Next Article in Journal
Combined Gaussian Local Search and Enhanced Comprehensive Learning PSO Algorithm for Size and Shape Optimization of Truss Structures
Previous Article in Journal
A New Steel-Joint Precast Concrete Frame Structure: The Design, Key Construction Techniques, and Building Energy Efficiency
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review of the Durability-Related Features of Waste Tyre Rubber as a Partial Substitute for Natural Aggregate in Concrete

1
State Key Laboratory of Eco-Hydraulics in Northwest Arid Region, Xi’an University of Technology, Xi’an 710048, China
2
China West Construction North Co., Ltd., Xi’an 710119, China
*
Author to whom correspondence should be addressed.
Buildings 2022, 12(11), 1975; https://doi.org/10.3390/buildings12111975
Submission received: 27 September 2022 / Revised: 20 October 2022 / Accepted: 3 November 2022 / Published: 14 November 2022
(This article belongs to the Section Building Materials, and Repair & Renovation)

Abstract

:
As the number of discarded tyres continues to increase, causing serious environmental problems, the need of recycling the waste tyre rubber become extremely urgent in worldwide. Today, there is an increasing focus on recyclable materials. The reuse of waste tyre rubber in concrete contributes to sustainable development. In the past 10 years, numerous experiments on the recovery of rubber from waste tyres to produce concrete products have been conducted. In this review, we conclude the major achievement of rubberized concrete (RC) durability, discuss and analyse the influence of rubber replacement rates, replacement patterns, particle size and treatment methods. Results show that an increase in rubber content can improve the chloride penetration resistance, acid and sulphate attack resistance, freeze–thaw resistance, and alkali–silica reaction damage resistance of concrete, and the content of 5–20% has a significant improvement effect. Rubber replacing fine aggregate is the best scheme for durability, followed by cement and coarse aggregate. In addition, the recommended rubber particle size is 0–3 mm. However, the rubber particle has adverse effects on abrasion resistance, impermeability, water absorption resistance and carbonation resistance. The pre-treatment of rubber or the addition of supplementary cementitious materials are effective and viable ways of improving the durability of RC. Further research is needed on the long-term durability of RC, as well as on ductility, energy absorption, and thermal and corrosion resistance.

1. Introduction

Concrete is used extensively and widely in construction projects; however, it can be a constant drain on resources and energy. In the context of sustainable development, recycled concrete is receiving increasing attention [1]. Therefore, the researchers focus on producing concrete from sustainable materials. For example, the use of pozzolanic material to produce concrete has a low environmental impact while improving performance [2]. In addition, many types of waste can be recycled, including waste tyres [3], waste glass [4], waste lathes [5]. Fibres are often used to improve the performance of concrete [6]. Studies have shown that recycled steel fibres from waste tyres or lathes can be effective in enhancing the mechanical properties of concrete [7,8], and recycled rubber from waste tyres is also a promising sustainable building material.
Tyres have grown exponentially caused by the rapid population growth and economic development, and substantial waste tyres have been generated due to being beyond their service life [9,10]. Burning and landfilling are the two easiest and cheapest disposal methods. Burning rubber increases CO2 emissions and releases toxic gases, such as styrene and butadiene, which seriously pollute the natural environment [11,12,13]. Rubber is difficult to decompose, and landfills seriously pollute the soil and make the land lose its vitality [14]. In recent years, recycling waste tyre to replace the cement or natural aggregates in concrete is a promising research direction [15,16,17]. Studies have shown that recycled tyre rubber and steel fibres can improve the compressive, splitting tensile strength [18], dynamic mechanical properties [19], ductility, and energy absorption capacity of concrete [20]. The recycling of waste tyre in the above way prevents environmental pollution and reduces CO2 emissions. Additionally, the use of natural aggregate in civil engineering can also be reduced, which is beneficial to protecting the ecological environment. This approach is resource-saving and environmentally friendly.
Rubber can replace some of the natural aggregates mixed into concrete, and the study found that rubber can significantly improve the dynamic mechanical properties of concrete [21]. However, the bonding ability between rubber aggregate (RA) and cement paste is weak, and microscopic analysis found micro cracks in the interface transition zone (ITZ) between RA and cement paste [22]. Therefore, rubber will reduce the compressive strength, tensile strength, and other mechanical properties of concrete [23]. Rubber can affect the workability of concrete due to its strong hydrophobicity [24]. Researchers have investigated the possibility of enhancing the adhesion of rubber particles by means of adhesives, activators, or other additives to effectively enhance the mechanical properties [25]. The durability of rubberized concrete (RC) can also be compromised due to the incompatibility of rubber with various concrete components. It has been found that pre-treatment of rubber can reduce the adverse effects on durability [26]. The rubber particle surface has been pretreated or precoated with NaOH [27], KMnO4, and NaHSO3 [28], cement mortar [29], silane coupling agent (SCA) [30], acrylic acid and polyethylene glycol [31], ethanol and acetone, carboxylated styrene–butadiene rubber (CSBR) latex [32,33] acids (H2SO4, HCl, HNO3, CH3CHOOH) [34,35,36], CaCl2, and H2O2 [34], organic sulphur compounds [37], limestone powder (LP), silica fume (SF) [38,39], ethoxyline resin, and styrene-butadiene-type copolymer [40,41]. Table 1 summarizes some improvement methods for RC by modifying rubber particles.
Some reviews have been published on RC with the field of fresh and hardened properties [10,21,22,24], which are summarized in Table 2. However, in these reviews, there was a lack of detailed description of durability. Some important properties, including water absorption, chloride penetration resistance, carbonization resistance, alkali–silica reaction (ASR) damage resistance, acid resistance, and sulphate attack, were less analyzed. Furthermore, the RC reported in the review was limited to untreated crumb rubber (CR), and it did not cover reports on RC with physical and chemical treated CR. Additionally, the previous review did not analyze in detail the impact of rubber particle size, replacement rate and replacement pattern on durability. In addition, the existing reviews lack RC durability reports in recent years, and most of their coverage is research in 2018 and before. Therefore, it is necessary to carry out the latest comprehensive review of concrete containing waste rubber in order to better understand the durability of this important construction material.
This paper covers the study of various methods to enhance the durability of RC in the past 10 years and summarizes the rubber particles treatment methods to enhance the durability. The effects of rubber particle size, replacement rate, and surface treatment method on RC durability are compared. In accordance with the different durability, the survey results are summarized and compared with the control group. The review provides theoretical and experimental basis for dealing with waste tyre rubber in a green and environmental way.

2. Classification of Rubber Particles

Waste tyre rubber mainly comes from automobile and truck tyres. The rubber composition of the tyre is shown in Table 3. Although the composition of car and truck tyres is different in various countries, their natural and synthetic rubber content, which is approximately 40–45%, is similar [21]. Automobile tyres are produced more than truck tyres; hence, the rubber used in research mainly comes from automobile tyres to facilitate the waste tyre rubber recycling [42]. In previous research, six types of rubber have been used for recycling waste tyre rubber as building materials; the term name, particle size range, and type of replacement for natural aggregate are indicated in Table 4. Several commonly used tyre RAs with different sizes are shown in Figure 1. Rubber particles production is mainly achieved by using two methods, one is mechanical grinding, the other is cryogenic grinding [43], and then screening of rubber particles made from waste tyres to replace the fine or coarse natural aggregate [44,45]. Tyre rubber is ground into rubber powder or rubber ash to replace a certain percentage of cement, rubber fibre, and CR replace partial fine aggregate [46,47].

3. Abrasion Resistance

The abrasion resistance of concrete may be defined as its ability to resist being worn away by rubbing, which is important in the use of concrete structures [49,50]. There are two common methods for testing the abrasion resistance of concrete [51]. One is the “underwater method” specified by ASTM C1138 [52]. The other is the “ring method” named by the Chinese hydraulic concrete test code SL 352-2006 [53]. As shown in Table 5, previous experimental research has indicated that rubber that replaces natural aggregate in ordinary concrete reduces abrasion resistance. Bisht and Ramana [54] observed that as fresh RC vibrated, the rubber particles moved to the surface of the test piece and accumulated, causing the surface strength to decrease. Weak cementing ability between CR aggregate and cement paste made the RC prone to pulling and cracking, resulting in friction cracks. The increase of rubber content leads a decrease of RC abrasion resistance. Gupta et al. [49] observed that rubber ash mixed into the concrete resulting the compressive strength decrease, leading to poor abrasion resistance. An increase in cement matrix porosity and the weak ITZ between RA and cement paste take primary responsibility for decreased of RC abrasion resistance; the former is due to an increase in water–cement ratio (w/c) that ensures the workability of RC, whereas the latter is caused by the weak binding capacity of rubber particles [50]. Turki et al. [55] detected the inside of RC via scanning electron microscopy (SEM) and found that RA is easier to separate from cement matrix compared with natural aggregate. As shown in Figure 2, the siliceous aggregate is closely combined with the cement matrix; however, there are obvious cracks between RA and cement matrix. Some studies show that due to the weak bonding between RA and cement matrix, rubber particles are easy to rise to the surface during vibration, resulting in uneven distribution, which will reduce the abrasion resistance of concrete. In addition, poor ITZ will reduce the mechanical properties and affect the abrasion resistance [22]. Therefore, guaranteeing concrete strength, enhancing the binding capacity of RAs and reducing the RC porosity are necessary to ensure RC abrasion resistance. Adding supplementary cementitious materials (SCMs), for example ground-granulated blast-furnace slag (GGBFS), metakaolin (MK), SF, and fly ash (FA), into RC can effectively enhance the abrasion resistance; SCMs can enhance the concrete strength by filling the internal space. In addition, the additional hydration reaction strengthens the internal structure of the RC [51,56,57,58,59].
Precoating or pretreating rubber particle surface before mixing into concrete can also improve the concrete abrasion resistance. Due to the hydrophobic property of RA, it can cause poor adhesion between cement paste and RA, which greatly reduces the mechanical properties, and limit the use of RC in a wide range. He et al. [28] reported that treatment of CR with KMnO4 and NaHSO3 solutions could make the surface of rubber particles highly hydrophilic, thus improving the binding capacity. Segre and Joekes [27] indicated that using NaOH to treat RA can enhance the rubber-matrix adhesion. Figure 3 show the state of contact surface with and without NaOH pretreatment. After treatment with NaOH, the crack width of ITZ is narrow, and the RA is difficult to detach [65]. Onuaguluchi [38] found that the two-stage approach of precoating rubber particles with LP and mixing SF into RC has also achieved a positive effect on the abrasion resistance. After LP precoating, the rubber surface becomes hydrophilic and acts as the reaction area of early hydration products. SF can generate a pozzolanic reaction on a rubber surface. The accumulation of hydration products on rubber surface changes the surface morphology and improves the binding capacity of RA [38]. Good binding capacity and ductility can make RA limit the formation and development of cracks caused by friction and reduce the stress concentration at the crack tip; hence, they can effectively improve abrasion resistance [66].
As shown in Table 5, the shape, size, and gradation of rubber particles also affect abrasion resistance. Fibre rubber particles have better ability to enhance abrasion resistance compared with granular rubber particles, which may be due to the fibre-holding effect preserving the integrity of cement paste [49,61]. The small particle size is better than that of larger one. Gesoğlu et al. [62] described that the combination of RA with different particle sizes can produce a good synergistic effect, Sukontasukkul and Chaikaew [64] also observed the same phenomenon. Well-graded RA makes the inner part of RC considerably compact [63].

4. Water Absorption and Permeability

As shown in Table 6, as rubber content increases, water absorption also increases. This is mainly due to the increase of porosity after rubber is mixed into concrete [63]. The hydrophobic CR particle surface, during the mixing process, rubber particles trap air, resulting in increased internal pores between fine aggregate and CR, as shown in Figure 4 [67]. The light weight and low elastic modulus of RA cause the RC low compactness. It will be more difficult to fill pores and provide adequate compactness. Thus, the compaction of RC specimens becomes difficult and leads to an increase in internal porosity [61,66,68]. A high w/c results in an increase in porosity inside RC, because the RC needs a high mixing water content to ensure effective workability [69]. Measures are taken to reduce vibration time to prevent light weight rubber particles from floating and avoid separation during the vibration process. Consequently, porosity increases due to the lack of sufficient vibration compactness time. When cement paste hardens, many closed cavities are formed inside concrete, as shown in Figure 5. Concrete with 0% CR has a denser internal structure compared with concrete mixed with RA. In concrete with 4% and 5.5% CR, closed cavities are a good water storage bag, which provides the possibility of high water absorption. Figure 6 depicts that ITZ is a porous and weak structure due to the poor cementation ability of rubber particle surface; such structure easily causes RA separate from cement paste [69]. The compressive strength is significantly reduced with rubber addition, which easily produces micro cracks in hardened cement matrix. The cracks around the rubber and in the cement matrix connect the originally closed cavities to form a good water seepage and storage channel, which help to increased water absorption [70,71].
The influences of the rubber replacement pattern and ratio on the water absorption of RC are shown in Table 6. Youssf et al. [72] described that CR (1.18–2.36 mm) partially as the replacement of fine aggregate. When the CR content is less than 10%, the water absorption of RC decreases with the rubber content increases. However, when the CR content is more than 10%, the opposite results occur. A total of 0–10% of fine rubber particles can effectively fill the internal pores and reduce the internal porosity of concrete owing to the rubber particle filling effect. The external water can be prevented from infiltrating into the internal pores of concrete due to the hydrophobicity of rubber particles. However, when excessive rubber (10–20%) is added, the bonding capacity between the cement paste and RA is poor and the internal cracks of RC increase, leading to an increase in water absorption [57,80]. Foamed concrete with 2% and 4% rubber contents is shown in the SEM diagram in Figure 7. Figure 7a illustrates that rubber particles fill in permeability channels, blocking the entry passage of water and improving the waterproof performance of foam concrete. As shown in Figure 7b, excessive rubber content leads to a dispersed cement matrix in foamed concrete; the poor cementation increases permeability channels and water absorption [81]. Thomas and Chandra Gupta [57] found that using three different particle size mixtures of CR (rubber powder 0.6 mm (40%) + CR 0.8–2 mm (35%) + CR 2–4 mm (25%) had a better filling effect compared with using a single-sized rubber alone.
Rubber particles size also has great influence on the water absorption of concrete. Yilmaz and Degirmenci [82] incorporated 20% of three types of rubber particles with different particle sizes (0–0.25, 0.25–0.5, and 0.5–1.0 mm) into concrete, and the water absorption was 30%, 26%, and 24%. As the rubber particle size increased, the water absorption decreased slightly. Li et al. [77] found the same phenomenon. RC with fine RA needs substantial water to achieve the same workability as the control group; rubber particles with large particle size are easy to mix because the surface is smooth and spherical [71]. However, an excessively large particle size has a negative effect. Sukontasukkul and Tiamlom [76] reported that when CR No.6 (0.5–3.35 mm) was incorporated into concrete; as the blending amount increased, the water absorption increased. By contrast, CR No.26 (0–0.5 mm) incorporated into concrete reduced water absorption. The rubber surface formed bubbles due to the non-polar rubber particles surface, and the internal porosity of RC increased. The volume of the captured bubble was larger, and the RA size increased, as shown in Figure 8. Although the surface of CR No.26 also formed bubbles, the volume of the bubbles was small. These small bubbles formed a separate closed space and did not act as a water seepage channel. Moreover, CR No.26 had a good filling effect, and can effectively fill the pores and reduce the porosity in RC. Similar results were obtained with rubber powder (0.1–1 mm) as a replacement for partial cement in concrete [74].
The major approaches to reduce the water absorption are to strengthen the cementation ability between cement paste and RA, reduce the internal porosity, and reinforce the internal compactness of RC [83,84]. The rubber particles surface pretreated with NaOH can effectively strengthen the cementation ability and reduce the porosity due to the weak ITZ [79]. Adding SF to RC can effectively reduce water absorption. SF promotes hydration reaction, thereby increasing hydration production. Hydration products are filled in the pores amongst cement matrix, rubber, and aggregate, effectively reducing the internal porosity of concrete [38]. The addition of SF can react with the NaOH in the pores of ITZ, make the internal structure compact and strengthen the cementation ability [67]. Mixing FA, MK, or nanosilica into RC also produces the same effect as SF incorporation [59,75,80,85,86].
Permeability is an important factor affecting the durability of concrete, and it directly influences freeze–thaw resistance, chloride penetration, and acid and sulphate resistance. Permeability reflects the ability of fluid to pass through concrete, and the depth of water penetration is adopted as the standard to evaluate permeability. Table 7 indicates that as the rubber content increases, the permeability increases. The factors affecting the permeability of RC are similar to those of water absorption, which can be owed to the porosity increase of the internal structure and the increase of internal macro- and micro cracks [63]. Rubber particles mixed in concrete easily float in the cement matrix owing to the light weight and elastic properties of RA; this condition causes poor compaction of RC, resulting in increased porosity and water permeability in RC [73,87]. The weak cementation ability between cement matrix and RA leads to cracks around rubber particles [42]. Cracks around rubber particles and in cement make it easy for external water to penetrate concrete and transfer in various parts [54]. Su et al. [87] reported that rubber particles with small particle size (0–0.5 mm) can reduce permeability better than rubber particles with large particle size (0.5–3 mm) as are placement for fine aggregate. The better filling effect of small rubber particles in comparison with that of large particles causes RC to have a dense internal structure [88]. RA with good gradation and composed of different sizes of rubber particles is compact because fine rubber particles can effectively fill the gap formed by large rubber particles [87]. Si et al. [89] observed that rubber particles pre-treated with NaOH as a replacement for fine aggregate in SCC can effectively reduce permeability. RC needs a high w/c to achieve the required workability; a high w/c ratio causes a bad influence on permeability resistance [63]. Gupta et al. [90] observed that when the temperature exceeded 150 °C, a pore network would be formed in RC due to rubber fibre decomposition. The depth of water penetration reached 49 mm, which resulted in the high water penetration of RC. When RC has impermeability requirements, it is unsuitable for working in a high-temperature environment.

5. Freeze–Thaw Resistance

As shown in Table 8, RAs replace natural aggregates in concrete can enhance the freeze–thaw resistance. As the rubber content increases, the freeze–thaw resistance becomes better. Turgut and Yesilata [91] used CR (particle size range: 0.075–4.75 mm) to replace fine aggregate in approximately 10% volume; after a freeze–thaw cycle test, it was 66.48% lower in the mass loss compared to the control group, the freeze–thaw resistance reached best with 30% CR content. Zhang et al. [89] pretreated CR with NaOH and Na2SiO3 mixed solution, it was found that using the treated CR instead of fine aggregate significantly improved the freeze–thaw resistance. When the replacement ratio was 5%, 10%, 15%, and 20%, the mass loss was reduced by 43.48%, 84.78%, 54.35%, and 71.74%, respectively. Rubber can be regarded as an air-entraining agent, making concrete achieve the same freeze–thaw resistance as air-entrained concrete [92]. Water enters into concrete under the action of hydraulic pressure and freezes at low temperature, expanding in volume, which is the major reason for the freeze–thaw damage [93,94]. The hydrophobic nature and small jagged shapes of rubber particle surface, as shown in Figure 9, tend to cause it to entrap air during the mixing process of RC, and an increase in ineffective pores improves the freeze–thaw resistance. When water is frozen and expanded in cement, effective pores can play the role of stress absorption mechanism, providing buffer space for water volume expansion [71]. The low rubber particles elastic modulus can absorb the volume expansion stress of water [95,96,97].
Table 8. Influence of rubber on the concrete freeze–thaw resistance.
Table 8. Influence of rubber on the concrete freeze–thaw resistance.
ReferenceTreatment MethodRA Type and Size (mm)RA Replacement Ratio (%)Replacement PatternConcrete TypeThe Ways of Freezing and ThawingFreeze–Thaw ResistanceML a, RDME, SL, DF, FTRG Compared to the Control Type
Turgut and Yesilata [91]UntreatedCR: 0.075–4.7510, 20, 30, 40, 50, 60, 70 by volumeFAGOCFreezing (−9 °C) and thawing (25 °C) in waterb66.48% ↓, 66.11% ↓, 71.30% ↓, 68.52% ↓, 59.63% ↓, 69.26% ↓, 70.37% ↓ (ML)
Gonen [98]UntreatedCR: 0.125–10.5, 1, 2, 4 by volumeFAGOCFreezing and thawing in 3% NaCl solution freezing34.55% ↓, 40.91% ↓, 73.64% ↓, 87.27% ↓ (ML)
UntreatedCR: 0.25–20.5, 1, 2, 4 by volumeFAGOCFreezing and thawing in 3% NaCl solution freezing38.18% ↓, 61.82% ↓, 70.91% ↓, 81.82% ↓ (ML)
Al-Akhras and Smadi [99]UntreatedRubber ash: 0–0.155, 10 by weightFAGOCFreezing in air and thawing in water211.11% ↑, 400.00% ↑ (DF)
Topçu and Bilir [100]UntreatedCR: 0–43.7, 7.3, 10.98 by weightFAGOCFreezing and thawing in water37.36% ↑, 114.82% ↑, 170.27% ↑ (SL)
Zhu et al. [93]UntreatedCR: 0.250.26, 0.5, 1.6 by weightCAG and FAGOCFreezing (−15 °C) and thawing (6 °C) in water75.00% ↑, 100.00% ↑, 87.50% ↑ (FTRG)
Paine [92]UntreatedCR: 0.5–1.52, 4, 6 by volumeFAGOCFreezing and thawing in water86.36% ↑, 75.00% ↑, 68.18% ↑ (RDME)
Liu et al. [40]Pre-coating with synthetic resinCR: 2.0–4.05 by volumeFAGOCFreezing (−16 °C) and thawing (3 °C) in water1.2% ↓ (SL)
Si et al. [101]Pre-treating with NaOH CR: 1.44–2.8315, 25, 35, 50 by volumeFAGOCFreezing (−18 °C) in air and thawing (4 °C) in water4.07% ↑, 0.19% ↑, 1.74% ↓, 3.10% ↓ (RDME)
Pham et al. [41]Pre-coating with styrene-butadiene-type copolymerCR: 0.65–330 by volumeFAGmortarFreezing (−18 °C) and thawing (4 °C) in water58.33% ↑ (RDME)
Wang et al. [102]Pre-treating with NaOHCR: 0.6–2.810, 15 by volumeFAGOCFreezing and thawing in water8.02% ↑, 0.10% ↑ (DF)
Zhang et al. [103]Pre-treating with NaOH and Na2SiO3CR: 0.15–4.755, 10, 15, 20 by volumeFAGOCFreezing (−20 °C) and thawing (5 °C) in water43.48% ↓, 84.78% ↓, 54.35% ↓, 71.74% ↓ (ML)
a Abbreviations: ML: Mass loss, RDME: Relative dynamic modulus of elasticity, SL: strength loss, DF: Durability factor, FTRG: Freezing–thawing resistance grade. b Evaluate standard change compared to the control type (%), corresponding to RA replacement ratio. Increase: ↑, Decrease: ↓.
Figure 9. CR particle at 80× magnification [104].
Figure 9. CR particle at 80× magnification [104].
Buildings 12 01975 g009
According to the Table 8, it can conduct that the concrete with smaller particle size CR have higher freeze–thaw resistance. Gesoğlu et al. [62] used rubber with different particle sizes (<0.5, 1–1.5 and 1.5–2 mm) to replace fine aggregate and found that the best frost resistance occurs when the rubber particle size <0.5 mm. Zhu et al. [93] reported that a 60-mesh (0.25 mm) rubber particle achieved the best freeze–thaw resistance. In case of maintaining rubber content, the larger the rubber particle size, the farther the distance amongst the rubber particles, such relationship is unbeneficial to the freeze–thaw resistance. Small-sized rubber particles has a high specific surface area, which is conducive to air entrance. These micro-scale pools can serve as a buffer space for water freezing in cement [62]. Large-sized rubber particles are unbeneficial to the combination of rubber and cement paste, resulting in substantial cracks inside concrete, which make it easy for water to penetrate concrete [103]. Additives are often used to enhance the freeze–thaw resistance of concrete [105]. In addition, pretreatment of rubber can also be effective in improving freeze–thaw resistance. Using NaOH to treat the RA and then incorporating it into concrete had better freeze–thaw resistance in comparison with RC than RA without NaOH treatment [101]. Liu et al. [40] used six modifiers to precoat the rubber particles’ surface and found that the precoating of synthetic resin could minimize the degree of mechanical strength reduction and achieve good freeze–thaw resistance. Precoating rubber particle with styrene-butadiene-type copolymer can better the bonding ability and effectively enhance the RC freeze–thaw resistance [41]. The negative effect of RA on concrete strength makes RC suitable for areas which do not have high strength requirements but need high freeze resistance, such as roads that need to undergo repeated freeze–thaw cycles in alpine regions [106]. However, the rubber mixed into SCC weakens the freeze–thaw resistance [100].

6. Acid and Sulphate Resistance

Acid solutions, such as H2SO4 or hydrochloric acid (HCl), penetrate concrete and react with Ca(OH)2 and calcium silicate hydrate (CSH) gel, causing physical and chemical damages [107]. As shown in Table 9, the rubber incorporated into concrete can enhance the resistance to H2SO4 attack. The resistance to H2SO4 corrosion of RC effected by different rubber contents are shown in Figure 10 and evaluated by changes in compressive strength after soaking in 3% H2SO4 solution. In the early 7 days of immersion, the initial concrete internal hydration reaction, causing the formation of ettringite and CSH, which can increase the compressive strength. After 180 days of immersion, a minimum compressive strength loss occurred at 5.5% CR, which indicated that the RA relieved the structural damage of concrete caused by H2SO4 corrosion [108]. Gupta et al. [107] found that rubber reacted with acid can enhance the rubber particles surface adhesion ability, which could exert the rubber bridge effect to suppress the generation of internal cracks and the separation of materials. Moreover, the rubber could alleviate the compressive stress generated by the expansion of the ettringite existing inside the cracks around the rubber particles [109]. HCl is less aggressive than H2SO4, as shown in Table 9. Gupta et al. [107] demonstrated that rubber incorporated into concrete enhanced resistance to HCl attack, and the combination of rubber fibre and rubber powder had improved results. Therefore, RC could be applied to an environment where acid corrosion occurs. Although high rubber content produces high resistance to H2SO4 attack, it can result in a reduction in the compressive strength of the concrete. Thomas et al. [110] observed that 4% CR satisfied serviceable strength and had good resistance to H2SO4 attack.
However, Rahimi et al. [111] reported that rubber incorporated into concrete reduced the resistance to H2SO4 attack, which is contrary to conclusions of previous research. Azevedo et al. [86] found that adding rubber to HPC greatly reduced the H2SO4 resistance. This phenomenon may be due to the increase in concrete permeability with the addition of rubber particles [112]. Using rubber fibre or rubber particles mixed with polyethylene terephthalate as fine aggregate could substantially increase the resistance to H2SO4 attack [111].
Table 9. Influence of rubber on the concrete acid and sulphate resistance.
Table 9. Influence of rubber on the concrete acid and sulphate resistance.
ReferenceTreatment MethodRA Type and Size (mm)RA Replacement Ratio (%)Replacement PatternConcrete TypeHA a, SA, and S Resistance ML, EX, and ACC Compared to the Control Type
Thomas et al. [110]UntreatedCR: 2–4 (25%) + 0.8–2 (35%) and rubber powder: 0.6 (40%)5, 10, 15, 20 by volumeFAGOCb(SA)2% ↓, 10.47% ↓, 11.41% ↓, 14.82% ↓ (ML)
Thomas et al. [113]UntreatedCR: 2–4 (25%) + 0.8–2 (35%) and rubber powder: 0.6 (40%)5, 10, 15, 20 by volumeFAGOC↑(SA)0.49% ↑, 0.61% ↓, 21.17% ↓, 22.26% ↓ (ML)
Azevedo et al. [86]UntreatedCR: 1–2.45, 10, 15 by weightFAGHPC↓(SA)7.41% ↑, 33.33% ↑, 50% ↑ (ML)
Gupta et al. [107]UntreatedRubber powder: 0.15–1.95, 10, 15, 20 by volumeFAGOC↑(HA)1.11% ↓, 3.33% ↓, 4.67% ↓, 5.00% ↓, (ML)
Particle size effectRubber fibres: width of 2–5, length up to 20, and rubber powder (10%)5, 10, 15, 20, 25 by volumeFAGOC↑(HA)2.25% ↓, 4.49% ↓, 5.62% ↓, 4.50% ↓, 3.37% ↓ (ML)
Hunag et al. [73]UntreatedCR: 0–4.710, 20, 30, 40 by weightFAGLow-strength lightweight aggregate concrete↑(S)7.02% ↓, 3.51% ↓, 1.75% ↓, 22.81% ↓ (ML)
Onuaguluchi and Banthia [114]UntreatedCR: 0.2–210, 15 by volumeFAGOC ↑(S)56.84% ↓, 52.63% ↓ (EX)
Thomas et al. [115]UntreatedCR: 2–4 (25%) + 0.8–2 (35%) and rubber powder: 0.6 (40%)5, 10, 15, 20 by volumeFAGHSC↓(S)18.78% ↑, 74.11% ↑, 113.71% ↑, 154.31% ↑ (ML)
Liu et al. [40]Untreated CR: 0.2–45, 10, 15, 20 by volumeFAGOC ↑(S)0.62% ↑, 1.35% ↑, 1.66% ↑, 2.39% ↑ (ACC)
Pre-coating with synthetic resinCR: 2–45 by volumeFAGOC ↑(S)1.84% ↑ (ACC)
Li et al. [116]Pre-treating with NaOHCR: 0.85–210 by volumeFAGOC↑(S)31.01% ↓ (ML)
a Abbreviations: ML: Mass loss, EX: Expansion, ACC: Anti-corrosion coefficient, HA: Hydrochloric acid, SA: Sulphuric acid, S: Sulphate. b Evaluate standard change compared to the control type (%), corresponding to RA replacement ratio. Increase: ↑, Decrease: ↓.
Sulphate reacts with hydration products to produce ettringite, which expands in volume to generate internal compressive stress that forms cracks inside concrete and causes the internal structure of cement to peel off and cause structural damage [108]. Table 9 shows that using RA to replace part of natural aggregate can enhance the resistance to sulphate attack of concrete. In accordance with previous experience, the content and size of rubber particles affect sulphate resistance. Using CR (0–4.7 mm) to replaced fine aggregate partially with different substitution levels (10%, 20%, 30%, and 40% by volume); after sulphate attack, mass loss was reduced by 7.02%, 3.51%, 1.75%, and 22.81%, respectively, compared with the control group [73]. With the rubber content increases, the ability of concrete to resist sulfate attack increases. Figure 11 depicts the resistance to sulphate attack in self-consolidating rubberized concrete (SCRC) effected by particle sizes (#30: 0.6 mm, #50: 0.3 mm) and different rubber contents. The weight loss of RC is less than that of the control group in the rubber content of 5–10%. When the content exceeds 10%, the opposite situation occurs. The best resistance to sulphate attack is achieved when 5% rubber powder (passes through a #30 sieve) is incorporated [117].
The incorporation of rubber into concrete can improve the resistance to sulphate attack, it can be attributed to the fact that compared with concrete without RA, the RC has better deformation ability, which can relieve the internal expansion stress caused by sulphate attack. Moreover, rubber particles can absorb expansion stress, prevent cracks from developing, and reduce the width of the cracks, thereby reducing the intrusion of sulphate ions into the interior of concrete [116]. The incorporation of fine rubber particles increases the micro-sized void content, which can act as the buffer space for the volume expansion of gypsum and ettringite [114]. Onuaguluchi and Banthia [114] reported that the RC incorporated with 10% SF gains great resistance to sulphate attack. This condition can be attributed to the reduced aluminate/Ca(OH)2 in the matrix and the compact interior of the RC; the intrusion of sulphate ions becomes difficult. Li et al. [116] found that the pretreatment with NaOH solution improved the surface roughness and hydrophobicity of CR, increased the adhesion between CR and cement mortar, and enhanced the blocking effect on sulphate attack. Liu et al. [40] pre-coated rubber particles with six modifiers, and all of them made the anticorrosion coefficients of RC larger than 97%. Synthetic resin is the best modifier, considering the effects of precoating on compressive strength and freeze–thaw resistance. However, RA replaces part of natural aggregate and blends into HSC reduces the resistance to sulphate attack, which may be attributed to the increased HSC water absorption with RA incorporation, making the aggressive solutions enter the internal structure easily [115].

7. Chloride Penetration Resistance

The rubber content and particle size effect the resistance of RC to chloride ion penetration. Table 10 demonstrates that rubber incorporated into concrete as a substitution for fine aggregate can effectively improve the chloride ion attack resistance. The rubber content increases causing chloride ion permeability decreases, owing to the impervious property of RA and the ability to increase the length and sinuosity of capillary channels [77]. However, excessive rubber content causes negative effects. Gheni et al. [118] replaced cement with rubber fibre powder (particle size < 0.075 mm) in five ratios of 5%, 10%, 15%, 20%, and 25%. When the content was less than 20%, the charge passed decreased, as the rubber powder content increased. When the amount exceeded 20%, the charge passed greatly increased. Thomas et al. [110] observed that the optimum amount of CR to replace fine aggregate is 7.5% for chloride penetration resistance. This condition can be attributed to the lack of compactness when excessive RA are mixed into concrete, leading to significant chloride penetration; it is similar to the reason for water absorption and permeability of RC [69,75]. Gupta et al. [107] described that rubber fibre (width: 2–5 mm, length: 20 mm) with hydrophobic nature and increasing crookedness and tortuosity acts as a good blocking mechanism between cement and chloride ion. Due to the small particle size, rubber powder (0.15–1.9 mm) can fill the internal pores of concrete, which help to enhance the resistance to chloride ion attack [99]. However, Fernández-Ruiz et al. [119] found that the rubber and cement lacked good cementing ability. When small particles of rubber powder (0.063–0.6 mm) were used in place of cement in concrete, the gap between them became a seepage channel, and increased rubber powder content led the chloride permeability coefficient to increase. The same phenomenon was observed when rubber powder (0–0.3 mm) was incorporated into SCC [77].
As shown in Table 10, several experimental studies have attempted to treat rubber particle surface with chemicals, such as NaOH, SCA, or CSBR latex, to enhance the RA surface adhesion ability. Mineral admixtures, such as SF or FA, are incorporated into RC to reduce internal porosity, thereby enhancing the resistance to chloride ion penetration. SF can effectively fill the pores at the joint of aggregate and cement, making the interior of concrete highly compact [47,68,120,121]. Dong et al. [122] found that pretreated CR particles with SCA and precoated with cement paste, the chloride ion penetration resistance can be improved significantly. As shown in Figure 12, after pretreating with SCA, the coupling agent can attach well on the rubber surface, which can make the cement matrix adhere to rubber particle better. Guo et al. [39] used NaOH or SCA to pretreat the surface of rubber particles and precoated the rubber particles with normal cement, SF blended with cement, or cement plus Na2SiO3. All of these measures could greatly increase electrical resistivity, which meant low chloride permeability of RC. NaOH treatment could reduce ITZ porosity, and SF could refine internal pores. After treatment with SCA, the rubber surface became hydrophilic, and it is easier to form hydrogen bonds on a rubber surface, thereby benefiting the cementation between them. After treatment with CSBR, a chemical bond amongst rubber particles was formed. The enhanced interfacial adhesion prevented the rubber particles from floating due to vibration during the rubber-concrete pouring process. Compared with RC without pretreatment, rubber pretreated with SCA and CSBR could be more uniformly dispersed in concrete [33,123,124].
Table 10. Influence of rubber on the concrete chloride penetration resistance.
Table 10. Influence of rubber on the concrete chloride penetration resistance.
ReferenceTreatment MethodRA Type and Size (mm)RA Replacement Ratio (%)Replacement PatternConcrete TypeChloride Penetration ResistanceD a, CD, CP, ER Compared to the Control Type
Gheni et al. [118]UntreatedRubber fibre powder: <0.0755, 10, 15, 20, 25 by volumeCementOCb20% ↓, 40% ↓, 50% ↓, 75% ↓, 200% ↑ (CD)
Thomas et al. [110]UntreatedCR: 2–4 (25%) + 0.8–2 (35%) and rubber powder: 0.6 (40%)2.5, 5, 7.5, 10, 12.5, 15, 17.5, 20 by volumeFAGOC7.7% ↓, 7.7% ↓, 7.7% ↓, 0 ↑, 7.7% ↑, 23.1% ↑, 23.1% ↑, 30.8% ↑ (D)
Al-Akhras and Smadi [99]UntreatedRubber ash: 0.155, 10 by volumeFAGOC72.27% ↓, 81.33% ↓ (CP)
Fernández-Ruiz et al. [119]UntreatedRubber powder: 0.063–0.62.5, 5, 10 by volumeCementOC5.62% ↑, 9.68% ↑, 21.58% ↑ (CD)
Bravo and Brito [69]UntreatedCR: <11.25, 10, 15 by volumeFAGOC18.67% ↓, 7.33% ↓, 6.67% ↑ (CD)
Gupta et al. [107]Untreated Rubber powder: 0.15–1.95, 10, 15, 20 by volumeFAGOC7.32% ↓, 8.54% ↓, 14.63% ↓, 24.39% ↓ (CD)
Sagawa et al. [110]UntreatedCR: 1–310, 15, 20 by volumeFAGOC1.61% ↓, 3.46% ↓, 22.35% ↓ (CD)
Li et al. [77]UntreatedCR: 1–2, 0–0.330 by volumeFAGOC21.44% ↓, 12.10% ↓(CP)
Thomas et al. [125]UntreatedCR: 2–4 (25%) + 0.8–2 (35%) and rubber powder: 0.6 (40%)2.5, 5, 7.5, 10, 12.5, 15, 17.5, 20 by volumeFAGHSC6.25% ↓, 6.25% ↓, 6.25% ↓, 0, 12.5% ↑, 18.75% ↑, 18.75% ↑, 25% ↑ (D)
Hall and Najim [85]UntreatedCR: 2–644 by volumeFAG and CAGSCC152% ↑ (CD)
Oikonomou and Mavridou [123]UntreatedCR: 0.09–12.5, 5, 7.5, 10, 12.5, 15 by weightFAGOC14.22% ↓, 16.76% ↓,
25.43% ↓, 30.25% ↓,
35.18% ↓, 35.85% ↓ (CP)
Gesoğlu and Güneyisi [75]UntreatedCR: 0.2–35, 15, 25 by volumeFAGSCC9.09% ↑, 13.64% ↑,
40.91% ↑ (CP)
Dong et al. [122]Pre-treating with SCA and pre-coating with cement.CR: 0.6–4.7515, 30 by volumeFAGOC26.9% ↓,13% ↓ (CD)
Guo et al. [39]Pre-treating with NaOH.CR: 1.5–2.815, 25, 35, 50 by volumeFAGOC35.71 ↑, 50.00 ↑, 55.71 ↑, 51.43 ↑ (ER)
Pre-treating with SCA and pre-coating with cement.CR: 1.5–2.815 by volumeFAGOC57.41 ↑ (ER)
Pre-treating with NaOH then pre-coating with cement.CR: 1.5–2.815 by volumeFAGOC50.00 ↑ (ER)
Pre-treating with NaOH then pre-coating with SF and cement.CR: 1.5–2.815 by volumeFAGOC71.43 ↑ (ER)
Pre-treating with NaOH then pre-coating with Na2SiO3 and cement.CR: 1.5–2.815 by volumeFAGOC52.86 ↑ (ER)
Li et al. [33]Pre-treating with SCA and CSBR.Rubber powder: <0.65, 10, 15, 20, 30 by volumeFAGOC35.84% ↓, 34.09% ↓,
9.15% ↓, 2.83% ↑,
16.34% ↑ (CD)
a Abbreviations: D: Depth of chloride penetration, CD: Coefficient of chloride diffusion, CP: Charge passed, ER: Electrical resistivity. b Evaluate standard change compared to the control type (%), corresponding to RA replacement ratio. Increase: ↑, Decrease: ↓.

8. Carbonation Resistance

Table 11 shows that RA increases the depth of carbonation by replacing natural aggregates incorporated into concrete. Gheni et al. [118] supposed that carbonation depth is closely related to concrete permeability, which is related to the distribution and size of internal pores; as the exposure time to the CO2 environment increases, the carbonation depth increases [49]. Concrete workability deteriorates due to rubber incorporation; the interior of concrete is difficult to compact during mixing, consequently forming internal pores and cracks that make it easy for CO2 to penetrate the concrete, and replacing coarse aggregate with rubber results in greater carbonation depth compared with replacing fine aggregate [69]. Thomas and Gupta [126] found that the reason for reduction in carbonation resistance is similar to that for reduction in permeability resistance. Both are attributed to the hydrophobic nature of RA, causing the cement matrix to be hard to adhere to the rubber particle surface stably. As a result, internal pores and cracks increase, which becomes a wide channel for CO2 to penetrate into concrete. Pham et al. [113] found that after the CR was pretreated with NaOH, the adhesion was improved and the penetration of CO2 was inhibited. In addition, the NaOH remaining on the rubber reacted with CO2, which was conducive to reducing the carbonation depth.
However, Thomas et al. [125] observed that with 2.5–12.5% rubber content, compared to the control group, the carbonation depth was smaller; when it was more than 12.5%, the depth of carbonation increases with increasing rubber content. When the content was 2.5–12.5%, the substituted fine aggregate had a similar particle size to rubber particles, causing the RC easier to mix, and the aggregate distribution inside was more uniform, which made the RC compact after mixing. Excessive rubber content leads to increased internal pores and lack of filling [113].

9. Alkali–Silica Reaction Damage Resistance

Previous research demonstrated that in concrete, using RA to replace partial nature aggregate can alleviate the damage caused by ASR. ASR occurs between the OH and active silica aggregate in the alkaline environment. It generates a hygroscopic gel in the pore of cement matrix, which causes internal pressures and structure rupture with gel volume expansion [127,128]. RA has lower elastic modulus than natural aggregate, and has good deformation ability, which can absorb the stress created by ASR gel volume expansion. Therefore, ASR damage can be alleviated to some extent [110]. Afshinnia and Poursaee [129] found that replacing natural fine aggregate with 16% and 24% RAs reduced the expansion of ASR by 43% and 39%, respectively. Rubber particles do not react in an alkaline environment and can prevent the development of cracks and dissipate the energy that caused cracks; as a result, fewer cracks and swelling were observed in RC compared with the control group [130]. The incorporation of synthetic polypropylene (PP) fibre into RC could help alleviate the swelling stress and reduce the ASR damage, The PP fibre have the bridging effect, which can maintain internal integrity [102]. After adding rubber (15%) and PP fibre (0.5%), the swelling caused by ASR was less than 0.1% at 14 days, which means very low possibility of ASR damage according to ASTM C1260 [131].
Pretreatment of the CR surface with NaOH solution, and then mixing the treated CR into self-compacting mortar (SCM) can alleviate the expansion of ASR [95]. ASR expansion of normal SCM and rubber-modified SCM is shown in Figure 13. The internal cracks of RC were reduced because the NaOH treatment strengthened the binding ability of cement matrix and CR. This condition resulted in a decrease in the mobility of alkali and water and helped reduce ASR gel. However, excessive rubber content reduced the stiffness of concrete and caused great deformation.

10. Conclusions

This work introduces approximately 90 research results found in the past 10 years, involves various durability of RC and treatment methods of rubber, the following conclusions are drawn:
  • Pre-treating and pre-coating of rubber can reduce the internal porosity of RC, the occurrence and development of cracks at ITZ of RC, and thus enhance the durability of RC. When selecting treatment materials, comprehensive consideration should be given to the effect of improving durability, the feasibility of operating procedures, cost consumption, and environmental impact. The pretreating and precoating processes have a promoting significance for the application of rubber to concrete structures.
  • Rubber particles reduce the abrasion resistance of concrete. There are two main reasons for reducing RC abrasion resistance, one is high porosity and the other is weak adhesion on the RA surface, and a rubber content of 5–10% has a slightly negative effect on abrasion resistance. By contrast, adding SCMs or pre-treatment of rubber can effectively improve the abrasion resistance. In addition, well-graded rubber particles contribute to improved abrasion resistance.
  • The non-hydrophilic nature properties of RA leading the poor binding ability, which make it easy to form pores and water seepage channel in RC, these interconnected channels help to increase the water absorption and permeability. Well-graded rubber particles can make RC denser than a single particle size. When the rubber content is less than 10%, the increase in water absorption is small or even decreases. When the rubber content is too large (greater than 15%), the water absorption increases significantly. Rubber with a small particle size (0–1 mm) can effectively fill the pores and water seepage channels, which can effectively reduce the water absorption and enhance the RC impermeability. Pre-treatment of rubber particles can effectively reduce water absorption and impermeability of RC.
  • The pores created by the rubber incorporated into the concrete can play a role in stress absorption, thereby enhancing the freeze–thaw resistance of concrete. The best freeze–thaw resistance is achieved when the rubber content is 25–30%. RC with small-sized rubber particles (0–1 mm) has high freeze–thaw resistance. Pretreating rubber with NaOH and precoating it with synthetic resin and styrene–butadiene-type copolymer can significantly enhance the RC freeze–thaw resistance. The rubber replaces partial natural aggregate in concrete can enhance the freeze–thaw resistance but negatively affect the concrete strength. RC is suitable for areas without high-strength requirements but with high freeze–thaw resistance requirements.
  • Using rubber particles in ordinary concrete can enhance the acid resistance, and the effect of well-graded rubber particles and fibre is significant. Rubber particles decrease concrete strength, 5–15% rubber content can meet the requirements of use strength and has high resistance to acid attack. Rubber can effectively improve the resistance to sulphate attack. Compared with concrete without RA, the RC has better deformation ability. Rubber can relieve internal expansion stress caused by sulphate attack. The best sulphate attack resistance is obtained when the rubber content is 5–10%. Synthetic resin is a good modifier for precoated rubber particles considering compressive strength change, resistance to sulphury acid, and freeze–thaw.
  • The water absorption resistance and impermeability have a closely relation with the chloride penetration resistance. The addition of an appropriate amount of rubber (5–20%) to concrete can effectively improve the resistance to chloride permeability. The effect of RA replaces fine aggregate is better than that of coarse aggregate. Rubber fibre and fine rubber particles have a better effect on enhancing resistance to chloride attack. Rubber particle size should not be larger than 3 mm. Pre-treating rubber particles with NaOH or SCA and pre-coating rubber with CSBR latex, Na2SiO3-mixed cement paste, or SF-mixed cement can effectively enhance the resistance to chloride ion penetration.
  • The addition of rubber particles to concrete increases carbonation depth. The principle of enhancing the carbonation resistance of RC is similar to that of reducing water absorption. Pre-treatment increases the adhesion of the rubber and increases the density of the RC, thereby inhibiting CO2 penetration and effectively reducing the depth of carbonation.
  • Rubber particles replace partial nature aggregate can alleviate the internal structure damage, caused by ASR. RA do not react in an alkaline environment and have good deformation ability; ASR gel expansion stress happening in concrete internal structure can be alleviated by flexible RA, which can prevent cracks from developing and dissipate the energy that caused cracks. Pretreating rubber with NaOH can enhance the resistance of RC to ASR damage. However, excessive rubber content (greater than 25%) leads to the deformation of RC, which is unbeneficial to structural services.
  • Through the analysis of the rubber particle size, replacement ratio, and replacement pattern, the recommended as follows: the replacement pattern FAG is preferred, followed by cement material and CAG, and the rubber particle size and replacement ratio are 0–3 mm and 5–20%, respectively. If auxiliary cementitious materials are added, the replacement ratio of rubber can be appropriately increased by 5–10%.

11. Further Research Needs

  • The treatment methods of rubber particles should be further studied, especially the related research of physical-chemical coupling treatment, so as to increase the bonding between rubber and cement-based, and then improve the durability.
  • More research on the improvement of concrete durability by rubber fiber can be carried out.
  • Combined with microstructure analysis, more research can be carried out on the durability of RC, especially the long-term durability.
  • Further research is needed on RC ductility and energy absorption.
  • The insulation, sound insulation, thermal resistance, and corrosion resistance of RC need to be further studied.

Author Contributions

Conceptualization, Y.L. and R.W.; methodology, Y.L.; software, J.C.; validation, Y.L., R.W. and Y.Z.; formal analysis, J.C.; investigation, X.T.; resources, Y.L.; data curation, J.C. and Y.Z.; writing—original draft preparation, J.C.; writing—review and editing, Y.L.; visualization, J.C.; supervision, Y.L.; project administration, R.W.; funding acquisition, Y.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Project funded by China Postdoctoral Science Foundation (2020M683687XB and 2022T150525), the National Natural Science Foundation of China (52009110 and 51879217), Natural Science Basic Research Program of Shaanxi (2021JM-331), Young Talent Fund of Association for Science and Technology in Shaanxi, China (20220416), and Young Talent fund of Association for Science and Technology in Xi’an City (095920211332).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Danish, A.; Mosaberpanah, M.A. A review on recycled concrete aggregates (RCA) characteristics to promote RCA utilization in developing sustainable recycled aggregate concrete (RAC). Eur. J. Environ. Civ. Eng. 2021, 26, 6505–6539. [Google Scholar] [CrossRef]
  2. Mohammed, Z.M.; Abdulhameed, A.A.; Kazim, H.K. Effect of Alkali-Activated Natural Pozzolan on Mechanical Properties of Geopolymer Concrete. Civ. Environ. Eng. 2022, 18, 312–320. [Google Scholar] [CrossRef]
  3. Xiao, Z.; Pramanik, A.; Basak, A.K.; Prakash, C.; Shankar, S. Material recovery and recycling of waste tyres—A review. Clean. Mater. 2022, 5, 100115. [Google Scholar] [CrossRef]
  4. Qaidi, S.; Najm, H.M.; Abed, S.M.; Özkılıç, Y.O.; Al Dughaishi, H.; Alosta, M.; Sabri, M.M.S.; Alkhatib, F.; Milad, A. Concrete Containing Waste Glass as an Environmentally Friendly Aggregate: A Review on Fresh and Mechanical Characteristics. Materials 2022, 15, 6222. [Google Scholar] [CrossRef] [PubMed]
  5. Karalar, M.; Özkılıç, Y.O.; Deifalla, A.F.; Aksoylu, C.; Arslan, M.H.; Ahmad, M.; Sabri, M.M.S. Improvement in Bending Performance of Reinforced Concrete Beams Produced with Waste Lathe Scraps. Sustainability 2022, 14, 12660. [Google Scholar] [CrossRef]
  6. Abdulhameed, A.A.; Al-Zuhairi, A.H.; Al Zaidee, S.R.; Hanoon, A.N.; Al Zand, A.W.; Hason, M.M.; Abdulhameed, H.A. The Behavior of Hybrid Fiber-Reinforced Concrete Elements: A New Stress-Strain Model Using an Evolutionary Approach. Appl. Sci. 2022, 12, 2245. [Google Scholar] [CrossRef]
  7. Aksoylu, C.; Özkılıç, Y.O.; Hadzima-Nyarko, M.; Isık, E.; Arslan, M.H. Investigation on Improvement in Shear Performance of Reinforced-Concrete Beams Produced with Recycled Steel Wires from Waste Tires. Sustainability 2022, 14, 13360. [Google Scholar] [CrossRef]
  8. Çelik, A.I.; Özkılıç, Y.O.; Zeybek, Ö.; Özdöner, N.; Tayeh, B.A. Performance Assessment of Fiber-Reinforced Concrete Produced with Waste Lathe Fibers. Sustainability 2022, 14, 11817. [Google Scholar] [CrossRef]
  9. Assaggaf, R.; Maslehuddin, M.; Al-Osta, M.A.; Al-Dulaijan, S.U.; Ahmad, S. Properties and sustainability of treated crumb rubber concrete. J. Build. Eng. 2022, 51, 104250. [Google Scholar] [CrossRef]
  10. Roychand, R.; Gravina, R.J.; Zhuge, Y.; Ma, X.; Youssf, O.; Mills, J.E. A comprehensive review on the mechanical properties of waste tire rubber concrete. Constr. Build. Mater. 2020, 237, 117651. [Google Scholar] [CrossRef]
  11. Li, X.; Ling, T.; Mo, K.H. Functions and impacts of plastic/rubber wastes as eco-friendly aggregate in concrete-A review. Constr. Build. Mater. 2020, 240, 117869. [Google Scholar] [CrossRef]
  12. Kaewunruen, S.; Li, D.; Chen, Y.; Xiang, Z. Enhancement of dynamic damping in eco-friendly railway concrete sleepers using waste-tyre crumb rubber. Materials 2018, 11, 1169. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Ince, C.; Shehata, B.M.H.; Derogar, S.; Ball, R.J. Towards the development of sustainable concrete incorporating waste tyre rubbers: A long-term study of physical, mechanical & durability properties and environmental impact. J. Clean. Prod. 2022, 334, 130223. [Google Scholar]
  14. Alyousef, R.; Ahmad, W.; Ahmad, A.; Aslam, F.; Joyklad, P.; Alabduliabbar, H. Potential use of recycled plastic and rubber aggregate in cementitious materials for sustainable construction: A review. J. Clean. Prod. 2021, 329, 129736. [Google Scholar] [CrossRef]
  15. Medina, N.F.; Medina, D.F.; Hernández-Olivares, F.; Navacerrada, M.A. Mechanical and thermal properties of concrete incorporating rubber and fibres from tyre recycling. Constr. Build. Mater. 2017, 144, 563–573. [Google Scholar] [CrossRef]
  16. Amiri, M.; Hatami, F.; Golafshani, E.M. Evaluating the synergic effect of waste rubber powder and recycled concrete aggregate on mechanical properties and durability of concrete. Case Stud. Constr. Mater. 2021, 15, 00639. [Google Scholar] [CrossRef]
  17. Dhivya, K.; Priyadharshini, K. Experimental study on strength properties of concrete with partial replacement of coarse aggregate by rubber tyre waste. Mater. Today Proc. 2021, 11, 578. [Google Scholar] [CrossRef]
  18. Samarakoon, S.S.M.; Ruben, P.; Pedersen, J.W.; Evangelista, L. Mechanical performance of concrete made of steel fibers from tire waste. Case Stus. Constr. Mater. 2019, 11, e00259. [Google Scholar] [CrossRef]
  19. 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. [Google Scholar] [CrossRef]
  20. Abdulhameed, A.A.; Hanoon, A.N.; Abdulhameed, H.A.; Banyhussan, Q.S.; Mansi, A.S. Push-out test of steel-concrete-steel composite sections with various core materials: Behavioural study. Arch. Civ. Mech. Eng. 2021, 21, 17. [Google Scholar] [CrossRef]
  21. Siddika, A.; Mamun, M.A.A.; Alyousef, R.; Amran, Y.H.M.; Aslani, F.; Alabduljabbar, H. Properties and utilizations of waste tire rubber in concrete: A review. Constr. Build. Mater. 2019, 224, 711–731. [Google Scholar] [CrossRef]
  22. Xu, J.; Yao, Z.; Yang, G.; Han, Q. Research on crumb rubber concrete: From a multi-scale review. Constr. Build. Mater. 2020, 232, 117282. [Google Scholar] [CrossRef]
  23. Ramarad, S.; Khalid, M.; Ratnam, C.T.; Chuah, A.L.; Rashmi, W. Waste tire rubber in polymer blends: A review on the evolution, properties and future. Prog. Mater. Sci. 2015, 72, 100–140. [Google Scholar] [CrossRef]
  24. Qaidi, S.M.A.; Dinkha, Y.Z.; Haido, J.H.; Ali, M.H.; Tayeh, B.A. Engineering properties of sustainable green concrete incorporating eco-friendly aggregate of crumb rubber: A review. J. Clean. Prod. 2021, 324, 129251. [Google Scholar] [CrossRef]
  25. Medina, N.F.; Garcia, R.; Hajirasouliha, I.; Pilakoutas, K.; Guadagnini, M.; Raffoul, S. Composites with recycled rubber aggregates: Properties and opportunities in construction. Constr. Build. Mater. 2018, 188, 884–897. [Google Scholar] [CrossRef]
  26. Assaggaf, R.A.; Al-Dulaijan, S.U.; Maslehuddin, M.; Al-Amoudi, O.S.B.; Ahmad, S.; Ibrahim, M. Effect of different treatments of crumb rubber on the durability characteristics of rubberized concrete. Constr. Build. Mater. 2022, 318, 126030. [Google Scholar] [CrossRef]
  27. Segre, N.; Joekes, I. Use of tire rubber particles as addition to cement paste. Cem. Concr. Res. 2000, 30, 1421–1425. [Google Scholar] [CrossRef]
  28. He, L.; Ma, Y.; Liu, Q.; Mu, Y. Surface modification of crumb rubber and its influence on the mechanical properties of rubber-cement concrete. Constr. Build. Mater. 2016, 120, 403–407. [Google Scholar] [CrossRef]
  29. Najim, K.B.; Hall, M.R. Crumb rubber aggregate coatings/pre-treatments and their effects on interfacial bonding, air entrapment and fracture toughness in self-compacting rubberised concrete (SCRC). Mater. Struct. 2013, 46, 2029–2043. [Google Scholar] [CrossRef]
  30. Huang, B.; Shu, X.; Cao, J. A two-staged surface treatment to improve properties of rubber modified cement composites. Constr. Build. Mater. 2013, 40, 270–274. [Google Scholar] [CrossRef]
  31. Haibo, Z.; Mifeng, G.; Liu, X.; Xuemao, G. Effect of rubber particle modification on properties of rubberized concrete. J. Wuhan Univ. Technol. 2014, 29, 763–768. [Google Scholar]
  32. Rivas-Vázquez, L.P.; Suárez-Orduña, R.; Hernández-Torres, J.; Aquino-Bolaños, E. Effect of the surface treatment of recycled rubber on the mechanical strength of composite concrete/rubber. Mater. Struct. 2014, 48, 2809–2814. [Google Scholar] [CrossRef]
  33. Li, G.; Wang, Z.; Leung, C.K.Y.; Tang, S.; Pan, J.; Huang, W.; Chen, E. Properties of rubberized concrete modified by using silane coupling agent and carboxylated SBR. J. Clean. Prod. 2016, 112, 797–807. [Google Scholar] [CrossRef]
  34. Youssf, O.; Hassanli, R.; Mills, J.E.; Skinner, W.; Ma, X.; Zhuge, Y.; Roychand, R.; Gravina, R. Influence of mixing procedures, rubber treatment, and fibre additives on rubcrete performance. J. Compos. Sci. 2019, 3, 41. [Google Scholar] [CrossRef] [Green Version]
  35. Abdulla, A.I.; Ahmed, S.H. Effect of rubber treated by acidic solution on some mechanical properties of rubberize cement mortar. Eng. Technol. J. 2011, 29, 2793–2806. [Google Scholar]
  36. Leung, C.K.; Grasley, Z.C. Effect of micrometric and nanometric viscoelastic inclusions on mechanical damping behavior of cementitious composites. Constr. Build. Mater. 2012, 35, 444–451. [Google Scholar] [CrossRef]
  37. Chou, L.H.; Lin, C.N.; Lu, C.K.; Lee, C.H.; Lee, M.T. Improving rubber concrete by waste organic sulfur compounds. Waste Manag. Res. 2010, 28, 29–35. [Google Scholar] [CrossRef] [PubMed]
  38. Onuaguluchi, O. Effects of surface pre-coating and silica fume on crumb rubber-cement matrix interface and cement mortar properties. J. Clean. Prod. 2015, 104, 339–345. [Google Scholar] [CrossRef]
  39. Guo, S.; Dai, Q.; Si, R.; Sun, X.; Lu, C. Evaluation of properties and performance of rubber-modified concrete for recycling of waste scrap tire. J. Clean. Prod. 2017, 148, 681–689. [Google Scholar] [CrossRef]
  40. Liu, H.; Wang, X.; Jiao, Y.; Sha, T. Experimental investigation of the mechanical and durability properties of crumb rubber concrete. Materials 2016, 9, 172. [Google Scholar] [CrossRef] [Green Version]
  41. Pham, N.P.; Toumi, A.; Turatsinze, A. Effect of an enhanced rubber-cement matrix interface on freeze-thaw resistance of the cement-based composite. Constr. Build. Mater. 2019, 207, 528–534. [Google Scholar] [CrossRef]
  42. Ganjian, E.; Khorami, M.; Maghsoudi, A.A. Scrap-tyre-rubber replacement for aggregate and filler in concrete. Constr. Build. Mater. 2009, 23, 1828–1836. [Google Scholar] [CrossRef]
  43. Flores, M.N.; Flores-Medina, D.; Hernández-Olivares, F. Influence of fibers partially coated with rubber from tire recycling as aggregate on the acoustical properties of rubberized concrete. Constr. Build. Mater. 2016, 129, 25–36. [Google Scholar] [CrossRef]
  44. Meddah, A.; Bensaci, H.; Beddar, M. Study of the effects of mechanical and chemical treatment of rubber on the performance of rubberized roller-compacted concrete pavement. Innov. Infrastruct. Solut. 2017, 2, 17–22. [Google Scholar] [CrossRef] [Green Version]
  45. Pham, N.P.; Toumi, A.; Turatsinze, A. Rubber aggregate-cement matrix bond enhancement: Microstructural analysis, effect on transfer properties and on mechanical behaviours of the composite. Cem. Concr. Compos. 2018, 94, 1–12. [Google Scholar] [CrossRef]
  46. Angelin, A.F.; Miranda, E.J.P.; Dos Santos, J.M.C.; Lintz, R.C.C.; Gachet-Barbosa, L.A. Rubberized mortar: The influence of aggregate granulometry in mechanical resistances and acoustic behavior. Constr. Build. Mater. 2019, 200, 248–254. [Google Scholar] [CrossRef]
  47. Gupta, T.; Chaudhary, S.; Sharma, R.K. Mechanical and durability properties of waste rubber fiber concrete with and without silica fume. J. Clean. Prod. 2016, 112, 702–711. [Google Scholar] [CrossRef]
  48. Li, Y.; Zhang, S.; Wang, R.; Dang, F. Potential use of waste tire rubber as aggregate in cement concrete—A comprehensive review. Constr. Build. Mater. 2019, 225, 1183–1201. [Google Scholar] [CrossRef]
  49. Gupta, T.; Chaudhary, S.; Sharma, R.K. Assessment of mechanical and durability properties of concrete containing waste rubber tire as fine aggregate. Constr. Build. Mater. 2014, 73, 562–574. [Google Scholar] [CrossRef]
  50. Gencel, O.; Gok, M.S.; Brostow, W. Effect of metallic aggregate and cement content on abrasion resistance behaviour of concrete. Mater. Res. Innov. 2015, 15, 116–123. [Google Scholar] [CrossRef]
  51. Kang, J.; Zhang, B.; Guangyu, L.I. The abrasion-resistance investigation of rubberized concrete. J. Wuhan Univ. Technol. 2012, 27, 1144–1148. [Google Scholar] [CrossRef]
  52. ASTM C1138; Standard Test Method For Abrasion Resistance of Concrete (Underwater Method). American Society for Testing and Materials: West Conshohocken, PA, USA, 2012.
  53. SL 352; Test Code for Hydraulic Concrete. The Ministry of Water Resources of the People’ s Republic of China: Beijing, China, 2006.
  54. Bisht, K.; Ramana, P.V. Evaluation of mechanical and durability properties of crumb rubber concrete. Constr. Build. Mater. 2017, 155, 811–817. [Google Scholar] [CrossRef]
  55. Turki, M.; Bretagne, E.; Rouis, M.J.; Quéneudec, M. Microstructure, physical and mechanical properties of mortar–rubber aggregates mixtures. Constr. Build. Mater. 2009, 23, 2715–2722. [Google Scholar] [CrossRef]
  56. Ridgley, K.E.; Abouhussien, A.A.; Hassan, A.; Colbourne, B. Evaluation of abrasion resistance of self-consolidating rubberized concrete by acousticemission analysis. J. Mater. Civ. Eng. 2018, 30, 04018196. [Google Scholar] [CrossRef]
  57. Thomas, B.S.; Chandra Gupta, R. Properties of high strength concrete containing scrap tire rubber. J. Clean. Prod. 2016, 113, 86–92. [Google Scholar] [CrossRef]
  58. Ozbay, E.; Lachemi, M.; Sevim, U. Compressive strength, abrasion resistance and energy absorption capacity of rubberized concretes with and without slag. Mater. Struct. 2010, 44, 1297–1307. [Google Scholar] [CrossRef]
  59. Mohammed, B.S.; Adamu, M. Mechanical performance of roller compacted concrete pavement containing crumb rubber and nano silica. Constr. Build. Mater. 2018, 159, 234–251. [Google Scholar] [CrossRef]
  60. Shen, W.; Shan, L.; Zhang, T.; Ma, H.; Cai, Z.; Shi, H. Investigation on polymer–rubber aggregate modified porous concrete. Constr. Build. Mater. 2013, 38, 667–674. [Google Scholar] [CrossRef]
  61. Silva, F.M.D.; Gachet Barbosa, L.A.; Lintz, R.C.C.; Jacintho, A.E.P. Investigation on the properties of concrete tactile paving blocks made with recycled tire rubber. Constr. Build. Mater. 2015, 91, 71–79. [Google Scholar] [CrossRef]
  62. Gesoğlu, M.; Güneyisi, E.; Khoshnaw, G.; İpek, S. Abrasion and freezing–thawing resistance of pervious concretes containing waste rubbers. Constr. Build. Mater. 2014, 73, 19–24. [Google Scholar] [CrossRef]
  63. 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. [Google Scholar] [CrossRef]
  64. Sukontasukkul, P.; Chaikaew, C. Properties of concrete pedestrian block mixed with crumb rubber. Constr. Build. Mater. 2006, 20, 450–457. [Google Scholar] [CrossRef]
  65. Kashani, A.; Ngo, T.D.; Hemachandra, P.; Hajimohammadi, A. Effects of surface treatments of recycled tyre crumb on cement-rubber bonding in concrete composite foam. Constr. Build. Mater. 2018, 171, 467–473. [Google Scholar] [CrossRef]
  66. Valadares, F.; Bravo, M.; Jorge, D.B. Concrete with used tire rubber aggregates: Mechanical performance. ACI Mater. J. 2012, 109, 283–292. [Google Scholar]
  67. Mohammed, B.S.; Anwar Hossain, K.M.; Eng Swee, J.T.; Wong, G.; Abdullahi, M. Properties of crumb rubber hollow concrete block. J. Clean. Prod. 2012, 23, 57–67. [Google Scholar] [CrossRef]
  68. Onuaguluchi, O.; Panesar, D.K. Hardened properties of concrete mixtures containing pre-coated crumb rubber and silica fume. J. Clean. Prod. 2014, 82, 125–131. [Google Scholar] [CrossRef]
  69. Bravo, M.; de Brito, J. Concrete made with used tyre aggregate: Durability-related performance. J. Clean. Prod. 2012, 25, 42–50. [Google Scholar] [CrossRef]
  70. Alsaif, A.; Bernal, S.A.; Guadagnini, M.; Pilakoutas, K. Durability of steel fibre reinforced rubberised concrete exposed to chlorides. Constr. Build. Mater. 2018, 188, 130–142. [Google Scholar] [CrossRef]
  71. Girskas, G.; Nagrockienė, D. Crushed rubber waste impact of concrete basic properties. Constr. Build. Mater. 2017, 140, 36–42. [Google Scholar] [CrossRef]
  72. Youssf, O.; Hassanli, R.; Mills, J.E. Mechanical performance of FRP-confined and unconfined crumb rubber concrete containing high rubber content. J. Build. Eng. 2017, 11, 115–126. [Google Scholar] [CrossRef]
  73. Hunag, L.; Wang, H.; Wu, Y. Properties of the mechanical in controlled low-strength rubber lightweight aggregate concrete (CLSRLC). Constr. Build. Mater. 2016, 112, 1054–1058. [Google Scholar] [CrossRef]
  74. Benazzouk, A.; Douzane, O.; Langlet, T.; Mezreb, K.; Roucoult, J.M.; Quéneudec, M. Physico-mechanical properties and water absorption of cement composite containing shredded rubber wastes. Cem. Concr. Compos. 2007, 29, 732–740. [Google Scholar] [CrossRef]
  75. Gesoğlu, M.; Güneyisi, E. Permeability properties of self-compacting rubberized concretes. Constr. Build. Mater. 2011, 25, 3319–3326. [Google Scholar] [CrossRef]
  76. Sukontasukkul, P.; Tiamlom, K. Expansion under water and drying shrinkage of rubberized concrete mixed with crumb rubber with different size. Constr. Build. Mater. 2012, 29, 520–526. [Google Scholar] [CrossRef]
  77. Li, N.; Long, G.; Ma, C.; Fu, Q.; Zeng, X.; Ma, K.; Xie, Y.; Luo, B. Properties of self-compacting concrete (SCC) with recycled tire rubber aggregate: A comprehensive study. J. Clean. Prod. 2019, 236, 117707. [Google Scholar] [CrossRef]
  78. Kashani, A.; Ngo, T.D.; Mendis, P.; Black, J.R.; Hajimohammadi, A. A sustainable application of recycled tyre crumbs as insulator in lightweight cellular concrete. J. Clean. Prod. 2017, 149, 925–935. [Google Scholar] [CrossRef]
  79. Meddah, A.; Beddar, M.; Bali, A. Use of shredded rubber tire aggregates for roller compacted concrete pavement. J. Clean. Prod. 2014, 72, 187–192. [Google Scholar] [CrossRef]
  80. Rezaifar, O.; Hasanzadeh, M.; Gholhaki, M. Concrete made with hybrid blends of crumb rubber and metakaolin: Optimization using response surface method. Constr. Build. Mater. 2016, 123, 59–68. [Google Scholar] [CrossRef]
  81. Wang, R.; Gao, P.; Tian, M.; Dai, Y. Experimental study on mechanical and waterproof performance of lightweight foamed concrete mixed with crumb rubber. Constr. Build. Mater. 2019, 209, 655–664. [Google Scholar] [CrossRef]
  82. Yilmaz, A.; Degirmenci, N. Possibility of using waste tire rubber and fly ash with Portland cement as construction materials. Waste Manag. 2009, 29, 1541–1546. [Google Scholar] [CrossRef]
  83. Pedro, D.; de Brito, J.; Veiga, R. Mortars made with fine granulate from shredded tires. J. Mater. Civ. Eng. 2013, 25, 519–529. [Google Scholar] [CrossRef]
  84. Ling, T.-C. Effects of compaction method and rubber content on the properties of concrete paving blocks. Constr. Build. Mater. 2012, 28, 164–175. [Google Scholar] [CrossRef]
  85. Hall, M.R.; Najim, K.B. Structural behaviour and durability of steel-reinforced structural plain/self-compacting rubberised concrete (PRC/SCRC). Constr. Build. Mater. 2014, 73, 490–497. [Google Scholar] [CrossRef]
  86. Azevedo, F.; Pacheco-Torgal, F.; Jesus, C.; Barroso de Aguiar, J.L.; Camões, A.F. Properties and durability of HPC with tyre rubber wastes. Constr. Build. Mater. 2012, 34, 186–191. [Google Scholar] [CrossRef]
  87. Su, H.; Yang, J.; Ling, T.; Ghataora, G.S.; Dirar, S. Properties of concrete prepared with waste tyre rubber particles of uniform and varying sizes. J. Clean. Prod. 2015, 91, 288–296. [Google Scholar] [CrossRef] [Green Version]
  88. Gesoğlu, M.; Güneyisi, E.; Khoshnaw, G.; İpek, S. Investigating properties of pervious concretes containing waste tire rubbers. Constr. Build. Mater. 2014, 63, 206–213. [Google Scholar] [CrossRef]
  89. Si, R.; Wang, J.; Guo, S.; Dai, Q.; Han, S. Evaluation of laboratory performance of self-consolidating concrete with recycled tire rubber. J. Clean. Prod. 2018, 180, 823–831. [Google Scholar] [CrossRef]
  90. Gupta, T.; Siddique, S.; Sharma, R.K.; Chaudhary, S. Effect of elevated temperature and cooling regimes on mechanical and durability properties of concrete containing waste rubber fiber. Constr. Build. Mater. 2017, 137, 35–45. [Google Scholar] [CrossRef]
  91. Turgut, P.; Yesilata, B. Physico-mechanical and thermal performances of newly developed rubber-added bricks. Energ. Build. 2008, 40, 679–688. [Google Scholar] [CrossRef]
  92. Paine, K.A.; Dhir, R.K. Research on new applications for granulated rubber in concrete. Proc. Inst. Civ. Eng. Constr. Mater. 2010, 163, 7–17. [Google Scholar] [CrossRef] [Green Version]
  93. Zhu, X.; Miao, C.; Liu, J.; Hong, J. Influence of crumb rubber on frost resistance of concrete and effect mechanism. Procedia Eng. 2012, 27, 206–213. [Google Scholar] [CrossRef] [Green Version]
  94. Li, Y.; Zhou, Y.; Wang, R.; Li, Y.; Wu, X.; Si, Z. Experimental investigation on the properties of the interface between RCC layers subjected to early-age frost damage. Cem. Concr. Compos. 2022, 134, 104745. [Google Scholar] [CrossRef]
  95. Nehdi, M.L.; Najjar, M.F.; Soliman, A.M.; Azabi, T.M. Novel eco-efficient two-stage concrete incorporating high volume recycled content for sustainable pavement construction. Constr. Build. Mater. 2017, 146, 9–14. [Google Scholar] [CrossRef]
  96. Baricevic, A.; Pezer, M.; Jelcic Rukavina, M.; Serdar, M.; Stirmer, N. Effect of polymer fibers recycled from waste tires on properties of wet-sprayed concrete. Constr. Build. Mater. 2018, 176, 135–144. [Google Scholar] [CrossRef]
  97. Alsaif, A.; Bernal, S.A.; Guadagnini, M.; Pilakoutas, K. Freeze-thaw resistance of steel fibre reinforced rubberised concrete. Constr. Build. Mater. 2019, 195, 450–458. [Google Scholar] [CrossRef]
  98. Gonen, T. Freezing-thawing and impact resistance of concretes containing waste crumb rubbers. Constr. Build. Mater. 2018, 177, 436–442. [Google Scholar] [CrossRef]
  99. Al-Akhras, N.M.; Smadi, M.M. Properties of tire rubber ash mortar. Cem. Concr. Compos. 2004, 26, 821–826. [Google Scholar] [CrossRef]
  100. Topçu, İ.B.; Bilir, T. Experimental investigation of some fresh and hardened properties of rubberized self-compacting concrete. Mater. Des. 2009, 30, 3056–3065. [Google Scholar] [CrossRef]
  101. Si, R.; Guo, S.; Dai, Q. Durability performance of rubberized mortar and concrete with NaOH-solution treated rubber particles. Constr. Build. Mater. 2017, 153, 496–505. [Google Scholar] [CrossRef]
  102. Wang, J.; Dai, Q.; Si, R.; Guo, S. Mechanical, durability, and microstructural properties of macro synthetic polypropylene (PP) fiber-reinforced rubber concrete. J. Clean. Prod. 2019, 234, 1351–1364. [Google Scholar] [CrossRef]
  103. Zhang, B.; Feng, Y.; Xie, J.; Lai, D.; Yu, T.; Huang, D. Rubberized geopolymer concrete: Dependence of mechanical properties and freeze-thaw resistance on replacement ratio of crumb rubber. Constr. Build. Mater. 2021, 310, 125248. [Google Scholar] [CrossRef]
  104. Richardson, A.E.; Coventry, K.A.; Ward, G. Freeze/thaw protection of concrete with optimum rubber crumb content. J. Clean. Prod. 2012, 23, 96–103. [Google Scholar] [CrossRef] [Green Version]
  105. Wang, R.; Hu, Z.; Li, Y.; Wang, K.; Zhang, H. Review on the deterioration and approaches to enhance the durability of concrete in the freeze–thaw environment. Constr. Build. Mater. 2022, 321, 126371. [Google Scholar] [CrossRef]
  106. Topçu, İ.B.K.; Demir, A. Durability of rubberized mortar and concrete. J. Mater. Civ. Eng. 2007, 19, 173–178. [Google Scholar] [CrossRef]
  107. Gupta, T.; Siddique, S.; Sharma, R.K.; Chaudhary, S. Behaviour of waste rubber powder and hybrid rubber concrete in aggressive environment. Constr. Build. Mater. 2019, 217, 283–291. [Google Scholar] [CrossRef]
  108. Bisht, K.; Ramana, P.V. Waste to resource conversion of crumb rubber for production of sulphuric acid resistant concrete. Constr. Build. Mater. 2019, 194, 276–286. [Google Scholar] [CrossRef]
  109. Sofi, A. Effect of waste tyre rubber on mechanical and durability properties of concrete—A review. Ain Shams Eng. J. 2018, 9, 2691–2700. [Google Scholar] [CrossRef]
  110. Thomas, B.S.; Gupta, R.C.; Panicker, V.J. Recycling of waste tire rubber as aggregate in concrete: Durability-related performance. J. Clean. Prod. 2016, 112, 504–513. [Google Scholar] [CrossRef]
  111. Rahimi, R.S.; Nikbin, I.M.; Allahyari, H.; Habibi, T.S. Sustainable approach for recycling waste tire rubber and polyethylene terephthalate (PET) to produce green concrete with resistance against sulfuric acid attack. J. Clean. Prod. 2016, 126, 166–177. [Google Scholar] [CrossRef]
  112. Bassuoni, M.T.; Nehdi, M.L. Resistance of self-consolidating concrete to sulfuric acid attack with consecutive pH reduction. Cem. Concr. Res. 2007, 37, 1070–1084. [Google Scholar] [CrossRef]
  113. Pham, T.M.; Elchalakani, M.; Hao, H.; Lai, J.; Ameduri, S.; Tran, T.M. Durability characteristics of lightweight rubberized concrete. Constr. Build. Mater. 2019, 224, 584–599. [Google Scholar] [CrossRef]
  114. Onuaguluchi, O.; Banthia, N. Long-term sulfate resistance of cementitious composites containing fine crumb rubber. Cem. Concr. Compos. 2019, 104, 103354. [Google Scholar] [CrossRef]
  115. Thomas, B.S.; Gupta, R.C.; Panicker, V.J. Experimental and modelling studies on high strength concrete containing waste tire rubber. Sustain. Cities Soc. 2015, 19, 68–73. [Google Scholar] [CrossRef]
  116. Li, Y.; Yang, X.; Lou, P.; Wang, R.; Li, Y.; Si, Z. Sulfate attack resistance of recycled aggregate concrete with NaOH-solution-treated crumb rubber. Constr. Build. Mater. 2021, 287, 123044. [Google Scholar] [CrossRef]
  117. Yung, W.H.; Yung, L.C.; Hua, L.H. A study of the durability properties of waste tire rubber applied to self-compacting concrete. Constr. Build. Mater. 2013, 41, 665–672. [Google Scholar] [CrossRef]
  118. Gheni, A.A.; Alghazali, H.H.; ElGawady, M.A.; Myers, J.J.; Feys, D. Durability properties of cleaner cement mortar with by-products of tire recycling. J. Clean. Prod. 2019, 213, 1135–1146. [Google Scholar] [CrossRef]
  119. Fernández-Ruiz, M.A.; Gil-Martín, L.M.; Carbonell-Márquez, J.F.; Hernández-Montes, E. Epoxy resin and ground tyre rubber replacement for cement in concrete: Compressive behaviour and durability properties. Constr. Build. Mater. 2018, 173, 49–57. [Google Scholar] [CrossRef]
  120. Noor, N.M.; Hamada, H.; Sagawa, Y.; Yamamoto, D. Effect of crumb rubber on concrete strength and chloride ion penetration resistance. J. Teknol. 2015, 77, 171–178. [Google Scholar]
  121. Gueneyisi, E. Strength development and chloride penetration in rubberized concretes with and without silica fume. Mater. Struct. 2007, 40, 953–964. [Google Scholar]
  122. Dong, Q.; Huang, B.; Shu, X. Rubber modified concrete improved by chemically active coating and silane coupling agent. Constr. Build. Mater. 2013, 48, 116–123. [Google Scholar] [CrossRef]
  123. Oikonomou, N.; Mavridou, S. Improvement of chloride ion penetration resistance in cement mortars modified with rubber from worn automobile tires. Cem. Concr. Compos. 2009, 31, 403–407. [Google Scholar] [CrossRef]
  124. Zhu, H.; Liang, J.; Xu, J.; Bo, M.; Li, J.; Tang, B. Research on anti-chloride ion penetration property of crumb rubber concrete at different ambient temperatures. Constr. Build. Mater. 2018, 189, 42–53. [Google Scholar] [CrossRef]
  125. Thomas, B.S.; Gupta, R.C.; Mehra, P.; Kumar, S. Performance of high strength rubberized concrete in aggressive environment. Constr. Build. Mater. 2015, 83, 320–326. [Google Scholar] [CrossRef]
  126. Thomas, B.S.; Gupta, R.C. Long term behaviour of cement concrete containing discarded tire rubber. J. Clean. Prod. 2015, 102, 78–87. [Google Scholar] [CrossRef]
  127. Siddique, R.; Naik, T.R. Properties of concrete containing scrap-tire rubber—An overview. Waste Manag. 2004, 24, 563–569. [Google Scholar] [CrossRef]
  128. Wang, R.; Zhang, Q.; Li, Y. Deterioration of concrete under the coupling effects of freeze–thaw cycles and other actions: A review. Constr. Build. Mater. 2022, 319, 126045. [Google Scholar] [CrossRef]
  129. Afshinnia, K.; Poursaee, A. The influence of waste crumb rubber in reducing the alkali–silica reaction in mortar bars. J. Build. Eng. 2015, 4, 231–236. [Google Scholar] [CrossRef]
  130. Ho, A.C.; Turatsinze, A.; Hameed, R.; Vu, D.C. Effects of rubber aggregates from grinded used tyres on the concrete resistance to cracking. J. Clean. Prod. 2012, 23, 209–215. [Google Scholar] [CrossRef]
  131. ASTM C1260; Standard Test Method for Potential Alkali-Reactivity of Aggregates Mortar Bar Method. American Society for Testing and Materials: West Conshohocken, PA, USA, 2003.
Figure 1. Various sizes of tire particles [48].
Figure 1. Various sizes of tire particles [48].
Buildings 12 01975 g001
Figure 2. Adherence of ITZ of mortars with 0% RA and RA mortars with 30% RA [55].
Figure 2. Adherence of ITZ of mortars with 0% RA and RA mortars with 30% RA [55].
Buildings 12 01975 g002
Figure 3. SEM images of fracture surface of cement test specimen. (a) Rubber particle; (b) Cement paste [27].
Figure 3. SEM images of fracture surface of cement test specimen. (a) Rubber particle; (b) Cement paste [27].
Buildings 12 01975 g003
Figure 4. Microstructure of CR concrete [67].
Figure 4. Microstructure of CR concrete [67].
Buildings 12 01975 g004
Figure 5. Optical microscope of CR concrete [54].
Figure 5. Optical microscope of CR concrete [54].
Buildings 12 01975 g005
Figure 6. Microstructure of RC containing 20% rubber ash [49].
Figure 6. Microstructure of RC containing 20% rubber ash [49].
Buildings 12 01975 g006
Figure 7. SEM images of foamed concrete incorporated with CR [81].
Figure 7. SEM images of foamed concrete incorporated with CR [81].
Buildings 12 01975 g007
Figure 8. No. 6 CR trapped air bubbles in water [76].
Figure 8. No. 6 CR trapped air bubbles in water [76].
Buildings 12 01975 g008
Figure 10. Change in compressive strength after soaking in 3% H2SO4 solution [108].
Figure 10. Change in compressive strength after soaking in 3% H2SO4 solution [108].
Buildings 12 01975 g010
Figure 11. Weight loss after erosion with sodium sulphate solution for SCRC [117].
Figure 11. Weight loss after erosion with sodium sulphate solution for SCRC [117].
Buildings 12 01975 g011
Figure 12. The condition of RA surface after two step surface treatment [122].
Figure 12. The condition of RA surface after two step surface treatment [122].
Buildings 12 01975 g012
Figure 13. ASR expansion of different types of SCM at different ages [89].
Figure 13. ASR expansion of different types of SCM at different ages [89].
Buildings 12 01975 g013
Table 1. Comparison of various treatment methods of rubber waste on concrete.
Table 1. Comparison of various treatment methods of rubber waste on concrete.
Treatment MethodMechanismAdvantagesDisadvantages
Pre-treating with NaOH solution [27]Remove zinc stearate from rubber particle surface, make rubber surface hydrophilic.High efficiency, widely used, and synergy with other methods.The compressive strength is not improved, even slightly reduced.
Pre-treating with KMnO4 [28]Oxidize rubber surface to make the surface hydrophilic.High efficiency, cheap, and alternative chlorinated oxidizer.Complicated operation and time consuming.
Pre-treating with NaHSO3 [28]Sulphonate rubber surface to make the surface hydrophilic.High efficiency and cheap.Complicated operation and time consuming.
Pre-coating with LP [38]Make the rubber surface hydrophilic and rough.Cheap and easy access to raw materials.The void content and water absorption increase.
Pre-coating with ethoxyline resin [40]Make the rubber surface hydrophilic and sticky.High efficiency.The freeze–thaw resistance decreases.
Pre-coating with emulsion [40]Increase the rubber elastic modulus of and make the rubber hydrophilic.High efficiency and simple operation.The compressive strength and axial compressive strength decrease.
Pre-coating with styrene-butadiene-type copolymer [41]Make the rubber hydrophilic and rough.High efficiency.The freeze–thaw resistance decreases slightly.
Pre-treating with SCA [30,33,39]Facilitate the chemical bonding between rubber and cement.High efficiency and easy to combine other methods.The dynamic elastic modulus decreases slightly.
Pre-coating with cement [30,39]Make the rubber hydrophilic and strengthen the elastic modulus.High efficiency and easy access to raw materials.Excessive cement coating is not conducive to the increase of density.
Pre-coating with Na2SiO3 [39]Promoting the generation of calcium silicate hydrate gel in ITZ.High efficiency and environment friendly.The air content increases.
Pre-treating with acetone [32]Facilitate the bonding between rubber and cement.High efficiency and simple operation procedure.The workability decreases.
Table 2. Properties summarized in previous review studies.
Table 2. Properties summarized in previous review studies.
ReferencesWorkabilityDensityCompressive StrengthSplitting Tensile StrengthFlexural StrengthModulus of Elasticity
[10]
[21]
[22]--
[24]
ReferencesAbrasion ResistanceWater AbsorptionPermeabilityFreeze–Thaw ResistanceAcid ResistanceSulphate Resistance
[10]-----
[21]--
[22]--
[24]-----
ReferencesChloride Penetration ResistanceCarbonation ResistanceAlkali–Silica Reaction Damage ResistanceShrinkageMicrostructureLong-Term Behavior
[10]------
[21]---
[22]----
[24]-----
Table 3. Composition of manufactured tires by weight [23,24,42].
Table 3. Composition of manufactured tires by weight [23,24,42].
MaterialsIn USAIn European Union
Truck TyreCar TyreTruck TyreCar Tyre
Natural rubber (%)27143022
Synthetic rubber (%)14271523
Carbon black (%)28282028
Steel (%)14–1514–152513
Others (textile, fillers, curatives, stabilizers, antioxidants, and antiozonants) (%)16–1716–171014
Table 4. Classification of the rubber particles [24,42].
Table 4. Classification of the rubber particles [24,42].
Term NameParticle Size RangeReplacement Type
Shredded tyre100–230 mm in width, 300–460 mm in lengthCoarse aggregate
Chipped tyre13–76 mmCoarse aggregate
Fibre rubber2–5 mm in width, 10–22 mm in lengthFine aggregate
Granulated crumb rubber0.5–9.5 mmCoarse aggregate or Fine aggregate
Crumb rubber0.425–4.75 mmFine aggregate or cement
Rubber powder/Rubber ash≤0.425 mmCement
Table 5. Influence of rubber on the concrete abrasion resistance.
Table 5. Influence of rubber on the concrete abrasion resistance.
Reference Treatment MethodRA a Type and Size (mm)RA Replacement Ratio (%)Replacement PatternConcrete TypeVariation in Abrasion ResistanceD, ML, CL, ARS Compare to the Control Type (%)
Bisht and Ramana [54]UntreatedCR: 0.64, 4.5, 5, 5.5 by weightFAGOC b1.27% ↑, 7.59% ↑, 11.39% ↑, 17.72% ↑ (D)
Ridgley et al. [56]UntreatedCR: 0–4.540 by volumeFAGOC 55.71% ↑ (ML)
Thomas and Gupta [57]UntreatedRubber powder: 0.6 (40%) and CR: 0.8–2 (35%) + 2–4 (25%)2.5, 5, 7.5, 10, 12.5, 15, 17.5, 20 by weightFAGHSC7.04% ↓, 19.72% ↓, 21.13% ↓, 19.72% ↓, 20.42% ↓, 27.46% ↓, 28.17% ↓, 32.39% ↓ (D)
Mohammed et al. [59]UntreatedRubber powder: 0.6 (40%) and CR: 1–3 (40%) + 3–5 (20%)10, 20, 30 by volumeFAGRCC10% ↓, 10% ↑, 12.5% ↑ (CL)
Shen et al. [60]UntreatedCR: 1.18–4.7518 by volumeCAGPolymer modified porous concrete13% ↓ (D)
Silva et al. [61]UntreatedCR: 1.18–2.3610, 20, 30, 40, 50 by weightFAGPaving block concrete1.37% ↑, 8.22% ↓, 16.44% ↓, 12.33% ↓, 17.81% ↓ (ML)
Gesoğlu et al. [62]UntreatedCR: 0.1–110, 20 by volumeCAGPervious RC57.78% ↓, 80% ↓ (D)
Thomas et al. [63]UntreatedRubber powder 0.6 (40%) and CR 0.8–2 (35%) + 2–4 (25%)2.5, 5, 7.5, 10, 12.5, 15, 17.5, 20 by weightFAGOC 8.51% ↓, 12.77% ↓, 2.13% ↓, 12.06% ↓, 13.48% ↓, 15.60% ↓, 16.31% ↓, 15.60% ↓ (D)
Sukontasukkul and Chaikaew [64]UntreatedCR: 1.2–510, 20 by weightFAG and CAGPedestrian block concrete303.3% ↑, 1376.7% ↑ (ML)
UntreatedCR: 0.16–1.210, 20 by weightFAG and CAGPedestrian block concrete223.3% ↑, 973.3% ↑ (ML)
UntreatedCR: 0.16–510, 20 by weightFAG and CAGPedestrian block concrete186.7% ↑, 756.7% ↑ (ML)
Gupta et al. [49]UntreatedRubber ash: 0.15–1.05, 10, 15, 20 by volumeFAGOC 5.04% ↑, 8.4% ↑, 16.8% ↑, 19.3% ↑ (D)
He et al. [28]Pre-treating with KMnO4 and NaHSO3Rubber powder: 0.4252, 4, 6 by weightFAGOC 5.1% ↑, 17.9% ↑, 41.1% ↑ (ARS)
Onuaguluchi [38]Pre-coating with LPCR: 0.9–35, 10, 15 by volumeFAGOC8% ↓, 19.9% ↓, 11.7% ↓ (ML)
Segre [27]Pre-treating with NaOH Rubber powder: 0–0.510 by weightFAGOC57.3% ↓ (ML)
a Abbreviations: RA: Rubber aggregate, CAG: Coarse aggregate, FAG: Fine aggregate, D: Wear depth, ML: Mass loss, ARS: Abrasion resistance strength, CL: Cantabro loss, OC: ordinary concrete. b Evaluate standard change compared to the control type (%), corresponding to RA replacement ratio. Increase: ↑, Decrease: ↓.
Table 6. Influence of rubber on the concrete water absorption.
Table 6. Influence of rubber on the concrete water absorption.
ReferenceTreatment MethodRA Type and Size (mm)RA Replacement Ratio (%)Replacement PatternConcrete TypeVariation in Water AbsorptionWater Absorption Compared to the Control Type
Youssf et al. [72]UntreatedCR: 1.18 and 2.3610, 20, 30, 40, 50 by volumeFAGOCa3.23% ↓, 12.90% ↑, 29.03% ↑, 32.26% ↑, 45.16% ↑
Bisht and Ramana [54]UntreatedCR: 0.64, 4.5, 5, 5.5 by weightFAGOC 12.57% ↑, 26.18% ↑, 41.88% ↑, 68.06% ↑
Hunag et al. [73]UntreatedCR: 0–4.7510, 20, 30, 40 by volumeFAGlightweight aggregate concrete1.43% ↑, 14.29% ↑, 28.57% ↑, 35.71% ↑
Thomas et al. [63]UntreatedRubber powder 0.6 (40%) and CR: 0.8–2 (35%) + 2–4 (25%)2.5, 5, 7.5, 10, 12.5, 15, 17.5, 20 by weightFAGOC16.67% ↑, 16.67% ↑, 33.33% ↑, 66.67% ↑, 66.67% ↑, 83.33% ↑, 100.00% ↑, 133.33% ↑
Onuaguluchi and Panesar [68]UntreatedCR: 0–2.35, 10, 15 by volumeFAGOC 12.73% ↑, 14.55% ↑, 27.27% ↑
Benazzouk et al. [74]UntreatedCR: 0–1.010, 20, 30, 40, 50 by volumeCementOC 52.33% ↓, 68.39% ↓, 74.61% ↓, 77.20% ↓, 80.83% ↓
Mohammed et al. [67]UntreatedCR: 0.1–0.510, 25, 50 by volumeFAGOC 5% ↑, 20% ↑, 45% ↑
Gesoğlu and Güneyisi [75]Untreated CR: 0.15–2.05, 15, 25 by volumeFAGOC 2.4% ↓, 6.2% ↓, 1.8% ↓
Mohammed and Adamu [59]UntreatedCR: 0.6 (40%), 1–3 (40%), 3–5 (20%)10, 20, 30 by volumeFAGOC 22.5% ↓, 13.4% ↓, 5.8% ↓
Girskas and Nagrockienė [71]Particle size effectCR: 2–45, 10, 20 by weightFAGOC 14.04% ↑, 30.66% ↑, 41.83% ↑
Thomas and Gupta [57]Particle size effectRubber powder: 0.6 (40%) and CR: 0.8–2 (35%) + 2–4 (25%)2.5, 5, 7.5, 10, 12.5, 15, 17.5, 20 by weightFAGOC 1.52% ↓, 4.55% ↓, 4.55% ↓, 3.03% ↓, 0%, 3.03% ↑, 7.58% ↑, 12.12% ↑
Sukontasukkul and Tiamlom [76]Particle size effectCR: 0.5–3.3510, 20, 30 by volumeFAGOC 15.38% ↑, 19.23% ↑, 42.31% ↑
Ganjian et al. [42]Particle size effectChipped rubber: 2.5–11.05, 7.5, 10 by weightCAGOC 2.38% ↑, 45.24% ↑, 64.29% ↑
Benazzouk et al. [74]Particle size effectRubber powder: 0–1.010, 20, 30, 40, 50 by volumeCementOC 52.33% ↓, 68.39% ↓, 74.61% ↓, 77.20% ↓, 80.83% ↓
Li et al. [77]Particle size effectRubber powder: 0–0.3, 1–230 by volumeFAGSCC10.9% ↓, 24.4% ↓
Kashani et al. [78]Pre-treating with NaOH,CR: 0.9–2.510, 20, 30 by weightTotal solid massLight weight cellular concrete14.29% ↓, 60.71% ↓, 66.07% ↓
Meddah et al. [79]Pre-treating with NaOH, followed pre-coating with resinGranulated CR: 2.5–55, 10, 15, 20, 25, 30 by volumeCAGRCC16.46% ↓, 41.77% ↓, 49.37% ↓, 60.76% ↓, 63.29% ↓, 70.89% ↓
Onuaguluchi [38]Pre-coating with LPCR: 0.9–35, 10, 15 by volumeFAGOC 23.2% ↓, 43.4% ↓, 49.9% ↓
a Evaluate standard change compared to the control type (%), corresponding to RA replacement ratio. Increase: ↑, Decrease: ↓.
Table 7. Influence of rubber on the concrete water permeability.
Table 7. Influence of rubber on the concrete water permeability.
ReferenceTreatment MethodRA Type and Size (mm)RA Replacement Ratio (%)Replacement PatternConcrete TypeVariation in Water PermeabilityPermeability Compared to the Control Type
Thomas et al. [63]UntreatedRubber powder: 0.6 (40%) and CR: 0.8–2 (35%) + 2–4 (25%)2.5, 5, 7.5, 10, 12.5, 15, 17.5, 20 by weightFAGOCa18.42% ↑, 13.16% ↑, 60.53% ↑, 105.26% ↑, 115.79% ↑, 115.79% ↑, 163.16% ↑, 163.16% ↑
Hunag et al. [73]UntreatedCR: 0–4.7510, 20, 30 by volumeFAGOC16.67% ↑, 50.00% ↑, 66.67% ↑
Bisht and Ramana [54]UntreatedCR: 0.64, 4.5, 5, 5.5 by weightFAGOC2.56% ↑, 7.69% ↑, 8.97% ↑, 19.23% ↑
Thomas and Gupta [57]UntreatedRubber powder: 0.6 (40%) and CR: 0.8–2 (35%) + 2–4 (25%)2.5, 5, 7.5, 10, 12.5, 15, 17.5, 20 by weightFAGOC0, 25% ↑, 75% ↑, 75% ↑, 100% ↑, 150% ↑,
150% ↑, 225% ↑
Wang et al. [89]Pre-treating with NaOH CR: 1.44–2.8315, 25 by volumeFAGSCC52.04% ↓, 54.08% ↓
a Evaluate standard change compared to the control type (%), corresponding to RA replacement ratio. Increase: ↑, Decrease: ↓.
Table 11. Influence of rubber on the concrete carbonation resistance.
Table 11. Influence of rubber on the concrete carbonation resistance.
ReferenceTreatment MethodRA Type and Size (mm)RA Replacement Ratio (%)Replacement PatternCarbonation ResistanceCarbonation Depth Compared to the Control Type
Gheni et al. [118]UntreatedRubber powder: <0.0755, 10, 15, 20, 25 by volumeCement a50% ↑, 75% ↑, 150% ↑, 50% ↑, 250% ↑
Thomas et al. [125]UntreatedCR: 2–4 (25%) + 0.8–2 (35%) and rubber powder: 0.6 (40%)2.5, 5, 7.5, 10, 12.5, 15, 17.5, 20 by volumeFAG0, 9.09% ↓, 9.09% ↓, 9.09% ↓, 0, 9.09% ↑, 18.18% ↑, 27.27% ↑
Gupta et al. [49]UntreatedRubber fibres: 2–3 width and 20 in length5, 10, 15, 20, 25 by volumeFAG2.68% ↑, 11.61% ↑, 15.18% ↑, 20.54% ↑, 25.00% ↑
Bravo and Brito [69]UntreatedCR: 15, 10, 15 by volumeFAG14.29% ↑, 21.43% ↑, 42.86% ↑
Pham et al. [113]Pre-treating with NaOHCR: 1–715, 30 by volumeFAG and CAG28.2% ↓, 16.7% ↓
a Evaluate standard change compared to the control type (%), corresponding to RA replacement ratio. Increase: ↑, Decrease: ↓.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, Y.; Chai, J.; Wang, R.; Zhou, Y.; Tong, X. A Review of the Durability-Related Features of Waste Tyre Rubber as a Partial Substitute for Natural Aggregate in Concrete. Buildings 2022, 12, 1975. https://doi.org/10.3390/buildings12111975

AMA Style

Li Y, Chai J, Wang R, Zhou Y, Tong X. A Review of the Durability-Related Features of Waste Tyre Rubber as a Partial Substitute for Natural Aggregate in Concrete. Buildings. 2022; 12(11):1975. https://doi.org/10.3390/buildings12111975

Chicago/Turabian Style

Li, Yang, Jiaqi Chai, Ruijun Wang, Yu Zhou, and Xiaogen Tong. 2022. "A Review of the Durability-Related Features of Waste Tyre Rubber as a Partial Substitute for Natural Aggregate in Concrete" Buildings 12, no. 11: 1975. https://doi.org/10.3390/buildings12111975

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