Next Article in Journal
Improved Optoelectronic Properties and Temporal Stability of AZO/Cu/AZO Films by Inserting an Ultrathin Al Layer
Previous Article in Journal
Shear-Induced Graphitization in Tongyuanpu Shear Zone, Liaodong Peninsula of Eastern China: Insights from Graphite Occurrences, Nanostructures and Carbon Sources
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effects of the Combined Incorporation of ZnO and TiO2 Nanoparticles on the Mechanical, Rheological, Thermal, and Healing Properties of a Dense Polymeric Asphalt Mixture

by
Jaqueline Wolfart
,
João Victor Staub de Melo
*,
Alexandre Luiz Manfro
,
Breno Salgado Barra
and
Rafael Cassimiro Barbosa
Department of Civil Engineering, Federal University of Santa Catarina, Rua João Pio Duarte Silva, Florianópolis 88040-970, SC, Brazil
*
Author to whom correspondence should be addressed.
Nanomaterials 2025, 15(23), 1779; https://doi.org/10.3390/nano15231779
Submission received: 4 November 2025 / Revised: 23 November 2025 / Accepted: 24 November 2025 / Published: 26 November 2025
(This article belongs to the Section Nanocomposite Materials)

Abstract

This study evaluated the combined incorporation of zinc oxide (ZnO) and titanium dioxide (TiO2) nanoparticles into a styrene–butadiene–styrene (SBS) copolymer-modified asphalt binder, aiming to increase thermal conductivity and healing potential while maintaining rheological performance. Nanocomposites containing ZnO + TiO2 (50/50 wt.%) were produced at dosages of 2–12 wt.% and subjected to the Rolling Thin Film Oven Test (RTFOT), thermal conductivity measurements, viscosity testing, and rheological characterization. A dense-graded asphalt mixture with the optimized dosage was evaluated through wheel-tracking, four-point bending fatigue and healing, and internal heating rate assessment under microwave radiation. The integrated results indicated 8.5 wt.% as the optimal dosage, providing a 106.3% increase in thermal conductivity and improving the high-temperature performance grade (PGH) from 76-XX to 82-XX. Non-recoverable creep compliance (Jnr) decreased by 21.1%, and viscosity at 135 °C increased by 41.8%, remaining below 3.0 Pa·s. In the asphalt mixture, healing capacity increased by 50.7%, and the internal heating rate by 50.0%, while the wheel-tracking rut depth decreased by 13.3%. These findings demonstrate that 8.5 wt.% ZnO + TiO2 simultaneously enhances heat conduction, healing efficiency, and resistance to permanent deformation, offering a promising solution for pavements subjected to high temperatures and heavy traffic.

1. Introduction

Metallic semiconductor nanomaterials, such as zinc oxide (ZnO) and titanium dioxide (TiO2), have been widely investigated due to their photocatalytic properties, particularly their ability to degrade environmental pollutants [1,2,3,4]. In pavement engineering, these materials have been incorporated into cementitious [5,6] and asphalt matrices [7,8,9,10] with the aim of reducing pollutant concentrations in urban environments. In asphalt binders and mixtures, typical dosages range from 2 to 15 wt.% [7,8,10,11,12,13].
Beyond their photocatalytic activity, the isolated use of nano-ZnO and nano-TiO2 in asphalt binders has been extensively studied for their effects on high-temperature rheological performance. Studies show that increasing the concentration of these nanomaterials raises the apparent viscosity [14,15], the high-temperature performance grade (PGH) [7,14,16], and the dynamic shear modulus (|G*|) [7,14,17,18,19], while reducing the non-recoverable creep compliance (Jnr) [7,14,20] and the phase angle (δ) [17,18,19], consistently indicating improvements in resistance to permanent deformation.
At the mixture scale, the incorporation of TiO2 has shown significant gains in indirect tensile strength, stiffness modulus, fracture energy, and rutting resistance [15,21,22]. Similarly, ZnO has been reported to enhance crack propagation resistance, fracture energy, moisture resistance, and permanent deformation performance [23,24,25].
Despite these advances, the literature predominantly investigates ZnO and TiO2 independently, and knowledge of their combined effects remains limited, particularly in SBS-modified systems. Recent studies have used both oxides, but with a primarily physicochemical or binder-level rheological focus. For example, Fu et al. [26] studied composite TiO2/ZnO–basalt fiber-modified binders, evaluating microstructure and rheology. Xie et al. [27] investigated an SBS/ZnO/TiO2 system aimed at improving UV resistance and low-temperature behavior through physico-mechanical testing and aging indices. Rocha Segundo et al. [28] combined low dosages of TiO2 and ZnO to impart photocatalytic activity to the binder, evaluating physicochemical behavior and rheology after aging. Zhang et al. [29] employed combinations of ZnO, TiO2, and polymers to simultaneously improve high- and low-temperature performance of base asphalts.
However, none of these studies evaluated ZnO + TiO2 nanocomposites in SBS-modified binders using methodologies oriented toward the functional performance of asphalt mixtures, such as four-point bending fatigue, intermediate-temperature rheology, microwave-induced internal heating, healing, or wheel-tracking permanent deformation. Thus, the integrated application of these performance-related analyses to such nanocomposites remains unexplored in the literature.
Another relevant gap concerns the behavior of polymeric asphalt binders. Although SBS improves resistance to permanent deformation and viscoelastic performance, its intrinsic healing capacity is limited because the molecular diffusion required for microcrack closure is often insufficient due to restricted internal heat redistribution. Considering the metallic and semiconducting nature of ZnO and TiO2 [30,31,32], as well as their thermal conductivity [33,34], it is hypothesized that these nanomaterials may enhance heat conduction and redistribution, promoting more favorable healing conditions in SBS binders. Furthermore, their role as physical reinforcement may contribute to increased structural stiffness, more uniform thermal response, and improved polymer dispersion, supporting enhanced performance under severe conditions [35,36].
Given this scenario, this study presents the first comprehensive evaluation of ZnO + TiO2 nanocomposites applied to SBS-modified binders, establishing an optimal dosage based on combined thermal and rheological criteria and demonstrating their multiscale effects on asphalt mixtures. The investigation examined whether dosages commonly used in photocatalytic applications could also enhance the healing potential of dense asphalt mixtures, thereby expanding the functional role of these nanomaterials beyond photocatalysis. This approach sought to develop a mixture capable not only of achieving high mechanical performance at elevated temperatures but also of leveraging these conditions to intensify the healing process, which represents an important strategy to mitigate failures such as rutting, particularly in scenarios where thermal healing and photocatalytic activity are most effective.
For this purpose, ZnO and TiO2 were incorporated at a 1:1 ratio (50/50 wt.%) and at concentrations ranging from 2 to 12 wt.% (in 2% increments) in a polymeric asphalt binder. Thermal conductivity and rheological tests were used to map the dose–response behavior, identify inflection points between reinforcement and potential penalties, and determine the optimal dosage range. This optimal dosage was then applied to mixtures subjected to wheel-tracking tests, intermediate-temperature rheology, fatigue, healing, and microwave-induced heating, providing quantitative evidence of the effectiveness of the ZnO + TiO2 combination for high-performance pavement applications.

2. Materials

The materials used in this investigation consisted of crushed mineral aggregates, a polymer-modified asphalt binder, and metallic oxide nanoparticles (ZnO and TiO2).
The crushed mineral aggregates were of granitic origin and were supplied by Sul Brasil Mineração (SBM), located in the municipality of Paulo Lopes, in the state of Santa Catarina, Brazil. These aggregates were selected for their suitable properties for use in asphalt mixtures, meeting the required specifications. Table 1 presents the main physical and mechanical characteristics of the aggregates, which are essential for asphalt mixture design.
The polymeric asphalt binder used in this research was manufactured and supplied by CBB Asfaltos, located in the municipality of Curitiba, in the state of Paraná, Brazil. It consists of a binder modified with 4% of a styrene-butadiene-styrene (SBS) copolymer, specifically the D1101 A grade, commercialized by Kraton Polymers (Houston, TX, USA). The modification was carried out on an industrial scale, ensuring homogeneity and quality of the final product. The main properties of the polymeric asphalt binder are presented in Table 2, while the technical characteristics of the incorporated SBS copolymer are detailed in Table 3.
The ZnO and TiO2 nanoparticles, both with dimensions in the nanometric range, were purchased from Nanostructured & Amorphous Materials, Inc., located in Houston, TX, USA. These nanomaterials were selected due to their specific physicochemical properties, such as high surface area, high purity, and controlled crystalline structure. Table 4 presents the main technical characteristics provided by the manufacturer. In addition to the information supplied by the manufacturer, field emission scanning electron microscopy (FEG-SEM) and thermal analyses, including thermogravimetry (TGA) and derivative thermogravimetry (DTG), were performed.
The morphological characterization of the nanoparticles was carried out using field emission scanning electron microscopy (FEG-SEM) to assess particle shape and size at the nanometric scale. The tests were conducted with an ultra–high-vacuum field emission scanning electron microscope, model JEOL JSM-6701F (Akishima, Tokyo, Japan). To obtain high-resolution and high-quality images, consistent operational parameters were adopted, including a magnification of 50,000× and an acceleration voltage of 10 kV. The resulting micrographs (Figure 1) reveal the morphological features of the ZnO and TiO2 nanoparticles.
Figure 1 shows the scanning electron micrographs of zinc oxide (ZnO) (Figure 1A) and titanium dioxide (TiO2) (Figure 1B) nanoparticles. Both materials exhibit predominantly spherical to ellipsoidal morphology. Additionally, both types of nanoparticles have dimensions below 100 nm, as evidenced by the micrographs, which is consistent with the characteristics reported by the manufacturer.
The thermogravimetric (TGA) and derivative thermogravimetric (DTG) analyses were conducted to assess the thermal stability of the nanoparticles, considering the thermal environment typical of asphalt mixture production and application stages. The analyses were performed using an STA 449 F3 Jupiter®—Netzsch (Selb, Bavaria, Germany) thermoanalyzer, following the procedures described in ASTM E2550 [61]. The mass loss curves as a function of temperature for both nanomaterials are presented in Figure 2. In the presented graphs, particular emphasis was given to the temperature range between 150 °C and 170 °C, since this interval corresponds to the thermal conditions commonly used during polymeric asphalt binder modification and asphalt mixture production and compaction. The evaluation of thermogravimetric behavior within this range is crucial, as the thermal stability of the nanomaterials determines their ability to withstand degradation or significant mass loss during processing. Such analysis ensures that the nanomaterials retain their functional properties and contribute effectively to the performance of the asphalt mixture throughout the entire production and application process.
As shown in Figure 2A, ZnO exhibited a mass loss of 1.0% at 150 °C and 1.2% at 170 °C. Similarly, Figure 2B shows that TiO2 exhibited a mass loss of 1.1% at 150 °C and 1.3% at 170 °C. These slight mass reductions, observed within the temperature range typical of processes involving asphalt mixtures, are exclusively associated with particle dehydration, that is, the release of physically adsorbed water from the surface, and not with chemical or structural degradation of the materials [62,63,64,65].Therefore, the results confirm the adequate thermal stability of the nanomaterials (ZnO and TiO2) within the operational temperature range relevant to asphalt paving applications.
In summary, the selected materials (granitic aggregates, polymeric asphalt binder with SBS D1101 A (Kraton Polymers, Houston, TX, USA), and ZnO and TiO2 nanoparticles) provide the foundation for the experimental program. The following sections describe the procedures for binder nanomodification, thermal and rheological evaluation for determining the optimal nanoparticle content, and, finally, the mechanical testing of the asphalt mixtures.

3. Methods

To achieve the objectives proposed in this research, an experimental program was structured in sequential and interdependent stages. The principal phases that comprised the experimental development are presented below: (Step 1) Nanomodification of the Polymeric Asphalt Binder; (Step 2) Analysis of Thermal Conductivity and Rheological Behavior to Define the Optimum Content; and (Step 3) Assessment of the Mechanical Performance of Asphalt Mixtures. The overall structure of the adopted methodology is presented schematically in Figure 3. In the subsequent sections, each stage composing the experimental procedure is described in detail to ensure the reproducibility and understanding of the investigative process employed.

3.1. Nanomodification of the Polymeric Asphalt Binder

The nanomodification of the polymeric asphalt binder was carried out by incorporating combined concentrations of ZnO and TiO2 ranging from 2%, 4%, 6%, 8%, 10%, to 12% by weight. The selection of this concentration range (2–12 wt.%, in 2% increments) was based on literature references that investigated the isolated use of these nanomaterials in studies focused on the photocatalytic degradation of pollutants in asphalt matrices [7,8,10,11,12,13]. Establishing this interval enabled the assessment of whether concentrations typically employed in photocatalytic applications could also enhance the self-healing potential of a dense asphalt mixture, thereby extending the functional performance of the nanomaterials beyond their conventional photocatalytic role.
A 1:1 mass ratio (50/50 wt.%) between ZnO and TiO2 was maintained across all addition levels. This proportion was adopted based on studies conducted in non-asphaltic matrices [66,67,68], which indicate that equimass ZnO–TiO2 formulations tend to exhibit superior photocatalytic performance compared with asymmetric ratios. It is also noteworthy that studies evaluating the combined use of ZnO and TiO2 in asphalt binders or mixtures remain scarce, which further justified the adoption of the 1:1 ratio as the reference condition in this research. The detailed composition corresponding to each concentration level used in this phase of the study is presented in Table 5.
The nanomodification protocol for the polymeric asphalt binder, established based on previous studies conducted by the research group involving nanoparticle-modified binders [35,36,69], was designed to ensure methodological consistency and direct comparability with those works. The procedure was structured into three main stages (Figure 4). In the first stage, the binder samples were preheated in an oven at 160 ± 5 °C (apparent viscosity between 0.510 and 0.777 Pa·s), a temperature widely employed in the literature for the nanomodification of asphalt binders [7,26,28,70]. In the second stage, the nanomaterials were manually added to approximately 200 mL of base binder under continuous stirring for 10 to 15 min, ensuring their initial wetting and pre-dispersion. In the third stage, high-shear mixing was performed using a 700 W Philco® electric mixer (Joinville, Santa Catarina, Brazil) operating at 6000 rpm. The process lasted 40 min, conducted in 1 min stirring cycles followed by 1 min of rest, resulting in 20 min of effective high-shear application. The total energy density applied to the system during the nanomodification process was 404.5 J/m3.

3.1.1. Bright-Field Microscopy

After the nanomodification process, bright-field microscopy was performed to investigate the distribution of the nanomaterials within the asphalt binder. Micrographs were obtained using an Olympus IX83 inverted microscope (Shinjuku, Tokyo, Japan), operating under diascopic bright-field illumination and equipped with a 40× objective lens. The unaged samples N0, N2, N4, N6, N8, N10, and N12 were analyzed. Sample preparation consisted of depositing and spreading a small amount of asphalt binder onto 26 mm × 76 mm glass slides to form a thin, homogeneous film. The samples were then covered with two 20 mm × 20 mm glass coverslips, each with a controlled thickness of 0.13 mm and 0.16 mm, to protect the material and enable the acquisition of higher-resolution images.

3.1.2. Short-Term Aging

Following the experimental procedure, short-term aging of the binder samples was performed using the Rotary Thin Film Oven Test (RTFOT) to simulate the thermal and oxidative conditions to which asphalt matrices are subjected during production and paving operations. The residues obtained after RTFOT were subsequently used for characterization and performance testing, including X-ray diffraction (XRD), Fourier-transform infrared spectroscopy (FTIR), thermal conductivity measurement, high-temperature performance grade (PGH) evaluation, assessment of susceptibility to permanent deformation, phase angle and dynamic shear modulus behavior at intermediate temperatures, as well as the linear amplitude sweep (LAS) test.
Short-term aging was carried out in accordance with ASTM D2872 [71], using a Rolling Thin Film Oven Test apparatus, model James Cox and Sons CS 325-B (Colfax, CA, USA), maintained at 163 °C for 85 min. The procedure was applied to the reference polymeric asphalt binder sample (N0) and to all nanocomposites developed with different nanomaterial contents (N2, N4, N6, N8, N10, and N12). In addition, mass loss (%) was determined for each sample, adopting a maximum acceptable limit of 1% [72]. Mass loss was calculated from the difference in sample mass before and after RTFOT aging, multiplied by 100 to express the result as a percentage. This evaluation is essential for verifying the volatilization of binder components and ensuring the thermal stability of the nanocomposites under temperature conditions representative of asphalt paving operations.

3.1.3. X-Ray Diffraction

The X-ray diffraction (XRD) technique was employed to investigate the structural characteristics of the asphalt binder and to verify the incorporation of the crystalline phases of ZnO and TiO2 into the polymeric asphalt binder matrix. The analyses were performed using a Rigaku Miniflex II diffractometer (Akishima, Tokyo, Japan). All matrices (N0, N2, N4, N6, N8, N10, and N12) were evaluated under short-term aging conditions (RTFOT), with one sample tested for each condition. For comparison purposes, the ZnO and TiO2 nanomaterials were also analyzed individually. The scans were conducted over the 2θ range of 5° to 90°, with a scanning rate of 0.025°/s.

3.1.4. Fourier-Transform Infrared Spectroscopy

Fourier-transform infrared spectroscopy (FTIR) was employed to investigate possible chemical interactions between the polymeric asphalt binder matrix and the incorporated nanomaterials, aiming to identify potential structural modifications or the formation of new chemical bonds resulting from the nanomodification process. The analyses were performed using a Bruker FT-IR VERTEX 70 spectrometer (Billerica, MA, USA), selected for its high sensitivity and spectral precision. Representative samples of each asphalt binder matrix (N0, N2, N4, N6, N8, N10, and N12), all previously aged through RTFOT, were analyzed. For comparison and identification of characteristic nanomaterial bands, ZnO and TiO2 were also tested individually. Spectra were collected from 400 to 4000 cm−1, covering the main vibrational bands of the polymeric asphalt binder and the nanomaterials. Each spectrum was obtained from 16 scans at a resolution of 4 cm−1, ensuring suitable optimization of acquisition time relative to data quality.

3.2. Analysis of Thermal Conductivity and Rheological Behavior to Define the Optimum Content

The analysis of thermal conductivity and high-temperature rheological behavior was conducted to assess the effects of nanomodification on the thermal and viscoelastic properties of the polymeric asphalt binder. These evaluations were fundamental for defining the optimized incorporation content of the metallic oxides (Section 3.2.3), ensuring a favorable combination between thermal enhancement and rheological compatibility under high-temperature conditions. The incorporation range was limited to a maximum of 12 wt.%, a value commonly used in heterogeneous photocatalysis formulations, allowing the combination of photocatalytic functionality with improved mechanical performance of the binder.

3.2.1. Thermal Conductivity

The thermal conductivity test aimed to verify whether the heat conduction capacity of the nanomaterials was effectively transferred to the polymeric asphalt binder after nanomodification. This property is fundamental for thermal healing, as higher conductivity favors heat redistribution within the matrix, increasing the local fluidity of the binder and facilitating the diffusion of bituminous molecules in damaged zones, resulting in greater healing efficiency under elevated temperatures [69].
Thermal conductivity was measured using a C-Therm Thermal Conductivity Analyzer (Fredericton, NB, Canada). Cylindrical samples (25 mm diameter, 1 mm thickness) were prepared for each binder matrix (N0, N2, N4, N6, N8, N10, and N12) in the short-term aged condition (RTFOT). Six sequential measurements were performed for each sample at a controlled temperature of 25 ± 0.5 °C to ensure reproducibility. As shown in Figure 5, a thin layer of thermal grease (Thermal Joint CompoundType 120, Wakefield Thermal, Nashua, NH, USA) was applied to the sample surface to improve thermal contact with the sensor. The sample was then positioned on the sensor and loaded with a standard metallic weight to ensure proper seating and minimize measurement interferences. After preparation, sequential readings were recorded, providing a detailed profile of the thermal.

3.2.2. Rheological Behavior at High Temperatures

The high-temperature rheological behavior of the asphalt matrices was evaluated using three tests: apparent viscosity, high-temperature performance grade (PGH), and multiple stress creep and recovery (MSCR). The procedures and parameters adopted for each test are described in the following subsections.
Apparent Viscosity
The apparent viscosity of the asphalt matrices was determined to evaluate their resistance to flow under different temperature and shear conditions, simulating real processing scenarios of the asphalt binder such as handling, mixing, and field application. This rheological parameter is essential to ensure the proper performance of the material during paving operations. The test followed ASTM D4402M [73] using spindle 21 and a Brookfield rotational viscometer, model RVDV-I+ (Middleboro, MA, USA). Seven formulations (N0, N2, N4, N6, N8, N10, and N12) were tested in the unaged condition. The apparent viscosity was measured at 135 °C, 150 °C, and 177 °C, temperatures commonly used to characterize binder flow behavior in asphalt industry. During testing, the rotational speed was adjusted to maintain torque within 10–98% of the instrument capacity, ensuring compliance with the standard validity criteria [73].
High-Temperature Performance Grade
The high-temperature performance grade (PGH) was determined to classify both the reference polymeric asphalt binder and the nanomodified composites with respect to their resistance to permanent deformation under elevated temperatures. The test followed ASTM D7175 [74], which specifies the procedure for evaluating the viscoelastic behavior of asphalt binders using oscillatory rheometry. A dynamic shear rheometer (DSR), model Discovery HR-2 from TA Instruments (New Castle, DE, USA), was used for the analyses. Formulations N0, N2, N4, N6, N8, N10, and N12 were evaluated in both unaged and short-term aged conditions (RTFOT). For each formulation, two specimens with a diameter of 25 mm and a 1 mm gap were tested. The PGH was determined by identifying the highest temperature at which the binder satisfies the |G*|/sin δ criterion defined in the standard. Additionally, the aging index was calculated as the ratio between |G*|/sin δ values of aged and unaged samples (Equation (1)), enabling assessment of the materials’ susceptibility to thermally induced oxidative hardening.
AI = | G * | / sin   δ after- RTFOT   | G * | / sin   δ before- RTFOT  
where
AI: aging index (dimensionless);
|G*|/sin δafter-RTFOT: |G*|/sin δ parameter after short-term aging (kPa);
|G*|/sin δbefore-RTFOT: |G*|/sin δ parameter before short-term aging (kPa).
Multiple Stress Creep and Recovery
The multiple stress creep and recovery (MSCR) test was performed to evaluate the susceptibility of asphalt binders to permanent deformation under repeated loading at high temperatures. This test allows analysis of the elastic recovery capacity and resistance to creep of the binder. The procedure followed ASTM D7405 [75], using a dynamic shear rheometer (DSR), model Discovery HR-2 from TA Instruments (New Castle, DE, USA). The reference binder (N0) and the nanocomposites N2, N4, N6, N8, N10, and N12, all previously subjected to short-term aging (RTFOT), were tested in duplicate using a 25 mm plate and a 1 mm gap. The tests were performed at 76 °C and 82 °C, temperatures selected based on the high-temperature performance grades (PGH) obtained in the previous section. These temperatures represent the upper performance limits for the base binder (PGH 76-XX) and for the nanomodified binders (PGH 82-XX), respectively.

3.2.3. Definition of the Optimized Content of ZnO + TiO2

To investigate the contribution of nanoparticles to the healing capacity and mechanical performance of the asphalt mixture, the incorporation content of ZnO and TiO2 nanoparticles in equal proportions (50/50 wt.%) within the polymer-modified asphalt binder was optimized. The determination of the optimum content was based on three main criteria: (i) enhancement of thermal conductivity, since this parameter directly influences the healing phenomenon in asphaltic materials by facilitating heat diffusion and binder mobilization in damaged regions; (ii) improvement of rheological parameters at high temperatures, aiming to maintain high performance of the mixture under these critical conditions where the healing process tends to be more active due to greater molecular mobility; and (iii) limitation of the apparent viscosity to a maximum of 3.0 Pa·s at 135 °C, as established by the SUPERPAVE methodology guidelines, ensuring adequate binder workability. This approach aimed to develop a mixture that not only exhibits strong mechanical performance under elevated temperatures but also leverages these conditions to enhance the healing process. Such a strategy seeks to prevent failures such as permanent deformation precisely under conditions where healing could be most effective. After determining the optimized nanoparticle content, complementary analyses were performed for this formulation, including the evaluation of phase angle (δ) and dynamic shear modulus (|G*|) behavior at intermediate temperatures, as well as the Linear Amplitude Sweep (LAS) test to characterize the fatigue resistance of the nanocomposite.
Phase Angle and Dynamic Shear Modulus Behavior at Intermediate Temperatures
The rheological behavior at intermediate temperatures was evaluated to characterize the viscoelastic response of the asphalt matrices through the phase angle (δ), dynamic shear modulus (|G*|), and their respective elastic (G′) and viscous (G″) components. These parameters provide detailed information on the elastic and viscous behaviors of asphalt binders, particularly within typical service temperature ranges. The tests followed AASHTO T 315 [76] using a Discovery HR-2 hybrid dynamic shear rheometer from TA Instruments (New Castle, DE, USA). One sample of the reference binder (N0) and one sample containing the optimal nanoparticle content (Noptimal) were tested in the short-term aged condition (RTFOT), using an 8 mm diameter and 2 mm gap geometry. The experimental protocol followed the same methodology applied in previous studies [77,78,79]. The analysis was conducted through a frequency sweep ranging from 0.1 Hz to 30 Hz to capture the material response at different loading scales. The test temperatures were set at 5 °C and 10 °C, with a strain amplitude of 0.1% (suitable for maintaining the linear viscoelastic regime under stiffer conditions), and at 15 °C, 20 °C, 25 °C, 30 °C, and 35 °C, with a strain amplitude of 1% (appropriate for temperatures where the binder exhibits lower stiffness).
Linear Amplitude Sweep
The linear amplitude sweep (LAS) test was conducted to evaluate the fatigue performance of the nanocomposite with the optimal nanoparticle content (Noptimal) by analyzing the viscoelastic response of the material under progressively increasing strain levels. This test allows the estimation of the service life of the asphalt binder under cyclic loading conditions. The procedure followed AASHTO T 391 [80], applying a linear strain ramp using a Discovery HR-2 hybrid dynamic shear rheometer from TA Instruments (New Castle, DE, USA). Six samples of each material (reference binder and optimized nanocomposite), all in the short-term aged condition (RTFOT), were tested using an 8 mm diameter and 2 mm gap geometry. The test was conducted at 20 °C, adopting a 35% reduction in |G*|sin δ as the failure criterion. The experimental data were processed using the “AASHTO T 391-20—Version 1.59” [80] spreadsheet from the Modified Asphalt Research Center (MARC), University of Wisconsin–Madison (USA), to determine the rheological parameters used to construct fatigue curves (Equation (2) [80]) and to calculate the asphalt binder fatigue factor (FF) (Equation (3) [81]).
Nf = A   ×   ( γ ) B
where
A and B: fatigue curve coefficients;
Nf: number of cycles until rupture;
γ : applied shear strain, expressed as a percentage (%).
FF = log   N f 1.25 × N f 2.5 2 ×   log 0.0250 0.0125
where
FF: asphalt binder fatigue factor;
log: logarithm to base 10;
Nf1.25: number of cycles until rupture for a shear strain of 1.25%;
Nf2.5: number of cycles until rupture for a shear strain of 2.5%.

3.3. Assessment of the Mechanical Performance of Asphalt Mixtures

The mechanical performance of the asphalt mixtures was evaluated through a series of laboratory tests, including resistance to permanent deformation, rheological behavior at intermediate temperatures (phase angle and dynamic modulus), fatigue resistance, healing capacity, and the assessment of internal heating rate under microwave radiation. The reference asphalt mixture produced with the reference polymeric binder (N0) is denoted as M1, while the nanomodified asphalt mixture, produced using the optimized ZnO + TiO2 nanocomposite (Noptimal) determined in Section 3.2.3, is designated as M2.
The binder content adopted for both mixtures was the same as that established by Manfro et al. [36], since the constituent materials, including mineral aggregates, gradation curve, and the reference asphalt binder, were identical to those used in the present study. This ensured uniform production conditions, eliminating the influence of binder content on the comparative analysis of the mechanical performance of the formulations. Both mixtures were produced with 4.44% asphalt binder. Sample compaction was performed using a SUPERPAVE gyratory compactor, model Servopac, manufactured by IPC Global® (Boronia, Victoria, Australia), currently part of the Controls Group. The compaction pressure was maintained at 600 ± 18 kPa, with an external angle of 1.25 ± 0.02° and a rotation speed of 30 ± 0.5 rpm. The compaction protocol followed the parameters defined for high traffic volume (≥30 million ESALs), adopting the following criteria: Ninitial = 9, Ndesign = 125, and Nmax = 205. The volumetric properties of the compacted mixtures met the requirements specified for a nominal maximum aggregate size of 19 mm, as defined for mixtures intended for extremely high traffic (prior to SUPERPAVE 5). The gradation used in the formulations is presented in Table 6 [82]. All mixture design and preparation procedures strictly followed the guidelines established in AASHTO M 323 [83], AASHTO R 35 [84], and AASHTO R 30 [85].
In the subsequent stage, the asphalt mixtures were compacted using the French BBPAC compaction table (MLPC®), developed by the company VECTRA (currently NextRoad), based in Fontaine-lès-Dijon, Bourgogne-Franche-Comté, France. The compaction process followed the specifications outlined in EN 12697—Part 33 [86]. After compaction, the specimens intended for the rutting test measured 50 cm × 18 cm × 5 cm. The slabs designated for intermediate-temperature rheological, fatigue, healing, and internal heating-rate tests were initially molded with dimensions of 60 cm × 40 cm × 9 cm. These slabs were later cut into specimens with final dimensions of 38.1 cm × 6.35 cm × 5.08 cm, in accordance with the geometries required for the respective mechanical tests. The compaction, preparation, and cutting procedures of the test specimens are schematically illustrated in Figure 6 to visually represent the experimental stages involved.

3.3.1. Permanent Deformation Resistance

The rutting test was conducted to determine the resistance of the nanomodified asphalt mixture (M2) to wheel-tracking deformation. The performance related to this property was evaluated according to the procedure established in EN 12697—Part 22 [87], using the Orniéreur (MLPC®) equipment developed by the company VECTRA (currently NextRoad), based in Fontaine-lès-Dijon, Bourgogne-Franche-Comté, France. For each asphalt mixture, two specimens were tested under a controlled temperature of 60 °C, with a longitudinal load of 5000 ± 50 N applied at a frequency of 1 Hz. Subsequently, the rut depth percentage of mixtures M1 and M2 was calculated according to Equation (4) for the following load cycles: 100, 300, 1000, 3000, 10,000, and 30,000.
Pi = 100   × Σ j mij m 0 j 15   ×   ES
where
Pi: mean percentage of sinking on the plate surface in cycle i (%);
j: permanent deformation reading point (points P1 to P15 according to Figure 7);
mij: depth measurement in cycle i, point j;
m0j: depth measurement in cycle 0, point j;
ES: plate height.
Figure 7. Representation of the allocation of points P1 to P15 for measuring the depth of the specimen rutting (adapted of Melo [82]).
Figure 7. Representation of the allocation of points P1 to P15 for measuring the depth of the specimen rutting (adapted of Melo [82]).
Nanomaterials 15 01779 g007

3.3.2. Rheological Behavior at Intermediate Temperatures

The rheological behavior of the asphalt mixtures at intermediate temperatures was evaluated through the analysis of the phase angle (δ) and the dynamic modulus (|E*|), which are key parameters for characterizing the viscoelastic response of the material under cyclic loading. The experimental procedure followed the requirements established in EN 12697—Part 26 [88], which specifies the determination of the viscoelastic properties of asphalt mixtures using four-point bending beam tests. The tests were carried out using the Pneumatic 4 Point Bending Apparatus manufactured by IPC Global® (Boronia, Victoria, Australia), currently part of the Controls Group. Two specimens were tested: one corresponding to the reference asphalt mixture (M1) and the other to the nanomodified mixture (M2). The test temperatures adopted were 0 °C, 5 °C, 10 °C, 15 °C, 20 °C, 25 °C, and 30 °C, covering a representative range of field operating conditions. For each temperature, different loading frequencies were applied (0.1 Hz, 0.2 Hz, 0.5 Hz, 1 Hz, 2 Hz, 5 Hz, 10 Hz, 15 Hz, and 20 Hz) to generate a broad response spectrum of the viscoelastic behavior of the mixtures [88,89].

3.3.3. Fatigue Resistance and Healing Capacity

The evaluation of fatigue resistance and healing capacity of the asphalt mixtures was performed in three sequential stages: (1) execution of the first fatigue test; (2) application of the healing protocol, according to the methodology described by [37]; and (3) execution of a second fatigue test to assess the residual life of the asphalt mixture after the healing process. The fatigue tests were conducted in accordance with the procedures established in EN 12697—Part 24 [90], using the Pneumatic 4 Point Bending Apparatus by IPC Global® (Boronia, Victoria, Australia). A total of 43 specimens were tested, each with dimensions of 38.1 cm × 6.35 cm × 5.08 cm, including 20 specimens from the reference asphalt mixture (M1) and 23 specimens from the nanomodified mixture (M2). The test parameters adopted were: temperature of 20 °C, sinusoidal loading frequency of 10 Hz, controlled microstrain levels between 225 µm/m and 293 µm/m, and the failure criterion defined by the standard as a 50% reduction in the initial dynamic modulus, with this value measured at the 100th load cycle [90]. The characteristic fatigue curve model for each mixture was fitted using Equation (5), while the fatigue factor of the asphalt mixtures (FFM) was calculated using Equation (6), as proposed by [91].
Nf = a   ×   ε b
where
Nf: number of load applications until the initial dynamic modulus decreases by 50%;
ε : specific tensile strain (microstrain);
“a” and “b”: constants.
FFM = 0.2 × [ log ( Nf 100 ) + log ( Nf 250 ) ]
where
FFM: fatigue factor of the asphalt mixture;
log: base-10 logarithm;
Nf100: number of load applications for the specific microstrain of 100 µm/m;
Nf250: number of load applications for the specific microstrain of 250 µm/m.
After completing the first fatigue test, the same specimens were subjected to a healing protocol developed by Schuster [37], with the aim of investigating the healing capacity of the asphalt mixtures. The protocol consisted of the following stages: (1) microwave heating for 140 s immediately after the completion of the first fatigue test, promoting partial mobilization of the binder constituents and activation of physicochemical self-healing mechanisms; (2) thermal conditioning in the climatic chamber of the fatigue testing equipment after heating, where the specimens were kept at rest for 3 h under a controlled temperature of 20 °C; (3) after 2 h and 45 min of rest, repositioning of the specimen in the fatigue testing apparatus; and (4) after the 3 h rest period, resumption of the fatigue test (second test) using the same parameters as the initial test (temperature, frequency, microstrain, and loading mode) until the failure criterion was reached, that is, a 50% reduction in the initial dynamic modulus, measured at the 100th cycle of the first fatigue test. Finally, based on the obtained results, Equations (7) and (8) were used to calculate the healing percentage and the normalized healing, respectively.
% H =   Nf   2 nd   fatigue   test Nf   1 st   fatigue   test × 100  
where
%H: healing percentage;
Nf 1st fatigue test: number of cycles until reaching 50% of the initial dynamic modulus;
Nf 2nd fatigue test: number of cycles after the healing procedure, until the stopping criterion is reached.
NH = % H E D
where
NH: normalized healing [1/(J/m3)];
%H: healing percentage;
ED: energy density (J/m3).

3.3.4. Assessment of the Internal Heating Rate of Asphalt Mixtures Under Microwave Radiation

This stage aimed primarily to evaluate the internal heating rate of the reference asphalt mixture (M1) and the nanomodified mixture (M2) when subjected to microwave heating. The objective was to investigate whether the incorporation of metallic oxides contributed to increasing the thermal efficiency of the asphalt matrix, thereby enhancing its response to heat application.
For this purpose, specimens (38.1 cm × 6.35 cm × 5.08 cm) of each mixture were preconditioned for at least 24 h at a constant temperature of 20 °C to ensure uniform initial thermal conditions. The specimens were then heated in a microwave oven (Electrolux, model ME28S, nominal power of 900 W, frequency of 2450 MHz, and internal volume of 28 L) (Manaus, Amazonas, Brazil) for six different exposure times: 30 s, 60 s, 90 s, 120 s, 150 s, and 183 s (three specimens for each heating duration). Immediately after heating, each specimen was longitudinally sectioned, and the internal temperature was measured using an infrared thermographic camera (FLIR® model B400, Wilsonville, OR, USA). From the collected data, mean internal temperature curves (Equation (9)) were plotted as a function of heating time, allowing for a comparative evaluation of the thermal performance of both formulations. Based on the results, it was possible to determine whether the presence of metallic oxides, due to their electromagnetic properties and thermal conductivity, enhanced the absorption and conversion of microwave energy into heat, promoting faster and deeper heating within the asphalt matrix. This characteristic is particularly relevant, as it helps explain potential improvements in healing capacity discussed in the previous section. A higher internal heating rate facilitates the mobilization of the asphalt binder and accelerates the processes of flow and recombination of molecular bonds, key factors for the healing of microcracks and the partial restoration of the functional properties of the asphalt matrix. Figure 8 illustrates the heating, sectioning, and thermal imaging stages.
MIT =     H + L 2
where
MIT: mean internal temperature (°C);
H: highest test specimen temperature (°C);
L: lowest test specimen temperature (°C).
Figure 8. Heating and temperature measurement procedure (A) test specimens with dimensions of 38.1 cm × 6.35 cm × 5.08 cm before the heating and breaking process (B) bipartite test specimen for measuring the internal temperature (C) thermographic recording with the FLIR® brand camera (Teledyne FLIR LLC, Wilsonville, OR, USA) and (D) thermographic image of the bipartite test specimen.
Figure 8. Heating and temperature measurement procedure (A) test specimens with dimensions of 38.1 cm × 6.35 cm × 5.08 cm before the heating and breaking process (B) bipartite test specimen for measuring the internal temperature (C) thermographic recording with the FLIR® brand camera (Teledyne FLIR LLC, Wilsonville, OR, USA) and (D) thermographic image of the bipartite test specimen.
Nanomaterials 15 01779 g008

4. Results and Discussion

This section presents and discusses the experimental results, beginning with the nanomodification of the polymeric asphalt binder, followed by the analysis of thermal conductivity and rheological behavior to determine the optimum nanoparticle content, and concluding with the evaluation of the mechanical performance of the reference and nanomodified asphalt mixtures.

4.1. Nanomodification of the Polymeric Asphalt Binder

The nanomodification process of the base polymer-modified asphalt binder (N0) resulted in the formulation of six nanocomposites with different incorporation levels of ZnO and TiO2, in a 50/50 wt.% ratio, designated as N2, N4, N6, N8, N10, and N12. After preparation, the nanocomposites were first examined using bright-field microscopy to assess the dispersion of the nanomaterials within the asphalt matrix. Subsequently, short-term aging (RTFOT) was performed to simulate the initial degradation conditions that occur during asphalt mixture production [71]. The nanocomposites were then characterized with respect to their internal structure, including the evaluation of structural and crystalline features by X-ray diffraction (XRD) and the identification of functional groups using Fourier-transform infrared spectroscopy (FTIR). These techniques provided a comprehensive characterization of the materials, enabling a deeper understanding of the nanomaterial incorporation process into the polymeric asphalt binder and offering essential information for interpreting the mechanical and rheological behavior of the matrices, discussed in the subsequent sections of this study.

4.1.1. Bright-Field Microscopy

The micrographs shown in Figure 9A–G indicate a uniform distribution of the nanomaterials within the matrix, even at higher concentrations, with no evidence of significant agglomerate formation. This behavior is particularly relevant, as proper dispersion of the nanomaterials can directly contribute to improved structural uniformity of the matrix and, consequently, to enhanced mechanical performance.

4.1.2. Short-Term Aging

The mass loss after short-term aging (RTFOT) is presented in Figure 10. It can be observed that all nanocomposites, regardless of the ZnO + TiO2 incorporation level, exhibited mass loss values below 1%, thus meeting the maximum limit established by ASTM D6373 [72]. This result indicates that nanoparticle modification did not compromise the stability of the asphalt binders during the initial aging process. The observed mass loss is associated with the volatilization of light compounds present in the asphalt binder, which tend to evaporate under the thermal conditions of the RTFOT [71]. However, variations among the different nanocomposites were minor and did not show any systematic trend related to the ZnO + TiO2 content. Such variations may be attributed to the inherent uncertainty of the testing method, which involves processes such as evaporation and air flow, both susceptible to slight operational fluctuations. Maintaining the mass loss within acceptable limits in all cases suggests that the presence of ZnO and TiO2 did not intensify volatile emissions.

4.1.3. X-Ray Diffraction

Figure 11 presents the X-ray diffraction (XRD) results obtained for the nanomaterials (ZnO and TiO2), the reference asphalt binder (N0), and the nanocomposites N2, N4, N6, N8, N10, and N12, all under short-term aging conditions (RTFOT). The obtained data allow for the characterization of the crystalline structure of the materials and the examination of the interaction between the oxides and the asphalt matrix after the incorporation process.
In Figure 11 the diffraction patterns of the TiO2 nanoparticles predominantly indicate the anatase phase (tetragonal structure), consistent with the JCPDS (Joint Committee on Powder Diffraction Standards) card 21-1272 [6,20,92,93]. For ZnO, the wurtzite phase (hexagonal lattice) is observed, in accordance with JCPDS 79-2205 [6,20,94,95]. The absence of additional reflections suggests high purity of the nanomaterials, with no evidence of impurities or secondary phases.
Regarding the reference polymeric asphalt binder (N0), the diffractogram exhibits a broad amorphous band at 2θ = 19° (γ band), associated with the presence of stacked aliphatic chains in the asphaltene structure [96,97], and a band at 2θ = 24° (002), related to the stacking of aromatic rings, also characteristic of asphalt binders [97]. A distinct peak at 2θ = 21° is also observed, attributable to paraffin wax [98], along with a band around 2θ = 43° (100), associated with the size of sheets formed by fused aromatic rings [96].
With the incorporation of ZnO and TiO2 nanomaterials into the polymeric asphalt binder (N0), a systematic reduction in the relative intensity of the characteristic asphalt binder bands (γ, 002, and 100) is observed as the incorporation level increases. This reduction is consistent with the dilution and attenuation effects of the diffracted signal of the binder caused by the presence of the added inorganic crystalline phases, without measurable changes in the positions (2θ) or full width at half maximum (FWHM) of the main bands. Thus, there is no evidence of detectable structural modifications in the mean stacking of the aromatic and aliphatic constituents of the binder.
In parallel, crystalline peaks attributed to the incorporated oxide phases emerge and become more pronounced. For ZnO, the most prominent regions correspond to hkl (100) and (101), consistent with the isolated material; for TiO2, the highest intensity occurs at (101), typical of the anatase crystalline phase, which is also dominant in the pure oxide sample. The clear and increasing presence of these crystalline phases in the nanocomposites, without significant overlapping or shifting of diffraction peaks, suggests a good physical distribution of the nanomaterials within the asphalt matrix, with no evidence of chemical reactions between the constituents. This indicates that the incorporation process preserves the structural integrity of the components, characterizing a predominantly physical interaction consistent with the objective of nanometric reinforcement. Such structural compatibility is essential to ensure the stability and functionality of the nanocomposites.
Finally, the observed reduction in the amorphous band (associated with the reference asphalt binder) may be correlated with improved dispersion of the SBS polymer. This behavior supports the hypothesis that the nanomodification process not only promotes the incorporation of nanomaterials but also contributes to a more uniform internal microstructure.

4.1.4. Fourier-Transform Infrared Spectroscopy

Figure 12 presents the Fourier-transform infrared (FTIR) spectra obtained for the nanomaterials (ZnO and TiO2), the reference asphalt binder (N0), and the developed nanocomposites (N2, N4, N6, N8, N10, and N12), all analyzed under short-term aging conditions (RTFOT). These results enable the comparison of the spectral signatures of the oxides and the polymeric matrix, as well as the evaluation of the effects of nanomaterial incorporation on the chemical structure of the asphalt binder.
To contextualize the isolated nanomaterials, Figure 12 shows that ZnO exhibits a peak at 439.68 cm−1, attributed to the stretching vibration of the Zn–O bond, characteristic of the wurtzite structure [99,100,101], as well as a peak at 1384.62 cm−1 associated with the symmetric stretching vibration mode of the carboxylate group (COO), possibly related to residual organic traces from synthesis or stabilization [102]. A broad band is also observed at 3413.34 cm−1 (O–H stretching), typical of hydroxyl groups or adsorbed water molecules on the surface of high–surface-area materials [99,101,102,103]. In turn, TiO2 exhibits a peak at 524.54 cm−1, corresponding to the Ti–O–Ti bridge stretching vibrations, consistent with the anatase/rutile phases [104,105,106], as well as a band at 1627.61 cm−1, associated with angular deformation (bending) and stretching vibrations of hydroxyl groups (–OH), commonly attributed to the presence of chemisorbed water on the surface of materials with high specific surface area [105]. In addition, the band observed at 3390.20 cm−1 corresponds to symmetric and asymmetric stretching vibrations of the –OH group, indicating the presence of hydroxyl groups strongly bonded to the TiO2 surface [104,106].
For the reference polymeric asphalt binder (N0) (Figure 12), the peaks located at 2916.01 cm−1 and 2850.44 cm−1 are attributed in the literature to the C–H stretching vibrations of aliphatic chains [107,108]. Furthermore, the peaks at 1458.01 cm−1 and 1373.15 cm−1 correspond to vibrations associated with the –CH2– and –CH3– groups, respectively [107,108,109,110]. The lower–intensity peaks observed at 968.15 cm−1 and 698.15 cm−1 are attributed to the butadiene chain in the SBS copolymer [111] and to C–H vibrations in monoalkylated aromatics, characteristic of polystyrene [109], confirming the presence of the copolymer in the matrix.
In the modified nanocomposites (Figure 12), a gradual reduction in transmittance is observed near 447.43 cm−1, proportional to the increase in incorporated ZnO + TiO2 content. This behavior is related to the characteristic bands of the nanomaterials at 439.69 cm−1 for ZnO and 524.54 cm−1 for TiO2, and indicates the presence of these metallic oxides in the matrix. It is important to note that the incorporation of ZnO and TiO2 did not cause significant changes in the main peaks of the polymeric asphalt binder matrix, indicating that no relevant chemical reactions occurred between the components. This chemical stability suggests a predominantly physical interaction. Similar results were reported by Neto et al. [112] and Filho et al. [108] when evaluating asphalt composites modified with metallic oxides.
Finally, the preservation of the characteristic peaks of the SBS copolymer, even after the addition of nanomaterials, reinforces the hypothesis that the nanomodification process does not compromise the integrity of the polymeric modifier. The XRD and FTIR results confirm that the incorporation of ZnO and TiO2 preserved the structural and chemical integrity of the system, which is a fundamental condition to ensure thermal and rheological consistency in the subsequent evaluation stages. Therefore, the analysis now turns to the functional performance of the nanomodified binders, focusing on properties directly associated with service behavior, such as thermal conductivity and rheological performance.

4.2. Analysis of Thermal Conductivity and Rheological Behavior to Define the Optimum Content

This section presents and discusses the results related to the thermal conductivity and high-temperature rheological behavior of the developed asphalt nanocomposites. The combined analysis of these parameters allowed the evaluation of the impact of nanomodification with metallic oxides on the performance of the asphalt binder, providing technical support for defining the optimized ZnO and TiO2 content to be used in a dense asphalt mixture.

4.2.1. Thermal Conductivity

The thermal conductivity of the nanocomposites was evaluated to verify the effective transfer of the thermal properties of the nanomaterials (ZnO and TiO2) to the polymeric asphalt binder. This parameter is highly relevant, as materials with higher thermal conductivity promote the healing process of the asphalt structure, making the localized heating required to induce binder flow and close microcracks more efficient. The improvement of thermal properties through the incorporation of various types of nanomaterials into asphalt binders has been reported in the literature [69,113]. The results obtained for the nanocomposites developed in this study are presented in Figure 13.
Figure 13 shows that as the nanomaterial content in the polymeric asphalt binder increases, the thermal conductivity also rises. For the highest incorporation level of ZnO + TiO2 analyzed (12%), the thermal conductivity reached 0.6 W/m·K, representing a 150% increase compared with the reference polymeric asphalt binder (0%).
A one-way Welch ANOVA analysis yielded p < 0.001, indicating a statistically significant difference among the ZnO + TiO2 incorporation levels with respect to thermal conductivity. The Games–Howell post hoc test (Table 7) reinforces this evidence, with most comparisons showing p < 0.001 (statistically significant differences), except for the comparison between N4 and N6, which yielded p = 0.003 (still within the threshold for statistical significance). Furthermore, the trend observed in the mean difference values suggests that increasing the concentration of these nanomaterials contributes to enhancing the thermal conductivity of the polymeric asphalt binder.
Finally, this improvement in thermal conductivity is relevant because it promotes more efficient heat redistribution within the asphalt matrix at elevated temperatures. This behavior may facilitate binder molecule mobility and favor the closure of microcracks, a mechanism associated with the healing potential of asphalt materials.

4.2.2. Rheological Behavior at High Temperatures

The following section presents and discusses the results of the rheological behavior at high temperatures, considering apparent viscosity, performance grade (PGH), and multiple stress creep and recovery (MSCR) tests.
Apparent Viscosity
The apparent viscosity was determined for both the polymeric asphalt binder (N0) and the nanocomposites (N2, N4, N6, N8, N10, and N12) at temperatures of 135 °C, 150 °C, and 177 °C (Figure 14).
At all test temperatures, a progressive increase in apparent viscosity was observed with the addition of ZnO + TiO2 nanoparticles to the polymeric asphalt binder. The analysis of the trend curves (Figure 14) shows that the highest percentage increase in apparent viscosity between the N0 composite (without nanomaterials) and N12 (with 12% ZnO + TiO2) occurs at 135 °C, with a 52.8% increase. At 150 °C and 177 °C, the observed increases were 47.7% and 43.5%, respectively. These results indicate that the effect of the nanomaterials on viscosity is more pronounced at lower temperatures, which may significantly impact the workability of the binder during production and application of the asphalt mixture.
The trend lines (Figure 14) show that at 135 °C, starting from a ZnO + TiO2 concentration of approximately 10.4%, the viscosity exceeds the 3 Pa·s limit established by the SUPERPAVE methodology [114] to ensure adequate pumpability and handling of the binder. Binders with viscosity values above this threshold may present operational challenges and require higher energy for pumping.
Therefore, the results demonstrate that the incorporation of ZnO + TiO2 nanomaterials directly affects the apparent viscosity of the polymeric asphalt binder. This behavior is consistent with previous studies that investigated the individual addition of ZnO and TiO2 to different asphalt matrices [14,15,62,70], which also reported viscosity increases as a result of nanometric modification.
From a functional performance perspective, this increase in viscosity can be beneficial under high-temperature service conditions, contributing to greater resistance to permanent deformation, as demonstrated by Neto et al. [14] and Cadorin et al. [62]. However, it is essential to define the nanomaterial dosage in a manner that ensures improved mechanical performance while preserving appropriate processing and application conditions.
High-Temperature Performance Grade
Figure 15 shows the influence of ZnO + TiO2 incorporation into the polymeric asphalt binder on the high-temperature performance grade (PGH), evaluated in both the unaged and short-term aged (RTFOT) conditions.
The incorporation of ZnO + TiO2 nanoparticles into the polymeric asphalt binder promoted a significant increase in high-temperature resistance, expressed by the rise in continuous grade. The observed increment rates were 0.30 °C/(%ZnO + TiO2) for the unaged binder and 0.23 °C/(%ZnO + TiO2) after short-term aging (RTFOT), as evidenced by the trend lines in Figure 15. This modification directly resulted in an improvement in the performance classification of the binder.
In the unaged condition (Figure 15), the transition of the high-temperature performance grade (PGH) from 76-XX to 82-XX occurred at a minimum concentration of 0.9% ZnO + TiO2. After RTFOT aging, this change was observed from 8.5% nanoparticle incorporation. These results indicate that the addition of nanomaterials contributes to increasing the critical failure temperature of the polymeric asphalt binder, enhancing its performance against permanent deformation under hot climates and/or heavy traffic loads. Similar findings were reported by Neto et al. [14] for ZnO and by Babagoli et al. [7] and Cadorin et al. [62] for TiO2, corroborating the effectiveness of these nanomaterials in improving the thermal performance of asphalt binders.
Based on the PGH values obtained, the aging index (AI) was also calculated to evaluate the susceptibility of the binder to aging. The index was determined specifically for samples N0 (reference), N6, and N12, considering the sensitivity of the test. Figure 16 presents the evolution of the AI as a function of temperature and ZnO + TiO2 concentration.
As shown in Figure 16, the progressive incorporation of ZnO and TiO2 nanoparticles into the polymeric asphalt binder promotes a systematic reduction in the aging index (AI), indicating lower susceptibility to short-term oxidative aging. This trend suggests that the nanomaterials act as stabilizing agents, enhancing binder durability by mitigating the detrimental effects of thermal aging.
The decrease in AI reflects a smaller variation in the critical performance temperature before and after aging (RTFOT), demonstrating an asphalt matrix less sensitive to thermo-oxidative transformations. This behavior was also observed by Neto et al. [14], who attributed the presence of ZnO to an antioxidant effect, reducing the formation of oxidative species and thereby limiting the degradation of binder constituents. Complementarily, Cadorin et al. [62] highlighted the role of TiO2 as a nanofiller that decreases the porosity of the asphalt matrix, increasing its impermeability and limiting both the volatilization of light components and the diffusion of oxygen, factors directly associated with the aging process. It is important to note that the reduction in AI was more pronounced up to approximately 6% ZnO + TiO2 incorporation, after which a stabilization and/or slight increase trend was observed.
Multiple Stress Creep and Recovery
The susceptibility of the asphalt matrices to permanent deformation was evaluated using rheological parameters obtained from the multiple stress creep and recovery (MSCR) test, with emphasis on the percentage of elastic recovery (%R3.2) and the non-recoverable creep compliance (Jnr3.2). These parameters were determined under a stress level of 3.2 kPa, applied at temperatures of 76 °C and 82 °C on samples aged using the RTFOT procedure. Figure 17 presents the %R3.2 values, while Figure 18 shows the corresponding Jnr3.2 results.
The percentage of elastic recovery (%R3.2), presented in Figure 17, showed absolute increases of 1.6 and 2.3 percentage points at 76 °C and 82 °C, respectively, when comparing the reference polymeric asphalt binder (N0) with the nanocomposite containing 12% ZnO + TiO2 (N12). Despite the observed increases, these values can be considered modest and do not exhibit a clearly defined trend as a function of the nanomaterial content. These results contrast with those reported by Günay and Ahmedzade [115], who observed an absolute increase of 1.7 percentage points in %R3.2 at 76 °C after incorporating 4% TiO2 at the nanoscale into an asphalt binder modified with 3% SBS. Similarly, Melo et al. [20] reported that the isolated addition of TiO2 and ZnO did not produce significant changes in the elastic response of a conventional asphalt binder (PGH 58-XX) under elevated temperatures.
Regarding the non-recoverable creep compliance (Jnr3.2), presented in Figure 18, a progressive reduction trend was observed with increasing ZnO + TiO2 nanoparticle content in the polymeric asphalt binder. This trend is evidenced by the regression curves, whose high coefficients of determination (R2) indicate a strong correlation between the increase in nanomaterial concentration and the reduction in binder deformability. Based on the trend line analysis, the addition of 12% ZnO + TiO2 (N12) resulted in decreases of 24.5% and 35.1% in Jnr3.2 at 76 °C and 82 °C, respectively, compared to the reference binder (N0). These results indicate a significant and systematic improvement in resistance to permanent deformation through the combined use of metallic oxide nanoparticles. Similar but isolated behaviors have been reported in the literature. Neto et al. [14] found that the addition of 7% ZnO to a conventional asphalt binder resulted in an 89.5% reduction in Jnr3.2 at 64 °C. Likewise, Günay and Ahmedzade [115] observed that adding 4% TiO2 to an SBS-modified binder reduced Jnr3.2 by 8.6% at 76 °C. The observed reduction in non-recoverable compliance can be attributed to the nanoscale reinforcement effect provided by the metallic oxides, which increases binder stiffness and decreases its susceptibility to creep under cyclic loading at high temperatures. This behavior reinforces the potential of nanomaterials as effective modifiers for designing asphalt binders with enhanced performance, especially for applications under heavy traffic and hot climate conditions.

4.2.3. Definition of the Optimized Content of ZnO + TiO2

To develop an asphalt mixture with optimized healing properties and superior rheological performance, an integrated and comparative analysis was conducted to determine the optimum incorporation content of nanomaterials (ZnO + TiO2) in the polymeric asphalt binder. The selection was based on thermal and mechanical performance criteria, including thermal conductivity (a key factor for the healing process), high-temperature performance grade (PGH), aging index, non-recoverable creep compliance, and apparent viscosity. As a result, the content of 8.5% ZnO + TiO2 was identified as the most suitable. At this concentration, the nanocomposite exhibited the following properties based on the trend line analysis:
  • Thermal conductivity of 0.5 W/m·K, representing a 106.3% increase compared with the reference polymeric asphalt binder;
  • Increase in PGH from 76-XX to 82-XX for both unaged and short-term aged (RTFOT) conditions;
  • Mean reduction of 7.2% in the aging index (AI), indicating higher resistance to oxidation;
  • Reduction in non-recoverable creep compliance (Jnr3.2): at 82 °C, the value decreased to 1.29 kPa−1, corresponding to a 24.9% reduction; at 76 °C, the value was 0.41 kPa−1, reflecting a 17.3% reduction;
  • Apparent viscosity of 2.870 Pa·s, which is 4.3% below the SUPERPAVE limit of 3.0 Pa·s (135 °C), ensuring that binder workability is maintained during mixing and application.
These results indicate that the 8.5% ZnO + TiO2 content yields the most favorable combination of thermal and mechanical enhancements, suggesting that this condition may be suitable for application in asphalt mixtures. The following sections present and discuss the effects of this optimized incorporation on the behavior of the polymeric asphalt binder at intermediate temperatures. The analysis includes the rheological parameters of dynamic shear modulus (|G*|) and phase angle (δ), as well as fatigue resistance evaluated through the LAS test.
Phase Angle and Dynamic Shear Modulus Behavior at Intermediate Temperatures
Table 8 presents the influence of the combined addition of 8.5% ZnO + TiO2 to the polymeric asphalt binder, compared with the reference binder (N0—0% ZnO + TiO2), under short-term aged conditions (RTFOT). The analysis is based on rheological parameters obtained from the dynamic shear test, including the dynamic shear modulus (|G*|), phase angle (δ), storage modulus (G′, elastic component), and loss modulus (G″, viscous component). These parameters provide a comprehensive understanding of the viscoelastic behavior of the binder at intermediate temperatures, allowing the evaluation of the effects of nanometric modification on the mechanical characteristics of the material.
According to the data presented in Table 8, the incorporation of 8.5% ZnO + TiO2 into the polymeric asphalt binder resulted in consistent rheological behavior across the entire temperature range analyzed, from 5 °C to 35 °C. Considering the mean loading frequencies (0.1 Hz to 30 Hz), a generalized increase was observed in the dynamic shear modulus (|G*|), with variations ranging from 100.5% to 108.1% relative to the reference binder (without nanoparticle addition). The storage (G′) and loss (G″) moduli also increased, ranging from 101.3% to 109.3% for G′ and from 99.4% to 105.3% for G″. Regarding the phase angle (δ), the results demonstrated stable viscoelastic behavior, with minimal variations between +0.1% and −2.1%. This slight reduction in δ indicates a minor shift toward greater elastic dominance, yet without producing a substantial change in the ratio between the elastic and viscous responses. In summary, the results indicate that modification with 8.5% ZnO + TiO2 significantly enhances binder stiffness at intermediate temperatures without meaningfully altering the nature of its rheological behavior.
Linear Amplitude Sweep
Table 9 presents the rheological parameters derived from the Linear Amplitude Sweep (LAS) test for the reference asphalt binder (N0) and the binder modified with 8.5% ZnO + TiO2 (Noptimal), both under short-term aged conditions (RTFOT). The fatigue life prediction equations, obtained based on viscoelastic models, were as follows: for the reference asphalt binder (N0), Nf = 10,350,246 γ−5·21, and for the optimized nanocomposite (Noptimal), Nf = 9,226,490 γ−5·24.
Based on the data presented in Table 9, the incorporation of 8.5% ZnO + TiO2 into the reference asphalt binder resulted in changes in the fatigue model parameters derived from the LAS test. Parameter A, which represents the initial fatigue resistance under small deformations, decreased by 10.9% in the nanomodified binder (Noptimal), suggesting a lower ability to delay the initiation of accumulated damage. Parameter B, associated with the sensitivity of the material to strain amplitude, increased by 0.6%, indicating that the nanomodified binder exhibits a greater dependence of fatigue life on the applied strain level. This behavior is corroborated by the number of cycles to failure (Nf), which was consistently lower for the binder containing nanomaterials across all tested strain amplitudes. The reductions ranged from 11.5% (for γ = 1.25%) to 17.8% (for γ = 15%), reflecting a decrease in durability under cyclic loading. Consequently, the fatigue factor (FF), used to estimate the overall binder performance, decreased by 1.0%, suggesting a negative impact of nanomaterial incorporation on damage tolerance. The reduction in fatigue life can be explained by the changes in viscoelastic properties previously identified at 20 °C, where a significant increase in the dynamic shear modulus (+102.9%) was observed, without a meaningful change in phase angle (−1.2%), indicating a stiffer behavior. Although this additional stiffness is beneficial for resisting deformation, it may compromise the ability of the binder to accommodate repeated stresses, particularly under fatigue conditions.
Despite the observed trends in the LAS test, it is important to note that this method is not yet widely consolidated within the scientific community as a definitive tool for predicting the contribution of the binder to the fatigue performance of asphalt mixtures, especially in systems modified with nanomaterials. The LAS test provides indirect estimates of fatigue resistance based solely on binder properties, disregarding essential aspects such as binder–aggregate interactions, gradation structure, composite matrix effects, and scale phenomena. These limitations make the results particularly sensitive to chemical and rheological modifications, as in the case of nanomodified binders. Therefore, the findings should be interpreted with caution and should not be used in isolation to predict the fatigue performance of asphalt mixtures.

4.3. Assessment of the Mechanical Performance of Asphalt Mixtures

With the optimized nanocomposite content established, the next step involved evaluating its behavior when applied in asphalt mixtures. In accordance with standard practice in binder-modification studies, only the optimum dosage was incorporated into the mixture, reflecting the condition previously identified as providing the most favorable combination of thermal and mechanical performance. Comprehensive experimental tests were conducted to investigate the effects of nanometric modification on the mechanical performance of the mixture. The procedures included: the permanent deformation test to assess resistance to permanent deformation; the characterization of viscoelastic behavior at intermediate temperatures; the fatigue test to estimate durability under repeated loading; the healing protocol to evaluate self-healing ability; and the assessment of the internal heating rate of the asphalt mixtures under microwave radiation

4.3.1. Permanent Deformation Resistance

The rutting performance was analyzed using the French Orniéreur traffic simulator, testing two specimens of the reference asphalt mixture (M1—0% ZnO + TiO2) and two specimens of the nanomodified asphalt mixture (M2—8.5% ZnO + TiO2). In Figure 19, the values represent the mean results of the tests, while the trend line and characteristic equation were derived from these data.
The results presented in Figure 19 illustrate the accumulated deformation (%) as a function of the number of loading cycles for two asphalt mixture formulations: the reference mixture (M1, with 0% ZnO + TiO2) and the nanomodified mixture (M2, with 8.5% ZnO + TiO2). Data analysis shows that, during the first four loading intervals (100, 300, 1000, and 3000 cycles), the reference mixture exhibited lower rutting levels. However, after 10,000 cycles, this trend reversed, and the mixture containing nanomaterials began to display lower accumulated deformation values. At the end of 30,000 cycles, the nanomodified mixture (M2) exhibited a rut depth of 3.9%, which was lower than the 4.5% observed for the conventional mixture (M1). This inversion over time indicates that, while the reference mixture shows better initial resistance to deformation, the nanomodified mixture presents a lower rate of permanent deformation progression.
The higher coefficient of M2 (0.83) explains its slightly greater initial deformation; however, its smaller exponent (0.15 < 0.20) indicates a lower rate of deformation growth with cycles, which explains why M2 outperforms M1 in the long term. Rheologically, although M2 exhibits higher |G*| and lower δ (greater stiffness and elasticity), as previously reported in this study, its stiffer internal network accumulates stress during the first loading cycles. As the load repetitions progress, this structure restricts creep and rut development, resulting in lower total deformation at the end of the test.
The superior performance of mixture M2 can be attributed to the improved rheological properties of the nanomodified binder, as previously demonstrated in this study. Notably, the increased apparent viscosity, higher |G*|/sin δ parameter values, and reduced non-recoverable deformation observed in MSCR tests all indicate greater binder stiffness and creep resistance, factors that directly contribute to mitigating plastic deformation in asphalt mixtures under repeated loading.
These findings are consistent with the literature. Kamboozia et al. [24], when analyzing porous mixtures with different ZnO contents, observed a progressive reduction in permanent deformation with increasing nanoparticle concentration. Similarly, Sadeghnejad and Shafabakhsh [21] reported significant improvements in rutting resistance in SMA-type mixtures modified with nano-TiO2, attributing this behavior to enhanced interaction between asphalt particles and nanomaterials, which increases internal cohesion and resistance to deformation.
Therefore, the results indicate that the combined addition of 8.5% ZnO + TiO2 to the polymeric binder can improve the permanent deformation resistance of the mixture, particularly under heavy traffic and high-temperature conditions, underscoring its potential for high-performance pavement applications.

4.3.2. Rheological Behavior at Intermediate Temperatures

Table 10 presents the results obtained for the phase angle and dynamic modulus at intermediate temperatures (30 °C to 0 °C) for the reference asphalt mixture (M1—0% ZnO + TiO2) and the nanomodified asphalt mixture (M2—8.5% ZnO + TiO2). The table reports the mean values, considering all loading frequencies (0.1 Hz to 20 Hz), for the dynamic modulus (|E*|), phase angle (δ), and the elastic (E1) and viscous (E2) components.
The data presented in Table 10 demonstrate that the incorporation of 8.5% ZnO + TiO2 into the polymeric binder produced a consistent increase in the dynamic modulus (|E*|) of the asphalt mixture throughout the entire temperature range tested (0 °C to 30 °C). The effect was most pronounced at higher temperatures, particularly at 30 °C, where the greatest increase in |E*| (+31.0%) was recorded compared to the reference mixture. The magnitude of this improvement gradually decreased with lower temperatures, showing increments of +21.1% (25 °C), +12.9% (20 °C), +9.5% (15 °C), +5.8% (10 °C), +3.8% (5 °C), and +4.5% (0 °C). This increase in |E*| reflects greater structural stiffness in the nanomodified mixture, attributed to the reinforcing effect of the metallic oxides on the matrix. Such behavior is particularly beneficial for reducing accumulated deformation under repeated loading, as previously observed in the permanent deformation test. Furthermore, these results are consistent with the rheological data of the modified binders obtained from the PGH and MSCR tests, in which the binder containing 8.5% ZnO + TiO2 also exhibited superior performance.
The phase angle (δ) (Table 10) showed reductions across the entire temperature range analyzed, varying between −4.6% (30 °C) and −10.3% (10 °C), indicating a shift toward more elastic behavior in the mixture. Reductions of −8.8% (5 °C) and −9.0% (0 °C) further confirm this trend at lower temperatures. The decrease in δ is associated with the predominance of the storage modulus (E1) over the loss modulus (E2), indicating increased resistance to viscoelastic deformation.
Overall, the results show that modification with metallic oxides promotes a rheological reinforcement, characterized by increased stiffness and greater elastic dominance, features that contribute to superior asphalt mixture performance against permanent deformation, fatigue, and thermal variations.

4.3.3. Fatigue Resistance and Healing Capacity

The fatigue performance of the asphalt mixtures was evaluated using the four-point bending beam test. The fatigue test was initially conducted until failure for both mixtures (reference—M1 and nanomodified—M2). Subsequently, the specimens used in the first fatigue test were immediately subjected to the healing protocol. Afterward, a new fatigue test was performed to assess the residual life of the samples following the healing procedure. The results obtained from the first fatigue test are presented in Figure 20, Table 11 and Table 12.
As shown in Figure 20 and Table 11, the reference asphalt mixture exhibited superior performance compared with the nanomodified mixture in terms of fatigue resistance. For the same strain level, the reference mixture withstood a higher number of cycles to failure (Nf), indicating a greater ability to resist fatigue damage in laboratory conditions. This result suggests that the modification with ZnO and TiO2 nanoparticles, at a concentration of 8.5%, reduced the fatigue resistance of the asphalt composite. The analysis of parameter b in the fatigue life equations, presented in Figure 20, revealed that the nanomodified mixture exhibited 6.7% lower strain susceptibility (b = −6.29) compared with the reference mixture (b = −6.74).
Table 11 presents the mean values of the initial dynamic modulus, measured at the 100th loading cycle. The nanomodified mixture (M2) showed a mean modulus of 7328 MPa (SD = 526 MPa), representing a 4.1% increase compared with the reference mixture (M1), which had a mean modulus of 7043 MPa (SD = 591 MPa). This increase in stiffness was corroborated by complementary analyses, such as the dynamic shear modulus of the nanomodified binder (Noptimal) and the dynamic modulus of the M2 mixture itself, both evaluated at 20 °C. However, the higher stiffness observed may be associated with the reduced fatigue resistance, since stiffer materials tend to have lower energy dissipation capacity and, consequently, greater susceptibility to cracking. This trend was also confirmed by the LAS test performed at 20 °C on the nanomodified binder, which demonstrated lower fatigue resistance.
On the other hand, the nanomodified asphalt mixture (M2), due to its higher stiffness, tends to experience lower levels of strain under loading in service conditions when compared to the reference mixture (M1). In laboratory tests conducted under controlled-strain loading, stiffer materials generally reach failure in fewer cycles. However, this response does not necessarily reflect their behavior in the field, where higher stiffness tends to limit the strains experienced by the material. Therefore, the lower number of cycles required to reach the same controlled strain does not, by itself, indicate reduced fatigue resistance, but rather reflects the change in the mechanical response of the mixture resulting from the increased stiffness.
Additionally, the fatigue factor of the mixture (FFM), presented in Table 12, indicates that the nanomodified mixture exhibited an FFM 4.4% lower than that of the reference asphalt mixture. This result reinforces that, at the concentration adopted, the incorporation of nanomaterials did not provide improvements in fatigue resistance at the same deformation amplitude. The results obtained contrast with those reported by Sadeghnejad and Shafabakhsh [21], who observed an increase in fatigue life with the incorporation of up to 1.2% TiO2 in Stone Mastic Asphalt (SMA) mixtures evaluated under indirect tensile tests. Similarly, Mousavi Rad et al. [23] observed improved fatigue performance with up to 8% ZnO in porous mixtures using four-point bending tests. This divergence can be attributed to three main factors: (i) differences in the binders used (conventional in previous studies versus SBS-modified in this study); (ii) variations in testing methodology, including the use of indirect tensile tests in Sadeghnejad and Shafabakhsh [21] and higher strain amplitudes (500 and 700 microstrains) used by Mousavi Rad et al. [23]; and (iii) the higher nanoparticle concentration adopted in this study (8.5%) compared with the maximum levels used in the previous studies (1.2% and 8%, respectively).
Regarding the healing capacity, Table 13 presents the results of the second fatigue test conducted for both asphalt mixtures (M1 and M2). The table also shows the calculated healing percentages, which allow for the evaluation of the effectiveness of the damage recovery process induced by fatigue in each mixture.
The results in Table 13 indicate that the nanomodified asphalt mixture exhibited superior performance compared with the reference mixture in terms of mean healing capacity. The mean healing percentage obtained for the modified mixture was 23.5%, compared to 15.5% for the reference mixture, representing a 51.6% increase attributed to the presence of nanomaterials. When the healing values are converted into their normalized form (Equation (9)), the results are 23.2 × 10−8 [1/(J/m3)] for the nanomodified mixture and 15.4 × 10−8 [1/(J/m3)] for the reference mixture. The improvement in healing under this metric corresponds to a 50.7% increase, reinforcing the evidence that the combined addition of ZnO and TiO2 has a positive impact on the healing capacity of the asphalt mixture.
The strain-range analysis reveals that this effect is more pronounced at lower microstrain levels. Specifically, within the range of 225 µm/m to 236 µm/m, the nanomodified mixture achieved a healing rate of 21.7%, while the reference mixture reached only 5.5%, representing a substantial increase of 294.6%. According to Nascimento [116], microstrain levels between 100 µm/m and 200 µm/m are commonly observed at the base of asphalt layers in Brazilian pavements, based on typical traffic-load simulations. Therefore, the data suggest a superior performance of the nanomodified mixture under service-level strain conditions.

4.3.4. Assessment of the Internal Heating Rate of Asphalt Mixtures Under Microwave Radiation

Following the healing procedure, the mean internal temperature of the asphalt mixtures M1 (0% ZnO + TiO2) and M2 (8.5% ZnO + TiO2) was measured as a function of exposure time to microwave heating, as illustrated in Figure 21. The purpose of this stage was to verify the hypothesis that the incorporation of nanomaterials would increase the internal heating rate of the mixture, thereby enhancing the healing process.
According to Figure 21, the reference asphalt mixture (M1—0% ZnO + TiO2) exhibited an internal temperature increase rate of 0.18 °C/s, whereas the nanomodified mixture (M2—8.5% ZnO + TiO2) reached a rate of 0.27 °C/s. This difference represents a 50.0% increase in the heating rate compared with the conventional mixture, demonstrating the direct impact of nanomaterials on the thermal response of the composite. Considering the total exposure time of 140 s to microwave heating, the mean internal temperature reached 50.2 °C for M1 and 60.8 °C for M2, corresponding to an increase of approximately 21.2% due to the addition of metallic oxides.
This behavior highlights the enhanced thermal efficiency of the nanomodified mixture, resulting from the greater absorption and heat conduction promoted by ZnO and TiO2 nanoparticles. However, because both ZnO and TiO2 are semiconductor oxides with microwave–dielectric interaction [117,118,119,120], the observed temperature rise cannot be attributed solely to improved thermal conductivity. Instead, it likely results from the combined effect of heat conduction and microwave-induced dielectric heating, which could not be isolated under the conditions of this study.
Therefore, the superior healing capacity observed in the M2 mixture can be attributed, at least in part, to the faster and more intense increase in internal temperature during the heating process, which facilitates the mobilization of the polymeric asphalt binder and promotes the recovery of the fatigue-damaged structure. Nevertheless, the contribution of dielectric heating must be acknowledged as a potentially significant mechanism that remains unquantified and constitutes a major limitation of the present interpretation.

5. Conclusions

This study aimed to optimize the combined incorporation of nano-ZnO and nano-TiO2 (50/50 wt.%) into a polymer-modified asphalt binder and to evaluate the effects of this nanocomposite on the mechanical performance and healing capacity of asphalt mixtures. The main conclusions are as follows:
  • The combined addition of ZnO and TiO2 at contents ranging from 2 to 12 wt.% (in 2% increments) enhanced the high-temperature performance of the polymer-modified binder, increasing its thermal conductivity and resistance to permanent deformation.
  • A dosage of 8.5 wt.% provided the most favorable combination of thermal and rheological properties, resulting in higher thermal conductivity, improved high-temperature stability, and reduced rutting susceptibility, while maintaining workability within the SUPERPAVE limit (≤3.0 Pa·s at 135 °C).
  • The asphalt mixture produced with the binder modified with 8.5 wt.% ZnO + TiO2 exhibited higher stiffness at intermediate temperatures, a 50.7% increase in normalized healing efficiency, and 13.3% greater resistance to permanent deformation, demonstrating the contribution of the nanoparticles to mechanical reinforcement and structural recovery.
  • The improvement in healing capacity was associated with a 50.0% increase in the internal heating rate, which accelerated binder mobilization, increased molecular mobility, and facilitated crack closure and the reconstruction of fatigue-damaged microstructures.
  • Regarding fatigue performance under the same imposed microstrain levels, the incorporation of ZnO + TiO2 into the asphalt mixture resulted in a 4.4% reduction in the fatigue factor. This behavior is attributed to the increased stiffness provided by the nanomaterials, which reduces the ability of the mixture to deform under loading.
  • Overall, the ZnO + TiO2 combination enhances the functional performance of asphalt materials, adding benefits in resistance to permanent deformation and healing capacity in addition to the well-known photocatalytic potential reported in the literature, making this nanocomposite a promising alternative for pavements subjected to high temperatures and heavy traffic.
Although this study satisfactorily addressed the behavior of the materials at high and intermediate temperatures, several aspects warrant further investigation. Future research should examine the performance of the nanocomposites and mixtures at low temperatures, evaluate the economic feasibility of incorporating nanomaterials, conduct comparative studies between microwave heating and equivalent conventional heating to isolate thermal-conductivity-related effects from dielectric heating, and investigate the influence of different ZnO:TiO2 ratios to determine whether asymmetric formulations can further enhance microwave-heating efficiency and the healing performance of the asphalt mixture. Additional efforts should also focus on assessing the potential release of nanoparticles during production and service and developing methods for directly characterizing nanoparticle dispersion in asphalt matrices. Together, these investigations would provide a more comprehensive understanding of the effects of ZnO and TiO2 on asphalt mixture performance and durability.

Author Contributions

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

Funding

This research was funded by National Council for Scientific and Technological Development—CNPq (Process: 130424/2022-6); the Coordination for the Improvement of Higher Education Personnel—Brazil (CAPES)—Financing Code 001; and the National Department of Transportation Infrastructure (DNIT), Brazil, through Agreement Term TED No. 702/2020.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Acknowledgments

The authors acknowledge the support of the following laboratories: Paving Laboratory (LABPav/UFSC), Laboratory of Polymerization Processes and Control (LCP/UFSC), Laboratory of Nanotechnology (NANOTEC/UFSC), Laboratory of Multifunctional Materials and Numerical Experimentation (LAMMEN/UFRN), Analysis Center (EQA/UFSC), and Laboratory of Development and Technology in Pavement (LDTPav/UFSC).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Lavudya, P.; Pant, H.; Srikanth, V.V.S.S.; Ammanabrolu, R. Mesoporous and Phase Pure Anatase TiO2 Nanospheres for Enhanced Photocatalysis. Inorg. Chem. Commun. 2023, 152, 110699. [Google Scholar] [CrossRef]
  2. Saini, R.; Kumar, P. Green Synthesis of TiO2 Nanoparticles Using Tinospora Cordifolia Plant Extract & Its Potential Application for Photocatalysis and Antibacterial Activity. Inorg. Chem. Commun. 2023, 156, 111221. [Google Scholar] [CrossRef]
  3. Chang, X.; Li, Z.; Zhai, X.; Sun, S.; Gu, D.; Dong, L.; Yin, Y.; Zhu, Y. Efficient Synthesis of Sunlight-Driven ZnO-Based Heterogeneous Photocatalysts. Mater. Des. 2016, 98, 324–332. [Google Scholar] [CrossRef]
  4. Ahmed, M.A.; Abou-Gamra, Z.M.; ALshakhanbeh, M.A.; Medien, H. Control Synthesis of Metallic Gold Nanoparticles Homogeneously Distributed on Hexagonal ZnO Nanoparticles for Photocatalytic Degradation of Methylene Blue Dye. Environ. Nanotechnol. Monit. Manag. 2019, 12, 100217. [Google Scholar] [CrossRef]
  5. De Melo, J.V.S.; Trichês, G.; Gleize, P.J.P.; Villena, J. Development and Evaluation of the Efficiency of Photocatalytic Pavement Blocks in the Laboratory and after One Year in the Field. Constr. Build. Mater. 2012, 37, 310–319. [Google Scholar] [CrossRef]
  6. Bica, B.O.; De Melo, J.V.S. Concrete Blocks Nano-Modified with Zinc Oxide (ZnO) for Photocatalytic Paving: Performance Comparison with Titanium Dioxide (TiO2). Constr. Build. Mater. 2020, 252, 119120. [Google Scholar] [CrossRef]
  7. Babagoli, R.; Nasr, D.; Ameli, A.; Moradi, M.R. Rutting and Fatigue Properties of Modified Binders with Polymer and Titanium Dioxide Nanoparticles. Constr. Build. Mater. 2022, 345, 128423. [Google Scholar] [CrossRef]
  8. Cadorin, N.D. Avaliação do Comportamento Reológico e da Eficiência Fotocatalítica de Nanocompósitos Asfálticos Produzidos com a Incorporação de nano-TiO2 e nano-ZnO. Bachelor’s Thesis, Civil Engineering, Federal University of Santa Catarina (UFSC), Florianópolis, Brazil, 2019. Available online: https://repositorio.ufsc.br/handle/123456789/202850 (accessed on 23 November 2025).
  9. Rocha Segundo, I.; Ferreira, C.; Freitas, E.F.; Carneiro, J.O.; Fernandes, F.; Júnior, S.L.; Costa, M.F. Assessment of Photocatalytic, Superhydrophobic and Self-Cleaning Properties on Hot Mix Asphalts Coated with TiO2 and/or ZnO Aqueous Solutions. Constr. Build. Mater. 2018, 166, 500–509. [Google Scholar] [CrossRef]
  10. Carneiro, J.O.; Azevedo, S.; Teixeira, V.; Fernandes, F.; Freitas, E.; Silva, H.; Oliveira, J. Development of Photocatalytic Asphalt Mixtures by the Deposition and Volumetric Incorporation of TiO2 Nanoparticles. Constr. Build. Mater. 2013, 38, 594–601. [Google Scholar] [CrossRef]
  11. Mousavi Rad, S.; Kamboozia, N.; Ameri, M.; Mirabdolazimi, S.M. Feasibility of Concurrent Improvement of Pollutants-Absorption Ability from Surface Runoff and Mechanical Performance of Asphalt Mixtures by Using Photocatalytic Nanomodified Porous Asphalt. J. Mater. Civ. Eng. 2023, 35, 04023248. [Google Scholar] [CrossRef]
  12. Qian, G.; Yu, H.; Gong, X.; Zhao, L. Impact of Nano-TiO2 on the NO2 Degradation and Rheological Performance of Asphalt Pavement. Constr. Build. Mater. 2019, 218, 53–63. [Google Scholar] [CrossRef]
  13. Hassan, M.M.; Dylla, H.; Asadi, S.; Mohammad, L.N.; Cooper, S. Laboratory Evaluation of Environmental Performance of Photocatalytic Titanium Dioxide Warm-Mix Asphalt Pavements. J. Mater. Civ. Eng. 2012, 24, 599–605. [Google Scholar] [CrossRef]
  14. Neto, V.F.D.S.; Lucena, L.C.D.F.L.; De Barros, A.G.; Lucena, A.E.D.F.L.; Filho, P.G.T.M. Rheological Evaluation of Asphalt Binder Modified with Zinc Oxide Nanoparticles. Case Stud. Constr. Mater. 2022, 17, e01224. [Google Scholar] [CrossRef]
  15. Babagoli, R.; Rezaei, M. Using Artificial Neural Network Methods for Modeling Moisture Susceptibility of Asphalt Mixture Modified by Nano TiO2. J. Mater. Civ. Eng. 2022, 34, 04022108. [Google Scholar] [CrossRef]
  16. Yunus, K.N.M.; Abdullah, M.E.; Ahmad, M.K.; Kamaruddin, N.H.M.; Tami, H. Physical and Rheological Properties of Nano Zinc Oxide Modified Asphalt Binder. MATEC Web Conf. 2018, 250, 02004. [Google Scholar] [CrossRef]
  17. He, Z.; Xie, T.; Li, Q.; Wang, P. Physical and Antiaging Properties of Rodlike Nano-ZnO–Modified Asphalt. J. Mater. Civ. Eng. 2021, 33, 04021316. [Google Scholar] [CrossRef]
  18. Xu, X.; Guo, H.; Wang, X.; Zhang, M.; Wang, Z.; Yang, B. Physical Properties and Anti-Aging Characteristics of Asphalt Modified with Nano-Zinc Oxide Powder. Constr. Build. Mater. 2019, 224, 732–742. [Google Scholar] [CrossRef]
  19. Shafabakhsh, G.H.; Ani, O.J. Experimental Investigation of Effect of Nano TiO2/SiO2 Modified Bitumen on the Rutting and Fatigue Performance of Asphalt Mixtures Containing Steel Slag Aggregates. Constr. Build. Mater. 2015, 98, 692–702. [Google Scholar] [CrossRef]
  20. Staub De Melo, J.V.; Manfro, A.L.; Barra, B.S.; Dell’Antonio Cadorin, N.; Borba Broering, W. Evaluation of the Rheological Behavior and the Development of Performance Equations of Asphalt Composites Produced with Titanium Dioxide and Zinc Oxide Nanoparticles. Nanomaterials 2023, 13, 288. [Google Scholar] [CrossRef] [PubMed]
  21. Sadeghnejad, M.; Shafabakhsh, G. Use of Nano SiO 2 and Nano TiO 2 to Improve the Mechanical Behaviour of Stone Mastic Asphalt Mixtures. Constr. Build. Mater. 2017, 157, 965–974. [Google Scholar] [CrossRef]
  22. Shafabakhsh, G.; Mirabdolazimi, S.M.; Sadeghnejad, M. Evaluation the Effect of Nano-TiO2 on the Rutting and Fatigue Behavior of Asphalt Mixtures. Constr. Build. Mater. 2014, 54, 566–571. [Google Scholar] [CrossRef]
  23. Mousavi Rad, S.; Kamboozia, N.; Anupam, K.; Saed, S.A. Experimental Evaluation of the Fatigue Performance and Self-Healing Behavior of Nanomodified Porous Asphalt Mixtures Containing RAP Materials under the Aging Condition and Freeze–Thaw Cycle. J. Mater. Civ. Eng. 2022, 34, 04022323. [Google Scholar] [CrossRef]
  24. Kamboozia, N.; Mousavi Rad, S.; Saed, S.A. Laboratory Investigation of the Effect of Nano-ZnO on the Fracture and Rutting Resistance of Porous Asphalt Mixture under the Aging Condition and Freeze–Thaw Cycle. J. Mater. Civ. Eng. 2022, 34, 04022052. [Google Scholar] [CrossRef]
  25. Fakhri, M.; Shahryari, E. The Effects of Nano Zinc Oxide (ZnO) and Nano Reduced Graphene Oxide (RGO) on Moisture Susceptibility Property of Stone Mastic Asphalt (SMA). Case Stud. Constr. Mater. 2021, 15, e00655. [Google Scholar] [CrossRef]
  26. Fu, Z.; Tang, Y.; Ma, F.; Wang, Y.; Shi, K.; Dai, J.; Hou, Y.; Li, J. Rheological Properties of Asphalt Binder Modified by Nano-TiO2/ZnO and Basalt Fiber. Constr. Build. Mater. 2022, 320, 126323. [Google Scholar] [CrossRef]
  27. Xie, X.; Hui, T.; Luo, Y.; Li, H.; Li, G.; Wang, Z. Research on the Properties of Low Temperature and Anti-UV of Asphalt with Nano-ZnO/Nano-TiO2/Copolymer SBS Composite Modified in High-Altitude Areas. Adv. Mater. Sci. Eng. 2020, 2020, 9078731. [Google Scholar] [CrossRef]
  28. Rocha Segundo, I.G.D.; Dias, E.A.L.; Fernandes, F.D.P.; Freitas, E.F.D.; Costa, M.F.; Carneiro, J.O. Photocatalytic Asphalt Pavement: The Physicochemical and Rheological Impact of TiO2 Nano/Microparticles and ZnO Microparticles onto the Bitumen. Road Mater. Pavement Des. 2019, 20, 1452–1467. [Google Scholar] [CrossRef]
  29. Zhang, H.; Su, M.; Zhao, S.; Zhang, Y.; Zhang, Z. High and Low Temperature Properties of Nano-Particles/Polymer Modified Asphalt. Constr. Build. Mater. 2016, 114, 323–332. [Google Scholar] [CrossRef]
  30. Janotti, A.; Van De Walle, C.G. Fundamentals of Zinc Oxide as a Semiconductor. Rep. Prog. Phys. 2009, 72, 126501. [Google Scholar] [CrossRef]
  31. Qumar, U.; Hassan, J.Z.; Bhatti, R.A.; Raza, A.; Nazir, G.; Nabgan, W.; Ikram, M. Photocatalysis vs Adsorption by Metal Oxide Nanoparticles. J. Mater. Sci. Technol. 2022, 131, 122–166. [Google Scholar] [CrossRef]
  32. Babu, U.J.R.; Hareesh, K.; Rondiya, S.R.; Nagaraju, D.H.; Mahendra, K. Synthesis and Characterization of Nitrogen and Phosphorus Co-Doped TiO2 Nanoparticle Anchored Graphitic Carbon Nitride Nanosheets: Photocatalytic Application on Dye Removal. Diam. Relat. Mater. 2023, 139, 110292. [Google Scholar] [CrossRef]
  33. Li, Y.; Fernández-Seara, J.; Du, K.; Pardiñas, Á.Á.; Latas, L.L.; Jiang, W. Experimental Investigation on Heat Transfer and Pressure Drop of ZnO/Ethylene Glycol-Water Nanofluids in Transition Flow. Appl. Therm. Eng. 2016, 93, 537–548. [Google Scholar] [CrossRef]
  34. Permanasari, A.A.; Kuncara, B.S.; Puspitasari, P.; Sukarni, S.; Ginta, T.L.; Irdianto, W. Convective Heat Transfer Characteristics of TiO2-EG Nanofluid as Coolant Fluid in Heat Exchanger. In Proceedings of the International Conference on Biology and Applied Science (ICOBAS), Malang, Indonesia, 13–14 March 2019; p. 050015. [Google Scholar]
  35. Broering, W.B.; De Melo, J.V.S.; Manfro, A.L. Incorporation of Nanoalumina into a Polymeric Asphalt Matrix: Reinforcement of the Nanostructure, Improvement of Phase Stability, and Amplification of Rheological Parameters. Constr. Build. Mater. 2022, 320, 126261. [Google Scholar] [CrossRef]
  36. Manfro, A.L.; Staub De Melo, J.V.; Villena Del Carpio, J.A.; Broering, W.B. Permanent Deformation Performance under Moisture Effect of an Asphalt Mixture Modified by Calcium Carbonate Nanoparticles. Constr. Build. Mater. 2022, 342, 128087. [Google Scholar] [CrossRef]
  37. Schuster, L. Avaliação da Contribuição da lã de aço Associada ao Nanotubo de Carbono na Reparação de Danos em Misturas Asfálticas Submetidas ao Aquecimento por Micro-Ondas. Master’s Thesis, Civil Engineering, Federal University of Santa Catarina (UFSC), Florianópolis, Brazil, 2020. Available online: https://tede.ufsc.br/teses/PECV1225-D.pdf (accessed on 23 November 2025).
  38. ASTM C131M; Standard Test Method for Resistance to Degradation of Small-Size Coarse Aggregate by Abrasion and Impact in the Los Angeles Machine. ASTM International: West Conshohocken, PA, USA, 2020.
  39. ASTM C127; Standard Test Method for Relative Density (Specific Gravity) and Absorption of Coarse Aggregate. ASTM International: West Conshohocken, PA, USA, 2024.
  40. ASTM D5821; Standard Test Method for Determining the Percentage of Fractured Particles in Coarse Aggregate. ASTM International: West Conshohocken, PA, USA, 2017.
  41. ASTM C1252; Standard Test Methods for Uncompacted Void Content of Fine Aggregate (as Influenced by Particle Shape, Surface Texture, and Grading). ASTM International: West Conshohocken, PA, USA, 2023.
  42. AASHTO T 176; Plastic Fines in Graded Aggregates and Soils by Use of the Sand Equivalent Test. American Association of State Highway and Transportation Officials: Washington, DC, USA, 2022.
  43. AASHTO T 112; Standard Method of Test for Clay Lumps and Friable Particles in Aggregate. American Association of State Highway and Transportation Officials: Washington, DC, USA, 2023.
  44. ASTM C88; Standard Test Method for Soundness of Aggregates by Use of Sodium Sulfate or Magnesium Sulfate. ASTM International: West Conshohocken, PA, USA, 2018.
  45. CBB Asfaltos. Certificado de Análise: Controle de Qualidade Número 54.421; CBB Asfaltos: Curitiba, Brazil, 2022. [Google Scholar]
  46. ABNT NBR 15166; Modified Asphalt—Test Method for Phase Separation. Associação Brasileira De Normas Técnicas: Rio de Janeiro, Brazil, 2004.
  47. ABNT NBR 6296; Semi-Solid Bituminous Products—Determination of Density and Relative Density. Associação Brasileira De Normas Técnicas: Rio de Janeiro, Brazil, 2012.
  48. ABNT NBR 6576; Betuminous Materials—Determination of Penetration. Associação Brasileira De Normas Técnicas: Rio de Janeiro, Brazil, 2007.
  49. ABNT NBR 6560; Asphaltic Binders—Determination of the Softening Point—Ring-and-Ball Method. Associação Brasileira De Normas Técnicas: Rio de Janeiro, Brazil, 2016.
  50. ABNT NBR 11341; Petroleum Products—Determination of the Flash and Fire Points by Cleveland Open Cup. Associação Brasileira De Normas Técnicas: Rio de Janeiro, Brazil, 2015.
  51. ABNT NBR 15086; Bituminous Materials—Determination of the Elastic Recovery by Ductilometer of Elastomeric Polymer and Ground Tyre Rubber Modified Asphalt. Associação Brasileira De Normas Técnicas: Rio de Janeiro, Brazil, 2022.
  52. ABNT NBR 14855; Asphalt Binder—Determination of Solubility in Trichloroethylene. Associação Brasileira De Normas Técnicas: Rio de Janeiro, Brazil, 2015.
  53. Kraton Polymers. KRATONTM D1101 A Polymer; Kraton Polymers: Houston, TX, USA, 2024. [Google Scholar]
  54. ISO 2781; Rubber, Vulcanized or Thermoplastic—Determination of Density. International Organization for Standardization: Geneva, Switzerland, 2018.
  55. ASTM D1895; Standard Test Methods for Apparent Density, Bulk Factor and Pourability of Plastic Materials. ASTM International: West Conshohocken, PA, USA, 2024.
  56. ISO 1133-1; Plastics—Determination of the Melt Mass-Flow Rate (MFR) and Melt Volume-Flow Rate (MVR) of Thermoplastics—Part 1: Standard Method. International Organization for Standardization: Geneva, Switzerland, 2022.
  57. ISO 37; Rubber, Vulcanized or Thermoplastic—Determination of Tensile Stress-Strain Properties. International Organization for Standardization: Geneva, Switzerland, 2024.
  58. ISO 868; Plastics and Ebonite—Determination of Indentation Hardness by Means of a Durometer (Shore Hardness). International Organization for Standardization: Geneva, Switzerland, 2003.
  59. Nanostructured & Amorphous Materials Titanium Oxide Nanopowder (TiO2, Anatase, 99+%, 10 Nm). Available online: https://www.nanoamor.com/inc/sdetail/45710 (accessed on 30 March 2025).
  60. Nanostructured & Amorphous Materials Zinc Oxide Nanopowder (ZnO, 99%, 20 Nm). Available online: https://www.nanoamor.com/inc/sdetail/19983 (accessed on 30 March 2025).
  61. ASTM E2550; Standard Test Method for Thermal Stability by Thermogravimetry. ASTM International: West Conshohocken, PA, USA, 2021.
  62. Dell’Antonio Cadorin, N.; Victor Staub De Melo, J.; Borba Broering, W.; Luiz Manfro, A.; Salgado Barra, B. Asphalt Nanocomposite with Titanium Dioxide: Mechanical, Rheological and Photoactivity Performance. Constr. Build. Mater. 2021, 289, 123178. [Google Scholar] [CrossRef]
  63. Zidi, Z.; Ltifi, M.; Ayadi, Z.B.; Mir, L.E.; Nóvoa, X.R. Effect of Nano-ZnO on Mechanical and Thermal Properties of Geopolymer. J. Asian Ceram. Soc. 2020, 8, 1–9. [Google Scholar] [CrossRef]
  64. Freitas, M.R. de Síntese e Caracterização de Nanopartículas de ZnO Dopado e Codopado com Metais de Transição. Master’s Thesis, Materials Science and Engineering, Federal University of Paraná (UFPR), Curitiba, Brazil, 2018. Available online: https://hdl.handle.net/1884/69956 (accessed on 23 November 2025).
  65. Song, Y.; Li, H.; Xiong, Z.; Cheng, L.; Du, M.; Liu, Z.; Li, J.; Li, D. TiO2/Carbon Composites from Waste Sawdust for Methylene Blue Photodegradation. Diam. Relat. Mater. 2023, 136, 109918. [Google Scholar] [CrossRef]
  66. Das, A.; Kumar, P.M.; Bhagavathiachari, M.; Nair, R.G. Hierarchical ZnO-TiO2 Nanoheterojunction: A Strategy Driven Approach to Boost the Photocatalytic Performance through the Synergy of Improved Surface Area and Interfacial Charge Transport. Appl. Surf. Sci. 2020, 534, 147321. [Google Scholar] [CrossRef]
  67. Munguti, L.; Dejene, F. Effects of Zn:Ti Molar Ratios on the Morphological, Optical and Photocatalytic Properties of ZnO-TiO2 Nanocomposites for Application in Dye Removal. Mater. Sci. Semicond. Process. 2021, 128, 105786. [Google Scholar] [CrossRef]
  68. Li, B.; Yuan, D.; Gao, C.; Zhang, H.; Li, Z. Synthesis and Characterization of TiO2/ZnO Heterostructural Composite for Ultraviolet Photocatalytic Degrading DOM in Landfill Leachate. Environ. Sci. Pollut. Res. 2022, 29, 85510–85524. [Google Scholar] [CrossRef]
  69. Spínola, J.R.; Staub De Melo, J.V.; Manfro, A.L.; Salgado Barra, B. Effects of the Addition of Industrial Graphene Nanoplatelets on the Permanent Deformation Susceptibility, Fatigue Resistance, and Healing Capacity of an Asphalt Mixture Modified with Styrene–Butadiene–Styrene Copolymer. J. Mater. Civ. Eng. 2025, 37, 04025048. [Google Scholar] [CrossRef]
  70. Saltan, M.; Terzi, S.; Karahancer, S. Mechanical Behavior of Bitumen and Hot-Mix Asphalt Modified with Zinc Oxide Nanoparticle. J. Mater. Civ. Eng. 2019, 31, 04018399. [Google Scholar] [CrossRef]
  71. ASTM D2872; Standard Test Method for Effect of Heat and Air on a Moving Film of Asphalt (Rolling Thin-Film Oven Test). ASTM International: West Conshohocken, PA, USA, 2022.
  72. ASTM D6373; Standard Specification for Performance-Graded Asphalt Binder. ASTM International: West Conshohocken, PA, USA, 2023.
  73. ASTM D4402 M; Standard Test Method for Viscosity Determination of Asphalt at Elevated Temperatures Using a Rotational Viscometer. ASTM International: West Conshohocken, PA, USA, 2023.
  74. ASTM D7175; Standard Test Method for Determining the Rheological Properties of Asphalt Binder Using a Dynamic Shear Rheometer. ASTM International: West Conshohocken, PA, USA, 2024.
  75. ASTM D7405; Standard Test Method for Multiple Stress Creep and Recovery (MSCR) of Asphalt Binder Using a Dynamic Shear Rheometer. ASTM International: West Conshohocken, PA, USA, 2024.
  76. AASHTO T 315; Standard Method of Test for Determining the Rheological Properties of Asphalt Binder Using a Dynamic Shear Rheometer (DSR). American Association of State Highway and Transportation Officials: Washington, DC, USA, 2022.
  77. Manfro, A.L. Influência da Incorporação de Nanopartículas de Carbonato de Cálcio no Concreto Asfáltico Quanto à Resistência à Deformação Permanente e aos Efeitos Deletérios da Ação da Água. Master’s Thesis, Civil Engineering, Federal University of Santa Catarina (UFSC), Florianópolis, Brazil, 2022. Available online: https://tede.ufsc.br/teses/PECV1267-D.pdf (accessed on 23 November 2025).
  78. Broering, W.B. Efeitos da Incorporação de Nanopartículas de Óxido de Alumínio na Condutividade Térmica e nas Propriedades Reológicas de Ligantes Asfálticos. Master’s Thesis, Civil Engineering, Federal University of Santa Catarina (UFSC), Florianópolis, Brazil, 2020. Available online: https://tede.ufsc.br/teses/PECV1237-D.pdf (accessed on 23 November 2025).
  79. Spínola, J.R. Avaliação da Influência de Grafeno no Desempenho Mecânico e na Reparação aos Danos por Fadiga de uma Mistura Asfáltica com Incorporação de Polímero. Master’s Thesis, Civil Engineering, Federal University of Santa Catarina (UFSC), Florianópolis, Brazil, 2023. Available online: https://tede.ufsc.br/teses/PECV1312-D.pdf (accessed on 23 November 2025).
  80. AASHTO T 391; Standard Method of Test for Estimating Fatigue Resistance of Asphalt Binders Using the Linear Amplitude Sweep. American Association of State Highway and Transportation Officials: Washington, DC, USA, 2020.
  81. DNIT 439 ME; Pavimentação—Ligante Asfáltico—Avaliação Da Resistência à Fadiga de Ligantes Asfálticos Usando Varredura de Amplitude Linear (LAS—Linear Amplitude Sweep)—Método de Ensaio. Departamento Nacional de Infraestrutura de Transportes: Brasilia, Brazil, 2022. Available online: https://www.gov.br/dnit/pt-br/assuntos/planejamento-e-pesquisa/ipr/coletanea-de-normas/coletanea-de-normas/metodo-de-ensaio-me/dnit_439_2022_me.pdf (accessed on 23 November 2025).
  82. Melo, J.V.S. de Desenvolvimento e Estudo do Comportamento Reológico e Desempenho Mecânico de Concretos Asfálticos Modificados com Nanocompósitos. Ph.D. Thesis, Civil Engineering, Federal University of Santa Catarina (UFSC), Florianópolis, Brazil, 2014. Available online: https://tede.ufsc.br/teses/PECV0931-T.pdf (accessed on 23 November 2025).
  83. AASHTO M 323; Standard Specification for Superpave Volumetric Mix Design. American Association of State Highway and Transportation Officials: Washington, DC, USA, 2022.
  84. AASHTO R 35; Standard Practice for Superpave Volumetric Design for Asphalt Mixtures. American Association of State Highway and Transportation Officials: Washington, DC, USA, 2022.
  85. AASHTO R 30; Standard Practice for Mixture Conditioning of Hot Mix Asphalt (HMA). American Association of State Highway and Transportation Officials: Washington, DC, USA, 2022.
  86. EN 12697-33; Bituminous Mixtures. Test Methods—Part 33: Specimen Prepared by Roller Compactor. European Committee for Standardization (CEN): Brussels, Belgium, 2022.
  87. EN 12697-22; Bituminous Mixtures. Test Methods—Part 22: Wheel Tracking. European Committee for Standardization (CEN): Brussels, Belgium, 2020.
  88. EN 12697-26; Bituminous Mixtures—Test Methods—Part 26: Stiffness. European Committee for Standardization (CEN): Brussels, Belgium, 2022.
  89. De Melo, J.V.S.; Trichês, G. Evaluation of Properties and Fatigue Life Estimation of Asphalt Mixture Modified by Organophilic Nanoclay. Constr. Build. Mater. 2017, 140, 364–373. [Google Scholar] [CrossRef]
  90. EN 12697-24; Bituminous Mixtures. Test Methods—Part 24: Resistance to Fatigue. European Committee for Standardization (CEN): Brussels, Belgium, 2018.
  91. Instituto de Pesquisas Rodoviárias (IPR); Departamento Nacional de Infraestrutura de Transportes (DNIT); Instituto Alberto Luiz Coimbra de Pós-Graduação e Pesquisa de Engenharia, Universidade Federal do Rio de Janeiro (COPPE/UFRJ). Execução de Estudos e Pesquisa para Elaboração de Método Mecanístico–Empírico de Dimensionamento de Pavimento Asfáltico: Manual de Utilização do Programa MeDiNa; DNIT/IPR: Rio de Janeiro, Brazil, 2020.
  92. Gómez-Rodríguez, P.; Van Grieken, R.; López-Muñoz, M.J. Influence of the TiO2 Crystalline Phase on the Performance of UVA/Brookite/Persulfate and UVA/Anatase/Persulfate Systems for the Degradation of Isothiazolinones in Aqueous Matrices. J. Environ. Chem. Eng. 2024, 12, 113110. [Google Scholar] [CrossRef]
  93. Tian, Y.; Zhang, W.; Li, X.; Yin, X.; Liu, Y.; Su, S.; Wang, Q.; Zhang, L. Enhanced Photocatalytic Activity of Sn3O4/TiO2 Heterostructures for Cr(VI) Reduction and Isoniazid Degradation. Mater. Sci. Semicond. Process. 2025, 192, 109441. [Google Scholar] [CrossRef]
  94. Kahandal, A.; Chaudhary, S.; Methe, S.; Nagwade, P.; Sivaram, A.; Tagad, C.K. Galactomannan Polysaccharide as a Biotemplate for the Synthesis of Zinc Oxide Nanoparticles with Photocatalytic, Antimicrobial and Anticancer Applications. Int. J. Biol. Macromol. 2023, 253, 126787. [Google Scholar] [CrossRef]
  95. Jakhrani, M.A.; Bhatti, M.A.; Tahira, A.; Shah, A.A.; Dawi, E.A.; Vigolo, B.; Nafady, A.; Saleem, L.M.; Haj Ismail, A.A.K.; Ibupoto, Z.H. Biogenic Preparation of ZnO Nanostructures Using Leafy Spinach Extract for High-Performance Photodegradation of Methylene Blue under the Illumination of Natural Sunlight. Molecules 2023, 28, 2773. [Google Scholar] [CrossRef]
  96. Han, Z.; Cong, P. Understanding Relationships between Aging Induced Variation of Asphaltene Aggregation Morphology and Asphalt Properties through Molecular Dynamics Simulation. Constr. Build. Mater. 2024, 420, 135610. [Google Scholar] [CrossRef]
  97. Jin, J.; Liu, S.; Chen, H.; Wen, Z.; Xiao, M.; Rao, R.; Zheng, J. Evaluation of Compatibility in Bio-Oil and Zinc Oxide Modified Asphalt to Facilitate Waste Recycling. J. Clean. Prod. 2024, 476, 143710. [Google Scholar] [CrossRef]
  98. Gebresellasie, K.; Lewis, J.C.; Shirokoff, J. X-Ray Spectral Line Shape Analysis of Asphalt Binders. Energy Fuels 2013, 27, 2018–2024. [Google Scholar] [CrossRef]
  99. Nagaraju, G.; Udayabhanu; Shivaraj; Prashanth, S.A.; Shastri, M.; Yathish, K.V.; Anupama, C.; Rangappa, D. Electrochemical Heavy Metal Detection, Photocatalytic, Photoluminescence, Biodiesel Production and Antibacterial Activities of Ag–ZnO Nanomaterial. Mater. Res. Bull. 2017, 94, 54–63. [Google Scholar] [CrossRef]
  100. Rajith Kumar, C.R.; Betageri, V.S.; Nagaraju, G.; Pujar, G.H.; Onkarappa, H.S.; Latha, M.S. Synthesis of Core/Shell (ZnO/Ag) Nanoparticles Using Calotropis Gigantea and Their Applications in Photocatalytic and Antibacterial Studies. J. Inorg. Organomet. Polym. Mater. 2020, 30, 3410–3417. [Google Scholar] [CrossRef]
  101. Worku, A.K.; Ayele, D.W.; Habtu, N.G.; Melas, G.A.; Yemata, T.A.; Mekonnen, N.Y.; Teshager, M.A. Structural and Thermal Properties of Pure and Chromium Doped Zinc Oxide Nanoparticles. SN Appl. Sci. 2021, 3, 699. [Google Scholar] [CrossRef]
  102. Xiong, G.; Pal, U.; Serrano, J.G. Correlations among Size, Defects, and Photoluminescence in ZnO Nanoparticles. J. Appl. Phys. 2007, 101, 024317. [Google Scholar] [CrossRef]
  103. Wahab, R.; Ansari, S.G.; Kim, Y.S.; Seo, H.K.; Kim, G.S.; Khang, G.; Shin, H.-S. Low Temperature Solution Synthesis and Characterization of ZnO Nano-Flowers. Mater. Res. Bull. 2007, 42, 1640–1648. [Google Scholar] [CrossRef]
  104. Facin, F.; Staub De Melo, J.V.; Costa Puerari, R.; Matias, W.G. Toxicological Effects of Leachates Extracted from Photocatalytic Concrete Blocks with Nano-TiO2 on Daphnia Magna. Nanomaterials 2024, 14, 1447. [Google Scholar] [CrossRef]
  105. Islam, T.; Roni, N.P.; Ali, Y.; Islam, R.; Hossan, S.; Rahman, M.H.; Zahid, A.A.S.M.; E Alam, N.; Hanif, A.; Akhtar, M.S. Selectivity of Sol-Gel and Hydrothermal TiO2 Nanoparticles towards Photocatalytic Degradation of Cationic and Anionic Dyes. Molecules 2023, 28, 6834. [Google Scholar] [CrossRef]
  106. Chawraba, K.; Medlej, H.; Noun, A.; Sakr, M.; Toufaily, J.; Lalevee, J.; Hamieh, T. Photocatalytic Epoxy Paint Based on TiO2 for the Decontamination of Water under Visible LED and Sunlight Irradiation. ChemistrySelect 2024, 9, e202303055. [Google Scholar] [CrossRef]
  107. Hesami, S.; Mahzari, M.; Sobhi, S.; Bay, M.A. Investigating the Rheological Properties and Microstructural Analysis of Nano-Expanded Perlite Modified Asphalt Binder. Case Stud. Constr. Mater. 2025, 22, e04208. [Google Scholar] [CrossRef]
  108. Filho, P.G.T.M.; Rodrigues Dos Santos, A.T.; Lucena, L.C.D.F.L.; De Sousa Neto, V.F. Rheological Evaluation of Asphalt Binder 50/70 Incorporated with Titanium Dioxide Nanoparticles. J. Mater. Civ. Eng. 2019, 31, 04019235. [Google Scholar] [CrossRef]
  109. Yan, C.; Xiao, F.; Huang, W.; Lv, Q. Critical Matters in Using Attenuated Total Reflectance Fourier Transform Infrared to Characterize the Polymer Degradation in Styrene–Butadiene–Styrene-Modified Asphalt Binders. Polym. Test. 2018, 70, 289–296. [Google Scholar] [CrossRef]
  110. Yu, J.; Vaidya, M.; Su, G.; Adhikari, S.; Korolev, E.; Shekhovtsova, S. Experimental Study of Soda Lignin Powder as an Asphalt Modifier for a Sustainable Pavement Material. Constr. Build. Mater. 2021, 298, 123884. [Google Scholar] [CrossRef]
  111. Wu, S.; Pang, L.; Mo, L.; Chen, Y.; Zhu, G. Influence of Aging on the Evolution of Structure, Morphology and Rheology of Base and SBS Modified Bitumen. Constr. Build. Mater. 2009, 23, 1005–1010. [Google Scholar] [CrossRef]
  112. Neto, V.F.D.S.; Minervina Silva, I.; Lucena, L.C.D.F.L.; Lucena, A.E.D.F.L.; De Medeiros Melo Neto, O.; Lima, R.K.B.D. Effect of Superficially Modified Zinc Oxide Nanoparticles as an Additive on the Rheological Performance of Asphalt Binder. Road Mater. Pavement Des. 2024, 25, 1211–1228. [Google Scholar] [CrossRef]
  113. Broering, W.B.; De Melo, J.V.S.; Manfro, A.L. Modification of the Asphalt Binder with Nano-Aluminum Oxide: An Alternative to Improve the Thermal Conductivity and the Rheological Properties of the Asphalt Matrix. J. Test. Eval. 2024, 52, 1109–1128. [Google Scholar] [CrossRef]
  114. Asphalt Institute. Asphalt Institute Asphalt Binder Testing. Manual Series N° 25 (MS-25); Asphalt Institute: Lexington, KY, USA, 2012. [Google Scholar]
  115. Günay, T.; Ahmedzade, P. Physical and Rheological Properties of Nano- TiO 2 and Nanocomposite Modified Bitumens. Constr. Build. Mater. 2020, 243, 118208. [Google Scholar] [CrossRef]
  116. Nascimento, L.A.H. Do Implementation and Validation of the Viscoelastic Continuum Damage Theory for Asphalt Mixture and Pavement Analysis in Brazil. Ph.D. Thesis, North Carolina State University, Raleigh, NC, USA, 2015. [Google Scholar]
  117. Horikoshi, S.; Matsubara, A.; Takayama, S.; Sato, M.; Sakai, F.; Kajitani, M.; Abe, M.; Serpone, N. Characterization of Microwave Effects on Metal-Oxide Materials: Zinc Oxide and Titanium Dioxide. Appl. Catal. B Environ. 2010, 99, 490–495. [Google Scholar] [CrossRef]
  118. Yang, Z.; Luo, F.; Hu, Y.; Duan, S.; Zhu, D.; Zhou, W. Dielectric and Microwave Absorption Properties of TiO2/Al2O3 Coatings and Improved Microwave Absorption by FSS Incorporation. J. Alloys Compd. 2016, 678, 527–532. [Google Scholar] [CrossRef]
  119. Ahmad, P.; Venkateswara Rao, A.; Suresh Babu, K.; Narsinga Rao, G. Particle Size Effect on the Dielectric Properties of ZnO Nanoparticles. Mater. Chem. Phys. 2019, 224, 79–84. [Google Scholar] [CrossRef]
  120. Sun, J.; Wang, W.; Yue, Q. Review on Microwave-Matter Interaction Fundamentals and Efficient Microwave-Associated Heating Strategies. Materials 2016, 9, 231. [Google Scholar] [CrossRef]
Figure 1. Scanning electron micrographs of (A) zinc oxide (ZnO) and (B) titanium dioxide (TiO2).
Figure 1. Scanning electron micrographs of (A) zinc oxide (ZnO) and (B) titanium dioxide (TiO2).
Nanomaterials 15 01779 g001
Figure 2. Thermogravimetric curves of (A) zinc oxide (ZnO) and (B) titanium dioxide (TiO2) nanoparticles.
Figure 2. Thermogravimetric curves of (A) zinc oxide (ZnO) and (B) titanium dioxide (TiO2) nanoparticles.
Nanomaterials 15 01779 g002
Figure 3. Diagram of the experimental method.
Figure 3. Diagram of the experimental method.
Nanomaterials 15 01779 g003
Figure 4. Nanomodification protocol of polymeric reference asphalt binder.
Figure 4. Nanomodification protocol of polymeric reference asphalt binder.
Nanomaterials 15 01779 g004
Figure 5. Test procedure for obtaining thermal conductivity: (A) highlighting the MTPS sensor of the Thermal Conductivity Analyzer equipment; (B) applying thermal grease to the sample to be tested; (C) placing the sample on top of the MTPS sensor; and, (D) positioning the weight on top of the sample and performing the test.
Figure 5. Test procedure for obtaining thermal conductivity: (A) highlighting the MTPS sensor of the Thermal Conductivity Analyzer equipment; (B) applying thermal grease to the sample to be tested; (C) placing the sample on top of the MTPS sensor; and, (D) positioning the weight on top of the sample and performing the test.
Nanomaterials 15 01779 g005
Figure 6. Production process of asphalt mixture specimens using the French wheel compactor. The (upper) sequence illustrates the procedure for producing specimens for the permanent deformation test, while the (lower) sequence represents the production of slabs for intermediate-temperature rheological, fatigue, healing, and internal heating-rate tests.
Figure 6. Production process of asphalt mixture specimens using the French wheel compactor. The (upper) sequence illustrates the procedure for producing specimens for the permanent deformation test, while the (lower) sequence represents the production of slabs for intermediate-temperature rheological, fatigue, healing, and internal heating-rate tests.
Nanomaterials 15 01779 g006
Figure 9. Bright-field microscopy images of samples in the unaged state: (A) reference asphalt binder N0 (0% ZnO + TiO2) and nanocomposites with varying concentrations of ZnO + TiO2: (B) N2 (2% ZnO + TiO2), (C) N4 (4% ZnO + TiO2), (D) N6 (6% ZnO + TiO2), (E) N8 (8% ZnO + TiO2), (F) N10 (10% ZnO + TiO2), and (G) N12 (12% ZnO + TiO2).
Figure 9. Bright-field microscopy images of samples in the unaged state: (A) reference asphalt binder N0 (0% ZnO + TiO2) and nanocomposites with varying concentrations of ZnO + TiO2: (B) N2 (2% ZnO + TiO2), (C) N4 (4% ZnO + TiO2), (D) N6 (6% ZnO + TiO2), (E) N8 (8% ZnO + TiO2), (F) N10 (10% ZnO + TiO2), and (G) N12 (12% ZnO + TiO2).
Nanomaterials 15 01779 g009
Figure 10. Mass loss of reference asphalt binder and nanocomposites produced after short-term aging simulation performed with Rolling Thin-Film Oven Test (RTFOT) [71,72]. Error bar: standard deviation.
Figure 10. Mass loss of reference asphalt binder and nanocomposites produced after short-term aging simulation performed with Rolling Thin-Film Oven Test (RTFOT) [71,72]. Error bar: standard deviation.
Nanomaterials 15 01779 g010
Figure 11. X-ray diffraction patterns of the nanomaterials (ZnO and TiO2) and of the samples in the short-term aging state of the reference asphalt binder N0 (0% ZnO + TiO2) and of the nanocomposites with varying concentrations of ZnO + TiO2: N2 (2%), N4 (4%), N6 (6%), N8 (8%), N10 (10%) and N12 (12%). Dashed vertical lines: indicate the 2θ positions corresponding to the main diffraction peaks of ZnO, TiO2, and the reference asphalt binder (N0).
Figure 11. X-ray diffraction patterns of the nanomaterials (ZnO and TiO2) and of the samples in the short-term aging state of the reference asphalt binder N0 (0% ZnO + TiO2) and of the nanocomposites with varying concentrations of ZnO + TiO2: N2 (2%), N4 (4%), N6 (6%), N8 (8%), N10 (10%) and N12 (12%). Dashed vertical lines: indicate the 2θ positions corresponding to the main diffraction peaks of ZnO, TiO2, and the reference asphalt binder (N0).
Nanomaterials 15 01779 g011
Figure 12. FTIR spectra of the nanomaterials (ZnO and TiO2) and of the samples in the short-term aging state of the reference asphalt binder N0 (0% ZnO + TiO2) and of the nanocomposites with varying concentrations of ZnO + TiO2: N2 (2%), N4 (4%), N6 (6%), N8 (8%), N10 (10%) and N12 (12%).
Figure 12. FTIR spectra of the nanomaterials (ZnO and TiO2) and of the samples in the short-term aging state of the reference asphalt binder N0 (0% ZnO + TiO2) and of the nanocomposites with varying concentrations of ZnO + TiO2: N2 (2%), N4 (4%), N6 (6%), N8 (8%), N10 (10%) and N12 (12%).
Nanomaterials 15 01779 g012
Figure 13. Thermal conductivity of polymeric asphalt binder and nanocomposites under short-term aging condition (RTFOT).
Figure 13. Thermal conductivity of polymeric asphalt binder and nanocomposites under short-term aging condition (RTFOT).
Nanomaterials 15 01779 g013
Figure 14. Apparent viscosity of polymeric asphalt binder and nanocomposites at temperatures of 135 °C, 150 °C and 177 °C.
Figure 14. Apparent viscosity of polymeric asphalt binder and nanocomposites at temperatures of 135 °C, 150 °C and 177 °C.
Nanomaterials 15 01779 g014
Figure 15. Influence of the incorporation of ZnO + TiO2 in the polymeric asphalt binder on the PGH classification at high temperatures, in the short-term aging (RTFOT) and unaged condition. Error bar: standard deviation.
Figure 15. Influence of the incorporation of ZnO + TiO2 in the polymeric asphalt binder on the PGH classification at high temperatures, in the short-term aging (RTFOT) and unaged condition. Error bar: standard deviation.
Nanomaterials 15 01779 g015
Figure 16. Corresponding aging index (AI) for each analyzed test temperature and incorporation content of ZnO + TiO2.
Figure 16. Corresponding aging index (AI) for each analyzed test temperature and incorporation content of ZnO + TiO2.
Nanomaterials 15 01779 g016
Figure 17. Behavior of recovery percentage (%R3.2) in relation to the content of ZnO and TiO2 incorporated in the polymeric asphalt binder. Error bar: standard deviation.
Figure 17. Behavior of recovery percentage (%R3.2) in relation to the content of ZnO and TiO2 incorporated in the polymeric asphalt binder. Error bar: standard deviation.
Nanomaterials 15 01779 g017
Figure 18. Behavior of non-recoverable compliance (Jnr3.2) in relation to the content of ZnO and TiO2 incorporated in the polymeric asphalt binder. Error bar: standard deviation.
Figure 18. Behavior of non-recoverable compliance (Jnr3.2) in relation to the content of ZnO and TiO2 incorporated in the polymeric asphalt binder. Error bar: standard deviation.
Nanomaterials 15 01779 g018
Figure 19. Accumulated deformation of reference (M1—0% ZnO + TiO2) and nanomodified (M2—8.5% ZnO + TiO2) asphalt mixtures in relation to the number of applied cycles.
Figure 19. Accumulated deformation of reference (M1—0% ZnO + TiO2) and nanomodified (M2—8.5% ZnO + TiO2) asphalt mixtures in relation to the number of applied cycles.
Nanomaterials 15 01779 g019
Figure 20. Fatigue curves: reference (M1—0% ZnO + TiO2) and nanomodified (M2—8.5% ZnO + TiO2) asphalt mixture.
Figure 20. Fatigue curves: reference (M1—0% ZnO + TiO2) and nanomodified (M2—8.5% ZnO + TiO2) asphalt mixture.
Nanomaterials 15 01779 g020
Figure 21. Mean internal temperature as a function of microwave heating time of asphalt mixtures.
Figure 21. Mean internal temperature as a function of microwave heating time of asphalt mixtures.
Nanomaterials 15 01779 g021
Table 1. Properties of mineral aggregates [37].
Table 1. Properties of mineral aggregates [37].
PropertiesTest MethodResults
Hardness (Los Angeles abrasion)ASTM C131 [38]18.64%
Absorption of coarse aggregateASTM C127 [39]1.43%
Angularity of coarse aggregateASTM D5821 [40]100%
Angularity of fine aggregateASTM C1252 [41]52.36%
Clay content (sand equivalent)AASHTO T 176 [42]71.95%
Bulk specific gravity (Gsb) of coarse aggregateASTM C127 [39]2.592 g/cm3
Apparent specific gravity (Gsa) of coarse aggregateASTM C127 [39]2.648 g/cm3
Deleterious materialAASHTO T 112 [43]0.00%
SoundnessASTM C88 [44]1.59%
Table 2. Results of characterization tests of polymer-modified asphalt binder [45].
Table 2. Results of characterization tests of polymer-modified asphalt binder [45].
PropertiesTest MethodResults
Phase separation (∆PA)NBR 15166 [46]0.3 °C
Density, 25 °CNBR 6296 [47]1.009
Penetration, 100 g, 5 s, 25 °CNBR 6576 [48]48 (0.1 mm)
Softening pointNBR 6560 [49]68 °C
Flash and fire pointsNBR 11341 [50]254 °C
Elastic recovery, 20 cm, 25 °CNBR 15086 [51]90.0%
Solubility in trichloroethyleneNBR 14855 [52]99.9% (mass)
Table 3. Typical properties of D1101 A copolymer [53].
Table 3. Typical properties of D1101 A copolymer [53].
PropertyTest MethodTypical Value
Specific gravityISO 2781 [54]0.94
Bulk densityASTM D1895 method B [55]0.4 kg/dm3
Melt flow rate, 200 °C/5 kgISO 1133 [56]<1 g/10 min
Tensile strengthISO 37 [57]33 MPa
300% modulusISO 37 [57]2.9 MPa
Elongation at breakISO 37 [57]880%
Hardness, shore A (30 s)ISO 868 [58]72
Table 4. Properties of ZnO and TiO2 nanoparticles [59,60].
Table 4. Properties of ZnO and TiO2 nanoparticles [59,60].
PropertiesZnOTiO2
Purity>99%>99%
Mean particle size20 nm10 nm
Morphology of particlesSphericalEllipsoidal/spherical
Specific surface area≥40 m2/g≥60 m2/g
Bulk density0.1–0.2 g/cm30.2–0.3 g/cm3
pH value6.5–7.55.5–7.0
Loss on drying (110 °C/2 h)≤1.0 wt.%≤2.0 wt.%
Loss on calcination (850 °C/2 h)≤3.0 wt.%≤5.0 wt.%
Table 5. Produced asphalt nanocomposites.
Table 5. Produced asphalt nanocomposites.
Number of Produced
Nanocomposites
Incorporation (by Binder Weight)Produced
Nanocomposite
Nanomaterial
ZnOTiO2
N0 (reference)0%0%0% (ZnO + TiO2)
N21%1%1% (ZnO + TiO2)
N42%2%2% (ZnO + TiO2)
N63%3%3% (ZnO + TiO2)
N84%4%4% (ZnO + TiO2)
N105%5%5% (ZnO + TiO2)
N126%6%6% (ZnO + TiO2)
Table 6. Granulometric composition of asphalt mixtures [82]. “No.” indicates sieve number and “in.” denotes aperture size in inches.
Table 6. Granulometric composition of asphalt mixtures [82]. “No.” indicates sieve number and “in.” denotes aperture size in inches.
Aggregate FractionSieves (ASTM Series)% Retained
Coarse aggregates1/2 in.22.5
3/8 in.16.2
No. 418.0
Fine aggregatesNo. 1019.0
No. 166.9
No. 304.8
No. 502.8
No. 1002.2
No. 2002.2
FillerFiller5.4
Table 7. Games-Howell post hoc test.
Table 7. Games-Howell post hoc test.
N0N2N4N6N8N10N12
N0Mean difference-−0.0683−0.0955−0.135−0.1895−0.2698−0.3163
p-value-<0.001<0.001<0.001<0.001<0.001<0.001
N2Mean difference -−0.0272−0.0667−0.1212−0.2015−0.248
p-value -<0.001<0.001<0.001<0.001<0.001
N4Mean difference -−0.0395−0.094−0.1743−0.2208
p-value -0.003<0.001<0.001<0.001
N6Mean difference -−0.0545−0.1348−0.1813
p-value -<0.001<0.001<0.001
N8Mean difference -−0.0803−0.1268
p-value -<0.001<0.001
N10Mean difference -−0.0465
p-value -<0.001
N12Mean difference -
p-value -
Table 8. Results of the dynamic shear modulus (|G*|), phase angle (δ), and elastic (G′) and viscous (G″) components of the reference (0% ZnO + TiO2) and nanomodified (8.5% ZnO + TiO2) asphalt mixtures, evaluated at different temperatures (35 °C to 5 °C) considering the mean across all loading frequencies (0.1 Hz to 30 Hz).
Table 8. Results of the dynamic shear modulus (|G*|), phase angle (δ), and elastic (G′) and viscous (G″) components of the reference (0% ZnO + TiO2) and nanomodified (8.5% ZnO + TiO2) asphalt mixtures, evaluated at different temperatures (35 °C to 5 °C) considering the mean across all loading frequencies (0.1 Hz to 30 Hz).
TemperatureParameterMean (0.1 Hz–30 Hz)% Variation
0% ZnO + TiO2
(N0)
8.5% ZnO + TiO2
(Noptimal)
35 °C|G*| (kPa)394802103.4%
δ (°)64.164.20.1%
G′ (kPa)178362103.1%
G″ (kPa)352715103.3%
30 °C|G*| (kPa)9852019105.0%
δ (°)61.661.3−0.4%
G′ (kPa)5121054106.0%
G″ (kPa)8401717104.3%
25 °C|G*| (kPa)26455419104.9%
δ (°)55.955.5−0.8%
G′ (kPa)16303377107.2%
G″ (kPa)20754225103.6%
20 °C|G*| (kPa)633212,846102.9%
δ (°)49.248.6−1.2%
G′ (kPa)44639116104.2%
G″ (kPa)44679001101.5%
15 °C|G*| (kPa)13,51027,091100.5%
δ (°)42.541.9−1.3%
G′ (kPa)10,47021,075101.3%
G″ (kPa)848216,91599.4%
10 °C|G*| (kPa)26,63055,413108.1%
δ (°)35.835.1−2.1%
G′ (kPa)22,30346,685109.3%
G″ (kPa)14,42529,615105.3%
5 °C|G*| (kPa)48,23299,115105.5%
δ (°)30.429.9−1.7%
G′ (kPa)42,46287,556106.2%
G″ (kPa)22,64446,015103.2%
Table 9. LAS test parameters calculated with the “AASHTO T 391-20—Version 1.59” [80] spreadsheet (MARC website) for the test temperature of 20 °C. Nf: number of cycles until rupture; FF: asphalt binder fatigue factor; γ: applied shear strain; SD: standard deviation; and CV: coefficient of variation.
Table 9. LAS test parameters calculated with the “AASHTO T 391-20—Version 1.59” [80] spreadsheet (MARC website) for the test temperature of 20 °C. Nf: number of cycles until rupture; FF: asphalt binder fatigue factor; γ: applied shear strain; SD: standard deviation; and CV: coefficient of variation.
ParameterN0 (0% ZnO + TiO2)Noptimal (8.5% ZnO + TiO2)% Variation
MeanSDCVMeanSDCV
A10,350,2461,753,85817.0%9,226,4901,399,26815.2%−10.9%
B5.210.010.1%5.240.020.4%0.6%
Nf (γ: 1.25%)3,236,134544,32516.8%2,865,510436,16115.2%−11.5%
Nf (γ: 2.50%)87,4171416.4%75,81711,72015.5%−13.3%
Nf (γ: 5.00%)236137916.1%200631715.8%−15.0%
Nf (γ: 7.50%)2864515.9%2403816.0%−16.1%
Nf (γ: 10.00%)641015.7%53916.2%−16.8%
Nf (γ: 12.50%)20315.6%16316.4%−17.3%
Nf (γ: 15.00%)8115.5%6116.5%−17.8%
FF (γ: 1.25–2.50%)1.720.021.3%1.710.021.1%−1.0%
Table 10. Results of the dynamic modulus (|E*|), phase angle (δ), and elastic (E1) and viscous (E2) components of the reference (M1—0% ZnO + TiO2) and nanomodified (M2—8.5% ZnO + TiO2) asphalt mixtures, evaluated at different temperatures (30 °C to 0 °C) considering the mean across all loading frequencies (0.1 Hz to 20 Hz).
Table 10. Results of the dynamic modulus (|E*|), phase angle (δ), and elastic (E1) and viscous (E2) components of the reference (M1—0% ZnO + TiO2) and nanomodified (M2—8.5% ZnO + TiO2) asphalt mixtures, evaluated at different temperatures (30 °C to 0 °C) considering the mean across all loading frequencies (0.1 Hz to 20 Hz).
Temperature ParameterMean (0.1 Hz–20 Hz)% Variation
0% ZnO + TiO2 (M1)8.5% ZnO + TiO2 (M2)
30 °C|E*| (kPa)1,536,6672,012,88931.0%
δ (°)42.140.1−4.6%
E2 (kPa)1,004,9441,230,10822.4%
E1 (kPa)1,160,8681,588,54736.8%
25 °C|E*| (kPa)2,774,7783,359,33321.1%
δ (°)36.634.5−5.9%
E2 (kPa)1,531,2541,711,36311.8%
E1 (kPa)2,304,0122,873,80224.7%
20 °C|E*| (kPa)5,293,6675,975,77812.9%
δ (°)27.425.2−7.9%
E2 (kPa)2,171,3832,280,8325.0%
E1 (kPa)4,802,0305,495,79714.5%
15 °C|E*| (kPa)8,268,1119,056,7789.5%
δ (°)19.717.9−9.2%
E2 (kPa)2,488,8402,502,4370.6%
E1 (kPa)7,852,3718,673,55110.5%
10 °C|E*| (kPa)11,605,00012,281,3335.8%
δ (°)14.012.5−10.3%
E2 (kPa)2,580,5232,463,172−4.6%
E1 (kPa)11,291,72512,008,6926.4%
5 °C|E*| (kPa)15,103,44415,672,7783.8%
δ (°)9.68.7−8.8%
E2 (kPa)2,350,0582,240,258−4.7%
E1 (kPa)14,903,35015,498,4574.0%
0 °C|E*| (kPa)18,560,55619,388,8894.5%
δ (°)6.45.9−9.0%
E2 (kPa)1,970,8551,901,843−3.5%
E1 (kPa)18,444,33819,288,8784.6%
Table 11. Results of the test specimens of the reference (M1) and nanomodified (M2) asphalt mixture tested for fatigue in the 4-point bending equipment. ε (µm/m): effective microstrain applied to the specimen; |E*| initial: initial dynamic modulus measured at the 100th loading cycle; Nf: number of applications of the request until the initial dynamic modulus is reduced by 50% and SD: standard deviation.
Table 11. Results of the test specimens of the reference (M1) and nanomodified (M2) asphalt mixture tested for fatigue in the 4-point bending equipment. ε (µm/m): effective microstrain applied to the specimen; |E*| initial: initial dynamic modulus measured at the 100th loading cycle; Nf: number of applications of the request until the initial dynamic modulus is reduced by 50% and SD: standard deviation.
Reference Asphalt Mixture
(M1)
Nanomodified Asphalt Mixture (M2)
ε (µm/m)|E*| Initial (MPa)Nfε (µm/m)|E*| Initial (MPa)Nf
2817741145,6112897401108,932
288798371,034289723670,574
289716350,745292714263,355
2887980109,200293725232,980
2897585196,106290777480,782
2806893197,173290816361,364
287793379,516291723632,584
2516226191,576290771659,148
2516610220,8232516270122,997
2507318290,6322506285226,283
2506209476,063251711588,356
2506375161,7442587987128,532
2586769199,6642576836133,059
2596911141,5922588212115,448
2257257641,2632597043224,878
2256444914,5782257171405,659
2256568665,9882256526134,329
2266769233,2452267647505,834
2266613391,7552257328222,345
2357503455,0502357647206,528
2257804354,816
2367571414,825
2347181149,794
Mean7043 Mean7328
SD591SD526
Table 12. Fatigue factor of reference and nanomodified asphalt mixtures.
Table 12. Fatigue factor of reference and nanomodified asphalt mixtures.
Asphalt Mixture0% ZnO + TiO2 (M1)8.5% ZnO + TiO2 (M2)
EquationNf = 4 × 1021 (ε)−6.74Nf = 2 × 1020 (ε)−6.29
Nf100123,448,26549,674,175
Nf250256,667155,987
FFM2.702.58
Table 13. Mean percentage of healing of reference (M1) and nanomodified (M2) asphalt mixtures. H: Healing.
Table 13. Mean percentage of healing of reference (M1) and nanomodified (M2) asphalt mixtures. H: Healing.
Reference Asphalt Mixture (M1)Nanomodified Asphalt Mixture (M2)
ε (µm/m)Number of
Cycles (Nf)
%Hε (µm/m)Number of
Cycles (Nf)
%H
1st
Fatigue Test
2st
Fatigue Test
1st
Fatigue Test
2st
Fatigue Test
281145,61131,63621.7289108,93220,94619.2
28871,03412,74918.028970,57411,53616.4
28950,74512,62724.929263,35523,49537.1
288109,20023,65721.729332,980884926.8
289196,10650,03025.529080,78214,26517.7
280197,17332,65116.629061,36411,40018.6
28779,51622,04227.729132,58410,49732.2
251191,57618,1079.529059,14818,76131.7
251220,82357,31926.0251122,99729,19923.7
250290,63253,25618.3250226,28352,83123.4
250476,06369,43514.625188,35626,90230.5
250161,74483235.2258128,53221,42716.7
258199,66458,94529.5257133,05943,62332.8
259141,59225,92818.3258115,44824,22121.0
225641,26318,9933.0259224,87842,73219.0
225914,57811,9381.3225405,65997,27824.0
225665,98830,1854.5225134,32937,32827.8
226233,24516,0086.9226505,834115,00022.7
226391,75531,7508.1225222,34544,69020.1
235455,05042,6259.4235206,52835,31817.1
225354,81650,81514.3
236414,825103,72425.0
234149,79433,56322.4
%H Mean15.5%H Mean23.5
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

Wolfart, J.; Staub de Melo, J.V.; Manfro, A.L.; Barra, B.S.; Barbosa, R.C. Effects of the Combined Incorporation of ZnO and TiO2 Nanoparticles on the Mechanical, Rheological, Thermal, and Healing Properties of a Dense Polymeric Asphalt Mixture. Nanomaterials 2025, 15, 1779. https://doi.org/10.3390/nano15231779

AMA Style

Wolfart J, Staub de Melo JV, Manfro AL, Barra BS, Barbosa RC. Effects of the Combined Incorporation of ZnO and TiO2 Nanoparticles on the Mechanical, Rheological, Thermal, and Healing Properties of a Dense Polymeric Asphalt Mixture. Nanomaterials. 2025; 15(23):1779. https://doi.org/10.3390/nano15231779

Chicago/Turabian Style

Wolfart, Jaqueline, João Victor Staub de Melo, Alexandre Luiz Manfro, Breno Salgado Barra, and Rafael Cassimiro Barbosa. 2025. "Effects of the Combined Incorporation of ZnO and TiO2 Nanoparticles on the Mechanical, Rheological, Thermal, and Healing Properties of a Dense Polymeric Asphalt Mixture" Nanomaterials 15, no. 23: 1779. https://doi.org/10.3390/nano15231779

APA Style

Wolfart, J., Staub de Melo, J. V., Manfro, A. L., Barra, B. S., & Barbosa, R. C. (2025). Effects of the Combined Incorporation of ZnO and TiO2 Nanoparticles on the Mechanical, Rheological, Thermal, and Healing Properties of a Dense Polymeric Asphalt Mixture. Nanomaterials, 15(23), 1779. https://doi.org/10.3390/nano15231779

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