Next Article in Journal
Urban Flood Risk and Economic Viability Analyses of a Smart Sustainable Drainage System
Next Article in Special Issue
Comparative Investigations into Environment-Friendly Production Methods for Railway Prestressed Concrete Sleepers and Bearers
Previous Article in Journal
Understanding Climate Hazard Patterns and Urban Adaptation Measures in China
Previous Article in Special Issue
Machine Learning Application to Eco-Friendly Concrete Design for Decarbonisation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Chemical, Thermal, and Rheological Performance of Asphalt Binder Containing Plastic Waste

by
Rosa Veropalumbo
1,
Francesca Russo
1,
Cristina Oreto
1,
Giovanna Giuliana Buonocore
2,
Letizia Verdolotti
2,*,
Herminio Muiambo
3,
Salvatore Antonio Biancardo
1 and
Nunzio Viscione
1
1
Department of Civil, Construction and Environmental Engineering, Federico II University of Naples, Via Claudio 21, 80125 Naples, Italy
2
Institute of Composites, Polymers and Biomaterials, National Research Council, P.le E. Fermi, 1, 80055 Naples, Italy
3
Chemistry Department, Eduardo Mondlane University, Maputo P.O. Box 257, Mozambique
*
Author to whom correspondence should be addressed.
Sustainability 2021, 13(24), 13887; https://doi.org/10.3390/su132413887
Submission received: 26 October 2021 / Revised: 4 December 2021 / Accepted: 6 December 2021 / Published: 15 December 2021
(This article belongs to the Special Issue Sustainable Organic Materials Used in the Construction Sector)

Abstract

:
In order to meet the environmental needs caused by large plastic waste accumulation, in the road construction sector, an effort is being made to integrate plastic waste with the function of polymer into asphalt mixtures; with the purpose of improving the mechanical performance of the pavement layers. This study focuses on the effect of a recycled mixture of plastic waste on the chemical, thermal, and rheological properties of designed asphalt blends and on the identification of the most suitable composition blend to be proposed for making asphalt mixture through a dry modification method. Thermo-gravimetric analysis, differential scanning calorimetry, and Fourier transform infrared spectroscopy analysis were carried out to investigate the effect of various concentrations and dimensions of plastic waste (PW) on the neat binder (NB). The frequency sweep test and the multiple stress creep and recovery test were performed to analyze the viscoelastic behavior of the asphalt blends made up of PW in comparison with NB and a commercial modified bitumen (MB). It has been observed that the presence of various types of plastic materials having different melting temperatures does not allow a total melting of PW powder at the mixing temperatures. However, the addition of PW in the asphalt blend significantly improved the aging resistance without affecting the oxidation process of the plastic compound present in the asphalt blend. Furthermore, when the asphalt blend mixed with 20% PW by the weight of bitumen is adopted into the asphalt mixture as polymer, it improves the elasticity and strengthens the mixture better than the mixture containing MB.

1. Introduction

The global plastics economy is mainly linear: plastics are produced, used, and generally disposed of with no recovery. In order to preserve the environment while meeting consumption demands, a global effort to shift the linear economy into a circular model must be made [1,2]. A circular economy model suggests the promotion and revalorization of plastics already in circulation [3]. To date one of the main environmental problems is pollution due to plastic waste: it degrades in a very long time and consequently ends up filling beaches, seas, and other natural places, when not correctly recycled or reused. Indeed, the COVID-19 pandemic has highlighted the need for single-use plastics since potential health risks and societal fears concerning virus-contaminated products introduce consumer fears of reuse [3]. Nevertheless, these socio-material challenges need a systematic approach to plastic waste management. On this basis, the sector of asphalt pavement construction has provided a good opportunity to recycle, reuse, and reduce plastic waste. For this reason, many researchers in the road construction sector are investigating the reuse of plastics/polymers into bituminous mixtures [4]. The employment of polymers as bitumen modifiers has been generally studied for decades. Polymer modification of bitumen is already practiced by the asphalt industry with the aim to improve the properties related to road pavement resistance. Indeed, it has been proved that the addition of polymer into binder enhances its stiffness at high temperature, its moisture resistance, and its fatigue life [5].
Some authors have shown that using a 4.5% of SBS polymer with high-vinyl content by weight of the binder increase the number of fatigue cycles to failure when compared to a commercial polymer-modified bitumen and a virgin bitumen at any strain level [6]. Furthermore, it was found that the modification of neat bitumen using a hybrid combination of styrene-ethylene/propylene-styrene (SEPS) and montmorillonite (MMT) clay improves not only its performance at high temperature but also the aging resistance due to a lower increase in carbonyl functional groups when exposed to oxidation [7].
As a consequence of the massive plastic waste that is generated worldwide, the use of waste plastics/polymers has been also investigated. According to Wu and Montalvo [8] and Vargas and Hanandeh [9], the major types of waste polymers used to modify bitumens are polyethylene teraphthalate (PET) [10,11], which not only reduced the amount of binder required for the optimization of an asphalt mixture but also produced a significant decrease in plastic deformations; high-density polyethylene (HDPE) [12], whose optimum content between 6 and 10% was found to improve the fatigue phenomenon where the number of loading cycles increases; polyvinyl chloride (PVC) [13], which increases the strength and stability of the mix; low-density polyethylene (LDPE) [14,15,16], which has been found to play important roles in asphalt’s low-temperature properties and hot storage stability; polypropylene (PP) [17], which produced increases of viscosity, softening point, flash point, and the decreasing of density when added to a commercial neat bitumen, penetration, and solubility, and polystyrene (PS) [18] that can be reduced 1% of bitumen in mixture replacing it producing an extra-rigidity on the final asphalt mixture. Other polymers that are attracting researchers’ attention are ethylene-vinyl acetate (EVA) [19,20,21], polyurethane (PU) [22,23,24], acrylonitrile butadiene styrene (ABS) [25], and polycarbonate [26].
Among the waste plastic polymers, polyethylene (PE) is considered the most effective for bitumen modification [27,28].
For example, Padhan and Sreeram [29] evaluated the effect of adding PE into a neat commercial-grade VG 10 binder with a 2:1 ratio to modify it. Better performances were obtained, in particular, the addition of PE exhibited higher softening point and rotational viscosity values, and lower penetration values than the neat binder.
Kakar et al. [30] showed that the addition of 5% PE by mass to the binder lead to an increase in the resistance to the rutting with respect to commercial polymer-modified binders.
Brasiliero et al. [31] analyzed the feasibility of using two percentages (6.5 and 12.5% by the weight of binder) of flakes made from recycled PE as binder modifiers. It was found that as the amount of PE increases the binder shows similar properties of a modified binder manufactured in the refinery, leading to an increase of the complex modulus and at the same time a reduction of the phase angle at high temperature.
Some studies were focused on the effect of mixing high-density polyethylene (HDPE) (melting point at 180 °C), low-density polyethylene (LDPE) (melting point at 160 °C), and polypropylene (PP) (melting point at 190 °C) on the viscoelastic performance of a traditional binder [32]. It has been shown that the addition of LDPE at loading above 2% by the weight of binder increase the viscoelastic behavior of the bitumen [33] while the addition of PE at loading between 2% and 8% by the weight of the binder [34] improves the fatigue performance of the binder with an increase of the failure values (Nf).
Nizamuddin et al. [5] reported that the addition of recycled LDPE as a modifier for bitumen showed significant improvements in the physical, chemical, rheological and thermal properties of base bitumen. Specifically, a concentration between 3 and 6% by the weight of bitumen is suggested to enhance sustainability and increase performance, thus favoring its use during construction. Other researchers have studied different plastic waste blends to find the optimum compositions for pavements that could withstand heavily loaded vehicles and hot climates. Particularly, several studies have been conducted on the reuse of PET widely used in the production of beverage containers and polyester textiles [35]. Merkel et al. [35] showed that additives made from deconstructed PET wastes, chemically modified through aminolysis to obtain terephthalic amides, improve the performance properties of asphalt at various experimental conditions.
It has already been proven that the mixtures containing waste plastic bottles have lower optimum asphalt content compared to the conventional mixture, which could reduce the amount of asphalt binder used in road construction [36]. When plastic wastes are used as modifiers into asphalt mixtures [37], the performance properties of the asphalt mixture are enhanced, in particular, the mixture turns to be less sensitive to moisture damage.
Given the above scenario, the aim of the present research is to investigate the effect of various concentrations and dimensions of plastic waste (PW), mainly composed of a mixture of PET and PE, on the structural, thermal, and rheological properties of asphalt blends, and, in a preliminary analysis, the feasibility to introduce PW into asphalt mixture through a dry method, to obtain performance similar to those of mixture containing commercial modified bitumen. A series of tests have been carried out as synthetized in the experimental program shown in Figure 1. The first phase of the research deals with the investigation of the main properties of the materials that are adopted for making the asphalt blends, and the asphalt mixtures, i.e., neat bitumen (NB) 50/70 penetration grade, modified bitumen 10/40–70 (MB), plastic waste, and limestone aggregates. After that, the asphalt blends containing PW were prepared, and jointly the basic properties in terms of ring and ball (EN 1427 [38]) and viscosity (EN 13702 [39]) were analyzed. Chemical characterization and a thermal degradation analysis were carried out on the asphalt blends to explore the effect of PW when mixed at a high temperature of 150 °C with NB. Subsequently, the main rheological properties were assessed, including the two binders (NB and MB), by means of a dynamic shear rheometer (DSR) and performing a frequency sweep (FS) test (EN 14770 [40]) and a multiple stress creep and recovery (MSCR) test (EN 16659 [41]). After the identification of the most suitable asphalt blend with PW among those analyzed, the same amount of PW corresponding to this blend was used for the preparation of an asphalt mixture optimized according to the Superpave procedures through the dry method, in compliance with the same volumetric properties of a traditional hot mix asphalt with NB and a hot mix asphalt prepared with MB, used as control mixes. On the three obtained mixtures, the indirect tensile strength (EN 12697-23 [42]) and the moisture damage (EN 12697-12 [43]) tests were performed to investigate the effect of PW within an asphalt mixture.

2. Materials and Methods

2.1. Materials

2.1.1. Binders

The neat bitumen (NB), 50/70 penetration grade, was obtained from an oil refinery in Southern Italy. The blends were prepared using NB while a modified bitumen 10/40–70 (MB) was used as a control binder. Table 1 gives a summary of the results of some tests performed on the two bitumens.

2.1.2. Plastic Waste

The plastic waste (PW) used as a modifier was obtained from the mechanical shredding of plastic bottles and made up of different plastic waste types, PET, HDPE, LDPE, and PE (see Figure 2a). The shredded plastic was washed and then dried by centrifugation and airstream ventilation for the removal of non-plastic residual material. The final size was obtained by an extruder equipped with a series of perforated plates whose diameter range from 4 to 0.125 mm.

2.1.3. Preparation of the Bituminous Blends

Two different samples of PW were adopted to obtain the modified bituminous blends: (a) PW with size/diameter ranging from 4 mm to 0.063 mm, and (b) PW with size/diameter ranging from 4 mm to 0.25 mm.
Therefore, three bituminous blends were prepared as follows:
  • NB mixed with 10% PW by the weight of NB whose size ranges between 4 and 0.25 mm, (B1).
  • NB mixed with 10% PW by the weight of NB whose size ranges between 4 and 0.063 mm, (B2).
  • NB mixed with 20% PW by the weight of NB whose size ranges between 4 and 0.063 mm, (B3).
An asphalt blend with NB plus 20% PW by the weight of NB whose size ranges between 4 and 0.25 mm was not studied, since the obtained compound was not workable.
The blend preparation consisted of the following steps [45]: first of all, the bitumen was preheated in a metal container at 150 °C for 1 h, and then poured into a steel container previously oven heated to 150 °C. A known amount of PW was slowly added to the bitumen and then mixed in the mixer at 4500 rpm speed for approximately 10 min (see Figure 2b), until a homogeneous compound was reached (see Figure 2c).
A preliminary analysis was made in terms of ring and ball (R&B) by comparing NB with the obtained three mixtures (B1, B2, and B3) to assess the effect derived from the addition of PW. As can be seen from the results reported in Table 2, R&B values of B1, B2, and B3 samples have been improved by 23, 20, and 88%, respectively, in comparison with NB.
As expected, the viscosity values (see Table 2) of the three blends B1, B2, and B3 are higher than those of NB, at both 100 °C and 135 °C. In particular, at 100 °C, the lowest increase (115%) was shown by the B2 sample while the highest (381%) was obtained by the B1 sample, while at 135 °C keeping constant the PW amount at 10%, B1 shows a lower viscosity value (68%) compared to B2.
It is worth noting that, both at 100 °C and 135 °C, the 10% PW blends show viscoelastic properties similar to the MB sample, in particular, the B2 sample at 100 °C and the B1 sample at 135 °C.
It is noticeable from the above results that the increase of the PW content with a narrower size range gives rise to blends more sensitive to the temperature and, therefore, more workable during the blending operation at temperatures around 150 °C.

2.1.4. Grading Curve with Limestone Aggregates

An asphalt concrete (AC) grading curve with a maximum size of 20 mm (AC20) for a binder layer was designed in this study by adopting four different sizes of limestone aggregates provided by a local quarry in southern Italy; Table 3 shows the mechanical and volumetric properties (apparent specific gravity (Gsa) and bulk specific gravity (Gsb)) of the limestone aggregates.
The design of the grading curve was defined according to the following requirements:
  • EN 13108-1 [51] (see Table 1 in EN 13108-1);
  • Special tender documents of southern Italy for a binder layer;
  • Superpave requirements, according to the control points, defined for a nominal maximum aggregate size of 19 mm, and a sand restricted zone to be avoided for compatibility problems [52];
  • Maximum-density gradation, according to the Fuller and Thompson equation, to be avoided (n = 0.45) [53].
The obtained design curve is reported in Figure 3.

2.1.5. Mix Design

According to the purpose of the study, to carry out the analysis on the asphalt mixture when PW is added using the dry method, it was necessary to optimize, in the first analysis, a hot mix asphalt with NB (HMANB) and a hot mix asphalt made up of MB (HMAMB), with the same volumetric properties [54]. The Superpave protocol was followed for the mix design phase, adopting a gyratory compactor to make cylindrical specimens (see Figure 4b) in compliance with EN 12697-31 [55].
The mixing procedure adopted to define the optimum binder content (OBC) for both HMANB and HMAMB was as follows:
  • The aggregates were oven heated at 180 °C, and since NB and MB showed different viscosity values (see Section 2.1.3) and therefore required different mixing temperatures, NB, and MB were preheated respectively at 150 and 185 °C.
  • The coarse aggregates and sand were blended with the bitumen for 4 min at the mixing temperature depending on binder type, after that the filler was added and the mixing continued for others 5 min (see Figure 4a).
After the mixing phase, the mixture was put in the oven for 2 h to reach a uniformity compaction temperature of 150 and 175 °C for HMANB and HMAPMB, respectively. To satisfy the Superpave requirement for a Va equal to 4% at Ndes, 100 and 112 were found to be the appropriate Ndes for HMANB and HMAMB respectively. The results of the mix design are shown in Table 4.

2.2. Methods

2.2.1. Chemical Characterization: Fourier Transform Infrared Spectroscopy-FTIR

FTIR spectra were recorded at room temperature by using an infrared spectrometer (model Frontier Dual Ranger, PerkinElmer, Waltham, MA, USA) in attenuated total reflectance (ATR) mode from 650 to 4000 cm−1 [2,56]. Spectra were recorded at 4 cm−1 resolutions and reported the average of 64 scans. It has been proved by several researchers [57] that infrared spectroscopy (FTIR) in ATR mode is an efficient and reliable method to investigate and monitor the chemical arrangement (i.e., oxidative phenomena) and structure of asphalt samples. Hence, it has been used to analyze NB, its composite samples, B1, B2, B3, and MB.
In the region ranging from 2000 to 600 cm−1, the main vibration peaks, which are indicators of asphalt binder oxidation level, absorb namely carbonyl (C=O, around 1700 to 1740 cm−1), sulfoxide functional groups (S=O, around 1030 cm−1), along with butadiene (−CH=CH−, 960 cm−1) and styrene (aromatic groups, around 690–750 cm−1) peaks [2].
In order to better understand the influence of PW filler on the chemistry of bitumen, we performed the FTIR characterization of the “neat” PW and of the PW after thermal treatment at 150 °C, which represents the average temperature used for the preparation of bitumen composites. All the spectra were baseline corrected by using Perkin Elmer Spectrum software.

2.2.2. Thermal Degradation Properties (TGA and DSC)

The thermal degradation behavior was evaluated by thermo-gravimetric analysis (TGA) and differential scanning calorimetry (DSC). TGA was carried out in air and in nitrogen flow (flow rate = 40 mL/min) using a TGAQ500 TA instrument, at a heating rate of 10 °C/min in the temperature range from 30 °C to 1000 °C, with sample mass of approximately 10 mg [58,59].
DSC analyses were conducted on Q1000 TA instruments. Samples were first heated from −80 °C to 200 °C and then subjected to cooling up to −80 °C, and finally reheated to 200 °C. The rates of heating and cooling were 10 °C/min.

2.2.3. Properties Investigation of Asphalt Binders and Mixtures

The rheological investigation of the asphalt binders was carried out by means of the frequency sweep (FS) test (EN 14770 [40]) and multiple stress creep and recovery (MSCR) test (EN 16659 [41]).
The FS test was carried out using frequency values falling within a range from 0.1 to 10 Hz (a total of 20 observations were made with a gap of 0.1 for frequencies from 0.1 to 1 Hz, and a gap of 1 Hz for frequencies from 1 to 10 Hz, passing through 1.59 Hz) at six test temperatures (0, 10, 20, 30, 40, and 50 °C). A Dynamic Shear Rheometer (DSR, Anton Paar Smart PAVE 102 type) was used.
A “25 mm plate-plate geometry” with a 1 mm gap was adopted to carry out the rheological analysis at test temperatures above 30 °C, and an “8 mm plate-plate geometry” with a 2 mm gap as the DSR configuration was used to investigate all the specimens at test temperatures below 30 °C.
The ratio between the maximum shear stress ( τ m a x ) and the maximum shear strain (γmax) determined the complex shear modulus (|G*|) (see Equations (1) and (2)), while the lag between τmax and γmax obtained the phase angle (δ).
| G * | = τ m a x γ m a x
G * = G 2 + G 2 = G * · c o s φ + G * · s e n φ
where G′ is the storage modulus and G″ is the loss modulus, both in kPa.
The MSCR test was carried out using DSR. Non-recoverable creep compliance (Jnr), which is one of the parameters that give a measure of the degree of resistance of a bituminous mastic and/or binder to permanent deformations under repeated loading–unloading cycles was established in accordance with EN 16659 [41]: two stress levels of 0.1 kPa and 3.2 kPa, and two test temperatures (40 and 50 °C) were used.
The mean value of non-recoverable creep compliance at the N-th cycle ( J n r N , see Equation (3)) and the mean value of the total creep compliance at the N-th cycle ( J t o t N , see Equation (4)) over ten cycles of creep and recovery was calculated for each temperature under 0.1 and 3.2 kPa.
J n r N = ε 10 N τ       kPa 1
J t o t N = ε 1 N τ       kPa 1
where ε 10 N is the strain value at the end of the recovery phase (after 10 s) of the N-th cycle, τ is the applied stress, 0.1 kPa and 3.2 kPa and ε 1 N is the strain value at the end of the creep phase (after 1 s) of the N-th cycle.
The indirect tensile strength (ITS) and the indirect tensile stiffness modulus (ITSM) were adopted for the mechanical characterization of the asphalt mixtures.
The ITS was carried out according to EN 12697-23 [42]. The cylindrical specimen to be tested is brought to the specified test temperature of 10 °C, placed in the compression testing machine between the loading strips, and loaded diametrically along the direction of the cylinder axis with a constant speed of displacement until it breaks. The ITS is the maximum tensile stress calculated from the peak load applied at break and the dimensions of the specimen, calculated by using Equation (5).
I T S = 2 P π D H
where ITS is the indirect tensile strength (GPa), P is the peak load (kN), D is the diameter of the specimen (mm), and H is the height of the specimen (mm).
A total of three specimens were tested and the final ITS value was obtained as a mean value of the three determinations.
The stiffness modulus conducted at three different temperatures of 10, 20, and 40 °C was calculated using an indirect tensile load configuration, in compliance with ANNEX C of the EN 12697-26. Since the standard requires that the peak load value shall be adjusted to achieve a target peak transient horizontal deformation of 0.005% of the specimen diameter when the test temperature is 10 °C, the test was conducted under strain control to keep constant the strain level below 50 με for each temperature. The ITSM was calculated using Equation (6).
I T M = F × ν + 0.27 z × h
where F is the peak value of the applied vertical load (N), z is the amplitude of the horizontal deformation obtained during the load cycle (mm), h is the mean thickness of the specimen (mm), and ν is Poisson’s ratio (assumed to be 0.35).

3. Results and Discussion

3.1. FTIR Characterization

In order to better understand the role of PW in the development of bituminous composites and on their chemistry, the FTIR characterization of “neat” PW and of PW after thermal treatment at 150 °C, which represents the processing temperature for the production of bitumen composites, has been performed. From the analysis of the FTIR spectra related to the PW powders before and after the thermal treatment (see Figure 5), the main characteristic absorption peaks correlated to the various plastic materials contained in the PW (PET, HDPE, LDPE, PP) have been identified. For instance, the C-H stretching vibration absorption, in the range 2964–2910 cm−1, which is from a bond uniquely identified in polymeric materials, along with the stretching vibrations of C-C and C-O-C bonds around 1016 cm−1 and 1100 cm−1 and the bending vibrations of CH2 at 1293 cm−3 were identified. Yet, the specific C=O (carbonyl functional group) vibration peaks, mainly related to the abundant presence of polyesters in the investigated PW, such as PET, were identified at around 1716 cm−1. Moreover, a further sharp peak at 3435 cm−1 was ascribed to the O-H and/or N-H stretching vibrations, attributed to the hydrolytic and oxidizing processes occurring on the surface of the PW [2,60]. The comparison of the spectra before (green curve) and after (red curve) the thermal treatment at 150 °C also shows an increase in the intensity of the carbonyl absorption group (C=O) in the thermally treated sample. This is mainly due to the oxidation of polyolefins (LDPE, HDPE, PP) exposed at high temperatures [2,60].
Figure 6 reports the FTIR spectra of neat and composite bitumen samples (MB, NB, B1, B2, and B3), and Table 5 summarizes the main vibrational assignments. The assignments around 2850 cm−1 and 2920 cm−1 were ascribed to the stretching vibrations of the aliphatic C-H bonds and the corresponding bending vibrations were identified at 1460 cm−1. Additionally, the C=C ring vibration bands from aromatic species were recorded at around 1605–1590 cm−1.
In the vibrational range 1500–1800 cm−1, MB showed some oxidized species which come from the aging of SBS moieties contained in the bitumen [57], i.e., the C=O vibration peaks, at 1741 cm−1, are correlated to the free carbonyl, and at 1690 cm−1, are associated to the conjugate carbonyl. This vibration peak was not observed for the NB, thus indicating that the NB is not oxidized. For the samples B1, B2, and B3, the stretching vibrations of free-carbonyl group C=O were recorded at lower wavenumbers, namely around 1716 cm−1. This peak can be correlated to the carbonyl stretching of PW as observed in the spectra reported in Figure 5. This means that the addition of PW in the bitumen blends does not affect the oxidation/aging process of bitumen.
Additionally, for the MB sample and for NB, a new characteristic oxidation peak at around 1030 cm−1 has been observed and identified as the S=O bond in sulfoxide. Instead, in the case of the B1, B2, and B3 blends, in addition to the S=O slight peak (~1033 cm−1), a vibrational peak at 1018 cm−1, attributed to the C-O-C bond of plastic waste, was also observed (Figure 6b).
Furthermore, the vibration peaks associated with the trans-di-substituted –CH=CH− butadiene block (968 cm−1) and to the C-H aromatic mono-substituted (styrene block, at 748 cm−1) have been highlighted in the spectra of MB samples and NB. At the same wavenumber values, a small shoulder has been also recorded for the investigated blends (B1, B2, and B3).
Moreover, a slight band at around 3400 cm−1 (correlated to the -OH-stretching) has been observed for MB samples (mainly for B3), likely related to the oxidation of the polybutadiene portion in the copolymer.
From the discussion above, it can be concluded that the addition of PW does not affect the oxidation process of SBS-based bitumen, but its presence significantly improves the aging resistance if compared to results obtained using other fillers such the crumb rubber, as also observed by other authors [60].

3.2. Thermal Analysis (TGA and DSC) of Neat Plastic Waste, Neat Bitumen, and Their Composites

To better evaluate the effect of PW on the thermal behavior of composite bitumen, thermal degradation properties of neat PW through TGA (in both air and N2 atmosphere) and DSC analysis have been carried out and the results are reported in Figure 7. In particular, TGA in a N2 atmosphere of the PW sample (Figure 7a) showed a degradation step in the range 360–420 °C with an onset temperature at 361 °C, and a corresponding weight loss of about 90%, that is likely related to the thermal degradation of PET. As already evidenced by FTIR analysis, DSC results (see Figure 7b) clearly show that the main component of the mixture is PET, with its typical melting temperature observed at 248 °C. In addition, from TGA under air atmosphere (reported in Figure 7c) the thermogram shows three separate portions of the mass loss curve: the first one correlated to the HDPE/PP mix (around 3 wt% of mass loss) and the second one to the PET (around 84 wt% of weight loss), both degrade quickly into char and volatiles at around 150–250 °C and 377–527 °C, respectively [61]. In the third part, further increase in temperature slowly decomposed (~13%) the char, which is combusted by reacting with oxygen. From this last analysis, it is possible to assume that in the PW mixture there are 96.5 wt% of PET and 3.5 wt% of HDPE/PP mix.
In Figure 8 the thermograms of the MB, NB, and their derivatives are reported. The DTG (see Figure 8b) was used to identify the point at which weight loss is more evident and to determine: (i) the onset degradation temperature (Tonset), (ii) the maximum decomposition temperature (Tmax), and (iii) the corresponding weight loss as well as (iv) the yield of residual char. The main data are summarized in Table 6. For all the investigated systems, an intense degradation event related to the random break/rupture process and subsequent pyrolysis of the components present in the bitumen samples was observed [62]. The DTG curves (see Figure 8b), showed that MB has the lowest thermal stability with respect to all the other samples, whereas the NB sample exhibited higher thermal stability with respect to its composites. This is due to the presence of the polymeric waste (PET-HDPE-LDPE-PP) which, with its Tmax equal to 427 °C, slightly decreases the Tmax value of NB from 456 °C to 442, 434, and 435 °C, for the three investigated blends B1, B2, and B3, respectively.
In Figure 9, DSC thermograms of all the investigated samples are reported. MB and NB samples showed glass transition temperature (Tg) values of −20 °C, 10 °C, and 60 °C which Özdemir et al. [63] associated, respectively, to the saturate, aromatic, and resin components of the bitumen. During the evaluation of possible interactions among the components in the bituminous blend, the obtained DSC results showed that the presence of PW does not bring any significant change in the Tg of the composites, whereas some differences can still be observed in the melting temperatures. Indeed, DSC curves revealed two sharp melting peaks (Tm = 246 °C and 256 °C) which indicates, as expected, the presence of plastic components in the blend and, particularly, the presence of different types of polymers in the PW including PET whose crystalline melting temperature is about ∼260 °C. It is worth noting that the blend of bitumen and PW is incompatible in nature [63]. They observed that, at low PW content (up to 5%), the presence of small melting peaks has been observed, indicating some level of compatibility. However, at higher loading content (i.e., 7%), two distinct and sharp melting peaks have been observed, thus indicating a high level of incompatibility of polymers in the bituminous matrix. Similar behavior was observed in the DSC curves of the samples investigated in the present work, thus confirming the incompatibility of the blends.

3.3. Performance Results of Asphalt Binder and Mixtures

In the first analysis, to evaluate the FS results the master curves have been designed (see Figure 10) [32].
In Figure 10 it can first be observed that the three blends show different behavior as the temperature increases if compared to NB. At low temperatures, the B2 and B3 blends behave in the same way as NB in terms of |G*|, while B1 showed a 47% higher |G*| than NB; at high test temperatures, the behavior of B2 and B3 is changed, resulting in a |G*| value respectively 80 and 333% higher compared to NB.
For test temperatures below 30 °C, B1 returned a |G*| value on average 77% higher than that of NB, while for test temperatures above 30 °C, B3 achieved a |G*| on average 333% higher than NB.
Comparing the three blends (B1, B2, and B3) it is possible to notice that the complex shear modulus of B1 for temperatures below 30 °C is on average 31 and 34% higher than B2 and B3 respectively; at test temperatures above 30 °C B1 is kept above B2 on average by 20%, while compared to B3 is reduced of 37%.
In order to identify if the performance of the three blends matched the modified bitumen (MB), a further comparison between MB and the three blends was carried out. It is noticeable from Figure 10 that MB returned for test temperatures below 40 °C a |G*| value on average 40% lower than the three asphalt blends, instead for temperatures above 40 °C, MB achieved the best performance, in particular |G*| is slightly higher of 5% than B3 that returned the highest |G*| values for high test temperatures in comparison to the remaining blends with PW (B1 and B2).
At high temperatures, the asphalt binders behave like viscous fluids with no capacity for recovering with no viscous component of |G*|, since δ = 0°. Under normal pavement temperature and traffic loading, bituminous binders act with the characteristics of both viscous liquids and elastic solids; when the binders are loaded, part of their deformation is elastic and part is viscous. Even if two binders are viscoelastic and have the same |G*|, the smaller δ value establishes which one is more elastic than the other, and the deformation could be recovered much more from an applied load.
Therefore, an analysis of the phase angle was carried out. First of all, as reported in Figure 10, it is noticeable that the phase angles of the three blends are on average 11% lower than NB; looking in more detail the single blends B1 and B2 returned, in particular, a constant decrease of 8% regardless of the temperatures, while B3 until 30 °C is 14% lower than NB decreasing even more of 19% and 29% at 40 and 50 °C, respectively. In this case, it is evident that the elasticity of the blends is given by the amount of PW that for B3 resulted higher than B1 and B2 and it is not dependent on the PW size since B1 and B3 returned the same δ values.
Comparing the blends with the modified bitumen at 0 °C, the δ values of B1, B2, and MB matched, while B3 returned a 9% lower phase angle values.
Since 30 °C resulted in a change of behavior for all the solutions so far analyzed, a Cole–Cole diagram at 30 °C was designed to evaluate in more detail the elastic component in the function of the viscous one (see Figure 11).
The elastic degree of the blends, as is shown in Figure 11, was 10%, 8%, and 21% higher than NB for B1, B2, and B3, respectively. The modified bitumen instead, as previously shown by phase angle analysis through the mater curve, returned the highest degree of elasticity between the two bitumen and the three blends.
The measure of the binder’s contribution to mixture permanent deformation is calculated through Jnr. The lower value of Jnr confirms the better performance of the asphalt binder sample against permanent deformation. The results reported in Table 7 reveal that by adding the PW the Jnr decreased compared to the neat binder; in fact, taking into account the average at both test temperatures (40 and 50 °C) and both stress levels (0.1 and 3.2 kPa), overall the Jnr values of the three blends (B1, B2, and B3) are on average 82% lower than NB.
Looking into detail at the Jnr value of the three blends at 40 °C, under both 0.1 and 3.2 kPa, B3 was 71% lower than B1 and B2; at the same time, B1 had a 17% lower Jnr value with respect to B2 at 0.1 kPa, while no substantial difference occurs between them at 3.2 kPa.
At the test temperature of 50 °C, under both stress levels, B3 showed lower Jnr values than B1 and B2, 72% and 65%, respectively. Comparing B1 and B2, B1 returned a Jnr value on average 22% lower.
A further comparison was made with the modified bitumen (MB), that compared to NB returned on average 92% lower Jnr values. Among all blends, B3 returned 55% and 2% lower Jnr value compared to MB under 0.1 kPa respectively at 40 and 50 °C.
Figure 12 shows two 3D scatter plots where the x-axis shows Jtot, the y-axis shows G′, and the z-axis shows Jnr at the test temperatures of 40 (see Figure 12a) and 50 °C (see Figure 12b). These graphs have been plotted to represent in a synthetic way the degree of elasticity of the blends and the binders, better conveyed through a link between the three characteristic parameters (Jnr, Jtot, and G′). As can be observed from the figure, all the asphalt blends exhibit behavior completely different from NB and near MB, both at 40 and 50 °C. B1 and B2 appear to be closer to modified bitumen, otherwise, the deviation of B3 compared to MB seems more marked compared to other mastics, due to the higher G′ value for B3. Therefore, among all mixtures, B3 is the most elastic, even more so than MB.
The data in Table 7 provide an exhaustive characterization of the blends since the Jtot reported were evaluated immediately before load removal. It is evident from the results that no substantial difference returned the three blends between Jnr and Jtot at the same temperature and stress level, confirming their ability to recover deformation already after one second of the end of the creep phase after load removal.
As can be seen from the results above described, B3 returned the most suitable solution to be compared to a modified bitumen as those objects of the present research.
From the investigation carried out on the asphalt mastics, B3 resulted as the most suitable blends to be reemployed into asphalt mixtures. Therefore, an internal laboratory protocol was pursued for mixing the asphalt mixture with the 20% PW by the weight of NB (HMAB3) as follows:
  • The aggregates were preheated for 4 h at 180 °C.
  • The PW stored at room temperature was added directly to the aggregates, before the bitumen. The first blending with coarse aggregates, sand, and PW was performed for 2 min to obtain a homogenous dispersion of all PW particles into the aggregates.
  • The next steps involved the addition of the NB, preheated for 1 h at 150 °C, and further mixing, for approximately 5 min until the aggregates were fully coated with bitumen.
  • The filler was then added, and further mixing was performed for 5 min.
  • After the mixing phase, the mixture was put in the oven for 2 h until it reached a uniform compaction temperature.
Using a Ndes equal to 100 the alternative asphalt mixture, HMAB3, did not exhibit the same volumetric properties as HMANB, since they differed in the Va values, in particular, HMAB3 did not satisfy the Superpave requirement for a Va of 4% at Ndes. Hence, the number of gyrations used for HMAB3 was changed and 120 was found to be the appropriate Ndes. The results of the Superpave optimization for HMAB3 are reported in Table 8a.
The first mechanical characterization of the asphalt mixtures was carried out through the ITS test, whose results are shown in Table 8b. HMAB3 showed the highest ITS at 10 °C than the other two mixtures, in particular, it was 47.4 and 30% higher than that of the HMAMB and HMANB, respectively. The asphalt mixture made up of MB revealed the lowest ITS at 10 °C (2.97 MPa). This is related to the MB used for making the asphalt mixture since no differences existed in terms of the grading curves, volumetric properties, or bitumen concentrations among all the optimized mixtures.
Table 8b illustrates the mean ITSM for each of the three asphalt mixtures at 10, 20, and 40 °C. The HMAB3 exhibited a higher ITSM at all test temperatures, in particular +11% at 10 °C, +18% at 20 °C, and +34% at 40 °C compared to HMANB. Comparing HMAB3 with HMAMB, a slight increase of 9% at both 10 and 20 °C was observed, while at 30 °C HMAB3 turned a higher ITSM value of 37%. Therefore, it emerges that, as the temperature increases, the beneficial effect produced by the plastic increases in terms of the final stiffness of the mixture.
In order to better analyze the effect of the introduction of PW in both asphalt binder and asphalt mixtures, all the parameters analyzed either by the mechanical characterization of the mixtures or by the rheological analysis of the asphalt mastics are summarized in Table 9, as a function of temperature. As it can be seen, for all the parameters analyzed it appears that the best performance is attributable to both asphalt mixtures and the binder containing PW (see the red values in Table 9). The highest stiffness shown by both the binder and the mixture containing PW (via ITSM and |G*|) is fully due to the plastic content because through the FTIR analysis it has been demonstrated that the mixing phase does not produce aging phenomena of bitumen that can stiffen it. In addition, this result was also validated by the analysis of the descriptive parameters of the elasticity of the asphalt mixtures and binders through G′, δ, Jnr, and Jtot which returned a higher elasticity in the case of B3 and HMAB3; this is obviously due to the solid presence of PW as it has emerged from the TGA and DSC analysis that at the mixing temperatures of 150 °C, used in this case study, do not produce degrading phenomena of PW and above all that, not all types of PW reach their melting point.

4. Conclusions

In this study, the effect of the different sizes and contents of plastic waste were assessed through the chemical, thermal, and rheological investigation of asphalt blends mixed with PW in comparison with a neat bitumen 50/70 and modified bitumen. The plastic waste was a mixture of different plastic waste types with a prevalence of PET and PE.
The results obtained lead to the following conclusions:
  • FTIR results highlight that the presence of PW does not affect the oxidation process of SBS-based bitumen, but significantly improves the aging resistance if compared to results obtained using other fillers such as crumb rubber.
  • From the rheological and mechanical investigation, it can be seen that as the temperature increases, there is an increase of the beneficial effect produced by the addition of PW in both asphalt blends and mixtures.
  • A total of 20% PW by weight of NB with a size range between 4 and 0.063 mm was found to be the blend that performs better than the modified bitumen since in terms of stiffness and elasticity; particularly, PW is beneficial to the high temperature achieving lower phase angle and Jnr values (averagely 30% than the other asphalt solutions investigated).
  • The adoption of PW as a modifier of the mixture through the dry method, shown by preliminary study on the mechanical characterization of the optimized asphalt mixture by means of Superpave method, was found to produce an increase of the ITS and ITSM at all test temperature in comparison to the asphalt mixtures made up of NB and MB, despite all mixtures having the same volumetric properties.
It was discovered that during the mixing phase of 150 °C, no melting temperatures and degradation process have been achieved by PW that increases the overall strength and elasticity of the innovative asphalt blends and mixtures.

Author Contributions

Conceptualization, R.V. and G.G.B.; methodology, R.V., C.O. and F.R.; validation, H.M., S.A.B. and F.R.; investigation, R.V., N.V. and L.V.; data curation, N.V., C.O., G.G.B. and L.V.; writing—original draft preparation, R.V. and L.V.; writing—review and editing, G.G.B. and F.R.; visualization, G.G.B. and F.R.; supervision, H.M., S.A.B. and F.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was developed within the Projects of National Interest-PRIN 2017 “Stone pavements. History, conservation, valorisation and design” (20174JW7ZL) financed by the Ministry of Education, University and Research (MIUR) of the Italian Government.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

The Italy/Mozambique Cooperation project Preparation And Characterization of Polyolefin Waste Blends/Bitumen Composites For Flexible Road Pavement funded by FIAM (Fundo para a Investigacao Applicada e Multissectiorial) is kindly acknowledged. The authors are grateful to Mario De Angioletti (CNR-IPCB), Fabio Docimo (CNR-IPCB), Mariarosaria Marcedula, and Alessandra Aldi (CNR-IPCB) for technical support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Oreto, C.; Veropalumbo, R.; Viscione, N.; Biancardo, S.; Russo, F. Investigating the environmental impacts and engineering performance of road asphalt pavement mixtures made up of jet grouting waste and reclaimed asphalt pavement. Environ. Res. 2021, 198, 111277. [Google Scholar] [CrossRef] [PubMed]
  2. Verdolotti, L.; Iucolano, F.; Capasso, I.; Lavorgna, M.; Iannace, S.; Liguori, B. Recycling and recovery of PE-PP-PET-based fiber polymeric wastes as aggregate replacement in lightweight mortar: Evaluation of environmental friendly application. Environ. Prog. Sustain. Energy 2014, 33, 1445–1451. [Google Scholar] [CrossRef]
  3. Schyns, Z.O.; Shaver, M.P. Mechanical recycling of packaging plastics: A review. Macromol. Rapid Commun. 2021, 42, 2000415. [Google Scholar] [CrossRef] [PubMed]
  4. Viscione, N.; Veropalumbo, R.; Oreto, C.; Biancardo, S.A.; Abbondati, F.; Russo, F. Additional procedures for characterizing the performance of recycled polymer modified asphalt mixtures. Measurement 2022, 187, 110238. [Google Scholar] [CrossRef]
  5. Nizamuddin, S.; Jamal, M.; Gravina, R.; Giustozzi, F. Recycled plastic as bitumen modifier: The role of recycled linear low density polyethylene in the modification of physical, chemical and rheological properties of bitumen. J. Clean. Prod. 2020, 266, 121988. [Google Scholar] [CrossRef]
  6. Gupta, A.; Lastra-Gonzalez, P.; Rodriguez-Hernandez, J.; González, M.G.; Castro-Fresno, D. Critical assessment of new polymer-modified bitumen for porous asphalt mixtures. Constr. Build. Mater. 2021, 307, 124957. [Google Scholar] [CrossRef]
  7. Hajikarimi, P.; Rahi, M.; Moghadas Nejad, F.; Babaii Ashourabadi, E.; Maniei, S.; MohammadGhasemi, P.; Fini, E.H. Introduction of Polymer Nanocomposites to Bitumen to Enhance its Thermomechanical Properties. J. Transp. Eng. Part B Pavements 2021, 147, 04021020. [Google Scholar] [CrossRef]
  8. Wu, S.; Montalvo, L. Repurposing waste plastics into cleaner asphalt pavement materials: A critical literature review. J. Clean. Prod. 2021, 280, 124355. [Google Scholar] [CrossRef]
  9. Vargas, C.; El Hanandeh, A. Systematic literature review, meta-analysis and artificial neural network modelling of plastic waste addition to bitumen. J. Clean. Prod. 2021, 280, 124369. [Google Scholar] [CrossRef]
  10. Leng, Z.; Sreeram, A.; Padhan, R.K.; Tan, Z. Value-added application of waste PET based additives in bituminous mixtures containing high percentage of reclaimed asphalt pavement (RAP). J. Clean. Prod. 2018, 196, 615–625. [Google Scholar] [CrossRef]
  11. Movilla-Quesada, D.; Raposeiras, A.C.; Silva-Klein, L.T.; 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]
  12. Arabani, M.; Pedram, M. Laboratory investigation of rutting and fatigue in glassphalt containing waste plastic bottles. Constr. Build. Mater. 2016, 116, 378–383. [Google Scholar] [CrossRef]
  13. Behl, A.; Sharma, G.; Kumar, G. A sustainable approach: Utilization of waste PVC in asphalting of roads. Constr. Build. Mater. 2014, 54, 113–117. [Google Scholar] [CrossRef]
  14. Ho, S.; Church, R.; Klassen, K.; Law, B.; MacLeod, D.; Zanzotto, L. Study of recycled polyethylene materials as asphalt modifiers. Can. J. Civ. Eng. 2006, 33, 968–981. [Google Scholar] [CrossRef]
  15. Punith, V.S.; Veeraragavan, A.; Amirkhanian, S.N. Evaluation of reclaimed polyethylene modified asphalt concrete mixtures. Int. J. Pavement Res. Technol. 2011, 4, 1. [Google Scholar]
  16. Bagampadde, U.; Kaddu, D.; Kiggundu, B.M. Evaluation of rheology and moisture susceptibility of asphalt mixtures modified with low density polyethylene. Int. J. Pavement Res. Technol. 2013, 6, 217–224. [Google Scholar]
  17. Sembiring, E.; Rahman, H.; Siswaya, Y.M. Utilization of polypropylene to substitute bitumen for asphalt concrete wearing course. Int. J. GEOMATE 2018, 14, 97–102. [Google Scholar] [CrossRef]
  18. 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]
  19. Garcia-Morales, M.; Partal, P.; Navarro, F.J.; Gallegos, C. Effect of waste polymer addition on the rheology of modified bitumen. Fuel 2006, 85, 936–943. [Google Scholar] [CrossRef]
  20. Bulatović, V.O.; Rek, V.; Marković, K.J. Rheological properties and stability of ethylene vinyl acetate polymer-modified bitumen. Polym. Eng. Sci. 2013, 53, 2276–2283. [Google Scholar] [CrossRef]
  21. Singh, M.; Kumar, P.; Maurya, M.R. Effect of aggregate types on the performance of neat and EVA-modified asphalt mixtures. Int. J. Pavement Eng. 2014, 15, 163–173. [Google Scholar] [CrossRef]
  22. Sun, M.; Zheng, M.; Qu, G.; Yuan, K.; Bi, Y.; Wang, J. Performance of polyurethane modified asphalt and its mixtures. Constr. Build. Mater. 2018, 191, 386–397. [Google Scholar] [CrossRef]
  23. Bazmara, B.; Tahersima, M.; Behravan, A. Influence of thermoplastic polyurethane and synthesized polyurethane additive in performance of asphalt pavements. Constr. Build. Mater. 2018, 166, 1–11. [Google Scholar] [CrossRef]
  24. Xu, C.; Zhang, Z.; Zhao, F.; Liu, F.; Wang, J. Improving the performance of RET modified asphalt with the addition of polyurethane prepolymer (PUP). Constr. Build. Mater. 2019, 206, 560–575. [Google Scholar] [CrossRef]
  25. Colbert, B.W.; You, Z. Properties of modified asphalt binders blended with electronic waste powders. J. Mater. Civ. Eng. 2012, 24, 1261–1267. [Google Scholar] [CrossRef]
  26. Pugsley, A.; Zacharopoulos, A.; Smyth, M.; Mondol, J. Performance evaluation of the senergy polycarbonate and asphalt carbon nanotube solar water heating collectors for building integration. Renew. Energy 2019, 137, 2–9. [Google Scholar] [CrossRef]
  27. Al-Hadidy, A.I. Effect of laboratory aging on moisture susceptibility and resilient modulus of asphalt concrete mixes containing PE and PP polymers. Karbala Int. J. Mod. Sci. 2018, 4, 377–381. [Google Scholar] [CrossRef]
  28. Iqbal, M.; Hussain, A.; Khattak, A.; Ahmad, K. Improving the Aging Resistance of Asphalt by Addition of Polyethylene and Sulphur. Civ. Eng. J. 2020, 6, 1017–1030. [Google Scholar] [CrossRef]
  29. Padhan, R.K.; Sreeram, A. Enhancement of storage stability and rheological properties of polyethylene (PE) modified asphalt using cross linking and reactive polymer based additives. Constr. Build. Mater. 2018, 188, 772–780. [Google Scholar] [CrossRef]
  30. Kakar, M.R.; Mikhailenko, P.; Piao, Z.; Bueno, M.; Poulikakos, L. Analysis of waste polyethylene (PE) and its by-products in asphalt binder. Constr. Build. Mater. 2021, 280, 122492. [Google Scholar] [CrossRef]
  31. Brasileiro, L.L.; Moreno-Navarro, F.; Martínez, R.T.; del Sol-Sánchez, M.; Matos, J.M.E.; del Carmen Rubio-Gámez, M. Study of the feasability of producing modified asphalt bitumens using flakes made from recycled polymers. Constr. Build. Mater. 2019, 208, 269–282. [Google Scholar] [CrossRef]
  32. Veropalumbo, R.; Russo, F.; Viscione, N.; Biancardo, S.A.; Oreto, C. Investigating the rheological properties of hot bituminous mastics made up using plastic waste materials as filler. Constr. Build. Mater. 2021, 270, 121394. [Google Scholar] [CrossRef]
  33. Dalhat, M.A.; Al-Abdul Wahhab, H.I. Performance of recycled plastic waste modified asphalt binder in Saudi Arabia. Int. J. Pavement Eng. 2017, 18, 349–357. [Google Scholar] [CrossRef]
  34. Hu, C.; Lin, W.; Partl, M.; Wang, D.; Yu, H.; Zhang, Z. Waste packaging tape as a novel bitumen modifier for hot-mix asphalt. Constr. Build. Mater. 2018, 193, 23–31. [Google Scholar] [CrossRef]
  35. Merkel, D.R.; Kuang, W.; Malhotra, D.; Petrossian, G.; Zhong, L.; Simmons, K.L.; Zhang, J.; Cosimbescu, L. Waste PET Chemical Processing to Terephthalic Amides and Their Effect on Asphalt Performance. ACS Sustain. Chem. Eng. 2020, 8, 5615–5625. [Google Scholar] [CrossRef]
  36. Moghaddam, T.B.; Karim, M.R.; Soltani, M. Utilization of waste plastic bottles in asphalt mixture. Eng. Sci. Technol. Int. J. 2013, 8, 264–271. [Google Scholar]
  37. Haider, S.; Hafeez, I.; Ullah, R. Sustainable use of waste plastic modifiers to strengthen the adhesion properties of asphalt mixtures. Constr. Build. Mater. 2020, 235, 117496. [Google Scholar] [CrossRef]
  38. European Standard. EN 1427 Bitumen and Bituminous Binders—Determination of the Softening Point—Ring and Ball Method; European Standard: Brussels, Belgium, 2015. [Google Scholar]
  39. European Standard. EN 13702 Bitumen and Bituminous Binders. Determination of Dynamic Viscosity of Bitumen and Bitu-Minous Binders by the Cone and Plate Method; European Standard: Brussels, Belgium, 2018. [Google Scholar]
  40. European Standard. EN 14770. Bitumen and Bituminous Binders—Determination of Complex Shear Modulus and Phase Angle—Dynamic Shear Rheometer (DSR); European Standard: Brussels, Belgium, 2012. [Google Scholar]
  41. European Standard. EN 16659. Bitumen and Bituminous Binders—Multiple Stress Creep and Recovery Test (MSCRT); European Standard: Brussels, Belgium, 2015. [Google Scholar]
  42. European Standard. EN 12697-23. Bituminous Mixtures—Test Methods—Part 23: Determination of the Indirect Tensile Strength of Bituminous Specimens; European Standard: Brussels, Belgium, 2018. [Google Scholar]
  43. European Standard. Bituminous Mixtures—Test Methods—Part 12: Determination of the Water Sensitivity of Bituminous Specimens; European Standard: Brussels, Belgium, 2018. [Google Scholar]
  44. European Standard. EN1426 Bitumen and Bituminous Binders—Determination of Needle Penetration; European Standard: Brussels, Belgium, 2015. [Google Scholar]
  45. Veropalumbo, R.; Russo, F.; Oreto, C.; Biancardo, S.A.; Zhang, W.; Viscione, N. Verifying laboratory measurement of the performance of hot asphalt mastics containing plastic waste. Measurement 2021, 180, 109587. [Google Scholar] [CrossRef]
  46. European Standard. EN 1097-2. Tests for Mechanical and Physical Properties of Aggregates—Part 2: Methods for the Determination of Resistance to Fragmentation; European Standard: Brussels, Belgium, 2018. [Google Scholar]
  47. European Standard. EN 933-4. Test for Geometrical Properties of Aggregates—Part 4: Determination of Particle Shape—Shape Index; European Standard: Brussels, Belgium, 2008. [Google Scholar]
  48. European Standard. EN 933-3. Tests for Geometrical Properties of Aggregates—Part 3: Determination of Particle Shape—Flakiness Index; European Standard: Brussels, Belgium, 2012. [Google Scholar]
  49. European Standard. EN 933-8. Tests for Geometrical Properties of Aggregates—Part 8: Assessment of Fines—Sand Equivalent Test; European Standard: Brussels, Belgium, 2015. [Google Scholar]
  50. European Standard. EN 1097-6. Tests for Mechanical and Physical Properties of Aggregates—Part 6: Determination of Particle Density and Water Absorption; European Standard: Brussels, Belgium, 2013. [Google Scholar]
  51. European Standard. EN 13108-1Bituminous Mixtures—Material Specifications—Part 1: Asphalt Concrete; European Standard: Brussels, Belgium, 2016. [Google Scholar]
  52. Asphalt Institute. Superpave Series No. 2 (SP-02); Asphalt Institute: Lexington, KY, USA, 1996. [Google Scholar]
  53. Fuller, W.B.; Thompson, S.E. The Laws of Proportioning Concrete. Trans. Am. Soc. Civ. Eng. 1907, 59, 67–143. [Google Scholar] [CrossRef]
  54. Russo, F.; Veropalumbo, R.; Biancardo, S.A.; Oreto, C.; Scherillo, F.; Viscione, N. Reusing Jet Grouting Waste as Filler for Road Asphalt Mixtures of Base Layers. Materials 2021, 14, 3200. [Google Scholar] [CrossRef] [PubMed]
  55. European Standard. EN 12697-31. Bituminous Mixtures—Test Methods—Part 31: Specimen Preparation by Gyratory Compactor; European Standard: Brussels, Belgium, 2019. [Google Scholar]
  56. Stanzione, M.; Russo, V.; Oliviero, M.; Verdolotti, L.; Sorrentino, A.; Di Serio, M.; Tesser, R.; Iannace, S.; Lavorgna, M. Synthesis and characterization of sustainable polyurethane foams based on polyhydroxyls with different terminal groups. Polymer 2018, 149, 149134–149145. [Google Scholar] [CrossRef]
  57. Yan, C.; Huang, W.; Ma, J.; Xu, J.; Lv, Q.; Lin, P. Characterizing the SBS polymer degradation within high content polymer modified asphalt using ATR-FTIR. Constr. Build. Mater. 2020, 233, 117708. [Google Scholar] [CrossRef]
  58. Verdolotti, L.; Di Maio, E.; Lavorgna, M.; Iannace, S. Hydration-induced reinforcement of rigid polyurethane-cement foams: Mechanical and functional properties. J. Mater. Sci. 2012, 47, 6948–6957. [Google Scholar] [CrossRef]
  59. Verdolotti, L.; Lavorgna, M.; Di Maio, E.; Iannace, S. Hydration-induced reinforcement of rigid polyurethaneecement foams: The effect of the co-continuous morphology on the thermal-oxidative stability. Polym. Deg. Stability 2013, 98, 64–72. [Google Scholar] [CrossRef]
  60. Liguori, B.; Iucolano, F.; Capasso, I.; Lavorgna, M.; Verdolotti, L. The effect of recycled plastic aggregate on chemico-physical and functional properties of composite mortars. Mater. Des. 2014, 57, 578–584. [Google Scholar] [CrossRef]
  61. Das, P.; Tiwari, P. Thermal degradation study of waste polyethylene terephthalate (PET) under inert and oxidative environments. Thermochim. Acta 2019, 679, 178340. [Google Scholar] [CrossRef]
  62. Naskara, M.; Chakia, T.K.; Reddy, K.S. Effect of waste plastic as modifier on thermal stability and degradation kinetics of bitumen/waste plastics blend. Thermochim. Acta 2010, 509, 128–134. [Google Scholar] [CrossRef]
  63. Özdemir, D.K.; Topal, A.; Gupta, J.; McNally, T. Aging effects on the composition and thermal properties of styrene-butadiene-styrene (SBS) modified bitumen. Constr. Build. Mater. 2020, 235, 117450. [Google Scholar] [CrossRef]
Figure 1. Summary of the experimental program.
Figure 1. Summary of the experimental program.
Sustainability 13 13887 g001
Figure 2. Blend preparation: (a) plastic waste powder with a size between 4 and 0.063 mm; (b) mixing phase; (c) final blend.
Figure 2. Blend preparation: (a) plastic waste powder with a size between 4 and 0.063 mm; (b) mixing phase; (c) final blend.
Sustainability 13 13887 g002
Figure 3. Grading curve for the binder layer.
Figure 3. Grading curve for the binder layer.
Sustainability 13 13887 g003
Figure 4. (a) Asphalt mixing phase and (b) gyratory compaction.
Figure 4. (a) Asphalt mixing phase and (b) gyratory compaction.
Sustainability 13 13887 g004
Figure 5. FTIR spectra of plastic waste, before and after thermal treatment at 150 °C (a) complete investigation range (b) magnification in the range 700–1900.
Figure 5. FTIR spectra of plastic waste, before and after thermal treatment at 150 °C (a) complete investigation range (b) magnification in the range 700–1900.
Sustainability 13 13887 g005
Figure 6. FTIR spectra of MB, NB, B1, B2, and B3 bitumen samples (a) all ranges; (b) range between 1800 and 600 cm−1.
Figure 6. FTIR spectra of MB, NB, B1, B2, and B3 bitumen samples (a) all ranges; (b) range between 1800 and 600 cm−1.
Sustainability 13 13887 g006
Figure 7. (a) TGA and DTGA in N2 atmosphere, (b) DSC analysis, and (c) TGA in air atmosphere of PW sample.
Figure 7. (a) TGA and DTGA in N2 atmosphere, (b) DSC analysis, and (c) TGA in air atmosphere of PW sample.
Sustainability 13 13887 g007
Figure 8. (a) Thermograms with the inset of two main sections of the graph zoomed in and (b) derivative thermograms of bitumen samples and their composites.
Figure 8. (a) Thermograms with the inset of two main sections of the graph zoomed in and (b) derivative thermograms of bitumen samples and their composites.
Sustainability 13 13887 g008
Figure 9. DSC curves of asphalt samples and their composites.
Figure 9. DSC curves of asphalt samples and their composites.
Sustainability 13 13887 g009
Figure 10. Comparing asphalt blends and binders in terms of master curves.
Figure 10. Comparing asphalt blends and binders in terms of master curves.
Sustainability 13 13887 g010
Figure 11. Cole֪–Cole diagram at 30 °C.
Figure 11. Cole֪–Cole diagram at 30 °C.
Sustainability 13 13887 g011
Figure 12. Scatter plot by comparing Jtot, G′, and Jnr: (a) 40 °C and (b) 50 °C.
Figure 12. Scatter plot by comparing Jtot, G′, and Jnr: (a) 40 °C and (b) 50 °C.
Sustainability 13 13887 g012
Table 1. Bitumen basic properties.
Table 1. Bitumen basic properties.
PropertiesUnitStandardBitumen
NBMB
Penetration at 25 °CdmmEN 1426 [44]6852
Softening point (R&B)°CEN 1427 [38]4687
Dynamic viscosity at 135 °CPa sEN 13702 [39]0.251.38
Dynamic viscosity at 100 °C4.109.18
Dynamic viscosity at 60 °C1851560
Table 2. Basic properties of studied materials: ring and ball and viscosity at 100 °C and 135 °C.
Table 2. Basic properties of studied materials: ring and ball and viscosity at 100 °C and 135 °C.
SampleR&BViscosity at 100 °CViscosity at 135 °C
(°C)(Pa × s)(Pa × s)
MB87.09.181.38
NB46.04.100.25
B156.819.731.56
B255.28.834.93
B386.315.622.45
Table 3. Aggregate main properties.
Table 3. Aggregate main properties.
Los Angeles (%)
EN 1097-2 [46]
Shape Index
(%)
EN 933-4 [47]
Flat Index
(%)
EN 933-3 [48]
Sand Equivalent
(%)
EN 933-8 [49]
Gsa
(g/cm3)
EN 1097-6 [50]
Gsb
(g/cm3)
EN 1097-6 [51]
Coarse aggregate
Limestone 10/1820.648n.d. *2.6942.686
Limestone 6/1220.1811n.d.2.7132.687
Fine aggregate
Limestone sandn.d.n.d.n.d.95.32.7182.679
Limestone fillern.d.n.d.n.d.n.d.2.7372.737
* n.d. = not determined.
Table 4. Superpave optimization: HMANB vs. HMAPMB volumetric properties.
Table 4. Superpave optimization: HMANB vs. HMAPMB volumetric properties.
MixtureVa
(%)
VMA
(%)
VFA
(%)
%Gmm at Ndes
(%)
Dust Portion
(-)
HMANB4.014.270.685.57%1.1
HMAMB4.114.471.895.90%1.1
Table 5. Assignments of the main IR bands of bitumen samples and their composites.
Table 5. Assignments of the main IR bands of bitumen samples and their composites.
AssignmentGroups Assignment Samples
MBNBB1B2B3
υst C-HAliphatic species29002900290029002900
υst C=OFree carbonyl1741n.d.171617161716
υst C=OConjugate carbonyl1685n.d.n.d.n.d.n.d.
υst C=CAromatic ring vibration16051601159015901590
δ C-HMethylene13731373137313731373
υst S=OSufoxide10331030103110311031
υst C-O-CEthern.d.n.d.101510151015
υst -HC=CH-Trans-butadiene block~965n.d.~968~968~968
δ C-HStyrene block~748~748~748~748~748
δ C=CAromatic ring vibration813815815815815
Table 6. Thermal properties of bitumen samples.
Table 6. Thermal properties of bitumen samples.
SampleTonset (°C)Tmax (°C)Loss Weight (wt%)Char (wt%)
PW36142790%10.20
MB20045284.0016.00
NB27645685.7514.25
NBat150°-10 min27845687.0013.00
B127544286.2013.80
B227543485.0015.00
B327643585.3014.30
Table 7. MSCR results: Jnr and Jtot both at 40 and 50 °C.
Table 7. MSCR results: Jnr and Jtot both at 40 and 50 °C.
SampleJnrJtot
40 °C50 °C40 °C50 °C
0.1 kPa3.2 kPa0.1 kPa3.2 kPa0.1 kPa3.2 kPa0.1 kPa3.2 kPa
NB0.16280.17221.12301.21190.16470.17391.12661.2125
MB0.01150.01850.05090.07870.01700.02190.06390.0903
B10.03090.04930.17460.29810.03420.05150.18800.3030
B20.03600.04620.24030.35750.03890.04840.25380.3621
B30.00510.02050.04990.12040.01520.02240.06140.1258
Table 8. (a) Superpave optimization: HMANB vs. HMAB3 volumetric properties and (b) ITS and ITSM results.
Table 8. (a) Superpave optimization: HMANB vs. HMAB3 volumetric properties and (b) ITS and ITSM results.
(a)
Asphalt MixtureVaVMAVFA%Gmm atNdesDust Portion
(%)(%)(%)(%)(-)
HMANB414.270.685.57%1.1
HMAB341471.897.64%1.1
(b)
Asphalt MixtureITS (MPa)ITSM (MPa)
at 10 °Cat 10 °Cat 20 °Cat 40 °C
HMANB3.37148538120773
HMAMB2.971539186131051
HMAB34.381649895821440
Table 9. Summary of the asphalt blends and mixtures results.
Table 9. Summary of the asphalt blends and mixtures results.
10 °C20 °C40 °C
Asphalt MixtureHMANBHMAMBHMAB3HMANBHMAMBHMAB3HMANBHMAMBHMAB3
ITS (MPa)3.3702.9704.380n.d.n.d.n.d.n.d.n.d.n.d.
ITSM (MPa)14,85315,39116,49881208613958277310511440
Asphalt BinderNBMBB3NBMBB3NBMBB3
|G*| (MPa)20.08013.28425.8073.6013.1815.9010.8530.9950.274
G′ (MPa)14.41010.44820.1641.8372.1503.6090.1510.5930.102
G″ (MPa)13.85418.17015.9313.0842.3394.6470.8390.7990.254
δ (°)47.93541.67540.00363.77649.51455.95482.03153.39866.808
Jnr (kPa−1)n.d.n.d.n.d.n.d.n.d.n.d.0.1670.0150.013
Jtot (kPa−1)n.d.n.d.n.d.n.d.n.d.n.d.0.1690.0190.019
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Veropalumbo, R.; Russo, F.; Oreto, C.; Buonocore, G.G.; Verdolotti, L.; Muiambo, H.; Biancardo, S.A.; Viscione, N. Chemical, Thermal, and Rheological Performance of Asphalt Binder Containing Plastic Waste. Sustainability 2021, 13, 13887. https://doi.org/10.3390/su132413887

AMA Style

Veropalumbo R, Russo F, Oreto C, Buonocore GG, Verdolotti L, Muiambo H, Biancardo SA, Viscione N. Chemical, Thermal, and Rheological Performance of Asphalt Binder Containing Plastic Waste. Sustainability. 2021; 13(24):13887. https://doi.org/10.3390/su132413887

Chicago/Turabian Style

Veropalumbo, Rosa, Francesca Russo, Cristina Oreto, Giovanna Giuliana Buonocore, Letizia Verdolotti, Herminio Muiambo, Salvatore Antonio Biancardo, and Nunzio Viscione. 2021. "Chemical, Thermal, and Rheological Performance of Asphalt Binder Containing Plastic Waste" Sustainability 13, no. 24: 13887. https://doi.org/10.3390/su132413887

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