Next Article in Journal
Development of WO3–Nafion Based Membranes for Enabling Higher Water Retention at Low Humidity and Enhancing PEMFC Performance at Intermediate Temperature Operation
Next Article in Special Issue
Analysis of Microplastics in Aquatic Shellfish by Pyrolysis–Gas Chromatography/Mass Spectrometry after Alkali Digestion and Solvent Extraction
Previous Article in Journal
Performance-Based Evaluation of Healing Efficiency on Mechanical Properties of Self-Healing Cementitious Materials Incorporated with PMMA/Epoxy Microcapsule
Previous Article in Special Issue
Mechanical Performance of Polymeric ARGF-Based Fly Ash-Concrete Composites: A Study for Eco-Friendly Circular Economy Application
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Performance Optimization Approach of Polymer Modified Asphalt Mixtures with PET and PE Wastes: A Safety Study for Utilizing Eco-Friendly Circular Economy-Based SDGs Concepts

1
State Key Laboratory of High-Efficient Mining and Safety of Metal Mines, University of Science and Technology Beijing, Ministry of Education, Beijing 100083, China
2
Department of Civil Engineering, Pakistan Institute of Engineering and Technology, Multan 66000, Pakistan
3
College of Engineering and Technology, American University of the Middle East, Egaila 54200, Kuwait
4
College of Transportation Engineering, Tongji University, Shanghai 201804, China
*
Authors to whom correspondence should be addressed.
Polymers 2022, 14(12), 2493; https://doi.org/10.3390/polym14122493
Submission received: 14 May 2022 / Revised: 12 June 2022 / Accepted: 12 June 2022 / Published: 19 June 2022
(This article belongs to the Special Issue Polymers and the Circular Economy Model)

Abstract

:
Eco-friendly waste utilization helps in the development of sustainable infrastructures. Recently, researchers have focused on the production of road infrastructures using the circular economy concept of human safety. The objective of this study is to investigate and explore the utilization of optimum polymer waste content for the development of polymer-modified asphalt mixtures using response surface methodology (RSM). RSM based on Box–Behnken design (BBD) was employed to optimize experimental design and included three factors: X1, polymer type; X2, polymer contents; and X3, testing day. The optimized responses determined by the RSM were as follows: MS of 42.98 kN, MF of 5.08 mm, and MQ of 8.66 kN/mm, indicating a favorable and consistent precision in comparison with experimental values. Moreover, the Marshall characteristics of samples prepared with PE were quite improved compared to PET. In conclusion, the incorporation of such polymer wastes in road construction is a sustainable and cost-effective way of improving their engineering properties. This study will help in the development of sustainable road infrastructures supporting human safety and environmentally friendly practices.

1. Introduction

Bitumen, also known as asphalt binder, is a petroleum by-product produced during the distillation of hydrocarbon chains. It has valuable properties, such as long-term durability, high stability, and water resistance, which allow it to be used as a road building material [1]. Bitumen is used with aggregates as a binding material in the production of asphalt mixtures for infrastructure projects. The performance parameters and ultimate stability of asphalt pavements are heavily reliant on the binder’s functionality. The degradation of the bitumen binder causes a wide range of road pavement failures such as cracking failures, rutting, and fatigue cracking [1,2]. Moreover, the increasing cost of bitumen has significant effects on the overall project cost.
Although the repairs and maintenance cost of road pavement are unfavorable for socioeconomic and climatic reasons, much effort is expended to minimize such aforementioned failures. A significant number of researchers have studied bitumen alteration in order to achieve improved durability and high stability and strength. In the past, countless initiatives have been explored, and scholars have effectively utilized numerous waste materials e.g., crumb rubber [3,4], nylon and polymer waste [5,6,7], polyethylene [8,9], polypropelene [10], recycled PET [11,12], polystyrene [13], and ordinary plastic bottles [14,15,16], etc. The positive enhancement of all these leading bituminous qualities that are crucial for road construction is recognized as a primary goal of all these investigations [17,18]. Polymer matrices of bitumen are one of the most suitable and widely used approaches among all reported modification approaches [19]. To produce polymer-based asphalt, the polymer is mixed into the bituminous mix by a chemical process (wet method) or mechanical mixing (dry process) as shown in Figure 1 [20,21]. The polymer and asphalt are directly incorporated at elevated temperatures for a specific duration of time in the wet procedure to ensure adequate chemical or physical reactions between ingredients. The wet process integrates plastic in the shape of flakes, pellets, or powder into warm bitumen when a polymer is processed from plastic wastes [22]. Plastics chosen for the wet method often require melting ranges just below the defined range, owing to standard blending temperatures between 160 to 170 °C [12,23].
Response surface methodology (RSM) is becoming a common practice in the construction sector as a more effective way of assessing and optimizing outcome measures [25]. Tan et al. [26] employed central composite design, also known as the CCD approach, in combination with the RSM model to find the best raw ingredient ratio for several asphalt mixture test parameters. The impacts of the prepared factors, such as grading, aggregation, and other factors, on the indirect tensile strength (ITS) of hot mix bituminous compositions using RSM were researched and assessed by Kavussi et al. [27]. Hamzah et al. [28,29,30] used the CCD of RSM to explore the effects of recovered aggregate ratio, compaction temperature, and bitumen percentage on the physical and mechanical parameters of asphalt pavement. In addition, Hamzah et al. [29] discovered that statistical methods built on RSM can accurately predict the impacts of aging on binder viscoelastic performance at extreme temperatures. RSM was also used and studied by another researcher [31] to assess the results of lime concentration and gradient on the ITS and its ratios in an asphalt mix under dry and wet situations. RSM-based statistical model was used to optimize the binder concentration and gradation of asphalt mixture using the tensile strength rate. RSM can be performed to effectively analyze bituminous binder and combination based on comprehensive literature [32].
The statistical method based on RSM to quantify the optimum polymer contents of PET and PE wastes in the pavement mixture is still not presented in the literature. Therefore, the main objective of this study is to optimize the polymer-modified asphalt mixture made with PET and PE wastes by employing the statistical approach of RSM. This statistical technique can measure the relationships and correlations among several input parameters by taking into consideration the contributions of two or more variables in an investigation. The results are examined and investigated theoretically, experimentally, and hypothetically using a statistical approach based on RSM to give a better understanding of these additives and polymer contents on Marshall test parameters. Moreover, this research aims to meet the eco-friendly, efficient, and economical requiremens for the development of road networks that will be helpful for all countries to utilize their plastic waste as a binder (additive) for the construction of extremely durable, inexpensive road systems, as well as reducing air and noise pollution in metropolitan areas.

2. Background

2.1. Literature Review

The flexible pavement comprises multiple materials, including bitumen (binder), filler, and fine and coarse aggregates. For the effective production of asphalt concrete mix, the percentage of each material appears to be most significant for efficient and economical construction objectives [33]. It is important to ascertain the fundamental properties of every ingredient for a better understanding and development of the asphalt mix. The load-bearing material of asphalt concrete is recognized as an aggregate because around 95% of aggregate by weight is involved in the production of asphalt mix [34]. In civil engineering projects, aggregates are generally utilized for almost every type of work, including concrete production, asphalt mix, railway ballast, road formation, etc.; they are in-active chemical materials that are blended with various binding substances to form concrete, asphalt mix, mortar, etc. [35,36,37].
Typically, each type of road deteriorates over time, principally due to countless considerations including severe weather situations, traffic loading, and asphalt and bitumen quality [38,39]. Worldwide, numerous scholars have investigated countless modifiers to enhance the quality and performance of bitumen through the utilization of a variety of modified bitumen blends. The literature shows a few studies that utilized additives such as fatty acid amides, sulfur, polyphosphoric acid, etc. [40,41,42,43]. Meanwhile, other polymers a utilized to enhance the characteristics of bitumen; the most efficient and widely utilized polymers for modification are recognized as styrene–butadiene rubber (SBR), styrene–butadiene–styrene copolymer (SBS), etc. [44,45,46,47]. For many experimental and theoretical analyses, polyethylene and ethylene–vinyl acetate have also been utilized [48]. The practical properties of normal bitumen have been enriched through the application of these valuable modifiers. Additionally, the costs of bitumen processing and road construction have been affected by the blending of additives with bitumen. The foremost aim of the entire process should be the evaluation of economic benefits [49]. Currently, the global enduring target of sustainability can be achieved through the application of waste raw materials or additives to systematically handle inadequate natural resources [50]. The aggregate is a key part of asphalt pavements obtained from natural resources. A large investigation was performed to study the utilization of recycled and solid waste materials as an alternative to aggregates through an assortment of raw materials such as glass, finely crushed waste bricks (FCWB), steel waste slag (SWS), crumbed rubber tires (CRT), etc. [45,51,52,53]. As mentioned above, several polymers successfully used to enrich the basic bitumen characteristics at an advanced level, such as polypropylene and polystyrene through useful reclaimed geo-membranes, have also been employed to stimulate anticipated attributes in bitumen [52,54,55,56,57]. The general properties of bitumen also improve through the accumulation of natural rubber [58]. Concerning the utilization of recycled materials, the recycled waste polyurethane foam has been efficiently employed as a substitute for bitumen in asphalt pavements; in addition, its durability has been examined for lightweight mortars [59,60,61,62,63].
In asphalt pavement, because of weak adhesion between bitumen and aggregates, the deterioration and destruction of road pavement surfaces occurs. After the advancement of adhesion, the phenomenon of stripping, concerning the surface tension among aggregate and binder, seems to be improved, reducing the surface deterioration. In the past, through statistical approaches, many asphalt pavement distress attributes, such as rutting, cracking, and shoving, etc. have been investigated as the cause of failure in these pavements [64,65]. The key cause of these distress attributes is the selection of inadequate mixing proportions, which leads to the deterioration of these pavements; otherwise, the best service quality of these pavements can be achieved through correct ingredient proportions [66,67]. In the case of hot mix asphalt pavement, the temperature variations have precisely impacted the viscosity of bitumen and then the adhesion properties are altered through this variation, which disturbs the bond between aggregates and bitumen [68]. The low stability of asphalt concrete leads to several types of distress, and the stability of specimen is identified through the maximum load resistance that can be borne by specimen before the failure [65]. Scientifically, the Marshall-based stability and tensile strength of specimens are related to each other, because the testing arrangement of the Marshall and the indirect tensile test appears to be similar, although the sample rested in the Marshall test head lies within a specified limit [69,70,71]. Conversely, countless constraints, such as the shape and size of the aggregates, weather variations, grade and content of bitumen, as well as loading situation of traffic, etc., also influence the stability of asphalt pavements. Through the application of stability tests, the rapid and efficient evaluation of the elastic behavior of asphalt concrete layers can be achieved [72]. Worldwide, this approach can be extensively utilized, although there are overwhelming constraints, and it is an expensive and time-consuming method [73,74,75].
The use of statistical methods based on Response Surface Methodology (RSM) to quantify the optimum polymer contents of PET and PE wastes in the pavement mixture is still not found in the literature. Therefore, the main objective of this study is to optimize the polymer-modified asphalt mixture made with PET and PE wastes by employing the statistical approach of RSM.

2.2. Application of Circular Economy Concepts and Sustainable Development Goals

Sustainable Development Goals, also known as SDGs, are the roadmap toward achieving greater global sustainability by 2030 [76]. SDGs are based on 17 global goals that are directly interlinked with 169 benchmarks. These goals are progressively being used as a foundation for organizing and promoting sustainability practices by both private and state sector stakeholders around the world. Similarly, the concept of the Circular Economy (CE) has recently become more popular to achieve sustainability at local and international levels [12,77]. Several drivers, such as states, municipalities, and many multinational organizations, are actively studying and promoting ways to practice CE initiatives on the global scale [78].
Few research papers were found when we searched for details about the connection between the CE and the SDGs [79,80]. The study examines the extent to which CE initiatives are significant for achieving the SDGs in developing countries based on an in-depth literature overview. The study concluded that CE concepts can be used as a toolbox and concrete action tactics for reaching a significant number of SDGs. The immediate relationship or collaborations among CE and SDGs are found in SDG 6: clean water and sterilization, SDG 7: affordable and clean energy, SDG 8: respectable work and financial development, SDG 12: sustainable utilization and creation, and SDG 15: life on the planet [79]. These five SDGs are explained one by one as follows:
  • The treatment of wastewater is a promising approach to the CE model based on the SDG 6 to address water pollution and increase current water supply–demand through recycling and re-use in the future.
  • SDG 7 is to ensure that everyone has access to affordable, efficient, renewable, and smart energy.
  • SDG 8 promotes sustainable economic growth and decent work for everyone.
  • SDG 12 is to ensure a sustainable framework of consumption as well as the production process.
  • SDG 15 is relevant to the restoration of natural capital and terrestrial ecosystems in a sustainable way.
The four other SDGs indirectly linked to the CE model are SDG 1 (no poverty), SDG 2 (zero hunger), SDG 11 (sustainable cities and communities), and SDG 14 (life below water). The production phase of circularity is based on four approaches (1 to 4) that attempt to maximize the usage of renewables while minimizing value loss across the life cycle. The consumption phases (such as sharing, re-using, repairing, re-manufacturing, and recycling) of circularity are based on six approaches (5 to 10) that prevent value leakage by rotating products and materials at their optimum utility levels. The three basic principles and ten corresponding strategies regarding circular economy [81] and SDGs are presented in Figure 2.

3. Materials and Methods

In this study, the research methodology has been divided and performed into two stages: polymer-modified bitumen and polymer-modified asphalt mixes. The evaluation of various plastic contents is performed by both fresh and mechanical testing. The engineering properties, such as Marshall stability flow and stiffness of polymer-modified asphalt mixtures, were comprehensively examined for the in-depth analysis of their behavior and performance under local environmental conditions. The research methodology of the current study is presented stepwise in the form of the flowchart in Figure 3.

3.1. Basic Materials

Bitumen is a sticky and highly viscous petroleum liquid, black in color, and taken from deep natural deposits as well as acquired through the fractional distillation of crude oil while boiling at 525 °C. It is composed of hydrocarbon compounds. Worldwide, around 70% of bitumen is utilized for asphalt pavements as a binder; the remaining 30% is utilized for many applications, including building protection, waterproofing, etc. [82,83]. In this investigation, 60/70 grade bitumen was used as a binder, and an optimal percentage content of 7% binder was found. For a detailed understanding of binder behavior, its physical properties were examined through the laboratory equipment according to the ASTM standard listed in Table 1 [84,85,86,87,88].
Sakhi Sarwar aggregate, which comprises 55% limestone and 45% sandstone, has been utilized as a coarse aggregate in this study. The coarse aggregate size varied from 20–4.75 mm, and the maximum size of fine aggregate or grains was <4.75 mm, which was from Chenab, Punjab. The detailed aggregate properties identified according to the ASTM standards are listed in Table 1 [88,89,90,91,92,93,94].

3.2. Bitumen Modification

The present study considers two waste materials for the modification of bitumen: polyethylene terephthalate (PET) and polyethylene (PE). The first additive, PET, was developed through the shredding of waste plastic beverage bottles into small pieces that are often found in garbage or water or even taken from waste dealers. The next additive, PE, was developed through the shredding of flexible plastic pipes locally available at an economical cost. Normally, after the long-term use of pipes, they become waste and people will dump them in garbage. Shredding these additives can transform them into a usable form [72]. The percentage substitution of bitumen with each additive ranged from 5–20% at each testing period. At each proportion of additive, bitumen, and additive were thoroughly mixed at a temperature of 160–170 °C.

3.3. Testing and Preparation of Specimens

For the Marshall stability analysis, the specimen was prepared in the transportation laboratory through the help of an asphalt mixer and the casting of modified and unmodified bitumen specimens was performed simultaneously. For the homogenous mixing of ingredients of the Marshall cake, the aggregates were first dried in an oven at a constant temperature of around 150–170 °C; then, for the removal of air voids, the bitumen was heated at a constant temperature of 160–170 °C. All these heating processes were performed according to the ASTM standards [56,95]. For one Marshall cake, approximately 1200 g of aggregate was taken from the aggregate blend, and the characteristics of the binder and aggregates utilized for the asphalt mix preparation are listed in Table 2. Through the asphalt mixer apparatus, the temperature variation during the mixing of ingredients has been managed and controlled. In this study, 60/70 grade bitumen was used and the mixing of asphalt ingredients was performed at an approximate temperature of 165 °C. After homogenization and the preparation of asphalt mix, the mix was laid into preheated metal molds, and mechanical compaction was performed at 100–140 °C, around 75 blows were applied on both sides of the specimen by using a standard compactor hammer for impact loading. Later, when the specimen cooled, it could be extracted from the mold with the help of a hydraulic extractor available in the transportation laboratory [95]. For testing purposes, the sample was submerged into a thermostatically controlled water bath at a constant temperature of 60 °C, for 30–40 min preceding the testing. In this study, at each percentage substitution of bitumen with additives, the specimens were tested over two periods, the minimum and maximum testing periods were 1 day and 30 days, respectively. The Marshall testing of all specimens at each period was performed according to the ASTM standard D-1559 for precise determination of stability and flow value [72,96]. The aggregate gradation curve is presented in Figure 4. Further, three samples were prepared for each blend of asphalt mix in each testing period, and the average of these was taken as the final value. Various statistical approaches were additionally employed to explore the performance of the produced asphalt mix [97,98,99,100,101,102].

3.4. Marshall Characteristics

The standard testing procedure was adopted for measuring the Marshall characteristics such as Marshall stability (MS) as the major variable, Marshall flow (MF) to help determine the testing performance, and the Marshall quotient (MQ) of both modified asphalt mixtures using the Marshall stability machine. The MQ is the ratio of stability to flow and is measured in kN/mm as an indicator of resistance to the cracking and deformation of asphalt mixtures.

3.5. Response Surface Methodology

Response surface methodology (RSM) is a statistical approach for boosting random research design, and it is used to explore the quantitative link among factors (independent) and response variables. Box–Behnken design (BBD) was utilized on the experimental asphalt mixtures dataset to find the optimum percentage of both polymer content and asphalt mix design. By introducing center points into the plan or system, the BBD creates a hybrid of two randomized block design strategies. The center runs, also known as NC, is a trial base in which the central point is (0, 0, …, 0), and has no less than three NCs with different sums of the k-factor. If three factors are used, the BBD design equals 12 + a center run as shown in matrix form below in Equation (1). A three-variable layout for the BBD is shown in Figure 5.
D = [ 1 1 0 1 1 0 1 1 0 1 1 0 1 0 1 1 0 1 1 0 1 1 0 1 0 1 1 0 1 1 0 1 1 0 1 1 0 0 0 ]
The experimental design in this study used the Box–Behnken approach to optimize the experimental dataset using RSM. The RSM was applied on 24 experimental runs. Each factor and the quadratic model vary considerably on three levels. The regression analysis and response surface graph were plotted using Design Expert Software (13 versions, Stat-Ease Inc., Minneapolis, MI, USA). In the second step, we identified the major effect and relationship among each variable on the response using the second-order quadratic polynomial equation as presented below (Equation (2)).
Y = β 0 + i = 1 n β i X i + i < j n β i j X i X j + j = 1 n β j j X j 2
Here, “Y” is the response variable; Xi and Xj are the independent variables; the mean response coefficient is represented by β 0 ; the number of factors is called “n”; and β i j is the linear relationship between independent variables. In this study, three factors were studied as independent variables: X1, Polymer Type; X2, Polymer content %; and X3, Day on which the Marshall stability was performed. The Marshall characteristics (such as Marshall stability, Marshall flow, and Marshall stiffness) are the response variables.

4. Results and Discussion

4.1. Performance Analysis of Polymer-Modified Bitumen

The primary step is to investigate the characteristics of modified bitumen according to ASTM standards. All the basic modified bitumen properties are examined through the help of laboratory equipment; the outcomes are presented in Figure 6, Figure 7, Figure 8 and Figure 9.
The key aim of this in-depth examination is to understand the behavior of modified bitumen at each percentage substitution. Because of the elastic–plastic nature of bitumen, basic testing is considered to be essential in this study to examine the alteration of binder (bitumen) at each proportion in both waste materials. The results demonstrate that as the percentage substitution of both modifiers increases, the softening point increases simultaneously. In contrast, the flash and fire point values lie within the standard ASTM range. At 10% substitution for either materials, the ductility values satisfy the standard limits; afterwards, the outcomes are below the specified limits. Hence, the analysis concludes that the values of waste material-modified bitumen lie close to the standard range specified by ASTM. It should be remembered that a higher melting point of bitumen is occasionally better in hot climate situations because it resists bitumen bleeding.
Figure 6 shows the softening point readings for five different matrix forms of modified bitumen as measured in °C. Overall, the readings of softening points increase with increased percentages. The softening point values at 0% are the lowest, around about 55, and are within the standard range, which ranges between 45–55 for 60/70 grade bitumen. Moreover, a significant rise in softening point values can be seen for the addition of both plastics. This value then rose sharply to reach a minimum of just about 23% at 20% replacement of PE waste content as compared to neat bitumen. A similar trend was observed in the case of PET-modified bitumen.
The above graph illustrates the flashpoint readings for five various matrix forms of modified bitumen as calculated in unit °C (Figure 7). Overall, the flashpoint values decreased as the percentage increased. The standard value for flashpoint is 232 °C. The flashpoint values at 0% were maximum for the control samples. Moreover, a dramatic fall in flashpoint values can be seen with addition of both plastics. This value then fell steadily to reach a minimum of just about 16% for 20% replacement of PE waste content as compared to neat bitumen. Similarly, a similar line pattern was seen in the case of PET-modified bitumen with values quite close to PE at a similar percentage level.
Figure 8 illustrates the ductility results for five various matrix types of modified bitumen as determined in units of cm @ 25 °C. Overall, the results of ductility fell for both plastic wastes over the percentages given. The standard value for ductility at 25 °C in cm is >75 cm. The ductility values at 0% were maximum for the pure binder with a penetration grade of 60–70. Additionally, a dramatic drop in ductility results can be seen for the case of both waste plastic modifiers. This value then fell steadily to reach a minimum of 59% for 20% replacement of PET waste content as compared to neat bitumen. An analogous pattern was found for PE-modified bitumen.
Figure 9 illustrates the penetration results for five various matrix types of modified bitumen as determined in units of °C. Overall, the results of penetration decreased for both plastic wastes as the percentages increased. The standard value for softening point at 25 °C is between 60–70. The penetration values at 0% were maximum for the pure binder of grades 60–70. Moreover, a dramatic fall in penetration results can be seen in the case of both plastic wastes addition. This value then fell steadily to reach a minimum of just about 42% at 20% replacement of PET waste contents as compared to the pure bitumen binder. Likewise, a comparable line pattern was observed for PE-modified bitumen, but the values were lower than PET before falling to 15%. This then increased to a peak of 1 degree higher than for PET.

4.2. Analysis and Optimization of Polymer Modified Bitumen Using RSM

The RSM was employed to explore the effects of polymer-modified asphalt mixtures under local climatic conditions on the Marshall stability indices, such as Marshall stability (MS), Marshall flow (MF), and Marshall quotient (MQ). By using Box–Behnken design (BBD), 24 experimental runs were designed and the overall series (runs) of experimental sets was considerably reduced. The experimental design and the Marshall characteristics of polymer-modified asphalt mixtures obtained from the experimental testing are described in Table 3. Three factors—PT: Polymer types i.e., PET (0) and PE (1), PC: Polymer contents (0–20%), and Testing days (1 and 30)—were used as an independent factors, and MS (kN), MF (mm), and MQ (kN/mm) were the response variables used in this study.

4.3. Statistical Analysis and ANOVA Results

The best-fitted regression model was recommended and adopted for additional analysis based on the above study design and experimental results using BBD. It was selected based on the R-square value, p-value, and F-value, between multiple response variables including linear, quadratic, 2FI, and cubic polynomial equations. In this study, the suggested model is 2FI which is used to perform the ANOVA analysis for each independent factor and their effect or interaction on the mechanical characteristics of polymer-modified asphalt mixtures measured in terms of response variables. The level statistical of significance was set as 0.05, meaning that models and factors are significant if the p-value is less than 0.05. In this research, p-values less than 0.05 were considered to be significant. The quadratic model appears to be the best, according to the RSM design, and these variables are significant. The ANOVA results are presented in Table 4.

4.4. Marshall Stability Analysis of Modified Bitumen

The ANOVA for the response variable MS (kN) is described in Table 4. The findings show that the 2FI model of MS has acceptable fitting values, such as a p-value < 0.0001. The model F-value of 605.04 implies the model is critical. A p-value less than 0.0500 indicates model terms are critical. In this case, A, B, C, AB, AC, and BC are significant variable model terms. Values above 0.1000 demonstrate that the model terms are not significant. The predicted R2 of 0.9895 agrees with the adjusted R2 of 0.9942; for example, one factor that matters is an adequate precision of 0.2, indicating the signal to noise proportion. A proportion more than 4 is attractive. Our ratio of 63.197 indicates an adequate signal. The regression coefficients of variables can be calculated using the least-squares approach. The appropriate Equation (3) concerning actual variables for MS (kN) is then determined by leaving out the irrelevant variables as follows:
Y 1 : + 4.72 + 0.6476 A + 0.0417 B + 0.5811 C + 0.6476 AB + 0.0706 AC 0.149 BC
Figure 10a shows the diagnostics of the predictive model, which shows an almost linear distribution of the data, suggesting a strong significance. From the ANOVA, the variables of polymer type (A), polymer content (B), and day (C) were significant in the model results i.e., p-value < 0.05, and the term polymer type had the most significant impact on Marshall stability, with a p-value of <0.0001 (which is also less than 0.05). Using the RSM approach, the response surface plots (2D and 3D) of MS are presented in Figure 10b and Figure 11 to explore relationships between the multiple factors and response variables (MS). The correlation among preparation variables and MS, and the interaction of setup factors, are displayed using the 2FI fitting model. Figure 10a shows that the stability of modified asphalt mixtures is increased with an increase in polymer content from 0 to 20% in the case of both polymer materials. The results for PE-modified materials were better as compared to PE wastes. The interactions of factors A (polymer type) and B (testing day) on the Marshall stability as shown in the 3D graph (Figure 10b) reveal an increase in values with the change in polymer type and testing days (i.e., after 30 days). Figure 10 shows the 3D plot between the interactions variables B (polymer content) and C (day), confirming improved MS with the change in both polymer contents after 1 and 30 days of testing.

4.5. Marshall Flow Analysis of Modified Bitumen

The ANOVA for the response variable MF (mm) is detailed in the second row of Table 4. The model F-value of 33,359.91 implies the model is significant. A p-value less than 0.0500 demonstrates model terms are significant. In this case, A, B, C, AB, AC, and BC are significant variable model terms. Values above 0.1000 indicate that the model terms are not significant. The predicted R2 of 0.9998 agrees with the adjusted R2 of 0.9999; for example, one factor that matters is an adequate precision of 0.2, indicating the signal to noise proportion. A proportion more than 4 is attractive (desirable). Our ratio of 412.739 indicates an adequate signal. The significant considerations in the model equation of MF, namely X1, X2, X3, X1X3, X1X2, and X2X3, were determined using the p-values. The appropriate Equation (4) concerning the actual variables for MS (kN) is then determined by leaving out the irrelevant variables as follows:
Y 2 : + 2.08 0.011 0.223 B + 0.3916 0.0109 AB 0.0039 AC 0.0314 BC
The condition with coded variables can be utilized to make forecasts about the reaction for given levels of each component. The coded condition is valuable for recognizing the general effect of the elements by contrasting the variable coefficients. Figure 12a presents the diagnostics of the predictive model, which shows a close dispersion of datasets, suggesting a good significance relation. Based on ANOVA results, the linear 2FI variables of polymer type (A), polymer content (B), and testing day (C) showed significant model results, i.e., p-value > 0.05, and these variables had strong significant effects on Marshall flow values with p-values of <0.05. Using the RSM approach, the response surface plots (2D and 3D) of MF responses are presented in Figure 12b and Figure 13 to explore relationships between the multiple variables and Y-response (MF). The correlation among design factors and MF (mm) and the interaction of setup factors are displayed using the fitting 2FI model. Figure 13a shows that flow values of modified asphalt mixtures are increased with an increase in polymer content from 0 to 20% in the case of both polymer materials. The results for PE materials showed stiffer properties compared to PE wastes. Therefore, PE-modified binders and asphalt mixtures showed better resistance to deformation.
After the detailed experimental work, the average outcomes of Marshall stability and flow values for both additive specimens after 1-day and 30-day periods are presented in Figure 11, Figure 12 and Figure 13. The quadratic model of MF has nonsignificant fitting values; the p-value is 0.4435, which is greater than 0.05. In brief, Figure 11, Figure 12 and Figure 13 show the variations in stability and flow values for each testing period scenario of each additive. In the 1-day testing scenario, for replacement by PET, the stability and flow values are increased at 5% replacement and then decreased, but acceptable according to ASTM standard. Comparatively, in the case of PE, the stability values increase; in fact, at 20% proportion there is an approximate 93% increase in stability, compared to the unmodified specimen, and then the flow value decreased. Conversely, in the 30-day testing scenario for PE, the stability values increase by around 47–93% in contrast to 1-day testing consequences, but all are satisfactory according to the ASTM standard, although the flow values are outside of the range. The same trend can be seen for the 30-day testing scenario for PET.

4.6. Marshall Quotient Analysis of Modified Bitumen

The Marshall quotient (MQ) is defined as the ratio of Marshall stability (kN) to flow (mm) and is used to determine rutting resistance. The ANOVA results of the Marshall quotient are presented in the third row of Table 4. The statistical results show that the quadratic model of MQ has acceptable fitting values, such as a p-value less than 0.05. The model performance based on the F-value of 849,978,602,645.14 suggests the model is significant. There is just a 0.01% chance that a F-value this enormous could happen because of clamor (noise). A p-value under 0.0500 indicates that model terms are significant. For this situation, A, B, C, AB, AC, BC, and ABC are significant variable model terms. Values more than 0.1000 indicate that the model terms are not significant. The significant considerations in the model equation of MQ, namely X1, X2, X3, X1X3, X1X2, and X2X3, were established using the p-values. The appropriate Equation (5) with actual variables for MQ (kN/mm) is then determined by exclusion of the irrelevant variables as follows:
Y 3 : = + 5.93 + 2.01 A + 1.81 B 1.05 C + 2.01 AB 0.506 AC 1.04 BC 0.507 ABC
The relationships for coded elements can be utilized to make forecasts about the reaction for given levels of each component. Of course, the elevated degrees of the variables are coded as +1 and the low levels are coded as −1. The coded condition is valuable for distinguishing the general effect of the elements by looking at the variable coefficients. Figure 14a shows the diagnostics of the predictive model, which shows an almost linear distribution of the data, suggesting a strong significance. As per the ANOVA model, the variables of polymer type (A), polymer content (B), and day (C) showed significant model results, i.e., p-value < 0.05, and the linear term of all variables also has a significant impact on MQ with a p-value of <0.05. Using the RSM approach, the response surface plots (2D and 3D) of the MQ variable are presented in Figure 14b and Figure 15 to explore the relationships between the multiple variables and MQ. The correlation among design factors and MQ (kN/mm) and the interaction of setup factors are displayed using the fitting quadratic regression model.
The effects of factors A (polymer type), B (polymer content), and C (testing days) on the Marshall quotient (MQ) is shown on the 3D graph (Figure 14). The experimental results showed that MQ values increased with an increase in waste plastic contents up to 20% of PE addition, as shown in the contour plot of Figure 14a. The results show that the waste polymer improves the Marshall stability of the asphalt mixtures, but also has a considerable detrimental impact on flow values, consistent with earlier findings. Based on experimental results, the increased MS and MQ of the polymer-modified asphalt could be associated with greater PET and PE dispersion in the asphalt through shear blending, which could provide higher stiffness and contribute to improved stability. The increased MQ values showed more stiffness characteristics and more resistance against deformation. The Marshall characteristics are thought to be primary measures of road performance, as better road performance is attributed to greater stability and reduced flow. We can conclude that the asphalt compositions treated with plastic wastes will have superior performance attributes based on the above findings.

4.7. Optimization of Polymer-Modified Asphalt Mixtures of PET and PE

Regarding the objective of the study, Figure 16 demonstrates the outcomes of optimizing polymer content (PET and PE), as much as is feasible, in Marshall properties. As previously stated, the prepared factors have diverse effects on the Y-variables for both polymer-modified asphalt mixtures and the respective influence significant levels are also diverse. A multi-response optimization approach was performed based on the best-suited 3D-RSM to establish the optimum combination of preparatory factors. The predicted and experimental values for multiple factors concerning corresponding responses are presented in Table 5. The BBD-based RSM can then be used to determine the best blend of prepared setups. The following was identified as the best blend of preparatory factors based on the laboratory settings.

4.8. Environmental Impact Analysis of Polymer-Modified Bitumen

The heat transfer of road pavements gives constant dynamical changes with periodic environmental conditions because of exposure to the actual environment. The heat is conducted via the base course, road foundation, and sub-grade during this change, which ultimately results in a dynamic and complex temperature profile. Hence, the temperature aspect of the road surface design should be considered as a temperature distribution zone under outside weather changes. As far as isotropic granular materials are concerned, the uneven heat flow in 3-D space obeys the conservation law of energy and Equation (6) of Fourier partial differential. In other words, the time-dependent heat phenomena of road pavement are defined by the differential equation given below.
ρ · c T t λ · ( 2 T x 2 + 2 T y 2 + 2 T z 2 ) Q = 0
Here, ρ = Density of asphalt material in unit kg/m3.
c = Specific heat constant of asphalt material in unit J·kg−1·k−1.
λ = Thermal conductivity in-unit W·m−1·K−1.
x, y, z = Three-dimensional coordinates.
Q = 0, which is the heat produced per unit volume.
In simultaneous analysis with the equation of the conduction of heat, the heat transfer of the pavement surface can be determined by the differential of Equation (1). The initial conditions of thermodynamics of the asphalt can generally be divided into two forms depending on the thermodynamic resistivity theory. When the temperature of the road surface is defined to be the boundary, the thermodynamic criterion can be considered to be the first form, such as the Dirichlet condition, and described as:
T ( x , y , z , t ) = T ¯ ( t )
Here, T ¯ is the assigned temperature of the pavement surface at °C.
After this heat flux volume at the demarcation line of the road pavement has been assigned, the thermodynamic flow pattern can be considered in a second form, such as the Neumann condition, and can be represented as:
λ ( T x n x + T y n y + T z n z ) = q ¯ ( t )
Here, n x , n y ,   and   n z are cosines directional, which are the normal outer to the road surfaces respectively and q ¯ stands for a density of heat flux in W/m. As for the road surface subjected to the open environment, q(t) on the surface of the road pavement comprises solar energy ( q x ), convection heat transfer ( q c ) and thermal irradiation ( q r ) as shown in Figure 17 and these can be calculated from the following equation.
q s = { 0 , 0 t < 12 c / 2 q 0 c o s   m ω ( t 12 ) , 12 c / 2 t 12 + 0 , 12 + c / 2 t 24 c / 2
Here, q 0 is the highest solar wave at noon, which is equal to 0.131 (m Q); Q represents the total daily radiation of the sun in units of J/m2; c represents the effective total daily solar radiation in hours; m is the coefficient of solar distribution, which is equal to 12/c; and ω stands for angular frequency (ω = 2Π/24 (rad). The heat transfer happens between both the asphalt road surface and the nearby air during the thermal conductivity. The convection heat transfer ( q c ) can be measured based on the Newton convection/cooling equation, which is given as:
q c = β · A · ( T T a )
Here, β is the coefficient of convection heat movement in unit W.m−2. K−1, which is equal to β = 3.0 υ + 5.6; υ stands for wind speed measured in unit m/s; A stands for contact area in m2; and Ta stands for air temperature measured in °C. While emitting radiation from the environment, the asphalt also emits heat energy from long-wave radiation. Equation (11) can be used to measure the thermal irradiation ( q r ) among the road surface and bordering air according to the Stefan–Boltzmann law.
q r = C s × e × [ ( T T * ) ( T a T * ) ]
Here, e is the coefficient of emissivity of the road surface (e = 0.81); C is the Stefan–Boltzmann constant C s = 5.6697 × 10 8   Wm 2 · K 4 .   T * is used for the absolute temperature at zero degrees and it is equal to −273 °C.
By combining Equations (9)–(11), the temperature of the road surface under the boundary conditions could be determined with the help of Equation (12).
λ ( T x n x + T y n y + T z n z ) = q s + β · A · ( T T a ) + C s × e × [ ( T T * ) ( T a T * ) ]

5. Conclusions

In this study, two waste polymer plastics, PET and PE, are employed as a bitumen modifier to optimize the structural and engineering properties of the asphalt mixtures using RSM based on Box–Behnken design (BBD). The effects of three parameters on the Marshall characteristics were studied and analyzed in this study. The key findings of the present study are concluded as:
  • The research suggests that employing domestic polymer waste, namely PET and PE plastic, as a modifier in 60/70 grade bitumen and pavement mixtures is a promising and novel strategy.
  • The current study investigated three parameters of Marshall characteristics for pure and polymer-modified asphalt binder and mixtures. There is only one parameter, MF, which is not optimized to within the standard range by RSM. Statistical analyses could not be performed for the MF response owing to a lack of data point diversity. This could be due to too large runs of data or range.
  • The optimization of responses is determined by RSM is as follows: MS is 42.98 kN, MF is 5.08 mm, and MQ (kN/mm) is 8.66, indicating favorable and consistent precision when compared with experimental values.
  • The incorporation of both waste polymers (PET and PE) dramatically reduces penetration (0.1 mm) and ductility (cm) measurements while increasing flashpoint (°C) and softening point (°C) readings compared with the corresponding standard range.
  • The polymer-modified mixtures had better stability strength than the control specimen, allowing them to withstand heavy traffic loads and local environmental conditions.
  • Similarly, the increased Marshall quotient (stability/flow) readings showed that asphalt mixtures modified with both polymer materials became stiffer and showed great resilience against permanent deformation after 1 and 30 days of testing when placed in local environmental conditions.
  • The use of polymer waste contents in asphalt mixtures during road construction will help to extend the expected lifespan of our roadways and assist in the diversion of millions of metric tons (MMT) of plastic wastes from landfills; both of these are positive outcomes.

6. Recommendations and Limitations

  • Proactive rules, monitoring, and assistance are needed, particularly in developing countries, to encourage the utilization of polymer waste (i.e., PET and PE) by road companies and organizations in the construction and maintenance of road projects.
  • Municipal waste management organizations should also establish appropriate waste disposal mechanisms to allow for the distinct collection and disposal of plastic, in addition to addressing this major current issue.
  • Future testing should include a variety of mixing circumstances to determine the strength and fatigue properties of PET- or PE-modified asphalt mixes. Moreover, future research could also include a comparison of PET concentration variations over time, as well as aging scenarios paired with UV exposure.
  • In addition, the effect of temperature variation on the road surface due to environmental conditions with new parameters is also recommended and will be studied in the future for a better understanding of the behaviour of asphalt modified with plastic wastes.

Author Contributions

Conceptualization: F.M. and Z.H.; methodology, F.M. and S.A.R.S.; software, F.M. and M.K.A.; validation, S.H. and M.A.; formal analysis, F.M. and M.K.A.; investigation, Y.G., F.M. and M.A.; Supervision, Y.Z. and Z.H.; writing—original draft preparation F.M.; writing—review and editing, M.A. and S.A.R.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research is funded by National Natural Science Foundation of China (No. 51974015 and No. 51474017), the Key Laboratory of Mining Disaster Prevention and Control (Shandong University of Science and Technology) (No. SMDPC202101) and the Fundamental Research Funds for the Central Universities (No. FRF-IC-20-01 and No. FRF-IC-19-013).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data will be available on suitable demand.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tahmoorian, F.; Samali, B.; Yeaman, J. Evaluation of structural and thermal properties of rubber and HDPE for utilization as binder modifier. Modif. Asph. 2018, 109, 109–127. [Google Scholar]
  2. Nizamuddin, S.; Boom, Y.J.; Giustozzi, F. Sustainable polymers from recycled waste plastics and their virgin counterparts as bitumen modifiers: A comprehensive review. Polymers 2021, 13, 3242. [Google Scholar] [CrossRef] [PubMed]
  3. Huang, J.; Wang, Q.-A. Influence of crumb rubber particle sizes on rutting, low temperature cracking, fracture, and bond strength properties of asphalt binder. Mater. Struct. 2021, 54, 54. [Google Scholar] [CrossRef]
  4. Xia, C.; Chen, M.; Geng, J.; Liao, X.; Chen, Z. Swelling and degradation characteristics of crumb rubber modified asphalt during processing. Math. Probl. Eng. 2021, 2021, 6682905. [Google Scholar] [CrossRef]
  5. Yin, J.; Wu, W. Utilization of waste nylon wire in stone matrix asphalt mixtures. Waste Manag. 2018, 78, 948–954. [Google Scholar] [CrossRef] [PubMed]
  6. Yin, J. The feasibility of waste nylon filament used as reinforcement in asphalt mixture. Int. J. Pavement Res. Technol. 2020, 13, 212–221. [Google Scholar] [CrossRef]
  7. Movilla-Quesada, D.; Lagos-Varas, M.; Raposeiras, A.C.; Muñoz-Cáceres, O.; Andrés-Valeri, V.C.; Aguilar-Vidal, C. Analysis of Greenhouse Gas Emissions and the Environmental Impact of the Production of Asphalt Mixes Modified with Recycled Materials. Sustainability 2021, 13, 8081. [Google Scholar] [CrossRef]
  8. Liang, M.; Xin, X.; Fan, W.; Zhang, J.; Jiang, H.; Yao, Z. Comparison of rheological properties and compatibility of asphalt modified with various polyethylene. Int. J. Pavement Eng. 2021, 22, 11–20. [Google Scholar] [CrossRef]
  9. Du, Z.; Jiang, C.; Yuan, J.; Xiao, F.; Wang, J. Low temperature performance characteristics of polyethylene modified asphalts—A review. Constr. Build. Mater. 2020, 264, 120704. [Google Scholar] [CrossRef]
  10. Habib, N.Z.; Kamaruddin, I.; Napiah, M.; Isa, M.T. Rheological properties of polyethylene and polypropylene modified bitumen. Int. J. Civ. Environ. Eng. 2011, 3, 96–100. [Google Scholar]
  11. Abuaddous, M.; Taamneh, M.M.; Rabab’ah, S.R. The potential use of recycled polyethylene terephthalate (RPET) plastic waste in asphalt binder. Int. J. Pavement Res. Technol. 2021, 14, 579–587. [Google Scholar] [CrossRef]
  12. Anwar, M.K.; Shah, S.A.R.; Alhazmi, H. Recycling and utilization of polymers for road construction projects: An application of the circular economy concept. Polymers 2021, 13, 1330. [Google Scholar] [CrossRef] [PubMed]
  13. Mahida, S.; Shah, Y.U.; Sharma, S. Analysis of the influence of using waste polystyrene in virgin bitumen. Int. J. Pavement Res. Technol. 2022, 15, 626–639. [Google Scholar] [CrossRef]
  14. Veropalumbo, R.; Oreto, C.; Viscione, N.; Biancardo, S.A.; Russo, F. Environmental assessment of asphalt mastics containing plastic bottles and jet grouting waste. Environ. Impact Assess. Rev. 2022, 93, 106736. [Google Scholar] [CrossRef]
  15. Mosa, A.; Jawad, I.; Salem, L. Modification of the properties of warm mix asphalt using recycled plastic bottles. Int. J. Eng. 2018, 31, 1514–1520. [Google Scholar]
  16. Mendonça, A.M.G.D.; Melo Neto, O.d.M.; Rodrigues, J.K.G.; Silvani, C.; Batista de Lima, R.K. Physicochemical and rheological effects of the incorporation of micronized polyethylene terephthalate in asphalt binder. Pet. Sci. Technol. 2022, 40, 822–838. [Google Scholar] [CrossRef]
  17. Boger, Z.; Guterman, H. Knowledge extraction from artificial neural network models. In Proceedings of the 1997 IEEE International Conference on Systems, Man, and Cybernetics; Computational Cybernetics and Simulation, Orlando, FL, USA, 12–15 October 1997; pp. 3030–3035. [Google Scholar]
  18. Alawi, M.H.; Rajab, M.I. Applications of neural network for optimum asphaltic concrete mixtures. In Proceedings of the 5th WSEAS International Conference on Simulation, Modelling and Optimization, Corfu, Greece, 17 August 2005; pp. 17–19. [Google Scholar]
  19. Zhu, J.; Birgisson, B.; Kringos, N. Polymer modification of bitumen: Advances and challenges. Eur. Polym. J. 2014, 54, 18–38. [Google Scholar] [CrossRef] [Green Version]
  20. Rezakazemi, M.; Vatani, A.; Mohammadi, T. Synergistic interactions between POSS and fumed silica and their effect on the properties of crosslinked PDMS nanocomposite membranes. RSC Adv. 2015, 5, 82460–82470. [Google Scholar] [CrossRef]
  21. Farahani, H.; Palassi, M.; Galooyak, S. Rheology investigation of waste LDPE and crumb rubber modified bitumen. Eng. Solid Mech. 2018, 6, 27–38. [Google Scholar] [CrossRef]
  22. Rahi, D.; Chandak, R.; Vishwakarma, A. Utilization of liquid plastic waste in bitumen for construction of roads. Int. J. Sci. Res. 2014, 3, 020010. [Google Scholar]
  23. Anwar, M.K.; Shah, S.A.R.; Qurashi, M.A.; Saeed, M.H.; Nisar, A.; Khan, A.N.; Waseem, M. Performance Evaluation of Modified Bitumen Using EPS Beads for Green and Sustainable Development of Polymer-Based Asphalt Mixtures. In Proceedings of the Multidisciplinary Digital Publishing Institute Proceedings, Online, 5–25 November 2020; p. 36. [Google Scholar]
  24. Brasileiro, L.; Moreno-Navarro, F.; Tauste-Martínez, R.; Matos, J.; Rubio-Gámez, M.d.C. Reclaimed polymers as asphalt binder modifiers for more sustainable roads: A review. Sustainability 2019, 11, 646. [Google Scholar] [CrossRef] [Green Version]
  25. Omranian, S.R.; Hamzah, M.O.; Valentin, J.; Hasan, M.R.M. Determination of optimal mix from the standpoint of short term aging based on asphalt mixture fracture properties using response surface method. Constr. Build. Mater. 2018, 179, 35–48. [Google Scholar] [CrossRef]
  26. Tan, Y.-Q.; Guo, M.; Zhang, L.; Cao, L.-P. Performance optimization method of composite modified asphalt sealant. J. Highw. Transp. Res. Dev. 2013, 7, 1–7. [Google Scholar] [CrossRef]
  27. Kavussi, A.; Qorbani, M.; Khodaii, A.; Haghshenas, H. Moisture susceptibility of warm mix asphalt: A statistical analysis of the laboratory testing results. Constr. Build. Mater. 2014, 52, 511–517. [Google Scholar] [CrossRef]
  28. Hamzah, M.O.; Golchin, B.; Tye, C.T. Determination of the optimum binder content of warm mix asphalt incorporating Rediset using response surface method. Constr. Build. Mater. 2013, 47, 1328–1336. [Google Scholar] [CrossRef]
  29. Hamzah, M.O.; Omranian, S.R. Effects of extended short-term aging duration on asphalt binder behaviour at high temperatures. Balt. J. Road Bridge Eng. 2016, 11, 302–312. [Google Scholar] [CrossRef]
  30. Hamzah, M.O.; Gungat, L.; Golchin, B. Estimation of optimum binder content of recycled asphalt incorporating a wax warm additive using response surface method. Int. J. Pavement Eng. 2017, 18, 682–692. [Google Scholar] [CrossRef]
  31. Khodaii, A.; Haghshenas, H.; Tehrani, H.K. Effect of grading and lime content on HMA stripping using statistical methodology. Constr. Build. Mater. 2012, 34, 131–135. [Google Scholar] [CrossRef]
  32. Haghshenas, H.; Khodaii, A.; Khedmati, M.; Tapkin, S. A mathematical model for predicting stripping potential of Hot Mix Asphalt. Constr. Build. Mater. 2015, 75, 488–495. [Google Scholar] [CrossRef]
  33. Brockenbrough, R.L.; Boedecker, J.K.J. Highway Engineering Handbook, 2nd ed.; McGraw Hill Professional: New York, NY, USA, 2003. [Google Scholar]
  34. Anwar, M.K.; Shah, S.A.R.; Sadiq, A.N.; Siddiq, M.U.; Ahmad, H.; Nawaz, S.; Javead, A.; Saeed, M.H.; Khan, A.R. Symmetric Performance Analysis for Mechanical Properties of Sustainable Asphalt Materials under Varying Temperature Conditions: An Application of DT and NDT Digital Techniques. Symmetry 2020, 12, 433. [Google Scholar] [CrossRef] [Green Version]
  35. Ali, F.; Khan, M.A.; Qurashi, M.A.; Shah, S.A.R.; Khan, N.M.; Khursheed, Z.; Rahim, H.S.; Arshad, H.; Farhan, M.; Waseem, M. Utilization of Pyrolytic Carbon Black Waste for the development of Sustainable Materials. Processes 2020, 8, 174. [Google Scholar] [CrossRef] [Green Version]
  36. Aodah, H.H.; Kareem, Y.N.; Chandra, S. Performance of bituminous mixes with different aggregate gradations and binders. Int. J. Eng. Technol. 2012, 2, 2049–3444. [Google Scholar]
  37. Qurashi, M.A.; Shah, S.A.R.; Farhan, M.; Taufiq, M.; Khalid, W.; Arshad, H.; Tayyab, M.; Shahzadi, G.; Waseem, M. Sustainable Design and Engineering: A Relationship Analysis between Digital Destructive and Non-Destructive Testing Process for Lightweight Concrete. Processes 2019, 7, 791. [Google Scholar] [CrossRef] [Green Version]
  38. Pérez-Acebo, H.; Mindra, N.; Railean, A.; Rojí, E. Rigid pavement performance models by means of Markov Chains with half-year step time. Int. J. Pavement Eng. 2019, 20, 830–843. [Google Scholar] [CrossRef]
  39. Bahia, H.U.; Hislop, W.P.; Zhai, H.; Rangel, A. Classification of asphalt binders into simple and complex binders. J. Assoc. Asph. Paving Technol. 1998, 67, 1–41. [Google Scholar]
  40. Smagulova, N.; Kairbekov, Z.; Ermek, A.; Yermoldina, E. Production of bitumens from coal sources modified by elementary sulfur. Adv. Mater. Res. 2012, 535, 1815–1818. [Google Scholar] [CrossRef]
  41. Masson, J. Brief review of the chemistry of polyphosphoric acid (PPA) and bitumen. Energy Fuels 2008, 22, 2637–2640. [Google Scholar] [CrossRef] [Green Version]
  42. Senior-Arrieta, V.; Córdoba-Maquilón, J.E. Mechanical characterization of porous asphalt mixes modified with fatty acid amides-FAA. Ing. Investig. 2017, 37, 43–48. [Google Scholar]
  43. Navarro, F.; Partal, P.; Martınez-Boza, F.; Gallegos, C. Thermo-rheological behaviour and storage stability of ground tire rubber-modified bitumens. Fuel 2004, 83, 2041–2049. [Google Scholar] [CrossRef]
  44. Bai, M. Investigation of low-temperature properties of recycling of aged SBS modified asphalt binder. Constr. Build. Mater. 2017, 150, 766–773. [Google Scholar] [CrossRef]
  45. Lu, X.; Isacsson, U. Modification of road bitumens with thermoplastic polymers. Polym. Test. 2000, 20, 77–86. [Google Scholar] [CrossRef]
  46. Sengoz, B.; Isikyakar, G. Analysis of styrene-butadiene-styrene polymer modified bitumen using fluorescent microscopy and conventional test methods. J. Hazard. Mater. 2008, 150, 424–432. [Google Scholar] [CrossRef] [PubMed]
  47. Valtorta, D.; Poulikakos, L.; Partl, M.; Mazza, E. Rheological properties of polymer modified bitumen from long-term field tests. Fuel 2007, 86, 938–948. [Google Scholar] [CrossRef]
  48. Moreno-Navarro, F.; Sol-Sánchez, M.; Jimenez del Barco, A.; Rubio-Gámez, M. Analysis of the influence of binder properties on the mechanical response of bituminous mixtures. Int. J. Pavement Eng. 2017, 18, 73–82. [Google Scholar] [CrossRef]
  49. Wang, K.; Yuan, Y.; Han, S.; Yang, Y. Application of FTIR spectroscopy with solvent-cast film and PLS regression for the quantification of SBS content in modified asphalt. Int. J. Pavement Eng. 2019, 20, 1336–1341. [Google Scholar] [CrossRef]
  50. Yousefi, A.A. Polyethylene dispersions in bitumen: The effects of the polymer structural parameters. J. Appl. Polym. Sci. 2003, 90, 3183–3190. [Google Scholar] [CrossRef]
  51. Su, N.; Chen, J. Engineering properties of asphalt concrete made with recycled glass. Resour. Conserv. Recycl. 2002, 35, 259–274. [Google Scholar] [CrossRef]
  52. Pérez, I.; Toledano, M.; Gallego, J.; Taibo, J. Mechanical properties of hot mix asphalt made with recycled aggregates from reclaimed construction and demolition debris. Mater. Constr. 2007, 57, 17–29. [Google Scholar]
  53. Chen, M.-Z.; Lin, J.-T.; Wu, S.-P.; Liu, C.-H. Utilization of recycled brick powder as alternative filler in asphalt mixture. Constr. Build. Mater. 2011, 25, 1532–1536. [Google Scholar] [CrossRef]
  54. García-Travé, G.; Tauste, R.; Sol-Sánchez, M.; Moreno-Navarro, F.; Rubio-Gámez, M. Mechanical Performance of SMA Mixtures Manufactured with Reclaimed Geomembrane–Modified Binders. J. Mater. Civ. Eng. 2018, 30, 04017284. [Google Scholar] [CrossRef]
  55. Huang, Y.; Bird, R.N.; Heidrich, O. A review of the use of recycled solid waste materials in asphalt pavements. Resour. Conserv. Recycl. 2007, 52, 58–73. [Google Scholar] [CrossRef]
  56. Al-Abdul Wahhab, H.; Dalhat, M.; Habib, M. Storage stability and high-temperature performance of asphalt binder modified with recycled plastic. Road Mater. Pavement Des. 2017, 18, 1117–1134. [Google Scholar] [CrossRef]
  57. Vila-Cortavitarte, M.; Lastra-González, P.; Calzada-Pérez, M.Á.; Indacoechea-Vega, I. Analysis of the influence of using recycled polystyrene as a substitute for bitumen in the behaviour of asphalt concrete mixtures. J. Clean. Prod. 2018, 170, 1279–1287. [Google Scholar] [CrossRef]
  58. Wen, Y.; Wang, Y.; Zhao, K.; Sumalee, A. The use of natural rubber latex as a renewable and sustainable modifier of asphalt binder. Int. J. Pavement Eng. 2017, 18, 547–559. [Google Scholar] [CrossRef]
  59. Junco, C.; Gadea, J.; Rodríguez, A.; Gutiérrez-González, S.; Calderón, V. Durability of lightweight masonry mortars made with white recycled polyurethane foam. Cem. Concr. Compos. 2012, 34, 1174–1179. [Google Scholar] [CrossRef]
  60. Gutiérrez-González, S.; Gadea, J.; Rodríguez, A.; Junco, C.; Calderón, V. Lightweight plaster materials with enhanced thermal properties made with polyurethane foam wastes. Constr. Build. Mater. 2012, 28, 653–658. [Google Scholar] [CrossRef]
  61. Tribout, C.; Husson, B. Use of treated sediments in road building techniques. Eur. J. Environ. Civ. Eng. 2011, 15, 197–213. [Google Scholar] [CrossRef]
  62. Carrera, V.; Cuadri, A.; García-Morales, M.; Partal, P. The development of polyurethane modified bitumen emulsions for cold mix applications. Mater. Struct. 2015, 48, 3407–3414. [Google Scholar] [CrossRef] [Green Version]
  63. Izquierdo, M.; Navarro, F.; Martínez-Boza, F.; Gallegos, C. Bituminous polyurethane foams for building applications: Influence of bitumen hardness. Constr. Build. Mater. 2012, 30, 706–713. [Google Scholar] [CrossRef]
  64. Vaitkus, A.; Paliukaitė, M. Evaluation of time loading influence on asphalt pavement rutting. Procedia Eng. 2013, 57, 1205–1212. [Google Scholar] [CrossRef] [Green Version]
  65. Tigdemir, M.; Kalyoncuoglu, S.F.; Kalyoncuoglu, U.Y. Application of ultrasonic method in asphalt concrete testing for fatigue life estimation. NDT E Int. 2004, 37, 597–602. [Google Scholar] [CrossRef]
  66. Roberts, F.L.; Kandhal, P.S.; Brown, E.R.; Lee, D.-Y.; Kennedy, T.W. Hot Mix Asphalt Materials, Mixture Design and Construction; NAPA Research and Education Foundation: Lanham, MD, USA, 1991. [Google Scholar]
  67. Ahmed, K.; Irfan, M.; Ahmed, S.; Ahmed, A.; Khattak, A. Experimental investigation of strength and stiffness characteristics of hot mix asphalt (HMA). Procedia Eng. 2014, 77, 155–160. [Google Scholar] [CrossRef] [Green Version]
  68. Mikolaj, J.; Remek, L.; Macula, M. Asphalt concrete overlay optimization based on pavement performance models. Adv. Mater. Sci. Eng. 2017, 2017, 6063508. [Google Scholar] [CrossRef] [Green Version]
  69. Movilla-Quesada, D.; Raposeiras, A.; Silva-Klein, L.; Lastra-González, P.; Castro-Fresno, D. Use of plastic scrap in asphalt mixtures added by dry method as a partial substitute for bitumen. Waste Manag. 2019, 87, 751–760. [Google Scholar] [CrossRef] [PubMed]
  70. Jiang, Y.; Deng, C.; Li, Q.; Liu, H. Effect of compaction methods on physical and mechanical properties of asphalt mixture. J. Mater. Civ. Eng. 2019, 31, 04019075. [Google Scholar] [CrossRef]
  71. Chen, W.-F.; Liew, J.R. The Civil Engineering Handbook; CRC Press: Boca Raton, FL, USA, 2002. [Google Scholar]
  72. Shah, S.A.R.; Arshad, H.; Waqar, A.; Saeed, M.H.; Hafeez, S.; Mansoor, J.; Sadiq, A.N.; Malik, M.A. Saving Energy in the Transportation Sector: An Analysis of Modified Bitumen Application Based on Marshall Test. Energies 2018, 11, 3025. [Google Scholar] [CrossRef] [Green Version]
  73. Baldo, N.; Manthos, E.; Miani, M. Stiffness modulus and marshall parameters of hot mix asphalts: Laboratory data modeling by artificial neural networks characterized by cross-validation. Appl. Sci. 2019, 9, 3502. [Google Scholar] [CrossRef] [Green Version]
  74. Baldo, N.; Manthos, E.; Pasetto, M. Analysis of the Mechanical Behaviour of Asphalt Concretes Using Artificial Neural Networks. Adv. Civ. Eng. 2018, 2018, 1650945. [Google Scholar] [CrossRef] [Green Version]
  75. Saeed, M.H.; Shah, S.A.R.; Arshad, H.; Waqar, A.; Imam, M.A.H.; Sadiq, A.N.; Hafeez, S.; Mansoor, J.; Waseem, M. Sustainable Silicon Waste Material Utilization for Road Construction: An Application of Modified Binder for Marshall Stability Analysis. Appl. Sci. 2019, 9, 1803. [Google Scholar] [CrossRef] [Green Version]
  76. United Nations. Sustainable Development Goals. About the Sustainable Development Goals. Available online: https://www.un.org/sustainabledevelopment/sustainable-development-goals/ (accessed on 13 May 2022).
  77. Hussain, S.; Zhou, X.; Shah, S.A.R.; Ahmad, N.; Anwar, M.K.; Basheer, M.A. Utilization of radium-bead material for road safety: An application of the circular economy concept. Polymers 2021, 13, 3708. [Google Scholar] [CrossRef]
  78. Knudsen, M.S.; Kaivo-oja, J. Bridging Industry 4.0 and Circular Economy: A New Research Agenda for Finland. Tulevaisuuden Tutkimuskeskuksen Blogi. Available online: https://ffrc.wordpress.com/2018/09/12/bridging-industry-4-0-and-circular-economy (accessed on 13 May 2022).
  79. Schroeder, P.; Anggraeni, K.; Weber, U. The relevance of circular economy practices to the sustainable development goals. J. Ind. Ecol. 2019, 23, 77–95. [Google Scholar] [CrossRef] [Green Version]
  80. Valverde, J.-M.; Avilés-Palacios, C. Circular Economy as a Catalyst for Progress towards the Sustainable Development Goals: A Positive Relationship between Two Self-Sufficient Variables. Sustainability 2021, 13, 12652. [Google Scholar] [CrossRef]
  81. PWC. The Road to Circularity. Available online: https://www.pwc.de/de/nachhaltigkeit/pwc-circular-economy-study-2019.pdf (accessed on 13 May 2022).
  82. Bitumen, S. The Shell Bitumen Industrial Handbook; Thomas Telford: Lodon, UK, 1995. [Google Scholar]
  83. Kemalov, R.A.; Kemalov, A.F.; Nam, F.K. Comprehensive analysis of electrochemical and structural-dynamic properties of bitumnious insulating materials. Indian J. Sci. Technol. 2016, 9. [Google Scholar] [CrossRef]
  84. ASTM D-D113-07; Standard Test Method for Ductility of Asphalt Materials. American Society for Testing and Materials: West Conshohocken, PA, USA, 2017.
  85. ASTM D5-2006; Standard Test Method for Penetration of Bituminous Materials. ASTM International: West Conshohocken, PA, USA, 2013.
  86. ASTM D98-18; Standard Test Method for Flash and Fire Points by Cleveland Open Cup Tester. American Society for Testing and Materials: West Conshohocken, PA, USA, 2014.
  87. ASTM D36-95; Standard Test Method for Softening Point of Bitumen (Ring-and-Ball Apparatus). American Society for Testing and Materials: West Conshohocken, PA, USA, 2014.
  88. Testing Aggregates. Part 105: Methods for Determination of Particle Shape. Section 105.1: Flakiness Index; British Standards Institution: London, UK, 1989. [Google Scholar]
  89. BS 812-105.2; Testing Aggregates: Methods for Determination of Particle Shape. Elongation Index of Coarse Aggregate; British Standards Institution: London, UK, 1990.
  90. Astm-C-127; Standard Test Method for Relative Density (Specific Gravity) and Absorption of Coarse Aggregate. Word Press: Montclair, NJ, USA, 1993.
  91. BS 812-110; Testing Aggregates—Methods for Determination of Aggregate Crushing Value (ACV). British Standards Institution: London, UK, 1990.
  92. BS 812-112; Testing Aggregates. Method for Determination of Aggregate Impact Value (AIV). British Standards Institution: London, UK, 1990.
  93. C131/C131M; Standard Test Method for Resistance to Degradation of Small-Size Coarse Aggregate by Abrasion and Impact in the Los Angeles Machine. ASTM International: West Conshohocken, PA, USA, 2018.
  94. C29/C29M-07; Standard Test Method for Bulk Density (Unit Weight) and Voids in Aggregate. ASTM International: West Conshohocken, PA, USA, 2007.
  95. Adedeji, A.; Grünfelder, T.; Bates, F.S.; Macosko, C.; Stroup-Gardiner, M.; Newcomb, D. Asphalt modified by SBS triblock copolymer: Structures and properties. Polym. Eng. Sci. 1996, 36, 1707–1723. [Google Scholar] [CrossRef]
  96. ASTM D5581; Standard Test Method for Resistance to Plastic Flow of Bituminous Mixtures using Marshall Apparatus. American Society for Testing and Materials-ASTM International: West Conshohocken, PA, USA, 2014.
  97. Kalyoncuoglu, S.F.; Tigdemir, M. An alternative approach for modelling and simulation of traffic data: Artificial neural networks. Simul. Modell. Pract. Theory 2004, 12, 351–362. [Google Scholar] [CrossRef]
  98. Garson, D.G. Interpreting Neural Network Connection Weights; Expert.ai: Modena, Italy, 1991. [Google Scholar]
  99. Gibson, N.H. A Viscoelastoplastic Continuum Damage Model for the Compressive Behavior of Asphalt Concrete. Doctoral Dissertation, University of Maryland, College Park, MD, USA, 2006. [Google Scholar]
  100. Goh, A.T. Seismic liquefaction potential assessed by neural networks. J. Geotech. Eng. 1994, 120, 1467–1480. [Google Scholar] [CrossRef]
  101. Kalantar, Z.N.; Karim, M.R.; Mahrez, A. A review of using waste and virgin polymer in pavement. Constr. Build. Mater. 2012, 33, 55–62. [Google Scholar] [CrossRef] [Green Version]
  102. Ling, M.; Luo, X.; Hu, S.; Gu, F.; Lytton, R.L. Numerical modeling and artificial neural network for predicting J-integral of top-down cracking in asphalt pavement. Transp. Res. Rec. 2017, 2631, 83–95. [Google Scholar] [CrossRef]
  103. Lapian, F.E.; Ramli, M.I.; Pasra, M.; Arsyad, A. The Performance Modeling of Modified Asbuton and Polyethylene Terephthalate (PET) Mixture Using Response Surface Methodology (RSM). Appl. Sci. 2021, 11, 6144. [Google Scholar] [CrossRef]
Figure 1. Polymer modification process of mixing by both wet and dry method [24].
Figure 1. Polymer modification process of mixing by both wet and dry method [24].
Polymers 14 02493 g001
Figure 2. (a) Three principles with ten corresponding strategies regarding circular economy [81]. (b) SDGs [76].
Figure 2. (a) Three principles with ten corresponding strategies regarding circular economy [81]. (b) SDGs [76].
Polymers 14 02493 g002
Figure 3. Flow chart of research methodology.
Figure 3. Flow chart of research methodology.
Polymers 14 02493 g003
Figure 4. Aggregates gradation curve.
Figure 4. Aggregates gradation curve.
Polymers 14 02493 g004
Figure 5. The layout of Box–Behnken design with three factors [103].
Figure 5. The layout of Box–Behnken design with three factors [103].
Polymers 14 02493 g005
Figure 6. Softening point values with varying plastic content percentages.
Figure 6. Softening point values with varying plastic content percentages.
Polymers 14 02493 g006
Figure 7. Flashpoint values with varying plastic content percentages.
Figure 7. Flashpoint values with varying plastic content percentages.
Polymers 14 02493 g007
Figure 8. Ductility values with varying plastic content percentages.
Figure 8. Ductility values with varying plastic content percentages.
Polymers 14 02493 g008
Figure 9. Penetration values with varying plastic content percentages.
Figure 9. Penetration values with varying plastic content percentages.
Polymers 14 02493 g009
Figure 10. Predicted vs. actual plot (a) for MS (kN); (b) Contour plot (2D) between the MS (kN) and Factors.
Figure 10. Predicted vs. actual plot (a) for MS (kN); (b) Contour plot (2D) between the MS (kN) and Factors.
Polymers 14 02493 g010
Figure 11. 3D response surface plots for the MS (kN) vs. (a) AB, (b) AC, (c) BC factors.
Figure 11. 3D response surface plots for the MS (kN) vs. (a) AB, (b) AC, (c) BC factors.
Polymers 14 02493 g011aPolymers 14 02493 g011b
Figure 12. (a) Predicted vs. actual plot for MF (kN); (b) contour plot (2D) between the MF (mm) and three factors.
Figure 12. (a) Predicted vs. actual plot for MF (kN); (b) contour plot (2D) between the MF (mm) and three factors.
Polymers 14 02493 g012
Figure 13. 3D response surface plots for the MF (mm) vs. (a) AB, (b) AC, (c) BC factors.
Figure 13. 3D response surface plots for the MF (mm) vs. (a) AB, (b) AC, (c) BC factors.
Polymers 14 02493 g013
Figure 14. (a) Predicted vs. actual plot for MQ (kN/mm) (b) contour plot (2D) between the MQ (kN/mm) and three factors.
Figure 14. (a) Predicted vs. actual plot for MQ (kN/mm) (b) contour plot (2D) between the MQ (kN/mm) and three factors.
Polymers 14 02493 g014
Figure 15. 3D-Response surface plots for the MQ (kN/mm) and three factors (a) AB, (b) AC, (c) BC.
Figure 15. 3D-Response surface plots for the MQ (kN/mm) and three factors (a) AB, (b) AC, (c) BC.
Polymers 14 02493 g015
Figure 16. Optimization values of responses vs. multiple variables using RSM.
Figure 16. Optimization values of responses vs. multiple variables using RSM.
Polymers 14 02493 g016
Figure 17. Effect of temperature variations on road surface due to climatic conditions.
Figure 17. Effect of temperature variations on road surface due to climatic conditions.
Polymers 14 02493 g017
Table 1. Physical properties of polymer-modified binders.
Table 1. Physical properties of polymer-modified binders.
TestUnitsResultsLimitsTest StandardRemarks
Bitumen (Binder)
Ductility @ 25 °Ccm95>75ASTM D-113ok
Penetration @ 25 °Cmm6460–70ASTM D-5
Flash Point°C302232 minASTM D-92
Softening Point°C5540–55ASTM D-36
Aggregates
Water Absorption%1.87<2ASTM C-127ok
Specific Gravity-2.272–3
Aggregate Crushing%24<30BS 812-3
Aggregate Impact Value%19<27
Los Angeles Abrasion%31<35ASTM C-131
Bulk Densitykg/m31500-ASTM C-29
Elongation Index%29.44<45BS 812-105.2
Flakiness Index%21<30BS 812-105.1
Table 2. Experimental design for Box–Behnken design (BBD)-based RSM.
Table 2. Experimental design for Box–Behnken design (BBD)-based RSM.
FactorsUnitsExperimental Levels in BBD-Based RSM
Low Level (−1)High Level (+1)
A: Polymer Type-01
B: Polymer Content%020
C: Testing DaysDay130
Table 3. Experimental design of polymer-modified asphalt mixtures using RSM.
Table 3. Experimental design of polymer-modified asphalt mixtures using RSM.
No.Factors (Independent)Response Variables
X1: PTX2: PC (%)X3: DayY1: MS (kN)Y2: MF (mm)Y3: MS (kN/mm)
1103029.217.14.11
200115.563.784.11
3020110.062.14.79
4003029.217.14.11
50203014.095.342.63
61203043.915.078.66
710115.563.784.11
8120130.032.0214.86
91203043.915.078.66
10103029.217.14.11
1100115.563.784.11
12003029.217.14.11
130203014.095.342.63
14120130.032.0214.86
15020110.062.14.79
1610115.563.784.11
170203014.095.342.63
18020110.062.14.79
191203043.915.078.66
20120130.032.0214.86
2110115.563.784.11
22003029.217.14.11
23103029.217.14.11
2400115.563.784.11
Table 4. ANOVA results for responses and factors of modified asphalt mixture.
Table 4. ANOVA results for responses and factors of modified asphalt mixture.
Responses-YFactorsSum of Sq.dFMean Sq.F-Valuep-ValueSignificant
Y1: MS (kN)Model28.9364.82605.04<0.0001Yes
A: Polymer Type10.07110.071263.17<0.0001*
B: Polymer Content0.041710.04175.230.0372*
C: Testing Days8.1018.101016.83<0.0001*
AB10.07110.071263.17<0.0001*
AC0.119510.119515.000.0015**
BC0.532510.532566.82<0.0001*
Residual0.1195150.0080
Y2: MF (mm)Model4.9060.817433,359.91<0.0001Yes
A: Polymer Type0.002810.0028115.96<0.0001*
B: Polymer Content1.1911.1948,744.70<0.0001*
C: Testing Days3.6813.681.502 × 105<0.0001*
AB0.002810.0028115.96<0.0001*
AC0.000410.000415.000.0015**
BC0.023710.0237967.11<0.0001*
Residual0.0004150.0000
Y3: MQ (kN/mm)Model337.86748.278.500 × 1011<0.0001Yes
A: Polymer Type97.18197.181.711 × 1012<0.0001*
B: Polymer Content78.79178.791.388 × 1012<0.0001*
C: Testing Days26.22126.224.618 × 1011<0.0001*
AB97.18197.181.711 × 1012<0.0001*
AC6.1616.161.085 × 1011<0.0001*
BC26.16126.164.608 × 1011<0.0001*
ABC6.1616.161.085 × 1011<0.0001*
Residual7.950 × 10−10145.678 × 10−11
Note: “*” p < 0.0001; “**” p > 0.0001.
Table 5. Optimal prepared factors values in BBD-based RSM (predicted) and experimentally.
Table 5. Optimal prepared factors values in BBD-based RSM (predicted) and experimentally.
Y: ResponsesOptimal Factors in BBD-Based RSMOptimal Y-Values
X1: PTX2: PC (%)X3: Testing DayPredictedExperimental
Y1: MS (kN)0.98203042.9843.84 ± 0.30
Y2: MF (mm)120305.084.61 ± 0.18
Y3: MQ (kN/mm)12029.58.758.68 ± 0.12
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mushtaq, F.; Huang, Z.; Shah, S.A.R.; Zhang, Y.; Gao, Y.; Azab, M.; Hussain, S.; Anwar, M.K. Performance Optimization Approach of Polymer Modified Asphalt Mixtures with PET and PE Wastes: A Safety Study for Utilizing Eco-Friendly Circular Economy-Based SDGs Concepts. Polymers 2022, 14, 2493. https://doi.org/10.3390/polym14122493

AMA Style

Mushtaq F, Huang Z, Shah SAR, Zhang Y, Gao Y, Azab M, Hussain S, Anwar MK. Performance Optimization Approach of Polymer Modified Asphalt Mixtures with PET and PE Wastes: A Safety Study for Utilizing Eco-Friendly Circular Economy-Based SDGs Concepts. Polymers. 2022; 14(12):2493. https://doi.org/10.3390/polym14122493

Chicago/Turabian Style

Mushtaq, Faizan, Zhian Huang, Syyed Adnan Raheel Shah, Yinghua Zhang, Yukun Gao, Marc Azab, Sajid Hussain, and Muhammad Kashif Anwar. 2022. "Performance Optimization Approach of Polymer Modified Asphalt Mixtures with PET and PE Wastes: A Safety Study for Utilizing Eco-Friendly Circular Economy-Based SDGs Concepts" Polymers 14, no. 12: 2493. https://doi.org/10.3390/polym14122493

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