1 Introduction

The final disposal of automobile tire debris is one of the planet's most significant environmental challenges. Nonetheless, there is a dearth of information about tire end-of-life management concerns [1]. Tire manufacturing has increased dramatically in recent decades as the automobile industry has expanded globally and vehicles have become the dominant source of mobility. As a result, there are currently vast supplies of old tires [2]. Despite the complexity of the tire disposal problem, people have devised innovative solutions. Researchers have updated the benefits of recycling and recovery schemes, as well as life cycle assessments. The prospect of converting old tires into a profitable resource should be studied [1, 3]. To dispose of worn tires, a variety of processes are employed, including landfilling, burning, pyrolysis, carbon black manufacture, and so on. Furthermore, the accumulation of tires presents many threats to public health, the economy, and the environment as a result of contamination of the atmosphere, water, and soil [4,5,6]. Often, the easiest, least expensive, and most effective approach to decomposing old tires is to burn them. However, due to the pollution that excessive smoke and rising temperatures generate in the area, several countries have outlawed this tactic [4, 7]. The high upfront expenditures associated with utilizing tires as fuel make it economically unappealing, even if it is technically possible. Although carbon black from rubber tires may be produced without the requirement for grinding or shredding, the quality and cost of carbon black from rubber tires are inferior to those of carbon black from petroleum oils [8, 9]. Because of their massive manufacturing volume and biodegradability challenges, recycling used tires is a critical concern shared by the scientific community and environmental groups [10]. Figures 1 and 2 depict the appearance of end-of-life tires (ELT) in several European nations in 2018, as reported by the European Tire and Rubber Manufacturers Association (ETRMA) [11]. The findings reveal that France, the United Kingdom, Germany, and Italy produce the most waste tires, with 449.5, 444.5, 439, and 372.5 kilotons, respectively. Malta, Estonia, and Cyprus are near the bottom of the list, with 2.3, 3.3, and 6.9 kilotons, respectively. Figures 3 and 4 indicate the treated proportion of ELT in European nations based on the same ETRMA data [11]. The data from ETRMA used to create a comparison between European countries and the distribution is shown upon the countries labeled with the data using Microsoft Excel Map. ETRMA explained that one way to handle ELT trash is to use it in civil engineering applications. Figure 5 examines the utilization of discarded tires in civil engineering areas throughout Europe between 2010 [12] and 2018 [11]. The comparison suggests that over the course of the two years, the amount of tire debris used in construction operations increased significantly, from 34 to 500 tons. The UK, however, has seen a decrease from 75 to 11.5 tons. A significant amount of concrete used in civil engineering can result from using up of discarded tires, according to research in this area, part of which is included in this article.

Fig. 1
figure 1

Distribution of ELT in kilotons around Europe in 2018 [11]. The data from the resource table is used to create a map of distribution using (Microsoft Excel Map) for better understanding

Fig. 2
figure 2

ELT arising in kilo tones for each European nation in 2018 [11]

Fig. 3
figure 3

Europe's 2018 percentage of treated ELT distribution map [11]. The data from the resource table is used to create a map of distribution using (Microsoft Excel Map) for better understanding

Fig. 4
figure 4

Proportion of each country's treated ELT in Europe in 2018 [11]

Fig. 5
figure 5

Comparing the usage of ELT in the European civil engineering industries between 2010 [12] and 2018 [11]

Multiple analyses took place to figure out the consequences of implementing crumb rubber tires in concrete on various traits of concrete [13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72] and to lessen the environmental risks of scrap tires [73]. The improved concrete properties and Environmental Features of Through Waste Materials encourage additional study into a form of green concrete [74], by replacing cement and aggregate in concrete with industrial by-product materials, tire rubber as an example [75, 76]. Because waste-based LWC requires less strength than traditional concrete for both structural and non-structural reasons, it may thus serve as a considerable replacement for conventional raw materials (cementitious material and aggregates) [77]. As a consequence of the requirement to minimize the quantity of resources needed to create concrete and to improve sections of its performance in ways that are economically and technologically advantageous, a range of waste products have been proposed as additions to cement-based materials [78]. Rubber accumulation may be viewed as a non-degrading material with detrimental environmental consequences. However, recycling the non-biodegradable material to use in concrete mixes would be a helpful approach to dispose of it [15, 79], as shown in Fig. 6. It is a challenge to recycle the nonbiodegradable wastes produced by mining, household, and industrial operations [80, 81]. The purpose of this study was to perform a complete assessment of the utilization of waste rubber in various types of concrete, such as Normal Concrete (NC) [17,18,19,20,21], Self-Compacting Concrete (SCC) [45,46,47,48], High Strength Concrete (HSC) [26, 29], Recycled Aggregate Concrete (RAC) [61, 71], Foamed Concrete (FC) [33], Aerated Cement (AC) [35], Cement Mortars (CM) [37], Concrete pedestrian block (CPB) [38], Pervious concrete (PC) [39], Self-compacting fibrous concrete (SCFC) [45], Concrete paving blocks (CPB) [63], and Fiber reinforced concrete (FRC) [72]. Some of the researchers used tire rubber as a fine aggregate replacement [17,18,19,20,21], coarse aggregate replacement [16, 28, 34, 37], or cement replacement [35]. Table 1 presents an analysis of the literature based on the rubberization dose in the combination, the replacement technique, and the mixture type. This study examined 60 prior research based on their applicability and used data and results to demonstrate the influence of rubber waste on various concrete qualities. The parameters discussed in this study include compressive strength, tensile strength, flexural strength, density, elastic modulus, and workability during slump, V-funnel, and L-box [82].

Fig. 6
figure 6

A. Crumb rubber [68], B. Chipped tire rubber [25], C. Tire–rubber particles [42], D. Crumb rubber aggregate [60], E. Fine aggregate and tire rubber mixture [36]

Table 1 Classification of the literature according to the rubberization dosage, utilization, and mixture type

2 Compressive strength

The compressive strength of concrete, as shown in Fig. 7, is the most essential characteristic. As a flexible building material, it is primarily utilized To resist stresses under compressive pressures at once. [83]. Compressive strength is calculated by dividing the maximum uniaxial load that concrete can support at a given rate by the cross-sectional area of a specimen loaded between two flat, rigid platens with a constant cross-sectional area, plane ends parallel to one another, and at a right angle to the specimen's axis [84]. The literature looked at the compressive behavior of rubberized concrete in a variety of scenarios, including using rubber to substitute fine and coarse aggregates. Toutanji [13] Depending on the volume percentage of rubber chips, it was discovered that the effects of replacing coarse mineral aggregate with rubber aggregate might result in compressive strength losses of up to 75%. In this experiment, rubber chips are utilized in volumetric concentrations of 0, 25, 50, 75, and 100%. Compressive strengths of 31.9, 19.6, 13.8, 9.9, and 7.5 MPa are recorded for the specimens. Whereas, in Wu et al. [25], instead of coarse particles, 100% chipped rubber was used to cast concrete samples, and compressed rubber concrete specimens with a 20% rubber replacement ratio had 35% greater concrete strengths than NAC samples. Compressed rubber concrete specimens with a 30% rubber replacement ratio provide comparable strength to natural aggregate concrete specimens. The results show that using waste chipped rubber in compressed rubber concrete at a rubber replacement ratio of 30% may be used in construction without affecting compressive strength. In another experiment, Sukontasukkul and Chaikaew [38] found that concrete blocks made with crumb rubber performed differently under compression force than their control. Rubber content and particle size both influence the compressive properties of concrete. Increased crumb rubber content resulted in lower strength and stiffness. Also, the results of Yilmaz and Degirmenci [24] supported the reduction of decreasing strength with increasing rubber dosage. Nonetheless, Eldin and Senouci [34] utilized rubber to replace pebbles, reducing compressive strength by up to 85%. When sand was replaced with crumb rubber, compressive strength decreased by up to 65%. However, there is an increase in 1 inch of hardened concrete compressive strength of 19.3% at 1 day, 9.3% at 7 days, and 6.8% after 28 days [61]. In Aslani and Gedeon [72], With a constant 20% fine rubber, both fibers lost compressive strength as compared to the control mix. Adding more did not raise the compressive strength of PP fiber; however, it did increase the steel strength. Also in Thomas et al. [26], by substituting up to 20% of the tiny particles in ABAQUS with crumb rubber, they aimed to create a concrete model. The specimens' compressive strength steadily declined, according to the findings of the comparison. Liu et al. [21] declared that the compressive strength decreased to 31.60, 29.99, 25.38, and 19.33 MPa, respectively, when they added crumb rubber in 1, 3, 5, and 10% ratios to the mixture.

Fig. 7
figure 7

Compressive strength test of rubberized concrete cube specimens [85]

For the fine and coarse aggregates, tire chips and crumb rubber were used, respectively in Güneyisi et al. [14], In line with this, rubber was added to some of the aggregate to create rubberized concrete mixes. The test findings showed a systematic fall in compressive strength from 0 to 50% of the total rubber content. Batayneh et al. [15] examined the possibility of substituting some of the minuscule particles used in concrete mixes with crumb tires. The results of the tests show that lightweight concrete still satisfies strength requirements even when using crumb tires lowers compressive strength. Aiello and Leuzzi [17] increased compressive strength by substituting a portion of the coarse and fine aggregate with rubber waste particles of the same size as the material they were replacing. The compressive strength of both types of rubberized concrete samples steadily reduced as the mineral aggregate was partially replaced with rubber shreds. When they compared samples with 50% and 75% by volume of replacement coarse aggregate, they discovered that compressive strength decreased by 54% and 62%, respectively. Furthermore, as compared to the control mixture, samples with 50% and 75% by volume of fine aggregate substitution exhibited 28% and 37% lower compressive strengths, respectively. Other studies [19,20,21,22,23, 27, 43] employed rubber as a replacement for fine particles and noticed a loss in compressive strength. The effect of rubber content on compressive strength is explained in Figs. 8, 9, 10.

Fig. 8
figure 8

Compressive strength per rubber content in different literatures [13,14,15, 17, 19, 20]

Fig. 9
figure 9

Compressive strength per rubber content in different literatures [21,22,23, 25]

Fig. 10
figure 10

Compressive strength per rubber content in different literatures [26, 27, 38, 43]

3 Tensile strength

The low tensile strength of concrete materials compared to their compressive strength has long been noted [86]. Tensile cracking of concrete reduces the utility and longevity of many projects, despite the fact that concrete is not designed to absorb stress. To establish structural theory under such circumstances, it is necessary to formulate the behavior of concrete under strain [87]. The following literature looks at how crumb rubber affects the tensile strength of concrete. Three distinct types of rubber were employed to partially replace aggregate in rubberized plain pervious concrete mixtures [39]. Furthermore, the usage of rubber reduced the concrete's overall splitting tensile strength. Initially, tire chips are utilized to replace 0, 59.8, and 119.6 kg/m3 of aggregate. Tensile strengths were recorded as 2.06, 1.60, and 1.14 MPa, respectively. In addition, the aggregate was replaced with 0, 48.6, and 97.3 kg/m3 of crumb rubber in the following combinations, yielding 2.06, 1.57, and 1.11 MPa, respectively. Finally, the researchers tried replacing the aggregate with 2.06, 1.56, and 1.49 MPa, respectively. The same study [39] discovered that rubber granules had a twofold influence on the tensile strength of pervious concrete. First, the tiny rubber particles worked to Apart the aggregates from the cement paste and one another, resulting in poorer bonding between the mixture's main components. Second, with pervious concrete, larger rubber particles act as reinforcing fibers, limiting strength loss. Aslani et al. [45] combined a rubber weight of 87.5 kg/m3 with steel fibers. The results show that adding steel fiber increased the bridging effect of the combinations, resulting in a high tensile strength. When heated at all tested temperatures, the SCRC combination with up to 0.75% steel fiber volume fraction increased tensile strength. Ling et al. [63] discovered that the splitting tensile strength varied between 0.72 MPa and 5.25 MPa when the rubber content and curing age increased from 0 to 30% and 1 day to 365 days, respectively. As the rubber aggregate content rose, the splitting tensile strength decreased.

For making crumb rubber, two distinct rubber sizes are used: grade 18 and grade 5 [48]. While the grade 5 CR has a 1-mm filter and goes through a 4-mm sieve, the grade 18 CR is a fine material that only passes through a 1-mm sieve. In several trials, the doses were 0, 5, 10, 15, 20, and 25 by volume. The control mix had the highest splitting tensile strength, which progressively declined as the rubber portion increased. The splitting tensile strength of grade 18, grade 5, and incorporated crumb rubber-containing concrete dropped by 18.47%, 23.86%, and 21.44%, respectively, at a concentration of 5%. Conversely, it fell by 80%, 94.64%, and 92.92% in 25% increments. Furthermore, Hesami et al. [50] found that the characteristics were more affected by coarse rubber particles than by microscopic ones. An investigation was carried out by Hesami et al. [51] to ascertain if a higher rubber content was linked to a lower tensile strength. The amount of rubber utilized to replace a rate of 15% of the fine-grained aggregate volume in the absence of fiber causes a 14.29% drop in tensile strength. Moreover, Youssf et al. [53] determined that increasing the number of rubber fibers by 2%, 3%, or 1% reduced the tensile strength of rubcrete by approximately 7%, 15%, and 15%, respectively. Because of its high Poisson's ratio and flexibility, rubber fiber has been offered as a method of removing cracks from concrete and boosting its tensile strength. The flexibility and surface qualities of the rubber indicate that there is insufficient cementitious material to fix the rubber to the surrounding material. At 15% and 30% rubber content, indirect tensile strength decreased by 14.5% and 28.5%, respectively. Also, more other experiments [55, 57, 66] recorded a reduction in tensile strength due to an increasing crumb rubber ratio. The application of tensile strength concrete of rubberized concrete specimens is shown in Fig. 11. In Figs. 12 and 13, the effect of the crumb rubber ratio on the tensile strength is explained.

Fig. 11
figure 11

Tensile splitting test on a cylindrical specimen [88]

Fig. 12
figure 12

Tensile strength per rubber content in different literatures [39, 48]

Fig. 13
figure 13

Tensile strength per rubber content in different literatures [51, 53, 55, 66]

4 Flexural strength

As an important mechanical property, the researchers tried to study the effects of different use conditions of rubber on the flexural strength of concrete, as shown in Figs. 14, 15, 16. Thomas and Gupta [29] observed that after 7 days, mixtures containing 0% and 2.5% crumbs of rubber had the strongest bending strength of 6.2 MPa, while those containing 17.5% and 20% crumbs rubber had the lowest of 4.6 MPa. The longest possible period is 28 days. The total amount of 2.5% crumb rubber stipulated the most significant value of 7.3 MPa, while the mixture comprising 20% crumb rubber imparted the lowest value of 5.5 MPa. Over 90 days, an identical pattern arose, with the highest and lowest pressures of 7.9 and 5.7 MPa, each. Following 90 days, the rubberized material (20% crumb rubber) indicated 28% lower flexural strength than the original control blend sample. Paine et al. [32] attempted to employ crumb rubber as a fine aggregate substitute in doses of 0, 2, 4, and 6%. When the rubber percentage ranges between 0 and 2%, flexural strength increases from 3.50 to 3.60 MPa, according to the researchers. However, when the rubber content increased, the flexural strength declined to 2.80 and 2.70 MPa, respectively. Similarly, Eiras et al. [33] discovered that as the rubber % increased, flexural strength decreased. The loss of strength is due to an insufficient contact between the CR and the cementing matrix. In another study, Benazzouk et al. [35] examined an aerated cement composite's physio-mechanical properties with rubber waste particles to create materials appropriate for cellular concrete applications. By substituting a novel proteinic air-entraining agent for air gaps and adjusting the quantity of rubber particles used instead of cement, we were able to artificially aerate the material. The findings of this study's flexural strength analysis showed that changing the porosity structural structure of the matrix significantly changed the behavior of the composite. The flexural strength of the ACRC sample is lost. The value decreases from 3.3 to 1.4 MPa. Because of the trapped air effect, this data indicates that the sample's porous structure has a bigger influence than the elasticity of the rubber aggregate. The data also reveal that, for a particular rubber volume ratio, the loss of flexural strength is smaller than the loss of compressive strength.

Fig. 14
figure 14

Flexural test arrangement [66]

Fig. 15
figure 15

Flexural strength per rubber content in different literature [50,51,52]

Fig. 16
figure 16

Flexural strength per rubber content in different literature [33, 35, 46, 48, 50]

In another study, Mishra and Panda [46] found that tire tread-rubberized concrete can be useful in specific circumstances when strength and mechanical properties are not the most critical factors. In the dosage of 0, 33.8, 67.5, 101.3, 135, 168.8, and 202.5 g, the tire represents 2.5 weight percent, 5 weight percent, 7.5 weight percent, 10 weight percent, 12.5 weight percent, and 15%wt of the concrete, resulting in flexural strengths of 9.00, 7.50, 5.70, 5.25, 4.30, 3.50, and 2.90 MPa, respectively. The samples were unable to provide the necessary strength using rubber concrete. It has been noted that the flexural strength of self-compacted concrete gets better with longer curing times. It is also observed that the flexural strength decreases as the fraction of rubber replacement increases. SCRC0 is more potent than the other mixtures. For example, at 7 days, 28 days, and 90 days, the strength difference in SCRC5 compared to the control specimen is 3.63%, 3.68%, and 2.96%, respectively, but in SCRC20, the strength change is 21.82%, 21.32%, and 17.74% at 7, 28, and 90 days, respectively. The implications of crumb rubber composition and size on the hardened traits of self-compacting concrete were investigated as well in a different investigation [48]. This study revealed that the highest net flexural strength of the reference mix was 5.6 MPa. The strongest net flexural strength was detected in SCC manufactured with grade 18 crumb rubber, while the lowest was reported in those established with grade 5 crumb rubber. The mixed crumb rubber, grade 5, and grade 18 in the control mix had a greater net flexural strength than the other combos, corresponding to the data. The data [49, 51] also declare the negative effect of the rubber content on flexural strength.

Medina et al. [50] examine the flexural characteristics of sustainable concrete composed of partly coated steel or plastic fibers and crumb rubber. On the other hand, CR is frequently used as aggregate in concrete. FCR is composed of fibers that are strewn with recycled tire rubber fragments. Revalorized as an addition, these rubber-coated fibers offer better-bending strengths than normal rubberized concrete manufactured with crumb rubber. Benazzouk et al. [52], two different kinds of tire rubber were utilized. CRA has smooth surfaces, whereas ERA has soft aggregates and alveolar surfaces. Studies on the flexural strength of CRAC and ERAC with varying rubber contents show that maximum values occur at volume ratios of roughly 20%. At 35% rubber content, the link between the rubber and cement matrix weakens, significantly lowering flexural strength. This decrease becomes more pronounced as the size of the rubber aggregate grows.

5 Density

Benazzouk et al. [35] found that rubber particles caused a linear drop in dry unit weight during their experiment. For samples of crumb rubber concrete and aerated crumb rubber concrete that include 50% rubber particles, respectively, the density values fall from 1910 kg/m3 to 1150 and 785 kg/m3. These two figures show reductions of up to 40% and 59%, in relation to one another. At concentrations of 2.5, 5, 7.5, 10, 12.5, and 15%wt tire rubber was used in place of fine aggregate Oikonomou and Mavridou [36]. The sample densities of 2.23, 2.11, 2.03, 1.94, 1.76, and 1.68 g/c m3 were observed, indicating a significant reduction. Additionally, crumb rubber dosages were raised to 0, 5, 10, and 15% in Onuaguluchi and Panesar [40], and the corresponding dry density values were 2402, 2360, 2340, and 2233 kg/m3. Xue and Shinozuka [41] stated that as the amount of rubberized concrete grew, the density of the concrete reduced because of the low specific gravity of the rubber crumbs. The typical concrete set's average density dropped to 2475 kg/m3, and for the 5%, 10%, 15%, and 20% rubber replacement ratios, it was 2374, 2273, 2171, and 2069 kg/m3, respectively. Aslani et al. [47] found that when the percentage of recycled material was increased, the density of the concrete decreased while the rubber content stayed constant at 65.62 kg/m3. When CR is used in place of some of the sand in the concrete, the unit weight decreases. According to Rezaifar et al. [64]'s estimation, the drop is around 4, 8.5, and 12.5% for 10, 20, and 30% CR replacement, respectively. MK raised the unit weight of the concrete somewhat, but not much. Benazzouk et al. [52] examined the disparities in density between two rubber varieties with differing particle sizes in an alternative experiment. Depending on the rubber volume ratio, losses of as much as 22% and 35% of the dry mass were seen for CRAC (smooth surfaces) and ERAC (soft aggregates with alveolar surfaces), respectively. It is crucial to take into account that a larger unit weight reduction is an indication of a smaller aggregate size. When expanded rubber is used, the effect is stronger. The necessity for water to achieve a regular functional consistency might be one reason for the difference in aggregate size. Studies by Liu et al. [56] and Aliabdo et al. [54] also reported the reduction of the density of concrete by increasing the rubber content, as shown in Fig. 17.

Fig. 17
figure 17

Concrete density per rubber content in different literatures [35, 41, 52, 54, 56]

6 Elastic modulus

Similarly, as Figs. 18, 19, 20 illustrate that further studies have looked into how the amount of rubber crumb in concrete alters its modulus of elasticity. According to Pacheco-Torgal et al. [37], the elasticity modulus decreases by 21.4% when the replacement volume of polymeric wastes (polyethylene terephthalate bottles and tire rubber) is increased by 40%. Furthermore, the results indicate that the elasticity modulus of mixes containing tiny PET particles is lower. Additionally, rubber usage decreased compressive strength in Gesoğlu et al. [39], It led to a considerable reduction in the elastic modulus of pervious concrete. made of porous concrete The size, volume, and compressive strength of the rubber particles affected the rubber's modulus of elasticity. A decrease in paste quantity after an increase in aggregate amount resulted in a statistically significant decline in static elastic modulus. According to Onuaguluchi and Panesar [40], the combination comprising coated crumb rubber had a maximum drop in static modulus of 25% as compared to the mixture incorporating crumb rubber, which had a maximum decrease of 29% as received. The loss of static modulus in similar combinations including coated crumb rubber and silica fume is limited to a maximum of 18%. In general, a number of variables, including the properties of the aggregate, the mix design, the curing environment, and the usage of mineral admixtures, influence the concrete's modulus of elasticity. According to Khaloo et al. [42], the ultimate strength and tangential modulus of elasticity of concrete are dramatically decreased when tire rubber particles are substituted for mineral aggregates. Because of the considerable reduction in ultimate strength, we do not advise using rubber concentrations higher than 25%. The mechanical qualities of tire rubber concrete may be enhanced by taking into account tire particle surface preparation. To find out how rubber's mechanical characteristics affect rubberized concrete's ultimate strength, more research is required. In keeping with the observed splitting and compressive tensile strengths, Hilal [48] also noted a decrease in the static elastic modulus as the rubber size and content rose. The results demonstrated that the maximum static elastic modulus was obtained when grade 18 crumb rubber was used to construct SCC. Rubber aggregate's decreased flexibility and bonding strength to the cement matrix cause it to gradually lose its elastic modulus when added to concrete, according to Medina et al. [50]. There is a larger decrease when rubber is used in place of all the stone aggregate.

Fig. 18
figure 18

Test setup for the dynamic modulus of elasticity using the elastic wave approach [89]

Fig. 19
figure 19

Elastic modulus per rubber content in different literature [37, 42, 55]

Fig. 20
figure 20

Elastic modulus per rubber content in different literature [48, 50]

7 Slump

The use of different-sized rubber particles as fine aggregates in concrete affects workability and water permeability more than the density and initial strengths of the concrete do [58]. A slump flow test used to examine the impact of rubber aggregates on the workability of self-compacting concrete [16] reveals that adding rubber aggregates significantly increased the superplasticizer concentration in order to maintain the slump flow within the desired range. Rubber aggregates form an interlocking structure that is resistant to the typical flow of concrete under its own weight, which explains the phenomenon. Raffoul et al. [44] showed that rubber levels ranging from 0 to 10% FR had no influence on the original mix's flow. Conversely, a 100% FR replacement reduced the flow by 30% and made up 45% of the total volume. The latter combination was severe, rigid, and inconsistent. Replacing fine aggregate with rubber particles has a significant impact on the slump flow diameter. The slump flow diameter decreases with an increase in the rubber particle replacement ratio. According to different research [62], as the rubber particle replacement ratio rises from 0 to 50%, the slump flow diameter of fresh SCRLC drops from 785 to 580 mm. There is around a 26% decrease in the slump flow diameter. SCRLC flowability and flow rate both rise with an increase in the rubber particle replacement ratio, according to changes in slump flow diameter. According to research [90], the slump of concretes containing crumb rubber aggregates significantly reduced as the quantity increased. For instance, the concrete's slump values dropped from 220 to 185 mm, yet the amount of crumb rubber increased from 0 to 25%. Concretes with 20 and 40% GGBFS replacement were found to have 200 mm slump values, and the amount of SP consumed decreased steadily as the GGBFS replacement ratio was increased. Waste tires had minimal effect on the workability of the amended concrete in Li et al. [28]'s study. In the Youssf et al. [59], concrete with 0%, 5%, 10%, and 15% SF concentrations sagged 1.22, 1.27, 2.53, and 7.75 times more when 20% rubber was added than when 0% rubber was used. Topçu and Bilir [31] looked at how slump workability decreased. The application of the slump test of rubberized concrete sample is shown in Fig. 21. The correlation between rubber content and concrete slump flow is seen in Figs. 22 and 23.

Fig. 21
figure 21

Slump flow test of rubberized concrete [62]

Fig. 22
figure 22

Slump flow diameter per rubber content in different literature [16, 44, 62, 67, 68]

Fig. 23
figure 23

Slump flow diameter per rubber content in different literature [30, 46, 51]

8 V-funnel

The researchers [30, 69,70,71] determined the relation between rubber content and workability during the V-funnel experiment, as shown in Fig. 24 and 25. Researchers Raj et al. [30] utilized the V-funnel test to measure the viscosity of SCRC mixtures; the time it took for the combinations to exit the V-funnel aperture is shown in the table. The concrete strength increased the amount of time needed for the concrete to empty the funnel, as seen by the V-funnel time. Regardless of the strength of the concrete, the investigation showed that variations in rubber composition increase over time. Rubber addition made SCC solutions appear more viscous, which makes sense considering that SCRC has an 18% longer V-funnel duration than SCC. The addition of crumb rubber altered the V-funnel test results in Younis et al. [69]. The V-funnel timings for mixes 10, 20, 30, and 40% were 9.8, 12.3, 15.4, and 18.3 s, respectively. As the replacement ratio of crumb rubber grows, the V-funnel time also rises. This indicates that, in comparison to the control mix, the mixes containing crumb rubber were more viscous. Additionally, the findings from Mohamadien et al. [70] and Popović et al. [71] demonstrate that a rise in rubber content ratio is accompanied by an increase in V-funnel time.

Fig. 24
figure 24

V-funnel test application of rubberized concrete [62]

Fig. 25
figure 25

V-funnel time per rubber content in different literature [30, 69,70,71]

9 L-box

The L-Box test as shown in Fig. 26, assesses the quantity of reinforcing that is obstructing the concrete's flow, as stated by Mishra and Panda [46]. The L-box data from Raj et al. [30] show that independent of compressive strength, the passing ability ratio falls with increasing rubber content and rises with concrete strength. Tire rubber crumb (TRC) is available in a range of shapes and sizes. Some have jagged edges, while others are more unevenly shaped. TRC reduces the amount of water required for cement hydration and aggregate lubrication by absorbing water from the slurry. The aggregate form is one distribution aspect that influences the concrete's consistency. Coarser aggregates require more energy to overcome internal friction and flow more easily. Furthermore, the fibers inhibit aggregate segregation and concrete flowability by penetrating the concrete and creating a three-dimensional web. As a result, internal friction rises and SCC's rheological properties decrease [51]. Furthermore, Lv et al. [62] realized that when the rubber particle substitute ratio increases, the L-box ratio decreases. When the replacement ratio varies between 0 and 50%, the rubber particle substitute ratio drops from 0.98 to 0.82. The L-box modifications show that as the rubber particle replacement ratio increases, so does the self-compacting rubberized lightweight concrete's transit capacity. As previously indicated, the form and surface features of rubber fragments may reduce the capacity of the new SCRLC path. These results and more are shown in Fig. 27.

Fig. 26
figure 26

V-funnel time per rubber content in different literature [62]

Fig. 27
figure 27

L box height ratio per rubber content in different literature [30, 62, 67, 68]

10 Analytical correlation

Globally, researchers studying engineering materials are trying to create oblique connections between each tangible property by using mathematical correlations, a knowledge of the relationship between the properties themselves, and the distribution of data among them. As seen in Figs. 28, 29, 30, 31, we attempted to establish a connection in this article between compressive, tensile, and flexural strength in order to comprehend the relationship between these previous mechanical parameters. The relationship between compressive and tensile strength is represented in Eq. 1, while the relationship between compressive and flexural strength is represented in Eq. 2.

$$ Ts = 0.0461*Cs + 0.9252 $$
(1)
$$ Fs = 0.086*Cs + 1.0297 $$
(2)
Fig. 28
figure 28

Data distribution between compressive and tensile strength from different literature [14, 15, 19, 37]

Fig. 29
figure 29

Linear regression between compressive and tensile strength [14, 15, 19, 37]

Fig. 30
figure 30

Data distribution between compressive and flexural strength from different literatures [13, 15, 19, 24]

Fig. 31
figure 31

Linear regression between compressive and flexural strength [13, 15, 19, 24]

11 Conclusions

In conclusion, the review of the mechanical and fresh properties of rubberized concrete underscores its potential as a sustainable solution in the construction industry. The globalization of the automotive sector has led to an accumulation of used tires, necessitating innovative recycling strategies. Our analysis of 60 previous studies reveals the multifaceted impacts of incorporating waste rubber into concrete mixtures. From our synthesis, it is evident that rubberized concrete offers several advantages, including comparable compressive, tensile, and flexural strengths, in addition, reduced density was detected by most scientists. When compared to conventional concrete. These properties make rubberized concrete a viable option for various construction applications, particularly in areas where strength, seismic resistance, and durability are not primary requirements. In Light of the previous studies, there are detected drawbacks of using rubber on concrete properties that we can address as follows:

  • In certain studies, it was shown that substituting rubber aggregate for coarse mineral aggregate might cause compressive strength reductions of up to 75%. For structural applications where compressive strength is crucial, such as in compression members, it renders this kind of concrete useless.

  • Since concrete is known to have a far lower tensile strength than compressive strength, adding rubber to the concrete is not a practical way to overcome this weakness, but it weakens and damages concrete's tensile strength under normal circumstances.

  • Ineffective bonding between the rubber particles and the concrete mixture results in a loss in strength, rendering this form of concrete unusable for structural applications where a great bond between concrete composites is required.

  • Decrease in workability according to the investigations of Slump, V-funnel, and L-box experiments. Rubber aggregates formed an interlocking pattern that opposes the normal flow of concrete under its own weight, which was the explanation given for the effect. This causes expected concrete kinds inappropriate for tight passageways and situations requiring highly workable concrete.

  • Rubberized concrete has unfavorable characteristics when it comes to possible durability problems and long-term performance. This indicates that it is unsuitable for use in circumstances requiring freeze–thaw resistance. furthermore, unsuitable in environments containing acid and chloride because of rubber's low resistance in these circumstances.

12 Recommendations for future work

Future research should focus on optimizing rubberized concrete mix designs, exploring alternative reinforcement methods, and conducting long-term durability studies to address these challenges effectively. Overall, our review contributes to a deeper understanding of the benefits and limitations of utilizing waste rubber in concrete production, paving the way for more sustainable practices and environmentally friendly construction materials. Continued research and practical implementation of rubberized concrete will play a crucial role in advancing the goals of sustainable development and resource conservation in the construction industry.

In order to understand the impacts of employing rubber as an aggregate substitute in zero-cement composites, it is important to consider the paucity of research on the effects of rubber particles in geopolymer concrete. In terms of binder and aggregate, it will be a huge step toward fully ecologically friendly concrete. On the other hand, in order to prevent concrete weakness and cracking, it is crucial to address the inadequate bonding between the rubber particles and the concrete mixture.

The analytical models are now trends in the field of concrete technology depending on Machine Learning through Artificial Neural Networks (ANN), Linear Regression (LR), Non-linear Regression (NLR), and Multi-linear Regressions (MLR). Prediction of different properties of rubberized concrete depending on the literature will be a great topic for future investigations and help for a better understanding of how this concrete type behaves.