Next Article in Journal
An In Vivo Proposal of Cell Computing Inspired by Membrane Computing
Next Article in Special Issue
Effect of Alumina Additives on Mechanical and Fresh Properties of Self-Compacting Concrete: A Review
Previous Article in Journal
The Prospect of Microwave Heating: Towards a Faster and Deeper Crack Healing in Asphalt Pavement
Previous Article in Special Issue
Processes and Properties of Ionic Liquid-Modified Nanofiller/Polymer Nanocomposites—A Succinct Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Experimental Study on the Effect of Basalt Fiber and Sodium Alginate in Polymer Concrete Exposed to Elevated Temperature

by
Seyed Esmaeil Mohammadyan-Yasouj
1,*,
Hossein Abbastabar Ahangar
2,
Narges Ahevani Oskoei
1,
Hoofar Shokravi
3,
Seyed Saeid Rahimian Koloor
4 and
Michal Petrů
5
1
Department of Civil Engineering, Najafabad Branch, Islamic Azad University, Najafabad 8514143131, Iran
2
Department of Chemistry, Najafabad Branch, Islamic Azad University, Najafabad 8514143131, Iran
3
Department of Structures and Materials, School of Civil Engineering, Faculty of Engineering, Universiti Teknologi Malaysia, Skudai 81310, Johor, Malaysia
4
Institute for Nanomaterials, Advanced Technologies and Innovation (CXI), Technical University of Liberec (TUL), Studentska 2, 461 17 Liberec, Czech Republic
5
Technical University of Liberec (TUL), Studentska 2, 461 17 Liberec, Czech Republic
*
Author to whom correspondence should be addressed.
Processes 2021, 9(3), 510; https://doi.org/10.3390/pr9030510
Submission received: 11 February 2021 / Revised: 1 March 2021 / Accepted: 8 March 2021 / Published: 11 March 2021

Abstract

:
Polymer concrete contains aggregates and a polymeric binder such as epoxy, polyester, vinyl ester, or normal epoxy mixture. Since polymer binders in polymer concrete are made of organic materials, they have a very low heat and fire resistance compared to minerals. This paper investigates the effect of basalt fibers (BF) and alginate on the compressive strength of polymer concrete. An extensive literature review was completed, then two experimental phases including the preliminary phase to set the appropriate mix design, and the main phase to investigate the compressive strength of samples after exposure to elevated temperatures of 100 °C, 150 °C, and 180 °C were conducted. The addition of BF and/or alginate decreases concrete compressive strength under room temperature, but the addition of BF and alginate each alone leads to compressive strength increase during exposure to heat and increase in the temperature to 180 °C showed almost positive on the compressive strength. The addition of BF and alginate both together increases the rate of strength growth of polymer concrete under heat from 100 °C to 180 °C. In conclusion, BF and alginate decrease the compressive strength of polymer concretes under room temperature, but they improve the resistance against raised temperatures.

1. Introduction

Concrete is an old building material, in use for centuries. Concrete is widely used in the construction of civil engineering structures such as buildings [1,2], bridges [3,4], and other civil structures [5,6]. Reinforced concrete structures are vulnerable to high-temperature conditions, which greatly shortens the service life and hinders its applications [7]. The properties of the mix design parameters have a significant effect on the performance of concrete exposed to elevated temperature [8,9,10,11]. Several research studies have been carried out to improve the thermal stability of concrete exposed to elevated temperatures using various materials, such as fillers, nanoparticles, and fibers [12,13,14,15].
Sodium alginate (SA) is a well-known natural polymer that is extracted from cell walls of brown algae [16,17]. Recently, SA has received much attention in concrete applications due to self-healing behavior [18,19]. The influence of the SA on mechanical and flexural properties is still unclear and different studies achieved different results. Abbas and Mohsen [20] indicated that the addition of SA increased the fresh and mechanical properties of concrete. However, Heidari et al. [21] reported a decrease in mechanical properties by using SA in the mix design. Mignon et al. [22] indicated that the addition of 1% calcium alginate can result in a 15% reduction in compression strength of concrete while it was 28% for SA. SA also improves the temperature resistance of concrete and enhances flame-, fire- and heat-resistance of cementitious composites [23]. Moreover, it was stated that the properties of the produced alginate gel are greatly affected by the polymer concentration and type [16,19]. Mignon et al. [22] showed that alginate can improve the internal curing of concrete effectively. Ouwerx et al. [19] showed that the concentration of SA in concrete can limit the elasticity of beads. Pathak et al. [16] prepared and compared different metal alginates and alginic acid exploited from fresh algae using the extraction method. It was reported that decomposition of cobalt alginate takes place at a higher temperature compared to sodium and calcium alginate. The distribution and pore size were significantly influenced by the presence of metal ions. The type of metal alginates may lead to variation in thermal behavior of the compounds. Table 1 shows a summary of the studies on the effect of alginates on mechanical parameters of concrete.
Table 1. A summary of the studies on the effect of alginates on mechanical parameters of concrete.
Table 1. A summary of the studies on the effect of alginates on mechanical parameters of concrete.
Reference.Alginate TypeWeight Fraction
(% of Cement)
Remarks
Mohammadyan-Yasouj et al. [24]NaAlg0.1
  • Alginate decreases the flexural, tensile, and compressive strengths of concrete.
Ouwerx et al. [19]Alginate gel bead-
  • Properties of alginate significantly depend on polymer type and concentration.
Pathak et al. [16]Alginic acid and metal alginates-
  • Presence of metal ions impacted on the pore size distribution of alginates.
  • Different thermal behavior was observed in different metal alginates due to structural differences.
Heidari et al. [21]Alginic acid0.5 and 1
  • Alginate decreases the flexural, tensile, and compressive strengths of concrete.
  • Alginate diminished the fresh properties of SCC.
Mignon et al. [22]NaAlg, CaAlg0.5 and 1
  • Alginate improves internal curing
  • CaAlg beads of 1% reduced 15% of the compressive strength.
  • NaAlg of 1% resulted in a 28% decrease in compressive strength due to increasing in the water uptake.
Polymeric concrete is a kind of concrete in which natural aggregates such as silica sand and gravel are bound together in a matrix with a polymer binder as a supplement or replacement of cement in concrete [25]. Polymeric concrete has higher mechanical strengths, chemical resistance, and ductility, compared with ordinary concrete (OC) [26,27]. Currently, three distinct types of polymeric concrete are studied that include polymer-Portland cement concrete (PPCC), polymer impregnated concrete (PIC), and polymer concrete (PC) [27].
According to American Concrete Institute (ACI 548.3R), in PC, the polymer is served as the sole binding material present in concrete [28]. However, filler materials such as silica fume and fly ash can be utilized along with aggregates to enhance mechanical properties and minimize building cost. PC can achieve about 80% of its 28 day compressive strength in one day, and compressive strength of 100 MPa can be attained [26,29]. Higher ductility after 60 °C [30] and no visible failure under compression for mixes containing epoxy or epoxy and filler of 30% of the total volume is also reported for PC [31]. An investigation by Golestaneh et al. [32] indicated an optimal amount of epoxy to reach the maximum compressive and flexural strength was 15% of the total volume, while Elalaoui et al. [33] suggested 13% to reach the highest physical and mechanical properties at the lowest cost. However, the higher value for the suggested optimal amount by Golestaneh et al. [32], even with 20% epoxy, may be due to the utilization of silica fume as filler to get the maximum tensile strength.
The results indicate that, compared to OC, polymeric concrete exhibited fewer cracks that are probably due to the bound effect between fiber and matrix [34]. PC has higher strengths than OC [34,35,36] and, similar to the OC, with the addition of fiber, reduction in its workability is unavoidable [34]. Acceptable interface adhesion between fiber and polymeric matrix for fiber loading of 15% of the total volume resulted in high tensile and flexural strengths [37]. Understanding the behavior of concrete exposed to various temperatures is considered an important factor in meeting the safety and service life objectives in which structures are intended and designed. Three different temperature ranges are generally used in concrete studies that include low temperature (<0 °C), medium temperature (0–50 °C), and high or elevated temperature (<50 °C). The thermal response of concrete to elevated temperature mainly depends on constituents’ characteristics and the mix composition [38]. Based on the applications and intended exposure condition, previous researches have considered various temperature ranges to evaluate the thermal response of concrete which can be classified as temperatures below 200 °C [39,40,41], below 600 °C [42,43,44], and below 1000 °C [45,46,47]. The value of 600 °C is generally selected because the interior of the concrete members will not exceed 600 °C in a short time. Higher temperatures (e.g., 1000 °C) are considered for the peak gas temperature of a fire that may reach a higher value [48], while the temperature lower than 200 °C is generally considered for members exposed to indirect fire heat.
The duration of exposure and heating rate are two other important factors that have been investigated in many research studies. Different heating durations are generally based on the recommendation by different specifications for maximum fire-resistance rating. For example, the fire resistance for column members in China fire design specifications is defined as 180 min (3 h) [48]. Alharbi et al. [49] indicated that duration of exposure has an effect on degradation of the stiffness while prolonged exposures can significantly reduce the strength of concrete. The heating speed caused by exposure to real fire is generally specified by fire standards (e.g., ISO 834 standard fire).
Investigation on the behavior of PC with basalt fiber has revealed that BF >1% reduces concrete strength [50]. On the other hand, the studies on epoxy/basalt polymer concrete showed enhancement of mechanical properties and more ductility under increasing temperature up to 100 °C [36]. Reis [30] observed that the epoxy mortars had higher sensitivity to temperature variations due to the heat distortion temperature of the resins [19]. Niaki et al. [51] demonstrated that basalt fiber improved the mechanical properties and increased the thermal stability of the PC subjected to different temperatures (up to 250 °C). Several research studies have been conducted to improve the durability of polymer concrete exposed to elevated temperatures [9]. Elalaoui et al. [52] proposed ammonium polyphosphate to be used as a constituent of the PC to improve the temperature resistance of the concrete. Niaki et al. [36] reported that adding crushed basalt aggregates into PC led to decreases in the highest maximum stress rapidly, but increases the yield displacement significantly due to temperature increase. Gorninski et al. [53] produced PC composites using waste alumina and showed that the addition of alumina flame retardant waste can efficiently improve the temperature resistance of the PC.
The effect of using BF on concrete has been carried out by many researchers, and several experiments were conducted to study the influence of adding different proportions, diameter, and length of fibers in concrete. Jiang et al. [54] demonstrated that BF addition can considerably improve the tensile and flexural strength and toughness index of fiber reinforce concrete. It was indicated that adding BF had no obvious increase in compressive strength. Compared to other types of fibers such as polypropylene and glass fiber (GF) higher performance for splitting tensile, flexural strengths, crack resistance, ductility post cracking flexural response have been reported for BF [54,55,56,57]. Almost all previous researchers are agreed upon the fact that an increase in the fiber content and/or length decreases the workability of concrete and increases porosity [58]. Kabay [59] showed that increasing the length and/or volume of fibers reduces concrete workability. It was observed that void content has a higher impact on abrasion compared to compressive and flexural strength. The inclusion of fibers reduced compression strength and abrasive wear of concrete. The highest flexure strength was achieved in mixtures with 60% w/c. Dias and Thaumaturgo [34] studied the effect of BF in geopolymer concrete. It was indicated that BF in geopolymer concretes improved the flexural and splitting tensile strength compared to geopolymer concretes; however, the addition of 1.0% BF resulted in a 26.4% reduction in compressive strengths of cement concretes.
The length and weight fraction of BF are two interactive factors that influence on mechanical and fresh properties of cementitious composites. Khan and Cao [60] investigated the mechanical properties of basalt-fiber-reinforced cementitious composites with four different fiber lengths (3, 6, 12, and 20 mm) and also with a combination of various lengths and contents. The results indicated that the compressive strength of basalt-fiber-reinforced cementitious composites containing fiber with the length of 6 and 12 mm is higher than that with 3 and 20 mm length. It was observed that short fibers in the matrix play a bridging role to limit the expansion of microcracks while longer fibers provided additional capacity to restrict the development of macrocracks. Amuthakkannan et al. [61] studied the effect of basalt fiber length and content on the mechanical properties of basalt-fiber-reinforced polymer matrix composites of the fabricated composites. It was shown that basalt fiber with the length of 10 mm and 50 mm provided better tensile strength than 4 mm and 21 mm in polymer matrix composites. It was indicated that shorter fiber lengths will create more fiber ends that act as stress concentration points, resulting in a faster failure at these points.
Almost all studies conducted on properties of fiber-reinforced cementitious materials indicated that increasing the fiber contents and/or length harms the workability and porosity of the fresh mix [58]. Reduction in the workability due to an increase in the fiber content requires higher water-to-cement ratio, which leads to higher porosity in the hardened concrete. Accordingly, fiber content is also a main parameter considered by previous researchers to control the fiber effect on the concrete properties. However, selecting an appropriate amount of fiber content to reach proper workability with expected mechanical properties for concrete is necessary. As shown in Table 2, the length of BF is varied from 3 to 30 mm while the fiber content is ranged from 0.05% to 5% of the volume fraction in the mixes. As it can be seen in the table, 2% is the most frequently used weight fraction in studies related to fresh and hardened concrete properties. Chopped fibers with shorter length are characterized with a more uniform distribution while longer lengths can better contribute with crack control.
Table 2. A summary of the studies on the effect of basalt fiber (BF) on mechanical parameters of concrete.
Table 2. A summary of the studies on the effect of basalt fiber (BF) on mechanical parameters of concrete.
Reference.Fiber TypeLength (mm)Weight Fraction (%)Remarks
Jiang et al. [54]Basalt
Polypropylene
12 and 22
4–19
0.05, 0.1, 0.3, & 0.5
-
Compressive strength ↑
-
Toughness & splitting tensile strength ↑
-
Flexural strength (0.3% BF) and (0.5% BF)
-
Tensile and flexure strength (22 mm length of BF)
Kabay [59]Basalt12 and 240.07 and 0.14
-
Length and/or volume of BF ↑ ⟹ workability ↓
-
Compression strength ↓
-
Abrasive wear (in the range of 2–18%) ↓
-
Flexure strength ↑
Ayoub et al. [62]Basalt251, 2, and 3
-
BF contents ↑ ⟹ compressive strength ↓
-
Tension strength ↑
-
No correlation between BF and E-value
Kizilkanat et al. [57]Basalt
Glass
12
12
0.25, 0.5, 0.75, and 1
-
BF contents ↑ ⟹ fracture energy ↑
-
Compressive strength ↑ (>0.25% fiber content)
-
Crack avoidance, tensile strength, and ductility ↑
-
GF ↑ ⟹ flexure strength
Fenu et al. [56]Basalt
Glass
12
12
3 and 5
-
No significant impact on dynamic in-crease factor
-
Energy absorption ↑
-
Static flexural strength and post-peak behavior ↑
Shafiqh et al. [55]Basalt
Polyvinyl Alcohol
25
30
1, 2, and 3
-
Strain attaining capacity ↑
-
PVA fibers ↑ ⟹ ductility, post-cracking flexural response, and toughness ↑
-
BF does not correlate with post-peak flex-ural behavior
Girgin [63]Basalt
Glass
24
24
2
-
Fibers ↑ ⟹ fracture energy ↑; workability ↓ wand E-value ↓
-
Compressive strength ↑ (0.5% and 0.75% fiber content)
-
Crack avoidance ↑, tensile strength ↑, and ductility ↑
Afroz et al. [9]Basalt (Chinese)
Basalt (Russian)
3
25
0.5
-
Stability in alkaline medium ↑
-
Protection against alkaline ion
-
Mechanical properties ↓
-
Long-term flexural strengths and splitting tensile ↑
Zhao et al. [64]Basalt181, 1.5, 2 & 2.5
-
Anti-impact deformation characteristics ↑
Katkhuda and Shatarat [65]Basalt180.1, 0.3, 0.5, 1, and 1.5
-
Flexural and splitting tensile compressive strength ↑
Sun et al. [66]Basalt6 and 121, 2, 3, 4, & 5
-
Compressive and splitting tensile strengths ↑; bending strength increase ↓
Ayoub et al. [62] stated that an increase in BF contents of the concrete mix leads to reduced compressive strength. However, 0.5% and 0.75% of BF content can slightly improve the compressive strength. The addition of 10% silica and BF can result in a higher tensile strength in high-performance concrete. BF content does not correlate with E-value. Fenu et al. [56] reported that the addition of GF and BF increased dynamic energy absorption at a high strain rate. Furthermore, it enhanced the post-peak behavior and static flexural strength of concrete. Shafiqh et al. [55] indicated that the addition of GF and BF enhanced strain capacity attaining of the concrete. Significant improvement was achieved in flexural response, ductility, and toughness by the inclusion of PVA fibers. However, no relation was found between BF addition and post-peak flexural behavior of concrete. Girgin 2016 [63] stated that the addition of GF resulted in a 35% reduction in the average strain capacity of concrete compared to the control specimens. The flexural strain capacity of BF was influenced by cement hydration in the matrix–fiber interface. Afroz et al. [67] studied the influence of BF fibers for sulfate and chloride resistance in concrete for long-term durability applications. It was shown that modified BF had better stability in the alkaline medium compared to non-modified ones. However, a slight reduction in the mechanical properties was reported for concretes reinforced with treated fibers.
Since polymer binders in polymer concrete are made of organic materials, they have a very low heat and fire resistance compared to minerals. Hence improving the thermal performance of the polymer concretes is of utmost importance for the construction industry. The limited information that exists in the literature indicates that application of SA in the concrete mixture could associate with self-healing of cracks and improving the fire, flame, and heat resistance of concrete while the inclusion of BF could enhance the energy absorption and ductile behavior of concrete that can prevent brittle fracture of structures. Several pieces of research have been conducted to study the effect of using SA and BF in cement concrete; however, such studies are limited when it comes to their applications in PC. To the author’s knowledge, this study is the first that investigated the strength characteristic of SA-based BF-reinforced PC subjected to elevated temperature. After an extensive literature review, two experimental phases, including the preliminary phase to set the appropriate mix design, and the main phase to investigate the compressive strength of samples after exposure to elevated temperatures of 100 °C, 150 °C, and 180 °C, were conducted. A summary of the research on PC/PPCC with and without fiber reinforcement is presented in Table 3 and Table 4, respecrively.

2. Materials and Methods

2.1. Materials

Materials including aggregates, BF, SA, and two-part epoxy resin for this experiment were provided by a local supplier. Basalt fiber, SA, and epoxy resin were productions originally from the U.S., Iran, and Germany, respectively. Triplicate cubic concrete samples of 50 × 50 × 50 mm were tested for the compressive strength of each mix composition. According to ASTM C117 and E11, the aggregates for the mortar were sieved through sieve number 4 to provide relatively fine aggregates of size less than 4.75 mm. The fineness modulus of the used aggregates was 2.72. Size distribution curve for aggregate used in the present research is shown in Figure 1. The review of the literature presented in Table 2 showed that the length of the BF considered in the table for production of polymer and cementitious composites is ranged from 6 to 45 mm. Shorter chopped lengths fibers are more effective for appropriate fiber distribution while longer chopped lengths are more effective in crack bridging action [68].
In order to achieve the objectives of this study and in light of the findings reported in previous studies, together with the literature review discussed in the introduction section, the average length of the chopped BF in the mortar mixtures is considered 10 mm for this study. Moreover, the amount of investigated fiber content reported in previous studies and literature review is ranged between 0.15% and 2% of the total volume of the PC. In this study, the content of the chopped BF in the mortar mixtures was considered 2% of the total volume, which is consistent with the literature. BF is another admixture in the composite to improve the thermal resistance of the composite. Nonetheless, the addition of BF into the mixtures may improve the fresh properties of the composite, it reduces the compressive strength of the hardened mortar specimens. In order to restrict the adverse effect and take advantage of BF in improving thermal resistance and rhetorical properties, a lower proportion of SA compared to that used in literature was incorporated. BF amount used in this study occupied only 0.1% of the total mortar volume. The geometrical and mechanical properties of the BF used in the present study are shown in Table 5. The fine fraction of aggregates 0–150 µm were excluded.
Two-part epoxy resin with a specific weight of 1.1 gr/cm3 was used in the experiment. In this process, a low viscous epoxy, known as saturant resin (Part A) and hardener (Part B), was mixed to obtain a homogeneous mixture. Epoxy resins are the most widely used PC compounds [69]. Epoxy concrete is composed of two parts of the epoxy and aggregate blend. The curing of the epoxy resin systems is an exothermic reaction. The aggregates were checked to be dry and free of dirt, debris, and organic materials using sieving and oven drying. The mix proportion of the saturant resin and hardener with the trading name of Araks FK20 was as per the guidelines of the manufacturer. Generally, epoxy formulations are skin sensitizers, and handling precautions should be taken carefully [69]. In order to determine the general proportions of selected materials in the experimental study, two amounts of 13% and 15% of the total mortar volume were used for epoxy content in the mix design stage. SA occupied 0.1% of the total mortar volume.
The curing time and temperature for PC are affected by parameters such as the epoxy formulation, mix proportions of PC composition, and mass volume. Generally, PC mortar cures in hours after casting while complete curing of the cement-based materials takes days or weeks.

2.2. Mix Designs and Preparation of Specimens

The experimental study was conducted in two phases, one aimed to obtain appropriate proportions and the other is to investigate the main objective of this research. In the first phase of the experimental program, two epoxy concrete mixes made of aggregate and epoxy resin without the inclusion of BF and SA were tested. The results of the compressive strength for the epoxy resin contents of 13% and 15% were 41.9 MPa and 39.7 MPa. For the second phase, different mix designs of PC were determined to have aggregate, epoxy resin, FB, and SA in the composition of the mix. The resin amount, in the PC mix designs, was constant and equal to 13% of the total concrete volume. The mix design for SA-based BF-reinforced mortar subjected to 0 °C and elevated temperatures of 100 °C, 150 °C, and 180 °C is presented in Table 6. Zero temperature in these tables that is room-curing temperature without heating was a reference before heating of samples. The hardened mortar specimens were exposed to three elevated temperatures of 100 °C, 150 °C and 180 °C for a duration of three hours. Then, the compressive strength of samples was tested and compared for each mix design under raised temperature conditions. To test compressive strength, PC samples were placed in the oven for three hours under the intended temperature after 7 days of curing. In order to prevent cooling shock, samples were left at room temperature to cool down gradually before applying compressive load in the testing machine.

2.3. Compressive Strength Testing

The compressive strength tests of the specimens were carried out according to ASTM C109-08. For each mixture, three 50 × 50 × 50 mm cubic specimens of each mix were tested after 3 days. In order to determine the ultimate load-carrying capacity under pure compression, the axial compressive load was applied continuously with a rate of 2.4 kN/sec according to ASTM C31 by a compression machine with 2000 kN capacity. The load was applied until the rupture of the specimen which means that the failure of the specimen at the point, the reading of the load in kN, and stress in MPa were recorded from the dial gauge.

3. Result and Discussions

Compressive tests were performed to investigate the temperature resistance of PC mortar and the effect of the inclusion of chopped BF and SA on the mechanical strength of PC. The results of the ultimate load-carrying capacity of the mortar specimens for different mix designs are shown in Table 7. The result shows that incorporation of BF and/or SA into PC reduced the compressive strength to 33.8, 34.3, and 35.3, which is consistent with the observation published by Dias and Thaumaturgo [34] and Niaki et al. [36], that BF reduces the compressive strength of PC.
Figure 2 shows the variation in the compressive strength of different mix designs compared to the reference state. The analysis of the results obtained by the compressive strength test of the samples maintained at room temperature revealed that the addition of BF and SA to the reference mix design decreased compressive strength. A lower decrease in compressive strength was observed for the mortar mix containing sodium alginate (SA) compared to those with only alginate. The samples with sodium alginate pose lower compressive strength due to the amount and size of voids in the structure of the hardened composite compared to the control mortar [24]. This weakening of compressive strength may ascribe to forming voids by the addition of the fibers into the matrix, which is in agreement with test results by Reis [30], Kabay [59], and Sun et al. [66]. However, the observation during uniaxial compressive loading indicated that the inclusion of FB leads to the absorption of a large amount of plastic energy before failure that is indicated also in other research studies. He and Lu [70] and Donker and Obonyo [71]. The analysis of the results obtained by the compressive strength test for polymer mortar containing BF and SA revealed that the addition of BF and SA to the reference mix design enhanced compressive strength growth in elevated temperatures. The cure of an epoxy resin is an exothermic process; however, the observation of an experimental investigation on curing temperature of epoxy resin conducted by Lahouar et al. [72] reported absorption of energy between 20 °C and 80 °C representing an endothermic peak due to possible phase change in curing state of the resin. Variations in the compressive strength after heating of the PC containing either BF or SA alone or both together indicate improving the effectiveness of these additives. The obtained results are not in agreement with those of either two works of Reis et al. [35] and [30], indicating a sharp increase (up to 50%) or drastic drop of compressive strength due to temperature elevation.
The negative percentage for variation of compressive strength of MPT specimens is in agreement with previous studies that indicated the adverse effect of elevated temperature on the hardened specimens. The increase in the variation of compressive strength in MPBT, MPAT, and MPBAT supports the finding of this research that the addition of BF and SA enhances the compressive strength of PC. The highest positive influence by raised temperature is corresponding to the mixture including SA alone that with an increase in the temperature from room temperature to 100 °C then 150 °C and finally to 180 °C, growth in the compressive strength became 21.08%, 29.45%, and 30.61%, respectively. A lower increase of the compressive strength of the BF-reinforced PC is due to the entrapped voids because of fiber addition. The obtained results about enhanced temperature resistance of SA-based mortar specimens also conform to the finding of DeBrouse [23].
Utilization of both BF and SA together in the PC has been with a lower increase from room temperature to 100 °C until 180 °C, compared to the individual application of these materials. However, it can be seen that heating of PC for 3 h under 180 °C resulted in strength reduction, while together or individual effectiveness of BF and SA in the PC is remarkable on the improvement in the compressive strength after heating for all the three temperatures. From Table 6, the highest compressive strength for samples of PC after 3 days of air curing was higher than the cement concrete samples. A comparison between the obtained compressive strength of polymer and cement concretes with BF and SA indicated that the elevated temperature had a greater influence on polymer concrete than that of cement concrete. The rate of increase was around 20% for polymer concrete, while it was around 5% for cement concrete for temperature rise from room temperature to 180 °C.
Polymer concrete samples after failure under compression test are shown in Figure 3. It should be demonstrated that PC samples exhibited no obvious crack on the surface when the applied load was released after failure under compression. The produced polymer concrete can be considered as a composition of two phases of polymer mortars and aggregates. Through applying the load displacement into the specimens, it was observed that the failure was initiated by the creation of interphase microcracks between polymer mortars and aggregates. By increasing the compressive load, longitudinal cracks were propagated along the jaws’ direction, which was similar to those that happen in normal concrete. It was observed that the time between crack initiation and total failure of the specimens was lower than that of cementitious concrete specimens [24,73].
The compressive load was applied until the mortar specimens of each mix reached their ultimate load-bearing capacity and failure occurred. High plastic deformation of the polymer concrete was observed corresponding to the applied load.

4. Conclusions

In order to investigate the effectiveness of the addition of BF with the natural source and/or SA as a waste material to reduce the strength degradation of polymer concretes after exposure to raised temperatures of 100 °C, 150 °C, and 180 °C, an experimental program was conducted in two phases including primary study and main study. The addition of BF and/or SA decreases concrete compressive strength under room temperature, but the addition of BF and SA each alone leads to compressive strength increase during exposure to heat. An increase in the temperature to 180 °C showed an almost positive effect on the compressive strength. The highest positive influence by raised temperature is corresponding to the mixture including SA alone that, with an increase in the temperature from room temperature to 100 °C then 150 °C and finally to 180 °C, growth in the compressive strength became 21.08%, 29.45%, and 30.61%, respectively. The addition of BF and SA both together increases the rate of strength growth of polymer concrete under heat from 100 °C to 180 °C. In view of these conclusions, the addition of BF and SA results in a decrease of compressive strength of polymer concretes under room temperature, but it is obvious that these materials are positive to improve the resistance of PC. However, the most important outcome of this research may be a natural source of BF and waste material SA to prevent PC compressive strength degradation under elevated temperature. Increase of compressive strength up to 30%, that when noticed to the cement amount with the energy-wasting process to be produced, is a very valuable finding.

Author Contributions

Conceptualization, N.A.O., S.E.M.-Y. and H.A.A.; methodology, N.A.O., S.E.M.-Y. and H.A.A.; software, N.A.O.; validation, N.A.O., S.E.M.-Y. and H.A.A.; formal analysis, N.A.O., S.E.M.-Y. and H.A.A.; investigation, N.A.O., S.E.M.-Y. and H.A.A.; resources, N.A.O., S.E.M.-Y., H.A.A., H.S., S.S.R.K., and M.P.; data curation, N.A.O.; writing—original draft preparation, N.A.O., S.E.M.-Y., H.A.A. and H.S.; writing—review and editing, N.A.O., S.E.M.-Y., H.A.A., H.S., M.P. and S.S.R.K.; visualization, N.A.O., S.E.M.-Y., H.A.A., H.S., S.S.R.K., and M.P.; supervision, S.E.M.-Y., H.A.A.; project administration, N.A.O., S.E.M.-Y., H.A.A., H.S., S.S.R.K., and M.P. funding acquisition, N.A.O., S.E.M.-Y., H.A.A., H.S., S.S.R.K., and M.P. All authors have read and agreed to the published version of the manuscript.

Funding

The research was supported by the Islamic Azad University, Najafabad Branch, and the Ministry of Education, Youth, and Sports of the Czech Republic and the European Union (European Structural and Investment Funds Operational Program Research, Development, and Education) in the framework of the project “Modular platform for autonomous chassis of specialized electric vehicles for freight and equipment transportation”, Reg. No. CZ.02.1.01/0.0/0.0/16_025/0007293.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mohammadyan-Yasouj, S.E.; Marsono, A.K.; Abdullah, R.; Moghadasi, M. Wide beam shear behavior with diverse types of reinforcement. ACI Struct. J. 2015, 112, 199–208. [Google Scholar] [CrossRef]
  2. Moghadasi, M.; Marsono, A.K.; Mohammadyan-Yasouj, S.E. A study on rotational behaviour of a new industrialised building system connection. Steel Compos. Struct. 2017, 25, 245–255. [Google Scholar]
  3. Shokravi, H.; Shokravi, H.; Bakhary, N.; Koloor, S.S.R.; Petrů, M. A Comparative Study of the Data-driven Stochastic Subspace Methods for Health Monitoring of Structures: A Bridge Case Study. Appl. Sci. 2020, 10, 3132. [Google Scholar] [CrossRef]
  4. Shokravi, H.; Shokravi, H.; Bakhary, N.; Heidarrezaei, M.; Koloor, S.S.R.; Petru, M. Vehicle-assisted techniques for health monitoring of bridges. Sensors 2020, 20, 3460. [Google Scholar] [CrossRef] [PubMed]
  5. Shokravi, H.; Shokravi, H.; Bakhary, N.; Heidarrezaei, M.; Rahimian Koloor, S.S.; Petrů, M. Application of the Subspace-Based Methods in Health Monitoring of Civil Structures: A Systematic Review and Meta-Analysis. Appl. Sci. 2020, 10, 3607. [Google Scholar] [CrossRef]
  6. Shokravi, H.; Shokravi, H.; Bakhary, N.; Koloor, S.S.R.; Petrů, M. Health Monitoring of Civil Infrastructures by Subspace System Identification Method: An Overview. Appl. Sci. 2020, 10, 2786. [Google Scholar] [CrossRef] [Green Version]
  7. Aslani, F.; Samali, B. Predicting the bond between concrete and reinforcing steel at elevated temperatures. Struct. Eng. Mech. 2013, 48, 643–660. [Google Scholar] [CrossRef] [Green Version]
  8. Shokravi, H.; Mohammadyan-Yasouj, S.E.; Rahimian Koloor, S.S.; Petrů, M.; Heidarrezaei, M. Effect of alumina additives on mechanical and fresh properties of self-compacting concrete: A review. Processes 2021. (under review). [Google Scholar]
  9. Mohammadyan-Yasouj, S.E.; Heidari, N.; Shokravi, H. Influence of Waste Alumina Powder on Self-compacting Concrete Resistance under Elevated Temperature. J. Build. Eng. 2021, 14. (under review). [Google Scholar]
  10. Awal, A.S.M.A.; Shehu, I.A. Performance evaluation of concrete containing high volume palm oil fuel ash exposed to elevated temperature. Constr. Build. Mater. 2015, 76, 214–220. [Google Scholar] [CrossRef]
  11. Ahmad, S.; Sallam, Y.S.; Al-Hawas, M.A. Effects of key factors on compressive and tensile strengths of concrete exposed to elevated temperatures. Arab. J. Sci. Eng. 2014, 39, 4507–4513. [Google Scholar] [CrossRef]
  12. Thirumurugan, S.; Anandan, S. Residual strength characteristics of polymer fibre concrete exposed to elevated temperature. Eng. J. 2015, 19, 117–131. [Google Scholar] [CrossRef] [Green Version]
  13. Fraternali, F.; Ciancia, V.; Chechile, R.; Rizzano, G.; Feo, L.; Incarnato, L. Experimental study of the thermo-mechanical properties of recycled PET fiber-reinforced concrete. Compos. Struct. 2011, 93, 2368–2374. [Google Scholar] [CrossRef]
  14. Rabehi, B.; Ghernouti, Y.; Boumchedda, K.; Li, A.; Drir, A. Durability and thermal stability of ultra high-performance fibre-reinforced concrete (UHPFRC) incorporating calcined clay. Eur. J. Environ. Civ. Eng. 2017, 21, 594–611. [Google Scholar] [CrossRef]
  15. 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. [Google Scholar] [CrossRef]
  16. Pathak, T.S.; San Kim, J.; Lee, S.-J.; Baek, D.-J.; Paeng, K.-J. Preparation of alginic acid and metal alginate from algae and their comparative study. J. Polym. Environ. 2008, 16, 198–204. [Google Scholar] [CrossRef]
  17. Engbert, A.; Gruber, S.; Plank, J. The effect of alginates on the hydration of calcium aluminate cement. Carbohydr. Polym. 2020, 236, 116038. [Google Scholar] [CrossRef]
  18. Wang, J.; Mignon, A.; Snoeck, D.; Wiktor, V.; Van Vliergerghe, S.; Boon, N.; De Belie, N. Application of modified-alginate encapsulated carbonate producing bacteria in concrete: A promising strategy for crack self-healing. Front. Microbiol. 2015, 6, 1088. [Google Scholar] [CrossRef] [Green Version]
  19. Ouwerx, C.; Velings, N.; Mestdagh, M.M.; Axelos, M.A.V. Physico-chemical properties and rheology of alginate gel beads formed with various divalent cations. Polym. Gels Netw. 1998, 6, 393–408. [Google Scholar] [CrossRef]
  20. Abbas, W.A.; Mohsen, H.M. Effect of Biopolymer Alginate on some properties of concrete. J. Eng. 2020, 26, 121–131. [Google Scholar]
  21. Heidari, A.; Ghaffari, F.; Ahmadvand, H. Properties of Self Compacting Concrete Incorporating Alginate and Nano Silica. Asian J. Civil Eng. Build. Hous. 2015, 16, 1–16. [Google Scholar]
  22. Mignon, A.; Snoeck, D.; D’Halluin, K.; Balcaen, L.; Vanhaecke, F.; Dubruel, P.; Van Vlierberghe, S.; De Belie, N. Alginate biopolymers: Counteracting the impact of superabsorbent polymers on mortar strength. Constr. Build. Mater. 2016, 110, 169–174. [Google Scholar] [CrossRef]
  23. DeBrouse, D.R. Alginate-Based Building Materials; World Intellectual Property Organization: Geneva, Switzerland, 2012. [Google Scholar]
  24. Mohammadyan-Yasouj, S.E.; Ahangar, H.A.; Oskoei, N.A.; Shokravi, H.; Koloor, S.S.R.; Petrů, M. Thermal Performance of Alginate Concrete Reinforced with Basalt Fiber. Crystals 2020, 10, 779. [Google Scholar] [CrossRef]
  25. Barbuta, M.; Rujanu, M.; Nicuta, A. Characterization of polymer concrete with different wastes additions. Procedia Technol. 2016, 22, 407–412. [Google Scholar] [CrossRef] [Green Version]
  26. Lokuge, W.; Aravinthan, T. Effect of fly ash on the behaviour of polymer concrete with different types of resin. Mater. Des. 2013, 51, 175–181. [Google Scholar] [CrossRef]
  27. Venkatesh, B.; Student, U.G. Review on performance of polymer concrete with resins and its applications. Int. J. Pure Appl. Math. 2018, 119, 175–184. [Google Scholar]
  28. Abd_Elmoaty, A.M. Self-healing of polymer modified concrete. Alexandria Eng. J. 2011, 50, 171–178. [Google Scholar] [CrossRef] [Green Version]
  29. Rebeiz, K.S. Precast use of polymer concrete using unsaturated polyester resin based on recycled PET waste. Constr. Build. Mater. 1996, 10, 215–220. [Google Scholar] [CrossRef]
  30. Reis, J.M.L. Effect of temperature on the mechanical properties of polymer mortars. Mater. Res. 2012, 15, 645–649. [Google Scholar] [CrossRef]
  31. Ferdous, W.; Manalo, A.; Aravinthan, T.; Van Erp, G. Properties of epoxy polymer concrete matrix: Effect of resin-to-filler ratio and determination of optimal mix for composite railway sleepers. Constr. Build. Mater. 2016, 124, 287–300. [Google Scholar] [CrossRef]
  32. Golestaneh, M.; Amini, G.; Najafpour, G.D.; Beygi, M.A. Evaluation of mechanical strength of epoxy polymer concrete with silica powder as filler. World Appl. Sci. J. 2010, 9, 216–220. [Google Scholar]
  33. Elalaoui, O.; Ghorbel, E.; Mignot, V.; Ouezdou, M. Ben Mechanical and physical properties of epoxy polymer concrete after exposure to temperatures up to 250 C. Constr. Build. Mater. 2012, 27, 415–424. [Google Scholar] [CrossRef]
  34. Dias, D.P.; Thaumaturgo, C. Fracture toughness of geopolymeric concretes reinforced with basalt fibers. Cem. Concr. Compos. 2005, 27, 49–54. [Google Scholar] [CrossRef]
  35. Reis, J.M.L. Mechanical characterization of fiber reinforced polymer concrete. Mater. Res. 2005, 8, 357–360. [Google Scholar] [CrossRef] [Green Version]
  36. Hassani Niaki, M.; Fereidoon, A.; Ghorbanzadeh Ahangari, M. Mechanical properties of epoxy/basalt polymer concrete: Experimental and analytical study. Struct. Concr. 2018, 19, 366–373. [Google Scholar] [CrossRef]
  37. Zhang, Y.; Yu, C.; Chu, P.K.; Lv, F.; Zhang, C.; Ji, J.; Zhang, R.; Wang, H. Mechanical and thermal properties of basalt fiber reinforced poly (butylene succinate) composites. Mater. Chem. Phys. 2012, 133, 845–849. [Google Scholar] [CrossRef]
  38. Mohamad, S.A.; Al-Hamd, R.K.S.; Khaled, T.T. Investigating the effect of elevated temperatures on the properties of mortar produced with volcanic ash. Innov. Infrastruct. Solut. 2020, 5, 1–11. [Google Scholar] [CrossRef]
  39. Abdel-Fattah, H.; El Hawary, M.M.; Falah, A. Effect of elevated temperatures on the residual fracture toughness of epoxy modified concrete. Kuwait J. Sci. Eng. 2000, 27, 27–39. [Google Scholar]
  40. Al-Salloum, Y.A.; Elsanadedy, H.M.; Abadel, A.A. Behavior of FRP-confined concrete after high temperature exposure. Constr. Build. Mater. 2011, 25, 838–850. [Google Scholar] [CrossRef]
  41. Behnood, A.; Ziari, H. Effects of silica fume addition and water to cement ratio on the properties of high-strength concrete after exposure to high temperatures. Cem. Concr. Compos. 2008, 30, 106–112. [Google Scholar] [CrossRef]
  42. Amin, M.; Tayeh, B.A. Investigating the mechanical and microstructure properties of fibre-reinforced lightweight concrete under elevated temperatures. Case Stud. Constr. Mater. 2020, 13, e00459. [Google Scholar] [CrossRef]
  43. Tang, Y.; Feng, W.; Feng, W.; Chen, J.; Bao, D.; Li, L. Compressive properties of rubber-modified recycled aggregate concrete subjected to elevated temperatures. Constr. Build. Mater. 2021, 268, 121181. [Google Scholar] [CrossRef]
  44. De Larissa, C.A.; dos Anjos, M.A.S.; de Sa, M.V.V.A.; de Souza, N.S.L.; de Farias, E.C. Effect of high temperatures on self-compacting concrete with high levels of sugarcane bagasse ash and metakaolin. Constr. Build. Mater. 2020, 248, 118715. [Google Scholar]
  45. Wang, W.-C.; Wang, H.-Y.; Chang, K.-H.; Wang, S.-Y. Effect of high temperature on the strength and thermal conductivity of glass fiber concrete. Constr. Build. Mater. 2020, 245, 118387. [Google Scholar] [CrossRef]
  46. Zhou, J.; Lu, D.; Yang, Y.; Gong, Y.; Ma, X.; Yu, B.; Yan, B. Physical and Mechanical Properties of HighStrength Concrete Modified with Supplementary Cementitious Materials after Exposure to Elevated Temperature up to 1000 °C. Materials 2020, 13, 532. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Zhang, H.; Li, L.; Yuan, C.; Wang, Q.; Sarker, P.K.; Shi, X. Deterioration of ambient-cured and heat-cured fly ash geopolymer concrete by high temperature exposure and prediction of its residual compressive strength. Constr. Build. Mater. 2020, 262, 120924. [Google Scholar] [CrossRef]
  48. Wang, Y.; Xu, T.; Liu, Z.; Li, G.; Jiang, J. Seismic behavior of steel reinforced concrete cross-shaped columns after exposure to high temperatures. Eng. Struct. 2021, 230, 111723. [Google Scholar] [CrossRef]
  49. Alharbi, Y.R.; Abadel, A.A.; Elsayed, N.; Mayhoub, O.; Kohail, M. Mechanical properties of EAFS concrete after subjected to elevated temperature. Ain Shams Eng. J. 2020. [Google Scholar] [CrossRef]
  50. Sarayu, K.; Gopinath, S.; Ramachandra, M.A.; Iyer, N.R. Structural stability of basalt fibers with varying biochemical conditions-A invitro and invivo study. J. Build. Eng. 2016, 7, 38–45. [Google Scholar] [CrossRef]
  51. Niaki, M.H.; Fereidoon, A.; Ahangari, M.G. Experimental study on the mechanical and thermal properties of basalt fiber and nanoclay reinforced polymer concrete. Compos. Struct. 2018, 191, 231–238. [Google Scholar] [CrossRef]
  52. Elalaoui, O.; Ghorbel, E.; Ouezdou, M. Ben Influence of flame retardant addition on the durability of epoxy based polymer concrete after exposition to elevated temperature. Constr. Build. Mater. 2018, 192, 233–239. [Google Scholar] [CrossRef]
  53. Gorninski, J.P.; Tonet, K.G.; Sokołowska, J.J. Use of polishing alumina as flame retardant in orthophthalic polyester resin matrix composites. In Proceedings of the Advanced Materials Research; Trans Tech Publications: Zürich, Switzerland, 2015; Volume 1129, pp. 209–216. [Google Scholar]
  54. Jiang, C.; Fan, K.; Wu, F.; Chen, D. Experimental study on the mechanical properties and microstructure of chopped basalt fibre reinforced concrete. Mater. Des. 2014, 58, 187–193. [Google Scholar] [CrossRef]
  55. Shafiq, N.; Ayub, T.; Khan, S.U. Investigating the performance of PVA and basalt fibre reinforced beams subjected to flexural action. Compos. Struct. 2016, 153, 30–41. [Google Scholar] [CrossRef]
  56. Fenu, L.; Forni, D.; Cadoni, E. Dynamic behaviour of cement mortars reinforced with glass and basalt fibres. Compos. Part B Eng. 2016, 92, 142–150. [Google Scholar] [CrossRef]
  57. Kizilkanat, A.B.; Kabay, N.; Akyüncü, V.; Chowdhury, S.; Akça, A.H. Mechanical properties and fracture behavior of basalt and glass fiber reinforced concrete: An experimental study. Constr. Build. Mater. 2015, 100, 218–224. [Google Scholar] [CrossRef]
  58. Sarkar, A.; Hajihosseini, M. Feasibility of Improving the Mechanical Properties of Concrete Pavement Using Basalt Fibers. J. Test. Eval. 2020, 48, 2908–2917. [Google Scholar] [CrossRef]
  59. Kabay, N. Abrasion resistance and fracture energy of concretes with basalt fiber. Constr. Build. Mater. 2014, 50, 95–101. [Google Scholar] [CrossRef]
  60. Khan, M.; Cao, M. Effect of hybrid basalt fibre length and content on properties of cementitious composites. Mag. Concr. Res. 2020, 1–12. [Google Scholar] [CrossRef]
  61. Amuthakkannan, P.; Manikandan, V.; Jappes, J.T.W.; Uthayakumar, M. Effect of fibre length and fibre content on mechanical properties of short basalt fibre reinforced polymer matrix composites. Mater. Phys. Mech. 2013, 16, 107–117. [Google Scholar]
  62. Ayub, T.; Shafiq, N.; Nuruddin, M.F. Mechanical properties of high-performance concrete reinforced with basalt fibers. Procedia Eng. 2014, 77, 131–139. [Google Scholar] [CrossRef] [Green Version]
  63. Girgin, Z.C.; Yıldırım, M.T. Usability of basalt fibres in fibre reinforced cement composites. Mater. Struct. 2016, 49, 3309–3319. [Google Scholar] [CrossRef]
  64. Zhao, Y.-R.; Wang, L.; Lei, Z.-K.; Han, X.-F.; Xing, Y.-M. Experimental study on dynamic mechanical properties of the basalt fiber reinforced concrete after the freeze-thaw based on the digital image correlation method. Constr. Build. Mater. 2017, 147, 194–202. [Google Scholar] [CrossRef]
  65. Katkhuda, H.; Shatarat, N. Improving the mechanical properties of recycled concrete aggregate using chopped basalt fibers and acid treatment. Constr. Build. Mater. 2017, 140, 328–335. [Google Scholar] [CrossRef]
  66. Sun, X.; Gao, Z.; Cao, P.; Zhou, C. Mechanical properties tests and multiscale numerical simulations for basalt fiber reinforced concrete. Constr. Build. Mater. 2019, 202, 58–72. [Google Scholar] [CrossRef]
  67. Afroz, M.; Patnaikuni, I.; Venkatesan, S. Chemical durability and performance of modified basalt fiber in concrete medium. Constr. Build. Mater. 2017, 154, 191–203. [Google Scholar] [CrossRef]
  68. Bi, J.; Huo, L.; Zhao, Y.; Qiao, H. Modified the smeared crack constitutive model of fiber reinforced concrete under uniaxial loading. Constr. Build. Mater. 2020, 250, 118916. [Google Scholar] [CrossRef]
  69. Committee, A.C.I. Guide for the use of polymers in concrete. ACI Comm. Rep. 2009, 548, 1–100. [Google Scholar]
  70. He, D.; Lu, Z. Experimental Study on Mechanical Properties of Chopped Basalt Fiber Reinforced Concrete. J. Henan Univ. Nat. Sci. 2009, 3, 320–322. [Google Scholar]
  71. Donkor, P.; Obonyo, E. Earthen construction materials: Assessing the feasibility of improving strength and deformability of compressed earth blocks using polypropylene fibers. Mater. Des. 2015, 83, 813–819. [Google Scholar] [CrossRef]
  72. Lahouar, M.A.; Caron, J.-F.; Pinoteau, N.; Forêt, G.; Benzarti, K. Mechanical behavior of adhesive anchors under high temperature exposure: Experimental investigation. Int. J. Adhes. Adhes. 2017, 78, 200–211. [Google Scholar] [CrossRef] [Green Version]
  73. Joshani, M.; Koloor, S.S.R.; Abdullah, R. Damage Mechanics Model for Fracture Process of Steel-concrete Composite Slabs. In Applied Mechanics and Materials; Trans Tech Publications Ltd.: Stafa-Zurich, Switzerland, 2012; pp. 339–345. [Google Scholar]
Figure 1. Size distribution curve for aggregate used in the present research.
Figure 1. Size distribution curve for aggregate used in the present research.
Processes 09 00510 g001
Figure 2. Failure of samples under compression test (all tests were conducted in triplicate).
Figure 2. Failure of samples under compression test (all tests were conducted in triplicate).
Processes 09 00510 g002
Figure 3. Failure of samples under compression test.
Figure 3. Failure of samples under compression test.
Processes 09 00510 g003
Table 3. Polymeric concrete.
Table 3. Polymeric concrete.
RemarksFiller and PolymerCodes & Sample Size (mm)Test Age (Day)Conducted Tests and Sample MonitoringRef.
Amount (% of the Total Volume (Vol%) or the Total Weight (wt %))DetailsName
  • 15% epoxy & 200% filler (15% fine silica, 25% medium-size silica, & 60% coarse silica powder) had maximum compressive & flexural strengths
  • Tensile strength was maximized with 20% resin & 200% filler
  • Mechanical strength of polymer concrete was 4–5-folds higher than cementitious concrete
100, 150, & 200 Vol%50–60, 600, & 1100 μm SizeSilica Fume
ASTM (C33-03, C579-01, C293-02, C496/C496/M, & C496-04)
  • 50 × 50 × 50 cube
  • 76.2 × 152.4 cylinder
-
  • Uniaxial compressive test
  • Uniaxial tensile test
  • Flexural compressive test
Golestaneh et al. 2010 [32]
10, 15, & 20 wt %Epoxy, Bisphenol A Epoxy Resin
  • Investigate the effect of temperature on the performance of epoxy and unsaturated polyester polymer mortars
  • Higher ductility after 60 °C
  • Epoxy more sensitive than polyester
  • From room temperature to 60 °C, flexural and compressive strength decreases drastically as temperature increases; a loss of more than 50% is reported
  • The epoxy mortars are more sensitive to temperature changes than the unsaturated polyester ones
88 Vol%245 μmOnly Coarse Aggregate
RILEM TC113/PC-2, RILEM TC113/PCM-8, ASTM (C39-05, C348-02)
  • 40 × 40 × 160 prism
  • 50 × 100 cylinder
7
  • Uniaxial compressive test
  • Flexural compressive test
Reis 2012 [30]
12 wt %Diglycidyl ether bisphenol A & an aliphatic amineHardener, RR515Epoxy Resin
  • Mechanical and physical properties of epoxy polymer concrete after exposure to temperatures up to 250 °C
  • Optimum polymer amount is 13%
  • 50% reduction in compressive strength after exposure to 250 °C
  • No change in flexure strength
  • The optimal polymer content is 13% which leads to obtain the highest physical and mechanical properties at the lowest cost
-(0–4) × 103 μmSand
RILEM CPT (PCM 2, PCM 8), NF EN (1097-6, 197-1), EN (12390-3, 206-1, 934-2)
  • 50 × 50 × 305 prism
  • 70 × 70 × 280 prism
  • 50 × 100 cylinder
  • 150 × 300 cylinder
  • 50 × 150 cylinder
7 & 28
  • Uniaxial compressive test
  • Flexural compressive test
  • Heating
  • Porosity
  • Scanning electron microscopy
  • E-value calculation
Elalaoui et al. 2012 [33]
-(4–10) × 103 μmGravel
6, 9, 13, & 16 wt %Bisphenol A diglycidyl ether resin, Eponal 371Epoxy resin
  • Effect of fly ash on the behavior of polymer concrete with different types of resin was investigated
  • Improve in ductility & increase in E-value by decreasing in fly ash content for all the mixes of PC concrete
  • Compressive strength as high as 100 MPa can be achieved using resins in polymer concrete
  • About 80% of the 28-day compressive strength can be achieved in 7 days for polymer concrete
  • Split tensile and flexural strengths decreased with the increasing fly ash content for all the mixes
0, 3, 10, 13, 20, 21, & 23 Vol% 15 μm (Type F)Fly Ash
ASTM (C128, D695, D3967), ISO 178
  • 9 × 16 × 160 prism
  • 50 × 100 cylinder
3, 7, 14, 21, & 28
  • Uniaxial compressive test
  • Uniaxial tensile test
  • Flexural compressive test
  • E-value calculation
Lokuge & Aravinthan 2013 [26]
20, 22, 30, 40, & 43 Vol%OrthophthalicPolyester Resin
BisphenolVinylester Resin
ThixotropicEpoxy Resin
  • Increase in density of polymer matrix by an increase in filler content which resulted in more & bigger voids
  • An increase in filler amount to 60% reduced the flexural strength by 70%.
  • No visible failure under compression was observed for the mixes with up to 30% filler.
  • An increase in filler amount leads to increased compressive modulus of elasticity.
  • Adding filler could help preserve structural performance through absorbing or blocking UV radiation before reaching chromophores on which the color of a polymer matrix is dependent
  • The mix with 30% filler has better mechanical properties while the mix with 50% filler would be the most economical choice. Both 30% and 50% filler mix have high durability for UV radiation
0, 10, 20, 30, 40, 50, & 60 Vol%Fire Retardant, Hollow Microsphere, Fly Ashlight-weight Filler
ASTM (D7028, D4065, C905, C580, & C579)
  • 60 × 14 × 4 prism
  • 80 × 10 × 10 prism
  • 25 × 25 cylinder
  • Uniaxial compressive test
  • Uniaxial tensile test
  • Flexural compressive test
  • Porosity
  • Ultraviolet
  • Density
  • Heat Generation
  • Glass transition temperature
Ferdous et al. 2016 [31]
Table 4. Polymer concrete (PC)/polymer-Portland cement concrete (PPCC) with fiber.
Table 4. Polymer concrete (PC)/polymer-Portland cement concrete (PPCC) with fiber.
RemarksBasalt and PolymerCodes & Sample Size (mm)Test Age (Day)Conducted Tests and Sample MonitoringRef.
Amount (% of the Total Volume (Vol%) or the Total Weight (wt %))Etra detailsDiameter (μm)Length (mm)Name
  • Geopolymer concrete showed higher strengths than Portland cement concretes
  • BF of 1.0% resulted in 26.4% & 12% reductions in compressive and splitting tensile strengths for Portland cement concretes
  • BF in geopolymer concretes resulted in higher flexural & splitting tensile strength than geopolymer concretes without fibers
  • Addition of fibers to all concretes tested caused increases in the VeBe time
  • Geopolymer concretes exhibited less cracks that is probably due to the bond between fiber and matrix
(0, 0.5, 1) Vol%-945Basalt Fibre (BF)
NBR (5738/93, 5739/94, 7222/94, 7222/94), ASTM (1993), ESIS (1992), Rilem Draft Recommendations (1989)
  • 150 × 300 cylinder
  • 150 × 150 × 500 prism (with 30 mm depth and 3 mm thickness notch)
28
  • Uniaxial compressive test
  • Uniaxial tensile test
  • Flexural compressive test
  • VeBe consistometer test
Dias & Thaumaturgo 2005 [34]
A/B = 1 (13.85 wt %) & 47.7 wt %A: SiO2/Al2O3 ratio = 5.35, (Na2O + K2O)/SiO2 = 0.209, B:SiO2/Na2O = 2.24, KOH = 14M--Geopolymer (Poly(Siloxo-Sialate))
  • Comparison between epoxy polymer concrete plain, reinforced with CF and GF
  • Increase in compressive strength of epoxy polymer (27.5–45.4% by GF & 36.1–55.1% by CF was observed
  • Slightly ductile failure of reinforced polymer, while unreinforced polymer showed a brittle failure
  • Comparing plain polymer concrete to ordinary concrete, compressive strength is 85% higher
  • When polymer concrete is reinforced, more than 100% increase in compression strength was observed
  • Epoxy polymer concrete (with and without fiber) proved to be an excellent alternative to concretes
1 & 2 Vol%-Chopped6Carbon Fiber (CF)
RILEM, ASTM (C469-94, C617-98)
  • 50 × 100 cylinder
7
  • Uniaxial compressive test
Reis 2005 [35]
6Glass Fiber (GF)
20% wt %Eposil 551 (Silicem®), Based on a diglycidyl-ether of bisphenol A, Aliphatic amine hardener --Epoxy Polymer
  • Maximum tensile and flexural strengths of BF/polyester poly butylene succinate (BF/PBS) were achieved at a fiber loading of 15 vol% and a good interface adhesion between the fiber and matrix observed in the composites
  • Impact strength of BF/PBS decreased with fiber addition primarily and then increased by increasing fiber
  • Thermal deflection temperature (HDT) of BF/PBS composite was significantly higher than the HDT of PBS resin
(0.0–15) Vol%-10continuous twistless rovingBF (coated by silane coupling agent)
GB/T(1040-1992, 9341, 1843, 1633, 1634)
  • Dumbbell shape
  • 4 × 10 × 80 prism (with 2 mm depth notch)
-
  • Uniaxial tensile test
  • Flexural compressive test
  • Impact test
  • Thermal stability
  • Scanning electron microscopy
Zhang et al. 2012 [37]
20–30 wt %Diglycidyl ether of bisphenol A resin (DGEBA), Low viscosity (CY 184), Polyaminehardener (Aradur ®2965)--Resin
  • BF was exposed to water, styrene butadiene rubber and superplasticizer that prevails in cementitious materials
  • BF >1% reduced concrete strength & basalt was degraded and degenerated by biochemical changes
  • In increased alkaline conditions, spalling of fibers was observed when exposed to the pH of 12 & further disintegration of fibers were noticed when exposed to polycarboxylate based superplasticizer and polymer matrix of styrene butadyene rubber at pH 12
  • Further studies are required to enhance the structural stability of BF in the concrete matrix
(0.5, 1, 1.5, & 2) Vol%-1324BF
-
  • 100 × 100 × 100 cube
  • 150 × 300 cylinder
14 & 28
  • Compression
  • Uniaxial tensile test
  • X-ray diffraction test
  • Fourier transform infrared
  • Scanning electron microscopy
  • Energy dispersive X-ray
Sarayu et al. 2016 [50]
97–85 vol%Mp 114 °C, Vicat softening temperature 96 °C--Polymer/Styrene Butadiene Rubber (SBR)
  • Mechanical properties of an epoxy/basalt polymer concrete at 24, 50, 75, & 100 °C were investigated
  • Increase in the amount of epoxy resin to 25 wt % enhanced mechanical properties of this concrete
  • Increase in temperature resulted in more ductility, but mechanical properties deterioration and stiffness loss
  • Decrease in aggregate size resulted in higher flexural & splitting tensile & lower compressive strengths
(0, 70, 72.5, 75, 77.5, & 80) Vol%-1–5-Crushed Basalt (aggregate)
  • ASTM (C39/C39M, C580, C496/C496M)
  • 50 × 100 cylinder
  • 25 × 25 × 54 prism
7 & 28
  • Uniaxial compressive test
  • Uniaxial tensile test
  • Flexural compressive test
Scanning electron microscopy
Niaki et al. 2017 [36]
Table 5. Properties of BF used in the experiment.
Table 5. Properties of BF used in the experiment.
Cutting Length (mm)10
Diameter (μm)17
Density (gr/cm3)2.65
Elastic modulus (GPa)93–110
Tensile strength (MPa)4100–4800
Elongation (%)1.3–3.2
Softening point (˚C)1050
Water absorption (%)<0.5
Table 6. Details of PC in the main study.
Table 6. Details of PC in the main study.
Name **BF (%) *SA (%) *Evaluation Temperature (°C)
MPT0--0
MPT100--100
MPT150--150
MPT180--180
MPBT00.2-0
MPBT1000.2-100
MPBT1500.2-150
MPBT1800.2-180
MPAT0-0.10
MPAT100-0.1100
MPAT150-0.1150
MPAT180-0.1180
MPBAT00.20.10
MPBAT1000.20.1100
MPBAT1500.20.1150
MPBAT1800.20.1180
* of the total weight; ** MPT, MPBT, MPAT, MPBAT are PC with no SA and BF, with BF only, with SA only, and with SA and BF, respectively.
Table 7. Compressive strength for the PC with/without SA and BF.
Table 7. Compressive strength for the PC with/without SA and BF.
Name f c ( MPa ) Variation to MPT0 (%)Variation after Heating (%)
7-Day
MPT043.7-- Processes 09 00510 i001
MPT10047.33+8.31+8.31
MPT15046.3+5.95+5.95
MPT18039.85−8.81−8.81
MPBT033.8−21.10- Processes 09 00510 i002
MPBT10039.33−10.00+16.36
MPBT15041.5−5.03+22.78
MPBT18041.15−5.83+21.75
MPAT034.3−21.51- Processes 09 00510 i003
MPAT10041.53−4.97+21.08
MPAT15044.4+1.60+29.45
MPAT18044.8+2.52+30.61
MPBAT035.3−19.22- Processes 09 00510 i004
MPBAT10038.3−12.36+8.50
MPBAT15042.9−1.83+21.53
MPBAT18044.5+1.83+26.06
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mohammadyan-Yasouj, S.E.; Ahangar, H.A.; Oskoei, N.A.; Shokravi, H.; Koloor, S.S.R.; Petrů, M. Experimental Study on the Effect of Basalt Fiber and Sodium Alginate in Polymer Concrete Exposed to Elevated Temperature. Processes 2021, 9, 510. https://doi.org/10.3390/pr9030510

AMA Style

Mohammadyan-Yasouj SE, Ahangar HA, Oskoei NA, Shokravi H, Koloor SSR, Petrů M. Experimental Study on the Effect of Basalt Fiber and Sodium Alginate in Polymer Concrete Exposed to Elevated Temperature. Processes. 2021; 9(3):510. https://doi.org/10.3390/pr9030510

Chicago/Turabian Style

Mohammadyan-Yasouj, Seyed Esmaeil, Hossein Abbastabar Ahangar, Narges Ahevani Oskoei, Hoofar Shokravi, Seyed Saeid Rahimian Koloor, and Michal Petrů. 2021. "Experimental Study on the Effect of Basalt Fiber and Sodium Alginate in Polymer Concrete Exposed to Elevated Temperature" Processes 9, no. 3: 510. https://doi.org/10.3390/pr9030510

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop