Properties of Recycled Steel Fibers-Incorporated Concrete: Comparison
Please note this is a comparison between Version 2 by Sirius Huang and Version 1 by Asad Zia.

Accumulating vast amounts of pollutants drives modern civilization toward sustainable development. Construction waste is one of the prominent issues impeding progress toward net zero. Pollutants must be utilized in constructing civil engineering structures for a green ecosystem. Large-scale production of industrial steel fibers (ISFs) causes significant damage to the goal of a sustainable environment. Recycled steel fibers (RSFs) from waste tires have been suggested to replace ISFs. Using RSFs can help reduce pollution due to ISFs production; on the other hand, it can improve the performance of the concrete at a low cost. 

  • concrete
  • mortars
  • fiber reinforced concrete
  • mechanical strength
  • recycled steel fibers
  • raw steel fibers
  • waste tires
  • pollution reduction

1. Background

1.1. Tires and Global Pollution

Great attention to the global increase in tires is required, especially in the European Union [1,2][1][2]. More than one billion tires are used for replacement every year globally, with more than half being abandoned and waiting to be disposed of [3]. According to the European Tire and Rubber Manufacturers Association (ETRMA2021), its members manufactured 4.24 million tons of tires in 2021, representing 70% of the global tire industry’s turnover [4]. The European Union’s Landfill Directive [5], which went into force in July 2006, requires all end-of-life tires (ELTs) to be recycled or reused. Following the implementation of this regulation, ETRMA estimates that 91% of end-of-life tires (ELTs) were collected and treated for material recycling and energy recovery in 2018 [6]. About 40 million worn tires are processed annually in the United Kingdom alone. Material recovery, which uses secondary materials from ELTs in construction, automotive, and civil engineering applications, was used to treat around 2 million tons (61.75 percent of total ELTs handled). Apart from this, in 2009, China’s tire production by rubber consumption accounted for around 70% of the country’s total rubber resource consumption, resulting in 233 million waste tires weighing approximately 8.6 million tons, equivalent to approximately 3 million tons of rubber resources [7]. According to Pilakoutas et al. [8], over a billion discarded tires are produced worldwide. The accumulation of these tires is a significant difficulty because tire component materials are exceedingly complex, making natural degradation impossible [9]. This necessitates the careful management of this massive quantity of trash. Waste tires can be managed using various methods, including material recovery, energy recovery, retreating, export, and landfill disposal [8]. According to ETRMA 2018 [10], approximately 72% of waste tires in Slovakia were recycled, with zero percent used in civil engineering, public works, or backfilling. As a result, investigating the potential application of waste tires or their ingredients in the construction industry to increase waste tire recycling in Slovakia and other countries is critical. Rubber content such as carbon black, steel insert, oil, and vulcanizing agents, such as inserts, synthetic yarns, and textiles, are among the components of a tire, with percentages of 46 and 48 percent, 25–28 percent, 10–12 percent, and 3–6 percent, respectively [11]. The amount of steel fiber removed from waste tires varies according to the tire type. Steel is used in tires for up to 15% of lightweight vehicles and up to 25% of trucks [8]. Liew and Akbar [12] reported that beneficial products from recycling tires consist of rubber, textile fibers, and steel fibers. Rubber makes up the majority of the tire, accounting for nearly 47–48 percent of the total weight, followed by black carbon (22 percent), steel cords (15–17 percent), textile fabric (5 percent), additives (8 percent), and zinc oxide and sulfur (1 percent each) [13,14][13][14].

1.2. Raw Steel Fiber from Waste Tires and Concrete Composites

Concrete is extensively used in building materials because of its high compressive strength, durability, and environmental compatibility. Concrete is used in various architectural projects, including foundations, walls, bridges, roadways, dams, and reservoirs [15]. However, its service life may be severely shortened in demanding conditions. Although concrete structures usually are designed and built to last at least 50 years, the sulfuric acid attack can cause them to deteriorate in just a few years. Repair and, in some cases, complete replacement of damaged structures are necessary when corrosion rates rise, which can be very costly and entail many social issues. Many research efforts have been made to improve concrete qualities for greater applicability due to these varied uses [16,17,18][16][17][18]. Industrial steel fiber reinforced concrete (ISFRC) has been beneficial in various applications throughout the last three decades, including tunnel linings, hydraulic structures, slabs, bridge decks, foundations, refractory concrete fiber shotcrete, and precast parts [8]. Adding 1% industrial steel fiber (ISF) to concrete automatically doubles the material cost [19]. As a result, steel fiber collected from discarded tires has become a viable option for use as reinforcement for cement composites. This advocates an environmentally beneficial method of dealing with some of the issues linked with the generation of waste tires. It also functions as a tool for improving building sector sustainability [20].
Pawelska-Mazur and Kaszynska [21] compared the energy consumption and CO2 emissions of the concrete mixture with raw steel fibers recovered from waste tires (RSF) to a concrete recipe with industrial steel fibers (ISF). It was concluded that the concrete recipe with RSF uses 31.3 percent less energy and emits 30.8 percent less CO2. However, due to intricate and energy-intensive manufacturing procedures, industrially manufactured steel fibers are the second component (after cement) that substantially impacts the natural environment. Therefore, energy consumption and greenhouse gas emissions could be reduced by substituting waste fibers obtained from the recovery of rubber from used tires for industrially manufactured steel fibers.
Qin and Kaewunruen [2] demonstrated (refer to Table 7 in Qin and Kaewunruen [2]’s paper) that the RSF is far cheaper than ISF. The average cost of ISF is over five times that of RSF. Furthermore, replacing the ISF with RSF can reduce the cost from 15.89 USD/m3 to 40 USD/m3. According to the findings, the contribution of employing RSF substitutes for ISF in terms of economy and environment is significant. Furthermore, RSF’s social role cannot be overlooked. Waste tires pose a significant environmental problem [3]. Human health and water resources are irreversibly harmed by the combustion or burying of waste tires [22]. As an environmentally acceptable resource in civil construction, scrap tires improve building cleanliness and extend the structure’s service life. Even though RSF’s performance is neither higher nor inferior to ISF concrete, it can still be a good substitute for ISF in economics, the environment, carbon emissions, and social development. According to concrete studies, replacing part of the ISF with some RSF is a good solution for waste tire disposal. Using RSF can produce similar advantages to using ISF on concrete [23].
Furthermore, 1.5 percent of RSF adds roughly 35 percent to construction budgets, and 1.5 percent of ISF contributes more than 50 percent [24]. In addition, the contribution of fiber to carbon emissions in concrete ranged from 15% (1.5 percent RSF) to 40% (1.5 percent ISF) [24]. According to a prior study, the RSF performs similarly to the ISF in enhancing concrete splitting and flexural strengths [25,26,27,28,29][25][26][27][28][29].

2. The Properties of RSF-Incorporated Concrete

The mechanical properties of the concrete incorporating raw steel fibers recovered from waste tires (RSFs) of previous research were demonstrated. Fundamentally, concrete is defined as a mixture of binder, fine aggregates, coarse aggregates, and water. Sometimes, additives or admixtures are added to obtain specific qualities. Hardened concrete’s three main mechanical properties are compressive strength, split tensile strength, and flexure strength. For percent comparison, the properties of the plain cement concrete of the concerned study were taken as a reference. In some cases, it is challenging to compare the strength properties of different research studies due to differences in the scales used by authors for adding the fibers to the concrete. This becomes very difficult when the complete set of parameters of the fibers is not provided for converting their weight to volume or for interconversion among different scales of adding the fibers. To deal with this issue, the strength properties are compared separately for concrete containing RSF’s content by volume fraction of concrete, RSF’s content added by weight fraction of the concrete, and RSF’s content added in kg/m3. For this purpose, the strength properties of each category were compared individually, and the effect of variation of fiber was reported. In addition, an optimized dose for RSF was demonstrated for each of the RSF’s contents: fiber by volume or weight of concrete or in kg/m3. Most previous research has assumed that the density of RSF is the same as that of industrial steel fibers (ISFs), presuming that the RSF is free of impurities/tire particles or that impurities/tire particles have been separated from the RSF before mixing with concrete. However, recycled steel fibers from waste tires (RSF) were utilized in some studies without separating impurities/tire particles [25]. As a result, the average density of RSF with impurities/tire particles was determined to be less (3014 kg/m3) than the industrial steel fibers (7200 kg/m3) [25]. In addition, it found that the density of the RSF varies from that of the commonly used density of steel, 7850 kg/m3, for inter-conversion of steel weight and volume [63][30]. Therefore, it becomes hard to calculate the exact amount of RSF by volume fraction of concrete if the RSF fibers are added to concrete in kg/m3, and the density of the fibers is not given. Therefore, the properties of the effect of the addition of RSF in kg/m3 are discussed in a separate section.

2.1. Compressive Strength of RSF Concrete

2.1.1. Compressive Strength of Concrete Containing RSF by its Volume Fraction

The compressive strengths (CS) of the concrete reinforced with raw steel fibers recovered from waste tires (RSF) were collected. The percent comparison of the CS is demonstrated in Figure 41. For percent comparison, the CS of the plain concrete of the concerned study was taken as a reference (equivalent to 100%). Additionally, the percent rise or decrease in the CS of RSF concrete is presented.
Figure 41.
Percentage comparison of the compressive strengths of concrete containing RSF by volume fraction.
The amount of fibers for specimens is given on the right axis of the graph, while an increase or decrease in CS to plain concrete is shown on the left y-axis of the graph. The red squares show the percentages of the fibers at which the CS was less than the PC. The rest of the different colors of the boxes represent a specific range of RSF doses used. The number of specimens is shown on the x-axis of the graph. Specimens 1 to 10 contain RSF from 0.13% to 0.30%. For specimen numbers 2, 3, 4, and 8, the CS decreased 18%, 11%, 15%, and 11%, respectively, from adding 0.15%, 0.19%, 0.19%, and 0.26% of RSF by volume fraction of concrete. The CS increased for specimens 1, 5, 6, 7, 9, and 10 by 3%, 13%, 20%, 23%, 12%, and 5% respectively. For RSF range from 0.46% to 0.50%, no significant decrease was reported except for specimens 12, 15, and 24, as shown in Figure 41. Moreover, contradictory results were reported for specimen numbers 12, 13, and 14, incorporating the same amount of 0.46% RSF. The possible reason is that for specimens 13 and 14, the authors used a planetary concrete mixer; thus, good fiber dispersion and better mixing were obtained [48,66][31][32]. On the other hand, better mixing was not achieved for specimen 12 due to the traditional mixing method [41][33] used. At 0.50% of RSF, all researchers found no improvement or slight increase in CS of RSF specimens 15 to 29 except a slight decrease of 3% and 4% for specimens 15 and 24, respectively. Only a few studies (specimens 30, 31, 32, and 33) tested the CS of the specimen incorporating RSF from 0.60% to 0.80% by volume fraction of concrete. Two studies reported increases of 2% and 15% for the incorporation of 0.75% RSF. At the same time, decreases of 25% and 13% were reported for specimens 30 and 33, respectively, at 0.60% RSF and 0.80% RSF, respectively. A rise in CS of the RSF specimens was reported at 1% RSF for specimens 34 to 48, except for specimens 34 and 39. Specimens 34 and 39 showed a decline of 8% and 1% in CS at 1% RSF. No decline in CS was reported by any studies for specimens 49 to 64 for concrete incorporating RSF from 1.25% to 5% RSF by volume fraction. The data given in Figure 41 is shown in tabular format for a specific range of RSF content, along with the diameters and lengths of the fibers. This can help conclude the influential factors for the increase and decrease in CS for the exact content of the RSF. The specimens incorporating RSF content of less than 1% are demonstrated in Table 21. The possible impact of the diameters and lengths can be described by comparing the results of the compressive strengths for the exact content of fibers in the context of RSF parameters. When advanced mixing techniques were not used, the concrete declined in compressive strength for short RSF (length was less than 31 mm) by more than 0.13%. A decrease in CS was reported for specimens 2, 3, 4, and 8, incorporating 0.13% to 0.26% RSF, by Aiello et al. [44][34]. The CS increased when RSF exceeded 0.19% for the exact dimension of the raw steel fibers, where special techniques were applied to control the fibers’ dispersion and homogeneity of the mix. Increases of 20%, 23%, and 12% in the CS of specimens six, seven, and nine for adding RSF of 0.23%, 0.23%, and 0.26%, respectively, could be correlated with using a planetary mixer and 0.17% more plasticizer.
Table 21. Results of compressive strength’s percent comparison and fiber properties of the concrete encompassing less than 1% of RSF by its volume fraction.
][47] and Shah et al. [60][48] noted increases of 2%, 5%, 4%, and 6% in CS for 0.10%, 0.25%, 0.25%, and 0.50% of RSF, respectively.
Table 43.
Results of compressive strength’s percent comparison and fiber properties of the concrete with RSF by its weight.
[48] for 0.50% RSF. The possible reason for improved CS could be better workability and uniformity achieved with reduced length (1 mm less) and diameters of the RSF utilized by Shah et al. [60][48]. A reduction in CS at 0.50% and 0.75% RSF can be associated with decreased workability due to excessive RSF within the same mix. The CS increased by 1.5% by incorporating 0.39% RSF in concrete by weight fraction. All authors reported an increase in CS by using 1% and 2% RSF except Shah et al. [60][48], who reported a 10% decline for specimen 7. Shah et al. [63][30] reported the same rhythm of decrease when RSF exceeded 0.25% of the concrete weight. At the same time, an increase in CS was reported for specimens 8 to 12, even for RSF having diameters of 100.16 mm. The maximum enhancement of 68 percent was noted in CS with 2% RSF, 29 mm long, and 0.20 mm thick [58][44]. Interestingly, Graeff et al. [59][50] and Gul et al. [46][49] reported an increase in CS for RSF with a minimum length of 13 mm and 7.62 mm, respectively. It can be deduced that an increase in CS can be achieved for the exact content of the RSF for varying dimensions of RSF. More RSF content can be used for small-size RSF compared to large-size RSF for achieving the same CS. For 2.5% RSF, the CS improved by 34% and 37% for specimens 13 and 14, while at 3% RSF, the CS decreased compared to 2.5% RSF concrete. A similar trend of decline in CS was reported for 3.5% RSF. However, it can be noticed that the short-length fibers could sustain a CS somewhat higher than the plain concrete at 2.5% RSF, while at 3% RSF, both types of RSF showed a decline in CS. Therefore, maximum strength can be attained for 2% to 3% RSF by keeping the diameters of the fibers below 7.62 mm for traditional concrete mixers. A decline of 17% and 7% were reported in CS for specimens 19 and 20 when diameters of RSF were 100.16 mm and 7.62 mm, respectively. Short RSF showed less decline in CS. Interesting results are reported when 4% and 6% RSF had reduced diameters and lengths added to concrete. It was investigated that the highest increase of 78% in CS was achieved using 29 mm long RSF, which had a diameter of 0.20 mm. When the same RSF (29 mm long and 0.20 thick) was increased from 4% by weight of concrete to 6%, the increase in CS reduced from 78% to 67% [58][44]. An increase of 22% was noted by Graeff [59][50] when incorporating 6% of RSF with 13 mm length and 0.20 mm thickness. On average, no decline of CS was reported for 4% to 6% of RSF with a length of 13 mm to 29 mm and a diameter of 0.20 mm.

2.1.3. Compressive Strength of Concrete Using RSF Content in Kg/m3

The compressive strengths (CS) of the concrete, containing raw steel fibers recovered from waste tires (RSF) in kg/m3, are demonstrated in Figure 74. The percentages of fibers where the CS was less than the PC are displayed in the red squares. A reduction in CS is noted for specimens 2, 3, 4, 5, 6, and 9 having RSF less than or equal to 20 kg/m3. However, a 3.2 percent increase was reported for specimen seven, incorporating 30 kg/m3 RSF. Specimens reinforced with RSF of 30 kg/m3 to 60 kg/m3 showed a considerable rise in CS. Specimens 9 and 11 showed a decrease of about 1% and 8% in CS at 40 kg/m3 and 60 kg/m3, respectively. The data given in Figure 74 is presented in tabular format along with the length and diameters of the fibers used to assess the effect of the dimension of the RSF on compressive strength (CS) for the same amount of RSF.
Figure 74.
Percentage comparison of the compressive strengths of concrete containing RSF in kg/m
3
.
The given data are shown for a particular range of RSF content in Table 54, along with the diameter and length of the fibers. For 5 to 30 kg/m3 of RSF, no significant increase (more than three percent) in CS of concrete was reported except for a decline in CS. A decline of 2% is noted in the CS of concrete for average RSF content of 15.71 kg/m3 (5 to 20 kg/m3). This is noticed for 50 mm to 52 mm-long fiber, whose average diameter ranges from 0.30 mm to 1.40 mm. Therefore, it can be suggested that the RSF content below 20 kg/m3 cannot help to improve the CS of the concrete.
Table 54.
Results of compressive strength’s percent comparison and fiber properties of the concrete with RSF in kg/m
3
.
. An inevitable decline was reported in STS when the content of RSF was lower than 1%. However, some specimens showed a substantial increase in STS even for RSF of less than 1%. The specimens 1, 5, 9, 10, and 11 incorporating RSF of 0.20%, 0.46%, 0.50%, 0.50%, and 0.80%, respectively, showed a decrease of 13%, 10%, 14%, 26%, and 10%, respectively, in the corresponding STS. On the other hand, specimens 2, 3, 4, 6, 7, and 8, containing 0.23%, 0.30%, 0.40%, 0.46%, 0.50%, and 0.50%, respectively, showed substantial improvement of 16%, 4%, 3%, 9%, 18%, and 43%, in respective STS. No improvement in STS was reported for specimen 11 at 0.60% RSF. A decline of 10% in STS was noted only for specimen 13 at 1% of RSF. For more than 0.60% of RSF, no decline was reported. The authors noted a significant improvement in STS by adding 1%, 1.50%, 2%, 3%, 4%, and 5% of RSF, as reflected in Figure 85. In addition, the STS of the plain concrete of the concerned study is taken as a reference (equivalent to 100%) to assess the percent rise or decrease in the STS of RSF concrete.
Figure 85.
Percentage comparison of split tensile strength of concrete containing RSF by volume fraction.
A certain decline was reported in STS when the content of RSF was lower than 1%. However, some specimens showed a substantial increase in STS even for RSF of less than 1%. The specimens 1, 5, 9, 10, and 11 incorporating RSF of 0.20%, 0.46%, 0.50%, 0.50%, and 0.80%, respectively, showed a decrease of 13%, 10%, 14%, 26%, and 10%, respectively, in the corresponding STS. On the other hand, specimens 2, 3, 4, 6, 7, and 8, containing 0.23%, 0.30%, 0.40%, 0.46%, 0.50%, and 0.50%, respectively, showed a substantial improvement of 16%, 4%, 3%, 9%, 18%, and 43%, in respective STS. No improvement in STS was reported for specimen 11 at 0.60% RSF. A decline of 10% in STS was noted only for specimen 13 at 1% of RSF. For more than 0.60% of RSF, no decline was reported. The authors noted a significant improvement in STS by adding 1%, 1.50%, 2%, 3%, 4%, and 5% of RSF, as reflected in Figure 85. The data regarding the STS of RSF-incorporated concrete and the properties of RSF used in each study are in Table 65. This can help analyze the factors that can affect the split tensile strengths of concrete for the same percentage of RSF. For example, the effect of the fibers’ diameter and length can easily be observed on the STS of the concrete incorporating RSF ranging from 0.20% to 0.50%. Furthermore, it can be observed that a significant decrease, 13%, was noticed even for lengthy RSF at 0.20% [37][35] when traditional mixers were used.
Table 65. Results of split tensile strength’s percent comparison and fiber properties of the concrete incorporating RSF by its volume fraction.

References

  1. Revuelta, D.; Carballosa, P.; García Calvo, J.L.; Pedrosa, F. Residual strength and drying behavior of concrete reinforced with recycled steel fiber from tires. Materials 2021, 14, 6111.
  2. Qin, X.; Kaewunruen, S. Environment-friendly recycled steel fibre reinforced concrete. Constr. Build. Mater. 2022, 327, 126967.
  3. Mohajerani, A.; Burnett, L.; Smith, J.V.; Markovski, S.; Rodwell, G.; Rahman, M.T.; Kurmus, H.; Mirzababaei, M.; Arulrajah, A.; Horpibulsuk, S.; et al. Recycling waste rubber tyres in construction materials and associated environmental considerations: A review. Resour. Conserv. Recycl. 2020, 155, 104679.
  4. ETRMA. European Tyre and Rubber Industry Statistics 2021; ETRMA: Brussels, Belgium, 2021.
  5. European Council. Council Directive 1999/31/EC of April 1999 on the Landfill of Waste. Off. J. Eur. Communities 1999, L182, 1–9.
  6. ETRMA. ELT-Management-Figures-2017-vf; ETRMA: Brussels, Belgium, 2019.
  7. Binglin, Y. Research status and prospect of waste steel fiber reinforced concrete. In Proceedings of the E3S Web of Conferences, Hangzhou, China, 13–15 August 2021; Volume 248.
  8. Pilakoutas, K.; Neocleous, K.; Tlemat, H. Reuse of tyre steel fibres as concrete reinforcement. Eng. Sustain. 2004, 157, 131–138.
  9. Gupta, T.; Sharma, R.K.; Chaudhary, S. Impact resistance of concrete containing waste rubber fiber and silica fume. Int. J. Impact Eng. 2015, 83, 76–87.
  10. ETRMA. ETRMA-2018. End of Life Tyres Management-Europe-Status; ETRMA: Brussels, Belgium, 2018.
  11. Bulei, C.; Todor, M.P.; Heput, T.; Kiss, I. Directions for material recovery of used tires and their use in the production of new products intended for the industry of civil construction and pavements. Proc. IOP Conf. Ser. Mater. Sci. Eng. 2018, 294, 012064.
  12. Liew, K.M.; Akbar, A. The recent progress of recycled steel fiber reinforced concrete. Constr. Build. Mater. 2020, 232, 117232.
  13. Scott, E. End-of-life Tyre REPORT, European Tyre & Rubber Manufacturers Association, 2015. Resour. Conserv. Recycl. 2015, 3.
  14. Evans, R. The Composition of a Tyre: Typical Components Creating Markets for Recycled Resources. Heliyon 2006, 5, e02139.
  15. Shannag, M.J. Characteristics of lightweight concrete containing mineral admixtures. Constr. Build. Mater. 2011, 25, 658–662.
  16. Atoyebi, O.D.; Odeyemi, S.O.; Orama, J.A. Experimental data on the splitting tensile strength of bamboo reinforced lateritic concrete using different culm sizes. Data Br. 2018, 20, 1960–1964.
  17. Zia, A.; Ali, M. Behavior of fiber reinforced concrete for controlling the rate of cracking in canal-lining. Constr. Build. Mater. 2017, 155, 726–739.
  18. Zia, A.; Khan, A.A. Effectiveness of bagasse ash for performance improvement of asphalt concrete pavements. SN Appl. Sci. 2021, 3, 502.
  19. Bedewi, N. Steel Fiber Reinforced Concrete Made with Fibers Extracted from Used Tyres. Master’s Thesis, Addis Ababa University, Addis Ababa, Ethiopia, 2009.
  20. Thomas, B.S.; Gupta, R.C.; Kalla, P.; Cseteneyi, L. Strength, abrasion and permeation characteristics of cement concrete containing discarded rubber fine aggregates. Constr. Build. Mater. 2014, 59, 204–212.
  21. Pawelska-Mazur, M.; Kaszynska, M. Mechanical performance and environmental assessment of sustainable concrete reinforced with recycled end-of-life tyre fibres. Materials 2021, 14, 256.
  22. Ayanoğlu, A.; Yumrutaş, R. Production of gasoline and diesel like fuels from waste tire oil by using catalytic pyrolysis. Energy 2016, 103, 456–468.
  23. Golpasand, G.B.; Farzam, M.; Shishvan, S.S. Behavior of recycled steel fiber reinforced concrete under uniaxial cyclic compression and biaxial tests. Constr. Build. Mater. 2020, 263, 120664.
  24. Mastali, M.; Dalvand, A.; Sattarifard, A.R.; Illikainen, M. Development of eco-efficient and cost-effective reinforced self-consolidation concretes with hybrid industrial/recycled steel fibers. Constr. Build. Mater. 2018, 166, 214–226.
  25. Samarakoon, S.M.S.M.K.; Ruben, P.; Wie Pedersen, J.; Evangelista, L. Mechanical performance of concrete made of steel fibers from tire waste. Case Stud. Constr. Mater. 2019, 11, e00259.
  26. Rutherford, S. Process for Recycling Vehicle Tires. U.S. Patent 5115983A, 26 May 1992.
  27. Bartolac, M.; Damjanovic, D.; Krolo, J.; Baricevic, A. Punching Shear Strength of Concrete Slabs Reinforced with Recycled Steel Fibres from Waste Tires. In Proceedings of the II International Conference on Concrete Sustainability ICCS16, Barselona, Spain, 13–15 June 2016.
  28. Peng, G.F.; Niu, X.J.; Long, Q.Q. Experimental study of strengthening and toughening for recycled steel fiber reinforced ultra-high performance concrete. Proc. Key Eng. Mater. 2015, 629–630, 104–111.
  29. Bjegovic, D.; Baricevic, A.; Lakusic, S.; Damjanovic, D.; Duvnjak, I. Positive interaction of industrial and recycled steel fibres in fibre reinforced concrete. J. Civ. Eng. Manag. 2013, 19, S50–S60.
  30. Centonze, G.; Leone, M.; Aiello, M.A. Steel fibers from waste tires as reinforcement in concrete: A mechanical characterization. Constr. Build. Mater. 2012, 36, 46–57.
  31. Caggiano, A.; Folino, P.; Lima, C.; Martinelli, E.; Pepe, M. On the mechanical response of Hybrid Fiber Reinforced Concrete with Recycled and Industrial Steel Fibers. Constr. Build. Mater. 2017, 147, 286–295.
  32. Shi, X.; Brescia-Norambuena, L.; Grasley, Z.; Hogancamp, J. Fracture Properties and Restrained Shrinkage Cracking Resistance of Cement Mortar Reinforced by Recycled Steel Fiber from Scrap Tires. Transp. Res. Rec. 2020, 2674, 581–590.
  33. Leone, M.; Centonze, G.; Colonna, D.; Micelli, F.; Aiello, M.A. Fiber-reinforced concrete with low content of recycled steel fiber: Shear behaviour. Constr. Build. Mater. 2018, 161, 141–155.
  34. Aiello, M.A.; Leuzzi, F.; Centonze, G.; Maffezzoli, A. Use of steel fibres recovered from waste tyres as reinforcement in concrete: Pull-out behaviour, compressive and flexural strength. Waste Manag. 2009, 29, 1960–1970.
  35. Rossli, S.A.; Ibrahim, I.S. Mechanical Properties of Recycled Steel Tire Fibres in Concrete; Technical Report; Universiti Teknologi Malaysia: Johor Bahru, Malaysia, 2012.
  36. Leone, M.; Centonze, G.; Colonna, D.; Micelli, F.; Aiello, M.A. Experimental Study on Bond Behavior in Fiber-Reinforced Concrete with Low Content of Recycled Steel Fiber. J. Mater. Civ. Eng. 2016, 28, 04016068.
  37. Groli, G.; Pérez Caldentey, A.; Soto, A.G. Cracking performance of SCC reinforced with recycled fibres—An experimental study. Struct. Concr. 2014, 15, 136–153.
  38. Siraj, N.; Kedir, N.S. Synthesis and characterization of pyrolised recycled steel fibers for use in reinforced concrete. Int. J. Eng. Sci. Manag. Res. 2017, 4, 21–32.
  39. Vistos, L.; Galladini, D.; Xargay, H.; Caggiano, A.; Folino, P.; Martinelli, E. Hybrid industrial/recycled SFRC: Experimental analysis and design. In Proceedings of Italian Concrete Days 2016. ICD 2016; Lecture Notes in Civil Engineering; Springer: Cham, Switzrland, 2018; Volume 10.
  40. Skarżyński, Ł.; Suchorzewski, J. Mechanical and fracture properties of concrete reinforced with recycled and industrial steel fibers using Digital Image Correlation technique and X-ray micro computed tomography. Constr. Build. Mater. 2018, 183, 283–299.
  41. Chen, M.; Si, H.; Fan, X.; Xuan, Y.; Zhang, M. Dynamic compressive behaviour of recycled tyre steel fibre reinforced concrete. Constr. Build. Mater. 2022, 316, 125896.
  42. Dorr, B.J.; Kanali, C.L.; Onchiri, R.O. Shear Performance of Recycled Tyres Steel Fibres Reinforced Lightweight Concrete Beam using Palm Kernel Shear as Partial Replacement of Coarse Aggregate. Int. J. Eng. Res. Technol. 2019, 12, 1818–1823.
  43. Mastali, M.; Dalvand, A. Fresh and Hardened Properties of Self-Compacting Concrete Reinforced with Hybrid Recycled Steel–Polypropylene Fiber. J. Mater. Civ. Eng. 2017, 29, 04017012.
  44. Younis, K.H. Mechanical Performance of Concrete Reinforced with Steel Fibres Extracted from Post-Consumer Tyres. In Proceedings of the 2nd Internationa Engineering Conference on Developments in Civil and Computer Applications, Erbil, Kudistan Region, Iraq, 20–21 February 2016.
  45. Köroğlu, M.A. Behavior of composite self-compacting concrete (SCC) reinforced with steel wires from waste tires. Rev. La Constr. 2019, 17, 484–498.
  46. Abdul Awal, A.S.M.; Kadir, M.A.A.; Yee, L.L.; Memon, N. Strength and Deformation Behaviour of Concrete Incorporating Steel Fibre from Recycled Tyre. In CIEC 2014, Proceedings of the International Civil and Infrastructure Engineering Conference; Springer: Singapore, 2015.
  47. Akhtar, T.; Ali, B.; Ben Kahla, N.; Kurda, R.; Rizwan, M.; Mohsin Javed, M.; Raza, A. Experimental investigation of eco-friendly high strength fiber-reinforced concrete developed with combined incorporation of tyre-steel fiber and fly ash. Constr. Build. Mater. 2022, 314, 125626.
  48. Shah, S.H.A.; Ali, B.; Ahmed, G.H.; Tirmazi, S.M.T.; El Ouni, M.H.; Hussain, I. Effect of recycled steel fibers on the mechanical strength and impact toughness of precast paving blocks. Case Stud. Constr. Mater. 2022, 16, e01025.
  49. Gul, A.; Alam, B.; Iqbal, M.J.; Ahmed, W.; Shahzada, K.; Javed, M.H.; Khan, E.A. Impact of Length and Percent Dosage of Recycled Steel Fibers on the Mechanical Properties of Concrete. Civ. Eng. J. 2021, 7, 1650–1666.
  50. Graeff, A.G.; Pilakoutas, K.; Neocleous, K.; Peres, M.V.N.N. Fatigue resistance and cracking mechanism of concrete pavements reinforced with recycled steel fibres recovered from post-consumer tyres. Eng. Struct. 2012, 45, 385–395.
  51. Sengul, O. Mechanical behavior of concretes containing waste steel fibers recovered from scrap tires. Constr. Build. Mater. 2016, 122, 649–658.
More