilized in
cement-based composites working at elevated temperatures. The available studies related to the effects
of NS on the thermal resistance of cement-based composites are summarized in Table 2. The inclusion
of nanosilica in the mix improves the resistance of cementitious composites to elevated temperatures,
since such composites retain their high residual compressive strength and demonstrate reduced crack
length and width, as compared to plain samples. It has been reported that nanosilica improves the
chemical stability of C–S–H (due to a reduction of calcium leaching from C–S–H) and increases the
average silicate chain length of C–S–H and the volume of high-density C–S–H. As a result, the thermal
degradation of the C–S–H phase during heating can be hindered [35–37]. Moreover, in the lower
temperature range (up to approximately 300–400 ◦ C), due to the high pozzolanic activity of NS,
nanoparticles promote the hydration of anhydrous cement grains during the internal autoclaving
process and as a result more C–S–H is produced. Subsequently, a higher strength increment is obtained
for samples containing NS than for plain cementitious composites. Also, due to the consumption
Nanomaterials 2018, 8, 465 6 of 33

of portlandite and the formation of additional C–S–H, nanosilica contributes to the filling of some
open pores, and therefore a reduction of surface cracks can be observed after thermal exposure [28,38].
Furthermore, Kumar et al. [39] have reported that the incorporation of NS decreases the thermal
conductivity of concretes and that heat transfer in NS-modified specimens is thus delayed. It means
that more time is required to reach the desired temperature at the core of specimens so that the rate of
thermal degradation of specimens is delayed, as compared to plain specimens.
Nanomaterials 2018, 8, 465 7 of 33

Table 2. Summary of available studies related to the effects of nanosilica (NS) on the thermal resistance of cement-based composites.

Size [nm] (Surface Area Heating Temperature


Amount [wt. %] Type of Material Tested Mechanical Properties Ref.
[m2 /g]) [◦ C]
15 1, 4 200–400–650–850–950 Cement paste Compressive strength [40]
30 ± 5 (195) 0.5, 1, 2, 3 300–600–800 Cement paste Compressive strength [41]
30 nm (180–300) 5 100–200–300–400–500 Cement paste Compressive strength [42]
12–15 (n/a) 1.5, 3 500–800 Cement paste and cement mortar Compressive strength [43]
n/a 2.5, 5, 7.5 400–700 Cement mortar Flexural strength, compressive strength [44]
n/a 2.5, 5, 7.5 400–700 Cement mortar Flexural strength, compressive strength [45]
n/a 1, 2, 3, 4, 5 200–400–600–800 Cement mortar Flexural strength, compressive strength [28]
30-70 (n/a) 3 200–400–600–800 Concrete Compressive strength, split-tensile strength [39]
45 (60) 1.5, 3, 4.5 400–600–800 Concrete Compressive strength, tensile strength [46]
16–20 (170–200) 5 200–500–800 Concrete Compressive strength [47]
5–8 (n/a) 1, 2, 3, 4, 5 100–200–400–600–800 Concrete Compressive strength [48]
n/a 1, 2, 3, 4, 5 200–400–600 Concrete Flexural strength, compressive strength [49]
30 (200 ± 10) 2 200–400–600–800 Concrete Uniaxial tensile strength [50]
Compressive strength, tensile splitting
30 (200 ± 10) 2 200–400–600–800 Concrete [51]
strength, flexural strength
Nanomaterials 2018, 8, 465 8 of 33

2.1. The Effect of Nanosilica on Mass Loss


Heikal et al. [40] have analyzed the mass loss (up to 950 ◦ C) of ordinary Portland cement (OPC)
pastes containing 1 and 4 wt. % of NS and cement pastes containing 25, 50 and 65 wt. % of cement
replaced with blast-furnace slag (GGBS) and a fixed amount of NS—4 wt. %. It was found that in the
case of OPC pastes, mass loss was higher for samples without an NS admixture, as a result of the higher
CH amount in the cement paste. Incorporation of 1 wt. % or 4 wt. % of NS contributed to a decrement
in the amount of calcium hydroxide (CH) and calcium carbonate (CC− ) in cement pastes, but the
lowest mass loss of OPC cement pastes was reported for the sample containing 1 wt. % of NS. When
GGBS and 4 wt. % of NS was introduced to the cement paste, a higher mass loss of GGBS-incorporated
specimens was reported, in the range of 200–400 ◦ C, as compared to OPC pastes. This was attributed to
the activation of GGBS (as a result of self-autoclaving conditions), nano-filling, and the high pozzolanic
activity of NS. This phenomenon resulted in the production of an additional C–S–H phase from the
liberated CH. However, when the temperature exceeded 400 ◦ C, the mass loss of GGBS-incorporated
cement pastes was generally lower than that of OPC pastes. Maheswaran et al. [43] have analyzed
the mass loss (based on thermogravimetric analysis) of cement pastes containing 20 wt. % of cement
replaced with lime sludge and an addition of nanosilica (1.5 wt. % and 3 wt. %) with a fixed w/b = 0.4.
After exposure to temperatures up to 400 ◦ C, the mass loss of samples containing NS was slightly
higher than that of plain samples, while in the case of exposure to a temperature range of 400 ◦ C to
600 ◦ C, the mass loss of samples containing NS was lower, as a result of CH consumption by the NS.
In another study [28], the effects of NS admixture in amounts from 1–5 wt. %, on the mass
loss of cement mortars (w/c = 0.5) containing three different types of aggregates (quartz, magnetite
and barite), were evaluated. The samples were subjected to elevated temperatures of up to 800 ◦ C.
No significant effect of the NS content or heating temperature on the mass loss of cement mortars was
observed, with the mass loss being highly related to the type of aggregate used. Bastami et al. [46]
have evaluated the effects of temperature (up to 800 ◦ C) on the mass loss of two types of high-strength
concretes (HSCs), with w/b = 0.25, containing fixed amounts (30 kg/m3 or 60 kg/m3 ) of SF or SF/NS.
In NS-modified concretes, SF was replaced with NS in amounts of 1.5, 3, and 4.5% by mass of cement
(1.41, 2.83 and 4.2 wt. % of binder mass, respectively). The results of these studies were inconclusive,
although at up to 400 ◦ C the mass loss of was slightly lower or negligible for samples containing NS.
After exposure to 600 ◦ C and 800 ◦ C, in comparison with plain concrete, a slightly lower mass loss was
observed for concretes containing 30 kg/m3 of SF/NS mixture, while in concrete specimens containing
60 kg/m3 of SF/NS mixture, the mass loss increased.
In a study presented by Shah et al. [47], the mass loss of three HSCs (w/b ranging from 0.25 to
0.3), containing 5 wt. % of NS, 10 wt. % of SF or a mixture of 5 wt. % NS and 5 wt. % SF, exposed to
200, 500 and 800 ◦ C, was evaluated. After exposure to 200 ◦ C, the mass loss values for all concretes
tested was relatively similar, varying from 2.47 to 3.17%. After exposure to 500 ◦ C, samples containing
NS exhibited a higher mass loss than SF-modified specimens, although the highest mass loss was
reported for samples containing a mixture of SF and NS. After exposure to 800 ◦ C, the lowest mass loss
was reported for NS-modified samples, while mass loss dramatically increased for samples containing
SF and a SF/NS mixture. Rathi and Modhera [48] have analysed the thermal resistance of HSCs with
cement replacements; from 10 up to 30 wt. % of FA (in increments of 5 wt. %) and with 1 to 5 wt. %
of NS (in 1 wt. % increments). For comparison, plain concretes without NS or FA were prepared.
The authors reported that in all the mixtures tested and under all testing temperatures (from 100 to
800 ◦ C), the mass loss was found to be smallest for concretes containing 3 wt. % of NS. Kumar et al. [39]
have evaluated the properties of two HSCs containing 0 and 3 wt. % NS (w/c = 0.29), exposed to
200, 400, 600 and 800 ◦ C. The authors reported that >200 ◦ C, the mass loss of NS modified concrete
was decreased slightly, as a result of the lesser amount of free water available to evaporate from the
concrete (consumed as a result of the formation of a higher quantity of hydration products). As a result,
after exposure to 400 ◦ C, the mass loss of NS-modified concretes increased slightly, because of the
dehydration of the higher amount of C–S–H, formed as a result of the presence of NS. Further exposure
Nanomaterials 2018, 8, 465 9 of 33

to 600 and 800 ◦ C, resulted in a slightly lower mass loss of NS-modified concretes, as a result of delayed
thermal degradation, attributed to the decreased thermal conductivity of NS-modified concretes.
From the references cited above, it can be concluded that the effect of NS on the mass loss of
cement-based composites is still controversial and is highly dependent on mix compositions as well as
on the type of composite tested (e.g., paste, mortar or concrete). It is difficult to draw a clear conclusion
about the influence of the incorporation of nanosilica on mass loss or the stability of cement composites
when subjected to high temperatures. Consequently, this issue requires further investigation, to
determine a distinct trend regarding the performance of various dosages of NS on the stability of
different cement-based materials.

2.2. The Effect of Nanosilica on the Mechanical Properties of Cement-Based Composites

2.2.1. Flexural Strength


The effect of NS on the flexural strength of cement-based composites is still in dispute, as
flexural strength is highly sensitive to micro-cracking during heating [28]. In work undertaken
by Ibrahim et al. [45], the effect of NS (in the amounts of 0, 2.5, 5 and 7.5 wt. %) on the flexural strength
of cement mortars, containing OPC and PP fibers exposed to high temperatures, was determined.
Samples were cured for 3, 7, and 28 days and then exposed to 400 and 700 ◦ C. In general, the authors
reported an improvement in flexural strength (after all curing periods) in unheated cement mortars,
when NS was present. After exposure to 400 ◦ C, the flexural strength of all the cement mortars
decreased, as a result of thermally induced stress leading to increased cracking and pressure in gel
pores. However, samples containing 2.5 wt. % and 5 wt. % of NS (after 28 days of curing) exhibited
higher flexural strength retention than plain control samples. On the other hand, in the case of samples
containing 7.5 wt. % of NS, flexural strength was slightly lower than that of the control samples. After
exposure to 700 ◦ C, all the samples lost almost all of their strength, and no significant effect of NS was
observed. In another study, carried out by the same authors [44], the effect of replacing FA with NS
(in the amounts of 0, 2.5, 5, and 7.5 wt. %), on the properties of high-volume fly ash cement mortars
(25, 35 and 45 wt. % of FA) exposed to high temperature, was evaluated. Samples were exposed
to 400 and 700 ◦ C. Similarly to the abovementioned studies [45], positive effects of NS-modified
samples on mechanical properties (after 3,7 and 28 days of curing), in unheated states, were observed.
The study showed that the partial replacement of FA with NS contributes significantly to the retention
of flexural strength loss, after exposure to 400 ◦ C. The best performance was exhibited by cement
mortar containing a mixture of 37.5 wt. % FA and 7.5 wt. % NS (total 45 wt. %). After exposure to
700 ◦ C, these samples lost most of their flexural strength, with their strength values varying between
20–30% of the initial strength.
In distinction to the abovementioned works, other studies have reported rather negligible effects
of NS on flexural strength [28,49,51]. Yan et al. [51] have analyzed the effect of NS on the flexural
strength of steel-fiber reinforced concrete (SFRC), exposed to temperatures of up to 800 ◦ C. They
reported an increment in the initial flexural strength of concrete containing NS, as well as a slightly
higher flexural strength after exposure to 200 ◦ C, compared to plain SFRC. Beyond 200 ◦ C, the effect of
NS seemed to be negligible. In another study [49], the effect of 200, 400, and 600 ◦ C on the flexural
strength of two types of concretes with w/c = 0.25 (containing 350 and 450 kg/m3 of cement), with
the addition of 15 wt. % of SF and 1–5 wt. % of NS, was examined. Adding NS improved the flexural
strength of unheated concrete specimens, although no significant changes (<10%) in residual flexural
strength, between specimens exposed to temperature, were reported. A negligible effect of NS addition
(1–5 wt. %), on the flexural strength of cement mortars containing different types of aggregates
(quartz, magnetite, and barite) exposed to temperatures up to 800 ◦ C, has also been reported in a study
undertaken by Horszczaruk et al. [28]. A slight variation (decrement) in the flexural strength of cement
mortars containing barite aggregate and nanosilica was reported, although this effect was attributed to
Nanomaterials 2018, 8, 465 10 of 33

the properties of the barite aggregate, which has low thermal resistance and cracking potential under
thermal stress.
It is difficult to draw any clear conclusions, regarding the effects of NS on the flexural strength
of cementitious composites, from the works cited above. While some authors have reported flexural
strength improvements [44,45] in the presence of NS under elevated temperatures, others have reported
only negligible effects [28,49,51]. As such, further work is required.
Nanomaterials 2018, 8, x FOR PEER REVIEW 9 of 32
2.2.2. Tensile Strength
It is difficult to draw any clear conclusions, regarding the effects of NS on the flexural strength
Tensile and split-tensile strength have been analyzed in several works [39,46,50,51]. In the work
of cementitious composites, from the works cited above. While some authors have reported flexural
done by Bastami et al. [46], the
strength improvements effect
[44,45] in of
the temperature
presence of NS(up under 800 ◦ C)temperatures,
to elevated on the tensile strength
others have of HSC
(w/b = 0.25), containing
reported 30 kg/m
only negligible 3 of SF replaced
effects [28,49,51]. As such,withfurtherNS in isthe
work amounts of 0, 1.5, 3 and 4.5 wt. %
required.
(by mass of cement), was determined. The best performance was observed for samples containing
2.2.2. Tensile Strength
4.5 wt. % of nanosilica. NS-modified samples exhibited higher tensile strength retention, as compared
Tensile and split-tensile strength have been analyzed in several works [39,46,50,51]. In the work
to plain concrete samples: from 8.14 to 3.14%, 41.19 to 36.17%, and 70.06 to 66.35%, after exposure
done by Bastami et al. [46], the effect of temperature (up to 800 °C) on the tensile strength of HSC
to 400, 600(w/b
and ◦ C, respectively.
800 containing
= 0.25), 30 kg/m3 of In anotherwith
SF replaced workNS in[39], the effects
the amounts of 3various
of 0, 1.5, and 4.5 wt. heating
% (by regimes
on the split-tensile
mass of cement), strength of HSCs The
was determined. (w/cbest=performance
0.29), containing
was observed 0 and 3 wt. %
for samples of NS, 4.5
containing were
wt. analyzed.
The samples were exposed to temperatures from 200 to 800 C, with 0 min, 1 h and 2 h to
% of nanosilica. NS-modified samples exhibited higher tensile ◦strength retention, as compared of constant
plain concrete samples: from 8.14 to 3.14%, 41.19 to 36.17%, and 70.06 to 66.35%, after exposure to
heating at400,
the600 desired temperature. It was observed that the split-tensile
and 800 °C, respectively. In another work [39], the effects of various heating regimes on the
strength of NS-modified
specimenssplit-tensile
increased up toof temperatures
strength HSCs (w/c = 0.29),of 400 ◦ C0 and
containing (in 3allwt.heating
% of NS, were regimes),
analyzed.while plain concrete
The samples
exhibited split tensile to
were exposed strength
temperaturesimprovement
from 200 to 800 only °C,up
with to0 200 ◦
min, 1C. Afterwards,
h and theheating
2 h of constant strength of the plain
at the
desired
concrete was lower temperature.
than that of It was observed that
the unheated the split-tensile
specimen. After exposure 400 ◦ C for 2specimens
strength oftoNS-modified h, the split-tensile
increased up to temperatures of 400 °C (in all heating regimes), while plain concrete exhibited split
strength oftensile
NS-modified HSC specimens was higher, by 13%, than that of the unheated specimens,
strength improvement only up to 200 °C. Afterwards, the strength of the plain concrete was
while the control
lower than HSCthat lost
of theapproximately
unheated specimen. 30% of exposure
After its strength.to 400The
°C for authors linked these
2 h, the split-tensile findings with
strength
the formation of a high-density
of NS-modified HSC specimens volume fraction
was higher, of C–S–H
by 13%, than that (which was thespecimens,
of the unheated result ofwhile
increased
the C–S–H
silicate chain length), which thus improved the stability of C–S–H. After exposure to 600 C and 800 ◦ C,
control HSC lost approximately 30% of its strength. The authors linked these findings with◦ the
formation of a high-density volume fraction of C–S–H (which was the result of increased C–S–H
split-tensile strength dropped significantly and the differences between the control and NS-modified
silicate chain length), which thus improved the stability of C–S–H. After exposure to 600 °C and 800
concrete decreased.
°C, split-tensile After exposure
strength dropped 800 ◦ C, theand
to significantly differences
the differencesbetweenbetween samples were
the control and<10%.
NS- In work
undertaken by Yanconcrete
modified et al. [51], the effect
decreased. After of 2 wt. %
exposure of NS
to 800 °C, admixture
the differences onbetween
the split-tensile
samples were strength
<10%. of SFRC,
In work
after exposure to undertaken
200, 400, 600,by Yan et al.
and ◦ C,the
800[51], waseffect of 2 wt. %
analyzed of NS admixture
(Figure 3). Up toon400 the ◦split-tensile
C, an increment in
strength of SFRC, after exposure to 200, 400, 600, and 800 °C, was analyzed (Figure 3). Up to 400 °C,
the split-tensile strength of concrete was observed, after which a loss of strength was reported. It was
an increment in the split-tensile strength of concrete was observed, after which a loss of strength was
observed that up to ◦ C, NS-modified SFRC specimens exhibited higher strengths than the control
600observed
reported. It was that up to 600 °C, NS-modified SFRC specimens exhibited higher strengths
SFRC specimen. The highest
than the control split-tensile
SFRC specimen. strength
The highest values
split-tensile of concrete
strength values ofspecimens were were
concrete specimens reported after
reported ◦ after exposure to 400 °C, with the NS-modified SFRC and
exposure to 400 C, with the NS-modified SFRC and control SFRC exhibiting strengths of 5.0 MPa and control SFRC exhibiting strengths
of 5.0 MPa and 4.5 MPa, respectively. After exposure to 600 °C, NS-modified SFRC exhibited a
4.5 MPa, respectively. After exposure to 600 ◦ C, NS-modified SFRC exhibited a strength similar to that
strength similar to that of the unheated sample, while the control SFRC had a lower strength than
of the unheated
that of thesample, while unheated
corresponding the control SFRC
sample. Afterhad800 a
°C,lower strengthstrength
the split-tensile than thatof allof
thethe corresponding
concretes
unheated sample.
was similar. ◦
After 800 C, the split-tensile strength of all the concretes was similar.

Figure 3. Split-tensile strength


Figure 3. Split-tensile of normal
strength concrete,
of normal steel-fiber
concrete, steel-fiber reinforced
reinforced concrete
concrete (SFRC)
(SFRC) and SFRCand SFRC
modified
modified with NS,with
as aNS, as a function
function of temperature.Redrawn
of temperature. Redrawn from from[51].
[51].
Nanomaterials 2018, 8, 465 11 of 33

In a subsequent work [50], the effect of 2 wt. % NS admixture on the axial tensile strength of
SFRC, after exposure to 200, 400, 600, and 800 ◦ C, was analyzed. The authors reported significant
strength improvement in NS-modified SFRC specimens, in unheated states, as compared to a control
SFRC. In all samples tested up to 400 ◦ C, strength improvements were reported. However, after
thermal exposure, the strength ratio between the NS-modified SFRC and the control SFRC decreased.
Nevertheless, NS-modified specimens were superior to control SFRC specimens, up to a temperature
of 800 ◦ C.
From the references cited above, it is safe to conclude that the inclusion of NS in cement-based
composites is beneficial for improving the tensile strength of composites working under elevated
temperature conditions. A significant improvement of tensile strength is observed when exposed to
temperatures up to 400 ◦ C, while further increments of temperature result in a decrease in the ratio
between NS-modified and plain specimens. In general, it can be stated that NS reduces the tensile
strength loss resulting from exposure to high temperatures, with different grades of change, depending
on the temperature applied.

2.2.3. Compressive Strength of Cement Pastes


The effect of NS on the compressive strength of cement pastes exposed to elevated temperatures
has been reported in several works [40–42,52]. Heikal et al. [52] analyzed the effects of 1–6 wt. % of NS
as a cement replacement in OPC pastes and pastes with GGBS, with specimens heated up to 1000 ◦ C.
Generally, it was observed that incorporation of NS in cement pastes helped to retain the strength of
cement pastes up to 1000 ◦ C, although a much more beneficial effect of NS was observed in cement
pastes containing GGBS and 4 wt. % of NS. In another work [40], the effects of NS in the amounts
of 1 and 4 wt. % and GGBS from 0, 25, 50, and 65 wt. %, as a cement replacement, on the thermal
resistance of cement pastes with or without superplasticizer, exposed to temperatures up to 950 ◦ C,
were analyzed. The study showed that the incorporation of superplasticizer, together with NS, has the
most beneficial effect in improving the thermal resistance of cement pastes. In the case of OPC paste,
1 wt. % of NS was established as the most beneficial amount for improving heat resistance up to 400 ◦ C.
In the case of GGBS-modified pastes, a positive effect of NS was observed up to 650 ◦ C. Lim et al. [42]
analyzed the effects of 5 wt. % of NS addition, on the thermal resistance of cement pastes exposed to
100–500 ◦ C, with two different cooling regimes: cooling down to room temperature and prolonged
heat treatment at 50 ◦ C for 3 days. The results were compared with plain OPC specimens: after cooling
down from all heating temperatures, NS-modified specimens exhibited higher compressive strength,
by 7 to 20%, as compared to control OPC specimens. The most noticeable difference in compressive
strength between samples was observed after exposure to temperatures of 400 ◦ C and 500 ◦ C, with the
values being 20% and 17%, respectively. When samples were cooled after prolonged heat treatment
regimes, a significant change in performance between NS-modified and plain OPC specimens was
observed, after exposure to 500 ◦ C. The study showed that plain OPC samples broke down completely
and lost their strength, while samples containing NS retained almost 80% of their initial strength.
El-Gamal et al. [41] have evaluated the effects of 0.5–3 wt. % NS admixture on the properties of OPC
cement pastes, exposed to temperatures of 300, 600, and 800 ◦ C (Figure 4). Two different cooling
regimes, gradually in air and water quenching, were evaluated. In the case of the first testing regime
(Figure 4a), all of the samples tested exhibited a slight strength increment after exposure to 300 ◦ C,
although the samples containing 0.5, 1, 2, and 3 wt. % exhibited strength improvements by 8, 9, 14,
and 12% (relative to the strength of unheated samples), while the control OPC exhibited only a 3%
improvement in strength. After exposure to 600 ◦ C, all the specimens tested, exhibited strength losses,
although samples containing 0.5, 1, 2, and 3 wt. % of NS retained 77, 75, 64, and 63% of their initial
strength (respectively); for the control sample only 51% of the initial strength was retained. After
exposure to 800 ◦ C, all the samples tested exhibited only 31–35% of their initial strength, with no
significant differences between samples being observed. In the case of a rapid cooling regime through
water quenching (Figure 5b), the drop in strength was observed at all the testing temperatures, while
Nanomaterials 2018, 8, 465 12 of 33

Nanomaterials 2018, 8, x FOR PEER REVIEW 11 of 32


at a temperature of 800 ◦ C the samples lost their strength and broke down. Nevertheless, samples
Nanomaterials 2018, 8, x FOR PEER REVIEW 11 of 32
containing NSafter
strength retained
exposure a higher
to 300compressive strength
°C and 600 °C, after
with the exposure
authors to 300 ◦that
concluding C and
the 600
◦ C,
optimal with
NS the
authors concluding
addition
strength value isthat
1 wt.the
after exposure %.tooptimal
300 °C NS
andaddition valuethe
600 °C, with is 1authors
wt. %.concluding that the optimal NS
addition value is 1 wt. %.

Figure 4. Compressive strength of cement pastes containing NS, after heating and cooling: (a)
Figure 4. Compressive strength of cement pastes containing NS, after heating and cooling: (a) gradually
gradually
4. in air, (b) suddenly in water. Reproduced containing
with permission fromheating
[41]. Copyright Springer
in air,Figure Compressive
(b) suddenly in water.strength of cement
Reproduced with pastes NS, after
permission from [41]. Copyright and cooling:
Springer (a) 2018.
Nature,
Nature, 2018.
gradually in air, (b) suddenly in water. Reproduced with permission from [41]. Copyright Springer
Nature, 2018.
2.2.4.2.2.4.
Compressive Strength
Compressive of of
Strength Cement
CementMortar
Mortar
2.2.4.effects
The Compressive
The of NS
effects Strength
on on
of NS thethe ofthermal
Cement
thermal Mortar of
resistance
resistance of cement mortarshas
cement mortars has also
also beenbeen analyzed
analyzed in several
in several
studies
studies [28,43–45].
[28,43–45]. In In
a a
studystudy undertaken
undertaken by
by Horszczaruk
Horszczaruk et
et al.
al.[28], the
[28], effect
the
The effects of NS on the thermal resistance of cement mortars has also been analyzed in several of
effect NS
of as
NS anas admixture
an admixture
(from
1 to 51 [28,43–45].
(fromstudies to 5 wt.
wt. %) on%) on
theathe
In compressive
compressive
study undertaken strength
strength of cement
of cementmortars
by Horszczaruk mortars
et containing
al. [28], the effecteither
containing NSquartz,
ofeither magnetite
as quartz, magnetite
an admixture
or barite aggregate
(fromaggregate
or barite 1 to 5 wt. %) and and
on exposedexposed
the compressive to temperatures up
strength of cement
to temperatures to 800
up to 800 °C,

mortars was determined.
containing
C, was The study
either quartz,
determined. showed
magnetite
The study showed
that
or thereaggregate
barite was an optimum
and amount
exposed to of NS (3 wt. up
temperatures %),towhich
800 was
°C, most
was beneficialThe
determined. for improving
study the
showed
that there was an optimum amount of NS (3 wt. %), which was most beneficial for improving the
compressive
that therestrengthstrength
was an optimum of cement
amountmortars
of NS containing quartz
(3 wt. %),quartz and
which and magnetite
was magnetite aggregate.
most beneficial The beneficial
for improving the
compressive of cement mortars containing aggregate. The beneficial
effect of NS was
compressive mainly
strength of noticeable
cement mortarsup to acontaining
temperature of 400
quartz and°C.magnetite
After thisaggregate.
point, the effect of NS on
The beneficial
effectcompressive
of NS was mainly noticeable up to a temperature 400 ◦ C. After
ofExamples this point, the effect of NS
effect of NS was mainly noticeable up to a temperature of 400 °C. After this point, the effect ofon
strength was less significant or negligible. of the effects of NS NS the
on
on compressive strengthofwas
compressive strength cementless significant or negligible. Examples of the effects of NS in on the
compressive strength was less mortars
significantwithorquartz and magnetite
negligible. Examples aggregates have
of the effects been
of NS plotted
on the
compressive strength
Figure 5. In the
compressive of cement
case ofofcement
strength mortars
cementmortars with
mortarscontaining quartz
with quartz and
barite magnetite aggregates
aggregate, aggregates
and magnetite the inclusion have
of NS
have been plotted
beencontributed
plotted in in
Figure 5. In the
to increased
Figure case of
cracking
5. In the cement
case ofpotential mortars containing
(as compared
cement mortars barite
to pristine
containing aggregate,
barite samples),
aggregate,with the inclusion
the compressive
the inclusion of NS contributed
strength of
of NS contributed
specimens
to increased
to increased after
cracking heating
crackingpotentialthus being
potential (as lower.
(ascompared
compared However,
to this phenomenon
to pristine
pristine samples),
samples), occurred
with
with thethe due to thestrength
compressive
compressive highstrength
water
of of
absorption
specimens
specimens of
afterafter the
heating barite
heating thus aggregate,
being
thus whichHowever,
beinglower.
lower. resulted inthis
However, undesired
this phenomenon
phenomenon compaction of the
occurred
occurred duecement
due matrix.
to high
to the the high Aswater
water
such, the
absorption
absorption ofeffect
of the
the ofbarite
NS is
barite dependentwhich
aggregate,
aggregate, on theresulted
which type of aggregate
resulted in in
undesired used.
undesired compaction
compaction of the of
cement matrix. As
the cement matrix.
such, the effect of NS is dependent on the type
As such, the effect of NS is dependent on the type of aggregate used. of aggregate used.

Figure 5. Comparison of residual compressive strengths of cement mortars containing quartz and
magnetite
Figure aggregate, with
5. Comparison and without NS. Reproduced
strengths of with permission from [28].quartz
Copyright
Figure 5. Comparison of of residual
residual compressive
compressive strengths ofcement
cementmortars containing
mortars containing and
quartz and
Elsevier, 2017.
magnetite aggregate, with and without NS. Reproduced with permission from [28]. Copyright
magnetite aggregate, with and without NS. Reproduced with permission from [28]. Copyright
Elsevier, 2017.
Elsevier, 2017.
Nanomaterials 2018, 8, 465 13 of 33

Ibrahim et al. [45] have analyzed the effects of temperature on the compressive strengths of OPC
mortars containing PP fibers and NS in the amounts of 0, 2.5, 5, and 7.5 wt. %. Their samples were
cured for 3, 7, and 28 days and then exposed to temperatures of 400 and 700 ◦ C. The compressive
strengths of the cement mortars increased after exposure to 400 ◦ C, but higher strength increments
were reported for specimens containing NS. Strength gains were much more noticeable in the early
stages of curing, as a result of accelerated pozzolanic reactions induced by heating in the presence
of NS, which led to the production of higher amounts of C–S–H phase than in plain mortar. After
exposure to 700 ◦ C, a remarkable reduction in compressive strength was observed in all specimens,
with the effect of NS being negligible. In another study [44], the authors detected a synergistic effect of
FA and NS in the production of heat-resistant mortar. High-volume fly ash cement mortars (w/b = 0.4),
containing 25, 35, and 45 wt. % of cement replaced with FA and 2.5, 5, and 7.5 wt. % of NS, were
produced. The specimens were exposed to 400 and 700 ◦ C, after 3, 7, and 28 days of curing. The authors
reported that after exposure to 400 ◦ C, all their specimens exhibited an increase in compressive strength,
although improvements were more noticeable for specimens containing NS. After exposure to 700 ◦ C,
the samples exhibited significant strength loss, but an optimal combination of FA and NS (37.5 wt. %
and 7.5 wt. %, respectively), enabled the production of mortar that exhibited compressive strength
comparable to that of the unheated specimen.
Results contrary to those presented above, have been reported by Maheswaran et al. [43]. In this
study, the compressive strengths of cement mortars (cured for 3, 7, and 28 days) and containing 20 wt. %
of cement replaced with lime sludge and NS (1.5 and 3 wt. %) or SF (3 and 6 wt. %), after exposure to
500 and 800 ◦ C, were determined. Generally, after exposure to 500 ◦ C (regardless of the time of curing),
specimens containing NS exhibited more strength loss than the control sample. The only exception
was mortar containing 1.5 wt. % of NS, cured for 3 days. In addition, samples containing NS showed
worse performance than those containing SF. After exposure to 800 ◦ C, specimens containing NS
exhibited worse mechanical performance than the plain control sample. However, their performance
was better than that of SF-incorporated specimens, which exhibited spalling and lost their strength,
while NS-incorporated samples exhibited higher strength and were not damaged.

2.2.5. Compressive Strength of Concrete


The effect of NS on the compressive strength of concretes, after thermal exposure, have also been
analyzed in several studies [39,46–49,51]. Kumar et al. [39] analyzed the effect of thermal exposure
(from 200 to 800 ◦ C) of HSCs (w/c = 0.29), containing 0 and 3 wt. % of NS, on compressive strength.
The samples were heated under three different regimes, with varying times of maintaining the desired
temperature (0 min, 1 h and 2 h). The authors observed that in NS-incorporated concretes, compressive
strength increased up to 400 ◦ C (in all heating regimes), while plain concrete exhibited compressive
strength improvements only up to 200 ◦ C. Beyond 200 ◦ C, the strength of the reference concrete was
lower than that of the unheated specimen. It is worth noting, that the relative compressive strength of
NS-modified HSC, after exposure to 400 ◦ C (with a 2 h heating regime), increased by 40%. This effect
was attributed to the improved stability of the C–S–H phase and higher volume of high-density C–S–H.
Further thermal exposure (600 and 800 ◦ C), led to a gradual strength decrement of the specimens,
with NS-modified samples exhibiting only slightly higher relative compressive strength than plain
HSC (with a difference less than 10%). Rathi and Modhera [48] analyzed the effects of HSCs with
cement replacements: from 10 wt. % to 30 wt. % of FA (increments of 5 wt. %), or from 1 to 5 wt. % of
NS (increment steps of 1 wt. %). For comparison, plain concretes without NS or FA were prepared.
The concrete samples were exposed to temperatures of up to 800 ◦ C. The study showed that the
incorporation of NS in the amount of 3 wt. % was beneficial in improving the thermal resistance of
concretes, up to 800 ◦ C. In the presence of NS, a compressive strength increment was observed, up
to 400 ◦ C. Moreover, the best performance among all the concretes tested, was exhibited by concrete
based on a mixture of FA (20 wt. %) and 3 wt. % of NS. Yan et al. [51] have characterized the effects of
2 wt. % of NS admixture on the compressive strength of SFRC exposed to 200, 400, 600 and 800 ◦ C.
Nanomaterials 2018, 8, 465 14 of 33

Nanomaterials 2018, 8, x FOR PEER REVIEW 13 of 32


The results were compared with plain SFRC and normal concrete, showing that the presence of NS
remarkably increases the compressive strength of concrete (by 41%), up to 400 ◦ C, as compared to
concrete, showing that the presence of NS remarkably increases the compressive strength of concrete
reference(byconcrete ◦ ◦
41%), up(Figure
to 400 °C, 6). asBeyond
compared400to reference
C, strength decreased
concrete (Figure and at 600400C °C,
6). Beyond thestrength
compressive
strengthdecreased
was comparable
and at 600 to °C that of the unheated
the compressive strengthspecimens.
was comparable Furthermore,
to that of theall the samples
unheated lost more
specimens.
than halfFurthermore, all the
of their initial samples after
strength, lost more than half
exposure toof800 ◦ C.
their initial strength, after exposure to 800 °C.

Compressive
Figure 6.Figure strength
6. Compressive of normal
strength concrete,
of normal concrete, SFRC andNS-modified
SFRC and NS-modified SFRC,
SFRC, as a function
as a function of of
temperature.
temperature. Redrawn
Redrawn fromfrom
[51].[51].

Bastami et al. [46] characterized the effects of replacing SF with NS, in two types of HSCs (w/b =
Bastami et al. [46]30characterized
0.25), containing kg/m3 and 60 kg/m the3 of
effects
SF andof replacing
exposed SF with NS,
to temperatures in 600,
of 400, twoandtypes
800 °C.of HSCs
(w/b = 0.25),
SF wascontaining
replaced with 30NS kg/m 3 and 60 of
in the amounts 1.5, 3,3 and
kg/m of SF and
4.5% by aexposed to temperatures
mass of cement (1.41, 2.83, andof 4.2
400, 600,
and 800 wt.◦ C.%SFofwasbinder mass, respectively).
replaced with NS in The study showed
the amounts that3,inand
of 1.5, both4.5%typesbyofaconcrete,
mass ofresidual
cement (1.41,
2.83, andcompressive
4.2 wt. %strength
of binder increased with NS content. For example, the strength loss of plain concrete (SF
mass, respectively). The study showed that in both types of concrete,
= 30 kg/m3), after exposure to 400, 600, and 800 °C was 15.3%, 48.4%, and 73.3%, while the
residualcorresponding
compressiveHSC strength increased with NS content. For example, the strength loss of plain
with 4.5 wt. % of NS lost its strength by 7.1%, 40.0%, and 67.9%, respectively.
concreteSherif
(SF =[49]30 analyzed 3
kg/m ),the after exposure to 400, 600, andof800 ◦ C was 15.3%, 48.4%, and 73.3%, while
thermal resistance of two types concretes, with different cement amounts
the corresponding
of 350 and 450 HSCkg/mwith
3 4.5 w/c
(fixed wt. =%0.25)of NSandlost its strength
an addition of SFby by 7.1%,
15 wt. 40.0%,
% and of and
NS 67.9%, respectively.
by 1–5 wt. %.
Their
Sherif [49] samplesthe
analyzed were
thermalexposed to temperatures
resistance of two types up ofto concretes,
600 °C. A with slightdifferent
compressive
cement strength
amounts of
improvement
350 and 450 kg/m3 (fixedwas observed
w/c = 0.25) after and
exposure to 200 °C;
an addition ofhowever,
SF by 15no wt.differences
% and ofbetween
NS by 1–5specimens
wt. %. Their
were reported. After exposure to 400 and 600 °C, ◦ the samples gradually lost their strength, however
samples were exposed to temperatures up to 600 C. A slight compressive strength improvement was
lower strength loss was reported for concretes containing NS. Also, with an increment of NS content,
observedtheafter exposure
strength loss was 200 ◦ C; After
to reduced. however,exposureno differences
to 400 and 600 between
°C, both specimens
types of plain were reported.
concrete (with After
exposurecement
to 400contents
and 600 ◦ C, the
of 350 andsamples
450 kg/mgradually
3) exhibited lost their
78.0%, 40.4%strength,
and 70.0%,however
36.2% lower
of theirstrength
initial loss
was reported for respectively,
strength, concretes containing NS. Also, with
while NS-incorporated concretean increment of NS content,
(5 wt. %) exhibited the strength
86.0%, 49.0%, 77.0%, and loss was
reduced.44.5%
After(respectively)
exposure toof400 theirand
initial
600 ◦ C, both types of plain concrete (with cement contents of 350
strength.
Contradictory
3 results have been presented by Shah et al. [47]. In their study, three HSC mixes
and 450 kg/m ) exhibited 78.0%, 40.4% and 70.0%, 36.2% of their initial strength, respectively, while
(w/b 0.25–0.3) containing 5 wt. % of NS, 10 wt. % of SF or a mixture of 5 wt. % of NS and 5wt. % of
NS-incorporated concrete (5 wt. %) exhibited 86.0%, 49.0%, 77.0%, and 44.5% (respectively) of their
SF, were prepared. The specimens were exposed to temperatures of 200, 500, and 800 °C. At the first
initial strength.
heating temperature (200 °C), a beneficial effect of NS was observed, with concrete containing 5 wt.
Contradictory results
% of NS exhibiting thehave
highestbeen presented
strength by Shah
improvement (by et al. However,
10%). [47]. In their study, three
after exposure to 500HSC
and mixes
800 °C, all
(w/b 0.25–0.3) the samples5gradually
containing wt. % oflost NS,their
10strength,
wt. % of with
SFtheorhighest strength
a mixture of 5loss
wt.being
% ofobserved
NS and for5wt. %
samples containing NS. After exposure to 800 °C, NS-modified
of SF, were prepared. The specimens were exposed to temperatures of 200, 500, and 800 C. At the samples lost 79% of their ◦
initial
strength, while specimens containing SF exhibited a strength loss of 49%. Incorporation of an NS/SF
first heating temperature (200 ◦ C), a beneficial effect of NS was observed, with concrete containing
mixture resulted in a 71% strength loss. The authors attributed this phenomenon to an excessive
5 wt. % of NS exhibiting
build-up the highest
of vapor pressure, which strength improvement
led to extensive cracking (by 10%). However,
in NS-incorporated after Moreover,
samples. exposure to 500
◦ C, all the samples gradually lost their strength, with the highest strength loss being observed
and 800 the authors reported a significant reduction of workability and problems with homogenous mixing,
for samples
whencontaining
NS was present NS.inAfter
the mix.exposure
This could 800 ◦been
to have C, NS-modified samples
the factor resulting in the lost 79%impact
negative of their of initial
NS present in the mix.
strength, while specimens containing SF exhibited a strength loss of 49%. Incorporation of an NS/SF
mixture resulted in a 71% strength loss. The authors attributed this phenomenon to an excessive
build-up of vapor pressure, which led to extensive cracking in NS-incorporated samples. Moreover,
the authors reported a significant reduction of workability and problems with homogenous mixing,
when NS was present in the mix. This could have been the factor resulting in the negative impact of
NS present in the mix.
Nanomaterials 2018, 8, 465 15 of 33

Based on the above-mentioned research, it is safe to conclude that the incorporation of NS


has a significant impact on the improvement of compressive strength, after exposure to elevated
temperatures, regardless of the type of composite (i.e., paste, mortar or concrete). The inclusion of
NS has a beneficial effect on compressive strength retention, even up to a temperature of 800 ◦ C.
However, in most of the works cited, NS-incorporated specimens exhibited a remarkable improvement
in compressive strength up to 400 ◦ C, after which the beneficial effect diminished gradually.

2.3. Microcracking and Spalling


As mentioned in the introduction, incorporation of selected SCMs, such as SF, represents
a significant threat related to the risk of increased cracking and spalling. Spalling is more likely
in HSCs (with lower water to cement ratio). Due to the filler effect, pozzolanic activity, and ultra-fine
particle size, the incorporation of SF results in concrete compaction and the refinement of the pore
structure, thus resulting in a cement matrix with low permeability. Heating of the concrete results in
a build-up of internal pressure, leading to extensive cracking. However, assuming NS action based on
previous findings regarding SF might be misleading, due to the different reactivity of NS, as well as
the lower amount of NS incorporated (usually lower than 5 wt. %).
Generally, in cement-based composites with standard w/c ratios, it is assumed that NS has
a beneficial effect in decreasing the number of surface cracks in cement-based composites. Due to an
increase in the amount of high-density C–S–H phase and the higher thermal resistance of the C–S–H
phase in a lower temperature range (up to 300–400 ◦ C), fewer cracks are observed in the cement matrix.
After exposure to a higher temperature range (above 400–450 ◦ C), the amount of surface cracks in
NS-modified cement-based composites has also been observed to be lower and that microcracks are
narrower and shorter. This effect is attributable to the fact that NS decreases the amount of CH in the
cement paste, because of its reaction with CH, resulting in the formation of an additional amount of
C–S–H. At a temperature of 400–450 ◦ C, decomposition of free Ca(OH)2 to CaO occurs. During cooling
and exposure to moisture, the dehydrated CH reforms rapidly, accompanied by a noticeable volume
increment, resulting in the further development of microcracks. However, due to a lower amount of
initial CH in NS-incorporated samples (before heating), fewer microcracks are observed after heating
in temperatures >400 ◦ C [28].
Bastami et al. [46] have analyzed the effect of NS, (as a SF replacement, in the amounts of 1.5, 3
and 4.5 wt. %) on spalling, in two types of HSC (w/b = 0.25), containing fixed amounts of silica fume;
i.e., 30 kg/m3 and 60 kg/m3 . The concretes were exposed to 400, 600, and 800 ◦ C, with the authors
reporting that the effect of NS on spalling was not easy to judge, but that a few general conclusions
could be drawn: up to 400 ◦ C, no spalling was observed and mass loss for all mixtures containing NS
was lower than that of the reference sample. After exposure to 600 ◦ C, spalling was observed for all the
samples tested. In the case of HSC with a fixed SF (or SF + NS) amount of up to 30 kg/m3 , the presence
of NS decreased spalling. Samples containing NS exhibited a lower mass loss, as compared to plain
samples. However, the presence of NS in HSC with a fixed SF (or SF + NS) amount of up to 60 kg/m3 ,
resulted in significantly increased spalling. For instance, the reference mixture exhibited a mass loss of
9.3%, while samples containing 1.5, 3 and 4.5 wt. % of NS exhibited mass losses of 22.9, 16.4, and 11.0%,
respectively. After exposure to 800 ◦ C, a similar trend in the spalling of the samples was observed.
The first type of concrete (with a lower SF/SF + NS content), exhibited lower spalling (lower mass loss),
while the concrete with 60 kg/m3 of SCMs exhibited higher spalling (higher mass loss), when NS was
present in the mix. Shah et al. [47] have studied the spalling of three HSCs (w/b = 0.25–0.3), containing
5 wt. % of NS, 10 wt. % of SF or a mixture of 5 wt. % of NS and 5 wt. % SF, exposed to 200, 500,
and 800 ◦ C. After exposure to 200 ◦ C, no visible cracks on the surface of the specimens were observed,
with the first spalling and cracking being observed only after exposure to 500 ◦ C. It was noted that
samples containing SF exhibited some fine surface cracks, while the samples containing NS exhibited
significant cracks. However, the highest cracking and spalling was observed in samples containing
a mixture of NS and SF. After exposure to 800 ◦ C, the authors reported that less cracks were observed
Nanomaterials 2018, 8, 465 16 of 33

in specimens containing SF, while extensive cracking (with cracks going deep into the cross-section)
was observed for samples incorporated with NS. Spalling occurred in samples containing a mixture of
NS and SF. The authors attributed this phenomenon to the extra dense microstructure produced by the
incorporation of a combination of SF and NS.
Horszczaruk et al. [28] analyzed the effects of the presence of NS on cement mortars containing
various aggregates (quartz, magnetite and barite). Generally, a positive effect, in terms of a reduction
in crack width and length was observed, but in the case of the barite aggregate, undesirable spalling
and cracking occurred. This was attributed to the high water absorption of the barite aggregate
itself, resulting in the accumulation of water in the aggregate and an undesired compaction of the
cement matrix. When NS was incorporated in the mix, a dense microstructure was produced. Due to
the high thermal conductivity of the barite aggregate, a fast diffusion process of absorbed water
associated with the aggregate, occurred. When the specimens were heated, the denser microstructure
of NS-modified mixes resulted in more extensive cracking, in comparison to the plain specimens.
Maheswaran et al. [43], have analyzed the effects of ternary blended OPC, containing lime sludge and
SF or NS, exposed to 500 and 800 ◦ C. No spalling in cement mortars was observed up to 500 ◦ C, while
beyond 800 ◦ C, spalling was observed in samples containing SF. On the other hand, the NS-modified
samples did not exhibit a spalling effect, as a result of the apparent pore-filling effect of NS and
enhanced particle packing in the mortar.
From the references cited above, it can be seen that most of studies have reported beneficial effects
of NS regarding crack reduction in cement-based composites, especially in the case of composites with
a higher w/c content. Compared to silica fume, the incorporation of NS reduces the risk of cracking
in cement-based materials with higher w/c ratios, after exposure to high temperatures. However,
the effect of NS on spalling in composites with denser matrices, such as HSCs, is still controversial and
requires further investigation.

3. The Effect of Carbon-based Nanomaterials on the Thermal Resistance of Cement-Based


Composites
Along with NS, particular research interest has also arisen in regard to the incorporation of
carbon-based nanomaterials in cementitious composites: mainly carbon nanotubes (CNTs), carbon
nanofibers (CNF), graphene oxide (GO), and graphene nanoplatelets (GNPs) [30]. Due to their
reinforcing ability (high tensile strength and high toughness), the incorporation of CNTs and CNFs,
even in very small amounts (usually less than 0.5 wt. %), significantly improves mechanical
performance, fracture characteristics, and the durability of cementitious composites. CNTs create
bridges between nano- and micro-cracks in the binder, thus increasing tensile strength and limiting
further crack propagation [23,53,54]. In addition, due to their remarkable mechanical and electrical
properties, CNTs are ideal for manufacturing self-sensing cement-based composites; whereby the
cement-based composites act as sensors able to detect their own state of strain or stress, as reflected in
changes to their electrical properties [55,56]. GO has also been introduced as an excellent reinforcement
for cementitious composites, due to its easier method of production (as compared to CNTs) and
higher solubility in aqueous cement matrices. In addition, graphene oxide has a higher surface
area than carbon nanotubes, and its surface is covered with higher amounts of functional groups.
Therefore, GO sheets exhibit higher reactivity (compared to CNTs), and thus increase the nucleation
area for the C–S–H gel [30,57,58]. The greater number of functional groups dispersed all across the
surface area, make this material ideal for incorporation in cementitious composites, as cement matrix
“nano-reinforcers”.
Nevertheless, interest related to the influence of carbon-based nanomaterials on the performance
of cement-based composites, under various thermal conditions, is relatively limited [59–62]. However,
the available data shows very promising results regarding the incorporation of carbon-based
nanomaterials, to produce heat-resistant cementitious composites. The studies available, related
to the effects of carbon nanomaterials, are summarized in Table 3.
Nanomaterials 2018, 8, 465 17 of 33

Table 3. Summary of available studies related to the effect of various nanomaterials on thermal resistance of cement-based composites.

Size and Specific Surface Area Tested Mechanical Optimal


Type of Nanomaterial Amount [wt. %] Heating Temperature [◦ C] Type of Material Ref.
(SSA) Properties * Amount [wt. %]
10–40 nm (diameter), 5–10 µm
MWCNT 0.02, 0.05, 0.1, 0.2 300 - 600–800 Cement paste Fc 0.1 [59]
(length), 93.81 m2 /g (SSA)
Carbon
85 nm CNS on 5–6 mm CF 0.55 200–400–600 Cement paste Fc - [61]
nanosphere-carbon fiber
>50 nm (diameter), length 20 µm,
Hydroxylated MWCNT 0.1, 0.2 200–400–600 Cement paste Ff , Fc 0.2 [63]
<400 (aspect ratio), >40 m2 /g (SSA)
Graphene oxide n/a n/a 400–600–800 Concrete Ft , Fc - [62]
Graphene sulfonate 1–2 nm (diameter),
0.1 200–400–600–800–1000 Concrete Fc , Fst - [60]
nanosheet 50–100 µm (length)
1 nm (diameter),300–500
Nanoclay 1, 2 200–400–600 Cement mortar Ff , Ft , Fc 2 [64]
(aspect ratio)
Nanoclay n/a 1, 2, 3 800 Cement mortar Fc 0.5-1 [65]
Nanoclay 10 nm (diameter), >600 m2 /g (SSA) 0.1, 0.3, 0.5 400–440–580–800 Concrete Fc , Fst 0.3 [66]
Nanoclay <10 nm (diameter) 0.1, 0.3, 0.5 300–440–500–580–800–1000 Concrete Fc 0.5 [67]
Nanometakaolin 100·50·10 nm, 48 m2 /g (SSA) 5, 10, 15 250–450–600–800 Cement paste Fc 15 [68]
Nanoalumina 40 ± 5 nm, 173 m2 /g (SSA) 0.5, 1, 2, 3 300–600–800 Cement paste Fc 0.5 [41]
Nanoalumina 15 ± 3 nm, 165 ± 12 m2 /g (SSA) 1, 2 250–450–600–800–1000 Cement mortar Fc 1 [69]
Nanoalumina 13 nm, 85–115 m2 /g (SSA) 1, 2, 3 100–200–300–400–600–800–1000 Cement paste Fc 1 [70]
Nano-iron oxide 10–20 nm, 50.5 m2 /g (SSA) 1, 2 200–300–400–600–800–1000 Cement paste Fc 1 [71]
Nano-iron oxide 14.6 nm 1, 2, 3 105–250–450–600–800 Cement paste Fc 1 [72]
Nanotitania 21 nm, 35–65 m2 /g (SSA) 1, 2, 3 100–200–300–400–600–800–1000 Cement mortar Fc 2 [73]
* Fc —compressive strength, Ff —flexural strength, Ft —tensile strength, Fst —split-tensile strength.
Nanomaterials 2018, 8, x FOR PEER REVIEW 17 of 32
Nanomaterials 2018, 8, 465 18 of 33

3.1. The Effect of Multi-Walled Carbon Nanotubes (MWCNTs)


3.1. The Effect of Multi-Walled Carbon Nanotubes (MWCNTs)
Amin et al. [59] have analyzed the effects of MWCNTs, in the amounts of 0.02, 0.05, 0.1, and 0.2
wt. Amin et al. [59] have analyzed the effects of MWCNTs, in the amounts of 0.02, 0.05, 0.1,
%, on the thermal resistance of OPC pastes and cement pastes containing up to 30 wt. % clay brick
and 0.2 wt. %,
waste (Homra) on the thermal
cement resistance The
replacement. of OPC pastes
cement and cement
pastes pastes
(w/b = 0.3) containing
were exposedup to to 30 wt. % clayof
temperatures
brick
up towaste (Homra)
800 °C (Figure cement
2). Thereplacement.
results showedThe cement
that an pastes
optimal (w/b = 0.3)ofwere
amount exposed(established
MWCNTs to temperatures
as 0.1
ofwt.
up%)to 800 ◦ C (Figure 2). The results showed that an optimal amount of MWCNTs (established as
in combination with clay brick waste resulted in a marked increase of residual compressive
0.1 wt. %) inincombination
strength, cement pastes with clay brick
exposed to waste resulted
elevated in a marked
temperatures, evenincrease
up to of residual
800 °C for compressive
certain paste
strength, in cement pastes exposed toinelevated temperatures, ◦
compositions (Figure 7b). However, the case of OPC pastes, even up to 800effect
the beneficial C for certain paste
of MWCNTs was
compositions
only visible (Figure
up to ◦7b).
300 However,
°C (Figure in the
7a).case
X-rayof OPC pastes,
powder the beneficial
diffraction effect
(XRD), of MWCNTs
differential was
scanning
only visible up(DSC),
calorimetry to 300and
C (Figure
scanning7a).electron
X-ray powder diffraction
microscope (SEM)(XRD),
analysisdifferential scanning
showed that MWCNTscalorimetry
do not
(DSC), andrate
affect the scanning electron
of hydration microscope
reactions. (SEM) analysis
The positive effect ofshowed
MWCNTs thatcan
MWCNTs do be
most likely notattributed
affect theto
rate ofpore-filling
their hydration reactions. Theaspositive
effect, as well to theireffect of MWCNTs
bridging can most
ability between likelyand
hydrates be attributed to their
across cracks.
pore-filling effect, as well as to their bridging ability between hydrates and across cracks.

Figure7.7.The
Figure Theeffect
effectofofcarbon
carbonnanotubes
nanotubes(CNT)
(CNT)content
contenton
onthe
therelative
relativeresidual
residualcompressive
compressivestrength
strength
of cement pastes containing 100 wt. % of ordinary Portland cement (OPC) (a) and 70 wt. % of OPC and
of cement pastes containing 100 wt. % of ordinary Portland cement (OPC) (a) and 70 wt. % of OPC
30 wt. % of Homra (b) after firing. Reproduced with permission from [59]. Copyright Elsevier, 2015.
and 30 wt. % of Homra (b) after firing. Reproduced with permission from [59]. Copyright Elsevier,
2015.
Similar conclusions have been drawn by Zhang et al. [63]. In their work, the effect of 0.1 and
0.2 wt.Similar conclusions have
% of hydroxylated been drawn
MWCNTs on theby Zhang et al. [63].
microstructure andInmechanical
their work,performance
the effect of 0.1 and 0.2
(flexural
wt. % of
strength andhydroxylated
compressiveMWCNTs strength) of oncement
the microstructure
pastes (w/c =and 0.4),mechanical
exposed toperformance
temperatures(flexural
up to
600 ◦ C, was
strength and compressiveXRD
determined. strength)
resultsof showed
cement pastesthat the (w/c = 0.4), exposed
inclusion of MWCNTs to temperatures
does not affectup tothe
600
°C, was determined.
hydration XRD results
kinetics of cement at room showed that thewhile
temperature, inclusion of MWCNTs
at elevated does notMWCNTs
temperatures, affect the hindered
hydration
kinetics
further of cement After
hydration. at room temperature,
exposure to 600while◦ C, noat influence
elevated temperatures,
of the MWCNTs MWCNTson the hindered
decompositionfurther
ofhydration.
hydrationAfter exposure
products was to 600 °C, noSEM
observed. influence of the
analysis MWCNTs
revealed thaton the
up todecomposition 400 ◦ C,
a temperatureofofhydration
products
CNT was aobserved.
exhibited bridgingSEM effectanalysis revealed
for the pores andthat up to
cracks in athetemperature
cement matrix,of 400while
°C, CNT
afterexhibited
exposure a
600 ◦ C, the
tobridging effect for the pores
MWCNTs and cracks
were mostly spalledin the cement
with matrix,
the matrix onwhile afterofexposure
the walls the porestoand 600cracks.
°C, the
MWCNTs
The positive were
effectsmostly spalled with
of the MWCNTs were the matrix both
reflected on the walls
in the of theand
flexural pores
the and cracks. The
compressive positive
strength of
effects pastes
cement of the MWCNTs
(Figure 8).were Up to 200 ◦ C,both
reflected in theenhancement
strength flexural and the wascompressive
attributable strength of cement
to the continued
pastes (Figure
existence of the8). Up to 200
bridging of °C,
CNT strength
in the enhancement
gaps and pores. was attributable
However, the to the continued
beneficial existence
effects of theof
the bridging
MWCNTs of CNT
on thermal in the gaps
resistance wereand mostpores. However,
pronounced afterthe beneficial
exposure to 400 ◦
effects
C. It of the MWCNTs
is worth noting that on
thermal resistance were
MWCNT-incorporated most pronounced
samples, exhibited similar after exposure
compressive to 400 °C. It isafter
strengths worth noting to
exposure that
200 ◦ C and
MWCNT-
400 ◦ C, while plain
incorporated samples,
cement exhibited similar compressive
paste exhibited gradual strength strengths afterexposure
loss after exposuretotothese 200 °C and 400 °C,
temperatures
while plain
(Figure 8b). cement
Therefore,pasteit exhibited
was carbon gradual
nanotubesstrengththatloss after exposure
contributed to prevention
to the these temperatures
of the loss(Figure
of
8b). Therefore,
compressive it wasafter
strength, carbon nanotubes
exposure to 400that◦ C.contributed
The authorsto the prevention
attributed of the losstoofthe
this phenomena compressive
potential
strength,
ability after exposure
of CNTs to work astochannels
400 °C. Thefor authors
the release attributed this phenomena
of autoclaving steam, thus to the potential damage
preventing ability of
CNTs
from to work as channels
high-pressure steam. After for the releasetoof600
exposure ◦
autoclaving steam, thus
C, the MWCNTs lostpreventing
their bridging damage from
ability, andhigh-
the
pressure steam. After exposure to 600 °C, the MWCNTs lost their bridging ability, and the samples
Nanomaterials 2018, 8, 465 19 of 33
Nanomaterials 2018, 8, x FOR PEER REVIEW 18 of 32

exhibited a similar compressive


samples exhibited strength. The
a similar compressive optimal
strength. value
The of MWCNTs
optimal in the pasteinwas
value of MWCNTs the estimated
paste was
Nanomaterials 2018, 8, x FOR PEER REVIEW 18 of 32
at 0.2 wt. %.
estimated at 0.2 wt. %.
exhibited a similar compressive strength. The optimal value of MWCNTs in the paste was estimated
at 0.2 wt. %.

Figure 8.
Figure Theeffect
8. The effectof
ofhydroxylated
hydroxylatedMulti-Walled
Multi-WalledCarbon
CarbonNanotubes
Nanotubes(MWCNTs)
(MWCNTs)on onthe
theflexural
flexural(a)
(a)
and Figure 8. The effect
compressive (b) of hydroxylated
strengths of Multi-Walled
cement pastes Carbonto
exposed Nanotubes
elevated (MWCNTs) on the
temperatures. flexural (a) with
Reproduced
and compressive (b) strengths of cement pastes exposed to elevated temperatures. Reproduced with
and compressive
permission [63].(b)
from [63]. strengths of
Copyright cement 2017.
Elsevier, pastes exposed to elevated temperatures. Reproduced with
permission from Copyright Elsevier, 2017.
permission from [63]. Copyright Elsevier, 2017.

Moreover, extensive SEM analysis ◦ C,


Moreover, extensive
Moreover, SEM
extensive SEM analysis(Figure
analysis (Figure9)
(Figure showed
9)9) showed
showed that
that uptoup
that
up toa temperature
atotemperature
a temperature of 400
of 400 °C,
400 the
of the
°C,
bridging phenomenon
the bridging
bridging phenomenon
phenomenonof CNT
of of
CNTin in
CNT theincement
the matrix
the cement
cement can
matrixmatrixbecan
can be observed, while
be observed,
observed, while at aat a temperature
while of 600
at a temperature
temperature of 600
°C °C◦ C
carbon
of 600 nanotubes
carbon areare
nanotubes
carbon notnot
nanotubes discernible.
arediscernible.
not discernible.

Figure 9. Scanning electron microscope (SEM) micrographs showing the bridging effect of MWCNTs
Figure 9. Scanning electron microscope (SEM) micrographs showing the bridging effect of MWCNTs
in a cement matrix before (20 °C) and after exposure to elevated temperatures of 200, 400, and 600 °C.
in a cement matrix
Reproduced before
with (20 ◦ C)
permission and[63].
from after exposure
Copyright to elevated
Elsevier, 2017. temperatures of 200, 400, and 600 C.

Reproduced
Figure with permission
9. Scanning from [63].(SEM)
electron microscope Copyright Elsevier,showing
micrographs 2017. the bridging effect of MWCNTs
in a cement matrix before (20 °C) and after exposure to elevated temperatures of 200, 400, and 600 °C.
Reproduced with permission from [63]. Copyright Elsevier, 2017.
Nanomaterials 2018, 8, x FOR PEER REVIEW 19 of 32
Nanomaterials 2018, 8, 465 20 of 33

3.2. The Effect of Carbon Nanospheres


3.2. The Effect of Carbon Nanospheres
Another approach for incorporating carbon-based materials to improve the thermal resistance
of cement-based
Another approach composites, has been proposed
for incorporating by Han
carbon-based et al. [61].
materials to In this study,
improve the authors
the thermal proposed
resistance of
a novel multiscale
cement-based reinforcement
composites, has been structure,
proposed built frometthe
by Han al.fast
[61].growth
In thisofstudy,
carbon thenanospheres
authors proposed (CNSs)
aon the multiscale
novel surface of carbon fibers (CFs).
reinforcement The fiber
structure, built surface
from thewas fastcovered
growthwith a uniform
of carbon CNS layer,
nanospheres with
(CNSs)
onanthe
average
surface thickness
of carbon of fibers
85 nm.(CFs).Afterwards,
The fibercement
surfacepastes
was (w/c
covered= 0.33)
withcontaining
a uniform0.55 CNS vol. % ofwith
layer, fiber
anwere prepared
average and subjected
thickness of 85 nm. toAfterwards,
temperatures of 200,
cement 400 and
pastes (w/c600 °C. Itcontaining
= 0.33) was observed 0.55 that
vol. as% ofthe
temperature
fiber increased,
were prepared andthe size of the
subjected cracks in the cement
to temperatures of 200,pastes
400 and increased.◦
600 C. CrackIt wassizes were smallest
observed that as
in temperature
the the samples containing
increased, the CNS-modified fibers. in
size of the cracks Thethestudy
cement showed
pastesthat, cementCrack
increased. pastessizescontaining
were
CNS-modified
smallest CFs, exhibited
in the samples containing betterCNS-modified
resistance to the thermal
fibers. The degradation
study showed (upthat,
to 600 °C) resulting
cement pastes
from exposure
containing to elevated
CNS-modified CFs,temperatures,
exhibited better with superiorto relative
resistance the thermal residual strength.
degradation (upThe ◦ C)
authors
to 600
attributed
resulting fromtheexposure
positiveto effect of CNS
elevated on the surface
temperatures, of CF, torelative
with superior two main phenomena:
residual strength.Firstly,
The authorsto the
improved interlocking between CNS-modified CFs and the cement
attributed the positive effect of CNS on the surface of CF, to two main phenomena: Firstly, to the matrix, as compared to pristine
CFs; and interlocking
improved secondly, to between the improved thermal resistance
CNS-modified CFs and theof CFs,
cementresulting
matrix, from a thin CNS
as compared tolayer that
pristine
oxidizes
CFs; first, thusto
and secondly, forming microchannels
the improved thermalwhich enable
resistance vaporresulting
of CFs, tension fromin theacapillaries
thin CNS layer to be more
that
easily alleviated
oxidizes first, thusand released.
forming microchannels which enable vapor tension in the capillaries to be more
easily alleviated and released.
3.3. The Effect of Graphene Oxide
3.3. The Effect of Graphene Oxide
Mohammed et al. [62] incorporated graphene oxide in normal strength concrete (w/c = 0.45) and
HSCMohammed
(w/c = 0.30),etexposing
al. [62] incorporated graphene up
them to temperatures oxide in normal
to 800 °C. In the strength
case ofconcrete (w/c = 0.45)
normal strength and
concrete,
HSC (w/c = 0.30), strength
the compressive exposingloss them to temperatures
(Figure up to 800 ◦ C.
10a) of GO-modified In the case
concretes wasofnoticeably
normal strength
lowerconcrete,
than that
the
of compressive
plain concrete, strength loss (Figure
especially 10a) of GO-modified
after exposure to 400 and 600 concretes was noticeably
°C (Figure 10a). Plain, lower
normalthanstrength
that of
plain concrete,
concrete especially
exhibited after exposure
a gradual strength loss to 400
after 600 ◦ C (Figure
andexposure to 400,10a).
600, Plain,
and 800 normal strength
°C, while concrete
GO-modified
exhibited
concrete aexhibited
gradual strength
only a slightloss after exposure
strength to 400,In600,
reduction. theandcase800 of ◦HSC,
C, while GO-modified
incorporation of GO concrete
in the
exhibited
mix wasonlyevenamoreslightefficient
strengththan reduction.
in the In theofcase
case of HSC,
normal incorporation
strength concrete.ofItGO in the mix
is known thatwas HSCs evenare
more
moreefficient
prone to than in the case
spalling and of normal strength
degradation underconcrete.
elevated Ittemperature
is known that HSCs arebut
conditions, more theprone to
authors
spalling
reported andthatdegradation
the spallingunder elevated
resistance temperature conditions,
of GO-incorporated HSC increased,but thesuch authors
thatreported
samples that werethe not
spalling
damaged, resistance
even upoftoGO-incorporated
800 °C and that strength HSC increased,
loss wassuch that However,
gradual. samples were in thenotcase
damaged,
of plaineven HSC,
up to 800 ◦ C were
specimens and that strengthabove
destroyed loss was 400 gradual.
°C. SimilarHowever,
trends in the case
were of plain
observed in HSC,
the casespecimens were
of the tensile
destroyed above ◦
400 C.although
Similar trends
strength of concretes, in thiswere
case,observed
the degree in of
thestrength
case of the tensile strength
degradation of concretes,
was higher (Figure
although
10b). in this case, the degree of strength degradation was higher (Figure 10b).

Figure 10. Compressive strength (a) and tensile strength (b) test results of graphene oxide
Figure 10. Compressive strength (a) and tensile strength (b) test results of graphene oxide (GO)-
(GO)-modified high strength concrete (GO-HSC), high strength concrete (HSC), GO-modified normal
modified
strength high (GO-NSC)
concrete strength concrete (GO-HSC),
and normal high strength
strength concrete (NSC) concrete (HSC),
after exposure GO-modified
to elevated normal
temperature.
strength concrete
Redrawn from [62]. (GO-NSC) and normal strength concrete (NSC) after exposure to elevated
temperature. Redrawn from [62].
Nanomaterials 2018, 8, 465 21 of 33

The authors attributed the beneficial effect of GO on mechanical properties to its ability of
minimizing temperature increase in the specimen and thus limiting the effects of temperature exposure,
as well as reducing cracks in the concrete. The authors reported that the incorporation of GO
contributed to alternation in the porosity of the cement matrix. The number of large pores (from 0.4 µm
to 10 µm) shifted to a number of small pores (<0.3 µm), which contributed to an increment in the
amount of effective gel pores. As such, porous structures with nano- and micro-scale channels were
produced, enabling the release of vapor pressure, helping to prevent extensive spalling in samples,
which was reflected in the mechanical properties of the concretes as well as in the decreased amount of
cracks [62].

3.4. The Effect of Graphene Sulphonate Nanosheets


Chu et al. [60] analyzed the effect of graphene sulfonate nanosheets (GSNSs), in the amount
of 0.1 wt. %, on the properties of sacrificial concretes (w/b = 0.33) exposed to temperatures of up
to 1000 ◦ C. Extensive SEM analysis and MIP tests showed, that the inclusion of GSNSs exhibited
reinforcing and toughening effects on the microstructures of the concretes tested, which was reflected
in a reduction of concrete porosity. The total porosity of concrete modified with GSNSs was reduced by
3.0–7.0% (depending on the heating temperature), as compared to plain concrete samples. Moreover,
the thermal gradient in the concrete modified with GSNSs decreased and as such, the thermal damage
rate dropped. The positive effect of GSNSs on microstructure translated in to an improvement in
split-tensile and compressive strength. The authors reported, that relative residual compressive
strength, as well as the splitting tensile strength of GSNS modified concretes, was higher than that of
plain concrete, at each tested temperature.

4. The Effect of Nanoclays and Calcinated Nanoclay on the Thermal Resistance of Cement-Based
Composites
Nanoclays (NCs) originate from naturally occurring clays. These clay particles are hydrous
silicates and can generally be described as fine-grained particles organized in sheet-like structures
stacked on top of each other [74]. Based on their chemical composition and morphology, NCs
have been classified into various groups, including montmorillonite, bentonite, kaolinite, hectorite,
and halloysite [31]. The incorporation of NCs has been gathering increasing attention in many fields
of engineering, which is attributable to their availability, environmental friendliness, and low cost,
as compared to other manufactured nanomaterials. The beneficial effects of NC admixtures on
the properties of cement-based composites has been widely reported. Due to their large surface
area to volume ratio, NC particles can facilitate chemical reactions. In the presence of silica and
aluminum, NC nanoparticles can exhibit a nucleation effect and high pozzolanic activity (as in case of
nanosilica), as well as a nano-filling effect, which leads to an increase in the overall performance of
concrete [24]. Previous studies have shown that the inclusion of NC contributes to the formation of
ill-crystalline CH and the formation of a C–S–H phase [75]. Chang et al. [76] have reported that the
incorporation of nano-montmorillonite results in higher CH consumption and higher production of
C–S–H. In addition, the study showed that NCs are non-combustible materials and as such, various
efforts to incorporate them as flame and fire protective admixtures in a vast range of materials,
including polymers, have been undertaken. [74]. Based on these observations, attempts to incorporate
NCs as cement admixtures for improving the thermal resistance of cement-based composites, has also
gathered researchers’ attention. Table 3 summarizes the studies related to the thermal resistance of
NC-modified cement-based composites.
Wang [67] have analyzed the effects of cement replacement, by 0.1, 0.3, and 0.5 wt. %, on the
thermal resistance of two types of concrete, with w/c = 0.4 and w/c = 0.5. In their study, compressive
strength and the thermal conductivity coefficient, under different temperature conditions (up to
1000 ◦ C), were analyzed. It was shown that in both types of concrete, when 0.1 wt. % of NC was
incorporated in the concrete, compressive strength loss was higher than that of plain concrete, during
Nanomaterials2018,
Nanomaterials 2018,8,8,465
x FOR PEER REVIEW 2221ofof3332

observed. In addition, the authors reported that the thermal conductivity coefficient of cement paste
heating0.1(Figure
withNanomaterials
wt. % 11).
NC However, when the amount
dropped, while of NC
replacement of was increased
cement to 0.3
with 0.3 wt. wt.
% and% and
0.5210.5
wt. wt. %,
% NC,
2018, 8, x FOR PEER REVIEW of 32
the highest compressive strength increase,
increased the thermal conductivity coefficient of paste.during heating and relative to ordinary concrete, was
observed. In addition,
observed. In addition, thethe
authors
authorsreported
reported that thethermal
that the thermalconductivity
conductivity coefficient
coefficient of cement
of cement paste paste
withwith
0.1 wt.
0.1 %wt.NC% NC dropped,
dropped,while
whilethe
thereplacement
replacement of of cement
cementwith
with0.30.3
wt.wt. % and
% and 0.5 %
0.5 wt. wt.NC, % NC,
increased
increased the thermal
the thermal conductivitycoefficient
conductivity coefficient of
of paste.
paste.

Figure 11. Compressive strength of nanoclay-modified concrete with w/c = 0.4 (a) and w/c = 0.5 (b)
after thermal exposure. Reproduced with permission from [67]. Copyright Elsevier, 2017.
Figure
Figure 11. Compressive
11. Compressive strengthofofnanoclay-modified
strength nanoclay-modified concrete
concretewith w/c
with w/c= 0.4 (a) and
= 0.4 w/c w/c
(a) and = 0.5 (b)
= 0.5 (b)
after thermal exposure. Reproduced with permission from [67]. Copyright Elsevier,
after thermal exposure. Reproduced with permission from [67]. Copyright Elsevier, 2017. 2017.
In a study undertaken by Irshidat et al. [64], the effect of hydrophilic montmorillonite, on the
thermalIn resistance of cementby
a study undertaken mortars
Irshidat(w/bet al.=[64],
0.55) thecontaining a NC cement
effect of hydrophilic replacementonrate
montmorillonite, the in the
In a study
thermal undertaken by Irshidat et al.= [64], the effect of hydrophilic montmorillonite, on the
amount of 1resistance
and 2 wt.of%, cement
exposedmortars (w/b
to 200, 400,0.55)
and 600 containing
°C, was a NC cement The
examined. replacement
authorsrate in the
determined the
thermal
amountresistance
of 1onand of cement mortars
2 wt. %,strength, (w/b
exposed totensile
200, 400, = 0.55)
and 600 °C,containing
was a NC
examined. Thecement replacement
authorsanddetermined rate
thein their
the
effect of NC flexural strength, and

compressive strength supported
amount of of
effect 1 and
NC 2onwt. %, exposed
flexural to tensile
200, 400, and 600 andC, was examined. Theand
authors determined the
study with XRD, DSC andstrength,
TGA analysis. strength,
The effect ofcompressive
NC on thestrengthcompressive supported
strengththeir of cement
effectstudy
of NC on XRD,
with flexuralDSC strength,
and TGA tensile strength,
analysis. The effectand of compressive strength andstrength
NC on the compressive supported their study
of cement
mortars was relatively limited and only a slight compressive strength improvement during heating
with mortars
XRD, DSC wasandrelatively limited and
TGA analysis. The only a slight
effect of NC compressive strength improvement
on the compressive during heating
strength of cement mortars was
was observed, when 2 wt. % NC was incorporated in the mix. The effect of NC was much more
relatively limited and only a slight compressive strength improvement during heating wasmore
was observed, when 2 wt. % NC was incorporated in the mix. The effect of NC was much observed,
pronounced regarding flexural and tensile strength
strength(Figure (Figure 12). After exposure to elevated
whenpronounced
2 wt. % NCregarding flexural in
was incorporated and
thetensile
mix. The effect of NC was 12).much
After exposure
more to elevated
pronounced regarding
temperatures,
temperatures,NC-incorporated
NC-incorporated samples exhibitedhigher
samples exhibited higher flexural
flexural strength
strength underunder all heating
all heating
flexural and tensile strength (Figure 12). After exposure to elevated temperatures, NC-incorporated
temperatures
temperatures (Figure
(Figure12a).
12a).After
After exposure
exposure to to200
200°C, °C, samples
samples containing
containing NC exhibited
NC exhibited only aonly a slightly
slightly
samples exhibited higher flexural strength under all heating temperatures (Figure 12a). After exposure
higher
higherflexural
flexural strength
strength than
than plain
plain mortar,
mortar, however,
however, when
when the the temperature
temperature
◦ C, samples containing NC exhibited only a slightly higher flexural strength than plain mortar,
increased
increased to 400 to°C,400
a °C, a
to 200significant differencebetween
between the the flexural
significant difference flexural strength
strength of of
unmodified
unmodified and and
modified mortars
modified
◦ C, a significant difference between the flexural
was was
mortars
however,
observed.when the temperature
Samples containing 11 increased
and 22 wt. to 400exhibited
observed. Samples containing and wt.%%NC NC exhibited 88% 88%andand 138% higher
138% flexural
higher strength
flexural strength
strength
than of unmodified
plain cement and Further
mortar. modified mortars
exposure to awas observed.
temperature of Samples
600 °C containing
caused a dramatic 1 and
strength2 wt. %
than plain cement mortar. Further exposure to a temperature of 600 °C caused a dramatic strength
NC exhibited
loss in cement88%mortars,
and 138% withhigher flexural
the specimen strength2than
containing wt. %plain
of NCcement
exhibitingmortar. Further
the highest exposure
flexural
loss in cement mortars, ◦with the specimen containing 2 wt. % of NC exhibiting the highest flexural
strength value.of
to a temperature The600beneficial effectaofdramatic
C caused NC was also reflected
strength in the
loss tensile strength
in cement mortars, of mortars
with the (Figure
specimen
strength value. The beneficial effect of NC was also reflected in the tensile strength of mortars (Figure
12b).
containing 2 wt. % of NC exhibiting the highest flexural strength value. The beneficial effect of NC
12b).
was also reflected in the tensile strength of mortars (Figure 12b).

Figure 12. Flexural strength (a) and tensile strength (b) of cement mortars, containing NC, exposed to
elevated temperature. Reproduced with permission from [64]. Copyright Elsevier, 2018.
Figure 12. Flexural strength (a) and tensile strength (b) of cement mortars, containing NC, exposed to
Figure 12. Flexural strength (a) and tensile strength (b) of cement mortars, containing NC, exposed to
elevated temperature. Reproduced with permission from [64]. Copyright Elsevier, 2018.
elevated temperature. Reproduced with permission from [64]. Copyright Elsevier, 2018.
Nanomaterials 2018, 8, 465 23 of 33

After exposure to 200 ◦ C, NC-modified samples exhibited a higher increment in tensile strength,
as compared to the plain mortar. After exposure to 400 ◦ C, the strength of the specimens declined
and the difference between samples was reduced; however, the highest strength was represented by
samples containing 2 wt. % of NC. Exposure to 600 ◦ C caused a total loss of strength in plain cement
mortar, while NC-incorporated samples still exhibited some strength. XRD analysis confirmed that
in the presence of 2 wt. % nanoclay, the consumption of CH reached its optimum value at 200 ◦ C,
whereas the additional formation of a C–S–H phase in the presence of NC reached its optimum level at
400 ◦ C. In addition, SEM analysis showed that the presence of NC decreased the density and width of
microcracks in the cement mortar, after heating, as compared to the plain mortar.
The beneficial effects of NC on the compressive, flexural and split-tensile strengths of concretes,
has also been reported by Ho et al. [66]. In their study, NC was added to concrete in 0.1, 0.3,
and 0.5 wt. %, with concrete specimens being exposed to temperatures of 400, 600, and 800 ◦ C.
It was found that the exposure of concretes with NC, up to a temperature of 400 ◦ C, barely affected
their compressive strength (loss by 1–2 MPa), while the control concrete lost 8.5 MPa of its initial
strength. Further exposure to elevated temperatures showed that NC-incorporated concrete exhibits
higher compressive strength than plain concrete and that the best performance is exhibited by concrete
containing 0.3 wt. % of NC. The tendency of the residual split-tensile strength of the concrete was
similar to that of compressive strength. Incorporation of NC successfully maintained the loss of
split-tensile strength after thermal exposure. Recently, Lee et al. [65] introduced the incorporation of NC
as an admixture, to produce high-strength nano-polymer modified fireproof cementitious composites
The authors produced fire resistant polymer-cement mortars (based on cement and ethylene-vinyl
acetate polymer powder), varying the amount of chamotte, silica fume, and nanoclay. NC was
incorporated into the initial mixes in the amounts of 1, 2, and 3 wt. %. The study showed that
amounts of NC > 1 wt. % negatively affected compressive strength of specimens in the unheated state.
To determine the thermal resistance of the cement mortars, their samples were exposed to 800 ◦ C.
Afterwards, statistical analysis was conducted and the tests were repeated in order to optimize the
NC content in the polymer-cement mortar. The study showed that for cost and performance reasons,
the amount of NC required to improve the thermal resistance of cement mortar is in the range of
0.5–1 wt. %.
Morsy et al. [68] incorporated calcinated nanoclay, namely nano metakaolin (NMK), obtained
from nanokaolin calcinated at a temperature of 750 ◦ C. They analyzed the effect of NMK cement
replacement in the amounts of 5, 10, and 15 wt. % of cement mortar. The pastes were exposed to
temperatures of up to 800 ◦ C. The study showed that after exposure to 250 ◦ C, the compressive strength
of all the tested samples increased, but that the highest strength increment was observed for samples
containing 10 and 15 wt. % of NMK (Figure 13a). According to the authors, this effect can be attributed
to the internal autoclaving effect and the pozzolanic reaction of amorphous aluminosilicate (present
in NMK) with CH, during cement hydration. It is resulting in the production of additional C–S–H,
with a low C/S ratio, with a calcium aluminate hydrate (C-A-H) phase being deposited in the pore
system. After further thermal exposure, the samples gradually lost their strength. However, the ones
containing a higher NMK content (10 wt. % and 15 wt. %) exhibited higher strength values than plain
and 5 wt. % samples, up to a temperature of 800 ◦ C. A beneficial effect of the NMK was also reflected
in the flexural strength of cement mortar (Figure 13b). After exposure to 250 ◦ C, samples containing
5, 10, and 15 wt. % of NMK exhibited strength increments by 3.4, 13.9, and 27% respectively, while
plain control samples exhibited a slight strength loss. This was attributed to an increase in the amount
of hydration products in the cement paste (due to an autoclaving effect), as well as because of the
filler effect of NMK, which acted as a fiber in the cement matrix. Exposure to higher temperatures
caused a gradual loss of flexural strength in the samples, although samples with a higher NC content
(10 and 15 wt. %) underwent only a minimal decrease, up to a temperature of 450 ◦ C. Exposure to
temperatures of 600 and 800 ◦ C caused a sharp strength loss in specimens, with samples containing
NMK in the amounts of 10 and 15 wt. % exhibiting the highest resistance to elevated temperature.
Nanomaterials 2018, 8, 465 24 of 33

Nanomaterials 2018, 8, x FOR PEER REVIEW 23 of 32


XRD and differential thermal analysis (DTA) confirmed the higher thermal stability of cement paste
matrix was
containing 15less
wt. damaged
% of NMK,(with
whilenarrower micro-cracks)
SEM analysis afterthe
proved that exposure
cement to 800 °C,
matrix wasthan
less the control
damaged
mortar.
(with ◦
narrower micro-cracks) after exposure to 800 C, than the control mortar.

Figure 13.Compressive
Figure13. Compressivestrength
strength(a)
(a)and
andflexural
flexuralstrength
strength(b)
(b)ofofcement
cementmortars
mortarscontaining
containingvarious
various
amount
amount of nano metakaolin (NMK), after thermal exposure. Reproduced with permission from[68].
of nano metakaolin (NMK), after thermal exposure. Reproduced with permission from [68].
Copyright
CopyrightElsevier,
Elsevier,2012.
2012.

5.5.The
TheEffect
EffectofofNanoalumina
Nanoaluminaon
onthe
theThermal
ThermalResistance
ResistanceofofCement-Based
Cement-BasedComposites
Composites
Nanoalumina(NA),
Nanoalumina (NA),along
alongwithwithnanosilica,
nanosilica,has hasgathered
gatherednoticeable
noticeableattention
attentionininthe thefield
fieldofof
modification of cement-based composites. Due to its pozzolanic
modification of cement-based composites. Due to its pozzolanic activity and high surface area, activity and high surface area,
aluminaparticles
alumina particlescan canreact
reactwith
withCH,CH,forming
formingadditional
additionalhydration
hydrationproducts
productssuchsuchasasC–A–S–H
C–A–S–Hand and
C–A–Hgel.
C–A–H gel.This
Thisphenomenon
phenomenonisisreflected
reflectedininaanoticeable
noticeableimprovement
improvementininmicrostructure
microstructureand andthethe
mechanical characteristics of cementitious composites [41]. Accordingly, NA has been introduced asas
mechanical characteristics of cementitious composites [41]. Accordingly, NA has been introduced
anadmixture
an admixturefor forimproving
improvingthe thethermal
thermalresistance
resistanceofofcementitious
cementitiouscomposites;
composites;the thestudies
studiesrelated
relatedtoto
thistopic
this topichave
havebeen
beensummarized
summarizedininTable Table3.3.
Heikaletetal.
Heikal al.[69]
[69]have
haveanalyzed
analyzedthe thefire
fireresistance
resistanceofofplain
plainand andsuperplasticized
superplasticizedcementcementpastespastes
containingNA
containing NAcement
cementreplacements
replacementsininthe theamounts
amountsofof1 1wt. wt.%%and and2 2wt.wt.%,%,and
anddetermined
determinedthe the
compressive strength, bulk density, total porosity and mass loss of samples
compressive strength, bulk density, total porosity and mass loss of samples exposed to temperatures exposed to temperatures
up 1000◦ C.
uptoto1000 °C.The
Thestudy studyshowed
showedthat thatcement
cementpaste pasteprepared
preparedwith withsuperplasticizer
superplasticizerexhibited
exhibitedbetter
better
dispersionofofNA
dispersion NAand andcement
cementgrains,
grains,thus
thusresulting
resultingininaamore moreefficient
efficienthydration
hydrationprocess
processand andaamore
more
homogenous cement paste microstructure. The result was cement
homogenous cement paste microstructure. The result was cement pastes with high bulk density pastes with high bulk density
valuesasaswell
values wellasas decreased
decreased porosities
porosities in all heating
heating temperatures.
temperatures.Another Anotherpositive
positiveeffect
effectofofNA NA on
the
on thehydration
hydrationprocess
processwas wasan anincrease
increase in in the compressive
compressive strength
strength of ofcement
cementpastes.
pastes.Specimens
Specimens
containing1 1wt.
containing wt.%%NA NAshowed
showedthe thebest
bestperformance
performanceamong amongall allthe
thesamples
samplestested,
tested,under
underall allheating
heating
temperature.This
temperature. Thisresult
resultwaswasattributed
attributedtotothe theenhancement
enhancementofofthe thehydration
hydrationofofunhydrated
unhydratedcement cement
clinker (self-autoclaving effect) and an improvement in the pozzolanic
clinker (self-autoclaving effect) and an improvement in the pozzolanic reaction of NA with CH, reaction of NA with CH,
resultingininan
resulting anadditional
additionalstrength
strengthgiving
givingC–A–S–H
C–A–S–Hphase. phase.
El-Gamaletetal.
El-Gamal al.[41]
[41]have
haveevaluated
evaluatedthe theeffects
effectsofof0.5,
0.5,1,1,2 2and
and3 3wt.wt.%,%,NANAadmixture
admixtureon onthethe
properties of OPC cement pastes, exposed to temperatures of 300,
properties of OPC cement pastes, exposed to temperatures of 300, 600 and 800 C. Two different 600 and 800◦ °C. Two different
coolingregimes,
cooling regimes,gradually
graduallyininair airand
andwater
waterquenching,
quenching,were wereevaluated
evaluated(Figure
(Figure14).
14).InInthe
thecase
caseofofthe
the
gradualair
gradual aircooling
coolingregimeregime(Figure
(Figure14a),
14a),all
allthe
thesamples
samplestested
testedexhibited
exhibitedaaslight
slightcompressive
compressivestrength
strength
increment,after
increment, afterexposure
exposuretoto300 300◦ C.
°C.Samples
Samplescontaining
containing0.5 0.5andand1 1wt.
wt.%%ofofNA NAexhibited
exhibitedthe thehighest
highest
strengthimprovement,
strength improvement,by by1515and
and18%18%respectively,
respectively,from fromamongamongallallthe thesamples
samplestested.
tested.ThisThiseffect
effect
was attributed to a self-autoclaving
was attributed to a self-autoclaving effect, resulting in the production of an additional
resulting in the production of an additional amount amount of of
C–
S–H phase,
C–S–H phase,asaswell
well asas
aa pozzolanic
pozzolanic reaction
reaction leading
leading toto
thetheproduction
production ofofananadditional
additional amount
amount of of
C–
A–H and
C–A–H andC–A–S–H.
C–A–S–H.After Afterexposure
exposureto to 600
600 °C,◦ C, all
all the specimens tested tested exhibited
exhibited aastrength
strengthloss,loss,
although,samples
although, samplescontaining
containingNA NAlost
lostonly
only15–21%
15–21%ofoftheir
theirinitial
initialstrength,
strength,while
whilethe
thecontrol
controlsamples
samples
lost around 50% of their initial strength. After exposure to 800 °C, all the samples tested exhibited a
dramatic loss (more than 60%) in their initial strength. However, the mixes containing 0.5 and 1% of
NA exhibited the lowest drops in strength; by 60 and 62%, respectively. In the case of the rapid
Nanomaterials 2018, 8, 465 25 of 33

lost around 50% of their initial strength. After exposure to 800 ◦ C, all the samples tested exhibited
Nanomaterials
a dramatic2018,
loss 8,(more
x FOR than
PEER 60%)
REVIEW
in 24 of1%
their initial strength. However, the mixes containing 0.5 and 32

of NA exhibited the lowest drops in strength; by 60 and 62%, respectively. In the case of the rapid
cooling regime by water quenching (Figure 14b), all the testing temperatures resulted in strength
cooling regime by water quenching (Figure 14b), all the testing temperatures resulted in strength losses
losses below the initial (unheated) strength, while at a temperature of 800 °C, samples degraded after
below the initial (unheated) strength, while at a temperature of 800 ◦ C, samples degraded after water
water quenching and lost their strength completely. Nevertheless, samples containing NS retained
quenching and lost their strength completely. Nevertheless, samples containing NS retained higher
higher compressive strengths at 300 and 600 °C. The authors concluded that the optimal value for
compressive strengths at 300 and 600 ◦ C. The authors concluded that the optimal value for improving
improving the fire resistance of cement pastes, is that of an NS addition of 0.5 wt. %.
the fire resistance of cement pastes, is that of an NS addition of 0.5 wt. %.

Figure 14.
Figure Compressivestrength
14. Compressive strength values
values of
ofcement
cementpastes
pastescontaining
containingnanoalumina
nanoalumina (NA),
(NA), after
after heating
heating
and cooling: gradually in air (a) and water quenching (b). Reproduced with permission from
and cooling: gradually in air (a) and water quenching (b). Reproduced with permission from [41]. [41].
Copyright Springer
Copyright Springer Nature,
Nature, 2017.
2017.

Farzadnia
Farzadnia et et al.
al. [70]
[70] have
have evaluated
evaluated the the effects
effects of
of elevated
elevated temperatures
temperatures (up (up toto 1000 ◦ C) on
1000 °C) on
cement
cement mortars
mortars(w/c (w/c= = 0.5), containing
0.5), containing 1, 21,and2 and3 wt. % cement
3 wt. % cementreplacement. They They
replacement. focused on mass
focused on
loss,
masspermeability,
loss, permeability,compressive strength,
compressive elasticelastic
strength, modulus,
modulus,energy absorption,
energy and brittleness.
absorption, The
and brittleness.
authors
The authorsreported thatthat
reported an optimal
an optimal amount
amount of NA
of NA(1 (1
wt.wt.
%)%) was
was beneficial
beneficialininimproving
improvingthe theresidual
residual
compressive
compressive strength
strength ofof mortars,
mortars, up up to
to aa temperature
temperature of of 800 ◦ C. The
800 °C. The addition
addition ofof 11 wt.
wt. %%NANAdecreased
decreased
the
the permeability
permeability (Figure
(Figure 15) 15) of
of samples
samples in in the
the unheated
unheated state,
state, as
as well
well asas after
after exposure
exposure to to 300 ◦ C and
300 °C and
600 ◦
600 °C,C, as
as compared
compared to to plain
plain cement
cement mortar.
mortar. Moreover,
Moreover, mortars
mortars containing
containing thethe optimum
optimum amountamount of of
NA (1 wt. %), exhibited a better effect on the relative modulus of elasticity, after
NA (1 wt. %), exhibited a better effect on the relative modulus of elasticity, after exposure to elevated exposure to elevated
temperature,
temperature, as ascompared
comparedtoto other
other NANA modified
modified specimens,
specimens, as wellas as
well ascement
plain plain cement
mortar. As mortar. As
reported
reported by the authors,
by the authors, in comparison
in comparison to control to samples,
control samples, incorporation
incorporation of NA reduced
of NA reduced the amount
the amount of CH,
of CH, contributing to an improvement in the transition zone and thus
contributing to an improvement in the transition zone and thus a denser and stronger ITZ. Compared a denser and stronger ITZ.
Compared
to C–S–H, to theC–S–H, the strength-contribution
strength-contribution potential potential
of CH is of CH is limited,
limited, becausebecause of its considerably
of its considerably lower
lower
surfacesurface area, Van
area, weak weak der Van derforces
Waals Waalsand forces and its tendency
its tendency to form an to form an oriented
oriented structure.structure.
The presenceThe
presence of a high amount of CH is associated with high porosity and bad
of a high amount of CH is associated with high porosity and bad durability, as it contains an extensive durability, as it contains
an extensive
network networkpores.
of capillary of capillary pores. Thus,
Thus, because becausedetrimental
it has several it has several detrimental
effects on cementeffects
mortaron cement
regarding
mortar regarding mechanical properties and durability, reducing CH
mechanical properties and durability, reducing CH content by the incorporation of NA is consideredcontent by the incorporation of
NA is considered to be advantageous. SEM analysis confirms that, in
to be advantageous. SEM analysis confirms that, in samples containing NA (1 and 2 wt. %), exposuresamples containing NA (1 and
2towt.
400%),◦ Cexposure
resulted in to fewer
400 °Chairline
resultedcracks,
in fewer hairline cracks,
as compared as compared
to plain to plain
cement mortar. cement mortar.
Moreover, C–S–H
Moreover, C–S–H in the matrix was found to have maintained its
in the matrix was found to have maintained its phase boundary and a less coarsened pore structure phase boundary and a less
coarsened
was observed. pore structure was observed.
Nanomaterials 2018, 8, 465 26 of 33
Nanomaterials 2018, 8, x FOR PEER REVIEW 25 of 32

Figure 15.
Figure 15. Permeability
Permeability coefficients
coefficients of
of cement
cement mortars
mortars containing
containing NA
NA and
and control
control samples
samples at
at 28,
28, 300
300
and 600
and 600 °C.
◦ Reproduced with
C. Reproduced with permission
permission from
from [70].
[70]. Copyright
Copyright Elsevier,
Elsevier, 2013.
2013.

6. The
6. The Effect
Effect of
of Nano-Iron
Nano-Iron Oxides
Oxides on
on the
the Thermal
Thermal Resistance
Resistance of
of Cement-Based
Cement-Based Composites
Composites
It has
It has already
already beenbeendemonstrated
demonstratedthat thatFe Fe2O O3 as well as Fe3O4 nanoparticles positively affect the
2 3 as well as Fe3 O4 nanoparticles positively affect the
mechanical and microstructural properties
mechanical and microstructural properties of cementitious of cementitious composites
composites when when incorporated
incorporated [77–80].
[77–80].
Fe 2O3 and Fe3O4 nanoparticles do not exhibit any noticeable chemical activity, and their positive effect
Fe2 O3 and Fe3 O4 nanoparticles do not exhibit any noticeable chemical activity, and their positive effect
on mechanical
on mechanical properties
properties and and durability
durability is is mostly
mostly attributable
attributable to to their
their nucleation
nucleation effect,
effect, as
as well
well as
as to
to
the refinement of the microstructure by the nano-filling effect. The effect of
the refinement of the microstructure by the nano-filling effect. The effect of nano-iron oxides on the nano-iron oxides on the
thermal properties
thermal properties of of cement-based
cement-based composites,
composites, has has been
been summarized
summarizedin inTable
Table3.3.
Heikal [71] analyzed the effect of Fe 2O3 nanoparticles, in the amounts of 1 and 2 wt. %, on the
Heikal [71] analyzed the effect of Fe2 O3 nanoparticles, in the amounts of 1 and 2 wt. %, on the
thermal properties of
thermal properties of cement
cement pastes
pastes exposed
exposed to to temperatures
temperatures of of up
up toto 1000
1000 ◦°C. It was
C. It was demonstrated
demonstrated
that the incorporation of a small amount of nanomaterial (1 wt. %) is
that the incorporation of a small amount of nanomaterial (1 wt. %) is beneficial for improving beneficial for improving the fire
resistance of cement pastes modified with these nanoparticles.
the fire resistance of cement pastes modified with these nanoparticles. The incorporation of Fe The incorporation of Fe2OO3
2 3
nanoparticles decreased
nanoparticles decreased specimens’
specimens’ loss loss in
in compressive
compressive strength
strength andand mass
mass after
after heating.
heating. Moreover,
Moreover,
the study showed that the presence of nano-Fe 2O3 is beneficial in diminishing crack length in cement
the study showed that the presence of nano-Fe2 O3 is beneficial in diminishing crack length in cement
pastes. Amer
pastes. Amer et et al.
al. [72]
[72] have
havestudied
studiedthe theeffects
effectsofoftemperatures
temperaturesup uptoto800800°C, onon
◦ C, thethe
performance
performance of
cement
of cement pastes
pastes with
with 1, 1,2, 2,and
and3 3wt.
wt.%%ofofnano-iron
nano-ironoxide.oxide. Bulk
Bulkdensity,
density, total
total porosity, and the
porosity, and the
compressive strength of cement pastes exposed to elevated temperatures
compressive strength of cement pastes exposed to elevated temperatures were studied. The study were studied. The study
showed that
showed that the
the bulk
bulk density
density of of nano-iron
nano-iron oxide
oxide incorporated
incorporated cement
cement pastes
pastes was
was higher
higher than
than plain
plain
cement pastes,
cement pastes, atat all
all firing
firing temperatures.
temperatures.As Asreported
reportedbybythe theauthors,
authors,this effect
this effectwas
wasattributable
attributable to
the fact that iron oxide nanoparticles can act as foreign nucleation sites,
to the fact that iron oxide nanoparticles can act as foreign nucleation sites, accelerating C–S–H gel accelerating C–S–H gel
formationand
formation andthusthus leading
leading to a compact
to a more more compact
and dense and dense Moreover,
structure. structure. ironMoreover, iron oxide
oxide nanoparticles
nanoparticles were found to exhibit a nano-filling effect, leading to a decrease
were found to exhibit a nano-filling effect, leading to a decrease in the total porosity of pastes, in the total porosity
beforeof
pastes, before and after exposure to elevated temperatures (Figure 16a).
and after exposure to elevated temperatures (Figure 16a). The study showed that an admixture of The study showed that an
admixtureoxide
nano-iron of nano-iron
has a noticeable oxide effect
has aonnoticeable effect after
strength values, on strength
all heating values, after all with
temperatures, heating
the
temperatures, with the optimal amount
optimal amount being found to be 1 wt. % (Figure 16b). being found to be 1 wt. % (Figure 16b).
Nanomaterials 2018, 8, 465 27 of 33
Nanomaterials 2018, 8, x FOR PEER REVIEW 26 of 32

Figure 16.
16. Total
Totalporosity
porosity(a)(a)
andand compressive
compressive strength
strength (b)cement
(b) of of cement
pastespastes containing
containing nano-iron
nano-iron oxide,
oxide, as a function of temperature. Redrawn
as a function of temperature. Redrawn from [72]. from [72].

7. The
7. The Effect
Effect of
of Nanotitania
Nanotitania on
on the
the Thermal
Thermal Resistance
Resistance of
of Cement-Based
Cement-Based Composites
Composites
Titanium dioxide
Titanium dioxide (NT) (NT) nanoparticles
nanoparticles find find their
their application
application mainlymainly in in the
the production
production of of
photocatalytically active cement-based surfaces, due to their ability to decontaminate
photocatalytically active cement-based surfaces, due to their ability to decontaminate pollutants via pollutants via
photocatalytic processes
photocatalytic processes [81,82].
[81,82]. Their
Their effect
effect on
on the
the hydration,
hydration, mechanical
mechanical properties,
properties, andand durability
durability
of cement-based
of cement-based composites
composites has has also
also been
been extensively
extensively studied
studied [24,83–86].
[24,83–86]. However,
However, the the incorporation
incorporation
of NT as an admixture for improving the thermal resistance of
of NT as an admixture for improving the thermal resistance of cement-based compositescement-based composites has not has
yet
gathered significant attention (Table 3). However, an extensive study
not yet gathered significant attention (Table 3). However, an extensive study undertaken by undertaken by Farzadnia et al.
[73], has demonstrated
Farzadnia et al. [73], has itsdemonstrated
potential application for theapplication
its potential above-mentioned purposes. They investigated
for the above-mentioned purposes.
the effect of NT as a cement replacement (in the amounts
They investigated the effect of NT as a cement replacement (in the amounts of 1, 2, 3 wt. %), on theofcompressive
1, 2, 3 wt. %), strength,
on the
modulus of elasticity,
compressive strength, energy
modulus absorption, andenergy
of elasticity, brittleness of cement
absorption, andmortars (w/c of
brittleness = 0.35),
cement exposed
mortarsto
elevated temperatures of up to 1000 °C. It was found that an optimal
◦ amount
(w/c = 0.35), exposed to elevated temperatures of up to 1000 C. It was found that an optimal amount of NT (i.e., 2 wt. %) was
beneficial
of NT (i.e.,in2 improving
wt. %) wasthe residualincompressive
beneficial improving the strength
residual of mortars,
compressiveup tostrength
a temperature of 600
of mortars, up°C.
to
Nevertheless, mortars ◦containing 1 wt. % exhibited a better effect on the relative
a temperature of 600 C. Nevertheless, mortars containing 1 wt. % exhibited a better effect on the modulus of elasticity,
after exposure
relative modulus to of
elevated temperatures,
elasticity, after exposure thanto other NT-modified
elevated temperatures, or plain
than cement mortars. Moreover,
other NT-modified or plain
cement mortars. Moreover, a slight decrease in the permeability of mortars modifiedofwith
a slight decrease in the permeability of mortars modified with 1 wt. % and 2 wt. % NT 1(exposed
wt. % and to
2temperatures
wt. % of NTof up to 300
(exposed to °C), was also reported.
temperatures of up to 300At ◦the
C),same time,reported.
was also after exposure
At the to 600 time,
same °C, allafter
the
samples containing ◦ NT exhibited increased permeability. SEM analysis
exposure to 600 C, all the samples containing NT exhibited increased permeability. SEM analysis showed that after exposure
to a temperature
showed that afterofexposure
400 °C, a toless coarse pore structure
a temperature was
of 400 ◦ C, present
a less in samples
coarse containing
pore structure was 1present
and 2 wt.in
% of NT, as compared to plain cement mortar. However, after exposure to
samples containing 1 and 2 wt. % of NT, as compared to plain cement mortar. However, after exposure800 °C, samples containing
NT800
to exhibited a less containing
◦ C, samples dense and compacted
NT exhibited microstructure
a less dense than plain cement
and compacted mortar.
microstructure than plain
cement mortar.
8. Discussion and Research Needs
8. Discussion and
This paper hasResearch Needs
presented a critical review of the existing research on the use of nanomaterials as
admixtures for improving
This paper the thermal
has presented a criticalresistance
review ofofthe
cementitious composites.
existing research on theInusesummary, it is clear
of nanomaterials
as admixtures for improving the thermal resistance of cementitious composites. In summary, it of
that the use nanomaterials has an outstanding effect on the improvement of the resistance is
cementitious
clear composites
that the use to elevated
nanomaterials temperature.effect
has an outstanding The onnano-sized featuresof of
the improvement thenanoparticles
resistance of
contribute to composites
cementitious modifications in the hydration
to elevated of cement,
temperature. the compaction
The nano-sized featuresdegree of its structure,
of nanoparticles and
contribute
improvement in the mechanical and fracture properties of cementitious composites,
to modifications in the hydration of cement, the compaction degree of its structure, and improvement thus improving
their
in themechanical
mechanical performance and durability
and fracture properties under elevated
of cementitious temperature
composites, conditions.
thus improving their mechanical
From all the nanomaterials examined, nanosilica
performance and durability under elevated temperature conditions. has been the most extensively studied material
and seems
From allto the
be the most advantageous
nanomaterials examined, in the production
nanosilica of heat-resistant
has been cementitious
the most extensively composites.
studied material
Researchers
and seems tohavebe theemphasized its beneficial
most advantageous influence
in the on theofhydration
production seeding
heat-resistant effect, the nano-filling
cementitious composites.
effect, as well
Researchers as its
have positive role
emphasized itsin encouraging
beneficial high on
influence pozzolanic activity,
the hydration resulting
seeding in the nano-filling
effect, retention of
strength
effect, as as
wellwell
as as
its apositive
reduction
roleininmicro-cracking.
encouraging high However,
pozzolaniccarbon nanotubes
activity, alsoinhave
resulting impressive
the retention of
mechanical properties and due to their tubular structure, exhibit bridging ability, thus restraining the
possibility of cracking in cement matrices, under thermal stress. Furthermore, other carbon-based
Nanomaterials 2018, 8, 465 28 of 33

strength as well as a reduction in micro-cracking. However, carbon nanotubes also have impressive
mechanical properties and due to their tubular structure, exhibit bridging ability, thus restraining the
possibility of cracking in cement matrices, under thermal stress. Furthermore, other carbon-based
nanomaterials have also come up in recent research works; for instance, graphene oxide, though its
incorporation requires further investigation. Other nanomaterials, such as nanoclay, nanoalumina,
nano-iron oxides, and nanotitania can also be used in cement-based composites. However, only
a few studies have been carried out in regard to utilizing these nanomaterials for improving the
thermal resistance of composites. Therefore, more research is required on the potential effects of these
nanoparticles on the thermal resistance of cementitious composites.
The following research needs can be drawn from the review presented above:

1. There is a lack of comprehensive, comparable work analyzing the properties of nanomaterials,


with similar mix designs, heating and cooling regimes and testing procedures, which would
make it possible to establish which nanomaterials have the most beneficial performance, in terms
of improving the thermal resistance of cementitious composites. In addition, studies regarding
some nanomaterials are very limited, and hence some findings from different studies are often
mutually exclusive or not fully explored.
2. Most studies are related to performance in cement pastes or cement mortars, with only very few
studies available regarding the thermal resistance of normal strength and high-strength concretes.
In addition, the effects of nanomaterials on the cracking and spalling potential of cement-based
composites are still undetermined. It is crucial that this issue addressed while considering the
further incorporation of nanomaterials in cementitious composites with compacted and dense
matrixes, such as HSCs.
3. No studies regarding optimization of mix design, by combining nanomaterials with other
SCMs, enabling a simultaneous decrease in production costs as well as an improvement in
the performance of cementitious composites under elevated temperature conditions, are available.
In addition, no evaluation of a potential combination of various nanomaterials, or incorporation
of molecular hybrids (i.e., core-shell structures) has yet been presented.
4. Most of the available studies evaluate the performance of nanomaterials based on laboratory
results. Therefore, it is necessary to verify performance in real scale tests, to obtain practical
experience in the application of nano-materials, thus enabling their further application in industry.
5. Evaluation of the long-term post-fire performance and durability of nano-modified cement-based
composites is absent. Most of the studies available are related to the effects of nanoparticles
on mechanical properties (mainly compressive strength), while less work has been devoted
to studying other properties, such as permeability, water absorption or porosity, after
thermal exposure.

To extract the full benefits of the incorporation of nanomaterials, future research should address
the abovementioned issues.

Author Contributions: P.S. outlined and mainly wrote the manuscript. M.A.E. contributed to the drafting of
the manuscript. D.S. revised, reorganized, and supervised the manuscript preparation process. All authors
contributed to the discussion, reviewed the manuscript, and accepted its final form.
Funding: This research was supported by the National Science Centre within project No. 2016/21/N/ST8/00095
(PRELUDIUM 11).
Acknowledgments: We acknowledge support by the Open Access Publication Funds of TU Berlin.
Conflicts of Interest: The authors declare no conflict of interest.
Nanomaterials 2018, 8, 465 29 of 33

Abbreviations
C–S–H calcium-silicate-hydrates
CH calcium hydroxide (portlandite)
β-C2 S β-dicalcium silicate
ITZ interfacial transition zone
OPC ordinary Portland cement
HSC high strength concrete
SFRC steel-fibre reinforced concrete
SCM supplementary cementitious material
SF silica fume
GGBS ground granulated blast-furnace slag
FA fly ash
NS nanosilica
NA nanoalumina
NC nanoclay
NMK nanometakaolin
NT nanotitania
CNT carbon nanotube
MWCNT multi-walled carbon nanotube
CF carbon fiber
CNF carbon nanofiber
GO graphene oxide
GSNS graphene sulphonate nanosheet
MIP mercury intrusion porosimetry
XRD X-ray diffraction
DSC differential scanning calorimetry
SEM scanning electron microscopy

References
1. Kodur, V. Properties of Concrete at Elevated Temperatures. ISRN Civ. Eng. 2014, 2014, 468510. [CrossRef]
2. Hager, I. Behaviour of cement concrete at high temperature. Bull. Pol. Acad. Sci. Tech. Sci. 2013, 61, 145–154.
[CrossRef]
3. Ma, Q.; Guo, R.; Zhao, Z.; Lin, Z.; He, K. Mechanical properties of concrete at high temperature—A review.
Constr. Build. Mater. 2015, 93, 371–383. [CrossRef]
4. International Federation for Structural Concrete. Fire Design of Concrete Structures—Structural Behaviour and
Assessment; State-of-Art Report; FIB: Lausanne, Switzerland, 2008.
5. Mohammed Haneefa, K.; Santhanam, M.; Parida, F.C. Review of concrete performance at elevated
temperature and hot sodium exposure applications in nuclear industry. Nucl. Eng. Des. 2013, 258, 76–88.
[CrossRef]
6. Arel, H.Ş.; Aydin, E.; Kore, S.D. Ageing management and life extension of concrete in nuclear power plants.
Powder Technol. 2017, 321, 390–408. [CrossRef]
7. Radzi, N.A.M.; Hamid, R.; Mutalib, A.A. A Review of Methods, Issues and Challenges of Small-scale Fire
Testing of Tunnel Lining Concrete. J. Appl. Sci. 2016, 16, 293–301. [CrossRef]
8. Marushchak, U.; Sanytsky, M.; Olevych, Y.; Vatulia, G.; Plugin, A.; Darenskyi, O. Effects of elevated
temperatures on the properties of nanomodified rapid hardening concretes. MATEC Web Conf. 2017, 116,
01008. [CrossRef]
9. Khoury, G.A. Effect of fire on concrete and concrete structures. Prog. Struct. Eng. Mater. 2000, 2, 429–447.
[CrossRef]
10. International Federation for Structural Concrete. Fire Design of Concrete Structures—Materials, Structures and
Modelling; FIB: Lausanne, Switzerland, 2007.
11. Xing, Z.; Beaucour, A.-L.; Hebert, R.; Noumowe, A.; Ledesert, B. Influence of the nature of aggregates on the
behaviour of concrete subjected to elevated temperature. Cement Concr. Res. 2011, 41, 392–402. [CrossRef]
Nanomaterials 2018, 8, 465 30 of 33

12. Tufail, M.; Shahzada, K.; Gencturk, B.; Wei, J. Effect of Elevated Temperature on Mechanical Properties of
Limestone, Quartzite and Granite Concrete. Int. J. Concr. Struct. Mater. 2017, 11, 17–28. [CrossRef]
13. Fernandes, B.; Gil, A.M.; Bolina, F.L.; Tutikian, B.F. Microstructure of concrete subjected to elevated
temperatures: Physico-chemical changes and analysis techniques. Rev. IBRACON Estrut. Mater. 2017,
10, 838–863. [CrossRef]
14. Haridharan, M.K.; Natarajan, C.; Chen, S.-E. Evaluation of residual strength and durability aspect of concrete
cube exposed to elevated temperature. J. Sustain. Cem.-Based Mater. 2016, 6, 231–253. [CrossRef]
15. Zhang, Q.; Ye, G.; Koenders, E. Investigation of the structures of heated Portland cement paste using various
techniques. Const. Build. Mater. 2013, 38, 1040–1050. [CrossRef]
16. Arioz, O. Effects of elevated temperatures on properties of concrete. Fire Saf. J. 2007, 42, 516–522. [CrossRef]
17. Lee, J.; Xi, Y.; Willam, K.; Jung, Y. A multiscale model for modulus of elasticity of concrete at high
temperatures. Cem. Concr. Res. 2009, 39, 754–762. [CrossRef]
18. Ozawa, M.; Morimoto, H. Effects of various fibres on high-temperature spalling in high-performance
concrete. Constr. Build. Mater. 2014, 71, 83–92. [CrossRef]
19. Yermak, N.; Pliya, P.; Beaucour, A.-L.; Simon, A.; Noumowé, A. Influence of steel and/or polypropylene
fibres on the behaviour of concrete at high temperature: Spalling, transfer and mechanical properties.
Constr. Build. Mater. 2017, 132, 240–250. [CrossRef]
20. Yuan, G.; Li, Q. The use of surface coating in enhancing the mechanical properties and durability of concrete
exposed to elevated temperature. Constr. Build. Mater. 2015, 95, 375–383. [CrossRef]
21. Szoke, S.S. Resistance to fire and high temperatures. Significance of Tests and Properties of Concrete and
Concrete-Making Material; ASTM International: West Conshohocken, PA, USA, 2006; pp. 169, 274–287.
22. Hager, I.; Mróz, K.; Tracz, T.; Hager, I.; Tracz, T. Concrete propensity to fire spalling: Testing and observations.
MATEC Web Conf. 2018, 163, 02004. [CrossRef]
23. Shah, S.P.; Hou, P.; Konsta-Gdoutos, M.S. Nano-modification of cementitious material: Toward a stronger
and durable concrete. J. Sustain. Cem.-Based Mater. 2015, 5, 1–22. [CrossRef]
24. Silvestre, J.; Silvestre, N.; de Brito, J. Review on concrete nanotechnology. Eur. J. Environ. Civ. Eng. 2015, 20,
455–485. [CrossRef]
25. Sikora, P.; Augustyniak, A.; Cendrowski, K.; Nawrotek, P.; Mijowska, E. Antimicrobial Activity of Al2 O3 ,
CuO, Fe3 O4 , and ZnO Nanoparticles in Scope of Their Further Application in Cement-Based Building
Materials. Nanomaterials 2018, 8, 212. [CrossRef] [PubMed]
26. Marushchak, U.; Sanytsky, M.; Mazurak, T.; Olevych, Y. Research of nanomodified portland cement
compositions with high early age strength. East.-Eur. J. Enterp. Technol. 2016, 6, 50–57. [CrossRef]
27. Marushchak, U.; Sanytsky, M.; Korolko, S.; Shabatura, Y.; Sydor, N. Development of nanomodified rapid
hardening fiber-reinforced concretes for special-purpose facilities. East.-Eur. J. Enterp. Technol. 2018, 2, 34–41.
[CrossRef]
28. Horszczaruk, E.; Sikora, P.; Cendrowski, K.; Mijowska, E. The effect of elevated temperature on the properties
of cement mortars containing nanosilica and heavyweight aggregates. Constr. Build. Mater. 2017, 137, 420–431.
[CrossRef]
29. Aggarwal, P.; Singh, R.P.; Aggarwal, Y.; Hussain, R.R. Use of nano-silica in cement based materials—A
review. Cog. Eng. 2015, 2, 127. [CrossRef]
30. Bastos, G.; Patiño-Barbeito, F.; Patiño-Cambeiro, F.; Armesto, J. Nano-Inclusions Applied in Cement-Matrix
Composites: A Review. Materials 2016, 9, 1015. [CrossRef] [PubMed]
31. Norhasri, M.M.; Hamidah, M.S.; Fadzil, A.M. Applications of using nano material in concrete: A review.
Constr. Build. Mater. 2017, 133, 91–97. [CrossRef]
32. Safiuddin, M.; Gonzalez, M.; Cao, J.; Tighe, S.L. State-of-the-art report on use of nano-materials in concrete.
Int. J. Pavement Eng. 2014, 15, 940–949. [CrossRef]
33. Singh, L.P.; Karade, S.R.; Bhattacharyya, S.K.; Yousuf, M.M.; Ahalawat, S. Beneficial role of nanosilica in
cement based materials—A review. Constr. Build. Mater. 2013, 47, 1069–1077. [CrossRef]
34. Sumesh, M.; Alengaram, U.J.; Jumaat, M.Z.; Mo, K.H.; Alnahhal, M.F. Incorporation of nano-materials in
cement composite and geopolymer based paste and mortar—A review. Constr. Build. Mater. 2017, 148, 62–84.
[CrossRef]
35. Mondal, P.; Shah, S.; Marks, L.; Gaitero, J. Comparative Study of the Effects of Microsilica and Nanosilica in
Concrete. Transp. Res. Rec. J. Transp. Res. Board 2010, 2141, 6–9. [CrossRef]
Nanomaterials 2018, 8, 465 31 of 33

36. Gaitero, J.J.; Campillo, I.; Guerrero, A. Reduction of the calcium leaching rate of cement paste by addition of
silica nanoparticles. Cem. Concr. Res. 2008, 38, 1112–1118. [CrossRef]
37. Jain, J.A.; Neithalath, N. Beneficial effects of small amounts of nano-silica on the chemical stability of cement
pastes exposed to neutral pH environments. In Proceedings of the ACI Fall 2009 Convention, New Orleans,
LA, USA, 8–12 November 2009; pp. 59–74.
38. Heikal, M.; El-Didamony, H.; Sokkary, T.M.; Ahmed, I.A. Behavior of composite cement pastes containing
microsilica and fly ash at elevated temperature. Constr. Build. Mater. 2013, 38, 1180–1190. [CrossRef]
39. Kumar, R.; Singh, S.; Singh, L.P. Studies on enhanced thermally stable high strength concrete incorporating
silica nanoparticles. Constr. Build. Mater. 2017, 153, 506–513. [CrossRef]
40. Heikal, M.; Al-Duaij, O.K.; Ibrahim, N.S. Microstructure of composite cements containing blast-furnace slag
and silica nano-particles subjected to elevated thermally treatment temperature. Constr. Build. Mater. 2015,
93, 1067–1077. [CrossRef]
41. El-Gamal, S.M.A.; Abo-El-Enein, S.A.; El-Hosiny, F.I.; Amin, M.S.; Ramadan, M. Thermal resistance,
microstructure and mechanical properties of type I Portland cement pastes containing low-cost nanoparticles.
J. Therm. Anal. Calorim. 2018, 131, 649–968. [CrossRef]
42. Lim, S.; Mondal, P. Effects of Nanosilica Addition on Increased Thermal Stability of Cement-Based Composite.
ACI Mater. J. 2015, 112, 305–316. [CrossRef]
43. Maheswaran, S.; Iyer, N.R.; Palani, G.S.; Pandi, R.A.; Dikar, D.D.; Kalaiselvam, S. Effect of high temperature
on the properties of ternary blended cement pastes and mortars. J. Therm. Anal. Calorim. 2015, 122, 775–786.
[CrossRef]
44. Ibrahim, R.K.; Hamid, R.; Taha, M.R. Fire resistance of high-volume fly ash mortars with nanosilica addition.
Constr. Build. Mater. 2012, 36, 779–786. [CrossRef]
45. Ibrahim, R.K.; Hamid, R.; Taha, M.R. Strength and Microstructure of Mortar Containing Nanosilica at High
Temperature. ACI Mater. J. 2014, 111, 163–170. [CrossRef]
46. Bastami, M.; Baghbadrani, M.; Aslani, F. Performance of nano-Silica modified high strength concrete at
elevated temperatures. Constr. Build. Mater. 2014, 68, 402–408. [CrossRef]
47. Shah, A.H.; Sharma, U.K.; Roy, D.A.B.; Bhargava, P.; Pimienta, P.; Meftah, F. Spalling behaviour of nano SiO2
high strength concrete at elevated temperatures. MATEC Web Conf. 2013, 6, 01009. [CrossRef]
48. Rathi, V.R.; Modhera, C.D. Effect of colloidal nano silica (CNS) on properties of high strength concrete at
elevated temperature. Int. J. Civ. Eng. Technol. 2017, 8, 618–628.
49. Sherif, M.A. Effect of Elevated Temperature on Mechanical Properties of Nano Materials Concrete. Int. J.
Eng. Innov. Technol. 2017, 7, 1–9.
50. Yan, L.; Xing, Y.; Li, J.; Zhang, J. Effect of nanosilica on the axial tensile strength of SFRC at high temperature.
Mag. Concr. Res. 2014, 66, 447–455. [CrossRef]
51. Yan, L.; Xing, Y.; Zhang, J.; Li, J. High-temperature mechanical properties and microscopic analysis of
nano-silica steel fibre RC. Mag. Concr. Res. 2013, 65, 1472–1479. [CrossRef]
52. Heikal, M.; Ali, A.I.; Ismail, M.N.; Ibrahim, S.A.N.S. Behavior of composite cement pastes containing silica
nano-particles at elevated temperature. Constr. Build. Mater. 2014, 70, 339–350. [CrossRef]
53. Yu, Z.; Lau, D. Evaluation on mechanical enhancement and fire resistance of carbon nanotube (CNT)
reinforced concrete. Coupled Syst. Mech. 2017, 6, 335–349.
54. Li, W.; Ji, W.; Torabian Isfahani, F.; Wang, Y.; Li, G.; Liu, Y.; Xing, F. Nano-Silica Sol-Gel and Carbon Nanotube
Coupling Effect on the Performance of Cement-Based Materials. Nanomaterials 2017, 7, 185. [CrossRef]
[PubMed]
55. Han, B.; Zhang, L.; Ou, J. Smart and Multifunctional Concrete Toward Sustainable Infrastructures; Springer:
Singapore, 2017.
56. Galao, O.; Baeza, F.J.; Zornoza, E.; Garcés, P. Carbon Nanofiber Cement Sensors to Detect Strain and Damage
of Concrete Specimens Under Compression. Nanomaterials 2017, 7, 413. [CrossRef] [PubMed]
57. Lu, L.; Ouyang, D. Properties of Cement Mortar and Ultra-High Strength Concrete Incorporating Graphene
Oxide Nanosheets. Nanomaterials 2017, 7, 187. [CrossRef] [PubMed]
58. Wang, L.; Zhang, S.; Zheng, D.; Yang, H.; Cui, H.; Tang, W.; Li, D. Effect of Graphene Oxide (GO) on the
Morphology and Microstructure of Cement Hydration Products. Nanomaterials 2017, 7, 429. [CrossRef]
[PubMed]
Nanomaterials 2018, 8, 465 32 of 33

59. Amin, M.S.; El-Gamal, S.M.A.; Hashem, F.S. Fire resistance and mechanical properties of carbon
nanotubes—Clay bricks wastes (Homra) composites cement. Constr. Build. Mater. 2015, 98, 237–249.
[CrossRef]
60. Chu, H.-Y.; Jiang, J.-Y.; Sun, W.; Zhang, M. Mechanical and thermal properties of graphene sulfonate
nanosheet reinforced sacrificial concrete at elevated temperatures. Constr. Build. Mater. 2017, 153, 682–694.
[CrossRef]
61. Han, T.; Wang, H.; Jin, X.; Yang, J.; Lei, Y.; Yang, F.; Yang, X.; Tao, Z.; Guo, Q.; Liu, L. Multiscale carbon
nanosphere–carbon fiber reinforcement for cement-based composites with enhanced high-temperature
resistance. J. Mater. Sci. 2015, 50, 2038–2048. [CrossRef]
62. Mohammed, A.; Sanjayan, J.G.; Nazari, A.; Al-Saadi, N.T.K. Effects of graphene oxide in enhancing the
performance of concrete exposed to high-temperature. Aust. J. Civ. Eng. 2017, 15, 61–71. [CrossRef]
63. Zhang, L.W.; Kai, M.F.; Liew, K.M. Evaluation of microstructure and mechanical performance of
CNT-reinforced cementitious composites at elevated temperatures. Compos. Part A Appl. Sci. Manuf.
2017, 95, 286–293. [CrossRef]
64. Irshidat, M.R.; Al-Saleh, M.H. Thermal performance and fire resistance of nanoclay modified cementitious
materials. Constr. Build. Mater. 2018, 159, 213–219. [CrossRef]
65. Lee, S.-J.; Kim, S.-H.; Won, J.-P. Strength and fire resistance of a high-strength nano-polymer modified
cementitious composite. Compos. Struct. 2017, 173, 96–105. [CrossRef]
66. Ho, C.M.; Tsai, W.T. Effect of Elevated Temperature on the Strength and Ultrasonic Pulse Velocity of Glass
Fiber and Nano-Clay Concrete. AMR Adv. Mater. Res. 2010, 163–167, 1532–1539. [CrossRef]
67. Wang, W.-C. Compressive strength and thermal conductivity of concrete with nanoclay under Various
High-Temperatures. Constr. Build. Mater. 2017, 147, 305–311. [CrossRef]
68. Morsy, M.S.; Al-Salloum, Y.A.; Abbas, H.; Alsayed, S.H. Behavior of blended cement mortars containing
nano-metakaolin at elevated temperatures. Constr. Build. Mater. 2012, 35, 900–905. [CrossRef]
69. Heikal, M.; Ismail, M.N.; Ibrahim, N.S. Physico-mechanical, microstructure characteristics and fire resistance
of cement pastes containing Al2 O3 nano-particles. Constr. Build. Mater. 2015, 91, 232–242. [CrossRef]
70. Farzadnia, N.; Abang Ali, A.A.; Demirboga, R. Characterization of high strength mortars with nano alumina
at elevated temperatures. Cem. Concr. Res. 2013, 54, 43–54. [CrossRef]
71. Heikal, M. Characteristics, textural properties and fire resistance of cement pastes containing Fe2 O3
nano-particles. J. Therm. Anal. Calorim. 2016, 126, 1077–1087. [CrossRef]
72. Amer, A.A.; El-Sokkary, T.M.; Abdullah, N.I. Thermal durability of OPC pastes admixed with nano iron
oxide. HBRC J. 2015, 11, 299–305. [CrossRef]
73. Farzadnia, N.; Abang Ali, A.A.; Demirboga, R.; Anwar, M.P. Characterization of high strength mortars with
nano Titania at elevated temperatures. Constr. Build. Mater. 2013, 43, 469–479. [CrossRef]
74. Nanoclay Reinforced Polymer Composites: Nanocomposites and Bionanocomposites; Jawaid, M.; Qaiss, A.E.K.;
Bouhfid, R. (Eds.) Springer: Singapore, 2016.
75. Morsy, M.S.; Alsayed, S.H.; Aqel, M. Effect of Nano-clay on Mechanical Properties and Microstructure of
Ordinary Portland Cement Mortar. Int. J. Civ. Environ. Eng. 2010, 10, 21–25.
76. Chang, T.-P.; Shih, J.-Y.; Yang, K.-M.; Hsiao, T.-C. Material properties of Portland cement paste with
nano-montmorillonite. J. Mater. Sci. 2007, 42, 7478–7487. [CrossRef]
77. Amin, M.S.; El-Gamal, S.M.A.; Hashem, F.S. Effect of addition of nano-magnetite on the hydration
characteristics of hardened Portland cement and high slag cement pastes. J. Therm. Anal. Calorim. 2013, 112,
1253–1259. [CrossRef]
78. Nazari, A.; Riahi, S. Computer-aided design of the effects of Fe2 O3 nanoparticles on split tensile strength
and water permeability of high strength concrete. Mater. Des. 2011, 32, 3966–3979. [CrossRef]
79. Shekari, A.H.; Razzaghi, M.S. Influence of Nano Particles on Durability and Mechanical Properties of High
Performance Concrete. Procedia Eng. 2011, 14, 3036–3041. [CrossRef]
80. Sikora, P.; Horszczaruk, E.; Cendrowski, K.; Mijowska, E. The Influence of Nano-Fe3 O4 on the Microstructure
and Mechanical Properties of Cementitious Composites. Nanoscale Res. Lett. 2016, 11, 182. [CrossRef]
[PubMed]
81. Sikora, P.; Augustyniak, A.; Cendrowski, K.; Horszczaruk, E.; Rucinska, T.; Nawrotek, P.; Mijowska, E.
Characterization of Mechanical and Bactericidal Properties of Cement Mortars Containing Waste Glass
Aggregate and Nanomaterials. Materials 2016, 9, 701. [CrossRef] [PubMed]
Nanomaterials 2018, 8, 465 33 of 33

82. Lucas, S.S. Influence of operating parameters and ion doping on the photocatalytic activity of mortars
containing titanium dioxide nanoparticles. Mater. Today Proc. 2017, 4, 11588–11593. [CrossRef]
83. Jalal, M.; Fathi, M.; Farzad, M. Effects of fly ash and TiO2 nanoparticles on rheological, mechanical,
microstructural and thermal properties of high strength self compacting concrete. Mech. Mater. 2013,
61, 11–27. [CrossRef]
84. Ma, B.; Li, H.; Mei, J.; Li, X.; Chen, F. Effects of Nano-TiO2 on the Toughness and Durability of Cement-Based
Material. Adv. Mater. Sci. Eng. 2015, 2015, 583106. [CrossRef]
85. Zhang, R.; Cheng, X.; Hou, P.; Ye, Z. Influences of nano-TiO2 on the properties of cement-based materials:
Hydration and drying shrinkage. Constr. Build. Mater. 2015, 81, 35–41. [CrossRef]
86. Pérez-Nicolás, M.; Navarro-Blasco, Í.; Fernández, J.M.; Alvarez, J.I. The Effect of TiO2 Doped Photocatalytic
Nano-Additives on the Hydration and Microstructure of Portland and High Alumina Cements. Nanomaterials
2017, 7, 329. [CrossRef] [PubMed]

© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

You might also like