Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Behavior of High-Strength Polypropylene Fiber-Reinforced

Self-Compacting Concrete Exposed to High Temperatures


José D. Ríos 1; Héctor Cifuentes, Ph.D. 2; Carlos Leiva, Ph.D. 3; Celia García, Ph.D. 4; and María D. Alba, Ph.D. 5
Downloaded from ascelibrary.org by MISSOURI, UNIV OF/COLUMBIA on 08/03/18. Copyright ASCE. For personal use only; all rights reserved.

Abstract: In this study we analyzed the use of high-performance structural concrete reinforced with polypropylene fibers in applications
requiring long exposure times to high temperatures, such as thermal energy storage systems. We analyzed the behavior of the concrete
at different temperatures (hot tests: 100°C, 300°C, 500°C and 700°C), cooled-down states (cold tests) and exposure times (6, 24, and
48 h). We also experimentally determined the thermogravimetric analysis, fracture behavior, compressive strength, Young’s modulus,
and tensile strength of concrete. Subsequently, we performed a comprehensive analysis of the thermal and mechanical behavior of
high-performance concrete under different thermal conditions. We applied longer exposure times to broaden the available results on the
behavior of high-performance fiber-reinforced concrete when subjected to high temperatures. Results show that, once thermal and moisture
equilibriums are reached, exposure time does not have any influence on mechanical properties. They also provide useful information about
the influence of high temperatures on the different parameters of fiber-reinforced concrete and its application for thermal energy storage
structures. DOI: 10.1061/(ASCE)MT.1943-5533.0002491. © 2018 American Society of Civil Engineers.
Author keywords: High-strength concrete; Polypropylene fibers; High temperatures; Exposure time; Fracture behavior.

Introduction variable periods of time during plant operation. One of the uses
of concrete in thermal storage systems is in steam collectors. These
The use of renewable energies as a main energy source is a topic of systems directly use steam as a means of storing thermal energy:
great interest for those countries lacking fossil energy sources, and when liquid water is heated and turned into steam, the steam goes to
for those seriously committed to the reduction of CO2 emissions. the turbine or is alternatively stored in concrete tanks or basins to be
For this reason, although renewable energy sources are already used when solar radiation is interrupted. In these cases, the concrete
being used, their technologies are the subject of ongoing research is subjected to high temperatures for long periods of time, being in
(Alotto et al. 2014; Alonso et al. 2016; Chauhan and Saini 2014). direct contact with the steam. Therefore, the heat loss factor, heat
Solar energy is one of the most widespread renewable energy op- storage power, and cost of the containing material are all key as-
tions worldwide because it is generally abundant, and the predic- pects. A concrete steam collector must be able to withstand a tem-
tion of solar radiation hours, and hence of energy production, is perature range between approximately 100°C and 300°C.
very easy when compared with other renewable energy sources Basically, energy storage systems must be composed of two
(Alonso et al. 2016). Additionally, the technology available for materials. One of them should have high heat storage capabilities,
the production of electricity through solar energy is also the only such as those of certain mixtures of molten salts or phase-change
one that allows for its storage. The development of a solar energy materials (Alonso et al. 2016; Herrmann et al. 2004; Laing et al.
storage system is therefore crucial to produce energy in a more uni- 2009; Pielichowska and Pielichowski 2014; Yang and Garimella
form way and would allow this renewable energy to be more com- 2010); the other one should be a structural material that functions
petitive against fossil energy sources. Storage system materials as a containment vessel for the former, and must be properly insu-
must not only be capable of bearing loads, but also, most impor- lated through the use of refractory materials to ensure it is not dam-
tantly, they must withstand the effects of high temperatures for aged by high temperatures. In this regard, concrete is emerging as a
material with a twofold application: on the one hand, it is one of the
1
most widely used structural materials because of its mechanical
Ph.D. Student, Grupo de Estructuras, Escuela Técnica Superior de properties, on-site workability, easy handling, worldwide availabil-
Ingeniería, Universidad de Sevilla, Sevilla, Andalucia 41092, Spain.
2 ity, and relatively low cost (Alonso et al. 2016); on the other hand,
Associate Professor, Grupo de Estructuras, Escuela Técnica Superior
de Ingeniería, Universidad de Sevilla, Sevilla, Andalucia 41092, Spain concrete has interesting heat storage capabilities due to its thermal
(corresponding author). ORCID: https://orcid.org/0000-0001-6302-418X. properties (Alonso et al. 2016). In this regard, examples are found
Email: bulte@us.es in the literature of heat storage technologies based on concrete as a
3 heat storage material that use a series of pipes filled with water as
Associate Professor, Grupo de Ingeniería de Residuos, Escuela Técnica
Superior de Ingeniería, Universidad de Sevilla, Sevilla, Andalucia 41092, a heat transfer fluid inside a concrete block (Alonso et al. 2016;
Spain. Laing et al. 2009) that contains phase-change materials (Fenollera
4
Grupo de Ingeniería de Residuos, Escuela Técnica Superior de et al. 2013; Memon et al. 2015).
Ingeniería, Universidad de Sevilla, Sevilla, Andalucia 41092, Spain. Normally, the maximum serviceability temperature of structural
5
Professor, Instituto de Ciencia de Materiales, Consejo Superior de
concrete is limited by the potential damage caused by the high
Investigaciones Científicas-Universidad de Sevilla, Sevilla 41092, Spain.
Note. This manuscript was submitted on November 9, 2017; approved
temperatures involved. This damage is manifested as microcracks
on May 2, 2018; published online on August 1, 2018. Discussion period originated by the pore pressure inside the concrete matrix, due to
open until January 1, 2019; separate discussions must be submitted for the evaporated water (Cifuentes et al. 2012; Zhang and Bicanic
individual papers. This paper is part of the Journal of Materials in Civil 2006). If the concrete has good matrix porosity, as happens in
Engineering, © ASCE, ISSN 0899-1561. the case of normal strength concrete, the evaporated water can flow

© ASCE 04018271-1 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2018, 30(11): 04018271


outside the concrete through the pore connection of the matrix. degradation than normal concrete when exposed to high tempera-
This effect, combined with medium-high temperatures, causes neg- tures, and is more affected by spalling and cracking (Alonso et al.
ligible effects of the pore pressure and thus results in no damage to 2016). The addition of polypropylene (PP) fibers reduces these
the concrete matrix (Cifuentes et al. 2012; Srikar et al. 2016; Zhang negative effects and improves the mechanical properties of concrete
and Bicanic 2006). However, in the case of dense concrete mixes at high temperatures (Chauhan and Saini 2014; Yang and Garimella
such as high-strength concrete or if the concrete is subjected to 2010). Fiber length is an important factor in enhancing the fracture
very high temperatures, the pore pressure thus originated can be behavior of concrete at room temperature, because longer fibers
very harmful for the concrete matrix, causing failure of the material provide higher resistance to pull-out failure (Herrmann et al. 2004).
through spalling (Cifuentes et al. 2012; Zhang and Bicanic 2006). Another important aspect to consider in these applications is
The only situation in which the behavior of structural concrete the study of the behavior of concrete over long exposure times
under high temperatures has been studied is in fire conditions at a constant temperature and in the subsequent cooling process.
Due to the usual limitations in the serviceability temperature of
Downloaded from ascelibrary.org by MISSOURI, UNIV OF/COLUMBIA on 08/03/18. Copyright ASCE. For personal use only; all rights reserved.

(Kang et al. 2016; Liu et al. 2008; Zoalfakar et al. 2016). In such
studies, the structural material has been analyzed in an accidental concrete, there are few studies on this topic in the literature, and the
ultimate limit state in accordance with the standards. Standardized exposure times explored are very moderate (Mohamedbhai 1986).
fire-resistance tests of structural materials normally involve gradu- Residual mechanical properties after cooling and exposure time are
ally heating the material to a very high temperature (over 700°C). two essential aspects to study because, in the applications de-
However, in the case of moderate temperatures such as the operat- scribed, concrete is exposed to high temperatures during long peri-
ing temperatures of solar energy storage systems (≤300°C), if the ods and is alternatively subjected to working and resting stages.
concrete has adequate porosity there is no apparent reason to expect The use of self-compacting concrete is widespread, due to its
a dramatic decrease in the strength of the concrete. Below 300°C, benefits for a more efficient production process and the high per-
no major decomposition of the hydration products of the cement formance of the final product. The industry of precast products
should be observed (Cifuentes et al. 2012), and any eventual dam- (Barros et al. 2007) is one of the areas in which this type of concrete
age should only be caused by the pore pressure originated by the is used due to its better fluidity, which improves the final quality of
evaporated water inside the concrete (Srikar et al. 2016). However, the product and reduces construction time. Self-compacting con-
crete, sometimes reinforced with PP or steel fibers, is also used
these pore pressures may well not be high enough to cause any
in high-quality concrete projects such as buildings and industrial
mechanical damage to the concrete; in addition, the energy sup-
structures that need to withstand high temperatures during opera-
plied to the concrete through the heating process contributes to
tion or in case of an accidental fire (Bamonte and Pietro 2016;
hydrate unhydrated cement particles, which actually counteracts
Noumowé et al. 2006; Srikar et al. 2016; Tao et al. 2010).
the effects of thermal damage to the concrete and increases its
In this study, we analyzed the influence of different tempera-
mechanical properties (Kodur 2014).
tures on the mechanical behavior of high-strength self-compacting
In these applications of structural concrete for thermal energy
concrete mixes, both plain and polypropylene fiber-reinforced
storage, the study of the fracture behavior of the concrete is key
(6 and 24 mm fiber lengths), during three different exposure times
to control any potential damage caused. Moreover, it is usually nec- (6, 24, and 48 h) in hot specimens, and only one exposure time
essary to use high-performance concrete with high strength capac- (48 h) in cold specimens (after heating). We paid special attention
ity, especially tensile strength, to prevent premature cracking due to to the fracture behavior of these concrete mixes. Specifically, we
pore pressure. However, this type of concrete undergoes greater conducted standardized tests to characterize the mechanical behav-
ior of the various mixes at different temperatures and exposure
times. These properties were directly measured in heated and cold
Table 1. Chemical composition (%) of the materials concrete specimens.
Chemical
composition PC FA LA CA SF
SiO2 20.69 86.50 86.97 83.35 95.63
Experimental Study
Al2 O3 5.24 5.83 5.56 4.50 0.43
Fe2 O3 2.46 1.33 1.39 2.26 0.10 Materials and Mix Design
MnO 0.10 0.03 0.02 0.05 0.02
MgO 1.50 0.13 0.19 0.36 0.39 A summary of the materials used in the present study and their
CaO 61.89 0.59 0.51 3.42 0.21 characteristics is shown below:
Na2 O 0.36 0.87 0.95 0.09 0.20 1. Portland cement (PC) Type II 32.5 B-L, in accordance with
K2 O 0.73 2.37 1.97 1.04 0.79 the EN 197-1 standard [EN197-1:2000/A3 (CEN 2008)]. Its
TiO2 0.28 0.13 0.15 0.23 <0.01 chemical composition is provided in Table 1. It is essentially
P2 O 5 0.17 0.07 0.02 0.05 0.08 SiO2 and CaO.
SO3 1.11 0.04 0.03 0.03 0.04 2. Fine aggregate (FA) with a maximum size of 2 mm. Its grading
LOI 4.76 1.34 1.51 3.59 1.84
distribution is shown in Table 2. Its chemical composition is
Specific gravity 3.18 2.69 2.58 2.48 2.30
shown in Table 1. FA was essentially SiO2 and Al2 O3 .

Table 2. Grading distribution


Sieve size (mm)
Material >16 10 8 4 2 1 0.5 0.25 0.125 0.063 <0.063
FA 0 0 0 0 0 22.20 35.82 26.15 10.33 3.74 1.10
LA 0 0 0 0 11.50 26.50 21.00 17.33 12.00 9.17 2.33
CA 0 19.30 40.13 38.31 2.26 0 0 0 0 0 0

© ASCE 04018271-2 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2018, 30(11): 04018271


Table 3. Mix design of all concrete types the actuator, the specimen and the supporting structure into the
Quantity (kg=m3 ) furnace.
W/PC
Type PC FA LA CA SF SP PPF ratio
Mass Loss and Structural Effects
D0 657 867 139 1,301 99 23 0 0.20
D1 657 867 139 1,301 99 23 1.2 (PPF-6) 0.20 The effect of temperature on the concrete was monitored by
D2 657 867 139 1,301 99 23 1.2 (PPF-24) 0.20 thermal analysis and variable temperature X-ray powder diffrac-
tometry (VTXRD) analysis. VTXRD patterns were recorded at the
CITIUS X-ray laboratory (University of Seville, Spain) by a
3. Limestone sand (LA) with a maximum size of 4 mm (Table 2). Bruker D8 Advance diffractometer (Bruker, Germany) fitted with
Its chemical composition is shown in Table 1. LA was composed a high-temperature camera (Anton Paar XRK 900, Austria) and a
essentially of SiO2 and Al2 O3 , and its chemical composition position-sensitive detector (Bruker Vantec PSD, Germany) in the
Downloaded from ascelibrary.org by MISSOURI, UNIV OF/COLUMBIA on 08/03/18. Copyright ASCE. For personal use only; all rights reserved.

was very similar to FA. 25°C–900°C temperature range at a heating rate of 10°C · min−1
4. Coarse aggregate (CA) with a maximum size of 10 mm (Table 2). (target: Cu; voltage: 40 kV; current: 40 mA; θ∶θ geometry com-
Its chemical composition is shown in Table 1. CA was composed bining divergent Göbel mirror configurations; detector: radial
essentially of SiO2 and Al2 O3 Soller slits; time scans: 5 min).
5. Silica fume (SF) manufactured by Sika, model S92-D, with Simultaneous thermogravimetric (TG/DTA) measurements
90% of particles under 0.1 μm and a SiO2 content above 90% were performed at the CITIUS Functional Characterization Service
(Table 1). SF is essentially SiO2. (University of Seville, Spain) on a TA (model STD-Q600) instru-
6. Superplasticizer (SP) (SIKA VISCOCRETE 20 HE), specific ment, using alumina as a reference. The specimens were put into
for low water/binder ratio mixes. platinum crucibles in a chamber containing air during the entire
7. Polypropylene (PP) fibers with a diameter of 33 μm and two heating period. The temperature was increased at a constant rate
different lengths, 6 (PPF-6) and 24 mm (PPF-24). The tensile of 10°C=min.
strength of the PP fibers was 450 MPa. A differential scanning calorimetric analysis (DSC) was carried
Table 2 shows the grading distribution of FA, LA, and CA. As out. The specimens, 5 mm in diameter and 3 mm thick, were placed
shown in this table, FA was the finest, with an average particle size in nonhermetic aluminum containers and subjected to a heating
of 0.48 mm; LA had an average particle size of 0.66 mm, and CA program of 2°C=min in a TA Instruments 2920 DSC from 40°C
had an average particle size of 6.72 mm. up to 300°C, using nitrogen as a purge gas (Vilches et al. 2005).
In this study, we analyzed three different self-compacting con-
crete mixes, differentiated only by their respective PP fiber content.
The first mix, called D0, was considered as the reference concrete Pore Size Distribution
mix and was not reinforced with PP fibers. The second and third Pore size was determined using a mercury intrusion porosimeter
concrete mixes, called D1 and D2, respectively, had a PP fiber with a measuring pressure range of 1.02 × 10−2 to 2.04 × 102 MPa.
content of 1.2 kg=m3 . Fiber length was 6 mm in the D1 concrete The selected contact angle was 141°, so the measurable pore size
and 24 mm in the D2 concrete. The concrete mix proportions are ranged from 0.007 to 144 pm. Pellet-shaped specimens about
specified in Table 3. 5 mm in size were used. All specimens were previously heated to
A laboratory mortar mixer was used to prepare the different 105°C until no moisture was left.
types of concrete. First, all dry constituents were poured into
the mixer and rotated for 5 min. Next, the PP fibers and the super-
Compressive Strength (f c )
plasticizer were added to the water and then poured into the mixer.
Finally, the mixer was left to rotate for a further 15 min until a fluid The compressive strength tests of the control concrete mix and the
and homogenous mix was obtained. Slump flow tests were con- hot and cold specimens were carried out in accordance with the
ducted on the self-compacting concrete (SCC) mixes in accordance EN12390-3 standard (CEN 2009a) in 50 × 50 × 50 mm-sized
with EFNARC guidelines (EFNARC 2005), with satisfactory cubes. These cubes were obtained by sawing off the 100 × 100 ×
results. The specimens were immediately cast, subsequently un- 440 mm cast prisms. All the prisms were made with the same con-
molded, and after 24 h submerged in a water bath at a temperature crete matrix to avoid any differences in the properties of the
of 20°C for 28 days. material. The specimens of each batch were all simultaneously
heated in the same furnace to ensure that they were all subjected
Heating Test to the same thermal cycle. The load was increased continuously at
a rate of approximately 0.5 MPa=s until the specimen ruptured
To explore the influence of temperature on the fracture behavior of completely. Four cube specimens were tested in each case.
PP fiber-reinforced SCC, experimental tests were performed on
specimens at high temperatures (hot tests) and also after cooling
Tensile Strength (f t )
them down (cold tests). Additionally, a first batch of specimens
of each mix (D0, D1, and D2) was tested at room temperature, The tensile strength of concrete is an important mechanical prop-
without undergoing any heating, to serve as control results. In the erty for the analysis of its fracture behavior. Cylindrical specimens
so-called hot tests, the second batch of specimens was heated at a 100 × 200 mm in size (diameter × height) were manufactured with
rate of 10°C=min until reaching five different temperatures (100°C, each concrete mix. Due to the complexity of direct-tension testing
300°C, 500°C, and 700°C), subjected to three different exposure in concrete, indirect tensile strength tests such as the Brazilian split-
times (6, 24, and 48 h) at each of these temperatures, and immedi- cylinder strength test were used (Planas 2007). The experimental
ately tested. In the so-called cold tests, the third batch of specimens tests were performed on specimens at room temperature (i.e., con-
was heated at the same rate and the same temperatures for 48 h and trol concrete and cold specimens). In the hot tests, due to limitations
then cooled for 24 h at room temperature, after which they were of the testing machine, tensile strength evolution was estimated
tested. The testing equipment used was specifically designed to from the maximum load value of three-point bending tests. It is
perform high-temperature tests, and made it possible to insert noted that the right determination of the tensile strength from

© ASCE 04018271-3 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2018, 30(11): 04018271


Downloaded from ascelibrary.org by MISSOURI, UNIV OF/COLUMBIA on 08/03/18. Copyright ASCE. For personal use only; all rights reserved.

Fig. 1. Three-point bending test of a notched specimen.

three-point bending tests requires a more sophisticated procedure From the P-δ curve, the fracture energy of concrete was calcu-
(see Guan et al. 2016, 2017; Hu et al. 2017; Wang et al. 2016); lated using Eq. (1) specified in the RILEM (1985), applying the
however, as the main objective is the analysis of the evolution corrections corresponding to the adjustment of the tail of the load-
of the tensile strength with temperature, a direct relationship be- displacement curve proposed by Guinea et al. (Cifuentes et al.
tween f t and the peak load of the three-point bending (TPB) tests 2017; Guinea et al. 1992).
was assumed. This simplified procedure was verified by using
WF
specimens at room temperature (i.e., control concrete and cold GF ¼ ð1Þ
specimens) and obtaining tensile strength values in agreement with BðD − aÞ
their corresponding experimental results. The procedure proved to where GF = fracture energy (N=m), W F = total work of fracture,
be a reasonable way to estimate this material property in hot tests, and BðD − a0 Þ is the ligament area. The total work of fracture is
given the difficulty of experimental testing. given by:
The Brazilian strength tests were carried out on the 100 ×
200 mm cylinders following the procedures recommended by the W F ¼ W m þ W nm ð2Þ
EN 12390-6 standard (CEN 2009b). Plywood strips 15 mm wide ×
4 mm thick were placed between the specimens and the actuator. W m = measured work of fracture corresponding to the total area
The displacement velocity of the machine actuator was 2.8 kN=s, below the P-δ curve recorded until the end of the test at δ u .
and four specimens were tested in each case. W nm = nonmeasured work of fracture, obtained by means of the
remote tail constant, A, taking into account the influence of cutting
the P-δ curve (Elices et al. 1992)
Determination of the Fracture Energy (G F )
2A
Notched beams 50 × 50 × 220 mm (width × depth × length ¼ W nm ¼ ð3Þ
δu
B × D × L) were used to determine the hot and cold fracture
energy. The fracture energy was assessed with stable TPB tests in
accordance with the RILEM work-of-fracture method (RILEM Young’s Modulus (E c )
1985). Certain modifications such as the recommendations pro- The Young’s modulus of the specimens tested at room temperature
posed by Guinea et al. (Elices et al. 1992; Guinea et al. 1992) were (i.e., control concrete and cold specimens) was experimentally cal-
considered to obtain a size-independent value. Four specimens culated in accordance with the EN 12390-13:2014 standard (CEN
were tested in each case. Specimen length was L ¼ 220 mm, so 2014). This was done by gradually loading a cylindrical specimen
specimen span length was S ¼ 200 mm. It was equal to four times in compression to approximately a third of its failure load and
the depth of the cross section; therefore, the influence of the span/ measuring the corresponding strain with 30 mm strain gauges.
depth ratio (Karihaloo et al. 2003) was avoided. Initial notch depth By contrast, it was impossible to use measuring instruments in the
was a ¼ 25 mm, so relative notch depth was α ¼ 0.5 in all cases. hot tests, so the Young’s modulus was calculated based on the com-
The notches were prepared with a diamond saw prior to the heating pliance method proposed by Jenq and Shah (Shah 1990) for these
process. specimens. This method requires previous knowledge of the
The three-point bending tests were performed in a 200 kN initial slope of the load-crack mouth opening displacement curve
dynamic machine, as illustrated in Fig. 1. The vertical displacement (P-CMOD), which was not measured in the hot specimens. The
at the midpoint in specimens at room temperature (i.e., control initial slope of the P-CMOD curve is related to the Young’s modu-
concrete and cold specimens) was measured with a linear voltage lus, as shown in Eq. (5). As a result, the initial slope of the load-
displacement transducer (LVDT) mounted on a rigid frame fixed to displacement curve (P-δ) is also related to material stiffness, so it
the specimen. In hot specimens, however, this measurement was is possible to obtain an initial slope of the P-CMOD curves and
obtained from the actuator position recording, because the high establish their relationship with the P-δ curves. In the hot tests, the
temperatures in the hot tests made it impossible to use any meas- P-CMOD curve was estimated from load-actuator-position curves,
uring instrument. Loading was conducted using an actuator, at a applying the aforementioned rate. The estimated P-CMOD was
rate of 0.01 mm=min until the peak load was reached; after this, used to calculate the Young’s modulus by means of Jenq and Shah’s
the velocity was increased to 0.05 mm=min during the postpeak compliance method (Shah 1990).
branch of the load-displacement (P-δ) curve. The supporting brack- According to these indications, after obtaining the initial compli-
ets were mounted on steel ball bearings to allow them to freely ance (Ci ) of the P-CMOD curves for all hot specimens, the elastic
rotate around the beam’s major axis. modulus of concrete was obtained using the following expression:

© ASCE 04018271-4 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2018, 30(11): 04018271


6SaV 1 ðα 0 Þ
Ec ¼ ð4Þ
Ci BD2

where V 1 ðα 0 Þ = geometric function given by

0.66
V 1 ðα 0 Þ ¼ 0.8 − 1.7α 0 þ 2.4α 02 þ
ð1 − α 0 Þ2
4D
þ ð−0.04 − 0.58α 0 þ 1.47α 02 − 2.04α 03 Þ
S
a þ HO
α0 ¼ ð5Þ
D þ HO
Downloaded from ascelibrary.org by MISSOURI, UNIV OF/COLUMBIA on 08/03/18. Copyright ASCE. For personal use only; all rights reserved.

and HO = clip gauge holder thickness. In this case, HO was 1 mm


and there was no accuracy loss if α 0 ≈ α.
Obviously, according to Eq. (4), materials with higher initial
P-CMOD compliance curves have lower elastic modulus values.

Fig. 3. Mineral composition of D0.


Results and Discussion

Mass Loss Fig. 4 shows the DSC analysis of the polypropylene fibers. The
The thermogravimetric analysis of D0 from room temperature to melting process took place over a range of temperatures between
1,000°C after 28 days is shown in Fig. 2. As observed, D0 exhibited 145°C and 180°C, as seen from the width of the melting endotherm.
three different weight losses. The first was observed at 100°C, The melting point, that is, the peak of the melting endotherm, was
represented a mass loss of 4.2% wt and was due to the moisture 165°C.
present in D0. The second, observed between 200°C and 500°C, As shown in Fig. 5, the fibers showed only one mass loss at
represented a mass loss of only 1.5% wt. It was due to the dehy- 460°C, because at that point they were transformed into water
dration of chemically bound water in several hydrates of the con- and CO2 .
crete (i.e., calcium silicate, carboaluminates, ettringite), although
primarily to the dehydration of calcium silicate hydrates at 450°C. Fracture Energy
The third was observed between 650°C and 750°C, represented
a mass loss of 3.5% wt and was caused by the decarbonation of The fracture energy values measured in each hot and cold test at the
calcium carbonate from the clinker and the filler (Alarcon-Ruiz different exposure times are presented in Fig. 6. In D0 hot tests
et al. 2005). [Fig. 6(a)], the fracture energy showed a decreasing trend at
The mineral composition of a D0 specimen portion at different 100°C. At that point, the concrete matrix still did not show any
temperatures (i.e., from room temperature to 900°C) after 28 days relevant microcracking, so the vapor could not be freely evacuated.
is shown in Fig. 3. The main mineral compounds at room temper- This led to high moisture gradients in the matrix, entailing a reduc-
ature were quartz, PDF 01-085-0335; (q), calcite: CaCO3 , PDF tion in the fracture energy (Zhang and Bicanic 2006). Above
01-083-0578; (c), C3 S: Ca3 SiO5 , PDF 00-055-0738; and (s), albite: 100°C, the fracture energy rose until a temperature of 500°C was
NaAlSi3 O8 , PDF 01-089-6427 (a). When the temperature of D0 reached, as shown by the increased area under the load-deflection
reached 700°C, the calcite disappeared, because it decomposed into curves [Figs. 7(b and c)], which increased along with the temper-
CO2 and CaO, as shown in Fig. 2. ature. Although the maximum load decreased due to microcracking
by spalling, deflection increased (Zhang and Bicanic 2006). At
700°C, a significant amount of microcracking occurred due to the

Fig. 2. TG of concrete: D0 (without PP fibers). Fig. 4. DSC of PP fibers.

© ASCE 04018271-5 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2018, 30(11): 04018271


Downloaded from ascelibrary.org by MISSOURI, UNIV OF/COLUMBIA on 08/03/18. Copyright ASCE. For personal use only; all rights reserved.

Fig. 5. TGA of the PP fibers.

Fig. 6. Evolution of the fracture energy with temperature for each exposure time: (a) D0 hot tests; (b) D1 hot tests; (c) D2 hot tests; and (d) cold tests.

decarbonation of the calcium carbonate (Figs. 2 and 3), signifi- concrete (Cifuentes et al. 2012, 2013). At low temperatures such as
cantly reducing the resistance of the material (Cifuentes et al. 2012; 25°C and 100°C, the condition of the PP fibers was not affected by
Tufail et al. 2017) and dramatically reducing the fracture energy. temperature (Fig. 4), and their main role was the stitching of even-
In the D1 and D2 hot tests performed on the PP fiber-reinforced tual cracks. At the aforementioned temperatures, the fibers gave the
mixes [Figs. 6(b and c), respectively], the fracture energy was concrete higher ductility, as inferred from the higher fracture energy
slightly higher at 25°C and 100°C because the fibers compensated values [Figs. 6(b and c)]. The fracture energy values increase in the
the effects of the high pressure and moisture gradients within the D1 and D2 hot tests at 300°C, because the PP fibers had partially

© ASCE 04018271-6 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2018, 30(11): 04018271


Downloaded from ascelibrary.org by MISSOURI, UNIV OF/COLUMBIA on 08/03/18. Copyright ASCE. For personal use only; all rights reserved.

Fig. 7. Load-deflection curves in D0 hot tests: (a) 100°C; (b) 300°C; (c) 500°C; and (d) 700°C.

melted at this point (Fig. 4). Thus, internal vapor was evacuated The deflection increased because the matrix had more pores due to
more easily, and the area under the load-deflection curves in- the melting of the fibers, which, in turn, led to higher ductility
creased. Due to the reproducibility of results, only the load- (Fig. 8) (Cifuentes et al. 2012). Both effects (i.e., a light decrease
deflection curves for D0 tests have been shown. At 500°C, the in maximum load and an increase in deflection) led to a higher
fracture energy was still high for the same reasons that affected fracture energy [Fig. 6(d)]. At 500°C and 700°C, the trend of the
D0. At 700°C, in addition to the decarbonation process, the PP load-deflection curves was the same as at 300°C (Fig. 9). The main
fibers were completely transformed into gaseous water and CO2 , difference was that the fibers were completely melted at that tem-
and the pore level in the concrete matrix significantly increased perature, so the damage level was higher (Fig. 5).
to 33 μm (the diameter of the fibers), as Fig. 10 shows. The use Regarding the D1 cold test [Fig. 6(d)] in comparison with the
of longer fibers increased the ductility of the material, as inferred D1 hot test, the maximum load increased at 100°C, as two effects
from the higher fracture energy values in Fig. 6(c) when compared happened simultaneously: the effect of temperature on the cement
with Fig. 6(b). Moreover, the longer fibers generated a greater net- paste, improving its hydration (Cifuentes et al. 2012; Zhang and
work of channels through which vapor could be more efficiently Bicanic 2006), and the reduction in internal pressure because there
discharged, reducing damage to the concrete and increasing the was more time for it to be released during the cooling process
fracture energy (Cifuentes et al. 2012). (Zhang and Bicanic 2006). As a result, the fracture energy was
Regarding the D0 cold test, the trend was similar to that of D0 higher in the D1 cold test [Fig. 6(b)]. At higher temperatures
hot tests, with a difference: additional microcracking occurred due (300°C, 500°C, and 700°C), the pore level and microcracking in the
to the cooling process, resulting in an earlier damage to the matrix cement paste increased, owing to the heating process and the addi-
[Fig. 6(d)] (Cifuentes et al. 2012; Zhang and Bicanic 2006). tional effect of the thermal gradient during cooling (Cifuentes et al.
Comparing the D1 and D0 cold tests, at 100°C the addition of 2012; Tufail et al. 2017; Zhang and Bicanic 2006). A similar effect
PP fibers increased the ductility of the material, as shown by the was observed by Tufail et al. (2017) through the compressive
growing deflection in Fig. 9. The D1 and D2 load-deflection curves stress-strain curves after high temperature exposition.
of cold tests have not been represented due to the reproducibility of The addition of longer fibers weakened the bearing capacity of
results. It is worth noting that they showed qualitatively a similar the material at room temperature and 100°C, as seen from the com-
trend with a more pronounced deflection. At 300°C, the maximum parison between D1 and D2 in hot and cold tests [Figs. 6(b–d),
load noticeably decreased as the PP fibers partially melted (Fig. 4). respectively]. At 300°C, when the fibers were partially melted,

© ASCE 04018271-7 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2018, 30(11): 04018271


Downloaded from ascelibrary.org by MISSOURI, UNIV OF/COLUMBIA on 08/03/18. Copyright ASCE. For personal use only; all rights reserved.

Fig. 8. Pore size distribution of D1 concrete.

behavior was similar to that reported in previous studies by Zhang


and Bicanic (2006). As regards D0 hot tests [Fig. 10(a)], when the
specimens were heated up to 100°C the compressive strength de-
creased, owing to vapor pressure in the concrete matrix (Cifuentes
et al. 2012; Zhang and Bicanic 2006). Between 100°C and 300°C,
the temperature affected the cement paste by improving its hydra-
tion, thus leading to an increase in the compressive strength. At
temperatures above 300°C, the matrix showed considerable micro-
cracking; as a result, the compressive strength decreased with the
increase in temperature.
In the D1 and D2 hot tests [Figs. 10(b and c)], in which the
specimens were reinforced with PP fibers (short and long fibers,
respectively), the results were qualitatively and quantitatively sim-
ilar to those obtained by the D0 [Fig. 10(a)]. The fibers did not
bring any improvement to the compressive mechanical properties
(Ramezanianpour et al. 2013). Moreover, the inclusion of fibers in
the matrix reduced the compressive strength at 700°C, when they
were completely melted.
Fig. 9. Load-deflection curves in D0 cold tests. Regarding the D0 cold tests [Fig. 10(d)], the evolution of the
results showed a decrease-increase-decrease trend, in keeping with
the findings of previous studies (Zhang and Bicanic 2006). At
100°C, the compressive strength was slightly greater than in the
the network of channels in the matrix was larger, which entailed a hot tests because this temperature did not generate high vapor pres-
noticeable reduction in spalling. As a result, the peak loads in the sure. At 300°C, the microcracking caused by the heating and cool-
load-deflection curves were higher. At 500°C and 700°C, when the ing processes prevented any increase in resistance, as was the case
fibers had completely melted, the damage level was similar in in the hot tests [Fig. 10(a)], leading to a decrease in the compressive
D1 and D2. strength (Nielsen and Bicanic 2003). At 500°C, the compressive
Exposure time to a given temperature had no influence on the strength decreased because the damage sustained during the heat-
fracture energy, because the core of the specimens reached the ing and cooling processes was more significant, due to the higher
target temperature in less than 6 h; hence, the thermal effects were temperature reached (Zhang and Bicanic 2006). At 700°C, the very
similar at the different exposure times (6, 24, and 48 h). high temperature caused the decomposition process of the cement
paste (Cifuentes et al. 2012; Zhang and Bicanic 2006) along with
microcracking during the cooling process.
Compressive Strength
As for the PP fiber-reinforced specimens, that is, D1 and D2
The compressive strength of all the mixes is shown in Fig. 10. The [Fig. 10(d)], the compressive strength was similar to that of D0
effects of temperature, the cooling process, and exposure time on at 100°C, although it decreased in some cases when the proportion
specimens with different PP fiber lengths in hot and cold tests can of fibers was very high, as shown in Fig. 10(d), due to the worm
be analyzed from these results. effect of fibers, leading to more pores in the concrete matrix
As shown in Figs. 10(a–c) (hot tests), a decrease-increase- (Ramezanianpour et al. 2013). At 300°C, the fibers were partially
decrease trend was observed in all three concrete types (D0, D1, melted, the fiber channels were not completely free, and vapor
and D2 mixes) at the three exposure times (6, 24, and 48 h). This discharge was consequently more difficult, generating spalling.

© ASCE 04018271-8 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2018, 30(11): 04018271


Downloaded from ascelibrary.org by MISSOURI, UNIV OF/COLUMBIA on 08/03/18. Copyright ASCE. For personal use only; all rights reserved.

Fig. 10. Evolution of compressive strength with temperature at each exposure time: (a) D0 hot tests; (b) D1 hot tests; (c) D2 hot tests; and (d) cold
tests.

Because of this, the concrete matrix sustained further damage, in The effect of the PP fibers was only slightly noticeable, but it is
comparison with room temperature and the same mixes in hot tests. worth noting that both the D2 mixes (reinforced with long fibers)
At 500°C and 700°C, the fibers were melted, and the behavior was and D1 mixes (reinforced with short fibers) achieved better results.
similar to that observed in the D0 cold tests. The reason for this is that the addition of PP fibers increases the
Exposure time to a given temperature had no influence on the ductility of the material, and this effect is more noticeable when
compressive strength, because the core of the specimens reached the PP fibers are longer (Gencel et al. 2011). At higher tempera-
the target temperature in less than 6 h; hence, the thermal effects tures, internal vapor pressure is reduced by means of the canal con-
were similar at the different exposure times (6, 24, and 48 h). nections generated by the melting PP fibers, so that vapor pressure
can be more efficiently evacuated (Cifuentes et al. 2012).
Young’s Modulus
The hot and cold tests showed a decreasing trend from 25°C to Tensile Strength
700°C, so rigidity always decreased. The decrease in Ec with tem- The trend observed in the hot tests [Figs. 12(a–c)] responds to a
perature in all mixes was due to the microcracking of the cement three-stage decrease-increase-decrease pattern, in accordance with
paste caused by the heating procedure [in hot tests, Figs. 13(a–c)] that proposed by Zhang and Bicanic (2006).
and the additional effect of the cooling process [in cold tests, The effect of the PP fibers was not very noticeable on the split-
Fig. 11(d)] (Cifuentes et al. 2012). Compared with the hot tests, ting tensile strength [Figs. 12(b and c)] in the tests conducted at
the Young’s modulus of the material at 25°C (Cifuentes et al. 25°C (Ramezanianpour et al. 2013). For all mixes in hot tests
2013; Gencel et al. 2011) was improved by adding PP fibers (D0, D1, and D2), the tensile strength decreased when the mixes
[Figs. 11(b and c)]. Nevertheless, there was a sharp drop of approx- reached 100°C, in keeping with the trend shown by the compressive
imately 50% at 100°C, which shows how the vapor pressure strength (Cifuentes et al. 2012; Zhang and Bicanic 2006). When the
quickly affected this property. However, from 100°C to 700°C temperature rose above 100°C, the tensile strength of concrete in-
the evolution of Ec was less marked and almost linear (Cifuentes creased because of the hydration effect caused by the temperature
et al. 2012; Tufail et al. 2017; Zhang and Bicanic 2006). As the on the cement paste. In fiber-reinforced mixes, a more noticeable
temperature rose, thermal microcracking took place, which entailed decrease took place between 500°C and 700°C due to dehydration,
a reduction in the rigidity of the concrete (Cifuentes et al. 2012). the decomposition of the cement paste (Cifuentes et al. 2012;
The effect of the fibers was barely noticeable at temperatures below Zhang and Bicanic 2006), and the melting of the fibers, which cre-
500°C. The highest values were obtained by the PP fiber-reinforced ated more voids in the concrete matrix than in the D0 mix (Bei and
concrete. Zhixiang 2016; Cifuentes et al. 2012; Sideris and Manita 2013).

© ASCE 04018271-9 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2018, 30(11): 04018271


Downloaded from ascelibrary.org by MISSOURI, UNIV OF/COLUMBIA on 08/03/18. Copyright ASCE. For personal use only; all rights reserved.

Fig. 11. Evolution of Young’s modulus with temperature at each exposure time: (a) D0 hot tests; (b) D1 hot tests; (c) D2 hot tests; and (d) cold tests.

Fig. 12. Evolution of splitting tensile strength with temperature at each exposure time: (a) D0 hot tests; (b) D1 hot tests; (c) D2 hot tests; and (d) cold tests.

© ASCE 04018271-10 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2018, 30(11): 04018271


Downloaded from ascelibrary.org by MISSOURI, UNIV OF/COLUMBIA on 08/03/18. Copyright ASCE. For personal use only; all rights reserved.

Fig. 13. Evolution of the characteristic length with temperature for each exposure time: (a) D0 hot tests; (b) D1 hot tests; (c) D2 hot tests; and
(d) cold tests.

The evolution of the results in the cold tests [Fig. 12(d)] showed higher characteristic length will also have a greater FPZ length
a decreasing trend, as in previous studies (Cifuentes et al. 2012; and thus a higher ductility (Bazant and Planas 1998).
Zhang and Bicanic 2006). The splitting tensile strength results Because all specimens were of the same size, there was no size
at 100°C were better than in the hot tests because vapor pressure effect, and it was consequently possible to directly analyze the
had more time to be evacuated during the cooling process. At ductility of the mixes by means of the characteristic length values.
temperatures above 100°C, the concrete matrix, which was already The characteristic length is defined as
damaged by the previous heating process, showed a further increase
Ec G F
in internal damage due to the cooling process, which generated lch ¼ ð6Þ
more microcracks and propagated already existing ones (Zhang f2t
and Bicanic 2006).
where ft = tensile strength and was, in this case, considered equal
The effect of the PP fibers on the splitting tensile strength was
to the splitting tensile strength fti .
more noticeable than their effect on the residual compressive
The hot tests [Figs. 13(a–c)] showed a decrease-increase trend,
strength, but the behavior was similar. The addition of PP fibers in which the lowest point was reached at 300°C. If the temperature
to the concrete matrix created a worm effect that increased the num- rose to 300°C, this affected the cement paste, improving its hydra-
ber of pores. This effect was more remarkable when the amount of tion and leading to an increase in the splitting tensile strength
fibers and their length increased. When concrete mixes reached [Figs. 12(a–c)] (Cifuentes et al. 2012; Zhang and Bicanic 2006)
temperatures above 300°C, the fibers were partially or completely and a decrease in the characteristic length. From 500°C to 700°C,
melted, and the proportion of voids in the concrete matrix signifi- the porosity level increased significantly, softening the material and
cantly rose. As a result, the splitting tensile strength decreased. This increasing its ductility. Postpeak softening significantly increased,
effect was more remarkable in mixes reinforced with longer fibers as shown in Figs. 7(c and d), which led to higher characteristic
[D2, Fig. 12(d)]. length and FPZ values. In the PP fiber-reinforced hot mixes
[Figs. 13(b and c)], the characteristic length was higher than in non-
reinforced specimens [Fig. 13(a)], as this increased when the PP
Characteristic Length
fibers were partially or completely melted, resulting in a softer con-
Characteristic length, lch , was assessed in both the hot and cold crete matrix with higher ductility. However, this effect became
mixes. This fracture parameter provides information about the in- more noticeable at temperatures above 500°C, because the porosity
trinsic brittleness of the cohesive material (Hillerborg et al. 1976; level was higher (Fig. 8).
Rosselló et al. 2006). It is also directly related to the fracture In the cold tests [Fig. 13(d)], at 100°C the trend was similar to
process zone (FPZ), although it does not indicate its exact length that of the hot tests, but the lowest characteristic length values were
[ACI Committee 446 (ACI 1991)]. Therefore, a material with reached because the damage was sustained during the cooling

© ASCE 04018271-11 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2018, 30(11): 04018271


process. At 300°C, the matrix of cold mixes was more damaged concrete to be used as thermal energy storage in solar thermal elec-
than that of hot tests at the same temperature. This led to a more tricity plants.” Cem. Concr. Res. 82 (Apr): 74–86. https://doi.org/10
ductile material with a higher characteristic length. At temperatures .1016/j.cemconres.2015.12.013.
above 300°C, the level of damage in the matrix increased (Fig. 8), Alotto, P., M. Guarnieri, and F. Moro. 2014. “Redox flow batteries for the
storage of renewable energy: A review.” Renewable Sustainable Energy
and the PP fibers melted. The addition of fibers increased ductility
Rev. 29 (Jan): 325–335. https://doi.org/10.1016/j.rser.2013.08.001.
at any temperature, as in the hot tests; this effect was more notice-
Bamonte, P., and G. Pietro. 2016. “High-temperature behavior of SCC in
able when longer fibers were used. compression: Comparative study on recent experimental campaigns.”
J. Mater. Civ. Eng. 28 (3): 4015141. https://doi.org/10.1061/(ASCE)
MT.1943-5533.0001378.
Conclusions Barros, J., E. Pereira, and S. Santos. 2007. “Lightweight panels of steel
fiber-reinforced self-compacting concrete.” J. Mater. Civ. Eng. 19 (4):
In this study, we explored the effects of temperature on the fracture 295–304. https://doi.org/10.1061/(ASCE)0899-1561(2007)19:4(295).
Downloaded from ascelibrary.org by MISSOURI, UNIV OF/COLUMBIA on 08/03/18. Copyright ASCE. For personal use only; all rights reserved.

behavior and other mechanical properties of high-strength polypro- Bazant, Z. P., and J. Planas. 1998. Fracture and size effect in concrete and
pylene fiber-reinforced self-compacting concrete, at a range of tem- other quasi brittle materials. Boca Raton, FL: CRC Press.
peratures up to 700°C and after cooling at three different exposure Bei, S., and L. Zhixiang. 2016. “Investigation on spalling resistance of
times. The following conclusions are drawn from the results: ultra-high-strength concrete under rapid heating and rapid cooling.”
• The irreversible process of weight loss increases with higher Case Stud. Constr. Mater. 4 (Jun): 146–153. https://doi.org/10.1016/j
temperatures. This process involves dehydration of residual .cscm.2016.04.001.
moisture at low temperatures, dehydration of chemically bound CEN (European Committee for Standardization). 2008. Cement. Part 1:
Composition, specifications and conformity creiteria from common
water at medium-high temperatures, and decarbonation of cal-
cements, AENOR. EN197-1:2000/A3. Brussels, Belgium: CEN.
cium carbonate at higher temperatures.
CEN (European Committee for Standardization). 2009a. Testing hardened
• The addition of fibers creates a network of channels in the ma- concrete. Part 3: Compressive strength of test specimens, AENOR.
trix that reduces internal pressure damage and spalling effects. EN12390-3. Brussels, Belgium: CEN.
However, it also reduces material strength at room temperature CEN (European Committee for Standardization). 2009b. Testing hardened
due to the increased number of pores created by the trapped air. concrete. Part 6: Tensile splitting strength of test specimens, AENOR.
• As long as they are not degraded, the fibers have a bridge effect EN12390-6. Brussels, Belgium: CEN.
on the crack front that stitches any eventual microcracking. This CEN (European Committee for Standardization). 2014. Testing hardened
improves the strength properties and ductility of the concrete in concrete. Part 13: Determination of secant modulus of elasticity in
both hot and cold conditions. compression, AENOR. EN12390-13. Brussels, Belgium: CEN.
• Longer fibers reduce the effects of spalling and create a greater Chauhan, A., and R. P. Saini. 2014. “A review on integrated renewable
network of channels for the evaporation of internal pressure. energy system based power generation for stand-alone applications:
Configurations, storage options, sizing methodologies and control.”
Yet, they also increase the number of voids in the matrix, which
Renewable Sustainable Energy Rev. 38 (Oct): 99–120. https://doi.org
in turn reduces its mechanical properties and causes a softening /10.1016/j.rser.2014.05.079.
of the material. Cifuentes, H., F. García, O. Maeso, and F. Medina. 2013. “Influence of the
• The cooling process generates microcracking due to the thermal properties of polypropylene fibres on the fracture behaviour of low-,
gradient that is more noticeable from medium-high tempera- normal- and high-strength FRC.” Constr. Build. Mater. 45 (Aug):
tures, where this thermal gradient is more abrupt, and the fibers 130–137. https://doi.org/10.1016/j.conbuildmat.2013.03.098.
are partially melted. These harmful effects add to those sus- Cifuentes, H., C. Leiva, F. Medina, and C. Fernández-Pereira. 2012.
tained during the heating process. “Effects of fibers and rice husk ash on properties of heated high-strength
• When the thermal gradient is more gradual, the cooling process concrete.” Mag. Concr. Res. 64 (5): 457–470. https://doi.org/10.1680
does not cause significant damage, and the fibers help to stitch /macr.11.00087.
any microcracking. This leads to higher mechanical properties. Cifuentes, H., M. Lozano, T. Holusova, F. Medina, S. Seitl, and A.
Fernandez-Canteli. 2017. “Modified disk-shaped compact tension test
• Exposure time has no effect when the temperature is uniformly
for measuring concrete fracture properties.” Int. J. Concr. Struct. Mater.
reached in the entire matrix. From that point, the behavior of 11 (2): 215–228. https://doi.org/10.1007/s40069-017-0189-4.
concrete is the same regardless of the length of time the speci- EFNARC (European Federation of National Associations Representing
mens are exposed to the temperature. Producers and Applicators of Specialist Building Products for Con-
crete). 2005. The European guidelines for self compacting concrete,
63. Farnham, UK: EFNARC.
Acknowledgments Elices, M., G. V Guinea, and J. Planas. 1992. “Measurement of the fracture
energy using three-point bend tests. Part 3: Influence of cutting the
The authors acknowledge the financial support provided to this P-d tail.” Mater. Struct. 25 (6): 327–334. https://doi.org/10.1007
study by the Spanish Ministry of Economy and Competitiveness /BF02472591.
under project BIA2016-75431-R. Fenollera, M., J. L. Míguez, I. Goicoechea, J. Lorenzo, and M. Á. Álvarez.
2013. “The influence of phase change materials on the properties of
self-compacting concrete.” Materials 6 (8): 3530–3546. https://doi.org
References /10.3390/ma6083530.
Gencel, O., C. Ozel, W. Brostow, and G. Martínez-Barrera. 2011.
ACI (American Concrete Institute). 1991. Fracture mechanics of concrete: “Mechanical properties of self-compacting concrete reinforced with
Concepts, models and determination of material properties. ACI polypropylene fibres.” Mater. Res. Innovations 15 (3): 216–225. https://
Committee 446. Detroit: ACI. doi.org/10.1179/143307511X13018917925900.
Alarcon-Ruiz, L., G. Platret, E. Massieu, and A. Ehrlacher. 2005. “The use Guan, J., X. Hu, and Q. Li. 2016. “In-depth analysis of notched 3-p-b
of thermal analysis in assessing the effect of temperature on a cement concrete fracture.” Eng. Fract. Mech. 165 (Oct): 57–71. https://doi
paste.” Cem. Concr. Res. 35 (3): 609–613. https://doi.org/10.1016/j .org/10.1016/j.engfracmech.2016.08.020.
.cemconres.2004.06.015. Guan, J., X. Hu, X. Yao, Q. Wang, Q. Li, and Z. Wu. 2017. “Fracture of
Alonso, M. C., J. Vera-Agullo, L. Guerreiro, V. Flor-Laguna, M. Sanchez, 0.1 and 2 m long mortar beams under three-point-bending.” Mater. Des.
and M. Collares-Pereira. 2016. “Calcium aluminate based cement for 133 (Nov): 363–375. https://doi.org/10.1016/j.matdes.2017.08.005.

© ASCE 04018271-12 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2018, 30(11): 04018271


Guinea, G. V., J. Planas, and M. Elices. 1992. “Measurement of the fracture Ramezanianpour, A. A., M. Esmaeili, S. A. Ghahari, and M. H. Najafi.
energy using three-point bend tests. Part 1: Influence of experimental 2013. “Laboratory study on the effect of polypropylene fiber on
procedures.” Mater. Struct. 25 (4): 212–218. https://doi.org/10.1007 durability, and physical and mechanical characteristic of concrete for
/BF02473065. application in sleepers.” Constr. Build. Mater. 44 (Jul): 411–418.
Herrmann, U., B. Kelly, and H. Price. 2004. “Two-tank molten salt storage https://doi.org/10.1016/j.conbuildmat.2013.02.076.
for parabolic trough solar power plants.” Energy 29 (5–6): 883–893. RILEM (Réunion Internationale des Laboratoires et Experts des Matériaux,
https://doi.org/10.1016/S0360-5442(03)00193-2. systèmes de construction et ouvrages). 1985. “TCM-85: Determination
Hillerborg, A., M. Modéer, and P. E. Petersson. 1976. “Analysis of crack of the fracture energy of mortar and concrete by means of three-point
formation and crack growth in concrete by means of fracture mechanics bend tests on notched beams.” Mater. Struct. 18 (106): 287–290. https://
and finite elements.” Cem. Concr. Res. 6 (6): 773–781. https://doi.org doi.org/10.1007/BF02498757.
/10.1016/0008-8846(76)90007-7. Rosselló, C., M. Elices, and G. V. Guinea. 2006. “Fracture of model concrete:
Hu, X., J. Guan, Y. Wang, A. Keating, and S. Yang. 2017. “Comparison Part 2. Fracture energy and characteristic length.” Cem. Concr. Res.
of boundary and size effect models based on new developments.” 36 (7): 1345–1353. https://doi.org/10.1016/j.cemconres.2005.04.016.
Downloaded from ascelibrary.org by MISSOURI, UNIV OF/COLUMBIA on 08/03/18. Copyright ASCE. For personal use only; all rights reserved.

Eng. Fract. Mech. 175 (Apr): 146–167. https://doi.org/10.1016/j Shah S. P.. 1990. “Determination of fracture parameters (KIc-s and
.engfracmech.2017.02.005. CTODc) of plain concrete using three-point bend tests.” Mater. Struct.
Kang, J., H. Yoon, W. Kim, V. Kodur, Y. Shin, and H. Kim. 2016. “Effect of 23 (6): 457–460.
wall thickness on thermal behaviors of RC walls under fire conditions.” Sideris, K. K., and P. Manita. 2013. “Residual mechanical characteristics
Int. J. Concr. Struct. Mater. 10 (S3): 19–31. https://doi.org/10.1007 and spalling resistance of fiber reinforced self-compacting concretes
/s40069-016-0164-5. exposed to elevated temperatures.” Constr. Build. Mater. 41 (Apr):
Karihaloo, B. L., H. M. Abdalla, and T. Imjai. 2003. “A simple method for 296–302. https://doi.org/10.1016/j.conbuildmat.2012.11.093.
determining the true fracture energy of concrete.” Mag. Concr. Res. Srikar, G., G. Anand, and S. Suriya Prakash. 2016. “A study on residual
55 (5): 471–481. https://doi.org/10.1680/macr.2003.55.5.471. compression behavior of structural fiber reinforced concrete exposed to
Kodur, V. 2014. “Properties of concrete at elevated temperatures.” ISRN moderate temperature using digital image correlation.” Int. J. Concr.
Civ. Eng. 2014: 1–15. https://doi.org/10.1155/2014/468510. Struct. Mater. 10 (1): 75–85. https://doi.org/10.1007/s40069-016
Laing, D., D. Lehmann, M. Fiß, and C. Bahl. 2009. “Test results of concrete -0127-x.
thermal energy storage for parabolic trough power plants.” J. Sol. Tao, J., Y. Yuan, and L. Taerwe. 2010. “Compressive strength of self-
Energy Eng. 131 (4): 041007. https://doi.org/10.1115/1.3197844. compacting concrete during high-temperature exposure.” J. Mater.
Liu, X., G. Ye, G. Deschutter, Y. Yuan, and L. Taerwe. 2008. “On the Civ. Eng. 22 (10): 1005–1011. https://doi.org/10.1061/(ASCE)MT
mechanism of polypropylene fibres in preventing fire spalling in self- .1943-5533.0000102.
compacting and high-performance cement paste.” Cem. Concr. Res. Tufail, M., K. Shahzada, B. Gencturk, and J. Wei. 2017. “Effect of elevated
38 (4): 487–499. https://doi.org/10.1016/j.cemconres.2007.11.010. temperature on mechanical properties of limestone, quartzite and gran-
Memon, S. A., H. Z. Cui, H. Zhang, and F. Xing. 2015. “Utilization of ite concrete.” Int. J. Concr. Struct. Mater. 11 (1): 17–28. https://doi.org
macro encapsulated phase change materials for the development of /10.1007/s40069-016-0175-2.
thermal energy storage and structural lightweight aggregate concrete.” Vilches, L. F., C. Leiva, J. Vale, and C. Fernánndez-Pereira. 2005. “Insu-
Appl. Energy 139 (Feb): 43–55. https://doi.org/10.1016/j.apenergy lating capacity of fly ash pastes used for passive protection against fire.”
.2014.11.022. Cem. Concr. Compos. 27 (7–8): 776–781. https://doi.org/10.1016/j
Mohamedbhai, G. T. G. 1986. “Effect of exposure time and rates of heating .cemconcomp.2005.03.001.
and cooling on residual strength of heated concrete.” Mag. Concr. Res. Wang, Y., X. Hu, L. Liang, and W. Zhu. 2016. “Determination of tensile
38 (136): 151–158. https://doi.org/10.1680/macr.1986.38.136.151. strength and fracture toughness of concrete using notched 3-p-b
Nielsen, C. V., and N. Bicanic. 2003. “Residual fracture energy of high- specimens.” Eng. Fract. Mech. 160: 67–77. https://doi.org/10.1016/j
performance and normal concrete subject to high temperatures.” Mater. .engfracmech.2016.03.036.
Struct. 36 (262): 515–521. https://doi.org/10.1617/13880. Yang, Z., and S. V. Garimella. 2010. “Thermal analysis of solar thermal
Noumowé, A., H. Carré, A. Daoud, and H. Toutanji. 2006. “High-strength energy storage in a molten-salt thermocline.” Solar Energy 84 (6):
self-compacting concrete exposed to fire test.” J. Mater. Civ. Eng. 974–985. https://doi.org/10.1016/j.solener.2010.03.007.
18 (6): 754–758. https://doi.org/10.1061/(ASCE)0899-1561(2006)18: Zhang, B., and N. Bicanic. 2006. “Fracture energy of high-performance
6(754). concrete at high temperatures up to 4508C: the effects of heating
Pielichowska, K., and K. Pielichowski. 2014. “Phase change materials for temperatures and testing conditions (hot and cold).” Mag. Concr.
thermal energy storage.” Prog. Mater Sci. 65 (Aug): 67–123. https://doi Res. 58 (5): 277–288. https://doi.org/10.1680/macr.2006.58.5.277.
.org/10.1016/j.pmatsci.2014.03.005. Zoalfakar, S. H., M. A. Elsissy, Y. B. Shaheen, and A. A. Hamada. 2016.
Planas, J. 2007. Experimental determination of the stress-crack opening “Multiresponse optimization of postfire residual properties of fiber-
curve for concrete in tension. RILEM Rep. No. 39. Bagneux, France: reinforced high-performance concrete.” J. Mater. Civ. Eng. 28 (10):
RILEM. 4016111. https://doi.org/10.1061/(ASCE)MT.1943-5533.0001622.

© ASCE 04018271-13 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2018, 30(11): 04018271

You might also like