Next Article in Journal
Study on Urban Spatial Expansion and Its Scale Benefit in the Yellow River Basin
Next Article in Special Issue
Integrating BIM-LCA to Enhance Sustainability Assessments of Constructions
Previous Article in Journal
Characteristics and Variations in Korea through the Lens of Net-Zero Carbon Transformation in Cities
Previous Article in Special Issue
Fatigue of Cold Recycled Cement-Treated Pavement Layers: Experimental and Modeling Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Physical, Rheological, and Permanent Deformation Behaviors of WMA-RAP Asphalt Binders

by
Kátia Aline Bohn
1,*,
Liseane Padilha Thives
1 and
Luciano Pivoto Specht
2
1
Postgraduate Program in Civil Engineering, Federal University of Santa Catarina, Florianópolis 88040-900, Brazil
2
Department of Transportation, Federal University of Santa Maria, Santa Maria 97105-900, Brazil
*
Author to whom correspondence should be addressed.
Sustainability 2023, 15(18), 13737; https://doi.org/10.3390/su151813737
Submission received: 17 August 2023 / Revised: 12 September 2023 / Accepted: 13 September 2023 / Published: 14 September 2023

Abstract

:
With the rapid global expansion of road networks, the asphalt industry faces several environmental challenges, such as material shortages, environmental concerns, escalating material costs, demand for eco-friendly materials, and the implementation of “Net Zero” policies. Given these challenges and recognizing the need to explore new solutions, this research evaluated asphalt binder samples incorporating Warm Mix Asphalt (WMA) and Reclaimed Asphalt Pavement (RAP), or WMA-RAP. The assessment focused on analyzing the physical, rheological, and permanent deformation characteristics of WMA-RAP samples containing 20%, 35%, and 50% recycled pavement. The study utilized a chemical surfactant-type WMA additive, Evotherm® P25. The findings showed that the WMA-RAP combination resulted in increased stiffness ranging from 247% to 380% and a reduced phase angle of 16% to 26% with an increasing RAP content from 20% to 50% at Tref 20 °C and 10 Hz. Furthermore, the penetration decreased from 20% to 47%, and the softening point increased from 7% to 17%. An improvement of 2 PGHs was observed by adding 35% and 50% RAP. Additionally, WMA samples containing up to 50% RAP presented more elevated permanent deformation resistance, supporting traffic levels of 64V or 70H. WMA-RAP binders allow mixture production at lower temperatures—an amount of 30 °C less—conserving energy and decreasing the need for new aggregate materials by incorporating recycled materials, thus minimizing the environmental impact.

1. Introduction

Rapid economic growth, population expansion, and increased industrial production have historically relied on resource depletion, resulting in environmental issues such as pollution and global warming. Recycling and reusing waste materials offer potential solutions to reduce resource dependence and pollution. Given the enormous global demand for asphalt in road construction, Reclaimed Asphalt Pavement (RAP) has gained significant attention as an environmentally friendly alternative. Researchers have been increasingly exploring the incorporation of RAP into asphalt mixes to enhance sustainability and reduce environmental impact.
During a highway’s service life, pavement restoration interventions are inevitable, leading to a substantial generation of RAP. Pavement recycling is a technique that repurposes RAP in new mixtures, as this material possesses significant economic value due to its aged asphalt binder while concurrently reducing the demand for new aggregates, thus conserving natural resources [1]. Hence, studying methods and alternatives that incorporate RAP into new pavements without compromising their performance becomes imperative, given that aged binders tend to lose their properties due to oxidation, potentially affecting their behavior in conjunction with other materials.
Some research on RAP’s physical, rheological, and predictive behavior has yielded results that advance state-of-the-art knowledge of recycled pavement performance [2,3,4,5]. The asphalt binder extracted from RAP exhibits distinct chemical and rheological properties compared to virgin asphalt binders [6]. Consequently, specific concerns arise regarding RAP utilization: the selection of base binder, shifts in mixing and compaction temperatures, susceptibility to temperature, resistance to permanent deformation, fatigue performance, and chemical alterations [1].
Studies [7,8] have demonstrated that asphalt mixtures containing RAP can achieve performance equivalent to conventional Hot Mix Asphalt (HMA) or even superior mechanical performance [9]. Recycled pavement encompasses both aged and rigid asphalt binder, and thus, the inclusion of RAP in new mixtures augments resistance to permanent deformation [2,10,11,12]. In investigating the viscoelastic behavior of recycled mixtures, Zheng et al. [13] observed an increase in the elastic component with higher RAP content and loading frequency, enhancing stability at elevated temperatures. Furthermore, dynamic modulus values are augmented with decreasing temperature and increasing frequency [13]. However, RAP within HMA mixtures could potentially experience secondary aging during high-temperature mixing stages.
Further research indicates that the successful production of asphalt pavements incorporating RAP necessitates carefully selecting milling types and production processes for the mixtures [14,15]. Apart from reducing the consumption of virgin materials, RAP incorporation also reduces the production cost of asphalt mixtures compared to conventional blends [16].
Given the global prominence of “Net Zero” policies in recent years, driven by climate concerns and the need to reduce greenhouse gas emissions to limit global warming to 1.5 degrees Celsius above pre-industrial levels, the field of pavement engineering has explored more sustainable technologies. These policies aim to balance emissions with their removal, making them crucial in climate change minimization [17,18,19]. One such technology is Warm Mix Asphalt (WMA), which reduces sample manufacturing and compaction temperatures, thus conferring environmental benefits such as reduced energy consumption during production and, consequently, diminished fuel usage and emissions [20,21,22].
At the same time, with this technology, the focus on recycling gains prominence, as the lower heating temperature of WMA enables higher RAP content [23,24,25,26,27], mitigating further degradation of the aged binder in RAP. The potential to incorporate a greater amount of RAP is due to the reduced mixing temperature, resulting in less binder aging and assisting in counteracting the aging of the old binder from the recycled material [27]. Research indicates that RAP can enhance the performance of WMA mixtures compared to HMA [28,29,30,31].
Utilizing WMA substantially reduces fume exposure during pavement operations compared to HMA [26,32,33], with a fume reduction of about 50% for every 12 °C reduction in temperature [33]. Additionally, the energy required during the compaction process is diminished, aiding in greenhouse gas emission reduction. A 30 °C reduction in aggregate heating temperature leads to a 15% or more reduction in fuel costs for mixture production, translating to a threefold reduction in emissions [34]. Compared to a new asphalt mixture comprising virgin aggregates and binders, the cost of asphalt mixtures with 50% RAP decreases by over 30%, and WMA additives reduce energy consumption by around 20% [30].
Typically, mixtures with high RAP content consist of 20% to 30% incorporated recycled material, contingent on the country and available milling type [35], with the vast majority applied in HMA mixtures. Thus, the common practice involves employing 25% RAP [36]. Some studies evaluating the feasibility of incorporating high RAP contents into warm mixtures suggest that WMA technologies could be used with 75% RAP [37], 90 to 100% RAP [26], and 100% RAP [38], thereby enhancing the workability of the new asphalt mixture. Enhanced workability is a key feature of the WMA additive [25], and it is also applicable to mixtures without RAP. More documented information is needed regarding the utilization of RAP in WMA mixtures compared to conventional ones, as well as data on the long-term field performance with high proportions of recycled pavement, which is necessary to investigate their durability throughout their service life.
Numerous studies have explored the performance of WMA-RAP mixtures, encompassing different RAP proportions and various WMA additives [37,39,40,41,42,43]. The literature highlights that an increase in RAP within WMA enhances the resistance to permanent deformation of mixtures [28,30,37,39,40,41,42,43,44] due to the heightened rigidity of the aged binder derived from RAP.
However, the resistance to permanent deformation of WMA-RAP mixtures remains a concern, as Zhao et al. [42] revealed a greater rutting depth than HMA mixtures. Nevertheless, Kim et al. [28] indicated that the field performance of RAP-containing mixtures after 5 to 6 years of service demonstrated favorable resistance to permanent deformation, with mean rutting depths below 6 mm.
The studies predominantly underscored the same benefits of WMA-RAP, highlighting its significant environmental and economic potential. Reducing production and compaction temperatures of asphalt mixtures through WMA additives [20,21,22] makes the process ecologically sound and sustainable, enabling more significant RAP usage [25,29] and yielding promising performance outcomes for new mixtures [24,30,41].
Considering the viscoelastic behavior, which depends on temperature, time, and load application rate [45,46,47], alterations in the physical and rheological properties of asphalt materials can arise due to adding additives and aged binders. Hence, it is crucial to initially investigate the type of RAP to be employed and its interaction with virgin binder and WMA additives to ascertain the proportion of aged binder that can be incorporated into these materials without compromising the final sample’s physical, rheological, and performance properties. The main challenge remains the extraction of aged asphalt binder from RAP [48] for more in-depth studies concerning its interaction with other materials.
This study evaluated the utilization of WMA-RAP asphalt binders, incorporating different proportions of RAP, to evaluate their physical, rheological, and permanent deformation performance behaviors. Focusing on the asphalt binder level of analysis is justified as it helps explore and develop higher-quality mixtures. Various factors influence the characteristics and performance properties of asphalt materials when combined with additives and other asphalt binders. Consequently, the study compared the WMA-RAP samples with reference samples, including HMA, WMA, and the original RAP prototype, to assess potential shifts in behavior. The objective was to determine how the increased proportion of aged RAP material affects sample properties and whether the use of WMA can mitigate the accelerated aging observed in materials during new mixture production. Additionally, the study aimed to identify an optimal RAP percentage based on the findings.

2. Materials and Methods

2.1. Materials

The materials employed in the research comprised conventional unmodified virgin asphalt (Neat Binder, Curitiba, Brazil), Warm Mix Asphalt (WMA) additive, and Reclaimed Asphalt Pavement (RAP) binder (Neat RAP Binder).
The RAP originated from the milling of a surface course with approximately 2 years of micro-coating over a layer with 9 years of service, from a federal highway in the southern region of Brazil, BR-287, between the cities of Mata and São Vicente do Sul, in Rio Grande do Sul State.
The asphalt binder used was conventional unmodified, classified by penetration grade of 50/70. Results of the asphalt binder characterization are presented in this research. WMA additive incorporated was of the chemical surfactant type, Evotherm® P25 (Ingevity–North Charleston/SC/USA). Figure 1 presents the materials utilized. In Table 1 are indicated the physical and chemical properties of Evotherm® P25.

2.2. Method

Initially, the asphalt binder was extracted from Reclaimed Asphalt Pavement (RAP) using a rotary evaporator and trichloro-ethylene (TCE) as the solvent, according to the standard BS EN 12697-3+A1 [50]. The extraction process consisted of two phases:
  • The first phase was conducted at a temperature of 90 ± 5 °C and a pressure of 40 ± 5 kPa. During this phase, the solution (solvent + asphalt binder) was transferred to the evaporation flask.
  • The second phase was carried out at a temperature of 160 ± 5 °C and a pressure of 2 ± 0.5 kPa to remove any residual solvent from the sample. The process was considered complete when no more bubbles were observed forming in the sample within the evaporation flask.
The results of the RAP binder characterization are presented in this research.
Relative to the total asphalt binder content of the sample, increments of 0.5% of Warm Mix Asphalt (WMA) additive were added. A total of six asphalt binder samples were evaluated. For comparison, two reference samples without RAP were tested: one hot (HMA_Ref) and one warm (WMA_Ref). Additionally, the sample consisting solely of RAP binder (RAP_Original) was also evaluated. The WMA-RAP samples were tested with 20%, 35%, and 50% RAP incorporation rates, designated as “RAP”, followed by the incorporated RAP content in percentage. The compositions of the studied samples are presented in Table 2.
The sample preparation protocol followed the guidelines of ASTM D4887 [51]. The asphalt binders were heated to a maximum temperature of 135 °C. In a glass beaker, proportions of conventional asphalt binder, RAP, and WMA additive were added and homogenized using a glass rod. The sample was maintained in the oven at 135 °C for 10 min and, after 5 min, mixed using a glass rod for 30 s. After this process, the sample was ready to perform the tests.
The samples were characterized and evaluated in their virgin/original condition and after short-term aging. The binders were aged using the Rolling Thin Film Oven (RTFO) equipment, following the ASTM D2872 [52] guidelines. The RTFO test was conducted at 133 °C for the additive-modified samples, which was 30 °C lower than the samples without additive, as per the manufacturer’s recommendation for temperature reduction when working with WMA additives.
To achieve the study objective of assessing the behavior of WMA-RAP samples in comparison to reference samples with varying RAP contents, with a focus on physical, rheological, and performance characterization related to permanent deformation, the conducted tests are detailed in Table 3.
Furthermore, the study aimed to evaluate how aging could impact the behavior of these samples and determine the optimal percentage of RAP to provide high-temperature performance pavements. Figure 2 illustrates the research flowchart.

2.2.1. High-Temperature Performance Grade

The High-Temperature Performance Grade (PGH) and continuous PGH were determined by analyzing the data obtained from the Dynamic Shear Rheometer (DSR), including the Dynamic Shear Modulus (|G*|) and Phase Angle (δ). The controlled deformation test followed the stop criteria and deformation values outlined in Table 4. Furthermore, the PGH was also determined using data from the Multi-Stress Creep and Recovery Test (MSCR) to establish the level of traffic stress absorbed by each asphalt binder.

2.2.2. Master Curves

The materials were also subjected to a frequency sweep in a linear ramp, ranging from 0.1 to 30 Hz, and a temperature sweep from 5 °C to 75 °C, which provided data for |G*| and δ. The test was conducted at 5 °C, 20 °C, and 35 °C using an 8 mm geometry and a 2 mm gap. For temperatures of 35 °C, 60 °C, and 75 °C, a 25 mm geometry and a 1 mm gap were employed. The test was carried out under a controlled deformation of 1%.
The δ value corresponds to the elastic-to-viscous response ratio during the shear process, indicating the relative amount of recoverable and non-recoverable deformation. The |G*| is a measure of the material’s overall resistance to deformation when subjected to repeated pulses of shear deformation. Both parameters reflect the linear viscoelastic characterization of the asphalt binders.
With the material responses under different temperatures and loading frequencies, master curves of |G*| and δ were constructed using the Time-Temperature Superposition Principle (TTSP) for a reference temperature (Tref) of 20 °C. The shifting factors (aT) were obtained using Equation (1) and adjusted using the Williams-Landel-Ferry (WLF) model, as shown in Equation (2).
a T T = f ( T ) f ( T r e f )                           a T T r e f = 1 ,
log a T = C 1 × T T r e f C 2 + T T r e f ,
where C1 and C2 represent the model’s shifting constants.
The master curves were modeled using the 2S2P1D model (2 springs, 2 parabolic elements, and 1 dashpot), as proposed by [61]. The model parameters are defined by Equation (3).
G * ω = G + G 0 G 1 + δ ( i ω τ ) k + ( i ω τ ) h + ( i ω β τ ) 1 ,
where G is the static modulus as ω → 0 (MPa), G0 is the glassy modulus as ω → ∞ (MPa), k and h are dimensionless constants of the parabolic elements, with 0 < k < h < 1, β is a dimensionless constant related to the linear dashpot viscosity, δ is a dimensionless shape constant, τ is the characteristic time, which varies only with temperature (s), and ω is the pulsation 2πf (s−1).

2.2.3. Multi-Stress Creep and Recovery Test (MSCR)

Following the Superpave classification [58], the MSCR test was conducted at the PGH temperature and at temperatures corresponding to −1PGH and +1PGH, ensuring that the results were obtained at the same temperature for the evaluated samples. The samples were tested for their virgin/original and RTFO conditions.
The creep and recovery scheme comprised 20 cycles at 0.1 kPa and 10 cycles at 3.2 kPa. Each cycle lasts 10 s, with 1 s of shear deformation and 9 s of rest. The lower stress level (first 20 cycles) falls within the linear viscoelastic (LVE) domain, while the higher stress level (last 10 cycles) is in the damage domain. The lower stress level corresponds to light traffic, and the higher stress level corresponds to heavy traffic conditions.
The initial 10 cycles at 0.1 kPa serve for sample conditioning, while the last 10 cycles at 0.1 kPa and the following 10 cycles at 3.2 kPa are employed to derive key critical parameters from the test. These parameters include the asphalt binder’s delayed elastic response, quantified by the percentage of recovery (%R), an evaluation of the material’s resistance to permanent deformation through non-recoverable compliance (Jnr), and an assessment of the material’s sensitivity to changes in stress level, determined by the percent difference in non-recoverable creep compliance between 0.1 kPa and 3.2 kPa (Jnrdiff).

3. Results and Discussions

3.1. Physical Characterization

3.1.1. Penetration

Figure 3 presents the average results from the penetration tests conducted on both virgin/original samples at 25 °C. It can be observed that the Warm Mix Asphalt (WMA) additive led to a reduction of approximately 12% in penetration when compared to WMA_Ref (49 10−1 mm) as opposed to HMA_Ref (56 10−1 mm). With more Reclaimed Asphalt Pavement (RAP) content, penetration decreased, indicating an increase in sample stiffness. Notably, RAP_Original exhibited significantly lower penetration, measuring only 7 10−1 mm, attributed to the aging resulting from its field history of field application. This aging effect influenced the reduction in penetration observed in the WMA-RAP samples (from 20% to 47%), with the reduction being directly proportional to the percentage of RAP added to the sample.
Regarding the retained penetration results after RTFO (Figure 4), it was evident that the WMA-RAP samples exhibited a higher percentage of retained penetration, less than 10%. On the other hand, the reference hot and warm samples experienced more significant penetration losses. These findings may indicate that the WMA samples, subjected to short-term aging at 133 °C, experienced lesser binder oxidation than the HMA sample. The RAP_Original sample, due to its advanced aging, presented a retained penetration of 86%. Thus, the WMA additive contributed to sustaining the physical properties after mixing and field compaction, maintaining similar retained penetration across the three studied WMA-RAP samples, regardless of the incorporated RAP ratio.
The RAP contributed to maintaining penetration values based on the results obtained for WMA-RAP samples after RTFO. The results of the WMA-RAP samples indicated that this finding could be even more significant when the WMA was incorporated, as the additive contributed to the maintenance of the penetration of similar value and mitigated the aging effect. Also, these samples were less susceptible to aging than those without RAP and WMA additives.
From the penetration results of both WMA_Ref and RAP_Original samples, it was possible to estimate the penetration value of the WMA-RAP samples based on the quantity of RAP incorporated. However, further testing on various types of RAPs is necessary to substantiate this indication.

3.1.2. Softening Point

Figure 5 shows the softening point test results. The HMA_Ref sample exhibited a softening point of 49 °C, while the WMA_Ref sample had a slightly higher value at 51 °C. As the proportion of RAP in the WMA-RAP samples increased, the softening point also increased, ranging from 7% to 17%. This rise indicated enhanced rigidity in the samples. However, it’s worth noting that this increase showed a slight decrease as the RAP content in the sample became more substantial.
The Increment of Softening Point (ISP), calculated as the difference between the softening point after short-term aging and the softening point of the samples in the virgin/original condition, is presented in Figure 6. HMA_Ref exhibited an increase of 4 °C, while RAP_Original presented an increase of 4.5 °C. WMA-RAP samples displayed a similar variation in the softening point after short-term aging, ranging from 2 °C to 2.5 °C, irrespective of the content of incorporated RAP. It could have occurred because the RFTO of WMA samples was set at 133 °C, indicating a reduced aging sensitivity of the WMA-RAP.
Similar to the penetration test results, it was possible to indicate that the WMA-RAP samples were less susceptible to aging and that short-term aging at a temperature of 133 °C significantly improved the physical properties of the samples. Furthermore, the softening point results suggest that the addition of RAP and WMA additives can enhance the asphalt binder’s resistance to permanent deformation. However, attention must be paid to the increase in the softening point of the virgin/original samples due to the addition of RAP compared to the other tests to assess its influence on the performance of these asphalt binders.

3.2. Rheological Characterization

3.2.1. Viscosity

The apparent viscosity was measured to evaluate the material’s workability at elevated temperatures. The tests were conducted using a Brookfield Viscometer equipped with spindle 21 at temperatures of 135 °C, 150 °C, and 177 °C, with rotations set at 20, 50, and 100 rpm, respectively. Figure 7 presents the obtained results.
The results reveal that the viscosity of HMA_Ref (360 cP) and WMA_Ref (305 cP) at 135 °C was quite similar, confirming the manufacturer’s indication that the use of the WMA additive does not significantly influence viscosity. As expected, at 135 °C, the viscosity increased, ranging from 44% to 181% for WMA-RAP samples with increasing RAP content. However, it’s important to note that the viscosity of RAP_Original (5310 cP) was notably higher than the other samples due to significant aging in the RAP. This sample would not meet the necessary viscosity limit for asphalt workability, which is limited to 3000 cP at 135 °C [57,58].
On the other hand, the other samples exhibit values below the limit, demonstrating that the use of the WMA additive with conventional virgin binder contributed to higher RAP contents while maintaining satisfactory workability, even with up to 50% RAP incorporation. Notably, the difference in viscosity among binders tended to decrease as the temperature increased.
Other studies have also concluded that WMA-RAP asphalt mixtures exhibit satisfactory workability [62]. Furthermore, according to Mejías-Santiago et al. [63], WMA-RAP mixtures showed adequate workability in comparison to HMA-RAP, with this characteristic decreasing as more RAP content was added to WMA mixtures.
The results of apparent viscosity after short aging are presented in Figure 8, and it was noted that all samples exhibit more elevated viscosity values after RTFO. Similarly to the virgin/original samples, viscosity increased with the addition of more significant RAP proportions. Table 5 presents the viscosity increments (RV), revealing that samples without WMA additives exhibited a higher susceptibility to aging.
In a broader context, the behavior of WMA-RAP samples was consistent, with slight differences in stiffness enhancement compared to the virgin/original condition, with an increase ranging from approximately 7% to 27% across all tested temperatures. In contrast, for HMA_Ref and RAP_Original, this variation ranged from 18% to 45%. It is worth highlighting that RV decreased as the test temperature increased.

3.2.2. High-Temperature Performance Grade

The PGH and continuous PGH of the evaluated samples were determined. To comprehend the sample behavior due to RAP incorporation, the criteria were made to present the PGH results in the virgin/original and RTFO condition states, considering the Superpave requirements (Table 4).
To assess the conformity of the samples with the specification of performance-graded asphalt binder [57], the mass change after RTFO was determined. As previously mentioned, for samples with WMA additive, the test was conducted at 133 °C. The obtained results can be found in Table 6. Under the limitations of ASTM D2872 [52], the mass change of the samples must not exceed 1%. All samples exhibited mass losses below the specified limit in the standard. The most significant variation was observed in the RAP_Original sample, which displays a change of 0.11%. The remaining samples vary from 0.01% to 0.04%, indicating minimal mass alteration when subjected to the short-term aging process.
In the virgin/original conditions, as can be seen in Table 7, PGH remained unchanged for the HMA_Ref and WMA_Ref (PGH 64 °C) but increased with the increment of RAP in the WMA-RAP samples. Continuous PGH enhanced from 69.9 °C to 79.4 °C with RAP increasing from 20% to 50%. The RAP binder exhibited a PGH exceeding 94 °C. However, testing at even higher temperatures was not feasible due to the technical limitations of the DSR equipment. The results of PGH virgin/original samples were higher than those after RTFO aging.
An improvement can be highlighted by analyzing the PGH results of the WMA-RAP samples after RTFO (Table 7). The samples with a 20% RAP increment showed a rise of 1PGH (64 °C), while those with 35% and 50% RAP content exhibited an increase of 2PGHs (70 °C) compared to the WMA_Ref (58 °C). The RAP_50% approached a PGH 76 grade. In terms of continuous PGH, there was an increase of 4.1 to 5.3 °C for WMA-RAP samples with each incremental percentage of RAP.
Comparing continuous PGH results for binders after RTFO with virgin/original samples, it was observed that samples subjected to short-term aging experienced a more substantial reduction in performance when the WMA additive was introduced. HMA_Ref and RAP_Original presented the slightest variation in continuous PGH, 1.7 °C and 0.9 °C, respectively. Conversely, the WMA_Ref presented a 3.5 °C difference, and the WMA-RAP samples obtained a difference of more than 4 °C (4.1 °C for RAP_20%, 4.7 °C for RAP_35%, and 4.2 °C for RAP_50%). This might imply that the WMA additive may have influenced this parameter, as including RAP does not necessarily represent an increase or decrease in this difference. Due to the inability to test PG at a temperature of 100 °C, the continuous PGH of the RAP_Original sample was an estimation.
Limited data are available regarding the high-temperature characteristics of WMA-RAP samples. However, some studies have shown promising results. For instance, Dong et al. [64] found that WMA-RAP samples with foam asphalt exhibited superior performance at high temperatures and greater temperature sensitivity compared to WMA alone. Similarly, Yu et al. [65] demonstrated that adding RAP increases the failure temperature of WMA samples with foam asphalt. Additionally, Almeida Junior et al. [66] observed an increase in PGH with greater binder aging. These findings suggest that WMA-RAP samples may have advantages in high-temperature performance.
In summary, the results of PGH suggest that the addition of WMA with 20% RAP has a minimal impact on the binder’s performance, with no deterioration compared to the reference sample. This implies that it is feasible to produce WMA-RAP binders that offer economic and sustainable advantages for paving without compromising high-temperature pavement performance compared to conventional binders. Moreover, incorporating higher RAP contents, such as 35% and 50%, can potentially enhance performance even further.

3.2.3. Linear Viscoelastic Characterization—Master Curves

The rheological properties describe the material’s behavior at different temperatures and loading frequencies. This study tested virgin/original samples and those subjected to RTFO. From TTSP, master curves for |G*| and δ were constructed for a TRef of 20 °C. The translation factors (aT) and the constants of the Williams-Landel-Ferry (WLF) model are presented in Table 8. The acquired data were fitted to the 2S2P1D rheological model, and the corresponding parameters are listed in Table 9.
The master curves of |G*| and δ for the virgin/original binders are presented in Figure 9 and Figure 10, respectively. These curves reveal that the dynamic modulus increased as the frequency rose. This increase can be attributed to reduced exposure to loading at higher frequencies and the limited time for reversible deformations to occur. At higher frequencies, the material exhibits predominantly elastic deformations, resulting in peak modulus values. In contrast, at lower frequencies, viscoelastic deformations become more pronounced due to the longer loading times, leading to lower dynamic modulus values and higher phase angle values.
Moreover, Figure 9 illustrates that the WMA_Ref sample obtained lower stiffness (|G*|) compared to the HMA_Ref sample across the entire frequency spectrum. This difference is attributed to the influence of the WMA additive on rheological properties, making the material less stiff. On the other hand, the master curves for the WMA-RAP samples showed increased stiffness as the RAP content increased. This stiffening effect was more pronounced with higher RAP content and increased loading frequency, ultimately enhancing the material’s stability at elevated temperatures. When analyzing stiffness at Tref 20 °C and 10 Hz, it is possible to measure that the stiffness increased by 247%, 314%, and 380%, respectively, for samples containing 20%, 35%, and 50% of RAP. The RAP_Original exhibited substantially higher stiffness than the other samples and a lower phase angle. This can be attributed to oxidation and aging that occurred in the field.
Figure 10 reveals that the lowest values of δ were observed in the RAP_Original sample, followed by samples with the highest to lowest increments of RAP and the reference binders. Under the conditions of Tref 20 °C and 10 Hz, the phase angle decreased by 16%, 21%, and 26% for samples containing 20%, 35%, and 50% of RAP, respectively. At lower frequencies, less than 1 × 10−4, the samples presented a similar behavior, approaching a purely viscous response (δ = 90°), except for the RAP reference sample, which exhibited a lower value. However, at intermediate and high frequencies, the addition of RAP led to a more elastic behavior rather than a tendency toward viscous flow. This resulted in increased energy storage within the material, allowing these materials to accumulate stress more rapidly during repeated deformation processes.
The master curves for the aged samples are shown in Figure 11 and Figure 12. When comparing the results of the virgin/original and aged samples, it can be noted that the curves exhibit greater stiffness, leading to an overall increase in the |G*| value for all binders (Figure 11) across the entire frequency spectrum. Relative to the δ value (Figure 12), there is a decrease in the phase angle due to the short-term aging process, indicating a higher degree of aging in the samples, which can be attributed to the RTFO procedure.
Furthermore, from the master curves, changes due to the aging effect in |G*| and δ were observed by the magnitude variations along the reduced frequency range under consideration. The rheological alterations occurred non-proportionally concerning the frequency spectrum and temperature. Regarding |G*|, the lower frequencies and higher temperatures undergo relatively more significant changes compared to the higher frequencies and lower temperatures for all evaluated samples. In contrast, for δ, a shift of nearly the entire curve towards lower phase angle values was stated, with more substantial alterations at higher frequencies and lower temperatures. These changes were more pronounced in the reference samples (HMA_Ref, WMA_Ref, and RAP_Original), indicating that the WMA-RAP samples were less sensitive to aging, likely due to the lower temperature of the RTFO tests.

3.3. Permanent Deformation

Multi-Stress Creep and Recovery (MSCR) Test

In order to predict permanent deformation, the MSCR test was conducted, and the parameters Jnr and R% were calculated. The test was performed at different temperatures for each sample, ranging from 58 °C to 94 °C, as presented in Table 10, ensuring the same tested temperatures for comparison.
Table 10 results showed that in all samples, the R% value was zero, confirming that the RAP binder originates from a conventional binder, similar to the virgin binder used in this study. Therefore, since R% represents a measure of recovery capacity typically found in polymers, no recovery indicated that the studied asphalts do not contain any polymers in their composition and do not regain their properties when subjected to stresses and deformations in the MSCR test. Thus, the WMA-RAP samples exhibited compromised elastic responses like the reference samples, which already displayed a null %R value.
The samples exhibited high Jnr0.1 and Jnr3.2 values (Table 10), attributed to the potentially poor performance of the virgin binder. This was more evident when comparing the data from the HMA_Ref and WMA_Ref samples, which showed similar results for the tests conducted at temperatures ranging from 58 °C to 70 °C. However, it is worth highlighting that the Jnr3.2 values were higher for the warm sample, indicating its lower resistance to permanent deformation than the hot sample.
The WMA-RAP samples exhibited improved characteristics (Table 10) regarding permanent deformation, with the influence of RAP contributing to this behavior. As the RAP content increased, the Jnr3.2 values decreased, indicating enhanced resistance to higher traffic levels (Figure 13). Similar findings were reported by Zaremotekhases et al. [67], who demonstrated that samples with RAP, at both testing levels, exhibited lower Jnr values than those without RAP, indicating the stiffening effect of RAP binder. According to Yu et al. [65], increasing amounts of RAP binder lead to a decrease in the Jnr value of WMA-RAP samples with foam asphalt and an increase in the %R value. Thus, a higher amount of RAP binder enhanced the asphalt’s resistance to permanent deformation. The Jnrdiff (%) values (Table 10) were within the acceptable range according to the standard, ranging from 1.8% to 8.2%, which is lower than the 75% limit [58], indicating minimal interference from the test levels of MSCR.
According to AASHTO M332 [58], based on the value obtained for Jnr3.2, the material can be classified for Standard “S” traffic (2.0 < Jnr3.2 < 4.5), Heavy “H” traffic (1.0 < Jnr3.2 < 2.0), Very Heavy “V” traffic (0.5 < Jnr3.2 < 1.0), and Extremely Heavy “E” traffic (0.0 < Jnr3.2 < 0.5). As seen in Table 10, adding 20% RAP increased the PGH from 58S (reference samples) to 64S. The addition of 35% RAP, a material with the same PG as the RAP_20% sample, was achieved but classified for Heavy traffic (64H). When the RAP proportion was increased to 50%, the material was suitable for areas with PGH 64 for Very Heavy traffic (64V). Additionally, the RAP_50% sample met PGH 70H, indicating its ability to withstand PGH 70 conditions for Heavy traffic. The RAP_Original binder exhibited high performance with low Jnr3.2 values, which could be suitable for situations with PGH 88H or even 94S. Considering that the adopted virgin binder sample is of the conventional type, this significant improvement in permanent deformation with WMA-RAP samples could cover a wide range of PGH levels and traffic conditions on Brazilian highways.
Due to the elevated Jnr3.2 values obtained in the results, the materials did not perform well in terms of permanent deformation at certain temperatures, mainly the virgin binder. On the other hand, testing the asphalt mixture is essential since the aggregate matrix plays a crucial role in permanent deformation. Thus, it is essential to evaluate the behavior of these materials through mixture production and mechanical tests to understand their performance.
Corroborating with the results obtained for softening point and PGH, it is evident that there is an improvement in the performance concerning permanent deformation with the addition of WMA and RAP. The greater the increment of recycled pavement material, the more pronounced the improvement in the behavior of the asphalt binder, which aligns with the resource-efficient and environmentally conscious principles of “Net Zero” policies. This study demonstrated that WMA-RAP blends present a viable alternative in locations requiring enhanced pavement performance under high-temperature conditions. As a result, these findings highlight the crucial role of WMA-RAP blends in mitigating carbon emissions and minimizing the broader environmental impact, perfectly aligning with the core tenets and objectives of “Net Zero” policies focused on sustainability.

4. Conclusions

This study aimed to investigate Warm Mix Asphalt (WMA) technology with Reclaimed Asphalt Pavement (RAP) addition and assess how these materials interact at the asphalt binder level. The economic benefits and sustainability aspects, considering greenhouse gas reduction policies and the rising global average temperature [20,21,22], make these types of materials promising solutions to be widely adopted in the coming years [29,30,31].
Moreover, it aligns with the United Nations’ Sustainable Development Goal 9, which centers on industry, innovation, and infrastructure. Specifically, this research contributed to the development of resilient infrastructure by creating advanced materials and technologies for longer-lasting and sustainable critical structures. Furthermore, it promoted sustainable practices within the construction industry and presented innovative solutions to infrastructure challenges, fostering innovation in the field. These points underscore the significance of the study.
Compared to reference samples, the research evaluated WMA-RAP samples and their physical, rheological, and permanent deformation behavior characteristics. The results demonstrated that increasing the RAP content significantly enhanced sample performance, highlighting the potential of WMA-RAP as a sustainable paving alternative.
The main findings of this study can be summarized as follows:
  • The interaction between the virgin binder, WMA additive, and RAP binder showed no negative behavior in high-temperature tests.
  • Physical assessment indicated a decrease in penetration (20% to 47%) and an increase in softening point (7% to 17%), including 20% to 50% RAP in the samples.
  • Apparent viscosity rose with higher RAP content at 135 °C (from 44% to 181%). However, the workability of all WMA-RAP samples remained suitable, staying below 3000 cP at 135 °C.
  • Samples with added RAP displayed increased PGH values compared to reference samples: 1 PGH at 20% RAP and 2 PGHs at 35% and 50% RAP. Additionally, continuous PGH analysis showed a temperature rise of 4.1 °C to 5.3 °C with each RAP content increase in WMA-RAP samples.
  • Master curves showed increased material stiffness and decreased phase angle, indicating aging when RAP binder was added. However, WMA-RAP samples displayed reduced sensitivity to aging across the entire frequency spectrum compared to reference samples, likely due to the RTFO being conducted at 133 °C.
  • MSCR results showed that increasing the RAP binder content improved resistance to permanent deformation in WMA-RAP samples. The Jnr3.2 value decreased as higher RAP contents were added, indicating the ability to withstand heavier traffic loads for the 64S, 64H, and 64V/70H categories, respectively, in samples with 20%, 35%, and 50% RAP.
  • WMA samples with up to 50% RAP exhibited satisfactory performance regarding permanent deformation. Despite increased aging due to higher RAP content, they showed higher PGH values, improved physical and rheological properties, and enhanced performance. Therefore, WMA-RAP samples demonstrated satisfactory behavior, even with significant RAP.
The findings indicate that permanent deformation resistance is not a barrier to utilizing WMA-RAP. This allows for lower-temperature mix production, reducing energy consumption, enhancing sustainability, and utilizing recycled materials. Nevertheless, chemical analysis is required for a more comprehensive understanding of material interactions. Additionally, evaluating fatigue resistance, water susceptibility, and testing other RAP binders should be considered for future research validation.

Author Contributions

Conceptualization, K.A.B., L.P.T. and L.P.S.; methodology, K.A.B.; validation, L.P.T. and L.P.S.; formal analysis, K.A.B.; investigation, K.A.B.; resources, L.P.T.; writing—original draft preparation, K.A.B.; writing—review and editing, L.P.T. and L.P.S.; visualization, K.A.B.; supervision, L.P.T.; project administration, L.P.T.; funding acquisition, L.P.T. and L.P.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received external funding in the form of a scholarship awarded to the first author. The scholarship was generously provided by the National Council for Scientific and Technological Development (CNPq), under Grant No. 432588/2016-7, and by the Coordination of Higher Level Staff Improvement—Brazil (CAPES) with the Finance Code 001.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data are shown in the manuscript.

Acknowledgments

The authors would like to thank CNPq/BRASIL, CAPES/BRASIL, and CBB Asfaltos for donating Neat Asphalt and WMA additive.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Singh, D.; Sawant, D. Understanding effects of RAP on rheological performance and chemical composition of SBS modified binder using series of laboratory tests. Int. J. Pavement Res. Technol. 2016, 9, 178–189. [Google Scholar] [CrossRef]
  2. Khosla, N.P.; Nair, H.; Visintine, B.; Malpass, G. Effect of reclaimed asphalt and virgin binder on rheological properties of binder blends. Int. J. Pavement Res. Technol. 2012, 5, 317–325. [Google Scholar]
  3. Stimilli, A.; Ferrotti, G.; Conti, C.; Tosi, G.; Canestrari, F. Chemical and rheological analysis of modified bitumens blended with artificial reclaimed bitumen. Constr. Build. Mater. 2014, 63, 1–10. [Google Scholar] [CrossRef]
  4. Xiao, F.; Putman, B.; Amirkhanian, S. Rheological characteristics investigation of high percentage RAP binders with WMA technology at various aging states. Constr. Build. Mater. 2015, 98, 315–324. [Google Scholar] [CrossRef]
  5. Wu, S.; Zhang, W.; Shen, S.; Li, X.; Muhunthan, B.; Mohammad, L.N. Field-aged asphalt binder performance evaluation for Evotherm warm mix asphalt: Comparisons with hot mix asphalt. Constr. Build. Mater. 2017, 156, 574–583. [Google Scholar] [CrossRef]
  6. Al-Qadi, I.L.; Elseifi, M.; Carpenter, S.H. Reclaimed Asphalt Pavement—A Literature Review; Illinois Center for Transportation: Rantoul, IL, USA, 2007. [Google Scholar]
  7. West, R.; Kvasnak, A.; Tran, N.; Powell, B.; Turner, P. Testing of Moderate and High Reclaimed Asphalt Pavement Content Mixes. Transp. Res. Rec. J. Transp. Res. Board 2009, 2126, 100–108. [Google Scholar] [CrossRef]
  8. Gennesseaux, M.M.L. Avaliação da Durabilidade de Misturas Asfálticas a Quente e Mornas Contendo Material Asfáltico Fresado. Ph.D. Thesis, Escola Politécnica da Universidade de São Paulo, São Paulo, Brazil, 2015. [Google Scholar]
  9. Sapkota, K.; Yaghoubi, E.; Wasantha, P.L.P.; Van Staden, R.; Fragomeni, S. Mechanical Characteristics and Durability of HMA Made of Recycled Aggregates. Sustainability 2023, 15, 5594. [Google Scholar] [CrossRef]
  10. Hussain, A.; Yanjun, Q. Effect of reclaimed asphalt pavement on the properties of asphalt binder. Procedia Eng. 2013, 54, 840–850. [Google Scholar] [CrossRef]
  11. Hossain, Z.; Zaman, M.; Solanki, P.; Ghabchi, R.; Singh, D.; Adje, D.; Lewis, S. Implementation of MEPDG for Asphalt Pavement with RAP; Final Report—University of Oklahoma; Oklahoma Transportation Center: Rantoul, IL, USA, 2013. [Google Scholar]
  12. Huang, S.; Turner, T.F. Aging characteristics of RAP blend binders: Rheological properties. Am. Soc. Civil Eng. 2014, 26, 966–973. [Google Scholar] [CrossRef]
  13. Zheng, K.; Xu, J.; Wang, J. Viscoelasticity of Recycled Asphalt Mixtures with High Content Reclaimed SBS Modified Asphalt Pavement. Sustainability 2023, 15, 2515. [Google Scholar] [CrossRef]
  14. Sabouri, M.; Bennert, T.; Daniel, J.S.; Kim, Y.R. Fatigue and rutting evaluation of laboratory-produced asphalt mixtures containing reclaimed asphalt pavement. Transp. Res. Rec. J. Transp. Res. Board 2015, 2506, 32–44. [Google Scholar] [CrossRef]
  15. Lo Presti, D.; Carrión, A.J.B.; Airey, G.; Hajj, E. Towards 100% recycling of reclaimed asphalt in road surface courses: Binder design methodology and case studies. J. Clean. Prod. 2016, 131, 43–51. [Google Scholar] [CrossRef]
  16. Correa, B.M.; Specht, L.P.; Schuster, S.L.; Almeida Junior, P.O.B.; Faccin, C.; Boeira, F.D.; Pereira, D.S.; Nascimento, L.A.H.; Brito, L.A.T. Fatigue Performance of Recycled Asphalt Mixtures: Viscoelastic Continuum Damage Approach and Cost Analysis. Sustainability 2023, 15, 8794. [Google Scholar] [CrossRef]
  17. Climate Action—European Commission. Available online: https://climate.ec.europa.eu/eu-action/climate-strategies-targets/2050-long-term-strategy_pt (accessed on 24 August 2023).
  18. Climate Action—United Nations. Available online: https://www.un.org/en/climatechange/net-zero-coalition (accessed on 24 August 2023).
  19. Climate Action Tracker. Available online: https://climateactiontracker.org/ (accessed on 24 August 2023).
  20. Wang, Q.-Z.; Chen, Z.-D.; Lin, K.-P.; Wang, C.-H. Estimation and Analysis of Energy Conservation and Emissions Reduction Effects of Warm-Mix Crumb Rubber-Modified Asphalts during Construction Period. Sustainability 2018, 10, 4521. [Google Scholar] [CrossRef]
  21. Espinoza, J.; Medina, C.; Calabi-Floody, A.; Sánchez-Alonso, E.; Valdés, G.; Quiroz, A. Evaluation of Reductions in Fume Emissions (VOCs and SVOCs) from Warm Mix Asphalt Incorporating Natural Zeolite and Reclaimed Asphalt Pavement for Sustainable Pavements. Sustainability 2020, 12, 9546. [Google Scholar] [CrossRef]
  22. Martinez-Soto, A.; Calabi-Floody, A.; Valdes-Vidal, G.; Hucke, A.; Martinez-Toledo, C. Life Cycle Assessment of Natural Zeolite-Based Warm Mix Asphalt and Reclaimed Asphalt Pavement. Sustainability 2023, 15, 1003. [Google Scholar] [CrossRef]
  23. Barthel, W.; Marchand, J.P.; Von Devivere, M. Warm Asphalt Mixes by Adding a Synthetic Zeolite. In Proceedings of the Eurasphalt & Eurobitume Congress, Viena, Austria, 12–14 May 2004. [Google Scholar]
  24. Prowell, B.D.; Hurley, G.C.; Crews, E. Field Performance of Warm Mix Asphalt at the NCAT Test Track. Transp. Res. Board J. Transp. Res. Board 2007, 1, 1–9. [Google Scholar]
  25. Chowdhury, A.; Button, J.W. A Review of Warm Mix Asphalt; Technical Report; Texas Transportation Institute: College Station, TX, USA, 2008. [Google Scholar]
  26. D’angelo, J.; Harm, E.; Bartoszek, J.; Baumgardner, G.; Corrigan, M.; Cowsert, J.; Harman, T.; Jamshidi, M.; Jones, W.; Newcomb, D.; et al. Warm-Mix Asphalt: European Practice. International Technology Scanning Program; Federal Highway Administration: Ashburn, VA, USA, 2008.
  27. Rubio, M.C.; Moreno, F.; Martínez-Echevarría, M.J.; Martínez, G.; Vásquez, J.M. Comparative analysis of emissions from the manufacture and use of hot and half-warm mix asphalt. J. Clean. Prod. 2013, 41, 1–6. [Google Scholar] [CrossRef]
  28. Kim, D.; Norouzi, A.; Kass, S.; Liske, T.; Kim, Y.R. Mechanistic performance evaluation of pavement sections containing RAP and WMA additives in Manitoba. Constr. Build. Mater. 2017, 133, 39–50. [Google Scholar] [CrossRef]
  29. Saberi, K.F.; Fakhri, M.; Azami, A. Evaluation of warm mix asphalt mixtures containing reclaimed asphalt pavement and crumb rubber. J. Clean. Prod. 2017, 165, 1125–1132. [Google Scholar] [CrossRef]
  30. Song, W.; Huang, B.; Shu, X. Influence of warm-mix asphalt technology and rejuvenator on performance of asphalt mixtures containing 50% reclaimed asphalt pavement. J. Clean. Prod. 2018, 192, 191–198. [Google Scholar] [CrossRef]
  31. Lu, D.X.; Saleh, M.; Nguyen, N.H.T. Effect of rejuvenator and mixing methods on behaviour of warm mix asphalt containing high RAP content. Constr. Build. Mater. 2019, 197, 792–802. [Google Scholar] [CrossRef]
  32. SABITA—Southern African Bitumen Association. Manual 32: Best Practice Guideline for Warm Mix Asphalt; SABITA: Cape Town, South Africa, 2011; ISBN 978-1-8974968-55-1. [Google Scholar]
  33. EAPA—European Asphalt Pavement Association. Asphalt. Available online: https://eapa.org/asphalt/ (accessed on 11 March 2019).
  34. Motta, R.S. Misturas Mornas: Redução de Emissão de Poluentes e Economia de Energia. In Workshop de Avaliação dos Objetivos da Produção de Asfaltos; WAOP: São Paulo, Brazil, 2011. [Google Scholar]
  35. Austroads. Maximising the Re-Use of Reclaimed Asphalt Pavement: Outcomes of Year Two: RAP Mix Design. AP-T286-15; Austroads: Sydney, NSW, Australia, 2015. [Google Scholar]
  36. Copeland, A. Reclaimed Asphalt Pavement in Asphalt Mixtures: State of the Practice. Report N°. FHWA-HRT-11-021; Federal Highway Administration: McLean, VA, USA, 2011.
  37. Mallick, R.B.; Kandhal, P.S.; Bradbury, R.L. Using warm–mix asphalt technology to incorporate high percentage of reclaimed asphalt pavement material in asphalt mixtures. Transp. Res. Board J. Transp. Res. Board 2008, 2051, 1–8. [Google Scholar] [CrossRef]
  38. Tao, M.; Mallick, R.B. Effects of Warm-Mix Asphalt Additives on Workability and Mechanical Properties of Reclaimed Asphalt Pavement Material. Transp. Res. Board J. Transp. Res. Board 2009, 2126, 151–160. [Google Scholar] [CrossRef]
  39. Lee, S.J.; Armikhanian, S.N.; Park, N.-W.; Kim, K.W. Characterization of warm mix asphalt binders containing artificially long-term aged binders. Constr. Build. Mater. 2009, 23, 2371–2379. [Google Scholar] [CrossRef]
  40. Rogers, W. Influence of Warm Mix Additives upon High RAP Asphalt Mixes. Ph.D. Thesis, Clemson University, Clemson, SC, USA, 2011. [Google Scholar]
  41. Hill, B.; Behnia, B.; Buttlar, W.G.; Reis, H. Evaluation of Warm Mix Asphalt Mixtures Containing Reclaimed Asphalt Pavement through Mechanical Performance Tests and an Acoustic Emission Approach. J. Mater. Civil Eng. 2013, 25, 1887–1897. [Google Scholar] [CrossRef]
  42. Zhao, S.; Huang, B.; Shu, X.; Woods, M. Comparative evaluation of warm mix asphalt containing high percentages of reclaimed asphalt pavement. Constr. Build. Mater. 2013, 44, 92–100. [Google Scholar] [CrossRef]
  43. Evotherm—Ingevity. Available online: https://www.ingevity.com/markets/asphalt-and-paving/evotherm/ (accessed on 22 March 2019).
  44. Nejad, F.M.; Azarhoosh, A.; Hamedi, G.H.; Roshani, H. Rutting performance prediction of warm mix asphalt containing reclaimed asphalt pavements. Road Mater. Pavement Des. 2014, 15, 207–219. [Google Scholar] [CrossRef]
  45. Bohn, K.A. Avaliação das Misturas Asfálticas Mornas com uso de Ligantes Convencional e Modificado por Polímero. Master’s Thesis, Programa de Pós-Graduação em Engenharia Civil, Centro de Tecnologia, Universidade Federal de Santa Maria, Santa Maria, Brazil, 2017. [Google Scholar]
  46. Di Benedetto, H.; Gabet, T.; Grenfell, J.; Perraton, D.; Sauzéat, C.; Bodin, D. Mechanical Testing of Bituminous Mixtures. In Advances in Interlaboratory Testing and Evaluation of Bituminous Materials; RILEM State-of-the-Art Reports; Partl, M.N., Bahia, H.U., Canestrari, F., Roche, C., Benedetto, H., Piber, H., Sybilski, D., Eds.; Springer: New York, NY, USA, 2013; Volume 9, pp. 143–256. [Google Scholar]
  47. Mangiafico, S. Linear Viscoelastic Properties and Fatigue of Bituminous Mixtures Produced with Reclaimed Asphalt Pavement and Corresponding Binder Blends. Ph.D. Thesis, L’École Nationale des Travaux Publics de L’État, Paris, France, 2014. [Google Scholar]
  48. Specht, L.P.; Babadopulos, L.F.A.; Benedetto, H.; Sauzeat, C.; Soares, J.B. Application of the theory of viscoelasticity to evaluate the resilient modulus test in asphalt mixes. Constr. Build. Mater. 2017, 149, 648–658. [Google Scholar] [CrossRef]
  49. Júnior, P.O.B.A.; Schuster, S.L.; Faccin, C.; Vestena, P.M.; Ilha, P.S.; Menegusso Pires, G.; Müller, E.I.; Pereira, D.S.; Specht, L.P. Rheological, permanent deformation and fatigue analysis for calibration of the recovery process of bitumens in the rotary evaporator. Road Mater. Pavement Des. 2023, 24, 1–26. [Google Scholar] [CrossRef]
  50. BS EM 12697-3:2013+A1:2018; Bituminous Mixtures—Test Methods—Part 3: Bitumen Recovery: Rotary Evaporator. European Standard, CEN: Brussels, Belgium, 2018.
  51. ASTM D4887; Standard Practice for Preparation of Viscosity Blends for Hot Recycled Bituminous Materials. American Society for Testing and Materials: West Conshohocken, PA, USA, 2016.
  52. ASTM D2872; Standard Test Method for Effect of Heat and Air on a Moving Film of Asphalt (Rolling Thin-Film Oven Test). American Society for Testing and Materials: West Conshohocken, PA, USA, 2012.
  53. ASTM D5; Standard Test Method for Penetration of Bituminous Materials. American Society for Testing and Materials: West Conshohocken, PA, USA, 2013.
  54. ASTM D36; Standard Test Method for Softening Point of Bitumen (Ring-and-Ball Apparatus). American Society for Testing and Materials: West Conshohocken, PA, USA, 2014.
  55. ASTM D4402; Standard Test Method for Viscosity Determination of Asphalt at Elevated Temperatures Using a Rotational Viscometer. American Society for Testing and Materials: West Conshohocken, PA, USA, 2015.
  56. ASTM D7175; Standard Test Method for Determining the Rheological Properties of Asphalt Binder Using a Dynamic Shear Rheometer. American Society for Testing and Materials: West Conshohocken, PA, USA, 2015.
  57. AASHTO M320; Standard Specification for Performance-Graded Asphalt Binder. AASHTO: Washington, DC, USA, 2017.
  58. AASHTO M332; Standard Specification for Performance-Graded Asphalt Binder Using Multiple Stress Creep Recovery (MSCR) Test. AASHTO: Washington, DC, USA, 2018.
  59. ASTM D7405; Standard Test Method for Multiple Stress Creep and Recovery (MSCR) of Asphalt Binder Using a Dynamic Shear Rheometer. American Society for Testing and Materials: West Conshohocken, PA, USA, 2015.
  60. AASHTO T350; Standard Method of Test for Multiple Stress Creep Recovery (MSCR) Test of Asphalt Binder Using a Dynamic Shear Rheometer (DSR). AASHTO: Washington, DC, USA, 2018.
  61. Olard, F.; Di Benedetto, H. General “2S2P1D” Model and Relation between the Linear Viscoelastic Behaviours of Bituminous Binders and Mixes. Road Mater. Pavement Des. 2003, 4, 185–224. [Google Scholar]
  62. Hettiarachchi, C.; Hou, X.; Wang, J.; Xiao, F. A comprehensive review on the utilization of reclaimed asphalt material with warm mix asphalt technology. Constr. Build. Mater. 2019, 227, 117096. [Google Scholar] [CrossRef]
  63. Mejías-Santiago, M.; Doyle, J.D.; Rushing, J.F. Accelerated pavement testing of warm-mix asphalt for heavy-traffic airfield. Transp. Res. Board J. Transp. Res. Board 2014, 2456, 11–20. [Google Scholar] [CrossRef]
  64. Dong, F.; Yu, X.; Xu, B.; Wang, T. Comparison of high temperature performance and microstructure for foamed WMA and HMA with RAP binder. Constr. Build. Mater. 2017, 134, 594–601. [Google Scholar] [CrossRef]
  65. Yu, X.; Dong, F.; Xu, B.; Ding, G.; Ding, P. RAP Binder Influences on the Rheological Characteristics of Foamed Warm-Mix Recycled Asphalt. J. Mater. Civil Eng. 2017, 29, 04017145. [Google Scholar] [CrossRef]
  66. Almeida, P.O.B.d., Jr. Reciclagem de Misturas 100% RAP: Investigação de Parâmetros para Avaliação e Dosagem de Agentes Rejuvenescedores com foco na Fadiga. Ph.D. Thesis, Programa de Pós-Graduação em Engenharia Civil, Centro de Tecnologia, Universidade Federal de Santa Maria, Santa Maria, Brazil, 2022. [Google Scholar]
  67. Zaremotekhases, F.; Sadek, H.; Hassan, M.; Berryman, C. Impact of warm-mix asphalt technologies and high reclaimed asphalt pavement content on the performance of alternative asphalt mixtures. Constr. Build. Mater. 2022, 319, 126035. [Google Scholar] [CrossRef]
Figure 1. Research materials.
Figure 1. Research materials.
Sustainability 15 13737 g001
Figure 2. Experimental flowchart of the asphalt binders analyzed.
Figure 2. Experimental flowchart of the asphalt binders analyzed.
Sustainability 15 13737 g002
Figure 3. Penetration test results of virgin/original samples.
Figure 3. Penetration test results of virgin/original samples.
Sustainability 15 13737 g003
Figure 4. Retained penetration results.
Figure 4. Retained penetration results.
Sustainability 15 13737 g004
Figure 5. Softening point test data in virgin/original condition.
Figure 5. Softening point test data in virgin/original condition.
Sustainability 15 13737 g005
Figure 6. Increment of Softening Point test results.
Figure 6. Increment of Softening Point test results.
Sustainability 15 13737 g006
Figure 7. Apparent viscosity results.
Figure 7. Apparent viscosity results.
Sustainability 15 13737 g007
Figure 8. Apparent viscosity after RTFO.
Figure 8. Apparent viscosity after RTFO.
Sustainability 15 13737 g008
Figure 9. Master curves of |G*| under virgin/original conditions.
Figure 9. Master curves of |G*| under virgin/original conditions.
Sustainability 15 13737 g009
Figure 10. Master curves of δ at virgin/original condition.
Figure 10. Master curves of δ at virgin/original condition.
Sustainability 15 13737 g010
Figure 11. Master curves of |G*| after RTFO.
Figure 11. Master curves of |G*| after RTFO.
Sustainability 15 13737 g011
Figure 12. Master curves of δ after RTFO.
Figure 12. Master curves of δ after RTFO.
Sustainability 15 13737 g012
Figure 13. Jnr3.2 data after RTFO with AASHTO M332 [58] limits.
Figure 13. Jnr3.2 data after RTFO with AASHTO M332 [58] limits.
Sustainability 15 13737 g013
Table 1. Evotherm® P25 properties.
Table 1. Evotherm® P25 properties.
Physical StateLiquid
ColorBrown
OdorMild
pH2.3 [Conc. (% w/w): 5%] (in 50:50 IPA/Water)
Melting Point<−16 °C
Boiling Point>100 °C
Flash PointClosed Cup: 181 °C
Relative Density0.99
SolubilityInsoluble in cold and hot water
ViscosityDynamic (room temperature): 487 cP
Source: Data adapted from supplier [49].
Table 2. Composition of the evaluated asphalt binder samples.
Table 2. Composition of the evaluated asphalt binder samples.
Asphalt Binder
Nomenclature
Sample Composition
% Asphalt Binder% RAP
HMA_Ref1000
WMA_Ref1000
RAP_20%8020
RAP_35%6535
RAP_50%5050
RAP_Original0100
Table 3. Tests and evaluated properties for the characterization of the samples and their respective standards.
Table 3. Tests and evaluated properties for the characterization of the samples and their respective standards.
Properties/TestsStandard
PenetrationASTM D5 [53]
Softening PointASTM D36 [54]
Viscosity (135 °C, 150 °C e 177 °C)ASTM D4402 [55]
Complex Modulus (G*) and Phase Angle (δ)–PGH and Master CurvesASTM D7175 [56], AASHTO M320 [57] and AASHTO M332 [58]
Elastic Recovery (%R) e Non-Recoverable Creep Compliance (Jnr)-MSCRASTM D7405 [59] and AASHTO T350 [60]
Table 4. Stop criteria and controlled deformation percentage used in the test.
Table 4. Stop criteria and controlled deformation percentage used in the test.
SampleG*/Senδ (kPa)Deformation (%)
Virgin/Original Binder1.012
RTFO Binder2.210
Source: Adapted from ASTM D7175 [56].
Table 5. Increment of Viscosity data.
Table 5. Increment of Viscosity data.
SampleRV135°CRV150°CRV177°C
HMA_Ref1.231.281.18
WMA_Ref1.301.211.16
RAP_20%1.271.241.18
RAP_35%1.181.181.07
RAP_50%1.211.171.10
RAP_Original1.451.411.25
Table 6. Mass change after RTFO.
Table 6. Mass change after RTFO.
Asphalt Binder SampleTemperature Test (°C)Mass Change
HMA_Ref1630.03%
WMA_Ref1330.01%
RAP_20%1330.03%
RAP_35%1330.04%
RAP_50%1330.03%
RAP_Original1630.11%
Table 7. PGH and continuous PGH data for samples in virgin/original and RTFO states.
Table 7. PGH and continuous PGH data for samples in virgin/original and RTFO states.
SampleVirgin/OriginalRTFO
PGH (°C)Continuous PGH (°C)PGH (°C)Continuous PGH (°C)
HMA_Ref6464.05862.3
WMA_Ref6464.25860.7
RAP_20%6469.96465.8
RAP_35%7074.67069.9
RAP_50%7679.47075.2
RAP_Original94 *96.8 *94 *95.9 *
* Estimated data, as the working limit of the DSR used in the research was 94 °C.
Table 8. Master curve parameters.
Table 8. Master curve parameters.
Sampleslog (aT)WLF
52035456075C1C2
Virgin/Original
HMA_Ref2.400.00−1.90−3.05−4.05−4.9014.99113.20
WMA_Ref2.650.00−1.95−2.85−3.90−4.7514.66114.70
RAP_20%2.250.00−2.05−3.25−4.45−5.3515.96109.09
RAP_35%2.300.00−2.00−3.30−4.45−5.4016.15109.47
RAP_50%2.250.00−1.95−3.25−4.43−5.4516.56112.10
RAP_Original2.450.00−1.85−3.15−4.50−5.6516.74107.92
RTFO
HMA_Ref2.500.00−2.00−3.10−4.15−5.0015.68117.46
WMA_Ref2.450.00−1.90−2.90−3.90−4.7015.09121.60
RAP_20%2.250.00−2.00−3.15−4.20−5.1515.97115.51
RAP_35%2.250.00−2.00−3.20−4.35−5.3015.83109.26
RAP_50%2.250.00−1.95−3.25−4.45−5.4516.56112.10
RAP_Original2.350.00−1.95−3.15−4.60−5.8016.93105.57
Table 9. 2S2P1D model parameters.
Table 9. 2S2P1D model parameters.
SamplesParameters
G00 (Pa)G0 (Pa)khδτ0β
Virgin/Original
HMA_Ref01.70 × 1090.250.659.02.50 × 10−560
WMA_Ref01.70 × 1090.300.688.02.50 × 10−540
RAP_20%04.00 × 1090.280.6513.02.50 × 10−5120
RAP_35%04.50 × 1090.230.6515.03.50 × 10−5180
RAP_50%05.00 × 1090.230.6617.05.50 × 10−5200
RAP_Original09.00 × 1090.200.6025.01.00 × 10−41400
RTFO
HMA_Ref03.80 × 1090.240.6514.01.30 × 10−5110
WMA_Ref02.20 × 1090.270.659.02.00 × 10−550
RAP_20%04.00 × 1090.240.6414.02.00 × 10−5130
RAP_35%04.50 × 1090.220.6516.03.00 × 10−5210
RAP_50%05.00 × 1090.220.6518.05.00 × 10−5260
RAP_Original09.00 × 1090.190.5826.01.20 × 10−43300
Table 10. MSCR test data after RTFO.
Table 10. MSCR test data after RTFO.
SamplePGHPGH Classification MSCRParameterTest Temperature (°C)
586470768894
HMA_Ref5858SJnr0.1 (kPa−1)2.6636.00512.355---
R0.1 (%)0.0000.0000.000---
Jnr3.2 (kPa−1)2.8136.25412.863---
R3.2 (%)0.0000.0000.000---
Jnrdiff (%)5.6064.1604.115---
WMA_Ref5858SJnr0.1 (kPa−1)3.5067.78916.137---
R0.1 (%)0.0000.0000.000---
Jnr3.2 (kPa−1)3.6668.16216.770---
R3.2 (%)0.0000.0000.000---
Jnrdiff (%)4.5694.7883.925---
RAP_20%6464SJnr0.1 (kPa−1)-3.5097.76016.220--
R0.1 (%)-0.0000.0000.000--
Jnr3.2 (kPa−1)-3.6958.17516.906--
R3.2 (%)-0.0000.0000.000--
Jnrdiff (%)-5.3095.3504.231--
RAP_35%7064HJnr0.1 (kPa−1)-1.9294.4129.432--
R0.1 (%)-0.0000.0000.000--
Jnr3.2 (kPa−1)-2.0034.66210.036--
R3.2 (%)-0.0000.0000.000--
Jnrdiff (%)-4.8965.6756.405--
RAP_50%7064V ou 70HJnr0.1 (kPa−1)-0.8432.0854.651--
R0.1 (%)-0.0000.0000.000--
Jnr3.2 (kPa−1)-0.9132.2164.968--
R3.2 (%)-0.0000.0000.000--
Jnrdiff (%)-8.2206.2656.814--
RAP_Original94 *88H ou 94SJnr0.1 (kPa−1)----1.4673.199
R0.1 (%)----0.0000.000
Jnr3.2 (kPa−1)----1.5523.257
R3.2 (%)----0.0000.000
Jnrdiff (%)----5.8041.802
* Estimated data, as the working limit of the DSR used in the research was 94 °C.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Bohn, K.A.; Thives, L.P.; Specht, L.P. Physical, Rheological, and Permanent Deformation Behaviors of WMA-RAP Asphalt Binders. Sustainability 2023, 15, 13737. https://doi.org/10.3390/su151813737

AMA Style

Bohn KA, Thives LP, Specht LP. Physical, Rheological, and Permanent Deformation Behaviors of WMA-RAP Asphalt Binders. Sustainability. 2023; 15(18):13737. https://doi.org/10.3390/su151813737

Chicago/Turabian Style

Bohn, Kátia Aline, Liseane Padilha Thives, and Luciano Pivoto Specht. 2023. "Physical, Rheological, and Permanent Deformation Behaviors of WMA-RAP Asphalt Binders" Sustainability 15, no. 18: 13737. https://doi.org/10.3390/su151813737

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